Introduction to Chemical Biology 128. Lecture 03. Reactivity and Arrow Pushing.

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments
Captions
[ Silence ] >> Welcome to week two of Chemistry 128: Introduction to Chemical Biology. I'm Professor Weiss. I'll be talking to you today about reactivity. OK. So last week we talked about the molecules that compose your cells. And our goal this week is to understand how those molecules interact with each other. There are two forms of this interaction. The first kind is that the molecules can decide to react with each other. They can start to form covalent bonds. Bonds can break. Bonds can form. So, we want to understand this property that we're going to call reactivity. And to understand this, we're going to look at arrows and the language of arrows which organic chemists have developed as a way of communicating this reactivity. I have to tell you, I think that this is one of the great achievements of organic chemistry. This is one of those accomplishments that all humans can be proud of because it reduces something that otherwise seems mysterious to a simple set of rules from which you can derive many, many reactions, essentially all reactions found on our planet. And to me, that's really exciting because that means that this language is universal and it's one that's very broadly applicable. And so, that's my bias going in is that I think this is really cool. OK. So we're going to have quick review of arrow pushing. And then I'm going to show you examples of applying this language of arrow pushing and this language of reactivity in chemistry to the chemistry that's found on our planet before life started. This is a type of chemistry called pre-biotic chemistry. Now, obviously there were no humans present to observe directly what was going on. However, we can infer what was going on in this pre-biotic period. And this is an artist conception of what the planet might have looked like for both fossil record and also from experiments that have tend to recreate the conditions that were found during that pre-biotic period, OK? So, we're going to be using what we've learned to really to look at syntheses of the molecules that compose a cell. And then the next topic we'll talk about this week is making molecules using a combinatorial approach. This is essential in chemical biology. This combinatorial approach takes place in your cells. It's one of the reasons why your immune system can very rapidly respond to foreign invaders. And it also is used in many chemical biology laboratories around the world. And so, for this reason, I have to introduce this concept of combinatorial chemistry and combinatorial biology to you this week. And then finally, we'll look at the second mode of molecules interacting with each other. Recall at the start of this, I said there were two modes. The first mode being reactivity that forms covalent reactions that results in covalent changes, bonds forming and breaking. The second mode is non-covalent interactions. This is when two molecules slide along side each other and decide to form a complex with each other. And the rules that determine whether or not this complex forms are also rules that we can understand. And importantly, this is also a really tough frontier for chemical biology. So, while I'll be able to tell you about the rules for reactivity and covalent bond breaking and bond forming reactions, I cannot reply with such certainty where we start talking about non-covalent interactions. There's a lot less that we understand and that makes it one of the challenges. But at the same time, it also makes it really exciting because that means there's opportunities for people like yourself to get out and do new experiments, to start to elucidate those types of rules. OK. So, I have some announcements before we go in that's kind of the overview. Let's zoom down and look at the particulars. First, for this week, I'd like you to read chapter two in the textbook, that's this book here. Now, there's occasional times where the treatment in the textbook is more advanced than what I'm talking to you about. For example, there's information about inversion of phosphate geometry-- phosphorus geometry. I'm not going to discuss that. And if I don't discuss it in the lecture, then don't get too hung up on it in the book, OK? So, simply skim the concepts that are not presented in the lecture, OK? So if I don't talk about in lecture, it's not important for the class in terms of our exams and what I'll be testing on. Simply skim through it. Homework. I'd like you to work chapter two problems, in particular, every odd problem. And there will be a worksheet to guide our discussions this week which will be posted to the class website. In addition, there will be one handout this week which will be posted to the course website. Please download this on Tuesday to skim through it. This handout is an example of a journal article report which I used to call a book report. And then on Thursday, I'll discuss it with you on further detail, OK? At this point, I would usually ask if you have any questions. If you do have questions then you can either e-mail me or the TAs. OK. So, let's review where we've been and then we'll get started on what I told you about at the beginning as our big picture. So, we want to understand the function of human cells at the level of atoms and bonds. This is the smallest unit that actually is meaningful to ask as chemists. And as I described to you last week, cells are bags of molecules. They are bags that are chock-full of molecules. The molecules are stuffed inside the cells. There is no elbow room. These things are jam-packed into the cell. So, because of that, we expect lots and lots of interactions which is our topic for this week. But I'm getting ahead of myself. Let me continue reviewing what we talked about in the previous week. First, we talked about the composition of a gene as an on/off switch with instructions. We talked about how molecules are synthesized in the cell using the template of DNA to a messenger RNA which then is translated into proteins, and then proteins and RNA carry out all the various instructions that are articulated by the DNA. We also discussed six types of organisms. But in this class, we're going to be generally talking about either bacteria or human cells, OK? And it turns out there's a lot of chemistry in just bacteria or human cells. So our goal in this week is to reduce the complexity of diagrams like this down to a few rules that chemists like ourselves can understand. OK. So, let's get started with what is life? What is this stuff? What are the molecules that compose cells? What are the rules that govern them? In 1948, the physicist, Erwin Schrodinger, wrote a very influential book called "What is Life?" I highly recommend this book to you. It's a slim little volume and it's a fun read. It's not particularly challenging. But the concepts that he presents to me are really earth-shattering. These are paradigm changing. What Schrodinger argued is that the molecules that govern your cells, that allow organisms like, you know, yeast and bacteria in humans that allow organisms to live, those molecules are governed by physical laws, by the same laws that we talked about in chemistry and physics classes. There was nothing special or unique about the molecules found in living organisms. They