You are on page 1of 23

ON Alt(n)-MODULES WITH AN ADDITIVE DIMENSION

WHEN n ≤ 6

BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

Abstract. Working in the general context of “modules with an ad-


arXiv:2405.05230v1 [math.GR] 8 May 2024

ditive dimension,” we complete the determination of the minimal di-


mension of a faithful Alt(n)-module and classify those modules in three
of the exceptional cases: 2-dimensional Alt(5)-modules in characteristic
2, 3-dimensional Alt(5)-modules in characteristic 5, and 3-dimensional
Alt(6)-modules in characteristic 3. We also highlight the remaining work
needed to complete the classification of the faithful Alt(n)-modules of
minimal dimension for all n; these open problems seem well suited as
projects for advanced undergraduate or master’s students.

1. Introduction
The setting of modules with an additive dimension was recently intro-
duced by Corredor and two of the present authors in an effort to unify and
extend the existing theory of G-modules in various contexts that support a
well-behaved notion of dimension, [CDW23]. For example, it treats simul-
taneously the classical setting of finite-dimensional vector spaces as well as
certain “tame” model-theoretic classes of modules, such as those possessing
finite Morley rank and those that are o-minimal. This setting provides a
promising context in which to undertake a study of model-theoretic repre-
sentation theory with the work of [CDW23], and to some extent the present
paper, establishing basic results. Notably, this level of generality does not
support character theory, nor basic tools such as Maschke’s theorem, but
“elementary” methods focusing on generators and relations still yield strong
(and perhaps clarified) conclusions.
In [CDW23], the authors classify the faithful Sym(n)- and Alt(n)-modules
of least dimension in this new context, assuming n is large enough, hence
solving the generic version of the following.
Main Problem. Among modules with an additive dimension, identify the
faithful Sym(n)- and Alt(n)-modules of least dimension for each n.
However, [CDW23] leaves open the topic for several small values of n; the
left side of Table A provides a summary of the situation for Alt(n)-modules.
These remaining cases are complicated by exceptional isomorphisms between
small instances of families of finite simple groups (e.g. Alt(5) ∼
= SL(2, 4)).
Via such isomorphisms, new geometric objects arise, and far from being

Date: May 9, 2024.


2020 Mathematics Subject Classification. Primary 20C30; Secondary 03C60, 20F11.
Key words and phrases. alternating group, finite-dimensional algebra, first-order rep-
resentation theory.
1
2 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

p
p 2 3 5 > 5 or 0
2 3 5 > 5 or 0 n
n
5 2 3 s 3
5 − − − −
6 − − − − 6 4 3 5 5
7 − s s s 7 4 s s s
8 − s s s 8 4 s s s
9 8 s s s 9 8 s s s
>9 s s s s >9 s s s s

Results of [CDW23] Incorporating our re-


sults

Table A. Minimal dimensions of faithful Alt(n)-modules


in characteristic p: “s” stands for (reduced) “standard” (see
§§ 1.1) and as a dimension translates to n − 2 if p is a prime
dividing n and as n − 1 otherwise. An entry in the table is
boxed if the modules of least dimension have been classified.

exotic, the resulting modules are geometrically natural. Our original fo-
cus for the present work was on Alt(n)-modules with n ≥ 5; however, our
intermediate results also address smaller n and the symmetric case.
In this paper, we make progress on the Main Problem by determining the
value of the minimal dimension of a faithful Alt(n)-module for all n that
are not treated in [CDW23] and also provide identification of the minimal
modules when (n, p) is (5, 2), (5, 5), or (6, 3). This is summarized on the
right side of Table A; precise results and additional cases are given in §§ 1.2.
A key take-away from our work echos that from [CDW23]: even though
the setting of modules with an additive dimension is quite general (with
known objects that are not in the category of vector spaces), the minimal
objects are indeed just the familiar ones.
1.1. Examples. Here we lay out some of the classical Sym(n)- and Alt(n)-
modules that provide the desired lower bounds on the dimension of a faithful
module and will be the targets for identification. We begin with the standard
module; notation mostly coincides with that in [CDW23].
Example 1.1 (Standard Module). Let perm(n, Z) = Ze1 ⊕ · · · ⊕ Zen be the
Z[Sym(n)]-module where Sym(n) permutes the ei naturally (by permuting
the subscripts). This is not a minimal Sym(n)-module; two submodules are:
P P
• std(n, Z) := { i ci eP
i | ci ∈ Z and ci = 0};
• Z(perm(n, Z)) := { i cei | c ∈ Z}.
We can easily generalize these modules to have coefficients in any abelian
group L. Indeed, view L as a trivial Sym(n)-module, and set perm(n, L) :=
perm(n, Z) ⊗Z L, which has submodules:
• std(n, L) := std(n, Z) ⊗Z L;
• Z(perm(n, L)) := Z(perm(n, Z)) ⊗Z L.
In the case when Z(perm(n, L)) ∩ std(n, L) 6= 0, we can get a ‘smaller’
module by considering the quotient r-std(n, L) = std(n, L)/(Z(perm(n, L))∩
ON Alt(n)-MODULES WHEN n ≤ 6 3

std(n, L)). We refer to std(n, L) as the standard module over L and the
quotient r-std(n, L) as the reduced standard module over L; these are
of course modules for Alt(n) as well.
Remarks 1.2.
(1) In the familiar linear setting when L = F+ is the additive group
of a field F , let ldimF denote the linear dimension over F . Then
ldimF (std(n, F+ )) = n − 1. Further, std(n, F+ ) = r-std(n, F+ ) if and
only if char F ∤ n; if char F | n, we have ldimF (r-std(n, F+ )) = n − 2.
(2) In general, std(n, L) = r-std(n, L) if and only if L does not have
elements of order dividing n.
Example 1.3 (Sign and Sign-tensored Modules). Let σ : Sym(n) → {±1}
denote the sign or signature homomorphism: σ(g) = 1 if and only if g
is an even permutation. We let sgn(n, Z) = Z be the Sym(n)-module with
action g(v) = σ(g) · v; we refer to this as the sign module. Though we will
not use it, we can “extend the scalars” to include any abelian group L by
letting sgn(n, L) = sgn(n, Z) ⊗Z L.
The sign module is clearly not faithful when n ≥ 3 (since Alt(n) acts
trivially), but we can use it to build an important, and typically faithful,
variant of the (reduced) standard module. For any abelian group L, we set
• stdσ (n, L) := sgn(n, Z) ⊗Z std(n, L)
• r-stdσ (n, L) := sgn(n, Z) ⊗Z r-std(n, L).
We refer to these as the sign-tensored (reduced) standard modules.
One effect of this construction is that an element of std(n, L) is centralized
(i.e. fixed) by a transposition τ if and only if the same element is inverted by
τ in stdσ (n, L). Note that the difference between std(n, L) and stdσ (n, L) is
only detectable by Sym(n); they are isomorphic as Alt(n)-modules (as are
r-std(n, L) and r-stdσ (n, L)).
The Sym(n)- and Alt(n)-modules of least dimension (in the classical set-
ting and in the one studied here) are always essentially of the form r-std(n, L)
or r-stdσ (n, L) provided n is large enough (see [CDW23]). But for small n,
there are various exceptional modules, some of which we highlight now.
Example 1.4 (Exceptional case: Alt(5)-modules of dimension 2). The main
point is that Alt(5) ∼ = SL(2, F4 ). Now, for F a field, we let Nat(n, F ) =
n
F be the natural module for SL(n, F ) with elements acting by matrix
multiplication. And, if L is any vector space over F , we may, as before,
extend Nat(n, F ) to the SL(n, F )-module Nat(n, L) = Nat(n, F )⊗F L. With
this notation, we have that Nat(2, L) is an Alt(5)-module whenever L is a
vector space over F4 .
Example 1.5 (Exceptional case: Alt(6)-modules of dimension 3). Here we
use that Alt(6) ∼
= PSL(2, F9 ). For F a field, we let Ad(n, F ) denote the
set of n × n matrices over F of trace 0, viewed as a PSL(n, F )-module with
elements acting by conjugation; this is the adjoint module for PSL(n, F ).
As usual, we also define the module Ad(n, L) = Ad(n, F ) ⊗F L for L any
vector space over F . So, in this notation, Ad(2, L) is an Alt(6)-module
whenever L is a vector space over F9 .
4 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

