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

A stochastic analysis approach to lattice Yang–Mills at strong coupling

Hao Shen Department of Mathematics, University of Wisconsin - Madison, USA [email protected] Rongchan Zhu Department of Mathematics, Beijing Institute of Technology, Beijing 100081, China [email protected]  and  Xiangchan Zhu Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100190, China [email protected]
Abstract.

We develop a new stochastic analysis approach to the lattice Yang–Mills model at strong coupling in any dimension d>1d>1, with t’ Hooft scaling βN\beta N for the inverse coupling strength. We study their Langevin dynamics, ergodicity, functional inequalities, large NN limits, and mass gap.

Assuming |β|<N232(d1)N|\beta|<\frac{N-2}{32(d-1)N} for the structure group SO(N)SO(N), or |β|<116(d1)|\beta|<\frac{1}{16(d-1)} for SU(N)SU(N), we prove the following results. The invariant measure for the corresponding Langevin dynamic is unique on the entire lattice, and the dynamic is exponentially ergodic under a Wasserstein distance. The finite volume Yang–Mills measures converge to this unique invariant measure in the infinite volume limit, for which Log-Sobolev and Poincaré inequalities hold. These functional inequalities imply that the suitably rescaled Wilson loops for the infinite volume measure has factorized correlations and converges in probability to deterministic limits in the large NN limit, and correlations of a large class of observables decay exponentially, namely the infinite volume measure has a strictly positive mass gap. Our method improves earlier results or simplifies the proofs, and provides some new perspectives to the study of lattice Yang–Mills model.

2010 Mathematics Subject Classification:
37A25; 39B62; 60H10

1. Introduction

The purpose of this paper is to apply stochastic analysis and ergodic theory for Markov processes to study the lattice Yang–Mills model with structure group G{SO(N),SU(N)}G\in\{SO(N),SU(N)\}. In particular, we will consider the Langevin dynamics of these models, and under explicit strong coupling assumptions, we will prove uniqueness of invariant measures in infinite volume, log-Sobolev and Poincaré inequalities, with some application in large NN limits of Wilson loops and exponential decay of correlations.

Lattice discretizations of the Yang–Mills theories were first proposed in the physics literature by Wilson [Wilson1974] which lead to well-defined Gibbs measures on collections of matrices. We refer to [Chatterjee18] for a nice review on the Yang–Mills model and its gauge invariant discretization as well as the fundamental questions for the model. Among the literature we only mention that approximate computations of the Wilson loop expectations as the size NN of the structure group becomes large was first suggested by ’t Hooft [tHooft1974], where the Yang–Mills Hamiltonian is multiplied by βN\beta N (known as the ’t Hooft scaling), which is closely related to our present article.

The problems we discuss in this paper have been of interest and studied for decades in mathematical physics. A closely related earlier paper is by Osterwalder–Seiler [OS1978], which showed that for the lattice Yang–Mills theory, when the coupling is sufficiently strong, the cluster expansion (or high-temperature expansion in statistical mechanics language) for the expectation values of local observables (i.e. bounded functions of finitely many edge variables) is convergent, uniformly in volume. The proof of this convergent cluster expansion was sketched in [OS1978] since it follows similarly as [GJS1973] for P(ϕ)2P(\phi)_{2} model (and also [MR389080]); in fact it is simpler than the P(ϕ)2P(\phi)_{2} model in [GJS1973] since the fields are bounded in lattice Yang–Mills theory. Moreover, as explained in [OS1978], the existence of a mass gap (exponential clustering) follows from convergence of the cluster expansion, so do existence of the infinite volume limit and analyticity of Schwinger functions in the inverse coupling. Uniqueness of infinite volume limit should also follow from cluster expansion, see e.g. [MR990999] for the case of the P(ϕ)2P(\phi)_{2} model. We also refer to the book [MR785937] for these expansion techniques and results. As for the large NN limits, in the recent papers, factorization property of the Wilson loop expectations was proved in [Cha, Corollary 3.2] and [Jafar] under the assumption that β\beta is sufficiently small.

Given the earlier work, we revisit these problems in this article for a number of reasons. First of all, the earlier work [OS1978] didn’t consider ’t Hooft scaling, but if we translate their results into ’t Hooft scaling where the Hamiltonian is multiplied by βN\beta N then their condition amounts to requiring βN\beta N to be small. However, to our best knowledge, under the ’t Hooft scaling βN\beta N uniqueness was not known for β\beta in a fixed small neighborhood of the origin when NN is arbitrarily large (see for instance the discussion after [Cha, Theorem 3.1]); this is the reason that [Cha] and [Jafar] formulated their large NN results on a sequence of NN-dependent finite volumes. One aim of this paper is to establish uniqueness of infinite volume measures for β\beta in a fixed and explicit small neighborhood of the origin which is uniform in NN, which allows us to prove the existence of a mass gap and large NN limits of Wilson loops directly in infinite volume for this range of β\beta.

Secondly, as another motivation of this paper, we develop new methods based on stochastic analysis and give new proofs to these results. In these methods, the curvature properties of the Lie groups are better exploited via the verification of the Bakry–Émery condition. In particular, this allows us to perform more delicate calculations and obtain more explicit smallness condition on inverse coupling. As another novelty we study the Langevin dynamics (or stochastic quantization) and we prove uniqueness of the infinite volume measures by showing that the dynamic on the entire d{\mathbb{Z}}^{d} has a unique invariant measure. To this end we employed coupling methods for our stochastic dynamics, which is a variant of Kendall–Cranston’s coupling. Such stochastic coupling arguments were used earlier in the stochastic analysis on manifolds, but to our best knowledge this appears to be the first time that such coupling arguments are used in the setting of statistical physics or lattice quantum field theory models with manifold target spaces. For our coupling arguments we will also need to introduce suitable weighted distances on the product manifolds, and in our calculations a subtle comparison between the weight parameter and the curvature plays a key role in order to obtain ergodicity.

