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

Characteristic Topological Invariants

Oliver Knill Department of Mathematics
Harvard University
Cambridge, MA, 02138
(Date: February 5, 2023)
Abstract.

The higher characteristics wm(G)w_{m}(G) for a finite abstract simplicial complex GG are topological invariants that satisfy kk-point Green function identities and can be computed in terms of Euler characteristic in the case of closed manifolds, where we give a new proof of wm(G)=w1(G)w_{m}(G)=w_{1}(G). Also the sphere formula generalizes: for any simplicial complex, the total higher characteristics of unit spheres at even dimensional simplices is equal to the total higher characteristic of unit spheres at odd dimensional simplices.

Key words and phrases:
Topological invariant, Green function, Higher characteristics

1. Summary

1.1.

We prove kk-point Green function formulas

wm(G)=xjGgm(x1,,xk)=xGkj=1kw(xj)wm(j=1kU(xj))w_{m}(G)=\sum_{x_{j}\in G}g_{m}(x_{1},\dots,x_{k})=\sum_{x\in G^{k}}\prod_{j=1}^{k}w(x_{j})w_{m}(\bigcap_{j=1}^{k}U(x_{j}))

for the mm’th characteristic of AGA\subset G

wm(A)=xAm,jxjAj=1mw(xj),w_{m}(A)=\sum_{x\in A^{m},\bigcap_{j}x_{j}\in A}\prod_{j=1}^{m}w(x_{j})\;,

where w(x)=(1)dim(x)w(x)=(-1)^{{\rm dim}(x)} and where GG is a finite abstract simplicial complex. The stars U(x)={y,xy}U(x)=\{y,x\subset y\} define a topological base of a finite non-Hausdorff topology on GG. For two arbitrary open sets U,VU,V, the valuation formula

wm(UV)+wm(UV)=wm(U)+wm(V)w_{m}(U\cup V)+w_{m}(U\cap V)=w_{m}(U)+w_{m}(V)

holds for all m1m\geq 1. This makes the higher characteristics wmw_{m} topological invariants: they agree for homeomorphic complexes. It also makes more general energized versions of wmw_{m} sheaf ready. For manifolds, wm(M)=w1(M)w_{m}(M)=w_{1}(M).

Refer to caption
Figure 1. These complexes G,H,KG,H,K are all homotopic to a wedge sum of a 2-sphere and a 1-sphere. The first two complexes G,HG,H are homeomorphic. The complex KK however is not homeomorphic to GG. We have Euler ω1(G)=ω1(H)=ω1(K)=1\omega_{1}(G)=\omega_{1}(H)=\omega_{1}(K)=1 and Betti vector b1(G)=b1(H)=b1(K)=(1,1,1)b_{1}(G)=b_{1}(H)=b_{1}(K)=(1,1,1). For Wu ω2(G)=ω2(H)=11,b2(G)=b2(H)=(0,0,11,1,1)\omega_{2}(G)=\omega_{2}(H)=11,b_{2}(G)=b_{2}(H)=(0,0,11,1,1) and ω2(K)=9,b2(K)=(0,0,11,3,1)\omega_{2}(K)=9,b_{2}(K)=(0,0,11,3,1). Then ω3(G)=ω3(H)=29,b3(G)=b3(H)=(0,0,0,31,3,2,1)\omega_{3}(G)=\omega_{3}(H)=-29,b_{3}(G)=b_{3}(H)=(0,0,0,31,3,2,1) and ω3(K)=23,b3(K)=(0,0,0,31,12,5,1)\omega_{3}(K)=-23,b_{3}(K)=(0,0,0,31,12,5,1).

1.2.

We use short hand wm(A)=XAm,XAw(X)w_{m}(A)=\sum_{X\in A^{m},\bigcap X\in A}w(X), summing over X=(x1,,xm)AmX=(x_{1},\dots,x_{m})\in A^{m} with XA\bigcap X\in A, meaning j=1mxjA\bigcap_{j=1}^{m}x_{j}\in A and w(X)=j=1m(1)dim(xj)w(X)=\prod_{j=1}^{m}(-1)^{{\rm dim}(x_{j})}. The Euler characteristic is w1(A)w_{1}(A) and the Wu characteristic is w2(A)w_{2}(A). kk-point Green function maps a k-point configuration XGkX\in G^{k} to gm(X)=w(X)wm(U(X))g_{m}(X)=w(X)w_{m}(U(X)) with U(X)=j=1kU(xj))U(X)=\bigcap_{j=1}^{k}U(x_{j})). The energy formula wm(G)=XGkgm(X)w_{m}(G)=\sum_{X\in G^{k}}g_{m}(X) involves open sets in GG and so is sheaf theoretic. Unlike in wm(G)w_{m}(G), there is no XG\bigcap X\in G assumption in the Green part. The closure B(x)B(x) of a star U(x)U(x) is called the unit ball. Its boundary S(x)S(x) is the unit sphere also known as link. The local valuation formula

wm(B(x))=wm(U(x))(1)mwm(S(x))w_{m}(B(x))=w_{m}(U(x))-(-1)^{m}w_{m}(S(x))\;

holds for all stars U(x)U(x), but in general not for arbitrary open sets UU with with B(X)=U(X)¯B(X)=\overline{U(X)} and S(X)=B(X)U(X)=δU(X)S(X)=B(X)\setminus U(X)=\delta U(X).

1.3.

The special one-particle case k=1k=1 gives for all m1m\geq 1 a Gauss-Bonnet or Poincaré-Hopf type formula

wm(G)=xGw(x)wm(U(x)),w_{m}(G)=\sum_{x\in G}w(x)w_{m}(U(x))\;,

which reduces the computation of wm(G)w_{m}(G) to local expressions. We can think of gm(x)g_{m}(x) as a “curvature” or interpret it as an “index”. Note that the Gauss-Bonnet type formula for open sets does not generalize if GG is replaced by an open set UU. It only works if GG is replaced by a smaller simplicial complex. The unit balls B(x)=U(x)¯B(x)=\overline{U(x)} for example are by themselves again simplicial complexes and we also have wm(G)=xGw(x)wm(B(x))w_{m}(G)=\sum_{x\in G}w(x)w_{m}(B(x)).

1.4.

For all complexes GG, the sphere formula

xGw(x)wm(S(x))=0\sum_{x\in G}w(x)w_{m}(S(x))=0

holds. We have seen that in [7, 10] for m=1m=1. It follows from the local valuation formula and the ball formula wm(G)=xGw(x)wm(B(x))w_{m}(G)=\sum_{x\in G}w(x)w_{m}(B(x)). This is can be useful for large complexes, where we can use the ball formula again for computing wm(B(x))w_{m}(B(x)), especially if many B(x)B(x) have the same type or are known. The sphere formula for example proves that if GG is a complex for which the characteristic wmw_{m} of all spheres S(x)S(x) is constant c0c\neq 0 then wm(G)=0w_{m}(G)=0. This applies for odd-dimensional manifolds or more generally for odd-dimensional Dehn-Sommerville d-manifolds 𝒳d\mathcal{X}_{d} [4] recursively defined by the property χ(G)=1+(1)d\chi(G)=1+(-1)^{d} and that all unit spheres satisfy S(x)𝒳d1S(x)\in\mathcal{X}_{d-1}. The induction starts with 𝒳1={0}\mathcal{X}_{-1}=\{0\}, where the void 0={}0=\{\} is the empty complex, which is also the (1)(-1)-sphere.

1.5.

Unlike manifolds, Dehn-Sommerville spaces 𝒳=𝒳\mathcal{X}=\bigcup\mathcal{X} define a monoid under the join operation G+H=GH{xy,xG,yH}G+H=G\cup H\cup\{xy,x\in G,y\in H\}. It contains the sphere monoid 𝒮\mathcal{S} of all spheres in which the (1)(-1) sphere 0 is the zero element. The join of a pp-spheres with a qq-sphere is a (p+q+1)(p+q+1)-sphere using induction and SG+H(x)=SG(x)+HS_{G+H}(x)=S_{G}(x)+H for xGx\in G or SG+H(x)=G+H(x)S_{G+H}(x)=G+H(x) for xGx\in G. It is natural to look at generalized dd-manifolds which are complexes for which all unit spheres are in 𝒳d1\mathcal{X}_{d-1}. This is a larger class than dd-manifolds but which like manifolds is no sub-monoid of the monoid (𝒢,+)(\mathcal{G},+) of all complexes.

