quantum simulation of molecular collisions with ...mgeller/pritchett.pdfa molecular collision with a...

4
Quantum Simulation of Molecular Collisions with Superconducting Qubits Emily J. Pritchett, 1 Colin Benjamin, 1, 2 Andrei Galiautdinov, 1 Michael R. Geller, 1 Andrew T. Sornborger, 3 Phillip C. Stancil, 1, 2 and John M. Martinis 4 1 Department of Physics and Astronomy, University of Georgia, Athens, GA 30602 2 Center for Simulational Physics, University of Georgia, Athens, GA 30602 3 Department of Mathematics and Faculty of Engineering, University of Georgia, Athens, GA 30602 4 Department of Physics, University of California at Santa Barbara, Santa Barbara, California 93106 (Dated: July 7, 2010) We introduce a protocol for the fast simulation of n-dimensional quantum systems on n-qubit quantum computers with tunable couplings. A mapping is given between the control parameters of the quantum computer and the matrix elements of Hs (t), an arbitrary, real, time-dependent n × n dimensional Hamiltonian that is simulated in the n-dimensional ‘single excitation’ subspace of the quantum computer. A time-dependent energy/time rescaling minimizes the simulation time on hardware having a fixed coherence time. We demonstrate how three tunably coupled phase qubits simulate a three-channel molecular collision using this protocol, then study the simulation’s fidelity as a function of total simulation time. A quantum computer can significantly reduce the re- sources necessary to simulate quantum mechanical sys- tems [1]. Typically, quantum simulation algorithms al- gorithms construct the simulated system’s time evolu- tion operator, energies and/or eigenstates from a uni- versal set of gates [2–8]. Alternatively, ultracold atoms, trapped ions, and liquid-state NMR have directly emu- lated the time evolution of certain other quantum sys- tems [9–11]. Recent experimental progress suggests that quantum simulation will be one of the first practical ap- plications of quantum computation [7–12]. In principle, an n-qubit quantum computer can store the state of any N =2 n dimensional quantum system, an exponential reduction in the resources necessary to store quantum information on a classical computer. However, simulation may require N 2 =2 2n elementary gates per time step unless the simulated Hamiltonian has special properties, e.g. locality [2, 13]. Even for these special Hamiltonians, fully digital quantum simulation often re- quires an excessive number of gates for current quantum computing technology [4, 6]. In this Letter, we show that a subspace of a tunable n-qubit quantum computer can emulate an arbitrary n- dimensional quantum system, trading an exponential re- duction in resources for simulations of a wider variety of Hamiltonians. This subspace simulates other quan- tum systems very different from the computer itself in an amount of time that is independent of n. By comparison, classical simulation of an n-dimensional quantum system requires n 3 elementary operations per time step. While the most efficient quantum simulation algorithms offer an exponential reduction in both qubits and elementary operations, they typically apply to specific, fundamental time-independent Hamiltonians, or those already simi- lar to that of the computer itself. We show that with a more modest polynomial reduction in resources, a sub- space of a tunable quantum computer can simulate any real, time-dependent Hamiltonian. We begin by outlining the theory behind our approach to simulation. First, we identify an n-dimensional invari- ant subspace suitable for quantum simulation. Then we define a time dependent energy/time rescaling that max- imizes the speed of the simulation within the constraints of the quantum computer. Finally, the control parame- ters of the quantum computer are given explicitly as a function of the matrix elements of H s (t). Our approach is tested by performing a simulation of a molecular collision with a circuit of tunably coupled Josephson phase qubits. Molecular collisions and elec- tronic structure calculations are widely studied as im- portant applications of quantum simulation techniques [6–8]. We show in detail how a superconducting circuit of three tunably coupled Josephson phase qubits simulates a three channel Na-He collision. Finally, we discuss the relationship between simulation fidelity and total simu- lation time for this particular example. An n-Dimensional Subspace of the full quantum com- puter’s Hilbert space, H, can emulate another quantum system at all times only if it is invariant to the time evo- lution generated by the computer’s Hamiltonian H qc (so that the subspace is well-isolated from the rest of H and evolves unitarily). We model H qc as H qc (t)= n X i=1 - i (t) 2 σ z i + 1 2 X i6=j g ij (t)J μν σ μ i σ ν j , (1) where i (t) are the uncoupled qubit energies, g ij (t)= g ji (t) are the pairwise qubit interaction strengths, J μν gives the relative size of the σ μ i σ ν j interaction, and μ, ν ∈{0,x,y,z} are summed over. While i (t) and g ij (t) may in general be time-dependent, the time-independent structure of qubit interaction is specified by J μν , a di- mensionless tensor that is typically fixed by a given ar- chitecture and is identical between each pair of qubits. In the weak coupling limit, |g ij |||J μν ||/ i 1, subspaces of H are invariant to time evolution generated by H qc if

