Skip to main content
Log in

A Cellular Topological Field Theory

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We present a construction of cellular BF theory (in both abelian and non-abelian variants) on cobordisms equipped with cellular decompositions. Partition functions of this theory are invariant under subdivisions, satisfy a version of the quantum master equation, and satisfy Atiyah–Segal-type gluing formula with respect to composition of cobordisms.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. We will give a short, working-knowledge introduction to the BV and the BV-BFV formalisms in this paper, but the reader is referred to the literature, especially [5], for more details.

  2. Besides casting the model into the BV-BFV setting, with cobordisms and Segal-like gluing, some of the important advancements over [27, 28] are the following: general regular CW complexes are allowed (as opposed to simplicial and cubic complexes); the new construction of the cellular action which is intrinsically finite-dimensional and in particular does not use regularized infinite-dimensional super-traces; a systematic, intrinsically finite-dimensional, treatment of the behavior w.r.t. moves of CW complexes—elementary collapses and cellular aggregations; understanding the constant part of the partition function (leading to the contribution of the Reidemeister torsion and the \(\bmod \,16\) phase); incorporating the twist by a nontrivial local system.

  3. In our presentation of this result (Theorem 8.6), the local unimodular \(L_\infty \) algebras are packaged into generating functions—the local building blocks \(\bar{S}_e\) for the cellular action. To be precise, the sum of building blocks \(\bar{S}_{e'}\) over all cells \(e'\) belonging to the closure of the given cell e is the generating function for the operations (structure constants) of the local unimodular \(L_\infty \) algebra assigned to e, in the sense of Sect. 8.2.1 and (134).

  4. In this paper we use the convention that the polarization is linked to the designation of boundaries as in/out. Thus we link out-boundaries to “A-polarization” and in-boundaries with “B-polarization”. This convention is entirely optional. On the other hand, the link between polarization (69) and the notion of the dual CW complex (Sect. 2.3) is essential for the construction.

  5. Partition functions are defined up to sign for the purposes of this paper, so that we don’t need to keep track of orientations on the spaces of fields and gauge-fixing Lagrangians.

  6. Superscripts pertain to the polarization \(p: {\mathcal {F}}_\partial \rightarrow {\mathcal {B}}_\partial \) (field A fixed on the out-boundary and field B fixed on the in-boundary) used to quantize the in/out-boundary.

  7. The role of these two conditions is to exclude trivial solutions to quantum master equation, e.g. \(S_X=0\), and also to have uniqueness up to homotopy for solutions satisfying the stated properties.

  8. One says that two solutions of the quantum master equation \(S_0\) and \(S_1\) are related by a canonical BV transformation (or “homotopy”) if they can be connected by a family of solutions \(S_t\) such that \(\frac{d}{dt} S_t=\{S_t,R_t\}-i\hbar R_t\) with \(R_t\) a degree \(-1\) “generator.” This definition implies that \(\frac{d}{dt}e^{\frac{i}{\hbar }S_t}=\Delta \Big (e^{\frac{i}{\hbar }S_t} R_t\Big )\) and hence \(\Delta \)-closed exponentials \(e^{\frac{i}{\hbar }S_1}\) and \(e^{\frac{i}{\hbar }S_0}\) differ by a \(\Delta \)-exact term.

  9. This assumption is made for convenience and can be dropped, see Remark 2.2 below.

  10. Throughout this paper we will be working in the piecewise-linear category. One can replace PL manifolds with smooth manifolds everywhere, but then instead of gluing of manifolds along a common boundary, one should talk about cutting a manifold along a submanifold of codimension 1 or work with manifolds with collars in order to achieve the correct gluing of smooth structures. For details on oriented intersection of chains in piecewise-linear setting, we refer the reader to [23].

  11. Recall that a CW complex is said to be regular if the characteristic maps from standard open balls to open cells \(\chi : \mathrm {int}(B^k) \xrightarrow {\sim } e\subset X\) extend to homeomorphisms of closed balls to corresponding closed cells \({\bar{\chi }}: B^k \xrightarrow {\sim } \bar{e}\subset X\). Another term for a regular CW complex is “ball complex.”

  12. The standard terminology is that for a cell e of any CW complex X, the star of e is the subcomplex of X consisting of all cells of X containing e. The link of e is the union of cells of \(\mathrm {star}(e)\) which do not intersect e.

  13. We can again use the construction of Remark 2.1. Cells \(\varkappa (e)\) are then defined exactly as in (3) and cells \(\varkappa _\partial (e)\) for boundary cells \(e\subset X_\partial \) are constructed as \(\varkappa _\partial (e)=\overline{\varkappa (e)}\cap \partial M\).

  14. In other words, we are asking for the intersection of X with a thin tubular neighborhood of the boundary to look like the product CW-complex \(X_\partial \times [0,1]\) intersected with \(X_\partial \times [0,\epsilon )\). Morally, even though there is no metric in our case, one should think of this property as an analog of the property of a Riemannian metric on a smooth manifold with boundary to be of product form near the boundary (Fig. 3).

  15. By convention, we endow \(M_{\mathrm {out}}\) with the orientation induced from the orientation of M, whereas \(M_{\mathrm {in}}\) is endowed with orientation opposite to the one induced from M. Thus, as an oriented manifold the boundary splits as \(\partial M=\overline{M_{\mathrm {in}}}\sqcup M_{\mathrm {out}}\) where the overline stands for orientation reversal.

  16. Observe that \(X^\vee \) is automatically of product type near the in-boundary of M, i.e. near the out-boundary of \({\widetilde{M}}\).

  17. Note that functoriality in this construction corresponds to flatness of \(\nabla _P\).

  18. Having in mind Poincaré duality, we are incorporating the degree shift by n in the definition of the dual (which is superfluous in purely algebraic setting). In our setup, the canonical pairing \(\langle ,\rangle : C^{*\;k}\otimes C^{n-k}\rightarrow \mathbb {R}\) has degree \(-n\).

  19. A special case of this situation is a flat Euclidean vector bundle, i.e. with fiberwise scalar product \((,)_E\) and a flat connection preserving it. In this case the structure group reduces to \(O(m)\subset SL_\pm (m)\).

  20. In the language of the variational bicomplex, \(\delta \) is the “vertical differential” mapping \(\Omega ^p({\mathcal {F}})\rightarrow \Omega ^{p+1}({\mathcal {F}})\). It is formal and we stress its distiction from the “horizontal differential” d—the cellular coboundary operator on cochains of X which does care about the adjacency of cells in X.

  21. We will use the sign convention where the graded binary operation \(\langle ,\rangle \) is understood as taking a cochain on \(X^\vee \) from the left side and a cochain on X from the right side. In other words, the mnemonic rule is that, for the sake of Koszul signs, the comma separating the inputs in \(\langle b,a\rangle \) carries degree \(-n\). This pairing is related to the one which operates from the left on two inputs coming from the right by \(\langle b,a \rangle =(-1)^{n\deg b}\langle b,a \rangle '\).

  22. Recall that, generally, to define the BV Laplacian on functions (as opposed to half-densities) on an odd-symplectic manifold \(\mathcal {M}\), one needs a volume element on \(\mathcal {M}\) [31]. In our case, the space of fields is linear, and so possesses a canonical (constant) volume element determined up to normalization. Since the BV Laplacian is not sensitive to rescaling the volume element by a constant factor, we have a preferred BV Laplacian.

