Tensor Networks for Lattice Gauge Theories beyond one dimension:
a Roadmap

Giuseppe Magnifico\orcidlink0000-0002-7280-445X Dipartimento di Fisica, Università di Bari, I-70126 Bari, Italy. Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Bari, I-70125 Bari, Italy. Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy.    Giovanni Cataldi\orcidlink0000-0002-9073-8978 Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy.    Marco Rigobello\orcidlink0000-0002-4544-3513 Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy.   
Peter Majcen \orcidlink0009-0009-7020-7246
Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy.
   Daniel Jaschke\orcidlink0000-0001-7658-3546 Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy. Institute for Complex Quantum Systems, Ulm University, D-89069 Ulm, Germany    Pietro Silvi\orcidlink0000-0001-5279-7064 Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy.    Simone Montangero\orcidlink0000-0002-8882-2169 Dipartimento di Fisica e Astronomia “G. Galilei”, Università di Padova, I-35131 Padova, Italy. Padua Quantum Technologies Research Center, Università degli Studi di Padova Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Padova, I-35131 Padova, Italy.
(July 3, 2024)
Abstract

Tensor network methods are a class of numerical tools and algorithms to study many-body quantum systems in and out of equilibrium, based on tailored variational wave functions. They have found significant applications in simulating lattice gauge theories approaching relevant problems in high-energy physics. Compared to Monte Carlo methods, they do not suffer from the sign problem, allowing them to explore challenging regimes such as finite chemical potentials and real-time dynamics. Further development is required to tackle fundamental challenges, such as accessing continuum limits or computations of large-scale quantum chromodynamics. In this work, we review the state-of-the-art of Tensor Network methods and discuss a possible roadmap for algorithmic development and strategies to enhance their capabilities and extend their applicability to open high-energy problems. We provide tailored estimates of the theoretical and computational resource scaling for attacking large-scale lattice gauge theories.

Gauge theories play a role of paramount importance in our understanding and description of the fundamental constituents of matter, their spectrum, and their interactions. At low energies, gauge theories characterize a large variety of collective phases of matter and phenomena, such as ferromagnetic superconductivity, spin liquids, topological order, and the fractional quantum Hall effect [1, 2, 3]. At high energies, as elegantly summarised in the Standard Model of particle physics, they are at the heart of the microscopical description of the building blocks of our universe, i.e. quarks, leptons and their interactions mediated by gauge bosons [4, 5].

A powerful approach to studying and simulating gauge theories in nonperturbative regimes lies in Lattice Gauge Theories (LGTs), in which the matter and the gauge degrees of freedom are discretized and regularized on a finite lattice. LGTs were originally introduced by Wilson to encode Quantum Chromodynamics (QCD) on a lattice, as a model for quark confinement beyond the perturbative regime [6, 7]. In LGTs, matter and antimatter fermionic fields are defined on lattice sites, whereas the gauge fields live on the links connecting nearest-neighbor sites. This approach has opened the doors to the application of powerful numerical methods, such as Monte Carlo (MC), to the simulations of LGTs on classical computers [8]. In the last decades, MC methods have provided a wide variety of significant results on LGTs in the context of high-energy physics, such as phases diagrams at equilibrium, characterization of the quark-gluon plasma, precise determination of the masses of quarks, mesons, and baryons, hadronic and nucleon form factors, hadronic spectra, and predictions for dark-matter models [9]. Furthermore, MC simulations of LGTs currently represent a powerful numerical tool to predict and interpret data from multiple large-scale high-energy experiments, such as the ones performed at the Large Hadron Collider (LHC).

Despite their impressive success, MC sampling methods are limited in some regimes of parameters of LGTs, such as in the presence of finite baryon chemical potentials, topological θ𝜃\thetaitalic_θ-terms, or for simulating out-of-equilibrium dynamics in real-time. In these cases, the notorious sign problems make the MC numerical approach ineffective and inaccurate [10].

Ranging from these problems, sign-problem-free methods based on Tensor Networks (TNs) have found significant applications in the simulations of LGTs in the last years [11]. TNs were originally introduced as a class of variational wave functions in the field of quantum many-body physics. They provide a compressed representation of physical states based on their entanglement content, capable of efficiently reproducing equilibrium properties, such as phase diagrams, and real-time dynamics of interacting quantum systems [12, 13, 14]. Nowadays, they represent one of the state-of-the-art numerical tools for simulating quantum many-body systems, even with strong correlations.

In the context of high-energy physics, TN methods have proven noteworthy achievements in simulating LTGs in (1+1) dimensions, for both Abelian and non-Abelian gauge groups [15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43]. They have recently found applications to Abelian LGTs up to (3+1) dimensions [44, 45, 46, 47, 48, 49], and also in simulating (2+1)-dimensional non-Abelian SU(2) models [50]. Despite their effectiveness in these first important applications, further and intensive developments are still required to tackle, with TN methods, high-energy physics problems at the center of current research efforts, such as large-scale non-Abelian LGTs and their continuum limits.

In this work, we present a general overview of TN methods for LGTs, and we discuss a possible roadmap in terms of algorithmic development and strategies to improve TN capabilities, toward the ambitious long-term goal of applying TNs to (3+1)-dimensional QCD.

Importantly, TN methods share a common language with quantum computers and simulators. Thus, these developments could also be relevant for encoding, validating, and benchmarking the current and future quantum computations and simulations of LGTs on experimental quantum hardware [51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61].

The paper is organized into sections as follows. In Sec. I, we give a general overview of TN methods, particularly the computational complexity of the most used algorithms for ground state computations and real-time dynamics. In Sec. II, we introduce the main concepts related to LGTs and the main strategies for simulating them with TN methods. In Sec. III, we present a possible roadmap of algorithmic development and optimization strategies crucial for making the TN approach competitive as a complementary method to MC techniques for simulating challenging LGT models. In Sec. IV, we draw our conclusions.

I Tensor Networks Overview

Refer to caption
Figure 1: Examples of tensor network structures: (a) Matrix Product States (MPS); (b) Projected Entangled Pair States (PEPS); (c) Tree Tensor Networks (TTN) for an underlying one-dimensional system; (d) TTN for an underlying two-dimensional square lattice. The physical links with local dimension d𝑑ditalic_d and the virtual links with bond dimension χ𝜒\chiitalic_χ are highlighted in all the figures.

In this section, we present a general overview of TN methods with a particular focus on Tree Tensor Networks (TTNs), the latter being particularly useful for LTGs in higher dimensions [44, 46, 50]. However, this work’s main concepts and ideas can be easily generalized and applied to other TN structures.

Consider a quantum many-body (QMB) system defined on a lattice of N𝑁Nitalic_N sites. The generic site j𝑗jitalic_j is described by a d𝑑ditalic_d-dimensional local Hilbert space jsubscript𝑗\mathcal{H}_{j}caligraphic_H start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, spanned by a local basis of vectors |i1idsubscriptket𝑖1𝑖𝑑{\ket{i}}_{1\leq i\leq d}| start_ARG italic_i end_ARG ⟩ start_POSTSUBSCRIPT 1 ≤ italic_i ≤ italic_d end_POSTSUBSCRIPT. The quantum states of the whole system live on the total Hilbert space =12Ntensor-productsubscript1subscript2subscript𝑁\mathcal{H}=\mathcal{H}_{1}\otimes\mathcal{H}_{2}\otimes\dots\otimes\mathcal{H% }_{N}caligraphic_H = caligraphic_H start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⊗ caligraphic_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⊗ ⋯ ⊗ caligraphic_H start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, that is the tensor product of the local Hilbert space of the lattice sites, with dimension dNsuperscript𝑑𝑁d^{N}italic_d start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. Then, any pure state of the system |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩ can be exactly expanded in terms of a complete basis set of \mathcal{H}caligraphic_H:

|ψ=i1,i2,,iN=1dci1,i2,,iN|i1,i2,,iN,ket𝜓superscriptsubscriptsubscript𝑖1subscript𝑖2subscript𝑖𝑁1𝑑subscript𝑐subscript𝑖1subscript𝑖2subscript𝑖𝑁ketsubscript𝑖1subscript𝑖2subscript𝑖𝑁\ket{\psi}=\sum_{i_{1},i_{2},\dots,i_{N}=1}^{d}c_{i_{1},i_{2},...,i_{N}}\ket{i% _{1},i_{2},...,i_{N}}\,,| start_ARG italic_ψ end_ARG ⟩ = ∑ start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_POSTSUBSCRIPT | start_ARG italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_ARG ⟩ , (1)

where |i1,i2,,iNketsubscript𝑖1subscript𝑖2subscript𝑖𝑁\ket{i_{1},i_{2},...,i_{N}}| start_ARG italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_ARG ⟩ represents the tensor product of the local basis vectors, i.e. |i1|i2||iNtensor-productketsubscript𝑖1ketsubscript𝑖2ketketsubscript𝑖𝑁\ket{i_{1}}\otimes\ket{i_{2}}\otimes\ket{...}\otimes\ket{i_{N}}| start_ARG italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ ⊗ | start_ARG italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ⟩ ⊗ | start_ARG … end_ARG ⟩ ⊗ | start_ARG italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_ARG ⟩. The coefficients of the linear combination ci1,i2,,iNsubscript𝑐subscript𝑖1subscript𝑖2subscript𝑖𝑁c_{i_{1},i_{2},...,i_{N}}italic_c start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_POSTSUBSCRIPT are in general complex scalars; their number scales as dNsuperscript𝑑𝑁d^{N}italic_d start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, i.e. they scale exponentially with the system size N𝑁Nitalic_N.

This is a fundamental limitation when solving a QMB problem on a classical computer since an exponential scaling with the number of degrees of freedom implies that the exact representation described in Eq. 1 is completely unfeasible from a computational and numerical point of view.

Nonetheless, a great variety of natural QMB systems happens to be described by ground and thermal equilibrium states that own little-to-moderate entanglement content. The physical states of these systems, instead of exploring the exponentially large dimension of the Hilbert space, live in a small corner of it, which can be efficiently targeted and parameterized. This property is formally described by the entanglement area law, fulfilled by low-energy states of local Hamiltonians [62]: the entanglement between a partition of the system and the rest is proportional to the area of the boundary between them, instead of its volume, as happens for the majority of states in the Hilbert space. Thus, obeying area law implies that the state contains much fewer quantum correlations than expected for a generic (or random) QMB state. Small corrections to the area law exist for instance close to a transition point for one-dimensional quantum systems (logarithmic corrections). However, the entanglement remains overall moderate [63]. From a theoretical point of view, the entanglement area law has been rigorously proven for (i) one-dimensional gapped local Hamiltonians, where the locality means that a lattice site interacts only with neighboring sites, without two-body all-to-all interactions [64, 65, 66]; (ii) for quantum states at thermal equilibrium, independently from the dimensionality of the system [67]. Even though rigorous proof for QMB systems in higher dimensions is lacking, several numerical and phenomenological shreds of evidence suggest that area law still holds in the presence of local interactions [68, 69, 62, 70, 71].

The area law has important implications on the TN simulation of quantum lattice models: indeed, it is possible to obtain an approximate but efficient representation capable of describing the main properties of these states if the entanglement content is low-to-moderate [62]; the condition applies for example to ground-states and first excited states of local Hamiltonians. TNs give a natural language for this representation, by replacing the complete tensor of rank-N𝑁Nitalic_N ci1,i2,,iNsubscript𝑐subscript𝑖1subscript𝑖2subscript𝑖𝑁c_{i_{1},i_{2},...,i_{N}}italic_c start_POSTSUBSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_i start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT end_POSTSUBSCRIPT of Eq. 1 with a chain of smaller tensors interconnected using auxiliary bond indices. The network keeps a number N𝑁Nitalic_N of physical indices of dimension d𝑑ditalic_d (one for each lattice site), whereas the dimension of the bond indices (called bond dimension) is a control parameter χ𝜒\chiitalic_χ that can be tuned in the numerical simulations and is related to the Schmidt decomposition [72, 18].

