Fusion inequality for quadratic cohomology
Abstract.
Classical simplicial cohomology on a simplicial complex deals with functions on simplices . Quadratic cohomology [6, 7] deals with functions on pairs of simplices that intersect. If is a closed-open pair in , we prove here a quadratic version of the linear fusion inequality [10]. Additional to the quadratic cohomology of there are five additional interaction cohomology groups. Their Betti numbers are computed from functions on pairs of simplices that intersect. Define the Betti vector computed from pairs with a and with pairs in with . We prove the fusion inequality for cohomology groups linking all five possible interaction cases. Counting shows for the f-vectors. Super counting gives Euler-Poincaré and for . 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 are bounded above by the spectrum of the quadratic Hodge Laplacian of .
Key words and phrases:
Quadratic cohomology1. In a nutshell
1.1.
We prove here that if is a sub-complex of a finite abstract simplicial complex and is the open complement [1, 12], there are besides the quadratic cohomology of five quadratic cohomology groups belonging to the five quadratic Hodge Laplacians with . They all satisfy spectral inequalities:
Theorem 1 (Spectral inequality).
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 . This result parallels the linear simplicial cohomology case [13], where and can not yet interact and is one of the Hodge Laplacians for simplicial cohomology.
1.3.
In the linear case, the Betti vectors satisfied the fusion inequality [10]. This linear fusion inequality had followed from the spectral inequality and the fact that cohomology classes are null-spaces of matrices. Counting gave for the -vectors and the Euler-Poincaré formula was for , seen directly by heat deformation using the McKean-Singer symmetry, rephrasing that is an isomorphism between even and odd parts of image of the Laplacian , implying for so that .
1.4.
In the quadratic cohomology case, where we look at functions on pairs of intersecting simplices, there is the cohomology of leading to and five interaction cohomologies. They each lead to Betti vectors. We call them ,, and . Besides pointing out that we have these new cohomologies, we give here a relation between them and the cohomology of . We call it the quadratic fusion inequality. We could use the heavier notation with dealing with functions on pairs with , but we prefer to stick to the simpler notation: one reason is that both deal with internal cohomology of and and do not involve simplices of the other set, while involve both sets and . So, while the quadratic Betti vectors are intrinsic and only depend on one of the sets or , the others are not. We prove:
Theorem 2 (Quadratic fusion inequality).
.
1.5.
The quadratic Betti vector belongs to , the vector belongs to all with , and belongs to all with , and belongs to all with . Note that if so that is accounted for in already. With the heavier notation, from the eight cases with only can occur.
1.6.
The quadratic characteristics are defined as (Wu characteristic) and and , we have which follows from the fact that -vectors for quadratic cohomology satisfy . Any of the cohomologies for satisfy the Euler-Poincaré formula, following heat deformation with or using the McKean-Singer symmetry.
1.7.
Why is this interesting? If is finite abstract simplicial complex that is a finite -manifold, and is an arbitrary function from to , then the discrete Sard theorem [11] assures that the open set is either empty or a -manifold in the sense that the graph with vertex set and edge set for which or is a discrete -manifold. Since is open, the complement is closed and so a sub-simplicial complex. All these cohomologies are topological invariants.
1.8.
For example, if is a discrete -sphere and is arbitrary, then is either empty, a knot or a link, a finite union of closed disjoint (possibly interlinked) -manifolds in the 3-sphere . The simplicial cohomology of and the quadratic cohomology of are not interesting: they are just circles: , where is the number of connected components of . The simplicial cohomology of 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 and the complement are completely unexplored. The inequality shows however that in general, more cohomology classes are created when splitting up into . Unlike in the linear case, we have now the possibility of particles (harmonic forms) that are functions of with with and also of function on with with .
1.9.