  23. In the context of classical abelian BF theory we could instead work with smooth functions on \({\mathcal {F}}\).

  24. The Euler–Lagrange equations describe the critical locus of S or, equivalently, the zero locus of Q.

  25. The superscript in \(\mu _{\mathcal {F}}^{1/2}\) stands for both the weight of the density and for the square root.

  26. The canonical BV Laplacian is related to the BV Laplacian \(\Delta =\Delta _{\mu _{\mathcal {F}}}\) on functions by \(\Delta _\mathrm {can}(f\, \mu _{\mathcal {F}}^{1/2})=\Delta (f)\,\mu _{\mathcal {F}}^{1/2}\), where \(f\in \mathrm {Fun}({\mathcal {F}})\).

  27. Here the second term on the r.h.s. is nondegenerately paired to the third term on the r.h.s. of (32) by Poincaré duality and vice versa; the first terms are paired between themselves. The map \({\mathbf {K}}^\vee _k: C^k(X^\vee ,E^*)\rightarrow C^{k-1}(X^\vee ,E^*)\) is defined as the dual (transpose) of \({\mathbf {K}}_{n-k+1}: C^{n-k+1}(X,E)\rightarrow C^{n-k}(X,E)\), up to the sign \((-1)^{n-k}\).

  28. That is, a product over cells of X of certain universal elementary factors, depending only on the dimension of the cell, see Lemma 5.6 below.

  29. See [5, Sect. 2.2.2] for details on fiber BV integrals.

  30. Convergence is due to the fact that, by construction, the point \((A,B)=0\) is an isolated critical point of the action S restricted to \({\mathcal {L}}\).

  31. We are using the natural notations \(a|_{\mathrm {in}}, a|_{\mathrm {out}}\) for the components of the image of a cochain a under restriction \(\iota ^*:C^\bullet (X)\rightarrow C^{\bullet }(X_{\mathrm {in}})\oplus C^\bullet (X_{\mathrm {out}})\), and likewise \(b|_{\mathrm {in}},b|_{\mathrm {out}}\) are the components of the image of b under \((\iota ^\vee )^*:C^\bullet (X^\vee )\rightarrow C^\bullet (X_{\mathrm {in}}^\vee )\oplus C^\bullet (X_{\mathrm {out}}^\vee )\).

  32. A-B” means that we stay in the setting of cobordisms and only allow attaching out-boundary (or “A-boundary”, for the polarization we are going to put on it in the quantization procedure to follow) of one cobordism to in- (or “B”-) boundary of the next one.

  33. We also understand that the orientation reversing identity map \(Y\rightarrow {\bar{Y}}\) acts on states by complex conjugation \(\phi (A_Y)\mapsto \overline{\phi (A_Y)}\), \(\psi (B_Y)\mapsto \overline{\psi (B_Y)}\). In particular, we have a sesquilinear pairing \({\mathcal {H}}^{(A)}_Y\otimes {\mathcal {H}}^{(B)}_Y\rightarrow \mathbb {C}\) (note that here Y has the same orientation in both factors), given by the formula (73) with \(\psi \) replaced by the complex conjugate \({\bar{\psi }}\).

  34. Recall that \(\bar{X}_{\mathrm {in}}\) is the CW complex \(X_{\mathrm {in}}\) endowed with the orientation induced from the orientation of X.

  35. This Lemma follows from the general treatment in [5], Sect. 2.4.1. For reader’s convenience, we give an adapted proof here.

  36. We are putting asterisks on boundary conditions \(A_2^*\), \(B_2^*\) at \(X_2\) to distinguish them from the components \(A_2^{II},B_2^I\) of bulk fields—coordinates on fibers of \({\mathcal {F}}_{II}\rightarrow {\mathcal {B}}_2^{(B)}\) and \({\mathcal {F}}_{I}\rightarrow {\mathcal {B}}_2^{(A)}\), respectively. In other words, in (84) we are counting each cell of \(X_2\), \(X_2^\vee \) twice: once as a boundary condition and once as a part of bulk fields.

  37. Indeed, for a k-cell \(e\subset X_2\) we have \(\int \mathcal {D}_\hbar B_{\varkappa _2(e)}^* \mathcal {D}_\hbar A_e^*\quad e^{-\frac{i}{\hbar } \langle B^*_{\varkappa _2(e)} , A^*_e \rangle _E}=(\underbrace{\xi _\hbar ^{n-\dim \varkappa _2(e)} \xi _\hbar ^{\dim e}}_{\xi _\hbar ^{k+1}\xi _\hbar ^k})^{\mathrm {rk} E}\cdot \left\{ \begin{array}{cl} (2\pi \hbar )^{\mathrm {rk}E} &{} \mathrm {if}\; k \;\mathrm {odd} \\ \left( \frac{i}{\hbar }\right) ^{\mathrm {rk}E} &{} \mathrm {if}\; k \;\mathrm {even}\end{array} \right. =1\)

  38. We use the term BV pushforward as a synonym for the fiber BV integral.

  39. Note that, using Poincaré duality on N, the factor \(\widetilde{\xi }_\hbar ^{H^\bullet (N)}\) in (101) can be expressed as \(\prod _k \left( \xi _\hbar ^k\right) ^{\dim H^{k-1}(N)}=\xi _\hbar ^{H^\bullet (N)[-1]}\).

  40. This setting for (continuum) BF theory is known as “canonical BF theory” in the literature, cf. e.g. [7].

  41. Here we are not introducing the twist by a local system and the notation is simply \(C^\bullet (X,\mathfrak {g}):=\mathfrak {g}\otimes C^\bullet (X,\mathbb {R})\)—cellular cochains with coefficients in \(\mathfrak {g}\). Likewise, for the chains, \(C_\bullet (X,\mathfrak {g}^*):=\mathfrak {g}^*\otimes C_\bullet (X,\mathbb {R})\).

  42. Our convention is that the leaves and the root are loose half-edges of the graph.

  43. Indeed, we already know by (109) that \(\{S^{(0)}_X,S^{(0)}_X\}=0\); we are left to check that \(\Delta S^{(0)}_X+\{S^{(0)}_X,S^{(1)}_X\}=0\). We calculate \(\Delta S^{(0)}_X=\Delta _{X_1} S^{(0)}_{X_1}+\Delta _{X_2} S^{(0)}_{X_2}-\Delta _{Y} S^{(0)}_{Y}= -\{S^{(0)}_{X_1},S^{(1)}_{X_1}\}_{X_1}- \{S^{(0)}_{X_2},S^{(1)}_{X_2}\}_{X_2} +\{S^{(0)}_{Y},S^{(1)}_{Y}\}_{Y}=-\{S^{(0)}_X,S^{(1)}_{X_1}+S^{(1)}_{X_2}-S^{(1)}_Y\}=-\{S^{(0)}_X,S^{(1)}_X\}\). (Here we indicate explicitly where we calculate the odd Poisson brackets and BV Laplacians; no index means X.)

