A mathematical theory of topological invariants of quantum spin systems
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 -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 , the category of fuzzy semilinear sets. We identify a Grothendieck topology on and show that infinitesimal symmetries of any gapped state of a quantum lattice system on can be described by a local Lie algebra over . We also study gapped states invariant under an action of a compact Lie group and define a -equivariant analog of . In Section 6 we construct invariants of -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 to a local symmetry of the state . We also explain how to generalize the construction of invariants to lattice systems defined on sufficiently nice (asymptotically conical) subsets of 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 be a manifold and be the category whose objects are open subsets of , and the set of morphisms from an open to an open is the singleton or the empty set depending on whether or . Composition of morphisms is uniquely defined. A pre-cosheaf on with values in a category is a functor . Thus for every inclusion of opens one is given a co-restriction morphism such that for any three opens one has . Pre-cosheaves (as well as pre-sheaves, which are functors from the opposite category of to ) can be used to describe local data on . This form of locality is rather weak, since it does not require to be expressible through and .
As an example, consider the Lie algebra of gauge transformations, i.e. the Lie algebra of smooth functions on with values in a finite-dimensional Lie algebra . It is a global object attached to . To “localize” it, for any open we define the Lie algebra to be the space of smooth -valued functions on whose closed support is contained in . In particular, . For any inclusion of opens we have a homomorphism of Lie algebras such that the Lie algebras assemble into a pre-cosheaf of Lie algebras on . This is a coflasque pre-cosheaf, ie. all its structure maps are injective.222The terminology comes from sheaf theory, where a pre-sheaf is called flasque if for any the restriction morphism is a surjection.
Continuing with the example, for any two opens the following sequence is exact:
(1) |
Here the first arrow is and the second arrow is . Exactness follows from the existence of a partition of unity for the cover of . In words, the exactness of the sequence (1) means that any element of can be decomposed as a sum of elements of sub-algebras attached to and modulo ambiguities which take values in the sub-algebra attached to .
More generally, the existence of a partition of unity implies that for any collection of opens , , the following sequence is exact:
(2) |
By definition, this means that the pre-cosheaf of vector spaces 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 is not a cosheaf of Lie algebras. Nevertheless, the following additional property of can be regarded as a form of locality of the Lie bracket: for any . We will call this ”Property I”, where ”I” stands for ”intersection”. In particular, elements of which are supported on non-intersecting opens and 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 of open subsets of with a more general structure which admits the notions of intersection, union, and cover.
The first thing to note is that the category is rather special: its objects form a pre-ordered set (i.e. the set of objects carries a relation which is reflexive and transitive), and the category structure is determined by the pre-order. The relation is also anti-symmetric: and implies . In other words, is a poset. In general, we will not require the pre-order to be anti-symmetric. From the categorical viewpoint, and means that and are isomorphic objects of the category , and as a general rule, it is not advisable to identify isomorphic objects.
For any pre-ordered set there is a natural notion of intersection and union. The intersection of can be defined as the greatest lower bound (or meet) of both and , i.e. a such that , , and for any we have . The meet of and is denoted . Similarly, the union of and can be defined as the smallest upper bound (or join) of both and . It is denoted . 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 is such that and exist for all .333If we turn 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 arbitrary (i.e. not necessarily finite) joins make sense.
Finally, to define covers of elements of , let us assume that for all .444The opposite relation is automatic, so this condition ensures that for all . This is equivalent to saying that the poset corresponding to is a distributive lattice. We will say that is a distributive pre-ordered set. This condition is certainly satisfied for . We say that a collection of elements of covers iff for all and . This definition ensures that if covers , then for any the collection covers .
In the case of the pre-ordered set , the standard topological definition of a cover allows to be infinite. In general, if admits only finite joins, needs to be finite. Also, we may or may not allow to be empty. This possibility only arises when is the smallest element of , i.e. for any . In the case of the pre-ordered set , the smallest element is the empty set , 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 to be empty.
From a categorical perspective, this notion of a cover equips any distributive pre-ordered set 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 to be empty, this Grothendieck topology is canonical.
For any we may consider the subset with the pre-order inherited from . 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 .
Given any distributive pre-ordered set , 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 with and with . Motivated by the above example, we introduce the following definition.
Definition 2.1.
A local Lie algebra over 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 be a local Lie algebra. Then for any the Lie algebra is an ideal in . One can equivalently define a local Lie algebra as a pre-cosheaf of Lie algebras over 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 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 be a pre-cosheaf of vector spaces on a distributive lattice , with co-restriction maps . Then is a cosheaf on (with the topology of finite covers) iff for any the following sequence is exact
(3) |
where and .
The following proof is essentially a restatement of the proof of Proposition 1.3 in [16], adapted to the site :
Proof.
We must show that for any and every covering , the sequence of vector spaces
(4) |
is exact, where and . 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 , and let be a cover of . Suppose lies in the kernel of . Let . An application of (3) with and shows that for some , and by right-exactness of (4) this shows that for some . Finally we write
where . Both terms above the first term lies in the image of by the inductive hypothesis, while the second term equals , which evidently is also in the image of . ∎
2.2 DGLA attached to a cover
Let be a distributive pre-ordered set. Let and be a cover of . Let be a pre-cosheaf of vector spaces over . The Čech chain complex is defined by
(where some linear order on has been chosen) with the differential is the Čech differential given by , where is the canonical map
Lemma 2.2.
If is a coflasque cosheaf of vector spaces over , the homology of the Čech complex is for and for .
Let be the augmented Čech complex.
Proposition 2.1.