The key advantage of passing from the exact representation of Eq. 1 to a TN representation is that the number of parameters in the TN is of the order O(poly(d)poly(N)poly(χ))𝑂poly𝑑poly𝑁poly𝜒O(\mathrm{poly}(d)\mathrm{poly}(N)\mathrm{poly}(\chi))italic_O ( roman_poly ( italic_d ) roman_poly ( italic_N ) roman_poly ( italic_χ ) ), e.g., O(Nχmax(χ,d)2)O(N\chi\max(\chi,d)^{2})italic_O ( italic_N italic_χ roman_max ( italic_χ , italic_d ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) for the TTN. The scaling with the system size is now polynomial and not exponential. In this way, we obtain an efficient representation of the quantum state in terms of computational complexity. It is worth noting that the bond dimension χ𝜒\chiitalic_χ determines the degree of entanglement and quantum correlations encoded in the TN, e.g. for χ=1𝜒1\chi{=}1italic_χ = 1 the TN describes a product state (no entanglement), whereas one recovers the exact but inefficient representation in the limit χdNless-than-or-similar-to𝜒superscript𝑑𝑁\chi{\lesssim}d^{N}italic_χ ≲ italic_d start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT. Tuning χ𝜒\chiitalic_χ properly allows interpolating between these two extreme regimes, efficiently reproducing the entanglement of the quantum state.

The most widely used TN architectures are represented in Fig. 1. Matrix product states (MPS) are an established ansatz for one-dimensional systems, in which the structure of the lattice is reproduced with a network of tensors, one for each lattice site [73]. As shown in Fig. 1(a), each tensor in the bulk of the network has three indices: one physical leg of dimension d𝑑ditalic_d representing the local degrees of freedom, and two virtual legs of dimension χ𝜒\chiitalic_χ connected to the neighboring sites. In open boundary conditions, the tensors at the boundaries have one trivial leg together with one physical and one virtual leg. MPS intrinsically satisfies area law and allows for efficient computation of scalar products between two states and physical observables. Currently, the MPS-based Density Matrix Renormalization Group (DMRG) stands as one of the most consolidated and accurate techniques for numerical simulations of one-dimensional QMB systems as well as quasi-one-dimensional systems such as ladder structures [74]. The computational complexity of this ground state searching algorithm for MPS is of the order O(Ndχ3+Nd2χ2)𝑂𝑁𝑑superscript𝜒3𝑁superscript𝑑2superscript𝜒2O(Nd\chi^{3}+Nd^{2}\chi^{2})italic_O ( italic_N italic_d italic_χ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_N italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) or O(Nd3χ3)𝑂𝑁superscript𝑑3superscript𝜒3O(Nd^{3}\chi^{3})italic_O ( italic_N italic_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_χ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ), depending if the algorithm optimizes a single MPS tensor at a time, or two-tensors simultaneously [74, 75, 76]. Two-site optimization is usually important to avoid getting stuck in local minima or meta-stable configurations during the energy variational minimization [77]. The subspace expansion is an intermediate approach with the benefits of the two-tensor update and a tunable computational cost in between the two approaches [18].

The generalization of the MPS ansatz to two- or higher-dimensional lattices is represented by Projected Entangled Pair State (PEPS) [72]. In this case, each tensor in the bulk has a physical leg of dimension d𝑑ditalic_d, and a number of χ𝜒\chiitalic_χ-dimensional virtual legs depending on the coordination number of the considered lattice. For example, this coordination number is four in the case of a two-dimensional square lattice, as shown in Fig. 1(b). PEPS directly encode in their structure the area law of entanglement, however, their exact contraction is an exponentially hard problem, meaning that PEPS can not be efficiently contracted for numerical computing, e.g. scalar products of states or physical observables [78]. To circumvent this problem, approximate contraction methods have been developed during the last years, and are still at the center of current research efforts [72]. But even with exploiting these approximate techniques, the computational complexity for ground state optimization remains quite high, e.g. of the order of O(Nd2χ8)𝑂𝑁superscript𝑑2superscript𝜒8O(Nd^{2}\chi^{8})italic_O ( italic_N italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_χ start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT ) [79, 80], limiting the maximum reachable bond dimensions (typical values are of the order χ10𝜒10\chi\approx 10italic_χ ≈ 10).

Another important family of TN ansätze is represented by Tree Tensor Networks (TTN), in which the wave function is decomposed into a hierarchical network of tensors that do not contain internal loops [81, 18]. This way, the network can be efficiently contracted and manipulated in polynomial time. A particular class of TTN is represented by binary tree tensor networks, reported in Fig. 1(c)-(d) for one- and two-dimensional lattices. In these structures, tensors in the lowest layer have two physical legs of dimension d𝑑ditalic_d (representing two lattice sites) and a virtual leg of dimension χ𝜒\chiitalic_χ, whereas, in the upper layers, they have three virtual legs of dimensions up to χ𝜒\chiitalic_χ. The network intrinsically encodes a renormalization procedure, in which, at each layer, two sites are mapped into a single effective one. In finite-range models, ground searching algorithms for binary TTN architectures display a numerical complexity of the order O(Nd2χ2+Nχ4)𝑂𝑁superscript𝑑2superscript𝜒2𝑁superscript𝜒4O(Nd^{2}\chi^{2}+N\chi^{4})italic_O ( italic_N italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_N italic_χ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ), see [18, 76]. This is a much more favorable scaling concerning equivalent algorithms for other TN structures, such as PEPS, that allows reaching quite large values of bond dimensions (χ500𝜒500\chi\approx 500italic_χ ≈ 500) [82]. The drawback of loopless structures, such as binary TTN, is that the area law may not be explicitly reproduced in dimensions higher than one [83], which becomes a limiting factor when large systems are addressed. The convergence and the precision of the numerical results obtained via a variational optimization of TTN can be analyzed and kept under control, e.g. by exploiting the large range of available bond dimensions. Furthermore, it is possible to explicitly encode the area law of high dimensional systems in the TTN ansatz by introducing an additional layer of independent disentanglers, acting on different couples of lattice sites and connected to the corresponding physical legs. This process augments the expressive power of TTN, and the resulting ansatz is known as augmented Tree Tensor Network (aTTN) [84]. The computational complexity of variationally optimizing an aTTN structure, which means optimizing both the tensors and the disentanglers, is of the order O(Nχ4d4+Nχd7)𝑂𝑁superscript𝜒4superscript𝑑4𝑁𝜒superscript𝑑7O(N\chi^{4}d^{4}+N\chi d^{7})italic_O ( italic_N italic_χ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + italic_N italic_χ italic_d start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT ). We point out that the scaling of the computational costs with the local dimension d𝑑ditalic_d is particularly severe in the case of aTTN due to the presence of the disentanglers layer.

Besides variational optimization for ground state searching, the previous TN families can also be exploited to simulate the real-time dynamics of local Hamiltonians via at least four methods [85]. One of the most widely used approaches, the Time Evolved Block Decimation (TEBD) algorithm, is based on a Suzuki-Trotter decomposition of the time evolution exponential [86]. The total evolution time is discretized in small time steps. The corresponding evolution operator is computed as products of local terms, such as two-body operators, and repeatedly applied to the TN wave function to generate the time-evolved state. Each application can determine an increase in the bond dimension of the network, so an optimized truncation is needed to maintain an efficient and manageable description of the quantum state. This truncation reduces the bond dimension back to χ𝜒\chiitalic_χ and is performed through a singular value decomposition that minimizes the distance between the evolved and the truncated state. In general, the TEBD method allows the simulation of the real-time dynamics for nearest-neighbor or finite-range Hamiltonians; one time-step with an MPS for a one-dimensional system with local interactions comes with a computational cost that is below a two-tensor sweep for the ground-state search algorithm.

Another method for simulating the evolution of quantum states via TN is the Time-Dependent Variational Principle (TDVP), which does not rely on the Suzuki-Trotter decomposition [87, 88]. In general, TDVP constrains the time evolution to the specific TN manifold considered, such as MPS or TTN, of a given initial bond dimension [89]. This is obtained by projecting the action of the Hamiltonian into the tangent space of the TN manifold and then solving the time-dependent Schrödinger equation within this manifold. This approach automatically preserves the energy and the norm of the quantum states during the time evolution. The TDVP algorithm and the variational ground state search rely both on a set of Krylov vectors and therefore have the same computational scaling for one time step compared to one sweep.

These algorithms represent important and efficient tools for simulating with TN the real-time dynamics of QMB systems. While equilibrium states satisfy the aforementioned area law, out-of-equilibrium time evolution can generate a linear growth of entanglement. In this case, the time-evolved state requires an exponential growth of the bond dimension as a function of the total time [90]. For this reason, TN methods are currently limited to studying the dynamics for low-to-moderate times, or close-to-equilibrium phenomena [91]. In this framework, further developments are extremely important to avoid or at least mitigate this barrier, by devising new algorithms or optimizing existing strategies [92, 93].

II Tensor Networks for Hamiltonian Lattice Gauge Theories

In the traditional MC approach to LGT, the action of a continuum gauge theory is regularized by working on a finite and discrete Euclidean spacetime — i.e., both space and (imaginary) time are discretized [8]. Instead, TN (and quantum) simulations typically rely on the Hamiltonian formalism, where time remains a real, continuous variable while D𝐷Ditalic_D-dimensional space is replaced by a cubic lattice ΛΛ\Lambdaroman_Λ. Simultaneously, Hamiltonian LGTs present features that distinguish them from other lattice models commonly simulated via TN [7]. We now discuss these properties in more detail and outline the main steps that have to be taken to exploit TN algorithms in LGT. We focus on matter-coupled LGTs of the Yang-Mills type, such as those routinely employed in high-energy physics to describe nature’s fundamental interactions; the paradigmatic example is lattice QCD, the SU(3) LGT describing quarks, gluons and their strong interactions.

II.1 LGT building blocks

Refer to caption
Figure 2: Graphical representation of the degrees of freedom of a 2D LGT: fermionic matter fields ψ^𝐱,αsuperscriptsubscript^𝜓𝐱𝛼absent\hat{\psi}_{\mathbf{x,\alpha}}^{\vphantom{dagger}}over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, defined on lattice sites, and gauge fields (the parallel transporter, U^𝐱,𝝁αβsuperscriptsubscript^𝑈𝐱𝝁𝛼𝛽\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha\beta}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT, and the chromo-electric fields, L^𝐱,𝝁νsuperscriptsubscript^𝐿𝐱𝝁𝜈\hat{L}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu}over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT and R^𝐱,𝝁νsuperscriptsubscript^𝑅𝐱𝝁𝜈\hat{R}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu}over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT), living on lattice links. A local gauge transformation at 𝐱𝐱\mathbf{x}bold_x acts on a matter site and all its attached links.

As depicted in Fig. 2, LGTs involve two types of degrees of degrees of freedom, matter and gauge, hosted respectively on lattice sites 𝐱Λ𝐱Λ\mathbf{x}\in\Lambdabold_x ∈ roman_Λ and links (𝐱,𝝁𝐱𝝁\mathbf{x},{\bf\it\mu}bold_x , bold_italic_μ), where 𝝁𝝁{\bf\it\mu}bold_italic_μ denotes one of the lattice basis vectors.

Matter fields — such as the electron or quark fields. As for these examples, we assume matter consists of Dirac fermions and employ staggered fermions [94] to tame the fermion doubling problem, i.e., the proliferation of propagating fermionic degrees of freedom on the lattice. Then, matter is represented by a multiplet ψ^𝐱,αsuperscriptsubscript^𝜓𝐱𝛼absent\hat{\psi}_{\mathbf{x,\alpha}}^{\vphantom{dagger}}over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT of anti-commuting fields, {ψ^𝐱,α,ψ^𝐲,β}=δαβδ𝐱𝐲superscriptsubscript^𝜓𝐱𝛼absentsuperscriptsubscript^𝜓𝐲𝛽subscript𝛿𝛼𝛽subscript𝛿𝐱𝐲\{\hat{\psi}_{\mathbf{x,\alpha}}^{\vphantom{dagger}},\hat{\psi}_{\mathbf{y,% \beta}}^{\dagger\,}\}=\delta_{\alpha\beta}\delta_{\mathbf{xy}}{ over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_y , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT } = italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT bold_xy end_POSTSUBSCRIPT. Assuming a single matter flavor in the fundamental representation of the gauge group 𝒢𝒢\mathcal{G}caligraphic_G, gauge transformations are generated by Q𝐱ν=α,βψ^𝐱,αλαβνψ^𝐱,βsubscriptsuperscript𝑄𝜈𝐱subscript𝛼𝛽superscriptsubscript^𝜓𝐱𝛼subscriptsuperscript𝜆𝜈𝛼𝛽superscriptsubscript^𝜓𝐱𝛽absentQ^{\nu}_{\mathbf{x}}=\sum_{\alpha,\beta}\hat{\psi}_{\mathbf{x,\alpha}}^{% \dagger\,}\lambda^{\nu}_{\alpha\beta}\hat{\psi}_{\mathbf{x,\beta}}^{\vphantom{% dagger}}italic_Q start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, where {λν}ν=1dim𝒢superscriptsubscriptsuperscript𝜆𝜈𝜈1dimension𝒢\{\lambda^{\nu}\}_{\nu=1}^{\dim\mathcal{G}}{ italic_λ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_ν = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_dim caligraphic_G end_POSTSUPERSCRIPT are the hermitian generators of 𝒢𝒢\mathcal{G}caligraphic_G — e.g., the electric charge for U(1) or the three spin matrices for SU(2). The generalization to multiple matter flavors is straightforward.

