Placeholder Image

Subtitles section Play video

  • J. MICHAEL MCBRIDE: So we've talked

  • about the main events in NMR

  • spectroscopy for determining structure.

  • That is, about the chemical shift and about spin-spin

  • splitting, that allows you to tell how many nearby

  • nuclei there are.

  • Today we're going to talk about some complications that

  • make the spectrum more complicated, and perhaps more

  • informative, in terms of high order effects, higher-order

  • effects in spin-spin splitting.

  • And then about using NMR to study, not structure, but to

  • study rates.

  • And we'll do some C-13

  • labeling stuff there.

  • So spin-spin splitting with other nuclei besides protons

  • next door, and we actually saw some of this last time with

  • deuterium splitting, with C-13.

  • splitting.

  • For example, if there's a signal like this and you blow

  • it way up by 30 times, then you can see that.

  • What gives these things that are so far apart 126 Hz.

  • Do you remember this from last--

  • what are the size of splittings you normally see

  • for protons that are on adjacent carbons?

  • They're the order of 7 Hz, right?

  • Here we have got 126.

  • What's the big thing that's going on?

  • What makes coupling so big?

  • It's C-13-H. In fact, 99% of the sample--

  • if you haven't specifically put C-13s

  • in there, are C-12

  • but about 1% are C-12.

  • So you see that.

  • Now here's a paper from 1959, which first noted something

  • interesting about these C13-H coupling constants, that they

  • relate to the amount of s-character in the carbon

  • atomic orbital that's involved in the bonds.

  • Remember, we said that the information has to be

  • communicated through the electrons because the action

  • of magnets through space

  • if you average over all angles of rotation in a liquid--

  • goes to zero.

  • So you communicate through the bonds.

  • The nice thing about...

  • that it's so strong for JC13-H, is that

  • C-13 has just one bond to get to the H. So it's a very short

  • path to communicate.

  • Now, but you remember, also, that to do that-- so that it

  • doesn't average out as you tumble--

  • you have to have the electrons in an s orbital.

  • So the hydrogen uses its s orbital, that's all it's got

  • essentially.

  • But carbon can use s or p. So the hybridization, as they say

  • here, relates to the coupling constant.

  • Because to communicate that electron must

  • be on the C-13 nucleus.

  • So this is a table from that paper.

  • It gives the coupling constant in units of per second.

  • Those were then changed to be called Hz, nowadays.

  • So here are a bunch of different compounds and

  • their JC13-H coupling constants.

  • If we make a histogram of them, you see that there's a

  • group between 120 to 140

  • one between 150 and 170

  • and another group down around 250.

  • So there are different groups involved here.

  • There are three different populations, we could draw the

  • lines between them.

  • And then if we look to see what the carbons are, that

  • bear these hydrogens that show that splitting, we could look

  • first at the very big ones down there.

  • And what's special about those carbons that are

  • splitting the hydrogens?

  • They're triple bonded.

  • So sp hybridization

  • 50% s in the bond to hydrogen.

  • The ones here, the green ones, are in aromatic systems, so

  • they have sp squared hybridization.

  • Except for cyclopropane.

  • Why would it have such a high coupling constant?

  • Why would it have a lot of s-character

  • in the bond to hydrogen

  • the carbon that's in the cyclopropane?

  • Because its bonds normally are tetrahedral-sp cubed

  • but when you try to make a smaller angle, more

  • p-character in the bonds to the other carbons, therefore

  • more s-character coming out to the hydrogens.

  • So we can actually see that hybridization change in the

  • NMR. And then these other are normal methyl groups, I speak

  • sp cubed hybridized carbons.

  • So here in NMR, we have direct evidence for this

  • hybridization stuff we talked about last semester.

  • So it's sp, sp squared, sp cubed.

  • Now here's a subtle point.

  • Perhaps we shouldn't be spending time on it, but it's

  • so interesting. And I suspect the curious among you may have

  • wondered about this.

  • Because we talked about coupling

  • goes through the bonds.

  • So from H to C to C to H. How about H to C to H?

  • That's a shorter path.

  • So methane, if you're looking at one hydrogen, how many

  • other hydrogens would you think it would be coupled to?

  • Three.

  • And they could be in 1:3:3:1.

  • So you should see a quartet for a proton of CH4.

  • But you don't, you just see a single sharp peak.

  • So where did the splitting go?

  • And this is related to what we talked about last time, where

  • a book said the splitting constant is zero.

  • And I said that's not true, the splitting constant is

  • high, maybe 10.

  • But it's not observable.

  • So why is it, why don't you see quartets for CH4?

  • OK, so here's one of these spectra-- actually if you go

  • up here and you can click on that and get taken to the

  • website of Chem 220 where they have problems. I'll work one

  • problem later in the class, if we have time.

  • But there are others you can practice on there.

  • So this is a compound, and I'll tell you its answer.

  • It's cinnamic acid.

  • It's related to the adelhyde related to this acid,

  • cinnamaldehyde is the smell of cinnamon.

  • So what do we expect when we look at the proton spectrum--

  • this is the proton spectrum of it-- what do we expect?

