1 >> Michael Freedman: So welcome, Microsoft regulars and our new visitors who have come from, in some cases, great distances. So thank you for being here and thank you for your willingness to teach us in a very rapid tutorial that Matthias and Krysta have scheduled about your respective subjects. I think it's probably correct to say that Microsoft is the largest curiosity-driven research institute in the world. Microsoft Research. And once our curiosity has taken to us some place, it's our tradition to understand the real world itch cations and see what's out there. And in the case of quantum computing, we do have a very history of curiosity research in this area for over ten particular, with this project, we're very interested in applications. How does the theoretical acceleration of translate into real world applications. strong and longstanding years. And in the real world certain problems So in a nutshell, what will quantum computers be able to do best if and when they're built. Now, many of you come from disciplines where perhaps you don't think of yourselves as quantum computer scientists, but I want to reassure you that anyone who does quantum mechanics professionally for a living is in a very good position to discover and design quantum algorithms. I think you do not need to be familiar with the formalism which the next talk will explain of quantum circuit model. So I don't think you need to think in terms of pictures like this, for example. It's too small to read, and I don't read this language anyway. And I'd like to give you some evidence for this, and the evidence is that very interesting ideas, you might call them breakthroughs in thinking about quantum algorithm, have historically, in many instance, come with their own model. People who know quantum mechanics and understand some aspect of it think of something that quantum mechanics can do. And later on, people translate that into the quantum circuit model, into, you know, pictures like that. But the initial inspiration may come with its own design. For example, the idea of using quantum computers to solve logical problems, satisfiability problems, max, cut problem, things like that. Problems that are typically NP hard, we don't expect to be able to solve them in polynomial time, but it's possible that quantum computers can give some interesting speed-up, perhaps a 2 square root in the exponential, who knows, over classical algorithms. And the idea here was introduced initially by Fahri, Goldston and Guttmann, and it, as most of you know, was to initialize Hamiltonian so that the ground state solved a rather trivial problem, like finding the ground state of a magnet and then evolving the Hamiltonian slowly so that the solution tracked in the ground state. It's a problematic proposal, and it has to do with the size of the spectral gaps how well it will work, but it's an idea that just required knowing the adiabatic theorem. It didn't require knowing what a quantum circuit model was. Another example from the same group, actually, is using scattering theory. You can think of it as quantum rather than a random walk in a tree. You send wave packets, perhaps, of light through some kind of optical tree. And depending on how you've designed the tree, the transmission probabilities can tell you something about solutions to game theoretic problems, such as values of a nan tree. And, again, this was within a week translated into the quantum circuit model, but the idea just came from knowing scattering theory. And a third example is from Aaronson's work. It's actually with a graduate student Kiphoff, whose name I should have included. And, you know, this is based on the old joke that in quantum mechanics, bosons have a much harder time of it than fermions. Because free fermion systems, the ground state will be Slater determinant. And we know from a computer science point of view that determinants are polynomially very easy to compute, and it's, you know, one of the great achievements in complexity theorem from 1979, valiant, that permanence, which is the positive version of determinants, this is where all the alternating signs are just forgotten and become plus signs, that those are sharp hard to compute. So the fact that the ground state of free bosonic systems will tell you something about permanence is kind of a wedge in to a new way of thinking about quantum computing and new architecture. It's not a complete slam dunk from that statement, because the thing you have to always think about in quantum complexity is that quantum computers are probabilistic machines in terms of their output. So just knowing that an exact solution is extremely difficult does not necessarily tell you that an approximate solution conveys a lot of 3 computational power. So that's why this paper of Aaronson and Kiphoff is 95 pages long instead of two. You know, you have to analyze the error. And this Hong-Ou-Mandel dip, this is like a two-photo on experiment that originally shows the correlations that they proposed leading a machine to exploit. So I'm actually not saying, and I wouldn't assert that any of these three ideas are practical. I mean, I don't know. I think they're all interesting, open questions. But it's just evidence that if you know quantum mechanics and you have problems to solve, you may think of ways of using quantum mechanics and they may really be out in left field with respect to the way quantum computer scientists are trained to write down algorithms. So there's a big coterie of people who will turn your idea into an algorithm if you have a good idea. So roughly speaking, the propose the applications of quantum mechanics seem to lie in three large baskets, you might say of increasing murkiness. So the sharp and clear applications, so the very mathematical ones, finding units in a number field, factoring numbers, discreet logarithm, things like this, these tend to be algorithms that you can analyze their computational costs prove fairly sharp bounds on their efficiency. And it's not the kind of thing where you need a lot of real world experience to learn how well these things work. Now, people have proposed using quantum computers for simulation and design of quantum mechanical systems, nano structure, maybe going way beyond the Hubbard models to studying coop rates, things like that. And that's the kind of thing where one really has to start doing the numbers and seeing whether it's realistic or farfetched, given the resources we're likely to have in quantum computing. And I'd say that maybe the third category is probably closest to the kinds of applications where you'd expect experience to be essential, rather than theory. So to figure out how important quantum computers are going to be in these areas, like general optimization and machine learning problems, it's probable that we need a quantum computer and we have to interact with it and learn gradually. It lacks sort of the mathematical crispness, in my opinion. But that's not to say that maybe some insights can't be gained early. So I'd like to sort of finish by showing you two slides I recently gave a public lecture last month at the [indiscernible] institute in Brussels, and 4 this is sort of typical of the way public lectures on quantum computing either begin or end. In this case, this was the way my ended. So it's full of all these kind of bold assertions. Somewhat shaky. Some caveats about, you know, what things we know how to do and where we might look for areas that classical methods have trouble with. Then I went on to list even more speculative applications and then I got a little modest and said we don't really know. But my goal for this meeting is to get past those first sentences, at least in one narrow area. I'll be very happy if, after we're done tomorrow, I can say sentences two through ten about one of those subjects. Thank you. >> Krysta Svore: Okay. Well, thank you, everyone for coming. And thanks, Mike, for our motivation and introduction for our meeting. I'm Krysta Svore, and I'm a member of the QuArC group, the quantum architectures and computation group here at MSR in Redmond. And because we're coming from different backgrounds, some of us are more on the classical side. Some of us are more on the quantum computing side. I thought I'd give just a very brief -- this is a 15-minute introduction to the quantum circuit model and how we think about quantum algorithms in that language. And a really brief introduction to phase estimation, which is one of the key hammers in our tool box that we would use for quantum algorithm, potentially, in this space and materials and chemistry. So I wanted to start just by introducing you to our group members. Sorry, Alan. I couldn't find a picture last night. So we have, we're a new group. We're about a year old here in MSR Redmond. And for those of you, obviously, I know the Microsoft people know us. But for those of you who are new and outside external to Microsoft, we have a new group here in quantum computation, and we're a sister group to station Q, and we focus more on the computer science, the quantum circuits, quantum algorithms and station Q focuses more on the theoretical physics. So I think we're a good team. And here's our members. So if you see them floating around. of folks are consultants and our key collaborators. The bottom group And our group focuses on designing quantum algorithms, so that's definitely one of the reasons for having this meeting. And we like to think about the implementation of small, medium and large-scale quantum computers. So what would we do with hundreds of qubits, thousands, and then maybe millions. 