Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Decoherence Limits the Cost to Simulate an Anharmonic Oscillator

Tzula B. Propp,1 Sayonee Ray,2 John B. DeBrota,1 Tameem Albash,1, 3, 4 and Ivan Deutsch1, 4
1
Center for Quantum Information & Control, University of New Mexico, Albuquerque, NM 87131, USA
2
IonQ Inc, 4505 Campus Dr, College Park, MD 20740, USA
3
Department of Electrical and Computer Engineering,
University of New Mexico, Albuquerque, NM 87131, USA
4
Department of Physics and Astronomy, University of New Mexico, Albuquerque, NM 87131, USA
We study how decoherence increases the efficiency with which we can simulate the quantum
dynamics of an anharmonic oscillator, governed by the Kerr effect. As decoherence washes out the
fine-grained subPlanck structure associated with phase-space quantum interference in the closed
quantum system, the open quantum dynamics can be more efficiently simulated using a coarse-
grained finite-difference numerical integration. We tie this to the way in which decoherence recovers
the semiclassical truncated Wigner approximation (TWA), which strongly differs from the exact
closed-system dynamics at times when quantum interference leads to cat states and more general
superpositions of coherent states. The regression to semiclassical dynamics become more pronounced
as the initial amplitude of the oscillator grows, with implications for the quantum advantage that
might be accessible in noisy devices.

I. INTRODUCTION [27] is difficult when the quasiprobability function devel-


ops substantial negativity [28]. Decoherence, however,
The macroscopic world is largely described by classical washes out subPlanck structure and the associated neg-
statistical physics even though the underlying fundamen- ativity by introducing a time scale over which the co-
tal description is quantum mechanical. As emphasized herence of a quantum state is lost [29]. This transition
in the seminal work of Zurek [1–8], decoherence helps us occurs more rapidly as the system becomes more macro-
understand the transition from the quantum-to-classical scopic [8]. Thus, the phase space representation exhibits
world: classical states are more robust to the coupling to the trade off between robustness and complexity — more
the environment whereas highly nonclassical states, such subPlanck structure and negativity lead to faster deco-
as macroscopic superposition “cat states” [9], are fragile herence, while those features that are robust to decoher-
in the face of decoherence [8, 10–12]. ence are essentially classical and more efficiently simulat-
This observation not only explains why quantum ef- able.
fects are largely unobserved in the classical world, but it In this work we consider how the phase space represen-
also represents the fundamental challenge of large scale tation quantifies the robustness versus complexity trade
quantum information processing as decoherence limits off in the context of quantum simulation. Instead of fo-
the quantum complexity one can harness at the macro- cusing on negativity itself (which precludes semiclassical
scopic scale. While in principle one can tame decoherence sampling), we will study how decoherence simplifies nu-
through quantum error correction [13–15], we do not yet merical simulation by erasing fine-grained structure. We
have the means to do so fault-tolerantly [16–19] except consider a canonical toy model – an anharmonic oscillator
at very small scales [20]. such as a single bosonic mode evolving under a Kerr non-
Nevertheless, one may hope to achieve a meaningful linearity. While a trivial model, it exhibits a wide range
quantum advantage without fault tolerant error correc- of well-studied nonclassical features, such as squeezing at
tion. A key question, thus, is how much complexity one short times [30, 31], the collapse and revival of quantum
can expect to harness in a noisy-intermediate scale quan- oscillations [32], and the generation of superpositions of
tum (NISQ) [21] device given that decoherence largely coherent states [33–37]. These effects have been studied
washes out the most nonclassical features. in a variety of platforms including in trapped ions [38]
One way to quantify the quantum-to-classical transi- and circuit QED [39]. While there are closed form solu-
tion is using the phase space representation of quantum tions for this integrable model [34, 35], including for the
mechanics. The nonclassical features of quantum states open quantum system [31], in practice determining the
are reflected in the negativity of the Wigner quasiprob- expectation values of some observables can be numeri-
ability distribution [22–24]. This is associated with cally intensive when the mean number of bosons is deep
the fine-grained subPlanck-scale structure that arises in in the macroscopic regime. We emphasize that formally
quantum dynamics beyond what is described by classi- nothing is computationally complex in this model, as all
cal flow in phase space [8]. These properties challenge quantities can be extracted with algorithms that scale
the classical simulation of macroscopic quantum dynam- polynomially with the mean boson number. Nonetheless,
ics; direct numerical propagation of the quasiprobabil- there remains a tradeoff between robustness and ease of
ity function scales poorly when trying to capture ultra- numerical simulation. Our goal is to study how deco-
fine features [25], and applying techniques such as the herence makes this system more “classical” and in doing
methods of characteristics [26] or Monte Carlo sampling so determine which observables correspond to properties
2

that are fragile to decoherence, and which observables thus can be expanded as discrete Fourier transforms,
can be more efficiently simulated for the open quantum
N −1
system. 1 X 2πnk
fe(o) [n] = √ ce(o) [k]ei N , (3)
The remainder of this article is structured as follows. N k=0
In Sec. II we derive the dynamics of the quantum state N −1
under pure unitary evolution, both in discrete and con- 1 X 2πnk
ce(o) [k] = √ fe(o) [n]e−i N , (4)
tinuous variable descriptions. In Sec. III we expand on N n=0
the continuous variable dynamics to include open system
effects using the Fokker-Planck equations. In Sec. IV we where the coefficients ce(o) [k] implicitly depend on both
study the time evolution of moments of the Wigner func- M and N . It follows that the time evolution operator at

tion which capture fine-grained information and discuss times tM,N , Û (tM,N ) = e−i N n̂(n̂−1) , is a superposition
the cost of simulating the open quantum system. We of phase space rotation operators
conclude with a summary and outlook in Sec. V.
N −1
1 X (2k+1)π
Û (tM,N ) = √ ce [k]ei N n̂ , N even, (5)
N k=0
N −1
II. CLOSED SYSTEM DYNAMICS 1 X 2kπ
= √ co [k]ei N n̂ , N odd. (6)
N k=0
We consider an anharmonic oscillator for a single Therefore, at the discrete times tM,N , the state revives
bosonic mode with a Kerr nonlinearity, governed by the to coherent superposition of N copies of the initial state,
Hamiltonian distributed in phase space on the circle with radius |α0 | at
multiples of 2π/N . We can find the expansion coefficients
κ †2 2 κ
Ĥ = â â = n̂(n̂ − 1). (1) by noting that
2 2
N −1
1 X πM k2
Here and throughout ℏ = 1. This Hamiltonian arises, ce [k] = √ fe [n + M k]ei N
N n=0
e.g., in quantum optics in the presence of an intensity-
πM k2
dependent index of refraction (Kerr effect) [40], and in = ce [0]ei N , (7)
atom optics for bosonic atoms undergoing cold collisions N −1
[41]. For concreteness, we will consider the bosons here 1 X πM k(k+1)
co [k] = √ fo [n + M k]ei N
to be photons. The dynamics of this system has been N n=0
well studied, and we review the salient results here. πM k(k+1)
= co [0]ei N , (8)
As photon number is conserved here, the Heisen-
berg equation of motion for the annihilation operator where the second equality follows from the periodicity of

dâ/dt = −iκ↠ââ is integrable, â(t) = e−iκtâ â â(0). In fe(o) [n].
the classical (mean field) limit, â(t) → αc (t), αc (t) = We thus find for an initial coherent state |α0 ⟩. Using
2
e−iκt|αc | αc (0), describing the rotation of the phasor at Parseval’s theorem, |ce(o) [0]|2 = 1, and neglecting the
an angular rate proportional to the amplitude squared overall phase ce(o) [0], the state at the revival time tM,N
(the classical Kerr effect). An arbitrary pure state evolves is
according to the trivial solution  2
E
 √1 PN −1 ei πMNk α0 ei (2k+M N

N even,
N k=0

|ψ(tM,N )⟩ = N −1 i πM k(k+1) 2kπ
E
 √1 P
α0 ei N
k=0 e N odd.
X N
|ψ(t)⟩ = cn e−iκtn(n−1)/2 |n⟩ . (2) N

n=0 (9)

For the case M = 1, N = 2, we obtain |ψ(t1,2 )⟩ =


We will be particularly interested in the case that the √1 (|iα0 ⟩ + i |−iα0 ⟩), a “cat state.” For general N and
initial state is a coherent state |α0 ⟩, in which case cn = 2
|α0 |2 | √ M , the state is a superposition of N coherent states as
e− 2 α0n / n!. originally analyzed by Bialynicka-Birula [43], which we
The evolution of the state is periodic returning to its will denote “kitten states” 1 . These states form a dense
initial condition when κt = M 2π, for integer M . There
are additional revivals due to quantum interference at
times κtM,N = (M/N )2π, where M and N are co-prime 1 The kitten states and the cat state generated by the Kerr inter-
integers. To see this consider the discrete functions, action all have this remarkable feature: each phase is such that
M πn2 M πn(n−1)
fe [n] ≡ e−i N and fo [n] ≡ e−i N . These are pe- the overall normalization constant is independent of the coherent
riodic with fe(o) [n] = fe(o) [n + N ] for N even (odd) and state amplitude α0 .
3

Figure 1: The Wigner function for an initial coherent state with amplitude α0 = 4 at t = 0 evolving under the Kerr
interaction. At short times we see characteristic squeezing by at longer times we see negativity in the Wigner
function and the formation of intricate subPlanck-scale structure. At times κtM,N = (M/N )2π, M and N coprime,
the state is a superposition of N coherent states, Eq. (9). For N ≫ α0 , the coherent states are not distinguishable
and we observe structures like the short time “banana” with minimal negativity in the top row. At later times,
distinguishable superpositions form. These include high-order kitten states which may appear multiple times (e.g.
the N = 9, M = 2 kitten state in the center of the third row) as well as the Schrödinger cat state at t = π/κ
(bottom right corner). Right before a distinguishable kitten state forms, we observe shuriken-like Wigner functions
(bottom left corner). In between t = π/κ and t = 2π/κ, the evolution is symmetric to the evolution between t = 0
and t = π/κ as the state progresses through the same states in reverse (not pictured for brevity). An animation
depicting the continuous dynamics is given in [42].

set and exist at all tM,N , but only come to the fore when with a quadratic nonlinearity is
the coherent states are distinguishably separated.
∂W
= {HW , W }M P
We are particularly interested in studying the dynam- ∂t
ics of the system using phase space representations. The
 
