Freshman Organic Chemistry

Video Lectures

Displaying all 37 video lectures.
Lecture 1
How Do You Know?
Play Video
How Do You Know?


Professor McBride outlines the course with its goals and requirements, including the required laboratory course. To the course's prime question "How do you know" he proposes two unacceptable answers (divine and human authority), and two acceptable answers (experiment and logic). He illustrates the fruitfulness of experiment and logic using the rise of science in the seventeenth century. London's Royal Society and the "crucial" experiment on light by Isaac Newton provide examples. In his correspondence with Newton Samuel Pepys, diarist and naval purchasing officer, illustrates the attitudes and habits which are most vital for budding scientists - especially those who would like to succeed in this course. The lecture closes by introducing the underlying goal for the first half of the semester: understanding the Force Law that describes chemical bonds.




Transcript



September 3, 2008



Professor Michael McBride: Okay. So for 150 years organic chemistry courses have tended to acquire a daunting reputation. So you need help. I know you're very able, but trust me, you need help. So where do you [get] the help? The PowerPoints are available on the Web. How many of you've already seen the PowerPoint for today, just so I have some idea? So about a quarter of you maybe. Okay, but anyhow, so your lecture notes are important, but you don't have to worry about getting everything down because you can download it from the Web. And I do it on a Mac, but I try my best to make it compatible with PCs and even with the free PC Viewer for PowerPoint. So you should be able to see it. But I don't see it on a PC. So if anything doesn't come through, let me know so that I can fix it. Okay then, in-class discussion is very important, and if you're really, really shy and can't participate in discussion in class, then email me a question. Okay?



There's the course website, which is our substitute for a text. It also includes the PowerPoint, and there's the link for it, and when go there you'll see it develop. The current website is mostly last year's course, so it'll change a little bit as we go along, but fundamentally it's the same. If you want to look ahead you'll see pretty much what's coming up. There'll be assigned problems and questions, and also there are previous exams and answers keys. All these things are on the course website so you'll get help there. But one thing that's really special is the course Wiki. This is the third year we've done it and the second year in a really systematic way. So you get assigned to do, to cover a couple frames of the PowerPoint. So those are the ones you really need to take careful notes on, and write them up, and help other people too; that's the nature of a Wiki, as you know. How many of you have participated in a Wiki? Well, by next week it will be all of you. Okay, but in order to get credit for it you have to get it by the night after the lecture. So for the lecture today you have to get it by late tomorrow night, 36 hours after the lecture. This is so other students can use it. Okay, in the spring there'll probably be a textbook. I haven't really decided yet. These things cost an arm and a leg. Maybe we can find one that's used, an older edition. It doesn't make any difference except to the publishers.



Okay, also there's personal help, like from me, and there's--and you can find my phone number, email and so on, on the website. Also the two graduate student TAs who are assigned to the course, who are Filip Kolundzic--Filip, back there--and Nathan Schley, not Schlay, Schley. So these are graduate students in chemistry and they run these discussion sections. Typically you have a 50-minute discussion section. But the way we run it in this course is that on two different nights a week there are two-hour sessions. You can come to any part of it you want to. You can go to both of them, you can go to four hours a week if you want to, or you can go to none at all if you want to. So really, for the bookkeeping purposes of the department you have to sign up for a section. Sign up for any section you want to and then come to what's useful for you. But also, the reason you pay the big bucks to come here, is not to hear me, it's to interact with the other students. That's a really big help. So, form study groups. And in fact you can get advice from previous people who've taken the course. That's on the Web. Also there's some of them, there's a list of them on the Web who would be happy to talk to you if you need it.



And we're blessed with three alumni, seniors who took this course as freshmen, who act as what are called peer tutors, and they'll run a session Sunday evening, from eight to ten p.m. is the current plan. We'll announce the rooms for these things on the website and probably by email to you as well. So let me introduce Tina Ho and Drew Klein and Justin Kim. So they'll be a big help to you too. So there's plenty of personal help, so use it.



These are the dates we're going to have exams. There are ten lectures and then an exam; nine lectures, exam; nine lectures, exam. Actually, if you check, you'll find that--and also you get 50 points for participation in the Wiki, and the total is 650 points, that's what your exam is based on. Actually this doesn't cover--it's nine lectures that are covered on the exam, but the previous Wednesday part of the lecture is going to be a guest lecturer that's going to be here just that day. So we're putting the exam off and it'll only cover the previous material; not that that's a big deal. Okay, and the semester grade is biased; that is, it's based on this, your total score here, out of 650 points. But if you're near a cutoff and you were very good about turning in your problem sets and so on, then we boost you up. We don't grade problem sets but it's worthwhile to do them, and they might make a difference.



So where are we going with this? What are the goals of our Freshman Organic Chemistry? In fact, if you click that in your PowerPoint you'll get taken to that site, but it's right on the website, you'll see it anyhow. First is to learn the crucial facts and vocabulary of Organic Chemistry--after all that's what we think we're here for--and to develop a theoretical intuition about how bonding works. This is the goal for, the primary goal of the first half of the fall semester is to learn how bonding works really, and that relates then to molecular structure; and also how bonding changes, and that of course is reactivity. But under the line there are a lot of other things that we do in Freshman Organic Chemistry that are arguably just as important, like make the scientific transition from school to university. In school they try to teach you what people know. In the University you try to develop new knowledge. So you need a different mindset for that, and we hope this course helps you develop that. So learn from Organic Chemistry, which is really in my view a model science, how to be a creative scientist.



So here's a creative scientist by anybody's measure, Louis Pasteur. And in the 1880s he said this in French, but in English it says, "Knowing to be astonished by something is the mind's first step toward discovery." Another way of putting that is that the characteristic comment on making a real discovery is not "Eureka", it's "Huh, that's funny." So that's what you really have to learn; learn enough about how chemistry works and form this picture in your mind that when something happens that doesn't fit, you know to be astonished so that you can discover something. That's exactly what Pasteur did, and we'll talk about that in the course. And even perhaps more important, to develop good taste, so that you can distinguish sense from nonsense; there's certainly more nonsense floating around than sense, and being able to tell the difference is important. The way you do it is to develop good taste by looking at a lot of good examples and then you're aware of how crummy the bad examples are. So we're going to try to emphasize good examples, and have fun. So and as we go along, if you have questions, break in. You'll do this much more as we go along, I know.



So the class really is mostly about theory, although we describe the basis for the theory and spend a lot of time trying to make it real. But we require Chemistry 126L, the lab. This is the only chemistry course that requires you to take a lab simultaneously. So I hope you're all enrolling in that because there'll be a certain day that you want to be able to take it. It's just one afternoon a week, three hours or whatever it is, but you want to get your first choice, so line up soon. You'll be accommodated but it's just more convenient if you get it arranged earlier.



But why? Because lab answers the really big question. And the big question was brought home to me by my son, John McBride, in his third year. This was the beginning of the third year, and his mother and I didn't know what was coming. For the next year, maybe 15, maybe 20 times a day, he said, "How do you know?" So here's John this last summer. He's now 38 and he has his own three-year-olds to say that to him. And he doesn't say, "How do you know?" anymore. He now says, "How do you know?" Okay? But that is the main question, how do you know what these things that they told you in school?



Well, there are four ways we can talk about of knowing, and two of them are shown on this manuscript from the Carolingian book painter. If we zoom in on the top frame, here's Moses on Mount Sinai. So the first way of knowing is divine authority. Here he's going to be the--here's the graduate student here getting the word. Here's the teaching assistant over on the left perhaps. [Laughter] Aaron, right? And then he comes down from Sinai, to see the class, the Children of Israel. So here's another kind of authority, which is human authority interpreting the scriptures. And here you can see the class; the guys are going like "huh." And the teaching assistant is off on the side still. But this doesn't make it. Science is not faith based. There may be other things you know that way but not science. Science ignores divine authority and it ignores human authority; not that they might not exist but they don't relate to science.



Now as you walked in today, did you notice these things over here? There's an Honor Roll of Chemists, and in fact we'll use that a lot this semester, and in particular one of the people on there is Michael Faraday, who started in a very humble way. He was a book binder's apprentice and he bound this book--not this particular copy but this book--which is called Conversations on Chemistry. He bound the first edition. This one is a later edition. So you see it's Conversations on Chemistry in which Elements of that Science are Familiarly Explained and Illustrated by Experiments. And who's the author? J. L. Comstock; he's actually not the author, he's the guy who stole it. He stole it from a woman, Mrs. Marcet, in England, who wrote this book, which was the most popular textbook--it was written for girls--but it was the most popular textbook in all chemistry, for the first half of the nineteenth century. It went through like 20-some editions. And here you see at the beginning it's a dialogue, a conversation between Mrs. B and Caroline and Emily, and it's fun to see this here, what Emily says at the beginning. "To confess the truth Mrs. B, I'm not disposed to form a very favorable idea of chemistry, nor do I expect to derive much entertainment from it." But in the long run, as you can imagine, they have a lot of fun with chemistry. It was a wonderful book, and still is. But he was binding it, and read it. And look what he says about this, as his introduction to be the leading experimental scientist of the nineteenth century:



"Do not suppose I was a very deep thinker or was marked as a precocious person. I was a very lively, imaginative person and could believe in the Arabian Knights as easily as the encyclopedia, but facts were important to me and saved me. I could trust a fact and always cross-examined an assertion. So when I questioned Mrs. Marcet's book by such little experiments as I could find means to perform, and found it true to the facts, as I could understand them, I felt I had got hold of an anchor in chemical knowledge and clung fast to it."



So the experiments were what did it. So the third way of knowing is by experimental observation. And here's Richard Feynman. How many of you've heard of Richard Feynman? He was a really great physicist, wrote a wonderful textbook as well as getting all sorts of prizes. He spoke to the National Science Teachers Association in 1966 saying,



"Learn from science that you must doubt the experts. Science is the belief in the ignorance of experts. When someone says, 'Science teaches such and such,' he's using the word incorrectly. Science doesn't teach it; experience teaches it. If they say to you, 'Science has shown such and such,' you might ask, 'How does science show it? How did the scientists find out? How, what, where?' Not science has shown, but this experiment or this effect has shown."



Now, why do we quote Feynman? Because he's an expert. [Laughter] Wrong. Though literally, expert, the etymology of expert, is it means someone who has done experiments. We quote him because what he says makes sense. So logic is the fourth way of knowing things. So the two ways that we know things in chemistry, or in science, are experiment and logic. And the lecture is a little bit more focused on logic and the lab is more focused on experiment, and you get an unbalanced view if you do one without the other.



Okay, so modern science got underway in the seventeenth century. There's the seventeenth century, 1600 to 1700. And 1638 was when New Haven Colony was founded, and 1701 was when Yale was founded. So that's when everything got underway, just when this enterprise was beginning here. Here we are. If you go back 100 years you get to quantum, quantization by Planck; and we'll talk about that. And if you go back another 100 years you get to Lavoisier and oxidation; and we'll talk about that. And if you get another 100 years you get to Newton and gravitation; and we'll talk a little bit about that. And if you go back a little more, another 100 years, you get to Copernicus and the revolution of the heavenly bodies, and Columbus and navigation, and Luther and the Reformation. And these things all have something in common. As Robert Hooke wrote, "The seventeenth century" (his age) "was an age, of all others, the most inquisitive."



All these things have to do with people inquiring into how people know things and finding out new things. And in particular an important figure was Francis Bacon and his Instauration. Now you may not know The Instauration so well, let's look at that. Here's Francis Bacon, there are his years. He was Elizabethan and Jacobean. He was almost exactly contemporary with Shakespeare, and with Galileo. He went to school, to university, at Cambridge. And here's a cartoon that shows him--it's a modern cartoon--imagining him in a class at Cambridge. Because he wrote of his tutors at Cambridge:



"They were men of sharp wits, shut up in their cells of a few authors, chiefly Aristotle, their dictator. All the philosophy of nature" (philosophy meant science in those days) "all the philosophy of nature, which is now received, is either the philosophy of the Grecians or that of the alchemists. The one is gathered out of a few vulgar" (that means ‘common' of course) "observations, and the other out of a few experiments of a furnace. The one never faileth to multiply words, and the other ever faileth to multiply gold."



So here's the book he wrote, The Instauration. That's the frontispiece for it. This picture's from the Beinecke Library; I went down and got a picture of the book. Notice it was published in 1620. What else happened then? That's when the Pilgrims came over, right? So the title of the book, rather small under his name and title as Lord Chancellor of England, is Instauratio Magna, which means the Great Restoration. Restoration of what? Of the way of knowing. A bigger, a part of it, it's called the Novum Organum, which is--and it develops the inductive scientific method, based on experiment, to replace Aristotelian deduction, which is you maybe did one experiment sometime, and then you reason everything from that. But he says no, you have to do more experiments. Now there's an interesting thing here. One of the devices on the title, on this frontispiece, is two pillars. What in the world are they doing there? Well they're the same pillars that you see on this. That's a piece of eight; you know, Treasure Island, pieces of eight? See it's eight reales, and it came from the silver of Mexico; it was minted in Mexico City. So and there you see the same pillars and on them it says plus ultra; more beyond. Beyond what? What are the pillars? Pardon me?



Student: Spain and Africa.



Professor Michael McBride
: Yes, it's Africa and Spain, but it's the Pillars of Hercules, which are the mouth of the Mediterranean, the old Classical World. So there's the Mountain of Moses in Morocco and the Mountain of Tarik, which is the name of Gibraltar. So here's the Mediterranean, the Classical World of Aristotle, and you can sail out into the New World and bring back silver, for example. There's danger of course. But look at what it says at the bottom. What will be brought back? Not just silver. Multi pertransibunt & augebitur scientia --"Many will pass through and knowledge will be increased." So we go beyond Aristotle into experimentally based science and knowledge will be increased.



So here's some quotes from The Instauratio Magna. "That wisdom which we have derived principally from the Greeks" (no offense, okay?) "is but like the boyhood of knowledge, and has the characteristic property of boys: it can talk but it cannot generate;" "…it is but a device for exempting ignorance from ignominy." That means it's a way of hiding your ignorance, and we'll see examples of that. We'll talk about, in Lecture 11, about correlation energy, and we'll talk in Lecture 32 about strain energy, and you'll see that both of these are just words that are used to hide our ignorance. "…the end which this science of mine proposes is the invention, not of arguments, but of arts" (ways of doing things). "…not so much by instruments" (although new instruments are important, like microscopes and so on) "as by experiments, skillfully and artificially devised for the express purpose of determining the point in question." (So artificial experiments designed to decide a question; experiments.) "And this will lead to the restoration of learning and knowledge."



So followers of Bacon established The Royal Society in 1662, just after Charles was restored to England, after the period of Cromwell. And there was a history written of The Royal Society, a book about this thick, published in 1667, only five years after it was founded. Why did they publish a history so soon? Well let's look at this, the frontis itpiece of this book. Here's the late Francis Bacon, who was said to be Artium Instaurator, the Restorer of the Arts. And here's the President of The Royal Society, the mathematician, Viscount Brouncker, and here in the middle, being crowned with laurel, is Charles II. Why do they have him up on a pedestal? Because they're hoping, as scientists have before and ever since, to extract some money out of the government to do their research. They actually never got it from Charles but it wasn't for lack of trying. And this is why they wrote the history, to try to make the case for being supported.



Okay, now let's look at all the good things that will come from science, from The Royal Society. In the background can you see what that is? We'll blow it up. Here there's a hint to it on the bookshelf. If you look really fine on the bookshelf you can see that some of them have writing on the spine. Do you see what that one is? Can anybody read it? What? What science book do you think they might have had?



Student: Copernicus.



Professor Michael McBride: Copernicus, right? So astronomy; that's a telescope in the back. Okay, or over here on the wall, what's that thing? It's a clock. Why is it shaped like a piece of pie?



Student: Because it has a pendulum inside.



Professor Michael McBride: Ah, because it has a pendulum inside. So horology, making good clocks. Okay, or here, what's that thing? It's hard for you to know.



Student: Solar.



Professor Michael McBride: It's a wind gauge. It has a vane inside that it blows on--and you can tell from how far it goes on the scale how strong the wind is. So meteorology. And back here on the pillar, what are those things for?



Students: A compass.



Professor Michael McBride
: For cartography. Now what do all these things have in common that they're going to do for Charles II? Astronomy--



Student: For the records.



Professor Michael McBride: Good clocks, meteorology, cartography, what--



[Students speak over one another]



Professor Michael McBride: They all have to do with navigation, with making England strong at sea. Now back here is another science, Chemistry. That doesn't seem to have anything to do with navigation. Why would Chemistry be important to Charles?



Student: War.



Professor Michael McBride
: Look at it.



Student: War.



Professor Michael McBride
: To make gun powder. There was a chapter about gun powder in The History of the Royal Society. Okay, and up here at the top is the motto, which is Nullius in Verba, which comes from this quote from Horace, and what it says is, "Lest you ask who leads me, in what household I lodge" (that is, what philosophy I advocate) "there is no master in whose words I am bound to take an oath. Wherever the storm forces me, there I put in as a guest." So it's the experiment, not the philosopher, that leads you to the conclusion. So Nullius in Verba is 'in the words of none.' And, in fact, the original name of The Royal Society was The Royal Society for the Improving of Natural Knowledge by Experiments.



Okay, so we're going to see, as the course goes along, important experiments that really decided questions, and in fact Bacon's most important kind of experiment was one that "finally decides between two rival hypotheses, proving the one and disproving the other." So you can do all sorts of experiments and just be collecting butterflies--no, I don't mean to insult people who collect butterflies, it's a fine thing to do. But there's something special about experiments that really are designed to answer a question. Now Bacon devised a name for such experiments, and they're based on this model, that you have a road that diverges and you need to know which way to go between these two hypotheses. What do you need, to know which way to go? You need a sign, or this was a cross that you mounted at a crossroads. So the Latin name for cross is crux. Do you see what they call the experiment? Crucial. That's the origin of the word crucial, right? It's the one that tells you which way to go. Okay, so here is Isaac Newton and he's holding something. Can you see what he's holding in his hand?



Student: A candle.



Professor Michael McBride: I'll give you a hint. It's a prism. Why is he holding a prism? Because that was his crucial experiment; and here's his diagram in that experiment with a prism in it. He called it the Experimentum Crucis, taking the word from Bacon. Okay, so light came in through a hole in the window, through a lens, and then got bent and dispersed into the different colors. So you get a spectrum here on this thing, the different colors. Now there's the question, how does the prism make color? Hooke and Descartes thought that light was a train of pulses and as it goes through something like a prism, or as it reflects from a thin layer of oil on water or something like that, that it changes the timing of the pulses and therefore changes the color. But Newton thought that the colors were pre-existing, and the prism just separates them. And this was his crucial experiment to decide between those two theories. You see what he did? He drilled a hole through the board and let through only the red light, and put a second prism there. And he wrote here, three times--he wrote it here, and also here, and also here--nec variat lux fracta colorem; which means "the broken light does not change its color." So this proved to Newton, at least at that time, that light is a substance, not a train of pulses. What do you think of that proof now? You think light is a substance or a train of pulses?



Students: Both.



Student: Neither.



Professor Michael McBride
: But at least you can see that in terms then, that was a crucial experiment, a really important experiment, and that's the kind of experiments we'll try to talk about in the course. Experiments are indispensable in Organic Chemistry. It's an empirical science based on observation, and that's why you have to take the lab. But so is logic--that was number three and number four of the ways of knowing--logic is important too. So believe what I tell you here only when it makes sense to you. Don't just cram it in, make sure it makes sense. But what if it doesn't? Now here's how to succeed in Chem. 125, and we'll take as our model science student Samuel Pepys. How many people have heard of Samuel Pepys? Have you heard of him as a scientist?



Student: I don't remember.



Professor Michael McBride: What did you hear of him as? What do you associate with Samuel Pepys?



Student: Newton-Pepys Problems.



Professor Michael McBride: Right in this period, in the heart of the science growth. But what you know him for was his diary, which tells all about life, everyday life in Restoration London. He was actually, as a sixteen-year-old, present when Charles I was beheaded in 1649. Now what's the connection? Do you know where this is? Anybody been there? Dixwell, Goffe and Whalley Avenues, those are named for three of the 50 judges that condemned Charles I to be beheaded, and they were the only ones that lasted very long, after the Restoration, and they lasted because they fled to New England and were hidden on West Rock. All these roads are heading west toward West Rock. Okay, so that's a tie-in to the same period. But anyway, he got his B.A. in Cambridge in 1654 and a Master's in 1656. And he got a good job, he became Clerk of the Acts for the Navy Board, which meant he was the guy that purchased everything for the Royal Navy, all the rope, all the tar, all the lumber and so on. And on July Fourth, 1662--it's the fourth of July but it's more than 100 years before that became relevant--he writes in his diary, "By and by comes Mr. Cooper, mate of the Royall Charles, of whom I intend to learn mathematiques, and do begin with him to-day, he being a very able man… After an hour's being with him at arithmetique (my first attempt being to learn the multiplication-table); we then parted till tomorrow."



So here was the guy doing all the purchasing for the Royal Navy and he didn't know multiplication, let alone division. But he worked hard at it. July ninth, five days later: He's "Up by four o'clock, and at my multiplicacion-table hard, which is all the trouble I meet withal in arithmetique." He can do the other things pretty well. July eleven: "Up by four o'clock and hard at my multiplicacion-table, which I am now almost master of." Christmas--so six months later: "…so to my office, practicing arithmetique alone and making an end of last night's book with great content till eleven at night and so home to supper and to bed." Or a year later--so he was motivated and he was diligent; that's good. A year later, on a Sunday: "…I below by myself looking over my arithmetique books and timber rule. So my wife arose anon and she and I all the afternoon at arithmetique," [laughter] "and she has come to do Addition, Subtraction and Multiplicacion very well, and so I purpose not to trouble her yet with Division…" Right? [Laughter] So he worked with a study partner, and that's crucial. And Isaac Newton--does anybody recognize this book?



Student: Yes, sure.



Professor Michael McBride: Right? The Mathematical Principles of Science, of Natural Philosophy. But there's an interesting thing on the title page. Samuel Pepys is the one who gave permission to publish that book, because he was the president of The Royal Society. Now six years later Pepys encountered a problem with dice. The reason was he went to coffee shops every night for dinner and they'd gamble, and people proposed various kinds of bets, and he couldn't figure out this one, so he wrote Newton for help. So this was the problem that he wrote to Newton, twenty-second of November. So A has six dice in a box, and he has to throw a Six by throwing it; B has 12 dice and he has to fling two Sixes; and C has 18 Dice but he has to get 3 Sixes to win. The question, "Whether B and C have not as easy a Taske at A, at even luck?" That is, if the dice aren't loaded or anything, who has the better bet? [Laughter]



How many people think that it's the same? Don't be shy. How many think it's A? How many think B? How many think C? How many don't really have any opinion at all? Good, that wins. Okay, so he wrote this letter to try to get help on his bets from Newton, and Newton replied, four days later, "What is the expectation or hope of A to throw every time one six, at least, with six dyes?" So you get two sixes you still win; that wasn't clear in the original statement. So he says, "If we formulate the question that way, it appears by an easy computation that the expectation of A is greater than B or C; that is, the task of A is the easiest." There's the answer. So Pepys replied on the sixth of December: "You give it in favour of the Expectations of A, & this (as you say) by an easy Computation. But yet I must not pretend to soe much Conversation with Numbers, as presently to comprehend as I ought to doe, all the force of that wch you are pleas'd to assigne for the Reason of it, relating to their having or not having the Benefit of all their Chances." So he wasn't ashamed to admit that he didn't really understand--that's crucial. "And therefore, were it not for the trouble it must have cost you, I could have wished for a sight of the very Computation. "



Can you show me how to figure it out? And he wanted that because somebody might change the terms of the bet and then he wouldn't know. He wanted really to understand. He insisted on proof. So this is two of the pages from Newton's correspondence of the letter that he wrote in response, and you can see that Pepys certainly got more than he had bargained for. "So to compute this I set down the following progressions of numbers." So you can go through all this and you get complicated quotients here, and it turns out that A has 31,031 chances out of 46,656, or 0.6651 chance of winning, and B has this, which is 0.6187; A wins. So is Pepys satisfied with this? Pepys writes back and he says, "Why?" Right? "I cannot bear the Thought of being made Master of a Jewell I know not how to wear." So he's willing to swallow his pride to search for really solid understanding. Now compare this with a comment we got at an end-of-semester evaluation in January a year ago. "I never went to his office hours for help because I felt like he would make me feel stupid, because he's superior to me in chemistry." I hope I'm superior to you in chemistry. [Laughter] I'm not superior as a person, but I hope I'm superior in chemistry. You're paying me the big bucks because of that. So swallow your pride and ask someone for help. Follow Pepys.



So read "Pepys and Newton" and get together to do problems for Monday, and contribute to the Wiki when you're asked to do so. So here are problems. For Friday the problems are optional but very helpful. One is find out which two class members have the rooms nearest you so that you can maybe use them as study partners. You don't have to use them, use anybody, but use someone, don't try to go it alone. Two. What are the three most common items of advice from course veterans? So if you click there, or go to the webpage, you can get anonymous advice; I was careful to make sure it's anonymous. So you can get all--much of it is contradictory, so your job is to look through it and try to get some idea of what they're telling you that's worth knowing. For Monday there are problems from that webpage, "Pepys & Newton." And let me tell you that you better get together with other people to do it, because there's a lot of stuff there. So get together in a group, parcel out who works on what, discuss what went on and so on. And then for a week from Friday there's stuff about drawing Lewis Structures, from another webpage about functional groups. So we'll get to that later, I just wanted you to know what's coming up. Incidentally, this thing about the problems set that has to do with the mathematics that Newton and Pepys were working on there, has to do with isotope ratios. You know, chlorine has a funny atomic weight. Why? What is the atomic weight, does anybody know?



Student: 35.5.



Professor Michael McBride: Thirty-five and a half. Why a half? Most of the elements are pretty near integers. It's because it's a mixture of isotopes. It's a quarter of one isotope and three-quarters of the other, and the average is thirty-five and a half; thirty-five and thirty-seven. But in fact it depends on where you get chlorine from, what the ratio is. There's thirty-five and a half for a standard atomic weight, but it ranges quite a bit, depending on where you get it. So by measuring these ratios, which is a lot like these odds in the betting, you can tell where things came from. You can tell where hydrocarbons came from sometimes that way. And those are the problems that you have to deal with. So it's not really very, very relevant to the course as a whole, but it's relevant to the Pepys and Newton thing and it's a fun problem. So that's what you've got to do for Monday.



Okay, and here are the assignments. I emailed all these people. I hope you're here, and that you picked up on what we talked about today so that you can improve the Wikis from last year. Last year people developed their Wikis from scratch. This time--which is a very valuable thing to do. So I was torn this year as to whether to have you do them from scratch or whether to modify last year. I figured modifying last year is better because then you have something to read earlier on, for those who are not working on developing it. So anyhow, but the hope is--so if you modify it--it has to be something significant, not put a comma in or something like that, but add something useful to understanding. If you do it within thirty-six hours, you get two points, and if what you do is really good, you get three points. So you go to that page and you click on those things and then you can read what--or you can either modify them or read it; it's a Wiki.



Now, we come to the question that we'll address more. We have another five minutes here. The question we're really dealing with now, having seen how to work in Chem. 125, is are there atoms and molecules; how do you know? And what force holds them together? Because if we knew that they're real, and we knew what force held them together, then in principle we might be able to calculate everything. So it would all be theory. So is it springs that hold atoms together? So Robert Boyle--notice he's right at that time, 1627 to 1691, and he's the oldest person that's on the Honor Roll outside the building here, he's over that way--so Robert Boyle wrote this first important book in chemistry, New Experiments Physico-Mechanical: Touching the Spring of Air. So you know Boyle's Law, how pressure and volume change, that air in a piston acts like a spring. So PV is a constant; that's Boyle's Law. So he developed this science on the basis of a new instrument, the pneumatical engine, which was built by Robert Hooke, the guy we quoted. And there's a picture of the air pump here with that crank on it, and so on to pump things in or out of that bulb. Now a couple of years later Hooke wrote this book, Lectures, De Potentia Restitutiva, or of Spring, explaining the power of springing bodies. And this is the beginning of that book. "The Theory of Springs, though attempted by divers eminent Mathematicians of this Age has hitherto not been Published by any. It is now about eighteen years since I first found it out, but designing to apply it to some particular use, I omitted the publishing thereof." So he kept it secret so no one would steal the idea. What did he hope to do with a spring, a really important technology that he could do with a spring? There was a coil in the spring that he depended on.



Student: Flying.



Student: Pendulum clock.



Professor Michael McBride
: A new kind of clock, one that could work better for navigation. It was his spring that actually allowed Harrison, 100 years later, to win the contest about making an accurate clock; some of you must know that story. So anyhow he had been hiding this so it wouldn't be stolen. In fact, it was stolen by Huygens. So "about three years since His Majesty was pleased to see the experiment that made out this theory, tried at White Hall, and also my spring watch. About two years since I printed the theory in an anagram at the end of my book on the description of helioscopes, viz." this. So this is what he published at the end of this book on helioscopes, and that is an anagram of how springs work, so that later he could prove that he knew it, if needs be, but no one could steal it in the meantime. And if you unscramble it, it's "Ut tensio sic vis; that is, the power of any spring is in the same proportion with the tension thereof. That is, if one power stretch or bend at one space, two will stress at two, three will bend at three, and so forward." So here's his figure that shows that. That's Hooke's Law, the force law. The force is proportional to the distortion. So here's his figure, and he's got plots of it here. Ut tensio (as the extension) sic vis (so the force). So it's linear, the force is how much you stretch it. And he has lots of different kinds of springs that do that. Here's this clock kind of spring, a coil spring, just stretching a wire. So the force is proportional to the extension. Or another way of saying it is, "The potential energy is proportional to the square of the extension." So it's a parabola. So that's Hooke's Law, and that's where we'll take up next time.



[end of transcript]

Lecture 2
Force Laws, Lewis Structures and Resonance
Play Video
Force Laws, Lewis Structures and Resonance


Professor McBride begins by following Newton's admonition to search for the force law that describes chemical bonding. Neither direct (Hooke's Law) nor inverse (Coulomb, Gravity) dependence on distance will do - a composite like the Morse potential is needed. G. N. Lewis devised a "cubic-octet" theory based on the newly discovered electron, and developed it into a shared pair model to explain bonding. After discussing Lewis-dot notation and formal charge, Professor McBride shows that in some "single-minimum" cases the Lewis formalism is inadequate and salvaging it required introducing the confusing concept of "resonance."




Transcript



September 5, 2008



Professor Michael McBride
: Okay, let's start up. So last time we got to the point of asking the question, are there atoms and molecules--and well, I anticipate that the answer is yes--and what force holds the atoms together? Because if we really understand the force that holds them, we've got chemistry all solved. So when you come to the idea of forces, in the first place, one goes back to Newton and Principia Mathematica. And he had the idea, or developed the idea, the mathematics of gravity. And the interesting thing about it was there was no mechanism for it. There were no springs or strings that connected the things. It was action at a distance.



So there were other ideas. Like, Descartes had the idea that gravity resulted from blocked repulsion, that the universe was full of particles zipping around and bombarding things. So like the moon would be bombarded by particles from all directions, but there'd be no net force on it because they'd all balance out. However, if you put the earth in the way, there'd be a shadow, and now you'd have more things hitting the moon from the right than from the left. So it would seem to be attracted to the earth. So that's a great mechanism. Now, it was important for Newton that the force be proportional to r2, to inverse r2, 1/r2. Does this work for that? If you pulled the earth back, how would you change things? What do you say?



Student: I don't know. It doesn't work for this.



Professor Michael McBride
: It doesn't work? Would it block more or fewer things if you pulled it back?



Students: Fewer.



Professor Michael McBride: How many fewer? Suppose you made it twice as far back, how many fewer would be blocked from hitting the moon?



Student: Wouldn't it be r2?



Professor Michael McBride: It would be 1/r2. You can think about that. So that aspect of it would work. How about--you know gravity is proportional to the mass of the things. How does that figure? You can think about that yourselves, okay? But anyhow, there were other ideas. But Newton didn't think about the mechanism, he thought about the law, that it was 1/r2 and proportional to the masses.



Now in 1717, he published the Second Edition of Opticks. He published the First Edition in 1704, and the reason he published it in 1704 was that Robert Hooke died in 1703. He'd actually--Newton had done this work in the 1660s and '70s, but he didn't publish it until Hooke died because he didn't want Hooke to get on his case, criticizing it. But in the Second Edition, in 1717, he put some additional material; you see, "with Additions. " And the additional material, even though the book is about optics, the additional material is about chemistry. And it's at the end, and posed not as hypotheses but as questions, because Baconian people, following Francis Bacon, weren't supposed to make hypotheses, but they could ask questions. So here are some of the questions. Question thirty-one, the one that has most to do with chemistry:



"Have not the small Particles of Bodies certain Powers, Virtues, or Forces by which they act at a distance, not only on the Rays of Light to reflect, refract and inflect them, but also upon one another for producing a great part of the Phaenomena of Nature?" (And this attraction between the particles or repulsion) "For it's well known that Bodies act upon one another by the Attractions of Gravity, Magnetism and Electricity; and these Instances shew the Tenor and Course of Nature, and make it not improbable but that there may be more attractive Powers than these. For Nature is very consonant and conformable to her self. How these Attractions may be perform'd, I do not here consider. What I call Attraction may be perform'd by impulse…"



Where did he get that idea? Where did he get the idea that there are impulses that can cause attraction?



Student: [Inaudible]



Professor Michael McBride: That was Descartes' idea, that we just explained. "…or by some other means unknown to me. I use that Word here to signify only in general any Force by which Bodies tend toward one another, whatsoever be the Cause. For we must learn from the Phaenomena of Nature what Bodies attract one another, and what are the Laws and Properties of attraction, before we enquire the Cause by which the Attraction is perform'd." (So get the law first so you can do mathematics on it, and then worry about how it works.)



"The Attractions of Gravity, Magnetism and Electricity, react to very sensible distances…" (He doesn't mean reasonable, he means ones that you can sense.) "…and so have been observed by vulgar Eyes." (You can see the distances over which electrostatic attraction works, obviously gravity, magnetism.) "And there may be others which reach to so small distances as hitherto to escape observation; and perhaps electrical Attraction may reach to such small distances, even without being excited by Friction." (So you know that friction can generate static electricity that'll make things attract, but maybe even without friction you can get things that'll attract a very short distance by electricity.) "The Parts of all homogeneal hard Bodies which fully touch one another, stick together very strongly." Did you ever have that experience of having two really flat things, put them together and they're very hard to get apart; what?



Student: Slides.



Student: Small things.



Professor Michael McBride: Microscope slides. Ever take out new microscope slides? They can be very hard to get apart. "And for explaining how this may be, some have invented hooked Atoms, which is begging the Question; and others tell us that Bodies are glued together by rest, that is, by an occult Quality, or rather by nothing, and others that they stick together by conspiring motions, that is by relative rest among themselves." (These are other theories that are sort of complicated; we don't need to go into it.) "I had rather infer from their Cohesion, that their Particles attract one another by some Force, which in immediate Contact is exceeding strong, at small distances performs the chymical Operations above mention'd, and reaches not far from the Particles with any sensible Effect." (So a very short-range, a very strong attraction.) "…reaches not far from the Particles with any sensible Effect."



So maybe it's even more dramatic than 1/r2. 1/r2 changes very rapidly when r gets very small, but maybe it's even a law that's more distance dependant than that, but only works at really, really small distances. "The Attraction" (and he's talking about between glass plates separated by a thin film of the Oil of Oranges; so you know, microscope slides, they put a little drop of something and stick them together and they're really hard to get apart) "may be proportionally greater" (the attraction) "and continue to increase until the thickness do not exceed that of a single Particle of Oil." (Imagine Newton, 350 years ago or 300 years ago here, thinking about a single particle of oil. Could he possibly measure that? We'll think about that later. Okay.) "There are therefore Agents in Nature able to make the Particles of Bodies stick together by very strong Attractions. And it is the business of experimental Philosophy to find them out." That's what we need to find, what is it that holds these particles together? "Experimental philosophy" means science. And we're going to engage in this business, and it'll take us about five weeks to figure out what it is that holds these things together.



Now, let's just look, from physics, at the amount of binding energy you get from different kinds of laws. First there's Coulomb's law; so it charges over the distance. I think you all know that one. Then there's magnetic interaction, two magnetic dipoles, and the energy is 1/r3, falls off very rapidly. Then there's the "strong" binding that holds particles together in the nuclei of atoms. And gravitation, of course. And the chemical bond. So here's a scale of energies, and we use, in this course, the old fashioned kilocalories/mol, which American organic chemists like to use. Some day it'll change to kilojoules/mol, but for me it's still kilocalories/mol. Okay, this is a logarithmic scale, so it covers a very wide range.



Now Coulomb, if you have a proton and an electron, as far apart as the distance of the carbon-carbon bond, that's 216 kilocalories/mol, right in the middle of this, or near the middle. Magnetic interaction is much weaker. If you have two electron spins, at the same distance from one another, their energy is what? One, two, three, four, five orders of magnitude smaller, 105th smaller. The "strong" binding that holds a proton and a neutron together is in the deuterium nucleus, that is within the nucleus, not at 1.54 Ångstroms, is 105th stronger than Coulombic interaction. And gravity is ever so much weaker. If you had two carbon atoms, that were at that distance, the energy would be 10-32 kilocalories/mol. So forget gravity as a way of holding things together. But then we have Coulombic up or down by 5 orders of magnitude.



Now what is a chemical bond; how strong is it? A chemical bond is about--the carbon-carbon bond is about 90 kilocalories/mol. So not so far from the Coulomb energy, from electrostatic attraction. So maybe this hints that electrostatic attraction has something to do with it. But maybe not, maybe it's something entirely new. And we haven't talked about kinetic energy. Does that have anything to do with the strength of the bond? So we'll get back to that later on. Okay, so is there a chemical force law, the thing that Newton was looking for? Now here's an interesting way to phrase that. Suppose you had a chain of atoms, a thread that was single atoms bonded to one another, and you started stretching it. It would stretch, and at some point it would pop.



Now let's just take three of the atoms and we'll put stiff rods on the outer ones and we'll start stretching it. So it'll stretch, stretch, stretch, stretch, pop. Now the question is how far do you stretch it before it pops? What determines how far you stretch it before it pops? Okay, this must have to do with the force law between atoms, which then will have to do with molecular structure. So it could be like springs, Hooke's law, Ut tensio, sic vis. Okay, so the force is proportional to the displacement, and the energy to the square of the displacement and the energy then, the potential energy, is a parabola. We define zero when they're at the standard, unstretched distance, and then you stretch it or compress it and the energy goes up, like a parabola. Now the slope of that parabola tells you what?



Student: The force.



Professor Michael McBride: The force. That's what's linear. As we stretch it further and further, the slope gets bigger. The force is proportional to the displacement. Okay, now there could be other kinds of force laws, like electrical charges, Coulomb, or gravity. So that's a different force law, inverse square force law, and the energy then is proportional to the displacement, in this way. So very different. Notice the zero of this one's at the top when--zero defined when they're very far apart. Here it's zero when they're at the standard distance. And so as you bring oppositely charged particles together, the energy goes down infinitely, as they get really close. Okay, and now if we look at the force here, as we go out the force gets smaller, the slope gets smaller.



Now suppose you had a particle that was connected to a spring on each side, as in this chain that we're stretching. And just to be general I made the second string--oh I say stronger, it's in fact weaker; it doesn't go up as rapidly--I'll have to change that. Okay now, but at some point the slopes of the red and the blue, at some distance, some position, will be equal and opposite. There. Okay, what's the force on the central particle at that point?



Students: Zero.



Professor Michael McBride: Zero. They balance. So if we added those two energies together to get the sum of the two, on this central particle, it looks like that. So there's a balanced minimum. There's a well defined position where the particle in the center of this chain will just sit there and it'll be hard to displace. So it can vibrate. Okay? Now how about in the other case? How about if it were electrical charges or gravity or something like that, an inverse square law that's holding things together? Then the second one is going to look like that, and that one is indeed stronger.



So they're two flanking bodies and we're interested in the position of the one in the middle. And again, there'll be some point where they balance, right there, where they're equal and opposite slopes. How will this differ from the first case? Can you see what's going to happen when we add these two together? Is it going to look more or less like the one on the left?



Student: [Inaudible]



Professor Michael McBride: It's going to look like that. And now there is no balanced minimum in the middle. There's not a position where the thing will just sit there, because it's always more strongly affected, and attracted, by the one it's closer to. There's some place where there's zero force but it's not a stable position. Because if it displaces ever so slightly, it'll keep on going. So that one is a--for the two of them--is a single minimum, and this one is a double minimum. And for the people who came early, there was a contest that if you won it, you could get an A in the course, and the contest is this. Here's a magnet, hanging here. And I'll stop it. Okay, and here are two magnets, the ones that will flank that one. And if I get it just right, it'll just sit there, the attraction to the two will balance it. So I'll put it on here, and try to get my A. So I'll get the string where it would balance right there. Ah. Okay? There is no stable position. Nobody got their A that way. Right? Because it's an inverse law, so it's always more strongly attracted to the closer thing. So you can't win with that, you can't get a balanced position. Now I'm going to get these off here. And back to the show.



So with springs you might be able to make a stable polyatomic molecule from point atoms--we saw a spring model before in the last lecture--but you can't do it with ions and you can't do it with magnets. However, Hooke's law can't do it--it can't be springs, because Hooke's law never breaks. So we need a different kind of force law. Does anybody know a force law--what it looks like--the form of the force between atoms or the energy for stretching a bond? You ever seen a picture of such a thing?



Student: [Inaudible]



Professor Michael McBride: It's called the Morse potential, or one form of it at least is called the Morse potential. And it's not something fundamental, it's not a law of nature. It was thought up by a physicist at Princeton in 1929, because it's mathematically convenient; so you could solve quantum mechanical problems if you used Morse potentials. And the idea is this. You have two neighbors and you look at the position of the one on the right and it could be there and the energy would be minimum; that's the bond distance, right? And if you move it further to the right, the energy goes up; move it to the left, the energy goes up. Okay, that's the Morse potential, and we're just holding the neighbor on the left fixed. Now we could put another neighbor on the other side and have another curve of the same sort; stretching the second bond. Or we could have it be a chain and have both neighbors there, and the sum of those two now has a minimum in the middle.



It's a single minimum, like we got with Hooke's law. So that means that atom in the middle would just stay there. And now we're going to grab those sticks on either side and pull the neighbors apart. So if we pull them apart, it's still a single minimum. And if we pull them apart further, it's still a single minimum, although it's very flat. And if we pull it further, it pops and it becomes a double minimum. And if we keep going the double minimum gets more pronounced. Now here's an interesting question. At what point--how far do you have to stretch it before it pops? That's the question we were asking, right? How can you look at those curves and tell how far apart you have to stretch it before the chain pops? Shai?



Student: The bottom is flat. So when the curve and the plot was flat, then that's as far as it could go because…



Professor Michael McBride: What do you mean flat? There's a name for a thing like that in graphs.



Student: Inflection.



Professor Michael McBride: It's the inflection point. It's when the curvature changes from being this way, like Hooke's law, to being this way, right? So when the inflection points cross, then the chain pops. Okay, so force laws are going to make a lot of difference in how atoms behave, and if really knew the force law we'd be in a great position for understanding chemistry. It snaps at the inflection point where it changes from a direct to an inverse dependence on distance for force.



Okay, but what are bonds? Newton said, "We'll look for the law." And it's sort of, the law is sort of like Morse, but we don't know where it came from, because we don't know how it works, we don't know what it really is. So can we find out what it really is? Well in the nineteenth century they discovered bonds. And this is a picture from 1861, we'll talk about this later. It's one of the first depictions that's recognizable by people today of bonds between atoms. So there are different numbers of these lines, valences, for different atoms. Hydrogen has one; carbon has four; oxygen two; nitrogen three.



Why do the elements differ? Why don't they all have the same valence? And even more complicated, sometimes nitrogen is three and sometimes it's five. For example, you have NH3 but you also have NH4Cl, where there are five things associated with the nitrogen. So how do you understand the valence? That was the challenge for people. They had figured out that there was valence, but why? How could you predict the valence of a new element? Well Gertrude and Robert Robinson published a paper in 1917 which showed this picture of ammonia, NH3. What's the loop? Pardon me?



[Students speak over one another]



Professor Michael McBride: Someone said--hold your hand up so that people can hear you. Okay yes.



Student: A lone pair on nitrogen.



Professor Michael McBride: A lone pair on nitrogen--wrong. It looks like the lone pair on nitrogen but the lone pair didn't come along that soon. What the loop means, if you look at it in the context, is what makes NH3 reactive? In this term, this is what they wrote: "It may reasonably be assumed that the partial dissociation is a stage in the complete process." So the bonds begin to break. This loop begins to break to become a weaker loop and two other partial valences, in this theory. And those valences begin to associate with partial valences from HCl. Right? The loop is a "latent" valence. You know what 'latent' means? Latent is the opposite of patent. Latent means hidden; patent is revealed. When you get a patent on something it means that you tell everybody how to do it but the government protects your right to do it for a certain period of time. Latent means it's hidden.



So there's a hidden valence, these loops, but they can become available, right? It must be assumed in some cases, as for example the combination of ammonia with hydrochloric acid. Now, so you can get a reaction and get the product, which has all five valences of nitrogen now. But how do you know there's not another loop on nitrogen, that you could break open and make it seven valent, or nine, or eleven? Or maybe you could break a loop into three. So might latent valence loop explain the trivalence and pentavalence of nitrogen, or the amine-HCl reactivity? The trouble is it's too slippery a concept. It explains everything. You could explain anything you wanted to. You could explain eighty-four valence of nitrogen or something. Anything that comes along, you'd say, "Ah ha, there are that many latent valences." But how do you know there are not more? Why do you have latent valences? When do you have them? When and why do you have partial dissociation? This thing didn't explain anything, or explained everything. Now, at this point I want you to smile so I can learn who you are.



[Professor McBride takes pictures of the class members]



Professor Michael McBride: So the electron was discovered in 1897. So maybe that has something to do with it. And the guy who had associated electrons with valence was G. N. Lewis. And this is a picture of him as a Harvard undergraduate in 1894. And this is eight years later, when he was an instructor at Harvard, and he then went to establish the Chemistry Department at the University of California at Berkley.



Student: Yay!



[Laughter]



Professor Michael McBride: But when he was at Harvard in 1902 he used these lecture notes, and what he was trying to explain was the periodic table, why you go across in eight and then another eight and another eight, and so on, and why you have electron transfers from some atoms to other atoms. And he said if the electrons in an atom are arranged at the corners of a cube, then you--eight is this very special number, because that occupies all the corners of the cube. So if there's a desire, for some reason, to complete octets, then you can explain periodicity and see why some atoms give up an electron to lose the outer shell and others gain one to complete the outer shell.



Okay, so but it also could explain how you get bonding, if you like to have octets; not just electron transfer but the formation of covalent bonds. So, for example, here's two chlorines. Each has only seven; the octet isn't complete. But if you bring them together and they share an edge, then both octets are complete. He doesn't say why octets should be so great, but if they're octets then you could explain bonds. How about double bonds? Suppose you had two oxygen atoms. What do you do now? Anybody see? What?



Student: Put the faces together.



Professor Michael McBride: Put faces together, share two edges, and then both have an octet. Now, suppose you want to put nitrogens together and form a triple bond. Now how do you do it?



Student: Squish it.



Professor Michael McBride: Push it hard, force it like a picture puzzle?



Student: Yes.



Professor Michael McBride: Put it in there.



Student: You need to push it.



Student: You did it.



Professor Michael McBride: It won't do it. But about ten years later he figured out how to do it. What you do is take the electrons--nitrogen has 5--but instead of using an octet, you contract it along opposing edges this way. So what was an octet, what was a cube, becomes what?



Student: A tetrahedron.



Professor Michael McBride
: A tetrahedron. And now if you take two tetrahedra, you can put a face together and complete both octets. But this wasn't such a great development. He did this, published this, in 1916. But the tetrahedral distribution of the bonds in carbon had been known by organic chemists for 40 years by that time, that the tetrahedron was a very important structure. So a good theory should be realistic and it should be simple. But there's a tension between these things. It should at least be as simple as possible, but that may not be very simple; that's why it takes us five weeks. In regard to the facts it should allow a number of properties. It should allow prediction. You should be able to say how will these atoms react; how will this molecule react; what valence should nitrogen have; how can it be both pentavalent and trivalent? A little below this on the scale of desirability is suggestion. A theory may--even if it doesn't give you the proper thing, may at least suggest some experiment that would be really good to do.



Below that is explanation--according to my theory all these known facts fit together, but who knows about any new thing. And lowest is classification and remembering; like Roy G. Biv allows you to remember the colors of the spectrum, but there's nothing fundamental about that. So even if you have a crummy thing it might allow you to at least remember some facts and organize them maybe. Okay, but those last two are not prediction, they're post-diction; you only use it to explain things that are already known as a way of remembering them. So from the number of valence electrons we'd like to predict how many atoms of different types come together to form molecules. That's constitution--the nature and the sequence of bonds--and the valence.



We'd like to know the structure of molecules, the distance between atoms, angles. We'd like to know how charge is distributed within molecules. Are some regions positive and others negative? We'd like to know what the energy of the molecule is. Is this arrangement of the atoms better or worse than an alternative arrangement? And we'd like to know about reactivity. So Lewis explains constitution, the nature and the sequence of bonds. "Nature" means like single, double, triple, and "sequence" means which atoms are connected to which. So he has electrons, valence, and also he added unshared pairs. So you count up how many, from the atomic weight or the atomic number or the position in the periodic table. You get how many valence electrons, how many of these outer octets are there. And then you can figure that the valence of hydrogen should be one, boron three, carbon four, but then down again, three, two, one.



So here's ammonia, NH3. It works fine and it has an unshared pair. But why do you have octets? Why not sextets? And why do you have a pair for hydrogen? Why not an octet for hydrogen also? Okay, so let's just apply it to HCN. So if carbon has four, that makes possible a single bond to hydrogen, a triple bond to carbon. Everybody has their octet except hydrogen that only wants two. So everybody's happy. And we can abbreviate it this way, and leave the dot. Lewis actually thought of drawing a colon, a pair of dots, to denote an unshared pair. So that's his notation. So NH3. Bingo! How about BH3? Just fine. But here's something new. The organic chemists would already have known this, they knew the valence numbers. The electrons just added something to it but didn't say anything really new. But here's something new, that BH3 reacts with NH3, to give a new bond. Now here's a puzzle. It's also true that BH3 reacts with BH3. That doesn't look like Lewis. And the bond is reasonably strong, it's half as strong as a carbon-carbon bond. So what's the glue that holds B2H6 together? And we'll come back to that fourteen lectures from now.



But there's another thing. There's bookkeeping that you do with Lewis structures to assign what's called formal charges; not real charges, just what charges you write in the formula, and you hope they mean something. And the bookkeeping scheme you use is that each atom is assigned half interest in the bonding pair. So if there's a bonding pair, for bookkeeping purposes you assign one electron to each. It may be they aren't evenly shared, but the bookkeepers don't care.



Incidentally, speaking of bookkeepers not caring, there's--people have had a lot of nonsense trying to get signed up for lab, because the people who run the system don't allow teachers to have access to the thing to see the problems you face. So I have no way of telling--I can't get on to your system. Has everybody figured out how to sign up for lab by now?



Student: Yes.



Professor Michael McBride
: If anybody still has problems, after class Dr. DiMeglio can take you to the computer center here and help you. So next year presumably it'll be easier. This is a new system this year. But anyhow, back to the bookkeepers here.



So these formal charges are not necessarily real charges, they're just what you write when you draw the formulas. Okay, so we split the BN electrons between the B and the N, and we split the hydrogen bonds as well. So nitrogen now has those four electrons, where it brought five to the table originally. So it has a positive charge. Tetravalent nitrogen is positive in the Lewis structure. Why? Because in forming a fourth bond, it gave up half-interest in one of its electron [pairs], the ones in the unshared pair. So it lost a charge, a negative charge, and becomes positive. By the same token, what else can you write in this structure?



[Students speak over one another]



Student: It's negative.



Professor Michael McBride: A negative charge on boron, because it got half-interest in a pair. But that's just bookkeeping. So when you write a Lewis structure you'd write plus on nitrogen and minus on boron. Now is it real? Is there a positive charge on nitrogen and negative charge on boron? Well we'll get later on to know about what a surface potential is. Surface potential is the energy a proton would have if it was on different points on the molecular surface. For this purpose you have to define what the surface is of BNH6. There's a way of doing that; we won't talk about it now, that's jumping the gun. But that's the surface, and it's color coded to show the energy a proton would have if it were on the surface at that point. So what you see is that the energy of the proton would be high, if it's here, where it's very blue. The energy of the proton would be low if it's in this region here. Is that consistent with what we were talking about?



Student: Yes.



Professor Michael McBride: Where would a proton be low in energy? Where there's negative charge, and it would be high in energy when there is positive charge. So this molecule, according to quantum mechanics, is positive on the left and negative on the right. And that's exactly what we have, it's consistent with what we have. And now Lewis also explains where pentavalent nitrogen comes from, because there's an ionic bond there too. Tetravalent nitrogen is plus. So there's just a Coulombic attraction between ammonium and chloride. So that's a great contribution. Now there are other things you can understand in terms of Lewis structures. Like you have an amine and it can react with an oxygen to give an amine oxide, which is positive on nitrogen, negative on oxygen. Or you can have a sulfide, which can react with an oxygen, to give a sulfoxide. Or what else can happen, with a sulfur, what additionally can happen?



Student: Another oxygen.



Professor Michael McBride
: It could do another oxygen, and give a sulfone, as it's called. So Lewis explains this. So there are problems to drill on Lewis structures. They're on the Web, go look at them. One of them is draw Lewis dot structures for HNC, in that order, H then N--we did HCN; but do HNC. And then try HCNO with CNO in all six linear orders; that is, nitrogen in the middle, carbon in the middle, oxygen in the middle, and then hydrogen on either end. And it's also possible to make a ring of CNO and put the hydrogen on any of the positions. So just practice drawing Lewis structures of those.



And start memorizing functional groups. We do precious little memorizing in this course, compared to most organic chemistry courses. We try to figure out how things work so that you can predict without memorizing it. But there are some things you have to know, like the names of groups. Otherwise it takes forever to talk about them. So here are the ones you have to learn, and on the fist exam, coming up in three weeks or whatever, I'll give you a question where you have to draw one of these structures from the name, or give the name for a structure.



These are all on the webpage; you don't have to copy them down. But those are simple functional groups. [Laughter] And here are--whoops, oh there you go, sorry. [Laughter] There. They're also conjugated functional groups. So you could run it yourself. Okay, now the final question here, for today, is a thing that comes up when you talk about Lewis structures, or structures in general, Lewis structures in particular, is equilibrium and resonance. How many people know what resonance is? What is it?



Student: Like the same molecule but different structure.



Professor Michael McBride: Speak up a little bit.



Student: It's the same molecule but a different structure.



Professor Michael McBride: The same molecule but a different structure. But isn't structure a molecule?



Student: But with that I guess--



Professor Michael McBride: You know like, let me just interject a minute here. Like, one of the problems is HCNO in any order. So there would be different molecules with C in the middle, N in the middle, O in the middle. Is that what you mean?



Student: No. I mean like it's--if you have a double or a triple bond, it's in a different location, it's around the central atom.



Professor Michael McBride
: Ah, now let's--actually resonance is one of those things where people try to hide ignorance from ignominy. And let me show you what I mean by that. So here's the structure for HCNO. And here it has all octets. If you go through you'll figure that out. It has charge separation though. You don't like to have that in your structures. You like not to have charges if you can avoid it. Obviously in BNH6 you can't avoid it. Okay, now but you can get rid of the charge separation by shifting that electron pair that's on oxygen to be shared between oxygen and nitrogen; then you don't have the charge on oxygen anymore. Everybody see that?



Student: Yeah sure.



Professor Michael McBride: What's the problem when you do that?



Nitrogen now has too many electrons. So you could shift a pair from--that's in the CN bond, onto the carbon, in your drawing you make. Okay, so you get this structure shown below which has all octets. But there's still charge separation, and in fact it's worse because the negative charge, instead of being on oxygen, is now on carbon. So there are these two different Lewis structures you can draw and each has its own properties, or good things and bad things.



Now what are the geometric implications of there being two structures? It could mean that in the middle one, in the lower one, nitrogen is exactly halfway between carbon and oxygen, because they are both double bonds. But in the top one, nitrogen is much closer to carbon than to oxygen. That could be the meaning of this. So that if you draw a picture of the energy, as a function of the position of the nitrogen, you could get two different structures, one where it's further to the left, where the nitrogen is further to the left you have this structure, and when the nitrogen's in the middle you have that structure. And you have to go up hill in energy, it'll click; as you start pushing the nitrogen it starts, it'll click, and it's the other one. Okay? That could be the way the atoms really behave. That's called a double minimum; we already saw a double minimum today.



But maybe in truth it's not this way. Maybe it's a single minimum. Maybe the best position for the nitrogen is neither to the left nor in the middle, but in between; maybe that's the lowest energy. And that is what resonance is. Resonance is when you--the true structure is in between the structures you draw. So really what it is, is a failure of the notation. Does everybody see how that's so? It's that the notation you use doesn't show the right structure, and when that happens there's said to be "resonance"; although actually all it is is that the true structure is a single minimum, not the double minimum you'd expect from your drawing. It's a failure of the way you draw.



And notice that for resonance you use a double headed arrow whereas if I back up a second here--equilibrium, when you have a double minimum, there are two arrows, two different structures. So the question is which is it? Is the real molecule a double minimum; or the real molecules a double minimum, are there two structures? Or is there in truth only one structure? So you need to choose. The choice between whether there's resonance or equilibrium must be based on experimental facts or on a theory that's better than Lewis's theory, to be able to know which is true; something that can distinguish between a single and a double minimum.



So equilibrium is two arrows, and there are two structures that go back and forth, and resonance is a single structure, that our notation isn't very good at drawing. So two real species, or one real species but two reasonable formulas we could draw for it, or that Lewis would draw for it. So really what it is, resonance is a failure of simplistic notation, and it's associated with unusual stability. Now when I say 'unusual,' that implies that you know what usual stability is. So compared to what is it unusual? And we'll address that question later on.



So let me just give you finally an example of equilibrium versus resonance. So this, as you'll learn when you learn functional groups, is a carboxylic acid, and you could have the hydrogen attached to this oxygen or it could be attached to the other oxygen. So these are not just the same thing upside down. It's the hydrogen being attached to one or being attached to the other of the two oxygens. And, in fact, it is that way. There are two structures and you go back and forth between them; two species.



But how about if you take the hydrogen atom away and have this thing that's called a free radical, where you don't have a complete octet on this oxygen? There are only two, four, six, seven electrons there. It has the ability to form another bond by sharing an electron. Is that two different species? Or is it one species--watch--with an intermediate geometry where you don't really have double and single bonds but something in between? And how do you know which it is? The only way is to do some experiment that tells you. And there is an experiment that tells you it's only one nuclear geometry. And we don't have time in the last thirty seconds to tell you how that proof works, but I'll tell you that the technique is electron paramagnetic resonance spectroscopy [Laughter], shows that in truth, although this is two species, this is one species. So there's equilibrium at the top and resonance at the bottom. And if you add an extra electron to make a carboxylate anion, infrared spectroscopy tells you that it's symmetrical too. There's resonance in this case. It's an intermediate structure. But that is "lore." You didn't predict that from Lewis structure ahead of time. It could've been either way, and in fact there were many people who thought it was one way or the other. It's only experiment that does it. So, next time we'll talk more about lore.



[end of transcript]

Lecture 3
Double Minima, Earnshaw's Theorem, and Plum-Puddings
Play Video
Double Minima, Earnshaw's Theorem, and Plum-Puddings


Continuing the discussion of Lewis structures and chemical forces from the previous lecture, Professor McBride introduces the double-well potential of the ozone molecule and its structural equilibrium. The inability for inverse-square force laws to account for stable arrangements of charged particles is prescribed by Earnshaw's Theorem, which may be visualized by means of lines of force. J.J. Thomson circumvented Earnshaw's prohibition on structure by postulating a "plum-pudding" atom. When Rutherford showed that the nucleus was a point, Thomson had to conclude that Coulomb's law was invalid at small distances.



Transcript



September 8, 2008




Professor Michael McBride: This is the slide from the end of last time, we didn't quite finish it. Equilibrium versus Resonance. Remember, equilibrium is a case where you have two different species, different structures, and you go back and forth between them, maybe fast, maybe slow. Like the hydrogen could be attached to the top or it could be attached to the bottom oxygen, and those are different, two different, what's called isomers; we'll get to that later on. Or you could imagine the same thing if you take the hydrogen off, you could have a short double bond to oxygen and a long single bond to oxygen, or they could exchange as to which oxygen was long and which was short. So you could have an equilibrium there. But it turns out you don't, that in fact it's not two different species, it's one species. It's a single minimum not a double minimum.



Now how do you know this? How do you know it's just one nuclear geometry with an intermediate bond distance? The only way you know is by experiment or by some really fancy calculation that you have to believe. A lot of people would believe experiments before calculations; some are the other way around. But there's evidence from a technique called electron paramagnetic resonance, or EPR, that shows that indeed this is one species, a single minimum. If you have an extra electron on it and you have a carboxylate anion, then again it's just one species, a single minimum. And there's evidence of that from infrared spectroscopy that we'll talk about next semester. But don't be disappointed that you're not able to predict this. A lot of really smart, experienced people couldn't predict it. This is lore. If you look it up in the Oxford English Dictionary it says that lore is, "That which is learned; learning, scholarship, erudition. Also, in recent use, applied to the body of traditional facts, anecdotes, or beliefs relating to some particular subject." So a lot of things are lore. You just have to learn them. You can't predict them ahead of time, they're way too subtle.



So don't be disappointed, because you haven't had enough time yet to get the lore; you're not supposed to know it yet. It would be nice if Lewis theory was so accurate and straightforward that if you draw two structures, there are two structures and it's a double minimum. But that's not true, and you have to know from something, and you don't have time to know yet. As time goes on maybe you'll figure it out. This is from a good textbook; it might be the one that we'll adopt next semester. It's a quote that says, "Empirical rules" (it's going to give) "Empirical rules for assessing the relative importance of the resonance structures of molecules and ions." That is, if you have two different pictures that you can draw for the thing, does it look halfway in between them, almost all like one, almost all like the other, or some fraction of the way between them. How far is it, one way or the other? These are two different pictures you draw to try to show different aspects of the same thing. But there is one real thing. The molecule doesn't know about resonance, it just knows what it is. The problem is with our notation. Okay, but anyhow here's what they say. So rules that will allow you to use this concept more productively by deciding which ones are better, which look more like the real thing. Okay, so it gives rules and it numbers them.



"(1) Resonance structures involve no change in the positions of nuclei; only electron distribution is involved." That is, when you draw these two structures you don't move the atoms, you just change where you draw a single bond or a double bond or a dotted bond or something like that. And in fact that's not even true because it's not--the electrons know where they want to be, it's the way we draw them that's uncertain. So when you draw two different resonance structures, you're not changing where the electrons are, you're just changing the lines you draw. Is that clear? It's our notation that's at fault.



"(2) Structures in which all first-row atoms have filled octets are generally important; however, resulting formal charges" (we talked last time about how you get formal charges) "and electronegativity differences" (and of course we have to know what electronegativity is, but you've heard about it at least) "can make appropriate nonoctet structures comparably important." So if you have a bad charge distribution, even though you have octets, it still might not be a very good structure.



"(3) The more important structures are those involving a minimum of charge separation" (so you don't want to have formal charge separation) "particularly among atoms of comparable electronegativity. Structures with negative charges assigned to electronegative atoms may also be important." So if you're going to get a charge, put it where it wants to be.



Now, look at this more carefully. Number two, it says 'generally important.' It doesn't say "always important," but "generally." That's not such a great rule because you have to know when there's going to be an exception. Or 'however;' this doesn't sound like the Ten Commandments graven in stone. Or "can make" a thing; not that "it will," but "it can." Or '"particularly" or "may also be important." These are all weasel words because the rules are not rules, and this is all lore again. So people write these rules but really the people who are writing the rules know the answers ahead of time and think "Ah, it generally works that way and mostly we can get away with it." But these are not rules like the rules you want to learn in physics. You want to learn these because it's sort of handy, but don't believe them. And notice at the top that it's "empirical" rules. It's not a fundamental theory, these are just ways of sort of correlating a bunch of the lore that's come in. And what's important is the experiments, not theories like this. And we'll get better theories later.



Okay, so the goal of this Lewis stuff was from the number of valence electrons it would be nice to be able to predict the constitution, that is, the valence numbers for the different atoms, how many atoms of one kind or another get together to make a molecule. Reactivity, we've seen a little bit of that at least, that unshared pairs can get together with vacant orbitals and make a new bond. That wasn't something that was part of the original rules of valence. And maybe something about charge distribution as well. Now let's look at the case of O2 and O3, and apply some of this stuff. So O2, you can have--complete the octets by bringing two oxygen atoms together and forming two bonds between them. So that looks pretty good, and it's a double bond, and we can draw it that way with just lines and forget the unshared pairs if we're not particularly concerned with them at a given time. Okay, now suppose you make O3. You could do the same kind of trick there and make a three-membered ring, or you could make it linear, or bent instead of a symmetrical three-membered ring. So you could have an equilateral triangle like that, or you could have it like this, have O2 and bring in another oxygen on one of the lone pairs. Now that's not an equilateral triangle; it's a triangle, or it could conceivably be a straight line, we're not quite sure what the geometric implications are. Okay, but we could call it an open structure as compared to the ring structure. And we can draw it like that, and the formal charges would be as shown. Why? Because that pair that originally belonged only to the original top oxygen of O2 is now shared with the other oxygen. So one of them loses half-interest in a pair, the other gains half-interest in the pair and it's plus, minus. So the trivalent oxygen is positive, as you would expect.



Now, what is the true structure of the molecule ozone, O3? Well you have to do some experiment to find that out or some high-faluting calculation that you believe; a quantum mechanical calculation. So it could be a ring or it could be an open structure. And notice, in the open structure you should have two resonance structures, because you could draw it that way, or you could draw it that way, and it could be two minima--it could be a double minima--double minimum situation, or it could be distorted one way or the other and click back and forth. Or it could be that the true structure is in between with equal bond distances. So we have to find out something about this.



Now one way of finding it out is to do calculations and draw a graph that shows what the answer is, and that's what we're going to do here. It's based on some fairly recent high-level calculations. But the problem is drawing the graph, because if you want to be able to show the structure, you have to--how many variables do you have to specify to say what the structure of O3 is? How many numbers would you have to have? It's three particles, right? So if you gave this distance and this distance and this distance, that would fix the triangle. Or you could give this distance and this distance and the angle here. That would also specify the triangle. But any way you slice it, you have to have three numbers to tell the structure, and then you have to have another number in your graph to say what the energy is, when it has that structure. So you have to plot four variables. And that's not trivial on a piece of paper, to plot a graph with four variables in it. So three distances plus energy, or two distances and an angle and energy you need to plot. So if you're good at plotting things and have had a lot of experience, great, but if you need a little warm-up exercise you could look at this webpage that you get by clicking up at the top there, and it uses these graphs to give you a little exercise in plotting things that are in many dimensions. But we're not going to do that in class, that's just to do on your own or with a discussion section. And we'll show the specific example of how we're going to do it in the case of O3.



So to specify O3 we need four dimensions. Now here we have a plot that shows one distance, a second distance and the angle. So that'll specify where the three oxygens are. But Rutenberg and his coworkers here, in 1997, when they did the fancy calculation, did it a little differently. They constrained it. They didn't allow everything to vary. They said we're going to require the two distances to be the same. So now instead of three variables, you have only two, and you can plot those two on a two-dimensional paper; you can plot two things. So here we're going to plot the position. If the two distances are equal, then if you know the position of the top right you know the position of the top left, because the original oxygen is at the origin 0,0. Everybody with me on that? So we're going to look only at where the top right oxygen is, and the other one will be someplace symmetrically related to it. So we're going to blow that up, and here's the plot they give then. So now any point on this will specify a precise geometry of the three oxygens. Everybody with me? We choose a point that tells where the top right one is. The bottom one is at 0,0, and the top left is just on the other side, in the corresponding mirror-image position. So any point on that thing will specify the structure. But now we still--but that's used up our two dimensions of the paper. How do we show the energy? Can you think of a three-dimensional graph you've ever used?



Student: Color codes.



Professor Michael McBride: Pardon me?



Student: Color codes.



Professor Michael McBride: You could color code it. You could make red really high and blue really low. That would be one way to do it. There are earlier ways, before printing made it--or computers made it easy to make colors.



Student: Contour lines.



Professor Michael McBride: Did you ever go hiking? Yes what?



Student: Contour lines.



Professor Michael McBride: Contour lines, like a geological map, right? Okay, so we can draw contour lines that show what the energy is for every one of these geometries. And so if that--notice that at that position, that X, it turns out to be what the red structure would be here, which is a ring; equilateral, okay? And this structure is open. And I chose those particular ones because when you do the fancy calculation, those turn out to be the ones that are at the bottom of valleys, low energy. And if you distort away from those, the energy goes up. Now you can go up--any direction you go from one of the bottoms of the valleys you go up in energy, but there's some particularly interesting positions. A particularly interesting position is that one, because that's the pass, that's the lowest you can go, the lowest energy that's required to go from one valley to the other. Everybody, I think, has probably done enough hiking to realize that that's the way you'd like to hike, or that if you spilled a cup of water up at the pass it would run down both ways.



Now in fact we can use that concept to make it even a simpler graph, where we don't need to draw contour lines, because we could take the steepest descent path--if you pour the water out at the pass and follow how it'll trickle down, it'll take the steepest way to go down. It turns out it crosses every contour perpendicularly. So you follow that and it would go down to the bottom, and then if you kept going, like if you dropped a marble or something, and it rolled down to the bottom it would keep going. So there is a particularly interesting path. Not that the true molecule would necessarily have to follow that path, but it's a well-defined path that gets from one valley to the other. So now what you could do is take a knife and slice this thing along that green line and unfold it so it's flat. Does everybody see how that would be?



Like I remember once when I was young we took a family vacation and the AAA sent you a map that showed where you would go, but it also showed--along a particular highway--and it also showed another map that showed how high you were all the time as you went along the roads. Got the idea? So this is exactly that kind of thing, where as you go along that green road there are different altitudes. So we could just--we could draw it this way. Does everyone see how that works? So it's not quite as specific about geometry anymore, the way the previous one was, but it shows how much energy you need as you go along. And a particularly important one--well there are two things that are really important. One is how high one valley is compared to the other, how much more stable one is, and the other is how much higher is the pass, how hard is it to get from one to the other?



Okay, so there now is, in a two-dimensional graph that you can draw on a piece of paper, is something that gives you this information. Okay, but this required that we choose R12 equals R23, in order to make this simplification. The guys who did the calculations said if we don't make R12 equal to R23, then it turns out that all the structures are still higher in energy. So these are the lowest energy structures, although they can't plot them on a piece of paper. So the lowest energy structure is this open form and it does have R12 equal to R23. It's a resonance structure, halfway between; not one double bond and one single bond but a symmetrical one. Okay, so it's a symmetrical single minimum, found by calculation. And there's experimental evidence that supports that too, but it's more complicated.



Now how about the charge distribution? What would the Lewis structures that we've drawn here predict about charge separation? Would there be charge anyplace, do you think, on the basis of this, in the real molecule, which is symmetrical? What do you think? Anybody got a suggestion? Yes?



Student: Well, it would be positive on the central oxygen because even converted, regardless of which of the two structures are on that side, that…



Professor Michael McBride: Both structures have it positive in the middle, so that looks like a good prediction.



Student: And the ends would be negative.



Professor Michael McBride: Right, and then partially negative but equally negative on the two ends. Now is that true? Well here's a picture that, based again on--so the prediction is positive in the middle, negative on the ends. Is it true? Well there's the structure, the minimum energy according to calculation, quantum-mechanical calculation. And we can draw what's called the surface potential of that structure, which again comes from calculation. So you put a proton--you define the molecular surface--and that is a little bit of a problem but we'll talk about that later, where the molecular surface is--but then you put a proton at a point on it; and we talked about this before, for ammonium chloride. You put a proton on and you find--or no, it was the BNH6 we talked about the surface potential. Here's the same thing for O3. And you'll notice the surface potential is high in the middle, a bad place to put a proton because there's positive charge there, and it's low on the two ends, at least some place on the two ends. So more or less the Lewis structure predicted this right. So that's great. We're not confident that it will always work but maybe the lore will build up that way; and it does. Okay, so charge distribution also.



So reactivity we saw was a special attribute of this nice Lewis theory, and charge distribution, at least qualitatively if not in detail, for O3 and also for the BNH6. Now how about specific distances and specific angles? We'll get into that later on as to whether you can do that from something like a Lewis structure. And how about the energy content? Well that's not so good actually, with just Lewis structures, but we'll test it later. Okay, so the Lewis-dot structures. It attempts to provide a physical basis for the valence rules, based on completing octets and sharing electrons for bonds. What it gives that's new is reactivity due to unshared pairs where both of the "hooks" in the bond come from the same atom, you might say. And it's convenient for electron bookkeeping. It certainly tells you what the molecular charge is. The formal atomic charges are qualitatively realistic, as we've just seen, at least in the case of O3. And there's this question of stability and resonance. But resonance actually is not something nature knows anything about. It's just a correction we try to apply to having made drawings based on Lewis theory. It's not something serious.



Now, but this leaves us with some serious questions. Why does Lewis theory work? What's so great about octets; why not have a different number? Or if you have a sestet, instead of an octet, how bad is it? Or how bad are structures that have charge separation? It said in those rules that you don't want to have charge separation. Well how bad is it? Suppose you had a choice, you either had to have a sextet instead of an octet, or you had to have charge separation. Which one would win? How bad is "bad" charge separation? Remember it said that if you put the negative charge on an electronegative atom, that's not so very bad, but how bad?



Now last year in the Wiki there came an interesting comment. Somebody said, "I have a question when drawing these structures. Is it more important to try to fill the octet or to have the lowest formal charge on as many atoms, especially carbon, as possible? And why?" That's a good question, because Lewis doesn't tell you that at all. And there's further the question, "Is this at all true?" Are there electrons--this is a really important question, we're going to spend some time on it--are there electrons in between the nuclei that hold them together, pairs of electrons? It'll take us several lectures to get to answer that question. Are there electron pairs between nuclei and are there unshared ones on some atoms? What's the nature of the force laws? That's what we're after, remember, here altogether.



Now, as to whether Lewis theory is right, there's a very fundamental theorem in physics that was developed in 1839 by Samuel Earnshaw, who was a tutor at Cambridge. And the statement of the theorem is that in systems governed by inverse-square force laws, things like gravity, magnetism, electrostatic interaction, there can be no local minimum or maximum of potential energy. He proved it mathematically. It's not our business to repeat his proof, but that's the statement of it. But we want to understand what that means. One thing that means is that if you have Coulombic interaction, positive/negative, you can't have a minimum energy structure that has a nucleus here and eight electrons at the corner of a cube, because there are inverse-square force laws and that can't be a minimum energy; if it distorted, it would keep going.



Now we can visualize Earnshaw's theorem here in terms of for electrostatics by the analogy that you've all seen of magnetic lines of force; everybody's seen something like this I think, right? And the idea, if it's electrostatic, rather than magnetic, the idea is that lines of force emerge from a positive charge and converge on a negative charge, and then you'd make them continuous by drawing these lines of force; which iron filings of course did in that case. And this was the idea of Michael Faraday, whom we met a little while back, and he thought these lines of force were real physical things. Most people don't think that now, they think they're just graphs that involve inverse-square force laws, but he thought they were real. And the neat thing about them is they not only show the direction of the force that a charged body would feel, because of the other charged body, the one that created the lines of force, they not only show the force, the direction of the force, they also show how strong the force is. And the strength of the force is by the density of lines. The denser the lines, the stronger the force.



Everybody's familiar with this idea? Speak up if things are not familiar because I'm assuming they are, if I say so. Okay? So you've seen that. Now let's just think about that a little bit. Suppose you look at the line density here. Through that little line, there are three lines of force that pass. So let's say that means there are three, the force is three at that--and it's obviously pointing to the right. Okay? Now suppose you checked it out at that distance. Stronger force or weaker force?



Students: Weaker force.



Professor Michael McBride: Weaker. In fact only one or one and a half, or something like that, lines of force are going through that. Now, how does it depend on distance, the number of lines going through this standard linked line, the blue one? Well you see that in flatland--we're just doing this in two dimensions, we'll get to three in just a second--in flatland, in two dimensions, the circumference through which all these lines of force pass, the length of the circumference is proportional to the radius. Right? So if you go out twice as far there's twice the circumference. So the density of lines passing through it is half as big. Everybody got that? Okay, so that means the force, which is proportional to the line density, must be proportional to 1/r. Okay? So then as you move out twice, there'll be half as many lines; move out three times, a third as many lines. Okay? But this is just in two dimensions. Now let's think about, how's it going to be different in three dimensions? Katelyn? Speak up a little bit, my hearing isn't so great. I said I went to my 50th high school reunion, right?



Student: Instead of a line it would be more like a square, some sort of an area.



Professor Michael McBride: Yes, it would be a two-dimensional area that this stuff's going to be. Now if you have a certain number of lines passing through at a certain distance, and you go out twice as far, what's the density of the lines going to be? Can you think? Anybody help? Yes?



Student: It'll be inverse-square.



Professor Michael McBride: Inverse-square, because now we're not talking about the circumference of a circle, we're talking about the area of a sphere as we go out. Right? So if we go to three dimensions here and take your area and move it out, the surface is going to be proportional to r2, as you go out. So the line density is going to be 1/r2. So if you have an inverse-square force law, then you can draw lines of force. The lines of force won't work if you don't have an inverse-square force law, because they have to drop off this way in order to give lines of force.



So in 3D such diagrams work only for inverse-square forces. You can't draw a thing like that for Hooke's law. Right? Now, so here's a whole bunch of charges, positive and negative, and the lines of force between them. And notice something interesting. The lines of force start on positive charges and end on negative charges. There's no other place in space where all the lines go away from it, or go toward it. Everybody see that? It's only at the charges that they all go away or all go toward. Right? And that has a very important meaning, that you can't have someplace off in free space where all the lines of force would converge. Notice if you were to have something that was positively charged, and it's at a minimum of energy, then any place you displace it, it'll get pushed back, if it's at a lowest point in energy. You push it any direction, it'll come back. That is, all the lines of force must converge on that point. Okay, everybody with me on that? But it can't be. The only place that all the lines of force converge is on a negative charge. It can't be any place--if you have inverse-square laws, you can't have someplace in space which is the lowest energy place for the charge to be, except on another charge. The same is true, you can't have a maximum either. That's a visualization of Earnshaw's theorem.



So if you have inverse-square force laws, or any combination of inverse-square force laws, like a combination of gravity and electrostatics and magnetics, you can't get a minimum energy structure, unless everything just falls together or blows apart. Okay? So Earnshaw's Theorem: "In systems governed by inverse-square force laws there can be no local maximum (or minimum) of potential energy in free space." Okay? And that's why I don't just float here, I have to be on the floor. Right? Did you ever see anything truly levitating, something that just sits in space and doesn't move, not touching anything else? That's been levitating for ten or fifteen years. It's not plugged into anything, it's just this thing right here. What do you conclude from seeing that, about the force laws that are involved?



[Students speak over one another]



Professor Michael McBride: There must be some force involved in that that's not an inverse-square force law; otherwise it wouldn't work. You can read about it on the Web, if you want to, it's not the business of this class. There is a non-inverse-square force law involved in that little magnet sitting there. The only stationary points allowed by Earnshaw's theorem are saddle points where it's flat in energy, for all directions, but you go one direction and then you go down, go the other direction and you go up. That's like a potato chip or a saddle. So you can have saddle points, but you can't have absolute minima or maxima of energy. There's the picture of that thing. Let me do the--this reminds me to get back to seeing who's here.



[Professor McBride takes attendance ]



Professor Michael McBride: Now we'll continue, got a few more minutes here. And yet it stands still. So there must be something that's not an inverse-square force law there. Okay, so J.J. Thomson, in 1897, discovered the electron. So the idea is maybe electrons have something to do with bonds; that's what brought Lewis into the game. But Thomson himself came up with what came to be called--and I don't know by who first, I've tried to find out and can't--the "plum-pudding" atom. Has anyone ever heard of that? Yes, okay. And I suspect that when people told you about the plum-pudding atom they sort of snickered. Am I right? This was sort of a naïve idea. It's not that naïve at all. Let me show you why.



Here's the book by J.J. Thomson called The Corpuscular Theory of Matter, and if you look in there you'll find out what he's talking about. He says--he has a model of electron configuration--he says: "Consider the problem of how to arrange 1, 2, 3, up to n corpuscles" (that's what he called electrons, he called them corpuscles); "Consider the problem as to how they would arrange themselves if placed in a sphere filled with positive electricity of uniform density." So that's the idea of the plum-pudding. You have this sphere of uniform positive density--why it should be like that nobody knows--but suppose you have a sphere, and you put the electrons in it, like plums in a plum-pudding, which is like a fruitcake in England. Okay, now notice that he said, "placed in a sphere filled with positive electricity." Why did he do that? Why didn't he just have--like Rutherford ultimately did it--a nucleus with positive charge and electrons around and about? Why did he put the electrons inside the positive charge? Zach?



Student: Was it lowest potential energy?



Professor Michael McBride: I can't hear very well.



Student: Lowest potential energy?



Professor Michael McBride: No. Yes?



Student: Well they were thinking [inaudible] maybe that they were going in.



Professor Michael McBride: No, no, they weren't thinking about things moving, they wanted them just sitting there.



Student: Maybe because on a macroscopic scale opposite charges attract. So maybe he might not…



Professor Michael McBride: Yes, but that's well known, Coulomb's law. Yes Keith--or Kevin, right?



Student: If you have them all, you have a positive sphere and if you have the negative corpuscle inside, then they'll cancel each other out and be a neutral body.



Professor Michael McBride: Yes, but that could be--it wouldn't have to be a big sphere of uniform density, you could have a little particle of positive charge, of the same charge, right? It'd be the same deal.



Student: They didn't know that the inverse existed, so they thought that…



Professor Michael McBride: No, none of this is it. It's what we've just been talking about. Yes?



Student: Isn't it about accounting for the bonding between the electrons and the protons and containing the electrons?



Professor Michael McBride: No, bonding hasn't come up yet. This is something very fundamental. Yes?



Student: He was a big fan of plum-pudding.



[Laughter]



Professor Michael McBride: He probably liked plum-pudding at Christmas.



Student: Did he have experimental evidence?



Professor Michael McBride: I can't hear.



Student: Did he have experimental evidence?



Professor Michael McBride: No. Yes?



Student: Because Earnshaw said that there should be no minimum or maximum…



Professor Michael McBride: Ah ha! Earnshaw said you can't have an energy minimum for separate particles, but if you put the negative ones inside the positive ones, then you could get a stable structure. It's Earnshaw's theorem that required it to be a plum-pudding. Okay? Now why a sphere, why not a doughnut or some other shape, a barbell or something, right? And he said that the positive charge is distributed in a way most amenable to mathematical calculation. He chose a sphere so it would be simple. You've heard about spherical cows and so on like that, that physicists like to calculate. Let's suppose there's a cow, let it be a sphere, right? So that's what he did. That's why it's a sphere.



Okay, we can solve the special case where the corpuscles are confined to a plane, if you do it in two dimensions--it's difficult mathematically in three dimensions--but you can do it in two dimensions. And he gives a picture in this book, like this, which is a solenoid magnet that attracts little needles that are magnetized, and those needles are stuck into corks, which float in the water, right? So they have to, the corks have to stay at the level of the surface of the water. So it's a two-dimensional problem. The needles are parallel, the magnets, so they repel one another. But the big magnet attracts them to the center. Okay? So what he does is toss a certain number of needles and corks in there and see what pattern they form. Okay? So here are some patterns. You can get these on the Web, at Greg Blonder's website here. So if you have just one, it goes to the center; no big deal. If you have two, you get a line; there's no big deal about that. Three make an equilateral triangle. Put in four, they make a square. Put in five, you make a pentagon. No one's surprised so far I suspect. Except that sometimes when you put in five and shake it up, you get a square with one in the middle. Can you see where this might be going?



Let's keep going. Okay, if you put in six, you get one inside a pentagon; seven, one inside a hexagon; eight, one inside a heptagon; nine, two inside; ten, two inside. And sometimes it's like that, and sometimes it's like that. These are experimental, right? And if you put in eleven, it's three inside, and then three inside nine; although sometimes you get two--pardon me, I'm screwing up here--sometimes you get--there are two different patterns--sometimes you get three inside eight, sometimes you get two inside nine, for that number. Okay, and then you can get four inside nine, four inside ten, five inside ten, and then after five, if you put in more, if you put in sixteen, you get one inside five inside ten. What is this reminding you of? Yes?



Student: Of orbitals.



Professor Michael McBride: The shell structure of atoms. Right? As you go down the periodic table you complete a shell. So it's a model of shells, and then you can get more and more and more. So this is what Thomson was thinking about. But that's just two dimensions. Right? Three dimensions is a bigger problem. But he could say, he was able to say something mathematically about eight. "The equilibrium of eight corpuscles at the corners of a cube is unstable." Even if you have the spherical charge--so it's not exactly Earnshaw--still you can't get eight at the corners of a cube. Now Lewis comes along and in 1923, as I've told you before, he writes: "I have ever since regarded the cubic octet as representing essentially the arrangement of electrons in the atom." This is long after Thomson had written this about not being able to do that. So was Lewis ignorant of Earnshaw's theorem? Because by 1923 they know that the nucleus is not a plum-pudding--that is the positive sphere--that it's a point. So was Lewis just naïve?



No, look what he wrote, in 1916. "The electric forces between particles which are very close together do not obey the simple laws of inverse-squares which holds at greater distances." So Coulomb's law breaks down. You don't have inverse-square. So then you don't have Earnshaw's theorem, and you don't have to worry, and you can get a structure if it's not an inverse-square force law. Okay? No trouble. But what is the force law? Well Thomson thought the same thing in 1923 in his book The Electron in Chemistry. He wrote: "If electron nuclear attraction were to vary strictly as the inverse-square, we know by Earnshaw's theorem that no stable configuration is possible with the electrons at rest or oscillating about positions of equilibrium… I shall assume that the law of force between a positive charge and an electron is expressed by this equation… Then a number of electrons can be in equilibrium about a positive charge without necessarily describing orbits around it." And look at the--we're going to end right now by looking at this equation. What's that bit of it, the first bit, the "one" part?



Students: Coulomb's law.



Professor Michael McBride: That's Coulomb's law. But he's got a correction to Coulomb's law, here at the end, c/r, where c is a distance--you divide it by r and you get a number--and c is a distance that's on the scale of atomic lengths. Right? And that means that as long as c is very small--pardon me, as long as--okay so have I got this right? Okay, so when distance r gets smaller than c, then the force changes sign. Okay? So what was attractive becomes repulsive, and then you can have the electrons sitting around the nucleus. Okay? So we'll see what happened three years later next time.



[end of transcript]

Lecture 4
Coping with Smallness and Scanning Probe Microscopy
Play Video
Coping with Smallness and Scanning Probe Microscopy


This lecture asks whether it is possible to confirm the reality of bonds by seeing or feeling them. It first describes the work of "clairvoyant" charlatans from the beginning of the twentieth century, who claimed to "see" details of atomic and molecular structure, in order to discuss proper bases for scientific belief. It then shows that the molecular scale is not inconceivably small, and that Newton and Franklin performed simple experiments that measure such small distances. In the last 25 years various realizations of Scanning Probe Microscopy have enabled chemists to "feel" individual molecules and atoms, but not bonds.



Transcript




September 10, 2008



Professor Michael McBride: Okay, so we saw last time that Lewis wasn't so dumb. He knew about Earnshaw's theorem, but he still thought there was an octet of electrons around the nucleus. How could he think such a thing? Because if you have inverse-square force laws you can't have static positions of charged things, unless they're right on top of one another or blown apart. So what did he think?



Student: It's not an inverse-square.



Professor Michael McBride: That it's not an inverse-square force law, that Coulomb's law breaks down when you get to a very small scale. And J.J. Thomson thought the same thing, and wrote here these two terms here. The first one is Coulomb's law indeed, but there's this other thing, c/r. So as long as r is very big you don't compare -- you don't care about that term, as long as it's very big compared to c. And c is a distance that's something like the size of atoms. So once r -- once the distance of the things that are interacting with one another, the electron and the nucleus say, once they get near the distance c, then this c/r thing becomes significant, if r becomes very small, and it even changes the sign. So what was attractive becomes repulsive. So that would do okay then for a structure.



That was in 1923. In 1926, just three years later, quantum mechanics came along and showed that Coulomb's law was just fine, nothing wrong with Coulomb's law. It goes to much, much, much, 1020th, smaller distances than the size of atoms. Coulomb's law still works. What was wrong was the way they treated kinetic energy, because kinetic energy quantum mechanics reformulates. So it gives you electron clouds. And so what's a cloud is not the positive charge -- remember, J.J. Thomson had a positive sphere of electricity in which he embedded the electrons -- what's a cloud is the electrons, and nuclei are put in it, to make molecules. So it really is -- he was right about plum-puddings, he just had the charges backwards. Not many people give him credit for that. So cubic octets of Lewis and ad hoc electrostatic force laws, like this c/r term in there, soon disappeared from conventional chemistry and physics; immediately in fact, I mean within a month say. But the idea of shared-pairs, and lone-pairs, that Lewis came up with, remained useful tools for discussing structure and bonding. So we still do it, and that's why you did problems today that had to do with that. Yes Niko?



Student: Yes. Can you explain the kinetic energy theorem again?



Professor Michael McBride: Yes, but I won't do it for a week still, because it takes a little while. Okay, so Earnshaw said there's no structure of minimum energy for point charges -- that is, no classical one -- but if you come to quantum mechanics and fiddle around with kinetic energy, then you can get it. But despite Earnshaw, might Lewis have been right? Might there still be shared-pair bonds and lone-pairs? What are bonds? And that's what we're going to try to find out. And how do you know, or how do you know? How are we going to know whether there are electron pairs that are shared between atoms and lone-pairs on atoms? How can we find out?



Student: Experiment.



Professor Michael McBride: Right. You've got to do an experiment. So what kind of experiment? Now these experiments are tough. Why are the experiments tough?



Student: It's so small.



Student: It's really small.



Professor Michael McBride: Because things are -- well we could see, or we could feel whether these electron pairs are there. But there's a problem, as you say, that it's inconceivably small, the thing we're looking for, to see or to feel. Okay? But is it really inconceivable how small they are? When I was in your place everybody thought yes, they're inconceivably small, we'll forget it. Right? But there are more recent techniques that suggest that you can really see things that small; at least you can think about them clearly. Now, the first idea of seeing them came from this book here. Actually this book was published -- this is a Third Edition of the book. The First Edition was 1909 and the Second was 1919, and this is the Third Edition, which is still in print. You can go on the web and buy this baby, Occult Chemistry, A Series of Clairvoyant Observations on the Chemical Elements, by Annie Besant and Charles Leadbeater. So these are the occult chemists who started practicing their trade in 1895. "Bishop" Charles Webster Leadbeater; he was a bishop of a church he founded that had -- he was the head bishop; there were like 100 bishops and 500 members of the religion. This is Annie Besant, who is one of the most interesting people in all of history. They closed the stock exchange in India the day she died, for example. This is in her -- the outfit she designed to lead the Boy Scouts in India. And she also proposed marriage to George Bernard Shaw. There are all sorts of interesting things about Annie Besant. But one of the things we're interested in is that she could see atoms. And then their helper, who became leader of the operation later on, was Jinarajadasa.



Okay, so this is a page from this book, from the paper that became the book in 1895, which shows what they could see for hydrogen, oxygen and nitrogen. Now there are a number of levels here. The very lowest phase is solid, H O N, then liquid, then gas. So this is what, when they went into sort of a trance, they could see. This is the gas. But then there's Ethereal 4, Ethereal 3, Ethereal 2 and Ethereal 1 phases as well, with higher and higher resolution. You can see these. So these, for example, these little dots that you see here are what get blown up here, and the little things inside this are what get blown up here, and the things inside this get blown up here, and the things inside this are this; and that's the same for all atoms, that's the fundamental unit. Okay, so here's hydrogen at the gas resolution level. So there are wheels in wheels in wheels here. And here's what oxygen looks like; it's these little helices. And now that's the thing you finally get to, the thing called anu. And that they actually ripped off from this book, The Principles of Color, which was published in America in 1878 by Edwin D. Babbitt. And here's his picture of this thing. And lo and behold, that's what they saw as the fundamental constituent of matter. It's fun to look at in detail. You can look at it at the Web, if you want to. It also has pictures of the Angel of Innocence here, and the psycho-magnetic curves surrounding your head.



Okay, now lo and behold the real confirmation of this thing was that if you counted up the anu in the different elements, there were eighteen anu in hydrogen, 290 in oxygen and 261 in nitrogen, and if you divide that by eighteen, you get the atomic weights, Q.E.D. Okay, so they have pictures in this book of a bunch of the atoms. This is helium, which has seventy-two anu. This is lithium with 127 anu. Not the bottom part; this is a model and this is the wooden base it sits on. Here's iron, 1008 anu. Here's neon, 360 anu. Now look at neon, and that is a 4f orbital electron density, calculated by quantum mechanics. Now the question is, why would you believe me, or believe a textbook, or believe a quantum mechanician, and not believe them? Because they said they saw it too. Okay? Well it takes time. We've talked about this before, the basis for scientific belief: evidence, you always cross-examine an assertion; logic; and taste, but taste matures with experience and at the beginning you don't know which people to believe and which not to believe. It's a problem in politics too. Okay, there's sodium, 418. They not only said they saw atoms, they also saw molecules. For example, here is sodium carbonate. What do you see in there? Two sodiums, right? Okay? In the back are pieces of a carbon atom. What else do you see?



Students: Oxygen.



Professor Michael McBride: Ah, there are three oxygen helices wrapped around the sodium and the carbon. And they wrote: "Note that this triangular arrangement of O3 has just been deduced by Bragg from his X-ray analysis of Calcite." We'll talk about Bragg and what he was doing next time. But they have experimental evidence that supports what they're reporting, independent experimental evidence. Here's their model of benzene, and they say about that, "Each of the three valencies of each Carbon are satisfied by Hydrogen, and the fourth valency, which some have postulated as going to the interior of the molecule, does actually do so." In fact, Annie Besant was the first woman to get a degree in Chemistry at the University of London; but that's another story. Her teacher was the guy who translated Das Kapital into English, and had a lot of other nefarious things that he did. Question six will have to do with that. Okay, so here's benzene and its resonance. These were structures that were first early proposed; we'll talk about that later on. But some people didn't like this idea of going back and forth between two things, and they thought that you just have the fourth valence of every carbon going into the middle, all satisfying one another. So this was what they said that their model confirmed, that the fourth valence of every carbon goes into the middle.



Okay, needless to say most people nowadays don't think too much of occult chemistry, even if you can still buy the book online. This portal you came through this morning. It says "Erected 1921". But here's the same place, in 1923, when the building was dedicated, and you notice those plaques weren't carved yet. It didn't say erected 1921. They were still finishing the building at the time the American Chemical Society had their national meeting here to celebrate the biggest academic building for chemistry in the world. Okay? So the slates were clean. Now if you go down into the crowd here, you see some interesting people. For example, there's G.N. Lewis with his Philippines' cigar in his hand. He always had a cigar; he smoked seventeen cigars a day. [Laughs] Groan.



Okay, but they were trying to decide how to decorate all these plaques around the building, and they wanted to put the names of distinguished American chemists. Had they done that, it would be pretty much a disaster because the people who were thought to be distinguished American chemists in those days were not so hot actually in retrospect. But they chose a wonderful group, and the idea came from Giuseppe Bruni from Bologna who had come to the meeting to speak. And he said, "Don't put living chemists there, put dead chemists." And that's what they did; in almost all cases they were dead. And it's a wonderful group; as we've already used Faraday over here. Okay, here's one of them that's over in that corner of the building. And van't Hoff, have you heard of van't Hoff? You'll have heard a lot of him by the end of this semester. And Gibbs?



Students: Yes.



Professor Michael McBride: Yes Gibbs, we'll talk about him too. And Mendeleev you've heard of. But I suspect that Crookes is not so familiar to you. This is Sir William Crookes. He was a Fellow of the Royal Society, and FTS. In 1861 he discovered the element thallium. He also developed the cathode ray tube, which he's holding there, which became the x-ray tube, and he invented the Cooke's radiometer. I bet every one of you has seen a Cooke's radiometer. Do you know? There's a picture of a Cooke's radiometer, to measure the intensity of light by how fast it spins. You've seen these things you buy in novelty stores that sit there and spin in the light. Okay? From 1913 to 1916 he was President of the Royal Society, the same thing that was founded back by Boyle -- remember? -- and Robert Hooke and those guys. In 1898 he was President of the British Association for the Advancement of Science, and he was also, in that year, President of the Society for Psychical Research. And FTS is Fellow of the Theosophical Society.



In his presidential lecture to the British Association for the Advancement of Science, one of the things he said was, "Telepathic research does not yet enlist the interest of the majority of my scientific brethren." But he's the guy who supplied the samples for Leadbeater and Besant to look at. He supplied them with lithium, chromium, selenium, titanium, vanadium, boron and beryllium samples so that they could draw these pictures of what they saw by clairvoyance. So you can read more on the Web if you want to about him, he's an interesting character. And there were a lot of semi-, well really serious scientists. Oliver Lodge was another one who invented radio and was head of the Physics Department at Bristol, I think [actually Liverpool and Birmingham], who was into communicating with the dead and stuff like that. So it wasn't obvious at the beginning who you should believe and who you shouldn't believe.



But to get back to the point, and forget clairvoyance for awhile, which I don't believe in, the question is, are molecules unobservably small for what Newton called "vulgar eyes"? Is there no way we can see something and measure it if it's that small? So consider water. A cubic centimeter of water is 1/18th of a mole, since the molecular weight is eighteen. That means it's 6/18 times 1023 molecules, in a cc of water. Now that's a really, really big number and it sounds like they must be just impossibly small, hopeless to try to see such a thing. But the difference is between the cube and the cube root, because when we measure distance we measure linear, not volume. So to get something about distances you take the cube root of the volume. So if we take the cube root of that, it's about three times 107. Now that's still a very big number, 30,000,000. Right? But it's not impossibly small. It's about -- 3 angstroms is this dimension, the size of a water molecule. Remember, we'll talk a lot about a carbon bond -- carbon-carbon bond, being about one and a half angstroms.



Okay, now think about 105. So this lecture room is about ten meters wide. One of my hairs is order of magnitude 100 microns in diameter, and that's a ratio of 105. Right? So 105 of my hairs sideways would go across this room. So that's a lot. Right? But it's not impossible to think of lining up hairs to go across the room, or putting them side by side. It'd take a while, but it's not impossible to think about that, a hair compared to the room. Right? But a molecule is about a nanometer, and a small atom is about 1/10 of a nanometer, or an angstrom. Right? And that ratio is 105. So the room to my hair is like my hair to a molecule, or even an atom, a big atom. So it's not impossibly small, it's just very, very small. In fact, a nucleus is 10 femtometers, which is another 105 down. So that means if the lecture room were an atom, the nucleus would be the diameter of a hair; very small, but not impossibly small to conceive of. Okay?



Now Newton, in Opticks, the book we looked at before, in 1717 wrote on page 369: "The thickness of the Plate where it appears very black, is three eights of the ten hundred thousandth part of an Inch." "Is" - that means he knew the size of it. Now how big is three-eighths of a ten-hundred-thousandth part of an inch? Well ten-hundred-thousand is a million; three-eights of a millionth of an inch is thirty water molecules. So somehow Newton was able to measure something, in 1717, that was the size of thirty water molecules. Now what tool could Newton possibly have had in 1717 that allowed him to measure something so very small? What?



Student: Glass ring.



Professor Michael McBride: Can't hear very well.



Student: Glass ring.



Professor Michael McBride: Glass springs?



Student: Ring, like you would --



Professor Michael McBride: That's right. Okay, so here it is, Newton's Rings, which really should be called Hooke's Rings, because he's the guy that in Micrographia published how they worked. But remember, they had different light theories. So here are two pieces of glass, two disks of glass. And I'll change the lights so we see light through them. Now I can't remember which is which, but let me guess that if I put this one over here, having turned it over -- okay. Oh I can see, yes you can see. See those patterns? You've seen things like that with microscope slides, colors. Okay? So these are two flats against one another, they get very close to one another. But if I turn the top one over, you see that it's a little bit -- let me -- so that's what we just saw. But notice that if I turn it over there's a little -- there turns out here at the edge to be a little bit of a gap, because the top one is just very slightly curved. And now let me see if I can see something. I can't see it yet, let me zoom in. We're going to need to focus on that. Ah, see that thing, see that.



[Technical adjustment]



There. Okay, so here's what you're seeing. In the middle is what's called Newton's Rings. They look like that. Different colors, right? And Newton could measure the thickness of the air gap that caused different colors. The colors repeat in higher orders. Okay? Now how could he know the distance? So he could associate every color with a distance. How could he know the distance? Because he could measure the diameter of the rings. And he said this: "Observation Six: The Diameter of the fixth Ring" (Right? One, two, three, four, five, six) "at the most lucid part of its Orbit was 58/100ths of an inch, and the Diameter of the Sphere on which the double convex Object-glass was ground was about 102 Feet; hence I gathered the thickness of the Air or Aereal Interval of the Glasses in that Ring." So in other words, here's the air gap he's trying to measure, which is the sixth ring out. He knows the diameter of that ring. He knows that they touch in the middle, the two glasses, and he knows that it's a sphere and it has a radius of fifty-one feet. So he can just use trigonometry to figure out that the air gap, if he knows that angle there, the air gap is 1.8 microns, right? So that's a lot bigger than an atom; but if you go into the first ring, or even closer, then you can measure it. Right? So all he needed was a spherical glass with a very large radius that he could measure his distances from. Okay, but here's a simpler measure of an even smaller distance, about 100 years later, and here's the guy who did it. You know who that is?



Students speak over one another]



Professor Michael McBride: Benjamin Franklin of course; we'll get back to the artist later on. So he published in the Philosophical Transactions of the Royal Society in 1774 -- the Society is now 110-years-old, 114-years-old -- this article on the stilling of waves by means of oil. Can you see what that means? What's that related to, a common saying?



Student: Separation?



Professor Michael McBride: Pouring oil on troubled waters; have you heard of that one? So the stilling of waves by means of oil. Extracted from sundry letters between Franklin and Brownrigg. So he wrote to Brownrigg, 1773: "I had, when a youth, read and smiled at Pliny's account of a practice among the seamen of his time, to still the waves in a storm by pouring oil into the sea; as well as the use made of oil by the divers…" Pearl divers would fill their mouth with oil and when they went down, if a breeze came up and ruffled the surface, which made it hard to see because the light wouldn't come through clearly, then they'd let oil out of their mouths that would float to the surface, stop the ripples, and they could see to get their pearls. This is what Pliny said. Right? But he said he smiled at that, what a quaint thing to think; it's like the occult chemists.



"I think that it has been of late too much the mode to slight the learning of the ancients. The learned, too, are apt to slight too much the knowledge of the vulgar. In 1757, being at sea in a fleet of ninety-six sail bound against Louisbourg," (off Cape Breton Island) "I observed the wakes of two of the ships to be remarkably smooth, while all the others were ruffled by the wind, which blew fresh. Being puzzled with the differing appearance, I at last pointed it out to our captain and asked him the meaning of it. 'The cooks,' said he, 'have I suppose just been emptying their greasy water through the scuppers, which has greased the sides of those ships a little.'" (So there's experimental evidence for what Pliny said.) "Recollecting what I had formerly read in Pliny, I resolved to make some experiments of the effect of oil on water when I should have the opportunity."



And when he had the opportunity was when he was in London. So here's London, and his experiment in 1762 was in that place, on Clapham Common in South London.



"At length being at Clapham, where there is on the common a large pond which I observed one day to be very rough with the wind, I fetched out a cruet of oil and dropped a little of it on the water. I saw it spread itself with surprising swiftness upon the surface; but the effect of smoothing the waves was not produced." (So the experiment was a failure.) "For I had applied it first on the leeward side of the pond where the waves were greatest; and the wind drove my oil back along the shore."



So it didn't spread on the surface. So what did he do?



Students: Go to the other side.



Professor Michael McBride: Go the other side, right? "I then went to the windward side where the waves began to form and there the oil, though not more than a teaspoon full, produced an instant calm over a space several yards square which spread amazingly and extended itself gradually till it reached the lee side, making all that quarter of the pond, perhaps half an acre, as smooth as a looking glass." So a teaspoon is five cubic centimeters, and he stilled half an acre, which is 2000 square meters, which means the thickness of this layer, to spread that many cubic centimeters over 2000 square meters, would be twenty-five angstroms, which is the length of a molecule of the fat that's in the oil. So Franklin had in fact measured the length of an oil molecule, in this way.



Now he didn't think he had measured it and didn't claim to have measured it. He said, "When put on the water it spreads instantly many feet around, becoming so thin as to produce the prismatic colours" (the Newton Ring's colors, right?) "for a considerable space, and beyond them so much thinner" (it became so thin that you didn't even get Newton's colors, right?) "so much thinner as to be invisible except in its effect of smoothing the waves at much greater distance." (So the technique was stilling the water; allowed you to measure how big the area was. Isn't that cool?) "It seems as if a mutual repulsion between the particles took place as soon as it touched the water." That is, they pressed one another apart and moved. So what he wouldn't have known was that they pushed one another apart but they stayed in contact with one another, to become a monomolecular thick layer over the water. So he didn't claim to have measured it, but indeed he did measure the size of an oil molecule. So molecules are very, very, very small, but they're not inconceivably small and there are ways you can measure them.



So we have this question, are there electron pairs? So let's try first feeling, with Scanning Probe Microscopy, or SPM. Can we feel individual molecules? Can we feel individual atoms? Can we see what bonds are by feeling them? So Scanning Tunneling Microscopy was invented in 1981 by these guys in Zurich, Gerd Binnig and Heinrich Rohrer. And here's the first publication. They worked at IBM and here's the first publication on the cover of their journal about a meeting about this new technique, Scanning Tunneling Microscopy. And they got the Nobel Prize within five years, or maybe it was even less; in very short order, 1986. Now how does that Scanning Tunneling Microscopy, which is one type of SPM scanning probe -- Scanning Tunneling is one way, and I'll mention that shortly. But first I'm going to talk about Atomic Force Microscopy. And for this you need a little chip, and I'll show you. The chip is in here, in this little bug box. Ah you can't see it here. Well there's the chip in the middle, right? And I'll pass it around so you can look at it. And if you tilt it so that the light from the ceiling glints off the gold, and you see it really shining, you may be able to see at the thin end these little v's; you can see here at the bottom little v's. The chip is 1.4 millimeters wide. It's a gold-coated silicon chip. And if you look at those with even higher magnification you see this -- and there's the size of a hair. You can actually see -- did you see them? They're not easy to see and maybe you will and maybe you won't, but they're there.



Okay, so there's the size of a hair. And in any given experiment you have a choice of five different cantilevers, as they're called, that you can use. Here's a higher resolution electron micrograph that shows the tip of one of these v's, and you see a little bulge here, at the bottom, and if you zoom in on that you see this; it's a little pyramid. And so at the bottom of that pyramid the radius of curvature of this tip is only about twenty nanometers, so about the size of twenty molecules. Right? It's round on the tip, it's not truly pointed and come to nothing; it's a little bit rounded. You can buy ones that are a third that size, their radius of curvature. Okay, and then you have this piece of glass. I said, "Do not touch" when you came in because that piece of glass costs $2000. Okay, so there's this piece of glass, and if I hold it up here there's a little chip on the bottom of the piece of glass. I know what I'm looking for. I don't know if you can see it or not. But anyhow it's held into this piece of stuff here. And those little tips are pointing up here, but in fact the thing is used upside down so they point down. So here's that thing mounted in an instrument, and you see there's a red light glowing in the glass, and the reason is it's being irradiated by -- so there's a blown up picture of one of these tips, pointing down. And laser light comes and bounces off the shiny gold on this tip, and gets reflected up to a detector. Now what would happen if you pushed up on the tip? It would deviate the reflection, right? It would go like this. And you have a very precise position detector for the laser light up where you see the light glowing. So you can tell the deviation of that tip, moving up and down, by a size that's less than an atom. Isn't that neat?



So now what you do in this is move this back and forth over the sample and watch the little needle go up and down by the reflection of the light. So if you had, if the surface was like this, they'll click, click, click, click, click, click, up, up, up, and just scan across. And it's sensitive, as I say, to less than one molecule change in height; although it doesn't feel an individual molecule because the tip, remember is twenty nanometers wide. So the width of the tip is like 100 or so atoms. Right? So it's touching a bunch of atoms at any one time. But if you come to a ledge, unlike a [smooth] crystal [face], it would be going at a certain height, touching a bunch, and then it would move up and touch a bunch more, and you'd be able to tell if it went up and down, if that distance was only one molecule; easy.



Here's, in fact, a picture taken by an undergraduate. In fact, it's a movie. So it's AFM traces. So what you do, it's like a TV where you go zoom, zoom, zoom, zoom, zoom, zoom, zoom, zoom, and then you go back to the top, zoom, zoom, zoom, zoom, zoom, say how high it is at every point along the scan, and then color code it to show how high. So the width of this is about 600 molecules, right? But the lines you see, which are ledges that go up and down; each of those ledges is one unit cell, 1.7 nanometers, one set of molecules. So these are -- and that, the thing at the top surface where it's flat, is absolutely flat, except for there's little holes there. And those holes were made by taking this tip and pressing down and scratching a little bit to knock some of the molecules away. So the larger pit is five nanometers deep, which means it's four layers of molecules that've been knocked off in there. And now this is underwater, or underwater in propanol. So the crystal is dissolving slowly. So those pits open up, as molecules come away. So watch. Right? Those are at one-minute intervals. So we're actually feeling individual molecular heights with this thing.



There's an interesting point here that I'm not going to dwell on, but the smallest one did not lose any molecules. If I back up, watch the smallest one. It always stays the same size as the others dissolve. That's an interesting puzzle isn't it? But at any rate, you can do it. And you can measure the rate at which different ledges dissolve. Some dissolve much faster than others. Here's scanning tunneling microscopy. This is done in a different way. It's not -- you don't reflect the laser light. What you do is detect the electrical conductivity through a layer. And if you have an atom there that is a good conductor of electricity, you get more current through and less and so on. So as you scan a real -- now this could be a really, really sharp tip where the tip is just one atom or a couple of atoms. So now as you go across you can feel things that are much smaller laterally. Okay? So this picture here, that thing right there, which is as you see a repeated pattern, that's one of those molecules, with 12 carbons and a bromine on the end. And here's a model that shows what the molecule looks like. The yellow is bromine, the brown are the two oxygens, and the green are carbon and the grey are hydrogens. They don't come colored that way, they come looking like this, as to conductivity. But you can see individual atoms. Right? Or one of the more dramatic early examples, in 1993, was at the IBM labs at Almaden in California, where they worked on iron atoms on a surface of copper, and they used this tip and changing the voltage to pick up atoms and move them where they wanted, then change the voltage and drop them there, then go back and get another one and do it. So they made this thing they called a quantum corral of whatever, fifty-eight or something iron atoms. Yes?



Student: Is that along the lines of when they moved the atoms --



Professor Michael McBride: Pardon me?



Student: Is this along the lines of when they moved the atoms to spell out IBM?



Professor Michael McBride: Oh yes, they also spelled out IBM. They knew which side their bread was buttered on too, just like the Royal Society did. Okay, so that's neat. So clearly you can see individual, or feel individual atoms, but you don't see the bonds. The bonds are smaller. It just looks like a bunch of balls, right? You don't see the electron pairs between them.



Here's SNOM. SNOM stands for Scanning Near-Field Optical Microscope. So this is in a sense seeing, but actually it's scanning. It's like feeling. So what you have here is a glass fiber that's drawn down to a very sharp point and coated with aluminum, and the hole at the bottom is just 100 nanometers; that's like 100 molecules, not really, really tiny like the scanning tunneling microscope. So what you do is you have something on the surface of a slide here that you want to probe with this, and you send light down the thing, and if there is a molecule where the light's coming out, through this tiny hole, if there's a molecule there, that will take the blue light and emit red light. Then you know when it's under the tip. Suppose there was just one molecule on the whole slide. You scan it around until suddenly red light comes out. Then you know that tip is pointed toward that particular molecule. Okay? So red light comes out and you have a detector that detects it, and you scan the sample back and forth, like a TV, and you find out where such molecules are. And here's a picture taken in that way. So this is a scale of microns, 2 x 2 microns; in fact it's a distorted scale, as I see here. Okay? And that arrow, doubled-headed arrow, is the wavelength of red light. So you're seeing things significantly smaller than the wavelength of the light, having done it in this way. You're not really looking with your eye, you're doing this scanning trick.



Okay, so scanning probe microscopies, like atomic force microscopy, scanning tunneling microscopy, scanning near-field optical microscopy, are really powerful, and you can see individual atoms. Right? The sharp points can resolve individual molecules, and even atoms, but not bonds. So it's not doing what we need to do for our purposes.



Now, that's -- I'm going to spend the last few minutes going over the problem that we were looking at with -- remember you were supposed to draw all the resonance structures for H, C, O, N isomers. I think you remember that. Okay, so let's just look at it. So what we're going to do is compare what we see with Lewis theory with things that have been calculated by reasonably good quantum mechanics; not the very, very highest level, but a pretty good level. Okay, so we want to try to make all the possible arrangements here. So we can put the three atoms, O, C, N, with any one of them in the middle, and then we can put hydrogen on either end. So then we'll do it with nitrogen in the middle and then with oxygen in the middle, so we'll have done all possibilities there. So here are two possibilities. Now are these good structures is what we have to decide? Okay, now it would be possible to shift electrons here and draw a different Lewis structure. Which of those two do you think is better?



Student: Probably the top one.



Professor Michael McBride: Why?



[Students speak over one another]



Professor Michael McBride: Okay, there's separation of charge on the bottom. Okay? So there's bad charge separation. The most electronegative atom, oxygen, in fact has a positive charge not a negative charge. Right? So that's -- not only is there a charge separation, it's in the wrong direction. But they're all octets that we've drawn in both structures. All atoms have octets. Now let's try the same trick over here. We can do it that way. How about that? Which of those is better? In both structures we could go through and count and see that they're all octets. Would you like me to do that, or is that something that's become second nature to you by now? Anybody want me to count them? Okay, you've got that. But now which of those two structures do you think is better?



[Students speak over one another]



Professor Michael McBride: So you say the top one's better, you say, because it doesn't have charge separation. Now, as compared to the case on the left, which do you think -- if you have to have charge separation, which one is better?



Student: The right.



Professor Michael McBride: The one on the right because at least you have it with the right atom, the oxygen being negative. You could've drawn the one on the top too, right? And that goes the other way. Notice that I've introduced something here that I haven't really talked about enough, the idea of curved arrows. Curved arrows show a shift of electrons, an electron pair. What would you draw if you wanted to show just one electron shifting, if you had to think up the notation?



Students: A dotted curve.



Professor Michael McBride: A dotted curve would be one possibility. Another possibility? Sophie?



Student: Half a head instead of a full head.



Professor Michael McBride: Ah, like a fishhook instead of the double arrow. And that's in fact what's used. If you want to shift just one electron you show a single barb, and if you do a pair of electrons you show a double barb, a curved arrow. So curved arrows don't show atoms moving. They don't show this atom goes to there. They show an electron pair moving, or sometimes a single electron moving, and you can use a double or a single barb to show which one. Okay, so get that straight, because that's often a problem. Okay, so the one on top is much worse than the one on the bottom. Okay, so there are our possibilities with C in the middle.



Now how about if you put N in the middle? Here are two structures with octets. Now -- oh pardon me, they don't all have octets, that one on the right has a sestet, because the carbon's making two bonds; so it gets three electrons out of those two bonds -- pardon me, it gets two electrons out of those two bonds, one from each bond, that's its share from the bookkeeping, plus an unshared pair. Right? That's only -- what's associated with it, it has only three… pardon me, I'm saying the wrong thing here. It has four electrons that it calls its own, from a bookkeeping point of view, one from each bond and two in the unshared pair. So it's not charged. It came in with four electrons, it's still got four electrons, but participates with only three pairs of electrons; so it's a sestet. So that's not so good on the right. The one on the left is not so good because it's charge-separated. At least it's on oxygen, the negative charge. We could draw this other one, which puts the negative charge on carbon. That's, we would say, not so good. Right? So that's a bad charge. Or we could do it that way. That looks like the best charge probably; better to put positive on carbon than on nitrogen. Okay, but that carbon has a sestet; the others are octet ones. Okay, we can do this other one over here. Now we've got an octet on the carbon, but we've got negative charge on the carbon. That's not such a great charge separation thing. Or we can do it this way, but again we got bad charge. Okay, so now if we put O in the middle we can do these structures, both of which you notice have -- what's bad about them?



[Students speak over one another]



Professor Michael McBride: Sestet there, sestet on the right too. There's a plus charge on oxygen; that's not so great. You can draw curved arrows and shift the electron pairs around. You've now got an octet on carbon but you got a sestet on nitrogen, and bad charge, negative on carbon; not so great. And over here you could do that. Sestet again, bad charge. Or you can make the three heavy atoms into a ring, and then you could put the hydrogen on any one of the -- coming out of any one of them. Now the one on the left, with my computer, I can't calculate it. I can't find an energy minimum that has that structure. So that one is very bad; which doesn't surprise anyone. It's got horrid angles for the bonds, although we haven't really said that we prefer one angle over another. But it also has bad charge separation. Okay? That one's a bad charge. So we could draw this other one here, but it's still got a positive on oxygen, a sestet. What do we have here? A sestet on carbon. Okay, so here are the bunch of isomers we've just gone through.



Now a couple of years ago, when I first talked about this in the class, and when these things had been accurately calc… or as accurately as people are likely to do, to calculate their true energy and structure -- I tested my colleagues in the department, asked them if they could rank these things as to what's the lowest and what's the highest energy and so on. And no one succeeded; in fact, only a few people get the -- know which the lowest one is, on the basis of their familiarity with Lewis theory. So the point is, don't be depressed when you are trying to do this. It takes a lot of lore, and even with all of the lore that's been gathered over many years, people can't use Lewis theory for this purpose because it's just not that good. Okay, but in fact the way I've drawn them there is their energy, according to these pretty good calculations. The very lowest energy is the C in the middle and H on the nitrogen, and then C in the middle and H on the oxygen. Now you can go back and look at what we said was good and bad about these things and convince yourself that "Ah! now you understand why those are on the bottom and these are on the top." But if it had come out differently you would have convinced yourself differently. Okay? So that's enough for today. Actually, there's a reference for this if you want to see it. The energies were published in the Journal of Chemical Physics in 2004.



[end of transcript]

Lecture 5
X-Ray Diffraction
Play Video
X-Ray Diffraction


Professor McBride introduces the theory behind light diffraction by charged particles and its application to the study of the electron distribution in molecules by x-ray diffraction. The roles of molecular pattern and crystal lattice repetition are illustrated by shining laser light through diffraction masks to generate patterns reminiscent of those encountered in X-ray studies of ordered solids.




Transcript



September 12, 2008



Professor Michael McBride: So as we ended last time, we said despite Earnshaw, which says you can't have these Lewis structures, might there really be shared-pair bonds and lone-pairs, and how do we know; we have to look, or feel. So last time we tried feeling with scanning probe microscopy -- AFM, STM, SNOM. They're really great. You can see atoms; you can see molecules, but you can't see bonds. Okay, so what's the key word here?



Students: Lux.



Professor Michael McBride: Lux, right? So maybe we'll see it, if we can't feel it. Now this is the entrance to this building, the old delivery entrance out on Prospect Street. And you know there's interesting stuff all over the building. Up here on top there's a fluted filter paper and a funnel, but hidden back in the shadows is something that's a little surprising to you, back there, which is a microscope. What science do you associate that with?



Students: Biology.



Professor Michael McBride: Biology, not chemistry, right? What's it doing? Well at least it's back in the shadows there. But maybe the eyeball up there says we can see the things we can't feel. Okay? This is a picture -- does anybody recognize this?



Student: Flea.



Professor Michael McBride: You know what it's from?



Student: Bubonic plague?



Professor Michael McBride: It's from Robert Hooke's great book, Micrographia, from 1665. It's a wonderful book. The page is about this big. And imagine the person who had fleas, seeing that for the first time. It's exquisite, this drawing. And he said in the book:



"But Nature is not to be limited by our narrow comprehension; future improvements of glasses may yet further enlighten our understanding, and ocular inspection may demonstrate that which as yet we may think too extravagant either to feign or to suppose."



So if he had a microscope that could do that in 1665, we must have a microscope now that's powerful enough to see bonds, right? And in fact we do. In fact, there's a brand new one that they promise will come online next Tuesday in a room about 100 yards over that way, and the trick it uses is the same one Newton used to measure the distance of the air gap for Newton's rings, or that Hooke explained by the pulses, right? And it's interference that comes from scattering. So you've seen oil on water, thin layers of oil on water, and you see rainbows in them. Why? Okay, so light comes in and scatters from two surfaces, that let's say are 200 nanometers apart. Okay? So the path difference between the one that reflects from the top and the one that reflects from the bottom, the difference is 400 nanometers; one goes 400 nanometers further than the other. Now suppose that happened to be one wavelength. Right? Then if that purple light that came in, one going further than the other, they would come out exactly one wave apart. So it would be as if it were a single wave and they would reinforce one another and you'd see the purple light strongly. Okay? But if the wavelength difference were half -- if the path difference were half the wavelength, like for red light --



Oh-oh. What do we do here, Elaine? Better get it quick, before something comes up. You're a good sport.



[Technical adjustments]



OK. Thank you.



Professor Michael McBride: If it's one wavelength path difference, then it reinforces and we get a nice strong wave coming out. Okay? But if it's half-a-wavelength difference, like for the red light, then they cancel. One is a maximum when the other is a minimum, and they're zero. Okay? So you get no scattering. So that's why you get different colors, because the oil is different thicknesses in different parts of the slick. Okay, so here's the new machine that's right over here that's supposed to come on -- new machines actually; there are two of them; it was a package deal. And that's Chris Incarvito who's the director of the Chemical Instrumentation Center and the proud owner of these two new machines. So the one he's looking at there is a user -- is to be operated by users. So it's just you sort of push buttons and you get where the atoms are in the molecule -- at least that's the hope -- and it costs about $200,000. Okay, so there's a little thread there that has a crystal on, but you need a magnifying glass to see it. You won't see it with your eyes; it's a tiny, tiny crystal. Okay? So here's an x-ray tube. X-rays come down the pipe there, hit the crystal and get scattered, and are detected by this CCD detector. Okay? And then from that information, those scattered rays, you get where the atoms are in the molecule; more precisely, as you'll see, where the electrons are in the molecule, or in the crystal.



The other machine that he's especially proud of costs $350K. And why is it more expensive? Because it collects more. Instead of using that little disc of a CCD, it has a curved image plate. Right? So when the x-rays come down here and hit the tiny crystal, many more of them get collected by the plate. So it's more efficient. So those will be coming soon. And if all you have to do is press a button and the structure comes out, that seems fine. But that's not the goal of this course, to press a button. We want to understand how such a thing works. Seeing individual molecules, atoms, and maybe bonds, there's a problem, and the problem is wavelength. Because you know in principle you can't resolve things that are closer together than the wavelength of light, and the wavelength of light we've been talking about here is 400-800 nanometers. Whereas what's the distance of a carbon-carbon bond? Anybody remember?



Student: 1.86.



Professor Michael McBride: One and a half angstroms, right? So three orders of magnitude smaller than the wavelength of light. So this is a problem, and it's why, as you'll see, you use x-rays, because they have short wavelengths. Now to understand this we'd have to know what light is. So what is light? You people have studied physics and so on. So what's light? Wilson?



Student: Who knows?



Professor Michael McBride: Who knows, okay.



Student: An electromagnetic wave.



Professor Michael McBride: It's an electromagnetic wave. Okay. Now I've seen -- indeed it is -- I've seen waves on the ocean, right? Does it mean that electromagnetic waves are like those waves? In what way is an electromagnetic wave a wave? What's wavy about it? Well you can make a graph that's a wave, that involves light. So here what we're going to plot is the force on a charge. The charge is at a particular position. We'll have the charge over there on the right; it's fixed there, and we're going to measure the force on it that'll make it accelerate up or down. Okay? Oops. It doesn't work quite the way it does on mine. Please let me know when that happens. No maybe it's going to work. [It does not.]



What that first thing did was went up and down, up and down, up and down. Okay? But if you plot it then as a function of time, it's a wave. Right? The field, at a particular point, as a function of time, is a wave. It goes up and down and up and down. Okay, now so we plot the field, or the force on a charge at one position as a function of time.



But there's another way of making a plot that also shows light as a wave, which is to show the field at different places at the same time. Right? So we're going to change it and now the horizontal axis is the position but it's at a particular instant. So as the thing goes along there are different forces at different positions, okay? And again, that will be a wave. So that's the sense in which light is a wave.



And what is it that we're plotting? We're plotting the electric field. So light is associated with an electric field that goes up and down. Right? And you can plot it in time or in space and get a wave. So that's the sense in which it's a wave. It's also, incidentally, a magnetic field; it's electromagnetic. Maxwell dealt with things like that. And there's a magnetic field perpendicular to the electric field, but we don't care about it, at least not until next semester when we talk about nuclear magnetic resonance, because the electric field is so much stronger in its effect on molecules. Now, its effect is on charges; on electrons, on protons, therefore on nuclei. Now, accelerated electrons scatter light. So here comes the light in. We'll see if this works. Okay, so it makes the electron go up and down, as the light goes by. But up and down moving of electrons, accelerating electrons, is what creates electric fields. Right? That's what an antenna is, is electrons moving back and forth.



So when the original light comes through and hits the electron and makes it vibrate up and down, that electron vibrating emits radiation, electromagnetic radiation, in all directions; all directions except this direction, and stronger the more you're perpendicular to the direction it's going up and down. But anyhow it scatters the light. So most of the light still goes through the direct beam, but a little bit of it gets scattered; much less would be scattered -- a single electron wouldn't scatter very much, you'd need lots of electrons to scatter a lot, and they have to be cooperating. Okay? Now you tell me, why are we interested in electrons scattering light? There are just as many protons in a sample as there are electrons. Why don't we worry about the protons scattering light?



Student: Electrons are a lot less massive.



Professor Michael McBride: So what?



Student: They move a lot more easily.



Professor Michael McBride: Ah, the electrons are what moves. You need the moving charge to generate the scattered light. The lightest positive charged things are -- of normal particles -- are a thousand times heavier than the electrons. So they don't move very much. It's the electrons that scatter the light. So with x-rays, you see electrons, not nuclei. Okay, they're too heavy. Now then, but you need a lot of electrons to scatter enough to see, and they have to cooperate. Right? So here we have ripples on a pond; you've all seen that kind of thing with a little bit of rain falling. And so suppose you had two spots that are generating ripples, two electrons let's say, and they give off circular waves here, in two dimensions, and they interfere with one another. But you can see a pattern is emerging, and when you're very, very far away from these, compared to the distance between the electrons, as indeed you are in these samples we're talking about where you have an infinitesimal crystal and the detector plate is a substantial distance away, right? The electrons are really close to one another. So you're way, way, way far away on the scale of the distance between the electrons. Okay?



And now you can see what pattern there is for high points and low points, for waves. Okay? Because you can see here that there's a pattern like that, that goes along the maxima. Everybody see that? And halfway in between those dashed yellow lines, there's no change at all; nothing's happening there. Right? And if you go out very, very far away, you can see that asymptotically these approach these straight lines that originate between the electrons, or on the crystal in a real case. So what you have coming out of the crystal are straight rays of light, x-ray light. Okay? And the angles at which they come out depend on where the electrons are. If the electrons are closer together or further apart or displaced in this direction, you'll get different patterns of the ripples. So somehow it must be possible, or there must be a connection between the positions of the electrons and where these rays are coming out. Right? And if you can go backwards, to go from where the rays come out, to where the electrons were, then you've solved the problem. Right?



You're not creating an image, the way you would with a normal microscope. You're trying to interpret the scattering. So the angular intensity distribution, at great distance, depends on the scatterer distribution at the origin; that is, in the crystal. Now, if you have normal light, and a lens like Hooke's microscope, then you can use these lenses to refocus it. So the same information is coming out. The sample there is emitting these rays, but a lens collects them and refocuses them to make an enlarged image that you then observe. The problem is that you can't do that with x-rays. People are making efforts now, with nanofabrication kind of things, to make things that will act like lenses for x-rays, and they've made some steps, but nothing like you would need to actually observe electrons in a crystal. Okay? Be sure to read -- there's a webpage -- have some of you read it so far? -- that has to do with what I'm lecturing on now, which will help you a lot I think; so look for that. So we're interested in seeing molecules, atoms, bonds, collectively, by x-ray crystallography. That is, we're not seeing the image of individual electrons, we're seeing the scattering that comes from all the electrons acting together. So we're not seeing them one at a time, we're seeing something collective about them. So that, for example, a real sample, or a sort of fake sample, of benzene, which had a bunch of -- six carbon atoms in a ring, might look like this at any given time. It's sort of regular but things are a little bit one way or the other, vibrating and so on.



Now you wouldn't see this in x-ray, because everything is cooperating in what you're getting. You'd see some sort of average of all of them, and it would be a little bit smeared because of that. Right? You don't see -- with scanning probe microscopy, with these sharp tips, you actually can feel individual atoms, and if one atom is someplace else, you'll see it, or feel it there. That's not true in x-ray. You see an averaged structure. And it's averaged in two ways. There's blurring, from motion and from defects. There's one benzene molecule missing there. It's time-averaged because it took you time to collect these, this information. With synchrotrons and really, really intense beams, people are trying to get faster and faster data collection, but still what you see is time-averaged on the scale of how fast atoms are moving in your sample. So it's time-averaged, what you get. And it's also space-averaged, right? It's as if you put them all on top of one another, that's what you would see, but some are displaced a little one way, the other, and so on. So it's a little fuzzy. Okay? And this is an advantage for scanning probe microscopy, which operates in real space, actually feels individual things.



Okay, so x-rays were discovered in 1895 by Röntgen, using Crookes' tube that we talked about. Okay? And he took a picture of his wife's hand there, Frau Röentgen's hand, in 1896. But what he sees is not a picture that you can blow up, like with a microscope, because it's just a shadow. All the bones do is stop the x-rays that are going through. So you don't get something that's enlarged. Right? You're not going to be able to use it the way you use medical x-rays. Okay. But in 1912 Max von Laue invented x-ray diffraction, which is this scattering and detect -- not trying to focus things, not using a shadow, but looking at these rays that come out and trying to figure out from them, something about what the atoms were. And that's his diffraction picture. That's what you'd see on that -- remember there was that round CCD plate that's on this new machine?



You might get a pattern that looked sort of like that. That was copper sulfate in 1912. But the real breakthrough by was William Lawrence Bragg, who was 22 years old. He had just graduated from Cambridge University when he determined the structure of a crystal using Laue's x-ray diffraction pattern. So he figured out how to go the other way, to go from the x-ray's scattered pattern to what the atoms were that were doing it. Right? And he did this in 1912 when he was 22-years-old, and he got the Nobel Prize in 1915. He's still the youngest Nobel Laureate. Right? His son gave me permission to use this picture. He's a very nice guy, lives in Cambridge. Okay, then of course we've come a long way from that. I think you probably recognize this picture, right? What is it?



[Students speak over one another]



Professor Michael McBride: It's the scattering from DNA. Is it a picture of DNA?



Students: No.



Professor Michael McBride: No it's not a double helix. What it is, is the way x-rays come out when they hit the double helix. Okay? And you have to figure backwards, and that's what Crick was able to do; and we'll discuss how he did that. Okay, so that's 1952, 40 years after Bragg's discovery. And then just in 2000, not so long ago, this has gone to the complete structure of the ribosome at 2.4 angstroms resolution, which was done here at Yale, in this building and the next building. Okay, and that's what it looks like. It has twenty-five nanometers across, 250 angstroms, as long as whatever number that is; lots of carbon-carbon bonds. And you see all those atoms. There are greater than 100,000 atoms, not counting the hydrogens. Incidentally, why is it easier to see other atoms than to see hydrogens?



Student: They have never counted them.



Professor Michael McBride: Because that's where the electrons are. Hydrogens come with only one electron. Other atoms come with lots of electrons. So it's hard to see hydrogens; much easier to see the other things. Okay, now what can electron diffraction show, x-ray diffraction show? That's what we want to know. Can it show molecules? Yes. Can it show atoms? Apparently; I just showed you a picture. But what we really want to know is whether it can show bonds and whether Lewis was right. So to understand this, we have to know how diffraction works. I mean, you could just be like the people that will go and punch the button on that machine, but that's not what I think you would be satisfied with. So I'll help you out. So like all light, x-rays are waves; they just are very short. So now I'm going to demonstrate with a machine here, which was designed for an overhead projector, but I believe it's going to work here.



Okay, so here are our waves. Okay, so here's a wave coming in. Okay? And when it gets to this position, it hits an electron here and an electron here; forget this one for now. Okay, everybody got me? At a given time, these two electrons are being pushed up; then they're going to be pushed down, as this wave passes by. Okay, now as those two get pushed up and down, they give off waves in almost all directions; all directions for our purposes. So they give off waves in all different directions. Right? Now notice the one, the part of that scattered x-ray that goes straight forward, in the same direction the original light came in, is bound to be in-phase with one another. Right? And we can test that with this line here, because if they're going right straight ahead, this one is a maximum when this one is a maximum. Everybody with me on that? And it'll keep that. So you'll get scattering by both of them cooperating, coming out of it straight ahead. But how about if it's at an angle? Okay, so I can push this up and change the angle. How about at that angle, how much light is going to be coming out from these two electrons?



Student: None.



Professor Michael McBride: None, they exactly cancel each other, right? But if I go to this angle, now they're just as strong as they were originally, right? And then it'll go weak, weak, weak, weaker, weakest, nothing; then stronger, stronger, stronger, stronger, very strong, weaker, weaker, weaker, weaker, weaker, zero. So there'll be a modulation -- as you go out at different angles here, going either up or down, it'll be strong, then nothing, then strong, then nothing, strong then nothing. Right? So now what will determine, what will determine how frequently these angles recur, at which reinforcement occurs? Is that clear, the question? What determines the angles? Yeah -- pardon me?



Student: The wavelength.



Professor Michael McBride: Ah, obviously the wavelength. What if the wavelength were very, very, very short?



Student: Lots of angles.



Professor Michael McBride: Then you'd get lots of them. You wouldn't have to do it very much before the next one came in, if they were very -- what else determines it, besides the wavelength. Yes John?



Student: The distance between the two of them.



Professor Michael McBride: The distance. That's what we're really interested in, is using this; knowing the wavelength of the x-rays, using this to measure distance. Okay? Now let's forget the one on the bottom here. Let's put this one in. Okay? Now -- oh okay, for reference let's look at the bottom a second. So the first reinforcement came here, between the top one and the bottom one. Everybody with me? How about for the top one and the middle one, where did the first one come? You had to go to a higher angle, there, to get reinforcement between the top and the middle. Right? So the closer things are together, the fewer the angles are. Everybody with me on that? Notice that's a reciprocal relationship. The closer things are in what we call 'real space,' the distance between the electrons, the further apart are the rays that come out, in angle. Okay? So it's backwards. Closer together, further apart, in what's called 'reciprocal space.'



Okay, now suppose you had a whole row of electrons that were evenly spaced. Okay? So here's the first, the second, the third. Now we'll take all three of them. Right? And what you notice is that as we go out it gets weaker, weaker, weaker, weaker, weaker, weaker, weaker, weaker, weaker. Now when we get to the angle here, where the first and the third were very strong, what do we see now? The one in the middle cancels the first one, and if it were a very long row of them, the fourth one would -- here the second one cancels the first one -- the fourth would cancel the third, the sixth would cancel the fifth, and so on, and you wouldn't get anything. Right? But then you'd get it again when you got to the second. The one that would've been the second angle, if it was just one and three, will be strong, because this one halfway in between will reinforce. Right? So closer together, further apart. And if you have a whole row of equally spaced things, they'll all be together, according to whatever the distance between successive ones is. Yes, Shai?



Student: What are the chances that electrons are going to be spaced evenly along --



Professor Michael McBride: Ah, what could cause electrons to be spaced exactly evenly?



Student: A crystal.



Professor Michael McBride: A crystal; x-ray crystallography, right? So that's why you use crystals. Okay, that's what we want to say here, I think.



[Technical adjustments]



Professor Michael McBride: Okay, so that's the wave machine. And if you want to do it in the privacy of your room, you can go to this website at Stonybrook and download something that allows you to run a Java applet that does sort of this kind of thing. This, if you blow it up in there, the waves look sort of funny there, at the atoms, and the reason is because you're measuring the phase perpendicular to different directions of the wave. That's just to help you out, if you try this and have -- and are confused by that. Okay, now there are -- so suppose you have just an arbitrary set of electrons, and the waves coming in and hitting them, what directions will you get reinforcement in? Well there are two directions that you're guaranteed, no matter what the spacing is, and that's this. Here the light comes in and goes out. So the direct beam, you have the same path, because one of them gets hit earlier, but has a longer path coming out, and the other one gets hit later but has a shorter path coming out. So they're guaranteed to be exactly in-phase if they're scattering straight ahead. Okay? But only a little of the light is getting scattered. Most of it is the direct beam. So you don't even notice the difference, from that. So that's not very exciting. But there's another angle at which you're sure that these two are going to be in-phase, and that's this angle. Now how do we get to that angle? I've lost some of the -- since you downloaded it there's some more stuff in here now. But -- so now I can't quite remember, let me just try.



