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

Fusion inequality for quadratic cohomology

Oliver Knill Department of Mathematics
Harvard University
Cambridge, MA, 02138
(Date: June 24, 2024)
Abstract.

Classical simplicial cohomology on a simplicial complex GG deals with functions on simplices xGx\in G. Quadratic cohomology [6, 7] deals with functions on pairs of simplices (x,y)G×G(x,y)\in G\times G that intersect. If K,UK,U is a closed-open pair in GG, we prove here a quadratic version of the linear fusion inequality [10]. Additional to the quadratic cohomology of GG there are five additional interaction cohomology groups. Their Betti numbers are computed from functions on pairs (x,y)(x,y) of simplices that intersect. Define the Betti vector b(X)b(X) computed from pairs (x,y)X×X(x,y)\in X\times X with xyXx\cap y\in X a and b(X,Y)b(X,Y) with pairs in X×YX\times Y with xyKx\cap y\in K. We prove the fusion inequality b(G)b(K)+b(U)+b(K,U)+b(U,K)+b(U,U)b(G)\leq b(K)+b(U)+b(K,U)+b(U,K)+b(U,U) for cohomology groups linking all five possible interaction cases. Counting shows f(G)=f(K)+f(U)+f(K,U)+f(U,K)+f(U,U)f(G)=f(K)+f(U)+f(K,U)+f(U,K)+f(U,U) for the f-vectors. Super counting gives Euler-Poincaré k(1)kfk(X)=k(1)kbk(X)\sum_{k}(-1)^{k}f_{k}(X)=\sum_{k}(-1)^{k}b_{k}(X) and k(1)kfk(X,Y)=k(1)kbk(X,Y)\sum_{k}(-1)^{k}f_{k}(X,Y)=\sum_{k}(-1)^{k}b_{k}(X,Y) for X,Y{U,K}X,Y\in\{U,K\}. As in the linear case, also the proof of the quadratic fusion inequality follows from the fact that the spectra of all the involved Laplacians L(X),L(X,Y)L(X),L(X,Y) are bounded above by the spectrum of the quadratic Hodge Laplacian L(G)L(G) of GG.

Key words and phrases:
Quadratic cohomology

1. In a nutshell

1.1.

We prove here that if KK is a sub-complex of a finite abstract simplicial complex GG and UU is the open complement U=GKU=G\setminus K [1, 12], there are besides the quadratic cohomology of GG five quadratic cohomology groups belonging to the five quadratic Hodge Laplacians L(X),L(X,Y)L(X),L(X,Y) with X,Y{U,K}X,Y\in\{U,K\}. They all satisfy spectral inequalities:

Theorem 1 (Spectral inequality).

λk(L)λk(L(G))\lambda_{k}(L)\leq\lambda_{k}(L(G))

1.2.

The assumption is that all eigenvalues are ordered in an ascending order and that they are padded left in comparision with the eigenvalues of GG. This result parallels the linear simplicial cohomology case [13], where UU and KK can not yet interact and LL is one of the Hodge Laplacians L(K),L(U)L(K),L(U) for simplicial cohomology.

1.3.

In the linear case, the Betti vectors satisfied the fusion inequality b(G)b(K)+b(U)b(G)\leq b(K)+b(U) [10]. This linear fusion inequality had followed from the spectral inequality and the fact that cohomology classes are null-spaces of matrices. Counting gave f(G)=f(K)+f(U)f(G)=f(K)+f(U) for the ff-vectors and the Euler-Poincaré formula was χ(X)=k(1)kfk(X)=k(1)kbk(X)\chi(X)=\sum_{k}(-1)^{k}f_{k}(X)=\sum_{k}(-1)^{k}b_{k}(X) for X{G,K,U}X\in\{G,K,U\}, seen directly by heat deformation using the McKean-Singer symmetry, rephrasing that DXD_{X} is an isomorphism between even and odd parts of image of the Laplacian LX=DX2L_{X}=D_{X}^{2}, implying str(Lk)=0{\rm str}(L^{k})=0 for k1k\geq 1 so that str(exp(tL(X))=str(1X)=χ(X){\rm str}(\exp(-tL(X))={\rm str}(1_{X})=\chi(X).

1.4.

In the quadratic cohomology case, where we look at functions on pairs of intersecting simplices, there is the cohomology of GG leading to b(G)b(G) and five interaction cohomologies. They each lead to Betti vectors. We call them b(K),b(U)b(K),b(U),b(K,U),b(U,K)b(K,U),b(U,K), and b(U,U)b(U,U). Besides pointing out that we have these new cohomologies, we give here a relation between them and the cohomology of GG. We call it the quadratic fusion inequality. We could use the heavier notation b(X,Y,Z)b(X,Y,Z) with X,Y,Z{K,U}X,Y,Z\in\{K,U\} dealing with functions on pairs (x,y)X×Y(x,y)\in X\times Y with xyZx\cap y\in Z, but we prefer to stick to the simpler notation: one reason is that both b(U),b(K)b(U),b(K) deal with internal cohomology of UU and KK and do not involve simplices of the other set, while b(K,U),b(U,K),b(U,U)b(K,U),b(U,K),b(U,U) involve both sets UU and KK. So, while the quadratic Betti vectors b(U),b(K)b(U),b(K) are intrinsic and only depend on one of the sets UU or KK, the others are not. We prove:

Theorem 2 (Quadratic fusion inequality).

b(G)b(K)+b(U)+b(K,U)+b(U,K)+b(U,U)b(G)\leq b(K)+b(U)+b(K,U)+b(U,K)+b(U,U).

1.5.

The quadratic Betti vector b(X)b(X) belongs to {(x,y),xyX}\{(x,y),x\cap y\in X\}, the vector b(K,U)b(K,U) belongs to all (x,y)K×U(x,y)\in K\times U with xyKx\cap y\in K, and b(U,K)b(U,K) belongs to all (x,y)U×K(x,y)\in U\times K with xyKx\cap y\in K, and b(U,U)b(U,U) belongs to all (x,y)U×U(x,y)\in U\times U with xyKx\cap y\in K. Note that xyKx\cap y\in K if x,yKx,y\in K so that b(K,K)b(K,K) is accounted for in b(K)b(K) already. With the heavier notation, from the eight cases b(X,Y,Z)b(X,Y,Z) with X,Y,Z{K,U}X,Y,Z\in\{K,U\} only b(K)=b(K,K,K),b(U)=b(U,U,U),b(K,U)=b(K,U,K),b(U,K)=b(U,K,K),b(U,U)=b(U,U,K)b(K)=b(K,K,K),b(U)=b(U,U,U),b(K,U)=b(K,U,K),b(U,K)=b(U,K,K),b(U,U)=b(U,U,K) can occur.

1.6.

