Mike Bassik: Multiplexing with CRISPR Screens

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments
Captions
(inspirational music) - I'm very excited to discuss my favorite topic in the whole world, which is high through put CRISPR Screens, and so I would encourage you guys to interrupt as often as you'd like to ask questions. I have a lot of material, but I'm happy to divert the flow to talk more about what you guys would be interested in. So, Luke set up a lot of important topics that, in general, what we're going to try to do now is to take some of those technologies that have been engineered to work fantastically well in individual instances, and now scale them up to do genome-scale screens, and so, first, I'm just going to give some general motivation for why one might want to do a genetic screen. Probably, many of you have either done one or thought about doing one, but there's lots of reasons you might consider doing this type of thing. I'll describe, partly, by way of historical context, but partly by way of describing how exactly it is that the Cas9 system has been so powerful at the functional genomics scale. Comparison to some RNAi technologies that were used early on in functional genomics, and in my own lab, we've done some direct comparison between RNAi and CRISPR that actually is somewhat eliminating in terms of understanding the difference between a knock down phenotype and a knock out phenotype, and what you can do with that information. I'll describe a number of different screens that have been done by a number of different labs. I'm gonna focus some on the screens that we've done, just because I can talk in more detail about them, but I'll describe some particular applications in terms of identifying the targets of different drugs and phenotypes for essential genes, which are very easy phenotypes to screen for. I'm gonna describe a little bit a continuation of what Luke set up, and using CRISPRi and a to do genome-scale screens and really elegant work that's been led by Luke and Max and others in the Weissman lab. And throughout, I'm gonna sort of try to give, actually, some tips and things that are just sort of practical, things that we've learned in the process of doing a lot of screens, as well as different design considerations that you might consider if you're thinking about setting up one of these screens, and then, the last bit, I'm gonna describe two relatively new applications of CRISPR screening that we've been exploring in my lab, but have also been used in a number of other labs. In particular, genetic interaction screens, where you're knocking out pairs of genes to look at synergies between different pairs of genes, and a scaling of the technology that Luke described at the very end of his talk, which is what can you do if you can mutagenize a large stretch of DNA. And then, at the end, I'll just describe some useful things for what do you do when you have a screen, the results from a screen. How can you actually use that information? So, just one really basic slide, I think, to kind of understand where screens have come from. Probably, this may be too basic for some of you, but before the sequencing of the genome, in the olden days, and actually still today, to some extent, the way you would do screens would be using forward genetics, and here, you take an organism. You mutagenize it. You select for some kind of a phenotype that you're interested in, say resistance to a drug, and then you find the gene that's responsible for that phenotype, usually by complementation or something like that. And this was, for a very long time, the way that people did genetics in an unbiased fashion. Of course, you can also study heritable traits and then find genes that way, but this is the way that people have been doing genetics for a long time. With the sequencing of the genome, a number of different technologies enabled reverse genetics, which is now that you know the sequence of those genes, you can either use, for example, an antisense oligonucleotide or RNAi, which was the technology of choice for about 10 or 15 years, where you can specifically code a double stranded RNA, a short, I should say, 21 nucleotide doubled stranded RNA that effectively incorporates into the cell's endogenous RNAi processing machinery, which, in the end, cases a degradation of a target RNA molecule, and so this, I should say, has been used for quite a long time. Still actually has some uses, as I'll describe to you, but the key idea here is that you're introducing whether by a chemically synthesized double stranded short RNA molecule, or a lenti-virally encoded hairpin that, in the end, gives you that same product. A double stranded RNA that causes degradation of its target. So, now that we have the ability to knock down specific genes and the sequence of a genome, you can do genome-scale reverse genetics, and that allows you to potentially find all the genes involved under your favorite process, so instead of just mutagenizing a plate of worms or bacteria and hoping that you find an interesting mutant that gives you the phenotype you'd like, you can really systematically scan through every possible gene, knock it out or knock it down, and ask what is the phenotype of that gene. And so, I say that's easier to saturate because, if you wanted to actually saturate the entire genome doing a chemical mutagenesis screen, you would need an enormous vat of worms or bacteria, and you know, thousands of people years to actually potentially get all of the available mutants that might give you all the phenotypes, but if you know the sequence, you can actually go in and just surgically knock them down or out one at a time. So, this is a much more efficient way of associating the gene with the phenotype, and now that we have not just the sequence of the genes but actually all these regulatory elements in the genome that Luke was describing, we can actually approach a systematic annotation of the genome, both for coding and non-coding areas. So, there's a lot of information on this slide, but I wanna just describe the two general strategies that have been used for doing genome-scale screens, so on the right, you'll see arrayed library screening. This entire slide is geared toward CRISPR screening, but a very similar overall workflow was used for RNAi for a long time, and the key differences are that, in this case, you have one well, one gene knock down, and in this case, you have a pooled library, where you have lenti-viral, usually, plasmids encoding either an shRNA or, later, a CRISPR guide RNA, and these are infected at low multiplicity, so that each cell gets one virus in a pool, and then you can subject this pooled population to selection, and so, in the end, in both cases, you end up with some selection strategy, and then, you know, informatics to separate the hits from the non-hits in these types of screens. One of the nice things about an arrayed screen is that you can potentially get much more high-content imaging, so you can put these plates, after you knock down one gene in a well into a microscope, take an image of what the organelles look like, for example, and really get an enormous amount of information for what's actually going on in that one well. You can't get, typically, that level of information in a pooled screen, but these have a number of advantages. They're much, much easier and cheaper to conduct, so I'll just tell you right now, everything I'm gonna talk about from here on out is in pooled format, because we love the pooled format, but I wanted to set this up because a lot of screens were done this way, and it's still actually possible to buy, even now, CRISPR guide arrayed in libraries, because there are some applications where you really want an arrayed one gene at a time, one well at a time screen. Okay, so, just to kinda go back to the history of these types of screens, so for a long time, screens were done, RNAi screens were done in an arrayed format, and typically, the workflow was sort of summarized in that slide before, but essentially, you have a stock plate, which has one siRNA in each well. You end up transfecting that into the corresponding well of cells through one or another strategy, and then you expose those plates to some kind of a drug, for example, so in this study from Michael White's lab, they were looking at modifiers of a chemotherapeutic drug called paclitaxel, and they could find a number of genes, including proteasome, different microtubule kinetochore-binding proteins, were modifiers of the toxic activity of this cancer drug, and so, the utility of this approach was can we find genetic modifiers that might be useful in the setting of cancer, where you use a drug like paclitaxel, which is a front line care chemotherapeutic in breast cancers, for example. Sometimes, that drug fails. Could you find another target that would actually improve the efficacy of that drug? So, in some cases, these worked. Here's a pooled screen. Now, each shRNA is encoded in a single lenti-virus. These are infected into cells, and the screen here is to find, in the context of a KRAS mutant cancer, genes which, when you knock them down, further sensitize to another kinase inhibitor, a MEK inhibitor, so MEK inhibitors have been this, you know, powerful therapeutic that have been used in some cancers. They almost always fail because the cancer finds another way around it, and so this was a synthetic lethal screen to try to find combinations of second targets that would help a MEK inhibitor to kill KRAS mutant cancer cells, which, in general, are thought to be un-druggable, so KRAS is this powerful oncogene that's mutated in a huge number of cancers and is thought to be difficult or impossible to drug, so if you could find other targets that would work in a KRAS cancer, you might have a good way of killing these cells, and so, in this pooled screen, they infected a entire genome-wide library, knocking down one gene at a time. They treated with this MEK inhibitor at a dose that would kill some of the cells, and then looked for things that, when you knock them down, actually further allow killing of the MEK inhibitor. So, this is what they found. The MEK inhibitor is called selumetinib, I think, and so what they found is that, in the presence of a control shRNA, the cancers still grow, but if you knock down BCL-X Long, and you use this inhibitor together, now you have synergistic killing of these cancer cells, so this is to show that, whether you're doing an arrayed RNAi screen or a pooled RNAi screen, the technology actually works, so I'll say, there have been enormous improvements in CRISPR screens. Some work from Luke and work from a number of other labs that I'll describe, where it's possible to really improve the specificity of these screens. Having said that, it's possible to do an RNAi screen and actually get useful results, and