Okay, so this blue line here is called the scattering vector, and that's how different the arrow coming out is from the arrow going in. Right? That's just a funny mathematical or geometric thing that people have defined; they call it the scattering vector. Right? Now this length is exactly the same as that length. This length is exactly the same as that length, and when you turn at a given angle, those two will -- the vector between those will be perpendicular to the direction they're going, and they'll be in-phase again. Right? And if you draw the line that connects the two points, notice the scattering vector is perpendicular to that line. So you say that this incoming wave was scattered perpendicular to this line. That's how a mirror works. Right? That angle is called the specular angle, because speculum is the Latin word for mirror. Right? So you're bound to get reinforcement at the specular angle, and you know that from having looked at mirrors.



Okay, now suppose there were another electron on that same line or plane. Okay? For the same reason it's guaranteed to be in-phase. So everything on that plane will scatter in-phase at that particular angle. They'll all reinforce one another. Great. So all electrons on a plane perpendicular to the scattering vector, scatter in-phase at the specular angle. Now suppose you have a whole bunch of electrons and you have to figure out how they're going to reinforce or cancel one another. There's a trick you can use. You see if you can discern a set of planes that are evenly spaced, that contain all, or almost all, of the electrons you're interested in. Okay, now if you look at this pattern, you can see that there's a set of planes, equally spaced, that passes through all of them. Can you perceive it? There.



Okay, there are three electrons on the first plane, four on the second, two on the third and one on the fourth. Okay? Now why is that handy? Okay, so there's the scattering vector, perpendicular to these planes, and all the electrons on any one plane will scatter in-phase with one another. Okay, so it's as if there were a single electron three times bigger, or four or two or one, on those successive planes, because they're all going to be in-phase with one another. So we could pretend they're all at one point, on any given plane. Okay? So now all we have to worry about is the phase relationship from one plane to the next. Got that? If they were in-phase from one plane to the next, as well as on the planes, then everything will be in-phase and you'll get really bright reflection coming out at that scattering angle. Okay? Now, so let's think -- so it's as if we had these collected electrons all lying on the scattering vector and we want to know whether they're in-phase with one another.



Okay, so here's light that comes in and out at a certain angle from the first one; three electrons worth of scattering from that, the ones that were on that plane -- okay? -- and four from the second plane. Now are the three and the four in-phase with one another? What condition would have to apply in order for those to be in-phase with one another? They have different path lengths, but if the paths differ by an integral number of wavelengths, then they'll be in-phase with one another and reinforce. Okay? So there's the path difference in red. If that happens to be -- suppose the wavelength of the light, and the angle of the scattering -- which also determines that path difference, as you can see -- suppose that the wavelength and the angle are such that that happens to be one wavelength? Okay, so those are in-phase with one another; all seven will be scattering together. Right? How about the next one, with two electrons, how about its phase? What was the condition we talked about, up here at the top?



Student: Evenly.



Professor Michael McBride: Evenly spaced. So what do you know about the next one?



Student: [inaudible]



Professor Michael McBride: It's going to be in-phase too, two waves behind the first one. Right? It'll be two wavelengths behind the first one, exactly, at some angle. Right? And the next one will be three wavelengths behind. So all these electrons are going to be scattering in-phase at that particular angle. So you get a really strong reflection coming out. Right? It'll be as if it were all the electrons working together. So the net in-phase scattering is as if there were ten electrons doing the scattering. That's great, right? Now suppose that the first path difference, instead -- suppose you had a different wavelength or a different angle. So the first, between the three and the four, suppose the difference in path was half a wavelength. Then how would it differ?



[Students speak over one another]



Professor Michael McBride: How about between the first and the second. Wilson?



Student: Cancel.



Student: The first and the second?



Professor Michael McBride: The first and the second, would they exactly cancel and get zero?



Student: There'd be one left.



Student: Zero.



Professor Michael McBride: Why?



Student: There's one --



Professor Michael McBride: Ah, there are three in the first and four in the second, right? So you got plus three but minus four. How about the next one? Speak up gang.



Students: Plus two.



Professor Michael McBride: Plus two. And the last one?



Students: Minus one.



Professor Michael McBride: Minus one. And what's the net scattering?



Student: Nothing.



Students: Zero.



Professor Michael McBride: Zero. Okay? So you can see how this is a neat trick to work. If you can see in the pattern a bunch of planes which would contain the electrons, then you can figure out in a particular direction, how strong the scattering would be. Now, we're going to do an experiment and it requires the room to be dark. And so I'm going to start turning the lights off and ask Filip to get into position to do stuff in the dark here. Okay, and turn this one off. Okay. Now I'll show you first what we're going to do. So pull out the little thing there. And there's a laser that's focused right here, but it's too bright, because the other things we're going to see are very dim, which is why we had to turn the lights off. So I'm going to put this black tube there, with a little hole in the end. Hopefully I can get it positioned so that most of the laser will go inside here and we won't see it very well. Let me get it positioned right. It's a moving target so it's not so easy.



Okay now, so here's the view from the ceiling. There's the screen, and this laser is coming from the back of the room and hitting the screen, right here. Okay? Now what Filip is going to do is put things called diffraction masks, which are just slides, thirty-five millimeter slides, and he's going to put them in the path of the light. And that's, the distance from Filip to here, is 10.6 meters; I measured it. Okay? And so put in the first slide please. And see this? Those are all deflections at different angles. And what's doing the scattering is a slide that looks like this. Okay? A jail window, right? Okay, so what it's giving is a row of dots. Now I'm going to ask Filip to do two things here. First, rotate the slide around this axis, so rotate it like this. What do you think will happen? You see, right? The row of spots is perpendicular to the direction of the lines. Okay? And now I'm going to ask him to do something else. I'm going to ask him to twist the slide like this, which changes the effective distance between these. And notice what happens as he makes them closer? Twisting makes them look closer together, right? What happens to the spots?



Student: Farther apart.



Professor Michael McBride: Does that surprise you?



Student: That's that reciprocal --



Professor Michael McBride: That's that reciprocal relationship. When things get closer together the angles get bigger. Okay, now things are going to get darker here. So I'm going to do some things. Turn that out and also get the room light. Well no, I want to show you something here first. Okay, so I'm going to show you what the masks are going to be and then we'll show you the effect from the masks. Okay, so the next -- so here's a question for you to work for homework. What is the spacing of the lines on Filip's slide? How far are the bars in the jail window apart on that slide, in order to give a spacing here -- oh I did the wrong, I was premature with that -- in order to get that 10.8 centimeter spacing here, at a distance of 10.6 meters, with 63 [ correction: 633] nanometer wavelength light? Okay? You can use that to find the distance between the bars on his slide, and I want you to do that, because if you can do it then you understand how this works. Okay, but the next one he's going to show -- not yet but he'll put it in -- is one that's a similar spacing, but of pairs of lines. There's a pattern being repeated, pairs of lines. Okay? And then he's going to go to a whole bunch of hexagons of dots, which we will call benzene, but it's not exactly like benzene would be, a benzene gas, because they're all oriented exactly the same way. Right? So it's oriented benzene; that's the next one we're going to look at. And then we're going to look at this one, which is also oriented benzene. Can you see the difference in the pattern between the one on the top and the one on the bottom? Those red lines are a hint.



Student: They're connected.



Professor Michael McBride: In the bottom they're all pairs of hexagons, that are distributed randomly, where in the top they're individual ones. Now then he's going to show you this one. How's that different?



Student: Quadrupled.



Professor Michael McBride: It's quartets of hexagons. And finally he's going to show you this one, which is a crystal of hexagons, and in a crystal they truly would be oriented. Okay? And then we're going to -- the pièce de résistance, at the end of this sound and light show, is going to be that. You know what that is? It's a light bulb filament, that he's going to hold up in the path of this; so a helix. Okay, now we're ready to do the trick. So we'll mute that and do this, and if you have a laptop open you'll need to close it because we got to -- it's not very bright, so we have to make it as dark as possible, and I have to find my way back there and close my laptop. Okay you ready Filip?



Teaching Fellow: Yes.



Professor Michael McBride: Okay, so first it's going to be pairs of jail bars. So how does it look different from what we saw before?



Student: Like triplets.



Student: Like different lights in there.



Professor Michael McBride: It's the same kind of dots but their intensity is modulated. They're not just evenly -- they're not just all the same or slowly dying away as you move out to right or left. They come in a pattern. Okay, now let's do the next one. This is the one that's benzene, but randomly, a random collection of benzene. So how would you describe that?



Student: A snowflake.



Professor Michael McBride: It looks like a snowflake. Oh, my laptop has come on. [Laughter] Go figure. I'm putting things to cover the lights up here. Okay, so it's a snowflake. Now we're going to look at pairs of benzenes. How is it different? It's still the snowflake.



[Students speak over one another]



Professor Michael McBride: But there are bars across it. Everybody see that?



Student: Yes.



Professor Michael McBride: Can we stop the green blink back there or put something across it? That's good. Okay, now we're going to do quartets of benzenes. How does that look? How would you describe that?



[Students speak over one another]



Professor Michael McBride: It's now got bars going in both directions, not just horizontaloid but also verticaloid; they're actually vertical I think, but not quite. Okay, now a lattice of benzenes. Isn't that great? So it takes the same -- it's the same underlying pattern, the snowflake, but it concentrates all the light that would've been spread out, into very, very fine points. Remember when it was just pairs, we put bars across, but it was the same amount of light coming out, so that it got focused sort of, and when we have a whole lattice it gets fabulously focused. I can see things quite far out from the middle here, because I'm so close to the screen. I'll hold my hand up but you can't see my hand. [Laughter] They go way out. So that's what a crystal is. The crystal takes the same underlying pattern, that comes from the molecule, and focuses it into intense points at many, many angles. So it's like looking at the scattering from a single molecule but looking at it through a pegboard. You know what a pegboard is? In a tool shop you used to put a pegboard up. It's a piece of masonite or something that had holes in it, regularly, and you could put little hooks in it to hang your hammer and your saw and whatnot. So it's as if you have, you're looking through such a pegboard at that underlying pattern. And now we're ready for the pièce de résistance, which is the light bulb filament, which is hard to see. This is why we really needed the lights out. We needed it for the snowflake too because it was so -- so what do you see for the light bulb filament?



Students: An X.



Professor Michael McBride: You see an X, right? Can people in the back see the X? And there are dots along the arms of the X. Okay? So that does it, you can open your laptops again now and we'll try to get some lights back on. Sorry for straining your eyes.



[Technical adjustments]



Professor Michael McBride: Okay, so there's the scattering from these things and we can -- now so let's try to understand it. So this was the slide with randomly positioned but oriented benzenes. It turns out that random positioning generates the same diffraction as a single pattern but gives it more intensity. If we'd just put one hexagon there you wouldn't have seen enough from it. Okay, so here's the pattern you got from isolated benzenes, which is that snowflake pattern. Right? And when we had a lattice made of those, you saw the same underlying pattern, but only at little -- only at regularly spaced spots that had to do with how far apart the hexagons were from one another, and much more intense, because all the light that would be scattered all over the screen here is focused in those little dots. Okay, now the thing that generated this snowflake was benzene, or what we're calling benzene, the hexagon. Okay, now can you see what we need in order to understand? What do you look for in order to see what directions will give you scattering? Ah sorry, I was having so much fun we went over. So we'll talk about this next week. Thanks.



[end of transcript]

Lecture 6
Seeing Bonds by Electron Difference Density
Play Video
Seeing Bonds by Electron Difference Density


Professor McBride uses a hexagonal "benzene" pattern and Franklin's X-ray pattern of DNA, to continue his discussion of X-ray crystallography by explaining how a diffraction pattern in "reciprocal space" relates to the distribution of electrons in molecules and to the repetition of molecules in a crystal lattice. He then uses electron difference density mapping to reveal bonds, and unshared electron pairs, and their shape, and to show that they are only one-twentieth as dense as would be expected for Lewis shared pairs. Anomalous difference density in the carbon-fluorine bond raises the course's second great question, "Compared to what?"




Transcript



September 15, 2008



Professor Michael McBride: So we want to understand diffraction patterns. It's clear that there's some relationship between the arrangement of the electrons that are doing the scattering and the pattern of the scattering you get. Right? At a particular angle there's a particular intensity, and it must depend on how those electrons are arranged. But how is the interesting question. So what we're going to do now is look just a little bit-- -- no one figures this out, in a real case; people have written fancy computer programs to do this. But so that it doesn't seem mysterious, I want you to get the idea of how it works. Okay, so the first slide we showed, or the third slide actually that we showed last time, was made of a bunch of these things we call benzenes, the collections of six dots, and they're all oriented the same way; that's not very reasonable except when they get regular in a crystal, which is where we're going. But anyhow, if you have a whole bunch of them, but randomly distributed, it turns out the intensity you get is stronger, proportional to the number of them that are doing the scattering. Right? But the pattern you get is the same you'd get from just one. So that's why we use this funny slide.



And this was the pattern that came out, this thing that looked a little bit like a snowflake, and I'd like you to understand why that hexagon of dots gives a snowflake. So we take the hexagon of dots, and what do we look for to see where scattering might be, to see--what's the--remember what the trick is? The trick is to look at things that lie on equally spaced planes. Try to choose a set of planes so that as many of the electrons as there are will fall on or near that set of evenly spaced planes. So And this is two dimensions, so we're looking for lines, or cross-sections of planes, right? So how can we draw lines on here, evenly spaced lines, so that all the dots will fall on them? Well we can do it this way, right? So they're evenly spaced, three of them, and all the electrons fall on or near them. Notice, they don't have to fall exactly on, because if they're displaced a little bit, the phase of the wave will change a little bit, compared to the others; but only a little bit. So it just has to be near, not exactly on. Okay, so there we have it. Now, so that means that you'll get scattering in that direction, where they will reinforce one another.



Let's go back to the display we were looking at before, here. Okay? We were talking about this just as class started. So a wave comes in, makes these things all move up and down, the electrons, here and here, at the same time. Actually, let's make all three of them vibrate for current purposes. Okay. Now we're interested at what angles light can come out and all these things will help each other out, reinforce one another. So as we go here and look at that vertical line, we see that they're not all in-phase at that angle. But as we increase the angle, when we get to say here, it looks not so bad. The top and the bottom are actually exactly in-phase, but the middle one is exactly out-of-phase, and it'll cancel out the top one. And if there were a whole row of these, the second one cancels the first, the fourth would cancel the third, the sixth would cancel the fifth, and so on; there wouldn't be anything coming out. And now we go a little bit away from that. And it's worth making a point here. Suppose that this were a very, very, very long line of scatterers, right? And the first two--let's get to this place here where they're all exactly in-phase; that'll obviously be a very strong angle. But suppose we looked at another angle that was close to that, but just a teeny bit off. So the first two are off by just 1° in-phase, and the next one's off by 2° in-phase; not much off, okay? So are we going to get pretty strong light coming out from that? Russell?



Student: If it's a very long string, you'll eventually get one that--



Professor Michael McBride: If it's a very long string, if the first one's 1° out, by the time we get to the 180th, it'll be 180° out, or exactly--so the 180th will cancel the first one, the 181st will cancel the second, the 182nd will cancel the third, and so on. Right? So it has to--if you have a really long string, it has to be exactly on, from one to the next, in order to get them so that you get something coming out. And what you get coming out then will be quite strong because it'll be the scattering from the whole bunch of them. Right? Where did you see that in last week's lecture, Friday's lecture?



Student: In the dots.



Professor Michael McBride: Right, you saw it that you got very fine dots when you had a really long row of things. Okay, so back to here. So there'll be things that are almost exactly in-phase, and it'll be a certain angle at which they'll be in-phase, and that angle is shown by the end of that arrow, and you get, therefore, a dot there, and also, by symmetry, the other five of the six dots; but you could draw the lines other angles the same way. Everybody with me on that? Okay, now so we look for electron density on or near evenly spaced planes. Now because it only has to be near, there are other directions we could draw it, so we get a set of planes where they're near but not on; like here, right? Now one is a little above the plane, two are a little below the plane--, right?-- Or vice-versa for the second line, right? But they're pretty close, and they're at a certain spacing. Now, that spacing is larger than the blue spacing. Does everybody see that? What does that mean about the angle at which the reflection will occur, or the scattering, where they'll be in phase?



Students: Smaller.



Professor Michael McBride: The angle will be smaller. So the part of the snowflake that comes from that is here, right? A smaller angle because it's a larger distance. And you can see immediately why that is, if we look at this thing, that if you look at the top to the bottom, where they're far apart, the lines are far apart, you get the first reflection there. But if the lines are closer together, like these two, then you have to go to a higher angle in order to get that first one, first coming into phase. Right? So that's that reciprocal relationship. Okay, so we get a set of six dots like that, and that's most of the pattern. But there are other places too you can see it. For example, you can take those lines. In that set of lines, every-- -- the center of every dot is on a line. There happens to be a line in the middle that doesn't have anything on it; but who cares? Right? The others are on evenly spaced lines. Right? So they'll all be in-phase. Where will that one come, that reflection, close in or far out?



Students: Far out.



Professor Michael McBride: Far out, and there you see it, way out at the edge. Can you see it there? Just a little bit with--I could turn the room lights down, but I don't think it's worth it. You can see that it's there. And so there are six of them out there as well. Okay, so you got the idea of how by looking for evenly spaced planes you can predict what the pattern will be for a certain arrangement of electrons. Now, notice those ones far out are very weak. That's because although the centers of the dots lie right on the lines, the lines are getting so closely spaced that there's part of the electron density that's not right on the lines, that in fact gets out to almost halfway between the lines. Right? And that part of the electron density will be canceling. So most of the electron density is all in-phase here, but some of it is out-of-phase, and as you get the lines closer and closer together, the fact that the electron distribution is finite, and not tiny points, means that things will get weaker and weaker as you go to higher angles. Okay, so there again we notice here that the closer spaced planes give higher angles. This is what we've just been talking about.



Now, when you take a slide with a regular lattice of these benzenes, then--remember, if you took just one, you get this snowflake pattern. If you take a whole row, across the top, like that-- -- oh pardon me, I meant to say that one will be very weak, you can hardly see it. What you need is a bunch of them to cooperate, to see something that you can see. Now first we could take the whole row along the top, and we're going to look only at scattering in the vertical direction; so I've narrowed in on the picture. But if we take that whole row along the top, remember, they'll all scatter in-phase at the specular angle, because--for scattering up and down. So now it gets more intense because they're all cooperating. For other directions they wouldn't all be in-phase from one hexagon to the next, but for scattering up and down, where they're all at exactly the same level, they'll all be exactly the same.



Okay, now suppose I add the next row. Will they be in-phase, for the same direction of scattering up and down, where that top row is all in-phase? So we would get the scattering shown here, right? Stronger than it was before because it's a bunch of these things now, right? But still not really strong. But let's add in the next row. Is that going to make them all stronger again? No, most of the time they'll be out-of-phase, from the next row, for scattering in this direction. Right? Most of the time they'll be out-of-phase, but for particular angles they'll be in-phase. So in fact what that will do, if we add the next row, we've made the picture brighter because we've got twice as many things scattering now, but there'll be bars across it, like that. It's always the same pattern, getting brighter and brighter as we add more of them, but at certain angles they're in-phase now and at certain angles they're out-of-phase, from one pattern to the next, or from one row to the next. Now, what's going to happen if we add the whole sheet of them? Can you see? Sam?



Student: Well you're only going to get very thin--



Professor Michael McBride: Ah, there'll only be really precise angles, at which they all interfere; otherwise the 180th will interfere with the first, the 181st with the second and so on. Right? So if we add the whole phase, the whole thing, then if we look at the pattern it's going to be brighter, right? But it'll be only dots, only at very precise angles. Right? And that's what we saw when we went to this final picture for scattering from a lattice. So what this means is that the lattice repeat concentrates the snowflake scattering into tightly focused spots, on a regular pattern. Right? So this is the diffraction from a 2-D lattice of benzenes, which is the same as the underlying snowflake that came from just one, but it's as if it were being viewed, as I said last time, through a pegboard. Right? All that scattering gets concentrated into a few points, very bright, and these regular points. Now some people seemed not to know what a pegboard was, so I took a picture of a pegboard over the weekend. How many people have never seen a pegboard before, just to calibrate me? Okay, most people have seen a pegboard. But you can put these hooks in it and then hang whatever you want on it, like tools or pictures or whatever you want to put on your wall. Okay, so it's like looking through the holes in that, at the snowflake pattern.



Now, there are two, quote, spaces that get talked about: direct space or real space; and diffraction space or reciprocal space. Okay? So the crystal is the real thing that you're interested in, and diffraction photo is in diffraction space; you see these dots that Lowy Laue saw, for example, for copper sulfate. Right? So in real space you have a unit cell structure; that is, the thing, the pattern that then gets repeated to make the crystal. But you have the pattern, right? And that gives rise to this fuzzy pattern, like a snowflake in reciprocal space. And then you have a crystal lattice, a regular repetition of these patterns, right? And that's as viewing through the holes in the pegboard. Okay? And remember that decreased spacing in real space corresponds to increased spacing in reciprocal space; which is of course why it's called "reciprocal". Okay, so that's what we want to understand about how it works. Right? There are patterns which give rise to patterns in reciprocal space, that fuzzy thing, and then if you have a repeating lattice of that, it gets concentrated into particular points; but it's the same underlying pattern. So there's the pattern, which is the structure of the molecule, and there's the lattice, which is the structure of the crystal.



Okay, now let's look at the light bulb filament, the last thing we looked at, and see why the scattering from that looks the way we do-- -- looks like this X with dots along it. Okay, so we take-- -- we want to find if there are electrons on this helix, right? We can find a set of evenly spaced, parallel planes like the yellow ones here. So we'll get scattering perpendicular to those planes. And sure enough there it is, that spot. But we could also take a set of planes, and it goes not only up but also down, both branches; both directions of that particular branch of the X, come from the same planes. Okay? But you could make planes that are twice as close to one another. Right? What will those give rise to?



Student: Further out.



Professor Michael McBride: Twice as far out; the sine is twice as big actually, not the actual spacing. But it's that one, the second one out along the branch. Okay? Now, will that be-- -- do you expect that to be brighter or not as bright as the first one?



Student: Not as bright.



Professor Michael McBride: Not--well it probably will be not as bright, and the reason is that for--there are two things that have to do with the brightness. One is how much density there is on the planes versus how much there is between the planes. Right? And we see here that--say take the original yellow one--there's a lot; all this stuff is nearly on the plane, and halfway between the plane there's not very much, just a thin cross-section of the wire. Okay? But as the planes get closer and closer, the fact that the wire has a finite thickness means that some parts of the wire, the same part of the wire, are out-of-phase compared to others. So because of the finite thickness you expect it to slowly fall off in intensity as you go out; but you'll then get the third one along the branch and so on. And you can do the same trick in the other direction. Right? So that's why it's an X. You can draw the analogous mirror image lines. Okay, now what does the X, the angle of the X tell you? The X could look like this, or it could look like this, right? What do you learn from that, the angle that the X makes?



Student: How tightly wound the helix is.



Professor Michael McBride: How tightly wound the helix is, because if the helix were very tight, then the planes would be like this, and if you grabbed the spring and stretched it out, they'd look like this. Right? So the angle of the X tells how tightly wound the helix is, tells the pitch of the helix. Okay? What does the spacing of the spots tell you? Did you do the homework I suggested for today, that you didn't have to hand in? Then you'll know what the spacing of the spots tells you. If you know the wavelength of the X-rays, or of the light in this case, what does the spacing of the spots tell you?



Student: The spacing of the plane that's--



Professor Michael McBride: It tells you what the actual distance is between successive turns of the helix, how far apart the planes are. So it tells the scale of the helix, right? The angle you get from the X, but how big the actual helix is you get from the spacing of those dots, if you know the wavelength and the distance that you're magnifying; in the case of the demonstration, the distance from the slide at the back of the room to the front of the room, for seeing this. And the spots weaken successively because of wire thickness, as we just explained, as you go on out. Now, that looks very much like Rosalind Franklin's Bb-DNA photograph from 1952. So we see the helix and we see how angled the successive turns of the helix are, what the pitch of the helix is, and from the spacing of the dots we get how big the actual helix is.



But there's something funny. The intensity doesn't just evenly decease as you go out. Instead of being strong, not quite as strong, less strong, weaker, very weak, as you go out, it goes weak, strong, strong, very weak, strong. Now why is that? Why do you get such a funny pattern of intensities? Let's look here at this thing again. Remember where, if we looked at the first, second, third and so on, we saw that they come in and out of phase. But let's suppose--so it would be, if you have things that are twice as close together here, as in the first case, what effect does that have on the actual pattern we see? The pattern we saw, if we had only every other one, was here the direct beam, first reflection, second reflection, third reflection, fourth reflection. How does it differ if we add one in between? Okay, here's the first reflection--can we see that, if we have all three. ? That's what would've been the first reflection if we didn't have the intermediate one. Do we see a reflection here, if we have a whole row of them? No, because every other one cancels out. So we won't see that one. How about the next one?



Student: It'll be very bright.



Professor Michael McBride: Yes, it'll be very bright, because there's more doing the scattering. How about if we go to the next one? No, nothing there, cancelled, right? How about the next one? Do you perceive a pattern emerging? What is it? What?



Student: It's successive; you see it, then you don't see it.



Professor Michael McBride: Can you say it in a more mathematical way? Suppose we numbered them. The one straight ahead we'll call zero; then there'll the first, one, two, three, four, five. What happens when we put one in between?



[Students speak over one another]



Professor Michael McBride: Is one there? This was one here, right? There was one for the top and bottom. Is that there?



Students: No.



Professor Michael McBride: How about two?



Students: Yes.



Professor Michael McBride: How about three?



Students: No.



Professor Michael McBride: No. How about four?



Students: Yes.



Professor Michael McBride: What's the--can you state it mathematically?



[Students speak over one another?]



Professor Michael McBride: Lexie, what do you say?



Student: [inaudible]



Professor Michael McBride: Can't hear very well.



Student: n+2?



Professor Michael McBride: Which ones disappear?



Student: Odd.



Student: The even.



Professor Michael McBride: The odd ones, one, three, five, seven, nine; the odd ones disappear but the even ones are still there, if we put an extra one in between. Now let's think about DNA. Here's a helix, right? But what do you say about DNA? Is it just a helix?



Students: Double helix.



Professor Michael McBride: It's a double helix, so there are two helices. Okay, so suppose we have--there are the successive turns and we get scattering at a certain angle--but suppose we make a double helix, now what'll happen? Lexie, you're going to help me again.



Student: Oh.



Professor Michael McBride: How different will it be if we have a second helix wound halfway between the first helix? What will happen to the--what we--from one helix we had one, two, three, four, five, six, seven, eight, nine angles, right? What happens now?



Student: You're missing even numbers.



Professor Michael McBride: Can't hear.



Student: You're missing even numbers.



Professor Michael McBride: Ah! All the odd ones disappear. Right? So what would you have expected to see in Rosalind Franklin's picture then? Right? It would be weak, strong, weak, strong, weak, strong, right? But generally getting weaker as you go out because of the finite thickness of the wire. Is that what we see? No, we saw--the first one is very weak, strong, strong, very weak, strong. Can you see any--how could that be? What is that telling you? Yes?



Student: They're not perfectly in-phase.



Professor Michael McBride: What do you mean they're not perfectly in-phase.



Student: They're not equally--



Professor Michael McBride: Ah. Maybe the second one isn't halfway between the turns of the first one. Right? What happens? So the planes are twice as close there and that would cancel every other reflection, all the odd ones. But suppose it's like that. Did you ever see a situation like that before where what repeats is not a line but a pair of lines? Angela?



Student: The second diffraction--.



Professor Michael McBride: The second diffraction pattern we looked at was of that sort, which showed the same dots but the intensity got modulated. Remember? Okay, so let's, instead of doing this where we have equally spaced ones, let's do one where they're not equally spaced, like this. Now--oops, it came loose here; here we go; oh no; there we go, okay. Now consider the top and the bottom one, which are in successive turns of the same helix, and this one is offset in the middle. But consider just the top and the bottom one. Okay, there's the first reflection, the second reflection, the third reflection, the fourth reflection, everybody see? Between the top and the bottom. Now tell me what happens if I add that other one offset this way. How about the first reflection there, is that going to be strong?



Students: No.



Professor Michael McBride: No, because this one is canceling it. And remember, there's also one like this below this one; remember, it's a really long row of these things. So it gets almost perfectly cancelled; not quite because it's not quite-- there would be exactly the minimum, and when the first and the third are in-phase, the second one is not exactly out-of-phase, but very close; very weak, that reflection's going to be. Now let's go to what would be the second reflection of the single helix there. How about there?



[Students speak over one another]



Professor Michael McBride: Yes, the second one now is helping out a little bit; not perfectly but it's--in fact, probably if we go just a little bit further maybe it'd be better. But anyhow there'd be a reflection there. Okay? Now how about if we go to the third? That's as good as the second one, right? And how about if we go to the fourth? Would the fourth--if they were evenly spaced, what would you expect for the fourth?



Student: Very strong.



Professor Michael McBride: Very strong, right? So let's go to the fourth; one, two, three, four. What should it be?



Students: Really weak.



Professor Michael McBride: Really weak. And how about the fifth?



Students: Strong.



Student: Really strong.



Professor Michael McBride: That's precisely what she saw, right? Weak, strong, strong, very weak, strong. So that's how much, this is how much it's offset from the center, in order to get that sequence of intensities. Everybody with me on this? Okay. So that's that repeated pair pattern that we saw before. And there you see that the outer coil of the DNA, which has phosphorous in it, with a lot of electrons, that's very much like this helix. Okay, much more electron density near these planes than between them, than halfway between. But there's also much more electron density on those, near those blue planes than in between, for the same reason, and that means that you get those spots on the side. That tells you the diameter of the helix. Okay, and the things top and bottom tell you the base stacking distance. Let me back up just a second to show you what we mean, right? So there are a bunch of electrons here, a bunch of electrons here, a bunch here, a bunch here, and that gives scattering in this direction. So all that information--the helix diameter, the fact that it's a helix, the fact that it's a double helix, the fact that it's an offset double helix, and the diameter of the cylinder, are all easily read out if you know what you're looking for; which Crick knew, because he had been studying helices of proteins, so he knew about--he knew from helix scattering.



Okay, so knowing the molecule's electron density, as we knew our artificial sort of benzene, or this DNA, it's easy to calculate the crystal's diffraction pattern, especially if you have a computer to help you out; but it's not an intellectual challenge, it's just work. Right? Using pretty heavy- duty math, people got the Nobel Prize for developing it; or a canned program, (that's what they got the Nobel Prize for), you can go the other way and go from the X-ray pattern to what it was that was causing it. Right? What we did was go from what's causing it to the pattern, but in fact you can go the other way now when you press a button on a machine. And you get out the electron density at every point in space. Okay, that's the yield from an X-ray structure. Why don't you get the nuclear positions, why do you get the electron positions?



[Students speak over one another]



Professor Michael McBride: Alex? Can't hear.



Student: It's too big to diffract.



Professor Michael McBride: Yes, it's too heavy, so it doesn't vibrate, the nuclei, so they don't function as an antenna, to give off the new waves the way the electrons do. Okay, so you get a result, which you can plot on a map. Now in the olden days when I started doing this kind of thing, the kind of map you got out was a big sheet of computer output paper, and it had numbers on it, and a given piece of paper was a given slice through the "unit cell', through the pattern, and numbers were printed out that said what was the electron density at that particular point on that level. So you had an x axis, y axis, and at this point it was seventy-five, forty-two, fifty-five, eighty-seven, or whatever the number is, and so on. And what you do is take a felt tip pen and connect dots of equal size, and you'd get a contour of that particular electron density. And here's where that was done for a particular compound.



This is taken from a book written at that time by Stout and Jensen. And you drew these by hands because you didn't have a computer that could draw--that could figure out where to draw it and so on. So here were atoms in a particular crystal structure. Then you had a whole bunch of sheets of paper for different slices, through the electron density; stack them up, right? Now sometimes the particular slice you're working in went right through the middle of an atom, so you got very high electron density. Sometime it just barely touched the atom, right? So you see here that that one, that atom, this particular slice, was very near the nucleus, so we got high electron density. But that one, the nucleus was not in this plane, so you just got a little glancing blow at it. Okay? But you could draw these things on acetate or plastic sheets and stack them up, right? So this was the model made for the first determination of the structure of penicillin; in fact, the potassium salt of penicillin. Where's the potassium atom? How can you distinguish potassium from nitrogen or carbon or hydrogen or oxygen?



Student: Electron density.



Professor Michael McBride: It'll have a lot more electron density. So do you see it? There's the potassium, lots of density. Okay, now here's a molecule that happened to be flat. So it was possible to slice through practically all the nuclei. It's not a very common type of molecule, but this one is, rubofusarin. Okay, now these contours are drawn at intervals of one electron per cubic angstrom. Right? So the first one is one electron per cubic angstrom, two electrons per cubic angstrom, three electrons per cubic angstrom and so on. Now there are both carbons and oxygens in this. How can you tell which ones are oxygens?



[Students speak over one another]



Professor Michael McBride: Yeah, Yoounjoou?



Student: They are more electronegative, so they should have more contour lines.



Professor Michael McBride: They should have more contour lines. So if we look here, we see that that one goes up to five electrons per cubic angstrom; one, two, three, four, five rings, right? That one goes up to seven electrons per cubic angstrom, and so do those. Right? So those are the oxygens. Okay? And there's a picture of the structure and you see indeed those atoms are oxygens. Now, what is our fundamental goal of looking at this stuff, what do we want to see?



Students: Bonds.



Professor Michael McBride: Bonds. Do you see bonds? What does it look like? It looks like just a bunch of balls, just atoms put there. Was Lewis right? Are there electron pairs halfway between the bonded atoms? No, it's just atoms. Okay. Why are there no hydrogens?



[Students speak over one another]



Professor Michael McBride: Because they have hardly any electrons compared to the others, right? So they're a very weak thing. Okay, now those are bonds. You can see them there. Where did those lines come from?



[Students speak over one another?]



Professor Michael McBride: Angela, what did you say?



Student: Someone drew them.



Professor Michael McBride: Yes, you drew them. How did you know where to draw the bonds?



Student: You connect the dots.



Professor Michael McBride: You connect the centers of the rings, right? Those are the bonds. Now how do you know what's a single bond, what's a double bond, what's a triple bond?



Student: More density?



Professor Michael McBride: You notice in the structure here, on the bottom right, there are double bonds and single bonds. How do I know which is which? Is there any way of knowing? Elizabeth?



Student: Maybe the distance between.



Professor Michael McBride: Ah, the distance! Double bonds should be shorter than single bonds. And, in fact, notice that here we have short, long and intermediate. Why intermediate?



[Students speak over one another]



Professor Michael McBride: Why do I draw a dotted line there?



Student: [inaudible]



Professor Michael McBride: Because we have this, look here; resonance, right? So there are one- and- a- half bonds. Right? Okay, but fundamentally what we see are spherical atoms, right? No bonds, no unshared pairs on oxygen, just atoms. So was Lewis way off base? Actually you can see the bonds with X-ray, but you don't look at total electron density the way we're looking at it. What you look at is difference density; you see how different is the picture I actually determine from what I would've expected if they were truly just spherical atoms, undistorted. How do you find--how would you make such a map? So you have these pieces of paper that say how much the electron density would be at all different points; experimentally, how much it is experimentally. Now I want to see how different is that from if I just had spheres for the atoms. What would I do? I have to find out what the spheres would look like, what density they would predict at all these different points, and then do what?



Student: Subtract.



Professor Michael McBride: Subtract, right? So that's the difference density. Sometimes it's called deformation because it's how much the electron distribution gets deformed when bonds get made. Okay, so you take the observed electron density, which is from X-ray, which is experimental, and you subtract atomic electron density, which you calculate somehow. You know what spherical atoms would look like; we'll get to that next, how you would know what that is. But anyhow, how do you know where to--when you're going to do the subtraction you have to make sure you position the atoms properly, to be subtracted, right? How do you know where to put them? How do you know where to put your spherical atoms so that you subtract it? Yes?



Student: On the center of the rings.



Professor Michael McBride: You put it on the center of those rings, right? Okay, so let's look at a case like this. This was done by Leiserowitz and Ziva Berkovitch-Yellin thirty-three years ago. So this is an interesting molecule because it has lots of different kinds of bonds in it: single bonds, double bonds--in fact three double bonds in a row--and also benzene rings that have these resonance bonds. Okay, so we're going to look first at one of the benzene rings. Okay? So this map that I'm showing you here is a difference density map. It's the total, minus what it would be for the atoms. But notice what was subtracted is just the carbon density; didn't subtract hydrogens. So let's look at the features that are big here. This is how the electron density changed in forming the molecule from atoms. Okay?



So first look up there. There are some very dense contours, right in the middle up there, and that's the hydrogen which was not subtracted, right? So it's the biggest game left in town, once you've subtracted the spherical carbons. Okay, so that's--and the total amount of intensity in that region is one electron, which is what you expect. Now the total amount of deformation density in this region, where there's a single carbon-carbon bond, is only 1/10th of an electron. What would Lewis have said? Two electrons, it's 1/10th of an electron. Okay, now let's slice that--and within the carbon-carbon bonds of the ring, the aromatic ones, it's bigger. Why bigger? Because it's partially double bond in the resonance structures, right? Okay, so it's 0.2 electrons for those ones in the ring. Now let's slice it and turn it so we can see its cross-section. And it's round; no surprise there. But let's slice the ones in the ring and turn them. It's oval, not round, the bonds are not round. Why?



Students: π bonds.



Professor Michael McBride: Ah, because they're partially made up of these p orbitals that point in and out of the ring, that distort it from being round. Okay, now let's look at the double bonds that are in the middle of this molecule, okay? First that one; we've already seen that bond before that was 0.1, had a round cross-section. But let's look at one of the double bonds. Not too surprisingly it has more electron deformation density in it; 0.2, right? And the one in the middle actually has 0.3. Now they're both double bonds, so probably they look similar, and in a sense they do. If we slice the first one, and turn it, it's oval for the same reason as before; there's a p orbital pointing in and out of the board, and if we look at it below, part of the bond is caused by overlapping, as we'll see, of those p orbitals. So it's got an oval, up and down shape. How about the next one? If we slice it and turn it, it's also oval in cross-section, but perpendicular. Right? What do you expect the next one would look like?



[Students speak over one another]



Professor Michael McBride: It's symmetrical; it'll look like the first one again. But here's something interesting. That cross-section of the middle one showed the deformation due to the pink but not the deformation showing--due to those neighboring blue ones. There must be something different about the interaction at the end and the interaction in the middle. Does everybody see that problem? You might think the middle one would be round or clover-leaf shaped, but it's not, it's distorted in the in-plane direction. But to understand this, you have to be patient and wait till next week, or later this week, when we start getting into quantum mechanics. But notice that it's as if the carbons formed bent bonds. So the bonds from those central carbons at the end go up and down, but the ones in the middle are in- plane. Now how many electrons are there in a bond? So here we're going to plot how many electrons vertically, and how long the bond is, which is related to whether it's single, double, triple. In fact, a single bond is there; a double bond and a triple bond have those distances. Now if you--and in fact you could also do a one-and-a- and a half bond, those bonds in the benzene ring. Now what would you expect this to be? How many electrons--if you were Lewis, how many electrons would you have in a single bond?



Students: Two.



Professor Michael McBride: Two. Double bond?



Students: Four



Professor Michael McBride: Four.



Students: Six.



Professor Michael McBride: Six.



Students: Three.



Professor Michael McBride: Three, okay one and a -and-a-half. Okay, so there's what you would expect for the plot, if Lewis was right:; two, four, six. Okay? Now we can take this molecule that we just looked at and see how much electron density there is in each of those bonds--and I've color-coded it--and it looks pretty darn good. Right? Hooray for Lewis. And in fact the people who did the work, Leiserowitz and Berkovitch-Yellin, looked at a bunch of other structures too and added those to the same plot, and they're not bad, except the scale is wrong. Instead of being two, four, six,, which I've faded out here, it's 0.1, 0.2, 0.3. So bonds are a very minor game in the distribution of electrons. Lewis, in a sense, was sort of right, there is electron density shifted between the atoms to hold them together, right? But much less than pairs of electrons. In fact, this amount of bonding density is about 1/20th of what Lewis might have said, but never did say, because he wasn't that naïve. Okay? So instead of getting two, you get a 20th of two; 0.1. Okay? So bonding density is about 1/20th of a Lewis. Incidentally, the unit "Lewis" is used only within this class. Okay?



So here's another example done by Dunitz in Switzerland and these people who were the--Schuyler Seiler and Schweizer, who were Swiss, who were the progeny of watchmakers and people that do very precise work; because to do this you have to have really precise experimental work, if you're going to subtract something from something else and have the difference be meaningful, when the numbers are almost precisely the same, right? Remember we looked at the total electron density. It looked just like spheres. So you have to be really good with this if you want to have this be meaningful; and those guys were really good. This was done, what, twenty-five years ago. So they took this molecule, which is highly symmetric, which is great. Remember it's got its resonance structure, so all the bonds around the middle are the same, top and bottom. Okay, so we need only look at a quarter of the molecule, or for convenience a little bit more than a quarter. So I'll blow it up here. And there is the electron difference density. So this is how the electrons shifted during formation of the bonds. Okay, so there's a carbon-carbon aromatic bond, a bond and a half, and if you turn it and look at it, it's distorted the way you expect, just as in the previous experiment we talked about. Here's another one, same deal. Now here is a single bond, carbon-carbon single bond. What's it going to look like in cross-section?



[Students speak over one another]



Professor Michael McBride: Round. Good. Now here is a carbon-nitrogen triple bond. This is something new. What's it going to look like in cross-section? Maybe like a clover-leaf, huh? Like that, it looks like that. It's round. It's not clover-leaf, and it's not a diamond, it's round in cross-section. So we'll address that later. What's that? That's not a bond.



Students: Unshared pair.



Professor Michael McBride: That's the unshared pair on Nitrogen. Now what's special here? What are we looking at? We're looking at the carbon-fluorine bond. Does it look like the carbon-carbon bonds, or carbon-nitrogen? How's it different? There's no there, there. Right? There aren't any electrons holding carbon to fluorine. What the heck is going on? Where is the C-F bond? This brings up the second great question in the course. What's the first question?



Students: How do you know?



Professor Michael McBride: How do you know, right? And the second question is illustrated by this film clip.



[Technical adjustments]



Come on baby. I worked hard on this.



Okay, I'll tell you what it says.



[Laughter]



[Technical Adjustments]



Professor Michael McBride: Okay, I can do it myself. (This is a disaster.)



"Catch some thieves Hobson, that's what we're here for."



"Very well sir."



Click, click, click, click, click, click, click, click, click, click. [Laughter]



"What do you think of him?"



"Compared with what sir?"



"Exactly."



Okay. [Laughter and applause] "What do you think of him?" "Compared with what sir?" "Exactly." That's the second question, compared with what? In anything quantitative, even many qualitative things, that's the question, and that's why we're in problems here looking for the C-F bond. We already saw this question when we talked about resonance. Remember, I asserted that when you have resonance the molecule is unusually stable. What should you have asked?



Student: Compared to what?



Professor Michael McBride: Compared to what? Right? In that case it turns out it's compared to observed or calculated energy. Compare the observed or calculated energy, (what really happens), to what you would expect if it were not resonant. And we'll talk later about how, on what basis, you might have an expectation. Now we're talking about difference electron density. So what do we compare?



[Students speak over one another]



Professor Michael McBride: We compare experimental electron density, (or it could be calculated, if we can use quantum mechanics to calculate it)--, but compare experimental electron density to what? What have we compared it to?



Student: [inaudible]



Professor Michael McBride: What it would be if it were just atoms. Right? But what kind of atoms, right? We compare the observed or calculated total electron density to the sum of electron densities for a set of undistorted atoms. Now what's an undistorted atom? Okay. Now, to avoid Pauli problems--and this isn't the time to discuss what those are, we need to subtract not an unbiased spherical fluorine atom, which if you count up how many electrons there are on fluorine, each quadrant of space around fluorine--of which there will be four-- each quadrant will have one and three three-quarters electrons in it. Right? You multiply that by four and you get seven electrons, which is the valence electrons of fluorine, right? But you can't have one and three three-quarters electrons from fluorine, and also an electron from carbon in the same place. That's too many electrons in the same place, right? So what you have to start with is not an undistorted spherical fluorine atom, but a fluorine atom that's prepared to make a bond, which has only one electron in the region where it's going to make a bond. So if you subtract a spherical one, you're subtracting too much from the region of the bond, right? So that's why you don't see any electron density in there. So, and if you want to understand this better than I've explained it, go on the Web and look at Professor Dunitz explaining it. We have a little movie of him talking about this particular problem. Okay? But at any rate the conclusion of all this is that bonding density is about 1/20th of a Lewis. All right? And it's time for me to quit. We'll go on with this next time.



[end of transcript]

Lecture 7
Quantum Mechanical Kinetic Energy
Play Video
Quantum Mechanical Kinetic Energy


After pointing out several discrepancies between electron difference density results and Lewis bonding theory, the course proceeds to quantum mechanics in search of a fundamental understanding of chemical bonding. The wave function ψ, which beginning students find confusing, was equally confusing to the physicists who created quantum mechanics. The Schrödinger equation reckons kinetic energy through the shape of ψ. When ψ curves toward zero, kinetic energy is positive; but when it curves away, kinetic energy is negative!




Transcript



September 17, 2008



Professor Michael McBride: Okay, so we just glimpsed this at the end last time. This is a crystal structure of a complicated molecule that was performed by these same Swiss folks that we've talked about, and notice how very precise it is. The bond distances between atoms are reported to ±1/2000th of an angstrom. The bonds are like one and a half angstroms. So it's like a part in a thousand, is the precision to which the positions of the atoms is known. Okay? But those positions are average positions, because the atoms are constantly in motion, vibrating. In fact, the typical vibration amplitude, which depends on -- an atom that's off on the end of something is more floppy than one that's held in by a lot of bonds in various directions, in a cage sort of thing. But typically they're vibrating by about 0.05 angstroms, which is twenty-five times as big as the precision to which the position of the average is known. Okay? So no molecule looks like that at an instant, the atoms are all displaced a little bit.



Now how big is that? Here, if you look at that yellow thing, when it shrinks down, that's how big it is, that's how big the vibration is. It's very small. But these are very precise measurements. Right? Now why did they do so precise measurements? Did they really care to know bond distances to that accuracy? Maybe for some purposes they did, but that wasn't the main reason they did the work very carefully. They did it carefully in order to get really precise positions for the average atom, so they could subtract spherical atoms and see the difference accurately. Okay? Because if you have the wrong position for the atom that you're subtracting, you get nonsense. Okay, and what this is going to reveal is some pathologies of bonding, from the point of view of Lewis concept of shared electrons. Okay, so here's a picture of this molecule. And remember, we had -- rubofusarin, which we looked at last time, had the great virtue that it was planar. So you could cut a slice that went through all the atoms. This molecule's definitely not planar, so you have to cut various slices to see different things. So first we'll cut a slice that goes through those ten atoms. Okay? And here is the difference electron density. What does the difference density show? Somebody? Yes Alex?



Student: It's the electron density minus the spherical --



Professor Michael McBride: It's the total electron density minus the atoms; that is, how the electron density shifted when the molecule was formed from the atoms. Okay, and here we see just exactly what we expected to see, that the electrons shifted in between the carbon atoms, the benzene ring, and other pairs of carbon atoms as well. It also shows the C-H bonds, because in this case the hydrogen atoms were subtracted. We showed one last time where the hydrogen atoms weren't subtracted. Okay, so this is -- there's nothing special here, everything looks the way you expect it to be; although it's really beautiful, as you would expect from these guys who do such great work. Now we'll do a different slice. This is sort of the plane of the screen which divides the molecule symmetrically down the middle, cuts through some bonds, cuts through some atoms and so on. So here's the difference density map that appears on that slice.



Now the colored atoms, on the right, are the positions of the atoms through which the plane slices, but the atoms are subtracted out; so what you see is the bonding in that plane. So you see those bonds, because both ends of the bond are in the plane, so the bonds are in the plane, and you see just what you expect to see. But there are other things you see as well. You see the C-H bonds, although they don't have nearly as much electron density as the C-C bonds did. Right? You also see that lump, which is the unshared pair on nitrogen. Right? But you see these two things, which are bonds, but they're cross-sections of bonds, because this particular plane cuts through the middle of those bonds. Everybody see that? Okay, so again that's nothing surprising. But here is something surprising. There's another bond through which that plane cuts, which is the one on the right, through those three-membered rings, right? And what do you notice about that bond?



Student: It isn't there.



Professor Michael McBride: It isn't there. There isn't any electron density for that bond. So it's a missing bond. This is what we'll refer to as pathological bonding, right? It's not what Lewis would have expected, maybe; we don't have Lewis to talk to, so we don't know what he would've thought about this particular molecule. So here's a third plane to slice, which goes through those three atoms, and here's the picture of it. And again, that bond is missing, that we saw before. Previously we looked at a cross-section. Here we're looking at a plane that contains the bond, and again there's no there there. Okay? But there's something else that's funny about this slice. Do you see what? What's funny about the bonds that you do see? Corey? Speak up so I can hear you.



Student: They're connected; they're not totally separate.



Professor Michael McBride: What do you mean they're connected separately?



Student: Usually you see separate electron densities, but they're connected.



Professor Michael McBride: Somebody say it in different words. I think you got the idea but I'm not sure everybody understood it. John, do you have an idea?



Student: The top one seems to be more dense than the bottom one.



Professor Michael McBride: One, two, three, four; one, two, three four five. That's true, it is a little more dense. That kind of thing could be experimental error, because even though this was done so precisely, you're subtracting two numbers that are very large, so that any error you make in the experimental one, or in positioning things for the theoretical position of the atoms, any error you make will really be amplified in a map like this. But it's true, you noticed. But there's something I think more interesting about those -- that. Yes John?



Student: The contour lines, they're connected, the contour lines between the top and the bottom bonds are connected. So maybe the electrons, maybe -- I don't know if they're connected.



Professor Michael McBride: Yes, they sort of overlap one another. But of course if they're sort of close to one another, that doesn't surprise you too much because as you go out and out and out, ultimately you'll get rings that do meet, if you go far enough down. Yes Chris?



Student: The center of density on the bonds doesn't intercept the lines connecting the atoms.



Professor Michael McBride: Ah. The bonds are not centered on the line that connects the nuclei. These bonds are bent. Okay? So again, pathological bonding; and in three weeks you'll understand this, from first principles, but you've got to be patient. Okay, so Lewis pairs and octets provide a pretty good bookkeeping device for keeping track of valence, but they're hopelessly crude when it comes to describing the actual electron distribution, which you can see experimentally here. There is electron sharing. There's a distortion of the spheres of electron density that are the atoms, but it's only about 5% as big as Lewis would've predicted, had he predicted that two electrons would be right between. Okay? And there are unshared pairs, as Lewis predicted. And again they're less -- but in this case they're even less than 5% of what Lewis would've predicted. But you can see them.



Now this raises the question, is there a better bond theory than Lewis theory, maybe even one that's quantitative, that would give you numbers for these things, rather than just say there are pairs here and there. Right? And the answer, thank goodness, is yes, there is a great theory for this, and what it is, is chemical quantum mechanics. Now you can study quantum mechanics in this department, you can study quantum mechanics in physics, you can probably study it other places, right? And different people use the same quantum mechanics but apply it to different problems. Right? So what we're going to discuss in this course is quantum mechanics as applied to bonding. So it'll be somewhat different in its flavor -- in fact, a lot different in its flavor -- from what you do in physics, or even what you do in physical chemistry, because we're more interested -- we're not so interested in getting numbers or solving mathematical problems, we're interested in getting insight to what's really involved in forming bonds. We want it to be rigorous but we don't need it to be numerical. Okay? So it'll be much more pictorial than numerical.



Okay, so it came with the Schrödinger wave equation that was discovered in, or invented perhaps we should say, in -- I don't know whether -- it's hard to know whether to say discovered or invented; I think invented is probably better -- in 1926. And here is Schrödinger the previous year, the sort of ninety-seven-pound weakling on the beach. Right? He's this guy back here with the glasses on. Okay? He was actually a well-known physicist but he hadn't done anything really earthshaking at all. He was at the University of Zurich. And Felix Bloch, who was a student then -- two years before he had come as an undergraduate to the University of Zurich to study engineering, and after a year and a half he decided he would do physics, which was completely impractical and not to the taste of his parents. But anyhow, as an undergraduate he went to these colloquia that the Physics Department had, and he wrote fifty years later -- see this was 1976, so it's the 50th anniversary of the discovery of, or invention, of quantum mechanics. So he said:



"At the end of a colloquium I heard Debye"(there's a picture of Debye)"say something like, 'Schrödinger, you're not working right now on very important problems anyway. Why don't you tell us something about that thesis of de Broglie?' So in one of the next colloquia, Schrödinger gave a beautifully clear account of how de Broglie associated a wave with a particle. When he had finished, Debye casually remarked that he thought this way of talking was rather childish. He had learned that, to deal properly with waves, one had to have a wave equation. It sounded rather trivial and did not seem to make a great impression, but Schrödinger evidently thought a bit more about the idea afterwards, and just a few weeks later he gave another talk in the colloquium, which he started by saying, 'My colleague Debye suggested that one should have a wave equation. Well I have found one.'"



And we write it now: HΨ=EΨ. He actually wrote it in different terms, but that's his way, the Schrödinger equation. The reason the one we write is a little different from his is he included time as a variable in his, whereas we're not interested in, for this purpose, in changes in time. We want to see how molecules are when they're just sitting there. We'll talk about time later. So within, what? Seven years, here's Schrödinger looking a good deal sharper. Right? And where is he standing? He's standing at the tram stop in Stockholm, where he's going to pick up his Nobel Prize for this. Right? And he's standing with Dirac, with whom he shared the Nobel Prize, and with Heisenberg, who got the Nobel Prize the previous year but hadn't collected it yet, so he came at the same time. Okay? So the Schrödinger equation is HΨ=EΨ. H and E you've seen, but Ψ may be new to you. It's a Greek letter. We can call it Sigh or P-sighi, some people call it Psee. Right? I'll probably call it Psi. Okay. And it's a wave function. Well what in the world is a wave function? Okay? So this is a stumbling block for people that come into the field, and it's not just a stumbling block for you, it was a stumbling block for the greatest minds there were at the time.



So, for example, this is five years later in Leipzig, and it's the research group of Werner Heisenberg, who's sitting there in the front, the guy that -- this was about the time he was being nominated or selected for the Nobel Prize. Right? So he's there with his research group, and right behind him is seated Felix Bloch, who himself got the Nobel Prize for discovering NMR in 1952. So he's quite a young guy here, and he's with these other -- there's a guy who became famous at Oxford and another one who became the head of the Physics Department at MIT. Bloch was at Stanford. So these guys know they're pretty hot stuff, so they're looking right into the camera, to record themselves for posterity, as part of this distinguished group; except for Bloch. What's he thinking about? [Laughter] What in the world is Ψ? Right? Now, in fact, in that same year, it was in January that Schrödinger announced the wave equation, and Ψ. Right?



And that summer these smart guys, who were hanging around Zurich at that time, theoretical physicists, the young guys went out on an excursion, on the lake of Zurich, and they made up doggerel rhymes for fun about different things that were going on, and the one that was made up by Bloch and Erich Hückel, whom we'll talk about next semester, was about Ψ. "Gar Manches rechnet Erwin schon, Mit seiner Wellenfuction. Nur wissen möcht man gerne wohl, Was man sich dabei vorstell'n soll." Which means: "Erwin with his Psi can do calculations, quite a few. We only wish that we could glean an inkling of what Psi could mean." Right? You can do calculations with it, but what is it? -- was the question. Okay? And it wasn't just these young guys who were confused. Even Schrödinger was never comfortable with what Ψ really means. Now if we're lucky, this'll play this time. So this is a lecture by Schrödinger, "What is Matter", from 1952.



[Short film clip is played]



Schrödinger's voice: "etwa so wie Cervantes einmal den Sancho Panza, sein liebes Eselchen auf dem er zu reiten pflegte, verlieren läßt. Aber ein paar Kapitel später hat der Autor das vergessen und das gute Tier ist wieder da. Nun werden sie mich vielleicht zuletzt fragen, ja was sind denn nun aber wirklich diese Korpuskeln, diese Atome - Moleküle. Ehrlich müßte ich darauf bekennen, ich weiß es sowenig, als ich weiß, wo Sancho Panzas zweites Eselchen hergekommen ist."



Professor Michael McBride: Right? So twenty-six years later Schrödinger still didn't really know what Ψ was. Okay? So don't be depressed when it seems a little curious what Ψ might be. Okay? First we'll -- like Schrödinger and like these other guys -- first we'll learn how to find Ψ and use it, and then later we'll learn what it means. Okay? So Ψ is a function, a wave function. What do you want to know, from what I've shown here? What is a function?



Student: A relationship.



Professor Michael McBride: Like sine is a function; what does that mean? Yes? I can't hear very well.



Student: You put an input into a function and you get an output.



Professor Michael McBride: Yes, it's like a little machine. You put a number in, or maybe several numbers, and a number comes out. Right? That's what the function does; okay, you put in ninety degrees and sin says one. Okay? So what do you want to know about Ψ first?



Student: What does it do?



Professor Michael McBride: What's it a function of? What are the things you have to put in, in order to get a number out? Okay? So it's different from the name. The wave functions have names. That's not what they're a function of. Right? You can have sine, sine2, cosine. Those are different functions, but they can be functions of the same thing, an angle. Right? So we're interested in what's it a function of; not what the function is but what's it a function of? So you can have different Ψs. They have names and quantum numbers give them their names. For example, you can have n, l, and m. You've seen those before, n, l, m, to name wave functions. Those are just their names. It's not what they're a function of. Or you can have 1s or 3dxy, or σ, or π, or π*. Those are all names of functions. Right? But they're not what it's a function of. What it's a function of is the position of a particle, or a set of particles. It's a function of position, and it's also a function of time, and sometimes of spin; some particles have spin and it could be a function of that too. But you'll be happy to know that for purposes of this course we're not so interested in time and spin. So for our purposes it's just a function of position. So if you have N particles, how many positions do you have to specify to know where they all are? How many numbers do you need? You need x, y, z for every particle. Right? So you need 3N arguments for Ψ. So Ψ is a function that when you tell it where all these positions are, it gives you a number. Now curiously enough, the number can be positive, it can be zero, it can be negative, it can even be complex, right, although we won't talk about cases where it's complex. The physicists will tell you about those, or physical chemists. Okay? And sometimes it can be as many as 4N+1 arguments. How could it be 4N+1?



Student: [inaudible]



Professor Michael McBride: Because if each particle also had a spin, then it would be x, y, z and spin; that'd be four. And if time is also included, it's plus one. Okay, so how are we going to go through this? First we'll try -- this is unfamiliar territory, I bet, to every one of you. Okay? So first we're going to talk about just one particle and one dimension, so the function is fairly simple. Okay? And then we'll go on to three dimensions, but still one particle, the electron in an atom; so a one-electron atom, but now three dimensions, so it's more complicated. Then we'll go on to atoms that have several electrons. So you have now more than three variables, because you have at least two electrons; that would be six variables that you have to put into the function to get a number out. Then we'll go into molecules -- that is, more than one atom -- and what bonding is. And then finally we get to the payoff for organic chemistry, which is talking about what makes a group a functional group and what does it mean to be functional, what makes it reactive? That's where we're heading. But first we have to understand what quantum mechanics is.



So here's the Schrödinger equation, ΗΨ=EΨ, and we're talking about the time-independent Schrödinger equation, so time is not a variable, and that means what we're talking about is stationary states. We don't mean that the atoms aren't moving, but just that they're in a cloud and we're going to find how is the cloud distributed. If a molecule reacts, the electrons shift their clouds and so on; it changes. We're not interested in reaction now, we're just interested in understanding the cloud that's sitting there, not changing in time. Okay, now the right part of the equation is E times Ψ, right? And E will turn out to be the energy of the system; maybe you won't be surprised at that. So that's quite simple. What's the left? It looks like H times Ψ. If that were true, what could you do to simplify things? Knock out Ψ. But HΨ is not H times Ψ. H is sort of recipe for doing something with Ψ; we'll see that right away. So you can't just cancel out the Ψ, unfortunately. Okay, so ΗΨ=EΨ. Oops sorry, what did I do, there we go. Now we can divide, you can divide the right by Ψ, and since it was E times Ψ, the Ψ goes away. But when you divide on the left, you don't cancel the Ψs, because the top doesn't mean multiplication.



Now I already told you the right side of this equation is the total energy. So when you see a system, what does the total energy consist of? Potential energy and kinetic energy. So somehow this part on the left, ΗΨ/Ψ, must be kinetic energy plus potential energy. That recipe, H, must somehow tell you how to work with Ψ in order to get something which, divided by Ψ, gives kinetic energy plus potential energy. So there are two parts of it. There's the part that's potential energy, of the recipe, and there's the part that's kinetic energy. Now, the potential energy part is in fact easy because it's given to you. Right? What's Ψ a function of?



Students: Position.



Professor Michael McBride: Position of the particles. Now if you know the charges of the particles, and their positions, and know Couloumb's Law, then you know the potential energy, if Couloumb's Law is right. Is everybody with me on that? If you know there's a unit positive charge here, a unit negative charge here, another unit positive charge here and a unit negative charge here, or something like that, you -- it might be complicated, you might have to write an Excel program or something to do it -- but you could calculate the distances and the charges and so on, and what the energy is, due to that. So that part is really given to you, once you know what system you're dealing with, the recipe for finding the potential energy. So that part of HΨ/Ψ is no problem at all. But hold your breath on kinetic energy. Sam?



Student: Didn't we just throw away an equation? There was an adjusted Couloumb's Law equation.



Professor Michael McBride: Yes, that was wrong. That was three years earlier, remember? 1923 Thomson proposed that. But it was wrong. This is what was right.



Student: How did they prove it wrong?



Professor Michael McBride: How did what?



Student: Did they prove it wrong or just --



Professor Michael McBride: Yes, they proved this right, that Couloumb's Law held, because it agreed with a whole lot of spectroscopic evidence that had been collected about atomic spectra, and then everything else that's tried with it works too. So we believe it now. So how do you handle kinetic energy? Well that's an old one, you did that already in high school, right? Forget kinetic energy, here it is. It's some constant, which will get the units right, depending on what energy units you want, times the sum over all the particles of the kinetic energy of each particle. So if you know the kinetic energy of this particle, kinetic energy of this particle, this particle, this particle; you add them all up and you get the total kinetic energy. No problem there, right? Now what is the kinetic energy that you're summing up over each particle? It's ½mv2. Has everybody seen that before? Okay, so that's the sum of classical kinetic energy over all the particles of interest in the problem, and the constant is just some number you put in to get the right units for your energy, depending on whether you use feet per second or meters per year or whatever, for the velocity.



Okay, but it turned out that although this was fine for our great-grandparents, it's not right when you start dealing with tiny things. Right? Here's what kinetic energy really is. It's a constant. This is the thing that gets it in the right units: (h2)/8(π2) times a sum over all the particles -- it's looking promising, right? -- of one over the mass -- not the mass mv2, but one over the mass of each particle -- and here's where we get it -- [Students react] -- times second derivatives of a wave function. That's weird. I mean, at least it has twos in it, like v2, right? [Laughter] That's something. And in fact it's not completely coincidental that it has twos in it. There was an analogy that was being followed that allowed them to formulate this. And you divide it by the number Ψ. So that's a pretty complicated thing. So if we want to get our heads around it, we'd better simplify it. And oh also there's a minus sign; it's minus, the constant is negative that you use. Okay, now let's simplify it by using just one particle, so we don't have to sum over a bunch of particles, and we'll use just one dimension, x; forget y and z. Okay? So now we see something simpler. So it's a negative constant times one over the mass of the particle, times the second derivative of the function, the wave function, divided by Ψ. That's kinetic energy really, not ½mv2. Okay? Or here it is, written just a little differently. So there's a constant, C, over the mass, right? And then we have the important part, is the second derivative. Does everybody know that the second derivative is a curvature of a function. Right? What's the first derivative?



Students: Slope.



Professor Michael McBride: Slope, and the second derivative is how curved it is. It can be curving down, that's negative curvature; or curving up, that's positive curvature. So it can be positive or negative; it can be zero if the line is straight. Okay. So note that the kinetic energy involves the shape of Ψ, how curved it is, not just what the value of Ψ is; although it involves that too. Maybe it's not too early to point out something interesting about this. So suppose that's the kinetic energy. What would happen if you multiplied Ψ by two? Obviously the denominator would get twice as large, if you made Ψ twice as large. What would happen to the curvature? What happens to the slope? Suppose you have a function and you make it twice as big and look at the slope at a particular point? How does the slope change if you've stretched the paper on which you drew it?



Student: It's sharper.



Professor Michael McBride: The slope will double, right, if you double the scale. How about the curvature, the second derivative? Does it go up by four times? No it doesn't go up by four times, it goes up by twice. So what would happen to the kinetic energy there if we doubled the size of Ψ every place?



Student: Stay the same.



Professor Michael McBride: It would stay the same. The kinetic energy doesn't depend on how you scale Ψ, it only depends on its shape, how curved it is. Everybody see the idea? Curvature divided by the value. Okay, now solving a quantum problem. So if you're in a course and you're studying quantum mechanics, you get problems to solve. A problem means you're given something, you have to find something. You're given a set of particles, like a certain nuclei of given mass and charge, and a certain number of electrons; that's what you're given. Okay, the masses of the particles and the potential law. When you're given the charge, and you know Couloumb's Law, then you know how to calculate the potential energy; remember that's part of it. Okay? So that part's easy, okay? Now what do you need to find if you have a problem to solve? Oh, for example, you can have one particle in one dimension; so it could be one atomic mass unit is the weight of the particle, and Hooke's Law could be the potential, right? It doesn't have to be realistic, it could be Hooke's Law, it could be a particle held by a spring, to find a Ψ. You want to find the shape of this function, which is a function of what?



Student: Positions of the particles.



Professor Michael McBride: Positions of the particles, and if you're higher, further on than we are, time as well; maybe spin even. But that function has to be such that HΨ/Ψ is the total energy, and the total energy is the same, no matter where the particle is, right, because the potential energy and the kinetic energy cancel out. It's like a ball rolling back and forth. The total energy is constant but it goes back and forth between potential and kinetic energy, right? Same thing here. No matter where the particles are, you have to get the same energy. So Ψ has to be such that when you calculate the kinetic energy for it, changes in that kinetic energy, in different positions, exactly compensate for the changes in potential energy. When you've got that, then you've got a correct Ψ; maybe. It's also important that Ψ remain finite, that it not go to infinity. And if you're a real mathematician, it has to be single valued; you can't have two values for the same positions. It has to be continuous, you can't get a sudden break in Ψ. And Ψ2 has to be integrable; you have to be able to tell how much area is under Ψ2, and you'll see why shortly. But basically what you need is to find a Ψ such that the changes in kinetic energy compensate changes in potential energy.



Now what's coming? Let's just rehearse what we did before. So first there'll be one particle in one dimension; then it'll be one-electron atoms, so one particle in three dimensions; then it will be many electrons and the idea of what orbitals are; and then it'll be molecules and bonds; and finally functional groups and reactivity. Okay, but you'll be happy to hear that by a week from Friday we'll only get through one-electron atoms. So don't worry about the rest of the stuff now. But do read the parts on the webpage that have to do with what's going to be on the exam. Okay, so normally you're given a problem, the mass and the charges -- that is, that potential energy as a function of position -- and you need to find Ψ. But at first we're going to try it a different way. We're going to play Jeopardy and we're going to start with the answer and find out what the question was. Okay? So suppose that Ψ is the sine of x; this is one particle in one dimension, the position of the particle, and the function of Ψ is sine. If you know Ψ, what can you figure out? We've just been talking about it. What can you use Ψ to find?



Student: Kinetic energy.



Professor Michael McBride: Kinetic energy. How do you find it? So we can get the kinetic energy, which is minus a constant over the mass times the curvature of Ψ divided by Ψ at any given position. Right? And once we know how the kinetic energy varies with position, then we know how the potential energy varies with position, because it's just the opposite, in order that the sum be constant. Right? So once we know the kinetic energy, then we know what the potential energy was, which was what the problem was at the beginning. Okay? So suppose our answer is sin(x). What is the curvature of sin(x), the second derivative?



[Students speak over one another]



Professor Michael McBride: It's -sin(x). Okay, so what is the kinetic energy?



Student: C/m.



Professor Michael McBride: C/m. Does it depend on where you are, on the value of x? No, it's always C/m. So what was the potential energy? How did the potential energy vary with position?



Student: It doesn't.



Professor Michael McBride: The potential energy doesn't vary with position. So sin(x) is a solution for what? A particle that's not being influenced by anything else; so its potential energy doesn't change with the position, it's a particle in free space. Okay? So the potential energy is independent of x. Constant potential energy, it's a particle in free space. Now, suppose we take a different one, sin(ax). How does sin(ax) look different from sin(x)? Suppose it's sin(2x). Here's sin(x). How does sin(2x) look? Right? It's shorter wavelength. Okay? Now so we need to figure out -- so it's a shortened wave, if a > 1. Okay, now what's the curvature? Russell?



Student: It's -a2 times sin(ax).



Professor Michael McBride: It's -a2 times sin(ax). Right? The a comes out, that constant, each time you take a derivative. So now what does the kinetic energy look like? It's a2 times the same thing. Okay? So again, the potential energy is constant. Right? It doesn't change with position. But what is different? It has higher kinetic energy if it's a shorter wavelength. And notice that the kinetic energy is proportional to one over the wavelength squared, right?; a2; a shortens the wave, it's proportional to a2, one over the wavelength squared. Okay. Now let's take another function, exponential, so ex. What's the second derivative of ex? Pardon me?



Student: ex.



Professor Michael McBride: ex. What's the 18th derivative of ex?



Students: ex.



Professor Michael McBride: Okay, good. So it's ex. So what's this situation, what's the kinetic energy?



Student: -C/m.



Professor Michael McBride: -C/m. Negative kinetic energy. Your great-grandparents didn't get that. You can have kinetic energy that's less than zero. What does that mean? It means the total energy is lower than the potential energy. Pause a minute just to let that sink in. The total energy is lower than the potential energy. The difference is negative. Okay? So the kinetic energy, if that's the difference, between potential and total, is negative. You never get that for ½m(v2). Yes?



Student: Does this violate that Ψ has to remain finite?



Professor Michael McBride: Does it violate what?



Student: That Ψ has to remain finite?



Professor Michael McBride: No. You'll see in a second. Okay, so anyhow the constant potential energy is greater than the total energy for that. Now, how about if it were minus exponential, e-x? Now what would it be? It would be the same deal again, it would still be -C/m, and again it would be a constant potential energy greater than the total energy. This is not just a mathematical curiosity, it actually happens for every atom in you, or in me. Every atom has the electrons spend some of their time in regions where they have negative kinetic energy. It's not just something weird that never happens. And it happens at large distance from the nuclei where 1/r -- that's Couloumb's Law -- where it stops changing very much. When you get far enough, 1/r gets really tiny and it's essentially zero, it doesn't change anymore. Right? Then you have this situation in any real atom. So let's look at getting the potential energy from the shape of Ψ via the kinetic energy. Okay, so here's a map of Ψ, or a plot of Ψ, it could be positive, negative, zero -- as a function of the one-dimension x, wherever the particle is. Okay? Now let's suppose that that is our wave function, sometimes positive, sometimes zero, sometimes negative. Okay? And let's look at different positions and see what the kinetic energy is, and then we'll be able to figure out, since the total will be constant, what the potential energy is. Okay? So we'll try to find out what was the potential energy that gave this as a solution? This is again the Jeopardy approach. Okay? Okay, so the curvature minus -- remember it's a negative constant -- minus the curvature over the amplitude could be positive -- that's going to be the kinetic energy; it could be positive, it could be zero, it could be negative, or it could be that we can't tell by looking at the graph. So let's look at different positions on the graph and see what it says. First look at that position. What is the kinetic energy there? Positive, negative, zero? Ryan, why don't you help me out?



Student: No



Professor Michael McBride: Well no, you can help me out. [Laughter] Look, so what do you need to know? You need to know -- here's the complicated thing you have to figure out. What is minus the curvature divided by the amplitude at this point? Is it positive, negative or zero? So what's the curvature at that point? Is it curving up or down at that point? No idea. Anybody got an idea? Keith? Kevin?



Student: It looks like a saddle point so it's probably zero.



Professor Michael McBride: It's not a saddle point. What do you call it?



Students: Inflection points.



Professor Michael McBride: A saddle point's for three dimensions. In this it's what? Inflection point. It's flat there. It's curving one way on one side, the other way on the other side. So it's got zero curvature there; okay, zero curvature. Now Ryan, can you tell me anything about that, if the curvature is zero?



Student: Zero.



Professor Michael McBride: Ah ha.



[Laughter]



Professor Michael McBride: Not bad. So that one we'll color grey for zero. The kinetic energy at that point is zero, if that's the wave function. Now let's take another point. Who's going to help me with this one? How about the curvature at this point right here?



Student: Negative.



[Students speak over one another]



Professor Michael McBride: It's actually -- I choose a point that's not curved.



Student: Ah.



Professor Michael McBride: It's straight right there. I assure you that's true. So I bet Ryan can help me again on that one. How about it?



Student: Zero.



Professor Michael McBride: Ah ha. So we'll make that one grey too. Now I'll go to someone else. How about there? What's the curvature at that point do you think? Shai?



Student: It looks straight, zero curvature.



Professor Michael McBride: It looks straight, zero curvature. So does that mean that this value is zero?



Student: Not necessarily, because the amplitude --



Professor Michael McBride: Ah, the amplitude is zero there too. So really you can't be sure. Right? So that one we're going to have to leave questionable, that's a question mark. How about out here? Not curved. So what's the kinetic energy? Josh?



Student: Questionable.



Professor Michael McBride: Questionable, right? Because the amplitude is zero again; zero in the numerator, also zero in the denominator; we really don't know. Okay, how about here? Tyler, what do you say? Is it curved there?



Student: Yes.



Professor Michael McBride: Curving up or down?



Student: Down.



Professor Michael McBride: So negative, the curvature is negative. The value of Ψ?



Student: Positive.



Professor Michael McBride: Positive. The energy, kinetic energy?



Student: Positive.



Professor Michael McBride: Positive. Okay, so we can make that one green. Okay, here's another one. Who's going to help me here? Kate?



Student: Yes.



Professor Michael McBride: Okay, so how about the curvature; curving up, curving down?



Student: It's curving down, that's negative.



Professor Michael McBride: Yes. Amplitude?



Student: Zero. So it should be green.



Professor Michael McBride: Ah, green again. Okay. How about here? Ah, now how about the curvature? Seth?



Student: I don't know.



Professor Michael McBride: Which way is it curving at this point here?



Student: Curving up.



Professor Michael McBride: Curving up. So the curvature is --



Student: Positive.



Professor Michael McBride: Positive.



Student: The amplitude is negative, so it's positive.



Professor Michael McBride: Yeah. So what color would we make it? Green again. Okay, so if you're -- you can have -- be curving down or curving up and still be positive; curving down if you're above the baseline, curving up if you're below the baseline. Right? So as long as you're curving toward the baseline, towards Ψ=0, the kinetic energy is positive. How about here? Zack? Which way is it curving? Curving up or curving down?



Student: It should be curving up.



Professor Michael McBride: Curving up, curvature is positive. The value?



Student: Positive.



Professor Michael McBride: Positive.



Student: I guess it'll be negative.



Professor Michael McBride: So it's negative kinetic energy there. Make that one whatever that pinkish color is. Okay? Here's another one, how about there? Alex? Which way is it curving at the new place? Here?



Student: Curving down.



Professor Michael McBride: Curving down, negative curvature.



Student: Negative amplitude.



Professor Michael McBride: Negative amplitude.



Student: Negative kinetic energy.



Professor Michael McBride: Negative kinetic energy; pink again. Is that enough? Oh, there's one more, here, the one right here. Okay?



Student: Negative.



Professor Michael McBride: Pardon me?



Student: Negative.



Professor Michael McBride: Negative, because it's -- how did you do it so quick? We didn't have to go through curvature.



Student: It goes away from the line.



Professor Michael McBride: Because it's curving away from the baseline, negative. Okay, pink. Okay, curving away from Ψ=0 means that the kinetic energy is negative. So now we know at all these positions whether the kinetic energy is positive, negative or zero, although there are a few that we aren't certain about. Right? So here's the potential energy that will do that. If you have this line for the total energy, right? Then here and here you have zero. Right? Also, incidentally, here and here, you have zero kinetic energy. With me? Okay. So no curvature, right? At these green places, the total energy is higher than the potential energy. So the kinetic energy is positive. Okay? At these places, the potential energy is higher than the total energy. So the kinetic energy is negative and the thing is curving away from the baseline. Right? And now we know something about this point. If the potential energy is a continuous kind of thing, then, although we couldn't tell by looking at the wave function, it's curving away from the baseline, but very slightly, right? It's negative kinetic energy there, and also on the right here is negative kinetic energy. And here we know, just by continuity, that at this point it must've been positive kinetic energy, even though we couldn't tell it by looking at the curve. There must be an inflection point when you go through zero, otherwise you'd get a discontinuity in the potential energy. Okay, so that one was green. Okay, now I have to stop.



[end of transcript]

Lecture 8
One-Dimensional Wave Functions
Play Video
One-Dimensional Wave Functions


Professor McBride expands on the recently introduced concept of the wave function by illustrating the relationship of the magnitude of the curvature of the wave function to the kinetic energy of the system, as well as the relationship of the square of the wave function to the electron probability density. The requirement that the wave function not diverge in areas of negative kinetic energy leads to only certain energies being allowed, a property which is explored for the harmonic oscillator, Morse potential, and the Columbic potential. Consideration of the influence of mass reveals an "isotope effect" on dynamics, on the energy, vibration frequency, and length of bonds.




Transcript



September 19, 2008




Professor Michael McBride: Okay, so last time we saw the jeopardy approach to solving the Schrödinger equation, which is to take the answer and find out what the problem was, and we saw that for a constant potential you got a sine wave. You could choose any length of sine wave you wanted, any wave period you wanted. It just turned out to be different energies and it had a relationship between the wavelength and the energy, the kinetic energy, or the total energy in that case. And we saw an exponential is -- corresponds to -- negative kinetic energy, which not very many people in the class liked, and not very many people ever like it, but suck it up because that's the way kinetic energy is for small particles. If you don't like to call it kinetic energy because you're so wedded to the fact that for big things it's ½mv2, then just consider it the difference between total energy and potential energy. Okay? But anyhow it's kinetic energy.



Okay, so we looked at the jeopardy approach, to start with the answer and get the question. But we'd a lot rather do it the other direction, to start with the problem and find out what the answer is. And you can do that by guessing the energy. If you have a sufficiently simple system, that is, one particle in one dimension; and this is a great way to get the idea of what's special about quantum mechanics, make it as simple as possible and then we'll go to realistic systems. Okay, so we can rearrange the Schrödinger equation. Remember, this is the Schrödinger equation: HΨ=EΨ; divide through by Ψ and split the HΨ/Ψ part into the potential energy part, V, which is a nothing, it's given to you, and the kinetic energy part, which is a negative constant that involves the inverse mass times the curvature of Ψ over Ψ. So it has to do with the shape of Ψ. And remember, if you multiply Ψ by a constant, you multiply the curvature by the same constant, so the kinetic energy stays the same. So it's only what we say, the shape, not really the amplitude, because that scales with the curvature.



Okay, but now we can rearrange this, this way. So we have a formula here for the curvature of Ψ. Now if you have a formula for the curvature, if you know how a curve is going to change, how a line is going to change its curvature [correction: slope], then you have a recipe for drawing it, if you know how to start. If you start with a certain height, and a certain slope -- suppose you know that, and you know the curvature, at that point, then you know if you take a little step in that direction what the next slope is going to be. Okay? Because that's how curved it is. And if you recalculate and know the curvature at that point, then you know what the curvature at the next point is going to be, and the next and the next and the next, and you can use this to just draw a curve that's bound to satisfy the Schrödinger equation because you're using the Schrödinger equation to generate it. Does everybody see this? This is really important. Do you want me to say it again, or you got it now? If you have a formula for the curvature -- that's how fast the slope is changing -- and if you know the initial height and the initial slope, then if you know the direction of the slope you know where the next point is going to be. So you go to that -- just think a tiny -- now this gets a problem in a computer. If the thing is really going like this and you take a finite step, the slope will have changed during your step and you won't get it just right. So you'll have to take smaller and smaller steps so that it doesn't change during the step. Chenyu?



Student: [inaudible]



Professor Michael McBride: Can't hear very well.



Student: Are you saying that the curvature is a second derivative?



Professor Michael McBride: Yes, the curvature is a second derivative, yes.



Student: How do we start with the correct slope?



Professor Michael McBride: We'll get to that. But if you have the initial slope, and the initial height, then you use the formula for the curvature to trace out the curve. Everybody got that now? Okay, so that's great, and it would be tedious to do this by hand, but that's what this little program, Erwin Meets Goldilocks, does is start the curve and draw it out, according to what -- if you know everything else in the equation. Okay, so, and remember it curves away from zero when the potential energy is greater than the total energy, and it curves towards zero when the potential energy is less than the total energy. Right? So the potential energy being greater than the total energy is a bad situation intuitively, right, it means the kinetic energy is negative, and it curves away from the baseline. And what's bad about that, or dangerous about that? Yes, Kevin? Pardon me?



Student: It's not bounded.



Professor Michael McBride: Yes, it'll keep going until it gets to infinity, if it's curving away. Okay, and that's not allowed, it has to not go to infinity. Okay, so now do we know everything else? Well we know m, the mass. We know the constant, C, and the V is given to us, the potential energy changes with position. Right? But we know how it changes with position; it's Coulomb's law typically, or we can do some other law if we want to. So this recipe then, you know everything on the right, except the total energy. Ψ you don't care about because you can scale it, you can make it anything you want, call it one. Okay? Then we can double it if we want to or whatever. So Ψ we know; m we know; C we know; V we know. We're going to guess E. So that means if we know the initial slope -- and for the problems we work on the initial slope will be zero, and you'll see why -- then we can use this formula to trace out the curve. Okay? And that's what the computer does for you. So people were worried last night about math. Don't worry about math. Let the computer calculate how these slopes change and so on. Right? All you have to do is think about the curvature. Okay? So you just have to think about the graph.



Okay, so we're going to talk now about nodes and quantization, in one dimension, using this program Erwin Meets Goldilocks. And there's a problem set for Monday, as you know, where you do this, and the answers are on the Wiki, but don't look at them before you've given it a good try. Okay, so let's -- there are simpler things to start with, like a constant potential energy or running into a wall or into a pillow more like. But the harmonic oscillator, or Hooke's law, is a very popular one to do; so we're going to do that. Okay, so here we're looking at the potential energy as a function of distance. And also on the -- so you see the distance goes from zero to two and a half angstroms and the potential energy goes from zero to 100 kilocalories per mole. You can change that if you want to on the computer. Okay? And the horizontal black line is a completely different plot. It's a plot of Ψ. That's the plot on top of this one where we're going to trace out the wavefunction. Okay? And that's zero of the wavefunction; here's zero of energy. Okay? And the reason we put them on top of one another is so that it's easy to see the influence of one on the other, of the wavefunction on the potential energy and vice-versa, and because they're functions of the same thing. What's Ψ a function of? What's it a function of?



Students: Position.



Professor Michael McBride: It's a function of the position. Position of what? The one particle we're talking about. Okay? And the potential energy is also a function of the position. Okay, so the way we're going to do this then is to guess an energy, let the computer trace out this curve, and see whether it's okay. What will tell us whether it's okay? How will we know if it's okay? It's got to be continuous. We're not going to have a problem with that because the computer is tracing it out this way; it's not going to be discontinuous. But the main thing is that it doesn't diverge, that it doesn't go to infinity. Okay, so let's just guess twenty-one kcal/mole. Why not? Okay, now you can already tell me something about what the curve is going to look like. Notice that at those two distances, between these two distances, the total energy is higher than the potential energy. What's the curve going to look like in that region? The kinetic energy is going to be positive, right? What will the curve look like? Pat?



Student: Spike up in between them.



Professor Michael McBride: Can't hear very well.



Student: It's going to spike up in between those two lines.



Professor Michael McBride: It may go up, it may go down. Yes? Pardon me? It's not -- would it be a sine curve? Does anybody think it wouldn't be a sine curve? Why not Russell?



Student: It does sinusoidal behavior when the potential is constant.



Professor Michael McBride: That's when the potential is constant. It's not constant through here. Right? So it won't be an exact sine curve. If this isn't very high, it might be close to a sine curve because it's sort of constant-ish, if this is a very small distance. But what characteristic must it have?



Student: It'll curve towards zero.



Professor Michael McBride: Say it again.



Student: It'll curve towards zero.



Professor Michael McBride: It'll curve towards zero. So it might -- it'll look something like this. Always -- it's curving and going up and down, maybe going high, but it's always curving to go back toward the baseline. It's never going to go to infinity. Okay? Now, there's a problem, because if you go the other way, the rest of the graph, the kinetic energy is negative because the potential's higher than the total. So in that region there's great danger that the line is going to go to infinity, because it's going to be curving away from the baseline. Okay? Now how could you possibly avoid that? How could you have it such that the curvature divided by the value of the wavefunction gets very big and the curvature away from the baseline, but still it doesn't go to infinity? How can the kinetic energy be big and negative, the curvature divided by the value is enormous, but it doesn't go to infinity? How can that be?



Student: Do we have to bound it?



Professor Michael McBride: Pardon me?



Student: Bound.



Professor Michael McBride: Bound?



Student: Yeah bound what we're looking at?



Professor Michael McBride: No, it's not a question of bound. This thing is, the particle, if it has this total energy, is bounded, it can't go all the way to the left or the right, that's true enough. Okay, but that's not the secret of how the wavefunction avoids going to infinity. Yeah?



Student: Make it a constant.



Professor Michael McBride: Pardon me?



Student: Make it a constant.



Professor Michael McBride: We can't -- make what a constant?



Student: Like -- another one a little bit.



Professor Michael McBride: Okay. Russell?



Student: You'd have very little curvature.



Professor Michael McBride: But -- you say very little curvature. But the kinetic energy is enormous. The curvature divided by the value of the wavefunction has to be enormous. Mike?



Student: You could make the value of the wave function go to zero.



Professor Michael McBride: Ah, if the value of the wavefunction is almost zero, then you can have an enormous -- pardon me, wait a minute -- then you can have a tiny numerator, hardly any curvature at all, and still the quotient can be big. Right? The kinetic energy can be enormous, and negative, if the denominator is near zero. So that means that if you're in those red regions, you know what the wavefunction has to look like. What does it have to look like? It has to get down near zero and stay there so that it doesn't have the curvature that'll -- enough curvature to pull it away. So it's got to look like that, on the left. Right? Because out here, that parabola is going way up; it's enormous, curvature divided by wavefunction, but if the wavefunction is essentially zero, still the curvature can be essentially zero. Okay? And the same on the right. So we know that the curve is going to look like this, something like that. Okay? And all we have to do is connect the pieces. Okay, so let's try it. We guessed twenty-one kcal/mole. This we did in our head, just by thinking about it. But let's let the program do it, and start at the left. And now do you see why it has to start with zero slope? Because if it's out in this region and has a finite slope, or an appreciable slope, and it's curving away, you're in trouble, because it's going to go to infinity. So we're going to use Erwin Meets Goldilocks, this program, and we're going to start at -- here goes the curve, and there's the part where the kinetic energy is negative, right? It hasn't gone to infinity because it started low enough and with low enough slope out at the left. Okay, now what's going to happen when it crosses that line?



[Students speak over one another]



Professor Michael McBride: What?



Student: The curvature changes.



Student: Now the curvature is going to be toward the baseline, as we guessed it would be; so like that. And now what's going to happen? Now it's going to start curving away from the baseline again. Uh oh, you don't win them all, right? Okay, now did we have too much energy or too little energy here? Notice that -- what did we want it to do on the right? On the left is where we started, that looks great, but on the right we're having trouble. What do you need to do in the middle in order for it to behave okay on the right? It has to not curve so much toward the baseline. Does everybody see what I'm saying? If it curved less, then it could join that other line we had originally. Okay, so we'll give it a little less energy. Let's try twenty. That's too hot, too much energy. Okay, so we'll guess twenty kcal/mole. Okay, here we go. Start out; almost exactly the same, it's nearly the same energy, so it doesn't surprise you that it looks about the same. And then it's going to curve in here, the way we expected, but a little bit less than it did before. Right? So now we're on the way to success. So what was that one?



Student: Too cold.



Professor Michael McBride: Too cold. And what do you need?



[Students speak over one another]



Professor Michael McBride: Someplace in between these, there's going to be one that's right. Okay? So it's between twenty and twenty-one. Now should we guess halfway between twenty and twenty-one, is that the reasonable thing to do? Which original guess was better? Which guess was better originally, and why? The red one or the blue one, which guess was better?



Student: The red one.



Professor Michael McBride: Why was the red one better?



Student: Because it curved close to the baseline.



Professor Michael McBride: It lasted longer before it went to infinity. So it'll be closer to the red than to the blue. And if you fiddle around -- and that's what I want you to do, because this is the way I think it gets into your head is by guessing numbers and saying "where should I make the next guess?" and so on. After you've done this a certain number of times -- but enough times that it really gets into your bones and your brain -- after you've done it enough times, then you can use the function that says "Solve" and it will do that guessing game for you. But do it yourself at the beginning. Okay, and it turns out that after you've done that, you get 20.74 da-da-da-da-da, like a lot of decimal points, decimal positions, and you get that one, just right. Okay? So that's the only math you need to do. Is it too hot or too cold or just right? Okay? And you have to know why it has to be that way, because of this curvature thing. Okay, so there we have it, 20.74 kcal/mole is a satisfactory energy, and it comes with this wavefunction, whatever that's good for. Okay? But the energy is what we were looking at, what energies are allowed for this system? Now, could there be a lower energy allowed for this system? Could you imagine a less curved wavefunction that would also have this property of not going to infinity? Can somebody guess what it would look like? Nick?



Student: You'd have one that would just go up.



Professor Michael McBride: Ah, it could just go up and then come back to the baseline -- it has to come back to the baseline -- but it could just go up rather than curving so much in the middle. Okay, so we guess a lower one. Could there be a lower energy Ψ? Yes indeed, there could be one at 4.15 -- it's actually .149-something-or-another, to a lot of decimal points. And there it is, exactly what Nick expected. Okay, so you got this. Could there be a still lower one?



Student: Yes, well in between the two.



Professor Michael McBride: Not in between. Can there be one lower than the blue?



Student: [inaudible]



Professor Michael McBride: Can you imagine a way to get it even lower than the blue, an even lower curvature than the blue? Yes, Angela? Pardon me?



Student: What if it curved, like the one --



Professor Michael McBride: Can't hear very well.



Student: If the blue one is going up, then there could be one going down.



Professor Michael McBride: Ah, there could be an upside-down blue one; absolutely true, there could be an upside-down blue one, but it's the same thing, because it's the blue one multiplied by minus one. Remember, you can multiply it by any number and it's still the same thing. Right? So it's exactly the same energy. It's not a different wavefunction, it's just multiplied by a constant. But you're right, that also does the trick, but it turns out to be the same thing. Can you imagine anything lower in energy than this? It's impossible, because if it had less curvature in the middle, it's bound to go to infinity when it got back in the other region. Okay, so that's -- the blue is the lowest one. Now could there be one in between these two?



Student: Yes.



Professor Michael McBride: How? It looks like -- how could you do it? Yes, Kevin?



Student: One peak is above the value and one peak --



Professor Michael McBride: Ah, one peak above and one below. Right? Could there be an energy in between? Yes indeed, there could be that one. Now could there be others in between? What happens as you add curvature? You have these nodes. A node is where the wavefunction is zero. It's not a node out here, where it's almost zero. A node is when it's exactly zero, and it has to be zero because it's going from being positive to being negative; obviously it has to go through zero. Out here, and out there, it's very, very, very small but it's not zero; that's not a node. A node is where it's zero because it's changing sign. Okay, now nodes are related to energy. Why? Notice that the lowest energy one, blue, had no node, the next one, the purple, had one node. The next one, the violet or whatever color it is, had two nodes. Why is there that relationship?



Student: I was noticing inflection points.



Professor Michael McBride: They are inflection points, but why do they get higher in energy when you have more nodes? Sam?



Student: Higher frequency.



Professor Michael McBride: Say it in terms of energy. Some people would say frequency is energy, that's true, but I want you to say it terms we've been talking about. How do you get more nodes? Yeah?



Student: A higher absolute value of curvature.



Professor Michael McBride: Yes, if the curvature gets bigger, you get more nodes, and curvature is the energy, the kinetic energy. Right? So more curvature, more nodes, more energy; or more energy, more curvature, more nodes. So we have zero nodes, one node, two nodes. Those are the lowest three you can get. And then there would be three nodes, four nodes, five nodes, six nodes, as many nodes as you want. Okay, now this is great. You can do all sorts of neat tricks for Monday, and what I want you to do with this program is just to fiddle around with it and make discoveries for yourself. And I'll show you some of the kinds of discoveries that you can do in the next slides here. But I warn you that if you have more dimensions, or more particles, things quickly become impossible. The only thing you can do this particular approach with is one particle and one dimension, because if you have more, then you don't know which curvature to change. You know the curvature -- remember, it's a sum of a bunch of curvatures, and knowing the kinetic energy, the difference between total and potential, doesn't tell you which curvature to change. If there's only one then you know that that's the one you have to do. So it's going to be a heck of a lot of work to do this for more complicated systems. Right? And that's not our purpose now. Our purpose is to understand how quantum mechanics works with this simple system. But the question is, is it going to be worth our effort to learn how it works if we're unable to do anything with it? Or another way of putting it is, "What's the reward, what are we going to get out of this? Is it worth spending some time on?"



Well, here's the reward for finding Ψ. [Laughter] The knowledge of everything. Okay? For example, you know what energies are allowed to a system; you know what the structures can be; you know how it behaves dynamically, how it moves; you know how bonds work -- that's what we're really after, remember? And you know what makes for reactivity. So there's a pretty good pot of gold at the end of this. So fasten your seatbelts and put your nose to the grindstone and all sorts of metaphors, and we'll do this for the next couple of weeks. Okay? First we'll look at allowed energies and structure. Okay now, of course, we already did this, right? No nodes, one node, two nodes. But what were the energies that were allowed? Okay, so here's a whole bunch of them. No nodes, one node, two nodes, three nodes, up to seven nodes or is it seven? Yes seven nodes. Now here are the energies: 4.15, 12.44, 20.74, 29.04, 37.34. Do you see any pattern to the energies? You might read the title if you have trouble seeing it. They're evenly spaced; after the first one they're evenly spaced. How about the first one compared with zero? How high is the first one if you know this spacing?



Students: Half.



Professor Michael McBride: It's half. The first one above zero is half as big as the spacing of the others. Right? So the energy is some constant, and that constant is what? 8.3 roughly, times an integer minus a half. Why do you have integers? Because nodes come in integers; you can't have half a node, right? It's zero nodes, one, two, three, four and so on. Okay? So that's pretty great. We know for springs, or harmonic oscillators, or an atom stuck on a bond, as long as the potential -- it doesn't vibrate so much that the potential deviates from looking like a parabola; at the bottom of the curve it looks like a parabola. So this is the way atoms should behave when they're attached by bonds -- approximately. Okay, now remember that doggerel poem from 1926: "We only wish that we could glean an inkling of what Ψ could mean." What have we used Ψ for so far?



Student: [inaudible]



Professor Michael McBride: We used it -- its not going to infinity -- to find out when we guessed the energy right. It was just a tool for telling us whether we had the right energy. But does Ψ really mean something, besides just being a handy tool for this purpose? And it was later, six months later, in 1926, that it was suggested what the meaning of it was. And it was in this paper by Max Born that the -- Ψ doesn't mean anything, but Ψ2 is probability density, or proportional to probability density. If you scale it right -- remember you can scale Ψ by -- multiply it by any number you want. If you scale it right, it is the probability density of the particle. Now here's what he wrote in German, and what it says is: "If one wishes to translate this result into physical terms, only one interpretation is possible, Ψ signifies the probability [of the structure]." But there's a footnote there, because he made a mistake, and he corrected it after the thing was already set in type. So he put a footnote here which says: "A correction added in proof: more careful consideration shows that the probably is proportional to the square of Ψ." Why couldn't it be -- why couldn't the probability be proportional to Ψ? Because Ψ can be negative! You can't have a negative… Well that's one reason. There were other reasons too. I don't know how embarrassed he was by this, but anyhow -- but since nobody else knew what Ψ was at all, I think he could be proud nonetheless, right? But anyhow, it's the square of Ψ, not Ψ itself, that gives probability density.



Now, six months later Albert Einstein wrote a letter to Max Born where he said something you've probably heard: "But an inner voice tells me that this is not the real thing. Theory yields a great deal, but it brings us no nearer to the secret of the Old One. Anyway I am convinced that He does not play dice." Right? So it's not probabilities. There's this fundamental determinism, Einstein thinks. But it turned out that in this respect Einstein was wrong. Right? It is probability.



Now, it's not probability, it's probability density, and we have to think a little bit about this, and we can think of it in -- by analogy, with mass density. Suppose you had a flask and it had mercury and water and oil in it, right? And you're going to ask the question, if the total mask in the flask is one kilogram, what fraction of that, or how much weight, is exactly one centimeter from the bottom? How many grams, or what fraction of the total do you think, would be exactly one centimeter from the bottom? Anybody got a guess? 5%, 1%, ½%, 10%?



Student: 12%.



Professor Michael McBride: 12%?



Student: Three.



Professor Michael McBride: There'll be -- this'll be an auction, right? Do I hear higher than 12%? Do I hear lower than 12%? Who's bidding? Yes, what do you say?



Student: Zero.



Professor Michael McBride: Zero?



Student: It's an exact value.



Professor Michael McBride: Why zero? Well obviously it has to do -- different heights would have different densities, right? So it'd be a lot more in mercury than it would if I drew the line in oil or in water, right? But you say zero. Why?



Student: Because it's an exact value. It won't have --



Professor Michael McBride: Right. A plane, or a line, or a dot doesn't have any volume. Right? You need a volume -- it's zero, you're right, you need a volume. You have to multiply density, grams per cubic centimeter by volume, how many cubic centimeters, in order to get mass. And the same is true of -- so that in this case the real question would be how much is between one centimeter and 1.01 centimeters, or something like that? You could ask that question, but you can't say how much is exactly one centimeter? So you have to multiply by volume, multiply density by volume, to get the amount. Right? And so Ψ2 is probability density. It's how probable it is to be in a small volume, a unit volume, around the point in question. So you go to some point in question, square Ψ, find the density, multiply it by the volume over which Ψ is constant -- because it doesn't vary real rapidly, so you could choose a really small one -- multiply, and that's the probability of being in that little block. Right?



Now, remember we said you could multiply, you could scale Ψ by any constant you wanted to, even minus one or -Ψ or whatever. Okay? Is there a good value, a preferred value to scale it by? You could scale it by such a number that when it's squared and you've multiplied by volume over every place, sum up all these little boxes, you get one. Why do you want to get one? Because the total probability is one that it's someplace, right? So that process is called normalization. We don't usually worry about that, right? We usually just take whatever numbers we want, that's convenient to plot on the graph, and know that we have to scale it such that if we want to calculate actual probability density that's going to sum to one. Okay, so here's an example where above we show two of these curves and below we show squares of those, not properly scaled together. There's, I think, a little more area in the red than in the blue, although maybe I'm bad. They should have the same area, if the thing were normalized. Okay, so that shows what it is. Now there's an interesting thing about that. If you have a pendulum, or a thing on a spring, and it's going back and forth, where does it spend most of its time?



Student: Extremes.



Professor Michael McBride: At the extremes, because it's going so slow there. Right? In the blue one where does it spend most of its time? Where does it spend most of its time? Where is it most dense if it's the blue wavefunction, the lowest wavefunction, the zero-node wavefunction? Obviously it spends its time in the middle, not at the extremes. But notice if you get to the seventh wavefunction; one, two, three, four, five, six, seven; the seventh wavefunction is going to have seven humps -- that then it's most probable at the extremes, and as it gets -- as you get higher and higher it gets more and more focused at the extremes, and that's why when you get to something that's really heavy it behaves the way you expect a pendulum or a thing on a spring to behave. Okay? This is not normalized. But notice something else that's really valuable here, really crucial, which is that this line is where the potential and the total are the same for the red curve. Right? So here is zero kinetic energy. In here the kinetic energy is positive, out here it's negative, and the thing can be out here. Right? So the thing exists in a region of negative kinetic energy. There's nothing special about that. Okay? It's just that that's the way things work. Kinetic energy can be negative. And in the case of the blue one, here was the point where the kinetic energy was zero, and all this out here is probability in a region of negative kinetic energy.



Now, what if we do it with a Morse potential? Notice if we look at the first one with a Morse potential, it looks the same as it would in a Hooke's-law harmonic oscillator. Does that surprise you? No, because the potential looks that way, pretty much. Right? But as you go higher and higher and get here to whatever that is, the twelfth or something like that, you can see that it's very different and it's spending a lot of time far out. It's shifting its average to the right. But that doesn't surprise you because the well moves out to the right, as you move up, in the Morse potential, as you move to higher energy, because low kinetic energy, which is what you have as you move to the right, means low curvature. Why does that make it high probability? Because as long as you have high curvature-over-wavefunction, whenever the wavefunction peeks up, it gets pulled down again; it can't stay up. But if you have very low curvature toward the baseline, very low kinetic energy, then it can go up and you don't mind, because the kinetic -- you don't need much curvature when the wavefunction is big. Okay, so it's like that, and then out in here it's pretty much like an exponential decay. Does that surprise you? Kate you look like it didn't surprise you.



Student: No it didn't.



Professor Michael McBride: Why not? When do you expect the wavefunction to be an exponential decay, under what condition?



Student: When the potential energy --



Professor Michael McBride: Can't hear very well.



Student: When the potential energy is constant.



Professor Michael McBride: Ah, when the potential energy is constant, and what else to be exponential? The sine wave is also constant.



Student: Right, but it has to be --



Professor Michael McBride: Right? What makes it exponential, rather than a sine wave? Yes?



Student: Negative kinetic energy.



Professor Michael McBride: Ah, the kinetic energy is negative, the potential's higher than the total. But that's exactly the case on the right, because there's the potential, there's the total, and the kinetic energy is going down; it's negative, the difference between the two. Okay? Now, here are a whole bunch of these for the harmonic -- for the Morse potential. How does it look different from Hooke's law? Does the spacing of energies look the same as it did in Hooke's law? Again, it's zero nodes, one, two, three, four and so on. But how is it different? Yes?



Student: [inaudible]



Professor Michael McBride: Ah, the spacing gets closer as you go up. So real bonds, which have a potential that looks more like this than like a spring, will have their energy levels get closer as you get more and more vibrational energy. Okay, they're not evenly spaced as they are for Hooke's law. As the energy increases, along with the number of nodes, the well widens more than it would for Hooke's-law parabola. So the wavelengths become longer, those nodes are spaced over a bigger distance, and the energies are lower than expected for Hooke's law, because you don't have as much curvature. You have the same number of nodes -- like the eighteenth level has the same number of nodes as it would in Hooke's law, but they're spaced over a wider distance, so there's less curvature, so the energy doesn't go up as much. Okay now, this is interesting. Look if you have the blue energy. Right? Now the wavefunction looks like this. It's no longer bounded. Does it surprise you that the wavefunction goes on out to infinity when the total energy is higher than that potential energy at the right? No. It means you have enough energy to break the bond. So the particle could just keep going. Okay, now what does this look like? In this region it looks like it does in a Hooke's law or something like that, more or less. What does it look like out at the right? It's a sine wave. Why? Sam, does that surprise you that it looks like a sine wave?



Student: Because the potential is just --



Professor Michael McBride: Ah, the potential is constant, effectively constant out there, and lower than the total. So it's a constant. It's about a sine wave because you have a constant positive kinetic energy when the total energy is above that dissociation limit. So there's the potential, there's the total, and now the kinetic energy is positive and about constant, so it's approximately a sine wave. Okay, now Coulombic potential is a more complicated one. The reason it's complicated is the energy goes to infinity when r becomes zero; when the positive and the negative are on top of one another it becomes infinitely favorable. So it's hard to do this numerically, and the program will screw up sometimes when you try to do Coulombic potential. And it's one-dimensional, and it turns out, curiously enough, that the system is simpler in three dimensions than in one dimension, for a Coulombic potential. Don't worry about the technical aspects of that. But at any rate, this is using the program for that, and you can see that the lowest -- that you have very, very high curvature when you get near zero because you have such high kinetic energy when the particles get very close to one another, and because the Erwin program is approximate, taking finite steps, you get in trouble often. Okay, but anyhow you get the idea that again as you get higher energy, you get waves that spread out. And they spread out a lot for the same reason they did in Morse potential, although here we're using a much lighter mass, we're using an electron. Right? But it spreads out much more because the well spreads out so much. Okay? So higher levels spread way out, and now the sequence of energies -- here's the lowest, the next and the next. You have a big spacing first and then it gets much, much smaller. It turns out that the energy is some constant divided by n2. So the number one level has an energy of -k, the next one is -¼k. The blue line is only a fourth as far down as the green, from zero, which is up at the top. Well zero is that grey line at the top, because it keeps going out to zero far out there. And the next one is 1/9th of the way from the top to the lowest level.



So the reward for finding Ψ was "the knowledge of everything." We've seen the allowed energy and structure. Now we're going to see another case where we get structure and we also see something about dynamics that's interesting, about the motion of the atoms. So here's -- suppose this is a proton attached by a bond to something that's heavy. So the heavy thing doesn't move, the hydrogen does the moving. Okay? And here's its wavefunction. This is the bond distance, the minimum energy. If you stretch the bond the energy goes up; if you compress the bond the energy goes up. And that's the wavefunction for the position of the hydrogen atom. Right? Now suppose we increase the mass and instead of talking about a hydrogen atom we talked about a carbon atom which had say fourteen, depending on which isotope we're using -- suppose we went to fourteen. How is that going to change things, if we have mass -- can we use the same function, if we use mass of fourteen? Notice that we're using the same potential. So the kinetic energy at any point is going to be the same for the two. How does mass enter into kinetic energy? Yes?



Student: Would it be more concentrated?



Professor Michael McBride: Can't hear.



Student: Would it be more concentrated?



Professor Michael McBride: Why do you say it would be more concentrated? What would make it more concentrated?



Student: Higher nuclear charge, so the electrons would want to--;



Professor Michael McBride: Yes, but you're not thinking about quantum mechanics. You're thinking that the higher -- well no, wait a second, we're not talking nuclear charge here. All we're talking is mass. It's held by a spring, we're thinking, okay? So electrostatics doesn't come into it, it's a spring that's stretching. Elizabeth?



Student: In that equation that you showed us before, the curvature is proportional to mass?



Professor Michael McBride: Yes, the kinetic energy has one over mass in it. So if you increase the mass, and keep the same wave, what are you going to do to the energy?



Student: Decrease.



Professor Michael McBride: You're going to decrease the energy. But it's the same energy as it was for hydrogen. You can't just decrease the energy, at a given position, if you know the total. Okay, so you want it to be -- to get the same energy it has to be more curved. Okay, everybody with me? To have the same energy, if you have a bigger mass in the denominator, you're going to have more curvature, fourteen times as much curvature, to compensate for the bigger mass. So it's going to curve more toward the baseline. Okay? So if we used that same energy and had the higher mass, so we got more curvature for the same energy, then it's going to curve a lot more, it's going to go through and go up to infinity, and it's even higher energy than the second wave would be. It's curved way too much. So how would you find the lowest energy, what would you do in your guessing? If you wanted to find the lowest energy what would you do? You'd lower your guess to make it less curved toward the baseline. Okay? And if you do that, so lower the energy, from that, and there we've got the answer. Right? So hydrogen has a much higher vibrational energy than carbon does for the same bond, to get the lowest energy. I think this will be clear to you when you have time to think about it. If not, ask questions. Okay? Now what if you had U-235 or a marble at the end of the same spring? What do you think its wavefunction would look like, the lowest energy wavefunction?



Student: A spike.



Professor Michael McBride: It's just a spike. The heavier it is, the narrower it gets. Right? So if you have something like a baseball held by a bond, it's going to sit right there. You're not going to be able to see that it's spread out. But electrons are so much lighter than nuclei that electrons really spread out. Okay, now so there are the probability distributions for hydrogen and for carbon, held by a single bond. Why is their average not in the same position? See what I mean? Elizabeth? Pardon?



Student: Acceleration has to do with mass.



Professor Michael McBride: No, no. It has to do -- yes?



Student: Because the electrons are not attracted equally.



Professor Michael McBride: No, no, it's -- remember, electrons aren't entering into this. We're just supposing that we have a nucleus held by a spring and that spring is the strength of a single-bond. Okay? What is the potential for a bond? Is it Hooke's law? What's a better potential for a bond than Hooke's law?



Students: Morse potential.



Professor Michael McBride: The Morse potential, which broadens out, remember, as it goes up. So since hydrogen goes up more, it samples further out to the right, so it's displaced to the right from the carbon. Okay? Now so let's -- here they are at their minimum energy position, and they vibrate. The higher energy is shifted to the right in the unsymmetrical Morse potential. And here's the half-maximum of the -- so it spends -- it's half as probable -- the probability density here is half what it is there. So when you say, how far does it spread out? This thing goes out to infinity. But we can talk about how wide it is by saying how wide it is at half-height. So the hydrogen at half-height is from here to here, but the carbon at half-height is only from there to there; that is, if you think of these things vibrating, the hydrogen has a bigger amplitude than carbon, because it's lighter. So the hydrogen vibrates by 0.1 angstrom, 9% of the distance. The carbon vibrates only half as much, 3% of the carbon-X distance, just because of the mass difference. And remember we said before, last time, that typically the atoms are vibrating by about 0.05 angstroms in the crystal. That's where this came from, it's this quantum mechanical thing. That's the least that they can -- and when they're in their lowest possible energy state they vibrate by that much, right? And that's how much it is, that little yellow dot after it shrank. So we've seen allowed energies and structure. We've seen dynamics, the vibration. Now, next time, we're going to look at dynamics, and also the payoff, bonding, how we understand bonding in one dimension.



[end of transcript]

Lecture 9
Chladni Figures and One-Electron Atoms
Play Video
Chladni Figures and One-Electron Atoms


After showing how a double-minimum potential generates one-dimensional bonding, Professor McBride moves on to multi-dimensional wave functions. Solving Schrödinger's three-dimensional differential equation might have been daunting, but it was not, because the necessary formulas had been worked out more than a century earlier in connection with acoustics. Acoustical "Chladni" figures show how nodal patterns relate to frequencies. The analogy is pursued by studying the form of wave functions for "hydrogen-like" one-electron atoms. Removing normalizing constants from the formulas for familiar orbitals reveals the underlying simplicity of their shapes.

 




Transcript



September 22, 2008




Professor Michael McBride: Okay, so we have the promise of the knowledge of everything, at least everything within a certain category, and we saw how we could see quantized energy last time, and structure, and structure-related dynamics, how nuclei vibrate and how it's different for heavier atoms. Today we're going to look again at a different kind of dynamics, and also at bonding. So in the one-dimensional world this is the payoff. We have to go to the three-dimensional world.



Okay, so you did problem sets for today and you're now familiar with things like this. What do you know about the blue curve and the red curve here? Anybody got a suggestion of what to do, or what they tell you about what the true solution is? Yes Angela?



Student: It looks like the true solution is somewhere between them.



Professor Michael McBride: I can't hear very well.



Student: It looks like the true solution will be somewhere between them, because it looks like --



Professor Michael McBride: Yes, the true solution should be between them. The red one is too hot and the blue one is too cold. Right? So there'll be, in between them, there'll be one that'll just come back down to the baseline for Hooke's law. Now the problem is that this isn't Hooke's law, because I didn't show you the whole picture. It's in fact a double minimum. Right? And those are both true solutions. So the correct single minimum energy would lie between those two, but for the double minimum there's one, there's a solution that's lower in energy and a solution that's higher in energy. Now what would happen if we moved the wells further apart? Here you see the separation of the two wells as 0.6 angstroms. Suppose we moved it apart to be 1.3 angstroms. Right? Now we have blue and red solutions. But now you notice the blue solution (or the red solution, they're the same in the left), they're both the same as the single minimum. So if you get the wells far enough apart they look like single minima with respect to their wavefunction. Okay?



Now, closer wells. If we move the wells closer together, we get a lower minimum energy than for the single well solution. Okay? And we get a higher energy for the next higher one. But they both look like the single minimum -- but they both started when the wells were far apart, looking like the single minimum solution, one with no nodes, one with one node. When the wells were far apart they had the same energy as the single minimum, but when the wells come together, the one with no nodes has less curvature, less kinetic energy and is lower energy than the single minimum, and the other one has one node, more curvature than the single minimum, and a higher energy; that is to say, there's a splitting between the energy of these two. If the wells are far apart, they're the same as the single minimum, both of them, the no-nodes and the one-node solution, are the same as the single minimum, but if you bring them together the no-nodes is lower in energy and the one-node is higher in energy.



So the energies split about the original single minimum structure. Now that's really important because that's what makes bonds. Right? So if we have two atoms or one-dimensional wells far apart, say A and B, and we get a solution in A and a solution in B, and if we want the solution for the whole thing, we just put them together that way, no nodes. Or we could flip B upside down, multiply it by minus one, add them together then and we get one node. But they have the same energy as the single ones would do. Right? But if we bring them together the energy gets lower, less curvature. Right? And that consists, or gives rise, to a stabilization of the particle. The particle is more stable when the wells are closer together. Right? What would you call that?



Student: A bond.



Professor Michael McBride: That's a bond. The energy is lower when they're close together. So that holds A and B together because the energy of the particle is lower when the wells are close together. So that's bonding. And part and parcel of this, the flipside of this, is that you have this wavefunction, which is higher. What will we call that?



Student: Antibonding.



Professor Michael McBride: Antibonding. So you get a combination of bonding and an antibonding combination because the kinetic energy changes when the wells come closer together. You get one that's less curved, one that's more curved. Okay, now that's bonding. How about dynamics? Well, suppose we have this double minimum and suppose we put a particle in and stop it there so it has zero -- that's our zero of energy -- and then we're going to let it go. What will it do if it's classical? It'll roll back and forth, right? Bingo, let it go. Roll across, to the same height, and roll back again, and then roll across again to the same height. With no friction it goes on forever, right? But here's what happens special, in real quantum mechanical systems -- oops, it went too far -- and now it's going to roll back and forth on the right; but every once in awhile it goes across, from one well to the other. Now, this is called tunneling, for obvious reasons. Right? But the word tunneling I hate. It's one of my pet peeves. It's misleading and mischievous, because it suggests there's something weird about the potential energy, that you can tunnel through and have a potential energy lower than you would've guessed from the potential energy curve. But that's not right. What happens is when it's in the middle you have negative kinetic energy. You have the potential energy that the curve shows, but the kinetic energy is negative. The potential energy can be higher than the total. And this happens in every bond and wavefunction, as you saw in your problem set, that the waves always go out into that forbidden region of negative kinetic energy where they curve away from the baseline. So there's the forbidden region, out to the left, to the right, and also in the middle for that energy. Right? And what really happens is that you go to negative kinetic energy and high positive energy to get over that hump, and then you get across to the other side.



Now how often does that happen? It turns out that the time to get from one well to the other, which I won't be able to prove to you because it requires time-dependent quantum mechanics, and we're talking about time-independent quantum mechanics, but let me just tell you that the answer is, if you know the energy difference between the blue and the red, that the rate of getting across is 5*10-14 seconds, divided by whatever that energy is, expressed in kilocalories/mole. Right? This is just an assertion, and it's based on time-dependent quantum mechanics. It's true, but I'm not telling you why. So you won't be satisfied, I hope. Okay? So the energy difference here is 1.4 kilocalories/mole. So how rapidly does a particle then get from -- if it started out in the left, how long do you have to wait, on average, until it's in the right one? You have to wait 5*10-14 divided by 1.4, which means about 4*10-14 seconds it takes to, quote, "tunnel"; although it really doesn't bore through, it goes over with negative kinetic energy. Okay, so there's something else about dynamics. And we're hoping soon to get to reactivity, but that'll come after the exam, after we talk about atoms and molecules.



So now it looks like we're in a pretty powerful position, at least with respect to one dimension, because with this Erwin program we can find satisfactory wavefunctions for any complicated potential. We could make it anything we would like and just follow the recipe, find out allowed energies, shapes of wavefunctions, probability density, rates of tunneling, all this kind of stuff, and we can rank all the wavefunctions by their energy, or their curvature - the number of nodes they have. Here's an unsatisfactory wavefunction. What do you notice that's bad about it, just to show that you've learned something? Yes?



Student: It doesn't start at the baseline.



Professor Michael McBride: It doesn't start at the baseline, on the left. How about on the right, does it look okay on the right?



[Students speak over one another]



Professor Michael McBride: Lucas?



Student: It tends to be going down.



Professor Michael McBride: Ah, it crosses the baseline, in the forbidden region. If it crosses the baseline while curving away, it's bound to go to negative infinity, or positive infinity if it crosses in the other direction. Right? And on the left, because it didn't get to the baseline when it became flat, it's going to keep curving and go up to infinity. So you've internalized a lot of this stuff now. Congratulations on doing well on the problem set. Okay? So that was a bad one. Okay, so it looks like we're in pretty good shape. We even can handle multiple minima and understand tunneling and so on. The unfortunate thing is this curve tracing recipe doesn't work if you have more dimensions. Remember, the reason it would work is if you knew the potential and the total energy, then you knew the kinetic energy, and you could assign it to curvature. Right? But if you have two dimensions, then there are two different curvatures you could assign it to, and you don't know how much to put in each of them. So it's no longer such a clear-cut problem if you have more dimensions, especially 3n dimensions. So when there are many curvatures, it's not clear how to partition the kinetic energy.



But Schrödinger had no trouble. The Schrödinger equation is what's called a differential equation, it involves derivatives. Right? And how many people have had differential equations? Okay, maybe a fifth of you I guess, or fewer. Right? So the rest of you are feeling, "Gee I wish I knew differential equations because then I could do this." But you don't have to, because people already have done it. You don't have to develop this from ground zero, right? And Schrödinger didn't solve the differential equations either. Right? He knew what the solutions were, because people had been studying them forever, and the reason they studied it was because of studying acoustic waves, and light also. And Chladni is the guy who originated this.



And here we have a book by Chladni called Acoustics. I got it out of the library. This is in fact the new, unchanged edition from 1830. The original edition was something like 1805 or something like that; actually we'll see it here I think. There's Chladni, Ernst Florens Friedrich Chladni, and this is the title page of that book, Acoustics; the original one was 1803, ours is 1830. Okay, now what he did was take plates of different -- he also did violin strings and timpani heads and things like that. He was interested in all kinds of acoustics. But the particular experiment we're interested in was he would take a plate and suspend it in the middle and touch it in different places, while he was bowing it with a violin bow to make it vibrate. Okay? So you bow it in different places, and he put sand on the surface of this thing to see what patterns would be formed when the thing vibrated -- this is why violins have the shape they have, so that they'll vibrate in certain ways when you bow to make them vibrate -- and the sand collects where the plate isn't vibrating, because when it's vibrating it shakes it away. So you get a picture of the pattern of vibration.



Now, I'm going to try to show this to you. But it might not work, and just in case I can go to the Web. So what we have here is a -- I can't play a violin, but as you know I can play a bell. In fact, I can play a bell in several ways; not only that, [Bell sound] I can do it this way too. [Bell sound] Now, that's not a very pleasing sound. Why? Why is it -- what's noise? I'm not very good at playing the bell. [Sound] Why does dry ice make a bell ring? This was discovered in London by an ice-cream vendor who had a bell on his cart, and he asked a physics teacher, a woman in London, why his bell rang when the dry-ice from his ice-cream hit it. [Sound] See, when you touch metal like that, it gives heat to the dry ice, which causes it to give off the pressure of gas, but when it gives off the pressure and the metal moves away, then heat's not being transferred, so it stops. So you get rapid pulses of CO2, which make the thing ring. But depending on how you touch it, you get noise, because there are many different patterns of vibration at different frequencies that are being generated. If I could do it just right, I would generate just one pure tone and then you'd say, "What a great player of bells he is."



Okay, now I'm going to try this, and I may fail in the same way, to do an experiment like Chladni's. But instead of using a violin bow -- I'm going to put sand on here -- instead of using a violin bow I'm going to use this piece of dry ice. Okay? And I have to -- if I just touch it, it won't do anything because I'll be touching sand and I won't get heat transfer. So I have to brush a little bit to get a clean part of the brass, and then let's see what happens here. [Sound] Now I'm doing the same thing I did with the bell, right? It's just noise. But let me try someplace else here. This is an empirical science. [Amazed sounds and laughter] Try something else here. Spread some around here. If you're lucky, this is really great. Hear that pure one? Maybe it's not sharp enough. We don't want to spend too much time doing this. [Sound] I want to try one more and then we're going to -- you can go to the website and see a much better once. Once I played just the lost chord, a really beautiful tone; and it's on, you can see it on the Web. [Sound] I'm trying to find where and how hard to touch so that I get a really pure tone. [Sound] Well, you can get patterns that way.



Okay, you can get much prettier patterns and prettier tones than I'm able to coax out of the thing right now, and you can look at it on the Web to see it. [projection adjustments] Okay, so these are some crude Chladni figures that I've made in the past, three rings; two rings with a line through the middle, vertically. That's the one we just made now, I think, isn't it? Isn't it the same number? A ring and three diameters. And there's a circle and four diameters. I actually made that in class last year. Okay, these are ones that Chladni did back in the late eighteenth century, taken from that book, right? And here, the first ones are just diameters, and the next ones are rings, together with a certain number of diameters, or two or more rings. Now let's see what that picture means. Below you see a vertical diameter and two rings. On top you see color coded as to whether the plane, at an instant of vibration, whether it's toward you or away from you, as it's vibrating; that's the pattern. Because those lines don't move, but either side of the line, one side moves up, the other side moves down. Okay? Or you could look at it this way. The dotted patterns and the circles and the line don't move as the rest of the thing deforms. This would be like a drum head on a timpani. Got the idea? So that's the pattern of motion.



Now Chladni was interested in sound, right? So he found what pitch each of these different patterns corresponded to. So this table shows the number of diameters that are nodes, the number of circular nodes, and what pitch it corresponds to, and the little lines may mean it's a little sharp or a little flat, something like that. For example, the one with two diameters and no rings is the pitch C. Got the idea? So then he tried to figure out if there's a mathematical relationship among these things. These were forty-seven patterns that he observed there. So here he says: "The pitch relationships agree approximately with the squares of the following numbers." So how many diameters, how many circles, and what he sees is the frequency is roughly the number of diameter nodes, plus twice the number of circular nodes, squared. This is his empirical observation. So, for example, you could have two diameters or one circle, both would give the number two, which you then square, to get something proportional to the frequency. Right? So there are two different ways to get the same pitch -- that's what's crucial -- either rings or lines. Or you can have a combination of rings and lines. So for higher frequencies you can get more and more combinations that give the same pitch. For example, the number eight, you can have four circles, three circles and two lines; two circles and four lines; a circle and six lines; or eight lines. All of them give the number eight and all give the same pitch, at least approximately. Okay? So this is what he -- but the lesson we want to take is that there are different ways of getting the same frequency, by combinations of these nodes.



Okay, now Chladni didn't solve his problem mathematically about how plates or strings or whatever vibrate. But there were great mathematicians working on this, like Bernoulli, Lagrange, Euler, who not only made a Communist Germany stamp, he also made a Swiss ten-franc note. Okay? Okay, so the Ψs for one-electron atoms -- now we're going from one dimension into three dimensions, for a real atom, electron and an atom -- so for one-electron atom there's going to be three variables, x, y, z for the electron, and the solutions, the waves we're going to get, involve what are called spherical harmonics, and they're 3D analogues of Chladni's 2D figures. These mathematicians could work in three dimensions as well as two. Okay, so a 3-dimensional H-atom wavefunction, Ψ, can be written as a product of three functions, an R function, a Θ function, and a Φ function. And these were available from other old-time mathematicians. The R function is called the Associated Laguerre Function, and it's named for Edmond Laguerre. And the Θ function is called the normalized Associated Legendre Polynomial, after Adrien-Marie Legendre. Okay? Now here are the solutions. Schrödinger didn't find these, he just looked them up. These guys had already done it, from acoustics, for 3-dimensional things vibrating. Right? So here's a table that you can use, the same way Schrödinger would've used it, to find out what wavefunctions for electrons in one-electron atoms actually look like. In all your books you've seen pictures of these things but you've never seen -- my guess is -- you've never seen the real thing, and you should wonder, "Are these pictures that people have shown me right?"



[projection adjustments]



Now how do you understand? This looks pretty complicated, this table, and in a sense it is. But if you look at it the right way it's really pretty simple. It relates to the position of an electron, shown on the right there, relative to the nucleus. And there's only one electron; it's a one-electron atom, otherwise we have more dimensions. But the nucleus can have any charge you want it to. So it's a one-electron atom of any nuclear charge, and we have the coordinates, x, y, z, and we have a potential. What's the potential law that's given to us, whose law?



Student: Coulomb's.



Professor Michael McBride: Coulomb's law. So it's 1/r2. And how do you write r2?; r2 is x2 + y2 + z2. Right, everybody know that? Everybody know that? Good, nod, yes; good, okay we're on the same page here. Okay, but that's a pretty complicated function, right? So the equations that involve that are going to be really, really complicated. But there's a clever way to get around this. Do you know what that way is? What's the clever way of getting around this? The clever way of getting around it is to change the coordinates of the system that you're using to one that's more natural for the problem. And the one that's natural for the problem is spherical polar coordinates. So the three dimensions are r, the distance, and θ, the angle down from the z axis, and φ, the angle rotated around from the x axis.



Student: Isn't it the other way around?



Professor Michael McBride: Well you can define it any way you want to. This is the way it's defined for these purposes. Okay? Now, so that simplifies the expression for potential enormously. Instead of being 1/√x2 + y2/c2, it's 1/r, or some constant over r. Okay? And what that means is -- and this is the mathematical payoff -- that the wavefunction can be written as a product of three functions, each of which is the function of only one variable. Instead of having x2 and y2 z2 all mixed together in this complicated function, you have a function only of r, a function only of θ, and a function only of φ, and you multiply them together and you get the solution to the Schrödinger equation for a one-electron atom. Now, that -- this table -- gives you those functions, the R function, the Θ function, and the Φ function. It's like a restaurant where you get to choose one from column A, one from column B, one from column C, right? Here's your appetizer, your main course and your dessert, right? So you choose one of these, one of these, one of these, multiply them together, and you have the true function. Right?



Now they look very, very complicated. You can't choose any combination you want to. Once you've chosen one -- once you've chosen your appetizer -- you can only have certain main courses, and once you've chosen that, you can only have certain desserts, to get solutions. Right? This is what these mathematicians had all figured out. Okay? And so how do we go about doing it? Well we can name the Ψ we're going to get by a nickname, like 1s or 2px or something like that, but you can also name it by the numbers that name these individual functions, and those are n, l and m, which I think you've encountered before probably. So let's make a try at it. No, actually notice something about the complicated form of the functions. First, every one of them has Z/a0 to the 3/2 power. Z is the nuclear charge; a0 is a unit of length of about half an angstrom about. Okay, so to the 3/2 power; that's pretty weird. But remember what you're going to do with the wavefunction? What do you do with wavefunctions, other than find energy?



Students: Probability.



Professor Michael McBride: Get probability density, and what do you do to do that Sam?



Student: Square it.



Professor Michael McBride: You square it. So when you square it, it's going to be (Z/a0)3, when you square it. Everybody with me on that? Okay, so what that means is you're going to get units that are Z3 per unit volume; because remember a0 is a distance. So a03 is a volume. So it's going to get a number -- the nuclear charge is a number Z, the atomic number is a number -- so it's going to get a number per unit volume, which is the right units to have for probability density. What is it per unit volume? Okay? So this is just the bit that scales it right so that you get the right density and the right units. So you can drop that out, that's not something very fundamentally interesting. Okay, let's try making up a 1s orbital. Where do you start? Can somebody tell me where to choose? It's not hard. They're labeled with red, right? If you want 1s, you choose the top one. Okay? And now if you've done that, n is one and l is zero. So when you come to the next one, l has to be zero. Now m can be either zero or one; or no, actually wait a second, if it's 1 -- no, if l is zero, then m has to be zero, it says in the second column. Does everybody see that? See up at the top? It says if l is zero, then m can be zero; lm. Okay? For m, pardon me, to be non-zero, like here, you have to have an l that's greater than zero. Okay? So to get a solution, if you chose n one and l had to be zero -- because remember that's the only choice here, you can't have a higher l -- then m is also going to be zero. So you've got to take this times this times this. Okay? Now how complicated are those functions? Okay, √2/2, that's just a constant, that's a nothing function. Or 1/√(2π), big deal. Why do you have these constants in there? Why do you care what constant you multiply the wavefunction by?



Student: No idea.



Professor Michael McBride: For what purpose?



Student: Normalization.



Professor Michael McBride: For normalization, right? If you want to get -- but if all you want is the shape, forget constants. So what's the real working part of the 1s wavefunction? What's the part that varies? We have here the (Z/a0)3/2; forget that; two, forget that, √2/2 forget that, in fact the 2s would cancel; √2 here would cancel that √2; so it's 1/√π times this, times what?; e(-ρ/2). That's the real working part of it. Right? When you square it, what are you going to get, for e (-ρ/2)? What's e(-ρ/2) squared? Yes Alex?



Student: e-ρ.



Professor Michael McBride: e-ρ. Okay, so the probability density is going to be e-ρ, times a constant. Okay, so here we got it. It's a constant times e(-ρ/2); when we square it, it's e-ρ. Now, there's something -- there are two things that are interesting here. One, we wanted to get a function of r, θ and φ, but we don't have r in it. Right? We have ρ instead. Why do we substitute r by ρ? There's a good reason for that. ρ is defined -- which is, indeed, the Greek version of r is ρ -- it's defined as r times a constant. The constant is twice the atomic number, divided by n, the quantum level n, times that distance a0. And why do you do it that way? It's because if you do it that way, then the same ρ works, no matter what the nuclear charge is, and no matter what n is. So you get the same -- so it makes the table very compact, because you use the same thing no matter what atom you're dealing with, if you work in units of ρ rather than r; that is scale r, to take into account the fact that you can have various nuclear charges. Okay, allows using the same e (-ρ/2) for any nuclear charge (Z) and any n. And notice, every one of these has e(-ρ/2) on it. Now, why does that make sense? Why is it no surprise that these things have e-r? Have we ever seen that before, e-x as a wavefunction? What kind of wavefunction -- what does that mean when you have e-r?



[Students speak over one another]



Professor Michael McBride: That's a constant negative kinetic energy, constant negative kinetic energy, which is going to be the situation for any nucleus and an electron. Once the electron gets pretty far away, the energy stops changing. Right? So the potential energy is constant, the kinetic energy is constant. Right? If the electron is bound to the nucleus, that means it can't just fly off to infinity, right? So it's below the ultimate potential energy. Then you have constant negative kinetic energy, when you're far from the nucleus, right? So e-ρ . So it doesn't surprise you that every one of these has e-ρ in it.



Okay, now let's just look at this scaling quickly. Okay, so here's e-ρ and ρ -- as a function of ρ. Okay? Now suppose -- so we can rearrange this. r is that. If we wanted to make a new plot, which has the horizontal axis being r, being distance instead of ρ. Okay now, so r for a hydrogen atom, the 1s orbital of a hydrogen atom, is 0.53 -- so n is one, a0 is half an angstrom, Z is one; so it's .53/2 times ρ. So we could just put a new scale on there. So here's half an angstrom for a hydrogen, here's one angstrom for a hydrogen. Everybody with me on this so far, how I did that? Right? I just went through and found out for any given ρ here, that I already had, one, two, three, four, five, e-ρ. For any give ρ, what would the r be if I were talking about a hydrogen atom? Okay?



And I just put a new scale on there. Now if I were talking about a carbon atom, which has a nuclear charge of plus six, then this number Z is going to be much bigger, it's going to be six times as big. What effect will that have on the scale? So suppose we're at five, for ρ. Instead of dividing by one, for Z here, I'm going to be dividing by six. So I'm going to get much shorter distances, real distances, rather than ρ. Right? Does that surprise you -- that the function is going to squash in if I have a plus six nuclear charge? Does it surprise you or not? It's what you expect. A bigger charge is going to suck the electrons in more. Okay? So if I do it for carbon, I get .53/12 instead of .53/2, and the scale is going to look like that. Instead of this being 0.5 here for hydrogen, it's 0.1, right? So then if I squash it in so that they're on the same scale -- those are different angstrom scales for carbon and hydrogen -- if I make them the same scale, the bigger nuclear charge sucks the 1s function in by a factor of six, and it looks like that. But now, of course, if I want the real, the normalized function, it's got to be higher, so that the total area is the same. I have the squared function here, e-ρ; remember that the wavefunction was e (-ρ/2); the density is e-ρ.



To get the same thing I'm going to have to multiply that by six. So it's going to look like that. So that's how the radial distribution, probability density distribution, looks different, for a one-electron atom with a plus one or a plus six nuclear charge. Carbon holds its core electrons in much more tightly than hydrogen does; no big surprise there. Okay. So here's something to think about. How would it differ if instead of talking about the 1s orbital of carbon, I was talking about the 2s orbital of carbon? Right? Now, ρ is going to be different, and here, instead of one, it's going to be twp. So it's going to change the scale by a factor of two. So the 2s are further out. Okay, so for Wednesday I want you to do these problems -- you can do them in groups if you want to -- some about Chladni Figures, and then some things about energy and some atomic orbital problems. Okay?



Now we already looked at the 1s. Let's look at some other atomic orbitals. Incidentally, this function, the Coulombic function, is simpler in three dimensions than it is in one dimension. It was that complicated thing in one-dimension, remember that had the cusp on it. It was a really complicated function, but it's really simple in three dimensions, it's just exponential in r. How does it vary with θ? As you come down from the axis, how does the wavefunction vary, the 1s? What does it say? How does it depend on θ?



Student: It doesn't.



Professor Michael McBride: It doesn't. How does it depend on φ?



Student: It doesn't.



Professor Michael McBride: It doesn't. So what shape does it have?



Student: A sphere.



Professor Michael McBride: It's spherically symmetric, it depends only on r. And the dependence on r is as easy as pie, it's just e-r (or ρ -- it depends on what units you measure r in). Okay, now let's look at 2s. Now tell me what to do. How do I write a 2s function from this table? First I have to choose the appetizer. Where do I go?



Student: Second one.



Professor Michael McBride: Second one down is 2s. Okay? So I take this one, and now 2s, that means l is zero. So where do I go; what direction do I go? Here. And then what direction do I go?



Student: Straight across.



Professor Michael McBride: Straight across, because m is zero. Okay. So here's -- there's 2s; multiply those together. Now what's interesting, there are a bunch of constants that are going to give me -- make it normalized. So that's a constant, that's a constant, just like they were before; this is a constant, that's a constant. But this part is interesting: (2-ρ) times e(-ρ/2). That's the function that we're interested in, (2-ρ) times e (-ρ/2). That has an interesting thing. What happens when ρ has the value two? It has the e (-ρ/2); that's just decaying, all of them have that. But how about the 2-ρ? That gives an interesting thing. What's its value when ρ is two?



Student: Zero.



Professor Michael McBride: Zero. So there's a node. What shape is the node? Is it a point, is it a line, is it a wiggly thing?



Student: A sphere.



Professor Michael McBride: It's a sphere, it's spherically symmetric again. So there's a spherical node. So inside -- and at a node, the wavefunction changes sign. So inside it's one sign, outside it's the opposite sign. Okay? Now how about a 2pz orbital? So where are we going to go? 2p, we'll start with this, and for pz, we do this one, and for z we do this one. So it's this times this times this. I think I have that circled here; let's see, yeah. Is this one interesting?



Student: No.



Professor Michael McBride: No. Is this one interesting?



Student: Yes.



Professor Michael McBride: What part of it is interesting?



Student: Cosine.



Professor Michael McBride: cos(θ) is interesting, right? That's real variation. And what part's interesting here? It's ρ times e(-ρ/2). Okay, so it's some constant times ρ times the cosine of θ times e(-ρ/2). You always have the e (-ρ/2); forget that. It just means it decays as you go out. Now ρ times cos(θ). What does that mean?; ρ, which is this distance, times the cos(θ), what is that? Russell? It's z. So I can simplify this a lot, if I mix my metaphors between polar and Cartesian coordinates, it's z times e(-ρ/2). That's the 2pz orbital. Now, can you guess what the 2px and the 2py look like? Can somebody guess the 2py? Josh?



Student: You replace x with --



Professor Michael McBride: Replace what?



Student: Replace z with x.



Professor Michael McBride: Replace z with x, you get the 2px orbital; replace z with y, you get the 2py orbital. Pretty straightforward functions. Now how about -- you've seen pictures of p orbitals. Now that you know what the functions look like, it's interesting to go back to the pictures you've seen and see what they mean. How do they relate to the actual formula? So you could plot -- remember, it's three parts, an R function, a Θ function, a Φ function. We could look at each of those functions separately, if we wanted to. Okay? We've been talking about the R function, how it would behave. But how would the Θ function behave in this, cosine θ? So we'll do a polar plot that shows the value of cosine θ as a function of θ. Everybody with me? Okay, so what is it when θ is zero; what's cosine of θ?



Student: One.



