Cohomology ring of tree braid groups and exterior face rings

By Jesús González and Teresa Hoekstra-Mendoza

Abstract

For a tree and a positive integer , let denote the -strand braid group on . We use discrete Morse theory techniques to show that the cohomology ring is encoded by an explicit abstract simplicial complex that measures -local interactions among essential vertices of . We show that, in many cases (for instance when is a binary tree), is the exterior face ring determined by .

1. Main results

For a finite graph and a positive integer , let denote the configuration space of ordered points on ,

The usual right action of the -symmetric group on is given by , , and stands for the corresponding orbit space, the configuration space of unlabelled points on . Both and are known to be aspherical Reference 1Reference 10; their corresponding fundamental groups are denoted by (the pure -braid group on ) and (the full -braid group or, simply, the -braid group on ). We focus on the case of a tree .

Besides its central role in geometric group theory, graph braid groups have applications in areas outside pure mathematics such as robotics, topological quantum computing and data science. Yet, there is a relatively limited knowledge of the algebraic topology properties of a graph braid group (or, for that matter, of a tree braid group), particularly concerning its cohomology ring structure.

Using discrete Morse theory techniques on Abrams’ cubical model for (reviewed below), D. Farley gave in Reference 4 an efficient description of the additive structure of the cohomology of . Later, and in order to get at the multiplicative structure, the Morse theoretic methods were replaced in Reference 5 by the use of a Salvetti complex obtained by identifying opposite faces of cells in . Being a union of tori, has a well understood cohomology ring. Yet more importantly, the projection map induces a surjection in cohomology. Farley’s main result in Reference 5 is a description of a set of generators for Ker, which yields a presentation for the cohomology ring of .

Although Reference 5 includes an algorithm for performing computations mod Ker, the price of not working at the Morse theoretic level is that Farley’s presentation includes many non-essential generators. As a result, calculations are hard to work with, in concrete examples, as well as in theoretical developments (cf. Remark 1.9). In particular, Farley-Sabalka’s conjecture Reference 7, Conjecture 5.7 that is an exterior face ring, suggested on the basis of extensive concrete calculations, was left open.

In this paper we combine Farley-Sabalka’s original Morse theoretic approach with Forman’s Morse-theoretic description of cup products to prove the integral version of Farley-Sabalka’s conjecture for a large family of trees. The statement in Theorem 1.1, which focuses on binary trees, i.e., on trees all of whose essential vertices have degree three, disproves Conjecture 5.17 in Reference 15 by exhibiting an infinite family of non-linear trees all of whose braid group cohomology rings are exterior face rings.

Theorem 1.1.

Assume is a binary tree. For a commutative ring with 1, the cohomology ring is the exterior face ring determined by a simplicial complex . Explicitly, is the quotient , where is the exterior graded -algebra generated by the vertex set of , and is the ideal generated by monomials corresponding to non-faces of .

As noted in Reference 7, p. 68, the isomorphism type of a complex as the one in Theorem 1.1 is well determined. We refer to as the -interaction complex of . A description of as an abstract simplicial complex is given in Definition 1.3. The explicit definition allows us, for instance, to easily deduce a concrete right-angled Artin group presentation for when is a linear binary tree (Example 1.6). This complements the inductive method in Reference 3 proving that linearity is a sufficient⁠Footnote1 condition for a tree to have right-angled Artin braid groups.

1

The condition is known to be necessary and sufficient.

The definition of applies for any tree and we show that the resulting combinatorial object encodes much of the ring structure of , whether is binary or not. Indeed, we generalize Theorem 1.1 in two directions. On the one hand, the ring-isomorphism assertion holds as long as is a tree with binary core (Theorem 6.4). Furthermore, we show that, for any tree , the vertices of can be thought of as giving an -basis of , while the cup-product-based rule sets a 1-1 correspondence between the family of ()-simplices of and an -basis of . More importantly, while cup squares are known to vanish in , certain (square-free) products are non-zero even when fails to be a face of (this can happen only if is not a tree with binary core). In any such case, we give a closed formula (Theorem 5.1) to write any such product as an -linear combination of basis elements, thus completing a full description of the cup-product structure in the cohomology of for any tree . Details are summarized in Theorem 1.7.

The techniques used in this work (discrete Morse theoretic approach to cup products) should be a valuable tool in understanding the algebraic topology properties of discrete models for other spaces, such as non-particle configuration spaces, as well as generalized (e.g., no--equal) configuration spaces.

Remark 1.2.

Ghrist’s pioneering work led to conjecture that any pure braid group on a graph would be a right-angled Artin group. In the case of full braid groups , Reference 13Reference 14 give two characterizations (one combinatorial and another cohomological) of the right-angled-Artin condition. For instance, for a tree, is a right-angled Artin group if and only if is the exterior face ring of a flag complex. Theorem 1.1 and its generalized version in Theorem 6.4 assert that, in the full braid group setting and for trees with binary core, Ghrist’s conjecture is true after removal of the flag requirement.

The description of the complex , as well as an explicit statement of Theorem 1.7, and a couple of explicit illustrations (Examples 1.5 and 1.6) of Theorem 1.1 require a few preparatory constructions. Unless otherwise noted, throughout the rest of the section stands for an arbitrary tree.

Fix once and for all a planar embedding together with a root (a vertex of degree 1) for . Order the vertices of as they are first encountered through the walk along the tree that (a) starts at the root vertex, which is assigned the ordinal 0, and that (b) takes the left-most branch at each intersection given by an essential vertex (turning around when reaching a vertex of degree 1). Vertices of will be denoted by the assigned non-negative integer. An edge of , say with endpoints and , will be denoted by the ordered pair , where . Furthermore, the ordering of vertices will be transferred to an ordering of edges by declaring that the ordinal of is . The resulting ordering of vertices and edges will be referred to as the -order.⁠Footnote2

2

This of course depends on the embedding and root chosen.

Let stand for the degree of a vertex of , so there are “directions” from . For a vertex  different from the root, the direction from that leads to the root is defined to be the -direction 0; -directions (if any) are then chosen following the positive orientation coming from the planar embedding. See Figure 1. For instance, if is not the root and the vertex incident to in -direction 0 is not essential (i.e. ), then . Likewise, if , then is the vertex incident to in -direction 1. It will be convenient to think of the only direction from the root vertex as 0-direction 1, in particular there is no -direction 0.

Fix essential vertices of T. The complement in of the set decomposes into components

where , , and for . The closure of each is a subtree of . is the component containing the root , while (for ) is the component whose closure contains and is located on the -direction . The set of “bounding” vertices of a component is defined to be the intersection of the closure of with . Note that for , however the root is not considered to be a bounding vertex of , just as no leave of (i.e., a vertex of degree 1 other than the root) is considered to be a bounding vertex of any .

Definition 1.3 (The -interaction complex of , ).
(a)

The vertex set of is the collection of all 4-tuples , where is a non-negative integer number, is an essential vertex of , and and are tuples of non-negative integer numbers satisfying the three conditions

, with

, where and

for at least one .

We stress that (i.e., the length of ) is one of the parameters determining the 4-tuple . For instance, if and , then and are two different elements in . The length of , on the other hand, is determined by and .

(b)

For with , , and so that , consider the components and and of as defined above. Then, for , the -local information of , denoted by , is defined by

and, for ,

Note that whenever .

(c)

The -interaction complex of is the abstract simplicial complex whose vertex set is and whose -simplices are given by families of vertices as in item (b) satisfying

for all and all relevant , and in such a way that, for every , Equation 2 is a strict inequality for at least one .

It is an easy arithmetic exercise (whose verification is left to the reader) to check that is indeed a simplicial complex.

Definition 1.3 is dictated by discrete Morse theoretic considerations—reviewed in later sections. Our choice for using angle brackets instead of parentheses for 4-tuples in will be justified later in the paper (Remark 6.2). More important at this point is to explain the role of as an object measuring “local interactions” between systems of “local informations” around essential vertices of . For starters, we refer to a vertex as a system of local informations around the essential vertex of . Indeed, as illustrated in Figure 2, we think of:

(i)

as the local information of in -direction 0,

(ii)

() as the local information of in -direction , and

(iii)

() as the local information of in -direction .

In these terms, Equation 1 gives a systematic way to spell out the information ingredients on a given family of systems of local informations. Likewise, item (c) in Definition 1.3 asserts that a family of systems of local informations around essential vertices of assembles a simplex of if, for each component of , the sum of the -local informations of vertices  bounding is suitably large, depending on and on the number of bounding vertices of .

Definition 1.4.

Let be a family of systems of local informations around essential vertices of . We say that interact strongly provided is a simplex of . We say that interact weakly provided Equation 2 holds for all relevant and but fails to be a simplex of —so that, in fact, Equation 2 is an equality for some and all . In all other cases, we say that do not interact.

Example 1.5.

Figure 3 shows three aspects of the smallest possible non-linear tree . The four essential vertices are labelled (following the -order) in the central picture. The fact that the 4-fold product

is a basis element follows from Theorem 1.1, as inspection in the picture on the right of Figure 3 reveals that the factors in Equation 3 interact strongly. Note that for each factor in Equation 3, and that the cases with a strict inequality in Equation 2 hold as required in the last clause of item (c) of Definition 1.3. Likewise, interaction analysis in the picture on the left exhibits the well known fact that is not flag (i.e., is not a right-angled Artin group): the three basis elements , and in have pairwise strong interactions (so their three double products are part of a basis of ), but the three basis elements do not interact (so their triple product vanishes).

Example 1.6.

Let be a binary tree whose essential vertices lie along a single embedded arc. Choosing the planar embedding shown in Figure 4, we see that has a right-angled Artin group presentation with generators , where is an essential vertex of and are non-negative integer numbers⁠Footnote3 satisfying and . In these terms, has a commutativity relation whenever and , where the former inequality refers to the -order resulting from the embedding. Note that the chosen planar embedding of rules out weak interactions.

3

Instead of writing the 1-tuples and , we have simply written and .

Theorem 1.7.

For any tree , any non-negative integer and any commutative ring with unit 1, there is a set-theoretic inclusion so that the faces of yield, via cup-product of their vertices, a graded basis of . For instance, the empty face corresponds to the unit . Furthermore, any product with vanishes (in particular cup-squares vanish), as do cup-products of non-interacting elements in .

The only piece of multiplicative information missing in Theorem 1.7, namely a description of cup-products of weak-interacting basis elements in , is fully addressed in Section 5 (see Theorem 5.1) through the concept of “interaction parameters” introduced in Section 4 (Definition 4.3).

Remark 1.8.

The only obstructions for realizing in Theorem 1.7 as the exterior face ring determined by are the non-vanishing products whose factors interact weakly. For trees with binary core, such weak-interacting non-trivial products are effectively ruled out in the final section of this paper (Theorem 6.4) by means of a suitable change of basis that adjusts the inclusion in Theorem 1.7.

Remark 1.9.

The results in this paper allow us to recover and generalize Scheirer’s main technical tool Reference 16, Lemma 3.6 for studying Farber’s topological complexity of . Extensions of Scheirer’s results will be the topic of a future publication.

In the rest of the paper we shall omit writing the coefficient ring in cohomology groups and associated (co)chain complexes.

2. Preliminaries

We start by collecting the ingredients and facts we need: cup-products in the cubical setting Reference 11Reference 12, reviewed in Section 2.1, Forman’s discrete Morse theory Reference 8Reference 9, reviewed in Subsection 2.2, and Farley-Sabalka’s gradient field on Abrams’ discrete model for (ordered and unordered) graph configuration spaces Reference 1Reference 2Reference 6Reference 13, reviewed in Subsection 2.3. This will set the notation we use in the rest of the paper.

2.1. Cup products in cubical sets

An elementary cube in is a cartesian product of intervals , where and . For simplicity, we write for a degenerate interval. We say that is an -cube if there are non-degenerate intervals among the cartesian factors of , say with . In such a case, the product orientation of is determined by (a) the orientation (from smaller to larger endpoints) of the non-degenerate intervals , and (b) the order , i.e., the order of factors in the cartesian product. Under these conditions, and for , set

Then, for a cubical set , i.e., a union of elementary cubes in , the boundary map in the oriented cubical chain complex is determined by

For instance, the oriented cubical boundary of the square can be depicted as

Example 2.1.

Let be a tree whose vertices and edges have been ordered as described in the previous section. Think of as cubical set. In fact, orient the edges of from the smaller to the larger endpoints and fix an orientation-preserving embedding of cubical sets, where elementary cubes in have product orientation. Thus, a vertex of becomes a 0-cube in , while an oriented edge in corresponds in to an oriented 1-cube , i.e., an elementary cube whose factors are degenerate, except for one of them.

Cup products in cubical cohomology are fairly similar to their classic simplicial counterparts. At the oriented cubical cochain level, there is a cup product graded map that is associative, -bilinear and is described on basis elements as follows. Firstly, for intervals and , let

Then, for elementary cubes and in , the cubical cup product of the corresponding basis elements⁠Footnote4 vanishes if either for some or if is not a cube in ; otherwise is up to a sign , the dual of the cube . Given our product-orientation settings, the sign is given by the usual algebraic-topology convention:

4

We shall omit the use of an asterisk for dual elements. The intended meaning will be clear from the context.

Remark 2.2.