  • Well, notice that we can draw resonance structures, which

  • will put positive charge on the carbon with that hydrogen,

  • so remove electron density from that hydrogen.

  • Or from that one, or that one, or that one.

  • So those all should be shifted downfield relative to where

  • they would otherwise be, relative to the ones that

  • aren't blue.

  • When we look at the compound, then, we see these really

  • complicated peaks in here. We throw up our hands trying

  • to understand that.

  • And then we see here two doublets

  • really nice, sharp, clean things.

  • So what do you think the two doublets are out here?

  • Let's ignore that bit for the time being and talk about

  • these two doublets.

  • Which protons do you think those come from?

  • Which one is this?

  • OK, let me start going around the compound.

  • The OH is way down further.

  • We're not showing that.

  • OK, so I go down.

  • I come to this one.

  • What do you think that is?

  • STUDENT: The one near the COOH.

  • PROFESSOR: No, that's the one up here.

  • This is where double bonds normally come.

  • Then you come here, and that one has a hydrogen on double-

  • bonded carbon, but it's down here because it's has lost

  • electron density.

  • Remember we drew a resonance structure where the blue

  • ones were more positive.

  • And then here, we have hydrogens that are on the

  • benzene ring, shifted down this way because of the ring

  • shift, the aromatic ring current that we

  • talked about last time.

  • And now there's a group of three and a group of two.

  • These are positive and are shifted down.

  • That's the group of two.

  • The group of three here is these two, plus this one.

  • Which would be shifted down because it's positive, but up

  • because it's far from this sp squared carbon.

  • Let's not worry about why that is.

  • But at any rate, that's the benzene ones.

  • But that's not what we're going to talk about here.

  • What we're going to talk about is the clean

  • signals. Here and here.

  • Now this is measured with a magnetic field of 7.05 Tesla.

  • That means it's a 300 MHz spectrometer.

  • So instead of using parts per million down here, we could

  • change to Hz if we wanted to, change

  • to a frequency scale.

  • And in terms of frequency, if we call zero the middle, then

  • that's lower-frequency to the top, higher-

  • frequency to the bottom.

  • It turns out that the ratio of the chemical shift difference,

  • delta--measured in Hz, not in parts per million

  • we want to be comparing apples with apples

  • here, Hz with Hz--

  • so the chemical shift from here to here is much bigger

  • than the coupling constant, J. It's 24.4 times as large, and

  • that's the pattern we see.

  • Now suppose we used a smaller magnetic field.

  • What's going to happen?

  • The chemical shift in parts per million

  • will stay the same.

  • But it'll be the same fraction of a smaller frequency. If we

  • use a smaller field.

  • The precession frequency gets smaller.

  • So if we measure in Hz, as we're doing here, they'll come

  • closer together

  • but J will stay the same, because it doesn't

  • depend on the field.

  • The magnets are not dependent on the

  • field, they're coupling.

  • So what will happen?

  • These two doublets will move together.

  • So at 2.89, it turns out the ratio is 10.

  • Or at point.

  • 0.86 Tesla, it would be 3.

  • Notice, incidentally, that this would be a much cheaper

  • spectrometer because it would have only 1/10 as big the

  • magnetic field as the other one.

  • But also, those other peaks are going to be in the middle.

  • Everything's going to be getting on top of one another,

  • and I'm just showing you what would happen

  • to this double doublet.

  • There it is at 0.29.

  • And finally, if we're 100 times smaller at 0.2 for the ratio

  • of chemical shift to J, measured in the same units, in

  • Hz, you get the purple pattern.

  • Now notice that the J is always the same, the splitting

  • between the two peaks

  • but what do you notice between the doublet?

  • What's curious about this?

  • I mean, I'm sure you all see.

  • Some of the peaks are disappearing and others are

  • getting stronger as they come together.

  • The outer peaks are getting very weak.

  • And in the limit where the chemical shift was zero, the

  • difference-- as in methane--

  • then the outer peaks disappear altogether and you

  • just see one peak.

  • Now notice that that part in the original spectrum was a

  • mess, because all those hydrogens of the benzene ring

  • have similar chemical shifts, and they're all coupled with

  • one another so it's more of this same thing

  • and all on top of one another, and very complicated. So we'll

  • forget that.

  • Let's think about what causes this change.

  • This is a so-called higher- order effect

  • not the simple thing, that you have a chemical shift, count

  • the neighbors and get 1:2:1, 1:3:3:1, 1:4:6:4:1

  • kind of thing.

  • So here we're plotting the energy in a big applied field

  • of different nuclear spin states.

  • And we're talking about just two nuclei.

  • So they could both be with the field; that

  • would be low in energy.

  • Or one with and the other against, or one against and

  • the other with.

  • And then they could both be against the applied field.

  • So you'd have these different energies, right?

  • And that's the 100 MHz, that's the energy that it

  • would take to get from here to here by flipping the red spin,

  • would be 100 MHz.

  • But all these look the same, whether you do a red line and

  • flip the red arrow, or do a blue line and

  • flip the blue arrow.

  • They're all the 100 MHz or whatever frequency, or 300

  • MHz in that spectrum we were looking at before.

