Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect...

11
Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1 Jelmer J. Renema, 1 Andreas Eckstein, 1 Antonio A. Valido, 2 Adriana Lita, 3 Thomas Gerrits, 3 Sae Woo Nam, 3 W. Steven Kolthammer, 1 Joonsuk Huh, 4 and Ian A. Walmsley 1 1 Clarendon Laboratory, Department of Physics, University of Oxford, Oxford OX1 3PU, UK 2 QOLS, Blackett Laboratory, Imperial College London, London, SW7 2AZ, UK 3 National Institute of Standards and Technology, 325 Broadway, Boulder, CO 80305, USA 4 Department of Chemistry, Sungkyunkwan University, Suwon 440-746, Korea (Dated: March 8, 2018) We study the impact of experimental imperfections on a recently proposed protocol for per- forming quantum simulations of vibronic spectroscopy. Specifically, we propose a method for quantifying the impact of these imperfections, optimizing an experiment to account for them, and benchmarking the results against a classical simulation method. We illustrate our findings using a proof of principle experimental simulation of part of the vibronic spectrum of tropolone. Our findings will inform the design of future experiments aiming to simulate the spectra of large molecules beyond the reach of current classical computers. Contribution of NIST, an agency of the U.S. government, not subject to copyright. INTRODUCTION Quantum chemistry is expected to benefit greatly from the development of quantum simulators and quantum computers, because the ability of classical computers to simulate quantum mechanical processes is severely lim- ited. For example, the calculation of molecular energies can in principle be done on a quantum computer involv- ing relatively few qubits [1]. It has recently been shown that the estimation of molecular vibronic spectra can in principle be done us- ing a quantum optics simulator [2, 3], whereas there is no known efficient classical algorithm for this task. Vibronic spectra, arising from simultaneous electronic and vibra- tional transitions in molecules, play an important role in determining the optical and chemical properties of those molecules. In addition to being useful for fundamental research in molecular physics and chemistry, calculating these spectra helps in assessing the performance of differ- ent molecules for applications in photovoltaics [4], biology [5], and other forms of industry [6]. The protocol for estimating vibronic spectra is consid- erably simpler than other quantum simulation protocols which require particle interactions or even full quantum computing. Since the vibrational modes of a molecule can be approximated as quantum harmonic oscillators, vibrational transitions can be mapped onto standard op- erations on quantum harmonic oscillators [7]. An exper- iment that implements simple transformations such as displacements, squeezing, rotations, and measurements in the Fock basis is sufficient to simulate this physics and therefore reconstruct the vibronic spectrum of a molecule. This protocol scales linearly in the number of vibrational modes to be simulated, does not require any post-selection, and can be implemented with read- ily available tools in several platforms. It is inspired by the boson sampling protocol [8] which has been demon- strated experimentally [9–11]. Various experimental platforms that make use of quan- tum harmonic oscillators have been used to simulate vi- bronic spectra, as shown by recent experiments using su- perconducting devices [12] and trapped ions [13]. More- over, the original theoretical proposal [2] suggests the use of quantum optics, in which each mode of the electromag- netic field is modelled as a quantum harmonic oscillator. The choice of platform imposes practical limitations on what can be achieved. Quantum optics is promising due to the availability of good sources of squeezing [14] and the large number of modes which can be manipulated, interfered, and measured [15, 16]. However, any experimental implementation of a quan- tum algorithm on a platform that does not have fault- tolerant architecture is necessarily degraded by imper- fections in the system operations. This is a potential limitation to the performance of all specialized quantum processors. In optical platforms, scattering and absorp- tion losses, mode mismatches, detector noise and unan- ticipated correlations may all contribute to less than ideal operation. The presence of experimental imperfections in any platform was not considered in the original proposal. These imperfections can be expected to affect the simu- lation, possibly reducing both the accuracy and precision of the results. In this work, we explore the impact of these imperfec- tions on a quantum optical simulation of vibronic spec- troscopy. We first describe in more detail the analogy between vibronic spectra and quantum optics using a specific transition in tropolone as an example. We then introduce the Gaussian state formalism, which we show can be used to quantify the impact of imperfections. Us- ing this formalism, we propose a method for adapting an experimental setup to account for its imperfections. We arXiv:1710.08655v3 [quant-ph] 7 Mar 2018

Transcript of Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect...

Page 1: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

Approximating Vibronic Spectroscopy with Imperfect Quantum Optics

William R. Clements,1 Jelmer J. Renema,1 Andreas Eckstein,1 Antonio A. Valido,2 Adriana Lita,3

Thomas Gerrits,3 Sae Woo Nam,3 W. Steven Kolthammer,1 Joonsuk Huh,4 and Ian A. Walmsley1

1Clarendon Laboratory, Department of Physics, University of Oxford, Oxford OX1 3PU, UK2QOLS, Blackett Laboratory, Imperial College London, London, SW7 2AZ, UK

3National Institute of Standards and Technology, 325 Broadway, Boulder, CO 80305, USA4Department of Chemistry, Sungkyunkwan University, Suwon 440-746, Korea

(Dated: March 8, 2018)

We study the impact of experimental imperfections on a recently proposed protocol for per-forming quantum simulations of vibronic spectroscopy. Specifically, we propose a method forquantifying the impact of these imperfections, optimizing an experiment to account for them,and benchmarking the results against a classical simulation method. We illustrate our findingsusing a proof of principle experimental simulation of part of the vibronic spectrum of tropolone.Our findings will inform the design of future experiments aiming to simulate the spectra of largemolecules beyond the reach of current classical computers.

Contribution of NIST, an agency of the U.S. government, not subject to copyright.

INTRODUCTION

Quantum chemistry is expected to benefit greatly fromthe development of quantum simulators and quantumcomputers, because the ability of classical computers tosimulate quantum mechanical processes is severely lim-ited. For example, the calculation of molecular energiescan in principle be done on a quantum computer involv-ing relatively few qubits [1].

