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

Thin Homotopy and the Signature of Piecewise Linear Surfaces

Francis Bischoff francis.bischoff@uregina.ca Department of Mathematics and Statistics, University of Regina, Regina, SK S4S 0A2, Canada  and  Darrick Lee darrick.lee@ed.ac.uk School of Mathematics and Maxwell Institute, University of Edinburgh, Edinburgh EH9 3FD, Scotland
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 22-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]

1. Introduction

The signature of a path 𝐱=(𝐱1,,𝐱n)C1([0,1],n)\mathbf{x}=(\mathbf{x}^{1},\ldots,\mathbf{x}^{n})\in C^{1}([0,1],\mathbb{R}^{n}) is a non-commutative power series

S0(𝐱)=IS0I(𝐱)eIk=0(n)k\displaystyle S_{0}(\mathbf{x})=\sum_{I}S_{0}^{I}(\mathbf{x})\,e^{I}\in\prod_{k=0}^{\infty}(\mathbb{R}^{n})^{\otimes k} (1.1)

whose coefficients are obtained by taking iterated integrals of the path:

S0I(𝐱)=0t1tk1𝑑𝐱t1i1𝑑𝐱tkik\displaystyle S_{0}^{I}(\mathbf{x})=\int_{0\leq t_{1}\leq...\leq t_{k}\leq 1}d\mathbf{x}^{i_{1}}_{t_{1}}\cdots d\mathbf{x}^{i_{k}}_{t_{k}} (1.2)

where I=(i1,,ik)[n]kI=(i_{1},...,i_{k})\in[n]^{k} is a multi-index and eI=ei1eike^{I}=e^{i_{1}}\otimes\ldots\otimes e^{i_{k}}, where {e1,,en}\{e^{1},\ldots,e^{n}\} is a basis of n\mathbb{R}^{n}. 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 MM arises. This leads to the thin fundamental group, π1th(M)\pi_{1}^{\operatorname{th}}(M), 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 GG, this states that there is an equivalence between the category of all GG-connections on MM and the category of smooth GG-representations of π1th(M)\pi_{1}^{\operatorname{th}}(M). In fact, the path signature can be seen as the holonomy of a universal translation-invariant connection on n\mathbb{R}^{n} valued in the free Lie algebra generated by n\mathbb{R}^{n}.

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 22-connection on n\mathbb{R}^{n}. This can be formulated as a homomorphism of crossed modules 𝐒:𝝉(n)𝐊^(n)\mathbf{S}:{\boldsymbol{\tau}}(\mathbb{R}^{n})\to\hat{\mathbf{K}}(\mathbb{R}^{n}), between a crossed module 𝝉(n){\boldsymbol{\tau}}(\mathbb{R}^{n}) of thin equivalence classes of surfaces in Euclidean space, and a crossed module 𝐊^(n)\hat{\mathbf{K}}(\mathbb{R}^{n}) of ‘formal surfaces’, which is defined by a formal integration of a free crossed module of Lie algebras 𝖐^(n)\hat{{\boldsymbol{\mathfrak{k}}}}(\mathbb{R}^{n}). 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

f:S22n\displaystyle f:S^{2}\to\mathbb{RP}^{2}\to\mathbb{R}^{n} (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 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V), which is functorial in the vector space VV. Because the formal integration 𝐊^(V)\hat{\mathbf{K}}(V) of the free crossed module 𝖐^(V)\hat{{\boldsymbol{\mathfrak{k}}}}(V) is also functorial in VV, this allows us to define the surface signature 𝐒PL:𝐏𝐋𝐊^\mathbf{S}_{\operatorname{PL}}:{\boldsymbol{\operatorname{PL}}}\Rightarrow\hat{\mathbf{K}} 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 𝐒PL:𝐏𝐋𝐊^\mathbf{S}_{\operatorname{PL}}:{\boldsymbol{\operatorname{PL}}}\Rightarrow\hat{\mathbf{K}} is the unique natural transformation extending the piecewise linear path signature. Furthermore, the smooth surface signature 𝐒:𝛕𝐊^\mathbf{S}:{\boldsymbol{\tau}}\Rightarrow\hat{\mathbf{K}} 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 𝐗\mathbf{X} be a piecewise linear surface such that its signature is trivial, S1(𝐗)=0.S_{1}(\mathbf{X})=0. Then 𝐗\mathbf{X} 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 𝐗\mathbf{X} 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 22-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 𝐗\mathbf{X} with vanishing signature has the property that

    𝐗ω=0for all compactly support 2-formsωΩc2(V).\displaystyle\int_{\mathbf{X}}\omega=0\quad\text{for all compactly support $2$-forms}\quad\omega\in\Omega^{2}_{c}(V). (1.4)
  • b)

    In Section 6, we complete the proof of injectivity. A closed surface 𝐗:S2V\mathbf{X}:S^{2}\to V which satisfies (1.4) has vanishing homology in its image C=im(𝐗)C=\operatorname{im}(\mathbf{X}),

    H2(𝐗)([S2])=0H2(C).\displaystyle H_{2}(\mathbf{X})([S^{2}])=0\in H_{2}(C). (1.5)

    To conclude that 𝐗\mathbf{X} is thinly null homotopic, it suffices to show that [𝐗]=0π2(C)[\mathbf{X}]=0\in\pi_{2}(C), and by Hurewicz, this would be true if π1(C)=0\pi_{1}(C)=0. In other words, the obstruction is the fundamental group of CC. Hence, by attaching sufficiently many discs to CC, 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 22-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 nn letters can be viewed as thin homotopy equivalence classes of lattice paths on n\mathbb{Z}^{n}: word reduction coincides with path reduction via retracing. In Section 2.4, we construct the group PL0(V)\operatorname{PL}_{0}(V) of piecewise linear paths on VV as the quotient of the free group on VV, 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]

In Section 5.2, we extend this construction to define a crossed module of piecewise linear surfaces 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V). The distinction between local and non-local cancellations is reflected in our construction. We first construct a pre-crossed module 𝐏𝐋~(V)\widetilde{{\boldsymbol{\operatorname{PL}}}}(V) 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 Peiff(V)\mathrm{Peiff}(V) 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 Peiff(V)\mathrm{Peiff}(V) 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 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V), 𝝉(V){\boldsymbol{\tau}}(V), and 𝐊^(V)\hat{\mathbf{K}}(V), 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 𝐗\mathbf{X}, 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 𝒞(𝐗)\mathcal{C}(\mathbf{X}) obtained by coning off the boundary of 𝐗\mathbf{X}. Schematically, this abelian component is given by the following formula

S1Γ(𝐗)=k01k!𝒞(𝐗)(idV)k,S_{1}^{\Gamma}(\mathbf{X})=\sum_{k\geq 0}\frac{1}{k!}\int_{\mathcal{C}(\mathbf{X})}(\mathrm{id}_{V})^{k},

where (idV)k(\mathrm{id}_{V})^{k} is viewed as a function on VV valued in the symmetric power Sk(V)S^{k}(V) (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 C1C^{1} 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 G=π1(C)G=\pi_{1}(C) of the image C=im(𝐗)C=\operatorname{im}(\mathbf{X}) of a surface 𝐗\mathbf{X}, and those which have further complexity. We show that thinly null homotopic surfaces in this latter case are classified by H3(G)H_{3}(G), 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, VV denotes a finite-dimensional real vector space VdV\cong\mathbb{R}^{d}, and we let 𝖵𝖾𝖼𝗍\mathsf{Vect} denote the category of finite dimensional vector spaces and linear maps. Furthermore, we always consider paths with sitting instants.

Definition 2.1.

A path 𝐱C([0,1],V)\mathbf{x}\in C([0,1],V) has sitting instants if there exists some ϵ>0\epsilon>0 such that

𝐱s=𝐱0and𝐱1s=𝐱1for all s[0,ϵ].\displaystyle\mathbf{x}_{s}=\mathbf{x}_{0}\quad\text{and}\quad\mathbf{x}_{1-s}=\mathbf{x}_{1}\quad\text{for all $s\in[0,\epsilon]$}. (2.1)

Unless otherwise specified, we assume all paths have sitting instants.

This ensures that when we consider composable paths 𝐱,𝐲C1([0,1],V)\mathbf{x},\mathbf{y}\in C^{1}([0,1],V), such that 𝐱1=𝐲0\mathbf{x}_{1}=\mathbf{y}_{0}, with some smoothness condition, the concatenation

(𝐱𝐲)s{𝐱2s:s[0,12]𝐲2s1:s[12,1]\displaystyle(\mathbf{x}\star\mathbf{y})_{s}\coloneqq\left\{\begin{array}[]{cl}\mathbf{x}_{2s}&:s\in[0,\frac{1}{2}]\\ \mathbf{y}_{2s-1}&:s\in[\frac{1}{2},1]\end{array}\right. (2.4)

preserves the smoothness condition 𝐱𝐲C1([0,1],V)\mathbf{x}\star\mathbf{y}\in C^{1}([0,1],V).

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 VV, all principal bundles can be assumed trivial and as a result, we use the following simplified definition of a connection.

Definition 2.2.

Let 𝔤\mathfrak{g} be a Lie algebra. A 𝔤\mathfrak{g}-connection on VV is a Lie algebra valued 11-form γ0Ω1(V,𝔤)\gamma_{0}\in\Omega^{1}(V,\mathfrak{g}). A connection is translation-invariant if it has the form

γ0=i=1dγ0idzi,\displaystyle\gamma_{0}=\sum_{i=1}^{d}\gamma_{0}^{i}\,dz_{i}, (2.5)

where γ0i𝔤\gamma_{0}^{i}\in\mathfrak{g} and ziz_{i} are linear coordinates on VV. The curvature of γ0\gamma_{0}, denoted κγ0Ω2(V,𝔤)\kappa^{\gamma_{0}}\in\Omega^{2}(V,\mathfrak{g}), is defined by the following formula

κγ0dγ0+12[γ0,γ0].\displaystyle\kappa^{\gamma_{0}}\coloneqq d\gamma_{0}+\frac{1}{2}[\gamma_{0},\gamma_{0}]. (2.6)
Definition 2.3.

Let GG be a Lie group with Lie algebra 𝔤\mathfrak{g}. Let γ0\gamma_{0} be a 𝔤\mathfrak{g}-connection and 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) a path. Consider the following differential equation for F0γ0(𝐱):[0,1]GF_{0}^{\gamma_{0}}(\mathbf{x}):[0,1]\to G,

dF0γ0(𝐱)tdt=dLF0γ0(𝐱)tγ0(d𝐱tdt),F0γ0(𝐱)0=e.\displaystyle\frac{dF_{0}^{\gamma_{0}}(\mathbf{x})_{t}}{dt}=dL_{F_{0}^{\gamma_{0}}(\mathbf{x})_{t}}\gamma_{0}\left(\frac{d\mathbf{x}_{t}}{dt}\right),\quad F_{0}^{\gamma_{0}}(\mathbf{x})_{0}=e. (2.7)

The parallel transport of γ0\gamma_{0} along 𝐱\mathbf{x} is defined to be

F0γ0(𝐱)F0γ0(𝐱)1.\displaystyle F_{0}^{\gamma_{0}}(\mathbf{x})\coloneqq F_{0}^{\gamma_{0}}(\mathbf{x})_{1}. (2.8)

The path signature is the parallel transport of the universal translation-invariant connection. This is a connection ζ0\zeta_{0} valued in the free Lie algebra generated by VV, 𝔨0𝖥𝖫(V)\mathfrak{k}_{0}\coloneqq\mathsf{FL}(V), and it can be understood as the identity endomorphism of VV

ζ0=idVVVΩ1(V,𝔨0),\displaystyle\zeta_{0}=\mathrm{id}_{V}\in V^{*}\otimes V\subset\Omega^{1}(V,\mathfrak{k}_{0}), (2.9)

where we view VV^{*} as the subspace of translation invariant 11 forms, and VV as the subspace of generators of the free Lie algebra. If {ei}i=1d\{e_{i}\}_{i=1}^{d} form a basis of VV, with dual basis {zi}i=1dV\{z_{i}\}_{i=1}^{d}\in V^{*}, viewed as coordinates on VV, then the connection has the following explicit expression

ζ0=i=1ddzieiΩ1(V,𝔨0).\displaystyle\zeta_{0}=\sum_{i=1}^{d}dz_{i}e_{i}\in\Omega^{1}(V,\mathfrak{k}_{0}). (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 𝔨0\mathfrak{k}_{0}, defined recursively as

LCS1(𝔨0)𝔨0andLCSr(𝔨0)[𝔨0,LCSr1(𝔨0)].\displaystyle{\operatorname{LCS}}_{1}(\mathfrak{k}_{0})\coloneqq\mathfrak{k}_{0}\quad\text{and}\quad{\operatorname{LCS}}_{r}(\mathfrak{k}_{0})\coloneqq[\mathfrak{k}_{0},{\operatorname{LCS}}_{r-1}(\mathfrak{k}_{0})]. (2.11)

Then, we define the nn-truncated free Lie algebra 𝔨0(n)\mathfrak{k}_{0}^{(\leq n)} as

𝔨0(n)𝔨0/LCSn+1(𝔨0).\displaystyle\mathfrak{k}_{0}^{(\leq n)}\coloneqq\mathfrak{k}_{0}/{\operatorname{LCS}}_{n+1}(\mathfrak{k}_{0}). (2.12)

As we assume that VV is finite dimensional, 𝔨0(n)\mathfrak{k}_{0}^{(\leq n)} is a finite-dimensional Lie algebra. Then, we can explicitly integrate 𝔨0(n)\mathfrak{k}_{0}^{(\leq n)} to the Lie group of exponential elements K0(n)K_{0}^{(\leq n)} of the truncated tensor algebra T0(n)(V)=k=0nVkT^{(\leq n)}_{0}(V)=\prod_{k=0}^{n}V^{\otimes k}. This is defined as follows

K0(n)(V){exp(x)T(n)(V):x𝔨0(n)}.\displaystyle K_{0}^{(\leq n)}(V)\coloneqq\left\{\exp(x)\in T^{(\leq n)}(V)\,:\,x\in\mathfrak{k}_{0}^{(\leq n)}\right\}. (2.13)

We denote the projective limit of these Lie groups, and their corresponding Lie algebras, by

K^0(V)limK0(n)(V)T((V))=k=0Vkand𝔨^0(V)=lim𝔨0(n)(V).\displaystyle\hat{K}_{0}(V)\coloneqq\lim_{\longleftarrow}K_{0}^{(\leq n)}(V)\subset T\mkern-0.25mu\mathbin{\left(\mkern-3.5mu\left({V}\right)\mkern-3.5mu\right)}=\prod_{k=0}^{\infty}V^{\otimes k}\quad\text{and}\quad\hat{\mathfrak{k}}_{0}(V)=\lim_{\longleftarrow}\mathfrak{k}_{0}^{(\leq n)}(V). (2.14)

Given a linear map f:VWf:V\to W, there is an induced group homomorphism K0(n)(V)K0(n)(W)K_{0}^{(\leq n)}(V)\to K_{0}^{(\leq n)}(W) for all nn\in\mathbb{N}. This induces a group homomorphism K^0(V)K^0(W)\hat{K}_{0}(V)\to\hat{K}_{0}(W). Hence, we get a well-defined functor

K^0:𝖵𝖾𝖼𝗍𝖦𝗋𝗉.\displaystyle\hat{K}_{0}:\mathsf{Vect}\to\mathsf{Grp}. (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 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V). For nn\in\mathbb{N}, consider the parallel transport of the universal connection ζ0\zeta_{0} of Equation (2.10) projected onto 𝔨0(n)\mathfrak{k}_{0}^{(\leq n)}. In particular, consider the following differential equation for S0(n)(𝐱):[0,1]K0(n)S_{0}^{(n)}(\mathbf{x}):[0,1]\to K_{0}^{(\leq n)},

dS0(n)(𝐱)tdt=S0(n)(𝐱)td𝐱tdt,S0(n)(𝐱)0=1.\displaystyle\frac{dS_{0}^{(n)}(\mathbf{x})_{t}}{dt}=S_{0}^{(n)}(\mathbf{x})_{t}\otimes\frac{d\mathbf{x}_{t}}{dt},\quad S_{0}^{(n)}(\mathbf{x})_{0}=1. (2.16)

The nn-truncated path signature of 𝐱\mathbf{x} is defined to be S0(n)(𝐱)S0(n)(𝐱)1S_{0}^{(n)}(\mathbf{x})\coloneqq S_{0}^{(n)}(\mathbf{x})_{1}. Then, we define the path signature of 𝐱\mathbf{x} to be the projective limit

S0(𝐱)limS0(n)(𝐱)K^0.\displaystyle S_{0}(\mathbf{x})\coloneqq\lim_{\longleftarrow}S_{0}^{(n)}(\mathbf{x})\in\hat{K}_{0}. (2.17)
Remark 2.5.

The completion of the tensor algebra is often considered analytically by defining appropriate norms on T(n)(V)T^{(\leq n)}(V), 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 𝐱,𝐲C1([0,1],V)\mathbf{x},\mathbf{y}\in C^{1}([0,1],V) are thin homotopy equivalent, denoted 𝐱th𝐲\mathbf{x}\sim_{\operatorname{th}}\mathbf{y}, if there exists an endpoint preserving smooth homotopy hC1([0,1]2,V)h\in C^{1}([0,1]^{2},V) between 𝐱\mathbf{x} and 𝐲\mathbf{y} such that

  • (homotopy condition) h0,t=𝐱th_{0,t}=\mathbf{x}_{t} and h1,t=𝐲th_{1,t}=\mathbf{y}_{t};

  • (thinness condition) rank(dh)1\operatorname{rank}(dh)\leq 1, where dhdh is the differential of hh.

Thin homotopy defines an equivalence relation on paths which is compatible with concatenation. The thin fundamental groupoid of VV is defined to be the set of equivalence classes

𝔗1(V)C1([0,1],V)/th.\displaystyle\mathfrak{T}_{1}(V)\coloneqq C^{1}([0,1],V)/\sim_{\operatorname{th}}. (2.18)

It is a groupoid over VV with product given by the concatenation of paths. Given a Lie group GG with Lie algebra 𝔤\mathfrak{g}, the parallel transport of a connection γ0Ω1(V,𝔤)\gamma_{0}\in\Omega^{1}(V,\mathfrak{g}) defines a groupoid homomorphism

F0γ0:𝔗1(V)G.\displaystyle F_{0}^{\gamma_{0}}:\mathfrak{T}_{1}(V)\to G. (2.19)

There is a natural action of the additive group VV on 𝔗1(V)\mathfrak{T}_{1}(V) by groupoid automorphisms given by translating paths. We define the thin path group to be the quotient

τ1(V)𝔗1(V)/trans\displaystyle\tau_{1}(V)\coloneqq\mathfrak{T}_{1}(V)/\sim_{\operatorname{trans}} (2.20)

and we note that it is a group. Our convention will be to use paths starting at the origin, 𝐱0=0\mathbf{x}_{0}=0, as representatives of the translation equivalence classes. There is a well-defined group homomorphism t:τ1(V)Vt:\tau_{1}(V)\to V given by sending a path 𝐱\mathbf{x} to its displacement 𝐱1𝐱0V\mathbf{x}_{1}-\mathbf{x}_{0}\in V. The thin fundamental groupoid can then be recovered as the corresponding action groupoid

𝔗1(V)Vτ1(V).\displaystyle\mathfrak{T}_{1}(V)\cong V\rtimes\tau_{1}(V). (2.21)

Given a linear map ϕ:VW\phi:V\to W and a path 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V), we can define a new path ϕ(𝐱)C1([0,1],W)\phi(\mathbf{x})\in C^{1}([0,1],W) by ϕ(𝐱)t=ϕ(𝐱t)\phi(\mathbf{x})_{t}=\phi(\mathbf{x}_{t}). This preserves thin homotopy classes of maps, and is equivariant with respect to translation. Thus, we obtain a thin path group functor

τ1:𝖵𝖾𝖼𝗍𝖦𝗋𝗉.\displaystyle\tau_{1}:\mathsf{Vect}\to\mathsf{Grp}. (2.22)

Because the universal connection ζ0\zeta_{0} is translation invariant, its parallel transport is invariant under both thin homotopy and translations. Therefore, the path signature defines a homomorphism S0:τ1(V)K^0(V)S_{0}:\tau_{1}(V)\to\hat{K}_{0}(V). Furthermore, by [27, Proposition 7.52], this fits into a natural transformation

S0:τ1K^0.\displaystyle S_{0}:\tau_{1}\Rightarrow\hat{K}_{0}. (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 C1C^{1} 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. (1)

    Reparametrizations: Given a path 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) and a reparametrization ϕ:[0,1][0,1]\phi:[0,1]\to[0,1], then 𝐱th𝐱ϕ\mathbf{x}\sim_{\operatorname{th}}\mathbf{x}\circ\phi.

  2. (2)

    Retracings: Given paths 𝐱,𝐲,𝐳C1([0,1],V)\mathbf{x},\mathbf{y},\mathbf{z}\in C^{1}([0,1],V), we say that a retracing is a path segment of the form 𝐳𝐳1\mathbf{z}\star\mathbf{z}^{-1}. Paths which differ by retracings are thin homotopy equivalent:

    𝐱𝐳𝐳1𝐲th𝐱𝐲.\displaystyle\mathbf{x}\star\mathbf{z}\star\mathbf{z}^{-1}\star\mathbf{y}\sim_{\operatorname{th}}\mathbf{x}\star\mathbf{y}. (2.24)

    The path 𝐱𝐲\mathbf{x}\star\mathbf{y} is called a reduction of 𝐱𝐳𝐳1𝐲\mathbf{x}\star\mathbf{z}\star\mathbf{z}^{-1}\star\mathbf{y}.

[Uncaptioned image]

We will focus mainly on thinly null-homotopic paths: paths that are thin homotopy equivalent to the constant path. This is because two paths 𝐱\mathbf{x} and 𝐲\mathbf{y} can then be defined to be thin homotopy equivalent if 𝐱𝐲1\mathbf{x}\star\mathbf{y}^{-1} 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 C1C^{1} paths, where there may be infinitely many retracings, using the notion of transfinite words.

Definition 2.7.

[13, Definition 3.1] Let EE be a set consisting of an alphabet, and let E1E^{-1} denote a formal inverse set. A transfinite word over EE is a function W:BEE1W:B\to E\cup E^{-1}, where BB is a totally ordered set, such that W1(e)W^{-1}(e) is finite for each eEE1e\in E\cup E^{-1}. 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 𝐱:[0,1]V\mathbf{x}:[0,1]\to V be a C1C^{1} path. There exists a collection of mutually disjoint subsets {An}n=0\{A_{n}\}_{n=0}^{\infty} such that

  1. (1)

    A0A_{0} is closed, AnA_{n} is open for n>0n>0, and n=0An=[0,1]\bigcup_{n=0}^{\infty}A_{n}=[0,1].

  2. (2)

    If tAnt\in A_{n}, then the set inverse 𝐱1(𝐱t)An\mathbf{x}^{-1}(\mathbf{x}_{t})\subset A_{n} and for n>0n>0, then |𝐱1(𝐱t)|=n|\mathbf{x}^{-1}(\mathbf{x}_{t})|=n.

  3. (3)

    𝐱t0\mathbf{x}^{\prime}_{t}\neq 0 if tAnt\in A_{n} for n>0n>0, while 𝐱t=0\mathbf{x}^{\prime}_{t}=0 for tA0t\in A_{0}^{\circ} (the interior).

  4. (4)

    Each AnA_{n} for n>0n>0 is a union of disjoint open intervals. The path 𝐱\mathbf{x} restricted to any such interval is an embedding. The image of any two such embeddings are either disjoint or identical.

Here, the set A0A_{0} is used to “collect” all regions of the interval in which the path is constant. Let BB denote the set of intervals from AnA_{n} for all n>0n>0, equipped with a linear order inherited from \mathbb{R}. Let E={𝐱(a):aB}E=\{\mathbf{x}(a)\,:\,a\in B\} be the set of embedded arcs (see (4) above) where we remove repetitions (up to reparametrization and switch in orientation). Let E1E^{-1} be a copy of EE denoting formal inverses. Then, we define the transfinite word associated to 𝐱\mathbf{x} to be the mapping W𝐱:BEE1W_{\mathbf{x}}:B\to E\cup E^{-1} sending each interval to its corresponding arc, taking into account the orientation. Any arc eEE1e\in E\cup E^{-1} has finite length, and thus W𝐱1(e)W_{\mathbf{x}}^{-1}(e) must be finite because 𝐱\mathbf{x} is C1C^{1}.

Definition 2.9.

The path 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) is word reduced if W𝐱W_{\mathbf{x}} 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 TT is an \mathbb{R}-tree if for any pair of points x,yTx,y\in T, all topological embeddings σ:[0,1]T\sigma:[0,1]\to T where σ0=x\sigma_{0}=x and σ1=y\sigma_{1}=y have the same image.

The following theorem collects known results relating various definitions of thin homotopy.

Theorem 2.11.

Let 𝐱:[0,1]V\mathbf{x}:[0,1]\to V be a C1C^{1} path with vanishing derivative at the endpoints. It is thinly null-homotopic or tree-like if any of the following equivalent definitions hold:

  1. (W1)

    Word Condition. The path 𝐱\mathbf{x} is word reduced.

  2. (H1G)

    Holonomy Condition. For a semi-simple Lie group GG with Lie algebra 𝔤\mathfrak{g}, the parallel transport of every smooth 𝔤\mathfrak{g}-connection ωΩ1(V,𝔤)\omega\in\Omega^{1}(V,\mathfrak{g}) along 𝐱\mathbf{x} is trivial222Each semi-simple GG is treated as an independent condition..

  3. (R1)

    Rank Condition. There exists an endpoint-preserving C1C^{1} homotopy h:[0,1]2Vh:[0,1]^{2}\to V from 𝐱\mathbf{x} to the constant path at 0 such that the rank of dhdh is 1\leq 1 everywhere.

  4. (I1)

    Image Condition. [3] There exists an endpoint-preserving homotopy h:[0,1]2Vh:[0,1]^{2}\to V from 𝐱\mathbf{x} to the constant path at 0 such that im(h)im(𝐱)\operatorname{im}(h)\subset\operatorname{im}(\mathbf{x}).

  5. (F1)

    Factorization Condition. [5] There exists a factorization of 𝐱\mathbf{x} through an \mathbb{R}-tree TT, namely 𝐱:[0,1]TV\mathbf{x}:[0,1]\to T\to V.

  6. (A1)

    Analytic Condition. [32] There exists a Lipschitz function h:[0,1]h:[0,1]\to\mathbb{R} such that h(t)0h(t)\geq 0 for all t[0,1]t\in[0,1], h(0)=h(1)h(0)=h(1) and if h(s)=h(t)=infsuth(u)h(s)=h(t)=\inf_{s\leq u\leq t}h(u), then 𝐱(s)=𝐱(t)\mathbf{x}(s)=\mathbf{x}(t). The function hh is called the height function.

  7. (S1)

    Path Signature Condition. The path signature S0S_{0} of 𝐱\mathbf{x} is trivial, S0(𝐱)=1S_{0}(\mathbf{x})=1.

Proof.

In [52, Theorem 3] it is proved that (W1), (H1G) for all semi-simple GG, and (R1) are equivalent. Furthermore, in the proof of [52, Theorem 2] which shows (W1) \implies (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 C1C^{1} 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 C1C^{1} setting.

In order to connect these two classes of results, we show that (S1) implies (H1G) for the specific case of G=SL2()G={\operatorname{SL}}_{2}(\mathbb{R}). Suppose 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) such that S0(𝐱)=1S_{0}(\mathbf{x})=1. Suppose to the contrary that there exists a smooth connection

ω=(ω1ω2ω3ω1)Ω1(V,𝔰𝔩2()),\displaystyle\omega=\begin{pmatrix}\omega_{1}&\omega_{2}\\ \omega_{3}&-\omega_{1}\end{pmatrix}\in\Omega^{1}(V,\mathfrak{sl}_{2}(\mathbb{R})), (2.25)

where ωiΩ1(V)\omega_{i}\in\Omega^{1}(V), such that the holonomy (with initial condition at the identity) is not trivial F0ω(𝐱)IF_{0}^{\omega}(\mathbf{x})\neq I. The holonomy can be expressed in terms of the iterated integrals of ω\omega, where specific entries are iterated integrals of ω1,ω2,ω3\omega_{1},\omega_{2},\omega_{3}. 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 C1C^{1} 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 VV, consider the free monoid (𝖥𝖬𝗈𝗇(V),)(\mathsf{FMon}(V),\star) on the underlying set of VV. We define the group of piecewise-linear paths as

PL0(V)𝖥𝖬𝗈𝗇(V)/\displaystyle\operatorname{PL}_{0}(V)\coloneqq\mathsf{FMon}(V)/\sim (2.26)

subject to the following relations:

  1. (PL0.1)

    (v,w)(v+w)(v,w)\sim(v+w) if vv and ww are linearly dependent;

  2. (PL0.2)

    (0)(0)\sim\emptyset, where 0V0\in V and \emptyset is the identity in 𝖥𝖬𝗈𝗇(V)\mathsf{FMon}(V) (the empty word).

With these relations PL0(V)\operatorname{PL}_{0}(V) is a group. Indeed, the inverse of (v1,,vk)PL0(V)(v_{1},\ldots,v_{k})\in\operatorname{PL}_{0}(V) is

(v1,,vk)1(vk,,v1).\displaystyle(v_{1},\ldots,v_{k})^{-1}\coloneqq(-v_{k},\ldots,-v_{1}). (2.27)

An important property of PL0(V)\operatorname{PL}_{0}(V) 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 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V) has a unique minimal representative

𝐱=(v1,,vn)min.\displaystyle\mathbf{x}=(v_{1},\ldots,v_{n})_{\min}. (2.28)

This is a word with the property that all vi0v_{i}\neq 0, and every consecutive pair (vi,vi+1)(v_{i},v_{i+1}) is linearly independent in VV. We use the subscript ()min(\cdot)_{\min} to denote the minimal representative.

Consider the map of sets ηV:VPL0(V)\eta_{V}:V\to\operatorname{PL}_{0}(V) given by the inclusion ηV(v)=(v)\eta_{V}(v)=(v). This has the property that if we restrict to a 1-dimensional subspace UVU\subset V, then ηV\eta_{V} is a group homomorphism. This gives rise to a universal property for PL0(V)\operatorname{PL}_{0}(V).

Lemma 2.13.

Let VV be a vector space and let GG be a group. Let f:VGf:V\to G be a map which restricts to a group homomorphism on subspaces UVU\subset V of dimension 1, and satisfies f(0)=ef(0)=e, where eGe\in G is the identity. Then, there exists a unique group homomorphism F:PL0(V)GF:\operatorname{PL}_{0}(V)\to G such that FηV=fF\circ\eta_{V}=f.

Proof.

Uniqueness of FF is immediate because PL0(V)\operatorname{PL}_{0}(V) is generated by VV. For existence, by the universal property of free monoids, there exists a unique monoid morphism f~:𝖥𝖬𝗈𝗇(V)G\tilde{f}:\mathsf{FMon}(V)\to G. First, note that f~(0)=f(0)=e=f~()\tilde{f}(0)=f(0)=e=\tilde{f}(\emptyset), where the second equality holds by assumption. Second, if v,wVv,w\in V are contained in a 1-dimensional subspace, we have

f~(v,w)=f~(v)f~(w)=f(v)f(w)=f(v+w)=f~(v+w).\displaystyle\tilde{f}(v,w)=\tilde{f}(v)\cdot\tilde{f}(w)=f(v)\cdot f(w)=f(v+w)=\tilde{f}(v+w). (2.29)

Therefore, this descends to a homomorphism F:PL0(V)GF:\operatorname{PL}_{0}(V)\to G such that FηV=fF\circ\eta_{V}=f.

Corollary 2.14.

The group of piecewise linear paths defines a functor

PL0:𝖵𝖾𝖼𝗍𝖦𝗋𝗉.\displaystyle\operatorname{PL}_{0}:\mathsf{Vect}\to\mathsf{Grp}. (2.30)
Proof.

Suppose ϕ:VW\phi:V\to W is a linear map. Then ηWϕ:VPL0(W)\eta_{W}\circ\phi:V\to\operatorname{PL}_{0}(W) restricts to a group homomorphism on each one-dimensional subspace. By the universal property in Lemma 2.13, there is a unique group homomorphism PL0(ϕ):PL0(V)PL0(W)\operatorname{PL}_{0}(\phi):\operatorname{PL}_{0}(V)\to\operatorname{PL}_{0}(W) such that PL0(ϕ)ηV=ηWϕ\operatorname{PL}_{0}(\phi)\circ\eta_{V}=\eta_{W}\circ\phi. This uniqueness implies functoriality. ∎

For later use, we will also consider the notion of the span of a path.

Corollary 2.15.

Let 𝒮(V)\mathcal{S}(V) be the set of linear subspaces of VV. There is a well-defined map

span:PL0(V)𝒮(V)given by𝐱span(v1,,vk),\displaystyle\operatorname{span}:\operatorname{PL}_{0}(V)\to\mathcal{S}(V)\quad\text{given by}\quad\mathbf{x}\mapsto\operatorname{span}(v_{1},\ldots,v_{k}), (2.31)

where (v1,,vk)min(v_{1},\ldots,v_{k})_{\min} is the minimal representative of 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V). If ϕ:VW\phi:V\to W is a linear map, then

span(PL0(ϕ)(𝐱))ϕ(span(𝐱)).\displaystyle\operatorname{span}(\operatorname{PL}_{0}(\phi)(\mathbf{x}))\subset\phi(\operatorname{span}(\mathbf{x})). (2.32)

The universal property of Lemma 2.13 provides an effective method for constructing homomorphisms out of PL0(V)\operatorname{PL}_{0}(V). For example, the identity map idV:VV\mathrm{id}_{V}:V\to V automatically extends to a homomorphism t:PL0(V)Vt:\operatorname{PL}_{0}(V)\to V, and the corresponding action groupoid VPL0(V)V\rtimes\operatorname{PL}_{0}(V) is the piecewise linear analogue of 𝔗1(V)\mathfrak{T}_{1}(V). Next, consider the map r0:Vτ1(V)r_{0}:V\to\tau_{1}(V) defined by r0(v)[𝐱v]r_{0}(v)\coloneqq[\mathbf{x}^{v}], where [𝐱v][\mathbf{x}^{v}] is the equivalence class of the path

𝐱v(t)ψ(t)v,\mathbf{x}^{v}(t)\coloneqq\psi(t)\cdot v,

where ψC([0,1],[0,1])\psi\in C^{\infty}([0,1],[0,1]) is a surjective map with sitting instants. Since r0r_{0} restricts to a homomorphism on 11-dimensional subspaces of VV, by Lemma 2.13, it induces a group homomorphism,

R0:PL0(V)τ1(V),\displaystyle R_{0}:\operatorname{PL}_{0}(V)\to\tau_{1}(V), (2.33)

which we call the realization map. In fact, as we vary VV, these maps fit into a natural transformation

R0:PL0τ1.\displaystyle R_{0}:\operatorname{PL}_{0}\Rightarrow\tau_{1}. (2.34)
Lemma 2.16.

For all V𝖵𝖾𝖼𝗍V\in\mathsf{Vect}, the realization map R0:PL0(V)τ1(V)R_{0}:\operatorname{PL}_{0}(V)\to\tau_{1}(V) is injective.

Proof.

Let 𝐯=(v1,,vk)minPL0(V)\mathbf{v}=(v_{1},\ldots,v_{k})_{\min}\in\operatorname{PL}_{0}(V) be a minimal representative, and suppose R0(𝐯)R_{0}(\mathbf{v}) is trivial. The path R0(v1,,vk)R_{0}(v_{1},\ldots,v_{k}) can only be thinly null homotopic if there are retracings. Since the representative is minimal, this cannot occur if k>0k>0. Hence 𝐯=0\mathbf{v}=\emptyset_{0}. ∎

Finally, consider the map s0S0r0:VK^0(V)s_{0}\coloneqq S_{0}\circ r_{0}:V\to\hat{K}_{0}(V), which can be expressed explicitly as

s0(v)=k=0vkk!.\displaystyle s_{0}(v)=\sum_{k=0}^{\infty}\frac{v^{\otimes k}}{k!}. (2.35)

Again, s0s_{0} restricts to a group homomorphism on 1-dimensional subspaces of VV, and so by Lemma 2.13, we obtain a homomorphism

SPL,0:PL0(V)K^0(V),\displaystyle S_{\operatorname{PL},0}:\operatorname{PL}_{0}(V)\to\hat{K}_{0}(V), (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 R0R_{0}, S0S_{0}, and SPL,0S_{\operatorname{PL},0} are natural transformations which factor as

SPL,0:PL0R0τ1S0K^0.\displaystyle S_{\operatorname{PL},0}:\operatorname{PL}_{0}\xRightarrow{R_{0}}\tau_{1}\xRightarrow{S_{0}}\hat{K}_{0}. (2.37)

The construction of PL0(V)\operatorname{PL}_{0}(V) 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 PL0(V)\operatorname{PL}_{0}(V).

Proposition 2.18.

Let P:AVP:A\to V be a map of sets with the property that P(a)P(a) and P(a)P(a^{\prime}) are linearly independent if aaa\neq a^{\prime} (in particular, P(a)0P(a)\neq 0 for all aAa\in A). Then the induced homomorphism P~:𝖥𝖦(A)PL0(V)\tilde{P}:\mathsf{FG}(A)\to\operatorname{PL}_{0}(V) is injective.

Proof.

Let P~:𝖥𝖦(A)PL0(V)\tilde{P}:\mathsf{FG}(A)\to\operatorname{PL}_{0}(V) be the unique homomorphism that restricts to ηVP\eta_{V}\circ P on AA. Let 𝐰𝖥𝖦(A)\mathbf{w}\in\mathsf{FG}(A) be an element, which can be expressed uniquely as a reduced word in AA1A\cup A^{-1}:

𝐰=(a1n1,a2n2,,arnr),\displaystyle\mathbf{w}=(a_{1}^{n_{1}},a_{2}^{n_{2}},...,a_{r}^{n_{r}}), (2.38)

where all nin_{i} are non-zero integers, and aiai+1a_{i}\neq a_{i+1} for 1ir11\leq i\leq r-1. Then

P~(𝐰)=(n1P(a1),n2P(a2),,nrP(ar)).\displaystyle\tilde{P}(\mathbf{w})=(n_{1}P(a_{1}),n_{2}P(a_{2}),...,n_{r}P(a_{r})). (2.39)

By assumption, each vector in the list is non-zero, and consecutive pairs are linearly independent. Hence, P~(𝐰)\tilde{P}(\mathbf{w}) is a minimal representative by Proposition 2.12. This implies that the kernel of P~\tilde{P} 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 𝖠𝖿𝖿\mathsf{Aff} be the category of affine spaces, whose objects are finite dimensional vector spaces, but whose morphisms are affine-linear maps. Given an affine space VV, define action groupoids

Π1,PL(V)=VPL0(V),Π^1(V)=VK^0(V),\displaystyle\Pi_{1,\operatorname{PL}}(V)=V\rtimes\operatorname{PL}_{0}(V),\qquad\widehat{\Pi}_{1}(V)=V\rtimes\hat{K}_{0}(V), (2.40)

where the actions are respectively defined by the homomorphism t:PL0(V)Vt:\operatorname{PL}_{0}(V)\to V and the truncation t:K^0(V)K0(1)(V)Vt:\hat{K}_{0}(V)\to K_{0}^{(\leq 1)}(V)\cong V, via the additive action of VV on itself. Along with the thin fundamental groupoid, these define functors from the category of affine spaces to the category of groupoids

Π1,PL,𝔗1,Π^1:𝖠𝖿𝖿𝖦𝗋𝗉𝖽.\displaystyle\Pi_{1,\operatorname{PL}},\ \ \mathfrak{T}_{1},\ \ \widehat{\Pi}_{1}:\mathsf{Aff}\to\mathsf{Grpd}. (2.41)

Furthermore, the maps R0R_{0}, S0S_{0}, and SPL,0S_{\operatorname{PL},0} give rise to natural transformations

S~PL,0:Π1,PLR~0𝔗1S~0Π^1\displaystyle\tilde{S}_{\operatorname{PL},0}:\Pi_{1,\operatorname{PL}}\xRightarrow{\tilde{R}_{0}}\mathfrak{T}_{1}\xRightarrow{\tilde{S}_{0}}\widehat{\Pi}_{1} (2.42)

which restrict to the identity on the objects of the groupoids.

Proposition 2.19.

There is a unique natural transformation

S~PL,0:Π1,PLΠ^1,\displaystyle\tilde{S}_{\operatorname{PL},0}:\Pi_{1,\operatorname{PL}}\xRightarrow{}\widehat{\Pi}_{1}, (2.43)

called the piecewise linear groupoid signature, which restricts to the identity on objects.

Proof.

It suffices to show that any natural transformation F:Π1,PLΠ^1F:\Pi_{1,\operatorname{PL}}\xRightarrow{}\widehat{\Pi}_{1} which is the identity on objects must be given by S~PL,0\tilde{S}_{\operatorname{PL},0}. First, let UU be a 11-dimensional vector space. Then t:K^0(U)Ut:\hat{K}_{0}(U)\to U is an isomorphism and hence Π^1(U)Pair(U)\widehat{\Pi}_{1}(U)\cong\mathrm{Pair}(U), the terminal object in the category of groupoids over UU. Therefore, we have equality of components FU=(S~PL,0)UF_{U}=(\tilde{S}_{\operatorname{PL},0})_{U}. Now let VV be a general vector space. Elements of Π1,PL(V)\Pi_{1,\operatorname{PL}}(V) can be factored into products of elements of the form (v,ηV(u))(v,\eta_{V}(u)), for v,uVv,u\in V. Hence, it suffices to show that FV(v,ηV(u))=(S~PL,0)V(v,ηV(u))F_{V}(v,\eta_{V}(u))=(\tilde{S}_{\operatorname{PL},0})_{V}(v,\eta_{V}(u)) for all such pairs. Given (v,u)(v,u), define the affine linear map

f(v,u):V,tv+tu.\displaystyle f_{(v,u)}:\mathbb{R}\to V,\qquad t\mapsto v+tu. (2.44)

Then Π1,PL(f(v,u))(0,η(1))=(v,ηV(u))\Pi_{1,\operatorname{PL}}(f_{(v,u)})(0,\eta_{\mathbb{R}}(1))=(v,\eta_{V}(u)), where (0,η(1))Π1,PL()(0,\eta_{\mathbb{R}}(1))\in\Pi_{1,\operatorname{PL}}(\mathbb{R}). Therefore, using the naturality of FF and S~PL,0\tilde{S}_{\operatorname{PL},0}, we obtain

FV(v,ηV(u))\displaystyle F_{V}(v,\eta_{V}(u)) =FVΠ1,PL(f(v,u))(0,η(1))=Π^1(f(v,u))F(0,η(1))\displaystyle=F_{V}\circ\Pi_{1,\operatorname{PL}}(f_{(v,u)})(0,\eta_{\mathbb{R}}(1))=\widehat{\Pi}_{1}(f_{(v,u)})\circ F_{\mathbb{R}}(0,\eta_{\mathbb{R}}(1)) (2.45)
=Π^1(f(v,u))(S~PL,0)(0,η(1))\displaystyle=\widehat{\Pi}_{1}(f_{(v,u)})\circ(\tilde{S}_{\operatorname{PL},0})_{\mathbb{R}}(0,\eta_{\mathbb{R}}(1)) (2.46)
=(S~PL,0)V(v,ηV(u)),\displaystyle=(\tilde{S}_{\operatorname{PL},0})_{V}(v,\eta_{V}(u)), (2.47)

where the equality in the second line uses F=(S~PL,0)F_{\mathbb{R}}=(\tilde{S}_{\operatorname{PL},0})_{\mathbb{R}}. ∎

Remark 2.20.

Piecewise linear paths are dense in C01([0,1],V)C^{1}_{0}([0,1],V), the C1C^{1} paths starting at the origin, equipped with the Lipschitz topology. Therefore, the groupoid signature S~0:𝔗1Π^1\tilde{S}_{0}:\mathfrak{T}_{1}\xRightarrow{}\widehat{\Pi}_{1} 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 𝐗C([0,1]2,V)\mathbf{X}\in C([0,1]^{2},V) has sitting instants if there exists some ϵ>0\epsilon>0 such that

𝐗u,t=𝐗0,t,𝐗1u,t=𝐗1,t,𝐗s,u=𝐗s,0,𝐗s,1u=𝐗s,1for allu[0,ϵ],s,t[0,1].\displaystyle\mathbf{X}_{u,t}=\mathbf{X}_{0,t},\quad\mathbf{X}_{1-u,t}=\mathbf{X}_{1,t},\quad\mathbf{X}_{s,u}=\mathbf{X}_{s,0},\quad\mathbf{X}_{s,1-u}=\mathbf{X}_{s,1}\quad\text{for all}\quad u\in[0,\epsilon],\,s,t\in[0,1]. (3.1)

Unless otherwise specified, we assume all surfaces have sitting instants.

For a surface 𝐗C([0,1]2,V)\mathbf{X}\in C([0,1]^{2},V), the bottom, right, top, and left boundary paths of 𝐗\mathbf{X} are

(b𝐗)s𝐗s,0,(r𝐗)s𝐗1,s,(t𝐗)s𝐗s,1,(l𝐗)s𝐗0,s.\displaystyle(\partial_{b}\mathbf{X})_{s}\coloneqq\mathbf{X}_{s,0},\quad(\partial_{r}\mathbf{X})_{s}\coloneqq\mathbf{X}_{1,s},\quad(\partial_{t}\mathbf{X})_{s}\coloneqq\mathbf{X}_{s,1},\quad(\partial_{l}\mathbf{X})_{s}\coloneqq\mathbf{X}_{0,s}. (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 𝐗\mathbf{X} using the counter-clockwise convention to be

𝐗=(b𝐗r𝐗)((t𝐗)1(l𝐗)1).\displaystyle\partial\mathbf{X}=(\partial_{b}\mathbf{X}\star\partial_{r}\mathbf{X})\star((\partial_{t}\mathbf{X})^{-1}\star(\partial_{l}\mathbf{X})^{-1}). (3.3)
[Uncaptioned image]

The sitting instants ensure that for horizontally and vertically composable smooth surfaces 𝐗,𝐘,𝐙C1([0,1]2,V)\mathbf{X},\mathbf{Y},\mathbf{Z}\in C^{1}([0,1]^{2},V), such that r𝐗=l𝐘\partial_{r}\mathbf{X}=\partial_{l}\mathbf{Y} and t𝐗=b𝐙\partial_{t}\mathbf{X}=\partial_{b}\mathbf{Z}, their horizontal and vertical compositions

(𝐗h𝐘)s,t{𝐗2s,t:s[0,12]𝐘2s1,t:s[12,1]and(𝐗v𝐙)s,t{𝐗s,2t:t[0,12]𝐙s,2t1:t[12,1]\displaystyle(\mathbf{X}\star_{h}\mathbf{Y})_{s,t}\coloneqq\left\{\begin{array}[]{cl}\mathbf{X}_{2s,t}&:s\in[0,\frac{1}{2}]\\ \mathbf{Y}_{2s-1,t}&:s\in[\frac{1}{2},1]\end{array}\right.\quad\text{and}\quad(\mathbf{X}\star_{v}\mathbf{Z})_{s,t}\coloneqq\left\{\begin{array}[]{cl}\mathbf{X}_{s,2t}&:t\in[0,\frac{1}{2}]\\ \mathbf{Z}_{s,2t-1}&:t\in[\frac{1}{2},1]\end{array}\right. (3.8)

are also smooth, so that (𝐗h𝐘),(𝐗v𝐙)C1([0,1]2,V)(\mathbf{X}\star_{h}\mathbf{Y}),(\mathbf{X}\star_{v}\mathbf{Z})\in C^{1}([0,1]^{2},V).

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 𝐗,𝐘C1([0,1]2,V)\mathbf{X},\mathbf{Y}\in C^{1}([0,1]^{2},V) are thin homotopy equivalent, denoted 𝐗th𝐘\mathbf{X}\sim_{\operatorname{th}}\mathbf{Y}, if there exists a corner-preserving smooth homotopy HC1([0,1]3,V)H\in C^{1}([0,1]^{3},V) between 𝐗\mathbf{X} and 𝐘\mathbf{Y} such that

  • (homotopy condition) H0,s,t=𝐗s,tH_{0,s,t}=\mathbf{X}_{s,t} and H1,s,t=𝐘s,tH_{1,s,t}=\mathbf{Y}_{s,t};

  • (thin homotopic boundaries) the four sides of the homotopy Hu,s,0,Hu,s,1,Hu,0,t,Hu,1,tH_{u,s,0},H_{u,s,1},H_{u,0,t},H_{u,1,t} are thin homotopies between the four boundary paths of 𝐗\mathbf{X} and 𝐘\mathbf{Y},

  • (thinness condition) rank(dH)2\operatorname{rank}(dH)\leq 2, where dHdH is the differential of HH.

Remark 3.3.

In this article, we use [0,1]2[0,1]^{2} 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. (1)

    Reparametrization. Given a surface 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V), and a corner and edge preserving reparametrization ϕ:[0,1]2[0,1]2\phi:[0,1]^{2}\to[0,1]^{2}, then 𝐗th𝐗ϕ\mathbf{X}\sim_{\operatorname{th}}\mathbf{X}\circ\phi.