5 And then we think, as Mike said, you don't need to know quantum circuit or how it will be implemented on the quantum device to design a quantum algorithm. So our group thinks of the quantum algorithms, but then we think about taking an algorithmic, you know, idea and actually converting it down to the device specifics. So we think about programming languages, optimizing the quantum circuit, converting the Hamiltonian into a quantum circuit that's going to be efficient, and questions like this. And then finally, in support of that goal of breaking down the algorithm into the device components and the device operations, we're building out a full comprehensive system architecture that will help in designing and programming a scaleable and, of course, fault tolerant quantum computer. And tomorrow, you'll have a great tutorial on liquid. And liquid is our architecture, our platform for programming and simulating quantum algorithms and quantum circuits and also Hamiltonians. So that will be tomorrow morning, and Dave Wecker here will be giving that then. So in an introduction to that, I'll introduce some of the language and then he'll show you how you can take this language and these tools and actually use it to program or simulate something. So to start with, I know all of us know quantum mechanics. So I'm going to try to just take quantum mechanics and just translate it into what we do for in the quantum computer and quantum circuit model. So with a qubit, and actually a qubit, it lives in a two-dimensional complex vector space. It has an inner product so it lives in a Hilbert space. And we can think of these two, for example, these two vectors form an orthonormal basis, so we typically think, we call this the Z basis. And in the quantum circuit model, we typically write everything in the Z basis. So the state vectors then are just one, zero, for the zero state and then the vector zero, one if we're talking about the one state. So in general, we can write any qubit, any two-dimensional quantum state here as a general linear super position of these two basis states. And so then the vector we use is alpha beta, where the alpha and beta are the amplitudes and alpha and beta are always complex numbers. And, of course, they have to satisfy the normalization condition. 6 So then we can think about if we move to multiple qubits, the size of that state space, and so, for example, a two qubit state, we have now the super position over four possible basis vehicle to and, again, we have all of their amplitudes and we again have to satisfy the normalization condition that the sum of the amplitudes squared has to equal one. And then we can, of course, move to a bigger state so where we have N qubits and, again, it's a linear super position over all of those orthonormal basis states. And again, it satisfies the normalization condition. I'm sure this is all repeat. So then let's think about the quantum evolution of this state, the system. So from quantum mechanics, we know that the time evolution of a closed system is going to evolve following a short Schroedinger's equation so here's a Schroedinger's equation we all know and love. And then we want to talk about how we can express this, and we know that we can solve Schroedinger's equation for two time points T1, T2. So we want to know how the state space evolves between time T1 and time T2. And so we start, you know, with our stateside, our quantum state at time T1, and it evolves as a unitary evolution, a unitary transformation U. And we can see the relation between if we define U to be E to the IH, the difference between the times over H bar, then we can see that the unitary evolution of the system matches the continuous time dynamics of Schroedinger's equation. So this is where we, you know, you come up with the Hamiltonian and then, you know, we want to actually find what the unitary is, and so this can be challenging, finding an efficient unitary that computes what we want. That computes this evolution. So in general, we want to know what these unitaries are and which ones we can use in practice and on the device. And in general, a unitary can be written, any unitary can be written as U equals E to the IK, where K is some Hermitian. So what unitaries are natural to consider when we actually want to build this device and make efficient quantum algorithms? So here are a few that we use in practice. Very common. So for single-qubit operators, so unitaries act on just a single qubit. The typical ones we know, you know, from quantum mechanics and physics, of course, we have the Pauli operators. So the Pauli gates, you know, in physics, we refer to them often as the sigma I, sigma X, 7 sigma Y, sigma Z. X, Y or Z. In the quantum circuit language, we often just call them I, And so we define, we have the identity, the X operator. So X operators a NOT, we call it NOT in traditional computer science language. The NOT operator takes the zero state, converts it to one state, and a one state and converts it to the zero state. And that's defined here by this two by two matrix. And then for the other two Pauli operators, we have sigma Y and sigma Z. Sigma Z changes the phase, and then sigma Y is a combination of X and Z so it does a bit flip and a phase flip with, also, this additional phase I. So we use these all the time in our quantum circuits. And then we, of course, have a phase gate. So if we want to think about any -applying an arbitrary phase theta to our quantum state, then we can do this with a rotation about the Z axis by theta, by angle theta. So in general, then, the matrix is defined as one zero zero E to the I theta, where theta is the angle about which we're rotating around the Z axis. So we all know the block sphere picture so here the zero state sits at the north pole, the one state sits at the south pole and then here's the Z axis, so when we talk about a phase gate, it's rotating there about the Z axis by angle theta. And there's some common Z rotation gates that we use a lot. The first being we've already seen the Pauli Z operator. This is, in fact, a Z rotation where theta is the angle pi. And the S gate, which is called the phase gate, the S gate is one zero zero I, this is just where the theta is now pi over two. So it's again a rotation about Z. And then finally, a very important gate, the T gate, which we call the pi over 8 gate, but theta equals pi over 4. So here, this is a rotation by angle pi over 4 about the Z axis. We call it the power over 8 gate, because you can factor out, in this form, one zero zero, either the I pi over four if we factor out the I pi over 8, then you'll be left with pi over 8s on the diagonal. So it's called the pi over 8 gate. And this is an important gate, because for a -- to have a universal single qubit set. Require something outside the Clifford group. So all of these gates, the Pauli gates are in the Clifford group, plus one more gate, the Hadamard gate is in the Clifford group. And this is not a universal set, the Clifford group. So we have to add an additional gate, and that's the T gate. One candidate is the T gate, which allows us to implement any single qubit unitary in practice. 