As the third motivation, it appears to us that some of the proofs in this paper are simpler. For instance, the large NN results on Wilson loops follow quickly from the Poincaré inequality, which simply comes from the Bakry–Émery condition. Our proof of exponential decay relies on some earlier ideas of Guionnet–Zegarlinski [GZ] together with our explicit bounds on commutators between derivatives and Markov generators on Lie groups. This seems to be simpler than cluster expansion, or at least provides some new perspectives.

1.1. Lattice Yang–Mills

We first recall the basic setup and definitions of the model.

Let ΛL=dL𝕋d\Lambda_{L}={\mathbb{Z}}^{d}\cap L{\mathbb{T}}^{d} be a finite dd dimensional lattice with side length LL and unit lattice spacing, and we will consider various functions on it with periodic boundary conditions. We will sometimes write Λ=ΛL\Lambda=\Lambda_{L} for short. We say that a lattice edge of d{\mathbb{Z}}^{d} is positively oriented if the beginning point is smaller in lexographic order than the ending point. Let E+E^{+} (resp. EE^{-}) be the set of positively (resp. negatively) oriented edges, and denote by EΛL+E_{\Lambda_{L}}^{+}, EΛLE_{\Lambda_{L}}^{-} the corresponding subsets of edges with both beginning and ending points in ΛL{\Lambda_{L}}. Define E=defE+EE\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}E^{+}\cup E^{-} and let u(e)u(e) and v(e)v(e) denote the starting point and ending point of an edge eEe\in E, respectively.

We write GG for the Lie group SO(N)SO(N) or SU(N)SU(N) and 𝔤\mathfrak{g} for the associated Lie algebra 𝔰𝔬(N)\mathfrak{so}(N) or 𝔰𝔲(N)\mathfrak{su}(N). Note that we always view GG as a real manifold (even for SU(N)SU(N)), and 𝔤\mathfrak{g} as a real vector space, and we will write d(𝔤)=dim𝔤d(\mathfrak{g})=\dim_{\mathbb{R}}\mathfrak{g}.

To define the lattice Yang–Mills theory we need more notation, for which we closely follow [Cha] and [SSZloop].

A path is defined to be a sequence of edges e1e2ene_{1}e_{2}\cdots e_{n} with eiEe_{i}\in E and v(ei)=u(ei+1)v(e_{i})=u(e_{i+1}) for i=1,2,,n1i=1,2,\cdots,n-1. The path is called closed if v(en)=u(e1)v(e_{n})=u(e_{1}). A plaquette is a closed path of length four which traces out the boundary of a square. Also, let 𝒫ΛL\mathcal{P}_{\Lambda_{L}} be the set of plaquettes whose vertices are all in ΛL\Lambda_{L}, and 𝒫ΛL+\mathcal{P}^{+}_{\Lambda_{L}} be the subset of plaquettes p=e1e2e3e4p=e_{1}e_{2}e_{3}e_{4} such that the beginning point of e1e_{1} is lexicographically the smallest among all the vertices in pp and the ending point of e1e_{1} is the second smallest.

The lattice Yang-Mills theory (or lattice gauge theory) on ΛL{\Lambda_{L}} for the structure group GG, with β\beta\in\mathbb{R} the inverse coupling constant, is the probability measure μΛL,N,β\mu_{\Lambda_{L},N,\beta} on the set of all collections Q=(Qe)eEΛL+Q=(Q_{e})_{e\in E_{\Lambda_{L}}^{+}} of GG-matrices, defined as

dμΛL,N,β(Q)=defZΛL,N,β1exp(𝒮(Q))eEΛL+dσN(Qe),{\mathord{{\rm d}}}\mu_{\Lambda_{L},N,\beta}(Q)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}Z_{{\Lambda_{L}},N,\beta}^{-1}\exp\Big{(}{\mathcal{S}}(Q)\Big{)}\prod_{e\in E^{+}_{\Lambda_{L}}}{\mathord{{\rm d}}}\sigma_{N}(Q_{e})\,, (1.1)

with {equ}[e:defS] S(Q) =defNβRe∑_p∈P^+_Λ_L Tr(Q_p), where ZΛL,N,βZ_{\Lambda_{L},N,\beta} is the normalizing constant, Qp=defQe1Qe2Qe3Qe4Q_{p}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}Q_{e_{1}}Q_{e_{2}}Q_{e_{3}}Q_{e_{4}} for a plaquette p=e1e2e3e4p=e_{1}e_{2}e_{3}e_{4}, and σN\sigma_{N} is the Haar measure on GG. Note that for p𝒫ΛL+p\in\mathcal{P}^{+}_{\Lambda_{L}} the edges e3e_{3} and e4e_{4} are negatively oriented, so throughout the paper we define Qe=defQe11Q_{e}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}Q_{e^{-1}}^{-1} for eEe\in E^{-}, where e1e^{-1} denotes the edge with orientation reversed. Also, Re{\mathrm{Re}} is the real part, which can be omitted when G=SO(N)G=SO(N).