One also has Alt(5) ∼ = PSL(2, F5 ), which gives rise to 3-dimensional


Alt(5)-modules of the form Ad(2, L) for L any vector space over F5 . These
modules turn out to be isomorphic to r-std(5, L) (see Remark 1.6 below).
1.2. Main results. We prepare to state our main results. As in [CDW23],
we use Mod(G, d) to denote the collection of all dim-connected G-modules
that carry an additive dimension; a proper definition of being a G-module
with an additive dimension is given in §§ 2.2 as is the definition of dim-
connectedness. Both notions are quite natural, and in the classical setting
of finite-dimensional vector spaces, all subspaces are automatically dim-
connected. We say that a module V has prime characteristic p if V is
an elementary abelian p-group; the notion of characteristic 0 in our more
general setting will be defined later (see Definition 2.15).
Our first theorem determines the bounds missing from the left side of Ta-
ble A and also gives information about the case when n = 4. The completed
Table A then follows from combining it with [CDW23] (see Remark 3.12).
Theorem 1 (Bounds). Let V ∈ Mod(Alt(n), d) be faithful.
(1) If n = 4, then d ≥ 2; if d = 2, then V has an Alt(4)-submodule of
characteristic 2 of positive dimension.
(2) If n = 5, then d ≥ 2; if d = 2, then V has characteristic 2.
(3) If n = 6, then d ≥ 3. Moreover, if d = 3, then V has characteristic
3, and if d = 4, then V has characteristic 2 or 3.
Our second theorem identifies the minimal Alt(n)-modules in three special
cases: those associated to the exceptional isomorphisms Alt(5) ∼= SL(2, F4 ),
∼ ∼
Alt(5) = PSL(2, F5 ), and Alt(6) = PSL(2, F9 ); the target modules were
defined in §§ 1.1. We use Mod(G, d, p) to denote the subclass of Mod(G, d)
consisting of those modules of characteristic p.
Theorem 2 (Recognition).
(1) If V ∈ Mod(Alt(5), 2, 2) is faithful, then V ∼
= Nat(2, L) for some
1-dimensional subgroup L ≤ V .
(2) If V ∈ Mod(Alt(5), 3, 5) is faithful, then V ∼
= Ad(2, L) for some
1-dimensional subgroup L ≤ V .
(3) If V ∈ Mod(Alt(6), 3, 3) is faithful, then V ∼
= Ad(2, L) for some
1-dimensional subgroup L ≤ V .
Remark 1.6. Since r-std(5, L) ∈ Mod(Alt(5), 3, 5), we indirectly see that
Ad(2, L) ∼
= r-std(5, L), so modules in Mod(Alt(5), 3, 5) are indeed standard
as indicated in Table A.
Though our main results focus on the alternating group, we also analyze
the full symmetric group in the smallest of cases. Lemmas 3.3 and 4.6 iden-
tify the minimal faithful Sym(3)-modules that are 3-divisible; Lemmas 3.7
and 4.7 do the same for Sym(4)-modules that are 2-divisible. Recall that
an abelian group V is p-divisible if for all v ∈ V there is a w ∈ W such that
v = pw; this is a strong form of “characteristic not p” (see Fact 2.10).
1.3. Remaining work. A quick glance at Table A shows that, although
the minimal dimensions of faithful Alt(n)-modules are known, classification
of the minimal modules remains open in certain cases. To solve the Main
Problem one is left with four independent subtasks.
ON Alt(n)-MODULES WHEN n ≤ 6 5

Remaining Task 1. Classify the minimal faithful Alt(5)-modules in char-


acteristics other than 2 and 5.
This leads to the following question, which is a current favorite of ours:
is every faithful module in Mod(Alt(5), 3, p) with p 6= 2, 5 derived from
an “icosahedron module” of dimension 3? Let φ ∈ C be the golden ratio.
Working in (Z[φ])3 , we may build a regular icosahedron I on the 12 ver-
tices arising from all cyclic permutations of (0, ±1, ±φ). Viewing Alt(5) as
the rotational symmetry group of I then allows us to define an action of
Alt(5) on (Z[φ])3 ; the resulting module we denote by Icosa(5, Z[φ]). We
know of no general recognition theorem for Icosa(5, Z[φ]) (analogous, say,
to Fact 4.8) and would be very happy to see one developed. An interesting
complication will be that not all finite fields have an analogue of φ. Do
note that in the classical setting Mod(Alt(5), 3, p) with p 6= 2, 5 contains
two faithful modules: the second √ being derived
√ from Icosa(5, Z[φ]) via the
Galois automorphism exchanging 5 and − 5.
Remaining Task 2. Classify the minimal faithful Alt(6)-modules in char-
acteristic not 3.
Here we would first ask if every faithful module in Mod(Alt(6), 4, 2) is of
the form r-std(6, L) or r-stdχ (6, L) where r-stdχ (6, L) denotes the module
obtained from r-std(6, L) via the outer automorphism χ ∈ Aut(Sym(6))
applied to Alt(6). And for p 6= 2, 3, the question is if every faithful module
in Mod(Alt(6), 5, p) is of the form std(6, L) or stdχ (6, L). Actually, we
should slightly modify this in characteristic 0 to allow for a faithful V ∈
Mod(Alt(6), 5, 0) to be dim-isogenous to std(6, L) or stdχ (6, L), which we
take to mean that there is a surjective Alt(6)-module homomorphism with
a (possibly nontrivial) kernel of dimension 0; this is typically what happens
in [CDW23, Theorem].
One strategy for approaching Task 2 would be to revisit the Extension
Lemma of [CDW23] and show that some version of it holds for n = 6 when
the characteristic is not 3. The main issue seems to be in adapting Claim 1
of the Extension Lemma (which makes significant use of n ≥ 7).
Remaining Task 3. Classify the minimal faithful Alt(n)-modules in char-
acteristic 2 for n = 7, 8, 9.
A key point for this problem is that Alt(8) ∼
= SL(4, 2); thus the faithful
modules in Mod(Alt(8), 4, 2) are expected to be of the form Nat(4, L) or
Natψ (4, L) where Natψ (4, L) denotes the module obtained from Nat(4, L)
via the inverse-transpose (outer) automorphism of SL(4, 2). The case of
n = 9 requires separate treatment and presents an increased challenge: here
there are at least three types of minimal modules [ABL+ 05].
Remaining Task 4. Classify the minimal faithful Sym(n)-modules in char-
acteristics not dividing n for n = 5, 6.
This last task highlights the remaining gap in determining the minimal
Sym(n)-modules. This is raised at the beginning of §§ 1.2 in [CDW23] where
known complications are also briefly mentioned.
6 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

2. Background and preliminary results


2.1. Modules. Although the setting of modules with an additive dimension
was introduced in [CDW23], many of our methods are classical. Here we
collect a handful of basic tools for analyzing G-modules that will be used in
our analysis; though typically quite classical, we include proofs to illustrate
the types of arguments used later. For now, we do not assume that there is
any notion of dimension around: a G-module is simply an abelian group V
and on which G acts as automorphisms.

Notation. Let G be a group. Assume V is a G-module and g ∈ G. We


view g as an element of Aut(V ), the automorphism group of V ; End(V )
denotes the endomorphism ring of V . When g has finite order m, we define
maps adg and trg in End(V ) as follows:
• adg = 1 − g;
• trg = 1 + g + · · · + g m−1 .
We refer to these as the adjoint and trace maps. They commute. If the
module V is clear from the context, we use notation Bg = adg (V ) and
Cg = trg (V ) for the images of V under these maps. When applying g to an
element v ∈ V , we typically write gv in place of g(v); however, we tend to
include the parentheses when applying the adjoint and trace maps.

Lemma 2.1. Let G be a group. Assume V is a G-module and g ∈ G has


finite order. Then
(1) for X ⊆ V , adg (X) = 0 if and only if g centralizes X;
(2) adg (Cg ) = 0, so Cg is centralized by g;
(3) trg (Bg ) = 0, so if |g| = 2, then Bg is inverted by g;
(4) every nontrivial element of Bg ∩ CV (g) has order dividing |g|;
(5) for all h ∈ G, h ◦ adg ◦h−1 = adhgh−1 and h ◦ trg ◦h−1 = trhgh−1 .

Proof. Set |g| = m. We proceed point by point.


(1) This follows from the definition of the adjoint map: adg (X) = 0 if
and only if for all v ∈ X one has v − gv = 0 (equivalently v = gv).

(2) For each v ∈ Cg there is w ∈ V such that v = 1 + g + · · · + g m−1 w,
so:

adg (v) = (1 − g) 1 + g + · · · + g m−1 w = (1 − gm ) w = 0w.

Thus, adg (Cg ) = 0, which is equivalent to g centralizing Cg by (1).


(3) Similar to the proof of (2), we may write each v ∈ Bg as v = (1 − g) w
for some w ∈ V , so

trg (v) = 1 + g + · · · + g m−1 (1 − g) w = (1 − g m ) w = 0w.

Thus, trg (Bg ) = 0. Further, if |g| = 2, then 0 = trg (v) = v + g · v if


and only if g · v = −v.
(4) Let v ∈ Bg ∩CV (g). Since v ∈ Bg , one has trg (v) = 0 by (3), meaning
(1+ g + · · · + gm−1 )v = 0. Also, v ∈ CV (g), so (1+ g + · · · + g m−1 )v =
v + gv + · · · gm−1 v = mv. Thus, mv = 0, which implies v has order
dividing m.
ON Alt(n)-MODULES WHEN n ≤ 6 7

(5) Observe that h ◦ adg ◦h−1 = h(1 − g)h−1 = 1 − hgh−1 = adhgh−1 ,


and similarly,
h ◦ trg ◦h−1 = h(1 + g + · · · + g m−1 )h−1
= (1 + hgh−1 + · · · + hg m−1 h−1 )
 m−1 
= 1 + hgh−1 + · · · + hgh−1 = trhgh−1 . 

Lemma 2.2. Let G be a group. Assume V is a G-module and g ∈ G has


finite order. Then
(1) Bg = Bg−1 and Cg = Cg−1 ;
(2) for all h ∈ G, h(Bg ) = Bhgh−1 and h(Cg ) = Chgh−1 .
In particular, if h centralizes or inverts g, then Bg and Cg are hhi-invariant.
Proof. Set |g| = m.
(1) As g is an automorphism of V , g(V ) = V , so also −g(V ) = g(−V ) =
g(V ) = V . Thus, Bg−1 = (1 − g−1 )(V ) = (1 − g −1 )(−g)(V ) =
(−g + 1)(V ) = Bg . Similarly, gm−1 is an automorphism of V , so
gm−1 (V ) = V . Thus, Cg−1 = (1 + g−1 + · · · + g−(m−1) )(V ) = (1 +
g−1 + · · · + g −(m−1) )gm−1 (V ) = (gm−1 + gm−2 + · · · + 1)(V ) = Cg .
(2) Using Lemma 2.1(5) and the fact that V = h−1 V since h is an
automorphism, we see that
hBg = h ◦ adg (V ) = h ◦ adg ◦h−1 (V ) = adhgh−1 (V ) = Bhgh−1 ,
and a similar calculation shows h(Cg ) = Chgh−1 .
For the final remark, if h centralizes or inverts g, then hBg = Bhgh−1 =
Bg±1 = Bg and similarly for Cg . So in this case, Bg and Cg are hhi-invariant.

