We provide a visual and intuitive introduction to effectively calculating in 2-groups along with explicit examples coming from non-Abelian 1- and 2-form gauge theory. In particular, we utilize string diagrams, tools similar to tensor networks, to compute the parallel transport along a surface using approximations on a lattice. We prove a convergence theorem for the surface transport in the continuum limit. Locality is used to define infinitesimal parallel transport, and two-dimensional algebra is used to derive finite versions along arbitrary surfaces with sufficient orientation data. The correct surface ordering is dictated by two-dimensional algebra and leads to an interesting diagrammatic picture for gauge fields interacting with particles and strings on a lattice. The surface ordering is inherently complicated, but we prove a simplification theorem confirming earlier results of Schreiber and Waldorf. Assuming little background, we present a simple way to understand some abstract concepts of higher category theory. In doing so, we review all the necessary categorical concepts from the tensor network point of view as well as many aspects of higher gauge theory.
I. INTRODUCTION
We use string diagrams to express many concepts in gauge theory in the broader context of two-dimensional algebra. By two-dimensional algebra, we mean the manipulation of algebraic quantities along surfaces. Such manipulations are dictated by 2-category theory and we include a thorough and visual introduction to 2-categories based on string diagrams. Such string diagrams, including their close relatives known as tensor networks, have been found to provide exceptionally clear interpretations in areas such as open quantum systems,1 foundations of quantum mechanics,2 entanglement entropy,3 and braiding statistics in topological condensed matter theory,4 to name a few.
We postulate simple rules for associating algebraic data to surfaces with boundary and use the rules of two-dimensional algebra to derive non-Abelian surface transport from infinitesimal pieces arising from a triangulation/cubulation of the surface. One of the novelties in this work is an analytic proof for the convergence of surface transport together with a more direct derivation of the iterated surface integral than what appears in Ref. 5, for instance. To be as self-contained as possible, we include discussions on gauge transformations, orientation data on surfaces, and a two-dimensional calculation of a Wilson cube deriving the curvature 3-form. We also review ordinary transport for particles to make the transition from one-dimensional algebra to two-dimensional algebra less mysterious.
Ordinary algebra, matrix multiplication, group theory, etc., are special cases of one-dimensional algebra in the sense that they can all be described by ordinary category theory. For example, a group is a type of category that consists of only a single object. Thanks to the advent of higher category theory, beginning with the work of Bénabou on 2-categories,6 it has been possible to conceive of a general framework for manipulating algebraic quantities in higher dimensions. In particular, monoidal categories and the string diagrams associated with them7 can be viewed as 2-categories with a single object. The special case of this where all algebraic quantities have inverses are known as 2-groups, with a simple review given in Ref. 8 and a more thorough investigation in Ref. 9. We do not expect that the reader is knowledgeable of these definitions and we only assume that the reader knows about Lie groups (even a heuristic knowledge will suffice since our formulas will be expressed for matrix groups).
While there already exist several articles,8,10,11,12 introducing the “conceptual” basic ideas of higher gauge theory and parallel transport for strings in terms of category theory and even a book by Schreiber describing the mathematical framework of higher-form gauge theories,13 there are few articles that provide explicit and computationally effective methods for calculating such parallel transport.14 Although Girelli and Pfeiffer explain many ideas, most results useful for computations are infinitesimal and it is not clear how to build local quantities from the infinitesimal ones.11 Baez and Schreiber15 focus on similar aspects as we do in this article, but our presentation is significantly simplified since we assume certain results on path spaces without further discussion, such as relationships between differential forms on a manifold and smooth functions on its path space, and therefore do not deal with the delicate analytical issues on such path spaces. Our goal is to provide tools and visualizations to perform more intuitive calculations involving mainly calculus and matrix algebra.
A. Some background and history
In 1973, Kalb and Ramond first introduced the idea of coupling classical Abelian gauge fields to strings.16 Actions for interacting charged strings were written down together with equations of motions for both the fields and the strings themselves. Furthermore, some of the quantization of the theory was discussed. The next big step took place in 1985 with the work of Teitelboim (aka Bunster) and Henneaux, who introduced higher form Abelian gauge fields that could couple to higher-dimensional manifolds.17,18 In Ref. 17, Teitelboim studied the generalization of parallel transport for higher dimensional surfaces and concluded that non-Abelian p-form gauge fields for p ≥ 2 cannot be coupled to p-dimensional manifolds in order to construct parallel transport. The conclusion was that the only possibilities for string interactions involved Abelian gauge fields. As a result, it seemed that only a few tried to get around this in the early 1980s. For example, the non-Abelian Stokes theorem came from analyzing these issues in the context of Yang-Mills theories and confinement19 (see also, for instance, Sec. 5.3 of Ref. 20). Although such calculations led people to believe defining non-Abelian surface parallel transport is possible, the expressions were not invariant under reparametrizations and they did not seem well-controlled under gauge transformations. Without a different perspective, interest in it seemed to fade.
