On Malle’s conjecture for nilpotent groups

By Peter Koymans and Carlo Pagano

Abstract

We develop an abstract framework for studying the strong form of Malle’s conjecture [J. Number Theory 92 (2002), pp. 315–329; Experiment. Math. 13 (2004), pp. 129–135] for nilpotent groups in their regular representation. This framework is then used to prove the strong form of Malle’s conjecture for any nilpotent group such that all elements of order are central, where is the smallest prime divisor of .

We also give an upper bound for any nilpotent group tight up to logarithmic factors, and tight up to a constant factor in case all elements of order pairwise commute. Finally, we give a new heuristical argument supporting Malle’s conjecture in the case of nilpotent groups in their regular representation.

1. Introduction

1.1. Malle’s conjectures

Let be a non-trivial, finite group and let be a number field. In this paper we are interested in the counting function

where denotes the relative discriminant of . As part of a broader conjecture, Malle Reference 37Reference 38 conjectured the following.

Conjecture 1 (Strong form of Malle’s conjecture.).

Let be a non-trivial, finite group. Then there exist constants , and such that

Malle Reference 37 proposed the following recipe for the constant

with being the smallest prime divisor of . In follow-up work Malle Reference 38 also included a proposal for

where denotes the set of conjugacy classes of and where two conjugacy classes and are equivalent if there exists such that . Here the action is given by with the cyclotomic character.

Malle does not give a recipe to compute the constant . In some cases the constant is a product of local densities. This is known for , and (and conjectured for ) and is commonly referred to as the Malle–Bhargava principle.

In case is allowed to be an arbitrary finite group, counterexamples are known to the strong form of Malle’s conjecture, see the work of Klüners Reference 27. In the counterexample of Klüners, Malle gives the wrong value of . However, the value of is still correct in Klüners’ counterexample. It is widely believed, but unproven, that Malle’s prediction for is correct for all non-trivial, finite groups (for this reason we will simply write from now on). Even such a result seems to be out of reach currently, since it implies a positive answer to the inverse Galois problem. Klüners’ counterexample led Türkelli Reference 51 to propose a corrected .

The strong form of Malle’s conjecture (including the more general situation where is not necessarily considered in its regular representation) has been verified in a limited amount of cases, see Reference 15Reference 16 for , Reference 53 for abelian, Reference 10 for , Reference 28 for generalized quaternion groups, Reference 5 for , Reference 6 for , Reference 9 for , Reference 22 for nonic Heisenberg extensions, Reference 39Reference 52 for direct products with and abelian. There is also the recent work Reference 4, which counts when the extensions are ordered by Artin conductor instead of discriminant.

It is worth mentioning that the weak form of Malle’s conjecture, which asserts that

is much better understood. There are no known counterexamples to the weak form, even when is allowed to be an arbitrary finite group. The weak form is known for nilpotent groups by the work of Klüners–Malle Reference 30. Alberts Reference 1Reference 2 and Alberts–O’Dorney Reference 3 made further progress in the solvable case.

1.2. Main results

We prove the strong form of Malle’s conjecture for a large family of nilpotent groups.

Theorem 1.1.

Let be a number field and let be a non-trivial, finite, nilpotent group. Let be the smallest prime divisor of and assume that all elements of order are central. Then there exists a constant such that

where is the Malle constant, which equals the number of elements of order divided by in this case.

In the process, we give a parametrization of -extensions that may prove fruitful for future investigations. We will say more about this parametrization in the next subsection.

Klüners Reference 29 has previously proven an upper bound of the correct order of magnitude in the situation of Theorem 1.1, namely

This follows from Reference 29, Theorem 1.4 and Reference 29, Lemma 5.13.

Let us construct some -groups satisfying the hypotheses of Theorem 1.1. Take for example any finite, abelian -group . Then there is an action of on by inversion, which gives an action of on by projecting first to . If one takes , then fulfills all the conditions of Theorem 1.1. Note that such can have arbitrarily large nilpotency class.

We remark that our techniques do not give an explicit handle on the constant guaranteed by Theorem 1.1. Even in the case where is cyclic of prime order it is a rather non-trivial task to provide an explicit value of the constant , see Reference 11. The following classical result of Wright Reference 53 is an immediate corollary of Theorem 1.1.

Corollary 1.2.

Let be a number field and let be a non-trivial, finite, abelian group. Let be the smallest prime divisor of . Then there exists a constant such that

where is the number of elements of order divided by .

Our proof is substantially shorter than Wright’s proof Reference 53 and makes limited use of class field theory. Our parametrization of -extensions immediately implies Theorem 1.3.

