Placeholder Image

字幕列表 影片播放

  • [ Silence ]

  • >> OK, so, hopefully that wasn't too painful.

  • I realize it might have been--

  • if you even put your name on the piece of paper,

  • you will get some points that I promise you.

  • In addition, next time I say that there's something I'd

  • like you go look up, I hope you take I mores [phonetic]--

  • you take it seriously.

  • But don't panic.

  • Hopefully, that's not too painful.

  • Let me just tell you very quickly the answers, OK?

  • So, last time we saw that half life of DNA was

  • on the order of 220 million years.

  • And I challenge you to go out and find

  • out what it-- had to measure it.

  • And at least one person came to my office hour with the solution

  • which is to heat the DNA so you can heat it up

  • and to the high temperature, measure the half life

  • at the higher temperature and extrapolate back

  • to the lower temperature.

  • And you don't have to provide me with a lot of equations

  • to show that that's true.

  • But if you just write heat on probably--

  • number two, I will accept it, OK?

  • But again, if you have your name on a sheet of paper

  • with your UCI ID number, we'll give--

  • you'd get at least some points, OK?

  • OK. All right, we're back to normal stuff now.

  • I want to pick up where we left off last time.

  • We were talking about structure of DNA and reactivity of DNA,

  • and we saw last time-- oh gees.

  • What happened here?

  • One moment.

  • I just [inaudible] this.

  • OK, we saw last time a lot about Watson-Crick based-pairing

  • and structures of DNA.

  • And it's B-DNA form.

  • Now, one thing I got asked about immediately

  • after the lecture was, you know,

  • ethidium bromide is commonly considered to be a carcinogen.

  • This molecule over here that we discussed as an interculator

  • of DNA is also a carcinogen.

  • It's a molecule that causes cancer and that's the reason why

  • when we work with it in the laboratory, over here,

  • we're exceptionally careful to keep it off of our skin, OK?

  • Now, yeah, it's a-- it articulates into DNA

  • and DNA interculators can caused,

  • can help to initiate carcinogesis, can--

  • it helps initiate cancer in the following ways.

  • Number one, it distorts the structure of DNA.

  • I showed you that on the previous slide over here

  • where I showed you how the structure of the DNA has

  • to unwind to accept interculator.

  • So it distorts the structure of DNA.

  • Number two, it places a hydrophobic functionality

  • in the center of the DNA.

  • This hydrophobic functionality can inappropriately attract

  • transcription factors to the DNA setting off incorrect

  • transcription, OK?

  • So those are two possible modes that this can start to cause--

  • this can start to initiate improper self responses

  • that will eventually lead to cancer.

  • And I want to talk about cancer today.

  • It's one of the prime topics in this class.

  • It's something we've already spend quite a bit

  • of time discussing in many different context

  • and we'll certainly be talking about it

  • in the context of DNA, OK?

  • So, any other questions about what we saw last time.

  • That was a really good question.

  • OK, last thought about ethidium bromide,

  • although it is definitely-- it is a cancer causing agent

  • and it's something you definitely do not want to ingest

  • or you definitely want to keep it off your hands.

  • I don't want to exaggerate its carcinogenic potential.

  • It's also used as an antibiotic in sheep.

  • I don't know if it's still is, but it was for a while.

  • It is-- It does have some other--

  • and so it's actually fed to sheep.

  • I don't know why sheep.

  • But in any case, as a tumor forming agent,

  • its activity is rather modest specially compared

  • to the other cancer causing small molecules that we're going

  • to see very shortly, OK?

  • All right, well let's dive right in this.

  • So last time, I was showing you that DNA likes

  • to form Watson-Crick based-pairing.

  • This is C binding to Gs.

  • A is binding to Ts and we saw that GCs,

  • Watson-Crick based-pairings have three hydrogen bonds

  • and ATs has only have two, OK?

  • Now, if we know that AT based-pairs are weaker than GCs,

  • there's a very simple that we can use to estimate the strength

  • of any two DNA strands taking together.

  • This rule is called the Wallace Rule and I'd

  • like you to memorize it, OK.

  • The Wallace Rule tells us

  • that the approximate melting temperature

  • for DNA sequence is equal to 2 times the number

  • of AT base pairs plus 4 times the number of GC base pairs

  • in degree C. This melting temperature is where 50 percent

  • of the DNA is no longer forming a double-stranded structure.

  • Let me explain.

  • So0 when you do a UV/Vis scan of DNA, this is what--

  • here's what you find, OK?

  • So this is absorbance on the Y axis and wave length,

  • the nanometers on the X axis.

  • And this is single-stranded DNA

  • and then this is double-stranded DNA in blue.

  • And to get this, you simply add higher temperature.

  • So at higher temperature, the DNA stands melt apart.

  • In other words, they separate out into the two single strands

  • and so at 82 degrees, this is the single-stranded

  • and 25 degrees, this is the double-stranded DNA.

  • Notice that the absorbance at 260 nanometers is higher

  • for the single-stranded DNA than the double-stranded DNA.

  • So you can use that change in absorbance to follow whether

  • or not your DNA is single-stranded

  • versus double-stranded.

  • And so, you can do this while at the same time,

  • you ramp up the temperature and at

  • about 50 percent-- and so here it is.

  • Pure double-stranded DNA, lower absorbance and then here it is

  • at a higher temperature

  • where it's entirely single-stranded DNA.

  • And the approximate 50 percent part

  • where 50 percent is melted is called the melting temperature.

  • And again, you can estimate what this melting temperature is

  • using this Wallace rule formula.

  • People in chemical biology laboratories uses Wallace rule

  • formula on a daily basis, OK, certainly in my laboratory

  • and certainly and probably five

  • or six other laboratories here at UCI Irvine.

  • So, I'd like you to memorize this rule.

  • It's incredibly useful.

  • Now, you're probably wondering, big deal.

  • So I know whether or not something melts, whether

  • or not it forms double stranded DNA.

  • If you know the temperatures

  • that it forms double-stranded DNA, you can start

  • to design structures made out of DNA.

  • Let me show you.

  • OK. So here are structures of DNA where it's single-strands

  • of DNA that are now hybridizing against each other

  • and forming elaborate patterns such as this pattern here.

  • And here's an atomic force microscope image

  • of this double-stranded DNA and you could see it's all--

  • it's forming this exact costly pattern

  • that was designed using something just a little bit more

  • complicated than the Wallace rule

  • which I asked you to memorize, OK?

  • It gets even better.

  • Check this out.

  • OK. So this is worked done by Paul Rothemund and colleagues

  • and coworkers at Caltech.

  • And he's using the Wallace rule to design DNA that folds

  • up into happy faces over here or check out this map

  • of the world written out of DNA that's been folded

  • up with itself.

  • OK. So that Wallace rule that I asked you

  • to memorize is actually pretty powerful.

  • You can develop whole structures of that stuff.

  • Now, exactly what the structures of DNA are going to be useful

  • for is not a hundred percent known at this point.

  • There are sort of a frontier

  • in chemical biology building structures out of DNA

  • and then trying to do something useful with them.

  • I've only seen one paper in 20 years of staring

  • at these beautiful pictures that has convinced me

  • that maybe there might be something useful about this.

  • And in that paper this DNA-- a DNA structure like this one,

  • not the happy face but something elaborate was used

  • as a delivery vehicle that bound to the surface of cells

  • and then dislodged drug therapeutic.

  • And it's possible that our future might feature many more

  • of these sort of examples of nanometer scale structures

  • that are designed by you, by the people on this room

  • to have specific properties such as binding

  • to specific cell types, unloading cargos

  • at specific times, et cetera.

  • This is a really exciting frontier and I encourage you

  • to think about it in your proposal preparation

  • because it's an area that's kind of wide open

  • for creative-- for creativity.