Let be a local Lie algebra over . Then for any and any cover of the 1-shifted augmented Čech complex has a natural structure of a non-negatively graded acyclic DGLA.
Proof.
Let be a cover of indexed by . Let be a vector space with a basis , and let , be the dual basis of . The 1-shifted augmented Čech complex of with respect to is naturally identified with a sub-complex of the DGLA , where is contraction with . 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 form a category whose morphisms are refinements. A refinement of a cover to a cover is a map such that . Refinements are composed in an obvious way.
Proposition 2.2.
For a fixed , the map which sends a local Lie algebra and a cover of to the DGLA is functorial in both and .
Proof.
Functoriality in is clear. Functoriality in is written Čech component-wise for :
for . ∎
We will call 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 in an obvious manner. The construction of the acyclic DGLA 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 (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 , is not required to be central and the derivation satisfies . We say that a graded local Lie algebra over with a terminal object is pointed if it is equipped with a distinguished central element of degree . Morphisms in the category of pointed graded local Lie algebras are required to preserve the distinguished element. Then the DGLA is an acyclic pointed DGLA, the distinguished central cycle being . 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 is functorial in both arguments.
3 Quantum lattice systems
3.1 Observables and derivations
We use the metric on , i. e. . For any we write and . They take values in extended non-negative reals . Thus and for any . For a nonempty set and we define while we set .
A quantum lattice system consists of a countable subset (“the lattice”) and a finite-dimensional complex Hilbert space for every . We make the following assumption on the lattice system666In [11] 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] was taken to be for simplicity. : there is a such that the number of points of in any hypercube of diameter is bounded by .
For any bounded nonempty let . For any there is an inclusion and the algebras form a direct system with respect to these inclusions. We extend this direct system to include the empty set by setting and letting the inclusion take . Each is a finite-dimensional -algebra with the operator norm, and the inclusions preserve this norm. The normed *-algebra of local observables is
The algebra of quasi-local observables is the norm-completion of ; it is a -algebra.
For any bounded , define the normalized trace as . For any bounded the partial trace is uniquely specified by the condition for any and . Besides forming a direct system with respect to inclusions, the spaces are also an inverse system with respect to the partial trace. extends to a normalized positive linear functional on , i.e. a state. We say that is traceless if . The space of traceless anti-hermitian elements of will be denoted . is a real Lie algebra with respect to the commutator. The Lie algebras form a direct system over the directed set of bounded subsets of , and its limit will be denoted . Equivalently, is the Lie algebra of traceless anti-hermitian elements of . Note that .
Definition 3.1.
A brick in is a non-empty subset of the form
(5) |
where , , and are -tuples of integers. We write for the set of all bricks in .
The set of bricks exhausts the collection of bounded subsets of 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 we have
Proof.
Any pair of points specifies a brick with and on opposing corners, and any brick can be specified this way (not uniquely). With the brick corresponding to and it is easy to see that , and so . Thus we have
and it remains only to bound the above sum. Let and . Using summation by parts we have
It is easy to check that and that , and so
which proves the Lemma. ∎
For any brick we define the following subspace of :
Each decomposes as a direct sum over bricks contained in , and for any brick the partial trace is the projection onto . Intuitively, consists of elements of which are not localized on any brick properly contained in .
Derivations of which appear in the physical context are typically only densely defined and have the form
where . We are now going to define for every a real Lie algebra which consists of derivations approximately localized on and such that all have a common dense domain. Moreover, they form a pre-cosheaf of Lie algebras over a certain category of subsets of .
Definition 3.2.
For any element of and every we let
(6) |
and define as the set of elements with for all .
If is empty, then for if and only if for all . Thus . We also denote .
It is easy to see that (6) is a norm on for each . We endow with the locally convex topology given by the norms (6) ranging over all . 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.
is a Fréchet space.
Proof.
The Hausdorff property follows from the fact that if for any then . To show completeness, suppose is Cauchy, i.e. that for any and any there is an such that . For any fixed this implies that is Cauchy in (with the operator norm) and thus converges to a limit . Let .
Fix and . For every , choose so that . For any , any , and any , we have
Taking shows that . Since was arbitrary, we have . Since was arbitrary, converges to in the topology of . ∎
Recall that for a nonempty we write . The norms obey the following dominance relation.
Lemma 3.2.
Let be subsets of the lattice and suppose that . Then for any we have
(7) |
In particular, and the inclusion is continuous.
Proof.
Let be any subset of the lattice. Then (7) follows from
where in the second line we used the triangle inequality and in the third we used . ∎
The above Lemma shows that the space only depends on the asymptotic geometry of the region in the following sense: if and for some then (as subsets of and are isomorphic as Fréchet spaces. In particular, for any non-empty bounded , the space coincides with .
To relate the spaces to the traditional -algebraic picture we prove the following:
Proposition 3.2.
Suppose is non-empty and bounded and let . Then the sum is absolutely convergent and defines a continuous dense embedding of into the subspace of traceless anti-hermitian elements of the algebra .
Proof.
We can assume without loss of generality that . We have
and by Lemma 3.1 the above sum is finite. This shows that the map is well-defined and continuous. Its image is dense in the space of traceless anti-hermitian elements of because it contains the dense subspace . Finally, to show that it is injective, writing we have
and so for all . ∎
Definition 3.3.
We let be the image of under the embedding of Prop. 3.2, with the Fréchet topology of .
In [11], the algebra of almost-local operators was defined as a subspace of where a countable family of norms similar to (6) takes finite values. Here we equivalently define as the set of elements of whose traceless hermitian and anti-hermitian parts live in , topologized as .
