Google Summer of Code 2015: outcomes

Guest post by Chao Xu

In the summer, I have extended the SAGE code base for matroids for Google Summer of Code. This post shows a few example of it’s new capabilities.

Connectivity

Let $M$ be a matroid with groundset $E$ and rank function $r$. A partition of the groundset $\{E_1,E_2\}$ is a $m$-separation if $|E_1|,|E_2|\geq m$ and $r(E_1)+r(E_2)-r(E)\leq m-1$. $M$ is called $k$-connected if there is no $m$-separation for any $m < k$. The Fano matroid is an example of $3$-connected matroid.

The Fano matroid is not $4$-connected. Using the certificate=True field, we can also output a certificate that verify its not-$4$-connectness. The certificate is a $m$-separation where $m < 4$. Since we know Fano matroid is $3$-connected, we know the output should be a $3$-separation.

We also have a method for deciding $k$-connectivity, and returning a certificate.

There are 3 algorithms for $3$-connectivity. One can pass it as a string to the algorithm field of is_3connected.

  1. "bridges": The $3$-connectivity algorithm Bixby and Cunningham. [BC79]
  2. "intersection": the matroid intersection based algorithm
  3. "shifting": the shifting algorithm. [Raj87]

The default algorithm is the bridges based algorithm.

The following is an example to compare the running time of each approach.

The new bridges based algorithm is much faster than the previous algorithm in SAGE.

For $4$-connectivity, we tried to use the shifting approach, which has an running time of $O(n^{4.5}\sqrt{\log n})$, where $n$ is the size of the groundset. The intuitive idea is fixing some elements and tries to grow a separator. In theory, the shifting algorithm should be fast if the graph is not $4$-connected, as we can be lucky and find a separator quickly. In practice, it is still slower than the optimized matroid intersection based algorithm, which have a worst case $O(n^5)$ running time. There might be two reasons: the matroid intersection actually avoids the worst case running time in practice, and the shifting algorithm is not well optimized.

Matroid intersection and union

There is a new implementation of matroid intersection algorithm based on Cunningham’s paper [Cun86]. For people who are familiar with blocking flow algorithms for maximum flows, this is the matroid version. The running time is $O(\sqrt{p}rn)$, where $p$ is the size of the maximum common independent set, $r$ is the rank, and $n$ is the size of the groundset. Here is an example of taking matroid intersection between two randomly generated linear matroids.

Using matroid intersection, we have preliminary support for matroid union and matroid sum. Both construction takes a list of matroids.

The matroid sum operation takes disjoint union of the groundsets. Hence the new ground set will have the first coordinate indicating which matroid it comes from, and second coordinate indicate the element in the matroid.

Here is an example of matroid union of two copies of uniform matroid $U(1,5)$ and $U(2,5)$. The output is isomorphic to $U(4,5)$.

One of the application of matroid union is matroid partitioning, which partitions the groundset of the matroid to minimum number of independent sets. Here is an example that partitions the edges of a graph to minimum number of forests.

Acknowledgements

I would like to thank my mentors Stefan van Zwam and Michael Welsh for helping me with the project. I also like to thank Rudi Pendavingh, who have made various valuable suggestions and implemented many optimizations himself.

References

[BC79] R.E Bixby, W.H Cunningham. Matroids, graphs and 3-connectivity, J.A Bondy, U.S.R Murty (Eds.), Graph Theory and Related Topics, Academic Press, New York (1979), pp. 91-103.

[Raj87] Rajan, A. (1987). Algorithmic applications of connectivity and related topics in matroid theory. Northwestern university.

[Cun86] William H Cunningham. 1986. Improved bounds for matroid partition and intersection algorithms. SIAM J. Comput. 15, 4 (November 1986), 948-957.

Extremal Matroids and Growth Rates

Guest post by Sandra Kingan

Growth rates are a popular topic on Matroid Union. Peter Nelson has written five posts on it: Growth Rates IIIIIIIVV.  Irene Pivotto has also written a post on it.