  2. (2)

    Folding. Given a surface 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V), we say that a fold is a region of a surface of the form 𝐗h𝐗h\mathbf{X}\star_{h}\mathbf{X}^{-h} or 𝐗v𝐗v\mathbf{X}\star_{v}\mathbf{X}^{-v}. Two surfaces which differ by folds are thin homotopy equivalent.

[Uncaptioned image]

However, thin homotopy of surfaces may yield non-local cancellations, as the following shows.

Proposition 3.4.

Consider a map 𝐗RPC1([0,1]2,n)\mathbf{X}^{RP}\in C^{1}([0,1]^{2},\mathbb{R}^{n}) which factors as

𝐗RP:[0,1]2𝑞S2𝑌2𝑔n,\displaystyle\mathbf{X}^{RP}:[0,1]^{2}\xrightarrow{q}S^{2}\xrightarrow{Y}\mathbb{RP}^{2}\xrightarrow{g}\mathbb{R}^{n}, (3.9)

where qq is a map which covers the sphere and sends the boundary of [0,1]2[0,1]^{2} to a basepoint S2*\in S^{2}, YY is the double cover obtained by identifying antipodal points, and gg is a smooth map. Then 𝐗RP\mathbf{X}^{RP} is thinly null homotopic.

Proof.

We represent 2=D2/\mathbb{RP}^{2}=D^{2}/\sim as the quotient of a disc D2D^{2} by the antipodal identification on the boundary. We compose the antipodal projection map YY with a map c:2S2c:\mathbb{RP}^{2}\to S^{2} which collapses the boundary circle to a single point, to obtain cY:S2S2c\circ Y:S^{2}\to S^{2} as in Figure 1.

Refer to caption
Figure 1. Collapsing map

The map cYc\circ Y has degree 0 and hence is null-homotopic by the Hurewicz theorem. Therefore, given a map g:S2ng^{\prime}:S^{2}\to\mathbb{R}^{n}, the map gcYqg^{\prime}\circ c\circ Y\circ q 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 𝐗RP\mathbf{X}^{RP} is thinly null-homotopic, it suffices to construct a thin homotopy between gg and gcg^{\prime}\circ c, for some map gg^{\prime}. To this end, let p:[0,1]D2p:[0,1]\to D^{2} be the map parametrizing the top boundary of the disc, and let γ:[0,1]2\gamma:[0,1]\to\mathbb{RP}^{2} be the corresponding loop in 2\mathbb{RP}^{2}, which generates the fundamental group π1(2)/2\pi_{1}(\mathbb{RP}^{2})\cong\mathbb{Z}/2. The loop gγ:[0,1]ng\circ\gamma:[0,1]\to\mathbb{R}^{n} is contractible. Let h:[0,1]2nh:[0,1]^{2}\to\mathbb{R}^{n} be a smooth homotopy that contracts γ\gamma to a point.

It is possible to modify gg on a collar neighbourhood of the boundary of D2D^{2}, so that its restriction to both the top half and to the bottom half of the disc agrees with hh. This can be pictured as in Figure 2

Refer to caption
Figure 2. The map g1g_{1}

Let g1:2ng_{1}:\mathbb{RP}^{2}\to\mathbb{R}^{n} denote the new map. It is obtained by adding a surface, namely the image of hh, to the original map gg. Since it sends the boundary of the disc D2D^{2} to a single point x0nx_{0}\in\mathbb{R}^{n} it factors as g1=gcg_{1}=g^{\prime}\circ c. Furthermore, the new map g1g_{1} is homotopic to gg through a homotopy which is restricted to lie within the image of g1g_{1}. Hence gg and g1g_{1} 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 gg 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 XRP:[0,1]2nX^{RP}:[0,1]^{2}\to\mathbb{R}^{n} be defined as in Proposition 3.4, with gg an embedding. There does not exist a null-homotopy H:[0,1]3nH:[0,1]^{3}\to\mathbb{R}^{n} of XRPX^{RP} such that im(H)im(XRP)\operatorname{im}(H)\subseteq\operatorname{im}(X^{RP}).

Proof.

Assume that such a homotopy HH exists. Since it is contained in the embedded surface im(XRP)\operatorname{im}(X^{RP}), it lifts to a null homotopy H^:[0,1]32\hat{H}:[0,1]^{3}\to\mathbb{RP}^{2} of fq:[0,1]22f\circ q:[0,1]^{2}\to\mathbb{RP}^{2}. This is a contradiction since fqf\circ q is a generator for π2(2)=\pi_{2}(\mathbb{RP}^{2})=\mathbb{Z}. ∎

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 𝔗(V)\mathfrak{T}(V) is an edge symmetric double groupoid where

𝔗0(V)V,𝔗1(V)C1([0,1],V)/th,and𝔗2(V)C1([0,1]2,V)/th.\displaystyle\mathfrak{T}_{0}(V)\coloneqq V,\quad\mathfrak{T}_{1}(V)\coloneqq C^{1}([0,1],V)/\sim_{\operatorname{th}},\quad\text{and}\quad\mathfrak{T}_{2}(V)\coloneqq C^{1}([0,1]^{2},V)/\sim_{\operatorname{th}}. (3.10)

The thin double group τ(V)\tau^{\square}(V) is an edge symmetric double group where

τ0(V)={},τ1(V)𝔗1(V)/trans,andτ2(V)𝔗2(V)/trans.\displaystyle\tau^{\square}_{0}(V)=\{*\},\quad\tau^{\square}_{1}(V)\coloneqq\mathfrak{T}_{1}(V)/\sim_{\operatorname{trans}},\quad\text{and}\quad\tau_{2}^{\square}(V)\coloneqq\mathfrak{T}_{2}(V)/\sim_{\operatorname{trans}}. (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 [𝐗],[𝐘]τ2(V)[\mathbf{X}],[\mathbf{Y}]\in\tau^{\square}_{2}(V) with representatives 𝐗,𝐘C1([0,1]2,V)\mathbf{X},\mathbf{Y}\in C^{1}([0,1]^{2},V), where 𝐗0,0=0\mathbf{X}_{0,0}=0 and 𝐘0,0=𝐗1,0\mathbf{Y}_{0,0}=\mathbf{X}_{1,0}, are horizontally composable if r𝐗thl𝐘\partial_{r}\mathbf{X}\sim_{\operatorname{th}}\partial_{l}\mathbf{Y}. Let hC1([0,1]2,V)h\in C^{1}([0,1]^{2},V) be a thin homotopy between r𝐗\partial_{r}\mathbf{X} and l𝐘\partial_{l}\mathbf{Y}. We define the horizontal composition of [𝐗][\mathbf{X}] and [𝐘][\mathbf{Y}] to be

[𝐗]h[𝐘][𝐗hhh𝐘].\displaystyle[\mathbf{X}]\star_{h}[\mathbf{Y}]\coloneqq[\mathbf{X}\star_{h}h\star_{h}\mathbf{Y}]. (3.12)

Next, we define the horizontal inverse of a surface in τ2(V)\tau^{\square}_{2}(V) by

𝐗s,t1𝐗1s,t𝐗1,0and[𝐗]1[𝐗1].\displaystyle\mathbf{X}^{-1}_{s,t}\coloneqq\mathbf{X}_{1-s,t}-\mathbf{X}_{1,0}\quad\text{and}\quad[\mathbf{X}]^{-1}\coloneqq[\mathbf{X}^{-1}]. (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,

𝐆=(δ:G1G0,:G0Aut(G1))\displaystyle\mathbf{G}=\left(\delta:G_{1}\rightarrow G_{0},\ \vartriangleright:G_{0}\rightarrow\operatorname{Aut}(G_{1})\right) (3.14)

is given by two groups (G0,),(G1,)(G_{0},\cdot),(G_{1},*), a group morphism δ:G1G0\delta:G_{1}\rightarrow G_{0} and a left action of G0G_{0} on G1G_{1} by group automorphisms, denoted elementwise by g:G1G1g\vartriangleright\cdot:G_{1}\rightarrow G_{1} for gGg\in G. These data are required to satisfy

δ(gE)=gδ(E)g1forgG0andEG1.\displaystyle\delta(g\vartriangleright E)=g\cdot\delta(E)\cdot g^{-1}\quad\text{for}\quad g\in G_{0}\quad\text{and}\quad E\in G_{1}. (3.15)

We say 𝐆\mathbf{G} is a crossed module of groups if it also satisfies the Peiffer identity

δ(E1)(E2)=E1E2E11forE1,E2G1.\displaystyle\delta(E_{1})\vartriangleright(E_{2})=E_{1}*E_{2}*E_{1}^{-1}\quad\text{for}\quad E_{1},E_{2}\in G_{1}. (3.16)

A (pre-)crossed module of Lie groups is the same as above, except G0G_{0} and G1G_{1} are Lie groups, and all morphisms are smooth. Given another (pre-)crossed module 𝐇=(δ:H1H0,)\mathbf{H}=(\delta:H_{1}\to H_{0},\vartriangleright), a morphism of (pre-)crossed modules f=(f1,f0):𝐆𝐇f=(f_{1},f_{0}):\mathbf{G}\to\mathbf{H} consists of group homomorphisms f0:G0H0f_{0}:G_{0}\to H_{0} and f1:G1H1f_{1}:G_{1}\to H_{1} such that, for all gG0g\in G_{0} and EG1E\in G_{1}, we have

δf1(E)=f0δ(E)andf1(gE)=f0(g)f1(E).\displaystyle\delta\circ f_{1}(E)=f_{0}\circ\delta(E)\quad\text{and}\quad f_{1}(g\vartriangleright E)=f_{0}(g)\vartriangleright f_{1}(E).

The categories of crossed modules of groups and Lie groups are respectively denoted 𝖷𝖦𝗋𝗉\mathsf{XGrp} and 𝖷𝖫𝖦𝗋𝗉\mathsf{XLGrp}.

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 𝖣𝖦𝗋𝗉\mathsf{DGrp}.

Theorem 3.8.

[8, Section 6.6] There is an equivalence of categories between 𝖣𝖦𝗋𝗉\mathsf{DGrp} and 𝖷𝖦𝗋𝗉\mathsf{XGrp}.

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 VV is the crossed module associated to the thin double group τ(V)\tau^{\square}(V). In particular, let

τ1(V)τ1(V)andτ2(V){𝐗τ2(V):𝐗s,0=𝐗0,t=𝐗1,tth0}.\displaystyle\tau_{1}(V)\coloneqq\tau^{\square}_{1}(V)\quad\text{and}\quad\tau_{2}(V)\coloneqq\left\{\mathbf{X}\in\tau^{\square}_{2}(V)\,:\,\mathbf{X}_{s,0}=\mathbf{X}_{0,t}=\mathbf{X}_{1,t}\sim_{\operatorname{th}}0\right\}. (3.17)

The group operation \star in τ2(V)\tau_{2}(V) is defined as the reversed444The group operation is defined via the equivalence between 𝖣𝖦𝗋𝗉\mathsf{DGrp} and 𝖷𝖦𝗋𝗉\mathsf{XGrp} in [8, Section 6.6], and is determined by the convention (starting point and orientation) used for the boundary \partial 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),

𝐗𝐘𝐘h𝐗\displaystyle\mathbf{X}\star\mathbf{Y}\coloneqq\mathbf{Y}\star_{h}\mathbf{X} (3.18)

and the inverse is given by the horizontal inverse (3.13). The crossed module boundary map δ:τ2(V)τ1(V)\delta:\tau_{2}(V)\to\tau_{1}(V) is given by the boundary \partial of the surface from (3.3)

δ(𝐗)(𝐗)=t(𝐗)1.\displaystyle\delta(\mathbf{X})\coloneqq\partial(\mathbf{X})=\partial_{t}(\mathbf{X})^{-1}. (3.19)

For a path 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) with 𝐱0=0\mathbf{x}_{0}=0, we define the degenerate surface σ𝐱:[0,1]2V\sigma^{\mathbf{x}}:[0,1]^{2}\to V by

σs,t𝐱𝐱s.\displaystyle\sigma^{\mathbf{x}}_{s,t}\coloneqq\mathbf{x}_{s}. (3.20)

Finally, we define a left action of τ1(V)\tau_{1}(V) on τ2(V)\tau_{2}(V) by

𝐱𝐗σ𝐱h(𝐱1+𝐗)h(𝐱1+σ𝐱1),\displaystyle\mathbf{x}\vartriangleright\mathbf{X}\coloneqq\sigma^{\mathbf{x}}\star_{h}(\mathbf{x}_{1}+\mathbf{X})\star_{h}(\mathbf{x}_{1}+\sigma^{\mathbf{x}^{-1}}), (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 ϕ:VW\phi:V\to W preserve thin homotopy equivalence of maps, and thus we obtain a functor

𝝉:𝖵𝖾𝖼𝗍𝖷𝖦𝗋𝗉.\displaystyle{\boldsymbol{\tau}}:\mathsf{Vect}\to\mathsf{XGrp}. (3.22)
Definition 3.10.

[8, Section 2.2] Let (C,c0)(C,c_{0}) be a based 2-dimensional CW complex. The fundamental crossed module of (C,c0)(C,c_{0}) is given by

𝝅(C,C1)(:π2(C,C1),π1(C1),),\displaystyle\boldsymbol{\pi}(C,C_{1})\coloneqq\left(\partial:\pi_{2}(C,C_{1}),\to\pi_{1}(C_{1}),\vartriangleright\right), (3.23)

where π1\pi_{1} is the fundamental group of the 11-skeleton C1C_{1}, and π2(C,C1)\pi_{2}(C,C_{1}) is the relative homotopy group of CC with respect to the 1-skeleton C1C_{1}. The later is defined to be the group of homotopy classes of maps 𝐗:[0,1]2C\mathbf{X}:[0,1]^{2}\to C such that the boundary of [0,1]2[0,1]^{2} is sent to C1C_{1}, the corners are sent to c0c_{0}, and such that the left, bottom and right boundary paths are null homotopic in C1C_{1}, following the convention of (3.17). The multiplication \star in π2(C,C1)\pi_{2}(C,C_{1}) is given by reversed horizontal concatenation 𝐗𝐘=𝐘h𝐗\mathbf{X}\star\mathbf{Y}=\mathbf{Y}\star_{h}\mathbf{X}, as in (3.18). This is equipped with the boundary map :π2(C,C1)π1(C1)\partial:\pi_{2}(C,C_{1})\to\pi_{1}(C_{1}) from (3.3) and an action of π1(C1)\pi_{1}(C_{1}) on π2(C,C1)\pi_{2}(C,C_{1}) given by

𝐱𝐗σ𝐱h𝐗hσ𝐱1,\displaystyle\mathbf{x}\vartriangleright\mathbf{X}\coloneqq\sigma^{\mathbf{x}}\star_{h}\mathbf{X}\star_{h}\sigma^{\mathbf{x}^{-1}}, (3.24)

where σ𝐱\sigma^{\mathbf{x}} 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

𝖌=(δ:𝔤1𝔤0,)\displaystyle{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright)

consists of Lie algebras (𝔤0,[,]0)(\mathfrak{g}_{0},[\cdot,\cdot]_{0}) and (𝔤1,[,]1)(\mathfrak{g}_{1},[\cdot,\cdot]_{1}), a morphism δ:𝔤1𝔤0\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0}, and an action \vartriangleright of 𝔤0\mathfrak{g}_{0} on 𝔤1\mathfrak{g}_{1} by derivations. In other words, the action satisfies

x[E,F]1=[xE,F]1+[E,xF]1and[x,y]0E=x(yE)y(xE)\displaystyle x\vartriangleright[E,F]_{1}=[x\vartriangleright E,F]_{1}+[E,x\vartriangleright F]_{1}\quad\text{and}\quad[x,y]_{0}\vartriangleright E=x\vartriangleright(y\vartriangleright E)-y\vartriangleright(x\vartriangleright E)

for all x,y𝔤0x,y\in\mathfrak{g}_{0} and E,F𝔤1E,F\in\mathfrak{g}_{1}. Furthermore, these data are required to satisfy

δ(xE)=[x,δ(E)]0for all x𝔤0andE𝔤1.\displaystyle\delta(x\vartriangleright E)=[x,\delta(E)]_{0}\quad\text{for all }\quad x\in\mathfrak{g}_{0}\quad\text{and}\quad E\in\mathfrak{g}_{1}. (3.25)

We say 𝖌{\boldsymbol{\mathfrak{g}}} is a crossed module of Lie algebras if it also satisfies the Peiffer identity

δ(E)E=[E,E]1for allE,E𝔤1.\displaystyle\delta(E)\vartriangleright E^{\prime}=[E,E^{\prime}]_{1}\quad\text{for all}\quad E,E^{\prime}\in\mathfrak{g}_{1}. (3.26)

Suppose 𝖍=(δ:𝔥1𝔥0,){\boldsymbol{\mathfrak{h}}}=(\delta:\mathfrak{h}_{1}\to\mathfrak{h}_{0},\vartriangleright) is another (pre-)crossed module of Lie algebras. A morphism of (pre-)crossed modules f=(f0,f1):𝖌𝖍f=(f_{0},f_{1}):{\boldsymbol{\mathfrak{g}}}\to{\boldsymbol{\mathfrak{h}}} consists of two Lie algebra morphisms f0:𝔤0𝔥0f_{0}:\mathfrak{g}_{0}\to\mathfrak{h}_{0} and f1:𝔤1𝔥1f_{1}:\mathfrak{g}_{1}\to\mathfrak{h}_{1} such that for all x𝔤0x\in\mathfrak{g}_{0} and E𝔤1E\in\mathfrak{g}_{1},

δf1(E)=f0δ(E)andf1(xE)=f0(x)f1(E).\displaystyle\delta\circ f_{1}(E)=f_{0}\circ\delta(E)\quad\text{and}\quad f_{1}(x\vartriangleright E)=f_{0}(x)\vartriangleright f_{1}(E). (3.27)

The category of crossed modules of Lie algebras is denoted 𝖷𝖫𝗂𝖾\mathsf{XLie}.

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 𝖌=(δ:𝔤1𝔤0){\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0}) be a crossed module of Lie algebras. A 2-connection valued in 𝖌{\boldsymbol{\mathfrak{g}}} over VV is a pair (γ0,γ1)(\gamma_{0},\gamma_{1}), such that γ0Ω1(V,𝔤0)\gamma_{0}\in\Omega^{1}(V,\mathfrak{g}_{0}) is a 𝔤0\mathfrak{g}_{0}-valued 11-form and γ1Ω2(V,𝔤1)\gamma_{1}\in\Omega^{2}(V,\mathfrak{g}_{1}) is a 𝔤1\mathfrak{g}_{1}-valued 22-form. The 1-curvature κγ0,γ1Ω2(V,𝔤0)\kappa^{\gamma_{0},\gamma_{1}}\in\Omega^{2}(V,\mathfrak{g}_{0}) and 2-curvature 𝒦γ0,γ1Ω3(V,𝔤1)\mathcal{K}^{\gamma_{0},\gamma_{1}}\in\Omega^{3}(V,\mathfrak{g}_{1}) are respectively defined as follows

κγ0,γ1=κγ0δγ1and𝒦γ0,γ1dγ1+γ0γ1,\displaystyle\kappa^{\gamma_{0},\gamma_{1}}=\kappa^{\gamma_{0}}-\delta\gamma_{1}\quad\text{and}\quad\mathcal{K}^{\gamma_{0},\gamma_{1}}\coloneqq d\gamma_{1}+\gamma_{0}\wedge^{\vartriangleright}\gamma_{1}, (3.28)

where κγ0\kappa^{\gamma_{0}} is the curvature of γ0\gamma_{0} from (2.6) and γ0γ1Ω3(V,𝔤1)\gamma_{0}\wedge^{\vartriangleright}\gamma_{1}\in\Omega^{3}(V,\mathfrak{g}_{1}) is defined by

(γ0γ1)(X,Y,Z)γ0(X)γ1(Y,Z)γ0(Y)γ1(X,Z)+γ0(Z)γ1(X,Y).\displaystyle(\gamma_{0}\wedge^{\vartriangleright}\gamma_{1})(X,Y,Z)\coloneqq\gamma_{0}(X)\vartriangleright\gamma_{1}(Y,Z)-\gamma_{0}(Y)\vartriangleright\gamma_{1}(X,Z)+\gamma_{0}(Z)\vartriangleright\gamma_{1}(X,Y). (3.29)

The 2-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) is called fake-flat (or semi-flat) if its 1-curvature vanishes, κγ0,γ1=0\kappa^{\gamma_{0},\gamma_{1}}=0. We will always work with fake-flat 2-connections.

We say that a 2-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) is translation-invariant if it has the form

γ0=i=1dγ0idziandγ1=i<jγ1i,jdzidzj,\displaystyle\gamma_{0}=\sum_{i=1}^{d}\gamma_{0}^{i}dz_{i}\quad\text{and}\quad\gamma_{1}=\sum_{i<j}\gamma_{1}^{i,j}dz_{i}\wedge dz_{j}, (3.30)

where γ0i𝔤0\gamma_{0}^{i}\in\mathfrak{g}_{0}, γ1i,j𝔤1\gamma_{1}^{i,j}\in\mathfrak{g}_{1}, and ziz_{i} are linear coordinates on VV. In this case, we can represent the 2-connection as a pair of linear maps

γ0L(V,𝔤0)andγ1L(Λ2V,𝔤1).\displaystyle\gamma_{0}\in L(V,\mathfrak{g}_{0})\quad\text{and}\quad\gamma_{1}\in L(\Lambda^{2}V,\mathfrak{g}_{1}). (3.31)

The curvature forms then simplify to

κγ0,γ1=12[γ0,γ0]δγ1and𝒦γ0,γ1=γ0γ1.\displaystyle\kappa^{\gamma_{0},\gamma_{1}}=\frac{1}{2}[\gamma_{0},\gamma_{0}]-\delta\gamma_{1}\quad\text{and}\quad\mathcal{K}^{\gamma_{0},\gamma_{1}}=\gamma_{0}\wedge^{\vartriangleright}\gamma_{1}. (3.32)

3.4. Surface Holonomy

Here, we discuss surface holonomy, the 2-dimensional generalization of parallel transport. For a surface 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V), the (s,t)(s,t)-tail path of 𝐗\mathbf{X} is

𝐱us,t{𝐗0,2ut:u[0,1/2]𝐗(2u1)s,t:u(1/2,1]\displaystyle\mathbf{x}^{s,t}_{u}\coloneqq\left\{\begin{array}[]{cl}\mathbf{X}_{0,2ut}&:u\in[0,1/2]\\ \mathbf{X}_{(2u-1)s,t}&:u\in(1/2,1]\end{array}\right. (3.35)
[Uncaptioned image]
Definition 3.13.

[44, Equation 2.13] Let 𝐆=(δ:G1G0,)\mathbf{G}=(\delta:G_{1}\to G_{0},\vartriangleright) be a crossed module of Lie groups with associated crossed module of Lie algebras 𝖌=(δ:𝔤1𝔤0,){\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright). Let (γ0,γ1)(\gamma_{0},\gamma_{1}) be a 2-connection valued in 𝖌{\boldsymbol{\mathfrak{g}}} and let 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V). Consider the following differential equation for F1γ0,γ1(𝐗):[0,1]2G1F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X}):[0,1]^{2}\to G_{1}

F1γ0,γ1(𝐗)s,tt=dLF1γ0,γ1(𝐗)s,t0sF0γ0(𝐱s,t)γ1(𝐗s,ts,𝐗s,tt)𝑑s,F1γ0,γ1(𝐗)s,0=eG1,\displaystyle\frac{\partial F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})_{s,t}}{\partial t}=dL_{F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})_{s,t}}\int_{0}^{s}F_{0}^{\gamma_{0}}(\mathbf{x}^{s^{\prime},t})\vartriangleright\gamma_{1}\left(\frac{\partial\mathbf{X}_{s^{\prime},t}}{\partial s},\frac{\partial\mathbf{X}_{s^{\prime},t}}{\partial t}\right)ds^{\prime},\quad F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})_{s,0}=e_{G_{1}}, (3.36)

where the tail path 𝐱s,t:[0,1]V\mathbf{x}^{s,t}:[0,1]\to V is defined in (3.35) and F0γ0(𝐱s,t)F_{0}^{\gamma_{0}}(\mathbf{x}^{s^{\prime},t}) is the parallel transport of γ0\gamma_{0} along 𝐱s,t\mathbf{x}^{s^{\prime},t}, as given in Definition 2.3. We define the surface holonomy of (γ0,γ1)(\gamma_{0},\gamma_{1}) along 𝐗\mathbf{X} to be

F1γ0,γ1(𝐗)F1γ0,γ1(𝐗)1,1.\displaystyle F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})\coloneqq F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})_{1,1}. (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 𝝉(V){\boldsymbol{\tau}}(V).

Theorem 3.14.

[44, Theorem 2.32] Let (γ0,γ1)(\gamma_{0},\gamma_{1}) be a translation-invariant 2-connection valued in 𝖌{\boldsymbol{\mathfrak{g}}}. Then the maps

Fγ0,γ1=(F1γ0,γ1,F0γ0):𝝉(V)𝐆\displaystyle F^{\gamma_{0},\gamma_{1}}=(F_{1}^{\gamma_{0},\gamma_{1}},F_{0}^{\gamma_{0}}):{\boldsymbol{\tau}}(V)\to\mathbf{G} (3.38)

define a morphism of crossed modules.

This theorem has two main implications.

  1. (1)

    Surface holonomy is invariant with respect to thin homotopies of surfaces.

  2. (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 22-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 𝖵𝖫\mathsf{VL} denote the comma category (id𝖥𝗈𝗋)(\operatorname{id}\downarrow\mathsf{For}) associated to the functors id:𝖵𝖾𝖼𝗍𝖵𝖾𝖼𝗍\operatorname{id}:\mathsf{Vect}\to\mathsf{Vect} and 𝖥𝗈𝗋:𝖫𝗂𝖾𝖵𝖾𝖼𝗍\mathsf{For}:\mathsf{Lie}\to\mathsf{Vect}. An object of 𝖵𝖫\mathsf{VL} is given by the data of a vector space VV, a Lie algebra 𝔤\mathfrak{g}, and a linear map s:V𝔤s:V\to\mathfrak{g}. A morphism f=(f1,f0):(s:V𝔤)(t:W𝔥)f=(f_{1},f_{0}):(s:V\to\mathfrak{g})\to(t:W\to\mathfrak{h}) consists of a linear map f1:VWf_{1}:V\to W and a Lie algebra morphism f0:𝔤𝔥f_{0}:\mathfrak{g}\to\mathfrak{h} such that tf1=f0st\circ f_{1}=f_{0}\circ s as linear maps. There exists a natural forgetful functor

𝖥𝗈𝗋:𝖷𝖫𝗂𝖾𝖵𝖫.\displaystyle\mathsf{For}:\mathsf{XLie}\to\mathsf{VL}. (3.39)

The free crossed module functor

𝖥𝗋:𝖵𝖫𝖷𝖫𝗂𝖾\displaystyle\mathsf{Fr}:\mathsf{VL}\to\mathsf{XLie} (3.40)

is defined to be the left adjoint. In the following, we describe this functor in detail. Given an object s:V𝔤s:V\to\mathfrak{g} of 𝖵𝖫\mathsf{VL}, we first build the free 𝔤\mathfrak{g}-representation on VV, which is U(𝔤)VU(\mathfrak{g})\otimes V, where U(𝔤)U(\mathfrak{g}) is the universal enveloping algebra of 𝔤\mathfrak{g}, where the 𝔤\mathfrak{g} action is given by

x(x1xk)v=(xx1xk)v,\displaystyle x\vartriangleright(x_{1}\otimes\ldots\otimes x_{k})\otimes v=(x\otimes x_{1}\otimes\ldots\otimes x_{k})\otimes v, (3.41)

and where the linear map ss is lifted to a morphism of 𝔤\mathfrak{g}-representations as follows

δ:U(𝔤)V𝔤,(x1xk)v[x1,[xk,s(v)]].\displaystyle\delta:U(\mathfrak{g})\otimes V\to\mathfrak{g},\quad(x_{1}\otimes\ldots\otimes x_{k})\otimes v\mapsto[x_{1},\ldots[x_{k},s(v)]\ldots]. (3.42)

We define the Peiffer subspace of U(𝔤)VU(\mathfrak{g})\otimes V to be

Pf(U(𝔤)V)span{δ(X)Y+δ(Y)X:X,YU(𝔤)V}.\displaystyle{\operatorname{Pf}}(U(\mathfrak{g})\otimes V)\coloneqq\operatorname{span}\left\{\delta(X)\vartriangleright Y+\delta(Y)\vartriangleright X\,:\,X,Y\in U(\mathfrak{g})\otimes V\right\}. (3.43)

By taking the quotient with respect to this subspace, we obtain the free crossed module of Lie algebras generated by ss,

𝖥𝗋(s)(δ:(U(𝔤)V)/Pf(U(𝔤)V)𝔤,),\displaystyle\mathsf{Fr}(s)\coloneqq\left(\delta:(U(\mathfrak{g})\otimes V)/{\operatorname{Pf}}(U(\mathfrak{g})\otimes V)\to\mathfrak{g},\vartriangleright\right), (3.44)

where the Lie bracket of X,Y(U(𝔤)V)/Pf(U(𝔤)V)X,Y\in(U(\mathfrak{g})\otimes V)/{\operatorname{Pf}}(U(\mathfrak{g})\otimes V) is defined by

[X,Y]=δ(X)Y=δ(Y)X.\displaystyle[X,Y]=\delta(X)\vartriangleright Y=-\delta(Y)\vartriangleright X. (3.45)

This satisfies the following universal property, which is proved in Theorem C.8.

Theorem 3.15.

Let (s:V𝔥)𝖵𝖫(s:V\to\mathfrak{h})\in\mathsf{VL}, 𝖌=(δ:𝔤1𝔤0,)𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright)\in\mathsf{XLie}, and

f=(f1,f0):(s:V𝔥)𝖥𝗈𝗋(δ:𝔤1𝔤0,).\displaystyle f=(f_{1},f_{0}):(s:V\to\mathfrak{h})\to\mathsf{For}(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright). (3.46)

Then there is a unique morphism of crossed modules

F=(F1,F0):𝖥𝗋(s)𝖌,\displaystyle F=(F_{1},F_{0}):\mathsf{Fr}(s)\to{\boldsymbol{\mathfrak{g}}}, (3.47)

such that F1ηs=f1F_{1}\circ\eta_{s}=f_{1} and F0=f0F_{0}=f_{0}.

Now consider the functor

:𝖵𝖾𝖼𝗍𝖵𝖫,\displaystyle\wedge:\mathsf{Vect}\to\mathsf{VL}, (3.48)

which sends a vector space V𝖵𝖾𝖼𝗍V\in\mathsf{Vect} to

sV:Λ2V𝖥𝖫(V),uv[u,v].\displaystyle s_{V}:\Lambda^{2}V\to\mathsf{FL}(V),\quad u\wedge v\mapsto[u,v]. (3.49)

Composing with the free crossed module functor gives

𝖐=𝖥𝗋:𝖵𝖾𝖼𝗍𝖷𝖫𝗂𝖾.\displaystyle{\boldsymbol{\mathfrak{k}}}=\mathsf{Fr}\circ\wedge:\mathsf{Vect}\to\mathsf{XLie}. (3.50)

Given a vector space VV, the free crossed module of Lie algebras generated by VV is the result of applying this functor, and is given by

𝖐(V)(δ:𝔨1(V)𝔨0(V),),\displaystyle{\boldsymbol{\mathfrak{k}}}(V)\coloneqq\left(\delta:\mathfrak{k}_{1}(V)\to\mathfrak{k}_{0}(V),\vartriangleright\right), (3.51)

where

𝔨1(V)=(T(V)Λ2V)/Pf(T(V)Λ2V).\displaystyle\mathfrak{k}_{1}(V)=(T(V)\otimes\Lambda^{2}V)/{\operatorname{Pf}}(T(V)\otimes\Lambda^{2}V). (3.52)

3.6. The Surface Signature

Following Kapranov [34], we consider the universal translation-invariant 2-connection. This is a 22-connection (ζ0,ζ1)(\zeta_{0},\zeta_{1}) over VV valued in 𝖐(V){\boldsymbol{\mathfrak{k}}}(V). The translation invariant 𝔨0(V)\mathfrak{k}_{0}(V)-valued 1-form ζ0Ω1(V,𝔨0(V))\zeta_{0}\in\Omega^{1}(V,\mathfrak{k}_{0}(V)) is the one defined in (2.10), and hence corresponds to the identity endomorphism of VV, ζ0=idV\zeta_{0}=\mathrm{id}_{V}. Similarly, the translation invariant 𝔨1(V)\mathfrak{k}_{1}(V)-valued 2-form ζ1\zeta_{1} can be understood as the identity endomorphism of Λ2V\Lambda^{2}V

ζ1=idΛ2VΛ2VΛ2VΩ2(V,𝔨1(V)).\displaystyle\zeta_{1}=\mathrm{id}_{\Lambda^{2}V}\in\Lambda^{2}V^{*}\otimes\Lambda^{2}V\subset\Omega^{2}(V,\mathfrak{k}_{1}(V)). (3.53)

Alternatively, ζ0\zeta_{0} and ζ1\zeta_{1} can be viewed as the linear maps given by including generators

ζ0L(V,𝔨0(V)),ζ1L(Λ2V,𝔨1(V)).\displaystyle\zeta_{0}\in L(V,\mathfrak{k}_{0}(V)),\qquad\zeta_{1}\in L(\Lambda^{2}V,\mathfrak{k}_{1}(V)). (3.54)

In coordinates, they are given by

ζ0=i=1ndzieiΩ1(V,𝔨0)andζ1=i<jdzidzj(eiej)Ω2(V,𝔨1).\displaystyle\zeta_{0}=\sum_{i=1}^{n}dz_{i}e_{i}\in\Omega^{1}(V,\mathfrak{k}_{0})\quad\text{and}\quad\zeta_{1}=\sum_{i<j}dz_{i}\wedge dz_{j}\,(e_{i}\wedge e_{j})\in\Omega^{2}(V,\mathfrak{k}_{1}). (3.55)

We follow a similar procedure as for path signatures and consider truncations of 𝔨1\mathfrak{k}_{1} via the 𝔨0\mathfrak{k}_{0}-lower central series of 𝔨1\mathfrak{k}_{1}, defined by

LCS1(𝔨0,𝔨1)=𝔨1andLCSr(𝔨0,𝔨1)=𝔨0LCSr1(𝔨0,𝔨1).\displaystyle{\operatorname{LCS}}_{1}(\mathfrak{k}_{0},\mathfrak{k}_{1})=\mathfrak{k}_{1}\quad\text{and}\quad{\operatorname{LCS}}_{r}(\mathfrak{k}_{0},\mathfrak{k}_{1})=\mathfrak{k}_{0}\vartriangleright{\operatorname{LCS}}_{r-1}(\mathfrak{k}_{0},\mathfrak{k}_{1}). (3.56)

This allows us to define the nn-truncated free crossed module as follows

𝖐(n)(δ:𝔨1(n)𝔨0(n),) where 𝔨1(n)𝔨1/LCSn+1(𝔨0,𝔨1).\displaystyle{\boldsymbol{\mathfrak{k}}}^{(n)}\coloneqq(\delta:\mathfrak{k}_{1}^{(n)}\to\mathfrak{k}_{0}^{(n)},\vartriangleright)\ \ \text{ where }\ \ \mathfrak{k}_{1}^{(n)}\coloneqq\mathfrak{k}_{1}/{\operatorname{LCS}}_{n+1}(\mathfrak{k}_{0},\mathfrak{k}_{1}). (3.57)

We can integrate 𝔨1(n)\mathfrak{k}_{1}^{(n)} as the Lie group of formal exponentials K1(n)K_{1}^{(n)}, and consider the projective limit K^1\hat{K}_{1}, where

K1(n){exp(x):x𝔨1(n)}andK^1limK1(n).\displaystyle K_{1}^{(n)}\coloneqq\{\exp(x)\,:\,x\in\mathfrak{k}_{1}^{(n)}\}\quad\text{and}\quad\hat{K}_{1}\coloneqq\lim_{\longleftarrow}K_{1}^{(n)}. (3.58)

This defines a crossed module of groups

𝐊^(V)=(δ:K^1(V)K^0(V),),\displaystyle\hat{\mathbf{K}}(V)=(\delta:\hat{K}_{1}(V)\to\hat{K}_{0}(V),\vartriangleright), (3.59)

with a corresponding crossed module of Lie algebras

𝖐^(V)=(δ:𝔨^1(V)𝔨^0(V),),𝔨^1(V)lim𝔨1(n)(V).\displaystyle\hat{{\boldsymbol{\mathfrak{k}}}}(V)=(\delta:\hat{\mathfrak{k}}_{1}(V)\to\hat{\mathfrak{k}}_{0}(V),\vartriangleright),\quad\hat{\mathfrak{k}}_{1}(V)\coloneqq\lim_{\longleftarrow}\mathfrak{k}_{1}^{(n)}(V). (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 K^1(V)\hat{K}_{1}(V) in a way that is similar to the construction of K^0(V)\hat{K}_{0}(V) 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 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V). For nn\in\mathbb{N}, the nn-truncated surface signature of 𝐗\mathbf{X}, denoted S1(n)(𝐗)S_{1}^{(n)}(\mathbf{X}), is defined to be the surface holonomy, as defined in Definition 3.13, of the 2-connection obtained by projecting the universal 2-connection (ζ0,ζ1)(\zeta_{0},\zeta_{1}) from (3.55) onto 𝖐(n)(V){\boldsymbol{\mathfrak{k}}}^{(n)}(V). The surface signature of 𝐗\mathbf{X} is defined to be the projective limit

S1(𝐗)limS1(n)(𝐗)K^1(V).\displaystyle S_{1}(\mathbf{X})\coloneqq\lim_{\longleftarrow}S_{1}^{(n)}(\mathbf{X})\in\hat{K}_{1}(V). (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

𝐒=(S1,S0):𝝉𝐊^.\displaystyle\mathbf{S}=(S_{1},S_{0}):{\boldsymbol{\tau}}\Rightarrow\hat{\mathbf{K}}. (3.62)

We note that the nn-truncated group K1(n)K_{1}^{(n)} is equipped with a natural topology induced by the Lie algebra 𝔨1(n)\mathfrak{k}_{1}^{(n)} 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 C01([0,1]2,V)C^{1}_{0}([0,1]^{2},V) be the space of based smooth surfaces 𝐗\mathbf{X} such that 𝐗0,0=0\mathbf{X}_{0,0}=0. The nn-truncated surface signature

S1(n):C01([0,1]2,V)K1(n)(V),\displaystyle S_{1}^{(n)}:C^{1}_{0}([0,1]^{2},V)\to K_{1}^{(n)}(V), (3.63)

where C01([0,1]2,V)C^{1}_{0}([0,1]^{2},V) 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 𝔨^1(V)\hat{\mathfrak{k}}_{1}(V). Using the representation theory of GL(V){\operatorname{GL}}(V), we produce an explicit expression for the 22-curvature of this abelianized 22-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 𝐆𝖷𝖫𝖦𝗋𝗉\mathbf{G}\in\mathsf{XLGrp} and 𝖌𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}\in\mathsf{XLie} be its associated crossed module of Lie algebras. A gauge transformation is a pair (θ,Θ)(\theta,\Theta), where θC(V,G0)\theta\in C^{\infty}(V,G_{0}) and ΘΩ1(V,𝔤1)\Theta\in\Omega^{1}(V,\mathfrak{g}_{1}). The two components act on a 2-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) valued in 𝖌{\boldsymbol{\mathfrak{g}}} in the following way

θ(γ0,γ1)\displaystyle\theta\cdot(\gamma_{0},\gamma_{1}) =(θd(θ1)+θγ0θ1,θγ1)\displaystyle=(\theta d(\theta^{-1})+\theta\gamma_{0}\theta^{-1},\theta\vartriangleright\gamma_{1}) (4.1)
Θ(γ0,γ1)\displaystyle\Theta\cdot(\gamma_{0},\gamma_{1}) =(γ0+δ(Θ),γ1dΘγ0Θ12[Θ,Θ]),\displaystyle=\Big{(}\gamma_{0}+\delta(\Theta),\gamma_{1}-d\Theta-\gamma_{0}\wedge^{\vartriangleright}\Theta-\frac{1}{2}[\Theta,\Theta]\Big{)}, (4.2)

where \wedge^{\vartriangleright} is defined in (3.29). As a convention, (θ,Θ)(\theta,\Theta) acts first by θ\theta, and then by Θ\Theta:

(θ,Θ)(γ0,γ1)=Θ(θ(γ0,γ1)).\displaystyle(\theta,\Theta)\cdot(\gamma_{0},\gamma_{1})=\Theta\cdot\big{(}\theta\cdot(\gamma_{0},\gamma_{1})\big{)}. (4.3)

By direct computation, the two components of the curvature are given by

κθ(γ0,γ1)=θκγ0,γ1θ1and𝒦θ(γ0,γ1)=θ𝒦γ0,γ1\displaystyle\kappa^{\theta\cdot(\gamma_{0},\gamma_{1})}=\theta\kappa^{\gamma_{0},\gamma_{1}}\theta^{-1}\quad\text{and}\quad\mathcal{K}^{\theta\cdot(\gamma_{0},\gamma_{1})}=\theta\vartriangleright\mathcal{K}^{\gamma_{0},\gamma_{1}} (4.4)

and

κΘ(γ0,γ1)=κγ0,γ1and𝒦Θ(γ0,γ1)=𝒦γ0,γ1κγ0,γ1Θ.\displaystyle\kappa^{\Theta\cdot(\gamma_{0},\gamma_{1})}=\kappa^{\gamma_{0},\gamma_{1}}\quad\text{and}\quad\mathcal{K}^{\Theta\cdot(\gamma_{0},\gamma_{1})}=\mathcal{K}^{\gamma_{0},\gamma_{1}}-\kappa^{\gamma_{0},\gamma_{1}}\vartriangleright\Theta. (4.5)

In particular, for a fake-flat 2-connection, where κγ0,γ1=0\kappa^{\gamma_{0},\gamma_{1}}=0, we obtain

κ(θ,Θ)(γ0,γ1)=0and𝒦(θ,Θ)(γ0,γ1)=θ𝒦γ0,γ1.\displaystyle\kappa^{(\theta,\Theta)\cdot(\gamma_{0},\gamma_{1})}=0\quad\text{and}\quad\mathcal{K}^{(\theta,\Theta)\cdot(\gamma_{0},\gamma_{1})}=\theta\vartriangleright\mathcal{K}^{\gamma_{0},\gamma_{1}}. (4.6)

An interesting feature of higher gauge theory is that given a fake flat 2-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) over a contractible space, one can always find a gauge transformation such that (θ,Θ)(γ0,γ1)=(0,γ1ab)(\theta,\Theta)\cdot(\gamma_{0},\gamma_{1})=(0,\gamma_{1}^{{\operatorname{ab}}}), with γ1ab\gamma_{1}^{{\operatorname{ab}}} valued in the kernel of δ:𝔤1𝔤0\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0} [49, Theorem 4.3] (see also  [23] and  [54] for similar results). We consider this construction for the universal 2-connection.

Let (ζ0,ζ1)(\zeta_{0},\zeta_{1}) be the universal 2-connection from (3.55). First, we define θC(V,K^0)\theta\in C^{\infty}(V,\hat{K}_{0}) by exponentiating a Lie algebra valued function ηC(V,𝔨0)\eta\in C^{\infty}(V,\mathfrak{k}_{0})

θ(x)=exp(η(x))whereη(x)=i=1dziei.\displaystyle\theta(x)=\exp_{\otimes}(\eta(x))\quad\text{where}\quad\eta(x)=\sum_{i=1}^{d}z_{i}e_{i}. (4.7)

Note that ζ0=dη\zeta_{0}=d\eta. Then, we have

θζ0\displaystyle\theta\cdot\zeta_{0} =Adexp(η)(dη)+exp(η)dexp(η)\displaystyle=\operatorname{Ad}_{\exp(\eta)}(d\eta)+\exp(\eta)d\exp(-\eta) (4.8)
=(exp(adη)+1exp(adη)adη)(dη)\displaystyle=\left(\exp(\operatorname{ad}_{\eta})+\frac{1-\exp(\operatorname{ad}_{\eta})}{\operatorname{ad}_{\eta}}\right)(d\eta) (4.9)
=(k=0adηkk!k=1adηk1k!)(dη)\displaystyle=\left(\sum_{k=0}^{\infty}\frac{\operatorname{ad}_{\eta}^{k}}{k!}-\sum_{k=1}^{\infty}\frac{\operatorname{ad}_{\eta}^{k-1}}{k!}\right)(d\eta) (4.10)
=k=0k+1(k+2)!adηk[η,dη]\displaystyle=\sum_{k=0}^{\infty}\frac{k+1}{(k+2)!}\operatorname{ad}_{\eta}^{k}[\eta,d\eta] (4.11)
=δ(k=0k+1(k+2)!adηkU),\displaystyle=\delta\left(\sum_{k=0}^{\infty}\frac{k+1}{(k+2)!}\operatorname{ad}_{\eta}^{k}U\right), (4.12)

where

U=i<jeiej(zidzjzjdzi)Ω1(V,𝔨1).\displaystyle U=\sum_{i<j}e_{i}\wedge e_{j}(z_{i}dz_{j}-z_{j}dz_{i})\in\Omega^{1}(V,\mathfrak{k}_{1}). (4.13)

Note that UU has the property that

δ(U)=[η,dη]anddU=2ζ1.\displaystyle\delta(U)=[\eta,d\eta]\quad\text{and}\quad dU=2\zeta_{1}. (4.14)

Now, if we let

Θ=k=0k+1(k+2)!adηkU,\displaystyle\Theta=-\sum_{k=0}^{\infty}\frac{k+1}{(k+2)!}\operatorname{ad}_{\eta}^{k}U, (4.15)

then we have

(θ,Θ)ζ0=0.\displaystyle(\theta,\Theta)\cdot\zeta_{0}=0. (4.16)

Thus, the transformed 2-connection is

ζ1ab(θ,Θ)ζ1=θζ1dΘ+12[Θ,Θ].\displaystyle\zeta_{1}^{\operatorname{ab}}\coloneqq(\theta,\Theta)\cdot\zeta_{1}=\theta\vartriangleright\zeta_{1}-d\Theta+\frac{1}{2}[\Theta,\Theta]. (4.17)

Note that because fake-flatness is preserved by gauge transformations, ζ1ab\zeta_{1}^{{\operatorname{ab}}} is valued in the completion of 𝔞1ker(δ:𝔨1𝔨0)\mathfrak{a}_{1}\coloneqq\ker(\delta:\mathfrak{k}_{1}\to\mathfrak{k}_{0}), which is an abelian Lie algebra.