Proposition 3.3.
is a dense sub-algebra of .
Proof.
contains all local observables and these are dense in . The fact that 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 is inner iff it is contained in for some bounded .
For any two bounded sets , a local observable is strictly localized on both and iff it is localized on their intersection, ie. . An analogous relation for the spaces does not hold in general777Indeed, for any two bounded the spaces and coincide and are nontrivial but if and are disjoint then ., but it does hold if we assume that and satisfy the following transversality condition:
Definition 3.5.
Let and . We say and are -transverse if
for all .
We will say and are transverse if they are -transverse for some .
Proposition 3.4.
If are -transverse then
(8) |
for all . In particular, is a topological pullback: it is the set with the topology of simultaneous convergence in and .
Proof.
The next proposition relates with and for any .
Proposition 3.5.
Proof.
Next we will show how to endow with the Lie algebra structure.
Proposition 3.6.
Let . For any and the sum
(11) |
is absolutely convergent for every . The resulting bracket satisfies the Jacobi identity and
(12) |
for some constant that depends only on .
To prove Proposition 3.6 we will need several lemmas. For any , define888This is well-defined since the intersection of an arbitrary number of bricks is either empty or a brick, so is the intersection of all bricks containing and . the join as the smallest brick that contains and .
Lemma 3.3.
For any with we have
(13) |
and for any we have
Proof.
Let be the projection onto the th coordinate. The following identites hold for any bricks :
When the results are clear. When they follow from the case via the above identities. ∎
Lemma 3.4.
Let and let and . Then unless and .
Proof.
The requirement that is clear, since and would commute otherwise. Suppose . Then is a brick that is strictly contained in , so ∎
We make the following definitions for the next lemma. For any and any brick write and .
Lemma 3.5.
For any , , and we have
Proof.
Let with and let be a brick. Pick an arbitrary point . We have
and so, since by Lemma 3.3 , we have
(14) |
It follows that for any we have
and so
(15) |
It remains to bound the sum on the right-hand side. Fix an arbitrary and let with and . Then by the second statement in Lemma 3.3 and the inequality for we have
(16) |
Using (16) we can bound the sum (15) as follows:
(17) |
∎
Now we are ready to prove Proposition 3.6.
Proof of Proposition 3.6.
Notice that by Lemma 3.2 we have so without loss of generality we set .
Let . By Lemmas 3.4 and 3.5 we have
with . This proves that (11) is absolutely convergent and establishes the bound (12).
Next, let us prove the Jacobi identity. Let . We have
(18) |
To show that this sum is absolutely convergent, we bound
Here we used Lemma 3.5 in the second and fourth lines and is a constant depending only on . Thus the sum (18) is absolutely convergent. In particular, using the fact that we have the following absolutely convergent expression
It is then easy to check that the Jacobi identity for the sum follows from the Jacobi identity for each term. ∎
Corollary 3.1.
Suppose .
-
i)
is a jointly continuous bilinear map from to .
-
ii)
If and are transverse, then this is jointly continuous bilinear map from to .
In particular since and are transverse, acts continuously on and this action is easily seen to extend to a continuous action of on the space . We denote the action of on by . By Prop. 3.6, it is given by
(19) |
It is not hard to check that for any and any we have and so the action of on is faithful. Thus, elements of for any may be identified with a subset of the Fréchet-continuous derivations of . By Proposition 3.2, the Fréchet-Lie algebra is identified with the Fréchet-Lie algebra of traceless anti-hermitian elements of acting by inner derivations.
3.2 Automorphisms
In this section we recall certain automorphisms obtained by exponentiating elements of following [11]. One can develop the theory of such automorphisms that are almost-localized on regions in 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 be a smooth map. It is shown in [11] that for any the differential equation
with the initial condition has a unique solution for all . Denote by the map taking to . It is a continuous automorphism of the Lie algebra that preserves the -operation.
Definition 3.6.
We call any automorphism of the form for some smooth map 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 , and that the set of LGAs forms a group under composition.
3.3 States
By a state we will mean a state on the quasilocal algebra . If is an inner derivation (see Def. 3.4), we define its -average as the evaluation of on the corresponding element of . The group of LGAs acts on states by pre-composition, which we denote . We say an LGA preserves a state if . We say an element preserves if for any , which is equivalent to the one-parameter group of automorphisms corresponding to a constant map preserving .
Definition 3.7.
For any define as the set of all elements of that preserve .
It is easy to check that is a closed subset of , and that if and preserve , then preserves . Thus the analog of Propositions 3.4 and 3.6 and Corollary 3.1 hold for the spaces . Proposition 3.5 on the other hand, does not hold for the spaces for a general state . 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 is gapped if there exists and such that for any one has
(20) |
Remark 3.1.
The meaning of this condition becomes more transparent if one recalls that any is a generator of a one-parameter group of -automorphisms of [11]. The condition (20) implies that is invariant under this one-parameter group of automorphisms [21], and that the corresponding one-parameter group of unitaries in the GNS representation of has a generator whose spectrum in the orthogonal complement to the GNS vacuum vector is contained in . The condition (20) also implies that is pure [22].
Remark 3.2.
If is a gapped state of and is an LGA, then 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 is gapped, the corresponding Hamiltonian is . Then there are linear functions
such that
-
i)
If preserves then preserves .
-
ii)
For every , , and we have
for some constants depending only on , and .
-
iii)
For every we have
Using the above Lemma we will prove the analog of Proposition 3.5 for the spaces .
Proposition 3.8.
To prove Proposition 3.8 we will need the following geometric result
Lemma 3.6.