∂HW ∂W ∂HW ∂W
equation of motion for the Wigner function, W (α, α∗ , t), = −i − (10)
∂α ∂α∗ ∂α∗ ∂α
 3
∂ HW ∂ 3 W ∂ 3 HW ∂ 3 W

i
− − ,
8 ∂ 2 α∗ ∂α ∂ 2 α∂α∗ ∂ 2 α∂α∗ ∂ 2 α∗ ∂α
4

where HW (α, α∗ ) = κ2 |α|2 (|α|2 − 2) + κ4 is the Weyl- to be normalized by n + m, e.g., {X̂ P̂ }sym = 12 (X̂ P̂ +
symbol of the Hamiltonian and {HW , W }M P is the Moyal P̂ X̂). Such correlation functions arise, e.g., in consid-
bracket [44] . The first line in Eq. (10) is {HW , W }P B , ering moments of a quadrature X̂θ = cos θX̂ + sin θP̂
the Poisson bracket, corresponding to the classical flow n
on phase space generated by HW . The second line repre- Roptically measuredn in homodyne detection, ⟨X̂θ ⟩(t) =
dXθ Pθ (Xθ , t)Xθ .
sents the nonclassical dynamics, which can lead to nega- For the closed system, the probability distribution
tivity of the Wigner function. These terms become non- which determines these moments ⟨X̂θn ⟩(t) is the square
negligible when the Wigner function develops subPlanck- of the wave function as a function of the quadrature Xθ
scale structure as this corresponds to a rapidly varying eigenvalue, which has an analytic form
function whose higher order derivatives are at least order
one. Neglecting these quantum dynamics is known as the P(Xθ , t) = |⟨Xθ |ψ(t)⟩|2 (13)
truncated Wigner approximation (TWA) [44], which well 2
|α0 |2 αn κtn2
√ 0 e−i( 2 +θ ) e−i 2
P∞ κt
describes the dynamics for times short times compared = e− 2 un (Xθ ) ,
n=0 n!
to the Ehrenfest time (that is, when corrections to the
Poisson bracket become nonnegligible). 2

Substituting HW for the Kerr Hamiltonian into where we have defined un (Xθ ) = An Hn (Xθ )e− 2 , with
Eq. (10), we have √ −1/2
An = ( π2n n!) and Hn (Xθ ) the nth Hermite poly-
nomial. The relevant number of terms to evaluate this
 
∂W ∂W ∂W
= −iκ(|α|2 − 1) α∗ ∗ − α formal expression grows as ∼ |α0 |4 . More generally, this
∂t ∂α ∂α
 3 3
 probability distribution can be obtained by marginalizing
iκ ∂ W ∗ ∂ W the Wigner function
− α 2 − α . (11)
4 ∂ α∂α∗ ∂ 2 α∗ ∂α Z
The first line is the TWA representing classical flow on P(Xθ , t) = dPθ W (Xθ , Pθ , t) (14)
phase space. Note −i(α∗ ∂α∗ − α∂α ) = X∂P − P ∂X is
the rotation operator on phase space,√where the quadra- where we define Pθ = Xθ+π/2 .
tures are defined by α = (X + iP )/ 2. The TWA has The higher order moments of the marginal capture
the expected form of a Kerr effect, that is, a rotation at the fine-grained structure in the Wigner function, and
angular rate κ|α|2 for |α| large. The evolution accord- generically this is hardest to simulate in the large |α0 |
ing to the TWA can be solved by the method of char- limit, when nonclassical subPlanck-scale structure devel-
acteristics, WTWA (α, α∗ , t) = WTWA (α(−t), α∗ (−t), 0), ops. However, as discussed above, this fine grained struc-
2
where α(t) = e−iκt(|α| −1) α is the classical flow. The ture is washed out by decoherence. Thus, our goal is to
TWA evolution of an initial coherent state represented determine how much of this complexity remains in the
as a Gaussian wavepacket is thus, WTWA (α, α∗ , t) = open quantum system and how the reduction of this com-
1 iκt(|α|2 −1)
π exp{−2|e α − α0 |2 }, as shown in the top row plexity leads to more efficient simulations.
of Fig. 2. For short times the nonlinear rotation “shears” The Kerr anharmonic oscillator is a useful system for
the distribution, leading to a squeezed Gaussian. At benchmarking the nonclassical dynamics and comparing
longer times the distribution becomes stretched and be- numerical simulations to the exact analytic solution. The
comes highly nonGaussian, but remains a positive prob- conservation of photon number implies normally ordered
ability distribution in the TWA. At these longer times correlation functions have a simple closed form for an
the corrections to the TWA become nonneglegible. Fig- initial coherent state,
ure 1 shows the Wigner function evolution for α0 = 4,
calculated using the numerical methods described in Ap- ⟨↵ âν ⟩(t) ≡ ⟨α0 |↵ (t)âν (t)|α0 ⟩, (15)
pendix A. For very short times, the TWA evolution ap-
proximately matches the exact evolution, but negativity where µ and ν are nonnegative integers. Using
κt
soon develops and revivals occur at the expected times, âν (t) |α0 ⟩ = α0ν e−i 2 ν(ν−1) α0 e−iνκt and ⟨α|β⟩ =
1 2 1 ∗ ∗
yielding kitten states and subPlanck structure. e− 2 |α−β| e− 2 (αβ −α β) one finds
While in principle the exact state is available in any κt
representation, in practice extracting measured values ⟨↵ âν ⟩(t) = (α0∗ )µ (α0 )ν ei 2 [µ(µ−1)−ν(ν−1)]
of observables becomes numerically intensive when α0 2
(1−cos[(µ−ν)κt]) i|α0 |2 sin[(µ−ν)κt]
× e−|α0 | e , (16)
is sufficiently large, which may describe realistic mean
photon numbers, e.g., ⟨n̂⟩ = |α0 |2 = 106 . We will specif- which tends towards zero exponentially with |α0 |2 except
ically focus on symmetrically-ordered correlation func- at special times when κt = M (µ−ν)π, with M an integer.
tions, which are moments of the Wigner function, Understanding quantum dynamics at these “recurrences”
will be the focus of Sec. IV.
Z
⟨{X̂ n P̂ m }sym ⟩(t) = dXdP X n P m W (X, P, t), (12) In contrast, no such simple form exists for direct calcu-
lation of correlation functions associated with the sym-
where the symmetrically-ordered product of X̂ and P̂ is metrically ordered expectation values. Symmetrically or-
5

dered expectation values are related to the normally or- The Wigner function for the cat state thus evolves un-
dered ones according to der the damping channel as

{↵ âν }sym = (17)


min(µ,ν) 1 γt
X µ!ν!
↵−γ âν−γ . W (X, P, t) = W0 (X, P − X0 e− 2 )
γ!(µ − γ)!(ν − γ)!2γ 2
γ=0 1 γt 2
+ W0 (X, P + X0 e− 2 ) + e−X0 γt sin(2XX0 )W0 (X, P ),
2
For very large order moments with large |α0 |, accurate
(22)
calculation of such sums becomes numerically intensive.

2 2 √
III. OPEN SYSTEM DYNAMICS where W0 (X, P ) = e−(X +P ) / π is the Wigner function
of the vacuum. The subPlanck-scale structure decays
We consider the simplest model of decoherence for the rapidly, at the rate X02 γ = 2 ⟨n̂⟩ γ. This can also be seen
single mode - the damped simple harmonic oscillator due directly from the phase space dynamics. Recall the Weyl
to photon loss, whereby the state evolves according to representation of the Lindbladian
the master equation
∂ ρ̂
= −i[Ĥ, ρ̂] + L [ρ̂], γ ∂2W
 
(18) γ ∂ ∂ ∗
∂t L [ρ̂]W = α+ ∗
α W+ (23)
2 ∂α ∂α 2 ∂α∂α∗
where γ

∂ ∂

γ ∂2W

∂2W

γ = X+ P W+ + ,
L [ρ̂] = − ↠âρ̂ + ρ̂↠â + γâρâ†

(19) 2 ∂X ∂P 4 ∂X 2 ∂P 2
2
is the Lindbladian describing photon loss in a zero tem-
perature reservoir. In the absence of Hamiltonian evo- which is a Fokker-Planck equation. The first order
lution,Pfor a general superposition of coherent states, derivative “drift” terms generate the decay of energy of
|ψ⟩ = k ck |αk ⟩ we have the closed form solution [40] the coherent state. The Laplacian diffusion terms lead
to the rapid washing out of subPlanck scale structure,
and thus decoherence. Indeed, we see the action of diffu-
γt
ED γt
eL t [|ψ⟩ ⟨ψ|] =
X
|ck |2 αk e− 2 αk e− 2
sion on the interference term in Eq. (22) gives, to lead-
k
∂2 ∂2
X 1−e−γt γt
ED γt ing order in X0 , γ4 ( ∂X 2 + ∂P 2 ) sin(2XX0 )W0 (X, P ) ≈
+ ck c∗k′ ⟨αk′ |αk ⟩ α k ′ e− 2 αk e− 2 . −X02 γ sin(2XX0 )W0 (X, P ), which shows that diffusion
k̸=k′
leads to decays of coherence with the same rate we ob-
(20) tained by different methods. The finer the subPlanck
scale structure, the faster decoherence washes it out [8] –
The amplitude of coherent states in the mixture decay as a manifestation of the robustness vs. complexity tradeoff.
expected, but the coherences in the superposition of co-
herent states decay more rapidly, at a rate that depends The equation of motion for the Wigner function under
on their overlap (distinguishability). In particular, for a the concurrent action of the Kerr Hamiltonian and pho-
∂t = {H, W }M B + L [ρ̂]W . While a
ton loss is given by ∂W
cat state, for t ≪ 1/γ,
formal solution exists for the Husimi representation [31],
1−e−γt γt 2
⟨−α0 |α0 ⟩ ≈ ⟨−α0 |α0 ⟩ = e−2|α0 | γt
, (21) no such solution exists for the Wigner function, and gen-
erally one must resort to numerical integration. As first
and this coherence decays at the rate 2|α0 |2 γ. For macro- studied by Stobinska et al. [37], given the invariance of
scopic α0 this implies that the coherence between the co- the Kerr interaction under rotation in phase space, this is
herent states is lost essentially instantaneously compared best done in polar coordinates by expressing the complex
to the time of substantial energy loss. amplitude α ≡ reiϕ , giving the equation of motion

∂2 1 ∂2
   
∂ ∂ κ 1 ∂ ∂
W (r, ϕ, t) = κ(r2 − 1) − + 2+ 2 2 W (r, ϕ, t)
∂t ∂ϕ 16 r ∂r ∂r r ∂ϕ ∂ϕ
∂2 1 ∂2
    
r ∂ γ 1 ∂
+ γ 1+ + + + 2 2 W (r, ϕ, t). (24)
2 ∂r 8 r ∂r ∂r2 r ∂ϕ
6

Figure 2: The Wigner functions of the closed system (top row) and open system (bottom two rows), evolving from
an initial coherent state with α0 = 4, are plotted at times corresponding to the first six-kitten state (first column),
the cat state (middle column), and the second six-kitten state (last column). In each inset, both the exact quantum
state and the state calculated using the truncated Wigner approximation (TWA) are shown on the left and right,
respectively. For nonzero coupling to the environment, the exact state more closely resembles that given by the
TWA, with the resemblance increasing with both decoherence strength γ (vertical axis) and time (horizontal axis).

