Optogenetics: Illuminating the Path toward Causal Neuroscience

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments
Captions
Good afternoon. Good afternoon and welcome to the 2019 Warren Alpert Foundation Prize Symposium. Today, we are celebrating the achievements of four pioneers whose collective work has propelled us closer to understanding the ultimate enigma in human biology, the brain. The discoveries that we honor today span the fields of genetics, neuroscience, physiology, bioengineering, and beyond. And together, these discoveries helped give birth to the field of optogenetics, a revolutionary approach that allows us to visualize and modulate neurons with once unimaginable power and precision by simply exposing them to light. The work of Edward Boyden, Karl Deisseroth, Peter Hegemann, and Gero Miesenbock has not merely transformed our ability to see the inner workings of the brain, it has brought us closer than ever to elucidating some of the deepest secrets of the mind, secrets such as the neural circuits that are involved in decision-making and behavior, as well as those involved in the development of neurologic and psychiatric disorders. To be sure, the advent of optogenetics as a discipline has stemmed from the collective work of many scientists over some decades, but the four recipients we're honoring today made pivotal discoveries and developed critical tools that have rendered optogenetics an indispensable technique in neuroscience. Now, the idea of tinkering with the nervous system to illuminate its functions has actually tantalized scientists for centuries. Think of the now classic frog experiments conducted by the 18th century Italian physician Luigi Galvani, who demonstrated that information in nerve cells is in the form of electricity of the electrical impulse. That notion that light then could be used to manipulate the nervous system is, in a sense, a centuries-old extension of Galvani's ideas. In the 1980s, Peter Hegemann set out to understand how green algae and other simple organisms sense light. And in the 1990s, Peter and his team identified and characterized the light-activated molecules in algae that enable them to respond to light. Peter then, along with Karl Deisseroth, deciphered the key principles, structure, and function of light-sensitive proteins. In 2002, the notion of optogenetic neural manipulation became a tangible reality thanks to work conducted by Gero Miesenbock. Gero demonstrated that it was indeed possible to use light to modify neural activity. Gero used light-sensing proteins from the eyes of fruit flies and genetically incorporated them into nerve cells. And the achievement did not merely render neurons sensitive to light, but also offered a way to control their activity with light. Gero's work de facto showed it was possible to use optogenetics as a tool to study and manipulate the brain. Karl Deisseroth and his team subsequently conducted a series of key experiments showing the light-sensing rhodopsin proteins-- the proteins first studied by Hegemann in single-cell organisms-- could be used to activate neurons in the mammalian brain, which Karl and his team discovered had the necessary chemicals to make these proteins functional. Karl continued to work on optogenetics, elucidating the structures of several light-sensitive ion channels and discovering multiple new optogenetic activators. All the while, he has continued to use these tools to make fundamental discoveries about the inner workings of the mammalian brain. Edward Boyden worked on critical early experiments in optogenetics. He was part of a team, along with Karl Deisseroth, Feng Zhang, and others, that in 2005 published a key discovery showing that the light-gated ion channels of algae previously studied by Peter Hegemann could be used to control neuronal firing. Ed then in his own independent laboratory went on to refine optogenetics, developing optogenetic activators to allow independent control of multiple cell types in the brain and used optogenetics to achieve neuronal silencing. Collaborating with pioneers in holographic microscopy, Ed developed tools that used optogenetics as a way to impact exact patterns of electroactivities on small groups of neurons mimicking natural firings. Now taken together, these discoveries have fundamentally reshaped the landscape of modern neuroscience. They have set the stage for optogenetics-based therapies that could one day be used to restore vision loss, preserve movement following spinal cord injury, or modulate the circuitry that fuels anxiety and depression, and many other applications. Recognizing the people behind this kind of transformative science, science that carries the promise to reshape how we understand, diagnose, and treat disease, is the Warren Alpert Foundation's reason for being. We would not be here today celebrating these momentous achievements without the vision of the Warren Alpert Foundation and its founder. Many of you will have heard the history behind the birth of the Foundation, but this rather mythical story bears repeating. In 1987, Warren Alpert came across a news article that described the work of Sir Kenneth Murray, a British scientist who had developed a vaccine against hepatitis B. Somewhat impulsively, Warren put down the paper, picked up the phone, and cold-called Murray. Now fortunately, Murray answered the phone. And Warren Alpert announced to him that he had won the Warren Alpert Foundation prize. The little detail that was missing, there wasn't the Warren Alpert Foundation yet. So Warren got to work. He contacted then Dean of Harvard Medical School Daniel Tosteson and asked him to help convene a panel of experts that could choose future award winners. And Dan said, yes. And here we are 31 years later. Over those three decades, the Foundation has awarded nearly $5 million to 69 scientists, 10 of whom have gone on to receive Nobel prizes. I am thrilled today to have with us members of the Warren Alpert Foundation's board of directors. I saw Fred Schiffman and Gus Schiesser in the back. And I apologize if other directors have shown and I haven't been able to greet them. But on behalf of the scientific community at Harvard Medical School and the world over, I certainly want to thank the Foundation for its support of science and discovery and for its indefatigable efforts to alleviate human suffering. I want to congratulate the winners of this year's Warren Alpert prize and to express my deepest admiration for their transformational achievements. Now, I'm going to turn the podium over to our symposium moderator, who is himself one of the preeminent neuroscientists of our time, Bernardo Sabatini. Bernardo received his undergraduate degree in biomedical engineering from Harvard. He went on to earn a PhD in neurobiology, as well as the MD degree from Harvard Medical School. He pursued postdoctoral training at the Cold Spring Harbor Laboratory. Bernardo is a Howard Hughes Medical Institute investigator, a member of the National Academy of Sciences, and currently, the Alice and Rodman Moorhead III Professor of Neurobiology in the Blavatnik Institute at Harvard Medical School. Bernardo and his team seek to uncover the basic mechanisms that underlie brain plasticity, a critical feature that allows mammalian brains to acquire new behaviors, to learn, and to adapt to new cognitive challenges. And the ultimate goal of Bernardo and his team's work is to define the perturbations in these processes that can give rise to neurologic and neuropsychiatric disorders. Please join me in welcoming Bernardo. Bernardo. [APPLAUSE] Thank you, George, and thank you, everybody, for coming to this wonderful occasion in which we're going to celebrate the arc of discovery and invention that led to the field of optogenetics. And as Dean Daley just mentioned, the excitement in optogenetics for the neurobiology field is that it brought systems neuroscience into a modern era in which causal experiments suddenly became possible. And so from the times when scientists first put electrodes into the brains of animals, they found that there were neurons that reflected very specific features of the environment of the animal, of the state of the animal, or of the motor action of the animal. And these beautiful studies over many decades led to precise theories as to how the brain could perform computation, store information, or generate motor action. But the problem was that all of these theories were very beautiful. In most cases, we lacked tools to really test them with high precision. And optogenetics has given us the kinds of gain and loss of function experiments to control neural activity that allows us to test if these theories are right and if the activity of particular neurons is necessary and sufficient to explain parts of behavior and signaling within the brain. Now as Dean Daley mentioned, it's been decades that people have wanted to use light to control neural activity and the literature is littered with many failed attempts to do so. And there's a very nice essay from Francis Crick, the codiscoverer of the structure of DNA, that he gave the Royal Society in the early '90s, in which he talks about how one would like to be able to manipulate the activity of cells in the brain remotely. And he said, the ideal signal would be light. This seems rather far-fetched-- sorry-- this seems rather far-fetched, but it is conceivable that molecular biologists could engineer a particular cell type to be sensitive to light in this way. And so the four people that we're honoring today are the ones that took that idea that was deemed far-fetched, and made it a reality. And what I like about the story of optogenetics is that it brings together science that's done by many different people with many different styles. And so we're going to see examples of a biophysicist who wanted to solve a problem that he encountered in nature, pure curiosity-driven science to understand how unicellular organisms detect and respond to light. We'll see other examples from somebody who wanted to solve problems in his own laboratory. He wanted to drive his own research forward, had roadblocks, and invented technologies to get beyond them. And we'll also see examples of people that took an idea and almost on an industrial scale, created dozens and dozens of permutations of that idea to push forward a field and endowed with us the tools that we need for optogenetics. So very different kinds of science that all came together to create this. Now, I've told you a little bit about why I'm excited about optogenetics, but I want to spend just a couple of more minutes on why you should be excited about optogenetics and why the Alpert family should care. And George touched on these things. And there are really two reasons. One is that optogenetics has allowed basic discovery into the brain that's now giving us, we hope, new ways to treat neuropsychiatric disease. And so we've been able to identify cells in the brains of animals that makes them continue to eat, for example, even though there's no caloric drive, and hence might be relevant to obesity and diabetes. We found cells in the brain that mediate anxiety, that mediate feelings of anhedonia, other cells that exacerbate or can correct the symptoms of Parkinson's disease or that drive forward processes related to Alzheimer's. And all of this work has led to a new appreciation within pharma that in order to treat neuropsychiatric diseases, one might have to target circuits instead of pursue specific molecules. And I think that that's going to be the wave of therapy for neuropsychiatric disease in the future. Second, as George mentioned, in the future, it's quite likely that humans with optogenetic manipulations of their brain will walk around, and they will use these manipulations to correct perturbed patterns of activity within the brain. There are already trials going on in which optogenetic actuators are put in the eye to restore light sensitivity to the retina of individuals who have lost their own endogenous light-sensitive cells. So today, we're going to hear from four leaders that we've chosen within this field, who over decades in their own laboratories have driven this field forward. Some, it turns out, have done it somewhat accidentally by, as I said, studying natural processes, and then realizing the importance of what was here, continue to drive the field forward. And other ones have made it their mission to simply solve this problem. We have a jam-packed day, so we're going to try to keep on time. We're not going to do questions and answers, so you'll have to find the speakers at the breaks if you want to talk to them. I'm going to keep the introductions short. I'm not going to list the hundreds of awards that these four have collectively won because today, it only matters that they won the Warren Alpert prize. So our first speaker is Dr. Peter Hegemann. He is the Hertie Professor of Neuroscience at the Humboldt University in Berlin, a very storied institution that's made amazing contributions over the last century. And he has really worked on optogenetics, I think, his entire scientific career. His thesis work was entitled "Purification and Characterization of the Functional Chloride Pumps-- Halorhodopsin," one of the key molecules that we still use today to manipulate the activity of cells. His first independent group at the Max Planck Institute in Martinsried was called photoreceptors of microalgae. So basically, his entire life, he's worked on this problem of how unicellular organisms sense light, detect light, and react to light. And through his curiosity-driven research, he has found, and characterized, and eventually identified with collaborators the central proteins that became the first wave of optogenetic activators that made all this possible. Peter, I look forward to hearing the story from your side. Thank you. [APPLAUSE] So dear Dr. Daley, dear Dr. Sabatini, thank you very much for the kind words of introduction. And to stand here is something really particular and a great honor for me. And I'd like to express my deepest gratitude to the Warren Alpert Foundation, especially to the selection committee and all the members that are involved in it. So today, my short talk, the first 15 minutes, I'd like to guide you through the history. And I don't want to bother you with all the biophysics we do. In the second part, I'd like to show you some examples of what we are doing now. So the first slide which you see here is the Brandenburg Gate. And this is a sign for separation of the country. And on the other hand, it's also a symbol for reunification. And it was my pleasure to be here on the October 3rd because it's exactly the day when Germany became united 29 years ago. So my first conclusion is walls never help to solve any problem. And so my second statement is when you start a new research project, it should start with a wonder. So you should wonder about something you see which you cannot explain. It can be a natural phenomena or it can be a disease which is completely unknown or unexplored. So here, you see that wave-- the orange waves-- that occasionally occur in the ocean, sometimes also here close to Boston. Or if you go to Northern Territories, you see you red snow. And this is called watermelon snow. And the question is how is it caused? And the reason is green algae or algae of the ocean are responsible for that, like these algae. And this is Alexandrium. This is one of the most toxic organisms you can imagine. And if you keep them in your laboratory, you need a good aeration, otherwise you die. And this is Chlamydomonas nivalis that is responsible for the watermelon snow. And it's also the reason why this is called watermelons because it tastes sweet. So my laboratory is working on a relative of this Chlamydomonas nivalis. And I forgot to say the last sentence. The real pleasure in science is to work on something, the outcome of which is totally open. So we work on Chlamydomonas reinhardtii, which is the green model organism. And when we started to study the behavior, after some years, we realized that it's not new at all because there was a publication in 1866 by Andrei Famintsyn, a Russian scientist. And he described in this article, the behavior of Chlamydomonas. And this is written in German, published in a French journal, and edited by Saint Petersburg University in Russia. And he used an assay, which you see here to the bottom-right, this population of Chlamydomonas. And you shine light on one side, and they move away from the light. And he asked already, what are the conditions under which it moves to or away from light? This person became famous because he established a botanical institute in Moscow University later on. And he didn't have the time to work on algae unfortunately anymore. The other two people I'd like to mention is the physicist Ken Foster-- he was a postdoc in Mike [INAUDIBLE] lab-- and a chemist Koji Nakanishi because they worked on green algae many, many years later. And they worked on a wide species that were unable to produce carotenoids and chlorophyll. And they found that these are not phototactic. And they added vitamin A, and they realized that they could increase the sensitivity of these algae by three orders of magnitude within one minute after addition of this vitamin A. And that was the starting point when I got interested in the work. And I spent some time in Ken Foster's lab. And his lab was totally chaotic. You couldn't do anything, but he was inspiring on the other hand as well. So we studied this species for a while, and then realized it was strongly dependent on the ionic conditions as already Famintsyn said 100 years before. And we tried to establish electrophysiology. And we were heavily discouraged by the Chlamydomonas community, because they said it will never work, except one person, Ursula Goodenough. Where is she? She should be in the audience. Oh, there she is, Ursula, wonderful. It's great to have you here. And she was the only person in the community who encouraged me, and provided a cell wall deficient algae, which allowed Hartmann Hart in my group to establish electrophysiology. So he sucked the cell into the pipette, and applied a short flash. And what he noticed is that there is a fast photos at the current, which means ion influx into the eye, and a slow flagella current, which is influx into the flagella. And he measured the wavelengths dependent, and demonstrating that it is a rhodopsin spectrum with 500 nanometer maximum. And we concluded that these photocurrents are mediated by rhodopsin. We learned over the years to record directly from the-- oh yeah, thank you-- directly from the eye by using a pitched pipette, and this greatly improved the sensitivity and the time resolution. And allowed us to record with a better time resolution, as you see here, and to conclude that there's no delay. And from these and other measurements, we concluded that the rhodopsin the ion channels are directly coupled. And a few years later, we made another conclusion that they form a light-gated ion channel together. And the channel conducts proton and calcium. And it's interesting that we'd never talked about sodium, because this algae live in absence of sodium, so quite different from the neuroscience situation that you're facing now. And these are one million charges in one of these photocurrents. And we concluded that the conductance is about 100 femtoseconds which is very small, and almost exactly the number that we know in these days. So we also noticed that there is a low intensity range and a high intensity range, so they have sensitivity dynamic range of four log unit. And so we established this model. There is a light-activated ion channels responsible for the upper light area from 1 to 100%, and a lower region, where it needs some amplification. And this has not been completely solved. We also measured flagella beating together with the photocurrents, and have demonstrated that the appearance of this action potential, which is so important for the later measurements, that these cause the switch from forward to backward swimming, so the trigger for completely different behavior. Then in parallel, we worked on the purification of the photoreceptor by using Karen Foster's mutinant, reconstituting it with radioactive retinal, purified the most abundant retinal protein, have shown that these in the eye spot, and proposed that this is a light-activated ion channel. No response from the community at all. After five years later, we got a-- four years later, we got a mail from this gentleman. And he ask us, "I have been interested for some time in potential methods by which mammalian neurons might be transfected with genes whose product would permit light-triggered and depolarization of action potentials." That was a very interesting conversation that we studied. And probably most of you know that this is Roger Chen, who became famous by the GFP studies, and he failed. And the reason why he failed is we sent him the wrong gene. [LAUGHTER] So at least, we have shown it ourselves that it was the wrong gene. The abundant retinal protein in Chlamydomonas is not the photoreceptor we were looking for. So, but, the on-off clearly showing the way to go remains yours. So this is my tribute to Roger Chen, who died much too early. It was always wonderful. He has no social behavior. When you met him, he continued the discussion at the sentence where we stopped three or four years earlier. [LAUGHTER] So then Sunir Katarya joined my lab, and he discovered in a Catoosa library, the first cDNA library for Chlamydomonas, two new genes, and they were related to rhodopsin to some extent. And they showed a seven transform in helix domain, and a long cytomal end, which is about 40% of the protein. And we decided to express it in oocytes, but didn't have the oocyte method established in the lab. So we teamed up with Georg Nagel in Frankfurt, and expressed this protein in frog oocytes by using two voltage clamp experiments. And we have shown immediately, after three weeks or so, that these are the photocurrent that we flagged in the photocurrent of Chlamydomonas. So it was immediately clear that this was the protein we are looking for. These are the original experiments with longer illumination times, with channelrhodopsin one, which shows a very strong pH dependent, and a model at cation dependent. And we called this protein channelrhodopsin, because they unify an ion channel and a sensory unit. And there are two of them, channelrhodopsin one and two. So why does Chen work with channelrhodopsin one unsuccessfully? Because the currents were too small. And then he didn't find a person in his lab to work on channelrhodopsin two. So then after this experiment, we expressed this in human embryo kidney cells, the channelrhodopsin two, because the channelrhodopsin one showed small photocurrents only. And the next thing is, what we found is that the seven transform in helix fragment is enough to trigger this photocurrent, and the 60% of the protein are unnecessary. So we had a very compact, small system which is a sensor and the ion