We review two works [Chandra et al., Publ. Math. l’IHÉS (published online, 2022) and Chandra et al., arXiv:2201.03487 (2022)] that study the stochastic quantization equations of Yang–Mills on two- and three-dimensional Euclidean space with finite volume. The main result of these works is that one can renormalize the 2D and 3D stochastic Yang–Mills heat flow so that the dynamic becomes gauge covariant in law. Furthermore, there is a state space of distributional 1-forms to which gauge equivalence approximately extends and such that the renormalized stochastic Yang–Mills heat flow projects to a Markov process on the quotient space of gauge orbits . In this Review, we give unified statements of the main results of these works, highlight differences in the methods, and point out a number of open problems.
I. INTRODUCTION
A. Yang–Mills theory
Yang–Mills (YM) theory plays an important role in the description of force-carrying particles in the standard model. An important unsolved problem in mathematics is to show that YM theory on Minkowski space-time can be rigorously quantized. We refer to Ref. 42 for a description of this problem, together with the surveys16,48 for related literature and problems.
Without loss of generality, we can take G ⊂ U(N) and for some N ≥ 1. In this case, the adjoint action is AdgA = gAg−1 and a possible choice for the inner product is ⟨X, Y⟩ = −Tr(XY), and one can rewrite .
The case d = 4 corresponds to physical space-time, but the task of constructing the probability measure μ makes sense for arbitrary dimension and even for a Riemannian manifold M in place of Rd. The -valued 1-forms in this case become connections on a principal G-bundle P → M, and one aims to define the probability measure μ on the space of connections on P.
The cases d = 2, 3 are considered substantially simpler than d = 4 as they correspond to super-renormalizable theories in quantum field theory (vs renormalizable for d = 4 and non-renormalizable for d ≥ 5). In the remainder of this article, we will focus on these dimensions and further restrict to finite volume, replacing Rd by the torus Td = Rd/Zd. The underlying principal bundle P is always assumed trivial and we keep in mind that all geometric objects (connections, curvature forms, etc.) can be written in coordinates (-valued 1-forms, 2-forms, etc.) once we fix a global section of P that identifies it with Td × G. The space of connections is an affine space, with the difference of two connections being a 1-form.
We write A ∼ B if there exists g such that Ag = B and write [A] = {B: B ∼ A} for the gauge orbit of A. In light of the above, the natural space on which to define the probability measure μ is not the space of 1-forms, but rather the quotient space of all gauge orbits. The space is a non-linear space if G is non-Abelian, which makes non-trivial even the construction of the state space on which the YM measure μ should be defined.
A number of works have made contributions to a precise definition of this measure. The most successful case is dimension d = 2, which includes R2 and compact orientable surfaces. The key feature that makes 2D YM special is its exact solvability, which allows one to write down an explicit formula for the joint distribution of Wilson loop observables; this property was observed in the physics literature by Migdal53 and later developed in mathematics; see, e.g., Refs. 17, 26, 29, 33, 44, 45, and 60.
In the Abelian case G = U(1) on Rd, one can make sense of the measure μ for d = 332 and d = 4.25 For any structure group G, a form of ultraviolet stability on Td was demonstrated for d = 4 using a continuum regularization in Ref. 51 and for d = 3, 4 using a renormalization group approach on the lattice in a series of works by Balaban2–4 (see also Ref. 28). However, a construction of the 3D YM measure and a description of its gauge-invariant observables, even on T3, remains open.
B. Stochastic quantization
Another approach to the construction of the YM measure was recently initiated in Refs. 11 and 12, which is based on stochastic quantization (see also Ref. 61 that treats scalar QED on T2). The basic idea behind this approach is to view the measure μ as the invariant measure of a Langevin dynamic. By studying this dynamic, one can try to determine properties and even give constructions of μ. The method was put forward in the context of gauge theories by Parisi–Wu57 and has recently been applied to the construction of scalar theories (see Refs. 1, 34, 40, and 54).
We point out right away that a difficulty in solving (1.1), even in the absence of noise, is that the equation is not parabolic. This is a well-known feature of YM theory and is connected with the infinite-dimensional nature of the gauge group: the YM equations are not elliptic and admit infinitely many solutions because if A is a solution, then so is every element of the gauge orbit [A].
