You are on page 1of 12

Large Deviations Theory Beyond the Kibble-Zurek Mechanism

1, 6, ∗
Federico Balducci ,1 Mathieu Beau ,2 Jing Yang ,1, 3 Andrea Gambassi ,4, 5 and Adolfo del Campo
1
Department of Physics and Materials Science, University of Luxembourg, L-1511 Luxembourg, G. D. Luxembourg
2
Department of Physics, University of Massachusetts, Boston, MA 02125, USA
3
Nordita, KTH Royal Institute of Technology and Stockholm University,
Hannes Alfvéns väg 12, SE-106 91 Stockholm, Sweden
4
SISSA - International School for Advanced Studies, via Bonomea 265, 34136 Trieste, Italy
5
INFN, Sezione di Trieste, Trieste, Italy
6
Donostia International Physics Center, E-20018 San Sebastián, Spain
The Kibble-Zurek mechanism (KZM) predicts that the average number of topological defects
generated upon crossing a quantum phase transition obeys a universal scaling law with the quench
time. Fluctuations in the defect number near equilibrium are approximately of Gaussian form,
in agreement with the central limit theorem. Using large deviations theory, we characterize the
universality of fluctuations beyond the KZM and report the exact form of the rate function in the
arXiv:2307.02524v1 [quant-ph] 5 Jul 2023

transverse-field quantum Ising model. In addition, we characterize the scaling of large deviations in
an arbitrary continuous phase transition, building on recent evidence establishing the universality
of the defect number distribution.

The Kibble-Zurek mechanism (KZM) is an important generated topological defects [33–35]. The full counting
paradigm in nonequilibrium statistical physics, describ- statistics of defects appears to be universal in classical
ing the dynamics across a continuous phase transition [1– and quantum systems. Specifically, in classical continu-
3]. The divergence of the equilibrium relaxation time in ous phase transitions, it has been found that the defect
the neighborhood of the critical point makes the criti- number distribution is binomial with an average density
cal dynamics necessarily nonadiabatic for large systems in agreement with the KZM [36–39]. Exact solutions in
and leads to the spontaneous formation of topological de- quantum integrable systems have shown that the kink
fects. Consider a phase transition from a high symmetry number distribution is Poisson-binomial [25, 34], a fea-
phase to a broken symmetry phase, induced by varying ture that can hold even when the system is coupled to
a control parameter g across its critical value gc in a fi- an environment [31, 32]. These predictions build on the
nite quench time τQ . The central prediction of the KZM conventional KZM but lie outside its scope, requiring ad-
is that the average defect density, generated during the ditional assumptions. We shall thus refer to them as
phase transition, displays a universal power-law depen- beyond-KZM physics.
dence as a function of the quench time. The KZM thus The average number of defects is an extensive quantity.
makes a quantitative prediction on the breakdown of adi- By contrast, the defect density is intensive, and its fluc-
abatic dynamics across a phase transition and holds both tuations near equilibrium are approximately Gaussian, in
in the classical and quantum regimes [1, 4–7]. agreement with the central limit theorem. Large devia-
Kibble’s pioneering work was motivated by cosmo- tions theory (LDT) addresses the probability of nontyp-
logical considerations regarding structure formation in ical events in which an intensive quantity deviates from
the early universe [8]. The prospect of exploring analo- its average value. The probability of such large devi-
gous phenomena in condensed matter systems was soon ations decays exponentially upon increasing the system
realized [9–11] and pursued experimentally [3, 12–14]. size, with a rate controlled by the so-called rate func-
The advance of quantum technologies has led to new tion [40, 41]. LDT provides a building block of statis-
tests of the KZM using quantum simulators in a variety tical mechanics in and out of equilibrium. As such, it
of platforms, including ultracold gases [15–21], trapped is a natural framework to explore beyond-KZM physics.
ions [22–26], and Rydberg gases [27, 28]. Recently, the To date, LDT has been used to describe the dynamics
KZM has been studied with quantum computing devices, of many-body quantum systems in the limit of sudden
such as quantum annealers [29–32]. The accumulated quenches when τQ → 0, e.g., to characterize the work
body of literature broadly supports the validity of KZM statistics of a given process [42–44].
in a wide variety of systems. In this Letter, we establish the universality of large
Experiments probing critical dynamics generally in- deviations beyond the KZM, after crossing a quantum
volve an ensemble of single experimental runs or individ- phase transition in a finite time. Specifically, we re-
ual realizations in which measurements are performed. port the exact rate function of the driven transverse-
As a result, they can access information beyond the av- field quantum Ising model (TFQIM), characterizing the
erage defect density and characterize the ensemble statis- statistics of large fluctuations away from the mean kink
tics. It is thus natural to ask whether there are universal density predicted by the conventional KZM. We further
signatures in the statistical properties of spontaneously generalize these results to characterize the universality of
2

large deviations in an arbitrary continuous phase transi- at the final time τQ . It exhibits a power-law scaling in
tion leading to point-like defects. the slow driving limit, i.e., to leading order in a 1/τQ
The transverse-field quantum Ising model. The expansion [1, 4–7]
TFQIM has been instrumental in generalizing the KZM s
from the classical to the quantum domain [4–7, 24], and 1 ℏ
assessing the universality of beyond-KZM physics, both ρKZM = ⟨ρ̂N ⟩ = , (5)
4π 2JτQ
in theory [34] and experiments [25, 31, 32]. Its Hamilto-
nian is given by in agreement with the celebrated, universal KZM power-
− ν
N
X law scaling ρKZM ∝ τQ 1+zν for the critical exponents
g(t)σlx + σlz σl+1
z ν = z = 1 of the TFQIM [3].
 