Gauge fields — such as the photon or gluon fields. They are represented by bosonic operators, namely the parallel transporter U^𝐱,𝝁αβsuperscriptsubscript^𝑈𝐱𝝁𝛼𝛽\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha\beta}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT and the generators of its left and right gauge transformations, L^𝐱,𝝁νsuperscriptsubscript^𝐿𝐱𝝁𝜈\hat{L}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu}over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT and R^𝐱,𝝁νsuperscriptsubscript^𝑅𝐱𝝁𝜈\hat{R}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu}over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT,

[L^𝐱,𝝁νmissing,U^𝐲,𝝁αβmissing]commutatorsuperscriptsubscript^𝐿𝐱𝝁𝜈missingsuperscriptsubscript^𝑈𝐲𝝁𝛼𝛽missing\displaystyle\commutator*{\hat{L}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}% \nu}missing}{\hat{U}_{\mathbf{y,{\bf\it\mu^{\smash{\prime}}}}}^{\vphantom{% dagger}\alpha\beta}missing}[ start_ARG over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT roman_missing end_ARG , start_ARG over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_y , start_ID bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT roman_missing end_ARG ] =δ𝐱𝐲δ𝝁𝝁γλαγνU^𝐱,𝝁γβ,absentsubscript𝛿𝐱𝐲subscript𝛿𝝁𝝁subscript𝛾subscriptsuperscript𝜆𝜈𝛼𝛾superscriptsubscript^𝑈𝐱𝝁𝛾𝛽\displaystyle=-\delta_{\mathbf{xy}}\delta_{{\bf\it\mu\mu^{\smash{\prime}}}}% \sum\nolimits_{\gamma}\lambda^{\nu}_{\alpha\gamma}\hat{U}_{\mathbf{x,{\bf\it% \mu}}}^{\vphantom{dagger}\gamma\beta}\,,= - italic_δ start_POSTSUBSCRIPT bold_xy end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT bold_italic_μ bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α italic_γ end_POSTSUBSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ italic_β end_POSTSUPERSCRIPT , (2a)
[R^𝐱,𝝁νmissing,U^𝐲,𝝁αβmissing]commutatorsuperscriptsubscript^𝑅𝐱𝝁𝜈missingsuperscriptsubscript^𝑈𝐲𝝁𝛼𝛽missing\displaystyle\commutator*{\hat{R}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}% \nu}missing}{\hat{U}_{\mathbf{y,{\bf\it\mu^{\smash{\prime}}}}}^{\vphantom{% dagger}\alpha\beta}missing}[ start_ARG over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT roman_missing end_ARG , start_ARG over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_y , start_ID bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT roman_missing end_ARG ] =+δ𝐱𝐲δ𝝁𝝁γU^𝐱,𝝁αγλγβν;absentsubscript𝛿𝐱𝐲subscript𝛿𝝁𝝁subscript𝛾superscriptsubscript^𝑈𝐱𝝁𝛼𝛾subscriptsuperscript𝜆𝜈𝛾𝛽\displaystyle=+\delta_{\mathbf{xy}}\delta_{{\bf\it\mu\mu^{\smash{\prime}}}}% \sum\nolimits_{\gamma}\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}% \alpha\gamma}\lambda^{\nu}_{\gamma\beta}\,;= + italic_δ start_POSTSUBSCRIPT bold_xy end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT bold_italic_μ bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_γ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_γ italic_β end_POSTSUBSCRIPT ; (2b)

In the Abelian U(1)U1\mathrm{U}(1)roman_U ( 1 ) case there is only one generator: the electric field, E^𝐱,𝝁=L^𝐱,𝝁=R^𝐱,𝝁superscriptsubscript^𝐸𝐱𝝁absentsuperscriptsubscript^𝐿𝐱𝝁absentsuperscriptsubscript^𝑅𝐱𝝁absent\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}}\mathop{=}\hat{L}_{\mathbf% {x,{\bf\it\mu}}}^{\vphantom{dagger}}\mathop{=}\hat{R}_{\mathbf{x,{\bf\it\mu}}}% ^{\vphantom{dagger}}over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, for which parallel transporters act as raising operators, [E^𝐱,𝝁missing,U^𝐱,𝝁]=U^𝐱,𝝁commutatorsuperscriptsubscript^𝐸𝐱𝝁absentmissingsuperscriptsubscript^𝑈𝐱𝝁absentsuperscriptsubscript^𝑈𝐱𝝁absent\commutator*{\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}}missing}{\hat% {U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}}}\mathop{=}\hat{U}_{\mathbf{x,% {\bf\it\mu}}}^{\vphantom{dagger}}[ start_ARG over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT roman_missing end_ARG , start_ARG over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT end_ARG ] = over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT.

In terms of the above ingredients, a possible discretization of a matter coupled Yang-Mills theory is defined by the Kogut-Susskind Hamiltonian [7]

HLGT=subscript𝐻LGTabsent\displaystyle H_{\rm LGT}=italic_H start_POSTSUBSCRIPT roman_LGT end_POSTSUBSCRIPT = c2a𝐱,𝝁α,β(s𝐱,𝝁ψ^𝐱,αU^𝐱,𝝁αβψ^𝐱+𝝁,β+h.c.)𝑐Planck-constant-over-2-pi2𝑎subscript𝐱𝝁subscript𝛼𝛽subscript𝑠𝐱𝝁superscriptsubscript^𝜓𝐱𝛼superscriptsubscript^𝑈𝐱𝝁𝛼𝛽superscriptsubscript^𝜓𝐱𝝁𝛽absenth.c.\displaystyle-\frac{c\hbar}{2a}\sum_{\mathbf{x},{\bf\it\mu}}\sum_{\alpha,\beta% }\quantity(s_{\mathbf{x},{\bf\it\mu}}\hat{\psi}_{\mathbf{x,\alpha}}^{\dagger\,% }\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha\beta}\hat{\psi}_{% \mathbf{x+{\bf\it\mu},\beta}}^{\vphantom{dagger}}+\text{h.c.})- divide start_ARG italic_c roman_ℏ end_ARG start_ARG 2 italic_a end_ARG ∑ start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( start_ARG italic_s start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x + bold_italic_μ , italic_β end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + h.c. end_ARG )
+mc2𝐱αs𝐱ψ^𝐱,αψ^𝐱,α𝑚superscript𝑐2subscript𝐱subscript𝛼subscript𝑠𝐱superscriptsubscript^𝜓𝐱𝛼superscriptsubscript^𝜓𝐱𝛼absent\displaystyle+mc^{2}\sum_{\mathbf{x}}\sum_{\alpha}s_{\mathbf{x}}\hat{\psi}_{% \mathbf{x,\alpha}}^{\dagger\,}\hat{\psi}_{\mathbf{x,\alpha}}^{\vphantom{dagger}}+ italic_m italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x , italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT
+cg22aD2𝐱,𝝁E^𝐱,𝝁2𝑐Planck-constant-over-2-pisuperscript𝑔22superscript𝑎𝐷2subscript𝐱𝝁superscriptsubscript^𝐸𝐱𝝁2\displaystyle+\frac{c\hbar g^{2}}{2a^{D-2}}\sum_{\mathbf{x},{\bf\it\mu}}\hat{E% }_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}+ divide start_ARG italic_c roman_ℏ italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_a start_POSTSUPERSCRIPT italic_D - 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
c2g2a4DTr(U^+U^),𝑐Planck-constant-over-2-pi2superscript𝑔2superscript𝑎4𝐷subscriptTrsuperscriptsubscript^𝑈absentsuperscriptsubscript^𝑈\displaystyle-\frac{c\hbar}{2g^{2}a^{4-D}}\sum_{\square}\mathrm{Tr}(\hat{U}_{% \mathbf{\square}}^{\vphantom{dagger}}+\hat{U}_{\mathbf{\square}}^{\dagger\,})\,,- divide start_ARG italic_c roman_ℏ end_ARG start_ARG 2 italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT 4 - italic_D end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT roman_Tr ( over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) , (3)

where c𝑐citalic_c is the speed of light, Planck-constant-over-2-pi\hbarroman_ℏ is the Plank constant, a𝑎aitalic_a is the lattice spacing, g𝑔gitalic_g is the gauge coupling, and m𝑚mitalic_m is the mass parameter. The first HLGTsubscript𝐻LGTH_{\rm LGT}italic_H start_POSTSUBSCRIPT roman_LGT end_POSTSUBSCRIPT term describes matter hop** between neighboring lattice sites, mediated by the gauge field; the second term is the mass-energy; s𝐱,𝝁subscript𝑠𝐱𝝁s_{\mathbf{x},{\bf\it\mu}}italic_s start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT and s𝐱subscript𝑠𝐱s_{\mathbf{x}}italic_s start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT are phases that arise due to the use of staggered fermions. The last two terms represent the (chromo)electric and (chromo)magnetic energy of the gauge field, respectively. The electric energy density is given by the Casimir operator:

E^𝐱,𝝁2=ν(L^𝐱,𝝁ν)2=ν(R^𝐱,𝝁ν)2.superscriptsubscript^𝐸𝐱𝝁2subscript𝜈superscriptsuperscriptsubscript^𝐿𝐱𝝁𝜈2subscript𝜈superscriptsuperscriptsubscript^𝑅𝐱𝝁𝜈2\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}=\sum_{\nu}(\hat{L}_{% \mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu})^{2}=\sum_{\nu}(\hat{R}_{% \mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu})^{2}\,.over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (4)

The magnetic energy density is a plaquette term that, on a cubic lattice, corresponds to a four-body interaction:

U^=α,β,γ,δU^𝐱,𝝁αβU^𝐱+𝝁,𝝁βγU^𝐱+𝝁,𝝁γδU^𝐱,𝝁δαsuperscriptsubscript^𝑈absentsubscript𝛼𝛽𝛾𝛿superscriptsubscript^𝑈𝐱𝝁𝛼𝛽superscriptsubscript^𝑈𝐱𝝁𝝁𝛽𝛾superscriptsubscript^𝑈𝐱𝝁𝝁absent𝛾𝛿superscriptsubscript^𝑈𝐱𝝁absent𝛿𝛼\hat{U}_{\mathbf{\square}}^{\vphantom{dagger}}=\;\sum_{\mathclap{\alpha,\beta,% \gamma,\delta}}\;\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha% \beta}\hat{U}_{\mathbf{x+{\bf\it\mu},{\bf\it\mu^{\prime}}}}^{\vphantom{dagger}% \beta\gamma}\hat{U}_{\mathbf{x+{\bf\it\mu^{\prime}}\!,{\bf\it\mu}}}^{\dagger\,% \gamma\delta}\hat{U}_{\mathbf{x,{\bf\it\mu^{\prime}}}}^{\dagger\,\delta\alpha}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_α , italic_β , italic_γ , italic_δ end_POSTSUBSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x + bold_italic_μ , start_ID bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β italic_γ end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x + start_ID bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_ID , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † italic_γ italic_δ end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , start_ID bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † italic_δ italic_α end_POSTSUPERSCRIPT (5)

where 𝝁𝝁{\bf\it\mu}bold_italic_μ and 𝝁𝝁{\bf\it\mu^{\prime}}bold_italic_μ start_POSTSUPERSCRIPT bold_′ end_POSTSUPERSCRIPT span the plaquette’s plane. Plaquette terms only exist in D>1𝐷1D>1italic_D > 1, contributing to the increased complexity of quantum and TN simulations of LGTs in higher dimensions [56].

II.2 Gauge field truncation

The link Hilbert space is the space of square-integrable functions on the gauge group, L2(𝒢)superscript𝐿2𝒢L^{2}(\mathcal{G})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_G ), which is infinite-dimensional for a continuous 𝒢𝒢\mathcal{G}caligraphic_G [95].

In one space dimension, gauge degrees of freedom are unphysical (absence of transverse polarizations) and can thus be integrated out, albeit at the price of introducing non-local interactions [96].