1. Gauge covariance
The reason why (1.3) is natural is that it is (formally) gauge covariant in law. To see this, it is convenient to work in coordinate-free notation and, for the moment, make an distinction between connections and 1-forms. Recall that the space of connections is affine modeled on the space Ω1 of -valued 1-forms [we are using the global section of P to identify ad(P)-valued forms with -valued forms]. For a connection A, recall further that dA: Ωk → Ωk+1 is a linear map from -valued k-forms to (k + 1)-forms with adjoint . Furthermore, gauge transformations g act on connections via A ↦ Ag and on forms via ω ↦ Adgω.
If we now assume ξ is a white noise and that the equations above make sense with the global section in time solutions, then Adgξ is equal in law to ξ by Itô isometry. This formal argument suggests that there is a coupling between two solutions A, B to (1.3) starting from gauge equivalent initial conditions such that A(t) ∼ B(t) for all t ≥ 0. In particular, the law of the projected process [A] on gauge orbits is equal to that of [B]. The projected process on gauge orbits is therefore well-defined and Markov, and its invariant probability measure is a natural candidate for the YM measure.
C. Main results
The basic objective of Refs. 11 and 12 is to make rigorous the above formal argument in the case of Td for d = 2 and d = 3, respectively. In particular, they aim to define a natural Markov process on gauge orbits associated with the YM Langevin dynamic (1.3). In this subsection, we give unified statements of the main results in Refs. 11 and 12. We will go into more detail and discuss the differences in proofs in Secs. II and III. We also discuss in detail these results in the simple case that G = U(1) in Sec. I D.
Recall that A represents a geometric object (a principal G-connection). On the other hand, the solution A to (1.4) at positive times is expected to be a distribution of the same regularity as the GFF. Therefore, a first natural question is whether there exists a state space large enough to support the GFF while small enough so that gauge equivalence extends to . One of the main results of Refs. 11 and 12 gives an answer to this question, which can be informally stated as follows.
There exists a metric space of -valued distributional 1-forms on Td, d = 2, 3, which contains all smooth 1-forms and to which gauge equivalence approximately extends in a canonical way. Furthermore, contains distributions of the same regularity as the GFF on Td.
The constructions of in Refs. 11 and 12 are rather different. In Ref. 11, (therein denoted by ) is a Banach space defined through line integrals, and gauge equivalence is determined by an action of a gauge group. In Ref. 12, is a non-linear metric space of distributions defined in terms of the effect of the heat flow; gauge equivalence is extended using a gauge-covariant regularizing operator (the deterministic YM flow). We describe these constructions further in Secs. II A and III A.
While Theorem 1.3 makes it seems like the 2D and 3D cases are on an equal footing, we actually know much more about in 2D than in 3D, e.g., the space of orbits in 2D comes with a natural complete metric, and is thus Polish, while we only know that is completely Hausdorff (and separable) in 3D.
We now turn to the question of solving (1.4). A natural approach is to replace ξ by a smooth approximation ξɛ and let the mollification parameter ɛ ↓ 0. The hope then is that the corresponding solutions A converge. Unfortunately, this is not, in general, the case and the stochastic partial differential equation (SPDE) requires renormalization for convergence to take place. The following result ensures that renormalized solutions to (1.4) exist, at least up to a potential finite time blow-up.
We call the ɛ ↓ 0 limit of A as in Theorem 1.6 the solution to (1.5) driven by ξ with bare mass .
The bare mass is used to parameterize the space of all “reasonable” solutions and is a free parameter at this stage. We will see below (Theorems 1.11 and 1.14) that there does exist a unique choice for , which selects a distinguished element of this solution space.
The operators are called the Bogoliubov-Parasiuk-Hepp-Zimmermann (BPHZ) constants and are given by (in principle, explicit) integrals involving χ and an arbitrary large scale truncation K of the heat kernel. While the solution to (1.5) in Definition 1.7 is independent of χ, it does, in general, depend on the choice of K used to define .
The point is a cemetery state and is added to to handle the possibility of finite time blow-up. Some care is needed to properly define {}) and the metric that one equips it with. This is done in Ref. 11 (Sec. 1.5.1), where, for a general metric space E, a metric space Esol of continuous paths with values in E ⊔{} is defined in which two paths are close if they track each other until the point when they become large. Our notation {}) here really means .
It turns out that in 2D, due to a cancellation in renormalization constants, converges to a finite value as ɛ ↓ 0 (see Theorem 2.5). Therefore, Theorem 1.6 in 2D remains true if replaced by any fixed , which is the formulation of Ref. 11 (Theorem 2.4). No such cancellation occurs in 3D and diverges at order ɛ−1.