channel together, so channelrhodopsins. And the conclusion was that channelrhodopsin can be functionally expressed in animal cells. So this conclusion stimulated many, or a number, not many yet, five laboratories, basically, to work on this. And the first publication came out where Karl Deisseroth and Ed Boyden, they demonstrated that this works in hippocampal neurons, and you can fire trains of light, and the response that you get is a train of action potentials. The next person was Hiromu Yawo in Sendai, he demonstrated that it works on brain slices. And this might be forgotten. And the third person was Stefan Herlitze. He has shown that this functioned then in an animal. And the first animal was not the mouse. It was a chicken embryo. And the third person was Alexander Gottschalk. He produced a cell line which continuously allowed channelrhodopsin to manipulate neurons. And the last person is Zhuo-Hua Pan, who demonstrated it in blind mice, that it can reconstitute vision. And he was also pretty late, so he has not been in the focus so far. But he belongs to the key people. And then later on, as you know, Karl and Ed have taken over the field, and delivered all those modifications and other things. And certainly, meanwhile, it's also heavily expressed in zebrafish and drosophila in the mouth, and you'll hear more about it later. So the technology is relatively simple. You take DNA from a microorganism, Chlamydomonas, for example, connect it with a promoter region from the cell of interest, pack it into a virus, inject the virus into the brain, and then you wait for a couple of weeks, and then you can replace a needle for the light guide, and then you can study the behavior, among other things. So what you need is a photoreceptor that is small and genetically encodable, a promoter element, a chromophore which is present in sufficient amount. And this was my biggest surprise, that the brain contains retinal in sufficient concentration to reconstitute the opsin efficiently. And certainly, response you can interpret, and probably many responses that are essential for a living organism in the wilderness, is not so easily identified in a mouse in a cage. So the specificity is a major issue, that you can target a single neuron. And in parallel, you can target another neuron with another actuator or inhibitor, and then you can study learning and memory, and sleep, and locomotor activity, and feeding, and certainly, since recently, vision, hearing, sexuality, autism, addiction, anxiety, Parksinson, and so on. And Karl and Ed will speak about this. So what remained for us? So we went back to our original starting point, that we wanted to understand the photoreceptor. And this is the current knowledge we have about the channelrhodopsin. This is mainly based on mutagenesis studies and biophysics, and also on the exostructure that has been provided by Osamo Diwaki in Japan. And he used a hybrid which was originally designed by Hiromu Yawo in Sendai. And certainly MD calculation that tell us where the water is most likely in the channel. And you see here the retinal chromophore which provides the light sensitivity, and green amino acids that are responsible for color tuning, and the brown amino acids that are responsible for conductance and ion selectivity. And we mutated all of them, and know more or less what they are doing. But the key elements of the protein are the gates, the central gate and the inner gate, which are closed in darkness. And you should keep in mind that the image we have is a dark state of the channelrhodopsin, so it means that it's closed. And what we still need is certainly information about the open state, which is not available at the moment. Also I'd like to bring to your attention that in contrast to other proteins, the sensory photoreceptors are highly dynamic. So they undergo thousands of conformational changes after light absorption, and only a few of them are detectable as absorption changes, because they have relation to the chromophore. And this can be monitored by a changing of the absorption wavelengths. This is 470 of the dark state, and then 500, and then 390, and 520. And this is a main conducting state. And then it decays the conducting state in 10 milliseconds, and it reverts to the dark state only on a seconds timescale. But most interesting part is not highlighted here, which is the initial state. And by using pump probe experiments, we studied them in detail together with Johann Kennis. And if you excite the cells from the electronic dark state to an excited state, the conformational changes occur on an energy landscape to a minimum. And then there is a dissection, clinical intersection between the excited state and the dark state. And here the decision is made to return to the dark state, or to go into the photo cycle product. So this is a very central point for the efficiency of this rhodopsins. So decision is made on a picosecond timescale in a range of 10 to the minus 12 seconds after the flash. Everything else, what comes later, is a dark activity. So we certainly looked at the chromophore to understand the system, and also to manipulate it in a sense that we can use it. And that was done with Karl many years ago, and with Ofer Yizhar, his postdoc, who is now at the Weizmann, and will come tonight. And I gave you a few examples. So if you mutated this residue, you get larger currents. If you [INAUDIBLE] this residue, you will get a shorter open state lifetime, but more importantly, you remove the voltage sensitivity of the protein. And that allows you to fire action potential with higher speed, higher frequency, to study, for example, interneurons. And the third example is this [INAUDIBLE] and if you mutated it, you slow down the photo cycle tremendously. And it goes from 10 milliseconds, to 100 seconds or so. And this can be used for continuous depolarization. So here is one example. You apply blue light. You excite the cells. It fires action potential. And then you apply green light. Then it goes back two states. So this step function rhodopsins were very useful for future experiments. But also the current itself was very difficult in a different channelrhodopsin. For example, this one is inactivating, and it's an invert rectifier. And this is hyperactivating during a light flash. It's again invert rectifying. And here's another species which is not invert rectifying. And here is another species from somewhere near Hawaii, recently discovered. It completely inactivates in continuous light, for whatever biological reason, we don't know. So then these properties are clustered in different evolutionary branches. And also, what came out, that the color tuning is very, very important. So we can collect different organisms to address different cells. So one important experiment or one important question is shown here, the question of inactivation. And this is a typical biophysical question, because it requires a deeper insight. So if you look at the photo cycle again, I have shown you the basic photo cycle, which is shown here again. And we recently found that there are two conductances, an early and a late conductance. And the first one, the early one, is proton-selective, and the second one is sodium-selective. And depending on the equilibrium, you get a more proton- or more sodium-selective photocurrent in your neuroscience experiment. Alternatively, so more or less competing with a third is this isomerization you get an soon-anti or anti-soon isomerization, and it produces a second dark state, which started its own photo cycle. And the open state in this photo cycle is only weakly conductance, and is more sensitive for protons. And this is the reason why in steady state light you get this reduced steady state level. This one here. So the system is more complicated than probably, as an applicant, you might imagine. And the question is how can we manipulate this selectivity? And we identified some years ago two key residues, one in the inner gate and one in the central gate. And if you replace these glutamate, which is conserved in most of the channelrhodopsins, you can manipulate the ratio between proton conductance, which is shown in red, and sodium conductance, which is shown in green here. And then you can combine certainly these different mutations to divide the two further in one direction. But you can also look at the y type channel rhodopsin. This is Johannes [INAUDIBLE] done, and compare the proton and the sodium selectivity. And you might end up with a PsChR, which is almost exclusively sodium-selective, or you should look at the Chrimson and CsChR, which is almost exclusively proton-selective. So only at neutral pH and low sodium, you get a current here. And at alkaline pH, you get no-- you get no current. So what we engineer in the lab, nature has done billion years before. You only have to find it. This is a problem. If you don't know what to look for, then you don't get anything. And Ed's lab has, for example, identified this Chrimson, which is very interesting for many reasons. It is almost purely proton-selective, so you get, at neutral pH and low sodium, large current, and no current at alkaline pH. But if you mutate this, this glutamate, at this position, you convert it into a sodium-selective channelrhodopsin. And what we concluded from this is that this selectivity photo is in this Chrimson at a completely different position, so close to the surface, whereas the central gate is not important at all, and not existing in this variant. So nature had developed many different possibilities how to control conductance and selectivity. And if you look at the crystal structure, you see the reason. So the water pore in this Chrimson is blocked at this position. Whereas it's a free flowing water in channelrhodopsin from Chlamydomonas. So unfortunately, we have not been successful to produce a potassium-selective channelrhodopsin. And this is on the list for a long time. And Karl and I came together again to work on it. But as long as this is not finished, we made a compromise, and established two component optogenetics. And in this case, we combined a soluble photo-activated enzyme, which is a cyclase that produces cyclic AMP, and that allowed us to activate a cyclic AMP activated potassium channel, which can induce hyperpolarization. Alternatively, we recently worked on a rhodopsin cyclase, which is also a rhodopsin with an unusual tail, and this tail is an enzyme directly coupled to the rhodopsin, never found before. And this produces cyclic GMP, and we can use cyclic GMP-activated potassium channels to hyperpolarize the cell. And here's one example by [INAUDIBLE],, a postdoc in my group. She used this blue light-activated photoreceptor enzyme, and combined it with a small potassium channel from a bacterium. And she got nicely hyperpolarization which was very efficient. And due to this amplification, you can use it in very low light intensities because it drives 10,000 charges after one photon absorption, which is much more, certainly, than the pump, for example, that only transport a single charge, or an ion channel like Chlamydomonas which transport maybe between 10 and 20. So this is, I'd like to show you a few examples. Francisca Schneider, a former student from my lab, she now has a group working on cardio optogenetics. And she tried this, and was able to inhibit action potentials from the heart cells, and also to inhibit heart beating in her model systems. So it works nicely. And here's an experiment in parameter neurons. And you see here the marvelous expression. And here is a small blue light flash, and it causes a long hyperpolarization in these cells. And certainly would be better, and probably more comfortable, to use cyclic GMP instead of AMP. And therefore recently, [INAUDIBLE] she established in my lab the functionality of the rhodopsin cyclases, interestingly, identified by a theoretical physicist. She became unemployed and moved to biology, and she discovered with her friends these rhodopsin cyclases. And this is good for controlling cyclic AMP, and also for hyperpolarizing cell. So these are only a few examples. And I'd like to summarize, the major player is still the light-activated cation channel. What we are still missing is a potassium-selective channel. It has been complemented by anion channels and pumps, and also, since some time, light-activated enzymes that complement the ion transporters. So this is still the major player. I could continue, but in the favor of time, I'd like to finish. My conclusion is algae have taken over the brain research. And if we continue to destroy the climate, they probably will take over the planet, and control it. What they have done over 3 billion years. And I still have hope that will not happen, that the human species will also survive for some time. And this is my group. And I'd like to express my gratitude to, many thanks to all my co-workers I had the privilege to work with, to my photoreceptor and neuroscience friends and colleagues, and to the Chlamydomonas community where all the business have started. And certainly, I had collaborators over the last three decades, and a few I would like to mention. First of all, Karl, who worked with us for the last, probably, 12 years or so. And he transferred all the knowledge to the neuroscience committee, and he had enough and a profound molecular knowledge, so it was always a pleasure to work with him. Thanks, Karl, for that. And Ofer Yizhar, is former student, now a group leader at the Weizmann. He established optogenetics at the Weizmann Institute, and Georg Nagel, who worked with us in previous times, and certainly many spectroscopies and crystallographers. And I'd like to thank you for your attention. [APPLAUSE] Thank you. Close. OK. Thank you so much for that wonderful talk. So we have two halves to the presentation today, and in each one, we have two of our award winners speaking. Interspersed between those talks, we have two postdoctoral fellows from laboratories in the Department of Neurobiology at Harvard. And so these are talks by junior scientists who are using optogenetics in their own research to advance the inner workings, in this case, of the mouse brain for both of them. So the first of these talks is going to be from Kimberly Reinhold. She did her undergraduate work here at MIT, and then went to University of California in San Diego to do a PhD with Mossimos Kanziani, and then I was lucky enough to have her join my laboratory here on the quadrangle. And she's going to tell us about her work using optogenetics to both activate and suppress neurons in the brain to undercover how a mouse learns a new skill. Kim? [APPLAUSE] Thank you. It's a real pleasure to be able to share a vignette of how we apply optogenetics. When I was in college I was required to take a PE class, a physical education class. So I signed up for squash because I had never before attempted a racquet sport. And the first day I showed up and I tried to hit the ball with the racquet and I was very, very bad. The instructor sent me to a court by myself to practice. And I did, I practiced. And I attended all the classes that semester. And at the end of the semester I still couldn't hit the ball with the racquet, but other students seemed to learn. Squash is an example of how we learn through trial and error. We learn to associate sensory inputs, like the ball flying at my head, with appropriate motor outputs. Maybe swinging the racquet, or in my case, running away. And we learn these associations through practice and feedback. Trial and error learning is a fundamental component of many different cognitive processes. Therefore it's vital we understand where in the brain and how it occurs. What do we know? Well, we know that people with Parkinson's disease have damage to the basal ganglia, a set of nuclei deep in the brain outlined in green. And we know that these people are impaired in trial and error learning. These people can learn things like episodic facts, people's names, the time of day when something happened. But they have deficits, both in motor learning-- like the squash example-- and also in purely cognitive trial and error learning tasks. Interestingly, people with damage to a different part of the brain, the temporal lobe, are amnesics. So these people can't learn things like the color of the experimenter's clothes, but they can learn, practice-based tasks, trial and error tasks. And so we see a dissociation that suggests that the basal ganglia specifically support trial and error learning. And this has been confirmed in a number of model species. To figure out what goes wrong in disease, it's important we understand how trial and error learning works in a healthy brain. So today I'll tell you about our work to try to do this. To try to nail down more precisely where in the brain trial and error learning is computed, and how. First I'll explain our approach. We've developed a task in which mice learn through trial and error. There are two stages-- learning, and then execution of the learned behavior. I'll ask if basal ganglia are needed after the mice have learned. Whether basal ganglia are needed during learning. And if they are, how? To ask the question, are basal ganglia needed, we'd like a way to shut off these structures and look for effects on behavior. Diseases and stroke do this, they shut off brain areas. And lesion studies do the same. But those changes are rarely specific to a single brain circuit, and unfortunately irreversible. There are other kinds of changes we can impose in the brain. And these have higher spatial and temporal precision, but what we're really looking for is a technique with high spatial precision that also has excellent temporal precision. For example, able to probe really fast cognitive processes, like the learning update that happens between swings of the racquet happening on seconds. And here, optogenetics fills the gap. Let me first tell you about the task we use. Mice, like humans, have basal ganglia that receives sensory inputs and project to motor outputs. To study this area we've designed a task where mice learn through trial and error to associate a sensory component with a motor component. The sensory component, while we could have chosen an external stimulus, like a flash of light, which would activate the eye and a number of visually responsive areas throughout the brain-- many of which themselves project to the basal ganglia-- now we have a lot of active brain areas in different regions, and it becomes difficult to follow the flow of neural activity through the brain from start to motor output. So we decided to play a trick. We restrict the cue to be the activation of just one brain area, the visual cortex. This is well studied, accessible, and has a specific projection into the basal ganglia. We can use viral and genetic tools to express an optogenetic protein specifically in these neurons, with cell bodies in the cortex that send their axons into the basal ganglia. And then we can implant a fiber through the skull of the mouse, shine blue light through that fiber into the brain, and that will activate the optogenetic protein channelrhodopsin, which you heard a bit about. These cation channels. And that causes the cell to depolarize and fire action potentials. So we turn on neural activity in the specific population. The motor component of the task is a reach to a food pellet. We make the mice do this in the dark so they can't see the food. And we've ensured that they also can't smell or whisk the food, so they have to use the forearm and the forepaw to see if the food pellet is there. So the task proceeds like this. The mice reach a lot. Often there's no food there, they're just hoping. The optogenetic cue turns on, and then they reach and there is food. So they have to learn through trial and error that the optogenetic cue predicts food. Mice learn to do this. Here I'll show you a movie where pellets come into position. Let's see if I can find a pointer. Thank you. So you'll see that pellets come into position. And then the cue, the optogenetic cue turns on. So keep your eyes on this blue circle when the light flashes. That means we're stimulating neurons in his brain. So the cue's about to turn on. There, it turned on. And he's learned that that means the food is now available, so he reaches and grabs it and eats it. Remember, this is happening in the dark, so he can't see the pellet. We're spying on him with an infrared camera. The light here is just a flashing blue light that has nothing to do with the cue. And it's just to ensure that the mouse doesn't reach the flashing lights. And we have a variable interval between pellet presentations to make sure he's not counting time. So he's learned this task. And we can plot his behavior across many cue presentations or trials on the y-axis versus time in seconds on the x-axis. And then put his reaching onto this, and see that sometimes he successfully grabs the pellet, other times he drops it, sometimes he misses it, and often he reaches and there's no pellet there. But what's important is that when we add up all the reaching across all of these trials and plot that as a histogram, with reaches on the y-axis and time on the x-axis, we see this huge increase in the frequency of reaching right after the cue. And this tells us that he has learned the association between cue and food. We've taught a number of mice to do this. Here's the mean and standard error. And we think that these animals are paying attention to the optogenetic cue, because when the cue turns on, they reach. This is the same thing I showed you on the last slide. But when on a random set of trials we omit the cue, the mice don't reach, even though the pellet is there. And when we omit the pellet but the cue still turns on, the mice do still reach. So it seems that animals can learn this cue response association. Are the basal ganglia needed? The basal ganglia are a good candidate to link the cue to the motor output, because these structures receive diverse sensory inputs and project to motor outputs. And because there is a direct pathway from cue activated cortex into the input nucleus of the basal ganglia called the striatum. And I now want to focus on this input nucleus, the striatum, and in particular, the subregion that receives that input from the cortex, that receives that cues signal. This is the dorsomedial tail of the striatum. And for the rest of the talk when I say striatum, I'm referring to this part specifically. Does striatum link the cue to the motor output? Does it trigger the motor output? Well, if this simple linear model is correct, then we should be able to see a change in neural activity here around the time of the cue. So we test this by recording neural activity in the striatum. We can acutely implant recording electrodes into the animal's brain and record the activity of neurons. We see action potentials here, or spikes. And then we can just draw a line every time we saw a spike. And represent a neuron's activity in this way, where the different rows are different cue presentations, and the x-axis is time. So you see