Geometric derivation of the finite master loop equation
Abstract.
In this paper we provide a geometric derivation of the master loop equation for the lattice Yang-Mills model with structure group . This approach is based on integration by parts on . In the appendix we compare our approach to that of [Cha19b] and [Jaf16] based on Schwinger-Dyson equations, and [SSZ22] based on stochastic analysis. In particular these approaches are all easily seen to be equivalent. The novelty in our approach is the use of intrinsic geometry of which we believe simplifies the derivation.
1. Introduction
The study of a Quantum Yang-Mills theory is of great importance for understanding the standard model of particle physics. Although the physically relevant theory takes place in Minkowski space, standard arguments indicate that it is sufficient to study the Euclidean/probabilistic theory. This paper will only deal with the Euclidean theory, for a general survey paper see [Cha19c].
Specifically, we will be dealing with the the lattice Yang-Mills model with structure group . One approach to ’solving’ the lattice Yang-Mills model is through computing Wilson loop expectations. Wilson loop expectations are known to satisfy an equation known as a master loop equation.
Master loop equations for lattice gauge theories were first published by Makeenko and Migdal [MM79] in 1979. However this derivation assumed a ’factorization property’ without proof. In recent years there have been several rigorous derivations of master loop equations for classical groups such as . For instance see [Cha19b], [Jaf16], [SSZ22], [PPSY23], and [CPS23].
The purpose of this paper is to provide a geometric derivation of the master loop equation for the lattice Yang-Mills model with structure group . This approach is based on integration by parts on , and is inspired by the derivations in [Cha19b] and [Jaf16]. In particular, the approaches of [Cha19b] and [Jaf16] rely on Schwinger-Dyson equations which they obtain from Stein’s idea of exchangeable pairs. As remarked in a talk of Chatterjee [Cha19a] (pointed out by Thierry Lévy), these Schwinger-Dyson equations are simply integration by parts on written in extrinsic coordinates. We show the equivalence of integration by parts and the Schwinger-Dyson equation in Appendix A.1 and Appendix B, and generalize this to any compact Lie group with Riemannian structure inducing the Haar measure. The main sections of the paper are devoted to deriving the master loop equation for directly from integration by parts, relying on the intrinsic geometry of in order to simplify the calculations of Chatterjee and Jafarov. Lastly in Appendix A.2 we compare our approach to the stochastic analyis approach in [SSZ22].
Acknowledgements
We thank Scott Sheffield and Sky Cao for many useful discussions.
2. Definitions and notation
2.1. Lattice gauge theory
In the following, is a finite two dimensional cell complex.
Definition 2.1.
A path is a sequence of oriented -cells such that their union is connected in . The set of all paths forms a groupoid , where two paths and may be concatenated if
The idea of a lattice gauge theory is to discretize a notion of a connection on a principal -bundle. By restricting paths to lie in a discrete set of -cells, we may identify a connection with its finite set of parallel transport maps. This leads to the following:
Definition 2.2.
Let be as before. A -connection or -gauge field is a homomorphism
. In other words, if is a path from to , and is a path from to , then
And
It’s clear from the definition that such a homomorphism is determined by its values on an oriented edge. Pick an arbitrary orientation for each edge. Let the set of such edges be denoted . Then we have
This characterization will be useful in defining lattice Yang-Mills measures.
Definition 2.3.
A path is called a loop if its image has empty boundary.
The most important observables in a lattice gauge theory are the Wilson loops.
Definition 2.4.
Let be an irreducible character on , and let be a loop. A Wilson loop is a functional of the form
Remark 2.1.
For the rest of the paper we will only consider Wilson loops defined with the character for , and for . The proofs do not immediately generalize to other irreducible characters.
Definition 2.5.
Let be any class function (for our purposes usually an irreducible character). A plaquette is a -cell in . Let the set of faces be denoted . Pick an arbitrary orientation, and let denote the set of positively oriented plaquettes. The lattice Yang-Mills measure with ’t Hooft coupling is the probability measure
Where is the Haar measure.
2.2. Path operations
The following is a list of operations on loops that will be relevant in defining the master loop equations. The terminology and notation is based on that of [Cha19b]. The master loop equation is based around integration by parts on an edge in . Thus, for every path, We define a set indexing the occurrences of in . Moreover, if , then is the orientation of that instance of .
Definition 2.6.
Let be a loop. We define
In other words, this is the string formed by excising , ordered starting from the edge after .
Definition 2.7 (Positive merger).