We now discuss the way in which solutions to (1.5) are gauge covariant in the sense described in Sec. I B 1. Remark that by inserting the counterterm , we are seemingly breaking the desired gauge covariance property discussed in Sec. I B 1 [in the notation of that section, should be written ]. However, the formal argument in Sec. I B 1 also breaks if we replace ξ by ξɛ because Itô isometry is not true for the latter.
A surprising fact is that if one chooses carefully, then the broken gauge covariance [due to the counterterm ] compensates in the ɛ ↓ 0 limit for the broken Itô isometry (due to the mollified noise ξɛ), and one obtains a solution to (1.5), which is gauge covariant in law.
It is not entirely trivial to make this statement precise, essentially because we do not know if (1.5) (with any bare mass) is global in time solutions. In particular, we do not know how to rule out that solutions to (1.5) with different gauge equivalent initial conditions a ∼ b blow up at different times, and this makes it unclear in what sense we can expect the projected process [A] on gauge orbits to be Markov.
To address this issue, it is natural to look for a type of process that solves (1.5) on disjoint intervals [ςj−1, ςj) and at time ςj jumps to a new representative of the gauge orbit . This should happen in such a way that A does not blow up unless the entire orbit [A] “blows up.” This class of processes is defined through generative probability measures in Refs. 11 and 12.
We say that a probability measure μ on the space of càdlàg functions {}) is generative with bare mass and initial condition if there exists a white noise ξ and a random variable A with law μ such that
A(0) = a almost surely;
there exists a sequence of stopping times ς0 = 0 ≤ ς1 ≤ ς2 < ⋯ such that A solves (1.5) driven by ξ with bare mass on each interval [ςj, ςj+1);
for every j ≥ 0, ; and
The point of this definition is to give a sufficiently general and natural way in which (1.5) can be restarted along gauge orbits. The following result from Refs. 11 and 12 ensures the existence of a canonical Markov process associated with (1.5) on the quotient space of gauge orbits , provided the bare mass is chosen in a precise way.
For every and , there exists a generative probability measure μ with bare mass and initial condition a.
There exists with the following properties. For all , if μ, ν are generative probability measures with initial conditions a, b, respectively, and bare mass , then the pushforward measures π*μ and π*ν are equal. In particular, the probability measure Px = π*μ, where μ is generative with bare mass and initial condition , depends only on x. Finally, are the transition functions of a time homogeneous, continuous Markov process on {}.
Theorem 1.11(b) makes no claims about the uniqueness of , but we conjecture that is indeed unique (which is not difficult to prove in the Abelian case; see Sec. I D).
Finally, we mention a result in Refs. 11 and 12 that is crucial for the Proof of Theorem 1.11(b) and that makes precise the coupling argument outlined in Sec. I B 1. This result makes a stronger statement about the constant for which uniqueness does hold. It can also be seen as a version of the Slavnov–Taylor identities for renormalization schemes that preserve gauge symmetries.
A mollifier χ is called non-anticipative if it has support in (0, ∞) × Rd.
What we would therefore like to show is that, for a special choice of , the solutions to (1.8) and (1.9) converge as ɛ ↓ 0 to the same limit. The identity Ag = B, which survives in the limit, would provide a coupling between (1.5) with initial condition A(0) and initial condition A(0)g(0) under which the two solutions are gauge equivalent, at least locally in time. It turns out that the following more general result is true.
It follows from Theorem 1.14 that the solutions to (1.8) and (1.9) indeed converge to the same limit as ɛ ↓ 0, provided we choose . The operator in Theorem 1.14 is exactly the operator appearing in Theorem 1.11(b); in 2D, we can give an explicit expression for [see (2.13)].
The value of in Theorem 1.14 is determined uniquely after we fix a choice for . However, recall from Remark 1.8 that is not unique or canonical—it is determined by χ and an arbitrary truncation of the heat kernel (see, e.g., Theorem 2.5). The final part of Theorem 1.14 states that the solution of (1.5) with bare mass is independent of χ and this choice of truncation.
E. Abelian case
We end this section by discussing the above results in the Abelian case, i.e., G = U(1) and . We consider here d ≥ 1 arbitrary. We will see in this case that
the constant in Theorem 1.11(b) is and is unique,
if Td is replaced by Rd, then uniqueness of fails (Remark 1.16),
(1.5) with bare mass is global in time solutions but no invariant probability measure (Remark 1.17)
Since Adg is now the identity, it is clear that a possible value for in Theorems 1.11 and 1.14 is . This is because if B(0) = A(0) + dω(0) and A, B solve (1.11) with , then B = A + dω for all times, where ω solves ∂tω = Δω.
