You are on page 1of 65

THE MOMENT PROBLEM FOR RANDOM OBJECTS IN A CATEGORY

WILL SAWIN AND MELANIE MATCHETT WOOD

Abstract. The moment problem in probability theory asks for criteria for when there ex-
ists a unique measure with a given tuple of moments. We study a variant of this problem for
arXiv:2210.06279v1 [math.PR] 12 Oct 2022

random objects in a category, where a moment is given by the average number of epimor-
phisms to a fixed object. When the moments do not grow too fast, we give a necessary and
sufficient condition for existence of a distribution with those moments, show that a unique
such measure exists, give formulas for the measure in terms of the moments, and prove that
measures with those limiting moments approach that particular measure. Our result applies
to categories satisfying some finiteness conditions and a condition that gives an analog of
the second isomorphism theorem, including the categories of finite groups, finite modules,
finite rings, as well as many variations of these categories. This work is motivated by the
non-abelian Cohen-Lenstra-Martinet program in number theory, which aims to calculate
the distribution of random profinite groups arising as Galois groups of maximal unramified
extensions of random number fields.

1. Introduction
The kth moment of a measure ν on the real numbers is defined to be X X k dν. The
R
classical moment problem asks whether, given a sequence Mk , whether there exists a measure
whose kth moment is Mk for all k, and whether that measure is unique. For the existence
part, a natural necessary and sufficient condition is known, whereas for uniqueness, multiple
convenient sufficient conditions are known [Sch17]. An important auxiliary result allows us
to apply the uniqueness result to determine the limit of a sequence of distributions from the
limit of their moments.
There have been recent works, in number theory, probability, combinatorics, and topol-
ogy that have used analogs of the moment problem to help identify distributions, but not
distributions of numbers–rather distributions of algebraic objects such as groups [EVW16,
Woo17, Woo19, Més20, NW21] or modules [LWZ19]. Even more recent works require dis-
tributions on more complicated categories, such as groups with an action of a fixed group
[BW17, Saw20], groups with certain kinds of pairings [LST20], groups with other additional
data [Woo18, Liu22], or tuples of groups [Lee22, NVP22]. The distributions that have been
identified this way include distributions of class groups of function fields, profinite comple-
tions of fundamental groups of 3-manifolds, sandpile groups of random graphs, and cokernels
of random matrices. These works have often required new proofs of moment problem analogs
in each different category considered.
The goal of this paper is to prove a theorem on the moment problem for distributions
of objects of a rather general category. We expect that future works of many authors will
consider distributions on new categories and we have tried to make our result as general as
possible so that it can be applied in these new settings. In a subsequent paper, we ourselves
will consider profinite groups G together with an action of a fixed group Γ and a class in
the group homology H 3 (G, Z/n), and will apply the main results of the current paper to
1
understand the distributions of class groups and fundamental groups of curves over finite
fields, and their influence under roots of unity in the base field.
In the categories in which we work, we give a criterion for there to exist a measure on
isomorphism classes of that category with a given sequence of moments, and a criterion for
the measure to be unique. We also give explicit formulas for the measure, when it exists, in
terms of the moments. This is a significant advance, as previous work has always relied on a
priori knowing a distribution with certain moments. Being able to construct a distribution
from its moments opens new avenues such as in our subsequent paper mentioned above.
Working at a high level of generality requires us to use some notation we expect to be
unfamiliar. Thus, for clarity, we begin with the statement of our main theorem in the
category of groups. In many cases of interest, the probability distributions of interest are on
the set of profinite groups (which includes finite groups), so we work there.
Let C be the category of finite groups. A level C is a finitely generated formation of
finite groups (a formation is a set of isomorphism classes of groups closed under taking
quotients and finite subdirect products). For example, the group Z/pZ generates the level
of elementary abelian p-groups. We consider the set P of isomorphism classes of profinite
groups that have finitely many (continuous) homomorphisms to any finite group. A group
X ∈ P has a pro-C completion X C which is the inverse limit of the (continuous) quotients
of X in C. We consider a topology generated on P generated by {X | X C ≃ F } as C and F
vary (and the associated Borel σ-algebra). For a finite group G, the G-moment of a measure
ν on P is defined to be Z
| Sur(X, G)|dν.
X∈P

We give a condition, well-behaved, below on a real numbers MG indexed by (isomorphism


classes of) finite groups that quantifies the notion that the MG “don’t grow too fast.” For
each surjection ρ : G → F , the quotients of G (up to isomorphism) that ρ factors through
form a poset ([Sta12, Section 3.7]). We let
P µ(ρ) = µ(F, G) be the Möbius function of that
poset. For F, G ∈ C, we let µ̂(F, G) = ρ∈Sur(G,F )/ Aut(F ) µ(ρ). Given such MG , for each
level C and finite group F , we define (what will be the formulas for our measures)
X µ̂(F, G)
(1.1) vC,F := MG .
G∈C
| Aut(G)|

Theorem 1.2. For each finite group G, let MG be a real number such that the sequence MG
is well-behaved.
(Existence): There is a measure on P with G-moment MG for each finite group G if and only if
the vC,F defined above are non-negative for each C and F .
(Uniqueness): When such a measure ν exists, it is unique and is given by the formulas
ν({X | X C ≃ F }) = vC,F .
(Robustness): If ν t are measures on P such that for each finite group G,
Z
lim | Sur(X, G)|dν t = MG ,
t→∞ X∈P

then a measure ν with moments MG exists and the ν t weakly converge to ν.


2
If one wishes to work with only finite groups, we have an analogous result, Theorem 1.6,
for each level C of finite groups. In Section 6, we give examples to show how to compute
these vC,F in practice.
We next describe the more general formalism of which Theorem 1.2 is a special case.
1.1. Main result. We first review some standard notation. Let C be a small category.
A quotient of G is a pair (F, πFG ) of an object F ∈ C and an epimorphism πFG : G → F .
Quotients (F, πFG ) and (F,′ πFG′ ) are isomorphic when there is an isomorphism ρ : F → F ′
such that πFG′ = ρπFG .
For G an object of C, quotients of G, taken up to isomorphism, form a partially ordered
set, where H ≥ F if there is h : H → F compatible with the morphisms from G.
A lattice is a partially ordered set where any two elements x, y have a least upper bound
(join) denoted x ∨ y, and a greatest lower bound (meet), denoted x ∧ y. A modular lattice
is one where a ≤ b implies a ∨ (x ∧ b) = (a ∨ x) ∧ b for all x.
We now introduce some less standard notation. For F a quotient of G, let [F, G] be the
partially ordered set of (isomorphism classes of) quotients of G that are ≥ F .
An object G ∈ C is called minimal if it has only one quotient (itself).
A set L of objects of C, stable under isomorphism, is called downward-closed if, whenever
H is a quotient of G, if G ∈ L then H ∈ L. It is called join-closed if, whenever H and F
are quotients of G and H ∨ F exists (in the partially ordered set of isomorphism classes of
quotients of G), if H ∈ L and F ∈ L then H ∨ F ∈ L.
For sets B, C of isomorphism classes of objects of C, we say B generates C if C is the
smallest downward-closed join-closed set containing B and every minimal object of C. A
level C is a set of isomorphism classes of objects of C generated by a finite set B.
An epimorphism G → F is simple if it is not an isomorphism and, whenever it is written
as a composition of two epimorphisms, one is an isomorphism (i.e. if the interval [F, G]
contains only F and G).
We can now define the type of categories to which our main theorem applies.
Definition 1.3. A diamond category is a small category C such that
(1) For each object G ∈ C, the set of (isomorphism classes of) quotients of G form a
finite modular lattice.
(2) Each object in C has finite automorphism group.
(3) For each level C of C and each G ∈ C, there are finitely many elements of C with a
simple epimorphism to G.
The terminology comes from the fact that (1) implies these categories satisfy the Diamond
Isomorphism Theorem (Lemma 2.1), a generalization of the Second Isomorphism Theorem.
We will see later, in Section 6, that the category of finite groups, the category of finite
commutative rings, and the category of finite R-modules for an arbitrary ring R, are all
examples of diamond categories. Diamond categories also include more exotic examples
such as finite quasigroups, finite Lie algebras, finite Jordan algebras, and more concrete
examples such as finite graphs with inclusions of induced subgraphs. Moreover, we will see
in Section 6 that one can make many modifications to a diamond category C to obtain a
new diamond category, such as one of objects of C with the action of a fixed finite group,
objects of C with an epimorphism to a fixed object, or objects of C with a choice of element
in some finite functorial set associated to the object.
3
We will now define the necessary notation to state our main theorem, which is on the
moment problem for measures of isomorphism classes of a diamond category. We encourage
the reader to consider the case when C is the category of finite groups first when trying to
understand the definitions and theorems. Some help with this is provided in Subsection 1.7.
For π : G → F an epimorphism, we define µ(π) = Pµ(F, G), where µ is the Möbius function
of the poset of quotients of H. We set µ̂(F, G) = π∈Epi(G,F )/ Aut(F ) µ(π).
For π : G → F an epimorphism, let Z(π) be the number of H ∈ [F, G] that satisfy the
distributive law H ∨ (K1 ∧ K2 ) = (H ∨ K1 ) ∧ (H ∨ K2 ) for all K1 , K2 ∈ [F, G]. We note that
F and G always satisfy the distributive law for all K1 , K2 , so Z(π) ≥ 2.
As in the classical theorems of analysis and probability saying that moments not growing
too fast determine a unique distribution of real numbers, we will need a notion of moments
not growing too fast. If for each G ∈ C/≃ (isomorphism class of C), we choose a real number
MG , we say the tuple (MG )G ∈ RC/≃ is well-behaved at level C if for all F ∈ C,
X X µ(F, G)
Z(π)3 MG
G∈C
| Aut(G)|
π∈Epi(G,F )

is absolutely convergent. (In Section 3 we will give formulas for µ(F, G) and Z(π) that are
easy to compute in practice.) We say (MG )G is well-behaved if it is well-behaved at level
C for all C. We will provide some explanation in Remark 1.9, after the statements of the
theorems in which the notion of well-behaved is used, to motivate why this is the rough
shape of what one needs for moments “not growing too fast”.
We define the level C quotient GC of G to be the maximal quotient of G that is in C, i.e.
the join of all level C quotients of G. By Lemma 2.8, we have that C is countable, and we
use the discrete σ-algebra for any measure on C.
Then our main theorem, below, says that for well-behaved moments, a distributions exists
with those moments if and only if certain formulas in the moments are positive, and when
a distribution exists with those moments, it is given by those certain formulas. Moreover,
this uniqueness is robust even through taking limits. We write Epi(F, G) for the set of
epimorphisms from F to G. If ν is a distribution on C/≃ , then G-moment of ν is defined to
be
Z
(1.4) | Epi(F, G)|dν.
F ∈C/≃

We also define
X µ̂(F, G)
(1.5) vC,F := MG .
G∈C
| Aut(G)|

Theorem 1.6 (Distributions on C from moments). Let C be a diamond category and let C
be a level. For each G ∈ C, let MG be a real number so that (MG )G is well-behaved at C.
(Existence): A measure νC on C with G-moment MG for all G ∈ C exists if and only if vC,F ≥ 0
for all F ∈ C.
(Uniqueness): If such a measure exists, then it is given by νC ({F }) = vC,F .
(Robustness): If νCt are measures on C such that for each G ∈ C,
Z
lim | Epi(F, G)|dνCt = MG ,
t→∞ F ∈C
4
then limt→∞ νCt ({F }) = vC,F .
1.2. The main result in the profinite case. If we use MG for each G ∈ C/≃ to define
measures νC as above for each level C, then by the uniqueness in Theorem 1.6, for levels
C ⊂ C ′ and the functor C ′ → C that sends X to X C , the measure νC ′ pushes forward to νC .
This compatible system of measures for each level gives rise to a measure ν. However, in
most of our main examples of interest, this measure ν is not on the category C, but rather
on pro-objects of C. We define a pro-isomorphism class X of C, to be, for each level C, an
isomorphism class X C ∈ C such that whenever C ⊂ C ′ we have that (X C )C ∼

= X C . (Here
′ ′ ′
X C ∈ C, so (X C )C takes the level C quotient of X C .)
The reader might wonder how this definition compares to the usual notion of a pro-
object. We will show in Lemma 5.7 that pro-isomorphism classes are naturally in bijection
with isomorphism classes of pro-objects satisfying an additional condition we call small
(Definition 5.4), which in the case where C is the category of finite groups, so that pro-
objects are profinite groups, restricts to the usual notion of small profinite groups (i.e groups
with finitely many open subgroups of index n for each n). When R is a Noetherian local
commutative ring with finite residue field, and R̂ its completion at its maximal ideal, then
pro-isomorphism classes of finite R-modules are naturally in bijection with isomorphism
classes of finitely generated R̂-modules (Lemma 5.10).
Let P be the set of pro-isomorphism classes of C. We define a topology on P with a basis
given by the open sets UC,F = {X ∈ P | X C ∼ = F } for each level C and F ∈ C.
For X a pro-isomorphism class, G ∈ C, and C a level containing G, we have that
Epi(X C , G) is independent of the choice of C. (This is because if D is the level gener-
ated by G, any epimorphism X C → G factors through (X C )D ≃ X D , and so Epi(X C , G) ∼
=
Epi(X D , G).) We define |Epi(X, G)| = Epi(X C , G) for any such C.
We take a σ-algebra (according to which we consider any measures) on P to be generated
by the UC,F for each level C and F ∈ C.
Now, we give the consequence of Theorem 1.6 for measures on P.
Theorem 1.7. Let C be a diamond category. For each G ∈ C/≃ , let MG be a real number
so that (MG )G is well-behaved.
(Existence): There exists a measure ν on P with G-moment MG for all G ∈ C if and only if for
every level C and every F ∈ C, we have vC,F ≥ 0.
(Uniqueness): If such a measure ν exists, then it is unique, and for every level C and F ∈ C, we
have ν(UC,F ) = vC,F .
If C has countably many isomorphism classes, then our σ-algebra on P is equivalent to
the Borel σ-algebra of the topology on P generated by the UC,F (because then there are only
countably many pairs C, F , so arbitrary unions of finite intersections of them are the same
as countable unions).
Theorem 1.8. Let C be a diamond category with at most countably many isomorphism
classes. For each G ∈ C/≃ , let MG be a real number so that (MG )G is well-behaved. If ν t is
a sequence of measures on P such that for all G ∈ C
Z
(Robustness): lim |Epi(X, G)|ν t = MG ,
t→∞ X∈P
then a measure ν as in Theorem 1.7 exists and as t → ∞ the ν t weakly converge to ν.
5
Remark 1.9. We explain some of the motivation for the definition of well-behaved. In both
Theorem 1.6 and Theorem 1.7, the sum
X µ̂(F, G)
MG
G∈C
| Aut(G)|

appears as the formula for a probability. For this formula to be well-defined, the sum must be
convergent, and since there is no obvious ordering of the terms, the sum must be absolutely
convergent. The condition that the MG be well-behaved is a strictly stronger condition than
this, because it requires the absolute convergence of the sum with an additional Z(π)3 factor.
This is technical restriction required by our method to prove the existence theorems, but we
do not know if it is necessary for the statement of the theorems to hold. However, for the
applications we are considering, the Z(π)3 factor is not large enough to turn the sum from
convergent to divergent, and thus causes no serious difficulty.
1.3. Motivation from number theory and prior work. When C is the category of
Fq -vector spaces for some prime power q, then the moments in our sense are just a finite
upper triangular transformation of the classical moments of the size of the group, since
| Hom(G, Fkq )| = |G|k . However, the bounds on the classical moment problem have not been
sufficient for many authors studying distributions on C, and they have had to develop their
own moment theorems. This includes work on the distribution of Selmer groups of congruent
number elliptic curves [Hea94], ranks of (non-uniform) random matrices over finite fields
[Ale99, Coo00], and 4-ranks of class groups of quadratic fields [FK06a, FK06b].
Cohen and Lenstra [CL84] made conjectures on the distribution of class groups of quadratic
number fields that have been a major motivating and explanatory force for work on class
groups, including that on 4-ranks mentioned above. Ellenberg, Venkatesh, and Westerland
[EVW16] showed that the analog of Cohen and Lenstra’s conjectures for Sylow p-subgroups
of class groups of imaginary quadratic fields, when Q is replaced by Fq (t), holds as q → ∞.
To do this, they found the limiting moments of the distributions of interest and showed
uniqueness and robustness in the moment problem for the particular distribution on abelian
p-groups conjectured by Cohen and Lenstra. This work introduced the idea that considering
averages of # Sur(−, G) was a useful way to access distributions on finite abelian groups. (For
finite abelian groups, these averages are a finite upper triangular transformation of classical
mixed moments of a tuple of integral invariants that determine the group, see [CKL+ 15,
Section 3.3].)
This approach has been followed by many papers, seeking to show function field (q → ∞)
analogs of the conjectures of Cohen and Lenstra and their many generalizations. In [Woo18],
the second author shows such an analog for real quadratic function fields, which follows from
a more refined result that gives the distribution of pairs (Pic0 , d∞ ), where Pic0 is the group of
degree 0 divisors on the associated curve and d∞ is an element of Pic0 given by the difference
between the two points above ∞. The quotient Pic0 /d∞ gives the desired class group. This
work included a moment problem uniqueness theorem for the category of pairs (A, s) where
A is an abelian p-group and a ∈ A. In [BW17], Boston and the second author, proved
function field analogs of some conjectures of Boston, Bush, and Hajir [BBH17] on p-class
tower groups of imaginary quadratic fields, which generalize the conjectures of Cohen and
Lenstra to a non-abelian analog. This work included a moment problem uniqueness theorem
for the category of p-groups with a generator inverting automorphism.
6
Liu, Zureick-Brown, and the second author [LWZ19] found the moments of the distribu-
tions of the fundamental groups of Γ-covers of P1Fq split at infinity, or equivalently of the
Galois groups of the maximal unramified extensions of totally real Γ-extensions of Fq (t), as
q → ∞. For the part of these groups of order relatively prime to q − 1 (the order of the
roots of unity in Fq (t)) and q|Γ|, they were able to construct an explicit distribution on
profinite groups with an action of Γ that had these moments. In particular their distribution
abelianizes to the distribution of finite abelian Γ-groups (i.e. groups with an action of Γ)
conjectured by Cohen and Martinet [CM90] for the class groups of totally real Γ-number
fields, and thus one can apply the uniqueness and robustness results on the moment problem
for finite abelian Γ-groups of Wang the the second author [WW21] to obtain a q → ∞ analog
of the Cohen-Martinet conjectures. The first author [Saw20] then showed a uniqueness and
robustness result on the moment problem for (not necessarily abelian) Γ-groups, which then
gave a q → ∞ theorem that the prime-to-(q − 1)q|Γ|-part of the fundamental groups of
Γ-covers of P1Fq approached the distribution constructed in [LWZ19].
For the part of the groups that has order not relatively prime to the roots of unity in
the base field, even though moments can be found with the methods of [LWZ19], in general
there is no known distribution with those moments. For the case of quadratic extensions,
Lipnowski, Tsimerman, and the second author [LST20] showed a q → ∞ function field result
for the class groups of quadratic function fields, even at primes dividing the order of roots
of unity in the base field. Their work considered additional pairings on the class groups
(a structure they call a bilinearly enhanced group), and proved uniqueness and robustness
theorem on the moment problem for bilinearly enhanced abelian groups.
In combinatorics, work to find the distribution of Sylow-p subgroups of cokernels of random
integral matrices [Woo19], the distribution of the entire cokernel of such matrices [NW21],
the distribution of Sylow-p subgroups of sandpile groups of Erdős-Rényi random graphs
[Woo17], and the distribution of Sylow-p subgroups of sandpile groups of d-regular random
graphs [Més20] has all involved applying a uniqueness and robustness theorem for the moment
problem on finite abelian groups.
Several recent works, particularly in probability, have been concerned with the distribution
of tuples of finite abelian groups. Boreico [Bor16, Corollary 3.8.5], while studying cokernels
of random integral matrices with uniform, bounded coefficients, proves uniqueness and ro-
bustness results for r-tuples (V1 , . . . , Vr ) where Vi is an Fqid -vector space for some (fixed)
di ∈ N. Nguyen and Van Peski [NVP22, Theorem 9.1] prove a uniqueness and robustness
result for r-tuples of abelian groups in order to prove new universality results on the joint
distribution of cokernels of products of random matrices. Lee [Lee22, Theorem 1.3] has
also proven a uniqueness and robustness result for r-tuples of abelian groups, in order to
apply it to several universality results for joint distributions of cokernels of random matrices
perturbed in various ways. These uniqueness and robustness results for tuples would also
follow from the argument in [WW21, Proof of Theorem 6.11] (since (R1 × · · · × Rr )-modules
correspond to tuples (M1 , . . . , Mr ), where Mi is an Ri -module). Lee [Lee22, Theorem 1.10]
also proves a uniqueness and robustness result for r-tuples of finite groups with an action of
a fixed group Γ, with an application to the joint distribution of the quotient of free profinite
groups by Haar random relations and perturbations of those relations.
Given the ever growing number of different categories that such moment problem results
are required in, we have made every attempt here to give a result that will be applicable as
7
broadly as possible. All the moment problem results mentioned above are special cases of
the main theorems in this paper (except that, unlike the current paper, the moment problem
results of [Saw20] do not always require absolute convergence of the sums involved).
In all of the above discussed work, a known explicit distribution was given (or was found by
trying various distributions) with certain moments, and so uniqueness and robustness were
the only aspects of the moment problem that were necessary. Yet there remained settings
where moments are known, but no distribution with those moments was known. In one
setting like this, the authors [SW22b] found the distribution of profinite completions of 3-
manifold groups in the Dunfield-Thurston model of random Heegaard splittings, constructing
the distribution itself from the knowledge of the moments. This required working on the
moment problem in the category of pairs (G, h), where G is a finite group and h ∈ H 3 (G, Z)
and proving theorems on existence, uniqueness, and robustness.
The current paper makes completely general the approach used in [SW22b] on reconstruc-
tion of a distribution from its moments, via explicit formulas, without any a priori knowledge
of a distribution with those moments. One key motivation is our forthcoming work, which
will give a q → ∞ theorem on the distribution of the part of fundamental groups of Γ-covers
of P1Fq that is relatively prime to q|Γ| (but not necessarily q − 1). In particular, this will give
a correction to the conjectures of Cohen and Martinet at primes dividing the order of roots
of unity in the base field, which will be provided in a short forthcoming paper [SW22a]. We
will obtain the desired distributions using the results of the current paper.
Before this paper, many papers constructed distributions as limits of distributions given
by random matrix models or models of random generators and relations. In this setting, it
is usually very easy to compute the limiting moments of these distributions. It is somewhat
more challenging, but doable, to compute the limiting distributions. As the categories got
more complicated, showing that the limiting distribution had those limiting moments be-
came increasingly more difficult. Now Theorem 1.6 can be used to not only compute the
limiting distributions of such models in a straightforward way from their moments, but also
immediately obtain that the limiting distribution has moments given by the limiting mo-
ments. Moreover, if one’s only goal is to have explicit formulas for a distribution with certain
moments, one no longer has to guess possible models that may have the desired moments,
since Theorem 1.6 gives those formulas directly.

