Algebraic invariants of graphs; a study based on computer exploration
Nicolas M. Thiéry
Not printed
Abstract:
We consider the ring In of polynomial invariants over weighted
graphs on n vertices. Our primary interest is the use of this ring
to define and explore algebraic versions of isomorphism problems of
graphs, such as Ulam's reconstruction conjecture.
There is a huge body of literature on invariant theory which
provides both general results and algorithms. However, there is a
combinatorial explosion in the computations involved and, to our
knowledge, the ring In has only been completely described for
n≤4.
This led us to study the ring In in its own right. We used
intensive computer exploration for small n, and developed
PerMuVAR , a library for MuPAD , for computing in invariant rings of
permutation groups.
We present general properties of the ring In, as well as results
obtained by computer exploration for small n, including the
construction of a medium sized generating set for I5. We
address several conjectures suggested by those results (low degree
system of parameters, unimodality), for In as well as for more
general invariant rings. We also show that some particular sets are
not generating, disproving a conjecture of Pouzet related to
reconstruction, as well as a lemma of Grigoriev on the invariant
ring over digraphs. We finally provide a very simple minimal
generating set of the field of invariants.
Introduction
Let K be a field of characteristic zero, n be a positive integer,
and {x{1,2}, ..., x{n-1,n}} be a set of ()
variables indexed by the pairs {i,j} of {1,...,n}. The symmetric
group Sn acts naturally on those variables by
Let K[x{i,j}] be the ring of polynomials in x{i,j}. We study the subring
In:=K[x{i,j}] Sn of the polynomials which remain invariant under
the action of Sn.
Our motivation comes from graph theory and in particular from graph
reconstruction. Pouzet [21, 22] formulated an
algebraic reconstruction conjecture for In, which implies Ulam's
famous reconstruction conjecture for weighted
graphs [2]. We disprove Pouzet's conjecture. Kocay
proposed a similar conjecture, by introducing the algebra of
subgraphs [19, 3]. This algebra is a
quotient of In; however this quotient is not graded, and we cannot
apply our method to disprove Kocay's conjecture.
The ring In can also be used to study the shape of sets of
vectors [1]. Our primary goal is to construct
complete systems of invariants (systems that separate weighted graphs
up to isomorphism), and in particular minimal generating sets of
In.
In § 1, we introduce the representation Gn of the
symmetric group Sn over the vector space Vn of weighted graphs
on n vertices, and the associated invariant ring In. We review
classical results and tools provided by invariant theory (finite
generation, grading, Hilbert series). Since Gn is a permutation
group, there is a combinatorial interpretation of the invariant ring,
and a reasonably fast algorithm for computing the Hilbert series. We
also review some general properties of minimal generating sets, and
the definition of the smallest degree bound β( In).
§ 2 is devoted to generating sets of In. We
provide a finite generating set. By studying the Hilbert series, we
show that two other sets are not generating, disproving Pouzet's
conjecture. We also prove that, for many common monomial orders,
In has no finite SAGBI basis.
Finding a good degree bound is crucial. In § 3, we
recall how Hironaka decompositions of In can be used to obtain the
degree bound β( In)≤()-µn, where µn
is a non-negative O(n) integer. We calculate µn by constructing
minimal multigraphs without odd automorphisms.
In § 4, we try to refine the degree bound by
constructing low degree systems of parameters. The study of the
Hilbert series for n≤21, combined with a conjecture of Mallows
and Sloane, suggests the existence of a system of parameters composed
of invariants of degrees 1,2,...,n, 2,3,..., (). This
would give β( In)≤()-µn. We
propose a natural construction for such a low degree system of
parameters, and check its validity for n≤5 using a Gröbner basis
computation. Unfortunately, this computation is intractable for
n≥6. Such a system of parameters seems to have nearly optimally
low degrees; therefore, this technique cannot be refined much further
in order to get better degree bounds.
§ 5 is devoted to the computation of minimal generating sets.
For n=4, a minimal generating set was first constructed by hand by
Aslaksen, Chan et Gulliksen [1]; it can now be
computed in a few seconds by invariant theory software (e.g. Kemper's
packages in Maple [16] or Magma [17]).
However, for n≥5, these software packages are unable to compute
even partial minimal generating sets. We wrote
PerMuVAR [32], a library of invariant theory routines
for MuPAD , which uses the usual
algorithms [29, 17], but is specialized for
permutation groups. This allows us to go a step further: for n=5, we
compute a partial minimal generating set, containing 57 polynomials
of degree ≤ 91. This suggests a much better degree bound:
β( In)=() -1.
In § 6, we prove that the invariant ring In is
Gorenstein when n is even. This fact could be used to accelerate the
computations of Hironaka decompositions [32].
We introduce in § 7 the chain product (a naive
interpretation of Stanley-Reisner rings [11]).
This allows for faster computations of generating sets at the expense
of non-minimality [32]: we obtain a generating set of
I5 containing about one thousand polynomials of degree ≤22.
In § 8 the projective limit I∞ is used to obtain
results about In; this includes the lower bound β( In)≥
⌊n/2⌋.
§ 9 presents various unimodality properties revealed
by computer exploration, for In as well as for more general
invariant rings.
Grigoriev [14] introduces a related invariant ring
over digraphs. In § 10, we apply the Hilbert series tool
of § 2 to disprove lemma 1 of
[14]. We also provide a simple counter-example.
Finally, in § 11, we study the field of
invariants. Grigoriev [14] gives a non-constructive
proof for the existence of a small generating set of the field of
invariants (the proof of the degree bound is incorrect though, since
it relies on lemma 1 of [14]). We construct such a
small generating set, composed of the elementary symmetric polynomials
and a very simple invariant of degree 2; to the contrary of
Grigoriev's assertion, it is not a complete system of invariants. We
also derive a minimality property of the invariant ring, by using
basic Galois theory on the field of invariants.
The results presented in this paper are part of the Ph. D.
thesis [30] of the author. We refer to this document for
the detailed proofs.
1 The invariant ring over graphs
1.1 Valuated graphs as a vector space
Let V be a K-vector space of finite dimension m, and G be a
finite subgroup of GL(V). Tacitly, we interpret G as a group of
m× m matrices or as a representation on V. Two vectors v
and w are isomorphic, or in the same G-orbit (for
short orbit), if σ⋅v=w for some σ∈ G.
Let n be a positive integer. We consider labelled, undirected graphs
on the vertices {1,...,n}, without loops, and whose edges are
weighted in K. A simple graph is a graph with weights in
{0,1}, and a multigraph is a graph with weights in N.
For any pair {i,j}, let e{i,j} be the simple graph with one single
edge {i,j}. The set of all graphs is a K-vector space Vn of
dimension m:=() with basis
{e{1,2},...,e{n-1,n}}. Indeed, any graph g can be
written uniquely as g:=∑ g{i,j}e{i,j}, where g{i,j} is the
weight of the edge {i,j}. Let {x{1,2},...,x{n-1,n}} be
the dual basis (x{i,j}(g) is the weight g{i,j} of the graph g on
the edge {i,j}).
Throughout the text, we denote objects attached to Vn by cursive
symbols, and objects attached to the generic vector space V by
ordinary symbols. Let Sn be the symmetric group of all
permutations of the n vertices. Our group Gn is the linear
representation of Sn defined on the basis of Vn by
σ⋅e{i,j}:=e{σ(i),σ(j)}. The notion of
isomorphism defined above coincides with the usual notion of
isomorphism of graphs. Orbits of labelled graphs are called
unlabelled graphs. Unless otherwise stated, all graphs are
unlabelled.