I'll describe a little bit more about this. So, this was taken further, and they actually took tumors in mice and treated with that same combination of now an avid drug that targets BCL-X Long and the MEK inhibitor, and could actually show tumor regression, so just to show that you can take an in vitro screen, use those findings to actually translate it to anti-tumor therapy. Okay, so, you know, while there shining examples of that type in the literature, where you could find something really useful, there were also these truly alarming examples of how terrible RNAi can be, and this is one of the reasons why people have been so excited about how dramatically improved the CRISPR screens are. This is sort of a classic example in literature of three different labs who all did an HIV screen. So, ostensibly, this is a virtually identical screen. I believe they used the same cell type. It was a screen for resistance for HIV infection. I think they used different RNAi libraries. There were three different labs, but what you see here is that the overlap in their findings was actually probably lower than you would expect by pure chance, which is actually sort of impossible if you think about it, but it's one of those cautionary tales that is actually often cited in reviews as an example of how bad RNAi can be. And just to be completely fair to these folks, it's not like they're, you know, terrible labs that don't know what the hell they're doing. It's that there are fundamental problems that actually need to be considered, I think, in any screen, not just RNAi screens, but CRISPR screens as well. RNAi is subject to massive off-target effects that are typically not seen in most CRISPR screens, although I'll give some caveats for that in a minute, so Luke showed this amazing figure, where you knock down with CRISPRi GFP, and that's the only gene in the whole genome that gets changed. You would never, ever see that in an RNAi screen, because if you put an RNAi species in a cell, of course, depending on how much shRNA or siRNA, you express hundreds of genes in addition to the one you're intending to target also change, so off-target effects are a major problem. With siRNA or shRNA, reagent heterogeneity is a major issue. It's still an issue with CRISPR guide RNAs, so we still don't know what is the perfect sequence of a guide RNA, but this is an even bigger issue with shRNAs. It's actually not that easy to find a good siRNA or an shRNA. At the same time, differences in methods for calling hits, cell lines, and the protocols that are done in different people's labs will always be a source of error, and so that should be kept in mind. I think, even now with, you know, the latest and greatest CRISPR libraries and people who are using, you know, what ostensibly should be ATCC-derived K562 cells, you sometimes only see 60% overlap between two labs that have done a nearly identical screen, so there are some factors that probably won't be solved in the near future that account for differences between labs. Okay, so enter CRISPR. So, by comparison to an shRNA screen or RNAi screen, where you're knocking down a transcription of an MRNA, with the initial CRISPR screens that were done, and of course, this is now complemented with the CRISPRi-type screens that Luke has done. You can take, for example, active Cas9, create an indels in genes and effectively cause functional gene deletion, so there's a huge difference now. You're causing an effective gene deletion at the DNA level versus a knock down of transcription at the RNA level, and I should say, a lot of the things that I'm describing in the context of RNAi are now applicable and actually even better with the CRISPRi system, so you might just substitute, in some cases, shRNA with CRISPRi when you're thinking about how you might conceivably compare a knock down phenotype with a knock out phenotype. Okay, so one actually useful figure. This is, I think, from Feng Zhang's lab, but basically, they're showing the type of differences in gene expression you might see in a knock out system versus a knock down system, so even though there are these nice examples in literature of when you target a gene with active Cas9 and get nearly complete deletion, in fact, the amount of cutting and editing of alleles you get is quite heterogeneous, so you know, at best, you're gonna get genes that were never edited, some heterozygotes, and you know, some complete deletions. This is, in fact, much more complicated in some cancer cell lines where you have more than two alleles. You may have a real spectrum of different editing phenotypes, and that accounts for some of the heterogeneity. If you saw in the slide that Luke showed from Bruce Conklin's induced nuclease, there's a lot more heterogeneity in that cell type than there is in the CRISPRi system, and that's partly because you can't really control what happens when you induce editing with active Cas9. This has important implications in a screen context, because you're now looking at a pooled population of cells and you have to think very carefully about what do I call a hit, if my population of 1,000 cells, each of which knocks down one gene, has an array of different alleles being knocked down. By contrast, RNAi or CRISPRi, while you may have not a complete knock down, would at least have a more uniform perturbation, so in principle, you could get an allelic series, so a uniform perturbation and different levels of knock down. This can actually be useful in some contexts. So, this is sort of one of the examples, and from one of the early screen papers, where you can see, here's an shRNA knock down, where you have different levels of knock down, or a CRISPR knock out. This looks amazing, but I should say, this is a single allele GFP. Turns out to be a very easy gene to knock out, because even frame shifts in this gene cause effective loss of fluorescence, but what you'll see, even in this, you know, seemingly near-perfect scenario, is this population of cells that actually weren't edited at all, so this is something that you need to consider in the context of a genome-scale screen. You will have some cells that are not edited. Okay, I should also say that both of these systems suffer from off-targets, and I thought I would just briefly kind of compare and contrast some of the useful properties of each of these systems, so CRISPR, in general, has, you know, greatly fewer off-targets. I have an asterisk there because I'm gonna show you that, at least for active Cas9, there are some very significant off-target effects that should be considered, especially in the context of a screen, namely that, if you have an active nuclease in a cell, it can cause DNA damage, and that DNA damage actually has a measurable growth phenotype, and potentially can be a modifier of other phenotypes that you're interested in studying. Having said that, if you look, for example, at what Luke had showed for that single GFP knock down, you know, it tends to be much, much more specific than RNAi, certainly. It's possible to completely delete genes. It's much easier to find effective guide RNAs. Not only can you delete genes, but, as Luke has elegantly shown, you can also activate and repress gene transcription. RNAi, well, has some limitations, as I pointed out. One of the really nice things is that there's a lot less engineering required, so one of the unfortunate kind of pains in the ass about doing a CRISPR screen is you have to get all of those components into the cells, so you have to get Cas9. You have to express it. You have to get a guide RNA. You have to express it at high enough levels. You know, there's a lot of reasons why you'd, obviously, want to do that, but I will say one of the nice things about an shRNA screen is that you actually only have to put in that one component, and you know, you still have to worry about expression levels, but it's less engineering. Because they don't actually induce DNA damage, even though they have their own toxicity due to like saturation of the micro-RNA machinery and the off-target effects, there's less nuclease toxicity in an shRNA screen. Perturbation can be more uniform. It's possible to target specific transcripts, although, of course, it does have some significant increase in off-target effects, so these are some pluses and minuses to keep in mind. Some of these apply to the CRISPRi system and the less toxic than the guide or than active Cas9 for sure, but actually, many of these disadvantages are not present in CRISPRi, so I think increasingly, people will be sort of shutting down shRNA screens, although, because of this type of thing, in some cases it's actually really nice to be able to pop in an shRNA and not have to worry about getting Cas9 and a guide in. Okay, so in some of the early work that I was doing together with Martin Kampmann and Jonathan Weissman's lab, we actually were, at that time, building what we had thought was a really nicely improved RNAi system, and so I'm gonna just briefly describe this, because it actually fit into what we then used to do the CRISPR screens, but essentially, we got around this problem of low on-target efficacy and the huge problem of off-target effects by just making a massively redundant library, so instead of the two or three shRNAs or siRNAs that people had used in the past, we just put, you know, 25 shRNAs, and if you use that many shRNAs, you're bound to have some that will work, and you have enough of them that, even if there's an off-target for one of them, you can ignore it because you look at the consensus phenotype of those shRNAs. And this greatly increased our ability to statistically significantly pull out hit RNAs and screens, and of course, it's possible because of the two key technologies that enable, really, a lot of the stuff that we're doing now, the ability to synthesize these pooled libraries in a complex micro-array synthesis platform, and the ability to sequence those libraries at the end. So, our platform that we built together with Agilent, of course, builds on work from Steve Elledge, Greg Hannan, George Church, and many others who had used micro-RNA based synthesis to not just make one shRNA or the exact same thing can be used for guide RNAs, but to synthesize now tens or hundreds of thousands of different guide RNAs or shRNAs in one shot, and you can clone these because the oligos are high enough fidelity directly into a lenti-viral vector in about a week, plus or minus. It allows you to do very rapid screening in pooled format, so instead of the arrayed format, which I should have mentioned also comes with an enormous cost and just time to actually screen through those libraries, these pooled screens are quite fast, and importantly, for our purposes, that the platform is very highly adaptable, so when, you know, after we had spent all this time building what we thought was a great shRNA