arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

11
Exploring Interacting Quantum Many-Body Systems by Experimentally Creating Continuous Matrix Product States in Superconducting Circuits C. Eichler 1* , J. Mlynek 1 , J. Butscher 1 , P. Kurpiers 1 , K. Hammerer 2,3,4 , T. J. Osborne 2 , A. Wallraff 1 1 Department of Physics, ETH Z¨ urich, CH-8093, Z¨ urich, Switzerland 2 Institute for Theoretical Physics, Leibniz University, 30167 Hannover, Germany 3 Institut f¨ ur Gravitationsphysik, Leibniz Universit¨ at, 30167 Hannover, Germany and 4 Max-Planck Institut f¨ ur Gravitationsphysik (Albert-Einstein Institut), 30167 Hannover, Germany Improving the understanding of strongly correlated quantum many body systems such as gases of interacting atoms or electrons is one of the most important challenges in modern condensed matter physics, materials re- search and chemistry. Enormous progress has been made in the past decades in developing both classical and quantum approaches to calculate, simulate and experimentally probe the properties of such systems. In this work we use a combination of classical and quantum methods to experimentally explore the properties of an interacting quantum gas by creating experimental realizations of continuous matrix product states - a class of states which has proven extremely powerful as a variational ansatz for numerical simulations. By systematically preparing and probing these states using a circuit quantum electrodynamics (cQED) system we experimentally determine a good approximation to the ground-state wave function of the Lieb-Liniger Hamiltonian, which describes an interacting Bose gas in one dimension. Since the simulated Hamiltonian is encoded in the mea- surement observable rather than the controlled quantum system, this approach has the potential to apply to exotic models involving multicomponent interacting fields. Our findings also hint at the possibility of experimentally exploring general properties of matrix product states and entanglement theory. The scheme presented here is applicable to a broad range of systems exploiting strong and tunable light-matter interactions. Progress in revealing relations between solid state physics and quantum information theory has constantly extended the range of quantum many body problems which are tractable with classical computers. One such successful approach is the density matrix renormalization group (DMRG), which was in- troduced by White in 1992 [1] and since then developed into a leading method for numerical studies of strongly interacting one dimensional lattice systems [2]. Later it was realized that the DMRG can be interpreted as a variational optimization over matrix product states (MPS) [3, 4]. The class of matrix product states [5, 6] naturally incorporates an area law for the entanglement entropy [7] and is thus ideally suited to parame- terize many-body states with finite correlation length [8]. An interesting connection between the MPS formalism and open quantum systems has recently been discovered [9], which led to the suggestion of using the high-level of experimental con- trol achievable over open cavity QED systems to create con- tinuous matrix product states [10] as itinerant radiation fields for the purpose of quantum simulations [11, 12]. In this letter we provide experimental evidence that this concept is indeed capable of determining the properties of strongly correlated quantum systems and offers promising perspectives, comple- mentary to existing digital and analog quantum simulation ap- proaches [13] explored with trapped atoms and ions [1418]. In clear distinction to previous experiments, here, we sim- ulate a continuous quantum field theory rather than a lat- tice model. In particular, we study the ground-state of the Lieb-Liniger model ˆ H = R dx ˆ H = R dx( ˆ N + ˆ T + ˆ W ) [19], which describes a gas of interacting bosons confined in a one-dimensional continuum [20], as schematically de- picted in Fig. 1a. Here, ˆ N = -μ ˆ ψ x ˆ ψ x is the potential en- ergy, ˆ T = x ˆ ψ x x ˆ ψ x is the kinetic energy of particles, and * Current address: Department of Physics, Princeton University, Princeton NJ 08544, USA. Correspondence to [email protected] m (a) x cMPS (b) control fields λ κ 1D transmission line φ(λ) H measure vary λ to minimize E λ E λ = 0 0.5 1 1.5 0 1 2 3 4 5 6 Τ Μs G 1 ΤΜs 1 2Π MHz 1.05 2Π MHz 1.48 2Π MHz 2.09 2Π MHz 2.95 0 0.5 1 1.5 0.0 0.2 0.4 0.6 0.8 1.0 1.2 Τ Μs g 2 Τ (c) (d) T W FIG. 1: Schematic of the interacting Bose gas and the principle of the quantum variational algorithm. (a), Bosonic particles with ki- netic energy ˆ T are propagating in one dimension along the x axis. Repulsion between particles mediates an interaction energy ˆ W . The particle density ρ of the gas is controlled by the chemical potential μ. (b), We experimentally simulate the ground-state of ˆ H by em- ploying a variational minimization procedure. A tunable cavity QED system is used to generate radiation fields emulating continuous ma- trix product states |φ(λ)i in a 1D transmission line. The average energy E λ = hφ(λ)| ˆ H|φ(λ)i of the simulated Hamiltonian is ex- perimentally determined from measured correlation functions. Ex- ternal control fields λ are used as variational parameters. (c),(d), Ex- amples of first-order G (1) (τ ) and second-order g (2) (τ ) correlation functions measured (dots) in superconducting circuits and simulated (solid lines) using a master equation approach for the indicated drive rates Ω, effective anharmonicity α/2π =5.2 MHz, and cavity decay rate κ/2π =2.2 MHz. arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

Transcript of arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

Page 1: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

Exploring Interacting Quantum Many-Body Systemsby Experimentally Creating Continuous Matrix Product States in Superconducting Circuits

C. Eichler1∗, J. Mlynek1, J. Butscher1, P. Kurpiers1, K. Hammerer2,3,4, T. J. Osborne2, A. Wallraff11 Department of Physics, ETH Zurich, CH-8093, Zurich, Switzerland

2 Institute for Theoretical Physics, Leibniz University, 30167 Hannover, Germany3 Institut fur Gravitationsphysik, Leibniz Universitat, 30167 Hannover, Germany and

4 Max-Planck Institut fur Gravitationsphysik (Albert-Einstein Institut), 30167 Hannover, Germany

Improving the understanding of strongly correlated quantum many body systems such as gases of interactingatoms or electrons is one of the most important challenges in modern condensed matter physics, materials re-search and chemistry. Enormous progress has been made in the past decades in developing both classical andquantum approaches to calculate, simulate and experimentally probe the properties of such systems. In thiswork we use a combination of classical and quantum methods to experimentally explore the properties of aninteracting quantum gas by creating experimental realizations of continuous matrix product states - a class ofstates which has proven extremely powerful as a variational ansatz for numerical simulations. By systematicallypreparing and probing these states using a circuit quantum electrodynamics (cQED) system we experimentallydetermine a good approximation to the ground-state wave function of the Lieb-Liniger Hamiltonian, whichdescribes an interacting Bose gas in one dimension. Since the simulated Hamiltonian is encoded in the mea-surement observable rather than the controlled quantum system, this approach has the potential to apply to exoticmodels involving multicomponent interacting fields. Our findings also hint at the possibility of experimentallyexploring general properties of matrix product states and entanglement theory. The scheme presented here isapplicable to a broad range of systems exploiting strong and tunable light-matter interactions.