Theorem 1.3.

Let be a number field and let be a non-trivial, finite, nilpotent group. Let be the smallest prime divisor of . Then

where is the number of elements of order in divided by .

Theorem 1.3 improves, in case of nilpotent groups in the regular representation, on recent work of Klüners–Wang Reference 31 and Klüners Reference 29 (see Proposition 5.8 and the text preceding it). The result in Theorem 1.3 is sharp up to logarithmic factors. In general we have the inequality . In fact, the upper bound in Theorem 1.3 matches Malle’s prediction precisely when we are in the situation of Theorem 1.1.

It should be possible to use our techniques to prove a more general version of Theorem 1.3 valid for arbitrary representations of nilpotent groups, but we shall not pursue this further here. Theorem 1.4 shows that we can achieve a sharp upper bound, up to a constant factor, provided that the elements of order commute with each other.

Theorem 1.4.

Let be a number field and let be a non-trivial, finite, nilpotent group. Let be the smallest prime divisor of . Suppose that all elements of order commute with each other. Then

The corresponding lower bound for follows from Reference 1, Corollary 1.7, where we take (the assumptions imply that is indeed abelian and normal). Earlier work of Klüners–Malle Reference 30 provides a slightly weaker lower bound correct up to logarithmic factors.

To give examples of groups where Theorem 1.4 applies (but Theorem 1.1 does not), consider an vector space of dimension and pick an ordered basis . Let act on by cycling the ordered basis and extending linearly. Then we can take , where acts on by first projecting to . We remark that is isomorphic to the small group from the GAP database and that is isomorphic to the small group . Note that can have arbitrarily large nilpotency class.

1.3. Method of proof

The main innovation of this paper is a new parametrization of -extensions, which is given in Section 2. The simplest case of the parametrization is that of multiquadratic fields over . The straightforward way to parametrize multiquadratic fields is the following. Let be a vector such that the are squarefree and linearly independent in . To each such vector , we associate the field . In this way we have created a map to multiquadratic fields of degree . This map has two convenient properties

the pre-images all have the same size (and are finite);

the map is surjective.

However, we will nevertheless argue that this does not provide a convenient way to count multiquadratic extensions. Indeed, the discriminant of is equal to up to factors of , where is the radical of an integer. Let us now consider the counting function

which is a good prototype for . To evaluate , we introduce the new variables

for every non-empty subset of . Then the variables are squarefree and pairwise coprime. From the variables we can reconstruct the using the formula

Under this change of variables, the sum becomes

which can be evaluated using classical techniques in analytic number theory. The crux of this idea is that the discriminant is simply

in the variables , while the discriminant is a more complicated function in terms of the variables . We remark that this change of variables also plays a role in the study of -Selmer ranks and -ranks, see for instance Reference 21Reference 24. Inspired by this, we parametrize multiquadratic fields by tuples of squarefree integers that are pairwise coprime. The parametrization map is then given by , where the are defined through equation Equation 1.2.

For arbitrary nilpotent extensions (but still over ) our parametrization keeps track of the image of the inertia subgroup of in our -extension for every prime . This involves the choice of an embedding . For arbitrary groups it is unfortunately not true that a homomorphism is determined by the restrictions .

Indeed, in general one needs to know the restriction of to for every embedding to make this conclusion. However, for nilpotent groups it suffices to fix one embedding , as a homomorphism is completely determined by the restrictions in this case, see Proposition 2.5. Our parametrization uses this property of nilpotent extensions in an essential way.

It is precisely for this reason that the source of our parametrization map is still tuples of squarefree integers that are pairwise coprime and are indexed by the elements in . We remark that this matches the multiquadratic case, since there is a bijection between non-empty subsets of and . This allows us to give a similar formula as above for the discriminant, see Proposition 2.10.

The fact that our parametrization is in terms of tuples of squarefree integers is also most convenient for analytic purposes. For example, it allows us to prove Theorem 1.3 using essentially the trivial bound, namely by ignoring all the embedding problems. This demonstrates the power of the parametrization, since it improves on the existing results in the literature by bounding trivially! In some sense one may view Theorem 1.3 as the correct trivial bound for Malle’s conjecture.

It also enables us to write down the various embedding problems explicitly, which we do in Section 3. We use this description to give a heuristic argument why Malle’s conjecture is correct for nilpotent extensions. We believe that this may also be of value to possible future work on this topic. Indeed, it makes clear exactly what the difficulties are in proving Malle’s conjecture for arbitrary nilpotent extensions. Roughly speaking, at every central extension we pinpoint a certain set of Frobenius elements that need to be equidistributed in a subgroup that we correspondingly identify. This explains how the constant arises from the construction of a nilpotent group as successive central extensions, see Section 3.