The representation of Gn on Vn splits into three irreducible
components: [n]⊕[n-1,1]⊕[n-2,2], where [n-2,2]
represents the irreducible representation of Sn indexed by the
partition λ=(n-2,2) of n
[1, 10]. The first component has
dimension 1 and corresponds to the vector space spanned by the
complete graph. The sum [n]⊕[n-1,1] of the first two components
is of dimension n, and corresponds to the vector space spanned by
the n stars E1,...,En, where Ei:=∑j≢ i
e{i,j}.
This representation is the natural representation of Sn by
permutation of E1,...,En. Let X1,...,Xn be the basis of
the dual, defined by Xi:=∑j≢ i x{i,j}. If g is a graph,
Xi(g) is the degree of the vertex i of g. Finally, the last
irreducible component [n-2,2] is the orthogonal of the two previous
components, that is the subspace of all 0-regular graphs
(graphs where each vertex as degree 0).
1.2 The invariant ring
Recall that if G acts on V, a complete system of invariants
is a set S of functions such that two elements v and w of V
are in the same orbit if and only if they give the same value to all
functions in S (i.e. p(v)=p(w) for all p∈ S). Our primary
goal is to construct, or at least find information about, complete
systems of invariants. We introduce the invariant ring of G which
provides a mechanical way to do this. We refer
to [27, 29, 5, 26, 17]
for classical literature on invariant theory of finite groups. Parts
of what follows are strongly inspired by [17].
Let (x1,...,xm) be a basis of the dual of V; for Vn, we take
(x1,...,xm):=(x{1,2},...,x{n-1,n}). Let
K[x1,...,xm] be the ring of polynomials over V. The action of
G on V extends naturally to an action of G on K[x1,...,xm]
by σ⋅ p:=p∘ σ-1. An invariant polynomial, or
invariant, is a polynomial p∈ K[x1,...,xm] such that
σ⋅ p=p for all σ∈ G. The invariant ring I(G) is
the set of all invariants. We call In:=I( Gn) the invariant
ring over graphs. Note that
σ⋅ x{i,j}:=x{i,j}∘σ-1=x{σ(i),σ(j)}.
Obviously, I(G) is a K-algebra. Hilbert's famous theorem states
that I(G) is finitely generated: there exists a finite set
S of invariants such that any invariant can be expressed as a
polynomial combination of invariants in S. We call S a
generating set. If no proper subset of S is generating, S
is a minimal generating set. Since I(G) is finitely
generated, there exists a degree bound d such that I(G) is
generated by the set of all invariants of degree at most d. We
denote by β(I(G)) the smallest degree bound.
There exist algorithms to compute (minimal) generating sets, and a
basic result of invariant theory states that they are complete systems
of invariants. However, this often leads to very intensive
computations, and rather large complete systems of invariants.
1.3 Invariant ring of a permutation group
The most famous invariant ring is the ring of symmetric polynomials
I( Sm), defined by the natural action of Sm on the
variables (x1,...,xm). The fundamental theorem of symmetric
polynomials [29, p. 2] states that I( Sm) is
generated by m algebraically independent symmetric polynomials, for
example the m elementary symmetric polynomials or the first m
symmetric power sums.
Gn is a permutation group, since it acts by permuting the
variables x{i,j}; thus, we also view Gn as a subgroup of the full
group Sm of the permutations of m variables. This results in
several convenient and powerful combinatorial interpretations of
invariants:
(i) A labelled multigraph g:=(g{1,2},...,g{n-1,n}) can
be identified with the monomial
xg:=x{1,2}g{1,2}...
x{n-1,n}g{n-1,n}. The exponential of g is the
polynomial xgˆ*:=∑hxh, where h belongs to the orbit
of g. The polynomial xgˆ* is invariant, and is well defined
even if g is unlabelled. The exponential therefore identifies
unlabelled graphs with some particular invariants. Moreover, the set
of all invariants xgˆ*, where g is a multigraph, is a vector
space basis of In. Note that the exponential differs from the
usual Reynolds operator * by a multiplicative factor:
xgˆ*=|Aut(g)|(xg)*, where |Aut(g)| is the size of
the automorphism group of g.
(ii) Let g1 and g2 be two multigraphs on n vertices. The
product xg1ˆ*xg2ˆ* is a linear combination of all
possible superpositions of g1 and g2 (with, at times,
counter-intuitive coefficients). For instance:
|
( |
 |
) |
|
|
× |
( |
 |
) |
|
|
=
|
( |
 |
) |
|
|
+ |
( |
 |
) |
|
|
+ |
( |
 |
) |
|
|
.
|
Let n'>n, and consider the multigraphs g1 and
g2 obtained by adding n'-n isolated vertices to g1
and g2. New superpositions, which fit in n' vertices but not in
n vertices, may appear in the product
xg1ˆ*xg2ˆ*. However, the use of a
modified Reynolds operator ensures that the coefficients in the linear
combination do not change. This makes the product somewhat independent
of n (see § 8).
(iii) If g and h are simple graphs, xgˆ*(h) counts the
number s(g,h) of subgraphs of h isomorphic to g. The
following invariants can be used to count respectively the number of
edges, the number of pairs of adjacent edges and the number of
Hamiltonian cycles of h:
( |
 |
) |
|
|
(h), |
|
|
( |
 |
) |
|
|
(h), |
|
|
( |
 |
) |
|
|
(h). |
|
Manipulations of the quantities s(g,h) are the cornerstone of
several results on reconstruction of graphs [2];
for the use of these algebraic considerations
see [23].
1.4 Grading, Hilbert series and degree bound
Powerful properties of an invariant ring are its grading and the
associated Hilbert series. As a K-vector space, I(G) is not
finite dimensional. However, since the action of G preserves the
degree of polynomials, I(G) decomposes into the direct sum of its
homogeneous components:
where I(G)d is the finite dimensional vector space of all
homogeneous invariants of degree d. The Hilbert series of I(G) is
the generating series of its dimensions:
H(I(G),z):= |
|
zd dim I(G)d.
|
For general finite groups of matrices, this series can be computed by
averaging over the group, through Molien's formula. However, since
Gn is a permutation group, the set of all invariants xgˆ*,
where g is a multigraph with d edges is a vector space basis of
In,d. Therefore, computing H( In,z) reduces to a Pólya
enumeration of multigraphs with respect to the number of
edges [27]. Recall that the conjugacy classes
Cλ of Sn are indexed by the partitions λ of
n. Let σ be a permutation of the vertices in Cλ. The
cycle type of the induced permutation of the edges is easily computed
from the cycle type of σ, i.e. from the partition
λ [15]. We denote by
l1(λ),...,lm(λ) this cycle type. Then,
where the sum is over all partitions λ of n. This provides
an algorithm whose complexity is about O(n4exp(n0.8)). Concretely,
we can compute H( In,z) for n≤21. It is sometimes useful to
consider the multigraded Hilbert series, where each grading
corresponds to one of the three irreducible components of the
representation Gn. We can compute this multigraded Hilbert series
for n≤15.