Following James Oxley’s book [Oxley, 2012] , the growth rate function of a minor-closed class $\mathcal M$, denoted by $h_{\mathcal M}(r)$, is defined as the maximum number of elements in a simple rank-$r$ matroid in $\mathcal M$ if the number is finite and infinity otherwise (p. 570). A simple matroid is extremal in a minor-closed class $\mu$ of matroid if $M\in \mu$, but $\mu$ contains no simple single-element extension of $M$ that has the same rank as $M$ (p. 572).

For example, if $\mathcal M$ is the class of graphs, then the complete graph of rank-$r$ denoted by $K_{r+1}$, for $r\ge 1$ is the rank-$r$ extremal matroid. The growth rate function is the size of $K_{r+1}$ which is $\frac {r(r+1)}{2}$.  On the other hand if $\mathcal M$ is the subclass of planar graphs, then we don’t know precisely the extremal planar graphs, but we do know the growth rate function. A straightforward graph theoretic argument (found in any graph theory textbook) shows that the number of edges of a “maximal” planar graph is $3r-3$. (In graph theory a maximal planar graph is defined as one to which if you add one more edge it becomes non-planar.)

Similarly, if $\mathcal M$ is the class of binary matroids, then the rank-$r$ projective geometry $PG(r-1, 2)$, for $r\ge 2$, is the rank-$r$ extremal matroid. The growth rate function is the size of $PG(r-1, 2)$ which is $2^r-1$.

Within the class of binary matroids lies the class of cographic matroids (matroids whose duals are graphic). Again we don’t know precisely the extremal cographic matroids, but we can prove that the growth rate function is $3r-3$ (as shown by Joseph Kung and further explained further in Pivotto’s post). Also in her post is the growth rate of series-parallel networks. Series-parallel networks are formed by starting with a single edge or loop and repeatedly adding edges in parallel or edges in series (turning one edge into two edges by putting a vertex on the edge). The growth rate of series parallel networks is $2r+1$. See [Oxley Section 5.4] and [Oxley 1987, Section 3] for the explanation.

The first extremal result is generally attributed to Paul Turán. The class he considered was the class of graphs with no $K_{s+1}$-subgraph. (Mantel proved the $K_3$ case earlier.) Turán proved that the rank $r$ extremal graph is a specific type of complete s-partite graph (see Wikipedia). The growth rate function is $\frac{(s-1)(r+1)^2}{2s}$ [Turan, 1941].

With respect to the growth rate function we are keen on knowing if it is linear or a higher order polynomial like quadratic or exponential. See Nelson’s posts for detailed explanations on this.

In 1963 G. A. Dirac determined the graphs without two vertex disjoint cycles [Dirac, 1963]. Excluding two vertex-disjoint cycles in a 3-connected graph is equivalent to excluding the prism graph $(K_5\backslash e)^*$ as a minor. So essentially Dirac determined the 3-connected graphs with no prism minor. He found that the 3-connected members of this class are $K_5$, $K_5 \backslash e$, the infinite family of wheel graphs $W_r$, for $r\ge 4$, and $K_{3,p}$, $K’_{3,p} $, $K”_{3,p} $ or $K”’_{3,p}$, for some $p\ge 3$.

Theorem 1.  A simple $3$-connected graph has no minor isomorphic to $(K_5\backslash e)^*$ if and only if it is isomorphic to $W_r$ for some $r\ge 3$, $K_5$, $K_5 \backslash e$, $K_{3,p}$, $K’_{3,p} $, $K”_{3,p} $ or $K”’_{3,p}$, for some $p\ge 3$.

Observe that there are two very distinct 3-connected infinite families here: $W_r$ and $K_{3, p}”’$. The size of $W_r$ is $2r+1$ for $r\ge 3$ and the size of $K_{3, r-2}”’$ is $3r-3$ for $r\ge 5$. Both of these 3-connected infinite families are growing linearly. I’ve heard the word “monarchs” used to describe $W_r$ and $K_{3, p}”’$; I think Jack Edmonds used this term. Monarchs is a good word to describe them.

Matroids that are not 3-connected may be pieced together from 3-connected matroids using direct-sums and 2-sums [Oxley, 2012, 8.3.1]. This result was first proven by Bixby in 1972. This is a terrific decomposition result that allows us to focus on just the 3-connected matroids in the family. Thus if we know the 3-connected matroids we can build the 2-connected ones via the operation of two sums.