It has recently been shown that the estimation ofmolecular vibronic spectra can in principle be done us-ing a quantum optics simulator [2, 3], whereas there is noknown efficient classical algorithm for this task. Vibronicspectra, arising from simultaneous electronic and vibra-tional transitions in molecules, play an important role indetermining the optical and chemical properties of thosemolecules. In addition to being useful for fundamentalresearch in molecular physics and chemistry, calculatingthese spectra helps in assessing the performance of differ-ent molecules for applications in photovoltaics [4], biology[5], and other forms of industry [6].

The protocol for estimating vibronic spectra is consid-erably simpler than other quantum simulation protocolswhich require particle interactions or even full quantumcomputing. Since the vibrational modes of a moleculecan be approximated as quantum harmonic oscillators,vibrational transitions can be mapped onto standard op-erations on quantum harmonic oscillators [7]. An exper-iment that implements simple transformations such asdisplacements, squeezing, rotations, and measurementsin the Fock basis is sufficient to simulate this physicsand therefore reconstruct the vibronic spectrum of amolecule. This protocol scales linearly in the numberof vibrational modes to be simulated, does not requireany post-selection, and can be implemented with read-ily available tools in several platforms. It is inspired by

the boson sampling protocol [8] which has been demon-strated experimentally [9–11].

Various experimental platforms that make use of quan-tum harmonic oscillators have been used to simulate vi-bronic spectra, as shown by recent experiments using su-perconducting devices [12] and trapped ions [13]. More-over, the original theoretical proposal [2] suggests the useof quantum optics, in which each mode of the electromag-netic field is modelled as a quantum harmonic oscillator.The choice of platform imposes practical limitations onwhat can be achieved. Quantum optics is promising dueto the availability of good sources of squeezing [14] andthe large number of modes which can be manipulated,interfered, and measured [15, 16].

However, any experimental implementation of a quan-tum algorithm on a platform that does not have fault-tolerant architecture is necessarily degraded by imper-fections in the system operations. This is a potentiallimitation to the performance of all specialized quantumprocessors. In optical platforms, scattering and absorp-tion losses, mode mismatches, detector noise and unan-ticipated correlations may all contribute to less than idealoperation. The presence of experimental imperfections inany platform was not considered in the original proposal.These imperfections can be expected to affect the simu-lation, possibly reducing both the accuracy and precisionof the results.

In this work, we explore the impact of these imperfec-tions on a quantum optical simulation of vibronic spec-troscopy. We first describe in more detail the analogybetween vibronic spectra and quantum optics using aspecific transition in tropolone as an example. We thenintroduce the Gaussian state formalism, which we showcan be used to quantify the impact of imperfections. Us-ing this formalism, we propose a method for adapting anexperimental setup to account for its imperfections. We

arX

iv:1

710.

0865

5v3

[qu

ant-

ph]

7 M

ar 2

018

Page 2: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

2

FIG. 1. Overview of the scheme for estimating vibronic spectra. A) A vibronic transition in a molecule such as tropolone(pictured, top) consists of a joint electronic and vibrational excitation. Depending on the energy of the absorbed photon,different vibrational states are excited, leading to complex spectra (bottom). The heights of the peaks depend on the overlapsbetween the ground state of the molecule and the excited vibrational states of the excited molecule. B) We model vibronictransitions using a harmonic approximation of the vibrational modes. The harmonic oscillators describing the excited state(in blue) are squeezed by S and displaced by D with respect to the ground state (in red). The overlaps between the differentFock states determine the heights of the spectral peaks. C) We simulate this process using a quantum optics experiment withsqueezing S and displacement D (top). Each optical mode is mapped onto a vibrational mode of the molecule; in our case weconsider two coupled vibrational modes of tropolone. The probabilities of measuring photon number outcomes using photonnumber resolving detectors (PNRD) are mapped onto the heights of the peaks in the spectrum (bottom).

also introduce a classicality criterion that can be used tobenchmark the performance of an imperfect simulation.Next, we perform a proof of principle experiment in whichwe simulate part of the vibronic spectrum of tropolone.This experiment highlights the impact of imperfectionsand illustrates our method for accounting for them. Fi-nally, we discuss our experimental results in light of ouranalysis.

THE 370 NM TRANSITION IN TROPOLONE

We first illustrate the connection between vibronicspectrocoscopy and quantum optics using the exampleof tropolone (C7H6O2), which is a molecule contributingto the taste and color of black tea [17] (see Fig. 1). Inthe 370 nm electronic transition in tropolone, the changein molecular configuration caused by the electronic exci-tation distorts the vibrational modes and couples themto each other. In the following, we focus on two of thesemodes which couple only to each other due to selectionrules [18], and use a harmonic approximation of the po-tential wells corresponding to the vibrational degrees offreedom. We can write the change in mass-weighted nor-mal coordinates (q1, q2) of the two modes under study as[18, 19]: (

q′1q′2

)=

(0.9 0.436−0.436 0.9

)(q1q2

)(1)

According to the Franck-Condon principle [20, 21], theintensity of a given vibrational transition is proportionalto the overlap between the wave function of its initialvibrational state and that of its final vibrational state. Ifthe initial vibrational state is the ground state and thefinal state hasm1 energy quanta in mode 1 andm2 energyquanta in mode 2, then this overlap can be written as:

P (m1,m2) =∣∣∣〈m1,m2| UDok |0, 0〉

∣∣∣2 (2)

where UDok is the operator implementing mode trans-formation 1, known as the Doktorov operator [22].P (m1,m2) is then the normalized intensity of the tran-sition at frequency m1ω1 +m2ω2, where ω1 = 176 cm−1

and ω2 = 110 cm−1 are the excited state vibrational fre-quencies of modes 1 and 2. P (m1,m2) is known as theFranck-Condon factor for this transition.

We now consider the quantum optics analogy to thevibrational transition described above. Equation 1 canbe interpreted in quantum optics as a Bogoliubov trans-formation between two optical modes. If the initial stateof these two modes is vacuum, then this transformationcan be achieved in quantum optics via two single-modesqueezing operations and a beam splitter [23]. P (m1,m2)becomes the probability of detecting m1 photons in mode1 and m2 photons in mode 2. An ideal quantum opticsexperiment that prepares two appropriate single modesqueezed vacuums (SMSV), interferes them on a beamsplitter with the appropriate reflectivity, and measuresthe resulting photon number distribution using photon-

Page 3: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

3

number resolving detectors can therefore be used to es-timate the Franck-Condon factors associated with the370 nm transition in tropolone.