The crux of the argument of Teitelboim is related to the fact that higher homotopy groups are Abelian. This is sometimes also known as the Eckmann-Hilton argument.8 However, Whitehead in 1949 realized that higher “relative” homotopy groups can be described by “non-Abelian” groups.21 In fact, it was Whitehead who introduced the concept of a crossed module to describe homotopy 2-types. This work was in the area of algebraic topology and the connection between crossed modules and higher groups was not made until much later. A review of this is given in Ref. 8. Eventually, non-Abelian generalizations of parallel transport for surfaces were made by Girelli and Pfeiffer using category theory and ideas from homotopy theory stressing that one should also associate differential form data with lower-dimensional submanifolds.11 Before this, most of the work on non-Abelian forms associated with higher-dimensional objects did not discuss parallel transport but developed the combinatorial and cocycle data building on the foundational work of Breen and Messing.22,10,23 This cocycle perspective eventually led to the field of non-Abelian differential cohomology.13,24,25 The idea of decorating lower-dimensional manifolds is consistent with the explicit locality exhibited in the extended functorial field theory approach to axiomatizing quantum field theories.26,27,28,29 Recently, in a series of four papers, Schreiber and Waldorf axiomatized parallel transport along curves and surfaces,30,5,31,12 building on an earlier work of Caetano and Picken.32
B. Motivation
We have already indicated one of the motivations of pursuing an understanding of parallel transport along surfaces, namely, in the context of string theory. Strings can be charged under non-Abelian groups and interact via non-Abelian differential forms. Just as parallel transport can be used to described non-perturbative effects in ordinary gauge theories for particles, parallel transport along higher-dimensional surfaces might be used to describe non-perturbative effects in string theory and M-theory. Yet another use of parallel transport is in the context of lattice gauge theory where it is used to construct actions whose continuum limit approaches Yang-Mills type actions.33
Higher form symmetries have also been of recent interest in high energy physics and condensed matter in the exploration of surface operators and charges for higher-dimensional excitations.34 However, the forms in the latter are strictly Abelian and the proper mathematical framework for describing them is provided by “Abelian” gerbes (also known as higher bundles)35,36 and Deligne cohomology. Higher non-Abelian forms appear in many other contexts in physics, such as in a stack of D-branes in string theory,37 in the ABJM model,38 and in the quantum field theory on the M5-brane.39 In fact, the authors of Ref. 38 show how higher gauge theories provide a unified framework for describing certain M-brane models and how the 3-algebras of Ref. 40 can be described in this framework. Further work, including an explicit action for modeling M5-branes, was provided recently in Ref. 41.
Although a description of the non-Abelian forms themselves is described by higher differential cohomology,13 parallel transport seems to require additional flatness conditions on these forms.8,15,11,10,13,42 For example, in the special case of surfaces, this condition is known as the vanishing of the fake curvature. Some argue that this condition should be dropped and the existence of parallel transport is not as important for such theories.43 However, our perspective is to take this condition seriously and work out some of its consequences. Indeed, since higher-dimensional objects can be charged in many physical models besides just string theory, parallel transport might be used to study non-perturbative or effective aspects of these theories, an important tool to understand quantization (see the discussion at the end of Ref. 44). Because it is not yet known how to avoid these flatness conditions, further investigation is necessary, with some recent progress by Waldorf.25,42
Therefore, because of the subject’s infancy, it is a good idea to devote some time to “calculating” surface transport explicitly to better understand how branes of different dimensions can be charged under various gauge groups. Here, we focus on the case of two-dimensional surfaces such as strings or D1-branes. However, we make no explicit reference to any known physical models. For these, the reader is referred to the other studies in the literature, for example, Ref. 41 and the references therein.
C. Outline
In Sec. II, we describe how categorical ideas can be used to express a mix of algebraic and geometric concepts. In Sec. II A, we review “string diagrams” for ordinary categories and how group theory arises as a special case of ordinary category theory. In Sec. II B, we define 2-categories and other relevant structures, providing a two-dimensional visualization of the algebraic quantities in terms of string diagrams. In Sec. II C, we specialize to the case where the algebraic data are invertible. We restrict attention to strict 2-groups, which is sufficient for many interesting applications.34,45,38,41,46
In Sec. III, we describe how gauge theory for 0-dimensional objects (particles) and 1-dimensional objects (strings) can be expressed conveniently in the language of two-dimensional algebra. In detail, in Sec. III A, we review how classical gauge theory for particles is described categorically. We include a review of the formula for parallel transport describing it in terms of one-dimensional algebra as an iterated integral obtained from a lattice discretization and a limiting procedure. In Sec. III B, we include several crucial calculations for gauge theory for 1-dimensional objects (strings) expressing everything in terms of two-dimensional algebra. In particular, we derive the local infinitesimal data of a higher gauge theory. To the best of our knowledge, these ideas seem to have first been analyzed in Refs. 22, 11, and 15, though our inspiration for this viewpoint came from Ref. 47. Furthermore, we use the rules of two-dimensional algebra to derive an “explicit” formula for the discretized and continuous limit versions of the local parallel transport of non-Abelian gauge fields along a surface. Although such a formula appears in the literature,15,5 we provide a more intuitive derivation as well as a useful expression for lattice computations. We provide a picture for the correct surface ordering needed to describe parallel transport along surfaces with non-Abelian gauge fields in Proposition 3.5 and the discussion surrounding this new result. We then proceed to prove that the surface ordering can be dramatically simplified in Theorem 3.8. In Remark 3.9, we show that our resulting formula agrees with the one given by Schreiber and Waldorf which was obtained through different means.5 In Sec. III C, we study the gauge covariance of the earlier expressions and derive the infinitesimal counterparts in terms of differential forms. In Sec. III D, we discuss the subtle issue of orientations of surfaces and how our formalism incorporates them. In Sec. III E, we again use two-dimensional algebra to calculate a Wilson cube on a lattice and from it obtain the 3-form curvature. We then study how it changes under gauge transformations showing consistency with the results of Girelli and Pfeiffer.11
Finally, in Sec. IV, we discuss some indication as to how these ideas might be used in physical situations and indicate several open questions.