Let and define . Then and are transverse and their intersection is .
Proof.
It is easy to check that . To prove transversality we will show
(22) |
for every . Suppose first that . Then
which implies (22). Suppose instead that , and let satisfy . Notice and so lies in the boundary of , which implies . Thus we have
where in the first and third lines we used the triangle inequality. Using and , this gives , which implies (22). ∎
Proof of Proposition 3.8.
The proof of Proposition 3.5 goes through unmodified except for the definition of , which needs to be changed to ensure that the image of consists of derivations that preserve . Suppose and . Define
Let (resp. ) be the splitting from Proposition 3.5 with the sets and (resp. and ). Define as
for . Using Prop. 3.7 and Lemma 3.6 and the fact that the commutator of two derivations that preserve preserves , one checks that takes to . Using Prop. 3.7 and the fact that and , we get , as desired. ∎
We showed that one can attach Fréchet-Lie algebras and to any . These form a pre-cosheaf over the pre-ordered set of subsets of . Moreover, Proposition 3.5, Corollary 3.1, and (when is gapped) Proposition 3.8 show that these spaces satisfy the cosheaf condition and Property I for transverse pairs . What prevents the functors and from forming local Lie algebras is the fact that pairs of subsets generally do not intersect transversely. This problem can be resolved by restricting to a suitable set of subsets of 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 is a subset of which can be defined by means of a finite number of linear equalities and strict linear inequalities. More precisely, a basic semilinear set in 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 will be denoted . Projections map to [23]. A map is called semilinear iff its graph is a semilinear subset of . The composition of two semilinear maps is a semilinear map [23].
Recall that we use the metric on .
Lemma 4.1.
The distance function is semilinear.
Proof.
The function is semilinear for any . The function is semilinear. If are semilinear, then is semilinear. Since the set of semilinear functions is closed under composition, this proves the lemma. ∎
Recall that for any set , we write and call this the -thickening of . It is easy to see that if is closed, then is also closed for any .
Lemma 4.2.
If is semilinear, is semilinear for any .
Proof.
Consider the set
By the previous lemma, is semilinear. On the other hand, is the projection to the first of . Since intersection and projection preserve the set of semilinear sets, the lemma is proved. ∎
Lemma 4.3.
If is convex, then is convex, for any .
Proof.
Suppose , and suppose satisfy and . Then for any we have ∎
A polyhedron in 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 is a polyhedron, then is a polyhedron, for any .
4.2 A category of fuzzy semilinear sets
Clearly, if for we have , then for any we have . Also, for any and any we have . Thus we can define a pre-order on by saying that iff there exists such that . We will call this pre-order relation fuzzy inclusion. Equivalently, can be made into a category, with a single morphism from to iff . One can turn the pre-ordered set 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 and call it the category (or pre-ordered set) of fuzzy semilinear sets.
Proposition 4.1.
has all pairwise joins: for any the join is given by .
Proof.
If and for some , then and , and thus . ∎
Let us show that has all pairwise meets, using the notion of transverse intersection from the previous section. Recall (Definition 3.5) that we say two sets are transverse if for some we have for all .
We need the following geometric result [25]999The proof in [25] is for the Euclidean distance, but since the Euclidean distance function and are equivalent (each one is upper-bounded by a multiple of the other), the result applies to as well. :
Lemma 4.4.
Let and be polyhedra . If then and are transverse.
Proposition 4.2.
For every there is an such that and are transverse.
Proof.
Corollary 4.2.
With as above, is a meet of and . In particular, has all pairwise meets.
Proof.
Since and (resp. and ) are isomorphic in , it suffices to show that is a meet of and . It’s clear that and . Now suppose satisfies and . Then there is an such that every satisfies . Since for some we have . ∎
Proposition 4.3.
The pre-ordered set is distributive.
Proof.
We need to show that for any we have . According to Corollary 4.2, there exists such that . Since we also have . On the other hand, for some , and for some . Thus for some . Since we have inclusions and , the lemma is proved. ∎
We can now equip 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 , every bounded closed semilinear set is an initial object (they are all isomorphic objects of the category ). 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 we may consider a full sub-category whose objects are such that . 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 are isomorphic objects of the corresponding category. More generally, any two elements of which coincide outside some ball in are isomorphic objects. Thus encodes the large-scale structure of . To make this explicit, we will show that the pre-ordered set is equivalent as a category to a certain poset of subsets of the “sphere at infinity” .
A cone in is a non-empty subset of which is invariant under , where . Every cone contains the origin . Cones in are in bijection with subsets of where is the group of positive real numbers under multiplication. If , we denote the corresponding cone . In particular, . If is a cone in , we will denote the corresponding subset of by .
Definition 4.1.
is a spherical polyhedron iff is a polyhedron and is contained in some open hemisphere of . is a spherical CS set iff it is a union of a finite number of spherical polyhedra. The set of spherical CS sets in is denoted .
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 . The map which sends to the equivalence class of is a homeomorphism which establishes a bijection between bounded polyhedra in and spherical polyhedra contained in . ∎
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 under inclusion. This poset has pairwise joins and meets given by unions and intersections, respectively.
Proposition 4.6.
The category is equivalent to the category .
Proof.