Given an integer d≥1, let K[I(G)<d] be the subalgebra of
I(G) generated by invariants of degree <d, and K[I(G)<d]d
its homogeneous component of degree d. Set s0(I(G)):=0 and
sd(I(G)):=dim I(G)d - dim K[I(G)<d]d. The generating
series s(I(G),z):=∑d=0∞zd sd(I(G)) is a
polynomial of degree β(I(G)).
A set S is homogeneous if its elements are also homogeneous.
The following lemma (valid for any graded algebra A, where A0 is
the ground field K) summarizes some general properties of
generating sets.
Lemma 1.1 Let S be a generating set of I(G).
(i) I(G) has a homogeneous minimal generating set composed of at
most |S|β(I(G)) invariants of degree at most β(I(G)).
(ii) Assume S is homogeneous, and let Sd be the set of all
invariants of S having degree d. Then, S is a minimal
homogeneous generating set if and only if for all d, Sd is a
vector space basis of a direct factor of K[I(G)<d]d in
I(G)d. In particular, |Sd|=sd(I(G)).
Proof:
(i) For each p∈ S and d, let pd be the homogeneous
component of degree d of p. Since I(G) is graded, it is
generated by the set {pd | p∈ S, 1≤ d≤β(I(G))}.
(ii) Use the grading and basic linear algebra.
QED
From (i), it is not very restrictive to consider only homogeneous
generating sets, since non-homogeneous generating sets are not much
smaller than homogeneous ones.
The Hilbert series provides a simple necessary condition to test if a
set S of homogeneous invariants is generating. The following
proposition is valid for any graded algebra A, where A0 is the
ground field K. We stress the importance of the homogeneity of the
invariants. A series s(z) is dominated by a series t(z) if
the coefficients of s(z) are upper-bounded term by term by the
coefficients of t(z).
Condition 1.2
Let S:=(p1,...,pt) be a homogeneous generating set, with
respective degrees (d1,...,dt). Then, the Hilbert series
H(I(G),z) is dominated by the series
Proof:
As a vector space, the homogeneous component I(G)d is generated
by the set of all the homogeneous products p1λ1...
ptλt whose degree d1λ1+...+dtλt is
d; those products are counted by the series F(d1,...,dt,z).
QED
This apparently weak condition is in fact very powerful. In
particular, it leads to the proof of theorem 2.3, and to
the disproving of Grigoriev's lemma 1 of [14] (see
§ 10).
2 Generating sets of In
For n=1,2,3, the invariant ring is the ring of symmetric
polynomials; the elementary symmetric polynomials form a minimal
generating set. For n=4, Aslaksen et al. [1]
constructed by hand the following minimal generating set:
ì í î |
|
( |
 |
) |
|
|
, |
( |
 |
) |
|
|
,
|
( |
 |
) |
|
|
,
|
( |
 |
) |
|
|
, |
( |
 |
) |
|
|
,
|
|
|
( |
 |
) |
|
|
,
|
( |
 |
) |
|
|
,
|
( |
 |
) |
|
|
, |
( |
 |
) |
|
|
|
ü ý þ |
. |
|
At about the same time, we had proven independently a similar result
through a Gröbner basis computation with CoCoA [4], using
theorem 2.7.9 of [29]. However, our set was not
minimal since we had not removed the invariant xcˆ* where c is
the complete graph. The set above can now be computed in about one
second with Kemper's implementation of Invar in
Magma [17].
We now provide a large, but reasonable, finite generating set of
In. A multigraph is quasi-connected if it
has, at most, one non-trivial connected component. For example, g1
below is quasi-connected, but not g2:
g1:= , |
|
g2:= . |
Proposition 2.1
(i) The homogeneous component In,d has for vector space basis
the set of all invariants xc1ˆ*⋯xckˆ*, where each
ci is a quasi-connected multigraph with ni non-isolated
vertices and di edges, and where n1+...+nk≤ n and
d1+...+dk=d.
(ii) The invariant ring In is generated by the set of all
homogeneous invariants xgˆ*, where g is a quasi-connected
multigraph with at most β( In) edges.
Proof:
(i) Let g be a multigraph with n vertices and k>1 non-trivial
connected components c1,...,ck. Let
c1,...,c1 be the quasi-connected
multigraphs on n vertices obtained by adding isolated vertices to
the ci. Obviously, n1+...+nk≤ n, and d1+...+dk=d.
Then:
where the hi are multigraphs with strictly less than k
non-trivial connected components. For example:
|
( |
 |
) |
|
|
=
|
( |
 |
) |
|
|
× |
( |
 |
) |
|
|
- |
( |
 |
) |
|
|
- |
( |
 |
) |
|
|
.
|
By induction on the number k of non-trivial connected components,
g is a linear combination of products
xc1ˆ*⋯xckˆ*. In fact, we just
inverted a triangular linear system with ones on the diagonal, and
uniqueness follows.
(ii) Use (i) and the definition of β( In).
QED
Obviously, in order to get a usable generating set, it is essential to
have a good bound for β( In).
We tried to use the technique of SAGBI basis. This is a powerful tool,
which generalizes Gröbner basis techniques for rings instead of ideals
[24]. The main drawback is that there exist
invariant rings with no finite SAGBI basis; this seems to be the case
for In, at least for many common monomial orders.
Theorem 2.2
There are no finite SAGBI basis for In if the monomial order is
either lexicographic, degree lexicographic, or degree reverse
lexicographic with the n-1 smallest variables corresponding to
adjacent edges.
Proof:
We prove in each case that there is an infinite number of
irreducible initial monomials (an initial monomial is irreducible if
it cannot be written as product of two smaller initial monomials).
For the lexicographic order, we can alternatively use Göbel's
characterization of permutation groups with finite SAGBI
basis [13].
QED
The following theorem states that some sets are not generating. (i)
disproves a tempting generalization of the fundamental theorem of
symmetric functions, whereas (ii) disproves Pouzet's
conjecture [21], which would have implied Ulam's
reconstruction conjecture.
Theorem 2.3
(i) For n≥5, the set of all invariants xgˆ*, where g is
a simple graph, do not generate In.
(ii) For 11≤ n≤18, the set of all invariants xgˆ*, where
g is a multigraph with at least one isolated vertex, do not
generate In.
Proof:
(i) For n=5,6,7,8, simple graphs can be counted with respect to
the number of edges using Pólya enumeration [15].
The coefficient of degree d=4 of the series S(d1,..., dt) is
strictly smaller than that of the Hilbert series. Therefore,
condition 1.2 applies. For n≥9, no new
isomorphism types of multigraphs with less than 4 edges appears,
so the coefficient of degree 4 of both series is the same as for
n=8. Condition 1.2 again applies.
(ii) By an argument similar to the proof of
proposition 2.1, we have only to consider
the set of all invariants xgˆ*, where g is a multigraph with
a unique non-trivial connected component, which is of size <n.
Those multigraphs can be counted from the total number of
multigraphs by using a technique similar to that described
in [15, § 4.2, p. 90]. For 11≤ n≤18,
computations of both series shows that
condition 1.2 fails.
QED
We could not check (ii) for n>18 since the computation were
intractable. However, the results for n≤ 18 strongly suggest that,
for d≈ 4n-24, the ratio between the coefficients of degree d
of the two series is bounded by an expression of the form
exp(-an)+0.17. This could probably be confirmed by an asymptotic
study, and we conjecture that (ii) is true for any n≥11.