Beyond one dimension, the removal is much more delicate, because it requires first decoupling the gauge field’s longitudinal component [97]. When impossible or inconvenient to remove, gauge degrees of freedom might have to be truncated to perform TN or quantum simulation. Among known truncation recipes are Quantum Link Models (QLM) [98, 99, 100, 101], which have been already adopted for quantum simulation of LGTs [102, 52, 103, 104, 105, 56, 60, 58, 106], finite subgroups [26, 30, 107], digitization of gauge fields [108], and fusion-algebra deformation [109].

Another adopted solution is truncating the spectrum of the electric energy density operator E^𝐱,𝝁2missingΘnormsuperscriptsubscript^𝐸𝐱𝝁2missingΘ\norm{\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}missing}\leq\Theta∥ start_ARG over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_missing end_ARG ∥ ≤ roman_Θ on each link [50, 110]. The cutoff is conveniently imposed in the irreducible representation (irrep) basis {|jmn}ket𝑗𝑚𝑛\{\ket{jmn}\}{ | start_ARG italic_j italic_m italic_n end_ARG ⟩ } [95] of L2(𝒢)superscript𝐿2𝒢L^{2}(\mathcal{G})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_G ), where E^𝐱,𝝁2superscriptsubscript^𝐸𝐱𝝁2\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is diagonal:

E^𝐱,𝝁2|jmn=C2(j)|jmn.superscriptsubscript^𝐸𝐱𝝁2ket𝑗𝑚𝑛subscript𝐶2𝑗ket𝑗𝑚𝑛\hat{E}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}\ket{jmn}=C_{2}(j)\ket{% jmn}\,.over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | start_ARG italic_j italic_m italic_n end_ARG ⟩ = italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) | start_ARG italic_j italic_m italic_n end_ARG ⟩ . (6)

Here m𝑚mitalic_m and n𝑛nitalic_n are indices in the j𝑗jitalic_j-irrep of 𝒢𝒢\mathcal{G}caligraphic_G and C2(j)subscript𝐶2𝑗C_{2}(j)italic_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_j ) is the quadratic Casimir of j𝑗jitalic_j [95]. In the strong coupling limit, where the electric energy term dominates HLGTsubscript𝐻LGTH_{\rm LGT}italic_H start_POSTSUBSCRIPT roman_LGT end_POSTSUBSCRIPT, this truncation is equivalent to an energy cutoff [110].

II.3 Gauss law and the dressed site

The most distinctive feature of gauge theories is arguably the presence of local constraints, analogous to the Gauss law of classical electrodynamics, relating the configuration of the gauge field to the spatial distribution of charges [111]. At the quantum level, Gauss law is the statement that only gauge invariant states are physical, namely, G^𝐱ν|Ψphys=0𝐱,νsuperscriptsubscript^𝐺𝐱𝜈ketsubscriptΨphys0for-all𝐱𝜈\hat{G}_{\mathbf{x}}^{\vphantom{dagger}\nu}\ket*{\Psi_{\text{phys}}}=0\;% \forall\mathbf{x},\nuover^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT phys end_POSTSUBSCRIPT end_ARG ⟩ = 0 ∀ bold_x , italic_ν, where G^𝐱νsuperscriptsubscript^𝐺𝐱𝜈\hat{G}_{\mathbf{x}}^{\vphantom{dagger}\nu}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT are the generators of local gauge transformations at 𝐱𝐱\mathbf{x}bold_x:

G^𝐱ν=Q^𝐱ν+q^𝐱ν+𝝁[L^𝐱,𝝁ν+R^𝐱𝝁,𝝁ν],superscriptsubscript^𝐺𝐱𝜈superscriptsubscript^𝑄𝐱𝜈superscriptsubscript^𝑞𝐱𝜈subscript𝝁delimited-[]superscriptsubscript^𝐿𝐱𝝁𝜈superscriptsubscript^𝑅𝐱𝝁𝝁𝜈\hat{G}_{\mathbf{x}}^{\vphantom{dagger}\nu}=\hat{Q}_{\mathbf{x}}^{\vphantom{% dagger}\nu}+\hat{q}_{\mathbf{x}}^{\vphantom{dagger}\nu}+\sum_{{\bf\it\mu}}[% \hat{L}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\nu}+\hat{R}_{\mathbf{x-{% \bf\it\mu},{\bf\it\mu}}}^{\vphantom{dagger}\nu}]\,,over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT = over^ start_ARG italic_Q end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT + over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT bold_italic_μ end_POSTSUBSCRIPT [ over^ start_ARG italic_L end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT + over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT bold_x - bold_italic_μ , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT ] , (7)

with q^𝐱νsuperscriptsubscript^𝑞𝐱𝜈\hat{q}_{\mathbf{x}}^{\vphantom{dagger}\nu}over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT representing eventual static background charges (typically vanishing). On the lattice, Gauss law provides a set of vertex constraints, each involving a lattice site and its 2D2𝐷2D2 italic_D neighboring links.

Due to Gauss law, the physical Hilbert space of LGTs is much smaller than the tensor product of all local sites and link Hilbert spaces. Properly exploiting gauge symmetries can thus significantly speed up numerical simulations [16]. Strategies that solve Gauss law by eliminating (partially or entirely) either the gauge fields or the matter fields have been developed. Nonetheless, such approaches come with specific limitations: the range of interaction has to be extended, moreover, integrating-out gauge fields become problematic in D>1𝐷1D>1italic_D > 1 [97], while the recipe for removing matter is a model (matter content) dependent [112].

Another possibility is to enforce Gauss law using a dressed site construction, sketched in Fig. 3 and outlined below. Dressed sites have local dimensions typically larger than those resulting from the aforementioned approaches, but they are obtained from a model-independent prescription which has the advantage of preserving the locality of the interactions [16, 110].

Refer to caption
Figure 3: Pictorial representation of the dressed site formalism adopted for TN simulations of LGTs: (a) starting from the original formulation of matter and gauge fields, (b) each truncated gauge link is split into two representations, one per half-link; each is equipped with a proper fermionic rishon mode ζ𝜁\zetaitalic_ζ. (c) All the half-links are absorbed into the attached matter site, forming a gauge-invariant dressed site (d) whose Hilbert space spans all the possible gauge singlets.

Rishons.

As a first step, every link is decomposed into a pair of left (L𝐿Litalic_L) and right (R𝑅Ritalic_R) rishon degrees of freedom, each associated with a Hilbert space spanned by the basis states |jmket𝑗𝑚\ket{jm}| start_ARG italic_j italic_m end_ARG ⟩, identifying |jmn|jmL|jnRket𝑗𝑚𝑛subscriptket𝑗𝑚𝐿tensor-productsubscriptket𝑗𝑛𝑅\ket{jmn}\hookrightarrow\ket{jm}_{L}\mathop{\otimes}\ket{jn}_{R}| start_ARG italic_j italic_m italic_n end_ARG ⟩ ↪ | start_ARG italic_j italic_m end_ARG ⟩ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ⊗ | start_ARG italic_j italic_n end_ARG ⟩ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT, and writing parallel transporters as rishon bilinears [50]:

U^𝐱,𝝁αβkζ𝐱,+𝝁L(k)αζ𝐱+𝝁,𝝁R(k)β.superscriptsubscript^𝑈𝐱𝝁𝛼𝛽subscript𝑘subscriptsuperscript𝜁𝐿𝑘𝛼𝐱𝝁subscriptsuperscript𝜁𝑅𝑘𝛽𝐱𝝁𝝁\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha\beta}\to\sum_{k}% \zeta^{L(k)\alpha}_{\mathbf{x},+{\bf\it\mu}}\zeta^{R(k)\beta\,\dagger}_{% \mathbf{x}+{\bf\it\mu},-{\bf\it\mu}}\,\,.over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT → ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_ζ start_POSTSUPERSCRIPT italic_L ( italic_k ) italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_x , + bold_italic_μ end_POSTSUBSCRIPT italic_ζ start_POSTSUPERSCRIPT italic_R ( italic_k ) italic_β † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_x + bold_italic_μ , - bold_italic_μ end_POSTSUBSCRIPT . (8)

Abelian link symmetries.

Physical gauge link configurations, i.e. those with the left and right rishons in the same irrep, are selected introducing a local link symmetry at the TN simulation level [16, 110]. Notice this is always Abelian, regardless of the Abelian or non-Abelian nature of the gauge group.

Gauge invariant computational basis.

Crucially, the gauge generators G^𝐱νsuperscriptsubscript^𝐺𝐱𝜈\hat{G}_{\mathbf{x}}^{\vphantom{dagger}\nu}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT now involve only the matter site at 𝐱𝐱\mathbf{x}bold_x and its 2D2𝐷2D2 italic_D neighboring rishons. Fusing these degrees of freedom in a composite site, Gauss law becomes an internal constraint that singles out the dressed site Hilbert space as their gauge invariant subspace:

dress=kerGmatt(rish)2D.subscriptdresskernel𝐺tensor-productsubscriptmattsuperscriptsubscriptrishtensor-productabsent2𝐷\mathcal{H}_{\text{dress}}=\ker G\subset\mathcal{H}_{\text{matt}}\otimes(% \mathcal{H}_{\text{rish}})^{\otimes 2D}\,.caligraphic_H start_POSTSUBSCRIPT dress end_POSTSUBSCRIPT = roman_ker italic_G ⊂ caligraphic_H start_POSTSUBSCRIPT matt end_POSTSUBSCRIPT ⊗ ( caligraphic_H start_POSTSUBSCRIPT rish end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT ⊗ 2 italic_D end_POSTSUPERSCRIPT . (9)

The expansion of the gauge singlet basis states of dresssubscriptdress\mathcal{H}_{\text{dress}}caligraphic_H start_POSTSUBSCRIPT dress end_POSTSUBSCRIPT in terms of the matter and rishon bases is computed via Clebsch-Gordan decomposition [110].

Defermionization.

It is possible to effectively eliminate fermionic matter, and treat dressed sites as large spins, for any gauge theory where the gauge field has a well-defined parity. Specifically, a local parity operator P𝐱,𝝁=P𝐱,𝝁subscript𝑃𝐱𝝁superscriptsubscript𝑃𝐱𝝁P_{\mathbf{x},{\bf\it\mu}}=P_{\mathbf{x},{\bf\it\mu}}^{\dagger}italic_P start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT such that P𝐱,𝝁2=1superscriptsubscript𝑃𝐱𝝁21P_{\mathbf{x},{\bf\it\mu}}^{2}=1italic_P start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 must satisfy {U^𝐱,𝝁αβ,P𝐱,𝝁}=0superscriptsubscript^𝑈𝐱𝝁𝛼𝛽subscript𝑃𝐱𝝁0\{\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}\alpha\beta},P_{\mathbf{x% },{\bf\it\mu}}\}=0{ over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT , italic_P start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT } = 0, as it happens for 2Nsubscript2𝑁\mathbb{Z}_{2N}roman_ℤ start_POSTSUBSCRIPT 2 italic_N end_POSTSUBSCRIPT, U(N), and SU(2N) [113, 112, 50]. In this case, it is possible to take fermionic rishons, and as a result, all physical (gauge invariant) dressed site operators are genuinely local, i.e. they mutually commute at a nonzero distance (as spins or bosons) [114]. In particular, this applies to the Hamiltonian HLGTsubscript𝐻LGTH_{\rm LGT}italic_H start_POSTSUBSCRIPT roman_LGT end_POSTSUBSCRIPT, making gauge-defermionization particularly convenient for higher-dimensional systems, where Jordan-Wigner strings result in long-range interactions [115].

II.4 Scaling of the local basis dimension

Table 1 lists the dressed site dimension d=dimdress𝑑dimensionsubscriptdressd=\dim\mathcal{H}_{\text{dress}}italic_d = roman_dim caligraphic_H start_POSTSUBSCRIPT dress end_POSTSUBSCRIPT associated with the first few nontrivial gauge truncations of three representative LGT models in D=2,3𝐷23D=2,3italic_D = 2 , 3 space dimensions. All the considered LGTs include dynamical matter, represented by one fermionic field multiplet in the fundamental representation of the gauge group. The \ellroman_ℓ-th row of Table 1 is obtained kee** only the first \ellroman_ℓ nonzero electric energy levels (i.e., using the \ellroman_ℓ-th smallest quadratic Casimir eigenvalue as cutoff ΘΘ\Thetaroman_Θ, see Sec. II.2). As Table 1 shows, the local dimension increases rapidly with \ellroman_ℓ. For 3-dimensional non-Abelian LGTs, dO(104)similar-to𝑑𝑂superscript104d\sim O(10^{4})italic_d ∼ italic_O ( 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) is reached already within the first two truncations, making the study of the untruncated limit prohibitive.