The first line is the closed-system evolution, where the of the state in the presence of decoherence as seen in
first term is the expected classical rotation in phase phase Fig. 2. We study both of these quantitatively in the sec-
depending on the amplitude squared, r2 (the TWA evo- tion to follow.
lution), and the second term is the nonclassical Hamilto-
nian flow. The second line is the Fokker-Planck equation
where the first term is the damping (drift radially inward IV. QUADRATURE MOMENTS
to the origin), and the second term is diffusion.
We study in detail the behavior of expectation values
Stobinska et al., used a finite difference method which
of powers of the phase quadrature operator
discretized the Wigner funcion in phase space in order
to numerically integrate this partial differential equation n  
X n
(PDE), but was limited to small α0 as the required grid ⟨X̂θn ⟩(t)
= cosn−m θ sinm θ⟨{X̂ n−m P̂ m }sym ⟩(t).
size would otherwise grow too large. This is a reflection m=0
m
of the need to capture the fine grained structure that de- (25)
velops in the Wigner function. However, as noted, we ex- These describe statistics of homodyne measurements,
pect the fine-grained structure to be limited by diffusion with higher-order moments corresponding to finer-scale
(decoherence), and thus a coarse grained approximation features of the Wigner function, and with their evolu-
to the Wigner function should give a good approximation tion governed by the quantum dynamics. As discussed
for the open quantum system. Similarly, since decoher- above, the distinction between the quantum and clas-
ence is washing out the features generated by nonclassical sical dynamics is due to quantum interference between
flow, we expect the TWA to give a better representation different photons numbers. For an initial distribution of
7

field amplitudes, as for an initial coherent state, under of ⟨↵ âν ⟩ has periodicity (that is, recurrence) at times
classical dynamics the quadrature moments will collapse
as a function of time due to the nonlinear phase shifts, 2π
Tµν = . (26)
whereas under closed-system unitary dynamics these mo- |µ − ν|κ
ments also exhibit a series of revivals at recurrence times
due to quantum interference, as discussed above. Away Tµµ is undefined, which is consistent with the Kerr inter-
from these times, the expectation value is well-described action being number preserving; there is never a recur-
by the semiclassical dynamics given by the TWA [44]. rence for powers of the number operator because it does
not change during the evolution. Since the expectation
value of any normally ordered operator can be expressed
as a sum of expectation values of the form of Eq. (16), any
We seek to understand the deviation of the exact quan- normally ordered operator will also exhibit recurrences at
tum dynamics from the semiclassical dynamics given by these times. Furthermore, the period Tµν of an expecta-
the TWA. In this section, we first derive the existence tion value ↵ âν (t) is unchanged by operator ordering
(n)
and times tM,N of these recurrences for the nth phase- since the recurrence depends only on the difference µ − ν
quadrature moment. Then, we show why the deviations and this difference is preserved by the commutation rela-
from the TWA occur solely at times near these recur- tions. Because of this, a symmetrized expectation value
rences (with the effect becoming more prominent in the {↵ âν }sym (t) also has the same period Tµν .
large-α0 limit), despite the dense set of highly-quantum The dependence of the recurrence times on the differ-
kitten states throughout the evolution. Finally, we study ence µ − ν and not µ and ν separately is also true for the
the quantum behavior of ⟨X̂θn ⟩(t) itself in terms of its open quantum system governed by Eq. (18). The open
deviation from semiclassical dynamics. system recurrences also exhibit independence of operator
ordering, so that normally ordered and symmetrically or-
dered operators with the same difference µ − ν have the
The existence of recurrence times can be seen from same recurrence times. We derive these results below
the expression for normally ordered correlation functions with further details given in Appendix B.
for the closed system given an initial coherent state with The Lindblad master equation Eq. (18) expressed in
amplitude α0 in Eq. (16). We observe that the magnitude the number basis is given by

d γ κ  p
ρnm (t) = − (n + m) + i (n2 − n) − (m2 − m) ρnm (t) + γ (n + 1)(m + 1)ρn+1, m+1 (t). (27)
dt 2 2

We can write a formal solution for the density matrix Eq. (28) that
elements as a generalized discrete Fourier series X
↵ âν (t) = µν
Pnm ρnm (t)δn−m,µ−ν
X
−γ̃n′ m′ t nm
ρnm (t) = Anm
n′ m ′ e δn′ −m′ ,n−m , (28) X
µν −γ̃nm t
n′ ,m′ = Rnm e δn−m,µ−ν , (30)
nm

where we define a complex function µν


where Pnm is a positive real combinatorial factor and
where we have defined new generalized Fourier coeffi-
γ ′ κ cients
(n + m′ ) + i (n′2 − n′ ) − (m′2 − m′ ) .(29)

γ̃n′ m′ =
2 2 µν
X µν ′
,n′
Rnm = Pn−m+n′ ,n′ An−m+n
nm . (31)
For the closed system, γ̃n′ m′ is purely imaginary, and n′

the Fourier coefficients Anm n′ m′ are all zero except for From this form of the expectation value, we make the
n′ = m, m′ = m with Anm nm = ρnm (0). The sum thus col- following three conclusions:
lapses to a single term. For the open system, all Fourier
coefficients with n′ < n and m′ < m are zero, and the 1. The recurrence times of operators depend only on
coefficients Anm the difference, µ − ν, and not the powers µ and ν
n′ m′ can be solved recursively by truncating
the Fock space (for details see Appendix B). directly.
In all cases the form of Eq. (28) implies that each set 2. The recurrence times are unchanged by operator
of elements ρnm (t) of the density matrix on a diagonal ordering. For symmetricly ordered and normally
“stripe,” with n − m a fixed constant, evolves indepen- ordered expectation values the combinatorial factor
µν
dently. We can use this fact to study the nature of the Pnm will be different, but otherwise the expressions
recurrences in the open quantum system. It follows from for the expectation value in Eq. (30) are the same.
8

3. Open system effects will not change the recurrence with the minus sign occurring for N -even and M -odd
times as the correlation function depends on the only3 . This property follows from the considerations.
same frequency components that contribute in the Up to normalization and an overall phase, the effect
closed quantum system. of â on an N -kitten state is to permute the phases
of the coherent states in the superposition, forming
Thus, the periodicity of the expectation value is inde- a cyclic group (ZN ) [45], so that all phases return to
pendent of operator ordering, and of being for an open or their original values after N applications [46]. Because
closed system, up to rescaling of those coefficients by the of this property, these states were originally denoted
real part of γ̃nm and a change of the value of the com- “generalized coherent states.” [43]. The same is true for
binatorial prefactor; both of these leave the recurrence the action of ↠on ⟨ψ(tM,N )|. Only when â and ↠are
time unchanged.2 This behavior extends to any anhar- applied a number of times whose difference is an integer
monic, single-mode bosonic system at finite temperature multiple of N do the phases fail to average out to zero,
as we show in Appendix B. giving rise to recurrence at the times predicted by the
To understand why away from these recur- period formula in Eq. (26). The recurrence becomes
rences the system is well-described by the TWA, more sharply peaked at this time only in the large-α0
consider the normally ordered expectation value limit, when the coherent states are approximately
⟨ψ(tM,N )| ↵ âν |ψ(tM,N )⟩, where |ψ(tM,N )⟩ is the N orthogonal.
kitten state at time tM,N defined in Eq. (9). This expec- We derive these properties for general kitten states
tation value is only sensitive to the number of kittens N (not only those generated by the Kerr interaction) in Ap-
when µ − ν = p N with p a non-zero integer. To see this, pendix C. The behavior of the expectation value is given
note from its form that âN |ψ(tM,N )⟩ = ±α0N |ψ(tM,N )⟩, by

µ

 n̂
 µ = ν,
ν
⟨ψ(tM,N )| ↵ âν |ψ(tM,N )⟩ = α0∗µ−ν n̂ µ − ν = p N, p ∈ Z± ,
0 + O α α ei 2π
 
µ − ν = p N, p ∈
/ Z,
0 0 N

(32)

where Z± is the set of non-zero integers, and Z is the at) recurrences Tµν become exponentially suppressed as
set of all integers including zero. Note that, for normally the coherent states in the final case of Eq. (32) become in-
ordered expectation values of a mixture of N coherent creasingly orthogonal. From our earlier expression for the
states, the same expression holds except that the last case normally ordered expectation values in the closed system,
becomes exactly zero for all α0 ; it is the superposition Eq. (16), we also see that this suppression is exponential
2 2
states that create oscillations around the recurrence times in time, with a dependence of e−|α0 | (t−Tµν ) . The quan-
for finite α0 . tum deviations from the TWA become more suppressed
We thus find that the effects of quantum coherence, as as |α0 | increases, including the initial short-time devia-
seen in the correlation functions, depend solely on the dif- tion from the TWA [32, 47] and at the recurrence times,
ference µ − ν, which is preserved by operator reordering. which become increasingly sharp.
In particular, for the symmetrically ordered moments un- We now have the full set of tools needed to consider the
der consideration here, when this difference appropriately expectation value of a moment of a quadrature operator
aligns with recurrence times we expect strong deviations ⟨X̂θn ⟩. Any such operator is decomposable into a sum of
from the TWA. At other times, the lack of alignment symmetrically ordered operators
leads to cancellation making phase relationships approx- n
imately irrelevant and we expect the TWA to be a good
X
⟨X̂θn ⟩ = Ci,n,θ ⟨{â†n−i âi }sym ⟩, (33)
approximation. Furthermore, as α0 increases, quantum i
deviations from the TWA at times near (but not exactly
where the index i runs along the even integers for n-even,
and along the odd integers for n-odd. As we established
earlier, for every symmetrically ordered expectation value
2 Note that this argument fails for ν = µ since these depend on ⟨{↵ âν }sym ⟩, there is a set of recurrence times pTµν
the diagonal of the density matrix, where γ̃jk is completely real. (with p ∈ Z+ ) where the expectation value is expected
However, this is to be expected, as these expectation values are to deviate from semiclassical dynamics. For expectation
simply powers of the number operator and exhibit no periodic- values of the operator X̂θn given in Eq. (33), there are a
ity/recurrence in the first place. (n)
3 This is due to an overall rotation in the location of kitten states set of times tM,N where at least one of the terms in the
when N is even and M is odd. expansion of X̂θn will exhibit recurrence, leading to an
9

