This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum reservoir probing of quantum phase transitions

Kaito Kobayashi    Yukitoshi Motome Department of Applied Physics, the University of Tokyo, Tokyo 113-8656, Japan
Abstract

Quantum phase transitions are highly remarkable phenomena manifesting in quantum many-body systems. However, their precise identifications in equilibrium systems pose significant theoretical and experimental challenges. Thus far, dynamical detection protocols employing global quantum quenches have been proposed, wherein transitions are discerned from global nonequilibrium excitations. In this work, we demonstrate that quantum phase transitions can be detected through localized out-of-equilibrium excitations induced by local quantum quenches. While the resulting dynamics after the quench are influenced by both the local quench operation and the intrinsic dynamics of the quantum system, the effects of the former are exclusively extracted through the cutting-edge framework called quantum reservoir probing (QRP). Through the QRP, we find that the impacts of the local quenches vary across different quantum phases and are significantly suppressed by quantum fluctuations amplified near quantum critical points, thereby precisely delineating phase boundaries. We demonstrate that the QRP can detect quantum phase transitions in the paradigmatic integrable and nonintegrable quantum systems, and even topological quantum phase transitions, all within the identical framework employing single-site observables.

INTRODUCTION

Quantum phase transitions emerge as quintessential manifestations of the interplay among quantum many-body phenomena [1, 2]. They are characterized by the universal description across microscopically different systems, which has been central to our understanding of quantum matters. Nevertheless, precisely detecting these transitions in equilibrium systems poses significant challenges both theoretically and experimentally. This has led recent investigations to focus on nonequilibrium dynamics, which is inherently related to quantum phase transitions through dynamical quantum critical phenomena. Indeed, nonequilibrium signatures of quantum phase transitions have been identified in a diverse range of systems, spanning integrable, nonintegrable, and topological systems [3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17]. A prevalent strategy in these explorations involves a global quantum quench, wherein the entire system is initially configured in a specific state and subsequently undergoes time evolution under a Hamiltonian with suddenly changed parameters. Recent experimental advancements in isolating and manipulating quantum systems have facilitated such sophisticated control in various platforms, including trapped ions [18, 19, 20, 21], ultracold atoms [22, 23, 24, 25], Rydberg atoms [26, 27, 28, 29], and nitrogen-vacancy centers [30].

A pertinent question then emerges regarding the feasibility of dynamically detecting equilibrium quantum phase transitions through local quantum quenches. Given that the local quench operation, in contrast to its global counterpart, selectively influences nontrivial excitations or quasiparticles in specific regions characteristic of the system [31, 32, 33, 34], the resulting dynamics may potentially unveil signatures of quantum phase transitions. However, the complexity lies in isolating and analyzing the response dynamics directly ascribed to the local quench operation. Since the dynamics is influenced by both the local quantum quench and the intrinsic dynamics of the quantum system itself, the latter obscures the former, thereby diminishing the potential benefit of local quenches in facilitating access to localized physics.

Refer to caption
Figure 1: Conceptualization of the QRP. a Schematic of the QRC, which harnesses the quantum system as a computational resource. The quantum state of the reservoir reflects the provided input information, and the final output is obtained by linearly transforming a read-out vector constructed through multiple measurements on various degrees of freedom. b Schematic representation of the QRP, which investigates the quantum system from the perspective of information processing capabilities. The input is provided via the local quench, the output is calculated from a local degree of freedom, and the estimation performance of the input signifies the impact of the local quench on that degree of freedom. c Preparation and local manipulation to construct the initial state |Ψkin|\Psi^{\mathrm{in}}_{k}\rangle. The gray spin undergoes the local quench induced by the yellow needle, while the remaining spins are initialized in the designated state |Φ|\Phi\rangle, illustrated here as the all-up state. d Diagrammatic representation of the initial quantum state [Eq. (1)]. The central spin is locally manipulated contingent upon the input value sks_{k} for each case (three cases are depicted). This local quench induces varied dynamics in the local operator Oi(t)O_{i}(t), which are reflective of the corresponding input values. e Dynamics of Oi(t)\langle O_{i}(t)\rangle summarized across three cases depicted in (d). f Schematic illustration of the determination coefficient Ri2(t)R_{i}^{2}(t) derived from Oi(t)\langle O_{i}(t)\rangle. An almost unity Ri2(t)R^{2}_{i}(t) signifies a precise estimation of the input value, indicating the pronounced influence of the local quench on the observed Oi(t)\langle O_{i}(t)\rangle [Eq. (4)]. Conversely, Ri2(t)R^{2}_{i}(t) approaches zero when Oi(t)\langle O_{i}(t)\rangle exhibits similar values that do not reflect each input, as exemplified by intersections of Oi(t)\langle O_{i}(t)\rangle.

In this paper, we reveal that signatures of quantum phase transitions manifest themselves in out-of-equilibrium local observables subsequent to local quantum quenches. To this end, we develop the methodology of “quantum reservoir probing” (QRP) [35], which selectively extracts the effects of the local quench from the dynamics influenced by multiple factors. Through the lens of the QRP, we observe that the local quantum quench exerts varied influences on a single-site operator depending on the inherent characteristics of quantum phases. Moreover, at the quantum critical points, these influences are notably suppressed due to the enhanced quantum fluctuations. These observations highlight the capability of our QRP to effectively detect quantum phase transitions, as exemplified in both integrable and nonintegrable models herein. Furthermore, notwithstanding the use of a single-site operator, we demonstrate that the QRP can identify topological quantum phase transitions, which are typically distinguished by the absence of local order parameters beyond the Landau-Gintzburg-Wilson paradigm. The broad applicability and design flexibility of our QRP establish it as a versatile tool for probing quantum phenomena, opening the door to further understanding of quantum many-body physics.

RESULTS

Framework of the QRP

Before going to the QRP,we provide an overview of a brain-inspired machine learning framework called reservoir computing [36, 37, 38]. In this paradigm, information is nonlinearly processed by networks of artificial neurons, termed the “reservoir”. A distinguishing feature is that all weight parameters within the reservoir part remain fixed, while only the output weight, which transforms the reservoir state into the final output, is trained utilizing a simple linear regression scheme. Consequently, optimization costs are markedly reduced compared to alternative schemes where the entire network weights are optimized. Importantly, due to this fixed nature, a physical system can embody the reservoir part; the quantum reservoir computing (QRC) harnesses a quantum system as the reservoir (Fig. 1a) [39]. The quantum reservoir system evolves under the influence of the input data, and a read-out vector, constructed through repeated measurements on various degrees of freedom, undergoes linear transformation to yield the final output.

The QRP is an inverse extension of the QRC, wherein the quantum system employed as the reservoir part is investigated by linking physical phenomena with information processing (Fig. 1b). As introduced below, the input is provided via the local quench, and the output is calculated from a local observable; the calculation objective is to estimate the original input value. Given that information is provided solely through the local quench, the estimation succeeds when the local quench exerts an influence on the observed degree of freedom, whereas it fails when the operator is entirely unaffected by the local quench. By systematically scanning local observation operators, the QRP selectively probes the impacts of the local quench on the quantum system through the input estimation performance.

To elaborate, the central feature underlying the QRP is the manipulation of a local degree of freedom, parametrized by a random input value sks_{k}. Our focus is on a local operator OO defined per site ii and time tt, denoted as Oi(t)O_{i}(t), with particular attention to the variations induced by the input variable sks_{k}. In the statistical analysis across various instances with different inputs subscripted by kk, the expectation value Oi(t)k\langle O_{i}(t)\rangle_{k} may display diverse characteristics that closely reflect the corresponding input values. In such cases, the value of sks_{k} can be precisely deduced from Oi(t)k\langle O_{i}(t)\rangle_{k} through straightforward calculations. Hence, the successful estimation of the original input value sks_{k} provides evidence that the local manipulation exerts an influence on the operator Oi(t)O_{i}(t) at that particular spatiotemporal coordinate. In contrast, the operator Oj(t)O_{j}(t^{\prime}) at a different spatiotemporal point may demonstrate consistent behavior across all instances, not accurately mirroring the local quench operation. In this latter case, the operator is predominantly governed by the inherent characteristics of the quantum system independent of the local quench. Therefore, the dependency of the operator Oi(t)O_{i}(t) on the input value sks_{k} explores how the local quantum quench influences the operator, while isolating it from other intrinsic phenomena. Remarkably, this linkage is highly sensitive to the Hamiltonian governing the quantum system. The operator’s dependency on the local quench operation thus serves as a witness to the properties of the quantum system, shedding light on the underlying quantum processes at play.

Here, we explain the specific procedures of the QRP on the one-dimensional (1D) quantum systems. The protocol consists of four steps: (i) the preparation of the initial state (Fig. 1c), (ii) the local quench on the central spin, parameterized by the input value sks_{k} (Figs. 1c and 1d), (iii) the observation of the dynamics of a local operator Oi(t)\langle O_{i}(t)\rangle under the Hamiltonian \mathcal{H} (Figs. 1d and 1e), and (iv) the statistical evaluation of the performance in estimating the input value sks_{k} from the observed dynamics, quantified via the determination coefficient Ri2(t)R_{i}^{2}(t) (Fig. 1f). We provide a comprehensive description of each step in the following.