3 Decomposition of Hironaka
The smallest degree bound β( In) and furthermore the polynomial
s( In,z) contains important information about the invariant
ring, which would be very useful when the computation of a minimal
generating set is intractable. But so far, we don't know how to
calculate them except by explicitly computing such a minimal
generating set.
Invariant theory provides only some bounds on β( In) and
s( In,z) [25, 6]. Noether's
theorem [29, p. 27] yields: β( In)≤ | Gn|=n!,
which is not very informative. A much better bound exists for
permutation groups: β( In)≤ ()=() [11]. However, our computations for
small n shows that this is still a rather loose bound. This section
introduces the tools that produce this bound, and possibly even better
bounds.
A set of m homogeneous invariants (θ1,...,θm) of
I(G) is called a homogeneous system of parameters or, for
short, a system of parameters if the invariant ring I(G) is
finitely generated over its subring K[θ1,...,θm].
That is, if there exist a finite number of invariants
(η1,...,ηt) such that the invariant ring is the sum of the
subspaces ηi.K[θ1,...,θm]. By Noether's
normalization lemma, there always exists a system of parameters for
I(G). Moreover, I(G) is Cohen-Macaulay, which means that
I(G) is a free-module over any system of parameters. So, if the set
(η1,...,ηt) is minimal for inclusion, I(G) decomposes
into a direct sum:
I(G)=⊕i=1t ηi . K[θ1,...,θm].
This decomposition is called a Hironaka decomposition of the
invariant ring. The θi are called primary invariants,
and the ηi secondary invariants (in algebraic
combinatorics literature, the θi are some times called
quasi-generators and the ηi
separators [11]). It should be emphasized
that primary and secondary invariants are not uniquely determined, and
that being a primary or secondary invariant is not an intrinsic
property of an invariant p, but rather express the role of p in a
particular generating set.
The primary and secondary invariants together form a generating
set. From the degrees (d1,...,dm) of the primary invariants
(θ1,...,θm) and the Hilbert series we can compute the
number t and the degrees (e1,...,et) of the secondary
invariants (η1,...,ηt) by the formula:
|
z |
|
+...+z |
|
=(1-z |
|
)⋯(1-z |
|
) H(I(G),z).
(1) |
Assuming d1≤...≤ dm and e1≤...≤ et, it can be
proved that:
|
t = |
|
,et = d1+...+dm - m - µ,β(I(G))≤ max(dm,et),
|
|
(2) |
where µ is the smallest degree of a polynomial p such that
σ⋅ p=det(σ) p for all σ∈ G [27, Proposition
3.8].
For example, if G is the symmetric group Sm, the m
elementary symmetric polynomial (or the m first symmetric power sums)
form a system of parameters, t=1, et=0 and η1=1. This is
consistent with the fundamental theorem of symmetric polynomials.
More generally, for Gn as well as for any permutation group, the
elementary symmetric polynomials still form a system of parameters.
This yields the following information on In.
Proposition 3.1
For n≥4, consider the system of parameters of In composed
of the elementary symmetric polynomials, and let (e1,...,et)
be the degrees of secondary invariants. Then,
|
et = |
æ è |
|
ö ø |
- µn = |
æ ç è |
|
ö ÷ ø |
- µn, |
|
|
where µn=0 if n is even, and µn=⌈3/4(n-1)⌉
otherwise.
For example, β( I4)≤15, β( I5)≤42 and
β( I6)≤104.
Proof:
We only have to check the value of µn. When n is even,
det(σ)=1 for all σ∈ Gn. Therefore, p:=1 verifies the
condition σ⋅ p=det(σ) p for all σ∈ Gn. We
note that this is generally true for any Gorenstein ring (see
§ 6 and [27, § 8]). When n is
odd, the sign of a permutation of the edges is the sign of the
corresponding permutation of the vertices. Then, the smallest degree
µn of a polynomial p such that σ⋅ p=signσ p
is the smallest number of edges of multigraph with no odd
automorphism. The following lemma completes the proof.
QED
Lemma 3.2
The smallest number of edges of a multigraph gn on n≥4
vertices without odd automorphism is ⌈3/4(n-1)⌉.
Proof:
Such multigraphs can be constructed for any n as follows:
-
g4:=
;
g5:=
;
g6:=
;
g7:=
;
- g4k is composed of k copies of g4 (3k edges);
- g4k+1 is composed of k copies of g4 and an isolated
vertex (3k edges);
- g4k+2 is composed of k-1 copies of g4 and one copy of
g6 (3k+1 edges);
- g4k+3 is composed of k-1 copies of g4 and one copy
of g7 (3k+2 edges).
The minimality of the number of edges of such multigraphs can be
proved by induction over n.
QED
So, the knowledge of a system of parameters and of the Hilbert series
provides both an upper bound on β( In), as well as bounds on
the coefficients of s( In,z). Unfortunately, our experience
has shown that generating sets composed of primary and secondary
invariants are far from minimal (see Figures 4
and 2), so those bounds are quite loose. Moreover, to our
knowledge, those bounds are the only obtainable information about a
minimal generating set, without actually computing it.
4 Low degrees systems of parameters
We now search for a low degrees system of parameters for In, in
order to improve the bound on β( In).
Equation (3) can guide our quest by suggesting
possible degrees. Indeed, there can only exist a system of parameters
of degrees (d1,...,dm) if the expression
(1-zd1)⋯(1-zdm) H(I(G),z) is a polynomial with positive
integer coefficients. It has even been conjectured by Mallows and
Sloane [20, 7] that the converse is
true: if (1-zd1)⋯(1-zdm) H(I(G),z) is a polynomial with
positive integer coefficient, then there exists a system of parameters
of degrees (d1,...,dm). A counter-example has been found, but
the conjecture still holds if the representation of G over V is
irreducible, or when using a multigraded Hilbert series (one grading
for each irreducible component) [7, p. 5].
By tweaking the Hilbert series for n≤21, and the multigraded
Hilbert series for n≤15, we find that the degree sequence
(1,...,n, 2,...,()) always produces a polynomial
with positive integer coefficient. We also proved that, for any n,
this degree sequence produces a polynomial. We are therefore somewhat
confident with the following conjecture:
Conjecture 4.1
For all n≥3, there exists a system of parameters for In of
degrees (1,...,n, 2,...,()). As a direct
consequence, β( In)≤ ()+()-µn.
For example, β( I4)≤9, β( I5)≤22 and
β( I6)≤60, which are much smaller degree bounds than those
provided by proposition 3.1.
Figure 1 displays the number of secondary
invariants depending on the system of parameters.
[System of parameters: symmetric power sums]
[Conjectured system of parameters]
Figure 1: Number of secondaries per degree
Next, we construct a reasonable system of parameters and check its
validity for n=3,4,5, which proves
conjecture 4.1 for those values. We note
that Dixmier [7] constructed a system of parameters
with degrees (2,3,4,5,6) for the representation [3,2] of S5,
and proved its validity by hand. By using the decomposition of the
representation G5 into [5]+[4,1]+[3,2], this also provides a
system of parameters with the expected degrees.