Figure 3: Deviation of the expectation value ⟨X̂θ6 ⟩(t) from its value calculated under the TWA is plotted for the
closed quantum system evolving under the Kerr interaction (that is, δ⟨X̂θ6 ⟩(t) ≡ ⟨X̂θ6 ⟩(t) − ⟨X̂θ6 ⟩TWA (t)). The initial
state is a coherent state with α0 = 6, and we normalize the expectation value by α06 . Different values of θ correspond
to different phase quadratures and give the same times of recurrence. The θ are chosen to correspond to angles
present in the 6 kitten state, which form at κt = π/3. Plots of the Wigner function at times κt = π/18, 2π/9, π/3,
and 2π/5 are included as insets. Notably, there is a dense set of times throughout evolution where the Wigner
function is a kitten state with highly quantum features, yet the TWA only fails to give the correct expectation value
(6)
for X̂θ6 near the beginning of evolution and near the first recurrence time t1,6 . At the times κt = 2π/9 and
κt = 2π/5, we do not observe a deviation from the TWA due to a mismatch between the symmetry of the
measurement operator X̂θ6 (which has the 6-fold symmetry of the cyclic group Z6 ) and the symmetries of the 5 and
9 kitten states (Z5 and Z9 , respectively). Near the recurrence
D time Eat κt = π/3, there are additional contributions

from high order kittens such that the terms of order O( α0 |α0 ei N ) in the last line of Eq. (32) become
non-negligible, which are exponentially suppressed in the large-α0 limit.

overall deviation from the TWA. These times are defined exhibit recurrence under the Kerr interaction4 . Note that
(n)
as the recurrence times that satisfy κtM,N = 2πM/N
where M, N ∈ Z+ and N is even(odd) for n even(odd),
and in both cases N ≤ n. The case with N = 0 comes
from terms with µ = ν, that is, operators that are propor- 4 Note that the number of recurrences between t = 0 and t = 2π κ
tional to powers of the number operator, and these never for an operator X̂θn can be compactly written in terms of the
Euler totient summary function, and thus scales quadratically: ∼
1
ξ(2)
n2 +O(nlogn), where ξ(2) = 2π 6
is the Riemann zeta function
evaluated at 2. In this way, the high spatial frequency features of
10

(n)
Figure 4: Averaged-deviation from the TWA (δ (t)) for α0 = 3 (left) and α0 = 6 (right) as defined in Eq. (35). In
both cases, the effect of increased decoherence is to suppress quantum deviations from the TWA. Increasing the
amplitude of the initial coherent state both magnifies the effects of decoherence and reduces the initial short-time
(n)
deviation from the TWA, and the oscillations become more sharply peaked around the recurrence times tM,N
[32, 47].

(n)
the set of recurrence times tM,N for any particular oper- which we normalize by |α0 |n so we can meaningfully com-
ator X̂θn is a strict subset of the times tM,N when kitten pare expectation values with differing α0 .
states form; for the latter tM,N , there is no restriction on As shown in Fig. 4, at times far from any recurrences
(n)
M and N other than they are positive integers and co- tM,N the behavior of the expectation values are well ap-
prime (hence, forming a dense set), while for the former proximated by the TWA (as predicted from the last two
(n) cases of Eq. (32)). This approximation becomes more
tM,N , we require N ≤ n and N to be even (odd) if n is
even (odd)5 . exact in the large-α0 limit, as seen with the increas-
Consider now the deviation of the exact evolution from ing sharpness of the recurrence with increasing α0 . In
that given by semiclassical TWA, this limit the coherent states in the kitten-states become
increasingly orthogonal, as detailed above, which also
δ⟨X̂θn ⟩(t) ≡ ⟨X̂θn ⟩(t) − ⟨X̂θn ⟩TWA (t). (34) serves to explain the decrease in initial deviations from
the TWA at short times as α0 increases, as has been de-
Whether the deviation from the TWA is maximal tailed elsewhere [32, 47]. In Fig. 4 we also observe that
(n)
precisely at tM,N depends on the particular phase- the TWA improves as γ increases, and for the same deco-
herence rate, the TWA improves for larger α0 as expected
quadrature as well as the choice of α0 (due to the fast
in the macroscopic limit.
α0 -dependent oscillation in Eq. (16)). We thus define
an average deviation from the TWA for our expectation Given the average deviation of the exact solution from
values of interest the TWA (Eq. (35)), we now define our quantitative
metric of interest. We numerically integrate Eq.(24) to
find the Wigner function and calculate expectation val-
Z
(n) 1 dθ
δ (t) ≡ δ⟨X̂θn ⟩(t)|, (35) ues based a finite difference method, as described in de-
|α0 |n 2π
tail in Appendix A. Given the symmetry, we decompose
the Wigner function into a Fourier series in the angular
variable and discretize the radial variable with resolution
the state characterized by higher-order moments are connected ∆r. The coarseness of this grid determines efficiency of
to high temporal frequency dynamics, exhibiting recurrence more the numerical integration. Thus, we define a cumulative
frequently. relative error induced by the phase space discretization
5 We do not require co-prime for the recurrence times as we did in the finite difference method,
for the Kitten times, as we do not need to avoid double counting.
Indeed, when there are multiple values of N and M correspond-
(n)
R (n) (n)
ing to the same tM,N , we observe a larger deviation from the τ
dt δ ∆r (t) − δ dr (t)
TWA at the recurrence time due to contributions from multiple ϵτ(n) (∆r) ≡ R (n)
. (36)
terms in the expansion in Eq. (33). τ
dt δ dr (t)
11

Figure 5: Error induced by finite grid effects on the calculation of ⟨X̂θ6 ⟩(t) measured relative to ∆rmin = 10−2 απ2 , for
0
α0 varying from 3 to 6, and averaging over 20 values of θ. We define the error in Eq. (36), and offset all data by
machine precision 10−24 . The maximum time tmax is chosen π
D Eto be 1.3 3κ so that, for an initial coherent state at
t = 0 = tmin , the full behavior of the first recurrence for X̂θ6 (t) illustrated in Fig. 3 is captured. We comment
that, for α0 = 3, the ratio of the error induced by discretization for the closed system to the largest γ simulation at
∆r = 0.1 is three orders of magnitude, whereas for α0 = 6 the same ratio also at ∆r = 0.1 is four orders of
magnitude.

Here dr is a suitably fine-grained radial grid size that well V. SUMMARY AND OUTLOOK
approximates the exact solution, τ is a time-window of
interest. This is plotted in Fig. 5 for the moment n = 6 In this work we have studied the effect of decoher-
with a variety of decoherence rates γ and a time window ence on the efficiency of simulating quantum dynamics
(6)
τ that includes the first recurrence t1,6 = π/(3κ). This in the case of an anharmonic Kerr oscillator, a standard
metric has the advantage that, if increasing the coarse- paradigm in quantum optics. As the decoherence facili-
(n)
grained grid size ∆r has the effect of over-estimating δ τ tates the quantum-to-classical transition and washes out
(n) quantum complexity, we expect more efficient represen-
at some times and under-estimating δ τ at other times,
tations are possible for the open quantum system. To
these effects will not cancel out.
quantify this, we studied quantum phase space dynam-
For the cumulative error, in all cases the error resulting ics represented by the Wigner function. Whereas the
from coarse-graining the grid is suppressed for larger de- closed-system quantum dynamics leads to fine-grained
coherence strengths. That this becomes more prominent subPlanck scale structure, decoherence acts to wash this
for larger α0 is made evident by carefully comparing the out, making the semiclassical description according to the
top and bottom plots in Fig. 5: the separation between TWA more accurate. This semiclassical dynamics is more
the lines corresponding to the closed and open system efficiently simulated by a coarse-grained finite-difference
increases with α0 at intermediate ∆r, especially for the integration of the Wigner function, when including the
smaller values of γ where the subPlanck-scale features are Fokker-Planck terms associated with decoherence.
not yet completely washed out. Since these fine-grained
For the specific cases of the Kerr oscillator we showed
features are generated more prominently for larger α0 , we
how semiclassical and quantum dynamics diverge due
expect the error to increase with α0 , which it indeed does
to quantum interference and the generation of superpo-
for the closed system. However, the error induced in cal-
sitions of coherent states. Without the fine structure
culations of the open system increases more slowly with
refeeding coherence to the system, evolution of the open
grid size ∆r. In the limit of asymptotically-strong deco-
system does not produce these high-order kitten states.
herence, we would expect the error to be completely flat
The result of this is a reduction of the quantum devia-
at intermediate ∆r since no subPlanck features arise.6

due to failure to capture even the semiclassical dynamics given


6 At sufficiently high ∆r there will always be discretization errors by the TWA, including the open system effect of energy loss.
12

tions from the TWA (Fig. 4) and a suppression of nu- nonlinear oscillators or qubits, and what is an efficient
merical errors induced by simulation of the system on a representation of the open quantum systems that would
coarser discrete grid for numerical integration (Fig. 5). allow for efficient classical simulations?
The tendency towards expectation value dynamics The current work gives some hints in this direction.
given by the TWA is well-known in the literature [47]. Phase space representations may provide for scalable ef-
Here, we have explained this in terms of the breaking of ficient simulations of open quantum systems when deco-
the Zn symmetry of the kitten states by decoherence con- herence is sufficiently large. Coarse graining and well-
current with Hamiltonian evolution. It remains an open chosen finite discretization, informed by semiclassical
question whether the source of this reduction in compu- dynamics is one potential method, as seen in recent
tational cost is a reduction in signal-to-noise ratio (such work [48]. Another potential method is mapping the
as is the case for depolarizing and dephasing noise mod- phase space dynamics to an underlying set of stochastic
els for qubits), or a nontrivial transition from quantum Langevin equations. While such a method cannot effi-
dynamics to classical dynamics. We conjecture that, for ciently capture all complex quantum dynamics, it may do
the anharmonic oscillator at zero temperature, the open so in the presence of decoherence [49]. In future work we
system is not a noisy simulation of the closed system but will study this to better understand when the quantum-
a different (and easier) problem entirely. to-classical transition is also a transition in computa-
Our work makes quantitative the intuition long estab- tional complexity.
lished in the early work on decoherence. Decoherence
reduces the complexity of the quantum state in phase
space, and it does so more prominently for more macro- VI. ACKNOWLEDGEMENTS
scopic initial states. This translates into more efficient
simulation of the open quantum system as the system size TzBP gratefully acknowledges helpful conversations
grows. While we have studied this for the toy problem with Jun Takahashi, Changhao Yi, and Chris Jackson.
of a single anharmonic oscillator, its implications for ob- This work was supported by National Science Founda-
taining a quantum advantage in NISQ devices is another tion Grant No. PHY-2116246, and is based upon work
open question. How does the rapidity of the regression to partially supported by the U.S. Department of Energy,
semiclassical dynamics with increasing system size gen- Office of Science, National Quantum Information Science
eralize for mulitpartite systems such as multiple coupled Research Centers, Quantum Systems Accelerator.