We note that in this example, Eq. 1 does not in-clude a displacement term. Such a term is present formany molecules. Accounting for this additional term ina quantum optics simulation would require an additionaldisplacement step in the state preparation process.

THEORY

Quantifying the Impact of Imperfections

In the absence of imperfections, equations 1 and 2would be used to determine which squeezing parametersand beam splitter reflectivity to use in an experimentaiming to simulate the spectrum of tropolone. How-ever, in the presence of imperfections these experimentalparameters will not yield the targeted photon numberstatistics. To obtain useful results from an experiment,we require a method for comparing the experimentallygenerated photon number statistics to those of the tar-get optical state.

In the limit of a large number of modes, we cannot di-rectly compare the experimentally generated state to thetarget state in the photon number basis because thereis no known efficient classical algorithm for calculatingthese photon number statistics. We therefore proposeto use the Gaussian state formalism [24, 25] to describethe optical states. This formalism applies to states thathave a Gaussian quasi-probability distribution in phasespace and to operations that preserve the Gaussian na-ture of these states, known as Gaussian transformations.Since the initial state (vacuum) and all the optical op-erations are Gaussian, as well as most realistic sourcesof imperfection, both the ideal target state and the ex-perimentally generated optical state can be described asmultimode Gaussian states. A Gaussian state can be ef-ficiently characterized experimentally in the phase spacedescription, and although the photon number statisticsof Gaussian states cannot be efficiently calculated, thefidelity between two such states can be [26].

The fidelity between the experimentally generatedstate and the target state can be used to bound theirdifference in photon number statistics. The fidelity Fbetween two states described by density matrices ρ1 andρ2 is related to the trace distance D by [27]:

D(ρ1, ρ2) ≤√

1− F (ρ1, ρ2)2 (3)

D is related to the maximum classical l1 distance betweendifferent possible measurement outcomes by:

D(ρ1, ρ2) = max{Em}

D(pm, qm) (4)

where the maximisation is over all sets of detectorPOVMs {Em} at the output of the network, and pm =tr(ρ1Em) and qm = tr(ρ2Em). If we consider the POVMsprojecting onto photon numbers, we then have that:

||P1 − P2|| ≤√

1− F (ρ1, ρ2)2 (5)

where P1 and P2 correspond to the photon number statis-tics associated with states ρ1 and ρ2.

Equation 5 now gives us an efficiently calculable mea-sure for bounding the distance between experimentalphoton number statistics and those of the ideal state.If ρ1 is the density matrix corresponding to the experi-mentally generated state and ρ2 is that of the ideal state,we can use this inequality to bound the error on an ex-perimental estimate of the Franck-Condon factors.

In addition to the error bound calculated in this way,we can also account both for the statistical error in esti-mating the Franck-Condon factors due to the finite num-ber of experimental samples and for small deviations froma Gaussian model of an experiment. If the statistical er-ror is bounded by εstat and the error caused by deviationsfrom a Gaussian model of an experiment is bounded byεG, then the distance between the estimated photon num-ber statistics Pexp derived from the set of measurementresults and the Franck-Condon factors P is bounded by:

||Pexp − P || ≤√

1− F (ρexp, ρideal)2 + εstat + εG (6)

where ρexp corresponds to the Gaussian description of theexperimentally generated state and ρideal is the densitymatrix of the target state.

Designing an Experiment

Using the metric described above, we propose amethod for determining, in the presence of imperfections,experimental parameters that yield a state that is as closeas possible to the target optical state. Typically, an ex-perimental setup has a certain number of experimentalparameters that can be controlled, such as the pumppower, which determines the squeezing, or the interfer-ence between the modes. There also are a certain numberof parameters that cannot be controlled such as the lossand the detector dark counts. The task of producing thestate with the highest fidelity to the target state can beformulated as an optimisation problem over the control-lable parameters given the presence of the uncontrollableparameters.

To optimize the fidelity, an accurate description ofthe experimentally generated state must first be formu-lated as a function of all the experimental parameters.This procedure can be done efficiently since the numberof steps required to produce this description is polyno-mial in the number of modes. Indeed, the squeezers and

Page 4: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

4

detectors can all be characterized independently. Effi-cient methods exist to characterize interferometers [28].Losses at the input and output of the interferometer canbe individually characterized using classical light. Modeoverlaps can be determined using their pair-wise Hong-Ou-Mandel dip visibility. The experimentally generatedstate can therefore in principle be accurately describedfor any combination of experimental parameters.

However, the numerical optimization procedure forfinding this optimal quantum state is not necessarily ef-ficient. The fidelity can be expected to be a nonlinearfunction of all the experimental parameters, so the opti-mization procedure is not straightforward. However, nu-merical optimization techniques can be used to at leastfind a local optimum in parameter space, which depend-ing on the desired accuracy of the simulation may besuitable.

Benchmarking against Classical Simulations

Quantum optical experiments aiming to estimate vi-bronic spectra are worthwhile if they outperform knownclassical algorithms. Although the idealised original pro-posal by Huh et al may outperform known classical al-gorithms, it is not a priori clear that an imperfect ex-periment designed according to the principles describedabove would also do so. Furthermore, while there is noknown efficient exact classical algorithm for calculatingvibronic spectra, some classical approximation strategiesdo exist. One case-by-case strategy involves guessingwhich transitions are likely to contribute the most to thespectrum and only calculating the corresponding Franck-Condon factors [29]. A quantum optics experiment withimperfections will only be worthwhile if such other ap-proximation strategies yield worse estimates of vibronicspectra than the experiment.

We propose the following efficient classical approxima-tion algorithm as a benchmark that experiments mustoutperform in order to produce better than classical esti-mates of vibronic spectra. This algorithm is conceptuallysimilar to the quantum simulation protocol, so that thesame analysis tools can be used in both cases.