First, the system is prepared in a specified initial state in step (i); for simplicity, a product state is assumed here (e.g., the all-up state in Fig. 1c). Subsequently, in step (ii), the central spin undergoes local manipulation based on the input value sks_{k}. The resultant state is expressed as

|Ψkin=|ψ(sk)|Φ,|\Psi_{k}^{\mathrm{in}}\rangle=|\psi(s_{k})\rangle\otimes|\Phi\rangle, (1)

where |ψ(sk)|\psi(s_{k})\rangle represents the state of the locally manipulated spin, while |Φ|\Phi\rangle denotes the prepared initial state for the remaining subsystem (Fig. 1d). The input value sks_{k} is randomly sampled from a uniform distribution, specifically sk[0,1]s_{k}\in[0,1]. For the spin-1/21/2 case, we consider the initial state as

|ψ(sk)=1sk|++sk|,|\psi(s_{k})\rangle=\sqrt{1-s_{k}}|+\rangle+\sqrt{s_{k}}|-\rangle, (2)

where |±=(±)/2|\pm\rangle=\left(\mid\uparrow\rangle\pm\mid\downarrow\rangle\right)/\sqrt{2}. This state can be prepared by applying a single-gate pulse or a local magnetic field 𝒉(2sk1,0,2sk(1sk))\bm{h}\propto(2s_{k}-1,0,-2\sqrt{s_{k}(1-s_{k})}).

We then let the system to evolve under the Hamiltonian \mathcal{H} in step (iii), resulting in the state |Ψk(t)=eit|Ψkin|\Psi_{k}(t)\rangle=e^{-i\mathcal{H}t}|\Psi_{k}^{\mathrm{in}}\rangle (see Methods section). The expectation value of a local operator Oi(t)O_{i}(t) is calculated as Oi(t)k=Ψk(t)|Oi|Ψk(t)\langle O_{i}(t)\rangle_{k}=\langle\Psi_{k}(t)|O_{i}|\Psi_{k}(t)\rangle; we omit the subscript kk unless necessary. By iterating this procedure across various instances with different sks_{k}, we generate multiple time series of the dynamics, as illustrated in Fig. 1e.

Finally in step (iv), we assess the influence of the local quantum quench on the local operator. As a criterion, we employ the precision in estimating the input value sks_{k} through the linear transformation of Oi(t)k\langle O_{i}(t)\rangle_{k}, following the standard methodology used in reservoir computing paradigm [36, 37, 38, 39, 35]. For a given instance kk, the output of this linear transformation is derived as

yi,k(t)=wi,O(t)Oi(t)k+wi,c(t),y_{i,k}(t)=w_{i,O}(t)\langle O_{i}(t)\rangle_{k}+w_{i,c}(t), (3)

where wi,O(t)w_{i,O}(t) and wi,c(t)w_{i,c}(t) represent the kk-independent coefficients. These weights are finely tuned for training instances so that the output yi,k(t)y_{{i,k}}(t) approximates the original input value sks_{k} as closely as possible over all kk, based on the least squares method. Subsequently in the testing phase, the estimation performance for unknown instances is quantified by the determination coefficient

Ri2(t)=cov2({yi,k(t)},{sk})σ2({yi,k(t)})σ2({sk}),R_{i}^{2}(t)=\frac{\mathrm{cov}^{2}(\{y_{{i,k}}(t)\},\{s_{k}\})}{\sigma^{2}(\{y_{{i,k}}(t)\})\sigma^{2}(\{s_{k}\})}, (4)

where cov\mathrm{cov} and σ2\sigma^{2} represent covariance and variance, respectively; see Methods section for more details. The value of Ri2(t)R^{2}_{i}(t) ranges between 0 and 11, with an approach towards 11 indicating that the outputs {yi,k(t)}\{y_{{i,k}}(t)\} closely reproduce the inputs {sk}\{s_{k}\} for all testing cases.

Figures 1e and 1f visually elucidate the relationship between the dynamics of Oi(t)k\langle O_{i}(t)\rangle_{k} and the corresponding Ri2(t)R^{2}_{i}(t). A high Ri2(t)R^{2}_{i}(t) value suggests that Oi(t)k\langle O_{i}(t)\rangle_{k} varies systematically in response to the input value sks_{k}. In contrast, Ri2(t)R^{2}_{i}(t) close to 0 indicates that Oi(t)k\langle O_{i}(t)\rangle_{k} is largely independent of sks_{k}, such as at the crossing points resulting from intrinsic oscillations in the quantum system. Therefore, Ri2(t)R^{2}_{i}(t) serves as an indicator of the extent to which Oi(t)O_{i}(t) is influenced by the local quench operation. Under different parameters of the Hamiltonian \mathcal{H}, the local quench operation exerts varying influences on the resultant dynamics, hence the behavior of Ri2(t)R^{2}_{i}(t) is instrumental in examining the unique properties of each quantum phase. It is worth emphasizing that restricting the transformation generating yi,k(t)y_{{i,k}}(t) from Oi(t)k\langle O_{i}(t)\rangle_{k} to a linear form is crucial. Any additional nonlinearity could lead to either an overestimation or underestimation of the effects of the local quench, while simultaneously impeding interpretability as the output is influenced by the extraneous characteristics of the transformation itself.

Refer to caption
Figure 2: Identifying the quantum phase transition in the transverse-field Ising model via the QRP. a Spatiotemporal representation of the performance for the estimation of the input value. The color represents the value of Ri2(t)R^{2}_{i}(t) derived from σix(t)\langle\sigma^{x}_{i}(t)\rangle. The upper band represents the schematic phase diagram regarding (g)\mathcal{H}(g), where the quantum critical point separating the ferromagnetically ordered phase and the quantum disordered phase exists at gc=1{g_{c}}=1. The system size is N=49N=49, and the time evolution is calculated using a bond dimension χ=256\chi=256; the result is quantitatively the same for χ=128\chi=128. b The early-time spin dynamics at the nearest-neighbor site from the central one under three different input values. c The estimation performance Ri2(t)R^{2}_{i}(t) utilizing σix(t)\langle\sigma^{x}_{i}(t)\rangle at the same site as (b). d The average R2¯\overline{R^{2}} over spatiotemporal indices (i,t)Λ×{0<t10}(i,t)\in\Lambda\times\{0<t\leq 10\}, with Λ\Lambda encompassing the central 1111 sites. The blue and pink lines represent the system sizes N=49N=49 and N=37N=37, respectively. The quantum critical point is denoted by the gray dashed line.

Transverse-field Ising model

Let us first apply our QRP protocol to the paradigmatic 1D Ising model with a transverse magnetic field under the open boundary condition. The Hamiltonian is given by

=Jiσizσi+1z+giσix,\mathcal{H}=-J\sum_{i}\sigma^{z}_{i}\sigma^{z}_{i+1}+g\sum_{i}\sigma^{x}_{i}, (5)

where σix{\sigma}_{i}^{x} and σiz{\sigma}_{i}^{z} are the xx and zz Pauli matrices at site ii, respectively. gg signifies the magnitude of the transverse magnetic field and J>0J>0 denotes the strength of the nearest-neighbor ferromagnetic interaction; we set J=1J=1 as our energy scale. This model manifests a quantum phase transition at the quantum critical point gc=1g_{c}=1 in the thermodynamic limit, which separates a ferromagnetically ordered phase for g<1g<1 from a quantum disordered phase for g>1g>1 [40, 1]. Notably, the model is solvable through the Jordan-Wigner transformation and the subsequent Bogoliubov transformation, where the spin system is mapped to the free quasiparticle picture [1].

The system is initialized in the all-up state. Following the local quantum quench, the resultant state |Ψkin|\Psi^{\mathrm{in}}_{k}\rangle in Eq. (1) is explicitly expressed as:

|Ψkin=|ψ(sk).|\Psi^{\mathrm{in}}_{k}\rangle=\mid\uparrow\uparrow\uparrow\dots\rangle\otimes|\psi(s_{k})\rangle\otimes\mid\uparrow\uparrow\uparrow\dots\rangle. (6)

Subsequently, we monitor the dynamics of the local spin σix(t)\langle\sigma^{x}_{i}(t)\rangle at each site ii under the Hamiltonian =(g)\mathcal{H}=\mathcal{H}(g). Repeating the procedure for many instances with varying sks_{k}, the estimation performance Ri2(t)R^{2}_{i}(t) in Eq. (4) is methodically evaluated.