Notation. Let G be a group acting on an abelian group V . We define
M
[G, . . . , G, V ] = adg1 ◦ · · · ◦ adgn (V ).
| {z }
n g1 ,...,gn ∈G

This agrees with the usual definition of the commutator [G, . . . , G, V ] (with
appropriate bracketing) in the corresponding group G ⋉ V .
Lemma 2.3. Let V be a nontrivial elementary abelian p-group. If V is a
G-module for G a finite p-group, then [G, V ] < V .
Proof. The main point is that if |g| = p, then adg (V ) < V . To see this,
suppose adg (V ) = V . Then (1 − g)V = V , so (1 − g)p V = V . But, the
characteristic is p, so V = (1 − g)p V = (1 − g p )V = 0, a contradiction.
We now proceed by induction on |G|. If |G| = p, then writing G = hgi,
we have that [G, V ] = [hgi, V ] = adg (V ) < V . Now suppose |G| > p; by
nilpotence of finite p-groups (a phenomenon lost for infinite p-groups), G
has some proper nontrivial normal subgroup N . By induction, we have
[N, V ] < V . Since N acts trivially on V /[N, V ], G acts on V = V /[N, V ]
as G/N . Again, by induction, we have [G/N, V ] < V , so [G, V ] < V . This
final point implies [G, V ] < V . 
We will also require a fundamental fact about so-called quadratic actions.
8 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

Definition 2.4. Let U be a group acting on an abelian group V . We say


that the action is quadratic if [U, U, V ] = 0 but [U, V ] 6= 0.
Lemma 2.5. If a group U acts faithfully and quadratically on an abelian
group V , then U is abelian.
Proof. Let g, h ∈ U be arbitrary. By assumption, 0 = adg (adh (v)) = adg (v−
hv) = v − gv − hv + ghv for all v ∈ V . Since this holds for all g and h, we
may switch the order to get 0 = v − hv − gv + hgv as well. Together, these
equations imply that ghv = 0 = hgv, so as the action is faithful, it must be
that gh = hg, as desired. 
2.2. Modules with an additive dimension. We now introduce modules
with an additive dimension; our definitions are taken directly from [CDW23].
We refer to [CDW23, Section 2] for more details and strongly encourage the
reader to review the motivation and various examples found there. Through-
out, we use ker f and im f for the kernel and image of a map f .
Definition 2.6. A modular universe U is a collection of abelian groups
Ob(U ) and homomorphisms Ar(U ) between them that satisfy the following
closure properties.
• [inverses] If f ∈ Ar(U ) is an isomorphism, then f −1 ∈ Ar(U ).
• [products] If V1 , V2 ∈ Ob(U ) and f1 , f2 ∈ Ar(U ), then V1 × V2 ∈
Ob(U ), and Ar(U ) contains f1 ×f2 , the projections πi : V1 ×V2 → Vi ,
and the diagonal embeddings ∆k : V1 → V1k .
• [sections] If W ≤ V are in Ob(U ), then V /W ∈ Ob(U ) and Ar(U )
contains the inclusion ι : W → V and quotient p : V → V /W maps.
• [kernels/images] If f : V1 → V2 is in Ar(U ), then ker f, im f ∈
Ob(U ), and for all W1 , W2 ∈ Ob(U ),
– if W1 ≤ ker f , the induced map f : V1 /W1 → V2 is in Ar(U );
– if im f ≤ W2 ≤ V2 , the induced map fˇ: V1 → W2 is in Ar(U ).
• [Z-module structure] If V ∈ Ob(U ), then Ar(U ) contains the sum
map σ : V × V → V and the multiplication-by-n maps µn : V → V .
The groups in Ob(U ) are called the modules of U and the homomorphisms
in Ar(U ) its compatible morphisms.
If V is a module in U and a group G acts on V by compatible morphisms
of U , we say that V is a G-module in U .
Definition 2.7. Let U be a modular universe. An additive dimension
on U is a function dim : Ob(U ) → N such that for all f : V → W in Ar(U ),
dim V = dim ker f + dim im f.
If (U , dim) is a modular universe with an additive dimension and V ∈
Ob(U ), we call V a module with an additive dimension, leaving U
implicit from context.
Definition 2.8. A module V with an additive dimension is called dim-
connected (for dimension-connected) if every proper submodule W < V in
U satisfies dim W < dim V .
Fact 2.9 ([CDW23, Connectedness Properties]). Let (U , dim) be a modular
universe with an additive dimension, and let V1 , V2 be modules in U .
ON Alt(n)-MODULES WHEN n ≤ 6 9

(1) If V1 is dim-connected and f : V1 → V2 is compatible, then im f is


dim-connected.
(2) If V1 and V2 are dim-connected, then so is V1 × V2 .
(3) If V1 , V2 ≤ V and V1 and V2 are dim-connected, then so is V1 + V2 .
Notation. Let G be a group. We use Mod(G) to denote the class of all dim-
connected G-modules with an additive dimension (each coming from its own
dimensional universe); Mod(G, d) denotes the subclass of those modules of
dimension exactly d.
Fact 2.10 ([CDW23, Divisibility Properties]). Let V ∈ Mod(G) for some
group G. If p is a prime, then V is p-divisible (as defined in § 1) if and only
if Ωp (V ) := {v ∈ V : pv = 0} has dimension 0.
Definition
P 2.11. If A1 , . . . , An are submodules
P Pof a module V , the sum
Ai Pis said to be quasi-direct if dim Ai = dim Ai , in which case we
write Ai = A1 (+) · · · (+) An .
Fact 2.12 ([CDW23, Coprimality Lemma]). Let V ∈ Mod(G) for a group
G. Suppose g ∈ G has prime order p and V is p-divisible. Then V =
Bg (+) Cg .
Remark 2.13. It follows from Fact 2.12 that, under the same assumptions,
dim CV (g) = dim Cg ; this is discussed in the Remarks following the proof
of [CDW23, Coprimality Lemma]. Since dim CV (g) has a dim-connected
submodule of equal dimension, it has only one, and we say Cg is the dc-
component of CV (g), see [CDW23, Remarks, page 11].
The next lemma presents the so-called Weight Lemma from [CDW23].
We include a proof here to illustrate some of the methods we use later; it
also presents a slightly different point of view than in [CDW23].
Lemma 2.14 (cf. [CDW23, Weight Lemma]). Let G be a group with K E G
a normal Klein 4-subgroup, and let V ∈ Mod(G). Write K = {1, α1 , α2 , α3 }.
Define Vα1 = trα1 ◦ adα2 ◦ adα3 (V ), Vα2 = adα1 ◦ trα2 ◦ adα3 (V ), and Vα3 =
adα1 ◦ adα2 ◦ trα3 (V ); set CK = trα1 ◦ trα2 ◦ trα3 (V ).
If V is 2-divisible, then V = Vα1 (+) Vα2 (+) Vα3 (+) CK , and if the nontrivial
elements of K are conjugate in G, then dim V = 3 · dim Vα1 + dim CK .
Proof. We have Vα1 = (1 + α1 )(1 − α2 )(1 − α3 )V (similarly for Vα2 and Vα3 )
and CK = (1 + α1 )(1 + α2 )(1 + α3 )V . Note that, since the αi commute for
all i, the order of the factors in each expression does not matter.
Bear in mind the dim-connectedness of Bαi and Cαi .
Suppose V is 2-divisible. Since |αi | = 2, we may apply Fact 2.12 to see
that V = Bαi (+) Cαi . Since Bαi is a dim-connected submodule of V , it is
also 2-divisible by Fact 2.10, so by Fact 2.12 applied with Bαi in place of
V , Bαi = adαj (Bαi ) (+) trαj (Bαi ). Similarly, Cαi = adαj (Cαi ) (+) trαj (Cαi ).
Thus, we find:
V = Bα3 (+)Cα3
= adα2 (Bα3 ) (+) trα2 (Bα3 ) (+) adα2 (Cα3 ) (+) trα2 (Cα3 )
= adα2 ◦ adα3 (V ) (+) trα2 ◦ adα3 (V ) (+) adα2 ◦ trα3 (V ) (+) trα2 ◦ trα3 (V ).
10 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