  44. The sign is chosen in such a way that \(D\mathsf {A}=d_{\partial e\rightarrow e}\mathsf {A}\) holds.

  45. Indeed, using the induction hypothesis we calculate

    $$\begin{aligned}&(D+Q^0_{\partial e})\left( \text{ r.h.s. } \text{ of } (123) \right) \\&\quad = \sum _{i=1}^{j-1}\sum _{l=0}^{i-1}Q^{i-l}_{\partial e}Q^{l}_{\partial e}\sigma _{j-i}+\sum _{i=1}^{j-1}Q_{\partial e}^i D \sigma _{j-i}-(-1)^n\sum _{i=1}^{j-1}\left\langle Q^i_{\partial e}\sigma _1, \frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _{j-i}\\&-(-1)^n \sum _{i=2}^{j-1} \left\langle (D+Q^0_{\partial e})\sigma _{j+1-i},\frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _i +(-1)^n \sum _{i=2}^{j-1} \left\langle \sigma _{j+1-i},\frac{\partial }{\partial \mathsf {A}_e} \right\rangle (D+Q^0_{\partial e})\sigma _i=:a+b+c+d+e \end{aligned}$$

    Further, we have

    $$\begin{aligned} a+b= & {} \sum _{i=1}^{j-1}Q^i_{\partial e}\left( (D+Q^0_{\partial e})\sigma _{j-i}+\sum _{l=1}^{j-i-1}Q_{\partial e}^l \sigma _{j-i-l}\right) \\= & {} -(-1)^n\sum _{i=1}^{j-1}\,\sum _{r,s\ge 2,r+s=j-i+1}Q^i_{\partial e}\left( \left\langle \sigma _r, \frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _s \right) \\= & {} -(-1)^n\sum _{i=1}^{j-1}\,\sum _{r,s\ge 2,r+s=j-i+1}\left\langle Q^i_{\partial e}\sigma _r, \frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _s +(-1)^n\sum _{i=1}^{j-1}\,\sum _{r,s\ge 2,r+s=j-i+1}\left\langle \sigma _r, \frac{\partial }{\partial \mathsf {A}_e} \right\rangle Q^i_{\partial e}\sigma _s \\=: & {} f+g \end{aligned}$$

    Next we note that \(c+d+f=\sum \nolimits _{r,s,t\ge 2, r+s+t=j+2}\left\langle \left\langle \sigma _r ,\frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _s, \frac{\partial }{\partial \mathsf {A}_e}\right\rangle \sigma _t\) and \(e{+}g=-\sum \nolimits _{r,s,t\ge 2, r+s+t=j+2} \left\langle \sigma _r ,\frac{\partial }{\partial \mathsf {A}_e} \right\rangle \left( \left\langle \sigma _s, \frac{\partial }{\partial \mathsf {A}_e}\right\rangle \sigma _t\right) \), and thus \(c+d+f+e+g=\sum \nolimits _{r,s,t\ge 2, r+s+t=j+2} \left\langle \sigma _r \sigma _s, \frac{\partial }{\partial \mathsf {A}_e}\frac{\partial }{\partial \mathsf {A}_e} \right\rangle \sigma _t=0\)—vanishes as a contraction of a symmetric and a skew-symmetric tensor. Thus we proved that \((D+Q^0_{\partial e})\left( \text{ r.h.s. } \text{ of } (123)\right) =0\).

  46. To see that the r.h.s. of (123) is annihilated by \(p_\mathcal {E}\), note that in \(H^\bullet (\mathcal {E})\), the weight coincides with the internal degree. On the other hand, the weight of (123) is \(j\ge 2\) while the internal degree is \(3-n\le 2\). Thus \(p_\mathcal {E}(\text{ r.h.s. } \text{ of } (123))\) is zero for degree reasons, except for the case \(n=1\) (i.e. e is an interval) with \(j=2\), where one sees explicitly that \(-Q^1_{\partial e}\sigma _1=-\frac{1}{2} \left[ \mathsf {A}_{[1]},\mathsf {A}_{[1]}\right] +\frac{1}{2} \left[ \mathsf {A}_{[0]},\mathsf {A}_{[0]}\right] \) vanishes on constant 0-cochains (we denoted the endpoints of the interval e by [0] and [1], as in Remark 8.4).

  47. Indeed, the most general construction of extension of a solution of QME from \(X_{k-1}\) to \(X_k\) is as in our proof of Theorem 8.6, where on each step of induction in j we can shift \(\sigma _j\rightarrow \sigma _j+(D+Q^{(0)}_{\partial e})(\cdots )\), also we can shift \(\bar{S}^{(1)}_e\rightarrow \bar{S}^{(1)}_e+Q_{\bar{e}}(\cdots )\). This amounts to a contractible space of choices. Thus the space of solutions of QME on \(X_k\) of form \(S_{X_{k-1}}+\bar{s}_{e}\), with a fixed solution \(S_{X_{k-1}}\) of QME on \(X_{k-1}\) and an indeterminate function \(\bar{s}_{e}\) satisfying ansatz (115), is contractible; in particular, it is path-connected.

  48. For this we assume \(\dim e\ge 1\); in the case of \(\dim e=0\) the induction step of the Lemma works trivially as \(\bar{S}_e\) is fixed uniquely by (b) of Theorem 8.6.

  49. This algebraic structure (and the example coming from Theorem 8.1) was introduced in [27, 28] under the name of a quantum\(L_\infty \) algebra. In [13] it was named a unimodular\(L_\infty \) algebra and was studied as an algebra over a particular Merkulov’s wheeled operad.

  50. This algebraic structure appeared independently and nearly simultaneously in [21], in the preprint of [8] and, in its Lie form, in the preprint of [27].

  51. Half-density \(\mu _{\mathsf {F}_X}^{\frac{1}{2}}\) is constructed using an a priori fixed density (i.e. a fixed normalization of the Lebesgue measure) \(\mu _\mathfrak {g}\) on \(\mathfrak {g}\), instead of the standard density on \(\mathbb {R}^m\) as in Sect. 5.2.

  52. By an abuse of notations, throughout Sects. 8 and 9 we are suppressing the superscript \(\mathrm {can}\) for the BV Laplacian on half-densities. Similarly, in Sect. 9 we will suppress this superscript for the operator \(\widehat{S}_\partial \) acting on half-densities.

  53. This is a general property of BV pushforwards for a change of gauge-fixing data in a smooth family, cf e.g. [5], Sect. 2.2.2. Note that by the discussion of Sect. 4.3, in our case the space of gauge-fixing data is contractible and in particular path-connected, thus any two choices of gauge-fixing can be connected by a smooth family. We are slightly abusing the term “canonical BV transformation”: for us it has two related meanings—for \(\Delta \)-closed half-densities, as in (137), and for actions (functions solving QME), as in (133). These meanings are equivalent for half-densities satisfying exponential ansatz \(Z=e^{\frac{i}{\hbar }S}\mu ^{1/2}\).

  54. Here we mean that we need to know \(l_n\)’s (Massey–Lie brackets) for a general Lie algebra of coefficients \(\mathfrak {g}\), which is tantamount to knowing the \(C_\infty \) operations (Massey products) on \(H^\bullet (X)\), see (139) below. In fact, one can recover the n-ary \(C_\infty \) operation \(m_n\) on \(H^\bullet (X)\) from \(l_n\) with \(\mathfrak {g}=\mathfrak {b}^+_{n+1}\) the algebra of upper-triangular matrices of size \(n+1\), simply from \({l_n(w_1\otimes t_{12},\ldots w_{n}\otimes t_{n\, n+1})} =m_n(w_1,\ldots ,w_n)\otimes t_{1\, n+1}\) with \(w_i\in H^\bullet (X)\) and \(t_{ij}\) the matrix with entry 1 at (ij)-th place and all other entries being zero. We stress that one does not recover the \(C_\infty \) structure on \(H^\bullet (X)\) by plugging \(\mathfrak {g}=\mathbb {R}\) into the formulae for operations \(l_n\) on \(H^\bullet (X,\mathfrak {g})\)—that would just kill all the operations.