1.6.

We have proven the energy formula in the case m=1,k=2m=1,k=2 already as χ(G)=x,yg(x,y)\chi(G)=\sum_{x,y}g(x,y), where

g(x,y)=w(x)w(y)χ(U(x)U(y))g(x,y)=w(x)w(y)\chi(U(x)\cap U(y))

are the matrix entries of the inverse of the connection matrix L(x,y)=1L(x,y)=1 if xyx\cap y\neq\emptyset and L(x,y)=0L(x,y)=0 else. In the case m=1,k=1m=1,k=1, we have seen the super trace expression χ(G)=w1(G)=xGw(x)χ(U(x))\chi(G)=w_{1}(G)=\sum_{x\in G}w(x)\chi(U(x)) involving g(x,x)=χ(U(x))g(x,x)=\chi(U(x)). We then generalized the energy formula to energized complexes, where w(x)w(x) was replaced by an arbitrary ring-valued function h(x)h(x). We had explored such energized complexes also in the case when hh was division algebra valued.

1.7.

The energy theorem can more generally be seen as a linear transformation which turns an interaction energy function hm:GmRh_{m}:G^{m}\to R on intersecting points to a Green function gm:GkRg_{m}:G^{k}\to R on k-tuples of open stars. The parameters mm and kk can be arbitrary positive integers. The energy

wm(A)=XAm,XAhm(X)w_{m}(A)=\sum_{X\in A^{m},\bigcap X\in A}h_{m}(X)

then defines kk-point potential values

gm(X)=w(X)wm(U(X))g_{m}(X)=w(X)w_{m}(U(X))

or XGkX\in G^{k}. We then still have for any m,k1m,k\geq 1 the relation

wm(G)=XGm,X>0hm(X)=XGkgm(X).w_{m}(G)=\sum_{X\in G^{m},\bigcap X>0}h_{m}(X)=\sum_{X\in G^{k}}g_{m}(X)\;.

Of course, for a general choice of the interaction energy hh, there is no topological invariance. Indeed, already requiring the weaker combinatorial property of being invariant under Barycentric refinements forces hh to be of the form h(x)=w(x)h(x)=w(x).

1.8.

The more general picture with energized local interaction hmh_{m} explains, why a positive semi-definite hh produces positive semi-definite gg: if hmh_{m} is zero except for some fixed X0X_{0}, where hm(X0)=1h_{m}(X_{0})=1, then gm:XGkgm(X)g_{m}:X\in G^{k}\to g_{m}(X) is positive semi definite form. Convex combinations of positive semi definite forms hmh_{m} in mm variables are mapped into convex combinations of positive semi definite forms gmg_{m} in kk variables. In the case m=1,k=2m=1,k=2 for example, where the Green function matrix g(x,y)g(x,y) is the inverse of the connection matrix L(x,y)L(x,y), we have a duality between positive definite quadratic forms. In this case, LL and gg are even isospectral. They gave isospectral positive definite integral quadratic forms. One can write L(x,y)=wm(V(x)V(y))L(x,y)=w_{m}(V(x)\cap V(y)) with V(x)={x}¯V(x)=\overline{\{x\}}. For |h(x)|=1,k=2,m=1|h(x)|=1,k=2,m=1 we had them inverses of each other. [5, 8, 6].

1.9.

As for references, we should add that the subject has a large cultural background starting with the Max Dehn and Poul Heegaard [2] who first defined abstract simplicial complexes in 1907. Together with finite topological spaces as first considered by Pavel Alexandrov [1] in 1937, this produces a powerful framework. We pointed out some reasons in [10, 9] why we do not want to look at geometric realizations. An example is that the geometric realization of the double suspension of a Poincaré homology sphere is homeomorphic to a 5-sphere even so the combinatorial complexes are obviously not homeomorphic in a discrete sense. To remain in the finite topology allows to avoid difficulties from the continuum.

1.10.

The mathematics of Dehn,Heegard or Alexandrov is not only mathematically simpler as it is only combinatorics of finite sets of sets only, we can also branch off into area, where the continuum is awkward. For example, we can look at the cohomology of open sets in a simplicial complex. This is richer than the cohomology of closed sets. Since the Betti vector of a single set {x}\{x\} (which by definition is an open set {x}\{x\} with locally maximal xx in a simplicial complex) is the basis vectors e|x|e_{|x|}, where |x||x| is the cardinality of the set, we can realize any Betti vector with open sets. We would not know how to realize a given Betti vector with a simplicial complex. In the continuum, the cohomology of an open set is usually only considered in the sense of a limit of cohomologies of compact subsets. Even the homotopy is different. An open interval {{1,2}}\{\{1,2\}\} for example has Euler characteristic 1-1 and is not homotopic to a single point {{1}}\{\{1\}\} which has Euler characteristic 11. These two spaces should not be considered homotopic, for any sensible definition of homotopy. But we can identify A={{1,2}}A=\{\{1,2\}\} for example with B={{2},{1,2},{2,3}}B=\{\{2\},\{1,2\},\{2,3\}\} which is also an open interval. The open sets A,BA,B are homeomorphic. Their boundaries δA=A¯A={{1},{2}}\delta A=\overline{A}\setminus A=\{\{1\},\{2\}\} and δB=A¯B={{1},{2}}\delta B=\overline{A}\setminus B=\{\{1\},\{2\}\} are both 0-spheres. Unlike for general finite set of sets (sometimes called multi-graphs), the cohomology and topology works well for both open and closed sets. Almost everything we describe here fails for general set of sets which are neither closed nor open.

2. Notations

2.1.

A finite abstract simplicial complex GG is a finite set of sets, closed under the operation of taking non-empty subsets. GG carries a finite topology 𝒪\mathcal{O}, in which the set of all stars U(x)={y,xy}U(x)=\{y,x\subset y\} is the topological base. If V=xxV=\bigcup_{x}x is the vertex set, the collection of vertex stars U(x)U(x) with x={v}x=\{v\} with vVv\in V are a topological subbase which when closed under intersection produces the topological base and when closed under intersection and union produces the topology. The closed sets in the topology are the sub-simplicial complexes. If xyx\subset y, then U(y)U(x)U(y)\subset U(x). A map f:GHf:G\to H between simplicial complexes is continuous if f1(U)f^{-1}(U) is open in GG if UU is open in HH. A simplicial map is a map from V(G)V(G) to V(H)V(H) that lifts to an order preserving map GHG\to H. Simplicial maps are continuous; however not all continuous maps are simplicial maps. An example is a constant map onto a simplex cc of positive dimension.

2.2.

The topology on GG is Kolmogorov (T0) but neither Fréchet (T1), nor Hausdorff (T2). As any finite topology, it is Alexandroff, meaning that there are smallest neighborhoods U(x)U(x) of every xGx\in G. These are the stars U(x)={yG,xy}U(x)=\{y\in G,x\subset y\}. The topology is Zariski type because the closed sets in the topology agreeing with sub-simplicial complexes of GG. In the case when GG is the Whitney complex of a graph (V,E)(V,E), formed by the vertex sets of complete subgraphs, then subgraphs define closed sets. Every GG again defines a graph, where Γ=(V,E)={(x,y),xy,or,yx}\Gamma=(V,E)=\{(x,y),x\subset y\;,{\rm or}\;,\;y\subset x\}. The Whitney complex GG of this graph is the Barycentric refinement G1G_{1} of GG. The topology has the desired connectivity properties of GG. The just described graph obtained from GG has the same connectivity properties than the topology of GG. The topology induced from the geodesic distance of the graph would produce the discrete topology on GG and render GG completely disconnected. The non-Hausdorff property is inevitable. In the case when GG comes from a graph Γ\Gamma, the graph Γ\Gamma is the Čech nerve of the topological subbase of vertex stars. The Čech nerve of the topological base of all stars is Γ1\Gamma_{1}, the Barycentric refinement of Γ\Gamma.

2.3.