We look at how α1 acts on the above four factors. Remember that α1 =


α2 α3 . Consider the first factor W = adα2 ◦ adα3 (V ). Notice that W ≤
Bα2 ∩ Bα3 , so both α2 and α3 invert W by Lemma 2.1(3). Then α1 = α2 α3
must centralize W , so for all w ∈ W , trα1 (w) = w + α1 w = w + w = 2w.
Thus trα1 (W ) = 2W , so by 2-divisibility, trα1 (W ) = W . This shows that
• adα2 ◦ adα3 (V ) = trα1 ◦ adα2 ◦ adα3 (V ) = Vα1 .
Similar calculations show
• trα2 ◦ adα3 (V ) = adα1 ◦ trα2 ◦ adα3 (V ) = Vα2 ;
• adα2 ◦ trα3 (V ) = adα1 ◦ adα2 ◦ trα3 (V ) = Vα3 ;
• trα2 ◦ trα3 (V ) = trα1 ◦ trα2 ◦ trα3 (V ) = CK .
Thus, V = Vα1 (+) Vα2 (+) Vα3 (+) CK .
Now suppose that α1 , α2 , α3 are conjugate in G. For each pair i 6= j there
exists g ∈ G such that gαi g−1 = αj , and for k ∈ / {i, j}, {gαj g−1 , gαk g −1 } =
{αi , αk } as K is normal. Then, similar to the proof of Lemma 2.2,
gVαi = g ◦ trαi ◦ adαj ◦ adαk (V )
= g ◦ trαi ◦ adαj ◦ adαk ◦g −1 (V )
= trgαi g−1 ◦ adgαj g−1 ◦ adgαk g−1 (V )
= trαj ◦ adgαj g−1 ◦ adgαk g−1 (V )
= trαj ◦ adαi ◦ adαk (V ) = Vαj .
Thus g(Vαi ) = Vαj , so dim Vαi = dim Vαj . Since V = Vα1 (+)Vα2 (+)Vα3 (+)CK ,
dim V = 3 · dim Vαi + dim CK . 
Definition 2.15. Let V be an abelian group. We say
• V has characteristic p, for p a prime, if V is an elementary abelian
p-group;
• V has characteristic 0 if it is divisible, i.e. if V is p-divisible for all
primes p.
Notation. Extending our earlier notation, we use Mod(G, d, p) to denote
the subclass of Mod(G, d) consisting of those modules of characteristic p
(possibly equal to 0).
Not all modules have a well-defined characteristic, but it turns out that
the “irreducible” ones do.
Definition 2.16. Let V ∈ Mod(G) for some group G. We say that V is
dc-irreducible (for “dimension-connected-irreducible”), if it has no non-
trivial, proper, dim-connected G-submodule.
Fact 2.17 (see [CDW23, Characteristic Lemma]). Let V ∈ Mod(G) for
some group G. If V is dc-irreducible, then V has a characteristic.

3. Bounding minimal dimensions


The bulk of our work occurs in this section. The main outcome will be a
proof of Theorem 1.
Lemma 3.1. Let V ∈ Mod(G, 1) for some group G. If g ∈ G has order 2,
then g either centralizes or inverts V .
ON Alt(n)-MODULES WHEN n ≤ 6 11

Proof. Consider Bg ; it has dimension 0 or 1. If dim Bg = 0, then as Bg


is dim-connected, Bg = 0, and Lemma 2.1(1) implies that g centralizes V .
Now suppose that dim Bg = 1; then dimension connectedness of V implies
Bg = V . Thus, in this case, Lemma 2.1(3) implies that g inverts V . 
Remark 3.2. Lemma 3.1 states that an element of order 2 must fix or
invert a dim-connected, 1-dimensional module. This is of course true for
any element of order 2, so if V ∈ Mod(G, 1) and a, b ∈ G both have order
2, then at least one of a, b, or ab must centralize V .
The next lemma (Lemma 3.3) as well as Lemma 3.7 appear as Claim 1
of [CDW23, First Geometrisation Lemma], but we include proofs here for
completeness and to add further detail.
Lemma 3.3. If V ∈ Mod(Sym(3), d) is faithful, then d ≥ 2.
Proof. Since V is dim-connected, d ≥ 1 as d = 0 would imply V = {0},
contradicting faithfulness. Now suppose d = 1. Let τ1 and τ2 be distinct
transpositions in Sym(3). By Remark 3.2, one of τ1 , τ2 , or τ1 τ2 fixes V .
This contradicts faithfulness, so d ≥ 2. 
Corollary 3.4. Let V ∈ Mod(Sym(3)). If γ ∈ Sym(3) has order 3, then
any dim-connected, 1-dimensional Sym(3)-submodule of adγ (V ) has charac-
teristic 3. In particular, if V is 3-divisible, then dim adγ (V ) 6= 1.
Proof. Let U be a dim-connected, 1-dimensional Sym(3)-submodule of Bγ =
adγ (V ). By Lemma 3.3, the action of Sym(3) on U is not faithful, so γ must
be in the kernel. In other words, γ centralizes U , so U ≤ CV (γ). Thus,
U ≤ Bγ ∩ CV (γ), which implies U has characteristic 3 by Lemma 2.1(4).
Now, if V is 3-divisible, then V has no subgroup of characteristic 3 of
positive dimension by Fact 2.10. In particular, cannot dim Bγ = 1 happen.

Lemma 3.5. If V ∈ Mod(Alt(4), d) is faithful, then d ≥ 2; if V is also
2-divisible, then d ≥ 3.
Proof. Let K = {1, α1 , α2 , α3 } be the (normal) Klein 4-group in Alt(4).
Then each αi has order 2, and α3 = α1 α2 . If d = 1, then by Remark 3.2,
some αi fixes V , so V is not faithful. Thus d ≥ 2.
Now suppose d = 2 and, towards a contradiction, also assume V is 2-
divisible. The three nontrivial elements of K are conjugate in Alt(4), so by
Lemma 2.14, dim V = 3 · dim Vα1 + dim CK . Since the dimension of V is 2,
the dimension of each Vαi is 0, so dim V = dim CK . As V is dim-connected,
V = CK ≤ CV (K), which contradicts faithfulness. 
Remark 3.6. Despite some effort, it remains unclear to us if a faithful
V ∈ Mod(Alt(4), 2) necessarily has characteristic 2. Lemma 3.5 (together
with Fact 2.10) shows that such a V must have exponent dividing 4, but
exponent equal to 4 has yet to admit it does not exist. Of course, exponent
2 does exist as can be seen via Alt(4) < Alt(5) ∼
= SL(2, F4 ).
We remark that in [CDW23] it is erroneously stated (in the Examples
following the Characteristic Lemma) that V = (Z/4Z)2 can be made a
faithful 2-dimensional Alt(4)-module with action given by any embedding of
12 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

Alt(4) into GL (2, Z/4Z). Certainly V can be made a faithful Alt(4)-module,


but, as we show, it cannot have dimension 2 (in the sense of Definition 2.7).
The group GL (2, Z/4Z) ∼ = (Alt(4) ⋊ Z/4Z) ⋊ Z/2Z (as can be seen, for
example, using GAP [GAP22]) contains a unique copy of Alt(4), which is
given by    
1 2 2 1
H= , .
2 1 1 1
Let α denote the first generator, and notice that 0 < Cα < Bα < V . Thus,
if this V is in Mod(Alt(4), d), we must have d > 2.
Lemma 3.7. If V ∈ Mod(Sym(4), d) is faithful, then d ≥ 3.
Proof. Since Sym(4) contains a copy of Sym(3), Lemma 3.3 implies that
d ≥ 2. Similarly, Lemma 3.5 can be used when V is 2-divisible to show that
d ≥ 3. Thus it remains to show that if V is not 2-divisible, then d 6= 2.
Assume that d = 2 and that V is not 2-divisible. Let W = 2V . Then
W is a dim-connected submodule of V . Since V is not 2-divisible, W 6= V .
Thus, as V is dim-connected, dim W ≤ 1.
If dim W = 1, consider the action of Sym(4) on W and V /W . By Re-
mark 3.2 applied to the transpositions, the product of any two transpositions
fixes W , so Alt(4) acts trivially on W . Similarly, Alt(4) acts trivially on
V /W . This implies that Alt(4) acts quadratically on V , so by Lemma 2.5,
Alt(4) is abelian, a contradiction.
Thus, dim W = 0. As W is dim-connected, 2V = W = 0, so V is an
elementary abelian 2-group. Let K = {1, α1 , α2 , α3 } be the Klein 4-group
in Alt(4). Since every v ∈ V satisfies v = −v, adαi (v) = (1 − αi )v =
(1 + αi )v = trαi (v), thus Bαi = Cαi . By faithfulness, Bαi 6= 0 and Cαi 6= V ,
so Bαi = Cαi must have dimension 1 for each i.
We claim that Bαi = Bαj for all i, j ∈ {1, 2, 3}. All elements of K
commute, so K acts on each Bαi (see Lemma 2.2). Thus, adαi (Bαj ) ≤ Bαj .
If adαi (Bαj ) has dimension 1, then Bαj = adαi (Bαj ) ≤ adαi (V ) = Bαi , so
Bαj = Bαi (as both groups are dim-connected of dimension 1). If instead
adαi (Bαj ) has dimension 0, then adαi (Bαj ) = 0, so αi fixes Bαj . Thus,
Bαj ≤ CV (αi ). Since we know each of Bαj , Bαi , and CV (αi ) have dimension
1, Bαj and Bαi must both be the dc-component of CV (αi ), which is unique
(see Remark 2.13). So, in all cases, Bαi = Bαj .
Since the K is normal in Sym(4) and Bαi = Bα1 for all i ∈ {1, 2, 3},
Lemma 2.2 shows that Bα1 is a Sym(4)-submodule of V . Considering the
action of Sym(4) on Bα1 and V /Bα1 , we find (as before) that Alt(4) acts
quadratically on V , implying that Alt(4) is abelian, a contradiction. 
Lemma 3.8. Assume n ≥ 5. Let V ∈ Mod(Alt(n), d) be faithful. If d is
the minimal dimension of a faithful Alt(n)-module, then V is dc-irreducible.
Proof. Let d be the minimal dimension of a faithful Alt(n)-module, and
choose V ∈ Mod(Alt(n), d). Assume V is not dc-irreducible, and let W be
a proper nontrivial dim-connected submodule of V . Then W and V /W are
Alt(n)-modules of dimension at most d − 1, so by assumption, Alt(n) does
not act faithfully on W nor V /W . Since n ≥ 5, Alt(n) is simple, so in fact,
Alt(n) acts trivially W and V /W , meaning that Alt(n) acts quadratically
on V . By Lemma 2.5, Alt(n) is abelian, a contradiction. 
ON Alt(n)-MODULES WHEN n ≤ 6 13