are simply molecules that are governed again by physical laws. So, this persuaded-- this book persuaded a generation of physicists to explore biology after World War II. This was an amazingly influential book. And this persuaded this generation that include great scientists like Francis Crick and Jim Watson and many others to explore biology and to do this by applying concepts from physics and concepts from chemistry. And the results are more or less what I presented to you last week when we talked a little bit about the structure of molecules. So, this is good news for us. The good news is everything that you've been learning about in chemistry classes before now applies to biology. There's nothing special about biology. There's no sort of life force that animates molecules found inside the cell. No. Rather, the same rules that you learned about in general chemistry, that you learned in organic chemistry, those apply to the molecules found inside your cell. OK. So, let's talk a little a bit about those molecules found inside your cells. So, our goal is to understand first the reactivity of those molecules. And then second, we'll talk about their non-covalent interactions. So, covalent interactions, reactivity first. In organic chemistry, you learned the powerful language of arrows which are a way of depicting the overlap of molecular orbitals. Let me remind you of some conventions of those arrows and the conventions of this language of organic chemistry. So, the first to this is that these arrows depict the overlap of molecular orbitals so that they show for example electrons in a highest occupied molecular orbital overlapping with the unoccupied lowest energy, molecular orbital of the second reactant of the reaction. OK. So, in this basic reaction we have an amine, we have a ketone and the two of these are going to be reacting with each other. So, if you take amine and you take ketone and you mix them together, we can predict in advance that our reaction will take place. And here's why. What we can predict is that the lone pair on the nitrogen is going to be highly reactive. Why is that? What's special about that lone pair? It happens to be very high in energy. It is a highest occupied molecular orbital. And it's going to want to react with the pi bond, the-- this carbonyl functionality of the ketone. What's so special about the carbonyl functionality of ketone? Well, it happens to be-- it happens to have a low energy unoccupied molecular orbital, OK? Now, let's break down what these molecular orbitals actually look like. What this looks like is the lone pair of this nitrogen on this nitrogen is found in an N orbital, so it's in a high energy state. It's the highest occupied molecular orbital, the HOMO. And it's going to be overlapping with the lowest energy molecular orbital of the carbonyl of the ketone which happens to be the antibonding orbital of the pi bond. OK. This is what it looks like in terms of molecular orbitals, and this is what it looks like on top in terms of organic chemistry and organic chemists speak. Good news, we organic chemists have agreed to the convention that we will depict complicated reactions like this one using the simplified descriptor, OK? And this is good news. I don't think anyone wants to spend lots of time on the test deriving what these molecular orbitals look like and trying to describe an antibonding orbital in terms of the lobes and so on and so forth, it would be just way too complicated. So, we're going to be using this description here. Now, the real challenge for us comes from the fact that the molecules that we talked about in biology often times have multiple functional groups. It's not a typical for a biomolecule to have say hundreds if not thousands of carbonyls or to have thousands upon thousands of different lone pairs. So the real challenge is for us to figure out which of those lone pairs and which of these carbonyls is actually going to engage in a reaction. And when that happens, we're going to fall back on orbitals to decide which of these is going to be most reactive. OK. So again, what we're going to be talking about is this overlap of molecular orbitals. That overlap of molecular orbitals, the filled-unfilled overlap leads to the formation of new bonds and consequence breakage of others, OK? So, when this lone pair overlaps, so if the antibonding orbital of the carbonyl, the pi-star orbital of the carbonyl, the result is a new covalent bond directed by this first arrow. Now, on the other hand, we know that this carbon can't have more than five bonds to it or can't have more than four bonds to it and so, five bonds would be disallowed. And so, for this reasons in consort with this formation of a new bond, there's breakage of the pi bond between carbon and oxygen of this ketone. This is good news, right? This totally makes sense because what we're doing is we're populating this antibonding orbital. And in doing so, we're making the orbit-- we're making that pi bond break, right? If you put electrons into an antibonding orbital, what does it do? The bond breaks hence the name antibonding, OK? So, this overlap, to me, it's kind of like the peanut butter and jelly of organic chemistry. We're always going to be talking about a HOMO, a highest occupied molecular orbital overlapping with the lowest unoccupied molecular orbital. And in the same way that peanut butter and jelly taste so good together, orbital overlap works so well. It is so complimentary in terms of reactivity. OK. So, let's get back to our challenge again. The challenge is in biology, we often times have many different puzzle of reactivities, we often times had many different possible reaction mechanisms that we can draw. Despite that plethora of possibilities, what we will see is that there's often times one and only one true mechanism, dominant mechanism for a particular set of molecules. And again, this is good news, OK? So, for example, let me show you sort of an easy case where we're going to be looking at reactions, two possible mechanisms. One that makes chemical sense and one that does not. And so, by dong this, we can start to eliminate a lot of different examples. OK. So, here is a clash of two possible wills. In this reaction, one possible mechanism has the lone pair attacking the antibonding orbital of the carbonyl and going through the transition state that's depicted down here. This reaction is an addition elimination reaction, it goes through this transition state in addition, and then in the elimination reaction, the chloride is eliminated giving us a substitution of nucleophile in place of chlorine, OK? Makes sense, fundamental reaction. A different type of reaction mechanism might look like this, where the nucleophile directly displaces the chloride. In doing so, the nucleophile, lone pair on the nucleophile is populating the sigma-star antibonding orbital of the bond between the carbon and the chlorine, OK? So, two possible mechanisms. One involves the pi-star orbitals, this one involves the sigma-star orbitals. And I guess I first blushed these two reaction mechanisms might both look totally legitimate and both equally valid. The problem is they aren't, that we can actually readily eliminate the reaction mechanism on the right that-- that is governed by the SN2 reaction. Instead, what we can do is actually very quickly decide that only the addition