The quadratic characteristics are defined as w(X)=x,yX,xyXw(x)w(y)w(X)=\sum_{x,y\in X,x\cap y\in X}w(x)w(y) (Wu characteristic) and w(X,Y)=xX,yY,xyKw(x)w(y)w(X,Y)=\sum_{x\in X,y\in Y,x\cap y\in K}w(x)w(y) and w(U,U)=xU,yU,xyUw(x)w(y)w(U,U)=\sum_{x\in U,y\in U,x\cap y\in U}w(x)w(y), we have w(G)=w(U)+w(K)+w(U,K)+w(K,U)+w(U,U)w(G)=w(U)+w(K)+w(U,K)+w(K,U)+w(U,U) which follows from the fact that ff-vectors for quadratic cohomology satisfy f(G)=f(K)+f(U)+f(K,U)+f(U,K)+f(U,U)f(G)=f(K)+f(U)+f(K,U)+f(U,K)+f(U,U). Any of the cohomologies for X{G,U,K,(U,K),(K,U),(U,U)}X\in\{G,U,K,(U,K),(K,U),(U,U)\} satisfy the Euler-Poincaré formula, following heat deformation with LXL_{X} or LX,YL_{X,Y} using the McKean-Singer symmetry.

1.7.

Why is this interesting? If GG is finite abstract simplicial complex that is a finite dd-manifold, and ff is an arbitrary function from GG to P={0,,k}P=\{0,\dots,k\}, then the discrete Sard theorem [11] assures that the open set U={xG,f(x)=P}U=\{x\in G,f(x)=P\} is either empty or a (dk)(d-k)-manifold in the sense that the graph with vertex set UU and edge set {(x,y)\{(x,y) for which xyx\subset y or yx}y\subset x\} is a discrete (dk)(d-k)-manifold. Since UU is open, the complement K=GUK=G\setminus U is closed and so a sub-simplicial complex. All these cohomologies are topological invariants.

1.8.

For example, if GG is a discrete 33-sphere and f:G{0,1,2}f:G\to\{0,1,2\} is arbitrary, then UU is either empty, a knot or a link, a finite union of closed disjoint (possibly interlinked) 11-manifolds in the 3-sphere GG. The simplicial cohomology of UU and the quadratic cohomology of UU are not interesting: they are just circles: b(U)=(l,l)b(U)=(l,l), where ll is the number of connected components of UU. The simplicial cohomology of KK however can be interesting and leads to knot or link invariants. It is well studied in the continuum as it is a knot invariant or a link invariant. Additional interaction cohomologies that take into account interaction between UU and the complement KK are completely unexplored. The inequality shows however that in general, more cohomology classes are created when splitting up GG into UKU\cup K. Unlike in the linear case, we have now the possibility of particles (harmonic forms) that are functions of (x,y)(x,y) with xU,yKx\in U,y\in K with xyKx\cap y\in K and also of function on (x,y)(x,y) with xU,yUx\in U,y\in U with xyKx\cap y\in K.

1.9.

The quadratic case we look at here would generalizes in a straightforward way to higher characteristics. One starts with m=1m=1, the linear characteristic which is Euler characteristic. The second m=2m=2 is quadratic characteristic or Wu characteristic going back to Wu in 1959. 111Historically, multi-linear valuations were considered in [14, 3, 6] and its cohomology in [8]. Quadratic cohomology is to Wu characteristic what simplicial cohomology is to Euler characteristic. In general, we would look for mm-tuples of points (x1,,xm)(x_{1},\dots,x_{m}) that do simultaneously intersect. There are then much more mm-point interaction cohomologies and b(G)b(G) is again bounded above by all possible cases of b(X1,,Xm)b(X_{1},\dots,X_{m}) with Xi{U,K}X_{i}\in\{U,K\} and the intersection in KK. The cases b(U,U,,U)b(U,U,\dots,U) has cases (x1,xm)Um(x_{1},\dots x_{m})\in U^{m} with xiU\bigcap x_{i}\in U which is part of bUb_{U} and a new part where xiK\bigcap x_{i}\in K. As in the case m=2m=2, the cohomology of b(K,K,,K)b(K,K,\dots,K) is part of b(K)b(K). We have to consider the Betti vector b(U)b(U) for functions on {(x1,,xm)Um,ixiU}\{(x_{1},\dots,x_{m})\in U^{m},\bigcap_{i}x_{i}\in U\} and b(U,,U)b(U,\dots,U) referring to functions on {(x1,,xm)Um,ixiK}\{(x_{1},\dots,x_{m})\in U^{m},\bigcap_{i}x_{i}\in K\}, where b(K,,K)=b(K)b(K,\dots,K)=b(K) in the notation used before. If all xix_{i} are in UU, then the intersection can be either in UU or KK, but if all xix_{i} are in KK, then the intersection must be in KK, not warranting to distinguish b(K,,K)b(K,\dots,K) and b(K)b(K). Despite the obvious duality UKU\leftrightarrow K, there is an asymmetry in that xU,yUx\in U,y\in U allows xyx\cap y to be in UU or KK while xK,yKx\in K,y\in K implies xyKx\cap y\in K: technically, a closed set KK is a π\pi-system while an open set UU is not unless it is \emptyset or GG.

Theorem 3.

b(G)b(U)+b(K)+Xi{U,K}b(X1,,Xm)b(G)\leq b(U)+b(K)+\sum_{X_{i}\in\{U,K\}}b(X_{1},\dots,X_{m}).

There would again be heavier notation b(X1,,Xm,X0)b(X_{1},\dots,X_{m},X_{0}) dealing with (x1,,xm)X1××Xm(x_{1},\dots,x_{m})\in X_{1}\times\cdots\times X_{m} with k=1mxkX0\bigcap_{k=1}^{m}x_{k}\in X_{0} and have b(G)Xi{U,K}b(X1,,Xm,X0)b(G)\leq\sum_{X_{i}\in\{U,K\}}b(X_{1},\dots,X_{m},X_{0}) taking into account that some of the cases like (K,,K,U)(K,\dots,K,U) are empty because X0=UX_{0}=U only is interesting if all X1==Xm=KX_{1}=\cdots=X_{m}=K. Again, like in the case m=2m=2, we have two cases b(U),b(K)b(U),b(K) which are intrinsic while the other cases involve simplices from both UU and KK.

1.10.