Lemma 3.9. If V ∈ Mod(Alt(5), d) is faithful, then d ≥ 2. Further, d = 2


is only possible when V has characteristic 2.
Proof. Since Alt(5) contains Alt(4), the first part of the statement is just
Lemma 3.5. Now assume d = 2. Then again by Lemma 3.5, V is not 2-
divisible, so the submodule 2V is proper in V . Since 2V is dim-connected,
Lemma 3.8 implies that 2V = 0, so V has characteristic 2. 
Remark 3.10. Note that there do exist faithful V ∈ Mod(Alt(5), 2, 2)
since Alt(5) ∼
= SL2 (F4 ), which acts naturally on the 2-dimensional F24 .
Lemma 3.11. If V ∈ Mod(Alt(6), d) is faithful, then d ≥ 3. Further, d = 3
is only possible when V has characteristic 3, and d = 4 is only possible when
V has characteristic 2 or 3.
Proof. By Lemma 3.8 and Fact 2.17, V has a characteristic q, and Fact 2.10
implies that V p-divisible for every prime p not equal to q.
Claim 1. d ≥ 3.
Proof of Claim. Consider the subgroup S of all permutations in Alt(6) that
either fix both 5 and 6 or interchange 5 and 6:
S = {ρ ∈ Alt(6) | {ρ(5), ρ(6)} = {5, 6}} .
It can be checked that S ∼
= Sym(4), so by Lemma 3.7, d ≥ 3. ♦
Claim 2. Suppose q 6= 3 and d ≤ 4. If γ, δ ∈ Alt(6) are disjoint 3-cycles,
then either
• Bγ = Bδ and Cγ = Cδ , or
• Bγ = Cδ and Cγ = Bδ .
Proof of Claim. As γ and δ commmute, hδi acts on Bγ by Lemma 2.2(2),
so we apply Fact 2.12 to see Bγ = adδ (Bγ ) (+) trδ (Bγ ). Let W1 := adδ (Bγ )
and W2 := trδ (Bγ ).
We first show that neither W1 nor W2 has dimension 1. We may assume
γ = (123) and δ = (456). Noting that (12)(45)δ(12)(45) = δ−1 , we find that
Σ = hδ, (12)(45)i is isomorphic to Sym(3). Every element of Σ centralizes
or inverts γ, so Lemma 2.2 implies that Σ acts on Bγ . Thus, we may apply
Corollary 3.4 to see dim W1 = dim adδ (Bγ ) 6= 1. A similar argument shows
that dim adγ (Cδ ) 6= 1, and as γ and δ commute, adγ (Cδ ) = (1 − γ)(1 + δ +
δ2 )(V ) = (1 + δ + δ2 )(1 − γ)(V ) = trδ (Bγ ) = W2 .
We next show W1 and W2 cannot both have dimension larger than 1.
Suppose they do. In this case, dim Bγ ≥ 4. However, W2 = trδ (Bγ ) ≤ Cδ ,
so we also find that dim Cδ ≥ 2. As γ and δ are conjugate, dim Cγ ≥ 2, so
Fact 2.12 implies that dim V = dim Bγ + dim Cγ ≥ 4 + 2, a contradiction.
Thus dim Wi = 0 for some i.
Suppose dim W1 = 0; then Bγ = W2 ≤ trδ (V ) = Cδ . As the action is
faithful, Lemma 2.1(1) implies that W2 = Bγ 6= 0, so as dim W2 6= 1, it
must be that dim W2 ≥ 2. Thus, dim Cδ ≥ 2, and as γ and δ are conjugate,
dim Cγ ≥ 2. Since dim Bγ + dim Cγ = dim V ≤ 4, it must be that dim Bγ =
dim Cγ = dim Cδ = 2. By dim-connectedness of Cδ , Bγ = Cδ , and by
conjugacy, Bδ = Cγ .
14 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

Next suppose dim W2 = 0; then Bγ = W1 ≤ adδ (V ) = Bδ , so Bγ ≤ Bδ .


Since γ and δ are conjugate, we find Bγ = Bδ . Now, adδ (Cγ ) ≤ Bδ = Bγ ,
and as δ acts on Cγ (since δ commutes with γ), we also have that adδ (Cγ ) ≤
Cγ . By Fact 2.12, adδ (Cγ ), which is contained in Bγ ∩ Cγ , has dimension
0. Another application of Fact 2.12 shows Cγ = adδ (Cγ ) (+) trδ (Cγ ), so
dim Cγ = dim trδ (Cγ ). By dim-connectedness of Cγ , Cγ = trδ (Cγ ), so Cγ ≤
Cδ . By conjugacy, Cγ = Cδ . ♦
Claim 3. Suppose q 6= 3 and d ≤ 4. Up to composing the action on V with
an automorphism of Alt(6), Bγ = Bδ for any disjoint 3-cycles γ, δ ∈ Alt(6).
Proof of Claim. By Claim 2, we only need to consider when Bγ = Cδ and
Cγ = Bδ . Let ε = γδ. We prove Bε = V . As Bγ = Cδ ≤ CV (δ), adε (Bγ ) =
(1 − γδ)(Bγ ) = (1 − γ)(Bγ ) = adγ (Bγ ). Since V is 3-divisible, Lemma 2.1(4)
implies that Bγ ∩ CV (γ) has dimension 0. Now, Bγ ∩ CV (γ) is the kernel of
the map adγ : Bγ → Bγ , so the image of the map has dimension equal to
dim Bγ . Thus, Bγ = adγ (Bγ ) = adε (Bγ ) ≤ Bε . A similar argument shows
Bδ ≤ Bε . We are assuming that Cγ = Bδ , so Bγ + Cγ ≤ Bε . By Fact 2.12,
V = Bγ + Cγ , so Bε = V as desired.
This also holds of any conjugate ε′ of ε. Now, there is an (outer) automor-
phism of Alt(6) that maps the conjugacy class of γ to the conjugacy class
of ε, see for example [JR82]. Up to composing with this automorphism, we
now have Bγ = V . But this also holds of any conjugate of γ, and δ is one
such. So we returned to Bγ = Bδ and we are done. ♦
Claim 4. If q 6= 3, then d ≥ 4.
Proof of Claim. By Claim 1, d ≥ 3, so assume d = 3. Let γ, δ ∈ Alt(6)
be disjoint 3-cycles. Using Claim 3, we consider V as an Alt(6)-module for
which Bγ = Bδ . Set B = Bγ = Bδ and ε = γδ.
We first show adδ (B) = B. Indeed, trδ (B) = trδ (Bδ ) ≤ Bδ ∩ CV (δ) has
characteristic 3 by Lemma 2.1(4), so it has dimension 0. We conclude by
coprimality.
We then show that Bε = B. Note that:
Bε = (1 − γδ)(V ) = [(1 − δ) + (1 − γ)δ](V )
≤ (1 − δ)(V ) + (1 − γ)δ(V ) ≤ Bδ + Bγ = B + B = B.
Now consider the decomposition B = adε (B) (+) trε (B). As in the proof
of Claim 2, we find that adε (B) and trε (B) are both invariant under the
action of hγ, (12)(45)i ∼
= Sym(3), so by Corollary 3.4, neither have dimension
1. Since dim B ≤ 3, one of adε (B) or trε (B) must have dimension 0. If
dim adε (B) = 0, then B = trε (B) ≤ Cε , so Bε ≤ B ≤ Cε ≤ CV (ε). By
Lemma 2.1(4), we find that Bε has characteristic 3, which contradicts the
fact that V is 3-divisible. Thus dim trε (B) = 0, so B = adε (B) ≤ Bε ≤ B,
implying B = Bε .
We next aim to show that γ and δ agree on B. By Lemma 2.1(3), trγ (B) =
trδ (B) = trε (B) = 0, so, restricted to B, we have the following equations.
1 + γ + γ2 = 0
1 + δ + δ2 = 0
1 + γδ + γ 2 δ2 = 0
ON Alt(n)-MODULES WHEN n ≤ 6 15

