\cftsetindents

part0.0in0.45in \cftsetindentssection0.285in0.3in \cftsetindentssubsection0.4in0.35in \cftsetindentssubsubsection0.75in0.35in

Energy-participation quantization of Josephson circuits

Zlatko K. Minev1,∗, Zaki Leghtas1,2, Shantanu O. Mundhada1,†, Lysander Christakis1,‡, Ioan M. Pop1,3, Michel H. Devoret1 1Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA 2Centre Automatique et Systèmes, Mines-ParisTech, PSL Research University, 60 Bd Saint Michel, 75006 Paris, France 3IQMT, Karlsruhe Institute of Technology, 76344 Eggenstein-Leopoldshafen, Germany Current address: IBM T.J. Watson Research Center, Yorktown Heights, New York 10598, USA; zlatko.minev@aya.yale.edu; www.zlatko-minev.com Current address: Quantum Circuit Incorporated (QCI), New Haven, CT 06511, USA Current address: Department of Physics, Princeton University, Princeton, NJ 08540, USA
(March 2, 2024)
Abstract

Superconducting microwave circuits incorporating nonlinear devices, such as Josephson junctions, are a leading platform for emerging quantum technologies. Increasing circuit complexity further requires efficient methods for the calculation and optimization of the spectrum, nonlinear interactions, and dissipation in multi-mode distributed quantum circuits. Here, we present a method based on the energy-participation ratio (EPR) of a dissipative or nonlinear element in an electromagnetic mode. The EPR, a number between zero and one, quantifies how much of the mode energy is stored in each element. The EPRs obey universal constraints and are calculated from one electromagnetic-eigenmode simulation. They lead directly to the system quantum Hamiltonian and dissipative parameters. The method provides an intuitive and simple-to-use tool to quantize multi-junction circuits. We experimentally tested this method on a variety of Josephson circuits, and demonstrated agreement within several percents for nonlinear couplings and modal Hamiltonian parameters, spanning five-orders of magnitude in energy, across a dozen samples.

I INTRODUCTION

Quantum information processing based on the control of microwave electromagnetic fields in Josephson circuits is a promising platform for both fundamental physics experiments and emerging quantum technologies (Devoret2013, ; GoogleSupremacy2019, ; Blais2020, ). Key to the success of this platform is the ability to quantitatively model the distributed quantized electromagnetic modes of the system, their nonlinear interactions, and their dissipation (see Fig. 1). This challenge is the subject of intensifying interest (Nigg2012, ; Bourassa2012, ; Solgun2014, ; Solgun2015, ; Smith2016, ; Gely2017, ; Malekakhlagh2017-Cutoff-Free, ; Pechal2017, ; Parra-Rodriguez2018, ; Parra-Rodriguez2018a, ; Ansari2018, ; Krupko2018, ; Malekakhlagh2018, ; Solgun2017, ; Petrescu2019, ; You2019-Koch, ; DiPaolo2019, ; Menke2021, ; Gely2020, ; Kyaw2020, ; Yan2020, ; Minev2021-lom, ; qiskit-metal-code, ), as experimental architectures (Barends2013, ; Minev2016, ; Brecht2016, ; Reagor2016-cavity, ; Gambetta2017, ; Rosenberg2017, ; Versluis2017, ; Naik2017, ; Krantz2019, ; Kjaergaard2019, ) and nonlinear devices (Josephson1962, ; Vijay2010, ; Kerman2010, ; Larsen2015, ; DeLange2015-nanowire, ; Janvier2015-atomic-point, ; Maleeva2018, ; Wang2018, ) scale in both complexity and diversity.

In this paper, we introduce a circuit quantization method based on the concept of the energy-participation ratio (EPR). We reduce the quantization problem to answering the simple question: what fraction of the energy of mode m𝑚m is stored in element j𝑗j? This leads to a constrained number between zero and one, the EPR, denoted pmjsubscript𝑝𝑚𝑗p_{mj} (Minev2019-Thesis, ). This ratio is the key quantity that bridges classical and quantum circuit analysis; we show it plays the primary role in the construction of the system many-body Hamiltonian. Furthermore, dissipation in the system is treated on equal footing by calculating the EPR pmlsubscript𝑝𝑚𝑙p_{ml} of lossy element l𝑙l in mode m𝑚m.

The EPR method deviates from previous black-box quantization work (Nigg2012, ; Solgun2014, ; Solgun2015, ), which uses the impedance-response matrix, denoted Zjj(ω)subscript𝑍𝑗superscript𝑗𝜔Z_{jj^{\prime}}\left(\omega\right), where j𝑗j and jsuperscript𝑗j^{\prime} index ports associated with nonlinear elements. For all pairs of ports, the complex function Zjj(ω)subscript𝑍𝑗superscript𝑗𝜔Z_{jj^{\prime}}\left(\omega\right) is calculated from a finite-element (FE) driven simulation in the vicinity of the eigenfrequency of every mode. Our method replaces these steps with a more economical FE eigenmode simulation, from which one extracts the energy participations pmlsubscript𝑝𝑚𝑙p_{ml} and pmjsubscript𝑝𝑚𝑗p_{mj}, needed to fully characterize both the dissipative and Hamiltonian properties of the circuit.

To test the method, we compared EPR calculations of circuit parameters to experimentally measured ones for 8 superconducting devices designed with the EPR method, comprising a total of 15 qubits, 8 readout and storage resonator modes, and one waveguide system. The results demonstrate agreement for Hamiltonian parameters spanning over five orders of magnitude in energy. Resonance frequencies were calculated to one percent accuracy, large nonlinear interactions, such as anharmonicities and cross-Kerr frequencies, to five percent, and small, nonlinear interactions to ten percent. This level of accuracy is sufficient for most current quantum information experiments.

Refer to caption
Figure 1: Conceptual overview. a Illustration of the physical model of an example quantum device, which comprises two three-dimensional (3D) cavities (grey enclosures), each housing several qubit chips (green boxes). A close-up view of one of the chips is depicted in the inset. The dotted box in the center of the chip schematically outlines a non-linear inductive sub-circuit, referred to as a Josephson dipole. b Results of a finite-element eigenmode analysis (FEesubscriptFEe\mathrm{FE}_{\mathrm{e}}) of the Josephson circuit linearized about its equilibrium. The m𝑚m-th mode eigenfrequency and electric and magnetic fields are ωmsubscript𝜔𝑚\omega_{m}, Em(r)subscript𝐸𝑚𝑟\vec{E}_{m}\left(\vec{r}\right), and Hm(r)subscript𝐻𝑚𝑟\vec{H}_{m}\left(\vec{r}\right), respectively, where r𝑟r denotes spatial position. Center inset: |Em|subscript𝐸𝑚\left|\vec{E}_{m}\right| profile (red: high; blue: low) for the fundamental mode of one of the 3D cavities. Additional FE driven simulations (FEdsubscriptFEd\mathrm{FE}_{\mathrm{d}}) are unnecessary; i.e., the impedance matrix Zjj(ω)subscript𝑍𝑗superscript𝑗𝜔Z_{jj^{\prime}}\left(\omega\right) is not calculated. c The Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}}, which includes nonlinear interactions to arbitrary order (see Results ), is computed directly from the eigenanalysis via the EPRs pmjsubscript𝑝𝑚𝑗p_{mj} and EPR signs smj=±1subscript𝑠𝑚𝑗plus-or-minus1s_{mj}=\pm 1 of the junctions, j𝑗j. Dissipative contributions due to a lossy element l𝑙l are similarly computed from the loss EPRs pmlsubscript𝑝𝑚𝑙p_{ml}; for linear dissipation, the EPR signs smlsubscript𝑠𝑚𝑙s_{ml} are unnecessary. Direct extraction of Hamiltonian and dissipative parameters from eigensolutions is unique to the EPR method. The geometry of the classical model is modified in an iterative search for the desired dissipative and Hamiltonian parameters (left-pointing arrow).

II RESULTS AND DISCUSSION

Refer to caption
Figure 2: Quantizing a simple circuit. a Illustration of a 3D cavity enclosing a transmon qubit chip. The cross symbol marks the location of a Josephson junction. Vertical blue arrows depict the electric field Em(r)subscript𝐸𝑚𝑟\vec{E}_{m}\left(\vec{r}\right) of the fundamental cavity mode, TE101. b Equivalent two-mode lumped-element- representation of the distributed circuit. Operators a^qsubscript^𝑎𝑞\hat{a}_{q} and a^csubscript^𝑎𝑐\hat{a}_{c} denote the qubit and cavity mode operators, respectively.

II.1 To quantize a simple circuit: qubit coupled to a cavity

In this section, we introduce the EPR method of quantum circuit design on a modest, yet informative, example: a transmon qubit coupled to a cavity mode (see Fig. 2). The transmon (Koch2007, ) consists of a Josephson junction shunted by a capacitance. It is embedded in the cavity, which we will consider as a black-box distributed structure. The Hamiltonian of this system H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} can be conceptually separated in two contributions (see Supplementary Section A2),

H^full=H^lin+H^nl,subscript^𝐻fullsubscript^𝐻linsubscript^𝐻nl\hat{H}_{\mathrm{full}}=\hat{H}_{\mathrm{lin}}+\hat{H}_{\mathrm{nl}}\;, (1)

where H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}} consists of all terms associated with the linear response of the junction and the resonator structure, and H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} consists of terms associated with the nonlinear response of the junction. Restricting our attention to the cavity and qubit modes of the otherwise black-box structure, the analytical form of the Hamiltonian follows from standard circuit quantization (Yurke1984, ; Devoret1995, ) (see Supplementary Section A4):

H^linsubscript^𝐻lin\displaystyle\hat{H}_{\mathrm{lin}} =\displaystyle= ωca^ca^c+ωqa^qa^q,Planck-constant-over-2-pisubscript𝜔𝑐superscriptsubscript^𝑎𝑐subscript^𝑎𝑐Planck-constant-over-2-pisubscript𝜔𝑞superscriptsubscript^𝑎𝑞subscript^𝑎𝑞\displaystyle\hbar\omega_{c}\hat{a}_{c}^{\dagger}\hat{a}_{c}+\hbar\omega_{q}\hat{a}_{q}^{\dagger}\hat{a}_{q}\;, (2)
H^nlsubscript^𝐻nl\displaystyle\hat{H}_{\mathrm{nl}} =\displaystyle= EJ[cos(φ^J)+φ^J2/2],subscript𝐸𝐽delimited-[]subscript^𝜑𝐽superscriptsubscript^𝜑𝐽22\displaystyle-E_{J}\left[\cos\left(\hat{\varphi}_{J}\right)+\hat{\varphi}_{J}^{2}/2\right]\;, (3)
φ^Jsubscript^𝜑𝐽\displaystyle\hat{\varphi}_{J} =\displaystyle= φq(a^q+a^q)+φc(a^c+a^c),subscript𝜑𝑞subscript^𝑎𝑞superscriptsubscript^𝑎𝑞subscript𝜑𝑐subscript^𝑎𝑐superscriptsubscript^𝑎𝑐\displaystyle\varphi_{q}\left(\hat{a}_{q}+\hat{a}_{q}^{\dagger}\right)+\varphi_{c}\left(\hat{a}_{c}+\hat{a}_{c}^{\dagger}\right)\;, (4)

where ωcsubscript𝜔𝑐\omega_{c} and ωqsubscript𝜔𝑞\omega_{q} are the angular frequencies of the cavity and qubit eigenmodes defined associated with H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}, respectively, and where a^csubscript^𝑎𝑐\hat{a}_{c} and a^qsubscript^𝑎𝑞\hat{a}_{q} are their annihilation operators, respectively. The Josephson energy EJsubscript𝐸𝐽E_{J} can be computed from the Ambegaokar-Baratoff formula adapted to the measured room-temperature resistance of the junction (Gloos2000, ). The junction reduced generalized flux φ^Jsubscript^𝜑𝐽\hat{\varphi}_{J} corresponds to the classical variable φJ(t)tvJ(τ)dτ/ϕ0subscript𝜑𝐽𝑡superscriptsubscript𝑡subscript𝑣𝐽𝜏differential-d𝜏subscriptitalic-ϕ0\varphi_{J}\left(t\right)\coloneqq\int_{-\infty}^{t}v_{J}\left(\tau\right)\,\mathrm{d}\tau/\phi_{0}, where vJ(τ)subscript𝑣𝐽𝜏v_{J}(\tau) is the instantaneous voltage across the junction (Yurke1984, ; Devoret1995, ), and ϕ0/2esubscriptitalic-ϕ0Planck-constant-over-2-pi2𝑒\phi_{0}\coloneqq\hbar/2e is the reduced flux quantum. The junction flux operator [Eq. (4)] is a linear, real-valued, and non-negative combination of the mode operators (see Supplementary Section A4), and in its expression, φc\varphi{}_{c} and φqsubscript𝜑𝑞\varphi_{q} are the quantum zero-point fluctuations of junction flux in the cavity and qubit mode, respectively. It is worth stating that the linear coupling between the cavity and qubit, commonly denoted g𝑔g (Koch2007, ), is fully factored in our analysis, and is implicitly handled in the extraction of the operators from the electromagnetic simulation.

Our principal aim is to determine the unknown quantities: ωc,ωq,φqsubscript𝜔𝑐subscript𝜔𝑞subscript𝜑𝑞\omega_{c},\omega_{q},\varphi_{q} and φcsubscript𝜑𝑐\varphi_{c}. As we will show, we extract and compute these quantities from an eigenanalysis of the classical distributed circuit corresponding to H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}. This includes the qubit-cavity layout, materials, electromagnetic boundary conditions, and a model of the junction as a lumped-element, linear inductor. The eigensolver returns the requested set of eigenmodes and their frequencies, quality factors, and field solutions. By running the eigensolver in the frequency range of interest, we obtain the hybridized cavity and qubit modes, whose eigenfrequencies ωcsubscript𝜔𝑐\omega_{c} and ωqsubscript𝜔𝑞\omega_{q} fully determine H^linsubscript^𝐻lin\hat{H}_{\text{lin}} (see Supplementary Section C for FE methodology).

To determine H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, we need the quantum zero-point fluctuations φqsubscript𝜑𝑞\varphi_{q} and φcsubscript𝜑𝑐\varphi_{c}, which are calculated from the participation of the junction in the eigenfield solutions. The participation pmsubscript𝑝𝑚p_{m} of the junction in mode m{c,q}𝑚𝑐𝑞m\in\left\{c,q\right\} is defined to be the fraction of inductive energy stored in the junction relative to the total inductive energy stored in the entire circuit,

pmInductive energy stored in the junctionTotal inductive energy stored in mode m,subscript𝑝𝑚Inductive energy stored in the junctionTotal inductive energy stored in mode 𝑚p_{m}\coloneqq\frac{\text{Inductive energy stored in the junction}}{\text{Total inductive energy stored in mode }m}\;, (5)

evaluated when only mode m𝑚m is excited. Thus, pmsubscript𝑝𝑚p_{m} can be computed from the electric Em(r)subscript𝐸𝑚𝑟\vec{E}_{m}(\vec{r}) and magnetic Hm(r)subscript𝐻𝑚𝑟\vec{H}_{m}(\vec{r}) eigenfields as detailed in Supplementary Section C2; r𝑟\vec{r} denotes spatial position. In the quantum setting, Eq. (5) links pmsubscript𝑝𝑚p_{m}, φ^Jsubscript^𝜑𝐽\hat{\varphi}_{J}, and the state of the circuit,

pm=ψm|12EJφ^J2|ψmψm|12H^lin|ψm,subscript𝑝𝑚quantum-operator-productsubscript𝜓𝑚12subscript𝐸𝐽superscriptsubscript^𝜑𝐽2subscript𝜓𝑚quantum-operator-productsubscript𝜓𝑚12subscript^𝐻linsubscript𝜓𝑚p_{m}=\frac{\langle\psi_{m}|\frac{1}{2}E_{J}\hat{\varphi}_{J}^{2}|\psi_{m}\rangle}{\langle\psi_{m}|\frac{1}{2}\hat{H}_{\mathrm{lin}}|\psi_{m}\rangle}\;, (6)

where |ψmketsubscript𝜓𝑚\left|\psi_{m}\right\rangle denotes a coherent state or a Fock excitation of mode m𝑚m. Note that normal-ordering must be used in Eq. (6); this correct treatment of vacuum fluctuations is detailed in Supplementary Section A6. Simplifying Eq. (6), one expresses the variance of the quantum zero-point fluctuations φcsubscript𝜑𝑐\varphi_{c} and φqsubscript𝜑𝑞\varphi_{q} as a function of the classical energy participations pmsubscript𝑝𝑚p_{m},

φc=2pcωc2EJandφq=2pqωq2EJ,\varphi_{c}{}^{2}=p_{c}\frac{\hbar\omega_{c}}{2E_{J}}\quad\text{and}\quad\varphi_{q}{}^{2}=p_{q}\frac{\hbar\omega_{q}}{2E_{J}}\;, (7)

which completely determines H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, and thus completes the description of the system Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}}. Here, φcsubscript𝜑𝑐\varphi_{c} and φqsubscript𝜑𝑞\varphi_{q} can be taken as positive numbers. As we will see in the next section, in the presence of multiple junctions, this is not always true.

Designing experiments with the EPR requires one to further extract from H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} the transition frequencies and nonlinear couplings between modes. Depending on the case, this can be done approximately or exactly using numerical or analytical techniques (DiPaolo2019, ). This task is easily achieved if H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} is a perturbation to H^linsubscript^𝐻lin\hat{H}_{\text{lin}} (Koch2007, ). In this limit, H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} for our qubit-cavity example can be approximated by the effective, excitation-number-conserving Hamiltonian, see Supplementary Eq. (B.2),

H^eff=(ωqΔq)n^q+(ωcΔc)n^cχqcn^qn^c12αqn^q(n^q1^)12αcn^c(n^c1^),subscript^𝐻effsubscript𝜔𝑞subscriptΔ𝑞subscript^𝑛𝑞subscript𝜔𝑐subscriptΔ𝑐subscript^𝑛𝑐subscript𝜒𝑞𝑐subscript^𝑛𝑞subscript^𝑛𝑐12subscript𝛼𝑞subscript^𝑛𝑞subscript^𝑛𝑞^112subscript𝛼𝑐subscript^𝑛𝑐subscript^𝑛𝑐^1\hat{H}_{\text{eff}}=\left(\omega_{q}-\Delta_{q}\right)\hat{n}_{q}+\left(\omega_{c}-\Delta_{c}\right)\hat{n}_{c}-\chi_{qc}\hat{n}_{q}\hat{n}_{c}\\ -\frac{1}{2}\alpha_{q}\hat{n}_{q}\left(\hat{n}_{q}-\hat{1}\right)-\frac{1}{2}\alpha_{c}\hat{n}_{c}\left(\hat{n}_{c}-\hat{1}\right)\;, (8)

where n^q=a^qa^qsubscript^𝑛𝑞superscriptsubscript^𝑎𝑞subscript^𝑎𝑞\hat{n}_{q}=\hat{a}_{q}^{\dagger}\hat{a}_{q} and n^c=a^ca^csubscript^𝑛𝑐superscriptsubscript^𝑎𝑐subscript^𝑎𝑐\hat{n}_{c}=\hat{a}_{c}^{\dagger}\hat{a}_{c} denote the qubit and cavity excitation-number operators, respectively, ΔqsubscriptΔ𝑞\Delta_{q} denotes the ‘Lamb shift’ of the qubit frequency due to the dressing of this nonlinear mode by quantum fluctuations of the fields, αqsubscript𝛼𝑞\alpha_{q} (αcsubscript𝛼𝑐\alpha_{c}) is the qubit (cavity) anharmonicity, and χqcsubscript𝜒𝑞𝑐\chi_{qc} is the qubit-cavity dispersive shift (cross-Kerr coupling). The Hamiltonian parameters can be calculated directly from the EPR, see Supplementary Section B,

αqsubscript𝛼𝑞\displaystyle\alpha_{q} =12χqq=pq2ωq28EJ,absent12subscript𝜒𝑞𝑞superscriptsubscript𝑝𝑞2Planck-constant-over-2-pisuperscriptsubscript𝜔𝑞28subscript𝐸𝐽\displaystyle=\frac{1}{2}\chi_{qq}=p_{q}^{2}\frac{\hbar\omega_{q}^{2}}{8E_{J}}\;, (9)
αcsubscript𝛼𝑐\displaystyle\alpha_{c} =12χcc=pc2ωc28EJ,absent12subscript𝜒𝑐𝑐superscriptsubscript𝑝𝑐2Planck-constant-over-2-pisuperscriptsubscript𝜔𝑐28subscript𝐸𝐽\displaystyle=\frac{1}{2}\chi_{cc}=p_{c}^{2}\frac{\hbar\omega_{c}^{2}}{8E_{J}}\;, (10)
χqcsubscript𝜒𝑞𝑐\displaystyle\chi_{qc} =pqpcωqωc4EJ.absentsubscript𝑝𝑞subscript𝑝𝑐Planck-constant-over-2-pisubscript𝜔𝑞subscript𝜔𝑐4subscript𝐸𝐽\displaystyle=p_{q}p_{c}\frac{\hbar\omega_{q}\omega_{c}}{4E_{J}}\;. (11)

Experimentally, the qubit Lamb shift can be obtained as Δq=αqχqc/2subscriptΔ𝑞subscript𝛼𝑞subscript𝜒𝑞𝑐2\Delta_{q}=\alpha_{q}-\chi_{qc}/2. Since a single EPR pmsubscript𝑝𝑚p_{m} determines the nonlinear interaction for each mode, the parameters χqcsubscript𝜒𝑞𝑐\chi_{qc} and αqsubscript𝛼𝑞\alpha_{q} are interdependent,

χqc=χqqχcc=2αqαc.subscript𝜒𝑞𝑐subscript𝜒𝑞𝑞subscript𝜒𝑐𝑐2subscript𝛼𝑞subscript𝛼𝑐\chi_{qc}=\sqrt{\chi_{qq}\chi_{cc}}=2\sqrt{\alpha_{q}\alpha_{c}}\;. (12)

As shown in Supplementary Section A7, the EPRs pcsubscript𝑝𝑐p_{c} and pqsubscript𝑝𝑞p_{q} obey the constraints

0pq,pc1andpq+pc=1.formulae-sequence0subscript𝑝𝑞formulae-sequencesubscript𝑝𝑐1andsubscript𝑝𝑞subscript𝑝𝑐10\leq p_{q},p_{c}\leq 1\quad\text{and}\quad p_{q}+p_{c}=1\;. (13)

These relations together with Eqs. (9) and (12) are useful to budget the dilution of the nonlinearity of the junction (see Supplementary Section B2) and to provide insight on the limits of accessible parameters (see Methods). Further, Eq. (13) is used to validate the convergence of the FE simulation.

Refer to caption
Figure 3: Schematic representation of the Josephson circuit and its nonlinear elements. a A simple example of a Josephson dipole—a Josephson tunnel junction. b An example of a composite junction, comprising four Josephson junctions in a ring, frustrated by an external magnetic flux ΦjextsuperscriptsubscriptΦ𝑗ext\Phi_{j}^{\mathrm{ext}} threading the loop. c Conceptual decomposition of a general Josephson dipole, denoted j𝑗j. For convenience, its potential energy function j(Φj;Φjext)subscript𝑗subscriptΦ𝑗superscriptsubscriptΦ𝑗ext\mathcal{E}_{j}(\Phi_{j};\Phi_{j}^{\mathrm{ext}}) can be Taylor expanded in a sum of nonlinear inductive contributions of increasing order ΦjpsuperscriptsubscriptΦ𝑗𝑝\Phi_{j}^{p}, with relative amplitude cjpsubscript𝑐𝑗𝑝c_{jp}, where p𝑝p denotes the index in the series. The energy function can be subjected to external bias parameters ΦjextsuperscriptsubscriptΦ𝑗ext\Phi_{j}^{\mathrm{ext}}, such as flux or voltages. d Schematic diagram of a general Josephson circuit conceptually resolved into a purely dissipative (left, κmlsuperscriptsubscript𝜅𝑚𝑙\kappa_{m}^{l}), linear (middle, H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}), and nonlinear (right, H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}) constitutions.

II.2 Quantizing the general Josephson system

The simple results obtained in the preceding section will now be generalized to arbitrary nonlinear devices enclosed in a black-box, distributed, electromagnetic structure. While such structures are frequently classified as planar (Blais2004, ; Wallraff2004, ; Barends2013, ; FYan2016, ) (2D), quasi-planar (Minev2013, ; Minev2016, ; Brecht2016, ; Rosenberg2017, ) (2.5D), or three-dimensional (Paik2011, ; Rigetti2012, ; Reagor2016-cavity, ; Axline2016, ) (3D), we will treat all classes on equal footing. The electromagnetic structure is assumed to be linear in the absence of the enclosed nonlinear devices. For simplicity of discussion, we can consider these devices to be inductive and lumped; distributed nonlinear devices, such as kinetic-inductance transmission lines (Yurke2006, ; HoEom2012, ; Vissers2015, ; Mortensen2016, ), can be thought of as a series of lumped ones.

The simplest nonlinear device comprises a single element, such as a Josephson tunnel junction [see Fig. 3(a)], an atomic-point contact (Koops1996, ; Bretheau2013, ), a nanobridge (Vijay2010, ; Peltonen2016, ), a semiconducting nanowire (Mooij2006, ; Ku2010, ; Abay2014, ; Larsen2015, ; DeLange2015-nanowire, ; Casparis2016, ), or another hybrid structure (Shim2014, ). A multi-element device, such as a SQUID (Zimmerman1966, ; Clarke2004, ), a SNAIL (Frattini2017, ) [see Fig. 3(b)], a superinductance (Manucharyan2012, ; Pop2014, ; Muppalla2017, ), or a junction array (Corlevi2006, ; Hutter2011, ; Bourassa2012, ; Masluk2012, ; Bell2012, ; Weibl2015, ; Macklin2015, ) refers to a subcircuit composed of purely inductive lumped elements. This subcircuit can also be subjected to external controls, such as voltage or flux biases.

The general nonlinear device that we now consider, referred to as a Josephson dipole, is any lumped, purely-inductive, nonlinear subcircuit with two terminals. The key characteristic of the Josephson dipole is that it possesses a characteristic energy function, which encapsulates all details of its constitution. For example, the two-terminal nonlinear device known as the symmetric SQUID (Clarke2004, ) is described by the energy function j(Φj;Φjext)=Ej(Φjext)cos(Φj/ϕ0)subscript𝑗subscriptΦ𝑗superscriptsubscriptΦ𝑗extsubscript𝐸𝑗superscriptsubscriptΦ𝑗extsubscriptΦ𝑗subscriptitalic-ϕ0\mathcal{E}_{j}\left(\Phi_{j};\Phi_{j}^{\mathrm{ext}}\right)=-E_{j}\left(\Phi_{j}^{\mathrm{ext}}\right)\cos\left(\Phi_{j}/\phi_{0}\right), where ΦjsubscriptΦ𝑗\Phi_{j} is the generalized flux across the device terminals (Yurke1984, ; Devoret1995, )Ejsubscript𝐸𝑗E_{j} is the effective Josephson energy, ΦjextsuperscriptsubscriptΦ𝑗ext\Phi_{j}^{\mathrm{ext}} is the external flux bias, and the subscript j𝑗j denotes the j𝑗j-th Josephson dipole in the circuit. The flux ΦjsubscriptΦ𝑗\Phi_{j} is defined as the deviation away from the value in equilibrium, as discussed below. To ease the notation, parameters such as ΦjextsuperscriptsubscriptΦ𝑗ext\Phi_{j}^{\mathrm{ext}} will be implicit hereafter. Similarly to the example of the single-junction transmon , the energy of a Josephson dipole can be separated in two parts. One part jlinsuperscriptsubscript𝑗lin\mathcal{E}_{j}^{\mathrm{lin}} accounts for the linear response of the dipole, while the other jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} accounts for the nonlinear response,

j(Φj)=jlin(Φj)+jnl(Φj),subscript𝑗subscriptΦ𝑗superscriptsubscript𝑗linsubscriptΦ𝑗superscriptsubscript𝑗nlsubscriptΦ𝑗\mathcal{E}_{j}\left(\Phi_{j}\right)=\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right)+\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right)\;, (14)

where

jlin(Φj)superscriptsubscript𝑗linsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) \displaystyle\coloneqq 12Ej(Φjϕ0)2,12subscript𝐸𝑗superscriptsubscriptΦ𝑗subscriptitalic-ϕ02\displaystyle\frac{1}{2}E_{j}\left(\frac{\Phi_{j}}{\phi_{0}}\right)^{2}\;, (15)

and where the constant Ejsubscript𝐸𝑗E_{j} sets the scale of the junction energy. This energy scale can be represented by the linear inductance Ljϕ0/Ejsubscript𝐿𝑗subscriptitalic-ϕ0subscript𝐸𝑗L_{j}\coloneqq\phi_{0}/E_{j} presented by the Josephson dipole when submitted to a small excitation about its equilibrium.

Frustrated equilibrium.

External biases can set up persistent currents in the circuit. These can alter the static [direct-current (dc)] equilibrium of the Josephson system. For example, frustrating a superconducting ring with a magnetic flux sets up a persistent circulating current in the ring. For a Josephson dipole in such a loop, the definition of the flux ΦjsubscriptΦ𝑗\Phi_{j} will differ in Eqs. (14) and (15) as a function of the equilibrium. Due to this adjustment, terms linear in ΦjsubscriptΦ𝑗\Phi_{j} are absent from Eq. (15) by construction. Supplementary Sections A8 and A9 discuss these equilibrium considerations in detail.

II.2.1 Quantum Hamiltonian.

Having conceptually carved out nonlinear contributions from the system Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}}, and collected them in the set of jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} functions, we define the linearized Josephson circuit to correspond to everything left over in the system. This linear circuit consists of the electromagnetic circuit external to the Josephson dipoles, combined with their linear inductances Ljsubscript𝐿𝑗L_{j}. We will use the eigenmodes of the linearized circuit to explicitly construct H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}}. The eigenmode frequencies and field distributions are readily obtained using a conventional finite-element solver (see Supplementary Section C). The Hamiltonian of the linearized Josephson circuit can thus be expressed as (see Supplementary Section A5)

H^linsubscript^𝐻lin\displaystyle\hat{H}_{\mathrm{lin}} =m=1Mωma^ma^m,absentsuperscriptsubscript𝑚1𝑀Planck-constant-over-2-pisubscript𝜔𝑚superscriptsubscript^𝑎𝑚subscript^𝑎𝑚\displaystyle=\sum_{m=1}^{M}\hbar\omega_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m}\;, (16)

where M𝑀M is the number of modes addressed by the numerical simulation, ωmsubscript𝜔𝑚\omega_{m} is the solution eigenfrequency of mode m𝑚m, and a^msubscript^𝑎𝑚\hat{a}_{m} the corresponding mode amplitude (annihilation operator), defined by the mode eigenvector. We emphasize that the frequencies ωmsubscript𝜔𝑚\omega_{m} will be significantly perturbed by the Lamb shifts ΔmsubscriptΔ𝑚\Delta_{m}, and should be seen as an intermediate parameter entering in the calculation of the rest of the nonlinear Hamiltonian,

H^nlsubscript^𝐻nl\displaystyle\hat{H}_{\mathrm{nl}} =j=1Jjnl=j=1JEj(cj3φ^j3+cj4φ^j4+)absentsuperscriptsubscript𝑗1𝐽superscriptsubscript𝑗nlsuperscriptsubscript𝑗1𝐽subscript𝐸𝑗subscript𝑐𝑗3superscriptsubscript^𝜑𝑗3subscript𝑐𝑗4superscriptsubscript^𝜑𝑗4\displaystyle=\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{nl}}=\sum_{j=1}^{J}E_{j}\left(c_{j3}\hat{\varphi}_{j}^{3}+c_{j4}\hat{\varphi}_{j}^{4}+\cdots\right) (17)
=j=1JEjp=3cjpφ^jp,absentsuperscriptsubscript𝑗1𝐽subscript𝐸𝑗superscriptsubscript𝑝3subscript𝑐𝑗𝑝superscriptsubscript^𝜑𝑗𝑝\displaystyle=\sum_{j=1}^{J}E_{j}\sum_{p=3}^{\infty}c_{jp}\hat{\varphi}_{j}^{p}\;, (18)
φ^jsubscript^𝜑𝑗\displaystyle\hat{\varphi}_{j} =m=1Mφmj(a^m+a^m),absentsuperscriptsubscript𝑚1𝑀subscript𝜑𝑚𝑗superscriptsubscript^𝑎𝑚subscript^𝑎𝑚\displaystyle=\sum_{m=1}^{M}\varphi_{mj}\left(\hat{a}_{m}^{\dagger}+\hat{a}_{m}\right)\;, (19)

where J𝐽J is the total number of junctions and φ^jΦ^j/ϕ0subscript^𝜑𝑗subscript^Φ𝑗subscriptitalic-ϕ0\hat{\varphi}_{j}\coloneqq\hat{\Phi}_{j}/\phi_{0}. In Eq. 17, we have introduced a Taylor expansion of jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}}, where the energy Ejsubscript𝐸𝑗E_{j} and expansion coefficients cjpsubscript𝑐𝑗𝑝c_{jp} are known from the fabrication of the Josephson circuit, see Fig. 3(c). For example, for a Josephson junction, the constant Ejsubscript𝐸𝑗E_{j} is just the Josephson energy, while cjpsubscript𝑐𝑗𝑝c_{jp} are the coefficients of the cosine expansion; i.e., cjpsubscript𝑐𝑗𝑝c_{jp} is 00 for odd p𝑝p and (1)p/2+1/p!superscript1𝑝21𝑝\left(-1\right)^{p/2+1}/p! for even p𝑝p. The expansion is helpful for analytics but does not need to be used in the numerical analysis of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, see Supplementary Section A.

The Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} is specified since the operators φ^jsubscript^𝜑𝑗\hat{\varphi}_{j} are expressed in terms of the mode amplitudes as a linear combination (see Supplementary Section A5). Here, φmjsubscript𝜑𝑚𝑗\varphi_{mj} are the dimensionless, real-valued, quantum zero-point fluctuations of the reduced flux of junction j𝑗j in mode m𝑚m. Determination of H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} is now reduced to computing φmjsubscript𝜑𝑚𝑗\varphi_{mj}. We achieve this by employing a generalization of the energy-participation ratio.

The energy-participation ratio pmjsubscript𝑝𝑚𝑗p_{mj} of junction j𝑗j in eigenmode m𝑚m is defined to be the fraction of inductive energy stored in the junction when only that mode is excited,

pmjsubscript𝑝𝑚𝑗\displaystyle p_{mj} Inductive energy stored in junction jInductive energy stored in mode mabsentInductive energy stored in junction 𝑗Inductive energy stored in mode 𝑚\displaystyle\coloneqq\frac{\text{Inductive energy stored in junction }j}{\text{Inductive energy stored in mode }m}
=ψm|12Ejφ^j2|ψmψm|12H^lin|ψm,absentquantum-operator-productsubscript𝜓𝑚12subscript𝐸𝑗superscriptsubscript^𝜑𝑗2subscript𝜓𝑚quantum-operator-productsubscript𝜓𝑚12subscript^𝐻linsubscript𝜓𝑚\displaystyle=\frac{\langle\psi_{m}|\frac{1}{2}E_{j}\hat{\varphi}_{j}^{2}|\psi_{m}\rangle}{\langle\psi_{m}|\frac{1}{2}\hat{H}_{\mathrm{lin}}|\psi_{m}\rangle}\;, (20)

which is a straightforward extension of Eq. (5), and is similarly computed using normal ordering (see Supplementary Section A6). The EPR pmjsubscript𝑝𝑚𝑗p_{mj} is computed from the eigenfield solutions Em(r)subscript𝐸𝑚𝑟\vec{E}_{m}(\vec{r}) and Hm(r)subscript𝐻𝑚𝑟\vec{H}_{m}(\vec{r}) as explained in Supplementary Section C3. It is a bounded, non-negative, real number, 0pmj10subscript𝑝𝑚𝑗10\leq p_{mj}\leq 1. A zero EPR pmj=0subscript𝑝𝑚𝑗0p_{mj}=0 means that junction j𝑗j is not excited in mode m𝑚m. A unity EPR pmj=1subscript𝑝𝑚𝑗1p_{mj}=1 means that junction j𝑗j is the only inductive element excited in the mode.

From the EPR pmjsubscript𝑝𝑚𝑗p_{mj}, one directly computes the variance of the quantum zero-point fluctuations (see Appendix A6),

φmj2=pmjωm2Ej.\boxed{\varphi_{mj}^{2}=p_{mj}\frac{\hbar\omega_{m}}{2E_{j}}\;.} (21)

Equation (21) constitutes the bridge between the classical solution of the linearized Josephson circuit and the quantum Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} of the full Josephson system, up to the sign of φmjsubscript𝜑𝑚𝑗\varphi_{mj}.

II.2.2 Universal EPR properties

The quantum fluctuations φmjsubscript𝜑𝑚𝑗\varphi_{mj} are not independent of each other, since the EPRs are submitted to three types of universal constraints—valid regardless of the circuit topology and nature of the Josephson dipoles. These are of practical importance, as they are useful guides in evaluating the performance of possible designs and assessing their limitations. As shown in Supplementary Section A7, the EPRs obey one sum rule per junction j𝑗j and one set of inequalities per mode m𝑚m,

m=1Mpmj=1and0j=1Jpmj1.formulae-sequencesuperscriptsubscript𝑚1𝑀subscript𝑝𝑚𝑗1and0superscriptsubscript𝑗1𝐽subscript𝑝𝑚𝑗1\sum_{m=1}^{M}p_{mj}=1\quad\text{and}\quad 0\leq\sum_{j=1}^{J}p_{mj}\leq 1\;. (22)

The total EPR of a Josephson dipole is a quantity that is independent of the number of modes—it is precisely unity for all circuits in which the dipole is embedded. It can only be diluted among the modes. On the other hand, a given mode can accept at most a total EPR of unity from all the dipoles. In practice, this sum rule can be fully exploited only if the bound M𝑀M reaches the total number of relevant modes of the system.

