Kibble–Zurek scaling in the quantum Ising chain with a time-periodic perturbation
Abstract
We consider the time-dependent transverse field Ising chain with time-periodic perturbations. Without perturbations, this model is one of the famous models that obeys the scaling in the adiabatic limit predicted by the quantum Kibble–Zurek mechanism (QKZM). However, it is known that when oscillations are added to the system, the non-perturbative contribution becomes larger and the scaling may break down even if the perturbation is small. Therefore, we analytically analyze the density of defects in the model and discuss how much the oscillations affect the scaling. As a result, although the non-perturbative contribution does not become zero in the adiabatic limit, the scaling does not change from the prediction of the QKZM. This indicates that the QKZM is robust to the perturbations.
I Introduction
The Kibble–Zurek mechanism (KZM) is a fundamental concept that explains the formation of topological defects during non-equilibrium phase transitions. The original theory was proposed in the context of cosmology, where the universe underwent a symmetry-breaking phase transition in the early stages of its evolution Kibble (1976, 1980). Since then, the KZM has been adapted to condensed matter systems, especially in the study of quantum phase transitions Zurek (1985, 1993, 1996). The KZM has been experimentally validated in a variety of systems, such as the superfluid helium experiments Hendry et al. (1994) and the superconductor experiments Monaco et al. (2002, 2003); Maniv et al. (2003).
The quantum Kibble–Zurek mechanism (QKZM), is an extension of the KZM that incorporates quantum effects. In the context of phase transitions, quantum corrections can lead to significant modifications in the physics near the critical point, giving rise to novel phenomena. The QKZM has been developed to investigate how these quantum corrections affect the predictions of the KZM and take into account the quantum fluctuations near the critical point. The QKZM has already been studied Dziarmaga (2010); Sinha et al. (2019); Sadhukhan et al. (2020); Dutta and Dutta (2017); Polkovnikov et al. (2011); Rossini and Vicari (2021); Cincio et al. (2007); Saito et al. (2007a); Sengupta et al. (2008); Sen et al. (2008); Dziarmaga et al. (2008); Divakaran and Dutta (2009); Honer et al. (2010); Zurek (2013); Cucchietti et al. (2007); Mukherjee et al. (2007); Nag et al. (2013); Del Campo and Zurek (2014); Dutta et al. (2015); Cherng and Levitov (2006); Mukherjee et al. (2008); Rams et al. (2019); Revathy and Divakaran (2020); Rossini and Vicari (2020); Hódsági and Kormos (2020); Białończyk and Damski (2020); Nowak and Dziarmaga (2021); Kou and Li (2022); Zurek et al. (2005); Kells et al. (2014); Heyl et al. (2013); Dziarmaga (2005); Coldea et al. (2010); Kinross et al. (2014); King et al. (2022); Sachdev (1999); Sun et al. (2022); Zeng et al. (2023); Yan et al. (2021); Fubini et al. (2007); Bermudez et al. (2009); Mukherjee and Dutta (2009) and observed in many experiments Ulm et al. (2013); Pyka et al. (2013); Monaco et al. (2006); Sadler et al. (2006); Chen et al. (2011); Griffin et al. (2012); Lamporesi et al. (2013); Navon et al. (2015); Braun et al. (2015); Chomaz et al. (2015); Anquez et al. (2016); Clark et al. (2016); Keesling et al. (2019); Baumann et al. (2011); Xu et al. (2014); Meldgin et al. (2016); Clark et al. (2016); Chen et al. (2020); Gardas et al. (2018); Gong et al. (2016); Cui et al. (2016); Bando et al. (2020); Li et al. (2023); Zamora et al. (2020). In the study of the QKZM, a theoretical approach based on the one-dimensional transverse field Ising model is sometimes used Nag et al. (2013); Del Campo and Zurek (2014); Revathy and Divakaran (2020); Cherng and Levitov (2006); Mukherjee et al. (2007, 2008); Kells et al. (2014); Rams et al. (2019); Rossini and Vicari (2020); Hódsági and Kormos (2020); Białończyk and Damski (2020); Nowak and Dziarmaga (2021); Kou and Li (2022); Zurek et al. (2005); Heyl et al. (2013); Dziarmaga (2005); Coldea et al. (2010); Kinross et al. (2014); King et al. (2022); Sachdev (1999); Sun et al. (2022); Zeng et al. (2023); Yan et al. (2021); Fubini et al. (2007); Bermudez et al. (2009); Dutta et al. (2015); Gardas et al. (2018); Chen et al. (2020); Bando et al. (2020); Cui et al. (2016); Gong et al. (2016); Li et al. (2023); Zamora et al. (2020). The system begins with all spins aligned which corresponds to the ground state at infinite past. As the system’s energy evolves linearly with time such as , a phase transition occurs, resulting in the emergence of defects. According to the QKZM, the density of defects is generally given by , where is the dimension of the system, is the dynamic exponent, and is the correlation length exponent. In the one-dimensional transverse field Ising model, these exponents are given by Sachdev (1999); Dziarmaga (2010).