The parametrization also plays an essential role in the proof of Theorem 1.1 and Theorem 1.4. Using the parametrization and some analysis, we reduce these theorems to counting multicyclic extensions with some extra local conditions. This idea also provides a new and short proof of Wright’s theorem Reference 53.

1.4. Related results

There is also the related problem of counting the number of degree extensions with bounded discriminant, which was first treated by Schmidt Reference 43. His upper bound was drastically improved by Ellenberg–Venkatesh Reference 18, Couveignes Reference 14 and Lemke Oliver–Thorne Reference 36.

Malle’s conjecture has strong ties with the Cohen–Lenstra conjectures Reference 12. There is the classical work of Davenport–Heilbronn Reference 16 on -torsion of class groups of quadratic fields, which was later extended by Reference 8 and Reference 49 in the form of a secondary main term. Davenport and Heilbronn obtain their results by counting certain -extensions.

Fouvry and Klüners Reference 20Reference 21 dealt with the -rank of quadratic fields building on earlier work of Gerth Reference 23, and Heath-Brown Reference 24 on -Selmer groups. There is also a rich literature on upper bounds for -torsion elements in class groups of which we mention Reference 33Reference 34 for multiquadratic extensions, Reference 7 for -extensions and Reference 45Reference 46 for monogenic -extensions. Furthermore, the average size of the -torsion of -extensions has been determined in Reference 25 conditional on a tail estimate. Over function fields Malle’s conjecture and the Cohen–Lenstra conjectures are better understood due to the results in Reference 17Reference 19.

Recently, Smith Reference 47Reference 48 dealt with the -part of class groups of quadratic fields, which was extended by the authors to the -part of class groups of degree cyclic fields Reference 32. His techniques can be adapted to give a lower bound for the number of -extensions of of the correct order of magnitude: this is another instance that highlights the clear ties between Malle’s conjecture and the Cohen–Lenstra conjectures. It seems plausible that this can be extended to an asymptotic once one extends the results on ray class groups of Pagano–Sofos Reference 42.

It is natural to wonder whether it is possible to combine Smith’s techniques with our parametrization to tackle Malle’s conjecture for all finite, nilpotent groups. It is the authors’ belief that this should be possible for a wide class of nilpotent groups, and we hope to return to this topic. However, one convenient property of is that the embedding problem for odd ramified primes often comes down to the condition that has residue field degree . In terms of our parametrization this is very convenient, since it gives good inductive control over the relevant Frobenius elements.

This fails badly for arbitrary nilpotent -extensions, where one may face “loops” in the following sense. We might simultaneously want to prove equidistribution of in a quotient of depending on , while we also need equidistribution of in a quotient of depending on . The authors currently do not know how to overcome this difficulty.

1.5. Organization of the paper

The paper is divided as follows. The core of the paper is Section 2, where we provide a new parametrization of -extensions. In Section 3 we give a new heuristic in support of Malle’s conjecture for nilpotent groups in their regular representation. Our main theorems are proven in Section 5 with Section 4 providing some analytic tools.

2. Arbitrary nilpotent groups and number fields

In our first subsection we restrict ourselves to the case that is a finite -group, but we work with an arbitrary number field . This is then extended to general nilpotent in the second subsection.

2.1. Finite -groups

We use the abbreviation throughout the paper. For a set and a prime number , we write for the free vector space on the set . Let us fix a separable closure of once and for all. All our number fields are implicitly taken inside this fixed separable closure. Furthermore, we write for the absolute Galois group of a number field . Similarly, for each prime we fix a separable closure of (which allows us to define for any extension of inside ) together with an embedding

This yields an inclusion

Let be a number field, picked inside . Denote by the set of all places of . For each finite place in , lying above a rational prime , the restriction of the map provides us with an inclusion

Denote by the residue field of at . Write

for the inertia subgroup.

Let now be any prime number. In what follows the group will always be implicitly interpreted as a Galois module with trivial action, whenever the notation suggests an implicit action of a group on . We denote by the image, via inflation, of in .

For a subset , containing all the archimedean places of , we consider the map

In case consists exactly of the archimedean places, we will denote the resulting map simply by . We start by recalling the following classical fact.

Proposition 2.1.

The abelian groups and are finite.

Proof.

