- 1Instituto de Física Teórica UAM-CSIC, Universidad Autónoma de Madrid, Madrid, Spain
- 2Institute of Fundamental Physics IFF-CSIC, Madrid, Spain
- 3Barcelona Supercomputing Center, Barcelona, Spain
The study of many-body quantum systems out of equilibrium remains a significant challenge, with complexity barriers arising in both state- and operator-based representations. Here, we review the recent approaches based on finding better contraction strategies for the full spatiotemporal tensor networks that encode the path integral of the dynamics, as well as the conceptual integration of influence functionals, process tensors, and transfer matrices within the tensor network formalism. We discuss recent algorithmic developments, highlight the complexity of influence functionals in various dynamical regimes, and present consistent results of different communities, showing how ergodic dynamics render these functionals exponentially difficult to compress. Finally, we provide an outlook on strategies to encode complementary influence functional overlaps, paving the way for accurate descriptions of open and closed quantum systems with tensor networks.
1 Introduction
Many-body systems out of equilibrium still defy our understanding, and our intuition of the origin of their complexity continues to evolve. In the case of closed quantum systems, where the many-body system is ideally isolated from the rest of the world constituting its environment, the dynamics is governed by the unitary evolution dictated by the Schrödinger equation. Such a unitary evolution can be applied to the state of the system or to its operators. This simple fact leads to different pictures about the difficulty of solving the dynamics. If the evolution is applied on states, in the Schrödinger picture, they become increasingly complex, and simple tensor networks ansätze struggle to describe them with polynomial resources, something that is known as the entanglement barrier (Calabrese and Cardy, 2005; Läuchli and Kollath, 2008; Dubail, 2017).
Contrarily, if the evolution is applied to operators, there are specific forms of evolution that can be solved in the Heisenberg picture. For example, for spin–1/2 systems, Clifford circuits map Pauli strings into themselves, and thus, the dynamics can be efficiently described (Gottesman, 1998). Additionally, the dynamics governed by integrable Hamiltonians are conjectured to generate only a small amount of local operator entanglement, and thus, describing the evolution of local operators in such systems has the same complexity as describing a local quench of states, allowing for an efficient description (Prosen and Žnidarič, 2007; Prosen and Pižorn, 2007; Bertini et al., 2020a; Bertini et al., 2020b; Giudice et al., 2022; Thoenniss et al., 2023). However, generic interacting systems generate again a barrier of operator entanglement and thus suffer from the same shortcomings as the simulation of the evolution of the states. Given the complexity barriers in both state and operator representations, a growing trend is to adopt an open-system perspective. Instead of attempting to describe the full system dynamics, this approach focuses on the evolution of few-body correlation functions (Bañuls et al., 2009; Müller-Hermes et al., 2012; Surace et al., 2019; White et al., 2018; Frías-Pérez and Bañuls, 2022; Paeckel et al., 2019). This perspective naturally connects the dynamics of closed quantum systems to those of open systems. From the open-system viewpoint, the few bodies involved in the correlation functions define the system, while the remainder of the many-body system acts as the environment.
Traditionally, significant progress in understanding open-system dynamics has been made by studying a small subsystem, such as an atom in a cavity or an impurity in a metal, and describing it with master equations. While the full system and environment evolve unitarily, the subsystem undergoes dissipative evolution driven by a trace-preserving quantum channel. This framework provides a significant simplification compared to the full system’s description as the subsystem often has a finite-dimensional Hilbert space, allowing the quantum channel to map between finite-dimensional density matrices. However, obtaining such master equations has historically relied on analytical approximations, such as weak coupling or memory-less environments, defining Markovian systems where the current state suffices to predict future states. Even in this context, the precise definition of Markovianity remains a subject of debate as seen by comparing, for example, the definition in Rivas et al. (2010) with the one in Dowling et al. (2024).
This work reviews recent advances in designing tensor network algorithms to study out-of-equilibrium dynamics. These algorithms unify the descriptions of closed and open dynamics with minimal approximations, particularly for one-dimensional many-body systems. The starting points of these approaches are the spatiotemporal tensor networks that encode the path integral of the dynamics. Following the original Feynman–Vernon idea (Feynman and Vernon, 1963), we define influence functionals once we identify a region as the system, while the rest constitutes the environment. Partial integration of the path integral over the spatiotemporal degrees of freedom of the environment gives rise to the influence functionals. Furthermore, in this framework, it is easy to show how these influence functionals naturally emerge as partial traces of process tensors, which were initially introduced in the quantum information community (Chiribella et al., 2008) to generalize quantum channels and study the role of specific gates in quantum circuits. We also show how these distinct concepts naturally integrate within the spatiotemporal tensor network framework. Additionally, finding the correct influence functional corresponds to solving an open-system dynamics problem, where spatial transfer matrices act as quantum channels driving the dissipative evolution of influence functionals. Specifically, the transfer matrices evolve the influence functionals in space; that is, they generate the influence functional of a smaller system (larger environment) from that of a larger region or that of a smaller environment by incorporating new sites into the environment. This paper briefly outlines the algorithms developed to achieve this task.
After introducing all these objects and the algorithms that we practically use to compress them in simple tensor networks, we also review the known results about the complexity of the influence functionals for different classes of dynamics. We will thus unveil how, for generic ergodic dynamics, influence functionals are exponentially difficult to compress. We conclude with an outlook on the potential to accurately describe the dynamics of open and closed quantum systems using tensor networks. This involves attempting to directly encode the overlaps of complementary influence functionals (e.g., left and right) rather than the individual influence functionals separately.