platform and the CRISPR revolution happened, we could just take a guide RNA and drop it into that exact same lenti-viral vector, and effectively use the same workflow to do now CRISPR screens instead of these shRNA screens. And because we can also change the algorithms that we use to design these shRNA or guide RNAs, we've tweaked the vector to make double guide RNA platforms, expression vectors, this platform has been quite valuable for making these changes. So, the general workflow for the pooled screens that I'm gonna spend most of the time talking about is we build a pooled lenti-viral library that code the guide RNAs. We infect them into cells so that each cell gets one guide RNA. These are typically already expressing Cas9. We then do a pooled screen, and I'll describe a number of different pooled screens that you can do but, you know, a standard option that we might do is to look for modifiers of a drug. I showed you two examples of different cancer drugs that have been used, so we do a lot of these in my lab, but a number of other labs have done this for a long time, to look for genetic modifiers of a drug. And in the end, you sequence these populations, and you compare the treated population with the untreated population. You count the abundance of the guide RNAs, and you can call hits. In our libraries, we use a huge number of negative controls so you can actually call hits. I think there's another lecture on statistics of calling hits and things like that, so I'm not gonna spend too much time actually describing how the hits are called, but suffice it to say, for most of the libraries that we use, we have a huge number of negative controls that allow us to carefully define what no phenotype looks like, and then you could compare the phenotypes of the guide RNAs targeting your favorite gene to that population and very clearly call hits from non-hits. At the end of the talk, I'm gonna describe some further work that we have done to take the output of this type of screen and actually do genetic interactions measurements, so at the end of screen, you might have 100 genes or 500 genes, and you would love to know, well, what do these things actually do, so we've come up with a strategy to repurpose tools that have been built in yeast to do systematic pair-wise interaction maps, and actually understand gene function in pooled format. Now, I'll describe that toward the end. So, some typical pooled screens that you might do and considerations, so as I mentioned, a standard den of workflow is to introduce a library, take your favorite drug, for example, and then look for modifiers of cell death. This tends to be incredibly easy to do, and so a lot of screens that people do out of the box are just growth-based screens, where you're looking for things that either increase or decrease the growth of a cancer cell by itself or modify the activity of a drug that'll allow you to do these sorts of screens. A number of other strategies have also been employed, so you can, for example, follow the activity of your favorite transcription factor by coupling it to either a fluorescent reporter, and then you do flow cytometry based screens, or you can couple it to expression of a chemical, a drug resistance marker, and then do viability based screens that are reporting on the activity of this promoter, so your favorite stress response, for example, you could couple to these different promoters and actually do sorting based screens based upon those pathways, and so, I think one kind of practical tip that's useful is, if you're doing these types of screens, it's useful to have a counter screen where you just look for modifiers of a general transcription or translation factor, so in addition to this, you might consider having just a constitutive promoter that drives mCherry expression, and then you could subtract everything that affects stress transcription or translation of mCherry as what you might call background in that screen, and then only look at things that specifically modify the expression of GFP, so this one probably useful constraint if you're thinking about these types of pooled screens. If you're doing these sort of death screens, we, in my lab, for example, have found about 50% killing works quite well. You can find modifiers in both directions, but depending on the type of screen that you're thinking about doing, sometimes people will use a 90% killing, the sort of apocalyptic death model, where everything dies except the five genes that are absolutely critical for a process. That can be a really useful strategy for, you know, getting rid of a lot of noise. On the other hand, you can lose a lot of things that way, so other people will screen at a 10% kill, and then you have much more dynamic range for things that sensitize to the drug, so depending on the particular paradigm that you're interested in, whether you're looking for synthetic lethal combinations, whether you're looking for critical transporters for the entry of a drug, for example, different levels of killing might be useful for those types of screens. Okay, I'm just gonna briefly go over this slide, because I still need to make one point, which is that, depending on the type of cell line that you're using, a very important consideration is just the promoter or the vector system you're using to express an shRNA, Cas9, or a guide RNA. We have collected, from ourselves and a number of other folks, including Luke, who's been incredibly generous with sending us some of these expression vectors, just an array of different promoters that have been quite useful because cell types differ enormously in the extent to which they will reliably express a guide RNA or a Cas9, and so one of the biggest issues in actually conducting these screens is establishing a cell line that reliably expresses Cas9, and so, you know, I would suggest, if you're actually thinking about doing this, talk to someone like Luke, who has enormous amount of experience in doing this, or myself or a number of other people, and you know, potentially just test a number of different constructs, because actually, trial and error can actually help enormously in finding an effective system for just expressing Cas9 and a guide. Okay, so these are sort of practical things to keep in mind. One other practical consideration I thought I would just mention is the importance of maintaining library representation, so these pooled libraries, you know, for the shRNAs we had 25 shRNAs per gene. Fortunately, for CRISPR guide RNAs, it's possible to really shrink the size of those libraries, 'cause it's much, much easier to find an effective guide RNA, and with the improved algorithms that Luke has described, from Max's work, for example, you can now shrink the library to maybe five guide RNAs. Some people are even trying one guide RNA, which I think is kind of ludicrous, but in principle, there are so much easier to find effective guide RNAs that you can really shrink the size of these libraries. Having said that, it's actually quite important to maintain effective representation of these libraries, so when we do a screen, we typically try to maintain 1,000 fold representation. This is the noise you see between two replicates at 1,000 fold. Actually, not too bad, but if it were possible to improve that coverage to like 50,000 fold, you can see the noise goes almost to zero, so it's something to keep in mind that the amount of coverage you get actually can really influence the noise in these types of screens, even with a really improved CRISPR library. Okay, so I'm now gonna describe some of the genome-scale screens that have been done, and a few of the first ones with CRISPR were using active Cas9 and were conducted by David Sabatini's lab, Feng Zhang's lab, and what they basically did was to take an oligo design, which, in effect, incorporated some of the same rules that Luke described in terms of optimizing guide RNA properties. A lot of this stuff has been published, and effectively targeted what were conserved axons, and they thought to go as five prime as you could with the idea that they wanted to make a guide library that would work on as many different transcript variances they could find, and if you put a stop code on as early as possible in the gene, you have the best chance of actually effectively deleting that gene's function. So, in general, this worked pretty well. This is sort of a V1 library, and I'll show you some other examples where some of these rules have been tweaked and improved upon, but the idea was you take an oligonucleotide synthesis platform, you generate a genome-scale library. I think, in this case, they did maybe three to six guides per gene, and then the workflow is basically exactly the same as what I described. You have a lenti-viral library. Each vector code for one guide. You infect it into cells, and then you can do a screen. And so, one of the first screens that they did was actually just a growth screen, so this is effectively called a screen for gene essentiality. You wanna find all the genes that are required for growth or viability in a cell. One of the interesting things you can see in this plot, so here, you're seeing that the ratio of cells compared day three to day 14, so this is, you know, growth over a two week period, and what you can see is that you now, after 14 days of growth, start to see this population of cells that are lower in abundance than they were at day three, and that is because they're dying, of course, so you knocked out a gene that was important for their growth. You can see what some of these genes are. They're key cellular processes, RNA processing, RNA binding, ribosome biology. These are core machineries that are absolutely required for cell growth. The interesting thing, actually, is that it's much easier to break something in a cell than it is to make it go faster. There are very, very few genes which, when you knock them out in a cancer cell, actually make those cells grow faster, compared to the thousands of genes, actually, depending on how you draw the threshold for growth phenotype that actually break or significantly impair cell growth. So, this was nifty. One of the other things that they did in this paper was to do a drug resistance screen, so I just charted growth screen where you're knocking out a gene and you're just looking at across time, at an early time point and a late time point. In this screen, what they did was they put the library in, and then they treated with this BRAF inhibitor that normally kills the cell, except if they knock out these genes, which, it turned out, are different, you know, direct members of this pathway or genes that had been previously shown to cause resistance to this BRAF inhibitor. So, one of the cool things that they highlighted in this work was that all four of the guide RNAs targeting this gene, for example. You're right, Luke. It's