1.2. Main Results

We will assume the following in our main results on lattice Yang–Mills.

Assumption 1.1.

Suppose that

K𝒮=def{N+2418N|β|(d1)>0,G=SO(N),N+2218N|β|(d1)>0,G=SU(N).\displaystyle K_{\mathcal{S}}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\begin{cases}\displaystyle\frac{N+2}{4}-1-8N|\beta|(d-1)>0,&\qquad G=SO(N)\;,\\ \displaystyle\frac{N+2}{2}-1-8N|\beta|(d-1)>0\;,&\qquad G=SU(N)\;.\end{cases}

Assumption 1.1 is equivalent to the following strong coupling assumption:

|β|<{132(d1)116N(d1),G=SO(N),116(d1),G=SU(N).\displaystyle|\beta|<\begin{cases}\displaystyle\frac{1}{32(d-1)}-\frac{1}{16N(d-1)},&\qquad G=SO(N)\;,\\ \displaystyle\frac{1}{16(d-1)},&\qquad G=SU(N)\;.\end{cases} (1.2)

Define the (product) topological space 𝒬=defGE+{\mathcal{Q}}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}G^{E^{+}}, which will serve as our infinite volume configuration space. By Tychonoff’s theorem 𝒬{\mathcal{Q}} is compact. For each a>1a>1 we define the distance ρ,a\rho_{\infty,a} on 𝒬{\mathcal{Q}} by {equ}[e:rho-inf] ρ_∞,a^2(Q,Q’)=def∑_e∈E^+1a—e—ρ^2(Q_e,Q_e’), with |e||e| being the distance from 0 to ee in d{\mathbb{Z}}^{d}. Here the distances for different choices of aa give equivalent topologies, and we just write ρ\rho_{\infty} when there’s no confusion. 𝒬{\mathcal{Q}} is then a Polish space w.r.t. ρ\rho_{\infty}. By standard results in topology, the topology induced by ρ\rho_{\infty} is equivalent with the product topology on 𝒬{\mathcal{Q}}.

We can easily extend the measure μΛL,N,β\mu_{\Lambda_{L},N,\beta} to the infinite volume configuration space 𝒬{\mathcal{Q}} by periodic extension, which is still denoted as μΛL,N,β\mu_{\Lambda_{L},N,\beta}. Namely, we can construct a random variable with law given by μΛL,N,β\mu_{\Lambda_{L},N,\beta} and extend the random variable periodically, and the law of the periodic extension gives the desired extension of measure. Since GG and 𝒬{\mathcal{Q}} are compact, {μΛL,N,β}L1\{\mu_{\Lambda_{L},N,\beta}\}_{L\geqslant 1} form a tight set.

We will consider the Langevin dynamic on 𝒬{\mathcal{Q}}, formally given by {equ}[e:YM-formal] dQ = ∇S (Q) dt + 2dB , with 𝔅=(𝔅e)eE+\mathfrak{B}=(\mathfrak{B}_{e})_{e\in E^{+}} being independent Brownian motions on GG. This is formal since we will need to “extend” 𝒮\nabla{\mathcal{S}} to infinite volume in a suitable sense. More precisely, the Langevin dynamic we consider is the following SDE system parametrized by eE+e\in E^{+}: {equs}[eq:YM in] dQ_e &= -12Nβ∑_p∈P,p≻e(Q_p-Q_p^*)Q_edt -12(N-1)Q_edt+2dB_eQ_e ,  if  G=SO(N) ,
dQ_e = -12Nβ∑_p∈P,p≻e( (Q_p-Q_p^*) - 1NTr(Q_p-Q_p^*) I_N) Q_edt
             -N2-1NQ_edt+2dB_eQ_e,     if  G=SU(N). Here B=(Be)eEB=(B_{e})_{e\in E} is a collection of independent Brownian motions on the Lie algebra 𝔤\mathfrak{g} of GG, and the terms linear in QeQ_{e} arise from Casimir elements of the Lie algebras; we will review these in Section 2.

Remark 1.1.

We note that the above SDE (in finite volume) was used earlier in [SSZloop] to derive the loop equations (i.e. Dyson–Schwinger or Makeenko–Migdal equations) for Wilson loops of the model (1.1). These loop equations also hold for any infinite volume tight limit of the measures, and in particular for the unique invariant measure for β\beta satisfying (1.2) as given in Theorem 1.2.

The study of a quantum field theory of the form (1.1) via a dynamic (LABEL:e:YM-formal) is also called stochastic quantization as first proposed by [Nelson66, ParisiWu].

We will prove that there exists a unique probabilistically strong solutions to SDE (LABEL:eq:YM_in) starting from any initial data in 𝒬{\mathcal{Q}} in Proposition LABEL:lem:4.7. Hence the solutions form a Markov process in 𝒬{\mathcal{Q}} and the related semigroup is denoted by (Pt)t0(P_{t})_{t\geqslant 0}.

Our first main result is as follows.

Theorem 1.2 (Uniqueness and ergodicity).

Under Assumption 1.1, the following statements hold.

(1) The invariant measure of the Markov semigroup (Pt)t0(P_{t})_{t\geqslant 0} for the Langevin dynamic (LABEL:eq:YM_in) is unique. We denote this invariant measure by μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta}.