This scaling is an estimate of the computational time for quantum annealing, since it corresponds to the probability of successfully obtaining the ground state. Therefore, it is important to investigate what happens to the scaling when the linear ramp is deviated or perturbations are added. The robustness to these changes has been investigated in several previous studies. For example, when the spin-spin coupling is changed alternately, the density of defects includes a factor that decays exponentially and is subject to large corrections Yan et al. (2021). Furthermore, the numerical simulation shows that the density of defects increases due to the effect of white noise. Fubini et al. (2007). It is also known to exhibit nontrivial behavior when oscillations are added as perturbations, but the effect of the perturbation on the scaling is not derived analytically Mukherjee and Dutta (2009).
This nontrivial behavior, caused by adding an oscillation term to a linear ramp, is also observed in other fields. The Franz–Keldysh effect, originally proposed in the 1950s Franz (1958); Keldysh (1958); Tharmalingam (1963); Callaway (1963), is an important phenomenon observed in semiconductors when subjected to strong electric fields. This analysis method is also applied to the dynamically assisted Schwinger mechanism Schützhold et al. (2008), the extension of the Schwinger mechanism Heisenberg and Euler (1936); Weisskopf (1936); Schwinger (1951) which explains the phenomenon in quantum electrodynamics where electron-positron pairs are generated in a vacuum by the application of an electric field. Recent research on the dynamically assisted Schwinger mechanism calculates the particle pair creation rate analytically using the Furry picture (FP) Furry (1951) for a system in which an oscillating electric field is perturbatively added to a strong constant electric field. It has been suggested that the perturbed electric field allows non-adiabatic contributions to appear Taya (2019); Huang and Taya (2019).
In this paper, we consider the transverse field Ising model which depends linearly on time, with time-periodic perturbations and investigate how the addition of oscillations affects the phase transition behavior in the QKZM framework. The analytical expression of the density of defects is derived using the Landau–Zener–Stükelberg–Majorana (LZSM) model Landau (1932); Zener (1932); Stückelberg (1932); Majorana (1932). The LZSM model describes a two-level system whose Hamiltonian has diagonal elements that are linearly dependent on time, while the off-diagonal elements are time-independent. The calculations are performed using the perturbation and FP formulation to derive analytical solutions with approximations. The perturbation approximation is valid for the non-adiabatic region, while the FP formulation is valid for the adiabatic region.
The structure of this paper is as follows. In Sec. II, we analyze the contribution of time-periodic perturbations to the two-level system for the calculations in the next section. In this section, we introduce the LZSM model and analyze the dynamics of the system when time-periodic perturbations are added, using the perturbation theory and the FP formalism. Furthermore, we confirm that these approximate solutions are in good agreement with numerical calculations. In Sec. III, we consider a time-dependent transverse field Ising chain with a time-periodic perturbation. We determine how the density of defects changes when a time-periodic perturbation is applied to the diagonal or off-diagonal elements, and compare the results with those obtained by the QKZM. In Sec. IV, we summarize the discussion so far.
II Transition Probabilities of Two-Level System
This section focuses on the treatment of the LZSM model in the presence of an external oscillation field to analyze many-body systems later. There have been some previous studies on this topic Mullen et al. (1989); Kayanuma and Mizumoto (2000). Here, we introduce the LZSM model first, and the perturbation theory and the FP formulation for the LZSM model. Finally, we evaluate the validity of these approximations.
II.1 LZSM model
The LZSM model is described by the two-level Hamiltonian
(1) |
In this model, if a state was an instantaneous eigenstate in the infinite past, the probability of transitioning to another instantaneous eigenstate in the infinite future is given by
(2) |
where the natural units are used. When is significantly larger than , the system is considered adiabatic, resulting in a small transition probability. Conversely, when is significantly smaller than , the system is characterized as non-adiabatic, leading to a large transition probability. Recent studies of the LZSM model have investigated the dynamics under various conditions, including the presence of external oscillating perturbations Mullen et al. (1989); Malla and Raikh (2018); Wubs et al. (2005, 2006); Saito et al. (2006, 2007b); Zueco et al. (2008); Ashhab (2014, 2016); Sinitsyn and Li (2016); Sun and Sinitsyn (2016); Kayanuma and Mizumoto (2000). The perturbation approach derives the approximate formula for the LZSM model when the diagonal elements of the Hamiltonian are small Mullen et al. (1989).