II. CATEGORICAL ALGEBRA
A. Categories as one-dimensional algebra
We will present categories in terms of what are known as “string diagrams” since we find that they are simpler to manipulate and compute with when working with 2-categories. Therefore, we will define categories, functors, and natural transformations in terms of string diagrams. Afterward, we will make a simplification and discuss special examples of categories known as groups.
A “category,” denoted by consists of
a collection of 1-d domains (also known as objects)
- (ii)
between any two 1-d domains, a collection (which could be empty) of 0-d defects (also known as morphisms)
- (iii)
an “in series” composition rule
- (iv)
between every 1-d domain and itself, a specified 0-d defect
These data must satisfy the conditions that
the composition rule is associative and
the identity 0-d defect is a left and right identity for the composition rule.
Let G be a group. From G, one can construct a category, denoted by consisting of only a single domain (say, red) and the collection of 0-d defects from that domain to itself consists of all the elements of G. The composition is group multiplication. The identity at the single domain is the identity of the group.
The previous example of a category is the one in which all 0-d defects are invertible.
Let and be two categories. A “functor” is an assignment sending 1-d domains in to 1-d domains in and 0-d defects in to 0-d defects in satisfying
the source-target matching condition
- (b)
preservation of the identity
- (c)
and preservation of the composition in series
There are several ways to think about what functors do. One, a functor can be viewed as a “construction” in the sense that one begins with data and from them constructs new data in a consistent way. Another perspective is that functors are invariants and give a way of associating information that only depends on the isomorphism class of 1-d defects. Another perspective that we will find useful in this article is to think of a functor as attaching algebraic data to geometric data. We will explore this last idea in Sec. III A and generalize it in Sec. III B. Yet another perspective is to view categories more algebraically and think of a functor as a generalization of a group homomorphism since the third condition in Definition 2.4 resembles this concept. We will explore this last perspective in in the following example:
Let G and H be two groups, and let and be their associated one-object categories as discussed in Example 2.3. Then, functors are in one-to-one correspondence with group homomorphisms f : G → H.
B. 2-Categories as two-dimensional algebra
2-categories provide one realization of manipulating algebraic data in two dimensions.
A “2-category,” denoted by , consists of
a collection of 2-d domains (also known as objects)
- (ii)
between any two 2-d domains, a collection (which could be empty) of 1-d defects (also known as 1-morphisms or domain walls)
- (iii)
between any two 1-d defects that are themselves between the same two 2-d domains, a collection (which could be empty) of 0-d defects (also known as 2-morphisms or excitations)
- (iv)
an “in parallel” composition (also known as horizontal composition) rule for 1-d defects
- (v)
an “in series” composition (aka vertical composition) rule for 0-d defects
- (vi)
an “in parallel” composition (aka horizontal composition) rule for 0-d defects
- (vii)
every 2-d domain R has both an identity 1-d defect and an identity 0-d defect
All composition rules are associative (this will be implicit in drawing the diagrams as we have).
The identities obey rules exhibiting them as identities for the compositions.
The composition in series and in parallel must satisfy the “interchange law”
These laws guarantee the well-definedness of concatenating defects in all allowed combinations.
Kitaev and Kong provide more examples of 2-categories in their discussion of domains, defects, and excitations in the context of condensed matter.49 In their language, we are viewing excitations as generalized defects.
The assignment F is such that all sources and targets are respected, i.e.,
- (b)
All identities are preserved (this is the “normalized” condition).
- (c)
For any 1-d defect f
- (d)
To every triple of parallel composable 1-d defects
If is the identity for all f and g in then F is said to be a “strict functor.”
Examples of weak functors abound. For example, projective representations are described by weak functors that are not strict functors as will be explained in the following example. Weak functors can also be used to define the local cocycle data of higher bundles.24 Since we will be working locally for simplicity, we will make little use of weak functors, but have included their discussion here for completeness and so that the standard definitions of higher bundles may be less mysterious.14,12,24 Strict functors will be used as a means of “defining” parallel transport along surfaces in gauge theory in Sec. III B. Natural transformations will be used to “define” gauge transformations of such functors and their infinitesimal counterparts will be “derived” from these definitions.
Let be two weak functors between two 2-categories. A “natural transformation” σ: F ⇒ G is an assignment sending k-d domains/defects of to (k − 1)-d defects of for k = 1, 2 satisfying the following conditions:
The assignment is such that
- (b)
To every pair of parallel composable 1-d defects
- (c)
To every identity 1-d defect idR, the equality
- (d)
To every 0-d defect
Such string diagram pictures facilitate certain kinds of computations54 (for instance, compare the definition of natural transformation in Fig. 10 of Ref. 54). Natural transformations between functors can be thought of as symmetries. For example, just as natural transformations of functors between ordinary categories can be used to describe intertwiners for ordinary representations, natural transformations of functors between 2-categories can be used to describe intertwiners of projective representations.
It will be important to compose natural transformations. This will correspond to iterating gauge transformations successively.
Technically, one should check this indeed defines a natural transformation. This is a good exercise in two-dimensional algebra. There are actually similar symmetries between natural transformations, called “modifications,” which we define for completeness.