(2) Furthermore, every tight limit of {μΛL,N,β}L\{\mu_{\Lambda_{L},N,\beta}\}_{L} is the same, and the whole sequence {μΛL,N,β}L\{\mu_{\Lambda_{L},N,\beta}\}_{L} converges to μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta} as LL\to\infty.

(3) Finally, the Markov semigroup (Pt)t0(P_{t})_{t\geqslant 0} is exponentially ergodic in the following sense: there exists a constant a>1a>1 such that for any ν𝒫(𝒬)\nu\in{\mathscr{P}}({\mathcal{Q}}) {equ}[e:WnuP] W_2^ρ_,a(νP_t,μ^ym_N,β) C(a) e^-~K_S t, t0, for some K~𝒮>0\widetilde{K}_{{\mathcal{S}}}>0 which only depends on the constant aa, dd, β\beta and GG (in particular NN).

Here W2ρ,aW_{2}^{\rho_{\infty,a}} is the Wasserstein distance w.r.t. ρ,a\rho_{\infty,a} given for any μ,ν𝒫(𝒬)\mu,\nu\in{\mathscr{P}}({\mathcal{Q}})

W2ρ,a(μ,ν)=definfπ𝒞(μ,ν)π(ρ,a2)1/2,\displaystyle W_{2}^{\rho_{\infty,a}}(\mu,\nu)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\inf_{\pi\in{\mathscr{C}}(\mu,\nu)}\pi(\rho_{\infty,a}^{2})^{1/2},

with 𝒞(μ,ν){\mathscr{C}}(\mu,\nu) being the set of couplings between μ\mu and ν\nu. Remark that K~𝒮\widetilde{K}_{\mathcal{S}} can be explicitly given by (LABEL:e:tildeK) below and gives a lower bound of spectral gap for (Pt)t0(P_{t})_{t\geqslant 0} in Wasserstein distance. In Theorem 1.4 we will see that K𝒮K_{\mathcal{S}} gives a lower bound of spectral gap in L2(μN,βym)L^{2}(\mu^{\textnormal{\tiny{ym}}}_{N,\beta}).

Remark 1.3.

The periodic boundary condition in the definition of {μΛL,N,β}L\{\mu_{\Lambda_{L},N,\beta}\}_{L} is not essential. By the same argument as in Theorem LABEL:th:in1 the tight limit of {μΛL,N,β}L\{\mu_{\Lambda_{L},N,\beta}\}_{L} when changing the periodic boundary condition to Dirichlet or other boundary conditions is also the invariant measure of the SDE (LABEL:eq:YM_in), hence, is the same as μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta}.

We remark that uniqueness for small β\beta could possibly also be proven using the method of Dobrushin, see e.g. [Dobrushin1970]. To this end one would also need to consider the related Wasserstein metric with respect to the Riemannian distance similarly as we do in this paper. However as we understand such an argument has not been carried out in detail for lattice Yang Mills in the literature. Here we give a proof based on a new idea which is a variant of Kendall–Cranston’s coupling used earlier in the stochastic analysis on manifold.

The idea for the proof of Theorem 1.2 is to use finite dimensional approximation, for which we construct a suitable coupling and find a suitable distance such that the associated Wasserstein distance between the two finite dimensional approximations starting from different initial distributions decays exponentially fast in time with uniform speed.

We define the cylinder functions Ccyl(𝒬)C^{\infty}_{cyl}({\mathcal{Q}}) by {equ}[e:cylQ] C^∞_cyl(Q) ={ F: F=f(Q_e_1,…, Q_e_n),n∈N, e_i∈E^+, f∈C^∞(G^n)}. We then obtain the following log-Sobolev inequality for μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta} based on Bakry–Émery’s criterion.

Theorem 1.4 (Log-Sobolev inequality).

Under Assumption 1.1, the log-Sobolev inequality holds for the measure μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta} in Theorem 1.2. Namely, for all cylinder functions FCcyl(𝒬)F\in C^{\infty}_{cyl}({\mathcal{Q}}) with μN,βym(F2)=1\mu^{\textnormal{\tiny{ym}}}_{N,\beta}(F^{2})=1,

μN,βym(F2logF2)2K𝒮eE+μN,βym(|eF|2).\displaystyle\mu^{\textnormal{\tiny{ym}}}_{N,\beta}(F^{2}\log F^{2})\leqslant\frac{2}{K_{\mathcal{S}}}\sum_{e\in E^{+}}\mu^{\textnormal{\tiny{ym}}}_{N,\beta}(|\nabla_{e}F|^{2}). (1.3)

This implies the Poincaré inequality, i.e. for all cylinder functions FCcyl(𝒬)F\in C^{\infty}_{cyl}({\mathcal{Q}}), {equ}[e:PoinYM] μ^ym_N,β(F^2)1KS_eE^+μ^ym_N,β(—_e F—^2)+μ^ym_N,β(F)^2, with e\nabla_{e} the gradient w.r.t. the variable QeQ_{e}.

Theorem 1.4 follows from Theorem 1.2 and Corollary LABEL:co:lo, which states the log-Sobolev inequality for every tight limit of (μΛL,N,β)L1(\mu_{\Lambda_{L},N,\beta})_{L\geqslant 1}. The RHS of (1.3) is the Dirichlet form associated with the Langevin dynamic (see Proposition LABEL:th:D). Hence, (1.3) holds for the functions in the domain of Dirichlet form by lower-semicontinuity.