II.2 perturbation theory
In this section, we consider the two-level time-dependent Hamiltonian
(3) | ||||
(4) | ||||
(5) |
which is the LZSM model with oscillations of magnitude and in the diagonal and off-diagonal elements, respectively. The initial state is assumed to be , where holds. The goal is to obtain the transition probability at the final time . We note that the transition probabilities were obtained approximately when either or is with perturbation theory Mullen et al. (1989). Changing the frame with the unitary operator
(6) |
the Schrödinger equation becomes
(7) |
where we define and
(8) | |||
Here, we used the formula
(9) |
where is the Bessel function of the first kind and the basis of the matrix is , where holds. We express the state in this basis as , and these variables satisfy
(10) | ||||
(11) |
The initial conditions on these variables can be regarded as and , and the transition probability can be expressed as . We introduce the dimensionless parameters and , and we define . By successive substitutions, we obtain the following result
(12) | ||||
(13) | ||||
(14) | ||||
(15) |
where we define the step function
(19) |
Here, we assume and are small enough that this approximation is valid in the non-adiabatic region. Finally, the transition probability is approximately
(20) |
We note that the transition probability must be periodic with because
(21) |
holds and the period is .
II.3 Furry Picture
Next, we decompose the Hamiltonian (3) as in
(22) | ||||
(23) | ||||
(24) |
where is the Hamiltonian of the LZSM model. Let be the time-evolution operator of and we define . If the is sufficiently small, we can approximate the time-evolution operator by first-order:
(25) |
Then, the transition probability becomes
(26) | ||||
(27) |
We note that we need to consider up to the second-order perturbation if we approximate the transition probability by the second-order of :
(28) | ||||
(29) |
However, we assume that the last term is negligible. This assumption is justified in the adiabatic limit because the term contains an exponentially small term in the limit. The above method is called the Furry picture formalism Furry (1951).
With this formalism, the transition probability in the adiabatic limit is
(30) | ||||
(31) | ||||
(32) | ||||
(33) |
where is the regularized confluent hypergeometric function of the first kind. We derive the probability in Appendix.A and the probability without the adiabatic limit is (106). The first-order approximation in this method is valid in the region where both are small. Unlike perturbation theory, this method does not treat as small values, but rather assumes that it takes on large values. In this way, this method calculates non-perturbative effects on .
II.4 numerical calculation
In this subsection, we compare the approximate formula (20) and (106) or (33) with the results of numerical solution of the Schrödinger equation.
First, consider the case of the non-adiabatic limit and the case . In this case, only the off-diagonal component has an oscillating term and the transition probability (20) can be expressed as
. |
The result (II.4) is also derived in the previous study Mullen et al. (1989). We note, however, that a factor of is missing in the third term of equation (7) in Mullen et al. (1989).
The numerical results in this case are shown in Fig. 1(a). In the region where is small, the numerical calculation and the approximate expression (II.4) are in good agreement. On the other hand, as increases, the contribution of , which is ignored in the approximate expression (II.4), increases, resulting in deviation from the numerical calculation.
Next, consider the case . From (20), the transition probability can be expressed as
. |
The numerical results in this case are shown in Fig. 1(b). Here, the sum was calculated in the range of . In this case, the sum is large enough that even if is not small, the numerical calculation and the approximate expression (II.4) are in good agreement. In Fig. 1(b), the transition probabilities show a simple behavior as increases. This corresponds to the region where is sufficiently small. In this limit, (II.4) yields
(34) |
This result is also derived in the previous study Mullen et al. (1989).
Finally, consider the case and . The numerical results for this case are shown in Fig. 1(c). In this case, the numerical calculation and the approximate formula (20) agree well even for large values of because the sums are sufficiently large.
Next, we show the validity of (106) in the adiabatic process. First, in the case of , the results of the numerical calculations are compared with those of the expression (106) in Fig. 2(a). It can be seen that in the region where is small, the results are in good agreement with the numerical calculations. The dashed line in the figure represents the LZSM transition probability (2). Although this probability is sufficiently small in the adiabatic limit, it can be seen that there are parameter regions where the transition probabilities are much larger than the LZSM transition probability for due to the effects of the oscillations.
Next, in the case of , the results of the numerical calculations are compared with those of (106) in Fig. 2(b). It can be seen that in the region where is sufficiently small, the results are in good agreement with the numerical calculations. In this case, as in the previous case, there are parameter regions where the transition probabilities are much larger than the LZSM transition probability due to the oscillations.
In addition, Fig. 2(c) compares the results of the numerical calculations with those of (106) in the case of and . In this case, we can see that (106) is in good agreement with the numerical calculation in the region where is sufficiently small due to the small value of .
Finally, we check that (33) is consistent with (106) in the adiabatic limit. The results are shown in Fig. 3. In this case, it can be seen that (106) is consistent with (33), especially in regions where is sufficiently small.

III TRANSVERSE ISING CHAIN with time-periodic perturbation
Next, we consider the transverse field Ising model which depends linearly on time, with time-periodic perturbations. For this model, there is a previous study that investigated the model numerically Mukherjee and Dutta (2009). However, this study only shows that the transfer matrix method Kayanuma and Mizumoto (2000) agrees with the numerical calculations. In the following, we consider the case where the perturbations are uniformly contained in the diagonal or off-diagonal elements.