2 Tensor network approach to the influence functional
In this section, we will see how the concept of influence functionals (IFs), introduced to describe how the environment affects a system within the path integral formulation of quantum mechanics (Feynman and Vernon, 1963), can be represented using tensor networks (TNs). The connection is made possible by the ideas of transverse contraction of the TN associated with the time evolution of a quantum system (Bañuls et al., 2009; Hastings and Mahajan, 2015; Frías-Pérez and Bañuls, 2022; Carignano et al., 2024) and of temporal matrix-product states (tMPSs) that can be used to describe these functionals.
To see this in more detail, let us take as a starting point the typical scenario of a quantum quench where one begins with a given initial state, usually a product state or the ground state of some Hamiltonian
When describing the dynamics of a one-dimensional closed system through tensor networks, one can integrate the Schrödinger equation by constructing the time-evolution operator
Because
We can then build the first-order Suzuki–Trotter decomposition,
or higher-order approximations, with a Trotter error of order
By decomposing each

Figure 1. Time runs from top to bottom. All quantities are represented as a brick-wall circuit (left column), with matrix-product operators (center column) and in an abstract representation (right column), where the black boundaries denote open indices. (a) A graphical representation of the first-order Suzuki–Trotter decomposition of the time-evolution operator

Figure 2. (a) By performing a singular value decomposition of elementary unitary gates derived from the Trotter expansion of the real-time dynamics induced by the two-qudit time-evolution operator
As a result of the discretization of the time evolution, any dynamical quantity we are interested in will be expressed in terms of numerous tensors, giving rise to a two-dimensional network that must be contracted. To give an example, the TN representing the time-dependent expectation value of a local operator acting on a few sites of the system,
The task of contracting two-dimensional TNs is, in general, exponentially hard (Schuch et al., 2007). The traditional prescriptions of quantum mechanics typically involve a contraction row by row along the temporal axis. In this case, the complexity of the contraction is dictated by the entanglement entropy of the time-evolved
by applying to them the columns associated with the neighboring site. In this way, the 2D TN equals the overlap between the two tMPSs representing the transverse contraction until adjacent columns. Thus, quantities like the Loschmidt echo are given by
Depending on the network structure associated with the desired dynamical quantity, further operations can be performed to carry out the transverse contraction in a more efficient way. In particular, if we are interested in computing the time evolution of the expectation values of a local operator or few-body correlators, it should be noticed that working out a transverse contraction may induce correlations between the forward and return contours if there is information partially traced out (see Figure 3a for a sketch). These correlations are long-ranged by default as a consequence of the setup, which makes the tMPS representation impractical. To overcome this problem, “folding” the network so that both contours are merged via a vectorization operation has been proposed (Bañuls et al., 2009):

Figure 3. For a setup including both forwards and backwards time evolution (cfr. Figure 1c), we show in (a) Graphical representation of how the correlations between contours are formed for a local entangled pair that propagates freely in opposite directions. The excitation marked in red remains in the contracted subsystem throughout the entire evolution so that it is traced in the transverse contraction. On the other hand, the excitation marked in blue leaves the subsystem and enters into the complement. Due to the spatial entanglement between the red and blue excitations, the transverse contraction leads to correlations between contours. (b) The associated long-range correlations are converted into short-range correlations through the folding operation.
Aside from providing a novel way to possibly circumvent the entanglement barrier, the idea of a transverse contraction can provide a natural bridge towards the formalism of open quantum systems and, more specifically, to the idea of an influence functional. To see this, let us recall that within the path integral formulation of quantum mechanics, open quantum systems can be studied by constructing the path integral of the system plus environment and, subsequently, integrating out the degrees of freedom of the latter. The result of this integration is what is called an influence functional (IF), corresponding to a function of the time-trajectories of the system.
The IF encodes the effects of the environment on the system and allows evolving the latter’s reduced density matrix. Being a function of time-dependent coordinates, the IF can be treated as a vector of the multi-time Hilbert space of the system (Petrat and Tumulka, 2014) and represented using a temporal MPS (Lerose et al., 2021a). In the context of our TN representation, the equivalence between the two becomes clear if we now observe that in the transverse contraction, we effectively trace out the degrees of freedom of a part of the many-body system (i.e., the environment), precisely encoding their influence on the rest (the system of interest) at different times through the resulting tMPSs. The folding operation provides another step in this direction, as its overlapping forward and backward time-evolution paths reproduce precisely the Schwinger–Keldysh contour, which is typically encoded in the IF construction.
Having established such a connection, one can leverage the powerful machinery of tensor networks to provide an efficient representation of these functionals encoding the dynamical properties of the system (Lerose et al., 2021a; Giudice et al., 2022; Lerose et al., 2023). We will elaborate further on this encoding in the following sections after introducing the relation of these objects with the process tensors.
3 The connection between process tensors and influence functionals
Process tensors, which we will define properly in the following, were originally introduced as a generalization of channels for operators. They were “introduced as a tool to optimize quantum circuit over a set of unknown gates for a given task” (Chiribella et al., 2008; Chiribella et al., 2009). They can also be connected with the idea of multiple-time states (Aharonov et al., 2007; Leifer, 2006; Leifer and Spekkens, 2013) discussed in the context of quantum foundations. In this section, we will introduce the concept of the process tensor, following the presentations of Dowling et al. (2024), Pollock et al. (2018), and Pollock (2018), and relate it to the tensorial objects encountered in Section 2. In the study of the evolution of open quantum systems, a significant challenge is that of generalizing the notion of Markovianity of classical stochastic processes (Pollock et al., 2018). A classical stochastic process is defined by the joint probability distribution of a stochastic variable
The generalization of these ideas to the quantum realm is not straightforward because the measurements needed to determine the control parameters X will disturb the process that we want to characterize. The multi-time process tensor formalism overcomes this problem by establishing a clear separation between the process and the control operations performed by the observer, leading to a definition of Markovianity that reduces to Equation 5 for classical processes.