Upload: others

Post on 24-May-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Quantum Simulation of Molecular Collisions with ...mgeller/pritchett.pdfa molecular collision with a circuit of tunably coupled Josephson phase qubits. Molecular collisions and elec-tronic

Quantum Simulation of Molecular Collisions with Superconducting Qubits

Emily J. Pritchett,1 Colin Benjamin,1, 2 Andrei Galiautdinov,1 Michael R.

Geller,1 Andrew T. Sornborger,3 Phillip C. Stancil,1, 2 and John M. Martinis4

1Department of Physics and Astronomy, University of Georgia, Athens, GA 306022Center for Simulational Physics, University of Georgia, Athens, GA 30602

3Department of Mathematics and Faculty of Engineering, University of Georgia, Athens, GA 306024Department of Physics, University of California at Santa Barbara, Santa Barbara, California 93106

(Dated: July 7, 2010)

We introduce a protocol for the fast simulation of n-dimensional quantum systems on n-qubitquantum computers with tunable couplings. A mapping is given between the control parametersof the quantum computer and the matrix elements of Hs(t), an arbitrary, real, time-dependentn×n dimensional Hamiltonian that is simulated in the n-dimensional ‘single excitation’ subspace ofthe quantum computer. A time-dependent energy/time rescaling minimizes the simulation time onhardware having a fixed coherence time. We demonstrate how three tunably coupled phase qubitssimulate a three-channel molecular collision using this protocol, then study the simulation’s fidelityas a function of total simulation time.

A quantum computer can significantly reduce the re-sources necessary to simulate quantum mechanical sys-tems [1]. Typically, quantum simulation algorithms al-gorithms construct the simulated system’s time evolu-tion operator, energies and/or eigenstates from a uni-versal set of gates [2–8]. Alternatively, ultracold atoms,trapped ions, and liquid-state NMR have directly emu-lated the time evolution of certain other quantum sys-tems [9–11]. Recent experimental progress suggests thatquantum simulation will be one of the first practical ap-plications of quantum computation [7–12].

In principle, an n-qubit quantum computer can storethe state of any N = 2n dimensional quantum system, anexponential reduction in the resources necessary to storequantum information on a classical computer. However,simulation may require ∼ N2 = 22n elementary gates pertime step unless the simulated Hamiltonian has specialproperties, e.g. locality [2, 13]. Even for these specialHamiltonians, fully digital quantum simulation often re-quires an excessive number of gates for current quantumcomputing technology [4, 6].

In this Letter, we show that a subspace of a tunablen-qubit quantum computer can emulate an arbitrary n-dimensional quantum system, trading an exponential re-duction in resources for simulations of a wider varietyof Hamiltonians. This subspace simulates other quan-tum systems very different from the computer itself in anamount of time that is independent of n. By comparison,classical simulation of an n-dimensional quantum systemrequires∼ n3 elementary operations per time step. Whilethe most efficient quantum simulation algorithms offeran exponential reduction in both qubits and elementaryoperations, they typically apply to specific, fundamentaltime-independent Hamiltonians, or those already simi-lar to that of the computer itself. We show that with amore modest polynomial reduction in resources, a sub-space of a tunable quantum computer can simulate anyreal, time-dependent Hamiltonian.

We begin by outlining the theory behind our approachto simulation. First, we identify an n-dimensional invari-ant subspace suitable for quantum simulation. Then wedefine a time dependent energy/time rescaling that max-imizes the speed of the simulation within the constraintsof the quantum computer. Finally, the control parame-ters of the quantum computer are given explicitly as afunction of the matrix elements of Hs(t).