We proceed first by defining a functor from to . 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 where for each , is the unit normal vector of the th supporting hyperplane and . By definition, is isomorphic to in . In other words, every polyhedron is isomorphic to a cone where is the matrix with as rows and we write for when for all . If spans , the set , which contains no antipodal points, is mapped to its corresponding spherical polyhedron. Otherwise, complete into a spanning set and decompose into a union of cones . 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 a fuzzy inclusion for some implies . It is easy to check that this functor is fully faithful and (essentially) surjective. ∎
Note that under this equivalence all bounded elements of correspond to . The canonical Grothendieck topology on corresponds to a slightly unusual Grothendieck topology on the poset of spherical CS subsets of : the one where empty covers of are not allowed. Consequently, a cosheaf of vector spaces on equipped with this topology need not map to the zero vector space.
4.4 Spherical CS cohomology
Let be a spherical CS cover of a spherical CS set . Spherical CS cohomology is defined to be the simplicial cohomology of the Čech nerve .
Definition 4.2.
Let be a spherical CS set. The spherical CS cohomology is defined as 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 the graded vector space is isomorphic to the singular cohomology .
Proof.
Every spherical CS cover can be refined to a cover by spherical polyhedra, so in the computation of it suffices to take colimit over such covers. All intersections of elements of a spherical polyhedral cover are contractible. Also, since every polyhedron is a geometric realization of a simplicial complex, is a cover of a simplicial complex by subcomplexes. By Theorem C of [28] for such the direct system of groups is constant and its limit is . ∎
5 Local Lie algebras over fuzzy semilinear sets
5.1 Basic examples
Let be a state of a quantum lattice system on . By Lemma 3.2 the maps sending to and are pre-cosheaves of Fréchet spaces on . We denote these pre-cosheaves by and . Putting together Lemma 2.1, Proposition 3.5, Corollary 3.1, and Proposition 3.8, we have
Theorem 5.1.
is a local Lie algebra over . If is gapped, then is a local Lie algebra over .
Our main object of study is the local Lie algebra attached to a gapped state of a lattice system .
A much simpler example of a local Lie algebra arises from a finite-dimensional Lie algebra and any subset . For any let be the space of bounded functions which decay superpolynomially away from . The subscript “al” stands for “almost localized”. More precisely, is the space of bounded functions , while is defined as a subspace of consisting of functions such that the following seminorms are finite:
Proposition 5.1.
The assignment is a local Lie algebra.
Proof.
It is easy to check that the assignment 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 the sequence
is exact. To show exactness at , we note that every can be written as a sum , where
and is defined by a similar expression with and exchanged. Using it is easy to check that and for all , and thus and .
5.2 Symmetries of lattice systems
If is countable, it can be viewed as a lattice in the physical sense, and the local Lie algebra over models infinitesimal gauge transformations of a lattice system on .
Definition 5.1.
A local action of a compact Lie group on a lattice system is a collection of homomorphisms such that the norms of the corresponding Lie algebra homomorphisms are bounded uniformly in .
A local action of on gives rise to a homomorphism from to the automorphism group of via which is smooth on . The corresponding generator is a homomorphism from to the Lie algebra of derivations of defined by
where is the traceless part of the generator of . The image of lands in , with . We will regard as a homomorphism of Fréchet-Lie algebras .
This morphism of Fréchet-Lie algebras can be lifted to a morphism of local Lie algebras over . Indeed, for any and any we let be a derivation of given by
It is easy to check that this derivation belongs to and that the above map is a continuous homomorphism . 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 of is said to be invariant under a local action of a compact Lie group if it is invariant under the corresponding automorphisms of .
Let be a gapped state of invariant under a local action of a Lie group . In that case the image of lands in . One may ask if this morphism of Fréchet-Lie algebras can be lifted to a morphism of local Lie algebras . If this is the case, then the symmetry of 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 -invariant gapped state we are going to define a graded local Lie algebra over which is a -equivariant version of the local Lie algebra . 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 is a local Lie algebra and is a locally finite supercommutative graded algebra with finite-dimensional graded factors ,101010A graded vector space is locally finite iff its graded components are finite-dimensional. then is a graded local Lie algebra. We denote it .
Fix a compact Lie group and a distributive pre-ordered set and consider the category of graded local Lie algebras over equipped with a -action. An object of this category is a graded local Lie algebra on which acts by automorphisms; morphisms are defined in an obvious manner. The first step is to define a functor from this category to the category of graded local Lie algebras over such that is the Lie algebra of -invariant elements of . It is clear how to define such a functor for coflasque pre-cosheaves of Lie algebras with Property I, but the pre-cosheaf will not be a cosheaf of vector spaces without further assumptions about and the -action.
Definition 5.3.
An action of on a pre-cosheaf of Fréchet spaces is smooth if for each the seminorms defining the topology of can be chosen to be -invariant and the map defining the action is smooth. An action of on a pre-cosheaf of graded Fréchet spaces is smooth is the -action on every graded component is smooth.
Proposition 5.2.
Let be a pre-cosheaf of graded Fréchet spaces over equipped with a smooth action of a compact Lie group . The assignment 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 the sequence
is exact. To show exactness at the rightmost term, let be a -invariant element of and let and be such that . Averaging the action map over with the Haar measure gives a linear map which is identity when restricted to . The co-restriction morphisms intertwine these maps. Thus which proves that 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 to be continuous. However, if it is smooth, becomes a -module and all elements in are annihilated by the -action. We will use this later on.
Example 5.1.
If acts locally on a lattice system, the action of on the local Lie algebra is smooth. If is a -invariant gapped state of such a lattice system, the action of on is smooth.
Example 5.2.
Consider the local Lie algebra over associated to a finite-dimensional Lie algebra (see Section 4). Assume that is the Lie algebra of a compact Lie group , then there is an obvious -action on :
This -action is smooth.
Example 5.3.
If is a local Fréchet-Lie algebra with a smooth -action and is a locally-finite supercommutative graded algebra on which acts by automorphisms, then the -action on is smooth.