Progress in revealing relations between solid state physicsand quantum information theory has constantly extended therange of quantum many body problems which are tractablewith classical computers. One such successful approach is thedensity matrix renormalization group (DMRG), which was in-troduced by White in 1992 [1] and since then developed intoa leading method for numerical studies of strongly interactingone dimensional lattice systems [2]. Later it was realized thatthe DMRG can be interpreted as a variational optimizationover matrix product states (MPS) [3, 4]. The class of matrixproduct states [5, 6] naturally incorporates an area law for theentanglement entropy [7] and is thus ideally suited to parame-terize many-body states with finite correlation length [8]. Aninteresting connection between the MPS formalism and openquantum systems has recently been discovered [9], which ledto the suggestion of using the high-level of experimental con-trol achievable over open cavity QED systems to create con-tinuous matrix product states [10] as itinerant radiation fieldsfor the purpose of quantum simulations [11, 12]. In this letterwe provide experimental evidence that this concept is indeedcapable of determining the properties of strongly correlatedquantum systems and offers promising perspectives, comple-mentary to existing digital and analog quantum simulation ap-proaches [13] explored with trapped atoms and ions [14–18].

In clear distinction to previous experiments, here, we sim-ulate a continuous quantum field theory rather than a lat-tice model. In particular, we study the ground-state of theLieb-Liniger model H =

∫dxH =

∫dx(N + T + W )

[19], which describes a gas of interacting bosons confinedin a one-dimensional continuum [20], as schematically de-picted in Fig. 1a. Here, N = −µψ†xψx is the potential en-ergy, T = ∂xψ

†x∂xψx is the kinetic energy of particles, and

∗Current address: Department of Physics, Princeton University, Princeton NJ08544, USA. Correspondence to [email protected]

m

(a)

x

cMPS(b)

control elds λ

κ

1D transmission line

φ(λ)

H

measure

vary λ to minimize Eλ

Eλ=

0 0.5 1 1.50123456

Τ Μs

G1 ΤΜs1

2Π MHz 1.052Π MHz 1.482Π MHz 2.092Π MHz 2.95

0 0.5 1 1.50.00.20.40.60.81.01.2

Τ Μs

g2Τ

(c) (d)

TW

FIG. 1: Schematic of the interacting Bose gas and the principle ofthe quantum variational algorithm. (a), Bosonic particles with ki-netic energy T are propagating in one dimension along the x axis.Repulsion between particles mediates an interaction energy W . Theparticle density ρ of the gas is controlled by the chemical potentialµ. (b), We experimentally simulate the ground-state of H by em-ploying a variational minimization procedure. A tunable cavity QEDsystem is used to generate radiation fields emulating continuous ma-trix product states |φ(λ)〉 in a 1D transmission line. The averageenergy Eλ = 〈φ(λ)|H|φ(λ)〉 of the simulated Hamiltonian is ex-perimentally determined from measured correlation functions. Ex-ternal control fields λ are used as variational parameters. (c),(d), Ex-amples of first-order G(1)(τ) and second-order g(2)(τ) correlationfunctions measured (dots) in superconducting circuits and simulated(solid lines) using a master equation approach for the indicated driverates Ω, effective anharmonicity α/2π = 5.2 MHz, and cavity decayrate κ/2π = 2.2 MHz.

arX

iv:1

508.

0647

1v1

[qu

ant-

ph]

26

Aug

201

5

Page 2: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

2

g(2)(0)G(1)(0) [ms-1] dω ω2 G(1)(ω) [ms-3]

W2 /W 0

2 [d

B]

a/2p [MHz]

12

34

5 0

5

10

150

50

100

150

200

W2 /W 0

2 [d

B]

a/2p [MHz]

12

34

5 0

5

10

150.0

0.5

1.0

W2 /W 0

2 [d

B]

a/2p [MHz]

12

34

5 0

5

10

150

10

20

30

40

(a) (b) (c)~

FIG. 2: Measured correlations in the variational space spanned by the anharmonicity α and the drive rate Ω. (a) Average photon flux G(1)(0)

proportional to the potential energy 〈N〉, (b)∫

dωω2G(1)(ω) proportional to the kinetic energy 〈T 〉, and (c) second-order correlator g(2)(0)

proportional to the interaction energy 〈W 〉. The smallest drive rate is Ω0/2π = 0.37 MHz.

W = v(ψ†x)2ψ2x is the interaction energy expressed in second

quantization by the field operator ψx. The Lieb-Liniger modelhas only one intensive parameter v = v/ρ, where ρ = 〈ψ†xψx〉is the average particle density and v the interaction strength.The ground-state energy of this model can be calculated ana-lytically using the Bethe ansatz [19]. The calculation of twopoint correlation functions requires the use of numerical meth-ods such as quantum Monte Carlo or DMRG [10]. The factthat this model is well understood makes it an ideal test caseto benchmark the as yet unexplored quantum variational algo-rithm recently proposed by Barrett et al. [12].

In the experiments presented here, we prepare continuousmatrix product states |φ(λ)〉 [10, 12] as microwave radiationfields propagating along a one-dimensional transmission line,see Fig. 1b. The radiation fields are generated by an ancillaryquantum system – in our case a tunable circuit QED system[21] – which is coupled with rate κ to the transmission line.Notably, any radiation field generated in this way is describedby a continuous matrix product state with a bond dimensiondepending on the number of participating ancillary energy lev-els [9]. We vary the quantum state |φ(λ)〉 by tuning a set oftwo external variational parameters λ = (α,Ω). Here, Ω isthe drive rate of a coherent field applied resonantly to the up-per eigenmode of the coupled system and α is the effectiveanharmonicity of the driven mode. Tunability of the effectiveanharmonicity α is experimentally achieved by employing aqubit of which both the frequency and its coupling to the cav-ity are adjustable in-situ during the experiment [22], see Ap-pendix A 5 for details.

The variational ground-state of the Lieb-Liniger Hamilto-nian H is found by measuring the expectation value Eλ =

〈φ(λ)|H|φ(λ)〉 = 〈N〉 + 〈T 〉 + 〈W 〉 and minimizing Eλwith respect to different variational states |φ(λ)〉 created inour experiment. The simulated Hamiltonian is thus solelydetermined by the measurement observable. Any model ofwhich the corresponding expectation values can be measured,is therefore accessible with this approach. Given the pho-tonic realization of |φ(λ)〉, the measurement of Eλ trans-lates into the measurement of photon correlation functions.