Our approach is tested by performing a simulation ofa molecular collision with a circuit of tunably coupledJosephson phase qubits. Molecular collisions and elec-tronic structure calculations are widely studied as im-portant applications of quantum simulation techniques[6–8]. We show in detail how a superconducting circuit ofthree tunably coupled Josephson phase qubits simulatesa three channel Na-He collision. Finally, we discuss therelationship between simulation fidelity and total simu-lation time for this particular example.

An n-Dimensional Subspace of the full quantum com-puter’s Hilbert space, H, can emulate another quantumsystem at all times only if it is invariant to the time evo-lution generated by the computer’s Hamiltonian Hqc (sothat the subspace is well-isolated from the rest of H andevolves unitarily). We model Hqc as

Hqc(t) =

n∑i=1

−εi(t)2

σzi +1

2

∑i6=j

gij(t)Jµνσµi ⊗ σ

νj , (1)

where εi(t) are the uncoupled qubit energies, gij(t) =gji(t) are the pairwise qubit interaction strengths, Jµνgives the relative size of the σµi ⊗ σνj interaction, andµ, ν ∈ {0, x, y, z} are summed over. While εi(t) and gij(t)may in general be time-dependent, the time-independentstructure of qubit interaction is specified by Jµν , a di-mensionless tensor that is typically fixed by a given ar-chitecture and is identical between each pair of qubits.In the weak coupling limit, |gij |||Jµν ||/εi � 1, subspacesof H are invariant to time evolution generated by Hqc if

Page 2: Quantum Simulation of Molecular Collisions with ...mgeller/pritchett.pdfa molecular collision with a circuit of tunably coupled Josephson phase qubits. Molecular collisions and elec-tronic

2

spanned by computational basis states having the samenumber of excited (tunable) qubits. The ‘single excita-tion subspace’, denoted as Hn, is an n-dimensional in-variant subspace spanned by |i

⟩n≡ |00..01i..0n

⟩for all

i = 1, 2..., n.The control parameters εi(t) and gij(t) directly control

the Hamiltonian thatHn simulates. We define Hn as Hqc

projected into the single excitation subspace,

Hn(t) ≡ PHqc(t)P† (2)

where P is an n×2n dimensional operator that projectsHonto Hn. Up to an additive energy shift, Hn has matrixelements

Hijn (t) ≡

{εi(t)− α

∑k 6=i gik(t), i = j

gij(t), i 6= j(3)

with α ≡ 2(Jzo + Jzz). We assume Jxx + Jyy 6= 0 andnormalize Jµν so that Jxx+Jyy = 1. In the weak couplinglimit, Hn is approximately invariant and generated byHn:

Un(t) ≡ PUqc(t)P†

' T e− i~∫ t0Hn(t

′)dt′ (4)

where T is the time-ordering operator. Hn generates Unexactly when no matrix elements of Hqc mix Hn with therest of H (i.e. J0x = J0y = Jzx = Jzy = 0). The (n2 +n)/2 parameters εi(t) and gij(t) independently controleach of the (n2 + n)/2 matrix elements of the real Hn

and can therefore be used to simulate any arbitrary, realHamiltonian in Hn.

While we can simulate Hs in Hn by choosing εi(t) andgij(t) so that Hn(t) = Hs(t) for all t, a direct mappingbetween Hamiltonians limits the computer to simulatingother quantum systems with similar energy scales overlengths of time within the computer’s coherence time.Fortunately, simulation of Hs only requires equality upto an overall phase between Un and the time evolutionoperator generated by Hs:

U(t) ≡ T e−i~∫ ttiHs(t

′)dt′

= eiφ(t)Un(tqc(t)). (5)

The time elapsed on the quantum computer, tqc(t), is astrictly increasing function of simulated time t, admit-ting a much less restrictive relationship between Hamil-tonians:

Hs(t) + c(t) = λ(t)Hn(tqc(t)). (6)

c(t) is a time-dependent, additive energy shift giving the

overall phase difference φ(t) = 1~∫ ttic(t′)dt′, and we have

introduced a positive, time-dependent energy/time scal-ing

λ(t) ≡ dtqc/dt. (7)

The energy/time scaling λ(t) determines the speed ofthe simulation. By carefully minimizing λ(t), we reducethe total simulation time and, consequently, the errordue to decoherence. λ(t) is bounded from below by ex-perimental constraints on the allowed values of controlparameters εi(t) and gij(t) as well as their maximumrates of change. Suppose qubit interaction strengths canvary in a range gij(t) ∈ [−gmax, gmax], and the uncoupledqubit energies can vary in a range εi(t) ∈ [εmin, εmax]. Forconvenience, we define a simulated energy Ei(t) anal-ogous to εi(t) when diagonal contributions from qubitinteractions are anticipated:

Ei(t) ≡ Hiis (t) + α

∑j 6=i

Hijs (t). (8)

Using this definition together with equations (3) and (6),we relate the control parameters of the quantum com-puter to the simulated energies in Hs(t):

gij(t) = Hijs (t)/λ(t)

εi(t) = [Ei(t)− c(t)]/λ(t). (9)

By choosing c(t) = Emax(t)− λ(t)εmax where Emax(t) isthe largest value obtained by the Ej(t) at a particular t,we force each εi to be as large as possible and thereforeminimize leakage out of Hn.

Each of the computer’s control parameters remainswithin its allowed range when λ(t) is larger than (n2 +n)/2 energy ratios at all times:

λ(t) ≥

{|Hij

s (t)|/gmax, i 6= j

∆Ei(t)/∆εmax

(10)

where ∆Ei(t) ≡ Emax(t) − Ei(t) and ∆ε ≡ εmax − εmin.λ(t) is also bounded by constraints on the speeds withwhich control parameters can change. Suppose vεi (tqc) ≡dεi(tqc)/dtqc and vgij(tqc) ≡ dgij(tqc)/dtqc can never belarger in magnitude than vεmax and vgmax respectively.Then for all t,

vgmax ≥1

λ2

∣∣∣∣dHijs (t)

dt− Hij

s (t)

λ

dt

∣∣∣∣ (11)

(and similarly for vεmax).To simulate Hs(t) in Hn, we first choose λ(t) as small

as both inequalities (10) and (11) allow, guaranteeing afast simulation within the experimental constraints of thequantum computer. We integrate over λ(t) to calculatetqc as a function of t:

tqc(t) =

∫ t

ti

λ(t′)dt′ + tqc(ti). (12)

With both λ(t) and tqc(t) known, we can explicitly mapthe matrix elements of Hs to the control parameters ofthe quantum computer:

εi(tqc(t)) = εmax + ∆Ei(t)/λ(t)

gij(tqc(t)) = Hijs (t)/λ(t). (13)

Page 3: Quantum Simulation of Molecular Collisions with ...mgeller/pritchett.pdfa molecular collision with a circuit of tunably coupled Josephson phase qubits. Molecular collisions and elec-tronic

3

To demonstrate our theory in detail, we describe threeJosephson phase qubits simulating a three-channel col-lision between a sodium and a helium atom. For threephase qubits with tunable inductive coupling,

Hqc(t) =

3∑i=1

−εi(t)2

σzi +1

2

∑i6=j

gij(t)Φ̂i ⊗ Φ̂j (14)

where Φ̂i is defined in terms of the matrix elementsϕjk =

⟨j|ϕ̂i|k

⟩of the local Josephson phase operator

in the computational basis of the ith qubit:

Φ̂i ≡ σxi +ϕ00 − ϕ11

2ϕ01σzi +

ϕ11 + ϕ00

2ϕ01σ0i . (15)

Both the εi and the ϕjk depend on Φx, the externallyapplied flux through the superconducting circuit. Ex-ternal flux bias is quantified by a dimensionless param-eter si(t) = Φx/Φ

∗x where Φ∗x is the qubit’s critical flux

bias, or alternatively, by the dimensionless well depth∆U/~ωp [14]. We consider external bias values for whichs ∈ [.89, .90] and ∆U/~ωp ∈ [13.7, 15.5]. In this range,

∆ε/h = 190MHz while Φ̂i ' σ1i + 11σ0

i varies little. Atunable mutual inductance independently controls thecouplings gij(t) between each pair of qubits. We haveassumed Josephson junction parameters I0 = 2.93 µA,C = 1.52 pF, and L = 808 pH.