As some simple applications of the Poincaré inequality, we show certain “susceptibility” bounds on the field QeQ_{e} and Tr(Qp){\rm Tr}(Q_{p}), see Corollaries LABEL:co:c and LABEL:co:cc. These examples demonstrate how to choose suitable functions in these functional inequalities to yield interesting bounds for the model.

Log-Sobolev and Poincaré inequalities in Theorem 1.4 follow by checking the Bakry–Émery criteria [MR889476, BakryLedoux] directly for the finite dimensional approximation on the product manifolds. As the Ricci curvatures of the target manifolds GG are given by positive constants, so are the Ricci curvatures of the configuration space (i.e. the product manifolds). The Hessian of the Hamiltonian could also be bounded by the Ricci curvatures in the strong coupling regimes.

As a corollary of Theorem 1.4 we obtain the following large NN properties of the Wilson loops. For the rest of this paper, by a loop we mean an equivalent class of closed paths (as defined in Section 1.1), where the equivalence relation \sim is given by cyclic permutations e1e2eneiei+1ene1e2ei1e_{1}e_{2}\cdots e_{n}\sim e_{i}e_{i+1}\cdots e_{n}e_{1}e_{2}\cdots e_{i-1} for any i{1,,n}i\in\{1,...,n\}, and it has no two successive edges of the form e1ee^{-1}e. We will always assume that a loop is non-empty, i.e. has positive number of edges. Given a loop =e1e2en\ell=e_{1}e_{2}\cdots e_{n}, recall that the Wilson loop variable WW_{\ell} is defined as {equ}[e:loop] W_ℓ=defTr(Q_e_1Q_e_2⋯Q_e_n) .

Corollary 1.5 (Large NN limit of Wilson loops).

Under Assumption 1.1, for every Wilson loop (LABEL:e:loop), writing Var{\mathrm{Var}} and 𝐄\mathbf{E} for the variance and expectation under the measure μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta} in Theorem 1.2, one has {equ}[e:varW] Var(1NW_)n(n-3)KSN,  G=SO(N); Var(1NW_)4n(n-3)KSN,  G=SU(N)  . In particular, we obtain the convergence {equ}[e:EW] WN-EWN0   as N→∞ in probability, and the factorization property of Wilson loops, i.e. for any loops 1,,m\ell_{1},\dots,\ell_{m}

limN𝐄W1WmNm=limNi=1m𝐄WiN.\displaystyle\lim_{N\to\infty}\mathbf{E}\frac{W_{\ell_{1}}\dots W_{\ell_{m}}}{N^{m}}=\lim_{N\to\infty}\prod_{i=1}^{m}\mathbf{E}\frac{W_{\ell_{i}}}{N}.

Corollary 1.5 is proven in Section LABEL:sec:3.2. Our proof is novel which is based on the Poincaré inequality (LABEL:e:PoinYM). Note that our formulation of the result is different from [Cha] and [Jafar] in which the factorization property of Wilson loops was obtained by taking a sequence of increasing finite lattices d=N=1ΛN{\mathbb{Z}}^{d}=\cup_{N=1}^{\infty}\Lambda_{N}, considering the correlations of Wilson loops over ΛN\Lambda_{N}, and taking infinite volume limit simultaneously as the large NN limit when sending NN\to\infty. In our approach, we work directly in infinite volume, which seems more natural. The subtlety here, as mentioned above and also explained in [Cha], is that the ’t Hooft coupling places NβN\beta instead of β\beta in front of the Hamiltonian so one would require NβN\beta to be sufficiently small to obtain the infinite volume limit, which would appear to be problematic when taking the large NN limit afterwards. However, thanks to our precise smallness condition on β\beta in (1.2), we can take β\beta small uniformly in NN. This also allows us to derive bounds on the variances of Wilson loops which are explicit in terms of NN. Our proof based on the Poincaré inequality which follows from the Bakry–Émery condition also appears to be simpler than the arguments in aforementioned previous work.

Furthermore, we obtain the following exponential decay property of the covariance. Consider fCcyl(𝒬)f\in C^{\infty}_{cyl}({\mathcal{Q}}) and we write Λf\Lambda_{f} for the set of the edges ff depends on. Let |Λf||\Lambda_{f}| denote the cardinality of Λf\Lambda_{f}. We define

|f|=defeΛfefL,\displaystyle|\!|\!|f|\!|\!|_{\infty}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\sum_{e\in\Lambda_{f}}\|\nabla_{e}f\|_{L^{\infty}},

where Ccyl(𝒬)C^{\infty}_{cyl}({\mathcal{Q}}) is introduced in (LABEL:e:cylQ) and e\nabla_{e} is introduced in Section LABEL:sec:3.1. We also write d(A,B)d(A,B) for the distance between A,BE+A,B\subset E^{+}, which is given by the nearest distance between the vertices in AA and BB.

Corollary 1.6 (Mass gap).

Suppose that Assumption 1.1 holds. Writing Cov{\mathrm{Cov}} for the covariance under the measure μN,βym\mu^{\textnormal{\tiny{ym}}}_{N,\beta} in Theorem 1.2. For f,gCcyl(𝒬)f,g\in C^{\infty}_{cyl}({\mathcal{Q}}), suppose that ΛfΛg=\Lambda_{f}\cap\Lambda_{g}=\varnothing. It holds that

Cov(f,g)c1d(𝔤)ecNd(Λf,Λg)(|f||g|+fL2(μN,βym)gL2(μN,βym)),\displaystyle{\mathrm{Cov}}(f,g)\leqslant c_{1}d(\mathfrak{g})e^{-c_{N}d(\Lambda_{f},\Lambda_{g})}\left(|\!|\!|f|\!|\!|_{\infty}|\!|\!|g|\!|\!|_{\infty}+\|f\|_{L^{2}(\mu^{\textnormal{\tiny{ym}}}_{N,\beta})}\|g\|_{L^{2}(\mu^{\textnormal{\tiny{ym}}}_{N,\beta})}\right),

where c1c_{1} depends on |Λf||\Lambda_{f}|, |Λg||\Lambda_{g}| and cNc_{N} depends on K𝒮K_{\mathcal{S}}, NN and dd.