Spatial correlations in the field ψx are mapped onto timecorrelations in the cavity output field aout(t) by identifyingψx = aout(t = x/s)/

√s, where the scale parameter s = x/t

acts as an additional variational parameter [12]. Entanglementin the matrix product states thus corresponds to entanglementbetween photons emitted from the cavity at different times.According to this correspondence, Eλ for the Lieb-LinigerHamiltonian is given by the first- and second-order correla-tion functions G(1)(τ) ≡ 〈a†out(τ)aout(0)〉 and G(2)(τ) ≡〈a†out(0)a†out(τ)aout(τ)aout(0)〉 [23]. More specifically, theaverage kinetic energy 〈T 〉 = s−3

∫dωω2G(1)(ω) is cal-

culated from the Fourier transform of the first-order corre-lation function G(1)(ω), the interaction energy is 〈W 〉 =s−2vG(2)(0) and the potential energy is given by the averagephoton flux 〈N〉 = −s−1µG(1)(0).

The presented variational approach thus crucially relies onthe ability to generate and probe a wide range of different(quantum) radiation fields with high efficiency. For fast andreliable correlation measurements [24] we have developed aquantum-limited amplifier which allows for phase-preservingamplification at large bandwidth and high dynamic range [25].The examples of measured correlation functions shown inFig. 1c-d illustrate their dependence on the drive rate Ω atconstant α. While G(1)(τ = 0) equals the total average pho-ton flux 〈a†outaout〉, the limit G(1)(τ → ∞) is proportionalto the square of the coherence of the field |〈aout〉|2. There-fore, G(1) increases with drive rate due to the enhanced pho-ton production rate. The normalized second-order correlationfunctions g(2)(τ) ≡ G(2)(τ)/(G(1)(0))2 show anti-bunchedbehavior (g(2)(0) < 1) for weak drive and Rabi type oscilla-tions when the drive rate Ω becomes larger than the decay rate[24]. Both the measured first-order and second-order correla-tion functions are in agreement with the results obtained frommaster equation simulations (black solid lines).

We have measured ∼ 103 such correlation functions overa wide range of variational parameters Ω and α by acquiringdata for one week. Without the employed parametric ampli-fier the time for measuring this set of data would have beenon the order of years because of the exponential scaling be-

Page 3: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

3

tween statistical error and correlation order [26]. Based onthis collection of time-resolved correlation functions we haveevaluated the three relevant terms G(1)(0),

∫dωω2G(1)(ω)

and g(2)(0) that enter the calculation of Eλ, see Fig. 2. Asexpected, the average photon flux G(1)(0) (Fig. 2a) increaseswith drive rate Ω and is suppressed for increasing anharmonic-ity α. The kinetic energy term

∫dωω2G(1)(ω) in panel b is

determined by the power spectral density G(1)(ω). Only thespectral weight of photons generated at finite detuning fromthe drive frequency (|ω| > 0) contributes to the integral. Inthe Bose gas picture such photons correspond to particles infinite momentum states and therefore carrying kinetic energy.The rate of scattering events from drive photons into photonswith finite ω increases with drive strength and has a non-trivialdependence on α. Finally, the second-order correlator g(2)(0)in Fig. 2c clearly reveals the crossover from antibunched ra-diation (g(2)(0) → 0) for large α and small Ω to coherentradiation (g(2)(0)→ 1) when lowering the anharmonicity.

Based on the three measured quantities shown in Fig. 2 anda chosen interaction parameter v we evaluate the energy Eλfor all prepared states |φ(λ)〉. We identify a local minimumin variational space (blue region), which corresponds to thevariational ground-state of the Lieb-Liniger model, see Fig. 3.When changing the interaction parameter v we find the energyEλ to be minimized by a different set of parameters (α,Ω) in

05

1015

1.2 2 3 4 5

Eλ-Emin

Ω2 /

Ω 02 [d

B]

α/2p [MHz]

0 max

v 0.63

max

0.31

v 1.36

max

0.12

v 2.18

max

0.1

v 4.

max

0.12

FIG. 3: Measured energy landscape for the Lieb-Liniger model. Eλcalculated from the measurement data shown in Fig. 2 relative to itsminimum Emin as a function of α and Ω. Interaction strength v in-creases from bottom to top as indicated. The color scale is adjusted ineach panel such that the maximal value max appears red. The min-ima from bottom to top are located at (Ω/Ω0)2 ∈ 13, 10, 7, 3 dBand α/2π ∈ 1.56, 2.25, 3.43, 5.04MHz as indicated by red dots.

variational space. While for large values of v the minimumappears in the anti-bunched region (lower right corner in toppanel), the minimum moves to the region where the radiationis mostly classically coherent when weakening the interactionstrength (upper left corner in bottom panel). The maximumand minimum value of v which can be explored in this wayis practically limited by the range of variational parametersα and Ω for which correlation functions have been acquiredexperimentally.

After identifying the variational ground-state for each in-teraction strength v as the respective minimum in the energylandscape, we further investigate its properties. We com-pare the experimental results with the numerically exact re-sults obtained from a variational matrix product state algo-rithm executed on a classical computer with bond dimensionD = 14 [10]. We followed the usual convention and rescaledall quantities so that they correspond to a particle density ofρ = 1 [10]. The Lieb-Liniger ground-state energy densityELL, which is the sum of kinetic energy and interaction en-ergy, increases with interaction strength and ideally convergestowards the Tonks-Giradeau limit lim

v→∞ELL = π2/3 indicated

as a dashed line in Fig. 8a. Given the small number of varia-tional parameters the experimental data (blue dots) reproducesthe characteristic dependence of ELL on v of the exact solu-tion (red solid line) quite well.

Importantly, having physical access to the ground-statewave functions |φ(λ)〉 we can also probe quantities beyondthe ground-state energy, such as two-point correlation func-tions. First-order correlation functions 〈ψ†xψ0〉 are obtainedfrom G(1)(τ) by converting time into spatial coordinates(Fig. 8b). As expected, we observe a decrease in correlationlength with increasing interaction strength v. Due to the ab-sence of spontaneous symmetry breaking in one dimension[27], the exact ground state of the Lieb-Liniger model doesnot exhibit Bose-Einstein condensation. The observed finitelimit lim

x→∞〈ψ†xψ0〉 is a characteristic feature of matrix product

states which do not support the U(1) symmetry of the modelfor finite bond dimensions.

The nontrivial nature of the ground-state in the presenceof interactions also becomes manifest in the particle-particlecorrelator 〈ψ†0ψ†xψxψ0〉 shown in panel c. With increasingv particles are more likely to repel each other leading to anti-bunching. Our experiments clearly resolve this crossover froma weakly into a strongly interacting Bose gas by accessingvariational wave functions for interaction parameters v overtwo orders of magnitude. While this general behavior is qual-itatively well reproduced, an accurate quantitative agreementwith the numerical results would require a larger number ofindependent variational parameters in the experimental real-ization. This becomes apparent when comparing the experi-mental results with numerical calculations based on continu-ous matrix product states of different bond dimensions,D = 2and D = 14, where the number of variational parameters is2D2. Correlation functions simulated with low bond dimen-sion D = 2 deviate from the exact results (D = 14) similarlyas the measured ones (Fig. 8d-g).

