Placeholder Image

字幕列表 影片播放

  • J. MICHAEL MDBRIDE: So, we'll finish up on

  • NMR spectroscopy, talking

  • about more dynamic things involving decoupling.

  • Then a little bit about C-13 NMR and double labeling

  • involving correlation.

  • And then we'll get onto another reaction,

  • electrophilic aromatic substitution.

  • OK, so proton decoupling.

  • Remember that the proton is four times as strong a magnet

  • as the Carbon-13, so it precesses much faster.

  • So it's easy to do a pulse that will tilt the Carbon-13

  • net magnetization down into the plane, so that it'll

  • precess and be an antenna that can tell us its local field

  • without doing the same thing to the proton.

  • And if we do that, then we see the C-13 NMR spectrum.

  • And there are two peaks, because of

  • the adjacent proton.

  • Right? it can be either up or down, 50:50

  • or almost exactly 50:50

  • so you see two peaks.

  • Now suppose at the same time you're doing that, when you do

  • this pulse and listen to the C-13 to see what its local

  • field is, you irradiate at 100 MHz.

  • Right?

  • That won't do anything to the C-13,

  • it's the wrong frequency.

  • But it'll now make the proton precess.

  • That is, if we look in the rotating frame, we can see a

  • weak field that's horizontal, about which the

  • magnetism will precess.

  • So the protons will go down, and then up, and then down,

  • and then up.

  • And what determines how fast it goes up and down?

  • How strong this field is.

  • The stronger it is, the faster it will precess.

  • And if it's fast enough, then you won't see a doublet for

  • the C-13 anymore, because you'll see the average.

  • Right?

  • So you'll see a single peak, you'll proton-decouple the

  • C-13 spectrum and see the C-13 as if the

  • protons weren't there.

  • OK, so it depends on how much power you put in.

  • Now here's a C-13 spectrum of a given compound.

  • Where we're listening to the C-13 and we see all these

  • different local fields, right?

  • But we're irradiating the proton at the same time, but

  • very, very weakly.

  • This number up there, 40 decibels, is an inverse ratio

  • of how strong that horizontal field is for the protons.

  • So it's 10-4th weaker than what you'll see you

  • later, 10,000 times weaker.

  • And if we look at the carbon, we see that some of the carbon

  • signals are split and some are not split.

  • The ones that are split are the ones that have

  • hydrogens on them.

  • So if we start from the left, we see there's one with no

  • hydrogens, then one with one hydrogen,

  • one with one hydrogen.

  • That red one is the solvent

  • CDCl3, right?

  • So we're looking at the carbon of the CDCl3.

  • But it has a deuterium on it, not a

  • hydrogen, so it's not split.

  • Then there's a carbon with two hydrogens, a carbon with three

  • hydrogens, it's a quartet.

  • Underneath, a carbon with two hydrogens, it's a triplet.

  • So very low power in the proton, that's what we see.

  • Now we're going to reduce the size of that.

  • And see what happens, is we start increasing the power.

  • So there's 40 dB, 20 dB,

  • and now things are beginning to change.

  • See how the patterns are changing their multiplicity

  • and their intensity?

  • Because it turns out that when you irradiate the protons by a

  • complicated mechanism it strengthens the signal of the

  • carbons to which they're coupled.

  • And that's called nuclear Overhauser enhancement, as

  • you'll see.

  • OK. So now you see the splitting is going away when

  • we get down to two decibels or one decibel.

  • Now we see that each carbon gives just a single peak, we've

  • decoupled the protons.

  • So now we have a peak for every carbon.

  • And notice, incidentally, that the CDCl3 didn't get stronger

  • as we did this.

  • Why?

  • Because what strengthened the signal was irradiating the

  • protons and CDCl3 doesn't have hydrogens on it.

  • So it doesn't get strengthened.

  • OK, notice also that the carbon without a hydrogen on

  • it, on the far left here, didn't get strengthened as

  • much as the others did.

  • OK, so that's decoupling, and it can be useful.

  • And it can be useful in several ways.

  • There's two things, you get rid of the spin-spin splitting

  • so you see just a peak for each carbon.

  • And also you strengthen the signals when

  • they have nearby protons.

  • And that could be bad if you're interested in measuring

  • the size of the peaks to get how many carbons there are,

  • because you can't count properly anymore because you

  • get different intensities with different

  • neighbors of protons.

  • But it's good if you want to determine structure and see

  • what hydrogens are how far away, and I'll

  • show you that shortly.

  • OK, now bear in mind that carbon-13 is only 1% in

  • natural abundance of carbon.

  • Of course, you can synthesize the molecule with expensive

  • carbon-13-containing precursors and put carbon-13

  • where are you want it in a molecule, and then the signal

  • from that carbon will be ever so much stronger.

  • If you put in pure carbon C-13 in some position, the signal

  • would be 100 times stronger.

  • In fact, 200 times stronger than without.

  • No, pardon me.

  • It would be 100 times stronger, because it would be

  • 100% instead of 1%.

  • OK, so here's a compound, neotame, it's

  • an artificial sweetener.

  • And it's the proton-decoupled C-13 spectrum.

  • So you see one peak for each carbon.

  • There are 20 carbons, there are 20 peaks.

  • But you can't go by their intensities.

  • Now, let's see what we can make of this.

  • It's proton-decoupled, so there's one peak per carbon

  • and, being carbon, it's pretty well spread out.