that this neuron has some activity around the time of the cue. And we've found a number of such neurons which seem to show changes at the time of the cue. So this might be triggering the reach. Second, if this linear model is correct, then when we shut off this striatum, the animals should no longer be able to do the cued reach. So we have an optogenetic approach to do this. There are output neurons in the striatum. But there is a second general class of neuron in this area, and these are locally projecting inhibitory neurons. So we can put a red activatable optogenetic protein into these cells, shine red light through two bilaterally implanted fibers, basically turn on these inhibitory neurons, and they act to shut off the neurons that project out the output neurons. And so essentially what we're doing here is performing a spatially and temporally precise loss of function, where we're shutting off specifically the part of striatum that gets the cue input, and then that sends output to the rest of the brain. And we want to know whether that output of striatum is needed for the cued reach. Importantly, previous work has shown that inactivating this part of the brain using drugs doesn't paralyze the animal's arm, so the animal can still move. What does red light do to neural activity? Let's take a look at this cell that we saw before. What we find is that turning on the red light prevents spiking in this neuron. And this is an inactivation lasting a second. But at the end of the red light, activity comes back. So unlike a lesion, this is reversible. We can suppress the activity of the cue responsive cells. You can see right here. And across all of the striatal projection neurons that we recorded, we see about an 86% reduction in activity. So the red light suppresses striatum. We can turn on the red light on a random set of trials at the time of the cue and ask if the mouse can still do the cued reach. So here's the cued reaching you've seen before. And now we want to know what happens when we eliminate striatum. Can the mouse still do that cued reach, or is cued reaching gone? What we see is that the mice are perfectly able to do the cued reaching. And in fact, there's no change in reaction time, and there's no change in the animal's ability to successfully grab the pellet and eat it. So we see no motor deficits at all. So it seems that striatum does not trigger the cued reach after learning, and there must be some other brain area that serves this function. We don't know yet. We have some ideas. It turns out these neurons that project to the striatum also have collaterals to the thalamus, the pons, and the superior colliculus. And we're investigating now whether one of those areas might be the link. But I began the talk by telling you that the basal ganglia are really important for trial and error tasks. So maybe the basal ganglia are needed during learning. In order to test this, we have to have a way to measure learning. The mice make cued reaches, but they also make spontaneous reaches before the cue even turns on, just hoping there is a pellet there. And we would expect that learning involves an increase in cued reaching, plotted here on the y-axis, and a decrease in non-cued reaching, plotted on the x-axis. And this is what we see. Here's an example learning trajectory from a single mouse. You can see it from the first day to the last day the animal increases cued reaching. And there is a little decrease in the non-specific reaching. So we can plot the direction of this learning change by putting on a vector from the first day to the last day. And then across mice we see that all animals learn by modifying behavior in this way. Interestingly, when we shut off the striatum for one second on every presentation of the cue, animals do not show the normal pattern of learning. Their behavior is disrupted. And here are the averages for those two groups. And so it seems the striatum is needed during learning. We can plot this data in a different way and combine the x and y-axes by asking, how much more cued reaching does the mouse do with respect to non-cued reaching. And this is a metric we call D prime. It doesn't matter, it's just a way to quantify learning. And I'm going to plot it on the y-axis, and then the day of training on the x. To give you a sense for what this means, low learning values-- learning index values-- mean no cued reaching. Middle values, the mouse is starting to do cued reaches. And at high values, the animal is performing really stereotyped cued reaches. So animals with the striatum intact learn this task. But when we inhibit striatum, mice don't learn. Maybe my striatum was asleep when I was trying to learn squash. Importantly, we haven't permanently damaged the mice in this red cohort, because we can perform a recovery experiment. So we stop the manipulation, and now these animals have the striatum intact, and they are able-- the same animals-- are able to learn. Maybe hope for my squash game yet. So it seems that striatum is needed to learn the cued reaching. Can we get any idea of how? One hypothesis is that the straight m might get information about the outcome of the animal's behavior in the past, and use that feedback to update the animal's behavior in the future. And that could happen at different time scales. This updating could be fast, as in the case of the squash racket swing. Where you swing, you're not very good, you swing again, you're a little better, but you're still not very good. And that update is very rapid on the timescale of seconds. Or we can imagine a student cramming for an exam. Learned a bunch of information, but that doesn't really get into memory until the little power nap begins. And maybe here the update is on a timescale of minutes to hours. So if we had a way to measure learning on a faster timescale, we could ask about the basal ganglia involvement on a faster timescale. I've showed you we can measure learning across days. If we had a way to similarly measure learning across trials within a day, then we could look at faster cognitive mechanisms. So now instead of looking at the probability of reaching, we're going to look at the animal's reaction time, which is the time delay to the first reach. And we're going to compare the reaction time early in the day to the reaction time 100 trials later. If the animal is improving, then his reaction time is speeding up. We're going to plot that improvement, that speed up, on the y-axis. On the x I'm going to show the change we expect if the animal is simply modifying the rate of reaching before that cue. So this is reaches before the cue, so it has nothing to do with cued reaching. And so we have a non-cued component and a cued component. We find that mice learn within a day, shown here. They increase cued reaching and decrease non-cued reaching. Does striatum store this accumulated learning within a day? If it does, then when we shut off the striatum at the end of the day-- which we can do with optogenetics very precisely, wait and then shut off striatum-- we expect that the animal's behavior will fall back to where it was at the beginning of the day. But if striatum does not store this learned accumulated performance change, then we expect that there will be little change in the cued reaching. And this is what we see. So it seems the striatum does not store the within day improvement. But if we disengage striatum early in the day and turn off the striatum on all trials, we see that the animals never seem to build up or accumulate that improved performance. And we can see incremental updates. When the mouse touches the pellet, he improves. But this incremental improvement is reduced when we shut off the striatum. So we favor the hypothesis that the striatum acts on a very fast timescale to update behavior. And so in conclusion, what I've showed you today is that activating a specific set of neurons in the brain is sufficient to teach mice a cued response. The striatum isn't needed after learning, but it is needed during learning. And we think it's needed specifically to provide these fast incremental updates. And so a picture emerges. This is just a cartoon, but we can imagine that the striatum is the arrow pushing the animal through behavior space. And that in healthy learning, these incremental updates add up to bring the animal to some optimal place. Maybe addiction is overcharged learning. Maybe there's a deficit in Parkinson's in that updating. Maybe Tourette's or OCD or wandering to the wrong part of behavior space, or the brain getting stuck in a local minimum. We think that perhaps the deficits that we see when we shut off striatum, while small on a fast time scale, could really add up over a long time to produce the really serious problems that we observe in basal ganglia dysfunction and disease. So I'd like to thank the people who contributed. Of course, all of the people who developed optogenetics, thank you. Bernardo Sabatini, my advisor, and Marci Iadarola, an excellent technician who's been working with me on this project. Thank you. [APPLAUSE] Thank you so much, Kim. So our next speaker is Dr. Edward Boyden from MIT. Ed is the Eva Tan professor of neurotechnology at MIT, and also a professor at the Media Labs, as well as an investigator of Howard Hughes Medical Institute. Ed did his undergraduate work at MIT, where apparently he studied quantum computing. That's what I learned from his CV. I was surprised about that. And then went to Stanford to do his PhD in neurobiology, working with Jennifer Raymond and Richard Chen, who is the brother of Roger Chen that we heard about earlier, somewhat ironically. While he was at Stanford he worked with Karl Deisseroth and Feng Zhang to first put channelrhodopsin into neurons and show, as we learned earlier, that one could control the activity of mammalian neurons with this tool. Afterwards he came to MIT. And I think his lab is really one of the broadest and yet consistently creative, in many areas, laboratories that I've ever seen. His group has worked, of course, on finding new and engineering new optogenetic actuators, and we've heard a little bit about that. His group has also developed robotic systems for doing electrophysiological analysis in the brain. He's worked on new amplifiers for electrophysiological analysis. And most recently, he's developed what almost seems like a comical approach, but is really incredible, which is that in order to look at tissue at higher resolution, he decided to not improve the microscope, but instead make the tissue bigger. And thus invented the field of expansion microscopy, which is really providing some remarkable insight into how complex tissues are organized. Ed, welcome. [APPLAUSE] Great. Well, first I'd like to express my gratitude to be here receiving this award with my good friends, collaborators, and colleagues. It's a tremendous honor. And I'm excited to get to talk today about optogenetics, but also how different technologies might fit together in this grand quest to understand the brain. The brain is so complicated that I think we need to think about an integrated toolset that lets us make maps of the brain, that lets us control the brain. And optogenetics, of course, is one of those key techniques. And then what you might call the opposite of optogenetics-- to watch the brain in action. And interestingly, tools in that nature are emerging from the optogenetic toolset itself. Why is it so hard? Well, the brain, amongst biological systems, exhibits extraordinary spatial complexity and temporal complexity. So if you think about it, you know, brain cells are enormous, right? They're centimeters in spatial extent. We have neurons a meter in length going down our spinal cord. And yet, if you care about the building blocks of neural computation, you care about axons and dendrites, synaptic connections, and biomolecules within brain cells. So how do you see across all those scales and control across all those scales of spatial extent? And there's also time. So of course, if you care about learning or memory or Alzheimer's disease or development or aging, these are your long term processes. They take hours to days to years. And the quantal building blocks of brain computations, though, are these millisecond time scale electrical pulses. So if we're thinking out ways to build tools to address Bring questions, you really have to think fundamentally about space and time. And so today, I'll tell you about sort of our thinking about these properties that yielded principles of how to discover and engineer tools focusing on optogenetics. But towards the end, I hope to also talk a little bit about how we're trying to develop strategies for imaging neural activity. And of course, if you can see and control neural activity, that's great. But it would be nice to have a map of the brain to know where in the brain to look or perturb neural activity. And my hope is that if we can stitch together these tool sets, comprehensive, emergent pictures of how brain circuits work might become more and more feasible. So I'll start with optogenetics. So you've already heard through the first couple talks the introduction of the idea. I first met Karl when I got to Stanford, and we started brainstorming about how would you control neural dynamics and sort of started going to the laws of physics. Could you use magnetic fields? Could you use light? And light, of course, would be great, as Francis Crick independently also outlined, because it's as fast as anything ever gets, and you can aim it at things. You have to bring light into the brain. And as much as people have brought electrodes into the brain for over a century, you can bring in optical fibers or other kinds of optical devices. This next question becomes, do you make a molecule that converts light to electricity, or do you find it? And as Peter already introduced, there's a family of microbial opsins the study of which goes back many decades which in single celled organisms will convert light into electrical signals. So the first of these to be characterized was actually a light-driven proton pump shown here in structural form. It's a seven transmembrane protein with an all trans retinal chromophore that absorbs the light. And you get these rapid conformational changes of what the discoverers named bacteriorhodopsin like a different proton pump. And that's found in halophilic archaea-- microbes that live in really salty water. Now a decade later, several groups found in the same species a light-driven chloride pump which they named halorhodopsin. And it shares some similarities but differs in certain specific key residues that makes it a chloride pump rather than a pump of positive charge. And then you've already heard about Peter and his colleagues who discovered the channelrhodopsin, these light-driven ion channels. Originally, the ones they found were cationic, and now we also know there are inhibitory ones that let negative charge through. So for me, one of the key interesting papers that got me interested in these opsins was from 1999 from [INAUDIBLE] colleagues. And at the time, these molecules have been characterized in these halophilic archaea. So they worked at very high salt concentrations. Here's electric current versus chloride. And you can see the peak is at a very high level of chloride. And if your brain is like mine, then the chloride is down around here. So this molecule wouldn't work very well. This is a halorhodopsin light-driven chloride pump from a specific species. But one of these molecules had, for whatever bizarre evolutionary reason, its peak function in the low chloride regime. And so this is actually one of the first molecules that Carl and I started collecting from colleagues. The first one that we tried out was, of course, the one that you've already heard about from Peter that was discovered by him and his colleagues-- the channel rhodopsin 2. And we put the gene into neurons, shined brief pulses of blue light just from a standard light source for seeing GFP at the time. And all of a sudden on the first try, what we found was that you could drive action potentials in cultured hippocampal neurons. And also, it didn't require the all-trans retinel, the chemical co-factor, to be added. For whatever strange reason, mammalian neurons made the chemical co-factor. So a lot of what we've been doing over the time since has been trying to figure out, well, what are the principles of finding these molecules and pushing their physical properties to the limits of speed and spectral sensitivity and all of the other parameters that we would like to achieve? And since this is a summary slide-- and I'll go through some of the examples of molecules in the following slides. But what we found is that members of all three of these classes-- the light-driven proton pumps, the light-driven chloride pumps, and light-driven ion channels-- can be found that actually are safe enough, effective enough, fast enough, and powerful enough that they can work in neurons which are, of course, a bit of a delicate environment with lots of complex physiology. So Brian [INAUDIBLE],, when they were working with me, looked at light-driven proton pumps and found that a member of this family from the archaerhodopsin class, if you genetically express it in neurons and then shine green or yellow light, will powerfully pump protons out of neurons and silence their activity quite powerfully. Halorhodopsins do inverse pumping but of a negative charge. So it has a similar physiological effect although different in a biophysical way. They pump chloride in response to green or yellow orange light and hyper-polarize the neuron, shutting it down. By deleting neural activity, you can look at the necessity of a set of neurons for a behavior or pathological state. And then the channelrhodopsin 2 we put into neurons back in 2004. And you can shine blue light on the neurons and let positive charge in. And you can activate them, letting you investigate what those neurons are sufficient to trigger. And I should mentioned that Amy Truong, when she was a grad student in our group, pushed halorhodopsins out into some of their different physical limits. And Nathan [INAUDIBLE],, when he was in our group, tried to and succeeded the same with channelrhodopsins. So light-driven proton pumps-- this is sort of surprising to us that this worked. We don't think of protons as very abundant in neurons or outside neurons. At neutral pH, they're orders of magnitude less concentrated than sodium and potassium and the other ions we often think about in neurophysiology. So to our surprise, we found this molecule, [INAUDIBLE] rhodopsin 3, when we put it into neurons, allowed us to make large photocurrents and to silence neural activity even in awake behaving mice. So this was really to knowledge the first near 100% nearly digital silencing of neural activity in awake behaving animals. And we find these molecules by either searching genomic databases or by sometimes doing our own genomic investigations. And there are different strategies that you can take. So one, of course, is if you find a molecule that you like, you can search locally in genomic space. Look at species related to the species that you found the molecule from and see if you can find improvements. So for example, there's an archaeal species for which the archaerhodopsin 3 came. And [INAUDIBLE] and Brian continued to search in species closely related. And the original molecule arch was powerful at silencing, but small [INAUDIBLE] ArchT was even more powerful. You can also search broadly in genomic space. So [INAUDIBLE] Brown had discovered that a species of fungus, actually, Leptosphaeria maculans, had a light-driven proton pump. And we got the gene, which we nicknamed Mac for short. And we found that it was also able to silence neural activity. And because Mac had a color shift-- this is the action spectrum, the current on the y-axis and color on the x-axis-- you could express Mac in a more redshifted opsin in two different neurons and then use two different colors of light to differentially affect them. So a Mac expressing neuron would be more silenced by blue light, and a molecule expressed as a more red light sensitive molecule will be more silenced by red light. That molecule that was more silenced by red light was actually that very same halorhodopsin that I mentioned earlier, which had been found in by [INAUDIBLE] and colleagues paper to be sensitive to light in a realm of salt concentration that was much lower than you might expect. And this molecule-- we published the first proof of concept neural silencing back in 2007. But it was a fairly weak molecule. The currents were not as impressive as we might have hoped. So we started thinking about the same genomic search properties and thought, well, could we find molecules that are light-driven chloride pumps that are much more powerful? And you could also find molecules that have a red shifted spectrum of activation. Now why would you want that? Well, when Amy [INAUDIBLE] was a PhD student working with me, we started thinking about the propagation of light in the brain. And of course, this is well known by many investigators long before us. But if you put blue, green, or yellow light into the brain there is absorption and scattering of the light. But if you go to red or even infrared light, there's less absorption. And that's one of the reasons why blood looks red, right? It doesn't absorb as much red light. And so the top are models and the bottom are actual measurements we made which suggested that redshifting molecules could be quite powerful. So as often begins, we started by looking through different genomes for candidate genes and stumbled across a class of molecules in this Crook's halorhodopsin class for the technical term, which did seem to have a redshifted spectrum with respect to the original. And then thanks to decades of structural work, both through point mutogenesis and crystallography, we made a couple point mutations that increased the current. And what we found was that if we expressed the gene for this molecule that we called Jaws because it came from the shark strain of [INAUDIBLE],, we could actually get light from a red laser, shine that through even an intact skull of an awake-behaving mouse, and shut down neurons many millimeters deep into the brain. Nathan [INAUDIBLE],, when he was a PhD student in the group, tried to do something similar but for the activators. And so he did a very large scale screen. He computationally looked at over 1,000 plant genomes in a project that's headed by [INAUDIBLE] Wong called the 1,000 plant project and identified over 60 new channelrhodopsins and expressed all of them looking for a function. As you can see from the red x's, a lot of them didn't work at all, but some of them did. And these are his screen currents in the red, the green, the blue. And he found exactly one of these after this enormous search that responded well to red light, which he named crimson. And crimson allows you to drive neural activity in response to red light. We made a point mutant crimson R that has better