Now extremal matroid is defined as the simple rank-$r$ matroid of maximum size. So 2-sums and even direct sums are allowed. But when the class contains $W_r$ and $K_{3, r-2}”’$ with $W_r$ growing at rate $2r$ and $K_{3, r-2}”’$ growing at rate $3r-3$, the latter dominates any rank-$r$ graph that can be constructed from direct sum or 2-sum. The proof is straightforward case-checking. Thus the growth rate of the class of graphs with no prism minor is $3r-3$ and the rank-$r$ extremal matroid is $K_{3, r-2}”’$. Even if we were not able to say precisely what the growth rate is, once the 3-connected families are found, and found to be growing linearly, piecing together using direct sum and 2-sum will still give a linear growth rate.

Moving on let’s take a look at Oxley’s 1987 result where he determined the 3-connected matroids with no 4-wheel minor [Oxley, 1987]. The main theorem of that paper is Theorem 2.1 and the key component of the main theorem is Theorem 2.2 which is stated below. Let $Z_r$ denote the $(2r+1)$-element rank-$r$ binary spikes. They are non-regular matroids represented by the binary matrix $[I_r| D]$ where $D$ has $r+1$ columns labeled $b_1, \dots b_r, c_r$. The first $r$ columns in $D$ have zeros along the diagonal and ones elsewhere. The last column is all ones.

Theorem 2. A 3-connected binary non-regular matroid has no minor isomorphic to $P_9$ or ${P_9}^*$ if and only if it is isomorphic to $F_7$, ${F_7}^*$, $Z_r$, ${Z_r}^*$, $Z_r \backslash b_r$, or $Z_r \backslash b_r$, for some $\ge 4$.

Observe that the 3-connected infinite family $Z_r$ is growing at the rate of $2r+1$ elements. In Section 3 of his paper Oxley states without proof (details are straightforward) that the growth rate of the class of binary matroids with no 4-wheel minor is $3r-2$ if $r$ is odd and $3r-3$ if $r$ is even (Theorem 3.1). Why the difference: $2r+1$ versus $3r-2$? The answer is given in the second part of the statement of the theorem, you can 2-sum $F_7$ with itself or $F_7$ with $Z_4$ or $F_7$ with $U_{2,3}$ and get slightly larger rank-$r$ matroids as compared to $Z_r$.

This is why it is sufficient to know the 3-connected members of a class, and for all practical purposes the definition of extremal matroid really should have had 3-connected in it. The definition of extremal matroid made in the 1950s did not forsee Bixby’s result. There’s probably no point changing any definition, but it must be understood clearly that if the 3-connected members of a class are known, then so is the growth rate.

This raises the question: What if we don’t know the 3-connected members, but we have a decomposition result? The first and most well-known such result in Matroid Theory is Paul Seymour’s regular matroid decomposition. He proved that if $M$ is a 3-connected matroid in $EX(F_7, F_7^*)$, then $M$ can be constructed by piecing together graphic and cographic matroids and $R_{10}$, which is a special 10-element matroid [Seymour, 1980].  The  3-connected regular matroids are not known precisely. However, in 1957 Heller showed that the growth rate of the class of regular matroids is $\frac {r(r+1)}{2}$. His proof predates Seymour’s decomposition theorem. I have not seen a proof of growth rate of regular matroids based on the Seymour’s decomposition theorem. Perhaps if someone has, a reference could be mentioned in the comments.

More recently, Kung, Mayhew, Pivotto, and Royle showed the growth rate of $EX(AG(3,2)$ is $\frac {r(r+1)}{2}$. See Nelson’s fifth post on growth-rates. In that post he said:

Given this (and Irene asked a question along these lines in her post), it is natural to ask for a characterisation of exactly which classes of matroids grow like the graphic ones:

In a paper titled “Growth rates of binary matroids with no $P_9^*$-minor,” I found another class that exhibits this behavior.

Theorem 3. Let $M$ be a simple rank-$r$ binary matroid with no $P_9^*$ minor. Then, $|E(M)| \le \frac {r(r+1)}{2}$ with this bound being attained if and only if $M\cong K_{r+1}$. Moreover, the growth rate function of non-regular matroids in $EX[P_9^*]$ is $4r-5$, for $r\ge 6$.

