1. Introduction
The concept of atomic decomposition has been of great interest in both philosophy and physics for centuries; its genesis presumably goes back to Leucippus and Democritus, 400 BC. During the last decades, different formal notions of atomicity have been formalized and studied in several areas of mathematics. The question of the existence of irredundant atomic decompositions in lattices with certain additional structures has been investigated since the middle of last century. In this regard, P. Crawley introduced a decomposition theory for nonsemimodular lattices [
1] and collaborated with R.P. Dilworth in the development of a decomposition theory for lattices without chain conditions [
2]. These contributions found further development in the work of M. Enre [
3], and, more recently, in the contributions of T. Albu [
4], which are inspired in applications to Grothendieck categories and torsion theories. A compilation of Dilworth’s theorems is presented in [
5]. In the context of commutative monoids and integral domains, atomic decompositions, also referred to as factorizations, were considered in the 1960s by L. Carlitz [
2] and P. Cohn [
3], in the 1970s by A. Grams [
6] and A. Zaks [
7], and in the 1980s by J. L. Steffan [
8] and R. Valenza [
4]. In 1990, D. D. Anderson, D. F. Anderson, and M. Zafrullah provided the first systematic investigation of atomic decompositions in the context of integral domains [
1]. From that point on, atomic decompositions have been systematically studied by many authors under the umbrella of factorization theory (see [
5,
9] and references therein). The reach of atomic decompositions has expanded significantly, influencing core developments in mathematics and computer science. In this regard, there are several open problems in algebra about the atomicity of certain classes of monoids [
7,
10,
11]. In probability, Mishura et al. investigated the atomic decompositions and inequalities for vector-valued discrete-time martingales [
8]. Along similar lines, X. Zhang investigated the atomic decompositions of Banach lattice-valued martingales and used them to study the relation of the martingale spaces [
12]. In the domain of ontologies, [
6,
9] used atomic decompositions to accurately represent the family of all locality-based modules of a given ontology.
This paper delves into atomic decompositions within the context of geometric lattices. To the best of our knowledge, no existing work comprehensively addresses this topic from the perspective presented here. When analyzing finite geometries, two key aspects must be considered: incidence, which deals with inclusion relationships between geometric objects (e.g., “a point lies on a line”), and measurement, which pertains to properties such as length, area, and angles. By focusing solely on incidence, we can establish fundamental properties that a finite geometry should satisfy. 1. There exists a base set of points, say, , and the geometric elements (e.g., points, lines, planes) are subsets of this base set. 2. Every point in is an element of the geometry, and itself is also considered an element of the geometry. 3. The intersection of any geometric elements is also an element of the geometry. 4. A rank or dimension function exists, which captures the hierarchical structure of the geometric elements. If we order these geometric elements by set inclusion, the resulting structure forms a lattice, where the points of the geometry correspond to the atoms of this lattice. This motivates the definition of a finite geometric lattice as an algebraic lattice , where ⪯ is the partial order relation, and ∨ and ∧ denote the join operator and the meet operator, respectively, which is semimodular (i.e., implies , where ⊏ stands for the covering relation (i.e., if and only if and for all , implies that either or )) and atomistic (i.e., every element different from the minimum element of the lattice can be expressed as the join of atoms (i.e., join irreducible elements of L)). To highlight the significance of geometric lattices in both algebraic and geometric contexts, it is worth noting two well-established facts: A geometric lattice is isomorphic to (1) the lattice of ideals of a finite, atomistic, semimodular, and principally chain lattice; and (2) the lattice of flats in a simple matroid.
A
finite matroid is a pair
where
S is a finite set and
is a subset of
such that (1) if
and
, then
; and (2) for every
, the maximal elements of
have the same cardinality. In a matroid
M, the rank of a subset
, denoted by
, is defined to be the maximum cardinality of those elements of
that are contained in
H, i.e.,
. A
k-flat of
M is a maximal subset of rank
k. The set of all flats of
M ordered by inclusion is known to be geometric lattice, which we shall denote by
.
M is said to be simple if there is no
satisfying
.
Theorem 1 ([
13]).
Let L be a finite lattice. L is geometric if and only if L is isomorphic to for some simple matroid M. An archetypal example of a geometric lattice is the set of all partitions of a finite set X with n elements, ordered by partition refinement.