Professor Michael McBride: One. Okay, so there it is, and we'll put a point there, okay? Now let's go to + and -30°. Cosine is 0.86, and we'll put points there. And let's go to 45°, it's the √½, 0.71, put points there. At 60°, it's ½; put points there. What is it if it's 90°? If it's 90°, the cosine is zero. So we got a point at the origin. Right? So you can see what the shape is. What if you go the other direction, what if you go beyond 90°, what happens to the cosine?



Student: It's negative.



Professor Michael McBride: Changes sign. So you've got another circle that's negative on the other side. So there's a p orbital. But I'm not showing the whole p orbital, I'm just showing how the angular part varies. So you've seen pictures like that of p orbitals, which is the Θ function. Or I could square it, in order to look at probability density, as a function of θ, and it would look like that. So you've seen pictures more or less like that. Or I could make a picture like that, which is a contour plot showing how big it is, on a slice, through the nucleus. Right? So it's ρ times e (-ρ/2) times cos(θ); that multiplied by a constant is the wavefunction. Okay, so you can see it's positive up at the top, negative at the bottom. And let's look at it a little bit. And, notice it doesn't depend on φ, the angle of rotation around here. So we can take that picture and rotate it to give a three-dimensional dumbbell. Okay? We're just looking at one slice here, in order to be explicit. Okay, now so there's a certain point there that has a certain ρ and a certain θ, and therefore it has a certain value. Now, let's find the maximum, of that function. Okay? So first it depends on cos(θ). What will θ be in order for this to be a maximum?



Student: Zero.



Professor Michael McBride: Zero. So it has to be along the axis. Okay? Now, so θ has to be equal to zero. Okay, now how do we find where it's going to be maximum with respect to ρ? We take the derivative with respect to ρ, set it equal to zero. So that's this, or simplified that, or simplified that; ρ=2. So there's the maximum. And if we put the constants in, we could find the true probability density anywhere; just plug in θ, plug in ρ, and square it, multiply it by the constant. That's the electron density at any given, at the point you're talking about.



Now, or you could have a computer do it for you, and that's where we're indebted to Dean Dauger, who's a physicist. He went to Occidental College and actually wrote this program while he was there, but then he got a Ph.D. at UCLA in Physics, and he wrote this for physicists, but we'll use it. Okay, here's a picture of Dean Dauger. He also juggles. And here's a picture of him at an Apple Developer Conference, and he's juggling, the guy on the left there. He also drops one of the pins, he's not perfect. Okay. So this is his program. He's at the Apple Developer Conference. It works only on an Apple, so if you don't have an Apple, borrow somebody's Apple and look at the program, because it's really fun. This is the screen of the program, and I don't have time to take you through it now, but you should try it. So look at the PowerPoint of this on the next few slides to learn what all the different information is that's being given here, and the kinds of questions you can ask. There's a 1s orbital, right? And it's as if the thing were really -- the electron were really colored, and it's a picture of how it would actually look if you could see electron density. But all, of course, he's done is used fancy computer graphics to plot the square of the wavefunction. Right? But it's a really nice program. We'll talk about this a little more next time.



[end of transcript]

Lecture 10
Reality and the Orbital Approximation
Play Video
Reality and the Orbital Approximation


In discussions of the Schrödinger equation thus far, the systems described were either one-dimensional or involved a single electron. After discussing how increased nuclear charge affects the energies of one-electron atoms and then discussing hybridization, this lecture finally addresses the simple fact that multi-electron systems cannot be properly described in terms of one-electron orbitals.




Transcript



September 24, 2008



Professor Michael McBride: So back to business. We're in a really great position here because we've gone from one dimension to three dimensions, to get something real, a real atom. And if you take that table from last time, that we got from the old-time mathematicians, and plug in according to what you can use together, you get the exact wave function that's a solution of the Schrödinger equation for the hydrogen atom, or any atom that has only one electron. It can have any nuclear charge you want, right? So it's the real thing. People have shown you pictures, which are some kinds of approximations of this, but you have the equation, the actual formula, for the true wave function, and no one has a better one than you have. That's the real McCoy. Okay?



And you can check these pictures that people have drawn to see if they're realistic graphs, or whether they're just some kind of artist's impression, or somebody who doesn't know anything's impression; maybe he's not even an artist. Right? So you can actually do these. And many of them are very simple functions, like e-r, right? You can hardly get easier than that. Okay, but you can also draw pictures that look just like what you could see if the electron had a color to it. They're done by ray tracing where the computer pretends it's your eye and sees how much density it would go through, if it went through this point or that point or that point, along a certain ray from your eye, and puts a color here that's proportional to how much electron density it sees. It's not the wave function; it's the square of the wave function, the electron density, that Atom-in-a-Box plots. Okay? So that's what it would really look like, if you could see it and if there weren't problems with wavelength and photons knocking electrons to kingdom come and so on.



That's what it would look like, right? But it'll be colored; it'll be colored according to the sign the wave function has. It's not the wave function; it's the square of the wave function. So it would be positive. But it's colored to say the wave function in this region would be positive or in this region negative. And I'll show it to you in just a second, but first let's -- or a few minutes -- but first let's describe it. So there's lots of information on the screen that you get with Atom-in-a-Box. How many of you have tried it? Okay, all of you should. Okay? Okay, first you have the Schrödinger equation. That always appears and it's sort of silly but if you need the Schrödinger equation for reference, there it is. Then up at the top is the particular function that's being plotted, and it's given as n, l and m, and also the nickname 1s, for this particular one, which is 1,0,0 for n, l and m. And then there's the formula that that particular wave function has, and you know what it is, it's a bunch of constants, which we can forget, but it's e-r. He's not using ρ here, notice. He uses actual r and puts in explicitly the Z and the ao, and the two is cancelled. Okay? And then he says what the energy would be, in units of electron volts. An electron volt is 23.06 times as big as a kilocalorie, which is what we usually talk in. Now, you can look at these pictures and ask questions, some fairly simple questions, like where is the electron density highest? The nucleus is in the center of this picture, not surprisingly. So this is not a hard question. Where's the electron density highest?



[Students speak over one another]



Professor Michael McBride: In the middle, right? No big surprise there. Okay, how do you know? You take the wave function, you square it and find out where it has a maximum, and e-distance has a maximum when the distance is zero; the value is one. It falls off from there. Okay, so you just, to find where the density is highest, you just square the wave function and maximize it. Okay, now but here's a slightly different question: What is the most likely distance? Can you see why that's a different question than where -- you might think the most likely distance is wherever the density is highest, but that's not quite true. Do you see why?



Student: [Inaudible]



Professor Michael McBride: Because what's most-- yes?



Student: It could be moving very quickly in the area.



Professor Michael McBride: No, it doesn't have to do with motion. Yes Lucas?



Student: There's almost no volume in the center.



Professor Michael McBride: There's almost no volume in the center. If you take--say if you talk about a point, then there's no volume at all. Okay? But take a little box and put it in the center, or actually a sphere of a certain radius, a small radius, 0.0001 Å say, and put it in the center, and see how much density is inside. Okay? But then you could look at one angstrom out, and now you're looking at a little spherical shell that's 0.0001 Å thick. Is everybody clear? So there could be different directions in which you have this electron density, which is at that particular distance from the nucleus. And the further out you go, the more volume this shell has. It's always the same thickness, but as you move out there's more stuff in it. You see what I mean? More volume in it. What's it proportional to? As you increase r, how does the volume of the shell increase? How does the surface of the sphere increase as you go out?



Student: r3.



Professor Michael McBride: No, volume is r3.



Student: r2.



Professor Michael McBride: It goes with r2. The surface goes with r2. Right? So you have to multiply the probability density by r2, to see at what radius you have the most stuff. Is that clear to everyone? Because you want to count everything that's at that radius, not only in that direction but also in that direction, that direction, that direction, and so on, and the sphere gets bigger as you go out. So you'll notice that what's given in this plot here on the bottom right is r2 times the wave function squared. Okay? So that'll tell you what's the most likely distance. Okay? So you have to do surface weighting by r2, and it's not at zero, it's at the unit that we use to measure distance in, a0, which turns out to be half an angstrom, 0.53 Å. Okay, now here's a 2s orbital. Right? Here we'll draw two different shells of the same thickness, right? We'll number them one and two, and we're going to ask the question, which one has higher density? [Technical adjustments of room lighting] Okay, which shell has higher density, one or two?



Student: One.



Professor Michael McBride: One, clearly, okay? Now, here's a different question -- okay, it shows one -- which shell contains more electron density, more stuff? Which one has more probability?



Students: Two.



Professor Michael McBride: It's not obvious, because one has higher density but two is much bigger. Okay? And in fact, the two has about three times the radius of one. Right? So how much more volume does that shell have?



Student: Nine times.



Professor Michael McBride: About nine times more volume; not of the sphere but of the shell. Okay? So it's about nine times the volume. So it turns out that actually two has more stuff, and you see that in this plot. Right? The red for two is bigger than the green for one. Right? So depending on what question you ask, you have to decide whether you have to weight it by r2 to get the answer you need. Okay, now here's another way of looking at a wave function. This happens to be the -- it's actually a 3D wave function; as you see, not from here, but from here, it's 3D. This is the picture in three dimensions, using that ray tracing, the way we talk about. But you can do it, if you click "Slice" down here, then you get a slice through it. Okay? So that's 2-dimensional, more like the contour plots we've been using. It's just a slice through the 3D thing. So this is a 2D, not a projection, but an actual slice of it, and with this slider here you can control where the slice is. It's in the middle there; so it goes through the nucleus. But if you want, you can slide it down and see what the slice would be if it were in front of the screen that you sliced this picture. Everybody got it? So you can look at different levels that way, if you want to, or you can pretend you're looking at the whole thing. So it can be near or far.



Okay, now let's look at these two different, the 2D and the 3D pictures, and look at the pattern of nodes, the shape and the energy of the 3d orbital. You see here that there are diagonal nodes in this 3d orbital. Okay? So that's 3d. Let's look at this one. Now we still have those two diagonal nodes. In fact, it turns out that the nodes are actually cones. This is a slice through a cone, on the top, and on the bottom. Okay? So the nodal surface -- bear in mind that in one dimension a node is a point; in two dimensions a node is a line; in three dimensions a node is a surface. Right? So we're slicing it and see an intersection of the conical surface here, but the actual node is a cone. Okay, but is that the only node that you see in that picture on the right?



Student: There's a circular--



Professor Michael McBride: Ah-ha, Angela says she sees a circle too. Right? So there are two kinds of nodes in this orbital. There's both a conical node, which we had in the previous one; the 3d orbital had a conical node. That's what makes it a d orbital, having a conical node. But this one also has a spherical node, which is a circle when we slice it. Okay? So is this higher energy or lower energy than the one we looked at before?



Students: Higher.



Professor Michael McBride: How do you know it's higher?



Student: Nodes.



Student: More nodes.



Professor Michael McBride: It has more nodes. Did you ever see a situation before where you can have two different kinds of nodes? Here you have both, in this section, both a circle and lines. You ever see anything like that before?



Student: Chladni plates.



Professor Michael McBride: Right. Anyhow it's higher energy, it's 4d, not 3d, as you said, and it's like this, right? You have both circular and diameter nodes, but this is three-dimensional so you have both sphere and cone. So can you guess what a 5d orbital would look like? We looked at a 3d orbital. This is a 4d orbital. How about a 5d orbital? Lucas?



Student: It would have a third layer of nodes.



Professor Michael McBride: Not a third layer of nodes. It would have how many more nodes?



Student: One.



Professor Michael McBride: One more node.



Student: A circle, a sphere.



Professor Michael McBride: And it would be -- in the section it would be another circle; or actually it's another sphere of a different diameter. Okay? So you see how you can combine cones in this case--or it can be planes too; you have planes for p orbitals--you can combine spheres with planes, or cones, and you can combine them in different ways to get the same energy. For example, you can have a sphere and a cone that has a certain energy--that's 4d, the one we're looking at here--or you could have two sets of -- forget the sphere but have another cone. Does that remind you of anything? And get the same energy. Or you can have a plane and a cone, or you could have a sphere and a plane, or you can have a sphere and a cone -- I said that already -- or you could have two spheres, or you could have two planes. [Correction: a conical surface is in fact a double cone and counts as is equivalent to two spheres (3d ≈ 3s) or a sphere and a plane (3d ≈ 3p).] All those are two nodes--does everybody see what I'm saying--and they all have the same energy. Does that remind you of anything? It's not exactly the same, but it should remind you of something. That's what Chladni saw; remember, you could get different combinations, in his case, of circles and lines, and get the same frequency. Right? So that's why the 2s and the 2p have the same energy, or the 3s, the 3p and 3d have the same energy. Okay, now let's talk about scaling -- H-like. What do we mean by H-like?



Student: One electron.



Professor Michael McBride: It's one electron; any nuclear charge, but one electron. Why incidentally do we want it to be one electron? Zack?



Student: Repulsion.



Professor Michael McBride: That will turn out to be the case, but even if that weren't the case, there's a good reason to start with one electron. Yeah?



Student: You get a Coulombic interaction between the electrons.



Professor Michael McBride: Yeah, it's true that we don't want -- well we'll get into big trouble when we get Coulomb interaction between electrons. It's absolutely right. But we want to walk before we try to run; work with three variables, not six variables, which you'd have with two electrons, or 3n if you had n electrons. Right? So we want simple functions at the beginning. So that's what an H-like wave function is, just one electron, just three variables; ρ, θ, φ are the convenient ones. Okay, but now we want to see how we scale it as we change the nuclear charge. We already started talking about that last time when we talked about ρ, but we're going to talk about size, also about electron density, and also about energy and how the energies of these one-electron atoms change as you change the nuclear charge. Okay, this is what we talked about last time, that if you make the table up, the formulas, in ρ, instead of in r, then you could use it for any hydrogen-like atom, because the distance scale gets multiplied by the nuclear charge; that is to say, if you have a given ρ -- that's a given size of the function we're talking about -- a given ρ, then you have that value at shorter r, if you have a higher Z; a tradeoff between Z and r. If you double Z, you get the same value at half the r. Does that make sense, that as you get Z bigger, the distances get smaller? Does that make sense? Why?



[Students speak over one another]



Professor Michael McBride: Because the nucleus pulls the electron in more. And notice that it's linear, right? If you double the nuclear charge, you shrink by a factor of two, in the distance. Okay? So increasing Z shrinks the wave function, makes r smaller for the same ρ; that is, the same height of the wave function, or the same position in the shape of the wave function, because you have to renormalize if you shrink the wave function. Okay, so let's compare hydrogen, carbon+6, and potassium+19. Right? So the distance scale will go 1, 1/6th, 1/19th. So this is what it looks like. If a hydrogen atom is that big, this is how big the carbon one-electron atom is, and that's how big the potassium one-electron is, 1/19th as the linear distance that's in hydrogen. Now, how much does the electron density change? I think that's the next thing; yeah, scaling electron density with Z. You can see that just from what we just talked about. If you shrink the distance by a factor of two, how much do you change the density? Factor of two, does the density get twice as high if you make the distance half as long? Ryan?



Student: 23.



Professor Michael McBride: It's got to be 23, because you shrink it this way and this way and also this way. So it goes up with Z3, the density. That means if you have a hydrogen atom, the electrons are every which-a-way, all over, but if you have even a carbon atom, six times the nuclear charge for the same 1s orbital, the density is 63 as big. What's 63? Thirty-six -- ~forty five -- over 200, right? So you have 200 times the electron density for the 1s electron in carbon, compared to the 1s electron in hydrogen. Where does that crop up in experiment -- that you have so much higher electron density for carbon than you do for hydrogen -- and other things are even more so, like potassium -- where does it crop up? Yeah?



Student: The contour plots.



Professor Michael McBride: Of what?



Student: Of the molecules, the bond lengths.



Professor Michael McBride: What experiment?



Student: I forget.



Professor Michael McBride: Close. What?



Student: I forget.



Professor Michael McBride: You can do it by calculation and you'd see much higher density. That's what we're talking about is theory, but where do you see it experimentally? Lucas?



Student: Refraction difference diagrams.



Professor Michael McBride: Not to do -- well you can do it in the difference, but if you look at the total electron density you don't see hydrogens, unless you look -- work -- really hard, because the density of electrons, which is what you're seeing, is so much higher for other atoms. Okay, so the normalization has to be one. So if you shrink it, and that's where that comes in. We talked about that Z/ao to the 3/2 that gets squared before, and so that the Ψ2 is proportional to Z3, and the density of the 1s electron, as you go across these atoms, goes from one to 216 to 7000; potassium is 7000 times higher. Remember where we saw that in X-ray, what a very dark atom that was. Okay, so it helps X-rays find heavier atoms more easily.



Now how about the kinetic energy? Right? So Ψ is a function of r, some kind of function of r, but it includes a Z; Z times r, because ρ is Z times r. So now, how does that influence the derivative, right? If you have a function of Zr, the derivative of it is Z times the function of Z, by the derivative of -- oh what's it?; blah-blah. I got that wrong, or do I have it right?



Student: That's right.



Professor Michael McBride: That's right, yeah I got it right. Okay, now about how about the second derivative?



Student: Z2.



Professor Michael McBride: It's going to have a Z2 in it, right? So the curvature is going to be Z2 times what the curvature was before, but the value of the wave function is the same at some -- the corresponding ρ. So what's going to happen to the kinetic energy? If you have a bigger nuclear charge, what's going to happen to the kinetic energy at any point, at a corresponding point? Right? It's going to be proportional to Z2, the curvature divided by the value of the wave function. Right? So the kinetic energy is going to really zoom up for a hydrogen-like atom, when you get a heavier atom. Right?



So let's summarize what we did so far. So distance will shrink according to Z, electron density will go up according to Z3, and the kinetic energy will go with Z2. Okay?



Now how about the potential energy, how will it change as we change the nuclear charge? Well if we change the nuclear charge; Coulomb's law includes the nuclear charge, right? So at the same distance you'll have Z times whatever it was for hydrogen, much higher, or indeed lower potential energy, more negative, further from zero. Okay, but at a fixed-- that's-- so at a fixed distance the potential energy will be proportional to Z; but the distance shrinks, right? So if you look at corresponding positions in the wave function, or you look at the average potential energy, it's going to go-- it's going to also be, take into account that the distance gets smaller, r gets smaller, because the thing got shrunken by a factor of Z. So what happens to the average potential energy? How does it scale with Z? Maria, you got an idea? At a given distance, the energy, according to Coulomb's law, will be multiplied by Z, because you've got a bigger nuclear charge, but every distance gets shrunk by a factor of Z. Right? So if you average everything, what are you going to get? There are two factors of Z that are increasing (or decreasing, making more favorable) the energy, the potential energy. Got that? Two factors. So what's it going to be?



Student: Z2.



Professor Michael McBride: Z2, right? So the potential energy--what did the kinetic energy go with? Remember how the kinetic energy scaled?



Student: Z2.



Professor Michael McBride: Z2. How does the potential energy scale?



Student: Z2.



Professor Michael McBride: How about the average potential energy, if I double the nuclear charge? How much lower is the -- obviously if I increase the charge, the average energy of the electrons is going to get lower, but for two reasons, because the nuclear charge is bigger, and because the electron is twice as close. Okay? So how does it scale, the potential energy average? Z2. So the kinetic energy scales with Z2, the potential energy scales with Z2. How about the total energy? This is a hard one.



Student: Z2.



Professor Michael McBride: Z2, right? Okay, so the potential energy goes as Z2, and the total energy, here's a big surprise, it goes with Z2 too. It also happens to be 1/(n2). Right? So the energy gets -- yeah, so as you go up in n, you get less stable, right? Now R turns out to be about 300 kilocalories/mole, if you want to really get what the energies are. And here is a plot of the energies. Here's the 1s, 2s or 2p, 3s or 3p or 3d, 4s, p, d, f and so on, right? But the distance from zero, the amount of stabilization, is proportional to 1/(n2); and we saw the same thing if we did-- in the one-dimensional Coulomb case, the energy was also 1/(n2). Okay. Notice it's independent of l and m; 3s is the same as 3p is the same as 3d. Does that square with what you have learned before?



Student: No.



Professor Michael McBride: What did you learn before? Lexy, did you learn anything about that before, about the energies of 3s, 3p, 3d? No. Anybody? Shai?



Student: s, p, d.



Professor Michael McBride: Yes, s is lower than p is lower than d. Now how can both these things be true, that you heard something that was true, and I'm telling you they're all the same?



Student: There are more electrons.



Professor Michael McBride: Ah-ha! What you heard about was real atoms, things that have many electrons. I'm telling you about if the atom has only one electron, then s and p and d are the same. Okay? But it'll be different when you have many electrons, and we'll see why shortly. Okay, because this is one-electron atoms. Okay, so we've -- this just summarizes -- the size goes with 1/Z, the electron density is proportional to Z3, and the energy goes with Z2, whether you're talking potential, kinetic or total energy. Okay, and it's also 1/n2. And the size happens to be n/Z as well. Okay, now here's a really pretty one, and to do this, to see this right, you want to see the real thing. So notice that this is a 2p orbital; n, l, m are 2,1,1. Everybody see that? I want to show you the real program that does that. [Technical adjustments to show 3d orbital in Atom-in-a-Box program] Okay, now what I want for the display -- this display is just as if electrons were all white, okay? But we want to color them according to the color of the -- [Technical Adjustments of Atom-in-a-Box format] -- we want to color them with the phase, plus, minus. That's 3d. The phase keeps changing. That has to do with time dependence, which we're not talking about. Okay? Now, that's a 3d orbital. We want to have -- what did we have before? We had 2p, right? So I'll make this instead of a three, a two. So there's a 2p orbital, a particular 2p orbital, 2,1,0. Which one is that? Notice you have a coordinate system next door there, and I can rotate this thing, I can grab it and rotate it; the coordinate system rotates. See that?



Student: Cool!



Professor Michael McBride: So I can look at it end-on if I want to, or look at it from the side. Which p orbital is that?



Students: 2pz.



Professor Michael McBride: How do you know?



[Students speak over one another]



Professor Michael McBride: Points along the z axis. Okay? Now, let's look at another 2p orbital, because we can make m one instead of zero. So the 2,1,0 is 2pz. Now what's 2,1,1? Let me look at it this way. That's time-dependence, right, and that's weird. And physicists love this sort of thing, because they deal with atoms in isolation, and atoms in isolation have angular momentum, things going around. But atoms, when they're in molecules, when they're with other atoms, or in other things that distort it, like in an electric field, they don't behave that way. So chemists use a different way of doing it, right? It's not that physicists are right and chemists are wrong, or that chemists are right and physicists are wrong, it's if you're dealing with isolated atoms, you use one thing; if you're dealing with atoms interacting with other atoms you should use something else. And the way you do this, dealing with something else -- I'll show you here. You do what's called superposition; that is, you add two functions in order to get a new function, and the two functions I'm going to add, when I click "Superposition" here, are 2,1,1, and the other one I'm going to use is 2,1,-1. And now what's that?



Students: 2py.



Professor Michael McBride: That's the 2py orbital. Okay, does everybody see what I did? So I can combine two of the physicist's orbitals to get the chemist's 2py orbital. Dean Dauger, who wrote this, is a physicist, so he writes physicist programs, right? But we can trick it to be chemistry by adding two of his together, or by subtracting them; and when you subtract them you have to put an i in -- imaginary. So I'm not going to do it right now. You can do it yourself. You can see down here, you can do the phase and so on; the phase ofp makes it imaginary. I won't do that now. We'll go back to the PowerPoint. [Technical Adjustment to change program] So superposition of that kind, when you add two functions together to get a new function, for whatever purpose, is a kind of what's called hybridization. Now there are other kinds of -- have you heard about hybrid orbitals? Right. What that is is adding two orbitals together to get a new orbital. And let's look about when that would be. As we said, this is the physicist's 2p, m=1, with orbital angular momentum, and it has complex numbers involved. And the chemist's 2p is that sum. Okay.



Now multiplying and adding wave functions. We've already seen a case where you multiply functions. Do you remember where we've seen things be multiplied?



[Students speak over one another]



Professor Michael McBride: Alison, do you remember? No.



Student: You multiplied the wave functions.



Professor Michael McBride: Multiplied which wave functions? Let me tell you one place we already multiplied was squaring a wave function is a kind of multiplication to get probability. But I want a different one.



Student: When we, for different variables --



Professor Michael McBride: Ah, for separating the variables, r, θ and φ; we had an R function, a Θ function, a Φ function, and we multiplied these pieces together to get the total wave function for the electron, the three-dimensional wave function. So that's one place you multiply, to get a one-electron wave function, which is -- we call an orbital. An orbital is a one-electron wave function; that's what an orbital is. Now, you can add orbitals to create a hybrid orbital; that's what we just did, we took two physicist's orbitals, added them together, and got a new orbital. Okay? 2py + 2pz is a hybrid orbital. An orbital is a one-electron wave function. What's it a function of? What variables do you put in, in order to get a value out? How many variables? Sophie?



Student: Three.



Professor Michael McBride: Three. What are they?



Student: Those.



Professor Michael McBride: The position of that electron, the one electron. If it's a one-electron orbital, you can write the orbital as many pieces -- here are two pieces, right? But it still is a function of just the position of one electron; it's not a function of six, or eighteen, or whatever number of variables, it's just a function of -- it's an orbital, a one-electron function, so it's a position of one electron that you plug in. Now, you can change orientation of orbitals by hybridization. Let's look at that. So here's the 2pz orbital. I've rotated the coordinate system so z is horizontal. Okay? And here's the 2py orbital, which we already made by adding two physicist's orbitals together. Okay? Now what happens -- and notice the 2py , surprise, is pointed along the y axis, 2pz was pointed along the z axis. What if I add the two together? What if I add 2pz and 2py? Now let's go back and look at 2pz a second. Notice it's red on the left, blue on the right; say positive numbers on the left, negative numbers on the right -- everybody with me? -- because it assigns numbers to different points in space, positive numbers on the left, negative on the right, let's say. Now this one assigns positive numbers on the top, negative numbers on the bottom. Suppose I add together, what am I going to get? Andrew?



Student: You get diagonal.



Professor Michael McBride: Ah, they should reinforce in the top left where they're both positive, and in the bottom right where they're both negative, but in the bottom left and the top right they should cancel. Okay? So if we add them together to make a weighted sum -- we don't have to add them 50:50, but we'll start by adding 50:50. Okay? So if we add them, a 50:50 mixture will give that, just what you said. Okay, so we changed the orientation of the p orbital by adding two. Okay, now suppose we didn't use 50%, suppose we used another mixture, suppose we used 75%, right? Or there's 50%, or 25%. Okay? So you can get any orientation you want by mixing -- within that plane, within the yz plane, by adjusting the ratio. Okay, there's 2py.



Now, what would happen with a 2s orbital if you put it in an electric field? You're changing the rules, you're changing the potential, because now the electron has -- doesn't just -- its energy doesn't depend just on its distance from the nucleus, it also depends on where it is in the field. So we're changing the definition of the potential energy. So it's a different quantum mechanical problem. But I don't want you to think about quantum mechanics, I just want you to think about what would happen if you put an electric field on. What do you think?



Student: It shouldn't --



Professor Michael McBride: Would it move the atom?



Student: Yes.



Professor Michael McBride: No, because the atom is neutral, overall. We'll do a hydrogen atom here. The atom's neutral, so it won't move. But what will happen?



[Students speak over one another]



Professor Michael McBride: The electrons will move one way and the nucleus will move the other, and if you've preserved the center of mass, then the nucleus will hardly move at all, and the electrons will do most of the moving. Okay. So you can change the shape by hybridization. We can get something that will work in an electric field, describe an atom in an electric field, by mixing the ones without electric field. Now let's look how we do that. We make an spn hybrid orbital. Have you ever seen that?; n is a number. Can you give me an example? two, sp2 or?



Student: sp3.



Professor Michael McBride: sp3 or sp, where n is one, right? Or sp0. Have you ever seen that?



Student: No.



Professor Michael McBride: sp0 is just s. This is how much p there is in it, after you square it. That is, it's a weighted sum, this thing you're using. You have different values of a and b, and it's weighted such that -- so when you square it, you get a2, a fraction of 2s squared; that's the distribution 2s would have; b2 of the way py would be when squared. But you also get a cross-term, 2ab times 2s times 2p. Okay, now so n, the number you put in to talk about sp2, sp3 and so on, is the square of b2/a2. So if you have an sp hybrid, it means equal parts of s and p, and if you have sp3 , it means three times as much p as s, after squaring. Okay, so it's what fraction you have of something that looks like 2s squared and 2p squared; and then there's also that part. Okay, so now let's just look at some examples here. Suppose n is zero; then you have just pure s. Okay? Now suppose you put an electric field that wants to shift the electrons. It would push a positive charge to the right but it will push electrons to the left. Okay, so let's -- here we went to 0.02, just a tiny amount of p. But see how it -- here's back to s, and there's sp0.02. It shifted the electrons to the left. What if I go to a higher value? 0.04, 0.06, 0.09, 0.11, 0.18, 0.25, 0.33. So what I'm doing is turning up a dial to make the electron -- electric field stronger -- and the electron is shifting, by changing the hybridization. Okay, 0.5, 1. There's an sp hybrid. Now what'll happen if I keep on going, make it sp2, for example. What do you think will happen? Guess.



Student: It'll get more and more to the right.



Professor Michael McBride: It'll keep going to the right. Wrong! Watch this. That's the maximum extension is sp. That's the most you can extend. Now watch what happens if I make it bigger; sp2, sp3, sp4, sp9, sp24, pure p. Right? Because both pure s and pure p are symmetrical. The 50:50 mixture is the one that shifts the most to the left.



Okay, now we're going to close here with trying to go beyond what we've already seen. What we've seen is one-electron atoms, hydrogen-like atoms. So we find that we can multiply pieces, R, Θ, Φ, to create this one-electron wave function, the orbital, and we can furthermore add these things to make hybrids that shift things around in useful ways. We can add orbitals to make an sp hybrid, for example, as what would happen in an electric field. But it's still an orbital, so it allows adjusting to new situations -- for example, an electric field -- while preserving the virtues of real solutions of the nuclear potential, because you never put on a field as strong as the field felt by the electron near a nucleus. Adding the external field by turning up your dial is just a small change. So you want things to look pretty much like what they looked originally, with the nucleus, just a little change, and you can do that by hybridization. Okay, the big deal is the nuclear potential at small distances.



Now, let's try to go to a two-electron wave function. Okay? So what's it a function of? What's a two-electron wave function going to be a function of, how many variables?



Students: Six.



Professor Michael McBride: Right, six, right? x,y,z for one electron, and x,y,z for the other electron, or r, θ, φ of electron one, and r, θ, φ of electron two, and they're on the same nucleus, so you measure the distances from the same point, but they're two different distances, one to electron one, another to electron two. They could be the same, but they don't need to be, they could be different. Okay? Now wouldn't it be neat -- that's the question mark -- if we could get that by multiplying two one-electron wave functions? Because we have one-electron wave functions cold; we look in the table and we've got the exact thing. If we could take two of those and multiply them together, so orbital A for electron one times orbital B for electron two -- remember, what these things are is just numbers. If you put in some value of position of electron one and some value of position of electron two, this first thing will give you a number, the second thing will give you a number. If you could multiply those two numbers together and get the number for the total wave function, that would be really great, because then you'd know how to do it. The question is, can you do that? Remember, an orbital is a one-electron wave function. So can we multiply orbitals in order to get a many-electron wave function? If we could -- and then, of course, we would square it to get probability density. Now it would be a different probability. In the first case it was probability of electron one being here. Now it's the joint probability of electron one is here, and electron two at the same time is here; that's the probability when you have six variables. Okay? Wouldn't that be great? It's just paradise, right?



Because if we had many electrons we could just multiply all those orbitals together, a certain orbital for electron one, another orbital for electron two, another one for electron three. We know how to write those; multiply them together, we get the thing. We find the total energy is whatever energy electron one had, plus whatever energy electron two had. That's great. We find the total electron density. We know how to find the electron density of each individual electron. So the total electron density should be just the sum of those. So if we were interested in the total electron density at this particular point, we'd see what the electron density of electron one is at that point, and what the electron density of electron two is at that point, and we'd add them together and that would give the total electron density, at that point. So the total electron density is a function of x, y, z. It's whatever this is, evaluated at x, y, z, plus whatever this, plus all the others. Absolutely wonderful. Right? And then you could write things. So the whole is the sum of the parts. Right? And you could write things like this at the top, that Neon has the configuration: two electrons in 1s, two electrons in 2s, two electrons in 2px, 2py. Right? Things like that. You've seen things like that. So those are the orbitals in which electrons are, to describe the neon atom. Right? Now, this is a problem in joint probability. So it's like tossing two coins to try to get to two heads. So I don't know if I can do these at the same time, but I have two quarters here. Need your back of your hand. Okay, so I got tails. What did you get?



Student: Heads.



Professor Michael McBride: How often?



Student: Half the time.



Professor Michael McBride: Half the time. Actually if I get tails and you get heads, that's a quarter of the time.



Student: Yes, that's a quarter of the time.



Professor Michael McBride: But a quarter of the time you'll get tails and I'll get heads. So heads/tails is half the time. How about heads/heads, how often?



Student: Heads/heads. Still a quarter.



Professor Michael McBride: A quarter of the time. Now I'll bet you, I'll bet you that I can get head/head half the time, actually doing the experiment.



Student: Yes.



Professor Michael McBride: You want to take the bet? I would advise against it.



[Laughter]



Student: I won't take it then.



Professor Michael McBride: Okay, here's what I'm going to do. Okay?



[Professor McBride tapes the quarters together]



Professor Michael McBride: Two tails, I lose.



[Laughter]



Professor Michael McBride: Whoops.



[Laughter]



Professor Michael McBride: Two tails, I lost again. But I think you can see that if I do it a lot of times I'm going to win half the time. Why? Why was the rule wrong?



Student: They're not independent.



Professor Michael McBride: Because they weren't independent, right? If they're independent, the probability of both A and B, that is, the probability that A will be here and at the same time B will be here, is the product of the probabilities; but if you do this, that's not true anymore. Right? So if you have a two-electron wave function, you square it to get the joint probability, that electron one will be here and electron two will be here at the same time. So is that valid, to multiply -- here I'm multiplying the individual probabilities that electron A is at a certain position, electron B is at another position. Do I get the joint probability that way? Well under what conditions do I get the joint probability that way?



Students: If they're independent.



Professor Michael McBride: If they're independent. Now, here's the big question. Are they independent?



Students: No.



Professor Michael McBride: Why not?



[Students speak over one another]



Professor Michael McBride: Because of Coulomb's law, they repel one another. There's no way they can be independent, they repel one another. So orbitals, one-electron wave functions, to express many-electron wave functions, cannot be right. Orbitals are a fiction, except for one-electron problems. Right? hydrogen-like atoms fine; we've got hydrogen atoms cold. But if you got more than one electron, orbitals won't work. But still we're going to use them.



[Laughter]



Professor Michael McBride: Right? And that's what you're going to find out after the exam. Okay, good luck on -- so that doesn't work -- good luck on the exam.



[end of transcript]

Lecture 11
Orbital Correction and Plum-Pudding Molecules
Play Video
Orbital Correction and Plum-Pudding Molecules


The lecture opens with tricks ("Z-effective" and "Self Consistent Field") that allow one to correct approximately for the error in using orbitals that is due to electron repulsion. This error is hidden by naming it "correlation energy." Professor McBride introduces molecules by modifying J.J. Thomson's Plum-Pudding model of the atom to rationalize the form of molecular orbitals. There is a close analogy in form between the molecular orbitals of CH4 and NH3 and the atomic orbitals of neon, which has the same number of protons and neutrons. The underlying form due to kinetic energy is distorted by pulling protons out of the Ne nucleus to play the role of H atoms.




Transcript



September 29, 2008




Professor Michael McBride: Okay, so what's coming for the next exam? Well we've been looking at atoms, and at the idea of orbitals for many-electron atoms, which we showed last time is wrong. So today we want to recover from the orbital approximation. Okay? Then we're going to look at molecules. And first we're going to look at them in a very unconventional way, from the point of view of molecules as plum-puddings, what's called the "United Atom" limit, to apply what we know from atoms to molecules. But then we're going to look at a very, very different way of looking at it, to try to understand bonds in terms of what are called linear combinations -- that means weighted sums -- of atomic orbitals to make molecular orbitals; but we're going to try understanding bonds.



And then these terms that don't mean much to you now but will mean a lot later on: "energy-match" and "overlap." And then we're going to get to reality -- all this stuff is theory -- but we're going to look at something real: at XH3 molecules, with different atoms for X, at their structure and at their dynamics, and how that ties in with our understanding of bonding. And then we'll go on to reactivity, which is of course the main goal. We'll talk about HOMOs and LUMOs, and you'll see what those are, and how to recognize functional groups and their reactivity; and that's the real prize, is to be able to -- you've been memorizing functional groups for this previous exam, but what I want you to be able to do is look at a molecule and recognize when it has a functional group, even if you've never seen it before, and how it might react, what it would react with. And then we're going to see how organic chemistry really developed, the real thing; how it developed from the time of Lavoisier.



Okay, so that's what we're going to. And here's where we were last time. We were wondering, when it comes to a two-electron wave function, might it be possible to write it as a product of one-electron wave functions? Because we know how to write 1-electron wave functions; we've got that table, we can write the real thing, at least for an atom. So if it were possible to write the six-variable, two-electron wave function as a product of one-electron wave functions, that would be a fantastic simplification, and all we'd have to do is square and we'd get the joint probability; not really that you care very much about the joint probability. I suspect you haven't stayed up late nights worrying about joint probability. But anyhow, if you want to handle things with more than one electron, you have to have a two-electron wave function. And if we could do it from one-electron wave functions, then we're really in a good position. But we ended the first quarter of the semester on a downer, the idea that there's no way electrons can be independent. They repel one another, so orbitals are fundamentally wrong. Okay? So forget that. And here's our paradise; it's gone. Okay? But are there tricks that would allow us to salvage orbitals, to use them, even though we know they're wrong?



Okay, well the first, the simple-minded trick, is Z-effective. We talked about how you could scale the one-electron wave functions, the atomic orbitals, for nuclear charge; if you increase the nuclear charge they contract; and we saw what different properties are proportional to. So if you have other electrons in the atom, it could be that they could be sort of approximated, as if they were just reducing the nuclear charge, right? There's a certain repulsion, as well as an attraction taking place, if there are other electrons there. But maybe we can just reduce the attraction, right? And things will be sort of -- at least corrected in the right direction. Right? So pretend that the other electrons, or some fraction of them, are concentrated at the nucleus. So from the point of view of an electron you're thinking about, they just reduce the nuclear charge; and that's a problem we know how to handle, what happens if you change the nuclear charge. Okay? So might there be an effective nuclear charge that we could use? So we pretend the other electrons just reduce the nuclear charge for the electron we're interested in, the orbital, the one-electron wave function that we want to find. So we're going to just to try to find a one-electron wave function in a many-electron problem. Okay? And this is what we know, that 1s looks like that and ρ scales with Z. So if I just reduce Z a little bit, maybe I get something at least that's corrected in the right direction.



Okay now, these guys, Clemente and Raimondi, who worked at IBM in the early days, when IBM was the only place that had computers that were as powerful as your laptop, or almost as powerful as your laptop, they did things like this, where they did good quality calculations, better than this, and then tried to match the wave functions they got, by adjusting Z, so they looked like the wave functions that you get with better approximations. Right? So they got a best fit to better calculation. So they found that helium -- which has two electrons; so it has a nuclear charge of two; Z=2 -- you get electrons distributed, more or less right, if you pretend that the nuclear charge is 1.69, instead of 2. Okay? So each electron sort of what's called 'screens' the electron from the nucleus. Right? Okay, now in the case of carbon, which has Z=6, you can also use Zeff, but it's different depending on which electron you're talking about. Why would that be so? Why would it be that you'd use a different screening constant, or a different effective Z, if you're talking about the 1s orbital, than if you're talking about the 2s orbital say? Why would it be different, how much you want to reduce the nuclear charge? Sherwin?



Student: [Inaudible].



Professor Michael McBride: Yes, the electrons that are outside, you don't care about. It's the electrons that are between you and the nucleus that are screening you from the nucleus. Right? So a 1s is already really close to the nucleus. There's very little other electron density that's inside it. Right? That's why in helium it wasn't screened very much. But in the case of carbon, the 1s electrons, which are way down near the nucleus, have an effective charge that's almost 6; 5.67. But the 2s have 3.22. Right? The 1s electrons are all down inside and making the nuclear charge appear small. Right? And to a certain extent the other, the 2p electrons, are also inside; to a certain extent, but not as much inside as the 1s is. So 3.22. And the Zeff for 2p is different than for 2s; it's 3.14. What do you notice about those two numbers? You might think they'd be more or less the same; and they are more or less the same. But 2s is slightly less screened. It sees more of the nucleus than the 2p does. Okay? So it looks like 2s gets more down inside than 2p does. Does that surprise you? Russell?



Student: No, because the 2p orbital dumbbell doesn't have anything at all.



Professor Michael McBride: Yes, the dumbbell doesn't have anything at the nucleus. Right? It's got a node at the nucleus. Whereas 2s has that first little core, down in -- the stuff that's inside its first spherical node is down quite close to the nucleus. Now we could look at that. There's what 2s looks like, and here's the probability distribution that we talked about: r2 times the radial function squared. So this stuff that's in here is inside. If we compare it with the 2p orbital, you see that this part is way down inside. It's not being screened by the 2p electrons. Right? In fact, we could also look at the 1s. This is how the 1s is distributed. Right? So the 1s will screen this. And of course it screens this part of the 2s, and it screens this of the 2p. But it's not really trivial to look at this and figure out which one would be more screened, the 2s or the 2p, because this part is further in than this part, but this part is way down inside. So it's a balancing act as to exactly which one would be bigger, and they're not very different. But as it turned out, according to Clemente and Raimondi and their calculations, the Zeff that you should use for 2s is a little bit bigger than the one you should use for 2p. Kevin?



Student: Why are there two peaks for 2s instead of 1?



Professor Michael McBride: Because there's a radial node; remember 2-ρ, the function is 2-ρ. So here's ao; 2 ao, when the distance is two; two minus two is zero; there's a node there. Okay? And it's been squared of course. The wave function changes sign, but here we square it. Any other questions about this? Okay, now let's look at sodium. So sodium has a nuclear charge of eleven, and you won't be surprised to see that Zeff for the 1s electrons of sodium is 10.63; almost 11, they are very little screened. 2s is 6.57. 2p is 6.8. But 3s is only 2 and a half, because it's further out still. So there are more electrons inside, hiding the nucleus. Everybody got the idea of what's going on here? You'll notice one thing sort of funny here. Do you notice what?



[Students speak over one another]



Professor Michael McBride: John?



Student: The 2p is higher than --



Professor Michael McBride: Ah, it turns around! This time it's vice-versa. Right? So it's a balancing act, and who knows? And there's nothing fundamental about this. This was just adjusted by Clemente and Raimondi, in order to get shapes that looked pretty much like better quality shapes. Okay, so it's a very subtle thing. But this is very crude. Right? Nobody argues that it's the last word. So isn't there a better way to do it? And there is a better way to do it, to get back to orbitals, even for many-electron problems, and that's called Self-Consistent Field, or SCF. And it's a recipe for calculating orbitals; for calculating better orbitals than you would get with an effective-Z. So first you go through all the electrons in an atom, or a molecule, and you find an approximate form of the orbitals; for example, you could use Zeff, to get something that's sort of approximate. Okay? So you have approximate wave functions, square them; you have approximate distributions for all the electrons. Now with Zeff what we were pretending was that a certain fraction of the other electrons are on the nucleus, and the rest of them we forget; that's pretty crude. Okay?



Now what we're going to do is pretend we know, at least approximately, how the other electrons are distributed in a cloud; the other electrons. We're interested in one electron, an orbital. Right? Now how will knowing how the other electrons are distributed help us find the orbital we're interested in, the one-electron wave function? What do you need in order to solve a quantum mechanical problem? You need the mass of the electron -- that's easy. What else do you need? The potential law; what its energy is, its potential energy at different positions. But if you know where the nucleus is, or nuclei, and you know the cloud of the other electrons, and assume they're just static clouds, then you can -- it's laborious, you need a computer to do it -- but you can calculate the potential that the electron you're interested in would have, at different positions, if there were this fixed cloud of other electrons and the nuclei. So you have the potential law, which means you can find your one-electron orbital. Okay? Does everybody see the strategy here? So we're going to do it one electron at a time. So we fix all the other electrons, all but one, from some crummy estimate, like Zeff or something like that, calculate the potential for the one electron we're interested in, and then we get -- use that potential to calculate an orbital for that one electron that's better than if we had used Zeff. Right? Because it's not just putting a certain fraction of the other electrons at the nucleus, it's treating them as a cloud. Okay? So now we have a much better guess for that one electron than we would've had earlier with Zeff, or whatever type of guess. What do you do next?



Student: Do it for every orbital.



Professor Michael McBride: Dana?



Student: You do it for every other orbital, in sequence.



Professor Michael McBride: Ah, one at a time. You now know how that one is distributed in the cloud, much better. Right? And you take all the others but one, fix them in their approximate clouds. Now you calculate a potential for the second electron, and do that trick. Okay, so we repeat steps two and three to improve the orbital for another electron. Then what do we do? Will, what would you do, at this point?



Student: See how it changes shape?



Professor Michael McBride: Well yes, it'll change shape. That second electron will have a better shape now. So you've got that. You've got a good shape for the first electron. Now you've got an improved shape for the second electron. What do you do?



Student: Start from step one or the first electron.



Professor Michael McBride: Or you go, if there's -- what if there are three electrons?



Student: Do all three again.



Professor Michael McBride: No, you do the third one.



Student: Oh yes, definitely.



Professor Michael McBride: Right? And then the fourth, fifth, sixth, until you get through all the electrons. And then?



Student: Start over.



Professor Michael McBride: Start over. So you improve all the orbitals, one by one, and then you cycle back to improve the first one again; go through them all. Then what do you do?



[Students speak over one another]



Professor Michael McBride: Start again. Then what do you do? This is why it's good to have a computer, right? Then what do you do? When do you stop?



[Students speak over one another]



Student: When they stop changing.



Professor Michael McBride: They'll stop changing; at least -- it'll be like Erwin meets Goldilocks where it's out in eighth decimal place that things are changing. So then you know you've got it as close as is reasonable to go. So you stop. And what do you call it, when you get to that situation when you stop? The system is self-consistent. Right? That's why it's a self-consistent field. Okay? So you quit when the orbital steps -- shapes stop changing. So now you have the right wave functions, the right orbitals. So now we've got the real thing. Right, or wrong? What could be wrong? We've got it self-consistent. Where's the weakness in the assumption on which we've been doing this? Kevin?



Student: Well you're assuming fixed positions for the other ones.



Professor Michael McBride: I'm assuming what?



Student: You're assuming fixed positions for the other ones.



Professor Michael McBride: You're assuming that all those other electrons, except the one you're working on, are fixed in clouds. But they're not fixed. Right? The electrons can move around. So depending on where your electron happens to be, those other electrons may change their shape, at any given instant. Okay, it's still wrong, because real electrons are not fixed in clouds; they keep out of each other's way by correlating their motion. Right? They don't move independently so that this one is a cloud and when this moves around it, nothing happens. As this one moves, when it gets near this place, this one gets away. Okay? So they keep out of other ways by correlating their motion. So the true energy must be lower, more favorable, than you calculate by self-consistent field, because the orbitals are able to get away from where you think they should be. Is everybody clear on why the limit goes in that direction; why the true energy is lower? Because the electrons are smarter than you are. They know to get out of each other's way; or at least than you pretended. Okay? So what do you do? You hide the residual error so that people won't embarrass you by saying, "What the heck were you thinking?" Right? You say that when you go to the Hartree-Fock limit -- which is a fancy name, named after two people that thought of doing this self-consistent field thing, or thought of a method for doing it -- when you get to the Hartree-Fock limit, that's a self-consistent field, but there's going to be an error because of correlation. So you give it a fancy name so people will think, "ah, what in the heck is that?"; they must really know a lot. Right?



[Laughter]



Professor Michael McBride: So what do you call the error? Do you call it error? No way. Sophie, do you know what you call it?



Student: Correlation energy.



Professor Michael McBride: Correlation energy. Right? There is no such thing as correlation energy. That's not a fundamental energy. It's not like Coulomb's Law or gravity or something like that. Right? It's just the error you make when you do self-consistent field, that the true energy is lower. Now if you want to measure correlation energy, what do you have to know?



Student: True energy.



Professor Michael McBride: You got to know the true energy, so that you know how bad your estimate is. Right? And where do you get the correct energy, or the true electron density that you have an error in?



Student: Experiment.



Professor Michael McBride: You get it from experiment. Or you get it from some whopping calculation that's not as simple as SCF, Okay? For example, there's a thing called "configuration interaction", which we won't know anything about, and don't need to. Right? But it's a much more complicated calculation. It takes into account the fact that electrons keep out of each other's way. Or there's another one called "density functional theory", which is also approximate. They're all approximations. You can't solve the real equation. But these are better approximations than just self-consistent field. Okay? But they're hard to think about, because they involve so much manipulation that it's hard to reason about them. Self-consistent field is easier to understand. Okay? So if we're really lucky though, correlation energy might be negligible; it might be so small we don't care about it, in which case we're golden, for practical purposes. We can use orbitals, self-consistent field orbitals, treat things as if they were independent electrons. Okay? And then we're in business. Okay? So we should think about the magnitude of energies and whether we care about correlation energy, this error that is orbital theory. What do I mean by saying orbital theory? What's an orbital?



[Students speak over one another]



Professor Michael McBride: A one-electron wave function. But we're trying to understand many-electron problems. Can we analyze many-electron problems as sums of one-electrons, as sums of things that come from orbitals? Is the whole equal to the sum of the parts, the parts being orbitals? We know that orbitals have to be fundamentally wrong, but if the error is -- if the correlation energy is really small, we don't care. So should we care about the error in orbital theory? So here is a scale, a logarithmic scale, of energy changes that occur when things happen, and how big are these energy changes. Let's start at the beginning. So we take a bunch of neutrons and protons and bring them together to make a nucleus, and energy comes out. Right? And the amount of energy that is given off when a C-12 nucleus is formed is 2*109 kilocalories/mol. How do I know? Because I look at the mass, the rest mass of the proton and the neutron, and I look at the mass of C-12, and mass was lost when it came together. Right? And the amount of mass lost, E=mc2, is 0.1 atomic mass units, which is 2*109 kilocalories/mol. So that's a lot of energy. Okay? Now so we got C+6. Now we're going to put two electrons on it, the 1s electrons of carbon. Right? Bingo! And that gives us 2*104 kilocalories/mol, given off when that happens. Okay?



Now that is our old, familiar friend. What do you notice about the ratio of these energies? 105, right? So it's 105 smaller, like the ratio of my hair to the width of the room. Right? So much, much smaller energy involved in putting electrons into the 1s shell of carbon, than in putting the carbon nucleus together. Okay? Then we're going to put the four valence electrons onto carbon, right? The 2s and 2p. And with that we get another 3*103. So it's an order of magnitude less. The 1s electrons are bound much more strongly than the 2s and 2p; and you know that, from this scaling, Z2/n2. Okay? Because you could use Zeff to guess how the 1s's are affecting the energies -- the nuclear charge for the 2s's. So you could lower the nuclear charge from six to four, because there are already two electrons way down in there; and then you have that n2 as well. So you could scale the energy and find that it's an order of magnitude less. Okay.



So then what are we going to do next, now that we have atoms? Going to put them together to make bonds. Okay. So we make four single bonds from carbon, but they're to other carbons. So for this one carbon we should count only half of each energy, right? Because of half of it we'll assign to the other carbon. So half of four single bonds -- a single bond is order of magnitude 100 kilocalories/mol. So that's about 200 kilocalories/mol. So another order of magnitude down, the energy in making bonds. Okay?



And then we can have non-bonded contacts. Now there are different kinds of non-bonded interactions between molecules. But typically, to be worth talking about, they're in the range of one to twenty kilocalories/mol; so another order of magnitude down, or more. Okay, and the weakest of all, the weakest bond known, or interaction known, attractive interaction, is between two helium atoms. It turns out that two helium atoms, although they don't form a bond, are attractive; you know, things at long distance are attractive and then they become repulsive. So the minimum energy distance is fifty-two angstroms -- right? -- thirty-some times as long as a normal bond. And the depth of that well, how favorable is that energy, is 2*10-6 kilocalories/mol; so that's nothing.



But all these things we're looking at down below here are all based on Coulomb's Law. The first one was not. Right? The nuclear binding energy is not Coulomb's Law. But all these others are Coulomb's Law. But they're all 105, up to 1015, times weaker than what goes on in the nucleus. And that means if you made any error, at all, in the energy of the nucleus, it would completely wipe out anything that had to do with Coulomb's energy. Right? Because it's so much bigger. So an infinitesimal error in nuclear energy, or change in nuclear energy during a reaction, would completely wipe out everything that we're talking about. Right? So this sounds like we would really worry about it, except -- so a 0.001% change in nuclear energy would overwhelm everything Coulombic. Right? But fortunately nuclear energy doesn't care about chemistry. If you change from one arrangement of atoms to another, the nuclear energy doesn't change at all. Right? So why is that good? You can just cancel it out. The starting material and products of any reaction have exactly the same nuclear energy. So if you're not a physicist, you can just forget nuclear energy, even though it's so enormous. Okay? Is that clear? So forget nuclear energy, we don't have to worry about it. All we have to worry about are these Coulomb things.



Now, so black that out. Okay. Now here's how big correlation energy is. It depends on what problem you're dealing with, how big the molecule is, what kind of atoms there are in the molecule. The error you make in SCF is different, for different cases. But for the kind of cases we're interested in, normal organic molecules, it's of the order of 100 kilocalories/mol. Now that's an error. Do you care about that error? The error in nuclear energy, an error, would have been enormous and completely wiped anything out. Is this enormous? Do we care about an error of 100 kilocalories/mol? Do we? Would you care about an error of -- Devin, what do you say? Would you care if you made an error of 100 kilocalories/mol? How big is that in the scale of things we're talking about?



Student: Doesn't seem too big.



Professor Michael McBride: Compared to what?



Student: That's the question.



[Laughter]



Professor Michael McBride: Yes. What are we interested in?



Student: Bonds.



Professor Michael McBride: Bonds. How big is it compared to bonds?



Student: Fifty times as big.



Professor Michael McBride: It's as big as a bond. A bond is 100 kilocalories/mol. That means that we're -- that's a disaster, right? -- unless what? The nuclear would've been a disaster too. Why isn't it? Errors in nuclear energy are not a disaster. Why?



[Students speak over one another]



Professor Michael McBride: Because they don't change. The starting material and the product have the same. So the same thing would be true here. If correlation energy didn't change, from one arrangement of atoms to another arrangement of the same atoms, then it would just cancel out and you wouldn't care. Everybody got that idea? Okay, so the correlation energy is about equal to the magnitude of a bond. But how big are changes in correlation energy, when you change the arrangement of atoms? Well changes tend to be about 10 to 15% as big as bond energy, or 10 to 15% as big as correlation energy. So correlation energy does change. You make different errors for different arrangements of the same atoms. But those errors are about 10 to 15% as big as a bond. Now do you care about correlation? You don't care as much. So you'll get approximate ideas. If you don't care within -- if you're satisfied with getting about 80% of the right answer, then you don't care. Okay? So that implies that orbital theory is fine, as long as what you're interested in is answering qualitative, rather than fine quantitative questions. Okay?



So if you want to get numbers really right, and the magnitude of the number really makes a difference to you, a few percent change, then you have to do something better than use orbitals. Okay? But if you just want to understand why bonds work, then orbitals are fine. Okay? Because the changes in correlation energy aren't that big.



So, but for these properties, for non-bonded contacts, especially for helium-helium, the correlation is the only game in town often. That's what holds helium atoms together. Helium atoms are nuclei, positively charged; electrons, negatively charged. So nucleus repels nucleus; electrons repel electrons; nucleus attracts electrons; nucleus attracts electrons. Right? At first, at fifty-three angstroms, or fifty-two angstroms, far apart, those essentially cancel. Right? But the motion of the electron around this nucleus correlates with the motion of the electron around this nucleus. They tend to be in-phase with one another. So at any given time you have plus, minus, plus, minus, and they attract one another. So precisely what holds helium to helium is correlation. So if what you're interested in is bonding, then use orbitals, fine. But if you're interested in non-bonded interactions, then correlation can be a big problem. Okay? But we're talking about bonds now, not correlation.



Okay, so orbitals can't be true. That's clear, because electrons influence one another, they repel one another. If you have just one electron, fine; more electrons, orbitals can't be true. But still we'll use them, to understand bonding and structure and energy and reactivity. And we know we won't get precise values, we'll be off by one to ten kilocalories/mol, but we'll get insight that's very useful. Okay. Now what gives atomic orbitals their shape? Why does this particular orbital have a node, for example? We know Coulomb's Law tends to attract the electrons to the nucleus. Why does it have a node and spread out? Because of kinetic energy, right? This curvature of wave functions that's required to solve Schrödinger. So it's kinetic energy that gives them their shape. Or if you double the nuclear charge, the thing gets half as big. That's Coulomb's Law, sucking it in. Right? So the potential energy scales the radius through the formula for ρ; we've seen that. But the kinetic energy is what creates nodes. So the 2s has that spherical node. Or this orbital here has a conical, two cones, and a spherical node; it's the 4d orbital. So kinetic energy is what creates the shapes. The charge just scales things in and out.



Now if we use orbitals, how should we calculate the total electron density? Well we have two one-electron wave functions. We know the density of electron one, at this position, by squaring its wave function. We know the density of electron two at that position, by squaring its wave function. How do we get the total electron density at that same position? How would you get it? You know how much of electron one is there, its probability density. You know two. How do you get the total? Ilana? If you know how much of electron one is there, and you know how much of electron two is there, how much total electron density is there? How would you get it? Can't hear very well.



Student: You'd just add them together.



Professor Michael McBride: Yes, add them together. A whole is the sum of its parts. Okay? So the total density is the sum of those two squared wave functions; just right. But notice it's a sum; it's not a product. This is not a question of joint probability. It's not what's the probability that electron one and electron two are there at the same time? That's not the question. The question is, what's the total probability of finding any electron there? Okay? So it's a sum, it's not a product. Okay? Now we had this question you looked at before: How lumpy is the nitrogen atom? This picture is taken from a recent organic text and they had a fancy graphic program or an artist or something that drew something that looks very realistic. Right? But we can check it because we know the formulas. And if we want to get the total electron density, we add the electron density of the electron that's in the px and the one that's in the py and the one that's in the pz orbital. So we square them, and we get this, and we sum it up to get the total electron density; and it's some constant times x2+y2+z2 times e-ρ ; after we've squared. Right? And how can you simplify that? What's x2+y2+z2? It's r2. Okay? So how does it depend on θ? How does it depend on φ? Elizabeth?



Student: It doesn't.



Professor Michael McBride: It doesn't depend. What does it look like?



Student: A sphere.



Professor Michael McBride: It's a sphere. It's a ball. It doesn't look like this thing. So it's spherical. So forget, for all the elegance of that picture, forget it; they're not showing you the truth. Okay, or this problem we had before of looking at the cross-section of the CN Triple Bond, where we sliced it and turned it and thought it might look like a cloverleaf, or maybe some kind of diamond shape or something. But in fact it's round. Why? Because (2px)2 +(2py)2 depends on x2+y2, which is r2, in two dimensions. Right? So it's cylindrical, doesn't depend on the angle around the bond axis.



Okay, so this is what we've seen. We've seen three-dimensional reality, hydrogen-like atoms. We talked about hybridization. We saw that orbitals are fundamentally wrong but that we can recover from the orbital approximation and use it, if we -- for example, with self-consistent field; as long as we're not interested in getting the very finest energy but can be satisfied with approximation. Now we're going to move on to something more interesting, which is molecules. We're going to look first at Plum-Pudding molecular orbitals, and then understanding bonds, and overlap, and energy-match.



Now there are lots of different ways of looking at the electron distribution. You know this story, presumably, about different blind men doing experiments on an elephant and getting completely different ideas. The same is true of the electrons in a molecule. So here's the electron density in a hydrogen molecule. It's calculated, but there are ways to get that kind of information experimentally as well. So, and you know how contours work; it's a lot more dense near the origin, near the nuclei. So which contour do we want to look at in order to understand it? Well we can choose our contour, and we get different pictures, different understanding, depending on which contour we choose. For example, we could choose, way down, a very high electron density. Right? And then we see just a set of two atoms; it just looks like two atoms. Right? We don't see the fact that it's a molecule. Where have we done this before?



Student: [Inaudible].



Professor Michael McBride: Andrew? Shai? Pardon me?



Student: Difference density.



Professor Michael McBride: Not difference density.



Student: Well, we were subtracting to get the difference density.



Professor Michael McBride: Yes. We looked at total electron density; it just looked like atoms. We had to do the difference in order to see that it was not just atoms. It was essentially just atoms. So if you look at high density, you see just a set of atoms; no excitement there. So that's the molecule as a set of atoms, just a set of atoms. Okay?



Or we could take a somewhat lower electron density contour and we'd see that the atoms are a little bit distorted. Or we could do a difference map, as Shai says, in order to see that they're a little bit distorted, because bonding distorts the shape of the atoms a little bit. But first -- and we'll, this is what we're going to look at shortly -- but first I want to get some insight by looking at the very lowest electron density. So that would be molecules formed from a set of atoms, rather than as a set of atoms. Right? There are little changes because of the bonds.



But how about if we look way out there? Now, if you don't look really close, it looks spherical, it looks just like an atom, or almost like an atom. And if you went even further out, it would get spherical, for all you could tell. Okay? So that's the molecule looking like an atom, with whatever number of electrons the molecule has. So that's the molecule as one atom. The only difference is that the nucleus, which for an atom would be in the middle, has been split, to give two nuclei, which of course distorts the shape of the electrons a little bit, if you move -- cut the nucleus in two and split it out. So we want to look at a few molecules, from this point of view, as single atoms -- because we know about atoms now -- but distorted by the fact that the nuclei has been split apart. So this is nuclei embedded in a cloud of electrons.



Now what gave those electrons their shape, the shape of the cloud in which this thing is embedded; what gave them their shape? We just talked about this a short time ago. What gives orbitals their shape? Dana? Can't hear very well.



Student: A desire to be as far apart as possible?



Professor Michael McBride: That's potential energy, right. But that's not what gives -- that's not what creates nodes. That just causes things to spread out. What gives them their characteristic shape of nodes, planar nodes, spherical nodes and so on? Elizabeth?



Student: Kinetic energy.



Professor Michael McBride: It's the kinetic energy. And the same kinetic energy considerations will apply in a molecule as in an atom. It's always curvature of the wave function divided by the wave function. So the electrons are dispersed by electron repulsion, and noded, given nodes, by the kinetic energy; and the kinetic energy also causes them to spread out, as we've seen before. Okay, but you see then that. Although Thomson was wrong about the atom, it wasn't a plum pudding of nuclei [correction: electrons] embedded in a cloud of positive charge. But a molecule is a plum pudding! It's nuclei embedded in cloud of negative charge, but that cloud of negative charge is given its form by kinetic energy largely; also potential energy, as you say. So it's backwards from what Thomson thought. But his idea wasn't a silly one, it just happened to be wrong, for atoms. So here's a piece of plum pudding; which you know is like fruitcake. So how do the plums distort the pudding? Right? We know what atoms would look like now. Right? But if you split the nucleus into several nuclei, and moved them around as the plums in the pudding, they'll change the shape of the electrons. How? That's what we want to look at today.



Okay, so first we're going to look at methane and ammonia, and we want to understand them visually. Why do the orbitals in these molecules have the shapes they do? Okay, so there are four pairs of valence electrons -- there are also two core electrons, 1s electrons on carbon and on nitrogen. So we want to compare the molecular orbitals to the atomic orbitals of neon, which has the same number of electrons; four electron pairs with n=2 in the neon atom. So there are the same number of electrons as neon. So they should, at a big distance, if you look at very low contours, it should look like a Neon atom. But what does it look like as you get closer? Okay, so here's a 1s orbital of the Neon atom. Right? And here's the 1s orbital of methane; it's that net that's shown, the red net. Everybody see it? And here's the -- the contour level that we're drawing is where the orbital, you know, which level are we choosing? We're choosing 0.001 electrons/cubic angstrom is the level we've chosen to draw, of the onion. Right? And here it is for nitrogen. The blue one there is the lowest molecular orbital. But they're just little spheres, right? They're like the 1s of neon. Okay, so the core orbitals are like the 1s of the carbon or nitrogen atom. They're tightly held, they aren't distorted very much because the nuclei have split apart, right? And they're boring. So forget them, we won't talk about them anymore. We'll focus on the valence orbitals where the bonding action will occur. Now, there are eight valence electrons. That means two to an orbital. So there'll be four molecular orbitals that have electrons in them, in methane. And they are arranged in energy this way. There's a lowest energy, and then there are three that have exactly the same energy. Do you remember what we call those when several orbitals have the same energy?



Students: Degenerate.



Professor Michael McBride: Degenerate. Okay? So there are three degenerate molecular orbitals for CH4 that have exactly the same energy. In the case of ammonia it's a little different. There's one lowest one, and then there are three, but one of them is higher than the other two; there are only two degenerate ones there. Incidentally, what does that remind you of, to have one that's a little bit lower and then three that are the same energy? Have you ever seen that before?



Student: 2p and --



Professor Michael McBride: Ah! 2s and 2p, right? There's one 2s and three 2p's, in an atom. Right? Here's a molecule and there's one that's lower and then three that are equivalent. Okay, now there's the lowest one, and it's a 2s orbital. It's the neon's 2s orbital that's been distorted by taking four protons out of the nucleus and pulling them out to be where the hydrogens are. Okay? So you can see how this has been distorted, by distorting the electron cloud into the direction where the protons have been pulled out. Do you see that? Okay, now you don't see the -- you have to remove the ball there, that's the carbon or the nitrogen, to see that little spherical node. It's way down near the origin that made it a 2s orbital. Right? That's the spherical node. And it doesn't look like a sphere because of the algorithm the computer uses to draw this, connecting dots with straight lines. Right? But it is more or less spherical; a little bit distorted by the fact that you've pulled these things out. Okay, now what do these molecular -- these are molecular orbitals. But that, you see, is a 2px orbital, that's been a little bit distorted. Right?



Notice that the difference between the CH4 and the NH3, there was a proton that went up in CH4, which pulled them up. But if you didn't have that, it didn't go up as high. Right? But then essentially this is the 2px orbital of the molecule, distorted from the atom by pulling protons out of the nucleus. Okay? Or that one, what's that one? It's 2py, but you have to rotate it to see it. If we rotate it 90 degrees here, you can see that it's a 2py orbital. And again, the CH4 is distorted at the top because two protons came out and stretched it out. Same thing in nitrogen. Now we go to the third of these 2p orbitals, which is different in nitrogen; for the nitrogen case, higher in energy. There they are. Why is that orbital higher in energy than the others in Nitrogen? Why isn't it so good? Because you have electrons up in that red region that don't have a proton there; it's not being stabilized. Here, the electron density that went up is around a proton. There the top lobe of the p orbital doesn't have a proton stabilizing it. So it's higher in energy. And it turns out to be more reactive. What do you call it?



[Students speak over one another]



Professor Michael McBride: That's the unshared pair. The high-energy reactive electrons are the ones that don't have a proton stabilizing them there. Now there are also vacant orbitals. Right? Remember, you could have any number of orbitals. We're looking at the ones that you can make from the 2s and 2p. But we have a number of atomic orbitals that we can use. We're looking at the lowest set, the ones that come just from 1s orbitals of Hydrogen, and 2s and 2p orbitals of carbon or nitrogen. But there are eight such orbitals: four, 2s and three 2p's, on the central atom; four on the hydrogen for CH4. So there'll be a number of other combinations you can make from mixing those, which won't have electrons to go in them. But let's look at them. So here are the next four orbitals made from the valence level orbitals of the atoms involved. So what you see on CH4 that this is the 3s orbital. It has a radial node. Right? Or pardon me, yes, a radial or a spherical node here, that surrounds the sphere. So it's red sign inside, blue sign outside.



The same thing is true of the NH3, but you have to rotate in order to see it. That's the lowest vacant, the lowest unoccupied molecular orbital. And if we rotate it here, you can see the inner part. And bear in mind that there are two spherical nodes. There's one that's really tiny, that's inside anything we can see here. Okay, then there's this orbital, which is the 3dx2-y2, looked at from a funny angle where it's not very clear what it is. But if we rotate all these by 90°, you can see that it's like that cross, or whatever you call it, of the 3dx2-y2. But why doesn't it look just like that? Why are they distorted? Why does it have this funny U-shape here. Instead of this lobe, this lobe, here and here, the lobes moved up there. Why? Angela? Pardon me?



Student: There are protons.



Professor Michael McBride: That's where the protons pulled them, the potential energy. But the fundamental shape, before it got distorted by the potential energy of where the protons are, the fundamental shape was from the kinetic energy. It was like the 3dx2-y2. Or how about this one? There you can see -- but it's clearer still if we rotate it -- that again it's that X-shaped thing. But the ones on top have been pulled out by the protons. Or this one. That's the one that has a doughnut around the middle. But the doughnut has been pulled down by the three hydrogens that aren't on the z axis. Or there it is, rotated. Okay? Now I hoped to get -- let's just start ethane and methanol, and then we can get to more interesting things in it next time. It has a lot more pairs of electrons, ethane and methanol. So we can compare those MOs to the AOs of Argon, which has the same number of electron pairs. Actually I should quit now and let you get to your next appointment.



[end of transcript]

Lecture 12
Overlap and Atom-Pair Bonds
Play Video
Overlap and Atom-Pair Bonds


This lecture begins by applying the united-atom "plum-pudding" view of molecular orbitals, introduced in the previous lecture, to more complex molecules. It then introduces the more utilitarian concept of localized pairwise bonding between atoms. Formulating an atom-pair molecular orbital as the sum of atomic orbitals creates an electron difference density through the cross product that enters upon squaring a sum. This "overlap" term is the key to bonding. The hydrogen molecule is used to illustrate how close a simple sum of atomic orbitals comes to matching reality, especially when the atomic orbitals are allowed to hybridize.




Transcript



October 1, 2008




Professor Michael McBride: Okay, let's get started. So last time we looked at methane and ammonia, and saw something interesting about molecular orbitals: we could give them the same name as atomic orbitals; that is, if we, especially if we look at a low electron density contour of the molecular orbitals, we see that what it looks like is an atom where the nucleus is split into pieces. Right? Now that splitting into pieces changes the potential energy for the electron. So you expect that to distort the orbital. The electrons will move in the directions that pieces of the nucleus have gone. But then, in addition to potential energy controlling the shape of orbitals, kinetic energy also does. And that's the same thing as it is in an atom. You have a thing with no nodes, a thing with one node -- and when you have one node you can have either a spherical node, or a planar node, and there can be three planar nodes; so a 2s and three 2p's. Exactly the same considerations apply in a molecular orbital as in an atomic orbital. You have the kinetic energy, which comes with curvature, which comes with nodes.



Now, as you split the nucleus up and pull pieces in different directions, it doesn't have the same symmetry it had when it was all together in the nucleus, a spherical kind of symmetry. So the nodes get distorted. But still you can see them there. And we saw them last time and went through all the occupied and vacant valance orbitals of ammonia and methane, and saw how they looked like atomic orbitals. This, not surprisingly, because it's so fundamental, the potential energy and the kinetic energy, applies to every system. It applies to you, viewed as a single atom, right? With a zillion electrons. Okay? But pieces have moved around, so the orbitals change. We'll look at two more complicated cases and then we'll get on to a different way of looking at bonding. So we'll look at ethane and methanol. And we use -- I didn't tell you last time, explicitly, where I got the molecular orbitals from. I got them from my laptop. There's a program -- the particular one is called Spartan, that I use -- and it calculates what molecular orbitals look like, using approximate molecular orbital (that is, Schrödinger equation kind of) theory.



Okay, so let's just look at the ones of ethane and methanol. Now both of these have seven pairs of valence electrons. There are also core electrons, and if we were looking at the orbitals for all the electrons, we'd include those. And exactly how we're going to count those -- you could do it one way or the other -- whether you consider the core electrons, the 1s electrons to just be part of the nucleus and then treat the rest as the electrons you're interested in; or whether you want to count the core electrons too. You can do it either way, but what you analogize to what depends on whether you count them as part of the nuclei. So anyhow, we're going to compare these molecular orbitals to the atomic orbitals of argon, which has also seven electron pairs. Okay, so there's the 2s orbital. I'm going to start with that, because I'm going to pretend that that red part is the core -- that there's a 1s which is core electrons. But in this case -- here's my pedantic note on this subject which I just added. So if you have -- before we had just one heavy atom, carbon in methane, nitrogen in ammonia. Right? So there was one 1s orbital. Right?



Now we have two heavy atoms, carbon and carbon in methane, carbon and oxygen in methanol. So there are two heavy atoms and therefore two boring core orbitals. So for purposes of making analogies, we'll use the atomic 1s orbital, of the atom that we're analogizing things to, to stand for all the molecular core orbitals. You can do it any way you want to. We're not really interested in that. We're interested in the valence orbitals. So whether I start with the 1s or the 2s depends on how I'm handling the core electrons; it's not a big deal. Anyhow, let's pretend it's the 2s of argon here, and we're going to compare it with this lowest valence level molecular orbital of ethane.



Now we can -- as I do this, on the left side of the pictures I'm going to show one view, and then, because it's more complicated than methane and ammonia were, I'm going to also show a picture rotated by ninety degrees. So I'm going to rotate around that axis, and on the right I'll show a different view of the same orbital. Okay, so that's the lowest valence level molecular orbital of ethane. And it's not very exciting; it's just a distorted sphere. And you can see the way in which it's distorted. It's distorted vertically, because the two carbons are pulled apart; so it's got sort of a narrow waist to it. And then it's pulled out, where each of the protons left the middle atom to come out and be hydrogens. Okay? Now if we look at methanol, it'll be a little different. Can you anticipate how it might be different if, on the bottom we'll have CH3 again, but on the top, instead of having CH3, we're going to have OH. How do you think it might be different, in the way it looks?



Student: It might be skewed more towards the OH because oxygen is --



Professor Michael McBride: Okay, so oxygen has a bigger nuclear charge. So that's going to pull the lowest energy orbital toward the oxygen. That'll be one thing. We expect it to be bigger, top-heavy. Okay? What else, about how the top will look? How will it be different from if it were a CH3? Yes, Alex? I can't hear very well.



Student: Oxygen has lone pairs.



Professor Michael McBride: Oxygen has lone pairs. Now, how is that going to change things? Or another way of saying the same thing is it has only one hydrogen up there. How's that going to change it, do you think? You have to speak up.



Student: It's not going to be symmetrical.



Professor Michael McBride: It's not going to be symmetrical. Which way is it going to be distorted to be unsymmetrical? You have to speak up.



Student: Towards the electron pair.



Professor Michael McBride: Toward the electron pair did you say? I couldn't hear.



Student: Away from hydrogen.



Professor Michael McBride: Away from hydrogen? Well how about -- the proton goes out; does it bring electrons with it, or does it repel electrons, the proton? Speak up.



Student: It brings them.



Professor Michael McBride: It pulls, so it should distort toward the hydrogen. So here's what it looks like. Right? It's top-heavy, as Angela said, and it's distorted out toward the hydrogen; there are no protons pulling it to the top left. And you see the same thing end-on there, on the right. Okay, this is the next orbital. What does that look like? Obviously you can peek in the middle and see. It's obviously a 2pz orbital, with the node across the middle. Okay? In both cases. Now the ethane case is symmetrical. The other case is unsymmetrical. Why is it unsymmetrical? Because the first orbital pulled electrons to the top. Right? So the next- in the next orbital, electrons aren't going to want to be at the top so much, because they're going to be repelled by the other electrons; so there'll be more toward the bottom, because the first ones went to the top. Okay? Okay, then here's the next orbital. You can see there are energies marching up here. The lowest one was s, then 2pz, now 2px, if we define the horizontal axis here on the left as the x axis. So again, it's what you expect, and it's pulled out, stretched vertically from being a dumbbell by where the nuclei went. And here's 2py, which we can see more clearly on the right, perpendicular to the 2px.



Okay? And notice that it's obviously so, that the methanol orbital will be less symmetric, in all cases, than the one for ethane. But still we can recognize the nodes, because they must go that way. It must be no nodes, one node of three different kinds, and so on. Okay, now this is 3s. You notice it has -- the node that we don't see, the one that's down near the nuclei, in fact two nodes down near the nuclei, one for each of the heavy atoms. But then this now has another spherical node, or approximately spherical node. So we have that extra red lump in the middle on the top, or blue, in the middle. Just to review, what's the difference between red and blue? On the top, the computer decided to draw it with red in the middle. On the bottom it decided to draw it with blue in the middle. What do those colors mean? Yes, Cathy?



Student: Positive and negative signs.



Professor Michael McBride: The mean positive and negative. So which one is right? Should it be positive in the middle or negative in the middle?



Student: Positive.



Professor Michael McBride: Positive is a guess. Wilson, what do you say?



Student: It doesn't matter.



Professor Michael McBride: Why not?



Student: Because it's not really positive or negative; it's just kind of a phase of it.



Professor Michael McBride: Yes, it's the sign of the wave function. But you can multiply the wave function by minus one, any constant, and it's just as good as it was before. So it's arbitrary, and the computer was arbitrary in choosing the colors. I think there's actually a function that I could've changed it, if I'd thought to do so, so they'd be the same. But actually it tells a story if I leave them different. Okay, so here's the next one. Now this is -- if you look on the left now, it's more clear where the nodes are, that that's a dxz orbital. The name 'xz' means that the product of x and z appears in the wave function. Right? So when both x and z are positive, then the product is positive; on the top right, red. Right? When x is negative, and z is positive, the top left, it's negative, the product of them. So that's why it has the name xz.



Okay, and there's dyz, which you see on the right, turned ninety degrees. And here's the dz2, which is that thing that has a doughnut that goes around the middle. Right? It's hard to see the doughnut. Can you see? It's blue on top. You can easily see the red on the top left, which is what's blue in the middle; the sign has changed. Right? But the doughnut is highly distorted, because as you go around, first a proton on the top pulls it up, then a proton on the bottom pulls it down, then up, down. So it's like a crown, the doughnut has been made into a crown around the end, by the protons pulling in that way. Okay. Then here's the 3pz. So it has the horizontal nodal plane, but also it has a spherical node, which you can see in either picture really. Okay?



Then here's the 3p y orbital, which again has that spherical node, but now the planar node, or approximately planar node, is vertical instead of horizontal. So say on the top right there you have red on the right and blue on the left. There's a vertical node over which it changes sign, going right to left. Or here's the 3px. Or here's the 3dxy; which you don't see so well here; well, but if you turned it down you would. And here's the 3dx^2-y^2; which again, to see it well, you'd have to turn it. And here's the 4f orbital. So you can see, especially say in the top ones, compared with the atomic orbital, that it's exactly the same general pattern of nodes, slightly distorted.



And incidentally, remember all the n equals whatever, n=3; all the orbitals had the same energy before. Now they don't have all the same energy. Notice that they all have different energies. None of them are degenerate. Why? Because if you broke the nucleus apart, how stable a thing that has that general shape, like a dumbbell or a cloverleaf or something, how stable the electrons are in those lumps depends on whether a proton got pulled into the lump. If there happens to be a proton in the lump, where kinetic energy, the node pattern, wants it to be, then that'll be unusually stable. If the protons have been pulled someplace where there is a node, because of kinetic energy, then it won't be stabilized. So that breaks the degeneracy of these different patterns that have the same number of nodes. It depends on where the -- how the potential energy changed.



Okay, now just finally and very quickly, I want to look at 1-flouoroethanol, which is a molecule that would have, I think, no stability at all, but you can put it into the -- as a practical matter, you're never going to put it in a bottle, but you can easily put it into the computer and calculate what its molecular orbitals would look like. And I did it to show you something that's very, very unsymmetrical, and has atoms of very different nuclear charge. Okay? So what will the very, very, very lowest orbital look like, do you think, for this thing? It has a fluorine, an oxygen, two carbons and five hydrogens. So what do you think? If you were an electron, and you had that set of nuclei, arranged this way, where would you want to go? Elizabeth?



Student: Very biased to the left, especially around the fluorine.



Professor Michael McBride: Yes. Now what would the very, very, very lowest one? It should be more toward the left, and really close to the fluorine. It would be a 1s orbital, and mostly on fluorine, because that's where most of the protons are, the most concentrated protons. So let's look. There's a smaller scale ball model, so that we can see really small orbitals, and there that's, you're absolutely right, it's essentially the 1s orbital of fluorine, is the very lowest orbital. What would be next? Suppose this one, you came up and this seat was already taken, now where would you go, the next electron? Zack?



Student: Oxygen.



Professor Michael McBride: To the oxygen, because that's the next highest concentration of protons. And next? Where will you go next? Steve, what do you say? The 1s of fluorine is taken. The 1s of oxygen is taken. Where do we want to go next? Pardon me?



Student: One of the carbons.



Professor Michael McBride: Ah, which carbon?



Student: The left carbon.



Professor Michael McBride: Why?



Student: Because it's closer to the fluorine than the oxygen.



Professor Michael McBride: And what does the proximity have to do with it? The proximity means that the electrons that are on that atom will have been drawn toward -- and we'll talk about this more later -- will have been drawn toward the fluorine and the oxygen. Therefore that atom will have fewer electrons, or lower electron density. Right? Therefore, it's a better place for other electrons to go. Okay? So you'd expect the next one to be the middle carbon. Right! And finally, of course, it's going to be the second carbon, the one that's remote from the electronegative, high nuclear charge atoms. Okay, now we've done all the 1s's. So we can look at what's interesting, the valence orbitals, the ones that are going to be involved in bonding; these don't have anything to do with bonding. Okay, so there's the first one. Russell, tell me something about the shape of this? Does it surprise you?



Student: No. It's highly distorted towards the fluorine.



Professor Michael McBride: Yes, it has no nodes; except it actually has little nodes down near the nucleus, because it's actually more like a 2s orbital in that respect, because of the core electrons. So it has tiny nodes that we don't see around the nuclei. But it has no nodes within the valence; big orbitals that we're looking at. So it's like an s orbital. Right? And it's big where the nuclear charge is big, as you say. Okay? What's the next one going to look like? Any ideas? Pardon me?



Student: Towards the oxygen.



Professor Michael McBride: Toward the oxygen. Will it also have no nodes? No. The next highest orbital has to have a node. Where do you think the node will be? What orbital will it look like, sort of? It'll be highly distorted. But what -- if this is the s orbital, right? What's going to be next? Elizabeth?



Student: It's as if it's switched but with a planar node in between.



Professor Michael McBride: Oh let's see. Right! So the first one was -- you were right Russell that this one should be big on oxygen. But notice it's like a p orbital, it has that horizontal node, because it's higher kinetic energy. Okay? And then this one, that's on the carbons, right? Mostly. But it's a py, it has a vertical node. So it's negative say on the fluorine and oxygen and positive on the carbons. Or if we rotate this one around a horizontal axis to look at it from the side, it looks like that, and you can see that it's a py kind of orbital. Elizabeth?



Student: I'm just asking a point of clarification. These aren't technically p orbitals, right? We're just saying they're analogous to them.



Professor Michael McBride: They're analogous to p orbitals; because it has to be the same. The orbitals must go in order of the number of nodes. Right? So the same thing is in an atom or in a molecule. You go from no nodes to one node. And there are three ways of getting one node: spherical or distorted sphere, planes, or distorted planes if the nuclei are pulled around; and so on. So they're really very much like the atomic orbitals. Okay, or this one. Now that's an interesting one because that's like a hybrid orbital. It looks like the orbital up here. It's a mixture of s with p. Did everybody see how that is? It's a big sort of blue lobe on the top and a small red one on the bottom. And also the next one is the other one, the hybrid that points the other direction, another combination of the s and the p, that points down. Okay? And then this one looks like dxy again. Right? It's sort of a cloverleaf with two nodes. Okay, so that's all I want to do with that, and we won't go any further with it. This is just an interesting way of looking at how -- that molecular orbitals are really just like atomic orbitals, and have energies for the same reason, except the potential energy gets screwed up by breaking the nucleus and pulling pieces around, but in an understandable way, and the nodes get distorted because of this.



Okay now, now we're getting into really -- we just looked at this view, at the single united atom view. But the other view is the one that's going to be more generalizable, and that's the one where we looked at bonding. Right? So you have to probe a little harder to get a qualitative understanding of what chemical bonds are. And that's what we're going to do now by choosing a higher contour with which to look at a molecule. Now, true molecular orbitals, to the extent that orbitals are true all together -- why aren't they true all together; why aren't orbitals true all together? Yes, Alex?



Student: Because you have multiple electrons.



Professor Michael McBride: Because you have many electrons; you can't have independent electrons, you can't have orbitals. But we're approximating things by orbitals, trying to take electron interaction into account in a sort of a left-handed way by Self-Consistent Field, or something like that. Because it's much easier if we can divide the whole into a bunch of parts, each of which we can understand. So, to the extent that molecular orbitals are true -- the kinds of things I've just been showing you, calculated with my laptop -- they extend over the whole molecule; they're not local. Right? Except like the 1s of fluorine was local, but mostly they go over the whole molecule. But bonds are thought of, and have always been thought of, as interactions between pairs of atoms. So we want to divide things completely differently and look at the bonds now, at pairwise LCAO molecular orbitals. Now what's an LCAO? It's a sum, or a linear combination. Right? A weighted sum of atomic orbitals. So here's an example. Ψ, which is an orbital -- what's it a function of?



Student: Position.



Professor Michael McBride: Position of what?



Student: One electron.



Professor Michael McBride: One electron. So it's a function of x1,y1,z1; we're talking just about electron one. So the wave function for electron one we say is 1/√2 times one atomic orbital plus another atomic orbital. Right? Now, have you ever seen adding orbitals like that before? That's what hybridization is; we added s and p. But this is different, because when we added s and p before, they were on the same nucleus, and we did it to get a new orbital for that particular nucleus for that atom; to distort it one way or the other, for example, or to rotate a p orbital. Right? But this is very different, because we're adding orbitals that are on different nuclei: A, nucleus A, and nucleus B. See the difference? Adding is just -- wave functions are numbers, we just add the numbers. But in the first case, hybridization, those two functions were on the same nucleus. Now they're on different nuclei, what we're adding together. Okay, now why is it sensible to think that you might get pairwise molecular orbitals that can be expressed like this? How do you interpret an orbital? Corey? What good is an orbital? What do you use it for?



Student: It's a one-electron wave function.



Professor Michael McBride: Well what do you use it for?



Student: For probability.



Professor Michael McBride: And how do you get probability? From the wave function -- if you have the wave function, how do you get the probability density?



Student: You square it.



Professor Michael McBride: You square it. So do you see why we have a 1/√2 in this? Because when we square it, that's going to be half, and we're going to get half of atomic orbital A squared, and of atomic orbital B squared; so it's a half of each of them. That's why we have 1/√2. So let's go on with this. Suppose we have a hydrogen molecule, and suppose that the nuclei are at a great distance from one another. So far they don't interact, or negligible interaction; they're very far apart. What would you expect the lowest energy, one-electron wave function to look like? One possibility is that the electron could sit exactly halfway between the two nuclei? Right? Is that a reasonable place, is that the low energy place for it to be? Lucas, what do you say?



Student: No, because the added probability density there is not the greatest.



Professor Michael McBride: Why not?



Student: It needs to be one or the other.



Professor Michael McBride: Why shouldn't the electrons sit -- if the two nuclei are this far apart; and I don't mean two angstroms apart, I mean two meters apart -- there's a proton here and a proton here. Is the electron most likely to be here, halfway between?



Student: No.



Professor Michael McBride: Where would it be most likely?



Student: Probably nearer to one of the two atoms.



Professor Michael McBride: And which of the two?



[Students speak over one another]



Professor Michael McBride: Suppose you took the long view. Suppose you averaged it over eighteen zillion millennia -- time averaged. Sometimes it would be near this one, sometimes it would be near this one. Okay? What would it be if you took a really long view?



Student: Both.



Professor Michael McBride: And half here, and half here. So the wave function, when you square it, you want it to be half looking like this atom, and half looking like this atom. Then you see the time average. Right? It might take a long time to achieve that average, because it'd take a long time for the electron to tunnel two meters, right? But in the very, very long time it would look like that. So we know what we want it to look like. What we want is that the probability density, the square of this one-electron wave function, should look half of the time like the atomic orbital A squared and half of the time like atomic orbital B squared. So on time average it's half of one and half of the other. Everybody with me? So that's what the -- yes, Nate?



Student: Why isn't there a 2A -- or a 2AB?



Professor Michael McBride: Oh, because I'm telling you what it looks like. It's got to look half like this and half like this. Right? So the square of it has to be this. So all we have to do to find the wave function is what, if we know what its square is? All we got to do is take the square root and we've got the wave function. Bingo! There's the square root. Right? So that is a reasonable way to write the wave function, 1/√2 (AOA + AOB). Claire, you have a question.



Student: From my small understanding, with the math that I've got -- and this may be wrong, correct me if I'm wrong. But don't you, if you have a bracket and two things inside it, you square outside of it, don't you have four things that come out of it.



Student: [Inaudible].



Student: You can't just put the square --



Professor Michael McBride: You want that?



Student: Yes.



[Laughter]



Professor Michael McBride: Okay, now Claire you're going to help me out. How big is -- that's a number. It's the product of -- atomic orbital A assigns numbers everywhere in space, everywhere in space. Atomic orbital B assigns numbers everywhere in space. So at some point in space atomic orbital B assigns a number and atomic orbital A assigns a number; and A times B is the product of those two numbers. How big is that product? How big is atomic orbital A here? This, a meter away from the proton?



Student: Not very big.



Professor Michael McBride: And how big is atomic orbital B there?



Student: Not very big.



Professor Michael McBride: Now how big is atomic orbital A here?



Student: Very big.



Professor Michael McBride: Okay. So now, how big an error do we make if we neglect A times B? Where do we make an error? Do we make an error here?



Student: No.



Professor Michael McBride: Do we make an error here?



Student: No.



Professor Michael McBride: Do we make an error here?



Student: Yes.



Professor Michael McBride: No! We make an error; yes, sure enough, we make an error. How big is the error?



Student: Not very big.



Professor Michael McBride: Ah, it's negligible, because it's at great distance. Okay, so at great distance we can forget that. So now it's easier to take the square root, right? Okay? The old fox up here, huh? [Laughter] Okay, now your problem is what happens if H2 is at the bonding distance? What if they're only an Angstrom apart? Now there should be a problem, because A times B is not going to be negligible everywhere now. Okay, so now that's going to come back. So now we got an error, right? Or do we? Let's think what that does. Okay, so we if approximate the molecular orbital as the sum of atomic orbitals, this way, then it looks very good near the nuclei. Because near A it looks like A; near B it looks like B. And if we want to square it to find the electron density, we do this. But if we then subtract, what the atoms would give for electron density. Now what does this remind you of, where we look at the total electron density and subtract the atoms?



Student: Difference density.



Professor Michael McBride: So we're actually looking for the difference density. Everybody with me on this? So we're going to subtract the atoms, which is ½A2 and ½B2. Right? If we subtract, what do we get? What do we get for a difference density? We get A times B. So we get the difference electron density, which is due to overlap. And what do I mean by overlap? I mean that is only important in regions where both of them have a finite value. Right? The product of A and B is negligible if A is zero; it's negligible if B is very, very small. Right? So it's only where they overlap, the two functions have simultaneous values, that you care. Okay, now that thing, the thing that's the bonding, the difference density, is really a byproduct -- that's a little bit of a pun because it's a product -- but it's a byproduct of squaring the sum -- of what Claire didn't like about it. So the very thing you didn't like is what's going to give rise to bonding density. Isn't that neat?



Okay, but notice that here we're multiplying A times B. But this is a completely different instance of multiplying from what we had before. Right? Before we multiplied two orbitals to try to get a two-electron wave function. This has nothing to do with this, because both of these are functions of the same electron; it's like one electron that we're squaring here. So this A times B, this product, this overlap, comes from the squaring. It was when we squared it that we got that. Okay? Now, because we have this extra term, we have not only ½(A2+B2), which would -- what's the probability of A2 summed over all space, or integrated, if it's normalized?



Students: One.



Professor Michael McBride: And this one?



Students: One.



Professor Michael McBride: And what's this whole quantity?



Students: One.



Professor Michael McBride: One, because of the half. But actually we've got it bigger than that, because we added the overlap term to it. Right? So it's actually not going to be half; it'll have to be something a little less than half, so that it'll sum up to one, if we want to normalize it. Okay? So there's going to be those less than halves there. Okay. Now what does that do? That shifts electron density, right? We're taking electron density away from where the nuclei are, from A2 and B2, and where's the electron density going? Because we're using less than half to begin with, we're taking electron density away. There's going to be -- we're subtracting more than was there at the beginning. Right? Which means that we're going to have negative electron density in the difference map. Electrons are going away from the atoms. Where are they going to?



Student: Into it.



Professor Michael McBride: Into the region where there's overlap. Right? So they go away from there, into the overlap region. So this is just like what we were seeing with X-ray. Okay, so that overlap, the A times B term, is what creates bonding. And we've seen this before. Remember when you have wells far apart, the wave function is the sum of the two; we saw this in one-dimension. Right? But if they come close together, you get a wave function that looks like that, which we looked before, just from the point of view of the energy, and saw that that would stabilize the particle, because it's got less curvature, less kinetic energy. Right? But also the electron density grows in the middle. Right?



So from the point of view of the electron distribution, that was the glue holding the atoms together. So it's held together, both because the energy goes down and because you put this glue in the middle, which is what causes the energy to go down. Okay? So that's bonding. And remember we also had this. So as the energy went up in the middle one, the energy is lower [correction: higher] here than it was in the atoms apart. So the nuclei push one another apart now, without the glue in the middle, and that was anti-bonding. So we've seen it before in one-dimension, but it's true in three-dimensions as well.



Now let's think about this again. So here's atom A. Now where is the square of that function significant? Is it significant there? No. Is it significant there? Yes, it's a little bit significant at least. How about there? A little bit. Right? Okay. Now suppose we have another atomic orbital there. Now, where is the product significant, of A times B? Okay? So is the product A times B significant there? No. Is it significant there? Nick, what do you say?



Student: No.



Professor Michael McBride: Why not? Speak up.



Student: It's very small near A.



Professor Michael McBride: What's very small?



Student: The value of A.



Professor Michael McBride: No, no. The value of A there is something -- we said that before, when we were looking only at A -- it's not very big but there's a significant value. But how about the product? Josh?



Student: The value of B is very small.



Professor Michael McBride: The value of B is very small there. So the product is going to be very small. How about there?



Student: There's going to be a change.



Professor Michael McBride: Ah, now they're equal; halfway between they're equal. So both of them are a little bit small, but their product is still significant. Right? Only in this region, where they overlap, is that product significant. Okay. So at the center, notice that the number ΨA assigns and the number ΨB assigns are the same number. So 2(ΨA ΨB) is as large as (ΨA)2+(ΨB)2; because ΨA times ΨB is the same as (ΨA)2. So the electron density is nearly doubled in the middle from what it would've been if it had just been two atoms. So that's the region of significant overlap, and that's what we care about. So the overlap integral, summing this product -- or integrating it -- over all that space, that's a certain density. Right? We squared in order to get that. Right? That's part of the density. So we sum that, the density that comes from that product, over all space, and that's called the overlap integral. If the atoms are very far apart, the overlap integral is essentially zero. If the atoms are close together, the overlap orbital will be finite, and the better the -- the more the orbitals overlap, the bigger the overlap integral, obviously. And that measures the net change that arises on bonding, the difference density, as we've just seen.



Now let's look at some theoretical examples here. So let's look at the total electron density as calculated for two -- adding two 1s orbitals of hydrogen at the appropriate distance for H2. This was this was done forty years ago and published in the Israel Journal of Chemistry. Okay now, and here's -- so on the left we have the total electron density that you'd calculate from that. A very simple thing: 1/√2(A+B). And you square it, and you get the density, and that's the density, contoured at 0.025 electrons/cubic ao, the unit of distance. Okay, now, on the right is the difference density. So from that thing on the left we've subtracted the atomic orbital, the atomic electron densities. And you see exactly what you expect. It builds up in the middle, where there's overlap, and at the expense of the atoms. Yes Russell?



Student: Shouldn't it be H2+ ion?



Professor Michael McBride: Oh, I think it's the H2 molecule, I'm sure it's the H2. They aren't so fantastically different, because the two electrons are in the same orbital; two electrons can be in the same orbital. So one'll be twice as big as the other. Qualitatively they'll look very similar. And I think this is the H2 molecule; but it might be the ion, I'm not sure. Okay, at any rate the contours on the right are much smaller. Remember, difference density is much smaller than total density. So what you see is that it's contoured at 0.004. So you have one, two, three, four, five contours. So you get up, in the middle, to 0.02 electrons per cubic angstrom [correction: ao]. That's how much bonding has changed things, at the maximum. Okay, now the energy that's calculated, with this very, very, very crude wave function, just one half -- 1/√2 times the sum of the two atomic orbitals. The energy, you calculate that, is 92.9% of the true energy. That's pretty darn good, right? But almost all of that energy that you calculate was already present in the separate atoms. We're not interested in the energy of the separate atoms, we're interested in how much it changes when you make a bond, which is a small difference between large numbers. So it turns out that although we're within 7% of the true total energy, this simple model only calculates about 50% of the change in energy that came from putting them together, right, which is much smaller.



Okay, so high accuracy is required to calculate a correct value of the bond energy. This simple thing won't do it. Well it's in the right direction, you're halfway there, so it's a pretty good start. Right? But to do the difference, as in the same way you needed high precision to do X-ray difference maps, you need better orbitals than this, if you want to calculate good bond energies. So you need to make the orbitals better. Okay? Now -- so but already we can take heart that the very crudest model shows most, 52%, of the energy of the bond, and it shows the electron density building up by 0.02 electrons per cubic bohr radius. And what we saw qualitatively was there was a shift from the atom to the bond, of electron density. Okay, now we can adjust the molecular orbital to get a better approximation of the true thing.



How will we know when we've adjusted it and it's gotten better? If we adjust it and get a lower, calculate a lower average energy -- I should've said a lower average energy, because if we don't have a true wave function we get different values for the total energy; the kinetic won't exactly offset the potential as you move from place to place. But if you get the lowest average energy, then that is, by definition almost, more realistic, because you can easily prove that the true energy is the lowest possible energy; that makes a certain amount of sense. The lowest possible calculated energy is the true energy. So if you change your wave function and get a lower average energy, you're closer to the truth. Okay? That's called the Variational Principle. Okay, so here we've changed the form of the molecular orbital. And how did we change it? What does a 1s orbital look like? Kate, do you remember the form for a 1s atomic orbital? I can't hear.



Student: A sphere.



Professor Michael McBride: Angularly it's a sphere. How does it change as you go out, do you remember? How does it depend on ρ? The same way they all depend on ρ. Anybody remember?



[Students speak over one another]



Professor Michael McBride: e-ρ. So it falls off exponentially at a certain rate. And that rate, how fast it falls off, is determined by the nuclear charge. Okay? Now one way to change that shape would be change how fast it falls off. Right? It wouldn't be correct for the atom anymore. We've already got the correct solution for the atom. But we could change the shape of the thing by changing how fast that exponent falls off. And we could vary that, in the molecule, using one half -- 1/√2(A+B). But those A+B are no longer true atomic functions, they're a little fatter, a little skinnier. Okay? And we can change how fat or skinny it is, until we get the lowest molecular energy. See, that's a way you can vary it and find the best value. And that's what was done here to optimize the exponent. And now you get a total electron density that looks essentially the same as it did before. And if you look at the difference density, how has it changed, if you do this? First, notice that the energy got lower. We're now to 73% of the lowering of the bond energy. So the total energy's gotten lower, it's better.



And how has the electron density changed? It got higher in the middle. Because what we did was spread the exponent out a little bit, so you had more overlap in the middle. Okay? So this wouldn't have been good for the single atom, to spread it out, but it gives a better function for the molecule. And it's still very, very simple. And what you see it did is it increases the bonding density and the bonding strength. You get a larger shift from the atoms to the bond. Now, how else could you change the shape of the atomic orbital in order to increase the overlap; some way other than making a single exponential and having it get fatter or thinner? Can you think of some other way? Here you have an atomic orbital, a sphere, and you want to change its shape so that it overlaps better over here. Right? How could you change the shape of an atomic orbital, without doing really gross damage to it, making it a cube or something like that, or a line? How could you change it so it looked pretty much still like an atom did, has a lot of the virtues of the atom, but is shifted over here? Sam? I can't hear.



Student: Can't you just allow the electrons to shift?



Professor Michael McBride: Yes, how am I going to write a function that allows the electrons to shift in the direction I want them to? Lexy?



Student: You could hybridize it.



Professor Michael McBride: Hybridize it! We could hybridize to shift the electrons. So that's the next one. So here we're going to -- instead of this we're going to hybridize it. Now, this particular calculation did the hybridization and also did a little self-consistent field calculation. And the hybridization left it 96.7% 1s. So essentially it's still a normal 1s orbital. But they added .6% of 2s, which expanded it a little bit, because 2s is -- goes further out than 1s. And they added 2.7% of 2p. Now why was 2p much more helpful than 2s? Lucas?



Student: It has those lobes.



Professor Michael McBride: Yes, so it takes density from one side and shifts it to the other. In fact, this is precisely what we saw before. Notice what it did was -- how much it increased the density in the middle. And notice now, it's not taking electron density away from the nuclei, which was a good place for electrons to be. Where does it take it away from?



Student: The left one.



Professor Michael McBride: Out beyond the nuclei, and shift it to the middle. So even before, when it was an atom -- here's the nucleus, a certain distance out here, and a certain distance out here, were the same in energy, for that atom. Now we've taken it from a place which is -- out here, and put it here. Right? Great idea. Well done Lexy. So there's what it looked like. Remember, if it's 100 percent at 1s, it looks like that, and if you change it to be hybridized that way, with Atom-in-a-Box, it looks like this. Right? It's not much shift, but it shifts from the left to the right, and gives better overlap. Okay, but it requires the atom to be a little bit less happy as an atom, because it's partly 2s and 2p now. The electrons are further from the nucleus, but you make up for that by having a better bond. Okay, now notice that didn't change the energy very much. It went from 73% to 76%. Right? We were already about the right energy, but it changed the density a lot, to be what we want it to be. Okay, so there's a much bigger shift, and it's now from beyond the nucleus into the bond. Right? And now we're going to do the last thing. We've done SCF already, but now you have to do a higher level calculation that'll do correlation, take the correlation of the electrons into account. And if you add some correlation calculation to this, you now get 90% of the bond. If you did complete correlation, you'd get 100% of the bond. What does it mean to do a calculation with complete correlation? It means you don't have an error anymore, you've used a really good calculation. So already at this level, that was done fifty years ago almost, you get 90% of the bond. And the bond density notice, what about the electron density?



Student: No change.



Professor Michael McBride: It hasn't really changed. All that happened was that the electrons kept apart from one another, but the average density was the same. So already, with just that hybridization, we got very close to the truth in electron distribution, and three-quarters of the way to the truth in how strong a bond is. So not a bad approximation. So hybridization, to give better overlap, is a great thing. So the density wasn't changed, but you got a much better energy. And how so? Because the electrons kept apart from one another, when you allow correlation.



Okay, so here's a pairwise atomic orbital: < 1/√2(A+B); and you can use hybridized orbitals. And the virtues are it's very easy to formulate and to understand. And it looks like atoms, especially when you get down near the nuclei. And you don't want that to change because that's the main event for electrons. You get much more energy forming atoms, as we saw before, than you do making new bonds, once you have the atoms already. Okay. It builds up electron density between the nuclei, through overlap, which is the source of bonding. It smoothes Ψ, to lower the kinetic energy. And then there's a pedantic footnote here that actually there's a thing called the Virial Theorem, which I'm not going to stress you with, but that little bit there, it happens to be true. But still we have the proper understanding of what's going on. Hybridizing AOs provides flexibility that gives you better overlap. And if you use all the H-like Atomic Orbitals, you have perfect flexibility, you can make any shape you want.



Okay, but we're going to keep it simple, use only 2s and 2p orbitals to hybridize, because that'll get you most of the way there and it's much simpler, rather than to try to mix 5f orbitals into it also. Okay, so that's great. And when we square it, we get this, which has the overlap part that helps us out. We have the atoms, plus the bond, which is the overlap, that product part. But we could've done the same thing to get the same product, as far as the atoms go, if we'd used a minus sign instead of a plus sign in combining things; although then we would've changed the sign of the overlap thing. It would become minus. So we would change it from being less than to being greater than, in order to have it be normalized at the bottom. Right? And that's the anti-bond. So we get both the bond and the anti-bond by doing this. And now we're going to go on, next time, to overlap, which we've already introduced, and also the concept of energy-match. And when you put these two together, you'll really understand bonding.



[end of transcript]

Lecture 13
Overlap and Energy-Match
Play Video
Overlap and Energy-Match


Professor McBride uses this lecture to show that covalent bonding depends primarily on two factors: orbital overlap and energy-match. First he discusses how overlap depends on hybridization; then how bond strength depends on the number of shared electrons. In this way quantum mechanics shows that Coulomb's law answers Newton's query about what "makes the Particles of Bodies stick together by very strong Attractions." Energy mismatch between the constituent orbitals is shown to weaken the influence of their overlap. The predictions of this theory are confirmed experimentally by measuring the bond strengths of H-H and H-F during heterolysis and homolysis.




Transcript



October 3, 2008



Professor Michael McBride: Okay, let's get started. The stuff we're doing these few days are really the focus of the first half of the semester. So pay attention and think about it. Okay, so the topics today are Overlap and Energy-Match, as determinants of how strong bonds are. So we'll start with overlap. We saw last time that overlap is what creates the difference density, what focuses some small amount of the electron density between the atoms and serves to hold them together. The overlap integral is the total amount of that product of A and B. Remember, you square (A+B) and you get A2 + B2 + 2AB. That 2AB term is the overlap. It's called overlap because it has values when both functions have appreciable magnitude, at the same point in space; they overlap. And summing that over all space, or integrating it, gives what's called the overlap integral. It'll depend on the distance. Obviously if they're very far apart, they don't give simultaneous values to the same point in space, because it's far from either one or the other, or from both. So it depends on distance. It also depends on hybridization, and that's what we're going to talk about first.



Okay, first let's just look for scale. Here are the 2s orbitals of carbon atom, and you can work out, with your formulas, that the diameter of the spherical node is 0.7 angstroms. Right? That means we have a distance scale here. Now, the distance between two carbons is roughly 1.4, 1.5, when they're forming a bond from one another -- between one another. Okay. So that means we can use that 0.7 angstrom scale to see how far we have to slide these together, in order to get them to bonded distance. If we superimposed the X's -- remember those diameters are each 0.7 -- so if we slide them together until we superimpose the X's, then they're 1.4 angstroms apart. So that's about the distance that these two orbitals are, in the proper scale, when they're in a carbon-carbon bond. Now, how big is the overlap integral, the sum of the product of A times B, over all space? Now I want you to guess, guess how big that integral is. And we have a hint. What would the overlap integral be if we moved them until they superimposed on one another exactly? That's the maximum we could get, if the distance were zero. So if pulled them all together, then A and B would be measured from the same center. So it would be the same thing as A2. Okay? So what would the integral of A2 be, over all space?



Student: One.



Professor Michael McBride: One; that's normalization. Okay, so the maximum value you can get when they're right on top of one another -- negatives multiply negatives to get positive; positive areas, multiply positive areas to get positive. Everything is optimum, is going to be one. Now I want somebody to give me an idea, a guess, of how big it's going to be here, or at least some considerations that would go into that. The maximum value you can get is one. How big do you think it would be here? What's something that would hurt it; that is, make it less than one? Russell?



Student: When there's very little overlap on the opposite sides.



Professor Michael McBride: Yeah, out beyond the red areas, there's very little overlap altogether. So you lose whatever overlap you'd get from there, in doing the normalization. Now, in the middle, right in the middle, you have blue on top of blue. That's good. But notice there's also a node that's going to interfere with some of the wave functions' overlapping, and there's going to be a little bit of blue on top of red. Everybody got that? So I just want people to guess. Okay, I know what I'll do. We'll do a poll. This is a democracy after all; we're about to get into November. So I'm going to start with 0.1, as what you think it would be. Right? The maximum, if you put them right on top of one another, is one. So I'm going to start with 0.1, and move up and see hands.



Okay, how many people think it's 0.1? 0.2? 0.3? 0.4? 0.5? 0.6? 7, 8, 9, one? Obviously not one. Now, I think, my sense of this is it peaked between 0.3 and 0.4, but there were some votes for 0.2, and I would've voted for 0.2, because it looks like you've lost an awful lot here. But in fact I personally am surprised that it's pretty big. It's 0.41. It's almost half as big as it would be if they were right on top of one another, despite the fact that you're losing great parts of it, that they don't overlap at all, and there's negative overlap where red's on top of blue. Still the fact is it's 0.41; so pretty big. Now that's going to depend on distance obviously. Right? So here's how it depends on distance. There's that point; at 1.4 angstroms it's 0.41. And I've put on here for reference the distance of a carbon-carbon single bond, the distance of a carbon-carbon double bond, and a carbon-carbon triple bond. And zero, notice, would be way down, out in the hallway someplace. Right? We're only looking at the region that's carbon-carbon bonds -- or maybe just about to the far wall. Okay? Here's half an angstrom; we got to go two more halves. Right? Okay, so here's how -- how do you think -- what should the line look like? Should it be sine wave? Should it be increasing as we go to the left? Should it be decreasing as we go to the left? What do you think? Some guess.



[Students speak over one another]



Professor Michael McBride: Should probably decrease. Because we know it's going to approach one. Right? And there it is. Yes, it decreases -- pardon me, it increases as you go to the left, and ultimately it'll go up to one, when you get out to the hallway. Okay? So that's s with s. Now let's look at some other orbitals that might overlap, and see how big their overlap integrals would be. Okay, suppose you have s on one, overlapping with p on the other. Now that p is called, for this purpose, π; π is the Greek version of p; in fact, in Greek you pronounce it 'pee,' not 'pie'. Right? So that's the Greek letter p. Am I right about that?



Student: It depends on classical or modern Greek.



Professor Michael McBride: Okay, which way is it?



Student: It'd be 'pee' in modern Greek; it'd be 'pie' in classical.



Professor Michael McBride: Uh-huh, okay, good. So anyhow, 'pie.'



[Laughter]



Professor Michael McBride: Now, so that symbol is used to talk about molecular orbitals, or parts of molecular orbitals; where p is used to talk about atomic orbitals. They correspond to one another. Now what is it that characterizes a p atomic orbital? Has a nodal plane. And in the same way a π orbital has a nodal plane that contains the nuclei that are in question. Right? So you see, if those two came together there'd be a nodal plane in the orbital on the right that contains the two nuclei. Okay? So that's called a π orbital. Okay? Now, but notice the contributions to the overlap. Below that plane, you're going to have blue overlapping with blue. So that's going to be a positive contribution. But above, you're going to have blue overlapping with red; that's going to be a negative contribution, and it's symmetrical. So what will the total be? What will the total overlap integral be; the sum of all the overlaps?



Student: One.



Professor Michael McBride: One?



[Students speak over one another]



Professor Michael McBride: It'll be zero, because for every contribution on the bottom that's positive, there'll be one on the top that's the same value and is negative. So it's going to be zero. So the one on the left is called σ. Now σ is the Greek equivalence of 's.' Right? So it means, for a molecular orbital, what s meant for an atomic orbital; there are no nodes that contain the nucleus. Okay? Okay, so if you have σ overlapping with a π, there's going to be this symmetry that causes a cancellation. So you can never get any overlap, by overlapping a σ orbital with a π orbital. Does that make sense to everyone? It does make sense, doesn't make sense? If you have one that doesn't have a horizontal node, that's the same top and bottom, and one that does have a node that's plus on the top, minus on the bottom, then you're going to get this cancellation everywhere, and zero. And that's called ‘orthogonal'. Okay, so you won't get any net interaction there. Okay, but if you turn the p orbital so that it is also σ -- notice it has a plane, a nodal plane, that contains that nucleus, but it doesn't contain the other nucleus you're interested in. So to call σ and π, we need a plane that contains both nuclei. That'll create this symmetry that cancels things out. Okay, so this one will give overlap. Now what would you expect for the trend of this one, as you go from great distance to small distance? If they're very, very far apart, how big will the overlap integral be; like ten meters apart?



Student: Zero.



Professor Michael McBride: Zero. Okay? Now, as you bring it together, as they start coming together, will you get positive or negative overlap?



Students: Positive.



Professor Michael McBride: Positive; right, the blue will overlap with the blue. And you keep coming, it'll get bigger and bigger. And then what will happen?



Student: [Inaudible].



Professor Michael McBride: Then the red is going to start coming in and -- well the blue is going to start, the blue on the right, from the p orbital, will start -- and going to overlap that red part on the left. And then, as it slides on over, till they get exactly on top of one another -- when they get exactly on top of one another, what will it be?



Students: Zero.



Professor Michael McBride: Zero, because it'll be canceling right to left. Okay, so here's a plot of that, for s-p σ overlap. Right? It's almost the same as s with s. In fact, it's a little bigger at great distance, a little smaller at small distance; except, at very small distance what will happen?



[Students speak over one another]



Professor Michael McBride: When you approach zero? It'll go to zero. Whereas the other one went to one. Right? The s with s went to one. So, but in the region we're interested in, the carbon-carbon bond distances, it's about the same. Okay, now how about pσ with pσ? Well we won't go through an elaborate guessing game. If they're very far apart it'll be zero. How about if they were on top of one another? It's obviously going to grow as blue gets on top of blue, and then as they get really close, the blue up from the right will be on top of the red, and the red on top of the blue. Right? So what will it approach? Zack, what do you say?



Student: Zero.



Professor Michael McBride: Approach zero? What will it be when they're right on top of one another? So red on blue here; blue on red here. Will the overlap be zero?



Student: Negative one.



Professor Michael McBride: Pardon me?



Student: Negative one.



Professor Michael McBride: It'll be minus one, because it would be like the orbital on top of itself -- that would be one -- except you've changed the sign, of one of them. Right? So every contribution will be negative. So that one starts there, and it's heading toward minus one. So it's going down. Okay? So that's pσ with pσ. Can you think of any other way we should talk about it? We've talked about s with s. We saw that s with pπ was orthogonal; s with pσ we've got there; pσ with pσ. Any others we need to think about?



Student: pσ with pπ?



Professor Michael McBride: How about pσ with pπ? Can anybody guess that one? Pσ is the same sign, top and bottom; pπ is opposite signs, top and bottom.



Student: So it's zero.



Professor Michael McBride: That's going to be zero. There'll be orthogonal. So have we done it now, or is there anything else to think about?



Student: π, π.



Professor Michael McBride: We could do pπ with pπ. Okay, so there's pπ with pπ. Okay? Now at great distance they'll be nothing. We'll bring them together, they'll begin to overlap. What will it approach as we go to zero?



Student: Zero.



Professor Michael McBride: Zero? If they're on top of one another, the product will be zero?



[Students speak over one another]



Professor Michael McBride: It'll be one, right? It'll be the orbital with itself, normalized. So that'll be one. Right? But now which do you think is going to be bigger, pσ with pσ, or pπ with pπ? Any guesses? Well, I'll give you the answer. Right? pπ with pπ is approaching one. At big distance, it's smaller than σ. Right? But at small distance, it must get much bigger, and in this region it's crossing; so it's bigger. Okay, so there are the kinds of overlaps of simple 2p orbitals. Now, there's a curiosity here. Over most of this range, more than half of the range, 2s overlaps with 2pσ better than either 2s with 2s or 2pσ with 2pσ. That just seems curious to me. You'd think that one of them would have a shape that gives better overlap. Right? So two of those together would do the best job. But it's not that way. It's that an s with a p is the best, over most of this range. I just think that's curious, and it has an important implication, which you'll see right now. What if we use hybrid orbitals, that are partly s and partly p? Now, since this one is partly s, it's s plus a certain amount of p. This one's s with a certain amount of p. The overlap will be s here with s here; s here with p here; s here with p here; and p here with p here. Right? There'll be four contributions: s with s; p with p; s with p; p with s. Right? So if we took sp3, for example -- and we can make, remember, four such orbitals -- then you get that. It's better than any of the pure hybrids. Right? And the reason is that it has s with s, and p with p; and also s with p, that comes in as a bonus, twice. Okay, now how about if we use, instead of sp3, how about if we use sp2? Remember what happens as we approach sp? The orbitals extend more. So what do you expect?



Student: More overlap.



Professor Michael McBride: More overlap. So there's sp2 with sp2, a little better. How about sp with sp? Better still. Could we get any better? How about s2p with s2p, or sp½ with sp½? Right? It's a little better at short distance, and a little not so good at long distance. So it optimizes around sp hybridization. That gives the best overlap. Why do we think now you want good overlap? Why would it be good to have good overlap?



Student: Keep the bonds strong.



Professor Michael McBride: Because that's what's building up the difference density, as we saw last time. Right? That's putting the glue in the orbital. Okay, so there's an interesting observation here, that hybrids overlap about twice as well as pure orbitals. Right? All the hybrids are roughly twice as good as the pure ones. Okay? And they're not very, very different from one another. But sp is about the best, over much of this range. The trouble with sp is you can only make two of them, because each involves 50% of s, and you only have one s to go in it; you have one s and three p's. So if you want to make just one orbital, fine, use sp, because you'll get good -- only one bond from the carbon; fine use sp and get the best overlap. Or maybe even s2p, depending on the distance. Right? But if you want to make more bonds, more than two bonds, then you're going to have to cut back on the s in each bond. Because you only have 100% to use, you can't use more than 50% in more than two bonds you're making. Okay. And what you see here is that sp3 isn't much worse than sp2. So why not make four -- right? -- because you can. Okay, anyhow, that's the hybridization of carbon. So, and the reason they do this, as I already said, is because they allow nearly full measure of s with p overlap, plus s with s, and p with p, when you have mixtures. So that all depends on the fact that s-pσ is so good in its overlap. Now, so sp gives the best overlap, but only allows two orbitals, with 50% in each; sp3 gives four, and nearly as much overlap. Okay?



Now we're going to look at the influence of overlap on molecular orbital energy. But we're going to use Erwin Meets Goldilocks, just for familiarity. So we're working here in just one-dimension, to get the idea of it. And we'll also then think in more general terms. And we've already done this, so it's partly review. So we'll assume that there's perfect energy match; that is, the two atoms we're talking about. Of course, atoms would be Coulombic potentials, not harmonic oscillator, Hooke's Law potentials. Right? But the idea of a double-minimum is the same. So we'll suppose that the atoms are equivalent. So the bottom of the well is the same on both sides. And guess an energy. And lo and behold we were right; we got the solution, and you've seen that one. Okay? But you can also get that one, one with one node. Right? And on the far left you can see that they have the same curvature-over-amplitude. So they have essentially the same energy. If the green energy is right, the red energy is the same deal. Right? They're ‘degenerate' orbitals, because they're far apart. Okay? No significant energy difference.



Now, suppose we increase the overlap. What will happen if we shorten the distance? You did this on a problem set with Erwin. Right? So you bring them closer together, and you have the one with no nodes, and the one with one node. And now there's a little -- right halfway between there's more electron density in one than there is in the other. Right? The antibonding one has zero and the bonding one has some buildup of electron density in between. But the energy hasn't changed visibly on this scale that we're using. It looks just the same. Okay? Now we increase it more and now there's a big difference in the middle. Right? And now the energies are beginning to change. The unfavorable orbital is going up and the favorable bonding orbital is going down in energy. And if we get it still closer together, we can see that there's now a big difference between them. Okay? So that increasing overlap creates the splitting that we talk about.



This is just review, as you know, but it's worth -- it's so important that it's good to mention it again. Now that overlap holds the atoms together. And this again is a picture we saw before, except we're going to do it with writing-in the equations. Remember, a good guess of the form of the molecular orbital is 1/√2 (A+B), or 1/√2 (A-B). And that will give two new orbitals, one better and one worse. And, in fact, if you think about normalization, it should be a little less than 1/√2, in the one that's going to have +2AB in it, when you square it, and a little less [correction: more] than 1/√2 when it's going to have a minus sign for the 2AB term. Okay? Now, because of that ‘a little bit greater' and ‘a little bit less', the energies aren't quite, as we saw them before, equally split, up and down. The one that goes down doesn't go down quite as much, because this is less than, and the one that goes up, goes up a little more because it's greater than. The mathematical requirement for that isn't immediately obvious to you, but it's true.



Okay, so there are the new orbitals you get when you bring the atoms together. Okay? And if you had more overlap, there'd be a bigger difference. So that's why overlap is important. The more overlap, the more splitting. We just saw that in one-dimension with Erwin Meets Goldilocks. Now, how many electrons do we have to put in these orbitals? That's going to make a big difference. So suppose we have just one electron for these two orbitals. When we bring the two wells together, when we bring the two atoms together, that electron will go down in energy, to the new molecular orbital. So that's going to be good. And if we tried to pull them apart, the electron would have to go back up in energy, and that would require energy, and that would oppose the breaking. So that's the bond strength. Right? You have to put an electron up that much. Now suppose there were two electrons. Now a second one goes down the same way. So it's even stronger. It won't, in truth, be quite twice as strong. Do you see why? Yes?



Student: Electron repulsion.



Professor Michael McBride: Because the electrons will repel one another. We haven't taken that into account when we think about the energies here. But it'll be certainly stronger, considerably stronger. Okay. Now suppose we had three electrons. What's going to happen?



Student: [inaudible].



Professor Michael McBride: One of them is going to have to go up instead of down. And, in fact, it will go up further than either of the others went down, even neglecting electron repulsion, right? Because of that 'greater than/less than' thing that shifted the pair up a little bit. Okay? So that's going to be even weaker than having one electron to hold them together, when you have three. How about if you have four?



Student: No more bonds.



Professor Michael McBride: Wilson?



Student: They'll be antibonding, so --



Professor Michael McBride: Yeah, you're now definitely losing, because the two that go up, go up more than the two that go down, go down. So that's bad. So we can summarize that here, the effect on the bond. If you have one electron, it's bonding; two electrons is strongly bonding; three electrons is weakly bonding, probably; and four electrons will definitely be antibonding. So that's why helium bounces off helium, until it gets fifty-two angstroms apart, and then there's this very, very, very weak bond, that we talked about, due to the correlation of the electrons. Right? But as soon as you get overlap with helium, it's repulsive, when you start to split these levels. Okay, now let's look at the other end. Why doesn't increasing overlap, as you bring things closer together, why doesn't that make these plum-puddings collapse? The electrons are getting more and more stable as you increase the overlap by bringing things closer together. So why doesn't it just collapse? In fact, the electron -- if a hydrogen molecule collapsed to helium -- it would be a funny helium because it would have no neutrons -- but if it collapsed, the electrons, indeed, would become 55% more stable. The electrons in helium are lower in energy than the electrons in H2. Right? So that would tend to pull them together, all the way, to distance zero. Why don't they? Elizabeth?



Student: The protons repel.



Professor Michael McBride: Because the protons will repel one another. That depends on 1/r. Right? And that increases much more dramatically. In fact, the amount by which the electrons become more stable is 650 kcal/mole, or would become more stable. Right? But the nuclear repulsion increases by that much by the time they get to 0.3 angstroms, let alone the last little bit, when 1/r gets enormous. Okay? So it's not worth it; unless you have glue to hold the neutrons together. Right? So on the sun you can take deuterium and fuse it into -- to make a helium -- you have neutrons there to hold the protons together -- and then you get 200 million kcal/mole; which is, of course, nuclear energy, not electronic energy. It is possible, but not under the chemical circumstances that we're talking about. Okay, so here's the Morse potential. Yes, Lucas?



Student: Have we been taking proton-proton repulsion into account before, with all our energies?



Professor Michael McBride: No, or only sort of subliminally. Right? Because if you take -- at first glance there's going to be nuclear-nuclear repulsion, and electron-electron repulsion, between the atoms; nuclear-electron attraction; nuclear-electron attraction. So to first approximation, those will all cancel out. Right? But then you get -- on top of that you have a little change from the fact that the nuclear-nuclear goes as 1/r and gets quite big, when they get really close, and that the electrons are coming down because of overlap. And we've been focusing on this overlap thing that causes the bonding, but obviously things aren't going to collapse because of the nuclear part. But we haven't been talking about it explicitly; you're right. Okay, so anyhow this is the form of the bonding potential, the Morse potential, which was just cobbled together artificially to make it convenient to calculate things and to have reasonable properties for a bond. But now we understand why it should look like that. This first part, the attractive part, is because the electron pair becomes more stable with increasing overlap, as the atoms come together. And then the nuclear repulsion becomes dominant, when it becomes too close. Right? So all this is Coulomb's Law. There's nothing special. There's not ‘correlation energy', which is some completely new physical phenomenon. Right? It's all Coulombic. Correlation energy is just, remember, an excuse for the error you're making, or a way to hide it.



Okay, so it all comes from Coulomb's Law for potential energy. But the funny thing was the kinetic energy of the electrons, according to Schrödinger. Right? Because as you make things -- as you concentrate things, you change the curvature as well. So you have to take that into account. But this curve provides the potential for studying molecular vibration; that is, once you know the form of this curve, then you can use quantum mechanics for the motion of the atoms -- not of the electrons but of the atoms -- and figure out how atoms vibrate; which we've already done, when we were doing Erwin Meets Goldilocks. Okay, but here, finally we understand the atom-atom force law. Right? That's what we've been doing the whole time, for four weeks or five weeks or whatever it's been, is trying to understand where -- what the force law is. Wouldn't Newton be happy? Remember, he said, "There are therefore agents in Nature able to make particles of bodies stick together by very strong attractions, and it's the business of Experimental Philosophy to find them out." And now we've found it out. Now, can we use this understanding then to gain greater purchase on chemistry?



So we've looked at overlap. But there's one other thing we have to look at, in order to apply it, and that's what we call energy-match. I don't think anybody else talks about energy-match, but we do. Okay, so here's what we talked about already. Right? And we have this splitting, which is about proportional to overlap. It's not numerically strictly proportional to overlap, but it certainly is close to proportional to overlap. Okay, now what would happen if the two atoms you started with were not the same, if one was a better place for electrons to be than the other; like there for example? So why would you -- remember what happened is you mix these. You get A+B and A-B. But why would you put any A together with B, if B is already much better than A? If they're equal, you don't care where it is, then fine, use a 50/50 mixture. But if one's better than the other, why use any of the inferior one? Why not just use the one that's good?



So we already actually saw this in the case of the 1s atomic orbitals, which did remain pure and unmixed, when we made the molecular orbitals of that funny molecule last time, the fluoroethanol. Remember, it looked like that. One of them was just an s orbital on F; another one was an s orbital on O; s orbital on one carbon, s orbital on the other carbon. The didn't mix. Right? So the compact 1s core orbitals did remain pure and unmixed, but the valence level atomic orbitals were heavily mixed. For example, the next one was biggest on fluorine, but also substantial on oxygen, and carbon even got into the act a little bit. Right? Now why is it that the core orbitals didn't mix and the valence orbitals did mix? What's different? The core orbitals are very small, very compact. These orbitals are bigger. So what happens? They overlap. Right? So it's overlap that causes things to mix. If there's no overlap they don't mix. If they overlap they mix.



So why use any of an inferior orbital? Well suppose the energy of A is much higher, less favorable, than that of B. Right? So now we make a sum, a weighted sum, a parts of A, b parts of B, and we square it. So we get a2 parts of the A density, b2 parts of the B density, and also this overlap term, 2ab. Now, can you profit from shifting electron density toward the inter-nuclear AB region, without paying too much of the high energy cost of A? That is, A2 is higher energy than B2. Right? The electron density around A is higher energy than the electron density about B. And if you mix A with B, you're going to get a certain amount of that bad, or less good, distribution, A2. But at the same time you're going to get some of AB, which puts stuff in the middle, at the expense of outside. That's good. Right? At a certain distance from the nuclei, you'd rather be between the two nuclei than out beyond one of them. So as you increase a, you increase -- oops sorry [technical glitch] -- you increase this bad part, but you also increase that good part. Can it be worthwhile to make a non-0? Which one is going to help?



But notice this. If you put only a small amount of A in, then the amount of A2 probability density, this bit here, is a2 parts of that; really, really tiny, if a is tiny. But the amount of ab you put in is a times b, not a2. Right? So it's much bigger. So you get this good stuff, without getting very much of this bad stuff. Right? For example, suppose you used, a was 0.03 and b was 0.98. So you square them, to find out how much -- [technical adjustment] -- you square them, and you find out that the amount of a2 you get is only 0.001. The amount of b2 is 0.96. And the amount of 2ab is 0.06. So you're getting sixty times as much of ab as you get of a2. So that's why it can be useful to put a little bit in. And if you're pedantic, you can look at the footnote there.



So let's look at Erwin Meets Goldilocks in the case where the two wells are not matched, where the energies are a little bit different; because that's what we're talking about. So here are non-degenerate atomic orbitals. So the one on the left is a little bit lower than the one on the right, but only a teeny bit. Okay? Now we're going to look at the wave functions here. So there's the lowest energy wave function. Right? And notice it's the sum, it's a weighted sum, with no nodes. But there's almost no mixing at all, just a really, really, really tiny amount of A in it. So essentially it's just the solution you'd get in the left well. Right? There'd be one with one node as well. What will it look like, can you guess? This is the one with zero nodes. What's the wave function with one node look like? Any guesses? Yes, Josh?



Student: It'll just have a node in the middle of that first potential and then be straight on the second one.



Professor Michael McBride: And be what?



Student: And be sort of straight on the second one.



Professor Michael McBride: Right. So the first one, the lowest one, is the one on the left. The second one, the next highest energy orbital, is the one on the right, with just a little bit of the one on the left in it; negligible amount of mixing. Right? And their energies, you might think their energies would be the same, because they're not mixing. But the green one is in this well, and the red one is in that well, and that well's higher. Okay? So it's got to be higher in energy. That spitting is due only to the original offset between the wells. There's no shift of it. Right? It's just you have A or you have B, and you have what you would expect for them. Now suppose we increase the overlap between these. What do you think is going to happen? You should get a little more mixing because AB is going to count for more, if you have overlap. Okay, so here they are close together. You still don't have very much mixing, just a little bit, and you haven't seen any appreciable change in the energies. You still have that same splitting that's due to the original offset. Okay?



Now we'll bring them still closer together. And now it's getting worthwhile to put it in the wrong well, because the AB term is so important. Right? So the antibonding energy is rising, the bonding energy is falling -- that's why they're antibonding and bonding -- and if we bring them still closer together, we get that. Right? Which, if you looked at it casually, they look like a single well. But notice that they're still quite unsymmetrical, they're still biased. The one without any nodes is still mostly in the left well, and the one with a node is mostly in the right well. But they're heavily mixed. Right? So what's happening is this increasing overlap, as you bring them together, is fighting the effect of the difference in the two wells. So when we had the core electrons in fluoroethanol, if we had tiny orbitals that didn't overlap, then they stayed pure. Right? But once we went to the valence level orbitals, that were much bigger, so that they overlapped, then they started mixing, like this.



Okay, now what if one partner is lower in energy than A? That's the case we're dealing with. But what will these ultimate energies be, when you have this competition between overlap, which is trying to mix them, and energy mismatch, which is trying to keep them separate from one another? So here's what you get. You get again splitting, but not as much shift as you got before. The lower level will look mostly like C, because C is better than A. The antibonding will look mostly like A, with just a little bit of C in it. And if they had had the same energy, you'd get a big energy shift because of overlap. But now you get a smaller energy shift, because the mismatch is fighting that. Right? You don't want to -- you lose as you try to mix, because of the difference between A and C, in this case. Now, so that one looks mostly like C, both in shape and in energy. That one looks mostly like A, both in shape and energy. But it's a little worse than A, and the first one is a little better than C.



Now how much smaller is that bonding shift? When they were mismatched, they didn't shift as much with the same overlap. How much less? Exactly where does it end up? Well that's not so crucially important to you, but there's a neat trick you can do to see that. Suppose we have that much energy mismatch. Okay? And the red dot is halfway between B and C, or between A and C, since A and B are the same energy. Now, so when they were perfectly matched, A and B perfectly matched in energy, that's how much splitting you got from that amount of overlap. So these are the two factors that go into it: how much splitting you get when they have the same energy, and how different the original energies are. How can you put those together, to find out how much shift you actually get? Well what we do is slide that over, and then bend down the blue one, so it's perpendicular to that, and make a rectangle around it, and draw the diagonal. And that diagonal will be the new energy difference between the new levels. Okay? You say, why do I make this construction? It's because there's a quadratic formula that comes in. So the diagonal is the square root of the squares of the two sides, the sum of the squares of the two sides. Okay, so we can rotate that; put a tack in there and rotate it. And that's going to be the new energy difference. Okay? And it's bigger than the original energy difference, because the diagonal has to be bigger than one of the sides. But because of the normalization, because one of them is a little less than -- one of them is a little smaller and one's a little greater -- that was the less than/greater than √2, 1/√2 thing -- it's going to shift up a little bit, both levels will shift up a little bit, like that. Right? I'll do it again so you see it. Both will shift up. So there are the new levels. That's how you do it. Okay?



Now, so for a given overlap, the bonding shift, how much the electrons changed when the atoms came together, is reduced if the energies aren't well matched, if you have the same overlap. You don't get as strong a bond, from that point of view. The energies didn't shift down as much, or up as much. But still A+C will be lower in energy than the original one was, A+B. Okay? And we had that construction, which shows us an interesting thing, which is the splitting is not very sensitive to the one of these contributions that's smaller. There are two contributions: the overlap part, that's the black arrow -- or pardon me, that's the blue arrow -- and the energy difference, the energy mismatch, which is the black arrow. But if one of those is very small, like here, the mismatch, which is black, is very big, and the overlap, the blue one, is very small. But you can see that the length of the diagonal is not going to be very sensitive to what the overlap is. You could increase that blue overlap, the width of the rectangle, which wouldn't change the diagonal very much. So it's sensitive to the one that's bigger. Okay, don't worry too much about that part. I just think it's fun.



Okay, so we can generalize from this. Mixing two overlapping orbitals gives one composite orbital that's lower in energy than either of the parents, and one that's higher in energy than either of the parents. Okay? The lower energy combination looks -- that is, its shape -- is most like the lower energy parent; and the same is true for energy. Right? And the higher one looks like the higher parent. For a given overlap, increasing energy mismatch decreases the amount of mixing, and decreases the magnitude of the energy shifts. Now what does this have to do with bonding? Okay, the AC electrons are clearly lower in energy. But that's not really what we're interested in. We're interested in the change of energy that comes when you break the bond, or make the bond. So which bond is stronger, AB or AC? How many think AB is stronger, given this scheme? How many think AC is stronger? It's clearly got lower energy electrons, AC. But this is a classic example of "compared to what?" Okay? How are we going to break the bond? Suppose what we're going to do is break it, to put two electrons in C, because that's the lower energy atom; or in the other case, to put two electrons in B. Okay? So both electrons go the same way. Can you see anything bad about that? Why might it be better to put one electron each way? Kevin?



Student: Well in terms of spin, they have to be --



Professor Michael McBride: Not spin. You can put two electrons in an orbital with opposite spin. So spin's not a problem, if you only have two. If you had three it would be a problem, but with two, spin is not a problem. But there's an obvious problem with having both the electrons go to one. Dana?



Student: They'll repel each other.



Professor Michael McBride: They'll repel each other. If you could put them on opposite ones, so much the better. Right? But it would be possible to do this, and it might plausible in the case of AC, where C is lower in energy than A; although the repulsion might make it better to go one on C and one on A. Okay, but anyhow, suppose you break it this way. Which one would be easier to break? Which requires less energy, red or blue?



Students: Red.



Professor Michael McBride: Red requires less energy. Okay, so AB is stronger, if you're forming A+ B-, if both the electrons are going the same way. Right? So mismatch aids heterolysis. Heterolysis, (uneven breaking) is to put them both on the same side, rather than one on each side (that's called homolysis, as you'll see in a second). So if you put them both on one side, then it's easier to break if they're mismatched. The bond is stronger if they're well matched. Okay? That's because that. But now suppose you do homolysis and put one on each. Now which one's easier?



Student: The blue one.



Professor Michael McBride: Now the blue is less energy than the red. Okay? So now the AC is going to be stronger. So mismatch hinders homolysis. Right? So you can't really say which bond is stronger, unless you say, "Compared to what -- how are you breaking it apart?" Okay?



Is all this true? So I've been weaving this fairytale for you, for the last couple of weeks, but is it true? Well, we can check it with experiment. For example, compare HH with HF. HH has perfect matching between the two atoms. In HF, there's a much bigger nuclear charge on F; electrons in the valence orbital are lower in energy on F than they are on H. Okay, so if we look at the σ orbital, the valence σ orbital, it's symmetrical in the case of H2 and quite unsymmetrical in the case of HF. It's big on F and small on H, just as we were speaking of. That's the σ, big on F. Right?



But there's also an antibonding orbital, σ*, which has a node. So that one is the contrary; it's big on H, small on F. This is just what we were talking about a few minutes ago. And the electrons, of course, are only in the σ, not in the σ*; there are only two electrons and that one's empty. And the * means that it's antibonding. Right? So it has the node between the nuclei, planar in the symmetrical case and bent in the case of HF; but it's still the node. Now, how about the experiment? Heterolysis, where you break, in order to get H+ H-, in one case, or H+ F- in the other case. In the case of HH, it costs 400 kilocalories to break the bond. In the case of HF, it costs only 373. Incidentally, this is a lot of energy, to break this bond, because you're putting both the electrons the same place. It can happen in solution, but it's not so easy in the gas phase, which is what we're talking about. Okay, so indeed we were right, that mismatch weakens the bond; for heterolysis, the HF bond is weaker.



