Placeholder Image

Subtitles section Play video

  • [ Silence ]

  • >> My name is Javert.

  • Was that at least as good as Russell Crowe?

  • OK, so we're ready to get started.

  • First a quick quote, "Chemistry is to biology,

  • what notation is to music."

  • To me this really grabs at the essence of chemical biology

  • in the sense that the notations

  • on a musical scale allow creativity.

  • They allow other reformers to interpret the works in new ways

  • and give the work context.

  • Chemistry does that in biology.

  • Chemistry gives us an opportunity for us

  • to be creative about biology and invent new ways

  • of thinking about biology.

  • It's sort of the underlying basis at the level of atoms

  • and bonds as I keep saying, for biology,

  • and to me in some way this really captures what we're

  • trying to do in this class.

  • OK. So this week, where it's already week 3

  • which is amazing-- Oh, hang on.

  • OK, so it's week 3 so we're up to chapter 3 and we're going

  • to be talking about DNA.

  • Our knowledge of DNA was really set in place

  • by the people in front of you.

  • These are the giants really in the field of structural biology

  • who determined structures of DNA in the 1950s.

  • This includes the great Rosalind Franklin whose very accurate

  • x-ray diffraction structures and her pictures

  • of the x-rays diffracting off the fibers of DNA set

  • in motion the determination of the structure.

  • She was working with Maurice Wilkins and two physicists,

  • Francis Crick and Jim Watson went

  • on to solve the structure of DNA.

  • And as we'll see in a moment,

  • really one of their key insights was at the level of atoms

  • and bonds in the sense that they discovered interesting

  • tautomerization of the DNA bases that made it possible

  • to have what we now call Watson-Crick base pairing

  • between the strands of DNA.

  • Getting a little ahead of myself but that's where we're going

  • in the next week or so.

  • So we're going to be finishing

  • up non-covalent interactions then talking

  • about DNA structure, DNA property

  • and finally DNA reactivity of small molecules.

  • This is a large chapter.

  • We have a lot to talk about, so bear with me.

  • Things are going to go not faster it's going

  • to be the same speed, but we're going to gloss

  • through a few topics that are less important and when we do,

  • this means then that you can focus your reading

  • and your study just on the level of detail

  • that we're covering in the class.

  • OK. Some announcements, in the textbook, read chapter 3.

  • Again, skim concepts not presented in lecture,

  • don't get too worked up about them.

  • Chapter 3 problems, do all of the odd numbered and all

  • of the asterisked problems.

  • In addition, I want to encourage you

  • to get involved here at UC Irvine.

  • This is super important.

  • Many of you I know aspire to become physicians or scientists

  • or pharmacists or whatever it is that you aspire to do.

  • All those big plans require preparation,

  • they require some evidence that you've gone beyond the ordinary

  • and I want to encourage you to do this.

  • OK. One way to get involved is to look

  • around for opportunities to volunteer.

  • This is one that's run by my friend who is one

  • of the founders, it's called the Social Assistance Program

  • for Vietnam.

  • If you go to this website, there are opportunities to volunteer

  • to spend two weeks in Vietnam, in a rural part

  • of Vietnam administering medicine, you know.

  • You'll probably not be of course, you know,

  • drilling people's teeth and, you know, doing open heart surgery

  • but you will get a unique opportunity

  • to actually see those types of things happening

  • and that's really important if you aspire

  • to that kind of career.

  • It provides evidence that you're qualified, that you're committed

  • and that you're someone who is altruistic.

  • All of those things professional and graduate schools look

  • for in your application.

  • You need to be doing those things now.

  • OK. And I'm on your side on this.

  • OK. I will help you get-- find those opportunities.

  • I'll bring them to your attention like this one.

  • And if there's something in particular that I can do

  • to connect you with, let me know

  • and I'll do my very best on your behalf.

  • OK. Along those lines, our laboratory always has openings

  • for talented undergraduates.

  • It's competitive but you have a chance to participate

  • at the full level of a graduate student.

  • Undergraduates in our laboratory are doing actual science.

  • They're publishing papers with us.

  • They're making discoveries and they're participating

  • as full members of the team.

  • OK. Here's how you apply,

  • send me a paragraph describing your career goals

  • and how research

  • in our laboratory would advance those career

  • and educational goals.

  • In addition, send me a copy of your college level transcripts.

  • This includes any transcripts at community colleges

  • if you're transferring.

  • Many of my best students are transfer students.

  • Send me those transcripts as well

  • and also send me three names and email addresses of TAs

  • who know you well in lab sections.

  • OK? And I'm going to email them and I'm going

  • to ask them what was this person like in the laboratory?

  • Were they, you know, the first one out of the room?

  • Were they last one out of the room?

  • Did they, you know, follow you

  • around the laboratory asking you,

  • "Does this look pink, does this look pink?"

  • or were they pretty independent, OK?

  • So I'll find out about that sort of thing

  • and then that's how I make a decision on who

  • to accept into the laboratory.

  • OK. And then of course the resume.

  • This is pretty standard, if you're interested

  • in doing research here at UC Irvine which I highly,

  • highly encourage you to do,

  • this is a pretty good way to go about it.

  • OK. This is an effective way to get noticed and to get

  • that job that you need.

  • OK. Any questions about these opportunities?

  • Why I think they're important and things like that?

  • OK. See me in office hours if there's something in particular

  • that you want from me and I'll try to hook you up.

  • OK, office hours this week, speaking of which.

  • Tomorrow I'll have my usual office hour,

  • 2.45 to 3.45, the usual location.

  • Thursday I'll have my office hour 11 to 1, usual location.

  • In addition, Mariam will have her office hour Fridays 1:30

  • to 2.30.

  • And, Kritika, could you raise your hand?

  • So, Kritika is our new TA.

  • She'll be joining in the team.

  • And Kritika, does this time work for you, Tuesdays 2:30 to 3:30?

  • Good. OK. And she'll be having her office hours Tuesday.

  • So notice that we spread out our office hours

  • so that there's one everyday of the week except Monday

  • because I know you're very busy

  • on Monday doing all kinds of things.

  • I hope you're having fun yesterday.

  • But yeah, so everyday of the week there's an office hour,

  • they're staggered so they're at different times,

  • so you can have your questions answered.

  • And again, Kritika is a graduate student in my laboratory.

  • She knows this material as well as I do.

  • She's really smart.

  • You can go to her office hour and get an answer.

  • That's as good as an answer that I will give you.

  • OK? And for that matter you can also email the TAs

  • with your abundant questions.

  • OK. I'm looking at you where I can find that person, OK,

  • there's like one person in the class,

  • he send me 10 emails a day but, you know, I will do my best

  • but you can also email the TAs as well.

  • OK. Oh, along those lines,

  • I sent you an email saying don't send me book

  • or potential journal articles.

  • And the reason is I must do-- I open my inbox and I had like 15

  • of those and I got to the point where I was bouncing messages

  • because the inbox was so full.

  • So, if you send me those, I can't do very much with them,

  • OK, it might clog my box.

  • So what I propose we do is instead of you emailing me them,

  • instead bring them to my office hours,

  • bring them to Kritika's office hours or Mariam's office hours

  • and ask your questions then.

  • OK. Now the standard question I get asked is,

  • is this article appropriate?

  • And my answer to that is if you follow the guidelines,

  • it will be appropriate.

  • Now in addition, when you're writing your summary,

  • your report on the journal article, focus on the aspects

  • of the article that fit the definition of chemical biology.

  • OK. So a paper in cell for example is going

  • to be a very meaty paper, it's going to cover about 10 pages.