  55. Recall, see e.g. [8] for details, that an \(A_\infty \) algebra is a \(\mathbb {Z}\)-graded vector space W together with a sequence of multilinear operations \(m_n:W^{\otimes n}\rightarrow W\), \(n\ge 1\), satisfying the quadratic associativity-up-to-homotopy identities. An \(A_\infty \) algebra \((W,\{m_n\})\) is called a \(C_\infty \) algebra if in addition operations \(m_n\) vanish on shuffle-products. We refer the reader to Appendix A.2 on how to construct the tensor product \(L_\infty \) structure on \(W\otimes \mathfrak {g}\), with W a \(C_\infty \) algebra and \(\mathfrak {g}\) a Lie algebra. The \(C_\infty \) structure on \(C^\bullet (X)\), for X a simplicial complex, coinciding with the one read off of the tree part of (106) and constructed via homotopy transfer from piecewise-polynomial forms by Kontsevich-Soibelman formula using Dupont’s chain homotopy operator was considered in [8].

  56. Here the remark (see [27, 28]) is that the perturbative evaluation of the integral (136) in the lowest order in \(\hbar \) corresponds to the homotopy transfer formula for \(L_\infty \) algebras to a subcomplex as a sum over (non-planar) rooted trees. This is the \(L_\infty \) version of Kontsevich-Soibelman formula for homotopy transfer of \(A_\infty \) algebras where the sum is over planar rooted trees. Also, one has that the homotopy transfer commutes with tensoring with a Lie algebra \(\mathfrak {g}\):

  57. For example, Pachner’s moves of triangulations of an n-manifold can be realized as a sequence of elementary expansions followed by a sequence of elementary collapses. Also, cellular subdivisions and aggregations can be realized as sequences of expansions and collapses.

  58. In more detail, the quantum master equation for \(S_h\) and for \(S_h-i\hbar \,\phi \) implies \(Q_h\phi =0\). Function \(\phi \) vanishes at \(\mathsf {A}=0\) (this is the point where we use the normalization of the half-densities \((\mu ^\hbar _{\mathsf {F}_Y})^{1/2},(\mu ^\hbar _{\mathsf {F}_X})^{1/2}\)) which implies, together with (144), that the obstruction to \(Q_h\)-exactness of \(\phi \) vanishes, i.e. we have \(\phi =Q_h \chi \) for some \(\chi \), a function of \(\mathsf {A}_h\) of degree \(-1\). Thus we have proved (143) for the generator \(R_t=-i\hbar \chi \).

  59. To obtain such a decomposition, order the cells \(e_1,\ldots , e_m\) of X in the order of non-decreasing dimension, as in the proof of Theorem 8.6. Then set \(X_k:=\left( \bigcup _{i\le k} Y|_{e_i}\right) \cup \left( \bigcup _{i>k} e_i\right) \).

  60. Note also that induction data (152) factors through integral cochains, \(C^\bullet (Y,\mathbb {Z}){\mathop {\rightsquigarrow }\limits ^{}}C^\bullet (X,\mathbb {Z})\).

  61. We introduce this term to avoid confusion with the cohomology of the BV Laplacian \(\Delta \) itself which is quite different (perturbative BV cohomology contains more information), see (i) vs. (vii) below.

  62. This follows from writing \(\mathsf {F}\) as an odd cotangent bundle \(T^*[-1]N\) of an evenly graded vector space \(N=\bigoplus _k V^{2k+1}\oplus \bigoplus _k (V^{2k})^*\) spanned by even components of \(\mathsf {A}\), \(\mathsf {B}\). Then \(\mathrm {Fun}(\mathsf {F}),\Delta \) is identified with the space of polyvector fields on N with differential given by the divergence operator; this complex in turn is isomorphic to the de Rham complex of N via odd Fourier transform [31]. Since N is a vector space, its de Rham cohomology is given by constant 0-forms on V; their odd Fourier transform gives \(\nu \cdot \mathbb {R}\).

  63. For this, we observe that the bound above can be improved to be uniform in k: structure constants \(C^{\mathbf {\Delta }^m}_{\Gamma _0,e_1,\ldots ,e_k}\) (for a simplex of fixed dimension m) have at most exponential growth in k, as follows from analyzing the explicit integral expressions for the structure constants arising from the construction (i) of the proof of Theorem 8.1. Thus, one has \(|S^{(0),k}|<\gamma ^k N^{2-k}\) for some \(\gamma \) depending on (AB) but independent of k.

  64. The argument we used to prove Lemma 8.28 uses Whitney forms, and it is not clear how to extend it to CW decompositions with arbitrary cells.

  65. Recall from Sect. 7 that we have, in fact, two models for the space of states, \({\mathcal {H}}_\partial \) (functions on \({\mathcal {B}}_\partial \)) and \({\mathcal {H}}_\partial ^\mathrm {can}\) (half-densities on \({\mathcal {B}}_\partial \)). The two models are isomorphic and the comparison goes via multiplication by the (appropriately normalized) reference half-density \(\psi \mapsto (\mu _{{\mathcal {B}}_\partial }^\hbar )^{1/2}\psi \).

  66. Note that in our case the cochains on \(X_\partial \) are twisted by a local system but the action (116) still satisfies the master equation, with the new definition (165) of \(A|_{\bar{e}}\), as can be seen by inspecting the proof of Theorem 8.6: we have quantum master equations on cells, where the local system is trivialized and this implies (by the gluing procedure (vi) of the proof of Theorem 8.1) that (116) is a solution of the master equation.

  67. In more detail, we have

    $$\begin{aligned} \widehat{S}_{\mathrm {in}}\circ e^{\frac{i}{\hbar }S}= & {} \left( \sum _{e\subset X_{\mathrm {in}}}\sum _{k\ge 1}\frac{1}{k!} \left\langle B_{\varkappa _{\mathrm {in}}(e)},l_k^e\left( \widehat{A}|_{\bar{e}},\ldots , \widehat{A}|_{\bar{e}}\right) \right\rangle \right) \circ e^{\frac{i}{\hbar }S} \\= & {} \left( \sum _{e\subset X_{\mathrm {in}}}\sum _{k\ge 1}\frac{1}{k!} \left\langle B_{\varkappa _{\mathrm {in}}(e)},l_k^e\left( \widehat{A}|_{\bar{e}}\circ \frac{i}{\hbar }S,\ldots , \widehat{A}|_{\bar{e}}\circ \frac{i}{\hbar }S\right) \right\rangle \right) \cdot e^{\frac{i}{\hbar }S} \\= & {} \left( \sum _{e\subset X_{\mathrm {in}}}\sum _{k\ge 1}\frac{1}{k!} \left\langle B_{\varkappa _{\mathrm {in}}(e)},l_k^e\left( A|_{\bar{e}},\ldots , A|_{\bar{e}}\right) \right\rangle \right) \cdot e^{\frac{i}{\hbar }S}= S_{\mathrm {in}}\cdot e^{\frac{i}{\hbar }S} \end{aligned}$$
  68. We talk here about “placing” elements of the graph \(\Gamma \) at cells and decorating them with particular tensors depending on the placement. In the end, to obtain the Feynman weight of the graph \(\varphi _\Gamma \), we sum over placements the contraction of the respective tensors. Sum over placements here is a cellular analog of configuration space integrals defining the weights of Feynman graphs in [5].

  69. Note that \(\varphi _\Gamma \) is a polynomial in the variables \((B_{\mathrm {in}},A_{\mathrm {out}},A_\mathrm {res},B_\mathrm {res})\) of degree \((V_{\mathrm {in}},V_{\mathrm {out}},V^A_\mathrm {res},V^B_\mathrm {res})\).

  70. Note that operators \(\widehat{S}_\partial ,\Delta _\mathrm {res}\) here act on half-densities. So, in the conventions of Sect. 7, we should be writing \(\widehat{S}_\partial ^\mathrm {can},\Delta _\mathrm {res}^\mathrm {can}\). We omit here the superscript \(\mathrm {can}\) to lighten the notation.