The technique in Oxley’s 1987 paper is Seymour’s Splitter Theorem which was used to obtain the precise 3-connected members. The technique in my paper is the Strong Splitter Theorem, which was joint work with Manoel Lemos, where we built on the Splitter Theorem. Using it I was able to find the 3-connected members for $EX({P_9}^*)$. This approach is completely different from the approach Kung, Mayhew, Pivotto, and Royle adopted to find $EX(AG(3,2)$.

This post is getting quite long. In my next post I will describe the complete structure of the class $EX(P_9^*)$ and explain how I can say precisely the growth rate of the 3-connected non-regular matroids is $4r-5$. Note that $4r-5$ is the growth rate of the rank-$r$ 3-connected family, but it is large enough that it dominates any rank-$r$ 2-connected matroid pieced together from small 3-connected matroids.

The next post will appear on my blog Graphs, Matroids, etc.

References:

[Heller, 1957] I. Heller (1957), On linear systems with integral valued solutions, Pacific J. Math. 7, 1351-1364.

[Kingan, 2015] S. R. Kingan, Growth rates of binary matroids with no $P_9^*$-minor,  arXiv:1412.8169.

[Kingan and Lemos, 2014] S. R. Kingan and M. Lemos (2014), Strong Splitter Theorem Annals of CombinatoricsVol. 18 – 1, 111 – 116.

[Kung, Mayhew, Pivotto, Royle, 2013] J. Kung, D. Mayhew, I. Pivotto, and G. Royle, Maximum size binary matroids with no AG(3,2)-minor are graphic,  http://epubs.siam.org/doi/abs/10.1137/130918915

[Oxley, 1987] J. G. Oxley (1987). The binary matroids with no 4-wheel minor, {\it Trans. Amer. Math. Soc.} {\bf 154}, 63-75.

[Oxley, 2012] J. G. Oxley (2012). {\it Matroid Theory}, Second Edition, Oxford University Press, New York.

[Seymour, 1980] P. D. Seymour (1980) Decomposition of regular matroids, {\it J. Combin. Theory Ser. B} {\bf 28}, (1980) 305-359.

[Turán, 1941] P. Turán (1941), “On an extremal problem in graph theory”, Matematikaiés Fizikai Lapok (in Hungarian) 48: 436–452.

 

Which biased graphs are group-labelled graphs? A postscript: Do we really need infinite groups?

Guest post by Daryl Funk

This post is meant as a short postscript to Irene’s post in March, Which biased graphs are group-labelled graphs? Since Irene’s post generated a fair number of comments, there may be some interest in taking a closer look at the construction used in Theorem 1 of that post. Let us quickly remind ourselves of the context, and the theorem.

A graph $G$ together with a collection $\mathcal{B}$ of its cycles — called balanced — is a biased graph. Cycles not in $\mathcal{B}$ are unbalanced. Orienting the edges of $G$ and labelling each edge with an element of a group gives us a group-labelled graph. A group-labelled graph naturally gives rise to a biased graph, as follows. Traverse each cycle via a simple closed walk, taking the product of the group elements labelling each edge as you go, taking the inverse of the element when traversing an edge in reverse. Declare as balanced just those cycles for which such a product is the identity. Thus, group-labelled graphs provide the primary examples of biased graphs. An arbitrary biased graph $(G,\mathcal{B})$ is group-labellable if there is a group and a labelling for which this procedure yields precisely its collection of balanced cycles $\mathcal{B}$. Hence the question of our title: Which biased graphs are group-labelled? Theorem 1 provides an answer.

Theorem 1. Let $(G,\mathcal{B})$ be a biased graph. Construct a 2-cell complex $K$ from $G$ by adding a disc with boundary $C$ for each $C \in \mathcal{B}$. Then the following are equivalent:

  1. $(G,\mathcal{B})$ is group-labellable.
  2. $(G,\mathcal{B})$ is $\pi_1(K)$-labellable (where $\pi_1(K)$ is the fundamental group of $K$).
  3. Every unbalanced cycle of $G$ is noncontractible in $K$.