[1] W. H. Zurek, “Pointer basis of quantum apparatus: Into [10] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A.
what mixture does the wave packet collapse?” Phys. Rev. Fisher, A. Garg, and W. Zwerger, “Dynamics of the dis-
D 24, 1516–1525 (1981). sipative two-state system,” Reviews of Modern Physics
[2] W. H. Zurek, “Decoherence and the transition from quan- 59, 1–85 (1987).
tum to classical,” Physics Today 44, 36–44 (1991). [11] A. Caldeira and A. Leggett, “Quantum tunnelling in
[3] W. H. Zurek and J. P. Paz, “Decoherence, chaos, and a dissipative system,” Annals of Physics 149, 374–456
the second law,” Physical Review Letters 72, 2508–2511 (1983).
(1994). [12] M. S. Kim and V. Bužek, “Schrödinger-cat states at finite
[4] J. R. Anglin, J. P. Paz, and W. H. Zurek, “Decon- temperature: Influence of a finite-temperature heat bath
structing decoherence,” Physical Review A 55, 4041– on quantum interferences,” Physical Review A 46, 4239–
4053 (1997). 4251 (1992).
[5] W. H. Zurek, “Decoherence, chaos, quantum-classical [13] P. W. Shor, “Scheme for reducing decoherence in quan-
correspondence, and the algorithmic arrow of time,” tum computer memory,” Phys. Rev. A 52, R2493–R2496
Physica Scripta T76, 186 (1998). (1995).
[6] J. P. Paz and W. H. Zurek, “Quantum limit of de- [14] D. Gottesman, A. Kitaev, and J. Preskill, “Encoding
coherence: Environment induced superselection of en- a qubit in an oscillator,” Physical Review A 64 (2001),
ergy eigenstates,” Physical Review Letters 82, 5181–5185 10.1103/physreva.64.012310.
(1999). [15] V. V. Albert and P. Faist, “The error correction zoo,”
[7] Z. P. Karkuszewski, C. Jarzynski, and W. H. Zurek, https://errorcorrectionzoo.org/.
“Quantum chaotic environments, the butterfly effect, [16] P. W. Shor, “Fault-tolerant quantum computation,” in
and decoherence,” Physical Review Letters 89, 170405 Proceedings of 37th Conference on Foundations of Com-
(2002). puter Science (IEEE, 1996) pp. 56–65.
[8] W. H. Zurek, “Sub-planck structure in phase space and [17] D. Aharonov and M. Ben-Or, “Fault-tolerant quantum
its relevance for quantum decoherence,” Nature 412, computation with constant error,” in Proceedings of the
712–717 (2001). Twenty-Ninth Annual ACM Symposium on Theory of
[9] V. Dodonov, I. Malkin, and V. Man’ko, “Even and odd Computing, STOC ’97 (Association for Computing Ma-
coherent states and excitations of a singular oscillator,” chinery, New York, NY, USA, 1997) pp. 176–188.
Physica 72, 597–615 (1974). [18] A. Y. Kitaev, “Quantum computations: algorithms and
error correction,” Uspekhi Matematicheskikh Nauk 52,
13

53–112 (1997). EPL (Europhysics Letters) 94, 54002 (2011).


[19] E. Knill, R. Laflamme, and W. H. Zurek, “Resilient [39] A. F. Kockum and F. Nori, “Quantum bits with joseph-
quantum computation: error models and thresholds,” son junctions,” in Fundamentals and Frontiers of the
Proceedings of the Royal Society of London. Series A: Josephson Effect, edited by F. Tafuri (Springer Interna-
Mathematical, Physical and Engineering Sciences 454, tional Publishing, Cham, 2019) pp. 703–741.
365–384 (1998). [40] D. F. Walls and G. F. Milburn, Quantum Optics, 2nd ed.
[20] R. A. et al., “Suppressing quantum errors by scaling a (Springer, Berlin, Germany, 2008).
surface code logical qubit,” Nature 614, 676–681 (2023). [41] M. J. Steel, M. K. Olsen, L. I. Plimak, P. D. Drummond,
[21] J. Preskill, “Quantum computing in the NISQ era and S. M. Tan, M. J. Collett, D. F. Walls, and R. Graham,
beyond,” Quantum 2, 79 (2018). “Dynamical quantum noise in trapped bose-einstein con-
[22] K. E. Cahill and R. J. Glauber, “Density operators and densates,” Phys. Rev. A 58, 4824–4835 (1998).
quasiprobability distributions,” Physical Review 177, [42] Tz. B. Propp, “Video of the Wigner function evolv-
1882–1902 (1969). ing in a Kerr medium generated using the fock-
[23] R. Hudson, “When is the Wigner quasi-probability den- space method described in the first appendix for
sity non-negative?” Reports on Mathematical Physics 6, the open and (approximately) closed system. Here,
249–252 (1974). we depict the rotation counter-clockwise as is of-
[24] A. Kenfack and K. Życzkowski, “Negativity of the wigner ten done in the literature [38], as opposed to the
function as an indicator of non-classicality,” Journal of (physical) clockwise rotation depicted elsewhere in the
Optics B: Quantum and Semiclassical Optics 6, 396–404 main text.” https://github.com/tzpropp/KerrVideos/
(2004). blob/main/WignerMovieFull.gif (2023).
[25] E. J. Heller, “Wigner phase space method: Analysis [43] Z. Bialynicka-Birula, “Properties of the generalized co-
for semiclassical applications,” The Journal of Chemical herent state,” Phys. Rev. 173, 1207–1209 (1968).
Physics 65, 1289–1298 (1976). [44] A. Polkovnikov, “Phase space representation of quan-
[26] C. Weedbrook, S. Pirandola, R. Garcı́a-Patrón, N. J. tum dynamics. Lecture notes. Boulder summer school,”
Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaus- (2013).
sian quantum information,” Reviews of Modern Physics [45] D. Nash, A friendly introduction to group theory (2016).
84, 621–669 (2012). [46] Tz. B. Propp, “Video illustrating effect of the op-
[27] J. Sellier, M. Nedjalkov, and I. Dimov, “An introduc- erator â6 on the phases in a 6-kitten superposi-
tion to applied quantum mechanics in the Wigner Monte tion state.” https://github.com/tzpropp/KerrVideos/
Carlo formalism,” Physics Reports 577, 1–34 (2015). blob/main/PhaseGif.gif (2023).
[28] I. Welland and D. K. Ferry, “Wavepacket phase-space [47] A. Polkovnikov, “Quantum corrections to the dy-
quantum Monte Carlo method,” Journal of Computa- namics of interacting bosons: Beyond the truncated
tional Electronics 20, 267–273 (2020). Wigner approximation,” Physical Review A 68 (2003),
[29] S. Habib, K. Shizume, and W. H. Zurek, “Decoherence, 10.1103/PhysRevA.68.053604.
chaos, and the correspondence principle,” Physical Re- [48] M. Roda-Llordes, D. Candoli, P. T. Grochowski,
view Letters 80, 4361–4365 (1998). A. Riera-Campeny, T. Agrenius, J. J. Garcı́a-Ripoll,
[30] G. J. Milburn, “Quantum and classical Liouville dynam- C. Gonzalez-Ballestero, and O. Romero-Isart, “Numeri-
ics of the anharmonic oscillator,” Phys. Rev. A 33, 674– cal simulation of large-scale nonlinear open quantum me-
685 (1986). chanics,” (2023).
[31] G. J. Milburn and C. A. Holmes, “Dissipative quantum [49] P. Deuar, A. Ferrier, M. Matuszewski, G. Orso, and
and classical Liouville mechanics of the anharmonic os- M. H. Szymańska, “Fully quantum scalable description
cillator,” Physical Review Letters 56, 2237–2240 (1986). of driven-dissipative lattice models,” PRX Quantum 2,
[32] A. Polkovnikov, “Phase space representation of quantum 010319 (2021).
dynamics,” Annals of Physics 325, 1790–1852 (2010). [50] R. Bank, W. Coughran, W. Fichtner, E. Grosse, D. Rose,
[33] A. Miranowicz, R. Tanaś, and S. Kielich, “Generation and R. Smith, “Transient simulation of silicon devices
of discrete superpositions of coherent states in the an- and circuits,” IEEE Transactions on Computer-Aided
harmonic oscillator model,” Quantum Optics: Journal of Design of Integrated Circuits and Systems 4, 436–451
the European Optical Society Part B 2, 253–265 (1990). (1985).
[34] K. Tara, G. S. Agarwal, and S. Chaturvedi, “Produc- [51] Tz. B. Propp and S. J. van Enk, “Quantum networks
tion of Schrödinger macroscopic quantum-superposition for single photon detection,” Phys. Rev. A 100, 033836
states in a Kerr medium,” Phys. Rev. A 47, 5024–5029 (2019).
(1993).
[35] S. J. van Enk, “Entanglement capabilities in infinite di-
mensions: Multidimensional entangled coherent states,”
Phys. Rev. Lett. 91, 017902 (2003).
[36] S. J. van Enk, “Decoherence of multidimensional entan-
gled coherent states,” Phys. Rev. A 72, 022308 (2005). Appendix A: Summary of Numerical Approach
[37] M. Stobińska, G. J. Milburn, and K. Wódkiewicz,
“Wigner function evolution of quantum states in the pres-
ence of self-Kerr interaction,” Phys. Rev. A 78, 013810 Our aim is to numerically solve the (linear) partial dif-
(2008). ferential equation (PDE) for the Wigner function in a
[38] M. Stobińska, A. S. Villar, and G. Leuchs, “Generation dissipative medium with a self-Kerr interaction in cylin-
of Kerr non-Gaussian motional states of trapped ions,” drical coordinates [37]
14
  


 1  1 1
(r2 − 1)∂φ ∂r2 + 2 ∂φ2

∂τ W (τ, r, φ) = −  ∂r +  ∂φ

 | {z } 16  r
|{z} | {zr }

mean field rotation
Gaussian shearing non-Gaussian rotation


 