While the expression (4.17) for the abelianized 2-connection (0,ζ1ab)(0,\zeta_{1}^{{\operatorname{ab}}}) is quite complicated, its 2-curvature can be computed explicitly. The 2-curvature of the universal 2-connection is

𝒦ζ0,ζ1=i<j<kBi,j,kdzidzjdzk𝔞1Ω3,cl(V)\displaystyle\mathcal{K}^{\zeta_{0},\zeta_{1}}=\sum_{i<j<k}B_{i,j,k}dz_{i}\wedge dz_{j}\wedge dz_{k}\in\mathfrak{a}_{1}\otimes\Omega^{3,\operatorname{cl}}(V) (4.18)

where

Bi,j,k=[ei,ejek][ej,eiek]+[ek,eiej]𝔞1𝔨1.\displaystyle B_{i,j,k}=[e_{i},e_{j}\wedge e_{k}]-[e_{j},e_{i}\wedge e_{k}]+[e_{k},e_{i}\wedge e_{j}]\in\mathfrak{a}_{1}\subset\mathfrak{k}_{1}. (4.19)

Then, applying the gauge transformation to the 2-curvature using (4.6), we obtain

𝒦ab𝒦0,ζ1ab\displaystyle\mathcal{K}^{{\operatorname{ab}}}\coloneqq\mathcal{K}^{0,\zeta_{1}^{{\operatorname{ab}}}} =θ𝒦ζ0,ζ1=m=01m!adηm(1i<j<knBi,j,kdzidzjdzk).\displaystyle=\theta\vartriangleright\mathcal{K}^{\zeta_{0},\zeta_{1}}=\sum_{m=0}^{\infty}\frac{1}{m!}\operatorname{ad}_{\eta}^{m}\left(\sum_{1\leq i<j<k\leq n}B_{i,j,k}\,dz_{i}\wedge dz_{j}\wedge dz_{k}\right). (4.20)

Furthermore, because the abelianized 1-connection is trivial, the Bianchi identity for the 2-connection [44, Proposition 2.20] shows that d𝒦ab=0d\mathcal{K}^{{\operatorname{ab}}}=0, and thus

𝒦ab𝔞^1^Ω^3,cl(V).\displaystyle\mathcal{K}^{{\operatorname{ab}}}\in\hat{\mathfrak{a}}_{1}\,\hat{\otimes}\,\hat{\Omega}^{3,\operatorname{cl}}(V). (4.21)

4.2. Polynomial Differential Forms and Currents

To understand 𝒦ab\mathcal{K}^{{\operatorname{ab}}}, we make use of the isomorphism between 𝔞1\mathfrak{a}_{1} 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 GL(V){\operatorname{GL}}(V)-representations. Here, we will consider VV to be a finite-dimensional 𝕂\mathbb{K}-vector space, for 𝕂=\mathbb{K}=\mathbb{R} or 𝕂=\mathbb{K}=\mathbb{C}. Polynomial differential forms and their completions are naturally graded by their total weight rr. We define

Ω¯m(V)r=Srm(V)Λm(V),Ω¯m(V)r=mΩ¯m(V)r,Ω^m(V)r=mΩ¯m(V)r,\displaystyle\overline{\Omega}^{m}(V)_{r}=S^{r-m}(V^{*})\otimes\Lambda^{m}(V^{*}),\quad\overline{\Omega}^{m}(V)\coloneqq\bigoplus_{r=m}^{\infty}\overline{\Omega}^{m}(V)_{r},\quad\hat{\Omega}^{m}(V)\coloneqq\prod_{r=m}^{\infty}\overline{\Omega}^{m}(V)_{r}, (4.22)

equipped with the usual exterior differential dm:Ω¯m(V)Ω¯m+1(V)d^{m}:\overline{\Omega}^{m}(V)\to\overline{\Omega}^{m+1}(V). The polynomial currents and their completions are defined by taking the graded dual of Ω¯\overline{\Omega}^{\bullet} as follows

Γ¯m(V)rΩ¯m(V)r=Srm(V)Λm(V),Γ¯m(V)r=mΓ¯m(V)r,Γ^m(V)r=mΓ¯m(V)r.\displaystyle\overline{\Gamma}_{m}(V)_{r}\coloneqq\overline{\Omega}^{m}(V)_{r}^{*}=S^{r-m}(V)\otimes\Lambda^{m}(V),\hskip 7.0pt\overline{\Gamma}_{m}(V)\coloneqq\bigoplus_{r=m}^{\infty}\overline{\Gamma}_{m}(V)_{r},\hskip 7.0pt\hat{\Gamma}_{m}(V)\coloneqq\prod_{r=m}^{\infty}\overline{\Gamma}_{m}(V)_{r}. (4.23)

To simplify notation, we will often write Γ¯Γ¯(V)\overline{\Gamma}_{\bullet}\coloneqq\overline{\Gamma}_{\bullet}(V) and Ω¯Ω¯(V)\overline{\Omega}^{\bullet}\coloneqq\overline{\Omega}^{\bullet}(V). Differential forms and currents are naturally equipped with the structure of GL(V){\operatorname{GL}}(V)-representations. An element gGL(V)g\in{\operatorname{GL}}(V) acts on a current α=u1ur(v1vk)Γ¯k(V)k+r\alpha=u_{1}\cdots u_{r}\otimes(v_{1}\wedge\ldots\wedge v_{k})\in\overline{\Gamma}_{k}(V)_{k+r} as follows

gα=(gu1)(gur)((gv1)(gvk)),\displaystyle g\gtrdot\alpha=(g\cdot u_{1})\cdots(g\cdot u_{r})\otimes\left((g\cdot v_{1})\wedge\ldots\wedge(g\cdot v_{k})\right), (4.24)

and acts on ωΩ¯(V)\omega\in\overline{\Omega}^{\bullet}(V) by pullback as follows

gω=(g1)ω.\displaystyle g\gtrdot\omega=(g^{-1})^{*}\omega. (4.25)

The exterior differential is GL(V){\operatorname{GL}}(V)-equivariant. Note also that the action of the subgroup of scalars 𝕂GL(V)\mathbb{K}^{*}\subseteq{\operatorname{GL}}(V) induces a grading on differential forms and currents which agrees with the weight grading defined above.

Although the polynomial forms Ω¯\overline{\Omega}^{\bullet} and currents Γ¯\overline{\Gamma}_{\bullet} are naturally dual to each other, the naive pairing between them is not GL(V){\operatorname{GL}}(V)-equivariant. Here, we will use the Cartan calculus to define the “correct” pairing which is implicitly used in [34]. In the following, we view V𝔛(V)V\subset\mathfrak{X}(V) as constant vector fields on VV; we note that these vector fields commute. These vector fields act on differential forms in two ways. First, for any vVv\in V, we can take the Lie derivative Lv:Ω¯m(V)Ω¯m(V)L_{v}:\overline{\Omega}^{m}(V)\to\overline{\Omega}^{m}(V), and for constant vector fields, this is commutative,

[Lv,Lu]=L[v,u]=0.\displaystyle[L_{v},L_{u}]=L_{[v,u]}=0. (4.26)

Second, we can take the interior product ιv:Ω¯m(V)Ω¯m1(V)\iota_{v}:\overline{\Omega}^{m}(V)\to\overline{\Omega}^{m-1}(V), which satisfies

ιvιu=ιuιv.\displaystyle\iota_{v}\iota_{u}=-\iota_{u}\iota_{v}. (4.27)

For kmk\leq m, we use these operations to define a map Pk,m:Γ¯k(V)Ω¯m(V)Ω¯mk(V)P_{k,m}:\overline{\Gamma}_{k}(V)\otimes\overline{\Omega}^{m}(V)\to\overline{\Omega}^{m-k}(V) where

Pk,m(u1ur(v1vk)ω)Lu1Lurιv1ιvkω.\displaystyle P_{k,m}(u_{1}\cdots u_{r}\otimes(v_{1}\wedge\ldots\wedge v_{k})\otimes\omega)\coloneqq L_{u_{1}}\ldots L_{u_{r}}\iota_{v_{1}}\ldots\iota_{v_{k}}\omega. (4.28)
Lemma 4.2.

For kmk\leq m, the map Pk,m:Γ¯k(V)Ω¯m(V)Ω¯mk(V)P_{k,m}:\overline{\Gamma}_{k}(V)\otimes\overline{\Omega}^{m}(V)\to\overline{\Omega}^{m-k}(V) is GL(V){\operatorname{GL}}(V)-equivariant. In particular, for gGL(V)g\in{\operatorname{GL}}(V), αΓ¯k(V)\alpha\in\overline{\Gamma}_{k}(V), ωΩ¯m(V)\omega\in\overline{\Omega}^{m}(V), we have

gPk,m(αω)=Pk,m((gα)(gω)).\displaystyle g\gtrdot P_{k,m}(\alpha\otimes\omega)=P_{k,m}((g\gtrdot\alpha)\otimes(g\gtrdot\omega)). (4.29)
Proof.

First, given gGL(V)g\in{\operatorname{GL}}(V) and vVv\in V, the pullback satisfies

gdω=d(gω)andg(ιvω)=ιg1(v)(gω).\displaystyle g^{*}d\omega=d(g^{*}\omega)\quad\text{and}\quad g^{*}(\iota_{v}\omega)=\iota_{g^{-1}(v)}(g^{*}\omega). (4.30)

Then, by the Cartan formula Lv=dιv+ιvdL_{v}=d\iota_{v}+\iota_{v}d, we have

g(Lvω)=Lg1(v)(gω).\displaystyle g^{*}(L_{v}\omega)=L_{g^{-1}(v)}(g^{*}\omega). (4.31)

Thus, by the definition of Pk,mP_{k,m} in (4.28), we have

gPk,m(u1ur(v1vk)ω)\displaystyle g^{*}P_{k,m}(u_{1}\cdots u_{r}\otimes(v_{1}\wedge\ldots v_{k})\otimes\omega) =Lg1(u1)Lg1(ur)ιg1(v1)ιg1(vk)gω\displaystyle=L_{g^{-1}(u_{1})}\ldots L_{g^{-1}(u_{r})}\iota_{g^{-1}(v_{1})}\ldots\iota_{g^{-1}(v_{k})}g^{*}\omega (4.32)
=Pk,m(g1(u1ur(v1vk))(g1ω)).\displaystyle=P_{k,m}\left(g^{-1}\gtrdot(u_{1}\cdots u_{r}\otimes(v_{1}\wedge\ldots v_{k}))\otimes(g^{-1}\gtrdot\omega)\right).

We also record a formula relating the exterior derivative and Pk,mP_{k,m}.

Lemma 4.3.

For u(v1vk)Γ¯k(V)u\otimes(v_{1}\wedge\ldots\wedge v_{k})\in\overline{\Gamma}_{k}(V), where uS(V)u\in S(V), and ωΩ¯m(V)\omega\in\overline{\Omega}^{m}(V), we have

dPk,m(u(v1vk)ω)\displaystyle dP_{k,m}(u\otimes(v_{1}\wedge\ldots\wedge v_{k})\otimes\omega) =i=1k(1)i1Pk,m(uvi(v1v^ivk)ω)\displaystyle=\sum_{i=1}^{k}(-1)^{i-1}P_{k,m}(uv_{i}\otimes(v_{1}\wedge\ldots\hat{v}_{i}\ldots\wedge v_{k})\otimes\omega) (4.33)
+(1)kPk,m(u(v1vk)dω)\displaystyle\quad+(-1)^{k}P_{k,m}(u\otimes(v_{1}\wedge\ldots\wedge v_{k})\otimes d\omega)
Proof.

This is immediate by the properties dLv=LvddL_{v}=L_{v}d, dιv=ιvd+Lvd\iota_{v}=-\iota_{v}d+L_{v}, and Lvιu=ιuLvL_{v}\iota_{u}=\iota_{u}L_{v}. ∎

Next, we assemble all of the Pk,mP_{k,m} into a map

P:Γ¯(V)Ω¯(V)Ω¯(V).\displaystyle P:\overline{\Gamma}_{\bullet}(V)\otimes\overline{\Omega}^{\bullet}(V)\to\overline{\Omega}^{\bullet}(V). (4.34)

For the final step in defining the equivariant pairing, consider the inclusion of the origin i:0Vi:0\to V, which is trivially GL(V){\operatorname{GL}}(V)-equivariant. This induces a trivially GL(V){\operatorname{GL}}(V)-equivariant pullback

i:Ω¯(V)Ω¯(0)=𝕂.\displaystyle i^{*}:\overline{\Omega}^{\bullet}(V)\to\overline{\Omega}^{\bullet}(0)=\mathbb{K}. (4.35)
Proposition 4.4.

The pairing

,iPk,m:Γ¯k(V)Ω¯m(V)𝕂,\displaystyle\langle\cdot,\cdot\rangle\coloneqq i^{*}P_{k,m}:\overline{\Gamma}_{k}(V)\otimes\overline{\Omega}^{m}(V)\to\mathbb{K}, (4.36)

is non-degenerate for k=mk=m, vanishes for kmk\neq m, and is GL(V){\operatorname{GL}}(V)-equivariant.

If necessary, we use ,m:Γ¯mΩ¯m𝕂\langle\cdot,\cdot\rangle_{m}:\overline{\Gamma}_{m}\otimes\overline{\Omega}^{m}\to\mathbb{K} to specify the degree. We can describe this pairing explicitly in coordinates. Let (z1,,zn)(z_{1},...,z_{n}) be a basis of VV^{*} and let (e1,,en)(e_{1},...,e_{n}) be the dual basis of VV. Given α=(α1,,αn)n\alpha=(\alpha_{1},\ldots,\alpha_{n})\in\mathbb{N}^{n}, we define zαz1α1znαnS(V)z^{\alpha}\coloneqq z_{1}^{\alpha_{1}}\cdots z_{n}^{\alpha_{n}}\in S(V^{*}) and similarly for eαS(V)e^{\alpha}\in S(V). Given an increasing index set I=(i1<<im)I=(i_{1}<...<i_{m}), we define dzI=dzi1dzimmVdz_{I}=dz_{i_{1}}\wedge...\wedge dz_{i_{m}}\in\wedge^{m}V^{*}. Similarly, we define eI=ei1eimmVe_{I}=e_{i_{1}}\wedge...\wedge e_{i_{m}}\in\wedge^{m}V. In this way, we obtain the bases

zαdvIandeαeI\displaystyle z^{\alpha}\otimes dv_{I}\quad\text{and}\quad e^{\alpha}\otimes e_{I} (4.37)

of Ω¯m(V)\overline{\Omega}^{m}(V) and Γ¯m(V)\overline{\Gamma}_{m}(V), respectively. In terms of these coordinates, the pairing satisfies

eαeI,zβdzJm=(1)m(m1)2α!δα,βδI,Jwhereα!=αi!.\displaystyle\langle e^{\alpha}\otimes e_{I},z^{\beta}\otimes dz_{J}\rangle_{m}=(-1)^{\frac{m(m-1)}{2}}\alpha!\,\delta_{\alpha,\beta}\delta_{I,J}\quad\text{where}\quad\alpha!=\prod\alpha_{i}!. (4.38)

Using the pairing, we can now define the codifferential

m:Γ¯m(V)Γ¯m1(V)bymα,ω=(1)mα,dm1ω\displaystyle\partial_{m}:\overline{\Gamma}_{m}(V)\to\overline{\Gamma}_{m-1}(V)\quad\text{by}\quad\langle\partial_{m}\alpha,\omega\rangle=(-1)^{m}\langle\alpha,d^{m-1}\omega\rangle (4.39)

for any αΓ¯m(V)\alpha\in\overline{\Gamma}_{m}(V) and ωΩ¯m1(V)\omega\in\overline{\Omega}^{m-1}(V). It is GL(V){\operatorname{GL}}(V)-equivariant. The closed forms and closed currents are defined as usual to be

Ω¯m,cl(V)ker(dm:Ω¯mΩ¯m+1)andΓ¯mcl(V)ker(m:Γ¯mΓ¯m1)\displaystyle\overline{\Omega}^{m,\operatorname{cl}}(V)\coloneqq\ker(d^{m}:\overline{\Omega}^{m}\to\overline{\Omega}^{m+1})\quad\text{and}\quad\overline{\Gamma}^{\operatorname{cl}}_{m}(V)\coloneqq\ker(\partial_{m}:\overline{\Gamma}_{m}\to\overline{\Gamma}_{m-1}) (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

Γ¯mclim(m+1:Γ¯m+1Γ¯m)coker(m+2:Γ¯m+2Γ¯m+1)(Ω¯m+1,cl),\displaystyle\overline{\Gamma}^{\operatorname{cl}}_{m}\cong\operatorname{im}(\partial_{m+1}:\overline{\Gamma}_{m+1}\to\overline{\Gamma}_{m})\cong{\operatorname{coker}}(\partial_{m+2}:\overline{\Gamma}_{m+2}\to\overline{\Gamma}_{m+1})\cong(\overline{\Omega}^{m+1,\operatorname{cl}})^{\vee}, (4.41)

where ()(\cdot)^{\vee} denotes the graded dual. Given α=m+1(β)Γ¯mcl\alpha=\partial_{m+1}(\beta)\in\overline{\Gamma}^{\operatorname{cl}}_{m}, where βΓ¯m+1\beta\in\overline{\Gamma}_{m+1}, and ω=dm(η)Ω¯m+1,cl\omega=d^{m}(\eta)\in\overline{\Omega}^{m+1,\operatorname{cl}}, where ηΩ¯m\eta\in\overline{\Omega}^{m}, the pairing between closed forms and closed currents is

α,ωcl(1)m+1α,ηm=β,ωm+1.\displaystyle\langle\alpha,\omega\rangle_{\operatorname{cl}}\coloneqq(-1)^{m+1}\langle\alpha,\eta\rangle_{m}=\langle\beta,\omega\rangle_{m+1}. (4.42)

In what follows, we will need the duality between closed 22-currents and closed 33-forms and it will be useful to have explicit dual bases. Given index list 𝐪=(q1,,qr)\mathbf{q}=(q_{1},\ldots,q_{r}), and indices i,j,ki,j,k, such that q1qri<j<kq_{1}\geq\ldots\geq q_{r}\geq i<j<k, define

γ𝐪,ijk=eq1eqreiejek(Γ¯3(V))r+3,ω𝐪,ijk=zαα!dzjdzk(Ω¯2)r+3,\displaystyle\gamma_{\mathbf{q},ijk}=-e_{q_{1}}...e_{q_{r}}\otimes e_{i}\wedge e_{j}\wedge e_{k}\in(\overline{\Gamma}_{3}(V))_{r+3},\qquad\omega_{\mathbf{q},ijk}=\frac{z^{\alpha}}{\alpha!}dz_{j}\wedge dz_{k}\in(\overline{\Omega}^{2})_{r+3}, (4.43)

where 𝜶=(α1,,αn){\boldsymbol{\alpha}}=(\alpha_{1},\ldots,\alpha_{n}), and αs\alpha_{s} denotes the number of times the index ss appears in (𝐪,i)(\mathbf{q},i).

Lemma 4.5.

There is a natural isomorphism

Γ¯2cl(Ω¯3,cl)\displaystyle\overline{\Gamma}^{\operatorname{cl}}_{2}\cong(\overline{\Omega}^{3,\operatorname{cl}})^{\vee} (4.44)

where, on the right-hand side, we are taking the graded dual. The closed currents 3(γ𝐪,ijk)\partial_{3}(\gamma_{\mathbf{q},ijk}) and closed forms d2(ω𝐪,ijk)d^{2}(\omega_{\mathbf{q},ijk}), for q1qri<j<kq_{1}\geq\ldots\geq q_{r}\geq i<j<k, give dual bases in weight r+3r+3.

Proof.

The space of closed weight r+3r+3 currents (Γ¯2cl(V))r+3(\overline{\Gamma}_{2}^{\operatorname{cl}}(V))_{r+3} is the irreducible representation of GL(V){\operatorname{GL}}(V) corresponding to the hook Young diagram of size (r+1,1,1)(r+1,1,1). It follows that the currents 3(γ𝐪,ijk)\partial_{3}(\gamma_{\mathbf{q},ijk}) give a basis. Taking the pairing in (4.42), we have

3(γ𝐪,ijk),d2(ω𝐩,abc)cl\displaystyle\langle\partial_{3}(\gamma_{\mathbf{q},ijk}),d^{2}(\omega_{\mathbf{p},abc})\rangle_{\operatorname{cl}} =γ𝐪,ijk,d2(ω𝐩,abc)3\displaystyle=\langle\gamma_{\mathbf{q},ijk},d^{2}(\omega_{\mathbf{p},abc})\rangle_{3} (4.45)
=αaα!eq1eqreiejek,zp1zprdzadzbdzc3\displaystyle=\frac{-\alpha_{a}}{\alpha!}\langle e_{q_{1}}...e_{q_{r}}\otimes e_{i}\wedge e_{j}\wedge e_{k},z_{p_{1}}...z_{p_{r}}dz_{a}\wedge dz_{b}\wedge dz_{c}\rangle_{3} (4.46)
=δ(𝐪,ijk),(𝐩,abc).\displaystyle=\delta_{(\mathbf{q},ijk),(\mathbf{p},abc)}. (4.47)

Finally, we conclude this section with the following relationship between the free crossed module from (3.51) and closed 22-currents; see [18, Appendix D] for further exposition.

Theorem 4.6.

[34] The symmetrization map ρ:T(V)Λ2VS(V)Λ2V=Γ¯2(V)\rho^{\prime}:T(V)\otimes\Lambda^{2}V\to S(V)\otimes\Lambda^{2}V=\overline{\Gamma}_{2}(V) is a GL(V){\operatorname{GL}}(V)-equivariant map which descends to define a Lie algebra map

ρ:𝔨1(V)Γ¯2(V)\displaystyle\rho:\mathfrak{k}_{1}(V)\to\overline{\Gamma}_{2}(V) (4.48)

which identifies the abelianization

𝔨1(V)[𝔨1(V),𝔨1(V)]Γ¯2(V).\displaystyle\frac{\mathfrak{k}_{1}(V)}{[\mathfrak{k}_{1}(V),\mathfrak{k}_{1}(V)]}\cong\overline{\Gamma}_{2}(V). (4.49)

Furthermore, the restriction of ρ\rho to 𝔞1(V)ker(δ:𝔨1(V)𝔨0(V))\mathfrak{a}_{1}(V)\coloneqq\ker(\delta:\mathfrak{k}_{1}(V)\to\mathfrak{k}_{0}(V)) defines an isomorphism of GL(V){\operatorname{GL}}(V)-representations

ρ:𝔞1(V)Γ¯2cl(V).\displaystyle\rho:\mathfrak{a}_{1}(V)\to\overline{\Gamma}^{\operatorname{cl}}_{2}(V). (4.50)

Given indices 𝐪=(q1,,qr)\mathbf{q}=(q_{1},\ldots,q_{r}), and i,j,ki,j,k, such that q1qri<j<kq_{1}\geq\ldots\geq q_{r}\geq i<j<k, define

B𝐪,ijk[eq1,,[eqr,Bijk]]𝔞1(V).\displaystyle B_{\mathbf{q},ijk}\coloneqq[e_{q_{1}},\ldots,[e_{q_{r}},B_{ijk}]\ldots]\in\mathfrak{a}_{1}(V). (4.51)

Under the isomorphism ρ\rho from Theorem 4.6, we see that

ρ(B𝐪,ijk)=3(γ𝐪,ijk).\displaystyle\rho(B_{\mathbf{q},ijk})=\partial_{3}(\gamma_{\mathbf{q},ijk}). (4.52)

Hence, we immediately conclude that B𝐪,ijkB_{\mathbf{q},ijk} define a basis of 𝔞1(V)\mathfrak{a}_{1}(V).

4.3. Computing the Abelianized Curvature

We will use the relationship between 𝔞1\mathfrak{a}_{1} and closed polynomial currents, along with their representation-theoretic properties, to explicitly compute the abelianized curvature 𝒦ab\mathcal{K}^{{\operatorname{ab}}}. As we will be using Schur’s lemma for complex representations, it is convenient to immediately consider the complexification VVV_{\mathbb{C}}\coloneqq V\otimes_{\mathbb{R}}\mathbb{C}. Because ρ\rho is defined over \mathbb{R}, this will not affect our final calculation of 𝒦ab\mathcal{K}^{{\operatorname{ab}}}.

Denote the complexified space of currents and forms by (Γ¯)(\overline{\Gamma}_{\bullet})_{\mathbb{C}} and Ω¯\overline{\Omega}^{\bullet}_{\mathbb{C}}, respectively. Because the isomorphisms from  Lemma 4.5 and  Theorem 4.6 are GL(V){\operatorname{GL}}(V)-equivariant, they preserve the weight grading, and therefore extend to the completions. Hence, we obtain an isomorphism of GL(V){\operatorname{GL}}(V_{\mathbb{C}})-representations ρ:𝔞^1(V)(Ω¯3,cl(V))\rho:\hat{\mathfrak{a}}_{1}(V_{\mathbb{C}})\to(\overline{\Omega}^{3,\operatorname{cl}}(V_{\mathbb{C}}))^{*}, which can be extended to

ρ^=ρid:𝔞^1(V)^Ω^3,cl(Ω¯3,cl)^Ω^3,cl.\displaystyle\hat{\rho}=\rho\otimes\operatorname{id}:\hat{\mathfrak{a}}_{1}(V_{\mathbb{C}})\,\hat{\otimes}\,\hat{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}}\to(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})^{*}\,\hat{\otimes}\,\hat{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}}. (4.53)

Under this isomorphism, the curvature from (4.18), which we denote by 𝒦=𝒦ζ0,ζ1\mathcal{K}=\mathcal{K}^{\zeta_{0},\zeta_{1}}, is sent to

ρ^(𝒦)=i<j<kρ(Bi,j,k)dzidzjdzk(Ω¯3,cl)^Ω^3,cl.\displaystyle\hat{\rho}(\mathcal{K})=\sum_{i<j<k}\rho(B_{i,j,k})dz_{i}\wedge dz_{j}\wedge dz_{k}\in(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})^{*}\,\hat{\otimes}\,\hat{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}}. (4.54)

By Lemma 4.5, the closed currents ρ(Bijk)=3(γijk)\rho(B_{ijk})=\partial_{3}(\gamma_{ijk}) and closed 33-forms dzidzjdzk=dωijkdz_{i}\wedge dz_{j}\wedge dz_{k}=d\omega_{ijk} form dual bases. Hence, ρ^(𝒦)\hat{\rho}(\mathcal{K}) is identified with the identity map for the weight 33 closed forms

ρ^(𝒦)=id(Ω¯3,cl)3EndGL(V)((Ω¯3,cl)3)(Ω¯3,cl)^Ω^3,cl.\displaystyle\hat{\rho}(\mathcal{K})=\operatorname{id}_{(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{3}}\in{\operatorname{End}}_{{\operatorname{GL}}(V_{\mathbb{C}})}((\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{3})\subset(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})^{*}\,\hat{\otimes}\,\hat{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}}. (4.55)

Here, EndGL(V)(W){\operatorname{End}}_{{\operatorname{GL}}(V_{\mathbb{C}})}(W) denotes the space of equivariant morphisms of WW, a GL(V){\operatorname{GL}}(V_{\mathbb{C}})-representation. Recalling that this is precisely the GL(V){\operatorname{GL}}(V_{\mathbb{C}})-invariant subspace of the representation End(W){\operatorname{End}}(W), we conclude that ρ^(𝒦)\hat{\rho}(\mathcal{K}) (and hence 𝒦\mathcal{K}) is GL(V){\operatorname{GL}}(V_{\mathbb{C}})-invariant.

Lemma 4.7.

The abelianized curvature 𝒦ab\mathcal{K}^{{\operatorname{ab}}} is GL(V){\operatorname{GL}}(V_{\mathbb{C}})-invariant. Therefore

ρ^(Kab)r=3EndGL(V)((Ω¯3,cl)r).\displaystyle\hat{\rho}(K^{{\operatorname{ab}}})\in\prod_{r=3}^{\infty}{\operatorname{End}}_{{\operatorname{GL}}(V_{\mathbb{C}})}\left((\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{r}\right). (4.56)
Proof.

We recall from Equation (4.20) that

𝒦ab=m=01m!adηm(𝒦).\displaystyle\mathcal{K}^{{\operatorname{ab}}}=\sum_{m=0}^{\infty}\frac{1}{m!}\operatorname{ad}_{\eta}^{m}(\mathcal{K}). (4.57)

Consider ad\operatorname{ad} as the action of S(V)𝔨0C(V,𝔨0)S(V^{*})\otimes\mathfrak{k}_{0}\subset C^{\infty}(V,\mathfrak{k}_{0}) on 𝔞1(V)Ω3\mathfrak{a}_{1}(V_{\mathbb{C}})\otimes\Omega^{3}_{\mathbb{C}}. By  Corollary C.10, we note that GL(V){\operatorname{GL}}(V_{\mathbb{C}}) acts on 𝖐{\boldsymbol{\mathfrak{k}}} in a way which preserves the crossed module structure. Therefore, GL(V){\operatorname{GL}}(V_{\mathbb{C}}) acts on both S(V)𝔨0S(V^{*})\otimes\mathfrak{k}_{0} and 𝔞1(V)Ω3\mathfrak{a}_{1}(V_{\mathbb{C}})\otimes\Omega^{3}_{\mathbb{C}}, and the action ad\operatorname{ad} is GL(V){\operatorname{GL}}(V_{\mathbb{C}})-equivariant:

gads(ω)=adgs(gω),\displaystyle g\gtrdot\operatorname{ad}_{s}(\omega)=\operatorname{ad}_{g\gtrdot s}(g\gtrdot\omega), (4.58)

where gGL(V)g\in{\operatorname{GL}}(V_{\mathbb{C}}), sS(V)𝔨0s\in S(V^{*})\otimes\mathfrak{k}_{0} and ω𝔞1(V)Ω3\omega\in\mathfrak{a}_{1}(V_{\mathbb{C}})\otimes\Omega^{3}_{\mathbb{C}}. The element η=i=1dviei\eta=\sum_{i=1}^{d}v_{i}e_{i} is GL(V){\operatorname{GL}}(V_{\mathbb{C}})-invariant, since it corresponds to the identity idVVVS(V)𝔨0\operatorname{id}_{V_{\mathbb{C}}}\in V^{*}\otimes V\subset S(V^{*})\otimes\mathfrak{k}_{0}. Because 𝒦\mathcal{K} is also GL(V){\operatorname{GL}}(V_{\mathbb{C}})-invariant, the same is true for adηm(𝒦)\operatorname{ad}_{\eta}^{m}(\mathcal{K}) and thus for 𝒦ab\mathcal{K}^{{\operatorname{ab}}} as well. In other words,

𝒦abr=3EndGL(V)((Ω¯3,cl)r).\displaystyle\mathcal{K}^{{\operatorname{ab}}}\in\prod_{r=3}^{\infty}{\operatorname{End}}_{{\operatorname{GL}}(V_{\mathbb{C}})}\left((\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{r}\right). (4.59)

The upshot of Lemma 4.7 is that the abelianized curvature can be decomposed as

ρ^(Kab)=r=3ρ^(Kab)r,\displaystyle\hat{\rho}(K^{{\operatorname{ab}}})=\sum_{r=3}^{\infty}\hat{\rho}(K^{{\operatorname{ab}}})_{r}, (4.60)

where each ρ^(Kab)rEndGL(V)((Ω¯3,cl)r)\hat{\rho}(K^{{\operatorname{ab}}})_{r}\in{\operatorname{End}}_{{\operatorname{GL}}(V_{\mathbb{C}})}\left((\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{r}\right). By [28, Section 8.2, Theorem 2], each representation (Ω¯3,cl)r(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{r} is irreducible. Therefore, by Schur’s lemma,

ρ^(Kab)r=λrid(Ω¯3,cl)r,\displaystyle\hat{\rho}(K^{{\operatorname{ab}}})_{r}=\lambda_{r}\operatorname{id}_{(\overline{\Omega}^{3,\operatorname{cl}}_{\mathbb{C}})_{r}}, (4.61)

for a constant λr\lambda_{r}\in\mathbb{C}. In the following theorem, we verify that this constant is λr=1\lambda_{r}=1.

Theorem 4.8.

The abelianized curvature is

ρ^(𝒦ab)=idEnd(Ω¯3,cl).\displaystyle\hat{\rho}(\mathcal{K}^{{\operatorname{ab}}})=\operatorname{id}\in{\operatorname{End}}(\overline{\Omega}^{3,\operatorname{cl}}). (4.62)
Proof.

The element ρ^(Kab)m\hat{\rho}(K^{{\operatorname{ab}}})_{m} has weight mm in the form component. Hence, it is given by

ρ^(1m!adηm(𝒦))=1m!ρ^adηm(1i<j<knBi,j,kdzidzjdzk).\displaystyle\hat{\rho}\left(\frac{1}{m!}\operatorname{ad}_{\eta}^{m}(\mathcal{K})\right)=\frac{1}{m!}\hat{\rho}\operatorname{ad}_{\eta}^{m}\left(\sum_{1\leq i<j<k\leq n}B_{i,j,k}\,dz_{i}\wedge dz_{j}\wedge dz_{k}\right). (4.63)

To determine the constant λm\lambda_{m}, it suffices to compute the coefficient of

dω1,,1,2,3=z1mm!dz1dz2dz3.\displaystyle d\omega_{1,...,1,2,3}=\frac{z_{1}^{m}}{m!}dz_{1}\wedge dz_{2}\wedge dz_{3}. (4.64)

This is easily seen to be

ρ(ade1m(B123))=ρ(B1,,1,2,3)=3(γ1,,1,2,3).\displaystyle\rho(\operatorname{ad}_{e_{1}}^{m}(B_{123}))=\rho(B_{1,...,1,2,3})=\partial_{3}(\gamma_{1,...,1,2,3}). (4.65)

By Lemma 4.5, this is the closed current dual to dω1,,1,2,3d\omega_{1,...,1,2,3}. Hence, λm=1\lambda_{m}=1. ∎

Corollary 4.9.

The abelianized curvature is

𝒦ab=𝐪i<j<kB𝐪,ijkd(ω𝐪,ijk)𝔞^1(V)^Ω^3,cl.\displaystyle\mathcal{K}^{{\operatorname{ab}}}=\sum_{\mathbf{q}\geq i<j<k}B_{\mathbf{q},ijk}\otimes d(\omega_{\mathbf{q},ijk})\in\hat{\mathfrak{a}}_{1}(V)\hat{\otimes}\hat{\Omega}^{3,\operatorname{cl}}. (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,

Ccl1([0,1]2,V){𝐗C1([0,1]2,V):𝐗=0}.\displaystyle C^{1}_{\operatorname{cl}}([0,1]^{2},V)\coloneqq\left\{\mathbf{X}\in C^{1}([0,1]^{2},V)\,:\,\partial\mathbf{X}=0\right\}. (4.67)

To this end, we make use of the relatively simple expression for the abelianized 22-curvature from Corollary 4.9 in order to simplify this calculation. The following lemma tells us that it suffices to integrate this 22-curvature over a volume.

Lemma 4.10.

Let 𝐗Ccl1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{cl}}([0,1]^{2},V) and let 𝒳C1([0,1]3,V)\mathcal{X}\in C^{1}([0,1]^{3},V) be a volume such that all boundaries are sent to 0 except the top, which is equal to 𝐗\mathbf{X},

𝒳0,t,u=𝒳1,t,u=𝒳s,0,u=𝒳s,1,u=𝒳s,t,0=0and𝒳s,t,1=𝐗s,t.\displaystyle\mathcal{X}_{0,t,u}=\mathcal{X}_{1,t,u}=\mathcal{X}_{s,0,u}=\mathcal{X}_{s,1,u}=\mathcal{X}_{s,t,0}=0\quad\text{and}\quad\mathcal{X}_{s,t,1}=\mathbf{X}_{s,t}. (4.68)

Let (0,γ1ab)(0,\gamma_{1}^{{\operatorname{ab}}}) be an abelian 2-connection valued in 𝖌{\boldsymbol{\mathfrak{g}}}, with 2-curvature 𝒦ab\mathcal{K}^{{\operatorname{ab}}}. Then, the surface holonomy is given by

F10,γ1ab(𝐗)=𝒳𝒦ab=[0,1]3𝒦ab(𝒳s,t,us,𝒳s,t,ut,𝒳s,t,uu)𝑑s𝑑t𝑑u.\displaystyle F_{1}^{0,\gamma_{1}^{ab}}(\mathbf{X})=\int_{\mathcal{X}}\mathcal{K}^{{\operatorname{ab}}}=\int_{[0,1]^{3}}\mathcal{K}^{{\operatorname{ab}}}\left(\frac{\partial\mathcal{X}_{s,t,u}}{\partial s},\frac{\partial\mathcal{X}_{s,t,u}}{\partial t},\frac{\partial\mathcal{X}_{s,t,u}}{\partial u}\right)\,ds\,dt\,du. (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 𝐗Ccl1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{cl}}([0,1]^{2},V) be a smooth closed surface based at the origin, let (γ0,γ1)(\gamma_{0},\gamma_{1}) be a 22-connection and let (θ,Θ)(\theta,\Theta) be a gauge transformation. Then,

F1(θ,Θ)(γ0,γ1)(𝐗)=(θ(0))1F1γ0,γ1(𝐗).\displaystyle F_{1}^{(\theta,\Theta)\cdot(\gamma_{0},\gamma_{1})}(\mathbf{X})=(\theta(0))^{-1}\vartriangleright F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X}). (4.70)

Putting this all together, we obtain the following result.

Theorem 4.12.

For any smooth closed surface 𝐗Ccl1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{cl}}([0,1]^{2},V) based at the origin, the surface signature is given by

S1(𝐗)=𝐪i<j<kB𝐪,ijk𝐗ω𝐪,ijk𝔞^1(V).\displaystyle S_{1}(\mathbf{X})=\sum_{\mathbf{q}\geq i<j<k}B_{\mathbf{q},ijk}\int_{\mathbf{X}}\omega_{\mathbf{q},ijk}\in\hat{\mathfrak{a}}_{1}(V). (4.71)

Furthermore, using ρ\rho to embed 𝔞^1(V)\hat{\mathfrak{a}}_{1}(V) into the space of formal 22-currents Γ^2(V)\hat{\Gamma}_{2}(V), the surface signature is given by the following expression

S1(𝐗)=αn,i<j1α!eαeiej𝐗zα𝑑zidzjΓ^2(V).S_{1}(\mathbf{X})=\sum_{\alpha\in\mathbb{N}^{n},\ i<j}\frac{1}{\alpha!}e^{\alpha}\otimes e_{i}\wedge e_{j}\int_{\mathbf{X}}z^{\alpha}dz_{i}\wedge dz_{j}\in\hat{\Gamma}_{2}(V).
Proof.

Let (ζ0,ζ1)(\zeta_{0},\zeta_{1}) denote the universal 22-connection and recall that its surface holonomy is denoted S1(𝐗)=F1ζ0,ζ1(𝐗)S_{1}(\mathbf{X})=F_{1}^{\zeta_{0},\zeta_{1}}(\mathbf{X}). Let (θ,Θ)(\theta,\Theta) denote the abelianization gauge transformation from (4.7) and (4.15) so that θ(0)=1\theta(0)=1. Given the surface 𝐗\mathbf{X}, let 𝒳\mathcal{X} be a volume as in Lemma 4.10. Then

S1(𝐗)\displaystyle S_{1}(\mathbf{X}) =F10,ζ1ab(𝐗)=[0,1]3𝒦ab(𝒳s,t,us,𝒳s,t,ut,𝒳s,t,uu)𝑑s𝑑t𝑑u\displaystyle=F_{1}^{0,\zeta_{1}^{{\operatorname{ab}}}}(\mathbf{X})=\int_{[0,1]^{3}}\mathcal{K}^{{\operatorname{ab}}}\left(\frac{\partial\mathcal{X}_{s,t,u}}{\partial s},\frac{\partial\mathcal{X}_{s,t,u}}{\partial t},\frac{\partial\mathcal{X}_{s,t,u}}{\partial u}\right)\,ds\,dt\,du (4.72)
=𝐪i<j<kB𝐪,ijk𝒳d(ω𝐪,ijk)=𝐪i<j<kB𝐪,ijk𝐗ω𝐪,ijk,\displaystyle=\sum_{\mathbf{q}\geq i<j<k}B_{\mathbf{q},ijk}\int_{\mathcal{X}}d(\omega_{\mathbf{q},ijk})=\sum_{\mathbf{q}\geq i<j<k}B_{\mathbf{q},ijk}\int_{\mathbf{X}}\omega_{\mathbf{q},ijk}, (4.73)

where the first equality follows from  Proposition 4.11 and the fact that θ(0)=1\theta(0)=1, 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 (id𝐗)(lblbl)(\mathrm{id}\otimes\int_{\mathbf{X}})(\sum_{l}b^{l}\otimes b_{l}), for (bl,bl)(b^{l},b_{l}) a dual pair of bases of 𝔞^1(V)\hat{\mathfrak{a}}_{1}(V) and Ω^3,cl\hat{\Omega}^{3,\operatorname{cl}}. Because 𝐗\mathbf{X} is closed, we have 𝐗𝑑μ=0\int_{\mathbf{X}}d\mu=0 for all μΩ1\mu\in\Omega^{1} by Stokes’ theorem. Therefore, using ρ\rho to embed 𝔞^1(V)\hat{\mathfrak{a}}_{1}(V) into Γ^2(V)\hat{\Gamma}_{2}(V), our expression for S1(𝐗)S_{1}(\mathbf{X}) also has the same form where now (bl,bl)(b^{l},b_{l}) are a dual pair of bases of Γ^2(V)\hat{\Gamma}_{2}(V) and Ω^2\hat{\Omega}^{2}. Our final expression for the signature then follows by using the dual bases from Equation 4.38. ∎

Corollary 4.13.

Let 𝐗Ccl1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{cl}}([0,1]^{2},V). Then, S1(𝐗)=0S_{1}(\mathbf{X})=0 if and only if

𝐗ω=0\displaystyle\int_{\mathbf{X}}\omega=0 (4.74)

for all compactly supported 2-forms ωΩc2(V)\omega\in\Omega^{2}_{c}(V).

Proof.

By  Theorem 4.12, the condition that S1(𝐗)=0S_{1}(\mathbf{X})=0 is equivalent to 𝐗ω=0\int_{\mathbf{X}}\omega=0 for all polynomial 22-forms. Since polynomial 2-forms Ω¯2\overline{\Omega}^{2} are dense in the compactly supported 2-forms Ωc2\Omega_{c}^{2}, we obtain the desired result. ∎

Remark 4.14.

Let 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) be a thinly null-homotopic path. The existence of a height function h:[0,1]h:[0,1]\to\mathbb{R} from the analytic condition (A1) is used to detect cancellations: whenever s<ts<t with h(s)=h(t)=infsuth(u)h(s)=h(t)=\inf_{s\leq u\leq t}h(u), the restriction 𝐱|[s,t]\mathbf{x}|_{[s,t]} 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 𝐗Ccl1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{cl}}([0,1]^{2},V). 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 PL0(V)\operatorname{PL}_{0}(V) 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 GG, let 𝖷𝖦𝗋𝗉(G)\mathsf{XGrp}(G) be the subcategory of crossed modules

𝐆=(δ:HG,)\displaystyle\mathbf{G}=(\delta:H\to G,\vartriangleright) (5.1)

where GG is fixed. Let 𝖲𝖾𝗍/G\mathsf{Set}_{/G} denote the category of set functions ρ:LG\rho:L\to G, where LL is a set and GG is the fixed group. There is a natural forgetful functor, along with a free functor as a left adjoint,

𝖥𝗈𝗋:𝖷𝖦𝗋𝗉(G)𝖲𝖾𝗍/Gand𝖥𝗋:𝖲𝖾𝗍/G𝖷𝖦𝗋𝗉(G).\displaystyle\mathsf{For}:\mathsf{XGrp}(G)\to\mathsf{Set}_{/G}\quad\text{and}\quad\mathsf{Fr}:\mathsf{Set}_{/G}\to\mathsf{XGrp}(G). (5.2)

The free crossed module generated by ρ:LG\rho:L\to G can be constructed as follows [8, Proposition 3.4.3]. Let 𝖥𝗋~(ρ)\widetilde{\mathsf{Fr}}(\rho) be the free group generated by G×LG\times L. It is equipped with a left GG-action by automorphisms which is defined on the generators by g(g,λ)=(gg,λ)g^{\prime}\vartriangleright(g,\lambda)=(g^{\prime}g,\lambda). Define a homomorphism δ:𝖥𝗋~(ρ)G\delta:\widetilde{\mathsf{Fr}}(\rho)\to G by the following map on generators

δ(g,λ)=gρ(λ)g1.\displaystyle\delta(g,\lambda)=g\cdot\rho(\lambda)\cdot g^{-1}. (5.3)

This is the free pre-crossed module on ρ\rho. Then, the free crossed module 𝖥𝗋(ρ)𝖥𝗋~(ρ)/Pf\mathsf{Fr}(\rho)\coloneqq\widetilde{\mathsf{Fr}}(\rho)/\sim_{{\operatorname{Pf}}} is the quotient by the Peiffer identity,

δ(g,λ)(g,λ)Pf(g,λ)(g,λ)(g,λ)1.\displaystyle\delta(g,\lambda)\vartriangleright(g^{\prime},\lambda^{\prime})\sim_{{\operatorname{Pf}}}(g,\lambda)\cdot(g^{\prime},\lambda^{\prime})\cdot(g,\lambda)^{-1}. (5.4)

The unit of the adjunction provides a map ηρ:L𝖥𝗋1(ρ)\eta_{\rho}:L\to\mathsf{Fr}_{1}(\rho) given by sending λL\lambda\in L to the equivalence class of (1,λ)(1,\lambda).

Next, let 𝖲𝖦\mathsf{SG} denote the comma category (id𝖥𝗈𝗋)(\operatorname{id}\downarrow\mathsf{For}) associated to the functors id:𝖲𝖾𝗍𝖲𝖾𝗍\operatorname{id}:\mathsf{Set}\to\mathsf{Set} and 𝖥𝗈𝗋:𝖦𝗋𝗉𝖲𝖾𝗍\mathsf{For}:\mathsf{Grp}\to\mathsf{Set}. An object of 𝖲𝖦\mathsf{SG} is given by the data of a set LL, a group GG, and a map of sets ρ:LG\rho:L\to G. A morphism f=(fL,fG):(ρ1:L1G1)(ρ2:L2G2)f=(f_{L},f_{G}):(\rho_{1}:L_{1}\to G_{1})\to(\rho_{2}:L_{2}\to G_{2}) consists of a set map fL:L1L2f_{L}:L_{1}\to L_{2} and a group homomorphism fG:G1G2f_{G}:G_{1}\to G_{2} such that the following diagram commutes

L1{L_{1}}G1{G_{1}}L2{L_{2}}G2.{G_{2}.}ρ1\scriptstyle{\rho_{1}}fL\scriptstyle{f_{L}}fG\scriptstyle{f_{G}}ρ2\scriptstyle{\rho_{2}} (5.5)

There exists a natural forgetful functor 𝖥𝗈𝗋:𝖷𝖦𝗋𝗉𝖲𝖦\mathsf{For}:\mathsf{XGrp}\to\mathsf{SG}. Free crossed modules of groups satisfy the following universal property.

Theorem 5.1.

Let (ρ:LH)𝖲𝖦(\rho:L\to H)\in\mathsf{SG}, 𝐆=(δ:G1G0,)𝖷𝖦𝗋𝗉\mathbf{G}=(\delta:G_{1}\to G_{0},\vartriangleright)\in\mathsf{XGrp}, and

f=(f1,f0):(ρ:LH)𝖥𝗈𝗋(δ:G1G0,).\displaystyle f=(f_{1},f_{0}):(\rho:L\to H)\to\mathsf{For}(\delta:G_{1}\to G_{0},\vartriangleright). (5.6)

Then there is a unique morphism of crossed modules

F=(F1,F0):𝖥𝗋(ρ)𝐆,\displaystyle F=(F_{1},F_{0}):\mathsf{Fr}(\rho)\to\mathbf{G}, (5.7)

such that F1ηρ=f1F_{1}\circ\eta_{\rho}=f_{1} and F0=f0F_{0}=f_{0}.

Proof.

Given the group homomorphism f0:HG0f_{0}:H\to G_{0}, we consider the pullback crossed module [8, Definition 5.1.1]

f0𝐆(δ:f0G1H,),f0G1{(h,g)H×G1:f0(h)=δ(g)},\displaystyle f_{0}^{*}\mathbf{G}\coloneqq(\delta:f_{0}^{*}G_{1}\to H,\vartriangleright),\quad f_{0}^{*}G_{1}\coloneqq\{(h,g)\in H\times G_{1}\,:\,f_{0}(h)=\delta(g)\}, (5.8)

and define the map f~1:f0G1G1\tilde{f}_{1}:f_{0}^{*}G_{1}\to G_{1} given by f~1(h,g)=g\tilde{f}_{1}(h,g)=g, which gives a morphism of crossed modules

f~=(f~1,f0):f0𝐆𝐆.\displaystyle\tilde{f}=(\tilde{f}_{1},f_{0}):f_{0}^{*}\mathbf{G}\to\mathbf{G}. (5.9)

We define a map u1(f):Lf0G1u_{1}(f):L\to f_{0}^{*}G_{1} by u1(f)(λ)=(ρ(λ),f1(λ))u_{1}(f)(\lambda)=(\rho(\lambda),f_{1}(\lambda)). This defines a map

u(f)=(u1(f),idH):(ρ:LH)𝖥𝗈𝗋(f0𝐆)\displaystyle u(f)=(u_{1}(f),\operatorname{id}_{H}):(\rho:L\to H)\to\mathsf{For}(f_{0}^{*}\mathbf{G}) (5.10)

in 𝖲𝖾𝗍/H\mathsf{Set}_{/H} such that 𝖥𝗈𝗋(f~)u(f)=f\mathsf{For}(\tilde{f})\circ u(f)=f. Then, using the universal property of 𝖥𝗋(ρ)\mathsf{Fr}(\rho) in 𝖷𝖦𝗋𝗉(H)\mathsf{XGrp}(H), there is a unique map

g:𝖥𝗋(ρ)f0𝐆\displaystyle g:\mathsf{Fr}(\rho)\to f_{0}^{*}\mathbf{G} (5.11)

which satisfies 𝖥𝗈𝗋(g)ηρ=u(f)\mathsf{For}(g)\circ\eta_{\rho}=u(f) in 𝖲𝖾𝗍/H\mathsf{Set}_{/H}. We define

F=𝖥𝗋(ρ)𝑔f0𝐆f~𝐆,\displaystyle F=\mathsf{Fr}(\rho)\xrightarrow{g}f_{0}^{*}\mathbf{G}\xrightarrow{\tilde{f}}\mathbf{G}, (5.12)

which is a morphism of crossed modules and it satisfies

𝖥𝗈𝗋(F)ηρ=𝖥𝗈𝗋(f~)𝖥𝗈𝗋(g)ηρ=𝖥𝗈𝗋(f~)u(f)=f,\displaystyle\mathsf{For}(F)\circ\eta_{\rho}=\mathsf{For}(\tilde{f})\circ\mathsf{For}(g)\circ\eta_{\rho}=\mathsf{For}(\tilde{f})\circ u(f)=f, (5.13)

so FF is the desired morphism. The map FF is unique because the map gg coincides with the unique factorization of the morphism FF 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 PL0(V)\operatorname{PL}_{0}(V), and define the group homomorphism

t:PL0(V)V,(v1,,vk)i=1kvi\displaystyle t:\operatorname{PL}_{0}(V)\to V,\quad(v_{1},\ldots,v_{k})\mapsto\sum_{i=1}^{k}v_{i} (5.14)

which defines the endpoint of a PL path. We define the group of piecewise linear loops by PL0cl(V)ker(t)\operatorname{PL}_{0}^{\operatorname{cl}}(V)\coloneqq\ker(t). Furthermore, we define the set of planar loops by

PlanarLoop(V){𝐱PL0cl(V):dim(span(𝐱))2},\displaystyle{\operatorname{PlanarLoop}}(V)\coloneqq\{\mathbf{x}\in\operatorname{PL}_{0}^{\operatorname{cl}}(V)\,:\,\dim(\operatorname{span}(\mathbf{x}))\leq 2\}, (5.15)

where we use the notion of span from (2.31) using the minimal representative of 𝐱\mathbf{x}. We note that the span of any nontrivial planar PL loop must be two-dimensional. There is a natural inclusion

ι:PlanarLoop(V)PL0(V).\displaystyle\iota:{\operatorname{PlanarLoop}}(V)\hookrightarrow\operatorname{PL}_{0}(V). (5.16)
Definition 5.2.