III.1 perturbation in the diagonal elements
We consider the time-dependent Hamiltonian
(35) | ||||
(36) |
where we impose the periodic boundary condition
(37) |
This Hamiltonian has symmetry and only the space to which the ground state belongs will be considered from now on. Here, we introduce the spinless fermion operators using the Jordan–Wigner(JW) transformation
(38) |
and we consider the Fourier expansion of the operators
(39) |
where . In the Heisenberg picture, these operators satisfy
(40) | |||
(41) |
The eigenvalues of the Hamiltonian in (III.1) are shown in Fig. 4. It can be seen that when , the energy gap is small, corresponding to the non-adiabatic region where non-adiabatic transitions occur, while in the other region, the energy gap is large, corresponding to the adiabatic region.
The initial state is set to be the ground state at : . The final time is set to and we calculate the expectation value
(42) | ||||
(43) | ||||
(44) |
where and we used the fact that the initial state can be written as . The solution of (III.1) can be expressed as
(45) |
where holds. Then, the expectation value becomes
(46) | ||||
(47) |
In the thermodynamic limit, the normalized expectation value can be expressed as
(48) | ||||
(49) |
In the model considered in this study, the phase transition points exist at times satisfying . However, as shown in (III.1), these phase transition points do not produce interference effects as discussed in previous studies Kou and Li (2022). Therefore, we will assume that for the current discussion.

First, we consider a non-adiabatic region. The non-adiabatic region corresponds to the situation , where . In this region, we obtain from (II.4)
(50) |
where we define . In the adiabatic limit where is sufficiently large, there is a non-adiabatic region only near . In the vicinity of , we obtain
(51) | ||||
(52) |
and in the vicinity of , we get
(53) | ||||
(54) |
We note that has the finite value near in this region if is large enough.
The distribution of is shown in Fig. 5. This figure shows that in addition to the transitions at predicted by the KZ mechanism, other transitions occur around them, which is the result from the time-periodic perturbation. We note that if is larger, the transitions in the adiabatic region occur closer to these vicinities as long as and are fixed.
From the above discussion, we obtain the expectation value approximately
(57) | ||||
(58) | ||||
(59) |
From (58), we see that the first term is proportional to . This corresponds to the part of the QKZM where perturbative oscillatory effects are added to the non-adiabatic transition. The second term is the contribution from the non-perturbative effect. In fact, this term is also proportional to . This can be seen as follows. First, for sufficiently large , has a value of only , so can be transformed to
(60) |
Transforming to and changing the upper bound of the integral to because the integrand function has no value at , we obtain
(61) |
This approximation and the original definition (59) are plotted in Fig. 6. The fact that both agree where is large indicates that this approximation is correct and that the non-perturbative contribution also shows a scaling of . It can also be seen numerically that the peak of appears where is satisfied. This can be interpreted as a result of the resonance phenomenon. Fig. 7 shows the dependence of the coefficient of in (61) on . It can be seen that the contribution of this non-perturbative term increases as the frequency and the amplitude increases.
To confirm that the derived equation (58) is correct as an approximation, we finally check it with numerical calculation when is finite as shown in the Fig. 8. Because the contribution of is large, it can be seen that the density of defects behaves differently from the case of no oscillations. As discussed, however, the scaling of does not change because this is the same in the non-adiabatic and adiabatic regions. This means that the QKZM is robust to the time-periodic perturbations. To verify that the finite discussed here is sufficiently large, we compared the integral (59) with the sum of (56) and the result is shown in Fig. 9. It can be seen that is sufficient to be regarded as the thermodynamic limit.





III.2 perturbation in the off-diagonal elements
Next, we consider another time-dependent Hamiltonian
(62) | ||||
(63) | ||||
(64) |
where we impose the periodic boundary condition.
As before, introducing spinless fermions by the JW transformation yields
(65) | |||
(66) |
First, we consider a non-adiabatic region: . In this region, the transition amplitude becomes
, |
where we define . We note that this expression becomes the same with (50) in the adiabatic limit if the amplitude is small enough. On the other hand, in the adiabatic region , we obtain
(67) |
These expressions of have values only at if is sufficiently large (Fig. 10), which is easy to see from the asymptotic expansion of . Furthermore, the first term in the absolute value in (67) is the largest contribution compared to the others because we focus on the regions at . This shows that the integral of scales with , as in . From the above discussion, the analytical expression for the density of defects can be expressed as in (58).
We check it with a numerical calculation when is finite. In the Fig. 11, we compared the density of defects obtained by numerically solving Schrödinger equation with an approximate analytical expression. As in the previous subsection, can be regarded as the thermodynamic limit. This figure shows that the density of defects behaves differently compared to the case without oscillations. However, the scaling of the non-perturbative contribution is , which is not different from the scaling predicted by the QKZM, indicating the robustness of the scaling. In addition, the resonance phenomenon was observed in the diagonal oscillation, but when there was oscillation in the off-diagonal elements, the resonance phenomenon was canceled out by the contribution of the second term in the absolute value of (67).