Particularly agreeable is the fact that a finite cartesian product of cubical sets comes equipped for free with the obvious structure of a cubical set. For instance, in the situation of Example 2.1, the cartesian power is a (product-oriented) cubical set in . In such a setting, an oriented cube in (where each is either a vertex or an edge of ) corresponds in to an oriented cube where, for each , at most one of the intervals is non-degenerate. These considerations, coupled with the fact that cubes of a single factor are at most one-dimensional, yield the next explicit description of cubical cup-products associated to and .

Proposition 2.3.

The cup product in of the duals of a pair of (oriented) cubes and in is given by the dual of

More generally, let be a (product-oriented) cubical subset of . The cup product in of the duals of a pair of cubes and in vanishes provided for some or, else, provided the cube is not contained in . Otherwise, the cup product is the multiple of the dual of , where

2.2. Discrete Morse theory

Let denote a finite regular cell complex with face poset ), i.e., is the set of (closed) cells of partially ordered by inclusion. For a cell , we write to indicate that is -dimensional. We think of the Hasse diagram of as a directed graph: it has vertex set , while directed edges (called also “arrows”) are given by the family of ordered pairs with . Such an arrow will be denoted as . Let be a partial matching on , i.e., a directed subgraph of whose vertices have degree precisely 1. The modified Hasse diagram is the directed graph obtained from by reversing all arrows of . A reversed edge is denoted as , in which case is said to be -collapsible and is said to be -redundant.

Discrete Morse theory focuses on gradient paths, i.e., directed paths in given by an alternate chain of up-going and down-going arrows,

A gradient path as the one on the left (right) hand-side of Equation 6 is called an upper (respectively, lower) path, and the gradient path is called elementary when , or constant when . The sets of upper and lower paths that start on a -cell and end on a -cell are denoted by and , respectively. Concatenation of upper/lower paths and is defined in the obvious way; for instance, any upper/lower path is a concatenation of corresponding elementary paths. An upper/lower path is called a cycle if in the upper case of Equation 6, or in the lower case. (By construction, the cycle condition can only hold with .) The matching is said to be a gradient field on  if has no cycles. In such a case, cells of that are neither -redundant nor -collapsible are said to be -critical or, simply, critical when is clear from the context. We follow Forman’s convention to use capital letters to denote critical cells.

It is well known that a gradient field on carries all the homotopy information of . For our purposes, we only need to recall how gradient paths recover (co)homological information. In the rest of the section we assume is a gradient field on .

Start by fixing an orientation on each cell of and, for cells , consider the incidence number of and , i.e., the coefficient (, since is regular) of in the expression of . Here is the boundary operator in the cellular chain complex . The Morse cochain complex is then defined to be the graded -free⁠Footnote5 module generated in dimension by the duals⁠Footnote6 of the oriented critical cells of . The definition of the Morse coboundary map in requires the concept of multiplicity of upper/lower paths. In the elementary case, multiplicity is given by

5

Cochain coefficients are taken in a ground ring , as we are interested in cup-products.

6

Recall we omit the use of an asterisk for dual elements.

and, in the general case, it is defined to be a multiplicative function with respect to concatenation of elementary paths. The Morse coboundary is then defined by

In other words, the Morse theoretic incidence number of and is given by the number of “mixed” gradient paths from to given as the concatenation of an arrow and a path , counted with multiplicity .

Gradient paths yield, in addition, a homotopy equivalence between and the usual cellular cochain complex . Indeed, the formulae

define (on generators) cochain maps and inducing cohomology isomorphisms and with .

2.3. Abrams discrete model and Farley-Sabalka’s gradient field

For a tree , think of as the cubical set described in Remark 2.2. Abrams discrete model for is the largest cubical subset of inside . In other words, is obtained by removing open cubes from whose closures intersect the fat diagonal. As usual, the symmetric group acts on the right of by permuting factors. The action permutes in fact cubes, and the quotient complex is denoted by . Following Farley-Sabalka’s lead, from now on we use the notation , and even , for a cube in (so each is either a vertex or an edge of ), and the notation , and even , for the corresponding -orbit. Beware not to confuse the parenthesis notation with a point of , or the braces notation with a set of elements of —even if all the ’s are vertices. The “coordinates” in a cube or in its -orbit are referred to as the ingredients of the cube. Closures of ingredients of cubes in and are therefore pairwise disjoint.

In his Ph.D. thesis, Abrams showed that is a -equivariant strong deformation retract of provided is -sufficiently subdivided in the sense that each path in between distinct vertices of degree not equal to 2 passes through at least edges. Such a condition will be in force throughout the paper, although it is not a real restriction because can be subdivided as needed without altering the homeomorphism type of its configuration spaces. The -equivariance of the strong deformation retraction above implies that is a strong deformation retract of . Consequently, we will switch attention from and to their homotopy equivalent discrete models and .

For a vertex of different from the root , let be the unique edge of of the form —recall this requires . Let be a cube either in or . A vertex-ingredient of is said to be blocked in if or, else, if replacing in the ingredient by the edge fails to render a cube in the corresponding discrete model; is said to be unblocked in otherwise. An edge-ingredient of a cube is said to be order-disrespectful in provided is of the form and there is a vertex ingredient in with and adjacent to (in particular must be an essential vertex); is said to be order-respecting in otherwise. Blocked vertex-ingredients and order-disrespectful edge ingredients in are said to be critical. Farley-Sabalka’s gradient field (on and ) then works as follows. Order the ingredients of a cube by their -ordering (as described in Section 1), and look for non-critical ingredients:

(i)

If the first such ingredient is an unblocked vertex in , then is redundant, and one sets , where is the cube obtained from by replacing by . We say that the pairing creates the edge . In this case is an order-respecting edge in , and all ingredients of smaller than are critical.

(ii)

If the first such ingredient is an order-respecting edge in , then is collapsible, and one sets , where is the cube obtained from by replacing by . Again, we say that the edge is created by the pairing . In this case is an unblocked vertex in , and all ingredients of smaller than are critical.

(iii)

If all ingredients of are critical, then is critical.

Definition 2.4.

For a vertex and a non-negative integer , let stand for the family of vertices . We think of as a size- stack of vertices supported by . Whenever we use such a stack of vertices, the -sufficiently subdivided condition on will assure the existence of the required vertices. Furthermore, for , let denote the vertex adjacent to that lies in -direction . For instance and , if is essential.

As illustrated in Figures 5 and 6, ingredients of a critical -cube are spelled out through

(a)

a stack of vertices supported by the root (here , i.e., can be empty);

(b)

pairwise different essential vertices of and, for each , an order-disrespectful edge with ;

(c)

for each and each , a stack of vertices supported by the vertex

subject to the requirements

(d)

some stacks might be empty, i.e., for all and . Yet, for each , there must exist an with (recall that is order-disrespectful);

(e)

, i.e., the total number of ingredients is .

The critical cube in the unordered discrete model determined by the above information will be denoted as

where and . Vertical bars are meant to stress the fact that each pair of parameters and is ordered and attached to . Other than that, Equation 10 is indeed a set formed by the triples and the singleton . Figure 6 illustrates a typical critical cube.

Remark 2.5.

In any arrow of Farley-Sabalka’s modified Hasse diagram, is an even face of , i.e., in the notation of Equation 4, for some .

Remark 2.6.

By construction, Farley-Sabalka’s gradient field in is -equivariant and, by passing to the quotient, it yields the corresponding gradient field in . Consequently, gradient paths can equivalently be analyzed in either the ordered or unordered settings. Indeed, a gradient path in corresponds to a -orbit” of gradient paths in . Due to the cup-product descriptions in Subsection 2.1, we find it more convenient to perform the gradient-path analysis at the level of the cubical set .

3. Gradient-path dynamics

Recall from Subsection 2.1 that the product orientation of a -dimensional cube in depends on (the orientation of edges—from the smaller to the larger vertex—in and on) the coordinate order , i.e. where , of the edge-ingredients. In particular, the quotient cube in inherits no well defined orientation. Definition 3.1 avoids the problem and is well suited for the analysis of gradient paths in .

Definition 3.1 (Gradient orientation, cf. Subsection 2.3 of Reference 5).

The listing , , of edge-ingredients of a -cube in or in is said to be in gradient order if , where the latter is the -ordering of vertices discussed in Section 1. The gradient orientation of is defined just as the product orientation, except that the gradient order of the edge-ingredients is used (rather than the coordinate order).

In the rest of the paper, and unless explicitly noted otherwise, we use gradient orientations. In doing so, the definitions of the cubes and in Equation 4 require a corresponding adjustment. Namely, if the edge-ingredients of a -cube are listed in gradient order as , then replacing the edge by the vertex or yields or , respectively. Remark 2.5 and the expression in Equation 5 for cubical boundaries then remain unaltered. A first advantage of gradient orientations is that the map induced at the cochain level by the projection involves no signs,

(Recall we omit asterisks for duals.) In view of Remark 2.6, the homotopy equivalences in Equation 9 satisfy:

Lemma 3.2.

The following diagram is commutative:

Remark 3.3.

The Morse differential in is trivial (see Reference 4 or Proposition 3.10). Therefore, for each , a graded basis of is given by the cohomology classes of the -images of the duals of the critical cubes Equation 10. By abuse of notation,⁠Footnote7 the -image⁠Footnote8 of the cohomology class so determined will also be denoted by the corresponding expression Equation 10. There is no loss of information because vertical maps in the previous diagram are injective and, more importantly, they induce injections in cohomology (the latter assertion follows from a standard transfer argument and the torsion-freeness of ).

7

The context clarifies the meaning.

8

We prefer to compute products in the ordered setting in view of the explicit descriptions in Subsection 2.1.

This section’s goal is the description of a cocycle in that represents a given cohomology class (Proposition 3.9). This requires the following discussion of dynamics for upper-paths that end at critical cubes.

Definition 3.4.

An edge-ingredient of a cube of is said to be

edge order-respecting in , written as is ”, if there are no edge-ingredients in with .

strongly order-respecting in , written as is ”, if is and there is no vertex-ingredient in with .

A Farley-Sabalka pairing that creates an edge-ingredient that is is said to be of sor type; otherwise, it is said to be of branch type. Likewise, is said to be of eor type if the edge-ingredient it creates is . An upper elementary path is said to be of falling-vertex type (sor type, branch type, respectively) provided ( is of sor type, is of branch type, respectively).

Note that if is the vertex-ingredient in that is responsible for a pairing , say creating the edge-ingredient of , then is obtained from by replacing the vertex by . In other words, in the falling-vertex type path , the vertex-ingredient “falls” to its predecessor . In particular, elementary paths of falling-vertex type have multiplicity 1.

Example 3.5.

Any edge-ingredient of is . On the other hand, for an essential vertex and a positive direction from , an edge-ingredient of is if and only if has no ingredient, neither vertex nor edge, in any of the components of lying in -directions . Furthermore, if is an edge-ingredient of a face of some cube of , then is if and only if is .

The final observation in Example 3.5 is freely used in the proof of:

Proposition 3.6.

Let be the gradient-order listing of the edge-ingredients of a -cube in .

(1)

If an arrow in the modified Hasse diagram for is of eor type, then is and, for any , is collapsible.

(2)

If the edge is , then there is no upper path starting at a face with and ending at a critical cube.

Proof.

(1) By definition, creates the edge-ingredient , which is assumed to be . Since ingredients of smaller than are critical, is in fact . Thus, for , is and, therefore, order-respecting in . On the other hand, for , and have the same ingredients smaller than , so that all ingredients in smaller than are critical. Thus, by definition, is collapsible for .

(2) Under the stated hypothesis, assume (for a contradiction) there is a gradient path

with , and critical. Then is and, in particular, is order-respecting in , which forces . Recursively, if is an edge-ingredient of both and (and so of ), and is , then is forced to be ( and, thus,) . It is not possible that is an edge-ingredient of all the ’s, for then would be , which is impossible as is critical. Let be the first integer () for which is not an ingredient of —so that is for . In particular, is order-respecting in . Thus, the vertex-ingredient of responsible for the pairing in Equation 12 satisfies and, in fact, , since is . On the other hand, since the edge created by is order-respecting in , and since is obtained from by replacing the edge by either or , the inequalities yield that

In particular, is not critical, so . Let be the vertex-ingredient of responsible for the pairing . By Equation 13, we get the first inequality in , so

( is an ingredient of ) ( is an ingredient of and, therefore, of );

( is unblocked in ) ( is unblocked in and, therefore, in ).

But, by definition, is the minimal unblocked vertex in , so , a contradiction.

Proposition 3.6 implies that upper paths ending at critical cubes have a forced behavior most of the time:

Corollary 3.7.

Let be an upper path in that ends at a critical cube. Any upper elementary factor of of sor type is of falling-vertex type.

Example 3.8.

Let us be specific about the dynamics of an upper path that ends at a critical 1-cube . By the -equivariance of the gradient field, we can assume with and

i.e., is the -orbit representative whose ingredients appear in the -ordering. By Corollary 3.7, the start of is forced to consist of falling-vertex elementary paths, where the vertices fall, each at a time, until they form the stack of vertices supported (and blocked) by the root. At that point arrives at the 1-cube and we see that must be positive, for otherwise would have reached a collapsible -cube. In particular must be an essential vertex and . Then, again by Corollary 3.7, it is the turn of vertices that are forced to fall, each at a time, until they form stacks of vertices blocked by in -directions . At that point arrives at a 1-cube of the form

Not all of the stacks are empty, so Equation 14 has as a critical edge-ingredient. The falling-vertex process is also forced by Corollary 3.7 on those vertices that are located in positive -directions (if any), and this takes to a 1-cube of the form

with all lying in -direction 0. Branching starts from this point on, with explicit options discussed in the next paragraph.

