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

A mathematical theory of topological invariants of quantum spin systems

Adam Artymowicz California Institute of Technology, Pasadena, CA 91125 Anton Kapustin California Institute of Technology, Pasadena, CA 91125 Bowen Yang Center of Mathematical Sciences and Applications, Harvard University, Cambridge, MA 02138
Abstract

We show that Hall conductance and its non-abelian and higher-dimensional analogs are obstructions to promoting a symmetry of a state to a gauge symmetry. To do this, we define a local Lie algebra over a Grothendieck site as a pre-cosheaf of Lie algebras with additional properties and propose that a gauge symmetry should be described by such an object. We show that infinitesimal symmetries of a gapped state of a quantum spin system form a local Lie algebra over a site of semilinear sets and use it to construct topological invariants of the state. Our construction applies to lattice systems on arbitrary asymptotically conical subsets of a Euclidean space including those which cannot be studied using field theory.

1 Introduction

The study of gapped phases of quantum matter at zero temperature is an important area of theoretical physics. Much conceptual progress has been made by assuming that gapped phases can be described by topological quantum field theory (TQFT). For example, the celebrated Quantum Hall Effect is captured by Chern-Simons field theory. However, the precise relation between gapped phases of matter and TQFTs is not understood. Recently, new mathematically rigorous approaches to classifying gapped phases of matter have been developed (see [1, 2, 3, 4] for the case of one-dimensional systems, [5, 6, 7, 8, 9, 10] for the case of two-dimensional systems, and [11] for systems in an arbitrary number of dimensions). They enable one to assign indices to gapped states of infinite-volume quantum systems invariant under symmetries. The main property of these indices, also referred to as topological invariants, is that they do not vary along suitably-defined continuous paths in the space of states. In some cases, the indices can be related to physical quantities, such as the zero-temperature Hall conductance, thereby explaining the robustness of the latter.

The methods of [7, 11, 8, 9] apply to arbitrary gapped states of infinite-volume quantum spin systems with rapidly decaying interactions and employ CC^{*}-algebraic techniques, some well-established and some relatively new. The construction of topological invariants in [11, 8] also uses some algebraic and geometric ingredients . The algebraic ingredient is a pointed (or curved) Differential Graded Lie Algebra (DGLA) and an associated Maurer-Cartan equation. The geometric ingredient is a collection of conical subsets of the Euclidean space triangulating the sphere at infinity. The appearance of these ingredients in the context of quantum statistical mechanics has not been motivated, and consequently the mathematical meaning of the invariants remains obscure.

The primary goal of this paper is provide a proper mathematical framework for the constructions of [11, 8] and to interpret topological invariants of gapped states as lattice analogs of ’t Hooft anomalies in Quantum Field Theory. The secondary goal is to generalize the construction in various directions. In particular, we show how to define topological invariants of lattice spin systems confined to well-behaved subsets of the lattice. This generalization makes explicit that the invariants take values in a vector space which is determined by the asymptotic geometry of the subset.

While our work concerns quantum lattice systems, we take inspiration from Quantum Field Theory (QFT). These two subjects are connected via the bulk-boundary correspondence. One aspect of this conjectural correspondence is that topological invariants of gapped states with symmetries are related to ’t Hooft anomalies of symmetries of the boundary field theory.111This assumes, of course, that a field-theoretic description of boundary degrees of freedom exists, which is far from obvious. It is usually said that ’t Hooft anomalies are obstructions to gauging a global symmetry of a QFT [12]. A possible mathematical interpretation of this statement is that an ’t Hooft anomaly is an obstruction to defining a local action of the group of gauge transformations on the algebra of local observables of a QFT. Assuming this interpretation, the presence of an ’t Hooft anomaly is a purely kinematic statement which involves neither the Hamiltonian nor the vacuum state of the field theory. It is not clear if conventional markers of ’t Hooft anomalies, such as anomalous Ward identities for vacuum correlators of currents, are implied by a kinematic statement. Proving or disproving this is currently out of reach because of gaps in the mathematical foundations of QFT. The mathematical theory of quantum lattice systems, on the other hand, is sufficiently mature and enables us to address the problem of ’t Hooft anomalies from the bulk side of the bulk-boundary correspondence. In this paper we show that topological invariants of gapped states of lattice systems, such as the zero-temperature Hall conductance, can be interpreted as obstructions to promoting a symmetry of a gapped state to a gauge symmetry. Dynamics enters this statement only though the state.

The main novelty of the paper is a new formulation of locality on a lattice. Building on ideas introduced in [7, 11, 8], we define for any (possibly unbounded) region of the lattice a space of derivations that are approximately localized on that region. For sufficiently regular regions, we show that these spaces behave as expected under natural operations like the commutator. To combine them into a single geometric object we show that they form a cosheaf on a certain site, i. e. a category with a Grothendieck topology. The utility of Grothendieck topologies in describing spaces of functions with a prescribed asymptotic behavior is well-known to analysts, see e.g. [13, 14], and here we apply the same idea in a non-commutative context. We call the resulting global geometric object a local Lie algebra. An on-site action of a Lie group on a lattice system yields a representation of a certain local Lie algebra by derivations, and we show that the invariants defined in [11] can be phrased as purely algebraic invariants of this representation. Namely, an obstruction exists to defining a representation which acts by state-preserving derivations, and this obstruction takes value in the homology of a certain DGLA associated naturally to the local Lie algebra of state-preserving derivations. The invariants defined in [11] then arise from a natural pairing of this homology with the Čech cohomology of the sphere at infinity. Aside from clarifying their mathematical meaning, this also shows that the invariants defined in [11] do not depend on certain choices present in their construction.

The content of the paper is as follows. In Section 2 we axiomatize the notion of an infinitesimal local symmetry by defining local Lie algebras abstractly in terms of cosheaves on a site. We also introduce the Čech functor which assigns a DGLAs to a local Lie algebra, and will serve as the main link between lattice geometry and algebraic topology. In Section 3 we turn to infinitesimal symmetries on the lattice, introducing the space of derivations approximately localized on a region, and prove the main properties of these spaces, some of which hold only for sufficiently regular regions. In section 4 we identify a suitable category of such regular regions in n{\mathbb{R}}^{n}, the category 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} of fuzzy semilinear sets. We identify a Grothendieck topology on 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} and show that infinitesimal symmetries of any gapped state ψ\psi of a quantum lattice system on n{\mathbb{R}}^{n} can be described by a local Lie algebra 𝔇alψ{\mathfrak{D}}^{\psi}_{al} over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. We also study gapped states invariant under an action of a compact Lie group GG and define a GG-equivariant analog of 𝔇alψ{\mathfrak{D}}^{\psi}_{al}. In Section 6 we construct invariants of GG-invariant gapped states. The construction is along the lines of [11, 8] and uses an inhomogeneous Maurer-Cartan equation. We show that these invariants are obstructions for promoting the symmetry GG to a local symmetry of the state ψ\psi. We also explain how to generalize the construction of invariants to lattice systems defined on sufficiently nice (asymptotically conical) subsets of n{\mathbb{R}}^{n} and show that their invariants take values in a space which depends on the asymptotic geometry of the subset. This goes beyond what one can access using the TQFT heuristics. In Appendix A we isolate some proofs necessary for Section 3, and Appendix B develops some properties of the inhomogeneous Maurer-Cartan equation.

We are grateful to Owen Gwilliam, Ezra Getzler, and Bas Janssens for discussions. The work of A. A. and A. K. was supported by the Simons Investigator Award. B. Y. would like to thank Yu-An Chen and Peking University for hospitality during the final stages of this work. A. K. would like to thank Yau Mathematical Sciences Center, Tsinghua University, for the same.

2 Local Lie algebras

2.1 Locality and (pre-)cosheaves

Let MM be a manifold and Open(M)Open(M) be the category whose objects are open subsets of MM, and the set of morphisms from an open UU to an open VV is the singleton or the empty set depending on whether UVU\subseteq V or UVU\nsubseteq V. Composition of morphisms is uniquely defined. A pre-cosheaf 𝔉{\mathfrak{F}} on MM with values in a category 𝒞{\mathcal{C}} is a functor 𝔉:Open(M)𝒞{\mathfrak{F}}:Open(M)\rightarrow{\mathcal{C}}. Thus for every inclusion of opens UVU\subseteq V one is given a co-restriction morphism eVU:𝔉(U)𝔉(V)e_{VU}:{\mathfrak{F}}(U)\rightarrow{\mathfrak{F}}(V) such that for any three opens UVWU\subseteq V\subseteq W one has eWVeVU=eWUe_{WV}\circ e_{VU}=e_{WU}. Pre-cosheaves (as well as pre-sheaves, which are functors from the opposite category of Open(M)Open(M) to 𝒞{\mathcal{C}}) can be used to describe local data on MM. This form of locality is rather weak, since it does not require 𝔉(UV){\mathfrak{F}}(U\cup V) to be expressible through 𝔉(U){\mathfrak{F}}(U) and 𝔉(V){\mathfrak{F}}(V).

As an example, consider the Lie algebra of gauge transformations, i.e. the Lie algebra 𝔊(M):=C(M,𝔤){\mathfrak{G}}(M):=C^{\infty}(M,{\mathfrak{g}}) of smooth functions on MM with values in a finite-dimensional Lie algebra 𝔤{\mathfrak{g}}. It is a global object attached to MM. To “localize” it, for any open UMU\subseteq M we define the Lie algebra 𝔊(U){\mathfrak{G}}(U) to be the space of smooth 𝔤{\mathfrak{g}}-valued functions on MM whose closed support is contained in UU. In particular, 𝔊()=0{\mathfrak{G}}(\varnothing)=0. For any inclusion of opens UVU\subseteq V we have a homomorphism of Lie algebras ιVU:𝔊(U)𝔊(V)\iota_{VU}:{\mathfrak{G}}(U)\rightarrow{\mathfrak{G}}(V) such that the Lie algebras 𝔊(U){\mathfrak{G}}(U) assemble into a pre-cosheaf 𝔊{\mathfrak{G}} of Lie algebras on MM. This is a coflasque pre-cosheaf, ie. all its structure maps ιVU\iota_{VU} are injective.222The terminology comes from sheaf theory, where a pre-sheaf :Open(M)opp𝒞{\mathcal{F}}:Open(M)^{opp}\rightarrow{\mathcal{C}} is called flasque if for any UVU\subseteq V the restriction morphism (V)(U){\mathcal{F}}(V)\rightarrow{\mathcal{F}}(U) is a surjection.

Continuing with the example, for any two opens U,VU,V the following sequence is exact:

𝔊(UV)𝔊(U)𝔊(V)𝔊(UV)0.\displaystyle{\mathfrak{G}}(U\cap V)\rightarrow{\mathfrak{G}}(U)\oplus{\mathfrak{G}}(V)\rightarrow{\mathfrak{G}}(U\cup V)\rightarrow 0. (1)

Here the first arrow is ιU,UV(ιV,UV)\iota_{U,U\cap V}\oplus(-\iota_{V,U\cap V}) and the second arrow is ιUV,UιUV,V\iota_{U\cup V,U}\oplus\iota_{U\cup V,V}. Exactness follows from the existence of a partition of unity for the cover 𝔘={U,V}{\mathfrak{U}}=\{U,V\} of UVU\cup V. In words, the exactness of the sequence (1) means that any element of 𝔊(UV){\mathfrak{G}}(U\cup V) can be decomposed as a sum of elements of sub-algebras attached to UU and VV modulo ambiguities which take values in the sub-algebra attached to UVU\cap V.

More generally, the existence of a partition of unity implies that for any collection of opens UiU_{i}, iIi\in I, the following sequence is exact:

i<j𝔊(UiUj)i𝔊(Ui)𝔊(iUi)0.\displaystyle\oplus_{i<j}{\mathfrak{G}}(U_{i}\cap U_{j})\rightarrow\oplus_{i}{\mathfrak{G}}(U_{i})\rightarrow{\mathfrak{G}}(\cup_{i}U_{i})\rightarrow 0. (2)

By definition, this means that the pre-cosheaf of vector spaces 𝔊{\mathfrak{G}} is a cosheaf of vector spaces. The cosheaf property is a compatibility of the pre-cosheaf with the notions of intersection and union of opens and expresses a stronger form of locality.

Note that the maps in the above exact sequences are not Lie algebra homomorphisms. Hence 𝔊{\mathfrak{G}} is not a cosheaf of Lie algebras. Nevertheless, the following additional property of 𝔊{\mathfrak{G}} can be regarded as a form of locality of the Lie bracket: for any U,V,U,V, [𝔊(U),𝔊(V)]𝔊(UV)[{\mathfrak{G}}(U),{\mathfrak{G}}(V)]\subseteq{\mathfrak{G}}(U\cap V). We will call this ”Property I”, where ”I” stands for ”intersection”. In particular, elements of 𝔊(M){\mathfrak{G}}(M) which are supported on non-intersecting opens UU and VV commute.

Symmetries of gapped states of quantum lattice spin systems are typically local only approximately. To phrase locality of symmetries in lattice systems in a similar language, one needs to replace the set Open(M)Open(M) of open subsets of MM with a more general structure which admits the notions of intersection, union, and cover.

The first thing to note is that the category Open(M)Open(M) is rather special: its objects form a pre-ordered set (i.e. the set of objects carries a relation \subseteq which is reflexive and transitive), and the category structure is determined by the pre-order. The relation \subseteq is also anti-symmetric: UVU\subseteq V and VUV\subseteq U implies U=VU=V. In other words, Open(M)Open(M) is a poset. In general, we will not require the pre-order to be anti-symmetric. From the categorical viewpoint, UVU\subseteq V and VUV\subseteq U means that UU and VV are isomorphic objects of the category Open(M)Open(M), and as a general rule, it is not advisable to identify isomorphic objects.

For any pre-ordered set (X,)(X,\leq) there is a natural notion of intersection and union. The intersection of U,VXU,V\in X can be defined as the greatest lower bound (or meet) of both UU and VV, i.e. a WXW\in X such that WUW\leq U, WVW\leq V, and for any WU,VW^{\prime}\leq U,V we have WWW^{\prime}\leq W. The meet of UU and VV is denoted UVU\wedge V. Similarly, the union of UU and VV can be defined as the smallest upper bound (or join) of both UU and VV. It is denoted UVU\vee V. For a general pre-ordered set, the meet and join may not exist for all pairs of objects. If they exist, they are unique up to isomorphism. We will assume that (X,)(X,\leq) is such that UVU\wedge V and UVU\vee V exist for all U,VXU,V\in X.333If we turn XX into a poset by identifying isomorphic objects, then this means that the poset is a lattice in the sense of order theory. The existence of all pairwise meets and joins implies the existence of all finite meets and joins. In the case of the pre-ordered set Open(M)Open(M) arbitrary (i.e. not necessarily finite) joins make sense.

Finally, to define covers of elements of XX, let us assume that U(VW)(UV)(UW)U\wedge(V\vee W)\leq(U\wedge V)\vee(U\wedge W) for all U,V,WXU,V,W\in X.444The opposite relation is automatic, so this condition ensures that U(VW)(UV)(UW)U\wedge(V\vee W)\simeq(U\wedge V)\vee(U\wedge W) for all U,V,WXU,V,W\in X. This is equivalent to saying that the poset corresponding to XX is a distributive lattice. We will say that XX is a distributive pre-ordered set. This condition is certainly satisfied for Open(M)Open(M). We say that a collection 𝔘={Ui}iI{\mathfrak{U}}=\{U_{i}\}_{i\in I} of elements of XX covers AXA\in X iff UiAU_{i}\leq A for all iIi\in I and AiIUiA\leq\bigvee_{i\in I}U_{i}. This definition ensures that if 𝔘{\mathfrak{U}} covers AA, then for any BAB\leq A the collection 𝔘B={UiB}iI{\mathfrak{U}}\wedge B=\{U_{i}\wedge B\}_{i\in I} covers BB.

In the case of the pre-ordered set Open(M)Open(M), the standard topological definition of a cover allows II to be infinite. In general, if XX admits only finite joins, II needs to be finite. Also, we may or may not allow II to be empty. This possibility only arises when AA is the smallest element of XX, i.e. AUA\leq U for any UXU\in X. In the case of the pre-ordered set Open(M)Open(M), the smallest element is the empty set \varnothing, and the standard choice is to allow the empty cover of the empty set. For a general pre-ordered set, it is up to us whether to allow II to be empty.

From a categorical perspective, this notion of a cover equips any distributive pre-ordered set (X,)(X,\leq) admitting pairwise meets and joins with a Grothendieck topology, thus making it into a site [15]. Apart from the option of allowing the labeling set II to be empty, this Grothendieck topology is canonical.

For any WXW\in X we may consider the subset XW={UXUW}X^{W}=\left\{U\in X\mid U\leq W\right\} with the pre-order inherited from (X,)(X,\leq). It is a distributive pre-ordered set in its own right. When equipped with its canonical Grothendieck topology, it can be regarded as a sub-site of the site associated to (X,)(X,\leq).

Given any distributive pre-ordered set (X,)(X,\leq), we can define the notion of a pre-cosheaf of vector spaces, a pre-cosheaf of Lie algebras, a cosheaf of vector spaces, and a coflasque pre-cosheaf of Lie algebras with Property I exactly as before, i.e. by mechanically replacing \cup with \vee and \cap with \wedge. Motivated by the above example, we introduce the following definition.

Definition 2.1.

A local Lie algebra over (X,)(X,\leq) is a coflasque pre-cosheaf of Lie algebras with Property I which is also a cosheaf of vector spaces over the corresponding site. A morphism of local Lie algebras is a morphism of the underlying pre-cosheaves of Lie algebras.

Remark 2.1.

Let 𝔉{\mathfrak{F}} be a local Lie algebra. Then for any UVU\leq V the Lie algebra 𝔉(U){\mathfrak{F}}(U) is an ideal in 𝔉(V){\mathfrak{F}}(V). One can equivalently define a local Lie algebra as a pre-cosheaf of Lie algebras over (X,)(X,\leq) which is a cosheaf of vector spaces and such that all co-restriction maps are inclusions of Lie ideals.

Remark 2.2.

In this paper all Lie algebras will be Fréchet-Lie algebras and all morphisms will be continuous. We define a local Fréchet-Lie algebra over (X,)(X,\leq) to be a coflasque pre-cosheaf of Fréchet-Lie algebras with Property I which is also a cosheaf of vector spaces.

The following Lemma will be useful in future sections to check the cosheaf property:

Lemma 2.1.

Let 𝔉{\mathfrak{F}} be a pre-cosheaf of vector spaces on a distributive lattice XX, with co-restriction maps ιU,V:UV\iota_{U,V}:U\to V. Then 𝔉{\mathfrak{F}} is a cosheaf on XX (with the topology of finite covers) iff for any U,VXU,V\in X the following sequence is exact

𝔉(UV)𝛼𝔉(U)𝔉(V)𝛽𝔉(UV)0,\displaystyle{\mathfrak{F}}(U\wedge V)\xrightarrow{\alpha}{\mathfrak{F}}(U)\oplus{\mathfrak{F}}(V)\xrightarrow{\beta}{\mathfrak{F}}(U\vee V)\to 0, (3)

where α=ιUV,UιUV,V\alpha=\iota_{U\wedge V,U}-\iota_{U\wedge V,V} and β=ιU,UV+ιV,UV\beta=\iota_{U,U\vee V}+\iota_{V,U\vee V}.

The following proof is essentially a restatement of the proof of Proposition 1.3 in [16], adapted to the site LL:

Proof.

We must show that for any UXU\in X and every covering U1Un=UU_{1}\vee\ldots\vee U_{n}=U, the sequence of vector spaces

i<j𝔉(UiUj)𝛼i𝔉(Ui)𝛽𝔉(iUi)0\displaystyle\bigoplus_{i<j}{\mathfrak{F}}(U_{i}\wedge U_{j})\xrightarrow{\alpha}\bigoplus_{i}{\mathfrak{F}}(U_{i})\xrightarrow{\beta}{\mathfrak{F}}(\vee_{i}U_{i})\to 0 (4)

is exact, where α=i<jιUiUj,UiιUiUj,Uj\alpha=\sum_{i<j}\iota_{U_{i}\wedge U_{j},U_{i}}-\iota_{U_{i}\wedge U_{j},U_{j}} and β=iιUi,U\beta=\sum_{i}\iota_{U_{i},U}. Exactness of the last three terms follows from an easy induction and the associativity of the join operation. For exactness of the first three terms of the sequence, we also proceed by induction: suppose the result holds for all covers of cardinality n1n-1, and let U1,,UnLU_{1},\ldots,U_{n}\in L be a cover of UU. Suppose (s1,,sn)i=1n𝔉(Ui)(s_{1},\ldots,s_{n})\in\bigoplus_{i=1}^{n}{\mathfrak{F}}(U_{i}) lies in the kernel of β\beta. Let V:=i=1n1UiV:=\bigvee_{i=1}^{n-1}U_{i}. An application of (3) with UnU_{n} and VV shows that sn=ιUnV,U(t)s_{n}=\iota_{U_{n}\cap V,U}(t) for some tVUn=(U1Un)(Un1Un)t\in V\wedge U_{n}=(U_{1}\wedge U_{n})\vee...\vee(U_{n-1}\vee U_{n}), and by right-exactness of (4) this shows that sn=i=1n1ιUiUn,Un(vi)s_{n}=\sum_{i=1}^{n-1}\iota_{U_{i}\cap U_{n},U_{n}}(v_{i}) for some vi𝔉(Ui)v_{i}\in{\mathfrak{F}}(U_{i}). Finally we write

(s1,,sn)=(s1+w1,,sn1+wn1,0)+(w1,,wn1,sn)\displaystyle(s_{1},\ldots,s_{n})=(s_{1}+w_{1},\ldots,s_{n-1}+w_{n-1},0)+(-w_{1},\ldots,-w_{n-1},s_{n})

where wi:=ιUiUn,Un(vi)w_{i}:=\iota_{U_{i}\cap U_{n},U_{n}}(v_{i}). Both terms above the first term lies in the image of α\alpha by the inductive hypothesis, while the second term equals i=1n1ιUiUn,Un(vi)ιUiUn,Ui(vi)\sum_{i=1}^{n-1}\iota_{U_{i}\wedge U_{n},U_{n}}(v_{i})-\iota_{U_{i}\wedge U_{n},U_{i}}(v_{i}), which evidently is also in the image of α\alpha. ∎

2.2 DGLA attached to a cover

Let XX be a distributive pre-ordered set. Let WXW\in X and 𝔘={Ui}iI{\mathfrak{U}}=\{U_{i}\}_{i\in I} be a cover of WW. Let 𝔉{\mathfrak{F}} be a pre-cosheaf of vector spaces over XX. The Čech chain complex C(𝔘,W;𝔉)C_{\bullet}({\mathfrak{U}},W;{\mathfrak{F}}) is defined by

Cn(𝔘,W;𝔉)=i0<<in𝔉(Ui0Uin),n0.C_{n}({\mathfrak{U}},W;{\mathfrak{F}})=\oplus_{i_{0}<\ldots<i_{n}}{\mathfrak{F}}(U_{i_{0}}\wedge\ldots\wedge U_{i_{n}}),\quad n\geq 0.

(where some linear order on II has been chosen) with the differential is the Čech differential :Cn+1Cn\partial:C_{n+1}\rightarrow C_{n} given by =j=0n(1)jλj\partial=\sum_{j=0}^{n}(-1)^{j}\lambda_{j}, where λj\lambda_{j} is the canonical map

𝔉(Ui0Uin)𝔉(Ui0Uij^Uin).\displaystyle{\mathfrak{F}}(U_{i_{0}}\wedge\ldots\wedge U_{i_{n}})\to{\mathfrak{F}}(U_{i_{0}}\wedge\ldots\widehat{U_{i_{j}}}\ldots\wedge U_{i_{n}}).
Lemma 2.2.

If 𝔉{\mathfrak{F}} is a coflasque cosheaf of vector spaces over XX, the homology of the Čech complex is 0 for n>0n>0 and 𝔉(W){\mathfrak{F}}(W) for n=0n=0.

Proof.

See [16], Corollary 4.3. Note that [16] uses the term “flabby” instead of “coflasque”. ∎

Let Caug(𝔘,W;𝔉)={C(𝔘,W;𝔉)𝔉(W)}C^{aug}({\mathfrak{U}},W;{\mathfrak{F}})=\{C_{\bullet}({\mathfrak{U}},W;{\mathfrak{F}})\rightarrow{\mathfrak{F}}(W)\} be the augmented Čech complex.

Proposition 2.1.

Let 𝔉{\mathfrak{F}} be a local Lie algebra over XX. Then for any WXW\in X and any cover 𝔘{\mathfrak{U}} of WW the 1-shifted augmented Čech complex C+1aug(𝔘,W;𝔉)C^{aug}_{\bullet+1}({\mathfrak{U}},W;{\mathfrak{F}}) has a natural structure of a non-negatively graded acyclic DGLA.

Proof.

Let 𝔘{\mathfrak{U}} be a cover of WW indexed by II. Let VV be a vector space with a basis eie_{i}, iIi\in I and let fi,iIf^{i},i\in I, be the dual basis of VV^{*}. The 1-shifted augmented Čech complex of 𝔉{\mathfrak{F}} with respect to 𝔘{\mathfrak{U}} is naturally identified with a sub-complex of the DGLA (𝔉(W)ΛV,)({\mathfrak{F}}(W)\otimes\Lambda^{\bullet}V,\partial), where \partial is contraction with ifi\sum_{i}f^{i}. It is easy to check that this sub-complex is closed with respect to the graded Lie bracket thanks to Property I. By Lemma 2.2, the resulting DGLA is acyclic. ∎

Covers of WXW\in X form a category whose morphisms are refinements. A refinement of a cover 𝔘={Ui}iI{\mathfrak{U}}=\left\{U_{i}\right\}_{i\in I} to a cover 𝔙={Vj}jJ{\mathfrak{V}}=\left\{V_{j}\right\}_{j\in J} is a map ϕ:JI\phi:J\rightarrow I such that VjUϕ(j)V_{j}\leq U_{\phi(j)}. Refinements are composed in an obvious way.

Proposition 2.2.

For a fixed WXW\in X, the map which sends a local Lie algebra 𝔉{\mathfrak{F}} and a cover 𝔘{\mathfrak{U}} of WW to the DGLA C+1aug(𝔘,W;𝔉)C^{aug}_{\bullet+1}({\mathfrak{U}},W;{\mathfrak{F}}) is functorial in both 𝔉{\mathfrak{F}} and 𝔘{\mathfrak{U}}.

Proof.