  • Remember, the range of hydrogen chemical shifts is

  • only about 10 parts per million.

  • The range for carbon is hundreds of

  • parts per million here.

  • 200 parts per million are shown.

  • Now why is there no C-13-C-13 splitting?

  • There's no proton splitting because it's decoupled.

  • But why no C-13-C-13 splitting?

  • STUDENT: The likelihood of two C-13s adjacent in the same

  • molecule is unlikely.

  • PROFESSOR: If it's only 1% natural abundance, and you're

  • a C-13 so that you're giving a signal, the chance that your

  • neighbor is C-13 is only 1%, so there would be tiny, tiny

  • peaks from those

  • you don't see them.

  • OK, so that's what Chris just said.

  • OK now, let's see if we can figure out which carbons these

  • are, that we're looking at.

  • And of course, this is how it was done empirically.

  • People put in known compounds and figured out what the

  • chemical shifts... what peaks

  • were from one compound to the next.

  • And then they figured out this comes in this range, this

  • comes in this range, this comes in this range.

  • So let us see, knowing what these peaks are.

  • OK first, there are these ones that are way

  • down close to 200.

  • They're numbers 1, 13, and 4.

  • So what do 1, 13, and 4 have in common that turns out to

  • make them way far downfield?

  • Ellen, what do you say?

  • What do 1, 13, and 4 have in common that makes

  • them different from the other carbons?

  • STUDENT: They're almost like esters.

  • Like, double bonding to an O.

  • PROFESSOR: Yeah, the double bond to an O. Carbonyl carbons

  • are way downfield.

  • OK, now we've got 7 and 8 to 12.

  • So there's 7 and there's 8, 9, 10, 11, 12.

  • Pretty clear what those are, right?

  • Those are the ones on a benzene

  • ring, aromatic carbons.

  • OK, then there are ones that are a little bit further down

  • than the others, and those are those blue ones.

  • What do they have in common?

  • Rahul, what do you say they have in common?

  • STUDENT: Tertiary?

  • PROFESSOR: They're not all tertiary, no. In fact, only

  • two of them have three things associated with them, the

  • others have one and two.

  • Lauren, can you see what the blue ones have in common?

  • In hydrogen, what makes things come a

  • little bit further downfield?

  • Yeah Amy?

  • STUDENT: Well, the hydrogen would be like deshielding from

  • an oxygen or something, like nitrogen or oxygen.

  • PROFESSOR: They all have an electronegative atom next to

  • them, right?

  • Nitrogen or oxygen.

  • So those are ones. A carbon that's attached to something

  • electronegative.

  • And then, finally, you have the others, which are attached

  • only to carbons.

  • OK, so you could begin to figure out these, and you can

  • get tables of what comes where.

  • And then you can start dealing with unknown compounds and

  • figure out what it's telling you.

  • But a nice thing about this proton-decoupled spectrum is

  • that there's just one peak for each carbon, and they're well

  • spread out.

  • So that it's good to do this kind of thing.

  • So those are the alkane carbons.

  • OK, now the power of correlation.

  • Often when you can measure two things on the same subject

  • simultaneously, you get a lot more information than

  • measuring one, or measuring the other, or measuring them

  • independently.

  • I'll show you that example with C-13 double labeling and

  • also in two-dimensional.

  • As it's called, nuclear magnetic resonance.

  • So first, C-13 double labeling.

  • And we'll talk about lanosterol biogenesis.

  • In fact, we already talked about it, if you look back at

  • lecture 52.

  • OK, so remember we got squalene from isopentenyl

  • pyrophosphate.

  • Then it curled around and we had this complicated thing

  • where it zipped up going one way, and then we had all these

  • rearrangements.

  • So as we said at that time, this is the source of

  • cholesterol and the steroid hormones.

  • And it's a very cute story, how all that rearrangement

  • took place

  • OK. But there's a question whether it's true.

  • And I said at that time, wait for NMR. So now we're at NMR,

  • so you can get the answer.

  • OK, so you can use Carbon-13-labeling to figure

  • this out.

  • So if you feed the plant that makes this stuff Carbon-13

  • labeled isopentenyl pyrophosphate--

  • in that particular carbon, that substituted carbon--

  • then you can see where that would appear in the squalene.

  • According to the way we linked them together before, it

  • should be those carbons.

  • Then when you curl it around and do all this stuff to make

  • lanosterol, those peaks should be labeled with C-13 and

  • therefore 100 times stronger than the other

  • peaks in that spectrum.

  • And indeed they are.

  • So that supports it.

  • Or you could do it a different way.

  • You could label the methyl group.

  • And if you label the methyl group, it turns out that those

  • methyl groups should be labeled in squalene.

  • And then if you curl it around, it should be those

  • methyl groups.

  • And indeed, they're 100 times stronger

  • than the other signals.

  • But much more interesting than this is to do double labeling.

  • So if you label both, you get that.

  • Big deal.

  • That's just the same information

  • we got before right?

  • Except done in one experiment rather than two.

  • And, in a sense this is not as valuable as these

  • two independent label experiments, because you don't

  • know but what two of those carbons might, in principle,

  • have changed places.

  • But the double labeling is much more important

  • if you do it dilute.

  • Now what do I mean dilute?

  • I've been talking as if we put 100% C-13 in here.

  • But suppose we don't put nearly 100%, suppose we only

  • put 10% in.