  • It's going to have, you know, eight or nine figures and some

  • of those figures, some of those experiments aren't exactly what

  • we will call chemical biology.

  • Don't focus on those.

  • Focus on the ones that are chemical biology related.

  • OK. Otherwise I don't know

  • that you know the definition of chemical biology.

  • OK. Any questions?

  • All right, guess what, we're heading into midterm season.

  • This is week 3, so week 4 is next week.

  • We will have a midterm next Thursday,

  • a week from this Thursday.

  • There will be a review session in advance

  • by the TAs, time to be announced.

  • Kritika will arrange for this.

  • The seating for the midterm will be assigned.

  • Mariam, you'll do the assignment.

  • It's really essential that you bring your UCI student ID.

  • We will check the IDs to make sure you're seated

  • in the right seat.

  • If you're not seated in the right seat, it will be treated

  • as an academic honesty infraction.

  • No notes, no calculators,

  • no electronic devices, you don't need them.

  • You're smart.

  • OK? Any questions about the upcoming midterm?

  • OK, now I know you want to know what will be on the midterm.

  • OK, so let me tell you, it will cover

  • through Tuesday's lecture one week from today

  • and so we will be about halfway

  • through chapter 4 on Tuesday, OK?

  • So plan to read through about halfway through chapter 4,

  • that's the chapter on RNA and that's where I expect to be

  • for Tuesday's lecture.

  • It's possible I might get behind but I'm going

  • to really try hard not to do that.

  • OK. All right, I will also post a practice midterm

  • to the website and you can use that along

  • with the discussion worksheets, the assigned problems as a guide

  • for what will be on the midterm.

  • OK? So the midterm will look very much like a compilation

  • of discussion worksheets, of assigned problems

  • and the practice midterm.

  • OK. And it will be about as long as the practice midterm as well.

  • So when the practice midterm comes out,

  • I'll post two versions.

  • One version will be blank, one version will be the key.

  • The blank version you should print out

  • and then give yourself an hour and 20 minutes and make sure

  • that you can handle it.

  • OK. And at the end of that,

  • then check your answers against the key.

  • But give yourself a real practice.

  • OK. That's pretty important, I think.

  • OK, so anyway, that's the plan, any questions

  • about the midterm coming up?

  • I know you will have lots of questions.

  • I look forward to hearing about them in my office hours.

  • All right, I want to go back to finish up our discussion

  • of non-covalent interactions and where we left off last time was

  • with charge-charge interactions.

  • I'm not ready to talk to you about interactions

  • between atoms that are uncharged.

  • OK. Neutral atoms that are interacting with each other.

  • These are described by a Lennard-Jones Potential

  • which is an equation

  • that describes how these neutral atoms would interact

  • with each other.

  • Another way of describing these neutral atoms,

  • another term that's used and probably one

  • that you encountered is London dispersion force.

  • OK. So when you have two say neon atoms that cozy up next

  • to each other, then they will interact

  • through a London dispersion potential or force

  • and that's what I'm describing here.

  • OK. So, it's just a couple of different ways

  • of saying the same thing.

  • This happens a lot in biology, not necessarily

  • between neon atoms but certainly between aliphatic side chains,

  • hydrophobic side chains in proteins, in interactions

  • with each other, interactions with lipids, at the--

  • a plasma membrane of the cell, and a whole host

  • of other non-covalent hydrophobic-hydrophobic

  • type interactions.

  • This turns out to be a very potent

  • and very strong force in biology.

  • OK. So we need to understand it better.

  • So the energy-- the potential energy

  • of a van der Waals interaction, yet another word to describe it,

  • is equal to-- is proportional to 1 over r

  • to the 12th minus 1 over r to the 6.

  • These terms, the sigma term deals with the diameter

  • in this epsilon ij, not so important, so let's ignore that.

  • Let's focus in on the 1 over r to the 12th term and 1

  • over r to the 6th term.

  • First, notice that it's minus 1 over r to the 6 and minus

  • in potential terms means more stable in energy,

  • lower on this Y-axis of potential energy over here.

  • OK. So that's going to be our attractive term.

  • Hydrophobic, if things attract each other, OK, not just due

  • to repulsion from water.

  • We'll talk about that next.

  • But hydrophobic things want to stick to each other

  • and they're going to do this

  • with an attraction that's proportional

  • to 1 over r to the 6.

  • The fact that it's 1 over r to the 6 as opposed to r to the 2nd

  • in the charged-charged interactions means

  • that this is a much shorter range attraction.

  • This attraction takes place on a very tiny distance scale.

  • OK. Now eventually the two atoms in this case as described here,

  • or two molecules-- or two molecules bang into each other

  • and go past the point where they're attracted to each other.

  • OK? And at that point, their electrons are trying

  • to overlap with each other.

  • That's really bad news, right?

  • We know by the Pauli exclusion principle

  • that that's not allowed and so in the same way

  • that my fingers are never going to fuse with each other,

  • just going to bang off of each other, the atoms push away

  • from each other and they push away from each other

  • with the repulsion force

  • or repulsion potential that's proportional

  • to 1 over r to the 12th.

  • OK? And so this means that this is extremely short ranged

  • and extremely sharp, right?

  • To the 12th power is a large number.

  • So this means that this really dramatically pushes apart the

  • atoms if they happen to get too close to each other.

  • It turns out that there are a whole series

  • of other non-covalent interactions that we find

  • in biology that actually contribute quite a bit

  • of non-covalent binding energy.

  • Here for example are the dispersion interactions

  • that we have discussed before on the previous slide

  • and so these include things

  • like aliphatic-aliphatic interactions,

  • but also aliphatic interacting with hydrophilic molecules.

  • So here is water interacting with methane.

  • They're going to interact with each other

  • and have some attraction.

  • These number here, minus 0.5

  • to minus 0.7 kcals per mole is pretty low.

  • OK. This is not a tremendously strong interaction.

  • Where it gets strong is when you have a molecule

  • that has a large number of functional groups.

  • Each one with 0.5 kcals per mole here, 0.5 here, 0.5 here,

  • and when you sum up across all of these, you're starting

  • to talk about big energy.

  • OK. Now just to give you an idea, you need to know one fact

  • that I think is really important.

  • And the fact is important enough that I'm going to try

  • to write it on the board over here in the corner.

  • The fact is that, a factor

  • of 1.4 kcals per mole will be a factor of 10 in--

  • at equilibrium constants.

  • OK. So, 1.4 kcals per mole is a magic number

  • in chemical biology.

  • OK. So look for 1.4 kcals per mole

  • because that tells you then that's favored tenfold

  • over nonbinding.

  • In other words, the interaction is going

  • to be 10 times more likely to form than not form.

  • OK? It's a factor of 10 in terms of equilibrium constants.

  • OK. So, if we're talking about something over here,

  • that's only 0.5, 0.7 kcals per mole, you have to start summing

  • up a whole bunch of these to get anywhere in terms

  • of enforcing the interaction.

  • On the other hand, some of these other interactions can be quite

  • strong and let's take a closer look at those next.

  • OK. So, for example, we've talked a little bit

  • about hydrogen bonding.

  • Hydrogen bonding of course has a donor and acceptor

  • and here's a range of strengths.

  • Hydrogen bonds vary enormously strength

  • from about 1 kcal per mole all the way to 7 kcals per mole.

  • The strength of the interaction depends enormously

  • on the identity of the donor and acceptor.

  • When the donor or-- and/or acceptors are charged,

  • if either one is a charged functionality, the strength

  • of this hydrogen bond goes up enormously.

  • And this kind of makes sense, right,

  • because remember earlier I described a hydrogen bond

  • as a kind of a special case of a charged-charged interaction

  • in which a hydrogen is being shared between two atoms.