  • So we're going to see just a single peak at first glance.

  • But if we home in and magnify things so we can see small

  • differences, so enlarge this center portion by a factor of

  • 10 to the fifth so we can see the difference is just a part

  • per million in the difference between the environment of the

  • blue and the environment of the red proton.

  • Then as we expand it, we see that they're

  • not quite the same.

  • It's harder to flip the blue proton than it is to flip the

  • red proton.

  • The blue proton is in a higher field because of the chemical

  • shift difference.

  • And if we call that difference delta/2, now you'll

  • notice that that made the red lines shorter by delta/2

  • and the blue lines longer by delta/2 doing that.

  • So now we're going to see a chemical shift of delta.

  • One is smaller by delta/2, the other larger my delta/2,

  • the difference is delta, the chemical shift.

  • So that's the chemical shift.

  • And now suppose the spins interact not only with the big

  • blue field, but also with each other.

  • And just suppose that their coupling through these

  • electrons is such that they prefer to be anti-parallel.

  • You could have the opposite sign of the coupling constant,

  • they could prefer to be parallel.

  • But let's suppose they prefer to be anti-parallel.

  • So the ones where they're anti-parallel will go down in

  • energy, the ones where they're parallel will go up in energy.

  • So let's do that.

  • And now what happened to the red and blue peaks?

  • Notice that this red transition went to lower

  • energy, but this one went to higher energy.

  • And this blue one went to lower energy, but this blue

  • one went to higher energy.

  • Everybody see that?

  • So what's going to happen?

  • The red peak is going to split.

  • There's going to be a lower and a higher red peak, there's

  • going to be a lower and a higher blue peak.

  • So it's going to look like that.

  • And that's going to be J. If these shifts are J/4,

  • this transition got longer by J/2, this one got shorter

  • by J/2.

  • So the difference between them is J now.

  • In terms of the energy of the things interacting with one

  • another, that's where we get chemical shift difference, and

  • spin-spin coupling J.

  • OK, so that looks just fine, and that's what's called the

  • first-order spectrum.

  • That's what we've been talking about.

  • But there are higher-order effects, and I think you'll

  • find this cute.

  • Because J does something else besides just shifting these

  • levels, it also mixes this state with this state.

  • And it mixes them into a sum and a difference.

  • And of course we've seen things just like this before.

  • That's what it looks like in the case of magnetism.

  • The sum has the two pointing opposite directions along z,

  • but not along x. And the difference has them pointing

  • opposite directions altogether.

  • When you do that, when you mix this with this.

  • You get the difference, which

  • turns out to be a little lower in energy. And the sum, which

  • is a little higher in energy.

  • And of course, that's just like mixing a p-orbital on

  • carbon with a p-orbital in oxygen in a C=O double bond.

  • You mix the two, their energies don't match, so this

  • one has more of that in it, this one has

  • more of that in it.

  • Oops, I wrote this wrong, it should be red and blue

  • I'll change that.

  • Everybody with me?

  • So I didn't tell you exactly why it couples to

  • this, but it does.

  • So this shifts those energies.

  • Notice that that makes the reds a little bit smaller,

  • both of the reds smaller.

  • And both of the blues larger.

  • So it shifted the reds to the right and the

  • blues to the left.

  • So delta isn't exactly what you thought it

  • was going to be.

  • That is, the difference between the average of those

  • two isn't exactly the chemical shift, because of this higher-

  • order effect.

  • But there's something much more interesting that happens.

  • That one is non-magnetic, obviously.

  • They cancel one another.

  • But how about the other one?

  • It's not magnetic along z, but it's just as magnetic along x

  • as either of those other states

  • the top one or the bottom one is.

  • So it has this resultant.

  • So it as magnetic perpendicular to the main

  • field as either the one that was where they're both against

  • the field, or the one where they're both with the field.

  • So this one is magnetic, and that one's not magnetic.

  • That is, this one can interact with the magnetic field of

  • light, but this one can't interact with the magnetic

  • field of light.

  • So if we look over here, we see that this one is weakened.

  • It's not interacting with light.

  • This one is interacting even more strongly with light.

  • Now what that means is this one is really strong. And this

  • one is weak now, in terms of interaction with light.

  • So this influences the intensity of the transitions.

  • Now what it does, is it makes the ones that involve this

  • level-- that's this red one and this

  • blue one get stronger--

  • and these get weaker.

  • Like that.

  • So what's going to happen over here?

  • This one, which is the shortest of all, and this one

  • which is the longest of all, this one and

  • this one get weak.

  • Pardon me, this is the longest of all.

  • This is the shortest of all, this red one.

  • So those get weak, but the inner ones get strong.

  • So it changes to look like that.

  • And the more you mix them... They're separated by delta or

  • delta/2, they're mixed by J. So the bigger J is compared

  • to delta, the more they'll mix,

  • the more they'll be exactly like this, which is when

  • they're completely mixed.

  • And the stronger these will become and the weaker these

  • will become

  • So as the delta gets smaller and smaller, in terms of Hz

  • as you change the magnetic field, the outer peaks shrink

  • and the inner peaks grow.