  • OK, so in addition to those sort of macro structures of DNA

  • that we saw, short stretches of DNA can also fold

  • and there's a couple of canonical structures

  • that we're going to see time and again.

  • One of these for example is called-- are called hairpins.

  • One of these is called the hairpin.

  • And so, these consist of a sequence

  • that folds back on itself.

  • Notice that it satisfies all

  • of the Watson-Crick base pairing requirements, G is to Cs.

  • A is to Ts and it forms something that looks kind

  • of like an old-fashioned hairpin.

  • OK. And it looks structurally like this.

  • This is the x-ray crystal structure of what it looks like.

  • And again, we're going to see this quite a bit.

  • OK. So DNA has a propensity to fold on itself.

  • It wants to form Watson-Crick base pairs.

  • It wants to form Watson-Crick base pairs with other sequences.

  • It wants to form base pairs with itself.

  • And so for this reason, DNA is rarely found.

  • It's sort of an unwound

  • which rarely found a single-stranded DNA

  • for one thing.

  • And furthermore, it's rarely found in the sort

  • of canonical B-DNA conformation that I've been showing you

  • where it says nice right-handed coiled double helix.

  • Rather, in cells, we typically find DNA in a wound

  • up configuration called a supercoil.

  • So a supercoil is where you take a coil

  • such as this old-fashioned telephone cord

  • which I'm sure is unfamiliar to everyone in this room.

  • But back in my day kids, we used to have this

  • and it would form the sort of supercoils

  • and it drove you nuts.

  • You know, you constantly be on the phone trying

  • to untangle the darn thing.

  • In fact, she [phonetic] was kind of a nice thing to do

  • because if you're on the phone with the tense conversation

  • or something, it gave you a task to take your mind off

  • of the annoyances that you're dealing with on the phone.

  • But anyway, so this is called supercoiling

  • and this is an example of a DNA plasmid which is a sequence

  • of DNA that forms a circle.

  • OK. So this is a nice plasmid DNA and again we rarely find it

  • in this sort of B-DNA where it's completely unwounded

  • configuration, rather it likes to twist up.

  • So here's a little twistiness.

  • Here's more twistiness, even more twistiness

  • and then finally even-- the most twistiness.

  • OK. So that's really the structures of DNA that we find.

  • For short sequences, we find it wound

  • up with itself to form hairpins.

  • I showed you that structure first.

  • For larger sequences like plasmids,

  • we find it supercoiling and twisting itself up.

  • And then for even larger sequences like you carry

  • out at genomes which I'll show you on the next slide,

  • it gets even more twisty than that.

  • OK. So-- oh, before I get to that,

  • here's why it has to get so twisty.

  • This is one of the benefits of having it coiling up on itself.

  • This is just the DNA in a single E. coli cell.

  • This is a classic picture of that DNA

  • where really toward [phonetic] the force microscopy was used

  • to lice the cell and spew out all of its DNA

  • and you could see it's just an enormous amount of DNA

  • for such a small cell.

  • And the human cells face a similar compaction problem,

  • right.

  • The human genome would be a rough--

  • roughly 1.7 meters if completely unwound.

  • So this property to want to wind on itself,

  • to coil up with itself is actually a really important one.

  • And so, when we look at human DNA,

  • we find that it's compacted up into chromosomes.

  • I'm sure-- hopefully, this is not an unfamiliar concept.

  • In fact actually, I showed you examples of chromosomes

  • from I believe it was guinea pigs when we talked

  • about bromouracil earlier this week.

  • In any case, so here're some images

  • of human chromosomes over here.

  • And here's how it's compacted up.

  • So, the DNA, here's the B-DNA

  • that we've been looking at over here.

  • I admit this is a terrible rendering because it looks

  • like it's a single helix.

  • Note that it's lacking the major groove.

  • This is-- or the lacking the minor groove.

  • This is one of my pet peeves

  • about artistic depictions of DNA.

  • But I'm going to let it slide by at this moment.

  • In any case, so here's the regular B-DNA.

  • The B-DNA is going to wrap

  • around the protein structure called the histone

  • and this is going to act like the spools for thread.

  • And then these histones are going to coil up

  • and those coils are going to coil up further

  • until eventually you get

  • to something that's massively compacted into a chromosome.

  • Now, the problem of course and this is the problem

  • if you're on the phone as well.

  • So imagine you're on the phone

  • with these old-fashioned telephone cords,

  • what you find is it has very hard to untangle the DNA

  • without disconnecting the cord and making a little break in it.

  • OK. And that's one of the annoying things

  • about supercoiling of old-fashioned telephone cords.

  • Similarly with DNA, where you have this one 1.7 meter long

  • object and a 20 microne long cell, there has to be a solution

  • to uncoil the DNA and the solution is

  • to make transient breaks.

  • So shortly breaks in the DNA using an enzyme called

  • DNA gyrase.

  • And here's an example of this.

  • This is a DNA gyrase that acts kind of like scissors, OK?

  • So, notice that it's a dimer.

  • The two arms down here can open up and the thing could just grab

  • onto the DNA and then introduce two breaks

  • on the DNA allow the super-- the DNA to relax, to uncoil

  • and then it gets rejoined.

  • This turns out to be a Achilles' heel

  • for bacterial cells, for cells in general.

  • In other words, it's a spot that can be targeted

  • with the antibiotics and we'll be talking about this.

  • We'll be talking quite a bit about different ways

  • that antibiotics works.

  • So antibiotics are pharmaceuticals, therapeutics

  • that are given to patients

  • to eliminate bacterial infections or fungal infections.

  • In this case, this is a really effective antibiotic that's

  • giving quite a bit.

  • This is the antibiotic Cipro which I'm sure many of you

  • in this classroom have taken at one point.

  • I read somewhere that like 85 percent of American women come

  • down with a UTI, a urinary tract infection at some point,

  • the first line of antibiotic used

  • against that is often Cipro.

  • And Cipro works by inhibiting the DNA gyrase of bacteria.

  • OK. Here's another one as well, another inhibitor as well.

  • OK, so let's talk a little bit more

  • about these bacterial plasmids because I want to transition

  • into a discussion of biotechnology and cutting

  • and pasting DNA in large scale.

  • So, oftentimes,

  • DNA is transferred amongst organisms using plasmids.

  • Plasmids are short circular stretches of DNA.

  • They need to have-- all plasmids need to have two properties, OK?

  • They must have sequences that encode two items.

  • Item number one is an origin of replication.

  • The origin of replication abbreviated ORI is the spot

  • that somehow convinces the cell that's taking up the plasmid

  • to start transcribe-- or start replicating that DNA, OK.

  • So that origin of replication kicks off replication

  • of the plasmid.

  • Without that, the plasmid would just be there

  • and it wouldn't get copied, and it wouldn't get passed

  • on to the next little guy, OK?

  • So that's absolutely essential.

  • The other essential thing is that the plasmid has

  • to confer some advantage onto the new host, OK?

  • In other words, the new bacteria has to take it

  • up and say, "Oh, yeah.

  • This is useful."

  • Otherwise, the plasmid will get quickly shunted aside

  • because cells are under a lot of pressure.

  • They have a lot of works to do and they have a limited amount

  • of resources, carbon, nitrogen, oxygen, things like that

  • that are available to do all of the priorities that they have.

  • OK, and that's kind of a long winded way to say that this has

  • to confer some resistance oftentimes

  • to some sort of antibiotic, OK?

  • So, resistance markers are sequences of DNA

  • that encode a protein that confers resistance.

  • OK. So for example, you can have a resistance marker

  • that encodes resistance to the antibiotic tetracycline

  • and this gene will work

  • by actively pumping tetracycline out of the cell, OK?