Barring degenerate cases, the $\pi_1(K)$-labelling constructed by Theorem 1 is a labelling by an infinite group. In practice, we often work with graphs labelled by a finite group — probably the most studied group-labelled graphs are those labelled by the group of order two (i.e. signed graphs). Moreover, we consider only finite graphs. The following therefore is a natural question:

Question 1. If $(G,\mathcal{B})$ is group-labellable, is $(G,\mathcal{B})$ labellable by a finite group?

Could it be that there are group-labelled graphs whose collections of balanced cycles may only be realised by a labelling using an infinite group?

Given any group-labelling of a biased graph $(G,\mathcal{B})$, say using elements of the group $\Gamma$, there is a homomorphism from the fundamental group $\pi_1(K)$ of the 2-cell complex $K$ we construct in Theorem 1, to $\Gamma$. This homomorphism makes explicit the connection between the topology of $K$ and group-labellings of $(G,\mathcal{B})$. It also enables us to easily answer Question 1.

In preparation for a description of this homomorphism, let us recall some additional terminology and notation from Irene’s original post. Let $G$ be a graph and let $\Gamma$ be a group. Orient the edges of $G$, and let $\phi: E(G) \to \Gamma$ be a function. We extend $\phi$ to the set of closed walks in $G$ by defining, for closed walk $W$ with edge sequence $e_1, e_2, \ldots, e_k$,
\[ \phi(W) = \phi(e_1)^{\textrm{dir}(e_1)} \phi(e_2)^{\textrm{dir}(e_2)} \cdots \phi(e_k)^{\textrm{dir}(e_k)} \]
where $\textrm{dir}(e_i)$ is 1 if $W$ traverses edge $e_i$ in the forward direction, and is $-1$ if $W$ traverses $e_i$ in the backward direction. Let $\mathcal{B}_\phi$ be the collection of cycles $C$ for which a simple closed walk $W$ traversing $C$ has $\phi(W)=1$. With this notation, $(G,\mathcal{B})$ is group-labellable if there is a group $\Gamma$ and labelling $\phi$ such that $\mathcal{B}_\phi = \mathcal{B}$. In this case we say the labelling $\phi$ realises $\mathcal{B}$, and that $(G,\mathcal{B})$ is $\Gamma$-labellable.

As Irene described near the end of her post, there is a fourth statement equivalent to statements 1, 2, and 3 of Theorem 1. To state it, we need just two more definitions. A rerouting of a closed walk $W$ is a closed walk $W’$ obtained from $W$ by replacing a subpath $P$ of $W$, say between vertices $u$ and $v$, with another $u$-$v$ path $P’$ that is internally disjoint from $P$. If $P \cup P’$ is a balanced cycle, then this is a balanced rerouting. Statement 4 below is equivalent to Statements 1, 2, and 3 of Theorem 1:

  1. No unbalanced cycle can be moved to a balanced cycle via a sequence of balanced reroutings of closed walks.

We may now describe the homomorphism $\pi_1(K) \to \Gamma$, where $K$ is the 2-cell complex constructed from $(G,\mathcal{B})$ and $(G,\mathcal{B})$ is $\Gamma$-labellable. Let $\phi : E(G) \to \Gamma$ with $\mathcal{B}_\phi = \mathcal{B}$. The fundamental group $\pi_1(K)$ of $K$ is constructed in our proof of Theorem 1 with presentation in terms of generators $g_e$, one for each edge $e$ not in a chosen spanning tree of $G$, and relations among these generators given by simple closed walks around the cycles in $\mathcal{B}$. Each edge in the spanning tree is labelled by the group identity 1, and for each cycle in $\mathcal{B}$ we add a relation corresponding to the product of the generators labelling the edges as they are traversed by a simple closed walk around the cycle (as described by Irene in her original post; of course, details may be found in our paper [DFP14]). Let us denote this $\pi_1(K)$-labelling by $\rho : E(G) \to \pi_1(K)$. For an element $g \in \pi_1(K)$ expressed as the word $g = g_{e_1}^{a_1} g_{e_2}^{a_2} \cdots g_{e_m}^{a_m}$ for some generators $g_{e_1}, g_{e_2}, \ldots, g_{e_m}$ and integers $a_1, a_2, \ldots, a_m$, define $\eta : \pi_1(K) \to \Gamma$ by
\[ \eta(g) = \eta \left( g_{e_1}^{a_1} g_{e_2}^{a_2} \cdots g_{e_m}^{a_m} \right) = \phi \left( \rho^{-1}(g_{e_1})^{a_1} \rho^{-1}(g_{e_2})^{a_2} \cdots \rho^{-1}(g_{e_m})^{a_m} \right) .\]

Theorem 2.The map $\eta:\pi_1(K) \to \Gamma$ is a group homomorphism.