Figure 2 encapsulates the results of the QRP under various magnetic fields gg. In Fig. 2a, we show the spatiotemporal representation of the estimation performance Ri2(t)R^{2}_{i}(t), which reveals a discernible light-cone structure emanating from the quenched central site. The broad distribution of nonzero Ri2(t)R^{2}_{i}(t) suggests that the local spin operator σix(t)\langle\sigma_{i}^{x}(t)\rangle across diverse spatial locations and temporal moments adequately reflect the impact of the local quantum quench. In the quantum disordered phase, where the ground state exhibits spins aligned along the xx direction, the effects of the local quench along the xx axis in spin space propagate akin to a spin wave, characterized by pronounced wavefronts of Ri2(t)R^{2}_{i}(t). In contrast, in the ferromagnetically ordered phase, the local quenching effects are preserved predominantly within the quasiparticles, manifesting as alterations in phase and amplitude. This influence is evidenced by the ripple-like patterns of the nonzero Ri2(t)R^{2}_{i}(t). An exception occurs near the quantum critical point at gc=1g_{c}=1, characterized by maximized fluctuations that lead to an effective breakdown of the quasiparticle picture. Here, the effects of the local quench are passed on the intrinsic fluctuations of the system, though such influences rapidly diminish as indicated by the noticeably small Ri2(t)R^{2}_{i}(t) after the passage of the wavefront. This observation highlights an anomalous suppression of the propagation of local quenching effects at the quantum critical point, which is distinct from other non-critical phases.

In Fig. 2b, the early-time dynamics of σix(t)\langle\sigma^{x}_{i}(t)\rangle at the nearest-neighbor spin from the central one is depicted for three distinct input values sks_{k} at different magnetic fields gg. At g=0.5g=0.5 and 1.51.5, the quasiparticles induce periodic spin oscillations, the phase and amplitude of which are dependent on the input sks_{k}. In contrast, at g=1.0g=1.0, the dynamics of σix(t)\langle\sigma^{x}_{i}(t)\rangle exhibits significant variation when the wavefronts pass at approximately t1t\simeq 1, thereafter stabilizing into a uniform pattern that does not reflect the influence of quench operation. Figure 2c displays the corresponding estimation performance at the same site (a vertical cross-section of the color map in Fig. 2a). Over broad temporal spans at g=0.5g=0.5 and 1.51.5, Ri2(t)R^{2}_{i}(t) becomes close to 11, which statistically underscores the input dependency of the phase and amplitude of the spin oscillations. Notably, their frequency is governed not by the local quench on the initial state, but rather by the intrinsic characteristics of the quasiparticles themselves. This is evidenced by Ri2(t)R^{2}_{i}(t) periodically approaching 0 (Fig. 2c) at points where oscillations of σix(t)\langle\sigma_{i}^{x}(t)\rangle intersect for different inputs (Fig. 2b). In contrast, at g=1.0g=1.0, while Ri2(t)R^{2}_{i}(t) nears 11 around t1t\simeq 1 where the wavefronts induce variations in the spin dynamics, the similarity of σix(t)\langle\sigma_{i}^{x}(t)\rangle in other temporal regions leads to Ri2(t)0R^{2}_{i}(t)\simeq 0, signifying the reduced influence of the local quench by the intrinsic fluctuations enhanced near the quantum critical point.

As described above, the effects of the local quantum quench on the resultant quantum dynamics of local operators manifest qualitatively different characteristics across various phases (Fig. 2a). The distribution of Ri2(t)R^{2}_{i}(t) thus serves as a marker for identifying the quantum phase transition. To enhance the clarity of the results, in Fig. 2d, we display the mean value of Ri2(t)R^{2}_{i}(t) over a specific subset of spatiotemporal indices {(i,t)}\{(i,t)\}, denoted as R2¯\overline{R^{2}}. This analysis focus on (i,t)(i,t) within the central eleven sites during the time interval 0<t100<t\leq 10 to minimize the effects from the edges of the system; selecting different subsets does not qualitatively alter the results (see the Supplemental Information). We note that R2¯\overline{R^{2}} represents a simple average of local quantities, and therefore, does not include nonlocal quantum effects. As gg approaches gc=1g_{c}=1 from the ferromagnetically ordered state, the fluctuations progressively intensify and interfere with the estimation of the original input, resulting in a decrease in R2¯\overline{R^{2}}. At the quantum critical point, R2¯\overline{R^{2}} reaches its minimum, which corresponds to the emergence of a dark region between the wavefronts observed in Fig. 2a for g=1.0g=1.0. Upon further augmentation of gg into the quantum disordered phase, the fluctuations weaken again and the effects of the quench operation become dominant, leading to an increase in R2¯\overline{R^{2}}. The pronounced dip in Fig. 2d is therefore a definitive indicator of the quantum phase transition occurring at gcg_{c}. Consequently, by focusing on the effects of the local quench, our QRP successfully detects the quantum phase transition only by measurements of single-site spin dynamics.


Anisotropic next-nearest-neighbor Ising model

Refer to caption
Figure 3: Detection of the quantum phase transition in the ANNNI model using the QRP. a Spatiotemporal representation of the estimation performance when employing σix(t)\langle\sigma^{x}_{i}(t)\rangle in the nonintegrable ANNNI chain characterized by the interaction parameter κ=0.5\kappa=0.5. The system size is N=49N=49, and the time evolution is calculated with a bond dimension of χ=128\chi=128. b The average R2¯\overline{R^{2}} over a subset (i,t)Λ×{0<t15}(i,t)\in\Lambda\times\{0<t\leq 15\}, where Λ\Lambda contains the central eleven sites. The upper band represents the schematic phase diagram, where the ferromagnetically ordered and quantum disordered phases are separated by the quantum critical point at approximately g1.6g\simeq 1.6.

To illustrate the applicability of our QRP framework to nonintegrable models, we extend the model to the anisotropic next-nearest-neighbor Ising (ANNNI) chain [41, 42, 43, 44]. The Hamiltonian is given by

=Jiσizσi+1zκiσizσi+2z+giσix,\mathcal{H}=-J\sum_{i}\sigma^{z}_{i}\sigma^{z}_{i+1}-\kappa\sum_{i}\sigma^{z}_{i}\sigma^{z}_{i+2}+g\sum_{i}\sigma^{x}_{i}, (7)

where we take J=1J=1 and κ\kappa represents the strength of the next-nerarest-neighbor interaction. The latter term makes the system nonintegrable by introducing the four-body interactions for the quasiparticles derived by the Jordan-Wigner transformation. This model hosts a rich phase diagram in the κ\kappa-gg plane [45]. Our investigation focuses on the case of κ=0.5\kappa=0.5, where the quantum phase transition occurs at gc1.6g_{c}\simeq 1.6 with the Ising universality, separating ferromagnetically ordered (g<gcg<g_{c}) and quantum disordered (g>gcg>g_{c}) phases. The initial state is prepared in the same manner as Eq. (6).

Figure 3a displays the spatiotemporal representation of Ri2(t)R^{2}_{i}(t) obtained from σix(t)\langle\sigma^{x}_{i}(t)\rangle, revealing a scenario similar to the transverse-field Ising model. For the majority of sites and times, the local spin σix(t)\langle\sigma_{i}^{x}(t)\rangle mirrors the input value sks_{k} employed in the local quantum quench, as evidenced by the nonzero values of Ri2(t)R^{2}_{i}(t). Above the critical field at gc1.6g_{c}\simeq 1.6, particularly for g=2.2g=2.2 in the right panel, the effects of the local manipulation propagate primarily through the excitation wavefronts, while the local spin oscillations exhibit weaker dependence on the local quench with small Ri2(t)R^{2}_{i}(t). Conversely, below the critical point the opposite behavior is noted in the left panel for g=1.0g=1.0. Furthermore, in the vicinity of the quantum critical point, the influence of the local quench manifest itself predominantly along the wavefronts, while in the other regions it is suppressed by the enhanced fluctuations, as demonstrated by the emergence of a discernible dark zone between the wavefronts. These fundamental dependencies of the local spin operators on the local quench retain a similarity to those observed in the integrable model presented in Fig. 2a.

In Fig. 3b, we present the mean value R2¯\overline{R^{2}}, calculated over the indices (i,t)(i,t) corresponding to the central eleven sites within the time frame 0<t150<t\leq 15 (see the Supplemental Information for other subsets). Analogous to the observation in Fig. 2d, R2¯\overline{R^{2}} demonstrates a pronounced decrease around the quantum critical point at g1.6g\simeq 1.6, precisely locating the boundary between the different spatiotemporal distribution patterns of Ri2(t)R^{2}_{i}(t) in the ferromagnetically ordered phase and the quantum disordered phase. This observation confirms that, even in nonintegrable systems, Ri2(t)R^{2}_{i}(t) is an effective marker for detecting quantum phase transitions.


Spin-1 XXZ model

The preceding sections have demonstrated the effectiveness of our QRP in identifying quantum phase transitions, particularly between conventional phases where the internal physics is well-characterized and understood within the Landau-Ginzburg-Wilson theory. Building on this foundation, we now aim to broaden our exploration to more complex quantum phenomena. In particular, we apply the QRP framework to the detection of topological quantum phase transitions, which lack local order parameters that can distinguish adjacent phases. This raises the question of whether the QRP framework, which solely utilizes a single-site spin operator, can still be effective in identifying these transitions beyond the Landau-Gintzburg-Wilson paradigm.