That's why you call it hydrofluoric acid, because it can easily break off an H. You don't call H2 hydrogenic acid or something like that. Right? But if you break it into two atoms, two H atoms or an H atom and an F atom, then it costs only 104 for HH, but 136 for HF. So the HF bond is stronger, if you're breaking to atoms. So that's the first verification, experimental verification, of what I've been talking to you about. And we're going to go on next time to talk about XH3, which will have lots of experimental data that will show that this stuff is sensible.



[end of transcript]

Lecture 14
Checking Hybridization Theory with XH3
Play Video
Checking Hybridization Theory with XH3


This lecture brings experiment to bear on the previous theoretical discussion of bonding by focusing on hybridization of the central atom in three XH3 molecules. Because independent electron pairs must not overlap, hybridization can be related to molecular structure by a simple equation. The "Umbrella Vibration" and the associated rehybridization of the central atom is used to illustrate how a competition between strong bonds and stable atoms works to create differences in molecular structure that discriminate between bonding models. Infrared and electron spin resonance experiments confirm our understanding of the determinants of molecular structure.




Transcript



October 6, 2008



Professor Michael McBride: Okay, so we're in the payoff period of having spent all this time on quantum mechanics, and we want to see first whether what we've done is realistic or not. So last time we looked at H-H and H-F, in terms of what the bond strengths were, and we saw that indeed one of them was more stable for breaking into ions and the other more stable with respect to breaking into atoms. But here's a more extensive reality check today, which involves a lot more experimental evidence. And it has to do with XH3; that is, the molecules BH3, CH3 and NH3. And we're going to look at their structure, and how they move, and see how that checks out with what we expect theoretically. So first, what do we expect theoretically? Well BH3, CH3, NH3 are the same, except boron has three valence electrons, carbon has four, and nitrogen has five. Now each of them uses three of their valence electrons to share with hydrogens to make the three XH3 bonds. But boron then has no leftover electrons, carbon has one and nitrogen has two. So we want to see what the implications of this are going to be for the structure and dynamic behavior of these molecules, and then we'll check it against reality. First we need to look at something that shows that the hybridization that you use to make bonds should be related to structure.



Now suppose there is one of these atoms that uses an sp hybrid orbital, that looks like that. You see the little dot where the nucleus is. And suppose it happens also to use some other hybrid orbital to make another bond. Now, you want to have, because of the Pauli Principle -- and I'm sorry, we don't have time to go into this now -- the orbitals that you're using must not overlap one another. Because if there's net overlap between two orbitals then -- and you have electrons in both of them -- those electrons are, to a certain extent, the same thing, if the orbitals overlap. If they don't overlap, then they're independent. So to have only two electrons per orbital, the orbitals must not overlap. So we can use that as a criterion to say something about what other orbitals could be used at certain angles. For example, suppose the other hybrid orbital that this atom was using was sp3. Now you'll notice that -- remember -- let's suppose the green and the blue are positive, and the red and the yellow are negative. Now how much overlap is there? Well there'll be positive overlap where green overlaps blue, and there'll be positive overlap where red overlaps yellow. But there'll be negative overlap where green overlaps red, or where yellow overlaps -- no, where blue overlaps red, or where green overlaps yellow. Okay? Is that right? Yeah, where blue overlaps red, or green overlaps yellow; that'll be negative. And this, there'll be some angle at which those things exactly cancel out and you get no net overlap, no overlap integral. And we're interested in knowing what that angle would be; that is, how should hybridization be related to angles, be related to structure? Well, so the question is, what's that angle? And there's a nice formula, that isn't too hard to work out, why it should be that way, that relates spm and spn, the values m and n, to the angle that should be between them, in order to have no overlap. So let's use that formula and look at several mn's, and what the angles is. Let's look at sp with sp; sp2 with sp2; sp3, sp3; and sp-anything with spinfinity. What's spinfinity? That says the ratio of p to s is infinite. So it's a pure p. So what angle would be pure p orbital make with any hybrid? Okay, so those are the things we're going to talk about. So first let's try the first one. You can do the math for me. Okay, so it's sp1, sp1; m and n are one. So the denominator's the √1, that's one. The cosine of the angle is minus one. What's the angle? What angle has a cosine of minus one?



Student: 180.



Professor Michael McBride: 180. Okay, good. So that would be linear. And you knew that, I bet, already. Didn't they teach you that in AP Chemistry, that if you have two sp orbitals, they're linear, or if you have bonds that point opposite direction you use sp hybrids for them. If they didn't, that's true anyhow. Okay? Now how about sp2, sp2? So we're going to have m and n are both two; √4 is two. So it's the angle whose cosine is minus one half. Anybody know what that is? I don't -- pardon me? Sixty? That's +½. It's 120, to be -½; it has to be greater than ninety to be negative. So that's going to be 120º, and that's trigonal. So sp2 hybrids make an angle of 120º. Now how about sp3 hybrids? That's √9; so minus a third. That one probably isn't stored in your head, but it's an interesting angle. It's 109.5º, which is the angle in a tetrahedron. Right? So that methane, which makes sp3 bonds at the carbon, has tetrahedral angles, 109.5º. Now, how about sp-anything, with n=infinity? What's the denominator?



Students: Infinity.



Professor Michael McBride: It's infinity. So what's the angle whose cosine is…



Student: 90º



Professor Michael McBride: Okay, 90º. Right? So that means -- that is what makes things σ and π. Right? A p orbital, with any other hybrid, has to be perpendicular to it. So you won't -- that's what we talked about, σ and π, last time. So the p orbital is π, and it's perpendicular to anything else, 90º, no matter what other hybrid. Okay, so now how about this particular one we have here where it's sp3 and sp1? That turns out to 125.3º. So that's what the angle should be, if the carbon were making one bond with sp and one with sp3. Right? But the important thing here is that's there's a relationship between hybridization of two orbitals on the same atom and what the angle should be between them. Yes, Angela?



Student: When would you have an sp3 and an sp hybrid; like be at angles?



Professor Michael McBride: Well we're going to have to decide what determines the hybridization. Right? And it could be different for two bonds from the same atom, and we'll see examples of that. So that will become answered soon. It's if they're bonding to different kinds of things, they may have different hybridizations; and we'll see an example of that, several examples. So we want to optimize the hybridization, that is to get -- yes, Lucas?



Student: Does different weightings of the hybridized orbitals change the angle, degrees?



Professor Michael McBride: Yes, there are two ways. You could change the angle by moving the nuclei around, for some other reason, whatever it is, and that would then change the hybridization of the carbon that's making the bonds. So it works both ways. If you change the hybridization, you change the angle; if you change the angle, you change the hybridization. Okay, now we want to optimize hybridization. What do mean, optimize? We mean we want to get the lowest possible energy. We want to choose the hybridization that'll give the lowest hybridization [correction: energy], and therefore a particular geometry, for XH3, certain angles. Okay, now let's think about that. We're going to have one of the valence electrons of the X atom in each of three bonding atomic orbitals, bonding to the hydrogen, and whatever's left over goes into the fourth atomic orbital, and we want to minimize the energy. Okay? Now let's first look at it from the point of view of the bonds. We want to make the strongest possible bonds. We want to maximize the overlap in the bonds. What hybrid gives the best overlap? sp. But we have to make three bonds; sp you can only make two bonds. So what's the best you can use if you want to make three bonds?



Student: sp2.



Professor Michael McBride: sp2. So no matter whether it's BH3, CH3 or NH3, all of them will overlap best in making three bonds, if you use sp2. So the bonds cry out, "Use sp2"; that'll maximize the overlap and minimize the energy of the electrons in the bonds. Okay? But that's not the only game in town, because there's also the other electron that's on the atom, if there is another electron. Okay? Because carbon has one and nitrogen has two other electrons. So they should get a say in it too. So the X atom, which has to accommodate whatever's left over, says, "Okay, go ahead, make three bonds, but get the lowest energy for the other electron, that I'm the only one responsible for." Right? So what's the lowest energy you could make for a pair of electrons, or for a single electron; what's the lowest energy valence orbital?



Students: s.



Professor Michael McBride: s. So the atom says, "Okay, make your three bonds, but make sure you don't let s orbital go begging." Right? Use the maximum you can of s orbital, because otherwise you're using high-energy atomic orbitals when you might not need to. Now sometimes these would agree and sometimes they would fight with one another, the bonds and the atom. Okay? So from the point of view of the atom, boron, with only three valence electrons, all being used in bonds, doesn't have anything else. So it would certainly want to use three sp2 and leave the p orbital vacant. That gets the maximum utility out of the s; puts the s where the electrons are. So that's what boron says, use three sp2's, for the bonds. What does carbon say -- no, pardon me, what does nitrogen say? Nitrogen, if it wanted to be the best atom, would put two electrons, the unshared pair, into the s orbital, and one into each of the remaining three p orbitals. Right? That gets the maximum use of out of s. You don't want to waste s by putting it where there's only one electron on the nitrogen atom. You want to use it where there are two electrons. Okay, so nitrogen says, "Well, I'd prefer to use three p orbitals to make my bonds, and put two electrons into the s orbital." Okay?



Now, what does carbon say? Carbon says, "Look, I got four electrons; three of them are going to be in some kind of hybrid, that's going to make the bonds, and one of them's going to be something else. But if I take the s out of one place, I'm going to put it into the other one. Over all the four, I'm going to use up all the s character. There's one electron in each of them. Who cares?" Right? So carbon really doesn't care, from the point of view of the carbon atom, about hybridization. It'll care for making -- for the bonds, but it doesn't care for the unshared electron. Okay, but now we have to reconcile these two considerations. So the boron will make the best bonds if it uses sp2 and it'll make the best atom if it uses sp2. So great, they agree. So boron should really want to use sp2 hybrids for its bonds. Remember, those are the ones that form 120º. So it'll be the pattern of a Mercedes star. Yes, Yoonjoo?



Student: Can you explain again why we use three p instead of like, the hybrid orbitals for nitrogen?



Professor Michael McBride: If you're just looking at a nitrogen atom, it's going to have two electrons in one of its orbitals, and one in each of three other orbitals. Right? So those two electrons, that orbital that holds two electrons, should be as low as possible in energy; and that's s. So put two in s and one in each of the three p's, and that'll give the lowest energy atom, with one electron in each of the four orbitals -- pardon me, with two in one and three in the other three; one in each of the other three. Okay? Okay, so but anyhow, with BH3 they agree, from both points of view: the strongest bonds and the best atom. So it'll certainly want to be flat, and 120º bond angles. Now how about CH3? CH3, from the point of view of the bonds, wants to be sp2, and the atom doesn't care. So fine, be sp2, be flat, be planar, with 120º bond angles. But it won't care quite as much as boron does. So carbon says it'll prefer to be planar, but not as strongly; so it'll be easier to bend it, because it doesn't have this consideration of leaving a p orbital vacant. Okay?



And now about nitrogen? Now you got a fight on your hands, right? Because the bonds want to be flat and 120º bond angles, but the atom wants to use p orbitals, 90º angles for the bonds; so like the corner of a cube would be the nitrogen, the bonds would be going out at 90º angles. So here there's going to have to be some compromise. Okay? So NH3, since it must compromise, won't be either perfectly flat or 90º angles, but something in between. It'll be pyramidal. Okay? Now are the predictions clear; and the reason for them? Yoonjoo, do you see now, why it's the way it is? Don't let me bully you into it, if you don't really understand. Dana?



Student: Can you explain the difference in what the atoms and the bonds want for nitrogen again?



Professor Michael McBride: Yes, what we want to do is get the lowest possible energy. The lowest energy for the bonds will be the best overlap. You're making three bonds. So sp2 will give the lowest energy, as far as the bonds go. But there also might be -- and that's the only game in town for boron. Right? But in carbon you have another electron you have to accommodate, and there's a question of what energy it will have, even though it's not being shared. And in nitrogen there are two of those other electrons. Now, the best orbital of an atom, for an electron to be in, the lowest energy, is s. And in nitrogen there are going to be two electrons that want to be in an s orbital. If they do that, they use up the s orbital, and the only thing you have left over to make the bonds is the p orbital. So now you've got a fight on your hands. Are you going to have sharp angles and have the unshared pair in a low-energy, s orbital but weak bonds, or are you going to have strong bonds, sp2, but put the unpaired electrons in a p orbital? Okay, Kevin, did you have a question or did that answer it? Okay. Yes? I can't hear very well.



Student: Is there such a thing as an sp3 orbital?



Professor Michael McBride: Yes. What does sp3 mean? It means a hybrid orbital, a mixture of s and p, with three times as much p as s. Okay? You could have that. You could have three sp3 bonds, for nitrogen, for example, and then an sp3 hybrid, in which the unshared pair lives. Okay? That would be a possibility, sure. And that's the kind of thing I'm talking about here, by saying it would be intermediate, between being 90º bond angles and 120º bond angles, flat. Any more questions? Yes?



Student: Why are we concerned about the atomic orbitals rather than molecular orbitals?



Professor Michael McBride: Ah, I'm going to get to that just shortly here. So, because we're talking about bonds. Bonds are local. Bonds don't go over the whole molecule. Molecular orbitals go over the whole molecule. And we've been talking thus far, mostly, in terms of molecular orbitals, but now we're shifting to talk about bonds, because what we wanted to do was talk about bonds. And I'll do a little about the philosophy of that shortly. Okay? So here are the predictions. BH3 should strongly want to be planar and be hard to bend. CH3 should want to be planar but be easy to bend. And NH3 should be easy to distort out of plane, so that it's not flat; in fact, not only should it be easy to distort, it should prefer to distort in its lowest energy form.



Okay, so now let's see if that's true. Is it true that that unshared pair will compete for s character, and therefore make nitrogen pyramidal? Now, how do we know? There's our prediction: distortion from the plane will weaken the bond and deprive the electrons of s character; so it doesn't want to bend. Here, distortion from the plane weakens the bonds, but it shifts s character to the lone electron, where at least it's not wasted, because there is at least an electron there. And in NH3, it weakens the bonds but it shifts s character from single electrons to a pair of electrons; which is something good. Okay, so are these predictions true? Well, we need experiment. How about X-ray for seeing the structure of BH3, CH3, NH3? How about NH3? Have you ever seen NH3? Have you ever smelled NH3? It's ammonia. So how about doing X-ray on it?



Student: No.



Professor Michael McBride: Why not? Shai?



Student: Because you can only use X-ray to see electrons --



Professor Michael McBride: Can't hear very well.



Student: You can only use X-ray to see electrons; those are the ones that are going to move when you ---



Professor Michael McBride: No, that's not the main problem. Yes?



Student: You can't make it a crystal.



Professor Michael McBride: You have to have a crystal to do X-ray crystallography, and NH3 is a gas. Now you might say, okay, we'll cool it way down to liquid nitrogen temperature and then maybe it'll crystallize. It'll be difficult to grow a crystal of ammonia, but it's conceivable we might be able to grow a crystal of ammonia. Okay? Can we use the same trick for CH3 and BH3? That one's a gas. Okay? CH3, unfortunately, if you get two of them together they react, to get C2H6. So you're never going to be able to cool them down and get a crystal. And the same thing is true of BH3. It forms B2H6. Right? So forget crystals for this. We can't use X-ray crystallography, the technique we've used so far to do this. But there are other techniques for finding structure, and the one we're going to talk about -- actually we're going to talk about two -- but we're going to talk about, not X-ray, we're going to talk about IR -- we're going to talk about spectroscopy, infrared spectroscopy, and electron spin resonance spectroscopy. Okay? We're going to talk more about infrared spectroscopy next semester, but I'll show you just enough to get through the question that we're interested in now. Does that remind you of anything that moves like this in a high wind?



Student: A weathervane.



Professor Michael McBride: It's like an umbrella, right? In a high wind it blows inside out. Okay, so this is sometimes called the "umbrella" vibration of XH3. And you can treat it with Erwin Meets Goldilocks, as if it's one-dimension; just how much motion of all these atoms, how much motion. You know what the masses of the atoms are. You know the kind of motion you're interested in. It's all coordinated. So you're changing really just one-dimension, in moving all these things in a coordinated way back and forth. Right? So you can treat it as an Erwin problem, with a fictitious mass that reflects the amount of motion of each of the four atoms, while that's happening. Okay, so here's BH3 doing that out-of-plane-bending vibration, the umbrella vibration, and CH3 doing the same thing with its electron. Now, it turns out that BH3 vibrates 34-trillion times a second. Okay? And that's called 34.2 terahertz; that's how often it vibrates. CH3 vibrates 18.7 terahertz; so only half as fast as BH3. Does this make sense? Suppose the springs that you're bending when you do this vibration -- suppose it's springs -- which set of springs is stiffer? The BH3 or the -- does a stiff spring make a vibration faster or slower?



Student: Faster.



Professor Michael McBride: Faster. So it says the springs are stiffer in BH3. It prefers to be planar more strongly than CH3 does. Right? Now, you might ask -- what might you ask at this stage, when I tell you that it vibrates that fast?



[Students speak over one another]



Professor Michael McBride: No, that's not the question I want.



Student: How can you tell?



Professor Michael McBride: How do you know? How do I know that it vibrates at that frequency? Well, so it would be a stiffer spring, but how do I know? Well, when it vibrates, the electric charge is moving back and forth. Right? So the direction of the -- there's a dipole there, +/-, and it's moving back and forth. And what does that create, when you have charges moving back and forth? Electromagnetic radiation. Right? So electromagnetic radiation is going to come out, as this thing vibrates, like that. And it comes out with a shorter wavelength for BH3 than for CH3. Right? And it turns out that the wavelength is such that for BH3 there are 1100 vibrations, waves, in a centimeter; short wavelengths. And it's longer wavelengths, only 600 waves in a centimeter for CH3. Now, we can also -- so different light would come out, infrared light, of these two; or different colors of light, infrared light, get absorbed. That's how you actually measure it. And remember, the frequency of light has to do with the energies of the photon. So in, if you take the quantum mechanical view that you go from one quantized state to another quantized state, when you absorb, or emit, light, then we know, from this frequency, that in one case the two energy levels you're talking about are separated by 3.26 kcal/mole, and in the other case by 1.73 kcal/mole; only about half as much. Is everybody on board with this so far? So it's from the absorption of infrared light that we know how fast they vibrate, and therefore how stiff the springs are. Okay?



Now just how stiff are they? Well we can use Erwin for this point of view. We have the amount of deformation, which is our one-dimension, as it goes back and forth, and we can assume that it's something like Hooke's Law -- that's an assumption -- the distortion, these springs. And then we can adjust the potential, how sharp the parabola is, of potential energy, until we get two energies that are spaced like that. Does everybody see what we're doing? So here we do it. On the left, we get these lowest two energy levels and we -- there's the parabola that does the trick. So that's a strong, a stiff spring, strong planar preference. And in the case of CH3, to get that lower energy separation, you need a broader parabola. It's easier to bend; the springs are softer. Okay? Everybody see how that worked? Okay, so infrared spectroscopy then tells us that first, that it prefers to be planar, and second, that it's stiffer for boron than it is for carbon; just as we predicted. Okay? Now how about for NH3? For NH3, you see two absorptions in the infrared, and they're very near one another; 968 and 932 wavenumbers; that is, waves per centimeter. If we translate that, it means that there are two energy levels -- there's a low energy level -- and then there are two that are very close to one another, in order to have these two different colors of infrared light being absorbed. How can we get this? How can we get -- this is not Hooke's Law. Remember what happens with Hooke's Law in the energy spacing? What happens? They're even. Remember? So this isn't anything like Hooke's Law. Where do you get two that are very similar to one another? Do you remember?



Students: Double minimum.



Professor Michael McBride: What?



Students: Double minimum.



Professor Michael McBride: A double minimum gives that. So NH3 must be a double minimum, like this. Right? And those must -- the blue and the red one, must be the one with three nodes and the one with two nodes. There's also one with zero nodes, and one with one node. But those are both down at the bottom. Right? They have the same energy, because there's not enough overlap between the two to make them split. Right? So the IR spectroscopy of ammonia tells you it's a double minimum, and also, in order to get that spacing, tells you how big the barrier is. It costs three kcal/mole to make NH3 planar. It prefers to be pyramidal. Okay? And that splitting, that "tunnel" splitting, that we talked about before, for those two levels, is thirty-seven wavenumbers. Remember, there's also the two that are down at the bottom, with zero and one node, and they're separated by only one wavenumber; they're very close to one another. But we can translate that into the tunneling frequency, when you're down in the lowest energy levels. How fast does it go back and forth? And you remember how to do that. If one wavenumber is two small calories/mole, not kilocalories, and we use that thing that we talked about in lecture nine, about how to go from splitting to tunneling rate; and it turns out that in the ground state, in the lowest energy state, ammonia is flipping back and forth 1011 times a second. And in fact for a while that was used as a fundamental clock; here was something that you really knew the frequency of. Hooke, when he was trying to devise a clock that could tell where you were in longitude, could've used ammonia, if he'd had the technology to do it, which would be an absolute frequency. Okay, so IR shows -- completely confirms what we said before, that BH3 is stiff and flat; CH3 is easily bent and flat; and NH3 is not bent, but can tunnel back and forth, by the umbrella motion.



Okay, so now the other technique I want to tell you about is electron spin resonance spectroscopy. And it turns out -- and we don't have time to go into this -- well you know already that electrons are magnetic, and some nuclei are magnetic, and you can get a magnetic interaction between the nucleus and the electron, which influences how hard it is to change the electron from being pointing one way to being pointing another way. That's a kind of spectroscopy; see what color of light you need to make an electron flip from being pointed one way to the other way. But it's influenced by the magnet, which is the nucleus. But there's a neat thing about it, which is that influence can occur only when the electron is inside the nucleus. Where, in a 1s orbital, is the highest electron density? Remember the formula for a 1s electron? What's the atomic orbital? It's a constant times what?



[Students speak over one another]



Professor Michael McBride: e-ρ. Where's it biggest?



Student: At zero.



Professor Michael McBride: Inside the nucleus is the most dense part of the electron. Right? So when the electron is inside the nucleus it feels this effect. When it's not inside the nucleus it does not feel the effect, for the purpose of this spectroscopy. Okay? So we can measure how much s character there is in a hybrid orbital that holds an electron by seeing how strongly the nucleus interacts with the electron; because it happens only to the extent that it's s. Right? So we could measure the percent of s character by measuring the light that's absorbed by the electron. Christopher?



Student: How could the electron ever actually be inside the nucleus? Wouldn't that have two particles occupying the same space?



Professor Michael McBride: But they're different particles. And in fact it's a little more complicated than that, because what's the kinetic energy, if the distance is zero? Coulombic energy is 1/r. Suppose r is zero. What's the kinetic energy?



Student: It's infinite.



Professor Michael McBride: Infinite. Right? Now, that means the electron is really zinging, which means it becomes relativistic and its mass changes and all sorts of wild things occur. So it's actually that which causes this interaction. So we'll leave that across the way here in Sloane Physics Lab. Okay? And take it from me, that the amount of this splitting that you see in the ESR spectrum depends on how much s character there is on the orbital. Okay? And you can worry about whether it's precisely inside the nucleus, or just so heavy and fast that it's behaving weirdly. Okay? But anyhow, that's the trick. Okay, so the singly-occupied molecular orbital, the one where there's only one electron, in CH3, will give rise to electron spin resonance, and the magnitude of this effect will tell you how much s orbital there is. Okay, so let's look at it. And this is what I just said. There's a line separation due to magnetic interaction, and it occurs only if the electron spends time on the nucleus, which happens only for an s orbital; so we can measure how much s orbital.



Okay, so here's a picture of planar CH3, and the orbital, the singly occupied molecular orbital, SOMO, that contains the electron. How much time is it spending on the nucleus? Zero; there's a node at the nucleus, it's in a p orbital. Now suppose we bend it, so it looks like that; or we could look inside and it looks like that. Now it's got some s character, in that orbital. Right? And if we make little boxes and blow them up to see what it looks like down near the nucleus, one of them has no density at the nucleus and the other has electron density at the nucleus. So you'll see magnetic interaction, in this case. If it's bent you'll see magnetic interaction, and the more bent, the smaller that angle from 120 -- remember, the smaller the angle, so it'll go down toward ninety; the more s character, there'll be more of it. Okay, so here's what you see for CH3. Bear in mind it has to be magnetic C, it has to be C-13 H3, not C-12, in order to see the spitting. But at any rate, the amount of this effect is thirty-eight gauss, which are the units it can be measured in. And that implies that, on average, it's 2% s. Now, remember we already said that its lowest energy form is flat, with no s. But it's constantly undergoing this umbrella vibration. So it's spending time with different amounts of s character, not zero, as it vibrates. And the average is 2%. Now suppose I made the hydrogens heavier, by making them deuterium. So I make the atoms heavier. Do you remember how that changes the shape of the wavefunction, if I make things heavier? Here it is for H is moving. What if it were deuteriums that are moving, how will the thing change?



Student: [inaudible]



Professor Michael McBride: It'll get narrower. So as it gets narrower, that means it's not bending as much, on average. And what should happen to the thirty-eight gauss, if it doesn't bend as much?



Student: [inaudible]



Professor Michael McBride: I'll let you think about it a second. It's only when it's bent that it has s character and you see this effect. If it were rigidly flat, the splitting would be zero; thirty-eight would be zero. What would happen if I change H to D, so that it doesn't bend as much, on average? Sophie, what would you say?



Student: [inaudible]



Professor Michael McBride: Can't hear.



Student: If it was more rigid it would --



Professor Michael McBride: If it were more rigid it would -- how big would the effect be? Bigger or smaller, the gauss?



Student: Smaller.



Professor Michael McBride: Pardon me?



Student: Smaller.



Professor Michael McBride: Thirty-six. Right? So again it confirms what we're saying, that it prefers to be flat. It's only when it bends that you get this effect, and it bends less with heavier atoms than with lighter atoms. Wonderful. So again it confirms what we said. There's what CD3 looks like. It's a little bit narrower and therefore a little bit less of this magnetic interaction. So this is a structural isotope effect. CH3 spends more time more bent than CD3, and thus uses more s character, and thus gives a bigger magnetic interaction. Okay, so again we've confirmed what we predicted. Now here's one that we didn't work on in predicting, is CF3. Now what's changing H to F going to do? Well, it's going to be heavier. So the same deal that would happen with deuterium should happen here. But there might be more to it than that, because now we have more protons and a bigger nuclear charge of the atoms that were H before. So they're going to attract the electrons in the bond more. So, but there's another thing too. Because the fluorines have a partial negative charge, and are bigger, they have more electrons altogether, they'll repel one another. Now what would -- if the things attached to the central carbon get bigger and repel one another, what effect should that have on it being bent? Should it make it easier or harder to bend?



Student: Harder.



Professor Michael McBride: If these things are big here? Easier to bend from flat, or harder, if these things are --



Students: Harder.



Professor Michael McBride: Harder, right? So valence electron pair repulsion, and all that, should say that if CH3 is flat, what about CF3?



Student: It's flatter.



Professor Michael McBride: It should really be flat, harder to bend, if these things are bigger. Okay. But there's another point of view that you can take, which is that fluorine has the lion's share of the bonding electron pair. The bonding electron pair, in each of the C-F bonds, is spending more time near fluorine than near carbon. So carbon, being a cagy sort of character, says, "Why should I waste my valuable s character trying to stabilize these electrons that are spending their time on fluorine anyway?" Right? "Why don't I use it to stabilize that other electron, that's only on me and not on fluorine?" Got the idea? So if carbon uses its s character to go to the odd electron, forgetting the ones that are mostly on fluorine anyway, then that would make it more bent. That would favor bending, if you put the s character in the odd electron rather than in the bonds. That'll favor using more p character in the bonds and therefore more bending.



So this is exactly the kind of thing that Bacon would have liked. Right? This is a crucial experiment. There are two opposite predictions, and you can do an experiment and find out which one is true. Is it that fluorine is bigger and negatively charged, so it makes it more flat? Or is it that the carbon says, "I want to put my electron, the only electron I'm really solely responsible for, in the orbital with the most s?" Right? So this is a real horserace and you place your bets. Okay? Now how many people think it's -- we'll guess ahead of time -- how many people think that fluorines are bigger and negatively charged, so it will be harder to bend away from planarity? And how many people think that it's the odd electron, that you want to put s character in? Well this is very close. Okay? So we'll have to look at experiment. Okay, so this is the singly-occupied molecular orbital of CF3. It's mostly on the central carbon, but a little bit on the fluorines. And if you look in the middle, you see that there's the s character on the carbon. And if we measure the splitting, it's 271 gauss. What was it for methane?



Students: Thirty-eight.



Professor Michael McBride: Thirty-eight. This is ten times almost, or whatever number of times, maybe like eight times more s character than there was in methane. What's the conclusion? It's much more bent, much more of the time, than CH3 is. Right? Okay, so that means it's 20% s. That means that it's sp4 in that, almost sp3. Right? So it's almost perfectly pyramidal, like methane would be. Right? So it's the character of that odd electron, wanting to be s, that's controlling the geometry. Okay, now we're going to get to what Christopher wanted, which is why the heck are we talking about bonds when we've been talking about molecular orbitals all the time? Well one reason is we wanted to talk about bonds. That's how we got into this whole thing, asking what bonds are. Okay? But there are three different operators here, and they have different motivation and methods. First there's the molecule deciding what to be; and then there's the computer doing quantum mechanics, deciding what the molecule should do; and then there's you, trying to understand all this. And these are different. And let's see the way in which they're different.



The experimental molecule, all it wants to know -- it doesn't care about Schrödinger -- all it wants to do is get the lowest possible energy. So it minimized the total energy, kinetic plus Coulomb energies, of the electrons and the nuclei, by just settling down into the most comfortable, lowest energy arrangement. Okay? That's what the molecule does, it just gets happy. Okay? And once it does that, then it has a certain structure, a certain arrangement of the nuclei and of the electron clouds. It has a total electron density, which you can measure by X-ray. It has certain energies that we showed measuring by IR; although you can measure them other ways as well. It has nuclear electron density, that we just measured by ESR. It has dipole moment, how much there's a separation of positive and negative, that you can -- that if it vibrates gives off IR and so on. So there are these real things that come from the molecule just doing what comes naturally. So it doesn't care about Schrödinger.



Now you have a computer that comes along, armed with Schrödinger's equation, and it knows it can't possibly solve the true Schrödinger equation, it's too complicated. So it'll do some approximation and try to get an approximate value. So it'll minimize the total energy, using the Schrödinger equation, with some realistic constraints; for example, using orbitals, or using configuration interaction, or using density functional theory, or some approximation. Okay? So it'll use like a limited set of atomic orbitals, only say 2s and -- it won't use the 5p orbital or the 5g orbital or something like that. Use Self-Consistent Field, or some correlation, or delocalized molecular orbitals come out. And its goal is to get a useful prediction of the properties; maybe the electron density, maybe the energy, maybe the vibration frequency, maybe absorption of, what colors of light get absorbed and so on. So the goal of the computer is to use any tricks it can get around -- it can do, and it doesn't care how hard it works -- it's not something you have to be able to do on a piece of paper -- just to get something useful. And it will be validated, the approximations you made, by experiment. If it agrees with experiment, then you can believe the computer. But that doesn't mean you understand what the computer did. Right? It's a really complicated calculation.



And that's different from what you want to do, because your goal is to understand structure and reactivity, with the simplest realistic model; not Schrödinger's fundamental equation, because you don't have the brain power to manipulate numbers that way and do all these integrals and so on; leave that to the computer. Your goal is to try to look at things and understand: ah! that should be reactive with this; that bond should be very stiff; that molecular structure should be very stiff because the, quote, "bonds" should be stiff. Okay? That's what you want to do is to develop an intuition that allows you to understand structure and reactivity. And you want the simplest, obviously, but realistic model. And that, to be both simple and realistic, may not be possible. Right? There's a certain tension there. Okay?



So what kind of things you'll talk about? You'll talk about localized bonds. You won't talk about these big things that are complicated and go over the whole molecule. It's interesting to see, for example, that those are analogous to atomic orbitals and that kinetic energy plays a part in them, but that doesn't empower you to predict the properties of things. Okay? So you want to talk about localized bonds, maybe pretend they're springs or something like that. And what makes a spring stiff or not stiff? And that, Christopher, is why we have been talking about springs and things here, rather than talking about molecular orbitals. So you talk about -- these are Lewis's ideas -- bonds, lone-pairs, hybridization. We talk about energy-match and overlap. And we'll be talking soon about HOMOs and LUMOs -- highest occupied molecular orbital, lowest unoccupied molecular orbital -- and what a very important thing those are.



So we're going to look at just little bits of things, to try to find out what's crucial and will allow us to understand. We want them to be realistic. We want them fundamentally to be based on something true, like quantum mechanics, not just on springs. But we have to make it simple enough to understand. So we're aiming for qualitative insight. And we'll know when we're right if we agree with experiment, and also if we agree with what the computer predicts; even though the computer doesn't know why it predicted it, it just grinds through and gets you a number. But we want to understand why the computer gets that number. Okay. So let me just do one thing quickly here. Pathological Bonding. Or let me let you go to your next class; we'll do Pathological Bonding next time.



[end of transcript]

Lecture 15
Chemical Reactivity: SOMO, HOMO, and LUMO
Play Video
Chemical Reactivity: SOMO, HOMO, and LUMO


Professor McBride begins by using previous examples of "pathological" bonding and the BH3 molecule to illustrate how a chemist's use of localized bonds, vacant atomic orbitals, and unshared pairs to understand molecules compares with views based on the molecule's own total electron density or on computational molecular orbitals. This lecture then focuses on understanding reactivity in terms of the overlap of singly-occupied molecular orbitals (SOMOs) and, more commonly, of an unusually high-energy highest occupied molecular orbital (HOMO) with an unusually low-energy lowest unoccupied molecular orbital (LUMO). This is shown to be a generalization of the traditional concepts of acid and base. Criteria for assessing reactivity are outlined and illustrated.




Transcript



October 8, 2008



Professor Michael McBride: Okay, so now we have a number of different perspectives we can take on understanding what holds molecules together; that is, on bonding. We can look at the real molecule, for example, with X-ray. Or we can do a computer calculation, and if we do a careful enough computer calculation, we can get something that we believe is pretty close to reality, like the total electron density I showed you. Right? And the goal of the computer is to approximate the Schrödinger equation. You can't do it absolutely, but you can, depending on how much computer power you want to throw at it, you can do it close enough for most purposes. Okay. But you're neither the molecule -- well you are a molecule -- but you're neither the kind of molecule we're studying, nor are you a computer. You're a chemist, and your goal is to understand things; let the computer do the heavy lifting mathematically and you understand it. Okay, so let's see if we can understand some of the things that we didn't understand when we were looking at the molecule really, with X-ray diffraction. Right? When we were looking, we looked at this molecule and said that the bonding, from the Lewis point of view, was pathological. And you remember how we saw that. We looked at a cross-section through those three atoms, and it looked like that in the electron difference density. And you remember what's funny about it? What's pathological? Dana?



Student: The bonds appear to be bent.



Professor Michael McBride: Some of the bonds are bent, and what else?



Student: And there's another bond that's missing.



Professor Michael McBride: And one bond is missing. So that bond isn't there in the electron density, the difference density. And those are bent. They don't lie on the line between the nuclei. Now, can we understand that, from the point of view of molecular orbitals, of Schrödinger's equation; why it should look like that in this particular case? So would a computer's molecular orbitals provide understanding? So I was just showing you, as the class assembled, how you get these molecular orbitals, and I was showing you some of them. There's the framework of the molecule superimposed on this map, and that triangle on top there, on the right, is what we're looking at. And so we could look at the thirty-second of the thirty-three occupied orbitals, the next to the highest occupied orbital, and it looks like that. That's the one we were just looking at, before we went on to the PowerPoint. And you see there is electron density in those -- the wavefunction has big value in the region where those bonds are bent. But there are thirty-three of these occupied orbitals, and they put electron density every which way, and it's very difficult to understand why this is the way it is, why it's bent. Even if you grant -- which is not easy to grant -- that this is the proper orbital to look at, and that indeed the electron density of the red and blue, big red and blue lobes, are not on the line between; we don't know why they are. Right? And there's so much other electron density in the molecule, with all complicated nodes, that it's very difficult to know where to look. Right? So the answer is no, it doesn't provide understanding. It gives far too complicated an answer. It doesn't tell you why. Even if you believe that it gives you the right numbers, it doesn't tell you why. Okay? But analysis in terms of pairwise -- that is, don't go over the whole molecule, just look at two atoms adjacent to one another, making a bond, so just a very local view -- pairwise bonding overlap of hybrid atomic orbitals explains these two pathologies. So let's look.



First, I will assert that the best overlap you can get between hybrid orbitals on these carbons -- if the nuclei are arranged in an equilateral triangle, so 60º angles -- the best hybrids you can get point at the angles shown by the red arrows here. They do it in order to get the best overlap. Now this seems funny, because the angle between the red arrows on the right carbon is greater than 90º. Why do I say it's funny that it's greater than -- you're not going to be able to make it 60º because two orbitals that made a 60º angle would not be independent orbitals. Obviously if they were right on top of one another, they'd be the same orbital. But you can't say they're not the same orbital if they're just a little bit away, because still they're mostly the same orbital. It's only once they get to 90º that they can be different. What would the orbitals be that would be independent of one another at 90º?



Students: p orbitals.



Professor Michael McBride: p orbitals, right? So 90º is the smallest angle you could get. Why don't you use 90º? Why do you use greater than 90º? Okay, that's the question. Why not use p orbitals and make the smallest angle you can? Does anybody see why that is? Okay here, there's a p orbital that points at 90º. You see its axis, the yellow axis, is a little bit inside the red arrow, and the blue axis, which would be another p orbital, perpendicular to that one, is a little inside the other red arrow; they make an angle of 90º. Got it? Now why don't the atoms use p orbitals to make the smallest possible angle? Can anybody see why that won't give the best overlap? How could you change the orbital so as to get better overlap?



[Students speak over one another]



Professor Michael McBride: Claire?



Student: The best overlap is between an s and a p orbital. So if it's two p orbitals, you don't have the best overlap.



Professor Michael McBride: Remember, that depends on the distance. And they're very close to one another. And it turns out it might be that at this angle -- if you used the s on one of the carbons, and the p on another -- it might be that you would get better overlap. I don't know that it's true, but it's not implausible. But what would be wrong about using this p orbital from that carbon, and on the top one of the triangle using an s, which might give better overlap? Can you see what's bad about that? If you use all the s of that top carbon there, you can't use any of it to make any of the other bonds from that carbon. Right? All you have left over on that one is p's, which give crummy overlap. So even though you might be able make this single overlap better, you couldn't make all the bonds strong from that top carbon. So you wouldn't get the best overall overlap. So that's a good suggestion. But there's another way to do it, which is to use a hybrid orbital. Now here we've put in a little bit less than 20% of s, to give the ratio sp4.1. Right? But notice that it really distorted; it made it much larger in the region where we're interested in it overlapping. Right? In the process, the axis moved out, to close to where the red arrow is. Right? But even though the bonds didn't point -- let's see, like this -- even though the bonds didn't point at one another, by extending they overlap more. Got the idea? So the hybridization, by shifting it to one side, gives better overlap, even though the angle might not be quite as good. So you have bent bonds, paradoxically enough, in order to get the best overlap. Isn't that funny? You'd think the best overlap would be when they'd point at one another. But you get it better when they make an angle because they stretch out.



Okay, so that explains the bent bonds. Now how about the missing bond? Well again we can see that in hybridization. Because what's left over, after you make those sp4.1's, three of them, pointing up from the bottom carbon, after you make three sp4.1's, what you have left over is an sp4.1's. And it points down, not up. If it pointed up, if the big blue lobe were up and the red down, then it would be part of those other three orbitals. The only way you can make it point up is to make the others point down. Right? And then they would have crummy overlap. So there's very poor overlap between the tiny lobes of these two. So there's not enough overlap to get a bond. You don't stabilize the electrons because you don't have overlap. Right? And rehybridizing this one, to strengthen this particular bond, by making it point up, would drastically weaken all the others, making them point down. Right? So the best overall overlap, for all the pairs that are involved in these bonds, is to have one bond not there, or to have very, very poor overlap so that the other three are very good. Sam?



Student: So when you hybridized the other two and you get the sp4.1's, shouldn't -- and I understand that shifts the electrons towards the other atom, but how does that shifting distort the angle?



Professor Michael McBride: Remember the angle between hybrids depends on m and n; spm and spn, we did that last time. If it's pure p, it's 90º. If it's sp3,-sp3, it's 109.5. If it's sp2,-sp2 it's 120; and so on. If it's sp4.1, it makes that certain angle that's shown there bigger than 90º. Okay? And the only reason for that is so that these hybrids will be different from one another; they won't be the same. Right? They won't overlap each other on the same atom, because to the extent they do, they're the same, and so they can't at all. They have to be orthogonal, if they're on the same atom.



Okay, now three views of BH3. First nature; the total electron density. I'm going over the same thing again and again here, with different examples. First look at total electron density; nature. Then we'll look at molecular orbitals; that's what a computer does. And then we'll look at bonds from hybrid atomic orbitals; and that's what we do, as chemists.



Okay, first we'll look at the electron cloud of BH3, calculated here, using that same program I just showed you. So we can look at the successive layers of the onion, the high electron density, lower, lower, lower, the various contours. So this lowest one is mostly 1s of boron, but a little bit of the 1s's of hydrogen too. And one of them doesn't even show up, because it's so small. All the hydrogens should be the same, but the graphing capability of the computer, for something that has as few points as this, the way it draws surfaces, doesn't draw them smooth. Right? It connects points, so it doesn't even show one of those. Nobody would ever make such a plot. Okay? That's 0.3 electrons per cubic angstrom. So suppose we go to 0.15 electrons per cubic angstrom, half that density. Now we're beginning to get smoother surfaces. And it's on the boron, it's 1s of the boron, and a little bit of the hydrogens as well. Right? And now we're going to lower the electron density, 0.05. And now there's an interesting feature you can see here. You see that dimple? Why is that? Why is there not as much electron density, at this level, around the boron? We're out in the valence-shell now; not 1s anymore, because the 1s is really down close, high density. It's because the hydrogen atoms are pulling the electron density away from the valence atoms [correction: orbitals] of Boron.



Often you talk about this in terms of electronegativity. But why is something electronegative? It's because of what the effective nuclear charge is. The effective nuclear charge of hydrogen, for the valence level, is one. Right? But for boron it should be higher. Boron is the fifth element. But it's got the 1s electrons shielding the nucleus. So it's not as high an effective nuclear charge as hydrogen does. So when you put the orbitals together, hydrogen is lower than boron, and the electrons, in those bonds, are mostly on the hydrogen. So you see that dimple in the middle, because there's not much valence electron density on the boron.



Now we go to 0.02 electrons per cubic angstrom; and finally to 0.002 electrons per cubic angstrom, which is defined as (for purposes of drawing what the van der Waals surface is, that is) how close molecules tend to get to one another at normal pressure in a liquid or in a crystal. Right? And that's, it's just found empirically, just by testing and measuring things, that if you draw a surface at 0.002 electrons per cubic Angstrom, you have about how close things would come to one another when you bump these surfaces into one another. Okay, so that's the van der Waals surface; that's a definition, right? And we showed once before electrostatic potential, where you show that same surface but color it, to show how the electrons are distributed, and the protons in the molecule. So where would a positive charge on the surface be happy? Where it's negative. That's red. And where would it be unhappy, the positive charge, because the molecule is positive there? That's the blue. So you can see what it is. It's red where the hydrogens are, because they're pulling the electrons, and it's blue where the boron is, because it's giving up electrons. Okay, so there's what the real molecule looks like.



Of course I didn't do it by an experimental measurement from BH3, because you can't crystallize BH3. I did it by a computer calculation. But I believe that what it gives is what you would see in an experiment, if you could do the experiment. Okay, so that's the total electron density. Now let's look at how the computer breaks that up into molecular orbitals in a calculation like this. So it looks at -- it makes symmetrical molecular orbitals, like Chladni things, with nodes; we've seen lots of those. So here's the framework, BH3, and we're going to look at the molecular orbitals the computer calculates. And we're going to make those molecular orbitals by adding up atomic orbitals, linear combination of atomic orbitals.



And which atomic orbitals will we use? The best would be if we used every atomic orbital. Then we could get perfect flexibility, but the calculation would take too long. So we use just a minimal number of atomic orbitals; namely the 1s orbital on boron, and also the valence orbitals, 2s, 2px, 2py, 2pz, and also the 1s orbitals of hydrogen. So there are one, two, three, four, five, six, seven, eight, there are eight atomic orbitals we're going to use, to make up our molecular orbitals. How many independent molecular orbitals can we make, if we start with eight atomic orbitals? We can make eight molecular orbitals, the same number you started with. If you tried to make more, they wouldn't be independent of one another. But you can make -- from eight, you make eight. Okay, so these are going to be the energies that'll turn out of the eight molecular orbitals that we're going to get, that the computer is going to get, starting from that. If you added more atomic orbitals, if you let it use 3s, 3p, 3d, and so on, then you'd get more molecular orbitals, but they'd be higher in energy, and less relevant. Okay? In fact, how many of these are relevant? How do we know how many we care about, the energy of? Russell?



Student: The total number of electrons.



Professor Michael McBride: How many electrons. We can put two electrons into each. Once we've got it filled up, we don't really care what other energies there could be; at least not for this purpose. Okay, well, it turns out that boron has five electrons, hydrogen has three; eight altogether. So there'll be eight electrons, four pairs, and that means four occupied orbitals. We don't care so much about the others, right? So those are the Occupied Molecular Orbitals, OMO. Okay? And then there'll be, also from this set of atomic orbitals, there'll be four Unoccupied Molecular Orbitals, UMOs. Okay, now those two are called -- nobody talks about OMO and UMO, but they talk about HOMO -- that's very important and we'll be talking about that -- that's the Highest Occupied Molecular Orbital. And also they're very interested in the Lowest Unoccupied Molecular Orbital; and you'll see why very shortly. So the LUMO and the HOMO are very important. In this particular case it turns out there are two HOMOs that have the same energy. Not surprisingly, because it's such a symmetrical molecule. Okay now, what do you think the very lowest orbital looks like? Katelyn, you're our expert. What?



Student: 1s.



Professor Michael McBride: It's the 1s orbital of boron. So there it is. No surprise there. Not to diminish your contribution, but it's not surprising. Okay, so there's 1s. That's the boron core. Now what's going to come next do you think? After we've got the 1s molecular orbital, what'll be next?



Students: 2s.



Professor Michael McBride: 2s. But it's a molecular orbital, it'll be made up of everything. So here it is, right? And it's all the hydrogens and also the boron 2s are contributing to that. Again there's a dimple in the middle, more of the hydrogens than of the boron. Why? Because the hydrogen 1s is lower than the boron 2s. The boron 1s was lower; that was the first one, right? Okay. And then, so there's that radial node, the spherical sort of node down the middle, showing it's a 2s. Right? And then there's a 2px molecular orbital, and a 2py molecular orbital. And that's four, and that's all eight electrons. So that's what we care about. They're the occupied molecular orbitals of BH3. Now how about the vacant orbitals? The lowest unoccupied orbital is the last 2p, the 2pz, that's pointing in and out of the plane of the molecule. And then we're going to have 3s, which you see has that node in the middle; so two spherical sort of nodes. And then a dx^2-y^2, and finally dxy. Right? So that's the molecular orbitals that the computer generates for this thing, using what's called a "minimal basis set"; that is, the lowest number of orbitals you can use to get -- that is, through the valence level; you don't put the higher ones in.



Okay, now that's how the computer does it. Right? And you can get good numbers out of the computer, like the energy of the molecule, or the shape of the molecule, where the nuclei would like to be to minimize the energy. That's fine, but for understanding it's not so good. What we want to do, as chemists to understand it, is to partition the total electron density, not into these molecular orbitals, but into atom-pair bonds, and anti-bonds, as you'll see, and lone pairs; the same stuff Lewis talked about. Okay. That is, we usually do this, but sometimes it doesn't work. And when doesn't it work? When it doesn't work, we must use more sophisticated orbitals, where we say there is a certain phenomenon going on. You know what that phenomenon is? When we can't use localized bonds?



[Students speak over one another]



Professor Michael McBride: That's what resonance is, when this approach doesn't work. Okay, remember that's where Lewis Theory had to get sort of complicated, Baroque. Okay, so there's that 2pz orbital, a vacant orbital. Right? Here's the boron core. But now we're doing it. We're happy to use a vacant orbital. That's part of Lewis's idea. So 2pz we'll share with the computer; both of us see that as a lowest vacant orbital. The boron core we understand quite well. Okay? But how are we going to treat those other three pairs of electrons, in the picture Lewis would draw? Where would Lewis put those three pairs of dots? Angela? Can't hear.



Student: He'd make them the bonds between --



Professor Michael McBride: Those would be the bonds, between B and H. Okay? So what will those look like in our crummy picture? Right? We're not as good, at least I'm not as good as the computer at drawing these things. But what I'd say, it's going to be an orbital that's big on hydrogen, smaller on boron; some hybrid sp2 hybrid of boron is going to overlap with the 1s orbital of hydrogen. It's going to be bigger on hydrogen than on boron. Why? Why should the lowest, the bonding orbital, be bigger on hydrogen than on boron? Lucas?



Student: The 1s is more shielding.



Professor Michael McBride: Yes. That's another -- that that orbital, the hydrogen orbital, is lower in energy than the boron it's mixing with. Right? So it's uneven mixing. Right? Poor energy match. So the one that went down looks mostly like the hydrogen, and the one that went up looks mostly like the boron. So we see one -- in fact, we see three of these, one pointing toward each hydrogen. Right? From our point of view they all have the same energy, they're all B-H bonds, shown as their energy in red there. Okay? But notice, you get the same total energy, as the computer would. The computer divides the pie completely differently. Right? But you get the same answer -- right? -- and the same geometry that the computer does, from the point of view of what direction do the hybrids point and so on. So we're getting the same structure, we're getting the same energy. You get the same total electron density too, if you add all the electrons together. Three bonds add together to give the -- the same way, remember, three p orbitals added up together to give a sphere. Right? Our three bonds add up to give the same energy as the computer. So who needs a computer? Right? Our bonds work just as well, if we view them right. Okay? And then there would be the anti-bond, the one that's -- what's left over, that's big on boron, small on H, and a node between hydrogen and boron; it's antibonding, the one that went up in energy when the atoms came together. And again there'd be three of these σ*'s, big on boron, small on H. They would all have the same energy. But again their average energy is the same as the average energy of unoccupied molecular orbitals that the computer gets.



So we're not doing so bad, following Lewis in this respect, looking at the electron density in terms of bonds rather than in terms of molecular orbitals. Okay? So for many purposes, localized bond orbitals are not bad, and they're certainly a heck of a lot easier to think about. But, beware of resonance. Right? Because sometimes this isn't going to work. So we'll have to understand how that works. So the localized bond orbital picture, pairwise molecular orbitals; that is, two atoms coming together to overlap. And also there'll be isolated atomic orbitals, like unshared pairs, or isolated vacant orbitals, like the pz orbital of boron that we just saw, that will be -- this picture will be our intermediate between hydrogen-like atomic orbitals, which we really understand -- you just look up in the table and you've got them, right? -- and computer molecular orbitals, which are too complicated really to understand in any kind of detail.



It's handy to sometimes -- for some purposes it's very handy to be able to calculate them and look at them, right? But you don't do it very often in your head. So when must we think more deeply? When you have these localized orbitals, but they mix, to cause reactivity; that is, a molecular orbital from this molecule mixes with a molecular orbital from this molecule, and one goes down and one goes up. That's what a reaction is. Or when you have two pairwise orbitals in the same molecule, that we thought of as bonds or vacant orbitals or something like that, and those interact with one another, within the same molecule. We were too simplistic in thinking of them as independent. They interact with one another. That's what resonance -- that's when resonance occurs is when you have important interactions within a molecule, between orbitals, or between what you thought were orbitals; but were thinking too simply.



So where are we? We've gone through atoms, we've gone through molecules, and looked at how we can understand the electron density in molecules. And now we're to the final payoff, which is reactivity. How can we understand reactivity in terms of SOMOs, Singly Occupied Molecular Orbitals, high HOMOs, and low LUMOs? And then how can we recognize functional groups? So even if we never saw them before, we know how reactive they're going to be, and what with; that's the payoff. Okay, so now we know that if you have two molecules and you're interested in their reacting, there'll be molecular orbitals of the two that come together, interact. And if the molecule gets -- if the energy goes down because of that, that'll be a favorable reaction. If the energy goes up, they'll bounce off one another. Okay? So, but how many molecular orbitals are there on a molecule?



[Students speak over one another]



Professor Michael McBride: Suppose methane, how many molecular orbitals does it have? It has an infinite number, right? The 13z orbital and so on. Okay? So we have to pick out which orbitals are going to be important in interacting with one another, or else we're just going to be at a loss, because there'll be so many. So you have these orbitals on molecule B, say, and those are occupied, and the ones above are vacant, unoccupied. And on molecule A you have a bunch of occupied and a bunch of unoccupied molecular orbitals. And now we're going to look at orbitals of B mixing with orbitals of A. Right? And which ones? We might look at those two interacting, for example, mix those two, between A and B. Or we could mix those two, or those two, or those two, or those two, or those two, or those two, or those two, or those two -- and it goes up infinitely. So this is a problem. Even if we understand how orbitals interact with one another, we're going to have to narrow our focus, if we want to deal with this. Okay? So there are myriad possible pairwise mixings.



Now there's one that's going to be obviously important. Suppose that molecule B has an odd number of electrons, and therefore a singly occupied molecular orbital, and the same is true of molecule A; just suppose that happens. Right? Then when they come together and mix, those two electrons will both go down in energy. So that's bound to be favorable, substantially favorable, and you'll get a new bond. That's like two hydrogen atoms coming together. So if there are SOMOs, on the two molecules -- and notice, suppose B had a SOMO but A didn't. But you're not just talking about two molecules, you're talking about two flasks. If you mix -- you had to have two B's that both had SOMOs. So A could be another B. So if you have any SOMOs, you can pair them; B with B, if A doesn't have any. Right? So if they exist, if there's a SOMO, then -- and that's true of many atoms, or things called free radicals, like the hydrogen atom, the chlorine atom, the methyl radical -- then they're going to be very reactive. Okay? And for that very reason they're not very common. You can't buy a bottle of methyl radicals because they will all have reacted with one another. The same for hydrogen atoms, the same for chlorine atoms. So that's an easy case to understand, when you have SOMOs. But it's very rare to encounter it, unless you're talking about flames, where lots of things are atoms and free radicals, as they're called. We'll talk about that a lot more at the end of the semester. Okay, now how about -- there are two that could mix. It's not bad energy match, right? And suppose they have good overlap. Do you care about that one? Why not? Lucas?



Student: It doesn't exist.



Professor Michael McBride: Oh it exists, it's an orbital.



Student: Oh, there are no electrons.



Professor Michael McBride: There are no electrons to have that energy. It's an available energy, but nobody availed themselves of it. So it doesn't contribute anything. Right? So there's nothing there. So forget that one. And you can forget most of them for this reason, of the infinite number. Okay, now how about this pair? Now they have -- let's suppose they have good overlap. They're not bad energy match. Right? So two electrons will go down and two will go up. Do you care? Steven?



Student: No, because since there's two that are higher energy, they're antibonding.



Professor Michael McBride: Yes. So they'd more or less cancel. Do they exactly cancel?



Students: No.



Professor Michael McBride: In which direction do they not cancel? Corey?



Student: The lower one is bigger, so it would be even worse.



Professor Michael McBride: The lower one is bonding, but bigger -- by bigger, you mean greater than 1/√2 kinds of thing; or no, pardon me, less than 1/√2, it's smaller, is what you should have said. Okay, it's smaller. So it doesn't go down as much as the upper one goes up. So pretty much they cancel. But to the extent they don't cancel, it's repulsive, it's bad to come together, which means the molecules will bounce off. Right? So when you have filled orbitals hitting one another, the molecules repel one another. Okay? So that's weak net repulsion. That's not going to explain a reaction. Now how about here? Now we're mixing a filled orbital with a vacant orbital. So two electrons go down. So you don't have this canceling. Is this going to be important, this one? Josh?



Student: No, bad energy match.



Professor Michael McBride: Ah, the energy match is so bad that you don't get very much going down; not enough to make up for the fact that if they'll come together, there are going to be a lot of other orbitals, molecular orbitals, overlapping, and they're going to be repulsive, the kind we just looked at. So this weak attraction is not going to be worth much. So it's negligible mixing because of bad energy match. Now how about here? Now that's not bad energy match; I mean, it's not great energy match, but it's certainly better than any other pair of occupied with vacant. Okay? So that means that pair will go down, right? And you'll get bonding, if the energy match is good enough that the amount that they go down is significant. Okay? So remember, when we tried the same thing the other direction, from the highest one on the right to the lowest one on the left, the energy match was not good. So that didn't help. But this one, where it's high on the left and low on the right, that one will help. Okay, so what it requires, to have a reaction, is that you have a high HOMO, an unusually high HOMO, on one molecule, and an unusually low LUMO on the other, so that you have good energy match, or at least as good as you can get. Right? Then you could get a reaction. And it turns out that there are simple names for things that have unusually high HOMOs and things that have unusually low LUMOs. And you know what those names are?



[Students speak over one another]



Professor Michael McBride: Something that has an unusually high HOMO is called a base, and something that has an unusually low LUMO is called --



Students: Acid.



Professor Michael McBride: An acid. Right? So that's a generalization of the idea of acid and base. Most mixing of MOs affects neither the overall energy, nor the overall electron distribution, for one or more of these reasons. Okay? First the electron occupancy can be four; that is, both were occupied, which means two go up and two go down, and the ones that go up, go up a little more. So that's not a bond. Or zero, and you don't care about the energy. Or the energy match is poor, so there's not much mixing all together. Or there's poor overlap. So even if the energy match were good, you wouldn't get anything out of it. But, if you have high HOMO and low LUMO mixing, then you get reactivity; unusually high HOMO and unusually low LUMO. And that's the secret we've been going for all semester.



So here's increasing generality in the concept of acid and base. The names were introduced by Lavoisier, in 1789, and we'll talk about that. And from his point of view, a base was an element, a substance that could be oxidized, and an acid was something that had been oxidized. Like sulfur was a base; oxidized sulfur, sulfuric acid, was an acid. Phosphorus was a base; oxidized phosphorus was phosphoric acid. Carbon was a base; oxidized carbon was carbonic acid. Okay? Then 100 years later, Arrhenius, in Sweden, said that an acid was something that was -- had the idea that there were ions, things that were charged, and that a thing that would give H+ was an acid, and a thing that would give OH- was a base. That's Arrhenius's Theory. And that was further generalized, in 1923, by Brønsted and Lowry -- Brønsted was a Dane and Lowry an Englishman -- who said that an acid was an H+ donor -- that's the same as Arrhenius -- but that a base was something that could accept H+; not only OH-, but ammonia, for example, could accept a base [correction: proton]. And Lewis then, in the same year, had the idea that an acid was something that could accept an electron pair, like BH3, and a base was something that could donate an electron pair, like ammonia. Right?



And other names for those acceptors and donors are electrophile, something that loves electrons, will accept them; and a nucleophile, something that loves a nucleus, is something that will give electrons to it. But in the 1960s this was generalized finally to where it is now, that an unusually low vacant orbital is something that can accept electrons; electrons like to be in a low energy orbital. So that is the ultimate definition of what an acid is. And by the same token, an unusually high HOMO has electrons that are eager to avail themselves of some vacant orbital, to lower their energy; and that is a base. Okay? So as so often in science, as the science matures, you get more and more generality to the idea. And HOMO and LUMO, unusually high HOMO, unusually low LUMO, are the ultimate form of acid-base; and that's what we'll be talking about. But they have to be unusual. Why do they have to be unusually high and unusually low?



[Students speak over one another]



Professor Michael McBride: Because otherwise you don't have good energy match, to change the electron energies when they come together. Okay, now here's what you should have asked. Unusual: compared to what? Okay? So here is what -- we're organic chemists. If we were some other kind of chemists in some other universe, dealing with different elements, we'd have different points of comparison. But what we deal with is things that have a lot of carbon and hydrogen in them. And by far our most common bonds are among -- between carbons, or between carbon and hydrogen. So it turns out that the valence orbitals of carbon, the sp hybrid orbitals of carbon, and the 1s orbital of hydrogen, are roughly the same in energy. For our purposes we'll consider them to be the same. Okay? So you bring two of these together, they overlap, and you get σ and σ*. And we've talked about that, the bonding and the antibonding orbital. So most of the occupied orbitals we're dealing with are σCC or σ CH, and most of the anti-bonding orbitals are σ*CH or σ*CC. So those are our point of comparison. Those are the plain vanilla electrons and holes. Okay?



Now let's look at other things and see whether they're unusual with respect to this. Now there will be several classes that they fall into. Right? So those are the usual LUMO and the usual HOMO, or the CC analogs. Now what could make things weird? One thing could make things weird is that if you had valence-shell atomic orbitals, that didn't get mixed with anything. So they're not σ or σ*, bonding or antibonding; they're the original orbital, in between these. So that would make things unusual. A second one would be a little bit in between these two. If there's no overlap, then you have unmixed orbitals. If there's a lot of overlap, you get bonds. But there could be poor overlap, in which case you don't go down as far with your bonding orbital, or up as far with your anti-bonding orbitals. So a second case would be poor overlap. A third case would be that the actual atomic energy, the atomic orbital energy that you're using, the atom you're using, is not like carbon and hydrogen. So it starts at a completely different place from where the standard orbitals start, before they make the molecular, the bonding orbitals. And fourth can be if there's charge around. Because obviously if you have a negative charge, it's not such a good place to put electrons; that means the orbitals are high in energy. If you have a net positive charge, it's a good place to put electrons; it's low in energy. So charge can make it. So we're going to show examples of these four contributors that will allow you to identify when a molecule has an unusual orbital energy. Right? Unusually high HOMO, unusually low LUMO. And then you know that it'll be a functional group, and you also know what it'll react with. Because if you have an unusually high HOMO, it'll react with an unusually low LUMO; and vice versa.



Okay, so first let's look at the first one, unmixed valence-shell atomic orbitals. So H+, Arrhenius's acid. Right? That's a vacant orbital. Right? No electron in it. So that's obviously unusually low. It's where this started, the 1s of hydrogen, before it went up or down. So if you had H+, it would be an unusually low vacant orbital. How about NH3? Right? It has an unshared pair, only on the nitrogen. Now, it likes being on nitrogen, compared to carbon or hydrogen. Why? What makes it better to be on nitrogen, in a valence-shell orbital than on -- the unshared pair -- than to be on an unshared pair on carbon or on hydrogen? Angela?



Student: There's more protons in nitrogen.



Professor Michael McBride: There are more protons in the nucleus. It's a lower energy orbital, to start with. Right? But it didn't mix with anything. There are N-H bonds, in ammonia. Right? And they mixed and went way down in energy, and a vacant orbital went up in energy. But the unshared pair didn't mix with anything. So it still has that original energy. So it's unusually high. Compared to what? Compared to what?



[Students speak over one another]



Professor Michael McBride: σCH or σCC. That's our point of comparison. Okay, or the flip side of that is BH3, which has a vacant orbital. It's high, to start with, compared with a hydrogen or a carbon, because boron doesn't have a very big nuclear charge. Right? And its nucleus is screened. Right? So it's unusually high for an atom. But it's still atomic, it didn't mix with anything and go up further. Right? So it's unusually low for a vacant orbital. So BH3 is an unusually low LUMO. So is it an acid or a base?



[Students speak over one another]



Student: An acid.



Professor Michael McBride: Which?



Student: Both.



Professor Michael McBride: I can't hear you.



Student: Both.



Professor Michael McBride: You've read through the lecture. You don't get to vote [laughs]. It's like H+, right? An unusually low LUMO. So it's an acid. How about NH3? It's a base, unusually high HOMO. Okay, how about water? Right? It has unshared pairs of electrons. Yes Lucas?



Student: Sort of a side question. Didn't we just say that they both had these properties because they had low nuclear charge?



Professor Michael McBride: No, nitrogen has high -- which has low nuclear?



Student: Nitrogen is, low nuclear charge compared to carbon.



Professor Michael McBride: Yes. [apparently did not understand the student's erroneous claim about nitrogen and thought he was speaking of hydrogen vs. boron]



Student: And BH3 is even lower, compared to carbon.



Professor Michael McBride: Yes.



Student: But the energies are completely different because it's -- is it just because of the unshared pair?



Professor Michael McBride: But remember, you have to bear in mind which row of the periodic table you're in. Right? H, true, it only has one proton, but you're talking a 1s orbital, low in energy. For carbon, nitrogen, oxygen, boron, you're talking 2s. Right? So you're changing the game.



Student: Between BH3 and NH3?



Professor Michael McBride: No, no, not between BH3 and NH3. NH3 is unusually high energy for an occupied orbital. BH3 is unusually low; although it's high, compared to nitrogen, it's low for a vacant orbital. It's an unusually low LUMO. Nitrogen's an unusually high HOMO. Now oxygen is lower than nitrogen. So water is a base. But it's not as good a base as ammonia, because its electrons aren't as high, because it has a bigger nuclear charge, and the same n; it's in the first complete row in the periodic table. Okay, OH- is even higher. It's a stronger base than water. Why? Why are its electrons higher in energy than water?



Student: Less protons.



Professor Michael McBride: Got less protons. Right? It's got a net negative charge, which makes all the electrons high in energy; and specifically the pair we care about, the highest one, unusually high, because it's got a negative charge. [Technical adjustment] Okay, CH3- is even more of a base, because carbon doesn't have as many -- as big a nuclear charge as oxygen. So when it has a negative charge it's a really high HOMO. Okay, I got to quit there. We'll go on to two, three and four next time.



[end of transcript]

Lecture 16
Recognizing Functional Groups
Play Video
Recognizing Functional Groups


This lecture continues the discussion of the HOMO/LUMO view of chemical reactivity by focusing on ways of recognizing whether a particular HOMO should be unusually high in energy (basic), or a particular LUMO should be unusually low (acidic). The approach is illustrated with BH3, which is both acidic and basic and thus dimerizes by forming unusual "Y" bonds. The low LUMOs that make both HF and CH3F acidic are analyzed and compared underlining the distinction between MO nodes that derive from atomic orbitals nodes (AON) and those that are antibonding (ABN). Reaction of HF as an acid with OH- is shown to involve simultaneous bond-making and bond-breaking.




Transcript



October 10, 2008




Professor Michael McBride: So this is where we were at the end last time, trying to decide when a high HOMO, or a low LUMO, is unusual in its energy; because, as the world is, things like to be at low energy. We'll talk about that later, why that's so. But that means that electrons tend to be in the occupied, the low -- occupy the low-energy orbitals and not the high-energy orbitals. So that empty orbitals are almost always higher than any occupied orbitals. So if you want to get good energy match, and therefore mixing, and therefore bonding, the occupied orbital that you're interested in must be unusually high, and the vacant orbital must be unusually low. So the task we have is to recognize what orbitals should be unusually high and what ones should be unusually low, and if we can find the unusual ones, then we'll know what's reactive, be able to identify functional groups. Okay, so unusual, compared to what? Compared to normal occupied and vacant orbitals. That is, in organic chemistry, or in most kinds of chemistry actually, carbon-carbon, carbon-hydrogen bonds are occupied orbitals; not unusually high, they're our standard of usual. And the antibonds, corresponding antibonds are unusual -- are the usual vacant orbitals.



So what can make things unusually low or unusually high? When the valence-shell orbitals that, if they bonded, would go down and up, don't bond with anything, don't mix with anything, then they'll be unusual. Second, if the orbitals don't overlap, then they don't go down and go up. Third, if you have an unusual atomic orbital. These are made of carbon and hydrogen, going down. But if you start with something that's unusual, unusually high or unusually low, then the things that come from it will be unusually high or unusually low. And finally, electrical charge. So right at the end last time we were looking at the first example: unmixed valence-shell atomic orbitals. So we saw that the simplest of all acids, H+, of course is unusually -- if it were occupied, it would be unusually high for an occupied orbital. Right? Because it hasn't gone down. But it's vacant, and it's unusually low, for a vacant orbital, because it hasn't gone up. And so on with these others: an unshared pair on nitrogen; or a little less so on oxygen. A vacant orbital on boron that's not bonding with anything is an unusually low LUMO; even though orbitals on boron are not unusually low, because the nuclear charge isn't very big for something that's using the second, n=2, quantum level orbitals, valence orbitals. It's an unusually low nuclear charge for that. So the orbitals are unusually high in energy, not being attracted by more protons. But it's low for a vacant orbital, if it hasn't mixed with anything.



Okay, and finally notice that some of these are charged: H+; plus means that's a good place to put an electron, so unusually low; OH-, unusually high because of the negative charge; CH3-, especially high. So this slide also shows the fourth point, that electrical charge makes a difference on how normal occupied and vacant orbitals are. Okay, so these unusually low vacant orbitals are acids, like H+. Unusually high occupied orbitals are bases, like OH-, as we just showed on the previous slide. That means that you can mix the high HOMO of OH-, with the low LUMO of H+, and two electrons, the ones that were on the OH-, go down in energy, and that makes a bond. And our notation for showing that is to draw a curved arrow, and it gives water.



Now think carefully about what a curved arrow means; it's not the same as a straight arrow. A curved arrow designates a shift of electron pairs. It doesn't show that this atom moves from here to here. It shows that a pair of electrons shifts, and that pair of electrons goes from being on the oxygen, to being between oxygen and hydrogen, to form the bond. So you start the curved arrow where the electron pair is at the beginning, and you end the arrow, put its point, where it's going to be in the product. Right? So it's not showing that an atom moves from here to there. It's showing that the electrons that were formerly on oxygen are now between oxygen and hydrogen. The effect of that is, of course, to pull all the hydrogen to the oxygen. But that's not what's being shown by the arrow. What's being shown by the arrow is how the electrons are moving in our picture. Okay, so there's a case.



Now here's another base with the same acid, and we can draw a curved arrow there and show making an ammonium ion out of ammonia. Or we could use the same base with a different acid, the low vacant orbital of BH3, and draw the same curved arrow, and make this anion, make the boron-oxygen bond. Or we could start with completely different ones. We could start with ammonia, not OH-, not Arrhenius's base, but ammonia; not Arrhenius acid, H+, but BH3. But it's exactly the same reaction. The high HOMO mixes with the low LUMO to form a new bond. Okay? So that's what curved arrows show, and don't confuse it by trying to draw molecules or atoms moving, and showing a curved arrow. If you want to show them moving, draw some other kind of arrow; draw dotted arrows or a straight arrow or a wiggly arrow, or something. But the curved arrows have a very specific meaning.



Now the second reason that things could be unusually high or low is because they don't have much overlap. Notice in the first case, it was just an extreme case of that. There was no overlap at all, the orbitals were just there. But even though the orbitals are mixed with something else, if the overlap is poor, they haven't changed very much, they still look very much like they did originally. Right? So a good example of not having very good overlap is the side-to-side overlap of p orbitals. The π overlap -- you remember from looking at those curves of amount of overlap for different orbitals versus distance -- the π overlap isn't very big. So even though the p orbitals start a little higher on carbon than they do on hydrogen, or normal bonds of carbon that have s character in them, they start a little higher, but they don't mix very much. Right? So the π orbital is unusually high, and the π* antibonding orbital is unusually low, because they didn't go down or up very much. Okay? Now, which of those do you think is more remarkable, in its highness or lowness? So for this reason, the carbon-carbon double bond could behave either as an acid or as a base. Right? High occupied orbital makes it a base; low vacant orbital makes it an acid. And indeed, it does have both kinds of reactivity. But which do you think is more pronounced? Which would be more familiar, how high the occupied one is, or how low the vacant one is? Lucas, what do you say?



Student: How low the vacant one is.



Professor Michael McBride: And why do you say so?



Student: Because electrons have to sort of -- they would likely move into a position of an energy as close as possible to it.



Professor Michael McBride: Yes, but on the other hand, the electrons of the π want to move into something else.



Student: There are lots and lots and lots of electrons going up.



Professor Michael McBride: Well there are lots -- there are even more vacant orbitals than there are electrons. Right?



Student: Because you also have starring electrons that will -- like in a normal carbon atom.



Professor Michael McBride: Let's just look at the picture. Yes Catherine?



Student: I think the HOMO is more unusual because it's farther away from the usual --



Professor Michael McBride: Ah, the HOMO is further above the normal bonds, than the LUMO is below the normal antibonds. Why?



Student: Because the p orbital started --



Professor Michael McBride: Can't hear very well.



Student: Because the p orbital started a little higher.



Professor Michael McBride: Ah, because they started a little higher. Right? So most reactions that we'll study, or at least many reactions that we'll study of the carbon-carbon double bond, it's reactive because of the high HOMO, not because of the low LUMO; although there are cases where the low LUMO makes it reactive. But it's this π orbital, here, that makes it particularly remarkable, because they started a little high. Okay, so the high HOMO makes it unusually reactive. Now how about the C=O double bond? Right? So it starts with the same p orbital on the carbon, which has π overlap and therefore not much shifting up and down, with a p orbital of oxygen, which of course is lower than that of carbon, because of the higher nuclear charge. Which of these will be most unusual? Somebody got a suggestion? So what would you think the characteristic reactivity of a C=O double bond -- would it be reactive mostly as a high HOMO -- that is, as a base -- or mostly as a low LUMO? Which is more unusual, compared with what we compare things with? Andrew?



Student: I think it's a low LUMO.



Professor Michael McBride: Why?



Student: First of all it can't -- it's further away from its origin.



Professor Michael McBride: And why is it further down, compared to what we usually compare? Why does the π*, C double bond O, why is that orbital so very low? Pardon me, one reason it's low is because the overlap isn't very much. So it didn't go up so much. Right? That's the same, true with a C-C double bond. But what's special about the C-O double bond Andrew?



Student: I thought the atoms are different, there's a --



Professor Michael McBride: Because of the bad energy match, it didn't go up very much from the carbon, or down very much from the oxygen. But the average is low, that it starts from. So what's special here is the π is not so unusual, not so far from here. But this one, π*, is very far from here. So what should make a carbonyl especially reactive is its low vacant orbital, because of poor overlap; and that it started with an oxygen, an unusually low orbital, atomic orbital. Okay? We saw another example here. The three-membered carbon ring is unusually reactive, for a carbon-carbon bond. That is, the electrons in that carbon-carbon bond are not shifted down so much, and the antibond is not shifted up so much, as in normal carbon-carbon bonds. Why? What's special about the bonds in the three-membered ring? Kevin?



Student: Well two of them are bent.



Professor Michael McBride: Ah, they're bent. And what does that mean?



Student: Greater than 90°.



Professor Michael McBride: Yes, but they're bent, so what?



Student: It overlaps, there are better overlaps.



Professor Michael McBride: Does it overlap as well? So it doesn't go up and down as much. Right? So another example of poor overlap. Okay, or you could have an unusual atomic energy orbital in the molecular orbital. For example, how about a C-F single bond? Right? So the overlap can be perfectly good. It's a regular old σ bond, straight on. Right? But the fluorine starts very low. There's bad energy match. So the π* isn't so very different from a vacant orbital on carbon; unusually low as a vacant orbital. So it should act as an acid. Things that have high HOMOs should come up and react with that σ* of C-F. It's an unusually low LUMO. Can you think of what the flip side of that -- how could you have something like this, a carbon bonding to something, that for the same or analogous reason would have an unusually high occupied orbital? What would you want the carbon to bond with? Choose an element.



Student: Lithium.



Professor Michael McBride: Lithium. Why lithium?



Student: Because --



Professor Michael McBride: Because you looked at the PowerPoint?



Student: No.



Professor Michael McBride: [Laughs]



Student: I didn't look at it.



Professor Michael McBride: Why lithium?



Student: Because its valence electron -- it's easy for the valence electron to be given away.



Professor Michael McBride: Why?



Student: Because it would be as stable as a lithium plus one ion.



Professor Michael McBride: Why? That's restating the same thing, right? What's different about lithium? Why are orbitals on lithium, as an atom, unusually high in energy, compared to carbon-hydrogen? Catherine?



Student: Because it has a low nuclear charge.



Professor Michael McBride: Because it has a low nuclear charge. Right? It's right at the left end of the table. Okay? So its orbitals aren't very stable because it doesn't have very high nuclear charge. So it's a high energy. And therefore it has bad energy match. Or it could've been lithium that I drew here. I happened to draw boron, which is, for the same reason, a lower nuclear charge than carbon; therefore higher energy as an atomic orbital. When they mix, what's special here is the occupied orbital, σ, which is unusually high, because the average started high. Okay. So boron. You may remember that in the first slide we showed, boron had a vacant orbital; that p orbital on boron is vacant. Unusually low. Therefore what, acid or base? Unusually low vacant orbitals, which does that make it?



Student: Acid.



Professor Michael McBride: Acid. So BH3 should be an acid. Right? Well what I just showed you is what Shai? On the previous slide; we'll go back there.



Student: That it had a big B-H; yes the B-H.



Professor Michael McBride: It has an unusually high HOMO, BH3. So what does that make it Shai?



Student: It would make it a base but --



Professor Michael McBride: So is it an acid or a base?



Student: Both.



Professor Michael McBride: Wilson?



Student: Both.



Professor Michael McBride: It's both. And what does that suggest, if the same molecule is both an acid and a base? What do acids react with?



Students: Bases.



Professor Michael McBride: So what if the same molecule is an acid and a base?



[Students speak over one another]



Professor Michael McBride: It's going to react with itself. Two of the molecules will react with one another. Okay, and that's why you can't do a crystal structure of BH3, because BH3 becomes B2H6. Let's look at how it happens. So there's the low LUMO that makes BH3 an acid. What's the high HOMO? What's the high HOMO of BH3? In localized terms. I don't mean these big things that go over the whole molecule. It's the B-H bond, right? Which was poorly matched in energy and so on. So there. Notice that the B-H bonding electrons are big on hydrogen, small on boron. We saw that last lecture. Okay? So they should overlap like this, right? And the vacant orbital on the B should stabilize these high-energy electrons in the B-H bond. But now, you could imagine many different orientations of the top molecule that would allow overlap. Do you know why I chose this particular orientation? Which one acted as the acid and which one is the base? The one on the bottom, the vacant orbital, acted as an acid. The one on the top acted as a base. So what other possibility is there? Yoonjoo?



Student: So then there also -- you could reverse the roles of them.



Professor Michael McBride: Ah ha! You could have -- so that would make -- notice incidentally, what I forgot to mention here, is that there are three nuclei being held together by that pair of electrons now. Originally the top molecule, that pair of electrons, mostly on hydrogen but partly on boron, held the hydrogen and the boron together. Now that same pair of electrons is attracted -- is helping form a bond with a boron, the bottom boron. So in fact that's an unusual kind of bond, because it's bonding three nuclei together, not just two. It's doing double, or perhaps triple duty. So we need a new symbol to talk about such bonds, and a reasonable one would be a bond that looks like that. It's a Y bond. We could make it blue. Okay? But now we have, as Yoonjoo said, the vacant orbital on the top and the high occupied orbital on the bottom, and they can do exactly the same thing. So we have two of these three-center, two-electron bonds. Right? Two electrons holding three atoms together. Right? Or we can draw it that way. Right? And that's the structure of B2H6. Two of the pairs of electrons are each bonding three atoms together instead of two. Any questions about this? Yes Alison?



Student: Is that what the dotted line means for --



Professor Michael McBride: Yes, the dotted just means it's a little different. You could draw the solid Y if you want to. There's no really standard notation for that. But it's clear what it means. It's just that a pair of electrons is being shared among three nuclei, rather than two, and their electron density is -- correspondingly, if you did a difference density map, if you could do an X-ray of this, you'd expect electron density to build up in the middle of all three. Yes Claire?



Student: This may seem like a stupid question but a high HOMO, we've talked about as a base, and a low LUMO we've talked about as an acid.



Professor Michael McBride: Yes.



Student: And the low LUMO is unoccupied. If you think about it, it's sort of receiving electrons.



Professor Michael McBride: Right.



Student: But generally, bases are supposed to receive electrons, instead of acids, and --



Professor Michael McBride: No, you got it backwards.



Student: Do I?



Professor Michael McBride: Yes.



Student: Oh okay.



Professor Michael McBride: Because H+ is an acid. Right?



Student: Right.



Professor Michael McBride: So it obviously can't give up electrons. There aren't any electrons. An acid accepts electrons. Okay? Think about it a little bit in the privacy of your room, and you'll see that. Okay?



Student: Does the bond between three nuclei only happen --



Professor Michael McBride: I can't hear very well.



Student: Does the bond between the three nuclei only happen here because of the geometry of the molecule?



Student: Yes. And to get that, you have to have overlap. So if you didn't have -- if the top BH3 were oriented so that the B-H bond were vertical and the boron was way up at the top, it wouldn't overlap the other one. So you have to -- always to get a bond, you have to have overlap, because otherwise the orbitals don't mix and you just have the original orbitals, as we've seen.



Student: Is this kind of bond very common?



Professor Michael McBride: Pardon me?



Student: Is this kind of bond very common?



Professor Michael McBride: No, it's not very common, because there are not many really low energy vacant orbitals running around. Right? Boron is a very special case. But a lithium can do the same thing. But lithium doesn't have energy, orbital energies as low as those of Boron, because it doesn't have as big a nuclear charge. So boron is particularly good at getting this kind of thing. And this answers the puzzle about Lewis structures that we raised in Lecture two, about how can BH3 react with BH3? Right? That's how it does it, by making three-center bonds.



Now here's a True and False quiz. On the basis of what you know, is it true that low energy molecular orbitals result in bonding? True or false? I don't think you trust me anymore. [Laughter] We've been saying that when you get those low orbitals -- things come together, you get a low orbital -- that results in a bond. Lots of overlap. [Laughter] I can't talk you into it? Good. That's false. What makes a bond is lowered energy orbitals. It's when things come together and the electrons get more stable. It's not how low they are, it's how much they're lowered by the coming together, because then pulling apart they have to go back up again. So it's not whether they're high or low, it's whether they get lowered. Okay? Now, compared to what? What do they have to get lower compared to?



Student: By the size.



Professor Michael McBride: Yoonjoo?



Student: So it's kind of like how in Erwin and Goldilocks you showed the antibonding and bonding. So wouldn't it be the reactants?



Professor Michael McBride: You can say something in fewer words than that. What do you compare to? When you say energy is lowered, and that makes a bond, what's it lowered, compared to? Christopher?



Student: The atoms in their standard states.



Professor Michael McBride: Well yes. It wouldn't necessarily be between atoms, it could be between two molecules, like BH3 with BH3. But you're right. What it's lowered compared to is the things it was before they came together. Right? Now these things, before they came together, one of them had electrons, one didn't. Those might've been very high. Right? So they came together. The electrons went substantially down, but still aren't so very low. They could've been very low-energy orbitals to begin with, but not gone down very much, because there was bad overlap say. So these, even though they're lower than the ones we talked about first, would not be so bonding. It's these that were bonding, because they came down a lot, when the mixing happened. So it's lowering. Compared to what? Compared to the separated components, before you made this bond. Right? So when things come together, and that results in the electrons going way down in energy, that's a strong bond. Yes, Chenyu? Pardon me?



Student: How far does it have to go down?



Professor Michael McBride: Well that's what we're going to have to learn. That's a question of lore. Right? And it has to be at least enough to overcome the fact that when they come together other electrons, other orbitals, filled orbitals are overlapping, which is net repulsive. So it may not be that there's an absolute criterion. Right? It may be that if there are a lot of other things opposing the coming together, filled orbital with filled orbital, then you have to have really enormous going down. So you can't make a simple answer to that. But we'll learn as we see examples. Okay. So now, HOMO/LUMO mixing, for reactivity and resonance. So reactivity means between molecules. So far we've been talking mostly about atoms coming together and forming a bond. But molecules have high orbitals and low orbitals, as in the case of BH3. The B-H was a molecular orbital, or not an atomic orbital. Right? So when things come together, orbitals are orbitals. If you have an unusually high energy and an unusually low energy, and they overlap and go down, that makes a bond. So that's between molecules. But it turns out that what resonance is, is HOMO/LUMO mixing, within a molecule.



Now you might say the molecules have certain molecular orbitals. How can you mix them? Right? The idea is that we made our first analysis on the basis of localized orbitals: σ,σ* here; σ,σ* here; not these big Chladni things that go over the whole thing. But it may be that this σ,σ*, and this σ,σ*, are near one another and overlap. So that when we made our initial analysis, and looked only at this and only with this, we didn't take into account that this one might interact with this one, and give still lower energy. When that kind of thing is important, that's when you have to draw other resonance structures. And I'll show you an example. But first I'm going to show you reactivity, and then we'll go on to resonance. Okay, now let's look at the frontier orbitals for H-F. Okay? So it has four valence electron pairs and five valence atomic orbitals; 2s, 2pxyz on fluorine, and a 1s on hydrogen, and four pairs of electrons. So there are going to be four occupied orbitals. And this is what the lowest orbital looks like. What does it look like?



[Students speak over one another]



Professor Michael McBride: Well it's a 2s orbital, the 1s being the core on fluorine. But it's made up of atomic orbitals. Is it exactly a 2s of fluorine? Does it look like sphere on fluorine? Alison, you're shaking your head.



Student: It's a little bit distorted.



Professor Michael McBride: And how did it get a little distorted? What did we mix with the F orbital?



Student: The H.



Professor Michael McBride: A little bit of the 1s on hydrogen. It's mostly the F on fluorine. Why? Because the fluorine's way down in energy. So the best combination is mostly this. But you can see that it's a little bit egg shaped, a little bit drawn out toward the hydrogen. Okay, so it's mostly a 2s of fluorine, but a little bit of 1s on hydrogen. Right? Now here's the next one. What's that mostly? Can you see?



[Students speak over one another]



Professor Michael McBride: Tyler, what do you say?



Student: I would say a 2p.



Professor Michael McBride: It looks like a 2p on fluorine. Does it look exactly like a 2p on fluorine, or is it just hard for you to see that it's not? It's very similar.



Student: I don't know, it looks pretty close, but the blue one might be a bit bigger.



Professor Michael McBride: Yes, the blue one is a little bit bigger, because it's got a little bit of the 1s of hydrogen lowering the energy of the 2p of fluorine. Okay? And now the next two orbitals are these. What are those? Steve, what do you say?



Student: The 2py and 2pz.



Professor Michael McBride: Yes, the 2py and 2pz of fluorine. Is there hydrogen in those too?



Student: Not very much.



Professor Michael McBride: Why not?



Student: Because they don't overlap with the one --



Professor Michael McBride: Ah, they're orthogonal. It's a π versus a σ. There's no overlap, therefore no mixing. So those are the occupied orbitals. And the 2p's have the same energy. Okay, and then remember there are going to be five molecular orbitals. And now you make the last one, which is going to have another node, with the leftovers. What's left over, after we made the occupied orbitals? What's it mostly left over? We used up the 2p's of fluorine to make those HOMOs. Those were pure, right? But what was left over from the bottom?



Students: 1s.



Professor Michael McBride: We used very little of 1s on hydrogen, and there's a little bit of 2p fluorine and 2s fluorine that we didn't use in the bottom ones. So they're back in the top. So what we have is mostly a 1s on hydrogen, but a little bit of some kind of sp hybrid on fluorine. Is that clear to everyone? So that's the vacant orbital. Is it unusual energy, that vacant orbital?



Student: It's low.



Professor Michael McBride: It's low. Is it unusually low? Kate, do you have an idea? What kind of criteria do we have for whether an orbital should be unusually low? What do we look for?



Student: We look for overlap and energy match.



Professor Michael McBride: Overlap. Energy match. How about this case, good overlap? Quite good overlap. It's a hybrid orbital on fluorine pointed right toward a hydrogen. Good overlap. How about the energy match? Kate?



Student: Sure.



Professor Michael McBride: Sure what? Sure it's good or sure it's bad? One of them's hydrogen. What's the other one?



Student: The other one is fluorine.



Professor Michael McBride: Where's fluorine, compared to hydrogen?



Student: Fluorine is going to be lower.



Professor Michael McBride: Why?



Student: It has greater nuclear charge.



Professor Michael McBride: Right. Okay, good energy match or bad energy match?



Student: Not great.



Professor Michael McBride: Not very good. So you don't get much mixing. And you know that already by looking at the picture, because you didn't mix it very much. It's almost all 1s of hydrogen. So it's unusually low for a σ*; it didn't go up very much from hydrogen. So indeed -- now it's got an unusually low vacant orbital. What does that make it, an acid or a base?



Student: An acid.



Professor Michael McBride: It's an acid. Are you surprised that H-F is an acid?



Student: No.



Professor Michael McBride: Why? What's its name?



Student: Hydrofluoric acid.



Professor Michael McBride: Hydrofluoric acid. And that's what makes it an acid. Arrhenius would say it's an acid because it gives up H+. We say it's an acid because it's an unusually low vacant orbital. Right? So notice that those three are made up of three atomic orbitals: the 1s of hydrogen, the 2s of fluorine, and a 2p orbital of fluorine. And they're in different mixtures in three of these. Usually we've looked at just two things, right? There's one atom, atomic orbital, and another one, and they mix. Right? A pair. Here there are three going in, to give three molecular orbitals. But still you can see quite easily why they should be the way they are; why the bottom ones are almost pure fluorine, and the top one is almost pure hydrogen. Okay. And so that top one is σ*; the LUMO, the unusually low LUMO. Lucas?



Student: How can we be sure that the 1s of the low nuclear-charge hydrogen is going to be that much different from the really high nuclear charge 2s of fluorine?



Professor Michael McBride: Yes, this you have to learn. And I'm going to show you very soon how we can tell that kind of thing. Okay, there's a different picture, that was made in 1973, of the same thing. These pictures were drawn just a year or two ago, but this is -- you can see the same thing. It's drawn at a different contour level. It was based on a different calculation. But you can see it's the same thing. And we'll notice that -- how many nodes does this thing have, that are clearly visible?



Students: Two.



Professor Michael McBride: Two nodes, right? Between the H and F. That one is antibonding. Right? When they came together they cancelled in the middle, rather than reinforcing. There's another node there. But notice that that didn't have anything to do with the bonding. That was already there in the atomic orbital you started with. So it didn't have anything to do with lowering. It didn't have anything to do with whether the thing was bonding or antibonding. The pink node had to do with whether it was bonding. That came when they came together. Right? It's unfavorable. So it's better for them to come apart. But if they come apart, you don't do anything with the blue node. It's still there, it's part of the atom. Okay? So you have to recognize that there are two kinds of nodes. There are nodes that were there already, atomic orbital nodes, and there are ones that are associated with the coming together, and that's what makes something bonding, or antibonding; in this particular case it's antibonding. Okay now let's look, instead of H-F, let's look at CH3-F. Sam?



Student: Why is it called a frontier?



Professor Michael McBride: Because you have occupied orbitals, and then you have vacant orbitals, and the ones you're interested in are the lowest of the vacant and the highest of the occupied, at this border between occupied and vacant orbitals. Okay, so this one has seven valence pairs of electrons. So you're going to occupy seven orbitals. And here's what they look like, the seven that are occupied. What's the very lowest one, mostly?



[Students speak over one another]



Professor Michael McBride: It's the 2s of fluorine. Right? And what's the next one? Well it's C-H bonds, all mixed together. Right? But also a little bit of a p orbital on fluorine. It has the blue, sort of, toward the front, and a little bit of that red behind, as part of the p orbital on fluorine. So it's a mixture of the C-H bonds and of the p orbital on fluorine. And then we have these others, some of them coming in pairs, and those are the HOMOs, because we have seven occupied. And let's look at them a little more closely and compare them, each one, with the one beneath it. And I'll draw another picture too, of that older kind, for making this comparison. And we're interested in what is it that went together, to make these orbitals, and why is one lower in energy, and the other higher? So what do you see on the top orbital, say this one here, the top left; what's that made up of? What is it on the left side? Sophie, what would you say the orbital is on the left side of that?



Student: I think it's 2p over there.



Professor Michael McBride: Right. This part here is a 2p π orbital of fluorine. Now what's this thing on the right here, the dash bit down here? If you just saw that, without any of the rest of it, what would you say that was?



Students: 1s.



Professor Michael McBride: It's more than a 1s on hydrogen; that would be spherical. It's, at a certain contour, the C-H bond. Do you see that? So this is electron density in here, bonding between C and H. And these two on the top are little bits of C-H bonds as well, mixed together. So what this orbital is, is a mixture between some combination of C-H bonds on the right, and the p orbital of fluorine on the left. Now is it a favorable, or an unfavorable combination, of the fluorine orbital with the C-H orbitals? This one up here. Is the interaction between the fluorine orbital and the C-H orbitals bonding or antibonding?



Students: Antibonding.



Professor Michael McBride: Becky, do you have an idea? Are they building up electron density in between, or having a node in between?



Student: A node.



Professor Michael McBride: There's a node in between. So that's antibonding. What's the orbital on the bottom, this one?



Student: Bonding.



Professor Michael McBride: That's the same components, but it's the bonding combination. So this is the favorable combination of those, and this is their unfavorable combination. So this lower-energy one is favorable, and the upper energy is unfavorable combination. How about here and here? Can you see what that is? What's on the left, here, that lump of red and this lump of blue? Eric? Here, this thing, it's very complicated. Okay, will you agree on that?



Student: Okay.



Professor Michael McBride: But part of it, this part on the left, there's a red lump in behind and a blue lump in front. You can see it maybe more clearly here. Red behind, and blue in front, surrounding the green fluorine atom. What is that orbital, that atomic orbital?



Student: Probably the p orbital.



Professor Michael McBride: Can't hear very well.



Student: Probably the p orbital, a different orientation.



Professor Michael McBride: It's the p orbital on fluorine that's pointing more or less toward you. Okay? So this is the p orbital of fluorine, that's mixing with C-H bonds here. Right? In fact, it's the same thing as this, turned on its side. Okay? So this is the bonding combination, built up between the fluorine and the C-H's. This is the antibonding one, with a node between the fluorine and the C-H's. So here we have a node, in this bottom one, but that node came from the atomic orbitals. That's what made it a p orbital here, on fluorine. That one, you don't get rid of, if you pull the fluorine away. That doesn't have anything to do with the bonding, that's just an atomic orbital node. Right? On the top, you again have the same atomic orbital node, because it's the same p orbital that's involved. But what about the top? Sam?



Student: You have another node but --



Professor Michael McBride: There's another node, that one, and that's an antibonding node. Right? The electrons in that orbital would get lower in energy if you broke the bond, if the fluorine came away.



So the pink nodes are the ones we're interested in; the antibonding nodes, not the ones that are just part of the atomic orbital. Okay. Now, we're getting to Lucas's question. So this thing down here is made up of fluorine and C-H. The one here is made up of fluorine and C-H. Right? This is the favorable combination, and this is the unfavorable combination. So get it right. So here are the two of them. They come together when the F comes up to the methyl. Right? And you get a favorable one and an unfavorable one. But the fluorine and the C-H's may not be at the same energy. How do you know which one's lower? That's your question. Is the fluorine lower, or is the C-H lower, or are they about the same? Now how are we going to tell? If the fluorine is lower, and they come together, what does the lower one look like, mostly?



Student: Fluorine.



Professor Michael McBride: Fluorine. If the fluorine's higher, and they come together, what does the low one look like?



Student: C-H.



Professor Michael McBride: Mostly C-H, right? So by looking at how big these are, we can tell which one is lower. So when we look here, we see that here I would say they're pretty similar, left to right. There's not much difference between a 2p orbital on fluorine and C-H σ bonds; not much difference. But to the extent they're different I would say, looking at this, that the C-H is a little bigger here, and the fluorine is a little bigger here. Would you say that? So which is lower, a C-H σ bond or the p orbital of fluorine?



Students: C-H.



Professor Michael McBride: C-H σ bond would be a little lower than fluorine; not much though, pretty similar. So they're about the same, as we say. But if you have to make a choice, the C-H is a little bit lower in energy; the fluorine's a little higher. Okay. Now there's also a vacant orbital, made up with the leftovers here, some of the leftovers. And what's that? Let's look at a different picture of it. So this has three nodes that are obvious. There's one that's near the fluorine atom, a node, plane, that goes back. The furthest to the left. Everyone see that? Is that an antibonding node, or is that an atomic orbital node?



Student: Atomic orbital.



Professor Michael McBride: That's part of the fluorine atomic orbital of some hybrid on fluorine. Right? And the same is true at the C-H end. There's a node that's an atomic orbital node of the carbon. Right? But what's important? Between those other two is what's between the fluorine and the carbon, and the CH3. And that's antibonding. So that, the LUMO is a σ* antibond, between C and methyl. And why is it unusually low? The question is whether the LUMO should be unusually low. Why is it unusually low? Is the overlap bad?



Student: It's fine.



Professor Michael McBride: No, the overlap's good; the hybrids are pointing right toward one another. But what makes it low in energy? So I'm not giving you specific problems on this, but look over those things and run your brain as to those four different things that make orbitals unusually high or unusually low. Because that's what you'll be doing next week when you're doing these Wikis, each of you, to decide some functional group, why is it functional? What makes it unusually high or unusually low; or maybe neither? Right? Why is the C-F bond, more properly the C-F σ* orbital, the antibonding orbital, why is it unusually low? Compared to what? What do you compare it to?



Student: C-H.



Professor Michael McBride: A C-H bond. How come C-F is lower, for the σ*? Alex?



Student: Energy mismatch.



Professor Michael McBride: Pardon me?



Student: Energy mismatch.



Professor Michael McBride: Yes, the fluorine is really low, their average is very low. So the vacant orbital is unusually low. It didn't go up much. Right? Where have you seen that before? H-F. The previous slide, H-F, was exactly the same. It's an acid. Remember Kate, you helped us with that. It's an acid because of the bad energy match between fluorine and hydrogen. This is the same thing, but it's the bad energy match between fluorine and carbon. Okay. So CH3-F is an acid for the same reason that H-F is an acid. Right? There's the low LUMO of H-F. Right? It's got that same antibonding node. Shai?



Student: Why is there no energy mismatch between carbon and hydrogen?



Professor Michael McBride: It just happens that it worked out that way. But it's true. Hydrogen is 1s. That makes it unusually low. But it doesn't have a very big nuclear charge. Carbon has a higher nuclear charge, but you're talking 2s and 2p. And it turns out those just cancel out. If that hadn't been the case, organic chemists -- like if boron happened to match hydrogen, then maybe our organic chemistry would be boron-hydrogen, not -- there would be borohydrates, not carbohydrates. Right? But that's the way things are.



Student: So in this picture we can tell that --



Professor Michael McBride: Actually there are other reasons it wouldn't be boron, because boron has these vacant orbitals and forms B2H6 and so on. But it just happens that that cancellation works that way. Lucas?



Student: Just by looking at this picture we can say that CH3, that there's poor energy match because the orbitals around fluorine are much smaller than those around CH3.



Professor Michael McBride: Let's look at the next slide here. Is it this one? No. We're going to get to that. If you looked at ones that were very badly mismatched, then the favorable and unfavorable combinations would be very dramatic. Actually, you've already seen that in H-F. Okay, we have one minute to start this. Okay, so we're going to look at how CH3-F behaves like H-F. Right? Both of them are acids. So first we'll look at H-F, and next time we'll go on to CH3-F. So here's H-F. So you have to bring -- here's a low vacant orbital. We've talked about that ad nauseam. Okay? Now you want to get good overlap. From what direction will another orbital come, in order to get good overlap, without the nuclei getting too close together? Obviously it'll come from off in the right, where this orbital is big, where you can get a lot of overlap without getting close to the nuclei. So if you had something with a high-energy pair of electrons, it would come and it would overlap, from the right; something like OH-. And you'd draw a curved arrow, to show those electrons; the high HOMO of OH-, being stabilized by the vacant orbital of H-F. And how would you draw the curved arrow? Where would it start, where would it end? To show the electrons of OH- forming a bond with -- OH bond?



[Students speak over one another]



Professor Michael McBride: You'd start from the pair of electrons that you're talking about, and you'd end between H and O. Right? But that means -- this is really important -- that means you're putting electrons, putting electron density, into this orbital. Right? What effect does putting electrons in that orbital have on the H-F bond? Okay, the curved arrows, blah-blah; we did that. Sherwin?



Student: It's like it pushes them out.



Professor Michael McBride: It's what? What do you call that orbital? What's the name of the orbital, that we're putting electrons into?



Student: 1s for --



Professor Michael McBride: It's σ*. It's mostly 1s on Hydrogen, but it's σ*. What does * mean?



Students: Antibonding.



Professor Michael McBride: What does it mean if you put electrons in?



Student: The bond breaks.



Professor Michael McBride: The bond will break. Right? It has that antibonding node. So electrons that go into that will get more stable if the bond breaks. So we're going to draw another curved arrow, like that. So we make a new bond between H and O, but we lose the bond between H and F. So there's our product. That's an acid-base reaction, and it showed H-F acting as an acid; not because it gave H+. H+ never appears in here. Right? What happened is you make a bond and break a bond to the hydrogen. So it's an acid-base reaction. And the same thing, we'll show next time, happens with CH3-F. Okay.



[end of transcript]

Lecture 17
Reaction Analogies and Carbonyl Reactivity
Play Video
Reaction Analogies and Carbonyl Reactivity


Continuing the examination of molecular orbital theory as a predictor of chemical reactivity, this lecture focuses on the close analogy among seemingly disparate organic chemistry reactions: acid-base, SN2 substitution, and E2 elimination. All these reactions involve breaking existing bonds where LUMOs have antibonding nodes while new bonds are being formed. The three-stage oxidation of ammonia by elemental chlorine is analyzed in the same terms. The analysis is extended to the reactivity of the carbonyl group and predicts the trajectory for attack by a high HOMO. This predicted trajectory was validated experimentally by Bürgi and Dunitz, who compared numerous crystal structures determined by X-ray diffraction.



Transcript



October 13, 2008



Professor Michael McBride: Okay, so this is what we looked at last time. We looked at an acid-base reaction from a new point of view, from the point of view of bringing up some HOMO to overlap with this unusually low LUMO. Why is it unusually low, the LUMO, σ*, of H-F? Why is it unusually low? Is the overlap bad?



[Students speak over one another]



Professor Michael McBride: Does the hybrid orbital on fluorine and the s orbital on hydrogen, does that have bad overlap, or does it overlap well? Points right at it, good overlap. So that's not the reason it's unusual. What's the reason that the energy of this vacant orbital is unusual? Dana?



Student: The unusual atomic energy of fluorine?



Professor Michael McBride: Right, because fluorine started very low; so that went down, when it mixed with hydrogen. Hydrogen went up, but it didn't go up very far, because the energy match was bad. So that's this LUMO, the unfavorable combination of fluorine and hydrogen. Which is it big on, hydrogen or fluorine, the one that went up? Hydrogen, right? The good one is mostly on fluorine; the bad one, σ*, mostly on hydrogen. And it has that node shown as an antibonding node; σ*, the star means antibonding. So in OH-, with its high energy occupied orbital -- now wait a second, why is an unshared pair on oxygen high energy? I thought oxygen had a big nuclear charge, which would make electrons stable, not have them be high. Why is OH- high? Obinna, do you have an idea about that? What is it about OH-, the unshared pair that makes the electrons unusually high in energy?



Student: Because it didn't mix.



Professor Michael McBride: Can't hear.



Student: It didn't mix.



Professor Michael McBride: Because it didn't mix with anything else. Right? It's not like a bond that went down. We're comparing it to C-H or C-C σ bonds that went down in energy. So even though oxygen started low, it didn't go down. Right? So it's left unusually high. And there's actually another reason that this one is unusually high. Josh?



Student: Negative charge on --



Professor Michael McBride: Negative charge on oxygen doesn't make electrons happy. Okay, so it comes along, mixes with that. In the process of putting electrons into that orbital, it makes the bond. But also it makes it antibonding between hydrogen and fluorine. So we draw that second curved arrow that shows the fluorine leaving with the electrons, that it already had most of. The electrons it leaves with -- not that you can really label electrons and say this one is this, this one is this. But remember, fluorine mixed with hydrogen went down. It already has most of those electrons that came from hydrogen. Now it has them all when it breaks away. It breaks away as an ion, F-. Okay, and that's an acid-base reaction; a more general interpretation of acid-base, than just H+ reacting with OH-.



Okay, now let's generalize a little bit further. So notice that besides creating a new bond, which would happen with H+, mixing the HOMO with this LUMO can break a bond, right? If the LUMO, shown there, has an antibonding node. Putting in electrons means that those electrons will be more stable if fluorine moves away. Right? So you not only make a bond, you also break a bond. So it's a make-and-break situation; whereas H+ plus OH- is just a make situation.



Okay, now let's look at F-CH3. Right? And look at the orbital, and I'm going to superimpose on it the corresponding LUMO of F-CH3. Watch it change. The carbon-fluorine bond is longer than the carbon-hydrogen bond. So it moved away. But the shape, especially around the fluorine, is almost identical. I'll back up and do the change again. Right? So it's going to have very similar properties to what H-F did. So if I move it over, you see it's going to be attacked by something coming here to overlap with that orbital. Some high HOMO will mix with the low LUMO. And here we show OH- doing it. And then what will happen? Remember, we have this antibonding node here. What happens when electrons go into that orbital? Kate?



Student: It breaks the bond.



Professor Michael McBride: Kate?



Student: It breaks. The fluorine is going --



Professor Michael McBride: Yes, the fluorine will take electrons and move away, like that. Right? So we get fluoride and methanol. Right? And this reaction, which we'll study at the beginning of next semester -- and we'll mention a little bit more as we go along this semester, but systematically we'll talk about it next semester -- is called SN2 Substitution. But notice, it's, from the point of view of orbitals, exactly the same make-and-break situation that you had with OH- reacting with H-F. It's the same reaction, just the names have been changed. Right? So you could, in fact, have called the reaction of OH- with H-F, you could've called that a substitution. It could've been SN2 Substitution. The difference would be that the atom on which the substitution takes place, the one that gets attacked and then loses the F-, is hydrogen rather than carbon. So you could call both reactions substitution, or you could call them both acid-base reactions. Right? So that's the way sciences, as I said last time, progress, by seeing generality, where things were thought to be different before.



Okay, so we have these two reactions, acid-base reaction, SN2 Substitution. What would happen if we had ethyl fluoride, sometimes known as ethanol -- improperly labeled, right? -- earlier today. Okay, so that's the LUMO of ethanol [correction: ethyl fluoride]. Now that doesn't look exactly like the LUMO we just calculated of ethanol [correction: ethyl fluoride]. But remember it was the LUMO+3. Wasn't it that we found? It was either plus two or plus -- it was plus three wasn't it? The LUMO+3, remember, had this characteristic, that it's a favorable mixture of two simple pairwise orbitals. That one is what? What's that orbital, if you just saw that?



Student: [inaudible]



Professor Michael McBride: Josh?



Student: It's the same thing as the other one.



Professor Michael McBride: Ah-ha, it's the same thing as up there. Right? It's σ* C-H, instead of C-F. Well it's certainly σ*, but it's of H-C, not of F-C. Right? Everybody got that? So but it's σ*, that's the important thing. It's the antibonding combination, σ* C-H. Okay, and what's the other thing that goes into this LUMO, or actually the LUMO+3? It's this. What's that? Right? Angela?



Student: It's the [inaudible]



Professor Michael McBride: Can you give it a name?



Student: It's the σ* of [inaudible]



Professor Michael McBride: If you name just two atoms, what is it? Speak up loudly so everybody can hear.



Student: σ* C-F.



Professor Michael McBride: σ* C-F. Everybody see that? Okay, and I say it's the favorable mixture of these two, because they're put together so as to be bonding between them. Right? So there were two vacant orbitals, σ* C-F and σ*C-H, and they mixed, and the favorable combination, where they overlapped favorably between those two, is the one we're looking at. It's still vacant, it's an unoccupied molecular orbital. But that's the one we're looking at here. Okay, now how should it react? Well that one looks like the top right, except that the colors are changed. Right? So how about doing SN2 substitution? Now Russell cringed when I did that. What's wrong with doing SN2 substitution?



Student: We have to get a favorable overlap with --



Professor Michael McBride: Ah-ha! what's hidden here is that bit. It's going to be hard to overlap with the blue on top, without having an unfavorable overlap with the red on the bottom. So there's another way you could do it. You could attack here, right? And get a big orbital, a big lobe, all of the same sign. So there you're going to get good overlap. So might it do that? So there are two possibilities. It could do the same as the top, it could do SN2 substitution, or it could put electrons into this LUMO by attacking from this second arrow position. And how do you know which one it does?



Students: Experiments.



Professor Michael McBride: You do experiment. So this is lore, right? And when we get to studying these reactions systematically, you'll see that both things can happen. They're both possibilities, and you can make little tweaks in solvent or base and things like that, what HOMO you're using, that'll make it go a little more one way or a little more the other. But that, of course, takes time to do all the experiments and to hear about them. But anyhow, there are these two possibilities.



Now let's suppose it does the second one. Now how did I set it up so that that would be favorable? I didn't use the true LUMO of the lowest energy, of the molecule, which I just calculated for you at the beginning of class. What I did was use that LUMO+3, but made it the LUMO. And the way I made it the LUMO was to stretch bonds. I stretched the C-H bond and I stretched the C-F bond. Right? And then if I -- in this particular phase of vibration of the atoms, this one is in fact, the LUMO. Okay, now why did I want to show you this? Look what happens if you do that bottom thing. Okay, so we bring OH- up, and the electrons go into this and make a bond between H and O. But there's a node here, that antibonding node. What does that mean, if you put electrons into this orbital? Lexy, what do you say?



Student: It'll break.



Professor Michael McBride: It'll break where?



Student: [inaudible]



Professor Michael McBride: Between the hydrogen and the carbon. Okay, so we're going to -- we'll draw a second curved arrow that shows those electrons pulling out. Okay? Now where do they go? Look here, there's an atomic orbital node there. That's not an antibonding node because it's the node of the p orbitals on the carbons. Everybody see that that's not antibonding? Because it doesn't go between the atoms, it goes through the atoms. So, in fact, if you look at that blue and red lobe in the middle, they're a little bit distorted. But what do they look like? If you look just at the red and blue in the middle? Lucas?



Student: p orbitals.



Professor Michael McBride: Not p. Greek.



Student: Oh π.



Professor Michael McBride: It's the π orbital, made up of 2p orbitals; the p orbital of the carbon on the top, the p orbital of the carbon on the bottom. And you put electrons in that and you get a double bond between the carbons. Okay? Okay, so there's a π bonding between the carbons. That means we can draw a second arrow that takes the electrons out of the bond that's breaking and puts them between the two carbons to make a double bond. Okay? But now there's another node in this LUMO; there, that antibonding node. What happens with respect to that, when we put electrons in this LUMO? Rick, what do you say?



Student: The C is going to break the bond between the C and the F.



Professor Michael McBride: The bond between C and F will break. How will we draw a curved arrow to show that?



Student: You would draw a curved arrow starting at the C and ending at the --



Professor Michael McBride: Not starting at the C. Which are the electrons that are leaving? You start with the electrons that are changing, right?



Student: Yes.



Professor Michael McBride: So where are they, in this diagram here? Here, tell me when to stop, when I hit the electrons?



Student: Stop.



Professor Michael McBride: Not here. That's the one that's building up electrons, right?



Student: There?



[Laughter]



Professor Michael McBride: These are the electrons that are leaving. And where do I draw the curved arrow to, where are they going? Okay. On to the F. So that's what happens. When the O-H comes up, you get water, you get ethylene, the double bond, and fluoride. So this is called, as you'll see -- just after we talk about SN2 substitution, we'll talk about E2 elimination. Right? And that's what this reaction is. Right? But that's just an old category. Right? If we want to make it -- what it is, is the same kind of HOMO/LUMO interaction, except this time we make two bonds and break two bonds, because of where the nodes were, in the LUMO that was being attacked. Right? There were two antibonding nodes. We break two bonds. Okay, does anybody have a question about this? Do you see how it works? Lucas?



Student: Why did you stretch C-H again?



Student: Why did I do it? Because I knew the answer. I knew that this reaction does this E2 elimination, because when I studied organic chemistry we did E2 elimination and never talked about orbitals at all. Okay? So I knew where I was going. I knew what really happened. So I said, how does that happen, that this is the orbital that actually gets attacked? And it's, in fact, the LUMO+3. Right? Why do the electrons go into that? And then I saw that in a certain phase of vibration of the molecule, when this bond is long, and when this bond is stretched, then it is the LUMO. So when the OH is attacking, the OH- is attacking, it's attacking at a time that those bonds are stretched at the same time. Right? And then they break. Okay? It's a little bit of chicken and the egg as to whether the electrons going in make them stretch, or their stretching allows the electrons to go in. Right? But anyhow, I knew what the answer was, which you wouldn't have known, if you had just looked at the orbitals for the minimum energy geometry of the molecule.



Okay, now let's look, let's generalize this to a complicated, or fairly complicated, reaction scheme, at least by our lights. So three ammonia molecules react with chlorine. That's called oxidation of ammonia, and it generates hydrazine, the nitrogen-nitrogen bond, or actually hydrazinium -- it's a cation -- chloride, and ammonium chloride. That looks pretty complicated. There's a lot of changes going on. How does it happen? Could we figure it out? I think you can. Okay, so in the starting material I want to find a high HOMO. Where do I find it? Sherwin, can you help us? What do I look -- I want to find a high HOMO in this starting material. The starting materials are ammonia and the chlorine molecule. So what about them could give you a high, unusually high-occupied orbital?



Student: The unmixed atomic orbital --



Professor Michael McBride: Say it clearly.



Student: The unmixed atomic orbital in ammonia.



Professor Michael McBride: Yes, the unshared pair of ammonia. Right? N has a bigger nuclear change than C. So you might expect the electrons to be low in energy. But it never mixed with anything, so it didn't go down to become a σ bond. So, unshared pair on ammonia. I'll take that. Okay. Now we have to react it with some -- and unshared pairs are often denoted n, little n. And I actually don't know why that is. It'd be fun to find out. Okay, now we need a low LUMO. Anybody got an idea for what a low LUMO might be in this system? What kind of things give rise to low LUMOs?



Student: Electronegative things.



Professor Michael McBride: Sam?



Student: Electronegative --



Professor Michael McBride: Electronegative nuclei. Right? Big nuclear charges. You know, people always say electronegative, but that's the cart and the horse. Why is something electronegative? That means it's a good place for electrons to go. Right? It's because the nuclear charge is big; that's what's fundamental. Okay, where do we find that here?



Student: In the chlorine.



Professor Michael McBride: The chlorine. Now which orbital, associated with the chlorines, is going to be unusually low in energy? Does it have a vacant atomic orbital? Bear in mind what's -- chlorine is in the last column, right? So it's just missing one electron. So a chlorine atom has three unshared pairs and one odd electron, which has come with the other chlorine to make that bond. Right? So after the Cl2 forms, where's the vacant orbital? Anybody see? We had two chlorines, each with this odd electron. They came together and you got a σ bond. Where's the low LUMO? Chris?



Student: It's going be like a 4s atomic orbital.



Professor Michael McBride: No. Kevin?



Student: σ*.



Professor Michael McBride: σ*, right? They each had one electron. Come together, they went down. But there was a σ* that went up. σ* is usually high energy, but not in this case because, as Sam pointed out, the nuclear charge is big, so it's low. So σ* on chlorine. Okay? So what kind of reaction are we going to have? If we mix σ* of chlorine with the unshared pair on nitrogen? Have you ever seen a reaction like this before, where you have a σ* mixing with an unshared pair? Katelyn?



Student: On the last slide.



Professor Michael McBride: On the last slide we saw lots of examples. Right? And so Katelyn, what happens? Tell me how to draw a curved arrow here.



Student: The electrons from the unshared -- no.



Professor Michael McBride: The electrons on the unshared pair of nitrogen, that's the high HOMO. Between what?



Student: The two Cl's?



Professor Michael McBride: Between the nitrogen and the chlorine. They mix with that vacant orbital. But the ones in the chlorine, now that's an antibonding orbital, σ*, so a chloride leaves. So it's exactly the same reaction, make and break, that we saw before. Okay? So we get ammonium chloride "chlorammonium" and Cl-, plus/minus, net neutral. We started neutral, we're still neutral. Okay? And now we have one of the products, the chloride up above we have. Okay, but we also have this molecule. Now it has a low LUMO, so it's reactive. What is the low LUMO? Yoonjoo?



Student: The N-H bond.



Professor Michael McBride: That's not what I would've said. It turns out to be right, but it's not what I would've said. There are going to be N-H -- it's got to be vacant, so what's the name of the orbital?



Student: σ*



Professor Michael McBride: σ* N-H. Are there any other σ*s here? Russell?



Student: Cl.



Professor Michael McBride: Cl. Which do you think is going to be lower, σ* N-Cl or σ* N-H? Kevin?



Student: I would guess N-Cl.



Student: I would have too. So the lowest LUMO is for some HOMO. Now what are we going to use for a HOMO? We have to have another reagent for this to react with. Well fortunately that σ* N-H is what we're actually going to use. Okay? So the high HOMO. We had three ammonias to begin with. So we can use another one; same deal again, right? So the best reaction, I think, would be for the unshared pair to attack the chlorine; at least that's something that you could imagine. If that happened, what would the product be? If NH3 attacked Cl, it would make an NH3-Cl bond, and lose what? What Russell?



Student: NH3.



Professor Michael McBride: So what about it?



Student: That's what it could make there.



Professor Michael McBride: It could happen 'til the cows come home, but the product is the same as the starting material, and no one cares. It would exchange which nitrogen is attached to the chlorine. But unless you label them with isotopes or something, you'd have no way of knowing. So that probably does happen, or at least I suspect it happens some of the time. Right? But you don't care. So now we resort to Yoonjoo's suggestion that we'll use a different one so we get a different product. Okay? So we'll attack the N-H bond. Now Corey, tell us how to draw a curved arrow to show this. Where's the curved arrow going to start?



Student: On the lone pair.



Professor Michael McBride: On the lone pair of the nitrogen. And where will it end?



Student: In between the N and the H.



Professor Michael McBride: In between the N and the H. And what else will happen?



Student: The bond will break.



Professor Michael McBride: Which bond? What's going to take off with the electrons? Can you see? Nothing. Well, that is, what you lose -- it becomes an unshared pair on nitrogen. N was N+, right? Now it gets the electrons that formerly it was sharing with H. So it ends on nitrogen. Right? So there's the product. And now I've highlighted in blue another one of the ultimate products. Right? So now we've made ammonium chloride. Right? We've made one of the products.



But we need the other one. But we have this product, from the second reaction. Now, what do we have for a low LUMO here? Nate, do you have an idea? You have only σ bonds, and an unshared pair on nitrogen, and three unshared pairs on chlorine. Right? But those are all occupied. What's a vacant orbital here? What kind of vacant orbitals do you get in molecules that have only σ bonds?



Student: The p orbitals.



Professor Michael McBride: No, the p orbitals are filled with electrons here. Right? So they're not vacant.



Student: You get a σ* then. You get σ*.



Professor Michael McBride: σ*. Which σ -- you have N-H σ* and you have N-Cl σ*. Which one do you think is lower?



Student: The last time it was the N-H.



Professor Michael McBride: No, it wasn't. The one that reacted last time was the N-H, but it wasn't the lowest. The N-Cl was the lowest. Why didn't we care last time? Because the product, when that reacted, you got the starting material back. So who cared? Right? So this time it's the N-Cl. Right? What's the high HOMO we're going to use?



Student: NH3.



Professor Michael McBride: Ah. Wilson?



Student: The NH3, we still --



Professor Michael McBride: Ah, we got a third one up there still. Right? So bring it in. And there are the curved arrows, and there's the product. Hydrazine -- hydrazinium, it's got a proton and there's a plus charge -- hydrochloride, hydrazinium chloride. Right? So that is the last product. So we've done it. So there are three of these reactions that take place, three cycles of make and break. And notice the difference among them, right? It's always NH3 that attacks. But first it attacks chlorine, then it attacks hydrogen, then it attack nitrogen. But it's the same reaction we talked about on the previous slide, right? Make and break. Okay, now we're going to look at four functional groups: The C double bond O, carbonyl; amide; carboxylic acid; and alkyl lithium. And after that, we'll have done what we wanted to do about functional groups. Although it will still be in your court to do functional groups, because you have these Wikis, to do the same kind of analysis on the functional groups that we're doing here, on these four functional groups. And those are due Thursday. And almost all of them are signed up for. There are just a few more to do. So if you want a choice, get there quickly, because I'll assign them this afternoon, I think, because we're almost at the end, and I have to make sure everything gets covered.



Okay, so the carbonyl group, C=O double bond. This is probably the most important functional group in organic chemistry. Why is it so important? (a) Because it's strong, so it gives stable compounds, low energy compounds. So you have a lot of carbonyl groups. But, paradoxically, it's not only strong, but it reacts easily. So it's easy to make things from carbonyls. Okay, that sounds contradictory, that it should be both stable and reactive; but it is, and we'll see why. So let's look at the shape of the frontier orbitals. There are six valence electron pairs in formaldehyde, shown here from the side. And we could show the molecular orbitals, right?; the Chladni style, plum-pudding molecular orbitals. So the lowest one is 1s, or 2s if you want to count the core. Right? Which is mostly on the oxygen, a little bit on carbon. And then there's the 2px, 2y and 2pz, as plum pudding orbitals. And then there's this 3s orbital and this 3dxy orbital.



But let's look a little more closely at that, from the point of view of pairwise interactions. What things interacted with one another to give these? So let's do the kind of pairwise mixing we've been talking about. There's the HOMO, an older style picture of the HOMO, the one that's called 3dxy, on the right. And this is drawn at a -- it was based on a different calculation and it's drawn at a different contour level. But you can see what it -- what is it mostly, the HOMO? If you had to give a simple name to it, what would you call it, if you neglected the little tiny pieces on the right? What does it look like? Sherwin?



Student: p orbital on oxygen.



Professor Michael McBride: It's a p orbital on oxygen, an unshared pair of oxygen, mostly, although it is mixed a little bit with something else. Now, to see what else it's mixed with, we can go down the list by three steps and look at that orbital, the one that's called 2py. Because it also has that little bit of oxygen in it, but more of the other stuff. Now let's look at nodes in these things. There's nodes that are nodes of atomic orbitals, of the oxygen and of the carbon; they bisect the molecule. Those don't have to do with bonding. Right? But the one on the top also has an antibonding node. Can you see where that is? Shai?



Student: Between the carbon and the oxygen.



Professor Michael McBride: Between the carbon and the oxygen, there's an antibonding node. So this is a favorable and an unfavorable combination of, in the front, oxygen, the front left is oxygen, and in the back right is some sort of C-H bond. So those molecular orbitals are made by combining 2p of oxygen with σ C-H; in fact two σ C-Hs. Okay? Now, which one is lower in energy, the unshared pair on oxygen, or σ C-H? Notice that the oxygen has something going for it, which is a big nuclear charge, but the C-H has something going for it, which is the bonding of the C-H, which lowered it from where carbon and hydrogen atomic orbitals would be. Which one is lower, the 2p of oxygen or σ C-H? Which wins? Can you tell by looking at this? Angela?



Student: σ C-H is lower because the one on the bottom has less nodes than the one on the top.



Professor Michael McBride: Right. The favorable combination, which looks mostly like the lower one, looks mostly like C-H; and the unfavorable one, the one that went up, looks mostly like O. So clearly the σ C-H is lower in energy than the 2p of oxygen. Okay, now let's look at the LUMO, up on top, in the plum pudding column, right? And here's another version of that. And you can see what it is. It's made up of a p orbital on oxygen and a p orbital on carbon. That's an atomic orbital node. It has nothing to do with the bonding. But there is an antibonding node between them. Right? So this is the π* orbital; not the occupied one but the vacant orbital, the LUMO. Right? There's poor overlap because it's π*; that's why it's unusually low. And it's also unusually low compared to C, C-C, because it's got an oxygen in it, a high nuclear charge. So poor energy match; 2p O is lower in energy than the 2p of carbon. Okay? And there -- we're going to show this on the next slide, those two -- but the top one, you see, is the favorable combination of 2p O and 2pz; the one of which this is the unfavorable combination. But it's not the HOMO, right? The HOMO is that unshared pair on oxygen, which is the dominant thing in 3dxy.



Okay, so now let's look at these. There's the C=O π bonding orbital. Now which one should that be big on? Don't look at it. Which one should the π -- the favorable combination of p on oxygen and p on carbon, which should it be big on? There's p on oxygen, p on carbon, they mix; π overlap, not too much overlap, so they don't go so much up and down. But what does this, the bonding one, what does it mostly look like?



Student: Oxygen.



Professor Michael McBride: It should look mostly like oxygen. When you look at this picture it doesn't look that way. That's because of which contour we drew. Because oxygen holds its electrons closer. So by the time you get out to a certain low -- so you get through most of the electron density of oxygen, as you go out one contour by the other of the onion, that's the oxygen; there are more electrons inside. So indeed it does have more electron density on oxygen than it does on carbon, but it doesn't show well in this particular contour. Anyhow, that's the π bonding. And this one is a different unshared pair on Oxygen. And that's mostly a p-rich, hybrid atomic orbital of oxygen, but it's got a little bit of the C-H bonding and it's a favorable combination; there's a little bit of favorable bonding between there. Okay, but the nodes that you see in this, in neither orbital are the nodes antibonding. They're always the nodes that go through atoms. They're part of atomic orbitals. These are bonding orbitals, the occupied ones, as you expect them to be.



Okay, but now let's look at that LUMO, that vacant orbital, the π* orbital. What direction should a high energy HOMO approach this, in order to mix with it? So it has to get good overlap as it comes in, and not run into the other nuclei, or the other electron density associated with other orbitals on the way in. Okay, well here are three possibilities. You could attack from within the plane, from the oxygen end; from within the plane toward the carbon; or from straight up, halfway between carbon and oxygen. How do those look as ways for an orbital to come in and overlap? Suppose it came from the right. Suppose you had a big ball, like OH-, coming in along that arrow from the right? You going to get good overlap? What would you call that situation?



Student: Orthogonal.



Professor Michael McBride: Orthogonal, right? So you're not going to get overlap -- how about if you come from the left, the arrow on the left?



Student: It's also --



Professor Michael McBride: That's also going to be orthogonal. How about if you come from the top? It's not precisely orthogonal, but it's pretty close to orthogonal. You have both positive and negative overlap. So in fact none of those are good. Now here are two other possibilities. You could attack the oxygen on the top, or you could attack the carbon on the top. Which of those do you think looks better?



Student: Carbon.



Professor Michael McBride: Carbon. Why? Angela?



Student: It's bigger.



Professor Michael McBride: Because it's bigger. Okay, so I'd say that one's smaller, so forget that. So come in from the carbon. Okay, so here's a better drawing of that orbital. And we want to come in actually from that direction, at a certain angle; not 90º, with respect to the oxygen-carbon bond. Now why come in from that angle? Why not come straight down on top?



Student: You get less overlap with the oxygen.



Professor Michael McBride: There are the nodes. Right? This is furthest from the node. So you can overlap the nodes you want to overlap with, without getting near the blue ones. Right? So you should come in at a certain angle. Now, there's a prediction. Is it true? So what we're going to do this is rotate this so the orientation corresponds to a graph, that appeared in a paper by Dunitz and Bürgi in 1983, when they were analyzing this. Now what did they do? They looked at X-ray crystal structures of many, many crystals that had both a nitrogen in it -- so the unshared pair -- and they had a carbon-oxygen double bond. And they looked to see, in many, many different structures, how those atoms are arranged, because they must be arranged in a low energy way next to one another. Okay? That's the way they would pack in the crystal.



Okay, so now how is this diagram made? Well notice that the carbon and oxygen are drawn there. So that's the carbon and oxygen, and they're made -- so all these different crystals, some have the group oriented this way, some this way, some this way, some this way. So you rotate all those crystals so that their carbons are superimposed with one another, in all the structures; the carbons are all there, and the oxygen is along this particular line, which lies in the plane of the screen. Right? And then there are things attached to it, which are R groups; an R group -- that is whatever group, hydrogen or carbon or something -- attached to the carbon. There's one in front of the plane and one behind the plane. Okay? So we rotate all these different structures around that carbon-oxygen bond. So those things are all one hiding the other one; the R in front hides the R behind.



And now we look, with that orientation, to see where the nitrogen is nearby. And those dots up on the top show in many different structures, labeled A, B, C, D, E, F, G, H, I -- no J -- K, L. Okay? In all those different structures where's the nitrogen? So in structure G, a particular compound, the nitrogen is where the blue dot is on top. The carbon is -- all the carbons are always there. That's the origin, right? Here. The oxygens are along this line. So the oxygen is along that line and at that position, that distance. And the R groups are bent down a little bit from being planar. One is here and the other one is behind. Okay? Okay, but now let's go G, H, I, K, L. G, H, I, K, L. Do you see what happens? Let's go back. K, I, H, G. These are five different crystal structures. But look again. So the nitrogen comes down, from one structure to the next. What happens to the R's, as the nitrogen comes down?



Student: They move away.



Professor Michael McBride: Right, the R's bend down. So the nitrogen -- the carbon, pardon me, in the middle, becomes pyramidal rather than planar. Back up. What happens to the oxygen as we do this? There's G, H, I, K, L. Right? Okay, so here's what happened. The nitrogen moves down, the Rs bend down, and the oxygen moves away. Why does the oxygen move away from the carbon, from a theoretical, orbital point of view? Where are you putting electrons? Where are you putting the electrons here, when you put them into this LUMO? Into an orbital that's antibonding between carbon and oxygen. So the oxygen stretches away from the carbon. Kevin?



Student: On the right side of the screen, the length units, what is that?



Professor Michael McBride: That's Ångstroms. So it's zero is the origin; that's where the carbon is. And there's one Ångstrom, 1, 2, 3. So they start three -- the nitrogen starts three Ångstroms from the carbon, and as it comes in it gets to be within about 1.6 or something. And I forget. I'm guessing that an actual bond between that nitrogen and carbon is probably the order of 1.3 or something; I'm not actually sure it is. So they haven't quite bonded. But it's following the pathway of reaction, when you look at a lot of these different structures. Okay? Now that angle, 110º, along the path that they follow, is called the Bürgi-Dunitz Angle; that's the guys who did the paper, Hans-Beat Bürgi and Jack Dunitz in Switzerland. So there is a bunch of what we might consider snapshots of a reaction actually taking place. And you can see that there's a well-defined angle from which the nitrogen approaches that LUMO, as it mixes with it.



Now, resonance is intramolecular HOMO/LUMO mixing. We've looked here at intermolecular, which is reaction. But next time we'll go on to this slide and look at the analogous thing that happens within a molecule. Okay.



[end of transcript]

Lecture 18
Amide, Carboxylic Acid and Alkyl Lithium
Play Video
Amide, Carboxylic Acid and Alkyl Lithium


Continuing the examination of molecular orbital theory as a predictor of chemical reactivity, this lecture focuses on the close analogy among seemingly disparate organic chemistry reactions: acid-base, SN2 substitution, and E2 elimination. All these reactions involve breaking existing bonds where LUMOs have antibonding nodes while new bonds are being formed. The three-stage oxidation of ammonia by elemental chlorine is analyzed in the same terms. The analysis is extended to the reactivity of the carbonyl group and predicts the trajectory for attack by a high HOMO. This predicted trajectory was validated experimentally by Bürgi and Dunitz, who compared numerous crystal structures determined by X-ray diffraction.




Transcript



October 13, 2008




Professor Michael McBride: Okay, so this is what we looked at last time. We looked at an acid-base reaction from a new point of view, from the point of view of bringing up some HOMO to overlap with this unusually low LUMO. Why is it unusually low, the LUMO, σ*, of H-F? Why is it unusually low? Is the overlap bad?



[Students speak over one another]



Professor Michael McBride: Does the hybrid orbital on fluorine and the s orbital on hydrogen, does that have bad overlap, or does it overlap well? Points right at it, good overlap. So that's not the reason it's unusual. What's the reason that the energy of this vacant orbital is unusual? Dana?



Student: The unusual atomic energy of fluorine?



Professor Michael McBride: Right, because fluorine started very low; so that went down, when it mixed with hydrogen. Hydrogen went up, but it didn't go up very far, because the energy match was bad. So that's this LUMO, the unfavorable combination of fluorine and hydrogen. Which is it big on, hydrogen or fluorine, the one that went up? Hydrogen, right? The good one is mostly on fluorine; the bad one, σ*, mostly on hydrogen. And it has that node shown as an antibonding node; σ*, the star means antibonding. So in OH-, with its high energy occupied orbital -- now wait a second, why is an unshared pair on oxygen high energy? I thought oxygen had a big nuclear charge, which would make electrons stable, not have them be high. Why is OH- high? Obinna, do you have an idea about that? What is it about OH-, the unshared pair that makes the electrons unusually high in energy?



Student: Because it didn't mix.



Professor Michael McBride: Can't hear.



Student: It didn't mix.



Professor Michael McBride: Because it didn't mix with anything else. Right? It's not like a bond that went down. We're comparing it to C-H or C-C σ bonds that went down in energy. So even though oxygen started low, it didn't go down. Right? So it's left unusually high. And there's actually another reason that this one is unusually high. Josh?



Student: Negative charge on --



Professor Michael McBride: Negative charge on oxygen doesn't make electrons happy. Okay, so it comes along, mixes with that. In the process of putting electrons into that orbital, it makes the bond. But also it makes it antibonding between hydrogen and fluorine. So we draw that second curved arrow that shows the fluorine leaving with the electrons, that it already had most of. The electrons it leaves with -- not that you can really label electrons and say this one is this, this one is this. But remember, fluorine mixed with hydrogen went down. It already has most of those electrons that came from hydrogen. Now it has them all when it breaks away. It breaks away as an ion, F-. Okay, and that's an acid-base reaction; a more general interpretation of acid-base, than just H+ reacting with OH-.



Okay, now let's generalize a little bit further. So notice that besides creating a new bond, which would happen with H+, mixing the HOMO with this LUMO can break a bond, right? If the LUMO, shown there, has an antibonding node. Putting in electrons means that those electrons will be more stable if fluorine moves away. Right? So you not only make a bond, you also break a bond. So it's a make-and-break situation; whereas H+ plus OH- is just a make situation.



Okay, now let's look at F-CH3. Right? And look at the orbital, and I'm going to superimpose on it the corresponding LUMO of F-CH3. Watch it change. The carbon-fluorine bond is longer than the carbon-hydrogen bond. So it moved away. But the shape, especially around the fluorine, is almost identical. I'll back up and do the change again. Right? So it's going to have very similar properties to what H-F did. So if I move it over, you see it's going to be attacked by something coming here to overlap with that orbital. Some high HOMO will mix with the low LUMO. And here we show OH- doing it. And then what will happen? Remember, we have this antibonding node here. What happens when electrons go into that orbital? Kate?



Student: It breaks the bond.



Professor Michael McBride: Kate?



Student: It breaks. The fluorine is going --



Professor Michael McBride: Yes, the fluorine will take electrons and move away, like that. Right? So we get fluoride and methanol. Right? And this reaction, which we'll study at the beginning of next semester -- and we'll mention a little bit more as we go along this semester, but systematically we'll talk about it next semester -- is called SN2 Substitution. But notice, it's, from the point of view of orbitals, exactly the same make-and-break situation that you had with OH- reacting with H-F. It's the same reaction, just the names have been changed. Right? So you could, in fact, have called the reaction of OH- with H-F, you could've called that a substitution. It could've been SN2 Substitution. The difference would be that the atom on which the substitution takes place, the one that gets attacked and then loses the F-, is hydrogen rather than carbon. So you could call both reactions substitution, or you could call them both acid-base reactions. Right? So that's the way sciences, as I said last time, progress, by seeing generality, where things were thought to be different before.



Okay, so we have these two reactions, acid-base reaction, SN2 Substitution. What would happen if we had ethyl fluoride, sometimes known as ethanol -- improperly labeled, right? -- earlier today. Okay, so that's the LUMO of ethanol [correction: ethyl fluoride]. Now that doesn't look exactly like the LUMO we just calculated of ethanol [correction: ethyl fluoride]. But remember it was the LUMO+3. Wasn't it that we found? It was either plus two or plus -- it was plus three wasn't it? The LUMO+3, remember, had this characteristic, that it's a favorable mixture of two simple pairwise orbitals. That one is what? What's that orbital, if you just saw that?



Student: [inaudible]



Professor Michael McBride: Josh?



Student: It's the same thing as the other one.



Professor Michael McBride: Ah-ha, it's the same thing as up there. Right? It's σ* C-H, instead of C-F. Well it's certainly σ*, but it's of H-C, not of F-C. Right? Everybody got that? So but it's σ*, that's the important thing. It's the antibonding combination, σ* C-H. Okay, and what's the other thing that goes into this LUMO, or actually the LUMO+3? It's this. What's that? Right? Angela?



Student: It's the [inaudible]



Professor Michael McBride: Can you give it a name?



Student: It's the σ* of [inaudible]



Professor Michael McBride: If you name just two atoms, what is it? Speak up loudly so everybody can hear.



Student: σ* C-F.



Professor Michael McBride: σ* C-F. Everybody see that? Okay, and I say it's the favorable mixture of these two, because they're put together so as to be bonding between them. Right? So there were two vacant orbitals, σ* C-F and σ*C-H, and they mixed, and the favorable combination, where they overlapped favorably between those two, is the one we're looking at. It's still vacant, it's an unoccupied molecular orbital. But that's the one we're looking at here. Okay, now how should it react? Well that one looks like the top right, except that the colors are changed. Right? So how about doing SN2 substitution? Now Russell cringed when I did that. What's wrong with doing SN2 substitution?



Student: We have to get a favorable overlap with --



Professor Michael McBride: Ah-ha! what's hidden here is that bit. It's going to be hard to overlap with the blue on top, without having an unfavorable overlap with the red on the bottom. So there's another way you could do it. You could attack here, right? And get a big orbital, a big lobe, all of the same sign. So there you're going to get good overlap. So might it do that? So there are two possibilities. It could do the same as the top, it could do SN2 substitution, or it could put electrons into this LUMO by attacking from this second arrow position. And how do you know which one it does?