Let and be loops.
Definition 2.8 (Negative merger).
Let and be loops.
Definition 2.9 (Positive split).
Let be a loop such that . If and , the positive split at is the pair of loops
Definition 2.10 (Negative split).
Let be a loop such that . If and , the positive split at is the pair of loops
Definition 2.11 (Positive twist).
Let be a loop such that . WLOG suppose and . If and , the positive twist of at is the loop
Definition 2.12 (Negative twist).
Let be a loop such that . Let and . WLOG let . Then the negative twist is
2.3. definitions and conventions
Recall the following:
The tangent space at is the space of matrices
There is a natural bi-invariant metric on :
Definition 2.13.
Let . Then
(Note that this is the restriction of half the Euclidean metric to )
With respect to this metric,
is an orthonormal frame on .
For lattice gauge theory, we use the lattice Yang-Mills measure defined by the character .
2.4. and definitions and conventions
Recall that
The tangent space at is the space of matrices
There is a natural bi-invariant metric on :
Definition 2.14.
Let or . Then
(Note that this is the restriction of half the Euclidean metric to )
For lattice gauge theory, we use the lattice Yang-mills measure defined by the class function
The case is complicated in that the character associated with the defining representation is now complex valued. Thus we need to make sense of gradients of functions
Definition 2.15.
Let . Define by
In other words, we extend the gradient in the natural way to a complex-linear map of smooth functions.
Similarly, we extend the metric.
Definition 2.16.
Let . Then
With this definition in mind, we can formulate the basis of the proof of the main theorem:
Lemma 2.1 (Laplacian integration by parts).
Let , with a compact Lie group. Equip with a bi-invariant metric. With respect to this metric,
In the sequel we will show that when and are Wilson loop correlation functions, laplacian integration by parts directly reduces to the master loop equation.
3. Main Results
The first main result is the master loop equation for
Theorem 3.1 ( master loop equation).
Let be a sequence of loops. Let be an edge that lies in at least one of . Let denote expectations with respect to the lattice yang mills measure. Let denote the set of positively oriented plaquettes containing . Finally, let be the set of occurrences of the edge in , the set of occurrences of , and . Then
(1) |
Theorem 3.2 ( and master loop equation).
Let be a sequence of loops. Let be an edge that lies in at least one of . Let for and when . Let , , , be as before. In addition, let be the set of plaquettes containing . Finally, let and . Then
4. analysis
The goal of this section is to prove theorem 3.1
4.1. Gradient identities for Wilson loops
Lemma 4.1.
Let be a Wilson loop. Let denote the set of occurrences of the edge in . Let denote the gradient with respect to the edge . Then
(2) |
Proof.
extends to a smooth function in a neighborhood of in the obvious way. We may thus first compute the euclidean differential:
For , we may rewrite . By orthogonality,
Thus, the euclidean differential is
Recall that the gradient is defined by the identity . Therefore, the euclidean gradient is
Recall that the tangent projection onto is
Setting , we finally arrive at the result:
∎
Lemma 4.2.
Let and be Wilson loops.
(3) |
Proof.
∎
4.2. Laplacian of Wilson loops
Lemma 4.3.
Let denote the tangent projection. Let and denote left and right multiplication, respectively. Then
(4) |
Proof.
∎
Lemma 4.4.
Let be a Wilson loop and the Laplace-Beltrami operator at the edge . Let be the number of occurrences of in . Then
(5) |
Proof.
Recall that for a vector field , where is the covariant derivative of . On a submanifold, the covariant derivative is the tangent projection of the ambient covariant derivative. Thus it suffices to compute
Where is the euclidean covariant derivative. Now, by lemma 4.1:
Now expanding further requires casework. In particular, in the first term, the th occurrence of is if and is otherwise. The opposite is true for the third term. Recall that
Let denote the linear map formed by substituting for the th occurrence of . We thus have
(6) |
We can now apply lemma 4.3, as every term in this sum is an operator of the form . We can first consider
We can write . Then and so the trace is
Similarly, for the case , . So we compute the trace of Which equals
4.3. master loop equation
Proof of theorem 3.1..
Let be a sequence of loops. We apply Laplace integration by parts to the Wilson loop correlation function:
On the left hand side, we have
Setting the two sides equal and rearranging terms gives the result:
This concludes the proof of the master loop equation. ∎
5. and analysis
We now move on to the proof of theorem 3.2
The procedure is mostly analogous. However unlike , not every element of these groups is conjugate to its inverse. Thus, the orientations of Wilson loops will become more relevant in the analysis. This is reflected in the fact that Wilson loops are now complex valued. We will introduce a parameter that vanishes when and is when .