If no vertices are left, then would have reached its final critical destination . Otherwise, is forced to fall until reaches, via some branch type pairing , the 2-cube depicted in Figure 7. At this point there are two options for . In the first option, is obtained from by replacing the recently created edge by , i.e., with an upper elementary path of falling-vertex type. In such a case, is forced to continue with the vertex falling until it is added to the stack of vertices blocked by the root 0. This leaves us at a situation similar to the one at the start of this paragraph. In the second option, is obtained from by replacing the edge by either of its end points. In such a case, is forced to continue:

(1)

with the falling of the vertices that are now unblocked in the neighborhood of (see Figure 7), until they form a stack of vertices blocked by —thus starting a critical situation around the edge —and, then,

(2)

with the falling of the vertices (if any) in -directions from to , which form (possibly empty) stacks of vertices blocked either by or —thus completing the critical situation around the edge .

Again, this leaves us at a situation similar to the one at the start of this paragraph, but now with the edge playing the role of the edge . The branching process in this paragraph then repeats, necessarily a finite number of times, until all vertices have been considered, when reaches its critical destination .

Proposition 3.9.

A cocycle in representing a 1-dimensional cohomology class in , with and , is given by

where the summation runs over

all permutations ,

all possible vertices in the component of in -direction 0,

all possible vertices in the components of in -directions from to so that, for , of the vertices lie in -direction ,

all possible vertices in the components of in -directions greater than so that, for , of the vertices lie in -direction .

Proof.

By construction, the representing cocycle we need is obtained by chasing, on the left square of the diagram in Lemma 3.2, the dual of the unordered critical cube whose ordered critical representative is

By Equation 9 and Equation 11,

where is the set of upper paths that start at a 1-cube and finish at a 1-cube of the form with . Let be the set of paths all of whose upper elementary factors are of falling-vertex type. Since for , the analysis in Example 3.8 shows that the summands in Equation 15 arise from the summands in Equation 16 with . It thus suffices to show

which will be done by constructing an involution such that every pair of paths and has the same origin but opposite multiplicities, i.e.,

—thus their contributions to Equation 17 cancel each other out. For a path , let denote the last elementary factor of that is not of falling-vertex type. In the notation of Example 3.8, is obtained from by replacing an edge by either or , and both options are possible. Then is defined so to start with the same factorization of into elementary paths, except for the elementary factor , for which the other end-point of is taken, and after which the rest of the elementary factors are of falling-vertex type—just like for . Note that the ending 1-cubes of and lie in the same -orbit, so . The required properties Equation 18 follow from (the construction and from) the fact that elementary paths of falling-vertex type have multiplicity 1.

The cancelation phenomenon in the previous proof allows us to give an easy gradient-path explanation of the main result in Reference 4: the vanishing of the Morse differential in . A variant of the cancellation phenomenon will also play an important role in our evaluation of cup products (Theorem 5.1). Thus, in preparation for that argument, we spell out the gradient proof of:

Proposition 3.10.

The Morse differential in vanishes.

Proof.

By Remark 2.6, it suffices to do the gradient path analysis directly at the level of . For a pair of unordered critical cubes and , let be the set of mixed gradient paths . By Equation 8, we only need to construct an involution so that, for every , . (Recall that the multiplicity of is the incidence number for multiplied by the multiplicity of the remaining upper path .) Let consist of the paths in all of whose upper elementary factors are of falling-vertex type. The definition of the restricted uses the two forms of replacing by a vertex the edge-ingredient at the start of the path. Likewise, for , the definition of the restricted uses the two forms of replacing by a vertex the edge ingredient at the last upper elementary factor that is not of falling-vertex type.

Propositions 2.3 and 3.9 immediately yield:

Corollary 3.11.

The product of two basis elements vanishes provided . In particular, squares of 1-dimensional elements in are trivial.

4. Cup products I: Upper gradient paths

The goal for this section and the next one is to get at a workable description of products

in . Associated to such a product, from now on we set , , , , and make free use of (i) the order-disrespectful edge encoded in the -th factor of Equation 19, of (ii) the conditions , and , and of (iii) the fact that, for each , and all of the and are non-negative, with not all of the being zero. Additionally, in view of Corollary 3.11, we can safely assume . Last, we use the shorthand

We start by tuning up the definition in Section 1 of the components of .

Definition 4.1 (Leaves and pruned trees).

Set and, for and ,

where stands for the interior of the edge . We think of each as a rooted but possibly pruned tree. Namely, in the notation of Section 1 and setting , the root of is , if or if with , whereas the root of is . Furthermore, the set of pruned leaves of is .

Remark 4.2.

Just as the sets give a partition of , the union of the trees agrees with the difference . Actually, each vertex of other than for , as well as each semi-open edge of not of the form with , belongs to a tree  for a unique .

Definition 1.4 is recast by the second part of:

Definition 4.3.
(1)

For a -tuple of integers , we write to mean that for all , reserving the expression to mean that with for at least one . Also, when , we write to denote a generic tuple of integers satisfying for all with in fact for at least one . We make no distinction between 1-tuples and integer numbers so, accordingly, we use instead of .

(2)

The interaction parameters , and of the factors in Equation 19 are given by

If , and for all , we say that the factors in Equation 19 interact weakly and, if in addition for some , we say that the factors in Equation 19 interact strongly. Otherwise, we say that the factors in Equation 19 do not interact.

Although not reflected in the notation, pruned trees and leaves depend on the essential vertices , while interaction parameters depend on the complete information encoded by the factors in Equation 19. Latter in the paper we will need to use pruned trees, their pruned leaves, as well as interaction parameters of subproducts of Equation 19. In such a case, we will use a notation of the type , , , , , as well as and in order to clarify the factors under consideration.

Next we adapt the expression in Equation 15 for usage within the -notation. In terms of the cocycle representative

in Proposition 3.9 for , Equation 19 is represented by the sum of all possible products

A number of vanishing such products can be ruled out as follows. Fix integers . Proposition 2.3 implies that if a product Equation 21 is non-zero, then must have , but cannot have , as one of its vertex ingredients. Likewise, must have , but cannot have , as one of its vertex ingredients. Actually, together with Remark 4.2, this shows that non-zero products Equation 21 are best organized (and easily evaluated—see below) by replacing each -representative

in Equation 20 by the one written in a “block” form . Here each tuple of ingredients starts with the relevant - or -information (if ), and continues with a repacking of the vertex ingredients of Equation 22 that lie in the trees for all relevant . In detail, for the -th factor in Equation 19 and each of the corresponding summands in Equation 22, let

(a)

be the tuple of vertex ingredients of Equation 22 that lie in , written in -order;

(b)

, where is the tuple of vertex ingredients of Equation 22 that lie in , written in -order;

(c)

If , , where is the tuple of vertex ingredients of Equation 22 that lie in , written in -order;

(d)

If , , where is the tuple of vertex ingredients of Equation 22 that lie in , written in -order.

Thus, summands in Equation 20 that have a chance to contribute with non-vanishing products Equation 21 to a cocycle representative of Equation 19 can be written as

where vertical bars are used interchangeably by commas, and are intended to make reading easier. Proposition 2.3 then implies that a product Equation 21, written as

is non-zero if and only if and for all relevant , in which case Equation 21 becomes

where is the permutation determined by the sequence of positions of the edges in the tuple

Note that the cube in Equation 23 is product-oriented (as required by Proposition 2.3), and that Equation 23 agrees with the gradient-oriented cube

since . This proves the first half of the next generalization of Proposition 3.9:

Proposition 4.4.

The product Equation 19 is represented in by the gradient-oriented cocycle

where the summation runs over all permutations and all possible tuples of vertices written in -order, taken from the corresponding pruned trees , and having the following lengths: Any block must have ingredients, while any block with must have ingredients for , and ingredients for . In particular, Equation 19 vanishes provided its factors do not interact.

Note that in Definition 4.3. This is compatible with the fact that cubes in Equation 24, if any, have ingredients. See also Corollary 4.5.

Proof.

It remains to prove the assertions about the sizes of blocks , and that all possible such blocks appear in Equation 24. As for the sizes, proceeding by induction on (with Proposition 3.9 grounding the argument), it suffices to consider a product with

where , and where the structure of the blocks is as specified in the proposition. Here we are assuming (a) that is a tuple of vertex ingredients written in -order and lying in -direction 0, (b) that any tuple with consists of vertex ingredients written in -order and lying in -direction , and (c) that any tuple with consists of vertex ingredients written in -order and lying in -direction . In addition, we make the conventions and , and assume the relations , and . Furthermore, signs and orientations will be ignored in the rest of the proof, as they have been carefully addressed in the discussion previous to this proposition. In particular, we can safely work at the unordered-cube level, thus ignoring the permutations and in Equation 25 and, instead, thinking of tuples of ingredients as sets of ingredients.

Consider the pruned trees and , as well as the pruned leaves and . There are three cases, depending on whether the edge belongs to , or to with and , or to with and , and the argument is virtually identical in each. We consider only the situation depicted in Figure 8, where the edge belongs to for some and some . In such a case we have

(i)

and , for as long as or ;

(ii)

;

(iii)

and for .

By Proposition 2.3, the product of the elements in Equation 25 vanishes unless

and

in which case

where is a shorthand for the sequence provided , whereas stands for the sequence

The induction is complete in view of items (i)–(iii) above and

which shows that has the prescribed cardinality. The inductive analysis makes it clear also that all blocks with the structure indicated in the proposition indeed appear in Equation 24.

Corollary 4.5.

The product Equation 19 agrees with the basis element provided the factors of Equation 19 interact strongly. Recall and .

Proof.

By the strong interaction hypothesis, a summand in Equation 24 that is the target of a lower gradient path must actually be critical (and must be constant) with ingredients equal to those associated to . The conclusion then follows from Equation 9 and Equation 11.

Lemma 4.6.

Fix essential vertices and take positive integer numbers and with for . Let , , , with , and , be non-negative integers satisfying . Then the system

has a unique solution of non-negative integer numbers satisfying the condition for each . If, in addition, for each there exists with , then the unique solution satisfies that, for each , there exists with .

Proof.

The two sets of equations with reduce to () and (). This also determines

The rest of the equations can be written as

for , and . The result then follows by induction since

where the second equality uses Equation 26.

Proof of Theorem 1.7.

Corollary 4.5 and Lemma 4.6 yield a set theoretic identification , where is the set of products Equation 19 whose factors interact strongly, and is the -dimensional basis of with basis elements

Together with Corollary 3.11 and Proposition 4.4, this completes the proof, where is identified with (the -preimage of) .

Note that the cohomology ring is generated by 1-dimensional classes, a fact already known from Reference 7. It is not true that a product Equation 19 vanishes when its factors interact but non-strongly. The description of such products relies on the dynamics of lower gradient paths.

5. Cup products II: Lower gradient paths

Let stand for a product Equation 19 whose factors interact strongly, so Corollary 4.5 applies. Choose an additional 1-dimensional basis element , of with and where the standard conditions and conventions are assumed, namely,

where , , , , and . Consider the interaction parameters and of the factors of (), as well as the first three interaction parameters , and of the factors of . This section is devoted to proving:

Theorem 5.1.

In the situation above, if the factors of interact but non-strongly, then

In the above expression we set , , , and . The summation in Equation 28 runs over all -tuples of non-negative integer numbers satisfying . The inner summation in Equation 29 runs over all -tuples of non-negative integer numbers and all non-negative integer numbers  satisfying . The inner summation in Equation 30 is empty if , otherwise it runs over all -tuples of non-negative integer numbers and all non-negative integer numbers satisfying .

Since summands in Equation 28, Equation 29, and Equation 30 are basis elements, Theorem 5.1 and the results in the previous section give a recursive method to effectively assess cup-products in .

Proof of Theorem 5.1 (Preparation).

We have seen that is represented in by the gradient-oriented cocycle

where the summation runs over all permutations and over all possible blocks of vertices written in -order, taken from the corresponding trees determined by the factors of , and having sizes as prescribed in Proposition 4.4 in terms of the relevant interaction parameters. The goal now is to identify the -image of Equation 31 which, by Equation 9, is the element in

Here is the set of lower paths starting at an -critical cube and finishing at a summand of Equation 31. We start by identifying (in the next two paragraphs) key characteristics of ending cubes for paths in .

Firstly, the condition forces one of the four configurations depicted in Figure 9. In any of those configurations, vertices with lie either on a component of in positive -direction or “below” the horizontal segment joining the root and . As a result, the equalities and hold for . The interaction hypotheses then yield

the -tuple consisting of zeros. This and Equation 27 rule out the two configurations on the right of Figure 9, as well as the one on the bottom left, since the equality is forced for those configurations. The only possible configuration, i.e., the one on the top left of Figure 9, will be assumed in the rest of the section.