Proof. Clearly $\eta$ is a homomorphism if it is well-defined. So suppose $g=$ $g_{n_1}^{a_{n_1}} g_{n_2}^{a_{n_2}} \cdots g_{n_m}^{a_{m_n}}$ $=$ $g_{k_1}^{a_{k_1}} g_{k_2}^{a_{k_2}} \cdots g_{k_p}^{a_{k_p}}$ are two words expressing the element $g \in \pi_1(K)$, where $g_{n_1}$, $g_{n_2}$, $\ldots$, $g_{n_m}$, $g_{k_1}$, $g_{k_2}$, $\ldots$, $g_{k_p}$ are generators and $a_{n_1}, a_{n_2}, \ldots$, $a_{n_m}, a_{k_1}, a_{k_2}, \ldots, a_{k_p}$ are integers. We show that
\[ \eta\left(g_{n_1}^{a_{n_1}} g_{n_2}^{a_{n_2}} \cdots g_{n_m}^{a_{m_n}} ( g_{k_1}^{a_{k_1}} g_{k_2}^{a_2} \cdots g_{k_p}^{a_{k_p}})^{-1}\right)=1,\]
and so that $\eta$ maps both words $g_{n_1}^{a_{n_1}} g_{n_2}^{a_{n_2}} \cdots g_{n_m}^{a_{m_n}}$ and $g_{k_1}^{a_{k_1}} g_{k_2}^{a_2} \cdots g_{k_p}^{a_{k_p}}$ to the same element of $\Gamma$. We have
\begin{multline*}
\eta \left(g_{n_1}^{a_{n_1}} g_{n_2}^{a_{n_2}} \cdots g_{m_n}^{a_{m_n}} ( g_{k_1}^{a_{k_1}} g_{k_2}^{a_{k_2}} \cdots g_{k_p}^{a_{k_p}} )^{-1}\right) \\
= \phi \left( \rho^{-1}(g_{n_1})^{a_{n_1}} \rho^{-1}(g_{n_2})^{a_{n_2}} \cdots \rho^{-1}(g_{n_m})^{a_{n_m}} \rho^{-1}(g_{k_p})^{-a_{k_p}} \cdots \rho^{-1}(g_{k_2})^{-a_{k_2}} \rho^{-1}(g_{k_1})^{-a_{k_1}} \right) \\
= \phi\left( \rho^{-1}(g_{n_1})\right)^{a_{n_1}} \phi\left(\rho^{-1}(g_{n_2})\right)^{a_{n_2}} \cdots \phi\left(\rho^{-1}(g_{n_m})\right)^{a_{n_m}} \phi \left(\rho^{-1}(g_{k_p})\right)^{-a_{k_p}} \cdots \\ \phi\left(\rho^{-1}(g_{k_2})\right)^{-a_{k_2}} \phi\left(\rho^{-1}(g_{k_1})\right)^{-a_{k_1}}.
\end{multline*}
In $\pi_1(K)$, $g_{n_1}^{a_{n_1}} g_{n_2}^{a_{n_2}} \cdots g_{m_n}^{a_{m_n}} g_{k_p}^{-a_{k_p}} \cdots g_{k_2}^{-a_{k_2}} g_{k_1}^{-a_{k_1}} = 1$; the relations in $\pi_1(K)$ that reduce this word to the identity correspond to a sequence of reroutings via balanced cycles. The word in $\Gamma$ given by $\eta$ has the same corresponding edge sequence. Since $\mathcal{B}_\phi = \mathcal{B}$, this word is therefore reduced to the identity by the same sequence of balanced reroutings. $\square$

Let’s now use this homomorphism to answer Question 1.

Theorem 3. There exists a group-labellable biased graph whose collection of balanced cycles is not realised by any labelling with any finite group.