H[g(t)] = −J , (1)
l=1 In any quantum state other than an eigenstate of
H(g = 0), the density operator ρ̂N will exhibit fluctua-
where J > 0 favors ferromagnetic alignment and g(t) tions of either classical or quantum nature. The probabil-
plays the role of an effective magnetic field. In the ity distribution function P (ρN ), characterizing the eigen-
fermionic representation, the Ising chain Hamiltonian be- value statistics of the kink-pair density operator, reads
comes [45]
X † P (ρN ) = ⟨δ(ρ̂N − ρN )⟩ , (6)
H[g(t)] = 2J ψk [τ z (g(t) − cos k) + τ y sin k] ψk , (2)
k>0 where ρN is the random variable associated with the
kink-pair-number operator ρ̂N . We aim at uncovering
in terms of the fermionic operators ψk† ≡ (c̃†k , c̃−k ) in via LDT the universality of large fluctuations of P (ρN )
momentum space. Here, τ x,y,z are Pauli matrices. We away from the mean, which the conventional KZM pre-
choose to work with periodic boundary conditions so that dicts.
the momentum is a good quantum number and takes the Large deviations theory beyond the KZM in the
values k = (2n + 1)π/N with n = −N/2, . . . , N/2 − 1, TFQIM. The central object in LTD is the scaled cu-
as discussed, e.g., in Refs. [45, 46]. Momentum conser- mulant generating function, associated with a random
vation restricts the formation of defects to kink-antikink variable ρN , depending on a large parameter N ,
pairs. Choosing the total number of spins N to be even
proves convenient since the number of kink pairs is then 1
N θρ̂N
λ(θ) = lim ln e . (7)
restricted to outcomes in the set {0, 1, 2, . . . , N/2}. Given N →∞ N
Eq. (2), the dynamics of the TFQIM can be reduced
to that of an ensemble of non-interacting two-level sys- The Gärtner-Ellis theorem states that when λ(θ) exists
tems [7]. for all θ ∈ R, then the random variable ρN satisfies the
Consider a quench, in a finite time τQ , from the para- large deviations principle [40, 41]
magnetic to the ferromagnetic phase
P ρN ∈ [ρ, ρ + dρ] ≈ e−N I(ρ) dρ,

(8)
 
t
g(t) = gc 1 − , (3) with the rate function I(ρ) given by the Legendre-Fenchel
τQ
transform
where g(0) = gc = 1 is the critical value of g, and  
we let t run from −3τQ to τQ . We will refer to τQ I(ρ) = sup θρ − λ(θ) . (9)
θ∈R
as the quench time. We choose g(τQ ) = 0 for simplic-
ity since the final Hamiltonian contains only the ferro- Deviations from the mean value are thus exponentially
magnetic term and commutes
PN with the kink-pair num- suppressed by the rate function I(ρ) weighted with
ber operator KN ≡ 14 l=1 1 − σlz σl+1 z
. This observ- the system size, and the random variable concentrates
able counts the number of kink-antikink pairs in a given around the mean in the thermodynamic limit.
quantum state and is extensive in the system size N [7]. Let us consider the application of the Gärtner-Ellis the-
The study of its eigenvalue statistics provided the ba- orem to the distribution of kink pairs generated across a
sis of previous studies exploring universality beyond the quantum phase transition in the TFQIM. In this case,
KZM [25, 31, 32, 34, 37]. We define an intensive kink-pair the defect density is a non-negative quantity. As a re-
density operator sult, I(ρ) is divergent for ρ < 0, and we focus on the
domain ρ ≥ 0. We note that in Fourier space, the oper-
N
KN 1 X ator associated with the density of kink pairs at the end
1 − σlz σl+1
z

ρ̂N ≡ = . (4)
N 4N of the quench is
l=1

The density of kink pairs, upon completion of the quench 1 X †


ρ̂N = γk (τQ )γk (τQ ), (10)
in Eq. (3), is given by the expectation value ρKZM = ⟨ρ̂N ⟩ N
k>0
3

where γk (τQ ) and γk† (τQ ) are the fermionic Bogoliubov 100
operators at the end of the quench, and the sum is re- τQ = 32 h̄J −1

stricted to k > 0 since the number of kink pairs equals the 80 τQ = 100 h̄J −1

number of right-moving kinks. Further, for free fermions τQ = 320 h̄J −1


60 τQ = 1000 h̄J −1
(with periodic boundary conditions), the time-dependent

¯ ρ̄)
τQ = 3200 h̄J −1
density matrix ϱ(t) retains the tensor product N structure

I(
during unitary time evolution, i.e., ϱ(t) = k ϱk (t). As 40 LZ
CLT
a result, the moment-generating function admits the ex-
20
plicit form

N θρ̂N Y h 0

i
e = Tr ϱk (τQ )eθγk (τQ )γk (τQ ) 0 2 4 6 8
k>0 ρ̄
Y
1 + eθ − 1 pk ,
 
= (11)
k>0 FIG. 1. Comparison of the scale rate function I(ρ̄) ¯ =
I(ρ)/ρKZM derived analytically with the numerically exact
where pk = ⟨γk† (τQ )γk (τQ )⟩ ∈ [0, 1] represents the prob- computation for a finite τQ and N . A TFQI chain is ini-
ability that the mode k is excited at the end of the tialized in its ground state g(−3τQ ) = 4gc . The equations of
protocol. This is the moment-generating function of a motion are integrated numerically up to time τQ , correspond-
ing to g(τQ ) = 0. From the final wavefunction, the cumulant
Poisson binomial distribution associated with the sum generating function λ(θ) is computed using Eq. (7) for finite
of N/2 independent random Bernoulli variables, each N , from which the scaled rate function I¯ is found with a
of which has probability pk for the occupation num- Legendre-Fenchel transform. As the quench time increases,
ber to be 1, corresponding to the formation of a kink- the agreement between numerics and the analytical solution
antikink pair, and probability (1 − pk ) for the occupa- based on the LZ approximation improves visibly, while the
tion number to be 0, corresponding to no defect for- agreement with the central limit theorem (CLT) prediction
¯
does not. The value at the origin is I(0) = ζ(3/2) follows
mation [34]. In addition, the value of pk can be esti-
from Eq. (14), while the minimum I¯ = 0 is attained at ρ̄ = 1
mated according to the Landau-Zener approximation [7], (diamond). Finite-size analysis reveals the convergence of the
pk = ⟨γk† (τQ )γk (τQ )⟩ ≈ exp(−2πJτQ k 2 /ℏ) near k = 0, numerically-evaluated rate function I¯ to the thermodynamic
dictating an exponential decay with increasing quench limit for N = 1000, which is used in this figure.
time and a Gaussian decay as a function of the wavenum-
ber. This behavior dictates the KZM scaling in a quan-
tum phase transition [6, 7, 47]. The explicit computation
of the scaled cumulant generating function, according to As a result, the rate function (14) is universal in the
¯
sense that I(ρ̄) = I(ρ)/ρKZM varies only with ρ̄ and is
Eqs. (7) and (11), in the limit N → ∞, yields
independent of the quench time τQ . This is the central
Z π
dk result of our work, which we elaborate and generalize in
ln 1 + eθ − 1 pk ,
  