Corollary 5.1.
Let be a graded local Fréchet-Lie algebra over with a smooth -action. The functor of -invariant elements maps to a graded local FréchetLie algebra over .
Definition 5.4.
Let be a local Fréchet-Lie algebra over with a smooth -action. The -equivariantization functor sends to the negatively-graded local Fréchet-Lie algebra defined by
In the cases of interest to us, the -action on a local Lie algebra over is infinitesimally inner, in the sense that the -module structure mentioned in Remark 5.1 arises from a homomorphism . In such a case, the graded local Lie algebra has an extra bit of structure: a central element in of degree . This element is simply re-interpreted as an element of . In the terminology of Section 2, is a pointed graded local Fréchet-Lie algebra over . It is easy to see that the -equivariantization functor respects this extra structure. That is, if is a morphism of local Fréchet-Lie algebras over commuting with infinitesimally inner smooth -actions on and , then maps the central element to the central element .
Example 5.4.
Let be a -invariant gapped state of a quantum lattice system with a local -action which on the infinitesimal level is described by . Consider the graded local Lie algebra with its smooth -action (Example 5.1) and its -equivariantization . The distinguished central element of is regarded as an element of .
Example 5.5.
Consider the graded local Lie algebra of Example 5.2. The degree component of is the space of -invariant bounded functions on with values in . The distinguished central element is the constant function on which takes the value .
Armed with the equivariantization functor, we can now explain our strategy for constructing obstructions for the existence of a local Lie algebra morphism which lifts the Fréchet-Lie algebra morphism . Suppose such a morphism exists. Applying the -equivariantization functor, we get a morphism of pointed negatively-graded local Fréchet-Lie algebras . For any CS cover of an application of the Čech functor gives a morphism of acyclic pointed DGLAs
Consequently, a obstruction for the existence of such a pointed DGLA morphism is an obstruction for the existence of . 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 -invariant gapped states
6.1 The commutator class
Let be a pointed DGLA, i.e. a DGLA with a distinguished central cycle . Assume it is a limit of an inverse system of nilpotent pointed DGLAs , .
Definition 6.1.
The commutator DGLA, denoted by , is defined to be the closure of the commutator subalgebra of . Namely, if and only if for any its projection to is a finite linear combination of commutators in .
Even if is acyclic, the DGLA is not necessarily acyclic. We would like to construct an obstruction to finding an element which satisfies and .
Definition 6.2.
Let be a pointed DGLA. A -twisted Maurer-Cartan element in is which satisfies
We will denote the set of -twisted MC elements of by . The map can be upgraded to a functor from the category of pointed DGLAs to the category of sets in an obvious way.
Let be pronilpotent pointed DGLA and . Then is a cycle of the DGLA .
Proposition 6.1.
Let be an acyclic pointed DGLA which is a limit of an inverse system of nilpotent acyclic pointed DGLAs , . Assume further that the structure morphisms are surjective and is central in . Finally, assume . Then is non-empty. Furthermore, the homology class of in for is independent of the choice of .
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 the commutator class of the acyclic pointed DGLA , for lack of a better name, and denote it . The assignment 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 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 over some , where is a local Lie algebra equipped with a smooth infinitesimally inner -action. Indeed, along with the graded algebra used in the construction of one can consider its quotient by the ideal . Replacing with the nilpotent graded algebra throughout, for any cover of the terminal object one gets a sequence of nilpotent acyclic pointed DGLAs labeled by . They assemble into an inverse system in an obvious manner, and its limit is the acyclic pointed DGLA 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 and and any CS cover of .
Proposition 6.2.
For any pointed DGLA obtained, as above, from the Čech functor with respect to a cover , is functorial in .
Proof.
Let be a refinement of . By definition, there exists a map with According to Prop. 2.2, there is a map from to . As is unaffected by , we deduce from that . ∎
Example 6.1.
Let be a compact Lie group, be its Lie algebra, and be the pointed local Lie algebra over of Examples (5.2) and (5.5). The distinguished central cycle in is the constant function on with value . Here is regarded as a -invariant element of . For any cover one can construct a twisted MC element as follows. Pick large enough so that the interiors of , , cover in the usual sense and pick a partition of unity , , subordinate to this open cover. For any and any let . It is easy to verify that this is a twisted MC element satisfying . Thus the commutator class vanishes in this case.
Let be a compact Lie group, be a gapped -invariant state of a lattice system on , and . The commutator class of the acyclic pointed DGLA is an obstruction for the existence of a morphism of local Lie algebras which is a lift of the Lie algebra morphism . Indeed, as explained in Section 5.3, if exists, it induces a morphism of pointed DGLAs which in turn induces a morphism in the category which maps the commutator class of to the commutator class of . 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 -invariant state because it takes values in a set which itself depends on . 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 and spherical CS cohomology classes of the sphere at infinity , which takes values in the algebra of -invariant symmetric polynomials on . This gives a useful invariant of a gapped phase which is also an obstruction to promoting the global symmetry of the state to a local symmetry. Additionally, we will show that the invariant does not depends on the choice of the cover and thus is essentially unique.
Keeping in view further generalizations, we work over a sub-site 111111This is the so-called overcategory whose objects are equipped with a morphism to . Due to coflasqueness, this amounts to a restriction to objects fuzzily included in . which depends on an arbitray . Let be the image of under the equivalence between and . We continue to denote by the local Fréchet-Lie algebra which maps to . For any cover of we denote by the corresponding cover of .
Definition 6.3.