8 So the final gate to mention is the Hadamard gate, and this gate is given by with one one one minus one plus this normalization, and this gate allows you to perform a super position so you can take the state zero to zero plus one over root two and the state one to zero minus one over root two. And now we can start introducing some quantum circuit diagrams, the smallest one in this so far. So normally, we denote the gate as a box, just with the name of the gate inside of it. And then this represents, the wire is the qubit. So this represents applying the Hadamard gate to whatever the state of the wire is to state psi. So then we can move to two qubit operators, and the most common two qubit operator is probably the controlled NOT gate. This is a controlled X gate, and what this gate does is it takes -- so it's a two qubit gate so we have qubits A and qubit B, and it takes qubit A and qubit B to A and the XOR of A and B. So essentially, it says if my first qubit is zero, I don't do anything to my second qubit. If my first qubit is one, then I'm going to flip, apply the X operation to my second qubit. And so the matrix here, you can see that in the lower right corner, we have the X gate. And in the upper left corner, we have the identity. So in general, if we want to apply any controlled U, which comes up a lot in the quantum chemistry algorithm, we have to apply a controlled phase gate, then if it's a single qubit phase, for example, we would have the identity here and then that U or that phase gate sitting in the bottom right. So now we can construct any controlled unitary operation. And the diagram for the controlled NOT is this here, where this is the control lean. So if this qubit is one, then we apply this, and that's the symbol for -- well, it's the XOR, which is why we're using it here, but you also might see it controlled with an X in a box on the bottom qubit line, which would match this notation here. So this is the notation for any controlled unitary operator U. And again, the filled-in circle means that qubit has to be in the state one for it to be applied. Another gate that comes up is the swap gate. Swap gate simply flips the state of the two qubits. So A, B goes to B, A. And this can be, for example, it can be implemented with a sequence of three controlled NOT gates. And the unitary is given by this matrix here. 9 So if you, you know, you can solve this and you can see that if you do three controlled NOTs, where the middle one is in the opposite direction, then you will see that you have a swap gate. So feel free to matrix multiply away. You can solve that. And then the notation is often these two Xs with a line in between as they were swapping the state of those two qubits. So that's a quick introduction to all of the quantum gates. Most of the quantum gates we use. And then I just want to go through and introduce phase estimation, because I think this is going to come up in a lot of, maybe a lot of the talks today. And so I think we all know this well. But in terms of the circuit notation which we use, which you're going to see a lot of tomorrow in Dave's talk and maybe some in Matthias's talk, or maybe not, definitely in Dave's talk, this will give you an idea of how we think about it in more of a computer science circuit way. So this is one of the -- or maybe the most common subroutine. Grover is a candidate as well, of course. But between Grover and phase estimation, those are our two main tools for designing quantum algorithms. We'd love to have more. So we should all be thinking about that, of course, but these are good things to look at when you're thinking about a quantum algorithm. So in the phase estimation problem, we have some unitary U, and then we have an eigenvector psi, eigenvector of U and then the eigenvalue of this is E to the 2 pi I phi. So that's, you know, U psi equals E to the 2 pi I phi psi. And the question we have when we're doing phase estimation is we don't know this phi value here. And so we want to say, you know, can we get an estimate on what phi is and to what precision. So we want to estimate this value phi. And there's a way, a known circuit to do this, and there's actually several known circuits. I'm presenting one. And there are different circuits that don't use some of these gates here that you can do more classical processing, and we can talk about that separately. But here, I'm just going to introduce one of the most -- probably the most standard that most people see in the textbooks. So here's the quantum circuit for performing phase estimation. So we start with a bunch of qubits all initialized to the zero state, and then we have our eigenvector psi. So 10 already, one of the assumptions of the subroutine is you can prepare psi in some efficient way. That's not always possible, of course. So we want to think about algorithms where we can actually prepare psi efficiently, and then so we have then a large super position state here in the top register. So we refer often to a set of qubits as a register, as we would in more traditional computer science. So this register is going to be controlling the application of these unitaries. So what happens is now these are in a large super position so part of that state or, you know, one part of the basis states that are in these super position are going to get the Us applied to them and the other one won't. So then we have a sequence of powers of U, controlled powers of U. we're going to apply the Fourier transform. And then So what happens, just at this state in the computation, after we've performed all of these controlled U operations, the different qubits, the different wires here contain the following states. So at this point, this qubit is in the state zero plus either the 2 pi I 2 to the zero phi one and so on down to the 2 to the N minus one phi. So you can see where this is starting to give you, if you can extract this information about the phase here, you can get information about the different bits so you can get different precision or accuracy on the estimate of phi. Request of so what we're trying to do is get information about each bit of phi, basically, here. And how do we get that out? Well, we know another trick, we can use the quantum Fourier transform in this case, we use the inverse of the quantum Fourier transform. And what does that do? So this large box here has a bunch of gates going on inside of it, which I'm not reviewing in this talk, but we can -- but it's a straightforward, efficient quantum circuit. It's polynomial time. Actually, quadratic. So we can -- what this box does is it takes this state, which is the state of the system at this point, and it's going to map it to this estimated version of phi and then it leaves our state psi, our eigen state, our eigenvector, it leaves it alone. So the algorithm ->>: Can I ask on this slide, what is the little slash with the M next to it? >> Krysta Svore: Oh, that's saying -- sorry. Yeah so in the circuit model, we 11 often do this to denote that there's -- this is an M qubit state. So this register has M qubits. This register has some other number of qubits based on what accuracy you need. >>: [indiscernible]. I think the two are unrelated. >> Krysta Svore: Yeah, sorry. We have N qubits in this register and M qubits in this register, in this diagram, yeah. And so if we were to just look -- so in an algorithmic -- if we're to write this out as an actual algorithm, then the input, while we require a black box or a quantum circuit that's going to allow us to perform controlled U to the J operations, so depending on the efficiency of this, this could be an efficient method or not. We require the ability some number where this we want -- so based on be and then, also, how to prepare this eigenvector psi, and then we also need is going to be set based on how accurate and how often two things. How accurate we want our estimate of phi to often we want it to succeed. >>: Just a think, the M in this slide is not the same as the N in the prior slide? >> Krysta Svore: >>: Oh, I probably switched notations, sorry. The N is the thing on the right of two plus one. >> Krysta Svore: Oh, I think I moved some stuff around, sorry. Okay. And then I changed a letter again. This should be M, I think. Sorry about that. Okay. So then the output is an approximation to phi. And then the run time of this, the QFT takes N squared time where N is now my N here and the oracle and it's plus the cost of this black box, this oracle. So I'm saying we can apply this, you know, controlled U to the J. If we can do that efficiently, then that's the cost. >>: There's no other [indiscernible]. >> Krysta Svore: Yeah, so you can do a little better than this, I think, in some implementations. We do have, yes, so you can do an approximate -- well, you're asking about the [indiscernible]. >>: [indiscernible]. This is the N squared? 