The control operations are completely positive trace non-increasing actions on the system, representing, for example, a unitary transformation (trace preserving) or a possible result from a measurement (trace decreasing). These operations are often called instruments. We might be interested in a set of available instruments at an instant in time, each of them to be chosen with a given probability. We represent the set of available instruments as Kraus operators
We consider that the control operations intervene at
with
where

Figure 4. The process tensor is represented as the gray shadowed area, containing the information on the initial state of the system (S) and its environment (E), together with their joint evolution under the unitaries
Let us proceed to show how the process tensor language naturally emerges in the study of the closed dynamics of a quantum spin chain. Imagine that we start with the simplest scenario obtained by considering the evolution of two constituents. In order to make a connection with Figure 4, we can assume that one plays the role of the system and the other of the environment. We also limit ourselves to the case of a two-times process tensor where an initial product state evolves under the unitary

Figure 5. Tensor network representation of (a) the process tensor
Among the possible choices of instrument sets, one is relevant for the connection we want to make here: it consists of using as instruments the rows (columns) of the left (right) unitary emerging from the decomposition of the evolution operator shown in Figure 2a,
with
The set
Such that now
This specific choice of instruments leads us to the first connection between the multi-time process tensor and the influence functionals. To see this, let us now consider a system of four qudits evolving under a brick-wall circuit based on the same two-body gates as before. Splitting the system in half, we consider that the left and right qudits undergo two evolution steps as before, but now the central left and right qudits are connected by one further step. We can interpret such step as the insertion of the instruments
Recall that both the left and right process tensors and the instrument sets
Suppose we are interested in the Loschmidt echo, Equation 2, of the four spins, which, according to the results of Section 2, can be computed as the overlap of left and right temporal vectors
We are assuming that the initial state of the four spins is a product state, such that we can assign a well-defined initial state to the left two spins,
We turn now to the network associated with another typical dynamical quantity of interest, namely, the expectation value of an operator. Focusing on the left half of our four-spin system, this network will contain both
If we vectorize
At this point, it is straightforward to extend these quantities to the many-body setting considered in Section 2. We start by considering the tensor network that describes the evolution of the system under
where

Figure 6. (a) Contracting a patch of the spatiotemporal tensor network gives rise to a tensor with both spatial and temporal open legs, represented by black solid lines. When cutting the network, we split the two-body gates joining the left and the right halves following the recipe of Figure 2. The resulting tensor can be identified with the spatiotemporal state
In the same way, the left influence functional associated with the calculation of local observables is also obtained from Equation 15 in the many-body setting. Figures 6c, d display the standard and folded versions of the influence functionals.
In the above formulas, we have made clear that the influence functional is, in general, a mixed state. If one is interested in the compressibility of such a mixed state in terms of the matrix-product operator, one must consider the operator entanglement of the influence functional, that is, the entanglement entropy of its vectorized form that we have indicated with
In the following section, we will discuss some of the algorithms developed for efficient compression of the influence functional as temporal MPS and the complexity of such an encoding.
4 Algorithms for obtaining a tensor network encoding of the influence functional, dissipative time evolutions
The transverse contraction framework introduced in Section 2 for the description of the dynamics of quantum many-body systems opens the way to an efficient encoding of process tensors and influence functionals in the compact form of temporal MPS. The best procedure to contract the network associated with time evolution can vary depending on the system we are considering and the truncation procedure we choose, as described in the following. For a finite chain, one possibility is to start from the sides, identify the left- and rightmost columns as temporal MPS, and progressively apply columns of tMPO to them until reaching the center, truncating at each step to prevent their bond dimension from growing exponentially. Alternatively, for an infinite homogeneous system, one could start growing the system from the center in the spirit of the infinite version of the density matrix renormalization group (DMRG) algorithm.
In this section, we explain the main aspects to take into account for these transverse algorithms and the compression procedure for these tMPSs. The steps described here can be applied to both pure and mixed states, corresponding to the “folded” picture introduced in Section 2. As such, in the following description whenever possible we will consider generic left vectors
4.1 Cost functions used for compressing tMPSs
Transverse contraction algorithms rely on the successive applications of tMPOs to the tMPSs associated with the left and right halves of the system, which we will refer to as
For this iterative process, the simplest prescription would be to compress both
and the same for
where
However, the transverse contraction framework offers more possibilities when it comes to the optimization of the left and right vectors. The first insight in this direction was presented in Hastings and Mahajan (2015), where it was realized that the imaginary time-evolution equations for the RDM of the right vector2 can be rephrased as a DMRG algorithm for a system mirrored across the bipartition, with a slightly modified Hamiltonian. This can be intuitively seen by relating the (temporal) RDM
and analogously for

Figure 7. (a) Reduced transition matrices used to compute the generalized temporal entropies from the path integral encoding a Loschmidt echo. (b) Reduced transition matrices used to compute the generalized temporal entropy in the context of a quantum quench. Notice that they encode partial overlap of the left and right influence functional in that context. (c) Shows the temporal entanglement defined as the entanglement of the vectorized left (or right) influence functional. Notice that when there is a reflection symmetry, (c) differs from (b) because it requires considering a transposition of the right influence functional before taking the partial overlap. Also, given that we are using the vectorized version of the influence functional, its entanglement actually represents the operator entanglement of the influence functional, which is the relevant version if we are interested in understanding how much we can compress the influence functional as an MPO.
An extended notion of entropy based upon the RTMs can be defined by substituting
The relationship of these interesting new quantities with the complexity of representing the associated tensor network is still the object of investigation (Hastings and Mahajan, 2015; Carignano et al., 2024; Carignano and Tagliacozzo, 2024). In fact, the compression procedure based upon RTMs is not straightforward: unlike the (reduced) DMs,
As such, we will focus on this type of truncation here while recalling that other methods based on bi-orthogonalization techniques can also be used when dealing with non-Hermitian transfer matrices (Wang and Xiang, 1997; Xiang, 1998; Zhong et al., 2024).