λ(θ) = (12) what follows. Taking the derivative with respect to θ,
0 2π
one finds at the supremum θ∗
which is a convergent integral.
For slow quenches, using a power-series expansion in ∗
1/τQ to leading order, or equivalently extending the up- eθ 
θ∗

ρ̄ = − Li1/2 1 − e . (15)
per limit of the integral in Eq. (12) to infinity, one finds eθ ∗ − 1

λ(θ) = −ρKZM Li3/2 1 − eθ ,



(13)
The function θ∗ (ρ̄) and the rate function scaled by the
in ¯
KZM density I(ρ̄) are found numerically. The rate func-
P∞terms of the polylogarithm function Liq (x) =
s=1 x s q
/s . The exact expression in the slow-quench tion is shown in Fig. 1. As the decay of the probability
limit, Eq. (13), shows that λ(θ) is differentiable for all density function P (ρ) is dictated by the rate function
values of θ, ensuring the applicability of the Gärtner- according to Eq. (8), the minimum of I(ρ̄)¯ at ρ̄ = 1 is
Ellis theorem. We verify that for θ = 0, λ(0) = 0, consis- associated with the most likely value of ρ̂N , which equals
tently with its definition. Further, for θ < 0, λ(θ) quickly the mean value ρKZM predicted by the KZM. Thus, LDT
approaches the constant value λ(−∞) = −ρKZM ζ(3/2). guarantees that the KZM prediction holds with maxi-
Indeed, λ(θ) is approximately constant for θ < 0 and is mum probability.
a monotonic function of θ. Concentration inequalities. Let us tackle the prob-
We define a dimensionless density of defects ρ̄ ≡ lem of large deviations from a complementary angle us-
ρ/ρKZM in terms of which ing concentration inequalities [48]. To bound the devia-
tions, we make use of the Chernoff bound, which reads
I(ρ) = ρKZM sup θρ̄ + Li3/2 1 − eθ .
 
(14)
θ∈R P (ρN > ρ) ≤ ⟨eθρ̂N ⟩e−θρ , for all θ > 0. The characteris-
4

tic function can be written as at different locations to be described by independent and


identically distributed discrete random variables Xi with
N π
 Z 
i = 1, . . . , N [36–39], where the outcome Xi = +1 cor-

θρ̂N
dk ln 1 + (eθ − 1)pk
 
e = exp (16)
2π 0 responds to the formation of a topological defect, and
≈ exp −N ρKZM Li3/2 1 − eθ .
 
Xi = 0 to no defect formation. The defect number distri-
bution takes the binomial form P (n) = N n
n p (1−p)
N −n
.
We thus find from the Chernoff bound that Numerical studies support this prediction in d = 1, 2 for
varying τQ [36–39]. Accordingly, the average number of
P (ρN > ρ) ≤ exp −ρKZM [θρ̄ + Li3/2 (1 − eθ )]

(17) − dν
topological defects is given by ρKZM = pN /Vd ∝ τQ 1+zν .
for all θ > 0. Tightening the above inequality by tak- The defectPdensity, an intensive random variable, is given
N
ing the supremum of the exponent, the right tail of the by ρN = i=1 Xi /Vd . We are interested in estimating
PN
distribution is bounded as the probability distribution of SN = i=1 Xi when N is
large. Using the Stirling approximation, one finds
P (ρN > ρ) ≤ exp[−N I(ρ)], (18)
1
P (SN = rN ) = p e−Vd ρKZM DKL (r∥p) , (20)
with I(ρ) given by Eq. (14). Likewise, the left tail is 2πr(1 − r)N
bound by the same term, P (ρN < ρ) ≤ exp[−N I(ρ)].
The logarithm of the two-sided Chernoff bound is the where DKL (r∥p) = r ln(r/p) + (1 − r) ln[(1 − r)/(1 − p)]
rate function. The above results establish the nature of is the Kullback-Liebler distance between two Bernoulli
large deviations of kink-antikink pairs formed across the distributions with success probabilities r and p. It sat-
quantum phase transition between the paramagnetic and isfies DKL (r∥p) ≥ 0, with the equality holding when r
the ferromagnetic phase of the TFQIM. Such results are equals p, which is the most probable value. Neglecting
generalizable to the family of quasi-free fermion models the prefactor, we thus find that for large N , fluctuations
in which the density of defects is given by the density of the defect number away from the mean are suppressed
of quasiparticles. In addition, the universal form of the exponentially with increasing N , i.e., P (SN = rN ) ≈
scaled cumulant generating function and the rate func- exp[−Vd ρKZM DKL (r∥p)]. In this sense, the defect num-
tion in the slow quench limit also hold when fast decaying ber distribution concentrates at the KZM prediction in
long-range interactions are considered [45], further con- the thermodynamic limit when Vd and N diverge. In-
firming their universality under fast-decaying long-range deed, in the spirit of LDT, we identify the rate function
deformations. We next turn our attention to an arbitrary
I(r) = ρKZM DKL (r∥p), (21)
continuous phase transition described by the KZM.
LDT beyond KZM: General scenario. Consider a sce- generalizing the findings for the TFQIM to arbitrary con-
nario of spontaneous symmetry breaking leading to the tinuous phase transitions. Similarly, I(r) dictates the
generation of point-like defects in d spatial dimensions. universal suppression of deviations away from the KZM
The KZM exploits the equilibrium scaling relations for prediction. For example, the right tail of the distribution
the correlation length ξ and the relaxation time τ , i.e., is bounded as
ξ0 τ0 P (SN ≥ rN ) ≤ exp [−Vd I(r)] (22)
ξ= , τ= , (19)
|ε|ν |ε|zν " dν
  1+zν #
pVd τ0
= exp − d DKL (r∥p) .
where ν and z are critical exponents and ξ0 and τ0 f ξ0 τQ
are microscopic constants. The dimensionless variable
ε = (gc − g)/gc quantifies the distance to the critical These results are fully consistent with LDT. Using
point gc , and vanishes at the phase transition. Lin- the dimensionless density of defect ρ̄ ≡ ρN /ρKZM =
1
PN
earizing the driving protocol in the neighbourhood of gc pN i=1 Xi , according to Sanov’s theorem in LDT [41]
as g(t) = gc (1 − t/τQ ), one identifies the quench time P (pρ̄ = r) = e−N DKL (r∥p) . Thus, P (ρ̄) = exp[−Vd I(ρ̄)]
τQ . The KZM sets the nonequilibrium correlation length where I(ρ̄) ≡ ρKZM DKL (ρ̄p∥p)/p. The tails of the dis-
ν
to be ξˆ = ξ0 (τQ /τ0 ) 1+zν . During the phase transition, tribution read then P (ρ̄ ≥ σ) ≤ exp[−Vd I(σ)] as in Eq.
the system is partitioned into protodomains of average (18).
volume ξˆd . A defect may be generated with an em- Discussion. The rate function governs the nature of de-
pirical probability p at the merging point between ad- viations away from the mean according to LDT. Knowl-
jacent domains. For point-like defects, the number of edge of the rate function relies on the full counting statis-
events for defect formation is estimated as N = Vd /(f ξˆd ), tics of topological defects beyond KZM, as the rate func-
where Vd is the volume in d spatial dimensions and f a tion is the Legendre-Frenchel transform of the scaled cu-
fudge factor of order one. As a result, it scales as N = mulant generating function. Using the exact solution of