totally impossible to see this thing, but in F2, for example, all four guides actually show the same phenotype. You know, they're considerably higher in the treated cells than they are in the untreated cells, and this is something you would basically never, ever see in an shRNA screen, so you would never see all four shRNAs giving you the same phenotype. Here, you see all four guide RNAs can cause significant resistance. And the other, you know, interesting observation is that you don't have 1,000 genes that are required for HIV resistance. You have eight genes or 20 genes or something like that, so you have a really narrow list of almost certainly right hits, and when they actually go in and validate these, you know, 90 plus percent of them validate, so I think it shows the power of these screens to get really clean phenotypes for gene knock outs. One caveat that I thought I would point out, from our own work, is that if you looked at this, you might say well, all four of these guides are working really well. In fact, I just said that, but in fact, there's a lot of heterogeneity in the performance of those guides, and you don't really see that until you really look carefully at the phenotypes for all the different guide RNAs, not just in a resistance screen where you're really driving all those things to cause, you know, an increase in growth, but in a drop out screen, so in this plot, what we're showing is actually the phenotype of a really critical essential gene, ABL. BCR AbL is the gene that drives K562 cells. If you knock it out, they're dead as a doornail. And what you can see here is that, of all the guide RNAs that target it, there's a huge range, and in fact, some of them are killed completely, and some of them really have a pretty mild phenotype, and so, it's important to consider that, with any screen, you're going to have a range of guide RNAs phenotypes. Okay, so one of the other things that we've done that's been quite useful with these types of screens is to actually do a drug target identification, so I showed you a couple of examples where people had looked for modifiers of a drug, and in most cases, those drugs already have a known target, so they're targeting some kinase, which is thought to be really important for cancer cell growth, and you wanna find a second target, which is important, for example, for, you know, a potential synergistic target for that drug or to understand how that particular cancer works. Another useful application of these screens has been to identify de novo, the target of a drug that does something interesting, but you actually have no idea how that drug works, so in this paper, together with the Cleary Lab, who had done a chemical screen, this is actually quite common in the pharmaceutical industry. People do, you know, giant chemical screens. They find some drug that magically kills leukemia cells, in this case, but not normal cells, so it looks extremely interesting, but they have no idea, actually, how this thing works. What we were able to do is to take an shRNA library and this is actually just a half genome screen, and systematically knock down one gene at a time, with the idea that, if you knock down the target of that drug, now that cell is much, much more sensitive to the drug than all the other cells, so you can see, with actually pretty remarkable, actually maybe opening too remarkable specificity, we were able to pull out what happened to be the molecular target of this drug, which is this NAD biosynthesis gene called NAMPT. You might say why didn't you find all the other NAMPT genes? So, one thing I should say is we probably should have just stopped doing screens entirely at this point. It was a really nice result. It's a half genome screen, and probably, because this was an early shRNA library, we had bad shRNAs for the rest of the components of that pathway, but it does illustrate the general principle that you can use these genome-scale screens or near genome-scale screens to do unbiased drug target identification if you don't know the target of a drug. So, in another example of a similar type, in my lab, we compared, we actually tried to find the target of this drug. This is now an anti-viral drug which targets the host cell, so there are a lot of different anti-viral drugs in the world. Typically, they target the virus itself, and so, if the virus mutates, you just need to find a new drug for that new virus. This is a pain, and so, if you were able to find drugs that actually target a host process that caused a protective phenotype against a range of viruses, that could actually be quite useful. And so, we set out to actually identify this target, which had been identified by GlaxoSmithKline in a similar, you know, large scale chemical screen for compounds that inhibit the replication of a range of different viruses, so they found this thing, GSK983, that looked like it was effective at inhibiting adenovirus SV40 and HPV-16, and they actually reported a kilogram-scale synthesis of this compound, so this is another common thing that happens in the pharmaceutical industry. The do a lot of work, probably millions of dollar worth of research on this drug. They synthesize kilograms of it, which means they're testing it in animal models, and then, inexplicably, they just drop it, and they have no idea what the compound did. Typically, that's why they drop it. They don't understand how it's toxic. We were able to, so I shouldn't, by we, I mean, Richard Deans, who's a chemistry student in the lab, together with Chaitan Khosla, who synthesized this compound, and we used the fact that it's toxic to actually find its target so we could add the drug. We knew that it caused some growth phenotype, and we could look for modifiers of it, and the reason I'm spending some time describing this setup to you is because what we did was to actually systematically compare an shRNA screen with a CRISPR screen, and it was actually very interesting to do this comparison, so you know, the screen workflow is exactly the same as what I had described to you. In both cases, you're infecting a lenti-viral library, in this case, one shRNA per cell, in this case, on guide RNA per cell, and we do parallel screens. One pool gets treated with the drug. One pool is untreated. At the end, you do sequencing and you compare and you ask which genes, when knocked down or knocked out, modify the activity of this drug. So, I'm gonna cut through an enormous amount of work to describe what is really, I think, the most interesting part of this story. So, what I'm showing you in this plot here is the phenotype in the shRNA screen on this axis and the CRISPR screen on this axis, and so, to make a very long story short, the target of this drug is this gene. It's DHODH, is a critical component of de novo perimetry nucleotide biosynthesis, so viruses, when they infect a cell, massively up-regulate nucleotide synthesis so they can support their own replication. If you can't do that, then the virus can't replicate, and so that's why this drug was actually an effective anti-viral. Unfortunately, it's also toxic, because the cells need those nucleotides to support their own DNA replication. It turns out, there's actually quite a nice therapeutic window there, where viruses need a lot more nucleotides than a standard cell does to stay alive, but the interesting thing here is that you can see, when you knock down this gene with an shRNA, you could actually, it's one of the strongest sensitizing factors to GSK983, and that's how we were able to find it, is because we could knock it down and show that it actually sensitized against the drug. In a CRISPR knock out screen, this is an absolutely essential gene, so if you delete it, those cells just die, so really, there was no phenotype in the CRISPR screen, because by the time we actually took that measurement, actually, you know, several weeks had passed and all those cells, for the most part, had dropped out. I think there's actually some interesting counterpoints to this idea. I think a number of people have now seen, in a CRISPR screen, you can sometimes see a modification of an essential gene by getting, you know, in some cases, hybrid morphs that give you, you know, deletion of one allele or two alleles, and it's telling it has five alleles, so there are some counter examples, but I think we've seen, in a number of cases, this very interesting difference between, you know, the utility of a knock down phenotype, which you might get with a shRNA screen or a CRISPRi screen and a complete knock out, where you're really eliminating all of the function of a gene. Now, conversely, these genes, which are negative regulars of mTOR, only showed up in a CRISPR screen, so the best shRNA in the world might give you 98% knock down, and if that is not enough to give you a phenotype, you won't see it, so these two negative regulators at mTOR, which, ironically, actually show up in just about every screen we do for some reason. mTOR is a very important gene that's involved in a lot of different processes. Only could be found with the CRISPR screens, so you really need complete deletion, so the interesting kind of take home from this is that sometimes it's good to have knock down and sometimes it's good to have knock out, and we've actually found, by comparing the two screens, sometimes you get a much more complete picture of the biology. Just as a side note, we found not only the de novo synthesis pathway, but the salvage pathway. Those cells have two ways of taking up nucleotides. They can either make them themselves, or they can take it up from the outside. We found both of these pathways in the screen, and we're now exploring a combination strategy where we block de novo synthesis and salvage to actually improve the activity of these compounds as anti-virals. One other side note, by the way, comparing shRNA knock down and CRISPR knock out screens is that we've actually, when we've done these types of comparisons with the same libraries side by side, it's not that one library is crap and the other one is good. They actually both do a pretty good job of calling essential genes, so if you take a gold standard essential gene set, so these are genes that have been found in like 100 cell lines to always be required for growth when you knock them down, in this case, with an shRNA library, and you say that's my gold standard, and so, whenever I find that, I'm right. If I don't find it, I'm wrong, and you draw this type of ROC curve, so up and to the left is better. You can see that they actually, both the CRISPR library and the shRNA library actually do a pretty good job of calling essential genes, but when you combine the two, you actually can do better, which means that there's non-redundant information in those two screens and it's actually useful to combine the two. Yeah? - [Attendee] So has anybody