A kite is a pair (𝐰,𝐛)(\mathbf{w},\mathbf{b}), consisting of a tail path 𝐰PL0(V)\mathbf{w}\in\operatorname{PL}_{0}(V) and a planar loop 𝐛PlanarLoop(V)\mathbf{b}\in{\operatorname{PlanarLoop}}(V). We define the set of kites by

Kite(V)PL0(V)×PlanarLoop(V).\displaystyle{\operatorname{Kite}}(V)\coloneqq\operatorname{PL}_{0}(V)\times{\operatorname{PlanarLoop}}(V). (5.17)

The pre-crossed module of piecewise linear surfaces is defined to be

PL~(V)(δ:PL~1(V)PL0(V),)wherePL~1(V)𝖥𝖦(Kite(V))/\displaystyle\widetilde{\operatorname{PL}}(V)\coloneqq(\delta:\widetilde{\operatorname{PL}}_{1}(V)\to\operatorname{PL}_{0}(V),\vartriangleright)\quad\text{where}\quad\widetilde{\operatorname{PL}}_{1}(V)\coloneqq\mathsf{FG}\big{(}{\operatorname{Kite}}(V)\big{)}/\sim (5.18)

is the quotient of the free group generated by kites subject to the following relations:

  1. (PL1.1)

    (𝐰1,𝐛1)(𝐰2,𝐛2)(𝐰1,𝐛1𝐮𝐛2𝐮1)(\mathbf{w}_{1},\mathbf{b}_{1})\star(\mathbf{w}_{2},\mathbf{b}_{2})\sim(\mathbf{w}_{1},\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}), where 𝐮=𝐰11𝐰2\mathbf{u}=\mathbf{w}_{1}^{-1}\star\mathbf{w}_{2}, if

    𝐛1𝐮𝐛2𝐮1PlanarLoop(V);\displaystyle\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}\in{\operatorname{PlanarLoop}}(V); (5.19)
    [Uncaptioned image]
  2. (PL1.2)

    (𝐰,0)1(\mathbf{w},\emptyset_{0})\sim\emptyset_{1}, where 0,1\emptyset_{0},\emptyset_{1} denote the empty words in PL0(V)\operatorname{PL}_{0}(V) and 𝖥𝖦(Kite(V))\mathsf{FG}\big{(}{\operatorname{Kite}}(V)\big{)} respectively; and

  3. (PL1.3)

    (𝐰,𝐱𝐛𝐱1)(𝐰𝐱,𝐛)(\mathbf{w},\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1})\sim(\mathbf{w}\star\mathbf{x},\mathbf{b}) for any 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V) such that 𝐱𝐛𝐱1PlanarLoop(V)\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}\in{\operatorname{PlanarLoop}}(V).

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 span(𝐮)span(𝐛1)=span(𝐛2)\operatorname{span}(\mathbf{u})\subseteq\operatorname{span}(\mathbf{b}_{1})=\operatorname{span}(\mathbf{b}_{2}). Furthermore, in Lemma B.7, we show that if (PL1.3) holds, then span(𝐱)span(𝐛)\operatorname{span}(\mathbf{x})\subseteq\operatorname{span}(\mathbf{b}).

With these relations, we note that (𝐰,𝐛)1=(𝐰,𝐛1)(\mathbf{w},\mathbf{b})^{-1}=(\mathbf{w},\mathbf{b}^{-1}), since

(𝐰,𝐛)(𝐰,𝐛1)(𝐰,𝐛𝐛1)=(𝐰,0)1,\displaystyle(\mathbf{w},\mathbf{b})\star(\mathbf{w},\mathbf{b}^{-1})\sim(\mathbf{w},\mathbf{b}\star\mathbf{b}^{-1})=(\mathbf{w},\emptyset_{0})\sim\emptyset_{1}, (5.20)

where here, 𝐮=𝐰1𝐰=0\mathbf{u}=\mathbf{w}^{-1}\star\mathbf{w}=\emptyset_{0}.

Remark 5.4.

As with the 1-dimensional case, we can represent elements of PL~1(V)\widetilde{\operatorname{PL}}_{1}(V) as words in Kite(V){\operatorname{Kite}}(V), without considering the formal inverses. In particular, we can make the equivalent definition of PL~1(V)\widetilde{\operatorname{PL}}_{1}(V) in terms of the free monoid, which we will use interchangeably,

PL~1(V)=𝖥𝖬𝗈𝗇(Kite(V))/.\displaystyle\widetilde{\operatorname{PL}}_{1}(V)=\mathsf{FMon}({\operatorname{Kite}}(V))/\sim. (5.21)

Next, we define the boundary map δ:PL~1(V)PL0(V)\delta:\widetilde{\operatorname{PL}}_{1}(V)\to\operatorname{PL}_{0}(V) on the generators by

δ(𝐰,𝐛)𝐰𝐛𝐰1,\displaystyle\delta(\mathbf{w},\mathbf{b})\coloneqq\mathbf{w}\star\mathbf{b}\star\mathbf{w}^{-1}, (5.22)

and the action of PL0(V)\operatorname{PL}_{0}(V) on the generators of PL~1(V)\widetilde{\operatorname{PL}}_{1}(V) by

𝐱(𝐰,𝐛)(𝐱𝐰,𝐛).\displaystyle\mathbf{x}\vartriangleright(\mathbf{w},\mathbf{b})\coloneqq(\mathbf{x}\star\mathbf{w},\mathbf{b}). (5.23)
Lemma 5.5.

The structure 𝐏𝐋~(V)=(δ:PL~1(V)PL0(V),)\widetilde{{\boldsymbol{\operatorname{PL}}}}(V)=(\delta:\widetilde{\operatorname{PL}}_{1}(V)\to\operatorname{PL}_{0}(V),\vartriangleright) is a pre-crossed module of groups.

In order to obtain a crossed module, we quotient by the standard Peiffer identity

δ(𝐰1,𝐛1)(𝐰2,𝐛2)Pf(𝐰1,𝐛1)(𝐰2,𝐛2)(𝐰1,𝐛11).\displaystyle\delta(\mathbf{w}_{1},\mathbf{b}_{1})\vartriangleright(\mathbf{w}_{2},\mathbf{b}_{2})\sim_{{\operatorname{Pf}}}(\mathbf{w}_{1},\mathbf{b}_{1})\star(\mathbf{w}_{2},\mathbf{b}_{2})\star(\mathbf{w}_{1},\mathbf{b}_{1}^{-1}). (5.24)
Definition 5.6.

The piecewise linear crossed module is defined by

𝐏𝐋(V)(δ:PL1(V)PL0(V),),wherePL1(V)PL~1(V)/Pf.\displaystyle{\boldsymbol{\operatorname{PL}}}(V)\coloneqq(\delta:\operatorname{PL}_{1}(V)\to\operatorname{PL}_{0}(V),\vartriangleright),\quad\text{where}\quad\operatorname{PL}_{1}(V)\coloneqq\widetilde{\operatorname{PL}}_{1}(V)/\sim_{{\operatorname{Pf}}}. (5.25)

There is a natural inclusion ξV,1:PlanarLoop(V)PL1(V)\xi_{V,1}:{\operatorname{PlanarLoop}}(V)\to\operatorname{PL}_{1}(V) defined by

ξV,1(𝐛)(0,𝐛).\displaystyle\xi_{V,1}(\mathbf{b})\coloneqq(\emptyset_{0},\mathbf{b}). (5.26)

In fact, for 2-dimensional vector spaces, this is an isomorphism.

Lemma 5.7.

If U𝖵𝖾𝖼𝗍U\in\mathsf{Vect} is 2-dimensional, then PlanarLoop(U)PL~1(U)PL1(U){\operatorname{PlanarLoop}}(U)\cong\widetilde{\operatorname{PL}}_{1}(U)\cong\operatorname{PL}_{1}(U) are isomorphic as groups.

Proof.

Since UU is two-dimensional, all loops are planar, and so PlanarLoop(U)=PL0cl(U){\operatorname{PlanarLoop}}(U)=\operatorname{PL}_{0}^{\operatorname{cl}}(U), which is a group. The map ξ~U,1:PlanarLoop(U)PL~1(U)\tilde{\xi}_{U,1}:{\operatorname{PlanarLoop}}(U)\to\widetilde{\operatorname{PL}}_{1}(U) is a group homomorphism by (PL1.1). In fact, it is an isomorphism with inverse given by the boundary map δ:PL~1(U)PL0cl(U)\delta:\widetilde{\operatorname{PL}}_{1}(U)\to\operatorname{PL}_{0}^{\operatorname{cl}}(U). Indeed, the identity δξ~U,1=id\delta\circ\tilde{\xi}_{U,1}=\operatorname{id} is immediate, and ξ~U,1δ=id\tilde{\xi}_{U,1}\circ\delta=\operatorname{id} follows by (PL1.3). Furthermore, the Peiffer identity (5.24) is already implied because it holds in PL0cl(U)\operatorname{PL}_{0}^{\operatorname{cl}}(U). ∎

For V𝖵𝖾𝖼𝗍V\in\mathsf{Vect} where dim(V)3\dim(V)\geq 3, PlanarLoop(V){\operatorname{PlanarLoop}}(V) is not a group. However, by Lemma 5.7, the restriction to a 2-dimensional subspace UVU\subset V is a group PlanarLoop(U)PL1(U){\operatorname{PlanarLoop}}(U)\cong\operatorname{PL}_{1}(U). For the remainder of the article, we will make the identification of crossed modules

(ι:PlanarLoop(U)PL0(U),)𝐏𝐋(U)=(δ:PL1(U)PL0(U),).\displaystyle(\iota:{\operatorname{PlanarLoop}}(U)\to\operatorname{PL}_{0}(U),\vartriangleright)\cong{\boldsymbol{\operatorname{PL}}}(U)=(\delta:\operatorname{PL}_{1}(U)\to\operatorname{PL}_{0}(U),\vartriangleright). (5.27)

Furthermore, the restriction of the inclusion ξV,1\xi_{V,1} in (5.26) yields a morphism of crossed modules

ξVU=(ξV,1U,ξV,0U):𝐏𝐋(U)𝐏𝐋(V),\displaystyle\xi^{U}_{V}=(\xi^{U}_{V,1},\xi^{U}_{V,0}):{\boldsymbol{\operatorname{PL}}}(U)\to{\boldsymbol{\operatorname{PL}}}(V), (5.28)

where ξV,0U:PL0(U)PL0(V)\xi^{U}_{V,0}:\operatorname{PL}_{0}(U)\to\operatorname{PL}_{0}(V) is the inclusion induced by UVU\hookrightarrow V. This leads to the following universal property.

Theorem 5.8.

Let VV be a vector space, and suppose

𝐆=(δ:GPL0(V),)\displaystyle\mathbf{G}=(\delta:G\to\operatorname{PL}_{0}(V),\vartriangleright) (5.29)

is a crossed module of groups. Let f:PlanarLoop(V)Gf:{\operatorname{PlanarLoop}}(V)\to G be a function such that f(0)=1f(\emptyset_{0})=1 and for any 2-dimensional subspace UVU\subset V, the restriction of ff to UU yields a morphism of crossed modules,

(f|U,ξV,0U):𝐏𝐋(U)𝐆,\displaystyle(f|_{U},\xi^{U}_{V,0}):{\boldsymbol{\operatorname{PL}}}(U)\to\mathbf{G}, (5.30)

where ξV,0U:PL0(U)PL0(V)\xi^{U}_{V,0}:\operatorname{PL}_{0}(U)\hookrightarrow\operatorname{PL}_{0}(V) is the inclusion from (5.28). Then, there exists a unique group homomorphism F:PL1(V)GF:\operatorname{PL}_{1}(V)\to G such that

(F,id):𝐏𝐋(V)𝐆\displaystyle(F,\operatorname{id}):{\boldsymbol{\operatorname{PL}}}(V)\to\mathbf{G} (5.31)

is a morphism of crossed modules and FξV,1=fF\circ\xi_{V,1}=f.

Remark 5.9.

Note that if dim(V)2\dim(V)\leq 2 then F=fF=f and there is nothing to prove.

Proof.

First, we define a map F~:𝖥𝖦(Kite(V))G\tilde{F}:\mathsf{FG}({\operatorname{Kite}}(V))\to G on the generators by

F~(𝐰,𝐛)=𝐰f(𝐛).\displaystyle\tilde{F}(\mathbf{w},\mathbf{b})=\mathbf{w}\vartriangleright f(\mathbf{b}). (5.32)

We verify that this descends to PL~1(V)\widetilde{\operatorname{PL}}_{1}(V) by checking the relations (PL1.1)-(PL1.3). For (PL1.2), we have

F~(𝐰,0)=𝐰f(0)=1\displaystyle\tilde{F}(\mathbf{w},\emptyset_{0})=\mathbf{w}\vartriangleright f(\emptyset_{0})=1 (5.33)

by assumption. Next, for (PL1.3), suppose 𝐰PL0(V)\mathbf{w}\in\operatorname{PL}_{0}(V), 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V) and 𝐛PlanarLoop(U)\mathbf{b}\in{\operatorname{PlanarLoop}}(U) is a nontrivial planar loop with span UU such that 𝐱𝐛𝐱1PlanarLoop(V)\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}\in{\operatorname{PlanarLoop}}(V). By Lemma B.7, 𝐱PL0(U)\mathbf{x}\in\operatorname{PL}_{0}(U). Then, since ff restricts to a morphism of crossed modules on UU,

F~(𝐰,𝐱𝐛𝐱1)=𝐰f(𝐱𝐛)=𝐰(𝐱f(𝐛))=(𝐰𝐱)f(𝐛)=F~(𝐰𝐱,𝐛).\displaystyle\tilde{F}(\mathbf{w},\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1})=\mathbf{w}\vartriangleright f(\mathbf{x}\vartriangleright\mathbf{b})=\mathbf{w}\vartriangleright(\mathbf{x}\vartriangleright f(\mathbf{b}))=(\mathbf{w}\star\mathbf{x})\vartriangleright f(\mathbf{b})=\tilde{F}(\mathbf{w}\star\mathbf{x},\mathbf{b}). (5.34)

Finally, we verify (PL1.1). Suppose (𝐰1,𝐛1),(𝐰2,𝐛2)Kite(V)(\mathbf{w}_{1},\mathbf{b}_{1}),(\mathbf{w}_{2},\mathbf{b}_{2})\in{\operatorname{Kite}}(V) are nontrivial kites such that 𝐛1𝐮𝐛2𝐮1PlanarLoop(V)\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}\in{\operatorname{PlanarLoop}}(V), where 𝐮=𝐰11𝐰2\mathbf{u}=\mathbf{w}_{1}^{-1}\star\mathbf{w}_{2}. By Corollary B.10, there exists UVU\subset V with dim(U)=2\dim(U)=2 such that 𝐛1,𝐛2,𝐮PL0(U)\mathbf{b}_{1},\mathbf{b}_{2},\mathbf{u}\in\operatorname{PL}_{0}(U). Then, using 𝐰2=𝐰1𝐮\mathbf{w}_{2}=\mathbf{w}_{1}\star\mathbf{u}, equation  (5.34), and the assumption that ff restricts to a homomorphism on PL0cl(U)\operatorname{PL}_{0}^{\operatorname{cl}}(U), we have

F~(𝐰1,𝐛1)F~(𝐰2,𝐛2)\displaystyle\tilde{F}(\mathbf{w}_{1},\mathbf{b}_{1})\cdot\tilde{F}(\mathbf{w}_{2},\mathbf{b}_{2}) =F~(𝐰1,𝐛1)F~(𝐰1𝐮,𝐛2)\displaystyle=\tilde{F}(\mathbf{w}_{1},\mathbf{b}_{1})\cdot\tilde{F}(\mathbf{w}_{1}\star\mathbf{u},\mathbf{b}_{2}) (5.35)
=(𝐰1f(𝐛1))(𝐰1f(𝐮𝐛2𝐮1))\displaystyle=(\mathbf{w}_{1}\vartriangleright f(\mathbf{b}_{1}))\cdot(\mathbf{w}_{1}\vartriangleright f(\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1})) (5.36)
=𝐰1f(𝐛1𝐮𝐛2𝐮1)\displaystyle=\mathbf{w}_{1}\vartriangleright f(\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}) (5.37)
=F~(𝐰1,𝐛1𝐮𝐛2𝐮1).\displaystyle=\tilde{F}(\mathbf{w}_{1},\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}). (5.38)

Therefore, the map F~\tilde{F} descends to a map F~:PL~1(V)G\tilde{F}:\widetilde{\operatorname{PL}}_{1}(V)\to G, which is the unique morphism of pre-crossed modules which satisfies F~ξ~=f\tilde{F}\circ\tilde{\xi}=f, where ξ~:PlanarLoop(V)PL~1(V)\tilde{\xi}:{\operatorname{PlanarLoop}}(V)\to\widetilde{\operatorname{PL}}_{1}(V) is defined by ξ~(𝐛)=(0,𝐛)\tilde{\xi}(\mathbf{b})=(\emptyset_{0},\mathbf{b}). The final step is applying the crossed module functor by quotienting out the Peiffer identity to obtain the desired unique morphism F:PL1(V)GF:\operatorname{PL}_{1}(V)\to G. ∎

We can also enhance this universal property by varying the base.

Corollary 5.10.

Let VV be a vector space and suppose 𝐆=(δ:G1G0,)\mathbf{G}=(\delta:G_{1}\to G_{0},\vartriangleright) is a crossed module of groups. Let f0:PL0(V)G0f_{0}:\operatorname{PL}_{0}(V)\to G_{0} be a group homomorphism and let f1:PlanarLoop(V)G1f_{1}:{\operatorname{PlanarLoop}}(V)\to G_{1} be a function such that f1(0)=1f_{1}(\emptyset_{0})=1 and for any 2-dimensional subspace UVU\subset V, the restriction of (f1,f0)(f_{1},f_{0}) to UU is a morphism of crossed modules,

(fU,1,fU,0):𝐏𝐋(U)𝐆.\displaystyle(f_{U,1},f_{U,0}):{\boldsymbol{\operatorname{PL}}}(U)\to\mathbf{G}. (5.39)

Then, there exists a unique group homomorphism F1:PL1(V)G1F_{1}:\operatorname{PL}_{1}(V)\to G_{1} such that

(F1,f0):𝐏𝐋(V)𝐆\displaystyle(F_{1},f_{0}):{\boldsymbol{\operatorname{PL}}}(V)\to\mathbf{G} (5.40)

is a morphism of crossed modules and F1ξV,1=f1F_{1}\circ\xi_{V,1}=f_{1}.

Proof.

We consider the pullback crossed module of 𝐆\mathbf{G} with respect to f0f_{0}, defined by

f0𝐆(δ:f0G1PL0(V),),f0G1{(𝐰,g)PL0(V)×G1:f0(𝐰)=δ(g)}.\displaystyle f_{0}^{*}\mathbf{G}\coloneqq(\delta:f_{0}^{*}G_{1}\to\operatorname{PL}_{0}(V),\vartriangleright),\quad f_{0}^{*}G_{1}\coloneqq\{(\mathbf{w},g)\in\operatorname{PL}_{0}(V)\times G_{1}\,:\,f_{0}(\mathbf{w})=\delta(g)\}. (5.41)

It is equipped with a morphism of crossed modules p=(p1,f0):f0𝐆𝐆p=(p_{1},f_{0}):f_{0}^{*}\mathbf{G}\to\mathbf{G}. There is a well-defined function f~:PlanarLoop(V)f0G1\tilde{f}:{\operatorname{PlanarLoop}}(V)\to f_{0}^{*}G_{1} given by sending a planar loop 𝐛\mathbf{b} to f~(𝐛)=(ι(𝐛),f1(𝐛))\tilde{f}(\mathbf{b})=(\iota(\mathbf{b}),f_{1}(\mathbf{b})). To see that this is well-defined, note first that f~(0)=(0,1)\tilde{f}(\emptyset_{0})=(\emptyset_{0},1). Furthermore, if 𝐛PlanarLoop(V)\mathbf{b}\in{\operatorname{PlanarLoop}}(V) is non-trivial, it determines a 22-dimensional subspace UU such that 𝐛PL0cl(U)\mathbf{b}\in\operatorname{PL}_{0}^{\operatorname{cl}}(U). Then, since (f1,f0)(f_{1},f_{0}) restricts to a morphism of crossed modules on 𝐏𝐋(U){\boldsymbol{\operatorname{PL}}}(U), we see that f0(ι(𝐛))=δ(f1(𝐛))f_{0}(\iota(\mathbf{b}))=\delta(f_{1}(\mathbf{b})). Note also that p1f~=f1p_{1}\circ\tilde{f}=f_{1}.

We wish to apply the universal property of  Theorem 5.8, and for this we must verify that for every 22-dimensional subspace UU, the restriction f~|U\tilde{f}|_{U} defines a morphism of crossed modules 𝐏𝐋(U)f0𝐆{\boldsymbol{\operatorname{PL}}}(U)\to f_{0}^{*}\mathbf{G}. Now, because (fU,1,fU,0)(f_{U,1},f_{U,0}) 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 f0𝐆f_{0}^{*}\mathbf{G}

(fU,1,fU,0):𝐏𝐋(U)(f~U,1,ξV,0U)f0𝐆𝑝𝐆.\displaystyle(f_{U,1},f_{U,0}):{\boldsymbol{\operatorname{PL}}}(U)\xrightarrow{(\tilde{f}_{U,1},\xi_{V,0}^{U})}f_{0}^{*}\mathbf{G}\xrightarrow{p}\mathbf{G}. (5.42)

But a simple calculation shows that f~|U=f~U,1\tilde{f}|_{U}=\tilde{f}_{U,1}, thus verifying the condition. As a result, by the universal property, there is a unique homomorphism F~:PL1(V)f0G1\tilde{F}:\operatorname{PL}_{1}(V)\to f_{0}^{*}G_{1} such that

(F~,id):𝐏𝐋(V)f0𝐆\displaystyle(\tilde{F},\operatorname{id}):{\boldsymbol{\operatorname{PL}}}(V)\to f_{0}^{*}\mathbf{G} (5.43)

is a homomorphism of crossed modules and F~ξV,1=f~\tilde{F}\circ\xi_{V,1}=\tilde{f}. Finally, we define F1=p1F~F_{1}=p_{1}\circ\tilde{F} to obtain our desired morphism (F1,f0)(F_{1},f_{0}).

Corollary 5.11.

The piecewise linear crossed module defines a functor

𝐏𝐋:𝖵𝖾𝖼𝗍𝖷𝖦𝗋𝗉.\displaystyle{\boldsymbol{\operatorname{PL}}}:\mathsf{Vect}\to\mathsf{XGrp}. (5.44)
Proof.

Let ϕ:VW\phi:V\to W be a linear map. Then, by Corollary 2.14, there exists a homomorphism PL0(ϕ):PL0(V)PL0(W)\operatorname{PL}_{0}(\phi):\operatorname{PL}_{0}(V)\to\operatorname{PL}_{0}(W). By  Corollary 2.15, PL0(ϕ)\operatorname{PL}_{0}(\phi) preserves the span of paths. Hence, if a path 𝐱\mathbf{x} is planar, then so is PL0(ϕ)(𝐱)\operatorname{PL}_{0}(\phi)(\mathbf{x}). As a result, we can also define a map fϕ,1:PlanarLoop(V)PL1(W)f_{\phi,1}:{\operatorname{PlanarLoop}}(V)\to\operatorname{PL}_{1}(W) by

fϕ,1:PlanarLoop(V)PL0(ϕ)PlanarLoop(W)ξW,1PL1(W).\displaystyle f_{\phi,1}:{\operatorname{PlanarLoop}}(V)\xrightarrow{\operatorname{PL}_{0}(\phi)}{\operatorname{PlanarLoop}}(W)\xrightarrow{\xi_{W,1}}\operatorname{PL}_{1}(W). (5.45)

Given a 2-dimensional subspace UVU\subset V, the restriction of the pair (fϕ,1,PL0(ϕ))(f_{\phi,1},\operatorname{PL}_{0}(\phi)) to UU is a morphism of crossed modules 𝐏𝐋(U)𝐏𝐋(W){\boldsymbol{\operatorname{PL}}}(U)\to{\boldsymbol{\operatorname{PL}}}(W). Hence, by Corollary 5.10, there is a unique homomorphism PL1(ϕ):PL1(V)PL1(W)\operatorname{PL}_{1}(\phi):\operatorname{PL}_{1}(V)\to\operatorname{PL}_{1}(W) such that

𝐏𝐋(ϕ)=(PL1(ϕ),PL0(ϕ)):𝐏𝐋(V)𝐏𝐋(W)\displaystyle{\boldsymbol{\operatorname{PL}}}(\phi)=(\operatorname{PL}_{1}(\phi),\operatorname{PL}_{0}(\phi)):{\boldsymbol{\operatorname{PL}}}(V)\to{\boldsymbol{\operatorname{PL}}}(W) (5.46)

is a morphism of crossed modules and such that PL1(ϕ)ξV,1=ξW,1PL0(ϕ)|PlanarLoop(V)\operatorname{PL}_{1}(\phi)\circ\xi_{V,1}=\xi_{W,1}\circ\operatorname{PL}_{0}(\phi)|_{{\operatorname{PlanarLoop}}(V)}. The component PL0(ϕ)\operatorname{PL}_{0}(\phi) is functorial by  Corollary 2.14, and the component PL1(ϕ)\operatorname{PL}_{1}(\phi) is functorial because of the uniqueness. ∎

Remark 5.12.

Let ϕ:UV\phi:U\hookrightarrow V be the inclusion of a 2-dimensional subspace into VV. Then the restriction of ξV,1U\xi_{V,1}^{U} from (5.28) is PL1(ϕ)\operatorname{PL}_{1}(\phi).

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 𝗙=(𝖥1,𝖥0):𝖵𝖾𝖼𝗍𝖷𝖦𝗋𝗉\boldsymbol{\mathsf{F}}=(\mathsf{F}_{1},\mathsf{F}_{0}):\mathsf{Vect}\to\mathsf{XGrp} is planar if

  • 𝗙(U)\boldsymbol{\mathsf{F}}(U) is trivial when dim(U)=0\dim(U)=0,

  • 𝖥1(U)\mathsf{F}_{1}(U) is trivial when dim(U)=1\dim(U)=1, and

  • δU𝖥:𝖥1(U)𝖥0(U)\delta^{\mathsf{F}}_{U}:\mathsf{F}_{1}(U)\to\mathsf{F}_{0}(U) is injective when dim(U)=2\dim(U)=2.

Lemma 5.14.

The piecewise linear crossed module 𝐏𝐋:𝖵𝖾𝖼𝗍𝖷𝖬𝗈𝖽{\boldsymbol{\operatorname{PL}}}:\mathsf{Vect}\to\mathsf{XMod} is a planar functor.

Proof.

The cases of dim(U)=0,1\dim(U)=0,1 are immediate, and the condition for dim(U)=2\dim(U)=2 is given by Lemma 5.7. ∎

The following lemma is the main result used in this section.

Lemma 5.15.

Let 𝗙=(𝖥1,𝖥0):𝖵𝖾𝖼𝗍𝖷𝖦𝗋𝗉\boldsymbol{\mathsf{F}}=(\mathsf{F}_{1},\mathsf{F}_{0}):\mathsf{Vect}\to\mathsf{XGrp} be a planar functor and let α0:PL0F0\alpha_{0}:\operatorname{PL}_{0}\Rightarrow F_{0} be a natural transformation with the property that αU,0(PL0cl(U))im(δU𝖥)\alpha_{U,0}(\operatorname{PL}_{0}^{\operatorname{cl}}(U))\subseteq\operatorname{im}(\delta^{\mathsf{F}}_{U}) when dim(U)=2\dim(U)=2. Then, there exists a unique natural transformation which extends α0\alpha_{0},

α=(α1,α0):𝐏𝐋𝗙.\displaystyle\alpha=(\alpha_{1},\alpha_{0}):{\boldsymbol{\operatorname{PL}}}\Rightarrow\boldsymbol{\mathsf{F}}. (5.47)
Proof.

Let 𝖵𝖾𝖼𝗍2\mathsf{Vect}_{\leq 2} denote the subcategory of vector spaces of dimension at most 22. We will first construct α\alpha on this subcategory. When dim(U)1\dim(U)\leq 1, the components αU,1\alpha_{U,1} are uniquely determined because PL1(U)\operatorname{PL}_{1}(U) is trivial. Now assume that dim(U)=2.\dim(U)=2. In this case, the component αU,1\alpha_{U,1} must make the following diagram commute

PL1(U){\operatorname{PL}_{1}(U)}𝖥1(U){\mathsf{F}_{1}(U)}PL0(U){\operatorname{PL}_{0}(U)}𝖥0(U).{\mathsf{F}_{0}(U).}αU,1\scriptstyle{\alpha_{U,1}}δUPL\scriptstyle{\delta^{\operatorname{PL}}_{U}}δU𝖥\scriptstyle{\delta^{\mathsf{F}}_{U}}αU,0\scriptstyle{\alpha_{U,0}} (5.48)

The image of the map αU,0δUPL\alpha_{U,0}\circ\delta^{\operatorname{PL}}_{U} is contained in the image of the injective map δU𝖥\delta^{\mathsf{F}}_{U}. Hence, the desired map αU,1\alpha_{U,1} exists and it is unique. The pair αU=(αU,1,αU,0)\alpha_{U}=(\alpha_{U,1},\alpha_{U,0}) is automatically a map of crossed modules because δUPL\delta^{\operatorname{PL}}_{U} and δU𝖥\delta^{\mathsf{F}}_{U} are injective.

To check naturality, let ϕ:UV\phi:U\to V be a linear map. We need to verify that 𝖥1(ϕ)αU,1=αV,1PL1(ϕ)\mathsf{F}_{1}(\phi)\circ\alpha_{U,1}=\alpha_{V,1}\circ\operatorname{PL}_{1}(\phi). Because 𝐏𝐋{\boldsymbol{\operatorname{PL}}} and 𝗙\boldsymbol{\mathsf{F}} are planar functors, this is trivially satisfied when either UU or VV has dimension less than 22. Hence, we assume that dim(U)=dim(V)=2\dim(U)=\dim(V)=2. Because α0:PL0𝖥0\alpha_{0}:\operatorname{PL}_{0}\Rightarrow\mathsf{F}_{0} is a natural transformation, we have F0(ϕ)αU,0=αV,0PL0(ϕ)F_{0}(\phi)\circ\alpha_{U,0}=\alpha_{V,0}\circ\operatorname{PL}_{0}(\phi). Then, using the fact that 𝐏𝐋(ϕ),𝗙(ϕ),αU{\boldsymbol{\operatorname{PL}}}(\phi),\boldsymbol{\mathsf{F}}(\phi),\alpha_{U} and αV\alpha_{V} are morphisms of crossed modules, we have

δV𝖥(𝖥1(ϕ)αU,1)=(𝖥0(ϕ)αU,0)δUPL=(αV,0PL0(ϕ))δUPL=δV𝖥(αV,1PL1(ϕ)).\displaystyle\delta^{\mathsf{F}}_{V}\circ(\mathsf{F}_{1}(\phi)\circ\alpha_{U,1})=(\mathsf{F}_{0}(\phi)\circ\alpha_{U,0})\circ\delta^{\operatorname{PL}}_{U}=(\alpha_{V,0}\circ\operatorname{PL}_{0}(\phi))\circ\delta^{\operatorname{PL}}_{U}=\delta^{\mathsf{F}}_{V}\circ(\alpha_{V,1}\circ\operatorname{PL}_{1}(\phi)). (5.49)

Hence, by injectivity of δV𝖥\delta^{\mathsf{F}}_{V}, we have 𝖥1(ϕ)αU,1=αV,1PL1(ϕ)\mathsf{F}_{1}(\phi)\circ\alpha_{U,1}=\alpha_{V,1}\circ\operatorname{PL}_{1}(\phi).

Next, we extend α\alpha to vector spaces VV of dimension at least 33. We will do this by applying the universal property of Corollary 5.10. Given the inclusion ιU:UV\iota_{U}:U\hookrightarrow V of a 2-dimensional subspace UU, we define the following morphism of crossed modules

sU=(sU,1,sU,0):𝐏𝐋(U)αU𝗙(U)𝗙(ιU)=ξVU𝗙(V).\displaystyle s_{U}=(s_{U,1},s_{U,0}):{\boldsymbol{\operatorname{PL}}}(U)\xrightarrow{\alpha_{U}}\boldsymbol{\mathsf{F}}(U)\xrightarrow{\boldsymbol{\mathsf{F}}(\iota_{U})=\xi^{U}_{V}}\boldsymbol{\mathsf{F}}(V). (5.50)

We assemble the sU,1s_{U,1} over all 2-dimensional subspaces UVU\subset V to obtain a map s1:PlanarLoop(V)𝖥1(V)s_{1}:{\operatorname{PlanarLoop}}(V)\to\mathsf{F}_{1}(V). Along with the homomorphism αV,0:PL0(V)𝖥0(V)\alpha_{V,0}:\operatorname{PL}_{0}(V)\to\mathsf{F}_{0}(V), this map satisfies the hypotheses of Corollary 5.10. Hence we obtain a unique homomorphism αV,1:PL1(V)𝖥1(V)\alpha_{V,1}:\operatorname{PL}_{1}(V)\to\mathsf{F}_{1}(V) such that

αV=(αV,1,αV,0):𝐏𝐋(V)𝗙(V)\displaystyle\alpha_{V}=(\alpha_{V,1},\alpha_{V,0}):{\boldsymbol{\operatorname{PL}}}(V)\to\boldsymbol{\mathsf{F}}(V) (5.51)

is a morphism of crossed modules and such that αV,1ξV,1=s1\alpha_{V,1}\circ\xi_{V,1}=s_{1}. In particular, αV𝐏𝐋(ιU)=𝗙(ιU)αU\alpha_{V}\circ{\boldsymbol{\operatorname{PL}}}(\iota_{U})=\boldsymbol{\mathsf{F}}(\iota_{U})\circ\alpha_{U}, meaning that α\alpha is natural with respect to inclusions of 22-dimensional subspaces. Since this property characterizes αV,1\alpha_{V,1}, we see that any natural transformation extending α0\alpha_{0} must be unique. Furthermore, it is straightforward to see that α\alpha is natural with respect to linear maps ϕ:UV\phi:U\to V where dim(U)2\dim(U)\leq 2. Indeed, this is immediate if dim(U)1\dim(U)\leq 1 because then PL1(U)\operatorname{PL}_{1}(U) is trivial. Assuming that dim(U)=2\dim(U)=2, factor ϕ\phi as

ϕ:U𝜓WιWV,\displaystyle\phi:U\xrightarrow{\psi}W\xrightarrow{\iota_{W}}V, (5.52)

where W=im(ϕ)W=\operatorname{im}(\phi). Therefore,

αV,1PL1(ιW)PL1(ψ)=𝖥1(ιW)αW,1PL1(ψ)=𝖥1(ιW)𝖥1(ψ)αU,1,\displaystyle\ \alpha_{V,1}\circ\operatorname{PL}_{1}(\iota_{W})\circ\operatorname{PL}_{1}(\psi)=\mathsf{F}_{1}(\iota_{W})\circ\alpha_{W,1}\circ\operatorname{PL}_{1}(\psi)=\mathsf{F}_{1}(\iota_{W})\circ\mathsf{F}_{1}(\psi)\circ\alpha_{U,1}, (5.53)

where the first equality follows from naturality with respect to inclusions and the second equality follows from naturality in 𝖵𝖾𝖼𝗍2\mathsf{Vect}_{\leq 2}.

The final step in the proof is to verify naturality of α\alpha for arbitrary linear maps ϕ:VW\phi:V\to W. We may assume that dim(V)3\dim(V)\geq 3. For each 2-dimensional subspace UVU\subset V, let ιU:UV\iota_{U}:U\hookrightarrow V be the inclusion and consider the following diagram

𝐏𝐋(U){{\boldsymbol{\operatorname{PL}}}(U)}𝗙(U){\boldsymbol{\mathsf{F}}(U)}𝐏𝐋(V){{\boldsymbol{\operatorname{PL}}}(V)}𝗙(V){\boldsymbol{\mathsf{F}}(V)}𝐏𝐋(W){{\boldsymbol{\operatorname{PL}}}(W)}𝗙(W).{\boldsymbol{\mathsf{F}}(W).}αU\scriptstyle{\alpha_{U}}𝐏𝐋(ιU)=ξVU\scriptstyle{{\boldsymbol{\operatorname{PL}}}(\iota_{U})=\xi^{U}_{V}}𝗙(ιU)\scriptstyle{\boldsymbol{\mathsf{F}}(\iota_{U})}αV\scriptstyle{\alpha_{V}}𝐏𝐋(ϕ)\scriptstyle{{\boldsymbol{\operatorname{PL}}}(\phi)}𝗙(ϕ)\scriptstyle{\boldsymbol{\mathsf{F}}(\phi)}αW\scriptstyle{\alpha_{W}} (5.54)

We know that the top and outer squares commute, and we wish to show that the bottom square commutes. For each UVU\subset V, consider the morphism of crossed modules

pU=(pU,1,pU,0):𝐏𝐋(U)αU𝗙(U)𝗙(ϕιU)𝗙(W).\displaystyle p_{U}=(p_{U,1},p_{U,0}):{\boldsymbol{\operatorname{PL}}}(U)\xrightarrow{\alpha_{U}}\boldsymbol{\mathsf{F}}(U)\xrightarrow{\boldsymbol{\mathsf{F}}(\phi\circ\iota_{U})}\boldsymbol{\mathsf{F}}(W). (5.55)

We assemble the pU,1p_{U,1} over all 2-dimensional subspaces into a function p1:PlanarLoop(V)𝖥1(W)p_{1}:{\operatorname{PlanarLoop}}(V)\to\mathsf{F}_{1}(W). Along with the homomorphism 𝖥0(ϕ)αV,0:PL0(V)𝖥0(W)\mathsf{F}_{0}(\phi)\circ\alpha_{V,0}:\operatorname{PL}_{0}(V)\to\mathsf{F}_{0}(W), this map satisfies the hypotheses of Corollary 5.10. Hence, there is a unique homomorphism P:PL1(V)𝖥1(W)P:\operatorname{PL}_{1}(V)\to\mathsf{F}_{1}(W) such that (P,𝖥0(ϕ)αV,0)(P,\mathsf{F}_{0}(\phi)\circ\alpha_{V,0}) is a morphism of crossed modules, and PξV,1=p1P\circ\xi_{V,1}=p_{1}. Hence, because both 𝖥1(ϕ)αV,1\mathsf{F}_{1}(\phi)\circ\alpha_{V,1} and αW,1PL1(ϕ)\alpha_{W,1}\circ\operatorname{PL}_{1}(\phi) 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 𝐗,𝐘C1([0,1]2,2)\mathbf{X},\mathbf{Y}\in C^{1}([0,1]^{2},\mathbb{R}^{2}) be a pair of maps with equal corners 𝐗i,j=𝐘i,j=ci,j\mathbf{X}_{i,j}=\mathbf{Y}_{i,j}=c_{i,j} for i,j={0,1}i,j=\{0,1\} and such that each boundary path (3.2) is thin homotopy equivalent i𝐗thi𝐘\partial_{i}\mathbf{X}\sim_{\operatorname{th}}\partial_{i}\mathbf{Y} for i{l,b,r,t}i\in\{l,b,r,t\}. Then 𝐗th𝐘\mathbf{X}\sim_{\operatorname{th}}\mathbf{Y}.

Proof.

Let hi:[0,1]22h_{i}:[0,1]^{2}\to\mathbb{R}^{2} by the thin homotopy between i𝐗\partial_{i}\mathbf{X} and i𝐘\partial_{i}\mathbf{Y} for i{l,b,r,t}i\in\{l,b,r,t\}. Then 𝐗\mathbf{X} is thin homotopy equivalent to the surface 𝐗~\tilde{\mathbf{X}}, where we glue the four thin path homotopies hih_{i} along the boundary as follows.

[Uncaptioned image]

Then, the linear homotopy between 𝐗~\tilde{\mathbf{X}} and 𝐘\mathbf{Y} is thin as it must have rank at most 22 and does not change the boundaries. Hence 𝐗~th𝐘\tilde{\mathbf{X}}\sim_{\operatorname{th}}\mathbf{Y}. ∎

Lemma 5.17.

For every planar loop 𝐛=(b1,,bm)minPL0cl(U)PlanarLoop(V)\mathbf{b}=(b_{1},\ldots,b_{m})_{\min}\in\operatorname{PL}_{0}^{\operatorname{cl}}(U)\subset{\operatorname{PlanarLoop}}(V), there exists a surface 𝐗𝐛C1([0,1]2,U)\mathbf{X}^{\mathbf{b}}\in C^{1}([0,1]^{2},U) whose boundary is in the thin homotopy class of R0(𝐛)R_{0}(\mathbf{b}).

[Uncaptioned image]
Proof.

For u,vUu,v\in U, define the function

𝐂u,v:[0,1]2Uby𝐂s,tu,vψ(t)(u+ψ(s)(vu)),\displaystyle\mathbf{C}^{u,v}:[0,1]^{2}\to U\quad\text{by}\quad\mathbf{C}^{u,v}_{s,t}\coloneqq\psi(t)\cdot(u+\psi(s)\cdot(v-u)), (5.56)

where ψ:[0,1][0,1]\psi:[0,1]\to[0,1] is a reparametrization with sitting instants. Define 𝐗𝐛C1([0,1]2,U)\mathbf{X}^{\mathbf{b}}\in C^{1}([0,1]^{2},U) by

𝐗𝐛=𝐂0,b^m1h𝐂b^m1,b^m2hh𝐂b^2,b^1h𝐂b^1,0,\displaystyle\mathbf{X}^{\mathbf{b}}=\mathbf{C}^{0,\hat{b}_{m-1}}\star_{h}\mathbf{C}^{\hat{b}_{m-1},\hat{b}_{m-2}}\star_{h}\ldots\star_{h}\mathbf{C}^{\hat{b}_{2},\hat{b}_{1}}\star_{h}\mathbf{C}^{\hat{b}_{1},0}, (5.57)

where h\star_{h} is defined in (3.8), and the concatenations are defined from left to right (as h\star_{h} is not associative). Note that the left, bottom, and right boundaries of 𝐗𝐛\mathbf{X}^{\mathbf{b}} are trivial, and it can be verified that 𝐗𝐛\mathbf{X}^{\mathbf{b}} has the correct boundary. ∎

Proposition 5.18.

There exists a unique natural transformation

𝐑=(R1,R0):𝐏𝐋𝝉,\displaystyle\mathbf{R}=(R_{1},R_{0}):{\boldsymbol{\operatorname{PL}}}\Rightarrow{\boldsymbol{\tau}}, (5.58)

called realization, which extends R0:PL0τ1R_{0}:\operatorname{PL}_{0}\Rightarrow\tau_{1} from (2.34).

Proof.

First, 𝝉(U){\boldsymbol{\tau}}(U) is trivial when dim(U)=0\dim(U)=0, and τ2(U)\tau_{2}(U) is trivial when dim(U)=1\dim(U)=1, as every surface in UU is thinly null homotopic by the linear homotopy. By Lemma 5.16, the crossed module boundary δ:τ2(U)τ1(U)\delta:\tau_{2}(U)\to\tau_{1}(U) is injective when dim(U)=2\dim(U)=2. Therefore, 𝝉{\boldsymbol{\tau}} is a planar functor. Furthermore, R0(PL0cl(U))im(:τ2(U)τ1(U))R_{0}(\operatorname{PL}_{0}^{\operatorname{cl}}(U))\subseteq\operatorname{im}(\partial:\tau_{2}(U)\to\tau_{1}(U)) when dim(U)=2\dim(U)=2 as every loop in UU can be filled in to obtain a surface by Lemma 5.17. Thus, by Lemma 5.15, there exists a unique extension 𝐑=(R1,R0)\mathbf{R}=(R_{1},R_{0}). ∎

Next, we will show that the path signature also extends uniquely to a natural transformation.

Theorem 5.19.

There exists a unique natural transformation

𝐒PL=(SPL,1,SPL,0):𝐏𝐋𝐊^,\displaystyle\mathbf{S}_{\operatorname{PL}}=(S_{\operatorname{PL},1},S_{\operatorname{PL},0}):{\boldsymbol{\operatorname{PL}}}\Rightarrow\hat{\mathbf{K}}, (5.59)

called the piecewise linear signature, which extends SPL,0:PL0K^0S_{\operatorname{PL},0}:\operatorname{PL}_{0}\Rightarrow\hat{K}_{0} from Proposition 2.17.

Proof.

First, 𝖐(U){\boldsymbol{\mathfrak{k}}}(U) is trivial when dim(U)=0\dim(U)=0, and 𝔨1(U)\mathfrak{k}_{1}(U) is trivial when dim(U)=1\dim(U)=1 since Λ2U=0\Lambda^{2}U=0. Now, consider a 2-dimensional vector space UU. By Lemma 4.5, we note that the kernel ker(δ:𝔨1(U)𝔨0(U))\ker(\delta:\mathfrak{k}_{1}(U)\to\mathfrak{k}_{0}(U)) is trivial. Thus

𝔨1(U)[𝔨0(U),𝔨0(U)]=LCS2(𝔨0(U))\displaystyle\mathfrak{k}_{1}(U)\cong[\mathfrak{k}_{0}(U),\mathfrak{k}_{0}(U)]={\operatorname{LCS}}_{2}(\mathfrak{k}_{0}(U)) (5.60)

as Lie algebras and the boundary map on groups δ:K^1(U)K^0(U)\delta:\hat{K}_{1}(U)\hookrightarrow\hat{K}_{0}(U) is injective, so 𝐊^\hat{\mathbf{K}} is a planar functor. Furthermore, we claim that SPL,0(PL0cl(U))im(δ:K^1(U)K^0(U))S_{\operatorname{PL},0}(\operatorname{PL}_{0}^{\operatorname{cl}}(U))\subset\operatorname{im}(\delta:\hat{K}_{1}(U)\to\hat{K}_{0}(U)). Indeed, the signature of a closed loop 𝐱PL0cl(U)\mathbf{x}\in\operatorname{PL}_{0}^{\operatorname{cl}}(U) must be trivial in degree 11, implying that log(SPL,0(𝐱))LCS^2(𝔨0(U))\log(S_{\operatorname{PL},0}(\mathbf{x}))\in\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(U)). But we have just seen that im(δ:𝔨^1(U)𝔨^0(U))=LCS^2(𝔨0(U))\operatorname{im}(\delta:\hat{\mathfrak{k}}_{1}(U)\to\hat{\mathfrak{k}}_{0}(U))=\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(U)). Therefore, SPL,0(𝐱)im(δ:K^1(U)K^0(U))S_{\operatorname{PL},0}(\mathbf{x})\in\operatorname{im}(\delta:\hat{K}_{1}(U)\to\hat{K}_{0}(U)). Thus, by Lemma 5.15, there exists a unique extension 𝐒PL=(SPL,1,SPL,0)\mathbf{S}_{\operatorname{PL}}=(S_{\operatorname{PL},1},S_{\operatorname{PL},0}). ∎

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 𝐑\mathbf{R}, 𝐒\mathbf{S}, and 𝐒PL\mathbf{S}_{\operatorname{PL}} are natural transformations which factor as

𝐒PL:𝐏𝐋𝐑𝝉𝐒𝐊^.\displaystyle\mathbf{S}_{\operatorname{PL}}:{\boldsymbol{\operatorname{PL}}}\xRightarrow{\mathbf{R}}{\boldsymbol{\tau}}\xRightarrow{\mathbf{S}}\hat{\mathbf{K}}. (5.61)

This factorization implies that the smooth surface signature is uniquely determined.

Theorem 5.21.

The smooth signature is the unique natural transformation

𝐒=(S1,S0):𝝉𝐊^,\displaystyle\mathbf{S}=(S_{1},S_{0}):{\boldsymbol{\tau}}\Rightarrow\hat{\mathbf{K}}, (5.62)

which extends S0:τ0K^0S_{0}:\tau_{0}\Rightarrow\hat{K}_{0} from (2.23) and such that S1:τ2(V)K^1(V)S_{1}:\tau_{2}(V)\to\hat{K}_{1}(V) is continuous with respect to the quotient topology on τ2(V)\tau_{2}(V) induced by the Lipschitz topology on C01([0,1]2,V)C^{1}_{0}([0,1]^{2},V).

Proof.