  • OK? So, if one of these happens to be charged, that's going

  • to be a much stronger charged-charged interaction.

  • Speaking of charged-charged interaction,

  • salt bridges are the coulombic potential

  • that we saw on Thursday.

  • These are the charged-charged interactions.

  • These vary also enormously depending upon the environment

  • that the salt bridge happens to find itself in, where a--

  • where water can shield this charge.

  • Water or counter ions can shield this charge,

  • decreasing it considerably

  • and making the interaction much, much weaker.

  • So, a salt bridging interaction, which is another way

  • of saying charged-charged interaction found

  • in a hydrophobic environment, say the interior

  • of a plasma membrane, is going

  • to be a much stronger interaction

  • than one that's found out in water where there's plenty

  • of water and counter ions to shield the charge.

  • OK, where those provide a counter against the charge.

  • Recall that those environmental terms are embodied by the 1

  • over 4pi epsilon term in the coulombic potential

  • that I showed you on Thursday.

  • OK. In addition, there's also dipole-dipole interactions

  • which are alignments of densities of charge

  • where we have a little bit more negative charge

  • on the oxygen over here.

  • The dipole is pointing in this way on the--

  • to the right on the upper acetone

  • and to the left on the lower acetone.

  • The two of these dipoles want to cancel each other out.

  • By cancelling each other out,

  • that will give you a more optimal interaction and that's

  • where some potential energy.

  • Finally, there's also a whole series of aromatic

  • or arene interactions.

  • And in general, this includes both face-to-face interactions

  • where you have two faces of a benzene ring

  • that are interacting with each other.

  • Notice in this picture over here

  • that the top benzene ring is offset

  • from the bottom one and this makes sense.

  • We're going to be looking at regions

  • of electron density interacting

  • with regions of electron poverty.

  • OK, that that's actually the basis for the interaction.

  • And so for that reason, we also see very commonly edge

  • to face interactions.

  • OK, so this is the one that we'll see in a moment

  • when we start looking at pi stacking in DNA.

  • But in addition, you can have an edge

  • of an aromatic system interacting with the face

  • of another aromatic system down here, and that's as strong,

  • right, it has the equivalent strength.

  • Even though you expect, you know, face-face to be ideal,

  • that's actually not what we see when we start looking

  • at large numbers of aromatic interactions.

  • We see this H-- edge to face interactions all of the time.

  • OK. And then finally there's some other ones

  • that are really bizarre and they include charged interacting

  • with the electron rich aromatic rings.

  • And this kind of makes sense, right?

  • You have something that's positively charged,

  • you have something that's very electron rich

  • in terms of the ring system.

  • So these cation-pi interactions

  • which is what this one is called are found pretty ubiquitously

  • in biology, oftentimes playing a commanding role,

  • playing a really key role in chemical biology.

  • OK. So, these are ones that I'd like you to memorize.

  • I'd like you to know something about their strengths,

  • which one is strong, which one is weak.

  • I don't want you to memorize the numbers per se but I want you

  • to know something and be conversant

  • on relative strengths.

  • OK, relative strengths matters.

  • OK. And one thing-- one last thing to keep in mind

  • if we're going for this 1.4 kcal per mole, again,

  • you can have a summation of a large number of interactions

  • to achieve that 1.4 kcals per mole or even more.

  • And I'll show you an example of that very shortly.

  • Now it turns out that it's actually a little bit tricky

  • to start comparing energetics when you design

  • in say the perfect cation-pi interaction.

  • What ends up happening is

  • that you get a complication due to water, OK?

  • So let's imagine that you had designed

  • in the perfect cation-pi interaction and in doing

  • so you put this positively charged thing that forces all

  • of the water around it

  • to rearrange itself or reorient itself.

  • It turns out that's actually a complicated thing

  • of the orient-- reorientation of water

  • but it cannot be neglected.

  • OK. So what we do is we make a very important simplifying

  • assumption, and I'll talk more about water on the next slide.

  • But before I do, water, since we just have to acknowledge

  • in advance, water can complicate everything, right?

  • It's present at 55 molar concentration in your cells

  • and we can't neglect it, OK.

  • It has its own energetics.

  • It's-- as I showed on this slide over here,

  • for example it's interacting with hydrophobic things.

  • So its own energetics are really complicated, OK,

  • and actually very hard for us to understand and pin down.

  • And so it's really difficult to estimate the entropy lost

  • or gained in an interaction due to that rearrangement of water

  • when you start making changes.

  • So what we'd like to do is compare things that are

  • as similar to each other as possible.

  • OK. This is the simplifying assumption

  • that I alluded to earlier.

  • Here for example is an example of that, OK?

  • So, here's two possible transitions states,

  • and transitions are two possible mechanisms.

  • Mechanism number one involves an SN2 reaction.

  • Mechanism number two involves the same molecule undergoing an

  • E2 elimination reaction.

  • And the key here is that the molecules are identical.

  • OK? That extreme similarity makes the comparison

  • between these two much easier to make.

  • OK. And so, for example if we're looking at two proteins,

  • we can look at empty protein versus ligand-bound protein.

  • But on the other hand, we're not trying to make all kinds

  • of changes to the protein structure over here.

  • Problem is proteins are rarely, you know, like looking like this

  • when the ligand is unbound.

  • So, these simplifying assumptions will start

  • to cause all kinds of problems.

  • Here's one though that works.

  • You can make a single change to the surface of a protein

  • and then compare the altered protein,

  • compare its interaction with a ligand.

  • So for example, we could change this isopropyl group

  • to a methyl group and then compare what's happened,

  • what's different in that receptor ligand interaction, OK?

  • So all you've done is to remove two methyl groups.

  • That's about as simple as it gets, right?

  • So that type of experiment is an easy one to make comparisons to.

  • OK. And again, by doing that, we're trying

  • to minimize how much the water has

  • to rearrange itself at that interface.

  • OK. It turns out actually this assumption works most

  • of the time.

  • So in short, being good scientist, not changing lots

  • of variables at the same time pays off in biology

  • because underlying everything we do is this complicated solvent

  • that we operate in called water.

  • Let's take a closer look at the structure of water.

  • OK, so here is water in ice and notice how neatly regular it is

  • and how nicely ordered it is.

  • And then here's water in a solution as liquid water.

  • And it's just crazy complicated.

  • First, notice that there, all these dots--

  • dotted lines are the hydrogen bonds.

  • These hydrogen bonds are pretty much maximized.

  • Water is not passing up any opportunities

  • to hydrogen and bond to itself.

  • OK, but the hydrogen bonds

  • in the liquid solution are nonoptimal.

  • OK, water in solution, each water molecule is jam-packed

  • with other water molecules

  • and oftentimes the hydrogen bonds are slightly distorted

  • or they don't have the right distances.

  • Those little distortions and that lack

  • of perfect distances makes the hydrogen bonds

  • in liquid water weaker than they are in solid water.

  • Furthermore, a molecule of water in its own, you know,

  • with a lot of other molecules of water is--

  • and behaving kind of like it's on a crowded dance floor, OK?

  • So, it's bouncing around wildly against these other, you know,

  • molecules that are nearby and interacting with lots and lots

  • of different molecules nearby,

  • constantly breaking interactions and forming new ones.

  • OK. So, water is actually very complex.

  • Weak and distorted hydrogen bonds, OK.

  • In addition, when water cozies up to hydrophobic surfaces,

  • it tends to form a very ordered structure that starts

  • to look a lot like the structure found in ice.

  • And this works by water form--

  • satisfies its propensity formed hydrogen bonds

  • by forming a clathrate-like structure.