Secondly, redundant summands in Equation 31 can be neglected, as none of those can be the destination of a lower path. Furthermore, Equation 33 shows that no summand in Equation 31 is critical. We thus focus on collapsible summands in Equation 31 which (in addition to their size and distribution properties summarized at the start of the proof) are forced to satisfy the following two properties: For one, ingredients of each that are smaller than  form a stack of vertices blocked by the root of . In addition, for any with smaller than , all ingredients of each () are blocked (this uses the fact that is not the zero tuple), so the tuple assembles a (unique, by block-size limitations) critical situation around . It follows that each summand in Equation 31 relevant for Equation 32 is collapsible by a branch-type pairing that creates the edge , as depicted in

Note that any -cube that has been identified on the right of Equation 34 as a potential destination of a path supports a gradient path with a critical -cube. For instance, start by replacing the edge in by , and let the rest of the path consist of falling-vertex elementary factors. It follows that the concatenation of and and, therefore, itself obey the rule in Corollary 3.7: any upper elementary factor of sor type is of falling-vertex type. Such a fact, together with cancellation phenomena similar to the one in the proof of Proposition 3.9, is used in the rest of the argument in order to analyze paths determining Equation 32. As in the proof of Proposition 3.10, the analysis can equivalently be done at the level of , which means that an ordered cube can be replaced by the corresponding orbit . Following the lead in Proposition 3.9, we first identify the actual sets of paths whose contributions in Equation 32 give Equation 28 Equation 29, and Equation 30.

The summation in Equation 28 arises from a set of paths having a single “lock” dynamics. Explicitly, each -tuple of non-negative integer numbers satisfying determines a lower gradient path that departs from the critical -cube

by replacing the edge by —this opens the lock. Then continues with the falling of the vertices that were blocked by , after which ends with the pairing that closes the lock by creating the edge required in Equation 34. Since both opening and closing locks are associated to the same face (the gradient-orientated -face), and since falling-vertex elementary paths have multiplicity 1, we see from Equation 7 that . Thus, yields Equation 28.

The set of paths is contained in a slightly larger subset which consists of paths , where runs over -tuples of non-negative integer numbers and runs over non-negative integers numbers satisfying and . Explicitly, starts by taking face (lock opening) of the critical -cube

Here and below we take the coordinate-wise sum of tuples. Then continues with the falling of the vertices that were blocked by , followed (if ) by the falling of the vertices , to finish with the falling of until it creates the required branch-type pairing Equation 34—which closes the lock. As in the case of  paths in have multiplicity . Likewise, there is the family consisting of paths , with and as above, except that the inequality is replaced by the strict inequality . Explicitly, starts by taking face (inverse lock opening) of the critical -cube

Then continues with the falling of , followed by the falling of the vertices that were blocked by , followed (if ) by the falling of the  vertices , to finish with the falling of until it creates the required branch-type pairing Equation 34—which closes the lock. Note that paths in have multiplicity .

Figure 10 summarizes dynamics of paths in (top) and paths in (bottom), with lock opening/closing represented by arrows. Note the shifting on the vertices falling from -direction , as well as on the vertices that make up at the end of the path. The key point is that if , the paths and share origin, so their contributions in Equation 32 cancel each other out. The only unmatched paths are those in with parameter , i.e., paths in  whose contribution in Equation 32 has been shown to yield (Equation 28).

By construction, consists of those paths in that start by taking a face with of a critical -cube with edges

and that evolve exclusively though falling-vertex elementary paths before reaching the required pairing Equation 34. Next we describe similar sets of paths contributing in Equation 32 with Equation 29 and Equation 30. In such sets of paths, an edge

plays the role of the edge in .

Paths with (, in the notation of Equation 35): If , set , otherwise consists of paths , where runs over -tuples of non-negative integer numbers and runs over non-negative integer numbers satisfying . Explicitly, if , then starts by taking face of the critical -cube

and evolves through falling-vertex elementary paths as depicted by the chart⁠Footnote9

9

As in the case of , the vertices falling from -directions 1 through are not shown in the chart.

before reaching the required pairing Equation 34. Both opening and closing locks of are associated to a (gradient-oriented) face, so that . The contribution in Equation 32 of the paths in thus gives raise to Equation 30. Note that no path that starts from the origin of a given by taking face —instead of —and that evolves through falling-vertex elementary paths can arrive at a summand of Equation 31. This is why the contribution to Equation 32 of the set of paths in the next paragraph does not cancel out terms in Equation 30.

Paths with (, in the notation of Equation 35): consists of paths , where runs over -tuples of non-negative integer numbers and runs over non-negative integer numbers satisfying . Explicitly, if , then starts by taking face of the critical -cube

and evolves through falling-vertex elementary paths as depicted by the chart

before reaching the required pairing Equation 34. Now , so the contribution in Equation 32 of the paths in gives raise to Equation 29. Again, no path that starts from the origin of a given by taking face —instead of —and that evolves through falling-vertex elementary paths can arrive at a summand of Equation 31.

Remark 5.2.

Since the closing-lock pairing Equation 34 must come from -direction , paths corresponding to cases with in Equation 35 have no contribution in Equation 32. Specifically, any path that starts from a critical cell with edges , where , by taking a face with , and that reaches the pairing Equation 34 through falling-vertex elementary paths, has a companion path that starts from the same critical cell by taking the face , and that also evolves through falling-vertex elementary paths until it reaches the closing-lock pairing Equation 34—so that and . Note that, in the ordered setting, and its companion path arrive at summands of Equation 31 whose ingredients differ only by a permutation (so as well). The phenomenon noticed in this remark is in fact the key to finishing the proof of the main result in this section.

Proof of Theorem 5.1 (Conclusion). Let stand for the set of paths analyzed up to this point, i.e., the paths in that (I) depart from a critical -cube with gradient-ordered edges , (II) start by taking the face or and (III) reach the ending branch-type pairing Equation 34 exclusively through falling-vertex elementary paths. It suffices to construct an involution , with , such that each pair of paths and shares origin and has opposite multiplicity. With this in mind, we first note that condition (II) is forced by conditions (I) and (III). Indeed, in any gradient path all of whose upper elementary factors are of falling-vertex type,

Therefore is partitioned into two sets, and , where the former set consists of the paths in that satisfy (III) without satisfying (I), and the latter set consists of the paths in that do not satisfy (III). We construct involutions and with the required properties.

For a path in , the observation in Equation 36 and the form of the closing-lock pairing imply that all edges , , must be ingredients of . The additional edge of the critical -cube must then have the form , with , which is then replaced by either or at the beginning of . Given the form of , must lie in -direction . Then, as in the proof of Proposition 3.10, the definition of is based on the two options for , as both lead to summands of Equation 31—unlike the situation in Remark 5.2, the ending cube of might fail to be in the -orbit of the ending cube of . Likewise, the definition of is based on the two forms of replacing by a vertex the edge ingredient at the last upper elementary factor that is not of falling-vertex type.

6. Exterior-face basis for trees with binary core

We have made a careful distinction between and in the previous sections so as to provide clear proof arguments. In this section we use the resulting algebro-combinatorial description of cup-products and have no need to make any further distinction between these isomorphic rings. Accordingly, we transfer the notation and descriptions of elements in back to . In particular, the notation and conventions in the paragraph containing Equation 19 will be carried over this final section, directly in the context of , with the simplifications discussed below.

Definition 6.1.

A tree is said to have binary core provided that, for each essential vertex of , at most two of the components of in -directions carry essential vertices (recall ).

Throughout this section, stands for a tree with binary core (e.g. an actual binary tree). In addition, we assume that the chosen planar embedding of has been adjusted so that, for any essential vertex  of ,

There are two reasons for sticking to such a hypothesis. For one, the existence of non-vanishing products whose factors are given by weak-interacting basis elements

with , i.e., the obstructions in Remark 1.8, is somehow restricted (cf. Example 1.6), while our description of the corresponding product is greatly simplified. Explicitly, in the setting and notation of Theorem 5.1, since the top left configuration in Figure 9 holds, Equation 37 forces , i.e., the edge must lie in the largest -direction, with then lying in the second largest -direction . In particular, the product takes the simpler form

where the sum runs over all -tuples of integer numbers with and .

The second advantage for working under the situation in Equation 37 is that, for and , any set of pruned leaves associated to a product Equation 19 is empty. As a result, the corresponding -local interaction is “vacuous” in the sense that the -instance of Equation 2 simplifies to —a condition which is certainly true. In fact, still in the context of Equation 19, there will be no local interactions in the positive -directions leading to a weak interaction situation as long as for some (cf. Equation 33). In particular, it makes sense to reset the notation for pruned leaves in the presence of Equation 37: we shall set and when , and (recall from Definition 4.1 that stands for the root of ).

Expression Equation 38 suggests redefining some of the basis elements , in the proof of Theorem 1.7. Namely, for the purposes of this section, if and , we set

where the summation runs over all -tuples with , otherwise we keep

Remark 6.2.

We use the angle-bracket notation since we have reserved the parenthesis notation for cubes in (as tuples of their ingredients). Additionally, the angle-bracket notation is intended to stress the change of basis in Equation 39.

A central task in this section is the analysis of the relationship between ordered⁠Footnote10 products

10

In the sense that .

We say that any of these products is a strong interaction product if the factors of the product on the right hand-side of Equation 40 interact strongly (in the sense of Definition 4.3).

Remark 6.3.

Corollary 4.5, Proposition 4.4 and Theorem 5.1 show that both products in Equation 40 are (possibly empty) linear combinations of basis elements

Such a linear combination will be written as

Here and below, a dot  stands for either an unspecified ring coefficient or an unspecified tuple⁠Footnote11 of integer numbers, , satisfying when the tuple immediately follows an essential vertex  (the context clarifies the option).

11

As in Definition 4.3, we make no distinction between integer numbers and 1-tuples.

Theorem 6.4.

Let be a tree with binary core, be a commutative ring with 1, and . Then . In detail: (i) An ordered product is non-zero if and only if it is a strong interaction product. (ii) Two ordered strong interaction products agree if and only if they have the same factors. (iii) A graded basis of is given by the set of ordered strong interaction products.

The crux of the matter in the proof of Theorem 6.4 is getting at a precise description of the conditions that have to be satisfied by some of the unspecified dot ingredients in

With this in mind, the product in Equation 41 will be denoted by throughout the section, setting

and , , for the corresponding interaction parameters. Furthermore, we set

where the latter expression stands for any triple with unspecified tuples in the second and third coordinates (subject to the usual restrictions). Additionally, the -th factor on the left hand-side of Equation 41 will denoted by . For instance, in terms of the notation set forth in Definition 4.3,

with a possibly empty summation, whereas Corollary 4.5 asserts that the second product in Equation 40 is trivial or agrees with under, respectively, the no-interaction or strong-interaction condition of the factors.

In the following results, some of which are true for general trees, we make free use of the notation and considerations above. Likewise, the use of cup-product descriptions in Sections 4 and 5, with the simplification in Equation 38, we will refer to generically as “interaction reasons”.

Lemma 6.5.
(1)

Assume (left configuration in Figure 11), then

(2)

Assume and with (right configuration in Figure 11) with (i.e. and if , then

Proof.

The first assertion follows by direct inspection of the expression

noticing that the only non-vacuous interaction occurs in the tree (so that and for ). The second assertion is proved in a similar way, noticing that this time non-vacuous interactions occur only either on or (or both). In any case, , for , while for .

A key situation with not covered by Lemma 6.5(2) is:

Lemma 6.6.

Assume . Then the product of with vanishes provided and .

Proof.

We proceed by induction on , where

is the unique strong-interaction factorization of noted in the proof of Theorem 1.7. Since for , the induction is grounded for by

where both summations run over tuples with . The inductive step then follows by noticing that, for ,

as

vanishes for with by interaction reasons.

Corollary 6.7.

If the factors on the left of Equation 41 do not yield a strong interaction product, then .

Proof.

By focusing on the factors of that are involved in a faulty interaction parameter, it suffices to consider three cases: , and . The first two cases are covered by Lemma 6.5. On the other hand, there are two options for the instances of the third case that are not covered by Lemma 6.5(2): either for some or for all —in both cases . In the latter option, the result follows from Lemma 6.6; in the former option we have while the condition is forced by the no-strong-interaction hypothesis, so that the result follows by interaction reasons in view of Lemma 6.5(1).

The proof of Theorem 6.4 will be complete once we set a one-to-one correspondence between the set of ordered strong interaction products and the graded basis of formed by the elements in Equation 10. With this in mind, we start with a two-step approach to the missing case in Lemma 6.5(2):

Lemma 6.8.

Assume with . Then

Proof.

Interactions occur only in , so , for , and for . By Corollary 6.7, only the case needs to be argued. Use Lemma 6.5(1) to write as

where (so ). The result then follows by direct inspection, though this time Equation 38 needs to be used in the analysis of the products giving rise to the terms in both summations of Equation 43.

Proposition 6.9.

Assume and , with and . Then

Here and below each expression is meant to represent a pair of unspecified tuples of integer numbers with , and such that and in the product ordering, i.e., for and , with at least one of the last inequalities being strict.

Proof.

By Corollary 6.7, it suffices to consider the case . Lemmas 6.5(1) and 6.8 allow us to write as the product of

with

where (so ). The result follows by inspection.

We are now ready to set up the strategy for completing the proof of Theorem 6.4. By Lemma 4.6, Remark 6.3 and Corollary 6.7, the goal reduces to describing, for fixed essential vertices , a partial ordering on the set of basis elements of such that any strong interaction product Equation 41 can be expressed by a congruence

modulo basis elements that are -smaller than . The partial ordering we need becomes apparent by writing either of the triples , and in Proposition 6.9 and Lemmas 6.8 and 6.5(2), respectively, as . Indeed, in such terms, the ()-conclusions in those results can be written as

Definition 6.10.