The next fundamental property concerns the orthogonality of the EPRs. Rewriting Eq. (21) in terms of the amplitude of the zero-point fluctuation we have

φmj=smjpmjωm/2Ej,subscript𝜑𝑚𝑗subscript𝑠𝑚𝑗subscript𝑝𝑚𝑗Planck-constant-over-2-pisubscript𝜔𝑚2subscript𝐸𝑗\varphi_{mj}=s_{mj}\sqrt{p_{mj}\hbar\omega_{m}/2E_{j}}\;, (23)

where the EPR sign smjsubscript𝑠𝑚𝑗s_{mj} of junction j𝑗j in mode m𝑚m is either +11+1 or 11-1. The EPR sign encodes the relative direction of current flowing across the junction. Only the relative value between smjsubscript𝑠𝑚𝑗s_{mj} and smjsubscript𝑠𝑚superscript𝑗s_{mj^{\prime}} for jj𝑗superscript𝑗j\neq j^{\prime} has physical significance (see Supplementary Figure S8). The EPR sign smjsubscript𝑠𝑚𝑗s_{mj} is calculated in parallel with the process of calculating pmjsubscript𝑝𝑚𝑗p_{mj}, from the field solution H(r)𝐻𝑟\vec{H}(\vec{r}), see Supplementary Section (C.8). We now obtain the EPR orthogonality relationship

m=1Msmjsmjpmjpmjsuperscriptsubscript𝑚1𝑀subscript𝑠𝑚𝑗subscript𝑠𝑚superscript𝑗subscript𝑝𝑚𝑗subscript𝑝𝑚superscript𝑗\displaystyle\sum_{m=1}^{M}s_{mj}s_{mj^{\prime}}\sqrt{p_{mj}p_{mj^{\prime}}}\, =0,absent0\displaystyle=0\;, (24)

valid when the sum from 111 to M𝑀M covers all the relevant modes, see Supplementary Eq. (A.60).

II.2.3 Excitation-number-conserving interactions

Thus, as announced, knowledge of the energy-participation ratios completely specifies H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, through Eqs. (17), (19), and (21). The Hamiltonian can now be analytically or numerically diagonalized using various computational techniques (DiPaolo2019, ). In this section, our focus will be now to explicitly handle the effect of the nonlinear interactions H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} on the eigenmodes. Before treating the case of a general nonlinear interaction, we focus on the leading-order effect of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} in the case of the ‘canonical’ Josephson system. In this case, the J𝐽J Josephson dipoles are all Josephson tunnel junctions, characterized by Eq. (A.14b), and the dispersive regime is satisfied for all pairs of modes k𝑘k and m𝑚m; i.e., ωkωmEjcjpφ^jpmuch-greater-thansubscript𝜔𝑘subscript𝜔𝑚subscript𝐸𝑗subscript𝑐𝑗𝑝delimited-⟨⟩superscriptsubscript^𝜑𝑗𝑝\omega_{k}-\omega_{m}\gg E_{j}c_{jp}\left<\hat{\varphi}_{j}^{p}\right> for p3𝑝3p\geq 3 and in the absence of strong drives. The leading-order nonlinear terms are the subset of p=4𝑝4p=4 terms that conserve excitation number. After normal ordering, see Supplementary Section B1, one finds the effective Hamiltonian

H¯^4=m=1MΔma^ma^m+αm2a^m2a^m2+n<mχmna^ma^na^ma^n,subscript^¯𝐻4Planck-constant-over-2-pisuperscriptsubscript𝑚1𝑀subscriptΔ𝑚superscriptsubscript^𝑎𝑚subscript^𝑎𝑚subscript𝛼𝑚2superscriptsubscript^𝑎𝑚absent2superscriptsubscript^𝑎𝑚2subscript𝑛𝑚subscript𝜒𝑚𝑛superscriptsubscript^𝑎𝑚superscriptsubscript^𝑎𝑛subscript^𝑎𝑚subscript^𝑎𝑛\hat{\overline{H}}_{4}=-\hbar\sum_{m=1}^{M}\Delta_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m}+\frac{\alpha_{m}}{2}\hat{a}_{m}^{\dagger 2}\hat{a}_{m}^{2}+\sum_{n<m}\chi_{mn}\hat{a}_{m}^{\dagger}\hat{a}_{n}^{\dagger}\hat{a}_{m}\hat{a}_{n}\;, (25)

which is a generalization of the one found in Eq. (8). In Eq. (25), we have introduced the Lamb shift ΔmsubscriptΔ𝑚\Delta_{m} of mode m𝑚m, the anharmonicity αmsubscript𝛼𝑚\alpha_{m} of the mode, and its total dispersive shift χmnsubscript𝜒𝑚𝑛\chi_{mn} (so-called cross-Kerr term) with a different mode, labeled n𝑛n. Each of these parameters is directly calculated from the EPRs. As shown in Supplementary Section (B.3a), for arbitrary m𝑚m and n𝑛n,

χmn=j=1Jωmωn4Ejpmjpnj,subscript𝜒𝑚𝑛superscriptsubscript𝑗1𝐽Planck-constant-over-2-pisubscript𝜔𝑚subscript𝜔𝑛4subscript𝐸𝑗subscript𝑝𝑚𝑗subscript𝑝𝑛𝑗\chi_{mn}=\sum_{j=1}^{J}\frac{\hbar\omega_{m}\omega_{n}}{4E_{j}}p_{mj}p_{nj}\;, (26)

while αm=χmm/2subscript𝛼𝑚subscript𝜒𝑚𝑚2\alpha_{m}=\chi_{mm}/2 and Δm=n=1Mχmn/2subscriptΔ𝑚superscriptsubscript𝑛1𝑀subscript𝜒𝑚𝑛2\Delta_{m}=\sum_{n=1}^{M}\chi_{mn}/2. Equation (26) implements, mathematically, the idea that the amplitude of these nonlinear couplings is the result of a spatial-mode scalar product of the EPRs. Remarkably, from Eq. (26) it is seen that the EPRs are essentially the only free parameters subject to design, when determining the nonlinear couplings, since ωmsubscript𝜔𝑚\omega_{m}, ωnsubscript𝜔𝑛\omega_{n}, and Ejsubscript𝐸𝑗E_{j} are generally tightly constrained by experimental considerations.

Equation (26) can be cast in matrix form by introducing the EPR matrix

𝐏(p11p1JpM1pMJ),𝐏subscript𝑝11subscript𝑝1𝐽subscript𝑝𝑀1subscript𝑝𝑀𝐽\mathrm{\mathbf{P}}\coloneqq\left(\begin{array}[]{ccc}p_{11}&\cdots&p_{1J}\\ \vdots&\ddots&\vdots\\ p_{M1}&\cdots&p_{MJ}\end{array}\right)\;, (27)

which we have found useful in handling large circuits, especially for those in excess of 100 modes. We also introduce the diagonal matrices of eigenfrequencies 𝛀diag(ω1,,ωM)𝛀diagsubscript𝜔1subscript𝜔𝑀\mathrm{\mathbf{\Omega}}\coloneqq\operatorname{diag}\left(\omega_{1},\ldots,\omega_{M}\right) and junction energies 𝐄𝐉diag(E1,,EJ)subscript𝐄𝐉diagsubscript𝐸1subscript𝐸𝐽\mathrm{\mathbf{E_{J}}}\coloneqq\operatorname{diag}\left(E_{1},\ldots,E_{J}\right), which lead to the matrix form of Eq. (26),

Kerr matrix:𝝌=4𝛀𝐏𝐄𝐉𝟏𝐏𝛀,Anharmonicity:αm=12[𝝌]mm,Lamb shift:Δm=12m=1M[𝝌]mm.Kerr matrix:missing-subexpression𝝌Planck-constant-over-2-pi4𝛀superscriptsubscript𝐏𝐄𝐉1superscript𝐏𝛀Anharmonicity:missing-subexpressionsubscript𝛼𝑚12subscriptdelimited-[]𝝌𝑚𝑚Lamb shift:missing-subexpressionsubscriptΔ𝑚12superscriptsubscriptsuperscript𝑚1𝑀subscriptdelimited-[]𝝌𝑚superscript𝑚\begin{array}[]{rcclc}\text{Kerr matrix:}&&\bm{\chi}&=&\frac{\hbar}{4}\mathrm{\mathbf{\Omega PE_{J}^{-1}P^{\intercal}}}\bm{\Omega}\;,\\ \text{Anharmonicity:}&&\alpha_{m}&=&\frac{1}{2}\left[\bm{\chi}\right]_{mm}\;,\\ \text{Lamb shift:}&&\Delta_{m}&=&\frac{1}{2}\sum_{m^{\prime}=1}^{M}\left[\bm{\chi}\right]_{mm^{\prime}}\;.\end{array} (28)

We have defined the symmetric matrix of dispersive shifts 𝝌𝝌\bm{\chi}, with elements [𝝌]mm=χmmsubscriptdelimited-[]𝝌𝑚superscript𝑚subscript𝜒𝑚superscript𝑚\left[\bm{\chi}\right]_{mm^{\prime}}=\chi_{mm^{\prime}}. Further discussion of the matrix approach and applications to p𝑝pth-order corrections is deferred to Supplementary Section B2, and the amplitude of an arbitrary multi-photon interaction stemming from the full H^nlsubscript^𝐻nl\hat{H}_{\text{nl}} is calculated in Supplementary Section B3.

II.2.4 EPR for dissipation in the circuit

The EPR method treats the calculation of Hamiltonian and dissipation parameters on equal footing. Unlike in the impedance method (Nigg2012, ), one can completely characterize both H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} and the effect of dissipative elements in the circuit from the eigenfield solutions, Em(r)subscript𝐸𝑚𝑟\vec{E}_{m}(\vec{r}) and Hm(r)subscript𝐻𝑚𝑟\vec{H}_{m}(\vec{r}). The list of dissipative elements include bulk and surface dielectrics (Martinis2005, ; Patel2013, ; Dial2016, ), thin-film metals (Vissers2012, ; Minev2013, ), surface interfaces (Wenner2011, ; Geerlings2012, ; Sandberg2013, ; Wang2015, ; Bruno2015-loss, ), and metal seams (Brecht2015, ). The energy-participation ratio of a dissipative element l𝑙l in mode m𝑚m will be denoted pmlsubscript𝑝𝑚𝑙p_{ml}. It is calculated similarly to pmjsubscript𝑝𝑚𝑗p_{mj}, as summarized in Supplementary Section D. The participation pmlsubscript𝑝𝑚𝑙p_{ml} and the quality factor Qlsubscript𝑄𝑙Q_{l} of the material of this element are used to estimate the total quality factor of mode m𝑚m in the standard way when the fields are not greatly altered by the dissipation (Qm1much-greater-thansubscript𝑄𝑚1Q_{m}\gg 1(Gao2008, ; Geerlings2013, ; Reagor2016-cavity, ; Brecht2017-micromachined, ),

Qm1=lpmlQl1.superscriptsubscript𝑄𝑚1subscript𝑙subscript𝑝𝑚𝑙superscriptsubscript𝑄𝑙1Q_{m}^{-1}=\sum_{l}{p_{ml}Q_{l}^{-1}}\;. (29)

Experimental values of Qlsubscript𝑄𝑙Q_{l} are found in the literature, and some are provided in Supplementary Section D. Equation (29) and the dissipative EPR pmlsubscript𝑝𝑚𝑙p_{ml} provide a dissipation budget for the individual influence of each dissipation mechanism in the system, providing a useful tool to optimize design layout for quantum coherence (Martinis2014, ).

Refer to caption
Figure 4: Comparison between theory and experiment over five-orders of magnitude in energy scale of the system Hamiltonian H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}} for eight distinct, multi-mode device samples, described in detail in the Methods, including 3D, flip-chip (2.5D), and 3D waveguide architectures incorporating readout and storage resonators and qubit modes. For each device, the dominant parameters in H^fullsubscript^𝐻full\hat{H}_{\mathrm{full}}, dressed frequencies ωmsuperscriptsubscript𝜔𝑚\omega_{m}^{\prime}, bare anharmonicities αmsubscript𝛼𝑚\alpha_{m}, and cross-Kerr interactions χmnsubscript𝜒𝑚𝑛\chi_{mn}, were measured and calculated using the EPR method with our open-source pyEPR package (Note1, ). Gray line is of slope one, representing ideal agreement between theory and experiment.

II.3 Comparison between theory and experiment

Applying the EPR method, we designed 8 superconducting samples to test the agreement between the EPR theory and experimental results. We tested several sample configurations, comprising 15 qubits, 8 cavity modes, and one waveguide in three different circuit-QED architectures. The samples were measured in a standard cQED setup, see Methods, at the 15 mK stage of a dilution unit, over multiple cool downs.

Six of the samples were each composed of 2 qubits and one 3D cavity, one sample was composed of 2 qubits and a waveguide, and one sample was a flip-chip, 2.5D system (Minev2016, ) consisting of a flip-chip qubit embedded in a two-mode whispering gallery mode resonator (Minev2013, ) (WGMR). The specifics of each sample are discussed in the Methods.

For each sample, we measured the circuit parameters of interest: dressed mode frequencies ωmΔmsubscript𝜔𝑚subscriptΔ𝑚\omega_{m}-\Delta_{m}, anharmonicities of qubits and high-Q cavities αmsubscript𝛼𝑚\alpha_{m}, cross-Kerr frequencies χmnsubscript𝜒𝑚𝑛\chi_{mn}, and input-output quality factors QCsubscript𝑄𝐶Q_{C} for any readout modes. Our measurement methodology is detailed in the Methods.

The measured parameters were compared to those calculated using the energy-participation method. The linearized Josephson circuit of each sample was modeled in Ansys High-Frequency Electromagnetic-Field Simulator (HFSS). Junctions were modeled as lumped inductors, whose nominal energy Ejsubscript𝐸𝑗E_{j} was inferred from room-temperature resistance measurements (Gloos2000, ). To account for the error bars of the measurement and the drift in resistance over time, Ejsubscript𝐸𝑗E_{j} was adjusted by no more than 10% to fit the measured qubit frequency. To minimize the number of free parameters, we neglect the small junction intrinsic capacitance CJsubscript𝐶𝐽C_{J} in our modeling. The tradeoff is a small and estimable systematic offset of the bare simulated mode anharmonicities. We estimate this correction to be on the order of 4% for a CJ=4fFsubscript𝐶𝐽4fFC_{J}=4\,\mathrm{fF}. From the eigenfield solutions, we calculated the EPRs pmjsubscript𝑝𝑚𝑗p_{mj} and the sign smjsubscript𝑠𝑚𝑗s_{mj} to construct H^^𝐻\hat{H} and extract its parameters. Detailed steps of the procedure can be found in Supplementary Section C. The results are presented in Tables 13.

Figure 4 summarizes the agreement of the measured and calculated sample parameters, which span five orders of magnitude in frequency. Accounting for CJsubscript𝐶𝐽C_{J}, we find that mode frequencies are calculated to one percent accuracy, large nonlinear interaction energies (namely, anharmonicity and cross-Kerr frequencies greater than 10 MHz) are calculated at the 5% level, and small nonlinear interaction energies agree at the 10% level. We highlight that we have used minimal, coarse adjustment to account for shifts in Ejsubscript𝐸𝑗E_{j}, and otherwise, by neglecting Cjsubscript𝐶𝑗C_{j}, the calculation is free from adjustable parameters.

The results of Fig. 4 demonstrate the accuracy and applicability of the EPR method. For each device, the EPR results are obtained from a single eigenmode simulation, using full automation of the analysis, provided by our open-source package pyEPR (Note1, ). For current standard applications, we find the agreement sufficient. Further improvements in accuracy would require improved ability to estimate the Josephson dipole energy Ejsubscript𝐸𝑗E_{j} and its intrinsic capacitance CJsubscript𝐶𝐽C_{J}. At the same level of accuracy, improvements in the precision and reproducibility of the implementation and assembly of the Josephson circuit design are needed, such as in chip-clamping techniques, precision machining of the device sample holder and input-output couplers.

Conclusion.

An intuitive, easy-to-use and efficient method is needed to design and analyze Josephson microwave quantum circuits. We have described in this article such a method, based on the distribution of the electromagnetic energy in the circuit and its participation in nonlinear and dissipative elements. This so-called EPR method offers physical insight helping the design process, and provides a simple link between the classical circuit and its quantum properties. By comparing our theory to 8 experimental devices incorporating Josephson junctions, we have shown that our method is accurate and applicable to a large range of quantum circuit architectures. It is directly applicable to a broader class of nonlinear inductive elements, such as weak-link nanobridges (Vijay2010, ; Peltonen2016, ), nanowires (Mooij2006, ; Abay2014, ; Larsen2015, ; DeLange2015-nanowire, ; Casparis2016, ), and kinetic-inductance thin-films (HoEom2012, ; Vissers2015, ; Maleeva2018, ). While best suited for weakly nonlinear systems, the EPR method is derived within circuit theory without approximations. It can be seen as arising from a change of basis adapted to nonlinear elements, as detailed in Supplementary Section A. In practice, the useful reach of the method is set by the numerical ability to include all relevant electromagnetic modes and to compute the spectrum of the extracted Hamiltonian (DiPaolo2019, ). We contribute an open-source package pyEPR 111See the pyEPR (pyEPR, ) code repository at http://github.com/zlatko-minev/pyEPR., which automates the EPR method, and was tested in the design of several further experiments (Leghtas2015, ; Minev2016, ; Mundhada2017, ; Touzard2017, ; Campagne-Ibarcq2018-RemoteEnt, ; Muppalla2017, ; Wang2019-cav-atten, ; Grimm2019, ; Minev2019Nature, ; Campagne-Ibarcq2019, ; Winkel2019, ; Winkel2019a, ).

III METHODS

III.1 Methods of the experiment

Device Frequency (MHz) Anharmonicity (MHz) Cross-Kerr (MHz) I-O coupling
ωD/2πsubscript𝜔D2𝜋\omega_{\mathrm{D}}/2\pi ωB/2πsubscript𝜔B2𝜋\omega_{\mathrm{B}}/2\pi ωC/2πsubscript𝜔C2𝜋\omega_{\mathrm{C}}/2\pi αD/2πsubscript𝛼D2𝜋\alpha_{\mathrm{D}}/2\pi αB/2πsubscript𝛼B2𝜋\alpha_{\mathrm{B}}/2\pi χDB/2πsubscript𝜒DB2𝜋\chi_{\mathrm{DB}}/2\pi χBC/2πsubscript𝜒BC2𝜋\chi_{\mathrm{BC}}/2\pi χDC/2πsubscript𝜒DC2𝜋\chi_{\mathrm{DC}}/2\pi QCsubscript𝑄CQ_{\mathrm{C}}
R9C1 4951 5664 9158 138 170 92. 4. 7 0. 4 5.20×1035.20superscript1035.20\times 10^{3}
4866 5691 9154 150 185 99. 4. 2 0. 55 7.40×1037.40superscript1037.40\times 10^{3}
-1.7% 0.5% -0.04% 8% 8% 7% -12% 27% 29%
R2C1 4823 5567 8947 150 192 64. 5 4. 8 0. 3 4.97×1034.97superscript1034.97\times 10^{3}
4770 5640 8950 161 211 67. 7 5. 88 0. 46 5.44×1035.44superscript1035.44\times 10^{3}
-1.1% 1.3% 0.03% 6.8% 9% 4. 7% 18% 35% 9 %
R7C1 4726 5475 8999 156 189 67. 4. 8 0. 34 2.68×1032.68superscript1032.68\times 10^{3}
4770 5640 8950 161 211 67. 7 5. 88 0. 46 3.07×1033.07superscript1033.07\times 10^{3}
0.9% 2.9% -0.55% 3.1% 10% 1% 18% 26% 13%
R3C2 4845 5620 8979 152 195 61 5. 1 0. 3 2.11×1032.11superscript1032.11\times 10^{3}
4770 5640 8950 161 211 67. 7 5. 88 0. 46 1.78×1031.78superscript1031.78\times 10^{3}
-1.5% 0.4% -0.3% 5.6% 7.6% 9. 9% 13% 35% -19%
R3C1 4688 5300 9003 148 174 85 5. 0. 33 2.43×1032.43superscript1032.43\times 10^{3}
4745 5265 8922 159 198 73 5. 1 0. 37 5.65×1035.65superscript1035.65\times 10^{3}
1.2% -0.7% -0.9% 6.9% 12.1% -16% 2% 9% 57%
DT3 6160 7110 9170 130 150 278 3. 2. 5 9.17×1039.17superscript1039.17\times 10^{3}
6100 7141 9155 140 177 312 3. 9 3. 1 7.33×1037.33superscript1037.33\times 10^{3}
-1.0% 0.4% -0.15% 7% 15% 11% 23% 19% -25%
Table 1: Two-qubit, one-cavity devices. Summary of measured and calculated Hamiltonian and input-output (I-O) coupling parameters for the six devices described in Methods. Indices D, B, C denote the dark, bright, and cavity modes respectively. The input-output quality factor to the readout cavity is denoted QCsubscript𝑄CQ_{\mathrm{C}}. For each device, the first (second) row quantifies the measured, m𝑚m, (bare calculated, c𝑐c) values. The third row quantifies the bare agreement, i.e., (cm)/c𝑐𝑚𝑐\left(c-m\right)/c. In the anharmonicity column, the bare agreement should be corrected by the systematic shift due to our choice to neglect the junction intrinsic capacitance in our modeling (see Methods). We evaluate the correction to be of order 4%, estimated by taking a nominal junction CJ=4fFsubscript𝐶𝐽4fFC_{J}=4\,\mathrm{fF}; hence, an overall corrected agreement of 4.3% for this column.
Device Frequency (MHz) Anharmonicity (MHz) Cross-Kerr (MHz) I-O coupling
ωQ/2πsubscript𝜔Q2𝜋\omega_{\mathrm{Q}}/2\pi ωS/2πsubscript𝜔S2𝜋\omega_{\mathrm{S}}/2\pi ωC/2πsubscript𝜔C2𝜋\omega_{\mathrm{C}}/2\pi αD/2πsubscript𝛼D2𝜋\alpha_{\mathrm{D}}/2\pi χQS/2πsubscript𝜒QS2𝜋\chi_{\mathrm{QS}}/2\pi χQC/2πsubscript𝜒QC2𝜋\chi_{\mathrm{QC}}/2\pi QCsubscript𝑄CQ_{\mathrm{C}}
WG1 4890 7070 7267 310 0.25 0.30 20×10320superscript10320\times 10^{3}
4820 7020 7340 325 0.29 0.33 16×10316superscript10316\times 10^{3}
-1.4% -0.7% 1.0% 4.6% 13% 9% -22%
Table 2: Flip-chip (2.5D), one-qubit, one-storage-cavity, one-readout-cavity devices. Summary of measured and calculated Hamiltonian and input-output (I-O) coupling parameters for the device described in Methods. Indices Q, S, C denote the qubit, storage, and readout cavity modes respectively. The input-output quality factor to the readout cavity is denoted QCsubscript𝑄CQ_{\mathrm{C}}. For each device, the first (second) row quantifies the measured, m𝑚m, (bare calculated, c𝑐c) values. The third row quantifies the bare agreement, i.e., (cm)/c𝑐𝑚𝑐\left(c-m\right)/c.

Device fabrication. Unless otherwise noted, samples were fabricated according to the following methodology. Sample patterns, both large and fine features, were defined by a 100 kV electron-beam pattern generator (Raith EBPG 5000+) in a single step on a PMAA/MAA (Microchem A-4/Microchem EL-13) resist bilayer coated on a 430 μ𝜇\mum thick, double-side-polished, c-plane sapphire wafer, grown with the edge-defined film-fed growth (EFG) technique. Using the bridge-free fabrication technique (Rigetti2009, ; Lecocq2011-bridge-free, ; Pop2012-Junction, ) the Al/AlOxsubscriptOx\text{O}_{\text{x}}/Al Josephson tunnel junctions were formed by a double-angle aluminum evaporation under ultra-high vacuum in a multi-chamber Plassys UMS300 UHV. The two depositions were interrupted by a thermal oxidation step, static 100 Torr environment of 85% argon and 15% oxygen, to form the thin AlOx barrier of the tunnel junction. Prior to the first deposition, to reduce junction aging (Pop2012-Junction, ), the exposed wafer surfaces were exposed to 1 minute oxygen-argon plasma cleaning, under a pressure of 3×1033superscript1033\times 10^{-3} mbar. After wafer dicing (ADT ProVecturs 7100) and chip cleaning, the normal-state resistance RNsubscript𝑅𝑁R_{N} of the Josephson junctions was measured to provide an estimate of the Josephson energy, EJsubscript𝐸𝐽E_{J}, of the device junctions. The junction energy was to first order estimated by an extrapolation of RNsubscript𝑅𝑁R_{N} from room temperature to the operating sample temperature, at approximately 151515 mK, using the Ambegaokar-Baratoff relation (Ambegaokar1963, ),

EJ=12hΔ(2e)2RN1,subscript𝐸𝐽12Δsuperscript2𝑒2superscriptsubscript𝑅𝑁1E_{J}=\frac{1}{2}\frac{h\Delta}{\left(2e\right)^{2}}R_{N}^{-1}\;, (30)

where ΔΔ\Delta is the superconducting gap of aluminum, e𝑒e is the elementary charge, and hh is Planck’s constant.

Frequency (MHz) Anharmonicity (MHz) Cross-Kerr (MHz)
ωD/2πsubscript𝜔D2𝜋\omega_{\mathrm{D}}/2\pi ωB/2πsubscript𝜔B2𝜋\omega_{\mathrm{B}}/2\pi αD/2πsubscript𝛼D2𝜋\alpha_{\mathrm{D}}/2\pi αB/2πsubscript𝛼B2𝜋\alpha_{\mathrm{B}}/2\pi χDB/2πsubscript𝜒DB2𝜋\chi_{\mathrm{DB}}/2\pi
6010 8670 85 180 278
5824 8878 97 206 281
-3.2% 2.3% 12% 13% 1.1%
Table 3: Two-qubit, one-waveguide devices. Summary of measured and calculated Hamiltonian parameters for the device described in Methods . Indices D and B denote the dark and bright modes, respectively. For each device, the first (second) row summarizes the measured, m𝑚m, (bare calculated, c𝑐c) values. The third row quantifies the bare agreement, i.e., (cm)/c𝑐𝑚𝑐\left(c-m\right)/c.

Sample holder. Sample holders were machined in aluminum alloy 6061, seams were formed using thin indium gaskets placed in machined grooves in one of the mating surfaces. Only non-magnetic components were used in proximity to the samples, molybdenum washers, aircraft-alloy 7075 screws (McMaster/Fastener Express) with less than 1% iron impurities, and non-magnetic SubMiniature version A (SMA) connectors.

Cryogenic setup. Samples were thermally anchored to the 15 mK stage of a cryogen-free dilution refrigerator (Oxford Triton 200) and were measured using a standard cQED measurement setup (Wallraff2004, ; Leghtas2015, ; Minev2019-Thesis, ). High-magnetic-permeability, μ𝜇\mu-metal (Amumetal A4K) shields together with aluminum superconducting shields enclosed all samples. Microwave input and output lines were filtered with Eccosorb CR-110 infrared-frequency filters (Rigetti2012, ; Geerlings2013, ), thermally anchored at the 15 mK stage. Output lines were additionally filtered with cryogenic isolators (Quinstar CWJ1019-K414) and 12 GHz K&L multi-section lowpass filters. Output lines leading up to the high-electron-mobility transistor (HEMT) amplifier (Low Noise Factory), anchored at 4 K, were superconducting (CoaxCo Ltd. SC-086/50-NbTi-NbTi PTFE).

Quantum amplifier. The output signal of a sample was processed by a Josephson parametric converter (JPC) anchored at the 15 mK stage and operated in amplification mode (Bergeal2010, ; Abdo2013, ), before routing to the HEMT. The JPC provide a typical gain of 21 dB with a typical noise-visibility ratio of 6 dB. See Ref. (Roy2017-Review, ) for a review of the parametric amplification.

Frequency and input-output (I-O) coupling measurements. Spectroscopic measurements were used to determine the frequencies of the resonator modes. Anharmonicities were determined in two-tone spectroscopy (Geerlings2013, ; Reagor2016, ). Cross-Kerr energies were determined from dressed dephasing measurements (Gambetta2006-dephasing, ; Gambetta2008-qm-traj, ). In particular, the dressed-dephasing measurement sequence consisted of first preparing the qubit in the ground state, then exciting it to the equator by a π/2𝜋2\pi/2 pulse. Subsequently, a weak readout tone excited the readout cavity of the qubit for a fixed duration, 10 times the readout cavity lifetime κrsubscript𝜅𝑟\kappa_{r}, after which we measure the qubit X and Y Bloch vectors, after waiting for a time 5/κr5subscript𝜅𝑟5/\kappa_{r} for any photons in the cavity to leak out. By varying the amplitude and frequency of the applied weak-readout tone, we could calibrate both the strength of our readout, in steady-state photon number in the readout cavity, and the value of the cross-Kerr frequency shift between the qubit and readout resonator. The values could be obtained from fits of the X and Y quadratures. For each sample, the coupling quality factor of the readout-cavity mode, denoted QCsubscript𝑄𝐶Q_{C}, was extracted from the spectroscopic response of the readout cavity at low photon numbers (Geerlings2013, ; Reagor2016, ), by measuring the scattering parameters, S21subscript𝑆21S_{21} or S11subscript𝑆11S_{11}.

To test EPR’s robustness to experimental variability and its applicability over wide range of experimental conditions, the presented samples were fabricated in multiple runs and measured in different cooldowns. Some devices were subjected to as many as 6 thermal cycles.

The Hamiltonian parameters and coupling energies for each sample were also calculated, following the EPR method presented in section on the general approach. In particular, we modeled the sample geometry and materials in a FE electromagnetic simulation, as explicated in Supplementary Section C. Our aim in writing this supplementary section has been to provide an easy access point to the practical use of the EPR method, which we hope will benefit the reader, and allow them to adopt it easily. Our choice of simulation software was the Ansys High Frequency Electromagnetic Field Simulation (HFSS), although we emphasize that the EPR ideas translate to any standard EM eigenmode simulation package. Further, we modeled the loss due to the input-output couplers in the simulation as 50Ω50Ω50\,\Omega resistive sheets, see Supplementary Section D4. The eigenmode analysis provided the calculated I-O quality factors and Purcell limits. All electromagnetic and quantum analyses, including the extraction of participations form the eigenfields and the numerical diagonalization of the Hamiltonian to extract its quantum spectrum, were performed in a fully automated manner using the freely available pyEPR package (Note1, ).

The mode quality due to the input-output coupling, QCsubscript𝑄𝐶Q_{C}, was set by the length of the I-O SMA-coupler pin. Its length inside the sample-holder box was measured at room temperature using calipers. This nominal length was used then used in the HFSS model to create a 3D model of the pin inside the sample holder. The quality factor QCsubscript𝑄𝐶Q_{C} was then obtained from the eigenmode eigenvalue. We remark that the measurement of the pin-length is accurate to no more than 20%; further, it can be affected by various idiosyncrasies, such as bending of the thin SMA center pin. Nonetheless, the predictions of the quality factors for low-Q modes were observed to be very reasonable estimates, and, similarly, the predicted Purcell limits for qubit and high-Q cavity modes were consistent with estimates from measurements.

Refer to caption
Figure 5: Two–qubit, one-cavity devices. a and b Not-to-scale diagram illustrating chip configurations A and B, respectively. Vertical blue arrow indicates cavity electric field orientation. Crosses mark the location of Josephson tunnel junctions. c Optical photograph of sample R2C1. Bottom half of aluminum sample holder is visible; top half is removed. The two-qubit chip (outlined by the dashed green box) is housed in the middle of the readout cavity (highlighted in blue). Cavity fundamental mode electric field profile E𝐸\vec{E} depicted by arrows. Input-output SMA pin coupler labeled I-O.

III.2 Devices

III.2.1 Two-qubit, one-cavity devices
Device description.

We measured 6 samples that were each comprised of two qubits and one cavity. The cavity was a standard, machined aluminum cavity (Paik2011, ). It housed either one or two sapphire chips, which were either patterned with transmon qubits or simply blank. Each transmon consisted of two thin-film aluminum pads connected by a Josephson junction. We tested two configurations of chips and patterns. Configuration A consisted of one chip with two orthogonal qubits, as depicted in Fig. 5(a). Similarly, configuration B consisted of one chip with two parallel qubits, depicted in Fig. 5(b). The two qubits were aligned parallel to each other; however, unlike configuration B, there was no galvanic connection between them. The results of the measurements are presented in Table 1.

Samples R1C9, R2C1, R7C1, and R2C1, R3C2 were fabricated in configuration A, sample DT3 was fabricated in configuration B. Three of the sample (R2C1, R7C1, R3C1) were fabricated simultaneously on the same sapphire wafer, all with nominally identical dimensions. Additionally, R2C1 and R7C1 were designed to have nominally the same Josephson junctions energy, Ejsubscript𝐸𝑗E_{j}. The rest of the samples (R1C9, DT3, and R3C2) were fabricated at different times and on different wafers. The dimensions of their transmons and the inductance of the junctions were designed to be different. Only sample R3C2 was designed to be very similar to the nominally identical sample R2C1 and R7C1, but with adjusted Ejsubscript𝐸𝑗E_{j}. For samples R2C1, R7C1, R3C1, and R3C2 a second, un-processed, un-patterned, blank sapphire chip was placed in parallel with the qubit carrying chip [see Fig. 5(c)] to purposefully lower the readout cavity frequency, thus bringing it within the JPC amplification band.

Configurations A and B were designed to test the ability of the EPR method to calculate the mixing between strongly coupled modes. The strong coupling was achieved in two distinct ways. First, configuration A used the spatial proximity of the two qubits to yielded a strong capacitive coupling between them, which resulted in large qubit-qubit mixing. Second, instead of spatial proximity, configuration B used a galvanic connection between the qubits to yield strong hybridization. Our two-qubit designs share some similar-in-spirit characteristics with the promising recent developments reported in Refs. Gambetta2011-Purcell, ; Srinivasan2011, ; Diniz2013, ; Dumur2015, ; Zhang2017, ; Roy2017-3qubits, , but our implementation is distinct and is designed to provide several unique advantages.

Mode structure and interesting physical insights.

Configuration A is characterized by strong capacitive coupling between the two transmons, which have different pad sizes, see Fig. 5(c), and hence different normal-mode frequencies. Due to the strong hybridization, each qubit normal mode consists of some excitation in the vertical and some in the horizontal transmon. With some foresight, we will label the vertical mode bright (B), and the horizontal dark (D). The bright-mode resonance is higher in frequency, and thus is closer to the resonance of the readout cavity mode (C). This smaller detuning made it a natural choice for designing stronger coupling between it, (B), and the readout mode (C). This was implemented by orienting the transmon design that participates in mode (B) vertical.

To understand this design choice, let us first consider the popular analogy (Koch2007, ; Devoret2007-HowStrongCanCouplingBe, ) between circuit-QED and cavity-QED, often used to discuss mode couplings. In this atomic analogy, the transmon qubit is analogous to a real atom inside the cavity. Thus, it can described by an electric dipole moment dBsubscriptd𝐵\vec{\mathrm{d}}_{B}. Meanwhile, its coupling, cross-Kerr, etc. to the cavity mode are derived from the electric-dipole coupling interaction. In particular, the coupling amplitude is proportional to dBEsubscriptd𝐵𝐸\vec{\mathrm{d}}_{B}\cdot\vec{E}, where E𝐸\vec{E} is the cavity electric field at the transmon junction. From this analogy, one can infer that the coupling is maximized when the two are parallel, dBEconditionalsubscriptd𝐵𝐸\vec{\mathrm{d}}_{B}\parallel\vec{E}, and one could hope to measure a strong cross-Kerr between the bright qubit and the cavity. This successful conclusion is true, but a coincidence. We will shortly discuss how this popular analogy fails spectacularly for the dark mode in configuration B. Instead, we will argue that a correct way to understand the nonlinear coupling between the two modes is through the participation ratio, which will provide the correct coupling for both configuration A and B.

Before proceeding to configuration B, we note one further useful features that configuration A exhibits. In particular, while the bright qubit mode can be Purcell limited (Houck2008-Purcell, ; Gambetta2011-Purcell, ), the dark mode is simultaneously Purcell protected. Thus, one can potentially achieve a high ratio in the I-O bath coupling of the two qubits.