  • So for example, here is a molecule of methane encapsulated

  • in one of these clathrates of water,

  • where clathrate is just simply a structure of water

  • that satisfies its desire to form hydrogen bonds with itself.

  • OK, or with other molecules of like kind, OK.

  • This really dramatically changes the strengths

  • of nearby non-covalent interactions, OK.

  • This does things to strengthen those non-covalent interactions

  • because every time one of those,

  • let's just say hydrophobic-hydrophobic

  • interactions breaks, then water has to slide

  • in between the now broken interaction

  • and form one of these clathrates.

  • OK, that formation of the clathrate, the formation

  • of an ordered structure cost energy.

  • It's a loss of entropy.

  • This is a more ordered structure than the structure

  • of disorganized water

  • that I showed you earlier in solution, OK?

  • So for this reason, hydrophobic--

  • hydrophobic molecules are driven against each other.

  • They want to find each other in water.

  • And this is sometimes referred to as a-- this is--

  • this is actually a water driven effect.

  • Forgetting the technical word for this.

  • OK, anyway, so-- oh sorry, it's--

  • it's sometimes referred to as a hydrophobic effect,

  • OK, in water.

  • OK, now let's take a closer look

  • at a receptor-ligand interaction,

  • now zooming in at the level of atoms and bonds.

  • This is a molecule called human growth hormone and, yes,

  • Lance Armstrong admitted to Oprah

  • that he took human growth hormone to win--

  • to help him recover basically from different stages

  • of the Tour De France during all seven of his victories,

  • and it really annoys me actually.

  • I could say a lot more about that

  • but I'm going to hold myself back.

  • OK, now when human growth hormone binds to its receptor

  • on the surface of cells, it's stimulating growth and recovery

  • of those cells, it's stimulating protein production, et cetera.

  • And when it binds to the surface of the cell

  • to the binding partner on the surface of the cell,

  • its receptor, then all of the region that's colored

  • in on this surface is buried, OK?

  • So in other words, human growth hormone binding protein binds

  • over here and then makes contact with each

  • of these colored atoms.

  • OK, everything that's in white here is still out in the water,

  • out in the solvent, it's not interacting

  • with the receptor at all.

  • Now, when I was at postdoc, I repeated a classic experiment

  • that was done by Jim Wells.

  • And Jim Wells and his co-workers found

  • that even though there are 19 residues that are buried

  • on the surface, there are 19 amino acids that are buried,

  • only the ones in red are actually contributing binding

  • energy, OK?

  • So, notice that all of these other stuff is in--

  • that is in blue that is buried is not

  • at all contributing any binding interaction.

  • So although there's interactions between these side chains

  • of these two proteins, there is no binding energy that's being

  • exchanged or gained by that interaction.

  • OK? So, just because two molecules find each other,

  • two functional groups find each other in space, does not ensure

  • that there's actually going to be a net gain in binding energy.

  • Because again, that net gain

  • in binding energy includes both the strength of interaction

  • but must also include the water ordering and disordering term

  • which we've been calling entropy earlier, OK?

  • So, in order for this interaction to take place,

  • you're going to be pushing out ordered water

  • and gaining some entropy in some places

  • and in other places losing some entropy.

  • OK, now when we look even more closely, let's just zoom

  • in on this red patch over here.

  • This red patch has been termed the hot spot of binding energy.

  • That's where the binding energy allowing these two molecules

  • to interact with each other is found.

  • OK, this is the essence of the non-covalent interaction

  • between human growth hormone and its binding partner,

  • human growth hormone binding protein.

  • And in green, these are the functional groups

  • that are found in this red patch.

  • OK? So the red patch is over here and now I'm showing you

  • that the functional groups were in green.

  • These are carbon atoms.

  • In blue, that's a nitrogen.

  • And in red, that's an oxygen.

  • OK, notice that the hydrophilic functionalities, the guanine--

  • of an arginine over here, a bunch of nitrogens,

  • another nitrogen over here, an oxygen, an oxygen over here.

  • Notice that those are around the periphery of this red region.

  • They're around the outside of this hotspot of binding energy.

  • The center of the hot spot is largely hydrophobic, OK?

  • Notice that it has lots and lots of carbons.

  • There's a benzene ring, it is smack in the center.

  • There's this aliphatic chain that's capped

  • by an amine functionality,

  • but nevertheless this is an aliphatic chain.

  • There's aliphatic functionalities over here

  • and over here and over here, et cetera.

  • OK, so in other words, the outside hydrophilic,

  • the inside hydrophobic.

  • And so, when molecules,

  • functional molecules find each other, this is a very common way

  • for them to interact with each other through a small set

  • of residues that form this hot spot of binding energy

  • which again kind of looks like a core sample through a protein.

  • Outside is hydrophilic, inside hydrophobic, OK?

  • Any questions so far?

  • OK, let's talk one last--

  • about one last section

  • of chapter 2 before we move on to chapter 3.

  • There's this concept that the biooligomers

  • on earth are highly modular.

  • We've discussed this before.

  • This also extends to the polyketides and the terpenes

  • which are composed of isoprenes and the polyketides

  • which are caused-- which are composed of either malonyl

  • or acetyl subunits that are strung together,

  • where the red bonds indicate that where the connection

  • between these modules such as the amino acids

  • as individual modules in a protein.

  • OK? And furthermore, this is also found in oligosaccharides

  • where you have this glycosidic bond

  • that connects the glycan fragments together.

  • There's also a numerical amplification and biosynthesis.

  • So, if there's only one or two copies

  • of DNA per cell depending upon whether it's a prokaryotic cell

  • or eukaryotic cell,

  • some prokaryotic cells are [inaudible] more than one

  • but let's just simplify it.

  • Then to RNA, each DNA is transcribed 10 to 50 times

  • and then each RNA is translated say 10 to 20 times.

  • So in the end you end up with this massive amplification

  • signal going through the cell, where with one copy

  • of DNA you can end up with millions of products

  • from some enzyme reaction down here.

  • Last thoughts, form follows function in biology.

  • These-- the bonds that join together,

  • the oligomeric subunits are--

  • have a strength that follows their function,

  • their functional requirements, OK.

  • And so, for example, when we look at the half-life of lipids,

  • we find that actually the ester bonds in a lipid have a halfway

  • on the order of a year or so.

  • OK, so esters, not so stable.

  • Compare that against DNA down here which has a half life

  • on the order of 220 million years.

  • OK, that's his half life for DNA.

  • And in retrospect, this kind of makes sense, right,

  • because DNA has to be a-- you know, has to be a biooligomer

  • for the life of the organism, OK?

  • And so, we're now at the point

  • where we're routinely taking advantage

  • of this tremendous stability of DNA to amplify DNA

  • from even extinct organisms like wooly mammoths, like species

  • of prototypical humans that haven't lived on the planet

  • for tens of thousands of years.

  • That sort of thing is going on right now

  • in laboratories talking advantage

  • of the tremendous stability of DNA.

  • Now your hair, which is a protein, has a lifetime

  • on the order of, you know, 300 years or so.

  • And you can see that, right?

  • We can find, you know-- we give-- well, anyway.

  • So, I guess it depends on the human that we're talking about.

  • My hair obviously doesn't exist that long.

  • But, you know, so certainly the lifetimes here are following

  • their function, right?

  • Proteins don't have to last this long.

  • Question? How does one get a PhD that's going to take you five

  • or six years studying and trying

  • to measure these half lives of 220 million years?

  • Anyone have any ideas how to do that experiment?

  • I can guarantee it to you, it's not like, you know,

  • you set up this test tube

  • and then you check it every 20 years, OK,

  • to see how much gets cleaved.