  • But suppose that those 10% that are

  • labeled are double labeled.

  • So we prepare a sample that's 100% labeled in two positions,

  • but then we mix it with 90% percent of unlabeled material

  • and feed it to the organism and get the stuff out.

  • OK, so what we're going to do is a dilute double label.

  • Now why do it dilute?

  • Because then, when label goes in, two labels go in.

  • But it's unlikely that any given squalene will have two

  • different positions double labeled, or two sets of

  • positions double labeled.

  • Because only 10% of the stuff is double labeled.

  • So when a squalene is double labeled, it'll be labeled in

  • just one pair of positions, a given molecule.

  • Like there.

  • So those two come labeled together.

  • The signal is 10 times stronger because

  • remember, it was 10%.

  • It's 10 times stronger than it was for the

  • ones that aren't labeled.

  • But what's special about it?

  • Can anybody see what's going to be special about the signal

  • we get from these molecules that look like that?

  • Ayesha, have you got an idea?

  • STUDENT: They're going to be the same more or less

  • because they're just an extra--

  • PROFESSOR: They're going to be the same what?

  • STUDENT: They're going to be the same more or less

  • because they're just extra carbons and

  • there's no special function.

  • PROFESSOR: I didn't quite understand you, but I was

  • listening for special words and I didn't hear them.

  • Sebastian, have you got a special word for us?

  • STUDENT: You get splitting?

  • PROFESSOR: Ah, they'll split each other, because now the

  • C-13s are next to one another.

  • Because they're not coming in at random 10%, they're coming

  • in as a pair.

  • So now when we look at the signal from that, it's going

  • to be a double doublet.

  • The blue and the red one are going to split each other.

  • Now other molecules in the same sample will turn out to

  • have been labeled elsewhere.

  • For example, so that proves that those entered as a unit.

  • Others will have come in there, and those will also be

  • a double doublet, show that they came in as a unit.

  • And if we go over to the other end of the molecule, those

  • will be a double doublet.

  • And those singles will be a double doublet.

  • Great.

  • What's going to be different?

  • Suppose we had that molecule where we had a methide shift.

  • So the carbon got away from its original neighbor.

  • Now it won't be next to a C-13 and it'll be just a pair of

  • singlets instead of a double doublet in a

  • proton-decoupled spectrum.

  • And the same thing will be true there.

  • So this double labeling absolutely proves that these

  • rearrangements took place.

  • So those two, incidentally, are

  • labeled in the same molecule.

  • But they're not adjacent to one another anymore, so they

  • don't show splitting.

  • And those are both labeled, the ones that are adjacent to

  • one another, but not in the same molecule.

  • So they don't do splitting either.

  • So this dilute double label experiment enhances the same

  • 12 peaks as the single label experiments did, the two

  • single label experiments.

  • But only eight of them show spin-spin splitting because

  • their bonds stay intact.

  • So this strongly confirms the story I told you before about

  • the rearrangement scheme.

  • OK, so this is the power of correlation.

  • And let's look at a different kind of power of correlation

  • in what's called two-dimensional NMR spectroscopy.

  • Now here's a two-dimensional NMR spectrum.

  • And the way it's collected is rather technical and we don't

  • have time to go into it.

  • But it shows that there are two different frequencies

  • being used.

  • And if you go along the diagonal of this plot, those

  • two frequencies are the same.

  • So what you see there, when you have just a certain

  • frequency going in, is the normal spectrum you see along

  • the diagonal.

  • But this is the spectrum of a protein.

  • So it's got a whole bunch of amino acids linked

  • together like that.

  • And we're in the range of the spectrum between 6 and 9,

  • where the N-H protons show up In this

  • particular kind of molecule.

  • So there are a whole bunch of them in this long chain and

  • they all overlap with one another, and it's very hard to

  • tell what's going on.

  • But if you don't have these two frequencies you're using

  • the same, you're looking at one while you're irradiating

  • another one.

  • And that affects the intensity of the one you're looking at.

  • That's this nuclear Overhauser enhancement that we talked

  • about, where you irradiate one, other things that it's

  • coupled to will change their intensity.

  • So what we're going to do is plot on the off-diagonal how

  • one frequency is influenced by having other frequencies in

  • there, in its intensity.

  • So that's called NOE, or nuclear Overhauser

  • enhancement.

  • And it comes, incidentally, by through-space magnetic

  • interaction.

  • Not the stuff like we talked about coupling where it goes

  • through bonds.

  • This works through space.

  • So it works if the two protons you're looking at-- the one

  • you're looking at and the one you're tickling to see if it

  • influences this one--

  • if they're within about 6A of one another.

  • So as I said, this is mostly these N-H protons.

  • And for one thing, one advantage of having these off-

  • diagonal peaks is they're much less congested than what's

  • along the diagonal than the normal spectrum.

  • So for example, we're looking here at the NH proton that's

  • at 7.25 parts per million.

  • But you can see that it interacts with

  • several other protons.

  • It's within 6A of signals that are at

  • 8.9, 8.3, 8.25, 7.7.

  • Now you don't measure the exact distance, all you know

  • is that it's close.

  • But this allows you to make a three-

  • dimensional map of the protein.

  • And the way you do it, is first you have to know which

  • NH goes with which R group.

  • So you see that that NH--

  • the one you're looking at, the red one--

  • is close to that blue one.

  • So it'll influence it.