  • So that's why methane has a single peak and you can't see

  • it splitting.

  • Even though the coupling is still there, you

  • just can't see it.

  • So the radio frequency light loses its handle on the

  • flanking peaks as delta becomes small compared to J,

  • as shown here.

  • So by increasing delta, measured in Hz, that is by

  • using an expensive, big field, you make things look more like

  • first-order spectra that are easy to interpret.

  • 1:2:1, 1:3:3:1, 1:4:6:1 and so on.

  • So that's the subtle aspect.

  • Now we're going to talk about how you use NMR to study

  • dynamics, to study rate

  • and what we mean by the NMR time scale, and

  • how decoupling works.

  • So we looked at this spectrum last time, this is a spectrum

  • of ethanol measured as it says over here by your own

  • TA, Phillip Lichtor.

  • If we look at these three peaks, we see there the OH,

  • the CH2, and the CH3, and we analyzed these last time and

  • saw that this is a triplet, this is a triplet, both being

  • split by this CH2

  • but this one is a quartet of doublets.

  • Now that was measured in DMSO solvent for a special reason.

  • Remember, that peak came from DMSO that had a little

  • hydrogen in it, as well as deuterium.

  • If we measured it in chloroform with a little bit

  • of acid in there--

  • or if Phil did, he measured this one as well--

  • notice the peaks shift a little bit.

  • Especially the OH peak shifts a lot, the others shift a

  • little bit, this one not too much.

  • What's much more interesting than these shifts is what the

  • peaks actually looked like, in terms of their splitting.

  • Now bear in mind this OH peak is here.

  • It's now over here, but I'm drawing it here.

  • So instead of a triplet, it's now a singlet.

  • And this one, instead of being a quartet of

  • doublets, is just a quartet.

  • No changes over here.

  • We've lost the splitting between the H and the CH2.

  • It's no longer split.

  • It's not split by the CH2, and the CH2 is not split by the H.

  • So what's going on when we did this?

  • In fact, what happens if we don't have the acid in there?

  • If we don't have the acid in there, this peak

  • moves way over here.

  • This one stays the same place, we're still

  • in chloroform solvent.

  • And this one stays the same place, but

  • the OH really changes.

  • But what's, again, much more interesting is what the

  • patterns look like.

  • They look like that now.

  • So this, instead of being either a triplet or a singlet,

  • is a really broad, low peak.

  • This one still looks a little bit like a quartet of

  • doublets, but it's much broadened.

  • This one is the same as it was before.

  • So something really funny is happening here.

  • Now let's look at the integrals.

  • So within experimental error, it's 1 to 2 to 3.

  • It's 0.96 to 2, to 3.04 in DMSO.

  • In CDCl3, it's 1.00 to 2.00 to 3.05, again with an

  • experimental error 1, 2, 3.

  • But the one that had a little bit of acid in it is much more

  • interesting.

  • It's 2 to 3, that's fine.

  • But this is bigger, 1.14.

  • So that peak-- this sharp peak here, the one that has

  • acid in it, is bigger.

  • There are more protons there than just one.

  • How did that happen?

  • It's an average of both the OH that's on the

  • alcohol and the H+.

  • Now let's see, how did that happen?

  • Well, H+ can go onto the alcohol, and then the H that

  • was on the alcohol can come back into the solution.

  • So we've exchanged those two H's.

  • And if that happens rapidly, there are H's moving all over.

  • Most of the time they're on alcohol but some time--

  • 14%--

  • there as H+ and what you're seeing is the average

  • of those, in this case.

  • So this is averaging.

  • The H's are zipping around not only between acid and alcohol,

  • but from one alcohol to another alcohol,

  • by way of the acid.

  • And what that means is, remember, the reason we see

  • splitting is that there are

  • these spin isomers.

  • There are different molecules in there.

  • Some molecules have the neighboring CH2 both up, some

  • have them both down, some have one up and one down.

  • But if you're zipping from molecule to molecule very

  • rapidly, you see the average. Which is nothing.

  • So when we have acid in there and it's averaging very fast

  • by going on and off and then going to different molecules,

  • then we see a sharp peak here, the splitting disappears.

  • Here, what's going on?

  • Well how about the ROH chemical shift, that thing?

  • It's the average of many hydrogen bonded structures,

  • because the hydrogen bonds keep getting made and broken

  • very rapidly.

  • So you average among many different hydrogen bonds, but

  • the H stays with the original molecule.

  • It's not going from molecule to molecule. It's just breaking

  • hydrogen bonds and going back and forth.

  • And how fast you do that and how much you do it depends on

  • concentration, on what the solvent is, on the

  • temperature, and so on.

  • But how about the spin-spin splitting between

  • the CH2 and the OH?

  • That can be the average involving H-exchange among

  • many molecules.

  • It's fast if you have H+.

  • It's slow in DMSO.

  • So in DMSO, you saw H's staying with their molecule,

  • and you see they're splitting.

  • In H+, they go from molecule to molecule fast. You

  • see the average.

  • And what happens if you have chloroform, but you don't have

  • the H+?

  • Some kind of averaging is beginning to

  • happen, it looks like.