12 >>: No, no, because the N here is different. >> Krysta Svore: Oh, right, yeah. Sorry about that. We're in the And then let see. So we have some success probability as well that The algorithm will succeed with some probability of success, and we that, you know, increase the number of qubits if we want to succeed for example. quantum. we use. can set more often, And then the procedure, so as we saw in the circuit, as we write this down in an algorithm, we would say we start with two quantum registers. This register, according to this slide, contains N qubits and then psi has M. And then that means -- after we perform the Hadamard gate, which we saw on the previous slide, then that takes us to a super position over all the states, plus our original eigenvector psi. And then we undergo this black box, this U to the J. >>: So when you say it takes [indiscernible] the source of the failure means when -- so where is the random [indiscernible]. >> Krysta Svore: Oh, so we measure all quantum algorithms are probabilistic. So here we're measuring this state. And so there's a chance we're going to measure the state, the zero. If we measure the zero here, let me go back one more. Let's say when we measure -- imagine we're measuring this state, for example. There's a chance we're going to measure zero and we don't get any information. So it's the measurement that has the probability associated with it maybe I should show it here, actually. >>: I think the -- >> Krysta Svore: We're measuring this. >>: I think the thing that's interesting is you have basically, you know, this real number which you're trying to [indiscernible] and the very last part of it, like that is a random choice. You pick the wrong one, then [indiscernible]. >> Krysta Svore: Yeah, so every bit that we're trying to approximate has a probability associated with it. We're not always exactly in this ring. So 13 we're not going to get exactly every bit with -- we're not actually always going to get the zero or one bit exactly. So we have a probability associated with each bit. >>: Let me just try to answer the question. I think the error comes from a mismatch of two discretization. If you have two bits in these top registers and the phase phi was at the K over N and then the denominator, then there would be no source of error. The problem is you're doing a discrete Fourier transform, trying to detect a unit complex number which is not necessarily in that group. So the Fourier transform won't be completely delta function. And then the end of the quantum computation is you sample the Fourier transform with respect to the norm squared. So you will be off a little bit. >> Krysta Svore: Right. Yeah, I should have probably put, you know, that typically, this is some number over N and we're sampling in the cyclic group N. So it might not be exactly N in the denominator as Mike just pointed out. So after we have the super position state, then we have to apply this black box that's U to the J. These are all the controlled U operations that were in the circuit diagram. And we may or may not be able to apply this efficiently. So this is something we really, I think, have to think about is how to do this. And I think Matthias and Dave will present some work on how you can approximate this for different algorithms. In shore's algorithm, for example, this is modular exponentiation. So the output after we apply this black box is this super position state here, then we 'ply, as we saw, the inverse QFT and we get out a sampling of phi, which we're trying to approximate. So this is a general trick, phase estimation, that I think we'll hear about a lot during the next two days. So now I'm going to turn it over to Matthias. >> Matthias Troyer: Okay. Thank you again. So I want to now tell you why we're meeting here. And the reason is that we want to turn the theory of this [indiscernible] application. Because once we build the quantum computer, then you want to use to it take care of something real. And so now we want to think about actually coding up something on a future machine and then I don't like oracles, because I don't know how to build an oracle. I don't have it, I can't buy it. So we need to build this oracle. And there I'm working in high performance computing so I have to think about 14 what I actually can compute on a machine. And then I'm running into problems. But first, I have two jobs. One is I'm a professor of computational physics at the ETH in Zurich, and one day a week, I work for Microsoft as a consultant on the quantum computing and this is a kind of a merger of both aspects. So what can quantum computers being used for? The first thing is factoring. We know it can be done if we build that big machine. The next thing, when you ask what else besides factoring would be interesting in simulating quantum systems, especially fermion systems, which are hard, and that has been around for 30 years since Fineman mentioned it first, and many of you worked on this [indiscernible] algorithms. So and we get the exponential speed up here, which is great, because we use quantum mechanics to evolve the quantum system. We get exponentially a speedup in the memory because we have qubits instead of [indiscernible] bits, and so the wave function can be stored in the N qubits, instead of here 2 to the N, the classical bits. And when we apply, the operator's also exponentially faster. So we have a huge speedup, which is good. But I always become cautious when talking to theoretical computer scientists just that an algorithm is NP does not mean I can ever compute it in finite time on my lifetime. So let's look at what has to be done if you want to simulate the material. And let's look at the Coulomb Hamiltonian, the general one. We have a [indiscernible] term. We have a one [indiscernible] tem. So we have one term with four operators and one with two. And you want to evolve it. And I don't have the oracle so I have to build the oracle. And to build the oracle, I have to evolve it for some time, and I can do that by splitting the time into small time steps, that is T, and using the Trotter scheme to evolve it. And now that's what I have to do. So what is the complexity? And this is a rough estimate. I have an order of N to the 4 terms when I have N basis functions. When I've run four time T, I have T over delta T Trotter steps and for each application, I need on the order of ten or so gate operations. But ten is the lower bound. So this is roughly the complexity. And when I'm saying N to the 4, then I'm getting scared. We know it's hard for classical machines. That's a big 15 problem. But they haven't found ways of reducing it. For quantum machines, you also have something that we have to address. So let me now not focus on how I get the [indiscernible] function. Let me just focus on if I have a weight function, how to get the energy or a gap or something. So let's just look at one aspect you want to get this energy and do it by the phase estimation that Krysta has shown. And so I want to evolve it for a certain time or a certain power and the time is the inverse of the accuracy that I need. So if I want six [indiscernible] places, I have to run for time this 2 pi times ten to the six. But it's other time steps. And now, what is the complexity of this? Essentially, I want to do about T over delta T times ten times N to the 4 and then choosing an accuracy of epsilon. So live this -- the Trotter time step if we can choose one of 0.1, this is a big one I think we can't go smaller. So then we're ending up with 10 to the 3 times N to the 4 over epsilon. Then talking to people in chemistry, when we wanted a least six digits, many people say ten. Let's keep it to six. There's a complex of 10 to the 9 times N to the 4. And now what is the run time if this is my algorithm? If an ion trap currently, the gate space is about ten microseconds per operation. In the super [indiscernible] qubit, one could aim for ten nanoseconds. So let's assume if a perfect decoherence free quantum computer that operates at 100 megahertz and the run time is ten times N to the four seconds. We're going to run it for a month. Let's assume for a single run. So that's how many seconds we have and how many operations we can do. Then the biggest problem size we can do is 22 spin orbitals. I can do them on my laptop in a few minutes. We can do much bigger on a classical measure. So we have a problem in this simple-minded way. So let's being a bit more optimistic. Let's think we can speed it up and get a really good run that operates at a gigahertz. And I'm going to run it for a year. Then we can do 75 spin orbitals. We've beaten the classical machine. What it shows you is that we have a problem. It's not just so simple as to say we can use a quantum computer and it is efficient in polynomial time. It's 16 just the oracle Krysta mentioned. The time evolution of the quantum system. If we do it in a quantum system, like in this pen here, it's efficient because it's analog. If we simulate it in digital hardware in a program, then we have -- I've got N to the 4 scaling. So the question now is what can we do? What important [indiscernible] are there around that we can do if we build something that can adjust just a bit more than the classical. We can do [indiscernible] maybe 60, 70 spin orbitals, which is a little bit bigger. Is there a big problem here. And the second question is we need new ideas to speed up those methods. In the simple-minded way it would be just be very, very hard and so those are the two things that we want to discuss tomorrow in detail. I first want to hear you talk about the state of the field and methods and then we want to open it up to brainstorming of what one can do to make it useful. We have a [indiscernible] so it's not hopeless, but one should not just say you'll be have a quantum [indiscernible] solve all problems of quantum chemistry. It's not that easy if you want your machine. So now ->>: [indiscernible] it seems to me like the real space to make in [indiscernible] is to determine what class of Hamiltonians can rank the sort of [indiscernible] time energy uncertainty. So we know shore's algorithm, just of thinking about it numbered