IV Conclusion
The QKZM is currently attracting attention, and scaling laws for models beyond the simple setting of an isolated system and linear sweep are also of interest. In this paper, we consider a model in which an oscillating external field is perturbatively added in addition to the usual linear linear sweep. In such a setting, it was found that it is necessary to consider not only a perturbative correction term for transitions in the non-adiabatic region, as in the usual QKZM, but also a non-perturbative correction in the adiabatic region. Moreover, although the power spectrum of transition probability is different between with and without oscillation, the non-perturbative correction term also scales as in the adiabatic limit, as in the usual QKZM, indicating the robustness of the QKZM with respect to the scaling law.
In the present study, the high symmetry in the model allows for the analytical discussion. The scaling laws of the QKZM have also been investigated for other models such as the spin glass model Dziarmaga (2006); Caneva et al. (2007); Suzuki (2011). The relation between symmetry and the effect of time-periodic perturbations on the QKZM is a subject for future work. Furthermore, we need to investigate the robustness of other quantities, such as kink-kink correlations Cincio et al. (2007); Del Campo (2018); Nowak and Dziarmaga (2021); Roychowdhury et al. (2021); Mayo et al. (2021); Dziarmaga and Rams (2022).
Acknowledgement
We thank H. Nakazato, M. Fujiwara, and G. Kato for helpful discussions.
Appendix A Furry Picture
In this section, we use the FP to obtain the transition probability (33). In the following discussion, we use these relations
(68) | ||||
(69) | ||||
(70) | ||||
(71) |
where is the parabolic cylinder function, is the regularized confluent hypergeometric function of the first kind, and is the confluent hypergeometric function of the second kind. We note that these relations are derived from the integral expressions of the special functions Gradshteyn and Ryzhik (2014) and applicable only when .
We consider the dimensionless Hamiltonian
(72) |
where we define and . The time-evolution operator for the LZSM Hamiltoinan () is
(73) | ||||
(74) | ||||
(75) | ||||
(76) | ||||
(77) |
We note that
(78) |
hold.
From this, the perturbation term can be written as
(79) | ||||
(80) | ||||
(81) |
where the argument was omitted. To obtain the transition probability, we need to calculate
(82) |
Hereafter, the argument is omitted. When , the diagonal part of the integral becomes
(83) | ||||
(84) | ||||
(85) | ||||
(86) |
Similarly, the off-diagonal part of the integral becomes
(87) |
In the adiabatic limit , the off-diagonal part can be written more simply as
(88) |
Next, we consider the case . In this case, the integrals of diagonal part and off-diagonal part becomes
(89) | ||||
(90) |
where we define
(91) | ||||
(92) | ||||
(94) | ||||
(95) | ||||
(96) | ||||
(97) | ||||
(98) | ||||
(99) | ||||
(100) | ||||
(101) | ||||
(102) | ||||
(103) |
In the adiabtic limit, the off-diagonal part becomes
(104) |
From the above discussion, the transition probability in the adiabatic limit under , and becomes
(105) | |||
(106) | |||
(107) | |||
(108) | |||
(109) | |||
(110) | |||
(111) |
where we use these relation
(112) | ||||
(113) |
which hold in .
References
- Kibble (1976) T. W. Kibble, J. Phys. A: Math. Gen. 9, 1387 (1976).
- Kibble (1980) T. W. Kibble, Phys. Rep. 67, 183 (1980).
- Zurek (1985) W. H. Zurek, Nature 317, 505 (1985).
- Zurek (1993) W. H. Zurek, Acta Phys. Pol. B 24, 1301 (1993).
- Zurek (1996) W. H. Zurek, Phys. Rep. 276, 177 (1996).
- Hendry et al. (1994) P. Hendry, N. S. Lawson, R. Lee, P. V. McClintock, and C. Williams, Nature 368, 315 (1994).
- Monaco et al. (2002) R. Monaco, J. Mygind, and R. Rivers, Phys. Rev. Lett. 89, 080603 (2002).
- Monaco et al. (2003) R. Monaco, J. Mygind, and R. J. Rivers, Phys. Rev. B 67, 104506 (2003).
- Maniv et al. (2003) A. Maniv, E. Polturak, and G. Koren, Phys. Rev. Lett. 91, 197001 (2003).
- Dziarmaga (2010) J. Dziarmaga, Adv. Phys. 59, 1063 (2010).
- Sinha et al. (2019) A. Sinha, M. M. Rams, and J. Dziarmaga, Phys. Rev. B 99, 094203 (2019).
- Sadhukhan et al. (2020) D. Sadhukhan, A. Sinha, A. Francuz, J. Stefaniak, M. M. Rams, J. Dziarmaga, and W. H. Zurek, Phys. Rev. B 101, 144429 (2020).
- Dutta and Dutta (2017) A. Dutta and A. Dutta, Phys. Rev. B 96, 125113 (2017).
- Polkovnikov et al. (2011) A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalattore, Rev. Mod. Phys. 83, 863 (2011).
- Rossini and Vicari (2021) D. Rossini and E. Vicari, Phys. Rep. 936, 1 (2021).