1.4. Strategy of proof. To understand the proof of Theorem 1.6, it is helpful to first
consider the special case where (MG )G is finitely supported, i.e. MG = 0 for G not iso-
morphic to any object in some finite list G1 , . . . , Gn . It follows that, if a measure ν
with moments MG exists, then ν({G}) = 0 unless G is isomorphic to one of G1 , . . . , Gn ,
since certainly MG ≥ ν({G}) as there is always at least one epimorphism from G to it-
self. Thus, in this case, the problem of existence and uniqueness of a measure can be
expressed as thePn existence and uniqueness of a solution to the finitely many linear equa-
tions MGi = j=1 |Epi(Gj , Gi )|ν({Gj }) in finitely many variables ν({Gj }), together with
the inequalities ν({Gj }) ≥ 0.
The linear equations can be expressed concisely in terms of a matrix A with Aij =
|Epi(Gj , Gi )|. The necessary observation to handle this case is that A is invertible. In
fact, it is possible to order G1 , . . . , Gn so that there are no epimorphisms from Gj to Gi if
j < i. This forces A to be upper-triangular, with nonzero diagonal, so an inverse always
8
exists. Thus the linear equations always have a unique solution, and a measure exists if and
only if that solution has ν({Gj }) ≥ 0 for all j.
A key observation that powers our proof is that one can explicitly calculate this inverse
µ̂(Gi ,Gj )
in great generality. In fact, the entry A−1 ij is given by the formula |Aut(Gj )| (Lemma 4.1).
Using this formula, we immediately obtain a guess vC,F for the unique measure with a
given moments (in Equation (1.4)). Proving that guess is correct, both that vC,F produces a
measure with those moments (Existence) and produces the only measure with those moments
(Uniqueness) requires only that we exchange the order of summation in a certain sum.
Using our assumption that the quotients of any object form a finite modular lattice, we note
that µ̂(Gi , Gj ) vanishes unless there exists a semisimple epimorphism from Gj to Gi . This
allows us to restrict attention to semisimple morphisms in sums involving µ̂. Furthermore, we
can calculate the exact value of µ̂ using known results on the structure of finite complemented
modular lattices. From this we can prove the sum relevant for (Uniqueness) is absolutely
convergent using the well-behavedness hypothesis, justifying the exchange of summation.
The sum relevant for (Existence), on the other hand, is trickier. It is not absolutely con-
vergent, under well-behavedness or even much stronger assumptions, in reasonable categories
like the category of finite groups. Instead we need a more intricate argument where we break
each sum into a series of multiple sums and exchange them one at a time, proving absolute
convergence only of certain restricted sums, which does follow from well-behavedness.
More precisely, we can place any level C into a series of levels C1 ⊂ C2 ⊂ · · · ⊂ Cn = C
where C1 consists (in the category of groups) only of the trivial group or (in general) only
of objects with no nontrivial quotients. We can express a sum over objects of level C as a
sum over objects H1 of level C1 , then over objects H2 of level C2 whose maximal level-C1
quotient is isomorphic to H1 , i.e. such that H2C1 ∼ = H1 , then over objects H3 of level C3 with
C2 ∼
H3 = H2 , and so on. Crucially, we can choose Ci to be “close together" in the sense that
for every H ∈ Ci , the natural epimorphism H → H Ci−1 is semisimple. This, again, lets us
restrict attention to semisimple morphisms, which are controlled by the well-behavedness
condition, when exchanging a given pair of sums.
The proof of (Robustness) requires, instead of an exchange of sums, the exchange of a
sum with a limit. We apply a similar strategy, breaking the sum into several smaller sums,
each over semisimple morphisms, and exchanging each sum with the limit in turn, using a
dominated convergence argument.
The approach of attacking problems in level C by an inductive argument using a sequence
of levels C1 , . . . , Cn , where each element of Ci has a semisimple morphism to an element of
Ci−1 , is something that we expect to be relevant to future problems about random groups and
other random objects. As a first example, for the problem of efficiently sampling from the
probability distributions we have constructed, we believe the best method will typically be
to choose a random object of level C1 , then a random lift to C2 , and so on, which reduces the
problem to a series of steps of sampling from a conditional probability distribution, which
will often be given by a simple probability distribution on an explicit set. For the exis-
tence problem, it is the combination of this inductive strategy with our Möbius calculations
which together allow us to restrict entirely to semisimple morphisms for the most difficult
calculations in the proof.
1.5. Organization of the paper. In Section 2, we show basic facts about modular lattices
and levels that will be used throughout the paper. In Section 3, we show how the Möbius
9
function and Z(π) can be given by simple formulas, and give examples in the categories of
groups, rings, and modules. In Section 4, we prove our main result, Theorem 1.6, which
operates at a single level. In fact, we prove slightly stronger results, in which we give the
minimal hypotheses separately for each piece of the theorem. In Section 5, we prove the
results regarding distributions on pro-isomorphism classes (Theorems 1.7 and 1.8), relate
pro-isomorphism classes to classical pro-objects, and give examples in groups and modules
to show what these are concretely. In Section 6, we give examples of the calculation of
measures from moments, including showing that moments of size O(|G|v ) are well-behaved
for any real v in the categories of groups or R-modules. Our examples give a blueprint for
how one evaluates, in practice, the sums defining vC,F and well-behavedness. We also show
many examples of diamond categories, including simple ones like (the opposites of) finite
sets and finite graphs, and how various modifications to diamond categories lead to new
diamond categories.

1.6. Further notation and conventions. We sometimes refer to objects in a level C as


level-C objects.
We write C/≃ for the set of isomorphism Pclasses of objects of C. Throughout the paper, we
sum over isomorphism classes. We write G∈C/≃ f (G), where f is some function of objects of
C that is invariant under isomorphism and we intend
P G to be an object of C. This is a slight
abuse of notation, which more precisely means [G]∈C/≃ f (G), where [G] is an isomorphism
class.
We write N for the non-negative integers.
We write f (x) ≪ g(x) to mean that there exists a constant C such that f (x) ≤ Cg(x) for
all values of x. The implicit constant may depend on certain parameters, and we will say
what it depends on.
We refer to quotients by the object, and leave the morphism implicit, whenever this does
not create ambiguity. If F is a quotient of G, we denote the implicit epimorphism by πFG .
We refer to the isomorphism classes of quotients by a choice of one of their elements.
Our notation [F, G] for the partially ordered set of quotients of G that are ≥ F agrees with
the usual notion of an interval in lattice theory. For H ∈ [F, G], note that there is a unique
map H → F implied by H ≥ F that is compatible with the G-quotient structure, and this
map is an epimorphism. A priori the notation [F, H] might either refer to the interval in
[F, G] or the the partially ordered set of quotients of H that are ≥ F , but we can see these
two options are in natural correspondence, as a quotient of H is also a quotient of G (using
the given map G → H), and the notions of isomorphism correspond. For F, H ∈ C, we only
write F ∨ H or F ∧ H when we already have a G-quotient structure on F and H for some G,
and we mean to take these operations in the partially ordered set of (isomorphism classes)
of quotients of G. If we need to remind the reader what G is, we write F ∨G H or F ∧G H.
The joins and meets in the lattice guaranteed by Definition 1.3 (1) have an interpretation
in terms of products and coproducts, when the latter exist. When F and H are quotients of
G, and the pushout of F and H over G exists, then it is the meet F ∧ H. When the meet
F ∧ H exists and the fiber product of F and H over F ∧ H exists and is a quotient of G,
the fiber product is the join F ∨ H. However, we will sometimes want to consider categories
where fiber products may not exist.
10
1.7. Translation of terms for the category of groups. We now summarize the meaning
of some of the notation and results in this section and the next section in the special case
where C is the category of finite groups, as a reference for readers.
For G → F a morphism of groups, the kernel is a group with an action of G by conjugation,
i.e. a G-group. We say a G-group is simple if it has no nontrivial proper normal G-invariant
subgroup. Using this notation, we see that a morphism G → F of finite groups is simple if
and only if the kernel is a finite simple G-group.
A morphism π : G → F is semisimple if and only if the kernel is a product of finite simple
G-groups, i.e. a product of finite normal subgroups with no G-invariant normal subgroup.
Here the “only if" direction is because for F a G-invariant normal subgroup of ker π and H
a complement in the sense that F ∩ H = {1} and F H = ker π, we have ker π ∼ = F ×H
as G-groups, and we can repeat this process until no nontrivial proper G-invariant normal
subgroups are left, thereby writing ker π as a product of finite simple G-groups and the “if"
direction follows from Lemma 2.4. Lemma 2.4 also includes the fact that a sub-G-group or
quotient G-group of a product of finite simple G-groups is itself the product of finite simple
G-groups.
The dimension of a group G is the length of its Jordan-Hölder filtration. The dimension
of a surjection G → F is the length of the initial segment of a Jordan-Hölder filtration of
G which includes the kernel of the surjection as one of its members. The dimension of a
semisimple morphism is the number of finite simple G-groups we can write the kernel as a
product of.
The diamond isomorphism theorem Lemma 2.1 in the case of groups follows from the
second isomorphism theorem.
1.8. Acknowledgments. The authors would like to thank Akshay Venkatesh for helpful
conversations related to this project. Will Sawin was supported by NSF grant DMS-2101491
while working on this paper. Melanie Matchett Wood was was partially supported by a
Packard Fellowship for Science and Engineering, NSF CAREER grant DMS-1652116, and
NSF Waterman Award DMS-2140043 while working on the paper. She was also a Radcliffe
Fellow during part of this work, and thanks the Radcliffe Institute for Advanced Study for
their support.
2. Preliminaries on lattices and categories
Throughout this section we let C be a diamond category.
The fundamental result about modular lattices is the following.
Lemma 2.1 (Diamond Isomorphism Theorem, [Bir67, Ch. I, Thm. 13 on p. 13]). If G and
H are two elements of a modular lattice, then the intervals [G ∧ H, H] and [G, G ∨ H] are
isomorphic as lattices. The isomorphism is given by the map · ∨ G : [G ∧ H, H] → [G, G ∨ H]
and the inverse map · ∧ H : [G, G ∨ H] → [G ∧ H, H].
We say an epimorphism F → H is semisimple if the lattice [F, G] is complemented (i.e.
for every H in [F, G], there exists K ∈ [F, G] such that H ∨ K = G and H ∧ K = F , i.e. K
is a complement of H). (When we refer to a morphism as simple or semisimple, we always
mean that it is an epimorphism.)
Remark 2.2. The definition of a complemented lattice is a natural generalization of the
notion of a semisimple representation, which can be defined as representations for which
11
each invariant subspace V has an invariant complement W , i.e. an invariant subspace such
that V ∩ W = {0} and V + W is the whole space.
A key advantage of the modular lattice property is that it gives the set of semisimple
morphisms a number of useful properties. A basic fact in lattice theory is the following.
Lemma 2.3. Let L be a finite modular lattice. Let 1 be the maximum element and 0 the
minimum element of L. The following are equivalent:
(1) L is complemented.
(2) Every interval in L is complemented.
(3) 1 is the join of the minimal non-zero elements of L.
(4) 0 is the meet of the maximal non-one elements of L.
Proof. For the first three, this is [Bir48, VIII, Corollary to Theorem 1 on p. 114], and the
equivalence between (1) and (4) follows from the equivalence between (1) and (3) and the
fact that the complemented modular lattices are manifestly stable under duality. 
Translated into morphisms, Lemma 2.3 immediately gives the following.
Lemma 2.4. Let π : G → F be an epimorphism in C. The following are equivalent:
(1) π is semisimple.
(2) For every factorization π = π1 ◦ π2 with π1 , π2 epimorphisms, π1 and π2 are semisim-
ple.
(3) G is the join of F1 , . . . , Fn with Fi → F simple.
(4) F is the meet of G1 , . . . , Gn with G → Gi simple.
G
For fixed H ≤ G, (i.e. for H a quotient of G via πH ),
(
X X 1 if G = H
(2.5) µ(H, F ) = µ(F, G) =
F ∈[H,G] F ∈[H,G]
0 if G 6= H,

by the defining property of the Möbius function and another well-known property ([Sta12,
Proposition 3.7.1] for f the characteristic function of H).
Semisimple morphisms play a crucial role in the study of the Möbius function.
Lemma 2.6 (Philip Hall). Let G be an object of C and F a quotient of G. If G → F is not
semisimple, then µ(F, G) = 0.
Proof. This follows from [Rot64, Corollary p. 349] but we give the simple proof below.
Let Λ be the sublattice of quotients H of G such that G → H is semisimple. By Lemma 2.4,
we have that Λ is meet-closed, and thus for every quotient H of G, there exists H ss , the
minimal semisimple quotient of G that is ≥ H. Then we have
X X
0= µ(H, G) µ(F, K) (each interior sum vanishes by defn. of µ)
H∈Λ K∈[F,H]
X X
= µ(F, K) µ(H, G) (rearrange sum)
K∈[F,G] H∈Λ∩[K,G]
i.e., H∈[K ss ,G]

= µ(F, G),
12
where the last equality holds because the inner sum vanishes unless K ss = G by (2.5), and
K ss = G if and only if K = G (as if K < G then K is less than some simple quotient of
G). 

Let π : G → F be an epimorphism. Let H be the join of all the minimal non-F elements
in the interval [F, G]. We call the epimorphism G → H the radical of π. Since H is the join
of the minimal non-F elements, H → F is semisimple, and it is the maximal semisimple
morphism through which π factors. We call H → F the semisimplification of π.
The finiteness in Definition 1.3 (1) ensures that we can write any epimorphism π as a
finite composition of simple morphisms. The modularity assumption means that, when we
do this, the number of simple morphisms depends only on π [Bir48, V, Theorem 3 on p. 68].
Definition 2.7. We define the dimension of π to be the number of simple morphisms it is a
composition of (or 0 if π is an isomorphism) and the dimension of G to be the dimension of
its morphism to its unique minimal quotient.
Note the dimension of an epimorphism G → H is the difference between the dimension of
G and the dimension of H. Also, if A < B, them dim A < dim B.
We will use many inductive arguments, which are all a form of induction on the dimension.
The first one is:
Lemma 2.8. For a level C, and F ∈ C, and d a natural number, there are finitely many
G ∈ C with an epimorphisms G → F of dimension d.
Proof. This follows by induction on d, with Definition 1.3 (3), and the induction step using
Definition 1.3 (3) and the fact that any epimorphism G → F of dimension d factors (in at
least one way) as G → G′ → F with G → G′ of dimension 1 (i.e. simple) and G′ → F of
dimension d − 1. 

The following will be very useful in inductive arguments.


Lemma 2.9. Let C be a level generated by a finite set B, and let B̄ ⊂ B be a subset containing
every proper quotient of an element of B, and C¯ the level generated by B̄. For G ∈ C, the
¯
natural morphism G → GC is semisimple.
¯
Proof. It suffices to show that the set of G with G → GC semisimple contains B, is downward-
closed, and join-closed and contains all minimal elements, as C is the smallest set with those
properties.
¯
To see that G ∈ B has G → GC semisimple, note that any maximal proper quotient H
¯ Since GC¯ is the maximal level-C¯ quotient of G, it must be ≥ H, thus must be
of G is in C.
¯
equal to G or H. Then [GC , G] is either [G, G] or [H, G], and either lattice is complemented,
¯
so G → GC is semisimple.
¯
To see, if G → GC is semisimple and H is a quotient of G, that H → H C̄ is semisimple,
¯
note first that GC ∧ H is level-C¯ and is a quotient of H, and thus is ≤ H C̄ . So [H C̄ , H] is an
¯ ¯ ¯ ¯
interval of of [GC ∧ H, H]. By Lemma 2.1, [GC ∧ H, H] is isomorphic to [GC , GC ∨ H], which
¯
is an interval of the complemented modular lattice [GC , G]. Since intervals in complemented
¯
modular lattices are complemented (Lemma 2.3), [H C̄ , H] is complemented and thus H → H C
is semisimple.
13
¯ ¯
To see, if G → GC and H → H C̄ are semisimple, that G ∨ H → (G ∨ H)C is semisimple,
¯ ¯
note that GC ∨ H C̄ is level-C n−1 and is a quotient of G ∨ H, and thus is ≤ (G ∨ H)C . Thus
¯
by Lemma 2.4, it suffices to show G ∨ H → GC ∨ H C̄ is semisimple.
¯ ¯ ¯ ¯ ¯ ¯
We have [GC ∨ H C , H C̄ ∨ G] = [GC ∨ H C , GC ∨ H C ∨ G], which, by Lemma 2.1 is isomorphic
¯ ¯ ¯
to [(GC ∨ H C̄ ) ∧ G, G]. The interval [(GC ∨ H C̄ ) ∧ G, G] is an interval in [GC , G] and thus is
¯ ¯ ¯ ¯
complemented. Thus [GC ∨ H C̄ , H C̄ ∨ G], and similarly [GC ∨ H C , GC ∨ H] are complemented.
¯ ¯
So both of G ∨ H C̄ and GC ∨ H are the join of minimal non-GC ∨ H C̄ elements (by Lemma
¯
2.3), which all lie in [GC ∨ H C̄ , G ∨ H], and thus the meet of all such elements is ≥ (G ∨
¯ ¯
H C̄ ) ∨ (GC ∨ H) = G ∨ H and thus must be equal to G ∨ H. This makes [GC ∨ H C̄ , G ∨ H]
¯
complemented and G ∨ H → GC ∨ H C̄ semisimple by Lemma 2.3.
¯
Clearly, all minimal G have G → GC semisimple, and this proves the lemma as originally
outlined. 
We have some basic facts about the interaction of lattice operations of taking the level-C
quotient.
Lemma 2.10. For any level C and A ≤ B objects of C, we have
(1) AC ≤ B C ,
(2) (A ∨ B C )C = B C , and
(3) A ∧ B C = AC .
Proof. Since AC ≤ A ≤ B, and AC is level-C, we have AC ≤ B C by definition of B C . Thus
B C = (B C )C ≤ (A ∨ B C )C ≤ B C , which gives the second statement. We have AC ≤ A, B C , so
AC ≤ A ∧ B C . Since A ∧ B C ≤ B C , we have that A ∧ B C is level-C. Then since A ∧ B C ≤ A,
we have that A ∧ B C ≤ AC , and the third statement follows. 
Lastly, we have a simple observation about the concept of a well-behaved tuple.
Lemma 2.11. Let (MG )G and (MG′ )G be tuples of nonnegative real numbers such that MG′ =
O(MG ) for all G, i.e. there exists a constant c such that MG′ ≤ cMG for all G.
If (MG )G is well-behaved at level C, then (MG′ )G is as well, and if (MG )G is well-behaved,
then (MG′ )G is as well.
Proof. This follows immediately from the definition since
X X |µ(F, G)| X X |µ(F, G)|
Z(π)3 MG′ ≤ Z(π)3 cMG
G∈C
| Aut(G)| G∈C
| Aut(G)|
π∈Epi(G,F ) π∈Epi(G,F )
X X |µ(F, G)|
=c Z(π)3 MG < ∞. 
G∈C π∈Epi(G,F )
| Aut(G)|

3. Calculation of the Möbius function


In this section, we explain how to calculate the Möbius function so as to obtain explicit
formulas from our main theorems in particular cases. We also explain how to estimate
the sum appearing in the definition of well-behaved, so as to check the well-behavedness
assumption. In view of Lemma 2.6, we may restrict attention to semisimple morphisms
π : G → H. For such morphisms, [H, G] is by definition a finite complemented modular
lattice. There is a powerful classification theorem for finite complemented modular lattices.
14
We begin by reviewing the classification theorem, then explain how to compute the Möbius
function in each case, and finally show how to apply these formulas in the categories of finite
groups, finite modules, and finite rings.
We begin by giving some examples of finite complemented modular lattices.
Example 3.1. The simplest example of a complemented modular lattice is the lattice of
subsets of a set, with the ordering given by inclusion. The complements here are the usual
notion of the complement of the set. The subsets of a finite set form a finite complemented
modular lattice.
Example 3.2. Another example of a complemented modular lattice is the subspaces of a
vector space, again ordered by inclusion. The complements of a subspace are the usual notion
of complement in linear algebra. The subspaces of a finite-dimensional vector space over a
finite field form a finite complemented modular lattice.
There is a relation between Examples 3.1 and 3.2: A one-dimensional vector space over
a field has only two subspaces, itself and the zero subspace, so the lattice of subspaces has
two elements. The product of n copies of this two-element lattice is the lattice of subsets of
an n-element set. A construction of finite complemented modular lattices which generalizes
both of the previous two cases is to take a finite product of a sequence of lattices, each
element of which is the lattice of subspaces of some vector space over some finite field.
Not all finite complemented modular lattices arise this way, so we need two additional
constructions.
Example 3.3. First, note that the lattice of subspaces of F2q consists of a minimal element
(the zero-dimensional subspace), a maximal element (the two-dimensional subspace), and
q + 1 incomparable elements (the one-dimensional subspaces). We can generalize this by
taking the lattice consisting of a minimal element, a maximal element, and q+1 incomparable
elements, where q is not necessarily a prime power.
Example 3.4. Second, note that the lattice of subspaces of F3q consists of a minimal element,
a maximal element, and elements corresponding to one-dimensional and two-dimensional
subspaces. The one-dimensional subspaces are naturally in bijection with points of the finite
projective plane P2 (Fq ), and the two-dimensional subspaces are naturally in bijection with
lines in this plane. We can generalize this by considering finite projective planes which do
not necessarily arise from vector spaces over finite fields. Since these are exactly the finite
projective planes that don’t satisfy Desargues’s theorem, they are called non-Desarguesian
planes. The smallest examples have 91 points and 91 lines.
We form a lattice from a finite projective plane by choosing the set of elements of the lattice
to consist of one minimal element, one maximal element, the set of points of the plane, and
the set of lines of the plane, with a partial ordering defined so that the minimal element is
≤ every other element, the maximal element is ≥ every other element, a point is ≤ a line if
the point lies on the line, and every other pair of elements is incomparable.
Every finite projective plane, including non-Desarguesian ones, has a number q, called the
order, so that every point lies on q + 1 lines, every line contains q + 1 points, the number of
points is q 2 + q + 1, and the number of lines is q 2 + q + 1 [AS68, Theorem 1 on p. 17].
These constructions, combined with subspaces of a finite projective space, can be produce
to construct every finite complemented modular lattice.
15
Theorem 3.5 ([Bir48, VIII, Theorem 6 on p. 120]). Every finite complemented modular
lattice can be expressed uniquely as a finite product of lattices of one of the following three
forms:
• The lattice of linear subspaces of an n-dimensional space over a finite field Fq , for
some n ≥ 1 and prime power q.
• A lattice consisting of one minimal element, one maximal element, and q + 1 incom-
parable elements, for q ≥ 2 not a prime power.
• A lattice consisting of one minimal element, one maximal element, the set of points
in P , and the set of lines in P , for P a non-Desarguesian finite projective plane, with
≤ defined as in Example 3.4.
For π : G → H semisimple, let ω(π) be the number of terms appearing in this product
representation of [H, G], and we call these the simple factors of the lattice.
Lemma
Qω 3.6. Let L be a finite complemented modular lattice, and write L as a product
i=1 Li as in Theorem 3.5.
Define q1 , . . . , qω and n1 , . . . , nω as follows: For each i, if Li is the lattice of linear subspaces
of Fqn , set qi = q and ni = n. (We can take qi arbitrary if ni = 1.) If Li has one minimal
element, one maximal element, and q + 1 incomparable elements, set qi = q and ni = 2. If
Li is obtained from a finite projective plane of order q, set qi = q and ni = 3.
Then for 0 and 1 the minimal and maximal elements of L, we have
ω
Y (ni )
µ(0, 1) = (−1)ni qI 2 .
i=1

Note that if ni = 1 then the formula does not depend on qi .