Students: Experiments.



Professor Michael McBride: You do experiment. So this is lore, right? And when we get to studying these reactions systematically, you'll see that both things can happen. They're both possibilities, and you can make little tweaks in solvent or base and things like that, what HOMO you're using, that'll make it go a little more one way or a little more the other. But that, of course, takes time to do all the experiments and to hear about them. But anyhow, there are these two possibilities.



Now let's suppose it does the second one. Now how did I set it up so that that would be favorable? I didn't use the true LUMO of the lowest energy, of the molecule, which I just calculated for you at the beginning of class. What I did was use that LUMO+3, but made it the LUMO. And the way I made it the LUMO was to stretch bonds. I stretched the C-H bond and I stretched the C-F bond. Right? And then if I -- in this particular phase of vibration of the atoms, this one is in fact, the LUMO. Okay, now why did I want to show you this? Look what happens if you do that bottom thing. Okay, so we bring OH- up, and the electrons go into this and make a bond between H and O. But there's a node here, that antibonding node. What does that mean, if you put electrons into this orbital? Lexy, what do you say?



Student: It'll break.



Professor Michael McBride: It'll break where?



Student: [inaudible]



Professor Michael McBride: Between the hydrogen and the carbon. Okay, so we're going to -- we'll draw a second curved arrow that shows those electrons pulling out. Okay? Now where do they go? Look here, there's an atomic orbital node there. That's not an antibonding node because it's the node of the p orbitals on the carbons. Everybody see that that's not antibonding? Because it doesn't go between the atoms, it goes through the atoms. So, in fact, if you look at that blue and red lobe in the middle, they're a little bit distorted. But what do they look like? If you look just at the red and blue in the middle? Lucas?



Student: p orbitals.



Professor Michael McBride: Not p. Greek.



Student: Oh π.



Professor Michael McBride: It's the π orbital, made up of 2p orbitals; the p orbital of the carbon on the top, the p orbital of the carbon on the bottom. And you put electrons in that and you get a double bond between the carbons. Okay? Okay, so there's a π bonding between the carbons. That means we can draw a second arrow that takes the electrons out of the bond that's breaking and puts them between the two carbons to make a double bond. Okay? But now there's another node in this LUMO; there, that antibonding node. What happens with respect to that, when we put electrons in this LUMO? Rick, what do you say?



Student: The C is going to break the bond between the C and the F.



Professor Michael McBride: The bond between C and F will break. How will we draw a curved arrow to show that?



Student: You would draw a curved arrow starting at the C and ending at the --



Professor Michael McBride: Not starting at the C. Which are the electrons that are leaving? You start with the electrons that are changing, right?



Student: Yes.



Professor Michael McBride: So where are they, in this diagram here? Here, tell me when to stop, when I hit the electrons?



Student: Stop.



Professor Michael McBride: Not here. That's the one that's building up electrons, right?



Student: There?



[Laughter]



Professor Michael McBride: These are the electrons that are leaving. And where do I draw the curved arrow to, where are they going? Okay. On to the F. So that's what happens. When the O-H comes up, you get water, you get ethylene, the double bond, and fluoride. So this is called, as you'll see -- just after we talk about SN2 substitution, we'll talk about E2 elimination. Right? And that's what this reaction is. Right? But that's just an old category. Right? If we want to make it -- what it is, is the same kind of HOMO/LUMO interaction, except this time we make two bonds and break two bonds, because of where the nodes were, in the LUMO that was being attacked. Right? There were two antibonding nodes. We break two bonds. Okay, does anybody have a question about this? Do you see how it works? Lucas?



Student: Why did you stretch C-H again?



Student: Why did I do it? Because I knew the answer. I knew that this reaction does this E2 elimination, because when I studied organic chemistry we did E2 elimination and never talked about orbitals at all. Okay? So I knew where I was going. I knew what really happened. So I said, how does that happen, that this is the orbital that actually gets attacked? And it's, in fact, the LUMO+3. Right? Why do the electrons go into that? And then I saw that in a certain phase of vibration of the molecule, when this bond is long, and when this bond is stretched, then it is the LUMO. So when the OH is attacking, the OH- is attacking, it's attacking at a time that those bonds are stretched at the same time. Right? And then they break. Okay? It's a little bit of chicken and the egg as to whether the electrons going in make them stretch, or their stretching allows the electrons to go in. Right? But anyhow, I knew what the answer was, which you wouldn't have known, if you had just looked at the orbitals for the minimum energy geometry of the molecule.



Okay, now let's look, let's generalize this to a complicated, or fairly complicated, reaction scheme, at least by our lights. So three ammonia molecules react with chlorine. That's called oxidation of ammonia, and it generates hydrazine, the nitrogen-nitrogen bond, or actually hydrazinium -- it's a cation -- chloride, and ammonium chloride. That looks pretty complicated. There's a lot of changes going on. How does it happen? Could we figure it out? I think you can. Okay, so in the starting material I want to find a high HOMO. Where do I find it? Sherwin, can you help us? What do I look -- I want to find a high HOMO in this starting material. The starting materials are ammonia and the chlorine molecule. So what about them could give you a high, unusually high-occupied orbital?



Student: The unmixed atomic orbital --



Professor Michael McBride: Say it clearly.



Student: The unmixed atomic orbital in ammonia.



Professor Michael McBride: Yes, the unshared pair of ammonia. Right? N has a bigger nuclear change than C. So you might expect the electrons to be low in energy. But it never mixed with anything, so it didn't go down to become a σ bond. So, unshared pair on ammonia. I'll take that. Okay. Now we have to react it with some -- and unshared pairs are often denoted n, little n. And I actually don't know why that is. It'd be fun to find out. Okay, now we need a low LUMO. Anybody got an idea for what a low LUMO might be in this system? What kind of things give rise to low LUMOs?



Student: Electronegative things.



Professor Michael McBride: Sam?



Student: Electronegative --



Professor Michael McBride: Electronegative nuclei. Right? Big nuclear charges. You know, people always say electronegative, but that's the cart and the horse. Why is something electronegative? That means it's a good place for electrons to go. Right? It's because the nuclear charge is big; that's what's fundamental. Okay, where do we find that here?



Student: In the chlorine.



Professor Michael McBride: The chlorine. Now which orbital, associated with the chlorines, is going to be unusually low in energy? Does it have a vacant atomic orbital? Bear in mind what's -- chlorine is in the last column, right? So it's just missing one electron. So a chlorine atom has three unshared pairs and one odd electron, which has come with the other chlorine to make that bond. Right? So after the Cl2 forms, where's the vacant orbital? Anybody see? We had two chlorines, each with this odd electron. They came together and you got a σ bond. Where's the low LUMO? Chris?



Student: It's going be like a 4s atomic orbital.



Professor Michael McBride: No. Kevin?



Student: σ*.



Professor Michael McBride: σ*, right? They each had one electron. Come together, they went down. But there was a σ* that went up. σ* is usually high energy, but not in this case because, as Sam pointed out, the nuclear charge is big, so it's low. So σ* on chlorine. Okay? So what kind of reaction are we going to have? If we mix σ* of chlorine with the unshared pair on nitrogen? Have you ever seen a reaction like this before, where you have a σ* mixing with an unshared pair? Katelyn?



Student: On the last slide.



Professor Michael McBride: On the last slide we saw lots of examples. Right? And so Katelyn, what happens? Tell me how to draw a curved arrow here.



Student: The electrons from the unshared -- no.



Professor Michael McBride: The electrons on the unshared pair of nitrogen, that's the high HOMO. Between what?



Student: The two Cl's?



Professor Michael McBride: Between the nitrogen and the chlorine. They mix with that vacant orbital. But the ones in the chlorine, now that's an antibonding orbital, σ*, so a chloride leaves. So it's exactly the same reaction, make and break, that we saw before. Okay? So we get ammonium chloride "chlorammonium" and Cl-, plus/minus, net neutral. We started neutral, we're still neutral. Okay? And now we have one of the products, the chloride up above we have. Okay, but we also have this molecule. Now it has a low LUMO, so it's reactive. What is the low LUMO? Yoonjoo?



Student: The N-H bond.



Professor Michael McBride: That's not what I would've said. It turns out to be right, but it's not what I would've said. There are going to be N-H -- it's got to be vacant, so what's the name of the orbital?



Student: σ*



Professor Michael McBride: σ* N-H. Are there any other σ*s here? Russell?



Student: Cl.



Professor Michael McBride: Cl. Which do you think is going to be lower, σ* N-Cl or σ* N-H? Kevin?



Student: I would guess N-Cl.



Student: I would have too. So the lowest LUMO is for some HOMO. Now what are we going to use for a HOMO? We have to have another reagent for this to react with. Well fortunately that σ* N-H is what we're actually going to use. Okay? So the high HOMO. We had three ammonias to begin with. So we can use another one; same deal again, right? So the best reaction, I think, would be for the unshared pair to attack the chlorine; at least that's something that you could imagine. If that happened, what would the product be? If NH3 attacked Cl, it would make an NH3-Cl bond, and lose what? What Russell?



Student: NH3.



Professor Michael McBride: So what about it?



Student: That's what it could make there.



Professor Michael McBride: It could happen 'til the cows come home, but the product is the same as the starting material, and no one cares. It would exchange which nitrogen is attached to the chlorine. But unless you label them with isotopes or something, you'd have no way of knowing. So that probably does happen, or at least I suspect it happens some of the time. Right? But you don't care. So now we resort to Yoonjoo's suggestion that we'll use a different one so we get a different product. Okay? So we'll attack the N-H bond. Now Corey, tell us how to draw a curved arrow to show this. Where's the curved arrow going to start?



Student: On the lone pair.



Professor Michael McBride: On the lone pair of the nitrogen. And where will it end?



Student: In between the N and the H.



Professor Michael McBride: In between the N and the H. And what else will happen?



Student: The bond will break.



Professor Michael McBride: Which bond? What's going to take off with the electrons? Can you see? Nothing. Well, that is, what you lose -- it becomes an unshared pair on nitrogen. N was N+, right? Now it gets the electrons that formerly it was sharing with H. So it ends on nitrogen. Right? So there's the product. And now I've highlighted in blue another one of the ultimate products. Right? So now we've made ammonium chloride. Right? We've made one of the products.



But we need the other one. But we have this product, from the second reaction. Now, what do we have for a low LUMO here? Nate, do you have an idea? You have only σ bonds, and an unshared pair on nitrogen, and three unshared pairs on chlorine. Right? But those are all occupied. What's a vacant orbital here? What kind of vacant orbitals do you get in molecules that have only σ bonds?



Student: The p orbitals.



Professor Michael McBride: No, the p orbitals are filled with electrons here. Right? So they're not vacant.



Student: You get a σ* then. You get σ*.



Professor Michael McBride: σ*. Which σ -- you have N-H σ* and you have N-Cl σ*. Which one do you think is lower?



Student: The last time it was the N-H.



Professor Michael McBride: No, it wasn't. The one that reacted last time was the N-H, but it wasn't the lowest. The N-Cl was the lowest. Why didn't we care last time? Because the product, when that reacted, you got the starting material back. So who cared? Right? So this time it's the N-Cl. Right? What's the high HOMO we're going to use?



Student: NH3.



Professor Michael McBride: Ah. Wilson?



Student: The NH3, we still --



Professor Michael McBride: Ah, we got a third one up there still. Right? So bring it in. And there are the curved arrows, and there's the product. Hydrazine -- hydrazinium, it's got a proton and there's a plus charge -- hydrochloride, hydrazinium chloride. Right? So that is the last product. So we've done it. So there are three of these reactions that take place, three cycles of make and break. And notice the difference among them, right? It's always NH3 that attacks. But first it attacks chlorine, then it attacks hydrogen, then it attack nitrogen. But it's the same reaction we talked about on the previous slide, right? Make and break. Okay, now we're going to look at four functional groups: The C double bond O, carbonyl; amide; carboxylic acid; and alkyl lithium. And after that, we'll have done what we wanted to do about functional groups. Although it will still be in your court to do functional groups, because you have these Wikis, to do the same kind of analysis on the functional groups that we're doing here, on these four functional groups. And those are due Thursday. And almost all of them are signed up for. There are just a few more to do. So if you want a choice, get there quickly, because I'll assign them this afternoon, I think, because we're almost at the end, and I have to make sure everything gets covered.



Okay, so the carbonyl group, C=O double bond. This is probably the most important functional group in organic chemistry. Why is it so important? (a) Because it's strong, so it gives stable compounds, low energy compounds. So you have a lot of carbonyl groups. But, paradoxically, it's not only strong, but it reacts easily. So it's easy to make things from carbonyls. Okay, that sounds contradictory, that it should be both stable and reactive; but it is, and we'll see why. So let's look at the shape of the frontier orbitals. There are six valence electron pairs in formaldehyde, shown here from the side. And we could show the molecular orbitals, right?; the Chladni style, plum-pudding molecular orbitals. So the lowest one is 1s, or 2s if you want to count the core. Right? Which is mostly on the oxygen, a little bit on carbon. And then there's the 2px, 2y and 2pz, as plum pudding orbitals. And then there's this 3s orbital and this 3dxy orbital.



But let's look a little more closely at that, from the point of view of pairwise interactions. What things interacted with one another to give these? So let's do the kind of pairwise mixing we've been talking about. There's the HOMO, an older style picture of the HOMO, the one that's called 3dxy, on the right. And this is drawn at a -- it was based on a different calculation and it's drawn at a different contour level. But you can see what it -- what is it mostly, the HOMO? If you had to give a simple name to it, what would you call it, if you neglected the little tiny pieces on the right? What does it look like? Sherwin?



Student: p orbital on oxygen.



Professor Michael McBride: It's a p orbital on oxygen, an unshared pair of oxygen, mostly, although it is mixed a little bit with something else. Now, to see what else it's mixed with, we can go down the list by three steps and look at that orbital, the one that's called 2py. Because it also has that little bit of oxygen in it, but more of the other stuff. Now let's look at nodes in these things. There's nodes that are nodes of atomic orbitals, of the oxygen and of the carbon; they bisect the molecule. Those don't have to do with bonding. Right? But the one on the top also has an antibonding node. Can you see where that is? Shai?



Student: Between the carbon and the oxygen.



Professor Michael McBride: Between the carbon and the oxygen, there's an antibonding node. So this is a favorable and an unfavorable combination of, in the front, oxygen, the front left is oxygen, and in the back right is some sort of C-H bond. So those molecular orbitals are made by combining 2p of oxygen with σ C-H; in fact two σ C-Hs. Okay? Now, which one is lower in energy, the unshared pair on oxygen, or σ C-H? Notice that the oxygen has something going for it, which is a big nuclear charge, but the C-H has something going for it, which is the bonding of the C-H, which lowered it from where carbon and hydrogen atomic orbitals would be. Which one is lower, the 2p of oxygen or σ C-H? Which wins? Can you tell by looking at this? Angela?



Student: σ C-H is lower because the one on the bottom has less nodes than the one on the top.



Professor Michael McBride: Right. The favorable combination, which looks mostly like the lower one, looks mostly like C-H; and the unfavorable one, the one that went up, looks mostly like O. So clearly the σ C-H is lower in energy than the 2p of oxygen. Okay, now let's look at the LUMO, up on top, in the plum pudding column, right? And here's another version of that. And you can see what it is. It's made up of a p orbital on oxygen and a p orbital on carbon. That's an atomic orbital node. It has nothing to do with the bonding. But there is an antibonding node between them. Right? So this is the π* orbital; not the occupied one but the vacant orbital, the LUMO. Right? There's poor overlap because it's π*; that's why it's unusually low. And it's also unusually low compared to C, C-C, because it's got an oxygen in it, a high nuclear charge. So poor energy match; 2p O is lower in energy than the 2p of carbon. Okay? And there -- we're going to show this on the next slide, those two -- but the top one, you see, is the favorable combination of 2p O and 2pz; the one of which this is the unfavorable combination. But it's not the HOMO, right? The HOMO is that unshared pair on oxygen, which is the dominant thing in 3dxy.



Okay, so now let's look at these. There's the C=O π bonding orbital. Now which one should that be big on? Don't look at it. Which one should the π -- the favorable combination of p on oxygen and p on carbon, which should it be big on? There's p on oxygen, p on carbon, they mix; π overlap, not too much overlap, so they don't go so much up and down. But what does this, the bonding one, what does it mostly look like?



Student: Oxygen.



Professor Michael McBride: It should look mostly like oxygen. When you look at this picture it doesn't look that way. That's because of which contour we drew. Because oxygen holds its electrons closer. So by the time you get out to a certain low -- so you get through most of the electron density of oxygen, as you go out one contour by the other of the onion, that's the oxygen; there are more electrons inside. So indeed it does have more electron density on oxygen than it does on carbon, but it doesn't show well in this particular contour. Anyhow, that's the π bonding. And this one is a different unshared pair on Oxygen. And that's mostly a p-rich, hybrid atomic orbital of oxygen, but it's got a little bit of the C-H bonding and it's a favorable combination; there's a little bit of favorable bonding between there. Okay, but the nodes that you see in this, in neither orbital are the nodes antibonding. They're always the nodes that go through atoms. They're part of atomic orbitals. These are bonding orbitals, the occupied ones, as you expect them to be.



Okay, but now let's look at that LUMO, that vacant orbital, the π* orbital. What direction should a high energy HOMO approach this, in order to mix with it? So it has to get good overlap as it comes in, and not run into the other nuclei, or the other electron density associated with other orbitals on the way in. Okay, well here are three possibilities. You could attack from within the plane, from the oxygen end; from within the plane toward the carbon; or from straight up, halfway between carbon and oxygen. How do those look as ways for an orbital to come in and overlap? Suppose it came from the right. Suppose you had a big ball, like OH-, coming in along that arrow from the right? You going to get good overlap? What would you call that situation?



Student: Orthogonal.



Professor Michael McBride: Orthogonal, right? So you're not going to get overlap -- how about if you come from the left, the arrow on the left?



Student: It's also --



Professor Michael McBride: That's also going to be orthogonal. How about if you come from the top? It's not precisely orthogonal, but it's pretty close to orthogonal. You have both positive and negative overlap. So in fact none of those are good. Now here are two other possibilities. You could attack the oxygen on the top, or you could attack the carbon on the top. Which of those do you think looks better?



Student: Carbon.



Professor Michael McBride: Carbon. Why? Angela?



Student: It's bigger.



Professor Michael McBride: Because it's bigger. Okay, so I'd say that one's smaller, so forget that. So come in from the carbon. Okay, so here's a better drawing of that orbital. And we want to come in actually from that direction, at a certain angle; not 90º, with respect to the oxygen-carbon bond. Now why come in from that angle? Why not come straight down on top?



Student: You get less overlap with the oxygen.



Professor Michael McBride: There are the nodes. Right? This is furthest from the node. So you can overlap the nodes you want to overlap with, without getting near the blue ones. Right? So you should come in at a certain angle. Now, there's a prediction. Is it true? So what we're going to do this is rotate this so the orientation corresponds to a graph, that appeared in a paper by Dunitz and Bürgi in 1983, when they were analyzing this. Now what did they do? They looked at X-ray crystal structures of many, many crystals that had both a nitrogen in it -- so the unshared pair -- and they had a carbon-oxygen double bond. And they looked to see, in many, many different structures, how those atoms are arranged, because they must be arranged in a low energy way next to one another. Okay? That's the way they would pack in the crystal.



Okay, so now how is this diagram made? Well notice that the carbon and oxygen are drawn there. So that's the carbon and oxygen, and they're made -- so all these different crystals, some have the group oriented this way, some this way, some this way, some this way. So you rotate all those crystals so that their carbons are superimposed with one another, in all the structures; the carbons are all there, and the oxygen is along this particular line, which lies in the plane of the screen. Right? And then there are things attached to it, which are R groups; an R group -- that is whatever group, hydrogen or carbon or something -- attached to the carbon. There's one in front of the plane and one behind the plane. Okay? So we rotate all these different structures around that carbon-oxygen bond. So those things are all one hiding the other one; the R in front hides the R behind.



And now we look, with that orientation, to see where the nitrogen is nearby. And those dots up on the top show in many different structures, labeled A, B, C, D, E, F, G, H, I -- no J -- K, L. Okay? In all those different structures where's the nitrogen? So in structure G, a particular compound, the nitrogen is where the blue dot is on top. The carbon is -- all the carbons are always there. That's the origin, right? Here. The oxygens are along this line. So the oxygen is along that line and at that position, that distance. And the R groups are bent down a little bit from being planar. One is here and the other one is behind. Okay? Okay, but now let's go G, H, I, K, L. G, H, I, K, L. Do you see what happens? Let's go back. K, I, H, G. These are five different crystal structures. But look again. So the nitrogen comes down, from one structure to the next. What happens to the R's, as the nitrogen comes down?



Student: They move away.



Professor Michael McBride: Right, the R's bend down. So the nitrogen -- the carbon, pardon me, in the middle, becomes pyramidal rather than planar. Back up. What happens to the oxygen as we do this? There's G, H, I, K, L. Right? Okay, so here's what happened. The nitrogen moves down, the Rs bend down, and the oxygen moves away. Why does the oxygen move away from the carbon, from a theoretical, orbital point of view? Where are you putting electrons? Where are you putting the electrons here, when you put them into this LUMO? Into an orbital that's antibonding between carbon and oxygen. So the oxygen stretches away from the carbon. Kevin?



Student: On the right side of the screen, the length units, what is that?



Professor Michael McBride: That's Ångstroms. So it's zero is the origin; that's where the carbon is. And there's one Ångstrom, 1, 2, 3. So they start three -- the nitrogen starts three Ångstroms from the carbon, and as it comes in it gets to be within about 1.6 or something. And I forget. I'm guessing that an actual bond between that nitrogen and carbon is probably the order of 1.3 or something; I'm not actually sure it is. So they haven't quite bonded. But it's following the pathway of reaction, when you look at a lot of these different structures. Okay? Now that angle, 110º, along the path that they follow, is called the Bürgi-Dunitz Angle; that's the guys who did the paper, Hans-Beat Bürgi and Jack Dunitz in Switzerland. So there is a bunch of what we might consider snapshots of a reaction actually taking place. And you can see that there's a well-defined angle from which the nitrogen approaches that LUMO, as it mixes with it.



Now, resonance is intramolecular HOMO/LUMO mixing. We've looked here at intermolecular, which is reaction. But next time we'll go on to this slide and look at the analogous thing that happens within a molecule. Okay.



[end of transcript]

Lecture 19
Oxygen and the Chemical Revolution (Beginning to 1789)
Play Video
Oxygen and the Chemical Revolution (Beginning to 1789)


This lecture begins a series describing the development of organic chemistry in chronological order, beginning with the father of modern chemistry, Lavoisier. The focus is to understand the logic of the development of modern theory, technique and nomenclature so as to use them more effectively. Chemistry begins before Lavoisier's "Chemical Revolution," with the practice of ancient technology and alchemy, and with discoveries like those of Scheele, the Swedish apothecary who discovered oxygen and prepared the first pure samples of organic acids. Lavoisier's Traité Élémentaire de Chimie launched modern chemistry with its focus on facts, ideas, and words. Lavoisier weighed gases and measured heat with a calorimeter, as well as clarifying language and chemical thinking. His key concepts were conservation of mass for the elements and oxidation, a process in which reaction with oxygen could make a "radical" or "base" into an acid.




Transcript



October 17, 2008




Professor Michael McBride: So we know something about all these powerful tools that had been developed: quantum mechanics around 1925, '26 actually; and then all these much more modern things, including scanning probe microscopy, which is really quite recent. But the most powerful tool in organic chemistry, for everyday practice, and certainly for anybody who's not a professional chemist and into quantum mechanics and stuff like that, is bonds; this amazing invention of bonds, and how people knew what things were like before any of these tools were dreamt of. Right? And, for that, we go back in time and look at these guys in the nineteenth century who invented it. And we begin with what's called -- whoops -- we begin with -- sorry, here we go again -- okay, we begin with the Chemical Revolution. Right? That's a name that historians of science have given this period, beginning with Lavoisier. So really that's where we'll start. But he didn't just spring from nowhere. There was a long tradition of some kinds of chemistry before that, and we're going to just review that very briefly. There's a background in ancient art and lore. Okay? So, for example, here's a mosaic from Montreale, in Sicily, of Noah, as making wine and then falling victim to it and having his shame hidden by his sons. Okay? And here's his flask, right?



Student: Oh no.



Professor Michael McBride: This is like 3000 years ago. The mosaic is about 1000 years; it's twelfth century, it's 800 or 900 hundred years old. Okay? Here is a Roman glass perfume vial, that's 2000 years old. So they made perfume, obviously, and extracted the stuff from flowers and whatever that would do that. Okay? There's the Chemical Research Building when it was five days old. Okay? And we already saw Francis Bacon who said: "All the philosophy of nature which is now received, is either the philosophy of the Grecians, or that of the alchemists." So the alchemists are in a sense our ancestors. He didn't have a very high opinion of them. He said: "The one is gathered out of a few vulgar observations" (that's the Greek philosophers) "and the other out of a few experiments of a furnace. The one never faileth to multiply words, and the other ever faileth to multiply gold." So even at this time, when science was beginning to get underway, modern science, alchemy was already in disrepute. But it did contribute some things. It didn't contribute much to theory, because the theory depended on Greek, well on the Greek antecedents and authority, rather than observation. But they did fiddle around. Here's a painting of an alchemist from 1663. And remember, Newton did more work in alchemy than he did in physics. Right? And he wrote an enormous amount but never published anything about it. There's a great website now, from the University of Indiana, of all Newton's alchemical works, and you can look at them.



But this was to be kept a secret. That was the idea of alchemy, that it was occult, it was hidden. Okay? And in fact there's a show that's coming up on alchemy at the Beinecke Library, in January, of their holdings in alchemy, and the title of the show is "The Book of Secrets." Right? And that's Harry Potter as well, right? [Laughter] So here's one of the things that's going to be shown. It's part of a really long scroll, which is the Visio Mystica of Arnold of Villanova, who was a thirteenth century alchemist who was into medicine and was considered the greatest medical authority of its time. This particular scroll was written in England in 1570. So a lot of it's in English. It says, "the red sea, the red lune [the red moon], the red sol [the red sun]." Right? All sorts of mystical things. And then, and you see here is the corpus, the body, which is earth; and here's the soul, which is oil; and here's the spirit or the air, the breath, which is water. So there was -- everything was a symbol for something else. Or if you look at the four corners of this, on the bottom right we have air. Right? "Eeyre," it says, "is hott and moyst." Right? Or on the top left is earth, which is cold and dry; the opposite of air. And down here is fire, which is hot and dry. And up there is water, which is cold and moist. Right? So this was supposed to be something profound and mysterious.



Another thing they're going to show is this book, On the Philosopher's Stone, which was written in the thirteenth century. The particular copy they have is from 1571. But it looks like somebody's organic chemistry text, after they've highlighted it in preparation for the exam. Every word is underlined. [Laughter] Right? And all in the margin are these fingers pointing to important things. Right? And if you look there, you'll see great words like alchemy there, you see elixir. I forget, there are a lot of key words. But fundamentally it's all nonsense, all the theory. The reason people kept it secret, really, was, I think, not to keep other people from finding it out, but to hide the fact that it was nonsense. That's just my own theory, so maybe that's wrong. Okay, but Paracelsus, in the early 1500s, was an alchemist, a traveling physician, and he developed what had been long before that, which was the Doctrine of Sympathies, and one aspect of that was in nature, antidotes for poisons are to be found near the source of the illness. Right? So, for example, you know what that is?



Student: Poison ivy.



Professor Michael McBride: Poison ivy. Right? But near poison ivy you're likely to find jewelweed, which is an antidote for poison ivy. Right? Or -- poison ivy is a New World thing, so that didn't interest the alchemists -- but this one certainly did. Willow -- Salix is the Latin name for willow -- which is found in malarial swamps. So you go into a swamp, you get malaria, but you also find the willow there. And the willow bark, extracting from the willow bark you can get Salicin, a glycoside, which is a sugar plus an aromatic thing there. Right? And if you hydrolyze that to get the sugar off, and oxidize that to make a carboxylic acid, you get this. What's that?



Students: Salicylic acid.



Professor Michael McBride: Right. Salicylic acid, from alix. Right? So it's good for fevers and so on, for your malaria. Okay, so that's a theory. This is another thing they're going to show, a Vade Mecum -- come along with me -- which is a lab manual that was kept by Caspar Harttung vom Hoff, in 1557, in Austria. So you can see he draws various -- what he's, he's reading various things, and writing extracts and notes to himself about it. It's like your pre-lab preparation. Right? Okay, and you notice who he's quoting up here? Arnold, that guy from the thirteenth century who did the thing we've showed first. Right? Arnold of Villanova. Okay, but he shows distillation apparatus. These things are called pelicans. Isn't that interesting? And here's a lamp. And here's somebody, he's filtering something through some kind of screen or grid there. Okay? So they developed tools that were of great use to chemistry, once chemistry got going. So there was a lot of -- even though the theory was nonsense -- there was a lot of practical background in preparing various elixirs and so on. So this was crucial. Now here is a lab that could easily be mistaken for an alchemical laboratory, but in fact it's an early chemical laboratory. And I'll show it to you here. It's this book, it's this picture here from 1777, from a book about Air and Fire. And it reports the discovery of a new element, in this book. And it's by Carl Wilhelm Scheele, who was in Uppsala; "Upsala," in Sweden. What do you think he discovered with light and fire -- or pardon me, with air and fire; Luft und dem Feuer?



Student: Oxygen.



Professor Michael McBride: Ah ha, we'll see. So that's his laboratory, Scheele's laboratory, or at least some artist's impression of it. And here he is. He's before the Chemical Revolution, but he's an important precedent, as you'll see, to the Chemical Revolution. He was by practice a pharmacist but he spent most of his time doing what is really chemical research. Here's a picture of a stamp, a Swedish stamp, showing Scheele. Except it's not Scheele. It turns out the costume he's wearing wouldn't have been developed until forty years after Scheele died. Okay, but he purified organic compounds, that weren't easy to purify; in particular, carboxylic acids. So he got an acid from -- that he called lactic acid; which we now know has that structure. Where did he get it?



[Students speak over one another]



Professor Michael McBride: From sour milk. Okay? So here's his paper about that, "On Milk and its Acid" from 1780. So he purified these acids as salts that he could crystallize. That was the method of purifying. So here's several reports in that paper. Item 7: "Bismuth, cobalt, antimony, tin, mercury, silver and gold were attacked by lactic acid, either by digestion" (that's just sitting there under it) "or by boiling. After standing over tin, the acid caused a black precipitate to form in a solution of gold and aqua regia." Right? So this is not mysterious writing. Right? It's talking in language that we can understand, even though it's translated from German. Iron and zinc were dissolved with a formation of flammable air. What do you suppose -- he reacted acid with zinc and he got flammable air. What do you suppose that was?



Student: Hydrogen gas.



Professor Michael McBride: Hydrogen, right? "The iron solution was brown and gave no crystallization, but the zinc solution crystallized." Why was that important?



Student: He could purify it.



Professor Michael McBride: Because he could purify it if it crystallized, and get just that salt. Right? So that's how he got pure samples of acids. Right? "With copper, our solution first took on a blue color, then green, finally dark blue, but did not crystallize"; unfortunately. And 10: "Lead dissolved after several days of digestion. The solution acquired a sweet, tart taste," [laughter] "but did not crystallize." Right? So what do you do when you don't have IR and NMR, right?



Student: Eat.



Student: Taste.



Professor Michael McBride: Sure. So he tasted all these things. Cyanide too. Okay, so he got citric acid. Where did he get that?



Student: Lemons.



Professor Michael McBride: From lemons, okay? And he got uric acid; obvious. He got tartaric acid. Tartaric acid, his discovery, turns out to one of the -- probably the single most important, maybe the second, no probably the most important compound in the nineteenth century, as you'll see. Okay? That comes from tartar, which is the deposit on the inside of wine casks, after you've fermented wine. Okay? Benzoic acid came from gum benzoin, a product of the Far East. Okay? And oxalic acid. Where do you think that came from? It came from rhubarb. Now why oxalic; why oxy? Oxy means sharp. So what does sharp have to do with it, with rhubarb?



Student: It has a sharp taste?



Professor Michael McBride: Yes, so you know oxy as a root, meaning sharp. Oxymoron doesn't mean a stupid ox. What it means is sharp, and moron means dull. So it's a sharp dullness, is an oxymoron. It's a self-contradictory word. Right? So what's sharp about rhubarb?



Student: Its taste.



Professor Michael McBride: Its taste; it tastes acidic. In fact, the word acidic comes from the Latin acidus; which comes from acre, to be sour; which comes from, the root is ac, which means sharp. So it's the same thing, acid and oxy. Okay? So look at all these things. They have what? The carboxylate group, which makes them acidic. Right? And we know why it makes it acidic; this is a review from last time, this functional group. It's not a carbonyl alcohol, it's a carboxylic acid. The high HOMO is stabilized in the acid. But it's even more stabilized when it's an anion, because you have a higher HOMO. So it changes the acidity, the ease of dissociation of H+, by a factor of 1011th. Right? Which depends on the energy difference between those two. So if you more stabilize the anion product than you stabilize the starting material, then you shift the reaction toward product, here by 1011th; a big change. Okay? But actually there's more to it than just that resonance, just that HOMO/LUMO interaction. There's a thing called "inductive effects" that we'll talk about later on. But a large part of it is due to that.



Okay, but you notice there's one exception here. Uric acid doesn't have a carboxylate group in it. So there's what it has. And notice that it has an unshared-pair on nitrogen. Like an amide, it's stabilized by a carbonyl; in fact, it's stabilized by two carbonyls, two adjacent LUMOs to stabilize it. Now if that were just stable, it wouldn't be a reason to get rid of it, to lose a proton. But the anion that you get if you lose the proton from nitrogen has a higher HOMO. So it's even more stabilized. The same trick as in carboxylic acid, but even more so, as you'll see here. The pKa of this compound is 5.8. It's pretty acidic. Right? But a normal amine, like ammonia losing a proton, has a pKa of 38. Right? So this is thirty-two powers of ten helped out, because it has such a high HOMO on the nitrogen, and two carbonyls to stabilize it. Okay, so uric acid is indeed an acid, like carboxylic acids.



Okay, now Scheele not only did these organic acids, he also discovered, or co-discovered, seven elements. They're listed here according to what row of the periodic table they're in. Notice down at the bottom here you have tungsten. Right? Tungsten comes from Swedish. He was Swedish. It's "tung" "sten," heavy stone. Being way down, it's got lots of protons and neutrons and is very, very dense. Right? So it's heavy -- the stones that it comes from are very heavy. Right? But by contrast, these up here are gases. And in fact that's what got the 19th Century chemistry going. That's what launched the Chemical Revolution, was the ability to work with gases. Because to be a gas, something has to be a small molecule and therefore simple, or at least relatively simple. So you had to start with simple things before you could get the complex ones, like salicylic acid, in terms of understanding.



Now Scheele, in 1771, had heated silver carbonate, and he found that he got CO2 out of them; he didn't know it was CO2, but the gas came out. Okay? And if you heated that still more, greater than 340 Celsius, then you get silver, and oxygen comes out from silver oxide; this gas, this feuerluft, fire air. That's what the book is about. Okay? So he wrote the book. But the book starts, as I'll show you here -- sorry, there we go -- the book starts with an introduction, a "vorbericht," which is translated, it says, from Swedish. And let's see where it says here. And it's by Torbern Bergman, written in 1777. He had this book written for two years, waiting for this preface to come. Bergman was a busy guy. Right? And during the time that this book was sitting, ready, the manuscript was sitting ready to be printed, Priestley, in England, discovered oxygen. So this book came out after Priestley. But there's no doubt that Scheele had discovered it earlier. His lab books from 1771 show it. And here, in 1774, is his draft of a letter that he wrote to France. And it begins -- or it actually begins with a couple of words on the previous sheet -- but it says: "…since I have no large burning glass, I beg you to try with yours..." Because he had to do this by heating things in an oven, at a really high temperature, which was hard to do. But if you could do it with a focused light of the sun to heat it, then it would be much more practical. And, in France, they had such a big magnifying glass, that would allow to do it. But that letter, although it was sent, was never answered. And you know who it was sent to, presumably. Lavoisier, the founder of the Chemical Revolution, and another discoverer of oxygen. Okay? Okay, so now we're going to talk about Lavoisier, who was -- wasn't a perfect person, but he was really very, very good. Okay, now the Chemical Revolution. You can say it started in 1789, the Chemical Revolution. And that's not the only revolution that started in France in 1789. Right? Do you know what this is? What?



Student: Tennis Court Oath.



Professor Michael McBride: It's the Tennis Court Oath, when the legislators, so to say, gathered to say that they wouldn't disband until the king granted them certain things; and you know what that led to, in 1789. The only guy that didn't agree was this guy here. He's the only one that didn't sign it. But at any rate, it was radical. Now there's an Indo-European word, that's the root of many words, called Werad; and it gives words in all sorts of languages. Like it means root. Okay? And Wurzel, in German, means root; and wort, like St. John's wort, is a root. Licorice; glukos rhiza, Greek, the rhiza is root; sweet root it means. Okay? Race; razza in Italian is the root of your being. Right? Rutabaga. Radix in Latin; and you know lots of words come from radix, like radish is a root. Or eradicate, what does that mean?



[Students speak over one another]



Professor Michael McBride: It means to pull it out by the roots. Okay? Or radical; something that's radical is something that goes right to the root, back to the very origin of something. And that word, used in that way, if you look in the Oxford English Dictionary, it was coined in mathematics in the 16th Century. The root of a number is its origin. Right? If you take the square root of a number, and multiply it by itself, you get the number. So it's the root of the number; the radical. Right? Or in politics, it was used in 18th Century in England, and in chemistry, in 18th Century in France, the idea of radical as the root of things, began to be used; which we'll see.



Okay, so 1787, radical was introduced as a political term, according to the Oxford English Dictionary, by J. Jebb, whoever he was; presumably a politician. Or in 1787, there was this radical document, "We the People." Right? But that same year, 1787, radical was introduced as a chemical term, by Louis Bernard Guyton de Morveau. And it was in the context of developing nomenclature for chemistry. So he, together with Berthollet and Fourcroy, developed a new method for nomenclature in chemistry. And here's a book. This is not the original French, but it's the first English translation, which you see comes from the Yale University Library, back when. It's from 1788. Okay? So A Method of Chemical Nomenclature, by Guyton de Morveau, Lavoisier, Berthollet and Fourcroy. So the fourth author of this new method of chemical nomenclature is Lavoisier. So there's Lavoisier with his wife. This is part of an enormous picture that's in the Metropolitan. It was commissioned by Lavoisier and his wife, who hired Jacques-Louis David to paint it. And they paid 7000 pounds, to the artist to paint it; which is the equivalent of $300,000 today. They were quite well-to-do. They had an income of the order of a million dollars a year, or the equivalent; it depends on how you translate numbers, of course. It's hard.



So here he is at the age of 45, in 1789, and he's working on drafting -- so these guys, when they had their portraits painted, always put something important in it. So what did he choose to have? He chose to be working on a manuscript, and the manuscript he's working on is the manuscript of this book, from 1789. It's called Traité Élémentaire de Chimie; the Elementary Treatise on Chemistry. And the other stuff he put in the picture is the equipment he used. So here you see various equipment from one of the plates at the end of the book. And you can see these items in the picture. There's that bell jar. There's that device. There's that big sixteen-pint flask, with a brass fitting on it. There's that valve that you attach to the bottom of the flask. And over here is a portfolio, and the portfolio says, down in the corner, "Paulze Lavoisier sculpsit." That means this was drawn by Paulze Lavoisier, who was his wife. She was his assistant in the laboratory, kept all the notebooks; read English for him, because he couldn't read English. So if he had to do anything with Priestley, she would read it to him. But she drew all these things. She studied with David, drawing, in order to be able to do this. Okay, she painted this portrait of their family friend. Who's that?



Student: Benjamin Franklin.



Professor Michael McBride: Yes, Benjamin Franklin. I showed you that picture earlier, said we'd refer to it again. So this is that particular plate in the book, and it relates to weighing a gas. It'll turn out that the most important thing for Lavoisier, and for the whole 19th Century -- all this development that led to bonds and their arrangement -- weighing was the key thing. But how do you weigh a gas? So you need gases so they're simple enough to deal with, and easy to purify. Right? But you need to weigh them. So how can you weigh a gas? Well you can collect a gas, as shown in this picture, by generating it in this retort G, and it comes and it bubbles up, displacing water or mercury, most often mercury, from a bell jar. You've done this kind of thing?



Student: Yes.



Professor Michael McBride: Some of you. But you can see how it would work; as it bubbles up the mercury comes down. Okay. So now you have the gas. Now we'll shift attention down to the bottom right here, and see how this works. So we have this bell jar on the bottom, which is fitted with valves on the top, is filled with mercury, and it's in a pool of mercury, and then this big sixteen-pint flask is evacuated. You use one of these pumps. Remember, 100 years before this, Hooke -- or 130 years before this -- Hooke made a great vacuum pump for Boyle. Boyle is the only person on the front of the building older than Lavoisier. Right? And that was dealing with gases and Boyle's Law, how pressure and volume relate to one another. Okay, so anyhow, he could evacuate that with a pump, then seal it off, turn the valves off. And he's got mercury in that thing. And now he puts a tip underneath it and generates gas. And it bubbles up and fills this container with a gas. And it's sitting in a mercury pool so that it's not communicating with the atmosphere, other than through the pressure, through the mercury pool.



Okay, so now he opens the valves. So the vacuum starts sucking up the mercury; that is, pulling the air in, up to a certain point. Right? Then the mercury stops rising. Okay? And now, at this point you know that the pressure of that gas is atmospheric pressure, less whatever the height of the mercury column is. Right? That's how a barometer works. So he knows how much gas, what volume of gas he has in A. He filled it with water first and weighed it to see what its volume was. Now he knows the volume, he knows the pressure. So he knows how much gas there would be, at atmospheric pressure. Right? And now, of course, he just turns off these things, unscrews one, the thing on top, and weighs it, and sees how much heavier it is than it was when it was evacuated. And that's how much the air weighs. And he knows how much volume, how much pressure. So he knows how much whatever gas he collected weighed. So he could weigh a gas. Pretty clever, huh? Okay, here he is working with one of these bell jars. Now these bell jars were filled with mercury. I don't know if you've -- here, would you help me out Wilson? Lift this up and show it to the class. But don't lift it high, hold it above the thing. Did it surprise you?



Student: Yes.



Professor Michael McBride: He said it surprised him. And you can come up afterwards, if you want to, and be surprised yourself, by lifting this up. But keep it over this, because people are so panicked about mercury nowadays. I heard on the way over to class, while I was bringing this, I heard there's a new law in the European Union that it's going to be illegal to transport mercury over international borders. Go figure. Anyhow, there's Lavoisier doing an experiment with this big thing of mercury. Right? He must've been a stout person. Okay? This is him in his library, with Madame Lavoisier taking his dictation, as he does his experiments. Okay, so here's the Traité Élémentaire de Chimie, and I'll show you the -- this actually is a facsimile, not the real thing. But here's the title page of the first volume. Okay. So you can look at that if you want to. And at the end, I'll just note here, at the end of the second volume, are all these pages, which are devices -- like here's this is the one we just looked at. Okay? So if you want to look at them, feel free. Don't handle the other one though; it's real.



Okay, so Elementary Treatise of Chemistry, Presented in a New Order -- this is the Revolution -- According to Modern Discoveries, With Figures; as I just showed you. And 1789 is the date, the same as the French Revolution. Okay, you notice he's a member of all different academies, including Philadelphia. Why in the world would he have been a member of the Scientific Academy of Philadelphia? He never went there.



Student: Ben Franklin was --



Professor Michael McBride: Right, his pal. Okay, 1789. So it has the most wonderful Introduction, called "Discours Préliminaire." And he says: "My only object, when I began this work" -- or, "I had no other object when I began the following work, than to extend and explain more fully the memoir, which I read at the public meeting of the Academy of Science in the month of April, 1787" (remember when radical was introduced and so on) "on the necessity of reforming and completing the nomenclature of Chemistry." So that's all he was trying to do was get a proper nomenclature that would be useful, in contrast to all this alchemical nonsense. Okay? "While engaged in this employment, I perceived, better than I had ever done before, the justice of the following maxims of the Abbé de Condillac and his System of Logic and some other works." So this is what Condillac said.



"We think, only through the medium of words. Languages are true analytical methods. Algebra, which is adapted to its purpose in every species of expression, in the most simple, most exact, and best manner possible, is at the same time a language and an analytical method. The art of reasoning is nothing more than language, well arranged."



So Lavoisier goes on to say:



"Thus, while I thought myself employed only in forming a Nomenclature, and while I proposed to myself nothing more than to improve the chemical language, my work transformed itself by degrees, without my being able to prevent it, into a treatise upon the Elements of Chemistry."



So in the process of reforming the language, he reformed the whole understanding of the science.



"The impossibility of separating nomenclature of a science from the science itself, is owing to this, that every branch of physical science must consist of three things; the series of facts which are the object of the science, the ideas which represent these facts, and the words by which the ideas are expressed. Like three impressions of the same seal, the word ought to produce the idea, and the idea to be a picture of the fact."



So all these things have to be harmonious. Right? Three impressions of the same seal.



"And, as ideas are preserved and communicated by means of words, it necessarily follows that we cannot improve the language of any science without at the same time improving the science itself; neither can we, on the other hand, improve a science, without improving the language or nomenclature which belongs to it. However certain the facts of any science may be, however just the ideas we may have formed of these facts, we can only communicate false impressions to others, while we want words by which these may be properly expressed."



Right? So clarity, as opposed to obscurity, was his goal; as opposed to Newton or the alchemists. Right? Facts, ideas and words, and they all have to tie into one another as impressions of the same seal. Okay? So he presented things, as he had advertised, in a new order; very different from any book on chemistry that had been written before. First was doctrine, -- that is, the theory -- the first part of the book, which is a two-volume book. So almost all of the first -- no, about two-thirds, I think, of the first volume are doctrine. And then nomenclature; that's what he had set out to do. And finally operations, how you can actually repeat this stuff for yourself, what devices you need. Of course, he was very wealthy and could employ people to manufacture all the equipment he needed; not everybody could do that. But he showed exactly how it was done and gave great -- it's easy to understand exactly what he did. Now, one of the first things he turned his attention to was elements. He says:



"…if by the name elements we mean to designate the simple, indivisible molecules" (molecule just means little thing, right?) "that make up substances, it is probable we do not know what they are." (They're just too small. Right?) "But if, on the contrary, we associate with the name of elements, or the principles of substances, the idea of the furthest stage to which analysis can reach, all substances that we have so far found no means to decompose are elements for us… they behave with respect to us like simple substances."



So it's an operational, not a philosophical, definition of element. If you can't break it apart, consider it an element until you can break it apart. We have elements here, the chemical elements. Are they elements, according to Lavoisier's definition? Like here we see cerium, praseodymium, neodymium, promethium, samarium, europium, gadolinium. Are they elements, according to Lavoisier?



Students: No.



Professor Michael McBride: Why not?



Students: You can break them.



Professor Michael McBride: You can break them apart, into nuclei and electrons. The nuclei you can break apart into protons and neutrons, and you can break these things apart into quarks and so on. But, for Lavoisier, or for chemists, those are elements, because you don't break them apart. Okay? So here's a table of simple substances, in the first English translation of the Traité Élémentaire de Chimie. So here's a table of the elements: "Simple substances belonging to all the kingdoms of nature, which may be considered as the elements of bodies." So what are the first two elements, things that you can't break apart?



Student: Light.



Professor Michael McBride: Light and heat are the first two elements -- which we don't see in Mendeleev's table, right? Because they're fundamentally different from the other elements, because they don't have any weight. You can weigh a gas but you can't weigh light, and you can't weigh heat. Okay? So he gives new names to these things -- light and caloric -- and what the old name was. Light used to be called light too, but in his new system he's going to keep the old name; or "caloric", he's going to use for what used to be called heat; or "the principle or element of heat"; or "fire"; or "igneous fluid"; or "the matter of fire or heat". Those were terms people had used before, but he's going to call them "caloric". Okay?



Now, if you have caloric, you must be able to measure it. If you can't weigh it, what you can do; you can use a "calorimeter" to measure it. And this is the calorimeter manufactured and used by Lavoisier; and also Laplace, a younger man who was his colleague, who became a great mathematician, as you probably know. So here's the thing. It's big. That's a three-foot rule. So this thing stood this high off the table, right? Okay? Now here's what it is. There's a lamp that's going to make fire. There's oil in the well of the lamp. You're going to measure how much heat you get out of burning that oil. So you put it inside this bucket, and you put the bucket in this mesh cage and put the lid on. Then you put that cage, and its lid, up into this can. Okay? And now you light the flame, in there. You want to measure how much heat it gives. How do you measure the heat? What you do is surround it by melting ice. So the heat will melt the ice. Now there's going to be a problem. Obviously the more heat, the more ice you melt. But that's not the only place heat's coming from. Where else will it come from, to melt the ice?



Students: Outside.



Professor Michael McBride: From outside. So this is where the thing is clever. So there's another can, outside that can, and you fill it with ice, which is an insulator, for the inside. So no heat comes from the -- any heat that comes from the outside, melts the outside ice, not the inside ice. Only the flame will melt the inside ice. Okay, and notice that the lids also are covered with ice too. So it's completely surrounded by ice, and then by another layer of ice. So your flame burns, burns, burns, burns. And you fill -- water comes up as they melt in both of them. And then you put that thing underneath and turn the tap, and see how much water was melted; only by the flame, right? And that measures how much heat. Pretty clever, huh?



Okay, so that's a fact, is measuring how much heat there is. But analysis in general is it. This is from the Oxford English Dictionary, which Yale has a subscription to. So you can look up words to your heart's content. It's a lot of fun. This year is the 80th anniversary of the Oxford English Dictionary. There was a symposium, down the hill, sponsored by the library, including the guy that wrote that The Professor and the Madman. Has anyone read that book about the Oxford English Dictionary? It's a wonderful short book, and the madman was a Yale graduate. Really, it's an interesting story. And the other guy was one who read the Oxford English Dictionary, cover to cover, 20 volumes, within one year. He wrote a book about that, during the last year. It was a fun thing. But anyhow, this is -- it's fun to look up things in the Oxford English Dictionary. And here's analysis. So you see it comes from ana, and I think it's lisein; but I can get some help on that.



Student: luein.



Professor Michael McBride: luein, okay. Anyhow, but it means "to loose," according to the Oxford English Dictionary. So it's to loose back; to take things apart is the sense of it. And generally it's the resolution or breaking up of anything complex into its various simple elements. So you can analyze a passage in literature. Okay? It's the opposite of synthesis. Okay? "The exact determination of the elements or components of anything complex; specifically in chemistry, the resolution of a chemical compound into its proximate, or ultimate, elements." Now, proximate and ultimate, what does that mean? You can see down in the historical uses of it what proximate and ultimate mean. 1791, this same time -- 1789, remember, was the Traité Élémentaire -- said, "the quantity of charcoal, which something yields by analysis." So you find out how much charcoal is in it. That's the word -- we now say carbon. So that's elemental analysis. That's ultimate analysis; take things all apart to the chemical elements, see how much of each one. But there's also this thing called proximate analysis, which you can see from this quote in 1831. "Sugar, starch and gum are proximate principles, and these we obtained by proximate analysis." So you can take some foodstuff and see what percentage of protein, what percentage of sugar, or what percentage of this, that and the other thing. So what you read on a candy bar, or something like that, that's proximate analysis, not elemental, not ultimate analysis. So both kinds are important in Lavoisier's work, as we'll see.



Okay, so we looked at light and caloric. Now let's look at a few of these elements, the ultimate elements here. He had the fact, the theory and the word for these things. So how about Azote? Why is that word -- that's the French name for -- the French still call nitrogen azote, and in this English translation it was called that in 1790 or '91. Okay, where does that word come from? What does the prefix a mean?



Student: Without.



Professor Michael McBride: Without. And how about zo? It's great we have somebody taking Greek.



Student: Life.



Professor Michael McBride: Without life. In what sense is that an appropriate name, a meaningful name for nitrogen? Alex?



Student: I think they performed the tests on nitrogen first and it was --



Professor Michael McBride: What kind of tests?



Student: They would suffocate like a bird in nitrogen.



Professor Michael McBride: Yes, if you put a mouse in an atmosphere of nitrogen, it's azote. Right? So that's what the name meant. It used to be called phlogisticated air, or gas, or mephitis, or the base of mephitis. Right? Azote is the name that Lavoisier decided to use for it. Or hydro-gen. How about it, help us out?



Student: Hdor would've been water.



Professor Michael McBride: Pardon me?



Student: Water.



Professor Michael McBride: What about water? Hydro is water. What's gen?



Student: To leave water.



Professor Michael McBride: It makes water. Right? So if you burn hydro-gen, you generate water. Right? So hydrogen. Okay. How about oxy-gen? What does it generate?



Students: Acid.



Professor Michael McBride: Acid, sourness. So oxygen is the element that generates sourness, that generates acid. And that is the key element in Lavoisier's theory, the oxygen theory of combustion. Okay, so oxygen, plus a base, or radical -- right? So these two terms meant the same thing to Lavoisier, the fundamental radical, the root of some substance, right? And you react it with oxygen and it makes the stuff into an acid. Can you think of an example of an element that you react with oxygen and it becomes an acid? Well let's just look in his table. Sulfur; you burn it and it becomes sulfuric acid, or sulfurous acid. Right? Phosphorous generates phosphoric acid. Carbon generates carbonic acid. Muriatic radical -- which we don't know; they haven't discovered the muriatic radical yet, the base of that acid -- but if you burn it and combine it with oxygen, you get muriatic acid. Does anybody know what muriatic acid is?



Student: Hydrochloric acid.



Professor Michael McBride: Hydrochloric acid. How much oxygen is in it?



Student: None.



Professor Michael McBride: None, right? But that was the theory, that you take a base, you react it with oxygen, you get an acid. So there must have been a muriatic radical. Okay? Unfortunately that part of it was wrong. The same for fluoric radical. Okay? But then there were also compound radicals, radicals that were only proximate, not ultimate; radicals that had several other elements in them. Right? And here were some of those, a list of those radicals, with the names that Lavoisier decided to use for them. And many of them, all the ones indicated by an arrow, are ones that Scheele had already discovered; like tartaric, citric, oxalic, benzoic, lactic. Lithic acid was another one that I didn't mention before, which comes from stones; see, it's from urinary calculus. Okay, so those were compound radicals. And that's the end of today's lecture.



[end of transcript]

Lecture 20
Rise of the Atomic Theory (1790-1805)
Play Video
Rise of the Atomic Theory (1790-1805)


This lecture traces the development of elemental analysis as a technique for the determination of the composition of organic compounds beginning with Lavoisier's early combustion and fermentation experiments, which showed a new, if naïve, attitude toward handling experimental data. Dalton's atomic theory was consistent with the empirical laws of definite, equivalent, and multiple proportions. The basis of our current notation and of precise analysis was established by Berzelius, but confusion about atomic weight multiples, which could have been clarified early by the law of Avogadro and Gay-Lussac, would persist for more than half a century.




Transcript



October 22, 2008



Professor Michael McBride: Okay, as you may remember, way back before the exam, we'd started looking at how things really happened, how people were able to figure out about bonds, and how atoms were arranged, and molecules reacted, before there were the powerful techniques that we -- that developed mostly in the last twenty-five or thirty years, as far as their practical application in organic chemistry; but how they found these out before that time. And it started with what we call the "Chemical Revolution", which was launched by Lavoisier, with this book, in 1789, the same year as the French Revolution, The Elementary Treatise of Chemistry. Which, you remember, he started work on just because he was interested in improving nomenclature, but found that there was such a tight coupling between nomenclature and the science -- that is, the facts and the theory -- that he couldn't improve any one of them without improving the others. So we saw last time that he developed a -- that in the process of developing a new nomenclature, he also developed instruments, so that he could weigh gases, because weighing turns out to be such an important feature in all of nineteenth century chemistry, no doubt the most important single technique. But he also could measure heat, which didn't have any weight. For that he developed a calorimeter together with Laplace, and we saw how that worked by melting ice. So now we're going to look at some of the other developments from this book, in particular the idea of oxidation.



There's no doubt that the most important element for Lavoisier was oxygen, because of its key role in his chemistry, so that all other elements were called bases, or radicals. Right? And they could react with oxygen to become new things. There were several degrees of oxidation, as shown in this table from the English translation of his book. So first you get an oxide, oxides of all the elements that are listed in the first table. For example, hydrogen when oxidized would give water; that's the second row. But the first row is caloric. This was sort of a stretch, at least it seems so from our perspective, that caloric, heat, when it gets oxidized, yields oxygen gas. But then there were further degrees of oxidation if you reacted with more oxygen. So the second degree of oxidation, according to his theory, was to give an acid, that the suffix was to be systematically o-u-s, so -ous. And those names we use now, like cuprous, ferrous, come directly from this work of Lavoisier. So that's the lower oxidation state of an acid. Right? So you have, for example, nitrous acid, carbonous acid, sulfurous acid, phosphorous acid, he said. And then you have a still higher degree of oxidation, the third degree, which is the -ic acid. So you have nitric acid, carbonic acid, sulfuric acid. And notice that many of these were not known; and many of them still are not known because the theory wasn't exactly right. For example, carbonous acid was not known at that time; you see in the second entry in the second row, second column. Okay? And then there was still a fourth degree of oxidation, which he called the oxygenated ic acid. He didn't use the peroxy acid, as we would often use nowadays. But he used, for example, oxygenated nitric acid, which was unknown. Oxygenated carbonic acid was unknown. Same for sulfuric or phosphoric. And oxygenated muriatic acid was -- used to be called dephlogisticated marine acid. But he has a more systematic name, showing its degree of oxidation. So these -ous and -ic names, that we still use, come directly from Lavoisier.



And probably the next most important tool he developed was one for doing elemental analysis. Remember, you could do proximate analysis, what other compounds are present in the substance you're studying. Or you could do ultimate analysis and find out how much of the elements there was in each. And this was how to do elemental analysis of an organic oil. So it has carbon, hydrogen, oxygen in it. Okay? So this big upside-down barrel, inside another barrel with water in it, is how he got air for the burning, because the upside-down barrel would be lowered into the water, which would create pressure in the oxygen that's trapped inside. It would then flow out through the pipe you see coming from the tank, through a dryer and into this big, vertical cylinder where the combustion takes place, fed by an oil supply which is siphoned in as needed. So there's a lamp inside, burning the oil with the air, and the vapors generated by the combustion come up and out the tube to the right. So there's the lamp. The gases come out and water gets condensed in the spiral tube, in here, and drips down and gets collected in the jar that you see. And the part that doesn't condense and get collected into the jar, gets absorbed in that horizontal tube that has a calcium chloride drying agent in it. So at the end of the experiment you weigh how much water there is in the jar and how much the calcium chloride has increased in weight, and that tells you how much water there was.



And what does that tell you about the compound? How much what? What element is being measured in the oil, by weighing the water? It's measuring the hydrogen. So he had to know what percent, by weight, water is, of hydrogen. Then if he measured the water that came from combustion, he measured how much hydrogen there was in the oil sample that was being burned. And then the gases continued to go through limewater, calcium hydroxide. So that would absorb CO2, as calcium carbonate. So those would increase in mass. And he had two of them, so whatever got through the first one would hopefully be collected in the next one. And then there's another drying tube, on further from that, because these solutions, as you pass gas over them, they lose a little bit of their water. Right? So they'd be losing weight, as well as gaining weight from the CO2 that's being absorbed. Right? But that water that's lost is reabsorbed in that final tube. So you weigh it at the end, and now you have -- you've corrected for whatever loss of water there was from the bulbs, and now you know what weight of CO2. And if know what fraction of CO2 is carbon -- and there's a homework problem for Friday that you can look at. You can work in groups on that, that shows you how he knew how much -- what weight percentage of CO2 was carbon. So having then measured how much CO2, he knew how much carbon there was in the oil. So he measured the hydrogen, he measured the carbon. So that's fine if it's a hydrocarbon oil. How about if there was oxygen in the oil? Notice, he's adding all kinds of oxygen from the air. So how's he going to know how much oxygen there was, if it was an alcohol or something that he was burning?



Student: [inaudible]



Professor Michael McBride: Sam? Can't hear you.



Student: The leftovers.



Professor Michael McBride: It's the leftovers. You see how much oil got burned, how much carbon, how much hydrogen, and you assume there's nothing else there, except oxygen. So you get oxygen by difference. And that's the way oxygen is almost always measured, is by difference. Right? So you assume -- and you have to be sure there are no other elements there, if you want that to be true, obviously. Okay, so this was a big operation and it took, you know, somebody to control the air supply to make sure it was constant. Notice, incidentally, there's a T here. So there was another one of these things over here, another one of these double barrels. Right? Why?



Student: [inaudible]



Professor Michael McBride: Ah, so when you run out of air in one, you switch the valve and take it from the other one; then you can go back to the first one. So you have an infinite supply of air, plenty to burn the oil. Right? But you need someone to control that, and you need someone else to make sure the lamp stays lit, that it gets the proper amount of oil siphoned in, and the airflow is right. You need someone else -- in the CO2 collectors, you just blow gas over the surface. Right? So it's got a little key on the top and a hook that goes down in. Somebody has to be standing there, doing this, to keep it stirred. Right? So that's stirred. So it takes at least three or four people to run this thing. Right? And it cost a lot of money to build. So no one else built one. In fact, it's not absolutely certain that Lavoisier ever used this, although it exists, right? Because he was busy with other things, having to do with the French Revolution, by the time this came along. But this was big science. Right? He was very wealthy, so he was able to do this kind of thing. But he showed, in this diagram, exactly how it could be done. So it was no secret, not like what the alchemists did. Was there a question here? No. Yes Lucas?



Student: Did they ever measure like the air supply they gave it?



Professor Michael McBride: Did they measure --



Student: Did they measure the amount of air that they gave into it? Did they know like the amount of the air?



Professor Michael McBride: I don't think so, because there wasn't any purpose; you didn't care how much air you put in. Because most of the air just flowed right on through, it was never collected. Right? It just carried the other gases along, where the water and the CO2 would be collected.



Student: Well I guess if they knew the content of the air, like they could just measure how much oxygen-



Professor Michael McBride: To get the oxygen? Conceivably; but that was never done. It would've been very difficult to collect all that air and measure it accurately. Okay now what if you have a substance that won't burn cleanly, so you can't measure these gases? Notice, incidentally, the role of gases in understanding things, because they provide easy separation. Right? And then since he was able to weigh gas, once, then he knows the gas density. So now all you have to do, from then on, in principle, is measure the volume of gas, and you know how much stuff there was. So gases played a really key role through all this.



But suppose you had a substance that wouldn't burn, that you couldn't convert to steam and CO2. What do you do then? For example, grape sugar. So here's chapter thirteen where he talks about le suc des raisins, the grape sugar; which will char, it won't burn cleanly. So what do you do? Well, as he says up at the top there, "tout le monde sait comment se fait le vin"; everyone knows how to make wine, cider and mead. Right? So can you see what he's going to do? So here's the device for that one, which is in many ways sort of similar, but you don't need an air -- you don't put the air in. What you have is a big flask at the beginning which has sugar, yeast and water. Right? So the yeast ferments the sugar and changes it into CO2. So you have -- it foams up; if you've even seen something fermenting. So you have to have something that catches the foam so it won't get into your device. And then you have something that will absorb water. Now the water is not just -- it's not like the previous one, where the water is all formed by the thing, because you got a whole bunch of water there at the beginning. Right? So that's not what you're trying to measure. What you're trying to measure is how much carbon, how much CO2 there is in the grape sugar. So you have these sodium hydroxide solutions, the base that will absorb the CO2, as a carbonate. So you weigh those before and after. And the excess gas, if there's another gas being evolved, will collect over the mercury in the far tube here. So he could measure the CO -- the carbon, in sugar, by converting it completely to CO2 through fermentation.



In speaking of fermentation here, he says it furnishes a means of analyzing sugar. So oxidation failed with air. Even with oxygen or sulfuric acid or mercuric oxide, you couldn't get clean combustion of sugar; you got charring instead. But as he says here -- this is the most interesting thing about his treatment of the data -- he said, "I could consider the materials subjected to fermentation, and the products of fermentation, as an algebraic equation, having to do with weights." Can you see what his assumption's going to be here?



Student: [inaudible]



Professor Michael McBride: Sam?



Student: The same weights on both --



Professor Michael McBride: The same weights on both side; conservation of mass is his assumption. Okay? So he can make an equation that the weight on the right and the weight on the left will be the same. Okay, so an algebraic equation. "And by in turn, supposing each of the elements of the equation to be unknown, I can derive a value, and thus correct experiment by calculation, and calculation by experiment." So what does he mean by that? So if you have this equation, a certain amount of sugar and a certain amount of oxygen gives a certain amount of CO2, for example. Right? So the weights on the two sides should be the same. He doesn't know one of them. Right? But if he knows all the others, and it's an algebraic equation, he can then solve for that one. So he says here: "I can derive a value and thus correct experiment by calculation." Have you ever heard of that being done? Lexie?



Student: [inaudible]



Professor Michael McBride: Would somebody -- would you do that in lab, correct experiment? Do you suppose there might be any nefarious people who would correct their experiments by calculation after finishing a lab?



Student: [laughs]



Professor Michael McBride: No, I can't imagine that would happen, right? Okay, but Lavoisier was upfront. Right? He was even, one might say, bragging about it, that he has this algebraic equation, so he can correct his experiments. Right? "I've often profited from this way of correcting the preliminary results of my experiments." No one told him that that's not the way science works. Right? That the experiments are the things you have to do. You have to adjust the theory to deal with the experiments. Okay? Remember, he was the treasurer for the firm that collected taxes for the French Crown. Right? So he was accustomed to bookkeeping, and everything had to balance, right down to the sou. Right? So the same thing he applied to chemistry. So Table One here is the materials of fermentation. He put in a certain amount of water, 400 pounds; a certain amount of sugar, 100 pounds; yeast. Now he knew that the yeast was wet. So there was a certain amount of water. He put ten pounds of yeast, but he knew by analysis, by proximate analysis -- right? -- not what the elements are but what the substances are there -- that in ten pounds of yeast, wet yeast, you'd have seven pounds, three ounces, six gros and forty-four grains. It turns that a gros is a 72nd of a grain, and eight gros make an ounce. So it's a gros is -- which is 28.35 grams. So, if you want, you can work through all this.



Do you think he had analyzed things that accurately, that he knew right down to the gros how much there was in there? No. There's no way he could've known that. But things had to balance, right? So he made an approximate thing, and then he carried it out to eighteen decimal points, the way you'd do on a calculator. Right? Okay, so he didn't know about significant figures yet either. Okay? So then in Table Two he figures in that much water, sugar, yeast, how much is there of -- in each one, how much is in water, how much is hydrogen, and how much is oxygen. So he knows that. So now he knows down to hundredths of gros how much of each of those elements there are in the 400 pounds of water he's using. Incidentally, he never used 400 pounds of water. He just rounded his numbers up, or multiplied his numbers, scaled them up to be that big. He used large quantities, but not that large. Okay.



So if you had 100 pounds of sugar, you'd have a certain amount of hydrogen, oxygen and charcoal - that is, carbon. And if you had a certain amount of the dry yeast you'd have that. And then he makes still a third table where he rearranges it, so he pulls the elements all together. So in the starting material, how much hydrogen, oxygen and carbon? Okay? So he's got 411 pounds, 12 ounces, 6 gros, 1.36 grains of oxygen -- grains and gros, got it; oh gros and grains, okay, right -- okay, that much oxygen and some amount of hydrogen and charcoal, and a certain amount of nitrogen that's in the yeast, because he knew that was there; which is just a small weight altogether, because there's not much yeast. And in all, 510 pounds exactly. Right? So his balance sheet was just like what a banker would want. Okay, and now he does -- he looks at the products, and he has a certain amount of CO2 that he collected; that's carbonic acid. And he knows how much oxygen and charcoal; there is that. A certain amount of water. A certain amount of alcohol, if dry. So he had to collect the alcohol and say how much alcohol and how much water was in it. How much acetic acid. How much residual sugar that had fermented. And how much dry yeast there was at the end. And he added them all together, and by God, there was 510 pounds, zero ounces, zero gros, and zero grains. Right? QED. Okay? And then he recapitulated it, according to how much oxygen, how much charcoal and how much hydrogen. And once again it balances out absolutely precisely. Okay, so he recapitulates it that way. And this is how he would do some of these things.



Okay? So this is a description of an experiment, on page eighty-eight to ninety-two of the Traité Élémentaire. Okay, so he's got a fire burning here, and a glass tube that runs through the middle of the charcoal that's burning. Right? And he's got a flask cemented to it, on the right, which is over a burner. And in the tube he puts twenty-eight grains of carbon -- not pieces but a weight of twenty-two grains. We saw what a grain was last time. That turns out to be 1.38 grams of carbon he's got inside the tube. Okay, and now do you think the carbon will evaporate, incidentally, if you put it in flame? Do you know how? How do you know it won't evaporate?



Student: It shouldn't.



Professor Michael McBride: Where do you get carbon? What did he call it?



Student: Charcoal.



Professor Michael McBride: He called it charcoal. You get charcoal by the flame, right? So with burning without enough oxygen. So obviously it's not going to evaporate, it's just going to sit there. Okay, so in flask A there he puts water. So there's water in it. And then he lights the fire underneath and the water begins to boil, and distills through the tube and collects over here in H. Right? There it comes. Okay? But did you notice what happened in the process? Want me to back up and do it again? What happened?



Student: The charcoal went away.



Professor Michael McBride: The charcoal went away. What happened? Well let's see what he found. So, the twenty-eight grains of carbon was gone. Right? And in the end he got the water that he had put in at the start, but not all the water. It was 85.7 grains short. So all the carbon and some of the water has gone. Where did it go? Well it went out the tube. So out the tube, he said, came 144 cubic inches -- and because he knows the density, he knows how much weight that is -- 100 grains, exactly, of carbonic gas, CO2; we would call it CO2 -- and 380 cubic inches of flammable gas. So one of them would be collected by limewater, the carbonic gas, and the other one would go right on through. What do you suppose flammable gas is?



Students: Hydrogen.



Professor Michael McBride: Hydrogen. So the oxygen went to the carbon, and the hydrogen became elemental hydrogen, in this process. Okay? Now, look at this, this is what's neat. So he got 100 grams -- 100 grains, pardon me, of carbonic gas, and 13.7 grains of hydro-gen, the stuff that gives water when you burn it, according to him. Right? Which is right, which is correct. And what do you notice, quantitatively?



Student: They match.



Professor Michael McBride: They exactly match, right down to the 10ths of grains. Okay? Now, if you recalculate, by modern theory, how much gas you should get from twenty-eight grains of carbon doing this, you should get 157 cubic inches and 313 cubic inches, instead of 144 and 380; you should get 103 grains of carbonic gas and 9.4 grains of flammable gas. So his experiments weren't perfect. Right? Which is not surprising, and there's no shame in that. But what would we now say he did wrong?



Student: He adjusted his results.



Professor Michael McBride: He adjusted his results to fit his theory. Okay? But this was early days. He thought, he actually said, on page ninety-two: "I have thought it best to correct by calculation and to present the experiment in all its simplicity." Now what's wrong with presenting an experiment in all its simplicity? You lose anything that's new, anything that you might have discovered in the experiment. Okay? But he was discovering a lot of stuff, so we'll forgive him that. Okay? He was certainly a great scientist.



So what were his contributions? Clarity, above all. Right? This idea of facts and ideas and words having to go together and improve together. And of the facts he worked on, there were these -- the apparatus I showed, and others as well. The important thing was quantitation, using numbers, which people hadn't used before -- not nearly as effectively at least -- concentrating on making things balance, using mass as the criterion for how much stuff you have. But also, in the case of gases, volume, but measuring the density, so that you can relate volume to mass, which is the fundamental quantity.



He found some new substances -- although that wasn't much of a contribution on his part; other people, like Scheele before him, had discovered lots and lots of new pure substances; that wasn't Lavoisier's strong suit -- and reactions. And then, with respect to ideas, the idea of what an element is; that it's not some platonic ideal, but it only means as far as we've so far been able to go in analyzing things and breaking them down. And if we can't break them down further, we'll assume for the time-being that they are elements, but maybe that they won't be in the long run. The idea of conservation of mass, of course, is key. Oxidation as the organizing principle. The idea of a radical, that an element is the base of something and that it can become an acid through oxidation. And the idea of salts, from mixing of elements. And with respect to words, the idea that names should be meaningful; in particular, the names for elements, oxidation state. So he developed these suffixes; -ous, -ic, -ide, -ite and -ate are all due to Lavoisier. And the idea of a salt composition, the way of naming them, that sodium hydrox-ide, for example, or sodium chlor-ide, or chlor-ate, or chlor-ite; which are different degrees of oxidation of the element.



But, like everyone who would follow, and everyone before, and everyone nowadays, he was plagued by lack of imagination. Right? That's always the short suit of everyone who's in science, or probably in anything. He said, at the end of this preface that we've talked about: "Chemistry's present progress, however, is so rapid, and the facts, under modern doctrine, have assumed so happy an arrangement, that we have ground to hope, even in our own times, to see it approach near to the highest state of perfection of which it is susceptible." Right? So people always think that; that there are a few things you see you have to do to fill in little chinks here and there, but otherwise we've got it now. Right? This is certainly what a lot of people tended to think about the human genome. Once you know what the human genome is, then all diseases are solved, everything else. Right? It's never been true and it almost certainly is not true now. And it certainly wasn't true in his time. He said, "even in our own times." He wrote this in 1789, and in 1794 he was guillotined. [Laughter] So it certainly didn't happen in his own time. The judge, Coffinhal, is reported to have said, although perhaps apocryphally, that "the Republic has no need of geniuses." And, in fact, all his equipment was seized for The People, including his 80 pounds of mercury that he carried around in this thing, to measure gases. That madness lasted only a couple of years; maybe not even that long. And his widow recovered the equipment, and it currently is in the Musée des Arts et Métiers in Paris. I've never seen it, but a number of students from the class have sent me postcards when they've gone to see it. Okay, and Lagrange, the guy that helped him build the calorimeter [correction: it was Laplace who collaborated on the calorimeter], wrote: "It took them only an instant to make this head fall, but 100 years may not suffice to reproduce one like it." So what we're going to look at is what the heads in the next 100 years were able to accomplish.



So today we're going to do the first twenty, or twenty-five, or thirty years after Lavoisier, to see where things went. So we've looked at Boyle and his gas laws, and Lavoisier. And now we're going to go on to Dalton. So here's John Dalton, who lived up near the Lake District in England. And he was an amateur meteorologist. So he was interested in keeping weather records and so on. And in particular he was interested in why, in the atmosphere, if you collect air at the top of a mountain, and down in the bottom of a valley, you get the same composition. Why don't the heavier gases sink and the lighter ones rise? Good question, right? Well people on the continent had proposed that different gases attract one another. So they tend to stick and therefore stay mixed. Right? But Dalton had a different idea. And this is a picture he drew of gases. The top is Azote, or nitrogen, and the bottom is hydrogen. So he said, "The atoms of one kind did not repel the atoms of another kind." It's not -- well you'll see in just a second.



Okay, so the center is his symbol for the atom; nitrogen in that case. And surrounding it, this sort of halo, is the "envelope" that's heat associated with that particular atom. So these rays come out. And if the rays meet one another, between adjacent atoms in the gas, then they repel one another; according to Dalton's theory. But if they don't match, as in hydrogen and nitrogen, then they don't repel one another. Right? So atoms of the same element tend to stay apart, but atoms of the other element are completely free to be inside there. They knew that the pressure went down, as you went up; that was no question, right? But why don't they stratify? It's because the different elements don't repel one another. So he substituted homorepulsion, that the same atoms repel one another, for what the Europeans used as heteroattraction. Okay? Obviously that turned out to be nonsense. But Dalton developed three principles, in terms of these atoms that he had started using for this purpose of something about the atmosphere. These principles have come to be known as Definite Proportions, Equivalent Proportions, and Multiple Proportions.



So Definite Proportions is that pure compounds always have the same weight ratio of their elements. So you don't get a sample, of sugar here, that's a certain percentage of carbon and hydrogen, and a slightly different percentage of carbon and hydrogen here. And he explains it by saying that these compounds are a given ratio of elements. So you have that same ratio all the time when you have a particular compound.



Now Equivalent Proportions is a pretty simple idea, but it looks complex. Okay, so suppose you have a certain weight of A reacts with a certain weight of B; okay, to form AB. Okay? And a certain amount of A reacts with a certain amount of C, to form AC. But if it's the same amount of A, then in some sense that weight of B and that weight of C are equivalent to one another. Okay? Because they react with the same weight of A; so they're equivalent. But now suppose you have another compound D that also reacts with b parts of B. A reacted with b parts of B. Right? What does that tell you, about equivalence?



[Students speak over one another]



Professor Michael McBride: That means that a parts of A is equivalent to d parts of D. Okay? So just suppose two grams of A reacted with three grams of B, and two grams of A reacted with eight grams of C. That means three grams of B is equal to eight grams of C, for purposes of combining. Okay? And suppose 1.5 grams of D also reacted with three grams of B. Then 1.5 grams of D is equal to one gram of A. Okay? Now what could you conclude? What's the final line going to be? What do you predict from this, if you have these kinds of equivalence; an equivalence between B and C and an equivalence between A and B? Can you see? Then you predict that if you react D and C, what's going to happen? What weight ratio will they be in?



Student: d parts of --



Professor Michael McBride: Okay, d parts of D will react with c parts of C. Right? Because the D is like the A in the beginning. The C is like the B in the beginning. So d parts of D should react with c parts of C. That's what multiple [correction: Equivalent] Proportions is. I'll show you an example. Okay? And finally Multiple Proportions. If the same two elements form several different compounds, the weight ratios are related by simple factors. And I'll show you an example of that. So these are all things that Dalton adduced as support for the idea that things were made of atoms, fundamental units.



Now Definite Proportions. Does the same compound always have the same ratio of the elements in it? The French School was divided on this question. Berthollet said, "Non!" Because he said they're metal alloys, or natural "organic" materials that vary in their analysis. Right? Okay? And Proust said, "Oui!" Right? And that defined what chemistry was; that chemists then began to deal only with things that obeyed this Law of Multiple [correction: Definite] Proportions. They didn't deal in fundamental ways with alloys or with complicated big molecules that had really complicated formulas, or couldn't be separated as mixtures and therefore got different analyses at different times. They assumed that anything that was pure had this -- obeyed the Law of Definite Proportions. So it was really more of a definition, than a law.



Okay, now Multiple Proportions. So there were oxides of carbon; you remember Lavoisier, 1789, defined 44% carbon and 56% oxygen in CO2. And again there are problems for you to work together to see how he did some of these things. So that's one of them, I believe. And then in 1801 carbonic acid, as it was called, was analyzed and found to be twenty-one grams of carbon to seventy-two grams of oxygen. Okay? Or in the case of nitrogen, there were several oxides that had these ratios, all as analyzed in 1810. Right? Now, you can look at the weight ratios, if you want to. Oxygen to carbon ratio in the first one is 1.27, the other is 2.57. Or you can look in the oxides of nitrogen, and they're 0.58, 1.27, 2.39. Right? Now, what do you notice about these? Do you learn anything from this? These guys got these results and sat down and scratched their heads. What did they come up with? Anything profound? Any relationships among these numbers, that you can see? Or are they just sort of random numbers? Yes. Nick?



Student: The ratio is kind of like --



Professor Michael McBride: Can you speak up a little bit?



Student: 1.27 is half of --



Professor Michael McBride: Ah, 1.27 is very close to half of 2.57. It's not exact, but there's going to be experimental error. Right? It's about one to two; it's actually 2.02. And if you look in the case of nitrogen, it's 1:2:4. Right? It's actually 2.19 and 4.12. So as always, when people develop things experimentally and observe regularities, there's a question of what's real and what's just experimental error; and you have no way of knowing at the beginning. We have a way of knowing now because we know what the true weights are. So they're integral values consistent with simple atomic ratios. So now we know what the percent errors were in these things. Okay, that the ratio of the error was about 5% or 4% off from what it should've been in the oxygen:nitrogen. One is minus 2%, the others are plus 11%. So within roughly 5% or so, they were doing a pretty job of the analysis. But they were able to see through to this Law of Multiple Proportions. And that's what you would expect, if these things were made up of identical molecules with simple ratios of numbers of atoms in them. So this tended to support Dalton's ideas of atoms.



Okay, so we've looked at Lavoisier. And now we're on further into elemental analysis, doing accurate elemental analysis, and the idea of atoms and dualism. And we'll see what that is right away. And to do that we're going to focus on Berzelius, but also talk about Gay-Lussac. Okay, so there's Faraday, whom we've talked about already; Berzelius and Gay-Lussac, and also Davy, in England. Okay, so here's Berzelius in Sweden. Born in 1779. Notice on the previous slide, all those guys were born within a year, 1778, 1779. Another guy that was born at the same time didn't make it to the list, but his statue is right below, who's Benjamin Silliman; he was exactly the same time. Okay, so here's Berzelius, who lived to be almost seventy years old, and did an awful lot of stuff. So he was very good at analysis, both of organic substances and of minerals. In a period of six years he accurately analyzed 2000 compounds for their elements. Okay? He developed really good atomic weights. Remember, we just saw that these previous ones were off by as much as 10 or 11%, but he got very good ones for fifty different elements. Okay? He studied electrolysis, which we will see, and used that to develop the theory of dualism and how -- understanding reactions as double decompositions; and we'll shortly see what that is. And he did a lot of teaching and writing. He was the most respected chemist worldwide. He chose in his picture -- remember, when these people got these pictures painted, they always chose some emblem to put in, and he chose his textbook that he wrote in 1803 [correction: 1808]. But he was a very important editor and summarizer of the year's progress in chemistry every year. And he developed the notation that we still use for composition. So you'll see that.



Okay, notation for composition. We're going to look at its evolution, from alchemy through Dalton to Berzelius, which is what we use, almost exactly. Okay, so in 1774, this is a table printed in Sweden of symbols for elements and compounds. And you see, for example, here's vitriolum cupri, or vitriolum coeruleum. Do you know what about cupri; that's the symbol there of copper, right? So why would it have been called coeruleum? Can you see any word that's related to that?



Student: Cerulean.



Professor Michael McBride: What's cerulean about it?



Student: The color.



Professor Michael McBride: I got some here. That's not it. That's not it. This is it. It's cerulean, right? Like the sky. Okay. Or iron, there's the alchemical symbol of iron. Or and vitriol means sulfate, right?; or mercury sulfate; okay, lead. Okay? So these are the chemical symbols. Notice particularly nitrum, at the top, that comes from nitrates. Okay, and we'll focus on that and see that that's exactly the same symbol that Dalton used for nitrogen. So he derived his symbols for atoms from the alchemical symbols; at least many of them. Not all of them. Here are some others. And here are the symbols that Berzelius used for the same elements, which you see are the ones we use now. Okay? So notice as -- so there were alchemical symbols at the beginning for Dalton, as you went across, but in the next row, once you get to iron, which is Fe in Berzelius, can you see what it is? It's a little hard to read at this resolution. It's a little circle with a letter inside. What's the letter for iron? Can you see it? It's in the middle of the second row. It's an 'I'. I'm sorry it's not such a good resolution. And the next one is Z and then C and the L. Why? Z?



Student: Zinc.



Professor Michael McBride: Zinc. C?



Student: Copper.



Professor Michael McBride: Copper. L?



Student: Lead.



Professor Michael McBride: Lead. And so on. Okay, now down below that are binary symbols. So you put two elements together, like H with O, H with N, N with O, H with C, O with C. Now here's Dalton's logic about these symbols. He says, "When only one combination of two bodies can be obtained, it must be presumed to be a binary one." That is, if you get only one compound, it must be one atom of one and one atom of the other. Right? So what would water be?



Student: H.



Professor Michael McBride: HO, or his circle with a dot in the circle. Right? "Unless some other cause appear to the contrary" that would show you that that's not what you should assume. Okay, the next row, the ternary ones: N2O, NO2, CO2, CH2. And what he said about that is: "When two combinations are observed, of the same elements, then they must be presumed to be a binary and a ternary." So you could have NO and NO2. What if there are three combinations of the same elements that give compounds? Do you see what he's going to observe, what he's going to suppose?



Student: Binary.



Professor Michael McBride: Right? "When three… a binary, and the other two ternary." Okay? Okay, and what about if there are four? "When four… one binary, two ternary, and one quaternary." Okay? So he's laying down rules of logic for what, the way atoms must behave. And it turns out, as we know, atoms have different ideas from Dalton's logic. There's nothing illogical about Dalton, it's just wrong. Okay, and this is Berzelius's notation for the same symbols that Dalton gave on the left. But you notice they're completely recognizable to them, except for one thing. What's different?



Student: Superscripts instead of subscripts.



Professor Michael McBride: Here was ours, and here is Berzelius's. The only difference is he used superscripts instead of subscripts. Okay? And he decided to base the names on Latin, so it would be international, whereas Dalton had used English. And the formulas, notice, are not structural. Berzelius's aren't, they're just written in a line and the number. These could've been assumed to be structural, that you thought the atoms were arranged that way. I don't know that Dalton really had that idea, but they sort of convey that idea. But Berzelius's don't, they're just a list of the elements and the ratio, by weight, of atoms in them. And he developed some more fancy abbreviations as well, where, for example, these little dots above elements would who show how much oxygen was associated with it. So that chromate potassique was K with an O, and chromium with three Os. Or down here he has some even fancier things. Notice the superscripts denote the number of atoms. And this benzoic acid should be H10C14O3. Now wait a second, a carboxylic acid has how many oxygens? Carboxylate group?



Students: Two.



Professor Michael McBride: Two. Here he's got three. The reason is that he felt there was water in these acids. So he heated the heck out of them to drive the water out, and what you actually were left with was the anhydride. So two acids had come together and lost water. Right? So the formula should be C[H]10C14O3, and he's got H12C15. So not so far off. And he's got these neat symbols for acetic, tartaric, citric and so on, acid. But those didn't catch on either. But his symbols for the elements did.



Now let's see what we got here, I had one more topic. Atomic weights and equivalents. Okay, these are Dalton's atomic weights that he used in 1808: hydrogen one, carbon five, nitrogen five, oxygen seven, and so on. Now as of 2004, this is what they should've been. What's wrong with Dalton? Why is he off so far, for example, in oxygen, which should be sixteen, and he has seven? How can he be off that much? Charles? Can't hear.



Student: Diatomic.



Professor Michael McBride: Yeah, what did he think water was?



Student: HO.



Professor Michael McBride: HO. We know it's H2O. So the ratio is off by a factor of two. Right? So all these are off -- most of these are off by a factor of two or three, depending on what the true formula we now know to be. Okay? But if you make that correction, this is what kind of errors he had in his atomic weights. There were roughly 10% or so errors. Right? And the question in silicon chloride. Thomas Thomson in Scotland thought it was SiCl. Gmelin in Germany thought it was SiCl2. Berzelius thought it was SiCl3 and Odling in England thought it was SiCl4. So you could have anything you wanted. Right? So which one is it? How can you tell which one it is? That's the problem. Okay, this is when Gay-Lussac in France gets into the act. So in 1804, he and the physicist Biot made a balloon ascent to a height of 7000 meters. That's way the heck up there. And in fact, that was a record for fifty years. And they have a big flask they're carrying. Why? What do they want to do with their big flask?



Student: The flask was to collect air.



Professor Michael McBride: Collect air at 7000 meters and see if it's different. Okay? Okay, so they established that the atmospheric composition is invariant with altitude; at least at the heights they could do. But that's not the most important thing he did, for our purpose. One of the really important things he did was figure out how to oxidize sugar, how to burn sugar cleanly. Remember, Lavoisier had to resort to fermentation. But he found if the source of the oxidation was sodium chlorate, then you could heat them together and actually get it right and get water and CO2 out of sugar. So that was a very important practical contribution. For the present purposes, he found that when you decompose water you get, to one volume of oxygen you get 1.9989 volumes of hydrogen. And if you do the same thing with ammonia, for a volume of nitrogen you get 3.08163 volumes of hydrogen, in his experiment. What did he conclude? What was Gay-Lussac's conclusion for this?



Student: H2.



Professor Michael McBride: What do you think he thought the values would've been if there was no experimental error? How much hydrogen would he have gotten?



Student: Two.



Professor Michael McBride: Two. And how much hydrogen would he have gotten from nitrogen?



Student: Three.



Professor Michael McBride: Three. Right? They're very, very close, just a few percent off. Right? And even closer in the case of water. So what did he think? He thought the number, that the volume of a gas was proportional to the number of atoms in it. Right? Whose name do you associate that with?



Student: Avogadro.



Professor Michael McBride: Avogadro. But both of them did it, independently, and Gay-Lussac is the one whose data I'm showing here. Okay, so this was an alternative to Dalton's Law of Greatest Simplicity. Dalton said if there's only one compound, it must be one to one. Gay-Lussac said look at the gas ratios. And that turned out to be correct; although it didn't get completely adopted for another sixty years. That's really funny. Okay, that's enough for now.



[end of transcript]

Lecture 21
Berzelius to Liebig and Wohler (1805-1832)
Play Video
Berzelius to Liebig and Wohler (1805-1832)

The most prominent chemist in the generation following Lavoisier was Berzelius in Sweden. Together with Gay-Lussac in Paris and Davy in London, he discovered new elements, and improved atomic weights and combustion analysis for organic compounds. Invention of electrolysis led not only to new elements but also to the theory of dualism, with elements being held together by electrostatic attraction. Wöhler's report on the synthesis of urea revealed isomerism but also persistent naiveté about treating quantitative data. In their collaborative investigation of oil of bitter almonds Wöhler and Liebig extended dualism to organic chemistry via the radical theory.

Transcript

October 24, 2008

Professor Michael McBride: Let's get started. Welcome to the parents who've made it today. Thank you for sending us your bright people to pay our salaries and learn from us. We've spent most of the semester so far figuring out what bond[ing] is, and how quantum mechanics works, and so on. But now we're trying to find out how it really happened -- how people discovered how chemistry in general, and organic chemistry in particular, worked. And we started with Lavoisier. And now we're in the next generation, which is Berzelius in Sweden. And, as we said last time, he was important for analysis. He got good atomic weights for fifty elements. We'll talk today about electrolysis and dualism. He was important for his teaching and writing. For the first half of the 19th Century he was the main man in chemistry; and his notation for composition that we talked about last time, that we still use.

Okay, so now we're going to talk about atomic weights and equivalents and what was involved there. We started this last time; remember, showed how Gay-Lussac had an alternative to Dalton. There was this question about what, even if you knew the proper ratios of elements, what you didn't know was the ratios of the atoms that were in them, because you didn't know whether an atomic weight should be doubled or halved. So they had the question about Silicon Chloride, was it SiCl, SiCl2. SiCl3, SiCl4? So how could you possibly settle something like this? Dalton did it by assuming that if you had only one compound, it had to be one-to-one. Gay-Lussac went with the volumes of the gases that were involved, which turned out ultimately to be correct, although it wasn't really widely accepted for about sixty years. But there was another neat way of doing it, depending on crystals. Mitscherlich, an Austrian, had this device on which you mount a crystal and reflect light off it through a telescope. So when you do that, you know precisely -- if it comes right up the middle of the telescope, you know exactly what direction the plane of that surface of the crystal is. So what you can do then is rotate it by a certain number of degrees, until some other face of the crystal reflects precisely, and then you can measure very, very accurately what the angle between those two faces is. So that was crystallography, in those days, before X-ray. Okay?

So here's his paper on "Relating Crystal Shape to Chemical Proportions." They sound like not closely related phenomena. But here's he's talking about arsenates and phosphates. And here, the pictures he drew -- they drew beautiful pictures of crystals in those days; if that's all you can do, you do it right. But he saw that those crystals were the same shape for corresponding arsenates and phosphates, this diammonium biarsenate and the diammonium biphosphate. Right? So he measured the angles. And here's his table of the angles corresponding. And you can see they're very, very close to one another. Those are the differences in tenths of degrees. So these crystals are precisely the same. Now what could he conclude from that? Anybody got an idea, where you can go from here? So you know that there's an arsenate and a phosphate with the same chemical composition. Right? And they give exactly the same shape. So the things that pack together must be exactly the same. So they must have the corresponding number -- however many arsenates there are -- arsenic atoms there are in one, there must be exactly the same number of phosphorous atoms in the other. Otherwise they could not possibly be exactly the same shape.

So now you know something. There was this need for relative atomic weights. Was Dalton right? Was water HO? Or was Gay-Lussac right, that it was H2O? Right? At least in this case, this thing called isomorphism, the same shape of the crystals, provided definitive atomic weight ratios for at least some of the atom pairs. So arsenic atoms must play exactly the same role in arsenates that phosphorous atoms play in phosphates. And therefore, if you have 100 grams of nitrogen, oxygen and hydrogen, that combines to give these salts, with 30.64 grams of phosphorous, or 78.11 grams of arsenic, it must be that arsenic atom is 2.55 times as heavy as the phosphorous atom. Right? That is, it can't be half that, or twice that. It has to be exactly that, because there have to be the same number to give exactly the same shape. Okay? So there was one way of determining this ratio. But it wasn't trivial. Okay?

Now here's Berzelius's Table of atomic and molecular weights from the year 1831. And if we zoom in on it a bit here, you can see he gave two columns for the atomic weights, one based on oxygen being 100 units, and the other based on H, with a bar across it, being one. Now, the bar across atoms means that the atoms are doubled. You can have it without a bar, which is a single atom; or the bar across, that means it's a doubled atom. So he had -- actually you notice, if you can go down the column there, that if you assume H2 as 1, then obviously H here is a half. Right? So these were the numbers he had in 1831. And we could compare them with what they would be nowadays, if you scaled to oxygen, as we define it now, to be 15.9994. And then you see that they're off by about 1%, up or down. How does that compare with what Dalton was getting? Do you remember how accurate he was? He was about 10%. So here in twenty-five years they've gotten tenfold more accurate in what the atomic weights are. Okay?

Now, combustion analysis was, and would continue to be, from the time of Lavoisier, right through the whole nineteenth century, the heavy carrier for organic chemistry. Okay, here's the way Berzelius did combustion analysis. So he started with a tube in which he put half a gram of organic substance that was going to be analyzed, three grams of sodium chlorate. Remember, that's what Gay-Lussac used to burn things with. Why that rather than air? Why did Gay-Lussac use it rather than air? Do you remember? Because a lot of things wouldn't burn in air, right? They would char. Okay, so but they used Gay-Lussac's oxidizing agent. And fifty grams of sodium chloride was ground up with this stuff, just so that the reaction wouldn't get out of hand, so it wouldn't explode. Okay, so those were put in there. Then he heated up the neck and drew it out, and plugged it into that bulb on the right. Right? So he joined it to that bulb, which would serve for collecting water, which is a combustion product, and a calcium chloride drying tube, which would get the rest of the water that didn't condense in the bulb. Okay? Then he assembled this thing along the bottom so that the gases that came out, after the water was removed, would exit and be collected in this bell jar, over mercury. And in that, there was a bulb that floated, which held potassium hydroxide; solid potassium hydroxide, which would absorb the CO2, thereby separating it from the oxygen that would come through. You get oxygen just by heating sodium chlorate. Okay? And then he closed the top with a piece of glove leather so that mercury, that it was floating on, couldn't -- little drops of mercury wouldn't get in. Because, as you know, from hefting it, that would change the weight a good deal. Okay?

So what he did was build a fire and move that barrier back, so it would heat from one end to the other, and the gases would come out and bubble up in the bell jar, and fill up a good deal of the bell jar. Right? And notice that the tube had been wrapped with metal to keep it from popping when it got hot in the fire, because there'd be pressure inside in order to bubble through the mercury. Okay, so then he cools it off. The gas contracts. Right? The KOH absorbs CO2. So you're left with oxygen in the bell jar, and the CO2 is absorbed on the potassium hydroxide. But he then -- so the mercury rises as the gas gets absorbed. But after the mercury stops rising, he waits twelve hours to make sure he's got all the CO2 absorbed, because it has to get through the glove leather, right? So it takes quite a while. It took much more than a day to do an analysis. So after it stops rising, he takes it apart, takes the glove leather off the top, weighs it, and finds out how much CO2 there was. So he's got the water and the CO2, therefore the hydrogen and the carbon. And this was a much better way of doing it than Lavoisier's, which required three or four people operating that big machine. Okay.

Electricity was another really important contribution from Berzelius. Okay? But it also was a contribution at the Royal Institution in London where -- this is a demonstration of N2O at the Royal Institution. But the guy holding the bellows there is Humphry Davy, who became the head of the Royal Institution, and was the head then of big science. So in 1807 and '8, he made some very important experiments. It was based on the work of Volta, who was just a little more than a decade earlier, who had invented this thing, shown on his desk here, which was the Voltaic Pile, which was a pile of copper and zinc discs. And when you put copper and zinc together, you can get 1.1 volts out of it. Okay? So this was going to be big science. So he put together twenty-four such pairs of things, which meant he could get twenty-six volts. Those were six inches square. Right? Then he made another -- there's the battery, incidentally, that Berzelius had, which is still in Stockholm. Okay? Then he put one with 100 such things, so he could get 110 volts. Then he made another one. They were only four inches square. Now 165 volts, right? And then he could hook them together in series. So he could get 301 volts across this thing. And then he tried doing experiments with it. So this is what he said: "I acted upon aqueous solutions of potash and soda…." You know what potash is?

Student: [inaudible]

Professor Michael McBride: Potassium hydroxide, and soda is sodium hydroxide. "…saturated at common temperatures, by the highest electrical power I could command, and which was produced by a combination of voltaic batteries belonging to the Royal Institution, containing twenty-four plates of copper and zinc, twelve inches square, 100 plates of six inches, and 150 of four inches square." So this is that battery I showed you. So he put 300-and-whatever-number of volts across this thing. "Though there was a high intensity of action, the water of the solution alone was affected, and hydrogen and oxygen disengaged with the production of much heat and violent effervescence." I can believe that. And so he wasn't about to give up. He tried again.

"The presence of water appearing thus to prevent any decomposition, I used potash" (so potassium hydroxide) "in igneous fusion." (So he put a flame to it and melted it.) "By means of a stream of oxygen gas from a gasometer applied to the flame of a spirit lamp, which was thrown on to a platina spoon containing potash, this alkali was kept for some minutes in a strong red heat, and in a state of perfect fluidity. The spoon was preserved in communication with the positive side of the battery"

I don't know how he held this thing. [Laughter]

"of the power of 100 of 6 inches, highly charged; and the connection from the negative side was made by a platina wire. By this arrangement, some brilliant phenomena were produced. The potash appeared a conductor in a high degree, and as long the communication was preserved, a most intense light was exhibited at the negative wire, and a column of flame, which seemed to be owing to the development of combustible matter, arose from the point of contact."

Right? Then he did another experiment.

"A small piece of pure potash, which had been exposed for a few seconds to the atmosphere, so as to give a conducting power to the surface, was placed upon an insulated disc of platina, connected with the negative side of the battery in the power of 250 of 6 and 4, in a state of intense activity; and a platina wire, communicating with the positive side, was brought in contact with the upper surface of the alkali. …small globules having a high metallic luster, and being precisely similar in visible characters to quick-silver..."

What's quick-silver? Live silver, right? Mercury, of course. Okay? "…some of which burnt with explosion and bright flame, as soon as they were formed, and others remained, and were merely tarnished, and finally covered with a white film which formed on their surfaces." What was this? He electrolyzed -- here he says potash, potassium hydroxide. What did he get?

Student: Potassium.

Professor Michael McBride: Potassium metal. So he discovered a new element. And from soda he got sodium; from potash he got potassium. Okay, now this was a great triumph for English science. But this was a period of great rivalry, 1807 and ‘8. The Peninsular War was going on in Portugal, Spain, against the French; in particular, against Napoleon. Now Napoleon had studied science a certain amount as a student at the École Militaire in Paris. In fact, he had been examined by Laplace, Lavoisier's colleague in measuring heat. So he wasn't about to knuckle under to England in this. So he got his crew together, including Gay-Lussac, and said, "Why haven't we discovered this?" And Gay-Lussac said, "We don't have the battery." He says, "You've got your battery."

So this was the battery that Gay-Lussac constructed accordingly. [Laughter] So it consisted of 600 one-kilogram copper plates and 600 three-kilogram zinc plates. So he had 2.6 tons of metal and in six different troughs here. And then he hooked them together in series. [Laughter] So he's now got 650 volts. Now, Napoleon was interested in science. So he came to visit the lab one day. And there's an account of this in the biography of Humphry Davy.

"With that rapidity, which characterized all his motions, and before the attendants could interpose any precaution, he thrust the extreme wires of the battery under his tongue and received a shock, which nearly deprived him of sensation. After recovering from its effects, he quitted the laboratory, without making any remark, and was never afterwards heard to refer to the subject."

[Laughter] Okay, so that was the end of the French competition with Humphry Davy. But Davy himself built one with 1000 plates; so 2200 volts. And with this he was able to prepare a bunch of elements. In addition to sodium and potassium, boron, magnesium, calcium and barium were all due to Sir Humphry Davy. But that wasn't really the most important thing for chemistry. It supplied more than new elements, it supplied an idea, and the idea was the organizing principle for what was called dualism. And dualism was the idea that there were things that were positive and things that were negative, and they would be attracted to one another. Not only would they be attracted to one another, but they could trade partners. Okay?

So remember we saw last time that the alchemists already had vitriolum cupri, or copper sulfate. So here we saw last time was copper sulfate. Let me see if I can get this thing going here. There we go. So there's copper sulfate, nice and blue. Let me get the light behind, that should be good. Okay, so we'll put some in here. So this is copper as a positive stuff, and sulfate as negative stuff. And now we'll take sodium hydroxide, and put a little bit in. So there's a precipitate. And actually it's much more complicated than people thought it was in those days. It gets quite dark blue of you put a lot in there. But now a whitish precipitate is going to form, which you don't see very well here; maybe it doesn't know how to focus on that. So it needs something for it to focus on. There. There you see the precipitate. Okay, so obviously something happened. Right? And according to Dualism, this is what happened. Right? You mix it with caustic soda and you have AB and CD; positive and negative on the left, sodium hydroxide, copper sulfate, and they change partners, and you get a precipitate. So you have -- and it's explained by electricity. And this then is what dualism was.

So, notice that this is our family tree, you'll remember, and now we've looked at elemental analysis, atoms and dualism. Notice that people are entered into this, not by when they made their contribution, because they made it at many different times, but by when they were born. So you can see what the generations were. So first we had Lavoisier, then Berzelius. And now we're into a new generation of people who are going to contribute when they get to be 20, 30-years-old down here. So the next thing we're going to talk about is urea and isomers, and the benzoyl radical, and the theory of substitutions and types. Okay? And our players are Jean-Baptiste André Dumas, in France, and Liebig and Wöhler in Germany. And they were all born right around 1800; between 1800 and 1803. Okay. And here they are on the outside of the building. So we're going to look at Wöhler and Liebig.

Now, here's Wöhler, writing to Berzelius in Sweden -- he's going to be the next generation -- saying, "Having developed the greatest respect for you through studying your writings, I have always thought it would be my greatest good fortune to be able to practice this science under the direction of such a man, which has always been my fondest desire." Those who have seen such letters will recognize that these as consistent up to the present day. He goes on to say: "Although I earlier had planned to become a physician…". Okay? And at the end he says, "With the greatest respect, Friedrich Wöhler from Frankfurt am Main." He was a great guy. They carried on a correspondence their whole lives; it's really entertaining, partly because Wöhler had sort of a lame-ish sense of humor and was a cartoonist. So, for example, in this letter to Berzelius in 1837 -- so this is ten years later -- he said, "To see this old friend [Palmstedt] again, especially here [in Göttingen], was a real delight. He was just the same old guy, with the sole exception that he no longer wears the little toupee swept up over his forehead as he used to do." So Wöhler drew this cartoon in there.

Okay, but what you know Wöhler for, probably, is urea. How many people have heard about Wöhler and urea? Okay. But the interesting thing is that the urea was not the main contribution. He also invented aluminum incidentally, or discovered aluminum. But urea wasn't really the main, for our purposes, wasn't the main contribution from this work of Wöhler. And you'll see that there are a couple of problems for next Wednesday, and readings on the web that support this, with original documents. So he wrote in 1828 to Berzelius: "Perhaps you still remember the experiment I carried out in that fortunate time when I working with you, in which I found that whenever one tries to react cyanic acid with ammonia, a crystalline substance appears which is inert, behaving neither like cyanate nor like ammonia." So that's what was funny, that it didn't behave like you would expect for ammonium cyanate. So the idea was it should be a double decomposition reaction, this dualistic thing, where you change partners. So you start with ammonium chloride, silver cyanate -- that was another way he did it -- and you should get ammonium cyanate and silver chloride, which is a precipitate. So then you've prepared ammonium cyanate. But you can test whether ammonium cyanate is the salt ammonium cyanate. One is to see whether it has the ammonium ion in it, and that you can tell by pulling the proton off and changing it back into ammonia, which you can smell. Okay? But when he treated it with sodium hydroxide, he didn't get ammonia. Okay? Or you can test… So it doesn't behave like an ammonium salt. Nor did it behave like a cyanate salt. If you treat it with acid you don't get the smell of cyanic acid. And if you treat it with lead, you don't get lead cyanate.

So it appears not to be either ammonia or cyanic acid anymore. Okay? But what it did do, when treated with nitric acid, was to give brilliant crystal flakes. So the crystals were what allowed him to identify it. And he knew that you got the same thing when you treated urea with nitric acid, just like those from urea and nitric acid. Right? So this raised a question. Berzelius wrote to Wöhler, in reply to that: "It is a unique situation that the salt nature so entirely disappears" -- it doesn't look like dualism anymore -- "when the acid and the ammonia combine, one that will certainly be most enlightening for future theory." So might ammonium cyanate actually be urea? Might urea be ammonium cyanate, with very curious properties? So Wöhler wrote back to Berzelius: "I recently performed a small experiment, appropriate to the limited time I have available" -- he was doing a lot of teaching at this time -- "which I quickly completed and which, thank God, did not require a single analysis." That was what was a pain in the neck to do.

So this is the data from Wöhler's letter, and the same thing was published as his paper in 1828. So it turned out that Dr. Prout, a physician in London, had already analyzed urea, and reported its composition in terms of nitrogen, carbon, hydrogen and oxygen; that is, its ultimate analysis. And Wöhler, without doing an experiment, knew what the analysis should be for ammonium cyanate, because he had Berzelius's atomic weights. So he could calculate for a substance of that formula what the percent by weight should be. And they agreed, as you can see, very, very well. So it looked like urea and ammonium cyanate had the same analysis, and therefore were the same thing. Okay? The discrepancies were less than 2%. Remember, the atomic weights were only good to about 1%. Okay, so therefore these things appear to be identical. In fact, it's interesting to compare them with the modern values, which you see here. What do you notice in the first row? When you compare the numbers in the first row? Russell?

Student: It seems like they're much the same.

Professor Michael McBride: What's the same?

Student: The nitrogen and --

Professor Michael McBride: Which is better, experiment or theory?

Student: They're the same.

Professor Michael McBride: What?

Student: They're the same.

Student: Experiment is better.

Professor Michael McBride: No, no, the first is Prout's analysis, the first column; the last column is Berzelius-Wöhler theory; and the middle is what we think is correct, because it's based on what we have now. What do you say?

Student: The experiment.

Professor Michael McBride: Ah ha! so there's a good lesson for us. Right? Experiment is better than theory. But if you look more closely, there's a much more interesting story here. So the moral is don't dry-lab; don't be like Lavoisier and try to make your experiments conform to theory. Right? So, because of that, Dr. Prout has a gotten a reputation among historians of science as a paragon of accuracy and honesty; the very model for lab work. Right? Okay, but let's look more carefully. You notice, for example, that he didn't make these things add up to 100%. Would you expect them to -- what would happen if it had been Lavoisier doing it, what would they add up to?

Student: 100%.

Professor Michael McBride: 100%. Exactly. Here it isn't. Why not?

Student: Being honest.

Professor Michael McBride: Because he's been honest; maybe. [Laughter] Okay. So notice, incidentally, that the theory didn't add up to 100%. Now how do you get those numbers? How would Wöhler have gotten those numbers? He decided how many atoms of nitrogen, hydrogen, oxygen there are in the compound. He'd see, according to the atomic weights, how much each of those would weigh. He'd add them all together. That would be the total. And then he'd divide each one by the total to get how many percent it was. Is that clear to everyone, how you go about it? So how do you do that and not get it to add to 100%? Well I recalculated it, using Berzelius's numbers, and that's what I got. Now you'll notice a couple of things here. One, here, you'll notice what?

[Students speak over one another]

Professor Michael McBride: That he just truncated numbers, he didn't round them off. Right? He just truncated the numbers, he didn't round them off. Right? So it's always going to be low; and indeed it is low, his sum. But that's not low enough. Right? What else do you notice? Angela?

Student: He switched them in the last --

Professor Michael McBride: He switched the numbers in the last row. He made an error. But the important thing is he didn't notice he made the error. He added it up and saw that it was 99.8, and thought, well what the heck, it's 99.8. Because this was early days of doing this kind of stuff. Okay? Now, so we come over to Dr. Prout. Is this experimental candor? Well you find out, if you add the numbers, it's actually 99.945. [Laughter] He added wrong. Okay? And now, but those things really agreed. Right? Now you wonder, how did old Dr. Prout do this, to measure it to 1 part in 5000? Right? It turned out he measured the volume of the gas, and what he reported was 6.3 cubic inches of gas. Right? So was he just lucky that it happened to be right to four significant figures? No. Because that's not how he did the experiment. What he reported was not actually experimental. He had a theory about atomic weights. His theory was that everything was made of hydrogen; which isn't so far wrong in terms of weight, right? Because a proton and a neutron have the same weight. A proton is hydrogen. So everything is the sum of protons and neutrons; for weight, the electrons don't count for much. Right? So it's not actually such a bad theory.

So in his system, carbon was six, oxygen eight, nitrogen fourteen; you know, that was what integer you multiply them by. Incidentally, hydrogen was called protyle, from, let's see, from (hyle prote, which means the first substance. Okay? So he had this theory. So what he did was do approximate analysis, get about what the ratio is, figure out what the ratio should be, in terms of small whole numbers, and then multiply them by his atomic weights. So these numbers are not really experimental numbers, they're more like Lavoisier. All right? So what he said was, it was one, one, two, one, and that gives those numbers. Okay? So Prout did dry-lab by making an approximate analysis and reporting results corrected by his theory. And so much for the paragon of accuracy and honesty, as far as we would see it. Okay, but his theory was better than Berzelius's experiments. That's interesting. Okay, so now ammonium cyanate to urea. Let's review that, from the point of view of what we've been talking about. So you have ammonium. What makes it reactive? Dana?

Student: High HOMO.

Professor Michael McBride: High HOMO or low LUMO?

Student: High HOMO and --

Professor Michael McBride: It's got a positive charge. That's one clue. Anybody got an idea? John?

Student: A low LUMO.

Professor Michael McBride: And what do you suppose that low LUMO is, in terms of localized orbitals? We're not going to talk about things that go over the whole molecule. There's not much to deal with here in ammonium. Yoonjoo?

Student: The LUMOs from the σ*,

Professor Michael McBride: Can't hear very well.

Student: The LUMOs from the σ*.

Professor Michael McBride: σ* of N-H. Okay, so that'll be the LUMO. How about a HOMO on the other side?

Student: Unshared pair.

Professor Michael McBride: There's a hint. Okay, the unshared pair on nitrogen with a negative charge. Okay? So we've got these, and we can do the make-and-break trick, attack the σ* with that. Right? So we've gotten back to ammonia and cyanic acid, a possible starting place for getting there. But if you have these two things, you could react them with one another. What makes ammonia reactive, NH3?

Student: The high HOMO.

Professor Michael McBride: The high HOMO, the unshared pair. And how about the cyanic acid? What do you see there that looks like a functional group? Don't worry too much about how big it has to be, just see something that looks familiar.

Student: [inaudible].

Professor Michael McBride: C=O double bond, π*, right? Okay, now so we can -- so one of these can attack the other, and we can arrange them like that and draw a curved arrow. And we could attack the C=O double bond. But you could also attack the C=N double bond. And, in fact, these aren't so different, because when you attack one, and make an unshared-pair on the other, in this case, for example, that high HOMO on the nitrogen, the N-, can be stabilized by mixing with the π* of the carbonyl. Right? So you could actually draw it either way. So it really doesn't make any difference which way you draw it, at the beginning. But what you can do is then have a further reaction, the HOMO on the bottom reacting with the σ* again; although they probably can't line up right, within a molecule. It's probably one molecule attacking another molecule to do this. Okay? And get the hydrogen transfer. And that is urea. Okay, so there's how you go from ammonium cyanate to urea. It's quite an easy chain of events. You could've done it the other way too, like that, and you'd get this compound. But those molecules, the urea and this one, can lose a proton, go back on, lose, go back on, one side or the other, and ultimately you get the one that's more stable, which is the one that has the carbon oxygen double bond; which is a very stable grouping, as we'll learn later, in terms of lore. You can actually understand it in terms of HUMO and LUMO mixing here too. But at any rate, that's what happened.

So the question arises, can ammonium cyanate exist, or does it always change? Is there such a salt, or does it always change to urea? And in fact Dunitz, our old pal and your great-uncle, in 1998, with Kenneth Harris, did an X-ray structure that showed that you can have ammonium cyanate. This was the first time it had ever been proved to exist. And there's what it looks like. But notice, the nitrogens, on the ammonia [correction: hydrogens on the ammonium] there, point toward oxygens, not toward nitrogens. So it's not -- from this arrangement, you couldn't have done that thing that we -- the mechanism that we drew in order to get from here to there.

Okay, so but the fact of getting urea, a product of animal metabolism, from purely (as far as they were concerned inorganic things, which is said to have done away with the idea of vitalism, was not, from my perspective, the most important contribution here. Because this is what Wöhler wrote toward the end of that paper, in 1828: "I refrain from all considerations which so naturally suggest themselves from this fact," -- that is, that ammonium cyanate and urea have the same analysis -- "especially in respect to the composition ratios of organic substances and in respect to similar elemental and quantitative compositions among compounds with very different properties, as may be supposed, among others, of fulminic acid and cyanic acid." We already did those; that was a problem early on when we were drawing Lewis structures. So two different arrangements of the same atoms. "…and of a liquid hydrocarbon and the olefiant gas."

The olefiant gas, the gas that will "make an oil", is ethylene, CH2CH2. But that's very similar to the -- that's exactly the same ratio of atoms as you have in a long hydrocarbon, if it has a double-bond in it. "It must be left to further investigation of many similar cases to decide what general laws can be defined therefrom." Now this is an interesting point, because when people have certain tools available, they tend to think that that's all the tools that could be available, and therefore it gives all the information that's possible. Right? So if what you can do is elemental analysis, you think that once you have the ratio of the elements, you know everything about the substance. But what he shows here is that there can be different substances that have the same elemental analysis, like urea and ammonium cyanate. So the possibility of isomerism showed there was more to be known than people thought, when all they could do was elemental analysis. So fulminic and cyanic acid was very important, as we'll see in just a second.

Now Berzelius wrote a paper just two years later, in 1830, "On the Composition of Tartaric Acid" -- remember we spoke before about how pervasive tartaric acid is in the nineteenth century in the development of organic chemistry--"and Racemic Acid" -- also called John's Acid from the Vosges--;"on the Atomic Weight of Lead Oxide, together with General Remarks on Substances that have the Same Composition but Different Properties." So he found that tartaric acid, and this John's acid from the Vosges, also called racemic acid, had exactly the same analysis, their salts did. Right? And here, in fact, is a picture of it. I have the actual stuff right here. This is from a cork from a wine bottle, and on top of it here, if we're lucky -- see. You can see those, and that's -- we took one of the crystals and put it on the X-ray. There's what they look like. They're beautiful little crystals. And they're calcium tartrate. So remember, you get -- tartaric acid grows from tartar on the walls of the wine barrels.

But this is what was important. He wanted to give a name for this phenomenon, things that had the same analysis but were clearly different.

"I thought it necessary to choose between the words homosynthetic and isomeric. The former is built from homos, equivalent and synthetos, put together; the latter from isomeres, has the same meaning, although it only properly says put together from the same pieces. The latter has the advantage with respect to brevity and euphony."

So it's shorter and it sounds -- it's easier to say. So from now on we're going to choose "isomer" to talk about this phenomenon. "By isomeric substances, I understand those which posses the same chemical composition and the same atomic…." By that he means molecular weight, but different properties. Okay, so that's where "isomers" came from. So it showed there was more to be discovered. There must be something about how these pieces are put together that's different. But how can you possibly find that out, just by doing reactions and weighing things? Okay? Same chemical composition but different properties. Okay, so there's more to chemistry than the analytical composition, which is what they were good at measuring. Now we know the importance of atomic arrangement, or structure. So there are four Cs that we'll organize things about it: Composition, Constitution, Configuration and Conformation. But you've got to be patient. It'll take a little while to get there; it was several more decades.

Okay, now these isomers of HCNO have been calculated to death in 2004, and we actually did a problem set and showed these earlier on. But here's what they are. There's cyanic acid. And notice, this is the same anion, but you pull the proton off one end and put it on the other. It's somewhat less stable, but they interconvert easily. Okay? There's the one, when you put nitrogen in the middle. It's considerably higher. That's called fulminic acid. And if you put the hydrogen on the other end of that one. Now the bottom one is what Wöhler was working with, but the top one was being studied by the other guy, Liebig; fulminic acid. And the name comes from, which means lightening. Because like so many young lads, Liebig enjoyed things that would blow up. Right? And he blew the window out of the pharmacy where he was serving as an apprentice. Right? And then he went to Gay-Lussac, to work on silver fulminate, which is not great stuff to work on, as people who know from fulminates know. But anyhow in 1824 he was in Paris working on that. And Gay-Lussac noticed the paper from Wöhler on the analysis of silver cyanate, and saw that they had the same analysis. So he got them together, Gay-Lussac did.

Now notice -- and we're going to talk much more about the interaction of those two guys -- but notice in this portrait of Liebig that there's a device down in the bottom right. Remember, people, when they got their portraits made, always put something they thought was significant there. So this is what he thought was significant. And here it is. It's not the actual one. This is made by our glassblower. And you've seen it before. For example, you come by this when you're hurrying to class. And notice right there, same thing. Right? Or, if any of you are student members of the American Chemical Society, that's on the logo. Because that was the most important device of the 19th Century for organic chemistry. It's called the five-bulb apparatus, the Fünfkugel Apparat. Okay? And it was used by Liebig for analysis. So here's how Liebig did his analysis. And you can see the five-bulb apparatus here. So he had a tube, much like Berzelius's, in which he put burning charcoal in there, and he'd heat it up and generate the gases, but on a much smaller scale than Berzelius did. Okay, it came through a drying agent, that collected the water that was a product. And then it came in to the Fünfkugel Apparat, which was the CO2 collector. So in here was sodium hydroxide. Right? And the way it worked was it was tilted a little bit. You notice that it's propped up here, with some pieces of wood, so that the bottom isn't flat. It's tilted like this. Okay? So the gas comes in. And because it's tilted, it bubbles from one to the next. Right? So it mixes well and absorbs. So you don't -- remember, Lavoisier needed a couple, a guy standing there to stir those things so that the CO2… This was self-collecting, right? And then at the end you cool the thing off. Now notice -- what do you notice about the bulbs here?

Student: The size.

Professor Michael McBride: They're different sized. And you had to put it this way, not this way. Why did you have to do it this way? Because what happened at the end? You stopped the fire, and it would suck back. So this had to be big enough to hold all the liquid. This one didn't. So they were a different size. You don't want to make both of them big, because you want to keep it as light as possible. It was made out of very thin glass. So there's only one that still exists, that Liebig had. So it was tilted for the reason we said. And here's the original one, which is still in Giessen, in Germany, where he was a professor then. Right? But this was a great advance because, with this, a single student could do three analyses in a day, rather than Lavoisier's big machine that required four people working forever; or Berzelius's, which required waiting overnight for the CO2 to be absorbed all, and so on. So this was the most important thing. And everybody, all the students, who wanted to be organic chemists, went to learn how to do this, in Giessen. So Liebig is almost everyone's great-great-great-great-great-whatever-grandfather in chemistry. Because all the people who went out to teach studied how to do this with him in Giessen.

So here's a portrait, a little bit later portrait, of Liebig, after he had gone to Munich to be a big wheel. And but again he chose in his portrait to have a picture of the Fünfkugel Apparat. But if you look at it carefully, you see something very interesting. Do you see what it is? As far as I know, no one ever noticed this before. It's hooked up backwards. How can he possibly have hooked it up backwards? He must've told hundreds of students over the years, "Look, when you put it together, you put it this way, not this way." It's like which tube in a condenser you connect to the water. Right? [Laughter] It's exactly the same kind of thing. Right? So I just think that's amazing, that he didn't notice that. But anyhow. There we got it backwards.

Okay, this was his teaching laboratory, which still exists in Giessen. It's a museum. It's interesting to go there. And there's, on the wall, there's this thing, and it shows all of his students and grand-students and great-grand-students who became distinguished professors, and the ones in red are ones who got the Nobel Prize; the ones with red on them. And you could extend it now. This stops in the middle of the 20th Century, and practically everyone's descended from Liebig, including you. Okay? This is his lab. And here, this is, here's Ortigosa, from Mexico, and he was supposed to be so good at analysis that he could do it better than Liebig did. Right? So he's the one that got to hold the five-bulb apparatus. And what do you notice about it? He's holding it upside down; the potassium hydroxide would run out. Right? And this guy is A.W. Hofmann, who -- there's this thing of him: "A master and shining teacher of chemistry, a triumphant discoverer of aniline and aniline dyes." And he also was the guy that was really good at languages, and devised a lot of the names we use nowadays. So we'll talk about that later. And here in the background, looking through a window into the lab, is it looks to be -- I'm not sure that that's what it is -- but it looks like a bust. Does it look like a bust to you? And if it's a bust, it's probably this bust, of Liebig. So there. [Laughter] It's fun to go there.

Okay, so in 1832, Liebig and Wöhler collaborated to develop the radical theory. And we'll just give an introduction to that and not go all the way through it. So Liebig and Wöhler remember had been introduced by Gay-Lussac. So they were of the same age and doing similar kinds of things. So they first met in Frankfurt in 1825. And five years later, when they were writing, they first used the familiar; you know, like the tu in French or du in Germany, if you're really great friends with someone, or within the family you use that. So they became great pals. And in 1832, Wöhler said: "I long to do some more significant work. Shouldn't we try to shed some light on the confusion about the oil of bitter almonds? But where to get material?" Right? So Baker, J.T. Baker didn't exist then. But if they were now, they could just order up a bottle of it. So can you smell this?

Student: Yeah.

Professor Michael McBride: Do you recognize the odor? Pass it around carefully, don't drop it. [Laughter] It's not that bad, but you don't want to drop it. See if you recognize the odor. Like almonds, the oil of bitter almonds. Okay? Okay, so in June of 1832, just the next month, Liebig wrote to his pal Wöhler and said,

"My poor deal Wöhler, how empty is every comfort against such a loss." (Because Wöhler's young wife had just died.) "When I think how content and happy you were during your move, what attachment and love you had for one another… The good wife, so young and full of life, so irreplaceable for her parents and for you. Come to us, dear Wöhler. Although we may not be able to give you comfort, we will perhaps be able to help you bear your grief. Staying in Cassel at this time would only be detrimental to your health. We need to be busy with something. I have just been able to get some amygdalin from Paris, and I am ordering 25 pounds of bitter almonds. You must not travel, you must busy yourself, but not in Cassel. I haven't had the courage to tell my wife yet." [Laughter] "Come to us, I expect you at the end of this week."

No, what he didn't want to tell her about was the death of Wöhler's wife, obviously. Okay. So then in July Wöhler says it came: "The oil of bitter almonds has come with the books from Paris. I've kept half of it and am herewith sending you the rest. I've already started all kinds of experiments with it, without being able to obtain any precise results. It seems to be a hard nut." This is an example of Wöhler's lame sense of humor. He did this all the time. "I'm coming soon to you and will be able to report." So they worked together for two weeks on this and developed the most important theory, on the basis of these experiments. Wöhler returned. He said: "Here I am again in my gloomy lonesomeness, not knowing how to thank you for all the love with which you took me in and kept me for so long. How happy I was to work with you from moment to moment. Herewith I'm sending the paper on oil of bitter almonds." So next time we'll talk about what they did with their oil of bitter almonds.

[end of transcript]

Lecture 22
Radical and Type Theories (1832-1850)
Play Video
Radical and Type Theories (1832-1850)


Work by Wöhler and Liebig on benzaldehyde inspired a general theory of organic chemistry focusing on so-called radicals, collections of atoms which appeared to behave as elements and persist unchanged through organic reactions. Liebig's French rival, Dumas, temporarily advocated radicals, but converted to the competing theory of types which could accommodate substitution reactions. These decades teach more about the psychology, sociology, and short-sightedness of leading chemists than about fundamental chemistry, but both theories survive in competing schemes of modern organic nomenclature. The HOMO-LUMO mechanism of addition to alkenes and the SOMO mechanism of free-radical chain reactions are introduced.




Transcript



October 27, 2008



Professor Michael McBride: So in 1832, as we saw at the end last time, Wöhler went to work for most of a month with Liebig, and together they worked on the oil of bitter almonds, which is making its way around -- it's benzaldehyde -- which is making its way around here. So there's the bottle that's coming around. What did you notice about it? Where is it? What do you notice about it Devin?



Student: [inaudible]



Professor Michael McBride: Yeah.



Student: About the smell?



Professor Michael McBride: What does it smell like? Is it almonds? Turn it around, look at the -- turn it upside down. What do you notice?



Student: There's a little bit of solid in the bottom.



Professor Michael McBride: Yeah, it's half solid. Benzaldehyde is the oil of bitter almonds; not the solid of bitter almonds. What does it say on the label, written by hand? Somebody's initials, J.W.



Student: J.W., 11/29/95.



Professor Michael McBride: So it was opened in 1995, thirteen years ago. When it was opened it was a liquid. Now it's half solid. How do you figure that? Well something must've happened. So what did it react with, in the bottle? Yoonjoo?



Student: Oxygen.



Professor Michael McBride: Oxygen. Okay? So oil of bitter almonds; they analyzed, C7H6O, and they found that it reacted with oxygen to get C7H6O2. Right? They also reacted it with bromine. The halogens were common, as we'll see, for reagents in those days. Got C7H5OBr. They reacted it with chlorine, and also that product with potassium iodide, or ammonia, or lead sulfide. And they got all these compounds. And what did they do with them when they got them? They smelled them. Right? What was the main thing they did?



Student: Tasted it.



Professor Michael McBride: No that's not the -- it's true that they probably tasted, that they tasted them, but that's not the main thing they did that gave them unique information.



Student: Weight.



Professor Michael McBride: What about the weight? What was the main technique that we've been talking about all the time? Dana?



Student: The analysis of combustion.



Professor Michael McBride: They did an elemental analysis; ultimate analysis of what the ratio of the elements. So that's what they had. Right? So big deal. What do you get from the analysis? What can you infer? Here were these guys; you could now compete with them, right? This is what they found out during that month of experimentation, and they came up with a theory that revolutionized organic chemistry, as of that time and for the next twenty years. So what did they notice? Claire?



Student: They noticed the carbon chains.



Professor Michael McBride: Pardon me?



Student: The carbon chains.



Professor Michael McBride: Well how would you know it's a chain? The numbers don't tell you it's a chain. What?



Student: The presence of the hydrocarbons.



Professor Michael McBride: Well it's, it's got hydrogen and carbon in it. That's true. But they also have oxygen and nitrogen, chlorine and other things. What could you make out of this? Brian?



Student: C7H5.



Professor Michael McBride: Yeah. All of them have at least C7H5. Some of them have more hydrogen than that, but nothing has less than C7 or less than H5. Can anybody carry that any further?



Student: C7H5 is non-reactive.



Professor Michael McBride: Pardon me?



Student: It's non -- the C7H5 is the unreactive part.



Professor Michael McBride: Yeah. What would they have called it; the thing that you have it there and then it reacts and gives this, that, or the other thing? What did Lavoisier call it?



Student: The radical.



Professor Michael McBride: The base, or the radical. But actually the radical is more than C7H5, the thing that persists. What else?



[Students speak over one another]



Professor Michael McBride: O. So C7H5O persists through all these transformations. So it looks like that's some sort of a core that gets modified. Okay? But it's there all the time. It's like a radical. Right? So it was called the benzoyl radical. They thought up that name at that time, and the idea of using the suffix -yl, to denote a radical. So if you denote the benzoyl radical by Bz -- you see that you started, the oil of bitter almonds is BzH, and then the acid is BzOH, and the acid chloride is BzCl and Br and I and NH2 and two Bz's, together with S, at the end. Okay? So a radical can be the base of more than just an acid. Right? Lavoisier had the idea that you react it with oxygen and you get an acid. But here you can react it with all sorts of different things and get different compounds. But the base is still there, the benzoyl radical. So this gave rise to the idea of organic dualism. Remember, we had this dualism last time. There were positive things and negative things and they could associate and trade partners. But maybe the difference in organic chemistry is that you have radicals, things that are plus or minus, but they're more complicated than just single atoms. There are combinations of elements that function in organic chemistry; and that's what makes it different from inorganic, according to this theory. So then the idea is to find all these organic radicals, that make organic chemistry special.



So during the 1830s these compound radicals were discovered everywhere. For example, in Germany, Liebig found acetyl. And Bunsen, in Heidelberg, found cacodyl; which is named because it smells so awful. Right? Or Berzelius in Sweden found ethyl. Or Piria in Italy got salicyl. And Dumas got a whole bunch of them, in Paris: methyl, cinnamyl, cetyl, ethylene. All these radicals were discovered. So the thought was this is the way to organize organic chemistry. And that theory, the dualism applied to organic theory, the radical theory, survives today in our nomenclature. Right? So, for example, we talk about ethyl chloride. That's not one word, it's two words, with a space. Right? And the reason is it's dualistic. It's a positive ethyl and a minus chlorine. Right? So it's two things that have come together.



Okay, now the -yl part, that Wöhler and Liebig thought up, comes from a Greek word, üleh, which means wood or matter. So it's the substance of stuff. Okay, so now ether was something that had been known for a long time, and it came from a Greek root which means to shine. So in the 1700s it was -- so it had been applied to the sky. It was transferred from the idea of shining, to the idea of the clear sky, and from that to a colorless liquid. So when they distilled something out of alcohol that had been treated with acid, and they got this clear stuff that was clear as the sky, they called it ether. Right? Which we call diethyl ether nowadays. So that's where "ether" came from. So hence eth-yl was the matter that appears in ether. So it was like two benzoyl radicals with sulfur. Right? You can have two ethyl radicals with oxygen. Right? And that was then eth-yl -- right? -- the material of ether. Okay?



How about methyl? Well meth comes from the Greek word meaning wine or spirit. Right? And the -yl, that same root, but a different meaning this time. Remember, it can mean matter -- that was the one before -- or it can mean wood, and in this case it means wood. So what does it mean, meth-yl, if the -yl means not the substance of but means wood?



Student: From wood.



Professor Michael McBride: Pardon me?



Student: From wood.



Professor Michael McBride: From wood. What from wood?



Student: Wine.



Professor Michael McBride: The spirits, from wood. So you've heard methanol called wood alcohol -- right? -- because you get it from distilling it. But the first word was methylene, and ene is the Greek feminine patronymic; it means "the daughter of." And ene, ine, one, all of those are like that; like Persephone, or Antigone means the "one who goes against her parents," and so on. So ene was the Greek -- so what did meth-yl-ene mean? It meant the daughter of wood spirits. So the theory was that wood spirits, that is, wood alcohol, was a combination of methylene plus water. So methylene thus, if wood alcohol is CH3OH, then methylene was CH2; which then you add water and you get wood alcohol. So that's where the name methylene came from. Right?



So then in 1840 -- that was 1835, so five years later -- they decided that they needed the radical CH3. So they named it methyl, from methylene. Right? And then ethylene came; that name came in 1852, because it was related to ethyl, which had already been named, the same way that methylene was related to methyl. Okay? So that's where these names came from. They all have their root in the radical theory. Now, that's C1 and C2, methyl and ethyl. How about C3 and C4, do you know what they're named? Roots? How about C3, do you know what C 3 alcohol is?



Students: Propyl.



Professor Michael McBride: Propyl. And C4?



Students: Butyl.



Professor Michael McBride: Butyl. So where do those names come from? Okay, C3H7 is propyl, and by the same reasoning as before, C3H6 is propylene; or propene, sometimes it shortened. Okay? And butyl and butylene, or butene. Okay, now butylene comes -- the C4 acid is butyric acid, which had already been named, because it's the stuff that makes rancid butter smell bad. Okay? So people worked it up and found butyric acid; so that's where butyl comes from. But how about propyl? It has a very much more interesting origin. Okay, so protos means first, and pion means fat. So propion was the first fat. In what sense? Well these carboxylic acids came from fats, right? And they were called fatty acids. And they behaved like fats, they dissolved in organic solvents. Right? But the very -- the ones with the fewest number of carbons, with just one carbon -- that's formic acid that you get by destructive distillation of ants, or acetic acid, from vinegar -- those are miscible with water. But the C3 is the first one that's not freely miscible with water; it behaves more like a fat. Right? So the first fatty acid is propionic acid. So propyl, propylene are the C3s. Okay, after that you get into numbers, the roots for numbers.



So Jean-Baptiste André Dumas was the successor of Gay-Lussac as the spokesman for French chemistry. You see, he was born on the 14th of July, a reasonable date for somebody to lead French chemistry, in 1800. So for eighty-four years -- well, for part of eighty-four years, he was the leader of French chemistry. There he is with some decoration on him. I'm not sure what it is. I'd like to find out. So he was the Post-Napoleonic guardian of the French tradition of chemistry. The French had what most people regard as a terrible system, which is they had chairs, you know, for professors. But you could have more than one chair. So a single individual could tie up three different appointments, and you didn't have as many people able to exercise their ingenuity in developing chemistry. So he had the chair at the Sorbonne; also the École Polytechnique and the École de Médicine. And he was a persistent opponent of Liebig and Berzelius. But in 1837, after this radical stuff got going -- and he had discovered four radicals himself, of which he was quite pleased -- he and Liebig happened to meet during a speaking tour in England, and they got in conversation, and Dumas decided, on the basis of that meeting -- though not Liebig -- that they were now great friends and could collaborate from here on in.



So he wrote this long thing, in flowery French, in 1837, "A Note on the Present State of Organic Chemistry." So this is five years after the Radical Theory began. He said: "Sixty years have hardly passed since the ever memorable time when this same assembly" -- he was speaking to the Paris Academy -- "heard the first discussions of the fertile chemical doctrine which we owe to the genius of Lavoisier. This short span of time has sufficed to examine fully the most delicate questions of inorganic chemistry, and anyone can easily convince himself that this branch of our knowledge possesses almost everything that it can with the methods of observation available." So check off inorganic chemistry, we've got that now. Okay? "There barely remain a few cracks here and there to fill in." So this is the persistent myopia of leaders of science. As I mentioned before, when we were speaking of Lavoisier, this certainly persists 'til today. So Dumas goes on: "In a word, how with the help of the laws of inorganic chemistry can" -- incidentally, this is all one sentence, of course. This is "in a word", right? So he really liked speaking, right?; flowery language. "How with the help of the laws of inorganic chemistry can one explain and classify such varied substances as one obtains from organic bodies, and which nearly always are formed only of carbon, hydrogen, and oxygen, to which elements nitrogen is sometimes joined?" So if it's all in the analysis of what atoms are there, how can you have so many different things? Right? "This was the great and beautiful question of natural philosophy, a question well designed to excite the highest degree of competition among chemists." Name two, right?



"For once resolved the most beautiful triumphs were promised to science. The mysteries of plants, the mysteries of animal life would be unveiled before our eyes; we would seize the key to all the changes of matter, so sudden, so swift, so singular, that occur in animals and plants; more importantly we would find means of duplicating them in our laboratories." (This would make a good research proposal, right?) "Well, we are not afraid to say it, and it is not an assertion which we make lightly: this great and beautiful question is today answered; it only remains to follow through on all the consequences which the solution entails. In fact to produce with three or four elements such varied combinations, more varied perhaps than those which make up the whole inorganic kingdom, nature has chosen a path as simple as it was unexpected; for with elements she has made compounds which behave in all their properties like elements themselves." (That is, radicals -- right? -- that persist through these reactions.) "And this, we are convinced, is the entire secret of organic chemistry."



Great and beautiful questions to answer. It only remains to follow through the consequences; compounds which behave like elements. "Thus organic chemistry possesses in its own elements, which sometimes play the role of chlorine or oxygen" -- what does he mean, play the role of chlorine and oxygen? How come radicals sometimes play the role of chlorine and oxygen? What would that role be?



Student: [inaudible]



Professor Michael McBride: Pardon me?



Student: Oxidation.



Professor Michael McBride: No. What roles do things play in this theory that he's using? Lucas?



Student: Plus and minus.



Professor Michael McBride: Plus and minus. So sometimes they're like chlorine or oxygen. What does that mean?



Students: Negative.



Professor Michael McBride: Sometimes they're negative, right? "And sometimes, on the contrary, they play the role of metals." Sometimes they're positive.