- Cincio et al. (2007) L. Cincio, J. Dziarmaga, M. M. Rams, and W. H. Zurek, Phys. Rev. A 75, 052321 (2007).
- Saito et al. (2007a) H. Saito, Y. Kawaguchi, and M. Ueda, Phys. Rev. A 76, 043613 (2007a).
- Sengupta et al. (2008) K. Sengupta, D. Sen, and S. Mondal, Phys. Rev. Lett. 100, 077204 (2008).
- Sen et al. (2008) D. Sen, K. Sengupta, and S. Mondal, Phys. Rev. Lett. 101, 016806 (2008).
- Dziarmaga et al. (2008) J. Dziarmaga, J. Meisner, and W. H. Zurek, Phys. Rev. Lett. 101, 115701 (2008).
- Divakaran and Dutta (2009) U. Divakaran and A. Dutta, Phys. Rev. B 79, 224408 (2009).
- Honer et al. (2010) J. Honer, H. Weimer, T. Pfau, and H. P. Büchler, Phys. Rev. Lett. 105, 160404 (2010).
- Zurek (2013) W. H. Zurek, J. Phys. Condens. Matter 25, 404209 (2013).
- Cucchietti et al. (2007) F. M. Cucchietti, B. Damski, J. Dziarmaga, and W. H. Zurek, Phys. Rev. A 75, 023603 (2007).
- Mukherjee et al. (2007) V. Mukherjee, U. Divakaran, A. Dutta, and D. Sen, Phys. Rev. B 76, 174303 (2007).
- Nag et al. (2013) T. Nag, A. Dutta, and A. Patra, Int. J. Mod. Phys. B 27, 1345036 (2013).
- Del Campo and Zurek (2014) A. Del Campo and W. H. Zurek, Int. J. Mod. Phys. A 29, 1430018 (2014).
- Dutta et al. (2015) A. Dutta, G. Aeppli, B. K. Chakrabarti, U. Divakaran, T. F. Rosenbaum, and D. Sen, Quantum phase transitions in transverse field spin models: from statistical physics to quantum information (Cambridge University Press, 2015).
- Cherng and Levitov (2006) R. Cherng and L. Levitov, Phys. Rev. A 73, 043614 (2006).
- Mukherjee et al. (2008) V. Mukherjee, A. Dutta, and D. Sen, Phys. Rev. B 77, 214427 (2008).
- Rams et al. (2019) M. M. Rams, J. Dziarmaga, and W. H. Zurek, Phys. Rev. Lett. 123, 130603 (2019).
- Revathy and Divakaran (2020) B. Revathy and U. Divakaran, J. Stat. Mech. Theory Exp. 2020, 023108 (2020).
- Rossini and Vicari (2020) D. Rossini and E. Vicari, Phys. Rev. Research 2, 023211 (2020).
- Hódsági and Kormos (2020) K. Hódsági and M. Kormos, SciPost Phys. 9, 055 (2020).
- Białończyk and Damski (2020) M. Białończyk and B. Damski, Phys. Rev. B 102, 134302 (2020).
- Nowak and Dziarmaga (2021) R. J. Nowak and J. Dziarmaga, Phys. Rev. B 104, 075448 (2021).
- Kou and Li (2022) H.-C. Kou and P. Li, Phys. Rev. B 106, 184301 (2022).
- Zurek et al. (2005) W. H. Zurek, U. Dorner, and P. Zoller, Phys. Rev. Lett. 95, 105701 (2005).
- Kells et al. (2014) G. Kells, D. Sen, J. Slingerland, and S. Vishveshwara, Phys. Rev. B 89, 235130 (2014).
- Heyl et al. (2013) M. Heyl, A. Polkovnikov, and S. Kehrein, Phys. Rev. Lett. 110, 135704 (2013).
- Dziarmaga (2005) J. Dziarmaga, Phys. Rev. Lett. 95, 245701 (2005).
- Coldea et al. (2010) R. Coldea, D. Tennant, E. Wheeler, E. Wawrzynska, D. Prabhakaran, M. Telling, K. Habicht, P. Smeibidl, and K. Kiefer, Science 327, 177 (2010).
- Kinross et al. (2014) A. Kinross, M. Fu, T. Munsie, H. Dabkowska, G. Luke, S. Sachdev, and T. Imai, Phys. Rev. X 4, 031008 (2014).
- King et al. (2022) A. D. King, S. Suzuki, J. Raymond, A. Zucca, T. Lanting, F. Altomare, A. J. Berkley, S. Ejtemaee, E. Hoskinson, S. Huang, et al., Nat. Phys. 18, 1324 (2022).
- Sachdev (1999) S. Sachdev, Phys. World 12, 33 (1999).
- Sun et al. (2022) H.-Y. Sun, Z.-Y. Ge, and H. Fan, arXiv preprint arXiv:2202.07892 (2022).
- Zeng et al. (2023) H.-B. Zeng, C.-Y. Xia, and A. Del Campo, Phys. Rev. Lett. 130, 060402 (2023).