We study here higher order chain complexes in finite geometries. Each of the situations is defined by a triple (X,D,R)(X,D,R), where XX is a finite set of nn elements, D=d+dD=d+d^{*} is a finite n×nn\times n matrix such that d2=0d^{2}=0 and where RR is the dimension function compatible with DD in the sense that the blocks of L=D2L=D^{2} have constant dimension. We can call this structure an abstract delta set because every delta set defines such a structure, but where instead of face maps, we just go directly to the exterior derivative dd. The advantage of looking at the Dirac setting is that DD can be much more general than coming from face maps. It could be a deformed Dirac matrix for example obtained by isospectral deformation D=[D+D,D]D^{\prime}=[D^{+}-D^{-},D] [5, 4], which keeps the spectrum of DD invariant but produces D=d+d+BD=d+d^{*}+B leading to new exterior derivatives d(t)d(t), a deformation which is invisible to the Hodge Laplacian as D2(t)=LD^{2}(t)=L is not affected. Since d(t)d(t) gets smaller, this produces an expansion of space. In general, also in the continuum, there is an inflationary start of expansion.

2. A small example

2.1.

Lets illustrate the quadratic fusion inequality in the case K2K_{2}:

2.2.

The linear simplicial cohomology is given by ([{1}{2}{1,2}],[001001110],R=[001])(\left[\begin{array}[]{c}\{1\}\\ \{2\}\\ \{1,2\}\end{array}\right],\left[\begin{array}[]{ccc}0&0&-1\\ 0&0&1\\ -1&1&0\\ \end{array}\right],R=\left[\begin{array}[]{c}0\\ 0\\ 1\end{array}\right]). Take the open-closed pair U={{1,2}}U=\{\{1,2\}\} and K={{1},{2}}K=\{\{1\},\{2\}\} leading to the abstract delta set structures ([{1}{2}],[0000],[00])(\left[\begin{array}[]{c}\{1\}\\ \{2\}\end{array}\right],\left[\begin{array}[]{cc}0&0\\ 0&0\\ \end{array}\right],\left[\begin{array}[]{c}0\\ 0\end{array}\right]) ([{1,2}],[0],[1])([\{1,2\}],\left[\begin{array}[]{c}0\\ \end{array}\right],[1]). The Betti vectors are b(G)=(1,0),b(U)=(0,1),b(K)=(2,0)b(G)=(1,0),b(U)=(0,1),b(K)=(2,0) and the f-vectors are f(G)=(2,1),f(U)=(0,1),f(K)=(2,0)f(G)=(2,1),f(U)=(0,1),f(K)=(2,0). The fusion inequality b(G)<b(K)+b(U)b(G)<b(K)+b(U) is here strict. Merging UU and KK fuses a harmonic 0 form in KK with the 11-form in UU. Betti vectors have been considered since Betti and Poincaré. Finite topological spaces were first looked at by Alexandroff [1]. For cohomology of open sets in finite frame works, see [10].

2.3.

If we look at quadratic cohomology for GG, where we have the abstract delta set (X,D,R)=(X,D,R)=

([{2}{2}{1}{1}{1,2}{2}{1,2}{1}{2}{1,2}{1}{1,2}{1,2}{1,2}],[0010100000101010000010100001100000101000010011110],[0011112]).(\left[\begin{array}[]{cc}\{2\}&\{2\}\\ \{1\}&\{1\}\\ \{1,2\}&\{2\}\\ \{1,2\}&\{1\}\\ \{2\}&\{1,2\}\\ \{1\}&\{1,2\}\\ \{1,2\}&\{1,2\}\\ \end{array}\right],\left[\begin{array}[]{ccccccc}0&0&-1&0&1&0&0\\ 0&0&0&1&0&-1&0\\ -1&0&0&0&0&0&-1\\ 0&1&0&0&0&0&1\\ 1&0&0&0&0&0&-1\\ 0&-1&0&0&0&0&1\\ 0&0&-1&1&-1&1&0\\ \end{array}\right],\left[\begin{array}[]{c}0\\ 0\\ 1\\ 1\\ 1\\ 1\\ 2\end{array}\right])\;.

The Hodge Laplacian L=D2=L0L1L2L=D^{2}=L_{0}\oplus L_{1}\oplus L_{2} has the Hodge blocks:

L0=[2002],L1=[2101121001211012],L2=[4]L_{0}=\left[\begin{array}[]{cc}2&0\\ 0&2\\ \end{array}\right],L_{1}=\left[\begin{array}[]{cccc}2&-1&0&-1\\ -1&2&-1&0\\ 0&-1&2&-1\\ -1&0&-1&2\\ \end{array}\right],L_{2}=\left[\begin{array}[]{c}4\\ \end{array}\right]

with Betti vector b(G)=(0,1,0)b(G)=(0,1,0) and f-vector f(G)=(2,4,1)f(G)=(2,4,1) and Wu characteristic w(G)=f0f1+f2=24+1=b0b1+b2=01+0=1w(G)=f_{0}-f_{1}+f_{2}=2-4+1=b_{0}-b_{1}+b_{2}=0-1+0=-1. The eigenvalues of L1L_{1} are (0,2,2,4)(0,2,2,4), the null-space is spanned by [1,1,1,1][1,1,1,1]. By accident L1L_{1} happens to be a Kirchhoff matrix of C4C_{4}. If K2K_{2} is seen as a 1-manifold with boundary δG\delta G 222We usually assume that manifolds with boundary have an interior. The Barycentric refinement of a complete graph Kd+1K_{d+1} would be a dd-manifold with boundary. we have w(G)=χ(G)χ(δ(G))=12=1w(G)=\chi(G)-\chi(\delta(G))=1-2=-1, illustrating that in general, for manifolds MM with boundary δM\delta M, the Wu characteristic is χ(M)χ(δM)\chi(M)-\chi(\delta M).

2.4.

Now to U={{1,2}}U=\{\{1,2\}\}, where we have the abstract delta set structure (X,D,R)=(X,D,R)=

([{1,2}{1,2}],[0],[2]).(\left[\begin{array}[]{cc}\{1,2\}&\{1,2\}\\ \end{array}\right],\left[\begin{array}[]{c}0\\ \end{array}\right],\left[\begin{array}[]{c}2\end{array}\right])\;.

with b(U)=(0,0,1)b(U)=(0,0,1) and f(U)=(0,0,1)f(U)=(0,0,1) and w(U)=1w(U)=1.
For K={{1},{2}}K=\{\{1\},\{2\}\} we have the abstract delta set structure (X,D,R)=(X,D,R)=

([{2}{2}{1}{1}],[0000],[00])(\left[\begin{array}[]{cc}\{2\}&\{2\}\\ \{1\}&\{1\}\\ \end{array}\right],\left[\begin{array}[]{cc}0&0\\ 0&0\\ \end{array}\right],\left[\begin{array}[]{c}0\\ 0\end{array}\right])\;

with b(K)=(2,0,0)b(K)=(2,0,0) and f(U)=(2,0,0)f(U)=(2,0,0) and w(K)=2w(K)=2. Obviously the intrinsic cohomologies of UU and KK are not yet giving a complete picture. The simplices in UU and KK can interact as we see next.