"Cyanogen, amide, benzoyl, the radicals of ammonia, of aliphatics, of alcohol, and analogous substances, these are the true elements with which organic chemistry operates. To discover these radicals, to study them, to characterize them, has been our daily study for ten years." (Okay? So they're plus and they're minus.) "Sometimes, none the less, our opinions have appeared to differ, and then, with each of us drawn on by the heat of our battle with nature, there arose between us discussions whose liveliness we both regret. Actually when we were able to discuss questions which separated us in several friendly meetings, we soon realized that we were in agreement on the principles…. We then understood that united we could undertake a task before which either of us in isolation would have recoiled…. We will analyze every organic substance... to establish reliably what sort of radical it refers to…."



So that is what everybody should do. Then you'll know all about organic chemistry.



"Each of us has, in fact, opened his laboratory to all young men who were motivated by true love of science; they have seen all, understood all. We have worked under their eyes, and have had them work under ours, in such a way that we are surrounded by young rivals, who are the hope of science, and whose work will be added to ours and mingle with ours, for it will have been conceived in the same spirit and carried out by the same method."



Right? So if everybody does what we say they should do, then we got it. Okay? "This is not an effort conceived for personal gain or in the interest of narrow vanity." Far be it from us. "No, and in a collaboration which is perhaps unheard of in the history of science, this is an undertaking in which we hope to interest every chemist in Europe." So everyone should work on this. So we'll go to the funding agencies and tell them this is the only kind of research you should fund; forget these other guys out there. This is true love of science, right? And conceived in the same spirit and carried out by the same methods. So this is megalomania, and doesn't show much imagination. Now, but there was a problem with dualism. So, for example, suppose you have benzoyl chloride, which remember was by reacting benzaldehyde or benzoyl hydride with chlorine, and you get benzoyl chloride and HCl as the other product. Now HCl is quite clear, it's H+ and Cl-. What's benzoyl chloride? It's obviously benzoyl+ and Cl-. And what problem does that create? What, Russell?



Student: The benzoyl is minus before, but hydrogen --



Professor Michael McBride: Ah ha. But how do you get benzoyl hydride, plus, plus? Right? So there's something weird going on. So this is a problem. In the 1840s and 1850s, the French discovered a competing theory -- or invented, I should say -- a competing theory called the substitution theory, or the type theory, or the unitary theory, as opposed to the dualistic theory, the plus, minus idea. So these began to compete with one another. And it started at a ball in the Tuileries Palace in 1830. This picture is from 40 years later, or 37 years later. But what happened is they got the ball started, all the people came dressed up fancy, and they began to cough and choke because the room was filled with some noxious gas. And when they discovered it came from the candles, they asked Dumas to look into it. And he identified the culprit as HCl, because it turned out that the candles were very white. The wax had been bleached with chlorine, and when they were burned HCl was given off. So the question is, what is it that holds chlorine? How did this fat, in the candles, fix chlorine gas? Well you're in a position to understand that now. We can think about mechanisms; in fact, two ways that the hydrocarbon could fix chlorine. Now, suppose we try -- there's chlorine that is being used to bleach the stuff. Let's try for a HUMO/LUMO approach. What makes chlorine reactive? High HOMO or low LUMO? What's unusual about chlorine?



Student: Low LUMO.



Professor Michael McBride: Pardon me?



Student: Low LUMO.



Professor Michael McBride: Why do you say so Claire? What is the low LUMO?



Student: It's the Cl-Cl bond.



Professor Michael McBride: Right, the σ* of Cl-Cl, which is low because chlorine has a high nuclear charge. Okay, now we need a HOMO to react with it. Now, one of the more interesting hydrocarbons, in this regard, is one that had been known already, for fourty years, to react with chlorine. And it was because of that reaction it was called the "olefiant gas," and was by this time known to be C2H4. Right? So we would write it with a double bond. "Olefiant," because ole, oil, and fiant, to make; so it's the stuff that makes oil. Right? And we'll see the reaction that makes oil here, its reaction with chlorine. So what makes the olefiant gas reactive? Kevin?



Student: Poor overlap.



Professor Michael McBride: Right, so poor overlap makes a high HOMO. And remember the name of it?



Student: π.



Professor Michael McBride: π. So the π electrons because of poor overlap are a high HOMO. So we can use those electrons to mix with the low LUMO; and again, one of these make-and-break situations. And chloride leaves and you get this thing, which has a positive charge. Now that thing itself is reactive. What do you make -- where is a low LUMO in this one? Pardon me?



Student: On the positive charge.



Professor Michael McBride: Speak up please.



Student: On the positive charge.



Professor Michael McBride: Where the positive charge is. There's a vacant orbital on carbon, an atomic orbital of carbon that's not shared. So there's a low LUMO. Where's a high HOMO? Have you got that too Virginia?



Student: On Cl.



Professor Michael McBride: Chlorine has unshared pairs, right. Bingo! So, in fact, both those things happen at once. It's not that one happens and then the other. Both those things happen at once. And you can see it by looking at molecular orbitals. So there's the HOMO of the ethylene or olefiant gas, and the σ* LUMO. So those things mix. The blue orbitals overlap and mix, shift electrons toward the other one; the chloride breaks away. But at the same time the HOMO of the chlorine mixes with the LUMO of the ethylene. So you're making two bonds at once, two pairs of electrons. So you make that three-membered ring, with two new bonds, and the chloride, as we said, breaks away. Okay, so we've got that substance now. And now it itself is reactive. Can you see what would be reactive about that cation intermediate? What are you probably looking for, a LUMO or a HOMO?



Student: A LUMO.



Professor Michael McBride: A LUMO. Angela, do you have an idea of what could be a LUMO here?



Student: The chlorine has a positive charge.



Professor Michael McBride: It's true that the chlorine has a positive charge. Does it have a vacant orbital, an unoccupied molecular orbital?



Student: No it doesn't.



Professor Michael McBride: No, it turns out it's got two unshared pairs. So it doesn't have any -- it's not like the carbon plus was. But the plus will make orbitals low in energy. So what's a vacant orbital of this thing? All it's got is σ bonds. But what makes -- suppose all you have in your molecule is σ bonds, but you want to have an unusually low energy vacant orbital. Right? The plus charge will help. But what orbital will you have?



Student: σ*.



Professor Michael McBride: σ*. Now you got two choices. You got carbon-carbon or carbon-chlorine. Which one's more likely to be low energy?



Student: Carbon-chlorine.



Professor Michael McBride: Why?



Student: Because the chlorine --



Professor Michael McBride: Say it --



Student: Chlorine has a high effective nuclear charge.



Professor Michael McBride: Right, chlorine has a high nuclear charge. So a σ* carbon-chlorine, would it be big on carbon or big on chlorine, Sam?



Student: Big on carbon.



Professor Michael McBride: Big on carbon, because the bonding orbital was big on chlorine. Right? This is the kind of stuff we're talking about. So σ*. And there it is. Okay, there's a localized σ*. Big on carbon, the black one; small on chlorine; and antibonding between them. Now what are you going to -- there's the low LUMO. What do you have for a high HOMO, to react with it? You have to think back to what's been happening. Sherwin?



Student: The chlorine one.



Professor Michael McBride: The chloride that you had at the beginning, that you lost in the first step. Okay, so we bring it over here. So it'll have good overlap. It comes up, makes a new bond; that make and break. And you get that. And that was the reaction in 1795 that resulted in ethylene being called the olefiant gas. Because this is the oil that was made, by reacting it with chlorine. So that was already a very old reaction at this time; the "oil of Dutch chemists", because it was four Dutch chemists who reported that oil. Okay, so there's one way that you can fix chlorine, make it part of a hydrocarbon molecule. It's addition to an alkene. So if the hydrocarbons are unsaturated, if they have some double-bonds, then they'll react with chlorine to fix the chlorine. So that's one possibility. But how about if you don't have a double bond? How about if you have methane? What's the problem now, in doing an analogous reaction? Sherwin?



Student: You don't have the π in there.



Professor Michael McBride: Pardon me?



Student: We don't have the π in there.



Professor Michael McBride: You don't have a π. You don't have a low LUMO. So you can't do a HOMO/LUMO reaction. Right? These are our model, saturated alkanes, like methane, our model of things that aren't unusually high or unusually low. So you have to have another trick. And here's the trick; that the chlorine-chlorine bond is weak. It's only 58 kilocalories per mole. And one of the reasons for that is that the chlorine has so many unshared pairs. So you mix -- if you were trying to form a π orbital in chlorine, you have two electrons in the p orbital here, two electrons in the p orbital here. They overlap. Is that going to be bonding, if you mix these two orbitals? You'll obviously mix them. When you mix two orbitals you get a lower one and a higher one. Will it be bonding? This one will be bonding but this one is anti-bonding. Everybody with me on this? Now, so Kate, what would you say? Is it going to be net favorable or unfavorable? Two electrons went down in energy, two electrons went up in energy. But the ones that went up, went up a little more than the down went down. So that's unfavorable. So having so many unshared pairs weakens the single bond. So chlorine has a weak bond. Now still it's worth 58 kilocalories per mole, which is plenty strong. It doesn't just break. You got to do something to help it to break. And what you can do is -- I've made it in this weird color, which is hard to see. Why? Anybody know?



Student: The color of chlorine.



Professor Michael McBride: Pardon me?



Student: It's the color of chlorine.



Professor Michael McBride: That's the color of chlorine. It absorbs visible light. Now how does it take on energy, when it absorbs visible light? Where does the energy go, in the molecule? Does anybody know? Dana?



Student: Electrons are promoted.



Professor Michael McBride: Can't hear very well.



Student: Electrons get promoted to higher --



Professor Michael McBride: Electrons go from orbitals to higher orbitals. So you can put -- and the next higher orbital is σ*. And what happens if you take an electron and put it in σ*? It breaks the bond. Right? That's what happened up at the top, when the chlorine broke, right? You put electrons into σ*. So you can do it with light, as well as with some HOMO attacking. Okay, so we have -- in fact, I said LUMO when I was talking about ethylene; I meant HOMO up above, I think, some three or four minutes ago. Okay, so the bond breaks. But it doesn't break into ions. It breaks one electron going each way, because that's easier. Okay? And notice that we draw curved arrows for that too, but you draw arrows with a single barb rather than a double barb, when it's just one electron rather than a pair of electrons that's executing the motion we're talking about. Okay, so now we have two chlorine atoms. And now we can do the trick with the chlorine atom, because we have this SOMO, and it can mix with the C-H bond to make a new bond; that is, one electron in the C-H bond now goes each way. One goes to complete the pair, to make HCl, and the other one is left on carbon. Right? So it's very much like the reaction above, but it's single electrons that are doing the moving instead of electron pairs. And the nice thing about this is you still have a radical. It must be so. If you start with something with an odd number of electrons, and react it with something with an even number of electrons, you must be left, at the end, with an odd number of electrons. Right? So CH3 is such a radical, and it can react with something, to break another bond, and it reacts with the weakest bond, chlorine. Right? So now you have methyl chloride -- you've incorporated chlorine into the alkane -- and you have a chlorine atom. Why is it neat, that you have a chlorine atom? What's great about that?



Student: [inaudible]



Professor Michael McBride: Dana, what did you say?



Student: That was what you started with.



Professor Michael McBride: That's what you needed at the beginning. That's why you used light, in order to get that. But you don't need any more light now. Right? That can go back and start over again. So it's a "chain" reaction. It's called a free-radical chain reaction. And so you can get lots of products from just one initial photon of light, that started this chain along. So these are two completely different ways. The first, the top is an "addition" of chlorine to an alkene, and the bottom is called free-radical "substitution" of chlorine for hydrogen, and involves SOMOs, rather than HOMOs and LUMOs. Chris?



Student: If you have a second chlorine radical from the first breaking…



Professor Michael McBride: Yeah.



Student: …why does it break a second chlorine molecule, rather than using the other --



Professor Michael McBride: Because they have to find one another. Right? They've gone off -- it'll take forever before they by chance encounter one another in solution. They'll react with many molecules. If you try to generate too many chlorine radicals, so the concentration gets high, then their concentration will drop again, or not get so high, because they find one another and combine. But as long as they're rare, they can survive. You know what the license plate of New Hampshire says on it? "Live free or die." Okay? I use that joke later on. Okay, 1830s to 1850s, we have this substitution or type or unitary theory. It doesn't involve the plus/minus stuff. Max?



Student: Is that kind of like how CFCs work?



Professor Michael McBride: Is it kind of like what?



Student: How CFCs destroy the ozone.



Professor Michael McBride: I couldn't hear clearly.



Student: Is that how CFCs break down the ozone?



Professor Michael McBride: Yeah, they involve -- that's a free-radical chain reaction, the ozone reduction. Yeah. We'll talk about that a little bit later, I hope. Okay, so there's more trouble for radicals, from Dumas in 1839. And that is that they had this -- remember, acetyl radical had been discovered by Liebig. So there was this great element that would survive from reaction to reaction. But here was a reaction with chlorine, of this kind that Dumas had been studying, where you start with acetic acid, acetyl OH, react it with chlorine, and you get a chloroacetyl. So the element has been changed. It's been transmuted. It doesn't go unchanged from reaction to reaction. Right? And, in fact, it goes even further. It can react again to give dichloroacetyl or trichloroacetyl. All the hydrogens can be substituted, as we now know by the kind of mechanism, the SOMO mechanism, we just studied; free-radical chlorination, chain reaction. So hydrogen can be substituted by an equivalent amount of halogen, or oxygen, right? But all these things you get, when you change a radical into something else, when you transmute it, have similar properties. All of these acids -- acetic acid, chloroacetic, dichloroacetic, trichloroacetic, are all acids; they all taste sharp and so on. Right? So they're similar. So they don't change the type. That's where this idea of Type theory came on. You get the same type of molecule, even after you have a substitution.



So by 1853, four types were recognized as prevalent. One was water, another was hydrogen, another was hydrochloric acid, and another was ammonia. So, for example, you could have these structures; and this drawing with a curly bracket like this was the notation used by the people who did Type theory. So you have water, hydrochloric acid, hydrogen and ammonia. And you could exchange, make exchanges, for the hydrogen. Right? Substitution. So, for example with ammonia, you can substitute ethyl for hydrogen and you could get ethylamine, diethylmanine, or triethylamine. But these were all basic. Why would we say they're basic, in the sense of acid base? Why do we say they're basic? They react with acids, why? Sherwin?



Student: The unshared pair.



Professor Michael McBride: Yeah, they all have the unshared pair on nitrogen, we would say. But they said they're just the same Type of molecule. Okay? Or you could have ethyl alcohol, which is of the water type. Or the potassium salt thereof, which they would say is the potassium-ethyl analog of water. Or of the HCl, you could have ethyl iodide, where you exchanged hydrogen with ethyl, and exchanged chloride with iodine. Okay, but in fact these two things could react with one another to give this ether; which is another thing that's still like water, clear liquid and so on. Okay? Fairly unreactive. So this particular reaction was named for the work, in 1850, of Williamson in England. So it's called the Williamson Ether Synthesis. And we'll talk about that again later. But it's quite an old reaction.



So how about the theory of these types? So notice it's unitary, not dualistic. They're just things that are holistic, right? Not plus/minus. Dumas said that molecules are like planetary systems, like the sun and its planets, "held together by a force resembling gravitation, but acting in accord with much more complicated laws." He didn't think it was gravity, but it was some force holds this assemblage together. And Williamson, Alexander William Williamson, who we just mentioned making ether, said this is something new. And notice it's because it's a very young guy that has a new idea, not like these imagination-starved older people. At this time Dumas was 40-years-old. He's 4/7ths as old as I am, right? So he was really getting over the hill. "A formula" -- the young guy says -- "may be used as an actual image of what we rationally suppose to be the arrangement of constituent atoms." This is entirely new. Formulas, at least since Dalton, were only what the elements were and what ratios they're in; not how they're arranged. Right? But he said, "We can think that they're like an orrery, which is an image of what we conclude to be the arrangement of our planetary system." Do you know what an orrery is? It's a thing like this, you know, where you have a mechanical model of the solar system and you turn a crank and the moon goes around the earth, the earth goes around the sun, and a bunch of moons go around Saturn and so on. Have you seen these devices? Okay, that was very popular. There was a show of Joseph Wright of Derby here at Yale last summer.



So butyl bromide, you remember, is a residue of radical dualism, right?; plus-butyl, minus-bromide. But there's another name for butyl bromide. Do you know what it is? You know the other name for butyl bromide? We haven't talked about systematic nomenclature yet, so you probably don't. But it's also called bromobutane. But bromobutane is not two words, it's one word. Right? And that's a relic of the unitary theory, the substitution theory, that it's butane in which a hydrogen has been replaced by bromine. Okay, so Berzelius, in 1838 when these things came along, said: "By reacting chlorine with ordinary ether [Dumas] produced a very interesting compound which he reckoned, according to the theory of substitutions, to be an ether in which 4 atoms of chlorine replaced 4 atoms of hydrogen." Right? So Dumas says that in these types you can replace hydrogen by chlorine. What would Berzelius think about that? What kind of theory is he advocating? Remember, we talked last time about Berzelius. He was the originator of dualism, plus/minus. What would he think of replacing hydrogen by chlorine? Lucas?



Student: It's impossible. If chlorine is minus, hydrogen is plus.



Professor Michael McBride: Right. So here's what he said: "An element as eminently electronegative as chlorine would never be able to enter into an organic radical. This idea is contrary to the first principles of chemistry." Okay? And in that same paper, in 1838, he talked about tartrate losing an atom of water -- but you mean it's a molecule -- at 190°. But the interesting thing about that, it's not transformation, but that this is the first time the letter R was used, to talk about a generic radical; "R" stands for radical. Right? So we still use that. So, so much of what we do nowadays derives from this period. Okay? So the principal journal at that time, in chemistry, was the Annalen der Chemie und Pharmacie, which was originally the Annalen der Pharmacie; but Chemie was added. And you see who put it out. It says: "With the collaboration of Dumas in Paris and Graham in London." And Graham is one of the guys out in front of the building here. Edited by Wöhler and Liebig; but actually it was Liebig, he was the one. And later the journal came to be called Liebig's Annalen. Right? So he was the one in charge, but he'd added these other people, just sort of for show.



Okay, so in 1840, there were a series of papers published. The first of them was by his so-called collaborator, Dumas, On the Law of Substitution and the Theory of Types. This was a 40-page paper. And it begins with this question, number one there: "Can one substitute the elements that play their role in any simple or compound substance equivalent for equivalent?" So can you make a new atom take the role of an old atom? And you won't be surprised that in 40 pages he concludes the answer is yes. But immediately following this is a note from the editor, which says "Remarks on the Previous Paper." And it's by J.L. at the bottom, Justus Liebig. And he says, begins: "I am a far cry from sharing the ideas that Monsieur Dumas has linked to the so-called laws of the substitution theory." And then there's another paper, by another Frenchman, Pelouze, on "The Substitution Law of Monsieur Dumas." And after that lengthy paper, there's a letter, "On the Law of Substitution and the Theory of Types," with a footnote that says it was a letter to Justus Liebig. And you notice it's dated Paris. And it's the only paper that's in French, in his particular issue. And there's another page. It's got some curious formulas in them that have a lot of chlorine. We'll come back to that in just a second. And it's by a chemist that no one had heard of before called S.C.H. Windler -- this letter to Liebig. And then the next paper is also "On the Reaction of Chlorine with the Chlorides of Ethanol and Methanol, and Several Points of the Ether Theory," by Regnault, who was a French chemist. Now this is that letter, translated.



"Paris, 1 March, 1840. Monsieur! I am eager to communicate to you one of the most striking facts of organic chemistry. I have confirmed the substitution theory in an extremely remarkable and completely unexpected manner. Only now can one appreciate the great value of this theory and foresee the immense discoveries that it promises to reveal."



So it starts -- it's got a complicated description of all the -- light that was used sometimes and the various distillations and crystallizations, and the crystals are described. But he started with manganese acetate, which had this formula with the acetyl radical in it, and was able to chlorinate -- to substitute the hydrogens with chlorine. So there's already a validation of the substitution theory. But it went further. A subsequent reaction exchanged the O in manganese oxide with chlorine, and a further reaction replaced the manganese itself with chlorine. And then the final reaction, the pièce de résistance, replaced the carbon with chlorine and the oxygen with chlorine, but still it preserved its type. Right? So it was the same kind of substance still, even though it was entirely chlorine.



"For all I know, in the decolorizing" (that is bleaching) "action of chlorine, hydrogen is replaced by chlorine, and the cloth, which is now being bleached in England, preserves its type according to the substitution laws. I believe, however, that atom-for-atom substitution of carbon by chlorine is my own discovery. I hope you will take note of this in your journal and be assured of my sincerest regards, etc. S.C.H. Windler."



And there's a footnote, which says: "I have learned that there is already in the London shops a cloth of chlorine thread, which is very much sought after and preferred above all others for night caps, underwear, etc." Now this is, you'll not be surprised a pun, because the name is pronounced schvindler or swindler. So where do you think this came from? Liebig didn't have enough sense of humor to do such a thing. Who was the joker?



Student: Wöhler.



Professor Michael McBride: Wöhler. So Wöhler had sent this in a letter to Berzelius, and Berzelius thought it was so fun that he forwarded it to Liebig. And to Wöhler's consternation, Liebig published it. Right? [Laughter] Which didn't make Dumas any happier. Right? So this was a -- Liebig at least thought it was funny, I guess.



Okay, so in 1849 Kolbe prepared the free methyl radical, the actual element, which had never been prepared before, CH3. He did it by electrolysis. So when you electrolyze acetic acid it turns out you can get hydrogen, hydrogen gas, and CO2, and what analyzed for CH3. Right? So he had prepared the actual radical. Now, of course, it was discovered ten years later that he hadn't prepared CH3. It had that analysis, but actually it was C2H6; it was the dimer of CH3. Okay, but at the time it was thought to be justification of the radical theory. So these two theories were competing with one another. And ironically, both the theories were supported by reactions that actually did involve radicals. So the oxygenation of benzaldehyde, the first reaction that generated benzoyl, did in fact involve benzoyl. It was a SOMO reaction in which a hydrogen atom was abstracted. And the chlorination, as we've already seen, of a hydrocarbon like methane, involves pulling a hydrogen atom off. And that electrolysis did indeed generate the methyl radicals, but they were so reactive, they found one another and dimerized to C2H6. So it's just sort of curious that all these reactions did indeed involve free radicals, but no one was truly aware of it. So next time we'll see what resolved all this.



[end of transcript]

Lecture 23
Valence Theory and Constitutional Structure (1858)
Play Video
Valence Theory and Constitutional Structure (1858)


Youthful chemists Couper and Kekulé replaced radical and type theories with a new approach involving atomic valence and molecular structure, and based on the tetravalence and self-linking of carbon. Valence structures offered the first explanation for isomerism, and led to the invention of nomenclature, notation, and molecular models closely related to those in use today.




Transcript



October 29, 2008



Professor Michael McBride: So we're going through the development of the theory and practice of organic chemistry, generation by generation. We started with Lavoisier, and then Berzelius and Gay-Lussac and Davy. And last time we did Dumas and Liebig and Wöhler. And now we're onto the next generation, people that were born around 1830. And the new developments are going to be valence and benzene. Remember, in this list, things that are in boldface are experimental and things that are theoretical are in normal face. So valence and benzene is experimental, but what we're interested in is the theory of benzene here. And the people who are associated with this are August Kekulé -- and you've all heard of Kekulé structures and Kekulé benzene, and he's memorialized on the front of the building here. He was born in 1829. But few of you probably have heard of Archibald Couper, who is not on the front of the building, and was born two years later in 1831. This is Couper as a student. He lived to a reasonable age, sixty-one-years-old, but everything he did, he did early. He was sort of a dilettante student. He was a native of Scotland and studied in Edinburgh, but also went to Berlin often. And he studied classics and metaphysics, logic, moral philosophy; went to a lot of concerts. But while in Berlin he got fascinated by chemistry.



So in 1855 he became a student chemist. And in 1856 he went to Paris to work with Wurtz, who was the other leading French chemist, and worked indeed on salicylic acid; as you have done. In 1858, while in Wurtz's lab, he got fired, for reasons that you'll see, and he went back to Scotland and tried to make a career there. But soon he had a mental breakdown, and the rest of his life he didn't do anything. Okay? But what he did in 1858 is really something. Okay? So he didn't get a stone on the front of our building. But on this building, which is "To Let," or at least was a few years ago, in Kirkintilloch, which is about eight miles northeast of Glasgow, in Scotland, there's a stone. And it says: "This plaque marks the birthplace of Archibald Scott Couper whose brilliant pioneering contributions to chemical theory have won for him international renown, and whose genius, stifled by an early illness, was denied the opportunity of consummation." When he was recovering from his first breakdown, he went fishing on this Endrick Water, north of Glasgow, and got sunstroke, which did him in for the rest of his life. And that's where he lived. His mother nursed him for his final 33 years in this house in Kirkintilloch.



So this is the paper. He tried to publish it in French. He submitted it to his boss, Wurtz, in the hopes that he would put it in the French journal, but Wurtz thought it was not such a -- it was a little rash as a paper, and put it in his desk to cool off for awhile. And in the meantime Kekulé published, which irritated Couper no little, and he told Wurtz what he thought about it, which was why he got sacked. But he also published it in English, and later it was published in French. But so this is the English version, not the first version he wrote; by Archibald Scott Couper. And the first sentence already is fascinating. "The end of chemistry" (that is, the goal of chemistry) "is its theory." Not many people would have said that. They would have had more practical motivations for chemistry. But Couper was a big thinker. He said:



"The guide in chemical research is a theory. It is therefore of the greatest importance to ascertain whether the theories at present adopted by chemists are adequate to the explanation of chemical phenomena, or at least, based upon the true principles which ought to regulate scientific research."



So what were the theories that were prevalent at this time, by 1850? Name one.



Students: Dualism.



Professor Michael McBride: Shai?



Student: Dualism.



Professor Michael McBride: Dualism; okay, the radical theory. That was prevalent in Germany. How about elsewhere, like France?



Student: Substitution.



Professor Michael McBride: The substitution, or type, or unitary theory. Okay, so those were the big theories.



"Among those which have been lately developed, there is one, on account of its apparent numerous merits, which particularly claims investigation, and respecting which we deem that it would not be unprofitable were either new proofs of its scientific value furnished or, on the contrary, should considerations be adduced establishing not only its inadequacy to the explanation, but its ultimate detriment to the progress of science. I allude to the system of types as advocated by Gerhardt."



So Gerhardt and Laurent, who were slightly second-level French chemists, were the ones who really came up with the idea of the type theory. But Dumas adopted and advocated it, and was the leader. So this was the theory that was going on in the lab where he was present, in Paris. So page 106 of this paper, on the French type theory, he said:



"Should the principle which is therein adopted be applied to common events of life, it will be found that it is simply absurd. Suppose that some one were to systematize the formation of letters into words that formed the contents of a book. Were he to begin by saying that he had discovered a certain word which would serve as a type" (remember there were four types; water, ammonia, HCl and hydrogen) "and from which by substitution and double decomposition all the others are to be derived, that he by this means not only could form new words, but new books, and books almost ad infinitum; he would state certainly an empirical truth."



That is, you can take some word and substitute letters, or pairs of letters, or sets of letters, for other letters, and get any word you want. Okay? "At the same time, however, this method would, judged by the light of common sense, be an absurdity. But a principle which common sense brands with absurdity is philosophically false and a scientific blunder." So this was not likely to make him popular in Paris. So here's what he was doing, actually, right? The Emperor's new clothes. He's a new guy on the block, right? He's only been a chemist for two years, only begun studying chemistry two years before, and says what everybody's doing is nonsense, in France. Well how about in Germany? That's where he was before, in Berlin. "I can only remark that it is not merely an unprofitable figure of language, but is injurious to science, inasmuch as it tends to arrest scientific inquiry by adopting the notion that these quasi elements…" What does he mean by the quasi elements; what are they?



Student: Compound radicals.



Professor Michael McBride: Compound radicals, right? "…contain some unknown and ultimate power, which it is impossible to explain." That is, okay, so you got radicals. What holds the radicals together? Right? "It stifles inquiry at the very point where an explanation is demanded, by putting the seal of elements, of ultimate powers, on bodies which are known to be anything but this." They're definitely not elements, because you can burn them, of course, and get water and CO2 and so on. Okay? So this was absolutely true; on both counts, both the French and the German theory. But it was stated pretty undiplomatically. Right? And especially for someone who'd only been studying chemistry for two years. Maybe it took someone who was only studying chemistry two years to see how stupid it was. Right? But at any rate, he got in trouble because of this. What did he offer as an alternative? To look at the properties of the elements themselves, not of radicals, not of these planetary systems that were the types. "Science demands the strict adherence to a principle in direct contradiction to this view. That first principle, without which research cannot advance a step, dare not be ignored; namely, that a whole is simply the derivative of its parts." So you have to understand the elements, and then you can understand assemblies of elements from that. That's not obviously true. In fact, I think you could make a pretty good argument that as you get things more complicated, they're not simply a sum of their parts. But that's what he said.



"As a consequence of this, it follows that it is absolutely necessary to scientific unity and research to consider these bodies as entirely derivative, and as containing no secret ultimate power whatever, and that the properties which these so-called quasi elements possess are a direct consequence of the properties of the individual elements of which they are made up."



So study the elements, not the radicals. And he began by saying he'll focus on carbon; which is not surprising for an organic chemist. "In applying this method, I propose at present to consider the single element, carbon. This body is found to have two highly distinguishing characteristics. (1) It combines with equal numbers of hydrogen, chlorine, oxygen, sulfur, etc." Would you agree with that? How could he say so? It's because he wasn't dealing with the right atomic weights, right? Okay? "It enters into chemical union with itself." So those are the two things: that it'll unite with itself, not just with other things, the way dualism would like you to do; and it combines with -- it has a certain combining power. "These two properties, in my opinion, explain all that is characteristic of organic chemistry."



So 1858 marks a new frontier in organic chemistry; a completely new revelation: The tetravalence and self-linking of carbon. That's points one and two, right? That it combines with a certain number of other things and it can link to itself. Right? Now here he shows the formula of methane, of CO2, and of carbon tetrachloride. But those are the atomic weights he's using, in order to get these formulas. Notice he says: "Here the whole four of hydrogen are not bound by a mutual affinity." You know that they're not all collected together because you can substitute them one by one; one chlorine, two chlorines, three chlorines. "For each element of hydrogen can be substituted for one of chlorine in regular series, beginning with the first and ending with the last. The atoms of oxygen are, on the contrary, united in pairs." What does he mean by that, that the oxygens are united in pairs? Why are they united in pairs in his formulas, drawn by those curly brackets, like they had in type theory? Yes Angela?



Student: Probably one oxygen, because the -- they were probably one oxygen because his weight is half of it.



Professor Michael McBride: Ah ha, because he had half the atomic weight. So they always came in pairs; to get oxygen sixteen, if he thinks oxygen is eight. "The atoms of oxygen are in pairs which will be more fully developed hereafter." Not in this paper. But there was no hereafter, because of his breakdown, okay? "And only for two atoms of oxygen two chlorines can be substituted; thus." So you can start with CO2, get a phosgene and carbon tetrachloride. "In the same manner with bodies that contain multiples of C2 united to hydrogen." So carbon is the same deal. It always comes in pairs, because its atomic weight is six instead of twelve. Okay. "Take the inverse of this. If the four atoms of hydrogen were bound together, we could evidently expect to form such bodies as H4 Cl4." Right? If the four were all held together and could make four bonds, then you could get H4 Cl4, or these other compounds. "One would naturally expect to find carbon substituted for chlorine, and find bodies like H4, Cl2, C2, and so on. These bodies are not only unknown, but the whole history of hydrogen might be investigated and not a single instance be found in favor of the opinion that it has any affinity for itself, when in union with another element."



Okay? Remember, getting all the chlorines together was Schwindler's idea. Okay, so here, now that he have the idea that these atoms associate together, it makes sense to talk about a structure. Right? What's connected to what's connected to what? You know, the toe-bone and the ankle-bone and all that stuff. Okay, so now you could write structural formulas. So these were the very first structural formulae, because before this there was no such thing as structure. Okay? Thus methylic alcohol has this formula, and ethylic alcohol has the other. So now do you see the theme here? What's he using to denote bonds?



[Students speak over one another]



Professor Michael McBride: Three dots are a bond. So he has a carbon bonded to three H's, and also to OOH. That's because oxygen is doubled, right? So we would write it this way. Or ethyl alcohol we would write that way. Or this compound. Now, this one has a printer's error in it. What is it, and what's the printer's error? What's the substance? Seth, can you read it off to me? Start from the left, the bottom left; what is it? What's the first group? There's a carbon. That's a single carbon. Yeah, methyl, CH3. Right? He doesn't draw the three bonds to hydrogen separately, but it's clearly C bonded to three hydrogens. What's next? Keep going. It's a C bonded to what? The top left. It's bonded to the methyl, of course. What else?



Student: Oxygen and hydrogen.



Professor Michael McBride: Bonded to oxygen, to a single oxygen; although he wrote two, right? And what else is that carbon bonded to?



Student: Hydrogen.



Professor Michael McBride: How many? How many hydrogens is it bonded to? The top left carbon.



Student: Two.



Professor Michael McBride: The two. So it's CH3CH2O. Now, can someone help me out with the rest of it? Ryan?



Student: Well wouldn't that only just be one O at the top because --



Professor Michael McBride: It should be one O, but he's got the different atomic weight. John?



Student: Shouldn't the C's on the far right be bonded?



Professor Michael McBride: Ah, the C's on the far right should be bonded to one another. The printer left that out. So it's symmetrical. It's diethyl ether. Okay? And it's the correct structure for it, except for the printer's error. Okay? Now, he says: "There is no reaction found where it is known that C2 is divided into two parts. It is only consequent therefore to write, as Gerhardt does, C2 simply as C, it being understood that the equivalent of carbon is twelve." So he changed it in this paper to be twelve. So now his formulas are going to be more like ours, without the C2. So he has a formula for glycerin, and one for glyceric acid. So let's do glyceric acid. What do you see up at the top? What's the top carbon? Max, what do you say? What's it bonded to, the carbon at the top, on the right?



Student: Bonded to a hydrogen.



Professor Michael McBride: One hydrogen.



Student: OH group.



Professor Michael McBride: An OH group.



Student: And another --



Professor Michael McBride: Pardon me?



Student: Another OH group.



Professor Michael McBride: Two OH groups, and the carbon below. Right? Because carbon can bond to carbon; that's the special thing about it. Okay, so here's the one on the left, in a formula nowadays, and there's the one on the right. And it turns out he guessed a bit wrong, because he put two OH's on the top part and none on the middle one; and it should be the other way around. So in truth he should've changed those two. Of course, he had no way of knowing it. And the same on the right. Okay? So he's pretty close with his structure. But probably the most remarkable thing is his structure for glucose, which is shown here, and which -- now, he doesn't show the vertical bonds between carbons. He uses this curly bracket. So that's a little reminiscent of type theory. But clearly that's what he meant was the carbons to be bonded to carbon. So that's the formula he wrote. And that's the correct structure for glucose, if it's hydrated. Okay?



So you can add -- glucose is an aldehyde, CHO on the bottom, double bond O. So you can add water to the aldehyde; that actually happens. And you get the structure he wrote here. So it's right, if the glucose is in water. Right? So try yourself. This is a good exercise for reviewing what we've been doing. Depending on how you define it, it's a two or a three-step sequence of HOMO/LUMO interactions that take water plus an aldehyde into a diol; which is what you have here. So try that one out. Okay? And in the French version of the same paper, which as we said appeared a little later, he drew it this way. And what do you notice is different now? There are two things different actually.



Student: Straight lines.



Professor Michael McBride: Yeah, now he's drawing single lines, straight lines for bonds. So this is the first time that was done; none of those dotted lines anymore. But there's a typo. Instead of drawing OH here--and notice--yeah, instead of drawing OOH here, for a second, he added--he's got the wrong formula, he's got H2 there. But this is really amazing, that he got the correct structure for such a complicated molecule, in the very first paper where the idea of structure, or of bonds altogether, was proposed. So this is really amazing. And we owe to Couper the idea of bonds, and the notation we use to do them. So here's Old Aisle Cemetery in Kirkintilloch in Scotland, yesterday. And there you see that this is the gravestone of Archibald Scott Couper. And down here is your thank-you note. [Students react] And those are the flowers--that's the heather you bought to put on his tombstone. So that was our secret thank-you note. And we thank Susan Frew for being our agent and delivering this yesterday. [Applause]



Okay. But on the front of the building Couper doesn't get any notice at all. It's only you guys that are in the know that can do that. Right? On the front of the building they say Kekulé. So let's talk about Kekulé. And we'll also talk about Hofmann, who we've mentioned before as the guy in the top-hat on the right end of that picture in Liebig's lab in Giessen. And Cannizzaro. Okay, this is a drawing that Kekulé made at the age of thirteen. By my standards, that's a pretty darn good drawing. And here's one he made at the age of eighteen. And he went to Giessen University to study architecture, but while he was there he went to Liebig's lectures and got fascinated by chemistry and became a chemist. Okay. He studied with Liebig and after he finished he wanted to go to Paris to study with the people there. Liebig, remember, had worked with Gay-Lussac. So he knew from Paris. What do you think he would tell Liebig, the young guy about to go off to Paris?



Student: Stay away from Dumas.



Professor Michael McBride: Stay away from Dumas perhaps, right? What he said was, "There you will broaden your horizons, there you will learn a new language, there you will learn to know the life of a great city, but there you will not learn chemistry." But he went anyhow. He also went to England and became widely acquainted with all the leading chemists of the day. Then he got a job in Heidelberg, where he was from 1856 to '58, and he did research on cacodyl; that's that free radical involving arsenic. A remarkable -- his student, who worked with him on that was Baeyer, who we'll talk about a lot later. That was Bunsen, who was the big cheese in Heidelberg, was the one Kekulé was associated with. And he did the work, not in the laboratory, but in the kitchen in his apartment, this work on arsenic.



So in 1857, while he was there, he proposed a new type, based on carbon. So carbon could have four things attached to it, which you could substitute one for another. Right? So this was where he got the idea of carbon being something special, the basis of a new type, and that it was tetravalent. And ultimately he got onto the idea, a year later, that carbon could link to itself. So he had proposed the marsh gas type before Couper's paper. But this other one was essentially simultaneous; just slightly later written but earlier published. Here are his Observations on Mr. Couper's New Chemical Theory.



"In fact, in two memoirs which have appeared in Liebig's Annalen, I have put forward different views, which, in my opinion, should furnish a clearer insight into the constitution of chemical compounds." [Clearer than Couper's.] "I may be allowed to indicate that [my first paper] lays down the principle...which I have called the basicity of atoms." [We'd say valence.] "If Mr. Couper thinks he has discovered the cause of this difference of basicity in the existence of a special kind of affinity, I am the first to admit that I have no right to contest his priority in this."



But he was quick to say that he was the one that had the idea first. And he went to quite a successful career, in contrast to Couper. Here he is in Ghent, the period during which he proposed the structure for benzene, with a bunch of people whom we will talk about later; or at least several of them we'll talk about later, as his students. And then he became the leader in Bonn, in a big new chemistry institute there. Here he is in 1872. And we'll talk a lot about this guy on the right here. So he became a really prime leader in chemistry.



Now, facts, ideas and words. Words means nomenclature. So these new ideas needed new words, and also not just words written out with letters, but notation; what symbols are you going to write for these things? And a further development to that is actual physical models that you could put together to show what you are talking about with molecules; if they have structure. So Hofmann was great with language. He actually went and became a leading professor in the Royal College of Chemistry in London, because Prince Albert -- you know, Victoria and Albert -- Albert was a German. So he had a lot of connections with German chemistry and went to Liebig. Liebig said, "Hofmann's a great guy, you should get him." So they got him. So he came to London, and was completely fluent in English, and very good at languages altogether. And he'd had the idea of systematizing the names of hydrocarbons. So he was going to base it on Latin roots. So four will be "quart," right? What do we say for four? Butane, right? We talked about that last time. So "quartane" is the start. And then you can start pulling hydrogens off and have names for the radicals. So you could pull one hydrogen off; that's "quartyl." Because yl, remember, is the root that means radical. And then you can pull a second hydrogen off; then you have "quartene." And then you pull a third hydrogen off, it's "quartenyl." And then "quartine, quartinyl, quartone, quartonyl, quartune, quartunyl." What was his system? How can you remember which one is "quartune"? Yeah?



Student: It's the vowels.



Professor Michael McBride: It's the vowels in order, a-e-i-o-u. And we preserve the a, the e and the i; although we write y, instead of i, for alkyne. Right? That's where it came from, from Hofmann. But the quart didn't stick, because but was well established by then. Okay, now how about Kekulé's -- so that's words. How about notation? So in the paper where he proposed benzene, in 1865, while he was Ghent, he says, in a footnote:



"For greater clarity I am presenting at the end of this note a table giving graphical formulae for most of the substances mentioned. The idea that these formulae are designed to express is rather well known now; so it will not be necessary to dwell upon it. I am keeping the form that I adopted in 1859 when expressing for the first time my views on the atomic constitution of molecules."



So 1859 is at the time he was proposing valence and so on; 1858, as I remember. So this is now five, six years later, and he's applying it to benzene. "This form is nearly identical with that which Monsieur Wurtz used in his beautiful lectures on chemical philosophy. It seems to me preferable to the modifications proposed by Messieurs Loschmidt and Crum-Brown." So let's first look at what Loschmidt did, and Crum-Brown. So here's what they said, to which Kekulé considers his superior. Well first let's think, what should a formula show? Right? What's the very first thing a formula should show? It has to correspond with the facts, right? So what facts do you deduce about a molecule? What should the formula show? Kevin?



Student: What elements are in the --



Professor Michael McBride: What elements are there: carbon, hydrogen, oxygen, chlorine, nitrogen, whatever. Okay, what else? Shai?



Student: The ratio of elements.



Professor Michael McBride: The ratio of the elements, right? Berzelius also had this. But what's going to be new? Composition: the elements and the number of atoms. What next; now, if we've got the new theory, the valence theory? Sherwin?



Student: The structure.



Professor Michael McBride: The structure. What do you mean by the structure?



Student: The relative positioning.



Professor Michael McBride: You mean like x,y,z coordinates? You could write numbers. Here's the origin; I'm going to give this 1.238 and so on. What?



Student: Then which atoms are bonded.



Professor Michael McBride: Ah, which ones are connected to one another. Right? Not their positions in space; but this is connected to this, this valence goes there, and so on. Okay, so Berzelius was already fine for composition. But now we also need to show Constitution -- that's the second C is constitution -- which means the nature and sequence of bonds. Now sequence is clear. This is bonded to this, bonded to this one. What do I mean by the "nature" of the bond? How could bonds differ in their nature?



Student: Strength.



Professor Michael McBride: Single, double, triple, right? So, and you need to be able to show isomers, which Berzelius couldn't show. Okay? So here's Lohschmidt. This is Lohschmidt's formula, 1861, for acetic acid. Can you understand it? Andrew, what do you say? No hope. What do you think the stuff on the left is?



Student: CH3.



Professor Michael McBride: Pardon me?



Student: CH3.



Professor Michael McBride: CH3, right? So the size is something about the weight, right? So it's CH3. What's it bonded to? C; same size. What's it bonded to?



Student: O.



Professor Michael McBride: O, H; and what's special?



Student: Double bond.



Professor Michael McBride: Double bond. So he's got a way of showing a double bond, right? So we can easily understand it. Okay, now here's Crum-Brown. Now that's benzoic acid in Crum-Brown's paper. So the COOH is even clearer than it was in Loschmidt, right? What's a double and single bond. The same thing we use now. Okay? What's different is the C6 part and five hydrogens attached to it. But we don't really know how the C6 is arranged. Although he also showed this one, for phenol, which is a benzene with a hydroxyl group on it, OH. So there you see a ring. It's double, single, double, single, double, single; and H is attached to all the carbons, except for one, which has OH. So that's easy to understand. Shai?



Student: Where does the idea of double bonds come from? I feel like it's just --



Professor Michael McBride: Because if you say carbon has four, then you have to figure out where the fourth one is. Right? If it's associated with three atoms, but can make four bonds; it must make two to one of them. That's the idea anyhow. Or it could be that they go to the center of the ring, the fourth one and so on. There are other ideas too. Okay, or here is a reaction in Crum-Brown's notation. So it's an aldehyde. You add the HOMO of cyanide, and a proton to the O, and you get that compound; a cyanohydrin, as it's called. So this is almost identical -- what difference is it from our current, how different is it? We don't draw circles around the elements anymore; that makes it faster to draw. Otherwise it's essentially the same. Okay? Or here, he says there are two kinds of alcohols: "true" alcohol, which can lose water and give an olefin; or olefin hydrate, which can lose water and give the same olefin. What's the difference between the true alcohol and the olefin hydrate? How are they isomers? Russell?



Student: The position of the OH.



Professor Michael McBride: Right, the position of the OH relative to the R. Okay? So he's explaining isomers here. So those are pretty good. But Kekulé says his is ever so much better. This is Kekulé's structure of benzene, from that same paper. Remember he says at the end, I'm going to give my superior formulas. So what in the heck is he talking about? Eric, you got any idea? So what do you think that is? Eric? No idea? Well you don't have much to work with in benzene. [Laughter] What do you have?



Student: Carbon and hydrogen.



Professor Michael McBride: It's C6H6. Right? So that's got to either be carbon or hydrogen. Which do you think?



Student: Hydrogen.



Professor Michael McBride: The top left there. No idea. Anybody got an idea? Why do you say carbon Kate?



Student: Yeah, because that one's showing double and single bonds all together.



Professor Michael McBride: Ah, it's showing up above here, the schematic one is showing double bonds between carbons and then a single bond; then a double bond and a single bond; then a double bond. Well let's go -- that's open chain, and this is closed chain. What are these arrows on the end?



[Students speak over one another]



Professor Michael McBride: Ah, those are the ones that loop back and attach to one another, somehow, to make a ring. Okay? Do you think he's trying to give -- to imply Cartesian coordinates, for the location of every atom, by where he draws these symbols? Is he trying to show x,y,z coordinates? What do you think Nate? Think he thinks he's showing that, or not?



Student: I don't know.



Student: Which Nate?



Professor Michael McBride: Either of you.



Student: He's got x,y. There, he's got an x,y.



Professor Michael McBride: Oh you could say, right, here's an x-axis and there's a y-axis. He says at a certain point it's that, that, that, that, that. Do you think he thinks that if he had a microscope and could see it, that's where those atoms would be, in fact? Is it physical position he's showing?



Student: They look kind of like [inaudible] to me.



Professor Michael McBride: Kelsey, you got an idea? Do you think he thinks he's showing where these atoms are actually located in space? What is he showing? Yeah Sherwin?



Student: The sequence.



Professor Michael McBride: What's bonded to what, and whether it's single bond or double bond. He's showing Constitution, the nature and sequence of bonds; not positions of atoms, certainly not. Okay, and then the little ones are hydrogens, right? So the length of one of these things is how many valences it has. Right? Four long, for carbon; only one long for hydrogen. And tangency then means a bond. Right? So this is C-H double bond; C-H single bond; C-H double bond, and so on. And then you get to the end, you go back to the beginning. Okay? Now he then shows chlorinated benzene, because chlorine is also monovalent. Right? So you can substitute for hydrogen, and then you could have dichlorobenzene too, or tri-, or tetra-, and so on. Right?



Now is he showing isomers here? That is, in this case, if you substituted this one for the chlorine, instead of that one; same thing or different things? If you believe that this was a picture of things in space, they'd be different. Right? But if all you're interested in is what's connected to what, by single and double bonds, then it's the same. You have the same pattern either way. How about for the dichloro? Would there be isomers of that?



Student: Yes.



Professor Michael McBride: Ah, you could get different sequence, right? They could be adjacent carbons that have chlorines, or next adjacent, or beyond that. Right? So you could use this system to talk about isomers. So how many isomers? Now what do you think Kekulé thought? What do you think he thought about how many isomers of chlorobenzene there are? Do you think he thought there was many? Just what -- guess. I'm going to show you. Okay, he thinks his system is superior. Now he also used his system -- in this same paper he did a lot of formulas. But he showed these compounds. Like this one. What's that group, on the left?



[Students speak over one another]



Professor Michael McBride: What's the group here? What is that? What's this atom? CH3. What's this? CH2. OH; O is two units long, right? So CH3CH2CH2OH. Alcool propylique; propyl alcohol. Get it? Now what's this one? CH3, CH3, CH, OH. It's alcool méthyle-éthylique; methyl ethylic alcohol. So this would be ethyl alcohol, the part on the top. But we've got a methyl substituted in it. And notice, he's having to change his formula, because now these are horizontal tangencies that are bonds, but this vertical tangency doesn't count as a bond. So it's getting a little complicated to explain to students. Right? Okay, what's the next one? CH3. What's this?



Student: C-O.



Professor Michael McBride: C double bond O. CH3. So that's our old friend acetone. And what's this one? CH3. CH-OH. Right? CH3. Alcool acétonique. So it's the alcohol you get by adding two hydrogens to acetone. Now what do you think about this? Do you see anything interesting? This one is CH3CHOHCH3. This one is CH3CHCH3OH. Sophie, what do you say?



Student: They're the same thing.



Professor Michael McBride: Ah, they're the same nature and sequence of bonds. Right? That one is what we know. That's that one. That one, that one. But the last, and number twenty-eight here, are the same thing, but he draws different formulas for them. So his system is not showing isomerism properly. Right? He drew these different. He didn't say that they were the same. Right? And they were thought to be different. Because when you do one -- when you prepare it one way and when you prepare it another way, you get a different boiling point say, or something. I actually don't know what the data was, but there was some difference that said there were these two different compounds. Did you ever get a wrong boiling point? So did they. Right? It just turned out the experiment was wrong. But that shows that his system is not as good as these other systems; whatever he said.



Okay, now how about physical models of molecules? Well here's a great patent that was issued by the U.S. Patent Office to Samuel M. Gaines of Glasgow, Kentucky, in 1868. And we can read about his description of what was so good about this. He says,



"Now, if a class thus instructed be at the recitation bench" [So here's our recitation bench, right?] "and one of them is requested to form carbonic acid, he will walk quickly to the table in front, where the box is open, and place on one of the shelves a cube marked 6, and two marked 8. Carbon, one, and oxygen,n two. If requested to form lime, he will place on the shelf a cube marked 20, (calcium,) and another one marked 8, (oxygen.) these being the elements of lime. These cubes being all placed in contact, and the pupil being asked, 'What have we now?' answers ‘carbonate of lime'." Right? Okay. "Another pupil may then be asked to give the proximate analysis for carbonate of lime. And the cubes before him will indicate the answer and fix it in his memory: ‘It is composed of one equivalent of carbonic acid, (22) and one of lime, (28)'." [You just pull them apart, and you see what the proximate analysis is.] "Another may be requested to give the ultimate analysis, and by the same means he is furnished with the answer. ‘Carbonic acid is composed of one atom of carbon, two of oxygen, and lime is composed of one atom of calcium and one of oxygen.' The intelligent educator will see at a glance, that in this chemical alphabet he has the means of at once arresting the attention of his class and converting what had hitherto been an irksome task into a pleasant recreation. The atomic theory, the characteristics of affinity, the law of multiple proportions, the nomenclature, isomerism, etc., may all be learned in half the time required heretofore, and learned so as not likely ever to be forgotten."



Now what do you think? Would this show the atomic theory that molecules are substances that are composed of atoms? Does it show it?



Student: Yes.



Professor Michael McBride: Yeah. Okay, how about the characteristics of affinity; single, double, triple bond, valence and so on, show that? Not in any obvious way that I can see. So I doubt that one. How about the law of multiple proportions? Remember what the law of multiple proportions was? You have the same elements -- if you have a parts of A and b parts of B, and so on, all that stuff, then what AD will be, and so on. [Here and in the following Professor McBride confuses the Law of Equivalent Proportions with the Law of Multiple Proportions, which the Gaines Models do show.] Yes Max?



Student: Why did Kekulé only show one oxygen, in his notation?



Professor Michael McBride: Pardon me?



Student: Did Kekulé know what the exact mass of oxygen was? Because he only had one in his notation.



Professor Michael McBride: Yes, but notice everything here is halved.



Student: Oh okay.



Professor Michael McBride: So it's okay, the right proportions, right? Okay? But multiple proportions, you have to know something about how many of one associates with how many of another; which isn't obvious here. Okay? Nomenclature? Well maybe, calcium oxide and so on, CO2. Isomerism, that you can have the same things differently arranged? Well obviously you can put the blocks on top of one another or beside one another. But how do you know how many you should do? Right? Not very clear to me that this system works well. So those models are good for some things, but not good for an awful lot of things.



Okay, now James Dewar in Scotland, a student, had the idea of using brass strips to show models. And, in fact, he could put these black balls on. He said, "To make the combination look like an atom, a thin round disc of blackened brass can be placed under the central nut." I think that's really cute, to "look like an atom," as if he knew what an atom looked like. Right? Okay, so then he drew -- he made these models from his strips of possible structures of benzene. Right? You can see where the atoms are, and the four valences on every carbon, and what they -- which direction they go. That particular one, that compound, that structure, was actually prepared about 100 years later, and when it was, it was called Dewar benzene. Right? That one's interesting. It's got a four-center bond, four carbons all sharing the same bond. Right?



Hofmann, the guy, remember, who was good at language and was in London; his last lecture in London was at the Royal Institution, the place where Faraday was, and where Davy had been and did the experiments on electrolysis. So he gave this public lecture where he used croquet balls for atom models, with sticks stuck in them, and he colored them, and the system he chose is the one we still use for what color the atoms are: white for hydrogen, red for oxygen, blue for nitrogen, black for carbon. Okay? So here's part of the lecture he gave. So he would build molecules before your very eyes. He'd start with a hydrogen, put an oxygen on, and another hydrogen; lo, he's made water. Right? Or he could make marsh gas, methane. Or he could show the oxides of HCl. So you start with HCl, but he could take it apart and put an oxygen in between, or put two oxygens, or three or four oxygens in between. You'd go from hydrochloric to hypochlorous to chlorous to chloric to perchloric. So this is like Lavoisier, going to different levels of oxidation.



Unfortunately, those aren't the structures. Right? What really happens is that the unshared pair of chlorine can share with oxygen to give this, which is the active ingredient in Clorox bleach. Or you can use another unshared pair of chlorine to put another oxygen. Or another for another, and get the oxidizing agent that Gay-Lussac used, the chlorate. Or you can put a last one. And this is an explosive. In fact, ammonium perchlorate, AP, is a component of military explosives. Notice that in Hofmann's thing you could've put fifteen oxygens in a row if you could have reached high enough. But you can only go to four here, because that's how many unshared-pairs chlorine has -- choride has to use.



Okay, but then he showed the chlorination of marsh gas. So monochlorinated, dichlorinated, trichlorinated and tetrachlorinated marsh gas. Now this is interesting because for the dichloro, you could imagine having two isomers with these models. Right? The chlorines could be near one another or opposite one another. Right? Now some years later there was an exhibition in London and they showed molecular models. By this time Hofmann was back in Germany. But they got a guy in Spain to make some Hofmann-style models. Manuel Gonzalez Rodriguez from Madrid made this set. And here's chloromethane made from that set. What's the difference between these -- it says "according to Hofmann" here, "Segun A.W. Hofmann". Right? What's the difference? Alex?



Student: Bonds on the carbon aren't just- aren't 90° from each other.



Professor Michael McBride: They're not 90° from one another. They're 109.5° from one another. It's a tetrahedron, not a square plane. Right? Now what does that say about isomers? So suppose we substitute another chlorine. There's a dichloro; there's a different dichloro; there's a different dichloro. How about it?



Student: No.



Professor Michael McBride: They're just rotated from another. They're not different, right? So depending on whether it's square plane or tetrahedral, you'll get different numbers of isomers. That's neat, because it gives you a tool for telling whether it's square plane or tetrahedral. Right? So you can get structure from analysis. So, but these constitutional models try to show the nature and sequence of bonds for Hofmann, not the arrangement in space; and not for Hofmann, and not for the guy that made these models. They weren't trying to explain isomers in that way. All they were trying to show was the same thing you draw in these pictures: the nature and the sequence of bonds; whether they're single or double, triple, and what's associated with what. So that comes a little later, trying to get to the arrangement in space. But the pièce de résistance of Hofmann's lecture was this: that if you take these croquet ball models and make the olefiant gas, which we talked about last time, it's got two spare valences, and if you add chlorine you can get the dutch liquid, the dichloroethane.



But remember what we saw last time. He called olefiant gas "unfinished" or "not saturated"; we now call it an unsaturated molecule. And it's reactive because of these dangling valences. And we now know why it's that way. That he's got the wrong structure; it should be CH2CH2, not CHCH3. And we know about HOMOs and LUMOs now. But that was what he really showed, was that the model had something completely new that no one anticipated. It could explain what, that's new? What new qualities has he associated with the models? What elements are there, what ratio they're in. Right? How they're linked to one another, single, double, triple bonds. What's he adding? Reactivity. Right? So it's something new. So the whole history is adding new things, new properties to the models. And there are more to come. Okay. Okay, that's it.



[end of transcript]

Lecture 24
Determining Chemical Structure by Isomer Counting (1869)
Play Video
Determining Chemical Structure by Isomer Counting (1869)


Half a century before direct experimental observation became possible, most structures of organic molecules were assigned by inspired guessing based on plausibility. But Wilhelm Körner developed a strictly logical system for proving the structure of benzene and its derivatives based on isomer counting and chemical transformation. His proof that the six hydrogen positions in benzene are equivalent is the outstanding example of this chemical logic but was widely ignored because, in Palermo, he was far from the seats of chemical authority.




Transcript



October 31, 2008

Lecture 25
Models in 3D Space (1869-1877); Optical Isomers
Play Video
Models in 3D Space (1869-1877); Optical Isomers


Despite cautions from their conservative elders, young chemists like Paternó and van't Hoff began interpreting molecular graphs in terms of the arrangement of a molecule's atoms in 3-dimensional space. Benzene was one such case, but still more significant was the prediction, based on puzzling isomerism involving "optical activity," that molecules could be "chiral," that is, right- or left-handed. Louis Pasteur effected the first artificial separation of racemic acid into tartaric acid and its mirror-image.




Transcript



November 3, 2008

Lecture 26
Van't Hoff's Tetrahedral Carbon and Chirality
Play Video
Van't Hoff's Tetrahedral Carbon and Chirality


With his tetrahedral carbon models van't Hoff explained the mysteries of known optical isomers possessing stereogenic centers and predicted the existence of chiral allenes, a class of molecules that would not be observed for another sixty-one years. Symmetry operations that involve inverting an odd number of coordinate axes interconvert mirror-images. Like printed words, only a small fraction of molecules are achiral. Verbal and pictorial notation for stereochemistry are discussed.




Transcript



November 5, 2008



Professor Michael McBride: At the end last time we saw this intemperate quote from Hermann Kolbe on the young van't Hoff: "The fancy trifles in it" (his book about the arrangement of atoms in space) "are totally devoid of any factual reality." The reason he thought so, Kolbe, was he knew there could be no way of knowing about the arrangement of atoms in space; they were just too small. But he was wrong. There was plenty of factual reality. What van't Hoff had done was to collect information about isomerism; in particular about optically active compounds, those that rotated the plane of polarized light, or twisted it.



In looking at the Wiki, I noticed some people talk about "bending" it. Bending would be like this. Right? That's not what happens to the plane of light as it goes. It goes and it twists. Right? It rotates; it doesn't bend. Bending is refractive index; you know, light goes into the surface of water and then bends. But this is twisting the direction that the electric vector oscillates.



Okay, so these optically active compounds that would do that, when in solution, were numerous, and van't Hoff was able to cite a number of examples; for example, lactic acid. But you remember who discovered lactic acid? Scheele already -- close to 100 years before this, Scheele had isolated it from sour milk. But Liebig then found the same analysis of the acid he got from meat… isolated it from meat. And Wislicenus, at this same time that van't Hoff is working -- in fact, the guy who had one of his helpers translate the work of van't Hoff into German for that publication and was therefore castigated by Kolbe -- Wislicenus showed that these things had the same connectivity, by chemical transformation; these two acids, the one from milk and the one from meat. Right?



If you look forty years later, in the Encyclopedia Britannica -- so this is a long time afterwards, right? -- about lactic acid, it says it's hydroxypropionic acid. "Two lactic acids are known, differing from each other in the position occupied by the hydroxyl group in the molecule; they are known respectively as α-hydroxypropionic acid." You remember what α means? It means on the carbon adjacent to a carbonyl group. Okay, so there's α-hydroxypropionic acid, which came from fermentation and is called inactive lactic acid. Why would they call it inactive; in what sense inactive? Can you think what active versus inactive would mean? It means whether it's able to rotate, to twist the plane of polarized light; whether it's optically active. Okay, inactive lactic acid -- and they give the formula, with an OH on the carbon next to the carbonyl of the acid group -- and 3-hydroxypropionic acid, or hydracrylic acid, which has the OH on the terminal carbon. Okay? Now these are clearly constitutional isomers. It's a different sequence of bonds. Okay?



But they go on to say, "Although on structural grounds there should be only two hydroxypropionic acids" (these two) "as a matter of fact four lactic acids are known. The third isomer" (in addition to the inactive one and the one that has the OH on the last carbon) "the third is sarcolactic acid." Sarco means from flesh. Okay, so that's the one that Liebig found in meat extract. In fact, Liebig was so interested in meat extract that he patented bouillon cubes, and they were known for a long time as Liebig's material or whatever, I forget exactly. But I once looked at a train menu from the late 19th century, in the restaurant car of the train, and they were serving Liebig's extract. Okay? So there's the third one is Liebig's, the sarcolactic acid, "and may be prepared by the action of Penicillium glaucum on the solution of ordinary ammonium lactate. It is identical with α-hydroxypropionic acid in almost every respect, except with regard to its physical properties." That's sort of an interesting statement, isn't it? That it's identical in every respect, except it's all different. Right? What was similar about it, do you think? In what way was it identical, if it had all -- its boiling point, everything was different; melting point, crystal form, everything was different, all the physical properties. But how could they possibly say "it is identical in almost every respect"? Kevin?



Student: Its formula.



Professor Michael McBride: What about its formula? Obviously they're isomers. So they have the same composition.



Student: The constitution's the same.



Professor Michael McBride: But the constitution is the same, the sequence of bonds. Right? "The fourth isomer, formed by the action of Bacillus laevolacti on cane sugar resembles sarcolactic acid in every respect, except in its action on polarized light." So one rotated it to the right, the other rotated it to the left, and in every other respect they were identical. So there were four -- [and of a] right, left and inactive. Okay? But here, forty years after, they don't give any explanation of why that should be so. Okay, tartaric acid we've seen a lot. Right? Scheele; Berzelius coined the name isomer to talk about it; Pasteur; Wislicenus again showed the constitution; which meant nothing to Pasteur. Why would the constitution mean nothing to Pasteur, the nature and sequence of bonds? Yeah?



Student: He believed the positions in space mattered, rather than the sequence of bonds.



Professor Michael McBride: Would Pasteur have written, at the time he did this work, "we don't know anything about the position of bonds"? I don't think he would've written that. You know why? Bonds wouldn't be invented for another ten years. There was no such thing as bonds, when Pasteur did his work in 1848. Bonds came in 1858. Right? So Pasteur didn't have the tools to think about this. Okay? So aspartic acid was an example of something that was optically active. Amyl alcohol could rotate light. Glucose could rotate light. Notice that this didn't mean they had both the one that could rotate one way and the one that could rotate the other way. They had something that could rotate light. In the case of lactic acid they had things that would rotate either way. Right? In the case of tartaric acid, after Pasteur, they had that. Okay? And also laevulose and lactose were another two that were known but were not cited by van't Hoff. Malic acid -- you know what that comes from? Where's the root, mal? It's apples, right? So it's sour apples. Okay, and most of the derivatives of these compounds, most things you could get by chemical transformations from these compounds were optically active, but not the stuff you got by dehydrating malic acid. If you got the double bond, so you got maleic and fumaric acid, two isomers, but they didn't -- neither of them rotated light, neither right or left. Right? So that was destroyed in this case. And also there was another example, which was replacing the OH by a hydrogen. So you have CH2CH2 in the middle, instead of CH(OH)CH2. And that particular experiment, van't Hoff was involved in; he was involved in by suggesting it to his fellow student, Bremer, who did it by treating it with hydriodic acid and phosphorous, which was a way of removing OHs and changing them to hydrogen, in those days. We'll see later that that's called reduction.



But at any rate there were these two derivatives of malic acid that were not optically active, although most derivatives were. Okay? Now, van't Hoff, as I said, was involved at least in suggesting these experiments. And, as we said, those two were inactive. In van't Hoff's obituary, a former student of his, Bancroft, who was by this time a professor at Cornell, wrote: "In his whole life he never made what would be called a very accurate measurement, and he never cared to. I remember his saying to me eighteen years ago, 'How fortunate it is there are other people who will do that sort of work for us.'" Right? So it's great to do experiments, but in some sense it's even greater if you can talk somebody else, who's competent, into doing them for you. Okay? And that was van't Hoff's approach.



Now, what van't Hoff noticed was that these compounds, that were optically active, all had something in common. They all had these red atoms in common. What's special about the carbons that are red, that exist in the ones that are optically active, but don't exist in the ones that are not optically active? What's special about the ones that are in italics in red? What makes them different? This is what van't Hoff noticed. Chris?



Student: Are they chiral carbons?



Professor Michael McBride: Well they are, but what did van't Hoff notice? Chiral didn't exist then, as a name. Yeah Andrew?



Student: They have four bonds [inaudible].



Professor Michael McBride: They all have -- they use their four bonds to bond to different groups. No two things are the same. Down here, this carbon that was red here, has two hydrogens on it. Here it has two bonds to another carbon. Right? But in the other cases they're all, these red carbons are unique in bonding to four different things. This one, for example, this carbon is bonded to hydrogen, to OH, to another carbon that goes out on this chain, and another carbon that goes out on that chain. And those two chains are different. This one has a carbonyl group; that has an OH. So the difference can be quite remote. But they're different. Okay, so that's what he noticed. And we now call these chiral carbons; or stereogenic, because they give rise to special stereo arrangement in space. So what he said was, "Every carbon compound which in solution can rotate the plane of polarized light contains one or more asymmetric carbon atoms." That's what he called these, that had four different things.



And he made models of carbon. This particular set he folded up, colored, pasted together, and gave to this friend, Bremer, who had done the experiment. And they're in the Museum Boerhaave in Leiden now. Now look at these things and see if you can figure out what some of them are. Here are tetrahedra, whose faces are colored. So you could make… if you had four different colors, that would denote being associated with four different things, and each face would correspond to a substituent. Okay? This one has colored vertices. So instead -- you could use a tetrahedron either way. You could use the faces to denote a neighbor, or you could use the points to denote where a bond is to the neighbor. And he did it both ways. What are these here? What are those models? Those aren't tetrahedra. Can you see why he might have made those models? Kevin, what do you say?



Student: To show pentavalence?.



Professor Michael McBride: No. We talked about van't Hoff last time, last lecture. Do you remember what he was talking about? Lucas?



Student: Ladenburg Benzene.



Professor Michael McBride: Pardon me?



Student: Ladenburg.



Professor Michael McBride: Ladenburg Benzene. A triangular prism. That's what these are. And you remember this is 1,2; 3,5; 4,6; or whatever it is, that are this way. If you color these two it's not superimposable on one where you color this one and this one. Or this case. That was AB or AB the other way. Those he said were absolutely different. So those are models he made with regard to this argument with Ladenburg. Okay, now here are his models of these two things you get by removing water -- pardon me -- from tartaric acid. So what he has is tetrahedra for the carbons. And what does he mean when he puts their edges together, like this? What's he using to denote a bond? There we go. What's denoting a bond? A shared vertex. Right? So here's a single bond, right? What's this? A double bond. That's like what Lewis did later -- remember when he brought cubes together to share an edge -- or a little bit like that at least. But now when you do that, when you share an edge, of the other two things on each carbon, they're different; this one red and not red. Right? You could put them together this way, but you could also put them together this way. And you can't get from one to the other without pulling them apart and putting them back together again. Right? So that would explain why they're two different isomers; maleic acid and fumaric acid. Right? That you get -- when you start with tartaric acid and pull off the water. Or start with malic acid. I was saying tartaric. You start with malic acid, pull off water, and you get either this or this. Okay, so he could explain that with his tetrahedral models. The red denotes a red vertex, obviously. Okay? Or if you have a single bond you could count isomers. Now what he says here is you don't get all that many isomers because it can freely rotate about the bond between the two. So the carbon atom is in the center of a tetrahedron, and its substituents are on the vertices. Right? So the top carbon has three things, R1, R2, R3. And what he shows here is you can rotate that freely. You don't count all these as separate isomers. Who would've thought you did count them as separate isomers? You may not remember the name. Where was he? What?



Student: The guy from [inaudible].



Professor Michael McBride: The guy from Palermo. Right? Paternó was his name, right? He tried to explain the existence of isomers. But van't Hoff says no, that's not true; you can freely rotate, so you can't count that. But you could have -- if you looked down on R1, R2, R3, you could look down and see them clockwise, from one to two to three, or you could see them counterclockwise from one to two to three. And you can't go back and forth between those by rotating, because they're in a certain order. So you do get isomers.



So he made this diagram to show that free rotation about the central bond results in rapid inter-conversion, and thus inseparability and irrelevance of the isomers that Paternó talked about. And they can be arranged clockwise or counterclockwise, which is not affected by the rotation about the central bond. And you can then count how many isomers there are, depending on whether R1, R2, R3 are the same as four, five, six, or are different from four, five, six. Okay?



But maybe the most spectacular thing he showed was that if you put three tetrahedra together like that, you get these forms, which -- what's the relationship between these two? Can you superimpose them and have them be exactly the same? Is there any relationship between them? Obviously if you just slid them together, R1 would be on top of R1 and two on two, but four and three -- this is a four here, it looks like a one -- four and three would not be on top of one another when you slid them together. And there's no way of twisting it so it'll be right. Right? But there's a special relationship between those. They're not completely unrelated. What symmetry relates them? Do you see? Zack?



Student: Right handed and left handed?



Professor Michael McBride: Yeah, they're mirror images, like right-handed and left-handed. Okay? So he predicted that if you had two double bonds in a row, with a common carbon in the middle, then you could get right and left-handed, where you can't get it with just a double bond, as in the maleic and fumaric acid. Those aren't "chiral", as we now say. But this wasn't shown until 1935. That's probably the longest that there's been a correct theory proposed, until the time that it was confirmed. Or at least I don't know of anything longer than that; sixty-one years after he predicted it, before it was demonstrated to be true. And it was proved, you see, in the labs of E.P. Kohler, the same guy we talked about last time. And here's the molecule he used, down on the bottom left. And we can enlarge it here. So there it is.



Now, that molecule, as you look at it, would be superimposable on its mirror image. It looks planar. Right? Now let's think about that. The bonds at that carbon should be tetrahedral. So one bond should be coming in and out, or one pair of bonds, and the other pair of bonds should lie in the plane. But if that's true, then the one on the right would have its four bonds coplanar -- right? -- all in the plane of the paper. So there's something wrong here. And you can see what's wrong if you think about the p orbitals that go into making a double bond. Right? What's wrong with that arrangement? Why don't the p orbitals like it?



Student: They're orthogonal.



Professor Michael McBride: They're orthogonal. They don't overlap, right? What you need to do is to twist it, like that, so that the groups, C10H7 and C6H5, come in and out of the plane. Right? And that's what then makes it not superimposable on its mirror image. Okay, now if you wanted to convert one mirror image to the other -- notice if you used the plane, something parallel to the screen, as a mirror, what comes out in front of the mirror would be the same on the left. But on the right C10H7 would be back in, and C6H5 would be coming out. Right? So if you want to interconvert these, you have to put ten where six is and six where ten is. You have to rotate around the bond. But to do that you have to break the bond, because it's a double bond. You'd have to go back to where the p orbitals don't overlap and you lose the double bond, and then you can keep rotating and get the other mirror image isomer. Right? But you have to break a bond to do that. So that's very tough.



Okay, now we were counting isomers in homework for today. And this slide goes through and counts them. I'm just going to begin it, and then hustle through the end, rather than go through it in great detail, because you've already worked on it. But this can be thought of as an answer key. Okay, so we want to count how many different mono-substituent positions there are in that compound. There's one. Okay? There's obviously a different one, two. Now about three, is that different, if you have just a mono-substitution. Is the blue different from the red? No, obviously because this thing has rotational symmetry, you could rotate it; put the red on top of the blue. So the blue's not a new one; forget blue. Okay, now how about that blue one, is that a new one? No, because the one in the left front would go, on that same rotation, go the right rear. So that one doesn't count. How about that one? Is that different or the same? Can you rotate it so the one on the front left becomes the one on the front right? No. So now we have a question. Is that something real about the models? Or is that just like the sausages, that Kekulé used, that it suggests a difference, when there's not a real difference? Okay? So, we'll put a question mark on that one. Okay, so there are two isomers, if we don't count the two on the bottom, and there's an additional isomer if we do count it. Okay? So there are either two isomers, or conceivably three isomers, of that one. Okay, and notice that this one, the third one in, has exactly the same arrangement among the hydrogens as the first one.



So the numbers would be the same: two and an additional one, if you count those extras. Now here, you have obviously three different positions you can have for mono -- top, middle, bottom. But also the ones on the right are not superimposable on the ones on the left, if this is three-dimensional. So you could have another three, if they count. And so on. I'll not go through these -- notice -- except to note in this one that there's obviously top here and these three. Now the question is, are those three different or not? And we're following van't Hoff here and saying it rotates rapidly around here. So on time-average you won't be able to put those in separate bottles, right? They're going to interconvert too rapidly to count. So there'll be three. But the two on the top are not really -- they're not superimposable by rotation, in the model. Okay? So the question is whether you have that additional one. And in prismane there's only one.



And then if you go disubstituted, if we start on the left, if you start with one on the top here, you can have the second one here, or here, or here. That's clearly three. But there might -- this one is not superimposable. That is, if you have this one already, this one and this one, as sites for the second, are not superimposable. Right? So maybe we need another one there and another one over here as well. So there might be an additional two. And again, this is the same as that one. And you can go through this -- and presumably you have done. But anyhow these are the answers I got. And if you found something different, let me know. Maybe you saw something I didn't see. Okay, so that's interesting.



Now mirror images. Here's me with a mirror image. Right? Now this is a question. Why does a mirror exchange right and left, but it doesn't exchange top and bottom? How does the mirror know which way gravity points? Or how about if I lay on my side when I looked in the mirror, would it still exchange left and right and not top and bottom? I'm talking about plane mirrors, not funhouse mirrors. That's an interesting problem, right? But notice, it doesn't change either, because top is on top and right is on right. It didn't change right and left. It left right on right. What did it change? Why does the mirror image look different from the real image, if it doesn't change right and left, and it doesn't change top and bottom?



Student: Front.



Professor Michael McBride: It changes front to back, or in and out. Right? Because my original self, you see the back of my head, and in the mirror image you see the front of me. So what a mirror changes is in and out, not right and left. Okay? But our intuition interprets that as rotation. Because if I look like this, and I go like this, you expect it to change, the right hand, because you're accustomed to people rotating. You're not expecting me to go pfff, like that and turn inside out. Right? [Laughter] Okay, so people don't invert. So our mind makes us think that the person we see in the mirror has turned 180°, so that the right should be on the left. Okay? And, of course, this is not a new observation. Through the Looking-Glass -- you know, Lewis Carroll, and Alice going through. And here are successive pictures in that book. There she is looking into the glass, and we turn the page, and here she is coming out, on the other side.



Now notice something about this. That's her right arm. Right? And in the previous picture it was also her right arm. That's not the way the mirror image should look. It should've changed the way it looks, when you go from back to front. Okay? Now, Lewis Carroll was a smart guy. He taught mathematics at Oxford. Right? But of course he didn't draw the picture. The picture was drawn by John Tenniel. And it turned out that John Tenniel was blind in one eye; a fact I only discovered yesterday, on Wikipedia. [Laughter] So is that the reason that he's unable to perceive depth, and draws the wrong thing? No, it's much more subtle than that. Notice that the name of the book is Through the Looking-Glass. She moved through the looking-glass and her right arm stayed right. But if she had been reflected in the looking-glass, if the title had been chosen to be In the Looking Glass, rather than Through the Looking-Glass, then when you looked at the mirror image it would've been the left arm. That's sort of cute. So Carroll is more subtle than you might give him credit for; or John Tenniel, or both. Okay? Or look even further in the book. This was, notice, 1872. What was going on in chemistry in 1872? Wislicenus was studying these isomers of lactic acid, the Scheele and the Liebig, and showing that they had the same constitution but were different. Okay? So look what happens here when she's talking to her kitty.



"Now, if you'll only attend Kitty, and not talk so much, I'll tell you all my ideas about Looking-glass House… Well then the books are something like our books, only the words go the wrong way. I know that because I've held up one of our books to the glass, and then they hold up the one in the other room… How would you like to live in a Looking-Glass House, Kitty? I wonder if they'd give you milk, there? Perhaps Looking-glass milk isn't good to drink."



Because in the college Carroll knew Vernon Harcourt, who was a chemist who was involved in this work, with Wislicenus. So almost certainly they'd had conversations about lactic acid, the mirror image form. Right? So there's a lot of stuff in Alice. So was he referring maybe to sarcolactic acid, the one that comes from meat instead of from sour milk? Anyhow, chirality comes from the Greek word "χÑ

Lecture 27
Communicating Molecular Structure in Diagrams and Words
Play Video
Communicating Molecular Structure in Diagrams and Words


It is important that chemists agree on notation and nomenclature in order to communicate molecular constitution and configuration. It is best when a diagram is as faithful as possible to the 3-dimensional shape of a molecule, but the conventional Fischer projection, which has been indispensable in understanding sugar configurations for over a century, involves highly distorted bonds. Ambiguity in diagrams or words has led to multibillion-dollar patent disputes involving popular drugs. International agreements provide descriptive, unambiguous, unique, systematic "IUPAC" names that are reasonably convenient for most organic molecules of modest molecular weight.




Transcript



November 7, 2008



Professor Michael McBride: We're talking about facts, ideas and words, in relation to stereochemistry. And the words have to be generalized, in the way we communicate things, to include models; things like this, right? [Technical Adjustments] Okay, so models, but also what you can draw on a piece of paper that will convey stereochemical structure to people. So the standard we use in drawing is wedges, to mean they're coming out at you, in the direction that they expand, and dashes mean they're going back into the board. Although if all you draw is dashes, it's not immediately obvious which end is going back, if it's a dash. Right? But if one end has nothing on it, it's clear, that it's going back from the carbon that's shown. So you have to use your stereochemical intuition sometimes to interpret that. Although if you draw a wedge, it's clear which way it's coming. And sometimes people draw wedges with dots, meaning they're going back in. But there are many ways that people make mistakes in drawing this. So it needs a little practice. That one's okay, and clearly corresponds to a tetrahedron with 109° degree bond angles, shown like this -- right? -- the blue one's coming out at you, the black- the yellow ones going back, green and red, up and down. Okay, so that drawing is fine. What's wrong with this drawing? Sherwin?



Student: It couldn't mean like a [inaudible]. The upper bond and the lower bond are --



Professor Michael McBride: I can't hear -- I still can't hear.



Student: Oh, the upper bond and the lower bond are [inaudible].



Professor Michael McBride: What are the bond angles? How about between the solid lines? 109.5°?



Student: 90°?.



Professor Michael McBride: Between the solid lines, not the wedges?



Student: Oh, 180°.



Professor Michael McBride: 180°. How about the other angles?



Student: 90°.



Professor Michael McBride: 90°, right? This is a square plane, drawn that way. The dash has to go the other way, and the lines that are in the plane of the board, or the piece of paper, have to be at 109°. Okay, so forget that one. I assure you that we will get these on exams, because we always have in the past; but this is warning. Okay, so that one's planar, forget that. How about this one? There I've got the wedges drawn the right way. Everything else right? No, the solid lines in the plane are 180°. That one's bad obviously. Not many people draw it that way. Okay? Ah, there's one that looks pretty good. Everybody agree? Any complaints? Lucas, what's the complaint?



Student: An acute angle there.



Professor Michael McBride: Pardon me?



Student: An acute angle, over there.



Professor Michael McBride: Ah, the angles between the normal lines and the wedge are all less than 90°. Well one of them is bigger than 90°, the one to the top. But the one to the bottom is less than 90°, if the normal line is in the plane of the screen. Right? That's an acute angle. Right? So the one on the top left is the one you want to go for. Okay? Now how about this? Do those look good? See any problems? Kate, what do you say? Looks fine to you. Tell me what the angle is here, between this bond and this one. Where would this bond point if it were 90° from this one, which is in the plane of the board?



Student: Straight out.



Professor Michael McBride: It would point straight out. Now if at 90° it would be coming straight out from that point. How about there? Is this angle greater than or less than 90°? Here's the carbon, here's the one that's in the plane of the board. Right? It's going like that. Here's the one -- that's 90°, right? Suppose I bend it up like this, which is what the wedge shows, now what's the angle?



Student: It's acute.



Professor Michael McBride: It's acute, right? So these are not right, right? The same is true at the bottom. But this structure is okay because, by convention, it doesn't mean to show the exact angle. It just means to show that it's going relatively out, right? But it wouldn't be -- the solid -- these three lines here, this one, this one, this one, the ones that are not wedges and not dashes, could not be in the plane of the board. They'd be 120° angles, or something like that. They couldn't be 109°, right? But in order to draw this, in a finite amount of time, when you're writing on paper -- people understand what you mean, in this case -- but if you draw that one, which means it's planar, people will be very much offended. Right? Okay, so this again, at a certain level, is lore. You have to be criticized before you find out where you're making mistakes. Okay? But anyhow, it's important to be sensitive about the fact that carbon is tetrahedral.



Now, these things are hard to do in three dimensions. And the Fischer projection was invented for that purpose in 1891. And the reason for its invention was interesting. He wrote:



"With the help of Friedländer's convenient rubber models, one can construct molecules of right-handed tartaric acid, left-handed tartaric acid, and inactive tartaric acid and lay them in the plane of the paper so that the four carbon atoms are in a straight line and the attached hydrogens and hydroxyls lie above the plane of the paper."



Now, here's the idea, that Friedländer's models used rubber tubing to connect these things. Right? So drawing this in two dimensions and making clear what you mean in three is not easy. But he says what you're going to do is take the four carbons that are in a chain -- this is tartaric acid, here's CO2, carbon with an OH, carbon with an OH and a carbon with O2 on it, right? So what he wants to do is to lay all four carbons in the plane of the paper, which you can't do. Right? But if the bonds are made of rubber, then you can bend the bonds -- right? -- and then you can make them all four, in the plane of the paper. Right? But what you have to understand is that in this convention you draw the four carbons vertical in a line; not horizontal. And then you understand that the bonds between them are severely bent. Right? So that the curious thing is that when you draw a Fischer projection, as is most convenient for sugars, for discussing the configuration of sugars, the bond, the four bonds for this carbon are understood that these two, the horizontal ones, come out toward you, as if they were wedges, and the other two from this carbon go back into the board. Okay?



This is his paper where he originally proposed it. Nowadays you wouldn't use dots here, which help make it clear that it's going back, you'd just draw lines. But understand that vertical lines mean they're going back into the board, and horizontal lines mean they're coming out at you. Okay, so that's the Fischer convention, that he invented in 1891. So this bond here is a very funny bond, because in bonding from this carbon to this carbon, it's going back into the board from this carbon, but also back into the board from that carbon, because it's bent. Okay? And it was because Friedländer made his models with rubber tubing that it occurred to Fischer that you could do this. And then all you have to do is draw straight lines -- you don't have to do wedges and so on -- and just put things in the right place.



So here he could draw right-handed tartaric acid, left-handed tartaric acid, and inactive or meso, as we call it now, tartaric acid. So the attached hydrogens and hydroxyls lie above the plane of the paper -- right? -- and the other carbons, vertical bonds, go back in from both of the atoms that they attach. Okay, so there are the wedges. And if you rotated that 90°, it would look like that. Very highly distorted, but you can do it with rubber and be clear about what the configuration is -- unambiguous. Okay, they go back from both carbons, and it was because you had rubber models that you could do that. Okay, so here are those models. And notice that you can rotate in-plane by 180°, and it means the same thing. So if I have this and it's bent like this, and I rotate it like that, 180°, it obviously means the same thing. Right? I can rotate a Fischer projection by 180° and it's still the same model. But I can't rotate it by 180° -- or by 90° this way. Why not? If I write it this way it won't mean the same thing. Why not? Shai?



Student: Because a number of horizontal bonds are going in.



Professor Michael McBride: Yeah, the horizontal bonds are going back in, and the vertical ones are coming out at you. You've reflected it in a mirror, each carbon. So you could rotate it 180°, but you can't rotate it 90°. Okay? So there, we'll rotate that 180°. That's still fine. And in fact what do you notice about it, other than the fact that the symbols are written backwards?



Student: The same thing.



Professor Michael McBride: It's the same thing it was above. Right? So rotating it 180° doesn't change it. Okay? How about the next one, if you rotate it 180°? Same as above? Yes. How about the last one? Okay. That one is now the mirror image. If you slide it up, like that, you can see that there would be a mirror there. Everybody with me? So that molecule is its own mirror image, whereas the first two -- and remember you can draw the mirror any direction and always get the same mirror image; xy, yz, xz, they all give the same mirror image, just differently oriented. Another way to draw the mirror one, that's easier, is to draw it right through the middle of the molecule horizontally, and then the top half is the mirror image of the bottom. Right? So it's obviously not a chiral molecule. It's superimposable on its mirror image, whereas the -- and that one's called meso, the name then; that was called meso-tartaric acid and the name was generalized to mean any molecule which is its own mirror image is "meso". Okay? And notice that these terms name relationships. They don't name an absolute molecule, these stereochemical terms, the ones like this. What's the relationship between those two? Are they identical, or are they just plain different, or are they different in the special way that they're mirror images of one another; the two on the top left?



Student: They're just different.



Professor Michael McBride: Okay? Can you rotate them and make them be on top of one another? No. If you rotate them, they become the same thing they were originally, right? That's what we already did. Are they mirror images of one another? Right. There's a vertical mirror between them, right? So those are non-superimposable mirror images. And the name given to things like that is "enantiomer". So the one on the left, the top left, is the enantiomer of the one that's in the middle. Caitlyn, what's the question?



Student: Okay. It's just the notation I guess. If the OOH were on the other side.



Professor Michael McBride: Oh no, but that's -- but all that means is the grouping C double bond OOH. That doesn't have any arrangement in space when it's just written with letters that way. Okay? That was just because I wanted to make strictly the rotation. Right? It's where the bonds point that makes a difference. Okay? Now, how about the relationship between the middle and the right? Are those identical; are they superimposable? Elizabeth, what do you think?



Student: No.



Professor Michael McBride: No you can't move those, or rotate them any way, and make them on top of one another. Are they mirror images?



Student: No.



Professor Michael McBride: So what are they?



Student: Diastereomers.



Professor Michael McBride: They're just plain different. They're diastereomers. So the one in the middle is the enantiomer of the one on the left but the diastereomer of the one on the right. So you don't say it's a diastereomer, or it's an enantiomer. You say it's a diastereomer of this and an enantiomer of that. Right? So these name relationships. Okay? Okay, how about those two, the first and the last? What's the proper relationship there? Maria?



Student: Diastereomer.



Professor Michael McBride: Right, diastereomer. Good, so we got the idea. Okay now, how many isomers do you have? Obviously if you have one chiral center you have right-handed and left-handed, if there are four different things on it. So there are two isomers. But suppose you have several chiral centers. That's not so obvious then how many you have. So if you have n stereogenic centers, as they're called, and each could be right or left, then you'd imagine the number of permutations would be 2n. Right? That would be expected. But that's not true because of meso compounds. And we'll illustrate that with a quote from van't Hoff's brochure, in 1877, where he said: "Next we consider a symmetrical formula" (where carbon, the first carbon, has one, two, three; the next one has two fours, and the last one has one, two, three, different substituents; and also the one where the two in the middle are different, four and five on the middle carbon.) "As is easily conceived from the foregoing discussion, they lead to only three isomers." Because you have this possibility of meso, that we just were talking about. Although this is with three centers possibly in the second one.



Now, Baeyer and Fischer were -- Baeyer, a student of Kekulé, was the teacher of Fischer. And they got together at a resort on the shore of the Mediterranean, in Italy, and were trying to figure out this particular question of how many isomers you have here. And they had read van't Hoff's thing. So they knew there were supposed to be three isomers. So they were sitting at breakfast and they had bread rolls and they had toothpicks. So they started making models of these and trying to count how many isomers. But they ran out of bread rolls, presumably, and couldn't answer the questions. They were completely flummoxed by trying to understand what van't Hoff was saying. So they gave up on using bread rolls. And that is what prompted Fischer to go back and figure out how to make Fischer projections. Okay?



Now let's look at that particular case using Fischer projections. So we have these three carbons that might have four different things on them. The one in the middle is a little bit tough to decide, because the ones on -- as to whether it has four different things. Obviously it has H, OH, and a carbon. But those carbons that are attached are identical constitutionally. They have the same nature and sequence of bonds, things attached to them, but they might or might not be the same from the point of view of stereochemistry. So it's not so obvious how this is going to work out. But it turns out that it's really easy, if you use Fisher projections. And we'll try it here. So here are the five carbons in a row, the top and the bottom one being COOHs, and you have the three carbons in the middle, each of which has an H and an OH on it, but they could be either way. So at first glance it looks like 23, or eight possible isomers. So let's check them out, instead of using bread rolls, using Fischer projections.



So we'll draw eight Fischer projections, and try to be systematic and draw all the possibilities, and then count them. Okay, so first of all take the bottom carbon, and we'll put in the top row all the OHs to the right, and in the bottom row all the OHs to the left. And then we'll do the analogous thing with the second row, except we'll make the first two to the right, the next two to the left, the next two to the right, the next two the left. Okay? And finally what are we going to do on the top row? Can you see how we're going to -- so that we get all eight? We'll go right, left, right, left, right, left, right, left. Okay, and now we've got all the possibilities. So now we have to go through them one by one and see if they're unique, or how they are related to the others, if they're related. Are they diastereomers? Are they enantiomers? Are they identical? And we're going to use these Fischer projections to do it. Okay, so we've got the first one. Okay? And now we look at the second one and compare it to the first one. Is it the same?



Student: No.



Professor Michael McBride: No. Are they superimposable?



Student: No.



Professor Michael McBride: No, that would make them the same. Are they mirror images?



Student: No.



Professor Michael McBride: No. So two is different, right? They're diastereomers. Okay, now about how the next one, three? Is it like any of the previous ones, or the mirror image of any of the previous ones?



Student: No.



Professor Michael McBride: No, three is unique. Okay, now how about four? Does it look like any of the preceding ones? Wilson?



Student: It's a rotation of two.



Professor Michael McBride: Ah, you can rotate it 180°; which you can allow, right? That'll keep the bonds pointing back, in the convention, right? And it'll look like two. So that one actually is two. So we'll blank it out. Now how about the first one in the bottom row here? Does it look like one? No. Does it look like three? No. Does it look like two?



Student: [Inaudible]



Professor Michael McBride: That's not so easy, right? But it's not superimposable. If you rotate it 180°, about the axis into the screen, it still doesn't look like two. Right? So it's different. Four. How about the next one? Virginia, you got an idea? Does this one here look like any of the previous ones, if we rotate it?



Student: Well, it's the mirror image of three.



Professor Michael McBride: Can't hear.



Student: Yeah, it looks like three.



Professor Michael McBride: Ah ha, it's like three. If we rotate it 180° it's like three. So forget that one. How about the next one? Does it look like anybody we've already seen?



Student: Four.



Professor Michael McBride: Four. Right, if you rotate it 180° it's four. Okay. And the last one?



Student: One.



Professor Michael McBride: Pardon me? Lexy, what do you say?



Student: It's the same as one.



Professor Michael McBride: It's one, if you rotate it 180°. Okay. So we've got these four that are different. Now what are the relationships? Okay, are any of them meso? Do any of them have mirror planes, so that their own -- they're their own mirror image. Angela?



Student: One.



Professor Michael McBride: One has a mirror image. The top is the mirror image of the bottom. There's the mirror. Any others?



Student: Three.



Professor Michael McBride: Three. Okay, but not two or four. Anything else that you want to note about two and four?



[Students speak over one another]



Professor Michael McBride: Sophie?



Student: They're mirror images.



Professor Michael McBride: They're mirror images of one another. Okay? So what we have are two meso isomers -- they're diastereomers of one another, right? -- and a pair of enantiomers. Each of those enantiomers is the enantiomer of one another, and the diastereomer of either of the previous ones, one and three. Okay, so with Fischer projections he could easily settle this question. And that became very important in the chemistry of sugars, where Fischer was deeply involved, as we'll see next semester.



Okay, now here's Halichondrin B, which comes from a marine sponge; as it says here. Okay? Now, you don't get very much of this out of a sponge. Right? But it was found that it was useful in anticancer therapy. So the hope was you could make it and then see if it will work as a good drug. So it was found that you don't need the whole molecule. If you cut it apart, to have only this portion we've seen here, then that part is just as good, medically, as the whole molecule. So that simplifies the synthesis problem quite a bit. But it's still non-trivial. So there's the active fragment of Halichondrin B. And SAR means Structure Activity Relationship. They take off this, take off that, trim it here and so on, change one thing to another thing, and see if you can get better activity than you get from just this thing itself; and especially looking to changes that would make it easier to synthesize. Okay? So they found out that this one, which is called E7389, that model, is about as easy as you can get to synthesize and still have really good activity. But it's not easy to synthesize. And you have the question that you have to get the right diastereomer because the other diastereomers are different. Now how big a problem is that going to be? Well it depends on how many stereogenic centers there are; how many carbons can generate two isomers, by being right or left-handed. So let's start up here at the left, and I'll go along this chain, like that, and you stop me when we get to a carbon that's chiral, that's stereogenic. Okay? So I'll start here, at the top left, and I'm going across. Stop me when I get to one. Oh, you didn't tell me.



Student: That one.



Professor Michael McBride: That one looks -- well you say oh that has only three things on it.



[Students speak over one another]



Professor Michael McBride: You have to have four different things, to be stereogenic. Does anybody think that one is stereogenic? Chenyu?



Student: It left the hydrogen out.



Professor Michael McBride: Can't hear.



Student: It doesn't have the hydrogen in it.



Professor Michael McBride: Ah, it doesn't show the hydrogen; the hydrogens aren't being shown. So there are, in fact, four different things on that third atom -- nitrogen, carbon, but the second carbon is stereogenic. Okay? So there's one. Now let's keep going. How about here, that one? No, that one has two hydrogens. How about this one?



Students: Yes.



Professor Michael McBride: That's stereogenic. How about this one?



Students: Yes.



Professor Michael McBride: Yeah. How about this one?



Students: Yes.



Professor Michael McBride: Okay? How about this one?



Students: Yes.



Professor Michael McBride: Okay, how about this one? This one? This one?



Students: Yes.



Professor Michael McBride: Ah, that's got four bonds already, no hydrogens there. Okay, that one's not. There? Here? Okay there. There? There? There? Yeah, that's got four. There?



Students: Yes.



Professor Michael McBride: Yeah. There?



Students: Yes.



Professor Michael McBride: Yeah, four different things, because you have to count the hydrogen. Here? Yeah, the hydrogen also. That one shows the hydrogen. Okay. Here? Yeah. Here? Here, no. Here? Well you say there are two oxygens. Right? Are they the same?



Student: No.



Professor Michael McBride: No, they're attached to different things. Right? So that one too. Okay, here? No, no. Here?



Student: Yes.



Professor Michael McBride: How many think so? Yeah, right, that one. The next one? This one?



Student: No.



Professor Michael McBride: Yes or no here?



Students: No.



Professor Michael McBride: No. Here?



Students: Yes.



Professor Michael McBride: Yes. Here?



Students: No.



Professor Michael McBride: Speak up.



Students: No. Yes. No. Yes. No. Yes.



Professor Michael McBride: And there's one more carbon here.



Students: No.



Professor Michael McBride: No. Okay. Well done. There are nineteen stereogenic centers. That means there are 219 possible isomers. Will any of them be meso? Are you going to get any mirror images within this molecule? No, it's completely unsymmetrical, right? It's not like tartaric acid. So there are going to be something like -- something greater than half a million configurational isomers. So it's not just a question of putting the right groups in the right place to get constitution. You have to put them stereochemically correct, to get one out of million, right? So this is a problem. But they figured out how to do it. And this, I just got this, this morning, off the web -- pardon me. It's a -- it'll say so at the top -- but this is the current report on clinical testing of this; so testing in humans. Some of them are in -- like if you look at the top one, it's in phase three. In phase one they test for toxicity with humans; obviously if it's toxic, you can't use it as a drug. The next one they test "does it work?" And the third they test whether it's better than current things for doing it, and what the proper dose should be and so on.



So this is phase three trials, and they're recruiting people to -- I'm sorry, I got -- pressed the wrong button -- they're recruiting subjects for it. They already have 1100 people in it. It started in June 2006, and this was updated last May, and it's to -- it doesn't have a completion date given there. This one has a completion date of March 2010. But in fact this is the objective, to see what the quality of life is; tumor response rate; duration of response; how many survive one, two and three years, in this breast cancer test. And it's being compared with another drug, as it says there. So here's what they're comparing in the two different things. Some people do this, some people do this. So the drug, this drug is injected 1.4 mg/m2, IV infusion over two to five minutes, on days one and eight, every twenty-one days. And it's being compared with this drug, which has 2.5 g/m2/day, administered orally twice daily in two equal doses, blah, blah, blah. Okay, so that's the test that's being undergone now. And the estimated primary completion date is September 2011. It's been going since June 2006. So it's going to be five years that they're doing this test, with 1100 people.



So these are really, really expensive things. And you notice it says it's being -- maybe it was on a previous slide -- that it's being funded by industry. Because they hope that it'll work and then they'll be able to recoup their costs and make a profit on the basis of selling the drug afterwards. Okay. If you go to this NIH website, clinicaltrials.gov, you find that they have 64,268 trials in 158 countries going as of today. So these are very, very long, expensive things, to try to get a drug that will work. But making that stuff was really something. But if they get a hit, then the company is on easy street, for a while at least. Like Lipitor. So here's a Lipitor pill, which is the world's best selling drug. In 2004 it sold almost eleven billion dollars. Right? But it raised a 10 billion dollar problem in stereochemical notation, that I'll show to you here. So this is a news report from 2005, dateline New York.



"Pfizer Inc. won a significant victory on Wednesday when a British judge upheld a key patent covering its blockbuster cholesterol drug Lipitor in the United Kingdom. But the medication still faces a similar yet more important case in the United States." (And that was decided in 2006, in the same way, in the U.S. Court of Appeals.) Okay, so, "Judge Nicholas Pumfrey" (in the High Court in London) "upheld the patent covering atorvastatin, Lipitor's active ingredient, but ruled that another patent was invalid. Indian pharmaceutical company Ranbaxy Laboratories Ltd." (a generic chemical company) "had challenged both patents, and was joined by Britain's Arrow Generics Ltd. against the second patent that was ruled invalid. Pfizer said the decision upholding the exclusivity of the patent covering atorvastatin until November 2011 was an important victory for scientists."



And perhaps also for stockholders and executives of Pfizer. Okay? Now, what is patented? Here's the patent that was under discussion. This is the U.S. version. There was also a British version. Okay, so it's patent number 4,681,893, from 1987. Okay? So the important part of a patent is the claims at the end, what's to be protected. And it said, "I claim a compound of structural formula 1." So this compound, okay? And notice that it has codes in here; like X can be any of those groups, linking the right and the left; R1 can be any of those groups; R2 and R3 can be any of these groups. Right? Still going. R4 can be any of these groups. Or it can be a hydroxyl -- instead of being this kind of ester, which is called a lactone, it forms a ring -- it could be the acid that you'd get from that by hydrolyzing it, or a salt of that acid. Okay, so all those things are covered in that patent. How many molecules are covered in the patent? If you made all the permutations of those groups, you'd find that it's greater than the number of protons, neutrons, and electrons in the solar system. So it's obviously impossible to make all the molecules. You'd need quite a few solar systems to have enough particles to do it with. Okay? But the patent is supposed to protect that.



Okay, now. So here -- but what is patented and what was being discussed was not what those groups are. That wasn't in contention. It's known that Lipitor is this stuff: the calcium salt, and it has those particular R groups on it. But that wasn't in question, the fact that it covered so many compounds. What was in question was what the stereochemistry is. Because Lipitor, the drug, is a single enantiomer. Right? Not its mirror image and not a mixture, not a racemate. Okay? But the question is, did this patent cover the single enantiomer? Which is what the picture shows; although it doesn't show it very well. Why not? If you were drawing this structure, on the top right, would you draw it that way? How about that carbon? Does it show its tetrahedron well?



Student: No.



Professor Michael McBride: It certainly doesn't show 109° angles, but it's -- and this dash should be up here, to be even close. How about that carbon? That's even worse, if you were trying to really show the angle, because that one's 180°. But at least it's unambiguous about that versus its mirror image, even though it's not very good. Right? It's not ambiguous. So the structure drawn in the patent was a single enantiomer, and it turned out to be the one that is Lipitor. Okay? But the text of the patent didn't do anything. It never resolved it and made it into one enantiomer. It just talked about it generically. And what it would be making is the racemate, not the single enantiomer. So does the patent cover Lipitor or doesn't it, was the question for the judge. Okay, so this is his opinion. He said:



"In the '633 patent," (when they talk about patents they just talk about the last three digits, not the big number) "it is absolutely clear from context throughout that formula (1) is being used to denote a racemate." (So what they talked about and what they prepared, tested and so on was both enantiomers. Okay?) "In my judgment, every time the skilled person sees formula 1 or formula X,"



Now, skilled person has a very special meaning. It means someone who is "skilled in the art", that knows about pharmaceutical chemistry and so on, knows what these compounds are, how to make them and that kind of thing, but is not creative. Because what you can patent is original creation. So the skilled person knows everything but can't figure anything out; if you see what I mean, can't devise anything new. Okay so the skilled person sees these formulas.



"He will see them with eyes that tell him that in that racemate," (which is described in the text) "there is a single enantiomer that is the effective compound," (So even though what's described is both, the person that isn't very swift, but is knowledgeable, knows that within that there'll be two mirror images, and one of them will do the trick, and not the other one. So you don't have to be creative to know that.) and he will know "that he can resolve the racemate using conventional techniques." (He doesn't have to invent something new, in order to get from that mixture, the one that will be the active one. Okay, so that's not creative. Okay?) "When one comes to claim 1, which echoes the purpose of the invention with its conventional reference to pharmaceutically acceptable salts, he will, in my judgment, continue to see the formulae in this light." (So the salt's the same deal, right?) "In my view, the claim covers the racemate and the individual enantiomers."



So the person without any creativity knows that yes, the patent described the racemate, but there'll be one in there that I can get, and that'll be the useful drug. And it doesn't take anything great to get there. We'll talk about that soon, how you would do that; why it's well-known. Okay? So this is good news for Pfizer, right? Their patent holds. They put out a press release saying this is a victory for science. Okay? But it's in a sense a pyrrhic victory, because they had a later patent, on the single enantiomer. Right? And now that patent is invalid. Because if the first patent covers it, you can't just every three years come up and say, "I patent this again," "I patent this again," I patent this again," and keep getting another fifteen years of exclusivity. Right? So once they were covered by the first patent, the second patent is invalid. Why do they care, if they're already covered by the first patent? Isn't that good news? Dana?



Student: Their fifteen years of exclusivity will run out.



Professor Michael McBride: Ah, they're going to run out. So this means that three years earlier their patent is going to run out -- right? -- at four or five billion dollars a year. That adds up. Okay? Now, so drug -- so there are conventions for how you name things, including stereochemistry. In fact, there was a stamp issued in Switzerland in 1992 commemorating the 100th year of the Geneva Convention, where chemists got together, once they knew stuff about how things were connected. Then they had to agree on how to describe it, on what the nomenclature and the notation should be. So the International Union of Pure & Applied Chemistry continues this. And you can go to these websites and learn about what the rules are for notation. So they had to devise rules for what you want to do when you make a name for a compound. So what is maybe the first thing you want to do? Nate?



Student: You need to --



Professor Michael McBride: No, not you.



Student: The other Nate?



Professor Michael McBride: Well let's see. Let's try somebody else. Andrew? No, no. Kate? Kate?



Student: [inaudible].



Professor Michael McBride: Kate, what was the first rule?



Student: They want the name to discuss composition, like which atoms are in it.



Professor Michael McBride: Well -- I didn't mean you anyhow. So what's the first rule about names?



[Students speak over one another]



Professor Michael McBride: John? Not you. [Laughter] Other Kate?



Student: They should describe only one.



Professor Michael McBride: Ah, they should be unique. There should only be one compound that has that name. You shouldn't say, "Take this," and they take the wrong John or the wrong Kate or the wrong Andrew. Right? So the very first thing should -- well it should be -- Kate, you were right first. It has to be clearly descriptive of what the compound is. And so it has to tell composition, constitution, configuration and conformation maybe, if you care, i.e., stereochemistry. And it can be like this. E7389 was the name of that compound. Forget trying to name it systematically, according to rules. Right? E is for Eisai, the name of the pharmaceutical company, and this is their 7,389th compound. Right? But at least everybody knows what that means. Right? Or you can look it up. But this is what we were just talking about. It has to be unambiguous. Right? One name must mean one structure, not some other structure as well. Okay, so unfortunately "amide" doesn't fit that bill, because it means the anion of an amine, but it also means that functional group we're talking about. It's the same name. You have to figure out from context which one you're talking about.



It also has to be unique. One structure is one name. So everybody will have the same name for the compound. Right? Which is more important, that it be unambiguous or that it be unique? Like you could have two names for the same compound; like bromobutane and butyl bromide. Right? And if you put a number in, you would know which -- where the bromine was. Okay? Or bromoethane and ethyl bromide. Two different names, right? So the names aren't unique. Right? But they're unambiguous. You know what compounds you're talking about. Why would you care whether you have just one agreed-on name by the people who get together in the International Union of Pure & Applied Chemistry? Why would care? What's wrong with having two names? Anybody see any disadvantage to having different names for the same compound? Sam? I could've called Sam too. Yeah, go ahead.



Student: It's hard to communicate about the stuff.



Professor Michael McBride: Not if everybody knows what you're talking about. Shai?



Student: Well like if something has a lot of names, eventually you're going to get to a point where someone doesn't know. You can say -- you're going to say a name and they're not going to know what you're referring to.



Professor Michael McBride: Or if you have to look it up, you have to choose the right name to look it up. So this was more important, when you didn't have computers to do the work for you, when you had to look in an index and get -- and be looking for the right name that everyone agreed, and the person who made up the index entered it there. Nowadays you can draw a structure often on a screen, and the computer will figure out what the official name is and then look it up. But anyhow, that -- so it's not as important that it be unique as that it be unambiguous. But if you're looking things up, it's very nice if you know that everybody will have named it a certain way. So you can also index things according to their composition, CHN, or using computer graphics, as we say. And, if possible, the name should be manageable. Right? It should be possible to write it fairly quickly. It should be easy to figure out; short, if possible, and pronounceable, if possible, so that you can talk about it.



Now let's look at systematic constitutional nomenclature, for this sort of intermediately complex compound. Compared to these others we've looked it's not much. Okay, so the first thing you do, these people decided, when you get together, is decide what the main chain of carbons is. Because that will provide the root with a Greek designation that says what the number of things is. So we have to find a carbon chain, and we'll find the longest carbon chain. We could take this one, or we could take that one. Which one is longer? The one, it could be an octane or it could be a nonane. Which one should we choose?



Students: Nonane.



Professor Michael McBride: Well we could've decided you take the shortest one. But that's sort of silly. There'd be a zillion shortest ones; or short ones, right? Or even shortest. The shortest is one carbon I guess. But there's only one longest one, right? Wrong. Because there's also that one. Same length. So now we have to decide when we're going to take the longest; how are we going to choose, if there are two of them? Well what you choose by is the number of substituents. Okay? Because that one has one, one, three, four, five substituents, and this one has four substituents. Which will you choose, the one with four or the one with five? You have to make the rule up. five or four?



Student: Five.



Student: Four.



Professor Michael McBride: Somebody said four. Why? Sophie you said four?



Student: Less naming to do.



Professor Michael McBride: There'd be less naming to do. Ah ha. But in fact they chose the most substituents, and you'll see why shortly. Okay, so you choose the most substituents. So you choose the one on the top, not the one in the middle. Now why? Sophie's not happy with this. But you'll see soon. Now you have to number the carbon atoms, to say where the substituents are. Now, we can start from either end. Right? If we start one, we could go one, two, three with green and continue; or one, two, three with red, from the other end, and continue. Now how are we going to decide which way to go? Kevin?



Student: Right. So you want in this case the Cl and the Br to have the lowest -- to be associated with the lowest number of carbons on the chain.



Professor Michael McBride: Okay, there are two ways to do it. You could make it so the overall sum of the numbers you're going to get -- because you'll get different numbers, according to which end you start from. You could sum them all up and see which one gives the lowest sum and choose that one; or the highest sum you could do. Or you could see which one gives the lowest number at the first number. So this, if you chose the red, you'd have 2-chloro. If you chose the green, you'd have 3-bromo. Right? And the other numbers would all be higher. Which one do you want to do, use the lowest sum of all the numbers, or the lowest number at the first difference? Lucas?



Student: Lowest number at the first difference.



Professor Michael McBride: Why?



Student: So you can actually build it as you go, instead of just saying --



Professor Michael McBride: Yeah, so you don't have to do all the work of going through the whole thing and adding them all up and so on. You just go 'til you get a difference and when you get a difference that's it. That's much quicker. Okay? So we'll go to the first difference, lowest number at the first difference. And now we have to say -- now we know we have the chain and its numbers, and now we say where the various substituents are. But first we have to name the substitutents. Now in the top you can see we have methyl, bromo, methyl, ethyl and chloro. Now had we followed Sophie, we would have to name that one, and that's not as easy to name. So the fewer substituents you have, the harder they are to name, because they get more complicated. Right? So that's why you choose to have the most substituents, so they'll be the easiest to name. You have to do more but it's easier. That particular one is called 1-chloroethyl. But why have to name complicated things if you can name simple things? Okay? So you choose the most substituents in order to get simpler names.



Then you alphabetize them. You don't put them on by order of their number, you alphabetize the names of the groups, and you count. So that compound is 7-bromo- (b is first) 2-chloro- (c) 3-ethyl-6,7-dimethylnonane. Now you might say dimethyl should be alphabetized by d, not by m. But that's not the way it's done. It's done by the group name and then the prefix that tells the number doesn't enter into the alphabetizing, usually; although some people do it the other way. Okay, so that's the question about the d. Okay.



Now, if you go to this website you can get help with nomenclature. Here's a compound that it goes through. And the minute you look at this and see the name at the bottom, the compound name, 4,5-dichloro-2,4-chloro-2-hydroxymethyl-5-oxo-hexyl-cyclohexane-1-carboxylic acid, you say "Wow!" And aren't you glad there are computers that can do this for you? But you can go through it and figure out why they did it that way. But that's sort of a parlor game. It's not fundamental chemistry. Okay? But there's some very useful non-systematic names for simple groups: isobutane, isopentane, neopentane. So you can have an isobutyl group or an isopentyl group or a neopentyl group, isopropyl, secondary-butyl, tertiary-butyl, neopentyl. So these names you just have to learn. Right? They're not systematic. And there's a nomenclature drill available on the course website. So it's good to know these so we can talk to one another. But in principle the systematic name is the way to go. Okay, that's it. But we haven't gotten to stereochemistry yet. That'll be tomorrow, or Monday.



[end of transcript]

Lecture 28
Stereochemical Nomenclature; Racemization and Resolution
Play Video
Stereochemical Nomenclature; Racemization and Resolution


Determination of the actual atomic arrangement in tartaric acid in 1949 motivated a change in stereochemical nomenclature from Fischer's 1891 genealogical convention (D, L) to the CIP scheme (R, S) based on conventional group priorities. Configurational isomers can be interconverted by racemization and epimerization. Pure enantiomers can be separated from racemic mixtures by resolution schemes based on selective crystallization of conglomerates or temporary formation of diastereomers.




Transcript



November 10, 2008




Professor Michael McBride: Okay, today in the Times there was a report of a clinical trial, which event, the kind of thing we talked about on Friday. This was of a statin, which is like Lipitor. It's a different one, it's Crestor. But they just reported that a study of 18,000 people -- so a very big clinical trial -- seemed to show that this, that it helped; it reduced the risk of heart attack by a half for people who didn't have high cholesterol, but had another risk, this C-reactive protein, CRP. And it said in the story this morning: "Like many clinical trials, the Jupiter Study was sponsored by a pharmaceutical company, in this case AstraZeneca. It makes the drug in the trial, rosuvastatin, which is sold as Crestor." And then it says: "Dr. Ridker, a co-inventor of a CRP test, said he first sought federal financing for the study and was turned down. He and other scientists interviewed for this article, except for Dr. Nabel, Dr. Gardner and Dr. Wolfe, have consulted for or received research money from stain makers." I guess I should say that I've occasionally gotten money from statin makers too, but not for this kind of thing.



Okay, so that's interesting and relevant to what we're talking about. But specifically what we're talking about now is nomenclature. And last time we talked about constitutional nomenclature, the rules people made up for naming the nature and sequence of bonds. But since arrangement in space is also important, you have to be able to have names for stereochemistry. And right now we're interested in configurational stereochemistry, the kind that can only be changed by breaking bonds -- right? -- whereas, as we'll see soon, conformational changes don't require breaking bonds, so typically are much easier to do.



So configurational isomers include the tartaric acids that we've been speaking about, that have these physical properties. One has a unique melting point. The other two are the same. But those two differ by the sign of the optical rotation of polarized light. So the first one is meso. Now how do we name these things? One way is just to name them according to the phenomenon, to what you observe. So sometimes those two tartaric acids that are enantiomers -- mirror images of one another but not superimposable -- sometimes they're referred to as small d, or small l, which stands for dextro, or levo, and means right or left. And all it means is the direction they rotate the plane of polarized light, which you observe experimentally. So there's nothing more arcane about that. Or you can say plus or minus, meaning that they rotate light to the right or rotate light to the left. And those are absolutely the same thing, right? Dextro and plus; levo and minus. Sometimes people use both names, but they're redundant. They both mean it rotates light one way or the other. So it just describes the phenomenon.



So there's nothing ambiguous about that. It's a perfectly good name. But it doesn't tell you what model you should make to show it, which mirror image is which. All you know is which way it rotates light. So that's phenomenological. But there can be other names that have to do with these Fischer projections we draw, as to exactly which one is which. So that one is clear enough. That one has a mirror image -- is its own mirror image. There's a mirror that goes through the middle of it. So the top is the mirror image of the bottom. It's a mirror image of itself. And that one's clearly the meso. Right? But how about the others? One of them will look one way and one the other. But is there any fundamental thing that we can understand that would tell us which will rotate light to the right and which will rotate light to the left? And I'll tell you that I can't say that, and most people I know -- in fact almost everybody I know -- can't say that. But there are a few people who think they can say that, and we'll hear about one of them soon, on Wednesday. But anyhow it's a question. Is it like -- should it be those for the right and the left-handed, the d and the l, or those? Which one is which?



Well Fischer decided that we just have to have to names for this and we have to be able to draw pictures, so let's make a guess as to which way it is. And he devised a new system of naming which doesn't depend on the phenomenon but depends on genealogy; that is, synthesis. You take a compound like glyceraldehyde, which is small d and +. What does that mean, d and +?



Student: The same thing.



Professor Michael McBride: For this? There's a particular bottle of glyceraldehyde, which has that constitution. Right? And we're going to label it d-(+). And what does that mean? Zack?



Student: Does it rotate light to the right?



Professor Michael McBride: That it rotates light to the right. But we don't know exactly how we should draw it -- right? -- the structure or its mirror image. So, and d and + here, remember, are redundant. They both just mean it rotates light to the right. But Fischer went further. He defined that particular glyceraldehyde as capital D. There's a different between capital D and small d. Capital D is the name he used, and that big D-glyceraldehyde, he guessed -- the one that rotates light to the right -- he guessed had that structure, which when drawn with a Fischer projection is that structure. So that's just a guess. So he said this now gives us a starting point. This glyceraldehyde, which I'm going to call capital D, and I'm guessing has that structure -- I might be 100% wrong, right? It might be the other one. But let's just make this guess and then we'll be able to talk about everything. So what could you talk about? You could do a multi-step synthesis in which the carbons connected by the arrows are the same carbon. That is, you redo the top half of the molecule by doing chemical reactions, the part that's in green. So CHO becomes that much more complicated thing. But you leave intact that carbon on the bottom. So it's the same in tartaric acid as it was in D-glyceraldehyde, whatever that was. And we'll guess it's this one, with the OH on the right. Is everybody with me, with the guess he made? So now if you can make that particular tartaric acid isomer from D, capital D, glyceraldehyde, then you say that that's D-tartaric acid, because you can make it from D-glyceraldehyde. Any questions about that? So it's just what you make from what. If you can make it from a D-compound, you call it a D-compound, and the mirror image would be the L-compound, capital L. Claire?



Student: Is he just swapping the things in the green squares? The arrows are a little confusing.



Professor Michael McBride: The swap -- the thing that's in the green square on the bottom becomes the stuff in the green square on the top. Obviously you have to add another carbon and you have to add a couple of hydrogens and some more oxygens. So you have to do some chemical transformations. But the important thing is that the thing on the top becomes the thing on the top, and the carbons connected by the arrow are precisely the same carbon.



Student: Right.



Professor Michael McBride: Okay? So that one you know because you know where it came from by genealogy. So that tartaric acid you'll call capital D, according to Fischer. But do you see a weakness in this nomenclature scheme? Yeah, Sherwin?



Student: It's dependent on the structure, your knowledge of the structure that he showed.



Professor Michael McBride: Well of course it could be 100% wrong. Everything could be the mirror image of what we thought it was. Right? That's clearly a problem, and that was well recognized by Fischer. But maybe I didn't understand your question -- did I?



Student: Yeah, that's what I was saying.



Professor Michael McBride: Okay. Does anybody see another weakness? Sam?



Student: Capital D isn't very descriptive.



Professor Michael McBride: It's certainly not descriptive in structure, except -- unless you know what d-glyceraldehyde is, then it's descriptive. But it's not descriptive alone. It's only relative, right? It's only relative to the d-glyceraldehyde. And you have to know how that relation works. For example, someone else might go into a lab and they want to prepare tartaric acid from d-glyceraldehyde. What might the other person do?



Student: Prepare it a different way.



Professor Michael McBride: Lucas, I can't hear.



Student: Prepare it a different way.



Professor Michael McBride: Prepare it a different way. So somebody else might put that carbon into the one on the top and change the brown ones. Right? There could be other transformations you do that would change the carbon on the bottom to the carbon on the bottom, and the carbon on the top is now the same as the one on glyceraldehyde. Right? In which case it would be exactly -- that one on the left would be named capital D, if you made it that way. Everybody see? The genealogy would just be different. Okay? So there's a fundamental problem here, that the capital D / capital L designation is ambiguous, without having some detailed synthesis recipe, or some agreement that people say we'll always make things this way rather than another way. Right? So it's not -- doesn't have meaning on its own. You have to know exactly how you made the transformation. So that's a pretty cumbersome way to do names. But that's what was done for sixty years. Right? Those were the names, capital D and capital L; for sugars, for example, for amino acids. Okay? And it's relative, and it's based on a guess.



But in 1950 things changed. Because in 1950 a man named Bijvoet, an X-ray crystallographer at Utrecht -- in fact working in the van't Hoff Laboratory, the same building in which van't Hoff did the things in 1874 -- Bijvoet determined the crystal structure of sodium rubidium salt -- remember, there are two carboxylic acids in tartaric acid. So the sodium rubidium salt of d (small d)-L-tartrate. What does the small d mean? Andrew?



Student: It reflects the light of the --



Professor Michael McBride: It rotates the plane of polarized light, or twists it to the right. And what does the capital L mean? Allison?



Student: Previously we said it rotated it to the left.



Professor Michael McBride: I couldn't hear very well.



Student: That previously we assumed that it would rotate it to the left.



Professor Michael McBride: No, that's not true. Or maybe I didn't understand "previously," what you mean.



Student: Like well, the model we were just talking about with Fischer.



Professor Michael McBride: Right, it's what we talked about with Fischer, that he was able to make this tartaric acid from l-glyceraldehyde, from the enantiomer of what we talked about before. So it was related by that particular synthesis to l-glyceraldehyde. So there's no conflict between saying d and L. Right? It rotates light to the right -- that's the small one -- but it's genealogically related to l-glyceraldehyde. Okay? But he was able to use a very special kind of X-ray technique, called anomalous dispersion, which we don't have time to talk about right now. But it was able to show exactly which way the atoms were oriented. And this is the picture he drew. Right? So clearly on the top is C, two Os, right?; a carboxylate group. On the bottom a carboxylate group. The middle two carbons each have an H and an O on them. The O is stippled, so you see what it is. And it's clear the way you draw it what's coming out -- the way he drew it, what's coming out toward you and what's going back in.



Now here's what Fischer had guessed, sixty years before, for L-tartrate, what he called L-tartrate. Was Fischer right or wrong? How about the top carbon that has OH on it, the HCOH? Is it the same as in the model that Bijvoet published? Does it agree or not? Yes, it agrees. The OH is coming out toward you and to the right, same as in Fischer's picture of the top carbon. Everybody got that? How about the bottom carbon? The bottom carbon, Fischer has it going to the left and Bijvoet has it going to the right; this one here. Right? That's going to the right from this carbon. Fischer's drawing it to the left. Do they disagree?



Student: No.



Professor Michael McBride: Who said something? Do they agree or disagree?



Student: Agree.



Professor Michael McBride: Seth, what do you say? What does Fischer mean by this bond here, the CH bond? Which is it, going out toward you, the H, or back in? Seth? The Fischer projection, which is shown here. It's got vertical bonds and horizontal bonds. Which one is coming out toward you, from the carbon? Okay. Somebody else? Yeah, Kate?



Student: In a Fischer projection the bonds are going into the board.



Professor Michael McBride: Which bonds?



Student: The carbon-carbon bonds.



Professor Michael McBride: The carbon-carbon bonds are going into the board; that's Fischer's convention, yeah.



Student: Because he twisted them so that everything else would be coming out of the board.



Professor Michael McBride: Yeah, so these two are coming out of the board. How about in Bijvoet's picture? For the second carbon -- here's the first one, here's the second one.



Student: The OHCs are going into the board.



Professor Michael McBride: Ah ha. So, to make it like Fischer, you have to rotate it so they're coming out of the board. So in fact they do agree. Right? It's just that Bijvoet's is made realistically, and Fischer's was made with these rubber tubing bonds and you had to bend it to get that way. But it's the same configuration. So Fischer was right. So these structures that had been drawn for sixty years were drawn correctly. Right? They could've -- it was a 50:50 chance that it could've been the other way around. But thank God he was right, so we don't have to go back and correct everything, and know whether things were written before 1950 or after 1950. Yeah?



Student: But don't you think that if you rotate that molecule, isn't the OH that's coming, that's like each -- his drawing coming out of the board, doesn't that go back in?



Professor Michael McBride: Okay, if you rotate this molecule by 180° around the vertical axis, then this carbon -- with respect to the carbon we're interested in here -- this carbon would be going back into the board and this carbon would be going back into the board. That's the way these are. Right? But in the process of that rotation, this OH, that was back and to the right, after 180° rotation about the vertical axis, will be out in front and to the left; which is the way it's drawn in the Fischer Projection.



Student: So the other OH group would be going back in?



Professor Michael McBride: Yeah, but you don't care. This is just to show the configuration, right? When you do that rotation, yeah, these will do the other. But the point is that Fischer's structure was right, with respect to each carbon, about what's going which way, which tetrahedron it is. Okay, so that's fine. And then Bijvoet wrote, or in the same paper he wrote: "The question of nomenclature is beyond the scope of our investigation." (He's just doing the X-ray to find out which way it is.) "The problem of nomenclature now concerns given configurations" (now you know which one it is you want to show) "and requires a notation which denotes these configurations in an unambiguous and if possible self-explanatory way." So you don't need to know how it was synthesized or anything. Now you know how it really is -- right? -- what name are you going to give to it so people can know, just from the name, which way it is.



Now, this is related to the idea of naming configurations involving double bonds. So we're going to first show you how you do that -- which actually came second -- and then we'll go back to see how they named the configuration at a tetrahedral carbon. So malic acid, from apples, has that structure shown, and if you heat it you lose water and get a double bond. But you get two isomers, maleic and fumaric acids. And you'll remember we showed that before, that van't Hoff made those with his models. And one of them has the two COHs near one another and one has them farther from one another. And we can draw them that way as well; near on the left and far on the right. And if you heat these molecules further, the one on the left can lose water and form an anhydride, but not the one on the right. Why not the one on the right? It's pretty obvious. Russell?



Student: They're too far apart.



Professor Michael McBride: They're too far apart. So you know which one they're close together and which one they're far apart. So that experiment proves which is which. It's not as hard as the case of the right and left-hand, where you had to do this very special X-ray technique. So it was known that maleic acid was the one that can form an anhydride, and therefore had them on the same side. And that was called cis, meaning on this side of, and the other one was called trans, across. So that was a perfectly good nomenclature for those two isomers of the dehydrated malic acid. Maleic was cis and fumaric acid was trans. So these names were used for a very long time, but they weren't really good names. And the reason, you can see, by looking at this molecule. Is this one, would you call it cis or trans? Same side or opposite side? Andrew, what do you say?



Student: I don't know.



Professor Michael McBride: Next Andrew. This is our nomenclature problem.



Student: I think trans.



Professor Michael McBride: Would you call it trans, you say?



Student: I would probably call it --



Professor Michael McBride: This one here.



Student: Okay, that one, I would call it cis.



Professor Michael McBride: Why?



Student: Because of the CH3's.



Professor Michael McBride: Because the CH3's are next to one another. So you might call it cis. On the other hand, if you wanted to say how this end is related to the acid, you'd say it's trans, right? So that means it's -- as an absolute nomenclature, it's very hard to generalize this because you have to know which one you're picking, to say it's cis or trans. It's fine to talk about the relative configuration, to say the methyl on the left is cis to the methyl on the right, or to say the methyl on the left is trans to the COOH. That's fine. But as to whether to say the molecule itself is cis or trans, you have to know which one you're talking about, on the right. So that's another thing where you're going to have to have some convention or something. And the way people decided to do it was this. You assign the groups at each end, take one end and then the other end of the double bond, and assign priority. So you say we're going to give higher priority to one group or the other, and that's the one we're going to use for the name. So on what basis might you assign a priority to a group; to say which group has higher priority, CH3 or H? Which one do you think should have higher priority?



Students: CH3.



Professor Michael McBride: CH3, okay. How about on this end, CH3 or COOH, which should have higher priority?



Students: COOH.



Professor Michael McBride: Why?



Student: It's heavier.



Professor Michael McBride: It's heavier. Okay. Now let's think about how that is going to work out. So we'll assign them by atomic number; or atomic weight, if we have two that are different isotopes of one another. Right? But we won't sum up the atomic numbers of all the groups that are in that; that is, we won't sum up C and O, and O and H, and C and 3 Hs. Why not sum them all up and then see which one has higher total atomic number? What's wrong with that? Perfectly well-defined.



Student: It's tedious.



Professor Michael McBride: It's tedious, it's cumbersome, to have to add it all up. So what rule are you going to do?



Student: Find a difference.



Professor Michael McBride: You'll go until you find a difference, go out until you find a difference. Once you find a difference you stop, so you don't have to go everywhere. Okay. So at the first difference; the same as we said with numbering last time, right? Okay, so on the left, in each case, you have a carbon compared to a hydrogen as the thing that's immediately attached. Carbon has higher priority, higher atomic number. On the right, it begins with a tie. Incidentally, we're going to have to deal with double bonds, COOH, at the bottom. Right? The way you do it, just by convention, there's nothing -- God didn't write this on the tablets, right? It's -- you just pretend that there are two of those atoms, when there's a double bond; just pretend.



Okay, so now we compare the top with the bottom. If we go to the first step, it's carbon versus carbon. It's a tie. So we're going to have to go further. When we go to the next level we see on the left it's oxygen versus hydrogen. Right? And oxygen is higher. We don't have to go any further, we've found a difference. So forget the rest of it. Everybody with me? Okay, so we might say the one on the left is trans and the one on the right is cis. However, people, for sixty years, or longer in this case, have been using the names cis and trans to mean one thing or another, and sometimes we'll agree with them and sometimes we don't; which is going to cause great confusion as to whether you're using a pre-1950 name or a post-1950 name, and at least half the compounds are going to be wrong. Right? So the names trans and cis have been polluted by previous usage. So we're going to have to have new names. Yeah Lucas?



Student: You don't really have -- in the second part of the double bond, as a substituent, and that itself is like cis or trans.



Professor Michael McBride: You'd mirror all -- you'd duplicate all those as well.



Student: Okay.



Professor Michael McBride: Okay. Okay, so this one was used with a German root, because the guys from Switzerland were doing this, entgegen, which means opposed. And the one on the right, so it's E when they're trans, when they're opposite one another. And when they're adjacent to one another it's called together, zusammen, Z. I've always thought this is a real pity that it wasn't the other way around. So I hope this doesn't confuse you. Because these are on the same side and these are on opposite sides. Right? But it's just backwards from that. So just remember that it's unfortunate. Okay? So Z means together and E means apart. Okay, that's the name we use for there.



Okay, now notice that in assigning priority, as we said, just to belabor this is a little bit, you proceed one shell at a time. So we saw a tie when it was carbon versus carbon, and when it was three oxygens versus three hydrogens, it was obvious that the top wins. But now let's go to this case. Okay, now carbon versus carbon, that's a tie. We go out further, three oxygens beat a carbon and two hydrogens. Right? But there are chlorines on the bottom. Right? The chlorine is really high priority. Right? But it's irrelevant, because the decision has already been made. So you respect earlier decisions. Right? You only go as far as you need to go, to see a difference. Okay, I think we've said that enough.



Now these are the guys that invented the scheme to name handedness. It's called the CIP Priority Scheme. And the C is R.S. Cahn, and the I is C.K. Ingold -- both of them are from Britain -- and the P is Vladimir Prelog, who's in Switzerland, although he's a native of Yugoslavia. Now, these were -- many people, particularly synthetic chemists, would consider Robert Robinson and R.B. Woodward the greatest organic chemists of the 20th century. They both thought they were probably the greatest individual one. Right? So I love this picture, taken by Jack Roberts of them at a seminar at MIT around this same time, because they sort of aren't looking at one other. Right? What's Robinson looking at? Robinson is looking at Ingold, whom he really despised, and vice-versa. All right? There was no love lost between Ingold and Robinson. And this is illustrated by an account that Prelog gave in his autobiography. He encountered Robinson in the airport in Zurich. Robinson was on his way to a meeting in Israel.



Robinson: "Hello Katchalsky. What are you doing here in Zurich?"



Prelog: "Excuse me, Sir Robert, I am only Prelog; I live here." (He got a Nobel Prize too, Prelog.) "I am only Prelog and I live here."



Robinson: "You know Prelog, your and Ingold's configurational notation is all wrong."



Prelog: "Sir Robert, it can't be wrong. It is just a convention. You either accept it or not."



Robinson: "Well then if it's not wrong, it's absolutely unnecessary."



Okay? So anyhow, the point of this slide is that it's a convention. There's not right and wrong about it. It's that they propose rules, and those are the rules that people have adopted. And that's how you give a name to the absolute configuration of a stereogenic carbon. And for fun, for Wednesday, you can try this exercise, going back to this medieval manuscript, and using the figures in it to devise your own convention to describe chirality. So you can put yourself in the position of Cahn, Ingold and Prelog. Okay, so here's the Cahn-Ingold-Prelog R/S nomenclature scheme for stereogenic centers. First you have to decide on each of these carbons -- which are mirror images, you see, of one another -- which atoms have which priority. So let's start on the left. What's the highest priority for a substituent on the central carbon? There are four groups. Which has the highest priority?



Student: OH.



Professor Michael McBride: Which one?



Student: OH.



Professor Michael McBride: Zack?



Student: I thought it was OH.



Professor Michael McBride: OH. Yeah O is the highest atomic number. What's the lowest?



Student: H.



Professor Michael McBride: H, right? There's also a D there. It's the same atomic number; hydrogen and deuterium. Which one should be higher?



Students: Deuterium.



Professor Michael McBride: Deuterium's heavier. If it's isotopes, you take the heavier one. Okay, so here's one, four, three, two. Right? That's the priorities; and analogously on the right. And now we have to decide whether they're -- which one to call right and which one to call left. And the way you do it, one way of doing it, is to take a thing like this where I have these red, yellow, green, blue; red, orange, yellow, green, blue, right; the order of the spectrum, red, yellow, green, blue. And I make a spiral that connects them, like that. And now, is that conventionally what you call a right-handed or a left-handed spiral? You know, if you look at it, if I turned it like this, it would move this way. Right? So screws are called right-handed and left-handed screws. And the reason is that the -- you know, toe-bone connected to the foot-bone kind of thing. The way your arm is put together, for right-handed people it's easier to drive a screw like this. Right? So that's conventionally called a right-handed screw. So this you would call right-handed. Okay?



Now, if you can't remember from how to turn a screwdriver, you can look at -- you can do this trick with your hand. You put your thumb along the direction of the lowest priority; hydrogen, in this case. So it's coming out of the board, right? And you curl your fingers. Right? And you notice it goes one, two, three, and then four. So that one on the left is left-handed. Everybody see how I did that? Curl your fingers, they go one, two, three, four. Okay? So that's a left-handed one. The one on the right, if you do it this way, you -- so that's. And you call it, not left, you call it S, for the Latin sinister. So it's not going to get confused with D and L and so on, that have already been used. It's a new way of doing it. So that one's S. And the one on the right, if I put my thumb so it goes back toward the H, and curl my fingers, I go from one to two to three. Okay? So that one is called R for rectus; Latin for right. Okay?



Now there's another -- there are lots of tricks to do this. If you don't like doing it that way, you can do it this way. Pretend that this is a steering wheel of a car, and the H is going back in. Right? So here I'm -- this one on the right -- I'm sitting here driving a car like this and the H is going away from me. Everybody got that? And I notice that I go one, two, three. Right? That's going to turn the car to the right. Everybody see that? So, and this one will turn the car to the left, if I turn the wheel to go from O to C to D. Okay? So that one's a left turn and that one's a right turn. So whatever works for you is fine. And I can assure you that on some test I'm going to look out and see people with their papers there and so on, and they're going to be going like this. Right? Because it's a handy way to do it. [Laughs] Handy. [Laughter] Okay, so let's try it on tartaric acid. Here, let's do the top carbon of tartaric acid. Okay? So we've got to decide on the priority of the groups here. Here's one group, another group, another group, another group. Which is the highest priority?



Student: OH.



Professor Michael McBride: OH is the highest, right? O is the highest. Now we got two carbons that are tied. Which one's going to win, top or bottom?



Student: Top.



Professor Michael McBride: Top you say Kate? Why?



Student: A tie.



Student: Because it has -- well we can pretend three oxygens, whereas the bottom one has a hydrogen, a carbon, and an oxygen.



Professor Michael McBride: And only one oxygen, right? So the one on the top has -- not more oxygens, but it's got the oxygens nearer. Okay? So we don't have to go as far. So this is going to be one, two, three, four. Okay, now let me see your hands operating. Put a thumb along the H, the direction the H goes, and curl your hands one, right, top, bottom, one, two, three, and tell me whether that's right or left-handed? Okay? How many think left? How many think right? This is a democracy, the rights have it; they're also correct. Okay, that one's one, four, two, three. Okay, that, it's right-handed. And the bottom one's also right-handed. Why didn't I have to take a lot of time to have you do this, to do the bottom one? Why does the bottom one follow, if you know the top one? Because you can rotate the thing 180° and it's the same, and that will change the top into the bottom. So whatever the top is, the bottom's the same thing. So the name of this is, in parentheses, (2R, -- carbon two is R; carbon three is also R), then dash (-), 2,3-dihydroxybutanedioic acid.



Okay, so now we have a scheme that does that. And notice, this is the gate into the new Chemistry building, just up Prospect Street, and it's got things that relate to the different branches of chemistry on it. And this one is those carbon tetrahedra with the spiral on it that says which handedness it is. There are both right and left-handed ones there. Okay. Now, racemization. If you start with a molecule, which maybe you've gotten from nature, that's all one hand, like lactic acid, from milk, (R)-lactic acid, there are processes which convert it into a 50:50 mixture of R and S. For example, suppose you had a base and it could pull off a hydrogen. What hydrogen would it pull off from this molecule, a base? Any idea? What is the most reactive group here?



[Students speak over one another]



Professor Michael McBride: Or another way of saying it, this is a high HOMO. What's the low LUMO going to be? Well there's several low LUMOS. There's the CO double bond, π*. That's a good one, right? However, it won't get you anyplace. You can put it on, come off again. Okay? Or there's σ* OH here, but there's also σ* OH here. But this one's better because it's next to the π*. Right? So this -- that's why it's called an acid. You already knew that. And you can lose the H+ and get down to this. Okay, and that'll go on off, on off, on off, a long time. That's fine, nothing special there. But once in a long while -- that's easy to do -- once in a long while you might pull off the wrong hydrogen; you might pull off this one. Right? So that's not nearly as easy. But when it happens it's interesting, for two reasons. One is, it's not as bad as you might've thought, because it's adjacent to this π*, which means that the vacant orbital here can stabilize that high HOMO. Right? Which means you can draw that resonance structure, with a double bond there. So it's hard, but it's not as hard as it might be.



