On Simplicial Complexes with Extremal Total Betti number and Total Bigraded Betti Number
Abstract.
We determine which simplicial complexes have the maximum or minimum sum of Betti numbers and sum of bigraded Betti numbers with a given number of vertices in each dimension.
Key words and phrases:
Total Betti number, bigraded Betti number, Sperner family2020 Mathematics Subject Classification:
05E45, 13F551. Introduction
For a finite CW-complex and a field , let
where and are the singular and reduced singular homology group of with -coefficients, respectively. Moreover, define
We call and the total Betti number and reduced total Betti number of with -coefficients, respectively. The difference between and is just . In addition, we call
the reduced Euler characteristic of , which is independent on the coefficient .
Convention: The coefficient field will be omitted when there is no ambiguity in the context or the coefficients are not essential. In fact, all the results obtained in this paper are independent on the coefficients.
The total Betti number of a topological space plays an important role in many theories in mathematics. For example, in the Arnold conjecture in symplectic geometry, the total Betti number of a symplectic manifold serves as the lower bound of the number of fixed points of any Hamiltonian symplectomorphism of ; see Abbondandolo [1, Chapter 6] for more information. Another example is from the Smith theory of cyclic group actions where the Smith inequality says that for a prime and a -action on a finite CW-complex , the fixed point set must satisfy ; see Allday and Puppe [2, Chapter 1]. In addition, an important conjecture in equivariant topology and rational homotopy theory due to Halperin [12] states that if a -dimensional torus can act almost freely on a finite-dimensional topological space , then . A parallel conjecture due to Carlsson [7] states that for a prime , if a -torus can act freely on a finite CW-complex , then .
In this paper, we focus our study on the total Betti number and total bigraded Betti number (see Definition 1.4) of a simplicial complex with a fixed number of vertices. Let be a simplicial complex whose vertex set is
Each simplex of is considered as a subset of . Let and denote the star and the link of in , respectively. For any subset , let
In particular, when , and define
Question 1: For a positive integer , which simplicial complexes have the maximum (reduced) total Betti number among all the simplicial complexes with vertices? When , the answer is just the discrete points. When , the answer is either the discrete points or the complete graph on vertices. A complete answer to Question 1 is contained in the following theorem.
Theorem 1.1 (Björner and Kalai [4, Theorem 1.4]).
Suppose is a simplicial complex with at most vertices. Then . Moreover, the following conditions are equivalent:
-
(i)
,
-
(ii)
,
-
(iii)
is the -skeleton of an -simplex, where if is even and or if is odd.
Note that Theorem 1.1 not only gives the answer to Question 1, but also reveal the connection between and the Euler characteristic of . The proof of Theorem 1.1 in [4] uses a nontrivial operation called algebraic shifting of a simplicial complex and a well-known theorem of Sperner in [18]. But if we only want to answer Question 1, we can use some easier argument as shown in Section 2.
Moreover, we can ask a more subtle question than Question 1 as follows.
Question 2: For each , which -dimensional simplicial complexes with vertices have the maximum (reduced) total Betti number among all the -dimensional simplicial complexes with vertices?
For a pair of integers , , we introduce the following notations:
-
•
Let denote the set of all simplicial complexes with vertex set .
-
•
Let denote the set of all -dimensional simplicial complexes with vertex set .
Moreover, define (with respect to a fixed coefficient field )
We consider , , and as partial ordered sets with respect to the inclusion of simplicial complexes. Clearly, if is an (maximal or minimal) element of , then is also an (maximal or minimal) element of .
In addition, for any , we use to denote the -dimensional simplex with vertex set . So the boundary of is a simplicial sphere of dimension . Moreover, for any ,
-
•
let denote the -skeleton of ;
-
•
let denote the minimal -dimensional subcomplex of that contains , which is unique up to simplicial isomorphism. Indeed, we can think of as the union of with the -simplex in where .
We will prove the following theorem in Section 2 which answers Question 2.
Theorem 1.2 (see Theorem 2.12).
The sets are classified as follows:
-
(i)
If or , then ;
-
(ii)
If , then ;
-
(iii)
If ,
-
–
when is odd, ;
-
–
when is even, .
-
–
By the above Theorem, the set is independent on the coefficients .
Remark 1.3.