By class field theory we have a canonical identification

where is the modulus consisting of all archimedean places. Therefore we have that

where the first factor can be dropped when is odd. Therefore the finiteness of follows from the finiteness of . The finiteness of is established in Reference 44, Theorem 5, eq. (14) and (16). In the notation of Shafarevich, we have

and the dual of his homomorphism is our . Combining Reference 44, Theorem 5, eq. (14) and (16) yields

which is a finite number defined on Reference 44, p. 129.

The following important fact falls as an immediate consequence of Proposition 2.1.

Proposition 2.2.

There exists a finite set of places , containing all archimedean places, such that is surjective.

Proof.

Write for the finite subset of archimedean places of . Thanks to Proposition 2.1 we can find a finite set of places such that the natural map

is surjective. This implies that for any vector

we can find and a global character such that

Therefore, since is entirely supported in , we conclude that

Hence we have shown that the map is surjective with this choice of , which is precisely the desired conclusion.

Of course if a set as in Proposition 2.2 works, then any larger set works as well. We fix once and for all a finite set as in Proposition 2.2, making sure that it also contains all places above . We denote by the subset of with

For , it follows from local class field theory that equation Equation 2.1 is equivalent to the condition

Write for the compositum of all finite Galois extensions of with a power of .

Proposition 2.3.

Let be a finite place coprime to such that . Then is unramified in every finite extension inside .

Proof.

It suffices to prove the proposition locally at . Take a positive integer and let be the unique unramified extension of of degree equal to with residue field denoted by . Then if we have a cyclic totally ramified degree extension of it follows from local class field theory and the fact that is coprime to

Now, if is a power of itself, we conclude that is already congruent to modulo contrary to our assumption that .

We remark that Proposition 2.3 certainly applies to any place . Let . Thanks to Proposition 2.2, there exists a character

such that has non-trivial coordinate precisely at and at no other places in . Fix once and for all such a choice of for each . By construction

is a linearly independent set. Furthermore by Proposition 2.1 we obtain that the subspace

has finite index. Additionally, there exist a positive integer and a basis

such that is a basis of . Fix once and for all such a choice of . We denote by

this fixed choice of a basis. Put

For each finite place we denote by

We have the following basic fact.

Proposition 2.4.

The group is pro-cyclic for each finite place coprime to .

Proof.

Let be a non-archimedean local field of characteristic and write for its residue characteristic. Let be a positive integer coprime to . Then every finite totally ramified extension of of degree equal to can be obtained as for a uniformizer of . For an elementary proof of this well-known fact see Reference 32, Proposition A.5. Applying this repeatedly to all finite unramified extensions of , we conclude that

is a pro-cyclic group, where is the inertia subgroup and is the wild inertia subgroup. Since is coprime to , we conclude that is a quotient of the pro-cyclic group , which gives in particular the desired conclusion.

Remark 1.

Since the group is a pro-cyclic pro- group, it is either isomorphic to a finite group of order a power of or isomorphic to . In case a primitive -th root of unity is in , then we claim that is infinite. Indeed, we have the infinitely ramified subextension

of (which is contained in exactly because is in ) given by any in with . Therefore we conclude that if possesses a non-trivial -th root of unity, then

for every finite place coprime to . Instead if is not in , we observe that Proposition 2.3 shows that the group is trivial in case is a finite place of that is coprime to and satisfies .

We fix once and for all a topological generator of for all in the following manner. Observe that for any topological generator of , since the character ramifies at . Hence we can always pick a generator with the normalization . We make such a choice of once and for all. In case , then the group is trivial by Proposition 2.3, and we declare .

Now it follows by construction that

for each and , where denotes the Kronecker delta function. Therefore we can complete the set

to a minimal set of generators

which is dual to the basis , i.e.

We denote by

Then we have the following.

Proposition 2.5.

The set is a set of topological generators of .

Proof.

By construction topologically generates , where is by definition the compositum of all degree cyclic extensions of . But since is a pro--group, we see that is also the quotient of by its Frattini subgroup. This implies the proposition, since a subset topologically generates if and only if it topologically generates modulo the Frattini subgroup.

Let be a finite Galois extension inside with Galois group . Take a -cocycle

with the requirement that . By the same argument as before, every class in can be represented by such a -cocycle . Consider the group

where the group law is

Our assumption on ensures that is the trivial element of .

Proposition 2.6.

Let be a prime number. Let be a number field and let be an extension with a finite -group. Suppose that is non-trivial in .

(a) The natural projection map can be lifted to a surjective homomorphism

if and only if is trivial in for each place that ramifies in . Moreover, if is a lift, then the -coordinate of is a continuous -cochain with

Conversely, given any such continuous -cochain with , the assignment

is an epimorphism lifting the canonical projection to an epimorphism . The two assignments are mutual inverses.