  71. Note that only \(i_{\mathcal {B}}\) (choice of cellular representatives for cohomology classes) is relevant for the construction of \(Z^\mathrm {r}= i_{\mathcal {B}}^* Z\) whereas \(p_{\mathcal {B}}\) and \(K_{\mathcal {B}}\) are manifestly irrelevant for \(Z^\mathrm {r}\); however, the whole package \((i_{\mathcal {B}},p_{\mathcal {B}},K_{\mathcal {B}})\) is involved in the construction of \(\widehat{S}^\mathrm {r}_\partial \).

  72. Sketch of proof: for Y an arbitrary cell decomposition of the cylinder \(\Sigma \times [0,1]\) (regarded as a cobordism from \(\Sigma \) to \(\Sigma \)) the reduced partition function \(Z^\mathrm {r}: {\mathcal {H}}^\mathrm {r}_{\Sigma }\rightarrow {\mathcal {H}}^\mathrm {r}_\Sigma \) is chain homotopic to identity (proven from the gluing property – (iv) of Theorem 9.7). Now let X and \(X'\) be two cellular decompositions of M. We can attach two cylinders at in- and out-boundaries of \(X'\) to obtain a cell decomposition \({\widetilde{X}}\) of \({\widetilde{M}}\) – a copy of M with collars attached at in- and out-boundary, such that \(\widetilde{X}_{\mathrm {in}}\simeq X_{\mathrm {in}}\) and \(\widetilde{X}_{\mathrm {out}}\simeq X_{\mathrm {out}}\). By the previous observation about cylinders yielding identity up to homotopy, and by gluing formula, we have \(Z_{\widetilde{X}}^\mathrm {r}\sim Z_{X'}^\mathrm {r}\) (where \(\sim \) stands for equality up to \((\frac{i}{\hbar }\widehat{S}_\partial ^\mathrm {r}-i\hbar \Delta _\mathrm {res})\)-exact terms). On the other hand, we can view \(\widetilde{X}\) and X as two cellular decompositions of M coinciding on the boundary, thus (iii) of Theorem 9.7 applies and we have \(Z_{\widetilde{X}}^\mathrm {r}\sim Z_X^\mathrm {r}\). Thus, we have \(Z_{X}^\mathrm {r}\sim Z^\mathrm {r}_{X'}\).

  73. More pedantically, (185) should be written as \({\mathbb {T}}_\mathrm {LES}(\tau (X,Y;E)\cdot \tau (X;E)^{-1}\cdot \tau (Y;E|_Y))=1\), with \({\mathbb {T}}_\mathrm {LES}\) the isomorphism (181) between the determinant line of the long exact sequence of cohomology of the pair and the standard line \(\mathbb {R}\). Likewise, (186) should be written as \({\mathbb {T}}_\mathrm {MV}(\tau (X\cup Y;E)\cdot \tau (X;E|_X)^{-1}\cdot \tau (Y;E|_Y)^{-1}\cdot \tau (X\cap Y;E|_{X\cap Y}))=1\) with \({\mathbb {T}}_\mathrm {MV}\) the isomorphism (181) for the Mayer–Vietoris long exact sequence.

References

  1. Alexandrov, M., Kontsevich, M., Schwarz, A., Zaboronsky, O.: The geometry of the master equation and topological quantum field theory. Int. J. Mod. Phys. A12, 1405–1430 (1997)

    Google Scholar 

  2. Alekseev, A., Mnev, P.: One-dimensional Chern–Simons theory. Commun. Math. Phys. 307(1), 185–227 (2011)

    Google Scholar 

  3. Cattaneo, A.S., Mnev, P.: Remarks on Chern–Simons invariants. Commun. Math. Phys. 293(3), 803–836 (2010)

    Google Scholar 

  4. Cattaneo, A.S., Mnev, P., Reshetikhin, N.: Classical BV theories on manifolds with boundary. Commun. Math. Phys. 332(2), 535–603 (2014)

    Google Scholar 

  5. Cattaneo, A.S., Mnev, P., Reshetikhin, N.: Perturbative quantum gauge theories on manifolds with boundary. Commun. Math. Phys. 357, 631–730 (2018)

    Google Scholar 

  6. Cattaneo, A.S., Mnev, P., Reshetikhin, N.: Perturbative BV theories with Segal-like gluing, arXiv:1602.00741 (math-ph)

  7. Cattaneo, A.S., Rossi, C.: Wilson surfaces and higher dimensional knot invariants. Commun. Math. Phys. 256(3), 513–537 (2005)

    Google Scholar 

  8. Cheng, X.Z., Getzler, E.: Transferring homotopy commutative algebraic structures. J. Pure Appl. Algebra 212(11), 2535–2542 (2008)

    Google Scholar 

  9. Cohen, M.M.: A Course in Simple-Homotopy Theory. Springer, Berlin (2012)

    Google Scholar 

  10. Dupont, J.: Curvature and Characteristic Classes, Lecture Notes in Math., no. 640, Springer, Berlin (1978)

  11. Getzler, E.: Lie theory for nilpotent L-infinity algebras. Ann. Math. 170(1), 271–301 (2009)

    Google Scholar 

  12. Gorodentsev, A.L., Khoroshkin, A.S., Rudakov, A.N.: On syzygies of highest weight orbits, arXiv:math/0602316

  13. Granåker, J.: Unimodular L-infinity algebras, arXiv:0803.1763 (math.QA)

  14. Gugenheim, V.K.A.M., Lambe, L.A., Stasheff, J.: Perturbation theory in differential homological algebra. I. Illinois J. Math. 33.4, 566–582 (1989)

    Google Scholar 

  15. Gwilliam, O.: Factorization algebras and free field theories. Northwestern University, Ph.D. diss. (2012). http://people.mpim-bonn.mpg.de/gwilliam/thesis.pdf

  16. Gwilliam, O., Johnson-Freyd, Th.: How to derive Feynman diagrams for finite-dimensional integrals directly from the BV formalism. In: Topological and Geometric Methods in Quantum Field Theory. AMS, pp. 175–185 (2018)

  17. Kadeishvili, T.: \(A_\infty \)-algebra structure in cohomology and rational homotopy type. Proc. A. Razmadze Math. Inst 107, 1–94 (1993). (in Russian)

    Google Scholar 

  18. Kadeishvili, T.: Cohomology \(C_\infty \)-algebra and rational homotopy type. Banach Center Publ. 85(1), 225–240 (2009)

    Google Scholar 

  19. Khudaverdian, H.: Semidensities on odd symplectic supermanifolds. Commun. Math. Phys. 247(2), 353–390 (2004)

    Google Scholar 

  20. Kontsevich, M., Soibelman, Y : Homological mirror symmetry and torus fibrations, arXiv:math/0011041 [math.SG]

  21. Lawrence, R., Sullivan, D.: A free differential lie algebra for the interval, arXiv:math.AT/0610949

  22. Manin, YuI: Gauge Field Theory and Complex Geometry. Springer, Berlin (1988)

    Google Scholar 

  23. McClure, J.E.: On the chain-level intersection pairing for PL manifolds. Geom. Topol. 10(3), 1391–1424 (2006)

    Google Scholar 

  24. Milnor, J.: A duality theorem for Reidemeister torsion. Ann. Math. 76, 137–147 (1962)

    Google Scholar 

  25. Milnor, J.: Whitehead torsion. Bull. AMS 72(3), 358–426 (1966)

    Google Scholar 

  26. Mnev, P.: Towards simplicial Chern–Simons theory, I, unpublished draft (2005) https://www3.nd.edu/~pmnev/Towards_simplicial_CS.pdf