looked at designing, basically, for want of a better word, crappy guide RNAs that are more likely to cause like a negative three, negative six, negative nine, so more likely to give you that intermediate knock down phenotype? So you make a slightly crappier protein, but you don't get an early stop on the line? - That's a really good question. So, in some cases, I think, if you're targeting the protein coding region, it's not a trivial problem to figure out where, exactly, you would put those guide RNAs to have a crappy effect. I'll show you a study where they actually scanned across a couple protein coding regions to find what would be good, and maybe in those examples, you could say well, this would probably be bad or relatively bad, so that you have an array of phenotypes. I think, in CRISPRi, it may be easier to conceive of a system where you would really try to, I don't know, Luke, how you guys would think of doing this, but step-wise, move away from the promoter until you realize that, you know, instead of having as efficient knock down as you can, you have, you know, a range of expression levels. There, it might be a little bit more uniform to do that kind of thing, but it's actually an important question, because sometimes you really would like to have a range of phenotypes. Yeah. By the way, feel free to interrupt. I realize I'm kind of rambling on here, so if any of you guys have any questions, feel free to shoot away. Okay, so one of the nice properties of this system for drug target identification is you don't need to modify the drug, so, you know, typically, if you wanna find the target of a drug, you'd have to stick a cross-linker on there. You potentially destroy its activity. It's not sensitive to the strength of the interaction, and you can find those targets in the biological context. Okay, so, in the examples I've discussed so far, typically, the guide RNAs have been targeted to a protein coding region, but there have been a number of other screens that have been done where people have started to look at the non-coding genome, so of course, in the context of a CRISPRi, you're typically targeting upstream of the transcription start site, but in a number of studies, people have now used active Cas9, and actually, increasingly, a number of other of these variants that Luke described to really start to probe, you know, what are the functions of not just protein-coding genes, but actually regulatory elements, and so this is a study that was done in Feng Zhang's lab, for example, where they tiled, you know, 100 KB upstream and downstream of this gene, which turned out to be, you know, one of these genes that they had found in their previous BRAF screen. And the workflow is effectively the same. You take this library of guide RNAs. Now, not targeting the protein-coding region, but targeting a very large region of upstream and downstream regulatory elements that control the transcription of that gene and ask for which of those elements might actually affect the transcription of a gene. Earlier on, actually, this group and Dan Bauer's group together with Stuart Orkin, did a quite similar screen. I happen to like this diagram better this way, but these guys actually did it first, to tile, again, the regulatory regions that are around BCL-11, so this is a gene which has an enhancer that actually affects the transcription of, again, one of these fetal hemoglobin genes. So, in this case, they're looking for modifiers of this BRAF inhibitor. In this case, they're looking for, when I knock out this element of a regulatory region, how do I affect the expression of this fetal hemoglobin, and this, I actually do in a flow cytometry based screen, so they do an antibody staining for fetal hemoglobin. They do cell sorting for either things that, when you knock out that element, either increase hemoglobin or decrease it, and then they actually call hits based upon which regulatory element modified the expression of that gene. And this is the type of data that they get out of it, so if they take, you know, thousands of guides and they tile them all along these regulatory regions, and they ask for which things actually affect the expression of those hemoglobin, they can pull out a locus like this, where they actually see differential expression, actually in human cells compared to mouse cells, where you have this specific region which has a GATA1 transcription factor binding site, so you can actually map important regulatory regions for the transcription of genes by using Cas9 to tile and, in principle, create deletions that prevent the binding of those transcription factors. And so, in the end, they can, you know, take these sort of data points, where they have modifications and expression of a gene, and then relate those to the binding sites that have been determined through chip-seq, as Luke described, where you can actually now try to functionally annotate which of these transcription factor binding sites or 3D contact loops are important for controlling gene expression. So, this is, you know, a significant departure from the typical protein-coding screens that have been done. Now, you're really asking quite complex questions in some cases, about which of these different regulatory control elements drive expression of a gene. Another really nice example comes from Luke, sorry, John and Max in Jonathan's lab, together with Daniel Lim. In this case, so I think Luke had described this briefly in his previous talk, but what they were able to do is to target long non-coding RNAs in cells, and so this is actually a uniquely, incredibly powerful application of the CRISPRi system, so if you think about it, CRISPRi has been very useful, and I'll show you some slides in a minute where the same screens have been done extremely well with CRISPRi, but what's nearly impossible to do with a active Cas9 screen is to actually target these non-coding RNAs, so these are, in some cases, very long RNAs, which are not coding for a protein, so you might put an active Cas9 in the middle of it, but what would that actually do to the function of a very long RNA for which, you know, particular elements may not be important. In this case, they were able to use CRISPRi to actually target the promoters of those long, non-coding RNAs, and screen for essential gene phenotypes, so some of this data, I think, we've showed, but the general principle was that they were able to do growth screens in a range of cell types targeting the expression of those different long non-coding RNAs. You know, the group were actually able to find some very interesting differences, different cell types that had different long non-coding RNAs that were essential for growth. So, this is an example of a screen in which they were scanning the protein coding region, so this is to the question that was asked earlier. How would you design a guide RNA that was especially bad. In this case, they were actually looking at sort of the opposite effect, so, you know, how would you design a guide RNA that was especially good for targeting this chromatin modifier, BRE4. So, what they did was to take guide RNAs really tiling the entire protein coding gene and look for guides which, when they knock it down, actually cause a growth phenotype in these cells, and the interesting thing that they found was that, when you put a guide RNA in some of these really conserved, you know, known functional elements of the protein, so these are actually some of the active elements of the protein. These tend to be much more impactful when you knock them out, and the thought is here that, you know, in some cases, you have a guide RNA that targets the middle of this region, and with a significant probability, you actually don't create an indel or a stop codon, whereas these guide RNAs, even with an in-frame deletion, may have enough of an effect that, you know, if you have a very highly conserved kinase domain, for example, you may lose a single amino acid, and that may have an enormous impact on the function of that protein. That would not be present if you were to lose a single amino acid here, so their rationale is, by targeting highly conserved, functional regions of proteins, you could actually have more effective gene deletion, and in some cases, they've actually found this allows you to map the functional domains of a protein, so if you didn't know that these were the important functional domains, you might actually find them by tiling across a gene and actually seeing, well, this is the place where the protein really hates to have an indel placed, or even an in-frame deletion. Okay, so a really important caveat that I think should be kept in mind for a lot of these screens with active Cas9 is that you are putting an active nuclease in the cell, massively over-expressing it, and expressing a guide RNA, so this is something that, in some sense, may be especially important in the context of a screen, where you're constitutively expressing Cas9 protein and a guide RNA over a long period of time. In some therapeutic applications, where you're transiently expressing a guide RNA or an RNP, it may not be as important to consider these types of off-targets, but actually, in this case, what a number of groups have found, this is a work from Bill Hahn or a group of Novartis. We've actually found quite similar things in our own data, and a number of groups have found this, including Sabatini early on. What they showed was, if you look at the behavior of guide RNAs across a chromosome in a cancer cell line where you have clearly amplified regions, guide RNAs that target these amplified regions are extremely toxic, whether or not there's a gene there, and this is because if you have enough copies of, you know, a piece of DNA, you can actually cause a significant amount of DNA damage. In some cases, so significant that the cell will just effectively die, because you have this massive, you know, DNA damage response, so this is especially important given that, you know, many of us are using these, you know, highly amplified and messed up cancer cell lines to do our work in, and K562 cells, for example, BCR-ABL locus is massively amplified. It's critical for growth. It turns out if you put a guide RNA in between, you know, the amplified genes, you can actually still see massive cell death, so this is something that needs to be kept in mind when you're doing these screens. We've actually found even a single cut has a measurable affect on growth, so I'm showing you a library that we built to try to make an effort at controlling for this effect, so we built our own sort of genome-wide CRISPR cutting library, where we had a set of negative controls that targeted what we thought should be safe regions in the genome, so these are regions that have no annotated, there are no genes there, there are no chromatin marks there, there are no regulatory