Thin Homotopy and the Signature of Piecewise Linear Surfaces
Abstract.
We introduce a crossed module of piecewise linear surfaces and study the signature homomorphism, defined as the surface holonomy of a universal translation invariant -connection. This provides a transform whereby surfaces are represented by formal series of tensors. Our main result is that the signature uniquely characterizes a surface up to translation and thin homotopy, also known as tree-like equivalence in the case of paths. This generalizes a result of Chen and positively answers a question of Kapranov in the setting of piecewise linear surfaces. As part of this work, we provide several equivalent definitions of thin homotopy, generalizing the plethora of definitions which exist in the case of paths. Furthermore, we develop methods for explicitly and efficiently computing the surface signature.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x1.png)
1. Introduction
The signature of a path is a non-commutative power series
(1.1) |
whose coefficients are obtained by taking iterated integrals of the path:
(1.2) |
where is a multi-index and , where is a basis of . The path signature was introduced by Chen [16], who proved that this invariant uniquely characterizes a path up to translation and thin homotopy [17]. Thin homotopy is an equivalence relation on paths that essentially consists in two basic equivalences: reparametrizations, and cancellation of retracings. This allows paths to be treated analogously to words in a free group, a fact which plays a crucial role in Chen’s proof of the injectivity of the signature.
Thin homotopy is a significant strengthening of the conventional homotopy relation in algebraic topology, from which the fundamental group of a manifold arises. This leads to the thin fundamental group, , which is defined as the group of loops in a manifold modulo the thin homotopy equivalence relation. This notion first arose in differential geometry and physics through the study of connections and the invariances inherent in their parallel transport. The thin fundamental group plays a key role in the generalization of the Riemann-Hilbert correspondence to the setting of non-flat connections due to [3, 11]. Given a Lie group , this states that there is an equivalence between the category of all -connections on and the category of smooth -representations of . In fact, the path signature can be seen as the holonomy of a universal translation-invariant connection on valued in the free Lie algebra generated by .
Since their introduction, Chen’s iterated integrals have become highly influential in many areas of geometry and topology (eg. [30, 31, 4, 1, 36, 37, 15, 14, 6]). More recently, the path signature was foundational in developing the theory of rough paths by Lyons in [41, 42], which has played a prominent role in the areas of stochastic analysis [26] and machine learning [45, 40]. In [32], Hambly and Lyons extended the injectivity of the path signature to bounded variation paths, while in [5], Boedihardjo et al. generalized this further to highly irregular rough paths. In this context, the thin homotopy equivalence relation is known as tree-like equivalence.
The motivation for the present paper is to generalize Chen’s injectivity result in another direction to the setting of surfaces. Surface holonomy is the generalization of parallel transport to surfaces, originally developed to study higher gauge theory [2, 43, 50]. The surface signature was introduced by Kapranov in [34] as the surface holonomy of a universal translation-invariant -connection on . This can be formulated as a homomorphism of crossed modules , between a crossed module of thin equivalence classes of surfaces in Euclidean space, and a crossed module of ‘formal surfaces’, which is defined by a formal integration of a free crossed module of Lie algebras . This construction has recently received interest in the rough paths literature [38, 18] where it forms the basis of a theory of rough surfaces. One of the key advantages of Kapranov’s notion of the surface signature over alternate approaches to generalizing the signature [29, 25, 24] is the fact that it preserves the concatenation structure of surfaces. This is crucial to proving the extension results in [38, 18], and leads to parallelizable computations for potential applications in machine learning.
In [34, Question 2.5.6], Kapranov poses the question of whether the signature characterizes a surface uniquely up to translation and thin homotopy. Thin homotopy for surfaces, first introduced by [12], is the equivalence relation whereby two surfaces are thinly equivalent if there is a homotopy between them which does not sweep out any volume. As in the case of paths, this includes generalized reparametrizations and cancellation of folds. However, there are also more general ‘non-local’ thin homotopies. For example, in Proposition 3.4 we construct a closed surface which factors through the real projective plane
(1.3) |
which is thinly null homotopic, even though it does not exhibit any folds to cancel. This example illustrates the crucial fact that surfaces do not admit unique reductions via folds. In contrast, every path can be uniquely reduced, up to reparametrization, to a path which does not contain any retracings. As the proofs of injectivity of the path signature rely on this property, these approaches cannot be immediately generalized to the case of surfaces.
In this paper, we focus on the special case of piecewise linear surfaces in order to avoid the analytical subtleties and focus on the underlying algebraic structures. In Section 5 we define a crossed module of piecewise linear surfaces , which is functorial in the vector space . Because the formal integration of the free crossed module is also functorial in , this allows us to define the surface signature as a natural transformation. This level of abstraction is immediately paid off by the following remarkable uniqueness result.
Theorem 1.1.
The piecewise linear surface signature is the unique natural transformation extending the piecewise linear path signature. Furthermore, the smooth surface signature is the unique continuous natural transformation extending the smooth path signature.
Our main result is the injectivity of the surface signature for piecewise linear surfaces.
Theorem 1.2.
Let be a piecewise linear surface such that its signature is trivial, Then is thinly homotopic to the constant surface.
Our proof of this result makes use of two main ideas:
-
a)
First, because of the injectivity of the path signature, any smooth surface with vanishing signature can be assumed to be closed. In [34], Kapranov suggests that the signature of a closed surface is given by its associated current. In Section 4, we use a gauge transformation to abelianize the universal -connection and give a proof of this fact in Theorem 4.12. As a result, we show in Corollary 4.13 that any closed surface with vanishing signature has the property that
(1.4) -
b)
In Section 6, we complete the proof of injectivity. A closed surface which satisfies (1.4) has vanishing homology in its image ,
(1.5) To conclude that is thinly null homotopic, it suffices to show that , and by Hurewicz, this would be true if . In other words, the obstruction is the fundamental group of . Hence, by attaching sufficiently many discs to , we can kill the fundamental group and thereby produce the desired thin null homotopy. In order to implement this, we make use of Whitehead’s theorem [55] that the fundamental crossed module of a -dimensional CW complex is free.
We remark that the above proof works for surfaces that are much more general than piecewise linear ones, but fails to work for general smooth surfaces because of the complicated nature of the image of a general smooth map. Our focus on piecewise linear surfaces is partially justified by the fact that they can be used to approximate general smooth surfaces. We hope to make use of this fact to prove a general injectivity result in future work.
For the remainder of this introduction, we highlight further results obtained in our paper.
Algebraic Models of Piecewise Linear Paths and Surfaces. Elements of the free group on letters can be viewed as thin homotopy equivalence classes of lattice paths on : word reduction coincides with path reduction via retracing. In Section 2.4, we construct the group of piecewise linear paths on as the quotient of the free group on , by certain relations to account for retracings. We show that this satisfies several properties analogous to free groups, such as the existence of minimal words and a universal property.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x2.png)
In Section 5.2, we extend this construction to define a crossed module of piecewise linear surfaces . The distinction between local and non-local cancellations is reflected in our construction. We first construct a pre-crossed module as the free group of planar regions, quotiented by relations which account for local fold cancellations, similar to the case of paths. Next, we quotient out the Peiffer subgroup to obtain a crossed module, which accounts for the remaining non-local cancellations. This crossed module also satisfies a universal property, which can then be used to prove Theorem 1.1. We remark that the injectivity of the signature further implies that the thin homotopy relation for piecewise linear surfaces is a formal consequence of the local fold cancellations and the Peiffer identity. Therefore, the normal subgroup contains within it the surfaces which are thinly null homotopic in a non-trivial way.
Computational Methods for Surface Signature. In recent years, the path signature has been used as a rich feature set for sequential and time series data [45]. Theoretical properties guarantee its effectiveness in approximating functions and characterizing measures on path space [20], thereby justifying its use in machine learning tasks. Furthermore, because the signature preserves concatenation of paths, it lends itself to parallelizable algorithms [35, 53]. While the study of the surface signature is still fairly new, recent work has studied theoretical properties of surface holonomy valued in matrix groups to provide features for 2-dimensional data [39]. However, there are two significant challenges in applying the surface signature itself. These are the lack of established computational methods and the absence of a canonical coordinate representation for surface signatures due to the opaque nature of the Peiffer identity111This should be compared with the path signature, which is valued in the tensor algebra: by choosing a basis on the underlying vector space, this induces a basis on the tensor powers..
In Section 5.4, we develop a natural decomposition of the surface signature into boundary and abelian components in both the piecewise linear and smooth settings. In particular, we find that the linear structure of a vector space induces canonical splittings of the crossed modules , , and , and show that the surface signature preserves the resulting decompositions. The main result of this section is Theorem 5.36, which gives an explicit formula for the signature of a surface. Given a surface , the boundary component simply records the path signature of the boundary, while the abelian component records the integrals of all monomial 2-forms over the closed surface obtained by coning off the boundary of . Schematically, this abelian component is given by the following formula
where is viewed as a function on valued in the symmetric power (see Remark 5.37). This elucidates the information contained in the surface signature, suggests methods for computing the signature, and provides explicit coordinates to represent the signature.
Characterizing Thin Homotopy. Thin homotopy equivalence for paths has been studied widely, and a variety of equivalent definitions have been proposed. Chen’s definition [17] involved reducing the path along retracings, which was generalized by Tlas [52] to paths. Geometric definitions in terms of the existence of homotopies which satisfy additional conditions were given in [11, 3], and a characterization based on holonomy was considered in [52, 46]. On the analytic side, the equivalent notion of tree-like equivalence was defined in terms of factorization through a tree or the existence of a height function [32]. Finally, due to the injectivity of the path signature, we obtain yet another definition. We summarize these results in Theorem 2.11.
This collection of equivalent definitions provides a plethora of distinct ways to understand thin homotopy of paths. While some of these definitions can be easily generalized to the case of surfaces, others require modification due to fundamental differences in the two dimensional setting: we must take into account non-local cancellations. In Theorem 6.26, we propose a generalization of each definition, and show that they are equivalent in the piecewise linear setting.
Finally, in Section 7, we highlight a connection between thin homotopies and group homology. This allows us to further classify the non-local cancellations into those due to conjugation by elements in the fundamental group of the image of a surface , and those which have further complexity. We show that thinly null homotopic surfaces in this latter case are classified by , leading to a new geometric interpretation of group homology.
Acknowledgments. We are very grateful to Camilo Arias Abad, who has met with us to discuss signatures and surface holonomy for countless hours over the past few years, has contributed several important ideas, and has significantly influenced the direction of this project. We would also like to thank Harald Oberhauser for several insightful discussions at the beginning of this project. The first author wishes to thank Tim Porter for introducing him to crossed modules and explaining various fundamental aspects of the theory. Furthermore, numerous stimulating conversations with Martin Frankland, Marco Gualtieri, and Jim Stasheff have influenced this work. F.B. is supported by an NSERC Discovery grant. D.L. was supported by the Hong Kong Innovation and Technology Commission (InnoHK Project CIMDA) during part of this work.
2. Thin Homotopy and the Path Signature
In this section, we provide some background on parallel transport and the path signature. We consider thin homotopy of paths, and relate several known definitions of thin homotopy. We then focus on the piecewise linear setting: while the signature of piecewise linear paths is well understood, we provide a novel functorial construction of the piecewise linear signature, which arises naturally using universal properties. This section serves as a guide to the results we will generalize to surfaces in the remainder of this article. Throughout this article, denotes a finite-dimensional real vector space , and we let denote the category of finite dimensional vector spaces and linear maps. Furthermore, we always consider paths with sitting instants.
Definition 2.1.
A path has sitting instants if there exists some such that
(2.1) |
Unless otherwise specified, we assume all paths have sitting instants.
This ensures that when we consider composable paths , such that , with some smoothness condition, the concatenation
(2.4) |
preserves the smoothness condition .
2.1. Parallel Transport and Path Signature
In this section, we define the path signature as the parallel transport of the universal translation invariant connection. Since we are working over a vector space , all principal bundles can be assumed trivial and as a result, we use the following simplified definition of a connection.
Definition 2.2.
Let be a Lie algebra. A -connection on is a Lie algebra valued -form . A connection is translation-invariant if it has the form
(2.5) |
where and are linear coordinates on . The curvature of , denoted , is defined by the following formula
(2.6) |
Definition 2.3.
Let be a Lie group with Lie algebra . Let be a -connection and a path. Consider the following differential equation for ,
(2.7) |
The parallel transport of along is defined to be
(2.8) |
The path signature is the parallel transport of the universal translation-invariant connection. This is a connection valued in the free Lie algebra generated by , , and it can be understood as the identity endomorphism of
(2.9) |
where we view as the subspace of translation invariant forms, and as the subspace of generators of the free Lie algebra. If form a basis of , with dual basis , viewed as coordinates on , then the connection has the following explicit expression
(2.10) |
Because the free Lie algebra is infinite dimensional, it is convenient to consider its truncations. These can be formally expressed via the lower central series of , defined recursively as
(2.11) |
Then, we define the -truncated free Lie algebra as
(2.12) |
As we assume that is finite dimensional, is a finite-dimensional Lie algebra. Then, we can explicitly integrate to the Lie group of exponential elements of the truncated tensor algebra . This is defined as follows
(2.13) |
We denote the projective limit of these Lie groups, and their corresponding Lie algebras, by
(2.14) |
Given a linear map , there is an induced group homomorphism for all . This induces a group homomorphism . Hence, we get a well-defined functor
(2.15) |
We now provide the standard definition of the path signature, and return to this in Section 2.4 on piecewise linear paths.
Definition 2.4.
Let . For , consider the parallel transport of the universal connection of Equation (2.10) projected onto . In particular, consider the following differential equation for ,
(2.16) |
The -truncated path signature of is defined to be . Then, we define the path signature of to be the projective limit
(2.17) |
Remark 2.5.
The completion of the tensor algebra is often considered analytically by defining appropriate norms on , or by considering a family of seminorms, as in [19]. In this article, as we do not need these analytic properties, we will consider formal completions.
2.2. The Thin Path Group
Parallel transport, and in particular the path signature, is invariant under thin homotopy equivalence. In this section, we construct a group of paths up to translation and thin homotopy equivalence, and show that the signature defines a homomorphism out of this group. The meaning of thin homotopy will be explained further in the next section.
Definition 2.6.
Two smooth paths are thin homotopy equivalent, denoted , if there exists an endpoint preserving smooth homotopy between and such that
-
•
(homotopy condition) and ;
-
•
(thinness condition) , where is the differential of .
Thin homotopy defines an equivalence relation on paths which is compatible with concatenation. The thin fundamental groupoid of is defined to be the set of equivalence classes
(2.18) |
It is a groupoid over with product given by the concatenation of paths. Given a Lie group with Lie algebra , the parallel transport of a connection defines a groupoid homomorphism
(2.19) |
There is a natural action of the additive group on by groupoid automorphisms given by translating paths. We define the thin path group to be the quotient
(2.20) |
and we note that it is a group. Our convention will be to use paths starting at the origin, , as representatives of the translation equivalence classes. There is a well-defined group homomorphism given by sending a path to its displacement . The thin fundamental groupoid can then be recovered as the corresponding action groupoid
(2.21) |
Given a linear map and a path , we can define a new path by . This preserves thin homotopy classes of maps, and is equivariant with respect to translation. Thus, we obtain a thin path group functor
(2.22) |
Because the universal connection is translation invariant, its parallel transport is invariant under both thin homotopy and translations. Therefore, the path signature defines a homomorphism . Furthermore, by [27, Proposition 7.52], this fits into a natural transformation
(2.23) |
2.3. Thin Homotopy of Paths
There are several different equivalent definitions for thin homotopy equivalence of paths. In the signatures and analysis literature, it is known as tree-like equivalence and has been extended to the setting of rough paths. In this section, we will review the equivalent definitions in the setting of paths. In particular, we will see that the path signature characterizes this equivalence relation. Roughly speaking, thin homotopy equivalence captures two main types of behavior:
-
(1)
Reparametrizations: Given a path and a reparametrization , then .
-
(2)
Retracings: Given paths , we say that a retracing is a path segment of the form . Paths which differ by retracings are thin homotopy equivalent:
(2.24) The path is called a reduction of .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x3.png)
We will focus mainly on thinly null-homotopic paths: paths that are thin homotopy equivalent to the constant path. This is because two paths and can then be defined to be thin homotopy equivalent if is thinly null-homotopic. Chen [17] originally defined thinly null-homotopic paths in the piecewise regular setting via path reductions as shown above. Tlas [52] generalized this definition to paths, where there may be infinitely many retracings, using the notion of transfinite words.
Definition 2.7.
[13, Definition 3.1] Let be a set consisting of an alphabet, and let denote a formal inverse set. A transfinite word over is a function , where is a totally ordered set, such that is finite for each . A word is reducible to the trivial word if and only if every finite truncation (mapping all but finitely many letters in the alphabet to the identity) reduces to the trivial word.