Note that ff and gg in the above corollary can be chosen to be Wilson loops, or functions of an arbitrary number of Wilson loops, which are of particular interest in physics. We also remark that exponential decay of correlations (together with certain center symmetry conditions) is also related to Wilson’s area law for Wilson loops (see [MR4278289, Theorem 2.4]). The proof of Corollary 1.6 is given in Section LABEL:sec:3.2.

We conclude this subsection by some brief comments on the challenges or subtleties in the proofs of the above results. One of the important ingredients in the proofs is to estimate the Hessian or the general second order derivatives for the interaction 𝒮{\mathcal{S}} defined in (LABEL:e:defS). Note that a term of the form NTr(Qp)N{\rm Tr}(Q_{p}) in the interaction 𝒮{\mathcal{S}} would “appear” to be of order N2N^{2}, which would be too large for us to obtain the desired results. In our proofs we will properly arrange terms and apply certain properties of the Lie groups and we will show that the relevant second order derivatives are actually at most of order |β|N|\beta|N. See the explanations before Lemma LABEL:lem:4.1 and the proof of Theorem 1.2 in Section LABEL:sec:uniqueYM for more details. This is one of the crucial reasons which allow us to compare the Ricci curvatures of the Lie groups and Hessians to verify the Bakry–Émery condition by choosing β\beta small, and also prove ergodicity using a suitable weighted distance.

1.3. Relevant literature and possible directions

The study of properties of lattice gauge theories recently attracts much interest. Besides the aforementioned work by [Cha] and [Jafar], [Chatterjee16] computed the leading terms of free energies, [MR3861073] provided an elaborate description of loop expectations in the planar setting, and [ChatterjeeJafar] derived 1/N1/N expansions in the SO(N)SO(N) case at strong coupling. Wilson loops (and also Wilson lines when coupled with Higgs) for gauge theories with finite structure groups were studied in [MR4107931, FLV20, FLV21, Cao20, forsstromAbelian, Adhikari2021]; see also [GS21] for the U(1)U(1) case. Moreover, exponential correlation decay for lattice gauge theories with finite abelian structure groups was obtained by [Forsstrom2021] using coupling argument, and for finite non-abelian structure groups this was proved by [DdhikariCao].

Our present article provides a new approach to study these models via stochastic analysis and dynamical perspective; see also [SSZloop] for a new derivation of loop i.e. Makeenko–Migdal equations for Wilson loops by such methods.

Remark 1.7.

Here by “stochastic analysis approach”, we do not mean the stochastic analysis approach for 2D Yang–Mills in continuum developed earlier by [GKS89] and [Driver89] (See Def. 3.3 therein) in which parallel translations (which are related with Wilson loops) are formulated as stochastic differential equations. See [MR3982691] and references therein for more recent literature in this direction. Our Yang–Mills SDE on the other hand is the stochastic dynamic for the connection fields on a lattice with fixed spacing, which is along the line of stochastic quantization.

We remark that the choice of constant positive curvature Lie groups SO(N)SO(N) and SU(N)SU(N) in this article is a technical simplification for demonstrating our method, and it should apply as well for other compact target spaces with constant or non-constant positive curvatures. For instance it should apply to a lattice SO(N)SO(N) Yang–Mills model coupled with a Higgs field Φ\Phi which takes values in a sphere in N\mathbb{R}^{N} (i.e. rotator model) via a gauge-covariant derivative term, whose action takes the form RepTr(Qp)+e|QeΦv(e)Φu(e)|2{\mathrm{Re}}\sum_{p}\mathrm{Tr}(Q_{p})+\sum_{e}|Q_{e}\Phi_{v(e)}-\Phi_{u(e)}|^{2}.

It would certainly be interesting to show if log-Sobolev inequalities still hold when the lattice spacing vanishes, in the situations where the continuum limits of these models are shown or expected to exist. In this direction, on the two dimensional torus, the continuum limit of lattice approximations of the Yang–Mills measures on 1-forms was recently obtained by Chevyrev [Chevyrev19YM], who also showed that a certain class of Wilson loop observables of this random 1-form coincide in law with the corresponding observables under the Yang–Mills measure in the sense of [Levy03]. Note that the Langevin dynamics for Yang–Mills models on the two and three dimensional continuous torus were recently constructed in [CCHS2d, CCHS3d] (see [ChevyrevReview] for a review of these results), and as mentioned in [CCHS2d] it would be interesting to show that the lattice dynamics of the type (LABEL:e:YM-formal) converge to the processes constructed in the above papers in two and three dimensions. For some of the recent progress along this direction, see the proof of log-Sobolev inequalities for the Φ2,34\Phi^{4}_{2,3} and sine-Gordon models [RolandPhi4, RolandSG], and the 1D nonlinear σ\sigma-model (see [AnderssonDriver, Hairer16, StringManifold]) for which the log-Sobolev inequality, ergodicity and non-ergodicity (depending on the curvature of the target manifolds) were obtained in [RWZZ17, CWZZ18].