Let 𝐓=(T1,T0):𝝉𝐊^\mathbf{T}=(T_{1},T_{0}):{\boldsymbol{\tau}}\Rightarrow\hat{\mathbf{K}} be a continuous natural transformation which extends S0S_{0}. Then 𝐓𝐑:𝐏𝐋𝐊^\mathbf{T}\circ\mathbf{R}:{\boldsymbol{\operatorname{PL}}}\to\hat{\mathbf{K}} is a natural transformation which extends SPL,0S_{\operatorname{PL},0}. By  Theorem 5.19, 𝐓𝐑=𝐒PL\mathbf{T}\circ\mathbf{R}=\mathbf{S}_{\operatorname{PL}}. Therefore, 𝐓\mathbf{T} and 𝐒\mathbf{S} agree on piecewise linear surfaces. Now let 𝐗C01([0,1]2,V)\mathbf{X}\in C^{1}_{0}([0,1]^{2},V). There exists a sequence of piecewise linear surfaces 𝐗k\mathbf{X}^{k} such that 𝐗kLip𝐗\mathbf{X}^{k}\xrightarrow{\operatorname{Lip}}\mathbf{X} in the Lipschitz topology. Indeed, let TkT^{k} be a triangulation of [0,1]2[0,1]^{2} such that the mesh size approaches 0. Define 𝐗s,tk=𝐗s,t\mathbf{X}^{k}_{s,t}=\mathbf{X}_{s,t} on all vertices (s,t)Tk(s,t)\in T_{k}, and extend linearly. Then, because 𝐗\mathbf{X} is smooth and [0,1]2[0,1]^{2} is compact, the derivatives of 𝐗k\mathbf{X}^{k} converge to those of 𝐗\mathbf{X}, and thus 𝐗kLip𝐗\mathbf{X}^{k}\xrightarrow{\operatorname{Lip}}\mathbf{X}. Because T1(𝐗k)=S1(𝐗k)T_{1}(\mathbf{X}^{k})=S_{1}(\mathbf{X}^{k}) for all kk, we must have T1(𝐗)=S1(𝐗)T_{1}(\mathbf{X})=S_{1}(\mathbf{X}) 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 𝐏𝐋{\boldsymbol{\operatorname{PL}}}, 𝝉{\boldsymbol{\tau}}, and 𝐊^\hat{\mathbf{K}} to crossed modules of groupoids defined on the category 𝖠𝖿𝖿\mathsf{Aff} 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. (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. (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 VV. 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 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V). To do this, we will make use of a universal property for the group of piecewise linear loops PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V), which we obtain by expressing PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V) as a quotient of a free group generated by triangular loops.

First, recall that the pair groupoid of VV, Pair(V)V{\operatorname{Pair}}(V)\rightrightarrows V, is the groupoid whose set of objects is VV and such that there is a unique morphism between any two objects. Hence, the space of morphisms is Pair(V)=V×V{\operatorname{Pair}}(V)=V\times V. By sending the pair (v,u)V×V(v,u)\in V\times V to the triangular loop with vertices (0,v,u)(0,v,u) we obtain a map of sets

ηV:Pair(V)PL0cl(V),ηV(v,u)=(v,uv,u).\displaystyle\eta_{V}:{\operatorname{Pair}}(V)\to\operatorname{PL}_{0}^{\operatorname{cl}}(V),\quad\quad\eta_{V}(v,u)=(v,u-v,-u). (5.63)

A simple computation shows that ηV\eta_{V} evaluates to the identity on every pair (v,u)(v,u) for which vv and uu are colinear, and is a groupoid homomorphism when restricted to any affine line V\ell\subset V. We use these two properties to define a monoid generated by the set of triangular loops. To this end, consider the quotient

Loop(V)𝖥𝖬𝗈𝗇(V2)/\displaystyle{\operatorname{Loop}}(V)\coloneqq\mathsf{FMon}(V^{2})/\sim (5.64)

of the free monoid by the relations

  1. (L.1)

    (v,u)(u,r)(v,r)(v,u)\star(u,r)\sim(v,r) if v,u,rv,u,r lie on an affine line, and

  2. (L.2)

    (v,u)(v,u)\sim\emptyset if vv and uu are linearly dependent.

Note that with these relations, Loop(V){\operatorname{Loop}}(V) is a group. Indeed, the inverse of the generator (v,u)(v,u) is (u,v)(u,v). The following theorem is proved in Section B.2.

Theorem 5.23.

There is a group isomorphism Loop(V)PL0cl(V){\operatorname{Loop}}(V)\cong\operatorname{PL}_{0}^{\operatorname{cl}}(V).

As a result, we obtain the following universal property for the group PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V), which is analogous to  Lemma 2.13 and Theorem 5.8.

Proposition 5.24.

Let VV be a vector space and let GG be a group. Let f:Pair(V)Gf:{\operatorname{Pair}}(V)\to G be a map which

  1. (1)

    restricts to a groupoid homomorphism on Pair(){\operatorname{Pair}}(\ell) for every affine line V\ell\subset V, and

  2. (2)

    restricts to the trivial homomorphism on Pair(L){\operatorname{Pair}}(L) for every subspace LVL\subset V with dim(L)1\dim(L)\leq 1.

Then there exists a unique group homomorphism F:PL0cl(V)GF:\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to G such that FηV=fF\circ\eta_{V}=f.

We can use this universal property to define a section of δ:PL1(V)PL0cl(V)\delta:\operatorname{PL}_{1}(V)\to\operatorname{PL}_{0}^{\operatorname{cl}}(V). Indeed, consider the map c:Pair(V)PL1(V)c:{\operatorname{Pair}}(V)\to\operatorname{PL}_{1}(V) defined by c(v,u)=(0,ηV(v,u))PL1(V)c(v,u)=(\emptyset_{0},\eta_{V}(v,u))\in\operatorname{PL}_{1}(V). This map verifies the equation δc=ηV\delta\circ c=\eta_{V} and satisfies the two conditions in Proposition 5.24. Therefore, it induces a homomorphism ConePL:PL0cl(V)PL1(V){\operatorname{Cone}}_{\operatorname{PL}}:\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to\operatorname{PL}_{1}(V) which satisfies δConePL=id\delta\circ{\operatorname{Cone}}_{\operatorname{PL}}=\operatorname{id}. As a result, the following short exact sequence splits

1PL1cl(V)PL1(V)𝛿PL0cl(V)1.\displaystyle 1\rightarrow\operatorname{PL}_{1}^{\operatorname{cl}}(V)\hookrightarrow\operatorname{PL}_{1}(V)\xrightarrow{\delta}\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to 1. (5.65)
Corollary 5.25.

The map

Φ:PL1(V)PL1cl(V)×PL0cl(V),Φ(𝐗)=(𝐗(ConePLδ(𝐗))1,δ(𝐗))\displaystyle\Phi:\operatorname{PL}_{1}(V)\to\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V),\quad\Phi(\mathbf{X})=(\mathbf{X}\star({\operatorname{Cone}}_{\operatorname{PL}}\circ\delta(\mathbf{X}))^{-1},\delta(\mathbf{X})) (5.66)

is a group isomorphism.

Proof.

Because the above splitting, PL1(V)\operatorname{PL}_{1}(V) is isomorphic to the semidirect product PL1cl(V)PL0cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V)\rtimes\operatorname{PL}_{0}^{\operatorname{cl}}(V). However, the conjugation action of PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V) on PL1cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V) is trivial since PL1cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V) is in the center of PL1(V)\operatorname{PL}_{1}(V). Thus PL1(V)\operatorname{PL}_{1}(V) is isomorphic to the direct product. ∎

The crossed module action in 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V) induces an action of PL0(V)\operatorname{PL}_{0}(V) on PL1cl(V)×PL0cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V) under the isomorphism Φ\Phi. In particular, for 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V) and 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V), we have

Φ(𝐱𝐗)\displaystyle\Phi(\mathbf{x}\vartriangleright\mathbf{X}) =((𝐱𝐗)ConePL(δ(𝐱𝐗)))1,𝐱δ(𝐗)𝐱1),\displaystyle=\Big{(}(\mathbf{x}\vartriangleright\mathbf{X})\star{\operatorname{Cone}}_{\operatorname{PL}}(\delta(\mathbf{x}\vartriangleright\mathbf{X})))^{-1},\mathbf{x}\star\delta(\mathbf{X})\star\mathbf{x}^{-1}\Big{)}, (5.67)

and the induced action on (𝐙,𝐛)PL1cl(V)×PL0cl(V)(\mathbf{Z},\mathbf{b})\in\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V), where 𝐗=Φ1(𝐙,𝐛)=𝐙ConePL(𝐛)\mathbf{X}=\Phi^{-1}(\mathbf{Z},\mathbf{b})=\mathbf{Z}\star{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{b}), is

𝐱(𝐙,𝐛)\displaystyle\mathbf{x}\vartriangleright(\mathbf{Z},\mathbf{b}) =((𝐱𝐙)(𝐱ConePL(𝐛))ConePL(𝐱𝐛𝐱1)1,𝐱𝐛𝐱1).\displaystyle=\Big{(}(\mathbf{x}\vartriangleright\mathbf{Z})\star(\mathbf{x}\vartriangleright{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{b}))\star{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1})^{-1},\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}\Big{)}. (5.68)

Define the suspension, SuspPL:PL0(V)×PL0cl(V)PL1cl(V){\operatorname{Susp}}_{\operatorname{PL}}:\operatorname{PL}_{0}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to\operatorname{PL}_{1}^{\operatorname{cl}}(V), by

SuspPL(𝐱,𝐛)=(𝐱ConePL(𝐛))ConePL(𝐱𝐛𝐱1)1.\displaystyle{\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{x},\mathbf{b})=(\mathbf{x}\vartriangleright{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{b}))\star{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1})^{-1}. (5.69)

Using the Peiffer identity, we can verify that, if 𝐜PL0cl(V)\mathbf{c}\in\operatorname{PL}_{0}^{\operatorname{cl}}(V), then

SuspPL(𝐜𝐱,𝐛)=SuspPL(𝐱𝐜,𝐛)=SuspPL(𝐱,𝐜𝐛𝐜1)=SuspPL(𝐱,𝐛)\displaystyle{\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{c}\star\mathbf{x},\mathbf{b})={\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{x}\star\mathbf{c},\mathbf{b})={\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{x},\mathbf{c}\star\mathbf{b}\star\mathbf{c}^{-1})={\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{x},\mathbf{b}) (5.70)

In particular, SuspPL(𝐱,𝐛){\operatorname{Susp}}_{\operatorname{PL}}(\mathbf{x},\mathbf{b}) only depends on the path 𝐱\mathbf{x} through its endpoint t(𝐱)Vt(\mathbf{x})\in V. If 𝐱=δ(ConePL(𝐱))PL0cl(V)\mathbf{x}=\delta({\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{x}))\in\operatorname{PL}_{0}^{\operatorname{cl}}(V), then by the Peiffer identity we have

𝐱(𝐙,𝐛)=(1,𝐱)(𝐙,𝐛)(1,𝐱)1=(𝐙,𝐱𝐛𝐱1).\displaystyle\mathbf{x}\vartriangleright(\mathbf{Z},\mathbf{b})=(\emptyset_{1},\mathbf{x})\cdot(\mathbf{Z},\mathbf{b})\cdot(\emptyset_{1},\mathbf{x})^{-1}=(\mathbf{Z},\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}). (5.71)

Hence, the action of PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V) on the PL1cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V)-component is trivial. As a result, we obtain the following expression for the action.

Lemma 5.26.

The action of PL0(V)\operatorname{PL}_{0}(V) on PL1cl(V)×PL0cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V) is given by the following formula

𝐱(𝐙,𝐛)=((t(𝐱)𝐙)SuspPL(t(𝐱),𝐛),𝐱𝐛𝐱1),\displaystyle\mathbf{x}\vartriangleright(\mathbf{Z},\mathbf{b})=\Big{(}(t(\mathbf{x})\vartriangleright\mathbf{Z})\star{\operatorname{Susp}}_{\operatorname{PL}}(t(\mathbf{x}),\mathbf{b}),\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}\Big{)}, (5.72)

where (𝐙,𝐛)PL1cl(V)×PL0cl(V)(\mathbf{Z},\mathbf{b})\in\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V), 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V), and we view t(𝐱)PL0(V)t(\mathbf{x})\in\operatorname{PL}_{0}(V) as the linear path from the origin to the endpoint of 𝐱\mathbf{x}.

5.4.2. Decomposition of the Kapranov Crossed Module.

Next, we study a canonical splitting of Kapranov’s Lie algebra 𝔨1(V)\mathfrak{k}_{1}(V) in order to obtain a decomposition of the group K^1(V)\hat{K}_{1}(V). We use the definitions in Section 4.2 for polynomial differential forms and currents. Let

E=ixixi\displaystyle E=\sum_{i}x_{i}\partial_{x_{i}} (5.73)

be the Euler vector field on the vector space VV. Let ιE:Ω¯k+1(V)Ω¯k(V)\iota_{E}:\overline{\Omega}^{k+1}(V)\to\overline{\Omega}^{k}(V) be the derivation of the de Rham complex given by interior product with EE. This operator is 𝔤𝔩(V)\mathfrak{gl}(V)-equivariant and satisfies ιE2=0\iota_{E}^{2}=0. The Lie derivative LE=[d,ιE]L_{E}=[d,\iota_{E}] acts on the weight rr subcomplex Ω¯k(V)r\overline{\Omega}^{k}(V)_{r} by multiplication by rr. Recall that we can extend the de Rham complex by putting a copy of the base field \mathbb{R} in degree 1-1. We can then extend ιE\iota_{E} by sending an element fΩ¯0(V)f\in\overline{\Omega}^{0}(V) to the value f(0)f(0). As part of the proof of the Poincare lemma, we obtain the following decomposition.

Lemma 5.27.

For each kk\in\mathbb{N}, we have a direct sum decomposition of 𝔤𝔩(V)\mathfrak{gl}(V)-represenations,

Ω¯k(V)=ker(d:Ω¯k(V)Ω¯k+1(V))ker(ιE:Ω¯k(V)Ω¯k1(V)).\displaystyle\overline{\Omega}^{k}(V)=\ker(d:\overline{\Omega}^{k}(V)\to\overline{\Omega}^{k+1}(V))\oplus\ker(\iota_{E}:\overline{\Omega}^{k}(V)\to\overline{\Omega}^{k-1}(V)). (5.74)

We dualize this to get a decomposition for currents. Recall that :Γ¯k(V)Γ¯k1(V)\partial:\overline{\Gamma}_{k}(V)\to\overline{\Gamma}_{k-1}(V) is the codifferential from (4.39). Similarly, the dual of ιE\iota_{E} is the operator e:Γ¯k(V)Γ¯k+1(V)e:\overline{\Gamma}_{k}(V)\to\overline{\Gamma}_{k+1}(V) defined by

e(α),ω=(1)kα,ιEω\displaystyle\langle e(\alpha),\omega\rangle=(-1)^{k}\langle\alpha,\iota_{E}\omega\rangle (5.75)

for αΓ¯k(V)\alpha\in\overline{\Gamma}_{k}(V) and ωΩ¯k+1(V)\omega\in\overline{\Omega}^{k+1}(V). It can be expressed explicitly as

e(u1urv1vk)=i=1ru1u^iuruiv1vk.\displaystyle e(u_{1}\cdots u_{r}\otimes v_{1}\wedge\ldots\wedge v_{k})=\sum_{i=1}^{r}u_{1}\cdots\hat{u}_{i}\cdots u_{r}\otimes u_{i}\wedge v_{1}\wedge\ldots\wedge v_{k}. (5.76)

Because ee is the dual of ιE\iota_{E}, we have e2=0e^{2}=0. Furthermore, =[e,]\ell=[e,\partial] acts on Γ¯k(V)r\overline{\Gamma}_{k}(V)_{r} by multiplication by the (negative) weight r-r. By the same argument as above, we obtain the following.

Lemma 5.28.

For each kk\in\mathbb{N}, we have a direct sum decomposition of 𝔤𝔩(V)\mathfrak{gl}(V)-representations,

Γ¯k(V)=ker(:Γ¯k(V)Γ¯k1(V))ker(e:Γ¯k(V)Γ¯k+1(V)).\displaystyle\overline{\Gamma}_{k}(V)=\ker(\partial:\overline{\Gamma}_{k}(V)\to\overline{\Gamma}_{k-1}(V))\oplus\ker(e:\overline{\Gamma}_{k}(V)\to\overline{\Gamma}_{k+1}(V)). (5.77)

As a particular case of this Lemma, we have the decomposition Γ¯2(V)=Γ¯2cl(V)ker(e)\overline{\Gamma}_{2}(V)=\overline{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\ker(e). Recall Theorem 4.6, which states that the abelianization map ρ:𝔨1(V)Γ¯2(V)\rho:\mathfrak{k}_{1}(V)\to\overline{\Gamma}_{2}(V) sends 𝔞1(V)=ker(δ)\mathfrak{a}_{1}(V)=\ker(\delta) isomorphically to Γ¯2cl(V)\overline{\Gamma}_{2}^{\operatorname{cl}}(V). Therefore, the ideal ρ1(ker(e))𝔨1(V)\rho^{-1}(\ker(e))\subset\mathfrak{k}_{1}(V) is a complement to 𝔞1(V)\mathfrak{a}_{1}(V) and hence is isomorphic to LCS2(𝔨0(V))=[𝔨0(V),𝔨0(V)]{\operatorname{LCS}}_{2}(\mathfrak{k}_{0}(V))=[\mathfrak{k}_{0}(V),\mathfrak{k}_{0}(V)] via the map δ\delta (cf. [47]). Working with completions, we define

^(V)ρ1(ker(e:Γ^2(V)Γ^3(V)))whereρ:𝔨^1(V)Γ^2(V).\displaystyle\hat{\mathcal{E}}(V)\coloneqq\rho^{-1}\Big{(}\ker(e:\hat{\Gamma}_{2}(V)\to\hat{\Gamma}_{3}(V))\Big{)}\quad\text{where}\quad\rho:\hat{\mathfrak{k}}_{1}(V)\to\hat{\Gamma}_{2}(V). (5.78)

Then δ:^(V)LCS^2(𝔨0(V))\delta:\hat{\mathcal{E}}(V)\to\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)) is an isomorphism, and we denote the inverse morphism as

𝔠:LCS^2(𝔨0(V))^(V)𝔨^1(V).\displaystyle\mathfrak{c}:\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V))\xrightarrow{\cong}\hat{\mathcal{E}}(V)\subset\hat{\mathfrak{k}}_{1}(V). (5.79)

In particular, the map 𝔠\mathfrak{c} satisfies δ𝔠=id\delta\circ\mathfrak{c}=\operatorname{id}, and should be viewed as the Lie algebraic analogue of the ConePL{\operatorname{Cone}}_{\operatorname{PL}} construction. Thus, we obtain the following decomposition.

Corollary 5.29.

There exists a canonical isomorphism of Lie algebras

Ψ:𝔨^1(V)Γ^2cl(V)LCS^2(𝔨0(V))defined byΨ(A)=(ρ(A𝔠δ(A)),δ(A)).\displaystyle\Psi:\hat{\mathfrak{k}}_{1}(V)\xrightarrow{\cong}\hat{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V))\quad\text{defined by}\quad\Psi(A)=\Big{(}\rho(A-\mathfrak{c}\circ\delta(A)),\delta(A)\Big{)}. (5.80)

Under this canonical decomposition of 𝔨^1(V)\hat{\mathfrak{k}}_{1}(V), the crossed module action of 𝖐^(V)\hat{{\boldsymbol{\mathfrak{k}}}}(V) induces an action of 𝔨^0(V)\hat{\mathfrak{k}}_{0}(V) on Γ^2cl(V)LCS^2(𝔨0(V))\hat{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)), which we can write out explicitly. For x𝔨^0(V)x\in\hat{\mathfrak{k}}_{0}(V) and A𝔨^1(V)A\in\hat{\mathfrak{k}}_{1}(V), we have

Ψ(xA)=(ρ(xA𝔠δ(xA)),δ(xA)).\displaystyle\Psi(x\vartriangleright A)=\Big{(}\rho(x\vartriangleright A-\mathfrak{c}\circ\delta(x\vartriangleright A)),\delta(x\vartriangleright A)\Big{)}. (5.81)

From the construction of 𝖐^(V)\hat{{\boldsymbol{\mathfrak{k}}}}(V) outlined in  Appendix C, we observe that the abelianization map ρ\rho satisfies ρ(xA)=π(x)ρ(A)\rho(x\vartriangleright A)=\pi(x)\rho(A), where VV acts via the product in the symmetric algebra S^(V)\hat{S}(V) of Γ^2=S^(V)Λ2(V)\hat{\Gamma}_{2}=\hat{S}(V)\otimes\Lambda^{2}(V). Therefore, the induced action on (γ,y)Γ^2cl(V)LCS^2(𝔨0(V))(\gamma,y)\in\hat{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)), where A=Ψ1(γ,y)=ρ1(γ)+c(y)A=\Psi^{-1}(\gamma,y)=\rho^{-1}(\gamma)+c(y), is

x(γ,y)=(π(x)γ+ρ(x𝔠(y)𝔠([x,y])),[x,y]).\displaystyle x\vartriangleright(\gamma,y)=\Big{(}\pi(x)\gamma+\rho(x\vartriangleright\mathfrak{c}(y)-\mathfrak{c}([x,y])),[x,y]\Big{)}. (5.82)

We define the Lie algebraic analogue of the suspension 𝔰:𝔨^0(V)×LCS^2(𝔨0(V))Γ^2cl(V)\mathfrak{s}:\hat{\mathfrak{k}}_{0}(V)\times\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V))\to\hat{\Gamma}_{2}^{\operatorname{cl}}(V) by

𝔰(x,y)=ρ(x𝔠(y)𝔠([x,y])),\displaystyle\mathfrak{s}(x,y)=\rho(x\vartriangleright\mathfrak{c}(y)-\mathfrak{c}([x,y])), (5.83)

which is the projection of ρ(x𝔠(y))Γ^2\rho(x\vartriangleright\mathfrak{c}(y))\in\hat{\Gamma}_{2} onto Γ^2cl\hat{\Gamma}_{2}^{\operatorname{cl}}. By applying the Peiffer identity, we can verify that for rLCS^2(𝔨0(V))r\in\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)), we have

𝔰(r+x,y)=𝔰(x,y).\displaystyle\mathfrak{s}(r+x,y)=\mathfrak{s}(x,y). (5.84)

Therefore, 𝔰(x,y)\mathfrak{s}(x,y) only depends on v=π(x)Vv=\pi(x)\in V. As a result, we obtain the following expression for the crossed module action.

Lemma 5.30.

The action of 𝔨^0(V)\hat{\mathfrak{k}}_{0}(V) on Γ^2cl(V)LCS^2(𝔨0(V))\hat{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)) is given by the following formula

x(γ,y)=(π(x)γ+𝔰(π(x),y),[x,y]),\displaystyle x\vartriangleright(\gamma,y)=(\pi(x)\gamma+\mathfrak{s}(\pi(x),y),[x,y]), (5.85)

where (γ,y)Γ^2cl(V)LCS^2(𝔨0(V))(\gamma,y)\in\hat{\Gamma}_{2}^{\operatorname{cl}}(V)\oplus\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)) and x𝔨^0(V)x\in\hat{\mathfrak{k}}_{0}(V).

Remark 5.31.

The suspension map 𝔰\mathfrak{s} in fact only depends on yLCS^2(𝔨0(V))y\in\widehat{{\operatorname{LCS}}}_{2}(\mathfrak{k}_{0}(V)) through its abelianization. To see this, consider the operator =[e,]\ell=[e,\partial] acting on currents Γ^2\hat{\Gamma}_{2}. It satisfies

id=1e+1e.\displaystyle\operatorname{id}=\ell^{-1}e\partial+\ell^{-1}\partial e. (5.86)

Since \ell acts as multiplication by the negative weight r-r, we can identify 1e\ell^{-1}e\partial as the projection onto ker(e)\ker(e), and 1e\ell^{-1}\partial e as the projection onto ker()=Γ^2cl\ker(\partial)=\hat{\Gamma}_{2}^{\operatorname{cl}}. Thus, we can express 𝔰\mathfrak{s} as

𝔰(x,y)=1e(π(x)ρ(𝔠(y))),\displaystyle\mathfrak{s}(x,y)=\ell^{-1}\circ\partial\circ e\big{(}\pi(x)\rho(\mathfrak{c}(y))\big{)}, (5.87)

so that 𝔰(x,y)\mathfrak{s}(x,y) depends only on the projection π(x)\pi(x) and the abelianization ρ(𝔠(y))\rho(\mathfrak{c}(y)).

Finally, the decomposition of  Corollary 5.29 induces a decomposition of the group K^1(V)\hat{K}_{1}(V)

K^1(V)=K^1Γ(V)×K^1(V),\displaystyle\hat{K}_{1}(V)=\hat{K}_{1}^{\Gamma}(V)\times\hat{K}_{1}^{\mathcal{E}}(V), (5.88)

where K^1Γ(V)ker(δ:K^1(V)K^0(V))\hat{K}_{1}^{\Gamma}(V)\coloneqq\ker(\delta:\hat{K}_{1}(V)\to\hat{K}_{0}(V)). Note that because K^1Γ(V)\hat{K}_{1}^{\Gamma}(V) is abelian, we may identify it with Γ^2cl(V)\hat{\Gamma}_{2}^{\operatorname{cl}}(V) with its additive group structure. Finally, because the constructions in this section are functorial, given a linear map ϕ:VW\phi:V\to W, the induced group homomorphism K^1(ϕ):K^1(V)K^1(W)\hat{K}_{1}(\phi):\hat{K}_{1}(V)\to\hat{K}_{1}(W) 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 PL1(V)\operatorname{PL}_{1}(V) and K^1(V)\hat{K}_{1}(V). We consider

SPL,1Φ1:PL1cl(V)×PL0cl(V)K^1Γ(V)×K^1(V),\displaystyle S_{\operatorname{PL},1}\circ\Phi^{-1}:\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to\hat{K}_{1}^{\Gamma}(V)\times\hat{K}_{1}^{\mathcal{E}}(V), (5.89)

where Φ1:PL1cl(V)×PL0cl(V)PL1(V)\Phi^{-1}:\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\operatorname{PL}_{0}^{\operatorname{cl}}(V)\to\operatorname{PL}_{1}(V) is the isomorphism from Corollary 5.25 given by Φ1(𝐗,𝐛)=𝐗ConePL(𝐛)\Phi^{-1}(\mathbf{X},\mathbf{b})=\mathbf{X}\star{\operatorname{Cone}}_{\operatorname{PL}}(\mathbf{b}).

Lemma 5.32.

The homomorphism SPL,1Φ1S_{\operatorname{PL},1}\circ\Phi^{-1} preserves the decompositions of both PL1(V)\operatorname{PL}_{1}(V) and K^1(V)\hat{K}_{1}(V).

Proof.

The restriction of SPL,1S_{\operatorname{PL},1} to PL1cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V) must be valued in K^1Γ(V)\hat{K}_{1}^{\Gamma}(V) since 𝐒PL\mathbf{S}_{\operatorname{PL}} is a morphism of crossed modules. To understand the restriction of SPL,1Φ1S_{\operatorname{PL},1}\circ\Phi^{-1} to PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V), consider the function

sV:Pair(V)ηVPL0cl(V)ConePLPL1(V)SPL,1K^1(V).\displaystyle s_{V}:{\operatorname{Pair}}(V)\xrightarrow{\eta_{V}}\operatorname{PL}_{0}^{\operatorname{cl}}(V)\xrightarrow{{\operatorname{Cone}}_{\operatorname{PL}}}\operatorname{PL}_{1}(V)\xrightarrow{S_{\operatorname{PL},1}}\hat{K}_{1}(V). (5.90)

Let (v,u)Pair(V)(v,u)\in{\operatorname{Pair}}(V), let UVU\subset V be a 2-dimensional subspace such that v,uUv,u\in U, and let ϕU:UV\phi_{U}:U\hookrightarrow V denote the inclusion. Because the PL surface signature is a natural transformation, we have

sV(v,u)=K^1(ϕU)sU(v,u).\displaystyle s_{V}(v,u)=\hat{K}_{1}(\phi_{U})\circ s_{U}(v,u). (5.91)

Furthermore, the factor K^1Γ(U)=0\hat{K}_{1}^{\Gamma}(U)=0 is trivial since UU is 2-dimensional, and this implies that sU(v,u)K^1(U)s_{U}(v,u)\in\hat{K}_{1}^{\mathcal{E}}(U). Then, because K^1(ϕU)\hat{K}_{1}(\phi_{U}) preserves the decomposition in (5.88), we conclude that sV(v,u)K^1(V)s_{V}(v,u)\in\hat{K}_{1}^{\mathcal{E}}(V). The map sVs_{V} satisfies the hypotheses of Proposition 5.24 and so by the universal property, there is a unique map

F:PL0(V)K^1(V)K^1(V)\displaystyle F:\operatorname{PL}_{0}(V)\to\hat{K}_{1}^{\mathcal{E}}(V)\subset\hat{K}_{1}(V) (5.92)

such that FηV=sVF\circ\eta_{V}=s_{V}. Recalling the definition of sVs_{V} from  (5.90), we see that (SPL,1ConePL)ηV=sV(S_{\operatorname{PL},1}\circ{\operatorname{Cone}}_{\operatorname{PL}})\circ\eta_{V}=s_{V} and hence

F=SPL,1ConePL=(SPL,1Φ1)|PL0cl(V),\displaystyle F=S_{\operatorname{PL},1}\circ{\operatorname{Cone}}_{\operatorname{PL}}=(S_{\operatorname{PL},1}\circ\Phi^{-1})|_{\operatorname{PL}_{0}^{\operatorname{cl}}(V)}, (5.93)

implying that SPL,1ConePLS_{\operatorname{PL},1}\circ{\operatorname{Cone}}_{\operatorname{PL}} is valued in K^1(V)\hat{K}_{1}^{\mathcal{E}}(V). As a result, we conclude that SPL,1Φ1S_{\operatorname{PL},1}\circ\Phi^{-1} respects the decompositions of both PL1(V)\operatorname{PL}_{1}(V) and K^1(V)\hat{K}_{1}(V). ∎

The signature satisfies the following equation

δSPL,1ConePL=SPL,0δConePL=SPL,0.\displaystyle\delta\circ S_{\operatorname{PL},1}\circ{\operatorname{Cone}}_{\operatorname{PL}}=S_{\operatorname{PL},0}\circ\delta\circ{\operatorname{Cone}}_{\operatorname{PL}}=S_{\operatorname{PL},0}. (5.94)

Hence, using the identification between K^1(V)\hat{K}_{1}^{\mathcal{E}}(V) and its image under δ\delta in K^0(V)\hat{K}_{0}(V), we conclude that (SPL,1Φ1)|PL0cl(V)=SPL,0(S_{\operatorname{PL},1}\circ\Phi^{-1})|_{\operatorname{PL}_{0}^{\operatorname{cl}}(V)}=S_{\operatorname{PL},0}. This implies the following decomposition of the surface signature.

Proposition 5.33.

The piecewise linear surface signature decomposes as

SPL,1=(SPL,1Γ,SPL,1):PL1(V)K^1Γ(V)×K^1(V),\displaystyle S_{\operatorname{PL},1}=(S_{\operatorname{PL},1}^{\Gamma},S_{\operatorname{PL},1}^{\mathcal{E}}):\operatorname{PL}_{1}(V)\to\hat{K}_{1}^{\Gamma}(V)\times\hat{K}_{1}^{\mathcal{E}}(V), (5.95)

where

SPL,1Γ(𝐗)=SPL,1(𝐗ConePL(δ(𝐗)1))andSPL,1(𝐗)=SPL,0(δ(𝐗)).\displaystyle S_{\operatorname{PL},1}^{\Gamma}(\mathbf{X})=S_{\operatorname{PL},1}(\mathbf{X}\star{\operatorname{Cone}}_{\operatorname{PL}}(\delta(\mathbf{X})^{-1}))\quad\text{and}\quad S_{\operatorname{PL},1}^{\mathcal{E}}(\mathbf{X})=S_{\operatorname{PL},0}(\delta(\mathbf{X})). (5.96)

5.4.4. Decomposition of Smooth Surface Signature

We can also consider a decomposition for the smooth thin crossed module

𝝉(V)=(:τ2(V)τ1(V),).\displaystyle{\boldsymbol{\tau}}(V)=(\partial:\tau_{2}(V)\to\tau_{1}(V),\vartriangleright). (5.97)

We denote the kernel by τ2cl(V)ker(:τ2(V)τ1(V))\tau_{2}^{\operatorname{cl}}(V)\coloneqq\ker(\partial:\tau_{2}(V)\to\tau_{1}(V)), and the group of thin homotopy classes of loops by τ1cl(V)im(:τ2(V)τ1(V))\tau_{1}^{\operatorname{cl}}(V)\coloneqq\operatorname{im}(\partial:\tau_{2}(V)\to\tau_{1}(V)). Recall our convention that these groups consist of thin homotopy classes of paths and surfaces based at the origin, so 𝐱0=𝐗0,0=0\mathbf{x}_{0}=\mathbf{X}_{0,0}=0. Now, we define a section of :τ2(V)τ1cl(V)\partial:\tau_{2}(V)\to\tau^{\operatorname{cl}}_{1}(V) by

Cone:τ1cl(V)τ2(V),Cone(𝐱)=((s,t)t𝐱s).\displaystyle{\operatorname{Cone}}:\tau_{1}^{\operatorname{cl}}(V)\to\tau_{2}(V),\quad{\operatorname{Cone}}(\mathbf{x})=\Big{(}(s,t)\mapsto t\mathbf{x}_{s}\Big{)}. (5.98)

This is well-defined under thin homotopy. Indeed, suppose 𝐱th𝐲\mathbf{x}\sim_{\operatorname{th}}\mathbf{y}, and let h:[0,1]2Vh:[0,1]^{2}\to V be a thin homotopy between 𝐱\mathbf{x} and 𝐲\mathbf{y} where h0,s=𝐱sh_{0,s}=\mathbf{x}_{s} and h1,s=𝐲sh_{1,s}=\mathbf{y}_{s}. Define a map H:[0,1]3VH:[0,1]^{3}\to V by

Hu,s,t=thu,s,whereH0,s,t=Cone(𝐱)s,tandH1,s,t=Cone(𝐲)s,t.\displaystyle H_{u,s,t}=th_{u,s},\quad\text{where}\quad H_{0,s,t}={\operatorname{Cone}}(\mathbf{x})_{s,t}\quad\text{and}\quad H_{1,s,t}={\operatorname{Cone}}(\mathbf{y})_{s,t}. (5.99)

Because rank(dh)1\operatorname{rank}(dh)\leq 1, we must have rank(dH)2\operatorname{rank}(dH)\leq 2. Hence, Cone(𝐱)thCone(𝐲){\operatorname{Cone}}(\mathbf{x})\sim_{\operatorname{th}}{\operatorname{Cone}}(\mathbf{y}). Then, by the same arguments as the PL setting in Corollary 5.25, we have the following.

Lemma 5.34.

The map

Φ:τ2(V)τ2cl(V)×τ1cl(V),Φ(𝐗)=(𝐗Cone((𝐗))1,(𝐗))\displaystyle\Phi:\tau_{2}(V)\to\tau^{\operatorname{cl}}_{2}(V)\times\tau^{\operatorname{cl}}_{1}(V),\quad\Phi(\mathbf{X})=\Big{(}\mathbf{X}\star{\operatorname{Cone}}(\partial(\mathbf{X}))^{-1},\partial(\mathbf{X})\Big{)} (5.100)

is a group isomorphism.

The map Cone{\operatorname{Cone}} extends the piecewise linear cone in the sense that ConeR0=R1ConePL{\operatorname{Cone}}\circ R_{0}=R_{1}\circ{\operatorname{Cone}}_{\operatorname{PL}}. This can be verified by checking equality after precomposing with ηV\eta_{V} and applying Proposition 5.24. Using this fact, we can show that the surface signature also preserves the decompositions of τ2(V)\tau_{2}(V) and K^1(V)\hat{K}_{1}(V). The main property to show is that cones are mapped to K^1(V)\hat{K}_{1}^{\mathcal{E}}(V) under the surface signature.

Lemma 5.35.

The map S1Cone:τ1cl(V)K^1(V)S_{1}\circ{\operatorname{Cone}}:\tau_{1}^{\operatorname{cl}}(V)\to\hat{K}_{1}(V) is valued in K^1(V)\hat{K}_{1}^{\mathcal{E}}(V).

Proof.

Recall that 𝔨1(V)=Γ2cl(V)(V)\mathfrak{k}_{1}(V)=\Gamma_{2}^{\operatorname{cl}}(V)\oplus\mathcal{E}(V). By projection, this decomposition also holds for truncations 𝔨1(n)(V)=Γ2cl,(n)(V)(n)(V)\mathfrak{k}_{1}^{(n)}(V)=\Gamma_{2}^{\operatorname{cl},(n)}(V)\oplus\mathcal{E}^{(n)}(V). Then, by [48, Section 2.7, Problem 4], the exponential group K^1,(n)(V)\hat{K}_{1}^{\mathcal{E},(n)}(V) is a closed subgroup in K^1(n)(V)\hat{K}_{1}^{(n)}(V).

Next, let 𝐱C1([0,1],V)\mathbf{x}\in C^{1}([0,1],V) be a smooth loop such that 𝐱0=𝐱1=0\mathbf{x}_{0}=\mathbf{x}_{1}=0. Then, by the same reasoning as the proof of Theorem 5.21, there exists a sequence {𝐱k}k=1\{\mathbf{x}^{k}\}_{k=1}^{\infty} of piecewise linear loops such that 𝐱kLip𝐱\mathbf{x}^{k}\xrightarrow{\operatorname{Lip}}\mathbf{x} as kk\to\infty. This implies that Cone(𝐱k)Cone(𝐱){\operatorname{Cone}}(\mathbf{x}^{k})\rightarrow{\operatorname{Cone}}(\mathbf{x}) in Lipschitz norm as kk\to\infty as well. Then, since Cone(𝐱k){\operatorname{Cone}}(\mathbf{x}^{k}) is the cone of a piecewise linear loop, Proposition 5.33 shows that S1(Cone(𝐱k))K^1(V)S_{1}({\operatorname{Cone}}(\mathbf{x}^{k}))\in\hat{K}_{1}^{\mathcal{E}}(V). Finally, since the (truncated) surface signature is continuous with respect to the Lipschitz topology (Proposition 3.18), we have S1(Cone(𝐱))K^1(V)S_{1}({\operatorname{Cone}}(\mathbf{x}))\in\hat{K}_{1}^{\mathcal{E}}(V). ∎

Thus, the analogue of Proposition 5.33 holds also in the setting of smooth surfaces. Combining this with the description of K^1(V)\hat{K}_{1}^{\mathcal{E}}(V) 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

K^1(V)Γ^2(V)×T((V)),\displaystyle\hat{K}_{1}(V)\to\hat{\Gamma}_{2}(V)\times T\mkern-0.25mu\mathbin{\left(\mkern-3.5mu\left({V}\right)\mkern-3.5mu\right)}, (5.101)

from the group K^1(V)\hat{K}_{1}(V) of formal surfaces to the product of Γ^2(V)\hat{\Gamma}_{2}(V), the vector space of formal 22-currents, and T((V))T\mkern-0.25mu\mathbin{\left(\mkern-3.5mu\left({V}\right)\mkern-3.5mu\right)}, the algebra of formal non-commutative power series. With respect to this embedding, the smooth surface signature decomposes as follows

S1=(S1Γ,S1):τ2(V)Γ^2(V)×T((V)).\displaystyle S_{1}=(S_{1}^{\Gamma},S_{1}^{\mathcal{E}}):\tau_{2}(V)\to\hat{\Gamma}_{2}(V)\times T\mkern-0.25mu\mathbin{\left(\mkern-3.5mu\left({V}\right)\mkern-3.5mu\right)}. (5.102)

Given a smooth surface 𝐗τ2(V)\mathbf{X}\in\tau_{2}(V), the two components of the signature are given by

  • the path signature of the boundary path of 𝐗\mathbf{X}

    S1(𝐗)=S0((𝐗))T((V)),\displaystyle S_{1}^{\mathcal{E}}(\mathbf{X})=S_{0}(\partial(\mathbf{X}))\in T\mkern-0.25mu\mathbin{\left(\mkern-3.5mu\left({V}\right)\mkern-3.5mu\right)}, (5.103)
  • the sum of surface integrals

    S1Γ(𝐗)=αn,i<j1α!(𝒞(𝐗)zα𝑑zidzj)eαeiej\displaystyle S_{1}^{\Gamma}(\mathbf{X})=\sum_{\alpha\in\mathbb{N}^{n},\ i<j}\frac{1}{\alpha!}\left(\int_{\mathcal{C}(\mathbf{X})}z^{\alpha}dz_{i}\wedge dz_{j}\right)e^{\alpha}\otimes e_{i}\wedge e_{j} (5.104)

    where, given linear coordinates ziz_{i} on VV, the sum runs over the set of all monomial 22-forms zαdzidzjz^{\alpha}dz_{i}\wedge dz_{j} and where eαeieje^{\alpha}\otimes e_{i}\wedge e_{j} are the dual polynomial 22-currents. Furthermore, the integrals are taken over

    𝒞(𝐗)=𝐗Cone((𝐗))1τ2cl(V),\displaystyle\mathcal{C}(\mathbf{X})=\mathbf{X}\star{\operatorname{Cone}}(\partial(\mathbf{X}))^{-1}\in\tau^{\operatorname{cl}}_{2}(V), (5.105)

    the closed surface obtained by coning off the boundary of 𝐗\mathbf{X}.

Remark 5.37.

The expression for the abelian component S1Γ(𝐗)S_{1}^{\Gamma}(\mathbf{X}) of the signature of a surface 𝐗\mathbf{X} from  Theorem 5.36 has the following simple coordinate-free description. Observe that the linear structure on VV allows us to write

Ω2(V)=C(V)Λ2V,\Omega^{2}(V)=C^{\infty}(V)\otimes\Lambda^{2}V^{*},

which in turn allows us to view the surface integral as a map

𝒞(𝐗):C(V)Λ2V.\int_{\mathcal{C}(\mathbf{X})}:C^{\infty}(V)\to\Lambda^{2}V.

The identity map on VV and its powers can be viewed as functions (idV)kC(V)Sk(V)(\mathrm{id}_{V})^{k}\in C^{\infty}(V)\otimes S^{k}(V). Therefore

𝒞(𝐗)(idV)kSk(V)Λ2V=Γ¯2(V)k+2.\int_{\mathcal{C}(\mathbf{X})}(\mathrm{id}_{V})^{k}\in S^{k}(V)\otimes\Lambda^{2}V=\overline{\Gamma}_{2}(V)_{k+2}.

Then, using the multinomial theorem, the abelian component of the surface signature can be shown to have the following expression

S1Γ(𝐗)=k01k!𝒞(𝐗)(idV)k.S_{1}^{\Gamma}(\mathbf{X})=\sum_{k\geq 0}\frac{1}{k!}\int_{\mathcal{C}(\mathbf{X})}(\mathrm{id}_{V})^{k}.

6. Thin Homotopy of Piecewise Linear Surfaces

In this section, our aim is to show the following injectivity result.

Theorem 6.1.

The map SPL,1:PL1(V)K^1(V)S_{\operatorname{PL},1}:\operatorname{PL}_{1}(V)\to\hat{K}_{1}(V) is injective.

Because of the factorization of SPL,1S_{\operatorname{PL},1} in Proposition 5.20, this result has two main implications. First, it shows that our algebraic construction of PL1(V)\operatorname{PL}_{1}(V) is rich enough to faithfully encode piecewise linear surfaces modulo thin homotopy.

Corollary 6.2.

The realization map R1:PL1(V)τ2(V)R_{1}:\operatorname{PL}_{1}(V)\to\tau_{2}(V) is injective.

In particular, this implies that there are two types of cancellations that can occur via thin homotopy of PL surfaces.

  1. (1)

    Local Cancellations (Folds). When two linear components with opposite orientation are adjacent, they are cancelled via the equivalence relation in (PL1.1).

  2. (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 S1:τ2(V)K^1(V)S_{1}:\tau_{2}(V)\to\hat{K}_{1}(V) to im(R1)\operatorname{im}(R_{1}) is injective.

Remark 6.4.

Consider the 2-truncated polynomial de Rham complex of VV

Ω¯0(V)Ω¯1(V)Ω¯2(V).\displaystyle\overline{\Omega}^{0}(V)\to\overline{\Omega}^{1}(V)\to\overline{\Omega}^{2}(V). (6.1)

In order to prove a de Rham theorem for this complex, the appropriate replacement for the singular chains in degree 22 appears to be the thin homotopy group of closed piecewise linear surfaces, PL1cl(V)ker(δ:PL1(V)PL0(V))\operatorname{PL}_{1}^{\operatorname{cl}}(V)\coloneqq\ker(\delta:\operatorname{PL}_{1}(V)\to\operatorname{PL}_{0}(V)) (or its smooth analogue). Indeed, there is a well-defined pairing given by integration

,:PL1cl(V)×Ω¯2(V)dΩ¯1(V),𝐗,ω=𝐗ω.\displaystyle\langle\cdot,\cdot\rangle:\operatorname{PL}_{1}^{\operatorname{cl}}(V)\times\frac{\overline{\Omega}^{2}(V)}{d\overline{\Omega}^{1}(V)}\to\mathbb{R},\quad\quad\langle\mathbf{X},\omega\rangle=\int_{\mathbf{X}}\omega. (6.2)

Recalling the isomorphism

coker(d:Ω¯1(V)Ω¯2(V))Γ^2cl(V),\displaystyle{\operatorname{coker}}(d:\overline{\Omega}^{1}(V)\to\overline{\Omega}^{2}(V))^{*}\cong\hat{\Gamma}_{2}^{\operatorname{cl}}(V), (6.3)

where Γ^2cl(V)\hat{\Gamma}_{2}^{\operatorname{cl}}(V) denotes the closed formal 2-currents on VV, the pairing induces a map PL1cl(V)Γ^2cl(V)\operatorname{PL}_{1}^{\operatorname{cl}}(V)\to\hat{\Gamma}_{2}^{\operatorname{cl}}(V), 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 𝐆=(δG:G1G0)\mathbf{G}=(\delta_{G}:G_{1}\to G_{0}) and 𝐇=(δH:H1H0)\mathbf{H}=(\delta_{H}:H_{1}\to H_{0}) be crossed modules of groups, and let F=(F1,F0):𝐆𝐇F=(F_{1},F_{0}):\mathbf{G}\to\mathbf{H} be a morphism of crossed modules. Suppose F0:G0H0F_{0}:G_{0}\to H_{0} and F1:ker(δG)H1F_{1}:\ker(\delta_{G})\to H_{1} are injective. Then F1:G1H1F_{1}:G_{1}\to H_{1} is injective.

Proof.

Let xG1x\in G_{1} such that F1(x)=1F_{1}(x)=1. Because FF is a morphism of crossed modules, we have δHF1=F0δG\delta_{H}\circ F_{1}=F_{0}\circ\delta_{G}, so F0(δG(x))=1F_{0}(\delta_{G}(x))=1. But since F0F_{0} is injective, this implies that xker(δG)x\in\ker(\delta_{G}). By injectivity of F1F_{1} on the kernel, we have x=1x=1. ∎

By  Lemma 2.16 and the injectivity of the path signature, SPL,0S_{\operatorname{PL},0} is injective. As a result,  Lemma 6.5 reduces the proof of Theorem 6.1 to showing that

SPL,1:PL1cl(V)K^1(V)\displaystyle S_{\operatorname{PL},1}:\operatorname{PL}_{1}^{\operatorname{cl}}(V)\to\hat{K}_{1}(V) (6.4)

is injective, where PL1cl(V)=ker(δ:PL1(V)PL0(V))\operatorname{PL}_{1}^{\operatorname{cl}}(V)=\ker(\delta:\operatorname{PL}_{1}(V)\to\operatorname{PL}_{0}(V)). Given 𝐗PL1cl(V)\mathbf{X}\in\operatorname{PL}^{\operatorname{cl}}_{1}(V) such that SPL,1(𝐗)=0S_{\operatorname{PL},1}(\mathbf{X})=0, our strategy will be to consider various representatives r(𝐗)r(\mathbf{X}) of 𝐗\mathbf{X} in order to show that 𝐗=0\mathbf{X}=0 in PL1(V)\operatorname{PL}_{1}(V).

6.1. Piecewise Linear Simplicial Complexes

In this section, we consider simplicial complexes CC which can be realized in a vector space VV, and explain how to construct maps from their fundamental crossed modules to 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V).

Definition 6.6.

A piecewise linear simplicial complex (PLSC) in VV is a finite ordered simplicial complex CC whose vertices lie in VV. In particular, the set of vertices C0VC_{0}\subset V is equipped with a total ordering. Each nn-simplex σ\sigma of CC 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. σ=[p0,,pn]\sigma=[p_{0},\ldots,p_{n}] of points piVp_{i}\in V, and hence, is equipped with a linear characteristic map Φσ:ΔσnV\Phi_{\sigma}:\Delta^{n}_{\sigma}\to V. Let |σ|V|\sigma|\subset V denote the image of Φσ\Phi_{\sigma}. It is the convex hull of the points of σ\sigma in VV. We say that an nn-simplex σ\sigma is degenerate if there exists some (n1)(n-1)-dimensional affine hyperplane UVU\subset V such that |σ|U|\sigma|\subset U, and we say that σ\sigma is non-degenerate otherwise. We say that CC is non-degenerate if all its simplices are non-degenerate.

We will abuse notation and use CC to refer both to the abstract simplicial complex as well as its geometric realization as a topological space, and we let CnC_{n} denote the nn-skeleton of CC. Furthermore, we use |C||C| to denote the piecewise linear realization of CC in VV, defined as the union of the subsets |σ||\sigma|

|C|σC|σ|V.\displaystyle|C|\coloneqq\bigcup_{\sigma\in C}|\sigma|\subset V. (6.5)

The complex is equipped with a continuous realization map Φ:C|C|.\Phi:C\to|C|.