5.1. Gradients of Wilson loops
Lemma 5.1.
Let be a loop.
(7) |
Proof.
As before, the differential is
Thus,
Recalling that the tangent projection is
We thus get
∎
Lemma 5.2.
Let be a loop.
Proof.
Recall that . Note that
Thus,
and the euclidean gradient is
Applying the tangent projection again finally gives for
∎
Lemma 5.3.
Let and be loops. Then
Proof.
Note that because , the first part of the gradient has algebra identical to that of the case. Thus
The first term after the mergers equals
Similarly, the second term after mergers is
Finally, the last term is
In total,
Now for . The first term is almost algebraically identical (the s cancel and s factor out). But the negative sign is gone. So
In total then we have
Now for the imaginary part.
Recall that in the definition of or , the orientation of the first term does not change. As a result,
Now for the remaining terms,
Similarly,
And finally,
In total,
By symmetry, is the same as with and . So
Now we need to examine the relationship between and and similarly for . For a negative merger, an orientation reversal occurs if . Otherwise it doesn’t happen. Similarly for a positive merger, an orientation reversal happens if . Thus we get
And so,
Thus we can conclude:
Note that
Completing the proof. ∎
This final lemma is required to account for terms from the measure.
Lemma 5.4.
Let and be loops. Then
Proof.
Most of the algebra has already been carried out. This inner product is
The first term is, as we’ve computed before,
The next term is
Putting them together, this gives
That last term can be simplified:
So in total,
∎
5.2. Laplacian of Wilson loops
Lemma 5.5.
Let and denote left and right-multiplication, respectively. Then
Proof.
We first compute the trace on . An orthonormal basis is , , , and .
Now, . Thus the trace on is
∎
Lemma 5.6.
Let be a Wilson loop and the Laplace-Beltrami operator at the edge . Then
Proof.
The laplacian does not involve any complex multiplication. We can therefore separately compute and .
so
Now, recall:
Taking the differential and accounting for orientations as in the proof,
For that last term,
Putting it all together,
The last two terms can be dropped, as both of their images lie in the normal bundle.
Next, we have
Next,
Next,
Finally,
Inserting these identities back in,
The last term can be simplified as follows: and So that term reduces to
where and . are the number of s and s respectively. So in total,
Now for the imaginary part of the laplacian, notice that no part of the algebra actually used that the matrices between occurrences of the edge belong to . We only needed that they are unitary. Thus, pick any edge that is not , and replace the matrix there with . This gives
And so, we have
∎
5.3. Master loop equation for and
Proof of theorem 3.2.
References
- [Cha19a] Sourav Chatterjee. Gauge-string duality in lattice gauge theories, Jan 2019.
- [Cha19b] Sourav Chatterjee. Rigorous solution of strongly coupled lattice gauge theory in the large limit. Communications in Mathematical Physics, 366(1):203–268, 2019.
- [Cha19c] Sourav Chatterjee. Yang–mills for probabilists. In Probability and Analysis in Interacting Physical Systems, pages 1–16, Cham, 2019. Springer International Publishing.
- [CPS23] Sky Cao, Minjae Park, and Scott Sheffield. Random surfaces and lattice yang-mills. arxiv preprint arXiv:2307.06790, 2023.
- [Hsu06] Elton Hsu. Stochastic analysis on manifolds. American Mathematical Society, 2006.
- [Jaf16] Jafar Jafarov. Wilson loop expectations in lattice gauge theory. arxiv preprint arXiv:1610.03821, 2016.
- [KS99] Ioannis Karatzas and Steven E. Shreve. Brownian motion and stochastic calculus. Springer, 1999.
- [MM79] Yu. M. Makeenko and Alexander A. Migdal. Exact Equation for the Loop Average in Multicolor QCD. Phys. Lett. B, 88:135, 1979. [Erratum: Phys.Lett.B 89, 437 (1980)].
- [PPSY23] Minjae Park, Joshua Pfeffer, Scott Sheffield, and Pu Yu. Wilson loop expectations as sums over surfaces on the plane. arxiv preprint arXiv:2305.02306, 2023.
- [SSZ22] Hao Shen, Scott A. Smith, and Rongchan Zhu. A new derivation of the finite master loop equation for lattice yang-mills. arxiv preprint arXiv:2202.00880, 2022.