The form of the degree sequence suggests starting from the () first symmetric power sums, and replacing the last n-1 degrees
(()+1,... ()) by some invariants of degree
2,...,n. Moreover, since the representation splits into
[n]⊕[n-1,1] and [n-2,2] (see § 1.1), we
get a system of parameters for the invariant ring of the whole
representation, by taking systems of parameters for the invariant
rings of each components and putting them together. Recall that the
first component is the natural representation of Sn by
permutations of the stars (E1,...,En). So, the invariant ring
over this component is the ring of symmetric polynomials in the dual
variables (X1,...,Xn), and the n first symmetric power sums in
the Xi form a system of parameters for this component. Note that,
up to a constant 2, the first symmetric power sum in the Xi is
equal to the first symmetric power sum in the x{i,j}. All this leads
to the following conjecture:
Conjecture 4.2
If n≥3, the following system of invariants is a system of
parameters for In.
x |
|
+...+x |
|
, ...,
x |
|
+...+x |
|
, |
|
X1+...+Xn, ..., X1n+...+Xnn. |
Proposition 4.3
Conjecture 4.2 holds for 3≤ n ≤5.
This is immediate for n=3, since the component [n-2,2] is
trivial. To test the conjecture for other small cases, we used the
following general characterization:
Caracterisation 4.4 [[29]]
A set of m homogeneous invariants (θ1,...,θm) is a
system of parameters if and only if v=0 is the only common zero
of the θi:
θ1(v)=...=θm(v)=0 ⇒ v=0
For n=4, this characterization is enough to prove
conjecture 4.2 by hand.
For n≥5, we can try to check conjecture 4.2 as
follow: compute a Gröbner basis for the θi, and verify that
for each variable x{i,j} there is a polynomial in the Gröbner basis
whose leading term is of the form x{i,j}k
(see [29, Subroutine 2.5.2]; this is both a necessary
and sufficient condition for the radical of the ideal generated by the
θi to be the irrelevant ideal).
For n=3,4, the direct computation of this Gröbner basis takes less
than one second, but for n=5 it seems to be intractable and fails.
However, some equivalent Gröbner basis can be computed in about one
minute, by using a suitable linear change of basis which respects the
decomposition of the representation. The verification of
characterization 4.4 is then straightforward.
For n=6, even with the same linear change of basis and using
FGb [9], the computation is intractable. In fact,
the growth of the Gröbner basis seems to follow nearly the worst
possible theoretical case [12].
Finally, as a by-product of the computation of minimal generating sets
with PerMuVAR , we checked that for n≤8 and for the low-degree
homogeneous components, the ring of invariants is indeed a free-module
over K[θ1,...,θm]. This gives some confidence in
conjecture 4.2, but so far we are unable to prove it.
If the above construction of a system of parameters is correct, we
believe it to be nearly optimal: there exist no general construction
of systems of parameters with lower degrees. Indeed, we calculated,
for small values of n, the smallest degree sequence allowed by the
multigraded Hilbert series. For 3,4,5, the degrees conjectured above
are optimal. For n≥6, it was usually possible to divide some of
the degrees by 2 or 3. For example, for n=6 and 7 the best
degree sequences are respectively (1,...,6,
2,...,7,8/2,9/3,10/2) and (1,...,7,
2,...,7,8/2,9,...,13,14/2,15). We have not noticed any
regularity in these case by case optimizations. Therefore, we don't
think this technique can be refined much further in order to get
better degree bounds.
5 Computing minimal generating sets
Since we can not get more a priori information on homogeneous minimal
generating sets of In, we proceed with explicitly computing them.
The computations are very intensive (even for n=5) but give some
feeling of the size and degree of minimal generating sets. The bounds
given by Hironaka decompositions seem very loose.
The basic principle of the classical algorithms is to construct
generating sets degree by degree, from 1 up to the best degree bound
known. Since the complexity of the computations involved usually
increases quickly with increasing degree, the quality of the degree
bound is crucial. One can take advantage of the existence of a
Hironaka decomposition by computing secondary invariants and, while
doing so, selecting the secondary invariants that are
irreducible (i.e. that cannot be expressed as products of lower
degree secondary invariants). The irreducible secondary invariants
together with the primary invariants form a minimal generating set
(some primary invariants may need to be removed).
Most software [16] relies on a precomputation of a
Gröbner basis of the system of parameters to greatly speed up the
rest of the computations. However, with In this precomputation is
very hard, if not impossible (see § 4). Software that
do not rely on this precomputation uses linear algebra on the
homogeneous component of degree d of the whole ring of polynomials,
whose dimension grows quickly with increasing d, and fail early.
Since our group is a permutation group, invariants can be stored as
linear combinations of invariants xgˆ*. This saves a lot of
memory (up to a factor of 1/| Gn| for monomials without symmetries,
which happens to be the case for most of them). This data structure
also allows for the same linear algebra operations inside the
homogeneous component of degree d of the invariant ring which is
considerably smaller. We therefore implemented our own invariant
theory software PerMuVAR which takes advantage of the particular
properties of permutation groups [32]. We chose the
computer algebra system MuPAD , which is freely available (but alas
not open source software), and allows modularity through object
oriented programming. Moreover, MuPAD 's dynamic modules will allow
for rewriting critical sections in a very efficient language like
C++.
A sketch of the algorithm follows. We denote by
〈θ1,...,θm〉d the homogeneous component of
degree d of the ideal generated by (θ1,...,θm) in I(G).
Algorithm 5.1 [Computing secondary invariants]
Input: a system of parameters (θ1,...,θm)
and a function nextInvariant(d) which iterates through a
set of invariants of degree d spanning I(G)d as a
vector space.
Output: (irreducible) secondary invariants.
Some comments about this algorithm are in order:
(i) In the last loop, the Hilbert series provides a stopping
condition, since the number of secondary invariants is known. To
maintain efficiency, the elements of L are mutually reduced by Gauss
elimination. Testing if p is in the vector space generated by L
amounts to reducing it modulo L; inserting it into L amounts to
further reducing the elements of L by p. Therefore, this algorithm
is essentially a step by step matrix inversion by Gauss elimination,
and the cost for each degree is about (dim I(G)d3.
(ii) The main waste of memory and time in this algorithm is the
explicit computation of the vector space basis L of
〈θ1,...,θm〉d. It would be nice to work
directly in the quotient of I(G) by the ideal
〈θ1,...,θm〉, as in the algorithm based on
a Gröbner basis precomputation. This approach is further developed
in [31].
(iii) By properly keeping track of the reductions in L, we
can determine which primary invariants should be removed in order to
obtain a minimal generating set. In addition, by properly choosing the
nextInvariant function, we can check whether a given set of
invariants is generating, and if not construct counter-examples.
For n=5, we could only compute a partial minimal generating set
S5 up to degree 10, whereas the best a priori degree bound is
β( In)≤22. However, s10( In)=0 (i.e. a minimal
generating set contains no invariant of degree 10), and
Figure 2 strongly suggested that sd( In)=0 for any
d≥10. This has been checked by Kemper [18], using
ad hoc computations, thus proving that S5 is a minimal
generating set (see § 9 for a possible alternative
approach). Therefore, β( I4)=5, and β( I5)=9.
Conjecture 5.2
If n≥4, β( In)=() -1.
Figure 2: sd( In): number of invariants per degree d in a minimal
generating set of In
6 The Gorenstein property
In this section, we show that the invariant ring In is Gorenstein
when n is even, which indicates several duality properties of
In. In particular, et = d1+...+dm - m and there are as many
secondary invariants of degree et-d and degree d. Actually, this
could be used to considerably speed up the construction of secondary
invariants [32].