A partition of a finite set X with elements is a collection of disjoint nonempty subsets of X whose union is X. That is, , with , for , and . The set of all partitions of the set X is endowed with the refinement of partitions. The partition is said to refine the partition if and only if every block of is the union of some blocks of (in particular, every block of is contained in some block of ). The notation means refines , which is also read as is finer than , and as is coarser than . If , and for every partition , implies that either or ; then, is said to cover the partition . Henceforth, we denote by the space of all partitions of X. We shall also use the notation to mean that the elements lie in the block of the partition .
The operations with partitions are another important ingredient of , which endow this set with a rich structure. For any two partitions and in , there can always be found partitions that refine both. The coarsest partition that satisfies this property is denoted by and is called the meet of and . The blocks of are all possible nonempty intersections of a block of and a block of . Similarly, for any two partitions and there can always be found partitions that are refined by both. The finest partition that satisfies this property is denoted by and is called the join of and . Two elements are placed in the same block of if and only if there is a sequence such that, for all , and are either placed in the same block of or in the same block of . Thus, the blocks of are the subsets of X that can be expressed both as unions of blocks of and as unions of blocks of , and are minimal with respect to this property. These operations convert into a lattice. Moreover, it is well-known that for any two partitions and , if covers , then covers , which makes an upper semimodular lattice.
There are two distinguished elements in the lattice : the finest partition, denoted by , in which all blocks are singletons, and the coarsest partition, denoted by , consisting of a single block equal to X. These serve as the neutral elements for the meet and join operations, respectively.
Moreover, is a ranked lattice, with rank function defined by , where is the number of blocks in the partition , and . This rank function measures the distance from the finest partition, , and organizes into layers based on the number of blocks.
Each covering relation corresponds to merging exactly two blocks, decreasing the number of blocks by one and increasing the rank by one. Every maximal chain has length and the rank increases uniformly along chains.
Furthermore, the partitions of X which cannot be but trivially decomposed as the join of partitions of X are called the atoms of . A partition is an atom if all its blocks are singleton except for exactly one, which contains exactly two elements. From now on, we shall denote by the atom whose only non-singleton block is , by the set of all atoms of , and by the set of those atoms of that refine the partition . It is straightforward to note that every partition is the join of all those atoms in . Thus, is also atomistic, and, therefore, a geometric lattice.
Several characterizations have been provided for partition lattices, including those by Ore [
14] and Sasaki and Fujiwara [
15]. Here, we have chosen to include a characterization that somehow encompasses the previous ones:
Theorem 2 (J. R. Stonesifer and K. P. Bogart [
16]).
Let L be a geometric lattice with a modular copoint m such that, for every point p, the interval is isomorphic to . Then, for , L is isomorphic to . In the scope of Theorem 2, a copoint should be interpreted as an element whose rank is one less than the rank of L. Additionally, m is said to be a modular element if, for every , the condition implies , for every . Also, 1 stands for the maximum element of L. Here is an interesting corollary of Theorem 2 for supersolvable geometric lattices—there exists a maximal chain of L such that, for every chain K of L, the sublattice generated by K and is distributive.
Corollary 1. If L is a supersolvable geometric lattice such that every interval has a characteristic polynomial which equals the characteristic polynomial of a partition lattice, then L is isomorphic to a partition lattice.
Young-jin Yoon rephrased Theorem 2 for combinatorial geometries [
17], thereby emphasizing the geometric relevance of this result. Given a geometry
G, a subset
T of the points of
G,
stands for the contraction of
G by
T—the geometry induced by the geometric lattice
on the set of all flats in
G covering the closure of
T,
. Moreover, for a geometry
G,
denotes the geometric lattice whose elements are the flats of
G with the set inclusion as the partial order.
Theorem 3 (YJ. Yoon [
17]).
If a geometry G has a modular copoint and, for every point , is isomorphic to , then is isomorphic to for . This paper focuses on geometric lattices that are isomorphic to a lattice of partitions, including the lattices of all equivalence relations on a finite set and the lattices of all factorizations of an integer q, ordered by inclusion. However, this work excludes other important examples of geometric lattices, such as the lattice of all subspaces of a vector space over a finite field.