theoretically, I think it is a Hamiltonian basis in time. >> Matthias Troyer: Yes. >>: I somehow can be Eisenberg's [indiscernible]. So I'd be curious if people have a sense of, like, what chemistry Hamiltonians would fit into that category that are not trivial. >> Matthias Troyer: That's one way of going there. The other way is can we simplify the [indiscernible] to get the scaling down. That's another way that one can go. We don't have to make it in one step or a few, but you can find some that you can do faster than N to the 4. These would be two ways of proceeding. >>: Well, there's this other thing that I talked to [indiscernible] a little 17 bit about, which is sort of a -- look at an ensemble of initial states, rather than a single initial state and essentially don't try to simulate the entire system for a long period of time with the sequential performance constraint that applied. Ned, do something much more clever about mixing between the states using a genetic algorithm or something else. This is like what we had with the metropolis algorithm after all. I mean, metropolis algorithm is a very serial algorithm when you implement the [indiscernible]. But, in fact, we have other techniques that don't rely on sequential simulation of the electron positions and so that might be a path. >> Matthias Troyer: Yes, so what we want to say actually is that we need to think about new methods, and we should think problem, and the problem is not solved at the moment. But we have all day tomorrow for brainstorming, and we're late already bit for your talk so let's start. Markus, thank you for coming here. >> Markus Rieher: Good morning, everybody. I name is Markus Rieher. For those of you, which is most of you, who don't know me, I'm a quantum chemist basically so I work with standard many particle methods, which are sold on classical computers to solve chemical problems. And I will actually come back to some of these problems which we had in your list in your first presentation. And I would like to discuss with you what are truly chemical problems. I mean, it's easy to say that, well, we would like to have desensitized solar cells with some properties and so on. I mean, let's look a bit more into the details here. And I happen to be here because I have been talking to Matthias for a few times in the past six months, and he already told me the story which he just presented to you. Say given there's a fixed not too large number of qubits this you could use for quantum computing in chemistry, what kinds of problems could you actually solve? And while I will focus mostly on chemical reactions, hence the title, and most of these chemical reactions will have to do with catalysis, because this is a problem which is really important in chemistry and it's kind of, you know, when you do [indiscernible] metric reactions, we need an equal number of reactants in order to get a product. That is already a challenge, but it's even more challenging to come up with some molecules that can, you know, you just need a few molecules in order to carry out a chemical reaction over and over again, meaning catalysis. 18 And now let's work under the assumption that you have only, well, I call it here a hundred one-electron states which you could construct your quantum in many particle state. These 100 one-electron states I have in mind are actually molecular orbitals. These are not qubits yet, because the molecular orbitals I'm talking about have four states, so it could be MT an molecular orbital, spin up, spin down and doubly occupied. So you would need two qubits in order to represent such a molecule orbital. And I just pick a hundred to have one number which is, I would say, sufficiently large to be interesting for -- to be competitive for standard methods on classical computers. Okay. And like I said, I will focus on chemical rather than physicohemical problems, right. So physicochemical problems are problems related to spectroscope, so where you have some sort of spectrum that you would like to compute. I'm after the chemistry. And the main target for us is the electronic energy and that is defined by the Born-Oppenheimer approximation, so you have to solve the stationary electronic [indiscernible] equation. Now, what would you need in order to describe catalytic reactions? So most of them, especially if they involve transition metal ions, I would say that you can set up the molecular model with 50 to 300 ions, okay. So you could do a huge deal of chemistry with such a model of 50 to 300 atoms. And this model, you would then, depending on the chemistry you're doing, embed into some environment and you would come up with an environment that is easy to model. For instance, a dielectric continuum, usually used to model solvents. You could use a quantum mechanical and quantum mechanical embedding and, for instance, one option is to have a frozen-density embedding where you take your active catalyst and you embed it into the electronic density some of surrounding, which is usually structured. And, of course, you have electrostatic embedding, and that's usually used in protein chemistry and you can set up a force field and into that force field, you embed your quantum system. >>: Can you ask you a question? 19 >> Markus Rieher: Sure. >>: When you say 50 to 300 atoms, is each atom requiring 30 qubits to model because of [indiscernible]? >> Markus Rieher: I'll come to that point. Before I come to it, let's look at this slide. So first of all, I mean, the nice point is for this kind of catalytic reactions, you can get away with a model with, say, 50 to 300 atoms. Could be much more, of course. You have chemical processes where you surely need more than 300 atoms. >>: Can you explain that, why you need 300 atoms? I just don't understand. >> Markus Rieher: Also, a slide. Give me a second. Before I tell you that, another important point to mention is that I would say that we get a way -- we don't need explicit nuclear dynamics. So what we can do is we look at stationary states of the Born-Oppenheimer potential energy surface so you take a structure and for a structure -- determined by the position of atomic nuclei, it's also an atomic structure problem and extract your chemistry from that. You don't need to move the nuclei around. For the reactions I'm interested in, of course, there are chemical problems where you need to do that. But for those problems I'm telling you about, you really don't need that. And these reactions, and that's the most important point, involves the breaking and forming of rather strong chemical bonds. So when you do that, this is when you get away with a model which has only a few hundred atoms. And what you do is you usually break only one or two atom-atom contact at a time. This is why it's possible. Still, the problem is this you need to consider a huge number of nuclear configurations and huge number of molecular structures. And for these nuclear configurations, you saw the [indiscernible] and that also means that -- yet you also need to do for a molecular and different charge in spin states. Now, coming back to your question, say if you take a hundred atoms, and this is a very low estimate, and each atom contributes 10 one-electron basis states, that means you end up with 1,000 molecular orbitals, and I told you we just want to use a hundred. So what can we do about that? That is a problem. The nice thing is that there are already ideas around for standard techniques 20 to circumvent the problem, and that has to do with the active space concept that was basically developed by a group and led by Roos and his coworkers. And the fide is the following. I mean, you know that you have many more molecular orbitals than you can treat already in a standard calculation. Now is it possible to select a reduce the said of orbitals which is totally sufficient in order to describe the chemistry? And this is called the complete active space concept. And so you select a few or the most essential molecular orbitals, which are needed to represent the total state only in these active orbitals. While in the standard methods, you always end up with a kind of a super position of many, many electron states, basis states usually with a fixed number of electrons and you expand your total state, your electronic functions into this basis. And the basis is, of course, constructed from those orbitals that you selected. Okay. Now, there are some drawbacks, of course, because, I mean, you select only a few orbitals and that ultimately has consequences on accuracy so usually, you can make sure that you get qualitatively correct wave functions. Sometimes they are even quantitatively correct, because some contributions might be lacking. And in quantum chemistry, we call them dynamical correlation. Also, it's not guaranteed that it will be possible to select the most relevant orbitals, but we will see that, say, a hundred molecular orbitals is a reasonably good number for the catalytic