  • How would you do this?

  • Yeah, how would you do it?

  • [ Inaudible Response ]

  • OK, a small amount of RNA.

  • And I would use a large amount

  • because very little is going to get degraded.

  • How would you do this though?

  • Yeah?

  • >> Maybe we can put it in a very decomposing environment?

  • >> OK. But then you wouldn't know

  • if the decomposing environment is different

  • than in the cell, right?

  • We want to know about the half lives in the cell, right?

  • Yeah?

  • >> Use a model organism.

  • >> Model organism.

  • No, I want to know what it's going to be--

  • what's going to last in, you know, in this cell

  • or this other one over here.

  • Question over here?

  • [ Inaudible Question ]

  • OK, you're definitely going to use radioactivity

  • because you need something that's super sensitive.

  • How would you do this?

  • [ Inaudible Response ]

  • OK, you're getting close.

  • What is your name?

  • >> Bryan.

  • >> Bryan? OK, Bryan is getting close.

  • So the suggestion was radioactivity.

  • Bryan's suggestion is you look for a tiny little quantity

  • and radioactivity gives you that sensitivity.

  • But are you going to do this

  • for 220,000 years or 220 million years?

  • >> No.

  • >> OK, so how are we going to do this experiment?

  • We have the sensitivity, we're going to look

  • for tiny little quantities and extrapolate back.

  • How are you going to model 220,000 years?

  • Yeah?

  • [ Inaudible Response ]

  • Carl, OK, look at fossils.

  • Yeah, and we do that.

  • Yeah?

  • [ Inaudible Response ]

  • OK C14.

  • [ Inaudible Response ]

  • OK.

  • Yes.

  • >> You would compare it to that.

  • You note the reactivity of one and then you compare it

  • to the one that you know.

  • >> Ahuh.

  • >> And it is-- and extrapolate that half life,

  • half life, half, half, half.

  • >> OK, so-- but the problem is you wouldn't know all the

  • conditions that's experienced over, you know,

  • say 100,000 years or something, right?

  • So, I mean, how do you--

  • you want to do this in a controlled circumstance.

  • You want to have everything just in a little test tube

  • where you know exactly what's been added

  • to the test tube, right?

  • But you don't want to wait around for 220,000 years

  • or 20 million years, what are you going to do?

  • OK, I'd like you to look this one up.

  • This is one that you should be able to design.

  • Look it up.

  • And then when we come back on Thursday we'll talk about this.

  • But I'd like everyone to have to look this up, OK?

  • This is important.

  • OK, let's talk--

  • let's summarized what we've been talking about in terms

  • of non-covalent interactions.

  • These are completely ubiquitous in biology.

  • Good news, we only have to learn two equations

  • which govern all interactions in chemical biology.

  • Those were the Coulomb's law for the charged-charged interactions

  • and the Lennard-Jones potential

  • for the uncharged interactions, OK?

  • And so if we know those two equations, we're set.

  • What's really-- and what's important to us is not

  • that we're going to be plugging in, you know, charge of this

  • and then, you know, radius of this.

  • What's important to us are the relationships,

  • the distance dependence,

  • the 1 over r squared versus 1 over r 6.

  • That type of distance dependence makes a big difference.

  • And knowing that sort of thing and having sort

  • of an intuitive grasp of that is going to be very important.

  • So-- and I'll just give you a quick example.

  • For example, we now know if DNA is negatively charged,

  • it's going to attract other charged ions to it

  • from great distances, right?

  • Because it's distance dependent, it's only 1 over r

  • to the 2nd power versus1 over R to the 6.

  • In addition, we've learned

  • that these non-covalent interactions are very sensitive

  • to the environment, the distance and the geometry.

  • Water is a really slippery molecule

  • to understand, to say the least.

  • Has a malleable structure

  • and it can dramatically alter the strength

  • of non-covalent interactions.

  • This makes it really tough for us to draw any generalities

  • because water is an intermediate lubricant between all

  • of these interactions and it plays a complicated

  • and sometimes hard to us-- it's hard for us to define role.

  • And there's still big arguments that are going

  • on in water chemistry to this day.

  • For example, there's an argument going

  • on about how many ions are found on the surface of water

  • or what's the pH at the surface of water.

  • And there's been a set of dueling papers

  • that have appeared that contradict each other.

  • The first paper had a title like the pH of the surface

  • of water is more acidic.

  • The next article-- the next article

  • by the competitor said the pH of water

  • at the surface is more basic.

  • And the two-- and these groups have been arguing backward

  • and forth and both making very reasonable arguments for years.

  • OK? The truth is what we found is actually it's somewhere

  • in between these two and you can actually see evidence

  • for either one and it turns

  • out to be a very minor effect that's not

  • so important in biology.

  • But the point is is that water itself is

  • such a complicated fluid

  • that we're still using the latest techniques to try

  • to understand it better.

  • It's not fully understood.

  • Hydrogen bonds have donors and acceptors

  • and they are also very susceptible to a competition

  • with water for those hydrogen bonds.

  • I would like you to know the approximate strengths,

  • the relative strengths, not the approximate

  • but the relative strengths, and distance dependence

  • of non-covalent interactions.

  • That's important.

  • OK, so that's a summary of chapter 2.

  • Any questions about chapter 2?

  • Yes, Chelsea.

  • [ Inaudible Question ]

  • Yeah, I really want you to know that.

  • OK, that's super important.

  • That's that Henderson-Hasselbalch equation.

  • That hopefully you learned in Chem 1,

  • you definitely need to know that.

  • Other questions?

  • OK, let's move on.

  • I want to talk to you about the structure of DNA.

  • This is the classic structure of DNA first proposed by Watson

  • and Crick in I believe 1952 or-- yeah, 1952, somewhere in there.

  • The structure of DNA has two strands running

  • in opposite directions to each other.

  • So they're anti-parallel to each other.

  • The strands are held together by phosphodiester bonds

  • which we'll look at more closely.

  • So, here's a schematic diagram of what the structure

  • of DNA looks like and here's a space filling view

  • where each one of these spheres is a van der Waals' sphere

  • to approximate where the atoms are,

  • where the outermost electrons of the atoms are.

  • One thing to notice is that DNA has two grooves, OK?

  • It has, yeah, the distance here

  • between these two strands is very close

  • versus the distance here between the two strands being much

  • further away.

  • These are going to be called the minor

  • and major grooves respectively.

  • And this is the origin of the fact that DNA is a double helix.

  • I think it's commonly thought that DNA is a double helix

  • because it's two relatively rod-shaped molecules

  • that are twisted with each other.

  • But that's actually not the case.

  • It's a double helix because it has a minor groove

  • and a major groove.

  • And I believe the next slide will show us that more closely.

  • OK. So, in blue, this is the major groove of DNA,

  • and in green, this is the minor groove.

  • In red, this is the phosphodiester backbone of DNA

  • that we've seen before, OK?

  • So again, notice that there are two helices

  • that are running parallel to each other, a major groove

  • and a minor groove, OK?

  • The structure of the bases is going to set up this major

  • and minor groove relationship.

  • As we will see shortly, DNA bases, base pairs form a U shape

  • and that U shape ensures that you're going to get a major

  • and a minor groove, where the inside of the U is going

  • to be this minor groove and the outside will be the

  • major groove.

  • But I'm getting a little bit ahead of myself.

  • The reason why this is important is as we'll see in a moment,

  • proteins like to interact with the major groove of DNA,

  • whereas they can't fit in to the much closer interstices

  • of the minor groove of DNA.