elimination reaction will work. So, returning to this possible-- this clash of two wills, when we look at a transition state or reaction coordinate diagram for the two possibilities which I think tells us which possibility is correct and which one is wrong. OK. So this transition-- this reaction coordinate diagram is depicted over here. So, in one reaction I showed on the previous slide, the mechanism is an SN2 reaction and on the right, this is the addition elimination reaction, OK? So, in this I acknowledge, this is a complicated diagram, bear with me. So, over here, these are the starting materials. This is the acid chloride. This is the nucleophile. And again, if this reacts through an SN2 reaction, you will get this left reaction coordinate. And if for reacts through an addition elimination reaction, you get the right coordinate. Now, two possibilities, small little hill, big hill. Which of these two is preferred? Small hill, big bill? All right. Now, let's just imagine, you're in electron, you have to decide which one would you prefer. Would you prefer tramping up the very, you know, steep key slope, or would you prefer the much shorter hill? OK. I will tell you also that electrons are lazy that they do not expend any extra energy than they need. And in doing so, they are going to prefer very strongly the tiny little hill or the much smaller hill of the additional elimination reaction to the SN2 reaction, OK? This makes sense. That's the way electrons live their lives. So, what this tells us is that yes, there are two possible reaction mechanisms for this reaction, yet only one is actually correct. The only one that's correct is this one on the right. The addition elimination reaction, one on the left has to go through a much higher energy SN2 reaction. OK. Now, I'm going to explain in greater detail in a moment why it is that the one on the right is preferred than the one on the left, OK? To understand that, I need to tell you about three possible components of orbital overlap. So, the energy in this interaction is proportional to three components, OK? And let me go back. Recall over here that on reaction coordinate diagrams, the Y axis depicts energy where a higher number up here indicates higher in energy. And again, electrons being lazy prefer lower energy, OK? So, that again is why the smaller hill is preferred to the bigger hill in terms of which side to go on, left side or right side. OK. Now, this energy is proportional to three components. Component number one are charge-charge interactions, OK? So, if these molecules happen to have plus charges and minus charges then, that will have some interactions, some Coulombic interaction. In addition, if the molecules have a repulsive interaction with each other, that will also contribute energy as well, OK? So, charge-charge interactions, these are govern by the social convention like opposites attract, OK? So, in social circles, opposites attract, I think is commonly accepted. It works as a formula for dating websites. It also works recently well as a formula for molecules as well. So happily social conventions mirror atomic formulas, OK? So, charge interactions are one possibility. If I go back, you could see that we don't really have any charge interactions operative in this mechanism as depicted here. Nucleophile is neutral, acid chloride also neutral, charge interactions off the table. Second term, repulsive interactions. So, this would be if the molecules have some sort of steric hindrance that prevents them from overlapping with each other. And this is a really important component in terms of preventing molecules from interacting. It's used extensively in biology. It's used extensively in enzymatic catalysis. Again, over here, that doesn't seem to be a possibility, right? The nucleophile has a wide open lone pair. The acid chloride similarly wide open, it has just a methyl group on this attached to it. So, there's really no repulsive interactions that are operative here. And by the way, just to remind you, the repulsive term of this equation is the term that allows this hammer to get pound-- to pound in the nail in this wall, OK? So, these repulsive interactions, that's basically the Pauli Exclusion Principle, that means that electrons cannot occupy that more than two electrons cannot occupy the same molecular orbital, OK? And so, for this reason hammer starts pounding on nail, nail goes into the wall to get away from hammer, OK? They don't, you know, suddenly merge with each other and magically start to create some sort of hybrid material, OK? Things don't happen that way. OK. So repulsive interaction is clearly important, not so operative in this reaction, right? These two can snuggle up as close as they want. There's no, you know, prevention of that by, you know, steric shrubbery. Last one, attractive interactions. This third term, I would describe as mysterious, right? This is not the term that we used to talking about. This attractive interaction is nothing more than the filled-unfilled overlap that I've been talking to you about today, OK? So, here reduce down to its terms is a different representation of the same equation one from up above. In other words, their reaction energy for a particular set of interactions is proportional to Coulomb's law which governs charge-charge interactions plus the steric terms, minus the filled-unfilled orbital overlap, OK? And it's this third term over here that governs whether or not the molecules actually get to form and to break bonds. OK. Now, here is the deal, the problem is that these three terms interact in a complicated way, OK? That if we go out and just, you know, start applying this equation to every possible social situation we find ourselves in, we're going to have trouble, OK? And I guess the most obvious thing is, you know, the opposites attract rule only carries you so far, OK? Before you get married to your, you know, snugly significant someone, it might be a good idea to find out whether that opposites attracting carries over to, you know, I don't know, temperature of the bedroom or something like that, OK? So, for this reason, this equation over here is a good deal more complicated. Why don't we take a look? OK. So, opposites attract. Here's an example, we have hydroxide, we have nitrogen. If they attract so much negatively charge hydroxide, positively charge nitrogen. Our first instinct might be to try to attempt to draw a bond, an arrow between the lone pair on this hydroxide and the positive charge on the nitrogen. That would be wrong, wrong and wrong. It will be totally wrong. And the problem is that this is wrong at every level. The results here would be a fifth bond to nitrogen. And nitrogen being in the first row in the periodic table cannot possibly handle such a large number of bonds. Remember, first row at the periodic table, carbon, nitrogen, oxygen cannot handle more than eight electrons around the atom, OK? That's four bonds. Five bonds totally wrong, OK? Another big problem with this that infuriates me is notice the arrow starting on the negative charge and moving to the positive, OK, that's wrong too because again, arrow is supposed to depict overlap of orbitals. I'm getting a little ahead of myself, OK? Here's the correct way to do it. The correct way to this is to show hydroxide attacking the carbon in and displacing the positively charge nitrogen in an