Remark 6.7.

Note that because a 1-simplex is a pair of distinct points [p0,p1][p_{0},p_{1}], piVp_{i}\in V, it is automatically non-degenerate.

Furthermore, we will require a notion of compatible complexes.

Definition 6.8.

Suppose CC is a PLSC in VV. We say that CC is compatible if for every pair of simplices σ,σC\sigma,\sigma^{\prime}\in C, their intersection in VV satisfies

|σ||σ|=|τ|\displaystyle|\sigma|\cap|\sigma^{\prime}|=|\tau| (6.6)

where τσσ\tau\subseteq\sigma\cap\sigma^{\prime} is a common subsimplex of both σ\sigma and σ\sigma^{\prime}, or τ=\tau=\emptyset in which case |σ||\sigma| and |σ||\sigma^{\prime}| are disjoint.

[Uncaptioned image]
Lemma 6.9.

Let CC be a compatible non-degenerate PLSC. Then the realization map Φ:C|C|\Phi:C\to|C| is a homeomorphism.

Proof.

Because CC is compact and |C||C| is Hausdorff, it suffices to prove injectivity. For each simplex σ\sigma, the map Φ:Δσ|σ|\Phi:\Delta_{\sigma}\to|\sigma| is injective since σ\sigma is nondegenerate. Hence, let pΔσp\in\Delta_{\sigma}, qΔσq\in\Delta_{\sigma}^{\prime} be points in different simplices such that x=Φ(p)=Φ(q)x=\Phi(p)=\Phi(q). Hence, |σ||\sigma| and |σ||\sigma^{\prime}| intersect non-trivially. Therefore, there is a subsimplex τ\tau of both σ\sigma and σ\sigma^{\prime} such that |τ|=|σ||σ||\tau|=|\sigma|\cap|\sigma^{\prime}|. Thus, there is a point rΔτr\in\Delta_{\tau} such that Φ(r)=x\Phi(r)=x. Because τσ\tau\subseteq\sigma is a subsimplex we have rΔσr\in\Delta_{\sigma}. By injectivity of Φ|Δσ\Phi|_{\Delta_{\sigma}}, we conclude that r=pr=p. Similarly, r=qr=q. Therefore, p=qp=q and Φ\Phi is injective. ∎

Let CC be a 2-dimensional PLSC. We denote the set of vertices by C0C_{0}, the set of 11-simplices by EE, and the set of 22-simplices by LL. The fundamental groupoid of the 11-skeleton, Π1(C1,C0)\Pi_{1}(C_{1},C_{0}), is free on the set of edges EE. We define a homomorphism

W~0:Π1(C1,C0)PL0(V)\displaystyle\widetilde{W}_{0}:\Pi_{1}(C_{1},C_{0})\to\operatorname{PL}_{0}(V) (6.7)

by sending an edge ϵ=[p0,p1]\epsilon=[p_{0},p_{1}] to the element ηV(p1p0)PL0(V)\eta_{V}(p_{1}-p_{0})\in\operatorname{PL}_{0}(V). Composing with the realization map R0:PL0(V)τ1(V)R_{0}:\operatorname{PL}_{0}(V)\to\tau_{1}(V), the edge ϵ\epsilon gets sent to R0W~0(ϵ)R_{0}\circ\widetilde{W}_{0}(\epsilon), the straight line path in VV from 0 to p1p0p_{1}-p_{0}. Let c0Vc_{0}\in V be a basepoint. Restricting W~0\widetilde{W}_{0} to the fundamental group based at c0c_{0} gives rise to a group homomorphism

W0:π1(C1,c0)PL0(V).\displaystyle W_{0}:\pi_{1}(C_{1},c_{0})\to\operatorname{PL}_{0}(V). (6.8)
Lemma 6.10.

Suppose CC is a PLSC whose 11-skeleton C1C_{1} is compatible and non-degenerate. Then the map W0:π1(C1,c0)PL0(V)W_{0}:\pi_{1}(C_{1},c_{0})\to\operatorname{PL}_{0}(V) is injective.

Proof.

Let γπ1(C1,c0)\gamma\in\pi_{1}(C_{1},c_{0}) be such that W0(γ)=0W_{0}(\gamma)=0. Then the realization R0W0(γ)τ1(V)R_{0}\circ W_{0}(\gamma)\in\tau_{1}(V) is a loop in |C1||C_{1}| which is thinly null homotopic in VV. By the image condition (I1), there exists a null homotopy hh whose image is contained in the image of R0W0(γ)R_{0}\circ W_{0}(\gamma), and therefore in |C1||C_{1}|. Because the map Φ:C1|C1|\Phi:C_{1}\to|C_{1}| is a homeomorphism by Lemma 6.9, the null homotopy hh can be lifted to C1C_{1}. As a result, γ=0\gamma=0. ∎

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 (C,c0)(C,c_{0}) be a connected based 2-dimensional CW complex with 1-cells EE and 2-cells LL. Fix a spanning tree TC1T\subset C_{1} of the 1-skeleton. For each 2-cell λL\lambda\in L, choose a basepoint cλC0c_{\lambda}\in C_{0}, and let ϕλπ1(C1,cλ)\phi_{\lambda}\in\pi_{1}(C_{1},c_{\lambda}) be the attaching map of the cell. Furthermore, let ωλΠ1(C1,C0)\omega_{\lambda}\in\Pi_{1}(C_{1},C_{0}) be the unique path in TT connecting c0c_{0} to cλc_{\lambda}. Define ρ:Lπ1(C1,c0)\rho:L\to\pi_{1}(C_{1},c_{0}) by ρ(λ)=ωλϕλωλ1\rho(\lambda)=\omega_{\lambda}\star\phi_{\lambda}\star\omega_{\lambda}^{-1}. Then, the fundamental crossed module of the pair (C,C1)(C,C_{1})

𝝅(C,C1)=(:π2(C,C1,c0)π1(C1,c0))\displaystyle\boldsymbol{\pi}(C,C_{1})=\Big{(}\partial:\pi_{2}(C,C_{1},c_{0})\to\pi_{1}(C_{1},c_{0})\Big{)} (6.9)

is the free crossed module on ρ\rho. Let η:Lπ2(C,C1,c0)\eta:L\to\pi_{2}(C,C_{1},c_{0}) denote the map which sends each 2-cell λL\lambda\in L to the corresponding generator η(λ)\eta(\lambda) of π2(C,C1,c0)\pi_{2}(C,C_{1},c_{0}).

Let CC be a 22-dimensional PLSC. We will use Whitehead’s theorem to construct a map 𝐖:𝝅(C,C1)𝐏𝐋(V)\mathbf{W}:\boldsymbol{\pi}(C,C_{1})\to{\boldsymbol{\operatorname{PL}}}(V). First, choose a basepoint c0Vc_{0}\in V and fix a spanning tree TC1T\subset C_{1} of the 11-skeleton. Let W0W_{0} be the homomorphism from  (6.8). By  Theorem 6.11, 𝝅(C,C1)\boldsymbol{\pi}(C,C_{1}) is free on the set of 2-cells LL of CC, and so by the universal property of Theorem 5.1, the map 𝐖\mathbf{W} is specified by a function f:LPL1(V)f:L\to\operatorname{PL}_{1}(V) such that δf=W0ρ\delta\circ f=W_{0}\circ\rho.

Each 2-simplex λL\lambda\in L has the form λ=[p0,p1,p2]\lambda=[p_{0},p_{1},p_{2}]. Let ϕλπ1(C1,p0)\phi_{\lambda}\in\pi_{1}(C_{1},p_{0}) be its boundary loop. Under W~0\widetilde{W}_{0} it gets sent to the following planar loop

𝐛λ=W~0(ϕλ)=(p1p0,p2p1,p0p2)PlanarLoop(V).\displaystyle\mathbf{b}_{\lambda}=\widetilde{W}_{0}(\phi_{\lambda})=(p_{1}-p_{0},p_{2}-p_{1},p_{0}-p_{2})\in{\operatorname{PlanarLoop}}(V). (6.10)

Let ωλΠ(C1,C0)\omega_{\lambda}\in\Pi(C_{1},C_{0}) be the unique path in TT connecting c0c_{0} to p0p_{0}, and let 𝐰λ=W~0(ωλ)PL0(V)\mathbf{w}_{\lambda}=\widetilde{W}_{0}(\omega_{\lambda})\in\operatorname{PL}_{0}(V). Then (𝐰λ,𝐛λ)Kite(V)(\mathbf{w}_{\lambda},\mathbf{b}_{\lambda})\in{\operatorname{Kite}}(V) is a kite and we define

f:LPL1(V),λ(𝐰λ,𝐛λ).\displaystyle f:L\to\operatorname{PL}_{1}(V),\qquad\lambda\mapsto(\mathbf{w}_{\lambda},\mathbf{b}_{\lambda}). (6.11)

It is then clear from the construction that δf(λ)=W0(ωλϕλωλ1)=W0ρ(λ)\delta\circ f(\lambda)=W_{0}(\omega_{\lambda}\star\phi_{\lambda}\star\omega_{\lambda}^{-1})=W_{0}\circ\rho(\lambda). We therefore obtain the desired morphism 𝐖\mathbf{W}.

Corollary 6.12.

Let CC be a 2-dimensional connected PLSC in VV. The maps ff and W0W_{0} above induce a morphism of crossed modules 𝐖=(W1,W0):𝛑(C,C1)𝐏𝐋(V)\mathbf{W}=(W_{1},W_{0}):\boldsymbol{\pi}(C,C_{1})\to{\boldsymbol{\operatorname{PL}}}(V) such that f=W1ηf=W_{1}\circ\eta, where η:Lπ2(C,C1)\eta:L\to\pi_{2}(C,C_{1}) 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 𝐗PL1cl(V)\mathbf{X}\in\operatorname{PL}_{1}^{\operatorname{cl}}(V) such that SPL,1(𝐗)=0S_{\operatorname{PL},1}(\mathbf{X})=0. The proof consists of three main steps.

  1. (1)

    Construct an appropriate representative r(𝐗)=(X1,,Xk)r(\mathbf{X})=(X_{1},\ldots,X_{k}) which induces a compatible PLSC CC (Definition 6.8). Then construct an element Yπ2(C)Y\in\pi_{2}(C) such that W1(Y)=𝐗W_{1}(Y)=\mathbf{X}.

  2. (2)

    The trivial signature condition will provide a matching of kites of opposite orientation, implying that under the Hurewicz map :π2(C)H2(C)\mathcal{H}:\pi_{2}(C)\to H_{2}(C), we have (Y)=0\mathcal{H}(Y)=0.

  3. (3)

    Construct a simply connected PLSC C^\widehat{C} such that CC^C\hookrightarrow\widehat{C}, and apply the Hurewicz theorem to show that W1(Y)=0W_{1}(Y)=0.

6.2.1. Construct Simplicial Model

Our first step is to construct a simplicial model for 𝐗\mathbf{X}. This is a PLSC CC whose fundamental crossed module contains an element YY which gets sent to 𝐗\mathbf{X} under the map W1W_{1} constructed in Corollary 6.12. Such a simplicial model will be induced by a triangulated representative of 𝐗\mathbf{X}. In order to define this representative, we begin by defining the set of marked kites

Kite×(V)𝖥𝖬𝗈𝗇(V)×PlanarLoop(V),\displaystyle{\operatorname{Kite}}^{\times}(V)\coloneqq\mathsf{FMon}(V)\times{\operatorname{PlanarLoop}}(V), (6.12)

where 𝖥𝖬𝗈𝗇(V)\mathsf{FMon}(V) is the free monoid generated by VV. 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

𝖥𝖬𝗈𝗇(Kite×(V))𝖥𝖬𝗈𝗇(Kite(V))PL1(V).\displaystyle\mathsf{FMon}({\operatorname{Kite}}^{\times}(V))\to\mathsf{FMon}({\operatorname{Kite}}(V))\to\operatorname{PL}_{1}(V). (6.13)
Definition 6.13.

A marked representative (or simply representative) of an element 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V) is a lift of this element to 𝖥𝖬𝗈𝗇(Kite×(V))\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)). A marked kite (𝐰,𝐛)Kite×(V)(\mathbf{w},\mathbf{b})\in{\operatorname{Kite}}^{\times}(V) is triangular if 𝐛=(b1,b2,b3)min\mathbf{b}=(b_{1},b_{2},b_{3})_{\min}. Therefore, a representative (X1,,Xk)𝖥𝖬𝗈𝗇(Kite×(V))(X_{1},...,X_{k})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) of 𝐗\mathbf{X} is said to be triangulated if each marked kite XiX_{i} is triangular.

We now explain how to construct a PLSC CC given the data of a triangulated representative. First, let X=(𝐰,𝐛)Kite×(V)X=(\mathbf{w},\mathbf{b})\in{\operatorname{Kite}}^{\times}(V) be a marked triangular kite. Therefore 𝐰=(w1,,wm)𝖥𝖬𝗈𝗇(V)\mathbf{w}=(w_{1},...,w_{m})\in\mathsf{FMon}(V) and 𝐛=(b1,b2,b3)minPL0cl(V)\mathbf{b}=(b_{1},b_{2},b_{3})_{\min}\in\operatorname{PL}_{0}^{\operatorname{cl}}(V). Define w^0=0V\hat{w}_{0}=0\in V, and given 1km1\leq k\leq m, define w^k=i=1kwi.\hat{w}_{k}=\sum_{i=1}^{k}w_{i}. Finally, define

b^1=w^m+b1andb^2=b^1+b2.\displaystyle\hat{b}_{1}=\hat{w}_{m}+b_{1}\quad\text{and}\quad\hat{b}_{2}=\hat{b}_{1}+b_{2}. (6.14)

The piecewise linear simplicial complex associated to XX, denoted Δ(X)\Delta(X), is given by

Δ0(X)\displaystyle\Delta_{0}(X) ={w^0,w^1,,w^m,b^1,b^2}\displaystyle=\{\hat{w}_{0},\hat{w}_{1},\ldots,\hat{w}_{m},\hat{b}_{1},\hat{b}_{2}\} (6.15)
Δ1(X)\displaystyle\Delta_{1}(X) ={[w^i,w^i+1]}i=0m1{[w^m,b^1],[b^1,b^2],[b^2,w^m]}\displaystyle=\{[\hat{w}_{i},\hat{w}_{i+1}]\}_{i=0}^{m-1}\cup\{[\hat{w}_{m},\hat{b}_{1}],[\hat{b}_{1},\hat{b}_{2}],[\hat{b}_{2},\hat{w}_{m}]\} (6.16)
Δ2(X)\displaystyle\Delta_{2}(X) ={[w^m,b^1,b^2]}.\displaystyle=\{[\hat{w}_{m},\hat{b}_{1},\hat{b}_{2}]\}. (6.17)

More precisely, we take the set of vertices to be C0=Δ0(X)VC_{0}=\Delta_{0}(X)\subset V, with the induced order. Since this is a subset of VV, any repetitions are automatically deleted. The simplicial complex is the union of Δ0(X),Δ1(X),\Delta_{0}(X),\Delta_{1}(X), and Δ2(X)\Delta_{2}(X) in the power set of C0C_{0}. As a result, any repetitions are removed, regardless of orientation or ordering in the above description. The 2-simplex associated to XX is the unique 2-simplex σX=[w^m,b^1,b^2]\sigma_{X}=[\hat{w}_{m},\hat{b}_{1},\hat{b}_{2}] of Δ(X)\Delta(X). Note that σX\sigma_{X} is non-degenerate and that Δ(X)\Delta(X) is connected.

Next, let r(𝐗)=(X1,,Xk)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})=(X_{1},...,X_{k})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) be a triangulated representative. The PLSC associated to r(𝐗)r(\mathbf{X}), denoted Δ(r(𝐗))\Delta(r(\mathbf{X})), is given by the union

Δ(r(𝐗))j=1kΔ(Xj).\displaystyle\Delta(r(\mathbf{X}))\coloneqq\bigcup_{j=1}^{k}\Delta(X_{j}). (6.18)

More precisely, the set of vertices C0VC_{0}\subset V is the union of the vertices from each Δ(Xj)\Delta(X_{j}), with the ordering such that the new vertices of Δ(Xj+1)\Delta(X_{j+1}) come after those of Δ(Xj)\Delta(X_{j}). The simplicial complex is then the union of each Δ(Xj)\Delta(X_{j}) in the power set of C0C_{0}. Note that Δ(r(𝐗))\Delta(r(\mathbf{X})) is connected because each complex Δ(Xi)\Delta(X_{i}) 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 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V).

Definition 6.14.

Let 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V). A compatible representative is a triangulated representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) such that Δ(r(𝐗))\Delta(r(\mathbf{X})) is compatible in the sense of Definition 6.8.

Remark 6.15.

Recall from Remark 6.7 that a 11-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 Δ(r(𝐗))\Delta(r(\mathbf{X})) is automatically non-degenerate.

We prove the following in Appendix D by carefully considering subdivisions.

Theorem 6.16.

There exists a compatible representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) of every 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V).

Now let 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V), fix a compatible representative r(𝐗)r(\mathbf{X}), and let CΔ(r(𝐗))C\coloneqq\Delta(r(\mathbf{X})) be the associated PLSC. Equip CC with the basepoint c0=0Vc_{0}=0\in V. All homotopy groups will be taken with respect to this basepoint. By Corollary 6.12, we may construct a morphism of crossed modules 𝐖=(W1,W0):𝝅(C,C1)𝐏𝐋(V)\mathbf{W}=(W_{1},W_{0}):\boldsymbol{\pi}(C,C_{1})\to{\boldsymbol{\operatorname{PL}}}(V). Since CC is compatible and non-degenerate, the map W0W_{0} is injective by Lemma 6.10. Next, we lift 𝐗\mathbf{X} to an element of 𝝅(C,C1)\boldsymbol{\pi}(C,C_{1}).

Proposition 6.17.

There exists an element Yπ2(C,C1)Y\in\pi_{2}(C,C_{1}) such that W1(Y)=𝐗W_{1}(Y)=\mathbf{X}. Furthermore, if SPL,1(𝐗)=0S_{\operatorname{PL},1}(\mathbf{X})=0, then Yπ2(C)Y\in\pi_{2}(C).

Proof.

The representative r(𝐗)=(X1,,Xk)r(\mathbf{X})=(X_{1},...,X_{k}) allows us to factor 𝐗\mathbf{X} into a product of kites in PL1(V)\operatorname{PL}_{1}(V). Hence, it suffices to show that each Xi=(𝐰i,𝐛i)X_{i}=(\mathbf{w}_{i},\mathbf{b}_{i}) is in the image of W1W_{1}. Each marked triangular kite XiX_{i} has an associated 2-simplex σXiC\sigma_{X_{i}}\in C and hence determines a generator η(σXi)π2(C,C1)\eta(\sigma_{X_{i}})\in\pi_{2}(C,C_{1}). Under the map W1W_{1} this generator is sent to

W1(η(σXi)si)=𝐯i𝐛i,\displaystyle W_{1}(\eta(\sigma_{X_{i}})^{s_{i}})=\mathbf{v}_{i}\vartriangleright\mathbf{b}_{i}, (6.19)

where si=±1s_{i}=\pm 1 is an appropriate sign and where 𝐯i=W~0(νi)\mathbf{v}_{i}=\widetilde{W}_{0}(\nu_{i}), for νiΠ1(C1,C0)\nu_{i}\in\Pi_{1}(C_{1},C_{0}) a path in C1C_{1} connecting c0c_{0} to the basepoint pip_{i} of 𝐛i\mathbf{b}_{i}. The element 𝐰i\mathbf{w}_{i} can also be realized as 𝐰i=W~0(ωi)\mathbf{w}_{i}=\widetilde{W}_{0}(\omega_{i}), for ωiΠ1(C1,C0)\omega_{i}\in\Pi_{1}(C_{1},C_{0}) a path in C1C_{1} connecting c0c_{0} to pip_{i}. Therefore, ωiνi1π1(C1)\omega_{i}\star\nu_{i}^{-1}\in\pi_{1}(C_{1}) and

W1((ωiνi1)η(σXi)si)=(𝐰i𝐯i1)(𝐯i𝐛i)=Xi.\displaystyle W_{1}((\omega_{i}\star\nu_{i}^{-1})\vartriangleright\eta(\sigma_{X_{i}})^{s_{i}})=(\mathbf{w}_{i}\star\mathbf{v}_{i}^{-1})\vartriangleright(\mathbf{v}_{i}\vartriangleright\mathbf{b}_{i})=X_{i}. (6.20)

This shows that a preimage YY exists. By the injectivity of the path signature and SPL,1(𝐗)=0S_{\operatorname{PL},1}(\mathbf{X})=0, we have δ(𝐗)=0\delta(\mathbf{X})=0. Thus, W0(Y)=0W_{0}(\partial Y)=0, and so Y=0\partial Y=0 by the injectivity of W0W_{0}. Hence Yπ2(C)Y\in\pi_{2}(C). ∎

6.2.2. Matching kites

Consider the Hurewicz map

:π2(C)H2(C).\displaystyle\mathcal{H}:\pi_{2}(C)\to H_{2}(C). (6.21)

Applying it to the element Yπ2(C)Y\in\pi_{2}(C) constructed in Proposition 6.17, we obtain an element (Y)H2(C)\mathcal{H}(Y)\in H_{2}(C). The next step is to show that this element is 0.

Let LL denote the set of 2-simplices of CC. The cellular chain complex of CC is given by

C2(C)=λLλ.C_{2}(C)=\bigoplus_{\lambda\in L}\mathbb{Z}\lambda.

Given the representative r(𝐗)=(X1,,Xk)r(\mathbf{X})=(X_{1},...,X_{k}), each marked triangular kite XiX_{i} has an associated 2-simplex σXi\sigma_{X_{i}} which is an element of LL. This defines a function α:[k]L\alpha:[k]\to L sending the index ii to σXi\sigma_{X_{i}}. Since the ordering on the triangular loop in XiX_{i} might not agree with the ordering of simplices in LL, we must also record this information as orientation data. This is a function s:[k]{±1}s:[k]\to\{\pm 1\} such that s(i)=+1s(i)=+1 if and only if the ordering on the triangular loop in XiX_{i} matches the fixed order on α(i)\alpha(i). We call the pair of functions (α,s)(\alpha,s) the simplex mapping of r(𝐗)r(\mathbf{X}) and we note that

(Y)=i=1ks(i)α(i).\displaystyle\mathcal{H}(Y)=\sum_{i=1}^{k}s(i)\alpha(i). (6.22)

The pair (α,s)(\alpha,s) is a simplex matching if (Y)=0\mathcal{H}(Y)=0, since this means that the simplices in r(𝐗)r(\mathbf{X}) are “matched up” in pairs of opposite orientation. By hypothesis, the signature of the realization of 𝐗\mathbf{X} is trivial, S1(R1(𝐗))=0S_{1}(R_{1}(\mathbf{X}))=0, and by Corollary 4.13, this implies that

R1(𝐗)ω=0\displaystyle\int_{R_{1}(\mathbf{X})}\omega=0 (6.23)

for all ωΩc2(V)\omega\in\Omega^{2}_{c}(V). We will use this fact to show that (α,s)(\alpha,s) is indeed a simplex matching.

Proposition 6.18.

Let Yπ2(C)Y\in\pi_{2}(C) be the element from Proposition 6.17. Then (Y)=0\mathcal{H}(Y)=0.

Proof.

Let ωΩc2(V)\omega\in\Omega^{2}_{c}(V) be a compactly supported form. Then

R1(𝐗)ω=R1(W1(Y))ω=i=1ks(i)α(i)ω.\displaystyle\int_{R_{1}(\mathbf{X})}\omega=\int_{R_{1}(W_{1}(Y))}\omega=\sum_{i=1}^{k}s(i)\int_{\alpha(i)}\omega. (6.24)

Given a 2-simplex σL\sigma\in L, let ωσΩc2(V)\omega_{\sigma}\in\Omega_{c}^{2}(V) be a compactly supported form with the property that σωσ=1\int_{\sigma}\omega_{\sigma}=1 and whose support is disjoint from all other simplices. Such a form exists by Lemma D.6. Then

i=1ks(i)α(i)ω=i:α(i)=σs(i).\displaystyle\sum_{i=1}^{k}s(i)\int_{\alpha(i)}\omega=\sum_{i\ :\ \alpha(i)=\sigma}s(i). (6.25)

This sum vanishes by  (6.23). Therefore (Y)=0\mathcal{H}(Y)=0. ∎

6.2.3. Show that W1(Y)=0W_{1}(Y)=0.

Our objective now is to apply the Hurewicz theorem to show that W1(Y)=0W_{1}(Y)=0. While CC may not be simply-connected in general, we can add 2-cells to kill off the fundamental group.

Lemma 6.19.

Let CC be a 2-dimensional, connected PLSC in VV. There exists a 2-dimensional PLSC C^\widehat{C} such that CC^C\subset\widehat{C} and π1(C^)=0\pi_{1}(\widehat{C})=0.

Proof.

Let C0={p0,,pn}C_{0}=\{p_{0},...,p_{n}\} be the set of vertices of CC and let EE be the set of 1-simplices. Choose a point xVC0x\in V-C_{0} such that for each edge ϵ=[pi,pj]E\epsilon=[p_{i},p_{j}]\in E the triple {x,pi,pj}\{x,p_{i},p_{j}\} is not contained in a line. For each vertex piC0p_{i}\in C_{0}, define a new 1-simplex ϵ^(pi)=[x,pi]\widehat{\epsilon}(p_{i})=[x,p_{i}], and for each edge ϵ=[pi,pj]E\epsilon=[p_{i},p_{j}]\in E, define a new 2-simplex λ^(ϵ)=[x,pi,pj]\widehat{\lambda}(\epsilon)=[x,p_{i},p_{j}]. Now define a new PLSC DD with vertex set D0=C0{x}D_{0}=C_{0}\cup\{x\}, set of 1-simplices E{ϵ^(pi)}piC0E\cup\{\widehat{\epsilon}(p_{i})\}_{p_{i}\in C_{0}}, and set of 2-simplices {λ^(ϵ)}ϵE\{\widehat{\lambda}(\epsilon)\}_{\epsilon\in E}. Extend the order on C0C_{0} to D0D_{0} so that xx comes after all other vertices. By construction, DD is contractible.

Now define C^\widehat{C} to be the union of CC and DD. This is a PLSC that contains CC and DD as subcomplexes. Furthermore, CC and DD intersect along C1C_{1}, the 11-skeleton of CC. Therefore, by the van Kampen theorem, π1(C^)=0\pi_{1}(\widehat{C})=0. ∎

Next, we prove a general relationship between the kernels of W1W_{1} and the Hurewicz map which will imply our main injectivity result.

Proposition 6.20.

Let CC be a 2-dimensional connected, compatible, and non-degenerate PLSC in VV. Let W1:π2(C,C1)PL1(V)W_{1}:\pi_{2}(C,C_{1})\to\operatorname{PL}_{1}(V) be the homomorphism defined in Corollary 6.12, and let :π2(C)H2(C)\mathcal{H}:\pi_{2}(C)\to H_{2}(C) be the Hurewicz map from (6.21). Then

ker()=ker(W1).\displaystyle\ker(\mathcal{H})=\ker(W_{1}). (6.26)
Proof.

First, we show ker()ker(W1)\ker(\mathcal{H})\subset\ker(W_{1}). Let C^\widehat{C} be the PLSC from Lemma 6.19 with inclusion map ι:CC^\iota:C\hookrightarrow\widehat{C}. The map ι\iota induces a map 𝝅(ι):𝝅(C,C1)𝝅(C^,C^1)\boldsymbol{\pi}(\iota):\boldsymbol{\pi}(C,C_{1})\to\boldsymbol{\pi}(\widehat{C},\widehat{C}_{1}). Applying Corollary 6.12 to C^\widehat{C}, we have another morphism of crossed modules 𝐖^=(W^1,W^0):𝝅(C^,C^1)𝐏𝐋(V)\widehat{\mathbf{W}}=(\widehat{W}_{1},\widehat{W}_{0}):\boldsymbol{\pi}(\widehat{C},\widehat{C}_{1})\to{\boldsymbol{\operatorname{PL}}}(V) such that the following diagram commutes,

𝝅(C,C1){\boldsymbol{\pi}(C,C_{1})}𝝅(C^,C^1){\boldsymbol{\pi}(\widehat{C},\widehat{C}_{1})}𝐏𝐋(V){{\boldsymbol{\operatorname{PL}}}(V)}.𝝅(ι)\scriptstyle{\boldsymbol{\pi}(\iota)}𝐖\scriptstyle{\mathbf{W}}𝐖^\scriptstyle{\widehat{\mathbf{W}}} (6.27)

Indeed, examining the construction of W0W_{0} in (6.8), it is clear that W^0π1(ι)=W0\widehat{W}_{0}\circ\pi_{1}(\iota)=W_{0}. The construction of W1W_{1} leading up to Corollary 6.12 depends on the choice of a spanning tree TC1T\subset C_{1}. Therefore, to ensure that W^1π2(ι)=W1\widehat{W}_{1}\circ\pi_{2}(\iota)=W_{1}, we choose the spanning tree T^C^1\widehat{T}\subset\widehat{C}_{1} to be an extension of TT. This implies that the chosen generators of π2(C,C1)\pi_{2}(C,C_{1}) are sent to the chosen generators of π2(C^,C^1)\pi_{2}(\widehat{C},\widehat{C}_{1}) under the map π2(ι)\pi_{2}(\iota).

Next, by naturality of the Hurewicz map, the following diagram commutes.

π2(C){\pi_{2}(C)}π2(C^){\pi_{2}(\widehat{C})}H2(C){H_{2}(C)}H2(C^),{H_{2}(\widehat{C}),}π2(ι)\scriptstyle{\pi_{2}(\iota)}\scriptstyle{\mathcal{H}}\scriptstyle{\mathcal{H}}H2(ι)\scriptstyle{H_{2}(\iota)} (6.28)

Let Yker()π2(C)Y\in\ker(\mathcal{H})\subset\pi_{2}(C). Because :π2(C^)H2(C^)\mathcal{H}:\pi_{2}(\widehat{C})\to H_{2}(\widehat{C}) is an isomorphism by the Hurewicz theorem, and (Y)=0\mathcal{H}(Y)=0, we have π2(ι)(Y)=0\pi_{2}(\iota)(Y)=0. Therefore, this implies that W1(Y)=W^1π2(ι)(Y)=0W_{1}(Y)=\widehat{W}_{1}\circ\pi_{2}(\iota)(Y)=0.

It remains to show that ker(W1)ker()\ker(W_{1})\subset\ker(\mathcal{H}). Suppose Yker(W1)Y\in\ker(W_{1}). Then, we have

W0(Y)=δW1(Y)=0.\displaystyle W_{0}\circ\partial(Y)=\delta\circ W_{1}(Y)=0. (6.29)

Since CC is compatible and non-degenerate, W0W_{0} is injective by Lemma 6.10. Hence it follows that Yπ2(C)Y\in\pi_{2}(C). Furthermore, the signature SPL,1W1(Y)=0S_{\operatorname{PL},1}\circ W_{1}(Y)=0 is trivial. Then, the same argument as Proposition 6.18 shows that (Y)=0\mathcal{H}(Y)=0.

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 𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V). In contrast to Proposition 2.18, Proposition 6.20 shows that this is not possible in general for crossed modules, as the kernel ker(W1)\ker(W_{1}) 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 SPL,1:PL1cl(V)K^1(V)S_{\operatorname{PL},1}:\operatorname{PL}_{1}^{\operatorname{cl}}(V)\to\hat{K}_{1}(V) is injective. Let 𝐗PL1cl(V)\mathbf{X}\in\operatorname{PL}_{1}^{\operatorname{cl}}(V) such that SPL,1(𝐗)=0S_{\operatorname{PL},1}(\mathbf{X})=0. Let r(𝐗)r(\mathbf{X}) be a compatible representative of 𝐗\mathbf{X}, C=Δ(r(𝐗))C=\Delta(r(\mathbf{X})), and Yπ2(C)Y\in\pi_{2}(C) be element from Proposition 6.17 such that W1(Y)=𝐗W_{1}(Y)=\mathbf{X}. By Proposition 6.18, we have (Y)=0\mathcal{H}(Y)=0, and finally by Proposition 6.20, W1(Y)=0W_{1}(Y)=0, so 𝐗=0\mathbf{X}=0. ∎

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 𝐗C1([0,1]2,V)\mathbf{X}\in C^{1}([0,1]^{2},V) such that

  • only the top boundary is non-trivial, l𝐗=b𝐗=r𝐗=0\partial_{l}\mathbf{X}=\partial_{b}\mathbf{X}=\partial_{r}\mathbf{X}=0, and

  • there exists a compatible PLSC TT in 2\mathbb{R}^{2} such that |T|=[0,1]2|T|=[0,1]^{2}, and the restriction of 𝐗\mathbf{X} to any 22-simplex σT2\sigma\in T_{2} is the composition of a smooth reparametrization ψσ:|σ||σ|\psi_{\sigma}:|\sigma|\to|\sigma| with sitting instants, and an affine linear function fσ:|σ|Vf_{\sigma}:|\sigma|\to V,

    𝐗||σ|:|σ|ψσ|σ|fσV.\displaystyle\mathbf{X}|_{|\sigma|}:|\sigma|\xrightarrow{\psi_{\sigma}}|\sigma|\xrightarrow{f_{\sigma}}V. (6.30)

We denote the space of smooth piecewise linear surfaces by CPL1([0,1]2,V)C_{\operatorname{PL}}^{1}([0,1]^{2},V).

The following lemma shows that CPL1([0,1]2,V)C^{1}_{\operatorname{PL}}([0,1]^{2},V) consists of surfaces which lie in the thin homotopy classes defined by the realization of PL1(V)\operatorname{PL}_{1}(V).

Lemma 6.23.

Let 𝐗CPL1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{PL}}([0,1]^{2},V). There exists some 𝐙PL1(V)\mathbf{Z}\in\operatorname{PL}_{1}(V) such that 𝐗\mathbf{X} is in the thin homotopy class of R1(𝐙)R_{1}(\mathbf{Z}).

Proof.

Let P=((1,0),(0,1),(1,0),(0,1))PL0(2)P=\Big{(}(1,0),(0,1),(-1,0),(0,-1)\Big{)}\in\operatorname{PL}_{0}(\mathbb{R}^{2}) be the boundary of the unit square. Let TT be a compatible PLSC in 2\mathbb{R}^{2} which satisfies the conditions of Definition 6.22, and note that it is a compatible triangulation of [0,1]2[0,1]^{2} in the sense of Definition D.2. Then, viewing (,P)Kite×(2)(\emptyset,P)\in{\operatorname{Kite}}^{\times}(\mathbb{R}^{2}) as a kite, the proof of Lemma D.5 implies there exists a compatible representative 𝐘=(Y1,,Yk)𝖥𝖬𝗈𝗇(Kite×(2))\mathbf{Y}=(Y_{1},\ldots,Y_{k})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(\mathbb{R}^{2})) of (,P)PL1(2)(\emptyset,P)\in\operatorname{PL}_{1}(\mathbb{R}^{2}). Then, applying the maps fσf_{\sigma} from Definition 6.22 to 𝐘\mathbf{Y}, we obtain an element 𝐙=(Z1,,Zk)PL1(V)\mathbf{Z}=(Z_{1},\ldots,Z_{k})\in\operatorname{PL}_{1}(V). Finally, by Lemma 5.16, the identity map I:[0,1]22I:[0,1]^{2}\to\mathbb{R}^{2} is in the thin homotopy class of R1(𝐘)τ2(2)R_{1}(\mathbf{Y})\in\tau_{2}(\mathbb{R}^{2}), so 𝐗\mathbf{X} is in the thin homotopy class of R1(𝐙)τ2(V)R_{1}(\mathbf{Z})\in\tau_{2}(V). ∎

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 𝐗CPL1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{PL}}([0,1]^{2},V), there exists a representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) by Lemma 6.23. We say that the surface 𝐗\mathbf{X} is word reduced if the representative r(𝐗)r(\mathbf{X}) is trivial in PL1(V)\operatorname{PL}_{1}(V). 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 𝖌=(δ:𝔤1𝔤0,)𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright)\in\mathsf{XLie} is holonomy nondegenerate if

  1. (1)

    there exists a semisimple Lie algebra 𝔥\mathfrak{h} such that 𝔥im(δ)\mathfrak{h}\subset\operatorname{im}(\delta), and

  2. (2)

    ker(δ)\ker(\delta) is nontrivial.

A crossed module of Lie groups 𝐆𝖷𝖫𝖦𝗋𝗉\mathbf{G}\in\mathsf{XLGrp} is holonomy nondegenerate if its associated crossed module of Lie algebras 𝖌𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}\in\mathsf{XLie} 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 𝐗\mathbf{X} and 𝐘\mathbf{Y} are equivalent if there is a homotopy that is constrained to lie in the union of the images of 𝐗\mathbf{X} and 𝐘\mathbf{Y}. 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 𝐗\mathbf{X} factors through such a complex is not enough. We must also require that the factorization is trivial in homology.

Theorem 6.26.

Let 𝐗CPL1([0,1]2,V)\mathbf{X}\in C^{1}_{\operatorname{PL}}([0,1]^{2},V) be a piecewise linear surface. It is thinly null-homotopic if any of the following equivalent definitions hold:

  1. (W2)

    Word Condition. There is a marked representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) of 𝐗\mathbf{X} which is trivial in PL1(V)\operatorname{PL}_{1}(V).

  2. (H2G)

    Holonomy Condition. For a holonomy nondegenerate 𝐆=(δ:G1G0,)\mathbf{G}=(\delta:G_{1}\to G_{0},\vartriangleright), the surface holonomy along 𝐗\mathbf{X} of every smooth fake-flat 𝖌{\boldsymbol{\mathfrak{g}}}-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) is trivial.

  3. (R2)

    Rank Condition. There exists a smooth thin homotopy H:[0,1]3VH:[0,1]^{3}\to V, as defined in Definition 3.2 from 𝐗\mathbf{X} to the constant path at 0.

  4. (I2)

    Image Condition. There exists a smooth homotopy H:[0,1]3VH:[0,1]^{3}\to V between 𝐗\mathbf{X} and the constant surface at the origin such that it satisfies (I1) on the boundary of [0,1]2[0,1]^{2} and such that

    im(H)i=0kim(H([0,1]2×{ti})),\displaystyle\operatorname{im}(H)\subset\bigcup_{i=0}^{k}\operatorname{im}(H([0,1]^{2}\times\{t_{i}\})), (6.31)

    for a collection of times 0=t0<t1<tk10=t_{0}<t_{1}\ldots<t_{k}\leq 1.

  5. (F2)

    Factorization Condition. The surface 𝐗\mathbf{X} has trivial boundary and, after modifying the boundary using a thin homotopy, there exists a 2-dimensional CW complex CC and a map Θ:CV\Theta:C\to V, which is smooth when restricted to each cell, such that 𝐗\mathbf{X} factors as

    𝐗:[0,1]2𝑞S2𝑓CΘV\displaystyle\mathbf{X}:[0,1]^{2}\xrightarrow{q}S^{2}\xrightarrow{f}C\xrightarrow{\Theta}V (6.32)

    and such that H2(f)([S2])=0H_{2}(f)([S^{2}])=0, where q:[0,1]2S2q:[0,1]^{2}\to S^{2} is a map which covers the sphere and sends the boundary to a basepoint S2*\in S^{2}.

  6. (A2)

    Analytic Condition. The surface 𝐗\mathbf{X} has trivial boundary, and for all compactly supported ωΩc2(V)\omega\in\Omega^{2}_{c}(V), we have 𝐗ω=0\int_{\mathbf{X}}\omega=0.

  7. (S2)

    Surface Signature Condition. The surface signature S1S_{1} of 𝐗\mathbf{X} is trivial, S1(𝐗)=1S_{1}(\mathbf{X})=1.

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 𝐗\mathbf{X} is closed, the equivalence holds in general.

Now we prove the equivalence of (H2G) for a fixed holonomy nondegenerate 𝐆=(δ:G1G0,)𝖷𝖫𝖦𝗋𝗉\mathbf{G}=(\delta:G_{1}\to G_{0},\vartriangleright)\in\mathsf{XLGrp} with associated 𝖌=(δ:𝔤1𝔤0,)𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright)\in\mathsf{XLie}. Because surface holonomy is invariant with respect to thin homotopies [2, 50, 44] as defined by (R2), it follows that (R2) \Rightarrow (H2G). Now, suppose (H2G) holds. Let 𝔥im(δ)\mathfrak{h}\subset\operatorname{im}(\delta) be the semi-simple subalgebra. Choose a linear section s:𝔥𝔤1s:\mathfrak{h}\to\mathfrak{g}_{1}. Given any 1-connection γ0Ω1(V,𝔥)\gamma_{0}\in\Omega^{1}(V,\mathfrak{h}), we can consider the 2-connection γ1=s(κγ0)Ω2(V,𝔤1)\gamma_{1}=s(\kappa^{\gamma_{0}})\in\Omega^{2}(V,\mathfrak{g}_{1}) which is fake-flat by definition. Because surface holonomy is a morphism of crossed modules by Theorem 3.14, we have

δF1γ0,γ1(𝐗)=F0γ0(𝐗).\displaystyle\delta F_{1}^{\gamma_{0},\gamma_{1}}(\mathbf{X})=F_{0}^{\gamma_{0}}(\partial\mathbf{X}). (6.33)

Hence, by the assumption of (H2G), this implies that F0γ0(𝐗)F_{0}^{\gamma_{0}}(\partial\mathbf{X}) is trivial for all γ0Ω1(V,𝔥)\gamma_{0}\in\Omega^{1}(V,\mathfrak{h}). Thus, (H1G) holds for G=HG=H, where HH is a Lie group with Lie algebra 𝔥\mathfrak{h}. This implies that the boundary of 𝐗\mathbf{X} is trivial, so 𝐗\mathbf{X} is closed. Next, let 𝔞=ker(δ:𝔤1𝔤0)\mathfrak{a}=\ker(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0}), which is nontrivial by assumption. Let 𝔞\mathbb{R}\subseteq\mathfrak{a} be a 11-dimensional subalgebra and consider an abelian 2-connection (0,Wab)(0,W^{{\operatorname{ab}}}), where WabΩ2(V,)Ω2(V,𝔞)W^{{\operatorname{ab}}}\in\Omega^{2}(V,\mathbb{R})\subseteq\Omega^{2}(V,\mathfrak{a}) is valued in the subalgebra. Let AA be the integration of 𝔞\mathfrak{a}. The subalgebra \mathbb{R} integrates to a subgroup which is either \mathbb{R} or S1S^{1}. Because the connection is abelian, the surface holonomy (Definition 3.13) is given by either

F10,Wab(𝐗)=𝐗WaborF10,Wab(𝐗)=exp(i𝐗Wab)S1.\displaystyle F_{1}^{0,W^{{\operatorname{ab}}}}(\mathbf{X})=\int_{\mathbf{X}}W^{{\operatorname{ab}}}\in\mathbb{R}\,\,\quad\text{or}\quad F_{1}^{0,W^{{\operatorname{ab}}}}(\mathbf{X})=\exp\left(i\int_{\mathbf{X}}W^{{\operatorname{ab}}}\right)\in S^{1}. (6.34)

Because WabW^{{\operatorname{ab}}} 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) \Rightarrow (W2) provided the desired factorization of 𝐗\mathbf{X} through a PLSC CC, where Θ\Theta is linear when restricted to each simplex. Furthermore in Proposition 6.17, we constructed a Yπ2(C)Y\in\pi_{2}(C) such that W1(Y)=𝐙W_{1}(Y)=\mathbf{Z}, and thus H2(f)([S2])=(Y)=0H_{2}(f)([S^{2}])=\mathcal{H}(Y)=0 by Proposition 6.18. Now, suppose (F2) holds. This implies that 𝐗\mathbf{X} is closed. Given a 22-form ωΩ2(V)\omega\in\Omega^{2}(V), integrating the pullback Θ(ω)\Theta^{*}(\omega) along the cells of CC defines a map ϕ:H2(C,)\phi:H_{2}(C,\mathbb{Z})\to\mathbb{R}. Hence, the fact that H2(f)([S2])=0H_{2}(f)([S^{2}])=0 implies that (A2) holds.

Next, we consider the image condition (I2), where it is immediate that (I2) \Rightarrow (R2). Now, suppose (S2) holds, and once again we consider the proof of Theorem 6.1. There, we show that 𝐗\mathbf{X} factors through a PLSC CC, and add 2-simplices L^={λ^1,,λ^m}\widehat{L}=\{\widehat{\lambda}_{1},\ldots,\widehat{\lambda}_{m}\} to build a simply-connected CW complex C^\widehat{C} in Lemma 6.19. The result of Theorem 6.1 implies that there is a smooth thin homotopy hh satisfying (R2) which is contained in the image of C^\widehat{C}. Suppose λi=[p0,p1,p2]\lambda_{i}=[p_{0},p_{1},p_{2}]. By choosing some 𝐰iPL0(V)\mathbf{w}_{i}\in\operatorname{PL}_{0}(V) which connects the basepoint p0p_{0} of λ^i\widehat{\lambda}_{i} to the origin, we can interpret each λ^iL^\widehat{\lambda}_{i}\in\widehat{L} as a kite λ^iKite(V)\widehat{\lambda}_{i}\in{\operatorname{Kite}}(V). Using Lemma 5.17 and translating the basepoint, we can realize each 22-simplex λ^i\widehat{\lambda}_{i} as a surface 𝐘iC1([0,1]2,V)\mathbf{Y}_{i}\in C^{1}([0,1]^{2},V). We can incorporate the tail paths by using the formula for the action given in (3.21) to obtain a surface 𝐗iC1([0,1]2,V)\mathbf{X}_{i}\in C^{1}([0,1]^{2},V) defined by

𝐗iσ𝐰ih𝐘h((𝐰i)1+σ𝐰i1).\displaystyle\mathbf{X}_{i}\coloneqq\sigma^{\mathbf{w}_{i}}\star_{h}\mathbf{Y}\star_{h}((\mathbf{w}_{i})_{1}+\sigma^{\mathbf{w}_{i}^{-1}}). (6.35)

Let 𝐱i𝐱i1\mathbf{x}_{i}\star\mathbf{x}_{i}^{-1} denote the top boundary of 𝐗ih𝐗i1\mathbf{X}_{i}\star_{h}\mathbf{X}_{i}^{-1}, and let Hi:[0,1]2VH_{i}:[0,1]^{2}\to V be a thin homotopy of paths from 𝐱i𝐱i1\mathbf{x}_{i}\star\mathbf{x}_{i}^{-1} (on the bottom face of HH) to 0 (on the top face of HH). Then, we define

𝐅i=(𝐗ih𝐗i1)vHiand𝐗^=𝐗h𝐅1hh𝐅m,\displaystyle\mathbf{F}_{i}=(\mathbf{X}_{i}\star_{h}\mathbf{X}_{i}^{-1})\star_{v}H_{i}\quad\text{and}\quad\widehat{\mathbf{X}}=\mathbf{X}\star_{h}\mathbf{F}_{1}\star_{h}\ldots\star_{h}\mathbf{F}_{m}, (6.36)

where concatenations are performed from left to right. Here, each 𝐅i\mathbf{F}_{i} is a surface which represents the fold 𝐗ih𝐗i1\mathbf{X}_{i}\star_{h}\mathbf{X}_{i}^{-1} such that the boundary of 𝐅i\mathbf{F}_{i} is trivial. Note that the image of 𝐗^\widehat{\mathbf{X}} is |C^||\widehat{C}| by definition, and since 𝐗\mathbf{X} and 𝐗^\widehat{\mathbf{X}} differ only by folds, there exists a homotopy between them which is contained in the image of 𝐗^\widehat{\mathbf{X}}, and whose boundary is contained in the image of 𝐗\partial\mathbf{X}. Then, the image of the sequence of homotopies

𝐗𝐗^𝐗0,\displaystyle\mathbf{X}\to\widehat{\mathbf{X}}\to\mathbf{X}\xrightarrow{h}0, (6.37)

is contained in the image of 𝐗^\widehat{\mathbf{X}}, 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 H3(G)H_{3}(G). On the one hand, this allows us to geometrically interpret H3(G)H_{3}(G) in terms of surfaces; on the other, it allows us to further classify thinly null homotopic behavior. Given a compatible PLSC CC in VV, we use Corollary 6.12 to obtain a morphism

𝐖=(W1,W0):𝝅(C)𝐏𝐋(V).\displaystyle\mathbf{W}=(W_{1},W_{0}):{\boldsymbol{\pi}}(C)\to{\boldsymbol{\operatorname{PL}}}(V). (7.1)

Recall from Proposition 6.20 that ker()=ker(W1)\ker(\mathcal{H})=\ker(W_{1}). Now, we aim to characterize ker()\ker(\mathcal{H}) 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 GG be a group and MM be a GG-module. Let ϵ:G\epsilon:\mathbb{Z}G\to\mathbb{Z} be the augmentation map defined by ϵ(g)=1\epsilon(g)=1 for all gGg\in G. The group of co-invariants of MM is defined by MGM/IMM_{G}\coloneqq M/IM, where I=ker(ϵ)I=\ker(\epsilon) is the augmentation ideal generated by g1g-1 for gGg\in G. In other words, MGM_{G} is the quotient of MM by elements of the form mgmm-g\cdot m for gGg\in G and mMm\in M.

