1 Introduction
Cohen, Lenstra, and Martinet [Reference Cohen and LenstraCL84, Reference Cohen and MartinetCM90] have given heuristics on the distribution of class groups of number fields that lead to the following conjecture. For a number field k, a transitive permutation group G, and a prime
$p\nmid |G|$
, there conjecturally exists a constant
$c_{k,G,p}>0$
such that

where
$E_k(G,X)$
is the set of extensions
$K/k$
with Galois closure group G (that is, the Galois group of the normal closure; see Section 1.5) and absolute discriminant
$ {\mathrm {Disc}}(K)\leq X$
. The limiting average in (1.1) was previously known for only two cases of G and p: for
$G=S_2$
and
$p=3$
by Davenport and Heilbronn [Reference Davenport and HeilbronnDH71] when
$k={\mathbb {Q}}$
and Datskovsky and Wright [Reference Datskovsky and WrightDW88] for general k, and for
$G=S_3$
and
$p=2$
by Bhargava [Reference BhargavaBha05] when
$k={\mathbb {Q}}$
(see also the work of Bhargava, Shankar, and Wang [Reference Bhargava, Shankar and WangBSW15] for the case of general k). This conjecture is particularly notable in light of the fact that the best general upper bound on
$| {\mathrm {Cl}}(K)[p]|$
is
$ {\mathrm {Disc}}(K)^{1/2+\epsilon }$
, and there has been a lot of recent breakthrough work (e.g., [Reference Heath-Brown and PierceHP17, Reference Pierce, Turnage-Butterbaugh and WoodPTBW20]) even to show that the average in (1.1) is bounded by
$X^{1/2-\delta }$
for some
$\delta>0$
, for certain G, despite the fact that it is conjectured to have constant limit.
Our first main result is that we find the average and prove the conjecture in (1.1) when G is any transitive
$2$
-group containing a transposition and
$p=3$
.
Theorem 1.1. For any m, let
$G\subset S_{2^m}$
be a transitive permutation
$2$
-group containing a transposition. For any number field k, there exists a constant
$c_{k,G,3}$
(given explicitly in Section 6) such that

The simplest new example that Theorem 1.1 provides is that the average size of
$3$
-torsion in the class groups of
$D_4$
-quartic fields over k tends to a constant; when
$k=\mathbb {Q}$
, this constant works out to be around
$1.42$
. It applies also to three permutation groups G of degree
$8$
, to
$26$
groups of degree
$16$
, and to at least as many groups in degree
$2^m$
as in degree
$2^{m-1}$
.
The restriction that G has a transposition is necessary for our proof, and is a natural condition in this setting. In particular, it is expected (for example, according to Malle’s conjecture [Reference MalleMal04]) that the groups G arising with positive density as Galois groups when fields of degree n are ordered by discriminant are precisely those with a transposition in their degree n permutation representation. This is known for
$n=4$
where there are
$\sim c_1 X$
quartic
$D_4$
fields [Reference Cohen, Diaz, Diaz and OlivierCyDO02] and
$\sim c_2X$
quartic
$S_4$
fields [Reference BhargavaBha05] of absolute discriminant bounded by X for certain constants
$c_1,c_2> 0$
, and
$O_\epsilon (X^{1/2+\epsilon })$
quartic fields with any other Galois closure group. Thus, the
$2$
-groups to which Theorem 1.1 applies – or, indeed, the groups susceptible to our method as a whole – are those expected to have positive density within the set of all fields of a given degree, which you might think of as “generic” Galois groups.
When G is a transitive permutation
$2$
-group with a transposition and
$K\in E_k(G,X)$
, then K has a unique index two subfield F, and we can also ask about the relative class group of
$K/F$
. Indeed for
$G=D_4$
Cohen and Martinet [Reference Cohen and MartinetCM87] have said this part of
$ {\mathrm {Cl}}_K$
is of particular interest. Our next main result gives the averages of
$| {\mathrm {Cl}}_{K/F}[3]|=| {\mathrm {Cl}}_K[3]|/| {\mathrm {Cl}}_F[3]|$
.