  • So, when the-- when the cell takes up this plasmid,

  • it's going to synthesize this pump that goes to the surface

  • of the cell and then every time it gets tetracycline it just

  • pumps it out of the cell furiously

  • and that allows the cell to live.

  • So only the cells that have the plasmid will survive an

  • onslaught of the antibiotic tetracycline

  • which again is a very common antibiotic

  • that I mentioned a few of you have encountered.

  • It's often used, for example, I believe for acne treatment.

  • OK. Here are some other classic examples of other antibiotics

  • that are used in my laboratory

  • and other chemical biology laboratories

  • as selection markers for drug resistance.

  • And the way this works is will coat the cells,

  • the bacterial cells on a plate and the plate has an agar

  • which I'll show you the structure very shortly.

  • It's isolated from seaweed.

  • It's a basically just a polymer and inside this agar plate,

  • we'll have some concentration of one to these antibiotics.

  • And so the only colonies and each one

  • of these circles is a colony that appears

  • on this plate are colony or bacteria cells that have taken

  • up the plasmid because now,

  • those cells are resistant to the antibiotic.

  • OK, so here is another way that this can work.

  • So, antibiotic that's commonly used is chloramphenicol.

  • Chloramphenicol inhibits the ribosome.

  • We talked about the ribosome before.

  • The drug resistance gene encodes an enzyme called chloramphenicol

  • acetyltransferase or CAT

  • and this enzyme transfers an acetyl group

  • to a primary hydroxyl of chloramphenicol, OK?

  • So, here's the acetyl group and acetyl-CoA and it's going

  • to get transferred to this primary hydroxyl in a reaction

  • that essentially disarms the chloramphenicol preventing it

  • from binding to the ribosome and allowing the cells on this plate

  • of chloramphenicol to live.

  • Third, a very common resistant marker, OK,

  • so I've shown you tetracycline.

  • We've talked about chloramphenicol.

  • Third one, the third one are beta-lactam antibiotics

  • of these sort.

  • Notice that this is a beta-lactam.

  • A lactam, of course, is a cyclic ring that has an amid bond

  • on it, and this is beta because it has two carbons, alpha, beta.

  • And so, that's the beta-lac--

  • that's the origin of the beta lactam nomenclature

  • which I know we talked about in 51C.

  • Hopefully, you encountered as well.

  • In any case, beta-lactamase is a--

  • enzyme encoded by the beta-lactamase gene

  • that confers the ability

  • to hydrolyze this amid bond that's part

  • of the beta-lactam ring.

  • And this is a very common gene that's found out on environment.

  • So, you can probably scoop up, you know,

  • some dirt over here just outside Rowland Hall

  • and you can readily find this beta-lactamase gene.

  • And so for this reason,

  • medicinal chemists are constantly making new

  • antibiotics that avoid that environmental drug resistant

  • that sort of omnipresent, OK.

  • So for example, here are two different kinds

  • of beta-lactam based antibiotics

  • and notice the structural differences.

  • This one has this benzoyl functionality over here.

  • This one has a carboxylate--

  • carboxylic acid and a phenol group instead.

  • And so, all of those little differences change,

  • affect the ability of the drug resistance enzyme,

  • the enzyme conferring drug resistance to bind

  • to the antibiotic and hydrolyzes them and bond.

  • Maybe this carboxylate sticks into the protein

  • and prevents the binding and that's a useful thing.

  • OK. So we're constantly on the hunt for new antibiotics

  • because the antibiotics we have seem to allow very rapid risk--

  • evolution of drug resistance and so,

  • there's a constant need really for our society

  • to develop new classes of antibiotics

  • that are more effective than the previous generation.

  • And in the last 10 years or so, there's been a real renaissance

  • of research in this area

  • to develop even more effective antibiotics.

  • OK. Let's get back to our discussion of DNA structure.

  • I showed you structure of plasmids.

  • Here's structure of a eukaryotic genes,

  • eukaryotic DNA that's wrapped around nucleosome, et cetera.

  • I don't have very much more to say about that.

  • Let's take a closer look however

  • at the structure of these histones.

  • So, the histones are these hexameric proteins shown here

  • in yellow and green where in green,

  • these are positively charged residues, OK?

  • So those are residues whose positive charge can interact

  • with negatively charged phosphodiester backbone

  • of the DNA.

  • Charge-charge interaction.

  • Nice long range interaction.

  • This wrapping up though basically hides the DNA

  • and prevents it from being transcribed.

  • When it's wrapped up around the histone it can't be a read

  • out and, you know, use for transcription.

  • And so, basically, whether or not the histone is wrapping

  • up things, it's-- it controls transcription

  • and controls packaging.

  • So these proteins over here are very tightly regulated

  • as to whether or not they're going to be binding to the DNA

  • and one easy way to do this regulation is

  • to acetylate the lysines side chains, OK, and I'll show you

  • that on the next slide.

  • OK. So first, this is the structure of lysine.

  • Lysine has a primary mean

  • and here's lysine within acetyl group.

  • What do you think the charge is if something has a primary mean

  • at neutral pH which is the pH roughly

  • of the cell approximately?

  • So it has a primary mean functionality

  • in neutral pH. What is its charge?

  • [ Pause ]

  • I'm a very patient guy.

  • [ Laughter ]

  • >> Positive one.

  • >> Positive.

  • >> Positive.

  • Positives.

  • Very good.

  • OK, good. So, when-- So if this is bear lysine,

  • it will have a positive charge and acetyl group over here,

  • it is back to neutral, OK?

  • So this guy, positive charge, acetyl group, neutral.

  • So that controls whether or not that lysine, the amino acid

  • of the protein interacts with the DNA.

  • If it's positive charge,

  • it's like a homing beacon for DNA, right?

  • The DNA is negatively charged.

  • Two of these want to stick together.

  • If it's acetylated however, it's not going--

  • it's going to be neutral and the two are not going to want

  • to interact with each other.

  • Here's one that's even wilder.

  • In this case, you're taking the primary amine

  • of lysine side chain and turning it into a secondary amine

  • or tertiary amine or even a quaternary amine.

  • And when you do this,

  • you're making the lysine side chain fixed

  • as a positive charged.

  • OK, now I should say, it's not fixed permanently.

  • It used to be thought that it is but now we know

  • that actually this is a reversible modification

  • as this is acetylation, OK?

  • So, this case, it's binding to DNA, binding to DNA and then

  • when you get rid of these methyl groups, it's back to--

  • it's still bonded to DNA but then it can get acetylated

  • so it's no longer binding to DNA.

  • OK. So there is a whole series of different modifications

  • to the surface side chains, the surfaces of the histones.

  • And all of these modifications have an important consequences,

  • OK.

  • So for example, some of these modifications

  • like these larger ones down here direct the histones

  • into the proteosome which is basically the garbage disposal

  • for the cell and so those get flushed away

  • and thrown into the trash.

  • And then others like this phosphorylation

  • of a hydroxyl functionality found on the surface

  • of the DNA can regulate the structure of the histone as well

  • and perhaps interfere with this binding

  • to the negatively charged DNA.

  • OK. So, all of this stuff is tightly choreographed,

  • there are enzymes that add each one

  • of these modifications highlighted

  • in blue on the slide.

  • And those enzymes are going to control its binding affinity

  • for DNA and in turn control whether or not the DNA is hidden

  • or available for transcription.

  • And you can imagine, this is very tightly choreographed

  • by the cell.

  • If anything gets in there to mess stuff up,

  • all kinds of havoc can be wreck, right?

  • Because the cell has to control, you know, turning on, you know,

  • specific genes has specific times, right.