The -th level of pruned leaves of the essential vertices is

The interaction level of the vertices is the largest such that . Furthermore, extending the notation introduced in Equation 42 and Equation 45, let denote the collection of blocks with , and let stand for any collection of blocks with . On the other hand, stands for any collection of blocks , with , satisfying:

and (the latter in the product ordering) for all , and

for at least one .

Note that the definition of is less restrictive than actually requiring to be a collection of blocks with . As in Proposition 6.9, the condition we want for is based on a strict product-order inequality. The reason for this becomes apparent in the proof of Proposition 6.12.

Example 6.11.

Lemma 6.5(1) gives in interaction level 1 (under a strong condition hypothesis). Likewise, Equation 45 becomes

in interaction level 2 (with , so consist of alone). In full generality:

Proposition 6.12.

Let be essential vertices having interaction level . If is a strong interaction product, then

Proof of Theorem 6.4 (Conclusion).

Partially order the set of basis elements

by means of a level-wise lexicographical comparison of their - and -ingredients. Then Equation 47 yields the required congruence Equation 44.

Proof of Proposition 6.12.

The argument is by direct computation, proceeding by induction on and with Example 6.11 grounding the induction. The real challenge consists of setting a suitable notation so arguments can be seen clearly. With this in mind, we start by checking the situation in the special case (so ), i.e., the generalization of Equation 46 to higher interaction levels. In such a situation

Accordingly, we reset notation and start level-number counting at 2 (rather than at 1) for , so as to make it compatible with that for . Thus, Equation 48 gets replaced by

Let be the essential vertices lying on the component of in -direction , while be the vertices lying on the component of in -direction (). Then, if , and stand for collections defined by all the vertices , we write

with , to denote the corresponding parts coming only from the vertices . Likewise, the case in Equation 50 stands for the parts that come from the vertices . For instance, . In these terms, we use induction to write as the product of the three expressions

and

where and . Note the compactified notation for the two summations running over , each of which really stands for sums of summations as in Equation 47. Note also that the interaction level of the vertices (or ) could be smaller than , in which case some of the corresponding collections of blocks are empty. Then, by direct inspection and interaction reasons (using Equation 38 when and the interaction parameter under consideration lies in -direction ), the product of the three expressions above takes the form Equation 47. This completes the proof when is a singleton.

In general, consists of, say, vertices , and we evaluate as the length- product

(This time there is no need to reset notation so as to get the analogue of Equation 49 to hold.) We have just seen that the -th factor in Equation 51 takes the form

where

and, for interaction levels larger than 1, a subindex ‘[]’ in a collection of blocks indicates that only blocks in positive -directions are to be taken. The required form Equation 47 for the product of all these expressions follows again from direct inspection—this time without requiring the use of Equation 38.

Figures

Figure 1.

The -directions from an essential vertex

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[ scale=.75, rotate = 180, xscale = -1] \tikzset{EdgeStyle/.style={->,font=\scriptsize,above,sloped,midway}} \node[style={circle, draw, scale=.6}] (1) at ( 2, 3.45) {$0$}; \node[style={circle, draw, scale=.6}] (2) at ( 7.4, 3.45) {$x$}; \node[style={circle, draw, scale=.6}] (3) at ( 9.58, 1.28) {}; \node[style={circle, draw, scale=.6}] (4) at ( 10.5, 2.72) {}; \node(5) at ( 10.25, 3.8) {}; \node[style={circle, draw, scale=.6}] (6) at ( 10.6, 4) {}; \node(7) at ( 8.25, 1.91) {}; \node(8) at ( 8.97, 2.6) {}; \node(9) at ( 9.1, 3.34) {}; \node(10) at ( 8.94, 4.04) {}; \node(11)[scale=.3] at (4, 3.45){}; \node(12)[scale=.3] at (6, 3.45){}; \node at (3.1, 3.24){\tiny$\;\;0$-direction 1}; \draw(4)[dotted] to[bend right] (6); \draw(1) -- (11); \draw(11)[dashed] -- (12); \draw(12) node[above]{\tiny\hspace{5mm}$x$-direction 0 \ }-- (2); \draw(2)--node[sloped,above,rotate=90]{\tiny$x$-direction 1}(3); \draw(4) --node[sloped,above,rotate=30]{\tiny$x$-direction 2} (2); \draw(6) --node[sloped, below, rotate=-20]{\tiny$x$-direction $d(x){-}1$} (2); \end{tikzpicture}
Figure 2.

The local information given by a vertex of

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[ scale=1.0, rotate = 180, xscale = -1] \tikzset{EdgeStyle/.style={->,font=\scriptsize,above,sloped,midway}} \node[style={circle, draw, scale=.6}] (1) at ( 2.8, 3.45) {$0$}; \node[style={circle, draw, scale=.6}] (2) at ( 7.4, 3.45) {$x$}; \node[style={circle, draw, scale=.6}] (3) at ( 9.58, 1.28) {}; \node[style={circle, draw, scale=.6}] (4) at ( 10.15, 2.75) {}; \node[style={circle, draw, scale=.6}] (0) at ( 10.15, 3.75) {}; \node(5) at ( 10.25, 3.8) {}; \node[style={circle, draw, scale=.6}] (6) at ( 9.82, 5.02) {}; \node(7) at ( 8.25, 1.91) {}; \node(8) at ( 8.97, 2.6) {}; \node(9) at ( 9.1, 3.34) {}; \node(10) at ( 8.94, 4.04) {}; \node(11)[scale=.3] at (4, 3.45){}; \node(12)[scale=.3] at (6, 3.45){}; \draw(0)--node[below]{\tiny$q_1\hspace{.7cm}$} (2); \draw(0)[dotted] to[bend right] (6); \draw(3)[dotted] to[bend right] (4); \draw(1) -- (11); \draw(11)[dashed] -- (12); \draw(12) node[above]{\tiny\hspace{1.3cm} $k$}-- (2); \draw(2)--node[sloped,above,rotate=90]{\tiny$p_1\hspace{1cm}$}(3); \draw(4) --node[sloped,above,rotate=30]{\tiny$p_{\red{r}}\hspace{.7cm}$} (2); \draw(6) --node[sloped, below, rotate=-65]{\tiny$q_{\red{s}}\hspace{.7cm}$} (2); \end{tikzpicture}
Figure 3.

Three different aspects of the minimal non-linear tree

Figure 4.

A planar embedding of a binary linear tree

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[x=.7cm,y=.7cm,every node/.style={circle, draw, scale=.5}, scale=1.0, rotate = 180, xscale = -1] \node[label={\red{root}}] (1) at ( 0, 3.2) {}; \node(2) at ( 1.5, 3.2) {}; \node(3) at ( 3, 3.2) {}; \node(4) at ( 8, 3.2) {}; \node(6) at ( 4.5, 3.2) {}; \node(8) at ( 4.5, 1.9) {}; \node(5) at ( 9.5, 3.2) {}; \node(7) at ( 1.5, 1.9) {}; \node(9) at ( 3, 1.9) {}; \node(14) at ( 8,1.9) {}; \node(15) at ( 9.5, 1.9) {}; \node(21) at ( 10.97, 3.2) {}; \draw[dashed] (2) -- (1); \draw[dashed] (3) -- (2); \draw[dashed] (6) -- (8); \draw[dashed] (6) -- (3); \draw[dotted] (6) -- (4); \draw[dashed] (5) -- (4); \draw[dashed] (7) -- (2); \draw[dashed] (3) -- (9); \draw[dashed] (14) -- (4); \draw[dashed] (15) -- (5); \draw[dashed] (21) -- (5); \end{tikzpicture}
Figure 5.

Critical ingredients blocked by the root () and by an order-disrespectful edge (, , , and )

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture} \node[style={circle, draw, scale=.6}, scale=1.0, xscale = -1] (1) at ( 1.3, 3) {0}; \node[scale=.2] (2) at ( 3.06, 3) {}; \node[scale=.2] (3) at ( 4.53, 3) {}; \node[style={circle, draw, scale=.5}] (4) at ( 6.0, 3) {$x_i$}; \node[style={circle, draw, scale=.6}] (5) at ( 6.6, 2.4) {}; \node[style={circle, draw, scale=.6}] (7) at ( 6.9, 3) {}; \node[style={circle, draw, scale=.6}] (8) at ( 7.8, 3) {}; \node[style={circle, draw, scale=.6}] (9) at ( 8.7, 3) {}; \node[style={circle, draw, scale=.6}] (10) at ( 6.8, 3.4) {}; \node[style={circle, draw, scale=.6}] (11) at ( 7.6, 3.8) {}; \node[style={circle, draw, scale=.6}] (6) at ( 8.4, 4.2) {}; \node[style={circle, draw, scale=.6}] (12) at ( 6.4, 3.9) {}; \node[style={circle, draw, scale=.6}] (13) at ( 2.3, 3 ) {}; \draw[very thin] (13) -- (1); \draw[very thin] (13) -- (2); \draw[dashed, very thin] (3) -- (2); \draw[very thin] (4) -- (3); \draw[very thin] (5) -- (4); \draw[very thin] (6) -- (11); \draw[very thick] (7) -- (4); \draw[very thin] (8) -- (7); \draw[very thin] (9) -- (8); \draw[very thin] (10) -- (4); \draw[very thin] (11) -- (10); \draw[very thin] (12) -- (4); \end{tikzpicture}
Figure 6.

A critical 3-cell

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[x=.6cm,y=.6cm] \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180, xscale = -1] (1) at ( 3.12, 2.86) {}; \node[fill=black,style={circle, draw, scale=.4},scale=1.0,rotate = 180,xscale = -1] (24) at (3.7, 2.86) {}; \node[fill=black,scale=.1] (2) at ( 6.09, 2.86) {}; \node[fill=black,scale=.1] (3) at ( 7.12, 1.51) {}; \node[fill=black,scale=.1] (4) at ( 7.66, 2.88) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180, xscale = -1] (5) at ( 7.14, 3.58) {}; \node[fill=black,scale=.1] (6) at ( 6, 4.14) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180, xscale = -1] (7) at ( 8.19, 1.12) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0, rotate = 180, xscale = -1] (8) at ( 8.69, 0.58) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0, rotate = 180, xscale = -1] (9) at ( 8.9, 1.12) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0, rotate = 180, xscale = -1] (10) at ( 8.7, 1.66) {}; \node[fill=black,style={circle, draw, scale=.4}, scale=1.0, rotate = 180, xscale = -1] (32) at (9.5,1.12){}; \node[fill=black,scale=.1] (11) at ( 8.66, 2.87) {}; \node[fill=black,scale=.1] (12) at ( 9.25, 2.35) {}; \node[fill=black,scale=.1] (13) at ( 9.26, 3.09) {}; \node[fill=black,scale=.1] (14) at ( 8.21, 4.35) {}; \node[fill=black,scale=.1] (15) at ( 8.92, 4.27) {}; \node[fill=black,scale=.1] (16) at ( 8.84, 4.76) {}; \node[fill=black,scale=.1] (17) at ( 6.79, 4.83) {}; \node[fill=black,scale=.1] (18) at ( 7.47, 5.15) {}; \node[fill=black,scale=.1] (19) at ( 6.98, 5.58) {}; \node[fill=black,scale=.1] (20) at ( 7.46, 1.01) {}; \node[fill=black,scale=.1] (21) at ( 6.95, 1.04) {}; \node[fill=black,scale=.1] (22) at ( 7.91, 2.39) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180,xscale = -1] (23) at ( 7.19, 4.08) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180,xscale = -1] (25) at ( 5.8, 5.14) {}; \node[fill=black,style={circle, draw, scale=.4}, scale=1.0, rotate = 180, xscale = -1] (33) at (5,6.1){}; \node[fill=black,scale=.1] (26) at ( 5.38, 4.7) {}; \node[fill=black,scale=.1] (27) at ( 8.41, 5.03) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180,xscale = -1] (28) at ( 7.67, 3.52) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180,xscale = -1] (29) at ( 6.14, 5.71) {}; \node[fill=black,style={circle,draw,scale=.4},scale=1.0,rotate = 180,xscale = -1] (30) at ( 5.33, 5.73) {}; \node[fill=black,scale=.1] (31) at ( 7.67, 1.71) {}; \node[fill=black,style={circle, draw, scale=.4}, scale=1.0, rotate = 180, xscale = -1] (34) at (8.2,3.46){}; \draw[very thin] (28) -- (34); \draw[very thin] (33) -- (30); \draw[very thin] (1)node[above]{\scriptsize0} -- (24); \draw[very thin] (24) -- (2); \draw[very thin] (32) -- (9); \draw[very thin] (3) -- (2); \draw[very thin] (4) -- (2); \draw[very thin] (5) -- (2); \draw[very thin] (6) -- (2); \draw[very thin] (7) -- (3); \draw[very thin] (8) -- (7); \draw[very thick] (7)node[below]{\scriptsize$x_3$} -- (9); \draw[very thin] (7) -- (10); \draw[very thin] (11) -- (4); \draw[very thin] (13) -- (11); \draw[very thin] (12) -- (11); \draw[very thin] (14) -- (5); \draw[very thin] (15) -- (14); \draw[very thin] (16) -- (14); \draw[very thin] (17) -- (6); \draw[very thin] (19) -- (17); \draw[very thin] (18) -- (17); \draw[very thin] (3) -- (21); \draw[very thin] (20) -- (3); \draw[very thin] (22) -- (4); \draw[very thin] (23) -- (5); \draw[very thin] (6) -- (25); \draw[very thin] (26) -- (6); \draw[very thin] (27) -- (14); \draw[very thick] (28) -- (5)node[below]{\scriptsize$x_2$}; \draw[very thick] (29) -- (25)node[right]{\scriptsize$x_1$}; \draw[very thin] (25) -- (30); \draw[very thin] (31) -- (3); \end{tikzpicture}
Figure 7.