elements, there are no, you know, long non-coding RNAs, so across 170 cell lines, there's nothing at all there that should be, what we think, functional, and we put a bunch of controls there with the idea that we're gonna see what the effect of cutting at one site would be in the genome. So, this is the work of David Morgens and Michael Wainberg in the lab. To cut to the chase, the sort of interesting observation here is, what I'm showing you here, is the distribution of guide RNAs for either gene targeting guides, so you can see, there's a range of phenotypes. Some of them are very highly deleterious when you knock them out. Some of them are, you know, slightly protective. If you compare the phenotype of a non-targeting guide RNA, so this is a guide RNA that we computationally predict should not have any match in the genome with tolerating up to three mismatches, versus one of these safe-targeting guide RNAs, which should have exactly one cutting site in the genome, you can see that the distribution of the safe guides is actually considerably broader than the non-targeting guides, which means that, actually, these have much more of an effect than non-targeting, and so, you can actually even see this when you look at the substitution position, so depending on the mismatch position, the effect of those safe targeting guides actually differs, which reflects known properties discovered by Jennifer and others about, you know, where a Cas9 can tolerate a mismatch. So, the take-home message is you should actually, if you're doing a Cas9 screen, or even an individual re-test experiment, if you're gonna use a control, use a guide that cuts somewhere in the genome, because there is a measurable affect on growth and some actually measurable DNA damage that occurs, so don't use a non-targeting guide, 'cause it's not really an appropriate control. - [Attendee] This effect probably depends a lot on what cell line you're using, because P53 mutants, for example, have more sensitivity to single breaks. - Yeah. (attendee speaking off microphone) That's definitely right. In fact, you can see that here, so we did it in, I believe, three different cell lines and you can see in some cell lines it didn't quite matter as much. It was not a perfect correlation with P53 status, but you're absolutely right. In some cases, it matters quite a bit more than others. Yeah. Okay, so fortunately, Luke gave a fantastically complete introduction to CRISPRi and a, so I'm not gonna describe this in any more detail than he just did, other than to say those significant problems with toxicity due to nuclease cutting can largely be eliminated by using either CRISPRi or CRISPRa to do these screens, so this is, because you're not using an active nuclease, you're using the dead Cas9 just as a targeting domain, you can really, to a pretty large extent, ignore those off-target cutting effects that have been, you know, a significant consideration. In some cases, it matters a lot more than other, but one of the major advantages of this approach is that you don't have to worry about the DNA damage you cause with an active nuclease. And so, as Luke mentioned, there are several different strategies to do this. I'm not gonna belabor that point, and the other thing that we found, even in our own screens, is that the point that Luke made about the importance of finding the position, the right position to put those guide RNAs is something that I think everyone can attest to, and so what I would say is, if you're interested in doing these CRISPRi experiments, the new and improved library is really significantly new and improved. It really actually works a lot better that the V1 library, and so I think one of the major improvements was just annotation of the transcription start sites, if I'm correct, so with an improved transcription start site annotation, A, you were able to do, and you know, the significant improvements in the algorithm that Max made, they can now very effectively effect transcription or oppression across the genome, and so these libraries work actually quite well. So, here's a couple of examples in two different papers, so in this one and the new one. So, one of the screens they did early on was one of my favorite toxins, the ricin toxin, near and dear to my heart. Don't eat it, but actually, it's really an amazing toxin, actually, so it gets endo-cytosed, it traffics from the endosome to the golgi to the ER. It gets unfolded, recognized as a misfolded protein by the ERAD machinery, retro-translocated back into the cytoplasm where it refolds, and only then can it kill the cell by de-purinating ribosomal RNA and shutting down translation, so it's just really a kind of amazing toxin that does all these neat trafficking and folding events, and we could use that to find all the genes in a genome-wide screen that were involved in its trafficking or processing. What was shown in this really nice figure is that you can actually see, depending on whether you knock down with CRISPRi or activate with CRISPRa, the expression of these previously demonstrated ricin hits. You can see an opposite phenotype, which is really cool, because it means that you can not only, you know, recapitulate some of the same biology that was found in other screens, but you can actually be even more confident. If you could flip the phenotype from repression to activation of sensitivity, for example, you can be even more confident in the phenotype of that gene. And this is just to show, using the same kind of gold standard essential genes in the new and improved library, you can see a near perfect recapitulation of essential genes with the CRISPRi library, so these libraries work really quite well, equivalently well to the best CRISPR active Cas9 libraries that we've used. Okay, so, in the last bit, I'm gonna describe a number of different sort of applications of these screens, so one really cool application that was just published, so you're gonna get a talk, I believe, tomorrow from Tom Norman, who worked together with Brett Adamson in Jonathan's lab, and together with Aviv Regev in a number of other labs. They had described very recently this kind of extraordinary, powerful redoubt for a screen, which I'm not gonna go into in any great depth, because I'm sure Tom will do it, but just to highlight, in the context of screens that you might do, it's now possible to not just look at a single gene phenotype, protection and viability or the activation of a reporter, but to look at, effectively, the modifications to the transcriptome, so a really high-dimensional multi-plex output to go with a multi-plex input of a genome-scale screen, so Tom will describe that in more detail, but I wanted to just put this in the context of screens, because it really, in principle, expands the range of phenotypes you could look at. So, yeah, in the last little bit, I wanna describe two applications that we've been pursuing in my lab for these types of screens. So, as I mentioned, everyone and their mother is doing some kind of a screen. You end up with a long list of hits, and you might say, well, I know what the first 10 do, but what about the next 50. How do I actually understand, at a systematic level, how these genes actually work together? And so, early on, when I was a post-doc together with Martin Kampmann in Jonathan's lab, we had done a genome-wide shRNA screen for ricin modifiers, and what you can see in this map is that I'm displaying the phenotypes of a single gene in combination with a series of second genes, and so, if it's yellow, that's better than expected, so the combination of the two genes, when knocked down, is a much better phenotype than you would have expected, and blue is significantly worse than you would have expected. And so, depending on the nature of that interaction, it can tell you something about how those genes actually work together, so the really cool thing about these maps is, if you cluster genes based upon the similarity of this pattern, what you can effectively do is reconstitute a lot of the biology of the cell, so we did a genome-wide ricin screen. We took a lot of the hits from that screen, and we clustered them together in this map by their function, and what you can see is that many of the genes, even if we didn't know what all those different trafficking steps were, we could have found them because here are all the ribosomal genes. Here are all the COP1 vesicle trafficking genes. But the really cool thing is if you have a gene of unknown function, so here is C5orf44. Previously, we had no idea what this gene actually did, but we can see that, genetically, it looks like a member of this vesicle tethering trap complex, and so one of the really nice things about this was it came up with a prediction for the function of this gene. Genetically, it has a pattern of interactions that looks like another complex that we do know the function for, so now we have a prediction for what this thing does, and we could actually show that, indeed, this is a physical member of this complex, and in fact, defines two new complexes, there are sub-complexes of TRAPP. So, these were very powerful tools for kind of a systems level understanding of what the hits from a screen would do. So, in my lab, we're interested in applying sort of a new application of this pair-wise genetic interaction screen, in this case, to look for, specifically, cancer drug combinations, and so why might you wanna do this? Of course, cancer drug resistance is a very widely appreciated problem. There's this famous picture that many of you have probably seen of this unfortunate patient who presented with a melanoma. He was treated and had a really nearly miraculous response to this BRAF inhibitor, which completely shrank all of his tumors, but as almost always the case in cancer, what ended up happening was, eventually, the patient relapsed and all the same tumors came back and were now resistant to this BRAF inhibitor, so there's an enormous interest in the pharmaceutical industry, and this has, of course, been an interest for a very long time in trying to find combinations of targets that would allow you to prevent the escape of these cancer cells to drug resistance. In addition, of course, it's very expensive to develop these kind of drugs, and so, if you could repurpose existing FDA approved drugs, there would be a lot of utility in that, and of course, testing all those combinations, even if you did know how they would work together is an enormous effort, and so what we wanted to do was to actually model drug action by using CRISPR to target the targets of FDA approved drugs, so we wanted to systematically look for drug combinations by targeting pairs of genes for which they had a corresponding target. This is the work of Kyuho Han and Edwin Jeng in the lab, very talented post-doc and a great graduate student. They were helped by David and Amy, and