Configuration B has two qubit modes, which we will also label dark (D) and bright (B). Since both transmons are designed with the exact same transmon pad geometry and junction energy EJsubscript𝐸𝐽E_{J}, see Fig. 5, we can expect that no single junction is preferred, due to the symmetry of the sample. This is in sharp contrast to the asymmetric energy distribution in configuration A. Returning to configuration B, we can estimate that in each qubit mode, both junctions participate equally and with near maximal allowed participation,

pD1=pD2=pB1=pB212.subscript𝑝D1subscript𝑝D2subscript𝑝B1subscript𝑝B212p_{\mathrm{D}1}=p_{\mathrm{D}2}=p_{\mathrm{B}1}=p_{\mathrm{B}2}\approx\frac{1}{2}\;. (31)

If the two transmons were well-separated spatially and not connected, they would be uncoupled. However, the galvanic connection between the two lower pads, see Fig. 5(a), results in a very strong hybridization and splitting between the nominally identical transmons. The result of the strong hybridization is a symmetric and antisymmetric combination of the two bare transmons. In other words, the hybridization results in a common mode, namely (B), where both junctions oscillate in-phase, and a differential mode namely (D), where both junctions oscillate out-of-phase. These phase relationships are captured by the signs:

sD1subscript𝑠D1\displaystyle s_{\mathrm{D}1} =1,absent1\displaystyle=1\;, sD2subscript𝑠D2\displaystyle s_{\mathrm{D}2} =1,absent1\displaystyle=-1\;, (32)
sB1subscript𝑠B1\displaystyle s_{\mathrm{B}1} =1,absent1\displaystyle=1\;, sB2subscript𝑠B2\displaystyle s_{\mathrm{B}2} =+1.absent1\displaystyle=+1\;. (33)

In an attempt to understand how these hybridized qubit modes will couple to the cavity mode (C), let us first consider the atomic analogy again. When the two junctions oscillate in phase, in the (B) mode, the net dipole moment of the bright mode, dBsubscript𝑑B\vec{d}_{\mathrm{B}}, must be large, since it is the sum of the two junction dipole contributions. Secondly, dBsubscript𝑑B\vec{d}_{\mathrm{B}} must be oriented in the vertical direction, parallel to the cavity electric field E𝐸\vec{E}. Hence, we would conclude that the bright mode coupling is large, dBE0much-greater-thansubscript𝑑B𝐸0\vec{d}_{\mathrm{B}}\cdot\vec{E}\gg 0, and there should be a strong cross-Kerr interaction between the cavity and bright qubit. Continuing the analogy in the case of the dark mode, we would deduce that the net dipole moment of the dark mode is zero, since the two junctions oscillate out of phase, and cancel each other’s contribution, dD=0subscript𝑑D0\vec{d}_{\mathrm{D}}=0. Thus, we should not expect any coupling between the dark qubit and the cavity mode, dDE=0subscript𝑑D𝐸0\vec{d}_{\mathrm{D}}\cdot\vec{E}=0. To the contrary of this conclusion, as can be seen in the measured results in Table 1, the nonlinear coupling of the dark and bright qubit to the cavity is nearly equal. The atomic analogy and the dipole argument have failed completely. We can understand the origin of this failure and how to arrive at the correct conclusion by using the energy-participation ratio. As embodied in Eq. (26), in the dispersive regime, the nonlinear coupling between two modes, in this case a qubit and cavity, is given by the overlap of the EPR distribution. In particular, the cross-Kerr amplitude between the dark qubit and the readout cavity mode is given by

χDCsubscript𝜒DC\displaystyle\chi_{\mathrm{DC}} =ωDωC4EJ(pD1pC1+pD2pC2),absentPlanck-constant-over-2-pisubscript𝜔𝐷subscript𝜔𝐶4subscript𝐸𝐽subscript𝑝D1subscript𝑝C1subscript𝑝D2subscript𝑝C2\displaystyle=\frac{\hbar\omega_{D}\omega_{C}}{4E_{J}}\left(p_{\mathrm{D}1}p_{\mathrm{C}1}+p_{\mathrm{D}2}p_{\mathrm{C}2}\right)\;, (34)

where both junctions have the same junction energy EJsubscript𝐸𝐽E_{J}, and ωBsubscript𝜔𝐵\omega_{B} (resp: ωCsubscript𝜔𝐶\omega_{C}) denotes the dark qubit (resp: cavity) mode frequency. The signs, used in the atomic dipole logic do not factor into the coupling, because the Josephson mechanics is fundamentally different. To obtain χBCsubscript𝜒BC\chi_{\mathrm{BC}}, one can replace the label ‘D’ with ‘B’ in Eq. (34). Then, it is easy to use Eq. (31) to show that the ratio of two Kerr couplings is not zero, but rather of order unity,

χBC/χDC=ωB/ωD.subscript𝜒BCsubscript𝜒DCsubscript𝜔𝐵subscript𝜔𝐷\chi_{\mathrm{BC}}/\chi_{\mathrm{DC}}=\omega_{B}/\omega_{D}\;. (35)
Failure of the conventional dipole approach.

We showed that although the heuristic atomic analogy seems seductively accurate, it fails completely in some cases to predict the nonlinear couplings. Instead, one can use the intuition and calculation method provided by the EPRs.

As an added note, we observe that Eqs. (32) and (32) embodies the orthogonality of the participations, see Eq. (24). We also remark that although the atomic analogy fails in the case of the nonlinear couplings, it can yield some guidance when considering the linear mixing of the modes, useful for discussing the Purcell effect. To illustrate, let us briefly extend the atomic analogy. The dipole-like coupling between the bright mode and the cavity suggests that the bright mode will inherit some coupling to the environment, mediated by the cavity. Thus, since dBE0much-greater-thansubscript𝑑B𝐸0\vec{d}_{\mathrm{B}}\cdot\vec{E}\gg 0, we can expect the bright qubit to potentially be Purcell limited. In contrast, since dDE=0subscript𝑑D𝐸0\vec{d}_{\mathrm{D}}\cdot\vec{E}=0, we could expect the dark qubit to be Purcell protected. Both of these qualitative Purcell predictions are valid, but to quantify them, we will use the EPR method and FE eigenmode simulation of the sample, as will be discussed shortly.

Experimental results.

Table 1 summarizes the results of the agreement between the measured and calculated Hamiltonian parameters for all two-qubit, one-cavity samples. The three modes in each sample are labeled dark (D), bright (B), and cavity (C); the reason for this convention is described above. In all samples, the qubits were designed to be in the dispersive regime with respect to the cavity, which was detuned by 2–4 GHz. However, in a large number of the samples, the two qubits were strongly hybridized, often necessitating higher-order nonlinear corrections to be included in the calculation. This strong hybridization was used as a test of the theory in this more challenging and fickle regime.

In total, for each sample we measured and calculated 8 frequency parameters and one dimensionless, coupling quality factor, QCsubscript𝑄𝐶Q_{C}, of the readout cavity mode. In particular, in the low-excitation limit, the nonlinear interactions among the modes were characterized by the effective dispersive Hamiltonian

H¯^/=ωDn^D+ωBn^B+ωCn^C12αDn^D(n^D1^)12αBn^B(n^B1^)χDBn^Dn^BχDCn^Dn^CχBCn^Bn^C,^¯𝐻Planck-constant-over-2-pisubscript𝜔𝐷subscript^𝑛𝐷subscript𝜔𝐵subscript^𝑛𝐵subscript𝜔𝐶subscript^𝑛𝐶12subscript𝛼𝐷subscript^𝑛𝐷subscript^𝑛𝐷^112subscript𝛼𝐵subscript^𝑛𝐵subscript^𝑛𝐵^1subscript𝜒𝐷𝐵subscript^𝑛𝐷subscript^𝑛𝐵subscript𝜒𝐷𝐶subscript^𝑛𝐷subscript^𝑛𝐶subscript𝜒𝐵𝐶subscript^𝑛𝐵subscript^𝑛𝐶\hat{\overline{H}}/\hbar=\omega_{D}\hat{n}_{D}+\omega_{B}\hat{n}_{B}+\omega_{C}\hat{n}_{C}\\ -\frac{1}{2}\alpha_{D}\hat{n}_{D}\left(\hat{n}_{D}-\hat{1}\right)-\frac{1}{2}\alpha_{B}\hat{n}_{B}\left(\hat{n}_{B}-\hat{1}\right)\\ -\chi_{DB}\hat{n}_{D}\hat{n}_{B}-\chi_{DC}\hat{n}_{D}\hat{n}_{C}-\chi_{BC}\hat{n}_{B}\hat{n}_{C}\;, (36)

where n^D,n^Bsubscript^𝑛𝐷subscript^𝑛𝐵\hat{n}_{D},\hat{n}_{B}, and n^Csubscript^𝑛𝐶\hat{n}_{C} denote the dark, bright, and cavity photon-number operator, respectively. The coupling of the resonator mode to the bath is given by the Lindblad superoperator term κC𝒟[a^C]ρsubscript𝜅𝐶𝒟delimited-[]subscript^𝑎𝐶𝜌\kappa_{C}\mathcal{D}[\hat{a}_{C}]\rho, where κC=ωC/QCsubscript𝜅𝐶subscript𝜔𝐶subscript𝑄𝐶\kappa_{C}=\omega_{C}/Q_{C}, and ρ𝜌\rho is the density operator.

We remark that all samples in configuration A demonstrated a large asymmetry in the Kerr coupling between the bright-to-cavity and dark-to-cavity coupling, χBCχDCmuch-greater-thansubscript𝜒𝐵𝐶subscript𝜒𝐷𝐶\chi_{BC}\gg\chi_{DC}. In contrast, samples in configuration B demonstrated near equal coupling, χBCχDCsubscript𝜒𝐵𝐶subscript𝜒𝐷𝐶\chi_{BC}\approx\chi_{DC}. In both configurations A and B, the dark mode was Purcell protected, we calculated a Purcell coupling factor of QPurcellD107much-greater-thansuperscriptsubscript𝑄Purcell𝐷superscript107Q_{\text{Purcell}}^{D}\gg 10^{7}, using the eigenmode method described in Supplementary Section D4. On the other hand, the bright mode was somewhat Purcell limited, QPurcellB106superscriptsubscript𝑄Purcell𝐵superscript106Q_{\text{Purcell}}^{B}\approx 10^{6}. From the relative Rabi amplitudes of the dark and bright qubits, we could verify the order of magnitude scaling calculated for the Purcell effect.

Refer to caption
Figure 6: Schematic representation (not to scale) of superconducting waveguide device DTW1. Two-qubit chip is placed λ/4𝜆4\lambda/4 from a short termination in the waveguide. Guided input-output waves are launched and monitored through the I-O port connecting to a standard SMA adapter (not shown).
III.2.2 Two-qubit, single-waveguide devices

We measured a two qubit sample inside of a waveguide. Figure 6 presents the setup, and depicts the sample, which was of the configuration B type presented in Fig. 5. The sample chip was positioned inside an aluminum WR90 waveguide. The waveguide was terminated in a short at one side, and attached to an impedance-matched SMA coupler port on the input-launcher side, which was used to drive and measure the waveguide. The chip was centered inside the cross-section of the waveguide, and placed λ/4𝜆4\lambda/4 away from the termination wall, at the measurement frequency. The rest of the experimental setup was identical to that described in section ‘Methods of the experiment’. The two qubit modes were labeled dark and bright, similarly to the samples discussed in the section ‘Two-qubit, one-cavity devices.’.

Table 3 presents the agreement between the measured and calculated key Hamiltonian parameters of the sample. These consist of the two mode frequencies, two qubit anharmonicities, and the strong cross-Kerr interaction between the two qubits.

Refer to caption
Figure 7: Illustration of flip-chip (2.5D) device WG1. a Depiction (not to scale) of chip stack consisting of two chips separated by a 100 \upmum\upmum\upmu\text{m} vacuum gap. The inner face of each chip supports part of the pattern of a multi-layer whispering gallery-mode resonator (WGMR) resonator (Minev2013, ). The lower layer contains a 2.5D, aperture transmon qubit (Minev2016, ) embedded in the WGMR. b Zoomed-in view of the lower layer of the aperture transmon. The cross marks the location of the Josephson tunnel junction device, which connects the lower center trace to an island embedded in the lower aperture. The capacitance to the top layer significantly participates in determining qubit parameters.
III.2.3 Flip-chip (2.5D), one-qubit, one-storage-cavity, one-readout-cavity devices

We also designed a multilayer planar (Minev2016, ) (2.5D) circuit-QED sample, depicted in Fig. 7, with the EPR method. It consisted of high-Q storage mode (S), one low-Q readout cavity (C), and one control transmon qubit (Q). The two cavity modes were formed in the footprint of a single whispering gallery mode resonator (Minev2013, ). The three modes were in the dispersive regime, and the storage mode was used to encode and decode quantum information, as well as to observe parity revivals. Details of the sample design have been reported in Ref. Minev2016, . The agreement between the measured parameters of the sample and those obtained by the EPR calculation methods are presented in Table 2.

Data availability.

Data are available from the authors on reasonable request.

Code availability.

The source code for the EPR method is open-sourced and can be found at http://github.com/zlatko-minev/pyEPR.

Acknowledgments.

We thank S.M. Girvin, R.J. Schoelkopf, S. Nigg, H. Paik, A. Blais, H.E. Türeci, F. Solgun, V. Sivak, S. Touzard, N. Frattini, S. Shankar, C. Axline, V.V. Albert, K. Chou, A. Petrescu, D. Cody, A. Eickbusch, and E. Flurin for valuable discussions, and the I. Siddiqi and B. Huard groups for using the early versions of pyEPR. This research was supported by the US Army Research Office (ARO) Grant No. W911NF-18-1-0212. Z.K.M. acknowledges partial support from the ARO (W911NF-16-1-0349). M.H.D. acknowledges partial support from the ARO (W911NF-18-1-0020) and the Air Force Office of Scientific Research (FA9550-19-1-0399). The view and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of the Army Research Office or the US Government. The US Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation herein.

Author contributions.

Z.K.M. conceived and developed the method and the theory, performed the experiments, and analyzed the data. Z.L. and M.H.D. contributed to the theory. I.M.P. and S.O.M. contributed to the measurement and fabrication of the devices L.C. assisted with the simulations. Z.K.M. and M.H.D. wrote the manuscript. All authors provided suggestions, discussed the results and contributed to the manuscript.

Competing interests.

The authors declare no competing interests.

Supplementary Information:
Energy-participation quantization of Josephson circuits

Supplementary Section A Theoretical foundation of the energy-participation method

Supplementary Table 1: Table of key symbols and relationships used in the derivation of the energy-participation-ratio method.
Basic Josephson circuit variables and parameters
Symbol Value Description
Φ0subscriptΦ0\Phi_{0} h/(2e)2𝑒h/(2e) Superconducting magnetic-flux quantum—the ratio of Planck’s quantum of electromagnetic action hh and the charge of a Cooper pair 2e2𝑒2e.
 ϕ0subscriptitalic-ϕ0\phi_{0} Φ0/(2π)subscriptΦ02𝜋\Phi_{0}/\left(2\pi\right) Reduced flux quantum.
 Φb(t)subscriptΦ𝑏𝑡\Phi_{b}\left(t\right) tvb(τ)dτsuperscriptsubscript𝑡subscript𝑣𝑏𝜏differential-d𝜏\int_{-\infty}^{t}v_{b}\left(\tau\right)\mathrm{d}\tau Generalized magnetic flux of circuit branch b𝑏b, at time t𝑡t. The instantaneous voltage vb(τ)subscript𝑣𝑏𝜏v_{b}\left(\tau\right) across the terminals of branch b𝑏b at time τ𝜏\tau is equivalently denoted Φ˙b(τ)subscript˙Φ𝑏𝜏\dot{\Phi}_{b}\left(\tau\right). In general, Φb(t)=tvb(τ)dτ+Φb()subscriptΦ𝑏𝑡superscriptsubscript𝑡subscript𝑣𝑏𝜏differential-d𝜏subscriptΦ𝑏\Phi_{b}\left(t\right)=\int_{-\infty}^{t}v_{b}\left(\tau\right)\mathrm{d}\tau+\Phi_{b}\left(-\infty\right), but we take the initial flux Φb()subscriptΦ𝑏\Phi_{b}\left(-\infty\right) to be zero, corresponding to the circuit in an equilibrium state (see Sec. A8). With this convention, ΦbsubscriptΦ𝑏\Phi_{b} is a deviation away from equilibrium. This variable is analogous to the elongation of a mechanical spring.
 j,J𝑗𝐽j,\,J The subscript or index j𝑗j denotes the j𝑗j-th Josephson dipole, where j{1,,J}𝑗1𝐽j\in\left\{1,\,\ldots,\,J\right\}, and J𝐽J denotes the total number of Josephson dipoles.
 Φj(t)subscriptΦ𝑗𝑡\Phi_{j}\left(t\right) Generalized magnetic-flux deviation of a non-linear device, the j𝑗j-th Josephson dipole.
 φj(t)subscript𝜑𝑗𝑡\varphi_{j}\left(t\right) Φj(t)/ϕ0subscriptΦ𝑗𝑡subscriptitalic-ϕ0\Phi_{j}\left(t\right)/\phi_{0} Reduced magnetic-flux variable of Josephson dipole j𝑗j.
 j(Φj)subscript𝑗subscriptΦ𝑗\mathcal{E}_{j}\left(\Phi_{j}\right) jlin(Φj)+jnl(Φj)superscriptsubscript𝑗linsubscriptΦ𝑗superscriptsubscript𝑗nlsubscriptΦ𝑗\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right)+\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right) Energy function of Josephson dipole j𝑗j. Typically split into two component: jlinsuperscriptsubscript𝑗lin\mathcal{E}_{j}^{\mathrm{lin}} associated with linear interactions and jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} associated with nonlinear ones.
 jlin(Φj)superscriptsubscript𝑗linsubscriptΦ𝑗\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) 12Ej(Φj/ϕ0)212subscript𝐸𝑗superscriptsubscriptΦ𝑗subscriptitalic-ϕ02\frac{1}{2}E_{j}\left(\Phi_{j}/\phi_{0}\right)^{2} Linear component of the energy function of junction j𝑗j, defined by the energy scale Ejsubscript𝐸𝑗E_{j}. Terms linear in ΦjsubscriptΦ𝑗\Phi_{j} are absent since we selected an equilibrium operating point of the circuit. The energy function jlinsuperscriptsubscript𝑗lin\mathcal{E}_{j}^{\mathrm{lin}} may never contain non-linear terms.
 jnl(Φj)superscriptsubscript𝑗nlsubscriptΦ𝑗\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right) j(Φj)jlin(Φj)subscript𝑗subscriptΦ𝑗superscriptsubscript𝑗linsubscriptΦ𝑗\mathcal{E}_{j}\left(\Phi_{j}\right)-\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) The nonlinear component of the energy function of junction j𝑗j. For certain situations, it may be favorable to select the partition of jsubscript𝑗\mathcal{E}_{j} such that jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} contains linear interactions. If possible, it is often convenient to expand jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} in a Taylor series around the circuit operating point: jnl(Φj)=Ejp=3cjp(Φjϕ0)psuperscriptsubscript𝑗nlsubscriptΦ𝑗subscript𝐸𝑗superscriptsubscript𝑝3subscript𝑐𝑗𝑝superscriptsubscriptΦ𝑗subscriptitalic-ϕ0𝑝\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right)=E_{j}\sum_{p=3}^{\infty}c_{jp}\left(\frac{\Phi_{j}}{\phi_{0}}\right)^{p}, where cjpsubscript𝑐𝑗𝑝c_{jp} are the dimensionless coefficients of the expansion.
 𝚽t(t)subscript𝚽t𝑡\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\left(t\right) (Φt1,Φt2,)superscriptsubscriptΦsubscriptt1subscriptΦsubscriptt2\left(\Phi_{\mathrm{t}_{1}},\,\Phi_{\mathrm{t}_{2}},\,\ldots\right)^{\intercal} Column vector consisting of the flux deviations of all circuit branches enclosed in the minimum spanning tree. The roman subscript tt\mathrm{t} denotes the tree. The flux of individual tree-branches are denoted Φt1,Φt2,subscriptΦsubscriptt1subscriptΦsubscriptt2\Phi_{\mathrm{t}_{1}},\,\Phi_{\mathrm{t}_{2}},\,\ldots
Linearized Josephson eigenmodes
 m,M𝑚𝑀m,\;M The label m𝑚m indexes sets associated with the eigenmodes of the linearized Hamiltonian linsubscriptlin\mathcal{H}_{\mathrm{lin}}. We consider m{1,,M}𝑚1𝑀m\in\left\{1,\,\ldots,\,M\right\}, where M𝑀M is the total number of eigenmodes of relevance determined by the context. That is, M𝑀M is either the total number of eigenmodes of linsubscriptlin\mathcal{H}_{\mathrm{lin}} or of the relevant eigenmodes included in the finite-element simulation.
 fullsubscriptfull\mathcal{H}_{\mathrm{full}} lin+nlsubscriptlinsubscriptnl\mathcal{H}_{\mathrm{lin}}+\mathcal{H}_{\mathrm{nl}} Hamiltonian function of the Josephson system. It can be partitioned into linsubscriptlin\mathcal{H}_{\mathrm{lin}}, a part comprised of purely linear terms (quadratic in 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}, or equivalently 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}}), and nlsubscriptnl\mathcal{H}_{\mathrm{nl}}, a part generally comprised of non-linear terms (higher-than quadratic in 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}, or equivalently 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}}).
 𝚽m,𝐐msubscript𝚽msubscript𝐐𝑚\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\,\mathrm{\mathbf{Q}}_{m} Column vectors (of length M𝑀M) whose elements are the eigenmode flux ΦmsubscriptΦ𝑚\Phi_{m} (considered as the generalized position) and charge Qmsubscript𝑄𝑚Q_{m} (considered as the generalized momentum) canonical variables, associated with the Hamiltonian linsubscriptlin\mathcal{H}_{\mathrm{lin}}.
 ωmsubscript𝜔𝑚\omega_{m} Eigenfrequency of the m𝑚m-th mode of linsubscriptlin\mathcal{H}_{\mathrm{lin}}, carries a dimension of circular frequency.
 𝛀𝛀\mathrm{\mathbf{\Omega}} Diag(ω1,ω2,)\operatorname{Diag}\left(\omega_{1},\omega_{2},\ldots\right)^{\intercal} Diagonal M×M𝑀𝑀M\times M matrix comprising the eigenfrequencies.
 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} 𝐄𝚽m𝐄subscript𝚽m\mathrm{\mathbf{E}}\mathrm{\mathbf{\Phi}}_{\mathrm{m}} The spanning-tree branch fluxes 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} are a linear combination of the eigenmode 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}} fluxes. The two are related by an affine transformation given by the properly-constructed eigenvector matrix 𝐄𝐄\mathrm{\mathbf{E}}, see Eqs. (A.38) and (A.37).
Quantum operators and energy-participation ratios
 H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}} m=1Mωma^ma^msuperscriptsubscript𝑚1𝑀Planck-constant-over-2-pisubscript𝜔𝑚superscriptsubscript^𝑎𝑚subscript^𝑎𝑚\sum_{m=1}^{M}\hbar\omega_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m} Hamiltonian operator corresponding to the Hamiltonian function linsubscriptlin\mathcal{H}_{\mathrm{lin}}, expressed in second quantization with respect to the eigenmodes of linsubscriptlin\mathcal{H}_{\mathrm{lin}}. The eigenmodes (expressed in terms of 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}} and 𝐐msubscript𝐐𝑚\mathrm{\mathbf{Q}}_{m}) define the bosonic amplitude (lowering) operators a^msubscript^𝑎𝑚\hat{a}_{m}.
 φ^jsubscript^𝜑𝑗\hat{\varphi}_{j} m=1Jφmja^m+H.c.superscriptsubscript𝑚1𝐽subscript𝜑𝑚𝑗subscript^𝑎𝑚H.c.\sum_{m=1}^{J}\varphi_{mj}\hat{a}_{m}+\text{H.c.} Operator corresponding to the reduced flux φjsubscript𝜑𝑗\varphi_{j} of Josephson dipole j𝑗j; φ^jsubscript^𝜑𝑗\hat{\varphi}_{j} admits a linear decomposition in terms of the mode operators a^msubscript^𝑎𝑚\hat{a}_{m}. The coefficients of this expansion are the quantum zero-point fluctuations φmjsubscript𝜑𝑚𝑗\varphi_{mj}.
 φmjsubscript𝜑𝑚𝑗\varphi_{mj} smjωmEjpmjsubscript𝑠𝑚𝑗Planck-constant-over-2-pisubscript𝜔𝑚subscript𝐸𝑗subscript𝑝𝑚𝑗s_{mj}\sqrt{\frac{\hbar\omega_{m}}{E_{j}}p_{mj}} The quantum zero-point fluctuation (ZPF) of the reduced magnetic flux ϕ^jsubscript^italic-ϕ𝑗\hat{\phi}_{j} of Josephson dipole j𝑗j due to mode m𝑚m; i.e., for a given mode, the ZPF magnitude gives the non-zero standard deviation of the magnetic flux in the ground state. The amplitude of the ZPF φmjsubscript𝜑𝑚𝑗\varphi_{mj} is determined by the physical structure of mode m𝑚m, and can be understood in terms of the energy-participation ratio pmjsubscript𝑝𝑚𝑗p_{mj} and its sign smjsubscript𝑠𝑚𝑗s_{mj}.
 pmj,pmlsubscript𝑝𝑚𝑗subscript𝑝𝑚𝑙p_{mj},\;p_{ml} Energy-participation ratios (EPR) of Josephson dipole j𝑗j and lossy element l𝑙l in mode m𝑚m, comprised between zero and unity, 0pmj,pml1formulae-sequence0subscript𝑝𝑚𝑗subscript𝑝𝑚𝑙10\leq p_{mj},p_{ml}\leq 1.
 smj,smlsubscript𝑠𝑚𝑗subscript𝑠𝑚𝑙s_{mj},\;s_{ml} Sign smjsubscript𝑠𝑚𝑗s_{mj} (resp., smlsubscript𝑠𝑚𝑙s_{ml}) of the energy-participation ratio pmjsubscript𝑝𝑚𝑗p_{mj} (resp., lossy EPR pmlsubscript𝑝𝑚𝑙p_{ml}). The value of the sign is either +11+1 or 11-1. See Supplementary Figure S8.

In this section, we derive the EPR method from first principles. We first review quantum electromagnetic circuit theory (Sec. A1), then use it to find the quantum eigenmodes of the circuit (Secs. A2A5). In Sec. A6, we define the EPR of a Josephson dipole in a quantum mode, and use it to find the quantum zero-point fluctuations (ZPF). The universal properties and sum rules of the EPRs are detailed in Sec. A7.

In Section A8, we detail use of the EPR method for biased systems—those that incorporate active elements (such as current and voltage sources) or external bias conditions that result in persistent currents (such as a frustration by an external magnetic flux).

The EPR derivation consists of a series of exact transformation. No approximation are made in deriving the EPR or in using it to find the quantum Hamiltonian of a general Josephson system. In this sense, the results are universal. Approximation are made in practice when using numerical methods.

A1 Review of electrical circuit theory, and the Josephson junction

We briefly recount the basic formulation of classical electrical theory. This formulation will be used as a stepping stone in the following derivation. We focus on the lumped-element regime. An electrical element is said to be in the lumped-element regime when its physical dimensions are negligible with respect to the electromagnetic wavelengths considered in the analysis. In other words, self-resonances and parasitic internal dynamical degrees of freedom of the element are neglectable.

Electrical theory describes the physical laws governing four basic manifestations of electricity. For each element, these are captured in the variables for voltage v(t)𝑣𝑡v\left(t\right), current i(t)𝑖𝑡i\left(t\right), charge Q(t)𝑄𝑡Q\left(t\right), and magnetic flux Φ(t)Φ𝑡\Phi\left(t\right), as functions of time t𝑡t. Maxwell’s equations and conservation of charge lead to the six universal relationships presented in Eqs. (A.1)–(A.3) (feynmanLec-Vol-II, ; Pozar, ).

The time-instantaneous voltage v(t)𝑣𝑡v\left(t\right) and generalized magnetic flux across an element are related by the lumped-element version of Faraday’s law of induction (Yurke1984, ; Devoret1995, ; feynmanLec-Vol-II, ; Pozar, ),

v(t)=ddtΦ(t)andΦ(t)=tv(τ)dτ+Φ().formulae-sequence𝑣𝑡dd𝑡Φ𝑡andΦ𝑡superscriptsubscript𝑡𝑣𝜏differential-d𝜏Φv\left(t\right)=\frac{\mathrm{d}}{\mathrm{d}t}\Phi\left(t\right)\quad\text{and}\quad\Phi\left(t\right)=\int_{-\infty}^{t}v\left(\tau\right)\,\mathrm{d}\tau+\Phi\left(-\infty\right)\;. (A.1)

By convention, the reference orientation of voltage is opposite to that of current. This is not the convention typically used in Lenz’s law; in electromagnetism, the current-density vector J𝐽\vec{J} and the electric-field vector E𝐸\vec{E} are usually projected on the same orientation.

From charge conservation, it follows that the time-instantaneous current i(t)𝑖𝑡i\left(t\right) and the charge Q(t)𝑄𝑡Q\left(t\right) having passed through the element, obey a relation similar to that of the voltage and flux,

i(t)=ddtQ(t)andQ(t)=ti(τ)dτ+Q().formulae-sequence𝑖𝑡dd𝑡𝑄𝑡and𝑄𝑡superscriptsubscript𝑡𝑖𝜏differential-d𝜏𝑄i\left(t\right)=\frac{\mathrm{d}}{\mathrm{d}t}Q\left(t\right)\quad\text{and}\quad Q\left(t\right)=\int_{-\infty}^{t}i\left(\tau\right)\,\mathrm{d}\tau+Q\left(-\infty\right)\;. (A.2)

All four variables have their support on a non-compact set; i.e., i,v,Q,Φ[,]𝑖𝑣𝑄Φi,\,v,\,Q,\,\Phi\in\left[-\infty,\infty\right].

Reference state and initial conditions.

In Eqs. (A.1) and (A.2), we now assume zero-valued initial conditions, Φ()=Q()=0Φ𝑄0\Phi\left(-\infty\right)=Q\left(-\infty\right)=0. In the case of a circuit frustrated by sources such that the equilibrium system state is non-zero, we can define the variables i,v,Q𝑖𝑣𝑄i,\,v,\,Q, and ΦΦ\Phi to denote deviations away from the global equilibrium of the circuit, as discussed in more detail in Sec. A8. By way of analogy, imagine a mechanical spring. The spring is stretched from its rest position to a new equilibrium by a second stretched spring. The deviations of the spring are measured from the equilibrium position of the combined system (this is what we mean by global equilibrium), not the spring in isolation.

Power and energy.

The instantaneous power p(t)𝑝𝑡p\left(t\right) delivered to an element and the total energy absorbed by the element (t)𝑡\mathcal{E}\left(t\right) are

ddt(t)=p(t)=v(t)i(t)and(t)=tp(t)dτ.formulae-sequencedd𝑡𝑡𝑝𝑡𝑣𝑡𝑖𝑡and𝑡superscriptsubscript𝑡𝑝𝑡differential-d𝜏\frac{\mathrm{d}}{\mathrm{d}t}\mathcal{E}\left(t\right)=p\left(t\right)=v\left(t\right)i\left(t\right)\quad\text{and}\quad\mathcal{E}\left(t\right)=\int_{-\infty}^{t}p\left(t\right)\,\mathrm{d}\tau\;. (A.3)

Given our convention for the orientation of i𝑖i and v𝑣v, power delivered to the element is positive if p(t)𝑝𝑡p\left(t\right) is positive. Passive elements (i.e., non-source elements) obey (t)()0𝑡0\mathcal{E}\left(t\right)-\mathcal{E}\left(-\infty\right)\geq 0 for all t𝑡t, and lossless elements convert all of their energy into stored electric or magnetic energy.

Capacitive and inductive elements.

A capacitive (resp., inductive) element is described by an algebraic relationship between charge and voltage (resp., flux and current). Let us introduce the simplest linear, passive, time-invariant elements. The simplest capacitor is defined by the constitutive relationship Q(t)=Cv(t)𝑄𝑡𝐶𝑣𝑡Q\left(t\right)=Cv\left(t\right), where C𝐶C is its capacitance—a positive, real constant. The dual relationship Φ(t)=Li(t)Φ𝑡𝐿𝑖𝑡\Phi\left(t\right)=Li\left(t\right) defines the simplest inductor, where L𝐿L is its inductance, also a positive, real constant. In terms of the flux across the element, the energies of the simple capacitor and inductor are

cap(Φ˙)=12CΦ˙2andind(Φ)=12LΦ2,formulae-sequencesubscriptcap˙Φ12𝐶superscript˙Φ2andsubscriptindΦ12𝐿superscriptΦ2\mathcal{E}_{\mathrm{cap}}\left(\dot{\Phi}\right)=\frac{1}{2}C\dot{\Phi}^{2}\quad\text{and}\quad\mathcal{E}_{\mathrm{ind}}\left(\Phi\right)=\frac{1}{2L}\Phi^{2}\;, (A.4)

respectively (and assuming Φ()=Q()=0Φ𝑄0\Phi\left(-\infty\right)=Q\left(-\infty\right)=0). Since C,L0𝐶𝐿0C,L\geq 0 and Φ(t)Φ𝑡\Phi\left(t\right)\in\mathbb{R} for all t𝑡t, we can verify that the total energy gained by these elements is always positive, as required for passive elements.

Josephson tunnel junction.

The chief non-linear element used in circuit quantum electrodynamics is the Josephson tunnel junction (Josephson1962, ; Devoret1995, ; Girvin2014, ), characterized by the flux-controlled inductive relationship i(t)=I0sin(Φ(t)/ϕ0)𝑖𝑡subscript𝐼0Φ𝑡subscriptitalic-ϕ0i\left(t\right)=I_{0}\sin\left(\Phi\left(t\right)/\phi_{0}\right), where ϕ0/2esubscriptitalic-ϕ0Planck-constant-over-2-pi2𝑒\phi_{0}\coloneqq\hbar/2e is the reduced magnetic flux quantum and I0subscript𝐼0I_{0} is the junction critical current. The energy function of the junction in terms of flux is

J(Φ)=EJ(1cos(Φ/ϕ0)),subscript𝐽Φsubscript𝐸𝐽1Φsubscriptitalic-ϕ0\mathcal{E}_{J}\left(\Phi\right)=E_{J}\left(1-\cos\left(\Phi/\phi_{0}\right)\right)\,, (A.5)

where EJI0ϕ0subscript𝐸𝐽subscript𝐼0subscriptitalic-ϕ0E_{J}\coloneqq I_{0}\phi_{0} denotes the Josephson energy. For small deviations about Φ=0Φ0\Phi=0, ignoring the constant energy offset,

J(Φ)12EJ(Φ/ϕ0)214!EJ(Φ/ϕ0)4+𝒪(Φ6).subscript𝐽Φ12subscript𝐸𝐽superscriptΦsubscriptitalic-ϕ0214subscript𝐸𝐽superscriptΦsubscriptitalic-ϕ04𝒪superscriptΦ6\mathcal{E}_{J}\left(\Phi\right)\approx\frac{1}{2}E_{J}\left(\Phi/\phi_{0}\right)^{2}-\frac{1}{4!}E_{J}\left(\Phi/\phi_{0}\right)^{4}+\mathcal{O}\left(\Phi^{6}\right)\;. (A.6)

To lowest order, the junction responds as a linear inductor with inductance LJEJ/ϕ02subscript𝐿𝐽subscript𝐸𝐽superscriptsubscriptitalic-ϕ02L_{J}\coloneqq E_{J}/\phi_{0}^{2} [compare this to Eq. (A.4)] It is useful to introduce the reduced magnetic flux φΦ/ϕ0𝜑Φsubscriptitalic-ϕ0\varphi\coloneqq\Phi/\phi_{0} of the junction.

Single- vs. multi-valued energy functions.

The Josephson junction exhibits a fundamental asymmetry. Inverting the current-flux relationship leads to a multi-valued function with an infinite number of branches; i.e., φ=sin1(i(t)/I0)+2πk𝜑superscript1𝑖𝑡subscript𝐼02𝜋𝑘\varphi=\sin^{-1}\left(i\left(t\right)/I_{0}\right)+2\pi k or φ=πsin1(i(t)/I0)+2πk𝜑𝜋superscript1𝑖𝑡subscript𝐼02𝜋𝑘\varphi=\pi-\sin^{-1}\left(i\left(t\right)/I_{0}\right)+2\pi k, where k𝑘k is some integer. For flux-controlled inductors, the multi-valued situation can be avoided by favoring a description in terms of flux, rather than charge.

Junction in a frustrated circuit.

Embedding the junction in a frustrated circuit can lead to a non-zero equilibrium value for φ𝜑\varphi. For deviations φφeq𝜑subscript𝜑eq\varphi-\varphi_{\mathrm{eq}} away from the equilibrium value φeqsubscript𝜑eq\varphi_{\mathrm{eq}}, see also Eq. (A.64),

J(φ)EJ[sin(φeq)(φφeq)+12cos(φeq)(φφeq)216sin(φeq)(φφeq)3]+𝒪(φ4).subscript𝐽𝜑subscript𝐸𝐽delimited-[]subscript𝜑eq𝜑subscript𝜑eq12subscript𝜑eqsuperscript𝜑subscript𝜑eq216subscript𝜑eqsuperscript𝜑subscript𝜑eq3𝒪superscript𝜑4\mathcal{E}_{J}\left(\varphi\right)\approx E_{J}\bigg{[}\sin\left(\varphi_{\mathrm{eq}}\right)\left(\varphi-\varphi_{\mathrm{eq}}\right)+\frac{1}{2}\cos\left(\varphi_{\mathrm{eq}}\right)\left(\varphi-\varphi_{\mathrm{eq}}\right)^{2}\\ -\frac{1}{6}\sin\left(\varphi_{\mathrm{eq}}\right)\left(\varphi-\varphi_{\mathrm{eq}}\right)^{3}\bigg{]}+\mathcal{O}\left(\varphi^{4}\right)\;. (A.7)

We can identify the differential inductance of the junction at φ=φeq𝜑subscript𝜑eq\varphi=\varphi_{\mathrm{eq}} to be LJ(φeq)=LJ/cos(φeq)subscript𝐿𝐽subscript𝜑eqsubscript𝐿𝐽subscript𝜑eqL_{J}\left(\varphi_{\mathrm{eq}}\right)=L_{J}/\cos\left(\varphi_{\mathrm{eq}}\right). In Sec. A8, we discuss how φ𝜑\varphi can be taken as a deviation away from the equilibrium value φeqsubscript𝜑eq\varphi_{\mathrm{eq}}, in effect mapping φφeqφmaps-to𝜑subscript𝜑eq𝜑\varphi-\varphi_{\mathrm{eq}}\mapsto\varphi, see Eq. (A.63). Also in Sec. A8, we discuss sources terms such as the one presented by EJsin(φeq)(φφeq)subscript𝐸𝐽subscript𝜑eq𝜑subscript𝜑eqE_{J}\sin\left(\varphi_{\mathrm{eq}}\right)\left(\varphi-\varphi_{\mathrm{eq}}\right). Such energy terms linear in φ𝜑\varphi turn the junction into an active component, capable of supplying current; i.e., we observe that the current relationship of the junction to lowest order I(t)=JΦ=I0sin(φeq)+LJ1(φeq)Φ(t)+𝒪(Φ2)𝐼𝑡subscript𝐽Φsubscript𝐼0subscript𝜑eqsuperscriptsubscript𝐿𝐽1subscript𝜑eqΦ𝑡𝒪superscriptΦ2I\left(t\right)=\frac{\partial\mathcal{E}_{J}}{\partial\Phi}=I_{0}\sin\left(\varphi_{\mathrm{eq}}\right)+L_{J}^{-1}\left(\varphi_{\mathrm{eq}}\right)\Phi\left(t\right)+\mathcal{O}\left(\Phi^{2}\right) contains the constant term I0sin(φeq)subscript𝐼0subscript𝜑eqI_{0}\sin\left(\varphi_{\mathrm{eq}}\right), which acts like a current source.

Relationship between the gauge-invariant superconducting phase difference of a tunnel junction and the reduced magnetic flux φ𝜑\varphi (compact vs. non-compact variables).