Let be two weak functors between two 2-categories and σ, ρ: F ⇒ G be two natural transformations. A “modification” m: σ ⇛ ρ assigns to every 2-d domain of a 0-d defect in such that the following conditions hold:
C. Two-dimensional group theory
A convenient class of 2-categories are those for which there is only a single 2-d domain and all defects are invertible under all compositions. Such a 2-category is called a “2-group.” 2-groups therefore only have labels on 1-d and 0-d defects. They can be described more concretely in terms of more familiar objects, namely, ordinary groups.
Examples of crossed modules abound.
Let G be any group, H ≔ G, τ ≔ idG, and let α be conjugation.
Let H be any group, G ≔ Aut(H), let τ(h) be the automorphism defined by τ(h)(h′) ≔ hh′h−1 for all h, h′ ∈ H, and set α ≔ idAut(H).
Let N be a normal subgroup of G. Set H ≔ N, τ the inclusion, and α the conjugation.
Let G be a Lie group, τ: H → G a covering space, and α conjugation by a lift. For instance, and the quotient map SU(n) → SU(n)/Z(n) give examples. Here, SU(n) is the set of n × n special unitary matrices and Z(n) is its center, i.e., elements of the form e2πik/nidn with
Let be the trivial group consisting of a single element, any Abelian group, the constant group homomorphism, and the unique map sending the identity of to .
It is “not” possible for H to be a non-Abelian group if G is trivial. In fact, for an arbitrary crossed module, (H, G, τ, α), ker(τ) is always a central subgroup of H.
We now use crossed modules to construct examples of 2-categories, specifically 2-groups.
This last class of examples of 2-groups from crossed modules will be used throughout this paper. In fact, all 2-groups arise in this way.
We now provide some examples of 2-groups along with weak functors between them to illustrate their meaning.
The following fact will be used in distinguishing two types of gauge transformations. It allows one to decompose an arbitrary gauge transformation into a composition of these two types.
III. COMPUTING PARALLEL TRANSPORT
In classical electromagnetism, or gauge theory, in general, the equations of motion dictate the dynamics. In particular, the field strength, not the gauge potential, appears in the equations of motion. The vector potential becomes relevant when formulating the equations of motion as a variational principle which is itself a reference point towards quantization.55,13 The exponentiated action and parallel transports of gauge theory are realized precisely in this intermediate stage of local prequantum field theory which lies between classical field theory and quantum field theory. We will focus on special 1-d and 2-d field theories, i.e., particle mechanics and string theory. The particle case is provided as a review as well as to set the notation. We will use the 2-dimensional algebra of Secs. II B and II C to explicitly compute parallel transport and its change under gauge transformations. The novelty here, compared with the results of Ref. 5, for instance, is the explicit calculations on a cubic lattice and a direct derivation of the formula for the parallel transport including convergence results. Although our main results are Propositions 3.5 and Theorem 3.8, the diagrammatic picture developed for how these gauge fields interact with combinations of edges and plaquettes in a lattice should be fruitful for future applications.
A. One-dimensional algebra and parallel transport
The solution to the initial value problem (IVP)
at time T with A(t) being a time-dependent n × n matrix is
where stands for time-ordering with earlier times appearing to the right, namely,
where f: {1, …, k} → {1, …, k} is any bijection such that
The choice of sign convention (3.1) is to be consistent with Refs. 14, 56, and 30. Be warned, however, as this sign will lead to different conventions for other related forms such as the curvature 2-form, the connection 2-form, and gauge transformation relations. Certain authors use other conventions.57,58 Yet another convention is to include an imaginary factor.47
The IVP (3.1) shows up in several contexts such as (a) solving Schrödinger’s equation with A(t) = iH(t) for a time-dependent Hamiltonian and ψ being a vector in the space on which H acts and (b) calculating the parallel transport along a curve in gauge theory, where A is the local vector potential, a matrix-valued (or Lie algebra-valued) differential form on a smooth manifold M. This integral goes under many names: Dyson series, Picard iteration, path/time-ordered exponential, Berry phase, etc.
Most of the calculations in this paper will follow this sort of logic. Although similar techniques were used in Refs. 11 and 15, we were largely motivated by the analysis in Ref. 47 and hope that our treatment will be accessible to a larger audience. More rigorous results can be found in Refs. 30, 5, 31, and 12.