[Vd /(f ξ0d )] (τ0 /τQ ) 1+zν . Assume defect formation events the critical dynamics in the TFQIM as a test bed, we
5

have explored the nature of large deviations in the num- M. J. Davis, and B. P. Anderson, Nature 455, 948
ber of topological defects generated across a quantum (2008).
phase transition driven at the finite rate and show that [16] G. Lamporesi, S. Donadello, S. Serafini, F. Dalfovo, and
the rate function is proportional to the KZM density of G. Ferrari, Nat. Phys. 9, 656 (2013).
[17] N. Navon, A. L. Gaunt, R. P. Smith, and Z. Hadzibabic,
kink pairs. The rate function exhibits a universal power- Science 347, 167 (2015).
law scaling with the quench time in which the phase tran- [18] M. Anquez, B. A. Robbins, H. M. Bharath, M. Bogus-
sition is crossed. We have further generalized these find- lawski, T. M. Hoang, and M. S. Chapman, Phys. Rev.
ings to account for the dynamics of arbitrary continuous Lett. 116, 155301 (2016).
phase transitions described by KZM. Assuming the in- [19] B. Ko, J. W. Park, and Y. Shin, Nat. Phys. 15, 1227
dependence of defect formation events during the phase (2019).
transition, we have shown that the probability density [20] C.-R. Yi, S. Liu, R.-H. Jiao, J.-Y. Zhang, Y.-S. Zhang,
and S. Chen, Phys. Rev. Lett. 125, 260603 (2020).
function of the defect number distribution decays expo- [21] L.-Y. Qiu, H.-Y. Liang, Y.-B. Yang, H.-X. Yang, T. Tian,
nentially with the number of defects predicted by KZM. Y. Xu, and L.-M. Duan, Sci. Adv. 6, eaba7292 (2020).
Our results are of broad interest in nonequilibrium quan- [22] S. Ulm, J. Roßnagel, G. Jacob, C. Degünther, S. T.
tum and classical statistical mechanics, connecting LDT Dawkins, U. G. Poschinger, R. Nigmatullin, A. Retzker,
with the breakdown of adiabatic dynamics, and should M. B. Plenio, F. Schmidt-Kaler, and K. Singer, Nat.
find broad applications in quantum simulation, quantum Commun. 4, 2290 (2013).
annealing, ultracold atom physics, and the study of crit- [23] K. Pyka, J. Keller, H. L. Partner, R. Nigmatullin,
T. Burgermeister, D. M. Meier, K. Kuhlmann, A. Ret-
ical phenomena. zker, M. B. Plenio, W. H. Zurek, A. del Campo, and
Acknowledgements. AdC thanks SISSA for its hospi- T. E. Mehlstäubler, Nat. Commun. 4, 2291 (2013).
tality during the early stages of this work. We thank Fed- [24] J.-M. Cui, Y.-F. Huang, Z. Wang, D.-Y. Cao, J. Wang,
erico Roccati for their feedback on the manuscript. This W.-M. Lv, L. Luo, A. del Campo, Y.-J. Han, C.-F. Li,
and G.-C. Guo, Sci. Rep. 6, 33381 (2016).
project was funded within the QuantERA II Programme
[25] J.-M. Cui, F. J. Gómez-Ruiz, Y.-F. Huang, C.-F. Li, G.-
that has received funding from the European Union’s C. Guo, and A. del Campo, Commun. Phys. 3, 44 (2020).
Horizon 2020 research and innovation programme under [26] B.-W. Li, Y.-K. Wu, Q.-X. Mei, R. Yao, W.-Q. Lian, M.-
Grant Agreement No. 16434093. AG acknowledges finan- L. Cai, Y. Wang, B.-X. Qi, L. Yao, L. He, Z.-C. Zhou,
cial support from the PNRR MUR project PE0000023- and L.-M. Duan, PRX Quantum 4, 010302 (2023).
NQST. For open access and in fulfillment of the obli- [27] A. Keesling, A. Omran, H. Levine, H. Bernien, H. Pich-
gations arising from the grant agreement, the authors ler, S. Choi, R. Samajdar, S. Schwartz, P. Silvi,
S. Sachdev, P. Zoller, M. Endres, M. Greiner, V. Vuletić,
have applied a Creative Commons Attribution 4.0 Inter-
and M. D. Lukin, Nature 568, 207 (2019).
national (CC BY 4.0) license to any Author Accepted [28] S. Ebadi, T. T. Wang, H. Levine, A. Keesling, G. Se-
Manuscript version arising from this submission. meghini, A. Omran, D. Bluvstein, R. Samajdar, H. Pich-
ler, W. W. Ho, S. Choi, S. Sachdev, M. Greiner,
V. Vuletić, and M. D. Lukin, Nature 595, 227 (2021).
[29] B. Gardas, J. Dziarmaga, W. H. Zurek, and M. Zwolak,
Sci. Rep. 8, 4539 (2018).