In superconductivity, the Josephson energy coupling two small superconducting islands has a cosine dependence on the gauge-invariant phase difference θjsubscript𝜃𝑗\theta_{j} of the superconducting phases of the two islands (tinkham2004-book, ; Clarke2004, ). This macroscopic variable is a phase angle—a compact variable in the half-open interval θ[0,2π[\theta\in[0,2\pi[; in contrast with the non-compact variable φ[,+]𝜑\varphi\in\left[-\infty,+\infty\right], which must be used in circuits where a superconducting wire connects the two sides of the junction.

In our treatment of the Josephson circuits so far, we have completely ignored this subtlety. Rather, we have based our discussion on the non-compact variable φ𝜑\varphi. The relationship between these two collective, macroscopic variables is θ=Φ/ϕ0mod2π𝜃moduloΦsubscriptitalic-ϕ02𝜋\theta=\Phi/\phi_{0}\mod 2\pi. Although the variable φ𝜑\varphi is non-compact, the associated wavefunction ψ(φ)𝜓𝜑\psi\left(\varphi\right) is submitted in practice to constraints (like confinement in one or a few potential wells) such that it is decomposed onto a basis set of wavefunctions that are indexed by a single discrete (rather than continuous) index; for example, the Fock basis ψn(φ)subscript𝜓𝑛𝜑\psi_{n}\left(\varphi\right), with n𝑛n\in\mathbb{N}. Quantum-mechanically, the representation of ψ(φ)𝜓𝜑\psi\left(\varphi\right) and ψ(θ)𝜓𝜃\psi\left(\theta\right) is therefore not very different since ψ(θ)𝜓𝜃\psi\left(\theta\right) is represented by the discrete rotor basis, ψk(θ)subscript𝜓𝑘𝜃\psi_{k}\left(\theta\right), with k𝑘k\in\mathbb{Z}.

A broken symmetry.

The compact support of θ𝜃\theta corresponds to a symmetry that is usually broken in most circuits—that of the impossibility to distinguish between different values of φ𝜑\varphi differing by 2π2𝜋2\pi. Losses associated with the junction, or coupling to other elements, such as in the RF-SQUID or fluxonium qubit (Manucharyan2009, ), render 2π2𝜋2\pi turns of φ𝜑\varphi macroscopically distinguishable—hence demanding a description in terms of non-compact support corresponding to a point on an open-ended line rather than a circle. For certain special cases, however, it may be advantageous to retain the compact support version. But even in such cases, one can start with the non-compact version of φ𝜑\varphi and recover the compact version by a limit procedure (Koch2007, ). Thus, in this article, we use only gauge-invariant phases with non-compact support; i.e., φ[,+]𝜑\varphi\in\left[-\infty,+\infty\right].

Flux-controlled inductor.

To generalize from the Josephson junction and introduce more general non-linear elements, consider the flux-controlled inductor. It is defined by an algebraic relationship i(t)=h(Φ(t),t)𝑖𝑡Φ𝑡𝑡i\left(t\right)=h\left(\Phi\left(t\right),t\right), where hh is a single-valued function. If hh is time-invariant, such as for the case of the Josephson tunnel junction, the energy function [see Eq. (A.3)] is

ind(Φ)=Φ()Φ(t)h(Φ)dΦ.subscriptindΦsuperscriptsubscriptΦΦ𝑡superscriptΦdifferential-dsuperscriptΦ\mathcal{E}_{\mathrm{ind}}\left(\Phi\right)=\int_{\Phi\left(-\infty\right)}^{\Phi\left(t\right)}h\left(\Phi^{\prime}\right)\,\mathrm{d}\Phi^{\prime}\;. (A.8)

Similar results can be obtained for current controlled inductors and for generalized capacitors.

A network of circuit elements.

An electrical circuit is an interconnected collection of circuit elements. The connectivity of the elements can be described by an oriented graph. Each branch in the graph corresponds to one element. The b𝑏b-th element is associated with the instantaneous voltage vb(t)subscript𝑣𝑏𝑡v_{b}\left(t\right), current ib(t)subscript𝑖𝑏𝑡i_{b}\left(t\right), charge Qb(t)subscript𝑄𝑏𝑡Q_{b}\left(t\right), flux Φb(t)subscriptΦ𝑏𝑡\Phi_{b}\left(t\right), and reduced flux φb(t)subscript𝜑𝑏𝑡\varphi_{b}\left(t\right). The universal relationships Eqs. (A.1)–(A.3) link the variables at each branch. Variables across different branches are linked by Kirchhoff’s laws.

Kirchhoff’s two universal circuit laws.

The following two laws are universal and topological in nature. They describe relationship among the branch variables, independent of the constitution of the branch elements. In other words, they apply to nonlinear, time-dependent, and even hysteric elements.

Kirchhoff’s voltage law (KVL) is the lumped-element manifestation of the Maxwell-Faraday equation, ×E=Bt𝐸𝐵𝑡\nabla\times\vec{E}=-\frac{\partial\vec{B}}{\partial t}. By applying Stokes’ theorem to the Maxwell-Faraday equation along an oriented loop of lumped elements (and a surface associated with the closed loop), one finds a relationship valid for any closed circuit loop: the oriented sum of the fluxes along the l𝑙l-th loop is equal to the external applied flux Φlext(t)superscriptsubscriptΦ𝑙ext𝑡\Phi_{l}^{\mathrm{ext}}\left(t\right) threading the loop,

bloopl±Φb(t)=Φlext(t),plus-or-minussubscript𝑏subscriptloop𝑙subscriptΦ𝑏𝑡superscriptsubscriptΦ𝑙ext𝑡\sum_{b\in\mathrm{loop}_{l}}\pm\Phi_{b}(t)=\Phi_{l}^{\mathrm{ext}}\left(t\right)\;, (A.9)

where the sum runs over all branches b𝑏b that form the l𝑙l-th loop. For a given branch b𝑏b, the positive (resp., negative) sign in Eq. (A.9) is selected if its flux reference direction aligns (resp., is opposite to) the loop orientation. Algebraically, KVL leads to a set of constraints among the network variables. Thus, in the context of a Lagrangian description of the circuit in which the generalized position variables are taken to be fluxes ΦbsubscriptΦ𝑏\Phi_{b}, the KVL conditions express a set of holonomic constraints that need to be eliminated in order to obtain a Lagrangian of the second kind (Landau1982, ).

Kirchhoff’s current law (KCL) is a statement of the conservation of charge: at every node in the circuit, the algebraic sum of all the current leaving or entering the node is equal to zero. Recast another way, for all branches b𝑏b connected to node n𝑛n,

bnoden±ib(t)=0.plus-or-minussubscript𝑏subscriptnode𝑛subscript𝑖𝑏𝑡0\sum_{b\in\mathrm{node}_{n}}\pm i_{b}(t)=0\;. (A.10)

The negative sign is chosen for branches whose current reference direction points toward the n𝑛n-th node. In the flux-based Lagrangian description of the circuit, the KCL algebraic conditions become the Lagrangian equations of motion.

Eliminating the KVL constraints using the method of the minimum spanning-tree.

The set of KVL algebraic equations defined in Eq. (A.9) reduce the number of independent branch fluxes ΦbsubscriptΦ𝑏\Phi_{b}. We can systematically choose a minimal set of independent branch fluxes using the minimum spanning-tree graph method (Devoret1995, ; Burkard2004, ; Girvin2014, ; Vool2017, ). For our derivation, it is not necessary to explicitly construct the tree. A set of branches from the graph can be selected to form a complete and minimum spanning tree. In general, there are many satisfactory tree sets of branches. Different trees are related by a simple algebraic transformation; similar to a basis change. The branches that belong to the spanning tree can be labeled t1,t2,subscriptt1subscriptt2\mathrm{t}_{1},\,\mathrm{t}_{2},\,\ldots The flux of the k𝑘k-th spanning-tree branch is denoted Φtk(t)subscriptΦsubscriptt𝑘𝑡\Phi_{\mathrm{t}_{k}}\left(t\right). In subscripts, roman (resp., italic) symbols denote labels (resp., variables). The spanning-tree can be organized in a column vector

𝚽t(t)(Φt1(t)Φt2(t)),subscript𝚽t𝑡matrixsubscriptΦsubscript𝑡1𝑡subscriptΦsubscript𝑡2𝑡\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\left(t\right)\coloneqq\begin{pmatrix}\Phi_{t_{1}}\left(t\right)\\ \Phi_{t_{2}}\left(t\right)\\ \vdots\end{pmatrix}\;, (A.11)

which serves the purposes of a basis for the description of the circuit. The branch-fluxes not in the spanning tree (links or chords of the graph) are obtained by a linear transformation of 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}. We define the energy functions and Lagrangian of the system in terms of 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} in the Sec. A3.

A2 The Josephson system and its non-linear Josephson dipoles

In the main text, under Quantizing the general Josephson system, we introduced the notion of a Josephson system—a general electromagnetic environment that incorporates nonlinear devices, referred to as Josephson dipoles. The Josephson system is treated as a distributed black-box structure.

Discretization.

We aim to model the Josephson system as realistically as possible. To account in detail for its physical layout, materials, boundary conditions, and dipole structures, we aim to leverage conventional electromagnetic analysis techniques, such as the finite-element (FE) method. The FE method subdivides the physical layout of the system. Using a set of basis functions, the electromagnetic circuit is discretized (Louisell1973, ; Jin2014, ). The discretized circuit can be represented by a lumped-element model. In principle, we can take the limit of infinite subdivision. The Josephson dipoles are assumed to be the only non-linear elements in the circuit. All other elements are linear, as representative of the linear nature of Maxwell’s equations.

Dissipation.

As the object of interest is the control of quantum information in the system, in this section, we focus on systems with low dissipation. This condition requires that the quality factor of all modes of relevance is high, Qm1much-greater-thansubscript𝑄𝑚1Q_{m}\gg 1, where m𝑚m is the mode index. In Sec. D, we treat dissipation as a perturbation to the lossless solutions.

Josephson dipole.

The Josephson dipole was introduced in the main text, under Quantizing the general Josephson system. For simplicity of discussion, here, we treat the dipole as a lumped, two-terminal, flux-controlled element. The j𝑗j-th Josepson dipole in the system is fully specified by its energy function j(Φj;Φj,ext)subscript𝑗subscriptΦ𝑗subscriptΦ𝑗ext\mathcal{E}_{j}\left(\Phi_{j};\Phi_{j,\mathrm{ext}}\right), see Eq. (A.8). The energy depends only on the magnetic flux Φj(t)subscriptΦ𝑗𝑡\Phi_{j}\left(t\right) across the terminals of the dipole and on any external parameters Φj,extsubscriptΦ𝑗ext\Phi_{j,\mathrm{ext}} that control the energy landscape; these can include a voltage bias, a current bias, and an external magnetic field bias. The index j𝑗j runs from unit to the total number J𝐽J of Josephson dipoles in the circuit.

This formulation is rather general. It encapsulate the wide span of Josephson dipoles discussed in the main text. These include simple devices, such as the Josephson tunnel junction or a nanowire, and also composite devices, such as SQUIDs, SNAILs, or more general sub-circuits. The underlying physical phenomenon giving rise to the low-loss, non-linearity of the dipole is immaterial.

Partition of the Josephson dipole energy-function.

It is always possible to partition j(Φj;Φj,ext)subscript𝑗subscriptΦ𝑗subscriptΦ𝑗ext\mathcal{E}_{j}\left(\Phi_{j};\Phi_{j,\mathrm{ext}}\right) into a linear and non-linear part,

j(Φj;Φj,ext)=jlin(Φj;Φj,ext)+jnl(Φj;Φj,ext).subscript𝑗subscriptΦ𝑗subscriptΦ𝑗extsuperscriptsubscript𝑗linsubscriptΦ𝑗subscriptΦ𝑗extsuperscriptsubscript𝑗nlsubscriptΦ𝑗subscriptΦ𝑗ext\mathcal{E}_{j}\left(\Phi_{j};\Phi_{j,\mathrm{ext}}\right)=\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j};\Phi_{j,\mathrm{ext}}\right)+\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j};\Phi_{j,\mathrm{ext}}\right)\;. (A.12)

This division is purely a conceptual one—the Josephson dipole cannot be physically divided into a linear and nonlinear part. By selecting an equilibrium point of the circuit and defining the branch fluxes ΦjsubscriptΦ𝑗\Phi_{j} as deviations away from the equilibrium [discussed in more detail in Sec. (A8)], the partitions take the concrete form

jlin(Φj)superscriptsubscript𝑗linsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) \displaystyle\coloneqq 12Ej(Φj/ϕ0)2,12subscript𝐸𝑗superscriptsubscriptΦ𝑗subscriptitalic-ϕ02\displaystyle\frac{1}{2}E_{j}\left(\Phi_{j}/\phi_{0}\right)^{2}\;, (A.13a)
jnl(Φj)superscriptsubscript𝑗nlsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right) \displaystyle\coloneqq j(Φj)jlin(Φj)subscript𝑗subscriptΦ𝑗superscriptsubscript𝑗linsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}\left(\Phi_{j}\right)-\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) (A.13b)
=\displaystyle= Ejp=3cjp(Φj/ϕ0)p,subscript𝐸𝑗superscriptsubscript𝑝3subscript𝑐𝑗𝑝superscriptsubscriptΦ𝑗subscriptitalic-ϕ0𝑝\displaystyle E_{j}\sum_{p=3}^{\infty}c_{jp}\left(\Phi_{j}/\phi_{0}\right)^{p}\;, (A.13c)

where Ejsubscript𝐸𝑗E_{j} is an overall scaling factor of the energy function, defined in Eq. (A.13a); we refer to it as the Josephson dipole energy scale. In Eq. (A.13c), we have introduced the Taylor series expansion of jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} around the equilibrium state of the circuit and have introduced the dimensionless expansion coefficients cjpsubscript𝑐𝑗𝑝c_{jp}. We stress that the expansion is not needed for the EPR method. It is merely a convenient tool for working analytically with weakly non-linear circuits, such as the transmon qubit. For notation simplicity, in Eq. (A.13) the dependance of jlin,jnl,Ejsuperscriptsubscript𝑗linsuperscriptsubscript𝑗nlsubscript𝐸𝑗\mathcal{E}_{j}^{\mathrm{lin}},\mathcal{E}_{j}^{\mathrm{nl}},E_{j}, and cjpsubscript𝑐𝑗𝑝c_{jp} on the external bias parameter Φj,extsubscriptΦ𝑗ext\Phi_{j,\mathrm{ext}} is made implicit, and we will continue to do so henceforth; in other words, keep in mind that EjEj(Φj,ext)subscript𝐸𝑗subscript𝐸𝑗subscriptΦ𝑗extE_{j}\coloneqq E_{j}\left(\Phi_{j,\mathrm{ext}}\right) and cjpcjp(Φj,ext)subscript𝑐𝑗𝑝subscript𝑐𝑗𝑝subscriptΦ𝑗extc_{jp}\coloneqq c_{jp}\left(\Phi_{j,\mathrm{ext}}\right).

Example of a Josephson dipole: the Josephson junction.

We illustrate the partitioning construction defined in Eq. (A.13) using the example of the Josephson tunnel junction. For an un-frustrated junction, it follows from Eq. (A.5) that

jlin(Φj)superscriptsubscript𝑗linsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right) \displaystyle\coloneqq 12Ej(Φj/ϕ0)2,12subscript𝐸𝑗superscriptsubscriptΦ𝑗subscriptitalic-ϕ02\displaystyle\frac{1}{2}E_{j}\left(\Phi_{j}/\phi_{0}\right)^{2}\;, (A.14a)
jnl(Φj)superscriptsubscript𝑗nlsubscriptΦ𝑗\displaystyle\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right) \displaystyle\coloneqq Ej[cos(Φj/ϕ0)+12(Φj/ϕ0)2],subscript𝐸𝑗delimited-[]subscriptΦ𝑗subscriptitalic-ϕ012superscriptsubscriptΦ𝑗subscriptitalic-ϕ02\displaystyle-E_{j}\left[\cos\left(\Phi_{j}/\phi_{0}\right)+\frac{1}{2}\left(\Phi_{j}/\phi_{0}\right)^{2}\right]\;, (A.14b)

where Ejsubscript𝐸𝑗E_{j} is the Josephson energy. The energy function jlinsuperscriptsubscript𝑗lin\mathcal{E}_{j}^{\mathrm{lin}} is associated with the linear response of the junction. It presents the inductance Lj=ϕ02/Ejsubscript𝐿𝑗superscriptsubscriptitalic-ϕ02subscript𝐸𝑗L_{j}=\phi_{0}^{2}/E_{j}. The energy jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} is associated with the response of non-linear inductor. The expansion coefficient of jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} as defined in Eq. (A.13c) are

cjp={(1)p/2+1p!for even p,0for odd p.subscript𝑐𝑗𝑝casessuperscript1𝑝21𝑝for even 𝑝0for odd 𝑝c_{jp}=\begin{cases}\frac{\left(-1\right)^{p/2+1}}{p!}&\text{for even }p\;,\\ 0&\text{for odd }p\;.\end{cases} (A.15)

In partitioning jsubscript𝑗\mathcal{E}_{j}, we included all of the linear response of the junction in jlinsuperscriptsubscript𝑗lin\mathcal{E}_{j}^{\mathrm{lin}}, leaving none for jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}}; i.e., jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} lacks quadratic terms in ΦjsubscriptΦ𝑗\Phi_{j}. However, this is not required. There are certain cases for which retaining some part of the linear response in jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}} is advantageous.

For notational ease, we now introduce the reduced flux φj(t)Φ(t)/ϕ0subscript𝜑𝑗𝑡Φ𝑡subscriptitalic-ϕ0\varphi_{j}\left(t\right)\coloneqq\Phi\left(t\right)/\phi_{0}.

Example of a Josephson dipole in a frustrated circuit.

Imagine a Josephson tunnel junction incorporated in a closed loop of several circuit elements. Suppose the loop supports the flow of a direct current. A current source in the path of the loop (or perhaps an external magnetic flux threading the loop) establishes a persistent current. The equilibrium flux φjsubscript𝜑𝑗\varphi_{j} of the junction shifts to a non-zero equilibrium value φeq,jsubscript𝜑eq𝑗\varphi_{\mathrm{eq},j}, as determined by the circuit equilibrium considerations (see Sec. A8). Applying the partition defined in Eq. (A.14) to Eq. (A.7), we find that in terms of the out-of-equilibrium flux deviation φjsubscript𝜑𝑗\varphi_{j},

j(φj;φeq,j)=12Ej(φeq,j)φj2+Ej(φeq,j)cj3(φeq,j)φj3+,subscript𝑗subscript𝜑𝑗subscript𝜑eq𝑗12subscript𝐸𝑗subscript𝜑eq𝑗superscriptsubscript𝜑𝑗2subscript𝐸𝑗subscript𝜑eq𝑗subscript𝑐𝑗3subscript𝜑eq𝑗superscriptsubscript𝜑𝑗3\mathcal{E}_{j}\left(\varphi_{j};\varphi_{\mathrm{eq},j}\right)=\frac{1}{2}E_{j}\left(\varphi_{\mathrm{eq},j}\right)\varphi_{j}^{2}\\ +E_{j}\left(\varphi_{\mathrm{eq},j}\right)c_{j3}\left(\varphi_{\mathrm{eq},j}\right)\varphi_{j}^{3}+\cdots\;, (A.16)

where Ej(φeq,j)=EJ/cos(φeq,j)subscript𝐸𝑗subscript𝜑eq𝑗subscript𝐸𝐽subscript𝜑eq𝑗E_{j}\left(\varphi_{\mathrm{eq},j}\right)=E_{J}/\cos\left(\varphi_{\mathrm{eq},j}\right), EJsubscript𝐸𝐽E_{J} is the j𝑗j-th junction Josephson energy, and cj3(φeq,j)=16sin(φeq,j)subscript𝑐𝑗3subscript𝜑eq𝑗16subscript𝜑eq𝑗c_{j3}\left(\varphi_{\mathrm{eq},j}\right)=-\frac{1}{6}\sin\left(\varphi_{\mathrm{eq},j}\right). We emphasize that φjsubscript𝜑𝑗\varphi_{j} denotes deviations away from the equilibrium; compare Eq. (A.16) to Eq. (A.7).

A3 Energy of the Josephson circuit and its Lagrangian

Capacitive energy.

The total capacitive energy capsubscriptcap\mathcal{E}_{\mathrm{cap}} of the Josephson system is simply the algebraic sum of the total energy of all its capacitive elements. Using Eq. (A.4), and summing over all capacitive branches, capbcap.12CbΦ˙b2subscriptcapsubscript𝑏cap12subscript𝐶𝑏superscriptsubscript˙Φ𝑏2\mathcal{E}_{\mathrm{cap}}\coloneqq\sum_{b\in\mathrm{cap.}}\frac{1}{2}C_{b}\dot{\Phi}_{b}^{2}, where Cbsubscript𝐶𝑏C_{b} is the capacitance of branch b𝑏b. Each of these fluxes can be expressed in terms of the linearly-independent spanning-tree fluxes 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}. The energy function is quadratic in 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}},

cap(𝚽˙t)=12𝚽˙tT𝐂𝚽˙t,subscriptcapsubscript˙𝚽t12superscriptsubscript˙𝚽tT𝐂subscript˙𝚽t\mathcal{E}_{\mathrm{cap}}\left(\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right)=\frac{1}{2}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{C}}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\;, (A.17)

where 𝐂𝐂\mathrm{\mathbf{C}} is the capacitance matrix of the circuit (Yurke1984, ; Devoret1995, ; Girvin2014, ). It follows from KVL and the constitutive relationships of the capacitors that 𝐂𝐂\mathrm{\mathbf{C}} is a positive-definite, real, symmetric (PDRS) matrix. In the continuous limit of space, the total capacitive energy capsubscriptcap\mathcal{E}_{\mathrm{cap}} can be found using Eq. (C.3).

Inductive energy.

The total inductive energy in the circuit indsubscriptind\mathcal{E}_{\mathrm{ind}} is similarly the algebraic sum of the total energy of all circuit inductive branches; i.e., indbind.b(Φb)subscriptindsubscript𝑏indsubscript𝑏subscriptΦ𝑏\mathcal{E}_{\mathrm{ind}}\coloneqq\sum_{b\in\mathrm{ind.}}\mathcal{E}_{b}\left(\Phi_{b}\right). In the Josephson system, inductive branches come in two distinct flavors: linear and non-linear. Physically, linear inductive branches are associated with the geometry and magnetic fields. We denote their total energy magsubscriptmag\mathcal{E}_{\mathrm{mag}}. The non-linear inductive branches (Josephson dipoles) are generally associated with the kinetic inductance of electrons. We denote their total energy kinsubscriptkin\mathcal{E}_{\mathrm{kin}}. Hence,

ind(𝚽t)=mag(𝚽t)+kin(𝚽t).subscriptindsubscript𝚽tsubscriptmagsubscript𝚽tsubscriptkinsubscript𝚽t\mathcal{E}_{\mathrm{ind}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)=\mathcal{E}_{\mathrm{mag}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)+\mathcal{E}_{\mathrm{kin}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)\;. (A.18)

The energy of the linear branches magsubscriptmag\mathcal{E}_{\mathrm{mag}} is the dual of Eq. (A.17). It is also a quadratic form,

mag(𝚽t)=12𝚽tT𝐋𝚽tmag1,subscriptmagsubscript𝚽t12superscriptsubscript𝚽tT𝐋subscriptsuperscriptsubscript𝚽t1mag\mathcal{E}_{\mathrm{mag}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)=\frac{1}{2}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{L}}{}_{\text{mag}}^{-1}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\;, (A.19)

where the inductance matrix 𝐋mag1\mathrm{\mathbf{L}}{}_{\text{mag}}^{-1} completely describes all linear, magnetic-in-origin inductances in the circuit (Yurke1984, ; Devoret1995, ; Girvin2014, ). Due to its nature, 𝐋mag1\mathrm{\mathbf{L}}{}_{\text{mag}}^{-1} is PDRS. In the continuous limit of space, the total magnetic inductive energy magsubscriptmag\mathcal{E}_{\mathrm{mag}} can be found using Eq. (C.4).

The total inductive kinetic energy of the circuit, associated with the non-linear dipoles, is

kin=j=1Jj(Φj)=j=1Jjlin(Φj)+j=1Jjnl(Φj).subscriptkinsuperscriptsubscript𝑗1𝐽subscript𝑗subscriptΦ𝑗superscriptsubscript𝑗1𝐽superscriptsubscript𝑗linsubscriptΦ𝑗superscriptsubscript𝑗1𝐽superscriptsubscript𝑗nlsubscriptΦ𝑗\mathcal{E}_{\mathrm{kin}}=\sum_{j=1}^{J}\mathcal{E}_{j}\left(\Phi_{j}\right)=\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{lin}}\left(\Phi_{j}\right)+\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right)\;. (A.20)

This energy is not stored in the magnetic fields. However, we can group the magnetic energy and the linear part of kinsubscriptkin\mathcal{E}_{\mathrm{kin}} together to express the inductive energy as a partition of linear and non-linear contributions

ind(𝚽t)subscriptindsubscript𝚽t\displaystyle\mathcal{E}_{\mathrm{ind}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right) =12𝚽tT𝐋𝚽t1+kin(𝚽t),absent12superscriptsubscript𝚽tT𝐋superscriptsubscript𝚽t1subscriptkinsubscript𝚽t\displaystyle=\frac{1}{2}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{L}}{}^{-1}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}+\mathcal{E}_{\mathrm{kin}}(\mathrm{\mathbf{\Phi}}_{\mathrm{t}})\;, (A.21a)
𝐋1\displaystyle\mathrm{\mathbf{L}}{}^{-1} 𝐋+mag112j=1JEj(Φj/ϕ0)2,\displaystyle\coloneqq\mathrm{\mathbf{L}}{}_{\text{mag}}^{-1}+\frac{1}{2}\sum_{j=1}^{J}E_{j}\left(\Phi_{j}/\phi_{0}\right)^{2}\;, (A.21b)
nl(𝚽t)subscriptnlsubscript𝚽t\displaystyle\mathcal{E}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right) j=1Jjnl(Φj),absentsuperscriptsubscript𝑗1𝐽superscriptsubscript𝑗nlsubscriptΦ𝑗\displaystyle\coloneqq\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right)\;, (A.21c)

where we have introduced the total inductance matrix of the circuit 𝐋1\mathrm{\mathbf{L}}{}^{-1} and the total non-linear energy function of the circuit nlsubscriptnl\mathcal{E}_{\mathrm{nl}}. For later use, we can obtain the series expansion of nlsubscriptnl\mathcal{E}_{\mathrm{nl}} in terms of that of jnlsuperscriptsubscript𝑗nl\mathcal{E}_{j}^{\mathrm{nl}}, defined in Eq. (A.13b),

nl(𝚽t)=j=1Jp=3Ejcjp(Φj/ϕ0)p.subscriptnlsubscript𝚽tsuperscriptsubscript𝑗1𝐽superscriptsubscript𝑝3subscript𝐸𝑗subscript𝑐𝑗𝑝superscriptsubscriptΦ𝑗subscriptitalic-ϕ0𝑝\mathcal{E}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)=\sum_{j=1}^{J}\sum_{p=3}^{\infty}E_{j}c_{jp}\left(\Phi_{j}/\phi_{0}\right)^{p}\;. (A.22)
Generalized coordinates for the Lagrangian.

We have expressed capsubscriptcap\mathcal{E}_{\mathrm{cap}} and indsubscriptind\mathcal{E}_{\mathrm{ind}} in terms of the independent set of spanning-tree fluxes 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}. We could equivalently have expressed capsubscriptcap\mathcal{E}_{\mathrm{cap}} and indsubscriptind\mathcal{E}_{\mathrm{ind}} in terms of the charge variables Qbsubscript𝑄𝑏Q_{b}. However, as discussed in Sec. A1, the flux-controlled Josephson dipoles present an asymmetry which favors treatment in the flux basis. We hence employ 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} as the generalized position coordinates in the Lagrangian description of the circuit, and 𝚽˙tsubscript˙𝚽t\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}} as the generalized velocity.

Lagrangian of the Josephson circuit.

The Lagrangian function of the Josephson circuit follows from KCL. Energy functions with 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} as their argument (resp., 𝚽˙tsubscript˙𝚽t\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}) play the role of potential (resp., kinetic) energies. The system Lagrangian is the difference of the total kinetic and potential energy functions,

full(𝚽t,𝚽˙t)subscriptfullsubscript𝚽tsubscript˙𝚽t\displaystyle\mathcal{L}_{\mathrm{full}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}},\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right) cap(𝚽˙t)ind(𝚽t)absentsubscriptcapsubscript˙𝚽tsubscriptindsubscript𝚽t\displaystyle\coloneqq\mathcal{E}_{\mathrm{cap}}\left(\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right)-\mathcal{E}_{\mathrm{ind}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right) (A.23a)
=lin(𝚽t,𝚽˙t)+nl(𝚽t),absentsubscriptlinsubscript𝚽tsubscript˙𝚽tsubscriptnlsubscript𝚽t\displaystyle=\mathcal{\mathcal{L}_{\mathrm{lin}}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}},\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right)+\mathcal{L}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)\;, (A.23b)

where we have partitioned the Lagrangian into a linear linsubscriptlin\mathcal{L}_{\mathrm{lin}} and nonlinear nlsubscriptnl\mathcal{L}_{\mathrm{nl}} part. Substituting in Eqs. (A.17) and (A.21),

lin(𝚽t,𝚽˙t)subscriptlinsubscript𝚽tsubscript˙𝚽t\displaystyle\mathcal{L}_{\mathrm{lin}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}},\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right) =\displaystyle= 12𝚽˙tT𝐂𝚽˙t12𝚽tT𝐋1𝚽t,12superscriptsubscript˙𝚽tT𝐂subscript˙𝚽t12superscriptsubscript𝚽tTsuperscript𝐋1subscript𝚽t\displaystyle\frac{1}{2}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{C}}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}-\frac{1}{2}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{L}}^{-1}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\;, (A.24a)
nl(𝚽t)subscriptnlsubscript𝚽t\displaystyle\mathcal{L}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right) =\displaystyle= nl(𝚽t)=j=1Jjnl(Φj).subscriptnlsubscript𝚽tsuperscriptsubscript𝑗1𝐽superscriptsubscript𝑗nlsubscriptΦ𝑗\displaystyle-\mathcal{E}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)=-\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{nl}}\left(\Phi_{j}\right)\;. (A.24b)

Equation (A.24) explicitly constructs the Josephson system Lagrangian. It partitions it into a linear and nonlinear part. In Sec. A4, we diagonalize linsubscriptlin\mathcal{L}_{\mathrm{lin}} to find the eigenmodes of the linearized system. In Sec. B, we treat the effect of nlsubscriptnl\mathcal{L}_{\mathrm{nl}}.

A4 Eigenmodes of the linearized Josephson circuit

In this sub-sectino, we diagonalize linsubscriptlin\mathcal{L}_{\mathrm{lin}} and find its eigenmodes, eigenfrequencies ωmsubscript𝜔𝑚\omega_{m} and eigenvectors (i.e., spatial-mode profiles). These intermediate result provide a key stepping stone on our path to quantizing the Josephson system and treating nlsubscriptnl\mathcal{L}_{\mathrm{nl}}. The process conceptually parallels that taken by the finite-element (FE) electromagnetic (EM) solver in an eigenanalysis of the linearized Josephson system.

The Lagrangian linsubscriptlin\mathcal{L}_{\mathrm{lin}} is the sum of two quadratic forms, see Eq. (A.24). We use the standard method for their simultaneous diagonalization (Landau1982, ), based on a series of principle-axis transforms. We then transform the Lagrangian into a diagonalized Hamiltonian.

Diagonalizing the inductance matrix.

Since the inverse inductance matrix 𝐋1superscript𝐋1\mathrm{\mathbf{L}}^{-1} is a PDRS matrix, we can diagonalize it with a real orthogonal matrix 𝐎𝐋subscript𝐎𝐋\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}, obeying 𝐎𝐋𝐎𝐋=𝐎𝐋𝐎𝐋=𝐈subscript𝐎𝐋superscriptsubscript𝐎𝐋superscriptsubscript𝐎𝐋subscript𝐎𝐋𝐈\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\intercal}=\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\intercal}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}=\mathrm{\mathbf{I}}, where 𝐈𝐈\mathrm{\mathbf{I}} is the identity matrix,

𝐎𝐋T𝐋1𝐎𝐋=𝚲𝐋1𝐈𝐋1,superscriptsubscript𝐎𝐋Tsuperscript𝐋1subscript𝐎𝐋superscriptsubscript𝚲𝐋1superscriptsubscript𝐈𝐋1\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\mathrm{T}}\mathrm{\mathbf{L}}^{-1}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}=\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{-1}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{L}}}^{-1}\;, (A.25)

where 𝚲𝐋1superscriptsubscript𝚲𝐋1\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{-1} is a diagonal matrix comprising the (dimensionless) eigenvalue magnitudes, and 𝐈𝐋subscript𝐈𝐋\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathrm{\mathbf{L}}}} is the identity matrix with physical dimensions of inductance. The eigenvectors of 𝐋1superscript𝐋1\mathrm{\mathbf{L}}^{-1} form the columns of 𝐎𝐋subscript𝐎𝐋\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}.

Employing Eq. (A.25) with Eq. (A.24), the system Lagrangian takes the suggestive form

full(𝚽t,𝚽˙t)=12𝚽˙tT𝐂𝚽˙t+nl(𝚽t)12(𝚽tT𝐎𝐋𝚲𝐋1/2)𝐈𝐋1(𝚲𝐋1/2𝐎𝐋T𝚽t),subscriptfullsubscript𝚽tsubscript˙𝚽t12superscriptsubscript˙𝚽tT𝐂subscript˙𝚽tsubscriptnlsubscript𝚽t12superscriptsubscript𝚽tTsubscript𝐎𝐋superscriptsubscript𝚲𝐋12superscriptsubscript𝐈𝐋1superscriptsubscript𝚲𝐋12superscriptsubscript𝐎𝐋Tsubscript𝚽t\mathcal{L}_{\mathrm{full}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}},\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}\right)=\frac{1}{2}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{C}}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{t}}+\mathcal{L}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)\\ -\frac{1}{2}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}^{\mathrm{T}}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{-1/2}\right)\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathrm{\mathbf{L}}}}^{-1}\left(\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{-1/2}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\mathrm{T}}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right)\;, (A.26)

which motivates the principle-axis transformation of the magnetic flux defined by

𝚽˘𝚲𝐋1/2𝐎𝐋T𝚽t,˘𝚽superscriptsubscript𝚲𝐋12superscriptsubscript𝐎𝐋Tsubscript𝚽t\breve{\mathrm{\mathbf{\Phi}}}\coloneqq\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{-1/2}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\mathrm{T}}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\;, (A.27)

where 𝚽˘˘𝚽\breve{\mathrm{\mathbf{\Phi}}} is a rotated and then scaled version of 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}. Under this transformation, the transformed Lagrangian function becomes

˘full(𝚽˘,𝚽˘˙)12𝚽˘˙T𝐂˘𝚽˘˙12𝚽˘T𝐈𝐋1𝚽˘+˘nl(𝚽˘),subscript˘full˘𝚽˙˘𝚽12superscript˙˘𝚽T˘𝐂˙˘𝚽12superscript˘𝚽Tsuperscriptsubscript𝐈𝐋1˘𝚽subscript˘nl˘𝚽\begin{split}\breve{\mathcal{L}}_{\mathrm{full}}\left(\breve{\mathrm{\mathbf{\Phi}}},\dot{\breve{\mathrm{\mathbf{\Phi}}}}\right)\coloneqq\frac{1}{2}\dot{\breve{\mathrm{\mathbf{\Phi}}}}^{\mathrm{T}}\mathrm{\mathbf{\breve{C}}}\dot{\breve{\mathrm{\mathbf{\Phi}}}}-\frac{1}{2}\breve{\mathrm{\mathbf{\Phi}}}^{\mathrm{T}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathrm{\mathbf{L}}}}^{-1}\breve{\mathrm{\mathbf{\Phi}}}+\breve{\mathcal{L}}_{\mathrm{nl}}\left(\breve{\mathrm{\mathbf{\Phi}}}\right)\;,\end{split} (A.28)

where ˘nl(𝚽˘)nl(𝚽(𝚽˘))subscript˘nl˘𝚽subscriptnl𝚽˘𝚽\breve{\mathcal{L}}_{\mathrm{nl}}\left(\breve{\mathrm{\mathbf{\Phi}}}\right)\coloneqq\mathcal{L}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}\left(\breve{\mathrm{\mathbf{\Phi}}}\right)\right) and

𝐂˘(𝚲𝐋1/2𝐎𝐋T)𝐂(𝐎𝐋𝚲𝐋1/2).˘𝐂superscriptsubscript𝚲𝐋12superscriptsubscript𝐎𝐋T𝐂subscript𝐎𝐋superscriptsubscript𝚲𝐋12\mathrm{\mathbf{\breve{C}}}\coloneqq\left(\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{1/2}\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}^{\mathrm{T}}\right)\mathrm{\mathbf{C}}\left(\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{1/2}\right)\;. (A.29)

Since the capacitance matrix 𝐂𝐂\mathrm{\mathbf{C}} is PDRS and it is transformed by a rotation and then a dilation, it follows that 𝐂˘˘𝐂\mathrm{\mathbf{\breve{C}}} is also PDRS. More generally, the eigenvalues of a matrix are invariant under a similarity transform, such as the one employed in Eq. (A.29).

Diagonalizing the capacitance matrix.

Since 𝐂˘˘𝐂\mathrm{\mathbf{\breve{C}}} is PDRS, we diagonalize it with a real, orthogonal transformation 𝐎𝐂˘subscript𝐎˘𝐂\mathrm{\mathbf{O}}_{\mathrm{\mathbf{\breve{C}}}}, such that

𝐎𝐂˘T𝐂˘𝐎𝐂˘=𝚲𝐂˘𝐈𝐂,superscriptsubscript𝐎˘𝐂T˘𝐂subscript𝐎˘𝐂subscript𝚲˘𝐂subscript𝐈𝐂\mathrm{\mathbf{O}}_{\mathrm{\mathbf{\breve{C}}}}^{\mathrm{T}}\mathrm{\mathbf{\mathrm{\mathbf{\breve{C}}}}}\mathrm{\mathbf{O}}_{\breve{\mathrm{\mathbf{C}}}}=\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{C}}}\;, (A.30)

where 𝚲𝐂˘subscript𝚲˘𝐂\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}} is the dimensionless, diagonal matrix constructed from the eigenvalues of 𝐂˘˘𝐂\mathrm{\mathbf{\mathrm{\mathbf{\breve{C}}}}} and 𝐈𝐂subscript𝐈𝐂\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{C}}} is the identity matrix with physical dimensions of capacitance.