Theorem 2.8.
[52, Theorem 1] Let be a path. There exists a collection of mutually disjoint subsets such that
-
(1)
is closed, is open for , and .
-
(2)
If , then the set inverse and for , then .
-
(3)
if for , while for (the interior).
-
(4)
Each for is a union of disjoint open intervals. The path restricted to any such interval is an embedding. The image of any two such embeddings are either disjoint or identical.
Here, the set is used to “collect” all regions of the interval in which the path is constant. Let denote the set of intervals from for all , equipped with a linear order inherited from . Let be the set of embedded arcs (see (4) above) where we remove repetitions (up to reparametrization and switch in orientation). Let be a copy of denoting formal inverses. Then, we define the transfinite word associated to to be the mapping sending each interval to its corresponding arc, taking into account the orientation. Any arc has finite length, and thus must be finite because is .
Definition 2.9.
The path is word reduced if is reducible to the trivial word.
This definition is instructive as it explicitly specifies a (possibly infinite) partition of the path, then matches up each constituent arc with an adjacent arc with the same image and opposite orientation. We will use a variation of this idea in our main injectivity proof for the surface signature later in the article. We now turn to other equivalent definitions of thin homotopy.
Definition 2.10.
A metric space is an -tree if for any pair of points , all topological embeddings where and have the same image.
The following theorem collects known results relating various definitions of thin homotopy.
Theorem 2.11.
Let be a path with vanishing derivative at the endpoints. It is thinly null-homotopic or tree-like if any of the following equivalent definitions hold:
-
(W1)
Word Condition. The path is word reduced.
-
(H1G)
Holonomy Condition. For a semi-simple Lie group with Lie algebra , the parallel transport of every smooth -connection along is trivial222Each semi-simple is treated as an independent condition..
-
(R1)
Rank Condition. There exists an endpoint-preserving homotopy from to the constant path at such that the rank of is everywhere.
-
(I1)
Image Condition. [3] There exists an endpoint-preserving homotopy from to the constant path at such that .
-
(F1)
Factorization Condition. [5] There exists a factorization of through an -tree , namely .
-
(A1)
Analytic Condition. [32] There exists a Lipschitz function such that for all , and if , then . The function is called the height function.
-
(S1)
Path Signature Condition. The path signature of is trivial, .
Proof.
In [52, Theorem 3] it is proved that (W1), (H1G) for all semi-simple , and (R1) are equivalent. Furthermore, in the proof of [52, Theorem 2] which shows (W1) (R1), a homotopy is constructed which satisfies the (I1) (see the top of [52, page 18]), so (W1), (H1G), or (R1) also imply (I1). Furthermore, if there exists a homotopy which satisfies (I1), this homotopy must also satisfy the rank condition (R1). Thus the first four conditions are equivalent. Furthermore, we note that the definition of a tree in [52, Definition 4] is a special case of our definition. Then, (W1), (H1G), (R1) or (I1) all imply (F1).
Now, [32] shows the equivalence of (F1), (A1) and (S1) in the case of bounded variation paths, which includes paths. Note that these three conditions do not use a homotopy (whose regularity we would need to consider). Thus, this implies these are equivalent in the setting.
In order to connect these two classes of results, we show that (S1) implies (H1G) for the specific case of . Suppose such that . Suppose to the contrary that there exists a smooth connection
(2.25) |
where , such that the holonomy (with initial condition at the identity) is not trivial . The holonomy can be expressed in terms of the iterated integrals of , where specific entries are iterated integrals of . This implies that there exists such an iterated integral which is nontrivial, and by following the proof of [17, Lemma 4.1], this implies that the signature is nontrivial, which is a contradiction. ∎
While these equivalent definitions are stated for paths, the tree condition (F1) can be extended to highly irregular rough paths, where this theorem has been generalized to show that the kernel of the path signature consists of tree-like rough paths [5]. We do not consider such rough paths in this article. Instead, we focus our attention on the piecewise linear setting where analogous questions regarding thin homotopy for 2-dimensional surfaces are still largely unexplored.
2.4. Group of Piecewise Linear Paths
In this section, we give an algebraic construction of the group of piecewise linear paths and the path signature. Given a vector space , consider the free monoid on the underlying set of . We define the group of piecewise-linear paths as
(2.26) |
subject to the following relations:
-
(PL0.1)
if and are linearly dependent;
-
(PL0.2)
, where and is the identity in (the empty word).
With these relations is a group. Indeed, the inverse of is
(2.27) |
An important property of is the existence of unique minimal representatives, analogous to the reduced word in a free group. The following result is proved in Section B.1.
Proposition 2.12.
An element has a unique minimal representative
(2.28) |
This is a word with the property that all , and every consecutive pair is linearly independent in . We use the subscript to denote the minimal representative.
Consider the map of sets given by the inclusion . This has the property that if we restrict to a 1-dimensional subspace , then is a group homomorphism. This gives rise to a universal property for .
Lemma 2.13.
Let be a vector space and let be a group. Let be a map which restricts to a group homomorphism on subspaces of dimension 1, and satisfies , where is the identity. Then, there exists a unique group homomorphism such that .
Proof.
Uniqueness of is immediate because is generated by . For existence, by the universal property of free monoids, there exists a unique monoid morphism . First, note that , where the second equality holds by assumption. Second, if are contained in a 1-dimensional subspace, we have
(2.29) |
Therefore, this descends to a homomorphism such that .
∎
Corollary 2.14.
The group of piecewise linear paths defines a functor
(2.30) |
Proof.
Suppose is a linear map. Then restricts to a group homomorphism on each one-dimensional subspace. By the universal property in Lemma 2.13, there is a unique group homomorphism such that . This uniqueness implies functoriality. ∎
For later use, we will also consider the notion of the span of a path.
Corollary 2.15.
Let be the set of linear subspaces of . There is a well-defined map
(2.31) |
where is the minimal representative of . If is a linear map, then
(2.32) |
The universal property of Lemma 2.13 provides an effective method for constructing homomorphisms out of . For example, the identity map automatically extends to a homomorphism , and the corresponding action groupoid is the piecewise linear analogue of . Next, consider the map defined by , where is the equivalence class of the path
where is a surjective map with sitting instants. Since restricts to a homomorphism on -dimensional subspaces of , by Lemma 2.13, it induces a group homomorphism,
(2.33) |
which we call the realization map. In fact, as we vary , these maps fit into a natural transformation
(2.34) |
Lemma 2.16.
For all , the realization map is injective.
Proof.
Let be a minimal representative, and suppose is trivial. The path can only be thinly null homotopic if there are retracings. Since the representative is minimal, this cannot occur if . Hence . ∎
Finally, consider the map , which can be expressed explicitly as
(2.35) |
Again, restricts to a group homomorphism on 1-dimensional subspaces of , and so by Lemma 2.13, we obtain a homomorphism
(2.36) |
which we call the piecewise linear path signature. It factors through the path signature by definition, and this fact holds at the level of natural transformations.
Proposition 2.17.
The maps , , and are natural transformations which factor as
(2.37) |
The construction of can be viewed as a linear algebraic analogue of the construction of a free group. In fact, the following result shows that free groups can be embedded into .
Proposition 2.18.
Let be a map of sets with the property that and are linearly independent if (in particular, for all ). Then the induced homomorphism is injective.
Proof.
Let be the unique homomorphism that restricts to on . Let be an element, which can be expressed uniquely as a reduced word in :
(2.38) |
where all are non-zero integers, and for . Then
(2.39) |
By assumption, each vector in the list is non-zero, and consecutive pairs are linearly independent. Hence, is a minimal representative by Proposition 2.12. This implies that the kernel of consists of the empty word, implying that the homomorphism is injective. ∎
We end this section by observing that the piecewise linear signature is unique, once we take into account the starting point of a path. Let be the category of affine spaces, whose objects are finite dimensional vector spaces, but whose morphisms are affine-linear maps. Given an affine space , define action groupoids
(2.40) |
where the actions are respectively defined by the homomorphism and the truncation , via the additive action of on itself. Along with the thin fundamental groupoid, these define functors from the category of affine spaces to the category of groupoids
(2.41) |
Furthermore, the maps , , and give rise to natural transformations
(2.42) |
which restrict to the identity on the objects of the groupoids.
Proposition 2.19.
There is a unique natural transformation
(2.43) |
called the piecewise linear groupoid signature, which restricts to the identity on objects.
Proof.
It suffices to show that any natural transformation which is the identity on objects must be given by . First, let be a -dimensional vector space. Then is an isomorphism and hence , the terminal object in the category of groupoids over . Therefore, we have equality of components . Now let be a general vector space. Elements of can be factored into products of elements of the form , for . Hence, it suffices to show that for all such pairs. Given , define the affine linear map
(2.44) |
Then , where . Therefore, using the naturality of and , we obtain
(2.45) | ||||
(2.46) | ||||
(2.47) |
where the equality in the second line uses . ∎
Remark 2.20.
Piecewise linear paths are dense in , the paths starting at the origin, equipped with the Lipschitz topology. Therefore, the groupoid signature is the unique continuous natural transformation which is the identity on objects.
3. Surface Holonomy and the Surface Signature
In this section, we introduce surface holonomy in the smooth setting, and discuss the surface signature introduced in [34], and further developed in [38, 18]. We begin with some conventions and notation for surfaces.
Definition 3.1.
A surface has sitting instants if there exists some such that
(3.1) |
Unless otherwise specified, we assume all surfaces have sitting instants.
For a surface , the bottom, right, top, and left boundary paths of are
(3.2) |
Furthermore, we define333Note that since the surface has sitting instants, the resulting boundary paths and boundary loop also have sitting instants. the boundary loop of using the counter-clockwise convention to be
(3.3) |
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x4.png)
The sitting instants ensure that for horizontally and vertically composable smooth surfaces , such that and , their horizontal and vertical compositions
(3.8) |
are also smooth, so that .
3.1. Thin Homotopy of Surfaces
We begin by considering the 2-dimensional notion of thin homotopy defined by generalizing the rank condition (R1).
Definition 3.2.
Two smooth surfaces are thin homotopy equivalent, denoted , if there exists a corner-preserving smooth homotopy between and such that
-
•
(homotopy condition) and ;
-
•
(thin homotopic boundaries) the four sides of the homotopy are thin homotopies between the four boundary paths of and ,
-
•
(thinness condition) , where is the differential of .
Remark 3.3.
In this article, we use to parametrize surfaces, which allows for simple definitions of horizontal and vertical concatenation, as in Equation (3.8). In Definition 3.2, we use corner-preserving maps. This will allow us, as in (3.12), to retain the simple formulation of concatenation for thin equivalence classes. We emphasize that we have chosen to use corner preserving maps because it simplifies the definition of the algebraic operations, but that it should be possible to relax this requirement.
Similar to the case of paths, thin homotopy equivalence encodes reparametrizations and local cancellations such as folding:
-
(1)
Reparametrization. Given a surface , and a corner and edge preserving reparametrization , then .
-
(2)
Folding. Given a surface , we say that a fold is a region of a surface of the form or . Two surfaces which differ by folds are thin homotopy equivalent.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x5.png)
However, thin homotopy of surfaces may yield non-local cancellations, as the following shows.
Proposition 3.4.
Consider a map which factors as
(3.9) |
where is a map which covers the sphere and sends the boundary of to a basepoint , is the double cover obtained by identifying antipodal points, and is a smooth map. Then is thinly null homotopic.
Proof.
We represent as the quotient of a disc by the antipodal identification on the boundary. We compose the antipodal projection map with a map which collapses the boundary circle to a single point, to obtain as in Figure 1.

The map has degree and hence is null-homotopic by the Hurewicz theorem. Therefore, given a map , the map is thinly null-homotopic. Indeed, this homotopy is a ‘fold-cancellation’ given by ‘opening’ the hole at the top of the sphere in order to contract the map to a point. Thus, to show that is thinly null-homotopic, it suffices to construct a thin homotopy between and , for some map . To this end, let be the map parametrizing the top boundary of the disc, and let be the corresponding loop in , which generates the fundamental group . The loop is contractible. Let be a smooth homotopy that contracts to a point.
It is possible to modify on a collar neighbourhood of the boundary of , so that its restriction to both the top half and to the bottom half of the disc agrees with . This can be pictured as in Figure 2

Let denote the new map. It is obtained by adding a surface, namely the image of , to the original map . Since it sends the boundary of the disc to a single point it factors as . Furthermore, the new map is homotopic to through a homotopy which is restricted to lie within the image of . Hence and are thinly homotopic to each other. This completes the proof. ∎
Proposition 3.4 shows that, in contrast to the case of paths, it is possible for an immersed surface to be thinly null homotopic. Indeed, it suffices to take the map from the Proposition to be an embedding. Hence, there are thin homotopies that do not only involve reparametrizations and fold cancellations. Furthermore, Proposition 3.4 also shows that, unlike in the case of paths, surfaces do not admit unique reductions via local cancellations. As a result, in order to see that a surface is thinly null-homotopic, it may be necessary to ‘backtrack’ by introducing new surfaces. This means that techniques used to show the injectivity of the path signature, which usually involve taking the reduced form of a path, will not generalize to surfaces.
Finally, one upshot of the present discussion is that generalizing the equivalent definitions of thin homotopy for paths from Theorem 2.11 to the case of surfaces is not entirely straightforward. Indeed, the naive generalizations may not be true. For example, the straightforward analog of the image condition (I1) fails, as the following corollary shows.
Corollary 3.5.
Let be defined as in Proposition 3.4, with an embedding. There does not exist a null-homotopy of such that .
Proof.
Assume that such a homotopy exists. Since it is contained in the embedded surface , it lifts to a null homotopy of . This is a contradiction since is a generator for . ∎
In addition, it is not immediately clear how to generalize the word condition (W1), the tree condition (F1), and the analytic condition (A1). One of the contributions of this article is to suggest a way to adapt all the conditions from Theorem 2.11 to the two-dimensional setting, which is stated in Theorem 6.26. However, our first task will be to study the generalization of the signature condition (S1) by first introducing the surface signature. We will begin by introducing the required algebraic structures and the concept of surface holonomy.
3.2. Double Groupoids and Crossed Modules of Groups
The algebraic structure inherent in the composition of surfaces is most naturally encoded using the formalism of double groupoids. However, in this paper, we prefer to use the equivalent concept of crossed modules, since they are more convenient to work with algebraically. In this section, we briefly recall this equivalence and construct the thin crossed module of surfaces.
Definition 3.6.
[44, Theorem 2.13] The thin fundamental double groupoid is an edge symmetric double groupoid where
(3.10) |
The thin double group is an edge symmetric double group where
(3.11) |
As in the case of paths, our convention will be to use surfaces based at the origin as representatives of the translation equivalence classes. Two thin homotopy classes with representatives , where and , are horizontally composable if . Let be a thin homotopy between and . We define the horizontal composition of and to be
(3.12) |
Next, we define the horizontal inverse of a surface in by
(3.13) |
The vertical operations can be defined in an analogous manner; see [44, Section 2.3.3] for details. In this article, we will primarily work with a related algebraic structure called a crossed module.
Definition 3.7.
A pre-crossed module of groups,
(3.14) |
is given by two groups , a group morphism and a left action of on by group automorphisms, denoted elementwise by for . These data are required to satisfy
(3.15) |
We say is a crossed module of groups if it also satisfies the Peiffer identity
(3.16) |
A (pre-)crossed module of Lie groups is the same as above, except and are Lie groups, and all morphisms are smooth. Given another (pre-)crossed module , a morphism of (pre-)crossed modules consists of group homomorphisms and such that, for all and , we have
The categories of crossed modules of groups and Lie groups are respectively denoted and .
Crossed modules of groups are equivalent to double groups, which are defined to be edge-symmetric double groupoids with thin structure and with a single object. We denote the category of double groups by .
Theorem 3.8.
[8, Section 6.6] There is an equivalence of categories between and .
Here we will consider two examples of crossed modules which will be used later.
Definition 3.9.
The thin crossed module of a vector space is the crossed module associated to the thin double group . In particular, let
(3.17) |
The group operation in is defined as the reversed444The group operation is defined via the equivalence between and in [8, Section 6.6], and is determined by the convention (starting point and orientation) used for the boundary given in Equation 3.3. We use the same convention as [44, 39, 38, 18], which results in the reversed ordering. horizontal composition (3.12),
(3.18) |
and the inverse is given by the horizontal inverse (3.13). The crossed module boundary map is given by the boundary of the surface from (3.3)
(3.19) |
For a path with , we define the degenerate surface by
(3.20) |
Finally, we define a left action of on by
(3.21) |
where we must translate the surfaces such that they are composable.
This indeed forms a crossed module by [8, Proposition 6.2.4]. Furthermore, linear maps preserve thin homotopy equivalence of maps, and thus we obtain a functor
(3.22) |
Definition 3.10.