Theorem 1.2. For any m, let
$G\subset S_{2^m}$
be a transitive permutation
$2$
-group containing a transposition. Let k be a number field, u be an integer, and
$E^u_k(G,X)$
be those extensions
$K\in E_k(G,X)$
such that
$ {\mathrm {rk}} \mathcal {O}_K^*- {\mathrm {rk}} \mathcal {O}_{F_K}^*=u$
, where
$F_K$
is the unique extension of k such that
$[K:F_K]=2$
. Then if
$E^u_k(G,\infty )$
is nonempty,

and this particular average value is predicted by the Cohen–Lenstra–Martinet heuristics.
Cohen, Lenstra, and Martinet actually gave a conjecture that predicts the average of
$f( {\mathrm {Cl}}_K)$
over G-extensions
$K/k$
for any function f not depending on certain “bad” Sylow subgroups of the class group. The only nontrivial cases of these conjectures previously proven were for averages of the two f of the form
$f( {\mathrm {Cl}}_K)=| {\mathrm {Cl}}_K[p]|$
mentioned above. Theorem 1.2 thus gives the first proof of the Cohen-Lenstra-Martinet conjecture for a function not of the form
$| {\mathrm {Cl}}_K[p]|$
(or trivial variations), and does this for infinitely many groups G. (When
$G=C_4$
and f is bounded and only depends on the class group of the quadratic subfield of the cyclic quartic, Bartel and Lenstra [Reference Bartel and LenstraBL20] showed that the average is computable in finite time to arbitrary precision, though the example they compute does not agree with the conjecture – see Section 7 for further discussion. There have also been some averages proven on the “bad” part of the class group including work of Fouvry and Klüners [Reference Fouvry and KlünersFK06] and Klys [Reference KlysKly20], and the recent groundbreaking work of Smith [Reference SmithSmi17] determining the entire distribution of
$ {\mathrm {Cl}}_K[2^\infty ]$
when
$G=S_2$
. Other work in this direction includes that of Gerth [Reference GerthGer84, Reference GerthGer87] and Koymans and Pagano [Reference Koymans and PaganoKP18].)
1.1 Methods
The condition that G contains a transposition implies that
$G=C_2\wr H$
for some H. So, for the extensions
$K/k$
we consider, K has a unique index two subfield F, and the overarching strategy of the paper is to average and then sum over each F. Accordingly, the work of Datskovsky and Wright on the average of
$| {\mathrm {Cl}}_K[3]|$
over quadratic extensions of a fixed F is a key input into our work. If we could use just the main term of the average in place of the average, the argument would be essentially straightforward. However, summing the error terms over F in a straightforward way leads to error that is larger than the main term.
The major obstacle we overcome in this paper is proving a sufficiently good tail bound so that we can sum over F. In particular, the challenge is to bound the contribution when the average is taken over a relatively small interval, given
$ {\mathrm {Disc}}(F)$
. Our first tool to do this is a heavily optimized upper bound on the average
$3$
-class number of quadratic extensions of a number field F, and the closely related number of
$S_3$
extensions of F, which we develop in Section 3. For very small or very large intervals, compared to
$ {\mathrm {Disc}}(F)$
, we use class field theory and the theory of the Shintani zeta function counting cubic orders due to Datskovsky and Wright [Reference Datskovsky and WrightDW88], respectively. While cubic extensions of a general number field are counted in [Reference Datskovsky and WrightDW88], in order to sum over different base fields, we must determine the explicit dependence of the count on the base field, which has not been done in previous work.
For intermediate intervals, however – summing
$| {\mathrm {Cl}}(K)[3]|$
for roughly
$ {\mathrm {Disc}} (K)\in [ {\mathrm {Disc}}(F)^{3.1}, {\mathrm {Disc}}(F)^6]$
– we require a new technique, which we call propagation of orders. In this range, the bound on cubic extensions given by the Shintani zeta function is too large. However, the Shintani zeta function we use counts nonmaximal orders as well as maximal orders. Given a maximal order, one can prove a lower bound on the number of nonmaximal orders inside it, up to a certain discriminant bound. Then, to count maximal orders of discriminant up to X, we count all orders of discriminant up to Z for some
$Z>X$
, and then account for the known overcounting. As Z gets larger the total count goes up but the known overcounting also improves, and it turns out in the intermediate range these trade-offs work out to give us an improved and sufficiently good bound.