Elements in GG also known as simplices or faces or simply called sets. GG is is a set of sets of VV and the topology is a set of sets in 𝒪\mathcal{O}. The closure A¯\overline{A} of an arbitrary set AGA\subset G is the smallest closed set in GG containing AA. We need to distinguish three different things: (i) the set xGx\in G as an element or point of GG, (ii) the subset A(x)={x}GA(x)=\{x\}\subset G and (iii) its closure K(x)={x}¯={y,yx}GK(x)=\overline{\{x\}}=\{y,y\subset x\}\subset G, which is the simplicial complex generated by {x}\{x\}. The set A(x)A(x) is open only if xx is a locally maximal simplex (meaning not contained in any larger simplex) and closed only if xx has dimension 0. In general, the set A(x)={x}GA(x)=\{x\}\subset G is neither open nor closed but xx always defines two natural sets, the open set U(x)U(x) and the closed set K(x)K(x). The open set U(x)U(x) is the smallest open set containing xx, the closed set K(x)K(x) is the smallest closed set containing xx. U(x)U(x) is the star and K(x)K(x) the core. The closure B(x)B(x) of U(x)U(x) is the unit ball, and its boundary S(x)=B(x)U(x)S(x)=B(x)\setminus U(x) of U(x)U(x) is the unit sphere. The dimension dim(x){\rm dim}(x) of xGx\in G is defined as |x|1|x|-1, where |x||x| is the cardinality of xx. We write X=(x1,,xk)GkX=(x_{1},\dots,x_{k})\in G^{k} to address a kk-tuple of points xjGx_{j}\in G, and define w(x)=(1)dim(x)w(x)=(-1)^{{\rm dim}(x)} and w(X)=j=1kw(xj)w(X)=\prod_{j=1}^{k}w(x_{j}). We write shorthand X\bigcap X for j=1kxj\bigcap_{j=1}^{k}x_{j}. Similarly, we write X\bigcup X for j=1kxj\bigcup_{j=1}^{k}x_{j}. We usually do not care about the order of the elements in XX but allow that the same element appears multiple times. The configuration X=(x,x,x,,x)X=(x,x,x,\dots,x) for example is a kk-point configuration in which all points are the same.

2.4.

We often look at functions h:GRh:G\to R, where RR is some algebraic object like a ring. It should have an additive structure that is commutative (with respect to addition). It could be \mathbb{Z} or a finite Abelian group for example, it can also have more structure like a vector space over a field or an operator algebra. Here we look at rings like R=R=\mathbb{Z} but the multiplicative structure of the ring does not really enter. It could be a division algebra like the quaternions for example, where the multiplication is not commutative. As we have expressions like w(x)wm(U(x))w(x)w_{m}(U(x)) in our main result, a multiplication with 11 or 1-1. If RR is an additive group, then r-r denotes the additive inverse of rr in the group. As the theme is part of a finitist approach to mathematics, we prefer finite objects. Even in the case r=r=\mathbb{Z}, the range of a function f:GRf:G\to R is always finite so that we still deal with finite objects, despite the fact that there is no a priori cap on the size of RR. Having a ring rather than only an additive group has the advantage that on can also look at objects like determinants (or in the non-commutative ring case Dieudonné determinants) which can play a role in the vicinity of what we do here. Having a ring is a familiar frame work in other parts of mathematics.

2.5.

Rather than have a fixed ring RR it is possible to attach a ring R(x)R(x) to every open set R(U)R(U). The ring R(x)R(x) is called the ring of sections. To fix the relations, one needs restriction maps. Because the open sets U(x)U(x) are minimal, the ring R(x)R(x) can be called the stalk at xx. Its elements are the germs at xx. Since xyx\subset y implies U(y)U(x)U(y)\subset U(x), a sheaf is determined by giving restriction maps r(x,y):U(x)U(y)r(x,y):U(x)\to U(y) satisfying pre-sheaf properties r(x,x)=Idr(x,x)=Id and r(y,z)r(x,y)=r(x,z)r(y,z)\circ r(x,y)=r(x,z) if xyzx\subset y\subset z. Having all the germs R(x)R(x) and the transition maps fixed, one already has the existence (called gluing) and uniqueness (called locality) which are necessary to have a sheaf and not only a pre-sheaf. An example is w(x)=(1)dim(x)w(x)=(-1)^{{\rm dim}(x)} with restriction maps r(x,y)=w(x)w(y)r(x,y)=w(x)w(y) for xyx\subset y. There is a unique extension of ww to all open sets w(U)=xUw(x)w(U)=\prod_{x\in U}w(x). In analogy to χ(U)=w1(U)=xUw(x)\chi(U)=w_{1}(U)=\sum_{x\in U}w(x), we have called this the Fermi characteristic ϕ(U)\phi(U) of UU [7]. It is equal to det(L){\rm det}(L) with the connection matrix L(x,y)=χ(K(x)K(y))L(x,y)=\chi(K(x)\cap K(y)), the inverse of the Green function matrix g(x,y)=w(x)w(y)χ(U(x)U(y))g(x,y)=w(x)w(y)\chi(U(x)\cap U(y)).

2.6.

For a general function h:GmRh:G^{m}\to R, where we write h(X)=h(x1,,xm)h(X)=h(x_{1},\cdots,x_{m}), we can define wm(G)=XGm,XGh(X)w_{m}(G)=\sum_{X\in G^{m},\bigcap X\in G}h(X) and more generally,

wm(A)=XAm,XAh(X)w_{m}(A)=\sum_{X\in A^{m},\bigcap X\in A}h(X)

for AGA\subset G for sets AA that are not necessarily simplicial complexes. No symmetry like that h(X)=h(Y)h(X)=h(Y) if XX and YY are permutations is assumed. We mostly take h(X)=j=1m(1)dim(x)h(X)=\prod_{j=1}^{m}(-1)^{{\rm dim}(x)} because this assures that wm(G)w_{m}(G) are combinatorial invariants, meaning invariant under Barycentric refinements. For kk arbitrary stars U(x1),,U(xm)U(x_{1}),\dots,U(x_{m}), define U(X)=j=1kU(xj)U(X)=\bigcap_{j=1}^{k}U(x_{j}) and ωm(U)=XGm,XGw(U(X))\omega_{m}(U)=\sum_{X\in G^{m},\bigcap X\in G}w(U(X)). The first characteristic w1(G)=xGw(x)w_{1}(G)=\sum_{x\in G}w(x) is the Euler characteristic of GG, the second characteristic w2(G)=x,y,xyGw(x)w(y)w_{2}(G)=\sum_{x,y,x\cap y\in G}w(x)w(y) is the Wu characteristic and the third characteristic is

w3(G)=x,y,z,xyzGw(x)w(y)w(z).w_{3}(G)=\sum_{x,y,z,x\cap y\cap z\in G}w(x)w(y)w(z)\;.

While for any subsets A,BA,B, we have for m=1m=1 the property w1(AB)=w1(A)+w1(B)w1(AB)w_{1}(A\cup B)=w_{1}(A)+w_{1}(B)-w_{1}(A\cap B), this valuation formula fails for m>1m>1 for general A,BA,B, even for closed A,BA,B in general. But wm(G)w_{m}(G) is a multi-linear valuation if extended to wm(G1,,Gm)w_{m}(G_{1},\dots,G_{m}) with GjGG_{j}\subset G arbitrary subsets of GG. Now Gjwm(G1,,Gm)G_{j}\to w_{m}(G_{1},\dots,G_{m}) with all other sets Gi,ijG_{i},i\neq j fixed, satisfies linearity ω(A)+ω(B)=ωm(AB)+ωm(AB)\omega(A)+\omega(B)=\omega_{m}(A\cup B)+\omega_{m}(A\cap B).

2.7.

The 1-point complex is defined to be contractible. If GG is a complex and both GU(x)G\setminus U(x) and S(x)S(x) are contractible, then GG is called contractible. Two complexes G,HG,H which can morphed into each other by homotopy reductions and extensions are called homotopic. This is an equivalence relation on the space of all complexes. Homotopy does not honor dimension and so is not topological. Homotopy preserves w1(G)w_{1}(G) but not wm(G)w_{m}(G) with m>1m>1. The higher characteristics are topological invariants in the sense that they are preserved under homeomorphisms, an other equivalence relation: HH is called a continuous image of GG if there exists a Barycentric refinement GmG_{m} and a continuous map f:GmHf:G_{m}\to H (of course using the finite topology defined above), such that f1(S(x))f^{-1}(S(x)) is homeomorphic to S(x)S(x) (inductively defined as the maximal dimension of S(x)S(x) is smaller than the maximal dimension of GG) and such that for every locally maximal xGx\in G of dimension dd, the complex f1(B(x))f^{-1}(B(x)) is a dd-ball. GG and HH are homeomorphic, if GG is a continuous image of HH and HH is a continuous image of GG. A dd-ball is a complex of the form GU(x)G-U(x), where GG is a dd-sphere. A dd-sphere is a dd-manifold GG such that GU(x)G-U(x) is contractible. A dd-manifold GG is a complex such that for all xx, the unit sphere S(x)S(x) is a (d1)(d-1)-sphere. The empty complex is the (1)(-1)-sphere.