[8, Section 2.2] Let be a based 2-dimensional CW complex. The fundamental crossed module of is given by
(3.23) |
where is the fundamental group of the -skeleton , and is the relative homotopy group of with respect to the 1-skeleton . The later is defined to be the group of homotopy classes of maps such that the boundary of is sent to , the corners are sent to , and such that the left, bottom and right boundary paths are null homotopic in , following the convention of (3.17). The multiplication in is given by reversed horizontal concatenation , as in (3.18). This is equipped with the boundary map from (3.3) and an action of on given by
(3.24) |
where is the degenerate surface defined in (3.20).
3.3. Crossed Module of Lie Algebras and 2-Connections
In order to define surface holonomy and the surface signature, we will require the infinitesimal version of crossed modules.
Definition 3.11.
A pre-crossed module of Lie algebras
consists of Lie algebras and , a morphism , and an action of on by derivations. In other words, the action satisfies
for all and . Furthermore, these data are required to satisfy
(3.25) |
We say is a crossed module of Lie algebras if it also satisfies the Peiffer identity
(3.26) |
Suppose is another (pre-)crossed module of Lie algebras. A morphism of (pre-)crossed modules consists of two Lie algebra morphisms and such that for all and ,
(3.27) |
The category of crossed modules of Lie algebras is denoted .
The notion of a 2-connection is defined in terms of a crossed module of Lie algebras.
Definition 3.12.
[43, Definition 2.16, Proposition 2.17] Let be a crossed module of Lie algebras. A 2-connection valued in over is a pair , such that is a -valued -form and is a -valued -form. The 1-curvature and 2-curvature are respectively defined as follows
(3.28) |
where is the curvature of from (2.6) and is defined by
(3.29) |
The 2-connection is called fake-flat (or semi-flat) if its 1-curvature vanishes, . We will always work with fake-flat 2-connections.
We say that a 2-connection is translation-invariant if it has the form
(3.30) |
where , , and are linear coordinates on . In this case, we can represent the 2-connection as a pair of linear maps
(3.31) |
The curvature forms then simplify to
(3.32) |
3.4. Surface Holonomy
Here, we discuss surface holonomy, the 2-dimensional generalization of parallel transport. For a surface , the -tail path of is
(3.35) |
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x8.png)
Definition 3.13.
[44, Equation 2.13] Let be a crossed module of Lie groups with associated crossed module of Lie algebras . Let be a 2-connection valued in and let . Consider the following differential equation for
(3.36) |
where the tail path is defined in (3.35) and is the parallel transport of along , as given in Definition 2.3. We define the surface holonomy of along to be
(3.37) |
In [44, Theorem 2.32], it is shown that the surface holonomy gives rise to a morphism of double groupoids. When the 2-connection is translation-invariant, we apply Theorem 3.8 to state a version of the theorem in terms of the crossed module .
Theorem 3.14.
[44, Theorem 2.32] Let be a translation-invariant 2-connection valued in . Then the maps
(3.38) |
define a morphism of crossed modules.
This theorem has two main implications.
-
(1)
Surface holonomy is invariant with respect to thin homotopies of surfaces.
-
(2)
Surface holonomy respects concatenation of surfaces.
3.5. Free Crossed Module of Lie Algebras
For the next two sections, we discuss the surface signature, which was originally introduced by Kapranov [34], and recently studied from the analytic perspective of irregular surfaces in [38, 18]. Generalizing the path signature, the surface signature is defined to be the surface holonomy of a universal -connection valued in a free crossed module. Here, we begin with an overview of the essential properties of free crossed modules of Lie algebras, and refer the reader to Appendix C for further details and proofs.
Let denote the comma category associated to the functors and . An object of is given by the data of a vector space , a Lie algebra , and a linear map . A morphism consists of a linear map and a Lie algebra morphism such that as linear maps. There exists a natural forgetful functor
(3.39) |
The free crossed module functor
(3.40) |
is defined to be the left adjoint. In the following, we describe this functor in detail. Given an object of , we first build the free -representation on , which is , where is the universal enveloping algebra of , where the action is given by
(3.41) |
and where the linear map is lifted to a morphism of -representations as follows
(3.42) |
We define the Peiffer subspace of to be
(3.43) |
By taking the quotient with respect to this subspace, we obtain the free crossed module of Lie algebras generated by ,
(3.44) |
where the Lie bracket of is defined by
(3.45) |
This satisfies the following universal property, which is proved in Theorem C.8.
Theorem 3.15.
Let , , and
(3.46) |
Then there is a unique morphism of crossed modules
(3.47) |
such that and .
Now consider the functor
(3.48) |
which sends a vector space to
(3.49) |
Composing with the free crossed module functor gives
(3.50) |
Given a vector space , the free crossed module of Lie algebras generated by is the result of applying this functor, and is given by
(3.51) |
where
(3.52) |
3.6. The Surface Signature
Following Kapranov [34], we consider the universal translation-invariant 2-connection. This is a -connection over valued in . The translation invariant -valued 1-form is the one defined in (2.10), and hence corresponds to the identity endomorphism of , . Similarly, the translation invariant -valued 2-form can be understood as the identity endomorphism of
(3.53) |
Alternatively, and can be viewed as the linear maps given by including generators
(3.54) |
In coordinates, they are given by
(3.55) |
We follow a similar procedure as for path signatures and consider truncations of via the -lower central series of , defined by
(3.56) |
This allows us to define the -truncated free crossed module as follows
(3.57) |
We can integrate as the Lie group of formal exponentials , and consider the projective limit , where
(3.58) |
This defines a crossed module of groups
(3.59) |
with a corresponding crossed module of Lie algebras
(3.60) |
Remark 3.16.
In [38], it is shown that by considering the free crossed module of associative algebras, one can construct the completion in a way that is similar to the construction of in the completed tensor algebra. Furthermore, one can also construct analytic completions in this setting by using Banach or topological algebras. As we do not require these details here, we will work with formal completions.
Definition 3.17.
Let . For , the -truncated surface signature of , denoted , is defined to be the surface holonomy, as defined in Definition 3.13, of the 2-connection obtained by projecting the universal 2-connection from (3.55) onto . The surface signature of is defined to be the projective limit
(3.61) |
As in the case of the path signature, there is a universal property [38, Theorem 4.29] which implies that the surface signature fits into a natural transformation
(3.62) |
We note that the -truncated group is equipped with a natural topology induced by the Lie algebra via the exponential. The continuity of the surface extension theorem from [38, Theorem 5.40] implies that the surface signature is continuous.
Proposition 3.18.
[38, Theorem 5.40] Let be the space of based smooth surfaces such that . The -truncated surface signature
(3.63) |
where is equipped with the Lipschitz norm, is continuous.
4. Abelianization
In this section, we establish our first characterization of the kernel of the surface signature. This is done in two steps. First, we consider a gauge transformation which converts the universal 2-connection of (3.55) into an abelianized 2-connection. This is a connection that is valued in the center of . Using the representation theory of , we produce an explicit expression for the -curvature of this abelianized -connection. Second, we relate the surface signature to an integral of this abelianized 2-curvature. Using these two results, we show in Theorem 4.12 that the surface signature of a closed surface is encoded by integration over all polynomial 2-forms.
4.1. Gauge Transformations and Abelianization
We begin with the general definition of gauge transformations for 2-connections.
Definition 4.1.
Let and be its associated crossed module of Lie algebras. A gauge transformation is a pair , where and . The two components act on a 2-connection valued in in the following way
(4.1) | ||||
(4.2) |
where is defined in (3.29). As a convention, acts first by , and then by :
(4.3) |
By direct computation, the two components of the curvature are given by
(4.4) |
and
(4.5) |
In particular, for a fake-flat 2-connection, where , we obtain
(4.6) |
An interesting feature of higher gauge theory is that given a fake flat 2-connection over a contractible space, one can always find a gauge transformation such that , with valued in the kernel of [49, Theorem 4.3] (see also [23] and [54] for similar results). We consider this construction for the universal 2-connection.
Let be the universal 2-connection from (3.55). First, we define by exponentiating a Lie algebra valued function
(4.7) |
Note that . Then, we have
(4.8) | ||||
(4.9) | ||||
(4.10) | ||||
(4.11) | ||||
(4.12) |
where
(4.13) |
Note that has the property that
(4.14) |
Now, if we let
(4.15) |
then we have
(4.16) |
Thus, the transformed 2-connection is
(4.17) |
Note that because fake-flatness is preserved by gauge transformations, is valued in the completion of , which is an abelian Lie algebra.
While the expression (4.17) for the abelianized 2-connection is quite complicated, its 2-curvature can be computed explicitly. The 2-curvature of the universal 2-connection is
(4.18) |
where
(4.19) |
Then, applying the gauge transformation to the 2-curvature using (4.6), we obtain
(4.20) |
Furthermore, because the abelianized 1-connection is trivial, the Bianchi identity for the 2-connection [44, Proposition 2.20] shows that , and thus
(4.21) |
4.2. Polynomial Differential Forms and Currents
To understand , we make use of the isomorphism between and the space of closed polynomial currents established by Kapranov [34]. In this section, we expand on the exposition in [34] and describe the spaces of polynomial differential forms and currents as -representations. Here, we will consider to be a finite-dimensional -vector space, for or . Polynomial differential forms and their completions are naturally graded by their total weight . We define
(4.22) |
equipped with the usual exterior differential . The polynomial currents and their completions are defined by taking the graded dual of as follows
(4.23) |
To simplify notation, we will often write and . Differential forms and currents are naturally equipped with the structure of -representations. An element acts on a current as follows
(4.24) |
and acts on by pullback as follows
(4.25) |
The exterior differential is -equivariant. Note also that the action of the subgroup of scalars induces a grading on differential forms and currents which agrees with the weight grading defined above.
Although the polynomial forms and currents are naturally dual to each other, the naive pairing between them is not -equivariant. Here, we will use the Cartan calculus to define the “correct” pairing which is implicitly used in [34]. In the following, we view as constant vector fields on ; we note that these vector fields commute. These vector fields act on differential forms in two ways. First, for any , we can take the Lie derivative , and for constant vector fields, this is commutative,
(4.26) |
Second, we can take the interior product , which satisfies
(4.27) |
For , we use these operations to define a map where
(4.28) |
Lemma 4.2.
For , the map is -equivariant. In particular, for , , , we have
(4.29) |
Proof.
First, given and , the pullback satisfies
(4.30) |
Then, by the Cartan formula , we have
(4.31) |
Thus, by the definition of in (4.28), we have
(4.32) | ||||
∎
We also record a formula relating the exterior derivative and .
Lemma 4.3.
For , where , and , we have
(4.33) | ||||
Proof.
This is immediate by the properties , , and . ∎
Next, we assemble all of the into a map
(4.34) |
For the final step in defining the equivariant pairing, consider the inclusion of the origin , which is trivially -equivariant. This induces a trivially -equivariant pullback
(4.35) |
Proposition 4.4.
The pairing
(4.36) |
is non-degenerate for , vanishes for , and is -equivariant.
If necessary, we use to specify the degree. We can describe this pairing explicitly in coordinates. Let be a basis of and let be the dual basis of . Given , we define and similarly for . Given an increasing index set , we define . Similarly, we define . In this way, we obtain the bases
(4.37) |
of and , respectively. In terms of these coordinates, the pairing satisfies
(4.38) |
Using the pairing, we can now define the codifferential
(4.39) |
for any and . It is -equivariant. The closed forms and closed currents are defined as usual to be
(4.40) |
respectively. By the Poincare lemma, the cochain complex of differential forms is acyclic, and thus by duality, the chain complex of currents is also acyclic. Hence, using the pairing, we have
(4.41) |
where denotes the graded dual. Given , where , and , where , the pairing between closed forms and closed currents is
(4.42) |
In what follows, we will need the duality between closed -currents and closed -forms and it will be useful to have explicit dual bases. Given index list , and indices , such that , define
(4.43) |
where , and denotes the number of times the index appears in .
Lemma 4.5.
There is a natural isomorphism
(4.44) |
where, on the right-hand side, we are taking the graded dual. The closed currents and closed forms , for , give dual bases in weight .
Proof.
The space of closed weight currents is the irreducible representation of corresponding to the hook Young diagram of size . It follows that the currents give a basis. Taking the pairing in (4.42), we have
(4.45) | ||||
(4.46) | ||||
(4.47) |
∎
Finally, we conclude this section with the following relationship between the free crossed module from (3.51) and closed -currents; see [18, Appendix D] for further exposition.
Theorem 4.6.
[34] The symmetrization map is a -equivariant map which descends to define a Lie algebra map
(4.48) |
which identifies the abelianization
(4.49) |
Furthermore, the restriction of to defines an isomorphism of -representations
(4.50) |
Given indices , and , such that , define
(4.51) |
Under the isomorphism from Theorem 4.6, we see that
(4.52) |
Hence, we immediately conclude that define a basis of .
4.3. Computing the Abelianized Curvature
We will use the relationship between and closed polynomial currents, along with their representation-theoretic properties, to explicitly compute the abelianized curvature . As we will be using Schur’s lemma for complex representations, it is convenient to immediately consider the complexification . Because is defined over , this will not affect our final calculation of .
Denote the complexified space of currents and forms by and , respectively. Because the isomorphisms from Lemma 4.5 and Theorem 4.6 are -equivariant, they preserve the weight grading, and therefore extend to the completions. Hence, we obtain an isomorphism of -representations , which can be extended to
(4.53) |
Under this isomorphism, the curvature from (4.18), which we denote by , is sent to
(4.54) |
By Lemma 4.5, the closed currents and closed -forms form dual bases. Hence, is identified with the identity map for the weight closed forms
(4.55) |
Here, denotes the space of equivariant morphisms of , a -representation. Recalling that this is precisely the -invariant subspace of the representation , we conclude that (and hence ) is -invariant.
Lemma 4.7.
The abelianized curvature is -invariant. Therefore
(4.56) |
Proof.
We recall from Equation (4.20) that
(4.57) |
Consider as the action of on . By Corollary C.10, we note that acts on in a way which preserves the crossed module structure. Therefore, acts on both and , and the action is -equivariant:
(4.58) |
where , and . The element is -invariant, since it corresponds to the identity . Because is also -invariant, the same is true for and thus for as well. In other words,
(4.59) |
∎
The upshot of Lemma 4.7 is that the abelianized curvature can be decomposed as
(4.60) |
where each . By [28, Section 8.2, Theorem 2], each representation is irreducible. Therefore, by Schur’s lemma,
(4.61) |
for a constant . In the following theorem, we verify that this constant is .
Theorem 4.8.
The abelianized curvature is
(4.62) |
Proof.
The element has weight in the form component. Hence, it is given by
(4.63) |
To determine the constant , it suffices to compute the coefficient of
(4.64) |
This is easily seen to be
(4.65) |
By Lemma 4.5, this is the closed current dual to . Hence, . ∎
Corollary 4.9.
The abelianized curvature is
(4.66) |
4.4. Abelianized Surface Signature
Now, we use the surface holonomy with respect to the abelianized 2-connection to compute the signature of closed surfaces,
(4.67) |
To this end, we make use of the relatively simple expression for the abelianized -curvature from Corollary 4.9 in order to simplify this calculation. The following lemma tells us that it suffices to integrate this -curvature over a volume.
Lemma 4.10.
Let and let be a volume such that all boundaries are sent to except the top, which is equal to ,
(4.68) |
Let be an abelian 2-connection valued in , with 2-curvature . Then, the surface holonomy is given by
(4.69) |
Proof.
This is a direct consequence of [44, Theorem 2.30]. ∎
Next, the following result shows how the gauge transformation acts on the surface holonomy.
Proposition 4.11.
[44, Corollary 4.9] Let be a smooth closed surface based at the origin, let be a -connection and let be a gauge transformation. Then,
(4.70) |
Putting this all together, we obtain the following result.
Theorem 4.12.
For any smooth closed surface based at the origin, the surface signature is given by
(4.71) |
Furthermore, using to embed into the space of formal -currents , the surface signature is given by the following expression
Proof.
Let denote the universal -connection and recall that its surface holonomy is denoted . Let denote the abelianization gauge transformation from (4.7) and (4.15) so that . Given the surface , let be a volume as in Lemma 4.10. Then
(4.72) | ||||
(4.73) |
where the first equality follows from Proposition 4.11 and the fact that , the second equality follows from Lemma 4.10, the third equality follows from Corollary 4.9, and the fourth equality follows from Stokes’ theorem. Note that this expression for the signature has the form , for a dual pair of bases of and . Because is closed, we have for all by Stokes’ theorem. Therefore, using to embed into , our expression for also has the same form where now are a dual pair of bases of and . Our final expression for the signature then follows by using the dual bases from Equation 4.38. ∎
Corollary 4.13.