Lemma 6.1
(i) Let σ be a permutation of the vertices, that is an
element of Sn, and σ be the corresponding
permutation of the edges in Gn. Then:
|
=sign(σ) |
|
if n is odd, |
|
=1 |
|
if n is even. |
(ii) If n is even, Gn is a subgroup of the special linear group
SL(V).
(iii) If n is odd, the representation of Sn on the
irreducible component [n-2,2] is a subgroup of SL(V).
Proof:
(i) If σ is a transposition of 2 vertices, then
σ exchanges n-2 pairs of edges, and
sign(σ)=(-1)n-2.
(iii) Take σ∈ Sn, and M the matrix of the
representation of σ on the component [n-2,2]. The
determinant of the representation of σ on Vn is
sign(σ), whereas the sign of the representation of
σ on the other component [n]⊕[n-1,1] is
sign(σ) (natural representation of Sn). Therefore,
det(M)=sign(σ)/sign(σ)=1, if n is odd.
QED
Then, Watanabee's theorem [27, § 8] applies.
Theorem 6.2
(i) When n is even, In is Gorenstein.
(ii) When n is odd, the invariant ring over the irreducible
component [n-2,2] is Gorenstein.
7 The chain product
We have discussed the power of the grading of the invariant ring. We
now define another product on the invariant ring In, called the
chain product, which preserves a finer grading and has a nice
computational behavior. Most algebraic properties of the invariant
ring with respect to the chain product transfer back to the usual
product. We only construct and use the chain product for In, but
it generalizes to any permutation group [32].
Let g be a multigraph. As in the following example, it can be
interpreted as a superposition of simple graphs
g1,g2,⋯,gk, where
g1⊇g2⊇⋯⊇gk:
Thus, g can be identified with the multichain (i.e. chain
with repetitions)
C(g):=g1⊇g2⊇⋯⊇gk of simple
graphs. The shape λ(g) of g is the decreasing
sequence of the sizes of the simple graphs in C(g). Here,
λ(g)=(5,3,3). A polynomial is called
finely-homogeneous if all its monomials have the same shape.
Since two monomials xg and xg' in the same
Sn-orbit have the same shape, any invariant decomposes into a
sum of finely-homogeneous invariants. Therefore, the shape defines a
fine grading on the invariant ring In, and we denote by
In,(5,3,3) the finely homogeneous component of In
for the shape (5,3,3).
The usual product does not preserve this grading:
|
( |
 |
) |
|
|
|
( |
 |
) |
|
|
=
|
( |
 |
) |
|
|
+
2 |
( |
 |
) |
|
|
+
2 |
( |
 |
) |
|
|
.
|
The chain product xg⋆xh of two monomials
xg and xh is the usual product of xg and
xh if the two multichains C(g) and C(h) can be merged
into another multichain, and zero otherwise. The chain product extends
to invariants, and yields for example:
The chain product preserves the fine grading of the invariant ring,
since the shape xg⋆xh can be obtained by merging the
shapes of g and h.
The invariant ring In together with the chain product is actually
isomorphic to the Stanley-Reisner ring of the poset of unlabelled
graphs on n vertices ordered by subgraph. Stanley-Reisner rings of
posets have been intensively studied, in particular by Garsia and
Stanton [11] to construct Hironaka
decompositions of invariant rings of certain permutation groups. We
did not succeed in using this theoretical framework to get a Hironaka
decomposition of In. The need of taking the elementary symmetric
polynomials as a system of parameters causes the main difficulty.
Indeed, this is not a low degree system of parameters and there are
too many secondary invariants. However, even our naive point of view
of the Stanley-Reisner ring as an alternative product on In yields
dramatic speed ups of the computations.
The following proposition is the heart of this technique.
Proposition 7.1 [[11]]
A Hironaka decomposition of In for the chain product, is also a
Hironaka decomposition of In for the usual product.
The key of the proof is that, if p and q are finely homogeneous,
the maximal finely homogeneous component of pq is exactly p⋆
q. The result follows by induction over the fine grading. We used the
same principle to prove a similar result on generating sets.
Proposition 7.2
A generating set of In for the chain product is a generating set
of In for the usual product.
In all our examples, however, minimal generating sets for the chain
product were far from being minimal for the usual product.
The elementary symmetric polynomials form a system of parameters for
the chain product. We do not know if there are other systems of
parameters, since the usual characterization from
proposition 4.4 does not apply for the chain
product. In particular, the symmetric power sums do not form a system
of parameters for the chain product. They are not even algebraically
independent since ∑ x{i,j}k=(∑ x{i,j})k. Given the size of the
minimal generating sets we computed, there are no systems of
parameters for the chain product with degrees as low as in
conjecture 4.1.
In [32], we describe how to use this product for
faster computations. Practically, we could push the computation of a
partial minimal generating set S for I5 up to the degree 22
instead of only 10. This is a significant progress, considering that
the dimension of I5,22 is 174403, whereas the dimension of
I5,10 is only 974. Unfortunately, we cannot use a low-degrees
system of parameters, so the degree bound is 42 instead of 22.
This means that there is still a lot of work to do to get a full
minimal generating set for the chain product. On the other hand, this
partial computation yields a generating set for the usual product,
since β( I5)≤22.
Proposition 7.3
The computed set S is a generating set of I5 for the usual
product. However, S has more than one thousand invariants of
degree up to 22.
To conclude, the usual product allowed us to compute a small set, with
is minimal, but not necessarily generating, whereas the chain product
allowed us to compute a set which is generating, but far from being
minimal.
8 The invariant ring for n=∞
In this section, we study the projective limit I∞ of the invariant
ring, and get back some information on In.
A multigraph g on n'≤ n non-isolated vertices can be identified
with a multigraph on n vertices by adding n-n' isolated vertices.
This defines xgˆ* in In. The set Bn of all invariants
xgˆ*, where g is a multigraph on less than n non-isolated
vertices, is obviously a vector space basis of In. For n'≤ n,
let Φn' be the linear projection from In to In'
which maps xgˆ* (in In) to 0 if g has strictly more than
n' non-isolated vertices, and to xgˆ* (in In')
otherwise. Our definition of the exponential (see
§ 1.3) makes it a surjective morphism of
graded algebra. The projective limit of In:
defines a graded algebra I∞, with a canonical vector space basis
B∞:={xgˆ*} indexed by the multigraphs g on a finite
number of non-isolated vertices.
Proposition 8.1
(i) I∞ is the free polynomial ring over C:={xgˆ* |
g is connected}.
(ii) The canonical morphism of graded algebra Φn:
I∞ In is an isomorphism up to the degree
⌊n/2⌋.
Proof:
(i) Following the proof of proposition 2.1
(ii), C generates I∞. Now, let g1,...,gk be k>0
connected multigraphs. In the product
xg1ˆ*⋯xgkˆ*, there is a term xhˆ* with
coefficient 1, where h is the disconnected multigraph whose
connected components are precisely the gi. This term is a marker
of the product xg1ˆ*⋯xgkˆ* in any non-trivial
polynomial combination of elements of S. The algebraic
independence follows.
Any multigraph with d edges and no isolated vertices has
less than 2d vertices. (ii) follows.
QED
Corollary 8.2
β( In)≥ ⌊n/2⌋.