Clearly (1.11) with does not have an invariant probability measure because evolves like a Brownian motion. In fact, any gauge equivalent generalization of (1.5) of the form ∂tA = ΔA + ξ + dω, where ω is adapted, will lack an invariant probability measure because the spatial mean is unaffected by dω. However, we do obtain an invariant probability measure for the projected process [A] because any 1-form is gauge equivalent to another 1-form B such that . This remark shows that the Markov process from Theorem 1.11 can have an invariant probability measure, while (1.5) with bare mass does not.
II. TWO DIMENSIONS
A. State space
The definition of the state space (denoted by in Ref. 11) is motivated by the desire to define holonomies and, thus, Wilson loops (for every element ). The construction is a refinement of that introduced in Ref. 17. Let , where . We think of as the collection of straight line segments ℓ = (x, v) in T2 of length at most . (The starting point of ℓ is x.)
The Banach space is defined as the completion of smooth -valued 1-forms under |·|α for some .
The metric Σ in Theorem 1.3 is then the usual metric Σ(A, B) = |A − B|α.
To motivate these norms, consider A = (A1, A2) a pair of i.i.d. GFFs. A simple calculation shows that, for all α < 1, E|A(ℓ)|2 ≲ |ℓ|2α. Furthermore, it follows from the Stokes theorem and the fact that dA is a white noise and that A(∂P) = ∫PdA and, hence, E|A(∂P)|2 = |P|. A Kolmogorov argument then implies that the GFF has a modification with |A|α < ∞ almost surely.
- For every and , one has , where is the path . The holonomy hol(A, ℓ) ∈ G, defined by hol(A, ℓ) = y1, where y solves the ordinary differential equation (ODE)is, therefore, well-defined by Young integration.31,50,63
More generally, hol(A, γ) is well-defined for any γ ∈ C1,β([0, 1], T2), where , and the map (A, γ) ↦ hol(A, γ) is Hölder continuous. In particular, classical Wilson loop observables are well-defined with good stability properties. The relation ∼ can be expressed entirely in terms of holonomies.
Since hol(A, γ) is independent of the parameterization of γ, it is natural to also measure the regularity of γ in a parameterization independent way. Such a notion of regularity is introduced in Ref. 11 (Sec. 3.2), which interpolates between C1 and C2 (akin to how p-variation is a parameterization invariant interpolation between C0 and C1).
- One has the embeddingswhere is the completion of smooth functions under |·|α. (Only the last of these is non-trivial; see Ref. 17, Proposition 3.21.) These embeddings are furthermore optimal in the sense that α/2 in Cα/2 cannot be decreased and α − 1 in Cα−1 cannot be increased. Remark also that , while , since α < 1, contains distributions that cannot be represented by functions, such as the GFF.
There exists a complete metric D on the quotient space of gauge orbits , which induces the quotient topology. To define D, we first define a new (but topologically equivalent) metric k on by shrinking the usual metric Σ(·, ·) = |· − ·|α in such a way that the diameter of every R-sphere goes to zero as R → ∞, but the distance between Sr and SR for large r ≤ R is of order , so, in particular, it goes to ∞ as R → ∞. Then, D is defined as the Hausdorff distance associated with the metric k on .
The space strengthens the definition of a Banach space introduced in Ref. 17; is defined in a similar way but with taken as the set of axis-parallel line segments. The main result of Ref. 17 is that if G is simply connected, then there exists a (non-unique) probability measure on such that the holonomies along all axis-parallel curves agree in distribution with those of the YM measure on T2 defined in Refs. 44 and 60. The proof of this result uses a gauge-fixed lattice approximation, which explains the restriction to axis-parallel lines.
B. Local solutions
It turns out that in 2D we can sharpen the statement of Theorem 1.6 as follows.
The ɛ ↓ 0 limit of A, which solves (1.5) with bare mass , depends on K and but not on χ.
If is simple (which one can assume without loss of generality; see Ref. 11, Remark 2.8), then λ < 0 is just a scalar.
Recall from Remark 1.10 that the convergence of to a finite limit is special to dimension 2 and is due to a cancellation between the diverging constants and .