  • You would not want, for example, you know, a muscle cell

  • to suddenly start growing, I don't know, neurons or some--

  • you know, the genes that are required for neuron growth,

  • neurite growth or something like that.

  • That would be really bad, OK?

  • So everything is very tightly choreographed at this level.

  • All right.

  • There are of course small molecules

  • that inhibit these histone deacetylases.

  • I shouldn't say of course.

  • This is actually a discovery that was made

  • by Jack Toden [assumed spelling] who is a graduate student--

  • when I was a graduate student, the same lab where I was.

  • This discovery is made when I was a graduate student

  • in the laboratory where I was getting that Ph.D. And in short,

  • this is what the-- an acetyl lysine surface looks

  • like of the histone.

  • So here it is.

  • Taking out here is an acetyl and then here's trapoxin

  • which looks remarkably like this lysine,

  • this acetyl lysine, right?

  • This look very, very similar.

  • Maybe a slightly different number of carbons

  • but it looks very, very similar.

  • And so, this is going to be a one possible way

  • to design inhibitors of enzymes which is

  • to mimic the substrates, OK?

  • So this is the starting material for the histone deacetylases,

  • the enzyme that chops off this acetyl group.

  • And then this compound over here is going

  • to inhibit that deacetylation.

  • There's more written about this in the book.

  • But in any case, this two look very similar and so,

  • that substrate mimicry, that mimicry

  • of the starting material is a very common way

  • to inhibit enzymes.

  • It works really well.

  • We're going to see that time and again throughout this class.

  • OK, I went to change gears now.

  • I've shown you all the cool things

  • that you can build out of DNA.

  • I want to talk to you next about how

  • to actually synthesize the DNA so you can build these things.

  • OK. If you want to make happy faces or maybe you want

  • to make the first frowny [phonetic] faces

  • out of DNA hasn't been done before to my knowledge.

  • You're going to have to know how to synthesize the DNA

  • so that you can make that happen.

  • OK, so I'll first-- I'll talk very briefly about DNA synthesis

  • in the lab but first I want to talk to you about DNA synthesis

  • by the enzyme DNA polymerase.

  • OK, so DNA polymerase takes a single-stranded piece

  • of DNA called the template and adds a second strand

  • of DNA to that template.

  • OK. Now, all DNA polymerase that's found

  • on the planet have a common mechanism.

  • And they all require a starting primer strand that gets the--

  • that gives it sort of a running start.

  • Without this running start, the enzyme doesn't know

  • where to begin and this is actually a very useful property,

  • right?

  • You don't wan DNA polymerase to come along

  • and start synthesizing random, you know, bits of pieces

  • of DNA here and there.

  • And it turns out that this is one that's been exploited quite

  • a bit.

  • And I'll show you some examples of that in a moment.

  • OK, so this starter is called a primer, the starting--

  • it forms again, double-stranded DNA with the targeted template

  • and then DNA polymerase lengthens this priming strand

  • in a five prime to three prime direction.

  • In other words, it grabs onto these three prime.

  • It adds the new five prime, et cetera, OK?

  • So this direction here is also common to all forms

  • of DNA polymerase found on the planet,

  • five prime to three prime.

  • The starting materials here are nucleotide triphosphates

  • structures-- so it's a nucleotides

  • that I showed you earlier

  • but with triphosphates attached to them.

  • OK. And basically what its doing is again it's taking the green

  • primer strand and lengthening it as shown

  • by this arrow over here.

  • So this is a classic experiment that was done

  • that applied this principle to crack the genetic code.

  • The genetic code is the code by which sequences

  • of DNA spell out amino acids, OK?

  • And this was back in the 50s and early 60s.

  • There's this enormous mystery about what

  • that code actually was.

  • OK. It was like this, you know, unsolved major, major problem

  • and Marshall Nirenberg-- I think it's Rockefeller.

  • I might be wrong about that.

  • All right.

  • Marshall Nirenberg used this property

  • to crack the genetic code.

  • What he did was he synthesized templates that were long strings

  • of particular DNA sequences, OK?

  • So he made a long string of As for example and then he looked

  • at not what was synthesized by DNA polymerase

  • but downstream what was made by ribosomes

  • when you give them a long string of As.

  • And then by doing that, he could figure

  • out what the genetic code was.

  • OK. So again, DNA polymerase requires a template

  • to lengthen the existing strand.

  • Only RNA polymerase can start from scratch, OK?

  • So RNA polymerase is kind of an exception to this rule.

  • DNA polymerase requires a priming strand reverse

  • transcriptase which takes RNA and synthesizes DNA,

  • we saw that earlier also requires a primer

  • and deoxynucleotide triphosphates.

  • RNA polymerase is kind of a special case.

  • By the way, any questions?

  • You guys feel free to interrupt

  • if there's anything that comes up, OK?

  • Anything that's unclear, you want to know more information

  • about it, don't hesitate to stop me, OK?

  • Yeah.

  • >> Does the primer get replicated as well or is it--

  • >> The primer gets extended but it doesn't get replicated

  • in the load that I'm showing you.

  • When we talk about PCR,

  • we'll show that it actually can get replicated, so, OK.

  • That was a good question.

  • Other questions?

  • Yeah.

  • >> If you wanted to hybridize [phonetic]

  • like a piece of a primer--

  • >> Yeah.

  • >> Lamination [phonetic] in it--

  • >> Yeah.

  • >> How many like base pairs do you usually need to--

  • >> OK, these are great questions.

  • OK. Awesome.

  • I'm glad you're asking.

  • I forgot to ask your names.

  • What was your name?

  • >> Paul [assumed spelling].

  • >> Paul and?

  • >> Anthony [assumed spelling].

  • >> Anthony.

  • OK. So Anthony's question is how many base pairs

  • of DNA should you have to get the--

  • to use as the primer to get DNA polymerase going.

  • OK. And it kind of depends, OK?

  • So, you want DNA polymerase to pick

  • up a specific gene and/or a specific sequence of DNA

  • within a complex mixture.

  • And so, if you want to pick up a specific gene

  • in the human genome, you need a primer that's

  • at least 18 base pairs in length.

  • OK, that's kind of a magic number.

  • OK? So, 18 base pairs means that you're uniquely encoding one

  • and only one gene in the human genome, OK?

  • Thanks for asking.

  • It's a good question.

  • On the other hand, if you want to do this at, you know,

  • a lower temperature, you can use the Wallace rule and get away

  • with maybe a shorter sequence.

  • OK, maybe you don't need such specificity.

  • Maybe your mixed-- your starting population is less complex.

  • OK? Thanks for asking Anthony and Paul.

  • OK. Let's move on.

  • OK, so here's a close up view of what I've been telling you

  • about in hand waving examples.

  • We're now zooming down to the level of atoms and bonds that,

  • of course, is what really thrills me.

  • So, here is the primer and DNA polymerase, not shown inside,

  • of course, this primer is forming a double-stranded DNA

  • to the template strand and also not shown is

  • that the template strand must have adenine.

  • Over here, the hybridized to the sliming [phonetic], OK?

  • In any case, the starting material used here is a

  • deoxynucleoside triphosphate.

  • Note that deoxy at the two prime hydroxyl.

  • OK. And here is the two prime-- or sorry.

  • Here is the triphosphate functionality.

  • The living group in this reaction is going

  • to be diphosphate which is use very commonly

  • in biology as a living group.

  • This is nature's tosylate or mesylate that you learned

  • about back in Chem-51C.

  • This thing works really well as a living group and it's one

  • of the reasons why we're going

  • to see it quite a bit as a living group.

  • The-- All DNA primaries, all enzymes that use diphosphate

  • as a living group absolutely require a dication to bind

  • to this diphosphate and their requirement is

  • for magnesium in DNA polymerase.