With these optimized bounds on
$S_3$
extensions of a single F, especially in shorter intervals, we begin to take the sum over F, and we can bound the sum as required except in one problematic range, which we call the critical range. In this range,
$ {\mathrm {Disc}}(K)\approx {\mathrm {Disc}}(F)^3$
. If
$ {\mathrm {Disc}}(K)$
is a power smaller or larger than this, then the methods above are sufficient. In the critical range, the best individual bounds we can prove on
$S_3$
extensions of F use class field theory and a trivial bound for
$| {\mathrm {Cl}}_{K/F}[3]|$
, and are too large when summed. Ellenberg and Venkatesh [Reference Ellenberg and VenkateshEV07] have shown that one can improve the trivial bound when K has enough small split primes. Ellenberg, Pierce, and the third author [Reference Ellenberg, Pierce and WoodEPW17] introduced the idea that in many situations, one can prove that most fields in a family have enough small split primes to apply [Reference Ellenberg and VenkateshEV07]. We use that approach, though there is one significant new aspect. As we sum
$| {\mathrm {Cl}}_{K}[3]|=| {\mathrm {Cl}}_{F}[3]|| {\mathrm {Cl}}_{K/F}[3]|$
over F, we require a significant saving over the trivial bound. We structure our proof as an induction, which allows us to assume an optimal (constant average) bound for the
$| {\mathrm {Cl}}_{F}[3]|$
factor. We then further need a nontrivial bound on the relative class number
$| {\mathrm {Cl}}_{K/F}[3]|= | {\mathrm {Cl}}_{K}[3]|/| {\mathrm {Cl}}_{F}[3]|.$
For this, in Section 2 we develop a relative version of the widely used lemma of Ellenberg and Venkatesh [Reference Ellenberg and VenkateshEV07], showing that one obtains a nontrivial bound on
$| {\mathrm {Cl}}_{K/F}[p]|$
when there are sufficiently many small split primes in
$K/F$
. To find that most fields have small split primes, in Section 4 we use a uniform zero density estimate for Hecke L-functions as in [Reference PastenPas17, Proposition A.2] or [Reference Thorner and ZamanTZ21, Theorem 1.2] to show that most of the Dedekind zeta functions do not have zeros in the relevant region. In Section 5, we put these results together to prove the desired tail bound, and in Section 6 we prove the main theorems.
Our methods apply more broadly than just to
$2$
-groups. We determine in Section 8 the average size of
$3$
-torsion in the class group of
$(C_2 \wr G)$
-extensions for many general classes of group G, where
$C_2 \wr G$
denotes the wreath product.
1.2 Previous work
There is a large body of previous work in arithmetic statistics and analytic number theory on the questions that arise in the this paper. There have been many remarkable achievements on counting cubic extensions, and relatedly averaging three torsion in class groups over quadratic extensions, after the work of Davenport and Heilbronn [Reference Davenport and HeilbronnDH71] and Datskovsky and Wright [Reference Datskovsky and WrightDW88] discussed above. Taniguchi and Thorne [Reference Taniguchi and ThorneTT13] and Bhargava, Shankar, and Tsimerman [Reference Bhargava, Shankar and TsimermanBST13] improved the error term in Davenport and Heilbronn’s count so far as to uncover a secondary term, and recent work by Bhargava, Taniguchi, and Thorne [Reference Bhargava, Taniguchi and ThorneBTT21] has improved the error term even further. Bhargava, Shankar, and Wang [Reference Bhargava, Shankar and WangBSW15] have given a geometry of numbers proof of the counting results of Datskovsky and Wright, opening the door to further applications. See also Thorne’s exposition [Reference ThorneTho11] of the different approaches that have been used to count cubic fields.