In summary, we have experimentally revealed connectionsbetween open quantum systems, the matrix product state vari-ational class, and quantum field theories that can be used forpractical quantum simulations. The presented quantum varia-tional algorithm is general in the sense that it can be applied

Page 4: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

4

v 0.06v 0.14v 0.38v 0.73

v 1.42v 3.09v 6.44

0.0

0.2

0.4

0.6

0.8

1.00 1 2 3 4 5

Ψx† Ψ

0

particle distance x

0.1 0.5 1 2 5 10

0.1

0.2

0.5

1

2

5

0.1 0.5 1 2 5 10

v

ELL

v(b)

0 1 2 3 4 50.0

0.2

0.4

0.6

0.8

1.0

particle distance x

Ψ0† Ψ

x† ΨxΨ

0

(a)

experimental

D=2

0 1 2 3 4 5

0.0

0.2

0.4

0.6

0.8

1.0

particle distance x

Ψx† Ψ

0

0 1 2 3 4 50.0

0.2

0.4

0.6

0.8

1.0

particle distance x

Ψ0† Ψ

x† ΨxΨ

0

0 1 2 3 4 50.0

0.2

0.4

0.6

0.8

1.0

particle distance x

Ψ0† Ψ

x† ΨxΨ

0

0 1 2 3 4 5

0.0

0.2

0.4

0.6

0.8

1.0

particle distance x

Ψx† Ψ

0

D=14(392 var. parameters)

experimental D=14

D=2(8 variational parameters)

(c)

(d)

(e)

(f)

(g)

FIG. 4: Comparison between experimental simulation and numerical result. (a) Lieb-Liniger ground-state energy density ELL vs. interactionparameter v at constant particle density ρ = 1 on a log/log scale. (b)-(c) Experimentally obtained first-order 〈ψ†xψ0〉 and particle-particlecorrelation functions 〈ψ†0ψ†xψxψ0〉 for the seven indicated interaction strengths v. (d)-(g) Corresponding numerical solutions using continuousmatrix product states with bond dimensions D = 2 and D = 14.

to any one-dimensional quantum field theory. Exploring inter-acting vector field models seems particularly appealing, sincethey are difficult to simulate on classical computers. Experi-mentally, this could be achieved by coupling tunable quantumsystems to multiple transmission lines each representing oneof the vector field components [28]. Higher accuracy in thesimulation will require more variational parameters and quan-tum systems with more degrees of freedom, which is achiev-able with tunable superconducting circuits. In this context, thecollective dissipation of multiple emitters coupled to the sametransmission line at finite distance [29] may turn out very use-ful. Extensions of the presented approach may be envisaged toalso explore dynamical phenomena and discrete lattice mod-els using experimentally created matrix product states.

We thank Christoph Bruder, Ignacio Cirac, TilmanEsslinger and Atac Imamoglu for discussions and comments.This work was supported by the European Research Coun-cil (ERC) through a Starting Grant, by the NCCR QSIT andby ETHZ. CE acknowledges support by Princeton Universitythrough a Dicke fellowship. TJO was supported by the ERCgrants QFTCMPS and SIQS, and by the cluster of excellenceEXC201 Quantum Engineering and Space-Time Research.

Appendix A: Experimental details

1. Measurement setup, sample fabrication andcharacterization

The experiments presented in the main text are performedwith a device consisting of a superconducting circuit, seeFig. 5a for details about the experimental setup. The sample

(Fig. 5b) consists of a λ/2 transmission line cavity with a res-onance frequency ωres/2π ≈ 7.3425 GHz of the fundamentalmode. The resonator has one output port which dominatesthe total decay rate κ/2π ≈ 2.2 MHz and one weakly cou-pled input port κin/κ ≈ 0.01 which is used for coherent driv-ing of the cavity field. The resonator is fabricated using pho-tolithography and reactive ion etching of a Niobium thin filmsputtered on a sapphire wafer. We have fabricated a supercon-ducting qubit (Fig. 5c) close to one end of the resonator. Bothits transition frequency ωge and coupling strength g to the res-onator are tunable by varying the magnetic fluxes threadingthe two SQUID loops [22, 30]. Flux control is achieved by acombination of a superconducting coil mounted on the back-side of the sample holder and a local flux line which couplespredominantly to one of the two SQUIDs. The currents feed-ing the coil and the flux line are generated by voltage biasedresistors at room temperature. The qubit is fabricated usingdouble-angle evaporation of aluminum on a mask defined byelectron beam lithography. The qubit decay and dephasingtimes are measured to be T1 ≈ 1.8µs and T2 ≈ 1.2µs. Theanharmonicity of the qubit is (ωef − ωge)/2π ≈ −80 MHz,where ωef is the transition frequency from the first excitedstate |e〉 to the second excited state |f〉 of the qubit.

The sample is mounted on the base plate of a dilution re-frigerator cooled down to a temperature of about 20 mK. Thequbit and the resonator are coherently driven through atten-uated charge control lines. The microwave radiation emit-ted from the cavity is guided through two circulators to aJosephson parametric dimer (JPD), which provides quantum-limited amplification at large bandwidth and dynamic range[25, 31, 32]. A directional coupler is used to apply and in-terferometrically cancel the reflected pump field. The ampli-

Page 5: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

5

T = 300K

+ -

10dB

+ -

LP lter

resonatordrive

20dB

20dB

20dB

qubitdrive

20dB

20dB

20dB

pump

20dB

20dB

20dB

20dB

+ f

20dB

HEMT

BP lter4-8 GHz

JPD amplier

LP lter

BP lter4-8 GHz

LO

RFLP lter

DCblock

ADC

FPGA board

LOfRF+ 25MHz

25MH

z

X(t)

P(t)

processing

RAM

4K

800mK

100mK

20mK

(a) Legend

circulator

isolator

attenuator

low passlter

fvariablephase shifter

+ splitter

IQ mixer

signalgenerator

amplier

directionalcoupler

band passlter

+ - batteryvoltage source

(b) 1 mm

150 mm

SQUID1 SQUID2

(c)

inout

(d) (e)

uxcontrol

chargecontrol

100 mm

20 mm

FIG. 5: (a) Schematic of the measurement setup. For details see text. (b) Optical micrograph of the sample. The second qubit gap in the leftpart of the chip is left empty. Enlarged images of the output capacitor (c), of the qubit (d), and of the input capacitor (e) are shown.

fied signal reflects back from the JPD, passes a bandpass (BP)filter, is further amplified by a high electron mobility transis-tor (HEMT) amplifier, and is down-converted to an interme-diate frequency (IF) of 25 MHz. After low-pass (LP) filteringand IF amplification the down-converted signal is digitized us-ing analog-to-digital conversion (ADC) and further processedwith field programmable gate array (FPGA) electronics.