  • Rather small molecules will fit into this minor groove and try

  • to largely avoid the less cozy major groove of DNA, OK.

  • So, almost immediately we can start to make some predictions

  • about where stuff binds just knowing

  • that DNA is a double helix, double by virtue of the fact

  • that it has two parallel helices,

  • minor and a major groove.

  • So, this DNA structure immediately sets up replication.

  • This is the original 1953 paper by Watson and Crick,

  • and this is the very last sentence of the paper

  • in which they had this incandescent understatement.

  • It has not escaped our attention,

  • it has not escaped our notice

  • that the specific pairing we postulated immediately suggest a

  • possible copying mechanism for the genetic material, OK?

  • So, if you have two strands of DNA running anti-parallel

  • to each other, you can simply separate out the two strands

  • and then get a perfect copy of one strand over here

  • and a perfect copy of the second strand over here, OK?

  • So, here's the parent strand of DNA and again,

  • here are the two new strands in orange and blue.

  • Note too that DNA forms a right-handed helix.

  • OK, does everyone see that you can trace out along the right--

  • with your right hand over here the structure of DNA?

  • I think it's worth trying that.

  • Whereas your left hand kind of slips off,

  • it doesn't trace it out effectively, OK?

  • Does everyone see that?

  • So, it's DNA is always a right handed helix.

  • You know, so this beautiful structure of DNA is one

  • that was solved by x-ray crystal structure.

  • Before then, there were a large number of wrong,

  • incorrect predictions about DNA structure, including by people

  • who I, you know, think the world of, I think are, you know,

  • absolute heroes in science.

  • For example, the great Linus Pauling

  • who proposed a triple helix of DNA

  • where the phosphodiester backbones would be in the center

  • of the molecule and the bases would be out on the outside.

  • This kind-- this is somewhat--

  • this is intellectually attractive if you don't think

  • about the fact that you have two parents.

  • But furthermore it's attractive

  • because at least the base pairs would be out here in space

  • where they can interact with transcription factors.

  • We now know of course that that's not correct.

  • Instead, we'll take a look in a moment

  • at where the transcription factors interact.

  • Before we do, let's zoom out a little bit, OK?

  • So, DNA in the cell is concentrated in two regions,

  • a nucleosome in the prokaryotic cell, so it's kind

  • of concentrated in the very center of an E. coli cell.

  • In a eukaryotic cell of course, DNA is found exclusive--

  • is found in the nucleus and also the mitochondria

  • but let's just focus on DNA that's in the nucleus for today.

  • The bases themselves are connected together

  • to form oligonucleotides through this phosphate,

  • this phosphodiester functionalities, OK?

  • So this is called a phosphodiester functionality.

  • The DNA also has a directionality associated

  • with it, OK.

  • So there are-- if we look closely

  • at this deoxyribose base, there is a 5 prime end,

  • there's a 5 prime hydroxy over here

  • and a 3 prime hydroxy over here.

  • And so, the convention is to always write DNA

  • in the direction from 5 prime to 3 prime.

  • In the same way that we read English going left to right,

  • DNA is always read out 5 prime to 3 prime.

  • This is a really important convention, OK.

  • Everyone on the planet follows this convention and I'm going

  • to hold you to it as well, OK, because if you read the DNA

  • in the opposite direction you get a different--

  • or different word coming out, OK.

  • It spells something else that might not be this--

  • it will almost certainly not be the same thing

  • and it might actually be, you know,

  • might actually cause a lot of trouble.

  • So we're always going to be reading this 5 prime

  • to 3 prime directionality, so this sequence here would be read

  • out as A, C, G and T, OK, where the structures of A, C,

  • G and T are shown here, OK.

  • Don't bother memorizing this-- sorry.

  • Don't bother memorizing the structures of these.

  • I'll simply give them to you on the midterm, OK?

  • So, at a graduate level, you should know this.

  • Mariam will need to know this for her orals exam, but the rest

  • of you are in luck because I'm not going to test you on them

  • at least for this class, OK.

  • And again, the-- the directionality matters a lot.

  • If there is a 5 prime phosphate,

  • this 5 prime phosphate is indicated

  • by a lower case p. Finally, last bit of nomenclature,

  • oligonucleotides that are connected together are often

  • referred to as oligos and that's how I'll describe them.

  • OK, now I realize oligos is not the most descriptive

  • nomenclature because it just simply means an oligomer

  • or something, but that's the convention

  • that we've been operating under 50 years, OK.

  • So oligos will refer to oligonucleotides.

  • Typically DNA oligonucleotides compose of deoxynucleic acid.

  • OK. Now, even though DNA is--

  • the bases of DNA are called bases, it turns out they're not

  • that basic and few are protonated

  • at physiological pH. It's--

  • this is kind of one of those historical anomalies.

  • Here's a bunch of pKa's,

  • for example starting with triethylamine.

  • Here is the pKa of the protonated triethylamine,

  • the conjugate acid of triethylamine, pKa of 10.8.

  • Here is the pKa of cytosine, thymine, adenine and guanine

  • and you can see none of these would be remotely

  • considered bases.

  • Whereas this one over here, triethylamine definitely a base,

  • OK, as evidenced by the fact

  • that its conjugate acid is, you know, 10.8 pKa.

  • OK. Question so far?

  • All right.

  • Now DNA of course is missing a 2 prime hydroxyl, OK.

  • So here is RNA, it has a 2 prime hydroxyl over here.

  • This 2 prime hydroxyl makes RNA considerably less stable

  • than DNA.

  • I didn't point this out-- let me go back to it--

  • when we talked earlier about half lives.

  • Let me just zoom back to that really fast.

  • The half life of RNA is considerably lower

  • than the half life of DNA, OK.

  • So, here is the half life of RNA, 220,000 years,

  • whereas the half life of DNA

  • at 220 million years is much, much greater.

  • OK, a thousand fold difference in stability

  • for the phosphodiester backbone of the DNA

  • versus the phosphodiester backbone of RNA.

  • This makes sense, OK?

  • The 2 prime hydroxyl of RNA sets you

  • up for hydrolysis using an intramolecular attack, OK?

  • So, here's again the structure of RNA.

  • Here's the 2 prime hydroxyl.

  • This 2 prime hydroxyl can act as a nucleophile

  • to attack the phosphodiester backbone

  • of the RNA setting up cleavage.

  • Does anyone want to see the mechanism of that?

  • OK, all right, let's take a quick look.

  • OK, so in this mechanism--

  • [ Pause ]

  • Let me just draw up the structures

  • and then I'll blank the board.

  • OK, one second.

  • OK. So, in this mechanism, here's our structure of--

  • [ Pause ]

  • OK, so here is our backbone structure of RNA

  • and I'm just going to draw this as base over here, OK?

  • OK, so if there is any base that's present,

  • let's just say hydroxide,

  • this can deprotonate the 2 prime hydroxyl, giving us an alkoxide.

  • [ Pause ]

  • -- adjacent

  • to the phosphodiester backbone of the DNA.

  • This neighboring alkoxide can now attack the backbone,

  • the phosphodiester backbone,

  • giving you a five-membered ring intermediate.

  • OK, which I'll show down here.

  • [ Pause ]

  • Five-membered ring intermediate

  • and this intermediate collapse leading to cleavage of the RNA.

  • OK. So here is that collapse.

  • OK. So, we're going to be making two strands of RNA

  • that are separated from each other.

  • [ Pause ]

  • OK, so here's one strand over here

  • and then here's the second strand down here.

  • [ Pause ]

  • OK. I'm going to just differentiate this

  • as base 1 and base 2.

  • OK. So, notice that the strand has actually cleaved apart.