Let . Then, if and only if
(4.74) |
for all compactly supported 2-forms .
Proof.
By Theorem 4.12, the condition that is equivalent to for all polynomial -forms. Since polynomial 2-forms are dense in the compactly supported 2-forms , we obtain the desired result. ∎
Remark 4.14.
Let be a thinly null-homotopic path. The existence of a height function from the analytic condition (A1) is used to detect cancellations: whenever with , the restriction is also thinly null homotopic. Due to the fact that cancellations in 2-dimensional thin homotopy can be non-local, a direct analogue of the height function does not exist. Instead, the condition in Corollary 4.13 uses the integration of compactly supported forms to detect both local and non-local cancellations which occur on a surface . Thus, we interpret this condition to be the generalization of (A1).
5. Piecewise Linear Surface Signature and Decompositions
In this section, we algebraically construct a crossed module of piecewise linear surfaces, extending from Section 2.4. This leads to an algebraic definition of the surface signature, and furthermore, a decomposition of the signature into abelian and boundary components.
5.1. Free Crossed Modules of Groups
We will begin by discussing free crossed modules of groups. This will be used in Section 6, but it also serves as motivation for the construction of the piecewise linear crossed module in the following section. Our aim is to show a global universal property, analogous to that for Lie algebras in Theorem 3.15.
Given a group , let be the subcategory of crossed modules
(5.1) |
where is fixed. Let denote the category of set functions , where is a set and is the fixed group. There is a natural forgetful functor, along with a free functor as a left adjoint,
(5.2) |
The free crossed module generated by can be constructed as follows [8, Proposition 3.4.3]. Let be the free group generated by . It is equipped with a left -action by automorphisms which is defined on the generators by . Define a homomorphism by the following map on generators
(5.3) |
This is the free pre-crossed module on . Then, the free crossed module is the quotient by the Peiffer identity,
(5.4) |
The unit of the adjunction provides a map given by sending to the equivalence class of .
Next, let denote the comma category associated to the functors and . An object of is given by the data of a set , a group , and a map of sets . A morphism consists of a set map and a group homomorphism such that the following diagram commutes
(5.5) |
There exists a natural forgetful functor . Free crossed modules of groups satisfy the following universal property.
Theorem 5.1.
Let , , and
(5.6) |
Then there is a unique morphism of crossed modules
(5.7) |
such that and .
Proof.
Given the group homomorphism , we consider the pullback crossed module [8, Definition 5.1.1]
(5.8) |
and define the map given by , which gives a morphism of crossed modules
(5.9) |
We define a map by . This defines a map
(5.10) |
in such that . Then, using the universal property of in , there is a unique map
(5.11) |
which satisfies in . We define
(5.12) |
which is a morphism of crossed modules and it satisfies
(5.13) |
so is the desired morphism. The map is unique because the map coincides with the unique factorization of the morphism through the pullback crossed module [8, Theorem 5.1.2]. ∎
5.2. Crossed Module of Piecewise Linear Surfaces
Here, we will generalize the construction in Section 2.4 and define a linear algebraic version of the free crossed module construction. Consider , and define the group homomorphism
(5.14) |
which defines the endpoint of a PL path. We define the group of piecewise linear loops by . Furthermore, we define the set of planar loops by
(5.15) |
where we use the notion of span from (2.31) using the minimal representative of . We note that the span of any nontrivial planar PL loop must be two-dimensional. There is a natural inclusion
(5.16) |
Definition 5.2.
A kite is a pair , consisting of a tail path and a planar loop . We define the set of kites by
(5.17) |
The pre-crossed module of piecewise linear surfaces is defined to be
(5.18) |
is the quotient of the free group generated by kites subject to the following relations:
-
(PL1.1)
, where , if
(5.19) -
(PL1.2)
, where denote the empty words in and respectively; and
-
(PL1.3)
for any such that .
Remark 5.3.
Here, we do not prescribe the planes on which the kites and their compositions are defined. However, assuming that the loops are non-trivial, in Corollary B.10, we show that if (PL1.1) holds, then . Furthermore, in Lemma B.7, we show that if (PL1.3) holds, then .
With these relations, we note that , since
(5.20) |
where here, .
Remark 5.4.
As with the 1-dimensional case, we can represent elements of as words in , without considering the formal inverses. In particular, we can make the equivalent definition of in terms of the free monoid, which we will use interchangeably,
(5.21) |
Next, we define the boundary map on the generators by
(5.22) |
and the action of on the generators of by
(5.23) |
Lemma 5.5.
The structure is a pre-crossed module of groups.
In order to obtain a crossed module, we quotient by the standard Peiffer identity
(5.24) |
Definition 5.6.
The piecewise linear crossed module is defined by
(5.25) |
There is a natural inclusion defined by
(5.26) |
In fact, for 2-dimensional vector spaces, this is an isomorphism.
Lemma 5.7.
If is 2-dimensional, then are isomorphic as groups.
Proof.
Since is two-dimensional, all loops are planar, and so , which is a group. The map is a group homomorphism by (PL1.1). In fact, it is an isomorphism with inverse given by the boundary map . Indeed, the identity is immediate, and follows by (PL1.3). Furthermore, the Peiffer identity (5.24) is already implied because it holds in . ∎
For where , is not a group. However, by Lemma 5.7, the restriction to a 2-dimensional subspace is a group . For the remainder of the article, we will make the identification of crossed modules
(5.27) |
Furthermore, the restriction of the inclusion in (5.26) yields a morphism of crossed modules
(5.28) |
where is the inclusion induced by . This leads to the following universal property.
Theorem 5.8.
Let be a vector space, and suppose
(5.29) |
is a crossed module of groups. Let be a function such that and for any 2-dimensional subspace , the restriction of to yields a morphism of crossed modules,
(5.30) |
where is the inclusion from (5.28). Then, there exists a unique group homomorphism such that
(5.31) |
is a morphism of crossed modules and .
Remark 5.9.
Note that if then and there is nothing to prove.
Proof.
First, we define a map on the generators by
(5.32) |
We verify that this descends to by checking the relations (PL1.1)-(PL1.3). For (PL1.2), we have
(5.33) |
by assumption. Next, for (PL1.3), suppose , and is a nontrivial planar loop with span such that . By Lemma B.7, . Then, since restricts to a morphism of crossed modules on ,
(5.34) |
Finally, we verify (PL1.1). Suppose are nontrivial kites such that , where . By Corollary B.10, there exists with such that . Then, using , equation (5.34), and the assumption that restricts to a homomorphism on , we have
(5.35) | ||||
(5.36) | ||||
(5.37) | ||||
(5.38) |
Therefore, the map descends to a map , which is the unique morphism of pre-crossed modules which satisfies , where is defined by . The final step is applying the crossed module functor by quotienting out the Peiffer identity to obtain the desired unique morphism . ∎
We can also enhance this universal property by varying the base.
Corollary 5.10.
Let be a vector space and suppose is a crossed module of groups. Let be a group homomorphism and let be a function such that and for any 2-dimensional subspace , the restriction of to is a morphism of crossed modules,
(5.39) |
Then, there exists a unique group homomorphism such that
(5.40) |
is a morphism of crossed modules and .
Proof.
We consider the pullback crossed module of with respect to , defined by
(5.41) |
It is equipped with a morphism of crossed modules . There is a well-defined function given by sending a planar loop to . To see that this is well-defined, note first that . Furthermore, if is non-trivial, it determines a -dimensional subspace such that . Then, since restricts to a morphism of crossed modules on , we see that . Note also that .
We wish to apply the universal property of Theorem 5.8, and for this we must verify that for every -dimensional subspace , the restriction defines a morphism of crossed modules . Now, because is assumed to be a morphism of crossed modules, by [8, Theorem 5.1.2] there exists a unique factorization through the pullback crossed module
(5.42) |
But a simple calculation shows that , thus verifying the condition. As a result, by the universal property, there is a unique homomorphism such that
(5.43) |
is a homomorphism of crossed modules and . Finally, we define to obtain our desired morphism .
∎
Corollary 5.11.
The piecewise linear crossed module defines a functor
(5.44) |
Proof.
Let be a linear map. Then, by Corollary 2.14, there exists a homomorphism . By Corollary 2.15, preserves the span of paths. Hence, if a path is planar, then so is . As a result, we can also define a map by
(5.45) |
Given a 2-dimensional subspace , the restriction of the pair to is a morphism of crossed modules . Hence, by Corollary 5.10, there is a unique homomorphism such that
(5.46) |
is a morphism of crossed modules and such that . The component is functorial by Corollary 2.14, and the component is functorial because of the uniqueness. ∎
Remark 5.12.
Let be the inclusion of a 2-dimensional subspace into . Then the restriction of from (5.28) is .
5.3. Piecewise Linear Surface Signature
In this section, we continue the generalization of Section 2.4 by extending the realization and piecewise linear signature natural transformations to the piecewise linear crossed module using the universal property in Corollary 5.10. In fact, we will show that for a certain class of planar functors, there is a unique natural transformation which extends the path constructions.
Definition 5.13.
A functor is planar if
-
•
is trivial when ,
-
•
is trivial when , and
-
•
is injective when .
Lemma 5.14.
The piecewise linear crossed module is a planar functor.
Proof.
The cases of are immediate, and the condition for is given by Lemma 5.7. ∎
The following lemma is the main result used in this section.
Lemma 5.15.
Let be a planar functor and let be a natural transformation with the property that when . Then, there exists a unique natural transformation which extends ,
(5.47) |
Proof.
Let denote the subcategory of vector spaces of dimension at most . We will first construct on this subcategory. When , the components are uniquely determined because is trivial. Now assume that In this case, the component must make the following diagram commute
(5.48) |
The image of the map is contained in the image of the injective map . Hence, the desired map exists and it is unique. The pair is automatically a map of crossed modules because and are injective.
To check naturality, let be a linear map. We need to verify that . Because and are planar functors, this is trivially satisfied when either or has dimension less than . Hence, we assume that . Because is a natural transformation, we have . Then, using the fact that and are morphisms of crossed modules, we have
(5.49) |
Hence, by injectivity of , we have .
Next, we extend to vector spaces of dimension at least . We will do this by applying the universal property of Corollary 5.10. Given the inclusion of a 2-dimensional subspace , we define the following morphism of crossed modules
(5.50) |
We assemble the over all 2-dimensional subspaces to obtain a map . Along with the homomorphism , this map satisfies the hypotheses of Corollary 5.10. Hence we obtain a unique homomorphism such that
(5.51) |
is a morphism of crossed modules and such that . In particular, , meaning that is natural with respect to inclusions of -dimensional subspaces. Since this property characterizes , we see that any natural transformation extending must be unique. Furthermore, it is straightforward to see that is natural with respect to linear maps where . Indeed, this is immediate if because then is trivial. Assuming that , factor as
(5.52) |
where . Therefore,
(5.53) |
where the first equality follows from naturality with respect to inclusions and the second equality follows from naturality in .
The final step in the proof is to verify naturality of for arbitrary linear maps . We may assume that . For each 2-dimensional subspace , let be the inclusion and consider the following diagram
(5.54) |
We know that the top and outer squares commute, and we wish to show that the bottom square commutes. For each , consider the morphism of crossed modules
(5.55) |
We assemble the over all 2-dimensional subspaces into a function . Along with the homomorphism , this map satisfies the hypotheses of Corollary 5.10. Hence, there is a unique homomorphism such that is a morphism of crossed modules, and . Hence, because both and satisfy these conditions, they must be equal. ∎
To use this result to extend the realization functor, we will require the following lemma.
Lemma 5.16.
Let be a pair of maps with equal corners for and such that each boundary path (3.2) is thin homotopy equivalent for . Then .
Proof.
Let by the thin homotopy between and for . Then is thin homotopy equivalent to the surface , where we glue the four thin path homotopies along the boundary as follows.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x10.png)
Then, the linear homotopy between and is thin as it must have rank at most and does not change the boundaries. Hence . ∎
Lemma 5.17.
For every planar loop , there exists a surface whose boundary is in the thin homotopy class of .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x11.png)
Proof.
For , define the function
(5.56) |
where is a reparametrization with sitting instants. Define by
(5.57) |
where is defined in (3.8), and the concatenations are defined from left to right (as is not associative). Note that the left, bottom, and right boundaries of are trivial, and it can be verified that has the correct boundary. ∎
Proposition 5.18.
Proof.
First, is trivial when , and is trivial when , as every surface in is thinly null homotopic by the linear homotopy. By Lemma 5.16, the crossed module boundary is injective when . Therefore, is a planar functor. Furthermore, when as every loop in can be filled in to obtain a surface by Lemma 5.17. Thus, by Lemma 5.15, there exists a unique extension . ∎
Next, we will show that the path signature also extends uniquely to a natural transformation.
Theorem 5.19.
There exists a unique natural transformation
(5.59) |
called the piecewise linear signature, which extends from Proposition 2.17.
Proof.
First, is trivial when , and is trivial when since . Now, consider a 2-dimensional vector space . By Lemma 4.5, we note that the kernel is trivial. Thus
(5.60) |
as Lie algebras and the boundary map on groups is injective, so is a planar functor. Furthermore, we claim that . Indeed, the signature of a closed loop must be trivial in degree , implying that . But we have just seen that . Therefore, . Thus, by Lemma 5.15, there exists a unique extension . ∎
By the uniqueness of these constructions, these natural transformations factor in the same way as for paths in Proposition 2.17.
Proposition 5.20.
The maps , , and are natural transformations which factor as
(5.61) |
This factorization implies that the smooth surface signature is uniquely determined.
Theorem 5.21.
The smooth signature is the unique natural transformation
(5.62) |
which extends from (2.23) and such that is continuous with respect to the quotient topology on induced by the Lipschitz topology on .
Proof.
Let be a continuous natural transformation which extends . Then is a natural transformation which extends . By Theorem 5.19, . Therefore, and agree on piecewise linear surfaces. Now let . There exists a sequence of piecewise linear surfaces such that in the Lipschitz topology. Indeed, let be a triangulation of such that the mesh size approaches . Define on all vertices , and extend linearly. Then, because is smooth and is compact, the derivatives of converge to those of , and thus . Because for all , we must have by continuity from Proposition 3.18. ∎
Remark 5.22.
The uniqueness results of Theorem 5.19 and Theorem 5.21 can be strengthen by keeping track of basepoints as in Proposition 2.19. Indeed, by upgrading , , and to crossed modules of groupoids defined on the category of affine spaces, we can show that the signature is uniquely characterized as the natural transformation extending the identity on objects.
5.4. Computational Methods for the Surface Signature
There are two main challenges in developing computational methods for the surface signature.
-
(1)
First, the standard definition of the surface signature via the surface holonomy equation in Definition 3.13 requires the solution to a complicated differential equation.
-
(2)
Second, the surface signature is valued in a group (or algebra) which lacks an evident natural choice of basis. This is in contrast to the path signature, which is valued in the completed tensor algebra and thus is equipped with a natural basis induced by a basis on the underlying vector space . Indeed, in [38] the surface signature is valued in a free crossed module of associative algebras. Even though this algebra is built out of tensor algebras, one must quotient by the Peiffer subspace. As a result, the choice of bases employed in [38] is non-canonical.
In the proof of Theorem 5.19, we observe through Lemma 5.15 that the signature of a planar surface is entirely determined by the path signature of its boundary. This already resolves the first problem mentioned above in the setting of piecewise linear surfaces: by leveraging the algebraic structure, we can compute the surface signature through composition of basic building blocks made up of the signatures of planar loops and their tail paths. In this section, we extend this approach by applying our previous results to develop a decomposition of the surface signature that enables tractable computational methods. Specifically, we demonstrate in Theorem 5.36 how the signature can be decomposed into a boundary component, valued in a subset of the usual completed tensor algebra, and an abelian component, valued in a vector space of formal currents. This resolves the second problem, as we can equip these vector spaces with canonical bases.
5.4.1. Decomposition of the PL Crossed Module.
We begin by constructing a decomposition of the piecewise linear crossed module . To do this, we will make use of a universal property for the group of piecewise linear loops , which we obtain by expressing as a quotient of a free group generated by triangular loops.
First, recall that the pair groupoid of , , is the groupoid whose set of objects is and such that there is a unique morphism between any two objects. Hence, the space of morphisms is . By sending the pair to the triangular loop with vertices we obtain a map of sets
(5.63) |
A simple computation shows that evaluates to the identity on every pair for which and are colinear, and is a groupoid homomorphism when restricted to any affine line . We use these two properties to define a monoid generated by the set of triangular loops. To this end, consider the quotient
(5.64) |
of the free monoid by the relations
-
(L.1)
if lie on an affine line, and
-
(L.2)
if and are linearly dependent.
Note that with these relations, is a group. Indeed, the inverse of the generator is . The following theorem is proved in Section B.2.