  • OK? So, actually a very common problem that I see

  • in my own libratory when a newbie shows up in the lab

  • and they have trouble with their DNA polymerase,

  • nine times out of 10 it's due the low concentrations

  • of magnesium.

  • OK? And there's a lot of ways

  • to get low concentrations of magnesium.

  • So, a little tip.

  • OK, so here's magnesium.

  • What is it doing?

  • Magnesium is a Lewis acid that's chelating to this diphosphate

  • and stabilizing its negative charge.

  • Doing this makes it a better living group, right?

  • This means that if it goes out into the solution,

  • it doesn't require some massive rearrangement of water,

  • it's already been stabilized.

  • It's at lower potential energy then it would otherwise be.

  • OK, so here's the role of DNA polymerase.

  • And I'll show you instructionally what it looks

  • like it a moment.

  • DNA polymerase brings together the three prime hydroxyl

  • of the priming-- the primer together

  • with this incoming nucleotide triphosphate and then sets

  • up a nucleophilic attack on the phosphorus

  • of the nucleotide triphosphate.

  • Note too that there is a second magnesium ion

  • in the active site.

  • This second magnesium ion does two things.

  • Number one, it helps to stabilize the alkoxide formed

  • when this three prime hydroxyl is deprotonated, OK?

  • So notice that it's forming an ion pair relationship

  • with this alkoxide.

  • Number two, it actually increases the nucleophilicity

  • of the lone pair that's going

  • to this nucleophilic attack over here.

  • OK, so what magnesium is doing here is it's making available--

  • better available this long pair for an attack

  • by helping promote the deprotonation of that hydroxide.

  • If the hydroxide is deprotonated,

  • there's more long pair that's available for the attack, right?

  • Make sense?

  • OK. This forms an intermediate.

  • There's a clasp [phonetic] of intermediate and not only gets

  • to what we saw on Tuesday when we looked at the hydrolysis

  • of DNA, exacts in mechanism.

  • That intermediate is depicted.

  • I'm showing arrows over here.

  • But again, we've looked at that intermediate before

  • so I fell comfortable living it off at this slide.

  • OK, any questions about this mechanism?

  • Yeah.

  • >> Where does the magnesium come from?

  • >> Great question, what is your name?

  • >> Nick [assumed spelling].

  • >> Nick, OK.

  • Nick's question is where does the magnesium come from?

  • Magnesium comes from the food you eat.

  • It comes from, you know, you added to the test tube.

  • So, typically we'll add magnesium chloride

  • or magnesium sulfate directly to the eppendorf tube.

  • Then you'll test tube that we use for these reactions, OK?

  • But in humans reading, you know all kinds

  • of food as magnesium in it.

  • OK. So, but it's absolutely essential, OK?

  • So without the magnesium this reaction does not go, OK.

  • It makes sense because I'm showing you what a key role

  • it place.

  • OK, let's look structurally at what this actually looks like?

  • This is an enzyme that again has a number

  • of different orthologs or homologs.

  • These are enzymes that do more or less the same thing.

  • Reverse transcriptase synthesizes DNA

  • from an RNA template.

  • The enzyme Taq is a DNA polymerase that's use quite a

  • bit in research for a tool called PCR,

  • which I'll talk about in a moment.

  • But all of these enzymes have a right handed structure, OK.

  • And here's what the structure looks like.

  • OK, so here's a right hand over here.

  • Here's my right hand.

  • And it's grabbing on to the DNA, OK?

  • So the DNA is in red and orange over here

  • and here is the enzyme grabbing onto this DNA.

  • Now what happens is during the synthesis, the DNA treads itself

  • through the crack form by my thumb and palm, OK?

  • And as at a certain-- when the crack nucleotide triphosphate

  • binds to the priming strand, the newly synthesize strand,

  • the enzyme can then close.

  • When it closes, the palm and the thumb get closer to each other,

  • palm, thumb and fingers.

  • OK. So it closes a little bit like this.

  • Each time that closes, that brings the magnesium's

  • up to the triphosphate setting up formation

  • of the covalent bond that I showed

  • on the previous slide, OK?

  • So each time the hand closes,

  • that's one nucleotide that's been out.

  • OK, so let's do this, right.

  • We have this one bond.

  • OK, now let's do a couple more, bond, bond, bond, OK?

  • Now here's the deal.

  • This enzyme is really cranking.

  • It's actually going to do a thousand up to a thousand

  • of this per second for some of these enzymes, OK.

  • So that's like, you know, too fast for it to see really.

  • OK. So this enzyme can really turn over very quickly

  • and actually this is actually something

  • that my libratory is directly observed.

  • We actually have watched one of this enzyme cranking over

  • and we've watch differences

  • as we add difference substrates to the enzyme.

  • It's really absolutely fascinating series

  • of experiments.

  • OK, so, all of these enzymes use a common mechanism.

  • Again, the enzyme doesn't close

  • until it gets the crack nucleotide triphosphate

  • that binds to it at that point it closes.

  • So in fact actually the rate determining step

  • for this enzyme is actually the rival

  • of the cracked nucleotide triphosphate.

  • A is to Ts or DATP to Ts, DCTP to Gs, et cetera.

  • OK, I'm going to tell you a little bit more about why I'm

  • such in love with this enzyme.

  • This is a 3D machine.

  • I like fast cars and I like fast enzymes.

  • This one is really amazing.

  • So check this out.

  • Imagine that double-stranded DNA was about a meter or so

  • in diameter, OK, running a long the length of this room, OK?

  • So we've got some DNA running through the room, all right.

  • If that was true, DNA polymerase would be about the size

  • of the FedEx delivery track.

  • OK. Including the polymerase

  • so there's some other replication machinery

  • that I'm living off for now that's involve as well.

  • But it would be the roughly the size

  • of FedEx delivery track pulling up right here, OK?

  • But here's the thing.

  • This delivery track would be racing along

  • at about 375 miles per hour.

  • And that's how fast DNA polymerase is going

  • in scale to the DNA.

  • Furthermore, it's making about a thousand covalent bonds per

  • second which is insanely, insanely fast.

  • And in addition, I haven't talk about this yet

  • but there are other subunits of DNA polymerase

  • that are providing an error checking

  • and a correction function such that the enzyme is making

  • to scale one error every a hundred or six miles or so, hey,

  • which is extraordinary, OK, 375 miles and hour

  • and one error every 106 miles.

  • OK, this is truly remarkable stuff.

  • You could read more in this reference down here.

  • OK, now, because this enzyme is so efficient

  • and so superbly specific

  • at getting the right Watson-Crick base pairing,

  • this has been used very commonly in lots and lots

  • of laboratories, chem-bio labs, molecular bio lab,

  • biochemistry labs, forensic laboratories,

  • all kinds of labs used DNA polymerase and they often use it

  • to amplify up copies of the DNA using a technique called the

  • polymerase chain reaction invented

  • by Kary Mullis [assumed spelling] amongst others.

  • The way this works is you start with some target sequence

  • of DNA shown here in purple.

  • Again, we'll call the target B template, OK?

  • So that's the template DNA that you're going to amplify.

  • Now, there are going to be three steps to this PCR reaction.

  • In step one, we hit the DNA

  • up to high temperature say 95 degrees.

  • And as we've discussed earlier today, DNA when it's heated

  • up the high temperature goes

  • from double-stranded to single-stranded.

  • It falls apart.

  • In step two, the solution is cool down

  • and that allows the primers shown here in green and blue

  • to a [inaudible], in other words hybridized

  • to the single-stranded DNA.

  • Note that these two purple strands don't find each other.

  • The concentration of template if very, very low.

  • In fact, you can get down to just a few copies of DNA.

  • So they never find each other.