Functoriality in 𝔉{\mathfrak{F}} is clear. Functoriality in 𝔘{\mathfrak{U}} is written Čech component-wise for 𝗉𝔙Ck(𝔙,W;𝔉){\mathsf{p}}^{\mathfrak{V}}\in C_{k}({\mathfrak{V}},W;{\mathfrak{F}}):

(ϕ𝗉k𝔙)i0,,ik:=j0ϕ1(i0)jkϕ1(ik)(𝗉k𝔙)j0,,jk,(\phi_{*}{\mathsf{p}}^{\mathfrak{V}}_{k})_{i_{0},\dots,i_{k}}:=\sum_{j_{0}\in\phi^{-1}(i_{0})}\cdots\sum_{j_{k}\in\phi^{-1}(i_{k})}({\mathsf{p}}^{\mathfrak{V}}_{k})_{j_{0},\dots,j_{k}},

for i0<i2<<iki_{0}<i_{2}<\cdots<i_{k}. ∎

We will call C+1aug(,W;)C^{aug}_{\bullet+1}(-,W;-) the Čech functor. We will need some variants of the Čech functor. First, we can define a graded local Lie algebra over a distributive pre-ordered set (X,)(X,\leq) in an obvious manner. The construction of the acyclic DGLA C+1aug(𝔘,W;𝔉)C^{aug}_{\bullet+1}({\mathfrak{U}},W;{\mathfrak{F}}) works in this case as well, except that it may have components in negative degrees.

Second, we define a pointed DGLA as a DGLA equipped with a distinguished central cycle of degree 2-2 (which we call the curvature). A morphism of pointed DGLAs is a DGLA morphism which preserves the distinguished central cycle.555The category of pointed DGLAs as defined here is a full subcategory of the category of curved DGLAs as defined in [17]. There, for a curved DGLA with curvature 𝖡{\mathsf{B}}, 𝖡{\mathsf{B}} is not required to be central and the derivation \partial satisfies 2=ad𝖡\partial^{2}={\rm ad}_{\mathsf{B}}. We say that a graded local Lie algebra 𝔉{\mathfrak{F}} over (X,)(X,\leq) with a terminal object TXT\in X is pointed if it is equipped with a distinguished central element 𝖡𝔉(T){\mathsf{B}}\in{\mathfrak{F}}(T) of degree 2-2. Morphisms in the category of pointed graded local Lie algebras are required to preserve the distinguished element. Then the DGLA C+1aug(𝔘,T;𝔉)C^{aug}_{\bullet+1}({\mathfrak{U}},T;{\mathfrak{F}}) is an acyclic pointed DGLA, the distinguished central cycle being 𝖡C1aug(𝔘,T;𝔉){\mathsf{B}}\in C^{aug}_{-1}({\mathfrak{U}},T;{\mathfrak{F}}). Of course, since the DGLA is acyclic, this central cycle is exact. The construction of a pointed DGLA from a pointed graded local Lie algebra and a cover of TT is functorial in both arguments.

3 Quantum lattice systems

3.1 Observables and derivations

We use the \ell^{\infty} metric on n{\mathbb{R}}^{n}, i. e. d(x,y):=maxi=1,,n|xiyi|d(x,y):=\max_{i=1,\ldots,n}|x_{i}-y_{i}|. For any U,VnU,V\subset{\mathbb{R}}^{n} we write diam(U):=supx,yUd(x,y)\operatorname{diam}(U):=\sup_{x,y\in U}d(x,y) and d(U,V):=infxU,yVd(x,y)d(U,V):=\inf_{x\in U,y\in V}d(x,y). They take values in extended non-negative reals [0,][0,\infty]. Thus diam()=0\operatorname{diam}(\varnothing)=0 and d(U,)=d(U,\varnothing)=\infty for any UnU\subset{\mathbb{R}}^{n}. For a nonempty set UU and r0r\geq 0 we define Ur:={xn:d(x,U)r}U^{r}:=\{x\in{\mathbb{R}}^{n}:d(x,U)\leq r\} while we set r=\varnothing^{r}=\varnothing.

A quantum lattice system consists of a countable subset Λn\Lambda\subset{\mathbb{R}}^{n} (“the lattice”) and a finite-dimensional complex Hilbert space VjV_{j} for every jΛj\in\Lambda. We make the following assumption on the lattice system666In [11] Λ\Lambda was assumed to be a Delone set, i.e. it was required to be uniformly filling and uniformly discrete. These assumptions were imposed on physical grounds. All the results proved in [11] hold under weaker assumptions adopted in this paper. In [8] Λ\Lambda was taken to be n{\mathbb{Z}}^{n} for simplicity. : there is a CΛ>0C_{\Lambda}>0 such that the number of points of Λ\Lambda in any hypercube of diameter dd is bounded by CΛ(d+1)nC_{\Lambda}(d+1)^{n}.

For any bounded nonempty XnX\subset{\mathbb{R}}^{n} let 𝒜(X):=jXΛHom(Vj,Vj){\mathscr{A}}(X):=\bigotimes_{j\in X\cap\Lambda}\mathrm{Hom}_{{\mathbb{C}}}(V_{j},V_{j}). For any XYX\subset Y there is an inclusion 𝒜(X)𝒜(Y){\mathscr{A}}(X)\hookrightarrow{\mathscr{A}}(Y) and the algebras 𝒜(X){\mathscr{A}}(X) form a direct system with respect to these inclusions. We extend this direct system to include the empty set by setting 𝒜()={\mathscr{A}}(\varnothing)={\mathbb{C}} and letting the inclusion 𝒜()𝒜(X){\mathscr{A}}(\varnothing)\hookrightarrow{\mathscr{A}}(X) take αα𝟏\alpha\mapsto\alpha\boldsymbol{1}. Each 𝒜(X){\mathscr{A}}(X) is a finite-dimensional CC^{*}-algebra with the operator norm, and the inclusions 𝒜(X)𝒜(Y){\mathscr{A}}(X)\hookrightarrow{\mathscr{A}}(Y) preserve this norm. The normed *-algebra of local observables is

𝒜=limX𝒜(X).\displaystyle{\mathscr{A}}_{\ell}=\varinjlim_{X}{\mathscr{A}}(X).

The algebra of quasi-local observables 𝒜{\mathscr{A}} is the norm-completion of 𝒜{\mathscr{A}}_{\ell}; it is a CC^{*}-algebra.

For any bounded XnX\subset{\mathbb{R}}^{n}, define the normalized trace tr¯:𝒜(X){\overline{\rm tr}}:{\mathscr{A}}(X)\to{\mathbb{C}} as tr¯(𝒜)=tr(𝒜)/dim(𝒜(X)){\overline{\rm tr}}({\mathcal{A}})=\operatorname{tr}({\mathcal{A}})/{\sqrt{\dim({\mathscr{A}}(X))}}. For any bounded XYX\subset Y the partial trace tr¯Xc:𝒜(Y)𝒜(X){\overline{\rm tr}}_{X^{c}}:{\mathscr{A}}(Y)\to{\mathscr{A}}(X) is uniquely specified by the condition tr¯Xc(𝒜)=tr¯(𝒜){\overline{\rm tr}}_{X^{c}}({\mathcal{A}}\otimes{\mathcal{B}})={\overline{\rm tr}}({\mathcal{A}}){\mathcal{B}} for any 𝒜𝒜(Y\X){\mathcal{A}}\in{\mathscr{A}}(Y\backslash X) and 𝒜(X){\mathcal{B}}\in{\mathscr{A}}(X). Besides forming a direct system with respect to inclusions, the spaces 𝒜(X){\mathscr{A}}(X) are also an inverse system with respect to the partial trace. tr¯{\overline{\rm tr}} extends to a normalized positive linear functional on 𝒜{\mathscr{A}}, i.e. a state. We say that 𝒜𝒜{\mathcal{A}}\in{\mathscr{A}} is traceless if tr¯(𝒜)=0{\overline{\rm tr}}({\mathcal{A}})=0. The space of traceless anti-hermitian elements of 𝒜(X){\mathscr{A}}(X) will be denoted 𝔡l(X){{\mathfrak{d}}_{l}}(X). 𝔡l(X){{\mathfrak{d}}_{l}}(X) is a real Lie algebra with respect to the commutator. The Lie algebras 𝔡l(X){{\mathfrak{d}}_{l}}(X) form a direct system over the directed set of bounded subsets of n{\mathbb{R}}^{n}, and its limit will be denoted 𝔡l{{\mathfrak{d}}_{l}}. Equivalently, 𝔡l{{\mathfrak{d}}_{l}} is the Lie algebra of traceless anti-hermitian elements of 𝒜{\mathscr{A}}_{\ell}. Note that 𝒜=𝟏(𝔡l){\mathscr{A}}_{\ell}={\mathbb{C}}{\mathbf{1}}\oplus({{\mathfrak{d}}_{l}}\otimes{\mathbb{C}}).

Definition 3.1.

A brick in n{\mathbb{R}}^{n} is a non-empty subset of the form

Y={(x1,,xn)|iximi,i=1,,n}\displaystyle Y=\{(x_{1},\ldots,x_{n})\ |\ \ell_{i}\leq x_{i}\leq m_{i},i=1,\ldots,n\} (5)

where (k1,,kn)(k_{1},\ldots,k_{n}), (1,,n)(\ell_{1},\ldots,\ell_{n}), and (m1,,mn)(m_{1},\ldots,m_{n}) are nn-tuples of integers. We write 𝔹n{\mathbb{B}}_{n} for the set of all bricks in n{\mathbb{R}}^{n}.

The set of bricks exhausts the collection of bounded subsets of n{\mathbb{R}}^{n} in the sense that any bounded subset is contained in a brick. In addition, the set of bricks satisfies the following regularity property:

Lemma 3.1.

For any jnj\in{\mathbb{R}}^{n} we have

Y𝔹n(1+diam(Y)+d(Y,j))2n2\displaystyle\sum_{Y\in{\mathbb{B}}_{n}}(1+\operatorname{diam}(Y)+d(Y,j))^{-2n-2} π44n(n+1)236\displaystyle\leq\frac{\pi^{4}4^{n}(n+1)^{2}}{36}
Proof.

Any pair of points x,ynx,y\in{\mathbb{Z}}^{n} specifies a brick with xx and yy on opposing corners, and any brick can be specified this way (not uniquely). With XX the brick corresponding to xx and yy it is easy to see that max(d(x,j),d(y,j))diam(X)+d(X,j)\max(d(x,j),d(y,j))\leq\operatorname{diam}(X)+d(X,j), and so (1+d(x,j))(1+d(y,j))(1+diam(X)+d(X,j))2(1+d(x,j))(1+d(y,j))\leq(1+\operatorname{diam}(X)+d(X,j))^{2}. Thus we have

Y𝔹n(1+diam(Y)+d(Y,j))2n2\displaystyle\sum_{Y\in{\mathbb{B}}_{n}}(1+\operatorname{diam}(Y)+d(Y,j))^{-2n-2} x,yn(1+d(x,j))n1(1+d(y,j))n1\displaystyle\leq\sum_{\begin{subarray}{c}x,y\in{\mathbb{Z}}^{n}\end{subarray}}(1+d(x,j))^{-n-1}(1+d(y,j))^{-n-1}
=(xn(1+d(x,j))n1)2,\displaystyle=\left(\sum_{x\in{\mathbb{Z}}^{n}}(1+d(x,j))^{-n-1}\right)^{2},

and it remains only to bound the above sum. Let f(k):=(1+k)n1f(k):=(1+k)^{-n-1} and g(k):=#(nBk(j))(1+2k)ng(k):=\#({\mathbb{Z}}^{n}\cap B_{k}(j))\leq(1+2k)^{n}. Using summation by parts we have

jΛ(1+d(x,j))n1\displaystyle\sum_{j\in\Lambda}(1+d(x,j))^{-n-1} k0f(k)(g(k+1)g(k))\displaystyle\leq\sum_{k\geq 0}f(k)(g(k+1)-g(k))
=limkf(k)g(k)k0g(k)(f(k+1)f(k)).\displaystyle=\lim_{k\to\infty}f(k)g(k)-\sum_{k\geq 0}g(k)(f(k+1)-f(k)).

It is easy to check that f(k)g(k)0f(k)g(k)\to 0 and that (f(k+1)f(k))(n+1)(1+k)n2-(f(k+1)-f(k))\leq(n+1)(1+k)^{-n-2}, and so

jΛ(1+d(x,j))n1\displaystyle\sum_{j\in\Lambda}(1+d(x,j))^{-n-1} 2n(n+1)k0(1+k)2\displaystyle\leq 2^{n}(n+1)\sum_{k\geq 0}(1+k)^{-2}
π22n(n+1)6\displaystyle\leq\frac{\pi^{2}2^{n}(n+1)}{6}

which proves the Lemma. ∎

For any brick YY we define the following subspace of 𝔡l(Y){{\mathfrak{d}}_{l}}(Y):

𝔡lY:={𝒜𝔡l(Y)|tr¯Xc(𝒜)=0 for any brick XY}.\displaystyle{{\mathfrak{d}}_{l}}^{Y}:=\{{\mathcal{A}}\in{{\mathfrak{d}}_{l}}(Y)\ |\ {\overline{\rm tr}}_{X^{c}}({\mathcal{A}})=0\text{ for any brick }X\subsetneq Y\}.

Each 𝔡l(Y){{\mathfrak{d}}_{l}}(Y) decomposes as a direct sum 𝔡l(Y)=XY𝔡lX{{\mathfrak{d}}_{l}}(Y)=\bigoplus_{X\subseteq Y}{{\mathfrak{d}}_{l}}^{X} over bricks contained in YY, and for any brick XYX\subseteq Y the partial trace tr¯Y\Xc{\overline{\rm tr}}_{Y\backslash X^{c}} is the projection onto ZX𝔡lZ𝔡l(Y)\bigoplus_{Z\subseteq X}{{\mathfrak{d}}_{l}}^{Z}\subseteq{{\mathfrak{d}}_{l}}(Y). Intuitively, 𝔡lY{{\mathfrak{d}}_{l}}^{Y} consists of elements of 𝔡l(Y){{\mathfrak{d}}_{l}}(Y) which are not localized on any brick properly contained in YY.

Derivations of 𝒜{\mathscr{A}} which appear in the physical context are typically only densely defined and have the form

𝖥:𝒜Y𝔹n[𝖥Y,𝒜],\displaystyle{\mathsf{F}}:{\mathcal{A}}\mapsto\sum_{Y\in{\mathbb{B}}_{n}}[{\mathsf{F}}^{Y},{\mathcal{A}}],

where 𝖥Y𝔡lY{\mathsf{F}}^{Y}\in{{\mathfrak{d}}_{l}}^{Y}. We are now going to define for every UnU\subset{\mathbb{R}}^{n} a real Lie algebra 𝔇al(U){\mathfrak{D}}_{al}(U) which consists of derivations approximately localized on UU and such that all 𝔇al(U){\mathfrak{D}}_{al}(U) have a common dense domain. Moreover, they form a pre-cosheaf of Lie algebras over a certain category of subsets of n{\mathbb{R}}^{n}.

Definition 3.2.

For any element 𝖥={𝖥Y}Y𝔹n{\mathsf{F}}=\{{\mathsf{F}}^{Y}\}_{Y\in{\mathbb{B}}_{n}} of Y𝔹n𝔡lY\prod_{Y\in{\mathbb{B}}_{n}}{{\mathfrak{d}}_{l}}^{Y} and every UnU\subset{\mathbb{R}}^{n} we let

𝖥U,k:=supY𝔹n𝖥Y(1+diam(Y)+d(U,Y))k\displaystyle\|{\mathsf{F}}\|_{U,k}:=\sup_{Y\in{\mathbb{B}}_{n}}\|{\mathsf{F}}^{Y}\|(1+\operatorname{diam}(Y)+d(U,Y))^{k} (6)

and define 𝔇al(U)Y𝔹n𝔡lY{\mathfrak{D}}_{al}(U)\subset\prod_{Y\in{\mathbb{B}}_{n}}{{\mathfrak{d}}_{l}}^{Y} as the set of elements 𝖥{\mathsf{F}} with 𝖥U,k<\|{\mathsf{F}}\|_{U,k}<\infty for all k0k\geq 0.

If UU is empty, then for k>0k>0 𝖥U,k<\|{\mathsf{F}}\|_{U,k}<\infty if and only if 𝖥Y=0{\mathsf{F}}^{Y}=0 for all Y𝔹nY\in{\mathbb{B}}_{n}. Thus 𝔇al()=0{\mathfrak{D}}_{al}(\emptyset)=0. We also denote 𝔇al(n)=𝔇al{\mathfrak{D}}_{al}({\mathbb{R}}^{n})={\mathfrak{D}}_{al}.

It is easy to see that (6) is a norm on 𝔇al(U){\mathfrak{D}}_{al}(U) for each k0k\geq 0. We endow 𝔇al(U){\mathfrak{D}}_{al}(U) with the locally convex topology given by the norms (6) ranging over all k0k\geq 0. Recall that a topological vector space is called a Fréchet space if it is Hausdorff, and if its topology can be generated by a countable family of seminorms with respect to which it is complete.

Proposition 3.1.

𝔇al(U){\mathfrak{D}}_{al}(U) is a Fréchet space.

Proof.

The Hausdorff property follows from the fact that if 𝖥U,k=0\|{\mathsf{F}}\|_{U,k}=0 for any k0k\geq 0 then 𝖥=0{\mathsf{F}}=0. To show completeness, suppose {𝖥m}m𝔇al(U)\{{\mathsf{F}}_{m}\}_{m\in\mathbb{N}}\subset{\mathfrak{D}}_{al}(U) is Cauchy, i.e. that for any k0k\geq 0 and any ϵ>0\epsilon>0 there is an NN\in\mathbb{N} such that m,mN𝖥m𝖥mU,k<ϵm,m^{\prime}\geq N\implies\|{\mathsf{F}}_{m}-{\mathsf{F}}_{m^{\prime}}\|_{U,k}<\epsilon. For any fixed Y𝔹nY\in{\mathbb{B}}_{n} this implies that {𝖥nY}\{{\mathsf{F}}_{n}^{Y}\} is Cauchy in 𝔡lY{{\mathfrak{d}}_{l}}^{Y} (with the operator norm) and thus converges to a limit 𝖥Y{\mathsf{F}}^{Y}. Let 𝖥:={𝖥Y}Y𝔹n{\mathsf{F}}:=\{{\mathsf{F}}^{Y}\}_{Y\in{\mathbb{B}}_{n}}.

Fix kk\in\mathbb{N} and ϵ>0\epsilon>0. For every =0,1,2,\ell=0,1,2,..., choose NN_{\ell}\in\mathbb{N} so that m,mN𝖥m𝖥mU,k<21ϵm,m^{\prime}\geq N_{\ell}\implies\|{\mathsf{F}}_{m}-{\mathsf{F}}_{m^{\prime}}\|_{U,k}<2^{-\ell-1}\epsilon. For any Y𝔹nY\in{\mathbb{B}}_{n}, any mN1m\geq N_{1}, and any M>1M>1, we have

𝖥mY𝖥Y\displaystyle\|{\mathsf{F}}_{m}^{Y}-{\mathsf{F}}^{Y}\| 𝖥mY𝖥N1Y+i=1M1𝖥NiY𝖥Ni+1Y+𝖥NMY𝖥Y\displaystyle\leq\|{\mathsf{F}}_{m}^{Y}-{\mathsf{F}}_{N_{1}}^{Y}\|+\sum_{i=1}^{M-1}\|{\mathsf{F}}_{N_{i}}^{Y}-{\mathsf{F}}_{N_{i+1}}^{Y}\|+\|{\mathsf{F}}_{N_{M}}^{Y}-{\mathsf{F}}^{Y}\|
ϵ(1+diam(Y)+d(U,Y))k+𝖥NMY𝖥Y.\displaystyle\leq\epsilon(1+\operatorname{diam}(Y)+d(U,Y))^{-k}+\|{\mathsf{F}}_{N_{M}}^{Y}-{\mathsf{F}}^{Y}\|.

Taking MM\to\infty shows that 𝖥mY𝖥Y(1+diam(Y)+d(U,Y))k<ϵ\|{\mathsf{F}}_{m}^{Y}-{\mathsf{F}}^{Y}\|(1+\operatorname{diam}(Y)+d(U,Y))^{k}<\epsilon. Since YY was arbitrary, we have 𝖥m𝖥U,k<ϵ\|{\mathsf{F}}_{m}-{\mathsf{F}}\|_{U,k}<\epsilon. Since kk was arbitrary, {𝖥m}\{{\mathsf{F}}_{m}\} converges to 𝖥{\mathsf{F}} in the topology of 𝔇al(U){\mathfrak{D}}_{al}(U). ∎

Recall that for a nonempty UnU\subset{\mathbb{R}}^{n} we write Ur:={xn:d(x,U)r}U^{r}:=\{x\in{\mathbb{R}}^{n}:d(x,U)\leq r\}. The norms U,k\|\cdot\|_{U,k} obey the following dominance relation.

Lemma 3.2.

Let U,VU,V be subsets of the lattice and suppose that UVrU\subseteq V^{r}. Then for any 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U) we have

𝖥V,k(r+1)k𝖥U,k.\displaystyle\|{\mathsf{F}}\|_{V,k}\leq(r+1)^{k}\|{\mathsf{F}}\|_{U,k}. (7)

In particular, 𝔇al(U)𝔇al(V){\mathfrak{D}}_{al}(U)\subseteq{\mathfrak{D}}_{al}(V) and the inclusion is continuous.

Proof.

Let YY be any subset of the lattice. Then (7) follows from

1+diam(Y)+d(Y,V)\displaystyle 1+\operatorname{diam}(Y)+d(Y,V) 1+diam(Y)+d(Y,Vr)+r\displaystyle\leq 1+\operatorname{diam}(Y)+d(Y,V^{r})+r
1+diam(Y)+d(Y,U)+r\displaystyle\leq 1+\operatorname{diam}(Y)+d(Y,U)+r
(r+1)(1+diam(Y)+d(Y,U)),\displaystyle\leq(r+1)(1+\operatorname{diam}(Y)+d(Y,U)),

where in the second line we used the triangle inequality and in the third we used UVrU\subseteq V^{r}. ∎

The above Lemma shows that the space 𝔇al(U){\mathfrak{D}}_{al}(U) only depends on the asymptotic geometry of the region UU in the following sense: if UVrU\subseteq V^{r} and VUrV\subseteq U^{r} for some r0r\geq 0 then 𝔇al(U)=𝔇al(V){\mathfrak{D}}_{al}(U)={\mathfrak{D}}_{al}(V) (as subsets of Y𝔡lY)\prod_{Y}{{\mathfrak{d}}_{l}}^{Y}) and are isomorphic as Fréchet spaces. In particular, for any non-empty bounded UnU\subset{\mathbb{R}}^{n}, the space 𝔇al(U){\mathfrak{D}}_{al}(U) coincides with 𝔇al({0}){\mathfrak{D}}_{al}(\{0\}).

To relate the spaces 𝔇al(U){\mathfrak{D}}_{al}(U) to the traditional CC^{*}-algebraic picture we prove the following:

Proposition 3.2.

Suppose UnU\subset{\mathbb{R}}^{n} is non-empty and bounded and let 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U). Then the sum X𝔹n𝖥X\sum_{X\in{\mathbb{B}}_{n}}{\mathsf{F}}^{X} is absolutely convergent and defines a continuous dense embedding of 𝔇al(U){\mathfrak{D}}_{al}(U) into the subspace of traceless anti-hermitian elements of the algebra 𝒜{\mathscr{A}}.

Proof.

We can assume without loss of generality that U={0}U=\{0\}. We have

Y𝔹n𝖥Y\displaystyle\sum_{Y\in{\mathbb{B}}_{n}}\|{\mathsf{F}}^{Y}\| 𝖥{0},2n+2Y𝔹n(1+diam(Y)+d(Y,0))2n2,\displaystyle\leq\|{\mathsf{F}}\|_{\{0\},2n+2}\sum_{Y\in{\mathbb{B}}_{n}}(1+\operatorname{diam}(Y)+d(Y,0))^{-2n-2},

and by Lemma 3.1 the above sum is finite. This shows that the map 𝔇al({0})𝒜{\mathfrak{D}}_{al}(\{0\})\to{\mathscr{A}} is well-defined and continuous. Its image is dense in the space of traceless anti-hermitian elements of 𝒜{\mathscr{A}} because it contains the dense subspace 𝔡l{{\mathfrak{d}}_{l}}. Finally, to show that it is injective, writing 𝒜:=Y𝔹n𝖥Y{\mathcal{A}}:=\sum_{Y\in{\mathbb{B}}_{n}}{\mathsf{F}}^{Y} we have

𝖥Y=tr¯Yc(𝒜)XYtr¯Xc(𝒜)\displaystyle{\mathsf{F}}^{Y}={\overline{\rm tr}}_{Y^{c}}({\mathcal{A}})-\sum_{X\subsetneq Y}{\overline{\rm tr}}_{X^{c}}({\mathcal{A}})

and so 𝒜=0𝖥Y=0{\mathcal{A}}=0\implies{\mathsf{F}}^{Y}=0 for all Y𝔹nY\in{\mathbb{B}}_{n}. ∎

Definition 3.3.

We let 𝔡al𝒜{{\mathfrak{d}}_{al}}\subset{\mathscr{A}} be the image of 𝔇al({0}){\mathfrak{D}}_{al}(\{0\}) under the embedding of Prop. 3.2, with the Fréchet topology of 𝔇al({0}){\mathfrak{D}}_{al}(\{0\}).

In [11], the algebra 𝒜a{\mathscr{A}}_{a\ell} of almost-local operators was defined as a subspace of 𝒜{\mathscr{A}} where a countable family of norms similar to (6) takes finite values. Here we equivalently define 𝒜a{\mathscr{A}}_{a\ell} as the set of elements of 𝒜{\mathscr{A}} whose traceless hermitian and anti-hermitian parts live in 𝔡al{{\mathfrak{d}}_{al}}, topologized as 𝒜a=𝟏(𝔡al){\mathscr{A}}_{a\ell}={\mathbb{C}}{\mathbf{1}}\oplus({{\mathfrak{d}}_{al}}\otimes{\mathbb{C}}).

Proposition 3.3.

𝒜a{\mathscr{A}}_{a\ell} is a dense sub-algebra of 𝒜{\mathscr{A}}.

Proof.

𝒜a{\mathscr{A}}_{a\ell} contains all local observables and these are dense in 𝒜{\mathscr{A}}. The fact that 𝒜a{\mathscr{A}}_{a\ell} is closed under multiplication is proven in [11]. ∎

In view of Prop. 3.2 it is natural to make the following definition.

Definition 3.4.

An element 𝖥𝔇al{\mathsf{F}}\in{\mathfrak{D}}_{al} is inner iff it is contained in 𝔇al(U){\mathfrak{D}}_{al}(U) for some bounded UU.