Omar Abdelghani, Department of Mathematics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA.
E-mail address: [email protected]
Ron Nissim, Department of Mathematics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA.
E-mail address: [email protected]
Appendix A Comments on Other Approaches
A.1. Integration by parts and exchangeable pairs
In this section we first show that the Schwinger Dyson equation in [Cha19b] is simply integration by parts on written in extrinsic coordinates by obtaining integration by parts from Stein’s method of exchangeable pairs. We then generalize this to all compact Lie groups with Riemannian structure inducing the Haar measure.
Recall
Definition A.1.
A pair of random variables is called an exchangeable pair if in distribution.
We have the following elementary lemma for exchangeable pairs,
Lemma A.1.
Let be a exchangeable pair and real valued Borel measurable functions. Then
(8) |
Proof.
See Lemma 6.1 from [Cha19b] ∎
Recall that has a Riemannian metric given by the inner product on . Thus has an orthonormal basis for . These basis vectors are simply the tangent vectors generated by rotations in the plane spanned by two basis vectors.
So working with the rotations at the identity, it follows that the Laplace Beltrami operator of of is simply given by
So
(9) |
where the expectation is taken over uniformly chosen from and uniformly from . Hence if is another smooth function we have
(10) |
where the expectation is now taken with respect to the Haar measure on , and . Since is compact and , the interchange of limit and expectation above follows from the bounded convergence theorem.
Similarly for
(11) |
Finally in [Cha19b], it is shown that form an exchangeable pair, thus combining (10) and (11), integration by parts on ,
follows as consequence of lemma 8. Moreover since lemma 8 with this Stein pair is Chatterjee’s starting point for deriving his Schwinger-Dyson equation (Theorem 7.1 [Cha19b]), we see that it is simply integration by parts written in extrinsic coordinates.
This derivation of integration by parts actually generalizes to any compact Lie group with Riemannian metric inducing the Haar measure. For such a Lie group , let be an orthonormal basis for the corresponding Lie algebra . We can write where . Now let be an element of selected from the Haar measure, and where is Bernoulli with and is uniform on all chose independently of each other.
Lemma A.2.
is an exchangeable pair
Proof.
By left invariance of the Haar measure, is also distributed according to the Haar measure. Moreover so in distribution since in distribution. ∎
A.2. Symmetrized Master Loop Equation and Langevin Dynamics
One immediate corollary of Theorem 3.1 and Theorem 3.2 is the symmetrized master loop equation (Theorem 1 [SSZ22]).
This corollary is derived in [SSZ22] through studying the following Langevin dynamics with invariant Yang-Mill’s measure.
(12) |
where is the Yang-Mills action, and is the intrinsic gradient on .
It can then be shown via integration by parts that the Yang-Mills measure is invariant under the dynamics (12), (Lemma 3.3 [SSZ22]).
The rest of the proof proceeds by applying Itô’s formula in to with evolving according to starting with the Yang-Mills measure as the initial distribution, and taking the expectation of both sides with respect the Yang-Mills measure.
is of course a semimartingale [KS99], and can thus be written in the form where is a martingale, and is a bounded variation process with . The equation [SSZ22] obtain is . But standard theory for stochastic analysis on manifolds (Chapter 3, [Hsu06]) tells us that the Langevin dynamics has infinitesimal generator , so . Thus the symmetrized master loop equation derived in [SSZ22] is equivalent to the following integration by parts
(13) |
where and are the intrinsic gradient and Laplace-Beltrami operator on respectively.
Appendix B Deriving the extrinsic integration by parts formula
Again let in the natural way.
Let be open, and smooth functions on . Then we can extend the integration by parts formula to functions .
Let be a vector field on . Note that for consistency we’re equipping with the metric . Thus,
the gradient for this metric is twice the usual gradient.
Recall that
so (because is an orthonormal frame on )
And so
Because this quantity is symmetric in and ,
We can simplify the form of the sums by noting that terms vanish anyway:
Then, by orthogonality the first term simplifies further
Next,
Noting that
Similarly,
Again using
Subtracting the two,
For the first term, by the symmetry of the summand in ,
For the second term, it can be regrouped into
For the first term,
This simplifies to
By orthogonality,
What remains is the term
This is quite nice because we can see that the unfamiliar term cancels. In summary then,
Now, recalling the integration by parts on :
On the LHS then, we have
And on the RHS we have
Rearranging,
Which is precisely the Schwinger dyson equation.