processes I have in mind. A little bit more notation. You will see this on the slides to come. CAS means complete active space. So that is a set of molecular orbitals in which you construct your total state in an exact manner and you're seeing this on Matthias's slide before. We call that in quantum chemistry full configuration interaction, full CI. So to be more precise, full CI usually refers to the exact solution in a total set of one particle states since we restrict our set of one particle states, we call this exact solution CAS-CI. If you do a full CI only within the active space that we have chosen. Now, since you do this restriction, you could also play around with relaxing the orbitals and then when we do that we call that CAS-SCF. So that's a CAS-CI method, a CAS-CI type wave function where we also optimize the orbitals. Now, two examples that should illustrate what kind of chemical problems I do 21 have in mind, which are important. One example is this one. solved. Well, despite 50 years of research ->>: That's an ancient reaction, though. >> Markus Rieher: >>: It's not yet Right, a very important one. You bet. >> Markus Rieher: So the point is the following. I mean, we have this in air, the nitrogen. Molecular nitrogen, 75 percent. In order to grow food, we need to conform to it ammonia. Usually, it's done in the industry through a process which is a hundred years old. It's called the [indiscernible] process running at elevated temperature and elevated pressure. It consumes two percent of the annual energy production. And it would be nice to really convert the nitrogen under ambient conditions, ambient temperature, ambient pressure, to ammonia. It has not been achieved by synthetic chemistry until the year 2003. And it is possible, though, because there's an enzyme doing this. Now, this is a system which first accomplished it in the lap, and it was published by the Schrock lab in M.I.T. in 2003, and it worked with [indiscernible]. You see here, I hope you are somehow used to chemical structures. So what you see here, this is -- or if you want to solve a chemical problem, you need to get used to it. But, of course, I'll explain this to you. So you see, there's a -- this is the active center. It's a metal. It's a molybdenum surrounded by four nitrogen atoms. Then in the fifth position here, there's a di-nitrogen binding. So this comes from it. Then all of this here is a barbed wire of atoms. Atoms taken from organic chemistry. It's basically a huge system. You see it in a ball and stick representation on the next slide. It's the first system to do, say, a few catalytic cycles. The turnover number that is only six so, say, after three cycles, it drops dead. And that is very bad for a catalyst. You need to achieve a million cycles, something like that. So we are in the position of the first system, which is synthetic, which can do the job, but it's totally inefficient. And since 2003, nothing has changed. 22 There's another change based on molybdenum that's available since two years which can also do the job, also drops dead after a few cycles, say four. Okay. And, of course you need ->>: [indiscernible]. >> Markus Rieher: That we could study with quantum chemical methods. The point is, you provide electrons through these compounds here and you have a proton source. And under these highly reductive conditions, what you do is you attach hydrogen atoms to these nitrogen atoms and the whole thing falls apart, because it's no longer stable. And that's kind of the problem. >>: [indiscernible]. >> Markus Rieher: >>: Yes. So do they use [indiscernible]. >> Markus Rieher: The mechanism is unknown. The active side is known only, basically, since last year. It took 20 years to clarify how the active side is really composed. The mechanism is not yet solved, and one reason is actually that standard quantum chemical methods have problems here for the reason that the number of active states that you could consider is too small. Of course, you can do DFT, but you never know how accurate that is. >>: Is it known how nitrogen fixing bacteria work? >> Markus Rieher: >>: No, that's what I just said. Do they have -- It's not known. The molybdenum. >> Markus Rieher: Yes, they do, but it's believed that it takes place on iron. So the active side of this enzyme consists of seven iron atoms and one molybdenum atom, and they are all bridged by sulphur. So in terms of selecting molecular orbitals, it's a mess. >>: But these are not exotic bonds? 23 >> Markus Rieher: No, no. No, this is -- for me, the conclusion is that it can be done chemically, and so maybe this is also a chemical -- this is a true chemical problem, which I don't know how to solve algorithmically because it's basically a combinatorial problem. The point is you need to know where to place your atomic nuclei, right? That's what I said back in the very beginning using the born-Oppenheimer approximation. But if you don't have an experiment which really tells you where to put them on the nuclei or from which nuclei the catalyst should be composed, well, there's not much that you can do. So and this was the first system that could do the job so everybody jumped on that system to understand how it works. But it's very hard and now impossible to really improve. For instance, we substituted this barbed wire here of atoms and whatever we did, it was worse than the original system. So this is a really hard problem. Well, this is the ball and stick representation of the catalyst. You see the molybdenum atom here in the center these are the nitrogen atoms. This is the [indiscernible] binding. And here you see what I call this organic barbed wire carbon atoms and hydrogen atoms. While we studied the system with basically DFT. This is the only method which is feasible, actually, but we don't know anything about the accuracy. So having more accurate data will be really helpful here. It also tells you that you need to come up with a structure and models that incorporate the acid. This is the acid here. Here you see the proton which is going to be transferred. This is actually a view on some of the reactions after the first ammonia molecule has been produced, because you see there's only one nitrogen atom missing here. And then you need to set up a molecular structure model. Well, we did it first with isolated molecules and then we let the acid approach to different reaction channels which is depicted here. So that is what I mean with you need to consider really many calculations for different nuclear configurations. And this is just a glimpse of the full complexity of the process. It's something that the calculations will be carried out with specific DFT, not knowing really what the accuracy is. Each node here is basically a different chemical species, and you see electrons added, protons added and things are going on, and that's only for the system which with the low turnover numbers. 24 So this is just understanding what's going wrong. It's not yet a solution. So this is one of my major points here. If you just run one calculation on one structure in chemical terms, it's pretty useless. I mean, you can solve a physical chemical problem. But a chemical problem, no way. But what could you do with a quantum computer? Well, the point is there are certain essential steps in the mechanism. For instance, one of them is the exchange of ammonia, which is produced by the newly incoming di-nitrogen ligate. And this is something that the Schrock system is really capable of doing. Many other systems are just there. They produce ammonia, but ammonia poisons the catalyst and then it takes -- you need to get rid of it and bind the new di-nitrogen ligate and this is something which is possible with Schrock system. So in this kind of chemistry, there are only four essential steps which must be accomplished by a catalyst. You need to be able to bind to di-nitrogen. Then you need to activate it, meaning you need to transfer the first electron, the first proton on to it. The rest is energetically downhill. You can forget about it. You need to get rid of the final ammonia molecule which is formed, and you need to increase the turnover number, so you need to deal with a huge amount of side reactions. And especially for the second step, a quantum computer could be helpful, because that is a step which involves electron and proton transfer and it's very hard to model that within DFT. And I'll just show you one example, which is a different system for the same purpose, which would benefit from such a simulation. You see it's a [indiscernible] system and the M stands for metal. The metal is [indiscernible], could be iron. And there's again organic barbed wire and the N2 molecule is clamped between the two metal fragments. The point is you can transfer through reduction or photo chemical activation an electron on to the N2 unit, and then you see that it will induce a structural change. And because of the structural change, protons will hop