The problem of proving nontrivial bounds for
$| {\mathrm {Cl}}_K[p]|$
has also been the topic of many important works. Pierce, Helfgott, and Venkatesh [Reference PiercePie05, Reference Helfgott and VenkateshHV06, Reference PiercePie06] proved the first nontrivial bounds (at a prime p not dividing the order of the Galois group), breaking the trivial bound in the case that K is quadratic field and
$p=3$
. This bound was improved by Ellenberg and Venkatesh [Reference Ellenberg and VenkateshEV07] to
$| {\mathrm {Cl}}_K[3]|=O_\epsilon ( {\mathrm {Disc}} K^{1/3+\epsilon })$
for quadratic fields K. Bhargava, Shankar, Taniguchi, Thorne, Tsimerman, and Zhao [BST+20] gave a nontrivial bound on
$| {\mathrm {Cl}}_K[2]|$
for all K using geometry of numbers. Ellenberg and Venkatesh [Reference Ellenberg and VenkateshEV07] also proved a nontrivial bound on
$| {\mathrm {Cl}}_K[3]|$
whenever
$[K:{\mathbb {Q}}]\leq 4$
. The second author gave a nontrivial bound on
$| {\mathrm {Cl}}_K[p]|$
when K is Galois with the Galois group a noncyclic and nonquaternion
$\ell $
-group or a nilpotent group where every Sylow-
$\ell $
subgroup is noncyclic and nonquaternion [Reference WangWan21, Reference WangWan20]. In the setting when p divides the order of the Galois closure group of K, Klüners and the second author [Reference Klüners and WangKW20] proved
$| {\mathrm {Cl}}_K[p]|=O_\epsilon ( {\mathrm {Disc}} K^{\epsilon })$
when K has Galois closure group a p-group, generalizing the classical result from Gauss’s genus theory for K quadratic and
$p=2$
.
Soundararajan [Reference SoundararajanSou00, pg.690] was the first to observe that nontrivial bounds for
$| {\mathrm {Cl}}_K[p]|$
could be proved for almost all imaginary quadratic K, or equivalently, on average over K. Heath-Brown and Pierce [Reference Heath-Brown and PierceHP17] reinvigorated the study of such average bounds for
$| {\mathrm {Cl}}_K[p]|$
, as well as bounds on higher moments, by using the large sieve. As mentioned above, Ellenberg, Pierce, and the third author [Reference Ellenberg, Pierce and WoodEPW17] proved nontrivial bounds for most fields in low degree families using a nontrivial bound of Ellenberg and Venkatesh [Reference Ellenberg and VenkateshEV07] conditional on the existence of small split primes, and precise field counting results with local conditions to prove such existence. Pierce, Turnage-Butterbaugh, and the third author [Reference Pierce, Turnage-Butterbaugh and WoodPTBW20] found ways to prove nontrivial bounds for many more families using zero density estimates for L-functions to prove the required existence of small primes, which is a strategy we use in this paper. This strategy has been improved and extended to further cases by An [Reference AnAn20], Frei and Widmer [Reference Frei and WidmerFW18, Reference Frei and WidmerFW21], Widmer [Reference WidmerWid18], Thorner and Zaman [Reference Thorner and ZamanTZ19b], and the first author, Thorner and Zaman [Reference Oliver, Thorner and ZamanLOTZ21].
1.3 Further questions
We expect the methods in this paper can be used in other settings. We have forthcoming work with Alberts which uses a similar inductive approach to prove many new cases of Malle’s conjecture [Reference MalleMal04] on the asymptotics of
$|E_k(G,X)|$
. It would be natural to apply our approach to the
$2$
-part of class groups of fields that can be obtained as cubic extensions of other extensions, using [Reference BhargavaBha05] as a base case.
It was noted by Bhargava and Varma [Reference Bhargava and VarmaBV16] that the average
$3$
-torsion in class group of quadratic extensions does not change when the fields are subject to finitely many local conditions at finite primes (see also [Reference WoodWoo18]). We expect that the average in Theorem 1.2 behaves similarly and is insensitive to local conditions, while the average in Theorem 1.1 behaves differently and is quite sensitive to local conditions, and it would be interesting to investigate whether this is the case. A subtlety in this question is in defining a notion of local conditions that works well in towers of fields.
The contrast between the constants in Theorems 1.1 and 1.2 naturally leads to the question: when ordering fields by a particular invariant, which averages does one expect to be as predicted by Cohen–Lenstra–Martinet? For example, does the analogue of Theorem 1.2 hold when the fields K are instead ordered by the Artin conductor of the representation of
$ {\mathrm {Gal}}(\bar {k}/k)$
induced from the sign representation of
$ {\mathrm {Gal}}(K/F_K)$
? (This is an irreducible representation, as we show in the proof of Theorem 7.1.) In the case of quartic
$D_4$
extensions of
$k=\mathbb {Q}$
, the fields themselves were recently counted by this Artin conductor in work of Alt
$\check{u}$
g, Shankar, Varma, and Wilson [Reference Altuğ, Shankar, Varma and WilsonASVW21].