2.5.

Now, we turn to the interactions of KK with UU

([{2}{1,2}{1}{1,2}],[0000],[11]).(\left[\begin{array}[]{cc}\{2\}&\{1,2\}\\ \{1\}&\{1,2\}\\ \end{array}\right],\left[\begin{array}[]{cc}0&0\\ 0&0\\ \end{array}\right],\left[\begin{array}[]{c}1\\ 1\end{array}\right])\;.
([{1,2}{2}{1,2}{1}],[0000],[11])(\left[\begin{array}[]{cc}\{1,2\}&\{2\}\\ \{1,2\}&\{1\}\\ \end{array}\right],\left[\begin{array}[]{cc}0&0\\ 0&0\\ \end{array}\right],\left[\begin{array}[]{c}1\\ 1\end{array}\right])\;

with b(K,U)=b(U,K)=(0,2,0)b(K,U)=b(U,K)=(0,2,0) and f(U)=(0,2,0)f(U)=(0,2,0) and w(U,K)=2w(U,K)=-2. There is no pair (x,y)U2(x,y)\in U^{2} such that xyKx\cap y\in K so that b(U,U)=(0,0,0)b(U,U)=(0,0,0).

2.6.

To summarize, we have

CaseBettifvectorCharacteristicU(0,0,1)(0,0,1)1K(2,0,0)(2,0,0)2(U,K)(0,2,0)(0,2,0)2(K,U)(0,2,0)(0,2,0)2(U,U)(0,0,0)(0,0,0)0G(0,1,0)(2,4,1)1.\begin{array}[]{|c|c|c|c|}\hline\cr Case&Betti&f-vector&Characteristic\\ \hline\cr U&(0,0,1)&(0,0,1)&1\\ K&(2,0,0)&(2,0,0)&2\\ (U,K)&(0,2,0)&(0,2,0)&-2\\ (K,U)&(0,2,0)&(0,2,0)&-2\\ (U,U)&(0,0,0)&(0,0,0)&0\\ G&(0,1,0)&(2,4,1)&-1\\ \hline\cr\end{array}\;.

The quadratic fusion inequality b(U)+b(K)+b(U,K)+b(K,U)+b(U,U)=(2,3,1)>b(G)=(0,1,0)b(U)+b(K)+b(U,K)+b(K,U)+b(U,U)=(2,3,1)>b(G)=(0,1,0) is here strict. The fusion has two 0-form-11-form mergers and one 11-form-22-form merger. The difference in the fusion inequality is is (1,1,0)+(1,1,0)+(0,1,1)=(2,3,1)(1,1,0)+(1,1,0)+(0,1,1)=(2,3,1).

2.7.

We see already in this small example, how the closed “laboratory” KK and the “observer space” UU are no more strictly separated, even so they partition the “world” GG. The “tunneling” between KK and UU is described using algebraic topology, expressed by cohomology groups. Unlike for simplicial cohomology which features homotopy invariance, there is only topological invariance. Already the Wu characteristic of contractible balls depends on the dimension. [For a dd-ball, the Wu characteristic w(M)w(M) is (1)d(-1)^{d} where dd is the dimension and illustrates that w(M)=χ(M)χ(δM)w(M)=\chi(M)-\chi(\delta M) in general for discrete manifolds MM with boundary δM\delta M and that for a dd-ball, the boundary is a d1d-1 sphere with Euler characteristic χ(δM)=1(1)d\chi(\delta M)=1-(-1)^{d}.] If we take a dd-ball in a dd-dimensional simplicial complex and replace the interior to get an other d-ball without changing the boundary, then the cohomology does not change because we can for any positive kk add add gauge fields (k-forms that are coboundaries) dgdg to render a cocycle zero in the interior (without changing the equivalence class) and use the heat flow to get back a harmonic form after doing the surgery in the interior. This implements the chain homotopy when doing a local homeomorphic deformation: move the field away from the “surgergy place”, do the surgery, then use the heat flow to “heal the wound” and get back harmonic forms.

2.8.

We have just given the argument for the following result:

Theorem 4.

All quadratic cohomology groups b(X),b(X,Y)b(X),b(X,Y) are topological invariants.

2.9.

For G=K3,K={{1}}G=K_{3},K=\{\{1\}\}, we can look at the complex G(U,K)=[{1}{1,3}{1}{1,2}{1}{1,2,3}]G(U,K)=\left[\begin{array}[]{cc}\{1\}&\{1,3\}\\ \{1\}&\{1,2\}\\ \{1\}&\{1,2,3\}\\ \end{array}\right] and D(K,U)=[001001110]D(K,U)=\left[\begin{array}[]{ccc}0&0&-1\\ 0&0&1\\ -1&1&0\\ \end{array}\right] with kernel spanned by [1,1,0][1,1,0]. For its Barycentric refinement and still K={1}K=\{1\} (on the boundary), and where we look at functions on X=[{1}{1,7}{1}{1,5}{1}{1,4}{1}{1,5,7}{1}{1,4,7}]X=\left[\begin{array}[]{cc}\{1\}&\{1,7\}\\ \{1\}&\{1,5\}\\ \{1\}&\{1,4\}\\ \{1\}&\{1,5,7\}\\ \{1\}&\{1,4,7\}\\ \end{array}\right], we have D(K,U)=[0001100010000011100010100]D(K,U)=\left[\begin{array}[]{ccccc}0&0&0&-1&-1\\ 0&0&0&1&0\\ 0&0&0&0&1\\ -1&1&0&0&0\\ -1&0&1&0&0\\ \end{array}\right] with kernel spanned by [1,1,1,0,0][1,1,1,0,0].

3. Quadratic cohomology

3.1.

Simplicial cohomology for a finite abstract simplicial complex GG is part of the spectral theory of the Hodge Laplacian L=D2L=D^{2} with Dirac matrix D=d+dD=d+d^{*}, where dd is the exterior derivative. Note that all these matrices d,D,Ld,D,L are n×nn\times n matrices if GG has nn elements. The matrix LL is a block diagonal matrix L=k=0dLkL=\oplus_{k=0}^{d}L_{k}. The kernels of the blocks LkL_{k} of LL are the kk-harmonic forms or kk-cohomology vector spaces. In this finite setting, this is linear algebra [2]. The dimensions bkb_{k} are the Betti numbers, the components of the Betti vector of GG.

3.2.