adolfo.delcampo@uni.lu [30] P. Weinberg, M. Tylutki, J. M. Rönkkö, J. Westerholm,
[1] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat- J. A. Åström, P. Manninen, P. Törmä, and A. W. Sand-
tore, Rev. Mod. Phys. 83, 863 (2011). vik, Phys. Rev. Lett. 124, 090502 (2020).
[2] J. Dziarmaga, Adv. Phys. 59, 1063 (2010). [31] Y. Bando, Y. Susa, H. Oshiyama, N. Shibata, M. Ohzeki,
[3] A. del Campo and W. H. Zurek, Intl. J. Mod. Phys. A. F. J. Gómez-Ruiz, D. A. Lidar, S. Suzuki, A. del Campo,
[4] A. Polkovnikov, Phys. Rev. B 72, 161201 (2005). and H. Nishimori, Phys. Rev. Research 2, 033369 (2020).
[5] W. H. Zurek, U. Dorner, and P. Zoller, Phys. Rev. Lett. [32] A. D. King, S. Suzuki, J. Raymond, A. Zucca, T. Lant-
95, 105701 (2005). ing, F. Altomare, A. J. Berkley, S. Ejtemaee, E. Hoskin-
[6] B. Damski, Phys. Rev. Lett. 95, 035701 (2005). son, S. Huang, E. Ladizinsky, A. J. R. MacDonald,
[7] J. Dziarmaga, Phys. Rev. Lett. 95, 245701 (2005). G. Marsden, T. Oh, G. Poulin-Lamarre, M. Reis, C. Rich,
[8] T. W. B. Kibble, J. Phys. A: Math. Gen. 9, 1387 (1976). Y. Sato, J. D. Whittaker, J. Yao, R. Harris, D. A. Lidar,
[9] T. W. B. Kibble, Phys. Rep. 67, 183 (1980). H. Nishimori, and M. H. Amin, Nat. Phys. 18, 1324
[10] W. H. Zurek, Nature 317, 505 (1985). (2022).
[11] W. H. Zurek, Phys. Rep. 276, 177 (1996). [33] L. Cincio, J. Dziarmaga, M. M. Rams, and W. H. Zurek,
[12] S. Deutschländer, P. Dillmann, G. Maret, and P. Keim, Phys. Rev. A 75, 052321 (2007).
Proc. the Natl. Acad. Sci. 112, 6925 (2015). [34] A. del Campo, Phys. Rev. Lett. 121, 200601 (2018).
[13] S. Maegochi, K. Ienaga, and S. Okuma, Phys. Rev. Lett. [35] A. del Campo, F. J. Gómez-Ruiz, and H.-Q. Zhang,
129, 227001 (2022). Phys. Rev. B 106, L140101 (2022).
[14] K. Du, X. Fang, C. Won, C. De, F.-T. Huang, W. Xu, [36] F. J. Gómez-Ruiz, J. J. Mayo, and A. del Campo, Phys.
H. You, F. J. Gómez-Ruiz, A. del Campo, and S.-W. Rev. Lett. 124, 240602 (2020).
Cheong, Nat. Phys. (2023), 10.1038/s41567-023-02112- [37] J. J. Mayo, Z. Fan, G.-W. Chern, and A. del Campo,
5. Phys. Rev. Research 3, 033150 (2021).
[15] C. N. Weiler, T. W. Neely, D. R. Scherer, A. S. Bradley, [38] A. del Campo, F. J. Gómez-Ruiz, Z.-H. Li, C.-Y. Xia,
6

H.-B. Zeng, and H.-Q. Zhang, JHEP 2021, 61 (2021). [45] See Supplementary Information.
[39] F. J. Gómez-Ruiz, D. Subires, and A. del Campo, Phys. [46] B. Damski and M. M. Rams, J. Phys. A: Math. Theor.
Rev. B 106, 134302 (2022). 47, 025303 (2013).
[40] R. Ellis, Entropy, Large Deviations, and Statistical Me- [47] B. Damski and W. H. Zurek, Phys. Rev. A 73, 063405
chanics, Classics in Mathematics (Springer, 2006). (2006).
[41] H. Touchette, Phys. Rep. 478, 1 (2009). [48] R. Vershynin, High-Dimensional Probability: An In-
[42] A. Gambassi and A. Silva, Phys. Rev. Lett. 109, 250602 troduction with Applications in Data Science, Cam-
(2012). bridge Series in Statistical and Probabilistic Mathematics
[43] J. Goold, F. Plastina, A. Gambassi, and A. Silva, “The (Cambridge University Press, 2018).
role of quantum work statistics in many-body physics,” in [49] D. Vodola, L. Lepori, E. Ercolessi, A. V. Gorshkov, and
Thermodynamics in the Quantum Regime: Fundamental G. Pupillo, Phys. Rev. Lett. 113, 1 (2014).
Aspects and New Directions, edited by F. Binder, L. A. [50] J. Yang, S. Pang, A. del Campo, and A. N. Jordan,
Correa, C. Gogolin, J. Anders, and G. Adesso (Springer Phys. Rev. Research 4, 013133 (2022).
International Publishing, Cham, 2018) pp. 317–336. [51] A. Dutta and A. Dutta, Phys. Rev. B 96, 125113 (2017).
[44] G. Perfetto, L. Piroli, and A. Gambassi, Phys. Rev. E [52] L. Pezzè, M. Gabbrielli, L. Lepori, and A. Smerzi, Phys.
100, 032114 (2019). Rev. Lett. 119, 250401 (2017).