Definition 7.2.

Let GG be a group and let (F,d)(F_{\bullet},d) be a projective resolution of \mathbb{Z} over G\mathbb{Z}G, where

FnF00\displaystyle\ldots\to F_{n}\to\ldots\to F_{0}\to\mathbb{Z}\to 0 (7.2)

is exact. Then, we define the group homology of GG by

Hn(G)Hn(FG).\displaystyle H_{n}(G)\coloneqq H_{n}(F_{G}). (7.3)

The following is an exact sequence, originally due to Hopf [33], which relates the homology and homotopy groups of a CW-complex CC with the homology of its fundamental group G=π1(C)G=\pi_{1}(C) through the Hurewicz map. We state a version from [7].

Theorem 7.3.

[7, Exercise 1, Section 2.5] Let CC be an nn-dimensional CW-complex such that πi(C)=0\pi_{i}(C)=0 for 1<i<n1<i<n for some n2n\geq 2. Let G=π1(C)G=\pi_{1}(C). Then, the sequence

0Hn+1(G)(πn(C))GGHn(C)Hn(G)0\displaystyle 0\to H_{n+1}(G)\to(\pi_{n}(C))_{G}\xrightarrow{\mathcal{H}_{G}}H_{n}(C)\to H_{n}(G)\to 0 (7.4)

is exact, where G:πn(C)GHn(C)\mathcal{H}_{G}:\pi_{n}(C)_{G}\to H_{n}(C) is the co-invariants functor applied to the Hurewicz map and Hn(C)H_{n}(C) is equipped with the trivial GG-action.

In our current setting of a 2-dimensional CW complex ZZ, the connectivity hypothesis is trivially satisfied, and thus we obtain the exact sequence

0H3(G)ϕ(π2(Z))GGH2(Z)H2(G)0.\displaystyle 0\to H_{3}(G)\xrightarrow{\phi}(\pi_{2}(Z))_{G}\xrightarrow{\mathcal{H}_{G}}H_{2}(Z)\to H_{2}(G)\to 0. (7.5)

The kernel of G:(π2(Z))GH2(Z)\mathcal{H}_{G}:(\pi_{2}(Z))_{G}\to H_{2}(Z) is exactly ϕ(H3(G))\phi(H_{3}(G)). Thus, we obtain the following characterization of ker(W1)\ker(W_{1}).

Proposition 7.4.

The kernel of W1:π2(C)PL1(V)W_{1}:\pi_{2}(C)\to\operatorname{PL}_{1}(V) is

ker(W1)=Iπ2(C)+Gim(ϕ),\displaystyle\ker(W_{1})=I\cdot\pi_{2}(C)+\mathbb{Z}G\operatorname{im}(\phi), (7.6)

where I=ker(ϵ:G)I=\ker(\epsilon:\mathbb{Z}G\to\mathbb{Z}) is the augmentation ideal.

Proof.

By Proposition 6.20, ker(W1)=ker()\ker(W_{1})=\ker(\mathcal{H}). Then, we can factor the Hurewicz map by

:π2(C)𝑞π2(C)GGH2(C).\displaystyle\mathcal{H}:\pi_{2}(C)\xrightarrow{q}\pi_{2}(C)_{G}\xrightarrow{\mathcal{H}_{G}}H_{2}(C). (7.7)

By (7.5), ker(G)=im(ϕ)\ker(\mathcal{H}_{G})=\operatorname{im}(\phi). Then, ker()=q1(im(ϕ))=Iπ2(C)+Gim(ϕ)\ker(\mathcal{H})=q^{-1}(\operatorname{im}(\phi))=I\cdot\pi_{2}(C)+\mathbb{Z}G\operatorname{im}(\phi). ∎

This result allows us to classify thinly null homotopic surfaces whose image lies in an embedding of CC. Suppose 𝐗:[0,1]2V\mathbf{X}:[0,1]^{2}\to V factors as

𝐗:[0,1]2𝑞S2𝑌C||V.\displaystyle\mathbf{X}:[0,1]^{2}\xrightarrow{q}S^{2}\xrightarrow{Y}C\xrightarrow{|\cdot|}V. (7.8)

If 𝐗\mathbf{X} is thinly null-homotopic, then W1(Y)=0W_{1}(Y)=0. Then, depending on how YY is killed based on (7.6), the thin homotopy exhibits different behaviors.

  1. (1)

    (Folds) Y=0π2(C)Y=0\in\pi_{2}(C): The map YY is null homotopic within its image in CC. This is the case if YY exhibits only folds.

  2. (2)

    (Nonlocal Path Conjugation) Y=0π2(C)GY=0\in\pi_{2}(C)_{G}: The map YY is null-homotopic within its image in CC up to conjugation by paths γπ1(C)\gamma\in\pi_{1}(C). This is an example of a nonlocal cancellation because the thin null-homotopy must move through the simply connected extension C^\widehat{C}.

  3. (3)

    (Nonlocal Surface Cancellation) G(Y)=0\mathcal{H}_{G}(Y)=0: The map YY is null homotopic within the simply connected extension C^\widehat{C}, and is classified by H3(G)H_{3}(G).

Example 7.5.

Consider the example of the group G=/2G=\mathbb{Z}/2 with the presentation x|x2\langle x\,|\,x^{2}\rangle, 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 n\mathbb{R}^{n} for sufficiently large nn. is C=2C=\mathbb{RP}^{2}. In this case, we have

H2(/2)=0,H3(/2)=/2,H2(2)=0,π2(2)=.\displaystyle H_{2}(\mathbb{Z}/2)=0,\quad H_{3}(\mathbb{Z}/2)=\mathbb{Z}/2,\quad H_{2}(\mathbb{RP}^{2})=0,\quad\pi_{2}(\mathbb{RP}^{2})=\mathbb{Z}. (7.9)

By (7.5), this implies that H3(/2)π2(2)/2=/2H_{3}(\mathbb{Z}/2)\cong\pi_{2}(\mathbb{RP}^{2})_{\mathbb{Z}/2}=\mathbb{Z}/2, and must represent a thinly null-homotopic surface which factors through 2\mathbb{RP}^{2}. Consider the surface 𝐗\mathbf{X} from Proposition 3.4. Note that the map Y:S22Y:S^{2}\to\mathbb{RP}^{2} from the factorization (3.9) represents a generator of π2(2)\pi_{2}(\mathbb{RP}^{2}), and is also nontrivial when we pass to the co-invariants π2(2)/2\pi_{2}(\mathbb{RP}^{2})_{\mathbb{Z}/2}. Thus, the nontrivial element in H3(/2)H_{3}(\mathbb{Z}/2) represents the thinly null homotopic surface 𝐗\mathbf{X}.

Appendix A Notation and Conventions

Symbol Description Page
Categories
𝖵𝖾𝖼𝗍\mathsf{Vect} category of finite-dimensional vector spaces 2
𝖫𝗂𝖾\mathsf{Lie} category of Lie algebras
𝖷𝖦𝗋𝗉,𝖷𝖫𝖦𝗋𝗉\mathsf{XGrp},\mathsf{XLGrp} category of crossed modules of groups (resp. Lie groups) 3.7
𝖷𝖫𝗂𝖾\mathsf{XLie} category of crossed modules of Lie algebras 3.11
𝖵𝖫\mathsf{VL} comma category (id𝖥𝗈𝗋)(\operatorname{id}\downarrow\mathsf{For}) associated to id:𝖵𝖾𝖼𝗍𝖵𝖾𝖼𝗍\operatorname{id}:\mathsf{Vect}\to\mathsf{Vect} and 𝖥𝗈𝗋:𝖫𝗂𝖾𝖵𝖾𝖼𝗍\mathsf{For}:\mathsf{Lie}\to\mathsf{Vect} 3.5
𝖲𝖦\mathsf{SG} comma category (id𝖥𝗈𝗋)(\operatorname{id}\downarrow\mathsf{For}) associated to id:𝖲𝖾𝗍𝖲𝖾𝗍\operatorname{id}:\mathsf{Set}\to\mathsf{Set} and 𝖥𝗈𝗋:𝖦𝗋𝗉𝖲𝖾𝗍\mathsf{For}:\mathsf{Grp}\to\mathsf{Set} 5.1
Crossed Modules
𝐏𝐋(V){\boldsymbol{\operatorname{PL}}}(V) piecewise linear crossed module 5.6
𝝉(V){\boldsymbol{\tau}}(V) thin crossed module (thin homotopy and translation equivalence classes) 3.9
𝝅(C,C1)\boldsymbol{\pi}(C,C_{1}) fundamental crossed module of CW-complex CC relative to the 1-skeleton C1C_{1} 3.10
𝖐(V),𝖐^(V){\boldsymbol{\mathfrak{k}}}(V),\hat{{\boldsymbol{\mathfrak{k}}}}(V) Kapranov’s free crossed module of Lie algebras and its completion 3.51, 3.6
𝐊^(V)\hat{\mathbf{K}}(V) formal integration of 𝖐^(V)\hat{{\boldsymbol{\mathfrak{k}}}}(V) as a crossed module of groups 3.6
Differential Forms and Currents
Ωk(V)\Omega^{k}(V) smooth differential kk-forms on VV
Ωck(V)\Omega^{k}_{c}(V) smooth compactly supported differential kk-forms on VV
Ω¯k(V),Γ¯k(V)\overline{\Omega}^{k}(V),\overline{\Gamma}_{k}(V) polynomial differential kk-forms and kk-currents on VV 4.2, 4.2
Ω^k(V),Γ^k(V)\hat{\Omega}^{k}(V),\hat{\Gamma}_{k}(V) completion of polynomial differential kk-forms and kk-currents on VV 4.2, 4.2
Holonomy and Signatures
γ0,γ1\gamma_{0},\gamma_{1} connection/2-connection 2.2, 3.12
ζ0,ζ1\zeta_{0},\zeta_{1} universal translation-invariant connection/2-connection 2.10, 3.55
κγ0\kappa^{\gamma_{0}} curvature of the connection γ0\gamma_{0} 2.6
κγ0,γ1,𝒦γ0,γ1\kappa^{\gamma_{0},\gamma_{1}},\mathcal{K}^{\gamma_{0},\gamma_{1}} curvature/2-curvature of the 2-connection (γ0,γ1)(\gamma_{0},\gamma_{1}) 3.12
F0,F1F_{0},F_{1} path/surface holonomy 2.3, 3.13
S0,S1S_{0},S_{1} smooth path/surface signature 2.4, 3.17
SPL,0,SPL,1S_{\operatorname{PL},0},S_{\operatorname{PL},1} piecewise linear path/surface signature 2.17, 5.19
R0,R1R_{0},R_{1} realization of PL paths/surfaces as thin homotopy equivalence classes 2.34, 5.18
Misc
PlanarLoop(V){\operatorname{PlanarLoop}}(V) planar piecewise linear loops in VV 5.2
Kite(V){\operatorname{Kite}}(V) set of kites Kite(V)PL0(V)×PlanarLoop(V){\operatorname{Kite}}(V)\coloneqq\operatorname{PL}_{0}(V)\times{\operatorname{PlanarLoop}}(V) in VV 5.2
Kite×(V){\operatorname{Kite}}^{\times}(V) set of marked kites Kite×(V)𝖥𝖬𝗈𝗇(V)×PlanarLoop(V){\operatorname{Kite}}^{\times}(V)\coloneqq\mathsf{FMon}(V)\times{\operatorname{PlanarLoop}}(V) 6.2.1
Loop(V){\operatorname{Loop}}(V) piecewise linear loops generated by triangular loops 5.4.1
Pair(V){\operatorname{Pair}}(V) pair groupoid of VV 5.4.1
ConePL,Cone{\operatorname{Cone}}_{\operatorname{PL}},{\operatorname{Cone}} piecewise linear and smooth cone map 5.4.1, 5.4.4
\mathcal{H} Hurewicz map
𝐖=(W1,W0)\mathbf{W}=(W_{1},W_{0}) morphism of crossed modules 𝐖:𝝅(C,C1)𝐏𝐋(V)\mathbf{W}:\boldsymbol{\pi}(C,C_{1})\to{\boldsymbol{\operatorname{PL}}}(V) 6.12
Δ(r(𝐗))\Delta(r(\mathbf{X})) PLSC associated to a representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) of 𝐗\mathbf{X} 6.13
(α,s)(\alpha,s) simplex mapping associated to compatible r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) 6.2.2
\star concatenation of paths / group operation in τ2(V)\tau_{2}(V) 2.1,4
h\star_{h}, v\star_{v} 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 PL0(V)\operatorname{PL}_{0}(V) have a unique minimal representative. This will be done by developing a rewriting theory on the free monoid 𝖥𝖬𝗈𝗇(V)\mathsf{FMon}(V) generated by VV. To simplify the notation, we will omit the symbol \star for monoid multiplication in this section. Now we consider rewriting on the free monoid 𝖥𝖬𝗈𝗇(V)\mathsf{FMon}(V) generated by VV. We define two rewriting steps corresponding to the relations for PL0(V)\operatorname{PL}_{0}(V). For arbitrary words 𝐚,𝐛𝖥𝖬𝗈𝗇(V)\mathbf{a},\mathbf{b}\in\mathsf{FMon}(V) and linearly dependent v,wVv,w\in V, we define

𝐚(v,w)𝐛(PL0.1)𝐚(v+w)𝐛and𝐚(0)𝐛(PL0.2)𝐚𝐛.\displaystyle\mathbf{a}(v,w)\mathbf{b}\xrightarrow{\ref{PL0.1}}\mathbf{a}(v+w)\mathbf{b}\quad\text{and}\quad\mathbf{a}(0)\mathbf{b}\xrightarrow{\ref{PL0.2}}\mathbf{a}\mathbf{b}. (B.1)

Given two words 𝐚,𝐛𝖥𝖬𝗈𝗇(V)\mathbf{a},\mathbf{b}\in\mathsf{FMon}(V), we write 𝐚𝐛\mathbf{a}\rightarrow\mathbf{b} for applying exactly one of the rewriting steps above and write 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} if we can get from 𝐚\mathbf{a} to 𝐛\mathbf{b} by a sequence of rewriting steps (possibly zero steps). Note that removing a 0 can always be realized by the (PL0.1) rewriting step, unless both 𝐚\mathbf{a} and 𝐛\mathbf{b} are empty words. Let l(𝐚)l(\mathbf{a})\in\mathbb{N} be the length of the word 𝐚\mathbf{a}. Note that if 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} then l(𝐛)l(𝐚)l(\mathbf{b})\leq l(\mathbf{a}). If this involves a non-zero number of rewrite steps, then l(𝐛)<l(𝐚)l(\mathbf{b})<l(\mathbf{a}), 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 𝐚=(v1,,vn)𝖥𝖬𝗈𝗇(V)\mathbf{a}=(v_{1},\ldots,v_{n})\in\mathsf{FMon}(V) and suppose that 𝐚𝐛\mathbf{a}\to\mathbf{b} and 𝐚𝐛\mathbf{a}\to\mathbf{b}^{\prime} are two rewriting steps. Then there is a word 𝐜\mathbf{c} such that 𝐛𝐜\mathbf{b}\xrightarrow{*}\mathbf{c} and 𝐛𝐜\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{c}.

Proof.

Since the rewriting step (PL0.2) only applies if the length of 𝐚\mathbf{a} is 11, we can assume that we are applying (PL0.1). Hence, the two rewriting steps have the form

(v1,,vi,vi+1,,vj,vj+1,,vn)(v1,,vi+vi+1,,vj,vj+1,,vn)\displaystyle(v_{1},...,v_{i},v_{i+1},...,v_{j},v_{j+1},...,v_{n})\to(v_{1},...,v_{i}+v_{i+1},...,v_{j},v_{j+1},...,v_{n}) (B.2)

and

(v1,,vi,vi+1,,vj,vj+1,,vn)(v1,,vi,vi+1,,vj+vj+1,,vn).\displaystyle(v_{1},...,v_{i},v_{i+1},...,v_{j},v_{j+1},...,v_{n})\to(v_{1},...,v_{i},v_{i+1},...,v_{j}+v_{j+1},...,v_{n}). (B.3)

We can assume that iji\leq j. If i=ji=j, then the two rewrite steps are identical. If i+1<ji+1<j then we take 𝐜=(v1,,vi+vi+1,,vj+vj+1,,vn)\mathbf{c}=(v_{1},...,v_{i}+v_{i+1},...,v_{j}+v_{j+1},...,v_{n}). If j=i+1j=i+1, then each pair vi,vi+1v_{i},v_{i+1} and vi+1,vi+2v_{i+1},v_{i+2} is contained in a common line. There are two cases here. First, we could have vi+1=0v_{i+1}=0. But then the two rewrites 𝐛\mathbf{b} and 𝐛\mathbf{b}^{\prime} are the same. Second, if vi+10v_{i+1}\neq 0, then vi,vi+1,vi+2v_{i},v_{i+1},v_{i+2} are all contained in the line span(vi+1)\mathrm{span}(v_{i+1}). Then we take c=(v1,,vi+vi+1+vi+2,,vn)c=(v_{1},...,v_{i}+v_{i+1}+v_{i+2},...,v_{n}). ∎

Lemma B.2.

Let 𝐚𝖥𝖬𝗈𝗇(V)\mathbf{a}\in\mathsf{FMon}(V) and suppose that 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} and 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b}^{\prime}. Then there is a word 𝐜\mathbf{c} such that 𝐛𝐜\mathbf{b}\xrightarrow{*}\mathbf{c} and 𝐛𝐜\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{c}.

Proof.

We prove this by induction on the length of the word 𝐚\mathbf{a}. We can assume that both 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} and 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b}^{\prime} 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 𝐚𝐚1𝐛\mathbf{a}\to\mathbf{a}_{1}\xrightarrow{*}\mathbf{b} and 𝐚𝐚1𝐛\mathbf{a}\to\mathbf{a}_{1}\xrightarrow{*}\mathbf{b}^{\prime}, then the result follows by induction since l(𝐚1)<l(𝐚)l(\mathbf{a}_{1})<l(\mathbf{a}), so that we can apply the induction hypothesis to 𝐚1\mathbf{a}_{1}. As a result, we may assume the rewrites to have the form

𝐚𝐚1𝐛, and 𝐚𝐚1𝐛.\displaystyle\mathbf{a}\to\mathbf{a}_{1}\xrightarrow{*}\mathbf{b},\qquad\text{ and }\qquad\mathbf{a}\to\mathbf{a}_{1}^{\prime}\xrightarrow{*}\mathbf{b}^{\prime}. (B.4)

and l(𝐚1)<l(𝐚)l(\mathbf{a}_{1})<l(\mathbf{a}) and l(𝐚1)<l(𝐚)l(\mathbf{a}_{1}^{\prime})<l(\mathbf{a}). By Lemma B.1, there is a word 𝐝\mathbf{d} such that 𝐚1𝐝\mathbf{a}_{1}\xrightarrow{*}\mathbf{d} and 𝐚1𝐝\mathbf{a}^{\prime}_{1}\xrightarrow{*}\mathbf{d}. Now l(𝐚1)<l(𝐚)l(\mathbf{a}_{1})<l(\mathbf{a}), and we have 𝐚1𝐛\mathbf{a}_{1}\xrightarrow{*}\mathbf{b} and 𝐚1𝐝\mathbf{a}_{1}\xrightarrow{*}\mathbf{d}. So by the induction hypothesis, there is a word 𝐞\mathbf{e} such that 𝐛𝐞\mathbf{b}\xrightarrow{*}\mathbf{e} and 𝐝𝐞\mathbf{d}\xrightarrow{*}\mathbf{e}. Similarly, l(𝐚1)<l(𝐚)l(\mathbf{a}_{1}^{\prime})<l(\mathbf{a}), 𝐚1𝐛\mathbf{a}^{\prime}_{1}\xrightarrow{*}\mathbf{b}^{\prime} and 𝐚1𝐝\mathbf{a}^{\prime}_{1}\xrightarrow{*}\mathbf{d}, so by the induction hypothesis, there is a word 𝐞\mathbf{e}^{\prime} such that 𝐛𝐞\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{e}^{\prime} and 𝐝𝐞\mathbf{d}\xrightarrow{*}\mathbf{e}^{\prime}. Finally, l(𝐝)l(𝐚1)<l(𝐚)l(\mathbf{d})\leq l(\mathbf{a}_{1})<l(\mathbf{a}) and 𝐝𝐞\mathbf{d}\xrightarrow{*}\mathbf{e} and 𝐝𝐞\mathbf{d}\xrightarrow{*}\mathbf{e}^{\prime}. So by the induction hypothesis, there is a word 𝐜\mathbf{c} such that 𝐞𝐜\mathbf{e}\xrightarrow{*}\mathbf{c} and 𝐞𝐜\mathbf{e}^{\prime}\xrightarrow{*}\mathbf{c}. Combining the rewriting steps we have

𝐛𝐞𝐜 and 𝐛𝐞𝐜.\displaystyle\mathbf{b}\xrightarrow{*}\mathbf{e}\xrightarrow{*}\mathbf{c}\qquad\text{ and }\qquad\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{e}^{\prime}\xrightarrow{*}\mathbf{c}. (B.5)

Corollary B.3.

If 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} and 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b}^{\prime} are rewritings such that 𝐛\mathbf{b} and 𝐛\mathbf{b}^{\prime} are minimal words, then 𝐛=𝐛\mathbf{b}=\mathbf{b}^{\prime}. In other words, any word reduces to a unique minimal word.

Proof.

By Lemma B.2, there is a word 𝐜\mathbf{c} such that 𝐛𝐜\mathbf{b}\xrightarrow{*}\mathbf{c} and 𝐛𝐜\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{c}. But 𝐛\mathbf{b} and 𝐛\mathbf{b}^{\prime} are minimal, implying that 𝐛=𝐜=𝐛\mathbf{b}=\mathbf{c}=\mathbf{b}^{\prime}. ∎

Proof of Proposition 2.12.

Let gPL0(V)g\in\operatorname{PL}_{0}(V). We claim that there is a unique minimal representative, namely a representative 𝐚=(v1,,vn)\mathbf{a}=(v_{1},...,v_{n}) such that each vi0v_{i}\neq 0 and each consecutive pair vi,vi+1v_{i},v_{i+1} are linearly independent. Note first that such a representative must exist: given an arbitrary representative of gg, 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 𝐚=(v1,,vn)\mathbf{a}=(v_{1},...,v_{n}) and 𝐛=(u1,,um)\mathbf{b}=(u_{1},...,u_{m}) are two minimal representatives. Since 𝐚\mathbf{a} and 𝐛\mathbf{b} are both words representing gg, there is a sequence of rewriting steps, and their inverses, going from 𝐚\mathbf{a} to 𝐛\mathbf{b}. By Lemma B.2, there is a word 𝐜\mathbf{c} such that 𝐚𝐜\mathbf{a}\xrightarrow{*}\mathbf{c} and 𝐛𝐜\mathbf{b}\xrightarrow{*}\mathbf{c}. But because 𝐚\mathbf{a} and 𝐛\mathbf{b} are minimal, it is not possible to apply any non-trivial rewriting step. Hence 𝐚=𝐜=𝐛\mathbf{a}=\mathbf{c}=\mathbf{b}. ∎

Remark B.4.

Note that the minimal representative of an element gPL0(V)g\in\operatorname{PL}_{0}(V) 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 η~V:Pair(V)Loop(V)\tilde{\eta}_{V}:{\operatorname{Pair}}(V)\to{\operatorname{Loop}}(V) be the natural map sending a generator (v,u)(v,u) 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 η~V:Pair(V)Loop(V)\tilde{\eta}_{V}:{\operatorname{Pair}}(V)\to{\operatorname{Loop}}(V) 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 Loop(V){\operatorname{Loop}}(V) has minimal representatives, analogous to Proposition 2.12.

Lemma B.5.

An element 𝐱Loop(V)\mathbf{x}\in{\operatorname{Loop}}(V) has a unique minimal representative

𝐱=((u1,v1),,(un,vn))min.\displaystyle\mathbf{x}=\Big{(}(u_{1},v_{1}),\ldots,(u_{n},v_{n})\Big{)}_{\min}. (B.6)

This is a word such that the following conditions are satisfied

  1. (1)

    for each ii, the vectors uiu_{i} and viv_{i} are linearly independent, and

  2. (2)

    for each i<ni<n, either viui+1v_{i}\neq u_{i+1} or viuiv_{i}-u_{i} and vi+1ui+1v_{i+1}-u_{i+1} are linearly independent.

Similar to (B.1), we define two rewriting steps in 𝖥𝖬𝗈𝗇(V2)\mathsf{FMon}(V^{2}). Let 𝐚,𝐛𝖥𝖬𝗈𝗇(V2)\mathbf{a},\mathbf{b}\in\mathsf{FMon}(V^{2}), let (v1,v2),(v2,v3)V2(v_{1},v_{2}),(v_{2},v_{3})\in V^{2} be such that v1,v2,v3v_{1},v_{2},v_{3} lie on a common affine line, and let (u1,u2)V2(u_{1},u_{2})\in V^{2} be such that u1,u2u_{1},u_{2} are linearly dependent. The two rewriting steps are

𝐚(v1,v2)(v2,v3)𝐛(L.1)𝐚(v1,v3)𝐛and𝐚(u1,u2)𝐛(L.2)𝐚𝐛.\displaystyle\mathbf{a}\star(v_{1},v_{2})\star(v_{2},v_{3})\star\mathbf{b}\xrightarrow{\ref{L1}}\mathbf{a}\star(v_{1},v_{3})\star\mathbf{b}\quad\text{and}\quad\mathbf{a}\star(u_{1},u_{2})\star\mathbf{b}\xrightarrow{\ref{L2}}\mathbf{a}\star\mathbf{b}. (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 𝐚𝐛\mathbf{a}\rightarrow\mathbf{b} and 𝐚𝐛\mathbf{a}\xrightarrow{*}\mathbf{b} 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 𝐚𝖥𝖬𝗈𝗇(V2)\mathbf{a}\in\mathsf{FMon}(V^{2}) and suppose 𝐚𝐛\mathbf{a}\to\mathbf{b} and 𝐚𝐛\mathbf{a}\to\mathbf{b}^{\prime} are two rewriting steps. Then, there is a word 𝐜\mathbf{c} such that 𝐛𝐜\mathbf{b}\xrightarrow{*}\mathbf{c} and 𝐛𝐜\mathbf{b}^{\prime}\xrightarrow{*}\mathbf{c}.

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 ηV\eta_{V} from  (5.63), this map verifies the assumptions of  Proposition 5.24. Using the universal property of the group Loop(V){\operatorname{Loop}}(V), there is a unique map F:Loop(V)PL0cl(V)F:{\operatorname{Loop}}(V)\to\operatorname{PL}_{0}^{\operatorname{cl}}(V) such that Fη~V=ηVF\circ\tilde{\eta}_{V}=\eta_{V}. It sends the element (v,u)(v,u) to the triangular loop ηV(v,u)=(v,uv,u)\eta_{V}(v,u)=(v,u-v,-u). This map is surjective since we can factor any element of PL0cl(V)\operatorname{PL}_{0}^{\operatorname{cl}}(V) as a product of triangular loops. It remains to show injectivity. Suppose 𝐱=((u1,v1),,(un,vn))min\mathbf{x}=\big{(}(u_{1},v_{1}),\ldots,(u_{n},v_{n})\big{)}_{\min} is the minimal representative of an element of Loop(V){\operatorname{Loop}}(V). Under FF, this gets sent to the equivalence class for

F(𝐱)=(u1,v1u1,v1,,un,vnun,vn).\displaystyle F(\mathbf{x})=\big{(}u_{1},v_{1}-u_{1},-v_{1},\ldots,u_{n},v_{n}-u_{n},v_{n}\big{)}. (B.8)

Let l()l(\cdot) denote the length of the minimal representative for both Loop(V){\operatorname{Loop}}(V) and PL0(V)\operatorname{PL}_{0}(V). While there may be cancellations for F(𝐱)F(\mathbf{x}) in PL0cl\operatorname{PL}_{0}^{\operatorname{cl}}, we must have l(𝐱)l(F(𝐱))l(\mathbf{x})\leq l(F(\mathbf{x})) because 𝐱\mathbf{x} is minimal in Loop(V){\operatorname{Loop}}(V). Indeed, the only reductions in F(𝐱)F(\mathbf{x}) can occur between vi-v_{i} and ui+1u_{i+1}. Suppose these vectors are linearly dependent. If viui+1v_{i}\neq u_{i+1}, then the word reduces to (,ui+1vi,)(...,u_{i+1}-v_{i},...) and there are no further reductions at this locus. If, on the other hand, vi=ui+1v_{i}=u_{i+1}, then the word reduces to (,viui,vi+1ui+1,)(...,v_{i}-u_{i},v_{i+1}-u_{i+1},...) and there are no further reductions at this locus. Hence, the minimal representative for F(𝐱)F(\mathbf{x}) will contain all viuiv_{i}-u_{i}.

If 𝐱ker(F)\mathbf{x}\in\ker(F), we must have l(𝐱)l(F(𝐱))=0l(\mathbf{x})\leq l(F(\mathbf{x}))=0. This implies that l(𝐱)=0l(\mathbf{x})=0 and thus 𝐱=\mathbf{x}=\emptyset. Hence, FF 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 𝐛PlanarLoop(V)\mathbf{b}\in{\operatorname{PlanarLoop}}(V) be a non-trivial planar loop with span UVU\subset V, let 𝐱PL0(V)\mathbf{x}\in\operatorname{PL}_{0}(V) be a path such that 𝐱𝐛𝐱1PlanarLoop(V)\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}\in{\operatorname{PlanarLoop}}(V) is planar, then 𝐱PL0(U)\mathbf{x}\in\operatorname{PL}_{0}(U).

Proof.

Let 𝐛=(u1,,up)\mathbf{b}=(u_{1},...,u_{p}) and 𝐱=(w1,,ws)\mathbf{x}=(w_{1},...,w_{s}) be minimal representatives. We will prove this by induction on the length ss. Consider

𝐱𝐛𝐱1=(w1,,ws,u1,,up,ws,,w1).\displaystyle\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1}=(w_{1},...,w_{s},u_{1},...,u_{p},-w_{s},...,-w_{1}). (B.9)

The only reductions can occur between wsw_{s} and u1u_{1}, or between upu_{p} and ws-w_{s}. Hence, if wsUw_{s}\not\in U, then the word is minimal. But then the span of 𝐱𝐛𝐱1\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1} contains UU and wsw_{s}, contradicting the assumption that 𝐱𝐛𝐱1\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1} is planar. Hence, wsUw_{s}\in U. Let 𝐲=(w1,,ws1)\mathbf{y}=(w_{1},...,w_{s-1}), which is minimal, and let 𝐜=ws𝐛(ws)\mathbf{c}=w_{s}\star\mathbf{b}\star(-w_{s}), which is a non-trivial planar loop with span UU. Then 𝐲𝐜𝐲1=𝐱𝐛𝐱1\mathbf{y}\star\mathbf{c}\star\mathbf{y}^{-1}=\mathbf{x}\star\mathbf{b}\star\mathbf{x}^{-1} is planar. By induction 𝐲PL0(U)\mathbf{y}\in\operatorname{PL}_{0}(U), and therefore, 𝐱=𝐲wsPL0(U)\mathbf{x}=\mathbf{y}\star w_{s}\in\operatorname{PL}_{0}(U). ∎

Lemma B.8.

Let 𝐛1,𝐛2PlanarLoop(V)\mathbf{b}_{1},\mathbf{b}_{2}\in{\operatorname{PlanarLoop}}(V) be non-trivial planar loops with spans U1,U2VU_{1},U_{2}\subseteq V. If 𝐛1𝐛2PlanarLoop(V)\mathbf{b}_{1}\star\mathbf{b}_{2}\in{\operatorname{PlanarLoop}}(V) is planar, then U1=U2U_{1}=U_{2}.

Proof.

Let 𝐛1=(u1,,us)\mathbf{b}_{1}=(u_{1},...,u_{s}) and 𝐛2=(v1,,vr)\mathbf{b}_{2}=(v_{1},...,v_{r}) be minimal representatives. Then

𝐛1𝐛2=(u1,,us,v1,,vr).\displaystyle\mathbf{b}_{1}\star\mathbf{b}_{2}=(u_{1},...,u_{s},v_{1},...,v_{r}). (B.10)

If v1U1v_{1}\not\in U_{1}, then this word is minimal. But then the span contains U1U_{1} and v1v_{1}, contradicting the assumption that the loop is planar. Hence v1U1U2v_{1}\in U_{1}\cap U_{2}. For the same reason, usU1U2u_{s}\in U_{1}\cap U_{2}. If v1v_{1} and usu_{s} are linearly independent, then U1=U2U_{1}=U_{2} and we’re done. Hence, we assume that they are colinear, allowing us to reduce further:

𝐛1𝐛2=(u1,,us1,us+v1,v2,,vr)\displaystyle\mathbf{b}_{1}\star\mathbf{b}_{2}=(u_{1},...,u_{s-1},u_{s}+v_{1},v_{2},...,v_{r}) (B.11)

Now if v2U1v_{2}\not\in U_{1}, then the word is minimal (possibly after removing us+v1u_{s}+v_{1} if this vector is 0). Because both 𝐛1\mathbf{b}_{1} and 𝐛2\mathbf{b}_{2} are non-trivial loops, we have s,r3s,r\geq 3, and both (u1,u2)(u_{1},u_{2}) and (v1,v2)(v_{1},v_{2}) are linearly independent pairs of vectors. Hence, the span of 𝐛1𝐛2\mathbf{b}_{1}\star\mathbf{b}_{2} contains U1U_{1} and v2v_{2}, again contradicting planarity of 𝐛1𝐛2\mathbf{b}_{1}\star\mathbf{b}_{2}. Thus, v1,v2U1v_{1},v_{2}\in U_{1}, implying that U1=U2U_{1}=U_{2}. ∎

Lemma B.9.

Let 𝐚,𝐛PlanarLoop(V)\mathbf{a},\mathbf{b}\in{\operatorname{PlanarLoop}}(V) be non-trivial planar loops with spans Ua,UbVU_{a},U_{b}\subseteq V. If 𝐚𝐛\mathbf{a}\star\mathbf{b} is a kite 𝐯𝐜𝐯1\mathbf{v}\star\mathbf{c}\star\mathbf{v}^{-1}, where 𝐯PL0(V)\mathbf{v}\in\operatorname{PL}_{0}(V) and 𝐜PlanarLoop(V)\mathbf{c}\in{\operatorname{PlanarLoop}}(V), then Ua=UbU_{a}=U_{b}.

Proof.

Suppose 𝐚𝐛=𝐯𝐜𝐯1\mathbf{a}\star\mathbf{b}=\mathbf{v}\star\mathbf{c}\star\mathbf{v}^{-1}, where 𝐜PlanarLoop(V)\mathbf{c}\in{\operatorname{PlanarLoop}}(V) with span(𝐜)=Uc\operatorname{span}(\mathbf{c})=U_{c}. If 𝐜\mathbf{c} is trivial, then 𝐛=𝐚1\mathbf{b}=\mathbf{a}^{-1} and we are done. Hence, we assume 𝐜\mathbf{c} to be non-trivial. Let

𝐚=(a1,,as)minand𝐛=(b1,,br)min,\displaystyle\mathbf{a}=(a_{1},\ldots,a_{s})_{\min}\quad\text{and}\quad\mathbf{b}=(b_{1},\ldots,b_{r})_{\min}, (B.12)

where s,r3s,r\geq 3 since they are both nontrivial loops. Consider a factorization of the minimal representative of the concatenation

𝐚𝐛=(a1,,as,bl,,br)min=𝐚𝐛,\displaystyle\mathbf{a}\star\mathbf{b}=(a_{1},\ldots,a_{s^{\prime}},b^{\prime}_{l},\ldots,b_{r})_{\min}=\mathbf{a}^{\prime}\star\mathbf{b}^{\prime}, (B.13)

where 𝐚PL0(Ua)\mathbf{a}^{\prime}\in\operatorname{PL}_{0}(U_{a}) and 𝐛PL0(Ub)\mathbf{b}^{\prime}\in\operatorname{PL}_{0}(U_{b}). To obtain the minimal representative, the only possible simplification is by applying (PL0.1) to (as,b1)(a_{s},b_{1}). If (as,b1)(a_{s},b_{1})\sim\emptyset, then we may apply (PL0.1) again to (as1,b2)(a_{s-1},b_{2}). If both steps are possible, this implies that as1,asUba_{s-1},a_{s}\in U_{b}, and thus Ua=UbU_{a}=U_{b}, 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 s2s^{\prime}\geq 2 and l2l\leq 2.

Next, let 𝐯=(v1,,vn)min\mathbf{v}=(v_{1},\ldots,v_{n})_{\min} and 𝐜=(c1,,cm)min\mathbf{c}=(c_{1},\ldots,c_{m})_{\min} be minimal representatives. We claim that it is possible to assume that vnUcv_{n}\not\in U_{c}. Indeed, let vkv_{k} be the first element from the right which is not in UcU_{c}, so that 𝐮=(vk+1,,vn)PL0(Uc)\mathbf{u}=(v_{k+1},\ldots,v_{n})\in\operatorname{PL}_{0}(U_{c}). Define 𝐯=(v1,,vk)\mathbf{v}^{\prime}=(v_{1},...,v_{k}), which is minimal, and let 𝐜=𝐮𝐜𝐮1PL0cl(Uc)\mathbf{c}^{\prime}=\mathbf{u}\star\mathbf{c}\star\mathbf{u}^{-1}\in\operatorname{PL}_{0}^{\operatorname{cl}}(U_{c}), which is a non-trivial planar loop with span UcU_{c}. Then 𝐯=𝐯𝐮\mathbf{v}=\mathbf{v}^{\prime}\star\mathbf{u} and hence

𝐯𝐜𝐯1=𝐯(𝐮𝐜𝐮1)(𝐯)1=𝐯𝐜(𝐯)1.\displaystyle\mathbf{v}\star\mathbf{c}\star\mathbf{v}^{-1}=\mathbf{v}^{\prime}\star(\mathbf{u}\star\mathbf{c}\star\mathbf{u}^{-1})\star(\mathbf{v}^{\prime})^{-1}=\mathbf{v}^{\prime}\star\mathbf{c}^{\prime}\star(\mathbf{v}^{\prime})^{-1}. (B.14)

Thus, since 𝐯𝐜(𝐯)1\mathbf{v}^{\prime}\star\mathbf{c}^{\prime}\star(\mathbf{v}^{\prime})^{-1} is also a kite, we may simply replace 𝐯\mathbf{v} with 𝐯\mathbf{v}^{\prime} and 𝐜\mathbf{c} with 𝐜\mathbf{c}^{\prime}, establishing the claim. The minimal representative for the kite is therefore

(v1,,vn,c1,,cm,vn,,v1)min.\displaystyle(v_{1},...,v_{n},c_{1},...,c_{m},-v_{n},...,-v_{1})_{\min}. (B.15)

So far, we have obtained two descriptions of the minimal word representative of the product 𝐚𝐛\mathbf{a}\star\mathbf{b}. Hence, we have equality of minimal words 𝐚𝐛=𝐯𝐜𝐯1\mathbf{a}^{\prime}\star\mathbf{b}^{\prime}=\mathbf{v}\star\mathbf{c}\star\mathbf{v}^{-1}. There are two cases to consider, depending on where the transition between 𝐚\mathbf{a}^{\prime} and 𝐛\mathbf{b}^{\prime} occurs relative to 𝐜\mathbf{c}.

  1. (1)

    First, the transition occurs in 𝐜\mathbf{c}. Then there are inclusions of words 𝐯𝐚\mathbf{v}\subset\mathbf{a}^{\prime} and 𝐯1𝐛\mathbf{v}^{-1}\subset\mathbf{b}^{\prime}. This implies that 𝐯PL0(UaUb)\mathbf{v}\in\operatorname{PL}_{0}(U_{a}\cap U_{b}). In this case, we have 𝐜=(𝐯1𝐚𝐯)(𝐯1𝐛𝐯)\mathbf{c}=(\mathbf{v}^{-1}\mathbf{a}\mathbf{v})\star(\mathbf{v}^{-1}\mathbf{b}\mathbf{v}), and by  Lemma B.8, this implies that Ua=UbU_{a}=U_{b}.

  2. (2)

    Otherwise, the transition occurs in 𝐯\mathbf{v}, implying the inclusion of words 𝐜𝐯1𝐛\mathbf{c}\star\mathbf{v}^{-1}\subset\mathbf{b}^{\prime}, or it occurs in 𝐯1\mathbf{v}^{-1}, implying the inclusion of words 𝐯𝐜𝐚\mathbf{v}\star\mathbf{c}\subset\mathbf{a}^{\prime}. In either case, we have 𝐯,𝐜PL0(Ui)\mathbf{v},\mathbf{c}\in\operatorname{PL}_{0}(U_{i}) for a common i=ai=a or bb. But then 𝐚𝐛=𝐯𝐜𝐯1\mathbf{a}\star\mathbf{b}=\mathbf{v}\star\mathbf{c}\star\mathbf{v}^{-1} is planar, and so by  Lemma B.8, Ua=UbU_{a}=U_{b}.

Corollary B.10.

Let 𝐛1,𝐛2PlanarLoop(V)\mathbf{b}_{1},\mathbf{b}_{2}\in{\operatorname{PlanarLoop}}(V) be non-trivial planar loops with spans U1,U2VU_{1},U_{2}\subseteq V. Let 𝐮PL0(V)\mathbf{u}\in\operatorname{PL}_{0}(V) be a path. If 𝐛1𝐮𝐛2𝐮1PlanarLoop(V)\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}\in{\operatorname{PlanarLoop}}(V) is a planar loop, then U1=U2=UU_{1}=U_{2}=U and 𝐮PL0(U)\mathbf{u}\in\operatorname{PL}_{0}(U).

Proof.

Let 𝐛3=𝐛1𝐮𝐛2𝐮1\mathbf{b}_{3}=\mathbf{b}_{1}\star\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}, which we assume to be a planar loop. If this loop is trivial, then 𝐮𝐛2𝐮1=𝐛11\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1}=\mathbf{b}_{1}^{-1}. Then, since 𝐮𝐛2𝐮1\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1} is planar, 𝐮PL0(U2)\mathbf{u}\in\operatorname{PL}_{0}(U_{2}) by Lemma B.7 and hence U1=U2U_{1}=U_{2}. Next, assuming that 𝐛3\mathbf{b}_{3} is non-trivial, let U3U_{3} be its span. Then 𝐛11,𝐛3\mathbf{b}_{1}^{-1},\mathbf{b}_{3} are non-trivial planar loops with the property that 𝐛11𝐛3=𝐮𝐛2𝐮1\mathbf{b}_{1}^{-1}\star\mathbf{b}_{3}=\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1} is a kite. Then U1=U3U_{1}=U_{3} by Lemma B.9 and 𝐮𝐛2𝐮1\mathbf{u}\star\mathbf{b}_{2}\star\mathbf{u}^{-1} is a planar loop. By  Lemma B.7, 𝐮PL0(U2)\mathbf{u}\in\operatorname{PL}_{0}(U_{2}). Hence U1=U2U_{1}=U_{2}. ∎

Appendix C Free Crossed Modules of Lie Algebras

In this section, we fix a field 𝕂\mathbb{K} of characteristic 0, either 𝕂=,\mathbb{K}=\mathbb{R},\mathbb{C}.

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 𝖵𝖾𝖼𝗍\mathsf{Vect} be the category of vector spaces over 𝕂\mathbb{K}, and 𝖫𝗂𝖾\mathsf{Lie} be the category of Lie algebras over 𝕂\mathbb{K}. There exists a natural forgetful functor

𝖥𝗈𝗋:𝖫𝗂𝖾𝖵𝖾𝖼𝗍,\displaystyle\mathsf{For}:\mathsf{Lie}\to\mathsf{Vect}, (C.1)

and here we describe the corresopnding left adjoint functor

𝖥𝖫:𝖵𝖾𝖼𝗍𝖫𝗂𝖾\displaystyle\mathsf{FL}:\mathsf{Vect}\to\mathsf{Lie} (C.2)

which sends a vector space VV to the free Lie algebra over VV, denoted 𝖥𝖫(V)\mathsf{FL}(V). Let T(V)T(V) be the tensor algebra, the free associative algebra over VV, equipped with the shuffle coproduct Δ:T(V)T(V)T(V)\Delta:T(V)\to T(V)\otimes T(V), which is the algebra map defined on VV by Δ(v)=v1+1v\Delta(v)=v\otimes 1+1\otimes v. We define 𝖥𝖫(V)\mathsf{FL}(V) to be the set of primitive elements of T(V)T(V),

𝖥𝖫(V)Prim(T(V))={wT(V):Δ(w)=w1+1w}.\displaystyle\mathsf{FL}(V)\coloneqq{\operatorname{Prim}}(T(V))=\{w\in T(V)\,:\,\Delta(w)=w\otimes 1+1\otimes w\}. (C.3)

Let BB be a basis of VV, and let 𝖥𝖫(B)\mathsf{FL}(B) be the free Lie algebra generated by the set BB. By [51, Theorem IV.4.2], there is an isomorphism U(𝖥𝖫(B))T(V)U(\mathsf{FL}(B))\cong T(V), where U()U(\cdot) is the universal enveloping algebra. Then, by [51, Theorem III.5.4], Prim(U(𝖥𝖫(B)))𝖥𝖫(B){\operatorname{Prim}}(U(\mathsf{FL}(B)))\cong\mathsf{FL}(B), allowing us to identify 𝖥𝖫(V)𝖥𝖫(B)\mathsf{FL}(V)\cong\mathsf{FL}(B). 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 𝖷𝖫𝗂𝖾(𝔤)\mathsf{XLie}(\mathfrak{g}) of crossed modules

(δ:𝔥𝔤,),\displaystyle(\delta:\mathfrak{h}\to\mathfrak{g},\vartriangleright), (C.4)

where 𝔤\mathfrak{g} is fixed and where morphisms are the identity on 𝔤\mathfrak{g}. Taking the underlying vector space of 𝔤\mathfrak{g}, we define the slice category 𝖵𝖾𝖼𝗍/𝔤\mathsf{Vect}_{/\mathfrak{g}} consisting of linear map s:V𝔤s:V\to\mathfrak{g}. There is a forgetful functor

𝖥𝗈𝗋:𝖷𝖫𝗂𝖾(𝔤)𝖵𝖾𝖼𝗍/𝔤,\displaystyle\mathsf{For}:\mathsf{XLie}(\mathfrak{g})\to\mathsf{Vect}_{/\mathfrak{g}}, (C.5)

and in this section, we will construct a free functor

𝖥𝗋:𝖵𝖾𝖼𝗍/𝔤𝖷𝖫𝗂𝖾(𝔤)\displaystyle\mathsf{Fr}:\mathsf{Vect}_{/\mathfrak{g}}\to\mathsf{XLie}(\mathfrak{g}) (C.6)

as a left adjoint to 𝖥𝗈𝗋\mathsf{For}. We will break up this construction into two steps by factoring the forgetful functor as

𝖷𝖫𝗂𝖾(𝔤)𝖱𝖾𝗉(𝔤)/𝔤𝖵𝖾𝖼𝗍/𝔤,\displaystyle\mathsf{XLie}(\mathfrak{g})\to\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}\to\mathsf{Vect}_{/\mathfrak{g}}, (C.7)

where 𝖱𝖾𝗉(𝔤)\mathsf{Rep}(\mathfrak{g}) is the category of 𝔤\mathfrak{g}-representations, and 𝖱𝖾𝗉(𝔤)/𝔤\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}} is the slice category, consisting of 𝔤\mathfrak{g}-equivariant maps s:V𝔤s:V\to\mathfrak{g}, where the target is the adjoint representation.