ξ ξ 2 1 
+ ξ + r∂r + ∂r2 + ∂r + 2 ∂φ2 W (τ, r, φ) , (A1)
2 } 8
| {z r r 

| {z 
}
drift diffusion

where we have defined dimensionless time τ = κt and where Ik (x) are the modified Bessel functions of the first
dimensionless decay ξ = κγ with boundary conditions kind. Because the coefficients of Eq. (A1) are indepen-
dent of φ and the Wigner function must be periodic in φ,
we choose to expand the Wigner function in a sine/cosine
2 −2|α0 |2 e−τ ξ
W (τ, 0, 0) = e , lim W (τ, r, φ) = 0 , series
π r→∞
(A2) ∞
X
and initial condition W (τ, r, φ) = a0 (τ, r) + (ak (τ, r) cos(kφ)
2 −2|α0 −reiφ |2 k=1
W (0, r, φ) = e . (A3) +bk (τ, r) sin(kφ)) . (A5)
π
Without loss of generality assuming α0 to be real, we The normalization condition can now be expressed as
have
Z 2π Z ∞ Z
2 −2(α20 +r2 ) 4αr cos φ
W (0, r, φ) = e e dφ dr rW (τ, r, φ) = 2π dr ra0 (τ, r) = 1 .
π" 0 0
∞ (A6)
#
2 −2(α20 +r2 ) X
= e I0 (4rα0 ) + 2 Ik (4rα) cos(kφ) , Plugging the expansion of the Wigner function into
π
k=1 Eq. (A1), we get a coupled system of ordinary differential
(A4) equations (ODE’s) to solve for each k value

   
1ξ ξ
∂τ a0 (τ, r) = ξ+ ∂r + ∂r2 a0 (τ, r) ,
r+ (A7)
4r2 8
k2
  
2 k 1 2
∂τ ak (τ, r) = k(r − 1) − ∂r + ∂r − 2 bk (τ, r)
16 r r
k2
    
ξ 1 ξ 2
+ ξ+ r+ ∂r + ∂r − 2 ak (τ, r) , (A8)
2 4r 8 r
k2
  
k 1
∂τ bk (τ, r) = −k(r2 − 1) + ∂r + ∂r2 − 2 ak (τ, r)
16 r r
k2
    
ξ 1 ξ
+ ξ+ r+ ∂r + ∂r2 − 2 bk (τ, r) . (A9)
2 4r 8 r

The equations for different k’s do not couple, but for each and the boundary condition at r → ∞ is given by
k the pair ak (τ, r) and bk (τ, r) are governed by coupled
ODE’s. The boundary condition at r = 0 is given by
lim a0 (τ, r) = 0 , (A11a)
r→∞
lim ak (τ, r) = 0 , k>0, (A11b)
r→∞
lim bk (τ, r) = 0 , k>0, (A11c)
r→∞
2 2 −τ ξ
a0 (τ, 0) = e−2|α0 | e , ak (τ, 0) = 0 , bk (τ, 0) = 0 ,
π
(A10) The initial condition is given by (again, assuming α0 is
15

real) We find that this provides a consistent expansion for the


functions at r = 0 even for τ > 0
1 −2(α20 +r2 ) π
Z
a0 (0, r) = e dφ e4rα0 cos(φ)
π2 −π
˜0 (τ ) + . . . ,
a0 (τ, r) = ã0 (τ ) + r2 ã (A14a)
2 2 2 ˜k rk+2 (τ ) + . . . ,
ak (τ, r) = ãk (τ )rk + ã (A14b)
= e−2(α0 +r ) I0 (4rα0 ) , (A12a)
π ˜
2 2 2
Z π bk (τ, r) = b̃k (τ )rk + b̃k rk+2 (τ ) + . . . . (A14c)
ak (0, r) = 2 e−2(α0 +r ) dφ e4rα0 cos(φ) cos(kφ)
π −π
4 2 2
The ODE’s are solved by discretizing the radial direc-
= e−2(α0 +r ) Ik (4rα0 ) , k > 0 , (A12b) tion into n equal steps
π
bk (0, r) = 0 , k > 0 . (A12c) rj = j∆r , j = 1, . . . n , (A15)
From the initial condition, we can find the leading be-
havior for the functions near r = 0 with ∆r = rmax /n, where rmax being some sufficiently
large r value. In our simulations, we take rmax = 2α0 .
˜0 + . . . ,
a0 (0, r) = ã0 + r2 ã (A13a) We will use the index j = 0 to denote the boundary
condition at r = 0. The derivatives discretized up to
˜k rk+2 + . . . .
ak (0, r) = ãk rk + ã (A13b) fourth order are given by

fj−2 (τ ) − 8fj−1 (τ ) + 8fj+1 (τ ) − fj+2 (τ )


∂r f (τ, r) → , (A16)
12∆r
−fj−2 (τ ) + 16fj−1 (τ ) − 30fj (τ ) + 16fj+1 (τ ) − fj+2 (τ )
∂r2 f (τ, r) → , (A17)
12∆r2

where fj (τ ) ≡ f (τ, rj ). Let us define the vector ⃗gk (τ ) = We use an absolute error tolerance of 10−9 and a relative
T
(ak,1 (τ ), . . . , ak,n (τ ), bk,1 (τ ), . . . bk,n (τ )) . We will then error tolerance of 10−6 . Our maximum temporal step size
write our ODE’s as a system of linear equations to be is chosen to be π/(100α02 ). We simulate pairs of (ak , bk )
solved up to a maximum k value of 60.
After solving for the time-evolved Wigner function
∂t⃗g0 (τ ) = A0⃗g0 (τ ) + ⃗b0 (τ ) , (A18a) up to some time τ , we can plot the Wigner func-
∂t⃗gk (τ ) = Ak⃗gk (τ ) , k > 0 . (A18b) tion in terms of (x, p) for fixed κt as was done in
the
P∞ main text by plotting the function a0 (τ, r) +
The vector ⃗b0 (τ ) arises from the discretization near r = 0: k=1 (ak (τ, r) cos(kφ) + bk (τ, r) sin(kφ)). Our φ grid is
when terms such as a0,0 (τ ) appear when evaluating the discretized in steps of π/200.
ODE at j = 1 and j = 2, we must replace them with the We can also calculate the expectation value of an ob-
initial condition servable  as (we drop the time dependence for now)
Z ∞ Z ∞
2 −2α2
a0,0 (τ ) = e , (A19) ⟨Â⟩ = dx dp W (x, p)Ã(x, p) , (A21)
π −∞ −∞

which then represent constants in the system of equa- where Ã(x, p) is the Weyl transform of the operator Ã
tions. (Recall that ak,0 (τ ) = 0 so this term does not
appear for k > 0.) The terms ak,−1 (τ ) and bk,−1 (τ ) that Z
y y
arise at j = 1 must be replaced by Ã(x, p) = dye−ipy/ℏ ⟨x + |Ã|x − ⟩
2 2
Z
ak,−1 (τ ) = (−1)k ak,1 (τ ) , bk,−1 (τ ) = (−1)k bk,1 (τ ) . u u
= dueixu/ℏ ⟨p + |Ã|p − ⟩ . (A22)
(A20) 2 2
This choice is to enforce the even/odd behavior from
In particular, the expectation value of arbitrary powers
Eq. (A14). We note that this term only appears from the
of X̂ and P̂ is given by
term 2ξ r∂r , and it is cancelled otherwise. We enforce the
Z ∞ Z ∞
boundary condition at infinity by setting to zero terms
beyond rmax , i.e. ak,n+1 (τ ) = ak,n+2 (τ ) = bk,n+1 (τ ) = ⟨X̂ m P̂ n ⟩ = dx dp xm pn W (x, p) , (A23)
−∞ −∞
bk,n+2 (τ ) = 0.
To solve the system of linear equations in Eq. (A18), Using our solution for the Wigner function in polar co-
we use TR-BDF2 [50], an implicit Runge-Kutta method. ordinates, this takes the form
16

∞ Z
X 2π Z ∞
m n m n
⟨X̂ P̂ ⟩ = dφ cos φ sin φ cos(kφ) dr rm+n+1 ak (r)
k=0 0 0
Z 2π Z ∞ 
m n m+n+1
+ dφ cos φ sin φ sin(kφ) dr r bk (r) . (A24)
0 0

The integral over φ can be calculated (numerically) by noting that

Z 2π m n   
2π X X m n
dφ cosm φ sinn φ cos(kφ) = (−1)kn (δm+n+k−2km −2kn ,0 + δm+n−k−2km −2kn ,0 ) ,
0 2m+n+1 in km kn
km =0 kn =0
(A25a)
Z 2π m n   
2π X X m n
dφ cosm φ sinn φ sin(kφ) = (−1)kn (δm+n+k−2km −2kn ,0 − δm+n−k−2km −2kn ,0 ) .
0 2m+n+1 in+1 km kn
km =0 kn =0
(A25b)

In this form, it is clear that


R ∞we need only to perform Appendix B: Derivation of Recurrence Time
radial integrals of the form 0 dr rp+1 ak (r), bk (r). Formulae and Zero-Temperature Fock-Space
Solution for the Open Quantum System

In Eq. (28), we observe that the density matrix for


a bosonic mode with an anharmonic, number-preserving
Hamiltonian has the formal solution
X (j,k)
ρjk (t) = Aj ′ k′ e−γ̃j′ k′ t δj ′ −k′ ,j−k , (B1)
j ′ ,k′

where γ̃j ′ k′ is a complex coefficient specific to each den-


sity matrix element, depending on both the Hamilto-
nian operator and Lindbladian superoperator’s elements
in the Fock basis. Such an equation is a solution to the
family of equations

d X  (n) + (n) −

ρjk (t) = −γ̃jk ρjk (t) + gjk ρj+n,k+n (t) + gjk ρj−n,k−n (t) , (B2)
dt n

which are generated by the Lindblad master equation trix separately, as depicted in Fig. 6 for a small number
with multiple arbitrary n-photon-loss/gain channels, de- of basis elements.
(n) ± If we specialize to the Kerr Hamiltonian defined in
scribed by a coupling gjk respectively. Recast in this
way, Eq. (B2) has the form of a set of coupled systems Eq. (1) and zero-temperature with amplitude damping
where population can leak out either to other systems or (that is, single photon loss only) giving rise to the Lind-
to an environment.7 Using techniques for such systems
[51], we can solve each diagonal stripe of the density ma-
the evolution of a single stripe of the density matrix, as depicted
in Fig. 6, with respect to an anharmonic Hamiltonian and open
system effects as energy moving between simple harmonic oscil-
7 Explicitly, we work with a time domain generalization of Eq. (28) lators which are coupled to each other and to an unmonitored
in [51], making a reindexing substitution ρjk → ci , which recasts environment.
17

where we reindexed j → j ′ − k ′ + j and contracted the


Kronecker product.
Note that it is the sums defining the Fourier coef-
ficients in Eq. (B7) that collapse to a single term for
the closed system, and not the sum in the expectation
value in Eq. (B6). Thus, there is a set of frequencies ωj
present in the time-dependent expectation value, where
ωj = Im[γ̃j,j+µ−ν ] for µ > ν and ωj = Im[γ̃j+ν−µ,j ]
for µ < ν (the two cases ensure that the indices on
γ̃j,k remain nonnegative). For any anharmonic number-
preserving Hamiltonian, all these frequencies are distinct.