Supplementary Information

FERMIONIC BASE FOR THE TRANSVERSE-FIELD QUANTUM ISING MODEL

To fix the notation, we briefly review how the transverse-field quantum Ising model (TFQIM) can be mapped onto
a set of non-interacting two-level systems [2]. Let us start from the chain Hamiltonian
N
X
g(t)σlx + σlz σl+1
z
 
H[g(t)] = −J . (23)
l=1

First, by applying on each site the unitary gate


 
1 z 1 −1
U = UHadamard σ = √ , (24)
2 1 1

H can be brought into the equivalent form


N
X
g(t)σlz − σlx σl+1
x
 
H[g(t)] = J . (25)
l=1

Second, we apply a Jordan-Wigner transformation and we introduce the fermionic operators

c†l := eiπΣl σl+ , cl := e−iπΣl σl− , σlz = 2c†l cl − 1, (26)

where Σl is the string


l−1
c†j cj .
X
Σl := (27)
j=1

The Hamiltonian (25) in terms of the fermionic operators reads


Xh † † i
H[g(t)] = −J cl cl+1 + c†l cl+1 − cl c†l+1 − cl cl+1 − 2g(t)c†l cl − N Jg(t). (28)
l

Next, we define the Fourier basis


1 X −ikl 1 X ikl †
c̃k := √ e cl , c̃†k := √ e cl , (29)
N l N l

where the symmetric definition of the Fourier transform (i.e., using the 1/ N prefactor) is essential to have the
commutation relations {c̃k , c̃†q } = δkq without spurious factors of N . One needs to impose the boundary conditions to
7

fix the values that k can acquire [46]. We assume periodic boundary conditions in the spin chain, i.e., ⃗σl+N = ⃗σl . In
the fermionic representation, there is a dependence on the number of particles. If it is even, a particle must acquire a
negative phase when performing a full circle. This means P + cl+N P + = −cl , where P + projects onto the even sector.
Similarly, P − cl+N P − = cl . We choose to work with an even number of particles, whence
1 X ikl 1 X ik(l+N )
cl+N = −cl =⇒ √ e c̃k = √ e c̃k =⇒ eikN = −1. (30)
N k N k

One can verify that the solution is


 
2π 1 N N N
k= n+ for n = − , − + 1, . . . , − 1. (31)
N 2 2 2 2

At this point, the fermionic Hamiltonian rewritten in momentum space reads


Xh    i
H[g(t)] = 2J −i sin k c̃†k c̃†−k + c̃−k c̃k + (g(t) − cos k) c̃†k c̃k − c̃−k c̃†−k . (32)
k>0

Alternatively,
X  g(t) − cos k −i sin k

c̃k


H[g(t)] = 2J c̃k , c̃−k (33)
i sin k −g(t) + cos k c̃†−k
k>0

and, defining
 
  c̃k
ψk† := c̃†k , c̃−k , ψk := , (34)
c̃†−k

the Hamiltonian acquires the compact form


X
H[g(t)] = 2J hk [g(t)], (35)
k>0

where

hk [g(t)] ≡ ψk† [(g(t) − cos k)τ z + sin k τ y ] ψk , (36)

and the Pauli matrices τ x,y,z act on the two entries of the ψk ’s. Notice that ψk has only positive k’s in the subscript.
Equation (35) is exactly the same as Eq. (2) in the main text.

DIAGONALIZATION OF THE TRANSVERSE-FIELD QUANTUM ISING MODEL

We want to diagonalize the Hamiltonian (2) through a Bogoliubov rotation. Thus, we change the basis for the
creation/annihilation operators as
    
γk (t) cos[θk (t)/2] −i sin[θk (t)/2] c̃k
† = . (37)
γ−k (t) −i sin[θk (t)/2] cos[θk (t)/2] c̃†−k

Again, only k > 0 are used, leaving an explicit minus sign where needed. The value of θk is fixed by the diagonalization
of hk , which yields the relation

g(t) − cos k
cos θk (t) = p . (38)
g 2 (t) + 1 − 2g(t) cos k

It follows that
  
X  ϵk (t) 0 γk (t)
H[g(t)] = γk† (t), γ−k (t) † , (39)
0 −ϵk γ−k (t)
k>0
8

10−1 −1/2
(a) 102 (b) (c)
∼ τQ

10 −3 0.3
N = 32 101
N = 56
10 −5 κ2/κ1

κq /κq0
hρ̂N i

N = 100 0.2
100

κq
κ3/κ1
N = 180
κ3/κ2
10−7 N = 320
−1 0.1
N = 560 10 κ1
10−9 N = 1000 κ2
N = 1800 κ3
10−2 0.0
0 1 2 3 4
10 10 10 10 10 100 101 102 103 100 101 102 103
τQJ/h̄ τQJ/h̄ τQJ/h̄

FIG. 2. (a) Average kink-pair density produced by the finite-time crossing of the phase transition. One can see the curves
−1/2
collapsing on the single power-law scaling ⟨ρ̂N ⟩ ∼ τQ as the quench time is increased. However, at finite N and for
quench times too large, the KZM breaks down since the dynamics becomes purely adiabatic. (b) Cumulants of the probability
distribution P (ρ). As already observed in Refs. [25, 31, 32, 34, 37], the cumulants scale with the same power law as a function of
the quench time, a hallmark of the underlying Poisson-binomial distribution. (c) The ratios between cumulants fastly approach
the universal values κ2 /κ1 = 0.293 . . . , κ3 /κ1 = 0.0334 . . . , κ3 /κ2 = 0.114 . . . , shown as dotted lines. To make this figure, the
same quench protocol of Fig. 1 was employed.

with
p
ϵk (t) = 2J g 2 (t) + 1 − 2g(t) cos k. (40)

Using {γ, γ † } = 1, one can rewrite the Hamiltonian as


 
1
γk† (t)γk (t)
X
H[g(t)] = ϵk (t) − , (41)
2
k

where the sum runs over both positive and negative k’s.