For any covering , any , and any define an evaluation
where . We adopt the convention that for bounded. The definition implicitly uses co-restriction of the coflasque cosheaf.
Lemma 6.1.
For any covering , any cycle , and any cocycle the derivation is inner.
Proof.
The chain complex
(23) |
is acyclic. Since is a cycle, for some Thus
where is the adjoint of defined by
The last expression is clearly inner because each is almost localized near some bounded region. The last equality depends on vanishing of
because when is unbounded where is the coboundary of . ∎
Remark 6.1.
The above lemma remains true if one replaces with where is any locally-finite graded vector space. Then the pairing takes values in the space of inner derivations (Definition 3.3) tensored with .
Remark 6.2.
The relation between and is as follows. They are equal for with unbounded. For with bounded, and is in general nonzero whereas . This discrepancy arises because the Grothendieck topology on the poset of spherical CS sets induced by the coherent topology on 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 are closed semilinear sets with a bounded meet then any derivation in is inner and thus has a well-defined average in any state on . More generally, if is a locally-finite supercommutative graded algebra and or some sub-algebra thereof, any element of for bounded has a well-defined average in any state of . The average takes values in . We will need the following more refined result:
Lemma 6.2.
Let and be closed semilinear sets such that is bounded. Let be a locally-finite supercommutative graded algebra and be a state. Then
Proof.
In what follows we will apply Lemma 6.1 to the graded local Lie algebra (see Remark 6.1). The evaluation of cycles of on spherical CS cocycles has especially nice properties when DGLA cycles belong to the commutator DGLA
(24) |
Proposition 6.3.
Proof.
The evaluations of cycles of the commutator DGLA on spherical CS cocycles can be used to construct topological invariants of -invariant gapped states on . We set . The cycle we use as an input is the commutator class of defined using the inhomogeneous Maurer-Cartan equation (Section 6.1). Since is a sub-algebra of with , Prop. 6.3 applies to such cycles.
Let be a CS cover of , and be a spherical CS cocycle of of degree with respect to the cover . Let be a -twisted MC element of . Evaluating on and taking into account that the grading is shifted by relative to the Čech grading, we get a -invariant element of where . Since is -invariant, is a -invariant element of . Equivalently, it is the value of a degree linear function on cocycles of valued in .
Theorem 6.1.
The function depends only on and the class of in the CS cohomology of .
Proof.
For a fixed covering , Prop. 6.1 and 6.3 implies is independent of . It depends on solely through its cohomology class. It remains to show invariance under refinement of cover.
Let be a refinement of with Cocycles and , where , are in the same CS cohomology class. From Prop.6.1 there exists -twisted MC element for the cover . Prop.6.2 then implies that is a -twisted MC element for the cover . An easy expansion shows
Since any -twisted MC element gives the same answer, we have shown that this contraction of interest depends only on the CS cohomology class of . ∎
By Proposition 4.7, the cohomology group is one-dimensioal for and vanishes otherwise. It is natural to take the class of to be a generator of . This generator is uniquely defined once an orientation of has been chosen. Thus for any even we obtain an invariant of gapped -invariant states on taking values in -invariant polynomials on of degree , and this invariant changes sign when the orientation of is changed. This is in agreement with Chern-Simons field theory.
6.3 The Hall conductance
For physics applications, the most important case is and . Let us specialize the construction of topological invariants to this case.
Let be a -invariant gapped state of a lattice system on . The generator of the action is . Let be the sub-algebra of -invariant elements of . Since is connected, this the same as the sub-algebra of elements of which commute with . More generally, for any . This is a local Lie algebra over . The graded local Lie algebra reduces in this case to where is a variable of degree .
Pick a cover of . To find a solution of the inhomogenenous Maurer-Cartan equation with , we write , where . To compute the topological invariant of a state on it is sufficient to solve for .
is a solution of . Explicitly, such that . Such exist because is a cosheaf. Then the component of the commutator class in is . The topological invariant of the state is obtained by evaluating it on a Čech 1-cocycle on corresponding to the cover and then averaging the resulting inner derivation:
Note that averaging over must be performed after the summation over because is not an inner derivation, in general.
The simplest CS cover of which can represent a nontrivial class in 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
So far we assumed that is an arbitrary countable subset of with a uniform bound on the number of points in a ball of radius . Suppose for some CS set and some . If this is the case, we will say the pair describes a quantum lattice system on . Then for any the component vanishes for any brick which does not intersect , and thus for any we have and . Thus the assignment can be regarded as a local Lie algebra over the site .
In particular, if a compact Lie group acts locally on the quantum lattice system on and preserves a gapped state , the generator of the action takes values in the sub-algebra . Consequently is a graded local Fréchet-Lie algebra over pointed by . Picking a cover of and applying the Čech functor gives an acyclic pointed DGLA whose commutator class can be evaluated on any CS cocycle of to give an inner derivation. Its -average is a -invariant polynomial on which depends only on the class of in . Thus a quantum lattice system on has a topological invariant which is a linear function of . It is easy to check that this linear function has degree .
For example, consider a -invariant gapped state of a quantum lattice system on affinely embedded in . The topological invariant defined in Section 6.2 is obtained by contracting with a 2-cocycle of (the sphere at infinity) and vanishes for dimensional reasons. On the other hand, by contracting with the 1-cocycle of one obtains the Hall conductance of this system.
For a more non-trivial example, for any consider a finite graph whose edges are geodesics and take to be the cone with base and apex at an arbitrary point of . The invariants of -invariant gapped states of quantum lattice systems on are labeled by generators of the free abelian group . This example goes beyond TQFT since 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
First let us characterize in terms of 0-chains. A 0-chain on is an element such that
(25) |
We say a 0-chain is supported on if whenever , and write for the set of -supported 0-chains, endowed with the norms (25) for .