If KK is a subcomplex of GG and UU is the open complement, then the separated system (K,U)(K,U) has a Laplacian LK,U=LKLUL_{K,U}=L_{K}\oplus L_{U} for which the energies λj(LK,U)\lambda_{j}(L_{K,U}) are less or equal than 2λj(LG)2\lambda_{j}(L_{G}) [13] implying that the separated system can not have more harmonic forms than GG. It can have more: if GG is a closed 2-ball for example and KK is the boundary 1-sphere then bG=(1,0,0),bK=(1,1,0)b_{G}=(1,0,0),b_{K}=(1,1,0) and bU=(0,0,1)b_{U}=(0,0,1). The closed part KK carries a trapped harmonic 1-form. It is fused with the 2-form present on UU, if K,UK,U get united to GG.

3.3.

A complex GG defines a delta set G=k=0dGkG=\bigcup_{k=0}^{d}G_{k}. The ff-vector f(G)=(f0(G),,fd(G))f(G)=(f_{0}(G),\dots,f_{d}(G)) has components fk(G)=|Gk|f_{k}(G)=|G_{k}|, the number of elements in GkG_{k}. The super trace of an n×nn\times n matrix LL 333We write the entries as L(x,y)L(x,y) is defined as str(L)=k=0d(1)kxGkL(x,x){\rm str}(L)=\sum_{k=0}^{d}(-1)^{k}\sum_{x\in G_{k}}L(x,x). Compare with the usual trace tr(L)=k=0dxGkL(x,x)=xGL(x,x){\rm tr}(L)=\sum_{k=0}^{d}\sum_{x\in G_{k}}L(x,x)=\sum_{x\in G}L(x,x). The Euler characteristic is χ(G)=xGw(x)\chi(G)=\sum_{x\in G}w(x). The Euler-Poincaré formula χ(G)=k(1)kfk=k(1)kbk\chi(G)=\sum_{k}(-1)^{k}f_{k}=\sum_{k}(-1)^{k}b_{k} follows directly from the McKean-Singer identity, stating that str(exp(itL))=χ(G){\rm str}(\exp(-itL))=\chi(G) for all tt which in turn follows from the fact that the Dirac matrix DD gives an isomorphism between even and odd non-harmonic forms. For t=0t=0, the super trace of the heat kernel is the combinatorial Euler characteristic, while for t=t=\infty, it is the cohomological Euler characteristic.

3.4.

Quadratic cohomology does not build on single simplices xGx\in G like simplicial cohomology but on pairs of intersecting simplices (x,y)G×G(x,y)\in G\times G. Define w(x)=(1)dim(x)w(x)=(-1)^{{\rm dim}(x)}. The quadratic analog of (linear) Euler characteristic χ(A)=xAw(x)\chi(A)=\sum_{x\in A}w(x) is the “Ising type” energy or Wu characteristic w(A)=x,y,xyAw(x)w(y)w(A)=\sum_{x,y,x\cap y\in A}w(x)w(y). It is an example of a multi-linear valuation. We also just call it quadratic characteristic, an example of higher characteristic. [9].

3.5.

The name “quadratic” is chosen because it is multi-linear and for m=2m=2 a quadratic valuation. Similarly as a quadratic form is a multi-linear map, linear in each argument, the quadratic characteristic w(A,B)=xA,yB,xyw(A)w(B)w(A,B)=\sum_{x\in A,y\in B,x\cap y\neq\emptyset}w(A)w(B) (or variants, where we ask the intersection to be in AA or BB) satisfies the valuation formula in each of the coordinates, like w(X,UV)=w(X,U)+w(X,V)W(X,UV)w(X,U\cup V)=w(X,U)+w(X,V)-W(X,U\cap V).

3.6.

Given an open-closed pair (U,K)(U,K), one can define quadratic cohomology on kk-forms. Forms are functions on Λ(X,Y)={(x,y)|xX,yY,xyX}\Lambda(X,Y)=\{(x,y)|x\in X,y\in Y,x\cap y\in X\} and Λ(X)={(x,y)|xX,yX,xyX}\Lambda(X)=\{(x,y)|x\in X,y\in X,x\cap y\in X\} and Λ(X,Y)={(x,y)|xX,yY,xy,xyK}\Lambda(X,Y)=\{(x,y)|x\in X,y\in Y,x\cap y\neq\emptyset,x\cap y\in K\}. The kk-forms are the forms on functions with dim(x)+dim(y)=k{\rm dim}(x)+{\rm dim}(y)=k.

3.7.

In the case of an open-closed pair, we have five different cohomologies U,KU,K, (U,K)(U,K), (K,U)(K,U), (U,U)(U,U). There is no case (K,K)(K,K) because the intersection of xK,yKx\in K,y\in K is in KK. The case (K,K)(K,K) is part of KK. The case (U,U)(U,U) looks at pairs such that the intersection in in KK while UU looks at pairs such that the intersection is in KK. We can have a disjoint union

Λ(G)=Λ(U)Λ(K)Λ(K,U)Λ(U,K)Λ(U,U).\Lambda(G)=\Lambda(U)\cup\Lambda(K)\cup\Lambda(K,U)\cup\Lambda(U,K)\cup\Lambda(U,U)\;.
444In the code we part we identify Λ(U,K)\Lambda(U,K) with Λ(K,U)\Lambda(K,U) so that we do not have to probe which of the entries is the closed set.

3.8.

The exterior derivative is inherited from the exterior derivative on products. It is df(x,y)=dxf(x,y)+w(x)dy(f,y)df(x,y)=d_{x}f(x,y)+w(x)d_{y}(f,y), where dx,dyd_{x},d_{y} are the usual simplicial exterior derivatives but with respect to the first or second coordinate. If we would look at this derivative on X×YX\times Y, the Hodge Laplacians are the tensor products of the Laplacians on XX and YY. Even if the set-theoretical Cartesian product X×YX\times Y is not a simplicial complex any more, we still have a cohomology. But now, we restrict this exterior derivative to pairs (x,y)(x,y) that intersect. We are not aware of such a construction in the continuum.

4. Spectral Monotonicity

4.1.

The proof of the quadratic fusion inequality Theorem (2) is analog to the linear case. The key is that in each case, the matrix LL is the square L=D2L=D^{2} of a matrix DD which has the property that a principal sub-matrix of DD has intertwined spectrum so that the left padded spectral functions of LL are monotone. This looks like a technical detail but it is important and at the heart of the entire story: the matrix LL does not have the property that taking away highest or lowest dimensional simplices produces principal sub-matrices which themselves come from a geometry. But the Dirac matrix DD does have the property. And since DD has symmetric spectrum with respect to the origin and D2=LD^{2}=L, we have also monotonocity for LL.

4.2.

Let us formulate the Cauchy interlace theorem a bit differently, than usual. The point is that if a principal submatrix BB of a self-adjoint matrix AA has the eigenvalues padded left when compared to the eigenvalues of AA then there is a direct comparison between all eigenvalues. This is very general and allows to talk about monotonicity rather than interlacing.

Lemma 1 (Left Padded Monotonicity).