B. Two-dimensional algebra and surface transport
Understanding higher form non-Abelian gauge fields has been a long-standing problem in physics, particularly in string theory and M-theory (see, for instance, the end of Refs. 63). Some progress is being made to answer some of these problems with the use of higher gauge theory (see Ref. 41 and the references therein). Although we do not aim to solve these problems, we hope to indicate the important role played by category theory in understanding certain aspects of these theories. We will show how 2-categories and the laws set up in Secs. II B and II C naturally lead to the notion of parallel transport along surfaces. This will also illustrate how explicit calculations can be done in 2-groups. Parallel transport will obey an important gluing condition analogous to the gluing condition for paths. Gauge transformations will be studied in Sec. III C. Furthermore, we will produce an explicit formula analogous to the Dyson series expansion for paths. Although an integral formula is known in the literature,5 the derivation there is not entirely direct nor is it obvious how the formulas are derived from, say, a cubic lattice approximation. A sketch is included in Ref. 15 in Sec. 2.3.2, but further analysis was done in path space, which we feel is more difficult. Indeed, the goal of that work was to relate gerbes with connection on manifolds to connections on their corresponding path spaces. Furthermore, although experts are aware of how bigons are related to more general surfaces, we explicitly perform our calculations on “reasonable” surfaces, namely, squares, for clearer visualization. Our method is more in line with the types of calculations done in lattice gauge theory.20
We feel it is important to express surface transport in a more computationally explicit manner using a lattice and “derive,” from the ground up, a visualization of the surface-ordered integral sketched in Fig. 15 in Ref. 14. This is done in Proposition 3.5, Theorem 3.8, and the surrounding text. Just as the G-valued parallel transport along paths in a manifold M is described by a functor crossed-module -valued parallel transport along surfaces should be described by a functor from some 2-category associated with paths and surfaces in M to the 2-group Ideally, such a 2-category should be a version of the (extended) 2-dimensional cobordism 2-category over the manifold M to mimic the ideas of functorial field theories. However, this has not yet been achieved in this form for general 2-groups. In fact, a version of this has only recently been achieved for the 1-dimensional case by Berwick-Evans and Pavlov.64 An earlier work on Abelian gerbes indicates this should be the case in general though this has not been fully worked out.65 Part of the reason is due to the fact that the representation theory for higher groups is a rather young subject.66
Fortunately, a related solution exists if one works with a 2-category of paths and homotopies. This 2-category is denoted by It is more natural to describe this category in terms of the Poincaré dual of string diagrams. Namely, objects of are points of M, 1-morphisms of are thin homotopy classes of paths in M, and 2-morphisms are thin-homotopy classes of “bigons” in M. A bigon is essentially a homotopy between two paths whose endpoints agree.
Two bigons Σ and Γ from paths γ to δ are “thinly homotopic” if there exists a smooth map of a 3-dimensional cube into M whose top face is Σ, whose bottom face is Γ, and similarly for the other face for the paths γ and δ along with their endpoints (all of these assume some constancy in a small neighborhood of each face). Furthermore, and most importantly, this map cannot sweep out any volume in M, i.e., its rank is strictly less than 3.
Let trivn be the generalization of the expression given in (3.40) for an n × n grid decomposition. Then, the sequence converges as n → ∞.
To prove this, one can introduce trivP for any partition P of the unit square. Then, one needs to show that this quantity has a well-defined limit over all partitions ordered by refinement. This argument has been deferred to Appendix B.
This argument has been deferred to Appendix B.
C. Gauge transformations for surface transport
In Sec. III A, we described gauge transformations as natural transformations of parallel transport functors for paths. In this section, we will use this as the “definition” of a gauge transformation and derive the corresponding formulas for differential forms. As before, let be a crossed module, be its associated 2-group, and M be a smooth manifold. A “(first order) gauge transformation” from a parallel transport functor to another is a (smooth) natural transformation triv ⇒ triv′. By Definition 2.16 and Proposition 2.30, such a natural transformation consists of a pair of (smooth) functions g: M → G and h: P1M → H satisfying the conditions described in that proposition. Namely, to every thin path
to every pair of composable thin paths
to every point x ∈ M,
and finally to any worldsheet,
viewed as a bigon from γδ to ζξ, the equality
holds [this follows from condition (d) in Definition 2.16]. Reading this diagram is a bit tricky without the arrows (recall Remark 2.9). More explicitly, this equality says
i.e.,
or equivalently by our earlier condition (3.87)
By Proposition 2.30, such a natural transformation can be decomposed as
A gauge transformation of the type (g, e) is typically called a (first order) “thin gauge transformation” and one of the type (e, h) is called a (first order) “fat gauge transformation.”69 Thus, Proposition 2.30 implies that an arbitrary gauge transformation of the first kind can be decomposed into a thin and fat gauge transformation. Using this, we can calculate infinitesimal versions of the functions g: M → G and h: P1M → H for small paths, i.e., for a point x ∈ M and a tangent vector at x. This was already done for g: M → G at the end of Sec. III A with result (3.17). For h: P1M → H, let t↦x(t) parametrize an infinitesimal path γ. Then, to lowest order in Δt,
for some 1-form Thus, plugging these expressions into (3.86), a fat gauge transformation from (A, B) to (A′, B′) infinitesimally gives
because g has been set to be the identity. Expanding out to lowest order in Δt gives
giving the relationship
for a fat gauge transformation. We already calculated what happens for a thin gauge transformation in Sec. III A. Using Proposition 2.30, combining (3.17) with this gives
for an arbitrary gauge transformation. The B field under an arbitrary gauge transformation changes according to (3.89). By substituting the necessary forms, this expression on the left-hand-side of (3.89) becomes (to avoid clutter, we have not explicitly written Δs and Δt)
where it is understood that all terms now are evaluated at (s, t), to lowest order. Meanwhile, the right-hand-side of (3.89) is
in components or
as an equation in terms of differential forms. We write such a gauge transformation as
This and (3.96) agrees with Proposition 2.10 of Ref. 5. We can express this purely in terms of A, B, g, and φ as
This will be useful later.
Now suppose (g, h), (g′, h′): triv ⇒ triv′ are two first order gauge transformations. A “second order gauge transformation” a: (g, h) ⇛ (g′, h′) is a modification from (g, h) to (g′, h′). By Definition 2.19, this consists of a function a: M → H fitting into
which, in particular, says
satisfying the condition that to any path
Expanding out this expression on infinitesimal paths gives
which to lowest order says
Note that if h ∈ H, the function αh: G → H is defined by G ∋ g ↦ αg(h) so that is the derivative. This result gives the condition (after multiplying by a−1 on the right)
on components and
as -valued differential forms. This and (3.103) exactly agree with Proposition 2.11 of Ref. 5. Physically, a second order gauge transformation is a gauge transformation for the 1-form φ, which itself is a type of field even though it appears as a (fat) gauge transformation for the 1-form and 2-form gauge potentials.