the idea is we generate a very large library, in this case, 500,000 pairs of guide RNAs that target combinations of genes which have drug targets, and the idea is can we pull out synthetic lethal combinations that might be effective in cancer. So, I'll skip over a lot of work that these guys did to express pairs of guide RNAs. There are now a number of different systems that get around some of the problems that Luke described to express pairs of guide RNAs, so we built one. The Weissman Lab built one. The Ventura Lab built one. I think, as Luke pointed out, it's nice to have some different options for how to do these things, 'cause one of the key problems is if you have identical promoters, you can't actually sequence the guide RNAs directly, and you get a lot recombination between the guides, so in the end, we built the system that can express pairs of guides that we can actually do deep sequencing on, directly sequence the guide pairs, and what we did was we took genes, which had a corresponding drug according to these drug databases. We selected ones that, by themselves, were not lethal because if a single gene kills a cell, you're not gonna find a synergy, and we took genes that had a mild phenotype. We selected the three best guide RNAs from our genome-wide screens, and we made a pair-wise combination library. In this case, targeting about 207 drug targets, which, in the end, gives you roughly 500,000 pairs, corresponding to 20,000 possible drug combinations. In the end, the workflow for this is exactly the same as all the screens that I've described to you. We now take a double guide library instead of a single guide library, and we infect that into cells, and we calculate interactions based upon how different is the expected pair from each individual gene when you knock it down, and we want to find the really rare synthetic lethal combinations that completely kill a cell only when you have those two genes knocked out. So, the data are highly reproducible. It doesn't matter if you flip the guide from one position to the other. Replicates are highly reproducible, and here's what the data look like. So, largely, you have a huge cloud around zero, so what I'm showing you here is the genetic interactions core. In one replicate, on this axis and another replicate on the other axis, and what you can see is that, by and large, most genes actually, in the set that we took, don't interact at all, so they have zero interaction. They're exactly what you would have expected, and this is probably not surprising, because we really took a random set of genes that have nothing to do with each other. They just happened to have a drug target, and we decided to look and see which of them, when you knock them out in combination, actually are synergistic, and the answer is very few. What's reassuring is, if you knock out a gene and then you use a second guide RNA against that same gene, now the phenotype of the double mutant is actually better than you'd expect. That's because once you knock out that gene, it doesn't matter if you target it with the second gene, for the most part. There's effectively the same phenotype, and so it's better than you would have predicted, but what's really nice is we could pull out this rare, actually quite potent, synthetic lethal combinations which, when you knock out the two genes together, those cells really have a hard time growing. So, in the end, we decided to focus on this combination of a Bcl-X long inhibitor and an Mcl-1 inhibitor. These are two anti-apoptotic genes, and you might say to yourself, well, of course two anti-apoptotic genes are gonna be synthetic lethal. In fact, cancer cells differ enormously in the extent to which they depend on particular anti-apoptotic genes for their survival, and so it was actually interesting that we could pull out this pair, that happened to also have two pretty effective drugs that can be used against them. We also did find a number of other, you know, pairs which were synthetic lethal. Some of these make perfect sense, 'cause they're two different isoforms of AKT, two different DNA-damaging, DNA damage repair pathways, and so we felt like we were getting, you know, legitimately synthetic lethal combinations. So, to make a long story short, we were able to show that the combination of these two drugs corresponding to the two genes that we found to be synthetic lethal in this screen were actually quite potently synergistic when we used them together. What was really cool, though, is we could actually see that, so here's K562 cells, where we're treating the cells with these two drugs in combination. You can see really potent synergy. When we use these either EBV transformed LCL cells or primary cells, so these are CD34 hematopoietic stem cells, there's actually much less synergy between those two drugs in this case, and so it was actually quite nice to see that the pair that we found were not just universal synthetic lethals. They were actually quite specific to cancer cells, and in this case, when we made imatinib-resistant cells, so these are K562 cells that have been grown, so in some sense, K562 is like the least interesting cancer to solve, because there's already a fantastic drug, imatinib, which targets BCR ABL driven cancers, but in some cases, you still get resistance to this drug, imatinib, and so if, when we generated it in the lab, an imatinib resistant version of K562 cells, it was actually really nice to see that this same combination of drugs was still lethal to those cells. So, this was kind of satisfying, because we took what was really an enormous library of guide pairs and we screened for very rare synthetic lethal combinations, which we could pull out and then show that the corresponding drugs were actually synthetic lethal in addition. So, that was nice, although, if you actually go back to the type of genetic interaction map that I showed at the beginning, if you did this same interaction map with the data that I just showed you from this drug map, you would see mostly black space, because, as I mentioned, there's really very low frequency of interaction, so if you tried to do the type of clustering based upon similarity of interaction patterns that I showed was useful in ricin, you get virtually nothing in terms of useful clusters coming out of an interaction map for this drug map, and this is not that surprising, because the frequency of just two random genes in the genome has been shown in, actually, a number of different studies in yeast, Drosophila, human cells, is actually quite low, 'cause the number of synthetic lethal combinations with any given gene is probably maybe 20 to 50 across the entire genome, depending on the genes. Some genes interact much more than others, but on average, the frequency of interactions is actually quite low. So, we went back to our favorite toxin, ricin, and we did a genome-wide CRISPR screen now, instead of an shRNA screen. You might say, well, that's incredibly boring. Why the hell would you do that? But it's because you can actually find a lot of new biology. Strikingly, a huge number of new genes come out of a CRISPR screen that we had, you know, done virtually identical thing with shRNA. But now, with more powerful knock out tools, we get a lot of new genes that come up. So, for example, all the genes that are involved in glycosylation of cell surface proteins, we didn't really see in the shRNA screen, but are among the strongest hits in the CRISPR screen, so this ricin is a lectin that has to bind to glycosylated residues on the cell surface. If you don't make that glycosylation residue, ricin can't get in, and so, when we knock those genes out, the cells behave as if they never saw ricin. It was kind of amazing. The interesting thing was, in this study, we could show that even those genes, so here are the same genes we've found previously, these trafficking genes, but then also, for example, all the glycosylation genes clustered next to each other, so we were happy to see that you could use the same guide pair platform to do these same types of interaction screens we had done before. Okay, so, in the last few minutes, I wanna tell you about this new tool that we built. So, Luke briefly outlined the strategy for doing mutagenesis using dCas9 to recruit deaminases, and I would like to describe, sort of, one of the applications that we've found for this, that's more of sort of a screening type application, and so, if you think about the way that people have used Cas9 to do mutagenesis, so traditionally, you know, work from Jennifer and a number of others have shown that you can create these indels, which will cause loss of function. Actually, people have used, as I showed you in these other tiling screens, the same active Cas9 to look for functional regions of proteins. However, another obvious application is to introduce by homologous recombination a library of variants, and so, a number of different labs, including work from Jason Drury's lab, have shown that you can take active Cas9, a single guide RNA, and then a library of oligonucleotide donors to introduce one mutation at a time into a specific site. What this allows you to do is then start with a library of point mutants that result from that donor being integrated, and then do some kind of a selection to find, you know, what are the particular point mutants that affect your protein of interest. And so, what we were trying to do is, instead of having to do this every time we wanted to make a new mutation, that is make a cut, make a new oligo-donor library, and introduce it one at a time at each spot you wanna make a mutation, what we tried to do was to make a diverse, functional mutant libraries in situ, without having to use an HR donor, and I emphasize functional here because we wanna use, we would like to avoid indels, so for a lot of protein evolution applications, you don't really wanna destroy the protein. You wanna find a new function, so you would like to avoid, if you can, introducing stop codons. So, we turn to the way this is done in nature, so with really remarkable specificity, during antibody development in B cells, your body has a way of targeting mutagenesis specifically to the variable region of an antibody, and really, almost nowhere else in the genome. It's kind of an amazing process, but you get quite diverse populations of point mutants developing in that diverse region of the antibody, and that is what allows you to have very high affinity antibodies against your favorite pathogen. This is because this enzyme, AID, generates mutations by de-animating cytosine, which is then repaired in a variety of different ways, and I won't really belabor that point, but just suffice it to say you can actually get quite a diverse population of mutations just resulting from this cytosine. In fact, even away from the actual cytosine you're removing the base