A portion of the 1-cube with its recently created edge

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[scale=1.0, rotate = 180, xscale = -1] \node[style={circle, draw, scale=.5}] (1) at ( 1.62, 3.2) {}; \node[scale=.1] (2) at ( 2.94, 3.2) {}; \node[scale=.1] (3) at ( 4.22, 3.2) {}; \node[fill=black,style={circle, draw, scale=.5}] (4) at ( 5.75, 3.2) {}; \node[style={circle, draw, scale=.1}] (5) at ( 5.6, 4.3) {}; \node[scale=.1] (7) at ( 7.73, 3.2) {}; \node[fill=black,style={circle, draw, scale=.5}] (8) at ( 6.4, 3.5) {}; \node[style={circle, draw, scale=.1}] (9) at ( 6.3, 4.19) {}; \node[fill=black,style={circle, draw, scale=.5}] (10) at ( 9.6, 3.2) {}; \node(21) at ( 9.4, 3.5) {$y$}; \node(22) at ( 5.6, 2.9) {$x$}; \node(22) at ( 6.95, 3.6) {\red{$x[d']$}}; \node[fill=black,style={circle, draw, scale=.5}] (22) at ( 10.2, 3.53) {}; \node[fill=black,style={circle, draw, scale=.5}] (16) at (10,2.8){}; \node[fill=black,style={circle, draw, scale=.5}] (17) at (10.4,2.4){}; \node[fill=black,style={circle, draw, scale=.5}] (18) at (10.8,2){}; \node[fill=black,style={circle, draw, scale=.5}] (12) at ( 11, 3.2) {}; \node[fill=black,style={circle, draw, scale=.5}] (13) at ( 10.7, 3.8) {}; \node[fill=black,style={circle, draw, scale=.5}] (14) at (10.3,3.2){}; \node[fill=black,style={circle, draw, scale=.5}] (15) at (11.7,3.2){}; \node[fill=black,style={circle, draw, scale=.5}] (19) at (9.6,2.7){}; \node[fill=black,style={circle, draw, scale=.5}] (20) at (9.6,2.2){}; \node[fill=black,style={circle, draw, scale=.5}] (30) at (9.6,1.7){}; \draw[very thin] (10) -- (19); \draw[very thin] (19) -- (20); \draw[very thin] (30) -- (20); \draw[very thin] (17) -- (18); \draw[very thin] (12) -- (15); \draw[very thin] (2) -- (1)node[below]{0}; \draw(3)[dashed, very thin] -- (2); \draw(4)[very thin] -- (3); \draw[very thin] (7) -- (4); \draw[dashed, very thin] (10) -- (7); \draw[very thick] (14) -- (10); \draw[very thin] (12) -- (14); \draw[very thin] (16) -- (10); \draw[very thin] (17) -- (16); \draw[very thin] (13) -- (10); \draw[dashed, very thin] (5) -- (4); \draw[very thick] (8) -- (4); \draw[dashed,very thin] (9) -- (4); \end{tikzpicture}
Figure 8.

The edge belongs to , so the path from to does not pass through an essential vertex

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[scale=1.0, rotate = 180, xscale = -1] \node[style={circle, draw, scale=.5}] (1) at ( 1.62, 3.2) {}; \node at ( 1.62, 3.5) {$0$}; \node[scale=.1] (2) at ( 2.94, 3.2) {}; \node[scale=.1] (3) at ( 4.22, 3.2) {}; \node[style={circle, draw, scale=.5}] (4) at ( 5.74, 3.2) {}; \node at (5.64, 3.5) {$\red{x_t}$}; \node[style={circle, draw, scale=.5}] (5) at ( 6.1, 2.2) {}; \node[style={circle, draw, scale=.5}] (6) at ( 6.7, 2.6) {}; \node[scale=.1] (7) at ( 7.73, 3.2) {}; \node[style={circle, draw, scale=.5}] (8) at ( 6.85, 3.6) {}; \node[style={circle, draw, scale=.5}] (9) at ( 6.3, 4.1) {}; \node[style={circle, draw, scale=.5}] (10) at ( 9.59, 3.2) {}; \node at ( 9.5, 3.5) {$x_{m+1}$}; \node[style={circle, draw, scale=.5}] (11) at ( 10.7, 2.6) {}; \node[style={circle, draw, scale=.5}] (12) at ( 11, 3.2) {}; \node[style={circle, draw, scale=.5}] (13) at ( 10.7, 3.8) {}; \draw[very thin] (2) -- (1); \draw[dashed, very thin] (3) -- (2); \draw[very thin] (4) -- (3); \draw[very thin] (7) -- (4); \draw[dashed, very thin] (10) -- (7); \draw[very thick] (12)node[right]{\red{$\,\overline{x}_{m+1}$}} -- (10); \draw[very thin] (11) -- (10); \draw[very thin] (13) -- (10); \draw[very thin] (5) -- (4); \draw[very thin] (6) -- (4); \draw[very thick] (8)node[right]{\red{\,$\overline{x}_t$}} -- (4); \draw[very thin] (9) -- (4); \end{tikzpicture}
Figure 9.

The four possible configurations with

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[scale=.8, rotate = 180, xscale = -1] \node[style={circle, draw, scale=.6}] (1) at ( -4, 1.0) {}; \node[style={circle, draw, scale=.6}] (2) at ( -2.5, 1) {}; \node[style={circle, draw, scale=.6}] (3) at ( -2.5, 2) {}; \node[style={circle, draw, scale=.6}] (4) at (-1.8,0.3) {}; \node[style={circle, draw, scale=.6}] (5) at ( -1.5, 1) {}; \node[style={circle, draw, scale=.6}] (6) at ( -2, 1.67) {}; \node[style={circle, draw, scale=.6}] (7) at ( 0, 1) {}; \node[style={circle, draw, scale=.6}] (8) at ( .7, .3) {}; \node[style={circle, draw, scale=.6}] (9) at ( 1, 1) {}; \node[style={circle, draw, scale=.6}] (10) at ( .5, 1.67) {}; \node[style={circle, draw, scale=.6}] (11) at ( -4.0, 4.5) {}; \node[style={circle, draw, scale=.6}] (12) at ( -2.5, 4.5) {}; \node[style={circle, draw, scale=.6}] (50) at ( -1.7, 5) {}; \node[style={circle, draw, scale=.6}] (13) at ( -2.5, 3.5) {}; \node[style={circle, draw, scale=.6}] (14) at ( -1.8, 3.8) {}; \node[style={circle, draw, scale=.6}] (15) at ( 0, 4.5) {}; \node[style={circle, draw, scale=.6}] (16) at ( .7, 3.8) {}; \node[style={circle, draw, scale=.6}] (17) at ( .7, 5.2) {}; \node[style={circle, draw, scale=.6}] (18) at ( 1, 4.5) {}; \node[style={circle, draw, scale=.6}] (22) at ( 3, 4.5) {}; \node[style={circle, draw, scale=.6}] (23) at ( 4.5, 4.5) {}; \node[style={circle, draw, scale=.6}] (24) at ( 4.5, 3.5) {}; \node[style={circle, draw, scale=.6}] (25) at ( 5.2, 3) {}; \node[style={circle, draw, scale=.6}] (26) at ( 4.5, 2.5) {}; \node[style={circle, draw, scale=.6}] (27) at ( 5.5, 3.5) {}; \node[style={circle, draw, scale=.6}] (28) at ( 6, 4.5) {}; \node[style={circle, draw, scale=.6}] (29) at ( 6.7, 5.2) {}; \node[style={circle, draw, scale=.6}] (30) at ( 7, 4.5) {}; \node[style={circle, draw, scale=.6}] (31) at ( 6.7, 3.8) {}; \node[style={circle, draw, scale=.6}] (32) at ( 5.2, 5.2) {}; \node[style={circle, draw, scale=.6}] (41) at ( 3, 1.0) {}; \node[style={circle, draw, scale=.6}] (42) at ( 4.5, 1) {}; \node[style={circle, draw, scale=.6}] (43) at ( 4.5, 0) {}; \node[style={circle, draw, scale=.6}] (44) at (5.2,0.3) {}; \node[style={circle, draw, scale=.6}] (45) at ( 5.5, 1) {}; \node[style={circle, draw, scale=.6}] (46) at ( 5, 1.67) {}; \node[style={circle, draw, scale=.6}] (47) at ( 7, 1) {}; \node[style={circle, draw, scale=.6}] (48) at ( 7.7, .3) {}; \node[style={circle, draw, scale=.6}] (49) at ( 8, 1) {}; \node[style={circle, draw, scale=.6}] (40) at ( 7.5, 1.67) {}; \draw[dashed, very thin] (42) -- (41)node[below]{\raisebox{-3mm}{$0$}}; \draw[very thin] (43) -- (42)node[below]{\raisebox{-2mm}{$x$}}; \draw[very thin] (44) -- (42); \draw[very thin] (46) -- (42); \draw[very thick] (45)node[below]{$\overline{x}$} -- (42); \draw[very thin, dashed] (47) -- (45); \draw[very thin] (48) -- (47)node[below]{\raisebox{-2mm}{$x_1$}}; \draw[very thick] (49)node[right]{$\overline{x}_1$} -- (47); \draw[very thin] (40) -- (47); \draw[dashed, very thin] (2) -- (1)node[below]{\raisebox{-3mm}{$0$}}; \draw[very thin] (3) -- (2)node[above]{\raisebox{.5mm}{$x$}}; \draw[very thin] (4) -- (2); \draw[very thin] (5) -- (2); \draw[very thick] (6)node[right]{$\overline{x}$} -- (2); \draw[dashed, very thin] (7) -- (5); \draw[very thin] (8) -- (7)node[below]{\raisebox{-2mm}{$x_1$}}; \draw[very thick] (9)node[right]{$\overline{x}_1$} -- (7); \draw[very thin] (10) -- (7); \draw[dashed, very thin] (12) -- (11)node[below]{\raisebox{-3mm}{$0$}}; \draw[dashed, very thin] (15) -- (12)node[below]{\raisebox{-2mm}{$x$}}; \draw[very thin](13) -- (12); \draw[very thin](50) -- (12); \draw[very thick] (14)node[right]{$\overline{x}$} -- (12); \draw[very thin] (16) -- (15)node[below]{\raisebox{-2mm}{$x_1$}}; \draw[very thin] (17) -- (15); \draw[very thick] (18)node[right]{$\overline{x}_1$} -- (15); \draw[very thin, dashed] (23) -- (22)node[below]{\raisebox{-3mm}{$0$}}; \draw[very thin, dashed] (24)node[left]{$x$} -- (23); \draw[very thin] (27) -- (24); \draw[very thick] (25)node[right]{$\overline{x}$} -- (24); \draw[very thin] (26) -- (24); \draw[dashed, very thin] (28) -- (23); \draw[very thin] (29) -- (28)node[below]{\raisebox{-2mm}{$x_1$}}; \draw[very thin] (31) -- (28); \draw[very thick] (30)node[right]{$\overline{x}_1$} -- (28); \draw[very thin] (32) -- (23); \end{tikzpicture}
Figure 10.

Dynamics of paths in (top) and (bottom)

Figure 11.

Configurations of essential vertices in Lemma 6.5

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \setlength{\extrarowheight}{0.0pt}\begin{tikzpicture}[scale=.8, rotate = 180, xscale = -1] \node[style={circle, draw, scale=.6}] (1) at ( -6, 4.5) {}; \node(2) at ( -4.5, 4.5) {}; \node[style={circle, draw, scale=.6}] at ( -3.9, 5.7) {}; \node[style={circle, draw, scale=.6}] at ( -3.9, 3.3) {}; \node[style={circle, draw, scale=.6}] at ( -3.6, 5.2) {}; \node[style={circle, draw, scale=.6}] at ( -3.6, 3.8) {}; \node[style={circle, draw, scale=.6}] at ( -3.5, 4.55) {}; \node at ( -3.5, 3) {$x_1$}; \node at ( -3.1, 3.7) {$x_2$}; \node at ( -3.3, 5.9) {$x_m$}; \node[style={circle, draw, scale=.6}] (3) at ( 1, 4.5) {}; \node(11) at ( 2.4, 4.5) {}; \node(12) at ( 3.1, 4.5) {}; \node[style={circle, draw, scale=.6}] (4) at ( 4.5, 4.5) {}; \node(5) at ( 6.3, 5.5) {}; \node(6) at ( 6.9, 4.5) {}; \node(7) at ( 5.3, 4) {}; \node(8) at ( 5, 3.8) {}; \node(9) at ( 4.6, 3.7) {}; \node(10) at ( 4.2, 3.7) {}; \draw[dotted, thick] (12) -- (11); \draw[dashed, very thin] (2) -- (1)node[below]{\raisebox{-3mm}{\scriptsize$x_0\,$(root)}}; \draw[very thin, dashed] (12) -- (4); \draw[very thin, dashed] (11) -- (3)node[below]{\raisebox{-3mm}{\scriptsize$x_0\,$(root)}}; \draw[very thin, dashed] (4)node[below]{\raisebox{-3mm}{$x_1$}} -- (5); \draw[very thin, dashed] (6) -- (4); \draw[very thin, dashed] (7) -- (4); \draw[very thin, dashed] (8) -- (4); \draw[very thin, dashed] (9) -- (4); \draw[very thin, dashed] (10) -- (4); \node[style={circle, draw, scale=.6}] at (6.6, 5.7) {}; \node[style={circle, draw, scale=.6}] at (6.7, 5.3) {}; \node[style={circle, draw, scale=.6}] at (6.3, 6) {}; \node[style={circle, draw, scale=.6}] at (7.3, 4.5) {}; \node[style={circle, draw, scale=.6}] at (7.2, 4.1) {}; \node[style={circle, draw, scale=.6}] at (7.2, 4.9) {}; \node at (7.5, 3.7) {$x_2$}; \node at ( 6.8, 6.3) {$x_m$}; \end{tikzpicture}

Mathematical Fragments

Theorem 1.1.