Figure 6: For open quantum systems with anharmonic, To see this, consider a Hamiltonian decomposed into
number-preserving Hamiltonians, each diagonal stripe the form
of the density matrix in the number basis (colored)
evolves independently.
X
Hanh = bk n̂k . (B8)
k
bladian in Eq. (19), the complex decay rates γ̃jk have the
form
γ κ
γ̃jk = (j + k) + i (j 2 − j) − (k 2 − k) ,

(B3) We see that, for an expectation value ↵ âν s (t) (as-
2 2
suming µ < ν) the frequencies for the closed system will
and the couplings are given by have the form
(1) +
p
gjk = γ (j + 1)(k + 1), (B4)
X
(n) ± ωj = bk (j k − (j + ν − µ)k ), (B9)
with all other gjk = 0.
k
Returning to the more general case of an anharmonic
number-preserving oscillator with arbitrary n-photon
loss and gain channels, we will now show that the two
results derived in the main text hold under this more which are negative in this case. For the harmonic case, b1
general case: recurrence times are independent of both is the only non-zero term, and the frequencies ωj within
operator ordering and open system effects. Here we de- a stripe are independent of j and thus identical (or when
fine s = +1 to be normally ordered, s = 0 to be symmet- µ = ν so that there simply is no frequency). In all other
rically ordered, and s = −1 to be anti-normally ordered. cases, each frequency depends on j and thus the sum in
We can decompose the expectation value of an s- Eq. (B6) will not result in cancellations of any particular
ordered product of powers of the creation and annihi- frequency, no matter what the Fourier coefficients Rjνµs′ k′

lation operators in terms of a single stripe of the density may be, provided they are nonzero.
matrix
That they are nonzero for the closed system is trivial;
the sum Rjνµs
X µνs
↵ âν s (t) = Pjk ρjk (t)δj−k,µ−ν , (B5) ′ k ′ reduces to a single term corresponding to

jk the initial conditions for a density matrix element ρj ′ k′ (0)


multiplied by a positive combinatorial factor.
(µνs)
with Pjk a positive and real combinatorial factor. Sub-
stituting in the solution for density matrix elements given To show that the Fourier coefficients are non-zero for
in Eq. (B1), we find the open system is the final ingredient to complete our
more rigorous argument. We begin by noting that the
effect of arbitrary n-photon gain and loss channels is to
X µνs
↵ âν s (t) = Rj ′ k′ e−γ̃j′ k′ t δj ′ −k′ ,µ−ν , (B6)
j ′ k′
modify the real part of γ̃jk as well as the coefficients
′ ′
Ajj ′ k−k

+j,j
.
where we have defined new Fourier coefficients for the
expectation value This latter point is seen by inserting the formal solu-
′ ′
tion for the density matrix elements Eq. (B1) into the
Pj ′ −k′ +j,j Ajj ′ k−k
X µνs
Rjµνs
′ k′ = ′
+j,j
. (B7) coupled equations Eq. (B2),
j
18

d X (jk)
ρjk (t) = − γ̃j ′ k′ Aj ′ k′ e−γ̃j′ k′ t δj ′ −k′ ,j−k (B10)
dt ′ ′
j k
X (jk) X  (n) + j+n,k+n (n) −

= −γ̃jk Aj ′ k′ e−γ̃j′ k′ t δj ′ −k′ ,j−k + gjk Aj ′ k′ + gjk Aj−n,k−n
j ′ k′ e−γ̃j′ k′ t δj ′ −k′ ,j−k ,
j ′ k′ n

where the first line comes from performing the time frequencies that are present in a closed system expecta-
derivative and the second line comes from direct sub- tion value are also present in the open system expectation
stituion into the right hand side of Eq. (B2). Making values, any periodicity of the closed system expectation
use of the linear independence of the functions e−γ̃j′ k′ t values must also be inherited by the open system expec-
to establish equality for each term, we rearrange to yield tation values (The effect of the real part of γ̃jk coming
an expression from open system effects is to rescale the Fourier coeffi-
cients). Thus, we have re-derived our results: recurrences
(jk) of the closed system are inherited by the open system ir-
Aj ′ k′ (γ̃j ′ k′ − γ̃jk )+
X  (n) + j+n,k+n  respective of operator ordering.
(n) −
gjk Aj ′ k′ + gjk Aj−n,k−n

j k ′ = 0. We now turn to our second task, deriving analytic solu-
n tions to the Kerr system’s evolution at zero temperature
(B11) with single-photon loss utilizing this Fock-space represen-
tation in terms of diagonal stripes of the density matrix.
(n) ± We begin by truncating the Fock space at some N ≫ ⟨n̂⟩,
Since the gjk are non-zero positive numbers, we im-
mediately see that including an additional n-photon loss such that for each terminating row and column of the
(jk)
or gain channel must modify the coefficient Aj ′ k′ . A density matrix
(jk)
corollary of this is that each coefficient Aj ′ k′ in a stripe d
(n) ±
ρN k (t) = −γ̃N k ρN k (t) ∀k,
is a function of every γ̃jk and gjk within that same dt
(jk) d
stripe. Now, consider if a single Aj ′ k′ were zero. This im- ρjN (t) = −γ̃jN ρjN (t) ∀j. (B12)
P (n) + j+n,k+n (n) −
dt
plies that gjk Aj ′ k′ + gjk Ajj−n,k−n
′ k′ = 0. This
n This allows the final density matrix elements of each di-
requires an incredible amount of fine-tuning, since each agonal stripe to be determined exactly
Ajj±n,k±n
′ k′ depends on all system parameters. Even if this
holds, for the diagonal elements of the density matrix, ρN k (t) = ρN k (0) e−γ̃N k t ∀k
the coefficients Aj±n,k±n
j ′ k′ are necessarily real and posi- ρjN (t) = ρjN (0) e−γ̃jN t ∀j. (B13)
tive and so this condition can never be met. Thus, every
(jk)
coefficient Aj ′ k′ in the sum in Eq. (B1) is non-zero. For simplicity, we focus on the bottom row of the density
Having shown that the Fourier coefficients Rjνµs matrix, whose elements are given exactly by the second
′ k ′ of the

expectation value are generally non-zero, we see that any line of Eq. (B13). From these, we will be able to recur-
expectation value of the form ↵ âν s (t) inherits a set of sively generate solutions to other density matrix elements
within the same stripe, and generate the rest of the den-
frequencies ωj present in the (µ − ν)th stripe of the den-
sity matrix by complex conjugation.
sity matrix, and that this set of frequencies is unchanged
To do this, we substitute our ansatz solution in terms
by open system effects; while each frequency may con-
of the generalized discrete Fourier transform Eq. (B1)
tribute differently, it is still present in the signal for all
into the simplified master equation for the Kerr system
finite time. This set is also unchanged by operator order-
at zero temperature
ing, as it must be, since different orderings correspond to
the same operator physically. Mathematically, this is be- d (1) +
cause the commutation relation [â, ↠] = 1 preserves the ρjk (t) = −γ̃jk ρjk (t) + gjk ρj+1,k+1 (t), (B14)
dt
difference in powers of the creation and annihilation op-
erators; powers of both are removed together. Because all yielding

X (jk) (1) + (j+1,k+1)



(γ̃j ′ k′ − γ̃jk ) Aj ′ k′ − gjk Aj ′ k ′ e−γ̃j′ k′ t δj ′ −k′ −(j−k) = 0. (B15)
j ′ ,k′

Here we proceed using the same methodology as we did in Eqs. (B10) and (B11): rearranging and isolating
19

each term in the sum. We now shift notation slightly and explicitly limit our
Since the frequencies within any individual stripe are attention to just the bottom left triangle of the truncated
distinct, each term in the sum is linearly independent matrix: let j → N − i and k → N − m − i, where N is
and we conclude the Fock space truncation defined previously, and m is
how far off diagonal the stripe is, ranging from 0 (the
(1) +
gjk diagonal) to N (the bottom left stripe consisting of a
(j,k) (j+1,k+1)
Aj ′ k ′ = Aj ′ k ′ . (B16) single element of the truncated density matrix), and i is
γ̃j ′ k′ − γ̃jk the position within the stripe from its bottom-rightmost
terminating element, ranging from 0 to N − m. In this
We now have a way to generate the Fourier coefficients in notation and making use of the truncated Fock space, we
terms of the next coefficient in a stripe. From Eq. (B13), can now write an analytic solution for the density matrix
we see that there is only a single nonzero term in the of the open quantum system.
Fourier expansion of the bottom row of the density ma- We begin by rewriting, in this new notation, our formal
trix. The coefficients are solution Eq. (B1)
(
(N,k) ρN k (0), j ′ = N, k ′ = k
Aj ′ k ′ = . (B17) i
0, otherwise X (N −i,N −m−i)
ρN −i,N −m−i (t) = AN −i′ ,N −m−i′ e−γ̃N −i′ ,N −m−i′ t ,
i′ =0
Furthermore, because of the unidirectionality of the cou- (B19)
pling in Eq. (B14) for zero temperature (that is, there is
no way for population to travel from lower Fock states to where we have made use of the result in Eq. (B18) that
higher ones), we find (N,N −m)
AN −i′ ,N −m−i′ = 0 for i′ > i to truncate the sum at i
instead of N − m. We also rewrite in the new notation
(j,k)
Aj ′ k′ = 0, j > j ′ , k > k ′ . (B18) the recursion relation Eq. (B16)