ADDITIONAL PLOTS

We perform additional numerical analysis of the kink statistics in the driven TFQIM, establishing its regime of
universality. In Fig. 2(a) we plot the average density of kink pairs. As predicted by the conventional KZM, the
−1/2
power-law scaling ⟨ρ̂N ⟩ ∝ τQ holds if quench times are not too long, i.e., before the onset of adiabaticity. From
this figure, it can be checked that the parameter values used to make Fig. 1 lie well within the KZ scaling region,
characterized by the universal power-law scaling with the quench time.
To further establish the regime governed by a universal power-law scaling, in Fig. 2(b)–(c) we compare the scaling of
the first three cumulants of the kink-pair probability distribution P (ρ), as a function of the quench time τQ . Making
use of the fact that the kink-pair distribution is Poisson-Binomial, the explicit expression for the first few cumulants
in the TFQIM reads [25, 31, 34]
X X X
κ1 = pk , κ2 = pk (1 − pk ), κ3 = pk (1 − pk )(1 − 2pk ). (42)
k>0 k>0 k>0

We obtain the excitation probabilities pk numerically without relying on the LZ approximation. The universal full
counting statistics of topological defects beyond KZM [25, 31, 32, 34, 37] dictates that all cumulants exhibit a universal
power-law scaling with the quench rate, with cumulant ratios being constant. Deviations from this universal behavior
are apparent for moderate quench rate satisfying JτQ /ℏ < 10, with high-order cumulants being more sensitive than
the mean to nonuniversal effects. Panels (b)–(c) of Fig. 2 identify the regime in which the universal power-law scaling
holds, not only for the mean but also for higher-order cumulants.
We complete this supplementary analysis with Fig. 3, where we explore the convergence of the rate function in LDT
from numerically-exact simulations using a finite system size. Specifically, we compared the numerically-evaluated
expression − N1 ln P (ρ̄) with the theoretical asymptotic value derived in the main text. Given the finite system size N ,
the random variable ρ̂N can assume only the discrete values 1/N, 2/N, . . . , 1/2. Thus, only a few data points fall into
9

τQ = 100 h̄J −1 N = 1000


60 60
N = 100 (a) τQ = 32 h̄J −1 (b)
N = 180 τQ = 100 h̄J −1
N = 320
τQ = 320 h̄J −1
40 40

− N ρ1KZ log P

− N ρ1KZ log P
N = 560
τQ = 1000 h̄J −1
N = 1000
LZ τQ = 3200 h̄J −1
LZ
20 20

0 0
0 2 4 6 0 2 4 6
ρ̄ ρ̄

FIG. 3. The rate function I¯ can be equivalently accessed by computing directly the logarithm of the probability density
function. However, even for moderately large N , the discreteness of the number of kink pairs results in few data points. To
make this figure, the same quench protocol of Fig. 1 was employed.

the relevant region of ρ̄, when the data is appropriately scaled. By contrast, the computation of I(ρ̄) ¯ makes use of
continuous values of ρ̄, leading to the better collapse of Fig. 1. As a result, the rate function obtained by numerically-
exact simulations converges to the thermodynamic limit for moderate system sizes N = 100 − 1000. Deviations from
the analytical expression derived in the main text simply arise from the fact that the probabilities pk deviate from
the LZ approximation for moderate quench times. This is consistent with the fact that the LZ and the validity of the
KZM power-law scaling are restricted to the limit of slow quenches.

UNIVERSALITY OF LDT BEYOND KZM IN SYSTEMS WITH FAST-DECAYING LONG-RANGE


INTERACTIONS

In this section, we first argue that near the critical region g = 1, the spectrum of excitations in the long-range
deformation of the TFQIM model, i.e., Eq. (43) is the same as the one in the TFQIM for α ≥ 2. Then we show
that the effect of the long-range is to renormalize the quench time by a positive constant. Therefore in the slow
quench limit, the universal KZM scaling and the rate function in the LDT are robust against fast-decaying long-range
interactions. As such, we expect the rate function Eq. (21) in the main text also holds for systems with fast decaying
long-range interactions.

The long-range Ising and Kitaev model

To explore the robustness of our predictions in long-range systems, we consider a long-range Hamiltonian
N N N N −1 l−1
X X JXX O
HLR [g(t)] = −J σjz σj+1
z
− Jg(t) σjx + κl,α (σjz σj+l
z
− σjy σj+l
y
) x
σj+n , (43)
j=1 j=1
2 j=1 n=1
l=2

where
(
1
lα 1 ≤ l ≤ N/2
κl,α = 1 (44)
(N −l)α otherwise,

As for the short-range case, we use periodic boundary conditions σ1a = σN


a
+1 . We note that HLR , like its short-ranged
counterpart, possesses a Z2 symmetry, i.e.,
! !
O O
x x
σi HLR σi = HLR . (45)
i i

As before, upon using the Hadamard gate to rotate the axes

σjx → −σjz , σjy → σjy , σjz → σjx , (46)


10

one obtains
N N N N −1 l−1
X X JXX O
HLR [g(t)] = −J σjx σj+1
x
+ Jg(t) σjz + κl,α (σjx σj+l
x y
− σjy σj+l ) z
(−σj+n ). (47)
j=1 j=1
2 j=1 n=1
l=2

Restricting to the even fermionic parity sector, we perform the Jordan-Wigner transformation defined in Eq. (26) and
obtain
N N   N −1 N −j
† † † 1
κl,α (cj cj+l + c†j+l c†j ).
X X X X
HLRK [g(t)] = −J (cj cj+1 + cj+1 cj ) + 2Jg(t) cj cj − −J (48)
j=1 j=1
2 j=1 l=1