4.2 Compression procedure
The compression of the tMPS is carried out by constructing the desired cost function, such as the RDM (Bañuls et al., 2009) or the RTM (Hastings and Mahajan, 2015; Carignano et al., 2024), and approximating it to a given rank for every possible bipartition. This is commonly achieved through singular value decompositions of the relevant objects.
While the truncation over RDM is amply covered in standard TN literature, let us briefly review the procedure for the case of an RTM. The notation used is as follows (see Figures 8a, b): given a tMPS, we will use Greek letters for the physical indices and Latin letters for the bond ones. In addition, given the bipartition introduced above, we define the matrices

Figure 8. Pictorial notation for the compression procedure assuming folded tMPSs. (a) Bipartition of a tMPS into top and bottom subsystems. (b) Definition of
With these definitions, we now proceed to explain the compression procedure. This is done by truncating the RTMs (Equation 19) associated with the different bipartitions, which are consecutively obtained by tracing the temporal sites from the
where
Given that
and repeat the process until all the bonds are truncated.
It should be noted that, depending on which dynamical quantity we are interested in as well as the cost function used, a meaningful truncation can require starting from

Figure 9. Two-dimensional TN representing the RTMs Equations 19, 22. (a) In case of not including the operator, the partial contraction from t to T gives
By following this prescription, the contraction of the bottom rows gives the Heisenberg evolution
This extra care concerning the directionality of the compression is a consequence of having both forward and return contours merged through the folding operation. For this reason, quantities with only the forward contour, such as the Loschmidt echo, do not present this problem. Indeed, for the Loschmidt echo starting from either the
5 On the complexity of process tensors and of the influence functional
Having elucidated the connection between temporal MPS, influence functionals, and process tensors, we can now employ the techniques described in the previous section in order to complete the picture of the cost of encoding influence functionals as tensor networks, filling some of the gaps in the current understanding of their complexity. We begin by recalling some results on tMPS complexity in the literature and then discuss their implications for process tensors.
5.1 Tensor network results based on generalized temporal entanglement
As discussed in Section 4, in the framework of transverse contractions, the generalized temporal entropy can be seen as a measure describing the complexity of a faithful description of the network contraction using tMPSs.
5.1.1 Expectation values of local operators
A first result on the complexity of the tMPS associated with the time evolution of the expectation value of a local operator was given by Carignano et al. (2024). As explained in Section 4.2, one can define folded tMPSs with the insertion of the operator
This is especially true for
where
Because the rank of
On the other hand, in the case of ergodic dynamics, one expects that the operator entanglement grows linearly with time (Prosen and Pižorn, 2007), implying an exponential cost in representing faithfully the IF for this problem. This was confirmed by Carignano et al. (2024), where some of the specific cases considered were found to saturate this bound.
5.1.2 Loschmidt echo
The transverse contraction framework also allows providing an analytical estimate of the computational complexity of evaluating the TN associated with the Loschmidt echo
The connection in the continuum limit was made again thanks to the path integral formulation, which maps the quench geometry to that of an infinite strip. CFT can then provide all the relevant information on the spectrum of the transfer matrix along the strip, as well as the generalized temporal entropy associated with a time-like cut, which can be mapped back to the reduced transition matrices

Figure 10. Examples of generalized second Rényi temporal entropies
5.2 Connections with other approaches
It should now be clear that one of the reasons the tensor network community has considered influence functionals and their complexity—measured in terms of temporal entropy—is the hope that these objects might be easier to encode than the evolution of states and operators, which are generally exponentially difficult to compress as simple tensor networks.
Recently, we have proposed using generalized temporal entanglement, computed from the path integral of the expectation value of local operators, as a measure of the cost of simulating these expectation values (Carignano et al., 2024). Prior studies have instead analyzed the entanglement of independent influence functionals
Numerical observations (Bañuls et al., 2009; Frías-Pérez and Bañuls, 2022) have shown that the temporal entanglement of the influence functional
Some authors have suggested that space–time duality might help overcome the entanglement barrier (Lerose et al., 2023). The evidence gathered here, however, points to the contrary: for ergodic systems, influence functionals seem to be exponentially difficult to compress using temporal matrix-product states (MPSs). This may seem counterintuitive as ergodic systems observed locally are expected to exhibit Markovian dynamics with little or no memory effects. Consequently, one might expect that the quantum channel governing the evolution of a local operator could be encoded using a temporal MPS with a small bond dimension, naturally leading to the exponential decay of correlations in time. However, we will review why this is not the case.
The first piece of evidence comes from Carignano et al. (2024), where it was shown that the bond dimension of a temporal MPS encoding the evolution of a local operator is upper-bounded by the bond dimension required to describe the operator’s evolution itself. Given the results in Prosen and Pižorn (2007) and subsequent works, the bond dimension necessary to describe operator evolution in the Heisenberg picture is expected to grow exponentially in time for ergodic systems. Such a growth has even been proposed as a definition of ergodicity. Consequently, the bond dimension of the temporal MPS encoding the influence functional is also bounded by an exponential function of time.
Building upon these insights, we incorporate findings from quantum information studies on the complexity of the process tensor in ergodic systems. Notably, the definitions of ergodicity and quantum chaos remain debated. Here, we adopt the operational definition introduced by Dowling et al. (2024), which generalizes the classical notion that chaotic systems exhibit extreme sensitivity to perturbations. In this framework, they define the “orthogonality of butterfly flutters,” meaning that two orthogonal choices of instruments applied to the same process tensor should produce orthogonal system-environment states. Furthermore, such orthogonality should not be rectifiable by a finite-depth unitary transformation, encapsulating the notion of information scrambling. Dowling et al. (2024) demonstrate that, under this definition, the process tensor has maximal entanglement for any space-time cut, implying that its temporal MPS representation must have an exponentially large bond dimension.