  • You can then hydrolyze this phosphodiester backbone,

  • this phosphodiester back

  • to a phosphomonoester using another equivalent of hydroxide.

  • [ Pause ]

  • And then finally, collapse

  • of this tetrahedral intermediate gives us the product.

  • OK. Questions about this mechanism?

  • All right, now notice again, if DNA lacks this 2 prime hydroxyl

  • over here, and I just want to make this totally explicit,

  • I'm going to label it 2 prime hydroxyl, 3 prime, 5 prime.

  • OK, so DNA lacks the 2 prime hydroxyl

  • and therefore does not have an opportunity

  • for this intramolecular nucleophilic attack

  • on the phosphodiester backbone.

  • So, for this reason, DNA is a thousand times more stable

  • than RNA, right, lacking this intramolecular nucleophile.

  • Makes sense?

  • Questions about this?

  • OK, let's go back.

  • Turns out that when you look at the liability of the bases,

  • we see actually a different trend.

  • OK. And actually I think I'm going to skip that.

  • OK, moving on.

  • OK, I'd like you to learn what I just told.

  • Don't worry so much about the base stability.

  • DNA bases are subject to important modifications.

  • These modifications have dramatic roles

  • on the phenotype of organisms, OK?

  • So, for example, methyl groups are often transferred to DNA.

  • I showed structures of DNA bases.

  • Again, they're subject to massive modification

  • by methyltransferases and other modifications.

  • So, for example, here's 5-methylcytosine over here,

  • 4-methylcytosine and then N6-methyladenine.

  • These modifications can dramatically alter

  • transcription levels.

  • They can set up the organism

  • to transcribe some genes more often, OK.

  • So, for example, lacking pigmentation,

  • the genes that encode pigmentation are

  • in my skin cells, my epidermal cells,

  • yet they're not transcribed very often.

  • And so, it's likely that my DNA has not been methylated

  • in those regions.

  • However, when I go out and spend a lot of time in the sun,

  • I'm getting additional little spots called freckles

  • which are resulting from methylation

  • of those DNA sequences which in turn then turns on transcription

  • of the pigmentation and results in freckles, OK?

  • So, the environment, the environment that you're exposed

  • to can alter these transcription patterns.

  • It's one of the ways that organisms like ourselves respond

  • to changes in the environment.

  • It's a very important way in fact.

  • And oftentimes this goes through methylation of DNA.

  • This DNA methylation is really as important as sequence

  • or genomics, and this is an area called epigenetics.

  • That's really an area of very active research that's taking

  • place in chemical biology.

  • OK. So, we've looked at structures

  • of the bases themselves, we've looked at structures

  • of the phosphodiester backbone,

  • let's start putting things together to start

  • to understand the structure of DNA.

  • The bases themselves are slightly U-shaped, OK.

  • So, here's a base between A and T, adenine and thymine.

  • Notice that this base is composed of two hydrogen bonds.

  • Here's a base of G and C which has three hydrogen bonds.

  • But notice more importantly that the bases are U-shaped

  • or equally importantly, OK, U-shaped here.

  • The inside of this U where the R is going

  • to be towards the ribose, the deoxyribose ring,

  • the inside of this U is going to form the minor groove

  • which I've showed you on an earlier slide.

  • The outside of the curvy part of this U is going

  • to form the major groove.

  • As you have these U's that are stacked on top of each other

  • and each one is slightly offset with each other, this is--

  • outside is going to result in a much bigger helix

  • than the inside over here, OK?

  • And here's what this looks like.

  • OK. So, here's a trace of the phosphodiester backbone

  • and then I've highlighted just one Watson-Crick base pair, OK.

  • And again, notice that it's U-shaped,

  • that there's more section traced out over on this side,

  • that will be the major groove,

  • and the inside will be the minor groove.

  • Furthermore, the green arrows define hydrogen bond donation

  • and acceptance by the base pair.

  • And notice that there is a pattern to this,

  • that there is an acceptor-acceptor donor, OK.

  • So, this is a donor acceptor donor over here.

  • So, there's actually a little bit of a pattern

  • to whether this is a G on this side and a C on this side or C

  • and G on the opposite sides.

  • So, in other words, A and T are not the same as T and A

  • because they're going to present a different pattern

  • of hydrogen bonds for molecular recognition where again,

  • the proteins are going to be--

  • the transcription factors are going to be interacting

  • over here in the major groove

  • and small molecules would be interacting

  • in this minor groove down here.

  • I should mention that there's also some protein DNA

  • interaction in the minor groove.

  • It tends to be more minor, however.

  • OK, let's take a close look at one example

  • of a transcription factor and how it works.

  • This is the transcription factor, Fos-Jun,

  • it consist of a leucine zipper which is two helices

  • that interact with the DNA like chopsticks, OK.

  • So, these are fitting neatly in the major groove.

  • It turns out the major groove has exactly the right size

  • to accommodate an alpha helical protein, OK.

  • So, this Fos-Jun is absolutely perfect.

  • It fits neatly in the major groove.

  • Now, these hydrogen bond donating functionalities are

  • going to then read out the sequence of the DNA and look

  • for a specific sequence of DNA to interact with,

  • trying to form complementary hydrogen bonds,

  • trying to form complementary van der Waals interactions

  • in this sequence, OK?

  • Let's take a closer look now

  • at the forces holding together the DNA double helix.

  • Earlier, I alluded to the fact

  • that AT base pairs form two hydrogen bonds

  • and GC base pairs form three.

  • Which one is stronger, just, you know,

  • from a crude approximation?

  • Yeah, three is stronger than two, right?

  • OK, so in addition to this, the DNA structure is held together

  • by pi stacking between the bases.

  • Again, this is a face-to-face interaction.

  • Typically not perfectly face-to-face,

  • rather it's typically offset.

  • And that offset needs the bases to stack not directly

  • on each other but slightly twisted from each other,

  • setting up this helical structure

  • that we're now familiar with.

  • In order for this base pairing to take place, the base pairing

  • that I showed on the previous slide,

  • you need a particular tautomer of these aromatic rings, OK.

  • And the first one that should strike you as funny is this one

  • over here, because you can imagine another resonant

  • structure that would make this C aromatic, right?

  • Notice that the C has--

  • is non-aromatic in this tautomer shown here, right.

  • It only has two pi electrons rather than the requisite six

  • that it would need to be aromatic.

  • OK, that's almost-- that's bizarre to begin with, OK?

  • So, what's going on here is that there is a preference

  • for this tautomer versus this one.

  • This one is actually thermodynamically more stable

  • and the reason for this is that the carbon-oxygen double bond

  • over here is quite strong.

  • I will tell you that I think any chemist looking

  • at this could not have predicted this in advance,

  • and in fact actually this tremendously slowed structure

  • determination of the original structure

  • of DNA back in the 1950s.

  • Watson and Crick were physicists and weren't as familiar

  • with the whole notion of tautomerization

  • as their chemical counterparts were racing

  • to solve the structure of DNA.

  • And so for them, this did not look funny whereas

  • to us I think it does look funny, right,

  • because it lacks aromaticity whereas a structure

  • on the left is aromatic.

  • Again, this happens to be just a little bit more stable

  • because of the strength of the carbon-oxygen double bond,

  • but I don't think anyone would have predicted that, OK?

  • I think now we, you know, with our 21st century guys,

  • we could predict it, but going back in time,

  • I don't think we could have predicted it so readily.

  • Similarly, over here, these amidines are actually going

  • to be more stable in the aromatic structure

  • than an amidine structure.

  • And in this case, that's due to the much poor overlap

  • between a carbon-nitrogen double bond

  • than a carbon-oxygen double bond, OK.