2. Controlling the qubit frequency and the coupling strength

We characterize the coupled cavity-qubit system by probingthe transmission coefficient of the cavity and fitting the datato the absolute square of the expression

t =Aκ

i(ωres − ω) + g2

i(ωge−ω)+γ/2 + κ/2, (A1)

which we obtain from input-output theory for the Jaynes-Cummings model [33]. Here, ω is the probe frequency, γ is

Page 6: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

6

7.335 7.340 7.345 7.3500

2

4

6

8

10

Frequency GHz

norm

aliz

edtr

ansm

issi

on

7.339 7.340 7.341 7.342 7.343 7.344 7.3450

2

4

6

8

10

Frequency GHz

tran

smis

sion

(a) (b)

norm

aliz

ed tr

ansm

issi

on

norm

aliz

ed tr

ansm

issi

on

w/2p w/2p

|t|2

|t|2

FIG. 6: (a) Measurements and fits of the absolute square of the transmission coefficient |t|2 for varying qubit frequency at approximatelyconstant coupling strength g/2π = 1.75 MHz. Individual data traces are offset from each other by one. (b) Transmission coefficient |t|2 forvarying coupling strength g at constant qubit frequency ωge ≈ ωres.

the qubit decoherence rate andA is a scaling factor. The probepower is chosen such that the average number of excitations ofthe coupled resonator qubit system is much smaller than one.In this case the qubit may be approximated by an harmonicoscillator. We determine the qubit detuning ∆ = ωres − ωgeand its coupling strength g to the cavity by fitting spectro-scopically obtained transmission data to the above model, seeFig. 6. The magnetic fluxes through the qubit SQUID loopsand with that the qubit parameters g and ∆ are controlled bya pair of voltages V1 and V2 applied to coil bias resistors. Forsmall g and ∆ we approximate the relation between (g,∆)and (V1, V2) by linear equations of the form(

g∆

)=

(m11 m12

m21 m22

)(V1 − V1,0V2 − V2,0

). (A2)

We determine the coupling matrix elementsmij and the offsetvoltages V1,0 and V2,0 by recording transmission data for pairs(V1, V2)k. For each set of data we extract the correspond-ing parameters (g,∆)k and perform a least-square fit to deter-mine the model parameters mij , V1,0 and V2,0. By invertingEq. (A2) we calculate the voltages V1 and V2 for a given set ofdesired qubit parameters (g,∆). In order to further fine-tunethe parameters, we have developed an automated calibrationalgorithm which minimizes the deviations from desired tar-get values by iteratively measuring, fitting and readjusting thecontrol parameters V1 and V2. With this control procedure weare able to independently set the qubit frequency and the in-teraction strength. We also use this procedure to monitor andcorrect for slow qubit frequency drifts occurring during longruns of the experiment.

We demonstrate individual control of qubit parameters byeither tuning the qubit frequency for approximately constantcoupling strength g (Fig. 6a) or by varying g for fixed qubitfrequency (Fig. 6b). For all sets of data we have turned on theJPD amplifier and have divided out its frequency dependentgain. For the measurements in Fig. 6b we have kept the qubitresonant with the cavity (∆ = 0) and have varied g. Thesemeasurements demonstrate the ability to tune the system from

the fast cavity (κ g γ) into the strong coupling regime(g γ, κ) [22].

3. Josephson parametric dimer amplifier

To measure higher order correlation functions efficientlywhile retaining a high level of linearity we employ a Joseph-son parametric dimer (JPD) amplifier. For details about theoperational principles of the JPD we refer the reader to [25].The gain G measured vs. signal frequency f is approximatelydescribed by a Lorentzian function (Fig. 7). The amplifierbandwidth at full width half maximum is approximately 35MHz. In order to increase dynamic range, we have chosen amoderate maximum gain of 17 dB. A measurement of the gainas a function of signal power Pin results in a 1 dB compres-sion point at about -107 dBm which corresponds to a photonflux of 3000 µs−1, see Fig. 7b. The largest photon flux whichis generated in the described experiments is less than 50 µs−1.The JPD amplifier is thus far away from its compression pointfor all measured correlation functions. The improvement indetection efficiency becomes apparent when measuring thenoise power spectral density Sδ in units of photons per Hzper second when the JPD amplifier is turned ON and when itis turned OFF. The effective noise level, which is referencedback to the input of the JPD amplifier is decreased by morethan an order of magnitude when it is turned on. The scalingof Sδ is based on a comparison between the frequency depen-dent gain Gδ and the JPD amplifier noise [34]. The deviationfrom the quantum limit is due to the additional noise from thefollowing HEMT amplifier which is comparable to the noiseat the output of the JPD. The equivalent detection efficiencyof the amplification chain is ηamp = 1/Sδ ≈ 50%.

Page 7: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

7

(a) (b)

150 130 110

0.1 11 10 100 103 104

10

12

14

16

18

20

P in dBm

photon ux Pin/hf Μs1

Gai

nd

B

7.28 7.34 7.40

5

10

15

Frequency f GHz

Gai

n Gd

B

10 5 0 5 100.5

21

51020

50100

2Π GHz

pow

er s

pect

ral d

ensi

ty S

detuning δ

δ

(c)

ON

OFF

FIG. 7: (a) Measured gain of the JPD amplifier (blue dots) and a Lorentzian fit (solid red line). (b) Measured JPD gain as a function of signalpower and equivalent photon flux. (c) Measured noise power spectral density Sδ referenced back to the input of the JPD amplifier in units ofphotons per Hz per second when the JPD is turned on (blue) and when it is turned off (red). The detuning δ is relative to the frequency ofa coherent test tone applied to the JPD input at frequency 7.35 GHz. The dashed line indicates the quantum noise limit for phase-preservingamplification and detection.

4. Calibration of drive rate and output power

We have calibrated the total gain of the detection chain in-cluding all cable losses in order to reference the measuredphoton flux back to the output of the cavity. Calibrating thetotal gain of the detection chain is equivalent to calibrating itsdetection efficiency ηtot. The detection efficiency is typicallylimited by losses between the cavity and the first amplifier(ηloss) and by noise added during the amplification process(ηamp), see Fig. 8. We compare the nonlinear response of the

(a)

(b)

hloss

JPD HEMT

hamp

htot

0 5 10 150.0

0.1

0.2

0.3

0.4

02 dB

a† a

anda2

ata

| a |2

FIG. 8: (a) The detection efficiency of the cavity output field is typi-cally limited by radiation loss, schematically represented as a beam-splitter with finite transmittivity ηloss, and by noise Sδ = η−1

amp addedin the amplification chain. The total detection efficiency is givenby the product ηtot = ηlossηamp. (b) Measurement (points) andfit (solid line) of the cavity photon number and absolute square ofthe coherent amplitude for the coupled cavity-qubit system driventhrough the cavity input port at rate Ω. The smallest drive rate isΩ0/2π = 0.37 MHz.

coupled cavity-qubit system with master equation simulationsto perform this calibration. We bias the qubit with parameters(∆, g)/2π = (3, 5.7) MHz and apply a drive field to the inputport of the cavity at rate Ω and resonant with the frequencyω+ = ωge + ∆/2 +

√g2 + ∆2/4 = 2π × 7.35 GHz of the

upper Jaynes-Cummings doublet state. For these settings wemeasure the coherent photon flux κ|〈a〉|2 and the total photonflux κ|〈a†a〉|2 emitted from the cavity for different drive ratesΩ. We fit these data sets to the results obtained from masterequation simulations leaving the the total gain factor of themeasurement chain and the absolute drive rate incident to thesample as free parameters (Fig. 8b). The equivalent detectionefficiency resulting from this fit is equal to the inverse of thescaled noise level and found to be ηtot = ηlossηamp = 0.27.Together with the estimate for the efficiency of the amplifi-cation chain ηamp stated in the previous section, we extract aradiation loss of 1 − ηloss = 0.46 between the cavity and theJPD, which is reasonable given the components and cablesconnecting the two stages.