(b) In case one has a lift as in part (a), then there is a unique one satisfying

Proof.

We start with part (a). We claim that a map is the first coordinate of a homomorphism

if and only if

Indeed, since is a homomorphism, we obtain

which is equivalent to

as claimed.

Now suppose that there exists with . We claim that is surjective. Let us first show that all characters must come from . If not, then the kernel of such a hypothetical character provides a splitting of , which implies that is trivial contrary to our assumptions. Hence, since is surjective, the image of generates modulo the Frattini of , and therefore equals .

Furthermore, we see that the lifting exists if and only if the inflation of to is trivial if and only if is trivial in for every place of . Thanks to Reference 32, Proposition 4.4, the vanishing at the finite places unramified in is already guaranteed: notice that in Reference 32, Section 4 the number is assumed to be an odd prime but Proposition 4.4 of Reference 32 also holds for with an identical proof. Let us now show is locally trivial at any unramified, infinite place of . If is an archimedean complex place, then this is clear. If is instead an archimedean real place and the extension is unramified at , then this means that for each place of above and thus the embedding problem is locally trivial at . This ends the proof of part (a).

We now prove part (b). The uniqueness follows at once from part (a) combined with the fact that is a system of topological generators for . Indeed, an epimorphism as in part (a) is entirely determined by its values on a set of topological generators. We next show the existence: here we will take advantage of the fact that is a minimal set of topological generators. Take a map satisfying

The resulting epimorphism corresponds to a finite extension. As such we conclude that for all but finitely many . Therefore the sum

is a well-defined element of . Hence, since and are dual to each other, we obtain that

vanishes at for all and at for all . This ends the proof of part (b).

We denote the unique -cochain as in part (b) of Proposition 2.6 by . In case is trivial as a -cocycle, then we choose . With this choice, we see that satisfies part (b) of Proposition 2.6, since the trivial character is the unique cyclic degree character vanishing at all .

Denote by the set , where is the set of squarefree integral ideals in entirely supported in . To an element we attach the character

Two pairs are said to be coprime in case are coprime ideals and there does not exist a such that . Formulated differently, the two pairs are coprime exactly when

Let be any finite set. We denote by the subset of consisting of vectors possessing pairwise coprime coordinates. We conclude this subsection by giving a bijection

which sends a vector to

where is the projection map on the -th coordinate. Let us prove that this map is indeed a bijection.

Proposition 2.7.

Let be a finite set. Then the map is a bijection.

Proof.

Assume without loss of generality that . To a vector in we attach a point

in as follows. For each we let divide the entry if and only if

Likewise for each we put if and only if

By construction is in . Using that is a system of topological generators, we deduce that

since the equality holds by construction when evaluated in an element of . Conversely let and let . There exists at most one such that . Suppose that such a exists. Then

which implies that

Hence and are mutual inverses, which finishes the proof of the proposition.

We now have the necessary tools to parametrize nilpotent extensions. We carry this out in the next subsection.

2.2. The parametrization in general

Let . A sequence of pairs

is called an admissible sequence if it satisfies the following inductive rules:

is the trivial group by convention. Furthermore, is an -group and is a -cocycle with for each ;

we have

for all ;

is the zero map if and only if the class of in is trivial.

For the remainder of this section we fix an admissible sequence . Set

The aim of this section is to construct a surjective map

which restricts to a bijection between

and . Furthermore, we explain how to read the ramification data on the right hand side from the left hand side of this parametrization.

We start by defining a map

as follows. Let be an element of . If is the trivial character, we declare . So we assume from now on that is non-trivial. Equivalently,

Hence is now a -cocycle on . If is not trivial in , then we declare . Now assume that is zero in ; we distinguish two cases. If is already a trivial -cocycle on , we have that and

if and only if and are linearly independent. In case and are linearly dependent, we set and otherwise we proceed. If instead is a non-trivial -cocycle on , we always have that

by Proposition 2.6.

Now we continue in this fashion inductively. At step we have either already assigned to • or we have obtained an epimorphism . Then we get a -cocycle , which gives a class in . If this class is non-trivial in , we send to •.

In case is trivial in , we distinguish two cases. If is already trivial in , we have that . Then

if and only if is linearly independent from the characters satisfying

Indeed, observe that in this case we have , therefore, since we have surjectivity if and only if we have surjectivity modulo the Frattini, we need to have that is linearly independent from the characters coming from . Thus our claim comes down to the claim that such characters are precisely spanned by the set

This is justified by the following observation. Let be a finite -group and let be a -cocycle. Then is trivial in if and only if the dimension of modulo its Frattini subgroup is one larger than that of modulo its Frattini subgroup. To see the non-trivial direction observe that if one takes a character that does not come from , then its kernel gives a splitting of the sequence.