Proof. Since Möbius function is multiplicative in a product of posets, we may reduce to the
case ω = 1, so assume L is one of the three forms of Theorem 3.5. If L is the lattice of
subspaces of a linear space, this is a classical result of Philip Hall [Rot64, Equation on p.
352].
In the other two cases, this is a straightforward explicit calculation, and is likely also
classical. 
Let us now see how this classification applies to the categories of groups, rings, and modules
(which we will check in Section 6 each form a diamond category).
Lemma 3.7. Let M be a module for a ring R. The lattice of submodules of M is comple-
mented if and only if M is a product of simple R-modules.
In this case, there must exist finite simple R-modules M
Q1ω, . . . , Mω , pairwise non-isomorphic,
and natural numbers n1 , . . . , nω ≥ 1, such that M ∼ = i=1 Mini . Let qi be the order of the
field of endomorphisms of Mi . Then the lattice of submodules of M is the product from i
from 1 to ω of the lattice of subspaces of Fnqii . The dual of the lattice of submodules is also
isomorphic to the same product.
Proof. By Lemma 2.3, the lattice of quotients of M is complemented if and only if M is a
join of simple modules, i.e. a quotient of the product of simple modules, but every quotient
of a (finite) product of simple modules is itself a product of simple modules by induction.
If M is a product of simple modules, then we can take
Qωone representative from each isomor-
phism class appearing in the product to write M ∼ = i=1 i M ni
. We can calculate the lattice
16
of ωi=1 Mini by noting that, by Jordan-Hölder,
Q
of submodules Q each submodule is isomorphic
to i=1 Miei for some ei ≤ ni , that injections ωi=1 Miei × ωi=1 Mini are parameterized by
Qω Q
tuples of full rank ni × ei matrices over Fqi , andQ two matrices give the same submodule if
and only if they can be related by elements of ωi=1 GLei (Fqi ), i.e. if and only if they have
the same image.
Finally, it is easy to check that this product lattice is isomorphic to its dual, by picking a
nondegenerate bilinear form on Feqi . 
Lemma 3.8. Let C be the category of finite modules over a fixed ring R. Then a morphism
π : M → N of finite R-modules is semisimple if and only if ker π is a product of simple
R-modules.
In this case, there must exist finite simple R-modules MQ1 , . . . , Mω , pairwise non-isomorphic,
and natural numbers n1 , . . . , nω ≥ 1, such that ker π ∼ ω
= i=1 Mini . Let qi be the order of the
field of endomorphisms of Mi . Then the lattice [N, M] is the product from i from 1 to ω of
(ni )
the lattice of subspaces of Fnqii . Thus µ(N, M) = ωi=1 (−1)ni qi 2 .
Q

Proof. Every module K ∈ [N, M] defines a submodule of ker π, the kernel of the natural
morphism M → K. Conversely, every submodule S ⊆ ker π gives a module M/S ∈ [N, M].
This bijection sends inclusion of submodules to the ≥ relation on [N, M], so π is semisimple
if and only if the lattice of submodules of ker π is complemented. By Lemma 3.8, this occurs
only if ker π is a product of finite simple R-modules. The lattice [N, M] is dual to the lattice
of submodules of ker π, which by Lemma 3.7 has the stated description. The Möbius formula
then follows from Lemma 3.6

Lemma 3.9. Let C be the category of finite rings. Then a morphism π : R → S of finite
rings is semisimple if and only if ker π is a product of finite simple R-bi-modules.
In this case, there must exist finite simple R-bi-modulesQM1 , . . . , Mω , pairwise non-isomorphic,
and natural numbers n1 , . . . , nω ≥ 1, such that ker π ∼ ω
= i=1 Mini . Let qi be the order of the
field of endomorphisms of Mi . Then the lattice [S, R] is the product of i from 1 to ω of the
(ni )
lattice of subspaces of Fnqii . Thus µ(S, R) = ωi=1 (−1)ni qi 2 .
Q
In particular, if π is a local morphism of finite local commutative rings with residue field
κ, then π is semisimple if and only if ker π ∼ = κn for some n, and, then [S, R] is isomorphic
n
to the lattice of subspaces of κn , so µ(S, R) = (−1)n |κ|( 2 ) .
Proof. Any ring T ∈ [S, R] defines a sub-R-bi-module of ker π, its kernel, and conversely, any
sub-R-bi-module of ker π gives a two-sided ideal of R, whose quotient is a ring in [S, R]. This
bijection sends inclusion of submodules to the ≥ relation on [S, R], so π is semisimple if and
only if the lattice of sub-bi-modules of ker π is complemented. Since a bi-module is simply
an R ⊗ Rop -module, we can apply Lemma 3.7 to show that this happens if and only if ker π
is a product of simple bi-modules. The lattice [S, R] is dual to the lattice of submodules of
ker π, which by Lemma 3.7 has the stated description, and the Möbius formula then follows
from Lemma 3.6
For commutative rings, the left and right actions of R on R are identical, so the same is
true for ker π, so we can express ker π and every submodule of it as modules rather than
bimodules. For local commutative rings with residue field κ, the unique simple module is κ,
and so the expression ωi=1 Mini simplifies to κn , and similarly for the Möbius expression. 
Q
17
Lemma 3.10. Let C be the category of finite groups. Then a morphism π : G → H of finite
groups is semisimple if and only if ker π is a product of finite simple G-groups.
In this case, there must exist pairwise non-isomorphic finite simple abelian G-groups,
V1 , . . . , Vn , and finite simple
∼ Qm G-groups N1 , . . . , Nm , as well as exponents e1 , . . . , en ≥
Qn non-abelian
ei
1, such that ker π = i=1 Vi × i=1 Ni . Let qi be the field of endomorphisms of Vi for i
from 1 to n. Then the lattice [H, G] is the product for i from 1 to n of the lattice of subspaces
of Feqii together with m copies of the two-element lattice.
(ei )
Thus µ(H, G) = (−1)m ni=1 (−1)ei qi 2 .
Q

Proof. The first claim is shown in Section 1.7. Q


To obtain an expression in terms of a product ni=1 Viei × m
Q
i=1 Ni , we take one represen-
tative from each isomorphism class of simple G-groups appearing in the product, and divide
into abelian and nonabelian. The nontrivial fact to check is that each isomorphism class of
nonabelian G-groups can only occur once. Indeed, if K1 × K2 is a sub G-group of ker π with
K1 , K2 isomorphic, then since the action of K1 ⊂ G by conjugation on K2 is trivial, then
the action of K1 on K1 by conjugation must be trivial, forcing K1 to be abelian, so K2 is
abelian as well.
To calculate the lattice, we note that every sub-G-group of ni=1 Viei × m
Q Q
i=1 Ni is the
product for each i from 1 to n of a sub G-group of Viei with the product for each i from
1 to m of either Ni or the trivial group [LW20, Lemma 5.5]. It follows the lattice is the
product of, for each i from 1 to n, the lattice of sub G-groups of Viei , with m copies of the
two-element lattice. By Lemma 3.7 applied to Z[G], the lattice of sub-G-groups of Viei is the
lattice of subspaces of Feqii , verifying the stated description of the lattice.
The Möbius formula then follows from Lemma 3.6. 
We now explain the meaning of Z(π) and how it can be used to bound the sum we will
need.
Lemma 3.11. For π : G → H a semisimple epimorphism, we have Z(π) = 2ω(π) .
Proof. In a product of lattices L1 ×L2 , an element (F1 , F2 ) ∈ L1 ×L2 satisfies the distributive
law in the definition of Z(π) if and only if F1 and F2 do, since both joins and meets in a
product are computed elementwise. Thus it suffices to show that Z(π) = 2 for each of the
lattices appearing in Theorem 3.5. This can easily be checked as 0 and 1 satisfy the law,
but for any other H we can take K1 and K2 to be two different complements of H and this
violates the distributive law. 
Lemma 3.12. Let C be a diamond category.
(1) For π : G → H semisimple , we have
X
|µ(K, G)| ≤ |µ(H, G)|Z(π)3 .
K∈[H,G]

(2) If C has the property that for any semisimple morphism G → H of level−C objects,
the number of simple factors in the lattice [H, G] is bounded by a bound depending
only on H and C, then for π : G → H with G of level C, we have
X
|µ(K, G)| ≪ |µ(H, G)|
K∈[H,G]
18
with the implicit constant depending only on H and C.
(3) If C has the property that for any semisimple morphism G → H of level−C objects,
the number of simple factors in the lattice [H, G] which are not isomorphic to the two-
element lattice is bounded by a bound depending only on H and C, then for π : G → H
with G of level C, we have
X
|µ(K, G)| ≪ |µ(H, G)|Z(π)
K∈[H,G]

with the implicit constant depending only on H and C.


From Lemmas 3.8 and 3.9, one can check that the condition of case (2) happens, for
example, for finite modules over a ring and finite local commutative rings. From Lemmas
3.9 and 3.10, one can check that the condition of case (3) holds for finite groups and for
arbitrary finite rings. The advantage of these statements is that they allow us to lower the
exponent of Z(π) in the hypotheses of our existence result.
If for each G ∈ C, we choose a real number MG , we say the tuple (MG )G ∈ RC is behaved
at level C if for all F ∈ C,
X X µ(F, G)
Z(π)a MG
G∈C
| Aut(G)|
π∈Epi(G,F )

is absolutely convergent, where either a = 3; C satisfies the hypotheses of Lemma 3.12 (2)
and a = 0; or C satisfies the hypotheses of Lemma 3.12 (3) and a = 1.
Proof. We handle case (1) first. Both sides depend only on the lattice [H, G] and are multi-
plicative under direct products of lattices. Since [H, G] is complemented modular, we may
reduce to checking each of the three forms of Theorem 3.5.
In the case of the lattice of subspaces of a n-dimensional vector space over Fq , the left side
is the sum over d from 0 to n of the number of (n − d)-dimensional subspaces of Fq times
d d
q (2) , and the right side is 23 q (2) by Lemmas 3.6 and 3.11. Since the number of subspaces of
dimension d is nd q , using the q-binomial theorem, the left side is

n   n−1 ∞ ∞
n d n n n n
q (2 ) = q ( 2 ) (1 + q −j ) ≤ q ( 2 ) (1 + q −j ) ≤ q ( 2 ) (1 + 2−j ) < q ( 2 ) 8,
X Y Y Y

d=0
d q j=0 j=0 j=0

giving the desired inequality.


For the other two cases, we note that the number of points of the lattice of each dimension
is given by the same formula in terms of n = 2, 3 and q, and the Möbius function is given
by the same formula in terms of q and the dimension, so the same calculation applies. This
finishes case (1).
In case (2), our assumption implies that Z(π) is bounded in terms of H and C, and so (2)
follows immediately from (1).
In case (3), we return to the argument of (1). We above showed that for any simple
complemented modular lattice,
X
|µ(x, 1)| ≤ 8|µ(0, 1)|.
x∈[0,1]
19
For the two-element lattice, the statement above is true with a constant of 2 instead of 8.
Multiplying these two statements, we see that
X
|µ(K, G)| ≤ |µ(H, G)|Z(π)4b ,
K∈[H,G]

where b is number of simple factors of [H, G] with more than two elements. By assumption,
4b can be absorbed into our implicit constant. 

The assumption of part (3) of Lemma 3.12 is satisfied for most examples arising in ordinary
mathematics, such as groups, rings, and modules. However, there are cases where it is not
satisfied. We give an example:
Example 3.13. Let C be the category whose objects are pairs consisting of a finite-dimensional
vector
L space V over F2 and a set S of nontrivial subsets of V such that the natural map
W ∈S W → V is an isomorphism, where morphisms (V1 , S1 ) → (V2 , S2 ) are given by maps
f : V1 → V2 such that f (W ) ∈ S2 ∪ {0} for all W ∈ S and if f (W ) = f (W ′) then W = W ′
or f (W ) = 0.
We can think of these as graded vector spaces, but instead of graded by integers, they are
graded by an arbitrary finite set.
It is not so hard to check a morphism is an epimorphism if and only if f is surjective, and
the poset of quotients of (V, S) is given by the product over W ∈ S of the poset of quotients
of W , and thus is a finite modular lattice. One can similarly check the other assumptions:
automorphism groups are subgroups of GLn (F2 ) and thus finite, and a simple morphism either
raises the dimension of one element of S by one or adds a new one-dimensional summand
in S, so there are finitely many simple morphisms.
By our description of the lattice of quotients, we say every object is a meet of its one-
dimensional quotients, and every one-dimensional object of C is isomorphic to (F2 , {F2}),
so every object is level-C for C = {(F2 , {F2 })}. Similarly, this meet property implies every
morphism is semisimple.
Thus the assumption of part (3) of Lemma 3.12 is not satisfied even for H = (0, ∅),
since [G, H] can be a product of an arbitrary finite number of lattices of quotients of finite-
dimensional vector spaces over F2 .

4. Proof of the main theorem


In this section, we prove Theorem 1.6. For some parts of the theorem, we state and prove
slightly stronger versions.
Throughout this section, we assume C is a diamond category. Any further assumptions
will be stated explicitly in our results, and in the theorems we will remind the reader of the
diamond assumption.
To compactify our formulas below, we define TF,H := µ̂(H, F ) (note the order switches),
and for π ∈ Epi(G, F ) we define Tπ := µ(π)/| Aut(F )|. We also let SG,F := | Epi(G, F )|/| Aut(F )|,
and I := C/≃ (isomorphism classes). Finally, whenever we consider (MG )G ∈ RI , we let
mG := MG /| Aut(G)|.
The main step in the proof of (Uniqueness) is the following lemma.

20
Lemma 4.1. Let G, H ∈ C, then
(
X X 1 if G = H
TF,H SG,F = TG,F SF,H =
F ∈I F ∈I
0 6 H.
if G =
Proof. The quotients of G that are (in C) isomorphic to F correspond exactly to the Aut(F )
orbits on Epi(G, F ) (which all have size |Aut(F )|). Thus, given G, H ∈ C, we have an
epimorphism
{F ∈ I, πFG ∈ Epi(G, F ), ρ ∈ Epi(F, H)} → {πH
G
∈ Epi(G, H), F ∈ [H, G]}
(F, πFG , ρ) 7→ (ρπFG , (F, πFG )),
where the fiber of (ρπFG , (F, πFG )) has size | Aut(F )|. Thus,
X X X 1 1 X
SG,F TF,H = µ(ρ)
F ∈I F ∈I G
| Aut(F )| |Aut(H)|
πF ∈Epi(G,F ) ρ∈Epi(F,H)
X 1 X
= µ(H, F )
G ∈Epi(G,H)
| Aut(H)|
πH F ∈[H,G]
(
1 if G = H
=
0 if G 6= H,
where the last equality is by (2.5). The second statement follows similarly, using the other
part of (2.5). 
We are now ready to prove (Uniqueness) with a slightly weaker hypothesis, which we
restate here.
Theorem 4.2. Let C be a diamond category, C be a level of C, and (MG )G ∈ RI be so that
for all F ∈ C, we have
X X
(4.3) |µ(F, G)mG | < ∞.
G∈C π∈Epi(G,F )

Suppose for all F ∈ C we have values pC,F ≥ 0, such that for all G ∈ C we have
X
SF,G pC,F = mG
F ∈C
Then, for all F ∈ C, X
pC,F = TG,F mG
G∈C

Proof. If we can exchange the order of summation, we have


X X X X X
TG,F mG = TG,F SH,G pH,C = pH,C TG,F SH,G = pC,F
G∈C G∈C H∈C H∈C G∈C

where the first equality is by assumption, the last is from Lemma 4.1 (using that anything
with an epimorphism from H is in C). To verify that we can exchange
P the order
P of summa-
tion, as in the middle equality, it suffices to check that the sum G∈C TG,FP H∈C SH,G pH,C
is absolutely convergent. Since SH,G pH,C ≥ 0, it suffices to show that G∈C TG,F mG is
absolutely convergent, which is precisely our assumption. 
21
Note that (Uniqueness) implies the “only if” direction of (Existence). Similarly, we could
prove the “if” direction of (Existence), conditional on absolute convergence of the below
sums,
!
X X X X
(4.4) SF,H TG,F mG = mG TG,F SF,H = mH .
F ∈C G∈C G∈C F ∈C

Unfortunately, in most cases of interest the absolute convergence of the above sums will
not hold (see Example 4.5). This requires us to make more delicate arguments to exchange
the order of summation one piece at a time, relying on further identities satisfied by the TG,F
as well as bounds on their size.
Example 4.5. One of the nicest examples of moments are that of the distribution µ on
profinite groups from [LW20], which has G-moment MG = 1 for every finite group G. (We
will see in Lemma 6.12 that these are well-behaved.) The category C here is the category of
finite groups. However, even in this case, the sum
XX XX 1
TG,F mG = TG,F
F ∈C G∈C F ∈C G∈C
| Aut(G)|
does not converge absolutely when C is generated by {F3 , S3 } (we write F2 , F3 for the cyclic
groups of order 2,3, respectively). (This sum is the left side of (4.4) in the special case
H = 1.) Indeed, to check that
XX
|TG,F || Aut(G)|−1 = ∞,
F ∈C G∈C

it suffices to prove a restricted sum is infinite, over only F = Fn2 and G = Fm n


3 ⋊ F2 for some
n and m and any action of Fn2 on Fm n
3 . Every representation of F2 over F3 is a sum of the
n
2 one-dimensional characters, and from this we can check every such G is level-C.
For such a G and F , any semisimple morphism G → F is the obvious projection Fm n
3 ⋊F2 →
n n
F2 composed with an automorphism of F2 , so
TG,F = µ(F, G).
For a given n, we can consider the set Hom(Fn2 , GLm (F3 )). We have that GLn (F2 ) ×
GLm (F3 ) acts on this set, by precomposition for the first factor, and conjugation for the
second factor. The orbits correspond to isomorphism classes of groups of the form Fm 3 ⋊ F2 ,
n

and the stabilizers Sa correspond to automorphisms of F3 ⋊a F2 that fix the subgroup Fn2


m n

(setwise). For a ∈ Hom(Fn2 , GLm (F3 )), the automorphisms of Fm n


3 ⋊a F2 act on the Sylow
m m im a
2-subgroups transitively, and there are [F3 : (F3 ) ] Sylow 2-subgroups. Thus the stabilizer
Sa has size | Aut(Fm n m m im a
3 ⋊ F2 )|/[F3 : (F3 ) ]. It follows by the orbit-stabilizer theorem that for
each n,

X
−1
X X |µ(F2n , F3m ⋊a F2n )|
|TG,F || Aut(G)| = .
G∈C m=0 n
[Fm m im a
3 : (F3 ) ]| GLm (F3 )|| GLn (F2 )|
a∈Hom(F2 ,GLm (F3 ))
G∼=F m n
3 ⋊F2
for some m

We add the subscript a to the ⋊ to clarify which action we mean.


The number elements of Hom(Fn2 , GLm (F3 )) giving multiplicity vector m1 , . . . , m2n when
decomposed into irreducible representations of Fn2 over F3 is | GLm (F3 )|/|ZGLm (F3 ) (im a)|,
22
where a ∈ Hom(Fn2 , GLm (F3 )) is any element with that multiplicity vector and ZGLm (F3 ) (im a)
Q n Qmi −1 mi
is the centralizer of im a in GLm (F3 ). Further, ZGLm (F3 ) (im a) has size 2i=1 j=0 (3 − 3j ).
Q n mi (mi −1)
The value of µ(F2n , F3m ⋊a F2n ) for such an action is given by 2i=1 (3 2 ) by Lemma 3.10.
Thus
∞ Q2n mi (mi −1)
X
−1
X 1 i=1 (3
2 )
|TG,F || Aut(G) | = P
m 2n Qm −1
| GLn (F2 )|3 i>1 i i=1 j=0 (3mi − 3j )
Q i
G∈C m1 ,...,m2n =0
G∼=Fm n
3 ⋊F2
for some m
∞ m(m−1)
!2n
1 X 3 2
> Qm−1 ,
| GLn (F2 )| m=0
3m j=0 (3m − 3j )
so !2n
∞ ∞ m(m−1)
XX X 1 X 3 2
|TG,F || Aut(G)|−1 > Qm−1 =∞
F ∈C G∈C n=0
| GLn (F2 )| m=0
3m j=0 (3m − 3j )
since m(m−1)

X 3 2 1
c := Qm−1 =1+ +··· > 1
m=0
3 m
j=0 (3m − 3j ) 6
2 n
and | GLn (F2 )| ≤ 2n , and hence for large n we have that c2 /| GLn (F2 )| ≥ 1.
4.1. Proof of (Existence). Example 4.5 shows that our proof of (Existence) will need to
be somewhat more complicated than the easy argument in (4.4) that requires interchanging
the order of summation between a sum over F ∈ C and a sum over G ∈ C. We will prove
(Existence) as a sequence of smaller identities, each based on rearranging a shorter sum that
does converge absolutely for well-behaved moments. The key is not to exchange a sum in G
with a sum over all F , but only a sum over F with a fixed pro-C¯ completion, for a regime C¯
just barely smaller than the regime C under consideration. In Example 4.5, this would mean
we only consider the sums for a fixed n, instead of summing over all n at once. After we
do this more limited exchange of summation, we will be able to cancel some terms before
proceeding further with the exchange of order of summation. We will see that our notion
of “well-behaved” moments is sufficient to allow us to do these more restrained exchanges of
the order of summation.
We say that an epimorphism G → E is relative level C if the induced map G → GC ∨G E
is an isomorphism.
Lemma 4.6. If E is level C then G → E is relative level C if and only if G is level C.
Proof. If G is level C, then GC ∨G E = G ∨ E = G. If G = GC ∨G E, since E is level C, we
have that G is level C. 
We now give an identity that we will use for the canceling of terms mentioned above.
Lemma 4.7. Let C be a level. Let G, E ∈ C and let φ : G → E be an epimorphism not of
relative level C. Let H ∈ C. Then
X X X X X X µ(π)
Tπ = = 0.
F ∈I F ∈I
| Aut(F )|
ρ∈Epi(F,E) π∈Epi(G,F ) ρ∈Epi(F,E) π∈Epi(G,F )
F C ≃H ρπ=φ F C ≃H ρπ=φ
23
Proof. We have that G → E not relative level C implies that G > GC ∨ E, and hence for
K ≤ GC , also that G > K ∨ E. Also, by Lemma 2.10 (3), for F ≤ G, we have F ∧ GC = F C .
Thus, for any epimorphism GC → H,
X X
0= µ(H, K) µ(F, G) (inner sums vanish since G > K ∨ E)
K∈[H,GC ] F ∈[K∨E,G]
X X
= µ(F, G) µ(H, K) (change order of summation)
F ∈[H∨E,G] K∈[H,F ∧GC ]
X
= µ(F, G) (inner sum vanishes unless F C = H).
F ∈[H∨E,G]
F C =H

We will see that, summed over the structures on H as a quotient of G (i.e. over Epi(G, H)/ Aut(H)),
this final sum is precisely the sum in the lemma.
There is an | Aut(F )| − 1 correspondence between pairs (F, π) of F ∈ I and π ∈ Epi(G, F )
and F ≤ G. Given π and φ, when F ≥ E then there is one choice of ρ ∈ Epi(F, E) such that
ρπ = φ, and otherwise there is no possible ρ such that ρπ = φ. Finally, F C = H in the sum
above means that F C , as a quotient of G, is isomorphic to H, and so obtain the sum over all
F ∈ I with F C isomorphic to H in C, as in the lemma, we need to sum over all G-quotient
structures on H up to isomorphism. 
Now we will see that well-behaved moments will provide absolute convergence on more
limited sums.
Proposition 4.8. Let C be a level generated by a finite set B, and let B̄ ⊂ B be a subset
containing every proper quotient of an element of B, and C¯ the level generated by B̄. Let
E ∈ C . and H ∈ C.¯ If (MG )G is behaved at the level generated by C and E, then
X X X X
Tπ mG < ∞.
G∈I F ∈I π∈Epi(G,F ) ρ∈Epi(F,E)
F C̄ ≃H ρπ rel. level C

Proof. Given G, F, π, ρ as in the sum, by Lemma 2.6, µ(π) = 0 unless π is semi-simple, so we


only consider terms with π semi-simple. We will work in the lattice of (isomorphism classes
of) quotients of G, and our first goal is to show that G → H ∨ E is semisimple. Lemma 2.9
¯ ¯
implies that GC → GC is semisimple, and Lemma 2.4 then implies GC → (E ∨ GC ) ∧ GC
¯ ¯ ¯
is semisimple, since GC ≤ (E ∨ GC ) ∧ GC . We have (E ∨ GC ) ∨ GC = E ∨ GC = G, where
the last equality is from ρπ relative level C. Then by the Diamond Isomorphism Theorem
¯ ¯ ¯ ¯
applied to E ∨ GC and GC , we have [E ∨ GC , G] = [(E ∨ GC ) ∧ GC , GC ], and thus G → E ∨ GC
¯
is semisimple. Since E ∨ GC and F are both (as G-quotients) the meet of quotients whose
¯
morphism from G is simple (by Lemma 2.4), the same is true for (E ∨ GC ) ∧ F . We have
H ∨ E = (GC ∧ F ) ∨ E (Lemma 2.10 (3))
= (E ∨ GC ) ∧ F (modular property).
Thus G → H ∨ E is semisimple.
We claim that, given E and H, there are only finitely many (isomorphism classes) of objects
that can be H ∨G E for some G, with G, F, π, ρ as in the sum. We can consider the level
C ′ generated by {H, E}, and note any H ∨G E is level-C ′ . By the Diamond Isomorphism
24
Theorem, we have that dim(H ∨G E → E) ≤ dim H. By Lemma 2.8, there are at most
finitely many (isomorphism classes of) K ∈ C ′ with an epimorphism K → E of dimension at
most dim H, which proves the claim. Let K1 , . . . , Kt be objects of C, ones from each such
isomorphism class of K, and note that Epi(Ki , E) is finite.
Let C˜ be the level generated by C and E. Since ρπ is relative level C, we have that G ∈ C.
˜
Then
X X X X
Tπ mG
G∈I F ∈I π∈Epi(G,F ) ρ∈Epi(F,E)
F C̄ ≃H ρπ rel. level C
t
X X X X X X X
≤ Tπ mG
i=1 ψ∈Epi(Ki ,E) G∈C˜ F ∈I π∈Epi(G,F ) ρ∈Epi(F,E) φ∈Epi(G,Ki )
F C̄ ≃H G→H∨G E semisimple G→H∨G E and φ
isom G-quotients,
ψφ=ρπ
t
X X X X X X
≤ Tπ mG
i=1 ψ∈Epi(Ki ,E) G∈C˜ φ∈Epi(G,Ki ) F ∈I π∈Epi(G,F )
φ semisimple φ=κπ for some κ
t
X X X X X |µ(F, G)MG |
=
i=1 ψ∈Epi(Ki ,E) G∈C˜ φ∈Epi(G,Ki ) F ∈[Ki ,G]
| Aut(G)|
φ semisimple
t
X X X X |µ(Ki , G)Z(φ)a MG |
≪ .
i=1 ψ∈Epi(Ki ,E) G∈C˜ φ∈Epi(G,Ki )
| Aut(G)|
φ semisimple

In the second inequality, we can drop the sum over ρ because ρ is determined as ψκ (and
κ is unique since π is an epimorphism). The last inequality is by Lemma 3.12, where a is
as in the definition of behaved. The implicit constant in the ≪ depends on the Ki and C,˜
which in turns depends on only C, H, and E. The final sum converges by the definition of
behaved and the face that the first two sums are finite. 
We combine Lemma 4.7 and Proposition 4.8 to give a general result from which (Existence)
will follow.
Proposition 4.9. Let C be a regime, and C¯ ⊂ C a sub-regime. Let E be an object of C, and
¯ If (MG )G is behaved at the level generated by C and E and vC,F ≥ 0 (defined in
let H̄ ∈ C.
(1.5)) for all F ∈ C, then
X X X X X X X X X
Tπ mG = Tπ mG
G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F ) H∈IC G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F )
F C̄ =H̄ ρπ rel. level C¯ H C̄ =H̄ F C =H ρπ rel. level C

(both sums converge and are equal).