Here, we study the spin-11 XXZ model with open boundaries [46, 47], which is defined by

=iJ(SixSi+1x+SiySi+1y)+λSizSi+1z,\mathcal{H}=\sum_{i}J(S^{x}_{i}S^{x}_{i+1}+S^{y}_{i}S^{y}_{i+1})+\lambda S^{z}_{i}S^{z}_{i+1}, (8)

where SiαS^{\alpha}_{i} denotes the α\alpha-component of spin-11 operator at site ii and λ\lambda represents the Ising anisotropy; we set the interaction strength J=1J=1. This model is renowned for hosting a prototypical symmetry-protected topological phase, known as the Haldane phase, observed for the anisotropy within 0<λ<λc0<\lambda<\lambda_{c} with λc1.18\lambda_{c}\simeq 1.18 [48, 49, 50]. The Haldane phase is distinguished by its protection under a Z2×Z2Z_{2}\times Z_{2} symmetry, and is characterized by nonlocal string order parameters [51, 52, 53]. With increasing the anisotropy, the topological quantum phase transition into the topologically trivial spin-11 Néel state occurs at the quantum critical point, which belongs to the two-dimensional (2D) Ising universality class [49, 54, 55].

The QRP framework can be naturally extended to the spin-11 XXZ model. Here, |Φ|\Phi\rangle is initialized as the Néel state, and the input spin state |ψ(sk)|\psi^{\prime}(s_{k})\rangle, the spin-11 extension of Eq. (2), is prepared by applying a local magnetic field 𝒉(sk1,0,sk(2sk))\bm{h}^{\prime}\propto(s_{k}-1,0,-\sqrt{s_{k}(2-s_{k})}). Explicitly, the resultant state |Ψkin|\Psi^{\mathrm{in}}_{k}\rangle is expressed as:

|Ψkin=|ψ(sk).|\Psi^{\mathrm{in}}_{k}\rangle=\mid\uparrow\downarrow\uparrow\downarrow\dots\uparrow\downarrow\rangle\otimes|\psi^{\prime}(s_{k})\rangle\otimes\mid\downarrow\uparrow\downarrow\uparrow\dots\downarrow\uparrow\rangle. (9)

Following this setup, the impact of the local quench on Six(t)\langle S^{x}_{i}(t)\rangle is evaluated through the estimation performance Ri2(t)R^{2}_{i}(t), analogous to the previous cases.

Refer to caption
Figure 4: Probing the topological quantum phase transition in the spin-1 XXZ model using the QRP. a Spatiotemporal representation of the estimation performance Ri2(t)R_{i}^{2}(t) employing Six(t)\langle S^{x}_{i}(t)\rangle for a system with N=49N=49 sites. The time evolution is calculated with a bond dimension of χ=128\chi=128. b The average R2¯\overline{R^{2}} over a subset (i,t)Λ×{10<t20}(i,t)\in\Lambda\times\{10<t\leq 20\}, where Λ\Lambda represents the entire system. The upper bar illustrates the phase diagram of (λ)\mathcal{H}(\lambda) with schematics composed of virtual S=1/2S=1/2 spins.

In Fig. 4a, we present the estimation performance Ri2(t)R^{2}_{i}(t) under the Hamiltonian \mathcal{H} for three different anisotropies: λ=1.0\lambda=1.0 in the Haldane phase, λ=1.3\lambda=1.3 in the Néel phase, and λ=1.18\lambda=1.18 at the vicinity of the quantum critical point. The spread of nonzero Ri2(t)R^{2}_{i}(t) values indicates that the operator Six(t)S_{i}^{x}(t) is reflective of the local quantum quench at various spatiotemporal points, similar to what is observed in S=1/2S=1/2 spin systems. Following an initial transient period, Ri2(t)R^{2}_{i}(t) shows a disturbed ripple-like pattern in the Néel phase, reminiscent of those in the ordered phases in Figs. 2a and 2b. In contrast, in the Haldane phase, Ri2(t)R^{2}_{i}(t) becomes more uniform, presumably due to the nonlocal sharing of quenching effects via the underlying singlet pairs. Around the critical point, a notable trend is observed wherein Ri2(t)R^{2}_{i}(t) progressively declines over time. This precisely mirrors the trend observed at the quantum critical point in conventional quantum phase transitions (Figs. 2a and 3a).

Figure 4b presents the mean value R2¯\overline{R^{2}}, which effectively distinguishes these three distinct distributions of Ri2(t)R^{2}_{i}(t). R2¯\overline{R^{2}} is computed as the average of Ri2(t)R^{2}_{i}(t) for the entire system within the time interval 10<t2010<t\leq 20. A key observation in Fig. 4b is the pronounced dip at approximately λλc\lambda\simeq\lambda_{c}, signifying the suppression of the impact of local quench. This dip thus marks the quantum critical point characterized by enhanced fluctuations, clearly delineating the boundary between the topological Haldane phase and the trivial Néel phase. Consequently, despite the absence of local order parameters in the Haldane phase, our QRP, based on the local quench and local observables, demonstrates itself as a potent instrument for detecting the topological quantum phase transition.


Extended Su-Schrieffer-Heeger model

Finally, we analyze the extended Su-Schrieffer-Heeger (eSSH) model [56, 57, 58, 59, 60, 61]:

=\displaystyle\mathcal{H}= Ji=1N/2(σ2i1xσ2ix+σ2i1yσ2iy+δσ2i1zσ2iz)\displaystyle J\sum_{i=1}^{N/2}(\sigma^{x}_{2i-1}\sigma^{x}_{2i}+\sigma^{y}_{2i-1}\sigma^{y}_{2i}+\delta\sigma^{z}_{2i-1}\sigma^{z}_{2i})
+\displaystyle+ Ji=1N/21(σ2ixσ2i+1x+σ2iyσ2i+1y+δσ2izσ2i+1z).\displaystyle J^{\prime}\sum_{i=1}^{N/2-1}(\sigma^{x}_{2i}\sigma^{x}_{2i+1}+\sigma^{y}_{2i}\sigma^{y}_{2i+1}+\delta\sigma^{z}_{2i}\sigma^{z}_{2i+1}). (10)

Here, JJ and JJ^{\prime} represent the strength of alternating nearest-neighbor interactions, and δ\delta denotes the anisotropy in the zz direction; we set J=1J=1. Our analysis focuses on the scenario at δ=0.5\delta=0.5, where a topological phase transition occurs around Jc1J^{\prime}_{c}\simeq 1, typically examined through many-body topological invariants [62, 60, 61]. For J<JcJ^{\prime}<J^{\prime}_{c}, the trivial dimer phase manifests, with spins forming singlet pairs with their nearest neighbors, while for J>JcJ^{\prime}>J^{\prime}_{c}, the topological phase emerges, characterized by the emergence of edge states. Importantly, these phases are entirely topological and devoid of local orderings on both sides of the quantum critical point, unlike the spin-1 XXZ model, which exhibits the local Néel order on one side.

Refer to caption
Figure 5: Detection of the topological quantum phase transition in the eSSH model using the QRP. a Color map of the estimation performance Ri2(t)R_{i}^{2}(t) employing σix(t)\langle\sigma^{x}_{i}(t)\rangle in the eSSH model with δ=0.5\delta=0.5. The system size is N=48N=48, and the time evolution is calculated with a bond dimension of χ=128\chi=128. b The average R2¯\overline{R^{2}} over a subset (i,t)Λ×{8<t20}(i,t)\in\Lambda\times\{8<t\leq 20\}, where Λ\Lambda contains the central 2424 sites. The upper band represents the phase diagram, where the topological and trivial phases are separated by the quantum critical point at approximately J1.0J^{\prime}\simeq 1.0.

The initial state |Ψk|\Psi_{k}\rangle is prepared as the ground state of \mathcal{H} in the topological phase with J=6.0J^{\prime}=6.0, subject to a weak magnetic field 𝒉(2sk1,0,2sk(1sk))\bm{h}\propto(2s_{k}-1,0,-2\sqrt{s_{k}(1-s_{k})}) applied locally to one of the central sites for encoding input information. To ensure that any variations in |Ψk|\Psi_{k}\rangle across different instances arise solely from differences in sks_{k}, a magnetic field is applied at both edge spins along the +z+z direction; this edge field resolves the ground state degeneracy associated with edge degrees of freedom. Both central and edge fields are set to a magnitude of 0.10.1 and deactivated following state preparation. Although the prepared state is not a product state and thus not representable in the form of Eq. (1), the effects of the local field are essentially confined to the central region, thereby presenting a virtually identical scenario to the local quench operation. Commencing from this initial state, we monitor the dynamics of the local spin σix(t)\langle\sigma_{i}^{x}(t)\rangle across multiple cases with different sks_{k}. The resultant discrepancies in the observed dynamics are attributed to the initial localized manipulation parameterized by sks_{k}, analogous to prior models.