For any two bounded sets U,VU,V, a local observable is strictly localized on both UU and VV iff it is localized on their intersection, ie. 𝒜(U)𝒜(V)=𝒜(UV){\mathscr{A}}(U)\cap{\mathscr{A}}(V)={\mathscr{A}}(U\cap V). An analogous relation for the spaces 𝔇al(U){\mathfrak{D}}_{al}(U) does not hold in general777Indeed, for any two bounded U,VU,V the spaces 𝔇al(U){\mathfrak{D}}_{al}(U) and 𝔇al(V){\mathfrak{D}}_{al}(V) coincide and are nontrivial but if UU and VV are disjoint then 𝔇al(UV)=0{\mathfrak{D}}_{al}(U\cap V)=0., but it does hold if we assume that UU and VV satisfy the following transversality condition:

Definition 3.5.

Let U,VnU,V\subseteq{\mathbb{R}}^{n} and C>0C>0. We say UU and VV are CC-transverse if

d(x,UV)Cmax(d(x,U),d(x,V))\displaystyle d(x,U\cap V)\leq C\max(d(x,U),d(x,V))

for all xnx\in{\mathbb{R}}^{n}.

We will say UU and VV are transverse if they are CC-transverse for some C>0C>0.

Proposition 3.4.

If U,VnU,V\subseteq{\mathbb{R}}^{n} are CC-transverse then

max(𝖥U,k,𝖥V,k)𝖥UV,k(C+1)kmax(𝖥U,k,𝖥V,k)\displaystyle\max(\|{\mathsf{F}}\|_{U,k},\|{\mathsf{F}}\|_{V,k})\leq\|{\mathsf{F}}\|_{U\cap V,k}\leq(C+1)^{k}\max(\|{\mathsf{F}}\|_{U,k},\|{\mathsf{F}}\|_{V,k}) (8)

for all k>0k>0. In particular, 𝔇al(UV){\mathfrak{D}}_{al}(U\cap V) is a topological pullback: it is the set 𝔇al(U)𝔇al(V){\mathfrak{D}}_{al}(U)\cap{\mathfrak{D}}_{al}(V) with the topology of simultaneous convergence in 𝔇al(U){\mathfrak{D}}_{al}(U) and 𝔇al(V){\mathfrak{D}}_{al}(V).

Proof.

The first inequality is true even without assuming transversality – it follows from Lemma 3.2. For the second, let Z𝔹nZ\in{\mathbb{B}}_{n} and choose x,yZx^{*},y^{*}\in Z so that d(x,U)=d(Z,U)d(x^{*},U)=d(Z,U) and d(y,V)=d(Z,V)d(y^{*},V)=d(Z,V). Then we have

d(Z,UV)\displaystyle d(Z,U\cap V) =infzZd(z,UV)\displaystyle=\inf_{z\in Z}d(z,U\cap V)
CinfzZmax(d(z,U),d(z,V))\displaystyle\leq C\inf_{z\in Z}\max(d(z,U),d(z,V))
CinfzZmax(d(z,x)+d(x,U),d(z,y)+d(y,V))\displaystyle\leq C\inf_{z\in Z}\max(d(z,x^{*})+d(x^{*},U),d(z,y^{*})+d(y^{*},V))
C(diam(Z)+max(d(Z,U),d(Z,V)),\displaystyle\leq C(\operatorname{diam}(Z)+\max(d(Z,U),d(Z,V)),

and thus

1+diam(Z)+d(Z,UV)\displaystyle 1+\operatorname{diam}(Z)+d(Z,U\cap V)\leq
(C+1)(1+diam(Z)+max(d(Z,U),d(Z,V))),\displaystyle\hskip 28.45274pt\leq(C+1)(1+\operatorname{diam}(Z)+\max(d(Z,U),d(Z,V))), (9)

which proves (8). ∎

The next proposition relates 𝔇al(UV){\mathfrak{D}}_{al}(U\cup V) with 𝔇al(U){\mathfrak{D}}_{al}(U) and 𝔇al(V){\mathfrak{D}}_{al}(V) for any U,VnU,V\subseteq{\mathbb{R}}^{n}.

Proposition 3.5.

For any U,VnU,V\subset{\mathbb{R}}^{n} consider the sequence of vector spaces

𝔇al(UV)𝛼𝔇al(U)𝔇al(V)𝛽𝔇al(UV)0,\displaystyle{\mathfrak{D}}_{al}(U\cap V)\xrightarrow{\alpha}{\mathfrak{D}}_{al}(U)\oplus{\mathfrak{D}}_{al}(V)\xrightarrow{\beta}{\mathfrak{D}}_{al}(U\cup V)\to 0, (10)

where α(𝖥)=(𝖥,𝖥)\alpha({\mathsf{F}})=({\mathsf{F}},-{\mathsf{F}}) and β(𝖥,𝖦)=𝖥+𝖦\beta({\mathsf{F}},{\mathsf{G}})={\mathsf{F}}+{\mathsf{G}}.

  1. i)

    The sequence (10) admits a right splitting, i.e. a map γ:𝔇al(UV)𝔇al(U)𝔇al(V)\gamma:{\mathfrak{D}}_{al}(U\cup V)\to{\mathfrak{D}}_{al}(U)\oplus{\mathfrak{D}}_{al}(V) with βγ=id\beta\circ\gamma=\operatorname{id}. In particular, it is exact on the right.

  2. ii)

    If UU and VV are transverse, then (10) is exact on the left.

Proof.

i)i). For any Y𝔹nY\in{\mathbb{B}}_{n}, define

χ(Y):={1 if d(Y,U)<d(Y,V)1/2 if d(Y,U)=d(Y,V)0 if d(Y,U)>d(Y,V).\displaystyle\chi(Y):=\left\{\begin{array}[]{cc}1&\text{ if }d(Y,U)<d(Y,V)\\ 1/2&\text{ if }d(Y,U)=d(Y,V)\\ 0&\text{ if }d(Y,U)>d(Y,V)\end{array}\right..

For any 𝖥𝔇al(UV){\mathsf{F}}\in{\mathfrak{D}}_{al}(U\cup V) define γ1(𝖥):=Y𝔹nχ(Y)𝖥Y\gamma_{1}({\mathsf{F}}):=\sum_{Y\in{\mathbb{B}}_{n}}\chi(Y){\mathsf{F}}^{Y}, and γ2(𝖥)=𝖥γ1(𝖥)\gamma_{2}({\mathsf{F}})={\mathsf{F}}-\gamma_{1}({\mathsf{F}}). Then it is not hard to show that γ1(𝖥)U,k𝖥UV,k\|\gamma_{1}({\mathsf{F}})\|_{U,k}\leq\|{\mathsf{F}}\|_{U\cup V,k} and γ2(𝖥)V,k𝖥UV,k\|\gamma_{2}({\mathsf{F}})\|_{V,k}\leq\|{\mathsf{F}}\|_{U\cup V,k} and that γ=(γ1,γ2)\gamma=(\gamma_{1},\gamma_{2}) is a right splitting of (10).

ii)ii). Follows immediately from Proposition 3.4. ∎

Next we will show how to endow 𝔇al(U){\mathfrak{D}}_{al}(U) with the Lie algebra structure.

Proposition 3.6.

Let U,VnU,V\subset{\mathbb{R}}^{n}. For any 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U) and 𝖦𝔇al(V){\mathsf{G}}\in{\mathfrak{D}}_{al}(V) the sum

[𝖥,𝖦]Z:=X,Y𝔹n[𝖥X,𝖦Y]Z\displaystyle[{\mathsf{F}},{\mathsf{G}}]^{Z}:=\sum_{X,Y\in{\mathbb{B}}_{n}}[{\mathsf{F}}^{X},{\mathsf{G}}^{Y}]^{Z} (11)

is absolutely convergent for every Z𝔹nZ\in{\mathbb{B}}_{n}. The resulting bracket [,][\cdot,\cdot] satisfies the Jacobi identity and

[𝖥,𝖦]U,k\displaystyle\|[{\mathsf{F}},{\mathsf{G}}]\|_{U,k} C3k𝖥U,k+4n+4𝖦V,k+4n+4\displaystyle\leq C3^{k}\|{\mathsf{F}}\|_{U,k+4n+4}\|{\mathsf{G}}\|_{V,k+4n+4} (12)

for some constant C>0C>0 that depends only on nn.

To prove Proposition 3.6 we will need several lemmas. For any X,Y𝔹nX,Y\in{\mathbb{B}}_{n}, define888This is well-defined since the intersection of an arbitrary number of bricks is either empty or a brick, so XYX\vee Y is the intersection of all bricks containing XX and YY. the join XY𝔹nX\vee Y\in{\mathbb{B}}_{n} as the smallest brick that contains XX and YY.

Lemma 3.3.

For any X,Y𝔹nX,Y\in{\mathbb{B}}_{n} with XYX\cap Y\neq\varnothing we have

diam(XY)diam(X)+diam(Y)\displaystyle\operatorname{diam}(X\vee Y)\leq\operatorname{diam}(X)+\operatorname{diam}(Y) (13)

and for any zXYz\in X\vee Y we have

d(z,X)diam(Y).\displaystyle d(z,X)\leq\operatorname{diam}(Y).
Proof.

Let πi:n\pi_{i}:{\mathbb{R}}^{n}\to{\mathbb{R}} be the projection onto the iith coordinate. The following identites hold for any bricks X,Y𝔹nX,Y\in{\mathbb{B}}_{n}:

πi(XY)\displaystyle\pi_{i}(X\vee Y) =πi(X)πi(Y),\displaystyle=\pi_{i}(X)\vee\pi_{i}(Y),
diam(X)\displaystyle\operatorname{diam}(X) =maxi=1,ndiam(πi(X)),\displaystyle=\max_{i=1,\ldots n}\operatorname{diam}(\pi_{i}(X)),
d(X,Y)\displaystyle d(X,Y) =maxi=1,nd(πi(X),πi(Y)).\displaystyle=\max_{i=1,\ldots n}d(\pi_{i}(X),\pi_{i}(Y)).

When n=1n=1 the results are clear. When n>1n>1 they follow from the n=1n=1 case via the above identities. ∎

Lemma 3.4.

Let X,Y,Z𝔹nX,Y,Z\in{\mathbb{B}}_{n} and let 𝖥𝔡l(X){\mathsf{F}}\in{{\mathfrak{d}}_{l}}(X) and 𝖦𝔡l(Y){\mathsf{G}}\in{{\mathfrak{d}}_{l}}(Y). Then [𝖥X,𝖦Y]Z=0[{\mathsf{F}}^{X},{\mathsf{G}}^{Y}]^{Z}=0 unless XYX\cap Y\neq\varnothing and ZXYZ\subseteq X\vee Y.

Proof.

The requirement that XYX\cap Y\neq\varnothing is clear, since 𝖥X{\mathsf{F}}^{X} and 𝖦Y{\mathsf{G}}^{Y} would commute otherwise. Suppose ZXYZ\nsubseteq X\vee Y. Then Z:=(XY)ZZ^{\prime}:=(X\vee Y)\cap Z is a brick that is strictly contained in ZZ, so [𝖥X,𝖦Y]Z𝔡lZ𝔡l(Z)={0}.[{\mathsf{F}}^{X},{\mathsf{G}}^{Y}]^{Z}\in{{\mathfrak{d}}_{l}}^{Z}\cap{{\mathfrak{d}}_{l}}(Z^{\prime})=\{0\}.

We make the following definitions for the next lemma. For any UnU\subset{\mathbb{R}}^{n} and any brick XX write d~(X,U):=1+diam(X)+d(X,U)\tilde{d}(X,U):=1+\operatorname{diam}(X)+d(X,U) and d~(X):=1+diam(X)\tilde{d}(X):=1+\operatorname{diam}(X).

Lemma 3.5.

For any UnU\subset{\mathbb{R}}^{n}, Z𝔹nZ\in{\mathbb{B}}_{n}, and k0k\geq 0 we have

X,Y𝔹nXYZXYd~(X,U)k4n4d~(Y)k4n4π816n(n+1)43k1296d~(Z,U)k.\displaystyle\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\tilde{d}(X,U)^{-k-4n-4}\tilde{d}(Y)^{-k-4n-4}\leq\frac{\pi^{8}16^{n}(n+1)^{4}3^{k}}{1296}\tilde{d}(Z,U)^{-k}.
Proof.

Let X,Y𝔹nX,Y\in{\mathbb{B}}_{n} with XYX\cap Y\neq\varnothing and let ZXYZ\subset X\vee Y be a brick. Pick an arbitrary point wXYw\in X\cap Y. We have

d(Z,U)\displaystyle d(Z,U) d(Z,w)+d(w,U)\displaystyle\leq d(Z,w)+d(w,U)
d(Z,w)+diam(X)+d(X,U)\displaystyle\leq d(Z,w)+\operatorname{diam}(X)+d(X,U)
diam(XY)+diam(X)+d(X,U)\displaystyle\leq\operatorname{diam}(X\vee Y)+\operatorname{diam}(X)+d(X,U)
2diam(X)+diam(Y)+d(X,U),\displaystyle\leq 2\operatorname{diam}(X)+\operatorname{diam}(Y)+d(X,U),

and so, since by Lemma 3.3 diam(Z)diam(XY)diam(X)+diam(Y)\operatorname{diam}(Z)\leq\operatorname{diam}(X\vee Y)\leq\operatorname{diam}(X)+\operatorname{diam}(Y), we have

d~(Z,U)\displaystyle\tilde{d}(Z,U) 1+3diam(X)+2diam(Y)+d(X,U)\displaystyle\leq 1+3\operatorname{diam}(X)+2\operatorname{diam}(Y)+d(X,U)
3(1+diam(X)+d(X,U)+diam(Y))\displaystyle\leq 3(1+\operatorname{diam}(X)+d(X,U)+\operatorname{diam}(Y))
3(1+diam(X)+d(X,U))(1+diam(Y))\displaystyle\leq 3(1+\operatorname{diam}(X)+d(X,U))(1+\operatorname{diam}(Y))
=3d~(X,U)d~(Y).\displaystyle=3\tilde{d}(X,U)\tilde{d}(Y). (14)

It follows that for any k>0k>0 we have

d~(Z,U)kd~(X,U)kd~(Y)k\displaystyle\tilde{d}(Z,U)^{k}\tilde{d}(X,U)^{-k}\tilde{d}(Y)^{-k} 3k\displaystyle\leq 3^{k}

and so

X,Y𝔹nXYZXYd~(X,U)k4n4d~(Y)k4n4\displaystyle\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\tilde{d}(X,U)^{-k-4n-4}\tilde{d}(Y)^{-k-4n-4}\leq
3kd~(Z,U)kX,Y𝔹nXYZXYd~(X,U)4n4d~(Y)4n4\displaystyle\hskip 28.45274pt\leq 3^{k}\tilde{d}(Z,U)^{-k}\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\tilde{d}(X,U)^{-4n-4}\tilde{d}(Y)^{-4n-4} (15)

It remains to bound the sum on the right-hand side. Fix an arbitrary zZz\in Z and let X,Y𝔹nX,Y\in{\mathbb{B}}_{n} with XYX\cap Y\neq\varnothing and ZXYZ\subset X\vee Y. Then by the second statement in Lemma 3.3 and the inequality 1+a+b(1+a)(1+b)1+a+b\leq(1+a)(1+b) for a,b0a,b\geq 0 we have

(1+d(z,X)+diam(X))(1+d(z,Y)+diam(Y))\displaystyle(1+d(z,X)+\operatorname{diam}(X))(1+d(z,Y)+\operatorname{diam}(Y))\leq
(1+diam(X))2(1+diam(Y))2.\displaystyle\hskip 42.67912pt\leq(1+\operatorname{diam}(X))^{2}(1+\operatorname{diam}(Y))^{2}. (16)

Using (16) we can bound the sum (15) as follows:

X,Y𝔹nXYZXYd~(X,U)4n4d~(Y)4n4X,Y𝔹nXYZXY(1+diam(X))4n4(1+diam(Y))4n4X,Y𝔹n(1+diam(X)+d(X,z))2n2(1+diam(Y)+d(Y,z))2n2(X𝔹n(1+diam(X)+d(X,z))2n2)2π816n(n+1)41296.\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\tilde{d}(X,U)^{-4n-4}\tilde{d}(Y)^{-4n-4}\leq\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}(1+\operatorname{diam}(X))^{-4n-4}(1+\operatorname{diam}(Y))^{-4n-4}\\ \leq\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\end{subarray}}(1+\operatorname{diam}(X)+d(X,z))^{-2n-2}(1+\operatorname{diam}(Y)+d(Y,z))^{-2n-2}\\ \leq\left(\sum_{\begin{subarray}{c}X\in{\mathbb{B}}_{n}\end{subarray}}(1+\operatorname{diam}(X)+d(X,z))^{-2n-2}\right)^{2}\leq\frac{\pi^{8}16^{n}(n+1)^{4}}{1296}. (17)

Now we are ready to prove Proposition 3.6.

Proof of Proposition 3.6.

Notice that by Lemma 3.2 we have 𝖦V,k+2n+2𝖦n,k+2n+2\|{\mathsf{G}}\|_{V,k+2n+2}\leq\|{\mathsf{G}}\|_{{\mathbb{R}}^{n},k+2n+2} so without loss of generality we set V=nV={\mathbb{R}}^{n}.

Let Z𝔹nZ\in{\mathbb{B}}_{n}. By Lemmas 3.4 and 3.5 we have

X,Y𝔹n[𝖥X,𝖦Y]Z\displaystyle\sum_{X,Y\in{\mathbb{B}}_{n}}\|[{\mathsf{F}}^{X},{\mathsf{G}}^{Y}]^{Z}\| =X,Y𝔹nXYZXY[𝖥X,𝖦Y]Z\displaystyle=\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\|[{\mathsf{F}}^{X},{\mathsf{G}}^{Y}]^{Z}\|
2𝖥U,k+4n+4𝖦n,k+4n+4X,Y𝔹nXYZXYd~(X,U)k4n4d~(Y)k4n4\displaystyle\leq 2\|{\mathsf{F}}\|_{U,k+4n+4}\|{\mathsf{G}}\|_{{\mathbb{R}}^{n},k+4n+4}\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Z\subset X\vee Y\end{subarray}}\tilde{d}(X,U)^{-k-4n-4}\tilde{d}(Y)^{-k-4n-4}
C3k𝖥U,k+4n+4𝖦n,k+4n+4d~(Z,U)k,\displaystyle\leq C3^{k}\|{\mathsf{F}}\|_{U,k+4n+4}\|{\mathsf{G}}\|_{{\mathbb{R}}^{n},k+4n+4}\tilde{d}(Z,U)^{-k},

with C=π816n(n+1)4648C=\frac{\pi^{8}16^{n}(n+1)^{4}}{648}. This proves that (11) is absolutely convergent and establishes the bound (12).

Next, let us prove the Jacobi identity. Let 𝖥,𝖦,𝖧𝔇al(n){\mathsf{F}},{\mathsf{G}},{\mathsf{H}}\in{\mathfrak{D}}_{al}({\mathbb{R}}^{n}). We have

[𝖥,[𝖦,𝖧]]W\displaystyle[{\mathsf{F}},[{\mathsf{G}},{\mathsf{H}}]]^{W} =X,Y𝔹nX,Y𝔹n[𝖥X,[𝖦X,𝖧Y]Y]W\displaystyle=\sum_{X,Y\in{\mathbb{B}}_{n}}\sum_{X^{\prime},Y^{\prime}\in{\mathbb{B}}_{n}}\left[{\mathsf{F}}^{X},\left[{\mathsf{G}}^{X^{\prime}},{\mathsf{H}}^{Y^{\prime}}\right]^{Y}\right]^{W} (18)

To show that this sum is absolutely convergent, we bound

X,Y𝔹nX,Y𝔹n[𝖥X,[𝖦X,𝖧Y]Y]W\displaystyle\sum_{X,Y\in{\mathbb{B}}_{n}}\sum_{X^{\prime},Y^{\prime}\in{\mathbb{B}}_{n}}\left\|\left[{\mathsf{F}}^{X},\left[{\mathsf{G}}^{X^{\prime}},{\mathsf{H}}^{Y^{\prime}}\right]^{Y}\right]^{W}\right\|
\displaystyle\leq 4X,Y𝔹nXYWXY𝖥XX,Y𝔹nXYYXY𝖦X𝖧Y\displaystyle 4\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ W\subset X\vee Y\end{subarray}}\|{\mathsf{F}}^{X}\|\sum_{\begin{subarray}{c}X^{\prime},Y^{\prime}\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ Y\subset X^{\prime}\vee Y^{\prime}\end{subarray}}\|{\mathsf{G}}^{X^{\prime}}\|\|{\mathsf{H}}^{Y^{\prime}}\|
\displaystyle\leq C4𝖦n,8n+8𝖧n,8n+8X,Y𝔹nXYWXY𝖥Xd~(Y)4n4\displaystyle C4\|{\mathsf{G}}\|_{{\mathbb{R}}^{n},8n+8}\|{\mathsf{H}}\|_{{\mathbb{R}}^{n},8n+8}\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ W\subset X\vee Y\end{subarray}}\|{\mathsf{F}}^{X}\|\tilde{d}(Y)^{-4n-4}
\displaystyle\leq C4𝖥n,4n+4𝖦n,8n+8𝖧n,8n+8X,Y𝔹nXYWXYd~(X)4n4d~(Y)4n4\displaystyle C4\|{\mathsf{F}}\|_{{\mathbb{R}}^{n},4n+4}\|{\mathsf{G}}\|_{{\mathbb{R}}^{n},8n+8}\|{\mathsf{H}}\|_{{\mathbb{R}}^{n},8n+8}\sum_{\begin{subarray}{c}X,Y\in{\mathbb{B}}_{n}\\ X\cap Y\neq\varnothing\\ W\subset X\vee Y\end{subarray}}\tilde{d}(X)^{-4n-4}\tilde{d}(Y)^{-4n-4}
<\displaystyle< .\displaystyle\infty.

Here we used Lemma 3.5 in the second and fourth lines and CC is a constant depending only on nn. Thus the sum (18) is absolutely convergent. In particular, using the fact that Y𝔹n[𝖦X,𝖧Y]Y=[𝖦X,𝖧Y]\sum_{Y\in{\mathbb{B}}_{n}}[{\mathsf{G}}^{X^{\prime}},{\mathsf{H}}^{Y^{\prime}}]^{Y}=[{\mathsf{G}}^{X^{\prime}},{\mathsf{H}}^{Y^{\prime}}] we have the following absolutely convergent expression

[𝖥,[𝖦,𝖧]]W=X,Y,Z𝔹n[𝖥X,[𝖦Y,𝖧Z]]W.\displaystyle[{\mathsf{F}},[{\mathsf{G}},{\mathsf{H}}]]^{W}=\sum_{X,Y,Z\in{\mathbb{B}}_{n}}[{\mathsf{F}}^{X},[{\mathsf{G}}^{Y},{\mathsf{H}}^{Z}]]^{W}.

It is then easy to check that the Jacobi identity for the sum follows from the Jacobi identity for each term. ∎

From Propositions 3.4 and 3.6 we immediately get

Corollary 3.1.

Suppose U,VnU,V\in{\mathbb{R}}^{n}.

  1. i)

    (𝖥,𝖦)[𝖥,𝖦]({\mathsf{F}},{\mathsf{G}})\mapsto[{\mathsf{F}},{\mathsf{G}}] is a jointly continuous bilinear map from 𝔇al(U)×𝔇al(V){\mathfrak{D}}_{al}(U)\times{\mathfrak{D}}_{al}(V) to 𝔇al(U)𝔇al(V){\mathfrak{D}}_{al}(U)\cap{\mathfrak{D}}_{al}(V).

  2. ii)

    If UU and VV are transverse, then this is jointly continuous bilinear map from 𝔇al(U)×𝔇al(V){\mathfrak{D}}_{al}(U)\times{\mathfrak{D}}_{al}(V) to 𝔇al(UV){\mathfrak{D}}_{al}(U\cap V).

In particular since {0}\{0\} and n{\mathbb{R}}^{n} are transverse, 𝔇al(n){\mathfrak{D}}_{al}({\mathbb{R}}^{n}) acts continuously on 𝔇al({0}){\mathfrak{D}}_{al}(\{0\}) and this action is easily seen to extend to a continuous action of 𝔇al(n){\mathfrak{D}}_{al}({\mathbb{R}}^{n}) on the space 𝒜a{\mathscr{A}}_{a\ell}. We denote the action of 𝖥𝔇al(n){\mathsf{F}}\in{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) on 𝒜𝒜a{\mathcal{A}}\in{\mathscr{A}}_{a\ell} by 𝒜𝖥(𝒜){\mathcal{A}}\mapsto{\mathsf{F}}({\mathcal{A}}). By Prop. 3.6, it is given by

𝖥(𝒜)=Y𝔹n[𝖥Y,𝒜].{\mathsf{F}}({\mathcal{A}})=\sum_{Y\in{\mathbb{B}}_{n}}[{\mathsf{F}}^{Y},{\mathcal{A}}]. (19)

It is not hard to check that for any X𝔹nX\in{\mathbb{B}}_{n} and any 𝒜𝒜(X){\mathcal{A}}\in{\mathscr{A}}(X) we have tr¯Xc(𝖥(𝒜))=[𝖥X,𝒜]{\overline{\rm tr}}_{X^{c}}\left({\mathsf{F}}({\mathcal{A}})\right)=[{\mathsf{F}}_{X},{\mathcal{A}}] and so the action of 𝔇al(n){\mathfrak{D}}_{al}({\mathbb{R}}^{n}) on 𝒜a{\mathscr{A}}_{a\ell} is faithful. Thus, elements of 𝔇al(U){\mathfrak{D}}_{al}(U) for any UnU\subset{\mathbb{R}}^{n} may be identified with a subset of the Fréchet-continuous derivations of 𝒜a{\mathscr{A}}_{a\ell}. By Proposition 3.2, the Fréchet-Lie algebra 𝔇al({0}){\mathfrak{D}}_{al}(\{0\}) is identified with the Fréchet-Lie algebra 𝔡al{{\mathfrak{d}}_{al}} of traceless anti-hermitian elements of 𝒜a{\mathscr{A}}_{a\ell} acting by inner derivations.