An atomic decomposition of a partition is a subset of such that the join of the atoms in equals . While itself forms an atomic decomposition of the partition , it typically includes superfluous atoms. The objective is often to generate a given partition by using the fewest possible atoms. An atom in an atomic decomposition is considered redundant if remains an atomic decomposition of . Thus, a decomposition is minimal if it contains no redundant atoms. If an atomic decomposition of the partition is not minimal, there exists an atom such that removing it, i.e., redefining , still results in an atomic decomposition of . This process of eliminating superfluous atoms can be repeated until a minimal decomposition is achieved. Since is finite, removing all redundant atoms requires only a finite number of steps. Consequently, every partition admits a minimal atomic decomposition. Henceforth, denotes the set all whose elements are the minimal atomic decompositions of .
Let us illustrate this notions with a simple example. The partition
of
admits the following atomic decompositions (among others):
and
Decompositions
and
are minimal, while decomposition
is not. To show this, let us lean on the following graphical illustration of
(see
Figure 1), where for each atom in the decomposition, we placed a segment of line connecting the only two elements in its only non-singleton block.
Notice that, according to the definition of join, the atoms in any atomic decomposition of should ensure the existence of a sequence of line segments for any two elements in the same block of . If the decomposition is minimal, then only one of such sequences exists for any two given elements lying in the same block of . It can be straightforwardly seen now that all atoms in and are needed, while in , we can remove one atom from among the two atoms , , and and one atom from among of the atoms , , and .
Contributions of the Article
Firstly, this article investigates some basic properties of the atomic decomposition in the lattice of partitions, including those introduced by D.D. Anderson, D. F. Anderson, and M. Zafrullah for the systematic study of commutative algebraic structures in factorization theory. Then, we focus on the key characteristic features of the function , which maps each partition of X into the number of minimal atomic decompositions of . Secondly, assuming that a subset of the set atoms is given, referred to as the set of red atoms, we derive a recursive formula for : the total number of partitions of X lying at the jth level of that can be expressed as the join of exactly s red atoms. (The term “red atoms” is used to emphasize that only a designated subset of atoms is considered, rather than the full set of atoms in the lattice.)
From a geometrical perspective, atomic decompositions can be understood as the process by which higher-dimensional structures, such as lines and planes, are formed by combining the smallest “building blocks” of the geometry, typically referred to as points. Emphasizing atomic decomposition in geometric lattices highlights the essence of constructing complex structures from simple, indivisible elements, reflecting both algebraic and geometric viewpoints.
To facilitate accessibility and understanding,
Table 1 compiles the main notations used throughout this article.
2. Basic Facts About Atomic Decompositions in
Our first theorem enables us to leverage well-known results from graph theory to describe fundamental properties of the atomic decomposition of partitions.
Along this paper we will use the following maps:
The map that takes a partition and returns the labeled non-directed graph on X, where if and only if .
The map that takes a labeled undirected graph on X and returns the partition .
Theorem 4. Let be an arbitrary partition of X. Then,
- 1.
There is a bijection between the set of all atomic decompositions of π and the set of all spanning subgraphs G of such that .
- 2.
There is a bijection between the set of all minimal atomic decompositions of π and the set of all spanning forests of .
Proof. Let be an atomic decomposition of the partition . We construct a labeled spanning subgraph of as follows: The vertex set of is X. Two vertices are connected by an edge in if and only if the atom . By construction, since each atom in is determined by a unique pair of elements forming its only nonsingleton block, is necessarily loop-free. Moreover, by virtue of definition of the join operator, the graph respects the block structure of , meaning that all edges connect elements within the same block of , and any two vertices in the same block of are connected by a path in . Thus, a subgraph C is a connected component of if and only if there is a block , , of such that C is a connected graph on the elements of . Therefore, is a spanning subgraph of such that .
Let be the function that assigns the spanning subgraph to the atomic decomposition of . We argue that is a bijection. Indeed, suppose and yield the same spanning subgraph of , i.e., . Then, and have identical edges, implying that and consist of the same atoms. Thus, , proving injectivity. Let us consider now a spanning subgraph of . Define as the set of atoms corresponding to the edges of . Then, is precisely , proving surjectivity.
To complete the proof, note that a decomposition is minimal if and only if is a forest. Any cycle in implies the existence of redundant atoms in , contradicting minimality. Conversely, removing superfluous atoms in a non-minimal decomposition corresponds to removing edges from , resulting in a forest. □
Corollary 2. The union of atomic decompositions of a partition π is also an atomic decomposition of this partition.