Figure 5a illustrates the estimation performance Ri2(t)R^{2}_{i}(t) under the Hamiltonian \mathcal{H} for three different values of JJ^{\prime}. In all scenarios, Ri2(t=0)R^{2}_{i}(t=0) acquires nonzero values only around the central region, confirming that the information regarding sks_{k} is indeed introduced locally through the application of the local magnetic field. During the subsequent time evolution, the effects of the initial localized manipulation begin to propagate throughout the system, where the distinctions between the phases become evident. Notably, in proximity to the quantum critical point at J=1.0J^{\prime}=1.0, Ri2(t)R^{2}_{i}(t) is markedly suppressed, displaying a discernibly dark spatiotemporal region in Fig. 5a. This behavior contrasts with the topological phase at J=1.3J^{\prime}=1.3, which exhibits a ripple-like pattern, and the trivial phase at J=0.7J^{\prime}=0.7, characterized by a strip-shaped pattern. In Fig. 5b, we illustrate the JJ^{\prime} dependence of R2¯\overline{R^{2}}, defined by averaging Ri2(t)R^{2}_{i}(t) over the central 2424 sites within the time frame 8<t208<t\leq 20. Consistent with previous models, R2¯\overline{R^{2}} exhibits a pronounced dip at the quantum critical point, which precisely delineates the boundary between the trivial and topological phases. Indeed, although the transition is entirely topological and the local orderings are absent, the quantum fluctuations are consistently enhanced at the critical point. The QRP locally detect such fluctuations as a witness of the topological quantum phase transition.

DISCUSSION

To summarize, we have proposed the QRP as an versatile tool for identifying the quantum phase transitions in a variety of quantum systems, including integrable, nonintegrable, and topological systems. The QRP selectively evaluates the effects of the local quench in a manner similar to the pump-probe paradigm. Here, the parametrized local quantum quench, acting as the “pump”, injects random disturbance into the dynamics. The estimation performance of this random input parameter, through the linear transformation of the single-site operator, serves as the “probe”. This pump-probe-like process informationally discerns the underlying relationship between the local quench and the resultant dynamics of operators. Using the QRP, we have observed that the local quench operation exerts varied influence depending on the nature of the quantum phases. At the quantum critical points, the introduced disturbance becomes subordinate to the maximally enhanced quantum fluctuations. A noticeable decline in the estimation performance observed therein serves as a marker for the transition between distinct quantum phases. These findings have established that local, nonequilibrium excitations bear the signatures of quantum phase transitions concerning global, equilibrium many-body ground states. The crux of our methodology lies in the suppression of local quenching effects on local dynamics due to amplified quantum fluctuations, the rigorous justification of which is left for future study.

Our investigations have focused on 1D quantum spin systems; however, the QRP methodology is applicable to a broader range of quantum systems in more than one dimension as well. Indeed, the conformal field theories governing quantum phase transitions in two and higher-dimensional systems exhibit significant differences from their 1D counterparts [63]. These differences manifest themselves as changes in nature of excitations, intrinsic dynamics, and critical behaviors near quantum critical points, all of which would be detectable through the QRP. Moreover, geometric frustrations in high-dimensional quantum spin systems, coupled with the inherent quantum fluctuations, give rise to a plethora of topological quantum phase transitions in a variety of systems. For instance, in certain systems, such transitions are associated with the emergence of anyons, which have promising potential for applications in topological quantum computing [64]. The examination of how anyons respond to local quantum operations through the QRP holds significant importance in light of the aforementioned applications. We leave the high-dimensional applications of the QRP for future investigations.

Finally, let us remark on the design flexibility of the QRP. A fundamental advantage of employing local quantum quenches, in contrast to global ones, lies in their capability to access and elucidate quantum phenomena induced by local operations, such as specific classes of excitations. Indeed, the concept of excitation in quantum systems is intrinsically linked to the relative relationship between local manipulations and initial states; thus, identical local quench operations can lead to varying outcomes based on the differences in the initial states. When one possesses prior knowledge regarding the physics of the targeted system, especially for discerning partially understood phenomena, the optimal selections for the initial state, local quantum quench, and read-out operators in the QRP become intuitively evident. In contrast, the true value of the QRP framework becomes markedly pronounced in the exploration of completely unknown quantum systems. Therein, the underlying physics remains undisclosed, rendering the interpretation of the complex dynamics observed a formidable task. Our QRP introduces a systematic methodology to address this complexity, categorizing operators by meticulously investigating their dependence on various local operations through a scan of the aforementioned conditions of the QRP. Furthermore, the experimental simplicity of the QRP is noteworthy, as it necessitates solely single-site measurements during the initial stages of the dynamics. Consequently, we believe that the QRP presents a promising methodology for eliciting novel insights into a broad spectrum of exotic quantum many-body phenomena, extending beyond quantum phase transitions.

METHODS

State preparation and time evolution

To construct the initial state, we define DMRG\mathcal{H}_{\mathrm{DMRG}} that yields |Ψkin|\Psi_{k}^{\mathrm{in}}\rangle as its ground state obtained from the density matrix renormalization group (DMRG) method [65, 66, 67]. For the transverse-field Ising model, the ANNNI model, and the spin-11 XXZ model, the initial states are prepared in product states as described in Eqs. (6) and (9). The Hamiltonian DMRG\mathcal{H}_{\mathrm{DMRG}} is correspondingly defined, consisting of the local up or down magnetic field along the zz-axis, which yields the all-up or Néel states as the ground state. A local magnetic field is applied to the central spin to prepare it as |ψ(sk)|\psi(s_{k})\rangle in the spin-1/21/2 systems and |ψ(sk)|\psi^{\prime}(s_{k})\rangle in the spin-11 system. For the eSSH model, we set DMRG=(J=6.0)\mathcal{H}_{\mathrm{DMRG}}=\mathcal{H}(J^{\prime}=6.0), applying the sks_{k}-dependent central field and an edge field in +z+z-axis with a strength of 0.10.1. The initial state is then obtained as the ground state of DMRG\mathcal{H}_{\mathrm{DMRG}} using the DMRG method, expressed as a matrix product state with a bond dimension χ\chi. For the time-evolution eit|Ψkine^{-i\mathcal{H}t}|\Psi_{k}^{\mathrm{in}}\rangle, we employ the time-dependent variational principle [68, 69, 70] with a bond dimension χ\chi and a time step Δt=0.005\Delta t=0.005. The tensor network calculations are implemented using the Julia version of the ITensor library [71, 72].


Statistical treatment in the QRP

From a uniform distribution [0,1][0,1], a set of 256256 input values {sk}\{s_{k}\} is sampled. Each input sks_{k} produces a distinct initial state |Ψkin|\Psi_{k}^{\mathrm{in}}\rangle, and the dynamics of operator Oi(t)k\langle O_{i}(t)\rangle_{k} is subsequently calculated. Among them, ltr=128l^{\mathrm{tr}}=128 instances are designated for training, while the remaining lts=128l^{\mathrm{ts}}=128 instances are reserved for testing. The linear transformation is represented using a 2D vector 𝒙i,k(t)=(Oi(t)k,1)\bm{x}_{i,k}(t)=\left(\langle O_{i}(t)\rangle_{k},1\right) and expressed as yi,k(t)=𝒙i,k(t)𝒘i(t)y_{i,k}(t)=\bm{x}_{i,k}(t)\bm{w}_{i}(t), with the weight vector 𝒘i(t)=(wi,O(t),wi,c(t))\bm{w}_{i}(t)=(w_{i,O}(t),w_{i,c}(t))^{\top}.

In the training phase, an (ltr×2l^{\mathrm{tr}}\times 2)-dimensional matrix Xitr(t)={𝒙i,k(t)}k=1ltrX_{i}^{\mathrm{tr}}(t)=\{\bm{x}_{i,k}(t)\}_{k=1}^{l^{\mathrm{tr}}} is constructed by summing up the observed dynamics. The output, represented as an ltrl^{\mathrm{tr}}-dimensional vector 𝒚itr(t)={yi,k}k=1ltr\bm{y}_{i}^{\mathrm{tr}}(t)=\{y_{i,k}\}_{k=1}^{l^{\mathrm{tr}}}, is calculated as 𝒚itr(t)=Xitr(t)𝒘i(t)\bm{y}_{i}^{\mathrm{tr}}(t)=X^{\mathrm{tr}}_{i}(t)\bm{w}_{i}(t). The weight vector is trained to closely match the output 𝒚itr(t)\bm{y}_{i}^{\mathrm{tr}}(t) with the original input 𝒔tr={sk}k=1ltr{\bm{{s}}}^{\mathrm{tr}}=\{s_{k}\}_{k=1}^{l^{\mathrm{tr}}}. The optimal solution, which minimizes the least squared error between these vectors, is given by

𝒘i(t)=Xitr(t)+𝒔tr,\bm{w}_{i}(t)={{X}_{i}^{\mathrm{tr}}}(t)^{+}{\bm{{s}}}^{\mathrm{tr}}, (11)

where Xitr(t)+{{X}_{i}^{\mathrm{tr}}}(t)^{+} denotes the Moore-Penrose pseudoinverse-matrix of Xitr(t){{X}_{i}^{\mathrm{tr}}}(t).