D. Orientations and inverses
It is well-known that given a path the parallel transport along the reversed oriented path γ−1 is the inverse
where is the (local) parallel transport functor. This can be viewed as a consequence of thin homotopy invariance and functoriality of parallel transport. Namely, although the paths γγ−1 and γ−1γ are “not” constant paths (the notation γ−1 is therefore a bit abusive), they are thinly homotopic to constant paths and hence give the same value on triv. Thus,
verifying (3.109). In this section, we will explore analogous results for reversing different kinds of orientations on bigons. We therefore include arrows for clarity
Technically, there is one more possibility given by However, this possibility is a combination of the above two, namely, The meaning of these different possibilities can be explained physically as follows. A string may be given the additional datum of an orientation. Furthermore, as it moves in time, it has an additional directionality. These two directional orientations are precisely encoded in the definition of a bigon/2-morphism. These different orientation reversals are given by time reversal for and spatial orientation reversal for Σ−1. The case corresponds to both reversals (the order of reversal does not matter since the operations commute). Note that the different orientations on a bigon can be expressed as an orientation of edges on the boundary and an orientation of the surface. The above bigons correspond to the following surfaces with associated orientations
respectively.
A necessary and sufficient condition for such orientations on surfaces and edges to give rise to a bigon is the following. Given a map of a polygon Σ into M, the boundary consists of the edges of the polygon. The union of the oriented edges consistent with the orientation of the polygon must be connected. Similarly, the union of the orientated edges with negative orientation with respect to the induced one from the polygon must also be connected. Then, the source of the bigon is the union of the consistent edges and the target is the union of the oppositely oriented edges. An example together with a nonexample are
respectively (blue, with arrows written using >, corresponds to an orientation agreeing with the induced one from the surface, while yellow, with arrows written using ≫, disagrees with that orientation).
Going back to the three bigons and their orientations at the beginning of this section, we notice that several of these bigons can be composed with one another. For instance,
and
after applying thin homotopies. Therefore, these bigons provide inverses in series and in parallel, respectively, of Σ. This, together with functoriality of triv and the inverses discussed in Example 2.27, implies
and therefore describes how parallel transport along surfaces changes under reversals in surface orientations and boundary orientations, respectively. For completeness, for the last possible orientation we have
E. The 3-curvature
In the following, we go through some additional calculations in higher gauge theory. Just as the curvature F of a 1-form connection A can be obtained by calculating the parallel transport along an infinitesimal loop, the 2-curvature of a 2-form connection (A, B) can be obtained by calculating the surface transport along an infinitesimal sphere, which on a cubic lattice corresponds to a cube. We will perform this calculation explicitly and study some properties of the resulting 3-form curvature. Similar analysis was done on a tetrahedron in the context of categorical higher gauge theory in Ref. 11. Another version was explored for cubes in Ref. 70 though the latter did not formulate their results using the notion of a 2-group.
Let (r, s, t) ↦ x(r, s, t) be an infinitesimal cube and consider the following domain for that cube along with the infinitesimal path that goes first along the r direction, then in the s direction, and finally in the t direction. Our convention is that (r, s, t) is a right-handed coordinate frame, i.e., dr ∧ ds ∧ dt is the volume form
Such a cube can be expressed as a bigon by the following composition of plaquette bigons that begin and end at the same path starting at the top left and moving clockwise:
The corresponding 2-group elements are given as follows: We begin with the first surface
where we use the shorthand notation
as well as
and similarly for the other terms. We also write ϵ instead of Δr, Δs, or Δt and use the derivatives to remind ourselves of the direction. We have also assumed for simplicity that our coordinates are centered at the origin and the lattice spacing is ϵ in each direction. Working out this diagram infinitesimally on the 0-d defect gives
to lowest order. As usual, rather than writing out the Δr, Δs, Δt, we use the number and type of derivatives appearing to keep track of the order. The other terms are given as follows:
The composition of all of these elements is given by the following diagram (with the light shaded blue squares depicting the faces on the cube):
The result of multiplying these out using two-dimensional algebra gives
This is yet another manifestation of two-dimensional algebra. The result of multiplying all these terms is given as follows, order by order. The zeroth order term is There are no first order terms. The second order terms are given by
by antisymmetry of Bμν in the μ and ν indices. Thus, the only nonzero terms are the zeroth and third order terms (up to third order). One type of the third order terms is given by
and vanishes again by antisymmetry of Bμν and commutativity of partial derivatives. The final result of expanding out (3.121) is therefore
In analogy to the curvature 2-form associated with a 1-form potential A obtained by calculating the holonomy along an infinitesimal square, we define this third order term to be the “3-form curvature” associated with the pair (A, B) and denote it by H (although this notation conflicts with our crossed module notation, it should cause no confusion in this section based on its usage). In terms of components, it is given by
and using differential form notation
This definition and result agrees with (3.28) of Ref. 11 and Lemma A.11 in Ref. 5. As was also pointed out in Ref. 11,
by the Bianchi identity. Since is a central Lie subalgebra of this means that H is a 3-form with values in an Abelian Lie algebra (see Remark 2.26). Under a first order gauge transformation as in (3.101), the 3-form curvature changes to
where the underlined terms cancel and the other terms with underbraces will be calculated and compared momentarily. At this point, it is useful to simplify some of these terms by applying and calculating the results in terms of commutators, etc. For example,
We can safely equate the terms inside the This is a helpful trick and we will use it to calculate all other terms. For instance, one can show using this trick that
Although this is a trick and not completely rigorous, it works for all of the calculations we will do. These formulas can all be found in Appendix A of Ref. 25 and can be derived more carefully. Since is a derivation,
which cancels with the term in (3.128). Furthermore, note that
for any -valued 1-form X and for any -valued 1-form Y. Although has been applied to derive these equalities, the expressions inside are still equal. This equality implies
One of the more cumbersome set of terms in (3.128) is
Combining this with the result preceding it gives just a single term Putting all of this together, we obtain
Finally, by the properties of crossed modules and by the vanishing of the fake curvature,
so that the above formula reduces further to
In particular, H is invariant under fat gauge transformations. This result agrees with what was discovered in Ref. 11.