It would be interesting to extend our methods to as wide a range of groups G as possible (e.g., by incorporating ideas from Sections 2 and 4 into an argument as in Section 8). For example, does an analogue of Theorem 1.1 hold for
$(C_2 \wr G)$
-extensions for arbitrary abelian groups G? Arbitrary p-groups?
Finally, as noted above, Malle’s conjecture predicts asymptotics for
$|E_k(G,X)|$
. These predicted asymptotics are of the form
$|E_k(G,X)| \sim c_{G,k} X^{1/a(G)} (\log X)^{b(k,G)-1}$
for specified integers
$a(G)$
and
$b(k,G)$
, and an unspecified positive constant
$c_{G,k}$
, as X tends to infinity. Our work raises the question, what is the least exponent
$\alpha = \alpha (G)$
such that
$|E_k(G,X)| \ll _{[k:\mathbb {Q}],\epsilon } {\mathrm {Disc}}(k)^{\alpha +\epsilon } X^{1/a(G)} (\log X)^{b(k,G)-1}$
for all k, all
$\epsilon>0$
, and all
$X \geq 1$
? Does such an exponent even exist? When G is abelian, this problem is closely related to bounding torsion subgroups of class groups. Some of the key technical results of this paper (see Section 3) provide the first such result for a nonabelian group G, namely
$G=S_3$
. What can be said for other nonabelian groups G? This question when
$G=S_3$
could naturally be approached by proving a discriminant-aspect subconvexity bound for the relevant Shintani zeta functions, which the methods of Section 3 fall short of proving. Can these or other methods be adapted to prove such a subconvexity result? We mention recent work of Hough and Lee [Reference Hough and LeeHL21] that establishes a subconvexity bound in t-aspect for the Shintani zeta functions over
$\mathbb {Q}$
, though we expect these methods do not easily apply to discriminant-aspect.
1.4 Technical remarks
The conjectures of Cohen, Lenstra, and Martinet in fact fix some archimedean data and only consider families of fields with this fixed data. In particular, for Galois
$\Gamma $
-extensions
$L/k$
, they fix the
$\Gamma $
-module structure of
$\mathcal {O}_L^* \otimes _{\mathbb {Z}} {\mathbb {Q}}$
, which is equivalent [Reference Cohen and MartinetCM90, Théorème 6.7] to fixing the representation
$\oplus _{v} \operatorname {Ind}_{\Gamma _v}^{\Gamma } {\mathbb {Q}}$
, where the sum is over infinite places v of k and
$\Gamma _v$
is the decomposition group of
$L/k$
at v, and we call this the enriched signature. The papers of Cohen, Lenstra, and Martinet are agnostic on the relative frequency of the various enriched signatures (and even whether such limiting frequencies exist). So, to be precise, to obtain (1.1) from their conjectures one must also assume such frequencies exist (which is natural to conjecture given all known counting results, for example, [Reference Davenport and HeilbronnDH71, Reference WrightWri89, Reference Cohen, Diaz, Diaz and OlivierCyDO02, Reference BhargavaBha05], and is conjectured by Malle [Reference MalleMal04]), and the conjectural value would be given as a weighted average of particular values for each enriched signature as conjectured by Cohen, Lenstra, and Martinet, where the weights are these frequencies. When one fixes an enriched signature, one can use the results of [Reference Wang and WoodWW21] to find the average value predicted by Cohen, Lenstra, and Martinet (see Section 7).
We prove a version of Theorem 1.1 with a specified enriched signature (in fact with slightly more specification at infinity) in Theorem 6.1. However, even when we do this, we do not apparently obtain the values conjectured by Cohen and Martinet in [Reference Cohen and MartinetCM90]. We show this computationally is the case in Section 7.2. This is because in this setting each F appears as the index
$2$
subfield for a positive proportion of K, and so the
$3$
-ranks of the particular F that appear with significant frequency bias the statistics, as in the work of Bartel and Lenstra [Reference Bartel and LenstraBL20]. When we only average
$ {\mathrm {Cl}}_{K/F}$
as in Theorem 1.2, this bias is removed. See Section 7 for further discussion of this phenomenon.