  • So it's clear that if you have H+ in

  • there, it can zip around.

  • What happens if you're just making and

  • breaking hydrogen bonds?

  • Well, you could have three alcohols hydrogen bonded

  • together in a circle.

  • And they could trade which hydrogen, so this is a

  • mechanism for exchanging hydrogens from

  • one alcohol to another.

  • And then as the trimer comes apart and the molecules become

  • another trimer and do this exchange, another, another,

  • another, this is a mechanism-a slow mechanism,

  • much slower than this H+ thing--

  • which can allow slow averaging of the Hs.

  • But if you have DMSO as the solvent, this doesn't happen.

  • Why?

  • Because DMSO forms very strong hydrogen bonds with the OH,

  • and it doesn't form these trimers in which the hydrogens

  • can shift from one molecule to the other.

  • If DMSO were a strong enough base, it could pull the H off

  • and be like an acid catalyst, it would be a base catalyst.

  • Pull it off one molecule, put it on another one.

  • But it's not that strong a base, it's just strong enough

  • to forbid forming those hydrogen bonded trimers.

  • So we can get this averaging.

  • But wait a second, we looked at this one a couple lectures

  • ago when we were looking at IR.

  • And we said that even though this was said to be an

  • anhydrate, it had these water peaks.

  • And we said there were two kinds of water peaks.

  • There are the ones that are hydrogen bonded here, and the

  • ones that are not hydrogen bonded.

  • That's two different peaks.

  • But shouldn't those average the same way they do in NMR

  • where as you go from molecule to molecule as the molecules come

  • apart and come together in

  • hydrogen bonding, you quickly average those?

  • Why don't you get averaging in the IR when you do get

  • averaging in the NMR?

  • Another question is, what's fast?

  • How fast does it have to be to see a blur, or to see the

  • average as a sharp average, rather than to see the two

  • distinct things?

  • This particular sample is a solid, so you could imagine

  • that these are held in place by the surrounding molecules.

  • But you see the same effect in solution, so that's not the

  • difference.

  • Now notice that in IR, we have wave numbers here to tell us

  • how fast this is vibrating and how fast this is vibrating.

  • And this one is 110 x 10 to the 10th Hz, and this OH

  • bond is 102 x 10 to the 10th Hz.

  • That is, the difference between them is

  • 10 to the 11th Hz

  • whereas the difference we were looking at in the alcohol

  • things were 100 Hz or a few hundred Hz, at most. So

  • this is a much bigger frequency difference here in

  • IR than it was in NMR. What does that

  • have to do with anything?

  • How long does it take to measure a frequency precisely?

  • How long is long for purposes of an average?

  • Well suppose you have two different swings.

  • I'm going to put these two guys on swings up to the

  • ceiling and they'll start swinging.

  • And they swing at a certain rate, but they won't swing at

  • exactly the same rate if they're a little bit

  • different, the two swings.

  • This is pretty fast swinging.

  • One of them swings a 20 Hz, and the other

  • swings at 21 Hz.

  • And I'm going to try to push them, but I'm not going to

  • adjust the frequency I'm pushing at.

  • I'm just going to sit here and go bum bum bum bum

  • bum bum bum bum bum.

  • And I'm going to do it a 20 Hz, bum bum bum bum bum.

  • So the guy that's swinging naturally at 20 Hz, I

  • always push him every time he comes, just right.

  • So the light field in the IR case, or the NMR case, matches

  • perfectly for 20 Hz but it doesn't match for

  • 21 Hz, of course.

  • If I just did it for a little while on the 21 Hz just for

  • a little while, it would be almost in phase and I'd be

  • pushing them.

  • But after a little while, we'd get out of phase,

  • and I'm pushing when

  • he's coming and then pulling when he's going and so on. So

  • it's going to get out of whack.

  • How long does it take to get out of whack?

  • How long do I have to be doing this in order to find out that

  • I'm not really doing anything with him?

  • Well notice that a one-second pulse samples the full range

  • of phases, from being completely in phase to coming

  • back to being completely in phase again.

  • So if I do it for one second, I get nothing.

  • It just cancels.

  • On this one, everything was golden.

  • But obviously, then, there's no net interaction with the

  • light in that one over the course of one second.

  • So a one second light pulse can clearly distinguish

  • between 20 Hz and 21 Hz.

  • Everybody with me?

  • That means... Suppose we have 22 Hz, a swing that I'm trying with

  • my light to hit

  • but I'm going at 20 Hz.

  • So that goes through two complete cycles there.

  • In fact, if it goes through one complete cycle

  • you can tell it's not working after only half a second.

  • That is, 1 over the difference in lambda.

  • That's long enough to sample the full range favorable and

  • unfavorable.

  • So I can more quickly tell the difference when the

  • frequencies are more different.

  • It depends on 1 over the frequency difference, that's

  • how long I have to go to know that I'm not going to get any

  • action out of it.

  • Now here's the interesting thing.

  • Suppose the frequency is changing in time.

  • It's averaging, a proton is jumping from one position to

  • another and changing its frequency.

  • Suppose that it goes at 22 Hz for a while, then at 20

  • Hz for a while-- it's in a different environment.