This lower bound is loose: for n≤5, we know that
β( In)≥()-1 and for 11≤ n≤18, it follows from
theorem 2.3 (ii) that β( In)≥ n-2. We expect
that refining this technique will yield much better lower bounds.
By (ii), the Hilbert series H( I∞,z) is the limit of the Hilbert
series H( In,z) as n goes to infinity, and by (i)
where nd is the number of connected multigraphs with d edges. We
do not know how to directly compute H( In,z), or whether there exists a
closed form formula. The only asymptotic studies we have seen in the
literature deal with n fixed and d going to
infinity [15].
9 Unimodality
A startling fact revealed by our computations of minimal generating
sets (MGS) lies in Figure 2, which shows the coefficients
of s( In,z). For n≤4 and most likely for n=5, this
polynomial is unimodal: the coefficients first increase
with the degree, and then decrease down to 0.
Conjecture 9.1
The polynomial s( In,z) is unimodal.
This would prove that the partial minimal generating set we computed
for n=5 is generating, and provide a very nice stopping condition
for algorithm 5.1.
To figure out which properties of Gn could be useful to prove this
conjecture, we extend it to general groups of matrices. A finite
subgroup G of GL(V) is MGS-unimodal if the polynomial
s(I(G),z) is unimodal.
Problem 9.2
Characterize MGS-unimodal groups.
Not all groups are MGS-unimodal. Indeed, let G be the subgroup of
GL(C2) generated by the matrix
where j and j are the two non-trivial third roots of
unity. Obviously, (x1x2, x13, x23) is a minimal generating set
of I(G), and s(G,z)=z+0z2+2z3, which is not unimodal.
Whereas the irreducible representations of the symmetric group are
thoroughly described, their invariants rings are barely known. In an
amazing but very technical paper [7], Dixmier has been
able to construct by hand minimal generating sets for several
irreducible representations of Sn, including all the irreducible
representations of S1,..., S5, except [3,1,1]. It follows
that the representations [n],[n-1,1], [2,2] and [3,2] are
MGS-unimodal, whereas the representations [2,1n-2] for n≥4
and [2,2,1] are not. This proves the existence non-MGS-unimodal
irreducible representations of the symmetric group.
The trivial group, the full symmetric group and multisymmetric
polynomials are MGS-unimodal. We checked with PerMuVAR that several
other small permutation groups are MGS-unimodal. It's tempting to
conjecture that all permutation groups are MGS-unimodal, since they
give rise to a lot of unimodality properties (see [28];
note that, as opposite to here, the corresponding series are always
either log-concave or symmetric). However, for n≥4, the
alternating group An is not MGS-unimodal. Indeed,
I( An) is generated by the elementary symmetric
polynomials of degrees 1,...,n together with the Van-der-Monde
determinant ∏i<j (xi-xj) of degree ().
Figure 1 also shows that, up to n=21, the
generating series of the secondary invariants is unimodal (except at
d=0 and possibly d=et), and very smooth.
Conjecture 9.3
Let G be a permutation group, and (θ1,...,θm) be a
system of parameters of I(G). Then, the generating series of the
secondary invariants is unimodal, except for d=0 and possibly
d=et.
We recall that this series can be computed directly from the Hilbert
series. Therefore a careful study of the Hilbert series might yield a
simple proof of this conjecture.
10 The invariant ring over digraphs
In [14], Grigoriev introduced a related invariant
ring, the invariant ring over digraphs (digraphs are directed
graphs, with loops). The definition is similar to the one for the
invariant ring over graphs, but there are n2 variables
(x1,1,x1,2,...,xn,n), indexed by the pairs (i,j) of
{1,...,n}. The action of Sn is then defined by
σ⋅ xi,j:=xσ(i),σ(j). In this section, we denote
by In the invariant ring over digraphs. More generally,
Grigoriev defined the invariant ring over oriented k-hypergraphs,
with nk variables indexed by k-uples of {1,...,n}.
Lemma 1 of [14] states that In is generated by
the invariants xgˆ*, where g is a simple digraph. The proof is
said to be an easy generalization of the usual proof of the
fundamental theorem of symmetric functions. This surprised us, since
we proved this was false in In (theorem 2.3 (ii)).
Therefore, we checked the condition 1.2, which failed
even for n=3 and degree 5. We then ran PerMuVAR to try to compute
a minimal generating set using only simple digraphs. It failed as
expected, and produced the two following very small invariants, which
are not generated by simple digraphs:
( |
 |
) |
|
|
, |
|
|
( |
 |
) |
|
|
. |
|
These counter-examples to lemma 1 of [14] can also
easily be checked by hand.
This was disappointing. Indeed, the invariant ring In is the
quotient ring of In by the ideal generated by xi,i=0 and
xi,j=xj,i. Therefore, we could have used lemma 1 to prove that
In is generated by the invariants where each variable appears with
degree at most 2. This would provide a pretty good degree bound
β( In)≤ 2(). Moreover, we could use this together
with computations of partial minimal generating sets to prove that
I5 is generated by the invariants xgˆ*, where g is a
multigraph with at least one isolated vertex, result of interest for
the reconstruction problem.
Most of the results on In apply as well for In, but since the
number of variables is greater, the computations are even harder than
for In, even if we ignore loops.
11 The field of invariant fractions
The field of invariants K(x{i,j}) Sn is the subfield of
all rational fractions of K(x{i,j}) which remain invariant under the
action of the group. The following classical lemma is valid for any
finite group of matrices.
Lemma 11.1
The field of invariant is exactly the field of fractions of the
invariant ring.
Proof:
By averaging over the group, write any invariant fraction as p/q where p and q are invariant polynomials.
QED
In [14], Grigoriev used basic Galois theory to prove
the existence of a generating set of the field of invariants
composed of m+1 invariants of degree less than m. The principle is
to first take the m elementary symmetric polynomials, and to
consider the subfield of symmetric fractions. Since the ground field
K has characteristic zero (this would be also the case for any
normal ground field, like a finite field), the primitive element
theorem applies: there exist a primitive element p which generates
the field of invariants over the field of symmetric fractions.
Therefore, the m elementary symmetric polynomials together with p
generates the field of invariants.
However, Grigoriev did not provide a way to construct such an element.
Moreover, the proof that it could be chosen of degree less than m
was incorrect, since it relied on lemma 1 of [14]
which we disproved in § 10.
Theorem 11.2
Let n≥4. The field of invariants over graphs (respectively over
digraphs) is generated by the elementary symmetric polynomials
together with:
p:= |
( |
 |
) |
|
|
, |
|
|
respectively
p:= |
( |
 |
) |
|
|
. |
|
Proof:
Key fact: in both cases a permutation of the edges belongs to the
group if and only if it leaves p invariant.
QED
Grigoriev also stated that such a generating set would be a complete
system of invariants. This is incorrect since, unlike a generating set
of the invariant ring, a generating set of the field of fraction is
not necessarily a complete system of invariants. For example, our
generating sets do not separate the following pairs of non-isomorphic
graphs:
In some cases, the field of invariants can be used to indirectly apply
Galois theory on the invariant ring.
Theorem 11.3
If n≢4,5,6,8, there is no intermediate invariant ring of matrix
group between the ring of symmetric polynomials (respectively the
ring of alternate polynomials for n even) and the ring of
invariants In.