In the testing phase, an (lts×2l^{\mathrm{ts}}\times 2)-dimensional matrix Xits(t)={𝒙i,k(t)}k=ltr+1ltr+ltsX_{i}^{\mathrm{ts}}(t)=\{\bm{x}_{i,k}(t)\}_{k=l^{\mathrm{tr}}+1}^{l^{\mathrm{tr}}+l^{\mathrm{ts}}} is similarly composed, yielding the output 𝒚its(t)=Xits(t)𝒘i(t)\bm{y}_{i}^{\mathrm{ts}}(t)=X_{i}^{\mathrm{ts}}(t)\bm{w}_{i}(t) using the trained weights. The estimation performance is assessed by comparing this testing output with the corresponding input 𝒔ts={sk}k=ltr+1ltr+lts{\bm{{s}}}^{\mathrm{ts}}=\{s_{k}\}_{k=l^{\mathrm{tr}}+1}^{l^{\mathrm{tr}}+l^{\mathrm{ts}}} using the determination coefficient

Ri2(t)=cov2(𝒚its(t),𝒔ts)σ2(𝒚its(t))σ2(𝒔ts).R^{2}_{i}(t)=\frac{\mathrm{cov}^{2}(\bm{y}^{\mathrm{ts}}_{i}(t),{\bm{{s}}}^{\mathrm{ts}})}{\sigma^{2}(\bm{y}^{\mathrm{ts}}_{i}(t))\sigma^{2}({\bm{{s}}}^{\mathrm{ts}})}. (12)

Here, cov\mathrm{cov} and σ2\sigma^{2} denote covariance and variance, respectively. The coefficient Ri2(t)R^{2}_{i}(t) statistically assesses the extent to which Oi(t)O_{i}(t) is influenced by the local quantum quench parameterized by sks_{k}. We have verified that a smaller dataset with ltr=100l^{\mathrm{tr}}=100 and lts=100l^{\mathrm{ts}}=100 produces quantitatively similar results.

As explained above, the QRP analyzes the variation of dynamics under different sks_{k} values. Ri2(t)R^{2}_{i}(t) should ideally be zero when no systematic variations dependent on the input sks_{k} are observed. However, in numerical simulations, Ri2(t)R^{2}_{i}(t) might accidentally exhibit nonzero values, sensitively reflecting the small deviations comparable to numerical errors. To avoid this numerical artifact, we introduce a threshold Δth\Delta_{\mathrm{th}} for the deviation of the operator across different instances. We employ the average absolute deviation Δi(t)\Delta_{i}(t), defined as

Δi(t)=1ltrk=1ltr|Oi(t)k1ltrk=1ltrOi(t)k|.\Delta_{i}(t)=\frac{1}{l_{\mathrm{tr}}}\sum_{k=1}^{l_{\mathrm{tr}}}\left|\langle O_{i}(t)\rangle_{k}-\frac{1}{l_{\mathrm{tr}}}\sum_{k=1}^{l_{\mathrm{tr}}}\langle O_{i}(t)\rangle_{k}\right|. (13)

When Δi(t)\Delta_{i}(t) is negligibly small, Ri2(t)R^{2}_{i}(t) calculated from the above Oi(t)\langle O_{i}(t)\rangle should approach zero. We thus manually set Ri2(t)=0R^{2}_{i}(t)=0 when Δi(t)<Δth\Delta_{i}(t)<\Delta_{\mathrm{th}}; we utilize Δth=105\Delta_{\mathrm{th}}=10^{-5}, and Δth\Delta_{\mathrm{th}} dependency is presented in the Supplemental Information. This procedure does not qualitatively alter our methodology and is experimentally feasible.

DATA AVAILABILITY

The data that support the findings of this study are available from the corresponding author upon reasonable request.

CODE AVAILABILITY

The simulation codes used in this study are available from the corresponding author upon reasonable request.