We briefly describe the ingredients in the Proof of Theorem 2.5, which is based on the theory of regularity structures. We only mention the overall strategy behind this theory and refer to Refs. 14, 31, and 37 for an introduction and more details. To solve an SPDE such as (1.4), one constructs a sufficiently large “regularity structure” and lifts the equation to a space of “modeled distributions” with values in the regularity structure. This construction, first introduced in Ref. 36, is done at a high level of generality in Ref. 8. One then constructs a finite number of stochastic objects from the noise called a “model”—these objects are essentially renormalizations of functions of forms (2.3) and (2.4)—the existence of which follows from Ref. 13. The point of construction is that the products A∂A, A3 and convolution with the heat kernel become stable operations on modeled distributions, and one can solve a fixed point problem for the “lifted” equation. Finally, one maps the resulting modeled distribution to a distribution on R × T2 via the “reconstruction operator” and identifies it with a solution to a classical renormalized PDE, at least for ɛ > 0. This final step is carried out systematically in Ref. 6. All these operations are done in a way that is stable as ɛ ↓ 0, thereby showing the desired convergence.
One of the contributions of Ref. 11 is to develop a framework in which the algebraic results from Refs. 6 and 8 can be transferred to a setting in which the noise and solution are vector-valued. References 6 and 8 provided a general method to compute the renormalized form of a system of scalar SPDEs, which, in principle, does apply to (1.4) by writing it as a system of scalar-valued equations using a basis. However, such a procedure is cumbersome and unnatural; it is more desirable to find a framework that preserves the vector-valued nature of the noise and solution, which is the purpose of Ref. 11 (Sec. 5).
The main idea behind the extension in Ref. 11 is to define a category of “symmetric sets” and a functor between this category and the category of vector spaces. This construction allows one to canonically associate partially symmetrized tensor products of vector spaces to combinatorial rooted trees that commonly appear in regularity structures. One of the main outcomes is a procedure to compute the renormalized form of equations like (1.4) without resorting to a basis.
It follows from the general theory of regularity structures that (2.2), for any , converges locally in time in Cα−1. To improve this to convergence in {}), one decomposes the solution A into A = Ψ + B, where Ψ solves the SHE ∂tΨ = ΔΨ + ξ with initial condition Ψ(0) = A(0) and B is in C1−κ for all κ > 0. One can then show by hand that (see Ref. 11, Sec. 4), which, together with the embeddings , shows that A converges to a maximal solution with values in .
C. Gauge covariance
Recall that Theorems 1.11 and 1.14 imply a form of gauge covariance for (1.5). Theorem 1.11(a) is a relatively straightforward consequence of Theorem 2.5. One defines the random variable A by solving (1.5) until the first time that |A(t)|α ≥ 2 + infB∼A(t)|B|α, at which point one uses a measurable selection to jump to a new small representative B of the gauge orbit [A(t)] for which |B|α < 1 + infa∈[A(t)]|a|α. These jump times define the increasing sequence of stopping times in item (ii). Items (i)–(iv) all follow readily from the construction.
The Proof of Theorem 1.11(b), which is the main statement of Theorem 1.11, requires more work. The idea is to use Theorem 1.14, which we admit for now, to couple the solutions to (1.5) with bare mass and initial conditions a ∼ b. Specifically, let ν be a generative probability measure with bare mass and initial condition , and consider any a ∼ b. Letting B and denote the random variable and white noise, respectively, corresponding to ν, it follows from Theorem 1.14 that, on the same probability space, there exists a càdlàg process A defined as above with (1.5) driven by and bare mass in such a way that B = Ag. Here, g is càdlàg with values in {} and jump times contained in those of A and B and solves (1.8) in between these jump times; see Fig. 1. This shows that the pushforward of ν to the orbit space is equal to the pushforward of the law of A from the Proof of Theorem 1.11(a), which proves Theorem 1.11(b).
Well-posedness for systems (2.7) and (2.8) as ɛ ↓ 0 is generally standard. However, a subtlety arises from the multiplicative noise term Adgξ, which is in C−2−κ and leads to problems in posing a suitable fixed point problem (−2 is the threshold regularity at which one cannot extend uniquely a distribution from R+ × Rd to R × Rd). This issue is handled by decomposing , where is the harmonic extension of g(0) to positive times. Then, vanishes at t = 0, which makes the product better behaved, while the product is shown to be a well-defined distribution in C−2−κ(R × Rd) using stochastic estimates.
To derive and solve the equation for , one substitutes ξ by ξδ and takes the limit δ ↓ 0 with ɛ > 0 fixed; this ensures that all objects are smooth for ɛ, δ > 0. This also explains the definition of as a limit in δ ↓ 0.