Proposition A.1.
Let be nonempty and let be its 1-thickening.
-
i)
If then for a -supported 0-chain with
-
ii)
If , then for any the sum
is absolutely convergent and defines a map with , where the constant depends only on .
Lemma A.1.
For any nonempty we have
Proof.
Choose and with , and choose with . Then since we have
and the Lemma follows. ∎
Proof of Proposition A.1.
. Suppose . Choose any total order on and for every let be the closest point to in , using the total order as a tiebreaker. For every , define
(26) |
Then either or and , and so
which by Lemma A.1 is bounded by .
. Suppose that is a -supported -chain. For any we have
It is not hard to show that for any brick the quantity is bounded by a constant depending only on , which shows . ∎
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
where is the set of traceless anti-hermitian operators strictly localized on the ball of radius around . To apply their results we prove the equivalence of these norms:
Lemma A.2.
For any and , the norms and obey
(27) | ||||
(28) |
where are constants depending only on .
Proof.
Suppose and let . Define . Then
Since means , we continue this as follows
where in the last line we used Lemma 3.1. This proves (27). To prove (28) suppose and let be any brick. Set . Then . Indeed, if then and , implying , which is impossible. Choose with . Since we have , and so
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 and . Suppose is gapped with Hamiltonian and gap , and write for the one-parameter family of LGAs obtained by exponentiating . There exists121212See for instance Lemma 2.3 in [9]. a function such that , and the Fourier transform131313We use the convention . is supported in the interval and satisfies . Let be the odd function which on the positive real line is given by . Then we define and as the following integral transforms:
Appendix B Inhomogeneous Maurer-Cartan equation
Let be a pointed pronilpotent DLGA which is a limit of an inverse system of nilpotent pointed DGLAs , . The set of -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 for all and the DGLA is the inverse limit of the corresponding system of DGLAs. For any let be the natural projection. Let . The sets of -twisted MC elements of will be denoted .
A -twisted MC element gives rise to a degree derivation of which squares to zero (twisted differential).
Lemma B.1.
If , then for all . Further, iff .
Proof.
Straightforward. ∎
Thus is the inverse limit of the system of sets , .
We are going to define an equivalence relations on and for all . This is done in the same way as for the ordinary (homogeneous) Maurer-Cartan equation [29, 30].
First, is a nilpotent Lie algebra, so there is a well-defined nilpotent Lie group with the group law given by the Campbell-Baker-Hausdorff formula. Similarly, is pronilpotent (i.e. is an inverse limit of a system of nilpotent Lie algebras), so the CBH formula defines a group .
Second, there are Lie algebra homomorphisms from (resp. ) to the Lie algebras of affine-linear vector fields on (resp. ). This homomorphism maps or to the affine-linear vector field
where or . Here we used the identification of the space of affine-linear vector fields on a vector space with the space of affine-linear maps . These homomorphisms exponentiate to actions of the groups and on and by affine-linear transformations. Explicitly, the actions are given by [29, 30]:
(29) |
Lemma B.2.
The actions of on (resp. on ) preserve the sets (resp. ).
Proof.
The proof in [29], Section 1.3, applies just as well in the inhomogeneous case. ∎
We say that elements of or 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 -twisted Deligne groupoid as the transformation groupoid for the action of on . Similarly, one can define ”reduced” Deligne groupoids for every .
We observe an easy but useful lemma.
Lemma B.3.
If , , are equivalent, then is equivalent to for all .
Now come the interesting statements. Assume from now on that the DGLAs and are acyclic, that is central for all , and that the morphisms are surjective for all .
Lemma B.4.
With the above assumptions, the set is non-empty if and only if is non-empty.
Proof.
The only if statement follows from Lemma B.1. To prove the if direction, we use induction on . Assume is non-empty. Let . Pick . Since is a -twisted MC element and , we must have
for some . We look for a solution of the -twisted MC equation of the form where . Taking into account that is central, the inhomogeneous MC equation reduces to . Since by assumption , and form a short exact sequence and the latter two are acyclic, so is . Hence such a exists. This completes the inductive step proving that for all . Moreover, we also proved that the morphisms are surjective. Therefore by Lemma B.1 is non-empty. ∎
The above lemma proves the first part of Prop. 6.1. Indeed when , is implied by acyclicity. Next we show that with the above assumption on and all -twisted MC elements are equivalent.
Lemma B.5.
Suppose and . Then for any one has
Proof.
See [29], Lemma 2.8. ∎
Lemma B.6.
Let , such that . Then and are equivalent.
Proof.
Let . By assumption, . Moreover, . Indeed:
By acyclicity of , we have for some . Then Lemma B.5 implies that maps to . ∎
Lemma B.7.
Let , be such that and are equivalent. Then and are equivalent.
Proof.
Let be an equivalence between and , i.e. Let be any lift of . Then . By Lemma B.6, is equivalent to , therefore is equivalent to . ∎
Proposition B.1.
For any all elements of are equivalent. All elements of are equivalent.
The first statement is proved by induction on , where the inductive step is Lemma B.7. The second statement follows by passing to the inverse limit in .
Theorem B.1.
Let . Then the homology class of in is independent of the choice of .
Proof.
Let and let be an equivalence between and . Since satisfy the inhomogeneous Maurer-Cartan equation and is a cycle, and are cycles as well. From (29) we have
where
Thus is -exact in . ∎
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 -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 -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. - and -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.