It is also meaningful to ask what kind of simplicial complexes with vertices have the minimum total Betti number. But the answer is just all the acyclic simplicial complexes which are numerous; see Kalai [16] for more discussion on this subject.
There is another family of integers associated to a simplicial complex called bigraded Betti numbers, which are derived from the Stanley-Reisner ring of . Recall the Stanley-Reisner ring of over (see Stanley [19]) is
where is the ideal in the polynomial ring generated by all the square-free monomials where is not a simplex of . Since is naturally a module over , by the standard construction in homological algebra, we obtain a canonical algebra from , where is considered as the trivial -module. Moreover, there is a bigraded -module structure on (see [19]):
where for each . The integers
are called the bigraded Betti numbers of with -coefficients. Note that unlike Betti numbers of , bigraded Betti numbers are not topological invariants, but only combinatorial invariants of in general.
Definition 1.4.
The total bigraded Betti number of with -coefficients is
By Hochster’s formula (see Hochster [14] or [19, Theorem 4.8]), the bigraded Betti numbers of can also be computed from the homology groups of the full subcomplexes of :
(1) |
So we can also express as
(2) |
This explains why we put “” in the notation .
Question 3: For each , which -dimensional simplicial complexes with vertices have the minimum total bigraded Betti number over a field among all the -dimensional simplicial complexes with vertices? We readily call such kind of simplicial complexes -minimal (over ).
It follows from Cao and Lü [6, Theorem 1.4] or Ustinovsky [20, Theorem 3.2] that there is a universal lower bound of (see (11)):
Definition 1.5 (Tight Simplicial Complex).
Let be a -dimensional simplicial complex with vertices. We call tight (over ) if .
We classify all the tight simplicial complexes in the following theorem.
Theorem 1.6 (see Theorem 3.10).
A finite simplicial complex is tight if and only if is of the form or for some positive integers and .
It is a convention to let and . Theorem 1.6 implies that the tightness of a simplicial complex is independent on the coefficient field .
From Theorem 1.6, we can easily deduce that if is tight, it is necessary that . In particular, the equality is achieved by when is even and by when is odd. Conversely, for any with , there always exists a tight simplicial complex . So if , the -minimal simplicial complexes in are exactly all the tight simplicial complexes.
But when , a -minimal simplicial complex in is never tight, and it seems to us that there is no very good way to describe -minimal simplicial complexes in general. One reason is that the full subcomplexes of a -minimal simplicial complex may not be -minimal. For example, by exhausting all the members of , we find that all the -minimal simplicial complexes in are and (see Figure 1) whose -value is . Note that none of the full subcomplexes of on four vertices are -minimal.

This example suggests that we may not be able to inductively construct all the -minimal simplicial complexes with vertices from the -minimal ones with vertices. In addition, it is not clear how to compute the minimal value of on when except exhausting all the members.
Moreover, we can ask the following question parallel to Question 1 for bigraded Betti numbers.
Question 4: For a positive integer , which simplicial complexes with vertices have the maximum total bigraded Betti numbers among all the simplicial complexes with vertices?
In Section 4, we prove the following theorem which answers Question 4. Let
(3) |
Theorem 1.7 (Theorem 4.2).
If is a simplicial complex with vertices, then
for any field , where the equality holds if and only if .
Remark 1.8.
Parallelly to Question 2, we can also ask for each , which simplicial complexes have the maximum total bigraded Betti number over a field among all the members of . But we do not know the complete answer of this question yet. The question can be turned into a very complicated combinatorial extremum problem.
The paper is organized as follows. In Section 2, we first write an easy argument using Mayer-Vietoris sequence to answer Question 1. Then we use the theory of algebraic shifting of simplicial complexes from [4] and some results on Sperner families from [17] and [11] to give a complete answer to Question 2. In Section 3, we classify all the tight simplicial complexes using some results from [20] and [21]. In Section 4, we give a complete answer to Question 4 via some combinatorial inequality proved in the appendix.
2. Simplicial complexes with the maximum total Betti number
In this section, we first give an alternative proof of Theorem 2.2 which answers Question 1. Our approach is different from [4].
Lemma 2.1.
For any two subcomplexes of , one always have
(4) |
Proof.
By the following Mayer-Vietoris sequence for :
we obtain:
Then sum it over all , we get the desired inequality. ∎
Note that the equality in (4) holds if and only if for all in the above Mayer-Vietoris sequence.