Theorem 5.23.
There is a group isomorphism .
As a result, we obtain the following universal property for the group , which is analogous to Lemma 2.13 and Theorem 5.8.
Proposition 5.24.
Let be a vector space and let be a group. Let be a map which
-
(1)
restricts to a groupoid homomorphism on for every affine line , and
-
(2)
restricts to the trivial homomorphism on for every subspace with .
Then there exists a unique group homomorphism such that .
We can use this universal property to define a section of . Indeed, consider the map defined by . This map verifies the equation and satisfies the two conditions in Proposition 5.24. Therefore, it induces a homomorphism which satisfies . As a result, the following short exact sequence splits
(5.65) |
Corollary 5.25.
The map
(5.66) |
is a group isomorphism.
Proof.
Because the above splitting, is isomorphic to the semidirect product . However, the conjugation action of on is trivial since is in the center of . Thus is isomorphic to the direct product. ∎
The crossed module action in induces an action of on under the isomorphism . In particular, for and , we have
(5.67) |
and the induced action on , where , is
(5.68) |
Define the suspension, , by
(5.69) |
Using the Peiffer identity, we can verify that, if , then
(5.70) |
In particular, only depends on the path through its endpoint . If , then by the Peiffer identity we have
(5.71) |
Hence, the action of on the -component is trivial. As a result, we obtain the following expression for the action.
Lemma 5.26.
The action of on is given by the following formula
(5.72) |
where , , and we view as the linear path from the origin to the endpoint of .
5.4.2. Decomposition of the Kapranov Crossed Module.
Next, we study a canonical splitting of Kapranov’s Lie algebra in order to obtain a decomposition of the group . We use the definitions in Section 4.2 for polynomial differential forms and currents. Let
(5.73) |
be the Euler vector field on the vector space . Let be the derivation of the de Rham complex given by interior product with . This operator is -equivariant and satisfies . The Lie derivative acts on the weight subcomplex by multiplication by . Recall that we can extend the de Rham complex by putting a copy of the base field in degree . We can then extend by sending an element to the value . As part of the proof of the Poincare lemma, we obtain the following decomposition.
Lemma 5.27.
For each , we have a direct sum decomposition of -represenations,
(5.74) |
We dualize this to get a decomposition for currents. Recall that is the codifferential from (4.39). Similarly, the dual of is the operator defined by
(5.75) |
for and . It can be expressed explicitly as
(5.76) |
Because is the dual of , we have . Furthermore, acts on by multiplication by the (negative) weight . By the same argument as above, we obtain the following.
Lemma 5.28.
For each , we have a direct sum decomposition of -representations,
(5.77) |
As a particular case of this Lemma, we have the decomposition . Recall Theorem 4.6, which states that the abelianization map sends isomorphically to . Therefore, the ideal is a complement to and hence is isomorphic to via the map (cf. [47]). Working with completions, we define
(5.78) |
Then is an isomorphism, and we denote the inverse morphism as
(5.79) |
In particular, the map satisfies , and should be viewed as the Lie algebraic analogue of the construction. Thus, we obtain the following decomposition.
Corollary 5.29.
There exists a canonical isomorphism of Lie algebras
(5.80) |
Under this canonical decomposition of , the crossed module action of induces an action of on , which we can write out explicitly. For and , we have
(5.81) |
From the construction of outlined in Appendix C, we observe that the abelianization map satisfies , where acts via the product in the symmetric algebra of . Therefore, the induced action on , where , is
(5.82) |
We define the Lie algebraic analogue of the suspension by
(5.83) |
which is the projection of onto . By applying the Peiffer identity, we can verify that for , we have
(5.84) |
Therefore, only depends on . As a result, we obtain the following expression for the crossed module action.
Lemma 5.30.
The action of on is given by the following formula
(5.85) |
where and .
Remark 5.31.
The suspension map in fact only depends on through its abelianization. To see this, consider the operator acting on currents . It satisfies
(5.86) |
Since acts as multiplication by the negative weight , we can identify as the projection onto , and as the projection onto . Thus, we can express as
(5.87) |
so that depends only on the projection and the abelianization .
Finally, the decomposition of Corollary 5.29 induces a decomposition of the group
(5.88) |
where . Note that because is abelian, we may identify it with with its additive group structure. Finally, because the constructions in this section are functorial, given a linear map , the induced group homomorphism preserves decomposition in Equation 5.88.
5.4.3. Decomposition of PL Surface Signature.
Next, we study how the surface signature factors with respect to the decompositions of both and . We consider
(5.89) |
where is the isomorphism from Corollary 5.25 given by .
Lemma 5.32.
The homomorphism preserves the decompositions of both and .
Proof.
The restriction of to must be valued in since is a morphism of crossed modules. To understand the restriction of to , consider the function
(5.90) |
Let , let be a 2-dimensional subspace such that , and let denote the inclusion. Because the PL surface signature is a natural transformation, we have
(5.91) |
Furthermore, the factor is trivial since is 2-dimensional, and this implies that . Then, because preserves the decomposition in (5.88), we conclude that . The map satisfies the hypotheses of Proposition 5.24 and so by the universal property, there is a unique map
(5.92) |
such that . Recalling the definition of from (5.90), we see that and hence
(5.93) |
implying that is valued in . As a result, we conclude that respects the decompositions of both and . ∎
The signature satisfies the following equation
(5.94) |
Hence, using the identification between and its image under in , we conclude that . This implies the following decomposition of the surface signature.
Proposition 5.33.
The piecewise linear surface signature decomposes as
(5.95) |
where
(5.96) |
5.4.4. Decomposition of Smooth Surface Signature
We can also consider a decomposition for the smooth thin crossed module
(5.97) |
We denote the kernel by , and the group of thin homotopy classes of loops by . Recall our convention that these groups consist of thin homotopy classes of paths and surfaces based at the origin, so . Now, we define a section of by
(5.98) |
This is well-defined under thin homotopy. Indeed, suppose , and let be a thin homotopy between and where and . Define a map by
(5.99) |
Because , we must have . Hence, . Then, by the same arguments as the PL setting in Corollary 5.25, we have the following.
Lemma 5.34.
The map
(5.100) |
is a group isomorphism.
The map extends the piecewise linear cone in the sense that . This can be verified by checking equality after precomposing with and applying Proposition 5.24. Using this fact, we can show that the surface signature also preserves the decompositions of and . The main property to show is that cones are mapped to under the surface signature.
Lemma 5.35.
The map is valued in .
Proof.
Recall that . By projection, this decomposition also holds for truncations . Then, by [48, Section 2.7, Problem 4], the exponential group is a closed subgroup in .
Next, let be a smooth loop such that . Then, by the same reasoning as the proof of Theorem 5.21, there exists a sequence of piecewise linear loops such that as . This implies that in Lipschitz norm as as well. Then, since is the cone of a piecewise linear loop, Proposition 5.33 shows that . Finally, since the (truncated) surface signature is continuous with respect to the Lipschitz topology (Proposition 3.18), we have . ∎
Thus, the analogue of Proposition 5.33 holds also in the setting of smooth surfaces. Combining this with the description of from Section 5.4.2, and Theorem 4.12 on the signature of closed surfaces, we obtain the following explicit formula for the surface signature, which is the main result of this section.
Theorem 5.36.
There is a canonical embedding
(5.101) |
from the group of formal surfaces to the product of , the vector space of formal -currents, and , the algebra of formal non-commutative power series. With respect to this embedding, the smooth surface signature decomposes as follows
(5.102) |
Given a smooth surface , the two components of the signature are given by
-
•
the path signature of the boundary path of
(5.103) -
•
the sum of surface integrals
(5.104) where, given linear coordinates on , the sum runs over the set of all monomial -forms and where are the dual polynomial -currents. Furthermore, the integrals are taken over
(5.105) the closed surface obtained by coning off the boundary of .
Remark 5.37.
The expression for the abelian component of the signature of a surface from Theorem 5.36 has the following simple coordinate-free description. Observe that the linear structure on allows us to write
which in turn allows us to view the surface integral as a map
The identity map on and its powers can be viewed as functions . Therefore
Then, using the multinomial theorem, the abelian component of the surface signature can be shown to have the following expression
6. Thin Homotopy of Piecewise Linear Surfaces
In this section, our aim is to show the following injectivity result.
Theorem 6.1.
The map is injective.
Because of the factorization of in Proposition 5.20, this result has two main implications. First, it shows that our algebraic construction of is rich enough to faithfully encode piecewise linear surfaces modulo thin homotopy.
Corollary 6.2.
The realization map is injective.
In particular, this implies that there are two types of cancellations that can occur via thin homotopy of PL surfaces.
-
(1)
Local Cancellations (Folds). When two linear components with opposite orientation are adjacent, they are cancelled via the equivalence relation in (PL1.1).
-
(2)
Non-local Cancellations. When two linear components with opposite orientation are not adjacent, they may sometimes be cancelled via the Peiffer identity in (5.24).
Second, it shows that the surface signature can detect whether two piecewise linear surfaces are thinly homotopic to each other. Hence, this generalizes the signature condition (S1).
Corollary 6.3.
The restriction of the surface signature to is injective.
Remark 6.4.
Consider the 2-truncated polynomial de Rham complex of
(6.1) |
In order to prove a de Rham theorem for this complex, the appropriate replacement for the singular chains in degree appears to be the thin homotopy group of closed piecewise linear surfaces, (or its smooth analogue). Indeed, there is a well-defined pairing given by integration
(6.2) |
Recalling the isomorphism
(6.3) |
where denotes the closed formal 2-currents on , the pairing induces a map , which we identify as the surface signature by Theorem 5.36. Therefore, Theorem 6.1 implies that this map is injective. Showing that this map is surjective, at least up to a finite degree in the weight grading, amounts to proving a generalization of Chow’s theorem [21] (see also [27, Theorem 7.28]). While we do not pursue this direction further, we leave this as an interesting avenue for future work.
To begin the proof of Theorem 6.1, we reduce the problem to the case of closed surfaces by a general result on crossed modules.
Lemma 6.5.
Let and be crossed modules of groups, and let be a morphism of crossed modules. Suppose and are injective. Then is injective.
Proof.
Let such that . Because is a morphism of crossed modules, we have , so . But since is injective, this implies that . By injectivity of on the kernel, we have . ∎
By Lemma 2.16 and the injectivity of the path signature, is injective. As a result, Lemma 6.5 reduces the proof of Theorem 6.1 to showing that
(6.4) |
is injective, where . Given such that , our strategy will be to consider various representatives of in order to show that in .
6.1. Piecewise Linear Simplicial Complexes
In this section, we consider simplicial complexes which can be realized in a vector space , and explain how to construct maps from their fundamental crossed modules to .
Definition 6.6.
A piecewise linear simplicial complex (PLSC) in is a finite ordered simplicial complex whose vertices lie in . In particular, the set of vertices is equipped with a total ordering. Each -simplex of is an ordered set555Our convention is to use square brackets to denote simplices. This is to distinguish between the point representations used for simplices as opposed to the edge representations for piecewise linear paths. of points , and hence, is equipped with a linear characteristic map . Let denote the image of . It is the convex hull of the points of in . We say that an -simplex is degenerate if there exists some -dimensional affine hyperplane such that , and we say that is non-degenerate otherwise. We say that is non-degenerate if all its simplices are non-degenerate.
We will abuse notation and use to refer both to the abstract simplicial complex as well as its geometric realization as a topological space, and we let denote the -skeleton of . Furthermore, we use to denote the piecewise linear realization of in , defined as the union of the subsets
(6.5) |
The complex is equipped with a continuous realization map
Remark 6.7.
Note that because a 1-simplex is a pair of distinct points , , it is automatically non-degenerate.
Furthermore, we will require a notion of compatible complexes.
Definition 6.8.
Suppose is a PLSC in . We say that is compatible if for every pair of simplices , their intersection in satisfies
(6.6) |
where is a common subsimplex of both and , or in which case and are disjoint.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x12.png)
Lemma 6.9.
Let be a compatible non-degenerate PLSC. Then the realization map is a homeomorphism.
Proof.
Because is compact and is Hausdorff, it suffices to prove injectivity. For each simplex , the map is injective since is nondegenerate. Hence, let , be points in different simplices such that . Hence, and intersect non-trivially. Therefore, there is a subsimplex of both and such that . Thus, there is a point such that . Because is a subsimplex we have . By injectivity of , we conclude that . Similarly, . Therefore, and is injective. ∎
Let be a 2-dimensional PLSC. We denote the set of vertices by , the set of -simplices by , and the set of -simplices by . The fundamental groupoid of the -skeleton, , is free on the set of edges . We define a homomorphism
(6.7) |
by sending an edge to the element . Composing with the realization map , the edge gets sent to , the straight line path in from to . Let be a basepoint. Restricting to the fundamental group based at gives rise to a group homomorphism
(6.8) |
Lemma 6.10.
Suppose is a PLSC whose -skeleton is compatible and non-degenerate. Then the map is injective.
Proof.
Let be such that . Then the realization is a loop in which is thinly null homotopic in . By the image condition (I1), there exists a null homotopy whose image is contained in the image of , and therefore in . Because the map is a homeomorphism by Lemma 6.9, the null homotopy can be lifted to . As a result, . ∎
Free crossed modules were originally developed by Whitehead [55] to provide an algebraic description of second relative homotopy groups. We consider a special case of Whitehead’s result for 2-dimensional CW-complexes, also discussed in [10, 9].
Theorem 6.11.
[55, 10] Let be a connected based 2-dimensional CW complex with 1-cells and 2-cells . Fix a spanning tree of the 1-skeleton. For each 2-cell , choose a basepoint , and let be the attaching map of the cell. Furthermore, let be the unique path in connecting to . Define by . Then, the fundamental crossed module of the pair
(6.9) |
is the free crossed module on . Let denote the map which sends each 2-cell to the corresponding generator of .
Let be a -dimensional PLSC. We will use Whitehead’s theorem to construct a map . First, choose a basepoint and fix a spanning tree of the -skeleton. Let be the homomorphism from (6.8). By Theorem 6.11, is free on the set of 2-cells of , and so by the universal property of Theorem 5.1, the map is specified by a function such that .
Each 2-simplex has the form . Let be its boundary loop. Under it gets sent to the following planar loop
(6.10) |
Let be the unique path in connecting to , and let . Then is a kite and we define
(6.11) |
It is then clear from the construction that . We therefore obtain the desired morphism .
Corollary 6.12.
Let be a 2-dimensional connected PLSC in . The maps and above induce a morphism of crossed modules such that , where is the map sending each 2-cell to its generator.
6.2. Proof of Injectivity
In this section, we will prove Theorem 6.1. Throughout this section, we fix such that . The proof consists of three main steps.
-
(1)
Construct an appropriate representative which induces a compatible PLSC (Definition 6.8). Then construct an element such that .
-
(2)
The trivial signature condition will provide a matching of kites of opposite orientation, implying that under the Hurewicz map , we have .
-
(3)
Construct a simply connected PLSC such that , and apply the Hurewicz theorem to show that .
6.2.1. Construct Simplicial Model
Our first step is to construct a simplicial model for . This is a PLSC whose fundamental crossed module contains an element which gets sent to under the map constructed in Corollary 6.12. Such a simplicial model will be induced by a triangulated representative of . In order to define this representative, we begin by defining the set of marked kites
(6.12) |
where is the free monoid generated by . In other words, these are kites for which we have additionally chosen the data of a representative of the tail path. There are naturally defined surjective monoid homomorphisms
(6.13) |
Definition 6.13.
A marked representative (or simply representative) of an element is a lift of this element to . A marked kite is triangular if . Therefore, a representative of is said to be triangulated if each marked kite is triangular.
We now explain how to construct a PLSC given the data of a triangulated representative. First, let be a marked triangular kite. Therefore and . Define , and given , define Finally, define
(6.14) |
The piecewise linear simplicial complex associated to , denoted , is given by
(6.15) | ||||
(6.16) | ||||
(6.17) |
More precisely, we take the set of vertices to be , with the induced order. Since this is a subset of , any repetitions are automatically deleted. The simplicial complex is the union of and in the power set of . As a result, any repetitions are removed, regardless of orientation or ordering in the above description. The 2-simplex associated to is the unique 2-simplex of . Note that is non-degenerate and that is connected.
Next, let be a triangulated representative. The PLSC associated to , denoted , is given by the union
(6.18) |
More precisely, the set of vertices is the union of the vertices from each , with the ordering such that the new vertices of come after those of . The simplicial complex is then the union of each in the power set of . Note that is connected because each complex is connected and contains the origin as a vertex.
The following definition will play an important role when we ‘match’ the kites in a representative of an element .
Definition 6.14.
Let . A compatible representative is a triangulated representative such that is compatible in the sense of Definition 6.8.
Remark 6.15.