  • They are like, you know, lost from each other

  • after the heating step.

  • But you have a high concentration of this green

  • and blue primers that can grab on to the crack sequence of DNA.

  • That motion targets DNA polymerase, drags DNA polymerase

  • to synthesize a specific stretch of DNA.

  • And then that's done in the third step

  • when the primers are extended using again DNA polymerase,

  • DNTPs, magnesium chloride and a temperature of 72 degrees.

  • To make this work, we use a special DNA polymerase

  • that likes to run at 72 degrees.

  • It's a type of polymerase called taq which is found in hot fence,

  • hot springs and it's in an organism that's found

  • in this hot springs that has evolved

  • to operate at this temperature.

  • And so, at 72 degrees, the enzyme starts cranking.

  • At the lower temperature, it's not working.

  • At the higher temperature, it stops working.

  • But in 72 degrees, it's loving [phonetic] life.

  • Again this is Celsius.

  • This is pretty warm, and its start synthesizing this black

  • strand of DNA.

  • If you do this process a whole bunch of times, each cycle,

  • you get a doubling of the amount of DNA

  • and so you do this 30 times.

  • You get a huge amplification for some target template of DNA.

  • Some target sequence of DNA.

  • OK. Makes sense?

  • Any questions about this?

  • I'm hoping I'm not telling you anything you don't already know.

  • PCR is now taught like high school and stuff like that now.

  • So, OK, summary.

  • Right hand role, we looked at species

  • about the magnesium two plus, stabilizing the nucleophile.

  • We've looked at this already.

  • Why don't we move on?

  • OK, so DNA polymerase is also a terrific target for inhibitors

  • and reverse transcriptase inhibitors have been very,

  • very important compounds for stopping synthesis of DNA.

  • There're many reasons why you'd want

  • to stop the synthesis of DNA.

  • To treat, for example cancer,

  • where cells are dividing uncontrollably

  • if you can shutdown the replication of DNA,

  • you have an effective way of stopping cancer.

  • And in fact actually, childhood leukemia's were stopped

  • in their tracks back in the late 70s

  • through the wonderful research of one

  • of my scientific heroes the great Gertrude Elion shown here.

  • Gertrude Elion was born in 1918.

  • Her parents wanted her to become a nurse, OK?

  • So they send her to college and they said,

  • "Go and become a nurse."

  • She actually wanted to become a chemist,

  • and when World War II broke out,

  • she was given her opportunity, OK?

  • So during World War II, the man were sent to the front to fight

  • and there are a lot of opportunities

  • that were available to women that weren't available before

  • that and she's one of those people

  • who took that opportunity.

  • She joined Burroughs-Welcome where she worked

  • with George Hitchings for her entire career.

  • She's one of this people

  • who spend her entire career at a single company.

  • And together with George Hitchings,

  • she discovered this class of compounds

  • that inhibits DNA polymerase and ended

  • up after having a major impact on childhood leukemia.

  • She is, you know, she is a true superstar of science.

  • OK, one last thought.

  • Gertrude Elion never received her Ph.D. She went

  • on to receive a Nobel Prize in the 80s for this work.

  • She did it through sheer force of will,

  • through her determination to contribute something.

  • And I highly, highly recommend in interview of her

  • and I'll leave and put it up on the board over here.

  • There's-- If you want to learn more about her,

  • this is terrific interview of her that's

  • in the documentary that I recommend.

  • OK, so the documentary is called Isaac--

  • can you see this OK, normally?

  • OK. "Isaac-Newton and Me"

  • and the director is the great Michael Apted.

  • I could actually have a whole class just

  • on Michael Apted's documentaries.

  • But in any case, she's interviewed in this documentary

  • and she talks about the incredible pride

  • and just the joy that she felt

  • when she would visit children's hospitals

  • and she would see kids being treated for the first time

  • with her compounds, and how transformative that was

  • in the life of these kids.

  • These are kids that were, you know, slated to diet

  • at very young age, just like the ages of 10 and 12

  • and that were suddenly getting cured by these compounds.

  • OK, let's take a closer look

  • and understand how these compounds work.

  • OK, so she invented a series of inhibitors of DNA polymerase

  • that look like this, OK?

  • So this is one called AZT.

  • It's also used very as a--

  • anti-HIV compound because it inhibits reverse transcriptase.

  • It looks kind of like a DNA base.

  • It has a deoxy at the 2' prime hydroxyl but in place

  • of the 3' prime hydroxyl, there is an azide, OK.

  • So what happens is this gets taken up and phosphorylated

  • to give a triphosphate and then DNA polymerase attempts

  • to use it as a substrate.

  • But what ends up happening is instead of a 3' prime hydroxyl,

  • there's an azide here and the azide caps the synthesis

  • of the new strand of DNA preventing it

  • from being lengthened, OK.

  • So similarly, this is a ddC another compound that's used

  • in the treatment of HIV and also leukemia and instead

  • of a 3' prime hydroxyl, it has a hydrogen there.

  • And so again, it gets used by DNA polymerase

  • and the polymerase can no longer lengthen the nascent strand

  • of DNA, OK, so both of these shutdown DNA syntheses.

  • I'm simplifying things a little bit.

  • There is another class compound, non-nucleoside analogs.

  • These are nucleoside analogs that are also used against HIV.

  • But that kind of gives you a taste

  • for what Gertrude Elion did.

  • She had a major, major impact in the fight against viruses

  • and in the fight against leukemia.

  • And I actually had the pleasure of meeting her once in life.

  • The day I was getting my Ph.D. at my graduation,

  • she was receiving an honorary Ph.D. from Harvard

  • and I just shook her hand.

  • That's about it.

  • I didn't have any profound conversation with her

  • which is to my regret.

  • But a truly remarkable woman, a true superstar of science.

  • I can not say enough about her.

  • I can have a whole lecture about her.

  • Why don't we move on?

  • OK, so, that's DNA synthesis in the cell--

  • oh, question over here.

  • Yeah.

  • >> So these nucleotides that she synthesized,

  • how did they deliver [inaudible] cancers cell--

  • >> Oh, this is such a good question

  • and I'm so glad you asked.

  • What is your name?

  • >> Bobbin [assumed spelling].

  • >> Bobbin?

  • >> Yes.

  • >> OK. So, Bobbin's question is how do you get the compounds

  • to the cancer cell.

  • What we're going to see time and again is

  • that cancer cells are actively eating up every little bit

  • of nutrient that they could find.

  • They will just be devouring stuff that's around them.

  • And so for this reason, they

  • and other dividing cells will more preferentially take

  • up drugs like these that are fed to the patient

  • that are injected into the patient.

  • OK. So, the problem of course is that there are other cells

  • in the body that are also dividing

  • and that will unfortunately take up this chemotherapeutics

  • and also end up dying because their DNA synthesis will

  • be impaired.

  • So, OK. Great question.

  • Thanks for asking.

  • OK. So, I want to-- I don't have very much to say

  • about chemical DNA synthesis.

  • I think it's an obviously amazing topic.

  • This is one of those areas of organic chemistry

  • that is a true triumph.

  • OK, so, we're going from strength to strength today.

  • In DNA synthesis in the laboratory has been so optimized

  • that we're at the point where we get 99.9 percent yields

  • for reactions, OK?

  • This is really the ultimate goal in the quest

  • to do organic synthesis.

  • And it was set in line by the great Gobind Khorana

  • who won a Nobel Prize but also Bob Letsinger

  • and Marvin Carruthers.

  • These guys deserve a Nobel Prize

  • because this really did kick off a revolution in biotechnology

  • that was made possible by the synthesis of DNA to make primers

  • which in turn allows PCR, which in turn allows smiley faces

  • out of DNA and all those other great discoveries

  • that we've been talking about.