Theorem 2.2.
Let be a simplicial complex with vertices. Then
Moreover, if and only if is isomorphic to the -skeleton of an -dimensional simplex, where if is odd and or if is even.
Proof.
Suppose the vertex set of is and consider as a subcomplex of . In the rest of the proof, we assume that .
Note that each permutation of determines a simplicial isomorphism , still denoted by . Let denote the image of under . Then by Lemma 2.1,
(5) |
Since is homeomorphic to , if is an element of , then both and must also belong to since their vertex sets are also . So in particular, if is a maximal (or minimal) element of , we must have (or ) for every permutation of , which implies . Then is the skeleton . An easy calculation shows
So reaches the maximum when and only when
It remains to prove that there are no other elements in except the skeleta of described in the theorem. Let be an arbitrary element of .
When is odd, since both the maximal and the minimal elements of are , can only be .
When is even, the maximal element of is while the minimal element is . So can only be or . If , then implies . But if , the situation is a bit complicated.
In the following, suppose is a minimal element of . If has only one -simplex, then clearly is less than , which contradicts our assumption that . So must have at least two -simplices. We have the following two cases.
-
Case 1:
If all the -simplices of are disjoint, then could only have two -simplices. But it is easy to check that is less than , which again contradicts the assumption .
-
Case 2:
Suppose and are two -simplices in which share exactly vertices where . Without loss of generality, we can assume .
Let be an arbitrary permutation of that preserves . Then contains , and hence . But by the minimality of in , we must have . In particular, belongs to . This implies that any -simplex in that shares exactly vertices with must also belong to . This is because the permutation can send the face to any other face of with vertices while mapping the face of to any other simplex with vertices in the complement of in . So can exhaust all the -simplices in that share exactly vertices with . Moreover, we can similarly prove that any -simplex in that shares exactly vertices with must also belong to .
Next, we prove that there exists a -simplex in which shares a -face with . Indeed, we can just trade the vertices with and obtain another -simplex which satisfies . So must belong to and we have .
From the above discussion, we see that if is even and is a minimal element in , we can always find a pair of -simplices and in which share a -face. Moreover, by the argument in Case 2, any -simplex of that intersects or at a -face must belong to . So without loss of generality, assume .
Claim: Every -simplex of must belong to , i.e. .
If a -simplex of is disjoint from , then . We can construct a sequence of -simplices , as follows:
, .
Since is a -simplex for every , we can prove that must all belong to by iteratively using the argument in Case 2.
Similarly, if , we can trade the vertices in with the vertices in one at a time to turn to . This proves the claim.
By the above claim, the minimal element of is which is also maximal. So consists only of . Therefore, when is even, must either be or . The theorem is proved. ∎
For any finite simplicial complex of dimension and a field , one can associate two integer vectors and , where is the number of -simplices of and is the -th reduced Betti number (over ) of (. A remarkable result proved in [4] tells us that a collection of nonlinear relations along with the linear relation given by the Euler-Poincaré formula completely characterize the integer vectors which can arise as and for a simplicial complex. But from the description in [4] of the range of the -vectors and -vectors of all the simplicial complexes with vertices, it is still not clear what the answer of Question 2 should be. Indeed, when we fix the dimension of the simplicial complex , the upper bound of will depend on and become a little complicated (see Corollary 2.13). So to obtain the answer of Question 2, we also need to use the machinery of shifting of simplicial complexes and the theory of Sperner family just as [4] did.
Now, let us recall some basic definitions that are needed for our argument.
Definition 2.3 (Sperner Family).
Let be a finite set. A Sperner family of is a set of subsets of that satisfies for distinct members of . Given a subset , a Sperner family of over is a Sperner family of where every member of has nonempty intersection with .
The following is a fundamental result of Sperner on the size of a Sperner family (see Sperner [18] or Anderson [3, Theorem 1.2.2]).
Theorem 2.4 (Sperner’s Theorem).
Let be a Sperner family of subsets of a finite set where . Then . If is even, the only Sperner family consisting of subsets of is made up of all the -subsets of . If is odd, a Sperner family of size consists of either all the -subsets or all the -subsets of .