Proof:
For n≢4,5,6,8, the group Gn is a maximal proper subgroup of
the symmetric group Sm (respectively the alternate group
Am for n even) [8]. Basic Galois
theory then proves the theorem for the field of invariants, and
lemma 11.1 transfers it back to the invariant ring.
QED
12 Conclusion
Invariant theory provides both very general tools and algorithms to
study the invariant ring In over graphs. Unfortunately, the
computer exploration of small cases appears to be very hard and shows
that those tools and algorithms lack accuracy and efficiency for our
particular invariant ring. However, we could still obtain a few
results, formulate conjectures related to In, and solve a problem
arising from graph theory.
Acknowledgments
This research was partially funded by the Région Rhône-Alpes. We
gratefully thank A. Garsia, J-C. Faugère and N. Wallach for the time
invested trying to compute a Gröbner basis for
conjecture 4.2. In particular, discussions with
A. Garsia where very helpful, and raised decisive ideas. Finally, we
would like to thank M. Pouzet for introducing and guiding us through
this beautiful subject.
References
- [1]
-
Aslaksen, H., Chan, S.-P., and Gulliksen, T.
Invariants of S4 and the shape of sets of vectors.
Appl. Algebra Engrg. Comm. Comput. 7, 1 (1996), 53--57.
- [2]
-
Bondy, J. A.
A graph reconstructor's manual.
In Surveys in combinatorics, 1991 (Guildford, 1991), vol. 166
of London Math. Soc. Lecture Note Ser. Cambridge Univ. Press,
Cambridge, 1991, pp. 221--252.
- [3]
-
Cameron, P. J.
Stories from the age of reconstruction.
Congr. Numer. 113 (1996), 31--41.
Festschrift for C. St. J. A. Nash-Williams.
- [4]
-
Capani, A., Niesi, G., and Robbiano, L.
Cocoa, a system for doing computations in commutative algebra.
Available via anonymous ftp from: cocoa.dima.unige.it.
- [5]
-
Cox, D., Little, J., and O'Shea, D.
Ideals, varieties, and algorithms, second ed.
Springer-Verlag, New York, 1997.
An introduction to computational algebraic geometry and commutative
algebra.
- [6]
-
Derksen, H., and Kraft, H.
Constructive invariant theory.
In Algèbre non commutative, groupes quantiques et invariants
(Reims, 1995). Soc. Math. France, Paris, 1997, pp. 221--244.
- [7]
-
Dixmier, J.
Sur les invariants du groupe symétrique dans certaines
représentations. II.
In Topics in invariant theory (Paris, 1989/1990), vol. 1478 of
Lecture Notes in Math. Springer, Berlin, 1991, pp. 1--34.
- [8]
-
Faradzev, I. A., Ivanov, A. A., Klin, M. H., and Woldar, A. J., Eds.
Investigations in algebraic theory of combinatorial objects.
Kluwer Academic Publishers Group, Dordrecht, 1994.
- [9]
-
Faugère, J.-C.
A new efficient algorithm for computing Gröbner bases (F4).
J. Pure Appl. Algebra 139, 1-3 (1999), 61--88.
Effective methods in algebraic geometry (Saint-Malo, 1998).
- [10]
-
Fulton, W., and Harris, J.
Representation theory, vol. 129 of Graduate Texts in
Mathematics.
Springer-Verlag, New York, 1991 - 1996.
A first course, Readings in Mathematics.
- [11]
-
Garsia, A. M., and Stanton, D.
Group actions of Stanley - Reisner rings and invariants of
permutation groups.
Adv. in Math. 51, 2 (1984), 107--201.
- [12]
-
Garsia, A. M., and Wallach, N.
Personnal communication, Feb. 1999.
- [13]
-
Göbel, M.
A constructive description of SAGBI bases for polynomial
invariants of permutation groups.
J. Symbolic Comput. 26, 3 (1998), 261--272.
- [14]
-
Grigoriev, D. J.
Two reductions of the graph isomorphism to problems for polynomials.
Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI)
88 (1979), 56--61, 237--238.
Studies in constructive mathematics and mathematical logic, VIII.
- [15]
-
Harary, F., and Palmer, E. M.
Graphical enumeration.
Academic Press, New York, 1973.
- [16]
-
Kemper, G.
The invar package for calculating rings of invariants.
IWR Preprint 93-94, University of Heidelberg, 1993.
- [17]
-
Kemper, G.
Computational invariant theory.
Queen's Papers in Pure and Applied Math. (Feb. 1998).
- [18]
-
Kemper, G.
Complete computation of a minimal generating set of the invariant
rings over graphs on 5 nodes.
personal communication, 2000.
- [19]
-
Kocay, W. L.
Some new methods in reconstruction theory.
In Combinatorial mathematics, IX (Brisbane, 1981). Springer,
Berlin, 1982, pp. 89--114.
- [20]
-
Mallows, C. L., and Sloane, N. J. A.
On the invariants of a linear group of order 336.
Proc. Cambridge Philos. Soc. 74 (1973), 435--440.
- [21]
-
Pouzet, M.
Quelques remarques sur les résultats de Tutte concernant le
problème de Ulam.
Publ. Dép. Math. (Lyon) 14, 2 (1977), 1--8.
- [22]
-
Pouzet, M.
Note sur le problème de Ulam.
J. Combin. Theory Ser. B 27, 3 (1979), 231--236.
- [23]
-
Pouzet, M., and Thiéry, N. M.
Algebraic invariants of graphs and reconstruction.
Discrete Mathematics (2001).
In preparation.
- [24]
-
Robbiano, L., and Sweedler, M.
Subalgebra bases.
In Commutative algebra (Salvador, 1988). Springer, Berlin,
1990, pp. 61--87.
- [25]
-
Schmid, B. J.
Finite groups and invariant theory.
In Topics in invariant theory (Paris, 1989/1990). Springer,
Berlin, 1991, pp. 35--66.
- [26]
-
Smith, L.
Polynomial invariants of finite groups. A survey of recent
developments.
Bull. Amer. Math. Soc. (N.S.) 34, 3 (1997), 211--250.
- [27]
-
Stanley, R. P.
Invariants of finite groups and their applications to combinatorics.
Bull. Amer. Math. Soc. (N.S.) 1, 3 (1979), 475--511.
- [28]
-
Stanley, R. P.
Log-concave and unimodal sequences in algebra, combinatorics, and
geometry.
In Graph theory and its applications: East and West (Jinan,
1986). New York Acad. Sci., New York, 1989, pp. 500--535.
- [29]
-
Sturmfels, B.
Algorithms in invariant theory.
Springer-Verlag, Vienna, 1993.
- [30]
-
Thiéry, N. M.
Invariants algébriques de graphes et reconstruction; une étude
expérimentale.
PhD thesis, Université Lyon I, June 1999.
N∘ d'ordre: 167-99.
- [31]
-
Thiéry, N. M.
Computing minimal generating sets of invariant rings of permutation
groups with SAGBI-Gröbner basis.
In International conference DM-CCG, Discrete Models -
Combinatorics, Computation and Geometry, Paris, July 2-5 2001 (2001).
Submitted.
- [32]
-
Thiéry, N. M.
PerMuVAR, a library for computing in invariant rings of
permutation groups.
Journal of Symbolic Computations (2001).
In preparation.
- 1
- Using ad hoc computations,
Kemper [18] checked recently that this system was
indeed a complete minimal generating set, thus proving that
β( I5)=9.
This document was translated from LATEX by
HEVEA.