To summarize, for any and any mollifier χ, the solutions B and to (2.9) and (2.10), respectively, converge to the same limit as ɛ ↓ 0 over a short random time interval [the argument in Ref. 11 that (B, g) and converge as maximal solutions uses non-anticipativity of χ].
To conclude the Proof of Theorem 1.14, it remains to show that (1.5) with bare mass defined by (2.13) is independent of χ and of K. Remark that χ can now be any mollifier, not necessarily non-anticipative. Independence of χ follows from the final part of Theorem 2.5 since is independent of χ. Independence of K follows from the fact that , which does not depend on the choice of K since K is always equal to the heat kernel near the origin.
The existence of with the above properties may appear as a bit of a miracle. Indeed, the fact that and converge to finite limits was easy to see because (G*G)(t, ·) = tG(t, ·) is a bounded function for the heat kernel G in 2D. On the other hand, the fact that converges to a finite limit, and, thus, that exists, is not a priori obvious because it relies on a cancellation between diverging BPHZ constants and in Theorem 2.5. The fact that is furthermore independent of χ relies on a cancellation between and and that (1.5) with bare mass is independent of K relies on the expression for .
III. THREE DIMENSIONS
We now discuss the main results of Ref. 12, which deals with the 3D theory.
A. State space
The construction of the state space in Ref. 12 proceeds in two steps. The first step is to define a space of initial conditions for a gauge-covariant regularizing operator. Abstractly, we will find a metric space of distributional 1-forms and a family of operators , such that
- for smooth A, B,(3.1)
is continuous for every t > 0.
If we can find such and , then we can extend gauge equivalence ∼ to by using (3.1) as a definition. Finally, we want to be sufficiently large to contain distributions as rough as the GFF.
The second step in the construction of is to augment with an additional norm, which ensures that a form of the bound (2.5) holds. This turns out to be critical in several places of the construction for the Markov process in Theorem 1.11(b).
We mention that the idea to use the YM flow to define a suitable space of distributional 1-forms was already suggested in Ref. 15 (see also Refs. 22, 30, 49, and 55 for related ideas in physics). We also point out that another state space that bears close similarity to was independently defined in Refs. 9 and 10 and was shown to support the GFF.
1. The first half
Since we can take δ < 1 in (3.5) [vs in (3.4)], this improved regularity ends up being enough to show that every Picard iterate of is well-defined when a is the GFF. It is, therefore, natural to make the following definition.
The space can be identified with a subset of because the map has a closed graph.
A standard argument with Young’s product theorem and estimates of the type (3.3) shows that the YM flow extends to in the following sense.
For every ball in centered at 0, there exists T > 0 such that for all t ∈ (0, T), the YM flow (3.2) extends to a continuous function (which is Lipschitz for any norm on C∞).
2. The second half
We now refine the space in order to obtain control on gauge transformations of the type (2.5), which proves crucial in the construction of the Markov process on gauge orbits associated with (1.5).
To motivate the norm |||·|||α,θ, recall that the estimate (2.5) relies on the identity (2.6), which, in turn, requires that line integrals and holonomies of A are well-defined. However, we saw that the GFF A cannot even be restricted to lines.
The metric space can be identified with a subset of and comes with the parameters (η, β, δ, α, θ), the possible range of which is given in Ref. 12 (Sec. 5); is the space appearing in Theorem 1.3 for d = 3.
We can now state Ref. 12 (Theorem 2.39), which is one of the main results of (Ref. 12, Sec. 2) and the motivation behind the norm |||·|||α,θ.
The Proof of Theorem 3.7 relies on two estimates: (i) the estimate (2.5) used in the 2D case (which, of course, holds in arbitrary dimension) and (ii) a “backward estimate” that controls the initial condition of a parabolic PDE in terms of its behavior for positive times [see Ref. 12, Lemma 2.46(b)]—this estimate is applied to the harmonic map flow-type PDE solved by h for which h(0) = g and for all t > 0. Theorem 3.7 is then obtained by suitably interpolating between estimates (i) and (ii).
One can show that mollifications of the SHE converge in probability in the space (see Ref. 12, Corollary 3.14). In particular, the SHE admits a modification with sample paths in .
Unlike in 2D, the action of on is not transitive over the orbits, and it is unclear if ∼, or some variant of it, is determined by the action of a group. This lack of a gauge group is responsible for the gap in our understanding of the quotient space in 3D vs 2D; see Remark 1.5.