  • Then at 22 Hz, then at 20 Hz, then at 22 hz, then

  • at 20 Hz for a while.

  • It's jumping back and forth randomly at a certain rate.

  • Now obviously, that's not going to work with the 22

  • Hz light.

  • It goes in and out of phase, so we won't get any absorption

  • of 22 Hz light.

  • Suppose we try 20 Hz light, so we'll move it up there.

  • Again, it goes in and out of phase.

  • No profit in that one.

  • But suppose I try 21 Hz.

  • Now it stays pretty much in phase, so

  • it's averaging properly.

  • And so what we will see is we have a very good match at 21 Hz,

  • the average frequency, and we'll see a single sharp peak.

  • 21 Hz will work, 20 Hz won't, 22 will. [correction: won't]

  • We'll see an average peak.

  • What's determining?

  • It's how long the pulse is compared to the

  • difference in frequency.

  • If the pulse is long enough, as long as the difference in

  • frequency, then we can tell it.

  • So when do peaks average?

  • When atoms don't stay put long enough to tell the difference

  • in frequency between the two sites that

  • they're changing between.

  • So if two peaks differ by 100 Hz, you have to count for

  • a hundredth of a second, roughly, to tell them apart.

  • But in the IR, they differed by 10 to the 11th Hz.

  • So the exchange of position would have to be 10 to the

  • 11th times a second in order to be able to

  • see separate peaks.

  • That's as fast as atoms vibrate.

  • They can't change their partners that rapidly.

  • So that's why in the IR you don't see averaging when you

  • do see it in the NMR. Because the NMR differences in

  • frequency are small, just a few Hz or a

  • few hundreds of Hz.

  • So if something happens in a hundredth of a second, you can

  • see separate peaks for it.

  • But you can't do that in the IR because the differences in

  • frequency are so large.

  • So this gives rise to what's called the NMR time scale.

  • Now suppose you had cyclohexane-d11.

  • It gives this spectrum.

  • It gives a spectrum for axial and equatorial hydrogens.

  • Incidentally, notice that that's sort of as if this were

  • like benzene with the ring current.

  • It gets shifted down when it's out the side, and shifted up

  • when it's along the axis.

  • At least, that's a way to remember it.

  • I'm not really sure that that's the explanation.

  • I'm not really sure that ring current is true.

  • But anyhow, you can remember which is which if you want to.

  • Now notice this is a special molecule, d11

  • just one proton.

  • And the reason is that you would get very complicated spin-spin

  • splitting if these are all hydrogen.

  • Because the axial will be split by that, this equatorial

  • will be split by that axial one.

  • Because they don't have the same chemical shift, so

  • they're going to split one another.

  • And then they'll also be split by these with different

  • coupling constants and so on.

  • So if we make all those deuterium, we get sharper

  • peaks so we can study it.

  • So it's d11.

  • But notice this is at minus 89 degrees centigrade.

  • And watch what happens if you start warming up.

  • The difference in chemical shift is point 0.48 ppm,

  • and this was measured in an old spectrometer in 1964

  • by Frank Bovey at Bell Labs.

  • And the difference here is 0.48 ppm

  • at 60 MHz,

  • or it's 29 Hz.

  • We have to put it in frequency. If we want to talk about this

  • kind of thing.

  • So 29 Hz difference in frequency.

  • Now we go from -89 dregrees to -68 degrees.

  • The peaks get broad.

  • To -63 degrees, broader still and moving in.

  • At 60 they coalesce and become a single

  • peak at -60 degrees.

  • At -57 degrees, it starts sharpening up, and sharpened

  • up still more at -49 degrees.

  • And what do you think it looks like at room temperature?

  • A single, really sharp peak

  • as sharp as these or sharper.

  • So it's a single sharp peak up there.

  • Even if you have C6H12 because it's flipping back and forth

  • rapidly enough that it's averaging the chemical shifts.

  • And once the chemical shifts become the same on average,

  • then you can't see the splitting anymore.

  • So cyclohexane gives a single sharp peak.

  • But this allows you to tell how fast, because you know

  • this frequency difference.

  • So it turns out that they coalesce to give a single peak

  • when it's at about that same--

  • the lifetime is about the same as that.

  • It's actually four times that, but for qualitative purposes

  • it's the same at the time they coalesce.

  • So it's about 100 times a second that it's flipping back

  • and forth at -60 degrees.

  • But you can measure this, you can fit these lines to

  • theoretical calculations for rates at other temperatures.

  • Or you can go to a different spectrometer which has the

  • same chemical shift difference, but a bigger

  • frequency difference.

  • And then get at what temperature you'd have that

  • frequency, or four times that frequency.

  • So that's a great way to measure the rates, and by

  • measuring the rates at different temperatures, you

  • can get what the barrier is.

  • And that's where we got the number from last semester. When

  • we were talking about conformational analysis.

  • We said that the barrier for flipping one chair to another

  • in cyclohexane is 10.5 kcal/mole, and

  • that's where you get it.

  • It's not easy to measure something like that, but NMR

  • allows you to do it.

  • So this is a handy way to measure really fast rates and

  • very small barriers that correspond

  • to these fast rates.