Once in awhile that'll happen, and usually it'll go back again. Right? But what's especially interesting -- well just as a footnote, it's possible to pull both of them off. But that would be very unusual, to make a di-anion, which is high in energy, because you're putting so many electrons close to one another. Okay? But what's relevant for our purposes is that that molecule is planar. So it's not handed. This one had a chiral carbon in it, but this one doesn't have an asymmetric carbon, a stereogenic carbon. So this one, the one in the middle, is neither right nor left. So if you go from right to something that isn't right or left, you can come back, put the hydrogen on. The hydrogen comes from the front of the screen, goes on to where the minus charge is, and you're back to the left. But the hydrogen could've come from the back of the screen, at the same carbon, and that would give the one on the right, and would be (S)-lactic acid. So after a long time of heating, with base, you can take one that's all R and have it become 50:50, R and S. And that process is called racemization. Why would it be called racemization? Because it makes a racemic mixture; a 50:50 mixture from something that was pure. Okay, that's interesting.



But what may be more interesting is -- well this is just a footnote on that. If you here, to go from (R,R)-tartaric acid, to meso-tartaric acid, you change just one such hydrogen. The one here has been taken off and put back on the other side, to give that one. So that's not racemization, you're not going to a 50:50 mixture of the two hands, right? So it has a different name. It's called epimerization. That's just for vocabulary. But what's more interesting than racemization is the reverse, to start with a 50:50 mixture and go to all one. And that's the main subject for the rest of what we're going to be talking about, before the exam.



Okay, so you start with (R,S) and you separate it into R and S, in separate vials. So you have just the one you want. Okay, one way we've already seen of doing that was Pasteur. A conglomerate is a mixture of crystals where each crystal is all one hand or all the other hand; but it's a 50:50 mixture of such crystals. So there's no net handedness to it. Right? But then if you, if you have a situation like that -- notice you wouldn't have that, if the crystals had a right-handed molecule, and then a left-handed, and then a right-handed, and then a left-handed; you couldn't pick it apart because you can't separate one molecule from another. Right? But if all crystals are right-handed, and some crystals are all left-handed -- right? -- then in principle you can pick them apart. And that's precisely what Pasteur did when he noticed it. Right? So you can separate them. And there are tricks that can help you do it even better, or more easily, when the crystals -- when the things crystallize as all right-handed or all left-handed. One is to start with molecules in solution and have only one form crystallize and have the other one stay in solution. Then all you have to do is filter it. Okay? And how can you do that? It's because -- and chiral-resolved poison is another one, which will keep one form from growing, so that only one form crystallizes.



How do you do that? Here's a crystal which is in solution. And suppose it's exactly at equilibrium; so it neither dissolves nor crystallizes. It's exactly at equilibrium, it's saturated. Okay, now the interior molecules of the crystal are very stable, because they have exactly the right environment. Right? They're more stable than the ones in solution. The ones on the surface are not so stable, because they don't have the neighbors -- all the neighbors they need. Right? But the average molecule is exactly the same as in solution. So it just sits there, it's at equilibrium with solution. Now, so that's the situation we have. That's the temperature and concentration, so that this crystal is exactly at equilibrium with solution. Now suppose that you had a smaller crystal. Would it also be at equilibrium with solution; if it were smaller? Or do you see any difference? The smaller one has a higher fraction of molecules on the surface. Okay? What does that mean? The molecules on the surface are less stable than the ones inside. If the average molecule here is precisely the same as the ones in solution, so it just sits there, how about this one? The average molecule is less stable than the ones in solution. So what will it do?



Student: It'll dissolve.



Professor Michael McBride: It'll dissolve. Okay, so a small one dissolves. This one just sits here. So that'll go away. And so as time goes on it gets smaller and smaller, the faces move in. How about this one, what will it do?



Student: It'll grow.



Professor Michael McBride: It'll grow, because it's more stable in solution. Right? Bingo. So that one grows, that small one shrinks. And this one in the middle is called the -- is metastable. We defined it as being stable, as just sitting there. Right? But it's metastable. If it gets any smaller it'll shrink. If it gets any larger it'll grow. Right? So you have to get a crystal that size, called the "critical size" before the crystal will grow. Right? So this suggests a way of getting only one form to grow. You have a solution, that's racemic, 50:50. Right? But there are no crystals in there that are big enough to grow. Can you see what you can do? You add a little fairy dust, right? Powdered crystal of one of the two forms, but have them be big enough to grow. So only those will grow. And then you can filter it and you have just that one. And that's done industrially, as a way of getting just one form, to separate the two forms; a means of resolution. Okay, and the other way is to add a poison that will absorb on one and keep it from growing. Okay, so that's one way that's actually used industrially, as well as in the lab. Another way is to form temporary diastereomers. Because remember, the right and the left are mirror images. So almost all their properties are the same; like their solubility. Okay? But suppose you put something else with it. Like here, let's shake hands Josh. Here, does that feel normal?



Student: No.



[Laughter]



Professor Michael McBride: It feels weird, right?



Student: Very weird.



Professor Michael McBride: That feels good. So right with right is different from right with left. Right? It's a different property. Okay? So if what I did was make diastereomers by adding something else to these molecules, which itself is only one hand, like right, now I'll have things that reacted that are right with right, and things that are reacted right with left. And what's the relationship between those new things? Are they enantiomers? Right with right, right with left, and now I'm comparing RR with RS. What's the relation? Are they mirror images?



Student: No.



Professor Michael McBride: No, they're not mirror images. They're just different. They're different solubility, different boiling point -- everything will be different -- different reactivity. Right? So on the basis of that I can separate them. But if I did it in such a way that now I can remove that thing I added, now after the separation I can go back and have the original things separated. Okay? So temporary diastereomers. One way to do it is chromatography. So the stuff that I put in the chromatography column can be all one hand. I put the new stuff on -- that's the packing of the column. I now run stuff through it. One absorbs more strongly than the other, doesn't move as fast. So one will come out the bottom of the column faster. That's one way of making temporary diastereomers; or making a compound with a chiral-resolved mate.



Okay, for example, this allenic compound that we saw before that proved van't Hoff was right, this is how it was actually resolved, how it was separated. So it says at the top they used alkaloids. An alkaloid is an organic base, isolated from plants, and the plants made only one hand of it. And so you use it to make a diastereomeric salt with a racemic mixture of right and left-handed acids. So now you have two salts -- right-right salt and right-left salt -- and they'll have different solubilities, and you can crystallize. So here's what happened. There's brucine; that's one of these alkaloids. And you see it has all those centers, stereogenic centers. So it's chiral, just one hand, because you got it from nature. And now you mix 4.7 grams of the acid and 5.2 grams of brucine, which means that there is about a 1:1 ratio. So you have both salts there. And now you crystallize, you get a solid salt, and it weighed 4 grams, which means you got 42% yield -- right? -- about half of the stuff out. And that stuff then you -- and you recrystallize it, to make sure it's pure, that salt you got out. The melting point didn't change, it was already pure. Liberate the acid by shaking it with hydrochloric acid and ether. Now you've got the acid; and it melted 145 to 146, and its rotation was +29.5°. And it recrystallized and didn't change it. It was pure. And it turns out that the other one melts 144 to 146°. And so it's the same melting point essentially, but exactly the opposite rotation. So they manage the separation by making salts that were diastereomeric -- so they had different solubility -- and then take the salts apart again. And Pasteur figured a way of doing this, in this story, which I suspect is not a true story.



Now I'm going to stop here. But we're going to have -- mostly what's on Wednesday's lecture is not going to be on the exam. But there's a little more here that I'm probably going to make available for the exam; that I'll talk about on Wednesday. And I'll have to think over just how much more it's going to be. Not too much more probably. Okay? So we'll see you on Wednesday. Most of what's on Wednesday won't be on the exam.



[end of transcript]

Lecture 29
Preparing Single Enantiomers and the Mechanism of Optical Rotation
Play Video
Preparing Single Enantiomers and the Mechanism of Optical Rotation


Within a lecture on biological resolution, the synthesis of single enantiomers, and the naming and 3D visualization of omeprazole, Professor Laurence Barron of the University of Glasgow delivers a guest lecture on the subject of how chiral molecules rotate polarized light. Mixing wave functions by coordinated application of light's perpendicular electric and magnetic fields shifts electrons along a helix that can be right- or left-handed, but so many mixings are involved, and their magnitudes are so subtle, that predicting net optical rotation in practical cases is rarely simple.




Transcript



November 12, 2008



Professor Michael McBride: So you remember Pasteur was very fond of racemic and tartaric acid. So when he'd travel around he would look for samples of tartaric acid. There are very detailed accounts that he published of these travels. I haven't found this one, so I'm not sure it's true. It may be apocryphal, but it was told to me by the guy who taught me organic chemistry. So I pass it onto you, for what it's worth, because it makes a good point. He said that Pasteur was in Alsace, the wine-producing region in northeast France, and in the town of Thann he found in a pharmacy a bottle of racemic acid that was moldy. So remember racemic acid's the 50:50 mixture. So, since he liked racemic acid -- and in fact it was becoming rare. It's a not a very common product; mostly you get the handed stuff. So he took it back and cleaned it up, got rid of the mold that was growing on it. And what he found was that after he had purified it, it was the unnatural, the left-handed, not the one you normally get. Can you see an explanation for that? Had the apothecary mislabeled it because he thought he could get a higher price for racemic than for tartaric acid? Or might he have been innocent? Zack?



Student: Was it the mold by any chance?



Professor Michael McBride: What did the mold do?



Student: Might have converted one into all.



Professor Michael McBride: Didn't convert.



Student: It didn't?



Professor Michael McBride: The mold ate the natural stuff, converted it into something completely different, leaving the unnatural one. Okay? Remember the smell of the carvones? Natural things, enzymes and so on, can distinguish between the two hands, and naturally the stuff that normally eats tartaric acid is going to eat normal tartaric acid, not the unusual one. So it left the unusual one. The penicillium glaucum had eaten the R,R. Okay, so that's one way of doing resolution, to get rid of one enantiomer from a racemic mixture, leaving the other one, because diastereomeric reactions have different rates. It's a right hand shaking a right hand versus a right hand shaking a left hand; they're just different. So one will go faster than the other. So if you react a racemate with some chiral reagent -- of course you have to use only one hand of that other thing you add in there -- but then you'll react with one more than the other. It could be even a catalyst so that it doesn't get consumed; an enzyme, for example. But that's not Nature's way, because it's very inefficient to make both and then destroy one. What nature does is prepare only one enantiomer. And you can do that, prepare only one, by starting with something that's already a single enantiomer and building on that. Or you can use a reagent, which is resolved, that will -- because diastereomeric rates are different -- will tend to produce one of the two enantiomers.



So here's an example of this molecule we've been talking about, Eisai-7389, with its nineteen asymmetric centers, that's made artificially, commercially. And this is where those nineteen centers came from. Five of them came from starting materials that they bought as a single hand. But the rest were all made. One of them was done by chromatography. They have chromatography columns about this big around, and this high, with chiral stuff inside. So you pass your thing through. One hand goes through a little faster than the other, because of its diastereomeric interaction with the packing; one comes through quicker. And they actually do that for one of the centers. But the others are all done by reactions that preferentially give one center rather than the other. Three of them are ones that everybody knew which one they would give. The other ones they just had to hope, and fiddle around with different reactions until they found one that would do the trick. So anyhow, they were able to generate these nineteen stereocenters.



Now we're left with a problem. That incidentally is all we're going to have for the exam on Friday. Okay? So now we're on to other stuff that will be covered on the final exam, and in particular return to this question about the tartaric acids. Remember there's d-(+) and l-(-). Could you have l-(+) and d-(-), of a different compound? Why not? What does d or l mean, in this context?



Student: [inaudible].



Professor Michael McBride: Andrew?



Student: It reflects light to the left.



Professor Michael McBride: It means which way it rotates the light. Right? And the plus means the same thing, as d. So it's redundant, this particular nomenclature. But there's a question mark because when -- Fischer just guessed, remember? And it could easily have been the opposite. So which way is it? Right? So if we knew, if we knew how optical activity worked, then by measuring the optical activity we'd know which of those structures is right. Okay? But we don't know that; at least I don't know that. Now there's a lot of knowledge about this, but it's a very, very tough problem. Fortunately there's a book by Laurence Barron called Molecular Light Scattering and Optical Activity, which goes into this stuff. And I see Professor Vaccaro --



Professor Laurence Barron: I'm here.



Professor Michael McBride: Oh, so we have a copy of the book. That's great. But even better than that, we have Laurence Barron.



Professor Laurence Barron: Or a copy of Laurence Barron.



Professor Michael McBride: Or a copy of Laurence Barron. [Laughter] Perhaps it's the mirror image of Laurence Barron. So he's going to tell us how this works.



[Technical instructions]



Professor Laurence Barron: Well good morning everybody. I was both delighted and dismayed, at the same time, when Professor McBride asked me to try and -- well to address you this morning. Delighted to have the opportunity to talk science to some of America's brightest budding young scientists, but also dismayed that he asked me to try and explain the molecular origin of optical rotation, optical activity. It's a very subtle, difficult, delicate problem, that's exercised some of the finest minds in physics and chemistry for the last hundred years. But anyway, let's see how it goes. Towards the end I think I may be presenting some stuff that's sort of beyond the boundaries of your current knowledge. But anyway, it can at least pass in front of your eyes. So chirality then means, as you well know, right- or left-handedness. It pervades much of modern science; from the physics of elementary particles, through organic stereochemistry, to the structure and behavior of the molecules of life; with a lot more besides. It comes up in what's called nonlinear optics, involving intense lasers; nanotechnology; materials, electrical engineering; pharmaceuticals; astrobiology; and origin of life. So it's a very important theme in modern science.



Now first of all though I'll tell you a little bit about Lord Kelvin. He was the first person to introduce the word chirality into science. He was professor of natural philosophy in Glasgow -- which is my home university -- through most of his career; well all of his career. He was one of the giants of physics of the nineteenth century. He's best known for inventing the absolute Kelvin temperature scale. Now he was originally, his original name was William Thomson, but then he became famous and became Sir William Thomson. Then he became even more famous, so they made him a Lord. And when you're made a Lord in the U.K., you choose your title from someplace that's dear to your heart, maybe your home area. He took his title from the name of the River Kelvin, which runs through the University Park in Glasgow. So whenever you use the absolute temperature scale now in the future, you can picture this idyllic scene.



Anyway, so he was the first to introduce the word chirality into science. And here's his definition, which you'll be familiar with: "I call any geometrical figure or group of points chiral, and say that it has chirality if its image in a plane mirror, ideally realized, cannot be brought into coincidence with itself." That was in his Baltimore Lectures. So he's just emphasizing the non-superimposabity of the mirror image, the enantiomers of a chiral molecule. But of course the whole subject started earlier with the wonderful work of Louis Pasteur, who showed mirror-image chiral molecules show optical rotation of equal magnitude but opposite sign; which was an epoch-making discovery. So you've come across then this, the fundamental manifestation of optical activity, which is natural optical rotation. You put linearly polarized light beam into a sample -- say a sample of an isotropic collection of chiral molecules, like a sugar solution -- and it will come out the other side with the plane of polarization rotated through some angle. And if you put in the mirror image version, you'll get an equal but opposite sense of optical rotation.



Now it's not to be confused with something called magnetic optical rotation, the Faraday effect. I'm just mentioning this to you. You probably haven't come across the Faraday effect yet, but you may do later on in your studies, or in your professional life. So I'll mention the Faraday effect. Faraday discovered, in 1846, that achiral samples -- no natural optical activity there -- if you apply a static magnetic field, parallel to the light beam, that will induce an optical rotation. And if you put in -- if you reverse the direction of the magnetic field, relative to the light beam, you'll get an equal and opposite sense of optical rotation. It would even work say for a sample of water, for instance. Any material will show a Faraday effect. But this has been a source of much confusion, in fact, to scientists.



Now Lord Kelvin was on the ball here. He knew all about it. He made a statement here. He said, "The magnetic rotation has neither right-handed nor left-handed quality." (That is to say no chirality; it's got nothing to do with chirality.) "This was perfectly understood by Faraday and made clear in his writings, yet even to the present day we frequently find the chiral rotation and the magnetic rotation of the plane of polarization classed together in a manner against which Faraday's original description contains ample warning." Well Lord Kelvin would be turning in his grave today, 100 years later, because you still see papers which involve the Faraday effect, in some way or other, and in the introduction they talk grandly of "inducing chirality with a magnetic field." That is completely wrong. Just for the record -- probably beyond your knowledge at the moment -- chiral phenomena, like natural optical rotation, they're characterized by what's called time-even pseudoscalar observables. A pseudoscalar is a number that changes sign under reflection or inversion; we'll leave it at that. So that's for the record. But well it turns out the essential symmetry characteristics of natural and magnetic optical rotation are completely different and you need different sorts of molecular quantum states, different characteristics to support them. But we won't pursue that.



Right, now back to natural optical rotation. So now we bring in circularly polarized light. So in order to detect molecular chirality, you must have some sort of chiral probe. Well what we're using here is right and left-circularly polarized light beams. They are actually mirror-image chiral systems. So they can be used as chiral probes. So here's a representation of a right-circularly polarized light beam. Now you know that light involves electromagnetic oscillations in space, and usually you just think of the oscillating electric vector of a light wave, and if it's linearly polarized, it's oscillating in one plane. There's actually also an oscillating magnetic field vector that oscillates perpendicular to the electric. We'll come back to that later. We're just looking at the oscillating electric vector here. So if it's linearly polarized, or plane polarized, it's just oscillating in a plane; but circularly polarized light, as well as an oscillation in this direction. You also have an electric [magnetic] vector oscillating perpendicular but 90° out of phase. And what happens is you get this circular polarization.



Now this picture here, this represents the instantaneous electric vectors at different points in space, in the direction of propagation, along the z direction. And so here we are, that's just showing -- that's just connected the ends of the instantaneous electric vectors. And there it is for left-circularly polarized. Anyway, you can see that at the very least those are chiral, those are helical, and those are mirror-image chiral systems. Now here we go. This is some extra fancy stuff that Professor McBride added to my [laughter] presentation. He doesn't like -- it was too simple and static. Anyway, so there we are. So now if you look at the wave, if you just look at the electric vector through -- in a fixed plane, as the light wave is propagating, there's the electric vector, and it will rotate in the fixed plane. So rotating clockwise, that defines right-circularly polarized light. And here we go, rotating counter-clockwise; that's left-circularly polarized light.



Okay, so we now have a chiral optical probe, circularly polarized light. Now chiral molecules respond slightly differently to right and left-circularly polarized light. I mean, an extreme example is say in the world of engineering. You can't fit a left-handed nut onto a right-handed bolt. That's an extreme example of different chiral interactions. Well it's obviously much more delicate than that here. But the point is right and left-circularly polarized light interact just slightly differently with a molecule of a given chirality. Anyway, there's a differential absorption of right and left-circularly polarized light. That corresponds to a phenomenon called circular dichroism, which is the basis of a widely used form of spectroscopy used to study chiral molecules. But now, what gives rise to optical rotation is a difference in refractive index towards -- of right and left-circularly polarized light beams.



Now linearly polarized light, you can describe it as a coherent superposition of right and left-circularly polarized waves of equal amplitude. Coherent means they're in-phase with each other. Rather than being random, they're oscillating in-phase with some fixed phase relationship. So, for example, now here's a linearly polarized light beam. But you can decompose it into a superposition of a left-circular and a right-circular component. You see when those are at the top they're reinforcing, and you've got the maximum linearly polarized vector up here. As they come away, left and right, they will tend to increasingly cancel, and that will decrease. And when they're this way and this way, you'll have zero electric field vector there, and as you come down it will increase again. So you can decompose a linearly polarized light into a coherent superposition of left and right waves. Now, refractive index corresponds to velocity through a medium. So if there's a difference in refractive index for right and left-circularly polarized light beams, that means there's a slight difference in the velocity of the right and left-circularly polarized light beams going through the medium. So the phase relations between the two contrarotating electric vectors will change. And you can easily see, this will give you a rotation in the plane of polarization. You see, if there's a difference in velocity, then at some instant this vector, this electric vectors of the left component will be here, and the right component here. And if you take the resultant, you see, it's no longer where it was. So this is a simple picture of how optical rotation develops, in terms of different refractive indices of the right and left components.



Now there's a picture in Atkins' Physical Chemistry. He tries to illustrate this there. He's got the linearly polarized beam coming through, and you've got the -- he's broken it down into the left and right, and he's saying the two velocities are slightly different, and that gives you the resultant optical rotation. You can easily develop an expression for the angle of rotation as a function of the difference in refractive indices for left and right-circularly polarized light beams. And it's also -- it's a function of the path length. Obviously the longer the path length, the more rotation will develop. And there's the wavelength there in the denominator. In fact, I mean, this is the secret; this path length, that's the secret of how you get a measurable rotation. Because this is an incredibly tiny effect. If it was just a single molecule event you're looking at, the polarization changes, you wouldn't see anything, they're so tiny. But you can build up this rotation over long path lengths, centimeters or even meters. In fact, if you go to Google, Google Images, and just Google 'circularly polarized light,' you'll find lots of sites there which describe polarized light, and they provide beautiful simulations, animations of this effect here. I didn't want to download any and try and show them here, but I would encourage you to go to that. But go and look in particular at this site here, enzim.hu. That comes from an institute of enzymology in Budapest. They've got some beautiful simulations of this optical rotation process and other more exotic polarization effects, as light propagates through matter.



Now, let's try and -- so that's just, what would you say, a phenomenological description. And in fact that's where most physical chemistry textbooks stop. And in fact Atkins' Physical Chemistry stops at that point, and then just there's just sort of hand waving saying okay, right and left-circularly polarized light interacts slightly differently with the chiral molecule; and they leave it at that. They don't attempt to try and give you a picture in molecular detail. But Professor McBride wanted me to try and attempt this. So just as a start, this is a simple picture, a simple scattering picture of optical rotation. Now a circularly polarized light wave 'bouncing' from one group to the other, as it scatters from a simple two-group chiral structure, will sample the chirality. So here we have a simple two-group chiral molecular structure. So we've got two achiral groups, held in a rigid, twisted arrangement. I think that's left-handed I've got there. And you can break down -- if you look at most chiral molecules, just look at the bonds, you can often break them down. You can see these sorts of two-group structures throughout. Anyway, so a particular model of optical rotation is that the light wave, it bounces from one group to the other, before coming off and getting involved in generating the optical rotation phenomenon. But you can see, it's sampling the chirality as it bounces from one to the other. So if the light beam is right-circularly polarized, that bouncing process will have a slightly different amplitude, as we say, from if it's left-circularly polarized. So that gives you a simple picture. It's worth mentioning that this is the basis of something called the "dynamic coupling model of optical activity" that was developed by somebody called John Kirkwood, who was chairman of this department in the 1950s.



Well now we really have to [laughter], we have to grasp the nettle at this point, because you just -- it's hopeless messing around with these simple models. [Laughter] They don't get you anywhere as regards prediction; you know, relating structure to sign and magnitude of optical activity. You have to go to the quantum-mechanics. So now this was the expression for the optical rotation, in terms of the refractive index difference for left and right-circularly polarized light. Now you can develop this using something called quantum-mechanical perturbation theory, and you develop an expression for the rotation angle in terms of this incomprehensible looking stuff. But let me just try and give you a feel for what it's telling us. The heart of it is this so-called rotational strength, which involves this so-called scalar product of an electric dipole and a magnetic dipole transition moment. So what we've got here is there's the ground state of the molecule n; j is some excited electronic state; and μ is the electric dipole operator that's connecting the ground to the excited state. And the light wave is interacting, is coupling with this electric dipole operator, and it's driving the transition. And it's this -- these electric dipole transitions, like this, that's behind conventional spectroscopy. You've done probably UV and infrared spectroscopy.



But now what we have in addition is the same transition, but now brought about through m; that's called a magnetic dipole operator, and that's activated by this oscillating magnetic component of the light wave. And so you have this so-called scalar product. So μ•m, that would be μxmx + μymy + μznz. Some of you have probably done vectors. Maybe others haven't. But okay. So that's the heart of it. But this is a very important feature here. We're summing over all excited states, j, all excited states. So the whole plethora of excited states of a chiral molecule come in here, and some can give you -- one particular excited state can give you a positive contribution to the optical rotation; another one a negative. You know? So it's a very subtle problem, and you have to consider them all carefully. So optical activity ultimately originates in interference between electric and magnetic dipole transitions during the light scattering process. That's at the heart of it.



Now, let me just show you how this works out. I can now give you a chemical, an interpretation, in terms of a real and highly important system in organic chemistry: the carbonyl chromophore. I believe you have come across the carbonyl chromophore. A lot of important organic molecules contain this carbonyl chromophore. It gives rise to a transition in the near ultraviolet at about 290 nanometers, and it's widely used in physical organic chemistry. It's called the n to π* transition. Now the carbonyl group itself -- here it is -- that's not chiral, that's got a plane of symmetry. So by itself it's not going -- there's going to be no optical activity there. But you see, in this particular molecule here, 3-methylcyclohexanone, that molecule overall is chiral. There's a chiral center there. So that carbonyl group is experiencing a chiral perturbation, a through space perturbation from the rest of the chiral molecule, and that induces chirality into the electronic transitions of the carbonyl group, and induces optical rotation and circular dichroism. Well let's look at this in a bit more detail, this famous n to π* transition. So here we've got the σ bonding orbital of the carbonyl group. There's the π bonding orbital, made up of two -- of the pπ on carbon and on oxygen. And here is a lone -- sorry, here is a py orbital. So those are px orbitals. That's the py orbital on the oxygen. Feed electrons in, and so you've got, in the ground state, you've got two there, two there. And you've got two electrons in the py orbital. So those are lone-pair electrons. So the lowest order, the lowest transition now is the py to π*, often called the n to π* transition. You're promoting one electron from the py orbital up to the π* orbital; and that's the origin of this transition. Well let's -- ah now.



Okay, now this transition, it happens to be fully magnetic dipole-allowed but completely electric dipole-forbidden. What happens is electric dipole character is induced by mixing -- well I'm giving the example of a higher oxygen dyz orbital into the π* orbital. Now Professor McBride has been messing around with my presentation here. Let's see what he's done. So here's the n orbital, the py orbital of the carbonyl group. There's the π*. Ah, now here we go. So this is now -- but wait -- yeah, so here you can see, in that n to π* transition, there's a net rotation of charge, a rotation of charge. And that's the essence of the magnetic dipole-allowed character. Magnetic dipole transitions involve a rotation of charge.



Professor Michael McBride: And the reason we messed about with them in this way is we've seen the mixing p orbitals changing the orientation, causing them to rotate that way.



Professor Laurence Barron: Right. So what's -- ah okay, now here we go again. Now here's a dyz orbital. Now one could consider a whole loads of other possible orbitals to mix in. But this is the simplest one, just to illustrate the idea here. So there's a dyz orbital. Now going from -- you see n to dyz, if you add those orbitals together, you tend to get a displacement of charge in the z direction. So now, you see, you have a combination of a rotation of charge with a displacement of charge, and rotation plus translation gives helicity. And here we go. There. So by mixing in a little bit of this dxz orbital, you get this -- which is electric dipole-allowed -- you get this helicity in the transition.



Now, in fact, the rotation of charge generates a magnetic dipole perpendicular to the plane of rotation, and it's pointing that way. So you'll get an mz component, a component of the magnetic moment along the z axis. Here you can see immediately you've got a component of the electric dipole along the z axis. So that generates a μzmz component, from this scalar product of μ and m, in the rotational strength. So this is sort of making a meal of it here now. So we're just putting down these so-called quantum-mechanical matrix elements. So here we've got n to π*, with a little bit of that mixed in, and that's fully magnetic dipole-allowed. Then we have n to π*, which is forbidden electric dipole, but mixing that in gives you a little bit of electric dipole character. And so you now get a non-zero contribution to the rotational strength.



However, you wouldn't want to go on and actually calculate the optical rotation of the carbonyl chromophore, in some situation from this. Because, as I said, there's many other excited electronic states you could have also considered, which may be giving opposite contributions to the rotational strength. You have to sum them all, and to do that you now have to go to modern ab initio quantum-chemistry, quantum-chemical calculations. There's wonderful programs out now, from Gaussian and Dalton; you can calculate all sorts of molecular properties, with quite good -- very good accuracy, in many cases. But this now turns it all into a black box procedure. You just feed in appropriate atomic orbitals and press buttons and turn handles, whatever you do for these calculations, and out will pop some physical quantity. So you can calculate this whole thing, ab initio, and taking in however many excited states in the sum are necessary. And you can calculate the sign and magnitude of the optical rotation, for a given absolute configuration. So you would feed into the calculation whether it's the S or the R absolute stereochemistry. So you'd put that in, and that will determine the sign of the optical rotation that comes out. And so -- I mean, in this particular case, this small molecule, the S absolute configuration, goes with the plus rotation and the R goes with the minus. So from the calculation then you can relate the sign of the optical rotation to the absolute stereochemistry.



But you wouldn't want to stake your life on it, because it's not completely reliable. It usually gives the right answer, for smaller molecules, but not always. And, for example, if this was -- as you know, and I think you'll hear more after my talk here. Many drugs now are chiral, and drug companies like to market now single enantiomer versions of the drug. So it's a tremendously important problem for them to know, with absolute certainty, the absolute configuration of the particular enantiomer they're marketing, because if they make a mistake and specify the wrong absolute configuration in their patent, they can literally lose billions of dollars. Anyway, there's the famous -- the importance of chirality in drugs is exemplified by the famous thalidomide case. But I won't elaborate that any more. I think Professor McBride is going to tell you some more of that in more detail. I should say that the cornerstone, the definitive method for determining absolute configuration has, for some years now, been something called the Bijvoet method of anomalous X-ray scattering, which again Professor -- have you told them about that?



Professor Michael McBride: Yes.



Professor Laurence Barron: Yeah. Yeah, that's the definitive method. But it's sort of cheating because you're actually seeing -- you're seeing it through X-ray eyes. But even then, you can occasionally make a mistake. But I should just mention also, there are some newer optical activity techniques involving something called vibrational optical activity. Here we've been looking at optical activity in electronic transitions, using visible light. But in recent years newer methods, measuring optical activity in vibrational transitions, have come along. And these are actually comparable with X-ray, anomalous X-ray scattering, for reliability of absolute configuration. That's because these calculations are much more reliable for these vibrational optical activity phenomena than electronic ones. Well that's about it. That's all I would like to say now. So thank you for listening.



[Applause]



Professor Michael McBride: Now that you've heard from the authority, do you have any questions about it? You are unwontedly quiet, like you usually are. Any questions? Yes?



Professor David Spiegel: I'm just wondering if the magnitude of the Faraday rotation can ever interfere with the measurement of optical rotation in the chiral material? That is to say, if one situates a polarimeter in proximity to an NMR spectrometer, for example?



Professor Laurence Barron: No. No, you wouldn't need to worry about that at all. No. You need a reasonable strength of magnetic fields just applied directly through the sample. Yeah.



Professor Michael McBride: Lucas?



Student: Just if I'm understanding. The n to π* is magnetically dipole-allowed, electric dipole-forbidden, and that's why then it goes to the xz as well, in order to get the electric dipole…



Professor Laurence Barron: The electric, yeah.



Student: …in there, that contribution.



Professor Laurence Barron: Yes.



Student: For my knowledge.



[Laughter]



Professor Michael McBride: Okay, thanks again Laurence.



Professor Laurence Barron: My pleasure.



Professor Michael McBride: Very good.



[Applause]



Professor Michael McBride: Thank you.



Laurence Barron: Okay.



Professor Michael McBride: So yeah, this is just a plug. This afternoon -- the reason Professor Barron is here is he's the Tetelman lecturer in Jonathan Edwards College. So he's giving a talk for the history majors and so on. But it'll be very interesting for anyone. And I think since we're into chirality, you would enjoy this, particularly this afternoon, in Davies Auditorium at 5:00. But there's the question, who cares? No offense. But who cares? Why do we care about chirality? Well Professor Barron hinted at it. Living things care, because they're chiral. Right? So which one they react with. Okay? The Food and Drug Administration cares, for the same reason, with respect to medications you might be taking. Drug companies care a lot, and their lawyers. And the US Patent Office cares a lot. Right? Which has generated a thing called a "chiral switch." Most drugs that used to be developed were developed as racemates, because it was difficult to separate the two hands. But if your patent runs out on the racemate, and you can resolve it and now sell just one of them, and if it's better, and the FDA will approve it, and you can convince people that that's the one they should buy, then you can go back to not having to compete with generic drug companies anymore and charge five dollars a pill instead of fifty cents a pill. Okay?



So this so-called chiral switch, to go from racemic to a single enantiomer, is a big movement nowadays. For example, consider this pain reliever. Let's figure out what it is. Do you remember what that group is: four carbons, in that sort of Mercedes structure? Do you know what that group's called? This, what we're doing for the rest of the lecture here out is actually review for the exam on Friday. You're not specifically responsible for it, but it's review. So you know that group?



Student: Isobutyl.



Professor Michael McBride: Isobutyl. Okay, and what acid is that, with three carbons, do you know? The first fatty acid?



Student: Propionic.



Professor Michael McBride: Propionic acid. And in the middle we have --



Student: Phenol?



Professor Michael McBride: Phenyl group. Okay? So what's the name of the drug?



Student: Ibuprofen.



Professor Michael McBride: Sherwin?



Student: Ibuprofen.



Professor Michael McBride: Ibu-pro-fen. Advil or Motrin. Okay, now in this case you have a chiral center there -- right? -- because there's a hydrogen on there that we don't see. And the S-form is a pain reliever; it's said to be so anyhow. And the R is inactive. Right? But it's marketed as a racemate. Right? One might try doing a chiral switch and selling only the S. But the trouble is that it very quickly racemizes inside you. So it wouldn't be doing any good, right? Because you'd do all the work of selling the S and then it would become R immediately, when you ate it. And here's another one. This is the one that Professor Barron referred to, which is a sedative; thalidomide. And there's the chiral center in that one, because there's an H on there. So the S-form is a sedative, but the R-form, at least it's said, is a teratogen; which means it makes monsters. [Laughter] And it's not so funny, because it was sold as a racemate from 1957 to '62 -- never in the U.S. because the FDA didn't -- they were lucky and never approved it. But in Europe it caused 10,000 birth defects; children born without arms, legs, things like that. So it was a tragedy. But this one also undergoes in vivo racemization. Okay?



You can find rate constants for these. It's curious that the rate of S going to R, and the rate of R going to S, are not the same. There's got to be more to it than that, that one has to be more stable than the other, of these mirror images; if that's true. It may be that the rates aren't exactly true. But if so, they must be bound to something that makes -- that's chiral -- that makes one of them more stable than the other. Anyhow, that's how fast they go back and forth. And this is how fast they get eliminated from the body; one gets eliminated much faster than the other. And if you put these things together, you can see how the concentration should vary with time, if it's going according to that rate. So the blue one, the one that's good for you, quickly drops off. The bad one, the teratogen, grows quickly. So you have maybe two hours of twenty-four hours where you got a lot more of the good one than the bad one. So essentially this drug is completely off-limits, especially for any women that could conceivably, under any circumstances, be pregnant. Okay? So but it's a wonderful drug for things like leprosy and so on. David here is an MD, as well as a professor of chemistry, and you probably know more about that. But it's a wonderful drug for certain things. Isn't that right?



Professor David Spiegel: That's right, that's right. It's still actually in common usage for anxiety disorders.



Professor Michael McBride: Yeah, but there are all sorts of warnings, in letters this big, about if you have a chance of getting pregnant, stay away from this baby. Right?



Professor David Spiegel: In fact, if you're a male patient taking it, and your wife is pregnant or likely to become pregnant, you're encouraged not to take it.



Professor Michael McBride: Wow.



Professor David Spiegel: Because of possible contamination.



Professor Michael McBride: Wow. Okay, so now here's a drug. We're going to look at the name of this one. It's a really whopper of a name, right? So let's just use this as a practice about nomenclature. Okay, so that thing there is called imidazole, and the 1H tells where -- which one has a hydrogen on it. Right? Okay, so that's position one. And that is the benzene ring; so benz. So it's benzimidazole, that group on the right. And you number it. Remember it was 1H; there's an H on the nitrogen one there. And you go around the ring and number in that conventional way. Okay, now but then you notice that there's something on the number five, and there's something on the number two carbon of that ring. On five, there's a methoxy group, which appears first among all the things in this -- named here; not the thing that's on two. Why is the methoxy group first? Do you remember what the rule is?



Student: [inaudible].



Professor Michael McBride: How do you arrange the substituent groups? Angela?



Student: Alphabetically.



Professor Michael McBride: Alphabetically; m is going to be the first one. Okay, so 5-methoxy, and then two is sulfinyl; that S with an O on it is called sulfinyl group. Okay, so it's 2-sulfinyl. But now there are all sorts of curly brackets and square brackets and so on, to tell what's attached to that. Okay, so attached to that is methyl group; the methyl is substituted. So it's methyl sulfinyl. Okay, that's the curly -- the square brackets. Okay, now what's on the methyl? There's a pyridine group; the benzene with a nitrogen, is pyridine. And it's substituted on the methyl at its own two position; the two position of the pyridine is what's attached to the methyl. So it's 2-pyridinyl, the end. But now on that, in the four position, is methoxy. And methoxy comes alphabetically before methyl; not before d, but before methyl, and you don't count the di. Right? So there's the name of that compound.



And this stuff is a drug. It's a gastric proton pump inhibitor. So it treats acid reflux disease. And it's the world's largest selling drug in the Year 2000; 6.2 billion dollars. And it's called omeprazole, or Prilosec. And you've seen probably, some of you at least, have seen boxes of Prilosec. I hope you don't have to take it, like I do occasionally. This is called OTC. We're going to be talking about that on Monday. Okay, so there it is. Now get your glasses up. Because can you see this? Can you see any sort of three-dimensions here? What's in front, or what's behind? Can anybody tell? You have any luck in seeing it in three-dimensions? Some people can't see it, but most of you, about 95% of you, should be able to see. It's not a really high quality three-dimensions. And because the computer -- the projector doesn't do it exactly right, for the colors. So one eye sees only one and one sees only the other. Anybody see? Okay, now what is chiral here? Can you see why this thing is chiral? There's no carbon with four different things on it. But what makes it handed? Can you see? In fact, there's no group that has four different things on it. Any suggestions? Lucas?



Student: Sulfur.



Professor Michael McBride: Sulfur. Sulfur has an unshared pair, an oxygen, and two different R groups. And which is pointing out toward you? Can you see it enough in three-dimensions, with the glasses, to see that, to see which is pointing out?



Student: The unshared pair.



Professor Michael McBride: Yeah, it's the unshared pair that's pointing out. So if you use your thumbs and recognize that the unshared pair has the lowest atomic number -- zero, right? -- then you can do it. And you'll find out that that particular one drawn there is the S enantiomer. Okay? So it's known -- omeprazole, Prilosec, Prilosec OTC, are the racemate. But this one that's drawn here is the S isomer, and it's called what? Esomeprazole; that's the name of it, right? Or Nexium. Right? And it's the S isomer, right? So this is the product of one of these chiral switches, where for a long time the stuff was marketed as a racemate and now they market it as a single enantiomer, Nexium. And we tried this one last time, so we're not going to spend time on it right now. That's what the class looks like when they're doing this. It's fun for me to see you. [Laughter] Okay?



But there are other ways of doing the stereoviewing. You can use a pair of periscopes, like this. Right? And the way that works of course -- well the point is, for each eye to see a different image. So if you want to try those and look at this after class, feel free. Or borrow them if you wish. Right? So what the eyes perceive is a superposition in the middle. It's like this. There are two pictures. But if you can do it with the glasses, you can do it probably just by looking at this picture, if you have a little while. But we don't have the little while now. The right eye will see that, the left eye will see that, and you see, in the middle -- there see, something in the same position but in slightly different projections. So the image in the middle of the three seems to be in 3-D. Okay.



Now we don't have time to go -- we're going to do a lot of curved arrow stuff, about how reactions happen in making omeprazole; and then, even more interesting, in the action of omeprazole that stops the stomach acid. But we'll have to wait 'til after the exam for that. Okay? Thanks again to Professor Barron.



[Applause]



[end of transcript]

Lecture 30
Esomeprazole as an Example of Drug Testing and Usage
Play Video
Esomeprazole as an Example of Drug Testing and Usage


The chemical mode of action of omeprazole is expected to be insensitive to its stereochemistry, making clinical trials of the proposed virtues of a chiral switch crucial. Design of the clinical trials is discussed in the context of marketing. Otolaryngologist Dr. Dianne Duffey provides a clinician's perspective on the testing and marketing of pharmaceuticals, on the FDA approval process, on clinical trial system, on off-label uses, and on individual and institutional responsibility for evaluating pharmaceuticals.




Transcript



November 17, 2008



Professor Michael McBride: Okay, let's start up. So remember we were talking about configuration, about handedness, which was introduced - the idea was introduced - in the context of a tetrahedral carbon. But you don't need a tetrahedral carbon in order to have handedness. And an example is shown by the oxidation of a sulfide to a sulfoxide. So let's first talk about the mechanism. What makes the sulfide reactive in this case? What do you see for an orbital that would make the compound on the top left? Russell?



Student: High HOMO.



Professor Michael McBride: Pardon me?



Student: High HOMO.



Professor Michael McBride: High HOMO, the unshared pair on sulfur. And how about the peroxy? Notice that's not a normal benzoic acid. It's got two oxygens in a row. What reactivity does that confer on it? Devin, you got an idea, two oxygens in a row? What's going to be unusual in terms of an orbital? We're looking, obviously, for a low LUMO.



Student: It's probably going to be a higher HOMO.



Professor Michael McBride: Yeah, it'll have -- it has unshared pairs, no doubt. But we want it to react with the sulfur that has unshared pairs. So we're looking for a low LUMO. Anybody got an idea? Russell?



Student: σ*, between --



Professor Michael McBride: σ* of the oxygens, because oxygen's got a big nuclear charge, right? So its orbitals are low. Okay, so we have the unshared pair on sulfur, σ* on the oxygen, and we get one of these where we make a bond and break a bond at the same time, which puts on oxygen onto the sulfur. And notice that, as you said Devin, the oxygen has unshared pairs, the n electrons there. But those won't give stabilization by mixing with an unshared pair on sulfur. But sulfur's in the next row of the periodic table. So it has a vacant d orbital that can overlap with that. So you can get partial stabilization there, of the unshared pair on oxygen, which you could denote by drawing a double bond. Right? So we sometimes draw a double bond to sulfur. And we can lose the proton from that, and we get what's called a sulfoxide. So there's the mechanism of making a sulfoxide. And notice that it's not planar at the sulfur. Even though the sulfur has only three substituents, it's not planar. So it's asymmetric. It's handed, it's chiral. Right? And obviously you could easily well have reacted the top unshared pair of the starting material, as well as the bottom one. So when you do the reaction you get a racemate, a 50:50 mixture of the sulfoxides. Okay, now we mentioned omeprazole last time. Here's the chemical structure of omeprazole. What will make it reactive? Anybody got a hint for what might make it reactive? Elizabeth?



Student: A lone pair on nitrogen.



Professor Michael McBride: It's got a lone pair on two nitrogens.



Student: Right.



Professor Michael McBride: Right? Okay. Now one way to discriminate between them is to protonate one of them. Right? Because when you do that, you now have this carbon-nitrogen double bond. Right? And the π* will be unusually low because of the positive charge that came from the proton. So acid is the stuff that's going to activate this. So now you've got a low LUMO and also the high HOMO, from the other unshared pair on nitrogen. So bingo, you can add, just as the nitrogen would add to a C=O double bond, like that. Then you get this thing with now four bonds to that central carbon. Protonate again, and now this compound has two positive charges. That's not so great, to have a dication. So you get rid of it. You can get rid of it by losing that original proton. But there's another proton you can lose, just as easily, in this way. Right? So lose the proton on the nitrogen and then have those electrons re-form the double bond to carbon, and the electrons go out on sulfur, and you get this thing. Now, that happens -- remember it's the acid coming along; the H+ is what caused this to happen to omeprazole. So if you have a pH, an acid pH between one and three, omeprazole will undergo this acid-catalyzed rearrangement, with a half-life of two minutes. So in no time at all it's converted to this form, which is the active form. Now active to what? Well suppose you have an enzyme that has a sulfur on it, that's got an unshared pair of electrons, that will make it reactive. What will make the drug, in this form, reactive? Anybody got an idea? Russell, I'm going to go to you again, because you answered the same question before. We just made this group here.



Student: Yes, that's right.



Professor Michael McBride: What's going to make it reactive?



Student: π* makes it --



Professor Michael McBride: Not π*, no.



Student: [Inaudible].



Professor Michael McBride: It's the same as O-O, right? Sulfur is right below oxygen.



Student: I meant to say σ*,



Professor Michael McBride: Can you see that? So it's σ*. Oh, that's what you meant to say.



Student: Yeah.



Professor Michael McBride: σ* S-O. Okay, so we got that low LUMO and the high HOMO on the sulfur coming in. So we can do exactly the same reaction as before, right? Make a bond, break a bond. OH- goes away. Right? And we have this thing where we've now formed a covalent bond with the enzyme. So the enzyme can't do its stuff anymore, right? Because it's tied up. Okay? So the pump, the pump which takes acid that's made in the cells that line the stomach and transports the acid into the stomach, doesn't work anymore, because that enzyme is what did the trick. So the pump enzyme is inactivated and the flow of HCl to the stomach is stopped. Okay? And, in fact, this is an interesting problem in the design of the drug, because it's acid that causes this to happen, in two minutes, to become the active form. But you don't want it to happen until it gets into the cells that line the stomach. So you take the pill orally, it goes into the stomach. What problem are you suffering from, when you put the -- when you take this pill?



Student: Acid.



Professor Michael McBride: Acid, in your stomach. So bingo, it's going to happen in the stomach, not in the cells in the wall of the stomach. So they have to coat it with something that'll make it get through the stomach first without doing its reaction, and so they say don't grind the pill up -- right? -- before you take it. Let it get through. Then it gets through the liver, into the bloodstream, and comes back to the wall of these cells in the stomach, and then it gets activated by acid and does this trick and stops the acid from pumping into the stomach. So that's the idea.



Okay, now should a chiral switch, to a single enantiomer -- omeprazole, Prilosec, is a racemate, it's a 50:50 mixture of the enantiomers; enantiomers, remember, at sulfur, not at carbon. Okay? Now, if you go to a single enantiomer, will the drug be twice as good, or at least better? Can anybody see a problem that you would have? Right? When something interacts with an enzyme, the enzyme is a single hand. So the complex, the reacting complex, between the enzyme and the stuff that you're reacting, is diastereomeric if that stuff is handed; if it's right or left-handed. The enzyme say is all right-handed. So you have right-right, or right-left. One of them will be better than the other. Right? So is that going to be something that'll be important here? If we make a single enantiomer of omeprazole -- remember, we talked about thalidomide, ibuprofen before. Will a single enantiomer likely help? There's an interesting observation about this compound, the active form, that makes it different from the original omeprazole. Do you see what it is? Incidentally, notice, that's the original omeprazole. All that that mechanism I showed you did was to change that bond from there to there, and the proton from nitrogen to sulfur. Let's go back again. That's all that happened during the activation. But it did something crucial with respect to stereochemistry. How about that original form? Was it chiral?



Student: Yes.



Professor Michael McBride: Why? Angela? What made it chiral? Yeah?



Student: The bond is going -- like it can either go inside or outside.



Professor Michael McBride: Right. The bond -- the oxygen on the sulfur could be either going in or out; the sulfur is pyramidal. What happened when we activated it?



Student: It gets bigger.



Professor Michael McBride: Now it's not chiral anymore. So this thing is not going to discriminate between anything, because it's not chiral anymore. Okay? So this thing is achiral. Okay? So it shouldn't make any difference which hand you're using at this stage. Now, so there's no difference after omeprazole has been activated by acid. Notice it's not activated by an enzyme, because if it were activated by an enzyme, then it could discriminate between the two hands and one would be more activated than the other. Right? So that could be a difference. But that's not true. It's just acid that does the activation and makes it achiral. So at first glance one would think it'd make no difference at all. Still, it could be that one enantiomer is better at getting through the digestive system and getting back to the stomach, in order to do the trick. So it's still possible. How could you tell whether it's better or not? This is getting pretty complicated -- right? -- all the different things it would go through to get from here to there. How would you find out whether it's worth using a single enantiomer? What would you do?



Student: Test it.



Professor Michael McBride: Test it. How?



Student: Clinical trial.



Professor Michael McBride: You'd do a clinical trial. Okay. So you need -- but in order to do that, you need a single enantiomer in order to do the laboratory and the clinical testing. And we're going to talk about that in just a second. But first let's talk just a bit about the economics of this. These things called proton pump inhibitors are the newest generation of things to treat acid reflux. And here's data from Wellmark, which is the Blue Cross/Blue Shield of Iowa and South Dakota; so a pretty small population area. And this is how many prescriptions they wrote for Prilosec in the period from 1999 to 2003; just went up by 250% to a quarter of a million. And if you go world -- so 15% of the members of Wellmark got this stuff, as of 2003; and 600 million worldwide. That's a big market. Right? So now here's cost comparison for things that treat acid stomach. So the first stuff is Tums and Rolaids and so on. Those cost three or four cents apiece. Okay, then there are these over-the-counter H2 blockers, like Zantac, which are now very cheap; like thirty-seven cents. But they didn't used to be, because they used to be covered by patent, and were prescription drugs. Okay. Then there are prescription versions you see of these H2 blockers, and Zantac, when it's for prescription, costs $4.27 apiece.



Okay, but now these proton pump inhibitors come along. And in 1988 Prilosec came on the market at $4.61 a pop. You take one every day. Okay? But in 2002 the patent ran out. So now generic people started selling the generic version of omeprazole for $2.76, and AstraZeneca, the manufacturer of Prilosec, said we can do better than that; we've been making this stuff for years, so we'll make an OTC version and have Proctor & Gamble sell it. And so in 2003 they introduced Prilosec OTC; which was actually exactly the same stuff but seventy-nine cents a pill. Right? But you can still get the prescription form, if you want the real thing, for $4.61 a pill. Okay, but so now you're losing a big market. If your pills are costing only something like what?; 1/6th or 1/7th of what they originally cost.



So in 2000, AstraZeneca introduced Nexium, which is a single enantiomer. Okay? And now that's $4.87 a pill. Right? And that process, to go from the R/S to the S, to go from the racemate to a single enantiomer, is called a "chiral switch", in the industry; to go from a racemate to a single enantiomer. Okay. Now this was touted, within AstraZeneca, as the most successful U.S. launch ever. And here you can see the graph of how much better it is than Viagra, Vioxx, Lipitor, Celebrex and so on. Over this little more than three-year period it went up to eight billion dollars in sales. Okay? And you can see that there's been a lot of integration of clinical and commercial enterprise at AstraZeneca. This is from a website that you can see there. In December 2003 they say:



"…the executive director and development brand leader" (at AstraZeneca) "adds the clinical and scientific proficiency as a research gastroenterologist. As Levine and his staff put together clinical development plans, such as additional indications or line extensions, they get commercial input at every stage."



Well I don't know if you noticed I drink water every once awhile because my mouth is dry. And just before -- when I learned that the class was going to be filmed this year, I realized that I'd been suffering from hoarseness since last spring, and I figured that wouldn't be good for the recording. So I went to see a doctor to find out if I could do anything about this, whether there was anything wrong. Now there's something you can do for this, and Nexium has been the most intensively advertised stuff ever. Right? That was for the public. This is for doctors. Okay? So if you go to that website, you can see a seven-part, seven-scene description of why Nexium is so much better than anything else. And what I'd like you to do for a homework problem for Wednesday is to go through that show and evaluate whether this series shows that Nexium is superior or not. Is it worth paying five times as much for a pill, or six times as much for a pill?



Now, this is the FDA-approved label for Nexium. And so they did clinical trials, just as we said they should do. Okay, so here they tried to heal erosive esophagitis. So what they tested, you can see, is the healing rate for this condition, from 40 mg of Nexium, 20 mg of Nexium -- so that's the single enantiomer in two different doses -- or the traditional omeprazole, the racemic stuff. So, "These were evaluated in patients with endoscopically diagnosed erosive esophagitis in four multicenter, double-blind, randomized studies. The healing rates after 4 and 8 weeks were evaluated and shown in the table below." Now suppose you were in charge of designing this test. How much of the racemate, omeprazole -- so 40 mg and 20 mg of Nexium -- but compared to omeprazole, how much would you have administered to compare with -- so there are the single enantiomer, 40 mg or 20 mg. How much omeprazole would you use?



Student: Forty and twenty.



Professor Michael McBride: And why? Lucas, what do you say?



Student: 20 mg, as little as possible.



Professor Michael McBride: [laughs] Why?



Student: To show the best distinction between the new stuff and the old stuff.



Professor Michael McBride: Okay. Does everybody agree with Lucas? Kate, what would you choose?



Student: Well I'd want to do both twenty and forty. But it looks like there's only one omeprazole.



Professor Michael McBride: Okay, but if you were designing the test, you'd do twenty and forty. We'll have an auction here. Do I hear eighty from anyone? Rick you --



Student: I would do eighty and forty.



Professor Michael McBride: eighty and forty. Why?



Student: Because that way -- because so omeprazole is a racemate. If you do eighty and forty of those, you would get forty and twenty of the active form.



Professor Michael McBride: Okay. Does everybody see what Rick's saying? If you want the same amount of what is ostensibly the active stuff, you should have double the amount of omeprazole. Now this is what the label says, what was actually used; 20 mg. You win Lucas. Okay? Now why? It's because that's the approved dose for that disease. Right? That's what the FDA approved, right? But we'll give twice or four times as much of what we think to be the active ingredient to the other people in the test. Okay, so that's the test. There were four different tests, and this shows it. But let's summarize it with a graph that'll show it I think more clearly.



So after four weeks and after eight weeks, these are what fraction of healing there was with the racemate, which is shown in open figures. And the shapes are different studies; remember, there were four different studies. And this is what you get if you use the single enantiomer, is 20 mg; which remember is twice the dose, if that one form is active. What would you conclude from this? Is it worth paying seven times as much, or six times as much?



Student: No.



Professor Michael McBride: Okay now -- or you could use 40 mg. Now I think we would probably agree that it's better, just eyeballing this thing, even if we aren't statisticians. Okay? So four times the dose does a better job. Now, then this was heavily advertised. You may have seen these things for the purple pill. So you get this grey-haired guy here who says,



"If I told you prescription Nexium heals acid reflux damage better, you'd want proof. And now your doctor has that proof. Recent medical studies prove Nexium heals better than the other leading prescription medicine." (Now notice, he says prescription medicine. The OTC stuff isn't a prescription medicine. So this is a different test.) "No wonder they call Nexium the healing purple pill. So call your doctor today, because if left untreated the damage could get worse."



So you can get a purple pill there. Okay, this is the test he was referring to, which compared esomeprazole, that's Nexium, to lansoprazole. But again it's 40 mg versus 30 mg. I don't know anything about lansoprazole, but that's what they tested anyhow. But notice incidentally, in the acknowledgements, this study was supported by a grant from AstraZeneca, Wayne, Pennsylvania. So these are not completely disinterested people, at least some of them, who are involved in this study. And it says, "So call your doctor today to learn about this." And fortunately we have my doctor here, Dr. Duffey. So she's going to give you her perspective [laughter] on Nexium and omeprazole. Thank you for coming.



Dr. Dianne Duffey: My pleasure Professor McBride. Thank you for having me.



[Technical adjustments]



Dr. Dianne Duffey: Well I can remember sitting in organic chemistry classe as a pre-med, wondering why it was I needed to learn how to make paint. So this is a real treat for me to be here today, to be able to talk to you about some clinical aspects of why it's important to study and do well in Professor McBride's class. Sorry. All right, so I'm a otolaryngologist, I'm an ear, nose and throat physician, and I work here at Yale. Some of you may recognize me from the health plan. And Professor McBride did give me permission today to talk a little bit about his case. So I'm not going to show you any confidential pictures, but I may allude to him here in the talk.



So first of all the disclosure. I don't have any financial interest in any of the drugs or companies discussed today. I'm not on any speakers' bureaus. So, in a word, I don't really have a vested interest in anything shown here today. But I will discuss some off-label or experimental uses of these compounds. And the opinions presented by me are mine and no one else's. So as you've heard testified here today, Professor McBride took Prilosec and actually did improve. So he sent me an email and was very excited that I could come and speak to you all today, and said, "By the way, my symptoms were improved." So Prilosec fixes symptoms of GE reflux disease, gastroesophageal reflux disease, and laryngopharyngeal reflux disease, which is the entity that I treat. GE reflux disease is in the domain primarily of the gastroenterologist. So I'm going to refer mostly to LPR today, laryngopharyngeal reflux.



So we've heard that Prilosec works. Or does it? We really need to know. How is it we know these drugs work? Someone talked about clinical trials today, and I'm going to focus a little bit more closely on that, because that's -- we rely on these, and they come primarily from pharma. Yale University, my clinic, doesn't really have the money to run extensive clinical trials on my patients. Sometimes some centers are able to do it, or you may have a grant to do something that looks at a drug. But primarily we're relying on what the pharmaceutical companies tell us in terms of data. So about Prilosec working in him, all we know is that his symptoms are improved -- they were in his body -- eating his diet, and living his life, and taking the drug at the prescribed dose. But there are a lot of variables that we may not know about, and some of these variables are what the pharma companies need to take into account, when they're doing these clinical trials, when they're designing them. Are we taking other patient variables into account? For example, diet; does he take a large number of herbal supplements we don't know about, for example? This could affect the pharmacodynamics. It could affect the pharmacokinetics of the drug in his body. Did he take the prescribed medication on an empty stomach, as it's supposed to be taken, so that the acid will activate it early? In other words, was he compliant? And these are all things that I need to be taking into account when I prescribe a drug to a patient.



For example, these studies were done in patient populations, but they may have been done on the West Coast, they may have been done in China. Is that patient population representative of my patient population? So when they're designing these trials, they try to control for as many patient individual variables as possible. In addition, these studies have to be statistically sound, because biostatistics drive these clinical trials and their design, so that if differences are actually observed between Prilosec and a placebo, or Prilosec versus esomeprazole, we have to be able to determine with reasonable certainty if these differences are real or if they're due to just chance alone. So that's where biostatistics comes in. And I submit -- and I think Professor McBride, who helped write some of these slides, probably also feels -- that there is some duty on the part, not only of the manufacturers, but of academic institutions, who are also running their own clinical trials, to actually design the studies so that it's easy to understand; the data are there and they're very clear, so that they can make also very legitimate head-to-head marketing comparisons between competitor compounds.



So I think this is very important. As a physician, I feel that I have a duty to really evaluate the literature critically. Sometimes I actually need to go back and pull the studies. I can look at the package insert, for these drugs, but how do I know -- how can I actually believe it? So sometimes I actually go back and I'll pull the study off PubMed, and read it. And we have to be able to ascertain the validity of the research that supports our choices as clinicians. There's a lot of marketing out there; as we heard, there's a lot of money to be made off these compounds. And I think I'm going to also propose today that we as a society, we as patients, we as members of the society, perhaps a society who may one day have a single-payer healthcare system, we need to be educated also. And this information's available. It's available on the web. You've already seen some of it from Professor McBride this morning. This website is a fantastic resource. If you have a question about any medication you're taking, you want to look at the chemical compound of it, the chemical structure and how they designed it, how they make it, how it's marketed, how it's distributed, all this is available on the website. So you can just plug in the drug that you're taking, or the drug of interest, and that information's all there. It's public domain. So use that as a resource.



But direct-to-patient marketing can be really effective. And this is also referred to as direct-to-consumer marketing. You saw some of it in the earlier slides, but all you have to do is turn on the evening news, and these ads are there, right? Every time there's a commercial break, you see some happy person walking around, talking about this drug. And you just hear the drug name. You don't hear anything more about it. You may not hear anything about the studies. But it can be very effective. So I think we need to look critically at some of the claims, and we need to think, is this really actually the right thing for me to be taking?



So my specialty -- so back to our clinical model -- my specialty is otolaryngology. I'm focusing on the larynx with laryngopharangeal reflux. The larynx is the voice box. It's where your vocal cords are. It's what we use when we're speaking. Ear, nose and throat is another name for my specialty. But we're talking about laryngopharangeal reflux, which is reflux, acid reflux, that primarily affects the voice box. It's under-diagnosed. It's also a significant source of morbidity and decreased quality of life -- hoarseness, feeling like you have a frog in your throat, sore throat -- and it's frequently associated with other forms of reflux disease, like GE reflux disease, gastroesophageal reflux disease. So it's a significant public health problem. And this is a very busy slide, but just to point out some statistics: that up to 10% of patients who present to the ENT practice, for any reason, may actually have symptoms or findings related to LPR, laryngopharyngeal reflux. It's also increasingly recognized as a problem that can be associated with non-allergic asthma, and a great number of these patients also present with a history of acid reflux from other -- where their symptoms are coming from other sources, such as the esophagus, for example. So reflux is a very big problem. I see a lot of patients. It's estimated that up to 40% of the adult population in this country may have reflux; and some studies say even more. So there's a lot of money to be made.



What is its treatment? In the year 2008, what we're doing is we're treating this with proton pump inhibitors. But I have to also put in a disclaimer that there was a recent meta-analysis, within the last couple of years, looking critically at the literature, and at a number of different studies. It evaluated about ten different studies, and there is also a significant placebo effect here. So the jury's not completely out, but there are plenty of studies, in my literature, that support it to use for the treatment of LPR. I'm not going to go into all that today. But the reality is that PPIs are FDA approved. They're out there. They're easy to get. All you have to do practically is ask your physician for a prescription.



So what are we trying to target? Here is a reasonably good-looking larynx. So these are the vocal cords. This is the anterior portion. This is the posterior portion, back where your esophagus is, and there's left and right. So this is what I see when I'm looking with a scope. I pass it very carefully through the nose and we get a very nice look at the larynx. And this is a patient with moderately severe laryngopharyngeal reflux. You can see that there's a lot of swelling here, compared to these other pictures. This area back here is all beefed up, and this area down here also looks quite reddened. So this is what we're treating with these proton pump inhibitors. And we're hoping to take a patient that looks like this, and turn their larynx into something more that looks like this normal larynx. You can see the sharpened edges here. So data is out there. We can follow this clinically, and that's usually what we do.



So another reality, when you're considering these drugs that are being developed, as we're hearing about today, and being prescribed, is that really only about one in a 1000 of these compounds that enter preclinical testing will actually make it to human testing, what we call clinical trials. And out of these only about a fifth may actually be deemed safe and effective enough by the FDA to gain FDA approval. FDA approval is the Holy Grail for a pharmaceutical company interested in developing drugs and getting these drugs on the market. Without FDA approval they can't market the drug. Without FDA approval they can't put out ads on the evening news. So this is an example of one of those documents that's available on the website that I showed you. And this is the approval letter for Prilosec OTC, which came out in 2003. It was written to Proctor and Gamble by the FDA. And I just want to point out this area down here. This is an approval letter for this drug, given at this dose, used for this indication. So they're very, very specific. The drug companies can't -- they're not allowed to market its use for other indications. But this is really, really important information for the company in order to be able to put this out and actually start to see some return on their investment for research and development. So again I just want to point out, this is for omeprazole twenty mg, for the treatment of frequent heartburn.



Now here's some information that -- you saw a different form of this earlier. Esomeprazole also has approval by the FDA for a number of different indications, but also for the treatment of GE reflux disease. And what I want to point out here is that in the design of these clinical trials they look very carefully, strategically, at how can we show that our drug is better than the competitor's? So clinical trials, when they're designing these, are generally broken down into three phrases: phase I, II and III. We have phase IV, which usually occurs after the FDA has given approval for a compound. Phase I is used in healthy volunteers. The endpoint is really safety. They want to know what are the side-effects. Is this a drug that we can actually give to the public? And they also use a lot of studies to determine metabolism and excretion of the drug at this point in development. And these are usually smaller studies, and they usually only need anywhere from twenty to 100 patients for these. And then once they pass into phase I -- oh sorry, let me say just a little bit about safety.



Some of the side-effects -- in the industry they call these adverse events. This is a terminology that the FDA uses as well, and it really is just something to mean a side-effect. Somebody had something that was out of the ordinary and they happened to be taking the drug. So you have to at least consider that it may have been considered by the -- it may have been caused by the drug. A serious adverse event would've led to some sort of damage to the patient, or extended care such as hospitalization or a surgery. And these AEs, adverse events, need to be reported to the FDA during the clinical trials. So they're watched very carefully, and they can't proceed through to the next phase of clinical studies if there are too many adverse events or, for example, if they have deaths; those usually tend to raise really big red flags. And again, all this information's available on that FDA website. So if you want to see how a drug was developed and what adverse events occurred during its development, you can see that there, for the most part. There's a lot of internal stuff that's not going to be on there, but for the most part they have to report these AEs to the FDA as they're going along.



In phase II they're looking for effectiveness or efficacy. The preliminary data generated generally is for effectiveness of the compound for a particular disease or a condition. So this is where we start to get into specifics about which dose of omeprazole am I going to choose if I'm going to compare this to esomeprazole? They can compare it also to placebo. They can compare it to a different drug. And again they're looking at adverse events, they're looking at safety, but they're also starting to now look, am I actually targeting the disease? The study sizes are usually bigger. And then finally, if it gets past phase II, you can get to phase III. And if anybody watches the stock market, looking at drug companies or biotech companies, phase III is where things really -- you'll hear about this. When Wall Street hears a whisper that a drug's not going to make it through phase III studies, or it may not get FDA approval, you can tank the price of the stock. It's really quite remarkable to watch, as a clinician. Because again I don't really have any vested interest in any particular company, but I just find it fascinating that even just a suggestion will actually impact things economically, so importantly.



So here again we continue to look at safety and effectiveness. We may study different patient populations. They may look abroad to do some of these studies. India is a place where a lot of studies are being done these days - Eastern Europe as well. They may look at different dosages, and they may combine this with other drugs. These are usually very large studies. By this time they've gotten to -- they want to see clinical differences that are so subtle that they have to have very large numbers of patients to make this statistically sound. So they cost -- these studies cost millions and millions of dollars too. So these companies are very, very invested. They want to make sure that they get return on their investment.



And then phase IV occurs after the FDA has approved a drug. These are post-marketing study commitments, and these are commitments by the sponsor, the person actually doing the studies and marketing the drug and selling it, that they will provide additional information to the FDA about the product safety, efficacy or perhaps its optimal use. And more recently we're hearing about phase zero trials in the cancer research literature -- my area of research is in cancer -- and phase zero trials are exploratory, first-in-human trials. So these are designed to speed up development of promising agents. As we know, cancer is a very big problem in this country, especially with the population aging, and it can kill very quickly. So these trials are designed to establish very early on whether an agent behaves in human subjects differently than it was expected from the pre-clinical studies. And that does happen sometimes. So you can imagine if you spent months and months planning a phase I trial, only to find out, when you put it into humans, that it reacts -- it doesn't behave the way you expected, based on how it looked in mice, you can lose a lot of money and you lose a lot of momentum. So these phase zero trials are being put more and more into play.



So back to our clinical model. So some of the studies that we already heard about from Professor McBride were to determine whether these drugs actually worked in patients with reflux or not. So, and I'm not showing all the studies, by the way. I just have to disclose that. I'm only showing you a few studies, just to demonstrate what type of data I have to actually take into consideration, as a clinician, when I'm looking critically at how the studies were done; whether I think the drug's actually going to be helpful for my patient or not. So this was in a group of patients, a particular patient population, with already established erosive esophagitis; sounds pretty bad. It's pretty uncomfortable. They studied these patients with esomeprazole and omeprazole, at similar doses, with similar numbers of patients in both arms. They found no statistically significant difference in the symptoms of heartburn, sustained resolution of heartburn symptoms in this group of patients. Now if you read the esomeprazole package insert, they state that they chose omeprazole twenty as the competitor dose, because it's the FDA approved dose for this indication; which it is. So then I went to the Prilosec package insert to find that they did look at 40 mg when they were looking at these types of symptoms. They found that there was really no improve -- no benefit to using the higher dose. So they went with the lower dose when they went for FDA approval. So that's, I think, partly why they chose 20 mg.



And then also, if you look more closely, there's also they lose some linearity between plasma concentration versus areas under the curve, which is plasma concentration in the patient over time. So it becomes less and less predictable, as you go up on the dose, how much drug is actually being seen physiologically for the patient. So once they get out of that predictable linearity, things become a little bit iffy, and you could potentially get into adverse effects. So I think that's part of the logic for why they chose that. But again, I'm not trying to make an exhaustive argument for what should've been done or against what was done; just to give you some idea for what we have to deal with as clinicians.



So then finally these are some of the three studies that Professor McBride has already alluded to, showing that there are variable results. This one showed a definite statistically significant improvement in healing of erosive esophagitis for esomeprazole over -- any dose of esomeprazole over omeprazole. But two other studies actually showed that there was no difference. So there's another bigger study as well that's not included here.



Professor Michael McBride: Actually it's the next one.



Dr. Dianne Duffey: Oh I'm sorry, yeah there it is, okay. Yeah, so again you just kind of have to look at the literature and say okay, I'm convinced or I'm not convinced. I don't know that there's any right or wrong. But the FDA also has a role to play here, and they really vet these studies quite well, and they're not allowed to market if they don't feel that -- if the FDA doesn't feel that there's sufficient evidence there to support it. So in my patient population, who have LPR, a large number of these patients have heartburn. They may or may not be treated for the heartburn already. I may be the first physician they see, and I may say, "Well do you also have heartburn symptoms?" Those would be GE reflux type of symptoms. So I can end up treating both of those if I put my patient on one of these PPIs for laryngopharyngeal reflux. So if I've just told you that these drugs are approved for a specific indication, maybe at a specific dose, how is it I'm able to use it for something that it's not really even approved for? And this is where we get into off-label use of drugs.



So once a drug is approved by the FDA for any indication, I as a clinician can write it, as a prescription, for another indication. But the FDA has put out these guidelines, such that good medical practice and the best interests of the patient should prevail here, so that we're using legally available drugs -- it's an important term -- according to the -- and devices -- according to my best knowledge and judgment as a clinician. If we use a product for an indication that's not in the approved labeling, I have the responsibility to be well-informed about the product, to base its use on firm scientific rationale, and on sound medical evidence. Now in my literature there's a lot of sound medical evidence supporting the use for PPIs in laryngopharyngeal reflux. But I just also --



Professor Michael McBride: Can I ask you a question about that?



Dr. Dianne Duffey: Yeah.



Professor Michael McBride: You say you have a -- it's not underlined there but you say there's a responsibility to maintain records of the product's use and effects.



Dr. Dianne Duffey: Uh-hum.



Professor Michael McBride: Does anyone come around and collect those records and try to make something of them?



Dr. Dianne Duffey: I've never been contacted by anybody.



Professor Michael McBride: Are people pretty conscientious about doing that?



Dr. Dianne Duffey: I think that we're conscientious about recording whether there've been side-effects at the given dose. We're conscientious about recording whether there's been improvement in the symptoms. But I don't have a detailed questionnaire, in general. So --



Professor Michael McBride: Yeah. Because one would think that this would be sort of like a wiki, that even though you can't afford to do a big study, that if you collect enough observations from enough people, you could make something out of it.



Dr. Dianne Duffey: We could, we could. But then you're in to experimenting, if you're doing it with the intention of actually showing or disproving a hypothesis.



Professor Michael McBride: Well but you said you have a responsibility to maintain the records. Why? [laughs]



Dr. Dianne Duffey: We do. No, I agree with that. I agree. But we're not really obligated to submit it to an IRB. If we were doing it with the intention of definitely showing a difference one way or another, then that would be considered an experiment. So I would be obligated to run that through the institutional review board, which is there for the protection of human subjects, and to make sure that any studies that we're designing are legitimate, that they're ethical, that the patient has been given the opportunity to ask questions. They can refuse, they can drop out of the study. So, but they're all important issues that I think would be in the realm of, at that point, doing actual experimentation. But it's a really important -- it's an important point.



But we can use these drugs for off-label marketed uses. Now the drug companies importantly are not allowed to market; they're not allowed to advertise for off-label uses of compounds. So again I have a duty, as a physician, to evaluate this literature critically and to really be able to validate or ascertain the validity of the research that supports my choices.



So just a very quick word about marketing. As I mentioned, my research area is in oncology. I've been involved in some clinical trials with oncologic drugs for head and neck cancer. And I get these emails all the time. Marketing is a really big deal when it comes to these compounds. Just to give you an idea, this one in particular is about how do you market your oncology products? And this is an actual meeting, to which I was invited. And one of the teasers here is that "in such a crowded and competitive market, the ability to differentiate your product has never been more important." So this just gives you some idea of what the thought process is and what the general culture is. And this is again a wordy slide, but just to point out -- and this comes from Nature News; it's fairly recent -- that industry is really starting shift attention now to other areas. So cardiovascular drugs, which cardiovascular disease is really the foremost killer in our country right now, but they're really shifting their attention now to drugs where perhaps they can make another blockbuster drug; and that's usually always what the impetus is, return on investment. So oncology drugs; immunology, which would be something for rheumatoid arthritis, for example; and neurology are going to be a very, very big focus of attention of pharma industry. So as you watch the evening news, or whichever channel you might be tuning into, you'll be hearing a lot more about this, and I think a lot more direct to patient marketing is going to take place. So again, this has been a lot of fun for me to participate, and I thank Professor McBride also for stimulating me to think a little bit more deeply about proton pump inhibitors and how I use them.



Professor Michael McBride: Great. Thank you.



[Applause]



Professor Michael McBride: Leave it on, there might be some questions. Are there some questions for Dr. Duffey?



Dr. Dianne Duffey: I'd be happy to try to field any questions, as long as they're not too hard. Yeah?



Student: I am an international student, and I have noticed in the marketing for drugs that they have on television here that they always give the full, like two paragraphs about the side-effects of the drugs, on television. Is that required, and why? How is that regulated?



Professor Michael McBride: Right, so the question is from -- you said you're an international student?



Student: Yeah.



Professor Michael McBride: She's from Australia.



Student: I'm not used to that at all. So it's sort of like whoa.



Dr. Dianne Duffey: Right. So she's noted when these ads come on TV that there's a very long disclaimer afterwards about the side-effects. And I'm not actually an expert on how the marketing of these drugs is done. I'm going to speculate here, but I'm pretty confident that this is correct, that the FDA mandates that they put that out there. So if you're going to make a claim about a compound and how great it is, you also better tell the other side of the story. You better tell them, "Listen, we might fix your reflux but you're going to get diarrhea." You know? And it's kind of funny actually, because if you look at these and you say, "Okay, I'm going to take that drug to fix my runny nose, but I'm going to get diarrhea and muscle pains and all these other things." So it really -- I think it's probably mandated by the FDA, and the FCC also may have something to do with that. But that would be my speculation. Yes?



Professor Michael McBride: Sam?



Student: I was wondering what your opinion is about -- I know recently one of the big things was like that Vioxx isn't safe. Do you think that clinical trials, they should do longer term studies? Because I think -- I don't know of other specific examples, but a lot of drugs like it comes through after a long time, long-term use problems --



Dr. Dianne Duffey: Right, and then we wonder as consumers, is this safe? If I'm going to keep taking this drug for years and years, am I safe, or are my progeny going to be safe? It's a good question. And that's where some of these phase IV studies -- they're called post-marketing studies -- come in. I was involved with one clinical trial where a phase IV study actually was mandated. And so for years afterwards that drug company had to continue to pour resources into contacting patients who had taken the drug -- "Are you still taking it? If you're still taking it, what side-effects are you having?" So those are mandated on some level. And the FDA really plays the key role there. And that's been my experience. But it definitely raises questions. And the Vioxx trial just was devastating for Merck. And I think a lot -- they didn't see a lot of that coming. So I think for the pharma companies -- correct me if I'm wrong; you may have more insights -- it's probably, it's a decision that they have to weigh. How much resources are they going to continue to pour in? And that would, on paper, it would take away from their profits. So I'm sure there's a lot of internal debate about it. But yeah, that's where we're thankful that we have the FDA looking out. A lot of countries don't have an organization such as the FDA. You can market anything you want, for almost any indication, and whatever dose goes. So I think we're very fortunate in that respect. I'm going to turn the mike over.



Professor Michael McBride: Well Dana's got a question.



Dr. Dianne Duffey: Oh sorry, I'm sorry.



Student: I've heard there can be cases where a drug was actually more effective for a certain off-label use than the use that it's actually FDA approved for, but it's been FDA approved for something that has a bigger patient population, so you can potentially make more money. If off-label uses can't be advertised, how do you as doctors find out about that and evaluate that?



Dr. Dianne Duffey: So the question is if an off-label use for a drug can't be advertised by the company, how do I as a physician hear about it? I would hear about it from other clinical trials that were done, maybe have -- may have been supported by the company; so as we saw an example of earlier. But I look for literature within my field. So laryngopharyngeal reflux is a perfect example of that. It works pretty well, for this drug, but you do have to take it for a longer time period, and sometimes you actually have to take it for higher doses than is recommended for GE reflux alone. But it works pretty well, and I hear patients all the time say that it works. I have other patients who say it doesn't work. So yeah, we just have to look at our own literature. And there are smaller studies; they may have fifty, seventy patients. But in an academic institution, somebody who, with an interest in that particular area, may run their own study, because they want that information and the information's not out there. So we look for good solid data to help guide our choices that way.



Professor Michael McBride: Kevin?



Student: As a physician, about off-label use, do you ever use drugs in an off-label use and take advantage of their side-effects, in that sense?



Dr. Dianne Duffey: Okay. I'll give you -- well yeah there was. So when -- I don't, I don't really have any drugs that I use that for. But there's one drug in my field; Botox, for example. Botox was being used to help weaken the vocal cords. Okay? It's been used for that for years, in patients who have speech problems and don't have good control over their vocal cords and the function. So you can use Botox injections. Then Botox injections came along for use in the face, and you read about this on page six all the time; it's hilarious. But what they've also noticed was that migraine headaches got a lot better. So you can use that, you can use Botox to treat migraine headaches. And there actually are phase III trials going on right now by a colleague of mine down in Manhattan, looking at exactly that question. So yeah, you can use them for -- and you can take advantage of some of these side-effects. But I don't know of any specific negative side-effects that were used. But that just happened. You may see some beneficial positive side-effects that the pharma companies didn't even know existed; they didn't even target that. So we continue -- so that's where it's important to do clinical trials, as a clinician. So I'm involved in clinical trials as well. And that's sort of an ongoing duty I think, certainly being an academic institution, that we need to also play an important role. Okay? Thanks.



Professor Michael McBride: So if you have a few more questions, come on up.



Dr. Dianne Duffey: Yeah.



Professor Michael McBride: I'm sure you have to get back to your practice. But the time's up now, and thanks again Dr. Duffey.



Dr. Dianne Duffey: My pleasure. Thank you.



[Applause]



[end of transcript]

Lecture 31
Preparing Single Enantiomers and Conformational Energy
Play Video
Preparing Single Enantiomers and Conformational Energy


After mentioning some legal implications of chirality, the discussion of configuration concludes using esomeprazole as an example of three general methods for producing single enantiomers. Conformational isomerism is more subtle because isomers differ only by rotation about single bonds, which requires careful physico-chemical consideration of energies and their relation to equilibrium and rate constants. Conformations have their own notation and nomenclature. Curiously, the barrier to rotation about the C-C bond of ethane was established by measuring its heat capacity.




Transcript



November 19, 2008




Professor Michael McBride: Last time we heard about some medical considerations regarding stereochemistry, and we're going to just start today with a little bit about legal considerations about stereochemistry. First, the doctor gave a disclosure last time about her connections. So I'll do the same thing. I've served as a scientific consultant or expert witness to a number of pharmaceutical companies, including Eisai; which is how I knew about that 7389. I take Lipitor, and served as an expert witness for a generic competitor in a case involving the validity of a Canadian Lipitor patent by Pfizer. And my only connection to AstraZeneca and omeprazole is as an occasional consumer of Prilosec. Okay?



This book here is very heavy. It's about the stereochemistry of organic compounds. And interestingly, it's from 1994; so it's less than fifteen years old; and it's still the bible. And only eight pages of it are about biological properties, and just one of those pages had anything to do with legal things; which had to do, incidentally, with cocaine. It turned out that somebody was on trial for cocaine and the defense had the idea that the prosecution had not tested the cocaine to see which enantiomer it was, [laughter] as to whether it was -- might have been artificial; because the enantiomer of cocaine, there's no law against. It also wouldn't work. But anyhow, they hadn't tested it. So they thought they could get their client off. I didn't actually look up the case to see whether they got off or not. But that's an interesting stereochemical application. But one that made a big difference was the court rejecting a suit about Nexium marketing. We saw last time, and you did a problem for today, about Nexium and whether the advertising seemed convincing to you; or fair, for that matter. There was a big suit about that, and this -- in 2005 there was a news report about that.



"A federal court in Delaware dismissed a class-action lawsuit" alleging AstraZeneca has had misleading marketing of Nexium and thus added billions to healthcare costs. And the "district court judge, Sue Robinson, rejected the suit brought by the Pennsylvania Employee Benefit Trust Fund on behalf of entities that foot the bill in healthcare plans. According to the health plan paying organizations, the big difference between the two drugs is not effectiveness, but advertising." (So you presumably looked at some of this and have your own opinion about that.) "By selling doctors and patients on the idea that patented Nexium is better than Prilosec, which faced generic competition, AstraZeneca was able to preserve billions in sales."



So, for example, the year Nexium was introduced, their advertising budget was 260 million dollars, and in 2005 it had declined slightly to 226 million dollars. But the sales were 5.8 billion dollars. So the advertising was negligible, compared to the cost of advertising [correction: compared to the income from sales]. "Judge Robinson said the courts should defer to the U.S. FDA in weighing the differences between drugs, and that since the FDA cleared Nexium's label, the lawsuit could not stand."



So that's interesting, it shows the importance of the FDA. Okay, back to chemistry now. In order to market Nexium, the single enantiomer, AstraZeneca had to, of course, prepare it as a single enantiomer, which meant either resolution -- so separation of the right and the left-handed forms -- or else preparation of only the one form, the (S)-omeprazole. I noticed that the doctor said eeso-meprazole. Maybe they all pronounce it that way, but the reason it got the name is because it's S; so I say esomeprazole. So how are you going to do it? Resolution. You could do a Pasteur conglomerate kind of thing. You could make temporary diastereomers and separate them, perhaps by crystallization. Or you could destroy one enantiomer, react the racemate with a resolved chiral reagent that would react with only one, or maybe a catalyst that would catalyze the destruction of only one, like an enzyme. Or you could do it nature's way and prepare only one enantiomer, either from a resolved chiral starting material, or using a resolved reagent or catalyst. And they tried all of these. So let's look at it.



First, resolution of omeprazole by chromatography. So they had silica -- which is what you use probably in your chromatography -- but they coated it with something that was a single enantiomer. Now how did they get something solid that was a single enantiomer that they could coat it with? They got it from nature. It was cellulose; trisphenylcarbamoylcellulose. And they did this in 1990, when they were looking forward to the time that they'd lose patent protection on the racemate. So there's cellulose and phenyl isocyanate; the thing with two double bonds in a row there. Now, what would make, do you think, phenyl isocyanate reactive? Do you see any functional groups there? Any ideas?



Student: A double bond of carbonyl.



Professor Michael McBride: Okay, carbonyl double bond. Good. And how about the cellulose? What's going to make it reactive?



Student: Hydroxyls.



Professor Michael McBride: Rick? What makes cellulose reactive, any functional groups?



Student: It looks like mostly hydroxy groups.



Professor Michael McBride: So unshared pairs on oxygen. Okay, so an unshared pair on oxygen can attack the carbonyl. Now, you could in fact use the C-O or C-N double bond. And, in fact, we've seen this before. This, remember, is how urea got made, from ammonia and the corresponding acid. This is just a phenyl derivate of that acid. So it does the same thing. And, in fact, you can do it with all the OHs in the group. So you put that carbamoyl group, as it's called, on all the oxygens. So that is the chiral stuff that was on the solid that the chromatography was conducted with.



Now, they weren't able to get very much. They did six injections with chromatography, and were able to get three 3 mg of the dextrorotatory and 4 mg of the levorotatory stuff separated by the rate at which they would go through this chiral column. Now that wasn't enough for any kind of human testing or anything. You need 20 mg, remember, for a human dose, for just one dose. But what they could do was measure how fast racemization was; because obviously if the stuff racemized immediately, it would be useless to have a single enantiomer. So they found out that the half-life for racemization was an hour at 75°, which they then extrapolated to what it would be at a body temperature to be 100 hours; so plenty of time. It's not going to racemize just spontaneously. So at least that was not a trouble. So this very first quick way of resolving it allowed them to see that it was maybe worthwhile continuing.



Okay, then they tried reversible formation of a crystalline mandelate ester. Now that red stuff there is mandelic acid, which is a natural substance. So you have that as the pure S enantiomer. And now you want to react that with the omeprazole, with racemic omeprazole. But you do it with an intermediate compound, formaldehyde, such that the formaldehyde, the carbonyl group, gets attacked from both sides, one at a time, by the nitrogen on the omeprazole and the oxygen on mandelic acid, to put a CH2 group between them. And we'll talk about that a lot, that kind of thing, when we get onto the chemistry of aldehydes and ketones. But anyhow, they could do that. Now they have what? They have two things. What's the stereochemical relationship? Because they had racemic omeprazole they had reacted from (S)-mandelic acid. What are the two -- you have two compounds, what's their relationship?



Student: Diastereomers.



Professor Michael McBride: Russell? They're diastereomers. So they're different. So you can crystallize them apart. Okay, so they did that. And then the important thing was that it's a reversible reaction. So they could go back again, adding water, and get the omeprazole back, after the separation. And now they have both the R and the S versions of them; both the d and l, or the dextrorotatory and the levorotatory forms of omeprazole, in separate bottles. And now they have hundreds of milligrams. So it's not enough to do a large-scale test, but they can do some biological testing. And with it, they found that the R was four times as active as the S, in rats that they studied. But lo and behold, when they got into humans, it turned out it was reversed, that the S was more active than the R. So that's why the drug is (S)-omeprazole.



Okay but that's -- but you're not going to do this to make enough to sell to the public, especially because you don't want to throw away half the material you make; both because you're polluting the environment and because you're wasting the money you put into making it. So they developed a chiral catalyst involving titanium. So there's a titanium bonded to four alcohols; or actually one of them bridges to another titanium. But anyhow, you can do something called ligand exchange. So the electrons in the bond go onto the oxygen. So you have RO- and Ti+; you've broken the bond. Now that goes away and a different one can come in, a different alcohol, and form a new bond. So that exchanges the ligands by losing one and putting on another one. Okay? Now you can do that trick with diethyltartrate. So two of the oxygens, the alcohol oxygens of tartaric acid, can exchange for two of the alcohols that are on the titanium, and you get this compound of titanium. Now what you really want is the titanium. That's what's the catalyst. So you can lose another RO- and you get this Ti+. And now's when the catalysis begins, because you bring in a peroxy compound, a hydroperoxide. So instead of an alcohol, there are two oxygens there. And that can do the same trick and combine with the titanium. So you have this compound. But now you have a different functional group there. You have the O-O bond. And what makes the O-O bond reactive? We've seen this before. Kevin?



Student: The high nuclear charge gives the low LUMO.



Professor Michael McBride: And what is the low LUMO?



Student: σ*.



Professor Michael McBride: Right. So σ* is going to be reactive there. Okay? And you can bring in a sulfur -- right? -- with its unshared pairs, and they'll react, make a bond, break a bond. Right? So you get the sulfur attached to the oxygen, and it's S+, since it has used its electrons to make that new bond. Now, that thing can come off, the same way the compound above it did; the RO came off, the sulfur compound could come off. And that's the sulfoxide; remember, that's the important group in the (S)-omeprazole; that's what sets things going. Okay, so we've got it. And now that compound on the lower left, the titanium compound, is the same as the one on the upper right. So we've made a cycle. Okay? So that means we can look at the cycle this way, a catalytic cycle. We bring in ROO-. It combines with the little machine. Then S comes in and RO- comes out. And then S double bond O comes out. And you get the catalyst back again. You can start at the top and go -- so you go round and round and round and round, and every time you have the peroxide come in, the sulfide come in, the RO- comes out, and the SO comes out. So essentially the ROO gives one of its Os, the red one, to the sulfur. Right? But it's catalyzed by this. Now why is it important for their purposes that this thing be involved in catalysis? We looked last time that you can put an oxygen on sulfur, just with peroxide. You don't need the catalyst to do that, although it's faster with the catalyst. Did they do it just to gain speed, or can you see something else? What's special about the catalyst? Lucas?



Student: Well it's one enantiomer.



Professor Michael McBride: Pardon? It's one enantiomer, right? It's a chiral oxidizing agent. The peroxide alone would've been an oxidizing agent, could've done the trick. But the chiral oxidizing agent means that it will discriminate between the two hands of the product. Right? It could attack one of the unshared pairs here, or it could attack the other. And those are enantiomers of one another. But when it's combined with this thing, that has tartaric acid on, which makes it chiral and of one hand, then those are diastereomers, when it attacks this one, or when this one does the attacking. Right? So they won't be 50:50.



Unfortunately, it was 50:50. It just happened to be the same rate. So did they give up? No, they started fiddling. And one thing they did was to add diisopropylethylamine. Why? Who knows? Right? But they discovered that when you do that, then you get what's called 94% enantiomer excess. That means the desired enantiomer is there at 97%, and there's 3% of the wrong enantiomer. Right? So there's an excess of 94% of the enantiomer you want. So now all you do is crystallize it, and it gets purified. You get rid of the 3%, and you have the stuff you want. And now you can sell it to the public as Nexium. So that's the way you do it practically, right? When you want to make a lot of stuff you make just one.



Okay, so that's what we're going to say about configuration. And now we have one "C" to go. We talked about composition, of course, from the time of Lavoisier; about constitution from the time of Couper and Kekulé. But then there are distinctions based on this bonding model. Note that constitution is already involving the bonding model. But something beyond constitution is how things are arranged in space. So configuration is obviously also a question of the model that you make for bonding, and conformation, which is the "C" we haven't talked about yet. And those last two are stereoisomers, because they have to do with arrangement in space; which constitution, as you remember, doesn't. So the difference between configuration and conformation, both of which have to do with arrangement in space, is that the configuration, you have to break bonds to go from one isomer to another, and in the conformation you just rotate about single bonds to get from one isomer to the other. And, in general, it's hard to break bonds and easier to rotate bonds. So typically configurational isomers last longer than conformational isomers; but not always, there can be special cases.



Now all things that you call isomers represent local energy minima. So the molecule can sit there at that geometry -- and of course it has to be vibrating, at least zero-point vibration. So the atoms are moving, but they're in a well; they stay in the same relative locations, more or less, right? They vibrate in there. But then they can get out of that well and go to a completely different well, and vibrate there. So not all the different phases of vibration are considered different molecules; that would be way too much, to call those all isomers of one another. Those are all one isomer, and these are all another isomer, but there's some barrier to get from one to the other; so you can have one or the other. And the barriers are typically higher for configuration and lower for conformation.



Okay, now we've seen that in the nineteenth century stereochemistry was qualitative. It was a question of counting isomers; not measuring something but just counting how many isomers you could get. Very qualitative, right? But conformation, which didn't come along until much later, which involves rotational isomerism about single bonds, is much more subtle, and it requires quantitative thought about equilibria, rates and energies. It's not just this qualitative counting of isomers. So notice that this conformational analysis came relatively late to the timeline we're talking about. All the other stuff we're talking about happened before -- most of the other stuff, with respect to structure -- happened before 1900. But it was only in 1950 that conformational analysis began, and then twenty-five years later that people started doing significant, what's called molecular mechanics modeling, treating molecules as if they were springs and so on. And we'll talk about that a little bit later.



So remember there you are, and we can trace your lineage back to P.D. Bartlett at Harvard, who was my Ph.D. supervisor, and his Ph.D. supervisor, J.B. Conant. Now, those guys were involved in a special flavor of organic chemistry which influences you. It's called physical-organic chemistry. It's interested not so much in making new compounds as in why compounds behave the way they do. So what you've suffered through this semester is because of these guys. Right? We talk much more about why things work -- how bonding works, how to recognize a functional group -- than we do with memorizing reactions that will get you from here to there. Okay, so Conant was one of the founders of this discipline, because he did a Ph.D. with two advisors, one of whom was a physical chemist and one who was an organic chemist. So if we go back a little bit here, you see the blue, the organic chemist, was Kohler, who we've talked about, and the red was T.W. Richards, the physical chemist, who himself studied with Wilhelm Ostwald, who together with van't Hoff and Arrhenius -- these three guys here, van't Hoff, Ostwald and Arrhenius -- those are the guys who founded physical chemistry as a discipline. Okay? So we have a lot more physical chemistry in this course than you normally would in an elementary organic course.



Now, if you go a little bit further in the family tree, you get G.N. Lewis as well, whom we've of course talked about. Now, in the late nineteenth century, when these guys were founding physical chemistry, organic chemists focused their efforts on molecular structure, on these things having to do with configuration and enantiomers and so on. But physical chemists focused on energy, sometimes to the exclusion of structure. In fact, Ostwald -- oh I forgot to bring his book, I was going to bring to show you; but I have a page from it here. He wrote lots and lots of books, Ostwald, who's like your four times great-chemical-grandfather, or something. This Principles of Chemistry -- which he gave the not too modest title An Introduction to all Chemical Textbooks. So this is what, if you read it, then you were ready to read any other chemical textbook. So it was a general view, in 1907. And on page 421, there's a very long footnote. And if you look at a little bit of this footnote, it says: "Dalton, who developed the law of combining weights on the basis of an hypothesis he had proposed about the composition of matter from atoms, at first took hydrogen as unity, since it had the smallest 'atomic weight', i.e., combining weight." And so there's the word 'atom'. And that's the only place, in 554 pages of this book, that the word atom occurs. It's in the subordinate clause of a footnote, because, at this time, Ostwald didn't believe in atoms. He believed that philosophically, since you couldn't see them, you shouldn't talk about them; so you should do things that you can deal with. And energy was such a thing, he thought. So physical chemists, these guys who founded physical chemistry, focused on -- of course, van't Hoff is interesting because he's the guy that had the idea of the tetrahedral carbon. So he didn't sign off on this, together with Ostwald. But Ostwald was much more influential, he had a much bigger school of students, and so he was -- it was only two years later that he finally accepted the existence of atoms. So I'm just drawing for you the contrast between what organic chemists were thinking about and what physical chemists were. And now, to deal with confirmation, we have to become a little physical chemist to see the influence of energy.



So here's the dedication of Sterling Chemistry Lab in April 1923, and there's a group of people that we'll look at. And all of these guys here were students of Ostwald. So he had a very broad influence. The one there is T.W. Richards; got the first American Nobel Prize in Chemistry. Okay? And he, remember is your something-grandfather. Okay? So then Donnan, the next guy -- well there's a guy from Canada, a guy from Britain, and the guy on the left there was a student of Richards, so a grand-student of Ostwald; with his cigar, as he always had. And this guy is Wilder Bancroft, who was at Cornell, and who founded and owned, for that matter, the Journal of Physical Chemistry. And he wrote an obituary of Ostwald, in 1933, in which he said: "Ostwald's gift for leadership showed itself in the way his pupils regarded him all through their lives. They usually believed what Ostwald said, even when they knew he was not right." Okay? And no doubt atoms were what he had in mind; that Oswald didn't believe in atoms, but his students still believed in atoms. Okay, but they all believed in energy.



Now, did anybody walk past here this morning? You recognize this place? Okay, you take a different path. Okay, but have you noticed what's off on the left there? "Here Stood the House of Josiah Williard Gibbs, Class of 1858, Professor of Mathematical Physics." So Yale has a really special connection to energy. There's Gibbs as a freshman, and here he is in later life. He spent his whole life here at Yale, except for one year he went to Germany. So physical chemists were, and still are, quantitative about equilibrium constants and rate constants. So big K is an equilibrium constant; small k is a rate constant. And they're related to energy. So you can call energy E, or if you're more particular you can call it H, for enthalpy, or G for Gibbs free energy; and entropy gets into the argument, as we'll see in a few lectures. So energy determines what can happen at equilibrium; that is, if you have this stuff, is it possible spontaneously to go to this stuff? If you have forever and ever for it to happen, is it conceivable that this stuff will go to this stuff? It can only happen if it's downhill in energy, in free energy. Okay? And the equilibrium constant is determined by -- is related to energy in this way. And here's something that's really handy, and you'll impress subsequent teachers and your roommates and everyone else is you remember this: that if you know the change in energy, in kilocalories per mole, you can say what the equilibrium constant is, because it's 10(-3/4 of whatever that is). So suppose the difference between this and this is 4 kcal/mole. What will the equilibrium constant be?



[Students speak over one another]



Student: It'll be 1000; 10(3/4 of 4); 3/4 of 4 is 3; 103 is 1000. So the equilibrium constant is going be 1000. Now whether it's 1000 or 1/1000 depends on which way you're going. But obviously it's going to favor the one downhill. So you have to keep your thinking cap on to get the right direction. But 10(-3/4 of ΔH), or -- yes, 10(-3/4) is the way to do it. Okay. But you can also say something about how fast it will go. This is just whether it's conceivable that it will go; the big K. The little k says how fast. Because rates -- this is an approximation, this isn't written on the tablets by the angel or someone. But for practical purposes, for our purposes, the rate in seconds -- that is, how many per second -- is 1013. So that's very fast; 1013 per second. But then it gets cut down by how big the energy barrier is you have to get across. Cutting down is exactly the same kind of thing you have in the big K, 10(-3/4 ΔH). So you could again -- that energy is called activation energy. It's how high the barrier is that you have to get over. And you can use the same trick of the 3/4ths. So it's, the rate constant is 10(13-(3/4 of whatever the energy)). So suppose you had a fourty kilocalorie energy barrier you had to get across; suppose you had a forty kilocalorie energy barrier you had to get across. Forty kilocalorie energy barrier, right? How fast would the rate constant be? Becky, why don't you help me? So the barrier is 40 kcal that you have to get across. So now how am I going to plug that in here?



Student: 1013th.



Professor Michael McBride: Okay, so 3/4ths of forty I need, right? So what's that?



Student: Ten -- or thirty.



Professor Michael McBride: Thirty. Okay, so it's 10(13-30). Okay, what's thirteen minus thirty?



Student: Negative twenty-seven [should be negative seventeen].



Professor Michael McBride: Okay, so it's 10-27 [correction: 10-17] per second. Do you think that's very fast, 10-27 [correction: 10-17] ? That's a pretty small number, right? So forget spontaneously, at room temperature, going over a barrier of 40 kcal/mole, for practical purposes. But you have this T here. So if you increase the absolute temperature, then you make that forbidding factor smaller and smaller and smaller, up in the exponent. So if you heat it up, you can make the reaction go. Okay, so anyhow, that's really helpful, to remember this 3/4ths trick. Okay, now conformation involves rotational isomerization about single bonds. So the question is, how free is single bond rotation? Paternó said it wasn't free at all; I can count isomers that way. I can have this isomer, and I can have this isomer, and I can have this isomer -- right? -- for purposes of counting. But van't Hoff says no, don't count isomers like that. Because why? Why shouldn't you count isomers like that?



[Students speak over one another]



Professor Michael McBride: Because they rotate too easily, right? It's like that in his picture. Right? So you shouldn't count this and this and this and this. Okay? Now, there are different kinds of models, something like these, and some of them you can rotate easily, but not others. Right? Which one's realistic? Right? Or this one you can rotate, it's just not easy to rotate -- right? -- it sticks, right? Or this one, remember, that Fischer used; which is good because you can bend it, for his purposes, for drawing two-dimensional Fischer projections. But forget rotating that. Right? It could vibrate, but it can't -- you can't rotate it, it'll always come back. Right? So the question is, which model is right? Can you rotate or can't you rotate, and how hard is it to rotate? Because that's going to determine how fast these things will be.



Okay, now notice that both of these authors drew their pictures with a conformation which we call eclipsed. Now eclipsed means these. It means this, that the ones on the top are right over ones on the bottom. Okay? And the same is true in van't Hoff's tetrahedra -- right? -- right above the others. Obviously you could rotate 60° on the top and everything would be staggered, not directly above. Okay? Notice, incidentally, that from the exam we looked at the Molecule of the Week, for last week, from the American Chemical Society. And notice how it's drawn. Right? It also is eclipsed. This red one here is directly opposite this red one here. This acid group is directly opposite that hydrogen, and this acid group is directly opposite that hydrogen; it's eclipsed, right? And here, this CO2 is directly above that CO2; it's eclipsed, right?



Now you need notation and names to talk about these things. And the projection that was invented for this purpose was invented by someone who sat in the seat that some of you are sitting in. I don't which. Maybe -- he probably, in the course of his undergraduate years and also getting his Ph.D., he probably sat in most of the seats in here; or many of them anyhow. But his name was Melvin Newman. You see he graduated from Yale in '29 and got his Ph.D. in '32, and then he went to Ohio State. And he invented a way of drawing conformations on paper. So there are two carbons, and the convention is, for the Newman projection, that you sight down the carbon-carbon bond. So I'm going to show this, the rotation around this central bond. Okay? So I look right along that bond, and there's one carbon in front and one carbon behind, and each of them has three things on it. Okay? So the one behind, here, has three hydrogens, and the one in front hides the one behind. Okay? But it has three other groups on it too. Everybody see how he's doing this; that that shows this? Well it shows it without the methyl groups here. Okay, and he calls that staggered, right? But you can also have them be right on top of one another, like this. Right? And that's called eclipsed. And that's the way it was shown on the previous slide by Paternó, by van't Hoff, and by the ACS operative who made the Molecule of the Week. Okay?



Now that's not so convenient, because you can't see the stuff in back. So sometimes people turn it a little bit that way, or sometimes they turn it a little bit that way; by which they don't mean that it's not eclipsed, they just mean I want to be able to show both things. Right? Now that's very cumbersome, and the reason you don't mind that it's cumbersome is that things are never eclipsed, they're always staggered. So all those previous pictures were wrong, including the Molecule of the Week, because conformations tend to be staggered, not eclipsed. Okay, but anyhow you can draw them this way. Okay, and that's just a conventional way of drawing something that's eclipsed; because, as you can see, it's not truly 100% eclipsed. Okay? Now, when you have substituents on those things, you can give different names to different phases of rotation. For example, they could be exactly opposite one another, like that; which if this were a double bond you would call trans, or E. Right? But for a single bond, you look at it with a Newman projection like this, and it's called anti. Right? Or you can have it rotated like this, and be what's called gauche. And I've never found out why it's called gauche. Right? But notice something about that. Is this, is it, anti, chiral; is that conformation chiral, or is it superimposable on it? Is there a mirror?



Student: Yes.



Professor Michael McBride: Right, there's a mirror right here. Right? So it's its own mirror image. How about gauche, is it chiral; is there a mirror here? There's a twofold axis here. If I rotate it like this, if I had you close your eyes and did that rotation, you couldn't tell whether I rotated it or not. Right? It's like a propeller. But if I take another one like this and make the mirror image of this gauche one, like this -- here, right? This is a mirror image. Everybody see? Right? But I can't superimpose them. Right? If I superimpose the back three, the front ones are not. Right? So these are chiral. So, with respect to gauche, we have enantiomers. Right? That one as well. And one of them we call minus and the other we call plus. Okay? And now if you have -- the analog for eclipsed is that you can have something that's fully eclipsed, like that, where the special groups are right on top of one another. Or it can be like that. Right? So and that can be one way or the other; and again it's twisted, like a propeller would be. And these are again cumbersome things to draw, with Newman projection. But you don't worry about their being cumbersome because you don't have to draw them usually, because they're not energy minima, they're not isomers, they're barriers between going from one to another. Okay? But again, those would be minus and plus.



And if you want to be really pedantic and meet in a committee of the Pure and Applied Chemistry people, the IUPAC, and decide precisely what you're going to call different things, then you get together and you scratch your head, and for different rotations of the thing in back, with respect to the thing in front, you have these different pedantic names: synperiplanar is the one that we call fully eclipsed; and you can have anticlinal and synclinal and synperiplanar, and plus and minus, and so on. But you never do that really. But it's there if you need it. Okay, so there's a notation, Newman projection, and there's a nomenclature, which can be sort of simple, like on the previous slide, or very complicated, like this, that allows you to discuss these things; to draw and discuss them.



Now, but here's the practical question. Is the threefold barrier -- so if you have ethane, like this, and you rotate this, the energy will go up and down; the energies will be different for different angles of rotation. But obviously there's a certain symmetry, that this, if it's all hydrogens, will be the same energy as this, will be the same energy as this. And for the eclipsed, this will be the same as this, will be the same as this. Okay? So there's a threefold barrier. Now the question is -- two, there are two questions. How big is the barrier? So are we going to be able to have isomers? And, what's the energy minimum?



Now let's just suppose, together with van't Hoff and Paternó and the person who made the Molecule of the Week, that the lowest energy -- if we plot energy versus phase of rotation here -- that the lowest energy is 0°; that's fully eclipsed, like this. Okay? That's 0°. And now we'll start turning it, and the energy may go up, when we go to a staggered, and then down and then up, down, up, down. Right? And then you're back where you started, once you've gone through 360°. Or is there a big barrier? Right? Is the barrier small or big? Now how can you tell? Here's an interesting thing, a way to tell, right? That if there's a big barrier, then the things are going to be quantized in it; the energies are being quantized. But the energy is -- if the barrier is very, very small, then almost any energy you have is above the barrier, and so it's not quantized. Remember, when you get above a barrier you can have energy, any energy you want. It's only when the thing is bounded on both sides that the wave function has to come down and reach zero. Okay? So if the barrier is big, the energy will only be certain amounts, but if the energy's very small, you can have any energy you want. And that you can tell.



That is, in fact, how it was discovered how big the barrier for rotation was, was by measuring something having to do with energy, with heat. Okay? And here's -- so the question is, is the energy quantized in this triple minimum? We've talked about double minima. This is actually a triple minimum. Okay, so they measured in 1936 the absorption of heat by ethane, going down to a very, very low temperature. This starts at 150 Kelvin, but I think they went further than that. So there's a thing called entropy; and we'll talk about that in different terms, a little bit -- in a week or two, after vacation. Okay? Vacation. [laughter] Okay, so there's a thing called heat capacity, Cp, which is how much heat you absorb. So entropy is defined, by people who do thermodynamics, in this way: that you sum up, or integrate, starting from zero Kelvin -- remember Kelvin? -- all the way up to whatever temperature you're interested in. You sum up how much heat is being absorbed; so Cp times [change in] temperature, that's how much heat is being absorbed at each temperature, as you go up. As you raise the temperature, more and more heat gets absorbed. But you always divide it by the temperature, as you go up. So if you absorb heat at very low temperature, that makes a much bigger difference than absorbing the same heat at high temperature, in this quantity called entropy. So don't worry about it, except by this definition for now. And this is very technical and not so very important to what we're doing. So don't get tied up in knots about it. But anyhow, if you have larger quantized spacings, then less heat gets absorbed -- and I'll show you that on the next slide -- and the heat that gets absorbed gets absorbed at a higher temperature. Right? So both those factors mean that the entropy, at any given temperature, is going to be smaller, because you've absorbed less stuff in the numerator, and you've absorbed it at a higher temperature; so you have a bigger denominator for all these contributions.



Okay, so the results in 1936 were this. If you're interested in what the entropy of ethane is, at 298 Kelvin, it turns out experimentally -- that's what they did out at Berkeley -- they found it was 54.8 plus or minus 0.2 entropy units. And then they calculated how high that would be, for different heights of this threefold barrier, and the quantization that's involved. And what they saw was that if the barrier were zero, that number, instead of 54.8, should be 56.4. If the barrier were 0.3, it should be 56.3. If the barrier's 3.1, it should be 54.6; which is within experimental error of the experiment. Right? So that showed that the barrier was about 3 kcal/mole. So it wasn't totally free rotation. It was just pretty fast. So suppose the barrier had been 4 kcal/mole, instead of three; just suppose it were four. How fast -- what would the rate constant be for the rotation? So the barrier is 4 kcal/mole, that you have to get over. How do I deal with that? Sam?



Student: It's this --



Professor Michael McBride: There's a trick, right?



Student: 10 to the --



Professor Michael McBride: 10 to the minus --



Student: -3/4. So 3/4 times four is three.



Professor Michael McBride: Okay, so it's going to be to the -- there's going to be a -3, up in the exponent. And what else is going to be in the exponent? Remember, for a rate?



Students: Thirteen.



Professor Michael McBride: 1013. So it's 10(13-3), which is ten to the --



Student: To the tenth.



Professor Michael McBride: 1010 per second. So it would take 10-10th seconds for it to rotate. That's pretty darn easy, at room temperature. Right? Pretty easy. But not perfectly free, and enough to make this difference in the entropy. Now, let me just see what time it is; okay, we have time to do this. Okay, so we're now going to use that big K, equilibrium constant. So suppose we have something that has -- this rotor say -- that has a barrier of such that the spacing between successive levels is 1 kcal/mole. Suppose it's a harmonic oscillator. We already did that. The lowest zero-point level, you got to have that energy, and then 1 kcal above that, 1 kcal, 1 kcal. Now what we're going to do is see how much energy, extra energy, above zero-point, is there at equilibrium, at any given temperature. So we have some stuff -- at equilibrium there's going to be stuff that has no extra energy, zero-point, a certain amount. But there'll be a certain amount that has 1 kcal/mole, that's in that first level. Now how do I know how much is in that first level, at equilibrium? How do I know the ratio of what's in the top to what's in the bottom, if the spacing is 1 kcal/mole? Is there a way of doing that? You don't have to do the numbers, but tell me what formula you would use, to get the equilibrium constant. Katelyn, do you remember? Sophie? How do you get the big K; how does it relate to energy? Anybody remember? We already did the hard one. Eric?



Student: 10-3/4ths.



Professor Michael McBride: 10-3/4ths of one. Okay. That's at room temperature. It's different at other temperatures. Okay, so we know how much there's going to be -- what fraction of the stuff, of the stuff that's here, is going to be one kilocalorie. Ah, it's going to be 10-3/4th, of this; so something like 1/10th of it. Right? 10-1 would be 1/10th. Okay? And then there'll be stuff here, and the amount here, relative to the amount here -- this is what we're doing. So that ratio, the ratio of 0:1 is going to be the same as the ratio of 1:2, because that one is one kilocalorie above this one. And the same for the third to the second, or the fourth to the third. So then we can sum all these up, knowing the ratios, and figure out what percentage there is of each of these. And it turns out that that 10(-3/4) is 5.6. So there's going to be 5.6 times as much here as there is here; 5.6 times as much here as there is here; 5.6 times as much here as there is here. And if you translate that into percentages, it'll be that there's 81.9% here and 1/5.6, 14.8% there and 2.7, and 0.5%, and 0.1%. But of course this stuff added only -- this 14% added only 1 kcal/mole. This one added two. This one added three. This one added four. So you sum all those up and you get how much energy -- how much heat was absorbed. This didn't absorb anything, it's still in the lowest. This absorbed a certain amount, 14.8 -- right? -- .148, and then twice that, 0.027, and so on. So you can sum it up and see how much heat would've been absorbed by the time you get out there.



Now, if the spacing had been very much smaller, if the spacing had been 0.025 kcal/mole, then we would've made -- how much is there of this that's in the lowest level; how much has one unit, 0.025; how much has 0.05, 0.075, and so on? And we do exactly the same trick; although of course you need a spreadsheet to do this with, because it's tedious. Right? But you would get this. And you see you absorb more heat if you have smaller steps. And now there's 4.2% here, and 0.8% at exactly 1 kcal/mole. But you have a lot of others in here. So 59% of the population is halfway out to the one; and then 33% is more or less one; and 6.6% is like that; and so on. So anyhow, you absorb more heat. And you absorb it at lower temperature too. Right? So both those things build up the entropy. So that one absorbs 364 kcal/mole more -- not kilocalories, small calories -- 364 more than the other one. And that is what gives that entropy difference. Okay, so that's how people knew that there was a barrier to rotation, was by measuring heat capacity. So we'll stop here, and next time we'll address the question of whether the -- now we know what the barrier is, but we don't know what the low energy form is, whether it's eclipsed or staggered.



[end of transcript]

Lecture 32
Stereotopicity and Baeyer Strain Theory
Play Video
Stereotopicity and Baeyer Strain Theory


Why ethane has a rotational barrier is still debatable. Analyzing conformational and configurational stereotopicity relationships among constitutionally equivalent groups reveals a subtle discrimination in enzyme reactions. When Baeyer suggested strain-induced reactivity due to distorting bond angles away from those in an ideal tetrahedron, he assumed that the cyclohexane ring is flat. He was soon corrected by clever Sachse, but Sachse's weakness in rhetoric led to a quarter-century of confusion.




Transcript



November 21, 2008



Professor Michael McBride: Okay, so we were talking last time about rotation in ethane, and the fact that there was a barrier, and that you can measure the barrier. And one way to measure the barrier, the first way that the barrier was measured, was remarkably enough, by the heat capacity of ethane; how much heat it takes to warm it to a certain temperature. So that showed that the barrier was about three kcal/mole, but it didn't say what the geometry of minimum energy was. Was it eclipsed, as shown by the red line here, so that the eclipsed form is the lowest in energy? Or was the favored form staggered? And people differed in their interpretation of that. Ultimately there was some diffraction study, electron diffraction, that showed that it was staggered.



But now the best word on this nowadays is probably from really high-quality calculations. And you can see from this paper, published in the Journal of Chemical Physics in 2003, that there's pretty good agreement. All these various abbreviations are various really high-quality computational results. You can see at the top, our experimental results, from heat capacity. It's about 1000 wavenumbers. A wavenumber is 2.86 small calories, not kilocalories, per mole. So there's heat capacity from infrared spectroscopy, from microwave, from Raman spectroscopy. All these experiments give very close to 2.9 kcal/mole, or about 1000 wave numbers -- right? -- within a couple of percent; even less than that probably. And then all these calculations give about 2.7 kcal/mole. And it doesn't make very much difference. The first one of these -- some of these aren't such very high-quality calculations, but some of them are very, very high quality, the best that can be done. And they all agree pretty much. So it's clear that the staggered form is favored over the eclipsed; despite what everyone wrote in the nineteenth century, when they always wrote them eclipsed, and what whoever it was that drew the picture for the American Chemical Society of their Molecule of the Week, last week, drew too. He drew it eclipsed; or she. Okay?



So anyhow, but then there's the question why? Why is it that the staggered form is lower in energy than the eclipsed? Why is there a three kilocalorie barrier? Well there's different points of view on that. One is that the eclipsed form is unstable; that for some reason having it eclipsed makes it unstable. The other one is that the staggered form is unusually stable. And what does this remind you of, in terms of a question?



Students: Compared to what?



Professor Michael McBride: Compared to what? What's unusual, right? Is what's unusual the eclipsed form or the staggered form? Well from the point of view of the eclipsed form, why would it be unstable? Well it could be that the hydrogens repel one another, because just in space; their van der Waals radii. Okay, but if you think about that, then if you had a methyl in place of one of those hydrogens, it's ever so much bigger. Therefore you should have a much bigger barrier to rotation, of a methyl group in propane. But, in fact, the barrier is almost exactly the same; 3.4, not ten or something like that. So the size of hydrogen doesn't seem to be so very important. Of course, protons repel one another, and the electrons that are in the bonds repel one another; there's electron repulsion. So that might be why you don't want to have it eclipsed; that you want to rotate it a little bit to make it staggered. But that's not easy to do in your head, because there are also attractions between a proton on one side, and the electron on the other, and it's not clear which of these is going to dominate. So anyhow, there's the possibility that it has to do with the eclipsed form being unusually destabilized -- right? -- because of repulsions; that could be.



The other point of view is that the staggered form is unusually stable. Now why could that be? It's that you have a σ on one side and a σ* on the other side of these C-H bonds; and vice-versa as well. And you can get overlap between this HOMO and LUMO in this way. And it turns out, surprisingly enough, that the overlap is bigger when they're anti than when they're eclipsed. You might think, intuitively, that there'd be more overlap when they're on the same side, than when they're on the opposite. But it doesn't take too much work to convince yourself that the other is in fact true, and that this is the better overlap. So you'll get better HOMO/LUMO interaction from one side to the other; and not only for these two hydrogens, but for the other two pairs of hydrogens as well, if the thing is staggered, so that the opposite ones are parallel to one another. So you can get this kind of mixing, HOMO/LUMO mixing, among σ bonds, which is called hyperconjugation. Conjugation is when normal double bonds, p orbitals overlap, from one to the other, and you get resonance structure. The name "hyperconjugation" was created to talk about the same phenomenon when it's σ bonds, rather than π bonds that are doing the trick; and it's usually much, much less important. We'll talk about more examples of this later on.



But at any rate, that's a very different point of view from the first one. Which is it, that the eclipsed is destabilized or that the staggered is unusually stable? Or maybe both. And, in fact, this is a little bit a scholastic argument analogous to how many angels could dance on the head of a pin; and you get people that debate about this. Right? So the point is that fundamentally there's quantum mechanics that controls this, but it doesn't say which one of these things is what. It doesn't divide it that way. So some people say one, some people say the other, and we'll just say maybe a little of both.



Okay, now here's a digression on the question of what's called "topicity"; which turns out to be relevant in a lot of biochemical applications. So it's worth mentioning at this point, this aspect of stereochemistry. Topicity relates to, for example, the question: are two protons equivalent? Now at first glance, that's sort of a stupid question. Obviously a proton is a proton, so they're equivalent. But suppose one is the H of an OH group in ethanol, and the other is an H of the methyl group. Now are they equivalent? Well protons clearly still are equivalent. Protons are protons. What's different about them? What's different is their environment. So it could be that the place they are is different, even though the protons intrinsically are the same, and therefore they could have different properties. Now, that happens to be the case in ethanol. An OH group has a pKa of about 16. It's not very acidic but it's possible to dissociate H+ and leave O-. Right? But an H that's on a methyl group doesn't do that. It has a pKa of about 50. So it's 1034th weaker as an acid, than the OH group. So there's clearly a difference between these two hydrogens because of the place they're in. Right? So that the OH on the alcohol group exchanges readily with acidic water; so you can put deuterium in and out of that, wash it in, wash it out; and you can't do that in the case of methane. So D+ can come in. Low LUMO, attacked by the high HOMO. You get that. And then unzip it the other way. The electrons go back on, H+ comes off. So you can do this exchange. That doesn't happen over on the other case because it's so hard to get the proton off, and the carbon has no unshared pair to pick up the new proton.



So these two protons, the red one and the blue one, the OH and the methyl-H, are called "heterotopic"; hetero means different, and the Greek topos means place. So they're in a different place. They have different properties because they're in a different place. So 'topicity' comes from the same root obviously, and it means the placeness of a group. It could be a hydrogen, it could be another group, two groups, that appear to be the same but are in different places. So this is a pretty obvious case. But there are cases that are less obvious. So protons within the blue group are homotopic, and the green protons are homotopic with one another. Right? They're in the same kind of place. At least they're in the same kind of place if we're discussing constitution, what they're bonded to, and what those things are bonded to and so on. Right? The nature and sequence of bonds are obviously the same between the two green or among the three blue hydrogens. And both of those are different from one another, and from the red hydrogen. Okay, so from the point of constitution, the blue ones are homotopic, the green ones are homotopic. The red one is its own -- right? -- but the green is heterotopic with respect to the red or the blue, and so on. So that isn't telling you much that you wouldn't already have realized easily.



But if you get into stereochemical topicity, it gets more subtle. So let's look at stereotopic relationships among these protons. First, we said that the blue ones were homotopic. But are they really homotopic? Let's consider those two hydrogens. Do they have the same environment? They're clearly the same with respect to constitution, what they're linked to -- right? -- but are they the same stereochemically? Are their environments superimposable, exactly on one another? Yes or no? So are they stereochemically homotopic? Corey, what do you say?



Student: I'd say yes.



Professor Michael McBride: You say they are?



Student: Yes.



Professor Michael McBride: If you took the top hydrogen away, and looked at everything else, and then made another model where you took the bottom one away, and looked at everything else, are those everything elses, the places, superimposable? No, they're not, because one has a hydrogen down here missing, and the other one has a hydrogen up there missing. You can't superimpose those unless you do a rotation. Right? So that's a question of conformation, of rotation, right? Okay, so those are diastereotopic from the point of -- that is, different places -- from the point of view of conformation, but not from the point of view of [laughs] constitution. Right? Sorry, I'm getting my c's confused. It's a long run, from the beginning of September to Thanksgiving. You'll find that this will have been the longest period, unrelieved period, of your time at Yale. It gets easier after this. Okay, so anyhow they're diastereotopic from the point of view of conformation, but homotopic from the point of view of constitution. How about those two? Do they have the same environment, those two hydrogens? Cathy, what do you say?



Student: Yes.



Professor Michael McBride: They're superimposable; that is, if you took the H in front away, you'd have a certain bunch of stuff left; that's its environment. And if you took the H in back away, and looked at the rest of the stuff, are those two remainders absolutely identical; superimposable?



Student: No.



Professor Michael McBride: What is the relationship between them? Obviously that was sort of a leading question.



Student: They're mirror images.



Professor Michael McBride: They're mirror images. So what would you call those two protons?



Student: Enantiomers.



Professor Michael McBride: Not enantiomers. It's their places that are different. Enantio-what?



Student: Enantiotopic.



Professor Michael McBride: Enantiotopic. So the ones at the top are enantiotopic, and the first ones we looked at were diastereotopic. Right? Now, do you care? Is that going to make a difference in chemistry? I say no, and the reason I say no is that it rotates so fast around a single bond -- we talked about that last time; 1010th per second. And when you rotate, you exchange one -- one takes the place of another, right? So they exchange places among themselves faster than almost anything could happen; 1010th per second. So who cares? Right? So in truth they're conformationally diastereotopic, but you don't care about conformation, in this case, because they would rotate so quickly. So they're not going to have experimentally distinguishable properties, probably, unless you can look pretty quick. Okay? Now, so these distinctions are only conformational and erased by rotation; in 10-12 seconds I said here. Okay? Yeah, that's more like it. Okay, now how about the two green ones? What's the relationship between them? Pat, what do you say? What's the relationship?



Student: They're enantiotopic.



Professor Michael McBride: They're enantiotopic. Now, are they enantiotopic with respect to conformation? Can you rotate one and give it exactly the same place as the other one?



Student: No.



Professor Michael McBride: Why not? What would happen if you tried to rotate, to put the back green one in the place of the front green one?



Student: It'd change the location of the OH.



Professor Michael McBride: It would change the location of the OH. Right? So how could you exchange their environments? The only way, we'd have to break a bond and put the other hydrogen someplace else. So what kind of enantiotopic are these? Is it configuration, conformation, or constitution, that you have to change? Can you do it just by rotation? Is it just conformation? No. Do you have to change what's bonded to what, to make one? No. So it's not -- so it's configuration, right? But the important thing is you have to break a bond to change them. Right? So that's not going to happen easily. So the difference between those two -- those are also enantiotopic, but configurationally enantiotopic. So that will last, that distinction, in principle, as long as the bonds endure. So there's quite a difference. Those two could be different. Now let's see if we can see a case where they're different. Now that carbon is not a stereogenic center. Why is it not a -- how do you recognized a stereogenic center? Nate?



Student: By what it's attached to -- it's bonded to four different things.



Professor Michael McBride: You guys shouldn't sit next to each other. I was talking to the other Nate.



Student: It's got two H's on it.



Professor Michael McBride: Can't hear.



Student: It's got two hydrogens on it.



Professor Michael McBride: It's got two hydrogens. It has four different things, right? And this has two of the same. But it's called, for this purpose of topicity stuff, "prochiral"; that is, it will become chiral if two of the things that are the same become different. For example, if one of them, instead of an H, were a deuterium, then it would be a chiral center. So that thing is called prochiral. It would be chiral if the enantiotopic atoms, or groups -- it could be chlorines or it could be methyl groups or whatever; if they were made different somehow, then it would be chiral. Now, so in ethanol, those two green groups are enantiotopic and configurationally. So now "toponymy" -- that's names of places; how do we name them? So we have things -- we have groups that are constitutionally homotopic. They have the same constitution, connected to the same things.



Okay, so consider those, the green hydrogen there, the light green one in front; and I'm drawing a Newman projection also to show it. We have to be able to give it a name so people will know what we're talking about, when we're -- if this turns out to be important for some reason. Okay, so the way we do it is to assign priority, if the hydrogens were different. Right? Then we'll be able to call it R or S, right? Because we already have a scheme for that. But the two hydrogens would have to be different. So let's look at the priority. OH at that carbon -- the prochiral carbon -- is obviously top, and the methyl, the carbon group attached to it, is second; but the two hydrogens come third and fourth. But which one should be third and which one should be fourth? Now here's the rule. You give higher priority to the one that you're naming. Right? So if we want to name this one, we'd give it higher priority than this one. So this one will be three and this one will be four. So here's the lowest priority. You can run one to four, in either direction, high to low or low to high. You get the same sense of handedness. So let's think about this. Here we have -- so if I put my left thumb coming out where this hydrogen is, the number four, and curl my fingers, I go one, two, three,. Okay? So this one would be left-handed, if it were higher priority. Everybody with me on that? So we'll call that, not S, we'll call that hydrogen pro-S, because it will become S -- okay? -- if we make this -- for purposes of naming, when we give it a higher priority. Okay? Now, suppose we look at the other hydrogen, that I've shown in the darker arrow here. Right? It obviously is going to be pro-R, because the same thing would run in the opposite direction. Okay, so pro-R and pro-S are the name that you can give to distinguish groups that are enantiotopic.



Now, is there a reactivity difference that we would care about between these groups? Suppose you bring up a chlorine atom, and it can attack one or it can attack the other. So suppose it attacks the one on the left. Right? Its SOMO mixes with one of the electrons of the H, and the other electron on the H goes on the carbon, and we break the bond and get HCl. This is the first step, remember, in free radical substitution. Or we could do it with the other one. What's the relationship between those two pathways? Can you see? The pathway that happens on the left and the pathway that happens on the right, are they the same?



Student: [Inaudible].



Professor Michael McBride: Well they're not superimposable, okay? So they're not precisely the same. But what are they? Sophie?



Student: They're mirror images.



Professor Michael McBride: They're mirror images of one another. Therefore they have the same energy. So one will be just as -- remember, we talked last time about how fast a reaction goes has to do with how high an energy you have to go to, along that path. Right? And these are going to be the same. So they'll have precisely the same rates. So who cares? Okay? So attacked by a reagent like a chlorine atom, those two versions of the attack, are mirror images, and thus they're identical in their rate. So who cares? But suppose that you use something that's handed, like a right hand, to pull it off? Okay? Now suppose you use the same right hand to pull the other one off? Those aren't mirror images of one another anymore. Right? What do you call -- what's the relationship between these two reactions? What would you call the relationship between them? Angela?



Student: They come out completely different.



Professor Michael McBride: Can't hear.



Student: They come out completely different. So they're diastereomers.



Professor Michael McBride: Right, they're diastereomeric paths. So they're different in energy. So one will be faster than the other. What does that tell you about the metabolism of ethanol? If you've got ethanol inside you, there are things that pull the hydrogen off and make aldehyde out of it. In fact, that's where the name "aldehyde" comes from. Aldehyde is an alcohol that's been dehydrogenated; al-de-hyde. It's interesting, isn't it? But anyhow, this can be -- would your body, if your body was trying to metabolize ethanol, could it distinguish between those two hydrogens?



Student: Yes.



Professor Michael McBride: Why? Who said yes? Sam?



Student: Because whatever enzyme is doing that is going to be a hand.



Professor Michael McBride: Because the enzyme that does it is a hand, a single hand. Right? So it will be able to distinguish between those. Now, can we prove that? Okay? So attacks by a resolved chiral reagent, like an enzyme, are diastereomeric and should have different rates. So horse liver alcohol dehydrogenase removes only the pro-R hydrogen from ethanol. Pardon me, if it removes only the pro-R hydrogen in this deuteroethanol, it should remove the hydrogen and never deuterium, in the case that it's deuterated like this. Right? Messed that up saying it, but you can read it better than I can. Okay? Now, so that would be a great experiment. If you had this compound, (S)-1-deuteroethanol, and you reacted it with this enzyme, if you always pulled off the H and left deuteroacetaldehyde, that would show it was specific. But if it weren't specific, then you'd sometimes get one and sometimes the other coming off, and you'd get different acetaldehydes. Some would have deuterium and some wouldn't, left behind. Okay? Now what's the problem in doing this experiment? Zack?



Student: How do you know that deuterium's right there?



Professor Michael McBride: How do you get the compound with the deuterium on only one side in the first place, so you can do the experiment? And this is where there's a clever way of doing it. Right? That would be the reaction, and you'd get only the deuterium left and never hydrogen left, if it were specific. So that's a good test, but where do you get the starting material? Now actually, the alcohol dehydrogenase is a catalyst, right? And a catalyst lowers the energy of the barrier you have to go through to go across; lowers the barrier. But that means it lowers the barrier in either direction; because you can go either direction across this. Right? So that means that you can -- this is an oxidation; removing hydrogen is an oxidation. But you could also do the reverse reaction; a reduction. That is, you could start with this. So what really does the oxidation -- LAD is a catalyst -- what does it is this NAD+ becoming NADH, taking H- away. Right? So if you ran this reaction backwards, and started with a deuterium there, then if the thing were specific, you'd put hydrogen only there, when you run it backwards; if it's specific taking it off, it'll be specific putting it on. Right? Now how would you know if it's specific, putting it on? How would you know if it did that? How would you know whether, when it does this reaction, it scrambles and puts it in both positions, and then when you run it backwards it scrambles and takes it off both positions, sometimes one, sometimes the other; or whether it's specific going on and specific coming off? Lucas, you got an idea?



Student: You can pull it off and check optical activity.



Professor Michael McBride: Ah, if you could measure the optical activity of this stuff, that would be a good way of doing it. But it doesn't have very much optical activity, if the only difference is between deuterium and hydrogen; not very much.



Student: This is just kind of guess, but if they did it to both things, wouldn't you get one with both and then one with only one? And then if it only went to one side, you'd only get one with one.



Professor Michael McBride: I think you said it.



Student: Probably not right.



Professor Michael McBride: But I don't know that everyone understood what you said. The idea is do both. Put on, and then take that product and take it off again. If it's specific going on, and specific going off, when you go through the whole cycle you'll get back where you started from. But if it mixes up going on and coming off, then sometimes it will put it on the wrong place, sometimes it will take it off the wrong place. And when you come back, sometimes you'll have H and sometimes you'll have D, in that starting material. Right? So you go both ways and see. And that works, right? By starting with the same catalyst and excess deuteride, you do that, to go that way. Okay? So if you do a full cycle, like that, then you come back -- pardon me -- you come back exactly where you started. But, this proves the specificity; but it doesn't say which one, it doesn't say whether it was pro-R or pro-S. For that you need to do other kinds of experiments that we don't have time to talk about. Okay? So that's really a neat experiment, that shows that topicity makes a difference, and that enzymes discriminate between stereotopic, enantiotopic groups, or atoms.



Okay, next subject is Baeyer Strain Theory, which was proposed in 1885. So that's ten years after van't Hoff. Okay? So here's the group in Munich, in 1893, and here, front and center, is the boss of the group, Adolf von Baeyer. And this picture hangs in the hallway out there, the original picture. You can look at it. And the reason it does is that this guy here is Henry Lord Wheeler. Did you ever see his name?



Student: [inaudible].



Professor Michael McBride: See how observant you are.



Student: [inaudible].



Professor Michael McBride: Pardon? Lexy, did you -- Pat?



Student: Is it down in the foyer to this building?



Professor Michael McBride: Yeah, when you come in, there are these names carved on the walls, which were the people who were, in the nineteenth, early-twentieth century, professors. So he was the first organic chemist at Yale, and he went to Munich, where a lot of people went to learn organic chemistry from Baeyer. Baeyer, remember, was the guy who fiddled with the bread rolls, together with Fischer, to make these models. He was also the guy that did experiments on arsenic in Kekulé's kitchen when he was a student. So Baeyer was the leading organic chemist of the time. Now, in 1885 he -- that's earlier than this picture; that picture was 1893, when Henry Lord Wheeler was there. Notice that he died at what; at thirty-seven, the age of thirty-seven [correction" forty-seven]. I don't know what he died of, in 1911 [correction: 1914]. So he wasn't a professor very long.



Okay, so this paper, eight years earlier, was about polyacetylene compounds; so a bunch of triple bonds arranged in a row. We've talked about double bonds in a row; you can have triple bonds in a row too. But you don't have them for very long because, as Baeyer reported in this paper, they explode. He had a collaborator named Dr. Homolka, who he thanks for doing some of the experiments in this paper. Dr. Homolka does not appear in the 1893 picture. [Laughter] I suspect that he just graduated, got his degree and went to be gainfully employed someplace. But who knows? Anyhow, polyacetylenes are explosive. And this got Baeyer thinking, why should just a regular old hydrocarbon be explosive? So he branched out in this paper, and the very first topic was "The Theory of Ring Closure and the Double Bond." Now that seems a funny thing to talk about; when you're talking about triple bonds, to talk about making rings, and talk about double bonds. But what he says is: "Ring closure" which was a popular synthetic goal, at that time, to make new kinds of compounds, was to try to make rings. They could make six-membered rings, they could make five-membered rings. In Bayer's lab they actually made four-membered rings and three-membered rings, right? Although that was quite a chore. Okay, so "Ring closure is apparently the only phenomenon that can supply information about the arrangement of atoms in space." This is ten years after van't Hoff. "Since a chain of five or six members can be closed easily, while one of more or fewer members is difficult or impossible, spatial factors are apparently involved." And you know, we already talked about a case like this, where you tell something about arrangement in space from the ability to form a ring. Do you remember what that is? Anybody? Yeah?



Student: Synthesis of mesitylene.



Professor Michael McBride: Can't hear very well.



Student: Synthesis of mesitylene.



Professor Michael McBride: Oh, the synthesis of mesitylene formed a ring. But that doesn't -- so he used that to say where the methyl groups were on the ring. And that's actually a good example, but it's not the one I was thinking about.



Student: Cyclopropane.



Student: Cyclopropane?



Professor Michael McBride: Dichloropropane?



Student: No.



Professor Michael McBride: Is that what you said?



Student: No, I said cyclopropane with the [inaudible].



Professor Michael McBride: It's not cyclopropane that I had in mind; although that was one of the compounds that was synthesized in Baeyer's lab for the first time, and found to be reactive. But there was another occasion. Remember maleic and fumaric acid; one was cis and one was trans? And the one that was cis could lose water to form a ring. Right? So that was what he's referring to here. Okay, so then he goes on to say: "The previously -- " (So there's a website you can click on to see this.) "The previously proposed general rules on the nature of carbon atoms are the following: I. Carbon is tetravalent." Who did that?



Student: Couper and Kekulé.



Professor Michael McBride: Couper and Kekulé, right? "The four valences are equivalent, shown by the fact there's only one monosubstitution product of methane." (We've talked about that.) "III. The valences are equivalently arranged in space to the corners of a regular tetrahedron." Who says that carbon is tetrahedral?



Student: van't Hoff.



Professor Michael McBride: van't Hoff. Also, incidentally, there was a Frenchman who did it simultaneously, called LeBel. Okay? "IV. The atoms or groups attached to the four valences cannot exchange places." (So you can't -- you have to break bonds to make new isomers, right? So if you get them one way, they'll stay.) "The evidence is that there are two tetrasubstitution products, abcd of methane." (So van't Hoff and LeBel's rule, that you can have configurational enantiomers, we would say.) "V. Carbon atoms can bond to one another with one, two, or three valences." (That is, you have single, double, triple bonds.) "VI. The compounds can form either open chains or rings." (That's what he's talking about here.) "I should like to add the following, to these generally accepted rules. So he's adding a new property, to the models, the same way van't Hoff did; van't Hoff added arrangement in space, to the models that people already drew. What Baeyer is adding is VII.) "The four valences of the carbon atom point to the directions connecting the center of the sphere to the corners of a tetrahedron." (That's what van't Hoff also said -- already said.) "But forming an angle of 109°28' with one another." (That's very quantitative and precise. That's the angle between the center and two vertices in a geometric regular tetrahedron. Okay?) But then this is what's interesting: "The direction of attachment can undergo alteration" (It doesn't have to be exactly that angle) "but a strain is generated, increasing with the size of the deflection." So if you have an angle other than 109°28', then there's going to be higher energy associated with strain; or, what he's actually talking about is not higher energy, but reactivity.



Okay, so he drew this picture, showing how much the angles are distorted from 109.5°. So if you consider ethylene to be a two-membered ring, with two bent bonds, the angle the bonds are going at is where it should be tetrahedral; and in fact they're collinear. That distortion of each bond is 54°44', and so on, for three, four; five is perfect, right? It only deviates by 44'. But the six-membered ring is stretched a little bit the other way. It has to go out, rather than sharper. Okay. So he says: "Dimethylene…" -- that's ethylene, the first one -- "…is indeed the weakest ring, which can be opened by HBr, bromine, or even iodine." (We talked about bromine, or chlorine attacking the ethylene already.) "Trimethylene…" (cyclopropane) "…is broken only by hydrogen bromide but not by bromine." (So it's not as reactive as the two-membered ring.) "Finally, tetramethylene and hexamethylene are difficult or impossible to break." They're pretty stable. So this strain causes reactivity, if the bond angles aren't right. So he's becoming more quantitative; unusually quantitative for an organic chemist at this period. Remember, the physical chemists -- this is just at the period when physical chemistry is coming into its own -- and they were the ones who did heat and energy and so on. All Baeyer is talking about is whether things will react or not. But he's using geometric precision.