Assume is a binary tree. For a commutative ring with 1, the cohomology ring is the exterior face ring determined by a simplicial complex . Explicitly, is the quotient , where is the exterior graded -algebra generated by the vertex set of , and is the ideal generated by monomials corresponding to non-faces of .

Definition 1.3 (The -interaction complex of , ).
(a)

The vertex set of is the collection of all 4-tuples , where is a non-negative integer number, is an essential vertex of , and and are tuples of non-negative integer numbers satisfying the three conditions

, with

, where and

for at least one .

We stress that (i.e., the length of ) is one of the parameters determining the 4-tuple . For instance, if and , then and are two different elements in . The length of , on the other hand, is determined by and .

(b)

For with , , and so that , consider the components and and of as defined above. Then, for , the -local information of , denoted by , is defined by

and, for ,

Note that whenever .

(c)

The -interaction complex of is the abstract simplicial complex whose vertex set is and whose -simplices are given by families of vertices as in item (b) satisfying

for all and all relevant , and in such a way that, for every , 2 is a strict inequality for at least one .

Definition 1.4.

Let be a family of systems of local informations around essential vertices of . We say that interact strongly provided is a simplex of . We say that interact weakly provided Equation 2 holds for all relevant and but fails to be a simplex of —so that, in fact, Equation 2 is an equality for some and all . In all other cases, we say that do not interact.

Example 1.5.

Figure 3 shows three aspects of the smallest possible non-linear tree . The four essential vertices are labelled (following the -order) in the central picture. The fact that the 4-fold product

is a basis element follows from Theorem 1.1, as inspection in the picture on the right of Figure 3 reveals that the factors in 3 interact strongly. Note that for each factor in 3, and that the cases with a strict inequality in Equation 2 hold as required in the last clause of item (c) of Definition 1.3. Likewise, interaction analysis in the picture on the left exhibits the well known fact that is not flag (i.e., is not a right-angled Artin group): the three basis elements , and in have pairwise strong interactions (so their three double products are part of a basis of ), but the three basis elements do not interact (so their triple product vanishes).

Example 1.6.

Let be a binary tree whose essential vertices lie along a single embedded arc. Choosing the planar embedding shown in Figure 4, we see that has a right-angled Artin group presentation with generators , where is an essential vertex of and are non-negative integer numbers⁠Footnote3 satisfying and . In these terms, has a commutativity relation whenever and , where the former inequality refers to the -order resulting from the embedding. Note that the chosen planar embedding of rules out weak interactions.

3

Instead of writing the 1-tuples and , we have simply written and .

Theorem 1.7.

For any tree , any non-negative integer and any commutative ring with unit 1, there is a set-theoretic inclusion so that the faces of yield, via cup-product of their vertices, a graded basis of . For instance, the empty face corresponds to the unit . Furthermore, any product with vanishes (in particular cup-squares vanish), as do cup-products of non-interacting elements in .

Remark 1.8.

The only obstructions for realizing in Theorem 1.7 as the exterior face ring determined by are the non-vanishing products whose factors interact weakly. For trees with binary core, such weak-interacting non-trivial products are effectively ruled out in the final section of this paper (Theorem 6.4) by means of a suitable change of basis that adjusts the inclusion in Theorem 1.7.

Remark 1.9.

The results in this paper allow us to recover and generalize Scheirer’s main technical tool Reference 16, Lemma 3.6 for studying Farber’s topological complexity of . Extensions of Scheirer’s results will be the topic of a future publication.

Equation (4)
Equation (5)
Example 2.1.

Let be a tree whose vertices and edges have been ordered as described in the previous section. Think of as cubical set. In fact, orient the edges of from the smaller to the larger endpoints and fix an orientation-preserving embedding of cubical sets, where elementary cubes in have product orientation. Thus, a vertex of becomes a 0-cube in , while an oriented edge in corresponds in to an oriented 1-cube , i.e., an elementary cube whose factors are degenerate, except for one of them.

Remark 2.2.

Particularly agreeable is the fact that a finite cartesian product of cubical sets comes equipped for free with the obvious structure of a cubical set. For instance, in the situation of Example 2.1, the cartesian power is a (product-oriented) cubical set in . In such a setting, an oriented cube in (where each is either a vertex or an edge of ) corresponds in to an oriented cube where, for each , at most one of the intervals is non-degenerate. These considerations, coupled with the fact that cubes of a single factor are at most one-dimensional, yield the next explicit description of cubical cup-products associated to and .

Proposition 2.3.

The cup product in of the duals of a pair of (oriented) cubes and in is given by the dual of

More generally, let be a (product-oriented) cubical subset of . The cup product in of the duals of a pair of cubes and in vanishes provided for some or, else, provided the cube is not contained in . Otherwise, the cup product is the multiple of the dual of , where

Equation (6)
Equation (7)
Equation (8)
Equation (9)
Equation (10)
Remark 2.5.

In any arrow of Farley-Sabalka’s modified Hasse diagram, is an even face of , i.e., in the notation of Equation 4, for some .

Remark 2.6.

By construction, Farley-Sabalka’s gradient field in is -equivariant and, by passing to the quotient, it yields the corresponding gradient field in . Consequently, gradient paths can equivalently be analyzed in either the ordered or unordered settings. Indeed, a gradient path in corresponds to a -orbit” of gradient paths in . Due to the cup-product descriptions in Subsection 2.1, we find it more convenient to perform the gradient-path analysis at the level of the cubical set .

Definition 3.1 (Gradient orientation, cf. Subsection 2.3 of Reference 5).

The listing , , of edge-ingredients of a -cube in or in is said to be in gradient order if , where the latter is the -ordering of vertices discussed in Section 1. The gradient orientation of is defined just as the product orientation, except that the gradient order of the edge-ingredients is used (rather than the coordinate order).

Equation (11)
Lemma 3.2.

The following diagram is commutative:

Example 3.5.

Any edge-ingredient of is . On the other hand, for an essential vertex and a positive direction from , an edge-ingredient of is if and only if has no ingredient, neither vertex nor edge, in any of the components of lying in -directions . Furthermore, if is an edge-ingredient of a face of some cube of , then is if and only if is .

Proposition 3.6.

Let be the gradient-order listing of the edge-ingredients of a -cube in .

(1)

If an arrow in the modified Hasse diagram for is of eor type, then is and, for any , is collapsible.

(2)

If the edge is , then there is no upper path starting at a face with and ending at a critical cube.

Equation (12)
Equation (13)
Corollary 3.7.

Let be an upper path in that ends at a critical cube. Any upper elementary factor of of sor type is of falling-vertex type.

Example 3.8.

Let us be specific about the dynamics of an upper path that ends at a critical 1-cube . By the -equivariance of the gradient field, we can assume with and

i.e., is the -orbit representative whose ingredients appear in the -ordering. By Corollary 3.7, the start of is forced to consist of falling-vertex elementary paths, where the vertices fall, each at a time, until they form the stack of vertices supported (and blocked) by the root. At that point arrives at the 1-cube and we see that must be positive, for otherwise would have reached a collapsible -cube. In particular must be an essential vertex and . Then, again by Corollary 3.7, it is the turn of vertices that are forced to fall, each at a time, until they form stacks of vertices blocked by in -directions . At that point arrives at a 1-cube of the form

Not all of the stacks are empty, so 14 has as a critical edge-ingredient. The falling-vertex process is also forced by Corollary 3.7 on those vertices that are located in positive -directions (if any), and this takes to a 1-cube of the form

with all lying in -direction 0. Branching starts from this point on, with explicit options discussed in the next paragraph.

If no vertices are left, then would have reached its final critical destination . Otherwise, is forced to fall until reaches, via some branch type pairing , the 2-cube depicted in Figure 7. At this point there are two options for . In the first option, is obtained from by replacing the recently created edge by , i.e., with an upper elementary path of falling-vertex type. In such a case, is forced to continue with the vertex falling until it is added to the stack of vertices blocked by the root 0. This leaves us at a situation similar to the one at the start of this paragraph. In the second option, is obtained from by replacing the edge by either of its end points. In such a case, is forced to continue:

(1)

with the falling of the vertices that are now unblocked in the neighborhood of (see Figure 7), until they form a stack of vertices blocked by —thus starting a critical situation around the edge —and, then,

(2)

with the falling of the vertices (if any) in -directions from to , which form (possibly empty) stacks of vertices blocked either by or —thus completing the critical situation around the edge .

Again, this leaves us at a situation similar to the one at the start of this paragraph, but now with the edge playing the role of the edge . The branching process in this paragraph then repeats, necessarily a finite number of times, until all vertices have been considered, when reaches its critical destination .

Proposition 3.9.

A cocycle in representing a 1-dimensional cohomology class in , with and , is given by

where the summation runs over

all permutations ,

all possible vertices in the component of in -direction 0,

all possible vertices in the components of in -directions from to so that, for , of the vertices lie in -direction ,

all possible vertices in the components of in -directions greater than so that, for , of the vertices lie in -direction .

Equation (16)
Equation (17)
Equation (18)
Proposition 3.10.

The Morse differential in vanishes.

Corollary 3.11.

The product of two basis elements vanishes provided . In particular, squares of 1-dimensional elements in are trivial.

Equation (19)
Definition 4.1 (Leaves and pruned trees).

Set and, for and ,

where stands for the interior of the edge . We think of each as a rooted but possibly pruned tree. Namely, in the notation of Section 1 and setting , the root of is , if or if with , whereas the root of is . Furthermore, the set of pruned leaves of is .

Remark 4.2.

Just as the sets give a partition of , the union of the trees agrees with the difference . Actually, each vertex of other than for , as well as each semi-open edge of not of the form with , belongs to a tree  for a unique .

Definition 4.3.
(1)

For a -tuple of integers , we write to mean that for all , reserving the expression to mean that with for at least one . Also, when , we write to denote a generic tuple of integers satisfying for all with in fact for at least one . We make no distinction between 1-tuples and integer numbers so, accordingly, we use instead of .

(2)

The interaction parameters , and of the factors in Equation 19 are given by

If , and for all , we say that the factors in Equation 19 interact weakly and, if in addition for some , we say that the factors in Equation 19 interact strongly. Otherwise, we say that the factors in Equation 19 do not interact.

Equation (20)
Equation (21)
Equation (22)
Equation (23)
Proposition 4.4.

The product Equation 19 is represented in by the gradient-oriented cocycle

where the summation runs over all permutations and all possible tuples of vertices written in -order, taken from the corresponding pruned trees , and having the following lengths: Any block must have ingredients, while any block with must have ingredients for , and ingredients for . In particular, Equation 19 vanishes provided its factors do not interact.

Equation (25)
Corollary 4.5.

The product Equation 19 agrees with the basis element provided the factors of Equation 19 interact strongly. Recall and .

Lemma 4.6.

Fix essential vertices and take positive integer numbers and with for . Let , , , with , and , be non-negative integers satisfying . Then the system

has a unique solution of non-negative integer numbers satisfying the condition for each . If, in addition, for each there exists with , then the unique solution satisfies that, for each , there exists with .

Equation (26)
Equation (27)
Theorem 5.1.

In the situation above, if the factors of interact but non-strongly, then

In the above expression we set , , , and . The summation in 28 runs over all -tuples of non-negative integer numbers satisfying . The inner summation in 29 runs over all -tuples of non-negative integer numbers and all non-negative integer numbers  satisfying . The inner summation in 30 is empty if , otherwise it runs over all -tuples of non-negative integer numbers and all non-negative integer numbers satisfying .

Equation (31)
Equation (32)
Equation (33)
Equation (34)
Equation (35)
Remark 5.2.

Since the closing-lock pairing Equation 34 must come from -direction , paths corresponding to cases with in Equation 35 have no contribution in Equation 32. Specifically, any path that starts from a critical cell with edges , where , by taking a face with , and that reaches the pairing Equation 34 through falling-vertex elementary paths, has a companion path that starts from the same critical cell by taking the face , and that also evolves through falling-vertex elementary paths until it reaches the closing-lock pairing Equation 34—so that and . Note that, in the ordered setting, and its companion path arrive at summands of Equation 31 whose ingredients differ only by a permutation (so as well). The phenomenon noticed in this remark is in fact the key to finishing the proof of the main result in this section.

Equation (36)
Equation (37)
Equation (38)
Equation (39)
Remark 6.2.

We use the angle-bracket notation since we have reserved the parenthesis notation for cubes in (as tuples of their ingredients). Additionally, the angle-bracket notation is intended to stress the change of basis in Equation 39.

Equation (40)
Remark 6.3.