SN2 reaction, OK? So these opposites attract business only carries us so far, OK? So that's our first problem. Is that this is-- that this really, this charge-charge interactions is very rare to provide an operative mechanism in organic chemistry and for that matter in bio organic chemistry. Really charge-charge interactions are very important for non-covalent bonding. Not so important for covalent bonding and in fact, potentially very, very misleading, so cautionary note. Instead, we need to turn to molecular orbital theory. Molecular orbital theory can explain the otherwise unexplained. And I'll give you one example of this before we go back to our canonical example that I showed you earlier, OK? So, for example, this methyl ester has a preference for the syn conformation versus the anti conformation. And to the first approximation, this should strike you as rather odd, right? Because in this case over here, the methyl group is as far as can be away from the lone pairs that populate the oxygen, right? Those lone pairs that stick up like Mickey Mouse ears above the oxygen. And so, this anti conformation showed to a first approximation appear to be the preferred orientation. But, you know, when we look closely at this and we can using various spectroscopic techniques, what we find is actually the dominant conformation is the syn conformation. And you can start to understand this if you think about overlap of molecular orbitals. OK. Here is-- you know, here's-- again, that's the syn conformation should appear to have some steric clash. But again, molecular orbitals explain why it is that it doesn't prefer that. OK. So, I keep talking about molecular orbitals. I think it's time for us to dive right in and start to dissect them and look at them in greater detail and let's get started. So, in molecular orbital theory, we're going to be talking about atomic orbitals. So, the atoms of a molecule each have atomic orbital associated with it, OK? So the nitrogen has some atomic orbital. The oxygen, the carbon, even the hydrogen has some little tiny molecular orbi-- or some little tiny atomic orbital associated with it. Those atomic orbitals are found in S, P, D and F orbitals, OK? That's where the electrons hang out. They hang out in shells or orbitals. I prefer the word orbital which describe their orbit as they orbit around the nucleus of the atom, OK? And remember those electrons, that's the business end of the atom. That's what endows it with functionality. That's what makes molecules the way they are, OK? Now, here's the thing. Often times, these electrons are not simply in either an S orbital or P orbital instead they typically hybridize into hybrids of S, P orbitals, OK? And we're used to this concept. These hybrid atomic orbitals are given the names SP3, SP2 and SP. Here's the important part, OK? So, this is review, I know that you've seen these hybrid atomic orbitals before. This is the part that matters to us as chemical biologist and bio organic chemist. The S-character of these hybrid atomic orbitals determines its stability. This totally makes sense, OK? So, an S orbital is a sphere where in the very center of the sphere is the nucleus of the atom. Nucleus is positively charge. The sphere defines the orbit of the electrons. And in the sphere, those electrons can cozy up as close as possible to the positively charged nucleus, OK? So, this is an-- a great example of opposites attract and the attraction equal stability. On the other hand, a P orbital as depicted up here has a nucleus, a node between the two lobes of the orbital, OK? So, the nucleus is right here in the center again, but that happens to be a zone of exclusion where the electrons are not allowed to exist rather the electrons in this orbital are hanging around either in this node up here or this other node down here. They're not in-- oh sorry, lobe. This lobe up here and/or this other lobe up here, they are not allowed to get up too close to the positively charge nucleus. And so, for this reason, the S-character of a hybrid atomic orbital determines the stability of that orbital, OK, of the electrons in that orbital. Conversely, the P-character defines the instability. It defines how reactive and how nucleophilic those electrons in that hybrid atomic orbital really are, OK? That's kind of like, you know, defining how unhappy the electrons are, OK? Happy electrons are found in the spherical S orbitals, unhappy electrons are found in P orbitals. And what happens when electrons are in unhappy situations? Well, they will move. They will do everything they can to find more stable orbitals for themselves. OK. So, these are the atomic orbitals, specifically the hybrid atomic orbitals over here, and P-character conveys-- confers reactivity and basicity. So, for example, if we look at a series of lone pairs found on carbon, what we find is that the higher the P-character the more reactive that result in lone pair will be, OK? And this could be dramatically illustrated in terms of basicity, OK? So, here's a lone pair in an SP3 hybridized orbital. Its pK is 50. Compare that against a lone pair in an SP or an SP2 hybridized orbital. The difference here is truly dramatic, OK? So, the pKa is only 41 in the case of the SP2 hybridized orbital and then, it's way down at 24 down in the SP hybridized orbital. This is an enormous difference, OK? Remember, pKas are a log scale. So in other words, this guy up here is 10 to the 26th times more reactive than this guy down here. And by more reactive, I mean, how avidly it's going to be reaching out and ripping hydrogens, ripping protons off of its neighbors, OK? And this tells us almost immediately that for example, you know, organic metallic compounds are going to be extremely avid at grabbing protons to the point where they're nearly-- they're incredibly flammable and nearly explosive. OK. Now, this 10 to the 26th times again is huge, right? That's a 1 followed by 26 zeros. It's such large number. It's hard to actually for us to even imagine it. OK. So, enormous difference is determined by this P-character, S-character. I hope by now, everyone who's listening to us and everyone in my class can explain why it is that these guys are so much more reactive than these guys. And it should make sense just from geometric considerations as depicted here. Now, these hybrid atomic orbitals recombine into molecular orbitals in molecules, OK? So, the hybrid atomic orbitals only carry us so far. More often, these hybrid atomic orbitals are shared between atoms and that sharing is what gives us bonds, OK? Now, these molecular orbitals are given the names sigma, pi and N, OK? So, this hybrid atomic orbitals form bonds with other atoms and that yields molecular orbitals. The energy of these molecular orbitals is defined very specifically. And there's no way around this. I basically just have to tell you, I'd like you to memorize this chart on this slide, OK? So, please memorize the order of this reactivity where a sigma molecular orbitals are lowest in energy, pi are higher in energy, N are even higher, OK? So these are the fields on molecular orbitals. These are molecular orbitals that have electrons in them. And these electrons are depicted by the up arrows and the down arrows, OK? That's a convention that you've seen before. OK. Now, sigma makes sense. Sigma are the molecular orbitals that define a single bonds, S