  • So the NMR application in dynamics is very important for

  • studying fast rate processes.

  • Let me see how much time.

  • Let me just go very quickly, I'm going to go really quickly

  • through this so we can get to a little bit

  • of something else.

  • But this was the kind of question that I

  • might ask on an exam.

  • The first part, Question 1, had to do

  • with making this molecule.

  • We've talked about making hemiacetals before.

  • But the relevant part here is Question 2, which was to

  • sketch what you'd expect for the NMR

  • spectrum of that compound.

  • So let's look at the different groups and you predict what

  • we're going to get in terms of chemical shift.

  • First, (a). How about its chemical shift, where do you

  • expect it on the spectrum?

  • The OH?

  • STUDENT: Downfield.

  • PROFESSOR: It'll be downfield because of the O, but it'll be

  • almost any place because of the hydrogen bonding and so

  • on, that has a big range.

  • And how about the splitting?

  • Should it be a singlet, a doublet, a triplet?

  • quartet, quintet?

  • 1:4:6:4:1?

  • What determines the splitting of this?

  • How many protons are on the next adjacent atom.

  • So we couple through this bond, this bond, to here.

  • These are too far away.

  • Those are removed by one, two, three, four bonds.

  • But to here, it's only one, two, three bonds.

  • So it should be a doublet, right?

  • Everybody agree it should be a doublet?

  • Sebastian, you shake your head slightly.

  • STUDENT: Well, it depends on what the solvent is.

  • PROFESSOR: Ah, in DMSO, it would be a doublet.

  • But mostly, you don't use that.

  • And in chloroform, it would be exchanging so you wouldn't see

  • the splitting.

  • So it could be almost any place between 1 and 6 for

  • hydrogen bonding, and the splitting is...

  • probably it would be a broad singlet

  • because of rapid OH exchange.

  • OK, now let's go to (b). Where should that

  • be, chemical shift?

  • Way at the right, way at the left, in the middle,

  • almost to the right?

  • STUDENT: Way at the right.

  • PROFESSOR: So it's hydrogens that are on a carbon and the

  • next atom is carbon too.

  • That's up at the right.

  • But it has a couple oxygens here.

  • So it'll be shifted down a little bit.

  • Slightly deshielded by two oxygen atoms on the

  • neighboring carbon.

  • So one might predict someplace between 1 and 2, and

  • there's a compound that's very like that, that is at 1.3.

  • OK, how about the splitting?

  • This one?

  • Now it has three bonds to get from here to here.

  • So that should be a splitting, there's no exchanging there.

  • So it should be a 1:1 doublet, split by--

  • and the J should be about 7 Hz.

  • OK, now how about (c), what should it look like?

  • What should its chemical shift be, relative

  • to, say, these hydrogens?

  • STUDENT: A doublet of quadruplets.

  • PROFESSOR: Derek what did you say?

  • DERRICK: I said a doublet of quadruplets.

  • PROFESSOR: Not the splitting, where should the chemical

  • shift be first?

  • DERRICK: In the middle.

  • PROFESSOR: It's attached to a carbon.

  • What's that carbon attached to.

  • Amy?

  • AMY: Oxygens.

  • PROFESSOR: They are oxygens.

  • So it'll be shifted down more than this one was.

  • Because it's closer to the oxygens.

  • So it's deshielded by two oxygen

  • atoms on the same carbon.

  • So it would be between 4.5- there's an analog

  • that's at 4.7.

  • Now how about, Derek, now you tell us about the

  • splitting of this one.

  • STUDENT: I think it's a doublet of quadruplets-- or a

  • quadruplet of doublets.

  • I don't know how to--

  • PROFESSOR: It would be -- if it were split by this, that makes

  • it a quartet.

  • STUDENT: Yeah.

  • PROFESSOR: And this would make it a doublet, except we already

  • said that that's probably exchanging.

  • It would depend on the solvent.

  • So probably it would just be a 1:3:3:1 quartet.

  • J about 7 Hz.

  • OK, let's keep going.

  • How about this one?

  • What's that going to be like?

  • Jack?

  • Chemical shift for that CH3?

  • STUDENT: 2.

  • PROFESSOR: Maybe about 2, because it's attached to a

  • carbon which bears oxygen.

  • So slightly deshielded by the oxygen.

  • That's rather like this one, which is attached to a carbon

  • which bore two oxygens.

  • So not as far down as that.

  • 1.1.

  • And how about the splitting of that one, Jack?

  • STUDENT: It would be a doublet.

  • PROFESSOR: Just a doublet, because

  • these are too far away.

  • 1:1 doublet, 7 Hz.

  • OK, now let's go to (e). What's it going to look like?

  • How about its chemical shift?

  • Debby?

  • STUDENT: It would be like 4?

  • PROFESSOR: Yeah, because it's got an

  • oxygen on the same carbon.

  • So it's 3.8.

  • Not bad.

  • And how about its multiplicity?

  • How many hydrogens splitting it three bonds away?

  • I can't hear very well.

  • STUDENT: It would be... I don't know.

  • PROFESSOR: Well, we already said that this one is going to

  • be a doublet split by that.