3. Energy theorem

3.1.

We assume m,k1m,k\geq 1 are integers. Define the mm’th order potential energy of the kk-point configuration XGkX\in G^{k} as gm(X)=w(X)wm(U(X))g_{m}(X)=w(X)w_{m}(U(X)). By design, it is zero if U(X)=U(xj)=U(X)=\bigcap U(x_{j})=\emptyset which is the case if at least one of the points is out of reach of the others. Even for non-intersecting x,yx,y it can happen that U(x)U(y)U(x)\cap U(y) is non-empty like if x,yx,y are zero-dimensional parts of a higher dimensional simplex zz, where zz is in U(x)U(y)U(x)\cap U(y). This is a case where x,yx,y can not be separated by open sets. Our goal is to have for any k1k\geq 1 and any m1m\geq 1, the mm’th characteristic can be expressed using kk-point Green function entries gm(x1,,xk)=j=1kw(xj)wm(j=1kU(xj))g_{m}(x_{1},\dots,x_{k})=\prod_{j=1}^{k}w(x_{j})w_{m}(\bigcap_{j=1}^{k}U(x_{j})). We think of this as the mm-th potential energy of the kk-point configuration X=(x1,,xk)X=(x_{1},\dots,x_{k}).

3.2.

Our main theorem tells that the mm’th characteristic is the total energy over all possible kk’point configurations.

Theorem 1 (Energy).

wm(G)=XGkgm(X)w_{m}(G)=\sum_{X\in G^{k}}g_{m}(X).

Proof.

The theorem is simpler to prove if formulated more generally like if

hm(x1,,xm)=j=1mw(xj)h_{m}(x_{1},\dots,x_{m})=\prod_{j=1}^{m}w(x_{j})

is replaced by a general function hm(x1,,xm)h_{m}(x_{1},\dots,x_{m}) of mm variables. The reason is that the map hmgkh_{m}\to g_{k} is linear so that we only need to verify the statement in the simplest possible case where hh is 11 only for a single configuration Z=(z1,,zm)Z=(z_{1},\dots,z_{m}) and 0 else. The left hand side is then 11. On the right hand side, we have to look at all X=(x1,,xk)X=(x_{1},\dots,x_{k}) for which ZU(X)=U(x1)U(x2)U(xm)Z\in U(X)=U(x_{1})\cap U(x_{2})\cap\cdots\cap U(x_{m}). As we will see however below, that condition will assure that ZU(xj)\bigcup Z\subset U(x_{j}) for all j=1,,mj=1,\dots,m. Therefore, XGkw(X)wm(U(X))=j=1kxjGw(xj)wm(U(xj))=1\sum_{X\in G^{k}}w(X)w_{m}(U(X))=\prod_{j=1}^{k}\sum_{x_{j}\in G}w(x_{j})w_{m}(U(x_{j}))=1. ∎

3.3.

A second major point is the sphere formula for S(X)=δU(X)=B(x)U(x)S(X)=\delta U(X)=B(x)\setminus U(x).

Theorem 2 (Sphere formula).

0=XGkw(X)wm(S(X))0=\sum_{X\in G^{k}}w(X)w_{m}(S(X)).

This means that the total energy of all unit spheres of even configurations is the same than the total energy of all unit spheres of odd configurations.

Proof.

The energy theorem also works of U(X)U(X) is replaced by B(X)B(X), which is the closure of U(X)U(X). The two equations

0=XGkw(X)wm(U(X))0=\sum_{X\in G^{k}}w(X)w_{m}(U(X))
0=XGkw(X)wm(B(X))0=\sum_{X\in G^{k}}w(X)w_{m}(B(X))

and the local valuation formula

wm(U(X))(1)mwm(S(X))=wm(B(X))w_{m}(U(X))-(-1)^{m}w_{m}(S(X))=w_{m}(B(X))\;

prove the theorem. ∎

3.4.

Let us add already a remark which however leads to an other story to which we hope to be able to write more about in the future: the theorem also works for the dual spheres X^=S(x1)S(xk)\hat{X}=S(x_{1})\cap\cdots\cap S(x_{k}) which is different from S(X)=δ(U(x1)U(xk))S(X)=\delta(U(x_{1})\cap\cdots\cap U(x_{k})).

0=XGkw(X)wm(X^).0=\sum_{X\in G^{k}}w(X)w_{m}(\hat{X})\;.

3.5.

The dual spheres X^\hat{X} play an important role in other places like in graph coloring. In the case when GG is a dd-manifold and if all points in XX are adjacent in the metric in which two points x,yS(x)x,y\in S(x) have distance 11, then X^\hat{X} is always a (dk)(d-k)-sphere. If k=1k=1 then X^=S(x)\hat{X}=S(x), which is in the manifold case a (d1)(d-1)-sphere. For k=2k=2, and x,yS(x)x,y\in S(x) we have (x,y)^=S(x)S(y)\hat{(x,y)}=S(x)\cap S(y) is a (d2)(d-2)-sphere. We have looked at this earlier in the context of graphs, where X=(v1,,vk)X=(v_{1},\dots,v_{k}) are the vertices of a complete graph KkK_{k} and so is associated to a vertex xx in the Barycentric refined graph. If GG is a dd-manifold, then by definition X^=x^\hat{X}=\hat{x} is a (dk)(d-k)-sphere. We have seen this as a duality because vxS(v)=x^\bigcap_{v\in x}S(v)=\hat{x} and vx^S(v)=x\bigcap_{v\in\hat{x}}S(v)=x. This can be seen as a duality between (k1)(k-1)-spheres (the boundary sphere of a simplex) and (dk)(d-k)-spheres. Shifting kk, there is a duality between kk-spheres and (d+1k)(d+1-k)-spheres.

3.6.

Still dwelling on that, moving cohomology from simplices to spheres could allow then to see Poincaré duality more elegantly, avoiding concepts like CW complexes which are necessary already in elementary setups like that the cube is the dual of the octahedron. When the cube is seen as a 2-sphere, it by definition is not a simplicial complex but only a more general CW complex if one wants to understand it as dual to the octahedron. So, if one wants to avoid the continuum (as we do), one usually goes to CW complexes. Poincaré himself already struggled quite a bit with the difficulty that the dual of a simplicial complex is not a simplicial complex. We will see the duality in the context of δ\delta-sets which generalize simplicial complexes too but more naturally than CW complexes. δ\delta sets are more general than simplicial sets, a construct which has a bit more structure than δ\delta sets. But simplicial sets as a subclass of δ\delta sets less accessible. Entire articles have been written just to explain the definition.

3.7.

Here is an other remark, maybe more for the mathematical physics minded: motivated by the Fock picture in which GkG^{k} are considered as kk-particle configurations, we could sum up the theorem over kk and get for example

wm(G)=k=112kXGkgm(X).w_{m}(G)=\sum_{k=1}^{\infty}\frac{1}{2^{k}}\sum_{X\in G^{k}}g_{m}(X)\;.

If we would replace w(x)w(x) with h(x)=(1)dim(x)/2h(x)=(-1)^{{\rm dim}(x)}/2 and set h(X)=jh(xj)h(X)=\prod_{j}h(x_{j}) and still use U(X)=jU(xj)U(X)=\bigcap_{j}U(x_{j}), if X=(x1,,xk)X=(x_{1},\dots,x_{k}) and setting gm(X)=h(X)wm(U(X))g_{m}(X)=h(X)w_{m}(U(X)), this would allow to write wm(G)=Xgm(X)w_{m}(G)=\sum_{X}g_{m}(X) where XX runs over all particle configurations ranging from single particles k=1k=1 to pair interactions k=2k=2, triple interactions k=3k=3 etc. The total energy of space is then the sum over all interaction energies overall possible particle configurations XX. If we think of gm(X)g_{m}(X) as the potential energy of the particle configuration XX, then wm(G)w_{m}(G) is the sum over all possible potential energies which particle configurations which can be realized in GG.

3.8.