Differently from the models typically encountered in condensed matter physics, the local dimension of LGTs can thus be a limiting factor for TN simulation — especially when d𝑑ditalic_d becomes comparable to commonly used TN bond dimensions (100χ500less-than-or-similar-to100𝜒less-than-or-similar-to500100\lesssim\chi\lesssim 500100 ≲ italic_χ ≲ 500 for TTN). In these cases, strategies aimed at further compressing the local computational basis are needed (see Sec. III.1). As just discussed, such truncation strategies are particularly crucial for high-dimensional LGTs.

Several numerical analyses suggest that, in some cases, a small-to-moderate truncation of the gauge group is enough for accurately approximating the continuum limits, at least for low-energy states [24, 26, 32, 116, 117, 118]. However, the optimal gauge truncation depends on the Hamiltonian parameters m𝑚mitalic_m and g𝑔gitalic_g.

\ellroman_ℓ d𝑑ditalic_d
(2+1)21(2+1)( 2 + 1 )-dimensions (3+1)31(3+1)( 3 + 1 )-dimensions
U(1) SU(2) SU(3) U(1) SU(2) SU(3)
1 35 30 164 267 178 3096
2 165 168 752 3437 3670 52476
3 455 600 3738 18487 35280 813438
4 969 1650 19878 64953 214958 17490134
5 1771 3822 43698 177155 967466 69232482
6 2925 7840 82128 408421 3509062 228461186
7 4495 14688 212496 835311 10828494 1245755754
8 6545 25650 333538 1561841 29473038 2782999996
Table 1: Dressed site Hilbert space dimension d𝑑ditalic_d for increasing number \ellroman_ℓ of allowed electric energy density levels in some 2- and 3-dimensional paradigmatic LGTs with dynamical matter and gauge groups U(1), SU(2), and SU(3).

As an example, we consider the (2+1)D U(1) LGT including dynamical matter (QED), whose Hamiltonian can be obtained from Sec. II.1:

HQEDsubscript𝐻QED\displaystyle H_{\text{QED}}italic_H start_POSTSUBSCRIPT QED end_POSTSUBSCRIPT =c2a𝐱[-iψ^𝐱U^𝐱,𝝁𝒙ψ^𝐱+𝝁𝒙\displaystyle=\frac{c\hbar}{2a}\sum_{\mathbf{x}}\Big{[}\text{-i}\hat{\psi}_{% \mathbf{x}}^{\dagger\,}\hat{U}_{\mathbf{x,{\bf\it\mu_{x}}}}^{\vphantom{dagger}% }\hat{\psi}_{\mathbf{x+{\bf\it\mu_{x}}}}^{\vphantom{dagger}}= divide start_ARG italic_c roman_ℏ end_ARG start_ARG 2 italic_a end_ARG ∑ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT [ -i over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , start_ID bold_italic_μ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x + start_ID bold_italic_μ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT (10)
(1)j𝐱+j𝐲ψ^𝐱U^𝐱,𝝁𝒚ψ^𝐱+𝝁𝒚+H.c.]\displaystyle-(-1)^{j_{\mathbf{x}}+j_{\mathbf{y}}}\hat{\psi}_{\mathbf{x}}^{% \dagger\,}\hat{U}_{\mathbf{x,{\bf\it\mu_{y}}}}^{\vphantom{dagger}}\hat{\psi}_{% \mathbf{x+{\bf\it\mu_{y}}}}^{\vphantom{dagger}}+\text{H.c.}\Big{]}- ( - 1 ) start_POSTSUPERSCRIPT italic_j start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT bold_y end_POSTSUBSCRIPT end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , start_ID bold_italic_μ start_POSTSUBSCRIPT bold_italic_y end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x + start_ID bold_italic_μ start_POSTSUBSCRIPT bold_italic_y end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + H.c. ]
+mc2𝐱(1)j𝐱+j𝐲ψ^𝐱ψ^𝐱𝑚superscript𝑐2subscript𝐱superscript1subscript𝑗𝐱subscript𝑗𝐲superscriptsubscript^𝜓𝐱superscriptsubscript^𝜓𝐱absent\displaystyle+mc^{2}\sum_{\mathbf{x}}(-1)^{j_{\mathbf{x}}+j_{\mathbf{y}}}\hat{% \psi}_{\mathbf{x}}^{\dagger\,}\hat{\psi}_{\mathbf{x}}^{\vphantom{dagger}}+ italic_m italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT ( - 1 ) start_POSTSUPERSCRIPT italic_j start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT + italic_j start_POSTSUBSCRIPT bold_y end_POSTSUBSCRIPT end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT
+g2c2a𝐱,𝝁E^𝐱,𝝁2c2g2aTr(U^+U^),superscript𝑔2𝑐Planck-constant-over-2-pi2𝑎subscript𝐱𝝁superscriptsubscript^𝐸𝐱𝝁2𝑐Planck-constant-over-2-pi2superscript𝑔2𝑎subscriptTrsuperscriptsubscript^𝑈absentsuperscriptsubscript^𝑈\displaystyle+\frac{g^{2}c\hbar}{2a}\sum_{\mathbf{x},{\bf\it\mu}}\hat{E}_{% \mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}2}-\frac{c\hbar}{2g^{2}a}\sum_{% \square}\mathrm{Tr}(\hat{U}_{\mathbf{\square}}^{\vphantom{dagger}}+\hat{U}_{% \mathbf{\square}}^{\dagger\,}),+ divide start_ARG italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_c roman_ℏ end_ARG start_ARG 2 italic_a end_ARG ∑ start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT over^ start_ARG italic_E end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG italic_c roman_ℏ end_ARG start_ARG 2 italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a end_ARG ∑ start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT roman_Tr ( over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) ,

Since the Abelian U(1) group has only one generator, the gauge fields E𝐱subscript𝐸𝐱E_{\mathbf{x}}italic_E start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT and U^𝐱,𝝁superscriptsubscript^𝑈𝐱𝝁absent\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT can be represented as

E𝐱,𝝁subscript𝐸𝐱𝝁\displaystyle E_{\mathbf{x},{\bf\it\mu}}italic_E start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT =σ𝐱,𝝁z()absentsubscriptsuperscript𝜎𝑧𝐱𝝁\displaystyle=\sigma^{z}_{\mathbf{x},{\bf\it\mu}}(\ell)= italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT ( roman_ℓ ) U^𝐱,𝝁superscriptsubscript^𝑈𝐱𝝁absent\displaystyle\hat{U}_{\mathbf{x,{\bf\it\mu}}}^{\vphantom{dagger}}over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT bold_x , bold_italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT =ζ𝐱,+𝝁()ζ𝐱+𝝁,𝝁()absentsubscript𝜁𝐱𝝁subscript𝜁𝐱𝝁𝝁\displaystyle=\zeta_{\mathbf{x},+{\bf\it\mu}}(\ell)\cdot\zeta_{\mathbf{x}+{\bf% \it\mu},-{\bf\it\mu}}(\ell)= italic_ζ start_POSTSUBSCRIPT bold_x , + bold_italic_μ end_POSTSUBSCRIPT ( roman_ℓ ) ⋅ italic_ζ start_POSTSUBSCRIPT bold_x + bold_italic_μ , - bold_italic_μ end_POSTSUBSCRIPT ( roman_ℓ ) (11)

where σz()superscript𝜎𝑧\sigma^{z}(\ell)italic_σ start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT ( roman_ℓ ) is the spin z𝑧zitalic_z-operator in the \ellroman_ℓ SU(2) irrep, while ζ()𝜁\zeta(\ell)italic_ζ ( roman_ℓ ) is a ×\ell\times\ellroman_ℓ × roman_ℓ ladder operator with Fermi statistics.

We focus on a single plaquette in open boundary conditions, as it provides the minimal setting allowing for both electric and magnetic effects. Then, to characterize the convergence in the gauge truncation \ellroman_ℓ, we consider a candidate observable O𝑂Oitalic_O and compute its ground state expectation value O^subscriptdelimited-⟨⟩^𝑂\langle\hat{O}\rangle_{\ell}⟨ over^ start_ARG italic_O end_ARG ⟩ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT for increasing \ellroman_ℓ. We iterate until the relative deviation between consecutive truncations drops below some threshold ϵtruncsubscriptitalic-ϵtrunc\epsilon_{\text{trunc}}italic_ϵ start_POSTSUBSCRIPT trunc end_POSTSUBSCRIPT: |O^O^1|<ϵtrunc|O^|subscriptdelimited-⟨⟩^𝑂superscriptsubscriptdelimited-⟨⟩^𝑂superscript1subscriptitalic-ϵtruncsubscriptdelimited-⟨⟩^𝑂superscript|\langle\hat{O}\rangle_{\ell^{*}}-\langle\hat{O}\rangle_{\ell^{*}-1}|<\epsilon% _{\text{trunc}}|\langle\hat{O}\rangle_{\ell^{*}}|| ⟨ over^ start_ARG italic_O end_ARG ⟩ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT - ⟨ over^ start_ARG italic_O end_ARG ⟩ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - 1 end_POSTSUBSCRIPT | < italic_ϵ start_POSTSUBSCRIPT trunc end_POSTSUBSCRIPT | ⟨ over^ start_ARG italic_O end_ARG ⟩ start_POSTSUBSCRIPT roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | for some superscript\ell^{*}roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. Fig. 4(a) shows the minimal truncation superscript\ell^{*}roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT at which the magnetic energy operator O^=ReU^^𝑂superscriptsubscript^𝑈absent\hat{O}=\real\hat{U}_{\mathbf{\square}}^{\vphantom{dagger}}over^ start_ARG italic_O end_ARG = start_OPERATOR roman_Re end_OPERATOR over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT is converged to ϵtrunc=105subscriptitalic-ϵtruncsuperscript105\epsilon_{\text{trunc}}=10^{-5}italic_ϵ start_POSTSUBSCRIPT trunc end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. We explore a grid of model parameters, whose extent has been chosen according to standard MC literature [119, 120, 121, 122, 8, 123, 124, 125, 126, 127, 128, 129, 130]. ReU^superscriptsubscript^𝑈absent\real\hat{U}_{\mathbf{\square}}^{\vphantom{dagger}}start_OPERATOR roman_Re end_OPERATOR over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT is used as a benchmark due to its relevance in the weak coupling regime, where the continuum limit of D<3𝐷3D<3italic_D < 3 lattice QED is located [8]. As Fig. 4(b) shows, superscript\ell^{*}roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT depends heavily on the coupling, growing asymptotically like g1similar-tosuperscriptsuperscript𝑔1\ell^{*}\sim g^{-1}roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ∼ italic_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as g𝑔gitalic_g is decreased, while m𝑚mitalic_m plays almost no role. An analogous inverse dependence of the minimal gauge truncation on the coupling is expected for non-Abelian LGT in arbitrary dimensions. Moreover, the continuum limit of non-Abelian LGTs in D3𝐷3D\leq 3italic_D ≤ 3 is also expected at g0𝑔0g\to 0italic_g → 0 [8], further substantiating the need to compress the local dimension in TN simulations of LGT whenever extrapolation to the continuum is in order.

Apart from the growth of the local dimension, extrapolation to the continuum is further complicated by the fact that the continuum limit of a lattice quantum field theory corresponds to a critical point of the underlying lattice model [131]. Close to criticality, quantum correlations are boosted and violations of the entanglement area law are expected [79]. The higher entanglement entropy in the proximity of the continuum (g,m1much-less-than𝑔𝑚1g,m\ll 1italic_g , italic_m ≪ 1 regime) is already captured by the single-plaquette analysis of Fig. 4(c). Continuum limits of LGTs are thus an area of potential advantage for quantum computation over classical methods, as the former is not limited by entanglement. Nonetheless, quantum computation is also affected by the need to relax gauge truncations when g0𝑔0g\to 0italic_g → 0, either by increasing the number of qubits used to encode a dressed site, which is at least log2(d)subscript2𝑑\lceil\log_{2}(d)\rceil⌈ roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_d ) ⌉, or by using hybrid devices, that have both qubits and bosonic modes [132].