The quadratic case we look at here would generalizes in a straightforward way to higher characteristics. One starts with , the linear characteristic which is Euler characteristic. The second 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 -tuples of points that do simultaneously intersect. There are then much more -point interaction cohomologies and is again bounded above by all possible cases of with and the intersection in . The cases has cases with which is part of and a new part where . As in the case , the cohomology of is part of . We have to consider the Betti vector for functions on and referring to functions on , where in the notation used before. If all are in , then the intersection can be either in or , but if all are in , then the intersection must be in , not warranting to distinguish and . Despite the obvious duality , there is an asymmetry in that allows to be in or while implies : technically, a closed set is a -system while an open set is not unless it is or .
Theorem 3.
.
There would again be heavier notation dealing with with and have taking into account that some of the cases like are empty because only is interesting if all . Again, like in the case , we have two cases which are intrinsic while the other cases involve simplices from both and .
1.10.
We study here higher order chain complexes in finite geometries. Each of the situations is defined by a triple , where is a finite set of elements, is a finite matrix such that and where is the dimension function compatible with in the sense that the blocks of 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 . The advantage of looking at the Dirac setting is that can be much more general than coming from face maps. It could be a deformed Dirac matrix for example obtained by isospectral deformation [5, 4], which keeps the spectrum of invariant but produces leading to new exterior derivatives , a deformation which is invisible to the Hodge Laplacian as is not affected. Since 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 :
2.2.
The linear simplicial cohomology is given by . Take the open-closed pair and leading to the abstract delta set structures . The Betti vectors are and the f-vectors are . The fusion inequality is here strict. Merging and fuses a harmonic form in with the -form in . 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 , where we have the abstract delta set
The Hodge Laplacian has the Hodge blocks:
with Betti vector and f-vector and Wu characteristic . The eigenvalues of are , the null-space is spanned by . By accident happens to be a Kirchhoff matrix of . If is seen as a 1-manifold with boundary 222We usually assume that manifolds with boundary have an interior. The Barycentric refinement of a complete graph would be a -manifold with boundary. we have , illustrating that in general, for manifolds with boundary , the Wu characteristic is .
2.4.
Now to , where we have the abstract delta set structure
with and and .
For we have the abstract delta set structure
with and and . Obviously the intrinsic cohomologies of and are not yet giving a complete picture. The simplices in and can interact as we see next.
2.5.
Now, we turn to the interactions of with
with and and . There is no pair such that so that .
2.6.
To summarize, we have
The quadratic fusion inequality is here strict. The fusion has two -form--form mergers and one -form--form merger. The difference in the fusion inequality is is .
2.7.
We see already in this small example, how the closed “laboratory” and the “observer space” are no more strictly separated, even so they partition the “world” . The “tunneling” between and 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 -ball, the Wu characteristic is where is the dimension and illustrates that in general for discrete manifolds with boundary and that for a -ball, the boundary is a sphere with Euler characteristic .] If we take a -ball in a -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 add add gauge fields (k-forms that are coboundaries) 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 are topological invariants.
2.9.
For , we can look at the complex and with kernel spanned by . For its Barycentric refinement and still (on the boundary), and where we look at functions on , we have with kernel spanned by .
3. Quadratic cohomology
3.1.
Simplicial cohomology for a finite abstract simplicial complex is part of the spectral theory of the Hodge Laplacian with Dirac matrix , where is the exterior derivative. Note that all these matrices are matrices if has elements. The matrix is a block diagonal matrix . The kernels of the blocks of are the -harmonic forms or -cohomology vector spaces. In this finite setting, this is linear algebra [2]. The dimensions are the Betti numbers, the components of the Betti vector of .
3.2.
If is a subcomplex of and is the open complement, then the separated system has a Laplacian for which the energies are less or equal than [13] implying that the separated system can not have more harmonic forms than . It can have more: if is a closed 2-ball for example and is the boundary 1-sphere then and . The closed part carries a trapped harmonic 1-form. It is fused with the 2-form present on , if get united to .
3.3.