for single, S for sigma. Pi, this defines double bonds and that's convenient, right? Pi looks kind of like a double bond. And the electrons in N orbitals are the lone pairs that are hanging out around the atoms, OK? So, when the N orbitals are present, those are going to be the highest occupied molecular orbitals. So, almost immediately, that clues us in that we need to pay attention to those lone pairs, OK? What about the unfilled molecular orbitals that we're going to encounter in chemical biology? These will be found in three molecular orbitals. And again, I need to ask you to memorize these-- the order of the energies, OK? The lowest in energy are P orbitals. P orbitals are exactly what I showed you a couple of slides ago, OK? That's them. These over here. This is what a P orbital looks like, it has a lobe and another lobe down here. P orbitals we find when we look at carbocations, OK? The empty hole that is the carbocation is a P orbital, OK? The other electrons that surrounds the carbon, that surround the carbon or the carbocation, those other electrons are in an SP2 hybridized atomic orbital. So, the remaining empty atomic orbital is a P orbital, OK? So, most of the time, we don't really have carbocations. The reason for this is that they are extremely reactive being so low in energy and so, for this reason in biology we very, very rarely find carbocations. In week eight, I'm going to show you an exception to this. But for now, let's keep in mind that we're just not going to see this very much. And again, the reason is biology takes place in water and carbocations react avidly with water. Pi-star, this is the antibonding complement partner to the pi orbital. And sigma-star is the antibonding complement to the sigma orbital. And again, we're seeing this relationship where pi-star is lower in energy than sigma-star. OK. So, here's what I need to tell you. Good news. You don't have to worry about where all those electrons are in a molecule. And it's really fabulous news, OK? If you just stop, take a moment, take a deep breath, pause and appreciate this. Because the molecules we talked about when we talk about biology are fiendishly, fiendishly complicated, OK? This goes back to the business that I talked about earlier of the hundreds, if not thousands of pi bonds, the thousands upon thousands of lone pairs. The good news is we get to simplify all of that complexity down to just worrying about the frontier orbitals, OK? So in other words, we only have to worry about the frontier highest occupied molecular orbital and the frontier lowest unoccupied molecular orbital, OK? In other words, all we have to do is focus in on the highest occupied lone pair or highest occupied molecular orbital that has a lone pair in this N orbital or/and also the lowest unoccupied molecular orbital over here. So in other words, if there's an available P orbital, it's going to react first. OK. If there's a carbocation, everything else will come to a halt and carbocation gets to stay in the sun, it gets to dance around, OK? If there is a lone pair, lone pair will be the dominant reactivity, OK? This is good news. OK. It simplifies everything. We just have to look for the highest energy, HOMO, and the lowest energy, LUMO, OK? What does this mean? What this means is that this highest occupied, a HOMO is the filled frontier orbital. And this is the orbital from whence all nucleophilicity, all basicity springs forth, OK? And I apologize for the kind of antiquated English but really that's how we think about this, OK? This is the orbital. That is the business end of this complicated molecule. It doesn't matter how many possible lone pairs it has. It doesn't matter how many different possible configurations it has. All that really matters is its highest high and its lowest low, OK? Again, this is majorly important because it simplifies things for us. OK. So this HOMO, highest occupied molecular orbital is the filled frontier orbital and it's a nucleophile in reactivity. OK. Now, the intrinsic nucleophilicity is governed by the energies of these molecular orbitals where again, the highest in energy is the N bonding, the non-bonding molecular orbital that has the lone pair and the lowest in energy are the electrons of the sigma or single bonds, OK? To reduce it down to simplest terms, we're never going to be really seeing reactions that start with sigma bonds, OK? It just doesn't happen in chemical biology. Most of our reactions are going to spring forth, are going to spring from lone pairs that are in non-bonding orbitals, occasionally electrons and pi bonds, but really-- we don't really have to worry about electrons and sigma bonds. We know they're there, you know they're there, they're there, but we don't have to get wrapped up in them. And this again is good news because there's a huge number of electrons in this complicated molecules that have, you know, thousands upon thousands of atoms. OK. What about the lowest energy unoccupied molecular orbital or the LUMO? This is the unfilled frontier orbital and the lowest energy unoccupied is the most available molecular orbital. This is the molecular orbital that's going to be the center of attention for reactivity, OK? And again, where you have these complex molecules, this is kind of like the funnel to which all reactivity zooms towards, OK? Now again, we need to know this order over here where P is lower in energy than pi-star which in turn is lower in energy in sigma-star. So, if we're given a choice of different sites for nucleophile to attack, nucleophile will choose every time to attack the P orbital because it's lowest in energy. And again, we see P orbitals when we look at carbocations. If there are no carbocations present which again, I had said earlier, is exceptionally common because carbocations are very, very rare in biology where biology takes place in water. So, if there are no carbocations present, we can eliminate this one and we start focusing on pi-star orbitals. If there are pi-star orbitals that are available for reaction, then it's likely that this will be dominant, the dominant reaction. Occasionally, you come across a molecule that doesn't have a pi-star in which case, then you might have an attack on a sigma-star. This is rare, OK? Especially it's depicted here. This is utterly wrong. It's depicted in the slide. I find it offensive, but I'm stuck with it. This might happen for example if there was a sulfur here, then you might have the sort reaction taking place. For now, let's keep in mind that we're going to probably be having reactions or electrophiles in our reactions are going to be molecular orbitals consisting of antibonding pi bonds, OK? So, it's the pi-star or antibonding pi molecular orbital. OK. I want to switch gears. If you have any questions about molecular orbitals or hybrid atomic orbitals, don't hesitate to shoot me an e-mail or talk to the TAs, come to my office hours, et cetera. We now have to talk-- I told you about what you do to decide what the reaction mechanism is. We now have to talk about how to actually tell me what that reaction mechanism is, OK? So, often times at chemistry we have some notion that molecules are reacting but we need a clear way of communicating that reactivity. So, organic chemists have developed this wonderful vocabulary using arrows. And so, let's take a closer look what those arrows are. The