Refer to caption
Figure 4: Exact diagonalization of a QED plaquette for a grid of masses and couplings, m[102,101]𝑚superscript102superscript101m\in[10^{-2},10^{1}]italic_m ∈ [ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ] and g2[101,101]superscript𝑔2superscript101superscript101g^{2}\in[10^{-1},10^{1}]italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∈ [ 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ]. (a,b) Minimal gauge truncation superscript\ell^{*}roman_ℓ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT required to reach a precision ϵtrunc=105subscriptitalic-ϵtruncsuperscript105\epsilon_{\text{trunc}}=10^{-5}italic_ϵ start_POSTSUBSCRIPT trunc end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT in the magnetic energy ReU^missingexpectation-valuesuperscriptsubscript^𝑈absentmissing\expectationvalue*{\real\hat{U}_{\mathbf{\square}}^{\vphantom{dagger}}missing}⟨ start_ARG start_OPERATOR roman_Re end_OPERATOR over^ start_ARG italic_U end_ARG start_POSTSUBSCRIPT □ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT end_POSTSUPERSCRIPT roman_missing end_ARG ⟩. (c) Corresponding entanglement entropy S𝑆Sitalic_S associated with a symmetric bipartition of the plaquette.

III Roadmap for advanced LGT simulations via tensor networks

As detailed in the previous sections, LGT models present some peculiar features that make TN simulations particularly challenging, especially for large system sizes and for studying the continuum limits in terms of gauge field truncation, lattice spacing, and volume.

State-of-the-art techniques, such as TTN algorithms, have been applied for simulating ground state properties of QED in (2+1)- and (3+1)-dimensions for small-to-intermediate sizes [44, 46]. Very recently, they have been also applied to the SU(2) Yang-Mills model in (2+1)-dimensions [50]. In all these simulations, small non-trivial representations of the gauge groups have been exploited, e.g. three electric field levels for QED and the first two irreducible representations of SU(2) for the Yang-Mills model.

Nowadays, lattice computations with MC-based techniques are performed on large lattices, e.g. of the order 643×128superscript64312864^{3}\times 12864 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT × 128 sites (space and time discretization), and with no truncation of the gauge fields. These large sizes are required to control finite-volume effects and to perform extrapolations toward the continuum limits [133]. In the last decades, the impressive progress in algorithmic development, high-performance optimizations, and the availability of increasingly powerful supercomputer facilities have played a major role in the advancement of MC-based LGT computations. Indeed, this progress has opened the doors to large-scale simulations, that currently represent the standard approach for studying non-perturbative phenomena in quantum field theory.

However, MC techniques are generally based on computations of path integrals in which the integrand functions are overall positive. Many physically relevant scenarios, such as finite baryon density regimes or real-time dynamics of quarks, give rise to a change in the sign of the integrands and highly oscillating behaviors. Thus, numerical evaluations suffer from the near cancellation of the opposite-sign contributions to the integrals. This is the essence of the infamous sign problem, a long-standing issue of LGT simulations with MC methods [10].

Hence the quest of conceiving, develo**, and optimizing alternative approaches that enable simulating these regimes, being the latter at the heart of many open problems related to our understanding of high-energy physics.

As described in the previous sections, TNs represent a promising complementary method, which found the first applications in simulating non-trivial instances of high-dimensional LGTs on small-to-intermediate lattice sizes. TNs are intrinsically sign-problem-free, enabling the simulation of both static properties at equilibrium, such as low-energy states, and real-time dynamics, even in the presence of finite chemical potentials or non-trivial topological terms. It is worth noting that, in addition to local observables and correlation functions, TNs allow the numerical computations of entanglement properties, such as entanglement entropies and central charges, that could potentially shield new light on LGT phenomena [32].

Nevertheless, TN simulations of high-dimensional LGTs still represent a challenging problem, especially for large lattice sizes or higher representations of gauge groups needed for analyzing continuum limits. In this framework, further and intensive developments are required to tackle TNs’ current problems related to LGTs, such as QCD’s non-perturbative effects on lattices of sizes comparable with MC simulations. In this regard, we note that sign-problem-free TN ansätze can also be used in combination with variational MC methods to tackle high-dimensional lattice gauge theories with arbitrary gauge groups [134].

In the following, we present a possible roadmap in terms of algorithmic development and optimization strategies that we foresee to be crucial for making the TN approach competitive as a complementary method to MC techniques. Therein, the first two sections, i.e. Secs. III.1 and III.2, can be approached with existing TN algorithms and a good intuition on LGT problems; then, Secs. III.3 and III.4 outline optimization for existing algorithms to leverage HPC systems; finally, we discuss new classes of ansätze to tackle finite temperature problems in Sec. III.5. In some parts, we focus on TTNs, but the vast majority of the presented concepts and techniques can be straightforwardly applied to other TN ansätze.

III.1 Local basis truncation

Refer to caption
Figure 5: Graphical representation of (a) the pseudo site DMRG (PS-DMRG) approach, in which a single site having large local dimension d𝑑ditalic_d is replaced with npssubscript𝑛𝑝𝑠n_{ps}italic_n start_POSTSUBSCRIPT italic_p italic_s end_POSTSUBSCRIPT pseudo sites with a smaller local dimension, e.g two; (b) the local basis optimization embedded in the MPS ansatz (LBO-MPS), so that an additional layer of tensors connected to the physical legs of the MPS performs the reduction from a large local basis with dimension d𝑑ditalic_d into an effective basis with smaller dimension dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

One of the main issues in simulating LGTs with TN methods lies in the very large local basis dimension d𝑑ditalic_d that one needs to handle to represent matter and gauge-field degrees of freedom properly. To some extent, this situation is very similar to some condensed matter models, e.g. the Holstein model, involving lattice fermions coupled with phonons [135], or bosonic systems coupled to optical cavities [136, 137]. Also in these cases, the local Hilbert space is in principle infinite-dimensional, and a truncation to a fixed cutoff is needed for performing TN simulations. For one-dimensional systems of this type, some efficient algorithms based on MPS have been developed in the last years [138], e.g. the pseudosite DMRG (PS-DMRG) method [139], and the DMRG with local basis optimization (DMRG-LBO) [140, 141, 142].

The key idea of the PS-DMRG is to replace a single site having large local dimension d𝑑ditalic_d with npslog2(d)subscript𝑛𝑝𝑠subscriptlog2𝑑n_{ps}\approx\mathrm{log}_{\rm\scriptscriptstyle 2}(d)italic_n start_POSTSUBSCRIPT italic_p italic_s end_POSTSUBSCRIPT ≈ roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_d ) pseudo sites of local dimension 2, as shown Fig. 5(a). Since a large class of optimization algorithms for TNs scale at least quadratically with the local dimension but linearly with the total number of sites, as detailed in Sec. I, one obtains a more efficient and manageable representation according to this procedure.

The price to pay lies in the range of the interactions: short-range interactions in the Hamiltonian are transformed into long-range operators due to the pseudo-site encoding of the local degrees of freedom. As a consequence, PS-DMRG should require a larger bond dimension and a larger number of variational steps to converge, and the benefits offered by the pseudo-sites could progressively fade for increasing values of npssubscript𝑛𝑝𝑠n_{ps}italic_n start_POSTSUBSCRIPT italic_p italic_s end_POSTSUBSCRIPT at fixed bond dimension. PS-DMRG has been applied to one-dimensional QMB systems with d𝑑ditalic_d up to O(100)𝑂100O(100)italic_O ( 100 ) [138].

The core of the DMRG-LBO algorithm, instead, lies in a local basis optimization protocol which enables a controlled and efficient truncation of the local Hilbert space [143]. For each site 𝐱𝐱\mathbf{x^{\prime}}bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT of the lattice, the optimized local basis is computed by starting from its reduced density matrix, i.e.

ρ𝐱=Tr𝐱𝐱|ψψ|,subscript𝜌𝐱subscriptTr𝐱𝐱𝜓𝜓\rho_{\mathbf{x^{\prime}}}=\mathrm{Tr}_{\mathbf{x}\neq\mathbf{x^{\prime}}}% \outerproduct{\psi}{\psi},italic_ρ start_POSTSUBSCRIPT bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = roman_Tr start_POSTSUBSCRIPT bold_x ≠ start_ID bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ID end_POSTSUBSCRIPT | start_ARG italic_ψ end_ARG ⟩ ⟨ start_ARG italic_ψ end_ARG | , (12)

where |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩ is a general state of the whole system and the trace is over all the degrees of freedom which do not involve the site 𝐱𝐱\mathbf{x^{\prime}}bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. By performing the diagonalization procedure, the eigenvalues λαsubscript𝜆𝛼\lambda_{\alpha}italic_λ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT and the eigenvectors vαsubscript𝑣𝛼v_{\alpha}italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT of ρ𝐱subscript𝜌𝐱\rho_{\mathbf{x^{\prime}}}italic_ρ start_POSTSUBSCRIPT bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT can be easily determined. The set of values λαsubscript𝜆𝛼\lambda_{\alpha}italic_λ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT represent the probabilities associated with the states vαsubscript𝑣𝛼v_{\alpha}italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. If λαsubscript𝜆𝛼\lambda_{\alpha}italic_λ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is small, e.g. below a certain numerical threshold, the related eigenvector vαsubscript𝑣𝛼v_{\alpha}italic_v start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT can be discarded from the local basis of the site 𝐱𝐱\mathbf{x^{\prime}}bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, with a controllable loss of accuracy for the state |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩. Thus, for reducing the local basis dimension of 𝐱𝐱\mathbf{x^{\prime}}bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT from d𝑑ditalic_d to a smaller value dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, an optimal choice is kee** the dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT eigenvectors of ρ𝐱subscript𝜌𝐱\rho_{\mathbf{x^{\prime}}}italic_ρ start_POSTSUBSCRIPT bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT with the largest probabilities. Then, the original state |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩ can be projected on the new basis, without losing the relevant physical information. Let us notice that the site 𝐱𝐱\mathbf{x^{\prime}}bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT can also be a generic unit cell of the system, composed of a certain number of lattice sites.

A crucial point of the LBO procedure is the knowledge of the original state |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩, generally the ground state of the system, that is not known prior. This issue can be overcome in several ways: by performing exact diagonalization of the system’s Hamiltonian with local dimension d𝑑ditalic_d for small lattice sizes, to determine |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩, and then truncating the basis from d𝑑ditalic_d to dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT for increasing values of the cutoff dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. From this procedure, we can obtain an optimized local basis ensuring a controlled approximation of the original ground state. This optimized basis can then be exploited in optimization algorithms for simulating larger lattice sizes, such as DMRG. Another strategy directly incorporates the LBO procedure in the tensor network ansatz, as shown Fig. 5(b): by inserting an additional tensor on each physical leg of the MPS, the large local basis with dimension d𝑑ditalic_d is transformed into an effective basis with smaller dimension dsuperscript𝑑d^{\prime}italic_d start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. In this way, the MPS tensors only see the effective basis in the optimization procedures, with a significant reduction of the computational costs described in Sec. I. This method has been used for both static DMRG and time evolution algorithms, such as TEBD, for the study of one-dimensional quantum impurity models and correlated electron-phonon systems [140, 141, 144].

In the context of LGTs, the dimension of the local basis d𝑑ditalic_d can easily go beyond values of the order of 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT, especially for higher-dimensional non-Abelian models and large representations of the gauge groups, as shown in Sec. II. In this scenario, numerical simulations with TNs are practically infeasible without an optimized scheme for truncating the local degrees of freedom. Techniques like PS-DMRG might offer some benefits for small system sizes, such as unveiling the most relevant degrees of freedom in the low-energy states, but it is difficult to scale them up for large sizes due to the long-range interactions induced between the pseudo-sites. Also, the local constraints imposed by Gauss law would become highly non-local when splitting a single site into multiple pseudo-sites representing the matter and link fields.

In principle, LBO-based procedures could instead represent a well-grounded route for addressing the problem. Their use in condensed matter, e.g. bosonic systems, is a rather consolidated approach, whereas their application to TN simulations of LGTs currently is an uncharted but promising territory. The main steps that we foresee as needed and important in this direction are the following:

  • (i)

    Employing exact diagonalization, testing the convergence of LBO procedures for one-dimensional systems, such as the φ4superscript𝜑4\varphi^{4}italic_φ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT-theory or the Schwinger model, for which analytical solutions, at least in some regimes of the phase diagram, and numerical results are widely available, also in the limit of no gauge field truncations. In this way, we can obtain valuable information about the scaling of the basis cutoffs concerning the final accuracy of the state representation for small system sizes.

  • (ii)

    Performing the same analysis on (2+1)-dimensional LGTs, such as QED or SU(N)𝑆𝑈𝑁SU(N)italic_S italic_U ( italic_N ) models, for one unit cell, like a single lattice-plaquette, to systematically study the effects of the magnetic interactions on the local degrees of freedom. Indeed, the QLM representation of LGTs, detailed in Sec. II, generally exploits the “electric field” basis, in which the electric field terms of the Hamiltonian and Gauss law are diagonal. In this scheme, the magnetic interactions correspond to non-diagonal operators which can increase the number of local electric states to include for an accurate description of the system, especially for small values of the coupling constant g𝑔gitalic_g of Sec. II.1, as highlighted in the numerical analysis in Sec. II.4.