kinetics. And you can even use light getting into the infrared. Here's 735 centimeter light driving a neuron to spike in a slice of the mouse visual cortex. So of course, you can use crimson and red light to activate large volumes and deep into tissue, but it's also found use in other areas that at the time, I didn't even really think about. [INAUDIBLE] research campus really wanted a better optogenetic activator for drosophila-- fruit flies. And the problem with fruit flies is if you use blue, green, or yellow light, they have a startle response. They kind of flail their arms and sort of freak out, I guess. But if you use crimson and red light, then the effect is minimized. And he was then able to elicit, through crimson and red light activation, behavior in drosophila. And so now it's in very widespread use in the fly community, for example. Nathan in also this screen found molecules that were very fast. So one that he named Cronos, which is a channelrhodopsin with very fast kinetics-- and so it's found uses in parts of neuroscience where kinetics is important, like in the auditory system, or in the stimulation of axons that have high firing rates. And interestingly, these two molecules, crimson and Cronos, have also been very valuable because they can be used together. So if you look at the photocurrent in the y-axis and the color on the x-axis, you can see that-- of course here is crimson, where it has a peak out here in the orange, and you can drive it in the red. Here's Cronos in the blue circles and then the original channelrhodopsin 2 in the black triangles. But they're all recruitable, at least to some extent, by blue light. So we made the observation that if we use dim blue light to drive Cronos, so dim that it wouldn't really recruit crimson, and we could then use bright red light to drive crimson, we could differentially control the spiking of independent populations. So groups have now used this to look at multiple synaptic inputs to the same cell or how a neuromodulatory pathway might affect a given excitatory pathway in the brain because it gives you differential control over these pathways. So a lot of the work [INAUDIBLE] done has been in the searching through genomes for interesting properties like extreme shift of color or extreme kinetic performance. But what about getting to the ultimate levels of spatial resolution? Could you get single cell, single synaptic event, or single spike control over circuitry? So we started collaborating with an expert on holographic neural stimulation-- Valentina Miliani at the Institute of Vision in France. And this is work that [INAUDIBLE] and [INAUDIBLE] triply spearheaded. So she builds microscopes that look like this where you have a laser and you bounce the laser off of a spatial light modulator and basically project a hologram into the brain-- a three dimensional sculpture of light, if you will. And so we decided, well, what if we could try to build opsins that were optimized for this purpose? Importantly, we also want to make sure the opsins are located just at the cell body and not on all the axons and dendrites. This is an idea that several groups [INAUDIBLE] Frank [INAUDIBLE] and [INAUDIBLE] Bolton had tried taking the original channelrhodopsin to and fusing peptides to them to get [INAUDIBLE] just at the cell body. Now, why is that? Well, even if you holographically drive a cell, you're going to hit the axons and dendrites of neighbors that come by. And by fusing a peptide to it, you can localize it to the cell body. So we thought, well, we have all these new opsins. What if we do a double screen and look for new opsins that enable very powerful control when you activate just a cell body and also peptides that will enable them to be targeted there? It turns out that one of the molecules that Nathan had found in his screen, CRCHR-- which, if we had known it was going to be cool, we would have come up with a better name-- is a very powerful molecule, about an order of magnitude higher occurrence than the molecules [INAUDIBLE] channelrhodopsin 2. And so we thought if we located just the cell body, that might help make up for the lack of current because you're depriving the axons and dendrites of all those currents that would normally come to there. And then [INAUDIBLE] found a peptide that when infused to COCHR targets it just to the cell body for expression. Here's a sea of green in this slice of cortex. And here, you can see cells that are spaced with darker intervals between them. So why is that helpful? Well, if you record a cell and then scan your holographic laser about, when the opsins are everywhere, about a third of the time, Valentina's team observed straight activation. But if you located the molecules just at the cell body, then the effect went down essentially to zero. So in summary, a lot of this quest has been an [INAUDIBLE] of luck, right? The molecules essentially out of the box had speeds and amplitudes and profiles that made them appropriate for controlling neural activity. And in recent years, we've really tried to push the tool box out to their physical limits of performance-- maximizing amplitude, accelerating speed, shifting colors, and improving spatial precision. But it's interesting to think about, well, what about the opposite? Can we learn from this experience and do anything for the opposite goal of imaging neural activity? Can we get neurons to light up when they are active? So in this case, of course, the natural world has not made us so lucky. There is no molecule that all by itself will convert neural activity into light with the right speed and safety profile and efficacy. You know, what we got lucky with in optogenetics did not translate to the inverse problem of imaging. So naturally we started thinking, well, if the natural world won't evolve these things, why don't we build a robot that will do the evolution in the laboratory? And so when [INAUDIBLE] and Erica Yung were postdocs in my group, we decided to try to build basically a robotic scientist. Why can't we build a robot that would kind of do what we do when we're screening for optogenetic tools but in an automated way? So how do you do that? Well, suppose you have a bunch of genes. They can be obtained from the wild, or they could be mutants of an old gene that you want to evolve in some direction. Some of the mutants might be more interesting for your goal, and some might be less. And then we transfect these genes into cultured mammalian cells that each cell gets one copy, a different mutant. And then you can use an automated microscope to scan around and look for cells and therefore molecules that have things of the right speed and safety profile and efficacy and all the things that we want for indicators that, for optogenetics, the natural world provided. Then we bring in a robotic arm developed by our collaborator, [INAUDIBLE]. And we can then pull out the cells and therefore the genes that are interesting. Now it turns out that Adam Cohen's group here at Harvard had made the serendipitous discovery that the molecule archaerhodopsin 3 that I'd mentioned earlier that we had discovered was a very powerful neural silencer was actually a weakly fluorescent voltage indicator. And his team went on to make a mutant called quasar 2 that was brighter than archaerhodopsin 3 but still quite dim and not very well localized to the membrane. And nevertheless, it's been useful in cultured neurons for imaging voltage. So we thought, why don't we try to do a very large scale-- and this might be one of the largest direct evolution screens ever done in mammalian cells, anyway. We did almost 10 million mutants in two rounds of evolution. And let's screen for multiple parameters. We want the molecule to be bright and well-localized and safe and photo-stable. Why screen for multiple parameters? Well, if you're trying to take a molecule and mutate it and then screen for better mutants, if you push it in one direction-- say, screening for brightness-- you might devolve it away from the other properties that you seek. Evolution doesn't pull any punches. It's just trying to get the job done. And so here, you can see brightness on the y-axis and localization to the membrane, which is sort of a proxy of safety as well as function. And each circle is a different cell containing a different mutant. And indeed, you can see cells in there from molecules that are very well-localized but, you know what, not that bright. And then there are molecules that are much brighter, but, hey, they're no better localized than the parent. So with this, we did a couple rounds of directed evolution and found a molecule that we named archon, which is well-localized to the membrane. In the lower left, it has good kinetics, which it inherited from the parent quasar 2. And on the right, has good changes in fluorescence and signal to noise. And so we gave it to groups like Bernardo [INAUDIBLE] group, who did some measurements of synaptic events in brain slice. They would stimulate in one layer of the cortex and image synaptic events in a different layer of the cortex and just focus on the lower left. And black is what you see when you record with a [INAUDIBLE]---- sort of a ground truth, if you will. And the magenta are the unaveraged traces imaged on a microscope. So it turns out that this is a red fluorescent molecule. You can shine 630 nanometer light, which is sort of the color of a laser pointer, and it'll emit redder light around 660 nanometers and longer. More recently, [INAUDIBLE] and Seth [INAUDIBLE] from Michelle Hahn's group have been expressing this in awake behaving mammals and being able to image normal activity in multiple regions of the brain-- motor cortex, visual cortex, stratum. These traces look like [INAUDIBLE] traces, but they're being imaged on a microscope. In this case, you can see in upper left an epifluorescent or one photon microscope. And you can even see a population of neurons in an awake behaving animal here in the mouse hippocampus and look at the dynamics of these cells in a local network. Importantly, since this molecule is a red fluorescent molecule, you can use it in conjunction with blue light optogenetics and drive neurons while you image them. So you can imagine [INAUDIBLE] perturb in a closed loop way neural activity while looking at the voltage of the cells as well. So in summary, we want to do for imaging what the natural world had done for optogenetics. And turns out that an optogenetic tool can be mutated to become a pretty useful fluorescent indicator of neural activity-- in this case, the voltage of the membrane. But ideally, you'd be able to image a circuit and perturb it with also some knowledge of how these cells are connected to other cells-- upstream cells that bring in inputs, downstream cells that have outputs. How do you know what the network looks like? And so the last couple minutes, I want to talk about some newer work we've been doing, which is, I think, going to be very helpful in building a pipeline for generating new hypotheses to be tested with optogenetics. And this is a way we developed to make maps of brain circuitry. Now why is this hard? Well, many people are using electron microscopy to make maps of brain circuitry, some here at Harvard who have pioneered the field of kinetomics, looking at large scale electron microscopy maps of the brain. But it's very hard to see molecular information with electron microscopy. There's also super resolution microscopy. Store microscopy was invented here at Harvard. But it's difficult to scale this up to large 3D structures because of the physical properties of super resolution microscopy. So starting with two then grad students, Fay Chen and Paul Tilburg-- and now that half our group works on this, we decided, well, what if instead zooming in on the brain, we could physically blow it up? What if you could install a dense spider web-like mash of swallowable material like the stuff in baby diapers around and between all the biomolecules of a cell, soften the specimen by treating it with chemicals, add water, and could you blow up the brain and make it bigger? And so this owes a debt to a bunch of old lines of research. Actually, my MIT colleague Toshi Tanaka, who unfortunately passed away relatively young from a heart attack-- but in the early 1980s, he was studying the physics of these highly swallowed polymers. So in this cartoon, you see the white polymer mesh. You add water, and it's drawn in through osmosis. The polymer swells. And it's a highly charged polymer, importantly. So the physical growth can be enormous in a very short amount of time. And he published this beautiful paper studying the sort of phase transition [INAUDIBLE] physics as the polymer increases by 1,000 fold in volume in a matter of minutes. You also have to get the polymer in, and there's also a long history to that. People like Peter [INAUDIBLE] were using uncharged hydrogels like polyacrylimide and taking specimens and embedding them in these polyacrylamide hydrogels to improve their imaging. So if you could synthesize this dense spiderweb-like mesh but make it a charged polymer, one could try to take a brain cell like the one on the left and pull the building blocks of life apart from each other to make something more like the one on the right-- the constellation of biomolecules hovering in space, but with their relative organization preserved. So how do we do it? Well, we had to invent a couple of chemistries. In this cartoon, the proteins are shown in brown. And we had to invent handles that would bind to DNA, RNA, proteins-- and now we're working, even, on sugars and lipids-- and put little anchors or handles on all of them, so we can apply force to them and pull them apart. Then we have to make the polymer. And so we use free radical polymerization to synthesize the polymer hydrogel mesh, except we use these charged monomers, sodium acrylate, to form a poly acrylate mesh. And the spacing between the polymer chains is very tiny, around the size of a biomolecule or so. And when these chains encounter the handles or anchors, they form a bond. Finally, we soften the tissue by adding in detergents or heat or even enzymes to chop things up. And then we add water. The polymers will swell-- as Tanaka had beautifully worked out the physics long ago-- but this time, the biomolecules will come along for the ride. So we published the initial discovery that we could evenly expand the biological system back in 2015. And Panel B is a piece of mouse brain. The polymer is very, very dense. So spacing is, again, at biomolecular scale. After the process, this piece of tissue grows until it's like the one on the right-- about 100 times bigger in volume. Now, by design, we made the mesh so dense and so evenly synthesized that we wanted it to be an even expansion process. But this is biology. It's not enough just to design it. You have to prove it. And so we and many others have been doing very detailed control experiments where we take a pre-expansion image with the classical method of nano-imaging like Storm. And then we take a pulse expansion image after we blow up the specimen and compare them. And the distortion is not zero, but it's really small-- maybe around a couple percent over length scales of tens to hundreds of microns. So here on the left, you can see a piece of the mouse brain-- the cortex and hippocampus. This is a Thy1-YFP mouse that Guoping Feng, and Josh Sanes, and colleagues made many years ago. And we're going to zoom in from top to bottom. That white square, we blow up. And you can see two cell bodies and some purplish dots that are synapses that we antibody stained. And we're going to zoom in again. And the purplish dots get blurry, because you've hit the resolution limit of our confocal microscope. That's all before expansion, but after we expand, you can now see cleanly the pre- and post-synaptic sides of these neural connections. Blue is an antibody against the pre-synaptic protein bassoon. Magenta is representing the image of-- taken with an antibody against Homer1a, a post-synaptic protein. And the distance between these two protein densities is the same that Catherine Dulac and Xiaowei Zhuang measured many years ago with STORM microscopy. Except now you can use hardware that already most groups already have. [INAUDIBLE] Group, we worked with to try to apply some light sheet microscopes that their group had invented to expanded brain tissues. And the effect is that we now have a several order of magnitude speed up over equivalent resolution competing technologies. It's just a matter of engineering to make the microscopes go faster. And so this was work that [INAUDIBLE],, and [INAUDIBLE] triply spearheaded across our group and Eric Betzig's group. Imaging mitochondria and lysosomes at the top. And myelin at the bottom. We can look at synapses, and dendritic architectures, and exome architectures across the thickness of the [INAUDIBLE] cortex. And our hope is that we might be able to have a 50,000-fold speed up just by further engineering, hopefully, not too many months from now. So the beauty here is that you can really image at scale across extended neural circuits, but without losing sight of the nano-architecture of what's in a brain. So here is the same color code as before. We have synaptic proteins in blue and magenta. And we now have YFP in yellow. And now we're kind of at a meter scale, but we can zoom in and get very close to individual synapses. And this is kind of a long movie. And in the interest of time, I'm going to skip to the part here where we're going to start sort of zooming out and seeing more of the context. And you can then again zoom in and see the detail. This is a movie that we made of an entire fruit fly brain where the dopaminergic neurons are expressing a [INAUDIBLE] protein. And I just like it because it feels like a roller coaster as you fly through it. We're going to go right through the ellipsoidg body right there. And now we're going to go out into the more lateral sides of the fruit fly brain. And I hope you can see that we can see individual axons and dendrites, but we can also zoom out and see the entire brain as well. So why is it helpful? Well, you can really start to look at the wires of the brain. Here's another Harvard Technology Brainbow from Jeff Whitman, and Ross Haynes, and others where you express fluorophores in combinations in brain cells. So this blue cell got one fluorophore delivered by a virus. This green cell got a different one. This aqua one might have gotten one copy of each. And if you zoom in on two axons-- we're here in the mouse hippocampus-- you quickly hit the resolution limit of the microscope. And it's hard to see these axons. It's a blur, right? What's this green banana shape? But after you expand, you can cleanly resolve the individual axons of this bundle. And so we and others are now trying to design machine-learning techniques to automatically trace neural circuits that are color coded with a strategy which they call a Brainbow and also to use expansion to give the resolution at scale. So to summarize, we discovered that you can physically magnify biological systems. And this technique has really started to become popular with people making discoveries published every week in a wide variety of species. Not just brain cells-- Giardia parasites. In the lower left, E. coli. In the upper right, planaria and kidney specimens. And the list goes on, and on, and on. So our last slide is-- what I really would like to see is if we can build these into a pipeline. Suppose with expansion mapping, you can make comprehensive maps of brain circuits. And then you can go in and observe the neural activity using fluorescent indicators of voltage and other signals that neurons perform. And then go in with optogenetics and use that to do a causal test of what a pattern means. Can we assemble this into a pipeline that could, who knows, maybe even yield computational models of how neural circuits work or how they go wrong in dysfunction? So I think of the knowledge along the way, all the people who have led specific projects, but I'll put up this slide, which I don't have time to go through. I will just acknowledge those people within the group at the top in our alumni who helped with these projects and an even longer list of people in the middle who collaborated with us to make this a reality. It's really an omnidisciplinary arena, neuroscience, nowadays. So I hope you can use these techniques in your group. We have a big culture of teaching. And feel free to email me if you have any questions. [APPLAUSE] Thank you very much, Ed. So now we have about a 20-minute break. And we'll come back with the second session. [UNINTELLIGIBLE CHATTER] [SIDE CONVERSATION] OK, if everybody can take their seats, please, we're ready to start again. Welcome back for the second half of our Alpert Prize Symposium. We're going to follow the same format with the two awardees bracketing talk by opposed talk. So our first talk is by Gero Miesenbock. Gero's the Waynflete Professor of Physiology at the University of Oxford and the director of the Center for Neural Circuits and Behavior. He's from Austria. And he did his medical degree at Innsbruck in which he studied really classical physiology and then took a remarkable reductionist turn and came to the United States to work at Yale with Jim Rothman on the mechanisms of vesicle secretory pathways in cells. And what I like very much about Gero is that when he faces a problem in biology-- and I think he's very much motivated by biological problems. When he faces a problem in biology, he creates a tool to solve it. So when he was in Jim's lab, he invented what I think is really the first GFP-based sensor of a cellular process-- or at least one of the first-- which was the SynaptopHluorin, in which he exploited the pH sensitivity and modified the pH sensitivity of GFP in order to create a protein whose fluorescence changed from the acidic environment to the extracellular environment in the secretory chain. This has been very important as a tool that's still used today to monitor, among other things, the release of neurotransmitter from neurons. When he set up his own laboratory, he chose something in between the secretory pathway and animal physiology-- or rather, mammalian physiology-- and studied the Drosophila nervous system and tried to understand how the brain of that smaller animal controls its behavior. Again, faced with problems, he invented several forms of optogenetics. One was to reconstitute, as we heard before, the entire visual transduction pathway of the fly in neurons to render those cells sensitive to light. It was a great demonstration. I don't think you really used it for biological discovery. So then he went on and invented yet a second approach, which was to exploit ion channels that were not present in the fruit fly for which he could design light-activated ligands. And he's used that to a great extent to make fundamental discoveries about the relationship between activity and behavior in the fruit fly. And I have great admiration for his work. I think he's going