Recall from Remark 6.7 that a -simplex in a PLSC is always non-degenerate. Furthermore, as noted above, the 2-simplex associated to a triangular kite is also non-degenerate. Therefore, the simplicial complex is automatically non-degenerate.
We prove the following in Appendix D by carefully considering subdivisions.
Theorem 6.16.
There exists a compatible representative of every .
Now let , fix a compatible representative , and let be the associated PLSC. Equip with the basepoint . All homotopy groups will be taken with respect to this basepoint. By Corollary 6.12, we may construct a morphism of crossed modules . Since is compatible and non-degenerate, the map is injective by Lemma 6.10. Next, we lift to an element of .
Proposition 6.17.
There exists an element such that . Furthermore, if , then .
Proof.
The representative allows us to factor into a product of kites in . Hence, it suffices to show that each is in the image of . Each marked triangular kite has an associated 2-simplex and hence determines a generator . Under the map this generator is sent to
(6.19) |
where is an appropriate sign and where , for a path in connecting to the basepoint of . The element can also be realized as , for a path in connecting to . Therefore, and
(6.20) |
This shows that a preimage exists. By the injectivity of the path signature and , we have . Thus, , and so by the injectivity of . Hence . ∎
6.2.2. Matching kites
Consider the Hurewicz map
(6.21) |
Applying it to the element constructed in Proposition 6.17, we obtain an element . The next step is to show that this element is .
Let denote the set of 2-simplices of . The cellular chain complex of is given by
Given the representative , each marked triangular kite has an associated 2-simplex which is an element of . This defines a function sending the index to . Since the ordering on the triangular loop in might not agree with the ordering of simplices in , we must also record this information as orientation data. This is a function such that if and only if the ordering on the triangular loop in matches the fixed order on . We call the pair of functions the simplex mapping of and we note that
(6.22) |
The pair is a simplex matching if , since this means that the simplices in are “matched up” in pairs of opposite orientation. By hypothesis, the signature of the realization of is trivial, , and by Corollary 4.13, this implies that
(6.23) |
for all . We will use this fact to show that is indeed a simplex matching.
Proposition 6.18.
Let be the element from Proposition 6.17. Then .
6.2.3. Show that .
Our objective now is to apply the Hurewicz theorem to show that . While may not be simply-connected in general, we can add 2-cells to kill off the fundamental group.
Lemma 6.19.
Let be a 2-dimensional, connected PLSC in . There exists a 2-dimensional PLSC such that and .
Proof.
Let be the set of vertices of and let be the set of 1-simplices. Choose a point such that for each edge the triple is not contained in a line. For each vertex , define a new 1-simplex , and for each edge , define a new 2-simplex . Now define a new PLSC with vertex set , set of 1-simplices , and set of 2-simplices . Extend the order on to so that comes after all other vertices. By construction, is contractible.
Now define to be the union of and . This is a PLSC that contains and as subcomplexes. Furthermore, and intersect along , the -skeleton of . Therefore, by the van Kampen theorem, . ∎
Next, we prove a general relationship between the kernels of and the Hurewicz map which will imply our main injectivity result.
Proposition 6.20.
Let be a 2-dimensional connected, compatible, and non-degenerate PLSC in . Let be the homomorphism defined in Corollary 6.12, and let be the Hurewicz map from (6.21). Then
(6.26) |
Proof.
First, we show . Let be the PLSC from Lemma 6.19 with inclusion map . The map induces a map . Applying Corollary 6.12 to , we have another morphism of crossed modules such that the following diagram commutes,
(6.27) |
Indeed, examining the construction of in (6.8), it is clear that . The construction of leading up to Corollary 6.12 depends on the choice of a spanning tree . Therefore, to ensure that , we choose the spanning tree to be an extension of . This implies that the chosen generators of are sent to the chosen generators of under the map .
Next, by naturality of the Hurewicz map, the following diagram commutes.
(6.28) |
Let . Because is an isomorphism by the Hurewicz theorem, and , we have . Therefore, this implies that .
It remains to show that . Suppose . Then, we have
(6.29) |
Since is compatible and non-degenerate, is injective by Lemma 6.10. Hence it follows that . Furthermore, the signature is trivial. Then, the same argument as Proposition 6.18 shows that .
∎
Remark 6.21.
In light of Proposition 2.18, a natural question is whether one can embed free crossed modules into the piecewise linear crossed module . In contrast to Proposition 2.18, Proposition 6.20 shows that this is not possible in general for crossed modules, as the kernel may be nontrivial.
Now, we can prove our main injectivity result.
Proof of Theorem 6.1.
By Lemma 6.5, it suffices to show that is injective. Let such that . Let be a compatible representative of , , and be element from Proposition 6.17 such that . By Proposition 6.18, we have , and finally by Proposition 6.20, , so . ∎
6.3. Equivalent Conditions
In this section, we complete the generalization of Theorem 2.11 for piecewise linear surfaces, and consider appropriate extensions of the remaining definitions of thin homotopy. We begin by defining the class of geometric surfaces that we will consider.
Definition 6.22.
A smooth piecewise linear surface is a surface such that
-
•
only the top boundary is non-trivial, , and
-
•
there exists a compatible PLSC in such that , and the restriction of to any -simplex is the composition of a smooth reparametrization with sitting instants, and an affine linear function ,
(6.30)
We denote the space of smooth piecewise linear surfaces by .
The following lemma shows that consists of surfaces which lie in the thin homotopy classes defined by the realization of .
Lemma 6.23.
Let . There exists some such that is in the thin homotopy class of .
Proof.
Let be the boundary of the unit square. Let be a compatible PLSC in which satisfies the conditions of Definition 6.22, and note that it is a compatible triangulation of in the sense of Definition D.2. Then, viewing as a kite, the proof of Lemma D.5 implies there exists a compatible representative of . Then, applying the maps from Definition 6.22 to , we obtain an element . Finally, by Lemma 5.16, the identity map is in the thin homotopy class of , so is in the thin homotopy class of . ∎
We now propose generalizations of the various definitions of thin homotopy for paths to the setting of smooth piecewise linear surfaces. The rank condition (R1) was generalized in Definition 3.2, the analytic condition (A1) was generalized in Remark 4.14, and the signature condition (S1) is generalized using the surface signature from Definition 3.17. Here, we comment on the remaining definitions before stating our generalization of Theorem 2.11.
First, we generalize the word condition (W1). Given a surface , there exists a representative by Lemma 6.23. We say that the surface is word reduced if the representative is trivial in . This is analogous to the condition that the transfinite word associated to a path is reducible to the trivial word. Next, to generalize the holonomy condition (H1G), we must consider a class of crossed modules that are analogous to semisimple Lie groups. We use the following condition.
Definition 6.24.
A crossed module is holonomy nondegenerate if
-
(1)
there exists a semisimple Lie algebra such that , and
-
(2)
is nontrivial.
A crossed module of Lie groups is holonomy nondegenerate if its associated crossed module of Lie algebras is holonomy nondegenerate.
Remark 6.25.
In the above definition, the first condition is used to ensure that the surface holonomy is rich enough to distinguish boundary paths, while the second condition allows us to distinguish between closed surfaces.
To formulate the image condition (I1), the naive generalization of the definition for paths suggests that two surfaces and are equivalent if there is a homotopy that is constrained to lie in the union of the images of and . However, as we saw in Corollary 3.5, this definition does not work because of nonlocal cancellations. Instead, we propose to use the transitive closure of this relation. Finally, our generalization of the factorization condition (F1) will make use of factorizations through 2-dimensional CW complexes. However, simply requiring that a surface factors through such a complex is not enough. We must also require that the factorization is trivial in homology.
Theorem 6.26.
Let be a piecewise linear surface. It is thinly null-homotopic if any of the following equivalent definitions hold:
-
(W2)
Word Condition. There is a marked representative of which is trivial in .
-
(H2G)
Holonomy Condition. For a holonomy nondegenerate , the surface holonomy along of every smooth fake-flat -connection is trivial.
-
(R2)
Rank Condition. There exists a smooth thin homotopy , as defined in Definition 3.2 from to the constant path at .
-
(I2)
Image Condition. There exists a smooth homotopy between and the constant surface at the origin such that it satisfies (I1) on the boundary of and such that
(6.31) for a collection of times .
-
(F2)
Factorization Condition. The surface has trivial boundary and, after modifying the boundary using a thin homotopy, there exists a 2-dimensional CW complex and a map , which is smooth when restricted to each cell, such that factors as
(6.32) and such that , where is a map which covers the sphere and sends the boundary to a basepoint .
-
(A2)
Analytic Condition. The surface has trivial boundary, and for all compactly supported , we have .
-
(S2)
Surface Signature Condition. The surface signature of is trivial, .
Proof.
We have proved the equivalence of (W2), (R2), and (S2) in Theorem 6.1. Moreover, Corollary 4.13 shows the equivalence between (A2) and (S2) for the case of closed surfaces. However, since (S2) implies that is closed, the equivalence holds in general.
Now we prove the equivalence of (H2G) for a fixed holonomy nondegenerate with associated . Because surface holonomy is invariant with respect to thin homotopies [2, 50, 44] as defined by (R2), it follows that (R2) (H2G). Now, suppose (H2G) holds. Let be the semi-simple subalgebra. Choose a linear section . Given any 1-connection , we can consider the 2-connection which is fake-flat by definition. Because surface holonomy is a morphism of crossed modules by Theorem 3.14, we have
(6.33) |
Hence, by the assumption of (H2G), this implies that is trivial for all . Thus, (H1G) holds for , where is a Lie group with Lie algebra . This implies that the boundary of is trivial, so is closed. Next, let , which is nontrivial by assumption. Let be a -dimensional subalgebra and consider an abelian 2-connection , where is valued in the subalgebra. Let be the integration of . The subalgebra integrates to a subgroup which is either or . Because the connection is abelian, the surface holonomy (Definition 3.13) is given by either
(6.34) |
Because is an arbitrary 2-form, either of these will imply (A2).
Next we consider the factorization condition (F2). Suppose (S2) holds. Our proof of Theorem 6.1 that (S2) (W2) provided the desired factorization of through a PLSC , where is linear when restricted to each simplex. Furthermore in Proposition 6.17, we constructed a such that , and thus by Proposition 6.18. Now, suppose (F2) holds. This implies that is closed. Given a -form , integrating the pullback along the cells of defines a map . Hence, the fact that implies that (A2) holds.
Next, we consider the image condition (I2), where it is immediate that (I2) (R2). Now, suppose (S2) holds, and once again we consider the proof of Theorem 6.1. There, we show that factors through a PLSC , and add 2-simplices to build a simply-connected CW complex in Lemma 6.19. The result of Theorem 6.1 implies that there is a smooth thin homotopy satisfying (R2) which is contained in the image of . Suppose . By choosing some which connects the basepoint of to the origin, we can interpret each as a kite . Using Lemma 5.17 and translating the basepoint, we can realize each -simplex as a surface . We can incorporate the tail paths by using the formula for the action given in (3.21) to obtain a surface defined by
(6.35) |
Let denote the top boundary of , and let be a thin homotopy of paths from (on the bottom face of ) to (on the top face of ). Then, we define
(6.36) |
where concatenations are performed from left to right. Here, each is a surface which represents the fold such that the boundary of is trivial. Note that the image of is by definition, and since and differ only by folds, there exists a homotopy between them which is contained in the image of , and whose boundary is contained in the image of . Then, the image of the sequence of homotopies
(6.37) |
is contained in the image of , and thus (I2) is satisfied. ∎
7. Thin Null Homotopy and Group Homology
In this section, we build a connection between thinly null homotopic surfaces and the group homology . On the one hand, this allows us to geometrically interpret in terms of surfaces; on the other, it allows us to further classify thinly null homotopic behavior. Given a compatible PLSC in , we use Corollary 6.12 to obtain a morphism
(7.1) |
Recall from Proposition 6.20 that . Now, we aim to characterize in terms of group homology. We provide a brief exposition to the required background on group homology, and refer the reader to [7] for further details.
Definition 7.1.
Let be a group and be a -module. Let be the augmentation map defined by for all . The group of co-invariants of is defined by , where is the augmentation ideal generated by for . In other words, is the quotient of by elements of the form for and .
Definition 7.2.
Let be a group and let be a projective resolution of over , where
(7.2) |
is exact. Then, we define the group homology of by
(7.3) |
The following is an exact sequence, originally due to Hopf [33], which relates the homology and homotopy groups of a CW-complex with the homology of its fundamental group through the Hurewicz map. We state a version from [7].
Theorem 7.3.
[7, Exercise 1, Section 2.5] Let be an -dimensional CW-complex such that for for some . Let . Then, the sequence
(7.4) |
is exact, where is the co-invariants functor applied to the Hurewicz map and is equipped with the trivial -action.
In our current setting of a 2-dimensional CW complex , the connectivity hypothesis is trivially satisfied, and thus we obtain the exact sequence
(7.5) |
The kernel of is exactly . Thus, we obtain the following characterization of .
Proposition 7.4.
The kernel of is
(7.6) |
where is the augmentation ideal.
Proof.
This result allows us to classify thinly null homotopic surfaces whose image lies in an embedding of . Suppose factors as
(7.8) |
If is thinly null-homotopic, then . Then, depending on how is killed based on (7.6), the thin homotopy exhibits different behaviors.
-
(1)
(Folds) : The map is null homotopic within its image in . This is the case if exhibits only folds.
-
(2)
(Nonlocal Path Conjugation) : The map is null-homotopic within its image in up to conjugation by paths . This is an example of a nonlocal cancellation because the thin null-homotopy must move through the simply connected extension .
-
(3)
(Nonlocal Surface Cancellation) : The map is null homotopic within the simply connected extension , and is classified by .
Example 7.5.
Consider the example of the group with the presentation , where the associated CW complex666While this CW complex is not a PLSC, we can refine it to a simplicial complex, and construct a piecewise linear map to for sufficiently large . is . In this case, we have
(7.9) |
By (7.5), this implies that , and must represent a thinly null-homotopic surface which factors through . Consider the surface from Proposition 3.4. Note that the map from the factorization (3.9) represents a generator of , and is also nontrivial when we pass to the co-invariants . Thus, the nontrivial element in represents the thinly null homotopic surface .
Appendix A Notation and Conventions
Symbol | Description | Page |
Categories | ||
category of finite-dimensional vector spaces | 2 | |
category of Lie algebras | ||
category of crossed modules of groups (resp. Lie groups) | 3.7 | |
category of crossed modules of Lie algebras | 3.11 | |
comma category associated to and | 3.5 | |
comma category associated to and | 5.1 | |
Crossed Modules | ||
piecewise linear crossed module | 5.6 | |
thin crossed module (thin homotopy and translation equivalence classes) | 3.9 | |
fundamental crossed module of CW-complex relative to the 1-skeleton | 3.10 | |
Kapranov’s free crossed module of Lie algebras and its completion | 3.51, 3.6 | |
formal integration of as a crossed module of groups | 3.6 | |
Differential Forms and Currents | ||
smooth differential -forms on | ||
smooth compactly supported differential -forms on | ||
polynomial differential -forms and -currents on | 4.2, 4.2 | |
completion of polynomial differential -forms and -currents on | 4.2, 4.2 | |
Holonomy and Signatures | ||
connection/2-connection | 2.2, 3.12 | |
universal translation-invariant connection/2-connection | 2.10, 3.55 | |
curvature of the connection | 2.6 | |
curvature/2-curvature of the 2-connection | 3.12 | |
path/surface holonomy | 2.3, 3.13 | |
smooth path/surface signature | 2.4, 3.17 | |
piecewise linear path/surface signature | 2.17, 5.19 | |
realization of PL paths/surfaces as thin homotopy equivalence classes | 2.34, 5.18 | |
Misc | ||
planar piecewise linear loops in | 5.2 | |
set of kites in | 5.2 | |
set of marked kites | 6.2.1 | |
piecewise linear loops generated by triangular loops | 5.4.1 | |
pair groupoid of | 5.4.1 | |
piecewise linear and smooth cone map | 5.4.1, 5.4.4 | |
Hurewicz map | ||
morphism of crossed modules | 6.12 | |
PLSC associated to a representative of | 6.13 | |
simplex mapping associated to compatible | 6.2.2 | |
concatenation of paths / group operation in | 2.1,4 | |
, | horizontal/vertical concatenation of (thin homotopy classes) of surfaces | 3.8, 4 |
Appendix B Piecewise Linear Paths and Loops
B.1. Minimal Representatives of PL Paths
In this section, we prove Proposition 2.12, which shows that elements in have a unique minimal representative. This will be done by developing a rewriting theory on the free monoid generated by . To simplify the notation, we will omit the symbol for monoid multiplication in this section. Now we consider rewriting on the free monoid generated by . We define two rewriting steps corresponding to the relations for . For arbitrary words and linearly dependent , we define
(B.1) |
Given two words , we write for applying exactly one of the rewriting steps above and write if we can get from to by a sequence of rewriting steps (possibly zero steps). Note that removing a can always be realized by the (PL0.1) rewriting step, unless both and are empty words. Let be the length of the word . Note that if then . If this involves a non-zero number of rewrite steps, then , so that rewriting strictly decreases the length of words. The only words for which we cannot apply a rewriting step are those described in Proposition 2.12. We will call these minimal words. It is clear that because rewriting decreases the length of words, the process must eventually terminate. Our goal is to show that it always terminates at the same minimal word.