3.2 Automorphisms

In this section we recall certain automorphisms obtained by exponentiating elements of 𝔇al(n){\mathfrak{D}}_{al}({\mathbb{R}}^{n}) following [11]. One can develop the theory of such automorphisms that are almost-localized on regions in n{\mathbb{R}}^{n} in a similar spirit to the above, but since we do not have much occasion to use them in this work, we opt instead for a more minimal development. Let 𝖥:𝔇al(n){\mathsf{F}}:{\mathbb{R}}\to{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) be a smooth map. It is shown in [11] that for any 𝒜𝒜a{\mathcal{A}}\in{\mathscr{A}}_{a\ell} the differential equation

ddt𝒜(t)=𝖥(t)(𝒜(t))\displaystyle\frac{d}{dt}{\mathcal{A}}(t)={\mathsf{F}}(t)({\mathcal{A}}(t))

with the initial condition 𝒜(0)=𝒜{\mathcal{A}}(0)={\mathcal{A}} has a unique solution 𝒜(t)𝒜a{\mathcal{A}}(t)\in{\mathscr{A}}_{a\ell} for all tt\in{\mathbb{R}}. Denote by αt𝖥\alpha^{\mathsf{F}}_{t} the map taking 𝒜{\mathcal{A}} to 𝒜(t){\mathcal{A}}(t). It is a continuous automorphism of the Lie algebra 𝒜a{\mathscr{A}}_{a\ell} that preserves the *-operation.

Definition 3.6.

We call any automorphism of the form αt𝖥\alpha^{\mathsf{F}}_{t} for some smooth map 𝖥:𝔇al(n){\mathsf{F}}:{\mathbb{R}}\rightarrow{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) a locally-generated automorphism, or LGA for short.

It is shown in [11] that every LGA extends to a continuous *-automorphism of the quasilocal algebra 𝒜{\mathscr{A}}, and that the set of LGAs forms a group under composition.

3.3 States

By a state ψ\psi we will mean a state on the quasilocal algebra 𝒜{\mathscr{A}}. If 𝖥{\mathsf{F}} is an inner derivation (see Def. 3.4), we define its ψ\psi-average as the evaluation of ψ\psi on the corresponding element of 𝔡al𝒜a{{\mathfrak{d}}_{al}}\subset{\mathscr{A}}_{a\ell}. The group of LGAs acts on states by pre-composition, which we denote ψα:=ψα\psi^{\alpha}:=\psi\circ\alpha. We say an LGA α\alpha preserves a state ψ\psi if ψα=ψ\psi^{\alpha}=\psi. We say an element 𝖥𝔇al(n){\mathsf{F}}\in{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) preserves ψ\psi if ψ(𝖥(𝒜))=0\psi({\mathsf{F}}({\mathcal{A}}))=0 for any 𝒜𝒜a{\mathcal{A}}\in{\mathscr{A}}_{a\ell}, which is equivalent to the one-parameter group of automorphisms tαt𝖥t\mapsto\alpha_{t}^{\mathsf{F}} corresponding to a constant map t𝖥t\mapsto{\mathsf{F}} preserving ψ\psi.

Definition 3.7.

For any UnU\subset{\mathbb{R}}^{n} define 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U) as the set of all elements of 𝔇al(U){\mathfrak{D}}_{al}(U) that preserve ψ\psi.

It is easy to check that 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U) is a closed subset of 𝔇al(U){\mathfrak{D}}_{al}(U), and that if 𝖥{\mathsf{F}} and 𝖦{\mathsf{G}} preserve ψ\psi, then [𝖥,𝖦][{\mathsf{F}},{\mathsf{G}}] preserves ψ\psi. Thus the analog of Propositions 3.4 and 3.6 and Corollary 3.1 hold for the spaces 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U). Proposition 3.5 on the other hand, does not hold for the spaces 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U) for a general state ψ\psi. To circumvent this, we will restrict to gapped states, where quasiadiabatic evolution [18, 19, 20] can be used to prove the analog of Proposition 3.5.

Definition 3.8.

A state ψ\psi is gapped if there exists 𝖧𝔇al(n){\mathsf{H}}\in{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) and Δ>0\Delta>0 such that for any 𝒜𝒜a{\mathcal{A}}\in{\mathscr{A}}_{a\ell} one has

iψ(𝒜𝖧(𝒜))Δ(ψ(𝒜𝒜)ψ(𝒜)ψ(𝒜)).\displaystyle-i\psi({\mathcal{A}}^{*}{\mathsf{H}}({\mathcal{A}}))\geq\Delta\left(\psi({\mathcal{A}}^{*}{\mathcal{A}})-\psi({\mathcal{A}}^{*})\psi({\mathcal{A}})\right). (20)
Remark 3.1.

The meaning of this condition becomes more transparent if one recalls that any 𝖧𝔇al{\mathsf{H}}\in{\mathfrak{D}}_{al} is a generator of a one-parameter group of *-automorphisms of 𝒜{\mathscr{A}} [11]. The condition (20) implies that ψ\psi is invariant under this one-parameter group of automorphisms [21], and that the corresponding one-parameter group of unitaries in the GNS representation of 𝒜{\mathscr{A}} has a generator whose spectrum in the orthogonal complement to the GNS vacuum vector is contained in [Δ,+)[\Delta,+\infty). The condition (20) also implies that ψ\psi is pure [22].

Remark 3.2.

If ψ\psi is a gapped state of 𝒜{\mathscr{A}} and α\alpha is an LGA, then ψα\psi^{\alpha} is also a gapped state. In [11] it was proposed to define a gapped phase as an orbit of gapped state under the action of the group of LGAs.

In Appendix A we prove the following.

Proposition 3.7.

Suppose ψ\psi is gapped, the corresponding Hamiltonian is 𝖧{\mathsf{H}}. Then there are linear functions

𝒥\displaystyle\mathcal{J} :𝔇al(n)𝔇alψ(n)\displaystyle:{\mathfrak{D}}_{al}({\mathbb{R}}^{n})\to{\mathfrak{D}}_{al}^{\psi}({\mathbb{R}}^{n})
𝒦\displaystyle\mathcal{K} :𝔇al(n)𝔇al(n)\displaystyle:{\mathfrak{D}}_{al}({\mathbb{R}}^{n})\to{\mathfrak{D}}_{al}({\mathbb{R}}^{n})

such that

  1. i)

    If 𝖥{\mathsf{F}} preserves ψ\psi then 𝒦(𝖥)\mathcal{K}({\mathsf{F}}) preserves ψ\psi.

  2. ii)

    For every k>0k>0, UnU\subset{\mathbb{R}}^{n}, and 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U) we have

    𝒥(𝖥)U,k\displaystyle\|\mathcal{J}({\mathsf{F}})\|_{U,k} Ck𝖥U,k+4n+3\displaystyle\leq C_{k}\|{\mathsf{F}}\|_{U,k+4n+3}
    𝒦(𝖥)U,k\displaystyle\|\mathcal{K}({\mathsf{F}})\|_{U,k} Ck𝖥U,k+4n+3\displaystyle\leq C^{\prime}_{k}\|{\mathsf{F}}\|_{U,k+4n+3}

    for some constants Ck,CkC_{k},C^{\prime}_{k} depending only on k,n,𝖧k,n,{\mathsf{H}}, and Δ\Delta.

  3. iii)

    For every 𝖥𝔇al(n){\mathsf{F}}\in{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) we have

    𝖥=𝒥(𝖥)𝒦([𝖧,𝖥]).\displaystyle{\mathsf{F}}=\mathcal{J}({\mathsf{F}})-\mathcal{K}([{\mathsf{H}},{\mathsf{F}}]).

Using the above Lemma we will prove the analog of Proposition 3.5 for the spaces 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U).

Proposition 3.8.

For any U,VnU,V\subset{\mathbb{R}}^{n} consider the sequence

𝔇alψ(UV)𝛼𝔇alψ(U)𝔇alψ(V)𝛽𝔇alψ(UV)0,\displaystyle{\mathfrak{D}}_{al}^{\psi}(U\cap V)\xrightarrow{\alpha}{\mathfrak{D}}_{al}^{\psi}(U)\oplus{\mathfrak{D}}_{al}^{\psi}(V)\xrightarrow{\beta}{\mathfrak{D}}_{al}^{\psi}(U\cup V)\to 0, (21)

where α(𝒜)=(𝒜,𝒜)\alpha({\mathcal{A}})=({\mathcal{A}},-{\mathcal{A}}) and β(𝒜,)=𝒜+\beta({\mathcal{A}},{\mathcal{B}})={\mathcal{A}}+{\mathcal{B}}.

  1. i)

    If ψ\psi is gapped then the sequence (10) admits a right splitting, and in particular it is exact on the right

  2. ii)

    If UU and VV are transverse, then (10) is exact on the left.

To prove Proposition 3.8 we will need the following geometric result

Lemma 3.6.

Let U,VnU,V\subset{\mathbb{R}}^{n} and define U:={xn:d(x,U)d(x,V)}U^{\prime}:=\{x\in{\mathbb{R}}^{n}:d(x,U)\leq d(x,V)\}. Then UU^{\prime} and UVU\cup V are transverse and their intersection is UU.

Proof.

It is easy to check that U(UV)=UU^{\prime}\cap(U\cup V)=U. To prove transversality we will show

d(x,U)4max(d(x,U),d(x,UV))\displaystyle d(x,U)\leq 4\max(d(x,U^{\prime}),d(x,U\cup V)) (22)

for every xnx\in{\mathbb{R}}^{n}. Suppose first that d(x,U)2d(x,V)d(x,U)\leq 2d(x,V). Then

d(x,U)\displaystyle d(x,U) 2min(d(x,U),d(x,V))\displaystyle\leq 2\min(d(x,U),d(x,V))
=2d(x,UV)\displaystyle=2d(x,U\cup V)

which implies (22). Suppose instead that d(x,U)>2d(x,V)d(x,U)>2d(x,V), and let yUy\in U^{\prime} satisfy d(x,y)=d(x,U)d(x,y)=d(x,U^{\prime}). Notice xUx\notin U^{\prime} and so yy lies in the boundary of UU^{\prime}, which implies d(y,U)=d(y,V)d(y,U)=d(y,V). Thus we have

d(x,U)\displaystyle d(x,U) d(x,y)+d(y,U)\displaystyle\leq d(x,y)+d(y,U)
=d(x,y)+d(y,V)\displaystyle=d(x,y)+d(y,V)
2d(x,y)+d(x,V)\displaystyle\leq 2d(x,y)+d(x,V)

where in the first and third lines we used the triangle inequality. Using d(x,y)=d(x,U)d(x,y)=d(x,U^{\prime}) and d(x,V)<d(x,U)/2d(x,V)<d(x,U)/2, this gives d(x,U)<4d(x,U)d(x,U)<4d(x,U^{\prime}), which implies (22). ∎

Proof of Proposition 3.8.

The proof of Proposition 3.5 goes through unmodified except for the definition of γ\gamma, which needs to be changed to ensure that the image of γ\gamma consists of derivations that preserve ψ\psi. Suppose U,VnU,V\subseteq{\mathbb{R}}^{n} and 𝖥𝔇alψ(UV){\mathsf{F}}\in{\mathfrak{D}}^{\psi}_{al}(U\cup V). Define

U\displaystyle U^{\prime} :={xn:d(x,U)d(x,V)},\displaystyle:=\{x\in{\mathbb{R}}^{n}:d(x,U)\leq d(x,V)\},
V\displaystyle V^{\prime} :={xn:d(x,V)d(x,U)}.\displaystyle:=\{x\in{\mathbb{R}}^{n}:d(x,V)\leq d(x,U)\}.

Let γU,V\gamma^{U,V} (resp. γU,V\gamma^{U^{\prime},V^{\prime}}) be the splitting from Proposition 3.5 with the sets UU and VV (resp. UU^{\prime} and VV^{\prime}). Define γ~=(γ~1,γ~2)\tilde{\gamma}=(\tilde{\gamma}_{1},\tilde{\gamma}_{2}) as

γ~i(𝖥)\displaystyle\tilde{\gamma}_{i}({\mathsf{F}}) :=𝒥(γiU,V(𝖥))𝒦([𝒥(γiU,V(𝖧)),𝖥])\displaystyle:=\mathcal{J}(\gamma_{i}^{U,V}({\mathsf{F}}))-\mathcal{K}([\mathcal{J}(\gamma_{i}^{U^{\prime},V^{\prime}}({\mathsf{H}})),{\mathsf{F}}])

for i=1,2i=1,2. Using Prop. 3.7 and Lemma 3.6 and the fact that the commutator of two derivations that preserve ψ\psi preserves ψ\psi, one checks that γ~\tilde{\gamma} takes 𝔇alψ(UV){\mathfrak{D}}_{al}^{\psi}(U\cup V) to 𝔇alψ(U)𝔇alψ(V){\mathfrak{D}}_{al}^{\psi}(U)\oplus{\mathfrak{D}}_{al}^{\psi}(V). Using Prop. 3.7 iii)iii) and the fact that γ1U,V(𝖥)+γ2U,V(𝖥)=𝖥\gamma_{1}^{U,V}({\mathsf{F}})+\gamma_{2}^{U,V}({\mathsf{F}})={\mathsf{F}} and γ1U,V(𝖧)+γ2U,V(𝖧)=𝖧\gamma_{1}^{U^{\prime},V^{\prime}}({\mathsf{H}})+\gamma_{2}^{U^{\prime},V^{\prime}}({\mathsf{H}})={\mathsf{H}}, we get γ~1(𝖥)+γ~2(𝖥)=𝖥\tilde{\gamma}_{1}({\mathsf{F}})+\tilde{\gamma}_{2}({\mathsf{F}})={\mathsf{F}}, as desired. ∎

We showed that one can attach Fréchet-Lie algebras 𝔇al(U){\mathfrak{D}}_{al}(U) and 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U) to any UnU\subset{\mathbb{R}}^{n}. These form a pre-cosheaf over the pre-ordered set of subsets of n{\mathbb{R}}^{n}. Moreover, Proposition 3.5, Corollary 3.1, and (when ψ\psi is gapped) Proposition 3.8 show that these spaces satisfy the cosheaf condition and Property I for transverse pairs U,VnU,V\subset{\mathbb{R}}^{n}. What prevents the functors 𝔇al{\mathfrak{D}}_{al} and 𝔇alψ{\mathfrak{D}}^{\psi}_{al} from forming local Lie algebras is the fact that pairs of subsets U,VnU,V\subset{\mathbb{R}}^{n} generally do not intersect transversely. This problem can be resolved by restricting to a suitable set of subsets of n{\mathbb{R}}^{n} that have well-behaved intersections. In the next section we identify one such set, the set of semilinear subsets, prove that they form a Grothendieck site, and discuss some properties of this site.

4 The site of fuzzy semilinear sets

4.1 Semilinear sets and their thickenings

A semilinear set in n{\mathbb{R}}^{n} is a subset of n{\mathbb{R}}^{n} which can be defined by means of a finite number of linear equalities and strict linear inequalities. More precisely, a basic semilinear set in n{\mathbb{R}}^{n} is an intersection of a finite number of hyperplanes and open half-spaces, and a semilinear set is a finite union of basic semilinear sets. The set of semilinear subsets of n{\mathbb{R}}^{n} will be denoted 𝒮n{\mathcal{S}}_{n}. Projections m×nm{\mathbb{R}}^{m}\times{\mathbb{R}}^{n}\rightarrow{\mathbb{R}}^{m} map 𝒮m+n{\mathcal{S}}_{m+n} to 𝒮m{\mathcal{S}}_{m} [23]. A map nm{\mathbb{R}}^{n}\rightarrow{\mathbb{R}}^{m} is called semilinear iff its graph is a semilinear subset of m+n{\mathbb{R}}^{m+n}. The composition of two semilinear maps is a semilinear map [23].

Recall that we use the \ell^{\infty} metric on n{\mathbb{R}}^{n}.

Lemma 4.1.

The distance function d:n×nd:{\mathbb{R}}^{n}\times{\mathbb{R}}^{n}\rightarrow{\mathbb{R}} is semilinear.

Proof.

The function (x,y)xiyi(x,y)\mapsto x_{i}-y_{i} is semilinear for any ii. The function ||:|\cdot|:{\mathbb{R}}\rightarrow{\mathbb{R}} is semilinear. If f,g:nf,g:{\mathbb{R}}^{n}\rightarrow{\mathbb{R}} are semilinear, then h=max(f,g):nh=\max(f,g):{\mathbb{R}}^{n}\rightarrow{\mathbb{R}} is semilinear. Since the set of semilinear functions is closed under composition, this proves the lemma. ∎

Recall that for any set UnU\subset{\mathbb{R}}^{n}, we write Ur:={xn:yU s.t d(x,y)r}U^{r}:=\{x\in{\mathbb{R}}^{n}:\exists y\in U\text{ s.t }d(x,y)\leq r\} and call this the rr-thickening of UU. It is easy to see that if UU is closed, then UrU^{r} is also closed for any r0r\geq 0.

Lemma 4.2.

If UU is semilinear, UrU^{r} is semilinear for any rr.

Proof.

Consider the set

Δr={(x,y)2n|d(x,y)r}.\Delta_{r}=\{(x,y)\in{\mathbb{R}}^{2n}|d(x,y)\leq r\}.

By the previous lemma, Δr\Delta_{r} is semilinear. On the other hand, UrU^{r} is the projection to the first n{\mathbb{R}}^{n} of Δr(n×U)n×n\Delta_{r}\cap({\mathbb{R}}^{n}\times U)\subset{\mathbb{R}}^{n}\times{\mathbb{R}}^{n}. Since intersection and projection preserve the set of semilinear sets, the lemma is proved. ∎

Lemma 4.3.

If UU is convex, then UrU^{r} is convex, for any r0r\geq 0.

Proof.

Suppose x,yUrx,y\in U^{r}, and suppose x,yUx^{\prime},y^{\prime}\in U satisfy d(x,x)rd(x,x^{\prime})\leq r and d(y,y)rd(y,y^{\prime})\leq r. Then for any t[0,1]t\in[0,1] we have d(tx+(1t)y,tx+(1t)x)t2d(x,x)+(1t)2d(y,y)r.d(tx+(1-t)y,tx^{\prime}+(1-t)x^{\prime})\leq t^{2}d(x,x^{\prime})+(1-t)^{2}d(y,y^{\prime})\leq r.

A polyhedron in n{\mathbb{R}}^{n} is an intersection of a finite number of closed half-spaces. A polyhedron is closed, but not necessarily compact. A closed semilinear set is the same as a finite union of polyhedra. Conversely, according to Theorem 19.6 from [24], a polyhedron can be described as a closed convex semilinear set. Combining this with the above lemmas, we get

Corollary 4.1.

If UU is a polyhedron, then UrU^{r} is a polyhedron, for any r0r\geq 0.

4.2 A category of fuzzy semilinear sets

Clearly, if for X,Y𝒮nX,Y\in{\mathcal{S}}_{n} we have XYX\subseteq Y, then for any r0r\geq 0 we have XrYrX^{r}\subseteq Y^{r}. Also, for any r,s0r,s\geq 0 and any U𝒮nU\in{\mathcal{S}}_{n} we have (Ur)sUr+s(U^{r})^{s}\subseteq U^{r+s}. Thus we can define a pre-order \leq on 𝒮n{\mathcal{S}}_{n} by saying that UVU\leq V iff there exists r0r\geq 0 such that UVrU\subseteq V^{r}. We will call this pre-order relation fuzzy inclusion. Equivalently, 𝒮n{\mathcal{S}}_{n} can be made into a category, with a single morphism from UU to VV iff UVU\leq V. One can turn the pre-ordered set (𝒮n,)({\mathcal{S}}_{n},\leq) into a poset by identifying isomorphic objects of the corresponding category, but for our purposes it is more convenient not to do so. On the other hand, every semilinear set is isomorphic to its closure, and we find it convenient to work with an equivalent category (or pre-ordered set) which contains only closed semilinear subsets. We will denote it 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} and call it the category (or pre-ordered set) of fuzzy semilinear sets.

Proposition 4.1.

𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} has all pairwise joins: for any U,V𝒞𝒮nU,V\in{\mathcal{C}\mathcal{S}}_{n} the join is given by UVU\cup V.

Proof.

If UWrU\subseteq W^{r} and VWsV\subseteq W^{s} for some r,s0r,s\geq 0, then UWmax(r,s)U\subseteq W^{\max(r,s)} and VWmax(r,s)V\subseteq W^{\max(r,s)}, and thus UVWmax(r,s)U\cup V\subseteq W^{\max(r,s)}. ∎

Let us show that 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} has all pairwise meets, using the notion of transverse intersection from the previous section. Recall (Definition 3.5) that we say two sets U,VnU,V\subset{\mathbb{R}}^{n} are transverse if for some C>0C>0 we have d(x,UV)Cmax(d(x,U),d(x,V))d(x,U\cap V)\leq C\max(d(x,U),d(x,V)) for all xnx\in{\mathbb{R}}^{n}.

We need the following geometric result [25]999The proof in [25] is for the Euclidean distance, but since the Euclidean distance function and d(x,y)=xyd(x,y)=\|x-y\|_{\infty} are equivalent (each one is upper-bounded by a multiple of the other), the result applies to d(x,y)d(x,y) as well. :

Lemma 4.4.

Let PP and QQ be polyhedra n{\mathbb{R}}^{n}. If PQP\cap Q\neq\varnothing then PP and QQ are transverse.

Proposition 4.2.

For every U,V𝒞𝒮nU,V\in{\mathcal{C}\mathcal{S}}_{n} there is an r>0r>0 such that UrU^{r} and VrV^{r} are transverse.

Proof.

Let U=iPiU=\cup_{i}P_{i} and V=jPjV=\cup_{j}P_{j} be a decomposition of UU and VV into finite unions of polyhedra and choose r>0r>0 so that PirQjrP_{i}^{r}\cup Q_{j}^{r} is nonempty for each pair i,ji,j. By Corollary 4.1 and Lemma 4.4 there are constants CPiQj>0C_{P_{i}Q_{j}}>0 such that

d(x,UrVr)Cmax(d(x,Ur),d(x,Vr))\displaystyle d(x,U^{r}\cap V^{r})\leq C\max(d(x,U^{r}),d(x,V^{r}))

for each pair i,ji,j. Since Ur=iPirU^{r}=\cup_{i}P_{i}^{r} and Vr=iQirV^{r}=\cup_{i}Q_{i}^{r}, for any xx\in{\mathbb{R}} there are indices ii^{*} and jj^{*} such that d(x,Ur)=d(x,Pir)d(x,U^{r})=d(x,P_{i^{*}}^{r}) and d(x,Vr)=d(x,Qjr)d(x,V^{r})=d(x,Q_{j^{*}}^{r}). Then we have

d(x,UrVr)\displaystyle d(x,U^{r}\cap V^{r}) =d(x,ij(PirQjr))\displaystyle=d(x,\cup_{ij}(P_{i}^{r}\cap Q_{j}^{r}))
d(x,PirQjr))\displaystyle\leq d(x,P_{i^{*}}^{r}\cap Q_{j^{*}}^{r}))
CPiQjmax(d(x,Pi),d(x,Qj))\displaystyle\leq C_{P_{i^{*}}Q_{j^{*}}}\max(d(x,P_{i^{*}}),d(x,Q_{j^{*}}))
=CPiQjmax(d(x,U),d(x,V)),\displaystyle=C_{P_{i^{*}}Q_{j^{*}}}\max(d(x,U),d(x,V)),

and so Ur,VrU^{r},V^{r} are CC-transverse for C:=maxi,jCPiQjC:=\max_{i,j}C_{P_{i}Q_{j}}. ∎

Corollary 4.2.

With U,V,rU,V,r as above, UrVrU^{r}\cap V^{r} is a meet of UU and VV. In particular, 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} has all pairwise meets.

Proof.

Since UU and UrU^{r} (resp. VV and VrV^{r}) are isomorphic in 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}, it suffices to show that UrVrU^{r}\cap V^{r} is a meet of UrU^{r} and VrV^{r}. It’s clear that UrVrUrU^{r}\cap V^{r}\leq U^{r} and UrVrVrU^{r}\cap V^{r}\leq V^{r}. Now suppose W𝒞𝒮nW\in{\mathcal{C}\mathcal{S}}_{n} satisfies WUrW\leq U^{r} and WVrW\leq V^{r}. Then there is an s>0s>0 such that every xWx\in W satisfies max(d(x,Ur),d(x,Vr))s\max(d(x,U^{r}),d(x,V^{r}))\leq s. Since d(x,UrVr)Cmax(d(x,Ur),d(x,Vr))d(x,U^{r}\cap V^{r})\leq C\max(d(x,U^{r}),d(x,V^{r})) for some C>0C>0 we have W(UrVr)CsW\subset(U^{r}\cap V^{r})^{Cs}. ∎

Proposition 4.3.

The pre-ordered set 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} is distributive.

Proof.

We need to show that for any U,V,W𝒞𝒮nU,V,W\in{\mathcal{C}\mathcal{S}}_{n} we have U(VW)(UV)(UW)U\wedge(V\vee W)\leq(U\wedge V)\vee(U\wedge W). According to Corollary 4.2, there exists r0r\geq 0 such that U(VW)Ur(VW)rU\wedge(V\vee W)\simeq U^{r}\cap(V\cup W)^{r}. Since (VW)r=VrWr,(V\cup W)^{r}=V^{r}\cup W^{r}, we also have U(VW)(UrVr)(UrWr)U\wedge(V\vee W)\simeq(U^{r}\cap V^{r})\cup(U^{r}\cap W^{r}). On the other hand, UVUrVr(Ur)s(Vr)sU\wedge V\simeq U^{r}\wedge V^{r}\simeq(U^{r})^{s}\cap(V^{r})^{s} for some s0s\geq 0, and UWUrWr(Ur)t(Wr)tU\wedge W\simeq U^{r}\wedge W^{r}\simeq(U^{r})^{t}\cap(W^{r})^{t} for some t0t\geq 0. Thus (UV)(UW)((Ur)s(Vr)s)((Ur)t(Wr)t)(U\wedge V)\vee(U\wedge W)\simeq\left((U^{r})^{s}\cap(V^{r})^{s}\right)\cup\left((U^{r})^{t}\cap(W^{r})^{t}\right) for some s,t0s,t\geq 0. Since we have inclusions UrVr(Ur)s(Vr)sU^{r}\cap V^{r}\subseteq(U^{r})^{s}\cap(V^{r})^{s} and UrWr(Ur)t(Wr)tU^{r}\cap W^{r}\subseteq(U^{r})^{t}\cap(W^{r})^{t}, the lemma is proved. ∎

We can now equip 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} with a Grothendieck topology of Section 2. There are two versions of it which differ in whether we allow empty covers of an initial object or not. In the case of the pre-ordered set 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}, every bounded closed semilinear set is an initial object (they are all isomorphic objects of the category 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}). Since such sets are not empty, we will disallow empty covers. This choice is also forced on us if we want certain pre-cosheaves to be cosheaves (see below). Note that we only consider non-empty closed semilinear sets.