We start by finding the classical optical state, definedas having a regular P-function in phase space [30], thatmaximises the fidelity to the target state within the spaceof all optical states. First, we note that the displacementoperation that occurs in the state preparation processcan in principle occur before the interferometer insteadof after, so that the vibronic spectroscopy experimentconsists of squeezed and displaced states sent into an in-terferometer. The fidelity being invariant under unitarytransformations, finding the closest classical state to theideal target state is equivalent to finding the closest clas-sical state to these initial displaced squeezed states. Theclosest single mode classical state to a single mode dis-

placed Gaussian state is a coherent state with the samedisplacement [31]. Therefore, the closest multimode clas-sical state to the target state is a multimode coherentstate with the same displacement. In the case of thetransition in tropolone discussed above, the closest clas-sical state is vacuum due to the absence of displacementin equation 1.

Next, we simulate sampling from the photon numberstatistics of this state. Since this state is classical, itsphoton number statistics can efficiently be sampled fromusing a classical algorithm [32]. Since this state is alsoGaussian, equation 5 can be used to estimate the tar-get vibronic spectrum to within some error bound. Tooutperform this classical algorithm, an experiment mustyield an error bound given by equation 6 that is smallerthan that yielded by the classical state.

This classical approximation strategy can also be usedas a classicality witness for the optical state: any ex-perimental state with a higher fidelity must have a non-regular P-function. We therefore use this best classicalstate as our benchmark for demonstrating a quantum ad-vantage in experiment. Any experimental optical statethat beats the witness is both a non-classical state andproduces a better approximation of vibronic spectra thanwould be possible with any classical state. Furthermore,we note that if an experimentally generated state beatsour classicality criterion, then currently known classicalsimulation algorithms based on the phase space descrip-tion of the state are generally not applicable [32, 33].

EXPERIMENT

Overview

We perform a proof of principle experiment to simulatethe transition in tropolone described by equation 1. Wechoose tropolone to illustrate our findings for the follow-ing reasons. First, we note that the vibronic transitiondescribed by equation 1 does not include a displacementterm, which is present in many other molecules. Dis-placements can be implemented in quantum optics usingclassical laser light and do not affect the classicality of astate, as defined by the regularity of its P-function [30].Furthermore, the squeezing parameters for the two modesin the ideal experiment are 0.19 and 0.72, which is quitelarge compared to most other molecules. The absence ofdisplacement and these large squeezing parameters allowus to focus our analysis on the quantum mechanical as-pects of the experiment. These factors also allow us tohighlight the impact of imperfections such as loss, whichsqueezing is strongly affected by, on an experiment.

Our setup (Fig. 2A) consists of a 100 kHz pulsed laserat 780 nm that pumps a periodically poled potassiumtitanyl phosphate (KTP) waveguide to produce a degen-erate two mode squeezed vacuum (TMSV) at 1560 nm

Page 5: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

5

FIG. 2. A) We approximate the Franck-Condon factors of tropolone experimentally by sending pump pulses at 780 nm intoa periodically poled KTP waveguide, separating the two orthogonally polarized downconverted modes at 1560 nm using apolarizing beam splitter (PBS), and rotating the polarization of one of the two modes using a half wave plate (HWP). Wethen couple these two modes into polarization-maintaining (PM) fiber, interfere them in a tunable PM beam splitter (BS), andmeasure them using two fiber-coupled transition-edge sensors (TES). B) We model this experiment as a two mode squeezedvacuum (TMSV) interfered on a beam splitter and measured by noisy photon number resolving detectors (PNRD), in thepresence of loss and noise produced by the non-overlapping parts of the two modes (both of which are modeled using beamsplitters, see Supplementary Materials). The squeezing parameter r of the TMSV and the transmissions of the beam splittersin our model of the experiment are shown in the table on the right.

[34], the two modes of which we separate and couple intothe two inputs of a fiber beam splitter. We then use twotransition-edge sensors (TES) [35] to sample from thephoton number statistics from the two outputs of thebeam splitter. The frequency of the occurrence of pho-ton numbers m1 and m2 in modes 1 and 2 gives a directestimate of the joint probability in equation 2 and thusof the Franck Condon factors for the transition towardsthe states with m1 and m2 vibrational quanta.

Our setup deviates in the following ways from the idealsetup described above. Firstly, we approximate the twoindependent SMSVs that are required in an ideal setupusing a TMSV, since TMSVs are experimentally simplerto generate and mode-match. The TMSV can be con-verted into two identical SMSVs using the beam splitterin our setup. Furthermore, the two modes are not exactlyidentical and so do not interfere perfectly on the beamsplitter; the Hong-Ou-Mandel dip [36] between the twomodes has a 94% visibility. Another significant imper-fection is the limited system efficiency from the photonsource to the detectors; approximately 60% of the gen-erated photons are not detected. This loss mostly comesfrom the low coupling efficiency from the photon sourceinto single mode fiber, and is expected to degrade thesqueezing. Finally, our TES detectors are noisy, with a0.2% dark count probability and a probability of 0.1%

of detecting pump photons that leak through our setup.Since photons from the pump have twice the energy of thedownconverted photons and TESs are energy-resolvingdetectors, these events register as two-photon events.

Characterization

To produce an optical state that maximizes the fidelityto the target state in the presence of imperfections, wefirst characterize the experimental setup in order to re-alise a model for the experiment within the Gaussianstate formalism. In the following, we explain how thischaracterization and modelling is performed. Figure 2shows the model for the experiment that we use as aresult of this characterization process.

Detectors

We model the TESs as photon number-resolving detec-tors that, despite having a very low intrinsic dark countrate, suffer from the following noise mechanisms.

Firstly, our TESs have some degree of inefficiency. Thisinefficiency is accounted for in our estimate of the totalsystem efficiency described in the following section.