Employing the orthogonality transformation 𝐎𝐂˘Tsuperscriptsubscript𝐎˘𝐂T\mathrm{\mathbf{O}}_{\mathrm{\mathbf{\breve{C}}}}^{\mathrm{T}} in a manner similar to the one used with 𝐎𝐋subscript𝐎𝐋\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}, we rotate (but do not scale) the coordinates for a second time. Under this second principle-axis transform, we define the eigenmode magnetic flux variable

𝚽m𝐎𝐂˘T𝚽˘,subscript𝚽msuperscriptsubscript𝐎˘𝐂T˘𝚽\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\coloneqq\mathrm{\mathbf{O}}_{\breve{\mathrm{\mathbf{C}}}}^{\mathrm{T}}\breve{\mathrm{\mathbf{\Phi}}}\;, (A.31)

in terms of which the Lagrangian linsubscriptlin\mathcal{L}_{\mathrm{lin}} is diagonal,

~full(𝚽m,𝚽˙m)12𝚽˙mT𝚲𝐂˘𝐈𝐂𝚽˙m12𝚽mT𝐈𝐋1𝚽m+~nl(𝚽m),subscript~fullsubscript𝚽msubscript˙𝚽m12superscriptsubscript˙𝚽mTsubscript𝚲˘𝐂subscript𝐈𝐂subscript˙𝚽m12superscriptsubscript𝚽mTsuperscriptsubscript𝐈𝐋1subscript𝚽msubscript~nlsubscript𝚽m\tilde{\mathcal{L}}_{\mathrm{full}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}\right)\coloneqq\frac{1}{2}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}^{\mathrm{T}}\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{C}}}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}-\frac{1}{2}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}^{\mathrm{T}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{L}}}^{-1}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\\ +\tilde{\mathcal{L}}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\right)\;, (A.32)

where the nonlinear part under the transformation is

~nl(𝚽m)nl(𝚽t(𝚽m)).subscript~nlsubscript𝚽msubscriptnlsubscript𝚽tsubscript𝚽m\tilde{\mathcal{L}}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\right)\coloneqq\mathcal{L}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\right)\right)\;. (A.33)
Equations of motion.

The Lagrangian equations of motion, ~𝚽mddt~𝚽˙m=𝟎~subscript𝚽mdd𝑡~subscript˙𝚽m0\frac{\partial\tilde{\mathcal{L}}}{\partial\mathrm{\mathbf{\Phi}}_{\mathrm{m}}}-\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial\tilde{\mathcal{L}}}{\partial\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}}=\mathrm{\mathbf{0}}, yield the harmonic eigenvalue equation d2dt2𝚽m+𝛀2𝚽m=𝟎superscriptd2dsuperscript𝑡2subscript𝚽msuperscript𝛀2subscript𝚽m0\frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}+\mathrm{\mathbf{\Omega}}^{2}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}=\mathrm{\mathbf{0}}, where 𝟎0\mathrm{\mathbf{0}} is the column vector of all zero elements and 𝛀2𝚲𝐂˘1𝐈ω2superscript𝛀2superscriptsubscript𝚲˘𝐂1superscriptsubscript𝐈𝜔2\mathrm{\mathbf{\Omega}}^{2}\coloneqq\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}}^{-1}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{\omega}}}^{2}. The identity matrix 𝐈ωsubscript𝐈𝜔\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{\omega}}} has physical dimensions of circular frequency. The generalized momentum canonical to 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}} is the vector of charge variables 𝐐m~𝚽˙m=𝚲𝐂˘𝐈𝐂𝚽˙msubscript𝐐m~subscript˙𝚽msubscript𝚲˘𝐂subscript𝐈𝐂subscript˙𝚽m\mathrm{\mathbf{Q}}_{\mathrm{m}}\coloneqq\frac{\partial\tilde{\mathcal{L}}}{\partial\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}}=\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{C}}}\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}.

Diagonalized Hamiltonian of the Josephson system.

The system Hamiltonian follows from the Legendre transform on ~fullsubscript~full\tilde{\mathcal{L}}_{\mathrm{full}}, full(𝚽m,𝐐m)=(𝚽˙m(𝐐m))T𝐐m~fullsubscriptfullsubscript𝚽msubscript𝐐msuperscriptsubscript˙𝚽msubscript𝐐mTsubscript𝐐msubscript~full\mathcal{H}_{\mathrm{full}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right)=\left(\dot{\mathrm{\mathbf{\Phi}}}_{\mathrm{m}}(\mathrm{\mathbf{Q}}_{\mathrm{m}})\right)^{\mathrm{T}}\mathrm{\mathbf{Q}}_{\mathrm{m}}-\tilde{\mathcal{L}}_{\mathrm{full}}, which we can partition into

full(𝚽m,𝐐m)=lin(𝚽m,𝐐m)+nl(𝚽m,𝐐m),subscriptfullsubscript𝚽msubscript𝐐msubscriptlinsubscript𝚽msubscript𝐐msubscriptnlsubscript𝚽msubscript𝐐m\mathcal{H}_{\mathrm{full}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right)=\mathcal{H}_{\mathrm{lin}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right)+\mathcal{H}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right)\;, (A.34)

where the linear and nonlinear parts of the Hamiltonian expressed in the eigenmode coordinates are

lin(𝚽m,𝐐m)subscriptlinsubscript𝚽msubscript𝐐m\displaystyle\mathcal{H}_{\mathrm{lin}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right) 12𝐐mT𝛀2𝐈𝐋𝐐m+12𝚽mT𝐈𝐋1𝚽mabsent12superscriptsubscript𝐐mTsuperscript𝛀2subscript𝐈𝐋subscript𝐐m12superscriptsubscript𝚽mTsuperscriptsubscript𝐈𝐋1subscript𝚽m\displaystyle\coloneqq\frac{1}{2}\mathrm{\mathbf{Q}}_{\mathrm{m}}^{\mathrm{T}}\mathrm{\mathbf{\Omega}}^{2}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{L}}}\mathrm{\mathbf{Q}}_{\mathrm{m}}+\frac{1}{2}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}^{\mathrm{T}}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{L}}}^{-1}\mathrm{\mathbf{\Phi}}_{\mathrm{m}} (A.35a)
=m=1M1L2ωm2Qm2+121LΦm2,absentsuperscriptsubscript𝑚1𝑀subscript1𝐿2superscriptsubscript𝜔𝑚2superscriptsubscript𝑄𝑚212subscript1𝐿superscriptsubscriptΦ𝑚2\displaystyle=\sum_{m=1}^{M}\frac{1_{L}}{2}\omega_{m}^{2}Q_{m}^{2}+\frac{1}{2}1_{L}\Phi_{m}^{2}\;, (A.35b)
nl(𝚽m,𝐐m)subscriptnlsubscript𝚽msubscript𝐐m\displaystyle\mathcal{H}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}},\mathrm{\mathbf{Q}}_{\mathrm{m}}\right) ~nl(𝚽m),absentsubscript~nlsubscript𝚽m\displaystyle\coloneqq-\tilde{\mathcal{L}}_{\mathrm{nl}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\right)\;, (A.35c)

where the diagonal eigenfrequency matrix 𝛀𝛀\mathrm{\mathbf{\Omega}} of the Hamiltonian linsubscriptlin\mathcal{H}_{\mathrm{lin}} is

𝛀𝛀\displaystyle\mathrm{\mathbf{\Omega}} 𝚲𝐂˘1/2𝐈ω=(ω1ωM),absentsuperscriptsubscript𝚲˘𝐂12subscript𝐈𝜔matrixsubscript𝜔1missing-subexpressionmissing-subexpressionmissing-subexpressionsubscript𝜔𝑀\displaystyle\coloneqq\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{\breve{C}}}}^{-1/2}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{\mathbf{\omega}}}=\begin{pmatrix}\omega_{1}\\ &\ddots\\ &&\omega_{M}\end{pmatrix}\;, (A.36)

and 1Lsubscript1𝐿1_{L} is unity carrying physical dimensions of inductance. The entries of 𝛀𝛀\mathrm{\mathbf{\Omega}} are the eigenmode frequencies ωmsubscript𝜔𝑚\omega_{m} of the linearized circuit. These correspond to the eigenfrequencies solved for by the FE eigenanalysis.

Eigenvectors of the Josephson system.

The eigenvector matrix 𝐄𝐄\mathrm{\mathbf{E}} relates the spanning-tree fluxes 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} to the eigenmode ones 𝚽msubscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{m}},

𝚽t=𝐄𝚽m.subscript𝚽t𝐄subscript𝚽m\mathrm{\mathbf{\Phi}}_{\mathrm{t}}=\mathrm{\mathbf{E}}\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\;. (A.37)

It is found by concatenating the principle-axis transformations defined in Eqs. (A.27) and (A.31),

𝐄𝐎𝐋𝚲𝐋1/2𝐎𝐂˘.𝐄subscript𝐎𝐋superscriptsubscript𝚲𝐋12subscript𝐎˘𝐂\mathrm{\mathbf{E}}\coloneqq\mathrm{\mathbf{O}}_{\mathrm{\mathbf{L}}}\mathrm{\mathbf{\Lambda}}_{\mathrm{\mathbf{L}}}^{1/2}\mathrm{\mathbf{O}}_{{\breve{\mathrm{\mathbf{C}}}}}\;. (A.38)

The eigenvector matrix 𝐄𝐄\mathrm{\mathbf{E}} is real and positive-definite, since it is the produce real, positive-definite transforms. It is dimensionless and, in general, non-symmetric. It is related to the square root of the inductance matrix, 𝐄𝐄=𝐋𝐈H1superscript𝐄𝐄superscriptsubscript𝐋𝐈H1\mathrm{\mathbf{E}}\mathrm{\mathbf{E}}^{\intercal}=\mathrm{\mathbf{L}}\mathrm{\mathbf{I}}_{\mathrm{H}}^{-1} and (𝐄1)𝐄1=𝐋1𝐈Hsuperscriptsuperscript𝐄1superscript𝐄1superscript𝐋1subscript𝐈H\left(\mathrm{\mathbf{E}}^{-1}\right)^{\intercal}\mathrm{\mathbf{E}}^{-1}=\mathrm{\mathbf{L}}^{-1}\mathrm{\mathbf{I}}_{\mathrm{H}}. The eigenvector matrix 𝐄𝐄\mathrm{\mathbf{E}} represents the eigenfield solutions found in the FE analysis. It is key in determining the quantum zero-point fluctuations of the mode and dipole fluxes, as shown in the following section.

A5 Quantizing the Josephson circuit

We quantize fullsubscriptfull\mathcal{H}_{\mathrm{full}} using Dirac’s canonical approach (Dirac1982-Book, ). Before taking passage from classical to quantum, we introduce the complex mode amplitude operator αmsubscript𝛼𝑚\alpha_{m}—the classical analog of the bosonic amplitude operator a^msubscript^𝑎𝑚\hat{a}_{m}. This provides a direct path to second quantization in the eigenmode basis of linsubscriptlin\mathcal{H}_{\mathrm{lin}}.

Complex action-angle variables.

We define the vector of action-angle variables 𝜶=(α1,,αM)T𝜶superscriptsubscript𝛼1subscript𝛼𝑀T\bm{\alpha}=\left(\alpha_{1},\ldots,\alpha_{M}\right)^{\mathrm{T}} by the non-canonical, complex transformation

𝜶(t)12𝛀(𝚽m(t)1H1/2+i𝛀𝐐m(t)1H1/2),𝜶𝑡12Planck-constant-over-2-pi𝛀subscript𝚽m𝑡superscriptsubscript1H12𝑖𝛀subscript𝐐m𝑡superscriptsubscript1H12\bm{\alpha}\left(t\right)\coloneqq\frac{1}{\sqrt{2\hbar\mathrm{\mathbf{\Omega}}}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{m}}\left(t\right)1_{\mathrm{H}}^{-1/2}+i\mathrm{\mathbf{\Omega}}\mathbf{Q}_{\mathrm{m}}\left(t\right)1_{\mathrm{H}}^{1/2}\right)\;, (A.39)

where 1Hsubscript1H1_{\mathrm{H}} is unity with dimensions of inductance. The normalization 1/2𝛀12Planck-constant-over-2-pi𝛀1/\sqrt{2\hbar\mathrm{\mathbf{\Omega}}} is chosen so that the Poisson bracket of the action-angles is {αm,αm}P=1/(i)δmmsubscriptsubscript𝛼𝑚superscriptsubscript𝛼superscript𝑚P1𝑖Planck-constant-over-2-pisubscript𝛿𝑚superscript𝑚\left\{\alpha_{m},\alpha_{m^{\prime}}^{*}\right\}_{\mathrm{P}}=1/(i\hbar)\delta_{mm^{\prime}}. In terms of the action-angles, the Hamiltonian remain diagonal,

lin=2(𝜶T𝛀𝜶+𝜶T𝛀𝜶).subscriptlinPlanck-constant-over-2-pi2superscript𝜶𝑇𝛀superscript𝜶superscript𝜶absent𝑇𝛀𝜶\mathcal{H}_{\mathrm{lin}}=\frac{\hbar}{2}\left(\bm{\alpha}^{T}\mathrm{\mathbf{\Omega}}\bm{\alpha}^{*}+\bm{\alpha}^{*T}\mathrm{\mathbf{\Omega}}\bm{\alpha}\right)\;. (A.40)

We have symmetrized linsubscriptlin\mathcal{H}_{\mathrm{lin}} to avoid operator order ambiguity. The flux is 𝚽m=𝛀1H2(𝜶+𝜶)subscript𝚽mPlanck-constant-over-2-pi𝛀subscript1H2superscript𝜶𝜶\mathrm{\mathbf{\Phi}}_{\mathrm{m}}=\sqrt{\frac{\hbar\mathrm{\mathbf{\Omega}}1_{\mathrm{H}}}{2}}\left(\bm{\alpha}^{*}+\bm{\alpha}\right).

Quantizing the action-angle variables.

Following Dirac’s prescription (Dirac1982-Book, ), we supplant Poisson brackets by commutators and promote observables to operators. The action-angles αmsubscript𝛼𝑚\alpha_{m} and αmsuperscriptsubscript𝛼𝑚\alpha_{m}^{*} promote into the ladder annihilation a^msubscript^𝑎𝑚\hat{a}_{m} and creation a^msuperscriptsubscript^𝑎𝑚\hat{a}_{m}^{\dagger} operators. Their commutator follows from the Poisson bracket (Louisell1961, ; Yurke1984, ; Devoret1995, ), i×{αm,αm}[a^m,a^m]maps-to𝑖Planck-constant-over-2-pisubscript𝛼𝑚superscriptsubscript𝛼𝑚subscript^𝑎𝑚superscriptsubscript^𝑎𝑚i\hbar\times\{\alpha_{m},\alpha_{m}^{*}\}\mapsto\left[\hat{a}_{m},\hat{a}_{m}^{\dagger}\right],

[a^m,a^k]=δmkI^,subscript^𝑎𝑚superscriptsubscript^𝑎𝑘subscript𝛿𝑚𝑘^𝐼\left[\hat{a}_{m},\hat{a}_{k}^{\dagger}\right]=\delta_{mk}\hat{I}\;, (A.41)

where [A^,B^]A^B^B^A^^𝐴^𝐵^𝐴^𝐵^𝐵^𝐴\left[\hat{A},\hat{B}\right]\coloneqq\hat{A}\hat{B}-\hat{B}\hat{A} is the commutator, δmksubscript𝛿𝑚𝑘\delta_{mk} is the Kronecker delta function, and I^^𝐼\hat{I} is the identity operator. Particles of the electromagnetic field (photons) are distinguishable bosons (Devoret1995, ); symmetrization of the many-body wavefunction is not required. Operators are in the Schrödinger picture, unless otherwise indicated by an explicit time argument.

Hamiltonian.

The quantized form of linsubscriptlin\mathcal{H}_{\mathrm{lin}}, see Eq. (A.40), is

H^lin=m=1Mωma^ma^m.\boxed{\hat{H}_{\mathrm{lin}}=\sum_{m=1}^{M}\hbar\omega_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m}\;.} (A.42)

This is Eq. (13) of the main text. The quantized form of nlsubscriptnl\mathcal{H}_{\mathrm{nl}} follows from combining Eqs. (A.21), (A.22), (A.24b) and (A.35c),

H^nl=j=1Jjnl(φ^j)=j=1Jp=3Ejcjpφ^jp.\boxed{\hat{H}_{\mathrm{nl}}=\sum_{j=1}^{J}\mathcal{E}_{j}^{\mathrm{nl}}\left(\hat{\varphi}_{j}\right)=\sum_{j=1}^{J}\sum_{p=3}^{\infty}E_{j}c_{jp}\hat{\varphi}_{j}^{p}\;.} (A.43)

where φ^jΦ^j/ϕ0subscript^𝜑𝑗subscript^Φ𝑗subscriptitalic-ϕ0\hat{\varphi}_{j}\coloneqq\hat{\Phi}_{j}/\phi_{0} is the reduced magnetic-flux operator for Josephson dipole j𝑗j. This is Eq. (14) of the main text. In the following, we decompose φ^jsubscript^𝜑𝑗\hat{\varphi}_{j} in terms of a^msubscript^𝑎𝑚\hat{a}_{m}; i.e., in second quantization with respect to the eigenmodes of linsubscriptlin\mathcal{H}_{\mathrm{lin}}. In writing Eq. (A.42), we have omitted the 12ωm12Planck-constant-over-2-pisubscript𝜔𝑚\frac{1}{2}\hbar\omega_{m} ground-state energy of every mode. In the limit of infinite discretization, the sum of these ground energies tends to infinity, a standard conceptual difficulty in quantum field theory. Since physical experiments observe changes in the energy of the field, the vacuum energy can be neglected (except for special cases; e.g., Casimir effect). To proceed, we introduce helpful notation.

Notation for vectors of operators.

A bold symbol typeset in roman, such as 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}}, denotes a vector or matrix, whose elements are constants or variables. Since variable, such as Φt1subscriptΦsubscriptt1\Phi_{\mathrm{t}_{1}} are promoted to quantum operators, Φ^t1subscript^Φsubscriptt1\hat{\Phi}_{\mathrm{t}_{1}}, we can accommodate vectors of operators in our notation with a hat symbol; e.g.,

𝚽^t=(Φ^t1Φ^t2)and𝚽^m=(Φ^m1Φ^mM).formulae-sequencesubscript^𝚽tmatrixsubscript^Φsubscriptt1subscript^Φsubscriptt2andsubscript^𝚽mmatrixsubscript^Φsubscriptm1subscript^Φsubscriptm𝑀\mathrm{\mathbf{\hat{\Phi}}}_{\mathrm{t}}=\begin{pmatrix}\hat{\Phi}_{\mathrm{t}_{1}}\\ \hat{\Phi}_{\mathrm{t}_{2}}\\ \vdots\end{pmatrix}\quad\text{and}\quad\mathrm{\mathbf{\hat{\Phi}}}_{\mathrm{m}}=\begin{pmatrix}\hat{\Phi}_{\mathrm{m}_{1}}\\ \vdots\\ \hat{\Phi}_{\mathrm{m}_{M}}\end{pmatrix}\;. (A.44)

The spanning-tree flux operator Φ^tksubscript^Φsubscriptt𝑘\hat{\Phi}_{\mathrm{t}_{k}} corresponding to the k𝑘k-th spanning-tree-branch flux variable ΦtksubscriptΦsubscriptt𝑘\Phi_{\mathrm{t}_{k}}. Similarly, the eigenmode flux operator Φ^mksubscript^Φsubscriptm𝑘\hat{\Phi}_{\mathrm{m}_{k}} corresponding to the k𝑘k-th eigenflux variable ΦmksubscriptΦsubscriptm𝑘\Phi_{\mathrm{m}_{k}}.

Zero-point fluctuations of the eigenoperators.

Inverting Eq. (A.39), one finds the vectors of eigenflux and eigencharge operators,

𝚽^msubscriptbold-^𝚽m\displaystyle\bm{\hat{\mathrm{\Phi}}}_{\mathrm{m}} \displaystyle\coloneqq 𝚽m𝐙𝐏𝐅(𝐚^+𝐚^),superscriptsubscript𝚽m𝐙𝐏𝐅superscriptbold-^𝐚bold-^𝐚\displaystyle\mathbf{\Phi_{\mathrm{m}}^{ZPF}}\left(\bm{\hat{\mathrm{a}}}^{\dagger}+\bm{\hat{\mathrm{a}}}\right)\;, (A.45a)
𝐐^msubscriptbold-^𝐐m\displaystyle\bm{\hat{\mathrm{Q}}}_{\mathrm{m}} \displaystyle\coloneqq i𝐐m𝐙𝐏𝐅(𝐚^𝐚^),𝑖superscriptsubscript𝐐m𝐙𝐏𝐅superscriptbold-^𝐚bold-^𝐚\displaystyle i\mathbf{Q_{\mathrm{m}}^{ZPF}}\left(\bm{\hat{\mathrm{a}}}^{\dagger}-\bm{\hat{\mathrm{a}}}\right)\;, (A.45b)

respectively, where the diagonal matrices of the quantum ZPF of the operators are

𝚽m𝐙𝐏𝐅superscriptsubscript𝚽m𝐙𝐏𝐅\displaystyle\mathbf{\Phi_{\mathrm{m}}^{ZPF}} \displaystyle\coloneqq 2𝛀1/2𝐈H1/2Planck-constant-over-2-pi2superscript𝛀12subscript𝐈superscriptH12\displaystyle\sqrt{\frac{\hbar}{2}}\mathrm{\mathbf{\Omega}}^{1/2}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{H}^{1/2}}
=\displaystyle= (ω121Hω121H),matrixPlanck-constant-over-2-pisubscript𝜔12subscript1Hmissing-subexpressionmissing-subexpressionmissing-subexpressionPlanck-constant-over-2-pisubscript𝜔12subscript1H\displaystyle\begin{pmatrix}\sqrt{\frac{\hbar\omega_{1}}{2}1_{\mathrm{H}}}\\ &\ddots\\ &&\sqrt{\frac{\hbar\omega_{1}}{2}1_{\mathrm{H}}}\end{pmatrix}\;,
𝐐m𝐙𝐏𝐅superscriptsubscript𝐐m𝐙𝐏𝐅\displaystyle\mathbf{Q_{\mathrm{m}}^{ZPF}} \displaystyle\coloneqq 2𝛀1/2𝐈H1/2.Planck-constant-over-2-pi2superscript𝛀12subscript𝐈superscriptH12\displaystyle\sqrt{\frac{\hbar}{2}}\mathrm{\mathbf{\Omega}}^{-1/2}\mathbb{\mathrm{\mathbf{I}}}_{\mathrm{H}^{-1/2}}\;. (A.46b)

We recall that the operator Φ^m=ΦmZPF(a^m+a^m)subscript^Φ𝑚superscriptsubscriptΦ𝑚ZPFsuperscriptsubscript^𝑎𝑚subscript^𝑎𝑚\hat{\Phi}_{m}=\Phi_{m}^{\mathrm{ZPF}}\left(\hat{a}_{m}^{\dagger}+\hat{a}_{m}\right) has a zero-mean, gaussian-distributed distribution in the ground state; i.e., its mean is 0|Φ^m|0=0quantum-operator-product0subscript^Φ𝑚00\left\langle 0\middle|\hat{\Phi}_{m}\middle|0\right\rangle=0 and its variance is 0|(Φ^m)2|0=(ΦmZPF)2quantum-operator-product0superscriptsubscript^Φ𝑚20superscriptsuperscriptsubscriptΦ𝑚ZPF2\left\langle 0\middle|\left(\hat{\Phi}_{m}\right)^{2}\middle|0\right\rangle=\left(\Phi_{m}^{\mathrm{ZPF}}\right)^{2}. The non-zero variance is representative of the ground state energy and the quantum zero-point fluctuations of the flux. The ZPF saturate the Heisenberg uncertainty bound, 𝚽m𝐙𝐏𝐅𝐐m𝐙𝐏𝐅=2𝐈superscriptsubscript𝚽m𝐙𝐏𝐅superscriptsubscript𝐐m𝐙𝐏𝐅Planck-constant-over-2-pi2𝐈\mathbf{\Phi_{\mathrm{m}}^{ZPF}}\mathbf{Q_{\mathrm{m}}^{ZPF}}=\frac{\hbar}{2}\mathrm{\mathbf{I}}. All mode ZPFs are positive, ΦmZPF,QmZPF>0superscriptsubscriptΦ𝑚ZPFsuperscriptsubscript𝑄𝑚ZPFsubscriptabsent0\Phi_{m}^{\mathrm{ZPF}},Q_{m}^{\mathrm{ZPF}}\in\mathbb{R}_{>0}.

Interpretation of the eigenmode ZPFs and effective mode inductances, and impedances.

The mode impedance Zm=ΦmZPF/QmZPF=ωm1Hsubscript𝑍𝑚superscriptsubscriptΦ𝑚ZPFsuperscriptsubscript𝑄𝑚ZPFsubscript𝜔𝑚subscript1HZ_{m}=\Phi_{m}^{\mathrm{ZPF}}/Q_{m}^{\mathrm{ZPF}}=\omega_{m}1_{\mathrm{H}} and ZPF ΦmZPF=ωm/2×1HsuperscriptsubscriptΦ𝑚ZPFPlanck-constant-over-2-pisubscript𝜔𝑚2subscript1H\Phi_{m}^{\mathrm{ZPF}}=\sqrt{\hbar\omega_{m}/2\times 1_{\mathrm{H}}} depend only on ωmsubscript𝜔𝑚\omega_{m}. These quantities are in general removed from direct physical meaning. Physical meaning is extracted by using them as computational tools to calculate the quantum ZPF of the branch fluxes and charges. The EPR method negates the need to explicitly compute them; rather, we only directly work with physically measure quantities, such as the eigenfields of the modes and the Josephson dipole EPRs, as discussed in the next sub-section.

From abstract to physical ZPFs.

The zero-point fluctuations of the spanning-tree fluxes follow from Eqs. (A.38) and (A.46),

𝚽t𝐙𝐏𝐅=𝐄𝚽m𝐙𝐏𝐅.superscriptsubscript𝚽t𝐙𝐏𝐅𝐄superscriptsubscript𝚽m𝐙𝐏𝐅\mathbf{\Phi_{\mathrm{t}}^{ZPF}}=\mathrm{\mathbf{E}}\mathbf{\Phi_{\mathrm{m}}^{ZPF}}\;. (A.47)

The (k,m)𝑘𝑚\left(k,m\right) element of 𝚽tZPFsuperscriptsubscript𝚽tZPF\mathrm{\mathbf{\Phi}}_{\mathrm{t}}^{\mathrm{ZPF}} is the ZPF of the k𝑘k-th spanning-tree branch due to mode m𝑚m and is either positive or negative. The overall sign is arbitrary. Hover, the relative sign between two branches k𝑘k and ksuperscript𝑘k^{\prime} in the same mode determines if they are excited by the mode in-phase or out-of phase with each other. The eigenvalue matrix 𝐄𝐄\mathrm{\mathbf{E}} is constructed canonically so as to correctly relate the fluctuations. The eigenvector matrix of the standard product 𝐂1𝐋1superscript𝐂1superscript𝐋1\mathrm{\mathbf{C}}^{-1}\mathrm{\mathbf{L}}^{-1}, obtained in the Lagrangian equations of motion, is not canonical and cannot be used in Eq. (A.47).

Second quantization of the dipole flux operator.

Since 𝐄𝐄\mathrm{\mathbf{E}} is real, the reduced-magnetic-flux ZPF amplitudes φmjsubscript𝜑𝑚𝑗\varphi_{mj} can be chosen to be real-valued numbers; Eq. 14(c) of the main text follows,

φ^jΦ^j/ϕ0=m=1Mφmj(a^m+a^m).\boxed{\hat{\varphi}_{j}\coloneqq\hat{\Phi}_{j}/\phi_{0}=\sum_{m=1}^{M}\varphi_{mj}\left(\hat{a}_{m}+\hat{a}_{m}^{\dagger}\right)\;.} (A.48)

A6 Energy-participation ratio (EPR)

We define energy-participation ratio in the context of quantum circuits. We motivate the omission of vacuum energy contributions and use the EPR to find the quantum zero–point fluctuations φmjsubscript𝜑𝑚𝑗\varphi_{mj}.

Definition of EPR in plain english.

The EPR pmjsubscript𝑝𝑚𝑗p_{mj} of Josephson dipole j𝑗j in eigenmode m𝑚m is the fraction of inductive energy allocated to the Josephson dipole when mode m𝑚m is excited.

Interpretation of the energy-participation ratio.

The EPR quantifies how much of the inductive energy of a mode is allocated to a Josephson dipole. The lowest possible participation is zero—the Josephson dipole inductor does not participate in the mode. When the mode is excited, none of the excitation energy flows to the dipole. The largest possible participation is unity. When the mode is excited, all of its excitation flows to the Josephson dipole; none of it is distributed to any other inductor. As we show in the following, a larger participation leads to larger quantum vacuum fluctuations.

Transmon-qubit example.

Consider the transmon qubit coupled to a readout cavity mode (see Methods). The Josephson junction has an EPR of near unity in the qubit eigenmode, since nearly all of the inductance of the transmon is due to the kinetic inductance of the tunnel junction. On the other hand, the EPR of the junction in the cavity mode is on the order of 102superscript10210^{-2}. The junction kinetic inductance contributes only on the order of one percent to the total inductance of the cavity, associated with the large cavity box. This leads to the large ZPF of the junction flux in the transmon mode, φqZPF1similar-tosuperscriptsubscript𝜑𝑞ZPF1\varphi_{q}^{\mathrm{ZPF}}\sim 1, and the therefore much smaller fluctuations of the junction in the cavity mode φqZPF101similar-tosuperscriptsubscript𝜑𝑞ZPFsuperscript101\varphi_{q}^{\mathrm{ZPF}}\sim 10^{-1} (see also Sec. A7).

Energy offset.

The EPR is defined in terms of the energy excitation of a mode, rather than in terms of its absolute energy.

Vacuum energy.

Mathematically, the vacuum energy is a consequence of the non-commutativity of a^^𝑎\hat{a} and a^superscript^𝑎\hat{a}^{\dagger}, which in effect adds a constant offset to the energy of each mode. In calculating the EPR, we take the reference energy of an element relative to its vacuum energy.

Equipartition theorem.

Classically, for a linear circuit, the energy of an eigenmode oscillates in time between inductive and capacitive energy. At periodic intervals given by π/ωm𝜋subscript𝜔𝑚\pi/\omega_{m}, the inductive energy of the mode is exactly zero. For this reason, we use the time-average energy of the Josephson dipole and the eigenmode. We recall that the time-averaged energy is half of the peak energy. This can be exploited in expressing the total inductive energy as half the total mode energy,

^ind¯=12H^lin¯=12mωma^ma^m¯.¯delimited-⟨⟩subscript^ind12¯delimited-⟨⟩subscript^𝐻lin12subscript𝑚Planck-constant-over-2-pisubscript𝜔𝑚¯delimited-⟨⟩superscriptsubscript^𝑎𝑚subscript^𝑎𝑚\overline{\left\langle\hat{\mathcal{E}}_{\mathrm{ind}}\right\rangle}=\frac{1}{2}\overline{\left\langle\hat{H}_{\mathrm{lin}}\right\rangle}=\frac{1}{2}\sum_{m}\hbar\omega_{m}\overline{\left\langle\hat{a}_{m}^{\dagger}\hat{a}_{m}\right\rangle}\;. (A.49)

We will discuss what states are used in calculating the expectation values below. The ground state energy is taken to be zero.

Calculating the EPR.

These ideas lead us to the following definition of the EPR in the quantum setting,

pmj^j,lin¯/^ind¯,subscript𝑝𝑚𝑗¯delimited-⟨⟩subscript^𝑗lin¯delimited-⟨⟩subscript^indp_{mj}\coloneqq\overline{\left<\hat{\mathcal{E}}_{j\mathrm{,lin}}\right>}/\overline{\left<\hat{\mathcal{E}}_{\mathrm{ind}}\right>}\;, (A.50)

where the over-line denotes a time average and the expectation value is taken over a state with an excitation only in mode m𝑚m. As discussed above, all energies are referenced to their ground-state expectation values, to disregard vacuum-energy contributions (for special notation to accommodate this, see remark on Wick ordering at the end of this section). Using Eqs. (A.49) and (A.42), we simplify the denominator, ^ind¯=m12a^ma^m¯delimited-⟨⟩subscript^indsubscript𝑚12superscriptsubscript^𝑎𝑚subscript^𝑎𝑚\overline{\left<\hat{\mathcal{E}}_{\mathrm{ind}}\right>}=\sum_{m}\frac{1}{2}\hat{a}_{m}^{\dagger}\hat{a}_{m}. The operator for the junction energy follows from Eq. (A.13a), ^j,lin=12Ej(φ^j2φ^j20)subscript^𝑗lin12subscript𝐸𝑗superscriptsubscript^𝜑𝑗2subscriptdelimited-⟨⟩superscriptsubscript^𝜑𝑗20\hat{\mathcal{E}}_{j\mathrm{,lin}}=\frac{1}{2}E_{j}\left(\hat{\varphi}_{j}^{2}-\left\langle\hat{\varphi}_{j}^{2}\right\rangle_{0}\right), where 0subscript0\left\langle\,\right\rangle_{0} indicates an expectation value over the ground state. Thus,

pmj=12Ejφ^j2¯12Ejφ^j20¯12mωma^ma^m.subscript𝑝𝑚𝑗¯delimited-⟨⟩12subscript𝐸𝑗superscriptsubscript^𝜑𝑗2¯subscriptdelimited-⟨⟩12subscript𝐸𝑗superscriptsubscript^𝜑𝑗2012subscript𝑚Planck-constant-over-2-pisubscript𝜔𝑚delimited-⟨⟩superscriptsubscript^𝑎𝑚subscript^𝑎𝑚p_{mj}=\frac{\overline{\left\langle\frac{1}{2}E_{j}\hat{\varphi}_{j}^{2}\right\rangle}-\overline{\left\langle\frac{1}{2}E_{j}\hat{\varphi}_{j}^{2}\right\rangle_{0}}}{\frac{1}{2}\sum_{m}\hbar\omega_{m}\left\langle\hat{a}_{m}^{\dagger}\hat{a}_{m}\right\rangle}\;. (A.51)

Using Eq. (A.48),

φ^j2φ^j20=m=1Mφmj2(2a^ma^m+a^m2+a^m2)+mmMφmjφmj(a^ma^m+a^ma^m+a^ma^m+a^ma^m).superscriptsubscript^𝜑𝑗2subscriptdelimited-⟨⟩superscriptsubscript^𝜑𝑗20superscriptsubscript𝑚1𝑀superscriptsubscript𝜑𝑚𝑗22superscriptsubscript^𝑎𝑚subscript^𝑎𝑚superscriptsubscript^𝑎𝑚2superscriptsubscript^𝑎𝑚absent2superscriptsubscriptsuperscript𝑚𝑚𝑀subscript𝜑𝑚𝑗subscript𝜑superscript𝑚𝑗subscript^𝑎𝑚subscript^𝑎superscript𝑚subscript^𝑎𝑚superscriptsubscript^𝑎superscript𝑚superscriptsubscript^𝑎𝑚subscript^𝑎superscript𝑚superscriptsubscript^𝑎𝑚superscriptsubscript^𝑎superscript𝑚\hat{\varphi}_{j}^{2}-\left\langle\hat{\varphi}_{j}^{2}\right\rangle_{0}=\sum_{m=1}^{M}\varphi_{mj}^{2}\left(2\hat{a}_{m}^{\dagger}\hat{a}_{m}+\hat{a}_{m}^{2}+\hat{a}_{m}^{\dagger 2}\right)+\\ \sum_{m^{\prime}\neq m}^{M}\varphi_{mj}\varphi_{m^{\prime}j}\left(\hat{a}_{m}\hat{a}_{m^{\prime}}+\hat{a}_{m}\hat{a}_{m^{\prime}}^{\dagger}+\hat{a}_{m}^{\dagger}\hat{a}_{m^{\prime}}+\hat{a}_{m}^{\dagger}\hat{a}_{m^{\prime}}^{\dagger}\right)\;. (A.52)
EPR for a Fock state excitation.

We can take expectation value in Eq. (A.51) for a Fock excitation of n𝑛n photons in mode m𝑚m, denoted |nmketsubscript𝑛𝑚\left|n_{m}\right\rangle. We recall that the many-body vacuum state is |0|vack=1M|0kket0ketvactensor-productsuperscriptsubscriptproduct𝑘1𝑀subscriptket0𝑘\left|0\right\rangle\coloneqq\left|\mathrm{vac}\right\rangle\coloneqq\prod_{k=1}^{M}\otimes\left|0\right\rangle_{k}, where |0ksubscriptket0𝑘\left|0\right\rangle_{k} denotes the single-particle, zero Fock state of mode k𝑘k; thus, |nm=1nm!(a^m)nm|vacketsubscript𝑛𝑚1subscript𝑛𝑚superscriptsuperscriptsubscript^𝑎𝑚subscript𝑛𝑚ketvac\left|n_{m}\right\rangle=\frac{1}{\sqrt{n_{m}!}}\left(\hat{a}_{m}^{\dagger}\right)^{n_{m}}\left|\mathrm{vac}\right\rangle. To take the expectation value of the Josephson dipole flux φ^j2delimited-⟨⟩superscriptsubscript^𝜑𝑗2\left\langle\hat{\varphi}_{j}^{2}\right\rangle\leavevmode\nobreak\ , we recall that Fock states have trivial time dynamic, and use Eq. (A.52). We observe that the expectation value of any product of two amplitude operators from different modes is zero. The only non-zero term in Eq. (A.52) is the a^ma^msuperscriptsubscript^𝑎𝑚subscript^𝑎𝑚\hat{a}_{m}^{\dagger}\hat{a}_{m} term. Thus, the EPR of the Josephson dipole for the Fock state |nmketsubscript𝑛𝑚\left|n_{m}\right\rangle is

pmj=Ejφmj212ωm.subscript𝑝𝑚𝑗subscript𝐸𝑗superscriptsubscript𝜑𝑚𝑗212Planck-constant-over-2-pisubscript𝜔𝑚p_{mj}=\frac{E_{j}\varphi_{mj}^{2}}{\frac{1}{2}\hbar\omega_{m}}\;. (A.53)

The EPR is independent of the excitation amplitude nmsubscript𝑛𝑚n_{m}.