  • But the blue one is then close to Hs that are in the R group,

  • so you can tell which R group that that blue H is nearby,

  • and which one the red H is near.

  • So the red is near the blue is near the R. So you know which

  • NH is near which R. And now the way you make a map, then,

  • in this way--

  • Professor Loria, who does this kind of

  • thing, uses this analogy--

  • it's like making a map of a town if all you have is a

  • telephone and a telephone book.

  • Sometimes there's a front page that has a map of the town,

  • then that's easy.

  • But if you don't have the map of the town, then what you do

  • is you just call someone at random.

  • Say, Hi, how are you today?

  • Blah blah blah.

  • Could you tell me who three of your near neighbors are?

  • And they tell you three of their near neighbors.

  • Then you call those people, and you say Hi, can you tell

  • me who three of your near neighbors are?

  • And they tell you.

  • And then you call the next one, Who are three of your

  • near neighbors?

  • So once you know who is near what, then you can see that

  • there will be a map that will allow you to figure

  • out what's near what.

  • Especially if you have some help, because you know in the

  • case of the protons, they have to fit with normal bond

  • distances, normal bond angles, and so on.

  • So if you do molecular mechanics together with this

  • knowing what's near what, you can get a very good three-

  • dimensional structure of the protein.

  • And the nice thing about this, from the point of view of

  • people who are actually working with proteins-- what's

  • the best way of getting the three-dimensional structure of

  • something like a protein or anything?

  • X-ray, but you need a crystal.

  • For this, you don't need a crystal.

  • So you can find the structure that way.

  • So with molecular mechanics constraint, you get the 3D

  • structure without a crystal.

  • Now here's a different kind of 2D correlation with NMR. It's

  • correlation not in space, but in time.

  • So this is the spectrum of a cation, that cation that's

  • shown at the right.

  • Notice it's benzene with methyl groups all around it,

  • with one extra methyl group and a positive charge.

  • Now in fact, you can get this in solution as long as you

  • don't have anything that reacts with cations in there,

  • that reacts with low LUMO, so you need very, very non-basic

  • solvents to do this.

  • So again, the diagonal running a different way this time--

  • the increase in parts per million is down to the right,

  • rather than what we usually see, increasing to the left.

  • But anyhow, it's color coded, according to these peaks that

  • show what they are in the molecule.

  • So the scale is slanted and backwards.

  • So you see that if you look at the resonance structures--

  • that one, that one, that one--

  • you can see that the blue and the green are near where

  • there's positive charge in the resonance structure.

  • So they're far downfield, very deshielded

  • compared to the others.

  • Now this can rearrange.

  • A methide can shift, so one of the red methyls on the top

  • shifted to the right.

  • So now you've changed which methyl groups

  • are in which locations.

  • So now the red one at the top, this one and this one are now

  • C, and will be shifted way downfield.

  • This one will be shifted downfield--

  • what was blue before in the D

  • position, and so on.

  • Then this blue one, which was downfield before is now an A,

  • it's up here.

  • So we've changed the locations, the local

  • frequencies, the local fields.

  • So that's a methide shift.

  • But is it a 1,2 methide shift?

  • Because it's possible that the methyl, instead of going just

  • to its neighbor, would move to the center of the ring, up

  • above the center of the ring and then it

  • could come down anywhere.

  • Wouldn't have to come down adjacent to where it started.

  • So you could have a 1,anywhere shift.

  • Now you can distinguish that and measure the rates with

  • this NMR spectrum, this two- dimensional NMR spectrum.

  • So again, we're seeing when one

  • frequency influences another.

  • Now if a given set of protons moves from one position to

  • another, then having been irradiating one influences

  • what happens where it ended up.

  • Because those same protons are now here.

  • So here you see off-diagonal peaks where

  • one has become another.

  • So we can see here, for example, that the proton that

  • was in position C becomes A. And that's what happened in

  • the blue, here, when we had that 1,2 rearrangement.

  • We can also see that D becomes B, and that you can see

  • happened here.

  • And you can see that B becomes C. But you notice there's not

  • A becomes B, A becomes D, or C becomes D. So you don't have

  • peaks for some of these patterns that

  • you could have done.

  • Now if you go through a whole sequence of the these, then

  • you can start any place and end any place else.

  • But it takes a longer time.

  • So depending on the time scale of this thing, you can tell

  • which things happen quickly.

  • And what happens quickly is a 1,2 shift, not

  • a 1 anywhere shift.

  • So again, that's information on mechanism.

  • It comes from two-dimensional NMR spectroscopy.

  • And it's proton versus proton correlation in time.

  • That's, incidentally, just another visualization of the

  • same map I showed before, that shows the peaks on the

  • diagonals and the off- diagonal peaks.

  • OK, now we've finished with NMR. We'll refer to it again,

  • but we've finished discussing it.

  • And we'll go on to the next important reaction, which is

  • electrophilic aromatic substitution.

  • We talked about electrophilic things before.

  • Remember, electro philic addition to an alkene.

  • Now you might think benzene is an alkene, so you can get

  • electro philic addition to the double bonds of benzene, but

  • you don't get addition.

  • You get substitution.

  • So here's an example of substituting

  • deuterium for hydrogen.

  • You put benzene and concentrated sulfuric acid

  • that has deuterium in it, and you get deuterium first in one

  • position then, of course, in other

  • positions in the benzene.