By taking the partial trace of the process tensor over its spatial degrees of freedom, we deduce that the influence functional—representing a maximally mixed state—is similarly encoded by an MPO with an exponentially large bond dimension. Additionally, the fully scrambled entanglement implies that this maximally mixed state cannot be factorized into independent mixed states at each time step, unlike in toy models built from swap tensors (Bañuls et al., 2009; Müller-Hermes et al., 2012). Furthermore, Dowling et al. (2024) establish that the butterfly flutter definition of ergodicity encompasses the linear growth of local operator entanglement as a special case. Thus, we complement the upper bound obtained by Carignano et al. (2024), which relies on local operator entanglement growth, with results on the structure of spatiotemporal entanglement in the process tensor, leading to the conclusion that, in general, influence functionals of ergodic systems exhibit linearly growing temporal entanglement and are consequently difficult to compress as MPOs.
Parallel findings in the many-body physics community further support these conclusions. Studies on matrix elements of Floquet circuits between initial and final singlet states (Ippoliti and Khemani, 2021; Lu and Grover, 2021; Ippoliti et al., 2022), which directly relate to
A significant implication of these results is that despite transverse evolution being dissipative (as discussed in earlier sections), its strength in an ergodic system is insufficient to induce an entanglement transition in its stationary states (Li et al., 2019), which always remain in a volume-law phase as if the evolution were unitary.
These findings suggest that encoding separate left and right influence functionals in ergodic systems is exponentially hard. However, in other contexts—such as many-body localized dynamics—transverse evolution can induce an entanglement transition, potentially simplifying process tensor encoding (Ippoliti and Khemani, 2021; Lu and Grover, 2021; Ippoliti et al., 2022).
Finally, these results raise concerns about the scalability of recent algorithms in open quantum systems. Many assume an environment initially in equilibrium with a thermal bath, which for integrable environments has been shown to yield highly compressible influence functionals (Makri and Makarov, 1995; Strathearn et al., 2018; Cygorek et al., 2022). However, it remains unclear whether this simplification arises from thermal equilibrium assumptions or from the non-ergodic nature of the environment, which typically consists of free oscillators. A similar question was explored by Ye and Chan (2021), where structured baths undergoing unitary dynamics were considered, demonstrating that the compressibility of the environment’s influence functional strongly depends on interaction strength. These observations reinforce the idea that strongly interacting ergodic systems possess complex influence functionals that are inherently difficult to compress.
6 Conclusion
In these notes, we have reviewed the connections among various objects studied in different quantum physics communities within the unifying framework of tensor networks. We have reviewed how both the return probabilities defining the Loschmidt echoes and the time evolution of local expectation values in a closed quantum many-body system can be encoded in the contraction of spatiotemporal tensor networks. Depending on the type of evolution and the system under consideration, such TNs may involve discretizing the path integral of a continuous system in space and time or simply contracting the quantum circuits that describe Floquet dynamics in lattice models. Regardless of these specific details, we have shown that contracting spatiotemporal patches of the tensor network results in the same objects: process tensors. These process tensors have been studied for different reasons in the quantum information community. In the context of tensor networks for characterizing out-of-equilibrium dynamics in many-body quantum systems, these objects are of interest due to the assumption that they might be easier to compress as matrix-product states than the full evolving states.
This view has so far been confirmed only for integrable systems, where we have shown that the complexity of the influence functional is upper-bounded by the local operator entanglement, which is conjectured to grow only logarithmically with time (Carignano and Tagliacozzo, 2024). Unfortunately, the same upper bound indicates that for generic ergodic dynamics, the complexity of the influence functional could grow exponentially with time. By exploiting the connection between influence functionals and process tensors reviewed here and using the definition of ergodicity proposed by Dowling and Modi (2024), we find that the upper bound obtained by Carignano and Tagliacozzo (2024) is generally saturated. This implies that influence functionals for ergodic systems are generically difficult to compress as temporal matrix-product states. These findings align with the results of Lu and Grover (2021) and Ippoliti et al. (2022), which analyze the scaling of temporal entropy in the context of the Loschmidt echoes for different classes of dynamics.
In parallel, in the open quantum systems community, there is a similarly optimistic impression that influence functionals of thermalized environments might be efficiently compressible as matrix-product operators (Strathearn et al., 2018; Cygorek et al., 2022). Initial results for Gaussian environments appear consistent with the observation that integrable systems yield simple environments. An intriguing question remains whether strongly interacting ergodic environments at equilibrium can also be easily compressible as temporal matrix-product operators, which, from our perspective, has not yet been convincingly demonstrated, as discussed by Park et al. (2024). We have also shown that extracting influence functionals requires a set of tensor network algorithms with strong connections to the density matrix renormalization group methods for open systems. In the thermodynamic limit, these algorithms are equivalent to those used to determine the stationary state of quantum channels, which is described by the spatial transfer matrix we defined.
The temporal entropy of the separated left and right influence functionals (the leading left and right eigenvectors of such a transfer matrix) turn out to not be gauge invariant. To address this, we have begun defining generalized temporal entropies that directly account for the overlap of the leading left and right eigenvectors, the left-right influence functionals (Carignano et al., 2024). These entropies are generally complex-valued; however, in some contexts, their real part exhibits only a mild growth over time (Carignano and Tagliacozzo, 2024). This suggests the possibility of encoding the overlap of left and right influence functionals using appropriate temporal matrix-product operators. Furthermore, we have recently shown that such generalized entropies can be measured in experiments by appropriately designed quenches on replicated systems (Bou-Comas et al., 2024). Generalized entropies thus give hope that the ergodic dynamics can still be efficiently simulated with tensor networks.