If is linearly dependent on these characters , we send to •. Otherwise we go to step .

Now suppose that is a non-trivial class of . Then we always obtain by means of Proposition 2.6 a new epimorphism

and we go to step . Continuing in this fashion we obtain either • or an element of

which is by definition . We put

We remark that is only defined in case is a Galois group. Fortunately, this small abuse of notation does not present any issues. Indeed, the epimorphism realizes the implicit identification between and the corresponding Galois group.

We additionally remark that in the construction of the map we have implicitly used that is set-theoretically defined to be with the identity element being , which is a consequence of our convention that -cocycles vanish on . Hence it makes sense to invoke the map .

Proposition 2.8.

The map

is a surjection, which restricts to a bijection between and surjective homomorphisms .

Proof.

This follows upon combining Proposition 2.6 and Proposition 2.7.

The parametrization allows us to read off very neatly the image of the elements from the tuples of squarefree ideals.

Proposition 2.9.

Let be an element of . Let . If for a (necessarily) unique then

If does not divide any of the elements of the vector , then

Proof.

For the case the conclusion follows immediately from the fact that the map in the proof of Proposition 2.7 is inverse to the map . Otherwise, we have that so that by definition. By construction of it follows that such do not divide any . Hence we are always in the second case of the proposition. Therefore the statement also holds for such .

We next read off the value of the discriminant under the bijection . For any continuous homomorphism of with values in some finite group, we denote by the relative discriminant (which is an ideal of ) of the corresponding extension. For a non-zero integral ideal in we write for the largest ideal dividing and entirely supported outside of .

Proposition 2.10.

Let be an element of

Then

Proof.

We show that the -adic valuation matches for any prime of . This is certainly true for the places in , but also for the places outside by Proposition 2.3. Now take a place in . Since is coprime to we know that

Thanks to Proposition 2.9 we deduce that is trivial in case does not divide any and equals in case divides . This is precisely the desired conclusion.

Our final goal for this subsection is to generalize Propositions 2.8, 2.9 and 2.10 to arbitrary finite, nilpotent groups. Recall that a finite group is nilpotent if and only if it decomposes as a direct product of its Sylow subgroups. Let be a positive integer. Let be distinct prime numbers. For each fix an admissible sequence

and write for the resulting -group. We put

To parametrize -extensions, we reduce to -groups by means of Proposition 2.11.

Proposition 2.11.

We have an identification

through the natural map.

Proof.

Let be any group and be any group homomorphism. We have to show that if is surjective for each , then is surjective. Observe that if each is surjective, then is divisible by for each . Since these values are coprime, we find out that is divisible by , which means precisely that is surjective.

Thanks to Proposition 2.11 we can now bundle together the various maps into one map

by simply taking the product map. Put

For every , let be a basis for the space of characters only ramified at . Define to be , where is the set of squarefree ideals supported outside . Let

be the set of tuples satisfying the following properties

writing for the natural projection map , we have that the are pairwise coprime;

the are pairwise coprime;

if divides and is a prime dividing the order of , then

For an element

and for we define

In this way we have created a very convenient bijection between

We need one additional piece of notation, namely we define

which we shall often implicitly view as a subset of . Proposition 2.12 generalizes Proposition 2.9.

Proposition 2.12.

Let be an element of

Take some . Let be the subset of such that if and only if there exists a (necessarily) unique with . Then we have

In particular if does not divide any of the elements , i.e. , then

Proof.

This follows at once from Proposition 2.8, applied to each -factor.

Proposition 2.13 generalizes Proposition 2.10. In the new coordinates we have a rather simple formula for the discriminant.

Proposition 2.13.

Notations as above. Let be an element of . Then

Proof.

This follows from Proposition 2.12 with exactly the same argument as used to establish Proposition 2.10 as a consequence of Proposition 2.9.

3. Local conditions and conjugacy classes

3.1. Some group theory

Let be a prime number and let be a finite -group given by an admissible sequence with . Our first goal is to study the formation of a conjugacy class in , through the various groups , with . For we denote by its conjugacy class. For each we write

for the natural projection map.

For now we take any finite -group , a -cocycle representing a class in , with , and an element . Denote the centralizer of by . Then lifting elements of to and taking the commutator with any lift of induces a homomorphism

which does not depend on the choice of lifts. Put

which is by definition a subgroup of . Let

be the natural projection map. Proposition 3.1 describes the relationship between and .