  • OK. Now, I just want you to scan this topic in the book.

  • Don't get too worked up about it.

  • In the end, we have these machines that looked like this

  • where you have a bunch of bottles down here

  • that inject-- that you can use.

  • You could program a computer up here to open--

  • to inject reagents directly into a flow cells

  • that have the DNA sequence that you're trying to synthesize.

  • So you do this using solid support base synthesis

  • where did-- the nascent strand of DNA is be--

  • is attached to some sort or bead and you basically flow

  • in the reagents one after another

  • and couple the correct nucleotide directly

  • onto the DNA sequence.

  • It's a little more complicated than that

  • but here's what you need to know.

  • If you want to synthesize any sequence

  • of DNA that's 150 base pairs, 150 bases or less,

  • you get on the web and you call up someone probably in Texas

  • or somewhere like that.

  • And there will be a whole warehouse of machines like this

  • and you enter into some form on this--

  • on their website exactly the sequence of DNA that you want.

  • And that sequence will get ported to a machine that looks

  • like this and there'll be this warehouse just filled

  • with these machines.

  • And then there's a bunch of technicians on roller blades

  • that are going to be running around

  • and keeping the machines fed with reagents.

  • You won't see any of this because at the end,

  • you'll get a FedEx package

  • with your DNA sequence probably in a couple of days.

  • Some of these I think are even overnight, right?

  • So overnight, you're going to get your sequence

  • of DNA perfectly cleaned up, purified, delivered to you

  • at 95 percent or 99 percent purity depending

  • on how much you decide to pay for it.

  • And, you know, it will be perfect every time.

  • You won't even have to think about the chemistry.

  • I love that.

  • That's the goal of organic synthesis.

  • The goal of organic synthesis to make this totally turn key

  • so that we can then use this thing

  • to answer biological problems

  • which is what I want to you about next.

  • OK, here's one example of using this, an amazing ability

  • to do DNA synthesis to address biological problems.

  • You can print sequences of DNA on microscope slides.

  • OK. So here's a machine that's nothing more fancy than I think

  • in this case it's an ink jet printing device.

  • OK. And it's going to be printing

  • out little oligonucleotide one after another

  • on this identical microscope slides.

  • So each square down here is a microscope slide, OK.

  • And across the surface of these microscope slides,

  • we're going to have a bunch of different sequences of DNA

  • that have been printed down onto specific spots, OK.

  • So what this is going to do is this is going

  • to give us an array of different sequences

  • of DNA each one capable

  • of hybridizing forming Watson-Crick base pairing

  • to a different other sequence of DNA or RNA.

  • OK. This is a technique called the DNA microarray.

  • OK. So, here's the way this works.

  • Let me just show you what it looks like, OK?

  • So now I'm zooming in on the microscope slide.

  • OK. So here's what it looks like.

  • Each spot over here is a different sequence of DNA

  • as you set up the hybridization such that you end

  • up with a fluorescently labeled sequence, OK.

  • So each-- again, each spot has a different sequence of DNA.

  • If there is a complimentary sequence in your sample,

  • then you will see fluorescence, OK?

  • So, in green, that tells you

  • that your sample has this particular sequence.

  • OK. And it's known exactly what the sequence is in green that's

  • down there, OK, because you've synthesized the compliment

  • to put it down right there.

  • OK, so-- I don't know.

  • Let's just say this is the gene that confers resistance

  • to beta-lactam antibiotics.

  • Yes, you have that gene because you're seeing green spots

  • right here.

  • In practice, this can be made a lot more complicated, OK?

  • Let's imagine now that we have two samples, one that we label

  • with red sequences and one that we label with green sequences.

  • OK. This allows us to compare all

  • of the different DNA sequences present

  • or RNA sequences present comparing them

  • against red versus green, OK?

  • And let's take a closer look.

  • OK, so each oligonucleotide hybridizes

  • to a different mRNA transcript where the one sample is labeled

  • in green and the second sample is labeled in red.

  • OK, here is the way this works.

  • On sample one, you use reverse transcriptase

  • to convert all these RNA into DNA

  • and you add the green fluorophore.

  • In sample two, you use reverse transcriptase and label all

  • of those with a red fluorophore, OK?

  • So sample one is green, sample two is red and then you add both

  • of those to the DNA microarray.

  • The ones that are in green are telling you, "Oh, that gene is

  • up regulated in those types of cells."

  • The ones that are on red is telling you it's up regulated

  • in the other types of cells and where red

  • and green overlap you see yellow, OK?

  • So, this allows you to compare experiment versus control

  • and the possibilities are endless.

  • For example, you can look at cancer versus non-cancer cells,

  • virus infected versus normal,

  • G1 phase of the cell cycle versus S-phase.

  • I'll explain these phases in a moment, young cells versus old,

  • drug treated versus no drug, et cetera.

  • And so in the end, you get these massive arrays

  • where you get a huge amount of data.

  • Each of these red spots

  • that could tell you, "Oh, that's a gene.

  • That's up regulated," and say,

  • "Cancer cells are virus infected cells."

  • And then the yellow ones,

  • those are ones you don't have to worry about.

  • The green ones though, right, because that's the same

  • in both types of cells.

  • The green ones however are ones that are up regulated in the--

  • let's say, the non-virus cells, OK down regulated

  • in the case of the cancer cells.

  • So you can see in one very simple experiment relatively

  • simple, little complicated, you get out enormous amount

  • of information about the transcription activity

  • of an entire cell.

  • Again, this is all set in place by DNA synthesis,

  • chemical DNA synthesis.

  • OK, so here's an example of this, a specific example.

  • The example I'm going to use is the small molecule FK-506 also

  • called tacrolimus.

  • This is a-- immunosuppressant that was discovered

  • by the chemical--

  • the pharmaceutical company Fujisawa hence the name FK has

  • an immunosuppressants.

  • It is given to patients who had an organ transplanted, OK?

  • So, this was kicked off sort of a revolution in the area

  • of liver transplantation for example

  • when it first became available in the 80s--

  • late 80s and early 90s.

  • OK, so if you give this drug to your patients

  • who have just had the liver transplant,

  • they will not reject the new transplant of liver, OK,

  • because their immune system is suppressed.

  • You can kind of see the problem with that approach.

  • All right.

  • I mean the immune system is suppressed that means

  • that if they get it cold,

  • they're going to be on big trouble.

  • Setting that aside, OK, there's other things that you can do.

  • Let's try to look at what pathways are changed

  • when we feed cells this compound.

  • In doing so, we could start to identify the pathways associated

  • with the immuno response.

  • And so here's a classic experiment done

  • by the company called Rosetta Inpharmatics now Merck

  • pharmaceuticals and in this experiment they have two kinds

  • of cells.

  • One kind of cells are green.

  • They were treated with the green fluorophore

  • so the transcripts are treated with the green fluorophore

  • and in the red-- so these have no drug added to them

  • and on red, those transcripts were treated

  • with the compound I showed

  • on the previous slide called FK-506.

  • In yellow, there's a whole bunch of different compounds.

  • Check this out.

  • In green over here, this is a-- this is a protein--

  • this is gene that encodes a protein that must be turned off

  • when the compound is added, OK?

  • Notice that it forms a bright green spot.

  • Over here, here's one that gets turned on.

  • It forms a red spot.

  • The yellow ones, the orange ones,

  • let's not worry about that so much.

  • OK. But you can imagine doing all kinds

  • of analysis of the stuff.

  • The bioinformatics side of these things, the computation

  • to analyze the stuff, fascinating, OK,

  • and a really exciting area of computer science research.

  • OK, does this make sense?

  • OK, so one experiment, you get the whole pathway that's being

  • targeted by this drug.