More generally, for any W𝒞𝒮nW\in{\mathcal{C}\mathcal{S}}_{n} we may consider a full sub-category 𝒞𝒮n/W{{{\mathcal{C}\mathcal{S}}_{n}}/W} whose objects are U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} such that UWU\leq W. This is a distributive pre-ordered set, and we will also have occasion to consider local Lie algebras on the associated site.

4.3 Spherical CS sets

Every two bounded elements of 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} are isomorphic objects of the corresponding category. More generally, any two elements of 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} which coincide outside some ball in n{\mathbb{R}}^{n} are isomorphic objects. Thus 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} encodes the large-scale structure of n{\mathbb{R}}^{n}. To make this explicit, we will show that the pre-ordered set 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} is equivalent as a category to a certain poset of subsets of the “sphere at infinity” Sn1S^{n-1}.

A cone in n{\mathbb{R}}^{n} is a non-empty subset of n{\mathbb{R}}^{n} which is invariant under xλxx\mapsto\lambda x, where λ0\lambda\geq 0. Every cone contains the origin 0. Cones in n{\mathbb{R}}^{n} are in bijection with subsets of Sn1=(n\{0})/+S^{n-1}=({\mathbb{R}}^{n}\backslash\{0\})/{\mathbb{R}}^{*}_{+} where +{\mathbb{R}}^{*}_{+} is the group of positive real numbers under multiplication. If ASn1A\subset S^{n-1}, we denote the corresponding cone c(A)c(A). In particular, c()={0}nc(\varnothing)=\{0\}\in{\mathbb{R}}^{n}. If KK is a cone in n{\mathbb{R}}^{n}, we will denote the corresponding subset of Sn1S^{n-1} by K^\hat{K}.

Definition 4.1.

ASn1A\subset S^{n-1} is a spherical polyhedron iff c(A)c(A) is a polyhedron and AA is contained in some open hemisphere of Sn1S^{n-1}. ASn1A\subseteq S^{n-1} is a spherical CS set iff it is a union of a finite number of spherical polyhedra. The set of spherical CS sets in Sn1S^{n-1} is denoted 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n}.

Every polyhedron is convex and thus contractible. This implies:

Proposition 4.4.

Any spherical polyhedron is contractible.

Proof.

Without loss of generality, we may assume that the spherical polyhedron is contained in the hemisphere S+n1=Sn1{xn>0}S^{n-1}_{+}=S^{n-1}\cap\{x_{n}>0\}. The map n1S+n1{\mathbb{R}}^{n-1}\rightarrow S^{n-1}_{+} which sends (x1,,xn1)(x_{1},\ldots,x_{n-1}) to the equivalence class of (x1,,xn1,1)(x_{1},\ldots,x_{n-1},1) is a homeomorphism which establishes a bijection between bounded polyhedra in n1{\mathbb{R}}^{n-1} and spherical polyhedra contained in S+n1S^{n-1}_{+}. ∎

Proposition 4.5.

The intersection of two spherical polyhedra is a spherical polyhedron. The union and intersection of two spherical CS sets is a spherical CS set.

Proof.

Clear from definitions. ∎

Spherical CS sets form a poset 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n} under inclusion. This poset has pairwise joins and meets given by unions and intersections, respectively.

Proposition 4.6.

The category 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n} is equivalent to the category 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}.

Proof.

We proceed first by defining a functor from 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} to 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n}. Notice, it is sufficient to define a functor on isomorphism classes of polyhedrons then extend by joins. Every closed polyhedron is a finite intersection of closed half spaces i=1k{nixbi},\bigcap_{i=1}^{k}\{n_{i}\cdot x\leq b_{i}\}, where for each i=1,ki=1,\ldots k, ninn_{i}\in{\mathbb{R}}^{n} is the unit normal vector of the iith supporting hyperplane and bib_{i}\in{\mathbb{R}}. By definition, i=1k{nixbi}\bigcap_{i=1}^{k}\{n_{i}\cdot x\leq b_{i}\} is isomorphic to i=1k{nix0}\bigcap_{i=1}^{k}\{n_{i}\cdot x\leq 0\} in 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. In other words, every polyhedron is isomorphic to a cone {Nx0},\{Nx\preceq 0\}, where NN is the matrix with nin_{i} as rows and we write y0y\preceq 0 for yny\in{\mathbb{R}}^{n} when yi0y_{i}\leq 0 for all i=1,ni=1,\ldots n. If {ni}i=1k\{n_{i}\}_{i=1}^{k} spans n{\mathbb{R}}^{n}, the set {Nx0}\{Nx\preceq 0\}, which contains no antipodal points, is mapped to its corresponding spherical polyhedron. Otherwise, complete {ni}i=1k\{n_{i}\}_{i=1}^{k} into a spanning set {n1,,nk,m1,,ml}\{n_{1},\dots,n_{k},m_{1},\dots,m_{l}\} and decompose {Nx0}\{Nx\preceq 0\} into a union of cones {Nx0}j{±mjx0}\{Nx\preceq 0\}\cap\bigcap_{j}\{\pm m_{j}\cdot x\leq 0\}. By the universal property of the join, an arbitrary closed semilinear set is isomorphic to a union of conical polyhedra. For functoriality, it is sufficient to note that for any pair of closed cones U,VU,V a fuzzy inclusion UVrU\subseteq V^{r} for some r0r\geq 0 implies UVU\subseteq V. It is easy to check that this functor is fully faithful and (essentially) surjective. ∎

Note that under this equivalence all bounded elements of 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} correspond to 𝒮𝒞𝒮n\varnothing\in{\mathcal{S}\mathcal{C}\mathcal{S}}_{n}. The canonical Grothendieck topology on 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} corresponds to a slightly unusual Grothendieck topology on the poset of spherical CS subsets of Sn1S^{n-1}: the one where empty covers of \varnothing are not allowed. Consequently, a cosheaf of vector spaces on 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n} equipped with this topology need not map \varnothing to the zero vector space.

4.4 Spherical CS cohomology

Let 𝔘{\mathfrak{U}} be a spherical CS cover of a spherical CS set AA. Spherical CS cohomology HˇCS(𝔘,A;)\check{H}^{\bullet}_{CS}({\mathfrak{U}},A;{\mathbb{R}}) is defined to be the simplicial cohomology of the Čech nerve N(𝔘)N({\mathfrak{U}}).

Definition 4.2.

Let AA be a spherical CS set. The spherical CS cohomology HˇCS(A,)\check{H}^{\bullet}_{CS}(A,{\mathbb{R}}) is defined as limHˇCS(𝔘,A;),\varinjlim\check{H}^{\bullet}_{CS}({\mathfrak{U}},A;{\mathbb{R}}), where the colimit is taken over the directed set of all spherical CS covers.

The following proposition connects spherical CS cohomology with singular cohomology using a functorial version of nerve theorems [26, 27].

Proposition 4.7.

For any spherical CS set AA the graded vector space HˇCS(A,)\check{H}^{\bullet}_{CS}(A,{\mathbb{R}}) is isomorphic to the singular cohomology H(A,)H^{\bullet}(A,{\mathbb{R}}).

Proof.

Every spherical CS cover can be refined to a cover by spherical polyhedra, so in the computation of limHˇCS(𝔘,A;)\varinjlim\check{H}^{\bullet}_{CS}({\mathfrak{U}},A;{\mathbb{R}}) it suffices to take colimit over such covers. All intersections of elements of a spherical polyhedral cover 𝔘{\mathfrak{U}} are contractible. Also, since every polyhedron is a geometric realization of a simplicial complex, 𝔘{\mathfrak{U}} is a cover of a simplicial complex by subcomplexes. By Theorem C of [28] for such 𝔘{\mathfrak{U}} the direct system of groups 𝔘HˇCS(𝔘,A;){\mathfrak{U}}\mapsto\check{H}^{\bullet}_{CS}({\mathfrak{U}},A;{\mathbb{R}}) is constant and its limit is H(A,)H^{\bullet}(A,{\mathbb{R}}). ∎

5 Local Lie algebras over fuzzy semilinear sets

5.1 Basic examples

Let ψ\psi be a state of a quantum lattice system on n{\mathbb{R}}^{n}. By Lemma 3.2 the maps sending U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} to 𝔇al(U){\mathfrak{D}}_{al}(U) and 𝔇alψ(U){\mathfrak{D}}_{al}^{\psi}(U) are pre-cosheaves of Fréchet spaces on 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. We denote these pre-cosheaves by 𝔇al{\mathfrak{D}}_{al} and 𝔇alψ{\mathfrak{D}}_{al}^{\psi}. Putting together Lemma 2.1, Proposition 3.5, Corollary 3.1, and Proposition 3.8, we have

Theorem 5.1.

𝔇al{\mathfrak{D}}_{al} is a local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. If ψ\psi is gapped, then 𝔇alψ{\mathfrak{D}}_{al}^{\psi} is a local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}.

Our main object of study is the local Lie algebra 𝔇alψ{\mathfrak{D}}^{\psi}_{al} attached to a gapped state ψ\psi of a lattice system (Λ,{Vj}jΛ)(\Lambda,\{V_{j}\}_{j\in\Lambda}).

A much simpler example of a local Lie algebra arises from a finite-dimensional Lie algebra 𝔤{\mathfrak{g}} and any subset Λn\Lambda\subset{\mathbb{R}}^{n}. For any U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} let 𝔤al(U){{\mathfrak{g}}_{al}}(U) be the space of bounded functions UΛ𝔤U\cap\Lambda\rightarrow{\mathfrak{g}} which decay superpolynomially away from UU. The subscript “al” stands for “almost localized”. More precisely, 𝔤al(n){{\mathfrak{g}}_{al}}({\mathbb{R}}^{n}) is the space of bounded functions Λ𝔤\Lambda\rightarrow{\mathfrak{g}}, while 𝔤al(U){{\mathfrak{g}}_{al}}(U) is defined as a subspace of 𝔤al(n){{\mathfrak{g}}_{al}}({\mathbb{R}}^{n}) consisting of functions f:Λ𝔤f:\Lambda\rightarrow{\mathfrak{g}} such that the following seminorms are finite:

pk,U(f)=supjΛ|f(j)|(1+d(U,j))k,k.\displaystyle p_{k,U}(f)=\sup_{j\in\Lambda}|f(j)|(1+d(U,j))^{k},\quad k\in{\mathbb{N}}.
Proposition 5.1.

The assignment U𝔤al(U)U\mapsto{{\mathfrak{g}}_{al}}(U) is a local Lie algebra.

Proof.

It is easy to check that the assignment U𝔤al(U)U\mapsto{{\mathfrak{g}}_{al}}(U) it is a coflasque pre-cosheaf of Fréchet-Lie algebras satisfying Property I. The only thing left to check is that it is a cosheaf of vector spaces. By Lemma 2.1, it is sufficient to show that for any U,V𝒞𝒮nU,V\in{\mathcal{C}\mathcal{S}}_{n} the sequence

𝔤al(UV)𝔤al(U)𝔤al(V)𝔤al(UV)0\displaystyle{{\mathfrak{g}}_{al}}(U\wedge V)\rightarrow{{\mathfrak{g}}_{al}}(U)\oplus{{\mathfrak{g}}_{al}}(V)\rightarrow{{\mathfrak{g}}_{al}}(U\vee V)\rightarrow 0

is exact. To show exactness at 𝔤al(UV)=𝔤al(UV){{\mathfrak{g}}_{al}}(U\vee V)={{\mathfrak{g}}_{al}}(U\cup V), we note that every f𝔤al(UV)f\in{{\mathfrak{g}}_{al}}(U\cup V) can be written as a sum fU+fVf_{U}+f_{V}, where

fU(x)={f(x),d(x,U)<d(x,V),12f(x),d(x,U)=d(x,V),0,d(x,U)>d(x,V),\displaystyle f_{U}(x)=\begin{cases}f(x),&d(x,U)<d(x,V),\\ \frac{1}{2}f(x),&d(x,U)=d(x,V),\\ 0,&d(x,U)>d(x,V),\end{cases}

and fV(x)f_{V}(x) is defined by a similar expression with UU and VV exchanged. Using d(x,UV)=min(d(x,U),d(x,V))d(x,U\cup V)=\min(d(x,U),d(x,V)) it is easy to check that pk,U(fU)pk,UV(f)p_{k,U}(f_{U})\leq p_{k,U\cup V}(f) and pk,V(fV)pk,UV(f)p_{k,V}(f_{V})\leq p_{k,U\cup V}(f) for all kk, and thus fU𝔤al(U)f_{U}\in{{\mathfrak{g}}_{al}}(U) and fV𝔤al(V)f_{V}\in{{\mathfrak{g}}_{al}}(V).

To show exactness at 𝔤al(U)𝔤al(V){{\mathfrak{g}}_{al}}(U)\oplus{{\mathfrak{g}}_{al}}(V), suppose f𝔤al(U)𝔤al(V)f\in{{\mathfrak{g}}_{al}}(U)\cap{{\mathfrak{g}}_{al}}(V). From the proof of Prop. 4.2, there exist r0,CUV>0r\geq 0,C_{UV}>0 such that d(x,UrVr)CUVmax(d(x,U),d(x,V))d(x,U^{r}\cap V^{r})\leq C_{UV}\max(d(x,U),d(x,V)). We may assume that CUV1C_{UV}\geq 1, in which case for any xnx\in{\mathbb{R}}^{n} and any kk\in{\mathbb{N}}

(1+d(x,UrVr))kCUVk(1+max(d(x,U),d(x,V))k.\displaystyle(1+d(x,U^{r}\cap V^{r}))^{k}\leq C^{k}_{UV}(1+\max(d(x,U),d(x,V))^{k}.

Therefore for any f𝔤al(U)𝔤al(V)f\in{{\mathfrak{g}}_{al}}(U)\cap{{\mathfrak{g}}_{al}}(V) we have

pk,UrVr(f)CUVkmax(pk,U(f),pk,V(f)).\displaystyle p_{k,U^{r}\cap V^{r}}(f)\leq C^{k}_{UV}\max(p_{k,U}(f),p_{k,V}(f)).

Since by Corollary 4.2 one can take UV=UrVrU\wedge V=U^{r}\cap V^{r}, we conclude that f𝔤al(UV)f\in{{\mathfrak{g}}_{al}}(U\wedge V). ∎

5.2 Symmetries of lattice systems

If Λn\Lambda\subset{\mathbb{R}}^{n} is countable, it can be viewed as a lattice in the physical sense, and the local Lie algebra 𝔤al{{\mathfrak{g}}_{al}} over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} models infinitesimal gauge transformations of a lattice system on n{\mathbb{R}}^{n}.

Definition 5.1.

A local action of a compact Lie group on a lattice system (Λ,{Vj}jΛ)(\Lambda,\{V_{j}\}_{j\in\Lambda}) is a collection of homomorphisms ρj:GU(Vj)\rho_{j}:G\rightarrow U(V_{j}) such that the norms of the corresponding Lie algebra homomorphisms 𝔤B(Vj){\mathfrak{g}}\rightarrow B(V_{j}) are bounded uniformly in jj.

A local action of GG on (Λ,{Vj}jΛ)(\Lambda,\{V_{j}\}_{j\in\Lambda}) gives rise to a homomorphism from GG to the automorphism group of 𝒜{\mathscr{A}} via gjΛAdρj(g)g\mapsto\otimes_{j\in\Lambda}{\rm Ad}_{\rho_{j}(g)} which is smooth on 𝒜a{\mathscr{A}}_{a\ell}. The corresponding generator 𝖰{\mathsf{Q}} is a homomorphism from 𝔤{\mathfrak{g}} to the Lie algebra of derivations of 𝒜a{\mathscr{A}}_{a\ell} defined by

𝖰:(a,𝒜)jΛ[𝗊j(a),𝒜],a𝔤,𝒜𝒜a,\displaystyle{\mathsf{Q}}:(a,{\mathcal{A}})\mapsto\sum_{j\in\Lambda}[{\mathsf{q}}_{j}(a),{\mathcal{A}}],\quad a\in{\mathfrak{g}},\ {\mathcal{A}}\in{\mathscr{A}}_{a\ell},

where 𝗊j{\mathsf{q}}_{j} is the traceless part of the generator of ρj\rho_{j}. The image of 𝖰{\mathsf{Q}} lands in 𝔇al(n){\mathfrak{D}}_{al}({\mathbb{R}}^{n}), with 𝖰(a)Y=jΛ𝗊j(a)Y{\mathsf{Q}}(a)^{Y}=\sum_{j\in\Lambda}{\mathsf{q}}_{j}(a)^{Y}. We will regard 𝖰{\mathsf{Q}} as a homomorphism of Fréchet-Lie algebras 𝔤𝔇al(n){\mathfrak{g}}\rightarrow{\mathfrak{D}}_{al}({\mathbb{R}}^{n}).

This morphism of Fréchet-Lie algebras can be lifted to a morphism of local Lie algebras 𝔤al𝔇al{{\mathfrak{g}}_{al}}\rightarrow{\mathfrak{D}}_{al} over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. Indeed, for any U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} and any f𝔤al(U)f\in{{\mathfrak{g}}_{al}}(U) we let 𝖰(f){\mathsf{Q}}(f) be a derivation of 𝒜a{\mathscr{A}}_{a\ell} given by

𝖰(f)(𝒜)=jΛ[𝗊j(f(j)),𝒜],𝒜𝒜a.\displaystyle{\mathsf{Q}}(f)({\mathcal{A}})=\sum_{j\in\Lambda}[{\mathsf{q}}_{j}(f(j)),{\mathcal{A}}],\quad{\mathcal{A}}\in{\mathscr{A}}_{a\ell}.

It is easy to check that this derivation belongs to 𝔇al(U){\mathfrak{D}}_{al}(U) and that the above map is a continuous homomorphism 𝔤al(U)𝔇al(U){{\mathfrak{g}}_{al}}(U)\rightarrow{\mathfrak{D}}_{al}(U). The physical interpretation is that a local action of a compact Lie group on a quantum lattice system can be gauged on the infinitesimal level.

Definition 5.2.

A state ψ\psi of 𝒜{\mathscr{A}} is said to be invariant under a local action of a compact Lie group GG if it is invariant under the corresponding automorphisms of 𝒜{\mathscr{A}}.

Let ψ\psi be a gapped state of 𝒜{\mathscr{A}} invariant under a local action of a Lie group GG. In that case the image of 𝖰:𝔤𝔇al(n){\mathsf{Q}}:{\mathfrak{g}}\rightarrow{\mathfrak{D}}_{al}({\mathbb{R}}^{n}) lands in 𝔇alψ(n){\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n}). One may ask if this morphism of Fréchet-Lie algebras can be lifted to a morphism of local Lie algebras 𝔤al𝔇alψ{{\mathfrak{g}}_{al}}\rightarrow{\mathfrak{D}}^{\psi}_{al}. If this is the case, then the symmetry GG of ψ\psi can be gauged on the infinitesimal level. In the next section we construct obstructions for the existence of such a morphism of local Lie algebras and show that zero-temperature Hall conductance is an example of such an obstruction.

5.3 Equivariantization

As a preliminary step, for any GG-invariant gapped state ψ\psi we are going to define a graded local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} which is a GG-equivariant version of the local Lie algebra 𝔇alψ{\mathfrak{D}}^{\psi}_{al}. Recall that a graded local Lie algebra is a cosheaf of graded vector spaces that is a pre-cosheaf of graded Lie algebras satisfying the graded analogue of Property I. For example, if 𝔉{\mathfrak{F}} is a local Lie algebra and A=kAkA=\prod_{k\in{\mathbb{Z}}}A_{k} is a locally finite supercommutative graded algebra with finite-dimensional graded factors AkA_{k},101010A graded vector space is locally finite iff its graded components are finite-dimensional. then U𝔉(U)AU\mapsto{\mathfrak{F}}(U)\otimes A is a graded local Lie algebra. We denote it 𝔉A{\mathfrak{F}}\otimes A.

Fix a compact Lie group GG and a distributive pre-ordered set XX and consider the category of graded local Lie algebras over XX equipped with a GG-action. An object of this category is a graded local Lie algebra 𝔉{\mathfrak{F}} on which GG acts by automorphisms; morphisms are defined in an obvious manner. The first step is to define a functor 𝔉𝔉G{\mathfrak{F}}\mapsto{\mathfrak{F}}^{G} from this category to the category of graded local Lie algebras over XX such that 𝔉G(U){\mathfrak{F}}^{G}(U) is the Lie algebra of GG-invariant elements of 𝔉(U){\mathfrak{F}}(U). It is clear how to define such a functor for coflasque pre-cosheaves of Lie algebras with Property I, but the pre-cosheaf 𝔉G{\mathfrak{F}}^{G} will not be a cosheaf of vector spaces without further assumptions about 𝔉{\mathfrak{F}} and the GG-action.

Definition 5.3.

An action of GG on a pre-cosheaf of Fréchet spaces 𝔉{\mathfrak{F}} is smooth if for each UXU\in X the seminorms defining the topology of 𝔉(U){\mathfrak{F}}(U) can be chosen to be GG-invariant and the map G×𝔉(U)𝔉(U)G\times{\mathfrak{F}}(U)\rightarrow{\mathfrak{F}}(U) defining the action is smooth. An action of GG on a pre-cosheaf of graded Fréchet spaces is smooth is the GG-action on every graded component is smooth.

Proposition 5.2.

Let 𝔉{\mathfrak{F}} be a pre-cosheaf of graded Fréchet  spaces over XX equipped with a smooth action of a compact Lie group GG. The assignment U𝔉G(U)=(𝔉(U))GU\mapsto{\mathfrak{F}}^{G}(U)=\left({\mathfrak{F}}(U)\right)^{G} is a cosheaf of graded Fréchet spaces.

Proof.

It is sufficient to prove this in the ungraded case. We need to show that for any U,VXU,V\in X the sequence

𝔉G(UV)𝛼𝔉G(U)𝔉G(V)𝛽𝔉G(UV)0,\displaystyle{\mathfrak{F}}^{G}(U\wedge V)\xrightarrow{\alpha}{\mathfrak{F}}^{G}(U)\oplus{\mathfrak{F}}^{G}(V)\xrightarrow{\beta}{\mathfrak{F}}^{G}(U\vee V)\to 0,

is exact. To show exactness at the rightmost term, let 𝖥{\mathsf{F}} be a GG-invariant element of 𝔉(UV){\mathfrak{F}}(U\vee V) and let 𝖥U𝔉(U){\mathsf{F}}_{U}\in{\mathfrak{F}}(U) and 𝖥V𝔉(V){\mathsf{F}}_{V}\in{\mathfrak{F}}(V) be such that ιUV,U𝖥U+ιUV,V𝖥V=𝖥\iota_{U\cup V,U}{\mathsf{F}}_{U}+\iota_{U\cup V,V}{\mathsf{F}}_{V}={\mathsf{F}}. Averaging the action map G×𝔉(U)G\times{\mathfrak{F}}(U) over GG with the Haar measure gives a linear map hU:𝔉(U)𝔉G(U)h_{U}:{\mathfrak{F}}(U)\mapsto{\mathfrak{F}}^{G}(U) which is identity when restricted to 𝔉G(U){\mathfrak{F}}^{G}(U). The co-restriction morphisms intertwine these maps. Thus ιUV,UhU(𝖥U)+ιUV,VhV(𝖥V)=𝖥\iota_{U\cup V,U}\circ h_{U}({\mathsf{F}}_{U})+\iota_{U\cup V,V}\circ h_{V}({\mathsf{F}}_{V})={\mathsf{F}} which proves that β\beta is surjective. Exactness in the middle term is proved similarly. ∎

Remark 5.1.

For the proof to go through, it suffices to require the map G×𝔉(U)𝔉(U)G\times{\mathfrak{F}}(U)\rightarrow{\mathfrak{F}}(U) to be continuous. However, if it is smooth, 𝔉(U){\mathfrak{F}}(U) becomes a 𝔤{\mathfrak{g}}-module and all elements in 𝔉G(U)𝔉(U){\mathfrak{F}}^{G}(U)\subset{\mathfrak{F}}(U) are annihilated by the 𝔤{\mathfrak{g}}-action. We will use this later on.

Example 5.1.

If GG acts locally on a lattice system, the action of GG on the local Lie algebra 𝔇al{\mathfrak{D}}_{al} is smooth. If ψ\psi is a GG-invariant gapped state of such a lattice system, the action of GG on 𝔇alψ{\mathfrak{D}}^{\psi}_{al} is smooth.

Example 5.2.

Consider the local Lie algebra 𝔤al{{\mathfrak{g}}_{al}} over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} associated to a finite-dimensional Lie algebra 𝔤{\mathfrak{g}} (see Section 4). Assume that 𝔤{\mathfrak{g}} is the Lie algebra of a compact Lie group GG, then there is an obvious GG-action on 𝔤al{{\mathfrak{g}}_{al}}:

(gf)(j)=Adgf(j),gG,jΛ.\displaystyle(g\cdot f)(j)={\rm Ad}_{g}f(j),\quad g\in G,j\in\Lambda.

This GG-action is smooth.

Example 5.3.

If 𝔉{\mathfrak{F}} is a local Fréchet-Lie algebra with a smooth GG-action and A=kAkA=\prod_{k\in{\mathbb{Z}}}A_{k} is a locally-finite supercommutative graded algebra on which GG acts by automorphisms, then the GG-action on 𝔉A{\mathfrak{F}}\otimes A is smooth.

Corollary 5.1.

Let 𝔉{\mathfrak{F}} be a graded local Fréchet-Lie algebra over XX with a smooth GG-action. The functor of GG-invariant elements maps 𝔉{\mathfrak{F}} to a graded local FréchetLie algebra 𝔉G{\mathfrak{F}}^{G} over XX.

Definition 5.4.

Let 𝔉{\mathfrak{F}} be a local Fréchet-Lie algebra over XX with a smooth GG-action. The GG-equivariantization functor sends 𝔉{\mathfrak{F}} to the negatively-graded local Fréchet-Lie algebra 𝔉𝐆{\mathfrak{F}}^{\bf G} defined by

U(𝔉(U)k=1Symk(𝔤[2]))G.\displaystyle U\mapsto\left({\mathfrak{F}}(U)\otimes\prod_{k=1}^{\infty}{\rm Sym}^{k}({\mathfrak{g}}^{*}[-2])\right)^{G}.