Notation. Given a Polish space EE, we write C([0,T];E)C([0,T];E) for the space of continuous functions from [0,T][0,T] to EE. We use 𝒫(E){\mathscr{P}}(E) to denote all the probability measures on EE with Borel σ\sigma-algebra.

Acknowledgments. We would like to thank Scott Smith, Feng-Yu Wang, Xicheng Zhang and Xin Chen for very helpful discussions. H.S. gratefully acknowledges financial support from NSF grants DMS-1954091 and CAREER DMS-2044415. R.Z. gratefully acknowledges financial support from the NSFC (No. 11922103). X.Z. is grateful to the financial supports by National Key R&D Program of China (No. 2020YFA0712700) and the NSFC (No. 12090014, 12288201) and the support by key Lab of Random Complex Structures and Data Science, Youth Innovation Promotion Association (2020003), Chinese Academy of Science. R.Z. and X.Z. are grateful to the financial support by the DFG through the CRC 1283 “Taming uncertainty and profiting from randomness and low regularity in analysis, stochastics and their applications”

2. Notation and preliminaries

In this section we collect some notation and standard facts about Riemannian geometry, Lie groups and Brownian motions.

Riemannian manifolds. Let MM be a Riemannian manifold of dimension dd. We denote by C(M)C^{\infty}(M) the space of real-valued smooth functions on MM. For xMx\in M we denote by TxMT_{x}M the tangent space at xx with inner product ,TxM\langle\cdot,\cdot\rangle_{T_{x}M}. For XTxMX\in T_{x}M, we write XfXf or X(f)X(f) for the differentiation of ff along XX at xx. For a smooth curve γ:[α,β]M\gamma:[\alpha,\beta]\to M the tangent vector along γ\gamma is defined by

γ˙tf=ddtf(γt),fC(M).\displaystyle\dot{\gamma}_{t}f=\frac{{\mathord{{\rm d}}}}{{\mathord{{\rm d}}}t}f(\gamma_{t}),\quad f\in C^{\infty}(M).

Let \nabla be the Levi-Civita connection, which is a bilinear operation associating to vector fields XX and YY a vector field YX\nabla_{Y}X. Recall that (YX)(x)(\nabla_{Y}X)(x) depends on YY only via Y(x)Y(x) for xMx\in M (e.g. [MR1138207, Remark 2.3]).

For fC(M)f\in C^{\infty}(M), we denote by f\nabla f the gradient vector field of ff. We also write Hess(f){\mathord{{\rm Hess}}}(f) for the Hessian. It can be calculated in the following ways {equ}[e:hess] Hess_f(X,Y) =defHess(f) (X,Y) = _X f, Y = X(Y f ) - (_X Y) f . It is a two-tensor: Hessf(φX,Y)=Hessf(X,φY)=φHessf(X,Y){\mathord{{\rm Hess}}}_{f}(\varphi X,Y)={\mathord{{\rm Hess}}}_{f}(X,\varphi Y)=\varphi{\mathord{{\rm Hess}}}_{f}(X,Y) for any φC(M)\varphi\in C^{\infty}(M) so Hessf(X,Y)(x){\mathord{{\rm Hess}}}_{f}(X,Y)(x) depends only on X(x)X(x) and Y(x)Y(x). Since Levi-Civita connection is torsion-free, Hess(f){\mathord{{\rm Hess}}}(f) is symmetric in X,YX,Y.

The Riemann curvature tensor (,){\mathscr{R}}(\cdot,\cdot) associated to vector fields X,YX,Y is an operator defined by

(X,Y)Z=X(YZ)Y(XZ)[X,Y]Z.\displaystyle{\mathscr{R}}(X,Y)Z=\nabla_{X}(\nabla_{Y}Z)-\nabla_{Y}(\nabla_{X}Z)-\nabla_{[X,Y]}Z.

Let {Wi}i=1d\{W_{i}\}_{i=1}^{d} be an orthonormal basis of TxMT_{x}M. The Ricci curvature tensor is defined by {equ}[e:ricci] Ric(X,Y)=_i=1^dR(X,W_i)W_i,Y_T_xM and is independent of the choice of {Wi}\{W_{i}\}. Note that Ric(X,Y)(x){\rm Ric}(X,Y)(x) depends on X,YX,Y only via X(x),Y(x)X(x),Y(x) for xMx\in M.

Let γ\gamma be a geodesic. A smooth vector field JJ is called a Jacobi field along γ:[0,t]M\gamma:[0,t]\to M if γ˙γ˙J+(J,γ˙)γ˙=0\nabla_{\dot{\gamma}}\nabla_{\dot{\gamma}}J+{\mathscr{R}}(J,\dot{\gamma})\dot{\gamma}=0. For any XTγ0MX\in T_{\gamma_{0}}M and YTγtMY\in T_{\gamma_{t}}M, there exists a Jacobi field JJ along γ\gamma satisfying J0=XJ_{0}=X and Jt=YJ_{t}=Y (c.f. [CE75, Section 1.5], [Wang, Section 0.4] ).

Lie groups and algebras. For any matrix MM we write MM^{*} for the conjugate transpose of MM. Let MN()M_{N}(\mathbb{R}) and MN()M_{N}(\mathbb{C}) be the space of real and complex N×NN\times N matrices.