Okay? So the question is, are six-membered rings, like this, cyclohexane, they should be, according to Baeyer, a little bit strained. Right? Because the angles should be -- you'd think the angles would be 120° instead of 109. So they're opened up a little bit too much. But Sachse, a young privatdocent -- that is, essentially, between a graduate student and a professor -- in 1890 published a paper that had this very funny picture, A,B,C,D,E,F, up on the top there. And people didn't understand what it was. But what it is, is a thing that if you trace it on cardboard and fold along the diagonals, you get this blue thing -- right?; if you fold it along the diagonals and paste the ends together. Right? And then if you paste van't Hoff tetrahedra on it, you get this thing. And you'll see that this is just like this. Everybody see that? It's using -- but this is a base on which to build such tetrahedral carbons, so as to make this structure. And what's interesting about it is that these angles are the normal tetrahedral angles. Right? They're not strained. Right? Why are they not 120°? Because the ring isn't flat, it's puckered. Okay? So that's what Sachse said. And he gave people this thing so they could make their own models and see it. Right? And he gave the directions for building this more complicated base, where you put it on and have a different form of cyclohexane. Okay? And that one, as you see, looks like this. Okay?



So in 1890 he knew exactly the score that we would know -- that we would get with our own models nowadays. Now, when this was abstracted in the journal that published abstracts of all chemical literature so people could go there and find out what was going on in chemistry, Julius Wagner, who abstracted it, said: "It is not possible to write an abstract of this paper, especially since the author's explanations are hardly understandable, without models." Now, one conclusion from that should be therefore go and build his models and look at them, and you'll see what he's talking about. But that's not what Wagner said, and it's not what people did. He said, "Forget that, it's nonsense" -- right? -- "you can't understand it." Okay. So Baeyer wasn't happy with this, disputing his theory. So he wrote, in the same year, 1890 -- Sachse wasn't too smart about this because he published that paper just as they were having a big celebration of Baeyer's contributions to aromatic chemistry in Berlin; there was this big, fabulous dinner with eighteen courses, or something like that, and all -- everybody, chemists from all over Germany gathered. And Sachse, this nobody, publishes a paper and says Baeyer's theory is wrong. Bad timing.



So Baeyer published a response quickly. He said: "A further proposal is that the atoms in hexamethylene…" (six-CH2 groups, cyclohexane) "…are arranged as in Kekulé's model." (Now Kekulé's model, remember, by this time was in fact a tetrahedron; which is what Sachse showed how you could do it. Right?) But he says, "…as in Kekulé's model, and that the arrangement of the atoms in space is one with minimum distortion of the valence directions." (That sounds like Sachse.) But what he says is: "Thus, the six carbon atoms must lie on one plane." (It's got to be planar.) "and six hydrogen atoms lie in equidistant parallel planes." (Six above, six below. All the CH2s are just like this around the ring. Right?) "Further, each of the twelve hydrogen atoms must have the same position relative to the other seventeen." (That is, they're homotopic; you can rotate it, if the plane is flat. Now how can he say this? It's not really like Kekulé's model. Okay?) "The experimental test of the correctness of this assumption is relatively easy; for example sufficient evidence is that there is a single isomer of hexahydrobenzoic acid."



That this, only one cyclohexane carboxylic acid. That is, if you had this, and put a COOH group here, or a COOH group here, what would be the relationship between those? This place and this place, what would you call -- are they homotopic? Are they enantiotopic, this place and this place? Mirror images? No, they're just different. They're diastereotopic, right? Now, but you could do a neat trick. So that's what he says. If it were like Sachse, then you'd have two isomers of a monosubstituted cyclohexane. But what he didn't take into account was this. This one is sticking up like that. Watch this. Notice the black one is sticking out, down at an angle, and the white one is sticking straight up. But watch this. Now this one's down at an angle and the black one's straight up. Did I break any bonds? What did I do?



Student: Rotated.



Professor Michael McBride: I just rotated about bonds. They're conformational; conformationally diastereotopic. Not configurationally. So they're different but they can interconvert easily. Right? So Baeyer says he's right because there's only one isomer, so you have to regard it as flat, even if it isn't; although he didn't say that. "Meanwhile, as long as our knowledge in the field is so incomplete, we must be satisfied that the above assumption is the most likely, and no known fact contradicts it." Okay?



Now, Sachse didn't take this lying down either. So he publishes a forty-one page paper in Zeitschrift für Physikalische Chemie, which is published by Ostwald, who hates Baeyer, and vice-versa; the physical chemists and the organic chemists didn't get along. So he publishes a paper in this other one where he gives all this trigonometry to explain what he means by this structure. Is this something calculated to appeal to organic chemists? Not very likely. Okay?



So this is edited by Ostwald, who didn't even believe in atoms, remember, and wrote disparagingly of his successor in Riga, who had been an organic chemist: "Scientifically, he had been brought up in the narrow circle of contemporary organic chemistry, and to him the arrangement in space of the atoms of organic compounds was the foremost of all conceivable problems." That wasn't what Ostwald thought was the best problem. So Sachse published again. Nobody responded to that first one, so he publishes again, the next year, thirty-four pages again. And here he gives more formulas; again, things that didn't appeal at all to anybody who was actually interested in the arrangement of atoms in space. And then he died at the age of thirty-one, in that year, and that was the end of Sachse, and the end of his theory. Right? And Baeyer wrote in 1905 -- so twelve years later: "Sachse disagreed with my opinion that larger rings are planar. He is certainly right from a mathematical point of view;" (Although I sort of doubt that Baeyer went through every line of this to check it. Right?) "Yet, in reality, strangely enough, my theory appears to be correct. The reason is not clear." Okay? So that was -- remember, Sachse was 1890. So what important lesson should we take from the tale of poor Sachse? Do you want me to give you that as a problem to think about over Thanksgiving? Fine.



Okay, so this, remember is what Sachse wrote, and he drew models like this. And -- I'll just conclude the punch line here -- in 1913 Bragg and Bragg determined the diamond structure, by X-ray diffraction of a crystalline diamond. Right? And five years later, Ernst Mohr published a diagram he drew of -- from that diamond structure. Braggs didn't draw the structure, but Mohr did five years later. And here's what he drew, and that is exactly what Sachse meant. Right? So the arrangement of the carbons in a six-membered ring, in diamond, is exactly what Sachse said; but he got nothing for it. Okay, have a good Thanksgiving.



[end of transcript]

Lecture 33
Conformational Energy and Molecular Mechanics
Play Video
Conformational Energy and Molecular Mechanics


Understanding conformational relationships makes it easy to draw idealized chair structures for cyclohexane and to visualize axial-equatorial interconversion. After quantitative consideration of the conformational energies of ethane, propane, and butane, cyclohexane is used to illustrate the utility of molecular mechanics as an alternative to quantum mechanics for estimating such energies. To give useful accuracy this empirical scheme requires thousands of arbitrary parameters. Unlike quantum mechanics, it assigns strain to specific sources such as bond stretching, bending, and twisting, and van der Waals repulsion or attraction.




Transcript



December 1, 2008




Professor Michael McBride: Okay, so as you remember, before we ended, we were talking about Sachse and how he explained everything in more detail than people wanted to hear about. Okay? And he knew exactly the story; that you had chairs of cyclohexane, and you had boat, and you could interconvert them by rotation about single bonds. But it was only after the Braggs had determined the X-ray structure of diamond, and Mohr, not equally importantly but very importantly, had drawn clear pictures of it, like this, that people understood what Sachse had been talking about twenty-eight years earlier. Okay? Because if we look in the middle there, we see that chair form of cyclohexane. Okay? We call it a chair. People like to give sort of silly, trivial names to things. Scientists are probably sort of geeks, actually; but that's okay. So there's a chair -- right? -- that looks like this chair. And that chair has a particular property, which is you can fold the back down and fold the legs up. Okay? But you can do the same thing, to do the ring flip, as we call it, in chair cyclohexane, by counter-rotation of two parallel bonds. So take these two bonds there, rotate it that way, and there rotate that way; counter-rotation, right? One rotates clockwise, the other counter-clockwise. And you rotate group four, at the bottom right, so that it goes up. Everybody see how the rotation does that? And I can do it with a model here. Right? So I rotate around this bond and this bond, hold this part here, and I rotate, and it goes up. Okay?



Now, so the product, after rotating that up, is a different conformation, which we call boat cyclohexane, or which people call boat cyclohexane, because it looks a little like a boat. Okay? And people who got very imaginative named various bonds on this; like the bowsprit, or the flagpole. But, in fact, this isn't really a conformation at all, because it's not a minimum of energy, it's a maximum of energy. You get lower energy -- this is the boat -- you get lower energy by twisting a little bit, like that. Right? So the boat is actually not a minimum of energy. Usually we reserve the name "conformation" for isomers that are minima in energy. Right? They can vibrate, but they're at a minimum of energy. That's not true of the boat. But the boat is what Sachse made his picture of, and it's easy to think about. So we often talk about the boat, even if it isn't true; if it really wants to twist a little bit. We'll discuss that a little bit more later, why it wants to twist. But anyhow, there's a boat.



Now, if you then did the same trick to the blue bonds, on the other side -- that is, counter-rotate so those go in and down, rather than in and up, as the red ones did on the left -- it goes down like that. And now you see what you have is a chair where everything that was up is down, and everything that was down is up. Okay? So we started with a boat here. I can flip it like that to make a capsized boat; or a chair, start with a chair. Flip it to make a capsized boat. Flip it again and I have another chair. But notice what happened. Here all these bonds that are pointing vertical, or down and vertical, are black, and the others are silver. Right? Everybody see that? But after I do this, they've changed; the silver ones point up and down, and the black ones point out. Right? So that's changed the environment. That's interchanged the environment of this one and this one, of the black and the metal one -- right? -- by doing that so-called ring flip.



Okay, now you should learn how to draw chair cyclohexanes. That's a very popular discipline among organic chemists, and it shows that you understand what's going on with the conformation, if you draw it right. And if you don't draw it right, it shows that you don't understand. So let's see how -- people say, "Oh, I'm not an artist. I can't really draw it right." But if you understand it, you can draw it right. The only thing you have to be capable of doing is drawing things that are parallel to one another. That's not too big a challenge. Okay, so notice that the carbon-carbon bonds are parallel in pairs. Right? So the red ones are parallel, the blue ones are parallel, and the green ones are parallel. Okay? Now that means when you draw the frame it looks like this, so that opposite ones are parallel to one another. And now the only challenge is to put the hydrogens on, or whatever the substituents are. Right? And for that purpose it's worth knowing -- noticing the symmetry of the ring. Right? There's a three-fold axis of symmetry, vertical, as they've drawn it here. Right? So you can rotate 120°, a third of a circle around that axis, and you can't tell that it happened; it's symmetry. Okay? So notice that some of the bonds are parallel; are called axial because they're parallel. But some of them are parallel up, and the intervening ones are parallel down, to keep the carbons looking tetrahedral.



Okay, so six of the hydrogen bonds are parallel to the axis of symmetry. So even if you drew the six-membered ring in some cockeyed direction, so that its mean plane was like this, or like this, or like this, or like this, still you can see where that axis points and make the axial bonds parallel to that axis. Okay, now you have the problem of drawing the last bond, which completes the tetrahedron. Notice that the equatorial bonds as they're called -- because it's sort of like the equator, relative to the axis -- they're not strictly horizontal. Right? They go a little bit opposite the way the red bond went on the same carbon. And notice in particular that they're parallel to the next-adjacent carbon-carbon bonds, which are shown in blue here. Okay? So they make a sort of an N, or a Z, with the next adjacent carbon-carbon bond. Okay?



Now, what o'clock -- if we start from that carbon at the back left -- what direction, starting from the carbon, do I draw a line, to draw the proper orientation of the hydrogen, at that carbon? You know, one o'clock, two o'clock, three o'clock, eight o'clock, seven o'clock, nine o'clock. What o'clock? Think about it a second, and then I'll ask you. Oh, we can do it like an auction. I'll go one, two, three, four, five, and you raise your hand when I get the right one. Okay, twelve o'clock? one o'clock? two o'clock? three o'clock? four o'clock? five o'clock? six o'clock? seven o'clock? eight o'clock? nine o'clock? ten o'clock? eleven o'clock? Okay, the answer is eight o'clock. Right? Well how do I know that? Because it's parallel to the next-adjacent C-C bond; there. Right? See the Z? So it's parallel to the next-adjacent C-C bond.



Okay, now how about there, what o'clock? one o'clock? two o'clock? three o'clock? four o'clock? five o'clock? Okay. It's about 1:30, right? -- sort of like one o'clock -- because it's parallel to the next-adjacent C-C bonds, which are pointing anti-parallel, right?; parallel, but in the opposite direction from their nearest carbon. Okay, try here. One o'clock? two o'clock? three o'clock? four o'clock? five o'clock? Ah, you're getting better, right? About four o'clock; 3:30 maybe. Okay, how about here? One, o'clock? two o'clock? three o'clock? Okay, two o'clock. Okay, how about here? One? Two? Three? Ah, I got some votes for one. Now I'm going to come back and ask what's wrong with one o'clock? Okay? There's what it is. It's actually seven o'clock. How is that related to one o'clock? It's anti-parallel, right? If you did one o'clock, it would have made a U with the next-adjacent bond, rather than an N or a Z. Do you see that, see where the mistake was? Okay, so if you understand this, if you understand that the hydrogen is anti to the next-adjacent C-C bond -- right? -- then you won't draw it wrong, you won't draw a U, which would be eclipsed. Right? You'd draw it anti, which makes a Z or an N. Okay, so if you understand it, you'll draw it. So it wouldn't be surprising if that were on some test sometime, to draw a cyclohexane sometime.



Okay, now this got interesting because in the 1940s and '50s, synthetic organic chemists were very interested in steroid hormones; in the '30s, '40s and '50s. Right? And it turned out that you had these two alcohols, which were called β and α. And beta is the one that came up, toward the viewer, and alpha is the one that goes back into the paper. But of course, that depends on how you drew the molecule. If you drew the molecule upside down, it'd be just the opposite. But synthetic organic chemists, who were interested in this kind of thing, weren't bothered about this, because they always drew it the same way. Right? So they knew what they meant by up and down. But that's like names that don't really tell you what the -- don't allow a novice to know what the structure is, to talk α and β. People still talk about α and β substituents, the ones that point out of the paper and into the paper, in steroids. But you have to know the lore. You have to know how people always draw the rings, in order to know what comes out and what comes in. So it's like cis and trans, a little bit, not knowing what's cis to what.



Baeyer, remember, said it didn't make a difference, because there was only one cyclohexane carboxylic acid. Right? Because if you had axial and equatorial ones, the rings could flip and they would interconvert and it'd all be the same. Right? So, but it's interesting that the β and the α isomers of these alcohols have different reactivity. They are different. Right? They behave differently. They don't interconvert. Right? And you can see why that would be so. What's the relation between the β and the α? Are they mirror images? Are their environments identical? Enantiotopic? Diastereotopic? There's an easy way to recognize this. They exist at a chiral center, because going around one carbon and going around the other are different paths; they encounter different things as you move along. Right? So it's a chiral center. But there are other chiral centers in the molecule. And when you have two chiral centers, then changing just one doesn't give the mirror image. It gives an epimer, not an enantiomer. So they are diastereotopic. Right?



Now, for these kinds of problems, Derek H.R. Barton invented what was called "conformational analysis", in 1950. He was interested in these steroid hormones, as many people were. Right? So conventionally these six-membered rings were labeled A, B, C, and then the five-membered ring D, so people could know what they were talking about. Right? And Barton redrew ring A. And notice that the β, α is configurationally diastereotopic. You'd have to break bonds to make the α into β. So it doesn't surprise us that they're different. Right? But let's look at just how they're different. So this is the picture, from his paper in 1950, that Barton drew to show that ring A; and the bracket there, that I've drawn in, shows where it would go on to ring B. I've truncated his picture. Now I want to ask you a question. So β goes up, relative to the mean plane of the ring, and α goes down. And he labeled these things e or p, for equatorial -- that's what we still call it, as I told you, the one that points more or less in the plane of the ring. And he called the others p, for polar, as if it were a globe and had an axis, the pole. But we call it now axial. So that's changed since he invented this. But I think you're probably sophisticated enough now, having had this practice, to recognize an error in what he drew. Do you see what's wrong with his pictures? Marty, what do you say? Are the equatorial bonds, do they make Z's with the carbon frame? That is, are they anti to the next adjacent carbon-carbon bond? Take an equatorial one, like this one here. Is this carbon-carbon bond anti-parallel, at the correct o'clock, relative to this bond?



Student: It needs to be more [inaudible].



Professor Michael McBride: Can't hear very well.



Student: It needs to be more like eleven-ish.



Professor Michael McBride: No, this one, notice, is exactly horizontal, and this one is exactly horizontal. So they're perfectly anti. Or take this one. Right? Here's its bond, and here's the next adjacent C-C bond, one to ten. And they're exactly antiparallel. Okay? So that's anti too. So the equatorial ones are fine. How about the axial ones? Do they look good? Catherine, what do you say? What does axial mean? What's axial about it? Maria?



Student: The molecule can be rotated around the axis.



Professor Michael McBride: Okay, what direction does that axis point, that you can rotate that six-membered ring around it and get the same thing? What o'clock does it point, from the center of the ring?



Student: So it should point like eleven o'clock.



Professor Michael McBride: Yeah, about eleven o'clock. There are the sets of three carbons that are related by that, and the axis goes at about eleven o'clock. What does that mean Maria, about the axial bonds?



Student: That they should also be pointing up.



Professor Michael McBride: They should be parallel to that axial direction. Right? So they really should be this direction -- right? -- parallel to that axis. So the very guy who invented it, in the very paper in which he did the invention, drew them wrong. That's not too surprising. Right? But it's interesting to note. So you don't have to feel bad when you draw it wrong the first few times. By the time we get to an exam, you'll know how to draw it right; better than Barton did, when he published his paper. Okay, he got the Nobel Prize, in 1969, for development of the concept of conformation and its applications in chemistry. So he didn't get the Nobel Prize for drawing the axial bonds wrong. Right? What he got the Nobel Prize was for the application, showing how important this was in chemistry; that those axial and equatorial groups -- hydrogens in this case; but they could be other groups -- have different chemistry. Right? And notice the date, 1950. All these things about stereochemistry were happening within plus or minus a few years of that. We've already talked about Bijvoet determining the absolute configuration of tartaric acid; Newman figuring out how to draw his projections, to show conformation; the Cahn-Ingold-Prelog rules; and, as we'll see shortly, the idea of molecular mechanics. Okay?



Now, what made Baeyer say everything was the same, although he didn't know it, was that the ring would flip, interconverting axial and equatorial, as we just looked. But can the ring -- if the ring can flip in this one, it would also interconvert axial and equatorial -- but can the ring flip? That's the question. So let's think a little bit more about how the ring-flip works. Notice that during a ring-flip, as in the top left there, what's equatorial becomes axial. So you look on the top left, there's an equatorial bond, and in the right that becomes an axial bond; well actually you'd call it a -- what is it? -- flagpole, in the boat. Right? But when you do the other flip of the blue one, that one that's purple is there. Right? So the one that was equatorial in the chair on the top left, after a flip becomes axial. And we already saw that here, that the ones that are equatorial become axial when you do the double ring flip. Okay? Now, and by the same token, the ones that are axial become equatorial. So if you start on the top left with those two green ones, which are anti to one another, axial -- notice they point a little to the right because the axis of -- the symmetry axis is not straight up and down. Right? After the first part of the flip to the boat looks like that; and the second part of the flip, they become equatorial, and gauche to one another, where they were anti originally. Right?



So now if you have fused rings -- that is, two six-membered rings that share a bond, as on the bottom there -- those two green ones can be part of the second ring; gauche to one another, as all the carbons are around the cyclohexane ring in its chair form. So they're gauche, and that's fine. But if you tried to flip it, those two green ones, with respect to the front ring, would become axial, as on the top left. And if they pointed axial, you couldn't possibly complete the ring. Right? So here's two chairs, here and here. Now if I could try to flip this one up -- I could flip this one down probably; there, down. Right? So I got it almost into a boat. But I can't even make it into a boat, let alone flip the other ring. And the reason is that these two are tying these two -- while I'm trying to flip this one, these two are holding this one and won't let it go. They can't become axial to one another. Right? Just because the carbons can't reach. Okay, so in what's called -- decalin means decahydro, ten hydrogens, on naphthalene; naphthalene is like benzene, except two of them. Right? So this is decalin, or ten hydrogens on naphthalene. And this one is called trans-decalin. Right? These two hydrogens are trans to one another. And in that one, I can't flip the rings. If I had cis-decalin, like this, then it's possible to flip the rings; which I could do like this. Let's see. Well ho, ho, ho. I think I did it there. Yeah, I did. Okay.



So you can flip the ring if it's cis, but not if it's trans. And the point is that you can't take two equatorial things and make them both axial in a ring, because those carbons get too far apart to be bridged by the rest of the ring. But if they're gauche to one another, then they can be gauche to one another on the other side and it can flip. It's fun to play. You know, we used to require people to buy models. Now you can do things with computers that make it that it's probably not worth your money to buy the models. But it's still a lot of fun, and you learn something with your hands. It's like when people have trouble in elementary school with arithmetic; the teacher gives them toothpicks to count with and so on. It's a higher level than that, but it's the same thing; you learn a lot with your hands. So it's fun to play with these. In fact, let me pass around here simplified models that are a chair. But you can do this trick of changing one into the other by rotating. And feeling it is fun. Right? So that you have the opportunity to feel it, I'll pass these around. There you go. There you go. Okay, the ring flip is impossible for trans-decalin. Okay, but so gauche because gauche is okay within the second ring of decalin, but not anti. Okay, so and you try with models if you're skeptical. If you want to come afterwards and fiddle with the decalins, you're free.



Okay now, nowadays people, much more often, encounter these things on animations. But this is a nice webpage that we have permission to use in the course; I mean, you could use it on your own anyhow, on the net if you wanted to. But it's a cautionary tale because just because it looks nice and works, doesn't mean it gives the right answers. So let me show you what I mean. Okay, so if you go to this website, you can click there to get it. And it's a nice tutorial about conformation. So they show here, for example, ethane, and if you click 'Play' -- let's see, I think I've got it animated here -- you can step from one position to the next and see what the shape looks like. Not that ethane is so very exciting, but you can see it go down and up, as it goes from eclipsed to staggered. And you can click on a point and the model will turn, to show you what it's like. I'll give an example of this later. And there's the staggered and there's the eclipsed. That's the barrier that goes across. And you see that that barrier is 5.2 kJ/mole, because they use -- they're modern and they use joules, whereas American organic chemists tend to use kilocalories. But anyhow, if you multiply that by 0.239, you get it in kilocalories, as what we've been talking about, which would say that the barrier is 1.24 kcal/mol. How big is the barrier in ethane? Does anybody remember from before Thanksgiving? Dana?



Student: Three.



Professor Michael McBride: It's three. Right? Just because they give a fancy chart with numbers on it, and just because it's been calculated by some quantum mechanical program that they got access to and could use, doesn't mean the numbers are going to be right. And they're not, they're off by a factor of more than two. Right? So don't believe everything you see. That's maybe the primary lesson that this whole unit tells you. Okay, it should be 2.9. So let the buyer beware in situations like that. If you don't pay anything for it, you probably get about that. But the pictures are still very nice, and the general shape is correct. So here's the rotation in propane, and it says the barrier is 5 kcal/mole. And you know it's something like 3.4, 3.3, instead. Here's butane, and again you can animate that and see the various staggered conformations and the fully eclipsed one. And we know the gauche is supposed to be 0.9 kcal/mole; which is not what this says. And forget the scale on the left.



We can use that 0.9 to tell how much of the gauche there should be at equilibrium. And let's rehearse this again. Remember, it's 10(3/4 ΔH), in kilocalories. So 3/4ths of 0.9 is 0.68. So it means the ratio of gauche to anti should be 1:4.7 (4.7 is 100.68). Okay? So 1:5, about. But it depends on what you're talking about, because that gauche is in fact chiral. Let's see, I didn't have it here. Well you can -- I do it with my hands. Okay, so here's anti. Okay? Here's gauche. Right? But gauche is chiral. It could also be this. Right? It could be one hand, or the other; the mirror image. It's like a propeller of gauche. Right? So when I say gauche, what do I mean? Do I mean gauche+, or do I mean gauche-, or do I mean both of them, taken together? Because obviously, if I include -- if I say gauche and mean both of them -- right? -- then I have to multiply this by two, and the ratio is 1:2.4, instead of 1:4.7. Okay? So there has to be this statistical factor taken into account as well, when you use a collective name like gauche. Okay, the eclipsed is 3.4 kcal/mole. That will tell how fast anti goes to gauche. So the difference in the well heights, tells how much gauche there is at the equilibrium, and the barrier height tells how fast you go, from one to the other. So how fast do you go from anti to gauche? Do you remember how to do that? Anybody remember? Chenyu, do you remember how to do it?



Student: How to do what?



Professor Michael McBride: How to find out how fast something is, if you know how big the barrier is you have to go across?



Student: I'm not sure.



Professor Michael McBride: No, but you'll know it for the final.



Student: I'm sure.



Professor Michael McBride: Okay? It's 1013th/second; pretty fast. But then slowed down by that same kind of equilibrium constant, 10(-3/4 of how big the barrier is). Now, if the barrier is 3.4 -- say it was four, say the barrier's four; it's about that. 10(3/4ths of 4) is 103rd. Right? So it's slowed down by 103. It's 1013th/second times 10-3. Right? So it's 1010th per second. Okay, there it is. Oops, I did it wrong. What did I do wrong here?



Student: You said four instead of 3/4ths.



Professor Michael McBride: Pardon me?



Student: You made it four instead of 3/4ths.



Professor Michael McBride: That's not the difference though. I think I've got a typo here. That can happen. Okay, anyhow, I think -- subject to thinking while I'm not on my feet -- that this should be about 1010th per second. Okay? Anyhow, you see how you would use it. Right? 1013th times -- whoops, oh I forgot to plug; no, I did plug it in. Oh, it's okay. Okay, so there we go. Okay? And the fully eclipsed barrier is about 4.4 kcal/mole. But that's hard to access experimentally. Why? Because experimentally you try to tell how fast one thing goes to another. Right? But it's possible to get from anti to gauche without going over that barrier, and it's possible to go from gauche to anti without going over that barrier. To go from gauche to gauche, do you have -- notice that 360° is the same as 0°. Right? So you can go from this gauche to this gauche, by going over that barrier. Right? So can you measure that rate and see what that barrier is? No, because there's an easier way to go. Instead of going like this, to get from one gauche to the other, you just go like this. Right? So there's an easier way to do it. So you can't measure the rate and know what that barrier is. That barrier is more or less irrelevant experimentally.



Okay, or here's the ring flip in cyclohexane. Now let me -- I'll use this one to animate it, to show you how. You can try this program yourself. Okay, so here we go. And there's the thing doing its ring-flip. And it's nice because you can grab it and rotate it around and see it from different points of view as it does its ring-flip. So there's flipping one end down, and then flipping the other end up, to go from one chair to the other. So you can fiddle with this. Or you can get models and try them too. Okay or, as I said, you can go -- we could stop that -- and you can go over here and see what it looks like at the halfway point. Now notice that that is a little bit like a boat. But it's a twisted boat, as I said, because the actual boat is up here at the top. In fact, even at the top it's not -- oh, pardon me, I said the wrong thing. It's not -- up at the top, it's sort of a half chair, half boat. Right? This top left carbon is still in the sort of chair-like form, or these five carbons are. But this one is halfway bent up, toward a boat. Okay?



Anyhow, you can fiddle with those things. Oh let's see -- what do I need to do here? Go back here. Okay. So there's the chair conformer, there's the flexible or twisted-boat conformer. Now that one is really fun. I'm going to do it here, and you people who have this can do it. Notice -- did you notice when you played with this, that the chair is sort of rigid. If you try to twist it, it's hard. Has everybody felt it, to feel that? But if you get into the boat form, like that, then it's quite flexible. You can do this, to your heart's content. Did you feel that? After you've passed it across, pass it back so everybody can feel how -- this is great, like the orb. Right? But the chair is quite rigid. Right? And as it goes from chair to boat, it clicks; like click. See it click? Okay. So there's a barrier; that's the point. But in the flexible form there's not a barrier, it just smoothly rotates. Okay, so there's that barrier of 11 kcal/mol. So it takes some time to get back and forth. Okay, and here's the flexible one, and you can animate that and watch it flex. Okay? So that's the twist-boat form.



Okay, now we're going to talk about the shape, strain energy, and molecular mechanics. The point is to talk about molecular mechanics. How do you get these energies? That particular animation used a quantum mechanical calculation. It didn't do a very good job on the energies, as we just showed. But is there an easier way to get these energies? And there is, and it's called molecular mechanics. Essentially it's just using Hooke's Law for the model, to calculate strain energies. So we already saw the torsional energy of ethane at 3 kcal/mole -- a threefold barrier, three minima, as you go across -- and the conformational energy of butane. And it's fun to remember those numbers: 0.9 for anti to gauche; 3.4 barrier between them, and then maybe 4.4, but who knows, for the other one. Okay, now remember 1950 is when all these things were happening. So in 1946 there was a paper by Frank Westheimer and Joseph Mayer from the University of Chicago, about using mechanics to calculate the energies of conformations. The particular thing they were interested in is when you have two benzene rings hooked together, so that they can rotate like this. Right? They can't rotate freely if you have things here that are sticking out, that would run into one another, as you tried to rotate them. Right? So you could get one that's twisted this way and one that's twisted this way, in a substituted biphenyl. And you can resolve them, and get optically active forms -- one that's twisted to the right, one that's twisted to the left -- even though you don't have a carbon that has four different things on it; because it can't rotate. Right?



But the question is, how big should that barrier be? And Westheimer, in fact, is your great-uncle. Right? Because he also studied with Conant and Kohler, at Harvard; the same as my teacher, Bartlett, did. Okay? So he was interested in physical-organic chemistry and did this. Incidentally, he's also the guy that figured out which hydrogen you pull off on ethanol; you know, that enzymes choose between the pro-R and the pro-S hydrogens. So anyhow, Westheimer did this in 1946 and figured if you have a bromine here and a bromine here, that when it's flat, and achiral, that's the transition state between being twisted one way and being twisted the other; that -- whoops, sorry -- that these two things would run into one another. That would mean the bonds might bend back a little bit or something; how much energy would that cost? So he tried to do it by what's called molecular mechanics. So these programs calculate the energy, and they can minimize it by adjusting angles to get the minimum energy, by treating these molecules as if they were mechanical. And to achieve useful precision they require a very large set of empirical force constants; that is, how strong are the various springs? And you adjust these arbitrarily on some test molecules, in order to get the best values, and then try to apply it to something else.



You can also do it nowadays by reliable -- unlike the ones we just talked about -- quantum mechanical calculations. But I want to show you how many parameters there are, when you do molecular mechanics. Okay, so this is so-called MM2 -- that's a particular molecular mechanics scheme -- and these are parameters. So there are sixty-six different atom types; fourteen different types of carbon, depending on what it's bonded to. Okay, so here are some of -- here are the fourteen different types of carbon. The carbon in an alkane, in an alkene, in an alkyne. That carbon in a carbon cation. The carbon in a carbonyl group. The carbon in a carboxylate; and so on. And these have different van der Waals radii; which is one of the things you have to put in. Okay? Now, then you have these various kinds of atoms hooked together, and you have a certain strength for stretching the spring. That is to say you need what the equilibrium bond distance is, the minimum energy bond distance, and how strong the force constant is, for stretching that spring. So notice there are 138 different bond stretchings. These are just the ones that involve where the first atom is of type one. Okay?



So you need to find out what all those force constants are. Okay. Then you can bend bonds, and there are 624 different strengths of springs for bending different types of bonds. Forty-one, shown here, involve alkane carbon, alkane carbon, and then some other group; it could be another alkane carbon. That's the very top one; one, one one is three alkane carbons. So that's the force constant for bending the bond. But then there are different equilibrium bond angles; all close to 109.5 -- remember, that's the tetrahedral angle -- but a little bit different, according to whether that particular carbon has two other R groups on it, or an R group and an H, or two Hs, in addition to the other two groups you're talking about actually bending the bond. So you get -- so there are -- so then, of course, you can twist bonds also. That's what conformation involves, after all. And there are 1494 different bond-twisting, and for each bond-twisting you have three parameters here. For example, V3 -- also V2, V1 -- for each of those 1500 whatever number it was of bonds. I'll just show you what that means. V3 means a threefold barrier. And it says its height is 0.093 kcal -- yeah, kcal/mole. So here it is. Right? It's maximum when it's eclipsed, it's minimum when it's staggered, maximum eclipsed, minimum staggered. So that's a threefold barrier, for rotating. But there's also, for a particular kind of set of three bonds, there's a twofold barrier, which looks like that. Right? And that one, you see, is 0.27 kilocalories high. And there's also a one-fold barrier, like that, which is, for that particular one, alkane, alkane, alkane, alkane, 0.2 kilocalories high. Now you add all these together, and that is the torsional contribution to the energy in butane; or anything where you have alkane, alkane, alkane, alkane. Right?



Now we know what it really looks like in butane. It looks like that. The scale is much bigger. So this is a rather minor contribution. Now why is it that the anti is the most stable, according to this scheme? It's because it's -- the van der Waals repulsion is least when these two carbons are as far apart from one another, rather than being eclipsed with one another. Right? But that then is biased -- you calculate that van der Waals repulsion, which depends on the radii of various things. Right? And you then tweak it by adding this to it, as well. Right? So you can see that these are pretty complicated -- I mean, for a computer it's not a problem. It's a lot easier than quantum mechanics, finding all these curvatures and stuff. But it's still a pretty intricate thing and an enormous numbers of parameters. After simplification, the so called MM3 scheme has more than 2000 arbitrarily adjustable parameters that you have to fiddle out by knowing a lot of different experimental results. And what would I have to make the numbers to make that right? And then it has to be also right with this one, this one, this one. So obviously you need at least 2000 different molecules to determine 2000 parameters, or 2000 different measurements at least.



So it's a very highly parameterized system. Contrast it with quantum mechanics, where there are no arbitrary parameters; all you have is the particle masses, their charges, and Planck's constant. Right? So there's nothing fundamentally correct about molecular mechanics. It's just a very complicated scheme that's been tweaked to try to give good results for the molecules that it's tweaked on the basis of. But that doesn't necessarily mean it'll work for anything else. But it does work pretty well. And the nice thing about it is by looking at the calculations, you can figure out why things happen the way they do. For example, here's ideal cyclohexane, the way we've been drawing it. And we can look at the various kinds of strain that there are in that. First we take an ideal cyclohexane; which means it has ideal bond lengths, bond angles, and is staggered, like this. Okay? And there's no strain for stretch, bend, or what's called stretch-bend. Why is there a term called stretch-bend? Because when you stretch a bond -- it might be easier to bend, or conceivably harder to bend when it's stretched. So you have to put another term in for that. Okay, but there is torsional energy, because there are gauche interactions here, like that one. But there are six of them, as you go around the ring. So they're not anti. So there's torsional energy.



And then there's what's called "non-1,4 van der Waals energy"; which in fact is favorable, it's attractive. Okay. And here you see an example: one, two, three four, five, those atoms, atom one and atom five there, are five atoms apart. Right? So it's non-1,4; it's 1,5. And they're at a distance where their interaction is attractive, according to van der Waals energy. But you can have 1,4 van der Waals energy, as here. There's one, two, three four; and that, you see, costs you -- Its strain of 6.3 kcal/mol -- not that particular one, but sum them all up. Okay? There's a very bad one here, you can see, between one and eight. Those are rather close together. Okay? So there's this much strain in that. Now, what a molecular mechanics program can do is adjust the geometry -- twist bonds, stretch them, bend them and so on -- to minimize the energy. Okay? And here's what happens. Whereas the total strain energy before deformation is almost 8 kcal/mole, after -- let's see, what did I do? There. After you've minimized it, those are the energies. It falls to 6.56. So it gets a kilocalorie 0.3 better. Okay? Now notice what happened. It stretched the bonds a little bit. It bent the bonds a little bit. Right? Let's see exactly what it did, by looking at the model. So there's the -- it's going to stretch and flatten the ring slightly, to reduce that bad van der Waals repulsion. Notice the 1,4 van der Waals before was 6.32, and it's cut down to 4.68. So watch what happens. So this is -- did you see it? Let's back up. See, it flattens and stretches, just a bit, to get -- to reduce those van der Waals repulsions. So this way I told you to draw cyclohexane, to get it the same way Mohr did, is the way organic chemists do it. But it's not quite right. Actually the rings flatten a little bit and the things that should be axial spread out just a little bit.



Okay, fine, so some things get better, some things get worse; overall it gets a little better. And it would be very hard to do that in your head. Right? Molecular mechanics is good to do that. Okay, now there you notice is a gauche butane within the cyclohexane. And we know how much gauche butane costs, compared to anti butane. How much strain is there in gauche butane? Remember gauche versus anti, I said that was worth remembering. It's 0.9 kcal/mol. Okay, but in fact, in the whole thing, there are six gauche butanes, because every bond is part of a gauche -- is the central bond of a gauche butane. So if you had six gauche butanes, that would be a strain of six times 0.9; 5.4 kcal/mol. Okay? And in fact that's rather close to 6.56. So that actually is way over-simplified. But it's a good mnemonic device for remembering how big it is. It's about six gauche butanes. Or suppose you had axial methylcyclohexane. Now you have much, much worse van der Waals interactions -- right? -- 6 and 8 kcal/mol of strain. But what's going to happen if I let it relax, if I run molecular mechanics and let it minimize its energy? Can you guess what's going to happen? How will the structure change? Virginia?



Student: The last methyl group.



Professor Michael McBride: Yeah, that methyl group on the top is really in trouble -- right? -- because of those non-1,4 van der Waals repulsion. But if you bend it back to the right a little bit, you'll reduce that. So here's what happens, it relaxes like that. And now notice that the non-1,4 van der Waals went from being bad, by 6 kcal/mol, to being good by 1.3, overall, summing them all up. Right? So it went from sixteen kilocalories of strain to only nine, or from seventeen to nine. Okay. And notice that here there are two more gauche butanes, in axial methylcyclohexane. There were already six in the cyclohexane ring. Now there are two more. So you'd guess -- whoops I went too fast -- you'd guess eight times 0.9; 7.2. It's 8.6. It's in the right ballpark. So it's roughly what you'd expect for eight gauche butanes. Okay?



Now, if you go from axial to equatorial, then it turns out that energy difference is 1.8 kcal/mole. Equatorial is 1.8 better than axial. Does that surprise you, that it's 1.8? Could you have guessed that it would be 1.8? Notice when it was axial, there were two gauche butane interactions. When it's equatorial, those two gauche butanes become anti butanes. And a gauche to anti is 0.9. There are two of them. So 1.8. Right? Two gauches become two antis. So that's axial to equatorial. And, in fact, you could put other groups there. Like instead of methyl group, it could be chlorine or bromine or ethyl or something like that. And for any of these, if you can measure the amount of equatorial and axial, you can get a measure of how big that group is. Right? So that, these so-called A values, the difference between energy when it's axial and when it's equatorial, is a nice rough measure of the effective group size. So we'll stop here, for now, and continue next time.



[end of transcript]

Lecture 34
Sharpless Oxidation Catalysts and the Conformation of Cycloalkanes
Play Video
Sharpless Oxidation Catalysts and the Conformation of Cycloalkanes


Professor Barry Sharpless of Scripps describes the Nobel-prizewinning development of titanium-based catalysts for stereoselective oxidation, the mechanism of their reactions, and their use in preparing esomeprazole. Conformational energy of cyclic alkanes illustrates the use of molecular mechanics.




Transcript



December 3, 2008



Professor Michael McBride: So you remember this slide from several lectures ago. What are they trying to make? Remember what the point of this synthesis was, this particular reaction? The idea was to do a catalytic reaction with something that was chiral as the catalyst, to do an oxidation on sulfur, to make a single enantiomer of omeprazole. And remember the idea was that the product, after you've gone through these three steps that we showed you before, that the product then comes back and is the catalyst again, and you can do another cycle. So, in fact, there's this thing called a catalytic cycle. So what comes in is a peroxide and a sulfide, and what comes out is an alkoxide and the product that you want, the oxidized product; where an oxygen has been added to that unshared pair, that high HOMO. And so the idea of the chiral oxidizing agent was so you'd get a single enantiomer of the product. But do you remember what happened?



Student: It was racemic.



Professor Michael McBride: Right, it gave racemic material. So what did they do? Do you remember? Pardon me?



Student: [inaudible].



Professor Michael McBride: Yeah, ethyldiisopropylamine they added; these researchers added, in 2000. And they found out that that gave a 94% enantiomer excess, but for no obvious reason. Now it would be great if we knew what the reason for this was. And it turned out that, in fact, this kind of oxidative catalytic reaction, to be stereospecific, to give a single enantiomer, was the source of the Nobel Prize in 2001. And if we click on this thing here, we can see -- should see -- yeah, there we go. So this is the Nobel website, and you can see that K. Barry Sharpless here got the Nobel Prize in 2001 for his work on chirally catalyzed oxidation reactions; so exactly this reaction. So if anyone can tell us how ethyldiisopropylamine does its trick, who should it be? Barry Sharpless, right? [Laughs]



[Applause]



[Technical adjustments]



Professor Michael McBride: There, we're all set.



Professor Barry Sharpless: Thanks Mike. Well I'd do anything for my friend Mike, because he saved me from publishing errors in JACS [Journal of the American Chemical Society] at least once. And he's saved a lot of us, who were willing to listen to him, because he's got a genius for seeing flaws in the beautiful stories that people tell, and then they realize that their beautiful stories are going to be too slick, usually; and Nature doesn't like things to be too slick. Well I'm going to just use -- I think the next slide should be --



[Technical adjustments]



Professor Barry Sharpless: Well this is something which we all possibly manipulate in-utero. It's interesting, because many things in our body, like a part can be on the left or right, and a mirror image -- I assume that means they're mirror-image related then. But this is a vein and two arteries that has a sheath on it, and it's the umbilical cord of a mammal. And it's got a nice screw axis; you know how we like to pull our hand down a screw axis. And then I have other reasons why I think I handled it. And then they take it away from you at birth; they cut it and they take it away. It becomes a narcissistically cathected object. And that means you're in trouble because you can't ever heal that wound. You want to get it back, but you can't heal it. You needed that mirror, that umbilical cord, back there when you were born. Anyway, that's why I went crazy trying to do asymmetric synthesis. And now, if you're a chemist, unless you're in deep space -- I don't know if they have any carbon atoms out there, but you got to be in a very high vacuum to get them. But this is the way we make organic compounds, right? But we don't really do it this way because we have to live on a condensed phase of the earth. So these are two -- but you put them together.



This is a point. Then you get a line. It's polyacetylene really. But let me -- give me poetic license of just having it be a line. And I'll take one unit out, and now if I add things across that line, I get into the plane, A and B, or I can add them cis or trans. And essentially you can make anything this way, almost, if you could do it by just surgically adding things. Now we're in the plane. And now the job in the plane -- and I've actually drawn a trans-disubstituted 2-butene. And it has a shape like this, right? It has a methyl group here, a methyl group here, and then hydrogens. So there's open space in these quadrants. And if you look at me coming at you this way, and if I turn around, you see it's going to be -- well you'll see it's going to be a mirror image situation. But I'm flat. So I have no chirality, as a starting point. I'm in the plane of flat. So then if I add hydrogen to it, I get nothing interesting. But if I add an atom, like oxygen, to it, like here, then I've pyramidalized it. And so I can either add the oxygen from my front, and go like that -- epoxide. Or I can add it from my back, and I go like that. And that's the other mirror-image compound. That's a really cool way. And it works. You can make epoxides that way.



And we have reagents today where you can recognize the front and the back of a flat object, that has pro-chirality, like this. And back in the early, in the mid -- early seventies, we saw this industrial reaction that was nice. You could take vanadium and molybdenum, and you could epoxidize olefins with these things called -- well I don't have the darn structure unfortunately -- but it's t-butylhydroperoxide. It's the alcohol group here, with a peroxide bond. And I think you've already learned that peroxide bonds are weak bonds. They have a low lying σ* orbital, and you can attack them on the back side and transfer atoms to make transfer to carbons, like a HOMO of an olefin. So what we noticed -- I noticed in one case that allylic alcohols were very much -- seemed to be special. It was just noticed by English chemists in the thirties that they seemed to react better than just normal alcohols. And so, but nobody had taken any really -- action on that.



Professor Michael McBride: We're not as fast as some people at getting to this stuff. So we should say that an allylic alcohol group is one that has an oxygen on the carbon next to the double bond.



Professor Barry Sharpless: Oh, oh yeah.



Professor Michael McBride: The OH is --



Professor Barry Sharpless: Oh right. I forget. The allylic alcohol -- allylic means -- allylic, you're alpha, one atom away from an olefin. So you could have anything there; nitrogen and oxygen. So you could have an OH, like in ethanol; that's an allylic alcohol. This is an allylic alcohol right -- where is it? I don't -- see, I'm sorry, I don't have the right structures. This is the allylic alcohol. It would be, if I go back one; I'll show you, not by going forward.



Professor Michael McBride: There's the allylic alcohol actually.



Professor Barry Sharpless: Yeah, but I got one -- like here, this is a carbon, that's a carbon, that's a carbon. By putting an OH group there, that's allylic alcohol. If I have another carbon, it's homoallylic. But it has a handle now, and that handle allows it to grab hold of the vanadium. The vanadium has all these alcohol -- these are esters. Imagine that's phosphorous. It is like phosphorous actually. But it's a yellowish liquid, and it exchanges its ligands like the wind, with alcohols. So they come in, they go out. It's a real dance. It's fast. You don't have to worry about that happening. Phosphates, we'd all be dead if that happened; we'd melt. But this is a transition metal. It goes to here. Now, and when I'm the metal, and I have oxygen here, and then I say have an arm. You can imagine my arm being a double bond, with π going like this. And here, on my belly, I'm going to have an atom -- and this is an atom like -- it should be red. Oxygen's red. But he knows why. I don't. Except for blood. So you come like this and you grab it, and it pops off me -- I'm the activated thing on the metal -- and it goes on to there. That's an epoxide; a three-membered ring. And if I attack it with a backhand, I would get the opposite result, in terms of attack. And that doesn't happen, because the geometry of the transition states and the projections of the center -- you can't get as close to a center shot on this thing, with a backhand.



So that's my way of starting to explain. We're going to explain a little bit more. Mike's going to do some acting later, and you'll see it a little bit more clearly -- the topology of this situation. But here it's cool because you get this template. It's like a mandrel. You grab hold of this oxygen atom, which is proto-oxygen. The metal has lots of empty orbitals. It's a d0 metal. It activates the thing this way, and then the olefin lone-pair, the HOMO, attacks the backside, and all this funny business goes on. But there's no mechanism. It's a no-mechanism transfer. It's one of these -- like you draw a bunch of arrows and everything happens sort of in a smooth way; no ions involved, unlike SN2 chemistry. Okay, slow step is then this step. And then it comes off and more stuff goes on and it goes around. So that's the trick we played here. The first idea was take advantage of fixing the thing. So once you've fixed this flat system that you want to make go pop up this way, or that way, now you have a chance of doing it, because the thing you're transferring from can recognize this kind of situation, from that one, by -- it's part of the same molecule temporarily. In fact, it's a covalent bond, right there, at this point. Too much, I always say too much. Okay, let's keep going.



Then came a great man from Japan. He's now a head professor at Kyoto, Oshima. He was a fantastic boiler room producer in titanium, osmium; you name it. I think that was where all the work was done for the Nobel Prize; it was done at MIT. But it happened at Stanford; just because I was there for a few years. But you know those early years in your career, when you do your work in the boiler room, are what produce you. Like your personality doesn't change after you're four years old -- right? -- or three years old; you're pretty much locked in, unfortunately, in some ways. This is vanadium system, and we put on this chiral machinery here, and it gave us 80% ee; that means enantiomeric excess was -- it was 90% one isomer -- enantiomer, the right hand say, and 10% the other. And this looks good. But it only works for that one substrate, and all the other types with it -- you can have like methyls there, or a methyl there, and you get nothing. So you have to get everything customized. This ligand, you got to get the right ligand, the -- and that's what you expect with Nature. We're taught by Emil Fisher -- the lock and key. Nature is supposed to be very, very touchy about changing her preferences. And that was what we expected. Well I didn't expect that. I expected that this type of olefin and this type, with something, that all these would need a class of catalysts. There are about six shapes of olefins. So that's what my dream was, to get maybe six catalysts. Because we didn't have the paraphernalia that nature has all around.



Well for years we tried. Ten years we didn't get anything. It was always this game of like almost no reliability. And then when we came to titanium -- we'd been working with vanadium and molybdenum. We came over to titanium, and we put in tartaric acid ester one day. This wasn't me, it was Katsuki. He should've gotten the Nobel Prize with me, because this means half of this -- I picked the right metal, at that time, and the right system, but he's the one who picked the right ligand. And let's go on to the next thing. You'll see, this is tartaric acid. This is so-called natural. It comes from grapes. It's the cream of tartar. When you protonate it, it becomes tartaric acid. Over here is the unnatural. But there's a plant in Africa, Bauhinia reticulata, that has about -- it's a huge area, around the Sahara. So there's maybe more of that so-called unnatural than there is of the natural. You know the story about tartaric acid. And it goes way back to Louis Pasteur, with the discovery of tetrahedral carbon; I mean then the precursor to it. Grapes, very sexy looking things. We always like -- humans seem to like to drink them, and eat them, and look at them. This is the famous recipe.



So that day, Katsuki took a tartrate and put it together with -- oh I think I forgot to say something here. Yeah, just quickly. Maybe I don't have it. No, it comes later. Okay. TBHP was the kind of oxygen we'd always been using. There's the titanium, which comes from titanium dioxide or titanium tetrachloride. Everything that's white in this world, all these -- that's titanium dioxide. It's the only white pigment anybody ever uses. And so that's titanium dioxide. It's very insoluble. But you can make it into this soluble derivative, if you do the right chemistry. And this is the wine acid. We put in diisopropyl tartrate as an ester, and we kept getting 90%, 100%, 95 -- every olefin we put in that was an allylic alcohol, that had this handle on it, gave Katsuki -- about a week of this, and we were just about dying. I was looking at him; he was looking at me. We didn't believe it, you know. So we had to try to kill this. But it was right. And so it meant that here we had some new principle. We could actually just take one catalyst and get them all; almost get them all. But they had to be allylic alcohols.



So this is the famous recipe discovered I think on an October day in 1980 at Stanford. I'd already decided to go back to MIT. I might not have decided to do that, if this had happened sooner. We get depressed when our research isn't working; I mean, whether we don't show it or not. It's the only thing that matters really to a real killer research guy who wants to understand what makes nature tick. And whatever else -- if that's going well, life is good. Otherwise nothing's good. So I went back to MIT, which I wouldn't have if I'd gotten this about three months earlier, probably. And here we are. There's Katsuki-san drinking tartaric aqueous -- it's about 5% aqueous tartaric acid -- in wine and champagne. And this was on the porch at the Mudd Building, which was where our lab was, on the top. And what I'm going to show here is why Katsuki was so important, in a nutshell. This is really a nice little substrate. It has this type of alcohol; it's trisubstituted. But you see the OH here? There's no OH over here. This is a more reactive olefin for the transfer of the oxygen, but it doesn't have the handle. So this one ends up winning by a factor of 200, in this system where the epoxidation requires the thing to bind and go in intramolecularly. And that's nice in its own right, but to get -- we got racemate all the time, of course, because we have no face selection. And titanium goes about ten times faster with tartrate. When you add it, this does something to make -- every other metal, years before Oshima had taken vanadium and molybdenum, which were the best for isolated double bonds. But you put tartrate in, it kills them. So my instinct would not have been to put tartrate in again.



This is the kind of thing -- everything in science, the rules change all the time. You wake up in the morning -- you should iterate your favorite desire every morning. It's going to look different to you in the shower. And you can never -- the atomic bomb was made by theoretical geniuses. But the guys who did it, actually they had aluminum foil on the source, and they kept it on there for about a year or two. Some guy said, "No, I don't think we need this. What about if we take it off?" That was the answer. I mean, these guys are like us. They just try this, try that, in the lab. Right? That's what gets really things done in this world. The theories are important. Nobody would even think about trying it without the theory. But -- okay, "A man in California just won a Nobel Prize for mixing paint and wine!" That's what they said, in the LA Times. It's sort of true. Titanium dioxide and tartaric acid. And there's the -- I like my students to know where Mother Earth is. And I like things to be cheap. I don't like to be far away from a river that's strong, or a power source; or in this case, this is tartaric acid. It comes in 100-pound bags. This is Spanish tartar, and this is Victor Martin, who's one of the heroes of this chemistry, from the Canary Islands. This comes to Pfizer. They make things out of it. But it comes in ships in 100-pound bags. And we still have that in our stockroom somewhere.



There's my daughter celebrating the synthesis of the unnatural sugars [laughter] by this method. And oh I got the wrong one, didn't I? I threw out the wrong -- yeah, anyway, you're going to have to interpret. This is the mirror-image of the picture that was taken, I guess; because you see the name over here is backwards. But she's looking at the l-sugars. These are the ones we don't make. There's eight of those, the hexoses. And these are the ones Emil Fisher made, in his famous work; about the best chemist who ever lived, except for van't Hoff, who was even better, because he did organic stereochemistry and physical chemistry, and I think he was the greatest chemist who ever lived; van't Hoff was. But anyway, Fisher was better than anybody who lived in this century, in terms of understanding and arguments and rationale of synthesis. And he made a lot of these sugars over here.



But this is Tito Simboli. She's a photographer, my wife's friend. This is taken in 1983. It was on the cover of Chemistry in Britain. And Hannah was seven. And anyway, Tito was a friend. So she took the picture. And this book though, is a huge book, which was gotten by Nancy Schrock, who's in charge of the MIT Archives and is a book restorer. And Nancy -- so here we got Nancy Schrock, we got Tito Simboli, whose husband is Dan McFadden; won the Nobel Prize in economics ten years later, and Dick won the Nobel Prize a couple of years ago, and I won it in 2001. So the picture is kind of connected. And there's Lewis Carroll, of course, looking-glass milk. And Emil Fisher. It's a heavy duty picture; for me anyway. [Laughs]



There's Nexium, and maybe we're getting -- I went too far into the other stuff. But here's the atom, the red atom, that's going to end up being transferred onto the sulfur. And well Mike showed -- told you that the diisopropylamine -- what would its role be? I hope I can go to the board now and show you a few things. But this catalyst is exactly the recipe. And it makes billions of dollars a year, this product. Right? But the patent is -- the Stanford patent is no longer valid. It's run out; even though it hadn't run out when they started. But they didn't believe that it might. This was an infringing of that patent actually in the beginning. But nobody in universities sues companies. They can't afford to. Stanford couldn't afford to.



Professor Michael McBride: Do you want to go to the board now?



Professor Barry Sharpless: Should I? Yeah.



[Technical adjustments]



Professor Barry Sharpless: Do you remember the vanadium picture? I'll make it titanium now. But there's this alcohol group and --



[Technical adjustments]



Professor Barry Sharpless: And on this we have this peroxide group that's bound, and it's bound datively there. It's got this tertiary-butyl group here. And then the alcohol is bound here, and well it goes sort of down. I'll just try to show some of the stereochemistry. And then underneath it's coming -- yeah, it's coming folded like this, underneath. I can't draw very well, but this is supposed to be coming with its lone pairs, it's π bonds, on the backside of this bond. So we're going to do an attack on there, and we'll break this bond, we'll bring the lone pair here in, and we get the epoxide. Now, but the thing about this is patent lawyers can be very creative -- like every field has creative people, and Yale is connected to this fellow, Bert Rowland. And he was in Palo Alto at the time, at a company, and he had just gotten famous, or almost infamous, because that patent was really aggressive; this Boyer-Cohen Patent. He wrote the Boyer-Cohen Patent. He's related to Mike, because he got his Ph.D. with Bartlett at Harvard, like Mike did. So that's a long time ago. That's a great-something relationship. And then comes Wiberg, who's here at Yale. He was at Seattle then, and he took a post-doc. He decided he liked chemistry in principle. He was smart; a physical-organic chemist. But he didn't like the lab; he wasn't any good in the lab and he didn't like to be there. So he became a patent attorney, and did very well. And he writes great patents. And so he read our paper, our first paper. We gave him the little communication. And then he got Katsuki and me over to his office, and he interviewed us for about ten minutes. The next day we got a patent. He never changed a word of it.



And I'll tell you why it's creative, I think. We're getting ready for the Nexium. But I could tell you about that. If we don't make it, it won't matter that much. But here's the main -- so this is going to be the oxygen; and they should be red but we have pink here. So does this one. But that one doesn't transfer. This is the one that transfers. So what Rowland said is you got a metal, which can grab things. Now the things it grabs, it could be X, it could be sulfur, or an amine group that it could grab. And then he just drew -- I think he drew a carbon with no -- it could be any carbon. It didn't have to have to have just CH2. And then he drew this, G. And then he drew -- yeah, he did draw an oxygen atom here, I guess, but it was activated. And so this oxygen, he didn't have a generality there. That was specific. That was the only thing that was specific, these two things. So this could be any kind of lone pair. An olefin is a lone pair. Sulfur has a lone pair. Phosphorous has a lone pair. Nitrogen has a lone pair. And sure enough, when I read that, I said, "My God, this gives us ideas." So we started doing amino alcohols. All these things work. You know? So this is from Bert Rowland, the patent attorney. He was a co-inventor of those other reactions. We never wrote any more patents. It's a good thing. Don't waste your time on patents. They'll just not use them until they expire. And we never -- we made enough to go on a vacation once a year, in a car, not too far away. [Laughter] Katsuki and I, we got -- people think we got rich. We got the most -- one year we got $20,000 each, or something. That was a really sharp maximum, from the bag-a-bug gypsy moth traps – Disparlure.



But okay, to show the transfer, I think maybe it would be nice for us, Mike and I, to show what features have to be over here in these ligands. You see, the tartrate has this feature. It grabs the titanium, like out front. Let's kind of put it out front here. So the titanium is out here. And down in the back here you have this ester group, and up in the back, over here, you have an ester group. And I can't remember if that's what we planned to do. But out here, on the front, it's coming -- the alcohol -- this G-group, and there's an oxygen that's hot over here, that can be transferred. So Mike's going to put that -- I'm going to put that on and --



Professor Michael McBride: I'm the olefin.



Professor Barry Sharpless: Yeah, you're the -- is that okay if you're the ole--



Professor Michael McBride: It's okay.



Professor Barry Sharpless: Maybe it's better if I'm the olefin.



[Laughter]



Professor Barry Sharpless: No, okay. I'll be the titanium.



Professor Michael McBride: Yeah, you're the titanium.



Professor Barry Sharpless: So see, I've got this oxygen that wants to go off, because it's a weak bond. And I've got to get it over to him. He's the allylic alcohol. But we're going to pretend he doesn't have to bind to me. We're just going to do this straight-out attack here. And yeah, it's like that. Right? So I'm blocked here, but I'm open in these quadrants.



Professor Michael McBride: So you're natural tartaric acid.



Professor Barry Sharpless: Is that right? Boy you're quick. I don't -- okay, so I'm natural?



[Laughter]



Professor Barry Sharpless: Okay, well wait a minute. And also, but I'm not like this, I'm like that. You see, I'm in a three-dimensional world, because I'm tartaric acid. So I'm sticking out here. So I'm a chiral object now. My mirror-image won't superimpose. But he's, yeah he's that way. He's a trans-olefin. [Laughter] And he looks like an Egyptian, like -- anyway.



Professor Michael McBride: I have a double bond here, right?



Professor Barry Sharpless: Yeah.



Professor Michael McBride: The question is whether you're going to come on here or on my back.



Professor Barry Sharpless: Actually the double -- we don't have this quite right. Can you get an arm down here? Because the double bonds is here.



Professor Michael McBride: No, no, the double bond is going this way.



Professor Barry Sharpless: Okay. You've got it going that way.



Professor Michael McBride: So here, right now.



Professor Barry Sharpless: Okay, so yeah double bond's going that way. Okay.



Professor Michael McBride: Thank you. So you --



Professor Barry Sharpless: Yeah, that's good.



Professor Michael McBride: Get your tartrate on.



Professor Barry Sharpless: Okay, because I could go like -- I may not -- it's like this, I'm sorry. Yeah, I got to go like that. I could go like that. That's the mirror-image. Okay, now I'm coming towards him.



Professor Michael McBride: But I want to go on my back.



Professor Barry Sharpless: Okay, you want it on your back? Oh my gosh. Oh I can't do it. See? But that's the other -- you need the other mirror-image for that one. So I come this way, and I can -- it says "Think Safety". Oh. So now you have to -- yeah, if he was a good yoga person, he could pyramidalize it a little better. [Laughter] Yet now when you put him that way, he's one enantiomer, and if you put him on the other way, that's the other enantiomer. Simple. It's real simple, isn't it? Thanks Mike.



Yeah, we just organized that ten seconds before we came over, as you can see. We didn't have any red ping pong balls. It's good with Velcro. You can do it with Velcro. We did it once with Velcro red balls. Okay now.



Professor Michael McBride: Back to the screen.



Professor Barry Sharpless: Yeah, I think we're -- oh no, I'm not back to the screen. Let's leave it alone. This is going to be a little --



[Technical adjustments]



Professor Barry Sharpless: Now if you look at the structure of the omeprazole. I won't put all the bells and whistles on it. It's an imidazole structure, benzimidazole, which has a benzene ring, two nitrogens. It's a five-membered aromatic heterocycle. And so you have a double bond and one H. And I can't remember what's on the sulfur. Is it an aromatic ring, or is it a benzyl? It's an aromatic benzyl thing, I think. And you might look at this and say, "This doesn't look anything like Bert Rowland's thing there." Because where are the anchor points that we need, the group that binds the metal? We have the lone pairs all right; you know, two lone pairs, rabbit ears. And there, if you put oxygen on one side, it's the same story as before. This isn't flat to start with. But now we have these pairs. If you put something here, you get one enantiomer. You put something there, you get the other. So the idea was to try to get -- and Kagan had done this. He's a fellow that could've won the Nobel Prize, I thought should of. But he's in France. And he had done a lot of great work in asymmetric chemistry. And he took the titanium tartrate catalyst of Katsuki, and he found if you put a little water in it, and you did the right things, and wished -- it wasn't as general, but he got very good asymmetric addition.



So titanium was already known to do this. But the feature that's important here is the pKa of this benzimidazole is probably -- I looked it up; I think it's below that of an alcohol. So that means it can go on the titanium, reversibly. So you see the titanium has alcohols on it, or ligands like this, or -- and this can exchange. So we can have that come off, and this N go on, for this heterocycle, and the benzene ring's down here. And now we have something like that, a covalent bond. And we've got the sulfur here, and we also have the alcohol. So there's an equilibrium where you could invoke that. But this is not as easy to probably do as an alcohol. It's more encumbered, and it's probably -- kinetics are slower for this, and off of oxygen, for the hydrogen transfers that are needed. So I hypothesize here that what the diisopropyl -- what the Hünig's base, we call it diisopropylethylamine is doing. This is such a hindered amine that it can't react with anything, but it's good for getting protons. So it can't itself get at the oxygen; which you would worry about. It would make an N-oxide. And it can't bind to the titanium anyway. Titanium hates nitrogen. If you have an oxygen around, it'll spit it out so fast it'll make your head spin. I mean, this thing is amazing hating nitrogen -- titanium, and silicon too, but especially titanium. So it's never going to get involved with titanium.



So what's it going to do? Well remember when you try to get chemistry going in a system like this, this amine has enough pKa power to pull this proton. And usually you could write a concerted mechanism. Like let me erase some of this and have some room here. Well I shouldn't erase that. I'll go over here. You got this benzimidazole, with a sulfur here, and I'm going to put the hydrogen up here. Whenever you engage compounds in reaction, acid-base reactions, you look for the basic sites and the acidic sites. There's the acidic site. You might think that's where the titanium goes. But usually the mechanism for these reactions involves a simultaneous loss of this proton, and attack here. This is the place you can attack. So you'll be attacking the titanium here, and you're going to get some help from -- you're going to get a lot of help from this amine, because you can put it someplace; just take it right off. But you'll have some of that ammonium salt around too, a little bit, transiently. And this will help you, because you need to move things around. That's why this reaction was not good without that amine. You need to scramble the system and keep it rolling, so that the things get on and off.



Catalysis often has this -- like there's six, maybe. There's always a loop, and if you look at any catalytic cycle, there'll be usually one step that's really bad news, until you get it fixed. And that step is where it controls the rate. If you get steady state, the slowest step -- this is a real democracy, catalysis. There's no step that's more important than any other. The catalyst goes from each step. The titanium is moving through those cycles. If there's a slow step, there's 99.9% of the titaniums that you need, stuck before this one mountain, that goes way up like Mount Everest. You got to drop that one down. If you get that one down, the rate goes way up. And if you get them all the same height, you're really rolling. And that's a sacred rule. The turnover-limiting step is everything, in catalysis. And a lot of times the turnover-ruling step is so high, they don't even know there is catalysis. If you can break that, you'll find a world of catalysis that didn't exist before. So that's what I think is exciting about catalysis; it's alive. As long as you have some energy to dissipate you're -- it's life itself.



Then I'll finish this quickly off, without any details. You can see what I'm going to do here. It's just that that hot oxygen atom, that's activated to transfer. This is the handle that would be the allylic alcohol. And so it's a dead ringer for the Katsuki -- for the Bert Rowland Patent, right? That's what I think. And the amine, yeah, the amine makes sense here too. And I guess that's --



Professor Michael McBride: So did that patent apply to this process?



Professor Barry Sharpless: Well Stanford made a little -- they brought over ten people, from research, and they brought over three lawyers from AstraZeneca; because Stanford thought maybe they should get something out of this. And I told the story like this, and they didn't -- they pretended it didn't make any sense at all. You know? And it was pretty unsatisfying. Bert Rowland was there. I don't know if he's still alive. He had cancer then. But so we didn't -- but they'd already made six billion dollars; because the catalyst patent was the only extant patent. But it did have this feature in it. And so they were right in the -- I think they're right in the face of the patent. But they didn't like that. And we didn't -- Stanford wasn't going to pursue it.



Another thing we -- I'd like to mention -- is they invited me over. This is before this meeting about the patent. I guess I started to notice the old titanium chemistry, and Katsuki and I talked, and we agreed that maybe we should raise our hand. But before that, what happened was I got invited to go to a hotel in Munich, the fanciest hotel, and they put me in this big room upstairs. They wanted me to say something about right- and left-handed medicines. And I had a little 15-minute spot -- 1000 or 1200 gastroenterologists, from Germany alone. And they were doing a story about how important drugs have to be optically pure. And that story doesn't wash too well here. And I didn't realize it, when it was going. But I went and I did my part. And it is true that many drugs are toxic in one form and good in the other. And I was able to make those points. But what happens here is -- oh, and also they gave me this suite on the top, that had like a spiral staircase to the bed, at the very top; and Madonna had stayed there the week before. [Laughter] That's okay for her. She doesn't have a prostate problem. [Laughter] I hated that place. I had to go -- they should've had a fire pole so I could get down to the bathroom.



Anyway. What happens is, as Mike told you, you just look at the facts in that system. And I noticed them just the day I was sitting there in the audience, and the press was interviewing us. And I whispered to the guy next to me, and he looked at me and frowned. It was about the weights. All right? I mean, the racemate 20 mgs, and optically pure. And I think the optically pure is substantially more reactive, in this case, because it gets in better. It's, of course, just a pro-drug, for the thing. But the thing that was missing from the experiments is the forty milligrams of racemate, which would give them equivalence against the one enantiomer of the twenty milligrams of pure. And that was missing, and they doubled the -- and it was just a sore thumb there for me. But I didn't say anything; nobody asked me, fortunately. Because if they had, I would've said something about it, I imagine. It's the way we are. Right, Mike? Mike doesn't -- he smiles, but he gets the hard evidence. In the middle of your talk sometimes he does this. But for me it's the best thing you can get.



Professor Michael McBride: [Laughs]



Professor Barry Sharpless: If you can get killed by a friend, it's the best way to get killed.



[Laughter]



Professor Barry Sharpless: Anyway, I'm finished. That's it.



Professor Michael McBride: Great.



[Applause]



Professor Michael McBride: Great. Thanks again.



Professor Barry Sharpless: Yeah, you're welcome.



Professor Michael McBride: We maybe have time for one question or so, before we get back to our normal business. Anybody got a question for Professor Sharpless? No. Well thanks again.



Professor Barry Sharpless: You're welcome.



Professor Michael McBride: Have a good trip back.



[Informal discussion between Professor McBride and Professor Sharpless]



Professor Barry Sharpless: Okay thanks. So long. Good luck.



[Applause]



Professor Michael McBride: Okay, back to our routine now.



[Informal discussion between Professor McBride and Professor Sharpless]



Professor Michael McBride: Oh, he forgot to show this picture. There he is with the royalty of Sweden, and his wife, and the other Nobel Laureates in Chemistry. Okay, and he was going to talk about carvone too. He wanted you to smell it. I told him you'd smelled it. But he had an idea for a novel based on carvone, which he didn't have time to get to. Okay, so we were talking about the conformation of rings. And we talked last time a lot about cyclohexane, and how it distorted -- remember, that it's not really ideal, with the axial bonds parallel to the axis. They spread out a little bit. The ring flattens a little bit to minimize the energy, as calculated by molecular mechanics. Now how about in a four-membered ring, instead of a six-membered ring? Well if we look at the, what molecular mechanics says about the source of strain in this system, you can see that the big contributors are bend -- that's what Baeyer had already talked about, about the ring strain; having 90° angles is not going to be good; so that's costing 13.5 kcal/mol -- and torsion. Because why is there such high torsional energy; 15 kcal/mol, almost? What's the source of that? Anybody see? As you go around the ring, everything is exactly eclipsed. Everybody see that? Every carbon-carbon bond is eclipsed. Right? So that, in molecular mechanics, sums up to almost 15 kcal/mol. Now, if you were -- how would you try to minimize the energy? Is this probably the lowest energy form, or can you think of changing it, so that it'll be lower? Any ideas? Kevin?



Student: Make it so the bond, in cyclobutane, aren't exactly parallel to the new bond.



Professor Michael McBride: Yeah, if you make it not exactly eclipsed, if you make it a little bit towards scatter- towards staggered, by twisting about the bonds, that'll lower that 15 kcal/mol. But, can you see what happens, if you begin to twist around the bonds? It sharpens the angles even more, from 90° to smaller than 90°. So here's what happens when you do it. The bending energy goes up, by 2.5 kilocalories per mole. Right? But the torsional energy goes down, by 3.5 kilocalories per mole -- right? -- if you don't bend it too far. Okay, and the other things stay pretty much the same. So this is a competition between torsion energy and bending energy, and the minimum is reached with a little bit of bending, in order to relieve some of that torsion.



Now how about if you had a five-membered ring? Now you notice you have -- notice the bending energy isn't bad, because as Baeyer pointed out, a pentagon, a regular pentagon, has just the right angles -- 109.5° about. Right? So there's hardly any bending. But there's still a lot of torsion, from the eclipsing of the carbons. So what you do, if you run the molecular mechanics program, is to cause bending energy to occur; that is, sharpen the angles a little bit, right? But you get rid of a lot of torsional energy by doing that. And this particular conformation -- which you can see how it's bent down; the black carbon, eight there -- is sometimes referred to as the envelope conformation of cyclopentane. See how it looks like the fold of the envelope bent down? Okay, so there again is a competition between bending and torsion.



Now, like plastic models, molecular mechanics is satisfying because it says not only what the structure should be, but why. What is it that makes the energy the way it is? Okay, so you remember last time we passed around the cyclohexanes. Remember how they clicked, to go from one form to another? So what's happening? The question is, what's the source of the barrier to the cyclohexane ring flip? So here's the chair cyclohexane, and when the ring starts to flip -- remember first you go toward a boat, by bending one of those carbons up, like that. Now what do you think the source of the energy is, that makes this hard to do? Why do you go to a maximum energy, as you do this, and then it clicks and it goes down again? Sherwin?



Student: Torsion.



Professor Michael McBride: Torsion. Now, that's a good point. There are, in fact, two bonds, two butanes, that become eclipsed. Right? Both of the ones on the right. This one is eclipsed, but also this one is eclipsed; those four. Right? And indeed, that's worth about 7 kcal/mol of going uphill. But there's an interesting point. You did this with the models -- right? -- and you felt it click. How do those models, those plastic models, know from torsional energy? Do you see what I mean? If you take ethane in those models, and spin it, it spins completely freely. There's nothing in the model. You could make it -- you could make the thing that goes into the tube, and the tube, a little bit triangular, so that as you tried to twist, it would go up in energy and then down again. Right? Do you see how you could do that in a model? But that's not the way those are made. They're cylinders, they rotate freely. So why did the model click? Pardon me?



Student: Bending.



Professor Michael McBride: Bending? Of what?



Student: The bonds.



Professor Michael McBride: Right, exactly. Look at this. Why does the plastic model click? Because when you're halfway across, it's becoming -- if it were flat, if the whole ring were flat, what would the carbon-carbon-carbon angles be, if the ring were flat? Sam?



Student: I don't know.



Professor Michael McBride: In a regular hexagon, what's the angle? Dana? You got to be careful holding your hand up, even if it's just to scratch your face. What's the angle in a regular hexagon?



Students: 120.



Professor Michael McBride: 120°. It wants to be 109. So it has to stretch out, if it becomes flat. And even when it becomes only partly flat -- right? -- this part of it is flat; this one's out of the plane. But that means in order to have this one in there, this bond gets compressed. Right? Because if these angles want to be 109, instead of 120, everything -- these are going to move; this one and this one will move closer together. So these are -- in this form, when it's half planar, these angles are bent out, and that angle is bent in. So it's the angle bending which in fact does it, does stress those plastic models, which causes it to click. So it's interesting to be able to look at models, or at these calculations of molecules as springs, and see why various things occur. So are molecular mechanics programs useful? Yes, definitely they're useful. Are they true? No. Okay, so we'll go on from there next time.



[end of transcript]

Lecture 35
Understanding Molecular Structure and Energy through Standard Bonds
Play Video
Understanding Molecular Structure and Energy through Standard Bonds


Although molecular mechanics is imperfect, it is useful for discussing molecular structure and energy in terms of standard covalent bonds. Analysis of the Cambridge Structural Database shows that predicting bond distances to within 1% required detailed categorization of bond types. Early attempts to predict heats of combustion in terms of composition proved adequate for physiology, but not for chemistry. Group- or bond-additivity schemes are useful for understanding heats of formation, especially when corrected for strain. Heat of atomization is the natural target for bond energy schemes, but experimental measurement requires spectroscopic determination of the heat of atomization of elements in their standard states.



Transcript



December 5, 2008

Lecture 36
Bond Energies, the Boltzmann Factor and Entropy
Play Video
Bond Energies, the Boltzmann Factor and Entropy


After discussing the classic determination of the heat of atomization of graphite by Chupka and Inghram, the values of bond dissociation energies, and the utility of average bond energies, the lecture focuses on understanding equilibrium and rate processes through statistical mechanics. The Boltzmann factor favors minimal energy in order to provide the largest number of different arrangements of "bits" of energy. The slippery concept of disorder is illustrated using Couette flow. Entropy favors "disordered arrangements" because there are more of them than there are of recognizable ordered arrangements.




Transcript



December 8, 2008



Professor Michael McBride: Okay, at the end last time we were looking at how you could possibly know the heat of atomization of graphite; how much energy it takes to put a carbon atom into the gas phase from graphite. Why would you want to know that? Why is that a key value? For what purpose? If you want to be able to know whether you can add together bonds and get the energy of a molecule, that means you start with separated atoms and you then see how much energy is given off when those come together to give a particular molecule, to make a whole bunch of bonds. Right? Now you can measure the energy of a molecule with respect to CO2 quite easily. How do you do that? Burn it. And you can measure the CO2 relative to graphite. How do you do that? How do you know the energy of a carbon in -- or the energy of CO2 relative to the energy of graphite plus oxygen? You burn graphite. Okay? But the last thing you need is the energy of graphite relative to the atom. And if you then knew that, then you could complete this scheme and know, just by burning things, whether you can add together bond energies to get the energy of a molecule.



Okay, so how do you get the energy from graphite to an atom? One way is spectroscopy, and we talked about that last time, that if you see what the minimum amount of energy you can put into CO2 and have it break into two atoms, that gives you the energy of the two atoms relative to CO2 -- pardon me, relative to CO I should be saying. Right? And then you burn CO and get it relative to CO2, and then you have everything you need. Okay? Okay, except that people who were very smart, Nobel Prize winners and so on, differed on interpreting this, because some of them thought that when you break [CO] into atoms, with the light, the atom you get is not the lowest energy state of the atom, but a higher energy state of the atom. So some of the energy you're putting in is going into making an excited atom. And the true energy of forming the minimum-energy carbon atom from graphite is lower than what you do spectroscopically. So the spectroscopic value is very precise. You can measure the position, the color of the light, very accurately. But they didn't know what it corresponded to, so they needed a different way to know what the energy of the carbon atom was relative to graphite.



Now Professor Sharpless said -- you know, he talked about increasing dimension of carbon; start with an atom, go to a line, polyacetylene and then to a bunch of double bonds, and finally to a saturated hydrocarbon -- he said that carbon atoms are very hard to come by. You need a really, really low vacuum to get it. Right? That is, the equilibrium constant for carbon atoms coming together to form bonds is very, very favorable. Right? So it's hard to get atoms. So how are you ever going to get atoms? Well let's think about that problem here. So we know that the equilibrium constant at room temperature is 10-(3/4)ΔE. So you can measure k. Then you know ΔE. So, but if the equilibrium you're trying to measure is between graphite and carbon atoms, and the energy to take a carbon atom out of graphite is 170 kilocalories/mol, then the equilibrium constant is 10-127. Right? And that means, since there are only something of the order of 1080 atoms in the universe, there would not be, at room temperature, a single carbon atom at equilibrium with graphite if everything in the universe was graphite. Right? So it's not a very favorable equilibrium constant. What could you do about it, if you wanted to measure the equilibrium constant, and in that way get the energy difference? Any knobs you can twist? Lucas?



Student: Temperature.



Professor Michael McBride: You could increase the temperature. Right? Because it's ΔE/kT. And remember, we're talking about room temperature. So if we're much higher than room temperature, then the exponent gets much smaller. So suppose we went to ten times room temperature, to 3000 Kelvin. Right? Then it would be, instead of 3/4ths, it would be 3/40ths, because the denominator would be ten times bigger. And now it would be 1 in 1013th would be an atom. And now that's a substantial number, if you're talking about Avogadro's number. So if you had something really, really sensitive, to measure atoms, you might be able to measure the atoms at equilibrium. But you have to establish equilibrium at something of the order of 3000°, or at least really, really hot. Right? So you could write the equilibrium constant, the concentration of atoms over the concentration of graphite, in this way. That has to do -- that is the heat of formation, atoms relative to graphite. Right?



Now, of course, exactly what one means by the concentration of graphite needs to be -- you scratch your head a little bit about that. Right? But you don't actually need to know it, because if you could measure the pressure of the C atoms at equilibrium with graphite, at very high temperature -- call that the pressure of carbon, right? -- that is some constant, and that constant will include whatever this concentration of graphite should be. So B. So multiply both sides by that concentration of graphite. Right? So we get some constant on the right, and then that heat of formation, in the exponent, that we want to find. So that means we could write -- if we took the log of both sides. You have the log of the pressure of carbon atoms, is the log of whatever B is -- that's some constant -- minus this other term. And what that means is that if you -- that that minus the heat of formation of the carbon atom, divided by R, is the slope of a plot of the pressure of carbon atoms, versus 1/T. Does everyone see that? In that equation, that equation says that there's an intercept, which is the log of b, and a slope, which is -ΔH formation to carbon, divided by R, if you plot against 1/T. Everybody with me on that? Nod if you see it. Okay. So all you need to do is plot the log of the pressure of carbon atoms, which will be very low, versus 1/T, at very high temperature -- right?; of the order of 3000 Kelvin.



So that's not easy to do. But it was done in 1955 by Chupka and Inghram. And this is the sketch of their instrument, and I'll show you how they did it. First there's a graphite cylinder. Okay? And around it is a can made out of tantalum. Now why tantalum? Because that's about the highest melting thing you can get; it's the highest melting metal. So 3293 Kelvin is its melting point. So you could really heat the heck out of this thing. Okay? And now you surround that with wires, and those wires are made of tungsten, because that's very high melting too. And so you connect wire -- you connect electricity to the tungsten, and also to the tantalum. So electrons boil off the tungsten and bombard the tantalum, and when the electrons hit it, they heat it up. So you can heat the heck out of this tantalum can by bombarding it with electrons. And that, of course, heats the graphite that's inside. Now, so inside there's going to be a gas of carbon, in equilibrium with the graphite. Now it won't just be C atoms. It'll also be C2, C3, C60, C70, and so on; but lots of different forms of carbon, as a gas. So you can't just measure the pressure in there; and indeed it would be tough to measure the pressure inside, at 3000°. But you at least have these things there at equilibrium.



Now what they did, Chupka and Inghram, was to drill a tiny hole, in the top of this thing. And that will allow a little bit of the gas to escape; not so much that you destroy the equilibrium inside, just a really slow leak. Right? So now, if you could measure the amount of these different carbon species leaking out, you could know what the pressure of them was inside the can. Okay? Now, so there's a beam of carbon species coming out, gaseous carbon species. This is all at very, very high vacuum. Right? And then up there, an electron beam comes across and hits these carbon things, and knocks electrons out of them. And that converts them into C+; C1, C2, C3, C60+. Okay? So you have a beam coming out of these charged carbon species. Now why do you want them to be charged? For two reasons. One, so you can detect them; because you can detect it when charge hits a plate. And the other one is so you can deflect the beams with a magnetic field. Right? So these various carbon species come out, and they go into a magnetic field, which puts a lateral force on them and causes them to bend. And, of course, the lighter they are, the easier they are to bend if they all have the same charge. So you're going to get curved trajectories. C3+, C2+, C1+ will be the most bent. And now if you put a detector at these different positions, you can see how much C1, how much C2, how much C3 there was. Okay? So that's how you're going to do it. But this machine has to be pretty special, because you're heating it so high. Lucas?



Student: How do you know the electron beam only knocks off one electron?



Professor Michael McBride: It can knock off two, but then you get it at a very different position. You can tell these things. This thing is called mass spectroscopy. But it knocks off one much more often than it knocks off two or three.



Student: And isn't this kind of the same problem as we had with the other thing, where the C is ionized?



Professor Michael McBride: Which other thing? I can't remember what problem --



Student: You said that we couldn't measure the other one because it was a high energy state, spectrographically.



Professor Michael McBride: Oh. Might it be an excited state?



Student: Yeah.



Professor Michael McBride: No because it's in equilibrium. [Note: the initial fragments are at equilibrium. No one cares if the ions are in their ground state. They are only being used to measure the abundance of fragments, and their only relevant properties are charge and mass.]



Student: Okay.



Professor Michael McBride: The excited state would be much, much less at equilibrium, because it's higher in energy. Right? That's a good point though; good that you thought about that. Now, so you have to put shielding around this stuff, so it doesn't melt everything. So you use tantalum, which is high melting, and a series of can-inside-a-can, like Russian dolls. Right? So the inside most one is very, very hot; then a little less hot, a little less hot. Okay? So you shield it. And now you need to know the temperature inside. And you do it by drilling a tiny hole through those shields so you can have what's called an optical pyrometer. That's just something that looks at the color of something that's really hot. And something that's hot gives -- even something that's cold -- gives off black-body radiation. And the color has to do with the temperature; you know, there's red heat, white heat, blue heat, and so on. So by measuring the color, you can see what the temperature was inside. And that window, that the light comes through, has to be made of quartz, not Pyrex glass. You know why?



Student: Pyrex would melt.



Professor Michael McBride: Because Pyrex glass would melt, even at that distance. Right? So you use quartz glass. Okay, so that's what you do. And here's a graph of the pressure of these various species, measured with that mass spectrometer, at different temperatures, measured as one over temperature here. So it goes from 2150 Kelvin to 2450 Kelvin, and the pressure increases. And it's plotted as the log of the intensity of the signal; that is, the pressure times temperature, because the pressure has to be corrected for temperature. Because if you have the same number of things giving the pressure, but they're hotter, they'll be pressing harder on things. It's the intensity of the signal times the temperature that you plot there. And from the slope, you can see up there at the top, that for C1 you get -- the slope says it's 171 kilocalories/mol; Q.E.D. Right?



Now you know which was the right one, measured by spectroscopy. It was the one that said 171. So this experiment, actually measuring the equilibrium between carbon atoms and graphite, by a really gargantuan kind of experiment, is what settled the question finally. So that when you look at the appendix of this book, Streitwieser, Heathcock and Kosower, you find that there are heats of formation for atoms and radicals, measured by spectroscopy and things like this. And there you see carbon, 170.9. And that was done by Professor Chupka, who was in this department and who used to come and tell people how he did this experiment. But as you can see, he passed away in 2007. So thanks to Professor Chupka. And the nice thing is, once that's done, it's done. Now you know it and you just plug it in when you burn your stuff and want to know what its energy is. You can get it relative to carbon atoms in the gas phase.



Okay, now how good are these spectroscopic experiments? Well this is a neat thing. The heat of atomization of methane, measured in the ways we've just been talking about, is 397.5 kilocalories/mol. Now that comes from mating a carbon atom with four hydrogen atoms; that is 397.5. Right? So we know what the average bond energy. There are four such bonds. So each one's worth 99.4 kilocalories/mol. So about 100 kilocalories/mol for a C-H bond; that's convenient to remember. Okay? But that's not how much it costs to take a single hydrogen atom away from methane. Taking a single hydrogen atom away from methane, the so-called "bond dissociation energy", which is the actual experimental energy it takes to do some particular process -- average bond energy is just an average, but the individual ones are not the same -- taking a hydrogen away from methane is 104.99, plus or minus 0.03 kilocalories/mol. Close, but not the same thing. And then you have CH3. If you take a second H off that, it's 110.4. The next one is 101.3, and the final one, taking H away from C, is only 80.9. Right? Now these are done by spectroscopy. And Barney Ellison may come and talk to us in the spring; he's often traveling through and talks about how he measured these things. But those are done by spectroscopy. But the neat thing about it is if you add all those four numbers together -- so, pardon me, I was going to say no individual bond equals the average. Right? But if you add them together you get 397.5, which is precisely the average.



So these are very good experiments. So we know, through heroic spectroscopy and this work of Chupka and Inghram, we know what these energies are, bond energies, and bond dissociation energies. So here are average bond energies in a table that you have at the end of this organic chemistry text. And it says a carbon-hydrogen bond is 99; and now you see where we get that. And you see that a carbon-carbon bond is 83. But the second carbon-carbon bond, in a double bond, is only 63. Right? Why is it weaker? Why is the second bond of a carbon-carbon double bond weaker than the first? Pardon me? Devin, what do you say?



Student: The overlap.



Professor Michael McBride: Yeah, you have bad overlap between the π electrons. In fact, the first, the single bond of a double bond, is probably stronger than a normal single bond. Can anybody see why that would be?



Student: More s character.



Professor Michael McBride: It's got more s character, better overlap; sp2-sp2. So the second one is probably more than 20 kilocalories weaker than the single one. But at any rate it's 146, that you add up to get a double bond, and 200 for a triple bond; which means the third bond is worth only 54 kilocalories/mol. Okay? And in C=O, it's about the same as C-C. So C-O is eighty-six. But the double bond, notice, is different in this case. Now the second bond is stronger than the first. So the carbonyl group is an especially stable group. Okay? So you have the question, can you sum up these average bond energies and get useful heats of atomization? So can you look at a structure and say how stable it's going to be?



Okay, so let's try it. So here's heats of atomization by additivity of average bond energies. So we have these average bond energies from the table. A C-C single-bond is 83. A C-H is 99. C double bond C is 146, 86, 111 and so on. And we're going to sum them all up to get the heat of atomization -- compare it with the actual heat of atomization. Okay, so for ethene, there are four C-H bonds, there's one C-C double bond. Add them up, and you get 542. The actual heat of atomization is 537.7. So there's an error of 4.3 kilocalories/mol, which is less than 1%. That's pretty good. But on the other hand, ethene probably entered in to determining these average bond energies. Right? So it's not 100% fair. Okay? How about cyclohexane? Now we have 6 carbon-carbon single bonds and 12 carbon-hydrogen single bonds: 1686 versus 1680.1. An error of only 5.9, less than half a percent error. So pretty good, by adding up bonds. Cyclohexanol. Remember we had quite a bit of trouble with these partly oxidized things before, when we were trying to just do it on the basis of the elements. But if you add bonds together, you get within 0.3% of the right value. Or if you do glucose, which has lots of oxygens in it, then you get within again less then 1%; 0.7% of the right value.



So this is pretty darn good; very impressive, very small errors, to predict these. But the question is, is it useful? How accurate does it have to be, to be useful? So why do you need to know the values? Because you want to know equilibrium constants. You want to know which direction a reaction will go, for example. Okay, so we know that the -- if we want to calculate an equilibrium constant, we can do it at room temperature with this 3/4ths trick. So the calculated equilibrium constant is whatever we're calculating here, for energy, between two things. We have two things: calculate the energy of these, the energy of these, compare the energies, and that'll give us the equilibrium constant, according to this formula. But notice I'm doing it on the basis of calculation. Now so that's -- whatever -- the calculated energy is whatever the true energy would be. But there's also some error in there. Okay? But if you add two exponents, that's the same as multiplying two things together. So the calculated equilibrium constant is the true equilibrium constant -- that's the first part, the part that it has with ΔH true -- times the part that has this exponent. Right? 3/4ths of the ΔH error.



So if you want the error to be small, that factor to be small, then ΔH error has to be small; small, not on a percentage basis, but absolutely it has to be small. That error, not the percentage error, determines this error factor. Right? So to keep the error less than a factor of ten, in the equilibrium constant, you need to know the equilibrium constants within 1.3 kilocalories/mol, so that 3/4ths of it will be one, and that would mean you'd be within a factor of ten. Everybody with me on this? So you need to do even better than this. You can't use the average bond energies and get something that's very useful, because if you're off by sixteen down here, in the case of sugar, that means you're off by a factor of 1012th in predicting the equilibrium constant, which wouldn't be acceptable probably. Okay?



So let's try it with the equilibrium between a ketone and the so-called enol, which is an isomer of a ketone in which a hydrogen has been taken from the methyl group on the right and put on the oxygen, and the double bond moved. Okay? So that's a very important equilibrium that we'll encounter when we talk about the chemistry of ketones. Let's see what the equilibrium constant -- should there be more ketone or more enol? Do you have a guess right at the outset of which one would be more stable? I would guess the ketone, because of what I just told you, that the C-O double bond is remarkably stable. Right? In the other case you have a C-C double bond. Okay, let's see. Now we could add together all the bonds. But most of them, most bonds are the same between the starting material and product. We only need to compare the ones that change. Okay? So we've highlighted in red the bonds that change between the two forms, the two isomers, because we're interested in the difference in energy between these two, to get the equilibrium constant. Okay, so these are the numbers I took from the table, that you see on the top left there: 179 for C-O double bond and so on. And I sum them up and that's 361, for those bonds. And the new set of bonds, in the enol, sum to 343. So the ketone indeed is more stable, it appears, by 18 kilocalories/mol. So 18 kilocalories/mol means that you have a factor of 1013th; the equilibrium constant is 1013.5. So it should lie, for practical purposes, entirely in the direction of the ketone. However, if you do it experimentally, you find that the equilibrium constant is only 107th, not 1014th. Right? So the true energy difference is 9.3 kilocalories; not the eighteen that we got by adding bonds together. Right? So that means we're going to have to deal with addressing why the enol is too stable. It's 9 kilocalories/mol too stable, compared to our predictions, on the basis of adding bonds together.



Now why? Well one thing is that we -- that those bonds that we cancelled, the C-H bonds that didn't change, in fact did change between the starting material and the product. Why could I say that they did change? In both cases, on both sides, there are single carbon-carbon, carbon-hydrogen bonds. How can I say they change? Angela?



Student: Well with the ketone, they're sp3 hybridized.



Professor Michael McBride: Ah ha.



Student: In the enol they're sp2 hybridized.



Professor Michael McBride: Right. They're changing hybridization. Actually, yeah, they go from sp3 to sp2, on the carbon, as you go across. And the sp2's on the right should be more stable. Okay? So the sp2-H should be stronger. So these things that I was saying cancelled do not actually cancel, if we take hybridization into account. So that's one factor. And there's another as well, which is you have that unshared pair, on the top right here, on the oxygen, is adjacent to a double bond. That means that this high HOMO can be stabilized by the π* low LUMO; it'll overlap. That isn't a possibility here, where the unshared pairs on the oxygen did not overlap with the π* orbital. So you get intramolecular HOMO/LUMO mixing in the enol that you don't get in the ketone; which will help stabilize the enol, with that -- we could draw that resonance structure. So those two things together make up that 9 kilocalorie error; or at least we can think -- they contribute to it at least. So constitutional energy, what we would get by adding bonds together, has to be corrected for various "effects", we'll call them, such as resonance, that's what we just looked at, like this HOMO/LUMO thing, such as hybridization changes, or such as strain, as in the case of axial methylcyclohexane, that we looked at. So there are lots and lots of these corrections that you have to apply to this model, where you add together bond energies in order to predict the energies of a particular molecule.



But for many cases now, you can do a pretty good job of predicting these things, and actually not do so bad at predicting relative energies of isomers, and therefore equilibrium constants. And these effects, of course, are a polite name for error. Right? They're correcting -- various ways of correcting errors that you think there should be in this scheme of just adding bonds together. Now, energy determines what can happen. Things always move toward equilibrium. Right? So if the ratio of two things is something, but the equilibrium ratio is different, the ratio will always move toward that, toward the equilibrium, if it's in isolation. But there's another equally important thing is how fast will it go there? And that, as we've seen before, can be approximated as 1013th/second, times this same kind of factor, relating to the barrier.



Now both of these things suggest that being low in energy is good. Right? You favor things that are low in energy. But you might ask why? That's not what people say about money. They don't say the less money you have the better. Right? Why the less energy the better? This is a really interesting case, and it has to do with statistics. And especially at Yale we should talk about this, because in 1902, when Yale celebrated it's bicentennial, they published a number of books showing off the scholarship of Yale; as you can see here. And the most important of those books, by about 500 miles, was this one: Elementary Principles in Statistical Mechanics and the Rational Foundation of Thermodynamics, by J. Willard Gibbs. So it's statistical mechanics. It's trying to understand the behavior of chemical substances, on the basis of statistics.



Now when you do this, you get exponents. And the organization of our presentation here is going to have to do with three different ways in which statistics enters into exponents, for purposes of doing equilibrium. So there's the Boltzmann Factor; that's what we've been talking about, the 10(3/4 ΔH), that's called the Boltzmann Factor. It includes the Boltzmann Constant. Then there are things that have to do with entropy, which often seems to be a very confusing topic. And finally there's a thing called the Law of Mass Action. And all of these things have exponents in them. And if you understand how the exponents behave, you understand what's going on. So let's look first at the Boltzmann Factor. So here's Ludwig Boltzmann, who committed suicide in 1906. And this is his important paper on "The Relationship between the Second Law of Thermodynamics and Probability Calculations" -- so statistics -- "Regarding the Laws of Thermal Equilibrium," in 1877. And his key equation is S = k lnW . So log relates to an exponential, and we'll see why that is. And here's his tombstone in the cemetery in Vienna. And you'll notice, up at the top there, S = k ln W. Okay?



So what Boltzmann considered was the implication of random distribution of energy. Suppose you have a certain amount of energy, in a system, but it's distributed at random. Right? So purely statistically. Then how should it be distributed? How much energy should any particular molecule have, is the question. And we can visualize this in a simple case, which is very like what he did, except he did it analytically and in a much bigger system. But just using four containers, which are like molecules, and each one can have a certain amount of energy in it. And we'll consider the energy to be -- to come in bits. He used that idea, that there were bits of energy to be distributed among molecules, or degrees of freedom within molecules. He didn't think that energy came in bits, but it made it possible to do the statistics, and then he just took the limit when these bits get very, very, very, very small, so that it becomes like a continuum of stuff, like a whole sand of energy bits.



But anyhow, let's just count up how many different ways there are of putting three different bits of energy -- or actually not different, they're all the same -- but three bits of energy into the red container. Okay? So if you -- how many complexions -- that's what, any particular arrangement he called a complexion -- how many different complexions have a certain number of bits in the first container? Well suppose you put all three of those energy bits into the first container. How many different ways are there of doing that? Just one. Okay? But suppose you put only two into the first container? Now how many different ways are there of arranging it so that there are two in the first container? How many different ways? There are three: 1, 2, 3. So there'll be three ways of putting two bits in the first container. How many of putting one in the first container? Well we put one there; there's 1, 2, 3, 4, 5, 6 ways of doing it. Okay? So there's six ways of putting one in there. And how many of putting none at all in there? 1, 2, 3, 4, 5, 6, 7, 8, 9, 10. Okay? So there are 10 ways of doing that. So let's make a graph and see how many ways -- what's the probability that you'll have a certain number of bits of energy in the first container? Okay, it looks like that: 10, 6, 3, 1. And what does that curve look like, if you made a plot of that? What type of curve does it look like? Is it a straight line? Anybody got a name for it, or something that looks a little bit like that?



Student: Exponential.



Professor Michael McBride: It's exponential decay. Now it's not truly exponential, in this case, of three bits among four containers. But if you do 30 bits among 20 containers, then it looks like that, and there is an exponential. Right? So what Boltzmann was able to do, to show mathematically, was that the limit, when you have very many, very small energy bits, is truly an exponential. So the probability of having a certain amount of energy, in a degree of freedom, is exponential: e-(whatever that energy is, divided by kT). So Boltzmann showed that that was the limit for lots of infinitesimal energy bits. And the idea behind it is quite clear. If all the complexions for a given total are equally likely -- and that's what he assumes; it's random, they can be any place they want to be -- then shifting energy into any one degree of freedom, of one molecule, is disfavored. Because when you put more in one molecule, there are fewer ways to distribute the rest among the others. So there'll be fewer ways, the more you put in this one. Is that clear? Because that's really the key concept. The more bits of energy you put in this one, the fewer different ways there are of permuting what's left among the others. Right? And it's exponential. Right? So if you have fewer ways among the others, then it's less likely. Right?



So it turns out that if you do this, the average energy is ½ kT, in this degree of freedom; which is to say that k, the Boltzmann Constant, relates temperature to average energy; which is to say that temperature is average energy. Temperature and average energy at equilibrium are the same thing -- okay? -- for each degree of freedom. You can put -- what did we mean by these little buckets into which we could put energy? We had a way of putting energy into the molecule, like stretching this bond, or stretching this bond, or bending some bonds, or torsion, or something like that. Now truly, we deal with quantum states. So you put energy into different quantum states and you count up the quantum states, to see how likely things are. Okay, so that's where the Boltzmann Factor is. That exponential, that e to the -- exponential ¾ ΔH comes just because that's what you expect. If you randomly distribute things, it'll come out that way.



Now how about the entropy factor? And this one is fun. Feynman, in his wonderful Lectures on Physics, says: "It is the change from an ordered arrangement to a disordered arrangement, which is the source of irreversibly." Have you heard this said, that entropy is disorder, and that you increase entropy in order to increase disorder? Okay, so that's what Feynman is saying here. "The change from an ordered arrangement to a disordered arrangement." Okay. Now here are two arrangements of the same number of dots. Which one is more ordered, left or right?



Students: Left.



Professor Michael McBride: Okay. You know I'm setting you up, right? So you're -- but I know what you would have voted for. Right? So I'm not going to ask you to vote. Okay? But look at the one on the right, from a different point of view. [Laughter] So what do you conclude? Which one is more ordered? The one on the right is just as ordered as the one on the left, but we didn't perceive the order. Okay? And that's like constellations; you know, the shepherds lay out and saw bears and dragons and things, in the sky -- right? -- and thought they were ordered. Okay? Now disorder, reversibility and Couette flow. Now I brought an experiment to do here, and it's -- but I'm not really sure it would work. So what I'm going to do -- because I didn't practice it before I came. I've done it before, but I didn't practice it today, and my pipette broke and I had to get a new one made before class. So if you want to see that, come after class and we'll try the actual experiment. But I'm going to show you a movie of it instead. Here. So this is the same thing here. What it is, is a -- well you'll see in the movie. I'll just start it up.



[Movie plays]



Professor Michael McBride: Okay, so it's a glass rod that goes up inside a glass cylinder. So there's like a doughnut inside, right? So I'm now going to pour Karo syrup in there; I brought Karo syrup with me to show you. So it's in that annulus between the rod and the cylinder. Okay? And now I'm going to take some yellow dye and put a strip of it -- first I'm going to mark it, so I can tell -- I'm going to rotate that outside cylinder, so you can see that it's rotating. And I'm putting a strip of yellow dye between the rod and the cylinder. Everybody see what I'm doing? And then I'm going to stir it up. And the way I'm going to stir it is by rotating the outside cylinder. Okay, so we'll zoom in and you'll see the watch, not very well, but to show that I'm not just running the movie backward or anything. Okay? Okay, now I start rotating the outside. So there's one rotation, two rotations, three rotations. So now it's all mixed up. And now watch. I'm unrotating. And it comes back. So if you want to see that happen, we'll try it after class here.



Student: Oh wow.



Professor Michael McBride: So you unmix things. That doesn't sound like entropy is working right. Okay? Now here's the way it actually happened. So there was the syrup, between the rod and the cylinder -- a look down from the top -- and we put a strip of ink in between, and then started rotating. And as we rotated, the ink spread out, like this. Right? Because the outer part moved, and the inner part didn't move, where it was in contact with the rod. So after I'd done three rotations, it looked like that. It wasn't really evenly mixed up. It's just that when we looked at it, it looked like it was mixed up -- right? -- when we looked through it. And now when we unrotate it, the whole thing -- nothing diffused and molecules didn't move at random. They just got spread out that way, but in a particular way. So it came back again. Okay? So the rotated state only seemed to be disordered. So that's the basis of the trick. Right? But that raises a very fundamental question. If disorder is in the mind of the beholder -- in this case, or in the case of that dinosaur, connect the dots -- if disorder is in the mind of the beholder, how can it measure a fundamental property, like entropy, if it depends on who's looking at it, to say whether it's disordered or not? Right? In fact, a disordered arrangement is an oxymoron, because arrangement is arrangement, and disordered is not arrangement. Right?



So how can you have a disordered arrangement, if the shepherd sees a dragon? Okay? The situation favored at equilibrium, by entropy, is one where particles have diffused every which-away, not into a coiled up piece of paper like the yellow thing, or not into a dinosaur. Every which-away; the key word is 'every'. That's what's statistical about it. A disordered arrangement is a code word for a collection of random distributions, whose individual structures are not obvious. So if a thing looks like, you know, a regular lattice like that, I say, "Ah ha, that's a regular lattice." But if it looks like this, I don't say it's exactly that; I say it's disordered, by which I mean I can't tell the difference between that one and this one, or this one, or this one. Right? So there are a whole bunch of those arrangements that I count together when I say 'disordered'. It's a collective word. So if all of them are equally likely, it's much more likely to have disordered -- many, many, many arrangements -- than the particular ordered ones that we're thinking about, even if they're all equally likely.



So that's the idea. It is favored at equilibrium because it includes so many individual distributions. So entropy is actually counting, in disguise. You count all these different arrangements, or all the different quantum levels, and the more you have, under a certain name, the higher the entropy associated with that name is. So, for example, a very common value of the entropy difference between two things is 1.377 entropy units. That seems a weird number, right? Now 1.377 happens to be R times the natural log of two. Now, consider the difference in entropy between gauche and anti-butane. Okay? So the equilibrium constant is e(-ΔG/RT). Do you remember what G is? That's the Gibbs Free Energy, which includes both heat, both the kind of things -- bond energy that we've been talking about -- and entropy is included in there too. So we can split it apart, into the part that has to do with heat -- or enthalpy, the ΔH between the two things, gauche and anti -- and TΔS, the part that has to do with entropy. I suspect you've seen this G=H+TS before; H-TS before. But let's just split it apart. Since they're in the exponent we can multiply two things together. So we have the first part, the one we've been dealing with, 3/4thsΔH. Right? And then we have the red part, that has to do with entropy. But you can simplify that. How can you simplify the part that has to do with entropy, right off the bat?



Student: Cancel the T's.



Professor Michael McBride: Cancel the T's. Okay, so it's actually ΔS/R. Now suppose that the value of ΔS is Rln2; which I said was a very common entropy difference. Right? Now you can simplify it further. Can you see how to simplify it further, for that particular entropy difference? Well obviously the R's cancel. And what's E raised to the power ln2?



Student: Two.



Professor Michael McBride: Two. So actually what that is, is our 3/4thsΔH times two. So when you see 1.377 entropy units, that's somebody who likes math telling you that's there a factor of two involved. That sounds more reasonable. Right? Two. Why should there be a factor of two, that favors gauche- over anti-butane? Yes?



Student: There are twice as many gauches.



Professor Michael McBride: There are twice as many gauches as there are anti's, because it can be right-handed or left-handed gauche. So you see what a crock this is, to say that the entropy difference between gauche- and anti-butane is 1.377 entropy units? It's just that there are twice as many of one as the other. Right? So that, the fact that ΔS occurs in an exponent, is just a complicated way of telling you that there's a statistical factor. You have to count how many of these things there are. Okay? Because you have two gauche butanes. So the conclusion; it just means a factor of two. And then that the equilibrium constant depends on temperature, because of ΔH, not because of ΔS. Often people think that because the free energy is H and TS, that therefore the entropy thing is changing as T changes. But in fact that's not true, because you divide by T to get anything out of it again. Right? So what really changes with temperature is the contribution due to ΔH. So sometimes that's just used to obscure what is fundamentally very simple.



Okay, we're going to stop here. And just so everybody's on the same page, we'll have the final lecture on Wednesday. But then I'll be here at class time on Friday too, and we can have a discussion then, to review for the exam. And I'm willing to have another one. I forget. When did I say? On Monday night. Now do people have -- there are not exams at night, are there; or are there? Does anybody have a -- is Monday night okay to have the review? It's probably the best time to have it, so you have a full day after that before the exam. So I'll get a room for next Monday night, a week from tonight, for a review session. But also I'll be here on Friday at lecture time. So we'll see you. If anybody wants to see this experiment, we'll do it.



[end of transcript]

Lecture 37
Potential Energy Surfaces, Transition State Theory and Reaction Mechanism
Play Video
Potential Energy Surfaces, Transition State Theory and Reaction Mechanism


After discussing the statistical basis of the law of mass action, the lecture turns to developing a framework for understanding reaction rates. A potential energy surface that associates energy with polyatomic geometry can be realized physically for a linear, triatomic system, but it is more practical to use collective energies for starting material, transition state, and product, together with Eyring theory, to predict rates. Free-radical chain halogenation provides examples of predicting reaction equilibria and rates from bond dissociation energies. The lecture concludes with a summary of the semester's topics from the perspective of physical-organic chemistry.




Transcript



December 10, 2008



Professor Michael McBride: Okay, so last time we were talking about exponents, and how statistics gives rise to them. And we looked last time at the Boltzmann factor, that e-H/RT or e-ΔH/RT, which tells you the temperature dependence of the equilibrium constant. It gives you the equilibrium constant; remember at room temperatures 3/4thsΔH in kilocalories. Right? And we saw last time that that comes from counting arrangements of a fixed number of energy bits. And the important thing is that they're random, that all of them are equally likely. [Technical adjustment] That has interesting philosophical implications I think. Do you remember at the start of the course we were talking about types of authority, and in particular, remember, in Hamlet, Act V, Scene II, he says, "There's a divinity that shapes our ends." So this was a traditional view in society. But then this guy came along, in 1859. Do you know who that is? Charles Darwin, with The Origin of the Species. So there was conflict between the religious fundamentalists and Darwinists that continues to this day.



But I think they really miss the point, because Darwin had the idea that things were developing and getting better all the time. Right? So there's a certain compatibility between that and the sensibilities of the fundamentalists. But also, not too long from then, this guy came along, Boltzmann, and he said everything is driven just by randomness. There's no goal for everything. That should be the real conflict, between religious fundamentalists and science, is Boltzmann. But so few people understand what Boltzmann did that there's no reason for the conflict. At any rate, we saw where the Boltzmann factor came from, in terms of statistics. Then there's the entropy factor, which we saw last time. And notice that, as I've written it here, in the top equation, the one with Boltzmann, there's an R, and the one at the bottom has a k. Sometimes you see R's and sometimes see k's. They're actually the same thing. The k is if you measure the energy per molecule, and R is if you measure the energy per mole. So all you have is Avogadro's number is included in one of them and not the other. They're really the same thing, these two exponentials. But at any rate, we can cancel the T's and know that that eS/k is a counting number, W; the number of molecular structures being grouped. And we saw that at the end of last time in terms of gauche- versus anti-butane. And remember that S = ln k [correction: S = ln W] -- which is the same equation, just the log on each side -- is on Boltzmann's tomb.



But let's look at one more example, in terms of cyclohexane. Do you remember the cyclohexane conorformers -- [technical adjustment] -- remember the cyclohexane conformers, if it's a chair, it's very rigid. Why is it rigid? Do you remember? Why is it hard to twist? With respect to the molecular model, why is it hard to twist? Eric? Why does this particular model, in which rotation around bonds is completely free, why is it hard to rotate to go like this? Why is there a click?



Student: [inaudible].



Professor Michael McBride: Pardon me?



Student: It has the optimal bending angles, 109.5.



Professor Michael McBride: Ah, it's the bending angle that has to flatten out when you do this. Remember we talked about that. But anyhow this one is very rigid. But if you click it, into the boat, then it's actually a twist-boat and it turns to all sorts of structures ever so easily. So there are many more accessible structures for a boat, than there are for a chair. The chair, you have this one and you have this one. Right? But for a boat, or a twisted, or the "flexible" form as it's often called, you have lots and lots of accessible structures. What's that going to mean in terms of entropy? That there are many more structures you're going to count for the flexible form than you count for the chair form, the rigid form. Okay? So if we look at -- this is a picture of the energy with the chair at the minimum of energy. The twist-boat is 5.5 kilocalories above, but it has very small barriers among different twist-boat forms. So we would say there are few structures that are chair cyclohexane, but many structures that are the boat. And furthermore, if we look at it in terms of quantum mechanics, which is the real way to do it, we notice that when there's a steep, stiff valley, there are very few energy levels that are accessible. Right? But if the barriers are low, then there are many, many quantum levels. So you count many more quantum levels for the flexible form than you do for the chair form.



What that means is from, whether the red classical view or the green quantum view, there's a big statistical factor, an entropy factor, which turns out to be about 7-fold favoring the twist-boat over the chair. The chair is still favored; it's favored because of energy, that Boltzmann factor, eΔH/kT. But this reduces that bias in favor of the chair -- which would otherwise be 103/4ths of 5.5; right? -- it reduces it by about a factor of 2000. So the equilibrium constant is only 14,000 -- instead of 14,000, it reduces by a factor of 7, to make it 2000. Okay, so that's -- but so we've been talking about entropy here in statistical terms. But remember earlier we talked about the fact that physical chemists actually measure entropy as something experimental, not just a counting exercise, and they do it beginning with something that's as close as possible to zero Kelvin. Because if you have a perfectly ordered crystal, at zero Kelvin, then there's only one structure you're talking about. Right? So the entropy is the log of one, which is zero, or log of 1/R; times R. Okay? So it's zero. But then as you warm it up, as you put heat in, you measure the increase in this thing called entropy by how much heat you put in, at each temperature, as you warm up. And we talked about that already. And you'll notice qualitatively how this works is that floppy molecules, like the flexible form of cyclohexane, that have many closely spaced energy levels, absorb more energy, and they absorb it at lower temperatures, and thus they have more entropy when they get warmed. So whether you do it experimentally, by measuring the heat that gets absorbed, or whether you measure it by counting quantum states, you get the same result; that a floppy molecule is favored in terms of entropy. We talked about that when we were talking about the barrier to ethane rotation.



Okay, so we've looked at the Boltzmann factor and this entropy factor, that come from counting. Notice that in the second case, truly we should be counting not structures but quantum states; how many states are there that count? And if you want to be really, really picky, you say the weighted number of quantum states; because ones that are higher in energy are not as populated as ones that are lower in energy. But we don't have to worry about that now. At any rate, those two factors, taken together, are what gives the equilibrium constant, which as we said before is e-ΔG/RT, free energy; that means it's both enthalpy, heat, and entropy, this counting thing. But there's more to it than that. There's also the statistics involved in the Law of Mass Action. And that comes from counting molecules per volume; that is, concentration. So you can count energy, you can count quantum states, and you can count molecules per volume. This is very easily understood.



The way it developed experimentally is that in the late 1700s there was an attempt to assemble a hierarchy of affinities; that some elements had more affinity than others. So if you had a compound that involved one element, and mixed it with an element that had higher affinity, that one with higher affinity would take away molecules from the one with lower affinity. But then it was found, in the early 1800s, that it doesn't always go in that direction, because concentrations can change and make things go the wrong way, toward something that's a very low affinity, if it has very high concentration. And by the middle 1800s, this was formulated as an equilibrium constant, K, which was understood as the balance between forward reactions and reverse reactions. You've seen that k1/k-1 is the equilibrium constant. So concentration comes into it, in this thing called the Law of Mass Action. If you have two A's that can go to A2, then you have this equilibrium constant, written as the concentration of A2 divided by the concentration of A2.



Now why the exponent? Where does that exponent two come from? And sometimes it's three. Remember? Again, it's statistical and very, very easy to see. Here I used Excel to place a bunch of circles randomly on a two-dimenstional plot. So it put 50 of these circles down. And we're going to count as dimers things where they overlap, where they touch one another. Okay? So there's -- if you have 50 particles, this particular realization of randomness gave one dimer. But suppose we add another 50. Right? Now, in addition to the original dimer, there are 8 more. Okay? And now suppose I add another 50, now there are 19 dimers. Or another 50; now there are 35 dimers. Or another 50, and now there are 59 dimers. Now let's make a plot of how many dimers there are, compared to how many particles there are, undimerized ones. So there's the number of particles and the number of dimers, plotted; the data there. And you see what it is. It's a parabola; that the number of dimers is proportional to the square of the number of monomers. Now why should that be so? It's because as you increase the concentration, you increase the number of units there; obviously. That's by definition. But you also increase the fraction of the units that have a near neighbor, that are touching. So not only do you increase the number that could be dimers, you increase the fraction that are dimers. So there are two ways in which increasing the concentration increases the dimer. Right? So it's a square. That's where the exponent comes in. It's purely statistical, for random distribution. And then, of course, energy can enter in, entropy can enter in. That will change the K. But the exponent on P -- right? -- is due just to this counting procedure. So you increase both the number and the fraction that are dimers.



Okay, so now we know about equilibrium, statistics and exponents. We have particle distribution, that's the law of mass action; that's that square we just talked about. We have energy distribution, ΔH, the Boltzmann factor. And we have counting of quantum states, which has to do with entropy. Okay, so free energy determines what equilibrium is, what can happen. Right? And that ΔG remember includes both entropy and energy. But it doesn't tell you at all how quickly it will happen, what the kinetics of the reaction are. So for that purpose we're going to try visualizing reactions, to see if we can get a picture of what determines rates. And we're going to look at two kinds of things. We're going to look at classical trajectories, and at the potential energy surface, and collective concepts.



So first let's start with the potential energy surface. Here's a very simple potential energy surface. It's just a two-dimensional diagram. So you've seen it before. We plot the distance between two atoms, A-B, and the potential energy as a function of the distance. So this is a Morse-type curve. Okay, so we put a ball on it, and let that ball roll to map out different distances, as it moves -- that's the horizontal axis -- and the energy corresponding to that distance. So it rolls back and forth -- right? -- mapping out the change in geometry horizontally, and the change in [potential] energy vertically. Now we're going to look at a three-dimensional surface, or actually two geometrical dimensions; plus the third dimension coming out at you, the contours, have to do with potential energy, of three particles now. But in order to describe the position of three particles, you need three distances; one to two, two to three, and three to one. Right? So that's too many dimensions to plot this way. So we're going to assume that they're all on a line. So we only have the distance between one and two, and the distance between two and three. So two coordinates will tell you what the arrangement of those three atoms are, if they're in just one dimension.



Okay, so a linear triatomic, A-B-C. On the horizontal distance we have the distance A to B, on the vertical axis, B to C, and the contours tell us how low the energy is; the darker, the lower the energy. Okay, so this specifies the structure. So the position of a point specifies not the position of an atom but the position of a set of atoms. Right? So as we move one point around, we're moving two atoms, or changing two distances. But we can denote it with one point here, in two geometric dimensions. Okay, so let's look at several regions. Let's look at that valley on the right. What is that valley? When a point is in that valley, what's it describing? It says that the distance between A and B is large. Right? So A is far away from B. What does it say about the BC distance? It's short. It's the normal distance for a BC molecule. So what is it; when a point is in that valley, what geometry is it describing? It's describing a BC molecule, with an A atom at a great distance from it, depending on how far out we go to the right. Does everybody see that? One position of the point tells us where both -- where all three atoms are, as long as they're on a line. Right? Now how about up there, on that plateau? What's happening there? What's the physical system? Maria, what do you say?



Student: All the atoms are separate from each other.



Professor Michael McBride: Yeah, the atoms are all separate from one another. C is far from B and A is far from B. So there the atoms are separated. And it's high in energy. You're not surprised that it takes energy to pull them apart. And there is a cliff. Right? If this were a hiking map, that would be a cliff. What does that cliff mean? Why is there a cliff there? Greg, what do you say? Why is there a cliff shown there? What's the geometry that points in that region described?



Student: Closer together.



Professor Michael McBride: What's close together? Is A close to B?



Student: Closer.



Professor Michael McBride: Is A close to B? I couldn't hear.



Student: Closer maybe to B.



Professor Michael McBride: No. Here's the distance A-B. It's way out there, where that cliff is. Right? So A's far away. But what's happening? B and C are getting very close to one another. So that's a collision between B and C; much shorter than their bond distance. Right? They're being scrunched together in that region, and that's why the energy goes up very rapidly, because they're running into one another; closer than they like to be. There's a ridge -- right? -- as you go down there. If you were hiking, that would be a ridge. And there's a special point on the ridge, which is marked by the red cross. What would you call that if you were hiking? If you were in the valley that's labeled yellow, and you were thinking of getting to the other valley, what would you call that cross, if you were hiking?



Student: A peak.



Professor Michael McBride: Not a peak. Right? It comes -- it gets high as it gets close to me here, and it gets high out on the plateau. It's actually low, as we go along the line, but it's high as we go perpendicular to the line. What would you call that if you were hiking? Some of you must hike.



Student: A pass.



Professor Michael McBride: Pardon me?



Student: A pass.



Professor Michael McBride: It's a pass, between one valley and the other valley. That's the way you would go if you were hiking, right? So you don't have to climb so high. Okay, so it's a pass. Or in chemistry we call it a transition state, or a transition structure. And it's like a potato chip. Right? It's a minimum in one direction, along the ridge, but it's a maximum in the path between the two valleys. Okay? So it's a very special point.



Okay, now let's look at this a little closer. Suppose we sliced this surface, we took a knife and sliced it along the red line there, horizontally, and then folded it back to look at the cross-section. Does everybody see what I'm saying? What would the cross-section look like? It's exactly that one we looked at before. So that involves stretching A and B, with C so far away that it's irrelevant. Right? So it's just stretching a diatomic molecule. Or if we sliced -- so that's vibration of AB, with a distant spectator of C. Okay? Now suppose we sliced it that way and pulled it back. Now what is it? It's a vibrating BC molecule, with A far away. Right? So again, by choosing points on this, we can describe any geometry of these three points we want, as long as they're on a line. But a single point is a whole structure.



Okay, now let's look at a trajectory. We roll a ball on this surface, and it rolls up toward the pass but doesn't make it and comes back like that. So that's a trajectory that was unreactive; it didn't make it into the product valley, it didn't get across the ridge. And physically what it means is that A gets closer to BC. While BC is vibrating, it's moving up and down. Right? So BC is vibrating. A comes in and then bounces off again. Right? So it was an unsuccessful attempt of A to take B away from C. Or suppose we do this one. A approaches, and now C flies away from vibrating AB. Does everybody see that? You want me to say it again? Right? So as it comes along, B and C are standing still. It's not moving vertically, right? So B and C are at their normal distance, their standard distance. A comes in, whops into it, gets really close, runs into that cliff, really close to B, because it's the distance A-B. Right? And then it starts -- AB is vibrating, and C is getting far away, as it goes down the product valley. So that's a reactive trajectory, where the first trajectory was unreactive. Notice that this is classical, rolling a marble on the thing, because A was approaching a non-vibrating BC. We know from quantum mechanics that BC can't just sit there. BC vibrates. So that's not a realistic trajectory, it's a classical model of a quantum mechanical system.



Now people have made surfaces like that for real systems, like H3, one hydrogen atom attacking a hydrogen molecule and taking one of the hydrogens away. Here it is. This was drawn. This is the first surface that was drawn, that I know of, by Henry Eyring at Princeton in 1935. And it had some -- it has crazy angles. And the reason it has that angle is so that a marble rolling on it will behave according to the way the masses of the thing -- so that the kinetics, the kinematics, are exactly right, so that it really does do the proper classical thing. If you roll a marble on here, it will trace Newton's Laws of Motion, for this H3 system. One thing that's unrealistic about it is that there's a minimum near the transition state. There's a little lake up there. That's just an artifact that came from the equation he used to approximate the surface. It's not real. There actually is a potato chip, not a lake up there at the pass. Some people have called it Lake Eyring. Okay?



But here's a more complicated surface that was -- it says it was done by C. Parr, in his Ph.D. thesis at Caltech in 1969. This is for H-H-Br; again linear, but a hydrogen atom attacking HBr to take away it's hydrogen. And this describes the construction of a model based on that surface, an actual physical model on which you can roll marbles. Right? So here it is, and the dimensions, that one inch corresponds to 1.4 inches is an angstrom, and so on. So they actually constructed physical models to do this. And, as it happens, we have that physical model here. Here, this is it. [Technical adjustment]



Okay, and we have some marbles. So notice what's different about this one. I'll hold it up so you can see it. There's the surface. Right? And here's a slice through one. Right? That's H2 with -- this valley is H2 with a Br atom. This valley is HBr, with an H atom. Right? So we can try rolling balls and see what happens here. So here I am far away. The H atom is far away from HBr. And I do that. And it's vibrating, but it doesn't succeed. Right? I can do it a little bit more. And most reactions aren't successful. Oh close. Okay? And as it comes out, it may be vibrating when it comes out, or it may not be vibrating when it comes -- oh.



What's the cross? The cross is that transition structure, right? -- or the transition state. Right? So you could do every possible trajectory. You could have different velocities at which A is colliding with BC, and you can have different amounts of vibration, and different phases of vibration. I could start it going this way, or I could start it going this way. And if I did every possible one of those, then I could average over all them and see how many of them succeeded, right? In reaction. And that would allow me to predict the rate of this process. [Technical adjustment]



Okay, so people have done that kind of thing. But studying lots of random trajectories provides too much detail. No one really cares about all those individual trajectories. What you care about is what fraction succeed. And if you could do that in a simpler way than rolling it a zillion times and averaging, then that would be a better approach. So it's better to summarize this thing statistically using collective terms; not individual paths but collective terms, such as enthalpy and entropy. And you can do it in this way. Okay, you have a steepest descent path from the pass. Suppose you slice this surface with a knife perpendicular to the screen, along that path, and then fold it out to make it flat, and look at the cross-section of it. Okay, and now we'll tip it up and it looks like that. Does everybody see how it looks like that?



Okay, we sliced along the path that goes up over the transition state. Okay, so we have potential energy as a function of distance along the reaction coordinate. But of course we're going to summarize a whole bunch of things here. Nothing rolls exactly along that path. Right? So we're going to lose some of the specificity of the reaction coordinate by grouping things. We're not going to take a trajectory, but a sequence of three species. The first species is the starting material; the second species is the transition state; and the third species is the product. And now instead of having an explicit meaning for every geometry along that path, what we're going to do is associate each of those starting materials, transition state and products, with a certain enthalpy and a certain entropy; that is, a certain free energy. So we have -- instead we're going to plot free energy vertically, incorporate entropy -- that is, how loose these things are -- as well as enthalpy there; energy. And now we just have three values of free energy: the free energy of the starting material; the free energy of the transition state; the free energy of the product.



And now we can use those -- but we've lost a lot of specificity in what the reaction coordinate is. It's just a sequence now of three species. But we can now use free energy to determine what can happen. We already could do that. But how rapidly? And you do that with the theory that Eyring made up when he drew that surface in 1935, called Transition State Theory. You assume that the rate constant -- and we talked about this before -- in units per second is 1013th. So things happen 1013th/second, times the concentration of the transition state. So the idea is that things that are in the transition state are moving at the rate 1013th/second. So if you know how many are at the transition state, then you know how many are going to product; 1013th times that, per second. Right? And you know how to do an equilibrium; we already do that. All we do is use that special double dagger, to mean the difference in energy going from the starting material to the transition state, the difference in free energy.



Now we can calculate that equilibrium constant. We know how much starting material there is; the Law of Mass Action. We know how much transition state there is in equilibrium with that starting material. Right? So that e-ΔG/RT is the equilibrium constant for getting the transition state; multiply it by the concentration of the starting material. You have the concentration of the transition state. Multiply it by 1013th, and you know how fast it's going. Right? So it assumes that there's a universal rate constant for transition states going to the product; 1013th/second. That's about how fast things vibrate. So things won't stay on the potato chip, they'll roll off, and the rate at which they roll off is 1013th/second. This is, of course, an approximation. It isn't correct. Because there's not true equilibrium between the transition state and the starting material. For true equilibrium the same number must be going in and out. Right? But when things get to the transition state, often they keep going, they don't come back. So it's not really a rigorous theory, but it's a very, very helpful theory for approximate purposes. And that's what I say here. Okay?



So using energies to predict equilibria and rates for one-step reactions; free radical halogenations. And let me look, we do have time to go through this stuff, I think; at least most of it. We've already seen this, that you break a chlorine molecule into two atoms. Then you do single electron curved arrows; take the hydrogen away from methyl -- from methane, to give methyl, and then it attacks chlorine -- we've seen this before -- and the chlorine atom comes back to constitute a chain reaction. Okay? So it's catalytic in radicals. Okay, now we break and make bonds in this. So we've already seen average bond energies that might tell us how hard it is to break a bond -- right? -- if that's what we have to do to get to the transition state. So we may be on the way to getting rates. But are these average bond energies real bond energies, or are they just a trick for reckoning the enthalpy, the total energy, the heat of formation or whatever, of a particular molecule? And we talked about this before and saw that mostly it is a trick; that individual bond energies change. Right?



So average bond energies are not what you want to use for this purpose; although they might be in the ballpark of the right numbers, but they're not the right numbers. What you really need are bond dissociation energies; the actual energy it takes to break a specific bond. And we've talked about how you could get that from spectroscopy. Those are real. The average bond energies are just a way of calculating molecular energy. So Appendix II of the Streitwieser and Heathcock book shows various specific bonds between the groups in column A and the atoms or groups in row A there, in the top row. And those were the best values when that book was published; the best values as of 2003 are a little bit changed from that, and were tabulated by Barney Ellison, whose picture I showed you the time before last.



So here's a table from Blanksby and Ellison, that shows a bunch of these things and how well they're known. The H2 bond dissociation energy is 104.206, plus or minus 0.003 kilocalories/mol. So it's very well-known. So some of them are very, very well-known, and some of them are known only approximately. Let's look at a few of them. So H2 104.2; HF 136; HCl 103; HBr 87; HI 71. So as you go down the halogens, the bonds get weaker and weaker. Why should that be? Why do larger halogens give weaker bonds? It's because they have poorer overlap with hydrogen, at normal bond distances, because their orbitals are very diffuse; they don't have high numbers in the region where the hydrogen is. And you have the energy match also is unfavorable. So if you have HF, with good overlap and a very low F, then you get lots of stabilization; the sum or those two red arrows is big, a strong bond. And if it's HI, the overlap isn't so good, and the energy match is better. But still the amount by which the electrons go down is not as much as with HF. So we can see qualitatively why that is. So less electron stabilization means a weaker bond. And we're going to talk about this more next semester. This is just to give you a flavor of it.



And here, in Table II, we see there's the same trend in bonding with methyl groups; that the fluorine-methyl bond is 115, but the iodine-methyl bond is only 58. But if you go across hydrogen with the different alkyls -- methyl, ethyl, isopropyl, t-butyl -- they're all very close to 100 kilocalories/mol. But not the same, and that will turn out to make a difference; as we'll see early next semester. And if you look at some of the other kinds of radicals, you see there are differences. There are special cases for vinyl, allyl -- remember allyl alcohol -- phenyl and benzyl. And I'll show you just a little bit about that. Are these unusual bond dissociation energy values due to unusual bonds or unusual radicals? That is, standard we have: here's a starting material; here's the bond; here's the product as the bond's broken -- for making the radicals. So we can get a strong bond, either by lowering -- by making the bond strong, or by making the radicals bad; either way increases the barrier. Right? Similarly we can make a small one, either by making the bond weak or making the radical stable. Is this something that it's meaningful to talk about?



Well let's look at it in the case of these special cases. So vinyl, you see, is 110 kilocalories/mol. It's the bond to H. So it's a very strong bond. Why? Is there something special about the vinyl radical? There's no special stabilization. There's no overlap between that singly occupied orbital, which is a σ orbital, and the double bond, which is a π orbital. There's nothing especially stable about the vinyl radical. But if you look at the bond in the starting material, the C-H bond, it has sp2 hybridization of the carbon. So that one's a strong bond. That one's hard to break, because the bond is unusually strong; not because the radicals are unusually unstable. Okay? Or if you look at phenyl, it's the same deal. You have a σ SOMO and the π bonds, the low LUMOs that might stabilize that singly occupied orbital's electron, are perpendicular or orthogonal to it. So again it's hard to break that bond. On the other hand, the allyl -- and remember allyl alcohol we were talking about -- here's an allyl radical. Now there's overlap, π overlap between the SOMO and the π* and the π. It turns out -- and we'll talk about this more next time -- that when you mix those two, the π* on the right and the SOMO in the middle, you get stabilization. You get the same stabilization by mixing the π with the SOMO, and if you mix both of them with the SOMO, you get that structure. But this isn't the time to talk about that. But at any rate, the normal, the starting bond, is normal. It's an sp3 C-H bond, a normal bond. But in this case the radical is unusually stable. So it's easy to break that one; 10 kilocalories easier than for normal C-H bonds. And the same for benzyl.



Okay, now possibility of doing a halogenation. Let's look first from the point of view of the equilibrium constant from this, at the difference in energy between starting material and product. So we look at what bonds are broken, and what bonds are formed. The red bonds are broken, the green bonds are formed. So there'll be a cost for breaking the red bonds and there'll be a return for making the green bonds. And let's see how big it is. Okay, we're going to do it for fluorine, chlorine, bromine and iodine. Okay, in every case we're breaking a C-H bond, 105 kilocalories/mol, and we're breaking a halogen-halogen bond. But those are different, for the different halogens; although interestingly not monotonic. It's not a smooth progression, it goes up and then down again. And the cost is the sum of those two; what it's going to cost to break those bonds.



How about making bonds? We're going to make the C-X bond and we're going to make the H-X bond. So the returns will be that. And now we see whether we can make a living doing this. What will the profit be? Right? So in the case of fluorine, 251 is much greater than the cost of 142. So it's a big -- it's really exothermic. Right? Chlorine, it's only 19. Bromine is only 9, and iodine is minus 12. What does that mean? It means you can't do this reaction with iodine. The equilibrium lies is in the wrong direction. Okay, so already the equilibrium constant and how energy relates to it tells you that something is impossible to do. But the others seem to be possible. But will they happen? How fast will they be? Do you have to wait until the end of the universe in order for this to happen?



Well let's look at how the rate, which will depend on the mechanism. The equilibrium constant doesn't depend on the path. It's just how high the starting material is and how high the product is. But how fast you get depends on the barriers you have to go across. So let's suppose -- let's just try breaking two bonds, changing partners, and forming two bonds. Okay? Then we have to -- to get to the barrier, we have to break two bonds. So that cost -- we're going to have to spend, in order to get to the barrier. So is this a plausible mechanism? How fast would the reaction be, if we used that mechanism? Well at room temperature, say 300 Kelvin, then it's 103/4ths of how high it is to get there; then times 1013th/second. But notice that 1013th times 10-106, which is 3/4ths of 142, is 10-93/second. 1093 seconds is a long, long time. I don't know how old the universe is, but I suspect that's longer. Right? So what about this mechanism? Plausible? Implausible? How could you make it happen? Shai?



Student: Change the temperature.



Professor Michael McBride: Pardon me?



Student: Change the temperature.



Professor Michael McBride: Change the temperature; because we can change that 3/4ths to something else by -- to a much smaller number -- by increasing the temperature. So no way to do that one. But if we go to 3000 Kelvin, then it's 250 reactions per second. So that's quite feasible, as long as nothing else happens; something else might be even faster. Right? But at least this mechanism could work, if you were in a flame say, or in Professor Chupka's oven, that we talked about last time.



Okay, now let's look at Eyring's H + H2 here. So we want to get from this valley to the other valley. How do we do it? This mechanism would be to break a bond and then -- so dissociation and then association. But uh, it's very slow to get up there, right? -- to get up to that plateau of breaking a bond. There's a much easier way to get from one valley to the other. What is it? Go through the pass. Right? So instead of doing that slow reaction, you can do this much faster reaction, making a new bond as you break the old one. So you don't have to pay all the cost of breaking the old one, and you're getting something back for it. So you can have a free radical chain substitution, where you have an X atom -- and we've talked about this before. Take the H away from R. Then the R group takes away X from X2. And it can just go round and round.



There's this machine, very much like the machine that Professor Sharpless talked about, that goes around and around. You feed in starting material, and the products come out. And you don't -- it's a catalyst; you don't have to go to such high energies as you normally would have to do. So the possibility of halogenation -- here we looked at it at equilibrium and saw that forget it with iodine; the others look okay. And for a mechanism with a reasonable rate, we could change those two columns, exchange them, so that we make the H-X, as we break CH3-H -- so the X atom helps you do that -- and that generates then a CH3 group, which helps the reaction on the right. So now, step one, how much energy do you have to put in? You're paying 105 but you're getting back 136. So that's great. Right? The first step seems plausible now; although you don't know how high the barrier is that you have to get from starting material to product. But at least it's much easier.



So the top one, it looks good all around. The chlorine and bromine will be a little touchy, on those first two steps, because they're uphill in energy. But step two is good in all cases. Okay, so what we need to do is be able to predict activation energy. But Rome wasn't built in a day, and that's going to happen next semester, to get into this and to talk in more detail about these catalytic reactions. Notice that this is exactly the kind of thing Sharpless was talking about. If you could find a way of lowering one of these barriers, then the catalytic cycle would work. He was saying, remember, that it was democratic; that all the steps, as you go around the cycle, have to be at the same -- go at the same rate, or else everything stops up, stops and waits before it tries to get over some big step. So you have to have all the steps be fast; and that's why he was saying that that diisopropylethylamine helped the formation of omeprazole.



Well so we've been doing organic chemistry this semester. Some people might doubt that. There'll be no doubt next semester that everyone will agree that what we do is organic chemistry. We're going to talk about the chemistry of functional groups and sugars and amino acids and all carbonyl groups and esters; all these things. But there's a reason that it's been a little different this semester. We've talked mostly about physical-organic chemistry. And the reason is because of where I came from; that I did my Ph.D. -- and in fact I took a course as an undergraduate with Paul Bartlett at Harvard. And as you've seen from our common ancestry, he was a physical-organic chemist; more interested in how reactions work than in what you can make. Okay, so there's this book that came out in 1939, went through at least three editions -- I have the Second and Third Editions here -- which is called The Nature of the Chemical Bond. And this was a fabulously influential book; even if it is the one that said it was 126 kilocalories/mol, to take carbon away from graphite, when we know it's 171. So Pauling was arguably the most influential chemist of the Twentieth Century. He got two Nobel Prizes.



[A tape is played]



Recording of Professor Jack Dunitz: At the time when I was reading that book I was wondering whether chemistry was really as interesting as I had hoped it was going to be. And I think I was almost ready to give it up and do something else. I didn't care very much for this chemistry which was full of facts and recipes and very little thought in it, very little intellectual structure. And Pauling's book gave me a glimpse of what the future of chemistry was going to be and particularly, perhaps, my future.



Professor Michael McBride: Right, okay. So that's what we've been doing the first semester, is this kind of thing, thinking about the chemical bond. We started with wondering, with Newton, whether there was an atomic force law, and then we looked to see if we could see bonds, or feel them. And then we tried to understand bonding and reactivity, through the Schrödinger equation; which most people believe is the fundamental way to understand things properly. And then we learned how chemists learned to treasure the molecular model. Things like this, that turned out to be really, really useful tools -- composition, constitution, configuration, and conformation -- and finally energy. Right?



So I hope that we've this semester, if we haven't done as much organic chemistry as some people would care for, that at least we've raised some big questions; like how does science know things? Right? Or, compared to what? Those are the really big questions. But even if we're only focusing on the specific content of the course, interested in bonds, here are two questions for you to think about. Were chemical bonds discovered or were they invented? To what extent are chemical bonds real? Or to what extent are they a figment of a chemist's imagination? This is the kind of thing we've been aiming at all semester. So you should be in a position to think about this now. Or would we even have chemical bonds, without our own chemical forbearers? So we've looked at how the idea of bonds and these different properties of bonds developed. But suppose chemistry was developing in some other solar system. Right? What would happen there? Suppose people discovered the Schrödinger equation, before they had the idea of bonds. Would bonds have been necessary? Is chemistry going to evolve, such that all you need do is put stuff into a computer, solve quantum mechanics, and forget about bonds? Okay? The answers to these questions aren't obvious, but they're good ones to think about. So this is the end, and good luck on the final. Are you fired up?



Students: Yeah.



Professor Michael McBride: Are you ready to go? Good. Thanks.



[end of transcript]