from, you can get mutations because you recruit error-prone polymerases, which actually incorporate other nucleotides where they shouldn't. So, this is work from Gaelen Hess, a very talented post-doc in the lab, who took dCas9, so we, being nerds, we called this CRISPR-X to generate mutations. Fans of the X-Men, but this is probably a mistake, but anyway, but the idea was you can recruit, as Luke described, using these mS2-fused variants of AID to the genome and create a diverse population of point mutants. So, Gaelen spent some time engineering this enzyme to get a hyperactive variant of it, so he chopped off the normal export signal, so in normal B cells, you actually want to keep this out of the nucleus most of the time, 'cause it's actually quite damaging to have this thing bopping around, removing cytosines, so it's normally exported. He chopped off the export signal, used a variant that had been identified in bacteria that had a higher activity, and then fused it to Cas9, and in the end, what you could do was to actually use a guide RNA to target a population of point mutants, really exactly where you put the point mutant, and what was satisfying to see is, so, David Lu, actually a few months before we were able to publish work, showed that you could use a different deaminase, so Luke pointed this out, so they used an apobec. We had been working on this for about a year and a half. It was actually kind of sad when they first published it, but I think what we actually ended up doing was quite different, and it's very interesting. So, what they showed was you can take this apobec and get really remarkably precise conversion of a C to a T, effectively with their new and improved version within like a five base pair window, so their proposed application was if you have a disease where a C is supposed to be a T, you can put in that deaminase and actually do a therapeutic editing. In our case, what we're doing is creating a window of actually reasonably diverse point mutants, with the improved version about 100 base-pair window of the guide RNA, and in fact, we get C to basically everything, G to basically everything, and even some measurable levels of As and Ts being converted, so you can generate a quite diverse library of point mutants exactly where you put the guide RNA. So, this was kinda nifty, because now you have this population of mutants, and you can actually start to do the type of classical genetic screens in a way that I described on the first slide, where you create a population of mutants, and now you can ask what is the thing that gives me a new function? So, one of the initial kind of applications that we thought about was to do evolution of GFP, because we know exactly what turns wild type GFP into EGFP. It's brighter and the mutations that do this are known, so we put guide RNAs around this locus, and GFP would cause mutants to be formed, and then we sorted for a GFP variance that had brighter fluorescence by flow cytometry, and after one round of sorting, we could already recover this known S65T mutation, so G to C, which is what's responsible for converting wild type GFP to EGFP, so it's actually, you know, of course we knew what should happen here, but it was satisfying to see. We could mutagenize this population and recover a variant that gave us a new function. More interesting was to do a tiling mutagenesis of PSMB5, so this is a protiozome sub-unit, and is the target of this chemotherapeutic drug, bortezomib, and so what we did was we took a library of guide RNAs targeting the entire protein coding region. We introduced this mutagenesis machinery, and then we pulsed with drug bortezomib, and asked for cells which now become resistant to this chemotherapeutic drug, and what you can see is we could recover in a PSMB5 targeting library, but not in a library targeting these safe regions in the genome, mutations that confer resistance, which was actually quite nice to see. In fact, what was interesting is this axon three is actually not expressed in K562 cells, and we saw no mutants coming out there. So, when we looked at where these mutations map onto PSMB5, so fortunately, the structure of PSMB5 is known, its structure in complex with bortezomib is known, and previously, people had identified in resistant cancer cells, mutations in the binding pocket of PSMB5 for bortezomib, and so we were happy to see that we could actually recover those same mutants that set right in the binding pocket, but we could also identify a number of other mutations that occur even on the complete opposite side of the protein, so would not really have been predicted from the structure, but in fact, clearly affect the binding somehow of this drug for its target, and so we're really kind of excited about using this to map drug protein interactions more generally, and you can imagine using this type of mutagenesis for a lot of different applications, so you know, the Lu lab is doing these sort of therapeutic editing applications. We're envisioning more, you know, evolution of protein function, potentially mapping of drug target interactions. Another interesting thing we observed with this technology is you can actually cause mutations, not only in a coding region, but actually upstream of the transcription factor binding site. Sorry, of the transcription start site. And that's potentially interesting because previously, AID had been thought to require transcription to induce mutations, but what we could show is that, as far as 3KB upstream of the transcription start site, we could put a guide RNA in this mutation machinery and still observe mutations. These are here in yellow. So, we think that this might also open up the space of mutagenizing regulatory elements that are not actually transcribed to try to find, for example, promoter variance transcription factor, binding site variance that would affect binding activity. So, we're kind of excited about the different applications for this: generation of different antibodies, evolution of enzymes and aptamers, and the nice thing about this system is you really only need a guide RNA, but you can generate a diverse library of mutants, effectively doing a screen by tiling a protein coding region, and I didn't mention this, but of course, you can multi-plex this by just targeting multiple sites at the same time, so you can express two guide RNAs in a cell. You can co-evolve two sites that would potentially have some new function if mutagenized together. Okay, does anyone have any questions so far? Otherwise, I have a couple of quick points on kind of retesting hits I can go through. Yeah? (attendee speaking off microphone) So, not too much. With the AID system? Yeah, so you do see some basal level that's slightly elevated, but really, you know, if you look, like, it's probably not worth going all the way back to make this point, but you can see, in this case, for example, by and large, you really only get mutations where you put the guide RNA, so yes, there's some, like, infinitesimally elevated level if you just do, you know, whole genome sequencing, for example, but you know, at the level of substantially enriched mutations, it's really not detectable outside of where you put the mutations. Yeah. Okay, so, the last two minutes, I wanted to just describe a couple of other strategies for kind of following up on hits from screens, so one of the things that we found to be useful, you do a whole genome screen. You may not wanna clone 100 different guide RNAs against your 100 top hits to actually see which of these are actually real, and so, one of the things we've found to be useful is to generate kind of a batch library. Here's for shRNA, but the same thing has been useful for guide RNA libraries. If you do a whole genome screen, you can now generate a mini-library, for example, of your favorite, you know, 100 genes, and retest them at much higher coverage and get much higher quality data, and that allows you to really, more efficiently process through the hits to get effective ones to follow up on. And of course, you can do this genetic interaction map strategy. We also find these competitive growth assays to be quite useful, so you can infect a cell with a guide RNA targeting your favorite gene and a fluorescent marker, mix them with a wild type population, and then study the behavior of those populations in a mixture, and you can, you know, for example, if this gene causes drug resistance, you expect it to take over and become, the population should be 100% green if you start out from a 50/50 ratio. So, this is another strategy we've used to kind of follow up on hits from these screens. Of course, you can tag genes. I think you guys are probably gonna be talking about gene tagging and gene knock outs in some of these workshops, but this is a gene We had no idea what it did from the screen. You can use CRISPR to knock in GFP and actually localize this thing in the genome and see where it goes, and of course, this is greatly facilitated by these RNP studies, which I think you guys are gonna learn about in the afternoon, so of these neutrals really make tagging a gene with GFP much easier. So, that's pretty much it. I have some references, a lot of people to thank for the work here. A lot of the early work was done together with people in the Weissman Lab, including Luke, Max, and Martin, and people in my lab, who did the work that I showed you, Richard, David, Michael, Edwin, and Gaelen. Thank you. So, I'll take any other questions you guys have. (applause) (attendee speaking off microphone) - [Attendee] So you say that AID is capable of using point mutations only at, basically, (speaking off microphone) areas that are transcribed. Does that also apply to things like long non-coding RNAs? - So, we think we can see mutations, even upstream of a transcription start site. So, we haven't tried long non-coding RNAs yet with this thing. When we've tried to put it in a completely dead region in the genome, you know, transcriptionally off compact chromatin, we really don't see any activity, so there probably is some requirement for at least accessibility. We don't really know yet how much transcription is required. You know, even at the promoter that I showed, where we can put a guide 3KB upstream of a transcription start, still see mutations, I can't guarantee you that there's zero transcriptional activity at that spot. There may be some, you know, transcription going in the opposite orientation. I would expect, if the long non-coding RNA is transcribed, you'd have a decent change of mutagenizing it, but we haven't tried it yet. (inspirational music)
Info
Channel: Innovative Genomics Institute – IGI
Views: 2,166
Rating: 5 out of 5
Keywords: Mike Bassik, CRISPR
Id: L8XH4Nb8B7E
Channel Id: undefined
Length: 84min 19sec (5059 seconds)
Published: Sat Nov 04 2017
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.