Transforming this hope into a concrete tensor network strategy is an active area of research. Currently, one of the main obstacles is to understand how to connect the scaling of a complex-valued pseudo-entropy with the complexity of the corresponding generalized transition matrix and its compressibility in terms of tensor networks. Regardless of these open questions, we have reviewed here how the field of partial contractions of spatiotemporal tensor networks encoding the path integrals of many-body quantum systems out of equilibrium, both in open and closed scenarios, is relatively young and very promising, having produced fertile connections and enhanced our understanding of out-of-equilibrium dynamics. We thus firmly believe that this field holds great potential for further breakthroughs.
While completing this manuscript, a new preprint (Milekhin et al., 2025) that addresses topics related to those discussed here has appeared on the arXiv.
Author contributions
SC-R: Visualization, Writing – original draft, Writing – review and editing. AB-C: Visualization, Writing – original draft, Writing – review and editing. JS: Writing – original draft, Writing – review and editing. EL: Writing – original draft, Writing – review and editing. LT: Writing – original draft, Writing – review and editing. SC: Writing – original draft, Writing – review and editing.
Funding
The author(s) declare that financial support was received for the research and/or publication of this article. We acknowledge support from the Proyecto Sinérgico CAM Y2020/TCS-6545 NanoQuCo-CM, the CSIC Research Platform on Quantum Technologies PTI-001, the Grant TED2021-130552B-C22 funded by MCIN/AEI/10.13039/501100011033 and by the “European Union NextGenerationEU/PRTR,” and Grant PID2021-127968NB-I00 funded by MCIN/AEI/10.13039/501100011033. ABC is supported by Grant MMT24-IFF-01, and SC is supported by Grant C005/24-ED CV1. The funding for these actions/grants and contracts comes from the European Union’s Recovery and Resilience Facility-Next Generation, in the framework of the General Invitation of the Spanish Government’s public business entity Red.es to participate in talent attraction and retention programs within Investment 4 of Component 19 of the Recovery, Transformation and Resilience Plan.
Acknowledgments
We acknowledge the discussion and collaboration on topics related to this review with Carlos Ramos Marimon, Jacopo De Nardis, and Guglielmo Lami.
Conflict of interest
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Generative AI statement
The author(s) declare that no Generative AI was used in the creation of this manuscript.
Publisher’s note
All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.
Footnotes
1Alternatively, one can decompose the
2As usual, the same reasoning can also be applied to
3See also Sirker and Klümper (2005) and Andraschko and Sirker (2014) for related works in this direction.
References
Aharonov, Y., Popescu, S., Tollaksen, J., and Vaidman, L. (2007). Multiple-time states and multiple-time measurements in quantum mechanics. Phys. Rev. A . Coll. Park. 79, 052110. doi:10.1103/PhysRevA.79.052110
Andraschko, F., and Sirker, J. (2014). Dynamical quantum phase transitions and the Loschmidt echo: a transfer matrix approach. Phys. Rev. B 89, 125120. doi:10.1103/PhysRevB.89.125120
Bañuls, M. C., Hastings, M. B., Verstraete, F., and Cirac, J. I. (2009). Matrix product states for dynamical simulation of infinite chains. Phys. Rev. Lett. 102, 240603. doi:10.1103/PhysRevLett.102.240603
Bertini, B., Kos, P., and Prosen, T. (2020a). Operator entanglement in local quantum circuits I: chaotic dual-unitary circuits. SciPost Phys. 8, 067. doi:10.21468/SciPostPhys.8.4.067
Bertini, B., Kos, P., and Prosen, T. (2020b). Operator entanglement in local quantum circuits II: solitons in chains of qubits. SciPost Phys. 8, 068. doi:10.21468/SciPostPhys.8.4.068
Bou-Comas, A., Marimón, C. R., Schneider, J. T., Carignano, S., and Tagliacozzo, L. (2024). Measuring temporal entanglement in experiments as a hallmark for integrability.
Calabrese, P., and Cardy, J. (2005). Evolution of entanglement entropy in one-dimensional systems. J. Stat. Mech. Theory Exp. 2005, P04010. doi:10.1088/1742-5468/2005/04/P04010
Carignano, S., Marimón, C. R., and Tagliacozzo, L. (2024). Temporal entropy and the complexity of computing the expectation value of local operators after a quench. Phys. Rev. Res. 6, 033021. doi:10.1103/PhysRevResearch.6.033021
Carignano, S., and Tagliacozzo, L. (2024). Loschmidt echo, emerging dual unitarity and scaling of generalized temporal entropies after quenches to the critical point. doi:10.48550/arXiv.2405.14706
Chiribella, G., D’Ariano, G. M., and Perinotti, P. (2008). Quantum circuit architecture. Phys. Rev. Lett. 101, 060401. doi:10.1103/PhysRevLett.101.060401
Chiribella, G., D’Ariano, G. M., and Perinotti, P. (2009). Theoretical framework for quantum networks. Phys. Rev. A 80, 022339. doi:10.1103/PhysRevA.80.022339