over to the di-nitrogen ligate, and that would accomplish the important second step. >>: What is that [indiscernible] stand for? >> Markus Rieher: Well, N2 is basically singlet, but the question is what the spin state of the metal fragment. So with respect to N2 as a molecule, it 25 pretty harmless. But the metal fragment is a problem. For instance if you have iron instead of [indiscernible] here, it's not clear it's a spin status. The second example that is from the hydrogen production business for clean energy production basically as it is pursued with an enzyme this time with a so-called iron iron hydrogenase. And I'm not talking about the process of forming H2, because there are many systems which can do that. The point is whenever you have such a system, it's sensitive to oxygen. So if you expose to it air, it drops dead. And you really need to do something about that. In biology, basically, three different evolutionary independent enzymes are known, which could do that and that is the modern iron hydrogenase, the di-iron hydrogenase and the nickel iron hydrogenase. Except for this system, for all the others, the problem is O2 inhibition. In the left -- for the system here on the left, you see here's the iron atom. The explanation is simple. It does not bind oxygen but it has also a bias to split the hydrogen. So it's not the most efficient at hydrogen producing catalyst. Well, if you want to study this, you need to consider different reactive oxygen species, three, four, five of them. You have different pathways, different spin coupling schemes, different charges need to be considered. We did that also with DFT and we considered more than 1,000 broken-symmetry DFT calculations. We don't know how accurate they are. The outcome was this. This is the most system in terms of hydrogen production. The central side of iron iron hydrogenase. FE, of course, means iron atom as a sulphur atom, and you see it has six iron sides. So four here and two in this sub cluster over here. It's actually the sub cluster which binds molecular oxygen from air, for instance, and then it produces reactive oxygen species, which are converted to an OOH radical and to this H2O2 oxidizing molecule, and these reactive oxygen species, then they attack the cluster and then it decomposes, and that's the reason why it drops dead. And from DFT, these are the energies for the process. We don't know how accurate that is. All we know that this is pretty much in line with what is experimentally seen. But we have no other independent way to test that. And for this, it would be nice to run calculations on a computer which could treat a hundred molecular orbitals. 26 Of course, if you understand that, then the next thing is you would increase your atomic model, say, to 700 atoms, because what you can do then is multi-genesis studies. So you can change the environment in order to make it more oxygen stable. Actually, this is something we are working also on, and in collaboration with experimental people who could do 600 mutants per day. Although they can do 600 mutants of the enzyme per day, the combinatorial possibilities here are so large that you really need theoretical calculations in order to guide you. >>: Sorry to interrupt, but my impression for this sort of work is that in some sense, the experimentalists feel like the calculations are almost good enough, right? You say -- you have a space of a gazillion, you give them, say, 10,000 things that look good to you. They make 60 mutants. Of those, 20 work, right? I hear theorists talk a lot about how it would be great if we could do something better. But the experimentalists are always like look, we've got these thousand things. We tried 600 of them. >> Markus Rieher: >>: It's true. So what do you tell your experimental -- >> Markus Rieher: Well, what you're describing is pretty much going on in the drug design industry, because the experimentalists don't care whether this or that proposition will be right or wrong, just to nail down the millions to a few thousand and test it experimentally. While I think when it comes to catalysis, you need to change the structure of the catalyst, and in order to do that he reliably, you really need a high accuracy, because it might be very hard to synthesize. It could take a year or two until you figure out how to do that. And when you're in such a situation, you really need to be able to trust your calculations and, well, DFT could work, but you never know. Okay. But this is as much about the two examples as I wanted to tell you. Just to give you an impression what a chemical problem really is. And, of course, there are also molecular properties we are interested in, and one property that we have spent some time on calculating is actually the spin density, because that was turned out to be very hard with standard methods. 27 Why is that so? Well, let me explain to you, this what you see here basically spin density distributions calculated with different standard methods all are CAS-SCF, calculations for a tri-atomic molecule, and you don't see the atoms. There's an iron atom here and there's a nitrogen atom here and an oxygen there. So it's FeNO. And the notation is that this seven means seven electrons and this other second seven means in seven molecular orbitals. And this means 11 electrons and nine molecular orbitals and so on. So this directly relates to the question how many molecular orbitals could you treat with the standard methods. The point is with the standard thought CAS-SCF, you can treat at most, I would say, 18 electrons and 18 orbitals. This is why you could treat a hundred orbitals, it would be great, or even 50. Here, you clearly have a problem, because you see, well, the point is this is a spin density and all these representations are spin density differences to be more accurate in my description, and you see that the spin density kind of oscillates around the referenced entity which I just picked as a CAS 11 and 14 reference. But we could go to other algorithms which are kind of not yet standard in chemistry, although they're quite standard in physics. And like the DMRG density matrix renormalization because with DMRG, we can really treat large active spaces. Here you see just one slide which should convince you that we can really converge a spin density within DMRG save for 13 electrons and 29 molecular orbitals, and then the reference spin density which we produce is this one. Now we can use that in order to compare with CAS-SCF and DFT just to see how wrong it is. This spin this spin is the comparison with CAS-SCF. So these are not the spin densities, but density differences. I forgot to tell you that -- let me go back. So is the spin density and blue means alpha spin excess and yellow means beta excess. These are spin density differences, and you see that the CAS-SCF calculations with this small number of active orbitals, 13, 14, 15, 16, is too much off. So this is the error in the spin density with one of the standard methods, CAS-SCF in quantum chemistry. 28 >>: What is the [indiscernible] 10 percent error, 20 percent error, 100 percent error? >> Markus Rieher: It's an iso surface, and I don't recall in absolute terms, but the error is too large. It's, of course, in the paper if you want to look it up. You can also take the difference to the DFT spin densities, and this is for eight standard density functions which are used in chemistry -- in quantum chemistry. And while they look like spin densities, but this is the error in DFT, okay? We can do that for larger systems. Before, I showed you tri atomic model system. Here we have a full-fledged transition metal complex. In order to describe the spin density here, you need to, for instance, 13 electrons and 42 orbitals. It also tells you if you can treat 50 molecular orbitals, meaning a hundred spin orbitals with respect to the spin density, this is the system you could study. The spin density is, of course, important for certain spectroscopic techniques in chemistry. Now, the final point I would like to discuss with you is, is there a way to choose the molecular orbitals from the huge set of molecular orbitals that you need to describe, that you need for the description of your molecules. Is there an automatic way to determine those orbitals which are relevant to the active space? And usually, that is done based on energetic criteria and mostly I would say, on chemical intuition. But recently, it turned out that concepts from quantum information theory are quite beneficial here. And the concept that we studied are the single orbital entropy and the mutual information. The single orbital entropy is computed from the eigenvalues, Omega of the reduced density matrix that we obtained when we trace out all the environment states with which a single orbital would interact. And then we end up with a four by four reduced density matrix and four eigenvalues and we can compute the single orbital entropy here. And in order to