An n-dimensional subspace can simulate a molecularcollision only after we project the full, many-body Hamil-tonian of the interacting electrons and nuclei into ann-dimensional basis. We construct the collision Hamil-tonian from Born-Oppenheimer energies and nonadia-batic couplings calculated previously for three molecu-lar channels: Na(3s) + He(1s2) [1 2Σ+] and Na(3p) +He(1s2)[1 2Π; 2 2Σ+] [15], labeled as |1

⟩s, |2⟩s

and |3⟩s

respectively. The energies are stored for fixed valuesof the internuclear distance R, which we assume takesstraight-line trajectories in a standard semiclassical ap-proximation: R(t) =

√b2 + v2t2 where v is the incoming

particle’s velocity and b is the impact parameter of thecollision.

Figure 1 outlines our simulation protocol for Hs(t) de-scribing a three-channel Na-He collision. The matrix el-ements of Hs(t) are displayed in Fig. 1(a) for a givensemiclassical trajectory R(t). Directly below, we plotthe energy/time scaling parameter λ(t) as a black curveenveloping the six energy ratios given in Eq. (10). Asmall λ(t) speeds the quantum computer through timeswhen the internuclear distance R is large, but as R de-creases (t → 0), a relatively small gmax value constrainsthe growing couplings. λ(t) increases over two orders ofmagnitude, creating a highly nonlinear relationship be-tween tqc and t, as shown in Fig. 1(c). This effectivelystretches the portion of the collision when internucleardistance is small over the entire simulation, as can beseen in the plot of the quantum computer’s control pa-rameters as a function of tqc in Fig. 1(d).

FIG. 1: (color online) Hs(t) describes a three channelNa-He collision with b = 0.5 and v = 1.0. (a) Ma-trix elements of Hs as a function of time in atomicunits (Eh = 27.21 eV and the atomic unit of time is2.419 × 10−8 ns). (b) The dimensionless time scalingparameter λ(t) envelopes six energy ratios (∆E3 = 0 forall t). We assume gmax/h = 2.0 MHz and ∆εmax/h =190 MHz. (c) Plot of tqc(t) for the case of tqc(ti) = 0,ti = −40 a.u.. (d) Control parameters that simulateHs(t) plotted as a function of tqc (ε3 = εmax for all tqc).

To study the fidelity of the simulation, we compare theexact and simulated time evolution operators, U(t) andUn(tqc(t)) respectively, by plotting (in Fig. (2)) transi-tion probabilities out of |1

⟩:

P1i(t) ≡ |⟨i|U(t)|1

⟩|2. (16)

Because the exact transition probabilities evolve differ-ently with t than the simulated evolve with tqc, we de-fine a time-dependent transition fidelity which accountsfor time scaling,

F (t) ≡ |s

⟨1|U†(t)Un(tqc(t))|1

⟩n|2, (17)

and a time-dependent leakage out of Hn,

L(t) ≡∑

⊥|⊥⟨i|Uqc(tqc(t))|1

⟩n|2 (18)

Page 4: Quantum Simulation of Molecular Collisions with ...mgeller/pritchett.pdfa molecular collision with a circuit of tunably coupled Josephson phase qubits. Molecular collisions and elec-tronic

4

FIG. 2: (a) (color online) Exact transition probabilitiesgenerated by Hs(t) shown in Fig. 1(a). (b) Transitionprobabilities simulated with parameter profiles given inFig. 1(d). Final simulation fidelity is 0.998.

where∑⊥ is the sum over all computational basis states

|i⟩⊥ orthogonal to Hn. In the upper part of Fig. (3),

fidelity and leakage are plotted together for four differentgmax values.