5. Drive scheme and variational parameters

In the original proposal for simulating the Lieb-Linigermodel with cavity QED it has been suggested to keep thequbit resonant with the cavity and use the qubit drive powerΩq and the coupling strength g as two variational parameters.We have experimentally realized this scheme and found that inthe limit of small coupling strengths g the total emission ratebecomes extremely small which in turn limits the signal tonoise ratio. This is because the effective emission bandwidthscales like g2/κ when g < κ and thus decreases quadraticallywith g. We have therefore developed an alternative schemefor which the photon emission rate remains proportional to κeven in the limit of small g. We have therefore made use ofthe ability to tune both the qubit frequency and the couplingstrength. Rather than keeping the qubit at fixed frequency weadjust for each value of g the detuning ∆ such that ω+ re-mains at constant frequency resonant with the drive frequencyω+ = ωd = 2π × 7.35 GHz. The spectroscopy data for the

Page 8: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

8

(a) (b)

(c)

7.25 7.26 7.27 7.28 7.29 7.3 7.31 7.32 7.33 7.34 7.350

5

10

15

20

25

30

Frequency GHz

spec

tros

copy

data

arb.

units

slice for bpkt10 5 10 15 20 25

80

60

40

20

0

g 2Π MHz

2ΠM

Hz

5 10 15 20 251

2

3

4

5

g 2Π MHz

anha

rmon

icityΑ

2ΠM

Hz

|g, 0

√–2|2+

|2–

2g|1+

|1–

x 2g

drive

D = 0 D < 0a

qubit spectroscopy res. spectroscopy coupling strength

detu

ning

probe frequency f [GHz]

FIG. 9: (a) Spectroscopy data for all values of g used in the quantum simulation experiment. The blue data points on gray background areobtained using resonator transmission measurements. The red data points are obtained using qubit spectroscopy measurements. The relativescaling and offsets are adjusted to display the data in the same plot. The inset shows the energy levels |n±〉 for the Jaynes Cummings model.(b) Pairs of (g,∆) extracted from the data shown in a using the fitting routine described before. (c) Effective anharmonicity of the upperbranch of the Jaynes Cummings ladder calculated from the pairs (g,∆) shown in b.

qubit bias points used in the quantum simulation experimentare shown in Fig. 9a and demonstrate constant ω+ over theentire range of coupling strengths. In order to keep ω+ con-stant, we compensate the larger splitting when increasing g bytuning the qubit further away from the cavity, see Fig. 9b.

For this specific tuning scheme we find that the effectiveanharmonicity α of the upper Polariton ladder decreases withincreasing coupling strength g, as shown in Fig. 9c. To il-lustrate this effect we show a schematic drawing of the en-ergy levels |n±〉 of the Jaynes Cummings model for the res-onant case (∆ = 0) and for the case of finite qubit detuning∆ < 0. The inverse proportionality between anharmonicityα and coupling strength g illustrates the appearance of anti-bunched behavior for the small values of g and the observedcoherent radiation for large g values for this bias scheme, seeFig. 2 of the main text. The drive rate Ω of a coherent fieldapplied to the cavity input port acts as a second variationalparameter for the quantum simulation.

6. Measurement of correlation function and master equationsimulation

We employ fast real-time signal processing performed withan FPGA at a clock rate of 100 MHz for the measurementof time-resolved correlation functions. The cavity output

field is processed, as described in section 1A. After digi-tization we multiply the sampled voltages with digital sineand cosine waves of frequency ωIF/2π = 25 MHz to ob-tain the quadrature components I(t) and Q(t), respectively.We then apply an FIR filter with an effective bandwidth ofΓ/2π = 10 MHz to the quadratures in time-domain to obtainthe filtered quadratures I(t) and Q(t), which in the follow-ing we write as the single complex valued amplitude S(t) =

I(t) + iQ(t). To measure the first-order correlation functionin S(t) we take the discrete Fourier transform (F) of M timetraces Si(t) of length 10.24 µs, multiply with their complexconjugate, and average

Γ(1)(τ) = F−1[

1

M

M∑i=1

F [Si(t)]F∗[Si(t)]

].

Similarly we extract the second-order correlation function bycalculating the absolute square of S(t) before Fourier trans-forming

Γ(2)(τ) = F−1[

1

M

M∑i=1

F [S∗i (t)Si(t)]F∗[S∗i (t)Si(t)]

].

We record each of these quantities with the drive field turnedon, giving Γ

(1)ON(τ),Γ

(2)ON(τ), and with the drive turned off,

giving Γ(1)OFF(τ),Γ

(2)OFF(τ). To avoid effects due to slow

Page 9: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

9

drifts we alternate between all four measurements every12.5 µs. As explained in detail in reference [26] and asdemonstrated in many experiments since then [24, 35, 36],we can use these four measurements to extract the corre-lation functions G(1)(τ) = κ〈a†(τ)a(0)〉 and g(2)(τ) =〈a†(0)a†(τ)a(τ)a(0)〉/〈a†a〉2 of the output field of the cavity.In these expressions, a (a†) is the annihilation (creation) oper-ator of the intra-cavity field. For the cases in which the aver-age photon number is small (〈a†a〉) 1 we find g(2)(0) val-ues which are systematically smaller than the correspondingmaster equation simulation. We attribute this to a weak ther-mal background radiation during the off measurements whichwe correct for [37]. Good agreement between the measuredand simulated correlation functions is found when correctingfor a thermal photon flux of nth ≈ 0.03/µs in the detectionband.

We compare these measurements with correlation functionsobtained from master equation simulations. For these simula-tions we describe the system by the Hamiltonian

Hsys/~ = (ωres − ωd)a†a+ (ωge − ωd)b†b

+αq2

(b†)2b2 + g(a†b+ ab†) (A3)

expressed in a frame rotating at the drive frequency ωd/2π =7.35 GHz. Here, b and b† are annihilation and creation op-erators for an excitation of the transmon. In addition to that,we account for qubit decay γ, qubit dephasing γφ and res-onator emission κ with standard Lindblad terms. Simulationsare run in a Hilbert space including 6 resonator and 3 trans-mon levels. In order to account for the finite detection band-width when simulating the second-order correlation functionwe employ the techniques described in [38]. In this approachwe introduce an ancillary mode c of frequency ωd, which isweakly coupled with rate ε/2π = 20 kHz to the cavity and

decays with a rate equal to the detection bandwidth Γ. Thesecond-order correlation function in c is then simulated basedon the total Liouvillian and taken as an estimate for the filteredcorrelation function of mode a.