understand what the entanglement between two orbitals is, we could calculate the mutual information. These concepts, by the way, have been introduced by Legeza almost ten years ago, and the mutual information by Rissler, Noack and White in chemical physics 2006. So the mutual information is computed from the single orbital entropies and from the two orbital entropy, 29 which you get in the pretty same way by using the eigenvalues of the two orbital reduced entity matrix. That way, you can see the two orbitals explicitly and you trace out all the environmental orbitals. Now, we can study these concepts and it turns out that we usually see three subsets of orbitals with high entanglement, medium entanglement, and weak entanglement. Turns out that they pretty much match the chemical intuition. So in chemistry, people have used, for instance, if you have active electrons in pi orbitals, you should correlate them with anti-bonding so-called pi orbitals and so on. So we have kind of a measure in our hands that we tend to replace chemical intuition here, which is really good if you want to pick, say, a hundred orbitals from the set of 1,000 that I had on one of my slides. And while he could go in more detail here. But the pattern in all the molecules that we have studied so far is pretty much the same. You always have these three classes of orbitals when it comes to orbital interactions. And we have seen that also for larger systems. That's basically everything I wanted to tell you about. So my conclusion here is the following. I think it is possible to define important yet unsolved chemical problems that would benefit from the availability of a quantum computer if it could treat 100 molecular orbitals. If it's only 50 molecular orbitals, well, I don't know whether it's worth the effort. It's important to know that whenever you do such a quantum simulation, you must be able to beat all the standard methods that are around, like coupled cluster, CAS-SCF, perturbation on top of CAS-SCF, DMRG or DFT. And I was also asked in this email by Leeanne to put up some questions. These are a few questions that I would have, because some of it might be already answered by Matthias' slide. So if you have this problem, how fast could it be solved on the quantum computer. Would it be really more efficient than the consisting schemes? Would it be possible to include the, what we call the dynamic correlation effect? So the fact that we have omitted most of the orbitals in the calculation, actually, and how difficult would it be to compute molecular properties, response properties. And that's it. Thanks for your attention. >>: Any more questions? We have time for discussion now. Mike? 30 >>: So can you describe how the information concepts like entropy and mutual entropy can be used to select the most important orbitals to keep track of. But classically, when you're [indiscernible] degrees of freedom, do you settle on one of your combinations in the initial degrees of freedom and not just choosing the best ones from the original basis by some principal value method. For example [indiscernible] the largest principal values and keep those. So I'm wondering whether something analogous is possible. linear combinations of orbitals. Can you select >>: Anyway, taking combinations of orbitals and occupancies, I mean, actually don't move the electrons one at a time either. Suppose you could say, okay, there are these quantum mechanical transitions described by unitaries among this, let's call it a basis per minute of occupancy the if sets of orbitals. And what we seek is a combination of those guys that we can model. I have no idea if that with a work, but, I mean, that's the kind of -- that extends not only the basis to the linear algebra, but also gets around this pushing electron timetable would be different. >> Markus Rieher: There's a lot of literature and knowledge available on how to choose these one-particle states and the point -- I didn't talk about this, but I considered it not to be a problem, because it's already knowledge available. For instance, you could pick the molecular orbital that comes out of a [indiscernible] calculation. But that is not what we did here. What we did is we used so-called natural orbitals that we computed from a small CAS-SCF calculation. So we run a CAS-SCF calculation, which also optimizes the orbitals and then we pick that linear combination which diagonalizes the first density matrix. >> Krysta Svore: >> Markus Rieher: This is already done. That's why I didn't talk about it. >>: I'm confused. It seems like this procedure, choosing the orbitals requires you use the bigger system already. If you're going to choose the orbital this way, you've got to be able to trace out everything else. So it seems like you've already solved the system here. >> Markus Rieher: That's right. The point is what we did here is we ran small DMRG calculations, small in a sense that it was a limited number of iterations. 31 Now, I think it's not a principal problem that you get this information posteriori. You can come up with a scheme to sample that knowledge in advance, I think. But in any case, it would be good to have such a measure which tells you whether you picked the right orbitals. Even if, for instance, this can be done only after the quantum computation. I don't think that's true. You could get the knowledge before with standard methods. >>: So [indiscernible] try to get it up to 50 is very hard to find. >> Markus Rieher: If you want to be competitive. guess Garnett agrees, don't you? >>: I don't know whether -- I Yeah, yeah. >>: I mean, in the knowledge that you are saying, these are the only -- these are only the [indiscernible] orbitals underlying the thousands of [inaudible]. >>: So that's the scale that we have to [indiscernible]. >> Markus Rieher: Come it comes to chemistry. >>: So you identify the [indiscernible]. But what about do you [indiscernible] techniques and measurement techniques these things more directly. [indiscernible] computation on the theoretical end but think about the experimental techniques that are developing later will be easier to use say, tricks that would let you [indiscernible] five years as an example of that. >> Markus Rieher: So I don't get your point. So in order to solve the chemical problem or to replace kind of the universal quantum computer by a system which just simulates directly. >>: I'm not thinking about simulation. I'm thinking about experiment now. Take techniques in experimental quantum conversation, use the more direct [indiscernible]. I guess ->>: Can I ask a question here? 32 >>: Sure. >>: So take these biological [indiscernible]. So we know people do these DFT calculations, a couple of things don't happen. One is they don't fold the right way so there's a huge hole and all this water comes in, kills the catalyst. The other thing which goes wrong is somehow, in some sense, the [indiscernible] activity of the active site is reduced for some reason. There's some [indiscernible] or whatever that screwed it up. So I think, so we -- what do you see as like the chemical, like the physical properties of these things that make them good catalysts? Like what do I need to worry about? Do I need to worry about readouts potential? Do I need to worry about the [indiscernible] of the surrounding? >> Markus Rieher: The point is that you need to figure out. >>: So Peter's question is can we use our abilities to use coherent spectroscopy as a way to, like, test ->> Markus Rieher: How would you do that? I mean, maybe I don't get you right, but the point is, in theory, you could study a system which is not yet experimentally. So how would you do that with an experiment? You see what I mean? >>: Yeah, but I think -- >>: Let me ask the question another way, which is if you could order experiments [indiscernible], what experiments would you ask them to do? things would you ask them to measure to best inform [indiscernible]. What >> Markus Rieher: To figure out whether the calculations are correct. Well, actually, what you really need in the first place is structures. So I don't think that you really need new experimental techniques. The point is it's very difficult to get the structures, because you need to, well, if you use x-ray defraction, you need a crystal. Difficult to get for certain systems. You could use more indirect techniques and, of course, that's why I basically had this point here, molecular properties. Of course, with standard techniques, we can compute molecular properties. For instance, whether you have N2 binding or not, vibrational spectroscopy is sufficient because it tells 33 you where the vibration is stretching vibration of N2 is. It's perfect. So I don't think that there are new techniques in spectroscopy that we really need. It's these systems are very difficult to treat experimentally, and that is basically a chance for us, in theory to really compete on an equal level with them. >>: Okay. Thank you.