arrows are going to be starting from highest energy occupied molecular orbitals, the HOMOs, and they're going to be ending on the lowest energy unoccupied molecular orbital, the LUMO, OK? This is a golden rule. This is a rule that always applies, OK? Your arrows, start on orbitals, they end on orbitals. They start on HOMOs, they end on LUMOs. And again, they're always going to start on the highest HOMO and the lowest LUMO and they'll end on the lowest LUMO. OK. So again, that lowest LUMO is the lowest energy unoccupied molecular orbital, that's the most available and in turn that's the most reactive. Now, the problem is again, we often times have many HOMOs, we have many LUMOs, what's an organic chemist to do? What a student supposed to do? So, when in doubt, refer to this idea of looking for the highest HOMO and the lowest LUMO. I can simplify it, cut it down to make it even easier for you. Most of the time, just start, put your pen on a lone pair, and start pushing electrons to end on the best electrophile, OK? It's not easy. If you're in doubt, you're stuck there at your desk during an exam, you don't know where to start, put the pen on the lone pair and just start drawing, OK? End the arrow on the best electrophile, nine times out of 10, 99 times out of 100, maybe even more, you'll get the answer right just by doing that. OK. So, I need to talk to you about some rules. We have rules because it's a language. And in order for us to be clear in what it is that we're communicating, we need to have some conventions. OK. The conventions we're going to follow in this course are the following. And by the way, before I present these conventions, I should tell you, I'm a stickler for these rules, OK? If you give me something that doesn't have-- doesn't follow these three rules, chances are even if it's correct conceptually, it won't get full credit, OK? And the reason for this is it's kind of like turning in an essay that has incorrect grammar to your English class or something like that. What's your English professor going to do? You know, give you an A for great ideas and a C for bad English? No. Your professor is probably going to give you a C overall because the goal was to communicate effectively. OK. So in the same way, when we speak using the language of arrows, we have to follow these conventions because this is what convinces us that we know what we're talking about, OK? So the conventions are arrows never indicate the motion of atoms. And this is one that if we simply think about it, actually it's kind of profound. I think that all of us are used to having arrows showing, you know, football player who's over here, let's say the quarterback moves back here and then gets behind this guy and then another arrow shows this guy moving forward. Those were the kind of arrows that you've been drawing, you know, I guess since you were able to draw arrows, OK, which is to show motion, to show the fourth intervention really, to show some element to time. Organic chemistry, we don't use arrows in that way. Rather, we're using arrows to depict overlap of orbitals. We're not depicting it in terms of time. We're depicting it in terms of thermodynamics, not kinetics. In other words, we're depicting an overlap of orbitals that's allowed, OK? So, arrows do not indicate the motion of atoms. Yes, it's true. The atoms must cozy up to each other. That's kind of understood. That's lurking in the background. But that's not really what arrows show you. Arrows never start or end on charge. Since these arrows are depicting the interaction of filled and unfilled molecular orbitals, charges are relevant, OK? Charge, formal charge is one of those nice conventions that makes the lowest structures so much easier to understand, yet the charge itself does not show you where the electrons are. It doesn't show you anything about the molecular orbital. And so, drawing an arrow from one formal charge to another is worthless, OK? So again, arrows never start or end on charge. Arrow instead-- here's the one that really embodies everything. Arrows begin with lone pairs, with pi bonds or sigma bonds and end on unfilled orbitals. OK. And I want you to be really precise about how you draw these things. That precision indicates that you understand what is going on, OK? And drawing your arrows to precisely end where it is that that pi-star orbital should be, you're telling me something, you're telling me a story, you're telling me where it is that those electrons are going to appear. And in doing that, you're describing to me the reaction that's talking place, OK? So I need to have all of these things taken care of when it comes time to, you know, for exams and things like that, OK? Make sense? OK. So, why don't we take a look at an example, OK? So an example is this very simple problem, OK? And the problem is we're going to have a lone pair on the nitrogen over here and we're going to do a simple substituted nucleophilic attack, substituted nucleophilic reaction that substitutes for chlorine this lone pair on nitrogen or this nitrogen, this amine. Everything looks good. This is a reaction very similar to the one I showed you at the very beginning of the class. Avoid the poisoned candied apple of simplicity rather a fall back on HOMOs and LUMOs. Let me show you what I mean. OK. And to do that, I need to roll up the screen. Here's what I mean by avoiding that tempting poisoned apple of simplicity, OK. So, in this example-- [ Pause ] In this example, there's a lone pair on the nitrogen. All right. OK. So here's our reaction again. And the simple-- and I would call it even simple minded possibility is for the nitrogen to simply displace the chloride as an SN2 reaction. All right. So the simple case, we have this reaction mechanism here. [ Pause ] OK. OK. This case, I would call an SN2 reaction, right? Substituted nucleophilic two reaction and this will give us-- [ Pause ] -- this guy over here. And then, this can lose a proton. OK. So I'll show this base. The base for example could be chloride. [ Pause ] OK. And this base can deprotonate. This nitrogen giving us the product. [ Pause ] OK. Now, what's wrong with this? This is totally, totally wrong and completely unacceptable. It appalls me. It's upsetting to me. What's so wrong about this? It's so appalling, is the fact that we're attacking a sigma-star orbital, OK? This is an attack on this sigma-star orbital over here, right? When we have several perfectly good pi-star orbitals that are available, right? So sigma-star is not the lowest energy unoccupied molecular orbital. Far from it. There's plenty of pi star molecular orbitals that are going to be lower in energy. Why don't we explore those as a possible reaction mechanism, OK? So, a different mechanism would start. OK. So I'll draw some lines through here, a different and more correct mechanism. Well, this time have the lone pair in that end, non-bonding orbital of nitrogen attacking the pi-star orbital of this alpha beta unsaturated carbonyl. Let me show you. OK. So here's the lone pair, it's now going to attack the pi-star molecular orbital, electrons bounce, bounce all the way to the electronegative oxygen, OK? So again, this is attack not on a sigma-star but it's attack on a pi-star molecular orbital, an antibonding orbital. OK. Now, why is this so much better? This is better because the pi-star is lower in energy than sigma-star. And