Proof. First, we consider finite sets of isomorphism classes of objects B0 ⊂ B1 ⊂ · · · ⊂ Bt
such that E ∈ Bt , and Bt is a subset of the level generated by C and E, and also that for all
i ≥ 1, every proper quotient of an element of Bi is in Bi−1 . Let Ci be the level generated by
Bi . We will first prove a general formula that implies our desired conclusion when C0 = C¯
25
and Cn = C for some n, and then check that a sequence B0 ⊂ B1 ⊂ · · · ⊂ Bt with these
properties actually exists.
For any i and j = i, i + 1, we let
X X X X
SH,i,j := Tπ mG .
G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F )
F Ci ≃H ρπ rel. level Cj

By Proposition 4.8, we have that the sum defining SH,i,j is absolutely convergent. Using
Lemma 4.7, we can subtract from SH,i,i+1 the terms such that ρπ is not level Ci and conclude
that
(4.10) SH,i,i+1 = SH,i,i .
Since the sum defining SH,i,i+1 is absolutely convergent, we can rearrange the order of
summation to obtain
X X X X X X X X X
Tπ mG = Tπ mG ,
G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F ) H ′ ∈Ci+1 G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F )
F Ci ≃H ρπ rel. level Ci+1 H ′ Ci =H F Ci+1 ≃H ′ ρπ rel. level Ci+1

i.e.
X
(4.11) SH,i,i+1 = SH ′ ,i+1,i+1 .
H ′ ∈Ci+1
(H ′ )Ci ≃H

We have
X X X X
SH,t,t = Tπ mG (Lemma 4.6)
G∈Ct F ∈I ρ∈Epi(F,E) π∈Epi(G,F )
F Ct ≃H
X X X
= TG,F mG (absolute convergence)
F ∈I ρ∈Epi(F,E) G∈Ct
F Ct ≃H
≥0 (vC,F ≥ 0).
Using (4.11) and (4.10), we inductively see that for all i, we have SH,i,i , SH,i,i+1 ≥ 0. This
non-negativity allows us to rearrange the following sum and obtain, for any H̄ ∈ C0 .
X X X
(4.12) SH ′ ,i+1,i+1 = SH ′ ,i+1,i+1.
H ′ ∈Ci+1 H∈Ci H ′ ∈Ci+1
(H ′ )C0 ≃H̄ H C0 ≃H̄ (H ′ )Ci ≃H

If we expand SH ′ ,i+1,i+1 with its definition, the resulting sums in (4.12) are not necessarily
absolutely convergent, as we can see in Example 4.5. This requires us to take a great deal
of care in this argument. However, the sums in (4.12) are sums of the same non-negative
terms SH ′ ,i+1,i+1 , and thus the sums both either equal some non-negative number or ∞.
Combining (4.12), (4.11), (4.10), we obtain
X X
(4.13) SH ′ ,i+1,i+1 = SH,i,i .
H ′ ∈Ci+1 H∈Ci
(H ′ )C0 ≃H̄ H C0 ≃H̄
26
From (4.13), we see inductively that for any n, we have
X X
(4.14) SH,n,n = SH ′ ,0,0 = SH̄,0,0 ,
H∈Cn H ′ ∈C0
H C0 ≃H̄ H C0 ≃H̄

and in particular the sum on the left hand side in (4.14) converges. If B0 generates C¯ and
Bn generates C then this establishes the proposition.
Now we will choose B0 ⊂ B1 ⊂ · · · ⊂ Bt so as the prove the proposition. If B̂ is the
closure under taking quotients of a set B of objects, then we note that B and B̂ generate
the same level. Thus to prove the proposition, we can assume without loss of generality that
C and C¯ are generated by finite sets B, B̄ respectively, both closed under taking quotients.
Let Bt = B\ ∪ {E}. We construct Bt−j for j ≥ 1 by removing from Bt−j+1 one element that
is not a proper quotient of any element of Bt−j+1 , and is not in B. Since B is closed under
taking quotients, this is possible unless Bt−j+1 = S, and eventually we get to that point and
Bn = B. Now we similarly construct a chain from Bn = B to B0 = B̄ such that Bi−1 ⊂ Bi
and every proper quotient of an element of Bi is in Bi−1 . Thus we have B0 ⊂ B1 ⊂ · · · ⊂ Bt
satisfying the assumptions at the start of the proof, and also such that B0 generates C¯ and
Bn generates C, which proves the proposition. 
We now can prove the “if” part of (Existence), which we restate here.
Theorem 4.15. Let C be a diamond category, C be a level of C, and let (MG )G ∈ RC be
behaved at C with vC,F ≥ 0 for all F ∈ C. Then νC ({F }) = vC,F defines a measure νC on C
with moments (MG )G .
Proof. Let E ∈ C. We apply Proposition 4.9 with C¯ containing only the minimal objects.
Note that an epimorphism F → E is relative level C¯ if and only if it is an isomorphism. The
left-hand side of Proposition 4.9 is
XX X X
Tπ mG = ME .
G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F )
ρπ an isom.

Since E ∈ C, we have that an epimorphism G → E is relative level C if and only if G ∈ C


(Lemma 4.6). The right-hand side of Proposition 4.9 is
XX X X X XX X X
Tπ mG = Tπ mG
H∈C G∈I F ∈I ρ∈Epi(F,E) π∈Epi(G,F ) H∈C G∈C ρ∈Epi(H,E) π∈Epi(G,H)
F C ≃H ρπ rel. level C
X X
= | Aut(E)| SH,E TG,H mG
H∈C G∈C
X
= | Aut(E)| SH,E vC,E .
H∈C

Thus Proposition 4.9 gives


X
| Aut(E)| SH,E vC,E = ME ,
H∈C

which proves the corollary. 


27
4.2. Proof of (Robustness). Now we prove (Robustness), which will eventual follow from
(Uniqueness) after some delicate limit switching arguments. Note that for Robustness it is
necessary that we work at a level (finitely generated formation) and not the whole category
(see [WW21, Example 6.14]).
We begin with a simple inequality, which we next use to prove a technical lemma to provide
us bounds we will need for exchanging limits and sums.
Lemma 4.16. Let π : G → F be a semisimple morphism of dimension d. For any k from 0
d

to d, the number of H ∈ [F, G] with the dimension of H → F equal to k is at least k .
Proof. This follows from the classification Theorem 3.5, because if the lower bound is true
for lattices L1 , L2 then it is true for their product, so it suffices to handle the three cases
of Theorem 3.5, which each follow from the fact that the q-binomial coefficients bound the
ordinary binomial coefficients for q ≥ 1. 
Remark 4.17. Lemma 4.16 could in many cases be replaced by a more exact calculation.
For example, in the case of groups one can exactly count the number of normal subgroups
intermediate between a kernel of a semisimple morphism π and the trivial subgroup. In this
case, Lemma 4.16 is sharp when the kernel is a product of nonabelian finite simple G-groups,
but is usually not sharp in the abelian case.
Lemma 4.18. Let C be a level. For all F ∈ C and n ∈ N, let pnC,F ≥ 0 be a real number.
Suppose that, for all G ∈ C, we have
X
(4.19) sup SF,G pnC,F < ∞.
n∈N
F ∈C

Then, for all H ∈ C, we have


X X X 
sup pnC,F < ∞,
n∈N
d∈N F ∈C ρ∈Epi(F,H)
dim ss(ρ)=d

where ss(ρ) denotes the semisimplification of ρ, the maximal semisimple K → H through


which ρ factors.
Proof. Let H ∈ C. Summing (4.19) over the set of G ∈ C and maps φ ∈ Epi(G, H) which
are semisimple of dimension ≤ 2 (which is finite by Lemma 2.8), we obtain
X X X
∞> sup SF,G pnC,F
n∈N
G∈C φ∈Epi(G,H) F ∈C
semisimple
dim φ≤2
X X X
≥ sup SF,G pnC,F
n∈N
G∈C φ∈Epi(G,H) F ∈C
semisimple
dim φ≤2

XX X X pnC,F
= sup
n∈N
G∈C F ∈C φ∈Epi(G,H) π∈Epi(F,G)
|Aut(G)|
semisimple
dim φ≤2
28
X X X X X 1
= sup pnC,F .
n∈N
F ∈C ρ∈Epi(F,H) G∈C φ∈Epi(G,H) π∈Epi(F,G)
|Aut(G)|
semisimple φπ=ρ
dim φ≤2

Note that given ρ as in the sum, if for π ∈ Epi(F, G) there exists a φ such that φπ = ρ, then
that φ is unique. Thus we can evaluate the inner sum above by
X X X 1
= |{G ∈ [H, F ] | G → H semisimple of dim. ≤ 2}.|
G∈C
|Aut(G)|
φ∈Epi(G,H) π∈Epi(F,G)
semisimple φπ=ρ
dim φ≤2

Let sρ : Rρ → H be the semisimplification of ρ. Then the above is equal to the number


of G ∈ [H, Rρ ] of dimension ≤ 2 over H, since an element of [H, F ] is semisimple over H if
and only if it is ≤ Rρ . By Lemma 4.16, this is at least
       
dim(sρ ) dim(sρ ) dim(sρ ) dim(sρ ) + 1
+ + = + 1.
2 1 0 2
This gives
  
X X dim(sρ ) + 1
∞ > sup pnC,F +1
n∈N
F ∈C ρ∈Epi(F,H)
2
  
XX X d+1
= sup pnC,F +1 .
n∈N 2
d∈N F ∈C ρ∈Epi(F,H)
dim sρ =d

Let N be the sum above. Then by restricting the sum over d in the definition of N to a
particular value, we obtain for any d ∈ N that
X X N
sup pnC,F ≤ d+1
 .
n∈N
F ∈C ρ∈Epi(F,H) 2
+1
dim sρ =d

This gives

X X X  X N
sup pnC,F ≤ d+1
  ≪ N < ∞,
d∈N
n∈N
F ∈C ρ∈Epi(F,H) d∈N 2
+1
dim sρ =d

as desired. 

If we have an epimorphism F → R0 , we can take its radical F → R1 , and then take the
radical of F → R1 , etc., which leads to the following definition. Given a sequence of maps
φ ρk ρ1
F → Rk → Rk−1 · · · → R0 , we say φ, ρk , . . . , ρ1 is a radical sequence if for all 1 ≤ i ≤ k,
we have that ρi+1 ◦ · · · ◦ ρk ◦ φ : F → Ri is a radical of ρi ◦ ρi+1 ◦ · · · ◦ ρk ◦ φ : F → Ri−1 .
Lemma 4.18 will allow us to inductively prove the following result using the Fatou-Lebesgue
theorem, moving the limit further and further past the sum as k increases.
29
Lemma 4.20. For some level C, for all F ∈ C and n ∈ N let pnC,F ≥ 0 be a real number.
Suppose that, for all G ∈ C, we have that (4.19) holds. Then, for any R0 ∈ C, for all natural
numbers k we have
X X X X pnC,F
lim sup SF,R0 pnC,F ≤ lim sup Qk .
n→∞
F ∈C d1 ,...,dk ∈N
n→∞
F,R1 ,...,Rk ∈C φ,ρk ,...,ρ1 rad. seq. i=0 |Aut(Ri )|
φ∈Epi(F,Rk )
ρi ∈Epi(Ri ,Ri−1 )
dim(ρi )=di

Proof. The proof is by induction on k. The case k = 0 is trivial. Now we assume the lemma
is true for k. An epimorphism F → Rk has a radical F → Rk+1 , where the object Rk+1 is
unique up to isomorphism, the map F → Rk+1 is unique up to composition with elements
of Aut(Rk+1 ), and the resulting map Rk+1 → Rk is semisimple. Thus we have
X X pnC,F X X X pnC,F
Qk = Qk+1 .
F,R1 ,...,Rk ∈C φ,ρk ,...,ρ1 rad. seq. i=0 |Aut(Ri )| dk+1 ∈N F,R1 ,...,Rk+1 ∈C b,ρk+1 ,...,ρ1 rad. seq. i=0 |Aut(Ri )|
φ∈Epi(F,Rk ) φ′ ∈Epi(F,Rk+1 )
ρi ∈Epi(Ri ,Ri−1 ) ρi ∈Epi(Ri ,Ri−1 )
dim(ρi )=di dim(ρi )=di

To complete the induction, it suffices to show that, for fixed d1 , . . . , dk ,


X X pnC,F X X pnC,F
lim sup Qk+1 ≤ lim sup Qk+1 ,
n→∞
dk+1 ∈N F,R1 ,...,Rk+1 i=0 |Aut(Ri )| dk+1 ∈N
n→∞
F,R1 ,...,Rk+1 i=0 |Aut(Ri )|
φ′ ,ρk+1 ,...,ρ1 φ′ ,ρk+1 ,...,ρ1

(where the sums are over the same sets as in the previous equation). Given d1 , . . . , dk ,
Lemma 2.8 says there are only finitely many possible isomorphism classes of objects that can
be R1 , . . . , Rk in the above sums, as dim(ρi ) = di . Thus the sum over R1 , . . . , Rk , ρ1 , . . . , ρk
can be freely exchanged with the lim sup. Fixing R1 , . . . , Rk , ρ1 , . . . , ρk , we apply the Fatou-
Lebesgue Theorem to the remaining sum, and the hypothesis is checked by applying Lemma 4.18
with ρ = ρk+1 φ′ and H = Rk . The condition that φ′ is the radical of ρk+1 φ′ determines Rk+1 ,
φ′ (up to simply transitive Aut(Rk+1 ) orbit), and then ρk+1 from ρ. The Fatou-Lebesgue
theorem thus gives the inequality above, and we conclude the lemma by induction. 
Lemma 4.20 will be useful because it turns out that at a particular level, radical sequences
cannot go on arbitrarily long without stabilizing.
Lemma 4.21. Let C be a level. Then there exists some k, depending on C, such that for
all tuples R0 , . . . , Rk ∈ C with semisimple ρi : Ri → Ri−1 and F ∈ C with φ ∈ Epi(F, Rk ), if
φ, ρk . . . ρ1 is a radical sequence, then φ is an isomorphism.
Proof. Let B be a finite set generating C, and let k be the maximal dimension of an element
of B. For 0 ≤ j ≤ k, let Bj be the set of all quotients of elements of C with dimension at
most j and let Cj be the level generated by Bj .
Let us check that Ri ≥ F Ci (as quotients of F ) for all i, by induction on i. It will follow
that Rk ≥ F Ck = F so Rk ≃ F , as desired.
The case i = 0 is automatic because F C0 is the minimal quotient of F and thus is ≤ R0 .
For the induction step, assume Ri ≥ F Ci . By Lemma 2.1, [Ri , F Ci+1 ∨ Ri ] is isomorphic
to [F Ci+1 ∧ Ri , F Ci+1 ], and the latter is a subinterval of [F Ci , F Ci+1 ] since F Ci ≤ F Ci+1 and
F Ci ≤ Ri . By Lemma 2.9, the natural morphism F Ci+1 → F Ci is semisimple for all i, so
30
[F Ci , F Ci+1 ] is a complemented modular lattice, and thus [F Ci+1 ∧ Ri , F Ci+1 ] is complemented
by Lemma 2.3, and hence [Ri , F Ci+1 ∨ Ri ] is as well. Because F Ci+1 ∨ Ri → Ri is semisimple,
we have Ri+1 ≥ F Ci+1 ∨ Ri , since Ri+1 is the maximal quotient that F → Ri factors through
that is semisimple over Ri . Thus, Ri+1 ≥ F Ci+1 , verifying the induction step. 
Combining Lemmas 4.20 and 4.21, we show that under the assumptions (4.19) (“lim sup
of moments is finite”) and that a limit distribution exists, we can take a limit before or after
taking moments (“limit moments exist and are moments of the limit”).
Lemma 4.22. Let C be a level and for all F ∈ C and n ∈ N let pnC,F ≥ 0 be a real number.
Suppose that, for each F ∈ C, we have that the limit limn→∞ pnC,F exists and for all G ∈ C,
we have
X
(4.23) sup SF,G pnC,F < ∞
n
F ∈C
Then for all G ∈ C we have
X X
(4.24) lim SF,G pnC,F = SF,G lim pnC,F .
n→∞ n→∞
F ∈C F ∈C

Proof. Choose k as in Lemma 4.21, the maximal dimension of an element of C. Lemma 4.20
with R0 = G gives
X
n
X X X pnC,F
lim sup SF,G pC,F ≤ lim sup Qk .
n→∞
F ∈C d1 ,...,dk ∈N
n→∞
F,R1 ,...,Rk ∈C φ,ρk ,...,ρ1 rad. seq. i=0 |Aut(Ri )|
φ∈Epi(F,Rk )
ρi ∈Epi(Ri ,Ri−1 )
dim(ρi )=di

Lemma 4.21 says that is in the above sum, F is determined by Rk and φ must be an
isomorphism. Then, Lemma 2.8 implies that, given d1 , . . . , dk , the remaining sums over
F, R1 , . . . , Rk , φ, ρk , . . . , ρ1 only have finitely many terms and can be exchanged with the
lim sup. Thus the right-hand side above is equal to
X X X 1 X X limn→∞ pnC,F
Qk lim pnC,F = .
i=0 |Aut(Ri )| n→∞ |Aut(G)|
d1 ,...,dk ∈N F,R1 ,...,Rk ∈C φ,ρk ,...,ρ1 rad. seq. F ∈C ρ∈Epi(F,G)
φ∈Epi(F,Rk )
ρi ∈Epi(Ri ,Ri−1 )
dim(ρi )=di

The equality above follows because if we take ρ = ρ1 ◦ · · · ◦ ρk ◦ φ, then, inductively each Ri−1
and ρi ◦ · · · ◦ ρk ◦ φ for 2 ≤ i ≤ k + 1 is determined, up to simply transitive Aut(Ri−1 ) orbit,
from the radical sequence condition, which then determines ρi−1 (from ρ and ρi ◦ · · · ◦ ρk ◦ φ).
Combining, we obtain
X X
lim sup SF,G pnC,F ≤ SF,G lim pnC,F .
n→∞ n→∞
F ∈C F ∈C
Fatou’s lemma gives
X X
lim inf SF,G pnC,F ≥ SF,G lim pnC,F ,
n→∞ n→∞
F ∈C F ∈C
and the lemma follows. 
Finally, this allows us to prove (Robustness) conditional on (Uniqueness).
31
Theorem 4.25. Let C be a diamond category and C be a level. Let (MG )G ∈ RC such that
the following implication holds:
• For any (qC,F )F ∈ [0, 1]CP
P
, if for all G ∈ C we have F ∈C SF,G qC,F = mG , then for all
F ∈ C, we have qC,F = G∈C TG,F mG .
For all F ∈ C and n ∈ N let pnC,F ∈ [0, 1]. Suppose that for each G ∈ C, we have
X
(4.26) lim SF,G pnC,F = mG .
n→∞
F ∈C

Then for all F ∈ C we have X


lim pnC,F = TG,F mG .
n→∞
G∈C

Proof. By a diagonal argument the pnC,F have a weak limit p∞C,F in n, i.e. there is a subsequence
nj
nj of N such that for all F , we have limj→∞ pC,F = p∞
C,F . By Lemma 4.22 and (4.26), for all
G ∈ C, X
SF,G p∞
C,F = mG .
F ∈C
By the assumed implication, this implies that for all F ∈ C,
X
p∞
C,F = TG,F mG .
G∈C

pnC,F
has p∞
P
The fact that in every weak limit of C,F equal to the same value G∈C TG,F mG
implies that X
lim pnC,F = TG,F mG .
n→∞
G∈C

Finally, we record how Theorem 1.6 follows from the three theorems of this section.
Proof of Theorem 1.6. Well-behavedness implies the assumption (4.3) of Theorem 4.2, which
gives (Uniqueness) and the “only if” part of (Existence). Theorem 4.15 gives the “if” part of
(Existence). Theorem 4.25, using (Uniqueness) to satisfy the hypothesis, gives (Robustness).


5. The pro-object measure


This section has three parts. In the first, we prove Theorem 1.7 and Theorem 1.8 on
measures of pro-isomorphism classes. In the second, we connect pro-isomorphism classes to
the classical notion of pro-objects. In the third, we discuss pro-objects concretely in concrete
categories such as rings and groups.
5.1. The pro-isomorphism class case. The following result is very useful for constructing
measures on inverse limits of discrete spaces such as P.
Lemma 5.1. Let ν be a finitely additive non-negative function on an algebra A of subsets
of a set X (with ν(X) = b finite). For each ℓ ∈ N, let Lℓ be a countable set of elements of
A whose disjoint union is X.
(1) Suppose for every element A ∈ A, there is some ℓ such that for all ℓ′ ≥ ℓ, the set A is
a countable disjoint union of elements of Lℓ′ .
32
(2) Suppose any sequence Ln ∈ Ln with Lj ∩ Li non-empty for all i, j has ∩j Lj non-empty.
(3) For i ≤ j, and Ui ∈ Li , and U
Pj ∈ Lj , then Ui ∩ Uj is either empty or Uj .
(4) Suppose for each ℓ, we have L∈Lℓ ν(L) = b.
Then ν is countably additive on A.
Proof. This proof follows the arguments proving [LW20, Theorem 9.1] from [LW20, Theorem
9.2], but we spell out the argument here since we are working in a much more general setting.
First we show the following claim, the analog of [LW20, Corollary 9.7]. Claim: If B ∈ B
and B is a disjoint union of Uℓ,j ∈ Lℓ (for a fixed ℓ), then
X
(5.2) ν(B) = ν(Uℓ,j ).
j

Since the elements of Lℓ are a countable partition of X, we also have that X \B is a countable
disjoint union of the Vℓ,i ∈ Lℓ not among the Uℓ,j . Hence for every M, by finite additivity
and the non-negativity of ν, we have
M
X M
[ M
X
ν(Uℓ,j ) ≤ ν(B) ≤ ν(X \ Vℓ,i ) = b − ν(Vℓ,i ).
j=1 i=1 i=1

Taking limits as M → ∞ gives



X ∞
X ∞
X
ν(Uℓ,j ) ≤ ν(B) ≤ b − ν(Vℓ,i ) = ν(Uℓ,j )
j=1 i=1 j=1

where the last equality is by assumption (4). This proves (5.2).


If we have disjoint sets An ∈ A with A = ∪n≥1 An ∈ A, by taking Bn = A \ ∪nj=1 Aj , to
show countable additivity, it suffices to show that for B1 ⊃ B2 ⊃ . . . (with Bn ∈ A) with
∩n≥1 Bn = ∅ we have limn→∞ ν(Bn ) = 0.
We can assume, without loss of generality, that for each ℓ ≥ 1, we have Bℓ = ∪j Uℓ,j , with
Uℓ,j ∈ Lℓ (using (1)). (We can always insert redundant Bi ’s if the required ℓ’s increase too
quickly.) We will show by contradiction that limℓ→∞ ν(Bℓ ) = 0.
Suppose, instead that there is an ǫ > 0 such that for all ℓ, we have ν(Bℓ ) ≥ ǫ. It follows
from (5.2) that for each ℓ we have a subset Kℓ ⊂ Bℓ such that ν(Bℓ \ Kℓ ) < ǫ/2ℓ+1 and Kℓ
is a finite union of Uℓ,j .
Next, let Cℓ = ∩ℓj=1 Kj . Then ν(Bℓ \ Cℓ ) < ǫ/2, since
ν(Bℓ \ Cℓ ) =ν(Bℓ \ Kℓ ) + ν(Kℓ \ Kℓ ∩ Kℓ−1 ) + · · · + ν(Kℓ ∩ · · · ∩ K2 \ Kℓ ∩ · · · ∩ K1 )
<ǫ/2ℓ+1 + ν(Bℓ−1 \ Kℓ−1 ) + · · · + ν(B1 \ K1 )
<ǫ/2ℓ+1 + ǫ/2ℓ + · · · + ǫ/22 .
So ν(Cℓ ) ≥ ǫ/2 for each ℓ and in particular it is non-empty. Note Cℓ+1 ⊂ Cℓ for all ℓ. Pick
xℓ ∈ Cℓ for all ℓ.
Note that Cℓ is a finite union of the Uℓ,j (using (3)). Pick an H1 ∈ L1 so that infinitely
many of the xℓ are in H1 (this is possible since all xℓ are in C1 and C1 is a finite union of
elements of L1 ), and then disregard the xℓ that are not in H1 . In particular note H1 ⊂ C1 .
Then pick H2 ∈ L2 so that infinitely many of the remaining xℓ are in H2 , and disregard the
xℓ that are not. Since all of the remaining xℓ are in H1 , we have H1 ∩ H2 is non-empty, and
also H2 ⊂ C2 . We continue this process and at each stage H1 ∩ H2 ∩ · · · HN is non-empty and
33
a subset of CN , and hence ∩j Hj is non-empty (by (2)) and gives an element of ∩j Cj ⊂ ∩j Bj ,
which is a contradiction. 
We will need one more result before we can apply Lemma 5.1 to prove Theorem 1.7 and
Theorem 1.8.
Lemma 5.3. Let C be a diamond category and C0 ⊂ C1 ⊂ · · · be levels. Let Gi for i ∈ N be
objects of C such that for j ≥ i, we have GCj i ≃ Gi . Then given any level C, there exists an
N such that for i, j ≥ N, we have GCi ≃ GCj .
Proof. We induct on the size of a quotient closed generating set B of C. If B is empty, then
GCm is the minimal quotient of Gm , and for n ≥ m, since Gm is a quotient of Gn , they have
the same minimal quotient. Let B ′ be a quotient closed subset of B with one fewer element
than B, and suppose the claim is true for the level C ′ generated by B ′ .