EPR for a coherent state excitation.

A coherent state excitation of mode m𝑚m is denoted |βmD^(βm,a^m)|vacketsubscript𝛽𝑚^𝐷subscript𝛽𝑚subscript^𝑎𝑚ketvac\left|\beta_{m}\right\rangle\coloneqq\hat{D}\left(\beta_{m},\hat{a}_{m}\right)\left|\mathrm{vac}\right\rangle, where D^^𝐷\hat{D} is the displacement operator of the m𝑚mth mode, D^(βm,a^m)exp(βma^mβma^m)^𝐷subscript𝛽𝑚subscript^𝑎𝑚subscript𝛽𝑚superscriptsubscript^𝑎𝑚superscriptsubscript𝛽𝑚subscript^𝑎𝑚\hat{D}\left(\beta_{m},\hat{a}_{m}\right)\coloneqq\exp\left(\beta_{m}\hat{a}_{m}^{\dagger}-\beta_{m}^{\ast}\hat{a}_{m}\right), and β𝛽\beta is a non-zero complex number. The coherent state time-evolution is a rotation, |βm(t)=|βm(0)eiωmtketsubscript𝛽𝑚𝑡ketsubscript𝛽𝑚0superscript𝑒𝑖subscript𝜔𝑚𝑡\left|\beta_{m}\left(t\right)\right\rangle=\left|\beta_{m}\left(0\right)e^{-i\omega_{m}t}\right\rangle. Using Eq. (A.51), we find the EPR to again be excitation independent and exactly equal to that given in Eq. (A.53).

Quantum fluctuations in terms of EPR.

The EPR for a Fock or coherent state excitation is given by Eq. (A.53). Inverting this expression, we find the ZPF in terms of the EPR,

φmj2=pmjωm2Ej.\boxed{\varphi_{mj}^{2}=p_{mj}\frac{\hbar\omega_{m}}{2E_{j}}\;.} (A.54)

In the case of a single Josephson dipole in the circuit, Eq. (A.54) reduces to Eq. (6) of the main text. The left-hand side of Eq. (A.54) gives the root-mean-square deviation of the quantum fluctuations of the reduced magnetic flux of Josephson dipole j𝑗j, referenced to its equilibrium value. The right-hand side of Eq. (A.54) comprises three classically-known parameters: the eigenmode frequency ωmsubscript𝜔𝑚\omega_{m}, the Josephson dipole energy scale Ejsubscript𝐸𝑗E_{j}, and the EPR pmjsubscript𝑝𝑚𝑗p_{mj}.

Sign of the EPR.

Solving Eq. (A.54) explicitly,

φmj=smjpmjωm2Ej,subscript𝜑𝑚𝑗subscript𝑠𝑚𝑗subscript𝑝𝑚𝑗Planck-constant-over-2-pisubscript𝜔𝑚2subscript𝐸𝑗\varphi_{mj}=s_{mj}\sqrt{p_{mj}\frac{\hbar\omega_{m}}{2E_{j}}}\;, (A.55)

where smjsubscript𝑠𝑚𝑗s_{mj} is the EPR sign, smj{1,+1}subscript𝑠𝑚𝑗11s_{mj}\in\left\{-1,+1\right\}. In practice, the sign is calculated using Eq. (C.8). Algebraically, the sign can be found from the eigenvector matrix 𝐄𝐄\mathrm{\mathbf{E}}. If the j𝑗jth row of the eigenvector matrix corresponds to the j𝑗jth spanning-tree branch flux and the m𝑚mth column corresponds to the m𝑚mth eigenmode, then  smj=sign([𝐄]mj)subscript𝑠𝑚𝑗signsubscriptdelimited-[]𝐄𝑚𝑗s_{mj}=\operatorname{sign}\left(\left[\mathrm{\mathbf{E}}\right]_{mj}\right), where [𝐄]mjsubscriptdelimited-[]𝐄𝑚𝑗\left[\mathrm{\mathbf{E}}\right]_{mj} is the (m,j)𝑚𝑗\left(m,j\right)th entry of the matrix.

EPR sign: sign freedom.

The value of an individual sign smjsubscript𝑠𝑚𝑗s_{mj} is completely arbitrary. It does not have measurable consequences in the same way that the amplitude of a standing mode can be taken to be positive or negative, indicating a π𝜋\pi-phase shift. The sign smjsubscript𝑠𝑚𝑗s_{mj} derives its physical meaning in relationship to other signs smjsubscript𝑠𝑚superscript𝑗s_{mj^{\prime}} of the same mode but different Josephson dipoles (see Supplementary Figure S8). If the signs of two dipoles are the same (resp., different), then the two dipoles oscillate in-phase (resp., out-of-phase). The sign of the EPR can flipped by either flipping the reference direction of the junction or the overall phase of the mode.

EPR for circuit design.

The EPR pmjsubscript𝑝𝑚𝑗p_{mj} in Eq. (A.54) is essentially the only free parameter in engineering the quantum ZPF and the amplitudes of the non-linear couplings. It is readily calculated from the classical eigenmode FE simulation of the distributed physical layout of the circuit, as detailed in Sec. C. In this way, the EPR serves as a bridge between the linearized classical circuits treated with FE solvers and the Josephson circuit in the quantum domain.

Remark: Equivalent formulation using Wick normal-ordered expectation value for energies.

Mathematically enforcing the energy to be referenced to that of the vacuum state can be equivalently accomplished by disregarding the commutation relationships in expectation values of the energy. The compact notation that accomplishes this was introduced in quantum optics and is also used in quantum field theory: for an operator O^^𝑂\hat{O}, one takes the (Wick) normal-ordered form of the operator (Gerry2005, ), denoted by a colon on either side of the operator, :𝑂::𝑂:{:\mathrel{\mspace{1.0mu}O\mspace{1.0mu}}:}. Thus, :a^a^:=a^a^{:\mathrel{\mspace{1.0mu}\hat{a}\hat{a}^{\dagger}\mspace{1.0mu}}:}=\hat{a}^{\dagger}\hat{a} and :(a^a^)2:=a^2a^2{:\mathrel{\mspace{1.0mu}\left(\hat{a}^{\dagger}\hat{a}\right)^{2}\mspace{1.0mu}}:}=\hat{a}^{\dagger 2}\hat{a}^{2}. The normal-ordered form of the operators should not be confused with the normal-ordering transformation 𝒩𝒩\mathscr{N} applied to the operator, which takes the commutation relations into account; e.g., 𝒩[a^a^]=a^a^+I^𝒩delimited-[]^𝑎superscript^𝑎superscript^𝑎^𝑎^𝐼\text{$\text{$\mathscr{N}$}$}\left[\hat{a}\hat{a}^{\dagger}\right]=\hat{a}^{\dagger}\hat{a}+\hat{I} and 𝒩[(a^a^)2]=a^2a2+a^a^𝒩delimited-[]superscriptsuperscript^𝑎^𝑎2superscript^𝑎absent2superscript𝑎2superscript^𝑎^𝑎\text{$\mathscr{N}$}\left[\left(\hat{a}^{\dagger}\hat{a}\right)^{2}\right]=\hat{a}^{\dagger 2}a^{2}+\hat{a}^{\dagger}\hat{a}. In the presence of non-linear interactions, the vacuum energy leads to observable effects, such as the Lamb shift, introduced in Eq. (7) of the main text. The definition given in Eq. (A.50) can be equivalently stated using Wick notation:

pmj:^j,lin:¯/:^ind:¯.p_{mj}\coloneqq\overline{\left<{:\mathrel{\mspace{1.0mu}\hat{\mathcal{E}}_{j\mathrm{,lin}}\mspace{1.0mu}}:}\right>}/\overline{\left<{:\mathrel{\mspace{1.0mu}\hat{\mathcal{E}}_{\mathrm{ind}}\mspace{1.0mu}}:}\right>}\;. (A.56)

A7 Universal EPR properties

The energy-participation ratios obey four universal properties. These properties are valid regardless of the circuit topology and nature of the Josephson dipoles. They follow directly from the normal-mode structure of the eigenmodes of Hlinsubscript𝐻linH_{\mathrm{lin}}, obtained in Sec. A4 and are linked to the ZPF, as discussed in Sec. A6. In the following, M𝑀M denotes the total number of modes.

The EPR is bounded.

The EPR is an energy fraction. It follows from its definition in Eq. (A.50) that it is a real number comprised between zero and one—since the Josephson dipole energy is always positive and equal to or smaller than the total inductive energy of the mode,

0pmj1.0subscript𝑝𝑚𝑗10\leq p_{mj}\leq 1\;. (A.57)
The total EPR of a dipole follows a sum rule—it is conserved while being diluted among the modes.

The total EPR of a dipole is conserved—it is exactly unity across all modes,

m=1Mpmj=1forj{1,,J},formulae-sequencesuperscriptsubscript𝑚1𝑀subscript𝑝𝑚𝑗1for𝑗1𝐽\sum_{m=1}^{M}p_{mj}=1\quad\text{for}\ j\in\left\{1,\ldots,J\right\}\;, (A.58)

which follows from Eqs. (A.38) and (A.51). Increasing the EPR of a Josephson dipole in one mode will proportionally reduce its participation across other modes. The EPR is neither created, nor destroyed. It is distributed among the modes. Adding additional modes to the circuit or removing modes from the circuit does not increase or decrease the total EPR of a dipole.

The total EPR of a mode is at most unity.

A single mode m𝑚m can at most have a total EPR of unity, and no less than zero,

0j=1Jpmj1form{1,,M}.formulae-sequence0superscriptsubscript𝑗1𝐽subscript𝑝𝑚𝑗1for𝑚1𝑀0\leq\sum_{j=1}^{J}p_{mj}\leq 1\quad\text{for}\ m\in\left\{1,\ldots,M\right\}\;. (A.59)

The upper bound is saturated when there are no linear inductors excited in the mode.

The vector EPRs of two dipoles are orthogonal.

For each dipole, we can define a vector EPR by the components {smjpmj|m=1,,M}conditional-setsubscript𝑠𝑚𝑗subscript𝑝𝑚𝑗𝑚1𝑀\left\{s_{mj}\sqrt{p_{mj}}\,|\,m=1,\ldots,M\right\}. The vector EPRs of two dipoles are orthogonal in the following sense

m=1Msmjsmjpmjpmjsuperscriptsubscript𝑚1𝑀subscript𝑠𝑚𝑗subscript𝑠𝑚superscript𝑗subscript𝑝𝑚𝑗subscript𝑝𝑚superscript𝑗\displaystyle\sum_{m=1}^{M}s_{mj}s_{mj^{\prime}}\sqrt{p_{mj}p_{mj^{\prime}}}\, =0forjj,formulae-sequenceabsent0for𝑗superscript𝑗\displaystyle=0\quad\text{for}\ j\neq j^{\prime}\;, (A.60)

where smjsubscript𝑠𝑚𝑗s_{mj} is the EPR sign.

Remark: Use of these four properties in quantum circuit design.

These universal EPR constraints are useful in the design of quantum circuits, especially for designs that require weak and strong nonlinear interactions simultaneously. For example, see Methods, where we discuss the impossibility of a design we targeted due to the EPR constraints. We subsequently use the constraints to obtain a best approximation of the design, and to gain insight into the range of possible non-linear couplings. In our experience, the EPR has allowed us to thus circumvent the need to run a prohibitively expensive finite-element sweep to explore the many possible design-parameter regimes.

A8 The biased Josephson system: equilibrium state in the presence of persistent currents

Refer to caption
Supplementary Figure S1: Example of a Josephson dipole embedded in frustrated loop. (a) Illustration of a Josephson tunnel junction (with Josephson energy Ejsubscript𝐸𝑗E_{j}) embedded in a distributed, superconducting ring subjected to an external magnetic-flux bias ΦextsubscriptΦext\Phi_{\mathrm{ext}}. (b) Lumped-element model a junction in an inductive ring (with inductance L𝐿L) frustrated by an external flux bias ΦextsubscriptΦext\Phi_{\mathrm{ext}}, a voltage source vssubscript𝑣𝑠v_{s}, and a current source issubscript𝑖𝑠i_{s}.

The magnetic flux across a Josephson dipole ΦjsubscriptΦ𝑗\Phi_{j}, introduced in Sec. A1, is defined as a deviation away from the equilibrium configuration of the Josephson system. A system that incorporates an active element—one that can act as a voltage or current source—or a system subjected to a frustrated constraint—such as found in a conducting loop frustrated by an external magnetic flux—can have a non-zero equilibrium state. Referencing the Josephson dipole and spanning-tree fluxes 𝚽tsubscript𝚽t\mathrm{\mathbf{\Phi}}_{\mathrm{t}} from equilibrium guarantees that the total inductive energy ind(𝚽t)subscriptindsubscript𝚽t\mathcal{E}_{\mathrm{ind}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t}}\right) is at a minimum for 𝚽t=𝟎subscript𝚽t0\mathrm{\mathbf{\Phi}}_{\mathrm{t}}=\mathrm{\mathbf{0}}.

Static-equilibrium conditions.

In static equilibrium, the net generalized forces (currents) at each node in the system vanish and, consequently, the generalized acceleration vanishes, d2𝚽tdt2=𝟎superscriptd2subscript𝚽tdsuperscript𝑡20\frac{\mathrm{d}^{2}\mathrm{\mathbf{\Phi}}_{\mathrm{t}}}{\mathrm{d}t^{2}}=\mathrm{\mathbf{0}}. The corresponding magnetic flux of the spanning-tree branches in static equilibrium 𝚽t,eqsubscript𝚽teq\mathrm{\mathbf{\Phi}}_{\mathrm{t,eq}} can be found by extremizing the Lagrangian with respect to the generalized position variables (Landau1982, ),

𝚽t(𝚽t,eq.)=ind𝚽t(𝚽t,eq)=𝟎.subscript𝚽tsubscript𝚽teqsubscriptindsubscript𝚽tsubscript𝚽teq0-\frac{\partial\mathcal{L}}{\partial\mathrm{\mathbf{\Phi}}_{\mathrm{t}}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t,eq.}}\right)=\frac{\partial\mathcal{E}_{\mathrm{ind}}}{\partial\mathrm{\mathbf{\Phi}}_{\mathrm{t}}}\left(\mathrm{\mathbf{\Phi}}_{\mathrm{t,eq}}\right)=\mathrm{\mathbf{0}}\;. (A.61)

Solving these conditions for the equilibrium flux 𝚽t,eqsubscript𝚽teq\mathrm{\mathbf{\Phi}}_{\mathrm{t,eq}} amounts to solving for the direct-current (dc) operating point of the circuit—a standard, classical problem; although one that is non-local in nature and, in the presence of Josephson dipoles, one that involves solutions to non-linear equations. In general, classical numerical methods can be used to find the equilibrium of the circuit, especially when the equilibrium equations are transcendental. However, for many practical situations, Eq. (A.61) simplifies or need not be evaluated at all, as discussed in Sec. A9.

Equilibrium flux of a Josephson Dipole in isolation vs. in a circuit.

In general, a Josephson dipole has a different equilibrium flux across its terminals when in isolation vs. when embedded in a system. When in isolation, the native equilibrium flux of the Josephson dipole is found from the simple local condition jΦj=0subscript𝑗subscriptΦ𝑗0\frac{\partial\mathcal{E}_{j}}{\partial\Phi_{j}}=0, where jsubscript𝑗\mathcal{E}_{j} is the energy function of the Josephson dipole and ΦjsubscriptΦ𝑗\Phi_{j} is the magnetic flux across the dipole. When embedded in a system, the equilibrium flux of the Josephson dipole, found from Eq. (A.61), is in general not a local property of the dipole anymore—but one of the system, as illustrated by the following example.

Refer to caption
Supplementary Figure S2: Mechanical analogy of a Josephson dipole in isolation vs. in a frustrated system. (a) Depiction of a mechanical spring in isolation. Without applied forces on the spring, the native equilibrium length of the spring is Leq0superscriptsubscript𝐿eq0L_{\mathrm{eq}}^{0}. Stretching or compressing the spring from equilibrium is measured by the deviation x𝑥x away from the native equilibrium. The spring constant is k𝑘k. (b) Depiction of two connected springs, with spring constant ksuperscript𝑘k^{\prime} and k′′superscript𝑘′′k^{\prime\prime}, constrained between two walls separated by a fixed distance Lextsubscript𝐿extL_{\mathrm{ext}}. The equilibrium lengths of the two springs in the system are Leqsuperscriptsubscript𝐿eqL_{\mathrm{eq}}^{\prime} and Leq′′superscriptsubscript𝐿eq′′L_{\mathrm{eq}}^{\prime\prime}. These differ from their equilibrium lengths in isolation; i.e., LeqLeqsubscript𝐿eqsuperscriptsubscript𝐿eqL_{\mathrm{eq}}\neq L_{\mathrm{eq}}^{\prime}. Deviations from the system equilibrium are denoted by x𝑥x.
Example: Josephson junction in a frustrated ring.

Consider a Josephson tunnel junction in isolation—not connected to any other elements. It’s energy function, see Eq. (A.5), is i(Φi)=EJcos(Φi/ϕ0)subscript𝑖subscriptΦ𝑖subscript𝐸𝐽subscriptΦ𝑖subscriptitalic-ϕ0\mathcal{E}_{i}\left(\Phi_{i}\right)=-E_{J}\cos\left(\Phi_{i}/\phi_{0}\right), where ΦisubscriptΦ𝑖\Phi_{i} is the total magnetic flux across the junction in isolation. (Here, the subscript i𝑖i denotes the junction in isolation.) The energy is minimized at the native equilibrium-flux value Φeqnat=0superscriptsubscriptΦeqnat0\Phi_{\mathrm{eq}}^{\mathrm{nat}}=0. Now, consider embedding this junction in a superconducting ring frustrated by an external magnetic flux ΦextsubscriptΦext\Phi_{\mathrm{ext}}, as depicted in Supplementary Figure S1(a). Suppose the geometric ring inductance is L𝐿L, see Supplementary Figure S1(b). The equilibrium condition of the circuit, given by Eq. (A.61), reduces to Kepler’s transcendental equation

L1Φi+ϕ01EJsin(Φi/ϕ0)=iseff,superscript𝐿1subscriptΦ𝑖superscriptsubscriptitalic-ϕ01subscript𝐸𝐽subscriptΦ𝑖subscriptitalic-ϕ0superscriptsubscript𝑖𝑠effL^{-1}\Phi_{i}+\phi_{0}^{-1}E_{J}\sin\left(\Phi_{i}/\phi_{0}\right)=i_{s}^{\mathrm{eff}}\;, (A.62)

where the effective dc current sourced to the loop is iseff=L1Φextsuperscriptsubscript𝑖𝑠effsuperscript𝐿1subscriptΦexti_{s}^{\mathrm{eff}}=L^{-1}\Phi_{\mathrm{ext}}. In the additional presence of a voltage source vssubscript𝑣𝑠v_{s} and current source issubscript𝑖𝑠i_{s} as depicted in see Supplementary Figure S1(b), the net effective current is iseff=L1Φext+issuperscriptsubscript𝑖𝑠effsuperscript𝐿1subscriptΦextsubscript𝑖𝑠i_{s}^{\mathrm{eff}}=L^{-1}\Phi_{\mathrm{ext}}+i_{s}.

A solution Φeq(Φext,Φs,is)subscriptΦeqsubscriptΦextsubscriptΦ𝑠subscript𝑖𝑠\Phi_{\mathrm{eq}}\left(\Phi_{\mathrm{ext}},\Phi_{s},i_{s}\right) of Eq. (A.62) can be obtained numerically or using Lagrange inversion. In the EPR method, the canonical flux deviation away from the selected equilibrium ΦeqsubscriptΦeq\Phi_{\mathrm{eq}} is then

ΦjΦiΦeq.subscriptΦ𝑗subscriptΦ𝑖subscriptΦeq\Phi_{j}\coloneqq\Phi_{i}-\Phi_{\mathrm{eq}}\;. (A.63)

The energy function of the junction now in terms of the deviation ΦjsubscriptΦ𝑗\Phi_{j} is

j(Φj)EJcos((Φeq+Φj)/ϕ0).subscript𝑗subscriptΦ𝑗subscript𝐸𝐽subscriptΦeqsubscriptΦ𝑗subscriptitalic-ϕ0\mathcal{E}_{j}\left(\Phi_{j}\right)\coloneqq-E_{J}\cos\left(\left(\Phi_{\mathrm{eq}}+\Phi_{j}\right)/\phi_{0}\right)\;. (A.64)

Using the series expansion of Eq. (A.64), given in Eq. (A.7), the Josephson dipole energy of the junction, as defined in Eq. (A.13), is EJ(Φext,Φs,is)EJcos(Φeq/ϕ0)subscript𝐸𝐽subscriptΦextsubscriptΦ𝑠subscript𝑖𝑠subscript𝐸𝐽subscriptΦeqsubscriptitalic-ϕ0E_{J}\left(\Phi_{\mathrm{ext}},\Phi_{s},i_{s}\right)\coloneqq E_{J}\cos\left(\Phi_{\mathrm{eq}}/\phi_{0}\right) and the leading-order non-linear coefficient is cj3(Φext,Φs,is)=16tan(Φeq/ϕ0)subscript𝑐𝑗3subscriptΦextsubscriptΦ𝑠subscript𝑖𝑠16subscriptΦeqsubscriptitalic-ϕ0c_{j3}\left(\Phi_{\mathrm{ext}},\Phi_{s},i_{s}\right)=-\frac{1}{6}\tan\left(\Phi_{\mathrm{eq}}/\phi_{0}\right).

Mechanical analogy: spring in isolation.

By way of analogy, a Josephson dipole is like a mechanical spring. Imagine a spring in isolation, see Supplementary Figure S2(a). No external forces act on it. Its native equilibrium length (when the spring is in isolation) is Leq0superscriptsubscript𝐿eq0L_{\mathrm{eq}}^{0}. Stretching the spring to length L𝐿L results in the force F=k(xi)𝐹𝑘subscript𝑥𝑖F=-k\left(x_{i}\right) on its ends, where  xiLLeq0subscript𝑥𝑖𝐿superscriptsubscript𝐿eq0x_{i}\coloneqq L-L_{\mathrm{eq}}^{0} denotes a deviation away from equilibrium. In general, k(xi)𝑘subscript𝑥𝑖k\left(x_{i}\right) is a non-linear function; however, in equilibrium, k(0)=0𝑘00k\left(0\right)=0. For a linear spring, the energy, i(xi)=12kxi2subscript𝑖subscript𝑥𝑖12𝑘superscriptsubscript𝑥𝑖2\mathcal{E}_{i}\left(x_{i}\right)=\frac{1}{2}kx_{i}^{2}, is minimized for x=0𝑥0x=0.

Mechanical analogy: spring in a frustrated system.

Imagine embedding the spring in a simple system as depicted in Supplementary Figure S2(b). The system imposes the KVL-like constraint condition L+L′′=Lextsuperscript𝐿superscript𝐿′′subscript𝐿extL^{\prime}+L^{\prime\prime}=L_{\mathrm{ext}}, where Lsuperscript𝐿L^{\prime} is the length of the second spring, and Lextsubscript𝐿extL_{\mathrm{ext}} is the total distance between the two walls. For linear springs, the length of the spring in the equilibrium of the system is Leq=[kLeq+k(LextLeq′′)]/(k+k)subscript𝐿eqdelimited-[]𝑘superscriptsubscript𝐿eqsuperscript𝑘subscript𝐿extsuperscriptsubscript𝐿eq′′𝑘superscript𝑘L_{\mathrm{eq}}=\left[kL_{\mathrm{eq}}^{\prime}+k^{\prime}\left(L_{\mathrm{ext}}-L_{\mathrm{eq}}^{\prime\prime}\right)\right]/\left(k+k^{\prime}\right), where ksuperscript𝑘k^{\prime} and Leq′′superscriptsubscript𝐿eq′′L_{\mathrm{eq}}^{\prime\prime} are the spring constant and native equilibrium length of the second spring, respectively. The energy of the spring is (x)=12kx2𝑥12𝑘superscript𝑥2\mathcal{E}\left(x\right)=\frac{1}{2}kx^{2}, where we omit constant energy terms and use xLLeq𝑥𝐿subscript𝐿eqx\coloneqq L-L_{\mathrm{eq}} to denote deviations away from the equilibrium length of the spring in the system (not the spring in isolation).

Remark: voltage source and equilibrium.

The voltage bias vssubscript𝑣𝑠v_{s} does not appear in the effective current bias issubscript𝑖𝑠i_{s} of the junction. The capacitor isolates it in dc from the rest of the circuit and prevents it from establishing a dc current in any loop. In Sec. A9, we discuss further situation in which a bias has no effect on the equilibrium.

Remark: voltage source and dc loops.

A dc voltage source is absent from the dc-conducting superconducting loop formed by the junction and inductor, since its magnetic flux grows linearly in time, Φs=vstsubscriptΦ𝑠subscript𝑣𝑠𝑡\Phi_{s}=v_{s}t, and it would hence supply a current to the loop that grows infinitely large. At some time, this would break the abstraction of the ideal voltage source or quench the superconductivity.

A9 The biased Josephson system: simplifications

Refer to caption
Supplementary Figure S3: Schematic of five types of elementary flux-biased Josephson circuits. (a) Example of a fully linear circuit (no Josephson dipoles). A superconducting ring is frustrated by an external magnetic flux ΦextsubscriptΦext\Phi_{\mathrm{ext}} and, through a mutual inductive coupling, a current source issubscript𝑖𝑠i_{s}. (b) Example of system incorporating a Josephson dipole (here, a Josephson tunnel junction) in a dc-conducting distributed loop frustrated by ΦextsubscriptΦext\Phi_{\mathrm{ext}} and issubscript𝑖𝑠i_{s}. (c) Same as panel (b), but with a Superconducting Nonlinear Asymmetric Inductive eLement (SNAIL) instead of a junction in the loop. The SNAIL, a composite Josephson dipole, contains internal nodes and loops. Its internal loop is frustrated by the external magnetic flux ΦextsuperscriptsubscriptΦext\Phi_{\mathrm{ext}}^{\prime}. (d) Example of Josephson tunnel junction embedded in a distributed non-dc-conducting loop. In the region of the gap, we define the segment of the line-contour for the magnetic flux ΦextsubscriptΦext\Phi_{\mathrm{ext}} to be the a minimal-length one across the gap—i.e., in terms of electrical schematics, we take the flux across the effective capacitance associated with the gap in the loop. (e) Example similar to (d) but with a SNAIL element instead of tunnel junction.

In Sec. A8, we reviewed the equilibrium conditions of the Josephson system. Here, we develop several common situations in which the analysis of these conditions simplifies from a global to a local one or one that need not be done at all.

A frustrated loop not incorporating Josephson dipoles.

If an external bias sets up a persistent current involving only linear parts of the circuit [no Josephson dipoles; see Supplementary Figure S3(a)], the equilibrium of the circuit does not need to be calculated, since the dc current will not affect the eigenmodes of the circuit. Put another way, the alternating-current (ac) response of a linear sub-circuit is independent of its dc configuration.

An open loop incorporating a Josephson dipole without internal loops.

If a Josephson dipole is only a member of loops in the circuit that cannot not support a dc current [see Supplementary Figure S3(d)], any external frustration or bias of the loop is impotent. Thus, the equilibrium flux of the Josephson dipole in isolation ΦeqnatsuperscriptsubscriptΦeqnat\Phi_{\mathrm{eq}}^{\mathrm{nat}} can be taken as the equilibrium flux of the Josephson dipole in the circuit. In other words, we can treat the dipole locally in the steady-state analysis—i.e., we can forget about the rest of the distributed circuit in which it is embedded. This is one of the most common cases of practical interest, and is the one of the transmon and SQUID-based qubit.

An open loop incorporating a Josephson dipole with internal loops.

A Josephson dipole may have frustrated loops internal to its structure [see Supplementary Figure S3(e)]. For example, a lumped-element SQUID, RF-SQUID, or SNAIL is subjected to an external magnetic flux ΦextsuperscriptsubscriptΦext\Phi_{\mathrm{ext}}^{\prime}. However, since a Josephson dipole is purely inductive and if it is not part of a global conducting loop, the equilibrium conditions of its internal nodes are solved locally.

A dc-conducting loop incorporating a Josephson dipole with no internal loops.

If a simple Josephson dipole with no loops internal to it (such as the Josephson tunnel junction; not a SQUID or SNAIL), is embedded in a dc-conducting distributed and frustrated loop, see Supplementary Figure S3(b), then the flux of the dipole in the equilibrium of the total circuit cannot be determined locally. Rather, calculating the equilibrium requires incorporating information about the geometric-inductance response of the distributed loop at dc. Classical simulation methods can be used to obtain an equilibrium point.

A dc-conducting loop incorporating a Josephson dipole with internal loops.

In the most general case, a Josephson dipole is part of a dc-conducting loop; see Supplementary Figure S3(c). If the loop is not frustrated, then the equilibrium of flux of the dipole can be taken as the native equilibrium flux of the dipole in isolation. However, in practice, even if no frustration is intended, there may be spurious magnetic flux that will frustrate the loop. If the loop is frustrated by sources (which can include other Josephson dipoles able to source current, such as a biased SNAIL), the equilibrium condition of the global loop should be calculated accounting for the geometric inductances.

Voltage offset.

A voltage offset will have also have a frustrating effect on the eigenspectrum of the Hamiltonian. However, for the purposes of calculating the potential energy minima for use in the EPR linearization, it is sufficient to consider only the frustration due to currents.

Multiple equilibrium states.

The equilibrium conditions can yield multiple solutions that locally minimize the energy function indsubscriptind\mathcal{E}_{\mathrm{ind}}. It is preferable to choose a lowest-energy, stable equilibrium state for the operating point of the circuit. In principle, any of the minima can be used in the EPR method as the operational point; the physics of all other minima can be reconstructed from the operating point. However, we note that the presence of multiple minima can lead to phase slips (Matveev2002, ; Manucharyan2012, ; Pop2012, ; Masluk2012, ; Astafiev2012, ; Rastelli2013, ).

Supplementary Section B Nonlinear interactions, effective Hamiltonians, and the EPR

In this section, we find the amplitude of a general nonlinear interaction due to H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} explicitly. In Sec. B1, we find the effective interaction Hamiltonian of weakly non-linear systems in the dispersive regime to leading-order. In Sec. B2, to more systematically handle large systems, we introduce the EPR matrix 𝐏𝐏\mathrm{\mathbf{P}} and use it find the Kerr matrix and the vectors of anharmonicities and Lamb shifts. In Sec. B3, we find the general, normal-ordered form of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} and an expression to calculate the amplitude of any many-body interaction contained within. In Sec. B4, we use these results to describe the parametric activation of nonlinearities in a pumped Josephson circuit.

B1 Excitation-number-conserving interactions of weakly-nonlinear systems

Josephson circuits that are weakly non-linear and in the dispersive regime, such as the transmon-cavity and transmon-transmon circuits, have, in the absence of drives, dominant non-linear interactions that conserve excitation numbers.

Perturbative and dispersive.

For such weakly nonlinear systems with small zero-point fluctuations φmj1much-less-thansubscript𝜑𝑚𝑗1\varphi_{mj}\ll 1, for all modes m𝑚m and junctions j, the Hamiltonian H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} exerts only a perturbative effect on the eigenspectrum of H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}. Hence, we treat its effect order-by-order using the expansion H^nl=j=1Jp=3Ejcjpφ^jpsubscript^𝐻nlsuperscriptsubscript𝑗1𝐽superscriptsubscript𝑝3subscript𝐸𝑗subscript𝑐𝑗𝑝superscriptsubscript^𝜑𝑗𝑝\hat{H}_{\mathrm{nl}}=\sum_{j=1}^{J}\sum_{p=3}^{\infty}E_{j}c_{jp}\hat{\varphi}_{j}^{p}, obtained in Eq. (A.43). For this approach, we focus on the regime in which the energy difference between two modes is much larger than the phase excitation of the Josephson dipoles, (ωkωm)Ejcjp|φ^jp|much-greater-thanPlanck-constant-over-2-pisubscript𝜔𝑘subscript𝜔𝑚subscript𝐸𝑗subscript𝑐𝑗𝑝superscriptsubscript^𝜑𝑗𝑝\hbar\left(\omega_{k}-\omega_{m}\right)\gg E_{j}c_{jp}\left|\hat{\varphi}_{j}^{p}\right| for all k,m,j𝑘𝑚𝑗k,m,j, and p𝑝p.

Example: transmon coupled to a readout cavity mode.

Recall the simple example circuit quantized at the beginning of the main text. A qubit (q𝑞q) is coupled to a cavity (c𝑐c). To leading order in p𝑝p, the circuit Hamiltonian H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} is, using Eq. (A.43),

H^nlsubscript^𝐻nl\displaystyle\hat{H}_{\mathrm{nl}} =\displaystyle= EJ24(φca^c+φqa^q+H.c.)4+𝒪(φ^J6),\displaystyle-\frac{E_{J}}{24}\left(\varphi_{c}\hat{a}_{c}+\varphi_{q}\hat{a}_{q}+\mathrm{H.c.}\right)^{4}+\mathcal{O}\left(\hat{\varphi}_{J}^{6}\right)\;, (B.1)

where EJsubscript𝐸𝐽E_{J} and φ^Jsubscript^𝜑𝐽\hat{\varphi}_{J} are the Josephson energy and flux operator of the tunnel junction, and φqsubscript𝜑𝑞\varphi_{q} and φcsubscript𝜑𝑐\varphi_{c} are the qubit and cavity mode ZPF, respectively.

Expanding the multinomial.

Expanding the p=4𝑝4p=4 multinomial term in Eq. (B.1), we find a weighted sum of all possible four-body interactions between the qubit and cavity; example terms include a^q(a^q)3subscript^𝑎𝑞superscriptsuperscriptsubscript^𝑎𝑞3\hat{a}_{q}\left(\hat{a}_{q}^{\dagger}\right)^{3} and a^ca^ca^qa^qsubscript^𝑎𝑐superscriptsubscript^𝑎𝑐subscript^𝑎𝑞superscriptsubscript^𝑎𝑞\hat{a}_{c}\hat{a}_{c}^{\dagger}\hat{a}_{q}\hat{a}_{q}^{\dagger}. Using the commutation relations [Eq. (A.41)], we normal order the polynomial. For example, the qubit-only operators, this yields

a^q4+4a^qa^q3+6a^q2a^q2+4a^q3a^q+a^q4+6a^q2+12a^qa^q+6a^q2+3I^.superscriptsubscript^𝑎𝑞44superscriptsubscript^𝑎𝑞superscriptsubscript^𝑎𝑞36superscriptsubscript^𝑎𝑞absent2superscriptsubscript^𝑎𝑞24superscriptsubscript^𝑎𝑞absent3subscript^𝑎𝑞superscriptsubscript^𝑎𝑞absent46superscriptsubscript^𝑎𝑞212superscriptsubscript^𝑎𝑞subscript^𝑎𝑞6superscriptsubscript^𝑎𝑞absent23^𝐼\hat{a}_{q}^{4}+4\hat{a}_{q}^{\dagger}\hat{a}_{q}^{3}+6\hat{a}_{q}^{\dagger 2}\hat{a}_{q}^{2}+4\hat{a}_{q}^{\dagger 3}\hat{a}_{q}+\hat{a}_{q}^{\dagger 4}\\ +\text{$6\hat{a}_{q}^{2}+12\hat{a}_{q}^{\dagger}\hat{a}_{q}+6\hat{a}_{q}^{\dagger 2}$}+3\hat{I}\;.
Higher-order nonlinearity yields lower-order coupling.

Due to the non-commutativity of the operators, the normal-ordering procedure results in terms that have lower polynomial order than those in original unordered expression; these new terms include a^qa^qsuperscriptsubscript^𝑎𝑞subscript^𝑎𝑞\hat{a}_{q}^{\dagger}\hat{a}_{q} and 3I^3^𝐼3\hat{I}. Such quadratic terms (e.g., a^qa^qsuperscriptsubscript^𝑎𝑞subscript^𝑎𝑞\hat{a}_{q}^{\dagger}\hat{a}_{q}, a^ca^csuperscriptsubscript^𝑎𝑐subscript^𝑎𝑐\hat{a}_{c}^{\dagger}\hat{a}_{c}, and a^ca^qsuperscriptsubscript^𝑎𝑐subscript^𝑎𝑞\hat{a}_{c}^{\dagger}\hat{a}_{q}) dress the modes of H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}} in a linear manner—both renormalizing their frequencies and hybridizing them. These new linear couplings cannot be a-priori straightforwardly included in H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}, since they occur only as a non-classical consequence of the ZPF in the non-linearity; they do not appear in a classical treatment of the Josephson system, which lacks the non-commuting operator aspect. Hence, their amplitudes are determined by the quantum ZPF of the eigenmodes. Remark: A self-consistent approach to include these linear couplings in Hlinsubscript𝐻linH_{\mathrm{lin}} is possible under certain assumptions (Solgun2017, ).

Rotating-wave approximation in the dispersive regime.