to tell us today about some biology including some wonderful work on the basis of sleep drive in fruit flies. Thank you. [APPLAUSE] Thank you very much, Bernardo, for this very, very kind introduction. It's obviously an enormous honor and a huge pleasure to be here. In fact, the honor and the pleasure is so large that I decided to share them with my doppelganger. This is Dr. Gero. [LAUGHTER] We have more in common than just a first name. He is also a scientist. And a mad one in this Japanese comic called Dragon Balls. He strives for world domination just like I do. And if you look carefully, you can see that his skull has been replaced with a transparent Plexiglas dome, of course, so that the function of specific genetically-targeted neural circuits in his brain can be controlled with light. And that's what today is all about. Now, what motivated the invention of optogenetics some 20 years ago was the idea that a technology like this would open three experimental doors for neuroscience that had previously been locked. The first of these doors was the ability to pinpoint the neuronal causes of behavior with much greater precision than what had been practical previously. And this idea, really, reflects my scientific upbringing as Jim Rothman's postdoc where the mantra that I was exposed to on a daily basis was reconstitution, reconstitution, reconstitution. In other words, if you are a biochemist, you want to understand how a biochemical process works. What you do is you purify the responsible actor as you put them back together. And you reconstitute the biologically processed from these pure components. So when I started my own lab, I thought, what would be the equivalent of a biochemistry constitution for a neuroscientist? And, of course, that equivalent is to metaphorically purify the electrical activity patterns that underpin our mental lives, play them back into a nervous system. And if you, in this way, can reconstitute perception, action, emotion, thought, then you have a credible claim that you really understand how these mental events are actually based in the physics of the nervous system. The second experimental toy that I thought optogenetics would unlock was the probing for neuronal connections, which is classically done in a painstaking way in paired electrode searches. And in more modern approaches, equally painstakingly, through large scale reconstructions of neural circuits. One alternative approach, of course, would be to replace one of the stimulating electrodes with a light beam that can be rastered across tissue. And then just listening with one electrode whenever the light beam hits a connected partner, and in this way, unravels synaptic connectivity. And the third experimental door, of course, is the test of mechanistic ideas. If you have a conjecture about how a system works, then, of course, the only way to figure out whether that conjecture is right or wrong is to interfere in a targeted fashion in the process. So for much of the rest of my talk today, I will relate some of our recent work on a biological problem in which optogenetics has indeed unlocked all three of these experimental doors for us. And that problem is the biological function and neuronal control of sleep. Sleep is one of the great biological mysteries. Each night, we disconnect ourselves from the world for seven or eight hours-- a state that leaves us vulnerable and unproductive. And yet, despite these risks and costs, we still have no clue as to what sleep is good for. We are trying to get at the biological role of sleep by understanding its neuronal regulation based on the premise that somehow the brain's sleep control systems must respond to molecular changes that are intimately linked to the core function of sleep. It's widely thought that there are two of these control systems in our brains that are symbolized on this classical diagram by two different forms of oscillation. The sine wave represents the well-understood circadian clock which oscillates in synchrony with predictable external changes that are caused by Earth's rotation. As such, it's a purely adaptive mechanism that makes sure we do our sleeping when it suits our lifestyles best. But understanding the clock is unlikely to speak to the deeper mystery of why we need to sleep in the first place. The solution to that mystery, we believe, will come from understanding the second control system-- the sawtooth oscillation that's superimposed on the circadian clock. And that sawtooth oscillation represents the sleep homeostat. The homeostat measures something that happens in our brains or our bodies while we're awake. That something accumulates or depletes-- logically, it's the same-- during waking. And when a certain threshold is reached, we go to sleep. The process resets itself while we're asleep, and then the cycle begins anew when we wake up on the next morning. We know a lot about the circadian clock. And this is really the Rosetta Stone that broke that problem open, the discovery by Seymour Benzer and his grad student Ron Konopka almost 50 years ago of fruit flies whose circadian clocks ran abnormally fast or slow. And from that discovery, then followed, through the work of many laboratories over the past five decades, a pretty complete molecular, cellular, and systems understanding of circadian timekeeping. This slide, in contrast, summarizes pretty much everything we know about-- [LAUGHTER] --the molecular basis of sleep homeostasis. And it's an overstatement, but maybe not too gross an overstatement. And my goal for the rest of the next 20 minutes or so will be to draw at least a few outlines on that blank canvas. $$$ Conceptually, we know how the homeostats must operate. It's a relaxation oscillator, a bi-stable system that switches between a fill and discharge mode where waking corresponds to fill mode where something called sleep pressure builds. Until a tipping point is reached, the system flips into discharge mode. And then the accumulated sleep pressure is dissipated. Now, at the end of my talk, I hope to propose a molecular interpretation of what sleep pressure is, when the brain it accumulates, and what the process are that underlie this bi-stability, this switching between a fill and a discharge mode. The story begins, like Seymour Benzer's and Ron Konopka's, in fruit flies with the discovery by a former postdoc in the lab, Jeff Donnelly, when he was actually a graduate student with Paul Shaw at Washington University in St. Louis of neurons in the brains of fruit flies that exert a powerful influence over sleep and waking. There's about two dozen of these neurons. They're labeled here by a promote enhance element called R23E10. So whenever you see that string of symbols, you know that a genetic manipulation is targeted selectively to these two dozen or so out of the 100,000 cells that make up the fruit fly's brain. The neurons project to this inverted V structure in the midline. This is one particular layer of the fan-shaped body of the central complex. Why such a small number of neurons can exert such a powerful influence over the probably most dramatic global states transitions we undergo on a daily basis is another mystery, but a topic for a different talk. Now, together, Jeff and I, discovered that these neurons represents the output arm of the sleep homeostat. The neurons themselves, I should say, they've originally identified in an activation screen where the brains of flies were randomly peppered with actuated molecules. So this is one example of an optogenetic genetic or, in his case, thermal genetic application where the neuronal substrates of behavior can be pinpointed in an almost classical forward genetic screen. So the way we do typically our experiments is that we [INAUDIBLE] fix a fly. We let it walk on a spherical treadmill, a little Styrofoam ball, whose rotations we read out with an optical computer mouse. And since there are no documented cases of somnambulism in flies, we know that whenever the ball is spinning, the fly must be awake. What you can't see is that the head capsule is actually open. And we've inserted a patch electrode into one of these 24 sleep control cells and expressed an optogenetic actuator in the entire population of neurons. So we can control the electrical activity of these neurons optically, and at the same, have one recording electrode as a measure of one member of that population. And this is now an experiment lasting for half an hour. You'll see that the fly starts out awake. It's moving along happily. The ball is spinning. These are the rates tick marks. And the sleep control neuron is completely silent. At about three or four minutes, we switch on the lights. You can see that the neuron whose activity we are recording begins to emit electrical impulses. And all movement virtually instantaneously stops. At about 19 minutes or so, we switched the lights off again. The sleep control neuron falls silent. And movement quickly resumes. So we have isolated a switch in the brain of the animal that allows us to toggle it into and out of sleep on command. Now, during many of such recordings, we found that when we targeted one of these sleep-inducing cells with our patch electrodes, these neurons were typically found in one of two states. In one state, shown here on the left, where the neuron behaves like you would expect a well-behaved neuron to act, you see that it responds to injections of depolarizing currents with action potentials whose frequency grows in a graded fashion with the amplitude of the injected current. The neuron on the right, in contrast, does not initially looked like a neuron at all. You can see we can still depolarized this cell to positive membrane voltages and still not squeeze a single electrical impulse out of that neuron. It's not just the active membrane properties still that have changed. It's also the passive membrane properties. If you compare the size of the voltage steps that are elicited by standard-sized current injections, you can see that the voltage deflections on the left are very large, suggesting that this neuron opposes the injected current with a large resistance, whereas the voltage deflections on the right are much, much smaller. Also, the neuron takes much less time on the right to settle into a new equilibrium membrane potential after a current step, whereas on the left, it takes quite much longer. So this combination of a short membrane time constant and the low-input resistance on the right is almost diagnostic of the opening of a current leak, or a current shunt. And I'll show you in a few minutes what the molecular basis of that current shunt is. Now, when we saw this, we, of course, immediately thought, well, maybe this is the mechanism of homeostatic sleep control. Maybe these neurons naturally switch between electrically active and silent state, depending on whether the fly is asleep or awake. And our sampling of flies, whose sleep histories have been manipulated, confirmed this prediction anecdotally. But of course, in order to really nail this point, to demonstrate that a neuron is capable of transitioning between these two states as a function of its sleep history, one would like to be able to control that translation directly. And in order to do that, one would need to know a signal that normally acts on these neurons, and actuates the switch. Now what might such a signal be? Well, a clue to the identity of that signal had come in the first experiments in which the behavior of an animal was controlled optogenetically. These experiments were done by my then-graduate student, Susana Lima, at Yale in 2004. And what Susana had done what she had expressed light-gated ion channels in all dopaminergic neurons in the brains of flies and then recorded their movement trajectories for two minutes before and after switching on dopaminergic activity in these animals. Here, you see examples of these movement trajectories in a circular arena of four animals before and after activation of the dopaminergic system, and you clearly see that dopamine has a highly arousing effect on flies, as it, of course, does on mammals. Most psychostimulants-- cocaine, amphetamine-- of course, acts by inhibiting reuptake of dopamine at the synapse, and thereby elevates synaptic dopamine levels. So one potential signal that should act on these sleep-inducing cells is an arousal-inducing dopaminergic projection. And there is, in fact, a class of dopaminergic neurons that extends their processes exactly into the same brain region that's also inhabitants by the sleep-inducing neurons. In fact, the two neurons shadow each other so closely that the question naturally arises whether they are synaptically connected. And optogenetic gives us the tool to probe for these connections by recording, with a patch electrode from one of the sleep-control neurons, while manipulating the activity of the putative presynaptic partner, the dopaminergic neurons. So this is now such an experiment where we start with our sleep-inducing neurons in the electrically active on state, and then we switch on the dopaminergic projections that innervate these neurons optogenetically. We can even predict what the effect of such an arousing dopaminergic signal on these neurons should be. It should, of course, silence them and switch them off. And this should be the mechanism that underlies awakening. And this is exactly what happens. You can see that after dopamine delivery, the neuron falls silent, the action potentials disappear. Also, the passive membrane properties change, the input resistance drops, the membrane time constant shortens, and importantly, if we hold the recording long enough, which we can do in some cases, we can see that these changes are completely reversible after an extended time frame. So this suspension of electrical activity is temporary. It's part of the normal duty cycle of the neurons' activity, and not an artifact that's brought on by our experimental manipulations. If we use the passive membrane properties' input resistance and time constant as a measure of the kinetics of these changes, we can see that the switch happens rapidly with the time constant of about one minute, which is, of course, way too fast to be accounted for by the production of new ion channels, but must instead, involve the modulation of the existing channel repertoire of these neurons. And we can also demonstrate that the action of dopamine on the sleep-control neurons is direct because we have discovered the dopamine receptor in these neurons that mediates the effect. And if we remove that receptor selectively from these neurons using RNEI, the neurons become resistant to the dopaminergic signal, the flies become unable to wake up. And literally, doze away their existences, spending 23 and 1/2 hours a day asleep. Now, the ability to control this excitability switch of the sleep-control neurons also gave us the means to dissect the underlying biophysical mechanism. In the interest of time, I will only summarize the results. What we discovered is that there are two potassium channels that get modulated antagonistically between the electrically active on state, which corresponds to the sleep state, and the electrically silent off state, which corresponds to the awake state. There's the classical voltage-gated Kv1-channel shaker, which gets upregulated during electrical activity and a [INAUDIBLE] potassium channel that we've discovered and termed "Sandman" that gets translocated into the membrane of sleep-inducing neurons when dopamine switches the electrical activity off. And it's the potassium current through this leak channel that underpins the shunt that you've seen in the electrophysiology. So that's responsible for the short-membrane time constant and the low-input resistance of these neurons. Now, knowing the biophysical basis of this transition between sleep and waking, then allows us to reframe the relatively wake biologically question, what is the biological purpose of sleep, into a mechanistically well-defined problem. We can ask, what signal or process switches the sleep-inducing dorsal fan-shaped body neurons on. And in fact, we can make our question even more mechanistically precise because we know the crucial role that is played by these two potassium channels. Any sleep-inducing signal that's sensed by these neurons must ultimately act by upregulating the shaker current and by driving the internalization of the Sandman channel that acts as a deterrent on the electrical output of these neurons. I'll focus, for the rest of today, on our understanding of the regulation of shaker, of which there has been more progress recently than in the Sandman control. Now shaker, like many voltage-gated potassium channels, is a beautiful structure composed of two different types of subunits. There is a pore-performing alpha subunit shown here in gray, to which is appended on the cytoplasmic side a beta subunit shown here in blue. Now if you zoom in closer on the beta subunit, you see that it actually has a small molecule cofactor bound shown here in red, which is the nicotinamide NADPH. This structure solved by Rod MacKinnon that revealed this enzymatic nature was not unexpected because when the first of these potassium-channel beta subunits were cloned some 25 years ago, the sequences suggested that they are actually enzymes, specifically oxidal reductases. And that then raised the question, are these molecules voltage-controlled enzymes, or are they redox-controlled ion channels? I will present you evidence that they are certainly the latter and that their ability to sense changes in cellular-redox chemistry is an integral component of the regulation of sleep, and perhaps even causally tied to the biological function of sleep. I will also argue that it may actually be the interplay between the pore of the channel and the active site of the enzyme that's the fundamental accounting principle that underlies sleep homeostasis. It was also noted quite early on that even if these molecules, these potassium-channel beta subunits, clearly look like aldo-keto reductases, they are terrible enzymes. They have very, very low turnover numbers. And one of the structural reasons is evident in this structure here. If you look carefully, you can see that the binding cleft, in which the NADPH sits, is almost closed in a latch-like fashion by a tryptophan residue that locks the cofactor in place. And it's this obstacle to cofactor exchange, that slows down the turnover of the enzyme. We think that this is an absolutely essential feature for the ability of these neurons to monitor changes in sleep pressure. Now in fruit flies, the Kv beta subunit is a protein called "hyperkinetic," which Chiara Cirelli and Giulio Tononi discovered more than 10 years ago. Causes insomnia, then mutates. It's just like mutations in the alpha-subunit shaker also lead to insomniac flies. But here, we've reproduce these experiments showing that homicide as hyperkinetic mutant flies are indeed insomniacs, and that we can rescue these insomnia of the hyperkinetic mutants by restoring wild-type Rod protein function just in these 24 sleep-regulatory neurons. So this points to these cells as the sleep-relevant site the action of the protein. Now surprisingly, if you use a rescue construct that carries a single-point mutation that allows normal expression, folding, association of the beta subunit with the channel, but abolishes its catalytic activity, the rescue no longer works. So the insomniac flies remain sleepless. This suggested to us that hyperkinetic sleep-regulatory role must be tied to its ability to bind this cofactor NADPH and sense changes in cellular-redox state. From that inference, then followed two predictions. The first one is that changes in redox chemistry are expected to accompany changes in sleep pressure, and second, that if we could somehow perturb the redox chemistry of these sleep-control neurons, that should have consequences for sleep. From this inference, and these two predictions, and our knowledge of intermediary metabolism, then also follows the conclusion that the dFB neurons probably monitor these redox processes as a gauge of energy metabolism because this is, of course, where redox chemistry is ultimately determined, specifically in the way electrons that are food-derived are handled in the electron-transport chain of mitochondria. So when we stumbled into this particular area of research, we certainly needed a little refresh in mitochondrial electron transport. And I suspect that something similar may be true for you. So here's a very simple refresher of mitochondrial electron transport. We have three proton-pumping complexes in the inner mitochondrial membrane-- one, three, and four. One accepts food-derived electrons, mostly from the Krebs cycle, but also, from the oxidation of fatty acids in the form of NADH. And these electrons are then handed off in a very carefully controlled fashion due to the explosive nature of the use of oxygen of combustion from one complex to the other using two mobile carriers-- ubiquinone, or Q, between complexes one and three and cytochrome c between complexes three and four. The proton gradient that's built up across the inner mitochondrial membrane is then, of course, used by the proton-powered turbine, the ATP synthase, which you see on the right, which phosphorylates ADP to ATP. So what you see here is a condition where the ATP demand is high. There is a high level of ADP, and there's a sufficient supply of NADH fuel. So demand and supply are in balance. But when that is not the case-- so when you have an overabundance of NADH, but ATP reserves that are full, and the proton motive force that is large-- then the ATP synthase slows down. Electrons still get stuffed into the transport chain at complex one, but they have nowhere to go. They accumulate, mostly in the ubiquinone pool, and start to transfer directly to molecular oxygen, and produce the oxygen-free radical, superoxide. So we would predict that these sleep-inducing neurons during the state of waking, when, as you remember, Sandman is inserted into the membrane, shunts their electrical activity-- so prevents them from producing energetically costly action potentials. But the animal being awake has just had its breakfast or its lunch and therefore, has ample caloric reserves that lead to exactly these conditions. That electrons are fed into the mitochondrial transport chain, but somehow, there is little demand for ATP synthesis, and so that should render these neurons particularly prone to mitochondrial oxidative stress. To test this idea, we filled the mitochondria of the sleep-inducing neurons with a protein called "MitoTimerm" which is the derivative of the green fluorescent protein, whose chroma four converts irreversibly from green to red as it's oxidized. So this is sort of an integrative indicator of mitochondrial oxidative burn. They express this protein in these 24 sleep-inducing neurons, and then imaged as [INAUDIBLE] dendritic fields. What you see here are two photon stacks through the dendtritic fields of twelfth [INAUDIBLE] flies, whose sleep histories