Multiplying the third equation by γ and subtracting it from the first yields
0 = 1 + γ 2 − γ 2 δ − δ2 = (1 − δ2 ) + γ 2 (1 − δ) = (1 + δ + γ 2 )(1 − δ).
Then, using that 1 + δ = −δ2 (from the second equation), we obtain
(γ 2 − δ2 )(1 − δ) = 0.
This equation holds on B, so we have (γ 2 −δ2 )◦adδ (B) = 0. But adδ (B) = B,
so γ 2 = δ2 on B. Squaring both sides, we find that γ = δ on B.
We also have Cγ = Cδ by Claim 2. Setting C = Cγ = Cδ , we see that
γ = δ on C since both maps centralize C by Lemma 2.1(2). Since V = B+C,
we now have that γ = δ on V , which is a contradiction to faithfulness. ♦
Claim 5. If q 6= 2 and q 6= 3, then d ≥ 5.
Proof of Claim. From our previous work, it remains to consider when d = 4.
Since V is 2-divisible, Lemma 2.14 shows that dim V = 3 · dim Vα + dim CK
where K ≤ Alt(4) is the Klein 4-group on {1, 2, 3, 4} and α is any nontrivial
element of K. As dim V = 4, dim CK is either 1 or 4. If dim CK = 4, then
V = CK ≤ CV (K), contradicting faithfulness, so dim CK = 1.
Now, the permutations (13)(56) and (12)(56) normalize K, so they act
on CK . Notice that they are conjugate under (123), which normalizes K
as well. By Remark 3.2, the product (123) = (13)(56)(12)(56) centralizes
CK . Then ad(123) (CK ) = 0, so Fact 2.12 implies that tr(123) (CK ) = CK . A
similar argument shows that trγ (CK ) = CK for any 3-cycle in Alt(4), so in
particular, CK ≤ C(123) ∩ C(234) .
By Claims 2 and 3, we may assume C(123) = C(456) and C(234) = C(156) .
Thus, CK ≤ C(123) ∩ C(234) ∩ C(456) ∩ C(156) , so by Lemma 2.1(2), CK is
centralized by each of (123), (234), (456), (156). We conclude that CK ≤
CV (h(123), (234), (456), (156)i). Now, direct computations show Alt(6) =
h(123), (234), (456), (156)i, so CK is centralized by all of Alt(6). In par-
ticular, CK is a proper nontrivial Alt(6)-submodule of V , so V is not dc-
irreducible. However, our assumptions together with the result of Claim 4
imply that d = 4 is the minimal dimension of a faithful Alt(6)-module, so
we have a contradiction to Lemma 3.8. ♦
This completes the proof of Lemma 3.11. 
We have everything needed for Theorem 1.
Proof of Theorem 1. Lemmas 3.5, 3.9, 3.11 prove almost everything. The
only point needing clarification is when V ∈ Mod(Alt(4), 2). In this case,
Lemma 3.5 implies that V is not 2-divisible, but this is equivalent to Ω2 (V ) =
{v ∈ V : 2v = 0} having dimension at least 1 by Fact 2.10. 
Remark 3.12. We briefly discuss how the right side of Table A now follows.
Letting d be minimal such that Mod(Alt(n), d, p) contains a faithful module,
Theorem 1 and the main result of [CDW23] combine to show that d is no
smaller than what is stated in Table A. (The exceptional cases for n = 7
and n = 8 are implied by the lower bound when n = 6.) That d is no larger
than stated follows from the examples in §§ 1.1, except when n = 7, 8. The
upper bound on d in these remaining cases comes from the natural action
of Alt(8) ∼= SL(4, 2) on a 4-dimensional vector space of characteristic 2
(discussed in §§ 1.3 following Task 3).
16 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

4. Recognizing minimal modules


We now work to identify various minimal modules; this will culminate in
a proof of Theorem 2.
4.1. Minimal modules for Sym(3) and Sym(4). We begin with recog-
nition results for the minimal, faithful Sym(n)-modules in the exception-
ally small cases of n = 3, 4. The main tool is the Recognition Lemma of
[CDW23], and to apply it, we mainly just need to show that a minimal,
faithful module is also dc-irreducible (with appropriate restriction on the
characteristic). However, our work here exposed a small issue with the
Recognition Lemma in the case when n = 4, which we first discuss and
correct.
Remark 4.1. The Recognition Lemma of [CDW23] is incorrect for n = 4.
To see why, first note that std(4, C+ ) and stdσ (4, C+ ) are both faithful and
irreducible Sym(4)-modules of dimension 3; let V denote either one of these
modules. Now, the condition “Alt({3, 4}) centralizes ad(12) (V )” is trivially
satisfied (for either choice of V ) because Alt({3, 4}) is itself trivial; this is
the issue with n = 4. Consequently, the Recognition Lemma implies that V
must be isomorphic to std(4, C+ ), but std(4, C+ ) ∼ 6 stdσ (4, C+ ).
=
When n = 3, the condition “Alt({3}) centralizes ad(12) (V )” also holds
trivially, but here we are saved by the fact that std(3, C+ ) ∼= stdσ (3, C+ ).
We now provide a corrected version of the Recognition Lemma, the proof
of which assumes familiarity with (and directly references) the original proof.
Lemma 4.2 (Corrected version of [CDW23, Recognition Lemma]). If n 6= 4,
the Recognition Lemma of [CDW23] holds as stated; if n = 4, the Recognition
Lemma holds under the following additional assumption: for any transpo-
sition τ , Bτ is either centralized or inverted by each element of Sym(τ ⊥ ),
where τ ⊥ is the set of fixed points of τ .
In particular, the Recognition Lemma holds when dim Bτ = 1 for τ any
transposition τ .
Proof. We adopt the notation of [CDW23]; set S = Sym(n) and A = S ′ =
Alt(n). Studying the proof of [CDW23, Recognition Lemma], there is only
one error; it occurs in the proof of Claim 1 when showing that B(ij) ∩ B(jk) is
S-invariant. We proceed to review the proof that B(ij) ∩ B(jk) is S-invariant
and correct the issue, which only appears when n = 4.
In [CDW23], it is first claimed that B(ij) ∩ B(jk) is S(ijk) -invariant, where
S(ijk) = Sym({i, j, k}). There is no error with this, and in particular, this
shows that B(ij) ∩ B(jk) is S-invariant in the case when n = 3.
The authors next claim that B(ij) ∩ B(jk) is Aj ⊥ -invariant, where Aj ⊥
denotes the subgroup of A consisting of those permutations fixing j. Their
reasoning is that A(ij)⊥ (defined to be Alt((ij)⊥ )) centralizes B(ij) and A(jk)⊥
centralizes B(jk), so Aj ⊥ centralizes B(ij) ∩B(jk) since Aj ⊥ = hA(ij)⊥ , A(jk)⊥ i.
Now, A(ij)⊥ does indeed centralizes B(ij) (and similarly for A(jk)⊥ ) by the
assumptions of the Recognition Lemma, so B(ij) ∩ B(jk) is centralized by
hA(ij)⊥ , A(jk)⊥ i. However—and here is the error—we only have that Aj ⊥ =
hA(ij)⊥ , A(jk)⊥ i when n ≥ 5. If n ≥ 5, we may fix a, b ∈ / {ijk}, and note
ON Alt(n)-MODULES WHEN n ≤ 6 17

that hA(ij)⊥ , A(jk)⊥ i contains all three cycles of the form (abc) for fixed
/ {ijk} and variable c 6= j, which do indeed generate Aj ⊥ . But if n = 4,
a, b ∈
A(ij)⊥ = A(jk)⊥ = 1 while Aj ⊥ is nontrivial.
Now, assuming n = 4, we show B(ij) ∩ B(jk) is Aj ⊥ -invariant under the
(stronger) assumptions of the present lemma. Here we have that B(ij) is
either centralized or inverted by each element of S(ij)⊥ and similarly for
(jk). Thus, every subgroup of B(ij) is S(ij)⊥ -invariant, and every subgroup
of B(jk) is S(jk)⊥ -invariant. We conclude that B(ij) ∩ B(jk) is invariant under
the action of hS(ij)⊥ , S(jk)⊥ i, and, crucially, it is easy to verify that Sj ⊥ =
hS(ij)⊥ , S(jk)⊥ i. Thus, B(ij) ∩ B(jk) is invariant under Sj ⊥ , hence also Aj ⊥ .
Finally, since B(ij) ∩ B(jk) is both S(ijk) - and Aj ⊥ -invariant, we find (as
in [CDW23]) that B(ij) ∩ B(jk) is S-invariant since S = hS(ijk) , Aj ⊥ i.
We now address the special case when dim Bτ = 1 for τ a transposition.
Under this assumption, Lemma 3.1 shows that Bτ is either centralized or
inverted by a transposition of Sτ ⊥ , so the same holds of all element in Sτ ⊥ .
Since transpositions in Sτ ⊥ are conjugate, they all act the same on Bτ , so Bτ
is centralized by Aτ ⊥ (as required by the original Recognition Lemma). 
Remark 4.3. The remainder of [CDW23] needs no further correction as
the Recognition Lemma is only used when n ≥ 7; this is fairly easy to track
as [CDW23] is structured so that the Recognition Lemma is essentially only
used in the proof of the main theorem in the final section (though it also
appears in a remark immediately preceding the proof of the Theorem).
We have yet to give a proper statement of the Recognition Lemma from
[CDW23], which we will need. This comes next, but we first introduce termi-
nology (adapted from the algebraic context) to streamline the presentation.
Definition 4.4. Let V and W be modules in some modular universe with
an additive dimension. A compatible homomorphism ϕ : V → W is called
a dim-isogeny (for “dimensional isogeny”) if it is surjective with kernel of
dimension 0.
We now give a simplified version of [CDW23, Recognition Lemma], which
incorporates the correction of Lemma 4.2.
Fact 4.5 (Special case of [CDW23, Recognition Lemma]). Let n ≥ 2. Sup-
pose that V ∈ Mod(Sym(n), d, p) is faithful and dc-irreducible; assume p
does not divide n. If dim B(12) = 1, then there is a Sym(n)-module dim-
isogeny ϕ : std(n, L) → V for some 1-dimensional subgroup L ≤ V . Further,
if V has no elements of order dividing n, then ϕ is an isomorphism.
Lemma 4.6. Let V ∈ Mod(Sym(3), 2) be faithful and 3-divisible. Then,
there is a Sym(3)-module dim-isogeny ϕ : std(3, L) → V for some 1-dimensional
subgroup L ≤ V ; further, if V has no elements of order 3, then ϕ is an iso-
morphism.
Proof. Assume V is as stated in the lemma; we aim to apply Fact 4.5.
To see that V is dc-irreducible, assume, towards a contradiction, that
W is a proper nontrivial dim-connected Sym(3)-submodule of V . Then
dim W = 1. Thus, W and V /W are both 1-dimensional Sym(3)-modules.
By Lemma 3.3, Sym(3) does not act faithfully on V /W , so the 3-cycles act
18 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