  • So you can go all the way to C6D6, that's how you prepare

  • deuterated benzene if you want to use it for a solvent for

  • something, for example.

  • You can use other electrophiles as well, not

  • just D+, you can use NO2+, Br+, sulphur+-, R+

  • and RC=O+.

  • All these things can substitute for hydrogen, and

  • we're going to talk about those.

  • Now, the mechanism is that the deuterium, D+, the

  • deuteron, adds to the ring.

  • But that, notice, changes the aromatic ring into just a

  • regular pentadienyl system.

  • So you lose a lot of resonance stabilization.

  • It happens, but it's not very stable and it'll easily come

  • off again if there's anything that can take a proton.

  • But if you have a very, very non-nucleophilic, non-basic

  • solvent, nothing to take it off, then it can stay on and

  • in fact it's observable.

  • You can get spectra of it.

  • And so in electrophilic addition to alkenes, the next thing that

  • happens is a nucleophile would add, and you'd

  • lose the double bond.

  • But here it's easier to lose the deuteron or, more

  • interestingly, the proton.

  • Losing the deuteron would take you back to

  • the starting material.

  • But losing the proton, the H, then takes you to the

  • substituted product.

  • And that's easier than adding a nucleophile, because you get

  • back to the aromatic stabilization, 30

  • kilocalories.

  • And remember, we just saw a spectrum

  • that looked like that.

  • When you added not H, but CH3+, in this case, to the

  • hexamethylbenzene.

  • So it's possible to see these things in NMR. So there's the

  • benzene, the D+ comes on, we get that.

  • Notice that it converts the aromatic ring into a chain,

  • destroying the aromaticity.

  • Let's look at the orbitals that are involved, just using

  • this simple Huckel program.

  • So benzene, we've looked at the orbitals of that on the

  • left, and pentadienyl we looked at before, too

  • and we see it on the right.

  • Now, the most interesting orbital in pentadienyl is the

  • one that's indicated in yellow over there.

  • The highest of the low three orbitals, and it's at 0.

  • And you know it's at 0, the same energy as a P orbital.

  • How do you know the energy of that molecular orbital,

  • consisting of three P orbitals is the same energy

  • as a normal P orbital?

  • Remember, if you make a double bond and put the two next to

  • one another, they overlap, one goes down, one goes up, it's

  • not to same energy.

  • But this one has the same energy as

  • an isolated P orbital.

  • How do you know that?

  • Notice how many nodes are there in the lowest one, the

  • one that's about minus two?

  • How many nodes for the pi orbitals?

  • One, two, three, four, five P orbitals connected in a chain.

  • How many nodes in the lowest combination?

  • STUDENT: Zero.

  • PROFESSOR: Zero, except for the node of the p

  • orbitals, of course.

  • So zero.

  • Then the next one, the one that's at about -1?

  • STUDENT: One.

  • PROFESSOR: One.

  • How about the ours?

  • STUDENT: Two.

  • PROFESSOR: OK, and so there are two nodes.

  • Red, white, red on the top.

  • Now, how do we know the energy of that?

  • How much overlap is there?

  • STUDENT: Zero.

  • PROFESSOR: None, they're not next to one another.

  • So it's the same as an isolated p orbital at the

  • Huckel approximation.

  • OK, so that one is very interesting, that SOMO.

  • It's non-bonding in the case where we have a single

  • electron in there, which is what's shown.

  • So that's the locus of that odd

  • electron in the free radical.

  • If we have 5 pi electrons, the fifth, the one that's in a

  • pair, is in those three positions.

  • Now suppose we're interested in the cation with five P

  • orbitals in a row.

  • So what we're going to do is just take that electron out.

  • And where we take it out is now going to

  • be a positive charge.

  • So that's the same as if we put another electron in, in

  • the anion, that would be the locus of the negative charge.

  • And in the cation, the locus of the positive charge.

  • And that's the same thing we'd get by drawing resonance

  • structures that show those same positions having charges

  • or free radicals in them.

  • So those are the positions of positive charge.

  • And we get the same thing drawing resonance pictures.

  • Now consider the influence on rate of substituting not

  • benzene, but a benzene that already has

  • some X group on it.

  • Now, a very common thing to have studied with this was

  • nitration, because it's such an easy reaction to do.

  • It's a very vigorous reagent, NO2+.

  • You make it by just having a mixture of nitric

  • and sulfuric acid.

  • You see how you get NO2+ from that mixture?

  • It's quite simple.

  • You get a proton from sulfuric acid, and it protonates the

  • OH, and then you lose water.

  • So you've got NO2+.

  • So it could be the thing that comes in and then a proton

  • comes off from the same position, and you've got the

  • substitution.

  • But now, when you have X in there, you can get three

  • different isomers.

  • Ortho, meta, and para, they're called.

  • Notice, incidentally, that NO2+ is an interesting

  • molecule because it's exactly the same as CO2, except it

  • has one more proton in the nucleus of the central atom.

  • Nitrogen instead of carbon.

  • So CO2 has a pair of electrons in that double bond, a pair of

  • electrons on the oxygen that are in the y orbitals,

  • pointing up and down.

  • It's also got x orbitals coming in and out that have

  • the same kind of thing.

  • But if we consider just the y orbitals, that'll be the LUMO.

  • There are two pairs of electrons, so the one with no

  • nodes has two electrons, the one with one

  • node has two electrons.

  • This is the one with two nodes, no electronics.