Proposition 3.1.

Let be as above this proposition. Then

The index equals if and only if the elements in are pairwise non-conjugate in . The index equals if and only if the elements of sit inside a unique conjugacy class in .

Proof.

Indeed, take a lift in . Observe that if we have

for some , then it follows that . From the left hand side we see that if the index is , then can take any possible value. If the index is instead equal to , then must be equal to .

We also have a similar proposition for the exponent of an element.

Proposition 3.2.

Let be as above. Then either consists entirely of elements with order equal to or it consists entirely of elements with order equal to .

Proof.

The class restricted to gives an element of . If the class is , then the sequence is split and we have that all the elements of have the same order as . If the class is non-zero, then the sequence has the shape

and hence all elements of have order times bigger than that of .

In case an as in Proposition 3.2 satisfies the second conclusion we say that is -stable. We now return to our previous setup. To we attach the following quantity

This quantity turns out to be the exponent of in the size of the conjugacy class . We call the ’s counted by the breaks for with respect to .

Proposition 3.3.

We have

Proof.

We have a filtration of subgroups

where we define

for every integer . We have that

Observe that

Therefore the desired conclusion follows at once from Proposition 3.1.

Proposition 3.4 provides the crucial link between group theoretic data and the local conditions imposed, through Proposition 2.6, on tuples in . We invoke the notation of Section 2. Let be a finite Galois extension inside . Let now and let be a -cocycle representing a class in , with . Let be a finite place coprime to . Recall that

is a pro-cyclic group, equipped with a canonical generator . We will fix once and for all the -image of a lift to of such an element. In this way we obtain an element in that we will denote also by : this slight abuse of notation will cause no confusion. As such we have naturally an element

Here denotes the normalizer of a subgroup in . For any non-trivial finite group , we denote by the smallest prime divisor of and by the subset of such that .

Proposition 3.4.

Let be a finite -group and let be a finite place coprime to . Assume that is an element of , which we shall also call . Then

Moreover, if is -stable, then

if and only if is trivial in .

Proof.

Let be any finite non-trivial group. Then we claim that implies . Indeed, conjugation induces a homomorphism

Since the latter group has size , and is the smallest prime divisor of , we have that is coprime to . Therefore the above homomorphism is actually trivial. This means exactly that as claimed. Hence we have already shown the first part of this proposition, namely that

Assume now that is also -stable. Observe that since lands in , we in particular conclude that ramifies in . It follows from Proposition 2.3 that

Also observe that the maximal pro- quotient of is isomorphic to

where acts by multiplication by on . Here is sent to , while a lift of is sent to . Since is -stable and lands in , we have that the lifting problem imposed by factors through

Since is modulo , the resulting quotient is simply

Recalling once more that is -stable, we see that the lifting problem is solvable if and only if the restriction of to

is in

This is precisely equivalent to asking

as was to be shown.

The final lemma of this subsection provides a simple way to compute the Malle constant for nilpotent .

Lemma 3.5.

Let be a non-trivial, finite group and let be a number field. Then we have

Proof.

Recall that there is a natural action of on

which sends a conjugacy class to with the cyclotomic character. By definition equals the number of orbits of this group action. But our action clearly factors through . We claim that the induced action of on is free, which implies the lemma.

So suppose that there exists such that

This implies that there exists a non-trivial such that and are conjugate, say

and hence by the argument given at the start of Proposition 3.4. We conclude that , which forces to be the identity as desired.

3.2. Interpretation of Malle’s constant

The goal of this subsection is to give a heuristic supporting Malle’s conjecture in the nilpotent case. Our heuristic is based on a combination of the parametrization given in Proposition 2.8 and the examination of the local conditions carried out in Subsection 3.1. To simplify the notation, we shall limit ourselves to the case where is an -group. We leave it to the reader to generalize the material below to arbitrary nilpotent groups .

Ignoring the finitely many bad places in , it follows from Proposition 2.8 and Proposition 2.10 that

should have order of magnitude

We now focus on the variables with . Upon combining Proposition 3.4 and Proposition 2.9, we see that the primes dividing impose a local condition only at the breaks for in the admissible sequence . The local conditions at the points that are not breaks are automatically satisfied in virtue of Proposition 3.4. Now pretend that the values

are jointly equidistributed at every break point . Then, in virtue of Proposition 3.4, we get the following sum

where the sum runs over all points in . Thanks to Proposition 3.3 the latter expression equals

where the sum still ranges over all points in . Standard analytic techniques, see Theorem 4.1, show that the sum in equation Equation 3.1 is asymptotic to

where is the Malle constant by Lemma 3.5. We remark that to turn this simple heuristic into an argument one also has to pay careful attention to the local conditions at the primes dividing variables outside of and to the local conditions at the primes in . These will affect the constant in the asymptotic. We finish this section by explaining the interplay between this heuristic and the proofs of our main theorems.