- Yan et al. (2021) B. Yan, V. Y. Chernyak, W. H. Zurek, and N. A. Sinitsyn, Phys. Rev. Lett. 126, 070602 (2021).
- Fubini et al. (2007) A. Fubini, G. Falci, and A. Osterloh, New J. Phys. 9, 134 (2007).
- Bermudez et al. (2009) A. Bermudez, D. Patane, L. Amico, and M. Martin-Delgado, Phys. Rev. Lett. 102, 135702 (2009).
- Mukherjee and Dutta (2009) V. Mukherjee and A. Dutta, J. Stat. Mech. 2009, P05005 (2009).
- Ulm et al. (2013) S. Ulm, J. Roßnagel, G. Jacob, C. Degünther, S. Dawkins, U. Poschinger, R. Nigmatullin, A. Retzker, M. Plenio, F. Schmidt-Kaler, et al., Nat. Commun. 4, 2290 (2013).
- Pyka et al. (2013) K. Pyka, J. Keller, H. Partner, R. Nigmatullin, T. Burgermeister, D. Meier, K. Kuhlmann, A. Retzker, M. B. Plenio, W. Zurek, et al., Nat. Commun. 4, 2291 (2013).
- Monaco et al. (2006) R. Monaco, J. Mygind, M. Aaroe, R. Rivers, and V. Koshelets, Phys. Rev. Lett. 96, 180604 (2006).
- Sadler et al. (2006) L. Sadler, J. Higbie, S. Leslie, M. Vengalattore, and D. Stamper-Kurn, Nature 443, 312 (2006).
- Chen et al. (2011) D. Chen, M. White, C. Borries, and B. DeMarco, Phys. Rev. Lett. 106, 235304 (2011).
- Griffin et al. (2012) S. M. Griffin, M. Lilienblum, K. T. Delaney, Y. Kumagai, M. Fiebig, and N. A. Spaldin, Phys. Rev. X 2, 041022 (2012).
- Lamporesi et al. (2013) G. Lamporesi, S. Donadello, S. Serafini, F. Dalfovo, and G. Ferrari, Nat. Phys. 9, 656 (2013).
- Navon et al. (2015) N. Navon, A. L. Gaunt, R. P. Smith, and Z. Hadzibabic, Science 347, 167 (2015).
- Braun et al. (2015) S. Braun, M. Friesdorf, S. S. Hodgman, M. Schreiber, J. P. Ronzheimer, A. Riera, M. Del Rey, I. Bloch, J. Eisert, and U. Schneider, Proc. Natl. Acad. Sci. USA 112, 3641 (2015).
- Chomaz et al. (2015) L. Chomaz, L. Corman, T. Bienaimé, R. Desbuquois, C. Weitenberg, S. Nascimbene, J. Beugnon, and J. Dalibard, Nat. Commun. 6, 6162 (2015).
- Anquez et al. (2016) M. Anquez, B. Robbins, H. Bharath, M. Boguslawski, T. Hoang, and M. Chapman, Phys. Rev. Lett. 116, 155301 (2016).
- Clark et al. (2016) L. W. Clark, L. Feng, and C. Chin, Science 354, 606 (2016).
- Keesling et al. (2019) A. Keesling, A. Omran, H. Levine, H. Bernien, H. Pichler, S. Choi, R. Samajdar, S. Schwartz, P. Silvi, S. Sachdev, et al., Nature 568, 207 (2019).
- Baumann et al. (2011) K. Baumann, R. Mottl, F. Brennecke, and T. Esslinger, Phys. Rev. Lett. 107, 140402 (2011).
- Xu et al. (2014) X.-Y. Xu, Y.-J. Han, K. Sun, J.-S. Xu, J.-S. Tang, C.-F. Li, and G.-C. Guo, Phys. Rev. Lett. 112, 035701 (2014).
- Meldgin et al. (2016) C. Meldgin, U. Ray, P. Russ, D. Chen, D. M. Ceperley, and B. DeMarco, Nat. Phys. 12, 646 (2016).
- Chen et al. (2020) Z. Chen, J.-M. Cui, M.-Z. Ai, R. He, Y.-F. Huang, Y.-J. Han, C.-F. Li, and G.-C. Guo, Phys. Rev. A 102, 042222 (2020).
- Gardas et al. (2018) B. Gardas, J. Dziarmaga, W. H. Zurek, and M. Zwolak, Sci. Rep. 8, 4539 (2018).
- Gong et al. (2016) M. Gong, X. Wen, G. Sun, D.-W. Zhang, D. Lan, Y. Zhou, Y. Fan, Y. Liu, X. Tan, H. Yu, et al., Sci. Rep. 6, 22667 (2016).
- Cui et al. (2016) J.-M. Cui, Y.-F. Huang, Z. Wang, D.-Y. Cao, J. Wang, W.-M. Lv, L. Luo, A. Del Campo, Y.-J. Han, C.-F. Li, et al., Sci. Rep. 6, 33381 (2016).