Page 6: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

6

FIG. 3. Characterization data for our experiment. A) Total losses for the two modes in our setup, estimated from ourtomography procedure for different pump powers, with the beam splitter set to full transmission. We use the average and thestandard deviation of the values in the shaded area for our estimate of the experimental parameters. B) Squeezing parameter,estimated from our tomography procedure, as a function of pump power. Within the shaded area, we find that the relationbetween squeezing parameter and pump power is very close to the theoretical relation P ∝ r2. C) Deviation between ourexperimental photon number statistics and the theoretical photon number statistics given by our model for the optical state,quantified by the l1 distance. D) Coincidence counts on our detectors measured as a function of the angle of the half waveplate (HWP) in our setup. We observe a clear Hong-Ou-Mandel dip [36].

Secondly, the binning procedure that we use to extractphoton numbers from the analogue output signal is sus-ceptible to electronic noise on our detectors. With a vac-uum input, there still is a 0.2% probability of wronglyregistering a one photon event. We approximate thisnoise as a dark count mechanism, which can be accountedfor within a Gaussian model by considering that our de-tector is sensitive to both a noiseless input signal andto an additional mode containing a thermal state with a0.2% single photon component.

Thirdly, some pump photons leak through our setupand make it to the detectors. Since TESs resolve theenergy of incoming photons, these pump photons arecounted as two-photon events. We find that our TESshave a 0.1% probability of detecting these pump pho-tons. This noise mechanism can be included within ourmodel by considering that our detector is sensitive to anadditional pump mode containing a weak coherent statewith a 0.1% single photon component.

This noise can be included in our estimate of the fi-delity. Both additional noise modes have a fidelity ofabout 0.998 to vacuum, and since the fidelity for a prod-uct state is the product of the fidelities, the total fidelityof these noise modes to vacuum is 0.9958. This fidelitymust then be multiplied by the estimated fidelity of theoptical state before the detectors in order to determinethe total fidelity.

Squeezing and loss

We characterize the total loss (including detector inef-ficiency) and squeezing in our system using a tomographytechnique similar to that described in [37]. We model ourphoton source as a perfect TMSV with additional loss inthe two modes, use our model for the detectors describedabove, and numerically find the squeezing parameter andthe distribution of the loss in both arms which yield thephoton number statistics that most closely match the ex-

perimental photon number statistics. We follow this pro-cedure for the beam splitter set first to 100:0, and then to0:100, in order to determine how the losses in the systemare distributed in both modes both before and after thebeam splitter. We note that since balanced losses canmathematically be commuted through the beam splitter,we need to consider losses in only one of the output portsof the beam splitter in our model.

To verify that our estimate of the losses is reliable, weperform this tomographic procedure for one of the twobeam splitter settings for several different pump powersranging from 10 µW to 300 µW. We find that at powersexceeding 100 µW our results are skewed as higher ordernonlinearities start affecting the pump and the downcon-verted modes. We therefore use the values for the lossthat are found in the low power region in Fig. 3, andestimate the error on these values to be ±2%. These val-ues are also consistent with the heralding efficiencies esti-mated from the photon number statistics in this plateauregion.

The tomography procedure yields the squeezing pa-rameter that we can directly used in our Gaussian model.The squeezing parameter r is related to the pump powerP by the following relation:

P ∝ r2 (7)

Once the squeezing parameter has been determined forone power, we can use equation 7 to determine thesqueezing for any power. To estimate the error on ourestimate of the squeezing parameter, we fit the squeezingparameters determined from our optimization procedureto a curve given by equation 7 in the plateau region. Thisfit is shown in Fig. 3. We estimate the error on our es-timate of the value of the squeezing parameter at lowpowers to be 0.01.

We also use our tomography results to estimate the er-ror εG stemming from our Gaussian approximation of theexperiment. Fig 3 shows the deviation, quantified by l1

Page 7: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

7

distance, between the measured photon number statisticsand those of the closest lossy TMSV measured with noisydetectors determined by our tomography technique. Atlow powers, this error is less than 10−3, so we considerthat our description of the optical state and of our de-tectors is satisfactory.

Distinguishability

We characterise the distinguishability δ between theoptical modes by using the depth of the Hong-Ou-Mandelinterference [36], shown in Fig. 3, measured by settingthe beam splitter to 50:50 and measuring the number ofcoincidence counts at the outputs as we rotate the HWPin our experimental setup. We use SNSPDs as our pho-ton detectors for this procedure due to their greater easeof operation. Considering that non-overlapping parts ofthe two modes can be modelled as noise photons, the ra-tio of noise photons to signal photons in the system isthen δ. We choose to model this noise as virtual beamsplitters of reflectivity δ, placed just after the squeezingoperation for both modes, between each mode and a vir-tual thermal state containing the same average numberof photons as the TMSV. The ±2% error on our esti-mate of δ comes from the error on the estimate of thedepth of the measured Hom-Ou-Mandel dip. We notethat our model for the noise given by the distinguisha-bility is only an approximation of the full description ofthis noise, which would require taking several additionalnon-interfering modes into account. However, our modelprovides a rough estimate of the contribution of this noisetowards the degradation of the fidelity.

Beam splitter reflectivity

The beam splitter reflectivity was set by blocking onemode in our experiment, setting the beam splitter to befully transmissive so that the maximum photon numberat the detector for that mode could be determined for agiven pump power, and then adjusting the beam split-ter until the average photon number at the detector wasthe desired fraction of the maximum. The error on ourestimate of the reflectivity was estimated at ±1%.

Optimization and Results

Finding the optimal state

The characterization procedure described previouslyyields a Gaussian model for the experiment. For anyvalue of the beam splitter reflectivity and squeezing pa-rameter, we now have a Gaussian description of the out-put state that accounts for the imperfections in the setup.

To find the values of the squeezing parameter r andbeam splitter reflectivity tBS that maximize the fidelityof the experimentally generated state to the target state,we use Matlab’s built-in fminsearch procedure. Giventhe small size of the parameter space, this routine canbe expected to find the global optimum for the fidelity.Fig. 2B shows the values of r and tBS that maximizethe fidelity within our model. This maximum theoreticalfidelity is 0.891.