  • (iii)

    Exploiting the optimized local bases obtained from exact diagonalization as input of TN simulations for larger sizes, testing the effects on the global ground state accuracy and in computing physically relevant quantities, such as the mass gap [130].

  • (iv)

    In the same spirit of LBO-MPS, implementing LBO protocols directly in TN ansätze that are more suitable for simulating high-dimensional LGT models, such as TTN. This step could be of great benefit in particular for aTTNs, which encode the area law for the entanglement but are severely limited by large local bases (see Sec. I).

By following and combining all these steps, we expect to reduce the effective local basis of LGT models, potentially enabling TN simulations for large representations of Abelian and non-Abelian gauge groups.

It is worth noting that constructing optimal bases for numerical and quantum computation of LGTs is an active area of research. Several approaches that have been recently proposed involve performing canonical transformations of the gauge degrees of freedom before truncation [107, 145, 146]. By exploiting a resource-efficient protocol of this type, Ref. [107] has shown that the number of local states required to reach a 1%percent11\%1 % accuracy level when computing the expectation value of the plaquette operators in two-dimensional (2+1)-dimensional QED can be reduced by more than 94%percent9494\%94 % compared to the unoptimized truncation. Integrating these approaches into TN algorithms could greatly benefit LGT simulations.

III.2 Tailored initial states

In TN algorithms for ground state searching, the optimization procedure generally starts from a random TN initial state |ψinitketsubscript𝜓init\ket*{\psi_{\mathrm{init}}}| start_ARG italic_ψ start_POSTSUBSCRIPT roman_init end_POSTSUBSCRIPT end_ARG ⟩, i.e. the tensors in the network are filled with random coefficients at the beginning. This strategy usually guarantees that the probability of overlap** with the true ground state |ψinit|ψgs|2superscriptinner-productsubscript𝜓initsubscript𝜓gs2|\innerproduct*{\psi_{\mathrm{init}}}{\psi_{\mathrm{gs}}}|^{2}| ⟨ start_ARG italic_ψ start_POSTSUBSCRIPT roman_init end_POSTSUBSCRIPT end_ARG | start_ARG italic_ψ start_POSTSUBSCRIPT roman_gs end_POSTSUBSCRIPT end_ARG ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is not vanishing. To reach a small error in the final energy, this procedure typically requires from 10 to 50 optimization sweeps for LGTs simulations, depending on the specific models, the Hamiltonian parameters, and the lattice size. Since the time for completing a sweep can be very long, especially for large bond dimensions, strategies for reducing the number of needed sweeps could be beneficial for scaling up system sizes. From this perspective, constructing appropriate states to be used as initial guesses can speed up the convergence, similar to the choice of the trial wave function for variational Monte Carlo simulations. We consider the following options:

  • (i)

    Physical insight: Initial states can be constructed by following physical intuition, at least in those regimes in which analytical or partial numerical results are available. For instance, initial guesses can be constructed by starting from the TN ground states numerically obtained for a lower representation to simulate large spin representations of the gauge fields.

  • (ii)

    Machine learning: Machine learning-assisted protocols can improve the construction of tailored initial states in the different regimes of the model parameters. For instance, feed-forward neural networks have been proposed as trial wave functions for quantum Monte Carlo simulations [147], and machine learning techniques have been used to feed TN simulations [148]. Similarly, neural networks might reveal great potential in constructing initial states for LGTs to be used in large-scale TN simulations, in which reducing the number of sweeps is a key point for feasibility.

  • (iii)

    Tensor network results: Following the idea of the physical insight, it is also possible to feed neighboring ground states as initial guesses into a ground state search. This option exists especially when scanning a phase diagram and varying parameters in a small increment such that the overlap between neighboring wave functions is sufficient; this overlap decreases for two points on the opposite sides of a quantum critical point. The same idea can be implemented by preparing an initial guess quenching from an easily accessible ground state to the target parameters; the quench itself does not have to be adiabatic or free of numerical errors, but must only have sufficient overlap with the ground state. The advantage here is that one quench can generate multiple initial guesses along the quench for different parameters.

III.3 Leverage HPC techniques for local optimization

We dedicate the two following sections to the numerical optimization of the TN algorithms. In this section, we give an overview of the topic and we discuss possible strategies to improve the optimization. The more technical steps are discussed in the following Sec. III.4.

To scale up TN simulations of LGTs, in particular regarding lattice sizes, another important factor is the number of optimization steps to be carried out. The number of optimization steps scales linearly with the number of sweeps as well as with the system size for MPS, PEPS, and TTN. The choice of the number of sweeps is set such that the algorithm reaches convergence when computing ground or low energy states. Let us briefly describe the general procedure for ground state searches: we focus on the TTN optimization, however, the main points described here can be applied to other TN structures, such as MPS or PEPS. For a complete and technical description of the algorithms and implementation details, see Ref. [18].

Consider a generic QMB Hamiltonian H𝐻Hitalic_H and a generic normalized state |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩, defined on the same Hilbert space. To numerically determine the ground state of H𝐻Hitalic_H, the following global minimization problem has to be solved:

min|ψ{E(|ψ)}=min|ψψ|H|ψ.subscriptket𝜓𝐸ket𝜓subscriptket𝜓bra𝜓𝐻ket𝜓\min_{\ket{\psi}}\left\{E({\ket{\psi}})\right\}=\min_{\ket{\psi}}\bra{\psi}H% \ket{\psi}\,.roman_min start_POSTSUBSCRIPT | start_ARG italic_ψ end_ARG ⟩ end_POSTSUBSCRIPT { italic_E ( | start_ARG italic_ψ end_ARG ⟩ ) } = roman_min start_POSTSUBSCRIPT | start_ARG italic_ψ end_ARG ⟩ end_POSTSUBSCRIPT ⟨ start_ARG italic_ψ end_ARG | italic_H | start_ARG italic_ψ end_ARG ⟩ . (13)

If |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩ is written in terms of the TTN ansatz introduced in Sec. I, the variational parameters are the coefficients in the TTN. In this case, the global optimization problem of Eq. 13 is broken down into a sequence of smaller optimizations, each of which involves only a minimal subset of tensors in the TTN. The algorithm solves the optimization via an eigenproblem searching for the eigenvector with the lowest eigenvalue. Without any loss of generality, one single tensor at a time is optimized in the simplest case, as shown in Fig. 6(a). In detail, the energy is computed by contracting the Hamiltonian between the TTN and its complex conjugate. Then, we start the optimization procedure from a target tensor T𝑇Titalic_T, by computing its environment, i.e. the network without the tensors T𝑇Titalic_T and Tsuperscript𝑇T^{\dagger}italic_T start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, which represents the effective Hamiltonian Heffsubscript𝐻effH_{\text{eff}}italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT for the local problem. At this stage, an eigenproblem of Heffsubscript𝐻effH_{\text{eff}}italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT is solved, and the tensor T𝑇Titalic_T is updated with the newly found ground state. The whole procedure is sequentially iterated for all the tensors in the network, performing an optimization sweep.

For each operation, efficient algorithms from linear algebra are typically used, e.g. the Arnoldi algorithm implemented in the ARPACK library [149, 150]. We recall that the numerical complexity of this procedure for a single TTN-tensor is O(d2χ2+χ4)𝑂superscript𝑑2superscript𝜒2superscript𝜒4O(d^{2}\chi^{2}+\chi^{4})italic_O ( italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_χ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) [76]. Therefore, a single optimization can be time-consuming when χ𝜒\chiitalic_χ is very large, e.g. χ1000𝜒1000\chi\approx 1000italic_χ ≈ 1000, as for simulating high-dimensional LGTs. We point out the established and promising future parallelization schemes for the single tensor optimization:

  • (i)

    opemMP: An efficient openMP implementation of the contraction between the effective operators with the tensor can speed up simulations. Moreover, the Arnoldi algorithm of ARPACK is optimized for large-scale linear algebra operations and supports intra-node multi-core parallelization based on openMP [151]; thus, ARPACK does not become a bottleneck in the openMP implementation. Nonetheless, many simulations remain expensive in computation time even with 64 or more cores available in HPC facilities; therefore, we consider more approaches beyond the well-established openMP path.

  • (ii)

    Accelerators: Graphics Processing Units (GPU) and Tensor Processing Units (TPU) offer a path to accelerate linear algebra routines, where both have demonstrated their usefulness: GPUs have reported speedups of up to a factor of 10 due to the efficient tensor manipulations [152, 153, 154, 155]; tensor processing units have shown great potential in large-scale simulations of several quantum systems, e.g. drastically reducing the computational time of DMRG calculations with very large bond dimensions from months to hours [156, 157, 158, 159]. Tensor processing units are application-specific integrated circuits originally introduced for machine learning; we consider the integration and tuning of TPUs therefore as a step after the successful integration of GPUs. While single GPUs can solve TTN-problems up to a bond dimension of χ<1000𝜒1000\chi<1000italic_χ < 1000, multi-GPU support is available for libraries; HPC systems typically provide hardware with four GPUs per node.

  • (iii)

    Multi-node approaches to local tensor optimizations: Both CPU and GPU algorithms can be further scaled by using multiple nodes. The underlying linear algebra routines of the local eigenvalue problem are parallelizable via libraries such as ScaLAPACK or MAGMA. Both libraries provide routines for distributed memory machines [160]; MAGMA supports moreover CPU and GPUs. In this way, the workload of the eigenproblem procedure can be split into several computation nodes. Then, it is important to analyze the performances as a function of the bond dimension χ𝜒\chiitalic_χ, to test the effectiveness of this approach against the latency of the inter-node communications.

  • (iv)

    Tuning of parameters and algorithms: accelerators developed for machine learning applications have excellent support for lower and real precision. Tuning parameters over the different sweeps is beneficial, e.g. increasing the precision towards the end of the sweep. This approach profits from faster single-precision implementations during the first sweeps. Selecting algorithms like random singular value decompositions can also bring benefits [161].

III.4 Sweeps and HPC parallelization

So far, we have parallelized single tensor optimizations within a sweep, but the sweep itself was sequential, i.e. serial. Recent works formulated parallel versions of MPS algorithms for ground state search and time evolution, e.g. via the Message Passing Interface (MPI) [162, 163, 164]. The main difference between the serial and parallel algorithms is the effective operators used in the optimization. In the serial version, the effective operators contain the information of the most recent version of all other tensors. This update is delayed in the parallel version, i.e. the tensor that entered the effective operator is not necessarily the one of the current sweep but can be the version of an earlier sweep.

Refer to caption
Figure 6: Effective operators and parallel tensor networks. (a) Procedure for optimizing a TTN to find the ground state of a QMB system: the energy is computed by contracting the Hamiltonian H𝐻Hitalic_H (yellow tensor) with the TTN, representing the state |ψket𝜓\ket{\psi}| start_ARG italic_ψ end_ARG ⟩, and its hermitian conjugate, representing ψ|bra𝜓\bra{\psi}⟨ start_ARG italic_ψ end_ARG |. The variational optimization starts from a target tensor T𝑇Titalic_T (red tensor), by computing its effective Hamiltonian Heffsubscript𝐻effH_{\text{eff}}italic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT and then solving the local eigenvalue problem for the latter. The tensor T𝑇Titalic_T is then updated with the newly found ground state, and the procedure iterates over all the tensors in the network (sweep). (b) The workload itself consists of optimizing each tensor held by the MPI thread t𝑡titalic_t, which requires effective operators calculated by other MPI threads. We dub delays ΔisubscriptΔ𝑖\Delta_{i}roman_Δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT the number of optimization cycles needed to obtain the information of tensors in the i-th MPI thread in another MPI thread via MPI communication. Matrix product states naturally split into sub-chains which communicate with one or two neighboring MPI threads to obtain updated effective operators. Delays for updates scale with the distance between two MPI threads along the chain. Each MPI thread can use threading or openMP, e.g. in a hybrid openMP-MPI approach. (c) Similarly, TTNs can be split into sub-trees for each MPI thread allowing for optimizing the sub-tree without communication with other MPI threads. Delays due to updating scale logarithmically as any distance in a tree network.

If the delay becomes an obstacle to convergence, there is the option to modify parameters during the sweeps. Typically, ten to fifty sweeps are necessary to converge to a solution. As the initial state is usually random, MPI can be used especially at the beginning. To ensure convergence, one can consider serial steps at the end; even a gradual reduction of the MPI processes as the sweeps proceed is possible and gradually reduces the delays.