Proof. This follows immediately from Theorem 4 and the fact that the union of spanning subgraphs of a graph is also a spanning subgraph of this graph. □
Corollary 3. The subset is an atomic decomposition of the partition π if and only if for every atom , if and only if there exists a sequence of distinct atoms in such that and . Furthermore, is minimal if and only if, for every atom , the sequence above is unique.
Proof. From Theorem 4, we know that is an atomic decomposition of if and only if is a spanning subgraph of . Since an atom refines (i.e., ) if and only if x and lie in the same block of , it follows that if and only if there exists a path in connecting x and . The edges of this path correspond to the desired sequence of atoms .
Furthermore, is minimal if and only if is a forest. In this case, any path between two vertices x and in is unique, ensuring that the sequence of atoms connecting x and is also unique. Conversely, if the sequence of atoms is unique for every pair x and such that , then must be a forest, and is minimal. □
Corollary 4. All minimal atomic decompositions of a partition π consist of the same number of atoms, which is , where stands for the number of elements in the ith block of π. Additionally, the number m of atoms in any atomic decomposition of the partition π satisfies Proof. The number of atoms in a minimal atomic decomposition of is determined by Theorem 4, which ensures that is a spanning forest of . Since every connected component of corresponds to a block of and is a tree, the number of edges (and hence atoms) in is precisely .
For an arbitrary atomic decomposition, the lower bound follows because it must contain at least the atoms of a minimal decomposition to preserve connectivity within each block. The upper bound is attained for the atomic decomposition , leading to the sum . □
Corollary 5. The union of different minimal atomic decompositions of a partition π is not minimal.
Proof. From Corollary 4, all minimal atomic decompositions of have the same number of atoms. This implies that the union of two distinct minimal atomic decompositions strictly contains both decompositions. Consequently, the union exceeds the size of a minimal decomposition, violating minimality. □
Our next result provides an analytical expression for the function .
Corollary 6. Let be the function which maps each partition π of X to the number of its minimal atomic decompositions. Then, for every partition , with and denoting the number of nonsingleton blocks of π, the following holds: Proof. The graph has k connected components, each being a complete graph on the elements of a block of . It is well known that the number of spanning forests of a graph is given by the product of the number of spanning trees of its connected components. Specifically, for , this is , where is the number of spanning trees of the ith connected component.
Using Cayley’s formula, which states that a complete graph on
vertices has
spanning trees, we deduce that if the
ith connected component of
contains
vertices, the number of spanning trees of that component is
. Finally, applying Theorem 4, which establishes a correspondence between minimal atomic decompositions of
and spanning forests of
, we conclude that
□
We now turn to analyzing the atomic decompositions of from an algebraic perspective. To this end, and for the sake of completeness, we shall briefly review some relevant algebraic concepts.
A monoid is a set equipped with an associative binary operation with an identity element. Thus, with the join operator ∨ is a commutative monoid. For the systematic study of the atomic decompositions in monoid, usually referred to as factorizations, D.D. Anderson, D.D. Anderson, and M. Zafrullah introduced the following properties:
A subset I of a monoid M is an ideal of M provided that . The ideal I is principal if for some . The monoid M satisfies the ascending chain condition on principal ideals (or ACCP) if each increasing sequence of principal ideals of M eventually stabilizes.
An atomic monoid M is said to be a finite factorization monoid (FFM) if every element admits only finitely many factorizations.
An atomic monoid M is said to be a bounded factorization monoid (BFM) if, for every element, there is an upper bound for the number of atoms (counting repetitions) in each of its factorizations.
An atomic monoid M is said to be half-factorial (HFM) if any two factorizations of the same element have the same number of atoms (counting repetitions).
An atomic monoid M is said to be unique factorization monoid (UFM) if every element has a unique factorization.
It is well-known that HFM implies BFM, which, in turn, implies ACCP, and, consequently, implies atomic. Additionally, UFM implies FFM, which, in turn, implies BFM.
Since is finite, it satisfies the ACCP condition. However, the join operator is idempotent, i.e., for every partition , we have , which means that we can add an atom to an atomic decomposition as many times as we wish. This behavior causes to fail all the other conditions.
In this context, however, it is more beneficial to disregard repeated atoms, as they are never necessary and may hinder our ability to identify distinguishing features among finite lattices. For this reason, we will consider variants of the conditions above in which repeated atoms are not allowed.