Consider only i such that GCi has stabilized to G, and note that there is a semisimple
epimorphism GCi → G by Lemma 2.9, and thus GCi is the join of all the simple X → G such
that GCi → G factors through X → G by Lemma 2.4. Given G, there are only finitely many
X → G such that X ∈ C by the definition of diamond . Some of these X eventually appear
in a Cj and some never do. Consider only i large enough such that all X ∈ C with a simple
X → G that will ever appear in a Cj appear in Ci . If X is not in any Ci , then it is not ever
a quotient of Gi for any i. Since X ∈ Ci , we have that all epimorphisms Gj → X for j ≥ i
factor through GCj i ≃ Gi . Thus GCj is the join of the same elements as GCi , which proves the
lemma. 
Proof of Theorem 1.7 and Theorem 1.8. Let ν be a measure on P (for the given σ-algebra)
with the specified moments. For a regime C let νC be the push-forward of ν along the map
X 7→ X C . For G ∈ C, we have | Epi(X, C
PG)| = | Epi(X , G)|, and thus νC has the same
moments. We have ν(UF,C ) = νC (F ) = G∈C µ̂(F, G)mG by Theorem 1.6, and so the sum
is non-negative. Since the 0-moment gives ν(P) = m0 < ∞, by Carathéodory’s extension
theorem and the determined values of ν(UF,C ), there is a unique measure ν with the given
moments. So we conclude (Uniqueness) and the “only if”Pof (Existence).
Now, we assume that for all C and G ∈ C, we have G∈C µ̂(F, G)mG ≥ 0, and we will
show a measure ν with the indicated moments actually exists. Theorem 1.6 implies that for
each level C a measure νC exists with the specified moments. The uniqueness in Theorem
1.6 implies for C ⊂ C ′ the measure νC ′ pushes forward to νC along the map F 7→ F C . Let A
be the algebra of sets generated by the UH,C . Let à be the algebra of sets that are an (at
most) countable union of UH,C for a single C (and varying H). (This is an algebra because
C has countably many isomorphism classes by Lemma 2.8, and for C ⊂ C ′ we have UH,C is a
countable disjoint union of the UH ′ ,C ′ for all H ′ ∈ C ′ with (H ′)C ≃ H.) We have A ⊂ Ã. We
define a pre-measure ν on à by letting, for any subset V ⊂ C,
!
[
ν UH,C = νC (V ) .
H∈V

If we have A ∈ Ã that can be defined at two different levels C, C ′ , i.e.


[ [
A= UH,C = UH ′ ,C ′ ,
H∈V H ′ ∈V ′
34
′′
then for any regime C ′′ ⊃ C, C ′ , let V ′′ = A∩C ′′ . We have that X ∈ A if and only if X C ∈ V ′′ .
We have that V ′′ is the preimage of V under the map F 7→ F C from C ′′ to C (and similarly
for V ′ , C ′ ). Thus the claim about push-forwards above gives
νC (V ) = νC ′′ (V ′′ ) = νC ′ (V ′ ) ,
and our definition above of ν on à is well-defined. A similar argument refining to a common
level shows that ν is finitely additive.
We have ν a finitely additive non-negative function on the algebra à of sets. We now wish
to show that it is countably additive. Suppose we have that A0 ∈ Ã is the disjoint union of
A1 , A2 , · · · ∈ Ã. Let Ai be defined at level Ci , and let C≤i be the level generated by Cj for
j ≤ i. Let Ā be the algebra of sets that are (necessarily at most countable) disjoint unions
of UH,C≤i for some fixed i. So A0 , A1 , . . . , ∈ Ā. Then we apply Lemma 5.1 to Ā, with Lℓ the
set of UH,C≤ℓ . To show hypothesis (2) in Lemma 5.1, we need to show that for a collection
C
Gi with Gj ≤i = Gi for j ≥ i, there is X ∈ P with X C≤i = Gi . We construct such an X by
letting X C be GCi for i sufficiently large such that this isomorphism class has stabilized (by
Lemma 5.3). We can see that (3) is satisfied since P for j ≥ i, we have X C≤i = (X C≤j )C≤i .
Since νC is a measure, we have that νC (C) = G∈C νC (G), and the claim about push-
forwards above gives that νC (C) = ν{0} (0). Thus hypothesis P (4) in Lemma 5.1 is satisfied
and we have that ν is countably additive on Ā. Thus i≥1 ν(Ai ) = ν(A0 ), which implies
ν is countably additive on Ã. Hence by Carathéodory’s extension theorem ν extends to a
measure on P. As above, the moments of ν are determined by the push-forwards νC , which
have the desired moments.
Finally, we prove Theorem 1.8. First, we will show that if I is countable, then every open
set in the topology generated by the UC,F is a disjoint union of sets of the form UC,F . We note
since there are countably many finite subsets of I, there are countably many level C1 , . . . ,
and let C≤i be the level generated by Cj for j ≤ i. We note that the topology is actually
generated by the UC≤i ,H , since UCi ,F is a union of UC≤i ,H over varying H. Given an open V ,
around each F ∈ I we consider the minimal iF for which UC ,F C≤iF ⊂ V . So V is a union of
≤iF
the UC ,F C≤iF over F ∈ I. If the two sets UC ,F C≤iF , and UC ,H C≤iH , intersect, then one
≤iF ≤iF ≤iH
is contained in the other, and by the minimality of iF and iH , the two sets must be equal.
This shows that V is a countable disjoint union of UC,F .
By Theorem 1.6, we have, for every level C and F ∈ C,
lim ν t (UC,F ) = ν(UC,F ).
t→∞

Since any open set in our topology is a disjoint union of basic opens, using Fatou’s lemma
and then the Portmanteau theorem proves Theorem 1.8. 

5.2. Pro-isomorphism classes and pro-objects. We now explain the relationship be-
tween pro-isomorphism classes and the usual notion of pro-object.
One defines a pro-object of C to be a functor F from a small category I to C, where the
category I is cofiltered in the sense that for any two objects of I there is an object with a
morphism to both, and for any two morphisms f, g between objects A, B ∈ I there is an
object C ∈ I with a morphism h : C → A such that f ◦ h = g ◦ h [KS06, Definition 6.1.1
and Definition 3.1.1]. We think of a pro-object as the formal inverse limit of the diagram
35
(I, F ). The morphisms between two pro-objects (I, F ) → (J, G) is given by the formula
[KS06, (2.6.4)]
Hom((I, F ), (J, G)) = lim lim Hom(F (i), G(j)).
←− −→
j∈J i∈I
It is not hard to define the identity morphism and composition for morphisms, making the
pro-objects into a category (the pro-completion of C.) We can view an object of C as a
pro-object using a category I with one object and only the identity morphism.
Definition 5.4. We say a pro-object is small if it arises (up to isomorphism) from a diagram
(I, F ) where every F (f ) is an epimorphism for every morphism f of I, and if for each G ∈ C,
the set of epimorphisms from (I, F ) to G is finite.
This definition is in analogy to the definition of small profinite groups (see [FJ08, Section
16.10], and also Lemma 5.8 below).
We now construct the bijection between isomorphism classes of small pro-objects and pro-
isomorphism classes. We first prove a quick lemma describing the quotients of a pro-object
Lemma 5.5. Let C be a diamond category and X be a small pro-object of C. The poset of
quotients of X in C is a modular lattice.
Proof. Let X arise from a diagram (I, F ) with all morphisms in the image F epimorphisms.
We first note that, for G ∈ C, we have Hom(X, G) = limi∈I Hom(F (i), G) so every morphism
−→
X → G arises from some morphism F (i) → G. Next observe that a morphism X → G is an
epimorphism if and only if it arises from an epimorphism F (i) → G, i.e. that Epi(X, G) =
limi∈I Epi(F (i), G). The case “only if" is immediate and “if" uses that all morphisms in the
−→
image of F are epimorphisms.
It follows that the poset of quotients of X is the forward limit over i ∈ I of the poset of
quotients of F (i). For F (f ) : F (i) → F (j) an epimorphism, the induced map from the poset
of quotients of F (j) to the poset of quotients of F (i) is an inclusion of modular lattices (in
particular, respects meets and joins). It follows that the limit is a modular lattice. 
Lemma 5.6. Let C be a diamond category and X be a small pro-object of C. Then, for any
level C, the object X has a maximal level-C quotient X C , i.e. an object of C of level C that
is a quotient of X such that every other such quotient factors through it.
Proof. Let B be a generating set of C and m the maximum dimension of an object of B,
and let Bn be the set of quotients of objects of B of dimension at most n. Let Cn be the
level generated by Bn . We will prove for all n that X has a maximal level-Cn quotient by
induction on n, and then the claim follows from the n = m case. For the base case, we note
that the cofiltered and modular lattice properties imply that all F (i) have the same minimal
quotient, and we see that this is X C0 , the maximal level-C0 quotient of X.
For the induction step, let X Cn−1 be a maximal level-Cn−1 quotient of X. Let K be a
level-Cn quotient of X. Then, by Lemma 2.9, K → K Cn−1 is semisimple. Since K Cn−1
is a level-Cn−1 quotient of X, we have K Cn−1 ≤ X Cn−1 , and thus K ∨ X Cn−1 → X Cn−1 is
semisimple by Lemma 2.1 for modular lattices. It suffices to show that there is a maximal
level-Cn -quotient of X with a semisimple morphism to X Cn−1 , as then K ∨ X Cn−1 , which has
those properties, will factor through it, and then K will as well.
To do this, consider a level-Cn quotient H of X such that H → X Cn−1 is semisimple of
fixed dimension d. There are finitely many possible isomorphism classes of such objects
36
H by Lemma 2.8. Each one admits finitely many morphisms from X by the smallness
assumption, so there are finitely many such quotients of X. In particular, taking d = 1,
there are finitely many level-Cn quotients of X with a simple epimorphism to X Cn−1 . If
H → X Cn−1 is semisimple of dimension d, then it factors through at least d different simple
morphisms Gi → X Cn−1 of quotients of K (by Lemma 4.16) and hence quotients of X, and
so the dimension of any semisimple morphism from a level-Cn quotient of X to X Cn−1 is
bounded in terms of X, C, and n. Consider the level-Cn quotients of X with a semisimple
morphism to X Cn−1 . There are finitely many possible dimensions, and then finitely many
of each dimension. Thus, such quotients form a finite set and the join of two elements of
this set is an element of this set (by Lemmas 5.5 and 2.3), it must have a maximal member,
completing the induction step. 

Lemma 5.7. Let C be a diamond category. For each pro-isomorphism class X of C, there
exists a small pro-object Y of C, unique up to isomorphism, such that Y C ∼ = X C for each
level C.
In other words, the map from the set of isomorphism classes of small pro-objects to P that
sends Y to (Y C ) is a bijection.

Proof. First, for each level C we chose an object in the isomorphism class X C , and by slight
abuse of notation we will now write X C for this object, and we make analogous choices
throughout this proof. We will next show the existence of an assignment to each pair C1 , C2
of levels with C1 ⊂ C2 an epimorphism πCC12 : X C2 → X C1 such that whenever C1 ≤ C2 ≤ C3
we have πCC12 ◦ πCC23 = πCC13 . Since there are finitely many epimorphisms X C2 → X C1 for each
C2 , C1 , to check that these infinitely many conditions can be satisfied, by the compactness of
an infinite product of finite sets it suffices to check that every finite subset of them can be
satisfied. Any finite set of equations of the form πCC12 ◦ πCC23 = πCC13 involves only finitely many
regimes Ci , and these finitely many regimes Ci are contained in a single regime C.
For each Ci , fix an epimorphism πi : X C → (X C )Ci realizing the definition of (X C )Ci . Then
for every i, j with Ci ⊂ Cj , there is a unique epimorphism πij : (X C )Cj → (X C )Ci such that
C
πij πj = πi . We define πCij for each i, j with Ci ⊂ Cj by composing πij with fixed isomorphisms
C
(X C )Ck ≃ X Ck chosen for each k. One can check these πCij satisfy all of the required equations
involving only Ci ⊆ C, completing the argument for existence of the πCC12 .
Any epimorphism X C1 → X C1 is an isomorphism by the finiteness of quotient posets. Thus
any epimorphism πCC12 : X C2 → X C1 realizes X C1 as the maximal level-C1 quotient of X C2 , since
it factors through the maximal level-C1 quotient of X C2 (which we know, by the definition
of a pro-isomorphism class, is abstractly isomorphic to X C1 ).
Define a category I to have one object for each level C and a single morphism C2 → C1
if and only if C1 ⊂ C2 . Let Y be the pro-object defined using the index category I and the
functor sending C to X C , and the unique morphism C2 → C1 to πCC12 . Let’s check that Y is
a small pro-object and that Y C ≃ X C . It suffices to check that Y C ≃ X C for each level C,
since taking C to be generated by H, the number of epimorphisms Y → H is then bounded
by the number of epimorphisms X C → H, which is finite.

To do this, let F be any level-C quotient of Y . Then F must be a quotient of X C for
some C ′ by definition of Y . Let C˜ be the level generated by C ′ and C. Every epimorphism
˜
X C̃ → F factors through πCC̃ : X C → X C by our observation above that this is the maximal
37
level C quotient of X C̃ . Thus every level-C quotient of Y factors through Y → X C , which
completes the proof of existence.
It remains to show uniqueness up to isomorphism. In other words, we fix a small pro-
object Z with Z C ∼ = XC ∼ = Y C for all C and must construct an isomorphism Y → Z. As
above, we can choose compatible maps πCC12 : X C2 → X C1 for all C1 ⊂ C2 . For each level C,
we choose a map τC : Z → Z C (which realizes Z C as the maximal level C quotient of Z),
and then define τCC12 : Z C2 → Z C1 for all C1 ⊂ C2 so that τC1 = τCC12 τC2 . Next we will choose
a system of isomorphisms X C → Z C (for each C) that are compatible in the sense that, for
C2 ⊂ C1 , they form a commutative diagram with πCC12 and τCC12 . This again is choosing an
element of an infinite product of finite sets satisfying infinitely many conditions that each
depend on only finitely many terms of the product, so it suffices to check that any finite
subset of the conditions can be satisfied. To do this, we chose a level C containing all levels
appearing in a finite list of conditions, fix an isomorphism φ : X C → Z C , and note that
since πCCi : X C → X Ci and τCCi φ : X C → Z Ci both give the maximal level-Ci quotient of X C ,
then they must be isomorphic, and there is a unique isomorphism φCi : X Ci → Z Ci such that
φCi πCCi = τCCi φ. Then these φCi are compatible in the above sense.
Now we construct the isomorphism Y → Z. Writing Z = (I, F ), we must for each i ∈ I
choose a level C and a map X C → F (i), in a way compatible with all the morphisms of I.
We can choose C to be the level generated by F (i). Then there is a unique epimorphism
Z C → F (i) commuting with τC and the natural map Z → F (i). Taking the composition
with φC constructed above, we have X C → F (i), and one can check these are compatible
with the morphisms of I.
To map Z → Y , we must for each regime C choose i ∈ I and a map F (i) → X C , compatible
with all inclusions between regimes. We have φC τC : Z → X C , so there exists i ∈ I with a
morphism F (i) → X C , and choose that i and morphism.
It is straightforward to check that these two maps are inverses, and so Y ≃ Z. 

5.3. Pro-objects concretely.


Lemma 5.8. Let C be a diamond category such that every morphism can be written as a
composition ιπ, where π is an epimorphism and ι is a monomorphism, and every object has
finitely many subobjects, up to isomorphism. Then for (I, F ) a pro-object of C, the following
are equivalent
• (I, F ) is small.
• (I, F ) has finitely many epimorphisms to each object of C.
• (I, F ) has finitely many morphisms to each object of C.
Proof. We will first show that every pro-object of C is isomorphic to one of the form (J, G)
where all morphisms in the image of G are epimorphisms.
Fix a pro-object (I, F ). For every j ∈ I mapping to i, we can express F (j) → F (i) in at
least one way as an epimorphism followed by a monomorphism. The monomorphism gives
a subobject of F (i). Call a subobject of F (i) arising this way a valid subobject. We are
interested in the valid subobjects that are minimal for the natural ordering on subobjects
(restricted to valid subobjects).
Let J be the category of triples of of an object i ∈ I, an object H ∈ C, and a monomor-
phism π : H → F (i), such that (H, π) is a minimal valid subobject of F (i). We take the
38
morphisms in J from (i1 , H1 , π1 ) → (i2 , H2 , π2 ) to be given by pairs of a morphism f : i1 → i2
and an epimorphism g : H1 → H2 such that F (f ) ◦ π1 = π2 ◦ g. Let G be the functor J → C
sending (i, H, π) to H and the morphism given above to g.
We claim that J is cofiltered and (J, G) is isomorphic to (I, F ).
To show this, we first check that for each i ∈ I there exists a minimal valid subobject of
F (i). This is because there is always at least one valid subobject F (i), and by assumption
the number of subobjects is finite, so there is always a minimal element.
Next we check that for a minimal valid subobject H of F (i), for j mapping to i such
that F (j) → F (i) factors through an epimorphism F (j) → H, for all k ∈ I mapping to
j, the given map F (k) → F (i) factors through an epimorphism F (k) → H. (We call such
a map j → i a witness of the validity of H.) Indeed, for every k mapping to j, the map
F (k) → F (i) factors through H. So, as a morphism, F (k) → H is the composition of an
epimorphism followed by a monomorphism. If that monomorphism is not an isomorphism
then it defines a proper subobject of H, hence a subobject of F (i) smaller than H, which
F (k) factors through, contradicting minimality. Thus that monomorphism is an isomorphism
so F (k) → H is an epimorphism.
We now check that J is cofiltered. Given two objects (i1 , H1 , π1 ) and (i2 , H2 , π2 ), we need
to find an object mapping to both. Choose j1 mapping to i1 witnessing the validity of H1
and and choose j2 mapping to i2 witnessing the validity of H2 . By the cofiltered property of
I, we can choose j3 mapping to j1 and j2 . By the above, we can find (j3 , H3 , π3 ) ∈ J. Let’s
check that any such (j3 , H3 , π3 ) maps to (i1 , H1 , π1 ) (lifting the chosen map f : j3 → i1 ). By
the above, F (f ) : F (j3 ) → F (i) factors through an epimorphism F (j3 ) → H1 . Taking g to
be the composition of this epimorphism with π3 , we see that π1 ◦ g = F (f ) ◦ π3. To see that g
is an epimorphism, we choose j4 such that F (j4 ) → F (j3 ) factors through H3 , and note that
F (j4 ) → H1 is an epimorphism since otherwise we would obtain a valid subobject smaller
than H1 , so H3 → H1 is an epimorphism, as desired. Then (j3 , H3 , π3 ) maps to (I2 , H2 , π2 )
by a similar argument, verifying the first part of the cofiltered assumption.
Now we check the second part of the cofiltered assumption. Consider two maps (f, g)
and (f ′ , g ′ ) from (i1 , H1 , π1 ) to (i2 , H2 , π2 ). Since I is cofiltered, we can pick j5 and a map
f51 : j5 → i1 such that f f51 = f ′ f51 , and further by the cofiltered assumption we can choose
such a j5 and f51 such that j5 maps to j1 and j2 . By the above, any (j5 , H5, π5 ) ∈ J maps
to (i1 , H1 , π1 ) via (f51 , h) for some h. We have
π2 gh = F (f )π1h = F (f )F (f51)π5 = F (f ′ )F (f51 )π5 = F (f ′)π1 h = π2 g ′ h.
Since π2 is a monomorphism, it follows that gh = g ′h and (f51 , h)(f, g) = (f51 , h)(f ′ , g ′),
which shows the second part of the cofiltered assumption.
Now we check that (I, F ) and (J, G) are isomorphic. To construct a morphism (I, F ) →
(J, G), we must choose for each (i, H, π) ∈ J a j ∈ I and a map F (j) → G(i, H, π) = H. We
can take j and this map to be a witness to the validity of (H, π). To construct a morphism
(J, G) → (I, F ), we choose for each i ∈ I a (j, H, π) ∈ J and a map H = G(j, H, π) → F (i).
We choose j = i and (H, π) a minimal valid subobject, and the map π. It is straightforward
to check these two maps are well-defined and are inverses.
We have now shown that every pro-object of C is isomorphic to one of the form (J, G)
where all morphisms in the image of G are epimorphisms.
Thus, a pro-object is small if and only if it has finitely many epimorphisms to each object
of C. It remains to show this condition is equivalent to having finite many morphisms to each
39
object of C. Having finitely many morphisms implies having finitely many epimorphisms,
so it suffices to show the reverse implication. Each morphism (I, F ) → G arises from a
map F (i) → G which factors as an epimorphism F (i) → H followed by a monomorphism
H → G. The morphism (I, F ) → H is manifestly an epimorphism. There are finitely
many monomorphisms H → G, up to isomorphism, by assumption, and for each H, finitely
many epimorphisms from (F , I), proving that there are then finitely many morphisms from
(I, F ) → G. This concludes the proof of the equivalence of the three conditions in the
lemma. 
Remark 5.9. Let C be the category of finite groups, which clearly satisfies the hypotheses of
Lemma 5.8. The category of pro-objects in the category of finite groups is equivalent to the
category of profinite groups [RZ00, Theorem 2.1.3]. A profinite group is small in our sense if
and only if it has finitely many surjections to any finite group by Lemma 5.8. By a standard
argument, this is equivalent to having finitely many open subgroups of index n for each n,
which is the usual definition of a small profinite group (e.g. see [FJ08, Section 16.10]). In
this case Lemma 5.6 says that a small profinite group has a maximal level-C quotient. This
case appeared previously in [SW22b, Lemma 8.8], while the case of a topologically finitely
generated profinite group is more classical [Neu67, Corollary 15.72].
Lemma 5.10. (1) Let R be a Noetherian local commutative ring with finite residue field,
let m its maximal ideal, and R̂ its completion. Then the category of small pro-finite
R-modules (i.e. pro-objects for the category of finite R-modules) is equivalent to the
the category of finitely generated R̂-modules. The sets US,L = {N ∈ P | N ⊗R S ∼ = L}
for S a finite quotient of R and K an R-module form a basis for the topology on the
set of isomorphism classes.
(2) Let R be a Noetherian commutative ring. The the category of small pro-finite R-
modules is equivalent to the product over maximal ideals m of R with finite residue
field of the category of small pro-finite Rm -modules (where Rm is the completion of R
at m). The topology on its set of isomorphism classes is given by the product topology
(of the topology described above).
Proof. This is a standard argument so we will be somewhat brief. We first consider the local
case. Given a finitely generated R̂-module N, we obtain a pro-finite R-module Ñ by taking
the inverse limit of the finite modules N/mk . This is small since N is finitely generated.
Conversely, any pro-finite R-module lim Ni gives a R̂-module N by taking the inverse limit,
and the smallness gives that N is finitely generated.
One can check these constructions give an equivalence of categories. By Lemma 6.1, the
sets US,L are exactly the usual basis for the topology on P.
Now let R be a general Noetherian commutative ring. The natural map from the direct
sum over maximal ideals m with finite quotient of the category of finite Rm modules to the
category of finite R-modules is an equivalence, and this can be used to check the rest of the
lemma. 
6. Examples and applications
We have already mentioned two fundamental examples of categories to which the main
theorem can be applied: finite groups and finite R-modules for a ring R. We will first
investigate these cases in more detail, showing particular distributions that can be obtained,
40
and proven to be uniquely determined by their moments, using the main theorem. We will
then provide a toolkit for constructing more general examples, which has two parts. First,
we give a general construction using the language of universal algebra that includes these two
fundamental examples as well as many more. Second, we explain how new examples can be
produced from old. We show that the product of any diamond two categories is a diamond
category, that the objects in any diamond category with an action of a fixed finite group
Γ form a diamond category, and similar statements allowing us to put on extra structure.
Finally, we end with a simple examples, such as the opposite of the category of finite sets,
and the opposite of the category of finite graphs, to show that there are interesting results
even in these categories.
6.1. Modules. In this subsection, we apply our main results to the category of finite R-
modules for a ring R. We will show how the constructed measures and the condition for well-
behaved can be computed more explicitly, and give some examples of important measures
and moments. First we see that finite R-modules form a diamond category and give another
interpretations of levels in this case.
Lemma 6.1. Let R be a ring (not necessarily commutative). Then
(1) The category of finite R-modules is a diamond category.
(2) Each level of the category of finite R-modules is the set of isomorphism classes of S-
modules for some finite quotient ring S of R, and conversely each set of isomorphism
classes of S-modules for some finite quotient ring S of R is a level of the category of
finite R-modules.
Proof. We prove the first two assumptions of the definition of a diamond category first. The
poset of quotients of a module is dual to the poset of submodules, which is easily checked to
be modular. Finiteness of automorphism groups follows from the finiteness of the modules.
Let T be a finite set of R-modules. The action of R on the objects in T factors through
a minimal finite quotient ring S. Let’s prove that the level C generated by T is equal to the
category of S-modules. Since fiber products exist in the category of R-modules, joins are
given by fiber products over the meet. The property of being an S-module is preserved by
quotients and fiber products, thus join-closed and downward-closed, so every module in C is
an S-module.
To show the converse, since every finite S-module is a quotient of S n , and S n is a fiber
product of copies of S, it suffices to show that S is level-C. Now S is a submodule of the
|N |
product over, for each Ni in T , the module Ni i via the map that records where s ∈ S sends
each element of Ni . So it suffices to show that C is closed under taking submodules. If K is
a submodule of N then K is the quotient of N ×N/K N by the diagonal N, so C is closed
under taking submodules as well.
To finish the proof of (1), we observe that, since S is finite, every S-module has finitely
many extensions by a simple S-module, meaning there are finitely many simple morphisms
from an object of level C.
To finish the proof of (2), we have already shown that every level is the category of S-
modules for some S and we must show the converse, but this is clear as we can take the level
generated by S. 
For S a finite quotient of R, we write CS for the level of finite S-modules. We now give
the sums defining vCS ,N , and the condition for well-behaved, more explicitly.
41
Proposition 6.2. Let R be a ring. Let S be a finite quotient ring of R, and let K1 , . . . , Kn
be representatives of the isomorphism classes of finite simple S-modules. Let q1 , . . . , qn be
the cardinalities of the endomorphism fieldsQn of Keii .
1
For
Qn N an S-module and α ∈ Ext (N, i=1 Ki ), let Nα be the associated extension of N
ei
by i=1 Ki .
Let (MN )N ∈ RCS . Then for N ∈ CS we have