For Lie groups SO(N)SO(N), SU(N)SU(N), we write the corresponding Lie algebras as 𝔰𝔬(N)\mathfrak{so}(N), 𝔰𝔲(N)\mathfrak{su}(N) respectively. Every matrix QQ in one of these Lie groups satisfies QQ=INQQ^{*}=I_{N}, and every matrix XX in one of these Lie algebras satisfies X+X=0X+X^{*}=0. Here INI_{N} denotes the identity matrix.

We endow MN()M_{N}(\mathbb{C}) with the Hilbert-Schmidt inner product {equ}[e:HS] X,Y = ReTr(X Y^*)   X,YM_N(C). We restrict this inner product to our Lie algebra 𝔤\mathfrak{g}, which is then invariant under the adjoint action. In particular for X,Y𝔰𝔬(N)X,Y\in\mathfrak{so}(N) or 𝔰𝔲(N)\mathfrak{su}(N) we have X,Y=Tr(XY)\langle X,Y\rangle=-\mathrm{Tr}(XY). Note that Tr(XY){\rm Tr}(XY)\in\mathbb{R} since we have Tr((XY))=Tr(YX)=Tr(XY){\rm Tr}((XY)^{*})={\rm Tr}(Y^{*}X^{*})={\rm Tr}(XY), and Tr(A)=Tr(A)¯{\rm Tr}(A^{*})=\overline{{\rm Tr}(A)} for any AMN()A\in M_{N}(\mathbb{C}).

Below GG is always understood as SO(N)SO(N) or SU(N)SU(N). Every X𝔤X\in\mathfrak{g} induces a right-invariant vector field X~\widetilde{X} on GG, and for each QGQ\in G, X~(Q)\widetilde{X}(Q) is just given by XQXQ since GG is a matrix Lie group. Indeed, given any X𝔤X\in\mathfrak{g}, the curve tetXQt\mapsto e^{tX}Q is well approximated near t=0t=0 by Q+tXQQ+tXQ up to an error of order t2t^{2}.

The inner product on 𝔤\mathfrak{g} induces an inner product on the tangent space at every QGQ\in G via the right multiplication on GG. Hence, for X,Y𝔤X,Y\in\mathfrak{g}, we have XQ,YQTQGXQ,YQ\in T_{Q}G, and their inner product is given by Tr((XQ)(YQ))=Tr(XY)\mathrm{Tr}((XQ)(YQ)^{*})=\mathrm{Tr}(XY^{*}). This yields a bi-invariant Riemannian metric on GG.

For any function fC(G)f\in C^{\infty}(G) and X𝔤X\in\mathfrak{g}, the right-invariant vector field X~\widetilde{X} induced by XX acts on ff at QGQ\in G by the right-invariant derivative {equ}[e:Xf] ~X f (Q)= ddt—_t=0 f(e^tX Q). We have {equ} ~[X,Y] = [~X, ~Y] ,   namely,   ([X,Y] Q ) f (Q) = [XQ, YQ] f(Q), where the [,][\cdot,\cdot] is the Lie bracket on 𝔤\mathfrak{g} on the LHS and the vector fields commutator on the RHS. Also, for the Levi-Civita connection \nabla we have {equ}[e:F27] _~X ( ~Y) =12 ~[X,Y] .

We refer the above facts to [AGZ, Appendix F], e.g. Lemma F.27 therein.

Brownian motions. Denote by 𝔅\mathfrak{B} and BB the Brownian motions on a Lie group GG and its Lie algebra 𝔤\mathfrak{g} respectively. The Brownian motion BB is characterized by

𝐄[B(s),XB(t),Y]=min(s,t)X,YX,Y𝔤.\mathbf{E}\Big{[}\langle B(s),X\rangle\langle B(t),Y\rangle\Big{]}={\mathord{{\rm min}}}(s,t)\langle X,Y\rangle\qquad\forall X,Y\in\mathfrak{g}. (2.1)

By [Levy11, Sec. 1.4], the Brownian motions 𝔅\mathfrak{B} and BB are related through the following SDE: {equ}[e:dB] dB = dB B = dB  B + cg2 B dt, where \circ is the Stratonovich product, and dB𝔅{\mathord{{\rm d}}}B\,\mathfrak{B} is in the Itô sense. Here the constant c𝔤c_{\mathfrak{g}} is determined by αvα2=c𝔤IN\sum_{\alpha}v_{\alpha}^{2}=c_{\mathfrak{g}}I_{N} where (vα)α=1d(𝔤)(v_{\alpha})_{\alpha=1}^{d(\mathfrak{g})} is an orthonormal basis of 𝔤\mathfrak{g}. Moreover, by [Levy11, Lem. 1.2], 111Note that in [Levy11, Lem. 1.2], the scalar product differs from (LABEL:e:HS) by a scalar multiple depending on NN and 𝔤\mathfrak{g}, so we accounted for this in the expression for c𝔤c_{\mathfrak{g}} above. {equ}[e:c_g] c_so(N) = -12(N-1),  c_su(N) = -N2-1N .

Denote by δ\delta the Kronecker function, i.e. δij=1\delta_{ij}=1 if i=ji=j and 0 otherwise. For any matrix MM, we write MijM^{ij} for its (i,j)(i,j)th entry. The following holds by straightforward calculations (see e.g. [SSZloop, (2.5)]): \minilabe:BB {equs}[2] d B^ij, B^kℓ⟩ &=12(δ_ikδ_j-δ_iδ