C.2.1. Step 1

Recall that 𝔤\mathfrak{g}-representations can be equivalently described as U(𝔤)U(\mathfrak{g})-modules, 𝖱𝖾𝗉(𝔤)𝖬𝗈𝖽(U(𝔤))\mathsf{Rep}(\mathfrak{g})\cong\mathsf{Mod}(U(\mathfrak{g})), for U(𝔤)U(\mathfrak{g}) the universal enveloping algebra of 𝔤\mathfrak{g}. Hence, the forgetful functor 𝖥𝗈𝗋:𝖱𝖾𝗉(𝔤)𝖵𝖾𝖼𝗍\mathsf{For}:\mathsf{Rep}(\mathfrak{g})\to\mathsf{Vect} can be viewed as restricting the U(𝔤)U(\mathfrak{g})-action along the map 𝕂U(𝔤)\mathbb{K}\to U(\mathfrak{g}). Its left adjoint is therefore the extension of scalars functor

𝖤:𝖵𝖾𝖼𝗍𝖱𝖾𝗉(𝔤),VU(𝔤)𝕂V.\displaystyle\mathsf{E}:\mathsf{Vect}\to\mathsf{Rep}(\mathfrak{g}),\quad V\mapsto U(\mathfrak{g})\otimes_{\mathbb{K}}V. (C.8)

We can slice this functor to obtain a left adjoint to 𝖥𝗈𝗋:𝖱𝖾𝗉(𝔤)/𝔤𝖵𝖾𝖼𝗍/𝔤\mathsf{For}:\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}\to\mathsf{Vect}_{/\mathfrak{g}}, which we also denote by E:𝖵𝖾𝖼𝗍/𝔤𝖱𝖾𝗉(𝔤)/𝔤E:\mathsf{Vect}_{/\mathfrak{g}}\to\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}. It sends an object s:V𝔤s:V\to\mathfrak{g} to the 𝔤\mathfrak{g}-equivariant morphism

U(𝔤)𝕂VidsU(𝔤)𝕂𝔤𝔤,\displaystyle U(\mathfrak{g})\otimes_{\mathbb{K}}V\xrightarrow{\operatorname{id}\otimes s}U(\mathfrak{g})\otimes_{\mathbb{K}}\mathfrak{g}\to\mathfrak{g}, (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 𝖥𝗈𝗋:𝖷𝖫𝗂𝖾(𝔤)𝖱𝖾𝗉(𝔤)/𝔤\mathsf{For}:\mathsf{XLie}(\mathfrak{g})\to\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}. Let (f:W𝔤)𝖱𝖾𝗉(𝔤)/𝔤(f:W\to\mathfrak{g})\in\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}} be an object, where WW is a 𝔤\mathfrak{g}-representation, and ff is an equivariant map. In order to define a crossed module, we must construct a Lie algebra structure based on WW, such that the Peiffer identity holds. In fact, we will use the Peiffer identity to define the Lie bracket. We define the Peiffer pairing ,:W×WW\langle\cdot,\cdot\rangle:W\times W\to W by

u,v=f(u)v+f(v)u.\displaystyle\langle u,v\rangle=f(u)\vartriangleright v+f(v)\vartriangleright u. (C.10)

This pairing is symmetric and bilinear, and satisfies xu,v=xu,v+u,xvx\vartriangleright\langle u,v\rangle=\langle x\vartriangleright u,v\rangle+\langle u,x\vartriangleright v\rangle. In other words, it defines a 𝔤\mathfrak{g}-equivariant map Pf:S2(W)W{\operatorname{Pf}}:S^{2}(W)\to W. We denote the image of this map by Pf(W){\operatorname{Pf}}(W), which is a subrepresentation of WW lying in the kernel of ff. We define

𝖰(W)W/Pf(W).\displaystyle\mathsf{Q}(W)\coloneqq W/{\operatorname{Pf}}(W). (C.11)

Next, we define a preliminary bracket

[,]:W×W𝖰(W)by[u,v]=f(u)v,\displaystyle[\cdot,\cdot]:W\times W\to\mathsf{Q}(W)\quad\text{by}\quad[u,v]=f(u)\vartriangleright v, (C.12)

which is bilinear and skew-symmetric. Furthermore, we have

[u,v,w]=f(u)v,w=f(u)v,w+v,f(u)wPf(W),\displaystyle[u,\langle v,w\rangle]=f(u)\vartriangleright\langle v,w\rangle=\langle f(u)\vartriangleright v,w\rangle+\langle v,f(u)\vartriangleright w\rangle\in{\operatorname{Pf}}(W), (C.13)

implying that the bracket descends to

[,]:Λ2𝖰(W)𝖰(W).\displaystyle[\cdot,\cdot]:\Lambda^{2}\mathsf{Q}(W)\to\mathsf{Q}(W). (C.14)

This leads to the desired crossed module of Lie algebras.

Lemma C.1.

The bracket [,][\cdot,\cdot] defined in (C.14) defines a Lie algebra structure on 𝖰(V)\mathsf{Q}(V), ff descends to an equivariant morphism of Lie algebras δ:𝖰(V)𝔤\delta:\mathsf{Q}(V)\to\mathfrak{g}, and the 𝔤\mathfrak{g}-action on 𝖰(V)\mathsf{Q}(V) is a derivation of [,][\cdot,\cdot]. In particular, (δ:𝖰(V)𝔤,)(\delta:\mathsf{Q}(V)\to\mathfrak{g},\vartriangleright) is a crossed module of Lie algebras.

Proof.

First, since Pf(W){\operatorname{Pf}}(W) is a subrepresentation of WW contained in the kernel of ff, it is clear that we have an induced equivariant map δ:𝖰(V)𝔤\delta:\mathsf{Q}(V)\to\mathfrak{g}. Then using the equivariance, we get

δ([u,v])=f(f(u)v)=f(u)f(v)=[f(u),f(v)]=[δ(u),δ(v)],\displaystyle\delta([u,v])=f(f(u)\vartriangleright v)=f(u)\vartriangleright f(v)=[f(u),f(v)]=[\delta(u),\delta(v)], (C.15)

showing that δ\delta is bracket preserving. To show that 𝔤\mathfrak{g} acts as a derivation of the bracket,

x[u,v]\displaystyle x\vartriangleright[u,v] =x(f(u)v)\displaystyle=x\vartriangleright(f(u)\vartriangleright v) (C.16)
=f(u)(xv)+[x,f(u)]v\displaystyle=f(u)\vartriangleright(x\vartriangleright v)+[x,f(u)]\vartriangleright v (C.17)
=f(xu)v+[u,xv]\displaystyle=f(x\vartriangleright u)\vartriangleright v+[u,x\vartriangleright v] (C.18)
=[xu,v]+[u,xv].\displaystyle=[x\vartriangleright u,v]+[u,x\vartriangleright v]. (C.19)

To show that [,][\cdot,\cdot] defines a Lie bracket, we must show that it satisfies the Jacobi identity. This is equivalent to showing that [u,][u,-] is a derivation of the bracket. But [u,]=f(u)[u,-]=f(u)\vartriangleright, which we have just shown is a derivation. Finally, to show that 𝖰(W)\mathsf{Q}(W) is a crossed module, we must check the Peiffer identity. But this holds by construction. ∎

The construction of 𝖰(W)\mathsf{Q}(W) described above is functorial, providing the functor

𝖰:𝖱𝖾𝗉(𝔤)/𝔤𝖷𝖫𝗂𝖾(𝔤).\displaystyle\mathsf{Q}:\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}\to\mathsf{XLie}(\mathfrak{g}). (C.20)
Lemma C.2.

The functor QQ is left adjoint to 𝖥𝗈𝗋:𝖷𝖫𝗂𝖾(𝔤)𝖱𝖾𝗉(𝔤)/𝔤\mathsf{For}:\mathsf{XLie}(\mathfrak{g})\to\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}.

Proof.

To exhibit the adjunction we describe the counit ϵ:𝖰𝖥𝗈𝗋id𝖷𝖫𝗂𝖾(𝔤)\epsilon:\mathsf{Q}\circ\mathsf{For}\to\operatorname{id}_{\mathsf{XLie}(\mathfrak{g})} and unit η:id𝖱𝖾𝗉(𝔤)/𝔤𝖥𝗈𝗋𝖰\eta:\operatorname{id}_{\mathsf{Rep}(\mathfrak{g})_{/\mathfrak{g}}}\to\mathsf{For}\circ\mathsf{Q} natural transformations. On the one hand, we have that Q𝖥𝗈𝗋=idXMod(𝔤)Q\circ\mathsf{For}=id_{\mathrm{XMod}(\mathfrak{g})}, because P(𝔥)=0P(\mathfrak{h})=0 when 𝔥𝔤\mathfrak{h}\to\mathfrak{g} arises from a crossed module. Therefore, we take ϵ\epsilon to be the identity. On the other hand, the unit natural transformation η\eta must have components

ηW:WW/Pf(W)\displaystyle\eta_{W}:W\to W/{\operatorname{Pf}}(W) (C.21)

which we take to be the natural projection maps. The counit-unit equations follow easily from the fact that Q(ηW)=idQ(W)Q(\eta_{W})=id_{Q(W)} and that η𝔥=id𝔥\eta_{\mathfrak{h}}=\operatorname{id}_{\mathfrak{h}}, when 𝔥\mathfrak{h} 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 𝖥𝗈𝗋:𝖷𝖫𝗂𝖾(𝔤)𝖵𝖾𝖼𝗍/𝔤\mathsf{For}:\mathsf{XLie}(\mathfrak{g})\to\mathsf{Vect}_{/\mathfrak{g}} has a left adjoint

𝖥𝗋=𝖰𝖤:𝖵𝖾𝖼𝗍/𝔤𝖷𝖫𝗂𝖾(𝔤),\displaystyle\mathsf{Fr}=\mathsf{Q}\circ\mathsf{E}:\mathsf{Vect}_{/\mathfrak{g}}\to\mathsf{XLie}(\mathfrak{g}), (C.22)

which sends a linear map s:V𝔤s:V\to\mathfrak{g} to the crossed module

𝖥𝗋(s)(δ:(U(𝔤)𝕂V)/Pf(U(𝔤)𝕂V)𝔤,).\displaystyle\mathsf{Fr}(s)\coloneqq\Big{(}\delta:(U(\mathfrak{g})\otimes_{\mathbb{K}}V)/{\operatorname{Pf}}(U(\mathfrak{g})\otimes_{\mathbb{K}}V)\to\mathfrak{g},\vartriangleright\Big{)}. (C.23)

This free crossed module satisfies the following universal property.

Corollary C.4.

The unit of the adjunction provides a distinguished map

ηs:V𝖰(U(𝔤)V)𝖵𝖾𝖼𝗍/𝔤\displaystyle\eta_{s}:V\to\mathsf{Q}(U(\mathfrak{g})\otimes V)\in\mathsf{Vect}_{/\mathfrak{g}} (C.24)

given by including the vector vVv\in V into the equivalence class of 1v1\otimes v. For any crossed module (δ:𝔥𝔤,)(\delta^{\prime}:\mathfrak{h}\to\mathfrak{g},\vartriangleright) with a map f:V𝔥f:V\to\mathfrak{h} in 𝖵𝖾𝖼𝗍/𝔤\mathsf{Vect}_{/\mathfrak{g}}, there exists a unique map F:𝖥𝗋(V)𝔥F:\mathsf{Fr}(V)\to\mathfrak{h} in 𝖷𝖫𝗂𝖾(𝔤)\mathsf{XLie}(\mathfrak{g}) such that

Fηs=f.\displaystyle F\circ\eta_{s}=f. (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 𝖤(V)\mathsf{E}(V)), and then quotienting out by the Peiffer identity (as we have done with 𝖰\mathsf{Q}). 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 𝖷𝖫𝗂𝖾\mathsf{XLie} 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

f=(f1,f0):𝖍=(δ:𝔥1𝔥0,𝔥)𝖌=(δ:𝔤1𝔤0,𝔤).\displaystyle f=(f_{1},f_{0}):{\boldsymbol{\mathfrak{h}}}=(\delta:\mathfrak{h}_{1}\to\mathfrak{h}_{0},\vartriangleright_{\mathfrak{h}})\to{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright_{\mathfrak{g}}). (C.26)

Our aim is to define the pullback crossed module f0𝖌f_{0}^{*}{\boldsymbol{\mathfrak{g}}} such that this map factors as

f:𝖍u(f)f0𝖌f~𝖌.\displaystyle f:{\boldsymbol{\mathfrak{h}}}\xrightarrow{u(f)}f_{0}^{*}{\boldsymbol{\mathfrak{g}}}\xrightarrow{\tilde{f}}{\boldsymbol{\mathfrak{g}}}. (C.27)

First, consider the composition

𝔥0f0𝔤0𝔤Der(𝔤1),\displaystyle\mathfrak{h}_{0}\xrightarrow{f_{0}}\mathfrak{g}_{0}\xrightarrow{\vartriangleright_{\mathfrak{g}}}{\operatorname{Der}}(\mathfrak{g}_{1}), (C.28)

which allows us to define the semi-direct product 𝔥0𝔤1\mathfrak{h}_{0}\ltimes\mathfrak{g}_{1}, equipped with a Lie algebra map

π:𝔥0𝔤1𝔤0,defined by(X,a)f0(X)+δ(a).\displaystyle\pi:\mathfrak{h}_{0}\ltimes\mathfrak{g}_{1}\to\mathfrak{g}_{0},\quad\text{defined by}\quad(X,a)\mapsto f_{0}(X)+\delta(a). (C.29)

Then, we define the pullback by

f0𝔤1ker(π)={(X,a):f0(X)=δ(a)}.\displaystyle f_{0}^{*}\mathfrak{g}_{1}\coloneqq\ker(\pi)=\left\{(X,a)\,:\,f_{0}(X)=-\delta(a)\right\}. (C.30)

By direct computation, one can check that the Lie bracket on f0𝔤1f_{0}^{*}\mathfrak{g}_{1} is given by

[(X,a),(Y,b)]=([X,Y],[a,b]),\displaystyle[(X,a),(Y,b)]=\left([X,Y],-[a,b]\right), (C.31)

which implies that there are natural Lie algebra maps

δ:f0𝔤1𝔥0,(X,a)Xandf~1:f0𝔤1𝔥1,(X,a)a\displaystyle\delta:f_{0}^{*}\mathfrak{g}_{1}\to\mathfrak{h}_{0},\quad(X,a)\mapsto X\quad\text{and}\quad\tilde{f}_{1}:f_{0}^{*}\mathfrak{g}_{1}\to\mathfrak{h}_{1},\quad(X,a)\mapsto-a (C.32)

which satisfy δf~1=f0δ\delta\circ\tilde{f}_{1}=f_{0}\circ\delta. Furthermore, one can check that the action

X(Y,b)([X,Y],f0(X)b)\displaystyle X\vartriangleright(Y,b)\coloneqq\left([X,Y],f_{0}(X)\vartriangleright b\right) (C.33)

satisfies the properties of a crossed module.

Lemma C.6.

The pullback defined by f0𝖌=(δ:f0𝔤1𝔥0,)f_{0}^{*}{\boldsymbol{\mathfrak{g}}}=(\delta:f_{0}^{*}\mathfrak{g}_{1}\to\mathfrak{h}_{0},\vartriangleright) is a crossed module of Lie algebras. Furthermore, f~=(f~1,f0):f0𝖌𝖌\tilde{f}=(\tilde{f}_{1},f_{0}):f_{0}^{*}{\boldsymbol{\mathfrak{g}}}\to{\boldsymbol{\mathfrak{g}}} is a morphism of crossed modules.

Then, the desired factorization is immediate.

Lemma C.7.

Let

f=(f1,f0):𝖍=(δ:𝔥1𝔥0,𝔥)𝖌=(δ:𝔤1𝔤0,𝔤)\displaystyle f=(f_{1},f_{0}):{\boldsymbol{\mathfrak{h}}}=(\delta:\mathfrak{h}_{1}\to\mathfrak{h}_{0},\vartriangleright_{\mathfrak{h}})\to{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright_{\mathfrak{g}}) (C.34)

be a morphism of crossed modules. There exists a unique map u(f)=(u1(f),id𝔥0):𝖍f0𝖌u(f)=(u_{1}(f),\operatorname{id}_{\mathfrak{h}_{0}}):{\boldsymbol{\mathfrak{h}}}\to f_{0}^{*}{\boldsymbol{\mathfrak{g}}} in 𝖷𝖫𝗂𝖾(𝔥0)\mathsf{XLie}(\mathfrak{h}_{0}) defined by

u1(f):𝔥1f0𝔤1,a(δ(a),f1(a)),\displaystyle u_{1}(f):\mathfrak{h}_{1}\to f_{0}^{*}\mathfrak{g}_{1},\quad a\mapsto(\delta(a),-f_{1}(a)), (C.35)

such that f=f~u(f)f=\tilde{f}\circ u(f), where f~\tilde{f} 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 𝖵𝖫\mathsf{VL} denote the comma category associated to the functors id:𝖵𝖾𝖼𝗍𝖵𝖾𝖼𝗍\operatorname{id}:\mathsf{Vect}\to\mathsf{Vect} and 𝖥𝗈𝗋:𝖫𝗂𝖾𝖵𝖾𝖼𝗍\mathsf{For}:\mathsf{Lie}\to\mathsf{Vect}. An object of 𝖵𝖫\mathsf{VL} is given by the data of a vector space VV, a Lie algebra 𝔤\mathfrak{g}, and a linear map s:V𝔤s:V\to\mathfrak{g}. A morphism f=(f1,f0):(s:V𝔤)(t:W𝔥)f=(f_{1},f_{0}):(s:V\to\mathfrak{g})\to(t:W\to\mathfrak{h}) consists of a linear map f1:VWf_{1}:V\to W and a Lie algebra morphism f0:𝔤𝔥f_{0}:\mathfrak{g}\to\mathfrak{h} such that tf1=f0st\circ f_{1}=f_{0}\circ s as linear maps. In this section, we will exclusively refer to the natural forgetful functor

𝖥𝗈𝗋:𝖷𝖫𝗂𝖾𝖵𝖫.\displaystyle\mathsf{For}:\mathsf{XLie}\to\mathsf{VL}. (C.36)

Furthermore, given (s:V𝔤)𝖵𝖫(s:V\to\mathfrak{g})\in\mathsf{VL}, the unit ηs\eta_{s} will refer to a morphism (ηs,id𝔤):(s:V𝔤)𝖥𝗈𝗋(𝖥𝗋(s))(\eta_{s},\operatorname{id}_{\mathfrak{g}}):(s:V\to\mathfrak{g})\to\mathsf{For}(\mathsf{Fr}(s)) in 𝖵𝖫\mathsf{VL}. We are now ready to state and prove the enhanced universal property.

Theorem C.8.

Let (s:V𝔥)𝖵𝖫(s:V\to\mathfrak{h})\in\mathsf{VL}, 𝖌=(δ:𝔤1𝔤0,)𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{g}}}=(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright)\in\mathsf{XLie}, and let

f=(f1,f0):(s:V𝔥)𝖥𝗈𝗋(δ:𝔤1𝔤0,).\displaystyle f=(f_{1},f_{0}):(s:V\to\mathfrak{h})\to\mathsf{For}(\delta:\mathfrak{g}_{1}\to\mathfrak{g}_{0},\vartriangleright). (C.37)

Then there is a unique morphism of crossed modules

F=(F1,F0):𝖥𝗋(s)𝖌,\displaystyle F=(F_{1},F_{0}):\mathsf{Fr}(s)\to{\boldsymbol{\mathfrak{g}}}, (C.38)

such that 𝖥𝗈𝗋(F)ηs=f\mathsf{For}(F)\circ\eta_{s}=f. In particular F1ηs=f1F_{1}\circ\eta_{s}=f_{1} and F0=f0F_{0}=f_{0}.

Proof.

Given the Lie algebra map f0:𝔥𝔤0f_{0}:\mathfrak{h}\to\mathfrak{g}_{0}, we consider the pullback crossed module f0𝖌f_{0}^{*}{\boldsymbol{\mathfrak{g}}}, and the corresponding morphism of crossed modules from Lemma C.6,

f~:f0𝖌𝖌.\displaystyle\tilde{f}:f_{0}^{*}{\boldsymbol{\mathfrak{g}}}\to{\boldsymbol{\mathfrak{g}}}. (C.39)

Mimicking the construction in Lemma C.7, let u1(f):Vf0𝔤1u_{1}(f):V\to f_{0}^{*}\mathfrak{g}_{1} be defined as u1(f)(v)=(s(v),f1(v)).u_{1}(f)(v)=(s(v),-f_{1}(v)). This defines a map

u(f)=(u1(f),id𝔥):(s:V𝔥)𝖥𝗈𝗋(f0𝖌)\displaystyle u(f)=(u_{1}(f),\operatorname{id}_{\mathfrak{h}}):(s:V\to\mathfrak{h})\to\mathsf{For}(f_{0}^{*}{\boldsymbol{\mathfrak{g}}}) (C.40)

in 𝖵𝖾𝖼𝗍/𝔥\mathsf{Vect}_{/\mathfrak{h}} with the property that 𝖥𝗈𝗋(f~)u(f)=f\mathsf{For}(\tilde{f})\circ u(f)=f. Now using the universal property of 𝖥𝗋(s)\mathsf{Fr}(s) in 𝖷𝖫𝗂𝖾(𝔥)\mathsf{XLie}(\mathfrak{h}) from Corollary C.4, there is a unique map

g:𝖥𝗋(s)f0𝖌\displaystyle g:\mathsf{Fr}(s)\to f_{0}^{*}{\boldsymbol{\mathfrak{g}}} (C.41)

in 𝖷𝖫𝗂𝖾(𝔥)\mathsf{XLie}(\mathfrak{h}) such that 𝖥𝗈𝗋(g)ηs=u(f)\mathsf{For}(g)\circ\eta_{s}=u(f), where ηs\eta_{s} is the unit from (C.24). Now, we define

F=𝖥𝗋(s)𝑔f0𝖌f~𝖌,\displaystyle F=\mathsf{Fr}(s)\xrightarrow{g}f_{0}^{*}{\boldsymbol{\mathfrak{g}}}\xrightarrow{\tilde{f}}{\boldsymbol{\mathfrak{g}}}, (C.42)

which is a morphism of crossed modules and it satisfies

𝖥𝗈𝗋(F)ηs=𝖥𝗈𝗋(f~)𝖥𝗈𝗋(g)ηs=𝖥𝗈𝗋(f~)u(f)=f.\displaystyle\mathsf{For}(F)\circ\eta_{s}=\mathsf{For}(\tilde{f})\circ\mathsf{For}(g)\circ\eta_{s}=\mathsf{For}(\tilde{f})\circ u(f)=f. (C.43)

Hence, FF is the desired morphism.

It therefore only remains to show that FF is unique. To this end, let GG be another such morphism. By the identity For(G)ηs=f\mathrm{For}(G)\circ\eta_{s}=f, we must have G0=f0G_{0}=f_{0}. Then, by Lemma C.7, GG factors as G=f~u(G)G=\tilde{f}\circ u(G), where

u(G)=(u1(G),id𝔥):𝖥𝗋(s)f0𝖌\displaystyle u(G)=(u_{1}(G),\operatorname{id}_{\mathfrak{h}}):\mathsf{Fr}(s)\to f_{0}^{*}{\boldsymbol{\mathfrak{g}}} (C.44)

is in the category 𝖷𝖫𝗂𝖾(𝔥)\mathsf{XLie}(\mathfrak{h}). Now consider the map l=𝖥𝗈𝗋(u(G))ηs:Vf0𝔤1l=\mathsf{For}(u(G))\circ\eta_{s}:V\to f_{0}^{*}\mathfrak{g}_{1} in 𝖵𝖾𝖼𝗍/𝔥\mathsf{Vect}_{/\mathfrak{h}}. It satisfies

For(f~)l=For(f~)For(u(G))ηs=For(f~u(G))ηs=For(G)ηs=f.\displaystyle\mathrm{For}(\tilde{f})\circ l=\mathrm{For}(\tilde{f})\circ\mathrm{For}(u(G))\circ\eta_{s}=\mathrm{For}(\tilde{f}\circ u(G))\circ\eta_{s}=\mathrm{For}(G)\circ\eta_{s}=f. (C.45)

This this implies that l=u(f)l=u(f), or that For(u(G))ηs=u(f)\mathrm{For}(u(G))\circ\eta_{s}=u(f). But now appealing to the universal property of 𝖥𝗋(s)\mathsf{Fr}(s), we see that we must have u(G)=gu(G)=g, the map defined above. Hence,

G=f~u(G)=f~g=F.\displaystyle G=\tilde{f}\circ u(G)=\tilde{f}\circ g=F. (C.46)

This establishes uniqueness. ∎

We can rephrase the above theorem as the existence of an adjunction.

Corollary C.9.

The forgetful functor 𝖥𝗈𝗋:𝖷𝖫𝗂𝖾𝖵𝖫\mathsf{For}:\mathsf{XLie}\to\mathsf{VL} has a left adjoint

𝖥𝗋:𝖵𝖫𝖷𝖫𝗂𝖾.\displaystyle\mathsf{Fr}:\mathsf{VL}\to\mathsf{XLie}. (C.47)

Restricting to the subcategory 𝖵𝖾𝖼𝗍/𝔤\mathsf{Vect}_{/\mathfrak{g}} this gives the functor 𝖰𝖤\mathsf{Q}\circ\mathsf{E} defined in Theorem C.3. Therefore, it is given by sending a linear map s:V𝔤s:V\to\mathfrak{g} to the crossed module 𝖥𝗋(s)\mathsf{Fr}(s) as defined in (C.23).

In (3.50), we consider a special class of such free crossed modules given by the functor 𝖐:𝖵𝖾𝖼𝗍𝖷𝖫𝗂𝖾{\boldsymbol{\mathfrak{k}}}:\mathsf{Vect}\to\mathsf{XLie}. In fact, this crossed module has additional structure, since there is a natural action of GL(V){\operatorname{GL}}(V) on VV. For a group GG, we define a GG-Lie algebra to be a Lie algebra 𝔥\mathfrak{h} which is also a GG-representation such that the action \gtrdot of GG on 𝔥\mathfrak{h} preserves the bracket,

g[x,y]=[gx,gy].\displaystyle g\gtrdot[x,y]=[g\gtrdot x,g\gtrdot y]. (C.48)
Corollary C.10.

For any V𝖵𝖾𝖼𝗍V\in\mathsf{Vect}, 𝖐(V){\boldsymbol{\mathfrak{k}}}(V) is a crossed module in the category of GL(V){\operatorname{GL}}(V)-Lie algebras.

Proof.

First, we note that 𝖥𝖫(V)\mathsf{FL}(V) has a natural GL(V){\operatorname{GL}}(V) action by extending the GL(V){\operatorname{GL}}(V) action on VV using (C.48). The vector space T(V)Λ2VT(V)\otimes\Lambda^{2}V also has a natural GL(V){\operatorname{GL}}(V) action via the usual tensor and wedge products of representations. The map 𝖤(sV):T(V)Λ2V𝖥𝖫(V)\mathsf{E}(s_{V}):T(V)\otimes\Lambda^{2}V\to\mathsf{FL}(V) is easily seen to be equivariant. Furthermore, the embedding 𝖥𝖫(V)U(𝖥𝖫(V))=T(V)\mathsf{FL}(V)\to U(\mathsf{FL}(V))=T(V) is equivariant, and thus the 𝖥𝖫(V)\mathsf{FL}(V) action on T(V)Λ2VT(V)\otimes\Lambda^{2}V is equivariant,

g(XA)=(gX)(gA),\displaystyle g\gtrdot(X\vartriangleright A)=(g\gtrdot X)\vartriangleright(g\gtrdot A), (C.49)

for gGL(V)g\in{\operatorname{GL}}(V), X𝖥𝖫(V)X\in\mathsf{FL}(V) and AT(V)Λ2(V)A\in T(V)\otimes\Lambda^{2}(V). It is also clear to check that the Peiffer subspace is invariant under the GL(V){\operatorname{GL}}(V) action, GL(V)Pf(T(V)Λ2V)Pf(T(V)Λ2V){\operatorname{GL}}(V)\gtrdot{\operatorname{Pf}}(T(V)\otimes\Lambda^{2}V)\subset{\operatorname{Pf}}(T(V)\otimes\Lambda^{2}V) so that the GL(V){\operatorname{GL}}(V) action descends to 𝖰(T(V)Λ2V)\mathsf{Q}(T(V)\otimes\Lambda^{2}V). The equivariance of δ\delta and the 𝖥𝖫(V)\mathsf{FL}(V) action \vartriangleright 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 r(𝐗)=(X1,,Xk)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})=(X_{1},\ldots,X_{k})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) and triangulations of their corresponding PLSCs Δ(r(𝐗))\Delta(r(\mathbf{X})). First, we show that triangulated representatives exist.

Lemma D.1.

There exists a triangulated representative r(𝐗)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)) of every 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V).

Proof.

It suffices to show that each marked kite X=(𝐰,𝐛)X=(\mathbf{w},\mathbf{b}) has a triangulated representative. Suppose 𝐛=(b1,,bk)minPlanarLoop(V)\mathbf{b}=(b_{1},\ldots,b_{k})_{\min}\in{\operatorname{PlanarLoop}}(V) is minimal with point representation 𝐛=[b^0,,b^m]\mathbf{b}=[\hat{b}_{0},\ldots,\hat{b}_{m}] where b^0=0\hat{b}_{0}=0 and b^k=i=1kbi\hat{b}_{k}=\sum_{i=1}^{k}b_{i}.

[Uncaptioned image]

Then, define

𝐛~r(b^r,br+1,b^r+1)PlanarLoop(V)andX~r(𝐰,𝐛~r).\displaystyle\tilde{\mathbf{b}}_{r}\coloneqq(\hat{b}_{r},b_{r+1},-\hat{b}_{r+1})\in{\operatorname{PlanarLoop}}(V)\quad\text{and}\quad\tilde{X}_{r}\coloneqq(\mathbf{w},\tilde{\mathbf{b}}_{r}). (D.1)

Then, we have 𝐗~=(X~1,,X~k2)(PL1.1)(X)\tilde{\mathbf{X}}=(\tilde{X}_{1},\ldots,\tilde{X}_{k-2})\sim_{\ref{PL1.1}}(X). The kites X~r\tilde{X}_{r} are either triangular, or else have 𝐛~r=0\tilde{\mathbf{b}}_{r}=\emptyset_{0}. In the later case X~r(PL1.2)1\tilde{X}_{r}\sim_{\ref{PL1.2}}\emptyset_{1} and hence can be removed from the list. Let 𝐗~\tilde{\mathbf{X}}^{\prime} be the resulting element of 𝖥𝖬𝗈𝗇(Kite×(V))\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)). It is a triangulated representative of the marked kite XX. ∎

Let r(𝐗)r(\mathbf{X}) be a triangulated representative of 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V). In general, there is no guarantee that this representative is compatible. Therefore, in order to prove Theorem 6.16, we must modify, or refine, r(𝐗)r(\mathbf{X}). We will do this by triangulating the PLSC Δ(r(𝐗))\Delta(r(\mathbf{X})).

Recall that a convex polygon in VV is the convex hull Conv(P)\mathrm{Conv}(P) of a finite set of points PP which are all contained in a 22-dimensional affine plane UVU\subset V. Such a polygon is non-degenerate if the points PP are not contained in any affine line LVL\subset V. By an edge in VV, we simply mean a line segment in VV connecting a pair of distinct points. Furthermore, given an edge or a polygon aa, we let aa^{\circ} denote its interior.

Definition D.2.

Let E={e1,,eq}E=\{e^{1},\ldots,e^{q}\} be a collection of edges in VV and let P={P1,,Pm}P=\{P^{1},\ldots,P^{m}\} be a collection of nondegenerate convex polygons. A compatible triangulation for (E,P)(E,P) is a compatible non-degenerate PLSC T(E,P)T(E,P) such that

  1. (1)

    for each 11-simplex ϵT(E,P)\epsilon\in T(E,P) and each edge eiEe^{i}\in E, if |ϵ|(ei)|\epsilon|^{\circ}\cap(e^{i})^{\circ}\neq\emptyset, then |ϵ|ei|\epsilon|\subseteq e^{i}.

  2. (2)

    for each 2-simplex τT(E,P)\tau\in T(E,P) and each polygon PiPP^{i}\in P, if |τ|(Pi)|\tau|^{\circ}\cap(P^{i})^{\circ}\neq\emptyset, then |τ|Pi|\tau|\subseteq P^{i}.

  3. (3)

    |T(E,P)|=(i=1qei)(i=1mPi)|T(E,P)|=(\bigcup_{i=1}^{q}e^{i})\ \cup\ (\bigcup_{i=1}^{m}P^{i}).

  4. (4)

    the edges in EE and in the boundaries of the polygons of PP are contained in |T(E,P)1||T(E,P)_{1}|, the realization of the 1-skeleton.

Lemma D.3.

Let T(E,P)T(E,P) be a compatible triangulation for a pair (E,P)(E,P). Given an edge in EE or a polygon in PP, denoted QQ, the subset

T(Q){σT(E,P):|σ|Q}\displaystyle T(Q)\coloneqq\{\sigma\in T(E,P)\,:\,|\sigma|\subseteq Q\} (D.2)

defines a compatible triangulation of QQ.

Proof.

First, we note that T(Q)T(Q) is indeed a PLSC, since for any σT(E,P)\sigma\in T(E,P) such that |σ|Q|\sigma|\subset Q, the faces of σ\sigma must also be contained in QQ. 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 T(E,P)T(E,P), and thus are inherited by T(Q)T(E,P)T(Q)\subset T(E,P). For condition (3), note first that |T(Q)|Q|T(Q)|\subseteq Q follows from the construction. If QQ is a polygon, then QQ is contained in the union of all 2-simplices of T(E,P)T(E,P). And by condition (2), the simplices which are not in T(Q)T(Q) have their interiors disjoint from QoQ^{o} and so cannot contribute any area. As a result Q|T(Q)|Q\subset|T(Q)|. A similar argument applies if QQ is an edge, using condition (1) instead. Condition (4) follows essentially for topological reasons. ∎

Lemma D.4.

Let EE and PP be, respectively, a set of edges and nondegenerate convex polygons in VV. Then there exists a compatible triangulation for (E,P)(E,P).

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 HH, affine lines LL, and vertices C0C_{0}.

  • Step 1

    The polygons in PP are supported on a finite set of affine planes, which we take to be HH.

  • Step 2

    The set of lines LL is constructed as follows. The boundary of each polygon PiPP^{i}\in P is a collection of line segments. Extend these to lines and add them to LL. Extend each edge in EE to a line and add it to LL. Finally, if two distinct planes Hi,HjHH_{i},H_{j}\in H intersect along a line, add it to LL.

  • Step 3

    To construct the set of vertices C0C_{0}, we start with the endpoints of edges in EE and the extremal points of the polygons in PP. If two lines in LL intersect at a point, we add this to C0C_{0}. If a line in LL intersects a plane in HH at a point, we add this to C0C_{0}. Finally, if two planes in HH intersect at a point, we add this to C0C_{0}.

Restrict attention to a single plane HiHH_{i}\in H. This plane will contain a subset LiLL_{i}\subseteq L of lines and a subset CiC0C_{i}\subseteq C_{0} of vertices. Suppose that there is a point pCip\in C_{i} which is not contained in any line from LiL_{i}. Then add a new line ll to LL which is contained in HiH_{i} and which contains the point pp. This will lead to new vertices in C0C_{0} arising from the intersections between ll and other lines from LL. However, there will not be further new vertices arising from intersecting ll with planes HjHiH_{j}\neq H_{i}. For this reason, adding ll will not introduce any new vertices which are contained within a plane HjHH_{j}\in H, but not within a line from LL lying in HjH_{j}. Therefore, by iterating this process, we may assume that every vertex from CiC_{i} is contained in some line from LiL_{i}.

Now observe that using the vertices in C0C_{0}, each line lLl\in L is decomposed into a finite number of closed segments ϵl\epsilon\subset l. Since each edge eiEe^{i}\in E is a subset of some line in LL lying between two vertices from C0C_{0}, it is decomposed into a union of such line segments ϵ\epsilon. Let ={ϵ1,,ϵp}\mathcal{E}=\{\epsilon^{1},...,\epsilon^{p}\} denote the collection of all finite line segments whose interiors intersect the interior of one of the edges from EE. Note that i=1pϵi=j=1qej\cup_{i=1}^{p}\epsilon^{i}=\cup_{j=1}^{q}e^{j}. Indeed, each ϵi\epsilon^{i} is contained in some eje^{j}, whereas each eje^{j} decomposes into a union of some ϵi\epsilon^{i}.

Next, consider the plane HiH_{i} and let LiLL_{i}\subset L denote the subset of lines that lie in HiH_{i}. Let Zi=HilLilZ_{i}=H_{i}\setminus\cup_{l\in L_{i}}l denote the complement of the set of lines LiL_{i}. It is an open set with the property that the closure of each bounded component is a convex polygon QQ in HiH_{i}. Since each PiPP^{i}\in P is the intersection of a collection of half-planes bounded by a subset of the lines lLil\in L_{i}, PiP^{i} is decomposed into a union of convex polygons QQ. Let 𝒫={Q1,,Qt}\mathcal{P}=\{Q^{1},...,Q^{t}\} denote the collection of all convex polygons in all planes HiH_{i} whose interiors intersect the interior of one of the polygons from PP. Note that the extremal points of QiQ^{i} lie in C0C_{0}, but that there may be additional points from C0C_{0} on the boundary. Note also that i=1tQi=j=1mPj\cup_{i=1}^{t}Q^{i}=\cup_{j=1}^{m}P^{j}. Indeed, each QiQ^{i} is contained in some PjP^{j}, whereas each PjP^{j} decomposes into a union of some QiQ^{i}.

We now construct a PLSC T(E,P)T(E,P) which will be our compatible triangulation.

  1. (1)

    Let T0VT_{0}\subset V be the collection of all vertices pC0p\in C_{0} which are either contained in an edge ϵi\epsilon^{i}\in\mathcal{E}, or in a convex polygon Qi𝒫Q^{i}\in\mathcal{P}. Endow T0T_{0} with an arbitrary order. This will form the set of vertices of our PLSC.

  2. (2)

    Given each convex polygon Qi𝒫Q^{i}\in\mathcal{P}, choose a triangulation that makes use of precisely the set of vertices in C0C_{0} lying on its boundary. Define Δ2(T)\Delta_{2}(T), the set of 22-simplices in T(E,P)T(E,P), to be the collection of all triples [p0,p1,p2][p_{0},p_{1},p_{2}] such |[p0,p1,p2]||[p_{0},p_{1},p_{2}]| is one of the triangles from a polygon QiQ^{i}.

  3. (3)

    We define Δ1(T)\Delta_{1}(T), the set of 11-simplices in T(E,P)T(E,P), to consist of all faces of all 22-simplices in Δ2(T)\Delta_{2}(T), as well as all pairs [p0,p1][p_{0},p_{1}] such that |[p0,p1]||[p_{0},p_{1}]| is one of the edges in \mathcal{E}.

By construction T(E,P)T(E,P) is a compatible non-degenerate PLSC whose piecewise linear realization satisfies

|T(E,P)|=(i=1pϵi)(i=1tQi)=(j=1qej)(i=1mPj).\displaystyle|T(E,P)|=(\bigcup_{i=1}^{p}\epsilon^{i})\ \cup\ (\bigcup_{i=1}^{t}Q^{i})=(\bigcup_{j=1}^{q}e^{j})\ \cup\ (\bigcup_{i=1}^{m}P^{j}). (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 r(𝐗)r(\mathbf{X}) and take its associated PLSC, Δ(r(𝐗))\Delta(r(\mathbf{X})), which may not be compatible. Then, we produce a compatible triangulation TT and use it to construct a new representative r(𝐗)r^{\prime}(\mathbf{X}). This new representative will have the property that its associated PLSC is a subcomplex of TT, thereby assuring that it is compatible.

The following lemma starts by treating the case of a triangular kite.

Lemma D.5.

Let X=(,𝐛)Kite×(V)X=(\emptyset,\mathbf{b})\in{\operatorname{Kite}}^{\times}(V) be a triangular loop, viewed as a marked triangular kite, and let |σX||\sigma_{X}| be its associated 2-simplex, viewed as a convex polygon in VV. Given a compatible triangulation T(|σX|)T(|\sigma_{X}|), there exists a triangulated representative r(X)=(X1,,Xk)r(X)=(X_{1},...,X_{k}) such that r(X)(X)r(X)\sim(X) in PL1(V)\operatorname{PL}_{1}(V), and such that Δ(r(X))T(|σX|)\Delta(r(X))\subseteq T(|\sigma_{X}|).

Proof.

The PLSC T(|σX|)T(|\sigma_{X}|) is homeomorphic to |T(|σX|)|=|σX||T(|\sigma_{X}|)|=|\sigma_{X}|, which is a 2-simplex. Its 1-skeleton Z~\tilde{Z} is a connected 11-dimensional simplicial complex which contains ZZ, the boundary of σX\sigma_{X}, as a subspace. Let c=0c=0 be the basepoint of σX\sigma_{X}. Because Z~\tilde{Z} is 1-dimensional, its fundamental group π1(Z~,c)\pi_{1}(\tilde{Z},c) is free. We will need a generating set which is indexed by the 2-simplices of T(|σX|)T(|\sigma_{X}|). For this, we choose a spanning tree TT of Z~\tilde{Z}. For each 2-simplex τ=[t0,t1,t2]\tau=[t_{0},t_{1},t_{2}], let ϕτ\phi_{\tau} be the boundary loop of τ\tau based at t0t_{0}, and let pτp_{\tau} be the unique path in TT connecting cc to t0t_{0}. Define

aτ=pτϕτpτ1π1(Z~,c),\displaystyle a_{\tau}=p_{\tau}\star\phi_{\tau}\star p_{\tau}^{-1}\in\pi_{1}(\tilde{Z},c), (D.4)

which form a generating set for π1(Z~,c)\pi_{1}(\tilde{Z},c). The inclusion ι:ZZ~\iota:Z\to\tilde{Z} induces a homomorphism

π1(ι):π1(Z,c)π1(Z~,c).\displaystyle\pi_{1}(\iota):\pi_{1}(Z,c)\to\pi_{1}(\tilde{Z},c). (D.5)

Therefore, given the generator γπ1(Z,c)\gamma\in\pi_{1}(Z,c), its image under π1(ι)\pi_{1}(\iota) admits a factorization

π1(ι)(γ)=aτi1±aτik±.\displaystyle\pi_{1}(\iota)(\gamma)=a_{\tau_{i_{1}}}^{\pm}\star...\star a_{\tau_{i_{k}}}^{\pm}. (D.6)

Now let W0:π1(Z~,c)PL0(V)W_{0}:\pi_{1}(\tilde{Z},c)\to\operatorname{PL}_{0}(V) be the homomorphism from  (6.8). Then W0(π1(ι))=𝐛W_{0}(\pi_{1}(\iota))=\mathbf{b} and

W0(aτ)=𝐰τ𝐛τ𝐰τ1,\displaystyle W_{0}(a_{\tau})=\mathbf{w}_{\tau}\star\mathbf{b}_{\tau}\star\mathbf{w}_{\tau}^{-1}, (D.7)

where 𝐰τ=W~0(pτ)PL0(V)\mathbf{w}_{\tau}=\widetilde{W}_{0}(p_{\tau})\in\operatorname{PL}_{0}(V) and where 𝐛τ=W~0(ϕτ)PL0cl(V)\mathbf{b}_{\tau}=\widetilde{W}_{0}(\phi_{\tau})\in\operatorname{PL}_{0}^{\operatorname{cl}}(V) is a triangular loop. Then, since all paths are contained in the plane spanned by |σX||\sigma_{X}|, we obtain a factorization in PL1(V)\operatorname{PL}_{1}(V),

(X)=((𝐰τi1,𝐛τi1±),,(𝐰τik,𝐛τik±)).\displaystyle(X)=((\mathbf{w}_{\tau_{i_{1}}},\mathbf{b}_{\tau_{i_{1}}}^{\pm}),...,(\mathbf{w}_{\tau_{i_{k}}},\mathbf{b}_{\tau_{i_{k}}}^{\pm})). (D.8)

Since pτp_{\tau} is a path in TT connecting vertices of T(|σX|)T(|\sigma_{X}|), we can choose a lift of 𝐰τ\mathbf{w}_{\tau} to 𝖥𝖬𝗈𝗇(V)\mathsf{FMon}(V) consisting of vectors in VV representing edges of T(|σX|)T(|\sigma_{X}|). This gives us the desired triangulated representative r(X)r(X). ∎

Finally, we treat the general case.

Proof of Theorem 6.16.

Let 𝐗PL1(V)\mathbf{X}\in\operatorname{PL}_{1}(V). By Lemma D.1, there exists a triangulated representative r(𝐗)=(X1,,Xn)𝖥𝖬𝗈𝗇(Kite×(V))r(\mathbf{X})=(X_{1},\ldots,X_{n})\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)). Let Δ(r(𝐗))\Delta(r(\mathbf{X})) be the associated PLSC, let E=|Δ1(r(𝐗))|E=|\Delta_{1}(r(\mathbf{X}))| denote the set of 1-simplices realized as edges in VV, and let P=|Δ2(r(𝐗))|P=|\Delta_{2}(r(\mathbf{X}))| denote the set of 2-simplices realized as convex polygons in VV. Then, by Lemma D.4, there exists a compatible triangulation CT(E,V)C\coloneqq T(E,V).

Given each marked triangular kite Xi=(𝐰i,𝐛i)X_{i}=(\mathbf{w}_{i},\mathbf{b}_{i}), the associated 2-simplex |σXi||\sigma_{X_{i}}| is a polygon in PP. By Lemma D.3, the subcomplex T(|σXi|)CT(|\sigma_{X_{i}}|)\subseteq C is a compatible triangulation of |σXi||\sigma_{X_{i}}|. Therefore, by Lemma D.5, there is a triangulated representative

r(𝐛i)=((𝐯1i,𝐜1i),,(𝐯qii,𝐜qii))\displaystyle r(\mathbf{b}_{i})=((\mathbf{v}_{1}^{i},\mathbf{c}_{1}^{i}),...,(\mathbf{v}_{q_{i}}^{i},\mathbf{c}_{q_{i}}^{i})) (D.9)

such that r(𝐛i)(,𝐛i)r(\mathbf{b}_{i})\sim(\emptyset,\mathbf{b}_{i}) in PL1(V)\operatorname{PL}_{1}(V) and such that Δ(r(𝐛i))T(|σXi|)\Delta(r(\mathbf{b}_{i}))\subseteq T(|\sigma_{X_{i}}|) (after shifting Δ(r(𝐛i))\Delta(r(\mathbf{b}_{i})) by the displacement of 𝐰i\mathbf{w}_{i}). Furthermore, 𝐰i=(u1,,uk)\mathbf{w}_{i}=(u_{1},...,u_{k}) consists of a collection of vectors in VV, which give rise to a subset of edges from EE. Each edge has the form [u^i,u^i+1][\hat{u}_{i},\hat{u}_{i+1}], where u^i+1=u^i+ui\hat{u}_{i+1}=\hat{u}_{i}+u_{i}. Because CC is a compatible triangulation, each such edge decomposes into a sequence of edges |ϵ||\epsilon|, such that ϵC\epsilon\in C is a 1-simplex. As a result, we may replace each uiu_{i} with a sequence of vectors (s1i,,srii)(s_{1}^{i},...,s_{r_{i}}^{i}) which sum to uiu_{i} and such that the corresponding edges [s^ji,s^j+1i][\hat{s}_{j}^{i},\hat{s}_{j+1}^{i}] are 1-simplices of CC. Let 𝐰i𝖥𝖬𝗈𝗇(V)\mathbf{w}_{i}^{\prime}\in\mathsf{FMon}(V) denote the resulting word and note that 𝐰i\mathbf{w}_{i} and 𝐰i\mathbf{w}_{i}^{\prime} are equivalent in PL0(V)\operatorname{PL}_{0}(V). Now define

r(Xi)=((𝐰i𝐯1i,𝐜1i),,(𝐰i𝐯qii,𝐜qii))𝖥𝖬𝗈𝗇(Kite×(V)).\displaystyle r(X_{i})=((\mathbf{w}_{i}^{\prime}\star\mathbf{v}_{1}^{i},\mathbf{c}_{1}^{i}),...,(\mathbf{w}_{i}^{\prime}\star\mathbf{v}_{q_{i}}^{i},\mathbf{c}_{q_{i}}^{i}))\in\mathsf{FMon}({\operatorname{Kite}}^{\times}(V)). (D.10)

By construction, r(Xi)r(X_{i}) is equivalent to XiX_{i} in PL1(V)\operatorname{PL}_{1}(V) and Δ(r(Xi))\Delta(r(X_{i})) is a subcomplex of CC. Finally, define r(𝐗)r(\mathbf{X})^{\prime} to be the product of the r(Xi)r(X_{i}), for i=1,,ni=1,...,n. This is a triangulated representative of 𝐗\mathbf{X} and its associated PLSC Δ(r(𝐗))\Delta(r(\mathbf{X})^{\prime}) is a subcomplex of CC. It is therefore non-degenerate and compatible. Hence r(𝐗)r(\mathbf{X})^{\prime} is a compatible representative for 𝐗\mathbf{X}. ∎

Next, we prove a lemma which is required for Proposition 6.18.

Lemma D.6.

Suppose CC is a 2-dimensional non-degenerate compatible PLSC whose set of 22-simplices is denoted by LL. For each 2-simplex σL\sigma\in L, there exists a compactly supported 2-form ωσΩc2(V)\omega_{\sigma}\in\Omega^{2}_{c}(V) such that σωσ=1\int_{\sigma}\omega_{\sigma}=1 and such that the support of ωσ\omega_{\sigma} is disjoint from all other 2-simplices,

supp(ωσ)|τ|=\displaystyle{\operatorname{supp}}(\omega_{\sigma})\cap|\tau|=\emptyset (D.11)

for all τL\tau\in L such that τσ\tau\neq\sigma.

Proof.

Let σL\sigma\in L and let HVH\subset V denote the 2-plane which supports σ\sigma. Let KHK\subset H denote a closed ball such that K|σ|K\subset|\sigma|^{\circ}. Consider a compactly supported 2-form νΩc2(H)\nu\in\Omega^{2}_{c}(H) on HH which is supported in KK and satisfies σν=1\int_{\sigma}\nu=1. Using a metric on VV, pullback ν\nu along the orthogonal projection π:VH\pi:V\to H to obtain a form π(ν)\pi^{*}(\nu). Then, choose a compactly supported smooth bump function ρ:V\rho:V\to\mathbb{R} such that ρ|K1\rho|_{K}\equiv 1 and which vanishes along each τL\tau\in L, τσ\tau\neq\sigma. Then, we define

ωσρπ(ν)Ωc2(V).\displaystyle\omega_{\sigma}\coloneqq\rho\ \pi^{*}(\nu)\in\Omega^{2}_{c}(V). (D.12)

This is the desired 22-form.

References

  • [1] Camilo Arias Abad and Florian Schätz. The A\infty 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 𝔐¯0,n\overline{\mathfrak{M}}_{0,n}. 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.