During the proof of Theorem 5.1 we simply ignore the local conditions, at the cost of losing track of the conjugation in : for this reason we get instead of .

Correspondingly for those for which and coincide we have that all the elements of are central, and consistently with Proposition 3.4 we have no local conditions coming from the variables in . In this case we are able to prove an asymptotic in Theorem 5.7.

Finally in the proof of Theorem 5.3, thanks to the fact that the elements of are pairwise commuting, we have to control the behavior of in the quotient for each dividing a variable in . This is very convenient, since the corresponding field is constructed out of the variables outside of , and those have a very large weight in the formula for the discriminant given in Proposition 2.10. As such, they can almost be treated as fixed, and the required joint equidistribution of Frobenius elements is provable by appealing to the Chebotarev density theorem. Hence in this case we can partially turn the above heuristic into a rigorous argument: since we control only the local conditions at the places dividing variables in , we naturally end up with an upper bound of the correct order of magnitude.

4. Analytic considerations

In this section we provide the analytic tools used to prove our main theorems. The material in this section is a generalization of the material in Montgomery–Vaughan Reference 41, Section 7.4 and is an application of the Selberg–Delange method. Let be a number field, let be an abelian extension of and let . Write for the group of non-zero fractional ideals of . For a squarefree ideal of we define for the number of prime divisors of and for the number of prime divisors of that are unramified in and satisfy . Given a complex number and a collection of prime ideals , we are interested in the sum

where is the Möbius function of . Write

for its Dirichlet series with coefficients . Then we have for

where is by definition . We assume that the Euler product

converges absolutely in the region for some constant . Then we approximate the Dirichlet series with

where

We recall that is by definition . Note that exists since the region is simply connected and does not vanish in this region. We choose our determination of the logarithm in such a way that it agrees with the real logarithm for real . It follows from equation Equation 4.2 that we have the fundamental relation

where is defined by an absolutely convergent Euler product in the region for some . In particular, if , then there exists some constant such that is a bounded non-zero holomorphic function on .

Theorem 4.1.

Let , , and as above. Then we have for all positive real numbers and all

where is a real constant depending only on , , , and .

Proof.

Since the proof is similar to Montgomery–Vaughan Reference 41, Theorem 7.17, we shall only sketch the necessary modifications. Set . An effective version of Perron’s formula, see Reference 41, Corollary 5.3, shows that

where we recall that is defined by equation Equation 4.1 and is a parameter at our disposal. We choose and estimate the error terms as in Reference 41. To do so, we need to have a good estimate for the sum

where we assume without loss of generality that is an integer greater than . The latter sum is estimated in Reference 41, Theorem 7.17 with Dirichlet’s hyperbola method.

We next move the path of integration. Note that has a branch point at if is not an integer. For this reason, we move the path of integration in such a way to avoid this branch point. Put , where is a small positive constant. Let be the polygonal path with vertices , let be the line segment from to , followed by a semicircle , and a line segment from to , and finally let be the polygonal path with vertices .

Let be the region enclosed by and the line segment from to . If is sufficiently small, then has no zeroes in the region by Reference 26, Theorem 5.10. Clearly, also has no zeroes in provided that is sufficiently small. Since the union of with the region is still simply connected, and are also well-defined in this region.

The main term comes from the integral over , and is extracted in exactly the same way as in the proof of Reference 41, Theorem 7.17. Finally, Montgomery–Vaughan estimate the integrals on the paths and by appealing to bounds for , see their Reference 41, Theorem 6.7. Hence we need to supply similar bounds for and . These can be derived by following the proof of Reference 41, Theorem 6.7, where we use Reference 26, Proposition 5.7, (2) as a replacement for Reference 41, Lemma 6.4.

Write for the Dirichlet convolution on . In Section 5 we will combine Theorem 4.1 with the following general lemma on convolutions.

Lemma 4.2.

Let be functions such that

for some real numbers , and . Then there is such that

Proof.

It follows from Dirichlet’s hyperbola method that

equals

The latter sum is at most . Since the first two sums in equation Equation 4.3 play a symmetric role, we shall only treat the first sum. The first sum equals

up to an error of size bounded by