differed. And you can see that in flies that have been kept forcefully awake-- that's just the top row-- there's a clear redshift of the MitoTimer fluorescence, suggesting that these sleep-deprived animals indeed suffer a larger degree of oxidative stress than well-rested flies at the bottom. Now what we also noted is when we measured the basal sleep of flies that expressed this reporter protein in the mitochondria, just of these 2,000 sleep-controlled neurons, is that there was a small, but significant, observer-effect present, namely flies that had MitoTimer in their mitochondria lost a small but significant average of about two hours of daily sleep. And we suggest that this reflects the fact that as MitoTimer is oxidized, it actually acts as a buffer for oxygen-free radicals. And it's the consumption of these oxygen-free radicals that's reflected in a reduction in sleep. To test this notion more clearly, we looked for better tools to do this. And probably, the best there is a plant-derived molecule. So this is another theme, I guess, today, that many of the best tools come from unexpected parts of the kingdom of life. So many plants have bifurcated mitochondrial electron-transport chains with a second terminal oxidase. Our terminal oxidase is complex IV. And plants have an alternative oxidation called AOX that taps directly into the ubiquinone pool and acts as an overflow valve when there are too many electrons accumulating in that pool. So it's not an uncouplet. It doesn't interfere with energy metabolism. It simply takes electrons that are surplus and detoxifies them by transferring them to molecular oxygen and producing water. So when we introduce this particular molecule into the inner mitochondrial membrane of these 24 neurons, you see that the sleep loss was dramatic, almost eight hours per day. So capping mitochondrial reactive oxygen-species production at the source indeed, seemed to ease the pressure to sleep. Now in animals, the typical antioxidant defenses are two enzymes-- superoxide dismutases. We manipulated both. I showed you the results with just one. Superoxide-- this mutates one. The cytoplasmic form, which exists in an antioxidant form, which has the predicted effect, namely a reduction in sleep-- but there is also a point mutation that turns the antioxidant into a prooxidant, and introducing this particular variant has the opposite effect, namely it increases sleep. But the increase in sleep is blocked if we remove either the potassium channel beta subunit hyperkinetic or the alpha subunit shaker from these neurons. So Altogether we think these behavioral and imaging results suggest that the potassium channel beta subunit indeed couples mitochondrial electron transport to sleep. Now, you probably ask yourselves, how can it be that am extremely short-lived agent, such as superoxide or hydrogen peroxide, lives a very short lifetime because it's so highly reactive, can serve as a signal that is conveyed from the mitochondria electron transport chain all the way to a potassium-channel beta subunit that's suspended from the plasma membrane. What we think is that we're probably missing a crucial biochemical link in this signaling chain. And we also have a hypothesis as to what this particular intermediate might be. We think it's lipid prooxidation products that are derived from mitochondrial-membrane lipids. These lipids are some of the most susceptible targets of oxygen-free radicals. And importantly, they produce compounds such as the 4-oxo-2-nonenal that you see, which are established potassium-channel beta-subunit substrates. The beta subunits are oxidal reductases, or specifically, aldo-keto reductases. So they take carbonyl compounds and reduce them to the alcohol. And the oxo-2-nonenal is obviously an aldehyde, so it has the correct chemistry to be reduced to the alcohol. And that reduction would then be coupled to the oxidation of NADPH to NADP+. There are several additional pieces of evidence that suggests that this is a likely candidate. One is that Rod MacKinnon, in the discussion of his structure, notices that the active site of the beta subunit is unusually hydrophobic. And he also notices that there is an ill-defined electron density in the active site. To me, this suggests that it's a lipid-- or a lipid mixture that's bound to the crystal. And, of course, since fatty acids are heterogeneous, the breakdown products that will be produced by them through pure oxidation. It will also be heterogeneous and not produce a clear diffraction pattern in the crystals. So the idea, then, is that the molecular signal that conveys rising sea pressure in these neurons is the progressive oxidation of the cofactor at the potassium channel beta subunit from NADPH to NADP+, and that that, somehow, is linked to the induction of sleep. So we, of course, looked for ways to test this idea causally and found an optogenetic tool-- or adapted an optogenetic tool-- that would allow us to do precisely that, and through a pulse of light, flip the redox state of the cofactor that's bound to potassium channel beta subunit. The tool that we used was developed by the late Roger Chen, who's also been mentioned several times already today, as a genetically encoded contrast agent for electron microscopy. The tool is called miniSOG, or small SuperOxide Generator. It's an engineered flavor protein which we have, in this case, anchored with a lipid modification in the leaflet of the plasma membrane. In close proximity to the potassium channel. And then, upon illumination, we expect this tool to oxidize the NADPH cofactor, either directly or via a locally produced lipid peroxidation in the media. And that, of course, then, if our idea is correct, should put flies to sleep. And as you can see in these experiments, it indeed did so. The crucial column to look for is the one in the center. In all cases, we measure sleep in individual flies-- each row is one fly-- for 30 minutes after an initial 9-minute exposure to blue light. And you can see that, compared to their parental controls, flies that have miniSOG go to sleep in much greater proportion and for longer than the parental controls. Once again, the effect is blocked by the removal of hyperkinetic. That's the fourth column from the left. But it's not blocked by the removal of an innocuous potassium channel, KD4 [INAUDIBLE].. Now the ability to set the redox chemistry of the cofactor directly throughout the genetic tool then also opens the door through-- to biophysical studies of what actually happens to the excitability of these neurons as we flip the state of the cofactor. So we're able to patch onto one of these sleep-inducing neurons, and then, again, after 9 minutes of illumination, measure either, in current clamp, its spiking behavior, or in voltage clamp, the characteristics of the voltage-gated potassium currents. This is an example of a neuron you can see that, clearly, after illumination, the spike rate increases. The input-output function steepens. The inter-spike interval contracts. So in other words, the neuron becomes much more vigorously electrically active. And the biophysical change that underpins all of this in the voltage-gated A-type potassium current is a lengthening of the inactivation time constant. So the potassium channel starts to inactivate more slowly with an oxidized cofactor than with a reduced cofactor. Now when Chuck Stevens and John Connor defined the A-type current in 1971, they included a modeling study in which they proposed that the A-type current is the main determinant of the inter-spike interval of persistently active neurons. And conceptually-- or intuitively-- the way to link inactivation kinetics to firing rate is in the following way-- you need a powerful A-type current to restore the membrane potential to its resting level after a spike. If your neuron is persistently active with each spike, you will push a certain fraction of your potassium channels into the inactivated state, and therefore make them unavailable during the next repolarization event. By slowing the inactivation, you keep a larger fraction of your channel population in the conducting, active state. And that allows you faster repolarization, and therefore higher spike rates, and in this particular physiological context, deeper or longer sleep. If we express just GFP and not miniSOG, you can see that light has no effect. But the changes that we saw upon illumination within cell experiments were also reflected in between cell recordings where we just compared the properties of neurons-- this is the bottom row now-- that either express the catalytically active or the catalytically dead rescue transgenes. So you can see that the catalytically active rescue transgenes, again, cause high spike rates and slowly inactivating A-type currents. And the same is also true for manipulations of the cells' ability to either prevent the production of reactive oxygen species or to induce them with the presence of this pro-oxidant version of SOT1. So this suggests that there may, in fact, be a direct mechanistic connection between rate of living and sleep, which is not entirely unexpected given the epidemiological evidence. Many things that cause oxidative stress have been implicated in aging and degenerative disease. And of course, chronic sleep deprivation has also been implicated as a cause of shortened lifespan. So possibly, this is the mechanism that might link these two important phenomena. So we've reached a stage where I can return to this conceptual animation and try to replace it, for you, with a molecular, mechanistic animation. On the next animation, that will be quite a bit going on, but I'll talk you, slowly, through it. So the crucial regulator that determines whether this sleep-inducing neuron is in fill or discharge mode, and the animal awake or asleep, is the sandman channel, shown here in yellow, which can be either in the plasma membrane or in intracellular vesicles. We know that dopamine drives the internalization. And we are feverishly working to find the signal that causes the endocytosis of the sandman channel. So when sandman is in the plasma membrane, spiking is blocked. And the cofactor of the potassium channel beta subunit population gets progressively oxidized to NADP+ as a reflection of the operation rate of the mitochondrial electron transport chain. Now I mentioned to you, before, that these beta subunits are probably the lousiest enzymes known to man. And of course, that's exactly the property you would desire if you were to construct a system like that. Because what you need is a biochemical memory that holds on to each oxidation event, and out of multiple of these events, then constructs an analog measure of the accumulated sleep pressure. If the enzyme was catalytically active, each oxidation would be fleeting and ephemeral. And your accumulated sleep pressure would disappear. Now through this ability of the beta subunit to communicate with the inactivation gate of the channel and to regulate the inactivation time constant, the same process also automatically determines the commensurate corrective action. Because it is the fraction of the hyperkinetic pool that's been oxidized that determines the kinetics of the A-type current, and therefore the spike rate of the neuron. Now one particularly important aspect of a system like this is, of course, that the accumulated sleep pressure somehow has to be dissipated when the animal actually goes to sleep. And what we think a particularly beautiful way of accomplishing this is by coupling the enzymatic activity of the beta subunit to the voltage-driven rearrangements of the alpha subunit. So as you see on the animation here, when sandman moves out of the membrane, the neuron becomes electrically active. The voltage sensors start to move. These conformational changes, we think, get transmitted to the beta subunit. And suddenly, an escape path for the oxidized cofactor opens up. NADP+ gets kicked out, gets replaced with NADPH. And the animal wakes up refreshed, with its cofactor pool replenished. Before I finish, I'd like to return, for just a minute, to the very Stone Age of optogenetics. When we started to work on the first optogenetic actuators, I became aware, through a citation alert, to the synapto-pHluorin paper that Ben Heidel mentioned, of a quotation that Ben Heidel also already mentioned in his introduction. And that showed to me that I was not the only scientist who had seen the need of technologies like this. And obviously, Francis Crick, in an essay entitled "The impact of molecular biology on neuroscience," which was published in the millennial issue of The Philosophical Transactions of the Royal Society, wrote, as Ben Heidel already said, "The next requirement is to be able to turn the firing of one or more types of neuron on or off in a rapid manner in the behaving animal. The ideal signal would be light. This seems rather far-fetched, but it is conceivable that molecular biologists could engineer a particular cell type to be sensitive to light in this way." So when we had the first experiments that turned that far-fetched possibility into a reality-- and these are these experiments. These are hippocampal neurons grown in culture, transfected with an opsin protein taken from the fly eye. Because the correct gene for general dobson had not yet been identified. And GFP-- so we then patch onto one of these transfected neurons. You can see that, in the dark, it sits around resting value. But as soon as we turn on the lights, there's a depolarizing step. And the neuron responds with a volley of action potentials. So when we had these experiments, I sent Crick a pre-print of our paper. And if you've read this wonderful book called The Eighth Day of Creation, which recounts the early days of molecular biology, you know that Crick was a prolific letter writer who steered the development of many fields through a vast correspondence. And his stylistic hallmarks as a correspondent were twofold. He was always encouraging, and he was also constructively critical. And that's exactly what I got. So he wrote back to me and said that he read the paper I had sent him with great interest and was excited to see that the system already works, at least to some extent. However, he realized, as I did, that it still needed improvement and that this was-- and that this would take further work. Unfortunately, Crick didn't live to hear how not just our experiments progressed, but those of many others. But I think it's fair to say that the improvements have come, thanks, in large measure, to my co-laureates and that, as a result of all these efforts, the way the neuroscientists go about our business has fundamentally changed. With that, thank you very much. And thank you too. [APPLAUSE] To the members of my group, I'd just like to mention a few of the key individuals. So the current crew is aligned to the left. And some of the notable alumni have been displaced by one tab stop to the right. Boris Zemelman was the postdoc who made the first optogenetic actuator, Susana Lima, the graduate student who used them in the behaving animal. And the recent work on sleep was done by, initially, two postdocs, Jeff Donlea and Diogo Pimentel, and the more recent work on the redox control of sleep by a postdoc, Anissa Kempf, and a grad student, Michael Song. Thanks. [APPLAUSE] Thanks, Carol, for that wonderful talk. So our next talk is from another postdoc. It's by Dr. Charlotte Arlt. She is originally from Germany, did her undergraduate work at the University of Cologne, and then went from there to do her PhD with Michael Houser at University College London, and then came to Harvard and joined the laboratory of Chris Harvey in the department of neurobiology. And she's going to tell us about how she uses light to read out and manipulate the activity of neurons in an effort to understand decision-making processeses in the brains of a mouse. [APPLAUSE] Thank you very much for the opportunity to share our recent work. It's truly an honor. And I'm excited to be able to share our work at this occasion. Coming back to the theme of racket sports, we make decisions in our everyday life all the time. If you think of being a tennis player like Roger Federer, for example, which I do on a daily basis, he has to make up his mind of hitting the ball to the left or to the right over and over again. But the process by which he arrives at a decision might very much depend on the context in which he makes it. So in this case, imagine you're Roger Federer, and you are in a training situation. Your coach is across the net. And the coach instructs you to always hit the ball to the exact same spot. So in this case, the mapping from sensation to action is very straightforward to you, guided by key instructions. But now imagine you're Roger Federer again, and you're playing in the Wimbledon final against Rafael Nadal. The ball is coming towards you in the exact same way as in training. And you might hit it to the exact same spot. But now what's guiding your decision is a complex model of your environment, including a model of your opponent and statistics of a match like this. So this combination of sensory input with some internal knowledge or experience to guide action is what we think of as cognition. And in the Harvey Lab, we would love to understand how this process is implemented in the nervous system. We understand this is a very ambitious question, so we try to tackle it by asking two concrete questions here in this talk. Firstly, we want to know what brain circuits actually mediate such seemingly simple decisions as hitting a ball to the left or to the right. And once we've identified these circuits, we can then ask, how does context actually affect the implementation of such a decision-making process in these very circuits? And what I want to tell you about today is how we search for such circuits mediating simple decisions. And to our surprise, we found that the identity and the involved brain areas mediating decisions would change depending on the context and the experience of the animal. So for the remainder of the talk, I'd like to tell you how we arrived at this conclusion. Wanting to study brain circuits for decision-making, we don't have tennis players. But we train mice to make decisions by running to the left or to the right in mazes. We think this is quite a naturalistic behavior for animals to display, because that's what they need to do in the wild to survive as well. And once we train mice to make decisions in that manner, we can then inhibit and silence different parts of their brain and ask, what brain areas are they actually using to guide their decisions? And you can imagine that experimental setups to manipulate different brain regions in the same mouse might be quite heavy and difficult for an animal to carry around. So instead of having the mouse run freely in the environment, we actually move the world around the mouse and keep the mouse stationary. So here you see an animal in such a virtual reality setup. It's running while being head-fixed. And the Styrofoam ball that it's running on, the movement of this ball is translated into movement of the virtual world that we project on a screen surrounding the animal's field of view. So animals are using the visual feedback that they get from this world to update their running patterns. We can use the system, now, to train mice to turn either left or right in a very simple Y-shaped maze. And we train them to use visual associations of cues that they see on the maze walls with rewarded turn directions. So in the top example, you see the mouse is seeing vertical cues on the side of the maze. And in this case, it has to run left at the end of the maze to get a reward once it gets there. In the bottom case, it sees the opposite trial type of vertical cue. And then it has to run to the right. So let's see what this actually looks like in action in a trained mouse. The animal in the left case encounters this horizontal bar, successfully runs to the left, where it's supposed to run, gets some visual feedback of the correctness of its choice. And then a little drop of milk is dispensed to reward it. And a few trials later, it encounters the opposite trial type, those vertical bars, chooses to run to the right, which is the correct decision in this case, and again, gets rewarded at the end. Now that we have animals making decisions in this virtual reality setup, we can expose their brains by removing the skin from the skull, and thereby gain optical access to the brain surface right underneath. We're using mice that express channelrhodopsin, what you've heard about a lot so far now, specifically in GABAergic neurons in neocortex. So these neurons are the small interneurons here, depicted in green, that inhibit the parameters population. And the parameter neurons are normally the ones that carry information from the local circuit out to different brain regions. So when we now shine light of the appropriate wavelength onto the skull, we can activate the interneurons through the skull. And those interneurons, in turn, inhibit the local population, thereby silencing a given volume of the brain. And we can do this with very high [INAUDIBLE] precision, again, as you've heard in the previous talks. Here the light is on, indicated by the blue bar. For just a few hundred milliseconds, the interneuron that you see in the top row there responds reliably and strongly. And the parameter neuron simultaneously recorded underneath is pausing and spiking at the same time. And at this point, I'd like to thank the pioneers of optogenetics, whom we're honoring here today. Because none of the experiments that I'm about to describe would have been possible without their contributions. And even after running these experiments for quite some time now, it's still astonishing to be able to remote control brain activity in a living mouse making decisions. What areas do we actually want to inhibit? We focused on a few candidate areas for decision-making, one of them being the parietal cortex. The parietal cortex gets sensory input for many different modalities, and in turn, projects to different areas implicated in action selection. So it's really at the intersection of sensation and action, and has been implicated in decision-making across species. Another region we're focusing on is the retrosplenial cortex, given that we're using a navigation-based decision-making task. Because the retrosplenial cortex is at the interface between subcortical systems for navigation, such as the hippocampus and the entorhinal cortex, and other cortical regions, including the parietal cortex. We can now couple our blue laser light to a system of mirrors whose position we can change very quickly to steer the laser beam onto different target locations on the skull. So in one trial, for example, we may be inhibiting parietal cortex on both hemispheres here while the animal is running down the maze and choosing to turn left or right. And then in the next trial, we can quickly change the mirror positions and steer the laser beam onto a different region-- here, retrosplenial