To leading-order, under the rotating-wave approximation (RWA), only excitation-number conserving interactions in H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} (those with an equal number of raising and lowering operators in each mode) contribute. Thus, Eq. (B.1) reduces to the effective, dispersive qubit-cavity Hamiltonian

H¯^p=4=mΔma^ma^m12αma^m2a^m2nm12χmna^ma^ma^na^n,subscript^¯𝐻𝑝4subscript𝑚Planck-constant-over-2-pisubscriptΔ𝑚superscriptsubscript^𝑎𝑚subscript^𝑎𝑚12Planck-constant-over-2-pisubscript𝛼𝑚superscriptsubscript^𝑎𝑚absent2superscriptsubscript^𝑎𝑚2subscript𝑛𝑚12Planck-constant-over-2-pisubscript𝜒𝑚𝑛superscriptsubscript^𝑎𝑚subscript^𝑎𝑚superscriptsubscript^𝑎𝑛subscript^𝑎𝑛\begin{split}\hat{\overline{H}}_{p=\mathrm{4}}=\sum_{m}-\hbar\Delta_{m}\hat{a}_{m}^{\dagger}\hat{a}_{m}-\frac{1}{2}\hbar\alpha_{m}\hat{a}_{m}^{\dagger 2}\hat{a}_{m}^{2}\\ -\sum_{n\neq m}\frac{1}{2}\hbar\chi_{mn}\hat{a}_{m}^{\dagger}\hat{a}_{m}\hat{a}_{n}^{\dagger}\hat{a}_{n}\;,\end{split} (B.2)

where m{q,c}𝑚𝑞𝑐m\in\left\{q,c\right\} is the mode label, ΔmsubscriptΔ𝑚\Delta_{m} is the effective Lamb shift, αmsubscript𝛼𝑚\alpha_{m} is the anharmonicity, and χmnsubscript𝜒𝑚𝑛\chi_{mn} is the total cross-Kerr frequency shift induced between modes m𝑚m and n𝑛n. The bar over H𝐻H denotes the RWA; the subscript p=4𝑝4p=4 denotes the highest the power of the Taylor expansion included in the effective Hamiltonian. The excitation spectrum of this Hamiltonian is illustrated in Supplementary Figure S4. Remark: Using first-order time-independent perturbation theory in place of the RWA yields the same result as the RWA.

Hamiltonian parameters.

The Hamiltonian parameters are obtained from the normal ordering and using Eq. (A.54),

χmnsubscript𝜒𝑚𝑛\displaystyle\chi_{mn} =j=1J1Ejφmj2φnj2=j=1Jωmωn4Ejpmjpnj,absentsuperscriptsubscript𝑗1𝐽superscriptPlanck-constant-over-2-pi1subscript𝐸𝑗superscriptsubscript𝜑𝑚𝑗2superscriptsubscript𝜑𝑛𝑗2superscriptsubscript𝑗1𝐽Planck-constant-over-2-pisubscript𝜔𝑚subscript𝜔𝑛4subscript𝐸𝑗subscript𝑝𝑚𝑗subscript𝑝𝑛𝑗\displaystyle=\sum_{j=1}^{J}\hbar^{-1}E_{j}\varphi_{mj}^{2}\varphi_{nj}^{2}=\sum_{j=1}^{J}\frac{\hbar\omega_{m}\omega_{n}}{4E_{j}}p_{mj}p_{nj}\;, (B.3a)
ΔmsubscriptΔ𝑚\displaystyle\Delta_{m} =12n=1Mχmn,absent12superscriptsubscript𝑛1𝑀subscript𝜒𝑚𝑛\displaystyle=\frac{1}{2}\sum_{n=1}^{M}\chi_{mn}\;, (B.3b)
αmsubscript𝛼𝑚\displaystyle\alpha_{m} =12χmm.absent12subscript𝜒𝑚𝑚\displaystyle=\frac{1}{2}\chi_{mm}\;. (B.3c)
Constraints in the Hamiltonian.

First, the structure of the non-linearity imposes certain constraints. For example, χmn2αmαnsubscript𝜒𝑚𝑛2subscript𝛼𝑚subscript𝛼𝑛\chi_{mn}\leq 2\sqrt{\alpha_{m}\alpha_{n}}, where the equality holds only for the one-junction case, J=1𝐽1J=1. Second, the universal EPR properties impose a strict limit on physically realizable designs, subject to Eqs. (A.57)–(A.60).

Degeneracies.

So far, we implicitly assumed the system does not have accidental degeneracies, such as those that occur when the frequency of a mode is an integer multiple of that of another, ωkz×ωmsubscript𝜔𝑘𝑧subscript𝜔𝑚\omega_{k}\approx z\times\omega_{m}, where z𝑧z\in\mathbb{Z}. In the case of such a degeneracy, additional terms survive the RWA; for example, if ωc=3ωqsubscript𝜔𝑐3subscript𝜔𝑞\omega_{c}=3\omega_{q}, then the non-excitation conserving term a^q3a^csuperscriptsubscript^𝑎𝑞3superscriptsubscript^𝑎𝑐\hat{a}_{q}^{3}\hat{a}_{c}^{\dagger} survives the RWA.

EPR sign & example.

Due to the even-power nature of the nonlinear interactions, the EPR sign drops out altogether from Eq. (B.3). The interaction strength between the qubit and cavity modes only depends on the overlap int the EPRs of the two modes alone. The significance of this is curiously showcased by Device DT3, presented in Methods. The qubit eigenmodes of DT3 are the equally-hybridized symmetric (bright, B) and antisymmetric (dark, D) combinations of the two bare transmon qubit modes. Naively, using the analogy of atoms in a cavity, one reasons that the symmetric mode couples to the cavity (C), |χBC|0much-greater-thansubscript𝜒BC0\left|\chi_{\mathrm{BC}}\right|\gg 0, and the antisymmetric modes does not, |χDC|=0subscript𝜒DC0\left|\chi_{\mathrm{DC}}\right|=0, due to the out-of-phase oscillation of the junction current dipole—a cancelation effect. However, from the EPR expression, it is seen that, to the contrary, the signs are irrelevant—no such cancelation is possible. Rather, according to the EPR method, since both modes have equal participation in the two junctions (pD1=pD2=pB1=pB2subscript𝑝𝐷1subscript𝑝𝐷2subscript𝑝𝐵1subscript𝑝𝐵2p_{D1}=p_{D2}=p_{B1}=p_{B2}), the couplings to the cavity are essentially equal, χDCχBCsubscript𝜒DCsubscript𝜒BC\chi_{\mathrm{DC}}\approx\chi_{\mathrm{BC}}; indeed, this is observed in the experiment (see the corresponding data table in the Methods section).

Generalizing to many modes.

Equations (B.2) and (B.3) were derived for the example of a qubit-cavity circuit; however, they generalize straightforwardly to M𝑀M modes under the simple extension m{1,,M}𝑚1𝑀m\in\left\{1,\ldots,M\right\}.

Refer to caption
Supplementary Figure S4: Excitation spectrum (not-to-scale) of a transmon-cavity circuit, corresponding to the effective dispersive Hamiltonian H^lin+H¯^p=4subscript^𝐻linsubscript^¯𝐻𝑝4\hat{H}_{\mathrm{lin}}+\hat{\overline{H}}_{p=4}, see Eq. (B.2). (a) The cavity and qubit frequencies in the ground state are ωcsuperscriptsubscript𝜔𝑐\omega_{c}^{\prime} and ωqsuperscriptsubscript𝜔𝑞\omega_{q}^{\prime}, respectively. The qubit anharmonicity is αqsubscript𝛼𝑞\alpha_{q}, and the qubit-cavity dispersive cross-Kerr shift is χqcsubscript𝜒𝑞𝑐\chi_{qc}. (b) Zoom-in on the qubit spectrum in the vicinity of ωqsuperscriptsubscript𝜔𝑞\omega_{q}^{\prime}. Each photon in the resonator loads the qubit frequency down by χqcsubscript𝜒𝑞𝑐\chi_{qc}. (c) Zoom-in on the cavity spectrum in the vicinity of ωcsuperscriptsubscript𝜔𝑐\omega_{c}^{\prime}.

B2 EPR matrices and many-body interactions

To more easily and systematically handle large-scale circuits and higher-order non-linearities, we introduce the EPR matrix 𝐏𝐏\mathrm{\mathbf{P}}, comprising the M×J𝑀𝐽M\times J EPRs pmjsubscript𝑝𝑚𝑗p_{mj} as its entries,

𝐏(p11p1JpM1pMJ).𝐏subscript𝑝11subscript𝑝1𝐽subscript𝑝𝑀1subscript𝑝𝑀𝐽\mathrm{\mathbf{P}}\coloneqq\left(\begin{array}[]{ccc}p_{11}&\cdots&p_{1J}\\ \vdots&\ddots&\vdots\\ p_{M1}&\cdots&p_{MJ}\end{array}\right)\;. (B.4)

The cross-Kerr interaction amplitudes χmn(p)superscriptsubscript𝜒𝑚𝑛𝑝\chi_{mn}^{(p)} due to the p𝑝p-th order terms of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} [for p=4𝑝4p=4, see Eq. (B.3a)] can similarly be organized in an M×M𝑀𝑀M\times M matrix,

𝝌p(χ11(p)χ1M(p)χM1(p)χMM(p)).subscript𝝌𝑝superscriptsubscript𝜒11𝑝superscriptsubscript𝜒1𝑀𝑝superscriptsubscript𝜒𝑀1𝑝superscriptsubscript𝜒𝑀𝑀𝑝\bm{\chi}_{p}\coloneqq\left(\begin{array}[]{ccc}\chi_{11}^{(p)}&\cdots&\chi_{1M}^{(p)}\\ \vdots&\ddots&\vdots\\ \chi_{M1}^{(p)}&\cdots&\chi_{MM}^{(p)}\end{array}\right)\;. (B.5)

It follows from Eq. (B.3a) that the leading-order Kerr matrix is

𝝌4=4(𝛀𝐏)𝐄j1(𝛀𝐏)T,\boxed{\bm{\chi}_{4}=\frac{\hbar}{4}\left(\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}}\right)\mathrm{\mathbf{E}}_{j}^{-1}\left(\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}}\right)^{\mathrm{T}}\;,} (B.6)

where 𝛀𝛀\mathrm{\mathbf{\Omega}} is the diagonal eigenfrequency matrix and

𝐄j1(E11EJ1)superscriptsubscript𝐄𝑗1matrixsuperscriptsubscript𝐸11missing-subexpressionmissing-subexpressionmissing-subexpressionsuperscriptsubscript𝐸𝐽1\mathrm{\mathbf{E}}_{j}^{-1}\coloneqq\begin{pmatrix}E_{1}^{-1}\\ &\ddots\\ &&E_{J}^{-1}\end{pmatrix} (B.7)

is the diagonal matrix of inverse Josephson energies.

To leading order, p=4𝑝4p=4, the vector of anharmonicities is the diagonal of 𝝌4subscript𝝌4\bm{\chi}_{4},

𝜶4(α1(4),,αM(4))T=12diag𝝌4,subscript𝜶4superscriptsuperscriptsubscript𝛼14superscriptsubscript𝛼𝑀4T12diagsubscript𝝌4\bm{\alpha}_{4}\coloneqq\left(\alpha_{1}^{(4)},\ldots,\alpha_{M}^{(4)}\right)^{\mathrm{T}}=\frac{1}{2}\mathrm{diag}\,\bm{\chi}_{4}\;, (B.8)

and the Lamb shift of mode m𝑚m is the m𝑚m-th row sum of the Kerr matrix.

𝚫4(Δ1(4),,ΔM(4))T=12𝝌4𝟏M,subscript𝚫4superscriptsuperscriptsubscriptΔ14superscriptsubscriptΔ𝑀4T12subscript𝝌4subscript1𝑀\mathrm{\mathbf{\Delta}}_{4}\coloneqq\left(\Delta_{1}^{(4)},\ldots,\Delta_{M}^{(4)}\right)^{\mathrm{T}}=\frac{1}{2}\bm{\chi}_{4}\mathrm{\mathbf{1}}_{M}\;, (B.9)

where 𝟏Msubscript1𝑀\mathrm{\mathbf{1}}_{M} denotes a column vector of length M𝑀M with all entries equal to unity.

Dilution of the nonlinearity.

The dilution of the nonlinearity of the Josephson dipole elements among the eigenmodes is neatly expressed in Eq. (B.6). The Josephson dipole energies 𝐄𝐉1superscriptsubscript𝐄𝐉1\mathrm{\mathbf{E_{J}}}^{-1} are diluted by 𝛀𝐏𝛀𝐏\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}} through a congruence transform to the Kerr coefficients. While the eigenfrequencies 𝛀𝛀\mathrm{\mathbf{\Omega}} weighs the contribution to each mode, the participation matrix 𝐏𝐏\mathrm{\mathbf{P}} dictates the dilution of the junction energies and their nonlinearity among the modes.

Higher-order nonlinear corrections and dilution.

Using the results of Sec. B3, the p𝑝p-th other correction to the Kerr matrix is

𝝌p=cp(𝛀𝐏)𝐄j1𝝋totp4(𝛀𝐏)T,subscript𝝌𝑝Planck-constant-over-2-pisubscript𝑐𝑝𝛀𝐏superscriptsubscript𝐄𝑗1superscriptsubscript𝝋tot𝑝4superscript𝛀𝐏T\bm{\chi}_{p}=\hbar c_{p}\left(\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}}\right)\mathrm{\mathbf{E}}_{j}^{-1}\bm{\varphi}_{\mathrm{tot}}^{p-4}\left(\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}}\right)^{\mathrm{T}}\;, (B.10)

where 𝝋tot=Diag(φ1,tot,,φJ,tot)subscript𝝋totDiagsubscript𝜑1totsubscript𝜑𝐽tot\bm{\varphi}_{\mathrm{tot}}=\operatorname{Diag}\left(\varphi_{1,\mathrm{tot}},\ldots,\varphi_{J,\mathrm{tot}}\right) is the diagonal matrix of the total ZPF fluctuation of the Josephson dipole reduced fluxes, defined in Eq. (B.18). The Kerr matrix incorporating corrections due to all orders of the nonlinearity, see Eq. (B.24), is 𝝌p=3𝝌p𝝌superscriptsubscript𝑝3subscript𝝌𝑝\bm{\chi}\coloneqq\sum_{p=3}^{\infty}\bm{\chi}_{p}. The congruence-transformation form of Eq. (B.10) is identical to that of Eq. (B.6); it governs the dilution of the nonlinearity in the same manner, subject to 𝛀𝐏𝛀𝐏\mathrm{\mathbf{\Omega}}\mathrm{\mathbf{P}}.

Similarly to the results of Eqs. (B.8) and (B.9), the p𝑝p-th order corrections to the anharmonicity and Lamb-shift vectors are

𝜶p=12diag𝝌pand𝚫p=cp(𝛀𝐏)𝝋totp2𝟏M.formulae-sequencesubscript𝜶𝑝12diagsubscript𝝌𝑝andsubscript𝚫𝑝subscript𝑐𝑝𝛀𝐏superscriptsubscript𝝋tot𝑝2subscript1𝑀\bm{\alpha}_{p}=\frac{1}{2}\mathrm{diag}\,\bm{\chi}_{p}\quad\text{and}\quad\mathrm{\mathbf{\Delta}}_{p}=c_{p}\left(\bm{\Omega}\mathrm{\mathbf{P}}\right)\bm{\varphi}_{\mathrm{tot}}^{p-2}\bm{1}_{M}\;. (B.11)

In general, the Lamb shift correction depends on the total ZPF frustration of the junctions.

B3 General many-body interactions

So far, we explicitly calculated the leading-order correction on the spectrum of H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}} due to H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}. We expressed the eigenmode interactions using normal-ordered many-body terms, such as χa^qa^ca^qa^c𝜒superscriptsubscript^𝑎𝑞superscriptsubscript^𝑎𝑐subscript^𝑎𝑞subscript^𝑎𝑐\chi\hat{a}_{q}^{\dagger}\hat{a}_{c}^{\dagger}\hat{a}_{q}\hat{a}_{c}. Here, we extend the analysis and compute the amplitude of any general normal-ordered term in H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}.

The form a general many-body interaction.

A general normal-ordered many-body interaction has the form

C𝜶,𝜷pa^1β1a^MβMa^1α1a^MαMC𝜶,𝜷p𝒂^𝜷𝒂^𝜶,superscriptsubscript𝐶𝜶𝜷𝑝superscriptsubscript^𝑎1absentsubscript𝛽1superscriptsubscript^𝑎𝑀absentsubscript𝛽𝑀superscriptsubscript^𝑎1subscript𝛼1superscriptsubscript^𝑎𝑀subscript𝛼𝑀superscriptsubscript𝐶𝜶𝜷𝑝superscriptbold-^𝒂bold-†absent𝜷superscriptbold-^𝒂𝜶C_{\bm{\alpha},\bm{\beta}}^{p}\hat{a}_{1}^{\dagger\beta_{1}}\cdots\hat{a}_{M}^{\dagger\beta_{M}}\hat{a}_{1}^{\alpha_{1}}\cdots\hat{a}_{M}^{\alpha_{M}}\eqqcolon C_{\bm{\alpha},\bm{\beta}}^{p}\bm{\hat{a}^{\dagger\beta}}\bm{\hat{a}^{\alpha}}\;, (B.12)

where C𝜶,𝜷psuperscriptsubscript𝐶𝜶𝜷𝑝C_{\bm{\alpha},\bm{\beta}}^{p} is an energy-dimensioned amplitude, determined by the p𝑝p-th-order of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} [see Eqs. (A.43) and (A.48)], and the multi-index M𝑀M-tuples

𝜶(α1,,αM)and𝜷(β1,,βM)formulae-sequence𝜶subscript𝛼1subscript𝛼𝑀and𝜷subscript𝛽1subscript𝛽𝑀\bm{\alpha}\coloneqq\left(\alpha_{1},\ldots,\alpha_{M}\right)\quad\text{and}\quad\bm{\beta}\coloneqq\left(\beta_{1},\ldots,\beta_{M}\right) (B.13)

describe the distribution of annihilation and creation operators involved in the interaction among the M𝑀M modes, respectively. The entires of the multi-index tuples 𝜶𝜶\bm{\alpha} and 𝜷𝜷\bm{\beta} are non-negative integers, αm,βm0subscript𝛼𝑚subscript𝛽𝑚subscriptabsent0\alpha_{m},\beta_{m}\in\mathbb{Z}_{\geq 0}. The right-hand side of Eq. (B.12) introduces the multi-index shorthand notation for the interaction terms.

Multi-index shorthand and operator powers.

The total number of lowering and raising operators in the expression of Eq. (B.12) is equal to the 1-norm of 𝜶𝜶\bm{\alpha} and 𝜷𝜷\bm{\beta},

|𝜶|m=1Mαmand|𝜷|m=1Mβm,formulae-sequence𝜶superscriptsubscript𝑚1𝑀subscript𝛼𝑚and𝜷superscriptsubscript𝑚1𝑀subscript𝛽𝑚\left|\bm{\alpha}\right|\coloneqq\sum_{m=1}^{M}\alpha_{m}\quad\text{and}\quad\left|\bm{\beta}\right|\coloneqq\sum_{m=1}^{M}\beta_{m}\;, (B.14)

respectively. For a given power p𝑝p of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, the total number of resulting operators is bounded, |𝜶|+|𝜷|p𝜶𝜷𝑝\left|\bm{\alpha}\right|+\left|\bm{\beta}\right|\leq p.

Expanding the H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} multinomials.

To arrive at Eq. (B.12) from Eq. (A.43), we first group the sum of the annihilation operators for a Josephson dipole, see Eq. (A.48), and define

A^jsubscript^𝐴𝑗\displaystyle\hat{A}_{j} m=1Mφmja^m;absentsuperscriptsubscript𝑚1𝑀subscript𝜑𝑚𝑗subscript^𝑎𝑚\displaystyle\coloneqq\sum_{m=1}^{M}\varphi_{mj}\hat{a}_{m}\;; (B.15)
H^nlsubscript^𝐻nl\displaystyle\hat{H}_{\mathrm{nl}} =p=3j=1JEjcjp(A^j+A^j)p.absentsuperscriptsubscript𝑝3superscriptsubscript𝑗1𝐽subscript𝐸𝑗subscript𝑐𝑗𝑝superscriptsubscript^𝐴𝑗superscriptsubscript^𝐴𝑗𝑝\displaystyle=\sum_{p=3}^{\infty}\sum_{j=1}^{J}E_{j}c_{jp}\left(\hat{A}_{j}+\hat{A}_{j}^{\dagger}\right)^{p}\;. (B.16)

Importantly, the commutator [A^j,A^j]=m=1M|φmj|2subscript^𝐴𝑗superscriptsubscript^𝐴𝑗superscriptsubscript𝑚1𝑀superscriptsubscript𝜑𝑚𝑗2\left[\hat{A}_{j},\hat{A}_{j}^{\dagger}\right]=\sum_{m=1}^{M}\left|\varphi_{mj}\right|^{2} is scalar-valued, which allows us to use the normal-ordering non-commutative binomial theorem to expand the p𝑝p-th power term of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} (Cohen1966, ; Blasiak2005, ; Reagor2015, ). Using the non-commuting expansion,

(A^j+A^j)p=k=0floorp2i=0p2kp!2kk!i!(p2ki)!×(φj,tot2)k(A^j)i(A^j)p2ki,superscriptsubscript^𝐴𝑗superscriptsubscript^𝐴𝑗𝑝superscriptsubscript𝑘0floor𝑝2superscriptsubscript𝑖0𝑝2𝑘𝑝superscript2𝑘𝑘𝑖𝑝2𝑘𝑖superscriptsuperscriptsubscript𝜑𝑗tot2𝑘superscriptsuperscriptsubscript^𝐴𝑗𝑖superscriptsubscript^𝐴𝑗𝑝2𝑘𝑖\begin{split}\left(\hat{A}_{j}+\hat{A}_{j}^{\dagger}\right)^{p}=\sum_{k=0}^{\mathrm{floor}\frac{p}{2}}\sum_{i=0}^{p-2k}\frac{p!}{2^{k}k!i!(p-2k-i)!}\times\\ \left(\varphi_{j,\mathrm{tot}}^{2}\right)^{k}\left(\hat{A}_{j}^{\dagger}\right)^{i}\left(\hat{A}_{j}\right)^{p-2k-i}\;,\end{split} (B.17)

where floorp2floor𝑝2\mathrm{floor}\frac{p}{2} gives the greatest integer that is less than or equal to p2𝑝2\frac{p}{2} and the total variance of the ZPF of the j𝑗j-th dipole is

φj,tot2superscriptsubscript𝜑𝑗tot2absent\displaystyle\varphi_{j,\mathrm{tot}}^{2}\coloneqq [A^,A^]=2Ej1m=1Mpmjωm.^𝐴superscript^𝐴Planck-constant-over-2-pi2superscriptsubscript𝐸𝑗1superscriptsubscript𝑚1𝑀subscript𝑝𝑚𝑗subscript𝜔𝑚\displaystyle\left[\hat{A},\hat{A}^{\dagger}\right]=\frac{\hbar}{2}E_{j}^{-1}\sum_{m=1}^{M}p_{mj}\omega_{m}\;. (B.18)

Since A^jsubscript^𝐴𝑗\hat{A}_{j} is the sum of operators that commute, see Eq. (B.15), we can now expand the powers of A^jsubscript^𝐴𝑗\hat{A}_{j} using the classical multinomial theorem,

(A^j)isuperscriptsubscript^𝐴𝑗𝑖\displaystyle\left(\hat{A}_{j}\right)^{i} =\displaystyle= |𝜶|=i(|𝜶|𝜶)φmj𝜶𝒂^𝜶,subscript𝜶𝑖𝜶𝜶superscriptsubscript𝜑𝑚𝑗𝜶superscriptbold-^𝒂𝜶\displaystyle\sum_{\left|\bm{\alpha}\right|=i}\left(\begin{array}[]{c}\left|\bm{\alpha}\right|\\ \bm{\alpha}\end{array}\right)\varphi_{mj}^{\bm{\alpha}}\bm{\hat{a}}^{\bm{\alpha}}\;, (B.21)

where the multi-index shorthand (φmj)𝜶m=1Mφmjαmsuperscriptsubscript𝜑𝑚𝑗𝜶superscriptsubscriptproduct𝑚1𝑀superscriptsubscript𝜑𝑚𝑗subscript𝛼𝑚\left(\varphi_{mj}\right)^{\bm{\alpha}}\coloneqq\prod_{m=1}^{M}\varphi_{mj}^{\alpha_{m}}, the multinomial coefficient is

(|𝜶|𝜶)|𝜶|!α1!αM!,𝜶𝜶𝜶subscript𝛼1subscript𝛼𝑀\left(\begin{array}[]{c}\left|\bm{\alpha}\right|\\ \bm{\alpha}\end{array}\right)\coloneqq\frac{\left|\bm{\alpha}\right|!}{\alpha_{1}!\cdots\alpha_{M}!}\;, (B.22)

and the sum condition |𝜶|=i𝜶𝑖\left|\bm{\alpha}\right|=i means that the sum includes terms with all possible tuples 𝜶𝜶\bm{\alpha} such that their 1-norm has the value i𝑖i.

The general normal-ordered form.

Combining Eqs. (B.16)–(B.21), we find the general normal-ordered many-body form of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} to all orders in p𝑝p and without approximations,

H^nl=p=3k=0floorp2i=0p2k|𝜷|=i,|𝜶|=p2kiC𝜶,𝜷p𝒂^𝜷𝒂^𝜶.subscript^𝐻nlsuperscriptsubscript𝑝3superscriptsubscript𝑘0floor𝑝2superscriptsubscript𝑖0𝑝2𝑘subscript𝜷𝑖𝜶𝑝2𝑘𝑖superscriptsubscript𝐶𝜶𝜷𝑝superscriptbold-^𝒂bold-†absent𝜷superscriptbold-^𝒂𝜶\boxed{\begin{split}\hat{H}_{\mathrm{nl}}=\sum_{p=3}^{\infty}\sum_{k=0}^{\mathrm{floor}\frac{p}{2}}\sum_{i=0}^{p-2k}\sum_{\begin{subarray}{c}\left|\bm{\beta}\right|=i,\\ \left|\bm{\alpha}\right|=p-2k-i\end{subarray}}C_{\bm{\alpha},\bm{\beta}}^{p}\bm{\hat{a}^{\dagger\bm{\beta}}}\bm{\hat{a}}^{\bm{\alpha}}\;.\end{split}} (B.23)

The energy-dimensioned amplitude of an interaction due to the p𝑝p-th power of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} for a general many body mode-interaction term is the sum of the individual junction contributions,

C𝜶,𝜷pp!𝜶!𝜷!k!2kj=1JEjcjpφmj𝜶φmj𝜷φj,tot2k,\boxed{C_{\bm{\alpha},\bm{\beta}}^{p}\coloneqq\frac{p!}{\bm{\alpha}!\bm{\beta}!k!2^{k}}\,\sum_{j=1}^{J}E_{j}c_{jp}\varphi_{mj}^{\bm{\alpha}}\varphi_{mj}^{\bm{\beta}}\varphi_{j,\mathrm{tot}}^{2k}\;,} (B.24)

where k12(p|𝜶||𝜷|)𝑘12𝑝𝜶𝜷k\coloneqq\frac{1}{2}\left(p-\left|\bm{\alpha}\right|-\left|\bm{\beta}\right|\right).

Equation (B.24) provides the amplitude for any mode interaction. Its exact value is calculated using the EPR to obtain the ZPF φmjsubscript𝜑𝑚𝑗\varphi_{mj}, using Eq. (A.55). Thus, we have analytically fully constructed H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, and, individually, every term contained within, from the FE simulations through the EPR.

Use cases.

Equation (B.24) can be used to explicitly calculate higher-order corrections to an effective Hamiltonian; for example, see Eqs. (B.10) and (B.11). Using Eq. (B.24), we can calculate mode parameters even when the numerical diagonalization of H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} becomes intractable—an issue that occurs at even moderate number of modes. Moreover, Eq. (B.24) can be used to engineer pumped multi-photon drive processes and to activate non-RWA interactions (Mundhada2018, ), as discussed in Sec. B4.

B4 The driven Josephson system: parametrically-activated interactions

Refer to caption
Supplementary Figure S5: Depiction of four-wave mixing in a Josephson junction used to parametrically activate a bilinear interaction between two modes. Off-resonant pumping of the Josephson circuit at ωp1subscript𝜔𝑝1\omega_{p1} and ωp2subscript𝜔𝑝2\omega_{p2} with effective straights ξ1subscript𝜉1\xi_{1}and ξ2subscript𝜉2\xi_{2}, respectively, can activate a specific four-wave-mixing process present in H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}. The pump frequencies ωp1subscript𝜔𝑝1\omega_{p1} and ωp2subscript𝜔𝑝2\omega_{p2} could be degenerate. The activation condition is determined by the storage- and readout-mode frequencies ωcsubscript𝜔𝑐\omega_{c} and ωrsubscript𝜔𝑟\omega_{r}, respectively; their corresponding mode operators are a^csubscript^𝑎𝑐\hat{a}_{c} and a^qsubscript^𝑎𝑞\hat{a}_{q}. (a), (b) The activated interaction is a beam-splitter-like conversion process, containing exactly one mode annihilation and one creation operator. The resonance condition is determined by ωcωrsubscript𝜔𝑐subscript𝜔𝑟\omega_{c}-\omega_{r}. (c), (d) Two-mode squeezing interaction, contains exactly two creation operators. The resonance condition is set by ωc+ωrsubscript𝜔𝑐subscript𝜔𝑟\omega_{c}+\omega_{r}. In the first column, panels (a) and (c), the resonance condition is set by ωp1ωp2subscript𝜔𝑝1subscript𝜔𝑝2\omega_{p1}-\omega_{p2}, whereas in the second column, panels (b) and (d), it is set by ωp1+ωp2subscript𝜔𝑝1subscript𝜔𝑝2\omega_{p1}+\omega_{p2}. Note, for each of the four diagram, the conjugate process (all arrow directions flipped) is also activated by the pumps.

The Josephson system can be subjected to strong external drives used to parametrically activate or enhance nonlinear couplings. We illustrate the use of Equation (B.24) in this context by calculating the amplitude of a excitation-swapping beam-splitter interaction.

Example: parametrically-activated beam-splitter interaction.

Motivated by the setup of device WG1, we aim to parametrically activate a beam-splitter (BS) interaction between two non-resonant modes of a Josephson system; for example, such an interaction can be used as a Q-switch (Leghtas2015, ; Pfaff2017, ). The system has a high-quality, storage-cavity mode (c𝑐c) and another low-quality, readout-cavity mode (r𝑟r). The two modes are far detuned and obey the conditions outlined at the start of Sec. B1. From the system H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, we aim to obtain the effective BS Hamiltonian

H^eff=ga^ca^r+ga^ra^c,subscript^𝐻effPlanck-constant-over-2-pi𝑔superscriptsubscript^𝑎𝑐subscript^𝑎𝑟Planck-constant-over-2-pisuperscript𝑔superscriptsubscript^𝑎𝑟subscript^𝑎𝑐\hat{H}_{\mathrm{eff}}=\hbar g\hat{a}_{c}^{\dagger}\hat{a}_{r}+\hbar g^{*}\hat{a}_{r}^{\dagger}\hat{a}_{c}\;, (B.25)

where g𝑔g is the rate excitation exchange.

Interaction in the rotating frame.

In the rotating frame with respect to H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}, defined by the transform U^(t)exp[it(ωca^ca^c+ωra^ra^r)]^𝑈𝑡𝑖𝑡subscript𝜔𝑐superscriptsubscript^𝑎𝑐subscript^𝑎𝑐subscript𝜔𝑟superscriptsubscript^𝑎𝑟subscript^𝑎𝑟\hat{U}\left(t\right)\coloneqq\exp\left[it\left(\omega_{c}\hat{a}_{c}^{\dagger}\hat{a}_{c}+\omega_{r}\hat{a}_{r}^{\dagger}\hat{a}_{r}\right)\right], the mode operators a^msubscript^𝑎𝑚\hat{a}_{m} acquire a harmonic time dependence, a^ma^m(t)U^(t)a^mU^(t)=a^meiωmtmaps-tosubscript^𝑎𝑚subscript^𝑎𝑚𝑡superscript^𝑈𝑡subscript^𝑎𝑚^𝑈𝑡subscript^𝑎𝑚superscript𝑒𝑖subscript𝜔𝑚𝑡\hat{a}_{m}\mapsto\hat{a}_{m}\left(t\right)\coloneqq\hat{U}^{\dagger}\left(t\right)\hat{a}_{m}\hat{U}\left(t\right)=\hat{a}_{m}e^{-i\omega_{m}t}, where m{c,r}𝑚𝑐𝑟m\in\{c,r\}. While a term in H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}} of the form a^ca^rsuperscriptsubscript^𝑎𝑐subscript^𝑎𝑟\hat{a}_{c}^{\dagger}\hat{a}_{r} exists, this terms is non-stationary, C(1,0),(0,1)p=4eit(ωrωc)a^ca^r+H.c.superscriptsubscript𝐶1001𝑝4superscript𝑒𝑖𝑡subscript𝜔𝑟subscript𝜔𝑐superscriptsubscript^𝑎𝑐subscript^𝑎𝑟H.c.C_{\left(1,0\right),\left(0,1\right)}^{p=4}e^{-it\left(\omega_{r}-\omega_{c}\right)}\hat{a}_{c}^{\dagger}\hat{a}_{r}+\text{H.c.}\;and is eliminated in the RWA in deriving the effective Hamiltonian.

Parametric activation and interaction rate.

A third mode of the system, indexed by p𝑝p, is off-resonantly driven at frequencies ωPsubscript𝜔𝑃\omega_{P} and ωPsuperscriptsubscript𝜔𝑃\omega_{P}^{\prime} with amplitudes ϵ1subscriptitalic-ϵ1\epsilon_{1} and ϵ2subscriptitalic-ϵ2\epsilon_{2}, respectively; within the RWA, the drive Hamiltonian is

H^pϵpeiωPta^p+ϵpe+iωPta^p.subscript^𝐻𝑝subscriptitalic-ϵ𝑝superscript𝑒𝑖subscript𝜔𝑃𝑡subscript^𝑎𝑝superscriptsubscriptitalic-ϵ𝑝superscript𝑒𝑖subscript𝜔𝑃𝑡superscriptsubscript^𝑎𝑝\hat{H}_{p}\coloneqq\epsilon_{p}e^{-i\omega_{P}t}\hat{a}_{p}+\epsilon_{p}^{*}e^{+i\omega_{P}t}\hat{a}_{p}^{\dagger}\;. (B.26)

Consider the following term contained in H^nlsubscript^𝐻nl\hat{H}_{\mathrm{nl}}, see Eq. (B.23),

C(0,1,2),(1,0,0)p=4a^r(t)a^c(t)a^p2,superscriptsubscript𝐶012100𝑝4superscriptsubscript^𝑎𝑟𝑡subscript^𝑎𝑐𝑡superscriptsubscript^𝑎𝑝2C_{\left(0,1,2\right),\left(1,0,0\right)}^{p=4}\hat{a}_{r}^{\dagger}\left(t\right)\hat{a}_{c}\left(t\right)\hat{a}_{p}^{2}\;, (B.27)

where the mode label order is (r,c,p)𝑟𝑐𝑝\left(r,c,p\right). Moving into a displaced and rotating frame defined by D^(ξp)exp(ξpa^pξpa^p)exp(iωPta^pa^p)^𝐷subscript𝜉𝑝subscript𝜉𝑝superscriptsubscript^𝑎𝑝superscriptsubscript𝜉𝑝subscript^𝑎𝑝𝑖Planck-constant-over-2-pisubscript𝜔𝑃𝑡superscriptsubscript^𝑎𝑝subscript^𝑎𝑝\hat{D}\left(\xi_{p}\right)\coloneqq\exp\left(\xi_{p}\hat{a}_{p}^{\dagger}-\xi_{p}^{*}\hat{a}_{p}\right)\exp\left(-i\hbar\omega_{P}t\hat{a}_{p}^{\dagger}\hat{a}_{p}\right), where ξPϵ/(ωPωp)subscript𝜉𝑃italic-ϵsubscript𝜔𝑃subscript𝜔𝑝\xi_{P}\coloneqq\epsilon/\left(\omega_{P}-\omega_{p}\right), the p𝑝p mode operator is shifted and rotated, a^pa^p(t)D^(ξp)1a^pD^(ξp)=(a^p+ξp)eiωPtmaps-tosubscript^𝑎𝑝subscript^𝑎𝑝𝑡^𝐷superscriptsubscript𝜉𝑝1subscript^𝑎𝑝^𝐷subscript𝜉𝑝subscript^𝑎𝑝subscript𝜉𝑝superscript𝑒𝑖subscript𝜔𝑃𝑡\hat{a}_{p}\mapsto\hat{a}_{p}\left(t\right)\coloneqq\hat{D}\left(\xi_{p}\right)^{-1}\hat{a}_{p}\hat{D}\left(\xi_{p}\right)=\left(\hat{a}_{p}+\xi_{p}\right)e^{-i\omega_{P}t} (Leghtas2015, ; Touzard2019, ). In this frame, expanding Eq. (B.27) yields the suggestive interaction term

C(0,1,2),(1,0,0)p=4a^ra^cξp2eit(ωrωc2ωP),superscriptsubscript𝐶012100𝑝4superscriptsubscript^𝑎𝑟subscript^𝑎𝑐superscriptsubscript𝜉𝑝2superscript𝑒𝑖𝑡subscript𝜔𝑟subscript𝜔𝑐2subscript𝜔𝑃C_{\left(0,1,2\right),\left(1,0,0\right)}^{p=4}\hat{a}_{r}^{\dagger}\hat{a}_{c}\xi_{p}^{2}e^{-it\left(\omega_{r}-\omega_{c}-2\omega_{P}\right)}\;, (B.28)

which is resonantly activated when ωrωc2ωP=0subscript𝜔𝑟subscript𝜔𝑐2subscript𝜔𝑃0\omega_{r}-\omega_{c}-2\omega_{P}=0, see Supplementary Figure S5. Under this condition, this term survives the RWA when deriving the effective Hamiltonian. We find the effective BS rate by casting Eq. (B.28) in the canonical BS form, Eq. (B.25),

g=C(0,1,2),(1,0,0)p=4ξp2.Planck-constant-over-2-pi𝑔superscriptsubscript𝐶012100𝑝4superscriptsubscript𝜉𝑝2\hbar g=C_{\left(0,1,2\right),\left(1,0,0\right)}^{p=4}\xi_{p}^{2}\;. (B.29)
General parametrically-activated interaction.