so, for this reason, this addition elimination reaction is greatly preferred, OK? This is actually the operative mechanism for this reaction. OK. You could continue on, I encourage you to do so, and in end you get this product over here. OK. There will be many times in this class. I'll kind of step up for you. I'm going to then let you finish it off on your own. I apologize for that. This is upper division organic chemistry, upper division chemical biology where at that point where I don't have to show you every step. There are fundamental steps that I want you to know, there are fundamental steps that I expect you to know, but I'm not going to show them to you during every lecture, OK? Rather, I want you to go home, I want you to fill them in in your notes. I want to make sure that you know them because on exam I will ask you to show me those steps. But on the other hand, I'm not going to dwell on them today, OK? We just don't have enough time to talk about them in the class of this length. OK. So, the lesson from this is clear. The lesson is don't be tempted by simplicity, instead look at the overlap of orbitals. Which one is a better overlap? Overlap of the sigma-star or overlap with the pi-star? Pi-star is lower in energy and it's therefore greatly preferred. OK. Let's move on. I have to lower this again. All right. Now, the other thing to make this work is that you also have to make sure that you're drawing the correct Lewis acid structure. For the most part, I don't think this is going to be a problem in Chem 128, but it is important that you set things up correctly, OK? If you are drawing for example five bonds to nitrogen, a lot of the reactivity of this in oxide is not going to be apparent because this is totally wrong. OK. Similarly, you know, in terms of the number of bonds that you draw, this helps you in terms of keeping track of things. For that matter, it's also essential for you to depict correctly the formal charge. Oh thanks, I'm sorry. This formal charge helps to guide us. For example, the negative charge on this carbon over here, that should look kind of funny to you, right? Covalence, that should look funny, that should be extremely reactive. So, formal charge helps to guide us in terms of drawing these correct mechanisms. I'll have a lot more to say about hydrogen bonds. I don't really care about dative bonds. We won't see them in this class. Don't-- let's not get into it today. I'll talk to you more about hydrogen bonds in a moment. OK. So, arrows start with bonds or lone pairs. And here are some correct depictions of arrows, OK? So in this case, where we're showing a bromide leaving for an elimination reaction. The bromide takes off. And notice that the arrow is starting at the carbon-bromine bond, OK? In other words, the electrons in that carbon-bromine bond decide to step out the door and leave with their friend, the bromine, giving us a bromide ion, OK? So, here's electrons that are starting with another bond. In this case, a pi bond. Here they are starting with a sigma bond, here they are starting with a pi bond, and here they are starting with a non-bonding lone pair, OK? All three of these cases are correct. Contrast that with these cases over here where I'm showing you arrows starting on charges. This again is deeply appalling and totally wrong. So, arrows do not start on atoms. So for example like this or like that, instead we want to draw them starting on the bonds themselves. This should make sense, right? Arrows are trying to depict the overlap of orbitals. They need to start where the electrons are. The electrons are found in these bonds. Electrons are not found in this negative charge. They're not really found around this bromide, instead we're talking about the electrons that are shared between bromine and carbon. Those are the electrons that matter. OK. Now, that's where they should start. Let's talk about where they should end. So, arrows need to end on atoms and bonds, OK? So, here's a lone pair attacking a proton. It's ending directly on that proton, OK? So, here is it ending on an atoms, the protons, here it is ending on the carbon of a carbocation, here it is ending on the hydrogen or the proton during in a beta elimination step. OK. So, atoms never terminate in empty space. So for example, when bromide is stepping out the door, the electrons don't simply hop out and then the door over opens into empty space. For that matter, this arrow would be wrong if it started at this carbon-bromine and then had the electrons just going off into empty space. That's not correct. The electrons don't get to walk off into empty space. That would be extremely high in energy and extremely repellant, rather the electrons get to end on this bromine atom giving us bromide, OK? So, arrows need to end on atoms. They will depict again this overlap of filled and unfilled orbitals. OK. Hydrogen for that matter is always attached to something. OK. I'm starting to get down to my pet peeves but this is one of those pet peeves that doesn't matter, OK? Hydrogen is not some atom that kind of like it is floating around next to the molecule, rather hydrogen is directly attached to some particular atom. And this matters a great deal because where it's attached will determine to large extent whether or not it's going to be acting as an acidic proton or perhaps not acidic at all, OK? So, these terms, proton, hydride, and hydrogen atoms are three different depictions of the hydrogen atom, and they have three different meanings. They have totally different meanings, OK? So, H plus is the proton, H minus is the hydride, and H radical is the hydrogen. They really-- we don't find them just kind of floating around like this in the chemistry that takes place inside cells, OK? Protons aren't just floating around inside the cell, rather they are always attached to something. Maybe they're attached to a water molecule to give you a hydronium ion. But they're not just kind of hanging out, they're doing something, OK? Hydrogens do not like being by themselves, OK? So, in other words, what you want to avoid is showing proton just kind of hanging out in space, waiting around for some lone pair electrons to attack it. That's not what happens. Hydrogen doesn't get to do that, OK? Furthermore, hydrogen radical also doesn't really occur nor does hydride really occur. OK. Rather in solution chemistry, we find species that could either donate a proton, donate a hydrogen radical, or donate a hydride, OK? So, what I propose you do is instead of depicting H plus as a reagent, instead depict H plus as catalytic "H plus." Those quotes, unquotes are going to tell us that yes, we mean H plus. But what we really mean is we mean H plus that's been picked up and delivered by some other species. In this case, that might mean attached to this methyl, this methanol molecule that's going to be its delivery character, or you can even write catalytic HA where in this case, it's HA that's attached to the conjugate base that's going to be delivering the proton, OK? Anyone of those is fine. It is important for you however to follow these conventions because they communicate to me that you know what molecular orbitals are being overlapped and it tells me whether or not, excuse me, you understand the chemistry that's involved with these reactions. OK. One second. OK. I want to conclude today's lecture by discussing