Appendix B: Theoretical aspects and data analysis

1. Calculation of the Lieb-Liniger energy from correlationfunctions

The expectation value to be minimized 〈H〉 = 〈T 〉+〈W 〉+〈N〉 is composed of the kinetic energy of the bosons 〈T 〉, itsinteraction energy 〈W 〉 and the potential energy 〈N〉. Accord-ing to the correspondence between the field operator ψx andthe time-dependent radiation field a(t) =

√sψx=st, each of

these expectation values is proportional to a specific measuredcorrelation function

〈T 〉 =1

s3

∫dωG(1)(ω)ω2,

〈W 〉 =v

s2G(2)(0),

〈N〉 = −µsG(1)(0) = −µρ. (B1)

Here, G(1)(ω) ≡ F [G(1)(τ)] is the Fourier transformof the first-order correlation function normalized such that∫

dωG(1)(ω) = G(1)(0). The energy terms in Eq. (B1) thusexplicitly depend on the scaling parameter s, which is to betreated as an additional variational parameter. We explicitlyminimize 〈H〉 with respect to s by solving ∂

∂s 〈H〉 = 0, whichresults in

s =3∫

dωG(1)(ω)ω2

−vG(2)(0) +

√(G(2)(0)

)2v2 + 3µG(1)(0)

∫dωG(1)(ω)ω2

. (B2)

Correspondingly we obtain an explicit expression for Eλ =

〈H〉, which only depends on the measured correlation func-tions and the model parameters µ, v. We find the variationalground state for a given set of model parameters (v, µ) byminimizing Eλ with respect to α and Ω.

2. Scaling transformation

We follow the usual convention and study the Lieb-Linigerground state subject to the constraint that its particle density ρis equal to one. Using the procedure described in Sec. B 1 wefind a ground state which generally does not obey this prop-erty. We therefore apply a scale transformation by adjustingthe chemical potential µ such that ρ→ 1. Under the followingtransformation

(µ, v)→ (µ, v) = (µy2, vy),

the ground state remains invariant up to a change in the param-eter s→ s = s/y, which immediately follows from Eq. (B2).Any variational ground state can therefore be transformed intoanother ground state satisfying ρ = 1, by choosing y appro-priately. We apply the following procedure to perform thisscale transformation:

• Chose parameter v, set µ = 1, and find the set of varia-tional parameters (smin,Ωmin, αmin) minimizing Eλ.

• Evaluate the particle density ρ = G(1)(0)/smin at thisminimum.

• Calculate the new chemical potential µ = µ/ρ2 and thenew interaction parameter v = v/ρ. The new scalingparameter becomes smin = sminρ = G(1)(0).

• The variational parameters (smin,Ωmin, αmin) specify

Page 10: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

10

the variational ground state for the model with interac-tion strength v and unit particle density.

Ground states for different interaction parameters are obtainedby starting the above procedure with a different value for v.The Lieb-Liniger energyELL at interaction strength v (Fig. 4aof the main text) is given by

ELL = 〈T 〉+ 〈W 〉

=1

s3min

∫dωG(1)(ω)ω2 +

v

s2ming(2)(0)

(G(1)(0)

)2=(G(1)(0)

)−3 ∫dωG(1)(ω)ω2 + vg(2)(0). (B3)

Correlation functions for the Lieb-Liniger model are directlyobtained from the measured correlation functions by identify-ing

〈ψxψ0〉 =G(1)(τ = x/smin)

G(1)(0), (B4)

and analogously for the second-order correlation function

〈ψ†xψ†0ψ0ψx〉 = g(2)(τ = x/smin). (B5)

3. Numerically exact solution

The exact solution [19, 39], via the Bethe ansatz, of theLieb-Liniger model at unit density is only possible for onespecific value of the interaction parameter, namely v = 2. Inorder to calculate properties of the model at unit density forother values of v it is necessary to take recourse to numericalmethods. We exploited a variational method over translationinvariant continuous matrix product states (cMPS) [10, 11,40]

|Ψ[Q,R]〉 ≡ tr

(P exp

[∫ ∞−∞

Q⊗ I +R⊗ ψ†x dx])|Ω〉,

(B6)where P exp denotes the path ordering of the argument fromleft to right for increasing values of x. The operators Q andR act on an auxiliary space CD, and |Ω〉 is the Fock vacuum.The variational parameters specifying the cMPS are preciselythe two D ×D matrices Q and R. The Lieb-Liniger hamilto-nian H (in the presence of a chemical potential) is comprisedof three terms H = T + W + N , and the expectation valuesof these three terms can be readily computed [40] in terms ofthe variational parameters according to

〈Ψ[Q,R]|T |Ψ[Q,R]〉 = tr([Q,R]†[Q,R]ρss

), (B7)

〈Ψ[Q,R]|W |Ψ[Q,R]〉 = v tr(R†

2R2ρss

), and (B8)

〈Ψ[Q,R]|T |Ψ[Q,R]〉 = −µ tr(R†Rρss

), (B9)

where ρss is the solution of the matrix equation

0 = −i[K, ρ] +RρR† − 1

2R†R, ρ, (B10)

and K = iQ + i2R†R. In order to find the variational mini-

mum of

E[v, µ;Q,R] ≡ 〈Ψ[Q,R]|H(v, µ)|Ψ[Q,R]〉

= tr(

[Q,R]†[Q,R] + vR†2R2 − µR†R

ρss

)(B11)

with respect to Q and R a tangent-plane method using thetime-dependent variational principle (TDVP) in imaginarytime was exploited [41]. This method proceeds as follows.Firstly, v and µ are selected. Then a value D as large as pos-sible is chosen. An initial guess |Ψ[Q0, R0]〉 for the groundstate results from random choice of Q0 and R0. Also a toler-ance η and a step size δ is selected. Set j = 0 and performthe following sequence of operations until the desired conver-gence is reached.

1. Calculate the gradient∇z|Ψ[Qj , Rj ]〉, where z is a 2D2

vector containing the entries of Qj and Rj in, say, lex-icographic order. Thus ∇z|Ψ[Qj , Rj ]〉 is a vector of2D2 cMPS states.

2. Calculate the gradient ∇zE[Qj , Rj ] of the energy ex-pecation value.

3. Calculate the inverse G−1 of the Gram matrix Gz,z′ ≡(∇z|Ψ[Qj , Rj ]〉)†∇z′ |Ψ[Qj , Rj ]〉.

4. Set zj+1 = zj − δ∑2D2

z′=1G−1z,z′∇z′E[Qj , Rj ].

5. Set j = j + 1, unpack zj+1 into the two matrices Qj+1

and Rj+1, and repeat step (1) until convergence of theenergy expectation values reaches the prespecified tol-erance η.