References

  • [1] Sachdev, S. Quantum Phase Transitions (Cambridge University Press, 2011), 2 edn.
  • [2] Zinn-Justin, J. Quantum Field Theory and Critical Phenomena: Fifth Edition (Oxford University Press, 2021).
  • [3] Bhattacharyya, S., Dasgupta, S. & Das, A. Signature of a continuous quantum phase transition in non-equilibrium energy absorption: Footprints of criticality on higher excited states. Sci. Rep. 5, 16490 (2015).
  • [4] Roy, S., Moessner, R. & Das, A. Locating topological phase transitions using nonequilibrium signatures in local bulk observables. Phys. Rev. B 95, 041105 (2017).
  • [5] Heyl, M., Pollmann, F. & Dóra, B. Detecting equilibrium and dynamical quantum phase transitions in ising chains via out-of-time-ordered correlators. Phys. Rev. Lett. 121, 016801 (2018).
  • [6] Dağ, C. B., Sun, K. & Duan, L.-M. Detection of quantum phases via out-of-time-order correlators. Phys. Rev. Lett. 123, 140602 (2019).
  • [7] Wei, B.-B., Sun, G. & Hwang, M.-J. Dynamical scaling laws of out-of-time-ordered correlators. Phys. Rev. B 100, 195107 (2019).
  • [8] Titum, P., Iosue, J. T., Garrison, J. R., Gorshkov, A. V. & Gong, Z.-X. Probing ground-state phase transitions through quench dynamics. Phys. Rev. Lett. 123, 115701 (2019).
  • [9] Dağ, C. B., Duan, L.-M. & Sun, K. Topologically induced prescrambling and dynamical detection of topological phase transitions at infinite temperature. Phys. Rev. B 101, 104415 (2020).
  • [10] Nie, X. et al. Experimental observation of equilibrium and dynamical quantum phase transitions via out-of-time-ordered correlators. Phys. Rev. Lett. 124, 250601 (2020).
  • [11] Haldar, A. et al. Signatures of quantum phase transitions after quenches in quantum chaotic one-dimensional systems. Phys. Rev. X 11, 031062 (2021).
  • [12] Bin, Q., Wan, L.-L., Nori, F., Wu, Y. & Lü, X.-Y. Out-of-time-order correlation as a witness for topological phase transitions. Phys. Rev. B 107, L020202 (2023).
  • [13] Robertson, J. H., Senese, R. & Essler, F. H. L. A simple theory for quantum quenches in the ANNNI model. SciPost Phys. 15, 032 (2023).
  • [14] Dağ, C. B., Uhrich, P., Wang, Y., McCulloch, I. P. & Halimeh, J. C. Detecting quantum phase transitions in the quasistationary regime of ising chains. Phys. Rev. B 107, 094432 (2023).
  • [15] Dağ, C. B., Wang, Y., Uhrich, P., Na, X. & Halimeh, J. C. Critical slowing down in sudden quench dynamics. Phys. Rev. B 107, L121113 (2023).
  • [16] Lakkaraju, L. G. C., Ghosh, S., Sadhukhan, D. & De, A. S. Framework of dynamical transitions from long-range to short-range quantum systems. arXiv preprint arXiv:2305.02945 (2023).
  • [17] Lakkaraju, L. G. C., Haldar, S. K. & Sen(De), A. Predicting a topological quantum phase transition from dynamics via multisite entanglement. Phys. Rev. A 109, 022436 (2024).
  • [18] Leibfried, D., Blatt, R., Monroe, C. & Wineland, D. Quantum dynamics of single trapped ions. Rev. Mod. Phys. 75, 281–324 (2003).
  • [19] Blatt, R. & Roos, C. F. Quantum simulations with trapped ions. Nature Phys. 8, 277–284 (2012).
  • [20] Richerme, P. et al. Non-local propagation of correlations in quantum systems with long-range interactions. Nature 511, 198–201 (2014).
  • [21] Neyenhuis, B. et al. Observation of prethermalization in long-range interacting spin chains. Sci. Adv. 3, e1700672 (2017).
  • [22] Kinoshita, T., Wenger, T. & Weiss, D. S. A quantum newton’s cradle. Nature 440, 900–903 (2006).
  • [23] Bloch, I., Dalibard, J. & Zwerger, W. Many-body physics with ultracold gases. Rev. Mod. Phys. 80, 885–964 (2008).
  • [24] Bloch, I., Dalibard, J. & Nascimbène, S. Quantum simulations with ultracold quantum gases. Nature Phys. 8, 267–276 (2012).
  • [25] Schreiber, M. et al. Observation of many-body localization of interacting fermions in a quasirandom optical lattice. Science 349, 842–845 (2015).
  • [26] Zeiher, J. et al. Many-body interferometry of a rydberg-dressed spin lattice. Nature Phys. 12, 1095–1099 (2016).
  • [27] Zeiher, J. et al. Coherent many-body spin dynamics in a long-range interacting ising chain. Phys. Rev. X 7, 041063 (2017).
  • [28] Bernien, H. et al. Probing many-body dynamics on a 51-atom quantum simulator. Nature 551, 579–584 (2017).
  • [29] Guardado-Sanchez, E. et al. Probing the quench dynamics of antiferromagnetic correlations in a 2d quantum ising spin system. Phys. Rev. X 8, 021069 (2018).
  • [30] Choi, S. et al. Observation of discrete time-crystalline order in a disordered dipolar many-body system. Nature 543, 221–225 (2017).
  • [31] Villa, L., Despres, J., Thomson, S. J. & Sanchez-Palencia, L. Local quench spectroscopy of many-body quantum systems. Phys. Rev. A 102, 033337 (2020).
  • [32] Villa, L., Thomson, S. J. & Sanchez-Palencia, L. Quench spectroscopy of a disordered quantum system. Phys. Rev. A 104, L021301 (2021).
  • [33] Villa, L., Thomson, S. J. & Sanchez-Palencia, L. Finding the phase diagram of strongly correlated disordered bosons using quantum quenches. Phys. Rev. A 104, 023323 (2021).
  • [34] Schneider, J. T., Despres, J., Thomson, S. J., Tagliacozzo, L. & Sanchez-Palencia, L. Spreading of correlations and entanglement in the long-range transverse ising chain. Phys. Rev. Res. 3, L012022 (2021).
  • [35] Kobayashi, K. & Motome, Y. Quantum reservoir probing of quantum information scrambling. arXiv preprint arXiv:2308.00898 (2023).
  • [36] Jaeger, H. & Haas, H. Harnessing nonlinearity: Predicting chaotic systems and saving energy in wireless communication. Science 304, 78–80 (2004).
  • [37] Tanaka, G. et al. Recent advances in physical reservoir computing: A review. Neural Networks 115, 100–123 (2019).
  • [38] Kobayashi, K. & Motome, Y. Thermally-robust spatiotemporal parallel reservoir computing by frequency filtering in frustrated magnets. Sci. Rep. 13, 15123 (2023).
  • [39] Fujii, K. & Nakajima, K. Harnessing disordered-ensemble quantum dynamics for machine learning. Phys. Rev. Appl. 8, 024030 (2017).
  • [40] Coldea, R. et al. Quantum Criticality in an Ising Chain: Experimental Evidence for Emergent E8 Symmetry. Science 327, 177–180 (2010).
  • [41] Selke, W. The annni model ― theoretical analysis and experimental application. Phys. Rep. 170, 213–264 (1988).
  • [42] Beccaria, M., Campostrini, M. & Feo, A. Density-matrix renormalization-group study of the disorder line in the quantum axial next-nearest-neighbor ising model. Phys. Rev. B 73, 052402 (2006).
  • [43] Chandra, A. K. & Dasgupta, S. Floating phase in the one-dimensional transverse axial next-nearest-neighbor ising model. Phys. Rev. E 75, 021105 (2007).
  • [44] Suzuki, S., Inoue, J.-i. & Chakrabarti, B. K. Quantum Ising Phases and Transitions in Transverse Ising Models (Springer, Heidelberg, 2013).
  • [45] Karrasch, C. & Schuricht, D. Dynamical phase transitions after quenches in nonintegrable models. Phys. Rev. B 87, 195104 (2013).
  • [46] Haldane, F. D. M. Continuum dynamics of the 1-d heisenberg antiferromagnet: Identification with the o(3) nonlinear sigma model. Phys. Lett. A 93, 464–468 (1983).
  • [47] Haldane, F. D. M. Nonlinear field theory of large-spin heisenberg antiferromagnets: Semiclassically quantized solitons of the one-dimensional easy-axis néel state. Phys. Rev. Lett. 50, 1153–1156 (1983).
  • [48] Botet, R. & Jullien, R. Ground-state properties of a spin-1 antiferromagnetic chain. Phys. Rev. B 27, 613–615 (1983).
  • [49] Botet, R., Jullien, R. & Kolb, M. Finite-size-scaling study of the spin-1 heisenberg-ising chain with uniaxial anisotropy. Phys. Rev. B 28, 3914–3921 (1983).
  • [50] Sakai, T. & Takahashi, M. Finite-size scaling study of s=1 xxz spin chain. J. Phys. Soc. Jpn. 59, 2688–2693 (1990).
  • [51] den Nijs, M. & Rommelse, K. Preroughening transitions in crystal surfaces and valence-bond phases in quantum spin chains. Phys. Rev. B 40, 4709–4734 (1989).
  • [52] Pollmann, F., Turner, A. M., Berg, E. & Oshikawa, M. Entanglement spectrum of a topological phase in one dimension. Phys. Rev. B 81, 064439 (2010).
  • [53] Pollmann, F., Berg, E., Turner, A. M. & Oshikawa, M. Symmetry protection of topological phases in one-dimensional quantum spin systems. Phys. Rev. B 85, 075125 (2012).
  • [54] Nomura, K. Spin-correlation functions of the s=1 heisenberg-ising chain by the large-cluster-decomposition monte carlo method. Phys. Rev. B 40, 9142–9146 (1989).
  • [55] Chen, W., Hida, K. & Sanctuary, B. C. Ground-state phase diagram of s=1s=1 XXZ\mathrm{XXZ} chains with uniaxial single-ion-type anisotropy. Phys. Rev. B 67, 104401 (2003).
  • [56] Su, W. P., Schrieffer, J. R. & Heeger, A. J. Solitons in polyacetylene. Phys. Rev. Lett. 42, 1698–1701 (1979).
  • [57] Su, W. P., Schrieffer, J. R. & Heeger, A. J. Soliton excitations in polyacetylene. Phys. Rev. B 22, 2099–2111 (1980).
  • [58] Hida, K. Crossover between the haldane-gap phase and the dimer phase in the spin-1/2 alternating heisenberg chain. Phys. Rev. B 45, 2207–2212 (1992).
  • [59] de Léséleuc, S. et al. Observation of a symmetry-protected topological phase of interacting bosons with rydberg atoms. Science 365, 775–780 (2019).
  • [60] Elben, A. et al. Many-body topological invariants from randomized measurements in synthetic quantum matter. Science Advances 6, eaaz3666 (2020).
  • [61] Meth, M. et al. Probing phases of quantum matter with an ion-trap tensor-network quantum eigensolver. Phys. Rev. X 12, 041035 (2022).
  • [62] Pollmann, F. & Turner, A. M. Detection of symmetry-protected topological phases in one dimension. Phys. Rev. B 86, 125441 (2012).
  • [63] Di Francesco, P., Mathieu, P. & Sénéchal, D. Conformal field theory. Graduate Texts in Contemporary Physics (Springer, Germany, 1997).
  • [64] Kitaev, A. Anyons in an exactly solved model and beyond. Ann. Phys. 321, 2–111 (2006). January Special Issue.
  • [65] White, S. R. Density matrix formulation for quantum renormalization groups. Phys. Rev. Lett. 69, 2863–2866 (1992).
  • [66] White, S. R. Density-matrix algorithms for quantum renormalization groups. Phys. Rev. B 48, 10345–10356 (1993).
  • [67] Schollwöck, U. The density-matrix renormalization group. Rev. Mod. Phys. 77, 259–315 (2005).
  • [68] Haegeman, J. et al. Time-dependent variational principle for quantum lattices. Phys. Rev. Lett. 107, 070601 (2011).
  • [69] Haegeman, J., Lubich, C., Oseledets, I., Vandereycken, B. & Verstraete, F. Unifying time evolution and optimization with matrix product states. Phys. Rev. B 94, 165116 (2016).
  • [70] Yang, M. & White, S. R. Time-dependent variational principle with ancillary krylov subspace. Phys. Rev. B 102, 094315 (2020).
  • [71] Fishman, M., White, S. R. & Stoudenmire, E. M. The ITensor Software Library for Tensor Network Calculations. SciPost Phys. Codebases 4 (2022).
  • [72] Fishman, M., White, S. R. & Stoudenmire, E. M. Codebase release 0.3 for ITensor. SciPost Phys. Codebases 4–r0.3 (2022).

ACKNOWLEDGEMENTS

This research was supported by a Grant-in-Aid for Scientific Research on Innovative Areas “Quantum Liquid Crystals” (KAKENHI Grant No. JP19H05825) from JSPS of Japan and JST CREST (No. JP-MJCR18T2). K.K. was supported by the Program for Leading Graduate Schools (MERIT-WINGS) and JSPS KAKENHI Grant Number JP24KJ0872. The computation in this work has been done using the facilities of the Supercomputer Center, the Institute for Solid State Physics, the University of Tokyo.

AUTHOR CONTRIBUTIONS

K.K. and Y.M. conceived the project. K.K performed the detailed calculations and analyzed the results under the supervision of Y.M. All authors contributed to writing the paper.

COMPETING INTERESTS

The authors have no conflicts of interest directly relevant to the content of this article.