  27. Mnev, P.: Notes on simplicial \(BF\) theory. Moscow Math. J. 9(2), 371–410 (2009)

    Google Scholar 

  28. Mnev, P.: Discrete \(BF\) theory, Ph.D. thesis, arXiv:0809.1160 (hep-th)

  29. Mnev, P.: Lecture notes on torsions, arXiv:1406.3705 (math.AT)

  30. Schwarz, A.S.: The partition function of a degenerate functional. Commun. Math. Phys. 67(1), 1–16 (1979)

    Google Scholar 

  31. Schwarz, A.S.: Geometry of Batalin–Vilkovisky quantization. Commun. Math. Phys. 155, 249–260 (1993)

    Google Scholar 

  32. Ševera, P.: On the origin of the BV operator on odd symplectic supermanifolds. Lett. Math. Phys. 78(1), 55–59 (2006)

    Google Scholar 

  33. Sullivan, D.: Infinitesimal computations in topology. Publ. Math. de l’IHÉS 47, 269–331 (1977)

    Google Scholar 

  34. Turaev, V.: Introduction to Combinatorial Torsions. Birkhäuser, Boston (2012)

    Google Scholar 

  35. Whitehead, J.H.C.: Simple homotopy types. Am. J. Math. 72, 1–57 (1950)

    Google Scholar 

  36. Whitney, H.: Geometric Integration Theory. Princeton University Press, Princeton (1957)

    Google Scholar 

Download references

Acknowledgements

We thank Nikolai Mnev for inspiring discussions, crucial to this work. We are grateful to the anonymous referee for insightful comments and questions that helped improve the paper. P. M. thanks the University of Zurich and the Max Planck Institute of Mathematics in Bonn, where he was affiliated during different stages of this work, for providing the excellent research environment.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pavel Mnev.

Additional information

Communicated by H. T. Yau

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This research was (partly) supported by the NCCR SwissMAP, funded by the Swiss National Science Foundation, and by the COST Action MP1405 QSPACE, supported by COST (European Cooperation in Science and Technology). P. M. acknowledges partial support of RFBR Grant No. 17-01-00283a.

Appendices

Appendix A. Determinant Lines, Densities, R-Torsion

1.1 Determinant lines, torsion of a complex of vector spaces

In what follows, line stands for an abstract 1-dimensional real vector space.

Definition A.1

For V a finite-dimensional real vector space, the determinant line is defined as the top exterior power \({\mathrm {Det}}\; V=\wedge ^{\dim V} V\). For \(V^\bullet \) a \({\mathbb {Z}}\)-graded vector space, one defines the determinant line as

$$\begin{aligned} {\mathrm {Det}}\; V^\bullet =\bigotimes _k ({\mathrm {Det}}\; V^k)^{(-1)^k} \end{aligned}$$

where for L a line, \(L^{-1}=L^*\) denotes the dual line.

Here are several useful properties of determinant lines.

  1. (i)

    The determinant line of the dual graded vector space is

    $$\begin{aligned} {\mathrm {Det}}\; V^*\cong \left( {\mathrm {Det}}\; V\right) ^{-1} \end{aligned}$$

    (with the grading convention \((V^*)^k=(V^{-k})^*\)). In the case of a vector space concentrated in degree 0, the pairing between \({\mathrm {Det}}\; V^*\) and \({\mathrm {Det}}\;V\) is given by

    $$\begin{aligned} \langle v_n^*\wedge \cdots \wedge v_1^*\;,\;v_1\wedge \cdots \wedge v_n \rangle ={\det }\langle v_i^*,v_j\rangle \end{aligned}$$

    with \(v_i\in V\), \(v_i^*\in V^*\) for \(i=1,\ldots ,n=\dim V\). Extension to the graded case is straightforward.

  2. (ii)

    Determinant line of the degree-shifted vector space is

    $$\begin{aligned} {\mathrm {Det}}\; V^\bullet [k]\cong \left( {\mathrm {Det}}\; V^\bullet \right) ^{(-1)^k}. \end{aligned}$$
  3. (iii)

    Given a short exact sequence of graded vector spaces \(0\rightarrow U^\bullet \rightarrow V^\bullet \rightarrow W^\bullet \rightarrow 0\), one has

    $$\begin{aligned} {\mathrm {Det}}\;V^\bullet \cong {\mathrm {Det}}\;U^\bullet \otimes {\mathrm {Det}}\;W^\bullet . \end{aligned}$$
    (180)

    In the case of non-graded vector spaces, the isomorphism sends

    $$\begin{aligned} \underbrace{(u_1\wedge \cdots \wedge u_{\dim U})}_{\in \; {\mathrm {Det}}\;U}\otimes \underbrace{(w_1\wedge \cdots \wedge w_{\dim W})}_{\in \;{\mathrm {Det}}\; W}\quad \mapsto \quad \underbrace{u_1\wedge \cdots \wedge u_{\dim U}\wedge w'_1\wedge \cdots \wedge w'_{\dim W}}_{\in \; {\mathrm {Det}}\;V} \end{aligned}$$

    where on the right, \(w'_i\) is some lifting of the element \(w_i\) to V. Extension to the graded case is, again, straightforward.

  4. (iv)

    If \(V^\bullet ,d\) is a cochain complex with cohomology \(H^\bullet (V)\), there is a canonical isomorphism of determinant lines

    $$\begin{aligned} {\mathbb {T}}:\; {\mathrm {Det}}\; V^\bullet \xrightarrow {\cong } {\mathrm {Det}}\; H^\bullet (V). \end{aligned}$$
    (181)

    Indeed, one applies property (180) to the two short exact sequences

    $$\begin{aligned} V_\mathrm {closed}^\bullet \hookrightarrow V^\bullet \xrightarrow {d} V^{\bullet +1}_\mathrm {exact},\qquad V^\bullet _\mathrm {exact}\hookrightarrow V^\bullet _\mathrm {closed}\rightarrow H^\bullet (V) \end{aligned}$$

    to obtain isomorphisms.

    $$\begin{aligned} {\mathrm {Det}}\; V^\bullet \cong {\mathrm {Det}}\;V^\bullet _\mathrm {closed} \otimes ({\mathrm {Det}}\;V^\bullet _\mathrm {exact})^{-1},\qquad {\mathrm {Det}}\;V^\bullet _\mathrm {closed}\cong {\mathrm {Det}}\;V^\bullet _\mathrm {exact}\otimes {\mathrm {Det}}\; H^\bullet (V) \end{aligned}$$

    which combine to (181).

All isomorphisms above are canonical (functorial).

It is convenient to work with determinant lines modulo signs, so that one can ignore the question of orientations and Koszul signs. We will use the notation \({\mathrm {Det}}\;V^\bullet /\{\pm 1\}\) for non-zero elements of the determinant line considered modulo sign; so the precise notation should have been \(({\mathrm {Det}}\;V^\bullet -\{0\})/\{\pm 1\}\).

Definition A.2

For \(V^\bullet ,d\) a cochain complex and \(\mu \in {\mathrm {Det}}\;V^\bullet /\{\pm 1\}\) a preferred element of the determinant line, defined up to sign, the torsion is defined as

$$\begin{aligned} \tau (V^\bullet ,d,\mu )={\mathbb {T}}(\mu )\quad \in {\mathrm {Det}}\;H^\bullet (V)/\{\pm 1\} \end{aligned}$$

with \({\mathbb {T}}\) as in (181).

The following Lemma has important consequences in the setting of R-torsion (Sect. A.2).