Here are special cases, written out in more detail. First of all, lets write down the definitions of Euler characteristic

w1(G)=xGw(x)w_{1}(G)=\sum_{x\in G}w(x)

and Wu characteristic

w2(G)=x,yG2,xyw(x)w(y).w_{2}(G)=\sum_{x,y\in G^{2},x\cap y\neq\emptyset}w(x)w(y)\;.

i) For m=1m=1, we have expressions for the Euler characteristic:
w1(G)=xw(x)w1(U(x))=xg1(x)w_{1}(G)=\sum_{x}w(x)w_{1}(U(x))=\sum_{x}g_{1}(x)
w1(G)=x,yw(x)w(y)w1(U(x)U(y))=x,yg1(x,y)w_{1}(G)=\sum_{x,y}w(x)w(y)w_{1}(U(x)\cap U(y))=\sum_{x,y}g_{1}(x,y)
w1(G)=x,y,zw(x)w(y)w(z)w1(U(x)U(y)U(z))=x,y,zg1(x,y,z)w_{1}(G)=\sum_{x,y,z}w(x)w(y)w(z)w_{1}(U(x)\cap U(y)\cap U(z))=\sum_{x,y,z}g_{1}(x,y,z)
w1(G)=x,y,z,wg1(x,y,z,w)w_{1}(G)=\sum_{x,y,z,w}g_{1}(x,y,z,w)
The first two expressions for k=1k=1 and k=2k=2 have appeared in [7]. The inverse of the matrix g1(x,y)g_{1}(x,y) is L(x,y)=1L(x,y)=1 if x=yx=y and L(x,y)=0L(x,y)=0 if xyx\neq y.
ii) Next come expressions for the Wu characteristic:
w2(G)=xw(x)w2(U(x))=xg2(x)w_{2}(G)=\sum_{x}w(x)w_{2}(U(x))=\sum_{x}g_{2}(x)
w2(G)=x,yGw(x)w(y)w2(U(x)U(y))=x,yg2(x,y)w_{2}(G)=\sum_{x,y\in G}w(x)w(y)w_{2}(U(x)\cap U(y))=\sum_{x,y}g_{2}(x,y)
w2(G)=x,y,zGw(x)w(y)w(z)w2(U(x)U(y)U(z))=x,yg2(x,y,z)w_{2}(G)=\sum_{x,y,z\in G}w(x)w(y)w(z)w_{2}(U(x)\cap U(y)\cap U(z))=\sum_{x,y}g_{2}(x,y,z)
w2(G)=x,y,z,wg2(x,y,z,w)w_{2}(G)=\sum_{x,y,z,w}g_{2}(x,y,z,w)
iii) And here are expressions for the third characteristic
w3(G)=xw(x)w3(U(x))=xg3(x)w_{3}(G)=\sum_{x}w(x)w_{3}(U(x))=\sum_{x}g_{3}(x)
w3(G)=x,yw(x)w(y)w3(U(x)U(y))=x,yg3(x,y)w_{3}(G)=\sum_{x,y}w(x)w(y)w_{3}(U(x)\cap U(y))=\sum_{x,y}g_{3}(x,y)
w3(G)=x,y,zw(x)w(y)w(z)w3(U(x)U(y))U(z))=x,y,zg3(x,y,z)w_{3}(G)=\sum_{x,y,z}w(x)w(y)w(z)w_{3}(U(x)\cap U(y))\cap U(z))=\sum_{x,y,z}g_{3}(x,y,z)
w3(G)=x,y,z,wg3(x,y,z,w)w_{3}(G)=\sum_{x,y,z,w}g_{3}(x,y,z,w)

Remarks: 1) The theorem works if h(X)=w(X)h(X)=w(X) is replaced by any RR-valued interaction function h:GkRh:G^{k}\to R. We would for example set w2(A)=x,y,xyAh(x,y)w_{2}(A)=\sum_{x,y,x\cap y\in A}h(x,y) and get this equal to x,yG2g2(x,y)\sum_{x,y\in G^{2}}g_{2}(x,y) with

g2(x,y)=w(x)w(y)w2(U(x)U(y))g_{2}(x,y)=w(x)w(y)w_{2}(U(x)\cap U(y))\;

or equal to xGg1(x)\sum_{x\in G}g_{1}(x) with g1(x)=w(x)w2(U(x))g_{1}(x)=w(x)w_{2}(U(x)).
2) The energy theorems in the case of k=1k=1 are of Gauss-Bonnet type because we attach a fixed curvature gm(x)g_{m}(x) to a point. We can actually interpret this also as a Poincaré-Hopf theorem. There is some duality here as U(x)U(x) can be seen dual to K(x)K(x).
3) The energy formula reduces the time for the computation of the characteristic substantially. Especially for k=1k=1, where we have only to compute the mm’th characteristic of the n=|G|n=|G| stars U(x)U(x).
4) Instead of self-interactions of all GjG^{j}, we could take open sets GjGG_{j}\subset G and get expressions

wm(G1,,Gk)=X,xjGjgm(X).w_{m}(G_{1},\dots,G_{k})=\sum_{X,x_{j}\in G_{j}}g_{m}(X)\;.

For example, wm(A,B,C)=xA,yB,zCgm(x,y,z)w_{m}(A,B,C)=\sum_{x\in A,y\in B,z\in C}g_{m}(x,y,z), where

gm(x,y,z)=w(x)w(y)w(z)wm(U(x)U(y)U(z)).g_{m}(x,y,z)=w(x)w(y)w(z)w_{m}(U(x)\cap U(y)\cap U(z))\;.

Think of this as the total m’th characteristic energy of the three sets A,B,CA,B,C.
5) We have seen that in order to prove the theorems, it is helpful to energize more generally h(x1,,xm)h(x_{1},\dots,x_{m}) and set wm(G)=XGm,X>0h(X)w_{m}(G)=\sum_{X\in G^{m},X>0}h(X). The Green function procedure which maps functions h:GkRh:G^{k}\to R to functions g:GmRg:G^{m}\to R with gm(X)=w(X)wm(U(X))g_{m}(X)=w(X)w_{m}(U(X)) is linear. It is therefore enough to verify the claim for basis elements, where h(x) is zero except at h(x1,,xk)=1h(x_{1},\dots,x_{k})=1.

4. Valuation Lemma

4.1.

The following theorem is a main reason, why higher characteristics are topological invariants. The following valuation formula does not work for closed sets in general already if m2m\geq 2.

Lemma 1 (Valuation).

wm(UV)=wm(U)+wm(V)wm(UV)w_{m}(U\cup V)=w_{m}(U)+w_{m}(V)-w_{m}(U\cap V) for all open U,VU,V and all m1m\geq 1.

4.2.

The key why this works is:

Lemma 2 (Patching).

Given XGkX\in G^{k} and U,VU,V are open, define z=Xz=\bigcap X. If zUVz\in U\cap V, then both x,yx,y are simultaneously in U,VU,V and so xUVx\in U\cap V and yUUy\in U\cap U.

Patching.

If z=Xz=\bigcap X is in UVU\cap V, then every simplex ww which contains zz is both in UU and VV and so in UVU\cap V so that ww is in UU as well as in VV and UVU\cap V. Because this works for any ww, it works for any of the points xjx_{j}. Since all xjx_{j}are in U,V,UVU,V,U\cap V also the union z=Xz=\bigcup X is in all three sets. ∎

Valuation.