B. Local solutions
We next explain how one proves Theorem 1.6 in 3D, which is done in Ref. 12 (Sec. 5). We do not restate the result here like we did in Sec. II B since we cannot make it substantially more precise.
Though primarily using the theory of regularity structures as before, there are two main additional challenges on top of the 2D case. The first is purely algebraic and concerns showing that the renormalization counterterms are of the form . The difficulty is that there are dozens of trees that potentially contribute to renormalization (vs just nine trees in 2D; see Ref. 11, Sec. 6.2.3).
By power counting, one can deduce that the renormalization is linear in A. To argue that one sees precisely requires a systematic approach to symmetry arguments, which is developed in Ref. 12 (Sec. 4) and which could of independent interest in other contexts.
To give an example of how this works, we argue that the renormalization is “block diagonal,” i.e., if the counterterm cAj appears in the Ai equation for j ≠ i, then c = 0. Indeed, if we flip the coordinate xi↦ −xi and, thus, ∂i↦ −∂i, together with Ai↦ −Ai and , while keeping all terms with indices j ≠ i the same, it is immediate that all the terms in the Ai equation (1.3) flip sign. Using the symmetry of the noise and invariance under the flip xi↦ −xi of the kernel K used to define , one can show that renormalized equation must possess the same symmetry, namely, all terms in the renormalized Ai equation must flip sign. Since we kept Aj the same, this shows that any factor cAj in the renormalized Ai equation must have c = 0.
The way one argues that the same appears for all i ∈ {1, 2, 3} and that is similar: one exploits symmetry under reflections xi ↔ xj and for the former and symmetry under constant gauge transformations Ai ↦ AdgAi and , where g ∈ G for the latter.
The second challenge is analytic and comes from the singularity of the initial condition in for κ > 0 (this was already encountered in 2D in a more mild form; see Remark 2.8). This singularity means, for example, that is ill-defined for generic distributions and . Similar to the discussion in Sec. III A 1, this type of product appears in the Picard iteration used to solve (1.5). As in Remark 2.8, this issue is addressed by decomposing , where Ψ solves the SHE ∂tΨ = ΔΨ + ξɛ, and solving for the “remainder” . One then shows with separate stochastic bounds that and converge in C−2−κ(R × T3) as ɛ ↓ 0.
A closely related issue not present in 2D is that of restarting the equation at some positive time τ > 0 to obtain maximal solutions. This is because, for ɛ > 0, A(τ) and ξɛ↾[τ,∞) see each other on a time interval of order ɛ2. Since the regularities of A(τ) and ∂Ψ add up to , this breaks the argument used to show that and converge as ɛ ↓ 0 when A(0) is independent of ξ. To restart the equation, one instead leverages that A(τ) for τ > 0 is not a generic element of but takes the form A(τ) = Ψ(τ) + R(τ), where R(τ) ∈ C−κ(T3). Since Ψ is defined globally in time, this decomposition allows one to restart the equation using the “generalized Da Prato–Debussche trick” from Ref. 6.
C. Gauge covariance
Finally, we describe the Proof of Theorem 1.11 in 3D. The Proof of Theorem 1.11(a) is similar to its 2D counterpart. The only appreciable difference is that the measurable selection is replaced by a Borel map , which preserves gauge orbits and such that whenever the right-hand side is finite. Here, is defined analogously to Σ but with a stronger set of parameters (η, β, δ, α, θ). This complication is due to a lack of any nice known properties of in 3D (e.g., polishness) (see Remarks 1.5 and 3.9), and we instead leverage compactness of the embedding .
The Proof of Theorem 1.11(b) is where we start to see a difference with the 2D case. First, admitting for now Theorem 1.14, we aim to prove that solutions to (1.5) with bare mass and gauge equivalent initial conditions can be suitably coupled. Specifically, one has the following result.
(Coupling). Suppose A solves (1.5) with bare mass and initial condition a. Then, for any b ∼ a, there exists on the same probability space a white noise and a process (B, g) such that and A(t)g(t) = B(t) for all t > 0 (before blow-up of A, B) and such that B solves (1.5) driven by with bare mass .
If ag(0) = b for some for as in Sec. III A 2, then this result follows almost immediately from Theorem 1.14. However, unlike the 2D case, it is now possible that b ∼ a, but no g(0) exists such that ag(0) = b, which leads to trouble in applying Theorem 1.14—we effectively have no initial condition for g in the PDE (1.8).