Given the errors in our model, we expect not to achievethis maximum fidelity in practice. To obtain a more rea-sonable estimate of the fidelity, we use a Monte Carlomethod to determine the most likely value of the fidelitythat we achieve as well as the error on this value. Wesimulate 100 states produced by our experiment, wherewe randomly select the experimental parameters from aGaussian probability distribution determined by the esti-mated mean and standard deviations given by our anal-ysis. The mean and standard deviation in the fidelity ofthese samples are used as our fidelity estimate and as theerror on this estimate. With this method, we revise ourestimate of the fidelity to the target state to 0.890(1).

Estimated Franck-Condon factors

We collected 1638370 samples over the course of about20 seconds from the photon number statistics of our statefrom which we estimate the Franck-Condon factors of thetransition under study. Table I compares our experimen-tal estimate to the exact theoretical target values numer-ically calculated using equations 1 and 2. By comparingour experimental values to the exact values, we find anerror of 0.206. This error is quite large; the followingsection will provide a detailed analysis of the sources oferror.

Frequency Experiment Ideal

0 0.9628 0.7731

ω1 0.0129 0

ω2 0.0127 0

2ω1 0.0035 0.1097

2ω2 0.0038 0.0041

ω1 + ω2 0.0035 0.0469

4ω1 < 10−4 0.0233

3ω1 + ω2 < 10−4 0.0200

TABLE I. Most significant Franck-Condon factors estimatedby our experiment and by simulations for an ideal experiment.

Error analysis and classicality

We apply equation 6 to our experiment to determinethe theoretical error bound using our analysis of the fi-

Page 8: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

8

delity and the observed deviation from Gaussian behaviorin our system. Given the large number of samples, we ne-glect the sampling error. We find a bound for the tracedistance of 0.455. Our experimental results are indeedwithin this bound.

We also apply our classicality criterion to this experi-ment. The classical state with the highest fidelity to thetarget state is vacuum, which has a fidelity of 0.879 tothe target state. Our experiment has a higher fidelity byabout 10 standard deviations and therefore satisfies ourclassicality criterion. Using vacuum, the classical approx-imation algorithm described above would yield an errorbound of 0.476, which is worse than what we achievedin experiment. However, we note that since vacuum is aGaussian state, then in the specific case where the closestclassical state is vacuum the difference in photon numberstatistics between the target optical state and the closestclassical state can be efficiently calculated, as opposed tosimply bounded using the fidelity.

DISCUSSION

In order to contribute to future designs of experiments,in this section we discuss our experimental results, and inparticular we analyse the main ways in which different ex-perimental imperfections contribute to the degradation ofthe fidelity. Figure 4 shows the effect of different sourcesof imperfection on the fidelity in our experimental setup.An ideal experiment consists of SMSVs and ideal detec-tors (black solid line). The additional dashed and dottedlines indicate the effect of additional imperfections thatmust be accounted for in our experiment. The orangedashed and dotted line indicates the effect of all the pa-rameters that we account for in our analysis, and theorange square corresponds to our experiment. The flatgreen solid line indicates the maximum fidelity that canbe achieved with the best classical state. These theo-retical curves can easily be derived using our Gaussianmodel for the experiment.

We see that our classicality criterion is relatively tol-erant of experimental imperfections to simulate transi-tions that involve large amounts of squeezing. An en-hancement over the best classical state can be achievedwith values of loss up to 90%, with noisy detectors, andwith the use of a TMSV instead of two SMSVs. For lossexceeding 90%, the noise on our detectors degrades thefidelity below that of the best classical state. For transi-tions in other molecules that involve less squeezing, suchas the S0(1A1g)→ S1(1B2u) transition in benzene or theSO−2 → SO2 transition studied by Shen et al [13], weexpect that our level of detector noise would prevent usfrom outperforming this criterion for any level of loss.

A comparison of our experimental results and of theideal theoretical results for a perfect experiment to simu-lations of other intermediate states is shown in Table II.

Frequency Experiment Ideal Lossy SMSVs Best SMSVs

0 0.9628 0.7731 0.9327 0.7631

ω1 0.0129 0 0.0377 0.0015

ω2 0.0127 0 0.0073 0.0015

2ω1 0.0035 0.1097 0.0136 0.1102

2ω2 0.0038 0.0041 0.0004 0.0046

ω1 + ω2 0.0035 0.0469 0.0053 0.0466

4ω1 < 10−4 0.0233 0.0004 0.0234

3ω1 + ω2 < 10−4 0.0200 0.0003 0.0199

Fidelity 0.890(1) 1 0.9068 0.9958

Error 0.206 0 0.195 0.005

TABLE II. Most significant Franck-Condon factors, fidelities,and errors estimated by our experiment, and by simulationsfor an ideal experiment, the best SMSVs for our level of loss,and the best SMSV in a lossless experiment but with noisydetectors.

The loss leads to an overestimate of the Franck-Condonfactor corresponding to vacuum, both for our experimentand for a theoretical lossy SMSV. Furthermore, whereasthe Franck-Condon factors for odd numbers of excitationsshould be 0 due to the difference in symmetry betweenthe ground state and the odd-numbered excited states,our experiment finds these to be non-zero due to photonswhich would correspond to higher order Franck-Condonfactors being lost. The use of a TMSV instead of two in-dependent SMSVs causes us to experimentally find pho-ton numbers that are roughly symmetric in both modes,to within the imbalance in the loss in the two arms. Wesee from our results that, in the case of tropolone, al-though the high squeezing and absence of displacementhas allowed us to highlight the issue of imperfections inan experiment, by the same token our simulations resultin a large error in estimating Franck-Condon factors.

CONCLUSION

We have proposed a method for accounting for exper-imental imperfections in quantum optical simulations ofvibronic spectroscopy, following the proposal by Huh etal. We have shown that the impact of these imperfectionscan be quantified, that an experimental setup can be ad-justed to account for the presence of these imperfections,and that a classicality benchmark that experiments mustoutperform to be worthwhile can be formulated. We il-lustrated these results using a proof of principle exper-iment that simulated part of the vibronic spectrum oftropolone.