Counting XX in UVU\cup V is the sum of the counts in UU and the counts in VV with a double count if XX is in the intersection. ∎

Remark:
1) For m>1m>1, this in general does not hold for closed sets nor for a mixture of closed and open sets if m>1m>1. We will see for general mm that for an open set A=U(x)A=U(x), a closed disjoint set B=S(x)B=S(x) and closed union B(x)B(x) and empty intersection we have ωm(U(x))ωm(S(x))=ωm(B(x))\omega_{m}(U(x))-\omega_{m}(S(x))=\omega_{m}(B(x)) but for odd mm that ωm(U(x))+ωm(S(x))=ωm(B(x))\omega_{m}(U(x))+\omega_{m}(S(x))=\omega_{m}(B(x)).
a) For G={(1),(2),(3),(12),(23)}G=\{(1),(2),(3),(12),(23)\}, the two closed sets
A={(1),(2),(12)},B={(2),(3),(23)}A=\{(1),(2),(12)\},B=\{(2),(3),(23)\} intersect in C={(2)}C=\{(2)\}. Now w2(A)=w2(B)=1w_{2}(A)=w_{2}(B)=-1 now ω2(A)+ω2(B)=w2(G)ω2(S)\omega_{2}(A)+\omega_{2}(B)=w_{2}(G)-\omega_{2}(S).
b) If GG is the octahedron complex which is the join C4{a,b}C_{4}\oplus\{a,b\} and A=GU(a),B=GU(b)A=G\setminus U(a),B=G\setminus U(b) which are both balls intersecting in the circle S={x1,x2,x3,x4,(x1x2),(x2x3),(x3x4),(x4x1)}S=\{x_{1},x_{2},x_{3},x_{4},(x_{1}x_{2}),(x_{2}x_{3}),(x_{3}x_{4}),(x_{4}x_{1})\} which is also closed. The pair (ax1)A(ax_{1})\subset A intersects with the pair (bx1)B(bx_{1})\subset B but these two pairs are both not in SS. Still w1(A)+w2(B)=w2(G)+w2(S)w_{1}(A)+w_{2}(B)=w_{2}(G)+w_{2}(S).
Example a) shows that looking for more general relations is rather pointless as it depends on mm and the dimension.
2) In the special case m=1m=1, where we deal with Euler characteristic, the valuation formula holds for all sets because there is no interaction between points. This is the reason that for m=1m=1, we have a homotopy invariant while for m>1m>1 we have topological invariants.

4.3.

The local valuation formulas link two closed and an open set:

Lemma 3 (Local valuation).
wm(B(x))=wm(U(x))(1)mwm(S(x))w_{m}(B(x))=w_{m}(U(x))-(-1)^{m}w_{m}(S(x))\;

This works more generally for the open sets U(X)=U(xj)U(X)=\bigcap U(x_{j}) and B(X)=U(X)¯B(X)=\overline{U(X)} and S(X)=B(X)U(X)S(X)=B(X)\setminus U(X), the boundary. The reason is that U(X)U(X) is a disjoint union of disjoint open U(x)U(x).

5. The case of manifolds

5.1.

For xGx\in G, the unit ball B(x)=U(x)¯B(x)=\overline{U(x)} and the unit sphere S(x)=B(x)U(x)S(x)=B(x)\setminus U(x) are both closed sets and so simplicial complexes. A complex GG is called a dd-manifold if every S(x)S(x) is a (d1)(d-1)-sphere. A dd-manifold is a dd-sphere, if there exists xx such that GU(x)G\setminus U(x) is contractible. In the case when GG is a dd-sphere, GU(x))G\setminus U(x)) is declared to be a dd-ball. A complex GG is called contractible if there exists xGx\in G such that both S(x)S(x) and GU(x)G\setminus U(x) are contractible. These definitions are recursive with respect to the dimension dd. The foundation is laid by assuming the void 0=0=\emptyset is the (1)(-1)-sphere and that the one-point complex called unit 1={1}1=\{1\} is the smallest contractible complex.

5.2.

A dd-manifold with boundary is a complex GG such that every S(x)S(x) is either a (d1)(d-1)-sphere or a (d1)(d-1) ball. The boundary δG={xG,S(x)ball}\delta G=\{x\in G,S(x)\;{\rm ball}\} is a closed subset of GG. (Just check that if xδGx\in\delta G and zxz\subset x then S(z)S(z) is a ball). It is a (d1)(d-1) manifold because for zδGz\in\delta G the set S(z)δGS(z)\cap\delta G is a sphere because B(z)δGB(z)\cap\delta G is a (d1)(d-1) unit ball. We have to distinguish here between “closed= no boundary” and “closed=topologically closed”. We also want to deal with the case of open manifolds without boundary like an open disk or topologically closed manifolds with boundary which are not closed as manifolds. Open manifolds model also infinite manifolds like in the continuum, where for example the contangent correspondence allows to see the space (1,1)(-1,1) to be naturally homeomorphic to \mathbb{R}. Let us use the notation UU for an open manifold without boundary, GG for the closure and SS the boundary. So, S=δGS=\delta G is the topological boundary.

5.3.

We can verify wm(B(x))=wm(U(x))(1)mwm(S(x))w_{m}(B(x))=w_{m}(U(x))-(-1)^{m}w_{m}(S(x)) in the case of manifolds without boundary.

Lemma 4 (Local data in manifold case).

wm(U(x))=(1)dmw_{m}(U(x))=(-1)^{dm}, wm(S(x))=1+(1)d1w_{m}(S(x))=1+(-1)^{d-1} and wm(B(x))=(1)d(m+1)w_{m}(B(x))=(-1)^{d(m+1)}.

Theorem 3 (Manifolds MM).

(i) For even-dimensional manifolds with or without boundary wm(M)=w1(M)w_{m}(M)=w_{1}(M). (ii) For odd-dimensional manifolds with boundary wm(M)=w1(M)(1)mw1(S)w_{m}(M)=w_{1}(M)-(-1)^{m}w_{1}(S).

Proof.

For even dimensional manifolds with boundary SS, we have wm(U(x))=1w_{m}(U(x))=1 in the interior and wm(U(x))=0w_{m}(U(x))=0 at the boundary, so that the formula wm(M)=xω(x)wm(U)w_{m}(M)=\sum_{x}\omega(x)w_{m}(U) immediately shows that wm(M)=xw(x)=w1(M)w_{m}(M)=\sum_{x}w(x)=w_{1}(M). This also works for odd-dimensional manifolds without boundary. For odd-dimensional manifolds with boundary SS, we have wm(B(x))=1w_{m}(B(x))=1 in the interior and wm(B(x))=(1)mw_{m}(B(x))=-(-1)^{m} at the boundary. So wm(M)=w1(M)(1)mw1(S)w_{m}(M)=w_{1}(M)-(-1)^{m}w_{1}(S) with or without boundary.
In the special case if the manifold has no boundary, we also can use the energy theorem wm(G)=xw(x)wm(U(x))w_{m}(G)=\sum_{x}w(x)w_{m}(U(x)), The previous lemma wm(U(x))=w(x)m=(1)dmw_{m}(U(x))=w(x)^{m}=(-1)^{dm} shows that if dd is even, then wm(U(x))=1w_{m}(U(x))=1 and so wm(G)=xw(x)=w1(G)w_{m}(G)=\sum_{x}w(x)=w_{1}(G). If dd is odd, then w1(G)=xw(x)=0w_{1}(G)=\sum_{x}w(x)=0 and this does not change when multiplying with the constant wm(x)=(1)dmw_{m}(x)=(-1)^{dm}. ∎

6. Topological invariance

6.1.

A simplex xx is called locally maximal if it is not contained in any larger simplex. This means that the star U(x)={x}U(x)=\{x\} of a locally maximal point therefore has only one point xx in GG. A map f:GHf:G\to H between simplicial complexes is called continuous if f1(U)f^{-1}(U) is open in GG if UU is open in HH. A complex HH is a topological image of GG if there exists a continuous f:GHf:G\to H such that f1(S(x))f^{-1}(S(x)) is homeomorphic to S(x)S(x) for all xx and f1(B(x))f^{-1}(B(x)) is a dim(x)dim(x)-ball for every unit ball B(x)B(x) with locally maximal xx in HH. If GG and HH are both topological images of each other, the two complexes are called homeomorphic. Also this definition is recursive. It relies on homeomorphism in dimension (d1)(d-1). Quantities which are constant on homeomorphism classes are called topological invariants.

6.2.

In comparison, a quantity is a combinatorial invariant if it stays the same under Barycentric refinements. We have in this context say what we mean with the Barycentric refinement of an open set. We can not just take the Whitney complex of the incidence graph as in the case of closed sets because the Whitney complex is a simplicial complex. With that definition, the Barycentric refinement of an open set would be a closed set. What we can do is to see an open set UU of a complex GG and define U1U_{1} as the complement of the Barycentric refinement (GU)1(G\setminus U)_{1} in the Barycentric refinement G1G_{1} of GG. Bott defined combinatorial invariant as a quantity that does not change when making Barycentric refinements. For all m1m\geq 1 we have:

Theorem 4 (Invariance).

wmw_{m} are topological invariants. wmw_{m} are combinatorial invariants.

Proof.