Considering the MPS scenario of a chain in Fig. 6(b), we split the chain into equal parts of (N/nMPIthreads)𝑁subscript𝑛MPIthreads(N/n_{\rm\scriptscriptstyle MPI-threads})( italic_N / italic_n start_POSTSUBSCRIPT roman_MPI - roman_threads end_POSTSUBSCRIPT ) sites. Each part of the chain communicates with its two neighbors apart from the two boundaries. The effective operators take into account the tensors of the same MPI process with zero delay as in the serial case. The tensors of the i-th neighboring MPI process have a delay of i𝑖iitalic_i. The worst-case scenario of the delay scales linearly with the number of MPI processes. The delay can be avoided by communicating the effective operators after each update through the chain, which is a quasi-serial step with no more than two MPI processes active at the same time.

The problem becomes more complicated for the TTNs suggested for LGT, but we expect a benefit for the parallelization of a TTN versus an MPS for higher dimensional systems. Fig. 6(c) shows an example of how each MPI process gets assigned a sub-tree within the complete TTN. Unlike the MPS, the number of neighboring MPI processes for communication is at least three and increases with (N/nMPIthreads)𝑁subscript𝑛MPIthreads(N/n_{\rm\scriptscriptstyle MPI-threads})( italic_N / italic_n start_POSTSUBSCRIPT roman_MPI - roman_threads end_POSTSUBSCRIPT ) tensors per thread. For simplicity, we assume equally shaped subtrees for all MPI processes. The delay of the tensor update is now 2log2(N)2subscriptlog2𝑁2\cdot\mathrm{log}_{\rm\scriptscriptstyle 2}(N)2 ⋅ roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_N ) in the worst-case scenario.

One-dimensional systems with nearest-neighbor interactions thus exhibit a delay of 1111 in the worst-case scenario in the MPS, while the delay is up to 2log2(N)2subscriptlog2𝑁2\cdot\mathrm{log}_{\rm\scriptscriptstyle 2}(N)2 ⋅ roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_N ) for the nearest neighbors in the center of the TTN. Rather, higher dimensional systems change this aspect, e.g., for an N×N𝑁𝑁N\times Nitalic_N × italic_N two-dimensional system mapped to 1D via a zig-zag map**. The MPS has a worst-case delay of the tensor update of N𝑁Nitalic_N for the slow index. In contrast, the TTN has the same log behavior and a maximum delay at the center of the TTN as 2log2(N2)2subscriptlog2superscript𝑁22\cdot\mathrm{log}_{\rm\scriptscriptstyle 2}(N^{2})2 ⋅ roman_log start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Thus, the worst-case delay is equal for 16×16161616\times 1616 × 16 systems; increasing N𝑁Nitalic_N further, TTNs exhibit smaller maximum delays during parallel sweeps. Moreover, the TTN is unaffected by the type of map** used; in contrast, the worst-case delay for the MPS grows to 2N2𝑁2N2 italic_N for the snake map** and to at least N2/2superscript𝑁22N^{2}/2italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 for the Hilbert curve [165]. Equal arguments hold for 3D systems and delays of N2superscript𝑁2N^{2}italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (MPS, zig-zag) versus 2log2(N3)2losubscriptg2superscript𝑁32\mathrm{log}_{\rm\scriptscriptstyle 2}(N^{3})2 roman_l roman_o roman_g start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_N start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) (TTN, any map**).

To get an intuition of what parallelized simulations can solve, we sketch out the specifications for a parallel simulation on the pre-exascale cluster Leonardo hosted by Cineca. We choose an MPI approach together with the GPUs. Leonardo has 3456 nodes with four GPUs totaling 13824 GPUs available for the complete cluster. Bond dimensions on the order of χ=450𝜒450\chi=450italic_χ = 450 consume 54GB of memory without effective operators (assuming double complex precision, 40 Lanczos vectors) and allow to solve the eigenproblem on the GPU without temporarily storing data on the CPU. We use the single-tensor per MPI-thread with χ=450𝜒450\chi=450italic_χ = 450 as a baseline where we extract a rough empirical estimate with Leonardo; in detail, we use a 2D quantum Ising model with 2subscript2\mathbb{Z}_{2}roman_ℤ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT symmetry in the vicinity of the quantum critical point and the initial tensor optimizations [166, 167]. Due to the delay of the tensor in the effective operators, the minimum number of sweeps must be beyond 24. Then, we consider the scaling of the TTN previously introduced and generate Table 2 with an overview of different system sizes and bond dimensions. These results provide a coarse-grained estimate, since plaquette terms, different symmetries, entanglement generation, and optimization time within later sweeps could further impact the computational time.

Our estimate predicts that a system of 256×256256256256\times 256256 × 256 sites takes about two months for bond dimension χ=450𝜒450\chi=450italic_χ = 450 on 1024 GPU nodes of Cineca’s leonardo. Future improvements are likely to bring this simulation time down, e.g., the next-generation GPUs in comparison to the A100 or further optimization in data movement. In contrast, a three-dimensional system with large entanglement and many sites requires three to four orders of magnitude improvement, where cluster size and other improvements have to come together to reach this challenge. Furthermore, Table 2 provides an estimate of the boundary for a potential quantum advantage in simulating lattice gauge theories with quantum computers or simulators.

System size χ𝜒\chiitalic_χ Factor Estimated walltime
64×64646464\times 6464 × 64 450450450450 Tbasesubscript𝑇baseT_{\rm\scriptscriptstyle base}italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 4.16days4.16days4.16~{}\mathrm{days}4.16 roman_days
64×64646464\times 6464 × 64 900900900900 16Tbase16subscript𝑇base16\cdot T_{\rm\scriptscriptstyle base}16 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 66.6days66.6days66.6~{}\mathrm{days}66.6 roman_days
256×256256256256\times 256256 × 256 450450450450 28Tbase28subscript𝑇base28\cdot T_{\rm\scriptscriptstyle base}28 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 116.5days116.5days116.5~{}\mathrm{days}116.5 roman_days
256×256256256256\times 256256 × 256 900900900900 448Tbase448subscript𝑇base448\cdot T_{\rm\scriptscriptstyle base}448 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 5.1years5.1years5.1~{}\mathrm{years}5.1 roman_years
16×16×1616161616\times 16\times 1616 × 16 × 16 450450450450 4Tbase4subscript𝑇base4\cdot T_{\rm\scriptscriptstyle base}4 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 16.6days16.6days16.6~{}\mathrm{days}16.6 roman_days
16×16×1616161616\times 16\times 1616 × 16 × 16 900900900900 64Tbase64subscript𝑇base64\cdot T_{\rm\scriptscriptstyle base}64 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 266days266days266~{}\mathrm{days}266 roman_days
64×64×6464646464\times 64\times 6464 × 64 × 64 450450450450 1984Tbase1984subscript𝑇base1984\cdot T_{\rm\scriptscriptstyle base}1984 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 23years23years23~{}\mathrm{years}23 roman_years
64×64×6464646464\times 64\times 6464 × 64 × 64 900900900900 31744Tbase31744subscript𝑇base31744\cdot T_{\rm\scriptscriptstyle base}31744 ⋅ italic_T start_POSTSUBSCRIPT roman_base end_POSTSUBSCRIPT 362years362years362~{}\mathrm{years}362 roman_years
Table 2: Estimated simulation time. We derive the baseline from a single-tensor optimization of a 64×64646464\times 6464 × 64 quantum Ising simulation with 2subscript2\mathbb{Z}_{2}roman_ℤ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT symmetry taking 7192s7192s7192\mathrm{s}7192 roman_s on a A100 GPU. Further, we assume that single-tensor update, one tensor and one GPU per MPI thread, and 50 sweeps for the baseline. To extrapolate to larger systems, we assume a scaling with 𝒪(χ4ND1)𝒪superscript𝜒4superscript𝑁𝐷1\mathcal{O}(\chi^{4}N^{D-1})caligraphic_O ( italic_χ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_N start_POSTSUPERSCRIPT italic_D - 1 end_POSTSUPERSCRIPT ) as well as seven (thirty-one) tensors per MPI thread for 256×256256256256\times 256256 × 256 (64×64×6464646464\times 64\times 6464 × 64 × 64) systems. The empirical scalings are approximately a factor of 2.32.32.32.3 for doubling the system size and 13131313 for doubling the bond dimension, which we obtain from smaller simulations with χ=225𝜒225\chi=225italic_χ = 225 and for 32×32323232\times 3232 × 32 qubits. The times are valid for any d<χ𝑑𝜒d<\chiitalic_d < italic_χ.

III.5 Finite temperature regime

To date, TN simulations of high-dimensional LGTs including dynamical matter are exploring zero temperature regimes, which are important to understand the low-energy properties of the models. To explore finite temperature phenomena, particularly relevant for open research problems such as the QCD phase diagram, technical and numerical challenges have to be tackled, e.g. devising and testing efficient TN algorithms for targeting quantum states at equilibrium. As suggested in the next paragraph, we foresee two possible paths toward finite temperature TN states.

Matrix product density operators (MPDOs) and locally purified tensor networks (LPTNs) provide already today the option to tackle finite temperature regimes via an imaginary time evolution [168, 169, 170]. Herein, the algorithm starts at the infinite temperature state and starts “cooling” the system via a specified number of time steps and specified step size to reach a given temperature. In its original formulation, both approaches are one-dimensional chains. Matrix product density operator can be formulated as TTN but faces some challenges in terms of the question of positivity [171] or integrating symmetries. In contrast, tree tensor operators (TTO) are the tree-equivalent of an LPTN; they are also a positive loopless representation of density matrices, recently introduced in [172]. However, TTOs cannot represent the infinite temperature state necessary for the imaginary time evolution approach. Instead, a possible use of the TTO employed in LGT simulations consists of a variational algorithm to target finite-temperature states or reconstruct open-system quantum dynamics, by efficiently compressing the relevant information. The TTO enables useful measures, e.g. computing the entanglement of formation as already shown for representative one-dimensional models at finite temperature [172].

IV Conclusions

The field of TN methods for LGTs has shown great potential in the last decade, during which significant efforts have been devoted to develo** numerical algorithms and strategies that can be seen as a complementary approach to MC simulations for high-energy physics. The validity of sign-problem-free TN algorithms has been proven for one-dimensional LGT models for both Abelian and non-Abelian scenarios. Recently, TN methods have also been applied to higher-dimensional LGTs with simple truncated gauge groups and small-to-intermediate lattice sizes. On the one hand, these results prove the effectiveness of TN methods for simulating LGTs even in regimes that are problematic for other numerical methods; on the other hand, they highlight the challenges that one needs to tackle to address state-of-the-art research problems, such as accessing the continuum limits or simulating high-dimensional QCD on large lattices.

In this work, after a general overview of TN methods and their use for LGT simulations, we have described these challenges, starting from the problem of the very large local basis required for complex gauge groups and then discussing several computational bottlenecks in terms of bond dimensions and system sizes.

To mitigate and possibly overcome these problems soon, we have presented a feasible roadmap, in terms of algorithmic developments and numerical strategies that might have a concrete impact in extending the range of applicability of TN algorithms to current research problems in high-energy physics. We foresee that all the presented steps could potentially have a key role in making the TN approach competitive as a complementary method to MC techniques for simulating high-dimensional LGTs, such as large-scale QCD.

Acknowledgements.
The authors acknowledge financial support from the following institutions: the European Union (EU) via the Horizon 2020 research and innovation program (Quantum Flagship) projects PASQuanS2 and Euryqa, via the NextGenerationEU project CN00000013 - Italian Research Center on HPC, Big Data and Quantum Computing (ICSC), and via the QuantERA projects T-NiSQ and QuantHEP; the Italian Ministry of University and Research (MUR) via the Departments of Excellence grants 2023-2027 projects Quantum Frontiers and Quantum Sensing and Modelling for One-Health (QuaSiModO), and via PRIN2022 project TANQU; the Italian Istituto Nazionale di Fisica Nucleare (INFN) via specific initiatives IS-QUANTUM and IS-NPQCD; the German Federal Ministry of Education and Research (BMBF) via project QRydDemo; the University of Bari via the 2023-UNBACLE-0244025 grant; the World-Class Research Infrastructure -- Quantum Computing and Simulation Center (QCSC) of Padova University. We acknowledge computational resources by the Cloud Veneto, CINECA, the BwUniCluster, the University of Padova Strategic Research Infrastructure Grant 2017: “CAPRI: Calcolo ad Alte Prestazioni per la Ricerca e l’Innovazione”, and ReCaS Bari.

References