Proof. Let $H$ be the Higman group
\[ H = \left< a, b, c, d \mid a^{-1} b a = b^2, b^{-1} c b = c^2, c^{-1} d c = d^2, d^{-1} a d = a^2 \right>. \]
The Higman group is infinite with no non-trivial finite quotients [H51] (the Wikipedia entry has a brief description). Construct a simplicial 2-complex $\mathcal{K}$ by identifying the points of edges marked $a$, $b$, $c$, and $d$, respectively, of five pentagons, oriented as shown in the figure.

ConstructingHigman1

Let $G$ be the graph consisting of the vertices and edges of the barycentric subdivision of $\mathcal{K}$ (the following figure shows the part of the graph obtained from the first pentagon above), and let $\mathcal{B}$ be the set of cycles of $G$ that are contractible in $\mathcal{K}$.

ConstructingHigman2

By construction, the fundamental group $\pi_1(\mathcal{K})$ is $H$. Hence the biased graph $(G,\mathcal{B})$ is $H$-labellable. Let $\Gamma$ be a group for which there is a labelling $\gamma : E(G) \to \Gamma$ realising $\mathcal{B}$. By Theorem 2, there is a homomorphism from $H$ to $\Gamma$. Since $H$ has no non-trivial finite quotient, and $(G,\mathcal{B})$ is not labellable by the trivial group, $\Gamma$ cannot be finite. $\square$

The graph $G$ constructed in the proof of Theorem 3 is pretty; it is shown below.

A Higman-labelled biased graph

References

[H51] Graham Higman. A finitely generated infinite simple group. J. London Math. Soc., 26:61-64, 1951.

[DFP14] DeVos, Matthew; Funk, Daryl; and Pivotto, Irene. When does a biased graph come from a group labelling? Advances in Applied Mathematics, Vol 61 (October 2014), pp 1-18.

Non-unimodality of 2-polymatroids

Guest post by Thomas Savitsky

A $k$-polymatroid is a generalization of a matroid in which the rank of an element is allowed to exceed one, but cannot exceed $k$.

Definition 1.
Let $S$ be a finite set. Suppose $\rho : 2^S \to \mathbb{N}$ satisfies the following four conditions:

  • if $X, Y \subseteq S$, then $\rho(X \cap Y) + \rho(X \cup Y) \le \rho(X) + \rho(Y)$
    (submodular);
  • if $X \subseteq Y \subseteq S$, then $\rho(X) \le \rho(Y)$ (monotone);
  • $\rho(\varnothing) = 0$ (normalized); and
  • $\rho(\{x\}) \le k$ for all $x \in S$.

Then $(\rho, S)$ is a $k$-polymatroid with rank function $\rho$ and ground set $S$.

So a matroid is a $1$-polymatroid. Here are a few examples of $2$-polymatroids.

  • If $(r_1, S)$, and $(r_2, S)$ are matroids, then $(r_1+r_2, S)$ is a $2$-polymatroid.
  • If $G = (V, E)$ is a graph, one may define a $2$-polymatroid $(\rho, E)$, where
    $\rho(X)$ equals the number of vertices incident to $X$.
  • Given an $m \times 2n$ matrix with entries in a field, one may define a representable $2$-polymatroid on $n$ elements by pairing up the columns in a obvious manner.

We became interested in $k$-polymatroids and thought it would be practical to have a catalog of the small ones at our disposal. We successfully adapted the canonical deletion
approach used by Mayhew and Royle (see [MR08]) to catalog matroids on nine elements to $2$-polymatroids. This first required developing a theory of single-element extensions of $k$-polymatroids. We then wrote code in the C programming language and interfaced with Brendan McKay’s nauty program and the igraph graph library. After a few days of execution time on a desktop computer, our program produced a catalog of all $2$-polymatroids, up to isomorphism, on at most seven elements.

By consulting our catalog, we produced Table 1, which lists the number of unlabeled $2$-polymatroids on the ground set $\{1, \ldots, n\}$ by rank.