  • So that's the vacant orbital, the LUMO of NO2+

  • just like CO2.

  • So NO2+- attacks, and we can measure the rate.

  • Now what we're going to do to measure the rate--we could do

  • careful work with a stopwatch and so on, but there's a

  • clever way to do it.

  • Which is to compare two molecules by doing them both.

  • Put both in there, and do the reaction for just a little

  • while and then kill it.

  • If you add base, you won't have the strong acid and the

  • reaction will stop.

  • And now see how much of each one becomes product.

  • So we get relative rates that way.

  • So we're doing relative rates compared to hydrogen being 1,

  • that is, benzene when X is hydrogen.

  • If X is methyl, it's 25 times faster.

  • That is, if you have 25 times as much benzene as toluene--

  • the one with methyl on it-- you'll get equal amounts of

  • the products at first, until you've

  • used up all the toluene.

  • So it's 25 times faster.

  • So methyl helps the reaction.

  • OH really helps it, if you use phenol it's 1,000 times faster

  • at getting the sum of all the products.

  • Now why are those so fast?

  • Because electron donation eases the formation of the

  • carbon cation intermediates.

  • Or, looked at a different way, the carbon cations are stable

  • because there are other electrons that become

  • stabilized when you get a cation there.

  • For example, the unshared pair on oxygen.

  • OK, now that's because we have positive

  • charge at those locations.

  • When X is there, that helps out.

  • Now if you put NO2 as X, so it's already there, then it's

  • 6 times 10-(-8), it's 10- (-7), 10

  • million times slower.

  • And the same if you have trimethyl ammonium with its

  • positive charge.

  • You don't want to make those cations if there's another

  • cation there, or if there's an NO2 group there.

  • So those are electron withdrawing, and they retard

  • the formation of cation intermediates.

  • The most interesting, as many times, is the halogen.

  • The halogen is slower than benzene, but not much slower.

  • Only 30 times slower.

  • And that's because both things are happening with halogen.

  • It's withdrawing electrons in the simga bond, but it has an

  • unshared pair that donates back, so

  • that it mostly cancels.

  • Why is NO2 electron withdrawing when OH is

  • electron donating?

  • O is more electron negative than nitrogen, so you might

  • think that OH should be more electron withdrawing.

  • An unshared pair on nitrogen is higher energy than the

  • unshared pair on oxygen

  • should be more electron donating.

  • So this seems, at first glance, to be sort of nuts,

  • that OH is 1,000 times faster and NO2 is 10

  • million times slower.

  • But it's pretty easy to understand when you look at

  • the orbitals that are involved.

  • So in that case of oxygen, there's this unshared pair on

  • the oxygen, which can overlap with the p orbitals on the

  • benzene ring.

  • It turns out, indeed, that the energy of the oxygen unshared

  • pair is about the same as pi electrons of a

  • carbon-carbon bond.

  • So that the unshared pair on oxygen has about that energy,

  • about -1 on the scale we're talking about here.

  • OK, so that's unusually high, as pi are, unusually high.

  • So when you have a vacant orbital next door, those

  • electrons get stabilized.

  • So it's a high HOMO, a good overlap with phenyl, and

  • therefore OH can donate its unshared pair.

  • OK, that's good.

  • How about NO2?

  • Notice that NO2 is allylic.

  • It's got three p orbitals on each

  • oxygen and on the nitrogen.

  • So it's going to have a low occupied molecular orbital

  • that gives good overlap with phenyl.

  • So that one is able to donate electrons because it has good

  • overlap, but it's not willing to share its electrons.

  • Its electrons are very low in energy.

  • They're not going to be much stabilized by having a vacant

  • orbital next door.

  • So you're not going to get anything out of that,

  • from the low one.

  • The next orbital of the NO2 allylic system has a node in

  • the middle.

  • That one.

  • That one has the energy shown in yellow there, it's going to

  • be very good.

  • It's got a good, high HOMO.

  • As high as the p orbital on oxygen, so it should be good

  • at donating electrons, at having its electrons

  • stabilized by the vacant orbital in the benzene ring.

  • But it's not, why not?

  • Can you see?

  • Why doesn't this orbital be stabilized by mixing with

  • orbitals that are on the benzene ring?

  • Because there's no overlap.

  • It doesn't have any size where it's adjacent

  • to the benzene ring.

  • So it has no overlap, so it's willing to give its electrons,

  • but it's not able because there's no overlap.

  • And finally, you have the third of the allylic

  • orbitals, that one.

  • Which is a low LUMO.

  • And it has good overlap with phenyl.

  • So it withdraws electrons, it's willing to take electrons

  • because it's unusually low, and it has overlap so it's

  • able to accept electrons.

  • So that's why nitro is electron withdrawing when OH

  • is electron donating.

  • it's because of that node in the middle that makes it

  • unable to give electrons to the benzene ring.

  • So NO2 is a pi accepter, whereas OH is a pi donor.

  • OK now, we just looked at the rates of these reactions, of

  • forming all three products together.

  • But we might be interested in the individual products,

  • ortho, meta, and para.

  • So let's look at that.

  • In the case of hydrogen, we'll call each of these 1.

  • And now we'll do mixtures of hydrogen with others--

  • but now we'll analyze the ones that already have an X in

  • them-- we'll analyze whether it's orth, meta, or para, and

  • how much of each one.