In the cases of interest to us, the GG-action on a local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} is infinitesimally inner, in the sense that the 𝔤{\mathfrak{g}}-module structure mentioned in Remark 5.1 arises from a homomorphism ρ:𝔤𝔉(n)\rho:{\mathfrak{g}}\rightarrow{\mathfrak{F}}({\mathbb{R}}^{n}). In such a case, the graded local Lie algebra 𝔉𝐆{\mathfrak{F}}^{\bf G} has an extra bit of structure: a central element in 𝔉𝐆(n){\mathfrak{F}}^{\bf G}({\mathbb{R}}^{n}) of degree 2-2. This element is simply ρ\rho re-interpreted as an element of 𝔉(n)𝔤[2]{\mathfrak{F}}({\mathbb{R}}^{n})\otimes{\mathfrak{g}}^{*}[-2]. In the terminology of Section 2, 𝔉𝐆{\mathfrak{F}}^{\bf G} is a pointed graded local Fréchet-Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. It is easy to see that the GG-equivariantization functor respects this extra structure. That is, if f:𝔉𝔉f:{\mathfrak{F}}\rightarrow{\mathfrak{F}}^{\prime} is a morphism of local Fréchet-Lie algebras over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} commuting with infinitesimally inner smooth GG-actions on 𝔉{\mathfrak{F}} and 𝔉{\mathfrak{F}}^{\prime}, then f𝐆f^{\bf G} maps the central element ρ𝔉𝐆(n)2\rho\in{\mathfrak{F}}^{\bf G}({\mathbb{R}}^{n})_{-2} to the central element ρ𝔉𝐆(n)2\rho^{\prime}\in{{\mathfrak{F}}^{\prime}}^{\bf G}({\mathbb{R}}^{n})_{-2}.

Example 5.4.

Let ψ\psi be a GG-invariant gapped state of a quantum lattice system with a local GG-action which on the infinitesimal level is described by 𝖰:𝔤𝔇alψ(n){\mathsf{Q}}:{\mathfrak{g}}\rightarrow{\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n}). Consider the graded local Lie algebra 𝔇alψ{\mathfrak{D}}^{\psi}_{al} with its smooth GG-action (Example 5.1) and its GG-equivariantization 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al}. The distinguished central element of 𝔇alψ,𝐆(n){\mathfrak{D}}^{\psi,\bf G}_{al}({\mathbb{R}}^{n}) is 𝖰{\mathsf{Q}} regarded as an element of 𝔇alψ(n)𝔤[2]{\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n})\otimes{\mathfrak{g}}^{*}[-2].

Example 5.5.

Consider the graded local Lie algebra 𝔤al𝐆{{\mathfrak{g}}^{\bf G}_{al}} of Example 5.2. The degree 2-2 component of 𝔤al𝐆(n){{\mathfrak{g}}^{\bf G}_{al}}({\mathbb{R}}^{n}) is the space of GG-invariant bounded functions on Λ\Lambda with values in 𝔤𝔤{\mathfrak{g}}\otimes{\mathfrak{g}}^{*}. The distinguished central element is the constant function on Λ\Lambda which takes the value id𝔤{\rm id}_{\mathfrak{g}}.

Armed with the equivariantization functor, we can now explain our strategy for constructing obstructions for the existence of a local Lie algebra morphism 𝔤al𝔇alψ{{\mathfrak{g}}_{al}}\rightarrow{\mathfrak{D}}^{\psi}_{al} which lifts the Fréchet-Lie algebra morphism 𝖰:𝔤𝔇alψ(n){\mathsf{Q}}:{\mathfrak{g}}\rightarrow{\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n}). Suppose such a morphism ρ\rho exists. Applying the GG-equivariantization functor, we get a morphism of pointed negatively-graded local Fréchet-Lie algebras ρ𝐆:𝔤al𝐆𝔇alψ,𝐆\rho^{\bf G}:{{\mathfrak{g}}^{\bf G}_{al}}\rightarrow{\mathfrak{D}}^{\psi,\bf G}_{al}. For any CS cover 𝔘{\mathfrak{U}} of n{\mathbb{R}}^{n} an application of the Čech functor gives a morphism of acyclic pointed DGLAs

C+1aug(𝔘,n;𝔤al𝐆)C+1aug(𝔘,n;𝔇alψ,𝐆).C^{aug}_{\bullet+1}({\mathfrak{U}},{\mathbb{R}}^{n};{{\mathfrak{g}}^{\bf G}_{al}})\rightarrow C^{aug}_{\bullet+1}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,\bf G}_{al}).

Consequently, a obstruction for the existence of such a pointed DGLA morphism is an obstruction for the existence of ρ\rho. In the next section we use the twisted Maurer-Cartan equation for pointed DGLAs to construct such obstructions and identify them as topological invariants of gapped states defined in [11].

6 Topological invariants of GG-invariant gapped states

6.1 The commutator class

Let (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}) be a pointed DGLA, i.e. a DGLA with a distinguished central cycle 𝖡𝔐2{\mathsf{B}}\in{\mathfrak{M}}_{-2}. Assume it is a limit of an inverse system of nilpotent pointed DGLAs (𝔐N,𝖡N)({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), NN\in{\mathbb{N}}.

Definition 6.1.

The commutator DGLA, denoted by [𝔐,𝔐]¯{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}}, is defined to be the closure of the commutator subalgebra of 𝔐{\mathfrak{M}}. Namely, 𝗊[𝔐,𝔐]¯{\mathsf{q}}\in{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}} if and only if for any NN its projection to 𝔐N{\mathfrak{M}}_{N} is a finite linear combination of commutators in 𝔐N{\mathfrak{M}}_{N}.

Even if 𝔐{\mathfrak{M}} is acyclic, the DGLA [𝔐,𝔐]¯{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}} is not necessarily acyclic. We would like to construct an obstruction to finding an element 𝗉𝔐1{\mathsf{p}}\in{\mathfrak{M}}_{-1} which satisfies 𝗉=𝖡\partial{\mathsf{p}}={\mathsf{B}} and [𝗉,𝗉]=0[{\mathsf{p}},{\mathsf{p}}]=0.

Definition 6.2.

Let (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}) be a pointed DGLA. A 𝖡{\mathsf{B}}-twisted Maurer-Cartan element in 𝔐{\mathfrak{M}} is 𝗉𝔐1{\mathsf{p}}\in{\mathfrak{M}}_{-1} which satisfies

𝗉+12[𝗉,𝗉]=𝖡.\displaystyle\partial{\mathsf{p}}+\frac{1}{2}[{\mathsf{p}},{\mathsf{p}}]={\mathsf{B}}.

We will denote the set of 𝖡{\mathsf{B}}-twisted MC elements of 𝔐{\mathfrak{M}} by MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}). The map (𝔐,𝖡)MC(𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}})\mapsto{\rm MC}({\mathfrak{M}},{\mathsf{B}}) can be upgraded to a functor from the category of pointed DGLAs to the category of sets in an obvious way.

Let (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}) be pronilpotent pointed DGLA and 𝗉MC(𝔐,𝖡){\mathsf{p}}\in{\rm MC}({\mathfrak{M}},{\mathsf{B}}). Then [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] is a cycle of the DGLA [𝔐,𝔐]¯{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}}.

Proposition 6.1.

Let (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}) be an acyclic pointed DGLA which is a limit of an inverse system of nilpotent acyclic pointed DGLAs (𝔐N,𝖡N)({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), NN\in{\mathbb{N}}. Assume further that the structure morphisms rN,N1:𝔐N𝔐N1r_{N,N-1}:{\mathfrak{M}}_{N}\rightarrow{\mathfrak{M}}_{N-1} are surjective and 𝔧N=kerrN,N1{\mathfrak{j}}_{N}=\ker r_{N,N-1} is central in 𝔐N{\mathfrak{M}}_{N}. Finally, assume [𝔐1,𝔐1]=0[{\mathfrak{M}}_{1},{\mathfrak{M}}_{1}]=0. Then MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) is non-empty. Furthermore, the homology class of [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] in [𝔐,𝔐]¯{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}} for 𝗉MC(𝔐,𝖡){\mathsf{p}}\in{\rm MC}({\mathfrak{M}},{\mathsf{B}}) is independent of the choice of 𝗉{\mathsf{p}}.

A proof of this result can be found in Appendix B. It uses some results from deformation theory. We will call the homology class of [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] the commutator class of the acyclic pointed DGLA (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}), for lack of a better name, and denote it 𝐜𝐨𝐦(𝔐,𝖡){\mathbf{com}}({\mathfrak{M}},{\mathsf{B}}). The assignment (𝔐,𝖡)(H([𝔐,𝔐]¯),𝐜𝐨𝐦(𝔐,𝖡))({\mathfrak{M}},{\mathsf{B}})\mapsto(H_{\bullet}({\overline{[{\mathfrak{M}},{\mathfrak{M}}]}}),{\mathbf{com}}({\mathfrak{M}},{\mathsf{B}})) is a functor from the full sub-category of the category of acyclic pointed DGLAs satisfying the conditions of Proposition 6.1, to the category 𝗉𝖵𝖾𝖼𝗍{{\mathsf{pVect}}_{\mathbb{Z}}} of pointed graded vector spaces.

The conditions of Prop. 6.1 apply to any pointed DGLA which is the value of the Čech functor on a graded local Lie algebra 𝔉𝐆{\mathfrak{F}}^{\bf G} over some (X,)(X,\leq), where 𝔉{\mathfrak{F}} is a local Lie algebra equipped with a smooth infinitesimally inner GG-action. Indeed, along with the graded algebra A=k=1Symk𝔤[2]A=\prod_{k=1}^{\infty}{\rm Sym}^{k}{\mathfrak{g}}^{*}[-2] used in the construction of 𝔉𝐆{\mathfrak{F}}^{\bf G} one can consider its quotient by the ideal JN=k=N+1Symk𝔤[2]J_{N}=\prod_{k=N+1}^{\infty}{\rm Sym}^{k}{\mathfrak{g}}^{*}[-2]. Replacing AA with the nilpotent graded algebra A/JNA/J_{N} throughout, for any cover 𝔘{\mathfrak{U}} of the terminal object TT one gets a sequence of nilpotent acyclic pointed DGLAs labeled by NN\in{\mathbb{N}}. They assemble into an inverse system in an obvious manner, and its limit is the acyclic pointed DGLA C+1aug(𝔘,T;𝔉𝐆).C^{aug}_{\bullet+1}({\mathfrak{U}},T;{\mathfrak{F}}^{\bf G}). It is easy to see that the remaining conditions of Prop. 6.1 are also satisfied. In particular, Prop. 6.1 applies to the pointed DGLAs associated the graded local Lie algebras 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al} and 𝔤al𝐆{{\mathfrak{g}}^{\bf G}_{al}} and any CS cover of n{\mathbb{R}}^{n}.

Proposition 6.2.

For any pointed DGLA (𝔐(𝔘),𝖡)({\mathfrak{M}}({\mathfrak{U}}),{\mathsf{B}}) obtained, as above, from the Čech functor with respect to a cover 𝔘{\mathfrak{U}}, MC{\rm MC} is functorial in 𝔘{\mathfrak{U}}.

Proof.

Let 𝔙={Vj}jJ{\mathfrak{V}}=\{V_{j}\}_{j\in J} be a refinement of 𝔘{\mathfrak{U}}. By definition, there exists a map ϕ:JI\phi:J\rightarrow I with VjUϕ(j).V_{j}\subset U_{\phi(j)}. According to Prop. 2.2, there is a map from 𝔐(𝔙){\mathfrak{M}}({\mathfrak{V}}) to 𝔐(𝔘){\mathfrak{M}}({\mathfrak{U}}). As 𝖡{\mathsf{B}} is unaffected by ϕ\phi_{*}, we deduce from 𝗉𝔙MC(𝔐(𝔙),𝖡){\mathsf{p}}^{\mathfrak{V}}\in{\rm MC}({\mathfrak{M}}({\mathfrak{V}}),{\mathsf{B}}) that ϕ𝗉𝔙MC(𝔐(𝔘),𝖡)\phi_{*}{\mathsf{p}}^{\mathfrak{V}}\in{\rm MC}({\mathfrak{M}}({\mathfrak{U}}),{\mathsf{B}}). ∎

Example 6.1.

Let GG be a compact Lie group, 𝔤{\mathfrak{g}} be its Lie algebra, and 𝔤al{{\mathfrak{g}}_{al}} be the pointed local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} of Examples (5.2) and (5.5). The distinguished central cycle in C+1aug(𝔘,n,𝔤al𝐆)C^{aug}_{\bullet+1}({\mathfrak{U}},{\mathbb{R}}^{n},{{\mathfrak{g}}^{\bf G}_{al}}) is the constant function on Λ\Lambda with value id𝔤{\rm id}_{\mathfrak{g}} . Here id𝔤{\rm id}_{\mathfrak{g}} is regarded as a GG-invariant element of 𝔤al(n)𝔤[2]{{\mathfrak{g}}_{al}}({\mathbb{R}}^{n})\otimes{\mathfrak{g}}^{*}[-2]. For any cover 𝔘={Ui}iI{\mathfrak{U}}=\{U_{i}\}_{i\in I} one can construct a twisted MC element 𝗊{\mathsf{q}} as follows. Pick r>0r>0 large enough so that the interiors of UirU_{i}^{r}, iIi\in I, cover n{\mathbb{R}}^{n} in the usual sense and pick a partition of unity χi\chi_{i}, iIi\in I, subordinate to this open cover. For any jΛj\in\Lambda and any iIi\in I let 𝗊i(j)=χi(j)id𝔤{\mathsf{q}}_{i}(j)=\chi_{i}(j){\rm id}_{\mathfrak{g}}. It is easy to verify that this is a twisted MC element satisfying [𝗊,𝗊]=0[{\mathsf{q}},{\mathsf{q}}]=0. Thus the commutator class vanishes in this case.

Let GG be a compact Lie group, ψ\psi be a gapped GG-invariant state of a lattice system on n{\mathbb{R}}^{n}, and U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n}. The commutator class of the acyclic pointed DGLA (C+1aug(𝔘,n;𝔇alψ,𝐆),𝖰)(C^{aug}_{\bullet+1}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,\bf G}_{al}),{\mathsf{Q}}) is an obstruction for the existence of a morphism of local Lie algebras ρ:𝔤al𝔇alψ\rho:{{\mathfrak{g}}_{al}}\rightarrow{\mathfrak{D}}^{\psi}_{al} which is a lift of the Lie algebra morphism 𝖰:𝔤𝔇alψ(n){\mathsf{Q}}:{\mathfrak{g}}\rightarrow{\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n}). Indeed, as explained in Section 5.3, if ρ\rho exists, it induces a morphism of pointed DGLAs Caug(𝔘,n;𝔤al)Caug(𝔘,n;𝔇alψ,𝐆)C^{aug}_{\bullet}({\mathfrak{U}},{\mathbb{R}}^{n};{{\mathfrak{g}}_{al}})\rightarrow C^{aug}_{\bullet}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,\bf G}_{al}) which in turn induces a morphism in the category 𝗉𝖵𝖾𝖼𝗍{{\mathsf{pVect}}_{\mathbb{Z}}} which maps the commutator class of 𝔤al{{\mathfrak{g}}_{al}} to the commutator class of 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al}. Since the former class vanishes (see Example 6.1), so must the latter.

6.2 Construction of topological invariants

The commutator class defined in the previous section is not a useful invariant of a gapped GG-invariant state ψ\psi because it takes values in a set which itself depends on ψ\psi. It is also not invariant under LGAs (defined in Section 3.2) and thus is not an invariant of a gapped phase (see Remark 3.2). In this section we define a pairing between the commutator classes of 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al} and spherical CS cohomology classes of the sphere at infinity Sn1S^{n-1}, which takes values in the algebra of GG-invariant symmetric polynomials on 𝔤{\mathfrak{g}}. This gives a useful invariant of a gapped phase which is also an obstruction to promoting the global symmetry GG of the state to a local symmetry. Additionally, we will show that the invariant does not depends on the choice of the cover 𝔘{\mathfrak{U}} and thus is essentially unique.

Keeping in view further generalizations, we work over a sub-site 𝒞𝒮n/W{{{\mathcal{C}\mathcal{S}}_{n}}/W}111111This is the so-called overcategory whose objects are equipped with a morphism to WW. Due to coflasqueness, this amounts to a restriction to objects fuzzily included in WW. which depends on an arbitray W𝒞𝒮nW\in{\mathcal{C}\mathcal{S}}_{n}. Let W^𝒮𝒞𝒮n{\hat{W}}\in{\mathcal{S}\mathcal{C}\mathcal{S}}_{n} be the image of WW under the equivalence between 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} and 𝒮𝒞𝒮n{\mathcal{S}\mathcal{C}\mathcal{S}}_{n}. We continue to denote by 𝔇al{\mathfrak{D}}_{al} the local Fréchet-Lie algebra which maps U𝒞𝒮n/WU\in{{{\mathcal{C}\mathcal{S}}_{n}}/W} to 𝔇al{\mathfrak{D}}_{al}. For any cover 𝔘{\mathfrak{U}} of WW we denote by 𝔘^\hat{\mathfrak{U}} the corresponding cover of W^\hat{W}.

Definition 6.3.

For any 𝔘={Ui}iI{\mathfrak{U}}=\{U_{i}\}_{i\in I} covering W𝒞𝒮nW\in{\mathcal{C}\mathcal{S}}_{n}, any 𝖿Cp(𝔘,W;𝔇al){\mathsf{f}}\in C_{p}({\mathfrak{U}},W;{\mathfrak{D}}_{al}), and any βCˇp(𝔘^,W^;)\beta\in\check{C}^{p}(\hat{\mathfrak{U}},\hat{W};{\mathbb{R}}) define an evaluation

𝖿,β=sIpβs𝖿s𝔇al(W),\langle{\mathsf{f}},\beta\rangle=\sum_{s\in I_{p}}\beta_{s}{\mathsf{f}}_{s}\in{\mathfrak{D}}_{al}(W),

where Ik:={i0<i1<<ikIk+1}I_{k}:=\{i_{0}<i_{1}<\dots<i_{k}\in I^{k+1}\}. We adopt the convention that βs=0\beta_{s}=0 for jsUj\bigwedge_{j\in s}U_{j} bounded. The definition implicitly uses co-restriction of the coflasque cosheaf.

Lemma 6.1.

For any 𝔘{\mathfrak{U}} covering W𝒞𝒮nW\in{\mathcal{C}\mathcal{S}}_{n}, any cycle 𝖿Cp(𝔘,W;𝔇al){\mathsf{f}}\in C_{p}({\mathfrak{U}},W;{\mathfrak{D}}_{al}), and any cocycle βCˇp(𝔘^,W^;)\beta\in\check{C}^{p}(\hat{\mathfrak{U}},\hat{W};{\mathbb{R}}) the derivation 𝖿,β𝔇al(W)\langle{\mathsf{f}},\beta\rangle\in{\mathfrak{D}}_{al}(W) is inner.

Proof.

The chain complex

i<j𝔇al(UiUj)i𝔇al(Ui)𝔇al(W)0\ldots\rightarrow\bigoplus_{i<j}{\mathfrak{D}}_{al}(U_{i}\wedge U_{j})\rightarrow\bigoplus_{i}{\mathfrak{D}}_{al}(U_{i})\rightarrow{\mathfrak{D}}_{al}(W)\rightarrow 0 (23)

is acyclic. Since 𝖿{\mathsf{f}} is a cycle, 𝖿=𝗀{\mathsf{f}}=\partial{\mathsf{g}} for some 𝗀sIp+1𝔇al(jsUj).{\mathsf{g}}\in\bigoplus_{s\in I_{p+1}}{\mathfrak{D}}_{al}(\bigwedge_{j\in s}U_{j}). Thus

𝖿,β=\displaystyle\langle{\mathsf{f}},\beta\rangle= 𝗀,β\displaystyle\langle\partial{\mathsf{g}},\beta\rangle
=\displaystyle= 𝗀,β\displaystyle\langle{\mathsf{g}},\partial^{*}\beta\rangle
=\displaystyle= sIp+1𝗀s(β)s\displaystyle\sum_{s\in I_{p+1}}{\mathsf{g}}_{s}(\partial^{*}\beta)_{s}
=\displaystyle= sIp+1,withjsUjbounded𝗀s(β)s,\displaystyle\sum_{\begin{subarray}{c}s\in I_{p+1},\ \mathrm{with}\\ \bigwedge_{j\in s}U_{j}\ \mathrm{bounded}\end{subarray}}{\mathsf{g}}_{s}(\partial^{*}\beta)_{s},

where \partial^{*} is the adjoint of \partial defined by

(β)i0ip=k=0p(1)kβi0ik^ip.\displaystyle(\partial^{*}\beta)_{i_{0}\dots i_{p}}=\sum_{k=0}^{p}(-1)^{k}\beta_{i_{0}\dots\widehat{i_{k}}\dots i_{p}}.

The last expression is clearly inner because each 𝗀s{\mathsf{g}}_{s} is almost localized near some bounded region. The last equality depends on vanishing of

sIp+1,withjsUjunbounded𝗀s(β)s,\displaystyle\sum_{\begin{subarray}{c}s\in I_{p+1},\ \mathrm{with}\\ \bigwedge_{j\in s}U_{j}\ \mathrm{unbounded}\end{subarray}}{\mathsf{g}}_{s}(\partial^{*}\beta)_{s},

because when jsUj\bigwedge_{j\in s}U_{j} is unbounded (β)s=(δˇβ)s=0(\partial^{*}\beta)_{s}=(\check{\delta}\beta)_{s}=0 where δˇ\check{\delta} is the coboundary of Cˇp(𝔘^,W^;)\check{C}^{p}(\hat{\mathfrak{U}},\hat{W};{\mathbb{R}}). ∎

Remark 6.1.

The above lemma remains true if one replaces 𝔇al{\mathfrak{D}}_{al} with 𝔇alA{\mathfrak{D}}_{al}\otimes A where AA is any locally-finite graded vector space. Then the pairing 𝖿,β\langle{\mathsf{f}},\beta\rangle takes values in the space of inner derivations 𝔡al{{\mathfrak{d}}_{al}} (Definition 3.3) tensored with AA.

Remark 6.2.

The relation between \partial^{*} and δˇ\check{\delta} is as follows. They are equal for sIp+1s\in I_{p+1} with jsUj\bigwedge_{j\in s}U_{j} unbounded. For ss with jsUj\bigwedge_{j\in s}U_{j} bounded, jsU^j=\bigcap_{j\in s}\hat{U}_{j}=\varnothing and (β)s(\partial^{*}\beta)_{s} is in general nonzero whereas (δˇβ)s=0(\check{\delta}\beta)_{s}=0. This discrepancy arises because the Grothendieck topology on the poset of spherical CS sets induced by the coherent topology on 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n} does not allow the empty cover of the empty set. It is this discrepancy that makes the evaluation of DGLA cycles on spherical CS cocycles non-trivial.

If U,VU,V are closed semilinear sets with a bounded meet then any derivation in [𝔇al(U),𝔇al(V)][{\mathfrak{D}}_{al}(U),{\mathfrak{D}}_{al}(V)] is inner and thus has a well-defined average in any state on 𝒜{\mathscr{A}}. More generally, if AA is a locally-finite supercommutative graded algebra and 𝔉=𝔇alA{\mathfrak{F}}={\mathfrak{D}}_{al}\otimes A or some sub-algebra thereof, any element of [𝔉(U),𝔉(V)][{\mathfrak{F}}(U),{\mathfrak{F}}(V)] for UVU\wedge V bounded has a well-defined average in any state of 𝒜{\mathscr{A}}. The average takes values in AA. We will need the following more refined result:

Lemma 6.2.

Let UU and VV be closed semilinear sets such that UVU\wedge V is bounded. Let AA be a locally-finite supercommutative graded algebra and ψ\psi be a state. Then ψ([𝔇alψ(U)A,𝔇al(V)A])=0.\psi([{\mathfrak{D}}^{\psi}_{al}(U)\otimes A,{\mathfrak{D}}_{al}(V)\otimes A])=0.

Proof.

It suffices to consider the case A=A={\mathbb{R}}. Let 𝖥𝔇alψ(U){\mathsf{F}}\in{\mathfrak{D}}^{\psi}_{al}(U) and 𝖦𝔇al(V){\mathsf{G}}\in{\mathfrak{D}}_{al}(V). Propositions 3.2 and 3.6 give

ψ([𝖥,𝖦])=XYψ([𝖥X,𝖦Y])=Yψ(𝖥(𝖦Y))=0,\displaystyle\psi([{\mathsf{F}},{\mathsf{G}}])=\sum_{X\cap Y\neq\varnothing}\psi([{\mathsf{F}}^{X},{\mathsf{G}}^{Y}])=\sum_{Y}\psi({\mathsf{F}}({\mathsf{G}}^{Y}))=0,

where the last equality is because 𝖥{\mathsf{F}} preserves ψ\psi. ∎

In what follows we will apply Lemma 6.1 to the graded local Lie algebra 𝔇alψA{\mathfrak{D}}^{\psi}_{al}\otimes A (see Remark 6.1). The evaluation of cycles of Caug(𝔘,W;𝔇alψA)C^{aug}_{\bullet}({\mathfrak{U}},W;{\mathfrak{D}}^{\psi}_{al}\otimes A) on spherical CS cocycles has especially nice properties when DGLA cycles belong to the commutator DGLA

[C+1aug(𝔘,W;𝔇alψA),C+1aug(𝔘,W,𝔇alψA)]¯.\displaystyle\overline{\left[C^{aug}_{\bullet+1}\left({\mathfrak{U}},W;{\mathfrak{D}}^{\psi}_{al}\otimes A\right),C^{aug}_{\bullet+1}\left({\mathfrak{U}},W,{\mathfrak{D}}^{\psi}_{al}\otimes A\right)\right]}. (24)
Proposition 6.3.

If a cycle 𝖿{\mathsf{f}} lies in the commutator DGLA (24), then ψ(𝖿,β)\psi(\langle{\mathsf{f}},\beta\rangle) depends only on the cohomology class of β\beta and the homology class of 𝖿{\mathsf{f}} in the DGLA (24).

Proof.