  • OK and I'm not telling you very about the pathway now.

  • Maybe we'll talk about it later.

  • Yeah. Question at the back.

  • [ Inaudible Question ]

  • OK, yeah. So, in practice you have a laser that scans

  • across the surface of this microarray

  • and then you have a CCD camera

  • that captures the intensity of each spot.

  • OK, I'm hoping this is blowing your mind

  • because it certainly blows mind.

  • OK, so now, I wanted to change gears a little bit and talk

  • to you about how to analyze DNA-- sequences of DNA.

  • Before I do that, I'm looking at the time here.

  • I'm going to run all the way up to the 10 before the hour.

  • But I just want to tell you I have some good news for you,

  • a little weekend treat.

  • Thanks for bearing with me of the quiz.

  • The midterm will only cover through chapter three.

  • There'll be no chapter four in the midterm, OK?

  • So when we come back on Tuesday, we'll be talking more about--

  • don't get ready to go.

  • I'm not going to stop now but I just want to let you know,

  • we'll be talking more about DNA and then we'll go into RNA.

  • But the midterm next Thursday, a week from today, will only cover

  • through chapter three, OK?

  • All right, let's get back to our discussion.

  • I want to switch gears very slightly.

  • Earlier I alluded to agarose which is used

  • in those Petri dish plates that I showed you earlier.

  • Agarose has the structure down here.

  • It's isolated from seaweed, and it is a polymer

  • that is tightly crosslet [phonetic] to form a notch, OK.

  • You can actually form little bricks form of the stuff

  • and then you can apply the DNA to one end of the agarose gel.

  • So here's your little brick called an agarose gel

  • and you add your DNA to one end of this gel

  • and then you use electrophoresis to push the DNA

  • through the polymer, OK.

  • We recall that DNA as negatively charged so it will be attracted

  • to the positively charged terminal

  • to possibly charge electrode, the cathode at one end of the--

  • of your electrophoresis apparatus.

  • OK, so again, you have this wires coming out.

  • They're going back to some source of electricity,

  • some power supply and that's pushing the DNA through the gel.

  • The DNA will get separated out then

  • on the basis of its size, OK?

  • So, bigger DNA is going to get caught

  • up in this complicated network

  • over here whereas the little pieces of DNA are going

  • to find it easier to flow through these little pores.

  • OK. So in practice, what this looks like is this.

  • I think I've already shown you one of these agarose gels.

  • This is the top of the gel where the big pieces are.

  • Here's the bottom of the gel where the little pieces are.

  • Again, the big pieces have not migrated as far

  • because they got stuck in the inner stitches of the DNA.

  • Frequently, we add what's called a DNA ladder to one lane

  • of the gel and that provides basically a ruler

  • that tells us what sizes of DNA we're looking at.

  • How is this visualized?

  • Why are these DNA pieces in this color over here lit up?

  • What do we add to the gel?

  • >> Ethidium bromide

  • >> Yes, ethidium bromide.

  • Thank you.

  • Yeah. We've added a diethidium bromide that concentrates

  • in the DNA and is fluorescent.

  • This works really well.

  • This tells us a lot about DNA length

  • and a little bit about its structure.

  • You can look at for example supercoiling using

  • this technique.

  • OK, yeah. Here's the ethidium bromide.

  • Here it is intercalated into DNA and this

  • at practice is what it really looks like.

  • It has this bright purple color.

  • In addition to separating on gels, we very commonly separate

  • out DNA using capillary electrophoresis.

  • And this is a much more effective way of separating

  • out DNA that it differs by a single base in size.

  • So you can take a large number of sequences of DNA

  • and then separate them out such that if you have, you know,

  • 100 meers [phonetic], 99 meers, 98 meers, 97 meers, 96,

  • et cetera all neatly separated

  • out using this technique called capillary electrophoresis.

  • It's the same electrophoresis technique

  • with a different mobility layer-- different mobility phase

  • but it's basically using electrophoresis again

  • but in a capillary.

  • OK. Here's why this is important.

  • You can use this gel based separation

  • as a way of sequencing DNA.

  • If you had some way of breaking it up into little pieces

  • where each piece differs by one base pair.

  • In practice, what we do is--

  • and I will only tell you about this one over here.

  • OK, this is the old fashion way to do this in the--

  • using this actually we use a acrylamide gels.

  • This is back when I was a post doc.

  • We no longer do this.

  • No one in lab does it this way.

  • We all do it this way.

  • Here's the way this works.

  • OK, what we do is we add a very low concentration

  • of dideoxy terminators.

  • This is exactly what a-- was invented by Gertrude Elion

  • on the Gertrude Elion slide that I showed you earlier.

  • This is missing, the 3' prime hydroxyl.

  • It has a hydrogen in place over here

  • and these 3' prime terminators missing this hydroxyl shutdown

  • DNA polymerase synthesis.

  • When they do, they also carry along a fluorophore

  • that has a specific wavelength associated with it, OK?

  • So you add a one percent concentration

  • of the dideoxy C inhibitor that has a green flurophore.

  • You add one percent that has G and has a red fluorphore.

  • You add one percent that has T and has a purple fluorophore

  • and you see where this is going.

  • So you have four dideoxy terminators.

  • Each one with its own fluorophore

  • and you set this fluorophores up so there's no overlap

  • of the wavelengths or a little overlap as you can get.

  • In the end, you separate out the sequences using capillary

  • electrophoresis and what you can do is actually read

  • out based upon the read, green or blue status of each

  • of those dice that you have a sequence

  • that says GAT, CTT, GTT, et cetera.

  • In our practice, we actually, you know, simply take the data,

  • feed it into a program and the computer gives us the sequence

  • at the end.

  • OK, so this stuff is massively automated.

  • OK, in fact, we're at the point where our lab simply sends

  • out the sequences and there's another lab.

  • It used to be on campus but now it's actually-- where is it?

  • >> I think its San Diego.

  • >> It's here?

  • >> San Diego.

  • >> San Diego.

  • OK. Yeah. So it's in San Diego that does all the sequencing

  • for us and it cost only a few dollars per sequence.

  • OK. And that's-- that price has dropped enormously.

  • OK, any questions about anything we've seen today?

  • Yeah, Anthony?

  • >> Well, I don't know if you said but I'm a little curious

  • about donor [inaudible] stuff there where you--

  • >> Oh, yeah, yeah, yeah.

  • Let's put that off.

  • >> OK.

  • >> OK. Yeah.

  • Fluorescents [phonetic] technology is interesting.

  • All right, last thing, you can use this DNA technology

  • for all kinds of things.

  • It's used very extensively.

  • You can even use it to program changes in organisms

  • like drosophila and you can use this to test hypothesis.

  • Little hard to tell but you see the extra eyeballs

  • that are growing out of this organism.

  • Eyeball at the stuck, eyeball over here, yeah.

  • So you could actually program-- you can insert sequences of DNA

  • into these organisms and [inaudible] them to get turned

  • on at specific times and use this some test whether

  • or not we understand what specific genes do.

  • And this is actually as fascinating story

  • that actually deals with a heat shock protein.

  • OK, there's a couple of ways of modifying DNA in organisms.

  • One way is to do it randomly using compounds

  • like this one whose mechanism we'll talk more

  • about on Tuesday.

  • So let's stop here.

  • When we come back next time,

  • we'll be talking more about this. ------------------------------f77944255c8c--

[ Silence ]

字幕與單字

單字即點即查 點擊單字可以查詢單字解釋

B2 中高級

化學生物學概論128.講座06.DNA與小分子的反應性。 (Introduction to Chemical Biology 128. Lecture 06. DNA Reactivity with Small Molecules.)

  • 187 18
    Scott 發佈於 2021 年 01 月 14 日
影片單字