Lemma A.3

(Multiplicativity of torsions with respect to short exact sequences). Let \(0\rightarrow U^\bullet \rightarrow V^\bullet \rightarrow W^\bullet \rightarrow 0\) be a short exact sequence of complexes, equipped with elements \(\mu _U,\mu _V,\mu _W\) in respective determinant lines, such that \(\mu _V=\mu _U\cdot \mu _W\). Then for the torsions we have

$$\begin{aligned} {\mathbb {T}}_\mathrm {LES}({\mathbb {T}}_U(\mu _U)\cdot {\mathbb {T}}_V(\mu _V)^{-1}\cdot {\mathbb {T}}_W(\mu _W))=1\quad \in \mathbb {R}/\{\pm 1\} \end{aligned}$$

where \({\mathbb {T}}_{U},{\mathbb {T}}_V,{\mathbb {T}}_W\) are the maps (181) for \(U^\bullet ,V^\bullet ,W^\bullet \). We denoted \(\mathrm {LES}\) the induced long exact sequence in cohomology \(\cdots \rightarrow H^k(U)\rightarrow H^k(V)\rightarrow H^k(W)\rightarrow H^{k+1}(U)\rightarrow \cdots \) viewed as an acyclic cochain complex, and

$$\begin{aligned} {\mathbb {T}}_\mathrm {LES}: {\mathrm {Det}}\; H^\bullet (U)\otimes ({\mathrm {Det}}\; H^\bullet (V))^{-1}\otimes {\mathrm {Det}}\; H^\bullet (W)\rightarrow \mathbb {R}\end{aligned}$$

is the corresponding isomorphism (181).

See [25] for details; cf. also [29] for discussion in the language of determinant lines.

1.2 A.2 Densities

Definition A.4

For \(\alpha \in \mathbb {R}\) and V a finite-dimensional real vector space, the space \({\mathrm {Dens}}^\alpha (V)\) of \(\alpha \)-densities on V is defined as the space of maps \(\phi :F(V)\rightarrow \mathbb {R}_+\) from the space of bases (frames) in V to positive half-line satisfying the equivariance property: for any automorphism \(g\in GL(V)\) and any frame \(\underline{v}=(v_1,\ldots ,v_{\dim V})\in F(V)\), one has

$$\begin{aligned} \phi (g\cdot \underline{v})=|\det g|^\alpha \cdot \phi (\underline{v}). \end{aligned}$$

\({\mathrm {Dens}}^\alpha (V)\) is a torsor over \(\mathbb {R}_+\) (viewed as a multiplicative group), and in the setting of \({\mathbb {Z}}\)-graded vector spaces, one defines

$$\begin{aligned} {\mathrm {Dens}}^\alpha (V^\bullet )=\bigotimes _k \left( {\mathrm {Dens}}^\alpha (V^k)\right) ^{(-1)^k} \end{aligned}$$

(tensor product is over \(\mathbb {R}_+\)); \(\alpha \) is called the weight of the density.

By default a “density” has weight \(\alpha =1\) (and then we write \({\mathrm {Dens}}\) instead of \({\mathrm {Dens}}^1\)), and a “half-density” has, indeed, \(\alpha =1/2\).

If \(\phi _\alpha \), \(\phi _\beta \) are two densities on \(V^\bullet \) of weights \(\alpha ,\beta \), then the product \(\phi _\alpha \cdot \phi _\beta \) is an \((\alpha +\beta )\)-density. Also, \(\phi _\alpha \) can be raised to any real power \(\gamma \in \mathbb {R}\) to yield a density \((\phi _\alpha )^\gamma \) of weight \(\alpha \cdot \gamma \). In particular, one has mutually inverse maps

$$\begin{aligned} {\mathrm {Dens}}^{1/2}V^\bullet \xrightarrow {(*)^2}{\mathrm {Dens}}\;V^\bullet ,\quad {\mathrm {Dens}}\;V^\bullet \xrightarrow {\sqrt{*}}{\mathrm {Dens}}^{1/2}V^\bullet . \end{aligned}$$

Evaluation pairing \(({\mathrm {Det}}\; V^\bullet /\{\pm 1\})\otimes {\mathrm {Dens}}\;V^\bullet \rightarrow \mathbb {R}_+\) induces a canonical isomorphism of \(\mathbb {R}_+\)-torsors

$$\begin{aligned} {\mathrm {Det}}\;V^\bullet /\{\pm 1\}\cong {\mathrm {Dens}}\;V^\bullet [1]. \end{aligned}$$

1.3 A.3 R-torsion

Let X be a finite CW-complex and \(Y\subset X\) a CW-subcomplex. Let

$$\begin{aligned} h:\pi _1(X)=\pi _1\rightarrow SL_{\pm }(m,\mathbb {R}) \end{aligned}$$
(182)

be some representation of the fundamental group of X by real matrices of determinant \(\pm 1\). It extends to a ring homomorphism \(h:{\mathbb {Z}}[\pi _1]\rightarrow \mathrm {Mat}(m,\mathbb {R})\) from the group ring of \(\pi _1\) to all real matrices of size m. Let \(p: \widetilde{X}\rightarrow X\) be the universal cover of X and denote \(\widetilde{Y}=p^{-1}(Y)\subset \widetilde{X}\). Consider the cochain complex of vector spaces

$$\begin{aligned} C^\bullet (X,Y;h)=\mathbb {R}^m\otimes _{{\mathbb {Z}}[\pi _1]} C^\bullet (\widetilde{X},\widetilde{Y};{\mathbb {Z}}) \end{aligned}$$
(183)

where on the right we have integral cellular cochains of the pair \((\widetilde{X},\widetilde{Y})\), which is a complex of free \({\mathbb {Z}}[\pi _1]\)-modules with elements of \(\pi _1\) acting on cells of \(\widetilde{X}\) by covering transformations, tensored with \(\mathbb {R}^m\) using the representation h. In \(C^\bullet (X,Y;h)\) one has a preferred basis of the form

$$\begin{aligned} \{v_i\otimes (\widetilde{e})^* \}_{1\le i\le m,\; e\subset X-Y} \end{aligned}$$
(184)

where \(\{v_i\}\) is the standard basis on \(\mathbb {R}^m\) (or any unimodular basis, i.e. such that the standard density on \(\mathbb {R}^m\) evaluates on it to \(\pm 1\)) and \(\widetilde{e}\) are some liftings of cells e of X not lying in Y to the universal cover; \((\widetilde{e})^*\) is the corresponding basis cochain.

Associated to the basis (184) by construction (29) is an element \(\mu \in {\mathrm {Det}}\;C^\bullet (X,Y;h)/\{\pm 1\}\), independent of the choices of liftings of cells \(e\mapsto \widetilde{e}\) and independent of the choice of unimodular basis in \(\mathbb {R}^m\).

Definition A.5

The R-torsion of the pair (XY) of CW-complexes, associated to the representation (182), is defined as the torsion (in the sense of Definition A.2) of the complex \(C^\bullet (X,Y;h)\) equipped with element \(\mu \):

$$\begin{aligned} \tau (X,Y;h)={\mathbb {T}}(\mu )\quad \in {\mathrm {Det}}\;H^\bullet (X,Y;h)/\{\pm 1\}. \end{aligned}$$

Torsion of a single CW-complex X is defined as \(\tau (X;h):=\tau (X,\varnothing ;h)\).

Of particular importance (and historically the most studied) is the acyclic case, when \(H^\bullet (X,Y;h)=0\). Then the torsion takes values in the trivial line and thus is a number (modulo sign).

Instead of choosing a representation h of \(\pi _1\), one can choose a cellular local \(SL_\pm (m)\)-system E on X, in the sense of Sect. 3, and define h as the holonomy of E. Cochain complex \(C^\bullet (X,Y;E)\) (dual to the chain complex \(C_\bullet (X,Y;E^*)\) constructed in Sect. 3) is isomorphic to (183). When we prefer to think in terms of a local system E rather than a representation h of \(\pi _1\) (e.g. when we consider restriction to a CW-subcomplex, or gluing of two complexes along a subcomplex), we will write the torsion as \(\tau (X,Y;E)\).