Supplementary Information for

Quantum reservoir probing of quantum phase transitions
Kaito Kobayashi and Yukitoshi Motome

Department of Applied Physics, the University of Tokyo, Tokyo 113-8656, Japan

(Dated: )

Corresponding author. E-mail: [email protected]


Supplementary Note 1: Average of the estimation performance over various subsets of spatiotemporal indices

In the main text, we have demonstrated the efficacy of the quantum reservoir probing (QRP) in identifying quantum phase transitions through a statistical analysis of the effects of a local quantum quench, as described by the estimation performance Ri2(t)R^{2}_{i}(t) in Eq. (5) in the main text. Figures 2d and 3b in the main text illustrate the signatures of quantum phase transitions and quantum critical points, utilizing the mean value of Ri2(t)R^{2}_{i}(t), which is calculated over selected spatiotemporal indices (i,t)(i,t) to minimize extraneous effects from the edges of the systems. Particularly, in the case of the transverse-field Ising model in Fig. 2d, we have presented the field dependence of the mean estimation performance R2¯\overline{R^{2}}, which is obtained by averaging Ri2(t)R^{2}_{i}(t) for indices (i,t)Λ×{0<t10}(i,t)\in\Lambda\times\{0<t\leq 10\}, where Λ\Lambda encompasses the central 1111 sites. Similarly for the ANNNI model in Fig. 3b, we have exhibited R2¯\overline{R^{2}} defined over (i,t)(i,t) drawn from the set Λ×{0<t15}\Lambda\times\{0<t\leq 15\}, where Λ\Lambda also contains central 1111 sites. It is imperative to note that the quantum phases manifest distinguishable characteristics in Ri2(t)R^{2}_{i}(t) (Figs. 2a and 3a in the main text), and R2¯\overline{R^{2}} serves as an ancillary tool to further clarify these characteristics. In other words, this averaging operation is implemented only to enhance the clarity of the results, and R2¯\overline{R^{2}}, defined as the simple average of local quantities, remains inherently local, and does not include nonlocal quantum effects. In this Supplementary Note, we present R2¯\overline{R^{2}} defined for various subsets of spatiotemporal indices (i,t)(i,t), thus demonstrating the robustness of the QRP in detecting quantum phase transitions independent of the employed spatiotemporal subset.

Refer to caption
Figure S1: Average of the estimation performance in the transverse-field Ising model over various subsets of spatiotemporal indices. a The average R2¯\overline{R^{2}} over subsets (i,t)Λ×{0<t10}(i,t)\in\Lambda\times\{0<t\leq 10\}, where Λ\Lambda encompasses the central 11 sites (blue), central 7 sites (pink), central 15 sites (green), and central 21 sites (orange). b The average R2¯\overline{R^{2}} over a subsets (i,t)Λ×{t}(i,t)\in\Lambda\times\{t\}, where Λ\Lambda encompasses the central 11 sites. The time window is 0<t100<t\leq 10 (blue), 2.5<t7.52.5<t\leq 7.5 (pink), 0<t7.50<t\leq 7.5 (green), and 5t155\leq t\leq 15 (orange). The system size is N=49N=49 in both (a) and (b).

Figure S1 displays the average R2¯\overline{R^{2}} in the transverse-field Ising model as a function of the field gg in the evolution Hamiltonian (g)\mathcal{H}(g). In Fig. S1a, we show R2¯\overline{R^{2}} computed over various spatial index sets Λ\Lambda, while the temporal index range is held constant at {0<t10}\{0<t\leq 10\}. Conversely, in Fig. S1b, we present R2¯\overline{R^{2}} calculated over fixed sets Λ\Lambda, which encompass the central eleven sites, while changing the temporal index range. While each curve corresponding to the subset {(i,t)}\{(i,t)\} exhibits distinct shapes, the qualitative characteristics are consistent: R2¯\overline{R^{2}} decreases as it approaches g=1.0g=1.0, reaches a minimum at gcg_{c} due to the strong fluctuations, and subsequently rises with further increasing gg. These patterns robustly indicate the occurrence of a quantum phase transition, with the quantum critical point at gcg_{c} demarcating the transition between the ferromagnetically ordered phase and the quantum disordered phase.

Figure S2 demonstrates the averaged estimation performance in the ANNNI model as a function of the field strength. Figure S2a illustrates R2¯\overline{R^{2}} calculated over various spatial index sets Λ\Lambda with the temporal index range consistently set at {0<t15}\{0<t\leq 15\}. In addition, Fig. S2b illustrates R2¯\overline{R^{2}} calculated over a fixed spacial indices set Λ\Lambda that includes the central eleven sites, while varying the temporal index range. Similar to the observations in the transverse-field Ising model, the qualitative behaviors of R2¯\overline{R^{2}} in the ANNNI model remain consistent across different subsets {(i,t)}\{(i,t)\}, with all results corroborating the occurrence of a quantum phase transition around gc1.6g_{c}\simeq 1.6. However, a notable feature in the ANNNI model is the broader shape of the dip in R2¯\overline{R^{2}} near the quantum critical point, which leads to variations in the identified field strength at which R2¯\overline{R^{2}} reaches its minimum depending on the specific subset employed. Nevertheless, it is anticipated that this variance would decrease as the system size increases. A larger system would reduce the influence of edge effects, which are the principal cause of the subset {(i,t)}\{(i,t)\} dependency in R2¯\overline{R^{2}}. Consequently, in a larger system, the quantum critical point is expected to be identified with greater precision and clarity.

Refer to caption
Figure S2: Average of the estimation performance in the ANNNI model over various subsets of spatiotemporal indices. a The average R2¯\overline{R^{2}} over subsets (i,t)Λ×{0<t15}(i,t)\in\Lambda\times\{0<t\leq 15\}, where Λ\Lambda encompasses the central 11 sites (skyblue), central 7 sites (pink), central 15 sites (green), and central 21 sites (orange). b The average R2¯\overline{R^{2}} over subsets (i,t)Λ×{t}(i,t)\in\Lambda\times\{t\}, where Λ\Lambda encompasses the central 11 sites. The time window is 0<t150<t\leq 15 (skyblue), 5t105\leq t\leq 10 (pink), 2.5t12.52.5\leq t\leq 12.5 (green), and 0<t200<t\leq 20 (orange). The system size is N=49N=49 in both (a) and (b).

Supplementary Note 2: Dependence of the estimation performance on the deviation threshold

As explained in the Methods section in the main text, we introduced the threshold for deviations in the dynamics over multiple cases in the evaluation of the estimation performance. The average absolute deviation Δi(t)\Delta_{i}(t), defined in Eq. (13) in the main text, is employed as a criterion. Ri2(t)R^{2}_{i}(t) should be zero when Δi(t)\Delta_{i}(t) is negligible at the level of the numerical error, however, it may accidentally become nonzero sensitively reflecting the small deviations. To circumvent such numerical artifacts, we manually set Ri2(t)=0R^{2}_{i}(t)=0 when Δi(t)\Delta_{i}(t) is smaller than the threshold Δth\Delta_{\mathrm{th}}. Figures in the main text are obtained under Δth=105\Delta_{\mathrm{th}}=10^{-5}, whereas the qualitative behaviors remain consistent with different Δth\Delta_{\mathrm{th}} as shown in the following.

In Fig. S3, we present Ri2(t)R^{2}_{i}(t) in the transverse-field Ising model for three different Δth\Delta_{\mathrm{th}}; the results with Δth=105\Delta_{\mathrm{th}}=10^{-5} are illustrated in Fig. 2a in the main text. In Fig. S3a, Ri2(t)R^{2}_{i}(t) without the threshold is presented, where Ri2(t)R^{2}_{i}(t) is depicted as obtained. Remarkably, we observe nonzero Ri2(t)R^{2}_{i}(t) even outside the light cone, likely due to numerical artifacts. In Figs. S3b and S3c, we introduce the threshold Δth=106\Delta_{\mathrm{th}}=10^{-6} and Δth=104\Delta_{\mathrm{th}}=10^{-4}, respectively. The unphysical nonzero values of Ri2(t)R^{2}_{i}(t) outside the wavefronts are eliminated, mitigating numerical artifacts. Furthermore, the introduction of the threshold narrows the “width” of the wavefront, particularly evident at g=0.5g=0.5. However, once the effects of the local quench have propagated (i.e., inside the wave fronts), the behavior of Ri2(t)R^{2}_{i}(t) changes minimally, and the phase-dependent behavior remains consistent. Therefore, the introduction of the threshold to the QRP does not qualitatively alter the outcomes.

Refer to caption
Figure S3: Dependence of the estimation performance on the deviation threshold in the transverse-field Ising model. The spatiotemporal representation of Ri2(t)R^{2}_{i}(t), as depicted in Fig. 2a in the main text. Ri2(t)R^{2}_{i}(t) is calculated with the threshold a Δth=0.0\Delta_{\mathrm{th}}=0.0 (no threshold), b Δth=106\Delta_{\mathrm{th}}=10^{-6}, and c Δth=104\Delta_{\mathrm{th}}=10^{-4}.