∞ n
!
1 X Y (−1)ei X
vCS ,N = MNα .
|Aut(N)| e ,...,e =0 i=1 |Hom(N, Ki )|ei ej=1 (qij − 1)
Qi
1
Qn ei
1 n α∈Ext (N, i=1 Ki )

We have that (MN )N is well-behaved at level CS if and only if the sum above is absolutely
convergent for every finite S-module N.
Proof. By definition we have
X X µ(π)ML
vCS ,N = .
L∈CS π∈Epi(L,N )
|Aut(N)||Aut(L)|

The µ(π) term vanishes unless π is semisimple,


Qn in which case ker π is a product of finite
simple S-modules, so we can write ker π as i=1 Kiei for a unique tuple e1 , . . . , en of natural
(ei )
numbers. In this case, µ(π) = ni=1 (−1)ei qi 2 by Lemma 3.8. Thus
Q

∞ n  X
1 X Y ei
ei ( 2 )
X ML
vCS ,N = (−1) qi .
|Aut(N)| e ,...,e =0 i=1 L∈C
|Aut(L)|
1 n s π∈Epi(L,N )
Q ei
ker π ∼
= ni=1 Ki

Now ker π ∼
Qn Qn
= i=1 Kiei if and only if there exists an injection f : i=1 Kiei → N such that
im f = ker π, and the number of such f ’s is
ei ei
!
n Y n n n Y n
e
q ( 2 ) (q j − 1) .
Y Y Y Y i
Y
e
Aut( Ki i ) = |Aut(Kiei )| = |GLei (Fqi )| = (q ei −q ei −j ) =


i=1 i=1 i=1 i=1 j=1 i=1 j=1

Thus
∞ n
!
1 X Y (−1)ei X X ML
vCS ,N = Qei j .
|Aut(N)| e ,...,e =0 i=1 j=1 (qi − 1) L∈Cs
|Aut(L)|
1 n π∈Epi(L,N )
Qn ei
f: i=1 Ki →L
Im f =ker π
f injective

Associated to eachQtriple L, f, π there is a class α ∈ Ext1 (N, ni=1 Kiei ), the class of the
Q
exact sequence 0 → ni=1 Kiei →f K →π N → 0. For this α we have ∼
Qn Nα = L so MNeαi = ML .
Next we check that each class α arises from exactly Aut(Nα )/ i=1 |Hom(N, Ki )| triples
L, f, π. Two triples give the same α, if and only if have the same L and are related by an
automorphism of L. Thus the number of different triples giving α is the number of orbits
of Aut(L) on maps f, π. The stabilizer is the group of automorphisms fixing the submodule
42
Qn
Kiei and its quotient module. The group of such automorphisms is Hom(N, ni=1 Kiei )
Q
i=1
with cardinality

Yn Y n
Hom(N, ei
Ki ) = |Hom(N, Ki )|ei


i=1 i=1

so the number of choices of (f, π) is Aut(Nα )/ i=1 |Hom(N, Ki )|ei , as desired.


Qn
Thus, we can replace the sum over L, f, π with a sum over α after multiplying by
n
Y
Aut(L)/ |Hom(N, Ki )|ei ,
i=1

which cancels the Aut(L) term and gives the desired statement.
The MN are well-behaved if and only if the sum defining vCS ,N with an additional Z(π)3
factor is absolutely convergent for all N. We have 2 ≤ Z(π) ≤ 2n , so that absolute conver-
gence happens if and only if the sum defining vCS ,N (in which n is constant) is absolutely
convergent for all N. 
Now we give a commonly occurring example of moments (e.g. the Cohen-Lenstra-Martinet
distributions have such moments [WW21, Theorem 6.2]), for which we find the measure
explicitly.
Lemma 6.3. Let R be a ring, C be the category of finite R-modules and u be a real number.
For each N ∈ C, let MN = |N|−u . Then (MN )N is well-behaved.
Let S be a finite quotient ring of R, and let K1 , . . . , Kn be representatives of the isomor-
phism classes of finite simple S-modules. Let q1 , . . . , qn be the cardinalities of the endomor-
phism fields of Ki . For N a finite S-module, we have
n Y ∞
!
Ext1 (N, Ki )

1 Y S
vCS ,N = 1− |qi |−j
|Aut(N)||N|u i=1 j=1 |Hom(N, Ki )||Ki |u

If u is an integer then vCS ,N ≥ 0 for all finite S-modules N.


Remark 6.4. If R is a DVR with residue field k, then for a finite S-module N, we have
Ext1S (N, k) ≤ |Hom(N, k)|, and then for all real u ≥ −1, we have vC ,N ≥ 0 for all finite

S
S-modules N.
Thus, by Theorem 1.7, for u ∈ Z (or u ≥ −1 when R is a DVR), there is a unique measure
on the set of isomorphism classes of pro-finite R-modules with moments |N|−u and it is given
by the formulas in Lemma 6.3. We write µu for this measure (when R is understood).
Proof. From Proposition 6.2 we have
∞ n
!
1 X Y (−1)ei X
vCS ,N = MNα .
|Aut(N)| e ,...,e =0 i=1 |Hom(N, Ki )|ei ej=1 (qij − 1)
Qi
1
Qn ei
1 n α∈Ext (N, i=1 Ki )

and
n
Y
−u −u
MNα = |Nα | = |N| |Ki |−ei u
i=1
43
so
∞ n
! n

1 X Y (−1)ei
1
Y
e

vCS ,N = Ext (N, Ki )
i

|Aut(N)||N|u |Hom(N, Ki )|ei |Ki |ei u ej=1 (qij − 1)
Qi

e1 ,...,en =0 i=1 i=1
∞ n
! n
1 X Y (−1)ei Y
Ext1 (N, Ki ) ei

= u
|Aut(N)||N| e ,...,e =0 i=1 |Hom(N, Ki )|ei |Ki |ei u ej=1 (qij − 1) i=1
Q i
1 n

∞ n
e !
(−1)ei Ext1 (M, Ki ) i

1 X Y
=
|Aut(N)||N|u e ,...,e =0 i=1 |Hom(N, Ki )|ei |Ki |ei u ej=1 (qij − 1)
Qi
1 n

n X ∞
e !
(−1)e Ext1 (N, Ki )

1 Y
=
|Aut(N)||N|u i=1 e=0 |Hom(N, Ki )|e |Ki |eu ej=1 (qij − 1)
Q

n Y ∞
!
Ext1 (N, Ki )

1 Y S −j
= 1− |qi |
|Aut(N)||N|u i=1 j=1 |Hom(N, Ki )||Ki |u
e
by the q-series identity ∞
Q∞
Q v −j
P
e=0 ej=1 (q j −1) = j=1 (1 + vq ) (from the q-binomial theorem).
This q-series sum is absolutely convergent because the denominator grows superexponen-
tially in e while the numerator grows only exponentially. All our manipulations preserve
absolute convergence in the reverse direction because we never combine two terms in the
sum with opposite signs – in fact, we never combine two terms except in the very first step,
where we group together identical terms.
|Ext1S (N,Ki )| 1
If u is an integer, then |Hom(N,K i )||Ki |
u is an integer power of qi since ExtS (N, Ki ), Hom(N, Ki ),

and Ki are all vector spaces over Fqi and hence have cardinality an integer power of qi . If
|Ext1S (N,Ki )|
|Hom(N,Ki )||Ki |u
is a positive integer power of qi then one of the terms in the product over j
vanishes, so the product is 0 and thus is nonnegative, and if it is a nonnegative integer power
of qi then all the terms in the product over j are positive, so the product is nonnegative.
Thus vCS ,N ≥ 0. 
Corollary 6.5. Let R be a ring and let C be the category of finite R-modules. Let (MN )N ∈
RC/≃ . If MN = O(|N|n ) for some real n, then (MN )N is well-behaved.
Proof. This follows immediately form Lemma 6.3, taking u = −n, and Lemma 2.11. 
Many special cases of both the uniqueness and robustness that follows from Corollary 6.5,
and the formulas for particular measures in Lemma 6.3 have appeared in previous work. For
example, Fouvry and Klüners prove uniqueness for the case R = Z/pZ [FK06a, Proposition
3] and give formulas for the probabilities when the moments are all 1 [FK06a, Proposition
2]. Ellenberg, Venkatesh, and Westerland prove uniqueness and give formulas for R = Zp
when all moments are 1, as well as prove robustness [EVW16, Section 8.1]. When R = Z,
the second author proved uniqueness and robustness given by Corollary 6.5, as well as giving
the formulas for moments |N|−u (see [Woo17, Theorem 8.3] and [Woo19, Section 3]). When
R is a maximal order in a semi-simple Q-algebra, Wang and the second author did the same
[WW21, Section 6]. For a R with finitely many (isomorphism classes of) simple R-modules
and such that Ext1R between any two finite R-modules is finite, the first author [Saw20,
Theorem 1.3] proved the uniqueness and robustness given by Corollary 6.5.
44
Under mild hypotheses, the measures from Lemma 6.3 have a nice description in terms of
a limit of random modules from generators and relations.
Lemma 6.6. Let R be a ring such that there are only countable many isomorphism classes
of finite R-modules (such as a Noetherian local commutative ring with finite residue field).
Let R̂ be the profinite completion of R.
As n → ∞, the distribution of the quotient of R̂n by n + u random relations taken from
Haar measure on R̂n weakly converges to µu (defined above as the measure on pro-finite
R-modules with moments |N|−u ).
Proof. A finitely generated R̂-module N naturally gives a small pro-finite R-module. Each
continuous finite quotient ring R̂/I is a finite quotient of R, and the functor from the category
of continuous finite quotients of R̂ taking R̂/I to the finite R-module R̂/I ⊗ N gives the
pro-object. Thus we can apply Theorem 1.8 to show that the distribution of the quotient
of R̂n by n + u random relations taken from Haar measure converges to µu as long as its
N-moment converges to |N|−u for all N.
This is true since, for each surjection π from R̂n to N, the probability that π descends to
the quotient of R̂n by n + u Haar random relations is |N|−n−u , since this is equivalent to
the probability that the images of the n + u random relations in N are all trivial. So the
expected number of surjections from the quotient of R̂n by n + u Haar random relations is
Sur(R̂n , N)|N|−n−u . As n goes to ∞, the ratio Sur(R̂n , N)/|N|n converges to 1 since it is
equal to the probability that n random elements of N generate N, which converges to 1.
Hence the expected number of surjections from the quotient of N converges to |N|−u , as
desired. 
For the rest of Section 6.1, we restrict attention to commutative rings.
When R is a local ring, the formulas for µu simplify even further, and we can make more
computations about the measure.
Lemma 6.7. Let R be a Noetherian local commutative ring with finite residue field k. We
have

!
Ext1 (L, k)

Y S
µu (US,L ) = |Aut(L)|−1 |L|−u 1− |k|−j−u
j=1
|Hom(L, k)|
for any nonzero finite quotient S of R.
Proof. Since R is a local ring, the only simple R-module is the residue field k, and for S 6= 0,
the residue field is an S-module, so we can take n = 1 and N1 = k in Lemma 6.3. In this
case, the formula of Lemma 6.3 simplifies to the stated one. 
One interesting set whose measure we can calculate in the local rings case is the set of
finite R-modules. This is a measurable set since it is a countable union of closed points.
Lemma 6.8. Let R be an infinite Noetherian local commutative ring with finite residue field
k. We have

0
 if u < 0
Q∞ −j−u
µu ({N | N finite}) = µu ({0}) = j=1 (1 − |k| ) if 0 ≤ u < dim R − 1

1 if u ≥ dim R − 1
45
where dim R is the Krull dimension. In the middle case, saying that the measure is µu ({0})
is equivalent to saying each nonzero finite R-module has measure 0.
Proof. Let N be a finite R-module. Then N has a minimal presentation in terms of d
generators and m relations for d = dim Hom(N, k) and m = dim Ext1R (N, k). We can choose
a finite quotient S of R such that N factors through S, as do all finite extensions of N by k,
so that Ext1S (N, k) = Ext1R (N, k). It follows that for an R module L, we have L ⊗R S ∼
=N

if and only if L = N, with “if" obvious and “only if" because otherwise some quotient of L
would be an extension of N by k and thus an S-module, a contradiction.
Then

Y
µu ({N}) = µu (US,N ) = |Aut(N)|−1 |N|−u (1 − |k|m−d−j−u )
j=1

which is 0 if m − d − j − u = 0 for some j > 0, equivalently, if m − d − u > 0. So we see the


probability of attaining N is nonzero only if m − d ≤ u.
We will now show that if N is a nonzero finite R-module, then m − d ≥ dim R − 1. The m
relations of N form an d×m matrix with entries in R, and let I be the ideal generated by the
d × d minors of that matrix. If I is the unit ideal, then N is trivial. If d − n ≤ dim R − 2 and
I is not the unit ideal, then I is contained in a prime of height dim R − 1 [BV88, Theorem
2.1], so a non-maximal prime ℘. If S is the fraction field of R/℘, then N ⊗R S ≃ S k for
some k > 1 since I ⊂ ℘ implies that that the relations defining N ⊗R S are rank at most
d − 1 over S. Thus, in this case N must be infinite, as if N were finite, then it would
be annihilated by some power of an element in the maximal ideal but not ℘, which would
contradict N ⊗R S ≃ S k
Thus, if u < dim R − 1, then we cannot have m − d ≤ u for N a nonzero R-module, and
so the probability of each nonzero finite R-module is zero.
For the zero module, we have d = m = 0 so the probability of the zero module Q∞is also
−u
zero for u < 0. If u ≥ 0, the probability of the zero R-module is given by 1 j=1 (1 −
0−0−j−u Q∞ −j−u
|k| ) = j=1 (1 − |k| ), finishing the proof of the middle case.
We finally consider the case u ≥ dim R − 1. In this case, we first check that the quotient
of R̂n by n + u random relations is finite with probability 1. Let Am be the random matrix
over R̂ who columns are the first m relations. Let It (m) be the ideal generated by t × t
minors of Am . We say a prime ℘ in R̂ of height h is bad if ℘ is not maximal, and if for
some h + 1 ≤ k ≤ n + h we have that the rank of Ak mod ℘ is < k − h, or equivalently,
Ik−h (k) ⊂ ℘. We will show by induction on h and k that the probability that there exists a
bad prime is 0.
We condition on the first k −1 columns of the matrix. If Ik−h (k −1) 6⊂ ℘, then Ik−h (k) 6⊂ ℘.
If h > 0, then we know by induction that I(k−1)−(h−1) (k − 1) is not contained in any prime
of height h − 1. Thus any prime of height h containing Ik−h (k − 1) is a prime of height 0
in R/Ik−h (k − 1). There are only finitely many such primes (given Ak−1 ). For each of these
finitely many primes ℘ of height h, the matrix Ak−1 mod ℘ has rank at least k − 1 − h with
probability 1 by the inductive hypothesis, and we choose some (k − 1 − h) × (k − 1 − h) minor
of Ak−1 that is non-zero mod ℘. We condition on all entries of Ak except for one entry in the
kth column and a row not in the chosen minor. Then there is at most one value that entry
can be mod ℘ so that Ak mod ℘ is rank < k − h. Since R̂/℘ is infinite (since it is dimension
at least 1), a Haar random element of R̂ takes that value mod ℘ with probability 0. Since
46
there are only finitely many primes of height h containing Ik−h (k − 1) (and the rest cannot
contain Ik−h (k)), this completes the induction.
Let N be the quotient of R̂n by n + dim R − 1 random relations. With probability 1, we
have that in the quotient by every non-maximal prime ℘ of R̂, the matrix A(n + dim R − 1)
is rank n, which implies that the localization N℘ = 0, which implies that the annihilator
IN of N is not contained in ℘. So, with probability 1, we have that R̂/IN is 0-dimensional
and hence finite since the residue is finite, and the finitely generated R̂/IN -module N is thus
finite.
To prove that µu of the set of finite R-modules is 1, it remains to prove that
Prob(R̂n /hn + u Haar random relationsi ∼
X
lim = N)
n→∞
N finite R-module

lim Prob(R̂n /hn + u Haar random relationsi ∼


X
= = N)
n→∞
N finite R-module
as we have just seen the left side is 1.
We do this using the dominated convergence theorem. The probability that R̂n modulo
n + u Haar random relations is isomorphic to N is bounded above by |Aut(N)|−1 |N|−u (see
the proof of Lemma 6.6), and is 0 if the minimal number of relations m to define N, minus
the minimal number of generators d, is greater than u.
On the other hand, if then m − d ≤ u, then
Y∞ ∞
Y
−1 −u m−d−j−u −1 −u
µu ({N}) = |Aut(N)| |N| (1 − |k| ) ≥ |Aut(N)| |N| (1 − |k|−j )
j=1 j=1

so in either case we have



Prob(R̂ /hn + u Haar random relationsi ∼
Y
n
= N) ≤ µu ({N}) (1 − |k|−j )−1
j=1

and this upper bound satisfies the assumption of the dominated convergence theorem since
X ∞
Y Y∞ Y∞
−j −1 −j −1
µu ({N}) (1 − |k| ) ≤ µu (P) (1 − |k| ) = (1 − |k|−j )−1 < ∞.
N finite R-module j=1 j=1 j=1


For a discrete valuation ring, we can give a simple condition for a collection of moments to
be well-behaved that is almost sharp. For an R-module N, we write ∧2 N for the quotient of
the R-module N ⊗ N by the R-module generated by elements of the norm n ⊗ n for n ∈ N.
Lemma 6.9. Let R be a discrete valuation ring with finite residue field. Let MN be an
assignment of a real number to each isomorphism class of R-modules. Then MN is well-
behaved if there exists ǫ > 0 and c such that MN ≤ c|N|1−ǫ |∧2 (N)| for all finite R-modules
N.
We then obtain corollaries for uniqueness of measures from Theorems 1.6 and 1.7. A
slightly weaker uniqueness result appears in [Woo17, Theorems 8.2 and 8.3], without the
|N|1−ǫ in the upper bound (though stated there over Zp , the argument works over other
DVR’s with finite residue field, e.g. see [WW21, Theorem 6.11]).
47
Lemma 6.9 is almost sharp because, if we set MN = |N||∧2 (N)| = Sym2 (N) then the

moments are attained for two distinct measures µ: The probability distribution the quotient
of R̂n by the columns of a Haar random n×n skew-symmetric matrix converges as n → ∞ for
n even to one measure, as n → ∞ for n odd to a different measure, which both have the
and
same moments Sym2 (N) . (See [CKL+ 15, Theorems 10 and 11] for an argument showing

moments in a different, but analogous case, and see [BKL+ 15, Theorems 3.9 and 3.11] for
the limiting distributions in the case R̂ = Zp .) So if we allow ǫ = 0 then MN ≤ c|N||∧2 (N)|
are not necessarily well-behaved since the conclusion of Theorem 1.7 is not necessarily true.

Proof. Let k be the residue field of R and q its cardinality. Let π be the uniformizer of R.
From Proposition 6.2, since k is the only simple R-module, it suffices to prove for each
finite R-module N that


!
1 X 1 X
c|Nα |1−ǫ ∧2 (Nα ) < ∞.

e Qe j
|Aut(N)| e=0 |Hom(N, k)| j=1 (q − 1)
α∈Ext1 (N,k e )

By the classification of modules over PIDs, we can write N ∼


Qn
= π=1 R/π di for some tuple
d1 , . . . , dn of positive integers. Then using the exact sequence 0 → Rn → Rn → N → 0,
we can write Ext1 (N, k e ) ∼ = k ne . Explicitly, we can represent an extension class as an e × n
matrix where the ith column is π di times a lift of the generator of R/π di from N to Nα .
We now describe ∧2 (Nα ). This has a filtration into submodules where F1 ⊂ ∧2 (Nα ) is
generated by all a ∧ b where a, b, ∈ k e , the second term F2 ⊂ ∧2 (Nα ) is generated by all a ∧ b
where a ∈ k e , and the last term F3 = ∧2 (Nα ). Then

F3 /F2 ∼
= ∧2 N,

while F2 /F1 is a quotient of N ⊗ k e ∼= k ne , and F1 is a quotient of ∧2 (k e ). Furthermore, if


a ∈ k e lies in the image of the matrix representing α then a ∈ πNα so for b ∈ k e , we have

a ∧ b ∈ πNα ∧ b = Nα ∧ πb = Nα ∧ 0 = 0,

and thus F1 is a quotient of ∧2 (k e / Im(α)). Thus

2 (e−rank(α))(e−rank(α)−1)
∧ (Nα ) = |F3 /F2 ||F2 /F1 ||F1 | ≤ ∧2 (N) q en ∧2 (k e / Im(α)) = ∧2 (N) q en q

2 .

We also note that |Nα | = |N|q e and |Hom(N, k)| = q n . Thus


48

!
1 X 1 X
c|Nα |1−ǫ ∧2 (Nα )

e Qe j
|Aut(N)| e=0 |Hom(N, k)| j=1 (q − 1)
α∈Ext1 (N,k e )

!
1 X 1 X (e−rank(α))(e−rank(α)−1)
c(|N|q e )1−ǫ ∧2 (N) q en q

≤ ne
Qe j
2
|Aut(N)| e=0 q j=1 (q − 1) α∈Ext1 (N,k e )

!
c|∧2 (N)||N|1−ǫ X 1 X (e−rank(α))(e−rank(α)−1)
= Qe j
(q e )1−ǫ q 2
|Aut(N)| e=0 j=1 (q − 1) α∈Ext1 (N,k e )

! n
c|∧2 (N)||N|1−ǫ X 1 X (e−r)(e−r−1)
= Qe j
(q e )1−ǫ q 2 #{α ∈ Hom(k n , k e ) | rank(α) = r}
|Aut(N)| e=0 j=1 (q − 1) r=0

! n ∞
! n
X 1 X
e 1−ǫ
(e−r)(e−r−1)
re
X 1 X e(e+1−2ǫ) r2 +r
≤CN Qe j − 1)
(q ) q 2 q = Q e j − 1)
q 2 + 2
e=0 j=1 (q r=0 e=0 j=1 (q r=0
∞ e(e+1−2ǫ) ∞
!
X q 2 1 X
≤CN′ Qe j − 1)
≤ Q ∞ −j )
q −ǫe < ∞,
e=0 j=1 (q j=1 (1 − q e=0

where CN , CN′ are constants only depending on c, N. 