IV. CONCLUSION
We have illustrated that 2-category theory can be implemented and used in such a way as to calculate parallel transport along two-dimensional surfaces, such as worldsheets of strings. We did this explicitly for gauge groups that are not necessarily Abelian via an approximation technique that can be implemented numerically. We have done this using string diagram techniques to facilitate 2-categorical techniques and bring higher category theory to a wider audience. Although Girelli and Pfeiffer have calculated infinitesimal gauge transformations and curvature forms via similar techniques11 and Schreiber and Waldorf provided a formula for the parallel transport along a surface,5 our infinitesimal methods give a much more explicit and direct construction of the iterated surface integral from elementary building blocks. This fills in some of the arguments sketched by Baez and Schreiber in Ref. 15, particularly in Sec. 2.3.2 (see also Sec. 11.4.1 of Schreiber’s thesis71). Schreiber and Waldorf’s integral in Ref. 5 was obtained from consistency conditions, and then they proved that it satisfies the necessary functorial properties expected of surface holonomy. In relation to other studies, such surface-ordered integrals have been used recently in constructing a Hochschild complex for surface transport.72 The novelty of our result is that we derived the formula for surface parallel transport from scratch using a discretization of our surface. To the best of our knowledge, this is the first appearance of such an explicit construction together with analytical results on convergence (Proposition 3.5) and a simplification providing a manageable surface-ordered integral by reducing the surface ordering to a single direction as opposed to two (this result is embodied in Theorem 3.8). By implementing string diagrams, we have also provided a more friendly visualization. Furthermore, we have avoided using path spaces explicitly and have simplified many arguments.
We hope that we have illustrated how two-dimensional algebra can be used for explicit calculations. If developed further, these ideas might be used to explain physical phenomena that utilize algebraic manipulations in more than one dimension more naturally. Such higher-dimensional algebra appears in many situations. For example, elements and molecules combine in a variety of ways forming complicated compounds, amino acids, and proteins. These are objects that use three dimensions to configure themselves. Therefore, a natural and faithful representation of them would involve a sort of 3-dimensional algebra. Another example occurs in painting. Given a painting, it is much simpler for us to “read” a 2-dimensional painting than to view all the pixels making it up in a straight 1-dimensional list. Both perspectives contain the same information-theoretical data, but the 2-d form is naturally and immediately recognized. As another even more speculative example, it is known that the entropy of a black hole is proportional to the surface area of its horizon. This may lead one to believe that the microstates of the theory can be expressed as living on a lower-dimensional world. This in turn then suggests the possibility that a two-dimensional algebra might be useful in describing some of the properties of these microstates. Although these ideas are entirely speculative, our point is that one can imagine that the one-dimensional algebra we have been using for so long is only the tip of an iceberg of algebraic structures. Higher category theory opens us to these other possibilities.
There are still many open questions in this relatively young field. One is how to construct useful actions in physics that model phenomena with non-Abelian higher form gauge fields and also the interactions with matter fields. Some recent progress in this direction has been made by Sämann and others—see Ref. 41 and the references therein. Work on the pure gauge field side was initiated in the work of Pfeiffer using a 2-categorical approach.10 To proceed, it seems that a better suited representation theory for 2-categories will be useful.66 Furthermore, characteristics for 2-groups73,74 and traces54,75 may need to be studied further to better understand what gauge invariant combinations are possible. Although the number of higher gauge theory examples are increasing76,77,39,69,38,14,41 some work is still required to solidify the role of higher gauge theory in lattice gauge theory and other areas of physics. Other lattice gauge theory approaches existed earlier78,79,80 with a renewed interest in Refs. 70 and 81, but it is not clear to us how these approaches to higher lattice gauge theory are related to the rest of the literature mentioned above.
In the realm of string theory and M-theory, beginning with early studies of Witten, Myers, and others,82,37 ideas for the construction of non-Abelian gauge theories on a stack of D-branes83 and their low energy effective actions were discussed. These effective actions are swarmed with higher form non-Abelian gauge fields, but the precise mathematical formulation is still lacking. Nevertheless, it is likely that non-Abelian differential cohomology13 is relevant based on the recent work on M5-branes in which it plays an essential role.39 Some arguments used to describe such effective actions are not always entirely straightforward and involve consistency conditions (such as T-duality37 and scattering amplitude calculations84) rather than direct derivations. It is therefore possible that a more thorough investigation may involve understanding non-perturbative effects, one of which is dictated by transport. On the other hand, due to the non-commutative nature of the normal coordinates to these branes,85 this may involve a modification of such transport to the setting of non-commutative geometry. These and many other ideas have also been briefly discussed in Ref. 71, and several such open questions can be found there.