  • So, all of these lead to the base pairs

  • with the hydrogen bonding preferences

  • that are shown here, OK?

  • Whereas for example, this is a non-aromatic ring

  • that could aromatic if it tautomerize,

  • but it doesn't prefer to be tautomerized.

  • Whereas this one over here seems to prefer to have an amidine

  • in this structure because of the strength

  • of a carbon-nitrogen double bond, OK?

  • And here is another example of that over here.

  • This one prefers aromatic

  • because carbon-nitrogen double bonds are relatively weak.

  • OK, pretty interesting.

  • Unnatural bases, however,

  • could dramatically shift these preferences for tautomerization,

  • and a good example of this is 5-bromouracil, OK.

  • So, if this compound here is fed to organisms,

  • what happens is an unusual tautomerization preference

  • where the enol form of bromo-U is actually more preferred

  • than it would be if there was no bromine over here, OK?

  • So, most of the time, it forms the regular base pair, however,

  • some of the time,

  • it can actually form the incorrect base pair

  • because it can actually more readily access this enol form

  • of the base, OK?

  • So, that's due to the electron withdrawing functionality

  • of bromine over here, OK?

  • That's changing this tautomerization preference.

  • The consequences of these are really dramatic.

  • Because the Watson-Crick base paring is not followed

  • as closely, what ends up happening is the DNA comes

  • out with all kinds of bizarre breaks and lesions, OK.

  • So here are chromosomes from a normal organism.

  • I think it's a hamster in this case.

  • And then here's chromosomes from hamsters that were exposed

  • to bromouracil and you could see they have all kinds

  • of bizarre shapes to them, things are incorrect, OK.

  • So, this causes cancer and breakages in DNA

  • which then eventually lead to cancer, cancer cells

  • and tumors in the organism, OK.

  • All right, so furthermore,

  • it turns out that we can test this--

  • the importance of the strengths of these hydrogen bonds

  • by synthesizing unnatural bases.

  • So, this is one of the great things about chemical biology.

  • If you have this hypothesis that something is important,

  • then you could test that hypothesis

  • by synthesizing compounds which are say missing

  • that key functionality.

  • So, from Watson and Crick, we expect to find

  • that hydrogen bonds are holding together the structure of DNA

  • and chemists went out and synthesized variants

  • of DNA bases that were lacking

  • that ability to hydrogen bond, OK?

  • Structures of these are shown here, OK?

  • So, for example, this compound here is simply a pyrene in place

  • of a base and it actually prefers to base pair

  • with a missing base over here, OK?

  • So, these guys over here, no hydrogen bonding,

  • no hydrogen bonding over here and yet these actually prefer

  • to pair with each other, OK.

  • So, you can actually have completely unnatural bases

  • missing hydrogen bonds that is-- are yet able to form base pairs

  • with each other preferentially.

  • What this tells us is that there's more going

  • on in DNA structure than simply hydrogen bonding.

  • Hydrogen bonding is a nice simplifying assumption

  • for our biochemical friends or molecular biology friends,

  • but in actuality, the pi stacking

  • of DNA is a driving interaction, the edge to edge interactions

  • of aromatic functionalities are also driving these interactions

  • between the strands of DNA.

  • And so, while we can do quite a bit with hydrogen bonding,

  • there's quite a bit more that's left to be explored.

  • OK, last thought, I've been showing you--

  • or it's not last thought, I've been showing you--

  • Oh, before I get to that, here's-- here for example is--

  • this illustration here emphasizes the importance

  • of pi stacking in here, OK?

  • So, one thing is that bigger bases tend to pi stack better,

  • for example, the guanine base tends

  • to pi stack better than say cytosine.

  • All right, in addition,

  • I've been showing you Watson-Crick base pairing

  • where it's a canonical base pair,

  • G's and C's have three hydrogen bonds,

  • A's and T's have only two.

  • Other kinds of hydrogen bonding possibilities are not only

  • possible but have been observed.

  • These were proposed by Karst Hoogsteen

  • and we observe this a lot in RNA structure.

  • We don't necessarily see this in DNA, but we definitely see this

  • in RNA and they're going to come up later.

  • So, I'll just show you the structures here.

  • This is an alternative to the usual AT base pair

  • and this is an alternative to the usual CG base pair.

  • This one being driven by a protonation event,

  • protonation of this nitrogen over here, OK?

  • So, this is actually-- these are sort of edge

  • to edge interactions rather than the sort

  • of neat more typical Watson-Crick base pair.

  • OK. Any questions about the structure of DNA?

  • Anything, whatsoever?

  • I want to change gears then and start talking

  • about how small molecules interact with DNA.

  • The first mode that small molecules can interact

  • with DNA is to actually slip into this pi stack of DNA.

  • So, aromatic compounds can slide into the pi stack of the DNA

  • and we're going to see the consequences

  • of this can be quite destructive.

  • Let's take a look at some examples, this is a class

  • of molecules called intercalators,

  • meaning that they intercalate into the pi stack of the DNA,

  • they get integrated into the DNA structure.

  • So, in order to fit into this pi stack,

  • these molecules must be also hydrophobic

  • and also aromatic, right?

  • They will form competing pi-pi stacking interactions

  • with the DNA and so they must also be aromatic.

  • Note too that in order to force the way into the pi stack,

  • these molecules force the DNA's double helix to slightly unwind

  • to accommodate the DNA intercalator.

  • Here are some examples of this.

  • These are examples of intercalators.

  • Notice that they are all aromatic compounds.

  • They're all flat and aromatic to slide into the pi stack.

  • Many of these molecules also have positive charge.

  • Positive charge is useful, right,

  • because DNA with the phosphodiester backbone

  • of the DNA is negatively charged.

  • This gives the molecule a way to be attracted in the DNA

  • through a long range charged-charged interaction,

  • right?

  • So these molecules are going to seek out DNA

  • like a homing missile.

  • And once they slide into the pi stack,

  • the consequences can be pretty bad

  • or actually fairly useful, OK?

  • Let me show you an example of a useful intercalation

  • over here on the right.

  • This is actually an agarose gel which is an important way

  • that chemical biology laboratories separate

  • out DNA structures.

  • Different DNA sequences can be separated out on the basis

  • of their size using these agarose gels.

  • I'll show you what that looks

  • like in a couple of slides from now.

  • To visualize the DNA, however, this molecule over here,

  • ethidium bromide is incorporated to the gel

  • and it gets concentrated into the DNA

  • by an intercalation interaction.

  • So, it slips into the pi stack of the DNA

  • and it's a fluorescent molecule,

  • many aromatic compounds are fluorescent.

  • We've talked about fluorescence before.

  • And so, you can actually shine UV light on the gel

  • and wherever you see this-- these pinkish bands,

  • that's where the DNA is present.

  • And so you can actually take a razorblade for example and cut

  • out the DNA of a particular size.

  • Here's a couple of more DNA intercalators.

  • Here's one that's designed to intercalate

  • and then have a little linker and then intercalate

  • down below the compound.

  • Here's what it looks like structurally,

  • so there's intercalator, linker,

  • intercalator up here for example.

  • I think that-- this it right over here.

  • These are also compounds that are used to treat cancer.

  • So, dynemicin, adriamycin are used as anticancer compounds

  • or some of the first rounds of anticancer compounds

  • that are used as chemotherapeutics.

  • And we'll talk more about their mechanism

  • of action later in the class.

  • We're not quite there yet.

  • OK. Let's stop here.

  • When we come back next time, we'll be talking more

  • about the structure of DNA. ------------------------------0b468b059f88--

[ Silence ]

Subtitles and vocabulary

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