Lemma B.1.
Let and suppose that and are two rewriting steps. Then there is a word such that and .
Proof.
Since the rewriting step (PL0.2) only applies if the length of is , we can assume that we are applying (PL0.1). Hence, the two rewriting steps have the form
(B.2) |
and
(B.3) |
We can assume that . If , then the two rewrite steps are identical. If then we take . If , then each pair and is contained in a common line. There are two cases here. First, we could have . But then the two rewrites and are the same. Second, if , then are all contained in the line . Then we take . ∎
Lemma B.2.
Let and suppose that and . Then there is a word such that and .
Proof.
We prove this by induction on the length of the word . We can assume that both and involve a non-zero number of rewriting steps, since otherwise the claim is obvious. Furthermore, if the first rewrite step is the same, meaning that and , then the result follows by induction since , so that we can apply the induction hypothesis to . As a result, we may assume the rewrites to have the form
(B.4) |
and and . By Lemma B.1, there is a word such that and . Now , and we have and . So by the induction hypothesis, there is a word such that and . Similarly, , and , so by the induction hypothesis, there is a word such that and . Finally, and and . So by the induction hypothesis, there is a word such that and . Combining the rewriting steps we have
(B.5) |
∎
Corollary B.3.
If and are rewritings such that and are minimal words, then . In other words, any word reduces to a unique minimal word.
Proof.
By Lemma B.2, there is a word such that and . But and are minimal, implying that . ∎
Proof of Proposition 2.12.
Let . We claim that there is a unique minimal representative, namely a representative such that each and each consecutive pair are linearly independent. Note first that such a representative must exist: given an arbitrary representative of , the process of applying the rewriting steps must terminate, since each step shortens the word. But rewriting can only terminate at a minimal word. To prove uniqueness, suppose that and are two minimal representatives. Since and are both words representing , there is a sequence of rewriting steps, and their inverses, going from to . By Lemma B.2, there is a word such that and . But because and are minimal, it is not possible to apply any non-trivial rewriting step. Hence . ∎
Remark B.4.
Note that the minimal representative of an element is also the unique representative with minimal length. This is because a shortest-length word must be minimal.
B.2. Piecewise Linear Loops
The goal of this section is to prove Theorem 5.23. Let be the natural map sending a generator to the corresponding element in the group. This map satisfies the conditions of Proposition 5.24. In fact, the universal property of this proposition holds for because of the defining relations. We will need to make use of this property in what follows.
The first step in proving Theorem 5.23 is to show that has minimal representatives, analogous to Proposition 2.12.
Lemma B.5.
An element has a unique minimal representative
(B.6) |
This is a word such that the following conditions are satisfied
-
(1)
for each , the vectors and are linearly independent, and
-
(2)
for each , either or and are linearly independent.
Similar to (B.1), we define two rewriting steps in . Let , let be such that lie on a common affine line, and let be such that are linearly dependent. The two rewriting steps are
(B.7) |
Note that the minimal words defined in Lemma B.5 are precisely the words which cannot be shortened using these rewriting steps. We follow the notation from Section B.1 and use and to denote a single and arbitrary (possibly zero) rewriting steps, respectively. The following lemma is proved in the same manner as Lemma B.1 by checking all cases, so we omit the details.
Lemma B.6.
Let and suppose and are two rewriting steps. Then, there is a word such that and .
Proof of Lemma B.5..
Using Lemma B.6 instead of Lemma B.1, we prove the analogue of Lemma B.2. Using this result, we then repeat the argument from the proof of Proposition 2.12. ∎
Now, we can use this to prove Theorem 5.23.
Proof of Theorem 5.23..
As noted below the definition of from (5.63), this map verifies the assumptions of Proposition 5.24. Using the universal property of the group , there is a unique map such that . It sends the element to the triangular loop . This map is surjective since we can factor any element of as a product of triangular loops. It remains to show injectivity. Suppose is the minimal representative of an element of . Under , this gets sent to the equivalence class for
(B.8) |
Let denote the length of the minimal representative for both and . While there may be cancellations for in , we must have because is minimal in . Indeed, the only reductions in can occur between and . Suppose these vectors are linearly dependent. If , then the word reduces to and there are no further reductions at this locus. If, on the other hand, , then the word reduces to and there are no further reductions at this locus. Hence, the minimal representative for will contain all .
If , we must have . This implies that and thus . Hence, is injective.
∎
B.3. Kites and Planar Loops
In this section, we collect several useful results about kites and planar loops.
Lemma B.7.
Let be a non-trivial planar loop with span , let be a path such that is planar, then .
Proof.
Let and be minimal representatives. We will prove this by induction on the length . Consider
(B.9) |
The only reductions can occur between and , or between and . Hence, if , then the word is minimal. But then the span of contains and , contradicting the assumption that is planar. Hence, . Let , which is minimal, and let , which is a non-trivial planar loop with span . Then is planar. By induction , and therefore, . ∎
Lemma B.8.
Let be non-trivial planar loops with spans . If is planar, then .
Proof.
Let and be minimal representatives. Then
(B.10) |
If , then this word is minimal. But then the span contains and , contradicting the assumption that the loop is planar. Hence . For the same reason, . If and are linearly independent, then and we’re done. Hence, we assume that they are colinear, allowing us to reduce further:
(B.11) |
Now if , then the word is minimal (possibly after removing if this vector is ). Because both and are non-trivial loops, we have , and both and are linearly independent pairs of vectors. Hence, the span of contains and , again contradicting planarity of . Thus, , implying that . ∎
Lemma B.9.
Let be non-trivial planar loops with spans . If is a kite , where and , then .
Proof.
Suppose , where with . If is trivial, then and we are done. Hence, we assume to be non-trivial. Let
(B.12) |
where since they are both nontrivial loops. Consider a factorization of the minimal representative of the concatenation
(B.13) |
where and . To obtain the minimal representative, the only possible simplification is by applying (PL0.1) to . If , then we may apply (PL0.1) again to . If both steps are possible, this implies that , and thus , completing the proof. Therefore, we assume that at most two rewriting steps are possible (i.e. one instance of (PL0.1) and one instance of (PL0.2)), and hence and .
Next, let and be minimal representatives. We claim that it is possible to assume that . Indeed, let be the first element from the right which is not in , so that . Define , which is minimal, and let , which is a non-trivial planar loop with span . Then and hence
(B.14) |
Thus, since is also a kite, we may simply replace with and with , establishing the claim. The minimal representative for the kite is therefore
(B.15) |
So far, we have obtained two descriptions of the minimal word representative of the product . Hence, we have equality of minimal words . There are two cases to consider, depending on where the transition between and occurs relative to .
-
(1)
First, the transition occurs in . Then there are inclusions of words and . This implies that . In this case, we have , and by Lemma B.8, this implies that .
-
(2)
Otherwise, the transition occurs in , implying the inclusion of words , or it occurs in , implying the inclusion of words . In either case, we have for a common or . But then is planar, and so by Lemma B.8, .
∎
Corollary B.10.
Let be non-trivial planar loops with spans . Let be a path. If is a planar loop, then and .
Proof.
Let , which we assume to be a planar loop. If this loop is trivial, then . Then, since is planar, by Lemma B.7 and hence . Next, assuming that is non-trivial, let be its span. Then are non-trivial planar loops with the property that is a kite. Then by Lemma B.9 and is a planar loop. By Lemma B.7, . Hence . ∎
Appendix C Free Crossed Modules of Lie Algebras
In this section, we fix a field of characteristic , either .
C.1. Free Lie Algebras and Representations
In this section, we begin with a brief discussion of the free Lie algebra generated by a vector space. Let be the category of vector spaces over , and be the category of Lie algebras over . There exists a natural forgetful functor
(C.1) |
and here we describe the corresopnding left adjoint functor
(C.2) |
which sends a vector space to the free Lie algebra over , denoted . Let be the tensor algebra, the free associative algebra over , equipped with the shuffle coproduct , which is the algebra map defined on by . We define to be the set of primitive elements of ,
(C.3) |
Let be a basis of , and let be the free Lie algebra generated by the set . By [51, Theorem IV.4.2], there is an isomorphism , where is the universal enveloping algebra. Then, by [51, Theorem III.5.4], , allowing us to identify . Hence, our notion of free Lie algebra coincides with the usual definition in terms of sets.
C.2. Free Crossed Modules of Lie Algebras
Now, we move on to consider free crossed modules of Lie algebras. Consider the subcategory of crossed modules
(C.4) |
where is fixed and where morphisms are the identity on . Taking the underlying vector space of , we define the slice category consisting of linear map . There is a forgetful functor
(C.5) |
and in this section, we will construct a free functor
(C.6) |
as a left adjoint to . We will break up this construction into two steps by factoring the forgetful functor as
(C.7) |
where is the category of -representations, and is the slice category, consisting of -equivariant maps , where the target is the adjoint representation.
C.2.1. Step 1
Recall that -representations can be equivalently described as -modules, , for the universal enveloping algebra of . Hence, the forgetful functor can be viewed as restricting the -action along the map . Its left adjoint is therefore the extension of scalars functor
(C.8) |
We can slice this functor to obtain a left adjoint to , which we also denote by . It sends an object to the -equivariant morphism
(C.9) |
where the second map is induced by the adjoint action.
C.2.2. Step 2
Next, we will construct the left adjoint to the forgetful functor . Let be an object, where is a -representation, and is an equivariant map. In order to define a crossed module, we must construct a Lie algebra structure based on , such that the Peiffer identity holds. In fact, we will use the Peiffer identity to define the Lie bracket. We define the Peiffer pairing by
(C.10) |
This pairing is symmetric and bilinear, and satisfies . In other words, it defines a -equivariant map . We denote the image of this map by , which is a subrepresentation of lying in the kernel of . We define
(C.11) |
Next, we define a preliminary bracket
(C.12) |
which is bilinear and skew-symmetric. Furthermore, we have
(C.13) |
implying that the bracket descends to
(C.14) |
This leads to the desired crossed module of Lie algebras.
Lemma C.1.
The bracket defined in (C.14) defines a Lie algebra structure on , descends to an equivariant morphism of Lie algebras , and the -action on is a derivation of . In particular, is a crossed module of Lie algebras.
Proof.
First, since is a subrepresentation of contained in the kernel of , it is clear that we have an induced equivariant map . Then using the equivariance, we get
(C.15) |
showing that is bracket preserving. To show that acts as a derivation of the bracket,
(C.16) | ||||
(C.17) | ||||
(C.18) | ||||
(C.19) |
To show that defines a Lie bracket, we must show that it satisfies the Jacobi identity. This is equivalent to showing that is a derivation of the bracket. But , which we have just shown is a derivation. Finally, to show that is a crossed module, we must check the Peiffer identity. But this holds by construction. ∎
The construction of described above is functorial, providing the functor
(C.20) |
Lemma C.2.
The functor is left adjoint to .
Proof.
To exhibit the adjunction we describe the counit and unit natural transformations. On the one hand, we have that , because when arises from a crossed module. Therefore, we take to be the identity. On the other hand, the unit natural transformation must have components
(C.21) |
which we take to be the natural projection maps. The counit-unit equations follow easily from the fact that and that , when arises from a crossed module. ∎
C.2.3. Universal Property
Now, we can put these two functors together to obtain a left adjoint to the forgetful functor in (C.5).
Theorem C.3.
The forgetful functor has a left adjoint
(C.22) |
which sends a linear map to the crossed module
(C.23) |
This free crossed module satisfies the following universal property.
Corollary C.4.
The unit of the adjunction provides a distinguished map
(C.24) |
given by including the vector into the equivalence class of . For any crossed module with a map in , there exists a unique map in such that
(C.25) |
Remark C.5.
This construction of the free crossed module is equivalent to other definitions, for instance in [22, Section 3.6], which first construct a free pre-crossed module (by taking the free Lie algebra of ), and then quotienting out by the Peiffer identity (as we have done with ). Our construction uses the fact that the Peiffer identity fully determines the Lie bracket, and bypasses the need to go through the free pre-crossed module.
C.3. Global Universal Property
In this this section, we will show that the free crossed module satisfies a stronger universal property within the entire category of crossed modules.
C.3.1. Pullback Property
We will start by describing a pullback construction that leads to a factorization of maps. Suppose we have the following map of crossed modules
(C.26) |
Our aim is to define the pullback crossed module such that this map factors as
(C.27) |
First, consider the composition
(C.28) |
which allows us to define the semi-direct product , equipped with a Lie algebra map
(C.29) |
Then, we define the pullback by
(C.30) |
By direct computation, one can check that the Lie bracket on is given by
(C.31) |
which implies that there are natural Lie algebra maps
(C.32) |
which satisfy . Furthermore, one can check that the action
(C.33) |
satisfies the properties of a crossed module.
Lemma C.6.
The pullback defined by is a crossed module of Lie algebras. Furthermore, is a morphism of crossed modules.
Then, the desired factorization is immediate.
Lemma C.7.
Let
(C.34) |
be a morphism of crossed modules. There exists a unique map in defined by
(C.35) |
such that , where is defined in Lemma C.6.
C.3.2. Global Universal Property
In this section, we will consider a more general universal property for free crossed modules. Let denote the comma category associated to the functors and . An object of is given by the data of a vector space , a Lie algebra , and a linear map . A morphism consists of a linear map and a Lie algebra morphism such that as linear maps. In this section, we will exclusively refer to the natural forgetful functor
(C.36) |
Furthermore, given , the unit will refer to a morphism in . We are now ready to state and prove the enhanced universal property.
Theorem C.8.
Let , , and let
(C.37) |
Then there is a unique morphism of crossed modules
(C.38) |
such that . In particular and .
Proof.
Given the Lie algebra map , we consider the pullback crossed module , and the corresponding morphism of crossed modules from Lemma C.6,
(C.39) |
Mimicking the construction in Lemma C.7, let be defined as This defines a map
(C.40) |
in with the property that . Now using the universal property of in from Corollary C.4, there is a unique map
(C.41) |
in such that , where is the unit from (C.24). Now, we define
(C.42) |
which is a morphism of crossed modules and it satisfies
(C.43) |
Hence, is the desired morphism.
It therefore only remains to show that is unique. To this end, let be another such morphism. By the identity , we must have . Then, by Lemma C.7, factors as , where
(C.44) |
is in the category . Now consider the map in . It satisfies
(C.45) |
This this implies that , or that . But now appealing to the universal property of , we see that we must have , the map defined above. Hence,
(C.46) |
This establishes uniqueness. ∎
We can rephrase the above theorem as the existence of an adjunction.
Corollary C.9.
The forgetful functor has a left adjoint
(C.47) |
Restricting to the subcategory this gives the functor defined in Theorem C.3. Therefore, it is given by sending a linear map to the crossed module as defined in (C.23).
In (3.50), we consider a special class of such free crossed modules given by the functor . In fact, this crossed module has additional structure, since there is a natural action of on . For a group , we define a -Lie algebra to be a Lie algebra which is also a -representation such that the action of on preserves the bracket,
(C.48) |
Corollary C.10.
For any , is a crossed module in the category of -Lie algebras.
Proof.
First, we note that has a natural action by extending the action on using (C.48). The vector space also has a natural action via the usual tensor and wedge products of representations. The map is easily seen to be equivariant. Furthermore, the embedding is equivariant, and thus the action on is equivariant,
(C.49) |
for , and . It is also clear to check that the Peiffer subspace is invariant under the action, so that the action descends to . The equivariance of and the action are also preserved. ∎
Appendix D Compatible Triangulations
In this appendix, we provide details on triangulations, which are used to prove Theorem 6.16 and Proposition 6.18. Here, we must carefully address the relationship between refinements of triangulated representatives and triangulations of their corresponding PLSCs . First, we show that triangulated representatives exist.
Lemma D.1.
There exists a triangulated representative of every .
Proof.
It suffices to show that each marked kite has a triangulated representative. Suppose is minimal with point representation where and .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/1543a99d-8867-4125-8c1d-a75f3e676435/x13.png)
Then, define
(D.1) |
Then, we have . The kites are either triangular, or else have . In the later case and hence can be removed from the list. Let be the resulting element of . It is a triangulated representative of the marked kite . ∎
Let be a triangulated representative of . In general, there is no guarantee that this representative is compatible. Therefore, in order to prove Theorem 6.16, we must modify, or refine, . We will do this by triangulating the PLSC .