First check that wmw_{m} is the same on all d-balls BB. This follows from the general manifold formula because BB is a manifold with boundary. For odd mm we have wm(B)=w1(B)w_{m}(B)=w_{1}(B). For even mm, we have wm(B)=w1(B)w1(S)w_{m}(B)=w_{1}(B)-w_{1}(S), where SS is the boundary. The fact that wmw_{m} is invariant under Barycentric refinement is treated in the next section about the topological product. Given two general G,HG,H which are homeomorphic. By definition, there is a continuous map from GnG_{n} to HH. Now use the Ball formula wm(H)=xw(x)wm(B(x))w_{m}(H)=\sum_{x}w(x)w_{m}(B(x)). We claim that this is the same than wm(G)=f1(B(x))w(x)wm(f1(B(x)))w_{m}(G)=\sum_{f^{-1}(B(x))}w(x)w_{m}(f^{-1}(B(x)))

6.3.

The property of being a Dehn-Sommerville space (as defined above again) is topological. If G,HG,H are homeomorphic and GG is Dehn-Sommerville, then HH is Dehn-Sommerville. The proof can be done by induction. Having verified it for dimension up to dd, then we have it for dimension d+1d+1 because there is a topological correspondence of unit spheres making sure that all unit spheres are Dehn-Sommerville.

6.4.

As an other question we should mention how to define homotopy for open sets. One idea two call two open sets U,VU,V to be homotopic in GG, if Uc=GUU^{c}=G\setminus U and VcV^{c} are homotopic closed sets. But we would like to have that if U,VU,V are homotopic, then the Betti vectors b(U)b(U) and b(V)b(V) should be the same. An open 3-ball for example has Betti vector (0,0,0,1)(0,0,0,1) and a closed 3-ball has Betti vector (1,0,0,0)(1,0,0,0). They should already not be considered homotopic because the Euler characteristic does not match. All this will hopefully be explored more in a future paper on the matter.

7. Topological product

7.1.

The topological product GHG\cdot H of two simplicial complexes G,HG,H is the Whitney complex of the graph in which the Cartesian product G×HG\times H are the vertices and where (a,b)(a,b) and (c,d)(c,d) are connected either ab,cda\subset b,c\subset d or if ba,dcb\subset a,d\subset c. It is a set of subsets of V=G×HV=G\times H and again a simplicial complex.

7.2.

Unlike other products like the Shannon product, the topological product is of a topological nature. It preserves manifolds and as we will see is compatible with all higher characteristics.

Lemma 5.

If GG and HH are manifolds, then GHG\cdot H is a manifold.

Proof.

Lets outline again the argument: we have to look at the unit sphere S((x,y))S((x,y)) of a point (x,y)(x,y) and show that it is a d1d-1-sphere. S((x,y))S((x,y)) is the union of two cylinders S(x)B(y)S(x)\cdot B(y) and B(x)S(y)B(x)\cdot S(y) which are glued together along S(x)S(y)S(x)\cdot S(y). In the case of a product of a 11-manifold GG and a 11-manifold HH for example, the unit spheres S(x,y)S(x,y) is a ”square” which is the union of S(x)×B(y)S(x)\times B(y) (left right parts) B(x)×S(y)B(x)\times S(y) (top bottom part) intersecting in 44 points S(x)×S(y)S(x)\times S(y). To see that this is a sphere, collapse B(y)B(y) to a point to see that S(x,y)S(x,y) is homotopic to the join of S(y)S(y) and S(x)S(x) which is a sphere. One then checks case by case that each point in S(x,y)S(x,y) has a d2d-2 sphere as unit sphere. For interior points in B(y)B(y) or B(x)B(x) this follows by induction. For points in S(x)×S(y)S(x)\times S(y) the unit sphere is a copy of two (d2)(d-2) balls glued along a (d3)(d-3) sphere and so a (d2)(d-2) sphere. ∎

7.3.

Similarly, one has by analyzing additionally the unit spheres of boundary points and knowing that the join of a ball with a sphere is a ball:

Lemma 6.

If G,HG,H are manifolds with boundary then GHG\cdot H is a manifold with boundary.

For example, the product of two closed intervals is a square, the product of an interval with a circle is a closed cylinder.

7.4.

The topological product is not associative. The complex G1G\cdot 1 is the Barycentric refinement G1G_{1} of GG so that G(11)=G1G\cdot(1\cdot 1)=G_{1} but (G1)1=G2(G\cdot 1)\cdot 1=G_{2} is the second Barycentric refinement.

7.5.

There is an algebraic Stanley-Reisner description (which is seen implemented in the code section). If the elements in V=xGxV=\bigcup_{x\in G}x are labeled with variables t1,,tqt_{1},\dots,t_{q}, then every xGx\in G can be written as a monomial t(x)=tj0tj2tjdim(x)t(x)=t_{j_{0}}t_{j_{2}}\dots t_{j_{{\rm dim}(x)}} and xyx\subset y is algebraically encoded by t(x)|t(y)t(x)|t(y). The complex GG is a polynomial xGt(x)\sum_{x\in G}t(x). In the same way, the complex HH is given by the polynomial yHt(y)\sum_{y\in H}t(y). The product GHG\cdot H is now represented by the product of these two polynomials. The vertex set of GHG\cdot H has |G||H||G||H| elements.

7.6.

For example, if G={{1},{2},{1,2}}G=\{\{1\},\{2\},\{1,2\}\} and H={{3},{4},{3,4}}H=\{\{3\},\{4\},\{3,4\}\}. The product complex GHG\cdot H is a complex for which the vertex set has 99 elements.

Theorem 5 (Product).

wm(GH)=wm(G)wm(H)w_{m}(G\cdot H)=w_{m}(G)w_{m}(H).

Proof.

We have wm(G)=x1,,xm,X>0h(x1)h(xm)=Xh(X)w_{m}(G)=\sum_{x_{1},\dots,x_{m},\bigcap X>0}h(x_{1})\cdots h(x_{m})=\sum_{X}h(X). Similarly, wm(H)=y1,,ym,Y>0h(y1)h(ym)=Yh(Y)w_{m}(H)=\sum_{y_{1},\dots,y_{m},\bigcap Y>0}h(y_{1})\cdots h(y_{m})=\sum_{Y}h(Y). Now,

wm(GH)\displaystyle w_{m}(G\cdot H) =\displaystyle= X,YX>0,Y>0h(x1)h(xm)h(y1)h(ym)\displaystyle\sum_{X,Y\bigcap X>0,\bigcap Y>0}h(x_{1})\cdots h(x_{m})h(y_{1})\cdots h(y_{m})
=\displaystyle= X,Yh(X)h(Y)\displaystyle\sum_{X,Y}h(X)h(Y)
=\displaystyle= Xh(X)Yh(Y)=wm(X)wm(Y).\displaystyle\sum_{X}h(X)\sum_{Y}h(Y)=w_{m}(X)w_{m}(Y)\;.

7.7.

Remarks:
1) A special case is if H=1={{1}}H=1=\{\{1\}\}, where G1=G1G_{1}=G\cdot 1 is the Barycentric refinement. And wm(G1)=wm(G)w_{m}(G_{1})=w_{m}(G) is a special important case.
2) As in the case m=1m=1, if GG is the Whitney complex of a finite simple graph, we would like to know about the behavior of the curvature when taking products. In the case of the Shannon product and m=1m=1, we have seen that the curvatures multiply.

8. Code

8.1.

Here is some code. As usual, one can copy-paste it from the ArXiv’s LaTex source. The procedures should be pretty self-explanatory, given that the Wolfram language allows to write mathematical procedures in a form resembling pseudo code. As for the topological product, we repeat the algebraic frame work as we have used it [3], (before even knowing about Stanley-Reisner rings).