trivially on V /W . Let γ ∈ Sym(3) be a 3-cycle. Since γ centralizes V /W ,


one has adγ (V /W ) = 0 so adγ (V ) ≤ W . By faithfulness of Sym(3) on V ,
adγ (V ) 6= 0, so as W is dim-connected of dimension 1, it must be that
adγ (V ) = W , which contradicts Corollary 3.4.
To apply Fact 4.5, it remains to show dim B(12) = 1. Of course, dim B(12) ≤
dim V = 2. If dim B(12) = 0, then (12) centralizes V , contradicting faith-
fulness. So suppose dim B(12) = 2. Then B(12) = V , so (12) inverts V .
By Lemma 2.2(2), (23) inverts V , so (123) = (12)(23) centralizes V , again
contradicting faithfulness. We conclude that dim B(12) = 1. 
Lemma 4.7. Let V ∈ Mod(Sym(4), 3) be faithful and 2-divisible. Then,
there is a Sym(4)-module dim-isogeny ϕ : V̂ → V with domain std(4, L)
or stdσ (4, L) for some 1-dimensional subgroup L ≤ V ; further, if V has no
elements of order 2, then ϕ is an isomorphism.
Proof. Assume V is as stated in the lemma; we again use Fact 4.5.
We first show V is dc-irreducible. If W is a proper nontrivial dim-
connected Sym(4)-submodule of V , then W and V /W are dim-connected,
Sym(4)-modules of dimension at most 2. By Lemma 3.7, Sym(4) does not
act faithfully on W nor V /W . Thus, the Klein 4-group K acts trivially
on both W and V /W . If we choose any nontrivial α ∈ K, then (as in the
proof of Lemma 4.6) adα (V ) ≤ W , so adα (adα (V )) ≤ adα (W ) = 0. Thus,
adα (V ) ≤ CV (α) by Lemma 2.1(1), and then Lemma 2.1(4) implies that
adα (V ) is an elementary abelian 2-group. Since V is 2-divisible, adα (V ) = 0
by Fact 2.10, so α acts trivially on V , a contradiction. We conclude that V
is dc-irreducible.
Next, exactly as in Lemma 4.6, we find that B(12) does not have dimension
0 nor 3, and if dim B(12) = 1, we directly apply Fact 4.5 to get the desired
result with V̂ = std(4, L).
Now assume dim B(12) = 2; by Fact 2.12, dim C(12) = 1. In this case, we
aim for the conclusion of the lemma in which V̂ = stdσ (4, L). Consider the
module V σ with the same underlying group V and Sym(4)-action defined
by g ∗ v = sgn(g)(g · v), where g · v is the original action of g on v. Then,
ad(12) (V σ ) = tr(12) (V ), so dim ad(12) (V σ ) = dim C(12) = 1. Thus, Fact 4.5
applies to V σ , and there is a surjective Sym(4)-module homomorphism ϕσ :
std(4, L) → V σ with the required properties. This then induces a Sym(4)-
homomorphism ϕ : sgn(4, Z) ⊗ std(4, L) → sgn(4, Z) ⊗ V σ , so as sgn(4, Z) ⊗
Vσ ∼= V , we have our desired morphism (found by composing ϕ with the
appropriate isomorphism). 
4.2. Minimal modules for Alt(5) in characteristic 2. Here we study
V ∈ Mod(Alt(5), 2, 2). In this setting, we have Alt(5) ∼= SL(2, F4 ), and we
aim to identify V as the natural SL(2, F4 )-module (see Example 1.4). The
main tool is the following (lovely) result of Timmesfeld; see also [Smi89].
Fact 4.8 (Special case of [Tim01, (3.7) Corollary]). Assume G ∼ = SL(2, F)
or G ∼= PSL(2, F) for F a field, and let U ≤ G be the image of the subgroup
of strictly upper triangular matrices. Suppose V is a faithful G-module such
that
(1) U acts quadratically on V ;
ON Alt(n)-MODULES WHEN n ≤ 6 19

(2) [G, V ] = V and CV (G) = 0.


L
Then, G ∼
= SL(2, F), and for some I, V = i∈I Vi with each Vi ∼
= Nat(2, F).
Lemma 4.9. Let V ∈ Mod(Alt(5), 2, 2) be faithful. Then V is an F4 -
vector space, and viewing Alt(5) as SL(2, F4 ), we have V ∼
= Nat(2, L) for
some 1-dimensional subgroup L ≤ V .
Proof. Let V be as stated, and view V as a G-module for G = SL(2, F4 ).
By Lemmas 3.8 and 3.9, V is dc-irreducible.
L
Claim 1. For some index set I, V = i∈I Vi with each Vi ∼
= Nat(2, F4 ).
Proof of Claim. To apply Fact 4.8, we first aim to show that the subgroup
U of strictly upper triangular matrices acts quadratically on V . Note that
U is an elementary abelian 2-group of order 4. By Lemma 2.3, [U, V ] < V ,
and since the action is faithful, [U, V ] 6= 0. Additionally, V is dim-connected
of dimension 2, so [U, V ] < V implies dim[U, V ] ≤ 1. Applying Lemma 2.3
again, we have that [U, U, V ] < [U, V ]. Since [U, V ] is also dim-connected,
dim[U, U, V ] = 0, and as [U, U, V ] is dim-connected, [U, U, V ] = 0.
We next show that [G, V ] = V and CV (G) = 0. The first follows immedi-
ately from the fact that V is dc-irreducible, so it remains to show CV (G) = 0.
As in Claim 1 of Lemma 3.11, Alt(5) (hence G) contains a subgroup S
isomorphic to Sym(3). Viewing V as an S-module, Lemma 4.6 tells us that
V is isomorphic std(3, L) for some 1-dimensional group L. In the current
setting, std(3, L), hence L, has characteristic 2, and as 2 does not divide 3,
no element of std(3, L) is fixed by all of S (also see the Remarks on page 4
of [CDW23]). Thus CV (S), hence CV (G), is trivial. ♦
Claim 2. V ∼ = Nat(2, L) for L = [U, V ] = CV (U ).
L
Proof of Claim. We now have V = i∈I Vi with each Vi ∼ = Nat(2, F4 ). Set
L = [U,
L V ]. It is easily
L verified that L = C V (U ) and, more precisely, that
L = C V (U ) = L i where L i = C V (U ) ∼
= (F 4 )+ . We equip L and
I i I i
each Li with the G-trivial action; then Vi ∼ = Nat(2, F4 ) ⊗ Li . Since direct
sums and tensor products commute, we then have:
!
M M M
V = Vi ∼= Nat(2, F4 ) ⊗ Li ∼
= Nat(2, F4 ) ⊗ Li = Nat(2, F4 ) ⊗ L,
I I I

which is the definition of Nat(2, L). ♦


This completes the proof. 
Remark 4.10. The equality CV (G) = 0 could also have been obtained as
follows: let V = V /CV (G), which has CV (G) = 0 by perfectness. Proceed
to identifying V . Then basic cohomology proves CV (G) = 0.
4.3. Minimal modules for Alt(5) in characteristic 5. We now investi-
gate V ∈ Mod(Alt(5), 3, 5). Here we use Alt(5) ∼ = PSL(2, F5 ) and work to
identify V as the adjoint module (as defined in Example 1.5). Identifica-
tion takes place via cubic actions, which were first analyzed in the style of
Timmesfeld by the second author [Del16] and later studied quite thoroughly
by Grüninger [Grü22].
20 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

Definition 4.11. Let U be a group acting on an abelian group V . We say


that the action is cubic if [U, U, U, V ] = 0 but [U, U, V ] 6= 0.
Fact 4.12 (Special Case of [Grü22, Theorem 8.1]). Assume G ∼ = PSL(2, F)
for F a field of characteristic not 2, and let U ≤ G be the image of the
subgroup of strictly upper triangular matrices. Suppose that V is a faithful
G-module such that
(1) U acts cubically on V ,
(2) CV (u) = CV (U ) for all nontrivial u ∈ U ,
(3) [G, V ] = V and CV (G) = 0, and
(4) V is square-free: there does not exist a non-trivial G-submodule W ≤
V for which U acts quadratically on W or on V /W .
L
Then, for some I, V = i∈I Vi with each Vi ∼ = Ad(PSL(2, F)).
The actual statement of [Grü22, Theorem 8.1] does not assume that the
acting group is known to be PSL(2, F); since we do, we may remove the
square-free condition, as we now show.
Corollary 4.13. Fact 4.12 holds without assuming that V is square-free.
Proof. Adopt all hypotheses of Fact 4.12 with the exception that V is square-
free. In particular, we may take G = PSL(2, F). Suppose V is not square-
free, and let W denote the relevant submodule.
First assume that U acts quadratically on W . Set W̌ = [G, W ]. As a
consequence of G being perfect, [G, W̌ ] = W̌ , and we also have CW̌ (G) = 0
since CV (G) = 0. But then Fact 4.8 applies to the action of G on W̌ implying
G∼ = SL(2, F) and contradicting that G ∼= PSL(2, F) in characteristic not 2.
Next, consider when U acts quadratically on V /W . Let W1 be the preim-
age in V of CV /W (G), and set V = V /W1 . We are assuming [G, V ] = V , so
[G, V ] = V as well. We now claim that CV (G) = 0. Let W2 be the preimage
in V of CV (G). Then W ≤ W1 ≤ W2 with [G, W1 ] ≤ W and [G, W2 ] ≤ W1 .
Thus, [G, G, W2 ] ≤ W , so as G is perfect, [G, W2 ] ≤ W , which implies that
W2 ≤ W1 and hence W2 = W1 . We may now apply Fact 4.8 to the action
of G on V and obtain the same contradiction as before. 
We are looking to identify the faithful modules in Mod(Alt(5), 3, 5) (and
later in Mod(Alt(6), 3, 3)), which by our previous work are dc-irreducible.
Thus, the condition [G, V ] = V is automatically met, but it is not imme-
diate that CV (G) = 0 since dc-irreducibility only forces dim CV (G) = 0
(and CV (G) might not be dim-connected). Nevertheless, it is easy to show
CV (G) = 0 in the specific cases we care about. A similar outcome was
obtained implicitly in [CDW23].
Lemma 4.14. Suppose Alt(4) ≤ G, and let V ∈ Mod(G, 3) be faithful. If
V has no elements of order 2, then CV (G) = 0.
Proof. Let K = {1, α1 , α2 , α3 } ≤ Alt(4) ≤ G be a Klein 4-subgroup. With
notation as in Lemma 2.14, we may write V = Vα1 (+) Vα2 (+) Vα3 (+) CK ,
and as the αi are conjugate in G, each Vαi has the same dimension. Since
dim V = 3, we thus have V = Vα1 (+) Vα2 (+) Vα3 . Now consider w ∈
CV (G). Writing w = v1 + v2 + v3 for vi ∈ Vαi , we may apply α1 to see
that v1 + v2 + v3 = α1 (v1 + v2 + v3 ) = v1 − v2 − v3 , which implies that
ON Alt(n)-MODULES WHEN n ≤ 6 21