A -subset of means a subset of order . Another very useful construction in the study of combinatorics of simplicial complexes is the shifting operation. A shifting operation is a map which assigns to every simplicial complex a shifted simplicial complex with the same -vector.
Definition 2.5 (Shifted Complex).
A simplicial complex with vertex set is called shifted if for every simplex where , any with and is also a simplex of .
A well-known (combinatorial) shifting operation, introduced by Erdös, Ko and Rado [10], has been of great use in extremal set theory. Later, another shifting operation was introduced by Kalai in [15] which preserves both the -vector and the -vector of a simplicial complex.
Theorem 2.6 (see [15] and [4]).
Given a simplicial complex on vertices and a field , there exists a canonically defined shifted simplicial complex on such that and for all .
Shifted complexes belong to a slightly larger class of simplicial complexes called near-cones (see [4]).
Definition 2.7 (Near-Cone).
A simplicial complex with vertex set is called a near-cone if for every simplex and , if and , then . For a near-cone , define
(6) |
A very nice property of a near-cone is that its total Betti number can be easily computed by counting the number of elements in .
Lemma 2.8 ([4, Lemma 4.2, Theorem 4.3]).
If is a near-cone on , then
-
(i)
is a Sperner family of ,
-
(ii)
every simplex is maximal in ,
-
(iii)
.
The following proposition tells one what kind of Sperner families on can be realized by of a near-cone .
Proposition 2.9.
Let be a Sperner family of . Then the following statements are equivalent:
-
(i)
there exists a -dimensional near-cone with vertex set contained in , such that
-
(ii)
there exists a subset such that no subset of belongs to and the order of each member of is no greater than .
Proof.
(i)(ii). First of all, since is a -dimensional near-cone, it has at least one -simplex, say . Then by the definition of in (6), no subset of belongs to since is a face of . Moreover, the order of each member of is no greater than since is -dimensional. (ii)(i). From the Sperner family , we define
(7) |
where is the simplicial complex generated by , that is, the minimal simplicial complex taking all the members of as its maximal faces. Then is clearly a -dimensional simplicial complex whose vertex set is a subset of . Moreover, for any and , we have , so is a near-cone. Finally, it is obvious that . ∎
Definition 2.10.
Let be a finite set with order . For a nonempty subset of , let denote the set of all subsets of that have nonempty intersection with . For any , let denote the collection of sets in of size .
The following lemma combines some results from Lih [17] and Griggs [11], where the is the ceiling function.
Lemma 2.11 (see [17, Theorem 2] and [11, Theorem 7.1]).
Let be a finite set of order . The maximal possible cardinality of a Sperner family of over a subset with is . Moreover, if and only if is one of the following cases:
-
(a)
;
-
(b)
, for odd and ;
-
(c)
, for even and .
In particular, if , then every member in has the same order.
Now, we are ready to prove the following theorem which answers Question 2.
Theorem 2.12.
The sets are classified as follows:
-
(i)
If or , then ;
-
(ii)
If , then ;
-
(iii)
If ,
-
–
when is odd, ;
-
–
when is even, .
-
–
Proof.
By the algebraic shifting construction in [4, Theorem 3.1], there is a unique shifted complex associated to where has the same -vector and -vector as . Moreover, is a near-cone. So by Lemma 2.8,
where is a Sperner family of . Moreover, since is shifted and , we can assume that is a -simplex of .
(i) The case is trivial. We do induction on . The case is clearly true. Let be an arbitrary element of . For any permutation of , since , we have
So we can deduce from (5) that .
Claim: , hence .
Assume that . Then since , we have , and hence . On the other hand, by our induction hypothesis .
-
•
When is odd, since , , a contradiction. So we must have .
-
•
When is even and , the same argument as the previous case applies. So the only remaining case is . Since is a Sperner family of , by Sperner’s theorem we have
where the equality holds if and only if consists of either all the -subsets or all the -subsets of . But since the dimension , must be the later case. This implies that and so . The claim is proved.
By the above claim, if is a minimal element of , then for every permutation of , which implies that . But is maximal in , so we can assert that .
(ii) and (iii) When , let
Observe that is a Sperner family of that satisfies the condition (2) in Proposition 2.9 with . Therefore, no subset of belongs to , which implies that every member of has nonempty intersection with . In other words, is a Sperner family of over . So by Lemma 2.11, the cardinality of satisfies:
where the equality holds if and only if is one of the three types of Sperner families listed in Lemma 2.11 with and . Then since computes , the Sperner families described in Lemma 2.11 will give us all possible members of .