As explained above, the boundary conditions at the level of the fermionic Hamiltonian become anti-periodic: cj =
−cN +j . Eq. (48) is the one-dimensional long-range Kitaev (LRK) chain with power-law decaying superconducting
pairing [49]. After a Fourier transform, Eq. (48) can be rewritten as
X
H[g(t)] = 2J hk [g(t)], (49)
k>0

where

hk [g(t)] ≡ ψk† [(g(t) − cos k)τ z + fαk τ y ] ψk ,


= [g(t) − cos k](c̃†k c̃k − c̃−k c̃†−k ) − ifαk (c̃k c̃−k − c̃†−k c̃†k ), (50)

and the function fαk is defined as


N −1
1 X
fαk ≡ κl,α sin(kl), (51)
2
l=1

with the quasi-momentum taking values k = ± 21 2π 3 2π 2π N 1


N , ± 2 N · · · , ± N ( 2 − 2 ). In the case of α = ∞, it holds that
f∞k = sin k, and the model reduces to the short-range Kitaev chain, i.e., the TFQIM after Jordan-Wigner-transforming
back.
Equation (49) can be further diagonalized through the Bogoliubov transformation in the form

ϵk (t)ψk† Uk† (t)τ z Uk (t)ψk ,


X
H[g(t)] = (52)
k>0

where
p
ϵk (t) = 2J [fαk ]2 + [g(t) − cos k]2 , (53)

and
 
cos [θk (t)/2] −i sin [θk (t)/2]
Uk (t) ≡ , (54)
−i sin [θk (t)/2] cos [θk (t)/2]
with
fαk [g(t) − cos k]
sin θk (t) = p , cos θk (t) = p . (55)
[fαk ] + [g(t) − cos k]2
2 [fαk ]2 + [g(t) − cos k]2
Denoting
   
γk (t) c̃k
† ≡ Uk (t) † , (56)
γ−k (t) c̃−k

the Hamiltonian can be rewritten as


 
X † 1
H[g(t)] = ϵk (t) γk (t)γk (t) − , (57)
2
k

where the summation is over all the modes, both positive and negative.
11

Properties of fαk

Let us list a few properties of fαk . First of all, fαk is odd in k, i.e.,

fαk = −fα,−k . (58)

Note that in the thermodynamic limit, the quasi-momentum k ∈ [−π, π]. While limk→±π fαk = 0, one should note
that limk→0 fα (k) does not necessarily vanish. In general, if κl,α decays slowly as l increases, fαk may become singular
at k = 0, as shown in Ref. [50]. In fact, one can express fαk in terms of the polylogarithm function

fαk = −i Liα (eik ) − Liα (e−ik ) .


 
(59)

/ N+ , fαk ∝ 1/k 1−α [49], which is adopted by


Using the expansion of polylogarithm function, one can show that α ∈
Ref. [51] to discuss KZM in the long-range Kitaev chain. We note that this scaling behavior can be alternatively
/ N+ . Thus, in view of Eq. (58),
proved using the Euler-Maclaurin formula for any α ≥ 0 without the constraint α ∈
fαk allows the following expansion near k = 0,
ξ0
fαk = + ξ1 k + O(k 3 ). (60)
k 1−α

LDT for the kink density with fast-decaying long-range interactions

The phase diagram of the long-range Kitaev chain is very rich [49, 52], including in particular topological phase
transitions. With the spectrum Eq. (53), for α > 1 one can easily verify that the gap closes only when g = ±1, with
the gapless mode being k = 0. Furthermore, according to Eq. (60), if one only considers α ≥ 2, then

fαk = k + O(k 3 ). (61)

This is the same behaviour of the spectrum of the TFQIM near k = 0, where fαk = sin k. Clearly, for both −1 < g < 1
and g > 1, one can continuously deform α from 2 to ∞ without closing the gap. As we will show below, the kink
density statistics and the large deviations theory are the same in the slow quench limit. We consider the ramp in the
main text, i.e.,
 
t
g(t) = gc 1 − , (62)
τQ

starting at an initial time before −τQ , outside the freeze-out region [3], and ending at τQ , whence g(τQ ) = 0. We
investigate the statistics of kink density and work with the adiabatic basis at t = −∞,

|0⟩, |k, −k⟩ = c†k c†−k |0⟩ = γk† (−∞)γ−k (−∞)|0⟩, (63)
|k, 0⟩ = c†k |0⟩ = γk† (−∞)|0⟩, |0, −k⟩ = c†−k |0⟩ = γ−k (−∞)|0⟩. (64)

For each mode Hamiltonian Jhk [g(t)] preserves the fermionic parity, and excitations remain in the subspace spanned
by |0⟩, |k, −k⟩. Projecting Eq. (50) onto this subspace, one obtains the effective Hamiltonian
 
[g(t) − cos k] −ifαk
Jhk [g(t)] = J . (65)
ifαk −[g(t) − cos k]

The excitation generated during the process can be characterized by the Landau-Zener formula, provided |k| < π/2:
 
D

E 2πJτQ 2
pk = γk (τQ )γk (τQ ) ≈ exp − fα (k) , (66)

where the average is over the state of the system at the final time t = τQ . The scaled cumulant generating function
becomes resembles that in the TFQIM
Z π   
1 θ 2πJτQ 2
λ(θ) = dk ln 1 + (e − 1) exp − fα (k) , (67)
2π 0 ℏ
12

where we have extended the integral up to π, an approximation with exponential accuracy. When α ≥ 2, according
to Eq. (60),

fαk = ξk + O(k 3 ). (68)

In the slow quench limit where τQ → ∞, one can perform a Gaussian approximation, leading to

−2πJτQ ξ 2 k 2
D E  
pk = γk† (τQ )γk (τQ ) = exp , (69)

and thus
π
−2πJτQ ξ 2 k 2
  
1
Z
λ(θ) = dk ln 1 + (eθ − 1) exp . (70)
2π 0 ℏ

Comparing the results for TFQIM, we see that the effect of the fast decaying long-range interactions is to renormalize
the quench time by a positive constant, keeping the structure of the scaled cumulant generating function otherwise
unchanged.

You might also like