A complex defines a delta set . The -vector has components , the number of elements in . The super trace of an matrix 333We write the entries as is defined as . Compare with the usual trace . The Euler characteristic is . The Euler-Poincaré formula follows directly from the McKean-Singer identity, stating that for all which in turn follows from the fact that the Dirac matrix gives an isomorphism between even and odd non-harmonic forms. For , the super trace of the heat kernel is the combinatorial Euler characteristic, while for , it is the cohomological Euler characteristic.
3.4.
Quadratic cohomology does not build on single simplices like simplicial cohomology but on pairs of intersecting simplices . Define . The quadratic analog of (linear) Euler characteristic is the “Ising type” energy or Wu characteristic . 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 a quadratic valuation. Similarly as a quadratic form is a multi-linear map, linear in each argument, the quadratic characteristic (or variants, where we ask the intersection to be in or ) satisfies the valuation formula in each of the coordinates, like .
3.6.
Given an open-closed pair , one can define quadratic cohomology on -forms. Forms are functions on and and . The -forms are the forms on functions with .
3.7.
In the case of an open-closed pair, we have five different cohomologies , , , . There is no case because the intersection of is in . The case is part of . The case looks at pairs such that the intersection in in while looks at pairs such that the intersection is in . We can have a disjoint union
3.8.
The exterior derivative is inherited from the exterior derivative on products. It is , where are the usual simplicial exterior derivatives but with respect to the first or second coordinate. If we would look at this derivative on , the Hodge Laplacians are the tensor products of the Laplacians on and . Even if the set-theoretical Cartesian product is not a simplicial complex any more, we still have a cohomology. But now, we restrict this exterior derivative to pairs 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 is the square of a matrix which has the property that a principal sub-matrix of has intertwined spectrum so that the left padded spectral functions of are monotone. This looks like a technical detail but it is important and at the heart of the entire story: the matrix 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 does have the property. And since has symmetric spectrum with respect to the origin and , we have also monotonocity for .
4.2.
Let us formulate the Cauchy interlace theorem a bit differently, than usual. The point is that if a principal submatrix of a self-adjoint matrix has the eigenvalues padded left when compared to the eigenvalues of 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 be a symmetric matrix and a principal submatrix, denote by the eigenvalues of and the eigenvalues of . Then for all .
Proof.
This follows directly from the interlace theorem and induction with respect to . Both induction assumption as well as the induction steps involve the interlace theorem. ∎

4.3.
We can now look at the map giving for each sub-complex 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 and 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 sub-matrices. Now, if we look at a Dirac matrix of a closed set and take a maximal simplex away, then we get monotonicity. The same happens if we take a minimal simplex away from an open set. Note that if we look at pairs belonging to some pair like and we take a maximal element away, then several pairs are removed from the complex on .
Theorem 5 (Spectral monotonicity).
For all we have
,
,
,
,
,
Proof.
If we add a locally maximal simplex to a given complex, the spectrum changes monotonically by interlace. For any vector , Define
As for the interlace theorem applied to as the Dirac matrix of is obtained from the Dirac matrix of by deleting the row and column belonging to the element which was added. The eigenvalues of the Dirac matrix of are now interlacing the eigenvalues of the Dirac matrix of . ∎
4.5.
In the quadratic case, taking a way a largest dimensional simplex (facet) will affect in general various pairs of simplices or . The effect is that the quadratic Dirac matrix of 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 which is block diagonal and is a matrix, the same size than . Lets call its eigenvalues . From the spectral inequalities for each block, we know
where are the eigenvalues of the quadratic Hodge Laplacian of . Therefore, there are at least as many 0 eigenvalues for the decoupled system than for , proving the inequality.
5. An example
5.1.
Here is an example with the Kite complex , 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 . The Dirac matrix is a matrix.
The Eigenvalues of the Laplacian are .
5.2.
Now lets take away the simplicies which do not involve the triangle . We have to select the rows and columns in . The Dirac matrix is
The eigenvalues of are now .
6. Code
6.1.
Here is the output of the above lines for simplicial cohomology
And here the output table for the quadratic cohomology part:


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.