- •
- •
- •
From the above discussion, we obtain the desired statements in (ii) and (iii). ∎
The following is an immediate corollary of the proof of Theorem 2.12.
Corollary 2.13.
For any -dimensional simplicial complex with vertices, the upper bound of is given by:
-
(i)
if ;
-
(ii)
if ;
-
(iii)
, if .
Moreover, the equalities in and hold if and only if belongs to as described in Theorem 2.12.
3. Simplicial complexes with the minimum total bigraded Betti number
For any -dimensional simplicial complex with vertices, there is a universal lower bound of that depends only on and . This lower bound was discovered through an interesting relation between and some canonical CW-complex associated to called real moment-angle complex of (see Davis and Januszkiewicz [9, p. 428–429] or Buchstaber and Panov [5, Section 4.1]). One way to write is
(8) |
where and is a copy of and indexed by , and denotes Cartesian product of spaces.
It is shown in [5, Section 4] (also see [6, Theorem 4.2]) that computes the cohomology groups of , which implies
(9) |
Moreover, by [6, Theorem 1.4] or [20, Theorem 3.2], there is a universal lower bound of for any simplicial complex :
(10) |
So for any -dimensional simplicial complex with vertices, we always have
(11) |
In the following, we study those simplicial complexes that make the equality in (10) hold, i.e. tight simplicial complexes (see Definition 1.5). The coefficient will be omitted in the rest of this section.
The following are some easy lemmas on the properties of .
Lemma 3.1.
If is a full subcomplex of , then .
Proof.
This follows from the formula (2) of and the simple fact that a full subcomplex of is also a full subcomplex of . ∎
Lemma 3.2.
Let be a simplicial complex with vertex set . Then . Moreover, if and only if is the simplex .
Proof.
We use the formula of in (2). Consider a minimal subset that does not span a simplex in (called a minimal non-face). If is not the empty set, then and . This implies
So if and only if all the minimal non-faces of are empty, i.e. is the simplex . ∎
Lemma 3.3.
For any finite CW-complexes and ,
where is the join of and .
Proof.
The equality follows from the Künneth formula of homology groups. In addition, by the homotopy equivalence
where “” is the smash product and “” is the (reduced) suspension, we obtain (with a field coefficient) that
(12) |
The second isomorphism in (12) follows from the relative version of the Künneth formula (see [13, Corollary 3.B.7]). Notice that is always path-connected and hence . Then it follows that . ∎
Lemma 3.4.
For any finite nonempty simplicial complexes and ,
Observe that if is a -dimensional simplicial complex with vertices, then for any , is a -dimensional simplicial complex with vertices and, by Lemma 3.4. So we obtain the corollary immediately.
Corollary 3.5.
A finite simplicial complex is tight if and only if is tight for all .
By Yu and Masuda [21, Proposition 2.1], any simplicial complex of the form is tight. So by Corollary 3.5, is also tight for all . Then one may ask whether simplicial complexes of the form or are all the tight simplicial complexes? We will see in Theorem 3.10 that the answer is yes.
For brevity, we introduce the following terms.
Definition 3.6.
For any positive integers and , we call the simplicial complex a sphere join and call a simplex-sphere join.
The following theorem proved in [21] will be useful for our proof.
Theorem 3.7 ([21, Theorem 3.1]).
Let be a simplicial complex of dimension . Suppose that satisfies the following two conditions:
-
(a)
is an -dimensional pseudomanifold,
-
(b)
the link of any vertex of is a sphere join of dimension ,
Then is a sphere join.
Recall that is an -dimensional pseudomanifold if the following conditions hold:
-
(i)
Every simplex of is a face of some -simplex of (i.e. is pure).
-
(ii)
Every -simplex of is the face of exactly two -simplices of .
-
(iii)
If and are two -simplices of , then there is a finite sequence of -simplices such that the intersection is an -simplex for all .
In particular, any closed connected PL-manifold is a pseudomanifold.
In addition, we will use the following inequality proved in [20, Theorem 3.2]:
(13) |
where is the number of vertices of , and
Lemma 3.8.
Let be a simplicial complex with vertices. If is tight, then
-
(i)
is pure.