Recall that a convex polygon in is the convex hull of a finite set of points which are all contained in a -dimensional affine plane . Such a polygon is non-degenerate if the points are not contained in any affine line . By an edge in , we simply mean a line segment in connecting a pair of distinct points. Furthermore, given an edge or a polygon , we let denote its interior.
Definition D.2.
Let be a collection of edges in and let be a collection of nondegenerate convex polygons. A compatible triangulation for is a compatible non-degenerate PLSC such that
-
(1)
for each -simplex and each edge , if , then .
-
(2)
for each 2-simplex and each polygon , if , then .
-
(3)
.
-
(4)
the edges in and in the boundaries of the polygons of are contained in , the realization of the 1-skeleton.
Lemma D.3.
Let be a compatible triangulation for a pair . Given an edge in or a polygon in , denoted , the subset
(D.2) |
defines a compatible triangulation of .
Proof.
First, we note that is indeed a PLSC, since for any such that , the faces of must also be contained in . The properties of being compatible and non-degenerate are automatically inherited. Next, we note that conditions (1) and (2) of Definition D.2 are properties of the simplices in , and thus are inherited by . For condition (3), note first that follows from the construction. If is a polygon, then is contained in the union of all 2-simplices of . And by condition (2), the simplices which are not in have their interiors disjoint from and so cannot contribute any area. As a result . A similar argument applies if is an edge, using condition (1) instead. Condition (4) follows essentially for topological reasons. ∎
Lemma D.4.
Let and be, respectively, a set of edges and nondegenerate convex polygons in . Then there exists a compatible triangulation for .
Proof.
The general idea behind this proof is to maximally subdivide all polygons and edges in order to obtain the compatible triangulation. The first step is to construct sets of affine planes , affine lines , and vertices .
-
Step 1
The polygons in are supported on a finite set of affine planes, which we take to be .
-
Step 2
The set of lines is constructed as follows. The boundary of each polygon is a collection of line segments. Extend these to lines and add them to . Extend each edge in to a line and add it to . Finally, if two distinct planes intersect along a line, add it to .
-
Step 3
To construct the set of vertices , we start with the endpoints of edges in and the extremal points of the polygons in . If two lines in intersect at a point, we add this to . If a line in intersects a plane in at a point, we add this to . Finally, if two planes in intersect at a point, we add this to .
Restrict attention to a single plane . This plane will contain a subset of lines and a subset of vertices. Suppose that there is a point which is not contained in any line from . Then add a new line to which is contained in and which contains the point . This will lead to new vertices in arising from the intersections between and other lines from . However, there will not be further new vertices arising from intersecting with planes . For this reason, adding will not introduce any new vertices which are contained within a plane , but not within a line from lying in . Therefore, by iterating this process, we may assume that every vertex from is contained in some line from .
Now observe that using the vertices in , each line is decomposed into a finite number of closed segments . Since each edge is a subset of some line in lying between two vertices from , it is decomposed into a union of such line segments . Let denote the collection of all finite line segments whose interiors intersect the interior of one of the edges from . Note that . Indeed, each is contained in some , whereas each decomposes into a union of some .
Next, consider the plane and let denote the subset of lines that lie in . Let denote the complement of the set of lines . It is an open set with the property that the closure of each bounded component is a convex polygon in . Since each is the intersection of a collection of half-planes bounded by a subset of the lines , is decomposed into a union of convex polygons . Let denote the collection of all convex polygons in all planes whose interiors intersect the interior of one of the polygons from . Note that the extremal points of lie in , but that there may be additional points from on the boundary. Note also that . Indeed, each is contained in some , whereas each decomposes into a union of some .
We now construct a PLSC which will be our compatible triangulation.
-
(1)
Let be the collection of all vertices which are either contained in an edge , or in a convex polygon . Endow with an arbitrary order. This will form the set of vertices of our PLSC.
-
(2)
Given each convex polygon , choose a triangulation that makes use of precisely the set of vertices in lying on its boundary. Define , the set of -simplices in , to be the collection of all triples such is one of the triangles from a polygon .
-
(3)
We define , the set of -simplices in , to consist of all faces of all -simplices in , as well as all pairs such that is one of the edges in .
By construction is a compatible non-degenerate PLSC whose piecewise linear realization satisfies
(D.3) |
The remaining conditions for a compatible triangulation are then easy to check. ∎
The strategy to prove Theorem 6.16 is to start with a triangulated representative and take its associated PLSC, , which may not be compatible. Then, we produce a compatible triangulation and use it to construct a new representative . This new representative will have the property that its associated PLSC is a subcomplex of , thereby assuring that it is compatible.
The following lemma starts by treating the case of a triangular kite.
Lemma D.5.
Let be a triangular loop, viewed as a marked triangular kite, and let be its associated 2-simplex, viewed as a convex polygon in . Given a compatible triangulation , there exists a triangulated representative such that in , and such that .
Proof.
The PLSC is homeomorphic to , which is a 2-simplex. Its 1-skeleton is a connected -dimensional simplicial complex which contains , the boundary of , as a subspace. Let be the basepoint of . Because is 1-dimensional, its fundamental group is free. We will need a generating set which is indexed by the 2-simplices of . For this, we choose a spanning tree of . For each 2-simplex , let be the boundary loop of based at , and let be the unique path in connecting to . Define
(D.4) |
which form a generating set for . The inclusion induces a homomorphism
(D.5) |
Therefore, given the generator , its image under admits a factorization
(D.6) |
Now let be the homomorphism from (6.8). Then and
(D.7) |
where and where is a triangular loop. Then, since all paths are contained in the plane spanned by , we obtain a factorization in ,
(D.8) |
Since is a path in connecting vertices of , we can choose a lift of to consisting of vectors in representing edges of . This gives us the desired triangulated representative . ∎
Finally, we treat the general case.
Proof of Theorem 6.16.
Let . By Lemma D.1, there exists a triangulated representative . Let be the associated PLSC, let denote the set of 1-simplices realized as edges in , and let denote the set of 2-simplices realized as convex polygons in . Then, by Lemma D.4, there exists a compatible triangulation .
Given each marked triangular kite , the associated 2-simplex is a polygon in . By Lemma D.3, the subcomplex is a compatible triangulation of . Therefore, by Lemma D.5, there is a triangulated representative
(D.9) |
such that in and such that (after shifting by the displacement of ). Furthermore, consists of a collection of vectors in , which give rise to a subset of edges from . Each edge has the form , where . Because is a compatible triangulation, each such edge decomposes into a sequence of edges , such that is a 1-simplex. As a result, we may replace each with a sequence of vectors which sum to and such that the corresponding edges are 1-simplices of . Let denote the resulting word and note that and are equivalent in . Now define
(D.10) |
By construction, is equivalent to in and is a subcomplex of . Finally, define to be the product of the , for . This is a triangulated representative of and its associated PLSC is a subcomplex of . It is therefore non-degenerate and compatible. Hence is a compatible representative for . ∎
Next, we prove a lemma which is required for Proposition 6.18.
Lemma D.6.
Suppose is a 2-dimensional non-degenerate compatible PLSC whose set of -simplices is denoted by . For each 2-simplex , there exists a compactly supported 2-form such that and such that the support of is disjoint from all other 2-simplices,
(D.11) |
for all such that .
Proof.
Let and let denote the 2-plane which supports . Let denote a closed ball such that . Consider a compactly supported 2-form on which is supported in and satisfies . Using a metric on , pullback along the orthogonal projection to obtain a form . Then, choose a compactly supported smooth bump function such that and which vanishes along each , . Then, we define
(D.12) |
This is the desired -form.
∎
References
- [1] Camilo Arias Abad and Florian Schätz. The A de Rham Theorem and Integration of Representations up to Homotopy. International Mathematics Research Notices, 2013(16):3790–3855, 2013.
- [2] John Baez and Urs Schreiber. Higher Gauge Theory: 2-Connections on 2-Bundles. arXiv:hep-th/0412325, December 2004.
- [3] J. W. Barrett. Holonomy and path structures in general relativity and Yang-Mills theory. International Journal of Theoretical Physics, 30(9):1171–1215, September 1991.
- [4] Jonathan Block and Aaron M. Smith. The higher Riemann–Hilbert correspondence. Advances in Mathematics, 252:382–405, February 2014.
- [5] Horatio Boedihardjo, Xi Geng, Terry Lyons, and Danyu Yang. The signature of a rough path: Uniqueness. Adv. Math., 293:720–737, April 2016.
- [6] Francis C. S. Brown. Multiple zeta values and periods of moduli spaces . Ann. Sci. Éc. Norm. Supér. (4), 42(3):371–489, 2009.
- [7] Kenneth S. Brown. Cohomology of Groups. Springer Science & Business Media, October 1982.
- [8] R. Brown, P.J. Higgins, and R. Sivera. Nonabelian Algebraic Topology: Filtered Spaces, Crossed Complexes, Cubical Homotopy Groupoids. EMS Series of Lectures in Mathematics. European Mathematical Society, 2011.
- [9] Ronald Brown. On the Second Relative Homotopy Group of an Adjunction Space: An Exposition of a Theorem of J. H. C. Whitehead. Journal of the London Mathematical Society, s2-22(1):146–152, August 1980.
- [10] Ronald Brown and Johannes Huebschmann. Identities among relations. Low dimensional topology, Ed. R. Brown and TL Thickstun, London Math. Soc. Lecture Notes, 46:153–202, 1982.
- [11] A. Caetano and R. F. Picken. An axiomatic definition of holonomy. International Journal of Mathematics, 05(06):835–848, December 1994.
- [12] A. P. Caetano and R. F. Picken. On a family of topological invariants similar to homotopy groups. Rend. Istit. Mat. Univ. Trieste, 30:81–90, 1998.
- [13] J. W. Cannon and G. R. Conner. The combinatorial structure of the Hawaiian earring group. Topology and its Applications, 106(3):225–271, October 2000.
- [14] Pierre Cartier. Jacobiennes généralisées, monodromie unipotente et intégrales itérées. Number 161-162, pages Exp. No. 687, 3, 31–52. 1988. Séminaire Bourbaki, Vol. 1987/88.
- [15] Pierre Cartier. Fonctions polylogarithmes, nombres polyzêtas et groupes pro-unipotents. Number 282, pages Exp. No. 885, viii, 137–173. 2002. Séminaire Bourbaki, Vol. 2000/2001.
- [16] Kuo-Tsai Chen. Iterated Integrals and Exponential Homomorphisms†. Proc. Lond. Math. Soc., s3-4(1):502–512, 1954.
- [17] Kuo-Tsai Chen. Integration of paths – a faithful representation of paths by noncommutative formal power series. Trans. Amer. Math. Soc., 89(2):395–407, 1958.
- [18] Ilya Chevyrev, Joscha Diehl, Kurusch Ebrahimi-Fard, and Nikolas Tapia. A multiplicative surface signature through its Magnus expansion. arXiv.2406.16856, June 2024.
- [19] Ilya Chevyrev and Terry Lyons. Characteristic functions of measures on geometric rough paths. The Annals of Probability, 44(6):4049–4082, 2016.
- [20] Ilya Chevyrev and Harald Oberhauser. Signature Moments to Characterize Laws of Stochastic Processes. J. Mach. Learn. Res., 23(176):1–42, 2022.
- [21] Wei-Liang Chow. Über Systeme von liearren partiellen Differentialgleichungen erster Ordnung. Mathematische Annalen, 117(1):98–105, December 1940.
- [22] Lucio Simone Cirio and João Faria Martins. Categorifying the Knizhnik–Zamolodchikov connection. Differential Geometry and its Applications, 30(3):238–261, June 2012.
- [23] G. A. Demessie and C. Sämann. Higher Poincaré lemma and integrability. Journal of Mathematical Physics, 56(8):082902, August 2015.
- [24] Joscha Diehl, Kurusch Ebrahimi-Fard, Fabian Harang, and Samy Tindel. On the signature of an image. arXiv:2403.00130, February 2024.
- [25] Joscha Diehl and Leonard Schmitz. Two-parameter sums signatures and corresponding quasisymmetric functions. arXiv:2210.14247 [math], October 2022.
- [26] Peter K. Friz and Martin Hairer. A Course on Rough Paths: With an Introduction to Regularity Structures. Universitext. Springer International Publishing, 2 edition, 2020.
- [27] Peter K. Friz and Nicolas B. Victoir. Multidimensional Stochastic Processes as Rough Paths: Theory and Applications. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2010.
- [28] William Fulton. Young Tableaux: With Applications to Representation Theory and Geometry. London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1996.
- [29] Chad Giusti, Darrick Lee, Vidit Nanda, and Harald Oberhauser. A topological approach to mapping space signatures. Advances in Applied Mathematics, 163:102787, February 2025.
- [30] V. K. A. M. Gugenheim. On Chen’s iterated integrals. Illinois J. Math., 21(3):703–715, 1977.
- [31] Richard Martin Hain. Iterated Integrals and Homotopy Periods. American Mathematical Soc., 1984.
- [32] Ben Hambly and Terry Lyons. Uniqueness for the signature of a path of bounded variation and the reduced path group. Ann. of Math., 171(1):109–167, 2010.
- [33] Heinz Hopf. Fundamentalgruppe und zweite Bettische Gruppe. Commentarii Mathematici Helvetici, 14(1):257–309, December 1941.
- [34] Mikhail Kapranov. Membranes and higher groupoids. arXiv:1502.06166 [math], February 2015.
- [35] Patrick Kidger and Terry Lyons. Signatory: Differentiable computations of the signature and logsignature transforms, on both {CPU} and {GPU}. In International Conference on Learning Representations, 2021.
- [36] Toshitake Kohno. Formal connections, higher holonomy functors and iterated integrals. Topology and its Applications, 313:107985, May 2022.
- [37] Rafal Komendarczyk, Robin Koytcheff, and Ismar Volić. Diagram Complexes, Formality, and Configuration Space Integrals for Spaces of Braids. The Quarterly Journal of Mathematics, 71(2):729–779, June 2020.
- [38] Darrick Lee. The Surface Signature and Rough Surfaces. arXiv.2406.16857, June 2024.
- [39] Darrick Lee and Harald Oberhauser. Random Surfaces and Higher Algebra. arXiv:2311.08366, November 2023.
- [40] Darrick Lee and Harald Oberhauser. The Signature Kernel, May 2023.
- [41] Terry Lyons. Differential equations driven by rough signals. Rev. Mat. Iberoam., 14(2):215–310, 1998.
- [42] Terry Lyons, Michael Caruana, and Thierry Lévy. Differential Equations Driven by Rough Paths. Éc. Été Probab. St.-Flour. Springer-Verlag, Berlin Heidelberg, 2007.
- [43] João Faria Martins and Roger Picken. On two-dimensional holonomy. Transactions of the American Mathematical Society, 362(11):5657–5695, 2010.
- [44] João Faria Martins and Roger Picken. Surface holonomy for non-abelian 2-bundles via double groupoids. Advances in Mathematics, 226(4):3309–3366, March 2011.
- [45] Andrew McLeod and Terry Lyons. Signature methods in machine learning. EMS Surveys in Mathematical Sciences, February 2025.
- [46] Claudio Meneses. Thin homotopy and the holonomy approach to gauge theories. Contemporary Mathematics, 775:233–253, 2021.
- [47] Christophe Reutenauer. Dimensions and characters of the derived series of the free Lie algebra. In Mots, Lang. Raison. Calc., pages 171–184. Hermès, Paris, 1990.
- [48] Wulf Rossmann. Lie Groups: An Introduction through Linear Groups. Oxford University Press, January 2002.
- [49] Christian Sämann and Lennart Schmidt. Towards an M5-Brane Model II: Metric String Structures. Fortschritte der Physik, 68(8):2000051, 2020.
- [50] Urs Schreiber and Konrad Waldorf. Smooth functors vs. differential forms. Homology, Homotopy and Applications, 13(1):143–203, January 2011.
- [51] Jean-Pierre Serre. Lie Algebras and Lie Groups: 1964 Lectures given at Harvard University. Springer, February 2009.
- [52] Tamer Tlas. On the holonomic equivalence of two curves. International Journal of Mathematics, 27(06):1650055, 2016.
- [53] Csaba Toth, Patric Bonnier, and Harald Oberhauser. Seq2Tens: An Efficient Representation of Sequences by Low-Rank Tensor Projections. arXiv:2006.07027 [cs, stat], 2020.
- [54] Theodore Voronov. On a non-Abelian Poincaré lemma. Proceedings of the American Mathematical Society, 140(8):2855–2872, August 2012.
- [55] J. H. C. Whitehead. Combinatorial homotopy. II. Bulletin of the American Mathematical Society, 55(5):453–496, May 1949.