  • Now if it's methyl, you see that two of them are very fast

  • but one of them is not fast, compared to what it was when

  • only hydrogen is on the ring.

  • How do you understand the two that are fast?

  • Why are those fast when X is methyl?

  • Linda, do you have an idea?

  • When X up here is methyl, this is good and this is good.

  • 39 and 46.

  • But this one is not any better.

  • Methyl wants to be where the cation is.

  • So it's especially good here, and especially good here.

  • But not here.

  • There are only hydrogens here, here, and here.

  • And it has the same rate as hydrogen.

  • OK, that's fine.

  • If you look at t-butyl, now again it's fast over here.

  • But it's not fast here.

  • So why don't you get this product when you have the R

  • group here, when it's t-butyl?

  • You get it when you have methyl, but

  • not when it's t-butyl.

  • Nathan, you got an idea?

  • STUDENT: Steric hindrance.

  • PROFESSOR: It's steric hindrance, it's too big.

  • It's hard to put this extra group here when you have

  • something big there already.

  • OK, so that's good.

  • Now there are two things about a substituent, X. Does it make

  • it go faster, or does it make it go slower, is it activating

  • or deactivating?

  • So it's activating, these are faster than benzene.

  • And how is it directing, what does it direct?

  • Its ortho and para, for the reason we saw.

  • Except that in the t-butyl case, it's not very much ortho

  • because of steric hindrance, as Nathan told us.

  • OK, so those are electron donating, and like to be

  • ortho/para, other things being equal.

  • Now here's one that's the same size as that neopentyl...

  • or t-butyl group,

  • the carbon with three methyls on it.

  • But notice it's very much slower.

  • And we saw that before.

  • But it's selectively slower, it's slowed much more here--

  • 0.6 times 10-(-8) than it is here...

  • 3 times 10-(-8).

  • So it's slow to get product, but when you get product this

  • is the one you get.

  • Meta.

  • You don't get ortho and para.

  • Especially you don't get ortho because

  • it's sterically hindered.

  • But you see, that's the same thing.

  • If you want to form these things when X has a plus

  • charge, don't put it here and don't put it here.

  • If you're going to do it, put it there.

  • Make it as far away as possible.

  • So that's meta.

  • And the same thing is true of nitro.

  • It's slowed down everywhere.

  • But it's not nearly as slowed down meta as it is here-- it's

  • slowed 10 times more here for ortho and whatever--

  • 300 times, right?

  • Yeah, 300 times slower when it's para.

  • And if you have an ester group, the same

  • kind of thing is true.

  • Here it's 10 times slower here than here, and twice slower

  • here than there.

  • So those are deactivating, for the reasons we talked about

  • before, and they're meta directing.

  • But again, halogen, is especially interesting.

  • Notice it's slowed down, it's deactivating, but it's

  • ortho/para directing.

  • So it shows both these characters, deactivating and

  • ortho/para.

  • Why?

  • It's deactivating because of the electron

  • withdrawal, which is sigma.

  • But if you're going to make it, at least make it where the

  • halogen's unshared pair can be stabilized by the vacant

  • orbital next door.

  • Now, how can you make the reaction work better, the

  • electrophilic substitution?

  • You can make the electrophile better.

  • So C+ would be a great nucleophile,

  • but you don't buy Cl+,

  • you have Cl2.

  • So what's the low LUMO here, in Cl2?

  • I've asked this maybe 15 times before in lecture.

  • STUDENT: Sigma*

  • PROFESSOR: Sigma*, OK.

  • But it's not nearly as low as Cl+.

  • One way to make it better is to treat it with a Lewis-acid

  • catalyst, like AlCl3, which has a vacant orbital on

  • aluminum, which can attack an unshared pair on the

  • chlorine like that.

  • And now you have a minus formal charge on the aluminum

  • and plus on the chlorine.

  • It's a better leaving group.

  • So now you can do it better.

  • So Cl2 won't be nearly as reactive as AlCl5.

  • Here, we can compare their LUMOs and see it's essentially

  • the same LUMO.

  • But it's lower in the case of AlCl5, so we can do this

  • reaction and it'll behave as if it were Cl+.

  • If you look at the surface charge, 0.29 positive there-- or

  • 29 there, and 38 on the one with AlCl3.

  • So it's a better electrophile.

  • So adding a Lewis acid makes these things work better.

  • So the leaving group is then AlCl4-,

  • rather than Cl-.

  • And remember before when we said that the rate of addition

  • to alkenes of bromine is, in some solvents, is second order

  • in bromine.

  • Because that second bromine molecule makes the leaving

  • group Br3-, rather than Br-.

  • It makes the reaction better, exactly the same idea here.

  • Another way to make the reaction better is to activate

  • the nucleophile.

  • So salicylic acid, which was studied by Kolbe the same

  • guy who excoriated van't Hoff about stereochemistry

  • invented a way to make salicylic acid.

  • So you start with phenol, you react it with CO2, which has

  • that vacant orbital we talked about, like

  • the one in NO2+.

  • So you could imagine that reaction.

  • Just like NO2+, except that CO2 isn't nearly as low

  • as NO2 plus with its extra proton.

  • But you can make the nucleophile better.

  • In the previous slide, we saw making the

  • electrophile better.

  • So here we have the unshared pair on oxygen,

  • which helps us out.

  • But we can make it ever so much better by putting base in

  • and making it a higher unshared pair.