After the above algorithm has terminated the cMPS corre-sponding to unit density (and at the rescaled interaction pa-rameter) is obtained via the rescaling procedure described inthe previous section. This method was used to calculate acMPS representation for the Lieb-Liniger ground state acrossa range of interaction parameters from v = 0 to v = 1000.The value D = 14 was used throughout as the results so ob-tained are indistinguishable from the known exact solutions atv = 0, 2,∞. For the other values of the interaction param-eter the accumulated errors were estimated and found to benegligible.

[1] S. White, Phys. Rev. Lett. 69, 2863 (1992).[2] U. Schollwock, Rev. Mod. Phys. 77, 259 (2005).[3] S. Ostlund and S. Rommer, Phys. Rev. Lett. 75, 3537 (1995).

[4] J. Dukelsky, M. A. Martn-Delgado, T. Nishino, and G. Sierra,EPL (Europhysics Letters) 43, 457 (1998).

[5] F. Verstraete, V. Murg, and J. Cirac, Advances in Physics 57,

Page 11: arXiv:1508.06471v1 [quant-ph] 26 Aug 2015

11

143 (2008), http://dx.doi.org/10.1080/14789940801912366 .[6] Z. Landau, U. Vazirani, and T. Vidick, Nat Phys 11, 566 (2015).[7] M. B. Hastings, Journal of Statistical Mechanics: Theory and

Experiment 2007, P08024 (2007).[8] J. Eisert, M. Cramer, and M. Plenio, Rev. Mod. Phys. 82, 277

(2010).[9] C. Schon, E. Solano, F. Verstraete, J. Cirac, and M. Wolf, Phys.

Rev. Lett. 95, 110503 (2005).[10] F. Verstraete and J. I. Cirac, Phys. Rev. Lett. 104, 190405

(2010).[11] T. J. Osborne, J. Eisert, and F. Verstraete, Phys. Rev. Lett. 105,

260401 (2010).[12] S. Barrett, K. Hammerer, S. Harrison, T. E. Northup, and T. J.

Osborne, Phys. Rev. Lett. 110, 090501 (2013).[13] I. Buluta and F. Nori, Science 326, 108 (2009),

http://www.sciencemag.org/cgi/reprint/326/5949/108.pdf.

[14] M. Greiner, O. Mandel, T. Esslinger, T. W. Hansch, andI. Bloch, Nature 415, 39 (2002).

[15] A. Friedenauer, H. Schmitz, J. T. Glueckert, D. Porras, andT. Schaetz, Nat. Phys. 4, 757 (2008).

[16] K. Kim, M.-S. Chang, S. Korenblit, R. Islam, E. E. Edwards,J. K. Freericks, G.-D. Lin, L.-M. Duan, and C. Monroe, Nature465, 590 (2010).

[17] B. P. Lanyon, C. Hempel, D. Nigg, M. Muller, R. Ger-ritsma, F. Zahringer, P. Schindler, J. T. Barreiro, M. Rambach,G. Kirchmair, M. Hennrich, P. Zoller, R. Blatt, and C. F. Roos,Science 334, 57 (2011).

[18] J.-P. Brantut, J. Meineke, D. Stadler, S. Krin-ner, and T. Esslinger, Science 337, 1069 (2012),http://www.sciencemag.org/content/337/6098/1069.full.pdf .

[19] E. Lieb and W. Liniger, Phys. Rev. 130, 1605 (1963).[20] B. Paredes, A. Widera, V. Murg, O. Mandel, S. Folling, I. Cirac,

G. V. Shlyapnikov, T. W. Hansch, and I. Bloch, Nature 429, 277(2004).

[21] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S. Huang,J. Majer, S. Kumar, S. M. Girvin, and R. J. Schoelkopf, Nature431, 162 (2004).

[22] S. J. Srinivasan, A. J. Hoffman, J. M. Gambetta, and A. A.Houck, Phys. Rev. Lett. 106, 083601 (2011).

[23] C. W. Gardiner and P. Zoller, Quantum Noise (Springer, 1991).[24] C. Lang, D. Bozyigit, C. Eichler, L. Steffen, J. M. Fink, A. A.

Abdumalikov Jr., M. Baur, S. Filipp, M. P. da Silva, A. Blais,and A. Wallraff, Phys. Rev. Lett. 106, 243601 (2011).

[25] C. Eichler, Y. Salathe, J. Mlynek, S. Schmidt, and A. Wallraff,Phys. Rev. Lett. 113, 110502 (2014).

[26] M. P. da Silva, D. Bozyigit, A. Wallraff, and A. Blais, Phys.Rev. A 82, 043804 (2010).

[27] P. C. Hohenberg, Phys. Rev. 158, 383 (1967).[28] F. Quijandrıa, J. J. Garcıa-Ripoll, and D. Zueco, Phys. Rev. B

90, 235142 (2014).[29] A. van Loo, A. Fedorov, K. Lalumiere, B. Sanders, A. Blais,

and A. Wallraff, Science 342, 1494 (2013).[30] J. M. Gambetta, A. A. Houck, and A. Blais, Phys. Rev. Lett.

106, 030502 (2011).[31] M. A. Castellanos-Beltran, K. D. Irwin, G. C. Hilton, L. R.

Vale, and K. W. Lehnert, Nat. Phys. 4, 929 (2008).[32] R. Vijay, M. H. Devoret, and I. Siddiqi, Rev. Sci. Instrum. 80,

111101 (2009).[33] C. W. Gardiner and P. Zoller, Quantum Noise 2nd ed (Springer,

Berlin Heidelberg, 1999).[34] C. Eichler, D. Bozyigit, C. Lang, M. Baur, L. Steffen, J. M.

Fink, S. Filipp, and A. Wallraff, Phys. Rev. Lett. 107, 113601(2011).

[35] D. Bozyigit, C. Lang, L. Steffen, J. M. Fink, C. Eichler,M. Baur, R. Bianchetti, P. J. Leek, S. Filipp, M. P. da Silva,A. Blais, and A. Wallraff, Nat. Phys. 7, 154 (2011).

[36] I.-C. Hoi, T. Palomaki, J. Lindkvist, G. Johansson, P. Delsing,and C. M. Wilson, Phys. Rev. Lett. 108, 263601 (2012).

[37] C. Lang, C. Eichler, L. Steffen, J. M. Fink, M. J. Woolley,A. Blais, and A. Wallraff, Nat. Phys. 9, 345 (2013).

[38] E. del Valle, A. Gonzalez-Tudela, F. P. Laussy, C. Tejedor, andM. J. Hartmann, Phys. Rev. Lett. 109, 183601 (2012).

[39] E. H. Lieb, Phys. Rev. 130, 1616 (1963).[40] J. Haegeman, J. I. Cirac, T. J. Osborne, and F. Verstraete, Phys.

Rev. B 88, 085118 (2013).[41] J. Haegeman, J. I. Cirac, T. J. Osborne, I. Pizorn, H. Verschelde,

and F. Verstraete, Phys. Rev. Lett. 107, 070601 (2011).