(1) +
(N −i,N −m−i) gN −i,N −m−i (N −i+1,N −m−i+1)
AN −i′ ,N −m−i′ = AN −i′ ,N −m−i′
γ̃N −i′ ,N −m−i′ − γ̃N −i,N −m−i
(1) +
i−1
!
Y gN −i+u,N −m−i+u (N,N −m)
= AN −i′ ,N −m−i′ , (B20)
γ̃ ′ ′ − γ̃N −i+u,N −m−i+u
u=1 N −i ,N −m−i

where in the second line we have applied the recursion are independent of i (the location within the stripe) and
relation i − 1 times and have assumed i > 0 and i′ ≤ i can form a matrix, which is spanned by the coordinates
(again, because of the result in Eq. (B18)). i′ and m. One can then recursively solve for the elements
Defining the function (N,N −m)
AN −i′ ,N −m−i′ for each i at t = 0 and making use of the
 i−1 (1) +
boundary condition of the known initial state
Q gN −i+u,N −m−i+u
i > 0,
fN mii′ = u=1 γ̃N −i′ ,N −m−i′ −γ̃N −i+u,N −m−i+u
1 i = 0,

i
(B21) X (N,N −m)
ρN −i,N −m−i (0) = fN mii′ AN −i′ ,N −m−i′ . (B23)
and substituting Eq. (B20) into our formal solution i′ =0
Eq. (B19), we find
ρN −i,N −m−i (t)
i
X (N,N −m) (B22) For i = 0 (the terminating row of the density matrix)
= fN mii′ AN −i′ ,N −m−i′ e−γ̃N −i′ ,N −m−i′ t .
i′ =0
we already have the solution for density matrix element
evolution in Eq. (B13) and have written the values of the
We have reduced the problem of the dynamics of a sin- coefficients in Eq. (B17). Since there is a single term only,
gle stripe of the density matrix elements ρN −i,N −m−i (t) (N,N −m)
we conclude for the i = 0 case AN,N −m′ = ρN,N −m (0).
(N,N −m)
to finding the matrix of coefficients AN −i′ ,N −m−i′ , which By then considering the i = 1 case and using the bound-
we can now derive from boundary terms, and evaluating ary condition (Eq. (B23)), we can recursively generate
(N,N −m)
Eq. (B22). Notably, now the coefficients AN −i′ ,N −m−i′ the coefficients for larger and larger i up to its maximum
20

value of N − m. For i = 1, we find where in the second line we have made use of the result
from i = 0. For the general case we find
(N,N −m)
ρN −1,N −m−1 (0) = fN m10 AN,N −m
(N,N −m)
+fN m11 AN −1,N −m−1
(N,N −m)
(N,N −m) ρN −1,N −m−1 (0) − fN m10 AN,N −m
→ AN −1,N −m−1 =
fN m11
(B24)


ρN,N −m (0) i = 0,
(N,N −m)  i−1

AN −i,N −m−i = (N,N −m) (B25)
 fN1mii
P
ρN −i,N −m−i (0) − fN mii′ AN −i′ ,N −m−i′ i > 0.
i′ =0

The coefficients generated from Eq. (B25) can then be µ = ν, b) µ − ν = pN , and c) µ − ν ̸= pN where p is a
substituted into Eq. (B22) to give the analytic solution nonzero integer.
for the full time evolution of the density matrix elements.
This recipe scales poorly (∼ α06 ), is tedious to derive, Case a) µ = ν: Since [ĤKerr , n̂] = 0, all operators pro-
fails numerically for the closed system (fN mii′ develops portional to powers of the number operator are conserved
singularities), and is memory-intensive. However, it has in the evolution |ψ(tM,N )⟩ = eiĤKerr tM,N |ψ(0)⟩. Thus we
the nice property that, since it front-loads the difficulty conclude ⟨↵ âν ⟩ = |α0 |2µ when µ = ν. This is an en-
of the problem into calculating the matrix of coefficients tirely semiclassical quantity (and easy to calculate, as it
A, it is intrinsically stable in time. With this recipe, we does not even require the semiclassical approximation of
have generated a movie giving the qualitative features of the unitary dynamics).
the evolution of the bosonic anharmonic oscillator for a
small value of α0 [42]. Case b) µ − ν = pN : First note that |ψ(tM,N )⟩ is an
Analytic solutions to non-linearly evolving open quan- eigenstate of âN with eigenvalue ±α0N , where the sign is
tum system dynamics have thus far been rare to find in negative only if N is even and M is odd. To see this, con-
quantum optics. It is our hope that this method, though sider applying âN to a single term in the superposition.
cumbersome, may inspire others to find more solutions. Each application of â multiplies the coherent state by the
iπ(2k+M ) iπ2k
complex number: α0 e N if N is even and α0 e N
if N is odd. After N applications the amplitude of this
Appendix C: Kitten State Expectation Values number is independent of k with the sign dependence
noted above.
Thus, if the difference of µ and ν is an integer multiple
In this section, we derive the three cases in Eq. (32) ex-
of N , we can utilize the eigenproperty to pull out µ −
plicitly, solidifying the relationship between the N -kitten
ν powers of α0∗ and are left with an expectation value
states and the operators X̂θn via their shared symmetry
of a power of the number operator as in the first case.
Zn (the n-fold cyclic group symmetry).
Explicitly,
Recall from Eq. (33) that an operator X̂θn is decom-
posable into a sum of symmetrically ordered operators
{↵ âν }sym , where the difference |µ − ν| ≤ n for each ⟨↵ âν ⟩ = (±1)p α0∗µ−ν |α0 |2ν µ − ν = p N, (C1)
term. By the results of Appendix B, recurrence times
are independent of operator ordering and open system where, again, the sign is negative if N is even and M is
effects, so it will suffice to study normally ordered oper- odd and positive otherwise.
ators ↵ âν of the closed quantum system. This quantity is M and N dependent and thus will
The question now is, given an N -kitten state |ψ(tM,N )⟩ not be captured by semiclassical calculations that are
as defined in Eq. (9), what is ⟨↵ âν ⟩ and when is the insensitive to the particular quantum state. Hence, the
semiclassical TWA a reasonable approximation to it? If, TWA will not give the correct value here.
for some µ and ν, this expectation is independent of M
and N , then that expectation value must be entirely cal- Case c) µ − ν ̸= pN : First we show the fol-
culable in terms of semiclassical behavior since it does lowing. For 2π|α0 | ≫ N , the expectation value
not depend on the quantum structure of the state (i.e. ⟨ψ(tM,N )| âq |ψ(tM,N )⟩ tends to zero for 0 < q < N and
which kitten state has formed). To answer this question, is exactly zero for all N in the infinite-|α0 | limit. Here
it is sufficient to consider the following three cases: a) the integer q plays the role of pN (which must be an in-
21

This result can also be understood geometrically, con-


sidering the unit circle and the cyclic group ZN as illus-
trated in Fig. 7 (as well as in the accompanying video
[46]); at each stage of the group action (rotating the
phase of each state vector by the phase of its coherent
state), an initially equal distribution over the support re-
mains equally distributed, even as the structure of the
support changes due to common factors within the roots
of unity.
Both this property and the eigenvalue relation above
will also apply to the associated mixed states distributed
uniformly in the same manner. The only difference is that
now this property holds exactly even outside the large-α0
Figure 7: The cyclic group Zn for n = 6. Initially, all limit. Since the density matrix is diagonal, there are no
states have the same phase, marked by the position on contributions from terms O(⟨αi |αj ⟩) as there are for the
the circle by a blue θ (as all the phases are measured superposition.
with respect to this un-moving point). After a single We now know the expectation value in Eq. (C2) is ap-
application of â, they will have phases equally proximately zero for q < N . From this, we can conclude
distributed over the circle at the location of the black that the expectation value ↵ âν where µ − ν = pN
dots. After a second application of â, they will now be with p ∈ / Z is also approximately zero using the eigen-
equally distributed around the three pink markers. state property. Taking µ < ν, we rewrite the expectation
After a third application, they will be equally value as ↵ âν−µ âµ . The action of ↵ to the left and
distributed around two points: θ and its antipode. âµ to the right result in matching rotations of the phases
After a fourth application, they will be equally in the kitten state and an overall multiplication of the
distributed around the three green markers. And after a state by |α0 |2µ . Thus, the expectation value reduces to
fifth application, they will be equally distributed around |α0 |2µ ⟨âν−µ ⟩, and we can make use of our results from
the black dots again before finally “recohering” after a Eq. (C2) with q = ν − µ to say that this zero, provided
sixth (nth) application at the dot with the blue theta. ν − µ < N . If instead ν − µ > N , we make use of the
At each stage the group action maintains an equal eigenstate property to pull out N powers of the annihi-
distribution over the support: the roots of unity lation operator â and replace them with α̃0N , repeating
corresponding to each way of factoring n. the process w times until there are q = ν − µ − wN < N
For a video describing this, see [46]. powers of â remaining in the expectation value and a pre-
factor α̃0wN . Since ν − µ is not an integer multiple of N ,
q = ν − µ − wN is not an integer multiple of N , and the
teger since µ − ν is an integer). For now we assume p < 1 expectation value is approximately zero. Similarly, for
so that q < N , but this restriction will be relaxed later. µ > ν, taking complex conjugation to give the associated
expressions.
The expectation value ⟨ψ(tM,N )| âq |ψ(tM,N )⟩ is Thus, for a sufficiently large |α0 |, the expectation value
rewritten in terms of an analytic function f (ω), such that is approximately zero and can be calculated semiclas-
sically. Note that when the number of kittens N be-
⟨ψ(tM,N )| âq |ψ(tM,N )⟩ = α̃0q f (ωN
q
) + O(⟨αi |αj ⟩), (C2) comes large enough such that 2π|α0 | ≲ N , their non-
orthogonality contributes to oscillations, giving rise to
the non-precision of the recurrences (due to neighboring
where ωN is an N th root of unity e2πi/N and we have high-N kitten states) as well as the deviations from the
iπM
defined α̃0 = e N α0 for N -even M -odd and α̃0 = α0 TWA at short times (also high-N kitten states) in Fig. 4.
otherwise. O(⟨αi |αj ⟩) goes to zero in the limit of orthog- To summarize, we have rederived the three cases of
onal coherent states (that is, 2π|α0 | ≫ N for coherent Eq. (32). We know that the recurrences for normally
states distributed uniformly around the circle with ra- ordered expectation values given in Eq. (26) that we cal-
dius |α0 |). f (ω) is the sum of a finite geometric sequence culated in the Heisenberg picture is a manifestation of
the symmetry of kitten states being that of the symmet-
1 − ωN ric group Zn , which is shared by operators ↵ âµ when
f (ω) = 1 + ω + ω 2 + ω 3 + · · · + ω N −1 = . (C3) µ − ν = pN , with p ∈ Z± . By the various results of the
1−ω
previous sections, we can also appreciate that this is true
of symmetric products of powers of ↠and â. The oper-
q
Since ω qN = 1q = 1, it follows that f (ωN ) = 0. ators X̂θn , when expanded in such products, will contain
Hence, the corresponding term in the expectation value terms with µ − ν = ±n, hence the relationships between
in Eq. (C2) is zero as well. the operators X̂θn and the N -kitten states when n = pN .

You might also like