Our results inform future efforts for performing largerscale simulations of vibronic spectra with current quan-tum optics technology, for example using recent advancesin fiber-loop based experiments [15, 16] and integratedphotonics [10]. We also note that our approach for deal-

Page 9: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

9

FIG. 4. Maximum achievable fidelity as a function of the fraction of light lost in a simulation of tropolone, for differentimperfections in an experimental setup. We simulate an ideal setup that is only affected by loss (black solid line), to which weadd our noisy detectors (blue dashed line), then replace the SMSVs by a TMSV (red dotted line), then add our measured partialoverlap between the modes (orange dashed-dotted line). The cross and the circle respectively indicate the best SMSV for ourlevel of loss and the best SMSV without loss but with noisy detectors, yielding the estimated Franck-Condon factors shown inTable 1. The orange square indicates the experimental parameters used for our experiment. For this analysis, we assume equalloss in both modes, hence the discrepancy between the orange square and the position of the orange dashed-dotted line.

ing with imperfections can be applied to the other plat-forms which have been proposed for vibronic spectra es-timation [12, 13]. We envisage that our work will beuseful for efficiently approximating the spectra of largemolecules that are outside the reach of classical comput-ers.

ACKNOWLEDGEMENTS

We thank David Phillips, David Drahi, Gil Trig-iner and Johan Fopma for experimental assistance, andAdrian Menssen, Benjamin Metcalf, Peter Humphreys,Raul Garcia-Patron, Jan Sperling, Oscar Dahlstenand Myungshik Kim for helpful discussions. W.R.C,J.J.R, A.E., W.S.K. and I.A.W. acknowledge sup-port from the European Research Council, the UKEngineering and Physical Sciences Research Council(project EP/K034480/1 and the Networked QuantumInformation Technology Hub), the Fondation WienerAnspach, and from the European Commission (H2020-FETPROACT-2014 grant QUCHIP). J.J.R. acknowl-edges support from NWO Rubicon. J.H. acknowl-edges support from the Basic Science Research Pro-gram through the National Research Foundation of Ko-rea (NRF) funded by the Ministry of Education, Scienceand Technology (NRF-2015R1A6A3A04059773).

ADDITIONAL INFORMATION

Competing financial interests

The authors declare no competing financial interests.

Supplementary Materials

Accompanies this paper. The underlying data andmodels used in this work are available upon reasonablerequest from the authors.

[1] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and M. Head-Gordon, “Simulated quantum computation of molecularenergies,” Science, vol. 309, no. 5741, pp. 1704–1707,2005.

[2] J. Huh, G. G. Guerreschi, B. Peropadre, J. R. McClean,and A. Aspuru-Guzik, “Boson sampling for molecular vi-bronic spectra,” Nature Photonics, vol. 9, pp. 615–620,2015.

[3] J. Huh and M.-H. Yung, “Vibronic boson sampling: Gen-eralized gaussian boson sampling for molecular vibronicspectra at finite temperature,” Scientific Reports, vol. 7,no. 7462, 2017.

Page 10: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

10

[4] J. Hachmann, R. Olivares-Amaya, S. Atahan-Evrenk,C. Amador-Bedolla, R. S. Sanchez-Carrera, A. Gold-Parker, L. Vogt, A. M. Brockway, and A. Aspuru-Guzik,“The Harvard clean energy project: large-scale compu-tational screening and design of organic photovoltaicson the world community grid,” The Journal of PhysicalChemistry Letters, vol. 2, no. 17, pp. 2241–2251, 2011.

[5] H. J. Butler, L. Ashton, B. Bird, G. Cinque, K. Curtis,J. Dorney, K. Esmonde-White, N. J. Fullwood, B. Gard-ner, P. L. Martin-Hirsch, M. J. Walsh, M. R. McAinsh,N. Stone, and F. L. Martin, “Using Raman spectroscopyto characterize biological materials,” Nature protocols,vol. 11, no. 4, pp. 664–687, 2016.

[6] M. Gross, D. C. Muller, H.-G. Nothofer, U. Scherf, D. Ne-her, C. Brauchle, and K. Meerholz, “Improving the per-formance of doped π-conjugated polymers for use in or-ganic light-emitting diodes,” Nature, vol. 405, no. 6787,pp. 661–665, 2000.

[7] J. Janszky, A. V. Vinogradov, I. A. Walmsley, andJ. Mostowski, “Competition between geometrical and dy-namical squeezing during a Franck-Condon transition,”Physical Review A, vol. 50, pp. 732–740, Jul 1994.

[8] S. Aaronson and A. Arkhipov, “The computational com-plexity of linear optics,” in Proceedings of the forty-third annual ACM symposium on Theory of computing,pp. 333–342, ACM, 2011.

[9] J. B. Spring, B. J. Metcalf, P. C. Humphreys,W. S. Kolthammer, X.-M. Jin, M. Barbieri, A. Datta,N. Thomas-Peter, N. K. Langford, D. Kundys, J. C.Gates, B. J. Smith, P. G. R. Smith, and I. A. Walmsley,“Boson sampling on a photonic chip,” Science, vol. 339,no. 6121, pp. 798–801, 2013.

[10] J. Carolan, C. Harrold, C. Sparrow, E. Martın-Lopez,N. J. Russell, J. W. Silverstone, P. J. Shadbolt, N. Mat-suda, M. Oguma, M. Itoh, G. D. Marshall, M. G. Thomp-son, J. C. F. Matthews, T. Hashimoto, J. L. O’Brien,and A. Laing, “Universal linear optics,” Science, vol. 349,no. 6249, pp. 711–716, 2015.

[11] H. Wang, Y. He, Y.-H. Li, Z.-E. Su, B. Li, H.-L. Huang,X. Ding, M.-C. Chen, C. Liu, J. Qin, J.-P. Li, Y.-M.He, C. Schneider, M. Kamp, C.-Z. Peng, S. Hfling, C.-Y.Lu, and J.-W. Pan, “High-efficiency multiphoton bosonsampling,” Nature Photonics, vol. 11, no. 6, pp. 361–365,2017.

[12] L. Hu, Y.-C. Ma, Y. Xu, W.-T. Wang, Y.-W. Ma, K. Liu,H.-Y. Wang, Y.-P. Song, M.-H. Yung, and L.-Y. Sun,“Simulation of molecular spectroscopy with circuit quan-tum electrodynamics,” Science Bulletin, 2018.