Cygorek, M., Cosacchi, M., Vagov, A., Axt, V. M., Lovett, B. W., Keeling, J., et al. (2022). Simulation of open quantum systems by automated compression of arbitrary environments. Nat. Phys. 18, 662–668. doi:10.1038/s41567-022-01544-9
Doi, K., Harper, J., Mollabashi, A., Takayanagi, T., and Taki, Y. (2023). Pseudo entropy in dS/CFT and time-like entanglement entropy. Phys. Rev. Lett. 130, 031601. doi:10.1103/PhysRevLett.130.031601
Dowling, N., and Modi, K. (2024). Operational metric for quantum chaos and the corresponding spatiotemporal-entanglement structure. PRX Quantum 5, 010314. doi:10.1103/PRXQuantum.5.010314
Dowling, N., Modi, K., Muñoz, R. N., Singh, S., and White, G. A. L. (2024). Capturing long-range memory structures with tree-geometry process tensors. Phys. Rev. X 14, 041018. doi:10.1103/physrevx.14.041018
Dubail, J. (2017). Entanglement scaling of operators: a conformal field theory approach, with a glimpse of simulability of long-time dynamics in 1+1d. J. Phys. A Math. Theor. 50, 234001. doi:10.1088/1751-8121/aa6f38
Feynman, R., and Vernon, F. (1963). The theory of a general quantum system interacting with a linear dissipative system. Ann. Phys. 24, 118–173. doi:10.1016/0003-4916(63)90068-X
Foligno, A., Zhou, T., and Bertini, B. (2023). Temporal entanglement in chaotic quantum circuits. Phys. Rev. X 13, 041008. doi:10.1103/PhysRevX.13.041008
Frías-Pérez, M., and Bañuls, M. C. (2022). Light cone tensor network and time evolution. Phys. Rev. B 106, 115117. doi:10.1103/PhysRevB.106.115117
Giudice, G., Giudici, G., Sonner, M., Thoenniss, J., Lerose, A., Abanin, D. A., et al. (2022). Temporal entanglement, quasiparticles, and the role of interactions. Phys. Rev. Lett. 128, 220401. doi:10.1103/PhysRevLett.128.220401
Gottesman, D. (1998). Theory of fault-tolerant quantum computation. Phys. Rev. A 57, 127–137. doi:10.1103/PhysRevA.57.127
Hastings, M. B., and Mahajan, R. (2015). Connecting entanglement in time and space: improving the folding algorithm. Phys. Rev. A 91, 032306. doi:10.1103/PhysRevA.91.032306
Ippoliti, M., and Khemani, V. (2021). Postselection-free entanglement dynamics via spacetime duality. Phys. Rev. Lett. 126, 060501. doi:10.1103/PhysRevLett.126.060501
Ippoliti, M., Rakovszky, T., and Khemani, V. (2022). Fractal, logarithmic and volume-law entangled non-thermal steady states via spacetime duality. Phys. Rev. X 12, 011045. doi:10.1103/physrevx.12.011045
Läuchli, A. M., and Kollath, C. (2008). Spreading of correlations and entanglement after a quench in the one-dimensional Bose–Hubbard model. J. Stat. Mech. Theory Exp. 2008, P05018. doi:10.1088/1742-5468/2008/05/P05018
Leifer, M. S. (2006). Quantum dynamics as an analog of conditional probability. Phys. Rev. A 74, 042310. doi:10.1103/PhysRevA.74.042310
Leifer, M. S., and Spekkens, R. W. (2013). Towards a formulation of quantum theory as a causally neutral theory of Bayesian inference. Phys. Rev. A 88, 052130. doi:10.1103/PhysRevA.88.052130
Lerose, A., Sonner, M., and Abanin, D. A. (2021a). Influence matrix approach to many-body floquet dynamics. Phys. Rev. X 11, 021040. doi:10.1103/physrevx.11.021040
Lerose, A., Sonner, M., and Abanin, D. A. (2021b). Scaling of temporal entanglement in proximity to integrability. Phys. Rev. B 104, 035137. doi:10.1103/PhysRevB.104.035137
Lerose, A., Sonner, M., and Abanin, D. A. (2021c). Influence matrix approach to many-body floquet dynamics. Phys. Rev. X 11, 021040. doi:10.1103/PhysRevX.11.021040
Lerose, A., Sonner, M., and Abanin, D. A. (2023). Overcoming the entanglement barrier in quantum many-body dynamics via space-time duality. Phys. Rev. B 107, L060305. doi:10.1103/physrevb.107.l060305
Li, Y., Chen, X., and Fisher, M. P. A. (2019). Measurement-driven entanglement transition in hybrid quantum circuits. Phys. Rev. B 100, 134306. doi:10.1103/PhysRevB.100.134306
Lu, T. C., and Grover, T. (2021). Spacetime duality between localization transitions and measurement-induced transitions. PRX Quantum 2, 040319. doi:10.1103/PRXQuantum.2.040319
Makri, N., and Makarov, D. E. (1995). Tensor propagator for iterative quantum time evolution of reduced density matrices. I. Theory. J. Chem. Phys. 102, 4600–4610. doi:10.1063/1.469508
Milekhin, A., Adamska, Z., and Preskill, J. (2025). Observable and computable entanglement in time. doi:10.48550/arXiv.2502.12240
Müller-Hermes, A., Cirac, J. I., and Bañuls, M. C. (2012). Tensor network techniques for the computation of dynamical observables in one-dimensional quantum spin systems. New J. Phys. 14, 075003. doi:10.1088/1367-2630/14/7/075003
Nakata, Y., Takayanagi, T., Taki, Y., Tamaoka, K., and Wei, Z. (2021). New holographic generalization of entanglement entropy. Phys. Rev. D. 103, 026005. doi:10.1103/PhysRevD.103.026005
Paeckel, S., Köhler, T., Swoboda, A., Manmana, S. R., Schollwöck, U., and Hubig, C. (2019). Time-evolution methods for matrix-product states. Ann. Phys. 411, 167998. doi:10.1016/j.aop.2019.167998