Corollary 4.5, Proposition 4.4 and Theorem 5.1 show that both products in Equation 40 are (possibly empty) linear combinations of basis elements

Such a linear combination will be written as

Here and below, a dot  stands for either an unspecified ring coefficient or an unspecified tuple⁠Footnote11 of integer numbers, , satisfying when the tuple immediately follows an essential vertex  (the context clarifies the option).

11

As in Definition 4.3, we make no distinction between integer numbers and 1-tuples.

Theorem 6.4.

Let be a tree with binary core, be a commutative ring with 1, and . Then . In detail: (i) An ordered product is non-zero if and only if it is a strong interaction product. (ii) Two ordered strong interaction products agree if and only if they have the same factors. (iii) A graded basis of is given by the set of ordered strong interaction products.

Equation (41)
Equation (42)
Lemma 6.5.
(1)

Assume (left configuration in Figure 11), then

(2)

Assume and with (right configuration in Figure 11) with (i.e. and if , then

Lemma 6.6.

Assume . Then the product of with vanishes provided and .

Corollary 6.7.

If the factors on the left of Equation 41 do not yield a strong interaction product, then .

Lemma 6.8.

Assume with . Then

Proposition 6.9.

Assume and , with and . Then

Here and below each expression is meant to represent a pair of unspecified tuples of integer numbers with , and such that and in the product ordering, i.e., for and , with at least one of the last inequalities being strict.

Equation (44)
Equation (45)
Example 6.11.

Lemma 6.5(1) gives in interaction level 1 (under a strong condition hypothesis). Likewise, Equation 45 becomes

in interaction level 2 (with , so consist of alone). In full generality:

Proposition 6.12.

Let be essential vertices having interaction level . If is a strong interaction product, then

Equation (48)
Equation (49)
Equation (50)
Equation (51)

References

[1]
Aaron David Abrams, Configuration spaces and braid groups of graphs, ProQuest LLC, Ann Arbor, MI, 2000. Thesis (Ph.D.)–University of California, Berkeley. MR2701024, Show rawAMSref\bib{MR2701024}{book}{ author={Abrams, Aaron David}, title={Configuration spaces and braid groups of graphs}, note={Thesis (Ph.D.)--University of California, Berkeley}, publisher={ProQuest LLC, Ann Arbor, MI}, date={2000}, pages={67}, isbn={978-0599-85818-3}, review={\MR {2701024}}, } Close amsref.
[2]
Jorge Aguilar-Guzmán, Jesús González, and Teresa Hoekstra-Mendoza, Farley-Sabalka’s Morse-theory model and the higher topological complexity of ordered configuration spaces on trees, Discrete Comput. Geom. 67 (2022), no. 1, 258–286, DOI 10.1007/s00454-021-00306-3. MR4356250, Show rawAMSref\bib{jorgetereyo}{article}{ author={Aguilar-Guzm\'{a}n, Jorge}, author={Gonz\'{a}lez, Jes\'{u}s}, author={Hoekstra-Mendoza, Teresa}, title={Farley-Sabalka's Morse-theory model and the higher topological complexity of ordered configuration spaces on trees}, journal={Discrete Comput. Geom.}, volume={67}, date={2022}, number={1}, pages={258--286}, issn={0179-5376}, review={\MR {4356250}}, doi={10.1007/s00454-021-00306-3}, } Close amsref.
[3]
Francis Connolly and Margaret Doig, On braid groups and right-angled Artin groups, Geom. Dedicata 172 (2014), 179–190, DOI 10.1007/s10711-013-9914-6. MR3253777, Show rawAMSref\bib{MR3253777}{article}{ author={Connolly, Francis}, author={Doig, Margaret}, title={On braid groups and right-angled Artin groups}, journal={Geom. Dedicata}, volume={172}, date={2014}, pages={179--190}, issn={0046-5755}, review={\MR {3253777}}, doi={10.1007/s10711-013-9914-6}, } Close amsref.
[4]
Daniel Farley, Homology of tree braid groups, Topological and asymptotic aspects of group theory, Contemp. Math., vol. 394, Amer. Math. Soc., Providence, RI, 2006, pp. 101–112, DOI 10.1090/conm/394/07437. MR2216709, Show rawAMSref\bib{MR2216709}{article}{ author={Farley, Daniel}, title={Homology of tree braid groups}, conference={ title={Topological and asymptotic aspects of group theory}, }, book={ series={Contemp. Math.}, volume={394}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2006}, pages={101--112}, review={\MR {2216709}}, doi={10.1090/conm/394/07437}, } Close amsref.
[5]
Daniel Farley, Presentations for the cohomology rings of tree braid groups, Topology and robotics, Contemp. Math., vol. 438, Amer. Math. Soc., Providence, RI, 2007, pp. 145–172, DOI 10.1090/conm/438/08451. MR2359035, Show rawAMSref\bib{MR2359035}{article}{ author={Farley, Daniel}, title={Presentations for the cohomology rings of tree braid groups}, conference={ title={Topology and robotics}, }, book={ series={Contemp. Math.}, volume={438}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2007}, pages={145--172}, review={\MR {2359035}}, doi={10.1090/conm/438/08451}, } Close amsref.
[6]
Daniel Farley and Lucas Sabalka, Discrete Morse theory and graph braid groups, Algebr. Geom. Topol. 5 (2005), 1075–1109, DOI 10.2140/agt.2005.5.1075. MR2171804, Show rawAMSref\bib{MR2171804}{article}{ author={Farley, Daniel}, author={Sabalka, Lucas}, title={Discrete Morse theory and graph braid groups}, journal={Algebr. Geom. Topol.}, volume={5}, date={2005}, pages={1075--1109}, issn={1472-2747}, review={\MR {2171804}}, doi={10.2140/agt.2005.5.1075}, } Close amsref.
[7]
Daniel Farley and Lucas Sabalka, On the cohomology rings of tree braid groups, J. Pure Appl. Algebra 212 (2008), no. 1, 53–71, DOI 10.1016/j.jpaa.2007.04.011. MR2355034, Show rawAMSref\bib{MR2355034}{article}{ author={Farley, Daniel}, author={Sabalka, Lucas}, title={On the cohomology rings of tree braid groups}, journal={J. Pure Appl. Algebra}, volume={212}, date={2008}, number={1}, pages={53--71}, issn={0022-4049}, review={\MR {2355034}}, doi={10.1016/j.jpaa.2007.04.011}, } Close amsref.
[8]
Robin Forman, A discrete Morse theory for cell complexes, Geometry, topology, & physics, Conf. Proc. Lecture Notes Geom. Topology, IV, Int. Press, Cambridge, MA, 1995, pp. 112–125. MR1358614, Show rawAMSref\bib{MR1358614}{article}{ author={Forman, Robin}, title={A discrete Morse theory for cell complexes}, conference={ title={Geometry, topology, \& physics}, }, book={ series={Conf. Proc. Lecture Notes Geom. Topology, IV}, publisher={Int. Press, Cambridge, MA}, }, date={1995}, pages={112--125}, review={\MR {1358614}}, } Close amsref.
[9]
Robin Forman, Discrete Morse theory and the cohomology ring, Trans. Amer. Math. Soc. 354 (2002), no. 12, 5063–5085, DOI 10.1090/S0002-9947-02-03041-6. MR1926850, Show rawAMSref\bib{MR1926850}{article}{ author={Forman, Robin}, title={Discrete Morse theory and the cohomology ring}, journal={Trans. Amer. Math. Soc.}, volume={354}, date={2002}, number={12}, pages={5063--5085}, issn={0002-9947}, review={\MR {1926850}}, doi={10.1090/S0002-9947-02-03041-6}, } Close amsref.
[10]
Robert Ghrist, Configuration spaces and braid groups on graphs in robotics, Knots, braids, and mapping class groups—papers dedicated to Joan S. Birman (New York, 1998), AMS/IP Stud. Adv. Math., vol. 24, Amer. Math. Soc., Providence, RI, 2001, pp. 29–40, DOI 10.1090/amsip/024/03. MR1873106, Show rawAMSref\bib{MR1873106}{article}{ author={Ghrist, Robert}, title={Configuration spaces and braid groups on graphs in robotics}, conference={ title={Knots, braids, and mapping class groups---papers dedicated to Joan S. Birman}, address={New York}, date={1998}, }, book={ series={AMS/IP Stud. Adv. Math.}, volume={24}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2001}, pages={29--40}, review={\MR {1873106}}, doi={10.1090/amsip/024/03}, } Close amsref.
[11]
Tomasz Kaczynski, Konstantin Mischaikow, and Marian Mrozek, Computational homology, Applied Mathematical Sciences, vol. 157, Springer-Verlag, New York, 2004, DOI 10.1007/b97315. MR2028588, Show rawAMSref\bib{MR2028588}{book}{ author={Kaczynski, Tomasz}, author={Mischaikow, Konstantin}, author={Mrozek, Marian}, title={Computational homology}, series={Applied Mathematical Sciences}, volume={157}, publisher={Springer-Verlag, New York}, date={2004}, pages={xviii+480}, isbn={0-387-40853-3}, review={\MR {2028588}}, doi={10.1007/b97315}, } Close amsref.
[12]
Tomasz Kaczynski and Marian Mrozek, The cubical cohomology ring: an algorithmic approach, Found. Comput. Math. 13 (2013), no. 5, 789–818, DOI 10.1007/s10208-012-9138-4. MR3105945, Show rawAMSref\bib{MR3105945}{article}{ author={Kaczynski, Tomasz}, author={Mrozek, Marian}, title={The cubical cohomology ring: an algorithmic approach}, journal={Found. Comput. Math.}, volume={13}, date={2013}, number={5}, pages={789--818}, issn={1615-3375}, review={\MR {3105945}}, doi={10.1007/s10208-012-9138-4}, } Close amsref.
[13]
Jee Hyoun Kim, Ki Hyoung Ko, and Hyo Won Park, Graph braid groups and right-angled Artin groups, Trans. Amer. Math. Soc. 364 (2012), no. 1, 309–360, DOI 10.1090/S0002-9947-2011-05399-7. MR2833585, Show rawAMSref\bib{MR2833585}{article}{ author={Kim, Jee Hyoun}, author={Ko, Ki Hyoung}, author={Park, Hyo Won}, title={Graph braid groups and right-angled Artin groups}, journal={Trans. Amer. Math. Soc.}, volume={364}, date={2012}, number={1}, pages={309--360}, issn={0002-9947}, review={\MR {2833585}}, doi={10.1090/S0002-9947-2011-05399-7}, } Close amsref.
[14]
Ki Hyoung Ko, Joon Hyun La, and Hyo Won Park, Graph 4-braid groups and Massey products, Topology Appl. 197 (2016), 133–153, DOI 10.1016/j.topol.2015.10.010. MR3426912, Show rawAMSref\bib{MR3426912}{article}{ author={Ko, Ki Hyoung}, author={La, Joon Hyun}, author={Park, Hyo Won}, title={Graph 4-braid groups and Massey products}, journal={Topology Appl.}, volume={197}, date={2016}, pages={133--153}, issn={0166-8641}, review={\MR {3426912}}, doi={10.1016/j.topol.2015.10.010}, } Close amsref.
[15]
Lucas Sabalka, On rigidity and the isomorphism problem for tree braid groups, Groups Geom. Dyn. 3 (2009), no. 3, 469–523, DOI 10.4171/GGD/67. MR2516176, Show rawAMSref\bib{MR2516176}{article}{ author={Sabalka, Lucas}, title={On rigidity and the isomorphism problem for tree braid groups}, journal={Groups Geom. Dyn.}, volume={3}, date={2009}, number={3}, pages={469--523}, issn={1661-7207}, review={\MR {2516176}}, doi={10.4171/GGD/67}, } Close amsref.
[16]
Steven Scheirer, Topological complexity of points on a tree, Algebr. Geom. Topol. 18 (2018), no. 2, 839–876, DOI 10.2140/agt.2018.18.839. MR3773741, Show rawAMSref\bib{MR3773741}{article}{ author={Scheirer, Steven}, title={Topological complexity of $n$ points on a tree}, journal={Algebr. Geom. Topol.}, volume={18}, date={2018}, number={2}, pages={839--876}, issn={1472-2747}, review={\MR {3773741}}, doi={10.2140/agt.2018.18.839}, } Close amsref.

Article Information

MSC 2020
Primary: 20F36 (Braid groups; Artin groups), 55R80 (Discriminantal varieties and configuration spaces in algebraic topology), 57M15 (Relations of low-dimensional topology with graph theory), 57Q70 (Discrete Morse theory and related ideas in manifold topology)
Keywords
  • Tree braid group
  • cubical cup-product
  • discrete Morse theory
  • Farley-Sabalka gradient field
Author Information
Jesús González
Departamento de Matemáticas, Centro de Investigación y de Estudios Avanzados del I.P.N., Av. Instituto Politécnico Nacional número 2508, San Pedro Zacatenco, México City 07000, México
jesus@math.cinvestav.mx
ORCID
Teresa Hoekstra-Mendoza
Departamento de Matemáticas, Centro de Investigación y de Estudios Avanzados del I.P.N., Av. Instituto Politécnico Nacional número 2508, San Pedro Zacatenco, México City 07000, México
idskjen@math.cinvestav.mx
MathSciNet
Journal Information
Transactions of the American Mathematical Society, Series B, Volume 9, Issue 34, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License (CC BY NC ND 4.0)
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.