Closure[A_]:=If[A=={},{},Delete[Union[Sort[Flatten[Map[Subsets,A],1]]],1]];
Boundary[A_]:=Complement[Closure[A],A];
Whitney[s_]:=If[Length[EdgeList[s]]==0,Map[{#}&,VertexList[s]],
Map[Sort,Sort[Closure[FindClique[s,Infinity,All]]]]];
UU[G_,x_]:=Module[{U={}},
Do[If[SubsetQ[G[[k]],x],U=Append[U,G[[k]]]],{k,Length[G]}];U];
VV[G_,x_]:=Module[{U={}},
Do[If[SubsetQ[x,G[[k]]],U=Append[U,G[[k]]]],{k,Length[G]}];U];
Basis[G_]:=Table[UU[G,G[[k]]],{k,Length[G]}]; Stars=Basis;
Cores[G_]:=Table[VV[G,G[[k]]],{k,Length[G]}];
SubBasis[G_]:=Module[{V=Union[Flatten[G]]},Table[UU[G,{V[[k]]}],{k,Length[V]}]];
UnitSpheres[G_]:=Module[{B=Basis[G]},
Table[Complement[Closure[B[[k]]],B[[k]]],{k,Length[B]}]];
UnitBalls[G_]:=Map[Closure,Basis[G]];
Cl[U_,A_]:=Module[{V=U},Do[V=Union[Append[V,
Union[V[[k]],A[[l]]]]],{k,Length[V]},{l,Length[A]}];V];
Topology[G_]:=Module[{V=B=Basis[G]},
Do[V=Cl[V,B],{Length[Union[Flatten[G]]]}];Append[V,{}]];
GraphBasis[s_]:=Basis[Whitney[s]];
Nullity[Q_]:=Length[NullSpace[Q]]; sig[x_]:=Signature[x];
sig[x_,y_]:=If[SubsetQ[x,y]&&(Length[x]==Length[y]+1),
sig[Prepend[y,Complement[x,y][[1]]]]*sig[x],0];
Fvector[G_]:=If[Length[G]==0,{},Delete[BinCounts[Map[Length,G]],1]];
Ffunction[G_,t_]:=Module[{f=Fvector[G],n},Clear[t]; n=Length[f];
If[Length[VertexList[s]]==0,1,1+Sum[f[[k]]*t^k,{k,n}]]];
BarycentricGraph[s_]:=ToGraph[Whitney[s]];
BarycentricComplex[G_]:=Whitney[ToGraph[G]];
dim[x_]:=Length[x]-1; w[x_]:=(-1)^dim[x];
Wu1[A_]:=Total[Map[w,A]]; Chi=Wu1; EulerChi=Wu1;
Wu2[A_]:=Module[{a=Length[A]},Sum[x=A[[k]]; Sum[y=A[[l]];
If[MemberQ[A,Intersection[x,y]],1,0]*w[x]*w[y],{l,a}],{k,a}]];Wu=Wu2;
Wu3[A_]:=Module[{a=Length[A]},Sum[x=A[[k]];Sum[y=A[[l]];Sum[z=A[[o]];
If[MemberQ[A,Intersection[x,y,z]],1,0]*
w[x]*w[y]*w[z],{o,a}],{l,a}],{k,a}]];
RingFromComplex[G_,a_]:=Module[{V=Union[Flatten[G]],n,T,U},
n=Length[V];Quiet[T=Table[V[[k]]->a[[k]],{k,n}]];
Quiet[U=G /.T]; Sum[Product[U[[k,l]],
{l,Length[U[[k]]]}],{k,Length[U]}]];
ComplexFromRing[f_]:=Module[{s,ff},s={}; ff=Expand[f];
Do[Do[If[Denominator[ff[[k]]/ff[[l]]]==1 && k!=l,
s=Append[s,k->l]], {k,Length[ff]}],{l,1,Length[ff]}];
Whitney[UndirectedGraph[Graph[Range[Length[ff]],s]]]];
TopologicalProduct[G_,H_]:=Module[{},
f=RingFromComplex[G,”a”];
g=RingFromComplex[H,”b”]; F=Expand[f*g]; ComplexFromRing[F]];

8.2.

The following lines illustrate some of the identities for random complexes.

G=Whitney[RandomGraph[{9,15}]]; G = Sort[G]; G = Map[Sort, G]; n = Length[G];
U1=Basis[G]; U2=Table[Intersection[U1[[k]],U1[[l]]],{k,n},{l,n}];
U3=Table[Intersection[U1[[k]],U2[[l,m]]],{k,n},{l,n},{m,n}];
V1=Cores[G]; V2=Table[Intersection[V1[[k]],V1[[l]]],{k,n},{l,n}];
V3=Table[Intersection[V1[[k]],V2[[l,m]]],{k,n},{l,n},{m,n}];
S1=Map[Boundary,U1]; S2=Table[Boundary[U2[[k,l]]],{k,n},{l,n}];
S3=Table[Boundary[U3[[k,l,m]]],{k,n},{l,n},{m,n}];
B1=Map[Closure,U1]; B2=Table[Closure[U2[[k,l]]],{k,n},{l,n}];
B3=Table[Closure[U3[[k,l,m]]],{k,n},{l,n},{m,n}];
Print[Checkenergyformulas␣␣␣];
Wu1[G]==Total[Table[w[G[[k]]]*Wu1[U1[[k]]],{k,n}]];
Wu2[G]==Total[Table[w[G[[k]]]*Wu2[U1[[k]]],{k,n}]];
Wu1[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu1[U2[[k,l]]],{k,n},{l,n}]]]
Wu2[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu2[U2[[k,l]]],{k,n},{l,n}]]]
Wu1[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu1[U3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
Wu2[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu2[U3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
Print[Checkenergyballformulas␣␣];
Wu1[G]==Total[Table[w[G[[k]]]*Wu1[B1[[k]]],{k,n}]]
Wu2[G]==Total[Table[w[G[[k]]]*Wu2[B1[[k]]],{k,n}]]
Wu1[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu1[B2[[k,l]]],{k,n},{l,n}]]]
Wu2[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu2[B2[[k,l]]],{k,n},{l,n}]]]
Wu1[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu1[B3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
Wu2[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu2[B3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
Print[Checksphereformulas␣␣␣␣␣␣␣␣␣␣];
0==Total[Table[w[G[[k]]]*Wu1[S1[[k]]], {k, n}]]
0==Total[Table[w[G[[k]]]*Wu2[S1[[k]]], {k, n}]]
0==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu1[S2[[k,l]]],{k,n},{l,n}]]]
0==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu2[S2[[k,l]]],{k,n},{l,n}]]]
0==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu1[S3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
0==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*w[G[[m]]]*Wu2[S3[[k,l,m]]],
{k,n},{l,n},{m,n}]]]
Print[Checklocalvaluationformula␣␣];
Map[Wu1,U1]==Map[Wu1,B1]-Map[Wu1,S1]
Union[Flatten[Table[Wu1[U2[[k,l]]]+Wu1[S2[[k,l]]]-Wu1[B2[[k,l]]],{k,n},{l,n}]]]
Map[Wu2,U1]==Map[Wu2,B1]+Map[Wu2,S1]
Union[Flatten[Table[Wu2[U2[[k,l]]]-Wu2[S2[[k,l]]]-Wu2[B2[[k,l]]],{k,n},{l,n}]]]
Print[Checkproduct␣␣];
G=Whitney[RandomGraph[{6,10}]]; H=Whitney[StarGraph[5]];
GH=TopologicalProduct[G,H];
{Wu1[G],Wu1[H],Wu1[GH]}
{Wu2[G],Wu2[H],Wu2[GH]}
(* A bit more time consuming to compute *)
Wu3[G]==Total[Table[w[G[[k]]]*Wu3[B1[[k]]],{k,n}]]
Wu3[G]==Total[Flatten[Table[w[G[[k]]]*w[G[[l]]]*Wu3[B2[[k,l]]],{k,n},{l,n}]]]

References

  • [1] P. Alexandroff. Diskrete Räume. Mat. Sb. 2, 2, 1937.
  • [2] M. Dehn and P. Heegaard. Analysis situs. Enzyklopaedie d. Math. Wiss, III.1.1:153–220, 1907.
  • [3] O. Knill. The Künneth formula for graphs.
    http://arxiv.org/abs/1505.07518, 2015.
  • [4] O. Knill. On a Dehn-Sommerville functional for simplicial complexes.
    https://arxiv.org/abs/1705.10439, 2017.
  • [5] O. Knill. The counting matrix of a simplicial complex.
    https://arxiv.org/abs/1907.09092, 2019.
  • [6] O. Knill. Energized simplicial complexes. https://arxiv.org/abs/1908.06563, 2019.
  • [7] O. Knill. The energy of a simplicial complex. Linear Algebra and its Applications, 600:96–129, 2020.
  • [8] O. Knill. Green functions of energized complexes.
    https://arxiv.org/abs/2010.09152, 2020.
  • [9] O. Knill. Finite topologies for finite geometries. https://arxiv.org/abs/2301.03156, 2023.
  • [10] O. Knill. The sphere formula. https://arxiv.org/abs/2301.05736, 2023.