Proposition 1. The lattice satisfies the ACCP, FFM, and BFM conditions and fails to satisfy HFM and UFM conditions. (All conditions here do not allow repeated atoms).
Proof. Given a partition , let denote the number of connected components of the graph . Theorem 4 states that the number of atomic decompositions of is equal to the number of spanning subgraphs of with connected components. Therefore, satisfies the FFM condition and, consequently, the BFM and ACCP conditions.
However, the number of spanning subgraphs of the depends on the sizes of the blocks of , which form a partition of the integer n into parts. Consequently, there exist partitions whose corresponding partitions of the integer n differ and, therefore, yield different numbers of atomic decompositions—for instance, an atom and any partitions that covers it. □
We introduce here a novel condition of the same kind, which enables us to distinguish among ranked lattices, although it does not extend to general monoids.
A ranked lattice L is said to satisfy the half-factorial ranked lattice (HFRL) condition if, for any two elements , the equality implies that x and have the same number of atomic decompositions. This property, for instance, distinguishes the lattice of partitions from the lattice of subspaces of a vector space over a finite field, since the former fails to satisfy it, while the latter does.
Lemma 1. fails to satisfy HFRL.
Proof. Any two partitions and at the same level of have the same number of blocks, which means that their respective sequences of block sizes correspond to partitions of the integer n into the same number of parts. However, we can always choose and with corresponding partitions of n that differ in a way that leads to a different number of atomic decompositions. For example, suppose and each have three blocks, with block size sequences 1+2+2 and 1+1+3. The number of atomic decompositions is 2 and 4, respectively. □
Aiming at investigating the other properties of minimal atomic decompositions by analyzing the behavior of the function along chains in , we conclude this section with two outcomes that shed some light on the topology of the lattice of partitions.
A chain in the lattice of partitions is a sequence of partitions in which every partition refines all partitions that succeed it in the sequence; i.e., . Notice that, in particular, for every pair of partitions such that , there is a chain in which every partition is covered by its immediate successor: . Let us suppose that and let and denote the two blocks of needed to be merged to obtain . Note that if and only if and . Let and be elements satisfying this condition and let us define the mapping that to every minimal atomic decomposition of assigns .
Proposition 2. The map is an injective function whose range lies within the set of all possible atomic decompositions of . Moreover, if we consider a different pair of elements and , the respective ranges of and do not overlap.
Proof. To verify that is an injective function, let us consider two atomic decompositions of and of the partition such that . By definition of , this implies that . Since both sets contain the same atoms, . Thus, is injective.
To prove that the ranges of and do not overlap for a different pair of elements and , observe the following: Every atomic decomposition returned by contains the atom but not the atom , and conversely, every atomic decomposition returned by contains the atom but not the atom . Since no atomic decomposition can simultaneously contain both and , the ranges of and are disjoint. □
It is worth remarking that not all minimal atomic decompositions of can be obtained this way from a minimal atomic decomposition of . Let and consider the partitions , and . Note that . One can straightforwardly verify that is a minimal atomic decomposition of which cannot be returned by any of the maps . In particular, note that none of the atoms in this decomposition refine .
The next corollary is basically a consequence of the fact that the composition of injective functions is an injective function as well. In spite of its simplicity, we decided to be emphatic about it here because we will need it later.
Corollary 7. Let and any maximal chain from π into . If , then the composition is an injection from the set of minimal atomic decomposition of π into the set of minimal atomic decompositions of .
The previous results offer valuable insights into retrieving minimal atomic decompositions of partitions as we move vertically within the Hasse diagram of . But what about horizontal movement? In other words, is there a relationship between the minimal atomic decompositions of partitions within the same layer of ? The following proposition addresses this question and provides some clarity.
Proposition 3. Let and be partitions of X such that there exists a permutation τ of the index set satisfying for every . Then, for every permutation of the elements of X such that, for any , x and lie in the same block of π if and only if and lie in the same block of , the mapping which assigns to every minimal atomic decomposition of π the minimal atomic decomposition of defined byis a bijection from the set of all minimal atomic decompositions of π onto the set of all minimal atomic decompositions of . Proof. The proof follows directly by constructing the inverse map of
. Define
, which maps each minimal atomic decomposition
of
back to a minimal atomic decomposition
of
, as follows:
where
is the inverse of
. Clearly,
reverses the mapping of
, and, thus,
is bijective. □
4. Constrained Atomic Decompositions
Let us now consider a subset of atoms, referred to as red atoms. (We refer to this distinguished subset as “red atoms” to highlight that our analysis focuses on a specific—albeit arbitrary—selection of atoms, rather than the entire set of atoms in the lattice.) The purpose of this section is to derive a recursive formula for —the total number of partitions lying at the jth level of which can be expressed as the join of s red atoms.