with you one other element of hydrogen atoms and that's the hydrogen bond. Everyone needs a favorite bond. My favorite bond is of course the great Sean Connery. But today, I'm going to be talking to you about a second favorite which is the hydrogen bond. OK. The hydrogen bond govern so much a biology that it is essential for us to really get to understand it correctly, OK? So, hydrogen bonds are actually largely a Coulombic interaction. They describe the sharing of a hydrogen atom between two partners. One partner is going to be our hydrogen bond donor and our second partner will be a hydrogen bond acceptor. And this hydrogen bond will be depicted by this dashed line, OK? So this dash line is/are going to be our convention for hydrogen bond. We're going to use this a lot, OK? Hydrogen bonds for example hold together the two strands of DNA. They make molecular recognition possible, the non-covalent interactions between molecules. So, hydrogen bonds absolutely essential to chemical biology. However, it turns out that the energy of the hydrogen bond is very sensitive to the environment and the geometry that's involved with the sharing of that hydrogen. The geometry in this case that I'm showing you over here is of a perfectly linear hydrogen bond which is the best possible example, OK? So, in this case, the lone pair on this water on this oxygen of water down here is perfectly positioned to share this hydrogen of this water up here and the oxygen, hydrogen and oxygen are lined up as a straight line. Often times that isn't the case, OK? So, for example, we can look at hydrogen bonds that are found in the active sites of enzymes and we find instead of having this neat straight line, we get a bendy line instead. That bendy line, that bent hydrogen bond, much, much weaker, OK? So, this is kind of the optimal geometry, optimal hydrogen bond acceptor which is a lone pair. Optimal hydrogen bond donor over here. And this is kind-- this will be our canonical hydrogen bond. Now, here's one of the problems. One of the problems amongst others is curved arrows. Curved arrows confound us when it comes time to talk about hydrogen bonds. The reason is curved arrows and hydrogen bonds simply don't mix. Curved arrows depict the overlap of filled and unfilled molecular orbitals, whereas hydrogen bonds are showing us a partnership of sorts between the donor and an acceptor. And there really isn't this sort of overlap that leads to covalent bond in the case of hydrogen bond. And this becomes tremendously confounding. So, for example, if you want to just show transfer of the proton on this nitrogen atom to the lone pair of the oxygen, you might be tempted to simply draw a hydrogen bond in here. And that would be utterly incorrect because this hydrogen bond is basically saying the hydrogen is somewhere between here and there, somewhere in the middle, somewhere in the sides, whereas over here in that case of the curly arrows, you're saying, no, it's going to pick this up, it's going to pick up the proton wholesale, hang on to it for a while and give you a positive charge on oxygen. These are two very different depictions. So, what we've-- what we were going to be doing in this class is showing those hydrogen transfers as an explicit step, OK? So, hydrogen bonds are going to be useful for us for talking about non-covalent interactions, but not useful at all for talking about covalent interactions, the reactivity that I've been talking to you about today. It turns out that hydrogen bonds that proton transfers of the sort that I showed in the previous slide, these sorts of proton transfers over here are extraordinarily fast, OK? There are often times diffusion controlled. In other words, they hit the speed limit of reactivity for reactions that take place in solution. That kind of speed limit and that kind of proton transfer ability is actually tremendously useful, OK? So, this is a diffusion controlled reactions. So, proton transfers to and from heteroatoms very, very fast. Proton transfers for that matter in the same way that hydrogen bonds require a linear geometry. Proton transfers also require linear geometries. And I can tell you that almost immediately this is going to annoy you. This takes away one of the conventions that you mislearned back in sophomore organic chemistry. I know it was cool back then to show a proton transfer as a neighboring oxygen, say picking up a proton over here on the nitrogen and you have this completely ridiculous and totally crazy four-membered ring transition state. It galls me to even say this. Can you imagine four atoms getting together to form, you know, some sort of very strained four-atom ring transition state? It's totally-- it's total insanity. Even more insane, notice that the geometry between oxygen, hydrogen and nitrogen is not perfectly linear, instead it's bent at a 90-degree angle. And this kind of proton transfers do not happen this way, instead they exclusively prefer a linear geometry. So, only linear geometries are going to count when we talk about proton transfers. And so, for this reason, I need to take this particular step out of your vocabulary, OK? And now, it was acceptable back in sophomore organic chemistry, it's no longer acceptable. So, acids and bases are required to catalyze proton transfers and tautomerizations. So instead of showing it like this, a much better alternative would not even alternative, the correct way to depict this would be to show the oxygen picking up a proton from a catalytic acid. And then, in turn the conjugate base of this acid acts as a base to deprotonate the neighboring positively charge nitrogen, the ammonium ion. OK. So, in all cases, we're going to see that acids and bases are required to catalyze these proton transfers and that turns out to be a general rule. And the good news for us in terms of chemical biology is that often times or all times, we can find abundant numbers of different molecules that are all too willing to volunteer to be those catalytic acids and bases and obvious examples would be for example water. Water can be a hydronium ion to act as a proton donor. Water can also act as a base to accept protons and become a hydronium ion. And since all biology takes place in water, feel free to use water as the catalytic acid and that catalytic base. OK. We've come quite a ways. I've shown you proton transfers, I've shown you how to draw arrows, we've talked about the rules that govern these electron transfers in terms of filled-unfilled overlap of molecular orbitals. We're now going to transition to looking at some examples of this. And I'm going to show you this on Thursday when we talk about the molecules found on earth that composed all living things. So, why don't we stop here. When we come back next time, I'll be showing you examples that apply the principles that we've talked about today. Thank you very much. [ Silence ] ------------------------------6c7738525c2e--
Info
Channel: UCI Open
Views: 13,056
Rating: undefined out of 5
Keywords: UCI, UC Irvine, OCW, OpenCourseWare, Weiss, Reactivity, Arrow Pushing, Chemical Biology, Orbital Overlap, Atomic Orbital, Molecular Orbitals, Hybrid Atomic Orbitals, Lewis Structures, Hydrogen Bonds, Mechanistic Arrow-Pushing
Id: IyU7Qyh9iIg
Channel Id: undefined
Length: 73min 30sec (4410 seconds)
Published: Wed Mar 06 2013
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.