Let AA be a symmetric n×nn\times n matrix and mm a principal m×mm\times m submatrix, denote by λ1λn\lambda_{1}\leq\cdots\leq\lambda_{n} the eigenvalues of AA and μnmμn\mu_{n-m}\leq\cdots\leq\mu_{n} the eigenvalues of BB. Then μkλk\mu_{k}\leq\lambda_{k} for all nmknn-m\leq k\leq n.

Proof.

This follows directly from the interlace theorem and induction with respect to mm. Both induction assumption as well as the induction steps involve the interlace theorem. ∎

Refer to caption
Figure 1. We see the sorted eigenvalues of a random real self-adjoint 300×300300\times 300 matrix A0A_{0}, then the eigenvalues of a 200×200200\times 200 principal submatrix A1A_{1} and then the eigenvalues of a 100×100100\times 100 principal submatrix A2A_{2} of A1A_{1}. The eigenvalues are padded left. The figure illustrates Lemma (1). The code which gave the output is listed below.
n=300; m=100; B=Table[20*Random[]-10,{n},{n}]; A0=Transpose[B].B;
A=A0; Do[A1=Transpose[Delete[Transpose[Delete[A,1]],1]]; A=A1,{m}];
A=A1; Do[A2=Transpose[Delete[Transpose[Delete[A,1]],1]]; A=A2,{m}];
T=Eigenvalues;S=Sort;{a,b,c}=PadLeft[{S[T[A0]],S[T[A1]],S[T[A2]]}];
ListPlot[{a,b,c},Joined->True,Filling->Bottom,PlotRange->All];

4.3.

We can now look at the map Kλ(K)K\to\lambda(K) giving for each sub-complex KK the spectral function ordered in an ascending way and padded left. The partial order on sub-simplicial complexes and the partial order on spectral functions are compatible:

Corollary 1.

The maps Xλ(X)X\to\lambda(X) and (X,Y)λ(X,Y)(X,Y)\to\lambda(X,Y) preserve the partial orders in the sense that if we remove maximal simplices from closed sets, or minimal simplicial from open sets, then the spectral functions can only get smaller.

4.4.

The same holds by iterating the process and take principal (nk)×(nk)(n-k)\times(n-k) sub-matrices. Now, if we look at a Dirac matrix of a closed set KK and take a maximal simplex xx away, then we get monotonicity. The same happens if we take a minimal simplex xx away from an open set. Note that if we look at pairs (x,y)(x,y) belonging to some pair like (K,U)(K,U) and we take a maximal element xx away, then several pairs (x,yi)(x,y_{i}) are removed from the complex on (K,U)(K,U).

Theorem 5 (Spectral monotonicity).

For all jnj\leq n we have λj(K)λj(G)\lambda_{j}(K)\leq\lambda_{j}(G),
λj(K)λj(G)\lambda_{j}(K)\leq\lambda_{j}(G),
λj(K,U)λj(G)\lambda_{j}(K,U)\leq\lambda_{j}(G),
λj(U,K)λj(G)\lambda_{j}(U,K)\leq\lambda_{j}(G),
λj(U,U)λj(G)\lambda_{j}(U,U)\leq\lambda_{j}(G),

Proof.

If we add a locally maximal simplex to a given complex, the spectrum changes monotonically by interlace. For any vector uu, u,Lu=u,D2u=Du,Du=Du2\langle u,Lu\rangle=\langle u,D^{2}u\rangle=\langle Du,Du\rangle=||Du||^{2} Define 𝒮k={Vn,dim(V)=k}\mathcal{S}_{k}=\{V\subset\mathbb{R}^{n},{\rm dim}(V)=k\}

λk(K)=minV𝒮kmax|u|=1,uVu,L(K)uminV𝒮kmax|u|=1,uVu,L(G)u=λk(G).\lambda_{k}(K)={\rm min}_{V\in\mathcal{S}_{k}}{\rm max}_{|u|=1,u\in V}\langle u,L(K)u\rangle\leq{\rm min}_{V\in\mathcal{S}_{k}}{\rm max}_{|u|=1,u\in V}\langle u,L(G)u\rangle=\lambda_{k}(G)\;.

As for the interlace theorem applied to DD as the Dirac matrix of KK is obtained from the Dirac matrix of LL by deleting the row and column belonging to the element xx which was added. The eigenvalues of the Dirac matrix DKD_{K} of KK are now interlacing the eigenvalues of the Dirac matrix DGD_{G} of GG. ∎

4.5.

In the quadratic case, taking a way a largest dimensional simplex (facet) xx will affect in general various pairs of simplices (x,y)(x,y) or (y,x)(y,x). The effect is that the quadratic Dirac matrix of GxG\setminus x is still a principal sub-matrix. We still have spectral monotonicity.

4.6.

To conclude the proof of Theorem (2), write down the decoupled Laplacian L(U)L(K)L(K,U)L(U,K)L(U,U)L(U)\oplus L(K)\oplus L(K,U)\oplus L(U,K)\oplus L(U,U) which is block diagonal and is a n×nn\times n matrix, the same size than L(G)L(G). Lets call its eigenvalues μk\mu_{k}. From the spectral inequalities for each block, we know

0μk5λk.0\leq\mu_{k}\leq 5\lambda_{k}\;.

where λk\lambda_{k} are the eigenvalues of the quadratic Hodge Laplacian of GG. Therefore, there are at least as many 0 eigenvalues for the decoupled system than for GG, proving the inequality.

5. An example

5.1.

Here is an example with the Kite complex G=K1,2,1G=K_{1,2,1}, where we have a complex with 2 triangles. We will see what happens if we take one of the triangles away. We look at the case (U,U)(U,U). The Dirac matrix D(U,U)D(U,U) is a 14×1414\times 14 matrix.

[0000100010000000000010010000000001000010000000000100010010000000000010001000000000100100000000000100010000000001100000000000100100000000000100100000000010000100000000010000110010100000000011010100].\left[\begin{array}[]{cccccccccccccc}0&0&0&0&-1&0&0&0&-1&0&0&0&0&0\\ 0&0&0&0&0&0&-1&0&0&-1&0&0&0&0\\ 0&0&0&0&0&-1&0&0&0&0&-1&0&0&0\\ 0&0&0&0&0&0&0&-1&0&0&0&-1&0&0\\ -1&0&0&0&0&0&0&0&0&0&0&0&1&0\\ 0&0&-1&0&0&0&0&0&0&0&0&0&1&0\\ 0&-1&0&0&0&0&0&0&0&0&0&0&0&1\\ 0&0&0&-1&0&0&0&0&0&0&0&0&0&1\\ -1&0&0&0&0&0&0&0&0&0&0&0&-1&0\\ 0&-1&0&0&0&0&0&0&0&0&0&0&0&-1\\ 0&0&-1&0&0&0&0&0&0&0&0&0&-1&0\\ 0&0&0&-1&0&0&0&0&0&0&0&0&0&-1\\ 0&0&0&0&1&1&0&0&-1&0&-1&0&0&0\\ 0&0&0&0&0&0&1&1&0&-1&0&-1&0&0\\ \end{array}\right]\;.