-
(ii)
For every simplex of , is tight.
Proof.
(i) By [20, Theorem 3.2], . Then since is tight, we must have , which implies that every maximal simplex of has the same dimension as . So is pure.
(ii) We do induction on . When , this is trivial. For any vertex of , let be the number of vertices in . By the proof of [20, Theorem 3.2] (note that the argument there works for any vertex of ), there is a subspace of with , where is homeomorphic to the disjoint union of copies of . So we have
Then since , we obtain
On the other hand, by our induction we have since . So is tight.
Now suppose we have proved that is tight for any simplex of with dimension less than . Let be a -simplex in and let be a vertex of . So is a -simplex and it is easy to see that
(14) |
By our assumption, we know that is tight. So by our preceding argument, we can assert that is also tight. This finishes the proof. ∎
Lemma 3.9.
If a simplicial complex is tight but not connected, then must be .
Proof.
If , then being tight implies that is isomorphic either to or to . If with and is not connected, let where and are two subcomplexes of that are disjoint. Then we can add a -simplex to with a vertex and and obtain a new simplicial complex . Observe that adding the -simplex to kills the generator of . So for any , it is easy to see that
So by the formula (2) of , we can deduce that . But since is tight, contradicting the inequality in (11). ∎
Now, we are ready to prove the following theorem which answers Question 3.
Theorem 3.10.
A finite simplicial complex is tight if and only if is of the form or for some positive integers and .
Proof.
By our preceding discussion, any sphere join or simplex-sphere join is tight. Conversely, suppose is tight and we do induction on . When , the statement is trivial. For , by Lemma 3.8 and Lemma 3.9, is connected, pure and is tight for every simplex of . Then since the number of vertices of is less than , our induction hypothesis implies that is either a sphere join or a simplex-sphere join.
Case 1: If for every vertex of , is a sphere join, then by the relation in (14), we can inductively prove that is also a sphere join for every simplex of . This implies that is a closed connected PL-manifold hence a pseudomanifold. So by Theorem 3.7, is sphere join.
Case 2: If there exists a vertex such that is a simplex-sphere join, let . Since is a pure -dimensional simplicial complex, . Take a vertex and consider the full subcomplex of . Note that
which has dimension . So the dimension of is , which implies
But removing a vertex can reduce at most by one. Then since is pure, . So . Then by (13), we obtain
But by Lemma 3.1, . So we have
(15) |
Moreover, by the formula (2) of , we obtain
So the equality (15) implies that for every that contains .
Claim: For any simplex of , is a simplex of .
We prove the claim by induction on . When , i.e. is a vertex, since . Assume that the claim is true when . If , then by induction is a simplex for every . If is not a simplex of , then is isomorphic to the boundary of an -dimensional simplex. But this contradicts the above conclusion that . So the claim is proved.
By the above claim, is a cone of with . It follows that
So by (15), , i.e. is tight. Then by our induction, is either a sphere-join or a simplex-sphere join, which implies that is a simplex-sphere join. The theorem is proved. ∎
4. Simplicial complexes with the maximum total bigraded Betti number
In this section, we give a complete answer to Question 4. First, we prove a lemma parallel to Lemma 2.1 on total bigraded Betti number.
Lemma 4.1.
For any two simplicial complexes with vertex set ,
Proof.
Theorem 4.2.
If is a simplicial complex with vertex set , then
for any field , where the equality holds if and only if .
Proof.
Let , which is a partially ordered set with respect to the inclusions of simplicial complexes.
Suppose is a minimal or a maximal element of . Then for every permutation of the vertex set of , it follows from Lemma 4.1 that
This implies that . So must be where . Observe that any nonempty full subcomplex of on vertices with is a simplex. So to compute via the formula (2), we only need to consider the full subcomplexes of with more than vertices. Moreover, it is easy to see that the reduced homology group of any nonempty full subcomplex of always concentrates at degree . Then an easy calculation shows that
So the theorem follows from Lemma 5.1 in the Appendix. ∎
Remark 4.3.
For a simplicial complex with vertex set , the following combinatorial invariants of were studied by Codenotti, Spreer and Santos [8]:
Formally, is some sort of weighted average of the -th Betti number of all the full subcomplexes of , which has a similar flavor as our .