The procedure illustrated with the BS example generalizes straightforwardly to the parametric activation of most other interactions and to finding their rates using Eq. (B.24). The procedure is useful for more complex and even cascaded processes (Mundhada2017, ; Mundhada2018, ) used in dissipation engineering (Kapit2017, ). It is reported that the limit of procedure is typically reached when ξPsubscript𝜉𝑃\xi_{P} approaches unity, and other activated nonlinear processes become non-negligible (Verney2019, ; Lescanne2019, ; Tripathi2019, ; Malekakhlagh2020, ).

Supplementary Section C Finite-element electromagnetic-analysis methodology

We detail the EPR methodology for the finite-element (FE) analysis of the Josephson circuit. In Sec. C1, we model a Josephson dipole in the FE simulation as a rectangular sheet with an inductive lumped-element boundary condition with inductance Ljsubscript𝐿𝑗L_{j}. In Secs. C2 and C3, we extract the EPR pmjsubscript𝑝𝑚𝑗p_{mj} and EPR signs smjsubscript𝑠𝑚𝑗s_{mj} from the result of an eigenanalysis simulation of the linearized Josephson circuit, corresponding to H^linsubscript^𝐻lin\hat{H}_{\mathrm{lin}}. This step completes the classical analysis part of the EPR method; from here, the Hamiltonian is fully specified, as described in Secs. A and B. The eigenresult also provides complete information on the dissipation and input-output coupling of the circuit, as described in Sec. D.

These steps, and the calculations detailed in this text, are automated by the open-source software package pyEPR 222See the pyEPR (pyEPR, ) code repository at http://github.com/zlatko-minev/pyEPR, which we offer to the community.

C1 Modeling the Josephson dipole

In the FE model of the Josephson circuit, we model a Josephson dipole as a simple rectangular sheet with a lumped-element boundary condition (Nigg2012, ), see Supplementary Figure S6. The sheet abstracts away the physical layout of the Josephson dipole and its wiring leads.

Idealization of the deep-sub-wavelength features.

This idealization of the Josephson dipole is justified in that its physical size is in the deep-sub-wavelength regime with respect to the eigenmodes of interest. For example, for the devices described in the Methods section, the separation between the Josephson dipole size and the mode wavelengths of interest is approximately five orders of magnitude. We hence treat a Josephson dipole in the FE model as lumped-element inductor with inductance Ljsubscript𝐿𝑗L_{j}, given by Eq. (A.13a); the linearization is taken with respect to the circuit equilibrium, see Sec. A8.

Electromagnetic model of the lumped inductance.

The j𝑗j-th Josephson dipole is modeled as a two-dimensional sheet Sjsubscript𝑆𝑗S_{j}, see Supplementary Figure S6. The sheet is assigned a surface-impedance boundary condition, which imposes E=Zs(n^×H)subscript𝐸parallel-tosubscript𝑍𝑠^𝑛subscript𝐻parallel-to\vec{E}_{\parallel}=Z_{s}(\hat{n}\times\vec{H}_{\parallel}) across the sheet, where Esubscript𝐸parallel-to\vec{E}_{\parallel} and Hsubscript𝐻parallel-to\vec{H}_{\parallel} are the tangential electric and magnetic fields of the sheet, respectively, n^^𝑛\hat{n} is the sheet normal vector, and Zssubscript𝑍𝑠Z_{s} is the complex-valued surface impedance corresponding to a total sheet inductance of Ljsubscript𝐿𝑗L_{j}. The hat symbol over n𝑛n denotes a unit vector in the context of electromagnetic fields; it is not to be confused with the hat notation used for quantum operators.

Reducing the model complexity: ignoring leads.

If the geometric inductance of the wire leads connecting a Josephson dipole to larger distributed surfaces (such as the pads of a transmon qubit) is negligible in comparison to Ljsubscript𝐿𝑗L_{j} and the wire leads are deeply sub-wavelength in size, then the wire leads can be ignored, as depicted in Supplementary Figure S6. If the wire leads have non-negligible inductance, then they can either be included in the simulation or their linear inductance can be included in the definition of the Josephson dipole. In practice, it is preferable to be able to abstract away the wire leads as much as possible, since their feature sizes are typically orders of magnitude smaller than other all other design features. The inclusion of such fine detail in the model can result in very large geometric aspect ratios, which in turn result in increased computational costs. However, while this finer detail is more representative of the physical design, we have found that it does not lead to noticeable corrections for typical cQED devices.

Refer to caption
Supplementary Figure S6: Illustration (not-to-scale) of the representation of a Josephson dipole in a finite-element (FE) simulation model. The two large grey rectangles on the edges of the illustration depict metal pads of a transmon qubit. The light-grey sheet Sjsubscript𝑆𝑗S_{j} in the center depicts the sheet used to model the Josephson dipole (and potentially its leads) in the FE model. The sheet is assigned a lumped-element inductive boundary condition, with inductance Ljsubscript𝐿𝑗L_{j} (sheet inductance symbolized by floating inductor-element symbol), corresponding to the Josephson dipole inductance with respect to the operating equilibrium point, see Sec. A9. The structural details of the Josephson dipole (location marked by brown cross) and its lead wires (brown wires connected to cross) can be abstracted away.
Performance tip: mesh operations.

The sheet Sjsubscript𝑆𝑗S_{j} is generally one of the smallest features of the FE model. Due to this small size but critical role, one can gently seed a higher level of mesh on Sjsubscript𝑆𝑗S_{j} to speed up the eigenanalysis. However, caution should be used to avoid seeding too heavy of an initial mesh, which can instead lead to poor convergence of the simulation. Convergence can be verified by plotting pmjsubscript𝑝𝑚𝑗p_{mj} as a function of the simulation adaptive pass number and by verifying the conditions detailed in Sec. A7.

C2 Calculating the EPR pmsubscript𝑝𝑚p_{m} in the case of a single Josephson dipole

If a Josephson circuit incorporates exactly one Josephson dipole, then we use the global electric and magnetic eigenmode energies to directly calculate the EPR pmsubscript𝑝𝑚p_{m} of the dipole in the mode.

Energy balance.

The time-averaged electromagnetic energy in a resonantly excited mode is equally split into an inductive indsubscriptind\mathcal{E}_{\mathrm{ind}} and capacitive capsubscriptcap\mathcal{E}_{\mathrm{cap}} contribution (Pozar, ). This detailed balance, ind=capsubscriptindsubscriptcap\mathcal{E}_{\mathrm{ind}}=\mathcal{E}_{\mathrm{cap}}, holds even in the presence of dissipation and defines the eigenmode condition. In the presence of a Josephson dipole, the inductive energy is split into a magnetic magsubscriptmag\mathcal{E}_{\mathrm{mag}} and a kinetic kinsubscriptkin\mathcal{E}_{\mathrm{kin}} contribution; ind=mag+kinsubscriptindsubscriptmagsubscriptkin\mathcal{E}_{\mathrm{ind}}=\mathcal{E}_{\mathrm{mag}}+\mathcal{E}_{\mathrm{kin}}. The magnetic contribution is associated with magnetic fields and geometric inductance. The kinetic contribution is associated with the Josephson dipole kinetic inductance and the flow of electrons, and their inertia. From the point of view of the FE analysis, the magnetic energy is stored in the magnetic eigenfields Hmsubscript𝐻𝑚\vec{H}_{m} and the kinetic energy is stored in the lumped-element boundary condition on Sjsubscript𝑆𝑗S_{j}. If lumped-element capacitive boundary conditions are absent from the model, then the capacitive eigenmode energy is stored entirely in electric eigenfields Emsubscript𝐸𝑚\vec{E}_{m}, cap=elecsubscriptcapsubscriptelec\mathcal{E}_{\mathrm{cap}}=\mathcal{E}_{\mathrm{elec}}; hence,

elec=cap=ind=mag+kin.subscriptelecsubscriptcapsubscriptindsubscriptmagsubscriptkin\mathcal{E}_{\mathrm{elec}}=\mathcal{E}_{\mathrm{cap}}=\mathcal{E}_{\mathrm{ind}}=\mathcal{E}_{\mathrm{mag}}+\mathcal{E}_{\mathrm{kin}}\;. (C.1)
Calculating EPR from the energy balance.

Using Eq. (C.1) and Eq. (5) of the main text, pm=kin/indsubscript𝑝𝑚subscriptkinsubscriptindp_{m}=\mathcal{E}_{\mathrm{kin}}/\mathcal{E}_{\mathrm{ind}}, the EPR is calculated from the ratio of global energy quantities,

pm=elecmagelec,\boxed{p_{m}=\frac{\mathcal{E}_{\mathrm{elec}}-\mathcal{E}_{\mathrm{mag}}}{\mathcal{E}_{\mathrm{elec}}}\;,} (C.2)

where the total magnetic- and electric-field energies are computed from the eigenfields phasors,

elecsubscriptelec\displaystyle\mathcal{E}_{\mathrm{elec}} =\displaystyle= 14ReVEmaxϵEmaxdv,14Resubscript𝑉superscriptsubscript𝐸maxitalic-ϵsubscript𝐸maxdifferential-d𝑣\displaystyle\frac{1}{4}\mathrm{Re}\int_{V}\vec{E}_{\text{max}}^{*}\overleftrightarrow{\epsilon}\vec{E}_{\text{max}}\,\mathrm{d}v\;, (C.3)
magsubscriptmag\displaystyle\mathcal{E}_{\mathrm{mag}} =\displaystyle= 14ReVHmaxμHmaxdv,14Resubscript𝑉superscriptsubscript𝐻max𝜇subscript𝐻maxdifferential-d𝑣\displaystyle\frac{1}{4}\mathrm{Re}\int_{V}\vec{H}_{\text{max}}^{*}\overleftrightarrow{\mu}\vec{H}_{\text{max}}\,\mathrm{d}v\;, (C.4)

where Emax(x,y,z)subscript𝐸max𝑥𝑦𝑧\vec{E}_{\mathrm{max}}(x,y,z) [resp., Hmax(x,y,z)subscript𝐻max𝑥𝑦𝑧\vec{H}_{\mathrm{max}}(x,y,z)] is the eigenmode electric (resp., magnetic) phasor, and ϵitalic-ϵ\overleftrightarrow{\epsilon} (resp., μ𝜇\overleftrightarrow{\mu}) denotes the electric-permittivity (resp., magnetic-permeability) tensor. The spatial integrals are performed over total volume V𝑉V of the device. The electric eigenfield is related to the phasor by

E(x,y,z,t)=Re[Emax(x,y,z)eiωmt].𝐸𝑥𝑦𝑧𝑡Redelimited-[]subscript𝐸max𝑥𝑦𝑧superscript𝑒𝑖subscript𝜔𝑚𝑡\vec{E}\left(x,y,z,t\right)=\mathrm{Re}\left[\vec{E}_{\mathrm{max}}\left(x,y,z\right)e^{i\omega_{m}t}\right]\;. (C.5)
Refer to caption
Supplementary Figure S7: Model of a transmon qubit overlaid with the surface-current eigendensity Jssubscript𝐽𝑠\vec{J}_{s}(r)𝑟\left(\vec{r}\right) of the qubit eigenmode, obtained using a FE simulation. Transmon pads depicted by the two large gray rectangles, separated by distance ljsubscript𝑙𝑗l_{j}. Small center rectangle represents the sheet model Sjsubscript𝑆𝑗S_{j} of the Josephson junction.

C3 Calculating the EPR pmjsubscript𝑝𝑚𝑗p_{mj} in the case of multiple Josephson dipoles

EPR pmjsubscript𝑝𝑚𝑗p_{mj} of the j𝑗j-th Josephson dipole.

In the case of multiple Josephson dipoles, the total kinetic energy kinsubscriptkin\mathcal{E}_{\mathrm{kin}} is itself split among the J𝐽J dipoles, see Eq. (A.20). We calculate the EPR pmjsubscript𝑝𝑚𝑗p_{mj} of junction j𝑗j in mode m𝑚m using the EPR definition, Eq. (A.50),

pmj=12LjImj/2ind,p_{mj}=\frac{1}{2}L_{j}I_{mj}{}^{2}/\mathcal{E}_{\mathrm{ind}}\;, (C.6)

where Imjsubscript𝐼𝑚𝑗I_{mj} is the peak value of the Josephson dipole current in mode m𝑚m. The current Imjsubscript𝐼𝑚𝑗I_{mj} is calculated from the integral of the mode surface-current density Js,m(x,y,z)subscript𝐽𝑠𝑚𝑥𝑦𝑧\vec{J}_{s,m}\left(x,y,z\right) over the dipole sheet Sjsubscript𝑆𝑗S_{j},

|Imj|=lj1Sj|Js,m|ds,subscript𝐼𝑚𝑗superscriptsubscript𝑙𝑗1subscriptsubscript𝑆𝑗subscript𝐽𝑠𝑚differential-d𝑠\left|I_{mj}\right|=l_{j}^{-1}\int_{S_{j}}\left|\vec{J}_{s,m}\right|\,\mathrm{d}s\;, (C.7)

where ljsubscript𝑙𝑗l_{j} is the length of the sheet, see Supplementary Figure S7.

EPR sign.

The EPR sign is calculated from the orientation of the current Imjsubscript𝐼𝑚𝑗I_{mj}. The absolute orientation is relative, as described in the main text, and is determined by defining a directed line DLjsubscriptDL𝑗\mathrm{DL}_{j} across Sjsubscript𝑆𝑗S_{j}. If the phasor Js,msubscript𝐽𝑠𝑚\vec{J}_{s,m} is aligned with DLjsubscriptDL𝑗\mathrm{DL}_{j} then smj=+1subscript𝑠𝑚𝑗1s_{mj}=+1; otherwise, smj=1subscript𝑠𝑚𝑗1s_{mj}=-1. Hence, we extract the sign using

smj=signDLjJs,mdl.subscript𝑠𝑚𝑗signsubscriptsubscriptDL𝑗subscript𝐽𝑠𝑚differential-d𝑙s_{mj}=\operatorname{sign}\int_{\mathrm{DL}_{j}}\vec{J}_{s,m}\cdot\mathrm{d}\vec{l}\;. (C.8)

The direction of the line merely establishes a convention for a positive EPR sign.

Enforcing energy balance.

The convergence of pmjsubscript𝑝𝑚𝑗p_{mj}, a quantity extracted from the local eigenfield solutions, can be enhanced by re-normalizing the set of mode EPR to ensure energy balance, see Eq. (C.1). If lumped-element capacitive boundary conditions are absent, then the EPR can be renormalized to ensure that their total sum is equal to the ratio j=1Jpmj=kin/indsuperscriptsubscript𝑗1𝐽subscript𝑝𝑚𝑗subscriptkinsubscriptind\sum_{j=1}^{J}p_{mj}=\mathcal{E}_{\mathrm{kin}}/\mathcal{E}_{\mathrm{ind}}, which is a globally calculated quantity, and is therefore expected to converge quicker.

The above calculations are automated by pyEPR (Note1, ).

C4 Remarks on the finite-element eigenmode approach

Finding the eigenfrequencies.

Rather than searching for the location of an unknown pole in an impedance response of the circuit and then honing in on it to perform finer sweeps, the eigenmode analysis returns the lowest M𝑀M modes above a minimum frequency of interest. This typically lifts the requirement for a prior knowledge of the mode frequencies.

Single simulation.

The FE eigenmode performs a single simulation from which complete information of the Josephson circuit is extracted. This is in contrast to the typical process flow used in an impedance analysis, which requires that mode frequencies be first identified so that a series of individual narrow-frequency-range impedance-response sweeps (one for each mode and junction) can be performed.

Closed-circuit optimization.

We have found that the above two feature (see remarks) can speed up the iterative refinement of a quantum design and can circumvents difficulties associated with finding and fitting unknown, narrow-line poles in an impedance analysis.

Supplementary Section D Dissipation budget and input-output coupling

In this section, we summarize the methodology used to fully characterize dissipation and input-output coupling in the Josephson system. Loss of energy results from material losses and radiative boundaries which guide energy away from the system. Additionally, control of the system is achieved by means of the radiative boundaries, such as input-output (I-O) coupling. The dissipation budget, comprising the individual loss-contribution bound of each lossy and radiative element, is extracted from the same eigensolution used to calculate the EPR pmjsubscript𝑝𝑚𝑗p_{mj}, in essentially the same way—by calculating the fraction of the m𝑚m-th mode energy stored in the l𝑙l-th lossy element—the lossy EPR pmlsubscript𝑝𝑚𝑙p_{ml}. While lossy EPR signs smlsubscript𝑠𝑚𝑙s_{ml} can also be calculated for each element, these are not needed for linear dissipation; however, their role in I-O coupling is detailed in Sec. D4.

D1 Dissipation budget

The dissipation budget comprises the estimated energy-loss-rate contributions due to each loss mechanism and each lossy object in the Josephson circuit (Pozar, ; Gao2008, ; Wang2009, ; Wenner2011, ; Zmuidzinas2012, ; Geerlings2013, ; Wang2015, ; Reagor2016, ; Gambetta2017a, ; McRae2020, ). We classily losses as capacitive, inductive, or radiative, and summarize their calculation here. Each loss mechanism is described by a corresponding lossy EPR pmlsubscript𝑝𝑚𝑙p_{ml} and an intrinsic quality factor Q𝑄Q. Assuming the dissipation is linear and the mode of interest is underdamped, the loss rates of each mechanism simply add; so, the total quality factor of the m𝑚m-th mode is (Zmuidzinas2012, ; Geerlings2013, ; Reagor2016, )

1Qtotal=1Qcap+1Qind+1Qrad,1subscript𝑄total1subscript𝑄cap1subscript𝑄ind1subscript𝑄rad\frac{1}{Q_{\text{total}}}=\frac{1}{Q_{\text{cap}}}+\frac{1}{Q_{\text{ind}}}+\frac{1}{Q_{\text{rad}}}\;, (D.1)

where Qcapsubscript𝑄capQ_{\mathrm{cap}} and Qindsubscript𝑄indQ_{\mathrm{ind}} are the total mode quality factors due to capacitive and inductive losses (i.e., those proportional to the intensity of the electric |E|2superscript𝐸2\left|\vec{E}\right|^{2} and magnetic field |H|2superscript𝐻2\left|\vec{H}\right|^{2}, respectively, see Secs. D2 and D3), and Qradsubscript𝑄radQ_{\mathrm{rad}} is the total radiative mode quality factor, see Sec. D4. In this section, the mode index m𝑚m is implicit.

Remark.

Beyond these intrinsic mechanisms, extrinsic factors, such as ionizing radiation, can play a significant role in understanding loss mechanisms, such as those due to quasiparticle in superconducting quantum circuits (Vepsalainen2020, ; Cardani2020, ).

D2 Capacitive loss

The total capacitive mode quality Qcapsubscript𝑄capQ_{\mathrm{cap}} is the weighted sum of intrinsic quality factors Qlcapsuperscriptsubscript𝑄𝑙capQ_{l}^{\text{cap}} (i.e., the inverse of the dielectric loss tangent) of all lossy dielectrics l𝑙l in the Josephson circuit (Zmuidzinas2012, ; Geerlings2013, ),

1Qcap=lpmlcapQlcap,1subscript𝑄capsubscript𝑙superscriptsubscript𝑝𝑚𝑙capsuperscriptsubscript𝑄𝑙cap\frac{1}{Q_{\text{cap}}}=\sum_{l}\frac{p_{ml}^{\text{cap}}}{Q_{l}^{\text{cap}}}\;, (D.2)

where pmlcapsuperscriptsubscript𝑝𝑚𝑙capp_{ml}^{\text{cap}} is the lossy energy-participation ratio of the l𝑙l-th dielectric in the mode—i.e., pmlcapsuperscriptsubscript𝑝𝑚𝑙capp_{ml}^{\text{cap}} is the fraction of capacitive energy stored in the dielectric element l𝑙l. We classify lossy capacitive elements as either bulk or surface. For example, the volume of three-dimension dielectric object, such as a chip substrate, is associated with bulk capacitive loss (Martinis2014, ; Dial2016, ; Kamal2016-anneal, ). On the other hand, the surface of the substrate, which may be a surface-dielectric layer is classified as a surface-loss element (Martinis2014, ; Wang2015, ). The lossy EPR for bulk capacitive loss is calculated from the eigenfield solutions,

pmlcap=1elec14ReVlEmaxϵEmaxdv,superscriptsubscript𝑝𝑚𝑙cap1subscriptelec14Resubscriptsubscript𝑉𝑙superscriptsubscript𝐸maxitalic-ϵsubscript𝐸maxdifferential-d𝑣p_{ml}^{\text{cap}}=\frac{1}{\mathcal{E}_{\mathrm{elec}}}\frac{1}{4}\operatorname{Re}\int_{V_{l}}\vec{E}_{\text{max}}^{*}\overleftrightarrow{\epsilon}\vec{E}_{\text{max}}\,\mathrm{d}v\;, (D.3)

where the integral is carried over the volume Vlsubscript𝑉𝑙V_{l} of the l𝑙l-th bulk dielectric object; the total electric energy elecsubscriptelec\mathcal{E}_{\mathrm{elec}} is defined in Eq. (C.3). The lossy EPR for a surface dielectric is approximated by

pmlcap,surf=1electlϵl4Resurfl|Emax|2ds,superscriptsubscript𝑝𝑚𝑙cap,surf1subscriptelecsubscript𝑡𝑙subscriptitalic-ϵ𝑙4Resubscriptsubscriptsurf𝑙superscriptsubscript𝐸max2differential-d𝑠p_{ml}^{\text{cap,surf}}=\frac{1}{\mathcal{E}_{\mathrm{elec}}}\frac{t_{l}\epsilon_{l}}{4}\operatorname{Re}\int_{\text{surf}_{l}}\left|\vec{E}_{\text{max}}\right|^{2}\,\mathrm{d}s\;, (D.4)

where the surface-dielectric layer thickness is tlsubscript𝑡𝑙t_{l} and its permittivity is ϵlsubscriptitalic-ϵ𝑙\epsilon_{l}.

D3 Inductive loss

Physically, inductive losses originate from the dissipative flow of electrical current in metals or through metal-metal seams. Supercurrent loss due to quasiparticles and vortices can be accounted for in an effective quality factor of the conducting surface. We denote the intrinsic inductive quality factor of a lossy object Qlindsuperscriptsubscript𝑄𝑙indQ_{l}^{\mathrm{ind}}. The bound on the total inductive-loss quality factor Qindsubscript𝑄indQ_{\text{ind}} of mode m𝑚m is a weighted sum of Qlindsuperscriptsubscript𝑄𝑙indQ_{l}^{\mathrm{ind}},

1Qind=lpmlindQlind,1subscript𝑄indsubscript𝑙superscriptsubscript𝑝𝑚𝑙indsuperscriptsubscript𝑄𝑙ind\frac{1}{Q_{\text{ind}}}=\sum_{l}{\frac{p_{ml}^{\mathrm{ind}}}{Q_{l}^{\mathrm{ind}}}}\;, (D.5)

where pmlindsuperscriptsubscript𝑝𝑚𝑙indp_{ml}^{\mathrm{ind}} is the lossy EPR for inductive element l𝑙l. We classify inductive losses as either surface, bulk, or seam.

Surface conductive loss (in the skin-depth).

The fraction of eigenmode energy stored in the skin depth λ0subscript𝜆0\lambda_{0} of metal surface l𝑙l, denoted surflsubscriptsurf𝑙\mathrm{surf}_{l}, if the lossy EPR of the surface, and is obtained from the eigenfield solutions,

pmlind,surf=1magλ0μl4Resurfl|Hmax,|2ds,p_{ml}^{\text{ind,surf}}=\frac{1}{\mathcal{E}_{\mathrm{mag}}}\frac{\lambda_{0}\mu_{l}}{4}\operatorname{Re}\int_{\text{surf}_{l}}\left|\vec{H}_{\text{max},\parallel}\right|^{2}\,\mathrm{d}s\;, (D.6)

where μlsubscript𝜇𝑙\mu_{l} is the magnetic permeability of the surface; typically, μl=μ0subscript𝜇𝑙subscript𝜇0\mu_{l}=\mu_{0}, and Hmax,\vec{H}_{\text{max},\parallel} is the magnetic field phasor parallel to the surface. For superconductors, pmlind,surfsuperscriptsubscript𝑝𝑚𝑙ind,surfp_{ml}^{\text{ind,surf}} is the kinetic inductance fraction (Gao2008, ; Zmuidzinas2012, ), commonly denoted α𝛼\alpha. The total magnetic energy magsubscriptmag\mathcal{E}_{\mathrm{mag}} is defined in Eq. (C.4).

Surface conductive loss: intrinsic quality.

In the normal state, a metal, such as copper or aluminum, has an intrinsic inductive quality factor of order unity (Pozar, )Qlind,surf1superscriptsubscript𝑄𝑙ind,surf1Q_{l}^{\text{ind,surf}}\approx 1. However, in the superconducting state, the lower bound on the quality factor is typically found to be in the range of several thousand (Reagor2013, )Qlind,surf103greater-than-or-equivalent-tosuperscriptsubscript𝑄𝑙ind,surfsuperscript103Q_{l}^{\text{ind,surf}}\gtrsim 10^{3}. The lower bound for thin-film superconducting aluminum has been measured to exceed Qlind,surf>105superscriptsubscript𝑄𝑙ind,surfsuperscript105Q_{l}^{\text{ind,surf}}>10^{5} (Minev2013, ).

Bulk magnetic loss.

The mode magnetic field can couple to bulk magnetic impurities present in the volume Vlsubscript𝑉𝑙V_{l} of lossy object l𝑙l. The lossy EPR for the bulk-inductive-loss mechanism of volume Vlsubscript𝑉𝑙V_{l} is

pmlind,bulk=1mag14ReVlHmaxμHmaxdv.superscriptsubscript𝑝𝑚𝑙indbulk1subscriptmag14Resubscriptsubscript𝑉𝑙superscriptsubscript𝐻max𝜇subscript𝐻maxdifferential-d𝑣p_{ml}^{\mathrm{ind,bulk}}=\frac{1}{\mathcal{E}_{\mathrm{mag}}}\frac{1}{4}\operatorname{Re}\int_{V_{l}}\vec{H}_{\text{max}}^{*}\overleftrightarrow{\mu}\vec{H}_{\text{max}}\,\mathrm{d}v\;. (D.7)

This coupling is typically negligible in current superconducting quantum circuits.

Seam loss.

The seam formed by pressing two metals together provides an electrical bridge for the current to flow across but also introduces a key dissipation mechanism (Brecht2015, ). A common example of a seam is the one formed by the two halves of the metal sample holders used to house a cQED chip. In the FE analysis, we model the seam by a line path seamlsubscriptseam𝑙\text{seam}_{l} tracing out the location of the seam at the surface of the two mating walls. The inductive lossy EPR participation for the seam is

pmlind,seam=1magλ0tlμl4Reseaml|Hmax,|2dl,superscriptsubscript𝑝𝑚𝑙ind,seam1subscriptmagsubscript𝜆0subscript𝑡𝑙subscript𝜇𝑙4Resubscriptsubscriptseam𝑙superscriptsubscript𝐻maxperpendicular-to2differential-d𝑙p_{ml}^{\text{ind,seam}}=\frac{1}{\mathcal{E}_{\mathrm{mag}}}\frac{\lambda_{0}t_{l}\mu_{l}}{4}\operatorname{Re}\int_{\text{seam}_{l}}\left|\vec{H}_{\text{max},\perp}\right|^{2}\,\mathrm{d}l\;, (D.8)

where the seam thickness is denoted tlsubscript𝑡𝑙t_{l}, its magnetic permeability μlsubscript𝜇𝑙\mu_{l}, and its the penetration depth λ0subscript𝜆0\lambda_{0}. It is convenient to rewrite the total seam loss due to seam l𝑙l in terms of an effective seam admittance gseamsubscript𝑔seamg_{\text{seam}}, defined in Ref. Brecht2015, ,

pmlind,seamQseam=1gseamseaml|Js×l|2dlωμ0all|Hmax|2dV.superscriptsubscript𝑝𝑚𝑙indseamsubscript𝑄seam1subscript𝑔seamsubscriptsubscriptseam𝑙superscriptsubscript𝐽𝑠𝑙2differential-d𝑙𝜔subscript𝜇0subscriptallsuperscriptsubscript𝐻max2differential-d𝑉\frac{p_{ml}^{\mathrm{ind,seam}}}{Q_{\text{seam}}}=\frac{1}{g_{\text{seam}}}\frac{\int_{\text{seam}_{l}}\left|\vec{J}_{s}\times\vec{l}\right|^{2}\,\mathrm{d}l}{\omega\mu_{0}\int_{\text{all}}\left|H_{\text{max}}\right|^{2}\,\mathrm{d}V}\;. (D.9)

D4 Radiative loss and input-output coupling

Refer to caption
Supplementary Figure S8: Schematic of a transmon-qubit circuit (mode operator a^^𝑎\hat{a}), comprising a Josephson tunnel junction (flux operator Φ^jsubscript^Φ𝑗\hat{\Phi}_{j}) and a capacitor, coupled to two input-output ports. The input and output fields of the left (resp., right) transmission line are a^in,1subscript^𝑎in1\hat{a}_{\mathrm{in},1} and a^out,1subscript^𝑎out1\hat{a}_{\mathrm{out},1} (resp., a^in,2subscript^𝑎in2\hat{a}_{\mathrm{in},2} and a^out,2subscript^𝑎out2\hat{a}_{\mathrm{out},2}). (a) Ports 1 and 2 both coupled to the top qubit node; hence, both port EPR signs smpsubscript𝑠𝑚𝑝s_{mp} are equal, sq1=sq2subscript𝑠𝑞1subscript𝑠𝑞2s_{q1}=s_{q2}. (b) Port 1 couples to the top node, but port 2 couples to the bottom qubit node. The port EPR signs smpsubscript𝑠𝑚𝑝s_{mp} have opposite signs, sq1=sq2subscript𝑠𝑞1subscript𝑠𝑞2s_{q1}=-s_{q2}.

The Josephson circuit incorporates radiative boundaries, typically purposefully introduced to serve as input-output ports of the circuit (Pozar, ), thus providing a means to perform system measurement and control. For cQED devices, the port exposes the circuit to an external transmission line conduit, such as a coaxial cable or a co-planar waveguide. The total radiative quality factor Qradsubscript𝑄radQ_{\text{rad}} of mode m𝑚m is the sum of the individual port contributions Qrad1=p=1PQmp1superscriptsubscript𝑄rad1superscriptsubscript𝑝1𝑃superscriptsubscript𝑄𝑚𝑝1Q_{\text{rad}}^{-1}=\sum_{p=1}^{P}Q_{mp}^{-1}, where P𝑃P is the total number of ports and Qmpsubscript𝑄𝑚𝑝Q_{mp} is the quality factor due to port p𝑝p. Below, we describe how port p𝑝p is modeled in the FE simulation to extract Qmpsubscript𝑄𝑚𝑝Q_{mp} from the eigensolutions.

Radiative energy loss.

Energy stored in mode m𝑚m can leak at a rate κmpsubscript𝜅𝑚𝑝\kappa_{mp} through port p𝑝p and be guided away by the transmission line. While for certain modes, such as readout ones, this coupling is desired, for other modes, such as a qubit one, this coupling is often considered spurious. In the case of a qubit mode, the energy loss to a readout port is seen as a manifestation of the Purcell effect (Houck2008-Purcell, ); while, for the readout mode of the same structure, the energy loss to the readout port sets the rate of information gain (Clerk2010, ). The rate κmpsubscript𝜅𝑚𝑝\kappa_{mp} is calculated from port EPR pmpsubscript𝑝𝑚𝑝p_{mp}, as detailed in the following for both wanted and spurious terms. The port EPR sign smpsubscript𝑠𝑚𝑝s_{mp} is calculated concurrently and is important for the system drive configuration.

FE model of the port.

In the presence of a port, the boundary of the Josephson circuit is somewhat ambiguous—we can include more or less of the port and conduit structure in the model. We choose to include a minimal but sufficiently large portion of these to faithfully model the disturbing effect of the boundary condition on the eigenmodes. For example, in the case of the qubit-cavity structure of Figure 1(a) of the main text, we include a short stub of the I-O coaxial cable in the FE model. The length of the stub can be determined from the effect of a sweep of its length on the target parameters. We have found that an alternative heuristic measure is to use the decay of the eigenfields inside the port structure and to make sure that the end of the port structure is at least several exponential decay lengths long (the field in the port structure decays exponentially since the eigenmodes are generally below its cutoff). The port structure termination surface Spsubscript𝑆𝑝S_{p} is treated as a resistive sheet with to model the effect of the transmission line guiding waves away from the Josephson circuit. The sheet effective resistance is Rpsubscript𝑅𝑝R_{p}. In the case of a 50 ΩΩ\Omega port line, Rp=50Ωsubscript𝑅𝑝50ΩR_{p}=50\,\Omega. While a resistive boundary condition can be assigned to the sheet, in the case of high-quality modes, it is possible to use a perturbative approach and to perform a lossless simulation of the FE model, from which the loss can be extracted (see also remark at end of this section).

Calculating the input-output (I-O) coupling rate κmpsubscript𝜅𝑚𝑝\kappa_{mp} from the port EPR pmpsubscript𝑝𝑚𝑝p_{mp}.

From the lossless eigensolutions of mode m𝑚m, the energy lost to the effective resistor Rpsubscript𝑅𝑝R_{p} of port p𝑝p during one mode oscillation period Tm=2π/ωmsubscript𝑇𝑚2𝜋subscript𝜔𝑚T_{m}=2\pi/\omega_{m} is

Pmp=12RpImp2Tm,subscript𝑃𝑚𝑝12subscript𝑅𝑝superscriptsubscript𝐼𝑚𝑝2subscript𝑇𝑚P_{mp}=\frac{1}{2}R_{p}I_{mp}^{2}T_{m}\;, (D.10)

where Impsubscript𝐼𝑚𝑝I_{mp} is the peak current across the port due to the excitation of mode m𝑚m; see Eq. (C.7). The total mode energy m(t)subscript𝑚𝑡\mathcal{E}_{m}\left(t\right) at time t𝑡t decays at rate

ddtm=pκmpm.𝑑𝑑𝑡subscript𝑚subscript𝑝subscript𝜅𝑚𝑝subscript𝑚\frac{d}{dt}\mathcal{E}_{m}=-\sum_{p}{\kappa_{mp}}\mathcal{E}_{m}\;. (D.11)

Hence, assuming a high quality mode, the energy loss during one oscillation period, between times t=0𝑡0t=0 and Tmsubscript𝑇𝑚T_{m}, is

m(Tm)=m(0)pTmκmpm(0).subscript𝑚subscript𝑇𝑚subscript𝑚0subscript𝑝subscript𝑇𝑚subscript𝜅𝑚𝑝subscript𝑚0\mathcal{E}_{m}(T_{m})=\mathcal{E}_{m}(0)-\sum_{p}{T_{m}\kappa_{mp}\mathcal{E}_{m}(0)}\;. (D.12)

This shows that the loss to port p𝑝p during one period is Pmp=Tmκmpm(0)subscript𝑃𝑚𝑝subscript𝑇𝑚subscript𝜅𝑚𝑝subscript𝑚0P_{mp}=T_{m}\kappa_{mp}\mathcal{E}_{m}(0), which we equate to the expression of Eq. (D.10) to find the I-O coupling rate in terms of quantities calculated from the eigenfields,

κmp=12RImp2m(0).subscript𝜅𝑚𝑝12𝑅superscriptsubscript𝐼𝑚𝑝2subscript𝑚0\kappa_{mp}=\frac{\frac{1}{2}RI_{mp}^{2}}{\mathcal{E}_{m}(0)}\;. (D.13)

Thus, the mode coupling quality factor is

Qmpωm/κmp=ωmm(0)12RImp2.subscript𝑄𝑚𝑝subscript𝜔𝑚subscript𝜅𝑚𝑝subscript𝜔𝑚subscript𝑚012𝑅superscriptsubscript𝐼𝑚𝑝2Q_{mp}\coloneqq\omega_{m}/\kappa_{mp}=\frac{\omega_{m}\mathcal{E}_{m}(0)}{\frac{1}{2}RI_{mp}^{2}}\;. (D.14)
Sign of the I-O participation.

If there are multiple ports, the port EPR sign smpsubscript𝑠𝑚𝑝s_{mp}, calculated using Eq. (C.8), is important in a manner similar to that explained for the case of the EPR sign smjsubscript𝑠𝑚𝑗s_{mj} in the case of multiple Josephson dipoles; see the main text. To illustrate, consider a simple transmon-qubit circuit coupled to two transmission lines in two different ways as depicted in the two panels of Supplementary Figure S8. While in both configuration the eigenmode frequency and quality factor is identical, the configurations are inequivalent. Consider driving both transmission lines with the same amplitude and in-phase, then the circuit of Supplementary Figure S8(a) is excited but that of Supplementary Figure S8(b) is not.

Remark on modeling the port termination as resistive vs. lossless.

The port sheet Spsubscript𝑆𝑝S_{p} can be treated as either resistive or lossless. In the former, the sheet is assigned a lumped-element boundary condition with impedance matching the port input impedance as seen from the system; typically designed to be Rp=50Ωsubscript𝑅𝑝50ΩR_{p}=50\,\Omega (Malekakhlagh2016-A2, ; Malekakhlagh2017-Cutoff-Free, ). The eigenresults fully account for the effect of the dissipation on the mode profile. These effects are negligible in the case of high-quality modes, but become significant as Qmsubscript𝑄𝑚Q_{m} approaches unity. For lossless treatment of Spsubscript𝑆𝑝S_{p}, suitable when Qm1much-greater-thansubscript𝑄𝑚1Q_{m}\gg 1, the termination simply assigned a perfectly conducting boundary condition. In both treatment, the eigenmode fields can be used to calculate the loss due to Rpsubscript𝑅𝑝R_{p}.

Reactive ports.

In addition to being resistive, ports can have a reactive component, typically occurring in the presence of non-idealities. For example, a port coupled to transmission line suffering from a down-line reflection will have some of the energy it leaks out to the line come back to it (Burkhart2020, ). The reactive part of the port structure thus houses energy in its internal modes. There can be treated by modifying the boundary condition on Spsubscript𝑆𝑝S_{p} or for a more general treatment can be accounted for by including the port, line, and scatterer structure in the FE model. This larger model will account for hybridization between the line modes and those of the system (Wang2019-cav-atten, ).

References