- Bando et al. (2020) Y. Bando, Y. Susa, H. Oshiyama, N. Shibata, M. Ohzeki, F. J. Gómez-Ruiz, D. A. Lidar, S. Suzuki, A. Del Campo, and H. Nishimori, Phys. Rev. Research 2, 033369 (2020).
- Li et al. (2023) B.-W. Li, Y.-K. Wu, Q.-X. Mei, R. Yao, W.-Q. Lian, M.-L. Cai, Y. Wang, B.-X. Qi, L. Yao, L. He, et al., PRX Quantum 4, 010302 (2023).
- Zamora et al. (2020) A. Zamora, G. Dagvadorj, P. Comaron, I. Carusotto, N. Proukakis, and M. Szymańska, Phys. Rev. Lett. 125, 095301 (2020).
- Franz (1958) W. Franz, Z. Naturforsch. A 13, 484 (1958).
- Keldysh (1958) L. Keldysh, J. Exp. Theor. Phys. 6, 763 (1958).
- Tharmalingam (1963) K. Tharmalingam, Phys. Rev. 130, 2204 (1963).
- Callaway (1963) J. Callaway, Phys. Rev. 130, 549 (1963).
- Schützhold et al. (2008) R. Schützhold, H. Gies, and G. Dunne, Phys. Rev. Lett. 101, 130404 (2008).
- Heisenberg and Euler (1936) W. Heisenberg and H. Euler, Z. Phys. 98, 714 (1936).
- Weisskopf (1936) V. Weisskopf, Mat. Fys. Medd. 14, 6 (1936).
- Schwinger (1951) J. Schwinger, Phys. Rev 82, 664 (1951).
- Furry (1951) W. Furry, Phys. Rev. 81, 115 (1951).
- Taya (2019) H. Taya, Phys. Rev. D 99, 056006 (2019).
- Huang and Taya (2019) X.-G. Huang and H. Taya, Phys. Rev. D 100, 016013 (2019).
- Landau (1932) L. Landau, Z. Sowjetunion 2, 46 (1932).
- Zener (1932) C. Zener, Proc. R. Soc. A 137, 696 (1932).
- Stückelberg (1932) E. C. G. Stückelberg, Helv. Phys. Acta 5, 369 (1932).
- Majorana (1932) E. Majorana, Nuovo Cimento 9, 43 (1932).
- Mullen et al. (1989) K. Mullen, E. Ben-Jacob, Y. Gefen, and Z. Schuss, Phys. Rev. Lett. 62, 2543 (1989).
- Kayanuma and Mizumoto (2000) Y. Kayanuma and Y. Mizumoto, Phys. Rev. A 62, 061401 (2000).
- Malla and Raikh (2018) R. K. Malla and M. Raikh, Phys. Rev. B 97, 035428 (2018).
- Wubs et al. (2005) M. Wubs, K. Saito, S. Kohler, Y. Kayanuma, and P. Hänggi, New J. Phys. 7, 218 (2005).
- Wubs et al. (2006) M. Wubs, K. Saito, S. Kohler, P. Hänggi, and Y. Kayanuma, Phys. Rev. Lett. 97, 200404 (2006).
- Saito et al. (2006) K. Saito, M. Wubs, S. Kohler, P. Hänggi, and Y. Kayanuma, Europhys. Lett. 76, 22 (2006).
- Saito et al. (2007b) K. Saito, M. Wubs, S. Kohler, Y. Kayanuma, and P. Hänggi, Phys. Rev. B 75, 214308 (2007b).
- Zueco et al. (2008) D. Zueco, P. Hänggi, and S. Kohler, New J. Phys. 10, 115012 (2008).
- Ashhab (2014) S. Ashhab, Phys. Rev. A 90, 062120 (2014).
- Ashhab (2016) S. Ashhab, Phys. Rev. A 94, 042109 (2016).
- Sinitsyn and Li (2016) N. A. Sinitsyn and F. Li, Phys. Rev. A 93, 063859 (2016).
- Sun and Sinitsyn (2016) C. Sun and N. A. Sinitsyn, Phys. Rev. A 94, 033808 (2016).
- Dziarmaga (2006) J. Dziarmaga, Phys. Rev. B 74, 064416 (2006).
- Caneva et al. (2007) T. Caneva, R. Fazio, and G. E. Santoro, Phys. Rev. B 76, 144427 (2007).
- Suzuki (2011) S. Suzuki, J. Phys. Conf. Ser. 302, 012046 (2011).
- Del Campo (2018) A. Del Campo, Physical Review Letters 121, 200601 (2018).
- Roychowdhury et al. (2021) K. Roychowdhury, R. Moessner, and A. Das, Physical Review B 104, 014406 (2021).
- Mayo et al. (2021) J. J. Mayo, Z. Fan, G.-W. Chern, and A. Del Campo, Physical Review Research 3, 033150 (2021).
- Dziarmaga and Rams (2022) J. Dziarmaga and M. M. Rams, Physical Review B 106, 014309 (2022).
- Gradshteyn and Ryzhik (2014) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Academic press, 2014).