  • If that's true--

  • if this one splits these, then these split that.

  • STUDENT: So do the top ones.

  • PROFESSOR: So do the top ones.

  • So how many altogether?

  • STUDENT: So it's six hydrogens, so its--

  • PROFESSOR: So it'll be seven lines in

  • the binomial intensity.

  • Seven line multiplet from the six atoms on

  • the neighboring carbons.

  • 7 Hz.

  • So that's going to be really, a whole bunch of them.

  • Seven lines.

  • OK, now how about (f)?

  • STUDENT: Like (d)?

  • PROFESSOR: Chris?

  • STUDENT: Same as (d).

  • PROFESSOR: Ah, same as (d). It's got the same connectivity as

  • (d), same hydrogen coupling and so on.

  • But it's slightly deshielded by the oxygen atom there-- so

  • I copied that from here.

  • And I copied this from here, and that's what you said.

  • But it's different from (d) because

  • here is a chiral center.

  • So one of those is pro-R and the other is pro-S. They're

  • different within another chiral center.

  • So in fact, when you look at a situation like that, there are

  • two that are really close to one another.

  • Two doublets.

  • They're about the same, but I caught you.

  • STUDENT: Can you see that?

  • PROFESSOR: Yeah, you do see that.

  • Not in every molecule.

  • Notice it's an isopropyl group, two CH3s on a CH

  • attached to something.

  • If the rest of the molecule is chiral,

  • those will be different.

  • CHRIS: By how much?

  • PROFESSOR: That depends on what the molecule is.

  • It's sometimes hardly observable.

  • But very often, you can see it.

  • I saw one, when I was a graduate student I got a

  • spectrum like that.

  • I said, what the heck is going on?

  • So that was a standard question I would ask on exams

  • down here of graduate students at that time, after

  • I learned it myself.

  • OK now, proton decoupling.

  • Let me just start this, and then we'll show the

  • example next time.

  • So we have JC13-H here.

  • Now notice that when we put this in a certain magnetic field,

  • the proton will precess at 100 MHz.

  • How about the C-13?

  • Will it precess at 100 MHz?

  • No, because it's a much weaker magnet.

  • It doesn't feel as much torque, it doesn't precess as

  • fast.

  • So the H has a steady field in that direction.

  • Or there are other ones that have exactly the same steady

  • field in the opposite direction.

  • Now C-13 is 25 MHz, it's only a quarter

  • as strong as a magnet.

  • So it precesses much more slowly, it's a smaller magnet,

  • and it can be either up or down.

  • So both these things are happening when we put it in a

  • magnetic field.

  • And here's the NMR spectrum, and we've seen that, and saw

  • the satellites--

  • this peak comes from C-12s.

  • This is from C-13s when they're down, and this is the proton

  • precession when the C-13 is up, reinforcing the applied field.

  • And at the same time, if we looked at the C-13 spectrum,

  • notice that essentially all of hydrogen is protium.

  • Only 0.015%, very little, is deuterium.

  • So you don't see that big peak in the middle here, because

  • all the hydrogen is either this way or this way.

  • And the magnitude of the splitting in both

  • cases is 125 Hz, that's fine.

  • So now, suppose we do something really cute.

  • Suppose we put two radio frequencies in.

  • One at 100 MHz, one at 25 MHz.

  • But what we're going to do--

  • the way we're going to do it is to irradiate 100

  • MHz all the time.

  • But just do a pulse, a 90 degree pulse. Of the C-13 to twist it

  • down into the plane, and then allow it to precess.

  • And that's what we're going to listen to

  • see if we got a signal.

  • We're going to listen to the C-13 while irradiating the H.

  • Now what effect does that have, then?

  • If you were in a frame rotating at 100 MHz,

  • you'd see that this little 100 MHz field is going to

  • cause the hydrogens to precess like this.

  • To go up, down, up, down, up, down.

  • What determines how fast they go up and down?

  • How fast they precess?

  • What determines that?

  • It's how big this field is.

  • How strong the field is.

  • The stronger the field, the faster the precession.

  • So that means H is going to be up, then down,

  • then up, then down.

  • And how fast it's going up, down, up, down, up, down

  • depends on how much power we put into our radio.

  • If we do it fast enough, we see the average of the C-13 and

  • decouple the spectrum.

  • We can see the C-13 spectrum without the H splitting it, by

  • making the H's quickly go up and down.

  • How fast do they have to go up and down in order to average

  • the 125 Hz splitting to be nothing, to

  • just be a single peak?

  • How fast?

  • Once you've had time to review the lecture, you'll be able to

  • answer that.

  • STUDENT: 125.

  • PROFESSOR: About 125 times a second, right?

  • The frequency difference is how fast you

  • have to flip to average.

  • So if you put enough power in on the hydrogen, you remove

  • the splitting from the C-13 spectrum, which can be handy,

  • as we'll show you next time.

  • So the decoupling power determines the rate of this

  • precession.

  • And this, we'll see next time.

J. MICHAEL MCBRIDE: So we've talked

Subtitles and vocabulary

Click the word to look it up Click the word to find further inforamtion about it