6.2. Groups. As in the example of modules, we show that the category of finite groups is a
diamond category, simplify the formulas for the measure for general moments, and simplify
them further and verify positivity for certain nice moments. The reader will notice the
proofs are quite similar, and indeed one can make analogous computations in many diamond
categories. Instead of carrying out these computations in further examples, we hope that
the examples given here will give the reader a blueprint for how to do the computations in
further examples of interest. One can see see [SW22b] and our forthcoming paper [SW22a]
for more examples of analogous computations in other cases.
Lemma 6.10. The category of finite groups is a diamond category.
Proof. The poset of quotients of a group is dual to the poset of normal subgroups, and it is
a classical fact (easy to check) that the normal subgroups form a modular lattice [Bir67, Ch.
I, Thm. 11 on p. 13] (a property that is manifestly stable under duality). The finiteness
of automorphism groups is trivial, and the last finiteness follows from [LW20, Corollary
6.13]. 
Proof of Theorem 1.2. Using Lemma 6.10, we have that Theorems 1.7 and 1.8 specialize to
Theorem 1.2, except that we need a translation from pro-isomorphism classes to profinite
groups with finitely many continuous homomorphisms to any finite group, which is provided
by Lemmas 5.7 and 5.8. 
We now introduce some notation that will help to classify simple surjections of finite
groups. Let G be a finite group. A [G]-group is a group H together with a homomorphism
from G to Out(H). A morphism of [G]-groups is a homomorphism f : H → H ′ such that
for each element g ∈ G, for each lift σ1 of the image of g from Out(H) to Aut(H), there is
a lift σ2 of the image of g from Out(H ′) to Aut(H ′) such that σ2 ◦ f = f ◦ σ1 . A [G]-group
49
isomorphism is then defined a [G]-group morphism with an inverse [G]-group morphism.
We write Aut[G] (H) for the group of automorphisms of G as an [H]-group. Note that G
acts on the set of normal subgroups of a [G]-group H, and we say a nontrivial [G]-group is
simple if it has no nontrivial proper fixed points for this action. If we have an exact sequence
1 → N → H → G → 1, then N is naturally a [G]-group. A normal subgroup N ′ of N is
fixed by the Out(G) action if and only if it is normal in H, and so N is a product of simple
H-groups if and only if it is a product of simple [G]-groups.
We can now give a formula for measures in the category of groups that is more cumbersome
to state than the general formula but will be easier to use in applications.
Proposition 6.11. Let C be a level in the category of groups and (MG )G ∈ RC .
Fix a finite group F . Let V1 , . . . , Vn be representatives of all isomorphism classes of irre-
ducible representations of F over prime fields such that Vi ⋊ F ∈ C. For each i, let Wi be the
subset of H 2 (F, Vi ) consisting of classes whose associated extension of F by Vi is in C and
let qi be the order of the endomorphism field of Vi . Let N1 , . . . , Nm be representatives of all
isomorphism classes of pairs of finite simple [F ]-groups N such that F ×Out(N ) Aut(N) is in
C.
For nonnegative integers e1 , . . . , en , f1 , . . . , fm and αi ∈ Wiei ⊆ H 2 (F, Vi )ei = H 2 (F, Viei ),
let Ge,f ,α be the fiber product over F of, for each i from 1 to n, the extension of F by Viei
associated to αi , with, for each i from 1 to m, F ×Out(Ni )fi Aut(Ni )fi .
Then we have
n
! m !
e
1 X Y (−1)ei |H 0 (F, Vi )| i Y (−1)fi X
vC,F = MGe,f ,α .
|Aut(F )| e∈Nn ,f ∈Nm i=1 |Vi |ei |H 1 (F, vi )|ei ej=1 (qij − 1) i=1 Aut[F ] (Ni ) fi (fi )!
Q i

Qn ei
α∈ W i=1 i

The MG are well-behaved at level C if and only if the similar sum


n
! m !
e
X Y |H 0 (F, Vi )| i Y 8 fi X
ei e Q e j f
MGe,f ,α
|V | |H 1 (F, v )| i i


n m i i (q
j=1 i 1) Aut [F ] (N i ) i
(fi )! Q e
e∈N ,f ∈N i=1 i=1 α∈ n W i
i=1 i

is absolutely convergent.
Proof. By definition we have
X X µ(π)MG
vC,F = .
G∈C π∈Epi(G,F )
|Aut(G)||Aut(F )|

The µ(π) term vanishes unless π is semisimple, in which case ker π is a product of finite
simple [F ]-groups, of which the abelian ones are irreducible representations of F over prime
fields, So, we can write ker π as ni=1 Viei × m fi
Q Q
i=1 Ni for unique tuple e1 , . . . , en , f1 , . . . fm
(ei ) Q
of nonnegative integers. In this case, µ(π) = ni=1 (−1)ei qi 2 · m fi
Q
i=1 (−1) by Lemma 3.10.
Thus
n  Ym
1 X Y ei
ei ( 2 )
X X MG
vC,F = (−1) qi (−1)fi ,
|Aut(F )| e∈Nn ,f ∈Nm i=1 i=1 G∈C
|Aut(G)|
π∈Epi(G,F )
Q ei Qm fi
ker π ∼
= ni=1 Vi × i=1 Ni

= here denotes an [H]-group isomorphism. Now ker π ∼


where ∼
Qn Qm f
= i=1 Viei × i=1 Ni i if
and only if there exists an [F ]-group isomorphism ι : i=1 Viei × m
Qn Q fi
i=1 Ni → ker π. By
50
[Saw20, Lemmas 2.3, 2.4, 2.5] and the easily-checked fact that bijective [F ]-morphisms are
[F ]-isomorphisms, when one such isomorphism exists, then there are exactly
ei
n m n
! m
ei Y
q ( 2 ) (q j − 1)
fi
Aut[F ] (Ni ) fi fi !
Y Y Y Y
|GLe (Fq )|
i i
Aut[F ] (Ni ) fi ! =
i=1 i=1 i=1 j=1 i=1

such isomorphisms. Thus


n m
1 X Y (−1)ei Y (−1)fi X X MG
vC,F = Q ei .
|Aut(F )| e∈Nn ,f ∈Nm i=1 j=1(q j − 1) i=1 Aut[F ] (Ni ) fi fi ! G∈C |Aut(G)|

π∈Epi(G,F )
Qn ei Qm fi
ι: i=1 Vi × i=1 Ni →G
Im ι=ker π
ι an [F ]-group morphism

For each triple G, π, ι we can quotient G by all the factors of ker π except ι(Viei ), obtaining
an extension of F by Viei , with an extension class αi ∈ H 2 (F, Vi )ei . We can quotient further
to obtain ei individual extensions of F by Vi , each of which must be level-C, hence αi ∈ W ei .
There is a natural isomorphism from G to the fiber product, over F , of each of the minimal
extensions of F obtained by quotienting G by all of the Vi and Ni factors except one. One
can check that the minimal extension by Ni is naturally isomorphic to Aut(Ni ) ×Out(Ni ) F ,
where G maps to Aut(Ni ) by its conjugation action and to F by the given map (using that
the center of Ni is trivial). Thus we have an isomorphism G → Ge,f ,α , respecting both the
map from ni=1 Viei × m fi
Q Q
i=1 Ni and the maps to F .
We next check that, for given e, f, each class α arises from exactly
n  ei
Y |H 0 (F, Vi )|
Aut(Ge,f ,α)
i=1
|Vi ||H 1 (F, Vi )|
triples G, π, ι. First, we note that each class α arises from G = Ge,f ,α and taking π, ι to be
the natural maps, as Ge,f ,α ∈ C. We take the group Ge,f ,α to be the representative of its
isomorphism class we use for the sum. From the last paragraph, we see that if Ge,f ,α , π ′ , ι′
gives α, then there is some automorphism β of Ge,f ,α such that ι′ = βι and π ′ = πβ. So
Aut(Ge,f ,α ) acts transitively on the non-empty set of triples G, π, ι that give α, and the
stabilizer is the subgroup of automorphisms preserving ι and π.
So it suffices to prove ei the number of automorphisms of Ge,f ,α that preserve ι and π
Qn |Vi ||H (F,Vi )|
1
is i=1 |H 0 (F,Vi )|
. To do this, note that any such automorphism sends g ∈ G
to the product of g with an element of ni=1 Viei × m fi
Q Q
i=1 Ni , which since the automor-
phism fixes Nifi and thus fi
Qnfixesei the action of g on Ni 1 must Qncommute with Nifi for all
i and thus must lie in i=1 Vi , giving a cocycle in C (F, i=1 Viei ), and every cocycle
gives such
Qn an automorphism. The number of cocycles is the size of the cohomology Qn group
Qn ei
1 ei 1
|H (F, i=1 Vi )| = i=1 |H (F, Vi )| times the number Q of coboundaries, which is i=1 |Vi |ei
e
divided by the number of cocycles in degree 0, which is ni=1 |H 0 (F, Vi )| i , giving the stated
formula.
Thus, we can replace the sum over G, π, ι with a sum over α after multiplying by
n   ei
Y |H 0 (F, Vi )|
Aut(Ge,f ,α) ,
i=1
|Vi ||H 1 (F, Vi )|
51
which cancels the Aut(G) term and gives the desired statement.
By definition, the assignment MG is well-behaved if and only if the sum, with an additional
Z(π)3 factor,Pm is absolutely convergent. We have Z(π) = 2
ω(π)
where ω(π) = #{1 ≤ iP≤ n |
m
ei > 0} + i=1 fi , by Lemma 3.10, and thus Z(π) is bounded by a constant 2n times 2 i=1 fi .
Thus,
Pm the MG are well-behaved if and only if the sum, with the signs removed and an
fi
8 i=1 factor inserted, is absolutely convergent. We may also ignore the Aut(F ) factor as
it is bounded. 
Lemma 6.12. Let C be the category of finite groups. Let u be a real number, and for each
G ∈ C, let MG = |G|−u . Then (MG )G is well-behaved.
Keeping the notation of Proposition 6.11, we have
n ∞  ! Ym −
|F |−u Y Y
1
|H 0 (F, Vi )||Wi | −j |Ni |u |Aut[F ] (Ni )|
vC,F = 1− qi e .
|Aut(F )| i=1 j=1 |Vi |u+1 |H 1 (F, Vi )| j=1

If u is an integer then vC,F ≥ 0 for all levels C.


Thus, by Theorem 1.7, for u ∈ Z, there is a unique measure on the set of isomorphism
classes of pro-finite groups with moments |G|−u and it is given by the formulas in Lemma 6.12.
For a fixed u, if one takes Xg to be the free profinite group on g generators, modulo g + u
independent Haar random relations, it is easy to compute that
lim E(Epi(Xg , G)) = |G|−u
g→∞

for every finite group G. So then Theorem 1.8 implies that the Xg weakly converge in
distribution to a probability measure on the space of profinite groups, indeed the one given
by the formulas in Lemma 6.12. This recovers [LW20, Theorem 1.1]. One can reconcile the
formulas in Lemma 6.12 with [LW20, Equation (3.2)] using observations from the proof of
Proposition 6.11.
Proof. We observe that
n
Y m
Y n
Y m
Y
−u ei fi −u −u −uei
MGe,f ,α = |Ge,f ,α | = (|F | · |Vi | · |Ni | ) = |F | · |Vi | · |Ni |−ufi
i=1 i=1 i=1 i=1

so by Proposition 6.11, (MG )G is well-behaved if and only if


n
! m !
0 ei −uei fi −ufi
X Y |H (F, Vi )| |Vi | Y 8 |Ni | X
|F |−u 1
|Vi |ei |H 1 (F, vi )|ei ej=1 (qij − 1) i=1 Aut[F ](Ni ) fi (fi )!
Q i

Qn ei
e∈Nn ,f ∈Nm i=1 α∈ i=1 Wi
Qn ei
is absolutely convergent. Since the length of the sum over α is i=1 |Wi | , this is absolutely
convergent if and only if
n
! m !
e
−u
X Y |H 0 (F, Vi )| i |Vi |−uei |Wi |ei Y 8fi |Ni |−ufi
|F |
|Vi |ei |H 1 (F, vi )|ei ej=1 (qij − 1) i=1 Aut[F ] (Ni ) fi (fi )!
Qi
e∈Nn ,f ∈Nm i=1

is. e
|H 0 (F,Vi )| i |Vi |−uei |Wi |ei
Now note that every term in the ratio |V |ei |H 1 (F,v )|ei Qei (qj −1) is exponential in ei , ex-
i i j=1 i

cept for ej=1 (qij − 1), which is superexponential, so the ratio decays superexponentially.
Qi
52
8fi |Ni |−ufi
Similarly, every term in f is exponential in fi except for (fi )!, which is super-
|Aut[F ] (Ni )| i (fi )!
exponential, so the ratio decays superexponentially. Thus, the term being summed decays
superexponentially in all the ei , fi variables, and hence the sum is absolutely convergent.
To calculate vC,F , we similarly obtain
n
! m !
e
|F |−u X Y (−1)ei |H 0(F, Vi )| i |Vi |−uei Y (−1)fi |Ni |−ufi X
vC,F = 1
|Aut(F )| e∈Nn ,f ∈Nm i=1 |Vi |ei |H 1 (F, vi )|ei ej=1 (qij − 1) i=1 Aut[F ] (Ni ) fi (fi )!
Q i

Qn ei
α∈ i=1 Wi

n
! m
!
e
|F |−u X Y (−1)ei |H 0 (F, Vi )| i |Vi |−uei |Wi |ei Y (−1)fi |Ni |−ufi
=
|Aut(F )| e∈Nn ,f ∈Nm i=1 |Vi |ei |H 1 (F, vi )|ei ej=1 (qij − 1) i=1 Aut[F ] (Ni ) fi (fi )!
Qi

n ∞
! m ∞
!
e
|F |−u Y X (−1)e |H 0 (F, Vi )| |Vi |−ue |Wi |e Y X (−1)f |Ni |−uf
=
|Aut(F )| i=1 e=0 |Vi |e |H 1 (F, Vi )|e ej=1 (qij − 1) i=1 f =0 Aut[F ] (Ni ) f f !
Q

n ∞  ! Ym −
|F |−u Y Y
1
|H 0 (F, Vi )||Wi | −j |Ni |u |Aut[F ] (Ni )|
= 1− qi e
|Aut(F )| i=1 j=1 |Vi |u+1 |H 1 (F, Vi )| j=1
P∞ e Q∞
by the q-series identity Q v −j
e=0 ej=1 (q j −1) = j=1 (1 + vq ) and the Taylor series for the
exponential function.
For u an integer, this is nonnegative because, on the one hand, the exponential terms are
|H 0 (F,Vi )||Wi |
nonnegative, and, on the other hand, |V |u+1 |H 1 (F,V )| is an integer power of q because every
i i
term is the cardinality of a vector space over Fqi raised to an integer power. (One can check
that C closed under fiber products and quotients implies that Wi is a vector space over the
endomorphism field of Vi .) If this power is positive then one of the terms in the product over
j is zero, making this factor nonnegative, and otherwise, each term in the product over j is
nonnegative, making this factor nonnegative. 
Corollary 6.13. Let C be the category of finite groups. Let (MG )G ∈ RC/≃ . If MG =
O(|G|n ) for some real n, then (MG )G is well-behaved.
Proof. This follows immediately form Lemma 6.12, taking u = −n, and Lemma 2.11. 
Remark 6.14. It may be possible to prove, following the calculations in [SW22b, §7], that
(MG )G is well-behaved as long as
!
|G|1−ǫ |H2 (G, Z)|
MG = O
|H1 (G, Z)|
for some ǫ > 0, but we do not pursue that here.
6.3. General tools for constructing examples. In this section, we give a general theorem
which gives a large class of examples of diamond categories, as well as tools for constructing
new diamond categories from old ones.
A Mal’cev variety [Smi76, 142 Theorem] is a variety in universal algebra (a set of abstract
n-ary operations for different natural numbers n together with a set of equations they satisfy)
such that some composition of the n-ary operations is an operation T (x, y, z) which satisfies
T (x, x, z) = z and T (x, z, z) = x. The most familiar example of such an operation is given
53
by T (x, y, z) = xy −1 z in the variety of groups, or in any variety which contains a group
operation. Given a Mal’cev variety, one has a category of finite algebras for that Mal’cev
variety. An object in this category is a finite set S with functions S n → S for each abstract n-
ary operation of the Mal’cev variety (for each n) such that the functions satisfy the equations
of the Mal’cev variety. A morphism in this category is a function S → S ′ that respects all the
operations. The congruences in any algebra for a Mal’cev variety (i.e. equivalence relations
respecting the algebra structure, analogous to normal subgroups of a group or ideals of a
ring) form a modular lattice [Bir67, Ch. VII: Theorem 4 and Theorem of Malcev].
Lemma 6.15. For any Mal’cev variety with operations of finitely many different arities, the
category of finite algebras for that Mal’cev variety, with morphisms given by surjections, is
a diamond category.
This includes the cases of finite rings, finite commutative rings, finite R-modules for a ring
R, finite groups, finite groups with the action of a fixed group Γ, finite quasigroups, finite
Lie algebras over a finite field, finite Jordan algebras over a finite field, etc.
Proof. In this proof, we use “algebra” to refer to an algebra for the given Mal’cev variety.
Since we choose morphisms to be surjections, every morphism is an epimorphism, and
the epimorphisms are dual to the lattice of congruences [Bir67, Ch. VI, Sec. 4] which is a
modular lattice as mentioned above. The lattice is finite since the number of congruences is
at most the number of partitions of the underlying set.
The automorphism group of a finite algebra is a subgroup of the permutation group of the
underlying set, and thus is finite.
The last finiteness step is more difficult. If A and B are both quotients of D, then the
elements of A ∨ B are equivalence classes of elements of D, where the equivalence is that
the equivalence defining A and the equivalence defining B both hold. Since we are in a
Mal’cev variety, the elements of A ∧ B are equivalence classes of elements of D, where the x
is equivalent to y if there is some z such that x and z are equivalent in A and z and y are
equivalent in B ([Bir67, Ch. VII: Theorem 4]). Thus, in the setwise fiber product A ×A∧B B
(which has a natural algebra structure), every element is in the image of the map from D.
Thus D → A ×A∧B B is a surjection of algebras, and the equivalence relation giving the
fibers is precisely the same as that defining the fibers of the surjection D → A ∨ B, and
hence A ∨ B = A ×A∧B B.
For any simple surjection between algebras, let the width be the maximum size of a fiber
of the underlying map of sets. For any algebra A, let the width of A be the maximum width
of any simple surjection between two quotients of A. Let n be the maximum width of an
element of a finite generating set of C.
Since A ∨ B = A ×A∧B B, we have that the sizes of the fibers of A ∨ B → A are all sizes
of fibers of B → A ∧ B and vice versa. We next will see that under certain conditions, the
width of E → F is the same as the width of E ∨ G → F ∨ G or E ∧ G → F ∧ G. Suppose
E → F is simple such that F ≥ G ∧ E (all as quotients of some object). Then ∨G gives
an isomorphism [G ∧ E, E] → [G, E ∨ G] by Lemma 2.1, and so E ∨ G → F ∨ G is simple.
Moreover, (F ∨G)∨E = E ∨G and (F ∨G)∧E = F ∨(G∧E) = F (by the modular property
and then the assumption that F ≥ G ∧ E). Thus, E ∨ G → F ∨ G is (F ∨ G) ∨ E → F ∨ G
and has the same width as E → (F ∨ G) ∧ E, which is E → F . Similarly, if E → F is simple
such that E ≤ G ∨ F , then E ∧ G → F ∧ G is simple has has the same width as E → F .
54
Our first step is to check that all level-C algebras have width at most n. By the definition
of level-C, it suffices to check that the set of algebras of width at most n is downward-stable
and join-stable. That it is downward-stable is clear by definition. For join-stable, consider a
simple surjection B → A of quotients of D ∨ E (in some lattice of quotients), where D and
E have width at most n. We have B ≥ (D ∧ B) ∨ A ≥ A so, by simplicity (D ∧ B) ∨ A, is
either equal to B or A.
If (D ∧ B) ∨ A = B, then B = (D ∧ B) ∨ A ≤ D ∨ A, so by the above D ∧ B → D ∧ A is
simple and has the same width as B → A. Since D ∧ B → D ∧ A is a simple surjection of
quotients of D, we conclude B → A has width at most n.
If (D ∧ B) ∨ A = A, then A ≥ (D ∧ B). By the above, B → A has the same width
as B ∨ D → A ∨ D. Moreover, since B ∨ D ≤ E ∨ D = E ∨ (A ∨ D), we have that
(B ∨ D) ∧ E → (A ∨ D) ∧ E is simple and also has the same width as B → A. However,
(B ∨ D) ∧ E → (A ∨ D) ∧ E is a simple map of quotients of E, and hence has width at most
n. Thus B → A has width at most n.
This concludes the proof that all level-C algebras have width at most n. It follows that, for
G of level-C, any H of level C with a simple surjection H → G has cardinality |H| ≤ |G|n.
If there are finitely many operations in the variety, we are done, as there are finitely many
algebra structure on any given finite set.
Otherwise, we need a slightly more complicated argument. Let K be the product of all
the elements of a finite generating set of C. Any equation (involving the algebra operations)
that holds in K holds in every level-C algebra, since equations are preserved by quotients
and fiber products. Since K is a finite set, there can only be finitely many distinct r-ary
maps K → K for each r, so sufficient equations hold in K so that there is some finite set of
operations, such that the equations holding on K, and hence for level-C algebras, determine
the value of every operation in terms of that finite set of operations. It then holds that there
are finitely many level-C algebras on sets of cardinality ≤ |G|n, completing the proof of the
final finiteness step. 
One we have a diamond category, there are many ways to modify it and get further
diamond categories.
Lemma 6.16. Let C1 and C2 be diamond categories. Then the product category C1 × C2 ,
whose objects are ordered pairs of an object of C1 and an object of C2 and whose morphisms
are ordered pairs of morphisms, is a diamond category.
Proof. We verify the three assumptions in order:
(1) The poset of quotients of an ordered pair (G1 , G2 ) is the product of the posets of
quotients of G1 and G2 , and the product of two finite modular lattices is a finite
modular lattice.
(2) The automorphism group of an ordered pair (G1 , G2 ) is the product of the automor-
phism groups, and the product of two finite groups is a finite group.
(3) Since the poset of quotients is the product of quotients, the product of two downward-
closed sets is downward-closed, and the product of two join-closed sets is join-closed.
Thus, the level-C objects are contained in the set of ordered pairs (G1 , G2 ) where G1
is level-C1 and G2 is level-C2 , for C1 the set of first terms of objects of C and C2 the
set of second terms. A morphism to (G1 , G2 ) is simple if and only if it is simple in
one factor and an isomorphism to the other, so the finiteness of simple morphisms
55
from level-C objects follows from the finiteness of simple morphisms from level-C1 and
level-C2 objects.

The following result will cover many kinds of additional structure we may add to a diamond
category to get another diamond category.
Proposition 6.17. Let F : C1 → C2 be a faithful functor that sends epimorphisms to epi-
morphisms. Assume:
• C2 is a diamond category.
• The natural map from the poset of quotients of G to the poset of quotients of F (G) is
the inclusion of a sublattice (i.e. we have H1 ≤ H2 if F (H1 ) ≤ F (H2 ) and the image
is closed under joins and meets)
• That there is some n such that F applied to any simple epimorphism gives a morphism
of dimension at most n.
• Given a level C of C1 , there are finitely many isomorphism classes of objects G ∈ C
with F (G) isomorphic to a given object in C2 .
Then C1 is a diamond category.
Remark 6.18. When applying Proposition 6.17, it is very helpful when checking the condi-
tions to note that F faithful implies that if F (π) is an epimorphism then π is an epimorphism.
Proof. We verify the three assumptions in order:
(1) The poset of quotients of G is a sublattice of the poset of quotients of F (G). A
sublattice of a finite modular lattice is modular since a modular lattice is defined by
an algebraic equation involving joins and meets.
(2) The automorphism group of G is a subgroup of the automorphism group of F (G),
hence a finite group.
(3) Let C be a level of C1 and let C ′ be the level of C2 generated by the application of
F to a generating set of C. Then the objects G ∈ C1 such that F (G) is level C ′ are
downward-closed and join-closed by the assumption on the natural map of posets,
so F (C) ⊂ C ′ . So it suffices to prove for an object H that there are finitely many
simple morphisms G → H with F (G) ∈ C ′ . By Lemma 2.8 and assumption, there
are finitely many possibilities for F (G) up to isomorphism. By assumption, there
are finitely many possibilities for G ∈ C up to isomorphism, and then each one has
finitely many surjections to H by the previous two properties.