The Eigenvalues of the Laplacian L(U,U)=D(U,U)2L(U,U)=D(U,U)^{2} are {4,4,4,4,2,2,2,2,2,2,2,2,0,0}\{4,4,4,4,2,2,2,2,2,2,2,2,0,0\}.

5.2.

Now lets take away the simplicies which do not involve the triangle (1,3,4)(1,3,4). We have to select the rows and columns in {1,2,3,4,7,8,9,11}\{1,2,3,4,7,8,9,11\}. The Dirac matrix is

[0000001000001000000000010000010001000000000100001000000000100000].\left[\begin{array}[]{cccccccc}0&0&0&0&0&0&-1&0\\ 0&0&0&0&-1&0&0&0\\ 0&0&0&0&0&0&0&-1\\ 0&0&0&0&0&-1&0&0\\ 0&-1&0&0&0&0&0&0\\ 0&0&0&-1&0&0&0&0\\ -1&0&0&0&0&0&0&0\\ 0&0&-1&0&0&0&0&0\\ \end{array}\right]\;.

The eigenvalues of LL are now {1,1,1,1,1,1,1,1}\{1,1,1,1,1,1,1,1\}.

6. Code

Generate[A_]:=If[A=={},{},Sort[Delete[Union[Sort[Flatten[Map[Subsets,A],1]]],1]]];
L=Length; Whitney[s_]:=Generate[FindClique[s,Infinity,All]]; L2[x_]:=L[x[[1]]]+L[x[[2]]];
(* Linear Cohomology *)
sig[x_]:=Signature[x]; nu[A_]:=If[A=={},0,L[A]-MatrixRank[A]];
F[G_]:=Module[{l=Map[L,G]},If[G=={},{},Table[Sum[If[l[[j]]==k,1,0],{j,L[l]}],{k,Max[l]}]]];
sig[x_,y_]:=If[SubsetQ[x,y]&&(L[x]==L[y]+1),sig[Prepend[y,Complement[x,y][[1]]]]*sig[x],0];
Dirac[G_]:=Module[{f=F[G],b,d,n=L[G]},b=Prepend[Table[Sum[f[[l]],{l,k}],{k,L[f]}],0];
d=Table[sig[G[[i]],G[[j]]],{i,n},{j,n}]; {d+Transpose[d],b}];
Hodge[G_]:=Module[{Q,b,H},{Q,b}=Dirac[G];H=Q.Q;Table[Table[H[[b[[k]]+i,b[[k]]+j]],
{i,b[[k+1]]-b[[k]]},{j,b[[k+1]]-b[[k]]}],{k,L[b]-1}]];
Betti[s_]:=Module[{G},If[GraphQ[s],G=Whitney[s],G=s];Map[nu,Hodge[G]]];
Fvector[A_]:=Delete[BinCounts[Map[Length,A]],1];
Euler[A_]:=Sum[(-1)^(Length[A[[k]]]-1),{k,Length[A]}];
(* Quadratic Cohomology *)
F2[G_]:=Module[{},If[G=={},{},Table[Sum[If[L2[G[[j]]]==k,1,0],{j,L[G]}],{k,Max[Map[L2,G]]}]]];
ev[L_]:=Sort[Eigenvalues[1.0*L]];
WuComplex[A_,B_,opts___]:=Module[{Q={},x,y,u},
Do[x=A[[k]];y=B[[l]];u=Intersection[x,y];
If[((opts==”Open” && Not[x==y] && L[u]>0 && Not[MemberQ[A,u]]) ||
(Not[opts==”Open”] && MemberQ[A,u])),
Q=Append[Q,{x,y}]],{k,L[A]},{l,L[B]}];Sort[Q,L2[#1]<L2[#2] &]];
Dirac[G_,H_,opts___]:=Module[{n=L[G],Q,m=L[H],b,d1,d2,h,v,w,l,DD}, Q=WuComplex[G,H,opts];
n2=L[Q]; f2=F2[Q]; b=Prepend[Table[Sum[f2[[l]],{l,k}],{k,L[f2]}],0];
D1[{x_,y_}]:=Table[{Sort[Delete[x,k]],y},{k,L[x]}];
D2[{x_,y_}]:=Table[{x,Sort[Delete[y,k]]},{k,L[y]}];
d1=Table[0,{n2},{n2}]; Do[v=D1[Q[[m]]]; If[L[v]>0,Do[r=Position[Q,v[[k]]];
If[r!={},d1[[m,r[[1,1]]]]=(-1)^k],{k,L[v]}]],{m,n2}];
d2=Table[0,{n2},{n2}]; Do[v=D2[Q[[m]]]; If[L[v]>0, Do[r=Position[Q,v[[k]]];
If[r!={},d2[[m,r[[1,1]]]]=(-1)^(L[Q[[m,1]]]+k)],{k,L[v]}]],{m,n2}];
d=d1+d2; DD=d+Transpose[d]; {DD,b}];
Beltrami[G_,H_,opts___]:=Module[{Q,P,b},{Q,b}=Dirac[G,H,opts];P=Q.Q];
Hodge[G_,H_,opts___]:=Module[{Q,P,b},{Q,b}=Dirac[G,H,opts];P=Q.Q;
Table[Table[P[[b[[k]]+i,b[[k]]+j]], {i,b[[k+1]]-b[[k]]},{j,b[[k+1]]-b[[k]]}],{k,2,L[b]-1}]];
Betti[G_,H_,opts___]:=Map[nu,Hodge[G,H,opts]];
Wu[A_,B_,opts___]:=Sum[x=A[[k]];y=B[[l]];u=Intersection[x,y];
If[(opts==”Open” && Not[x==y] && L[u]>0 && Not[MemberQ[A,u]]) ||
(Not[opts==”Open”] && MemberQ[A,u]),
(-1)^L2[{x,y}],0],{k,L[A]},{l,L[B]}];
Fvector[A_,B_,opts___]:=Module[{a=F2[WuComplex[A,B,opts]]},Table[a[[k]],{k,2,L[a]}]];
s = CompleteGraph[{1,2,1}]; G = Whitney[s]; K = Generate[{{1,4}}]; U=Complement[G,K];
Print[”LinearCohomology”];
{bU,bK,bG}=PadRight[{Betti[U],Betti[K],Betti[G]}];
{fU,fK,fG}=PadRight[{Fvector[U],Fvector[K],Fvector[G]}];
Print[ Grid[{
{”Case”, ”Betti”,”F-vector”,”Euler”}, {”U”, bU,fU, Euler[U]},
{”K”, bK,fK, Euler[K]}, {”G”, bG,fG, Euler[G]},
{”Compare”,bU+bK-bG,fU+fK-fG, Euler[U]+Euler[K]-Euler[G]}}]];
Print[”QuadraticCohomology”];
{bU,bK,bKU,bUK,bUU,bG}=PadRight[{Betti[U,U,”Closed”],Betti[K,K,”Closed”],
Betti[K,U,”Closed”],Betti[U,K,”Closed”], Betti[U,U,”Open”],Betti[G,G,”Closed”]}];
{fU,fK,fKU,fUK,fUU,fG}=PadRight[{Fvector[U,U,”Closed”],Fvector[K,K,”Closed”],
Fvector[K,U,”Closed”],Fvector[U,K,”Closed”], Fvector[U,U,”Open”], Fvector[G,G,”Closed”]}];
Print[ Grid[{ {”Case”,”Betti”,”F-vector”,”Wu”},{”U”,bU,fU,Wu[U,U,”Closed”]},
{”K”,bK,fK,Wu[K,K,”Closed”]},{”UK”,bKU,fKU,Wu[K,U,”Closed”]},{”KU”,bKU,fKU,Wu[K,U,”Closed”]},
{”UU”,bUU,fUU,Wu[U,U,”Open”]},{”G”, bG, fG, Wu[G,G,”Closed”]},
{”Compare”,bU+bK+bKU+bKU+bUU-bG,fU+fK+fKU+fKU+fUU-fG,
Wu[U,U,”Closed”]+Wu[K,K,”Closed”]+2Wu[K,U,”Closed”]+Wu[U,U,”Open”]-Wu[G,G,”Closed”]}}]];