The following two properties are consequences of the multiplicativity of the algebraic torsion with respect to short exact sequences of cochain complexes (Lemma A.3).

  1. (A)

    For \(X\supset Y\) a pair of CW-complexes, one has

    $$\begin{aligned} \tau (X;E)=\tau (X,Y;E)\cdot \tau (Y;E|_Y). \end{aligned}$$
    (185)

    The formula makes sense because \({\mathrm {Det}}\;H^\bullet (X;E)\cong {\mathrm {Det}}\;H^\bullet (X,Y;E)\otimes {\mathrm {Det}}\;H^\bullet (Y;E|_Y)\), since the determinant line of the long exact sequence in homology of the pair (XY) (regarded itself as a complex) is \({\mathrm {Det}}\;H^\bullet (X,Y;E)\otimes ({\mathrm {Det}}\;H^\bullet (X;E))^{-1} \otimes {\mathrm {Det}}\;H^\bullet (Y;E|_Y)\) and, on the other hand, is the trivial line, by (181) applied to the long exact sequence.

  2. (B)

    For \(Z=X\cup Y\) a CW-complex represented as a union of two intersecting subcomplexes, one the gluing (inclusion/exclusion) formula

    $$\begin{aligned} \tau (X\cup Y;E)=\tau (X;E|_X)\cdot \tau (Y;E|_Y)\cdot \tau (X\cap Y;E|_{X\cap Y})^{-1}. \end{aligned}$$
    (186)

    The reason why l.h.s. and r.h.s. can be at all compared is as in (A), but one replaces the long exact sequence of a pair by Mayer–Vietoris sequence.Footnote 73

In the acyclic case (i.e. when all relevant cohomology spaces vanish), (185, 186) are equalities of numbers.

Theorem A.6

(Combinatorial invariance of R-torsion). If \((X',Y')\) is a cellular subdivision of the pair (XY), then

$$\begin{aligned} \tau (X',Y';h)=\tau (X,Y;h). \end{aligned}$$

For the proof, see e.g. [25]. The case \(Y=Y'=\varnothing \) is due to Reidemeister, Franz and de Rham.

The combinatorial invariance theorem implies in particular that, for M a compact PL manifold with two different cellular decompositions X and Y, one has \(\tau (X;h)=\tau (Y;h)\). Thus in this case it makes sense to talk about the R-torsion of a manifold M, \(\tau (M;h)\), forgetting about the cellular subdivision.

Theorem A.7

(Milnor [24]). If M is a piecewise-linear compact oriented n-manifold with boundary \(\partial M=\partial _1 M\sqcup \partial _2 M\), one has

$$\begin{aligned} \tau (M,\partial _1 M;h)=(\tau (M,\partial _2 M;h^*))^{(-1)^{n-1}} \end{aligned}$$

where \(h^*\) is the dual representation to h.

Note that the l.h.s. belongs to \({\mathrm {Det}}\; H^\bullet (M,\partial _1 M;h)\) while the r.h.s. belongs to \(({\mathrm {Det}}\; H^\bullet (M,\partial _2 M;h^*))^{(-1)^{n-1}}\) (modulo signs); these determinant lines are canonically isomorphic due to Poincaré-Lefschetz duality \(H^k(M,\partial _1 M;h)\cong (H^{n-k}(M,\partial _2 M;h^*))^*\). Thus it does make sense to compare the two torsions.

Corollary A.8

For M a closed even-dimensional manifold and h such that \(H^\bullet (M;h)=0\), the torsion is trivial, \(\tau (M;h)=1\).

Appendix B. Two Points of View on “\(C_\infty \otimes \mathrm {Lie}=L_\infty \)

In connection with Remark 8.17, we recall two ways to see the \(L_\infty \) algebra structure on the tensor product of a \(C_\infty \) algebra and a Lie algebra.

Given a \(C_\infty \) algebra W with multlinear operations \(m_n:W^{\otimes n}\rightarrow W\), and given a Lie algebra \(\mathfrak {g}\), one can construct the tensor product \(L_\infty \) algebra structure on the graded vector space \(W\otimes \mathfrak {g}\) by defining

$$\begin{aligned} l_n(w_1\otimes \alpha _1,\ldots ,w_n\otimes \alpha _n)=\sum _{\sigma \in S_n}\pm m_n(w_{\sigma _1},\ldots ,w_{\sigma _n})\otimes (\alpha _{\sigma _1}\cdots \alpha _{\sigma _n}) \end{aligned}$$
(187)

with \(w_1,\ldots ,w_n\in W\) and \(\alpha _1,\ldots ,\alpha _n\in \mathfrak {g}\) arbitrary elements. The sum on the r.h.s. is over permuations \(\sigma \) of \(1,\ldots ,n\). Here the product of \(\alpha _i\)’s is seen as a product in the universal enveloping algebra \(U\mathfrak {g}\). The \(C_\infty \) property of the operation \(m_n\) (vanishing on shuffle-products) implies that the result lands in \(W\otimes \mathfrak {g}\subset W\otimes U\mathfrak {g}\).

Another way to present the same tensor product \(L_\infty \) structure on \(W\otimes \mathfrak {g}\) is as follows. The \(C_\infty \) operations \(m_n\) can be written in the form

$$\begin{aligned} m_n(w_1,\ldots ,w_n)=\sum _{T,\pi } m_n^T\circ \pi ^{-1}(w_1\otimes \cdots \otimes w_n) \end{aligned}$$
(188)

where the sum runs over binary rooted trees T with n leaves (viewed up to graph automorphism; for each T we fix arbitrarily a “standard” planar realization) and their planar realizations \(\pi \); \(m_n^T\in \mathrm {Hom}(W^{\otimes n},W)^{\mathrm {Aut}(T)}\) are some multilinear operations invariant w.r.t. automorphisms of T acting by permutations of factors in \(W^{\otimes n}\) (with appropriate signs); \(w_1,\ldots ,w_n\in W\) are arbitrary vectors; \(\pi ^{-1}(\cdots )\) is understood as a permutation of factors in \(W^{\otimes n}\) corresponding to going from the planar representative \(\pi \) to the “standard” representative of T. Then the tensor product \(L_\infty \) algebra structure on \(W\otimes \mathfrak {g}\) is given by

$$\begin{aligned}&l_n(w_1\otimes \alpha _1,\ldots , w_n\otimes \alpha _n)\nonumber \\&\quad =\sum _{\sigma \in S_n}\sum _T \pm \frac{1}{|\mathrm {Aut}(T)|} m_n^T(w_{\sigma _1},\ldots ,w_{\sigma _n}) \otimes \mathrm {Jacobi}_T(\alpha _{\sigma _1},\ldots \alpha _{\sigma _n}) \end{aligned}$$
(189)

with \(\mathrm {Jacobi}_T(\cdots )\) the nested commutator determined by the tree T.

Here the first point of view on the tensor product (187) is more direct and does not require splitting \(m_n\) into pieces \(m_n^T\) possessing different symmetries. However, we wanted to also present the second point of view (189) since it compares directly to the tree part of (115) and explains how to construct the corresponding \(C_\infty \) algebra (via (188)).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cattaneo, A.S., Mnev, P. & Reshetikhin, N. A Cellular Topological Field Theory. Commun. Math. Phys. 374, 1229–1320 (2020). https://doi.org/10.1007/s00220-020-03687-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-020-03687-3

Navigation