2(v2 + v3 ) = 0. Since V has no elements of order 2, v2 + v3 = 0. Similarly,


using α2 and α3 , we find that v1 + v3 = v1 + v2 = 0. This implies that
2vi = 0 for each i, which, as before, leads to vi = 0, as desired. 
We proceed with identification of V ∈ Mod(Alt(5), 3, 5).
Lemma 4.15. Let V ∈ Mod(Alt(5), 3, 5) be faithful. Then V is an F5 -
vector space, and viewing Alt(5) as PSL(2, F5 ), we have V ∼
= Ad(2, L) for
some 1-dimensional subgroup L ≤ V .
Proof. Set G = PSL(2, F5 ), and view V as a G-module. Let U ≤ G be the
image of the subgroup of strictly upper triangular matrices. Lemmas 3.8
and 3.9 show that V is dc-irreducible.
L
Claim 1. For some index set I, V = i∈I Vi with each Vi ∼ = Ad(2, F5 ).
Proof of Claim. We aim to verify the hypotheses of Fact 4.12. By Corol-
lary 4.13, we may ignore that V be square free, and by dc-irreducibility
together with Lemma 4.14, we already know [G, V ] = V and CV (G) = 0.
Further, U is cyclic of prime order, so we certainly have CV (u) = CV (U ) for
all nontrivial u ∈ U .
It remains to show that U acts cubically on V . As in the proof of
Lemma 4.9, we may repeatedly use Lemma 2.3 together with the fact that
dim V = 3 to see that [U, U, U, V ] = 0, and since the action is faithful,
[U, V ] 6= 0. Now, if [U, U, V ] = 0, then U acts quadratically on V , but then
Fact 4.8 implies that G ∼ = SL(2, F5 ), a contradiction. We conclude that U
acts cubically. ♦
Claim 2. V ∼
= Ad(2, L) for L = [U, U, V ] = CV (U ).
Proof of Claim. We may assume Vi = Ad(2, F5 ) in the decomposition V =
L
i∈I Vi . From this, it is readily verified that [U, U, V ] = CV (U ), which we
denote by L. We may then argue as in Claim 2 of Lemma 4.9 to obtain
V ∼
= Ad(2, L). ♦
This completes the proof. 

4.4. Minimal modules for Alt(6) in characteristic 3. Following our ap-


proach for Mod(Alt(5), 3, 5) above, we now identify V ∈ Mod(Alt(6), 3, 3).
Lemma 4.16. Let V ∈ Mod(Alt(6), 3, 3) be faithful. Then V is an F9 -
vector space, and viewing Alt(6) as PSL(2, F9 ), we have V ∼
= Ad(2, L) for
some 1-dimensional subgroup L ≤ V .
Proof.LOur proof is similar to that of Lemma 4.15. We will only show that
V = i∈I Ad(PSL(2, F9 )) for some index set I; the proof that V ∼ = Ad(2, L)
then follows using the same proof as in Claim 2 of Lemma 4.15.
As before, set G = PSL(2, F9 ), view V as a G-module, and let U ≤ G
be the image of the subgroup of strictly upper triangular matrices. By
Lemmas 3.8 and 3.11, V is dc-irreducible.
All conditions of Fact 4.12 are easily verified as in the proof of Lemma 4.15
with the exception of (2), so we now work to show CV (u) = CV (U ) for
nontrivial u ∈ U . The subgroup U is a Sylow 3-subgroup of G, so we
22 BARRY CHIN, ADRIEN DELORO, JOSHUA WISCONS, AND ANDY YU

may choose an isomorphism of G with Alt(6) that maps U to h(123), (456)i.


Identifying G with Alt(6) in this way, consider the subgroup
S = {ρ ∈ Alt(6) | ρ ({5, 6}) = {5, 6}} ,
which is isomorphic to Sym(4).
We first show that CV (γ) is dim-connected and of dimension 1 for every
3-cycle γ. By Lemma 4.7, V is isomorphic as an S-module to either std(4, L)
or stdσ (4, L) for some 1-dimensional L ≤ V . Computing in perm(4, Z) =
Ze1 ⊕ · · · ⊕ Ze4 and using that V has characteristic 3, we find that
ad(123) ◦ ad(123) (V ) = he1 + e2 + e3 i ⊗ L = CV ((123));
thus, (using Fact 2.9) we see that CV ((123)) is dim-connected and of di-
mension 1. As any three cycle γ is conjugate to (123), CV (γ) is also dim-
connected of dimension 1.
We next show that CV ((123)) = CV ((456)). Consider here the group
T = {ρ ∈ Alt(6) | ρ ({1, 2}) = {1, 2}, ρ(3) = 3} ,
which contains (456) and is isomorphic to Sym(3). Every element of T either
centralizes or inverts (123), so T acts on CV ((123)). Since dim CV ((123)) =
1, Lemma 3.3 shows that this action is not faithful, implying that (456) acts
trivially on CV ((123)). Thus, CV ((123)) ≤ CV ((456)), so as both centraliz-
ers are dim-connected of dimension 1, they must in fact be equal.
We currently have that CV ((123)) = CV ((456)) = CV (h(123), (456)i),
so it remains to show that if ε ∈ h(123), (456)i is a product of disjoint 3-
cycles, then we also have CV (ε) = CV (h(123), (456)i). Certainly the former
centralizer contains the latter, so it suffices to show that CV (ε) is dim-
connected of dimension 1. This is quickly seen using that (123) is conjugate
to ε via an outer automorphism α of G (see [JR82]): replacing S by its image
of S under α in our analysis above shows that CV (ε) is indeed dim-connected
of dimension 1. 

Acknowledgements
Work on this project began in earnest during 2021 while the first and last
authors were master’s students at California State University, Sacramento.
Theorem 1 was the initial goal. Theorem 2 had been suggested by the second
author for some time, and with delay, it was picked up and realised in 2023.
The work of the first and third authors were partially supported by the
National Science Foundation under grant No. DMS-1954127. The first au-
thor would like to thank his wife and life partner Alana for making a world
with him.

References
+
[ABL 05] Rachel Abbott, John Bray, Steve Linton, Simon Nickerson, Simon Norton,
Richard Parker, Ibrahim Suleiman, Jonathan Tripp, Peter Walsh, and Robert
Wilson. Atlas of finite group representations, version 3. Available online at
http://brauer.maths.qmul.ac.uk/Atlas/v3, 2005.
[CDW23] Luis Jaime Corredor, Adrien Deloro, and Joshua Wiscons. Sym(n)- and Alt(n)-
modules with an additive dimension. J. Algebra, 623:1–33, 2023.
[Del16] Adrien Deloro. Symmetric powers of NatSL(2, K). J. Group Theory, 19(4):661–
692, 2016.
ON Alt(n)-MODULES WHEN n ≤ 6 23

[GAP22] The GAP Group. GAP – Groups, Algorithms, and Programming, Version
4.12.1, 2022.
[Grü22] Matthias Grüninger. Cubic action of a rank one group. Mem. Amer. Math.
Soc., 276(1356):vi+147, 2022.
[JR82] Gerald Janusz and Joseph Rotman. Outer automorphisms of S6 . Amer. Math.
Monthly, 89(6):407–410, 1982.
[Smi89] Stephen D. Smith. Quadratic action and the natural module for SL2 (k). J.
Algebra, 127(1):155–162, 1989.
[Tim01] Franz Georg Timmesfeld. Abstract root subgroups and simple groups of Lie type,
volume 95 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 2001.
Email address: barry_chin@live.com

Sorbonne Université and Université de Paris, cnrs, imj-prg, F-75005 Paris,


France
Email address: adrien.deloro@imj-prg.fr

Department of Mathematics and Statistics, California State University,


Sacramento, Sacramento, CA 95819, USA
Email address: joshua.wiscons@csus.edu

Department of Mathematics and Statistics, California State University,


Sacramento, Sacramento, CA 95819, USA
Email address: andyyu@csus.edu

You might also like