To this end, it is useful if we can determine first the number of partitions in that can be expressed as the join of red atoms. Henceforth, let denote the set of partitions of X that can be generated with red atoms. In other words, the partition if and only if there are red atoms such that (regardless the number of red atoms k and the level at which lies).
Let denote the set all whose elements are the spanning subgraphs of , and let us consider the equivalence relation ∼ on given by the following: For all , if and only if .
Proposition 4. is equipotent to the quotient set .
Proof. Let G be a spanning subgraph graph of and consider the partition . Since the edges of correspond to the atom in , can be expressed as the join of red atom. Let us consider now the map that to a spanning graph G of assigns the partition . We argue that is surjective. Indeed, take any partition and define the spanning subgraph G of whose edges correspond to those red atoms that refine . Since we can recover by taking the join of such red atom, we can conclude . The results now follow from the first isomorphism theorem. □
Now, we shall focus on computing the number of equivalence classes in .
Lemma 2. If is a forest, then the quotient set consists of exactly equivalence classes.
Proof. We will start by proving that if with edge sets and , respectively, are spanning subgraphs of for which there exists an edge satisfying and , then and are not equivalent.
Indeed, since is a forest, the only path connecting to in is the edge . This implies that if any spanning subgraph of contains a path connecting to , that path must be the edge . Therefore, there is a path in connecting to but fails to contain such a path. This means that and are in the same connected component of but in different connected components of , which proves that the respective induced partitions and are different. Consequently, and are not equivalent. Thus, every spanning subgraph is just equivalent to itself, which means that the quotient set has as many equivalent classes as spanning subgraphs . It follows, then, that such a number is . □
Lemma 3. If is a (simple) cycle, then, for any , if , and and , and are not equivalent.
Proof. We shall divide the proof in two cases.
Case 1: . Then, the graph is a forest, and and are also spanning subgraphs of . Combining Lemma 2 with the fact that , we conclude that and are not equivalent.
Case 2: . Let us suppose that and are two of the edges that were removed from to obtain , i.e., . Then, there is no path in connecting x and . However, since , and, hence, there is a path in connecting x and . This proves that and are not equivalent. □
The next result is obvious.
Lemma 4. If is a cycle, then every spanning subgraph of such that is equivalent to .
Combining Lemmas 3 and 4, we get the following result.
Proposition 5. If is a cycle, then the quotient set consists of exactly equivalence classes.
Proof. From Lemma 4, we know that the equivalence class of consists of all those spanning subgraphs of obtained from by removing at most one edge are equivalent, which add up . From Lemma 3, every spanning subgraph of obtained from by removing more than one is only equivalent to itself. Since there is a total of spanning subgraphs, there are of such equivalence classes. Therefore, total number of equivalence classes is . □
Theorem 6. Let , , be pairwise edge-disjoint spanning subgraphs of such that
- 1.
, i.e., . (Here, by the sum of two graphs and , we mean the graph .)
- 2.
For every cycle C in , there is such that C is a subgraph of .
Then,
Proof. Note that if is a spanning subgraph of , then our hypothesis ensures that we can find such that, for every , , such that .
Let us consider now two arbitrary spanning subgraphs and of , and let and be their respective decompositions satisfying the condition above, i.e., for every , , , and , . We argue that if there is an index such that are are not equivalent, then and are not equivalent either. Indeed, if are are not equivalent, then there are elements x and in X which are connected by a path in one of these graphs but not in the other. Without loss of generality, let us suppose that is the subgraph containing a path that connects x and . This implies that x and are connected by path in . Now, notice that x and cannot be connected by a path in because such a path must be contained in one of the , for some . But this either contradicts the fact that , which is edge-disjoint with , so it cannot contain the same path we used for , or, if it were a different path, the fact that every cycle of is contained in one of the ’s. Therefore, and are not equivalent. The desired formula now follows immediately. □
Corollary 8. If all the cycles of are pairwise edge-disjoint, and T is the spanning subgraph all whose edges are those that lie in no cycle of , thenMoreover, by virtue of Lemma 2 and Proposition 5, Formula (10) can be rewritten aswhere denotes the edge set of the cycle . Corollary 8 can be generalized as follows. Consider the equivalence relation on the set of all cycles of given by the following: Two cycles, C and , are equivalent if there exists a sequence of cycles such that any two consecutive cycles in the sequence have common edges (at least one). Henceforth, denotes the equivalence class of the cycle C and denote the sum of the cycles equivalent to the class representative C.