One such structure is adding the action of a finite group, where F is the forgetful functor.
Lemma 6.19. Let C be a diamond category. Let Γ be a finite group. Define the category
Γ-C in which objects are pairs of an object G ∈ C and a homomorphism Γ → Aut(G) and
morphisms are epimorphisms in C compatible with the action of Γ. Then Γ-C is a diamond
category.
Proof. We verify the conditions of Proposition 6.17 for the forgetful functor F : Γ-C → C
forgetting the Γ action, and then the lemma will follow. We have that F is faithful and takes
all morphisms to epimorphisms (by definition of the morphisms in Γ-C). We have assumed
56
that C is a diamond category. We can check that the poset of quotients of an object G
with an action of Γ is the subposet of Γ-invariant elements of the poset of quotients of G.
Since the join and meet of Γ-invariant elements is Γ-invariant, the Γ-invariant elements of
the poset of quotients of an object of C is a sublattice as required.
A simple morphism of objects with an action of Γ has an underlying morphism of objects
of dimension at most |Γ|. Indeed, let G → H be a morphism of objects, and suppose we
have compatible actions of Γ on G and H. The induced morphism of objects with an action
of Γ is simple if there is no K ∈ [H, G] that is Γ-invariant other than H and G. Take any
element of [H, G] that is simple over H and take the join of its orbit under Γ. Since this
join depends only on the orbit, it is invariant under Γ, and it cannot be H, so it must be G.
Thus G is the join of at most |Γ| objects simple over H and thus must have dimension at
most |Γ| over H.
Since Γ and the automorphism group of any object of C are finite, there are finitely many
isomorphism classes of objects in Γ-C mapping to an object of C. 
Several previous works have considered Γ-groups. If we let C be the category of finite
Γ-groups whose order is relatively prime to |Γ|, Liu, Zureick-Brown and the second author
proved the existence of, and gave explicit formulas for, a measure on the pro-category with
moments [G : GΓ ]−u [LWZ19, Theorem 6.2]. The first author [Saw20, Theorem 1.2] proved
uniqueness and robustness in this setting for moments that are O(|G|n) for some n. The
results of the current paper, along with following the “blueprints” above allow one to give a
different proof of these explicit formulas and uniqueness and robustness results.
Another structure is the addition of a map to a fixed object.
Lemma 6.20. Let C be a diamond category. Let K be an object of C. The category K-Ext
of K-extensions whose objects are objects of C together with an epimorphism to K and where
morphisms are epimorphisms of C respecting the epimorphisms to K is a diamond category.
This construction is a variant of the notion of slice category, restricted to epimorphisms.
Applied to the category of finite groups, this produces the category of groups with a
surjection onto a fixed finite group K. This is closely related to the category of groups with
an action of K, as if G is a group with an action of K then G⋊K is a group with a surjection
onto K.
Proof. We verify the conditions of Proposition 6.17 for the forgetful functor F taking an
object of C together with an epimorphism to K to the underlying object, and then the lemma
will follow. This is faithful and takes all morphisms to epimorphisms (by construction of the
category). We have assumed that C is a diamond category. For G an object and f : G → K
a surjection, the poset of quotients of (G, f ) is simply the interval [K, G], which is indeed
a sublattice. A morphism (G, f ) → (H, g) is simple if and only if G → H is simple, so F
applied to a simple morphism is always simple. There are finitely many isomorphism classes
of pairs (G, f ) for any G because there are finitely many epimorphisms f : G → K. in C
since C being a diamond category implies that quotient lattices and automorphism groups
are finite. 
We can also add any kind of finite data that can pushforward along morphisms.
Lemma 6.21. Let C be a diamond category. Let G be a functor from C to the category of
finite sets. Then the category (C, G) of pairs of an object G ∈ C together with an element
57
s ∈ G(G), with morphisms (G, sG ) → (H, sH ) given by morphisms f : G → H in C such that
G(f )(sG ) = sH , is a diamond category.
One example of such a functor on the category of finite groups is provided by the kth
group homology for any fixed k. The k = 3 case was important in [SW22b]. The “bilinearly
enhanced groups" studied by Lipnowski, Tsimerman, and the first author in [LST20] can also
be expressed this way, as the set of bilinearly enhanced group structures on a finite abelian
group is a functor from the category of finite abelian groups to finite sets. Our results then
can be used to give a new proof of [LST20, Theorem 8.17], which proves uniqueness and
robustness for finite abelian ℓ-groups with bilinearly enhanced group structures. It should
also be possible to give a new proof of the existence result and formulas of [LST20, Theorem
8.14] using our main results and following the “blueprints” above. In forthcoming work, the
authors will use the results from this paper in the category of Γ-groups with additional finite
data as in Lemma 6.21 (a class in H 3 (G, Z/n)) to give conjectures about the distribution of
Galois groups of maximal unramified extensions of Γ-number fields, as well as q → ∞ results
in the function field case.
Note that, if C has fiber products, then the category of pairs constructed here need not,
unless F (G ×K H) = F (G) ×F (K) F (H). For example, the category of finite groups together
with a kth homology class does not have fiber products. It is for this and similar reasons
that we avoid using fiber products and use joins instead, even though the familiar examples
like groups, rings, and modules have fiber products.
Proof. We verify the conditions of Proposition 6.17 for the (faithful) forgetful functor F
taking a pair (G, sG ) to G, and then the lemma will follow.
If a morphism f : G → H in C such that G(f )(sG ) = sH is an epimorphism (G, sG ) →
(H, sH ), F is another object of C, and h1 , h2 are two morphisms H → F such that h1 ◦ f =
h2 ◦ f . Then
G(h1 )(sH ) = G(h1 )(G(f )(sG )) = G(h1 ◦f )(sG ) = G(h2 ◦f )(sG ) = G(h2 )(G(f )(sG )) = G(h2 )(sH ).
Let sF = G(h1 )(sH ) = G(h2 )(sH ). Then h1 , h2 are two morphisms (H, sH ) → (F, sF ) whose
compositions with f are equal, so they are equal, thus h1 = h2 . Hence f is an epimorphism
G → H in C.
We have assumed that C is a diamond category. The poset of quotients of (G, sG ) is iden-
H
tical to the poset of quotients of G since each quotient H has a unique element G(πG )(sG ) ∈
H
G(H) compatible with the morphism πG from G. A morphism (G, sG ) → (H, sH ) is simple
if and only if G → H is simple, and the dimension of such morphisms is bounded by one.
Since G is valued in finite sets, there are finitely many isomorphism classes of pairs (G, sG )
for a fixed G. 
Moreover,in the situation of Lemma 6.21, we can easily related well-behavedness between
(C, G) and C.
Lemma 6.22. Let C be a diamond category. Let G be a functor from C to the category of
finite sets.
Let (MH )H ∈ R(C,G)/≃ . Then (MH )H is well-behaved (for (C, G)) if and only if ( s∈G(G) M(G,s) )G ∈
P

RC/≃ is well-behaved (for C).


Note that an analogous statement need not be true for other notions of “categories of
objects with extra structure" described in Proposition 6.17.
58
Proof. The key observation is that for π : (G, sG ) → (H, sH ) a map of ordered pairs aris-
ing from a map f : G → H, the interval [H, G] is naturally isomorphic to the interval
[(H, sH ), (G, sG )] and thus Z(π) = Z(f ) and µ(H, G) = µ((H, SH ), (G, sG )). We write f∗ for
G(f ). P
We first show that if ( s∈G(G) M(G,s) )G is well-behaved then (MH )H is as well. For C a
level of the category of pairs, writing C ′ for the level generated by the underlying groups of
a finite set of generators of C, and (F, sF ) ∈ C, we have
X X |µ((F, sF ), (G, sG ))|
Z(π)3 M(G,sG )
| Aut((G, sG ))|
(G,sG )∈C π∈Epi((G,sG ),(F,sF ))
X X |µ((F, sF ), (G, sG ))|
= Z(π)3 M(G,sG )
| Aut((G, sG ))|
(G,sG )∈C f ∈Epi(G,F )
f∗ sG =sF
X X |µ(F, G)|
= Z(f )3 M(G,sG )
| Aut((G, sG ))|
(G,sG )∈C f ∈Epi(G,F )
f∗ sG =sF
X X X |µ(F, G)|
= Z(f )3 M(G,sG )
| Aut(G)|
G∈C ′ sG ∈G(G) f ∈Epi(G,F )
(G,sG )∈C f∗ sG =sF
X X X |µ(F, G)|
≤ Z(f )3 M(G,sG )
| Aut(G)|
G∈C ′ sG ∈G(G) f ∈Epi(G,F )
X X X |µ(F, G)| X
= Z(f )3 M(G,sG ) < ∞,
| Aut(G)|
G∈C ′ sG ∈G(G) f ∈Epi(G,F )) sG ∈G(G)

showing that (MH )H is well-behaved.


For the converse, we take C a level of C, fix F ∈ C, and let C be the level of (C, G) generated
by F , and every object with a simple morphism to F , together with all possible values of
sG . Closure of C under joins implies that for every G ∈ C with a semisimple morphism to F
and sG ∈ G(G), we have (G, sG ) ∈ C. We now assume that (MH )H is well-behaved and have
X X |µ(F, G)| X
Z(f )3 M(G,sG )
G∈C
| Aut(G)|
f ∈Epi(G,F ) sG ∈G(G)
X X |µ(F, G)| X X
= Z(f )3 M(G,sG )
G∈C f ∈Epi(G,F )
| Aut(G)|
sG ∈G(G) sF ∈G(F )
f∗ sG =sF
X X X X |µ(F, G)|
= Z(f )3 M(G,sG )
| Aut(G)|
sF ∈G(F ) G∈C sG ∈G(G) f ∈Epi(G,F )
f∗ sG =sF
X X X |µ((F, sF ), (G, sG ))|
≤ Z(π)3 M(G,sG ) < ∞,
| Aut((G, sG ))|
sF ∈G(F ) (G,sG )∈C π∈Epi((G,sG ),(F,sF ))
P
showing that sG ∈G(G) M(G,sG ) is well-behaved. 
59
We have seen some examples, but there are many more constructions allowed by Propo-
sition 6.17. Proposition 6.17 will be useful because the variations of extra data that may
need to be considered are endless. Liu [Liu22] works with Γ-groups with a certain kind of
homology element, and the current authors plan to work in a similar setting in forthcoming
work. Proposition 6.17 still applies in these settings. We hope the results of this paper and
the “blueprints” above can be used to prove Conjectures 6.1 and 6.3 in [Liu22], which are on
existence, uniqueness, and robustness for measures with certain moments.
6.4. Finite sets and finite graphs. We now give some very simple examples of diamond
categories, which still can lead to interesting results.
Lemma 6.23. The opposite of the category of finite sets is a diamond category.
Proof. This follows from Lemma 6.15 because the opposite of the category of finite sets is
equivalent to the category of finite Boolean algebras, [Awo06, Prop. 7.30] which is a Mal’cev
variety.
It can also be checked directly. A quotient object in the opposite of the category of finite
sets is a subobject in the category of finite sets, i.e., a subset, and subsets of a fixed finite
set form a modular lattice. The automorphism group of a finite set is finite, and simple
morphisms to a finite set S are injections of S into a set of cardinality |S| + 1, of which there
exists only one up to isomorphism. 
Since isomorphism classes of finite sets are parameterized by natural numbers, our main
theorem in this context gives a statement about when probability distributions on natural
numbers are determined by their moments.
Proposition 6.24.
P Let2n−mM0 , M1 , . . . be a sequence of nonnegative numbers such that for all
Mn
m ≥ 0, we have n≥m m!(n−m)! < ∞.
Then there exists a sequence νm of nonnegative real numbers such that
∞ n−1
!
X Y
(m − i) νm = Mn
m=0 i=0
n−m M
for all n ≥ 0 if and only if if and only if n≥m (−1)
P n
m!(n−m)!
≥ 0 for all m ≥ 0.
n−m
If such νm exists, then we have νm = n≥m (−1) Mn
P
m!(n−m)!
for all m ≥ 0.

PThis
∞ Qis a version of the classical moment problem. Note that the factorial
n−1 P∞moments
n
m=0 i=0 (m − i) νm carry the same information as the power moments m=0 m νm
since we can write either in terms of the other using the Stirling numbers.
Proof. Let C be the opposite of the category of finite set. First note that every object of this
category is level-C where C is generated by the one-element set, because joins correspond to
unions of sets and every set is the union of one-element sets.
Since the isomorphism classes of finite sets are parameterized by natural numbers, a mea-
sure on C is the same thing as a measure on N, i.e. a sequence νm of nonnegative real
numbers. We write [n] for the object corresponding to a set with n elements.
The epimorphisms from [m] to [n] are the injections from an n-element
Qn−1 Qn−1 set to an m-
element set. The number of these is i=0 (m − i), so Epi([m], [n])  = i=0 (m − i). Thus the
[n]-moment of the measure νm is given by ∞
P Qn−1
m=0 i=0 (m − i) νm .
60
For an epimorphism π : [n] → [m], the interval [[n], [m]] is the lattice of subsets of an
n-element set containing a fixed m-element set, i.e. the lattice of subsets of an n − m-
element set, which is the product of n − m copies of a two-element lattice. This implies
n
µ([n], [m]) = (−1)n−m , and µ̂([n], [m]) = (−1)n−m m , as well as Z(π) = 2n−m .
Thus ∞ n
(−1)n−m m

X X (−1)n−m Mn
vC,m = Mn = .
n=0
n! n≥m
m!(n − m)!
Since [[n], [m]] is a product of copies of the two-element lattice, we can use a = 1 in the
definition of “behaved.” We have

X X |µ([n], [m])| X X |(−1)m−n | n−m
Z(π)Mn = 2 Mn
| Aut([n])| n=0
n!
[n]∈C π∈Epi([n],[m]) π∈Epi([n],[m])
∞ ∞
X 2n−m X 2n−m Mn
= Mn = m! < ∞,
n=m
(n − m)! n=m
m!(n − m)!
so our moments are behaved. Then, the “if” of the first statement follows from Theorem 4.15.
The final statement of the proposition follows from Theorem 4.2. The “only if” of the first
statement follows from the final statement. 
If G is a contravariant functor from finite sets to finite sets, we can apply Lemma 6.21 to
obtain a new category to work in, starting from the opposite category of finite sets.
For example, if G(S) is given by the set of relations on S which are reflexive and symmetric
but not necessarily transitive, then we obtain the opposite of the category of finite simple
graphs. We represent each graph by its set of vertices together with the relation of two
vertices either connected by an edge or identical, and morphisms of graphs G1 → G2 given
by maps from the vertices of G2 to the vertices of G1 that sends every pair of vertices
connected by an edge to either a pair of vertices connected by an edge or the same vertex.
(One could also use the opposite convention where the relation never includes a vertex and
itself.)
Epimorphisms in this category correspond to inclusions of induced subgraphs. We can
thus express MG /|Aut(G)| as the expected number of induced subgraphs isomorphic to G
in a random graph.
A similar notion of moments has been considered in graph theory, with work surveyed in
[LS10]. These works studied the probability that a subgraph induced on a set of vertices
sampled uniformly at random from an infinite graph (a graphon) is isomorphic to a fixed
one. Because these moments are associated to a single graphon rather than a probability
measure on the set of finite graphs, the existence and uniqueness statements obtained for
the moment problem in that context differ from those that can be obtained from Theorem
1.6
There are many models of random graphs, which one could hope to check by our method
are uniquely determined by their moments, and one could possibly even create new models
of random graphs from suitably chosen moments. Note, however, that any model of random
graphs with a fixed number of vertices is automatically determined by its moments, so the
main interest is in models of random graphs with varying number of vertices (perhaps by
first generating the number of vertices using an existing distribution on the natural numbers
such as a Poisson or exponential distribution).
61
For now we give just one example:

Lemma 6.25. For λ a positive real number and α ∈ [0, 1], there exists a unique prob-
ability measure on the set of isomorphism classes of finite graphs such that the expected
number of induced subgraphs isomorphic to a given finite graph G is λ#V (G) α#E(G) (1 −
#V (G)
α)( 2 )−#E(G) /|Aut(G)|, which is an Erdős-Rényi random graph, with edge probability α,
on a number of vertices given by a Poisson distribution with mean λ.

Proof. Let C be the opposite of the category of finite simple graphs as described above.
First, we check that the given random graph has Gth moment (in this category) MG =
#V (G)
λ#V (G) α#E(G)(1−α)( 2 )−#E(G) . If we condition on the number n of vertices of the random
graph, then there are n(n − 1) · · · (n − #V (G) + 1) ways to give an injection from the vertices
#V (G)
of G to the vertices of our random graph, and each of these has α#E(G) (1 − α)( 2 )−#E(G)
probability of realizing G as an induced subgraph. The expectation of n(n − 1) · · · (n −
#V (G) + 1) for n Poisson with mean λ is λ#V (G) . Thus the expected number of inclusion
maps from G to our random graph, as an induced subgraph, (which is the Gth moment), is
#V (G)
λ#V (G) α#E(G)(1 − α)( 2 )−#E(G) . Two different realizations of G as an induced subgraph
of our random graph have the same image, i.e. correspond to the same induced subgraph,
if and only if they differ by an automorphism of G, and so the average mentioned in the
lemma statement is always | Aut(G)|−1 times the Gth moment.
For uniqueness, we apply Theorem 4.2. We have seen above that C is a diamond category,
and we can check that the whole category is generated, as a level C, by a graph on one vertex
and the path on two vertices, as each graph is the join of such graphs. As noted in the proof
of Lemma 6.21, the poset [F, G] is the same as the poset [V (F ), V (G)] in the opposite of
the category of finite sets, so µ(F, G) = (−1)#V (G)−#V (F ) . Let F be a fixed graph with m
vertices and e edges. We now check the hypothesis of Theorem 4.2, and have

X X |µ(F, G)MG |
G∈C i:F →G
|Aut(G)|
inclusion as an
induced subgraph
X X X |µ(F, G)MG |
=
n≥m G on n i:F →G
n!
labelled vertices inclusion as an
induced subgraph
n
X X X λn α#E(G) (1 − α)( 2 )−#E(G)
= .
n≥m G on n i:F →G
n!
labelled vertices inclusion as an
induced subgraph

There are n!/(n − m)! ways to map to vertices of F to n labelled vertices. For a graph on
those n vertices to include F as an induced subgraph according to one of these maps, the
edges between the m vertices corresponding to F are determined, but all other edges can
n
− m 
appear or not appear. Thus given one of these n!/(n − m)! maps, there are ( 2 ) e′( 2 ) graphs
G on n vertices with e′ + e edges such that the given map realized F as an induced subgraph.
62
Thus, the above expressions are equal to
X  n − m  λn αe′+e (1 − α)(n2 )−e′ −e
 
X n! 2 2

n≥m
(n − m)! ′ e′ n!
e
n X n − m  ′
 
m λ n m
=αe (1 − α)( 2 )−e αe (1 − α)( 2 )−( 2 )−e

X
2 2
(n − m)! e′ e′
n≥m
m X λn
=(1 − α)( 2 ) < ∞.
n≥m
(n − m)!

This verifies the hypothesis of Theorem 4.2, and thus we have a unique distribution of finite
graphs with the given moments. 

References
[Ale99] A. N. Alekseı̆chuk. Conditions for the uniqueness of the moment problem in the class of $q$-
distributions. Diskretnaya Matematika, 11(4):48–57, 1999.
[AS68] Abraham Adrian Albert and Reuben Sandler. An Introduction to Finite Projective Planes. Holt,
Rinehart and Winston, 1968.
[Awo06] Steve Awodey. Category Theory. Oxford Logic Guides. Oxford University Press, Oxford, 2006.
[BBH17] Nigel Boston, Michael R. Bush, and Farshid Hajir. Heuristics for $p$-class towers of imaginary
quadratic fields. Mathematische Annalen, 368(1-2):633–669, 2017.
[Bir48] Garrett Birkhoff. Lattice Theory, Revised Edition (2nd), volume 25 of AMS Colloquium Publica-
tions. American Mathematical Society, 1948.
[Bir67] Garrett Birkhoff. Lattice Theory. Colloquium Publications (American Mathematical Society) ;
Volume 25. American Mathematical Society, Providence, Rhode Island, third edition. edition,
1967.
[BKL+ 15] Manjul Bhargava, Daniel M. Kane, Hendrik W. Lenstra, Jr., Bjorn Poonen, and Eric Rains.
Modeling the distribution of ranks, Selmer groups, and Shafarevich-Tate groups of elliptic curves.
Cambridge Journal of Mathematics, 3(3):275–321, 2015.
[Bor16] Iurie Boreico. Statistics of Random Integral Matrices. PhD thesis, Stanford University, United
States – California, 2016.
[BV88] Winfried Bruns and Udo Vetter. Determinantal Rings, volume 1327 of Lecture Notes in Mathe-
matics. Springer-Verlag, Berlin, 1988.
[BW17] Nigel Boston and Melanie Matchett Wood. Non-abelian Cohen–Lenstra heuristics over function
fields. Compositio Mathematica, 153(7):1372–1390, July 2017.
[CKL+ 15] Julien Clancy, Nathan Kaplan, Timothy Leake, Sam Payne, and Melanie Matchett Wood. On
a Cohen–Lenstra heuristic for Jacobians of random graphs. Journal of Algebraic Combinatorics,
pages 1–23, May 2015.
[CL84] Henri Cohen and Hendrik W. Lenstra, Jr. Heuristics on class groups of number fields. In Number
Theory, Noordwijkerhout 1983 (Noordwijkerhout, 1983), volume 1068 of Lecture Notes in Math.,
pages 33–62. Springer, Berlin, 1984.
[CM90] Henri Cohen and Jacques Martinet. étude heuristique des groupes de classes des corps de nombres.
Journal für die Reine und Angewandte Mathematik, 404:39–76, 1990.
[Coo00] C. Cooper. On the distribution of rank of a random matrix over a finite field. Random Structures
& Algorithms, 17(3-4):197–212, October 2000.
[EVW16] Jordan S. Ellenberg, Akshay Venkatesh, and Craig Westerland. Homological stability for Hurwitz
spaces and the Cohen-Lenstra conjecture over function fields. Annals of Mathematics. Second
Series, 183(3):729–786, 2016.
[FJ08] Michael D. Fried and Moshe Jarden. Field Arithmetic, volume 11 of Ergebnisse Der Mathe-
matik Und Ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results
63
in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics].
Springer-Verlag, Berlin, third edition, 2008.
[FK06a] Étienne Fouvry and Jürgen Klüners. Cohen–Lenstra Heuristics of Quadratic Number Fields. In
Florian Hess, Sebastian Pauli, and Michael Pohst, editors, Algorithmic Number Theory, number
4076 in Lecture Notes in Computer Science, pages 40–55. Springer Berlin Heidelberg, January
2006.
[FK06b] Étienne Fouvry and Jürgen Klüners. On the 4-rank of class groups of quadratic number fields.
Inventiones mathematicae, 167(3):455–513, November 2006.
[Hea94] D. R. Heath-Brown. The size of Selmer groups for the congruent number problem. II. Inventiones
Mathematicae, 118(2):331–370, 1994.
[KS06] Masaki Kashiwara and Pierre Schapira. Categories and Sheaves. Springer Berlin Heidelberg, 2006.
[Lee22] Jungin Lee. Mixed moments and the joint distribution of random groups, October 2022.
[Liu22] Yuan Liu. Non-abelian Cohen–Lenstra Heuristics in the presence of roots of unity.
(arXiv:2202.09471), February 2022.
[LS10] László Lovász and Balázs Szegedy. The graph theoretic moment problem.
https://arxiv.org/abs/1010.5159, 2010.
[LST20] Michael Lipnowski, Will Sawin, and Jacob Tsimerman. Cohen-Lenstra heuristics and bilinear
pairings in the presence of roots of unity. arXiv:2007.12533 [math], July 2020.
[LW20] Yuan Liu and Melanie Matchett Wood. The free group on n generators modulo n + u random
relations as n goes to infinity. Journal für die reine und angewandte Mathematik, 2020(762):123–
166, May 2020.
[LWZ19] Yuan Liu, Melanie Matchett Wood, and David Zureick-Brown. A predicted distribution for Galois
groups of maximal unramified extensions. arXiv:1907.05002 [math], July 2019.
[Més20] András Mészáros. The distribution of sandpile groups of random regular graphs. Transactions of
the American Mathematical Society, 373(9):6529–6594, September 2020.
[Neu67] Hanna Neumann. Varieties of Groups. Springer Berlin Heidelberg, 1967.
[NVP22] Hoi H. Nguyen and Roger Van Peski. Universality for cokernels of random matrix products,
September 2022.
[NW21] Hoi H. Nguyen and Melanie Matchett Wood. Random integral matrices: Universality of surjec-
tivity and the cokernel. Inventiones mathematicae, October 2021.
[Rot64] Gian Carlo Rota. On the foundations of combinatorial theory I. Theory of Möbius Functions.
Zeitschrift für Wahrscheinlichkeitstheorie und Verwandte Gebiete, 2(4):340–368, January 1964.
[RZ00] Luis Ribes and Pavel Zalesskii. Profinite Groups. Springer Berlin Heidelberg, 2000.
[Saw20] Will Sawin. Identifying measures on non-abelian groups and modules by their moments via re-
duction to a local problem. arXiv:2006.04934 [math], June 2020.
[Sch17] Konrad Schmüdgen. The Moment Problem, volume 277 of Graduate Texts in Mathematics.
Springer, Cham, 2017.
[Smi76] Jonathan D. H. Smith. Mal’cev Varieties. Lecture Notes in Mathematics, Vol. 554. Springer-
Verlag, Berlin-New York, 1976.
[Sta12] Richard P. Stanley. Enumerative Combinatorics. Volume 1, volume 49 of Cambridge Studies in
Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 2012.
[SW22a] Will Sawin and Melanie Matchett Wood. Conjectures for distributions of class groups of G-
extensions of fields with roots of unity. forthcoming, 2022.
[SW22b] Will Sawin and Melanie Matchett Wood. Finite quotients of 3-manifold groups.
https://arxiv.org/abs/2203.01140, 2022.
[Woo17] Melanie Matchett Wood. The distribution of sandpile groups of random graphs. Journal of the
American Mathematical Society, 30(4):915–958, 2017.
[Woo18] Melanie Matchett Wood. Cohen-Lenstra heuristics and local conditions. Research in Number
Theory, 4(4):41, September 2018.
[Woo19] Melanie Matchett Wood. Random integral matrices and the Cohen-Lenstra heuristics. American
Journal of Mathematics, 141(2):383–398, 2019.
[WW21] Weitong Wang and Melanie Matchett Wood. Moments and interpretations of the Co-
hen–Lenstra–Martinet heuristics. Commentarii Mathematici Helvetici, 96(2):339–387, June 2021.
64
Department of Mathematics, Columbia University, 2990 Broadway, New York, NY 10027
USA
Email address: sawin@math.columbia.edu

Department of Mathematics, Harvard University, Science Center Room 325, 1 Oxford


Street, Cambridge, MA 02138 USA
Email address: mmwood@math.harvard.edu

65

You might also like