Minimizing gmax||Jµν ||/εmin, either by decreasing gmax

or by increasing εmin, reduces leakage and thus improvessimulation fidelity. In this example, we find simulationfidelity more sensitive to the cutoff in gmax because leak-age is most prominent when the interatomic distances aresmall (t → 0) and the diabatic couplings between chan-nels are the dominant terms. By reducing gmax we alsoincrease λ(t) and thus the total simulation time, as stud-ied in the lower plot of Fig. (3). To increase fidelity from.9990 to .9999 we need to increase the simulation timeby a factor of ∼ 3, a relationship that is independentof n. While not introducing specific models of decoher-ence, we note that high fidelity simulations are possibleon superconducting qubits with coherence times around100 ns.

When applied to molecular collisions, our approach toquantum simulation requires classical overhead to projectthe fundamental, time-independent, many-body Hamil-tonian into an R-dependent, n-channel Hs. The quanti-ties of physical interest, cross sections, are obtained byintegrating the final transition probabilities over manysemiclassical trajectories with different impact parame-ters, which requires no further classical overhead. A clas-sical simulation of transition probabilities requires ∼ n3

elementary operations per time step for a single impactparameter, thus cross section calculations are computa-tionally intensive for large n. Alternatively, simulationtime is independent of n using our protocol, so once theR-dependent Hs has been calculated, cross sections canbe obtained quickly.

In summary, we have presented a straightforward pro-tocol for quantum simulation that can be implementedwith currently available superconducting quantum com-puting technology. While a promising application ofquantum computation, current quantum simulation pro-tocols require a threshold number of gates and qubits

FIG. 3: (color online) (a) Fidelity and leakage as afunction of simulation time for four different gmax val-ues, with all other parameters the same as in Fig. 1.(b) Final simulation fidelity versus total simulationtime for varying gmax. The gmax value is referenced bythe shade of the data point.

that prohibits fully digital quantum simulations from be-ing demonstrated on available quantum computers. How-ever, we have shown how quantum computers of only afew qubits can simulate arbitrary quantum systems ac-curately and quickly even before they reach the regimeof fault tolerant quantum computation.

It is a pleasure to thank Joydip Ghosh for interest-ing discussions. This work was partially supported byNSF grants PHYS-0939849 and PHYS-0939853 from thePhysics at the Information Frontier Program.

[1] R.P. Feynman, Int. J. Theor. Phys. 21, 467 (1982).[2] S. Lloyd, Science 273, 5273 (1996).; J. Dodd, et al.,

Phys. Rev. A. 65 040301 (2002).[3] S. Wiesner, quant-ph/9603028 (1996).; C. Zalka, Proc. R.

Soc. London Sect. A 454, 313-322 (1998).; B. Boghosisanand W. Taylor, quant-ph/9701019 (1997).; D.S. Abramsand S. Lloyd, Phys. Rev. Lett. 79, 2586 (1997).; L.-A.Wu, et al., Phys. Rev. Lett. 89, 057904 (2002).

[4] K. Brown, et al., Phys. Rev. Lett. 97, 050504 (2005).[5] A. T. Sornborger and E. D. Stewart, Phys. Rev. A 60,

1956 (1999).[6] I. Kassal, et al., PNAS 105, 18681 (2008); A. Aspuru-

Guzik, et al., Science 309, 1704 (2005).[7] B. P. Lanyon, et al., Nature Chemistry 2, 106 (2010).[8] J. Du, et al., Phys. Rev. Lett. 104, 030502 (2010).[9] M. Greiner, et al., Nature 415, 39 (2002).; S. Trotzky, et

al. Science 319, 295 (2008).[10] D. Leibfried, et al., Phys. Rev. Lett. 89, 247901 (2002);

A. Friedenauer,et al., Nat. Phys. 4, 757 (2008).[11] S. Somaroo, et al., Phys. Rev. Lett. 82, 5381 (1999).[12] I. Buluta and F. Nori, Science 326, 108 (2009).[13] A. Barenco, et al. , Phys. Rev. A 52, 34573467 (1995).[14] J. M. Martinis, et al., (to be published); J. Clarke and

F.K. Wilhelm, Nature 453, 1031 (2008).[15] C. Y. Lin, et al., Phys. Rev. A 78, 052706 (2008).