Corollary 9. Let be all the cycles of and let T be the spanning subgraph all whose edges are those that lie in no cycle of , then Given a collection
of spanning subgraphs of
and
, i.e.,
, we shall consider the following subcollections:
Proposition 6. Let be a collection of spanning subgraphs of and let be an edge for which . If, for every graph containing a path p connecting x and , there is a graph equivalent to G, then Proof. Notice first that for any graph , the corresponding partition places x and in different blocks, while the partition corresponding to a graph places x and in the same block. Thus, the every graph can only be equivalent to graphs lying in . This implies that the equivalence class of a in the entire set is the same as its equivalence class in the subset . On the other hand, every graph containing a path p that connects x and is equivalent to some graph ; the remaining classes admit a representative from . This proves the result. □
Corollary 10. satisfies the premises of Proposition 6 for every edge . Therefore, By virtue of Proposition 4, the previous results can be rephrased in term of atomic decomposition. We shall focus on Proposition 6.
Given a red atom
and a collection
of nonempty subsets of
, we shall consider the following subcollections:
Note that for .
Proposition 7. Let be a collection of nonempty subsets of and let be a red atom for which . If, for every such that , there is such that , then Proof. Notice that if , then, by definition, for some . If , by hypothesis, there is such that . Thus, that is refined by the atom actually belongs to . On the other hand, if is not refined by , then . Therefore, the proposition follows. □
Henceforth, we adopt a more detailed notation for certain subsets of . We denote by the set of all partitions of X that can be expressed as the join of exactly s atoms in ; that is, , for ). Here, the symbol ∗ indicates that the level (or rank) at which the partition lie is arbitrary and unspecified. We further define as the subset of consisting of those partitions lying at the jth level of , that is, partitions of rank j. It is important to observe that the fact that does not preclude from also belonging to , for some . Analogously, we define the sets and .
Proof. Let J be an arbitrary size-s subset of . We claim that, for every atom , if , then there is such that , , and . Indeed, if , we may simply take . Otherwise, let be the graph whose vertex set is X and whose edges correspond to the atoms in J. Since , there is a path p in connecting x and . Consider the graph obtained from G by adding the edge , that is, setting . Observe that has exactly one more edge than and contains the cycle formed by the path p together with the edge e. Let be another edge of this cycle. Then, setting , the graph obtained from by removing the edge has as many edges as . In addition, since maintains the same connectivity as , . Letting be the subset of corresponding to the edges of , we conclude that . The result now follows from Proposition 7. □
We shall conclude this section by establishing a recursive formula for the . Here, recursive formula refers to expressing as linear combination of instances of the same formula involving smaller values for the parameters, where at least one of the following reductions occurs: (1) the set X is reduced; (2) the set is reduced; (3) the number if j is reduced; and (4) the number s is reduced.
Let us assume that is an arbitrary red atom. If, in addition to the edge , there exist distinct paths in , each connecting x and with respective lengths , then for a t-dimensional multi-index , where for each , we construct the set by selecting, from each path , the edge at position . The goal is to construct a minimal cut in which every path in interrupted exactly once. If an edge is removed from a path, and that same edge belongs to another path under consideration, then the latter path must not be interrupted again. Thus, contains at most t different edges.
Now, if is a -dimensional multi-index, that is, , where each , , is a t-dimensional multi-index, then we define . W will denote the dimension of any multi-index by .
Furthermore, given , a subset of path indexes, define as the graph obtained from by collapsing the entire union of the paths , into a single new vertex . That is, all vertices and all edges that belong to any of the path , , are identified with the vertex . Every edge in such that b is a vertex of is replaced in by the edge , redirecting all connections to the collapsed component. Every edge in that is disjoint from the union of paths is retained unchanged in .
Let denote the vertex set of , and let be the set of atoms of that correspond to the edges of . Additionally, we denote by the join of the atom of that correspond to the edges of , and by the total number of such edges.