cortex. We choose the order of these target locations randomly. And we also interleaf them with controlled trials where we steer the laser beam outside of cortex. We have another control trial in some outer sensory cortex where we inhibit a local volume that supposedly isn't involved in making these types of visual association decisions. So given that we have a system now where animals are making decisions in this virtual reality, and we can inhibit different brain areas, we can finally ask, what area is actually necessary for making this simple type of navigation decision? And we do this by subselecting trials where the laser beam was in a particular location and then quantifying the average performance of the animal in those trials. So we quantify performance as fraction correct where 1, or 100%, means the animal is making no mistakes. And 0.5, or 50%, means the animal would perform at chance level, either just guessing randomly or continuously running to the same side. But you see here, in the control case, the performance is very high, far away from chance level. The animal is making very few mistakes, meaning, one, that it knows this task very way, but two, also that it's not distracted by blue light in general. When we now inhibit some outer sensory cortex, we see a very similar picture, indicating that the animal is not relying on spiking activity in that area to guide its navigation decision. When we inhibit retrosplenial cortex, we see a very different picture now where, every time, on average, when we inhibit this brain area, we are causing the animal to make mistakes in this type of decision-making task, indicating that it's actually relying on activity in this area to guide its decisions. And now quite surprisingly, when we inhibit parietal cortex, the animal can still perform the task very well, indicating that activity in this area seems to be dispensable in this setting, and the animal can do quite well without it. So having identified, now, the retrosplenial cortex as an area that's mediating these simple types of decisions, we can now go ahead and modify the context in which the simple decision is made. So we create a flexible context now where, on top of the two associations that I've showed you before, sometimes the animal has to make the opposite choice given the same visual cue. In a given experimental session, we introduce these two pairs of associations in blocks of tens of trials. And once an animal has been trained in this setting for quite some time, it can perform quite well, mainly making mistakes, really, at the change points of these blocks. So after a switch of the association block, the animal makes a few mistakes, because it's still using the old association. Then it realizes it doesn't get rewarded this way, and updates its strategy, and uses the new pair of associations. So towards the end of each of these blocks, the animals are performing at a very high percent correct. And their decisions, outwardly, look very similar to the ones that animals made in the simple context. And let me just show you how similar those decisions look. In the left case, you see an animal that was trained in the simple context. It encounters this horizontal cue and has to turn to the left. In the right, you see an animal that was trained in the flexible context, encounters the same trial type, also has to turn to the left. And when you just look at these movies side by side, they really look identical, indicating that the types of decisions, outwardly, these animals make are very similar. So now as a sanity check, we first wanted to see if, again, retrosplenial is actually mediating decisions in the rightward case. So we inhibited the same targets as I showed you before, but now specifically, towards the end of these blocks, where the animal is performing at very high fraction correct, meaning it knows the association well. So again, in the control setting, or with some outer sensory cortex inhibition, the animal is performing the task very well. Again, with retrosplenial cortex inhibition, we induce many errors. But now you see the drop is quite large, actually close to 50%, meaning the animal is almost performing at chance level. So in this flexible context, it seems to especially rely on activity in this part of the brain to guide its decisions. But now, quite surprisingly, when we inhibit parietal cortex, we also see a very large drop in performance. And again, we didn't see this type drop in performance in the simple context. So in the flexible context now, it's not just retrosplenial cortex guiding the animal's decisions, but the animal is also relying on this additional brain area, the parietal cortex, which, just to remind you, in the simple context, for the very same outward decision, it didn't need. And given that we were quite stunned by this result, we wanted to make sure that we understand what's going on. It seems like the current cognitive context dictates whether this parietal area is also necessary for decision-making. Now if that's true, we should be able to take the exact same mouse, change the cognitive context it's experiencing, and thereby change the number of brain areas that are involved in the decision. So to test this, as a sanity check, we took an animal that was trained in the flexible context. And again, to remind you, with parietal cortex inhibition, we see a very large performance drop. So the animal is using this part of the brain. And then we transition this animal to the simple context for 14 days, where it's not experiencing any association switches. In the control case, of course, the performance stays very high. But now, stunningly, when we inhibit parietal cortex, we continue to see the strong effect on performance, suggesting that the brain persists in using this decision-making area for this very simple decision even weeks after we switched the animal from the flexible to the simple context. And we thought this was quite stunning, especially, again, when we compare the lack of effect in the animals that were trained in the simple context here. Because again, animals trained in the simple context who have never seen the flexible context do not rely on this brain area, the parietal cortex, at all to guide their decisions. So in addition to the current context dictating what brain areas are used to make decisions, context can also have a very long-lasting impact on what brain areas are used to guide the same simple decision. Finally, we wondered whether, maybe, we see this very large effect of context. Because perhaps we're looking at a very extreme case here. Or maybe it's some special case where we're comparing this simple context to a flexible context. It could be that, perhaps, the flexible context is especially demanding for the nervous system, where the same cue has to be mapped to opposite choices. So we created another context where the animal doesn't have to reverse its choices, but we just create a more diverse context. So we keep the two associations that you've seen before. And then we add two more associations with different visual cues. But now, importantly, the mapping from the visual cue to the choice of the animal that's rewarded is constant. So finally, we ask what brain areas the animal relies on in this type of decision-making. And again, we see not only retrosplenial cortex, but also parietal cortex being used by the animal to guide its decisions. So it seems that context, in general, has a very strong impact on what brain areas are used for decision-making. And context can be changed in various ways. You can increase diversity of context, or you can make the context more flexible. And probably, there are many different ways on top of these two variations we've shown you that would change the number of brain regions involved in decision-making. So we think we've seen something quite interesting about the brain here, namely that it can implement the exact same decision from the outside using completely different brain circuits. And this suggests that the brain is actually tremendously flexible and shaped in very profound ways by context and experience. Because we're not just talking about changing the synaptic connections between individual neurons here. We're talking about using entirely different sets of brain regions for the very same decision. And we think this might have some important implications for how we want to study cognition as systems neuroscientists. Because we have shown that behavioral task design, details, and training history really matter. If we have two animals that perform this outwardly seemingly identical task, they might actually be relying on different brain regions to do so depending on their experience. So we need to control for experience much more carefully. But in addition, we are also suggesting that perhaps we should leverage this diversity to create more diverse and naturalistic laboratory settings in which we study decision-making. Now where do we take this work next? With Sofia Soares, in the Harvey Lab, we've built a special microscope that allows us to image different brain areas simultaneously. So we're currently using this approach to look at activity in all those decision-making areas that I've talked to you about previously. And we're asking how activity in those areas, but also across those areas, may differ depending on the cognitive context in which animals are making the seemingly same decisions. And another interesting research direction taken by other lab members is trying to actually get to what's the inner workings of those individual areas. So all the optogenetic approaches that I've shown you today were quite coarse, silencing entire brain regions. But Selmaan, in lab, has developed a method where he can activate individual neurons one at a time while monitoring the activity of the surrounding tissue. So here you see, from left to right, he's intentionally, with lights, activating one neuron at a time in the living brain. And he can use this technique to ask questions about the function, and micro-circuit architecture, and about computations that a given circuit may be performing. And Dan Wilson, in the Harvey lab, has taken this approach a step further by now activating 10 neurons simultaneously, here shown with the blue arrow. And he's doing this as animals perform decision-making tasks. So he can ask what the causal link is between the activity of individual neurons that he can functionally identify, for example, neurons that respond to a visual a cue that the animal is using to guide its choice. So you can link activity of those neurons, now, to the network activity, but also to the performance of the animal, trying to draw causal links between activity of individual neurons and cognition. And with that, I'd like to thank everyone who's contributed to this work, first and foremost, Roberto Barroso-Luque. He was a research technician in the Harvey Lab whom I very closely collaborated with on this project and who is now off to grad school. But he really helped to push it forward and push it in all different directions, to the scale that you saw today. I'd also like to thank Chris Harvey, who's been a tremendous advisor on all aspects of the project, from training mice all the way up to preparing this talk. I really value his input, and advice, and his passion for science in general. I'd also like to thank Selmaan, who originally designed this flexible decision-making task that inspired the whole project, really, and then the whole Harvey Lab community for fun, scientific discussions and great feedback, our funding sources, the research and instrumentation core, and the medical school community in general. [APPLAUSE] Thank you, Charlotte. That was wonderful. So our last speaker is Dr. Karl Deisseroth from Stanford. He is the Chen professor and chair in the departments of psychiatry and bioengineering, an interesting combination which really defines his career. Karl did his undergraduate work here at Harvard, majoring in biochemistry, and then went to Stanford to do a combined MD/PhD. He also worked with Dick Chen, as did Ed, and studied the coupling between activity of neurons, calcium entry, and cellular processes. And this is where I first got to know him. Because I was doing similar work here with Wade Regier. After finishing graduate school and the MD, he did clinical training in psychiatry and is still a practicing psychiatrist. As we've already mentioned, he, along with Ed and Feng, were the first to put channelrhodopsin into mammalian neurons and show that they could control the excitability of those cells using light. Since then, his laboratory has really led a steady drumbeat of developments in optogenetics over the years, in which he's produced literally dozens of different kinds of optogenetic actuators that we can use to manipulate the activity of cells. His lab has also produced light-activated G protein-coupled receptors, step function options, and many, many other tools. Separately, his laboratory also invented the light clearing-- the brain clearing approach CLARITY that's now being used ubiquitously to look at the structure of the brain in intact organs that don't need to be sliced. This has become a very powerful technology. And as you'll see in a minute, in addition to inventing technologies, Karl's lab has constantly used these to make fundamental discoveries about the organization of the mouse brain, and I think, led by his own experience as a psychiatrist, has really begun to reveal how the animal not only makes decisions in a normal state, but also how this goes wrong in some pathological states. Karl, thank you. [APPLAUSE] All right, thank you, Renardo. Very grateful for this tremendous honor. And heartfelt congratulations to my fellow prize winners. This is a wonderful moment. So thank you for all you've done over the years. I want to-- since going last, you have heard a lot about things. So I'm going to move, as quickly as I can, to the present day without spending too much time on the past. I do want to talk a little bit more, in even greater detail, about the inner workings of the channelrhodopsin protein itself. The progress of this field has been very rapid. As recently as 2011, we did not know much about the inner structure of channelrhodopsin. But things have progressed very quickly over the ensuing six years. We now know a great deal. Here's that retinol binding pocket that Peter mentioned. This is the ion pore, lined with charged and polar residues, including these five glutamates E1, E2, E3, E4, E5. And attaining this level of understanding of the protein is, of course, exciting in its own right for people who care about proteins, and molecules, and amazingly, elegant natural machines like this, but of course, also has led us to be able to change them, change their properties fundamentally in ways that really matter and are useful. For example, we were able to make them faster, as Peter mentioned-- so here, going up to 200 Hertz, spiking with a very fast mutant, as we described in 2010, getting red-light-driven spiking, as we did in 2011, together with Peter Hegemon and Ofar Izar, who's here today, getting this bistable operation with the step function tools, flipping cells into and out of excitable states, and then making the channelrhodopsins inhibitory, and then making that, in turn, be bistable inhibition as well. And all of these stemmed from molecular modeling, structural determinations, and great deal of work, most of it in collaboration with Peter Hegemon and many other very talented colleagues. I want to touch on two aspects of this scientific journey that were particularly useful and relevant to modern neurobiology. Of course, a big part of it was getting these three crystal structures. When we got the 2012 one in collaboration with [INAUDIBLE] Cato, Feng Zhong, Ofar Izar, Shar Ramakrishnan, and Osama Nurekee, we saw, right away, it was a dimer of two 7-transmembrane proteins. Each one had its own retinal, its own pore. But we also-- in seeing the pore, we, for the first time, had the opportunity to change it. It had been prominently hypothesized that the pore might lie-- rather than within each monomer, might lie at the interface of a dimer, or even a trimer. This turned out to be wrong. But of course, not even knowing where the pore was, it would be very hard to re-engineer it. And we were able to do that. In looking at the inner lining of the pore from our structure, we could see that it was largely lined with polar, but also residues that would be predicted to give rise to negative surface electrostatic potential in the internal lining of the pore and in the inner and outer vestibules. And this led to an idea to change the ions' selectivity. This wild-type channelrhodopsin was a non-selective cation channel, as you've heard, fluxing sodium, potassium, not calcium and protons. And Andre Barent and Su Li, in my lab, worked hard to change that inner lining to be more positive. And they succeeded. Against all odds, this came out, along with a beautiful paper from Peter Hegemon-- a similar result and different mutations, both ending up in creating this chloride conducting channelrhodopsin capability, which allows one to deliver blue-light-based inhibition of spiking. And together with Peter, we optimized these further and created the step function inhibition forms of these inhibitory chloride conducting channelrhodopsins. So these have been, now, widely used. For example, together with Will Allen, in my lab-- Will is now here as a Harvard fellow-- we were able to use this fast IC++, the next-generation version of the chloride channel, to identify the causal roles of neurons that are involved in the fundamental survival drive of thirst. Now this is just one example. Getting to the inhibitory chloride connecting channelrhodopsins in 2014 was one step. But then a very interesting thing happened. The following year, John Spudich's lab identified naturally occurring chloride conducting channelrhodopsins from Gulliardia theta. And just last year, we were able to get the crystal structures of both the naturally occurring chloride channel and the one that we had produced together with Peter. And this gave us a very interesting insight into both the natural and the designed chloride conducting channelrhodopsins, in particular that both the engineered one and the one that nature had developed, in fact, used this principle of surface electrostatic potential within the pore, and also at the surface vestibules of the channel pore, to exclude, in this case, likely, anions and [INAUDIBLE] cations, in this case, to create a anion conducting pore. So this revealed-- and we were able to both convert anion conducting channelrhodopsins to give them cation selectivity, take cation conducting channelrhodopsins, give them anion selectivity, all based on this structure-based analysis of the pore. So at this point, we understand the pore, at least to some extent. And as you'll see later, that's even helped us screen for, and identify, and understand new kinds of opsins that have new kinds of functionality. But I want to talk about color selectivity as a very important step first. And this red-light-driven spiking, in part, depended on discovery of a red-light-driven channelrhodopsin, which was work, again, in collaboration with Peter and with Ofar Izar, but work led by Feng Zhong in my lab. In 2008, he found this red-light-driven channelrhodopsin from a multicellular green algae called Volvox carteri. And this enabled, ultimately-- although we didn't realize that it would have this impact at the time, this ultimately enabled us to even move beyond what Crick's initial concept of the utility of light control might be. And this has been shown a couple times already. But I want to focus on a different aspect of it, of Crick's initial statement. He very clearly focused on types, types of neurons-- engineer a cell type. And indeed, this is very useful. And indeed, this is how the vast majority of optogenetics has been done around the world, allowing one to turn on or off genetically targeted cell types. But he didn't, even in this piece, describe a control of multiple individual neurons, which is what, ultimately, the red-light-driven channelrhodopsin have done a great deal to enable. Of course, in the very first experiments, we controlled single cells. But this was in-- with a readout of a patch clamp, a pipette in culture. And here, showing some of those very early experiments, here is the small group back then. Here's myself, and Ed, and Feng-- the good old days. With Mike Greenberg in the audience, it's nice to point out that the initial readout of membrane depolarization was CREB Ser133 phosphorylation. I was initially a patch clamper, but I also did a lot of work that ultimately followed up on Mike's identification and creation of reagents that allowed us to study this very interesting phosphorylation event. So let the word go forth, from the Warren Alpert Symposium, that Mike Greenberg helped launched optogenetics. So thank you, Mike. [LAUGHTER] But then, of course, Ed's gorgeous spiking recordings, Fong's elegant design of viruses, and his design of fiber-optic interfaces to allow us to control in behaving animals-- and that led to this initial control of mammalian behavior in 2007, where this illumination of supplementary motor cortex M2 on the one-sided animal causes the animal to rotate in the opposite direction. As soon as the little blue light turns off, the animal stops rotating. Now even this was with control of types of cells. In this case, the photosensitive cell population was layer 5 cortical neurons. And we went on from there to target deep hypocretin neurons in the lateral hypothalamus-- again, work led by Feng Zhong and Antoine Adamantidis-- but all of this cell types. And what ultimately turned out to be particularly useful in opening the door to single-cell was a derivative of the initial Volvox channelrhodopsin that Ofar, and Peter, and myself, and several in our group described in 2011 called C1V1. It's a chimera of channelrhodopsin-1 from minimonis and Volvox channelrhodopsin-1. Roheat Prekash, in my lab was, able to express this in 2012, here using a patch clamp electrode and loose patch configuration in a awake and living mouse, and doing raster scanning to photon illumination just above the cell, not getting spiking, within the cell, getting spiking, and just below the cell, not getting spiking-- so single-cell resolution control in vivo, in mammals. And this was work in collaboration with Adam Packer and Rafa Usta. And it was back to back with a paper, also collaborative, between our two groups showing the first spatial light modulator, a liquid-crystal-based holographic control of single cells. But that was in culture. It took a number of years between 2012 and the present to actually translate this into the control of mammalian behavior by control of multiples-- individually specified cells. The path to this led through all optical experiments-- so using the red-light-driven aspect of the Volvox-driven channelrhodopsins and the blue-light-actuated calcium sensors, like