  • Now you have a big push to do it, and you get salicylic

  • acid, which where aspirin comes from, from

  • putting CO2 on phenol.

  • Now you could imagine the same thing for NH2.

  • It's a higher unshared pair than the

  • one on oxygen, normally.

  • So it would do this pushing too, and you

  • could nitrate aniline

  • aminobenzene, it's called

  • aniline.

  • And in fact, this was the proposed and, in fact, used in

  • 1944 as a rocket fuel

  • a mixture of nitric acid and aniline.

  • And there's a little story I'm going to tell here about a

  • good friend of mine, Dennis Tanner, who was a professor

  • for many years at the University of Alberta.

  • But when he was a post-doc, he was at Stanford.

  • And he's a very hard worker.

  • And he was working in the lab on the Fourth of July.

  • Not many people were working on the Fourth of July, but

  • several people were, and they passed each other in the hall,

  • and they decided they should have some fun, too.

  • So they decided that each person would put on a display

  • out in the parking lot.

  • So Dennis had heard about this being a self-igniting rocket

  • fuel, so he thought it would be fun to mix to nitrate

  • aniline, to do the nitration of aniline.

  • So he got concentrated nitric acid and put it in a beaker,

  • and he got aniline and put it in a separatory funnel and

  • hung it above the beaker.

  • And he crouched down and grabbed the thing and twisted

  • it, and started to run.

  • But the instant it hit, before he got more than two paces,

  • there was an enormous fireball that burned off his eyebrows.

  • So indeed, this is a very, very active nitration

  • reaction, as Dennis proved.

  • But you can tame it. if you put an acetyl group on the

  • aniline, then you lower the energy of the unshared pair.

  • The pi* of the carbonyl mixes with it and lowers it.

  • Now you don't have so much of a push anymore.

  • And now, the nitration still takes place.

  • It's still ortho/para directing, but it doesn't burn

  • your eyebrows off.

  • So the acetylation of aniline makes this nitration

  • controllable.

  • So you can start with aniline, react it with acetyl chloride

  • and pyridine, which is often used as a catalyst for putting

  • acetyl groups on, and we'll talk about that later.

  • And now you can do the nitration.

  • And you see it's ortho/para directing, not meta directing.

  • So the unshared pair is still doing its work, but not as

  • dramatically, as in Dennis's case.

  • But then if you really wanted to do the nitration of

  • aniline, now you take the acetyl group back off again.

  • So you have the product, so you tamed the nitration.

  • Now I just want to introduce the biographical part of this,

  • and we'll talk about the rest of it next time.

  • So the Friedel-Crafts reaction is one of the most important

  • electrophilic substitution reactions.

  • In fact, it's one of the most important reactions

  • altogether.

  • And it was named after Charles Friedel

  • and James Mason Crafts.

  • Friedel was a Frenchman.

  • Notice he's seven years older than

  • Crafts, who was an American.

  • So this is a great example of where America stood in

  • international science in the last part of the 19th century.

  • So these are the publications of James Mason Crafts from

  • 1862, when he was a student, until 1915 when he died.

  • So some of his papers were on organic chemistry, and they

  • didn't have Friedel as a co-author.

  • And at different ages,

  • he published a number as a student there, as you'll see.

  • But then kept publishing.

  • Now here's his trajectory.

  • He was born in Boston, he studied at Harvard.

  • But then to learn real science, he went to Europe.

  • First he went to Germany for just a short period.

  • He worked with Bunsen, actually.

  • Then he went to France and worked with Wurtz, the

  • successor of Dumas.

  • And he stayed there for four years.

  • And then he came back, he actually inspected mines in

  • Mexico for a year--

  • and California.

  • Then he went to Cornell--

  • which had just been founded--

  • to be the chemistry professor for a couple of years.

  • Then he went to MIT.

  • And then he was ill, but also tired of not doing research.

  • And he went back to Paris.

  • He went back for just two years, but he

  • stayed for 17 years.

  • And now he was working with his pal, Friedel, whom he'd

  • met as a student in Wurtz's lab.

  • Then he came back to MIT and, in fact, he was president of

  • MIT for two years around 1900.

  • Now here are his papers that are not organic chemistry.

  • They're about thermometry, about physical chemistry.

  • And notice he did most of the publication during this second

  • period in Paris when he was doing only research.

  • But if you look at his papers with Friedel, there are an

  • enormous number.

  • During this period--when he was in America,

  • he published almost nothing.

  • When he was in Paris, he published a lot, both as a

  • student when he started working with Friedel, and

  • especially after he came back in these 17 years.

  • So he published 100 papers during these years, 99 papers

  • during those years.

  • So it wasn't possible to do the kind of research in this

  • country that you could do over there.

  • And then in 1877, just after he had gone back to recover

  • and work with his pal Friedel at the School of Mines, they

  • published a really important paper and most of those red

  • publications after that are about this

  • Friedel Crafts reaction.

  • And they got into it because they had been students

  • together working with Wurtz.

  • And we'll talk about this then next time.

J. MICHAEL MDBRIDE: So, we'll finish up on

字幕與單字

單字即點即查 點擊單字可以查詢單字解釋

B1 中級

24.高階效應、動力學和核磁共振時間尺度。 (24. Higher-Order Effects, Dynamics, and the NMR Time Scale)

  • 76 7
    Cheng-Hong Liu 發佈於 2021 年 01 月 14 日
影片單字