Theorem 7. , and for every , . Additionally, for and any atom , we have the following:
- 1.
If the only path in connecting x and is the edge , thenwhere . - 2.
If, in addition to the edge , there are different paths whose lengths are , respectively, connecting x and in , then
- (a)
, for .
- (b)
- (c)
For ,where .
Proof. The join of s atoms lies at most at the sth level of . Therefore, for every , and consequently, .
According to Corollary 11,
Let us suppose that . If e is the only path connecting the elements x and , then e is a cut-edge and, hence, the sets and are equipotent, and their common cardinality is the same as that of the set , namely, . Indeed, if , then . Moreover, if and , then given that x and belong to different connected components of . On the other side, since for every , fails to refine , the atom is never part of the s red atoms used to decompose as the join of s red atoms. Hence, the cardinality of is the same as that of . This proves 1.
However, notice that in the case that there are other paths in connecting x and , if , with , atoms it is not possible to complete a path other than e connecting x and . This means that e is still a cut-edge in every subgraph of corresponding to the chosen size-s subset of red atoms containing e. In this case, the same equality holds, which proves 2.(a).
To prove 2.(b), notice that if , then, in addition to e, there is at least another path connecting x and . To ensure that cycles are not formed in the corresponding graph, we need to interrupt every path different from e connecting x and . For each t-dimensional multi-index , let us consider , where is as defined above. Notice that is a minimum cut that separates x and . Thus, . Indeed, if , then Proposition 7 ensures that is the join of a size-j subset of that contains . Since has rank j, the corresponding graph does not contain cycles, and, therefore, e is a cutting edge of this email. Removing e results into a partition . Obviously, if and , then the corresponding partitions and are different. Notice also that the absence of cycles in means that all paths connecting x and have been interrupted, and, hence, , for some minimum cut . Analogously, we can conclude that . Statement 2.(b) now follows from the Inclusion–Exclusion principle to each of the unions above.
Note first that if
, with
, then
for some subset
such that
and
.
Let be the graph defined as follows: an edge if and only if or there exists such that is an edge of the path .
Among all subsets satisfying , , and , let us denote by the subset that contains the greatest number of cycles in . If more than one such subset exists, we choose among them the one(s) containing cycles with the smallest index values.
Let be the set of all partitions such that the corresponding set contains at least one cycle that passes through both x and .
For each , define as the set of indices of the paths that are common to both and . That is, records which indexed paths contribute edges to the subgraph induced by .
For every nonempty subset
, define the set
Note that for all
, with
, we have
, so these sets form a disjoint partition of
.
Therefore, we obtain the identity
To compute the cardinality of , note that in every partition , all vertices in the union lie in the same block. Therefore, is in bijection with the set of partitions of into blocks such that the associated graph contains no cycle passing through the collapsed vertex .
Let denote the number of vertices in the graph . The corresponding partitions lie at level in the lattice .
To interrupt each of the cycles in that pass through , we proceed as described above: Construct a minimal cut by removing exactly one edge from each cycle.
Thus, by the Inclusion–Exclusion principle, we obtain
Let be the set of partitions for which the only path in connecting x and is the edge e.
We claim that
if and only if for every
,
where
denotes the set of atoms used in the decomposition of
that belong to
.
Indeed, suppose for some . Then, , and so contains a path from x to different from e, contradicting the assumption that .
Similarly, if for some , then all but one edge of are present in , and a redundant edge could be added to complete a cycle passing through both x and , again contradicting .
However, if for all , then even though redundant edges might exist, completing a cycle would change the block structure of , yielding a different partition. Thus, such must lie in .
To compute the cardinality of , we proceed by compressing the edge e and analyzing the resulting graph , where e is replaced by a single vertex . In this graph, we must interrupt every cycle passing through twice in order to prevent alternate connections between x and .
Let
be a minimal set constructed by selecting exactly two edges from each such cycle in
. Then,
where
is the vertex set of the compressed graph and
denotes the set of atoms corresponding to the edges in
.
Applying the Inclusion–Exclusion principle, we obtain
Therefore, since , the computations above yield the total cardinality of .
Finally, to compute the cardinality of , the set of partitions in which x and lie in different blocks, we must interrupt all paths connecting x and .
Let
be a minimal cut as defined earlier, obtained by removing exactly one edge from each such path. Then, using the Inclusion–Exclusion principle again, we obtain
This completes the proof. □