Table 1: The number of unlabeled $2$-polymatroids.
rank $\backslash$ $n$ 0 1 2 3 4 5 6 7
0 1 1 1 1 1 1 1 1
1 1 2 3 4 5 6 7
2 1 4 10 21 39 68 112
3 2 12 49 172 573 1890
4 1 10 78 584 5236 72205
5 3 49 778 18033 971573
6 1 21 584 46661 149636721
7 4 172 18033 19498369
8 1 39 5236 149636721
9 5 573 971573
10 1 68 72205
11 6 1890
12 1 112
13 7
14 1
total 1 3 10 40 228 2380 94495 320863387

 

Surprisingly, the number of $2$-polymatroids on seven elements is not unimodal in rank. In contrast, matroids are conjectured to be unimodal in rank, and the catalog of matroids with nine elements supports this. By the way, the symmetry in the columns in Table 1 is accounted for by a notion of duality for $2$-polymatroids.

Note that one can obtain the analogue of Table 1 for labeled $2$-polymatroids by computing the automorphism group of each $2$-polymatroid and then using the Orbit-Stabilizer relation. This allowed us to confirm the results of our enumeration through another means. By interpreting a $2$-polymatroid as a solution to a certain integer programming program, the number of labeled $2$-polymatroids can theoretically be computed by integer programming software. Fortunately, the software package SCIP was up to the task when $n \le 7$.

See [Sa14] for more details on all of the above.

Now recall that a matroid $M$ is paving if it contains no circuit of size less than $r(M)$. If both $M$ and $M^{*}$ are paving, then $M$ is sparse-paving. If $M$ is sparse-paving, then one can show that that every set of size less than $r(M)$ is independent and that the dependent $r(M)$-subsets are circuit-hyperplanes; furthermore, the symmetric difference of any two circuit-hyperplanes must be at least $4$. In fact, sparse-paving matroids are characterized by these properties.

It is conjectured that almost all matroids are sparse-paving.

The ideas in the remainder of this post were communicated to me by
Rudi Pendavingh.

We first mention the following background item. Let $S = \{e_1, e_2, \dots, e_n\} \cup \{f_1, f_2, \dots, f_n\}$ be a set of size $2n$. Suppose $(r, S)$ is a matroid. We will pair up the elements of $S$ to define a $2$-polymatroid as follows. Define $S’ = \big\{\{e_1, f_1\}, \{e_2, f_2\}, \dots, \{e_n, f_n\}\big\}$, and define $\rho : S’ \to \mathbb{N}$ by
$$\rho\big(\big\{\{e_{i_1}, f_{i_1}\}, \{e_{i_2}, f_{i_2}\}, \dots, \{e_{i_m}, f_{i_m}\}\big\}\big) = r(\{e_{i_1}, f_{i_1}, e_{i_2}, f_{i_2},\dots, e_{i_m}, f_{i_m}\}).$$
Then $(\rho, S’)$ is a $2$-polymatroid on $n$ elements with $\rho(S’) = r(S)$. Furthermore, every $2$-polymatroid on $n$ elements may be obtained in this manner from a matroid on $2n$ elements. See Section 44.6b of Schrijver’s Combinatorial Optimization or Theorem 11.1.9 of Oxley’s Matroid Theory for details.

Now assume that $r$ is a sparse-paving matroid. If $r(S)$ is odd, then $\rho$ does not detect any of the circuit-hyperplanes of $r$; namely,
\begin{equation*}
\rho(X) =
\begin{cases}
2|X| & \text{if} \ 2|X| < r(S),\\ r(S) & \text{if} \ 2|X| > r(S).\\
\end{cases}
\end{equation*}
To illustrate, all the rank-$7$ sparse-paving matroids on 14 elements map, in this manner, to a single rank-$7$ $2$-polymatroid on seven elements. However, if $r(S)$ is even, then the circuit-hyperplanes of $r$ are picked up by $\rho$. Perhaps this observation, combined with the conjecture that almost all matroids are sparse-paving, makes the non-unimodality of $2$-polymatroids appear more reasonable.

References

[MR08] Dillon Mayhew and Gordon F. Royle.
Matroids with nine elements.
J. Combin. Theory Ser. B, 98(2):415–431, 2008.
doi

[Sa14] Thomas J. Savitsky.
Enumeration of 2-polymatroids on up to seven elements.
SIAM J. Discrete Math., 28(4):1641–1650, 2014.
doi