the GCaMP series that you've heard a fair bit about already today. And we had combined-- in a experiment, together with David Tank, we'd combined a Volvox-derived opsin together with readouts of calcium signals, blue-light-driven calcium signals, the GCaMPs, in 2014. But it was not with behavior as a readout. It was simply showing that you could do all optical interrogation of neural circuits, but not affecting behavior. It was in behaving animals, but not affecting behavior. It took some time, even from that point, to get to the point where we could exert control over mammalian behavior at the level of multiple single cells. This was work from earlier this year led by Josh Jennings, and Tina Kim, and Jim Marshall in a lab. And what we did was target orbital frontal cortex. This is a part of the mammalian brain that, in human beings, if it's lesioned, gives rise to a syndrome called orbital frontal syndrome, where you can have serious disregulations in feeding behavior and in social behavior. This is a structure that, of course, does many other things as well. It's involved in value-based decision-making. But we were interested in understanding and leveraging our potential for single-cell resolution control. Our goal was to see if we could study the interaction and competition of two primary drives, primary survival drives, feeding and social interaction, within this structure. And these cells, as is common in systems neuroscience, are very often observed to be active during behaviors, but not necessarily known to be causally involved in those behaviors. And so that was a big and open question. So what we did was use a GRIN-lens-based optics to give us access to this somewhat deep cortical structure and exert control over the single cells while also getting readouts via calcium imaging of the activity of those individual cells. And under this system, we can quite readily identify feeding cells. As this trace progresses at each of the gray bars, that's when a little droplet of a high-caloric reward is delivered. And we can see cells that quite reliably respond at the feeding droplet delivery. And so these are, as we call them, feeding cells. These are cells that always-- or almost always-- respond when this high-caloric reward is given. So this is their naturally occurring activity, and this is very useful. They're sprinkled in among other cells that don't have these properties. And then we can come and give optogenetic control, leveraging our ability to exert this single-cell resolution control. And we can do this as shown here. The non-targeted cells called NT here, these are cells that are next to stimulated cells. And this gives you a flavor of the spatial resolution. You can see the non-targeted cells, the NT cells, are active in their own right, doing their own thing. But they're not activated when we activate, optically, the feeding cells that are right next to them, absolutely cheek by jowl with them in this structure. So we have single-cell resolution control over these cells. And then the question is, do we have-- are these cells causally involved in the behavior? And so here's the result of that experiment. If you-- in this case, driving-- we found if we just drove 20 to 25 of these feeding cells, we could enhance and extend the feeding response to the droplet where each little tick is a lick delivered by the animal. Did a number of controls. For example, if you leave the Volvox-derived opsin out, you don't get that response. So this is-- that's good news for the experiment. It shows that it's not an artifact of the light, for example. But of course, you might ask, does it matter that you're targeting the feeding cells? What if you had targeted other cells too? And what if they were important cells to the animal cells involved in appetitive drive. And so here, we wanted to identify the socially responsive cells. And we did that experiment as shown here. Now this is all done in the head-fixed configuration. So you might ask, what, really, is head-fixed social behavior? And it's maybe not quite the same as natural social behavior. But in this case, there definitely is a conspecific social interaction, an interaction with another member of the species, a juvenile, same-sex mouse that can move freely around this chamber and occasionally come in here, where there's an extended period of whisking and sniffing. And indeed, there are social cells that are active when this happens. And those are not the same as the feeding cells. You might also ask, maybe these are just surprise cells, or novelty cells. And that was an important thing. When I'm very surprised or shocked, I might not eat as much. And so that's an important distinction to know. Are these truly social cells, if that's what we're interested, or something else? And so here we 3D-printed a mouse and had it pop up in quite a shocking, almost a horror movie type fashion. I'll play this movie here. [VIDEO PLAYBACK] So the mouse is here, waiting. Oh. [LAUGHTER] And what's amazing is the social cells absolutely do not respond at all, even to that shocking stimulus. There are other novel object cells that respond. It's going to come up again here. And other controls give us confidence that these are related, in some way, to social interaction. And then the question became-- [END PLAYBACK] --if you drive those cells, what happens? Does it increase, or decrease, or do nothing to the feeding response? And what we found, at least for the first couple minutes, was there was, in fact, a suppression of the feeding response by driving the social cell. So here, in orbital frontal cortex, there is a interaction between cells observing, or related to, primary survival drives. If you drive other cells that are non-feeding or non-social-- in SNF cells, you see no effect at all. So it does appear to be at least somewhat a specific aspect of the feeding cells-- of the social cells in their effect on feeding. So this was exciting, because we were able to take these Volvox-derived opsins and finally do an experiment that we'd wanted to do from the very beginning, which was to exert control at the multiple individually specified cell level over mammalian behavior. But of course, we've been wanting to push this to yet further and further levels. One limitation that, at first, seems almost like a physical limitation is one has to be very careful with the amount of light power that one delivers into living tissue. These are powerful lasers that we're using to generate these spots and drive these cells. And if we start wanting to control more and more cells, we may, if we're not careful, enter into regimes where we're delivering too much heat or damaging cells in other ways. So we've been working hard on this, working on ways to deliver spots of light over broader fields of view, which is good in its own right as well, and also finding, and identifying, and using opsins that need vastly less light in order to give rise to still fast, triggered responses. And this has come along quite quickly as well. Along with device development, we've been producing very large spatial light modulators that can give rise to holograms projected over very large swaths of the mammalian brain, up to a 1-by-1-millimeter area, which, for example, can cover most of the visual cortex. These look like this. And these now have enabled us to do the following sort of experiment. What's shown here are six-- 1, 2, 3, 4, 5, 6-- squares depicting 1-by-1-millimeter areas of primary visual cortex in an awake, behaving mouse. And these six areas are listed this way because we're going from superficial to deep. So this is in three dimensions. And all the red, circled cells are cells that are, first of all, coexpressing both GCaMP and a new opsin that allows us to deliver much less light while still controlling many cells that I'll tell you about in a moment. But also, these are cells that have been picked out by us by virtue of their naturally occurring activity. So we're able to individually specify these visual cortex cells by first identifying the ones we want to control by presenting visual stimuli to the animal and picking out the cells that respond in the way we want. In this case, the circled cells are all responding, are all cells that we've picked out because they respond to one orientation of a drifting grading, but not another. So with this configuration, we can do the following sorts of experiments. For example, here are two cells that are more than a millimeter apart. By setting up our holograms, we can truly stimulate, simultaneously, two cells that are set this far apart, or dozens, tens, or even hundreds of individually specified cells at once in three dimensions. And the opsin that we use is a very interesting and strange one. This is the first one that has this unusual property. This is a cation conducting channelrhodopsin. But its primary sequence phylogeny puts it closer to the pumps, actually, than to the other channelrhodopsins, which, by itself, was interesting. We found this in a collaboration with Susuma Yoshizawa and Hideaki Kato. It's from a marine organism. And we called it ChRmine, because we used a structure-guided genome mining approach, and it's a channelrhodopsin. And carmine, I learned from my lab manager, is a deep red color, which, I didn't know that before. You learn a lot in this business. And I learned a lot about colors. But it was a name that actually turned out to be-- although it looks somewhat sinister, it's actually quite a good opsin. Don't say "crime." It's ChRmine. So what this allowed us to do, it was-- its key properties, it's red-light-activated. But it has extraordinarily high photocurrents, more than 5 nanoamps per cell. Very often, of course, that's actually too much. And so we can back off on light power, which is what we ultimately want, and give rise to currents that drive spiking at extremely low light power densities. And that lets us in turn control many cells, tens to hundreds. And I'll show you how we're able to use this to control individually specified cells, look at population dynamics elicited in cortex, and look at behavior. ChRmine has-- although we don't have a structure for ChRmine yet, it's got these very large photocurrents in the red. But also, by homology modeling, we think there are some interesting features to its likely internal structural design-- not yet proven. What's interesting is that most of its predicted surface electrostatic potential is localized, we think, toward the inner and outer vestibules and not so much in the channel lining itself, which may reduce the effective electrostatic stickiness of the-- inside of the channel and allow higher ion fluxes while still giving it selectivity by affecting the access of ions to the vestibule. We are working hard on getting the structure. We don't have that yet. But very large photocurrents, very low light sensitivities-- and this allows us to do this following sort of experiment where we can have an awake, alert animal. We present visual stimuli to it, for example, vertical or horizontal drifting gratings. We can find all the cells that respond to vertical stimuli, all the cells respond to horizontal stimuli, show that they're selective, and pick those out over 3D across a visual cortex. We can also identify nonselective cells-- we call these the random population-- to see that it matters that we're picking out cells that are of a particular orientation or not. And then we can come in and control the cells of one orientation or another, or random cells, both while imaging thousands of cells across visual cortex, looking at the elicited population dynamics, the internal representations, if you will, of the visual stimuli, and also, later, as I'll show you, looking at behavior. What we first found was something-- we didn't know if it would be the case. We didn't know if-- by stimulating a few cells, we didn't know what would happen to the rest of cortex. Would there be no broad, generalized response? If one were to stimulate 10, or 20, or 100 individual cells, would those be, by and large, mostly the only cells activated? Or would we recruit large numbers of other cells? And if we recruited many other cells, what cells would those be? What would be the patterns that they gave rise to? Would it look like something natural, naturalistic, as if the animal were seeing the vertical or horizontal stripes? Or would it be some-- would it be some other aberrant pattern of activity? And so to do this, we look at the population dynamics. It's very high-dimensional in principle, but you can reduce this into a lower-dimensional space with a principal component analysis. And here are the trajectories that these thousands of cells in visual cortex take when the animal is looking at vertical or horizontal stimuli. Here's one mouse and another mouse. And what we can see is this is the natural response to visual stimuli. And this is the response to optogenetic stimulation of just 20 or so individually specified cells of the same orientation. And what you can see is that the trajectories in this principle component space resemble those that are seen during natural visual stimuli. And those are not seen with random cell stimulation or no stimulation. And it's not just us looking at this and saying, huh, those look sort of similar. You can train a classifier and see that the classifier can automatically identify which orientations were the cells that you were stimulating based on the population dynamics of the response. So this was reassuring in many ways. It was nice to see that population dynamics at this regional level elicited by properly targeted optogenetics resembled those that are given rise to by natural stimuli. It also was quite interesting that stimulating just 20 or so cells-- and these are untrained animal cells that haven't been worked hard with light patterns for long periods of time that might have gone through plasticity. These are in animals that are not behaving. We're just looking at the population responses. And stimulating just 20 or so cells can give rise to this broad recruitment of hundreds of neurons among the thousands that we're imaging from in this 3-dimensional space of visual cortex, which, by itself, was interesting. And in this paper, together with Surya Ganguli, an outstanding computational neuroscientist at Stanford, we've begun to explore what this means that cortical circuits seem to exist in this critically excitable regime. But of course, we also did want to see if we could affect behavior as well. And so for these sorts of experiments, we take animals, and we train them to respond to one orientation of the visual stimulus or another. And we can make the job challenging by reducing the contrast, for example, of the grating. And we can generate psychometric curves. We train the animals on high contrast, which, they learn the task very well and perform at high levels. You've seen this d-prime measure before. But even after training, they can't do well at low contrast, at 2% contrast. And 10% is fairly intermediate, and they can perform all right on that. We found a couple of things. First, by optogenetic stimulation of a population of cells that was concordant with a weak-contrast visual stimulus, we could improve the animal's behavioral performance. We could help it to detect, significantly more reliably, what the correct orientation was if we gave stimuli concordant with the orientation of the visual stimulus. But then we even took away the visual stimulus completely. And so in darkness, we asked, just by stimulating a few cells, can we get the animal to respond as if it is seeing the visual stimulus? And the answer is it can. And what's more, we could titrate down the cell number to remarkably low levels. And this is both in terms of behavioral measures and in terms of looking at the population dynamics, the classifier automatically detecting and classifying the nature of the stimulus. And you can see number of neurons stimulated here. We can actually drop that even below 20, to just a handful of cells where we can still detectably see both the correct behavior and the population dynamics response. Looks like layer 5 cells are a little more potent than layer 2, 3 cells and that you can get down to fewer of them to get a comparable level of behavioral response. So this, by itself, is also quite interesting and leads to a lot of interesting questions about how cortex is set up to allow this sort of ignition, or critical dynamics, to be present and to not cause any problems. Now this, of course, was also interesting. Because we were seeing naturalistic dynamics elicited by a few cells, this was broad, and 3-dimensional, and covered most of visual cortex. But of course, we would like to know, even brain-wide, does properly targeted optogenetics elicit naturalistic brain-wide responses as well? And although we can't see, in real time, the activity all across the mammalian brain with optical tools, we can get such measures with electrical tools. And this is work led by Will Allen in the lab, who, as you can tell, has many talents. And he led this experiment where we used neuropixels probe, which are probes which are very long-shank, high-density electrical recording devices, placed in different trajectories in different animals. In this experiment, there was one neuropixels probe per animal. But by using a temporally precise task, we're able to combine the results across many different animals. With known trajectories, we clear the brains. We see where the trajectory of the neuropixels probe was, align that to the Allen Brain Institute Atlas. And we can build up, in this way, a brain-wide understanding of the populations of cells that are active during behaviors. And as a first step, we picked, probably, the simplest possible behavior we could imagine, which is just an animal licking for water when thirsty. And this is, by design, the simplest possible task. Because we wanted to see, across the whole brain, what would be the representation of this task. And starting from the simplest possible task was a good place to start. And in this go, no-go task, the animal has learned that one odor means that there will be water coming. Another odor means there will not be water coming. There's an onset, an offset of the odor, and then an onset of reward, three vertical lines. And here's just the behavior. The animals learn to lick for the go odor and not for the no-go odor. And this is over many trials, the animal eventually becoming sated and no longer licking, even for the go odor. First question is, what happens to my computer? Let's see. There we go. First question is what happens across the brain without optogenetics? What is the brain-wide representation? And an interesting finding-- these are all different brain regions, color coded. Here those three dotted lines indicating the different phases of the task. Red means more active, and blue means less active. You can see there's recruitment of virtually the whole brain by this very simple task. In fact, more than half of all the neurons that we recorded from were statistically modulated by the operation of this task. That was an interesting finding in itself, which has its own implications, which we maybe can talk about later. But then the question was, what happens if we optogenetically drive, in a properly targeted way, a deep population of thirst neurons? And here we targeted the same pathway that we had identified using IC++, the engineered chloride conducting channelrhodopsin, to implicate a particular population of thirst eliciting neurons. And so targeting the subfornical organ input to the MnPO thirst neurons, you can get the sort of behavioral result, animal licking to the go cue, becoming sated. And then when you drive the thirst neurons, this tiny population deep in the brain, you can restore this triggered survival drive behavior, licking for water. And so then the question is, what's happening across the brain in this setting? And the answer, remarkably, is that it's very similar to the natural state. All across the brain, tens of thousands of neurons-- here's the natural licking for water state. Here's the sated state. And here's this optogenetically induced state resembling the natural state to quite a remarkable degree. So this is good news, of course, for everybody studying optogenetics, that if you target things right, you elicit naturalistic dynamics both locally and across the entire brain. And also, it raises a lot of interesting questions analogous to the ones I mentioned in visual cortex. What does it mean for the controllability of the brain? How is it structured to allow small populations of cells to elicit such broad responses? This clearly is a very elegant design, one that probably can go wrong, and is beautiful when it goes right. And I'll wrap up there. But I will show this last slide, which, by chance, was the first slide that Peter showed at the very beginning. I think it's a useful slide to reflect on. Because all the exciting advances that we've made in understanding the brain, and mammalian behavior, and behavior of many species and circuits across biology, in many ways, is deeply rooted in botany and in basic science of studying plants. And so it's a nice story, I think, for us to keep in mind, thinking about the value, and the importance, and the need to support basic science. And I'll take a moment at the end to thank all my amazingly talented students and postdoctoral fellows. I mentioned many of them along the way. The crystal structure work-- I mentioned the stuff that came out very recently was led by Yun Kim, a graduate student in my lab. I think I forgot to mention him early on-- extremely talented. He might be coming here as well. We'll see. But many other very talented students and postdocs along the way-- the work that he did was also important for identification of ChRmine and for-- the paper on eliciting the visual responses in mice was led by Jim Marshall, along with Tim Machado and Sean Kerin in the lab. And all my many collaborators around the world over the years-- first and foremost, Peter Hegemon, but many others as well-- and again, the amazing, talented people who've worked with me back at Stanford, Ed, and Feng, and many others, it's been a wonderful time, a lot of astonishing progress. And it's been a pleasure to share it with you. So thank you. [APPLAUSE] Wow. OK, thank you very, very much-- an incredible afternoon of great science. I want to thank Bernardo for his moderating this. I want to recognize, again, the members of the Warren Alpert Foundation. I had forgotten to mention early, our former dean, Joe Martin is here, also a member of the foundation board. And finally, to congratulate the winners on this spectacular display of science-- thank you very much. And I look forward to seeing you all here next year again. [APPLAUSE] [BACKGROUND CONVERSATION]
Info
Channel: Harvard Medical School
Views: 24,504
Rating: 4.5983262 out of 5
Keywords:
Id: hOKLMjuwlLo
Channel Id: undefined
Length: 234min 38sec (14078 seconds)
Published: Thu Oct 03 2019
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.