6.1.

Here is the output of the above lines for simplicial cohomology

CaseBettiF-vectorEulerU{0,0,0}{2,4,2}0K{1,0,0}{2,1,0}1G{1,0,0}{4,5,2}1Compare{0,0,0}{0,0,0}0.\begin{array}[]{|c|c|c|c|}\hline\cr\text{Case}&\text{Betti}&\text{F-vector}&\text{Euler}\\ \hline\cr\text{U}&\{0,0,0\}&\{2,4,2\}&0\\ \text{K}&\{1,0,0\}&\{2,1,0\}&1\\ \text{G}&\{1,0,0\}&\{4,5,2\}&1\\ \hline\cr\text{Compare}&\{0,0,0\}&\{0,0,0\}&0\\ \hline\cr\end{array}\;.

And here the output table for the quadratic cohomology part:

CaseBettiF-vectorWuU{0,0,0,0,0}{2,8,12,8,2}0K{0,1,0,0,0}{2,4,1,0,0}1UK{0,0,2,0,0}{0,4,8,2,0}2KU{0,0,2,0,0}{0,4,8,2,0}2UU{0,0,0,2,0}{0,0,4,8,2}2G{0,0,1,0,0}{4,20,33,20,4}1Compare{0,1,3,2,0}{0,0,0,0,0}0.\begin{array}[]{|c|c|c|c|}\hline\cr\text{Case}&\text{Betti}&\text{F-vector}&\text{Wu}\\ \hline\cr\text{U}&\{0,0,0,0,0\}&\{2,8,12,8,2\}&0\\ \text{K}&\{0,1,0,0,0\}&\{2,4,1,0,0\}&-1\\ \text{UK}&\{0,0,2,0,0\}&\{0,4,8,2,0\}&2\\ \text{KU}&\{0,0,2,0,0\}&\{0,4,8,2,0\}&2\\ \text{UU}&\{0,0,0,2,0\}&\{0,0,4,8,2\}&-2\\ \text{G}&\{0,0,1,0,0\}&\{4,20,33,20,4\}&1\\ \hline\cr\text{Compare}&\{0,1,3,2,0\}&\{0,0,0,0,0\}&0\\ \hline\cr\end{array}\;.
Refer to caption
Figure 2. The Dirac matrix DD and the Hodge Laplacian L=D2L=D^{2} in the linear case for the kite graph GG. The splittings are given by the f-vector f(G)=(4,5,2)f(G)=(4,5,2). There are 4 points, 5 edges and 2 triangles in GG.
Refer to caption
Figure 3. The Dirac matrix DD and the Hodge Laplacian L=D2L=D^{2} in the quadratic case for the kite graph. The splittings are given by the f-vector f(G)=(4,20,33,20,4)f(G)=(4,20,33,20,4). The space of 11-forms (intersecting points) is 44-dimensional, the space of 22-forms (intersection of a point with an edges) is 2020-dimensional, the space of 33-forms (intersection of two edges or a triangle-point has dimension 3333), the space of 44 forms (intersection of an edge and triangle) is 2020-dimensional, the space of 55 forms (intersection of two triangles) is 44-dimensional.

References

  • [1] P. Alexandroff. Diskrete Räume. Mat. Sb. 2, 2, 1937.
  • [2] B. Eckmann. Harmonische Funktionen und Randwertaufgaben in einem Komplex. Comment. Math. Helv., 17(1):240–255, 1944.
  • [3] B. Grünbaum. Polytopes, graphs, and complexes. Bull. Amer. Math. Soc., 76:1131–1201, 1970.
  • [4] O. Knill. An integrable evolution equation in geometry.
    http://arxiv.org/abs/1306.0060, 2013.
  • [5] O. Knill. Isospectral deformations of the Dirac operator.
    http://arxiv.org/abs/1306.5597, 2013.
  • [6] O. Knill. Gauss-Bonnet for multi-linear valuations.
    http://arxiv.org/abs/1601.04533, 2016.
  • [7] O. Knill. The cohomology for Wu characteristics.
    http://arxiv.org/abs/1803.06788, 2017.
  • [8] O. Knill. The cohomology for Wu characteristics.
    https://arxiv.org/abs/1803.1803.067884, 2018.
  • [9] O. Knill. Characteristic topological invariants.
    https://arxiv.org/abs/2302.02510, 2023.
  • [10] O. Knill. Cohomology of Open sets. https://arxiv.org/abs/2305.12613, 2023.
  • [11] O. Knill. Discrete algebraic sets in discrete manifolds. https://arxiv.org/abs/2312.14671, 2023.
  • [12] O. Knill. Finite topologies for finite geometries. https://arxiv.org/abs/2301.03156, 2023.
  • [13] O. Knill. Spectral monotonicity of the Hodge Laplacian. https://arxiv.org/abs/2304.00901, 2023.
  • [14] Wu W-T. Topologie combinatoire et invariants combinatoires. Colloq. Math., 7:1–8, 1959.