Suppose 𝖿Cpaug(𝔘,W;𝔇alψA){\mathsf{f}}\in C^{aug}_{p}({\mathfrak{U}},W;{\mathfrak{D}}^{\psi}_{al}\otimes A) and 𝖿=𝗀{\mathsf{f}}=\partial{\mathsf{g}} for some 𝗀{\mathsf{g}} in the commutator DGLA. Then the vanishing of ψ(𝖿,β)A\psi(\langle{\mathsf{f}},\beta\rangle)\in A follows from the proof of Lemma 6.1 and Lemma 6.2. Thus it remains to show that ψ(𝖿,δˇb)=0\psi(\langle{\mathsf{f}},\check{\delta}b\rangle)=0 for any bCp1(𝔘,W^)b\in C^{p-1}({\mathfrak{U}},\hat{W}) if 𝖿{\mathsf{f}} is a cycle of the commutator DGLA. Indeed, since 𝖿,b=𝖿,b=0\langle{\mathsf{f}},\partial^{*}b\rangle=\langle\partial{\mathsf{f}},b\rangle=0, we have

𝖿,δˇb=sIp,withjsUjunbounded𝖿s(b)s=sIp,withjsUjbounded𝖿s(b)s,\displaystyle\langle{\mathsf{f}},\check{\delta}b\rangle=\sum_{\begin{subarray}{c}s\in I_{p},\ \mathrm{with}\\ \bigwedge_{j\in s}U_{j}\ \mathrm{unbounded}\end{subarray}}{\mathsf{f}}_{s}(\partial^{*}b)_{s}=-\sum_{\begin{subarray}{c}s\in I_{p},\ \mathrm{with}\\ \bigwedge_{j\in s}U_{j}\ \mathrm{bounded}\end{subarray}}{\mathsf{f}}_{s}(\partial^{*}b)_{s},

and the ψ\psi-average of each term in the latter sum vanishes by Lemma 6.2. ∎

The evaluations of cycles of the commutator DGLA on spherical CS cocycles can be used to construct topological invariants of GG-invariant gapped states on n{\mathbb{R}}^{n}. We set W=nW={\mathbb{R}}^{n}. The cycle we use as an input is the commutator class of C(𝔘,n;𝔇alψ,𝐆)C_{\bullet}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,\bf G}_{al}) defined using the inhomogeneous Maurer-Cartan equation (Section 6.1). Since 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al} is a sub-algebra of 𝔇alψA{\mathfrak{D}}^{\psi}_{al}\otimes A with A=k=1Symk𝔤[2]A=\prod_{k=1}^{\infty}{\rm Sym}^{k}{\mathfrak{g}}^{*}[-2], Prop. 6.3 applies to such cycles.

Let 𝔘{\mathfrak{U}} be a CS cover of n{\mathbb{R}}^{n}, and β\beta be a spherical CS cocycle of W^=Sn1\hat{W}=S^{n-1} of degree ll with respect to the cover 𝔘^\hat{\mathfrak{U}}. Let 𝗉{\mathsf{p}} be a 𝖰{\mathsf{Q}}-twisted MC element of C+1aug(𝔘,n;𝔇alψ,𝐆)C^{aug}_{\bullet+1}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,\bf G}_{al}). Evaluating [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] on β\beta and taking into account that the grading is shifted by 11 relative to the Čech grading, we get a GG-invariant element of 𝔡alψSymk(𝔤[2]){{\mathfrak{d}}^{\psi}_{al}}\otimes{\rm Sym}^{k}({\mathfrak{g}}^{*}[-2]) where 2k3=l2k-3=l. Since ψ\psi is GG-invariant, ψ([𝗉,𝗉],β)\psi(\langle[{\mathsf{p}},{\mathsf{p}}],\beta\rangle) is a GG-invariant element of Symk(𝔤[2]){\rm Sym}^{k}({\mathfrak{g}}^{*}[-2]). Equivalently, it is the value of a degree 33 linear function on cocycles of Cˇ(𝔘^,)\check{C}^{\bullet}(\hat{\mathfrak{U}},{\mathbb{R}}) valued in Sym𝔤[2]{\rm Sym}^{\bullet}{\mathfrak{g}}^{*}[2].

Theorem 6.1.

The function βψ([𝗉,𝗉],β)\beta\mapsto\psi(\langle[{\mathsf{p}},{\mathsf{p}}],\beta\rangle) depends only on ψ\psi and the class of (𝔘,β)({\mathfrak{U}},\beta) in the CS cohomology of Sn1S^{n-1}.

Proof.

For a fixed covering 𝔘{\mathfrak{U}}, Prop. 6.1 and 6.3 implies ψ([𝗉,𝗉],β)\psi(\langle[{\mathsf{p}},{\mathsf{p}}],\beta\rangle) is independent of 𝗉{\mathsf{p}}. It depends on β\beta solely through its cohomology class. It remains to show invariance under refinement of cover.

Let (𝔙,ϕ)({\mathfrak{V}},\phi) be a refinement of 𝔘{\mathfrak{U}} with VjUϕ(j).V_{j}\subset U_{\phi(j)}. Cocycles (𝔘,β)({\mathfrak{U}},\beta) and (𝔙,ϕβ)({\mathfrak{V}},\phi^{*}\beta), where (ϕβ)j0,,jk=βϕ(j0),,ϕ(jk)(\phi^{*}\beta)_{j_{0},\dots,j_{k}}=\beta_{\phi(j_{0}),\dots,\phi(j_{k})}, are in the same CS cohomology class. From Prop.6.1 there exists 𝖰{\mathsf{Q}}-twisted MC element 𝗉{\mathsf{p}} for the cover 𝔙{\mathfrak{V}}. Prop.6.2 then implies that ϕ𝗉\phi_{*}{\mathsf{p}} is a 𝖰{\mathsf{Q}}-twisted MC element 𝗉{\mathsf{p}} for the cover 𝔘{\mathfrak{U}}. An easy expansion shows

[ϕ𝗉,ϕ𝗉],β=[𝗉,𝗉],ϕβ.\langle[\phi_{*}{\mathsf{p}},\phi_{*}{\mathsf{p}}],\beta\rangle=\langle[{\mathsf{p}},{\mathsf{p}}],\phi^{*}\beta\rangle.

Since any 𝖰{\mathsf{Q}}-twisted MC element gives the same answer, we have shown that this contraction of interest depends only on the CS cohomology class of (𝔘,β)({\mathfrak{U}},\beta). ∎

By Proposition 4.7, the cohomology group HCSl(Sn1,)H^{l}_{CS}(S^{n-1},{\mathbb{R}}) is one-dimensioal for l=n1l=n-1 and vanishes otherwise. It is natural to take the class of (𝔘,β)({\mathfrak{U}},\beta) to be a generator of HCSn1(Sn1,)H^{n-1}_{CS}(S^{n-1},{\mathbb{Z}}). This generator is uniquely defined once an orientation of n1{\mathbb{R}}^{n-1} has been chosen. Thus for any even nn we obtain an invariant of gapped GG-invariant states on n{\mathbb{R}}^{n} taking values in GG-invariant polynomials on 𝔤{\mathfrak{g}} of degree (n+2)/2(n+2)/2, and this invariant changes sign when the orientation of n{\mathbb{R}}^{n} is changed. This is in agreement with Chern-Simons field theory.

6.3 The Hall conductance

For physics applications, the most important case is G=U(1)G=U(1) and n=2n=2. Let us specialize the construction of topological invariants to this case.

Let ψ\psi be a U(1)U(1)-invariant gapped state of a lattice system on n{\mathbb{R}}^{n}. The generator of the U(1)U(1) action is 𝖰𝔇alψ(n){\mathsf{Q}}\in{\mathfrak{D}}^{\psi}_{al}({\mathbb{R}}^{n}). Let 𝔇alψ,𝖰{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al} be the sub-algebra of U(1)U(1)-invariant elements of 𝔇alψ{\mathfrak{D}}^{\psi}_{al}. Since U(1)U(1) is connected, this the same as the sub-algebra of elements of 𝔇al{\mathfrak{D}}_{al} which commute with 𝖰{\mathsf{Q}}. More generally, for any U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} 𝔇alψ,𝖰(U)=𝔇al(U)𝔇alψ,𝖰{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}(U)={\mathfrak{D}}_{al}(U)\cap{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}. This is a local Lie algebra over 𝒞𝒮n{\mathcal{C}\mathcal{S}}_{n}. The graded local Lie algebra 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al} reduces in this case to 𝔇alψ,𝖰[[t]]{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}\otimes{\mathbb{R}}[[t]] where tt is a variable of degree 2-2.

Pick a cover 𝔘={Ui}iI{\mathfrak{U}}=\{U_{i}\}_{i\in I} of n{\mathbb{R}}^{n}. To find a solution 𝗉{\mathsf{p}} of the inhomogenenous Maurer-Cartan equation with 𝖡=𝖰t{\mathsf{B}}={\mathsf{Q}}\otimes t, we write 𝗉=k=1𝗉ktk{\mathsf{p}}=\sum_{k=1}^{\infty}{\mathsf{p}}_{k}\otimes t^{k}, where 𝗉kCk1(𝔘,n;𝔇alψ,𝖰){\mathsf{p}}_{k}\in C_{k-1}({\mathfrak{U}},{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}). To compute the topological invariant of a state on 2{\mathbb{R}}^{2} it is sufficient to solve for 𝗉1{\mathsf{p}}_{1}.

𝗉1{\mathsf{p}}_{1} is a solution of 𝗉1=𝖰\partial{\mathsf{p}}_{1}={\mathsf{Q}}. Explicitly, 𝗉1={𝖰i𝔇alψ,𝖰(Ui)}iI{\mathsf{p}}_{1}=\{{\mathsf{Q}}_{i}\in{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}(U_{i})\}_{i\in I} such that iI𝖰i=𝖰\sum_{i\in I}{\mathsf{Q}}_{i}={\mathsf{Q}}. Such 𝖰i{\mathsf{Q}}_{i} exist because 𝔇alψ,𝖰{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al} is a cosheaf. Then the component of the commutator class in C1(U,n;𝔇alψ,𝖰)C_{1}(U,{\mathbb{R}}^{n};{\mathfrak{D}}^{\psi,{\mathsf{Q}}}_{al}) is {[𝖰i,𝖰j]}i<j\{[{\mathsf{Q}}_{i},{\mathsf{Q}}_{j}]\}_{i<j}. The topological invariant of the state ψ\psi is obtained by evaluating it on a Čech 1-cocycle β\beta on S1S^{1} corresponding to the cover 𝔘^\hat{\mathfrak{U}} and then averaging the resulting inner derivation:

σ(β)=ψ(i<jβij[𝖰i,𝖰j]).\displaystyle\sigma(\beta)=\psi\left(\sum_{i<j}\beta_{ij}[{\mathsf{Q}}_{i},{\mathsf{Q}}_{j}]\right).

Note that averaging over ψ\psi must be performed after the summation over i,ji,j because [𝖰i,𝖰j][{\mathsf{Q}}_{i},{\mathsf{Q}}_{j}] is not an inner derivation, in general.

The simplest CS cover of 2{\mathbb{R}}^{2} which can represent a nontrivial class in HCS1(S1,)H^{1}_{CS}(S^{1},{\mathbb{R}}) is made of three cones with a common vertex. In this case the construction of the invariant reduces to that in [7, 11]. It was shown there that the resulting invariant is proportional to the zero-temperature Hall conductance as determined by the Kubo formula.

6.4 Topological invariants of gapped states on subsets of n{\mathbb{R}}^{n}

So far we assumed that Λ\Lambda is an arbitrary countable subset of n{\mathbb{R}}^{n} with a uniform O(rn)O(r^{n}) bound on the number of points in a ball of radius rr. Suppose ΛWϵ\Lambda\subset W^{\epsilon} for some CS set WW and some ϵ>0\epsilon>0. If this is the case, we will say the pair (Λ,{Vj}jΛ)(\Lambda,\{V_{j}\}_{j\in\Lambda}) describes a quantum lattice system on WW. Then for any 𝖥𝔇al{\mathsf{F}}\in{\mathfrak{D}}_{al} the component 𝖥X{\mathsf{F}}^{X} vanishes for any brick XX which does not intersect WϵΛW^{\epsilon}\cap\Lambda, and thus for any U𝒞𝒮nU\in{\mathcal{C}\mathcal{S}}_{n} we have 𝔇al(U)=𝔇al(UW){\mathfrak{D}}_{al}(U)={\mathfrak{D}}_{al}(U\wedge W) and 𝔇alψ(U)=𝔇alψ(UW){\mathfrak{D}}^{\psi}_{al}(U)={\mathfrak{D}}^{\psi}_{al}(U\wedge W). Thus the assignment U𝔇alψ(U)U\mapsto{\mathfrak{D}}^{\psi}_{al}(U) can be regarded as a local Lie algebra over the site 𝒞𝒮n/W{{{\mathcal{C}\mathcal{S}}_{n}}/W}.

In particular, if a compact Lie group GG acts locally on the quantum lattice system on WW and preserves a gapped state ψ\psi, the generator of the action 𝖰:𝔤𝔇alψ{\mathsf{Q}}:{\mathfrak{g}}\rightarrow{\mathfrak{D}}^{\psi}_{al} takes values in the sub-algebra 𝔇alψ(W){\mathfrak{D}}^{\psi}_{al}(W). Consequently 𝔇alψ,𝐆{\mathfrak{D}}^{\psi,\bf G}_{al} is a graded local Fréchet-Lie algebra over 𝒞𝒮n/W{{{\mathcal{C}\mathcal{S}}_{n}}/W} pointed by 𝖰𝔇alψ,𝐆(W){\mathsf{Q}}\in{\mathfrak{D}}^{\psi,\bf G}_{al}(W). Picking a cover 𝔘{\mathfrak{U}} of WW and applying the Čech functor gives an acyclic pointed DGLA whose commutator class can be evaluated on any CS cocycle β\beta of W^\hat{W} to give an inner derivation. Its ψ\psi-average is a GG-invariant polynomial on 𝔤{\mathfrak{g}} which depends only on the class of (𝔘,β)({\mathfrak{U}},\beta) in H(W^,)H^{\bullet}(\hat{W},{\mathbb{R}}). Thus a quantum lattice system on WW has a topological invariant which is a linear function of H(W^,)Sym𝔤[2]H^{\bullet}(\hat{W},{\mathbb{R}})\rightarrow{\rm Sym}^{\bullet}{\mathfrak{g}}^{*}[2]. It is easy to check that this linear function has degree 33.

For example, consider a U(1)U(1)-invariant gapped state of a quantum lattice system on W=2W={\mathbb{R}}^{2} affinely embedded in 3{\mathbb{R}}^{3}. The topological invariant defined in Section 6.2 is obtained by contracting with a 2-cocycle of S2S^{2} (the sphere at infinity) and vanishes for dimensional reasons. On the other hand, by contracting with the 1-cocycle of W^=S1\hat{W}=S^{1} one obtains the Hall conductance of this system.

For a more non-trivial example, for any n>2n>2 consider a finite graph ΓSn1\Gamma\subset S^{n-1} whose edges are geodesics and take WW to be the cone with base Γ\Gamma and apex at an arbitrary point of n{\mathbb{R}}^{n}. The invariants of U(1)U(1)-invariant gapped states of quantum lattice systems on WnW\subset{\mathbb{R}}^{n} are labeled by generators of the free abelian group H1(Γ,)H^{1}(\Gamma,{\mathbb{Z}}). This example goes beyond TQFT since WW need not be smooth or even locally Euclidean.

Appendix A 0-chains

In this section we use the results of [11] to derive the properties of LGAs which we used in Section 3.

A.1 0-chains on n{\mathbb{Z}}^{n}

First let us characterize 𝔇al(U){\mathfrak{D}}_{al}(U) in terms of 0-chains. A 0-chain on n{\mathbb{Z}}^{n} is an element 𝖺={𝖺j}jnjn𝔇al({j}){\mathsf{a}}=\{{\mathsf{a}}_{j}\}_{j\in{\mathbb{Z}}^{n}}\in\prod_{j\in{\mathbb{Z}}^{n}}{\mathfrak{D}}_{al}(\{j\}) such that

𝖺k:=supjn𝖺j{j},k<.\displaystyle\|{\mathsf{a}}\|_{k}:=\sup_{j\in{\mathbb{Z}}^{n}}\|{\mathsf{a}}_{j}\|_{\{j\},k}<\infty. (25)

We say a 0-chain 𝖺{\mathsf{a}} is supported on UU if 𝖺i=0{\mathsf{a}}_{i}=0 whenever iUi\notin U, and write C0(U)C_{0}(U) for the set of UU-supported 0-chains, endowed with the norms (25) for k0k\geq 0.

Proposition A.1.

Let UnU\subset{\mathbb{R}}^{n} be nonempty and let U1:={xn:d(x,U)1}U^{1}:=\{x\in{\mathbb{R}}^{n}:d(x,U)\leq 1\} be its 1-thickening.

  1. i)

    If 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U) then 𝖥=𝖿{\mathsf{F}}=\partial{\mathsf{f}} for a U1U^{1}-supported 0-chain 𝖿{\mathsf{f}} with 𝖿k2k𝖥U,k\|{\mathsf{f}}\|_{k}\leq 2^{k}\|{\mathsf{F}}\|_{U,k}

  2. ii)

    If 𝖿C0(U){\mathsf{f}}\in C_{0}(U), then for any Y𝔹nY\in{\mathbb{B}}_{n} the sum

    (𝖿)Y:=jnU𝖿jY\displaystyle(\partial{\mathsf{f}})^{Y}:=\sum_{j\in{\mathbb{Z}}^{n}\cap U}{\mathsf{f}}_{j}^{Y}

    is absolutely convergent and defines a map :C0(U)𝔇al(U)\partial:C_{0}(U)\to{\mathfrak{D}}_{al}(U) with 𝖿U,kC𝖿k+2n+1\|\partial{\mathsf{f}}\|_{U,k}\leq C\|{\mathsf{f}}\|_{k+2n+1}, where the constant C>0C>0 depends only on nn.

Lemma A.1.

For any nonempty U,YnU,Y\subset{\mathbb{R}}^{n} we have

1+d(Y,U1n)+diam(Y)2(1+d(Y,U)+diam(Y))\displaystyle 1+d(Y,U^{1}\cap{\mathbb{Z}}^{n})+\operatorname{diam}(Y)\leq 2(1+d(Y,U)+\operatorname{diam}(Y))
Proof.

Choose yYy\in Y and uUu\in U with d(y,u)=d(Y,U)d(y,u)=d(Y,U), and choose znz\in{\mathbb{Z}}^{n} with d(u,z)1d(u,z)\leq 1. Then since zU1nz\in U^{1}\cap{\mathbb{Z}}^{n} we have

d(Y,U1n)\displaystyle d(Y,U^{1}\cap{\mathbb{Z}}^{n}) diam(Y)+d(y,z)\displaystyle\leq\operatorname{diam}(Y)+d(y,z)
diam(Y)+d(y,u)+d(u,z)\displaystyle\leq\operatorname{diam}(Y)+d(y,u)+d(u,z)
diam(Y)+d(Y,U)+1,\displaystyle\leq\operatorname{diam}(Y)+d(Y,U)+1,

and the Lemma follows. ∎

Proof of Proposition A.1.

i)i). Suppose 𝖥𝔇al(U){\mathsf{F}}\in{\mathfrak{D}}_{al}(U). Choose any total order on U1nU^{1}\cap{\mathbb{Z}}^{n} and for every Y𝔹nY\in{\mathbb{B}}_{n} let j(Y)j^{*}(Y) be the closest point to YY in U1nU^{1}\cap{\mathbb{Z}}^{n}, using the total order as a tiebreaker. For every iΛi\in\Lambda, define

𝖿i:=Y𝔹nj(Y)=i𝖥Y.\displaystyle{\mathsf{f}}_{i}:=\sum_{\begin{subarray}{c}Y\in{\mathbb{B}}_{n}\\ j^{*}(Y)=i\end{subarray}}{\mathsf{F}}^{Y}. (26)

Then either 𝖿iY=0{\mathsf{f}}_{i}^{Y}=0 or d(Y,U1n)=d(Y,j)d(Y,U^{1}\cap{\mathbb{Z}}^{n})=d(Y,j) and 𝖿iY=𝖥Y\|{\mathsf{f}}_{i}^{Y}\|=\|{\mathsf{F}}^{Y}\|, and so

𝖿ik\displaystyle\|{\mathsf{f}}_{i}\|_{k} =supY𝔹n\{}(1+diam(Y)+d(Y,{i})k𝖿iY\displaystyle=\sup_{Y\in{\mathbb{B}}_{n}\backslash\{\varnothing\}}(1+\operatorname{diam}(Y)+d(Y,\{i\})^{k}\|{\mathsf{f}}_{i}^{Y}\|
=supY𝔹n\{}(1+diam(Y)+d(Y,U11))k𝖥Y,\displaystyle=\sup_{Y\in{\mathbb{B}}_{n}\backslash\{\varnothing\}}(1+\operatorname{diam}(Y)+d(Y,U^{1}\cap{\mathbb{Z}}^{1}))^{k}\|{\mathsf{F}}^{Y}\|,

which by Lemma A.1 is bounded by 2k𝖥U,k2^{k}\|{\mathsf{F}}\|_{U,k}.

ii)ii). Suppose that 𝖿{\mathsf{f}} is a UU-supported 0-chain. For any k0k\geq 0 we have

𝖿Y\displaystyle\|\partial{\mathsf{f}}^{Y}\| jnU𝖿jY\displaystyle\leq\sum_{j\in{\mathbb{Z}}^{n}\cap U}\|{\mathsf{f}}_{j}^{Y}\|
𝖿k+2n+1jnU(1+diam(Y)+d(j,Y))k2n1\displaystyle\leq\|{\mathsf{f}}\|_{k+2n+1}\sum_{j\in{\mathbb{Z}}^{n}\cap U}(1+\operatorname{diam}(Y)+d(j,Y))^{-k-2n-1}
𝖿k+2n+1(1+diam(Y)+d(U,Y))kjnU(1+diam(Y)+d(j,Y))2n1\displaystyle\leq\|{\mathsf{f}}\|_{k+2n+1}(1+\operatorname{diam}(Y)+d(U,Y))^{-k}\sum_{j\in{\mathbb{Z}}^{n}\cap U}(1+\operatorname{diam}(Y)+d(j,Y))^{-2n-1}
𝖿k+2n+1(1+diam(Y)+d(U,Y))k(1+diam(Y))njn(1+d(j,Y))n1\displaystyle\leq\|{\mathsf{f}}\|_{k+2n+1}(1+\operatorname{diam}(Y)+d(U,Y))^{-k}(1+\operatorname{diam}(Y))^{-n}\sum_{j\in{\mathbb{Z}}^{n}}(1+d(j,Y))^{-n-1}
𝖿k+2n+1(1+diam(Y)+d(U,Y))k(1+diam(Y))njn(1+d(j,Y))n1\displaystyle\leq\|{\mathsf{f}}\|_{k+2n+1}(1+\operatorname{diam}(Y)+d(U,Y))^{-k}(1+\operatorname{diam}(Y))^{-n}\sum_{j\in{\mathbb{Z}}^{n}}(1+d(j,Y))^{-n-1}

It is not hard to show that for any brick YY the quantity (1+diam(Y))njn(1+d(j,Y))n1(1+\operatorname{diam}(Y))^{-n}\sum_{j\in{\mathbb{Z}}^{n}}(1+d(j,Y))^{-n-1} is bounded by a constant CC depending only on nn, which shows 𝖿YC𝖿k+2n+1\|\partial{\mathsf{f}}^{Y}\|\leq C\|{\mathsf{f}}\|_{k+2n+1}. ∎

Proposition A.1 will allow us to apply the results of [11] on 0-chains. The results in [11] are phrased in terms of the norms

𝖺x,kKS:=supr>0(1+r)kinf𝖻𝔡(Br(x))𝖺𝖻\displaystyle\|{\mathsf{a}}\|^{KS}_{x,k}:=\sup_{r>0}(1+r)^{k}\inf_{{\mathsf{b}}\in{\mathfrak{d}}(B_{r}(x))}\|{\mathsf{a}}-{\mathsf{b}}\|

where 𝔡(Br(x)){\mathfrak{d}}(B_{r}(x)) is the set of traceless anti-hermitian operators strictly localized on the ball of radius rr around xnx\in{\mathbb{R}}^{n}. To apply their results we prove the equivalence of these norms:

Lemma A.2.

For any xnx\in{\mathbb{Z}}^{n} and k>0k>0, the norms x,kKS\|\cdot\|^{KS}_{x,k} and {x},k\|\cdot\|_{\{x\},k} obey

𝖺x,kKS\displaystyle\|{\mathsf{a}}\|^{KS}_{x,k} C𝖺{x},k+2n+2\displaystyle\leq C\|{\mathsf{a}}\|_{\{x\},k+2n+2} (27)
𝖺{x},k\displaystyle\|{\mathsf{a}}\|_{\{x\},k} 8kC𝖺x,kKS\displaystyle\leq 8^{k}C^{\prime}\|{\mathsf{a}}\|^{KS}_{x,k} (28)

where C,CC,C^{\prime} are constants depending only on kk.

Proof.