5. Appendix
Lemma 5.1.
For , reaches the maximum when and only when .
Proof.
The cases can be checked by hand, so we assume in the rest of the proof. It is easy to verify that
(16) |
So , which implies that
Then we can deduce that
-
•
if , then is maximal only when or ;
-
•
if , then is maximal only when or ;
-
•
if , then is maximal only when or (since in this case).
So to prove the Lemma, we only need to prove: for any ,
-
(a)
;
-
(b)
;
-
(c)
.
Note that by (16),
We directly compute
(17) |
So we obtain
Moreover, by the Cauchy-Schwarz inequality:
So to prove (b) and (c), we only need to show for and , which is easy to check. Similarly,
where we use the result of (5) with substituted by and the relation . Then again by the Cauchy-Schwarz inequality, we obtain
So to prove (a), we only need to show for , which is also easy to check. This finishes the proof. ∎
References
- [1] A. Abbondandolo, Morse theory for Hamiltonian systems, Chapman Hall/CRC Res. Notes Math., 425, 2001.
- [2] C. Allday and V. Puppe, Cohomological methods in transformation groups, Cambridge Studies in Advanced Mathematics 32, Cambridge University Press, 1993.
- [3] I. Anderson, Combinatorics of finite sets, Dover Publications, Inc., Mineola, NY, 2002.
- [4] A. Björner and G. Kalai, An extended Euler-Poincaré theorem, Acta Math. 161 (1988), no. 3-4, 279–303.
- [5] V. M. Buchstaber and T. E. Panov, Toric Topology. Mathematical Surveys and Monographs, Vol. 204, American Mathematical Society, Providence, RI, 2015.
- [6] X. Cao and Z. Lü, Möbius transform, moment-angle complexes and Halperin-Carlsson conjecture, J. Algebraic Combin.35 (2012), no. 1, 121–140.
- [7] G. Carlsson, Free actions and a problem in commutative algebra, Lecture Notes in Math., 1217, Springer, Berlin, 1986, 79–83.
- [8] G. Codenotti, J. Spreer and F. Santos, Average Betti numbers of induced subcomplexes in triangulations of manifolds, Electron. J. Combin. 27 (2020), no.3, Paper No. 3.40.
- [9] M. W. Davis and T. Januszkiewicz, Convex polytopes, Coxeter orbifolds and torus actions. Duke Math. J. 62 (1991), no. 2, 417–451.
- [10] P. Erdös, C. Ko and R. Rado, Intersection theorems for systems of finite sets, Quart. J. Math. Oxford Ser. (2) 12 (1961), 313–320.
- [11] J. Griggs, Collections of subsets with the Sperner property, Trans. Amer. Math. Soc. 269 (1982), no. 2, 575–591.
- [12] S. Halperin, Rational homotopy and torus actions, In: Aspects of Topology. London Math. Soc. Lecture Note Series, 93. Cambridge Univ. Press, Cambridge (1985), 293–306.
- [13] A. Hatcher, Algebraic topology, Cambridge University Press, Cambridge, 2002.
- [14] M. Hochster, Cohen-Macaulay rings, combinatorics, and simplicial complexes, Ring Theory II (Proc. Second Oklahoma Conference), B. R. McDonald and R. Morris, eds. Dekker, New York (1977), 171–223.
- [15] G. Kalai, Characterization of -vectors of families of convex sets in , I. Necessity of Eckhoff’s conditions, Israel J. Math. 48 (1984), no. 2-3, 175–195.
- [16] G. Kalai, -vectors of acyclic complexes, Discrete Math. 55 (1985), no. 1, 97–99.
- [17] K. W. Lih, Sperner families over a subset, J. Combin. Theory Ser. A 29 (1980), no. 2, 182–185.
- [18] E. Sperner, Ein Satz über Untermenge einer endlichen Menge, Math. Z., 27 (1928), 544–548.
- [19] R. Stanley, Combinatorics and commutative algebra, 2nd edition. Birkhäuser Boston, 2007.
- [20] Y. M. Ustinovsky, Toral rank conjecture for moment-angle complexes, Mathematical Notes 90 (2011), no. 2, 279–283.
- [21] L. Yu and M. Masuda, On descriptions of products of simplices, Chinese Ann. Math. Ser. B 42 (2021), no. 5, 777–790.