Park, G., Ng, N., Reichman, D. R., and Chan, G. K. L. (2024). Tensor network influence functionals in the continuous-time limit: connections to quantum embedding, bath discretization, and higher-order time propagation. Phys. Rev. B 110, 045104. doi:10.1103/PhysRevB.110.045104
Petrat, S., and Tumulka, R. (2014). Multi-time wave functions for quantum field theory. Ann. Phys. 345, 17–54. doi:10.1016/j.aop.2014.03.004
Pižorn, I., and Prosen, T. (2009). Operator space entanglement entropy in $XY$ spin chains. Phys. Rev. B 79, 184416. doi:10.1103/PhysRevB.79.184416
Pollock, F. A., Rodríguez-Rosario, C., Frauenheim, T., Paternostro, M., and Modi, K. (2018). Non-Markovian quantum processes: complete framework and efficient characterization. Phys. Rev. A 97, 012127. doi:10.1103/PhysRevA.97.012127
Pollock, F. A., Rodríguez-Rosario, C., Frauenheim, T., Paternostro, M., and Modi, K. (2018). Operational markov condition for quantum processes. Phys. Rev. Lett. 120, 040405. doi:10.1103/PhysRevLett.120.040405
Prosen, T., and Pižorn, I. (2007). Operator space entanglement entropy in a transverse ising chain. Phys. Rev. A—Atomic, Mol. Opt. Phys. 76, 032316. doi:10.1103/physreva.76.032316
Prosen, T., and Žnidarič, M. (2007). Is the efficiency of classical simulations of quantum dynamics related to integrability? Phys. Rev. E 75, 015202. doi:10.1103/PhysRevE.75.015202
Rivas, Á., Huelga, S. F., and Plenio, M. B. (2010). Entanglement and non-markovianity of quantum evolutions. Phys. Rev. Lett. 105, 050403. doi:10.1103/PhysRevLett.105.050403
Schuch, N., Wolf, M. M., Verstraete, F., and Cirac, J. I. (2007). Computational complexity of projected entangled pair states. Phys. Rev. Lett. 98, 140506. doi:10.1103/physrevlett.98.140506
Sirker, J., and Klümper, A. (2005). Real-time dynamics at finite temperature by DMRG: a path-integral approach. Phys. Rev. B 71, 241101. doi:10.1103/PhysRevB.71.241101
Sonner, M., Lerose, A., and Abanin, D. A. (2021). Influence functional of many-body systems: temporal entanglement and matrix-product state representation. Ann. Phys. 435, 168677. doi:10.1016/j.aop.2021.168677
Strathearn, A., Kirton, P., Kilda, D., Keeling, J., and Lovett, B. W. (2018). Efficient non-Markovian quantum dynamics using time-evolving matrix product operators. Nat. Commun. 9, 3322. doi:10.1038/s41467-018-05617-3
Surace, J., Piani, M., and Tagliacozzo, L. (2019). Simulating the out-of-equilibrium dynamics of local observables by trading entanglement for mixture. Phys. Rev. B 99, 235115. doi:10.1103/PhysRevB.99.235115
Tang, W., Verstraete, F., and Haegeman, J. (2025). Matrix product state fixed points of non-hermitian transfer matrices. Phys. Rev. B 111, 035107. doi:10.1103/PhysRevB.111.035107
Thoenniss, J., Lerose, A., and Abanin, D. A. (2023). Nonequilibrium quantum impurity problems via matrix-product states in the temporal domain. Phys. Rev. B 107, 195101. doi:10.1103/PhysRevB.107.195101
Tirrito, E., Robinson, N. J., Lewenstein, M., Ran, S. J., and Tagliacozzo, L. (2018). Characterizing the quantum field theory vacuum using temporal matrix product states. arXiv preprint arXiv:1810.08050.
Verstraete, F., and Cirac, J. I. (2006). Matrix product states represent ground states faithfully. Phys. Rev. B 73, 094423. doi:10.1103/PhysRevB.73.094423
Vidal, G. (2004). Efficient simulation of one-dimensional quantum many-body systems. Phys. Rev. Lett. 93, 040502. doi:10.1103/physrevlett.93.040502
Wang, X., and Xiang, T. (1997). Transfer-matrix density-matrix renormalization-group theory for thermodynamics of one-dimensional quantum systems. Phys. Rev. B 56, 5061–5064. doi:10.1103/physrevb.56.5061
White, C. D., Zaletel, M., Mong, R. S. K., and Refael, G. (2018). Quantum dynamics of thermalizing systems. Phys. Rev. B 97, 035127. doi:10.1103/PhysRevB.97.035127
Xiang, T. (1998). Thermodynamics of quantum heisenberg spin chains. Phys. Rev. B 58, 9142–9149. doi:10.1103/physrevb.58.9142
Yao, J., and Claeys, P. W. (2024). Temporal entanglement profiles in dual-unitary Clifford circuits with measurements. doi:10.48550/arXiv.2404.14374
Ye, E., and Chan, G. K. L. (2021). Constructing tensor network influence functionals for general quantum dynamics. J. Chem. Phys. 155, 044104. doi:10.1063/5.0047260
Keywords: open quantum systems, closed quantum dynamics, tensor networks, influence functionals, process tensor, temporal entropies, generalized entropies
Citation: Cerezo-Roquebrún S, Bou-Comas A, Schneider JT, López E, Tagliacozzo L and Carignano S (2025) Spatio-temporal tensor-network approaches to out-of-equilibrium dynamics bridging open and closed systems. Front. Quantum Sci. Technol. 4:1568471. doi: 10.3389/frqst.2025.1568471
Received: 29 January 2025; Accepted: 06 March 2025;
Published: 12 May 2025.
Edited by:
Jorge Yago Malo, University of Pisa, ItalyReviewed by:
Tomohiro Hashizume, University of Hamburg, GermanyPeter Kirton, University of Strathclyde, United Kingdom
Copyright © 2025 Cerezo-Roquebrún, Bou-Comas, Schneider, López, Tagliacozzo and Carignano. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: Luca Tagliacozzo, bHVjYS50YWdsaWFjb3p6b0BpZmYuY3NpYy5lcw==