With Lemma 3.10 in hand, together with the fact that g in its statement cannot blow up before Σ(A, 0) + Σ(B, 0) blows up (again due to Theorem 3.7), it is relatively straightforward to prove Theorem 1.11(b) like we did in the 2D case. See, in particular, the discussion around Fig. 1.
Both of these facts are again proven by introducing the new variables h = (dg)g−1 and U = Adg and writing the corresponding equations for h, U. This time, instead of a direction computation with just two trees as in (2.11), a more complicated strategy relying on power counting and symmetry arguments is necessary. The main insight, which helps with this argument, is that the trees appearing in the “lifted” B and equations are obtained by attaching (or grafting) the trees appearing in the U equation onto the trees appearing in (1.5) (see Ref. 12, Sec. 6.2.1).
The analytic theory for B and equations is somewhat more involved than what we saw in Secs. II C and III B, but the general strategy is the same: we decompose the solution into the initial condition, globally defined singular terms, and a better behaved remainder and solve for the last of these. Restarting these equations, specifically the equation for , is also not straightforward because, unlike in Sec. III B, we are outside the scope of the generalized Da Prato–Debussche trick of Ref. 6 since the multiplicative noise means that the most singular part of the solution is not just the SHE. Therefore, an entirely separate fixed point problem needs to be written for the restarted equation (see Ref. 12, Sec. 6.6), which leverages that the new initial condition comes from a modeled distribution defined for earlier times. The proof that B and converge to the same limit uses the same ɛ-dependent norms on regularity structures introduced in Ref. 11.
The bounds (3.10) are actually used in the short time analysis of (3.7) and (3.8) since they allow us to relate B and to a simpler equation with additive noise through the above mechanism (B is related through pathwise gauge transformations and is related through equality in law; see Ref. 12, Sec. 6.6).
Finally, to prove that exists and is independent of the non-anticipative mollifier χ, one argues in a similar way except, as earlier, we appeal to equality with (1.5) in law and use instead that two limiting solutions with different bare masses cannot be equal in law.
These final statements mimic exactly what we saw in Sec. II C where for non-anticipative χ and is given by (2.13). By analogy, (3.15) should hold for any mollifier, not necessarily non-anticipative, but this is not necessarily the case for . Furthermore, as in 2D, we expect that for non-anticipative χ.
IV. OPEN PROBLEMS
We close with several open problems that we believe to be of interest.
Does the Markov process on gauge orbits in Theorem 1.11 possess a unique invariant measure? The existence of the invariant measure should imply uniqueness due to the strong Feller property38 and full support theorem for SPDEs.39 Furthermore, the invariant measure for d = 2 is expected to be the YM measure associated with the trivial principal G-bundle on T2 constructed in Refs. 44, 45, and 60. For d = 3, this would provide the first construction of the YM measure in 3D, even in finite volume.
Can the analysis in Refs. 11 and 12 be extended to infinite volume Td ⇝ Rd? This is non-trivial even for d = 2, although the YM measure on R2 is arguably simpler.26
Can one extend these results beyond the case that the underlying principal bundle P → Td is trivial? For non-trivial principal bundles, one can no longer write connections as globally defined 1-forms, which complicates the solution theory.
For d = 3, can one modify the construction of the state space in Sec. III A so that the gauge equivalence ∼ is determined by a gauge group or a similar structure? This would yield a notion of gauge equivalence conceptually closer to the classical space of gauge orbits and would carry a number of technical advantages (see Remarks 1.5 and 3.9 and the start of Sec. III C).
Taking G to be one of the classical groups, e.g., G = U(N), what is the behavior of the dynamic as N → ∞? In 2D, the associated YM measure is known to converge to a deterministic object called the master field,19–21,46 which is governed by the Makeenko–Migdal equations;52 see Ref. 47 for a survey. No such result is rigorously known in 3D (the measure at finite N has not been constructed). It would be interesting if one can use stochastic quantization to recover some of the known results in 2D and obtain new results in 3D; see Ref. 62 where the Langevin dynamic is used to derive the finite N master loop equation on the lattice.
ACKNOWLEDGMENTS
It is a great pleasure to thank Ajay Chandra, Martin Hairer, and Hao Shen for many interesting discussions and explorations while carrying out the reviewed works.
AUTHOR DECLARATIONS
Conflict of Interest
The author has no conflicts to disclose.
Author Contributions
Ilya Chevyrev: Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.