ACKNOWLEDGMENTS
We express our sincere thanks to Urs Schreiber and Radboud University in Nijmegen, Holland, who hosted us for several productive days in the summer of 2012 during which a preliminary version of some ideas here were prepared and presented there. We also thank Urs for many helpful comments and suggestions. We would like to thank Stefan Andronache, Sebastian Franco, Cheyne Miller, V. P. Nair, Xing Su, Steven Vayl, Scott O. Wilson, and Zhibai Zhang for discussions, ideas, interest, and insight. Most of this work was done when the author was at the CUNY Graduate Center under the NSF Graduate Research Fellowship Grant No. 40017-01-04 and during a Capelloni Dissertation Fellowship. This work is an updated version of a part of the author’s Ph.D. thesis.88
NOMENCLATURE
category/2-category
- G
group
one-object groupoid
- σ: F ⇒ G
natural transformation from functor F to G
crossed module
one-object 2-groupoid
- A
1-form potential
- M
smooth manifold
identity matrix
- triv(γ)
local transport along a path γ
path groupoid of M
- g
(thin) gauge transformation
path 2-groupoid of M
- triv(Σ)
local transport along a bigon Σ
- B
2-form potential
- F
2-form curvature of A
- φ
fat gauge transformation
- H
3-form curvature of (A, B)
differential crossed module
APPENDIX A: DIFFERENTIAL LIE CROSSED MODULES
Here, we briefly review the infinitesimal version of a Lie crossed module (H, G, τ, α) including the many relations that these maps satisfy. We also make some comments on differential forms with values in and This information can also be found in many articles on the subject of higher gauge theory such as Refs. 15 and 11, especially Waldorf’s concise one page formula sheet in Appendix A of Ref. 25. Martins and Miković also have an exceptionally clear and thorough exposition in Sec. 2.1 of Ref. 69.
is the derivative of τ: H → G at the identity and is a Lie algebra homomorphism since τ is a Lie group homomorphism. Notice that α can be equivalently described as a function α: G × H → H that is a group homomorphism in each component separately. As a result, for any fixed g ∈ G, αg: H → H is a Lie group homomorphism and hence has a derivative at the identity denoted by This map, besides being a Lie algebra homomorphism, satisfies the additional property that
for all and g ∈ G. Similarly, although α: G × H → H is not a group homomorphism, it is smooth and its derivative is a well-defined linear map. It is a derivation once the coordinate is fixed, i.e.,
for all and also satisfies
for all and Finally,
and
for all and
Once combined with differential forms, the maps and are extended in the appropriate way (see Ref. 56, Part II, Chap. 3, in the section on the Bianchi identity for details on differential forms with values in Lie algebras). For instance, is a graded derivation in its second coordinate. To clarify the notation used throughout, consider differential forms When we write expressions such as or , we mean the following. First, let be a basis for and be a basis for Then
where a summation over repeated indices is assumed and where Aa, φb ∈ Ω1(M) and Fa, Bb ∈ Ω2(M) for all indices. Then by definition,
and similarly for any other forms. Because we use Lie algebra valued forms, the bracket is graded. For instance,
but
since φ is a 1-form and B is a 2-form. The last two equalities follow if we think of our Lie algebras as coming from matrix Lie algebras, which we often do. The general formula is
where |ω| and |η| are the degrees of the forms and Other properties are derived as needed in calculations in the article.
APPENDIX B: SURFACE PRODUCT CONVERGENCE
This appendix serves to prove the convergence of the surface-ordered product (3.50) as n → ∞ and to also prove upper bounds on the kth order terms when expanded out. For this, we will first relax our conditions and work with arbitrary partitions of the unit square. We will follow the conventions of Munkres and use the results there without further Ref. 86. The surface-ordered product is well-defined for each partition P and will be denoted by trivP. We will also use the notation γs,t to denote the path defined after Theorem 3.8. It will be helpful to define the function
APPENDIX C: PROOF OF CONFIGURATIONS LEMMA
This appendix serves to give a rigorous proof of Lemma 3.7. For the proof of this lemma, it is useful to rewrite Sn,k as
where it is understood that any sum operation on the left acts on everything to the right. Before working out this summation to obtain a more explicit formula, for each define the function
Explicitly, this simplifies to Ref. 87
where δqp is the Kronecker delta function and Br is the Bernoulli number defined, for instance, by the coefficients in the power series expansion (thought of as a formal power series in the variable x)
The first few of these Bernoulli numbers are
while the first few ϕn are
Examining ϕn(p) a little more, one immediately notices the crucial result
Now, Sn,k can be written as a polynomial in the ϕn’s
where
and so on (a more explicit formula will be given momentarily). For example, one obtains the following expressions for small values of k:
Looking back at the expressions for Sn,k, one sees that there is a recursion relation for Sn,k. Setting Sn,0 ≔ 1, this recursion relation reads
This recursion relation can be used to express Sn,k purely in terms of the ϕn’s and is given by
where it is understood that the sum terminates earlier if any of the j’s are larger than 1. For example, if there are s of them, then
Therefore, fixing k, one obtains
To see this, first notice that jr = 1 for all r ∈ {1, …, s} when s = k in which case the denominator and numerator in (C14) are equal and the limit is 1. However, when s < k, by the formula for ϕn(p) in (C3) and the asymptotics of this given in (C7),
Hence,
Finally going back to Rn,k and using this fact gives