Suppose 𝖺{x},k+2n+1<\|{\mathsf{a}}\|_{\{x\},k+2n+1}<\infty and let r>0r>0. Define 𝖻:=XBr(x)𝖺X{\mathsf{b}}:=\sum_{X\subset B_{r}(x)}{\mathsf{a}}^{X}. Then

𝖻𝖺𝖺{x},k+2n+2XBr(x)(1+diam(X)+d(x,X))k2n2\displaystyle\|{\mathsf{b}}-{\mathsf{a}}\|\leq\|{\mathsf{a}}\|_{\{x\},k+2n+2}\sum_{X\nsubseteq B_{r}(x)}(1+\operatorname{diam}(X)+d(x,X))^{-k-2n-2}

Since XBr(x)X\nsubseteq B_{r}(x) means diam(X)+d(x,X)r\operatorname{diam}(X)+d(x,X)\geq r, we continue this as follows

(1+r)k𝖺{x},k+2n+2XBr(x)(1+diam(X)+d(x,X))2n2\displaystyle\leq(1+r)^{-k}\|{\mathsf{a}}\|_{\{x\},k+2n+2}\sum_{X\nsubseteq B_{r}(x)}(1+\operatorname{diam}(X)+d(x,X))^{-2n-2}
C(1+r)k𝖺{x},k+2n+2\displaystyle\leq C(1+r)^{-k}\|{\mathsf{a}}\|_{\{x\},k+2n+2}

where in the last line we used Lemma 3.1. This proves (27). To prove (28) suppose 𝖺x,kKS<\|{\mathsf{a}}\|^{KS}_{x,k}<\infty and let XX be any brick. Set r:=(diam(X)+d(x,X))/4r:=\lfloor(\operatorname{diam}(X)+d(x,X))/4\rfloor. Then XBr(x)X\nsubseteq B_{r}(x). Indeed, if XBr(x)X\subseteq B_{r}(x) then d(x,X)rd(x,X)\leq r and diam(X)2r\operatorname{diam}(X)\leq 2r, implying diam(X)+d(x,X)3r\operatorname{diam}(X)+d(x,X)\leq 3r, which is impossible. Choose 𝖻𝔡(Br(x)){\mathsf{b}}\in{\mathfrak{d}}(B_{r}(x)) with 𝖺𝖻(1+r)k𝖺x,kKS\|{\mathsf{a}}-{\mathsf{b}}\|\leq(1+r)^{-k}\|{\mathsf{a}}\|_{x,k}^{KS}. Since XBr(x)X\nsubseteq B_{r}(x) we have 𝖻X=0{\mathsf{b}}^{X}=0, and so

𝖺X\displaystyle\|{\mathsf{a}}^{X}\| =(𝖺𝖻)X\displaystyle=\|({\mathsf{a}}-{\mathsf{b}})^{X}\|
4n𝖺𝖻\displaystyle\leq 4^{n}\|{\mathsf{a}}-{\mathsf{b}}\|
4n𝖺x,kKS(1+r)k\displaystyle\leq 4^{n}\|{\mathsf{a}}\|_{x,k}^{KS}(1+r)^{-k}
4n+2k𝖺x,kKS(1+diam(X)+d(x,X))k,\displaystyle\leq 4^{n+2k}\|{\mathsf{a}}\|_{x,k}^{KS}(1+\operatorname{diam}(X)+d(x,X))^{-k},

where in the second line we used [11, Proposition C.1]. ∎

A.2 Proof of Lemma 3.7

We begin by describing the construction of 𝒥\mathcal{J} and 𝒦\mathcal{K}. Suppose ψ\psi is gapped with Hamiltonian 𝖧{\mathsf{H}} and gap Δ\Delta, and write τt\tau_{t} for the one-parameter family of LGAs obtained by exponentiating 𝖧{\mathsf{H}}. There exists121212See for instance Lemma 2.3 in [9]. a function wΔ:w_{\Delta}:{\mathbb{R}}\to{\mathbb{R}} such that wΔ(t)=O(|t|)w_{\Delta}(t)=O(|t|^{-\infty}), and the Fourier transform131313We use the convention f^(ξ)=f(t)eiωt𝑑t\widehat{f}(\xi)=\int f(t)e^{-i\omega t}dt. wΔ^\widehat{w_{\Delta}} is supported in the interval [Δ/2,Δ/2][-\Delta/2,\Delta/2] and satisfies wΔ^(0)=1\widehat{w_{\Delta}}(0)=1. Let WΔW_{\Delta} be the odd function which on the positive real line is given by WΔ(|t|)=|t|wΔ(u)𝑑uW_{\Delta}(|t|)=-\int_{|t|}^{\infty}w_{\Delta}(u)du. Then we define \mathcal{F} and 𝒢\mathcal{G} as the following integral transforms:

𝒥(𝖥):=wΔ(t)τt(𝖥)𝑑t,\displaystyle\mathcal{J}({\mathsf{F}}):=\int w_{\Delta}(t)\tau_{t}({\mathsf{F}})dt,
𝒦(𝖥):=WΔ(t)τt(𝖥)𝑑t.\displaystyle\mathcal{K}({\mathsf{F}}):=\int W_{\Delta}(t)\tau_{t}({\mathsf{F}})dt.
Proof of Lemma 3.7.

𝒥\mathcal{J} and 𝒦\mathcal{K} correspond to 𝒥𝖧,wΔ\mathscr{J}_{{\mathsf{H}},w_{\Delta}} and 𝒥𝖧,WΔ\mathscr{J}_{{\mathsf{H}},W_{\Delta}} in Section 4.1 of [11]. Part i)i) follows from the definition of 𝒦\mathcal{K} and the fact that 𝖧{\mathsf{H}} preserves ψ\psi. Part ii)ii) follows from Proposition A.1 and Lemma A.2, together with [11, Lemma F.1] (specifically line (177) therein). Part iii)iii) is [11, line (72)]. ∎

Appendix B Inhomogeneous Maurer-Cartan equation

Let (𝔐,𝖡)({\mathfrak{M}},{\mathsf{B}}) be a pointed pronilpotent DLGA which is a limit of an inverse system of nilpotent pointed DGLAs (𝔐N,𝖡N)({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), NN\in{\mathbb{N}}. The set MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) of 𝖡{\mathsf{B}}-twisted MC elements has an additional equivalence relation. This section revolves around this additional structure leading eventually to a proof of Prop. 6.1.

We have morphisms rN,K:𝔐N𝔐Kr_{N,K}:{\mathfrak{M}}_{N}\rightarrow{\mathfrak{M}}_{K} for all N>KN>K and the DGLA 𝔐{\mathfrak{M}} is the inverse limit of the corresponding system of DGLAs. For any NN\in{\mathbb{N}} let rN:𝔐𝔐Nr_{N}:{\mathfrak{M}}\rightarrow{\mathfrak{M}}_{N} be the natural projection. Let 𝔧N=kerrN,N1{\mathfrak{j}}_{N}=\ker r_{N,N-1}. The sets of 𝖡N{\mathsf{B}}_{N}-twisted MC elements of 𝔐N{\mathfrak{M}}_{N} will be denoted MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}).

A 𝖡{\mathsf{B}}-twisted MC element 𝗉{\mathsf{p}} gives rise to a degree 1-1 derivation 𝗉=+ad𝗉\partial_{\mathsf{p}}=\partial+{\rm ad}_{\mathsf{p}} of 𝔐{\mathfrak{M}} which squares to zero (twisted differential).

Lemma B.1.

If 𝗉MC(𝔐N,𝖡N){\mathsf{p}}\in{\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), then rN,K(𝗉)MC(𝔐K,𝖡K)r_{N,K}({\mathsf{p}})\in{\rm MC}({\mathfrak{M}}_{K},{\mathsf{B}}_{K}) for all K<NK<N. Further, 𝗉MC(𝔐,𝖡){\mathsf{p}}\in{\rm MC}({\mathfrak{M}},{\mathsf{B}}) iff 𝗉N=rN(𝗉)MC(𝔐N,𝖡N)N{\mathsf{p}}^{N}=r_{N}({\mathsf{p}})\in{\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N})\ \forall N\in{\mathbb{N}}.

Proof.

Straightforward. ∎

Thus MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) is the inverse limit of the system of sets MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), NN\in{\mathbb{N}}.

We are going to define an equivalence relations on MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) and MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}) for all NN. This is done in the same way as for the ordinary (homogeneous) Maurer-Cartan equation [29, 30].

First, 𝔐N,0{\mathfrak{M}}_{N,0} is a nilpotent Lie algebra, so there is a well-defined nilpotent Lie group exp(𝔐N,0)\exp({\mathfrak{M}}_{N,0}) with the group law given by the Campbell-Baker-Hausdorff formula. Similarly, 𝔐0{\mathfrak{M}}_{0} is pronilpotent (i.e. is an inverse limit of a system of nilpotent Lie algebras), so the CBH formula defines a group exp(𝔐0)\exp({\mathfrak{M}}_{0}).

Second, there are Lie algebra homomorphisms from 𝔐0{\mathfrak{M}}_{0} (resp. 𝔐N,0{\mathfrak{M}}_{N,0}) to the Lie algebras of affine-linear vector fields on 𝔐1{\mathfrak{M}}_{-1} (resp. 𝔐N,1{\mathfrak{M}}_{N,-1}). This homomorphism maps 𝖺𝔐0{\mathsf{a}}\in{\mathfrak{M}}_{0} or 𝔐N,0{\mathfrak{M}}_{N,0} to the affine-linear vector field

ξ𝖺(𝗉)=[𝖺,𝗉]𝖺,\displaystyle\xi_{\mathsf{a}}({\mathsf{p}})=[{\mathsf{a}},{\mathsf{p}}]-\partial{\mathsf{a}},

where 𝗉𝔐1{\mathsf{p}}\in{\mathfrak{M}}_{-1} or 𝔐N,1{\mathfrak{M}}_{N,-1}. Here we used the identification of the space of affine-linear vector fields on a vector space VV with the space of affine-linear maps VVV\rightarrow V. These homomorphisms exponentiate to actions of the groups exp(𝔐0)\exp({\mathfrak{M}}_{0}) and exp(𝔐N,0)\exp({\mathfrak{M}}_{N,0}) on 𝔐1{\mathfrak{M}}_{-1} and 𝔐N,1{\mathfrak{M}}_{N,-1} by affine-linear transformations. Explicitly, the actions are given by [29, 30]:

𝗉exp(𝖺)𝗉=exp(ad𝖺)(𝗉)+1exp(ad𝖺)ad𝖺(𝖺).\displaystyle{\mathsf{p}}\mapsto\exp({\mathsf{a}})*{\mathsf{p}}=\exp({\rm ad}_{\mathsf{a}})({\mathsf{p}})+\frac{1-\exp({\rm ad}_{\mathsf{a}})}{{\rm ad}_{\mathsf{a}}}(\partial{\mathsf{a}}). (29)
Lemma B.2.

The actions of exp(𝔐0)\exp({\mathfrak{M}}_{0}) on 𝔐1{\mathfrak{M}}_{-1} (resp. exp(𝔐N,0)\exp({\mathfrak{M}}_{N,0}) on 𝔐N,1{\mathfrak{M}}_{N,-1}) preserve the sets MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) (resp. MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N})).

Proof.

The proof in [29], Section 1.3, applies just as well in the inhomogeneous case. ∎

We say that elements p1,p2p_{1},p_{2} of MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) or MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}) are equivalent if they belong to the same orbit of these actions.

Remark B.1.

By analogy with the homogeneous case, one can define a 𝖡{\mathsf{B}}-twisted Deligne groupoid as the transformation groupoid for the action of exp(𝔐0)\exp({\mathfrak{M}}_{0}) on MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}). Similarly, one can define ”reduced” Deligne groupoids for every NN\in{\mathbb{N}}.

We observe an easy but useful lemma.

Lemma B.3.

If 𝗉iMC(𝔐N,𝖡N){\mathsf{p}}_{i}\in{\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), i=1,2i=1,2, are equivalent, then rN,K(𝗉1)r_{N,K}({\mathsf{p}}_{1}) is equivalent to rN,K(𝗉2)r_{N,K}({\mathsf{p}}_{2}) for all K<NK<N.

Now come the interesting statements. Assume from now on that the DGLAs 𝔐N{\mathfrak{M}}_{N} and 𝔐{\mathfrak{M}} are acyclic, that 𝔧N𝔐N{\mathfrak{j}}_{N}\subset{\mathfrak{M}}_{N} is central for all N>1N>1, and that the morphisms rN,N1r_{N,N-1} are surjective for all N>1N>1.

Lemma B.4.

With the above assumptions, the set MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) is non-empty if and only if MC(𝔐1,𝖡1){\rm MC}({\mathfrak{M}}_{1},{\mathsf{B}}_{1}) is non-empty.

Proof.

The only if statement follows from Lemma B.1. To prove the if direction, we use induction on NN. Assume MC(𝔐N1,𝖡N1){\rm MC}({\mathfrak{M}}_{N-1},{\mathsf{B}}_{N-1}) is non-empty. Let 𝗉N1MC(𝔐N1,𝖡N1){\mathsf{p}}_{N-1}\in{\rm MC}({\mathfrak{M}}_{N-1},{\mathsf{B}}_{N-1}). Pick 𝗉~NrN,N11(𝗉N1)\tilde{\mathsf{p}}_{N}\in r_{N,N-1}^{-1}({\mathsf{p}}_{N-1}). Since 𝗉N1{\mathsf{p}}_{N-1} is a 𝖡N1{\mathsf{B}}_{N-1}-twisted MC element and rN,N1(𝖡N)=𝖡N1r_{N,N-1}({\mathsf{B}}_{N})={\mathsf{B}}_{N-1}, we must have

𝗉~N+12[𝗉~N,𝗉~N]=𝖡N+𝗊N\displaystyle\partial\tilde{\mathsf{p}}_{N}+\frac{1}{2}[\tilde{\mathsf{p}}_{N},\tilde{\mathsf{p}}_{N}]={\mathsf{B}}_{N}+{\mathsf{q}}_{N}

for some 𝗊N𝔧N{\mathsf{q}}_{N}\in{\mathfrak{j}}_{N}. We look for a solution of the 𝖡N{\mathsf{B}}_{N}-twisted MC equation of the form 𝗉N=𝗉~N+𝖻N{\mathsf{p}}_{N}=\tilde{\mathsf{p}}_{N}+{\mathsf{b}}_{N} where 𝖻N𝔧N{\mathsf{b}}_{N}\in{\mathfrak{j}}_{N}. Taking into account that 𝔧N{\mathfrak{j}}_{N} is central, the inhomogeneous MC equation reduces to 𝖻N=𝗊N\partial{\mathsf{b}}_{N}=-{\mathsf{q}}_{N}. Since by assumption 𝔧N,𝔐N{\mathfrak{j}}_{N},{\mathfrak{M}}_{N}, and 𝔐N1{\mathfrak{M}}_{N-1} form a short exact sequence and the latter two are acyclic, so is 𝔧N{\mathfrak{j}}_{N}. Hence such a 𝖻N{\mathsf{b}}_{N} exists. This completes the inductive step proving that MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N})\neq\varnothing for all NN. Moreover, we also proved that the morphisms MC(𝔐N,𝖡N)MC(𝔐N1,𝖡N1){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N})\rightarrow{\rm MC}({\mathfrak{M}}_{N-1},{\mathsf{B}}_{N-1}) are surjective. Therefore by Lemma B.1 MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) is non-empty. ∎

The above lemma proves the first part of Prop. 6.1. Indeed when [𝔐1,𝔐1]=0[{\mathfrak{M}}_{1},{\mathfrak{M}}_{1}]=0, MC(𝔐1,𝖡1){\rm MC}({\mathfrak{M}}_{1},{\mathsf{B}}_{1})\neq\varnothing is implied by acyclicity. Next we show that with the above assumption on 𝔐N{\mathfrak{M}}_{N} and 𝔐{\mathfrak{M}} all 𝖡{\mathsf{B}}-twisted MC elements are equivalent.

Lemma B.5.

Suppose 𝖺,𝖻𝔐N,0{\mathsf{a}},{\mathsf{b}}\in{\mathfrak{M}}_{N,0} and 𝖺𝔧N{\mathsf{a}}\in{\mathfrak{j}}_{N}. Then for any 𝗉𝔐N,1{\mathsf{p}}\in{\mathfrak{M}}_{N,-1} one has

exp(𝖻+𝖺)(𝗉)=exp(𝖻)(𝗉)𝖺.\displaystyle\exp({\mathsf{b}}+{\mathsf{a}})({\mathsf{p}})=\exp({\mathsf{b}})({\mathsf{p}})-\partial{\mathsf{a}}.
Proof.

See [29], Lemma 2.8. ∎

Lemma B.6.

Let 𝗉iMC(𝔐N,𝖡N){\mathsf{p}}_{i}\in{\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), i=1,2i=1,2 such that rN,N1(𝗉1)=rN,N1(𝗉2)r_{N,N-1}({\mathsf{p}}_{1})=r_{N,N-1}({\mathsf{p}}_{2}). Then 𝗉1{\mathsf{p}}_{1} and 𝗉2{\mathsf{p}}_{2} are equivalent.

Proof.

Let 𝗊=𝗉2𝗉1𝔐N,1{\mathsf{q}}={\mathsf{p}}_{2}-{\mathsf{p}}_{1}\in{\mathfrak{M}}_{N,-1}. By assumption, 𝗊𝔧N{\mathsf{q}}\in{\mathfrak{j}}_{N}. Moreover, 𝗊=0\partial{\mathsf{q}}=0. Indeed:

𝗊=12[𝗉2,𝗉2]12[𝗉1,𝗉1]=[𝗊,𝗉1]+12[𝗊,𝗊]=0.\displaystyle\partial{\mathsf{q}}=\frac{1}{2}[{\mathsf{p}}_{2},{\mathsf{p}}_{2}]-\frac{1}{2}[{\mathsf{p}}_{1},{\mathsf{p}}_{1}]=[{\mathsf{q}},{\mathsf{p}}_{1}]+\frac{1}{2}[{\mathsf{q}},{\mathsf{q}}]=0.

By acyclicity of 𝔧N{\mathfrak{j}}_{N}, we have 𝗊=𝖺{\mathsf{q}}=\partial{\mathsf{a}} for some 𝖺𝔧N,0{\mathsf{a}}\in{\mathfrak{j}}_{N,0}. Then Lemma B.5 implies that exp(𝖺)exp(𝔐N,0)\exp({\mathsf{a}})\in\exp({\mathfrak{M}}_{N,0}) maps 𝗉2{\mathsf{p}}_{2} to 𝗉1{\mathsf{p}}_{1}. ∎

Lemma B.7.

Let 𝗉~iMC(𝔐N,𝖡N)\tilde{{\mathsf{p}}}_{i}\in{\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}), i=1,2,i=1,2, be such that 𝗉1=rN,N1(𝗉~1){\mathsf{p}}_{1}=r_{N,N-1}(\tilde{{\mathsf{p}}}_{1}) and 𝗉2=rN,N1(𝗉~2){\mathsf{p}}_{2}=r_{N,N-1}(\tilde{{\mathsf{p}}}_{2}) are equivalent. Then 𝗉~1\tilde{{\mathsf{p}}}_{1} and 𝗉~2\tilde{{\mathsf{p}}}_{2} are equivalent.

Proof.

Let 𝖺𝔐N1,0{\mathsf{a}}\in{\mathfrak{M}}_{N-1,0} be an equivalence between 𝗉1{\mathsf{p}}_{1} and 𝗉2{\mathsf{p}}_{2}, i.e. exp(𝖺)𝗉2=𝗉1.\exp({\mathsf{a}})*{\mathsf{p}}_{2}={\mathsf{p}}_{1}. Let 𝖺~𝔐N,0\tilde{{\mathsf{a}}}\in{\mathfrak{M}}_{N,0} be any lift of 𝖺{\mathsf{a}}. Then rN,N1(exp(𝖺~)𝗉~2)=rN,N1(𝗉~1)r_{N,N-1}(\exp(\tilde{{\mathsf{a}}})*\tilde{{\mathsf{p}}}_{2})=r_{N,N-1}(\tilde{{\mathsf{p}}}_{1}). By Lemma B.6, exp(𝖺~)𝗉~2\exp(\tilde{{\mathsf{a}}})*\tilde{{\mathsf{p}}}_{2} is equivalent to 𝗉~1\tilde{{\mathsf{p}}}_{1}, therefore 𝗉~2\tilde{{\mathsf{p}}}_{2} is equivalent to 𝗉~1\tilde{{\mathsf{p}}}_{1}. ∎

Proposition B.1.

For any NN\in{\mathbb{N}} all elements of MC(𝔐N,𝖡N){\rm MC}({\mathfrak{M}}_{N},{\mathsf{B}}_{N}) are equivalent. All elements of MC(𝔐,𝖡){\rm MC}({\mathfrak{M}},{\mathsf{B}}) are equivalent.

The first statement is proved by induction on NN, where the inductive step is Lemma B.7. The second statement follows by passing to the inverse limit in NN.

Theorem B.1.

Let 𝗉MC(𝔐,𝖡){\mathsf{p}}\in{\rm MC}({\mathfrak{M}},{\mathsf{B}}). Then the homology class of [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] in H([𝔐,𝔐])H_{\bullet}([{\mathfrak{M}},{\mathfrak{M}}]) is independent of the choice of 𝗉{\mathsf{p}}.

Proof.

Let 𝗉,𝗉MC(𝔐,𝖡){\mathsf{p}},{\mathsf{p}}^{\prime}\in{\rm MC}({\mathfrak{M}},{\mathsf{B}}) and let 𝖺𝔐0{\mathsf{a}}\in{\mathfrak{M}}_{0} be an equivalence between 𝗉{\mathsf{p}} and 𝗉{\mathsf{p}}^{\prime}. Since 𝗉,𝗉{\mathsf{p}},{\mathsf{p}}^{\prime} satisfy the inhomogeneous Maurer-Cartan equation and 𝖡{\mathsf{B}} is a cycle, [𝗉,𝗉][{\mathsf{p}},{\mathsf{p}}] and [𝗉,𝗉][{\mathsf{p}}^{\prime},{\mathsf{p}}^{\prime}] are cycles as well. From (29) we have

𝗉𝗉=𝖺+k=1ad𝖺k(𝗉)k!k=1ad𝖺k(𝖺)(k+1)!,\displaystyle{\mathsf{p}}^{\prime}-{\mathsf{p}}=-\partial{\mathsf{a}}+\sum_{k=1}^{\infty}\frac{{\rm ad}^{k}_{\mathsf{a}}({\mathsf{p}})}{k!}-\sum_{k=1}^{\infty}\frac{{\rm ad}^{k}_{\mathsf{a}}(\partial{\mathsf{a}})}{(k+1)!},

where

𝖿:=k=1ad𝖺k(𝗉)k!k=1ad𝖺k(𝖺)(k+1)![𝔐,𝔐]¯.\displaystyle{\mathsf{f}}:=\sum_{k=1}^{\infty}\frac{{\rm ad}^{k}_{\mathsf{a}}({\mathsf{p}})}{k!}-\sum_{k=1}^{\infty}\frac{{\rm ad}^{k}_{\mathsf{a}}(\partial{\mathsf{a}})}{(k+1)!}\in{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}}.

Thus [𝗉,𝗉][𝗉,𝗉]=2(𝗉𝗉)=2𝖿[{\mathsf{p}},{\mathsf{p}}]-[{\mathsf{p}}^{\prime},{\mathsf{p}}^{\prime}]=2\partial({\mathsf{p}}^{\prime}-{\mathsf{p}})=2\partial{\mathsf{f}} is \partial-exact in [𝔐,𝔐]¯{\overline{[{\mathfrak{M}},{\mathfrak{M}}]}}. ∎

References

  • [1] Yoshiko Ogata. A Z2-Index of Symmetry Protected Topological Phases with Time Reversal Symmetry for Quantum Spin Chains. Communications in Mathematical Physics, 374(2):705–734, 2019.
  • [2] Yoshiko Ogata. A 2{\mathbb{Z}}_{2}-index of symmetry protected topological phases with reflection symmetry for quantum spin chains. arXiv e-prints, page arXiv:1904.01669, April 2019.
  • [3] Chris Bourne and Yoshiko Ogata. The classification of symmetry protected topological phases of one-dimensional fermion systems. arXiv e-prints, page arXiv:2006.15232, June 2020.
  • [4] Anton Kapustin, Nikita Sopenko, and Bowen Yang. A classification of invertible phases of bosonic quantum lattice systems in one dimension. Journal of Mathematical Physics, 62(8):081901, 2021.
  • [5] Yoshiko Ogata. A H3(G,𝕋)H^{3}(G,{\mathbb{T}})-valued index of symmetry protected topological phases with on-site finite group symmetry for two-dimensional quantum spin systems. arXiv e-prints, page arXiv:2101.00426, January 2021.
  • [6] Nikita Sopenko. An index for two-dimensional SPT states. Journal of Mathematical Physics, 62(11):111901, 2021.
  • [7] Anton Kapustin and Nikita Sopenko. Hall conductance and the statistics of flux insertions in gapped interacting lattice systems. J. Math. Phys., 61(10):101901, 24, 2020.
  • [8] Adam Artymowicz, Anton Kapustin, and Nikita Sopenko. Quantization of the Higher Berry Curvature and the Higher Thouless Pump. Comm. Math. Phys., 405(8):Paper No. 191, 2024.
  • [9] Sven Bachmann, Alex Bols, and Mahsa Rahnama. Many-body Fu-Kane-Mele index. arXiv e-prints, page arXiv:2406.19463, June 2024.
  • [10] Sven Bachmann, Matthew Corbelli, Martin Fraas, and Yoshiko Ogata. Tensor category describing anyons in the quantum Hall effect and quantization of conductance. arXiv e-prints, page arXiv:2410.04736, October 2024.
  • [11] Anton Kapustin and Nikita Sopenko. Local Noether theorem for quantum lattice systems and topological invariants of gapped states. J. Math. Phys., 63(9):Paper No. 091903, 31, 2022.
  • [12] Gerard ’t Hooft. Naturalness, chiral symmetry, and spontaneous chiral symmetry breaking. NATO Sci. Ser. B, 59:135–157, 1980.
  • [13] Masaki Kashiwara and Pierre Schapira. Ind-sheaves. Astérisque, 271, 2001.
  • [14] Stéphane Guillermou and Pierre Schapira. Construction of sheaves on the subanalytic site. Astérisque, 383:1–60, 2016.
  • [15] Saunders Mac Lane and Ieke Moerdijk. Sheaves in geometry and logic. Universitext. Springer-Verlag, New York, 1994. A first introduction to topos theory, Corrected reprint of the 1992 edition.
  • [16] Glen Bredon. Cosheaves and homology. Pacific Journal of Mathematics, 25(1):1–32, April 1968.
  • [17] J. Chuang, A. Lazarev, and Wajid Mannan. Cocommutative coalgebras: homotopy theory and Koszul duality. arXiv e-prints, page arXiv:1403.0774, March 2014.
  • [18] M. B. Hastings. Lieb-Schultz-Mattis in higher dimensions. Physical Review B, 69(10):104431, 2004.
  • [19] Alexei Kitaev. Anyons in an exactly solved model and beyond. Annals of Physics, 321(1):2–111, 2006.
  • [20] Tobias J Osborne. Simulating adiabatic evolution of gapped spin systems. Physical review A, 75(3):032321, 2007.
  • [21] Ola Bratteli and Derek W. Robinson. Operator algebras and quantum statistical mechanics. 1. CC^{\ast}- and WW^{\ast}-algebras, symmetry groups, decomposition of states. Texts and Monographs in Physics. Springer-Verlag, New York, 2nd ed. edition, 1987.
  • [22] Anton Kapustin and Nikita Sopenko. Anomalous symmetries of quantum spin chains and a generalization of the Lieb-Schultz-Mattis theorem. arXiv e-prints, page arXiv:2401.02533, January 2024.
  • [23] Lou van den Dries. Tame topology and o-minimal structures, volume 248 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1998.
  • [24] R. Tyrrell Rockafellar. Convex analysis, volume No. 28 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1970.
  • [25] Fedor Petrov. Bounding distance to an intersection of polyhedra. MathOverflow. URL:https://mathoverflow.net/q/463857 (version: 2024-02-10).
  • [26] Karol Borsuk. On the imbedding of systems of compacta in simplicial complexes. Fundamenta Mathematicae, 35:217–234, 1948.
  • [27] Jean Leray. Sur la forme des espaces topologiques et sur les points fixes des représentations. Journal de Mathématiques Pures et Appliquées, 24:95–167, 1945.
  • [28] Ulrich Bauer, Michael Kerber, Fabian Roll, and Alexander Rolle. A unified view on the functorial nerve theorem and its variations. Expositiones Mathematicae, 41(4):125503, 2023.
  • [29] William M. Goldman and John J. Millson. The deformation theory of representations of fundamental groups of compact Kähler manifolds. Inst. Hautes Études Sci. Publ. Math., 67:43–96, 1988.
  • [30] Marco Manetti. Lie methods in deformation theory. Springer Monographs in Mathematics. Springer, Singapore, 2022.