[13] Y. Shen, Y. Lu, K. Zhang, J. Zhang, S. Zhang, J. Huh,and K. Kim, “Quantum optical emulation of molecu-lar vibronic spectroscopy using a trapped-ion device,”Chemical Science, 2018.

[14] L.-A. Wu, H. Kimble, J. Hall, and H. Wu, “Generation ofsqueezed states by parametric down conversion,” Physi-cal Review Retters, vol. 57, no. 20, p. 2520, 1986.

[15] P. L. McMahon, A. Marandi, Y. Haribara, R. Hamerly,C. Langrock, S. Tamate, T. Inagaki, H. Takesue,S. Utsunomiya, K. Aihara, R. L. Byer, M. M. Fejer,H. Mabuchi, and Y. Yamamoto, “A fully-programmable100-spin coherent Ising machine with all-to-all connec-tions,” Science, 2016.

[16] T. Inagaki, Y. Haribara, K. Igarashi, T. Sonobe, S. Ta-mate, T. Honjo, A. Marandi, P. L. McMahon, T. Umeki,K. Enbutsu, O. Tadanaga, H. Takenouchi, K. Aihara, K.-

i. Kawarabayashi, K. Inoue, S. Utsunomiya, and H. Take-sue, “A coherent Ising machine for 2000-node optimiza-tion problems,” Science, vol. 354, no. 6312, pp. 603–606,2016.

[17] J. C. Giddings, Advances in chromatography, vol. 14.CRC Press, 1976.

[18] W. Smith, “The strong Duschinsky effect and the inten-sity of transitions in non-totally symmetric vibrations inthe electronic spectra of polyatomic molecules,” Journalof molecular spectroscopy, vol. 187, no. 1, pp. 6–12, 1998.

[19] F. Duschinsky, “The importance of the electron spec-trum in multi atomic molecules. Concerning the Franck-Condon principle,” Acta Physicochim. URSS, vol. 7,pp. 551–566, 1937.

[20] E. U. Condon, “Nuclear motions associated with elec-tron transitions in diatomic molecules,” Physical Review,vol. 32, no. 6, p. 858, 1928.

[21] J. Franck and E. G. Dymond, “Elementary processes ofphotochemical reactions,” Trans. Faraday Soc., vol. 21,no. February, pp. 536–542, 1926.

[22] E. V. Doktorov, I. A. Malkin, and V. I. Man’ko, “Dy-namical symmetry of vibronic transitions in polyatomicmolecules and the Franck-Condon principle,” Journal ofMolecular Spectroscopy, vol. 64, no. 2, pp. 302–326, 1977.

[23] S. L. Braunstein, “Squeezing as an irreducible resource,”Physical Review A, vol. 71, no. 5, p. 55801, 2005.

[24] G. Adesso, S. Ragy, and A. R. Lee, “Continuous vari-able quantum information: Gaussian states and be-yond,” Open Systems & Information Dynamics, vol. 21,no. 01n02, p. 1440001, 2014.

[25] C. Weedbrook, S. Pirandola, R. Garcıa-Patron, N. J.Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaus-sian quantum information,” Reviews of Modern Physics,vol. 84, no. 2, p. 621, 2012.

[26] L. Banchi, S. L. Braunstein, and S. Pirandola, “Quantumfidelity for arbitrary gaussian states,” Physical ReviewLetters, vol. 115, no. 26, p. 260501, 2015.

[27] M. A. Nielsen and I. L. Chuang, Quantum computationand quantum information. Cambridge university press,2010.

[28] A. Laing and J. L. O’Brien, “Super-stable tomog-raphy of any linear optical device,” arXiv preprintarXiv:1208.2868, 2012.

[29] H.-C. Jankowiak, J. L. Stuber, and R. Berger, “Vi-bronic transitions in large molecular systems: Rigorousprescreening conditions for Franck-Condon factors,” J.Chem. Phys., vol. 127, p. 234101, 2007.

[30] U. Leonhardt, Measuring the quantum state of light,vol. 22. Cambridge university press, 1997.

[31] P. Marian, T. A. Marian, and H. Scutaru, “Quantify-ing nonclassicality of one-mode Gaussian states of theradiation field,” Physical Review Letters, vol. 88, no. 15,p. 153601, 2002.

[32] S. Rahimi-Keshari, T. C. Ralph, and C. M. Caves, “Suffi-cient conditions for efficient classical simulation of quan-tum optics,” Physical Review X, vol. 6, no. 2, p. 021039,2016.

[33] S. Rahimi-Keshari, A. P. Lund, and T. C. Ralph, “Whatcan quantum optics say about computational complexitytheory?,” Physical Review Letterrs 114, 060501 (2015),2015.

[34] A. Eckstein, A. Christ, P. J. Mosley, and C. Silberhorn,“Highly efficient single-pass source of pulsed single-modetwin beams of light,” Physical Review Letters, vol. 106,

Page 11: Approximating Vibronic Spectroscopy with …Approximating Vibronic Spectroscopy with Imperfect Quantum Optics William R. Clements, 1Jelmer J. Renema, Andreas Eckstein, Antonio A. Valido,2

11

no. 1, p. 013603, 2011.[35] A. E. Lita, A. J. Miller, and S. W. Nam, “Counting near-

infrared single-photons with 95% efficiency,” Optics ex-press, vol. 16, no. 5, pp. 3032–3040, 2008.

[36] C. K. Hong, Z.-Y. Ou, and L. Mandel, “Measurementof subpicosecond time intervals between two photons by

interference,” Physical review letters, vol. 59, no. 18,p. 2044, 1987.

[37] A. Worsley, H. Coldenstrodt-Ronge, J. Lundeen,P. Mosley, B. Smith, G. Puentes, N. Thomas-Peter, andI. Walmsley, “Absolute efficiency estimation of photon-number-resolving detectors using twin beams,” Opticsexpress, vol. 17, no. 6, pp. 4397–4412, 2009.