1.5 List of notations
k: a number field, for us a subfield of a fixed algebraic closure
$\bar {{\mathbb {Q}}}$
$\mathcal {O}_k$
: the ring of algebraic integers in k
$ {\mathrm {disc}}(F/k)$
: relative discriminant ideal in
$\mathcal {O}_k$
, for an extension
$F/k$
$ {\mathrm {Nm}} I$
: norm
$|\mathcal {O}_k/I|$
of an ideal I in
$\mathcal {O}_k$
and extend to any fractional ideals in
$\mathcal {O}_k$
multiplicatively
$ {\mathrm {Nm}}_{F/k} I$
: an ideal in
$\mathcal {O}_k$
that is
$(I\cap \mathcal {O}_k)^{f(I| I\cap \mathcal {O}_k)}$
when I is a prime ideal in
$\mathcal {O}_F$
and f is the inertia degree of I over
$I\cap \mathcal {O}_k$
, and defined for any fractional ideals in
$\mathcal {O}_F$
multiplicatively
$ {\mathrm {Disc}}(F/k)$
:
$ {\mathrm {Nm}}( {\mathrm {disc}}(F/k))$
(When k omitted,
$k={\mathbb {Q}}$
, so
$ {\mathrm {Disc}}(F)$
denotes the absolute value of the discriminant)
$ {\mathrm {Gal}}(F/k)$
: the Galois closure group of
$F/k$
, defined as the Galois group of
$\tilde {F}/k$
as a permutation group, where
$\tilde {F}$
is the Galois closure of F over k (in
$\bar {{\mathbb {Q}}}$
) acting on the embeddings of
$F\rightarrow \tilde {F}$
that fix k pointwise
G-extension: for a permutation group G, an extension
$L/k$
of number fields (in
$\bar {{\mathbb {Q}}}$
), and a choice of isomorphism
$\phi : {\mathrm {Gal}}(L/k)\simeq G$
as permutation groups (see Definition 6.7)
$2$
-extension: a G-extension for some
$2$
-group G
$E_k(G,X)$
: the set of G-extensions
$F/k$
with
$ {\mathrm {Disc}}(F)\le X$
$N_k(G,X)$
: the number of G-extensions
$F/k$
with
$ {\mathrm {Disc}}(F/k)\le X$
(note relative discriminant, versus absolute above)
$ {\mathrm {Cl}}_F$
: class group of F
$ {\mathrm {Cl}}_{F/k}$
: relative class group of
$F/k$
, when
$k=\mathbb {Q}$
it is the usual class group of F (see Section 2)
$A[\ell ]$
:
$\{ [\alpha ]\in A \mid \ell [\alpha ] = 0\in A \}$
for an abelian group A
$h(F/k)$
,
$h(F)$
: the size of
$ {\mathrm {Cl}}_{F/k}$
,
$ {\mathrm {Cl}}_{F}$
$h_{\ell }(F/k)$
,
$h_{\ell }(F)$
: the size of
$ {\mathrm {Cl}}_{F/k}[\ell ]$
,
$ {\mathrm {Cl}}_{F}[\ell ]$
$\pi _k(X)$
: the number of prime ideals
$\mathfrak {p}$
in k with
$ {\mathrm {Nm}}( {\mathrm {\mathfrak {p}}})<X$
$\pi _k(Y; F, \mathcal {C})$
: the number of unramified prime ideals
$\mathfrak {p}$
in k with
$ {\mathrm {Nm}}_{k/{\mathbb {Q}}}( {\mathrm {\mathfrak {p}}})<Y$
and with Frobenius at
$ {\mathrm {\mathfrak {p}}}$
in the conjugacy class of
$\mathcal {C}$
in
$ {\mathrm {Gal}}(F/k)$
$r_1(k)$
: the number of real embeddings of k
$r_2(k)$
: the number of pairs of complex embeddings of k
$ {\mathrm {rk}} \mathcal {O}_K^*$
: the unit rank, that is,
$\dim _{{\mathbb {Q}}} (\mathcal {O}_K^*\otimes {\mathbb {Q}}$
)
$\zeta _k(s)$
: Dedekind zeta function of k
$\kappa _k$
: the residue
$\mathrm {Res}_{s=1}\zeta _k(s)$
$ {\mathrm {Rg}}_k$
: the regulator of k
$f(X) \sim g(X)$
:
$\lim _{X\to \infty } f(X)/g(X) = 1$
or
$\lim _{X\rightarrow \infty } f(X)=\lim _{X\rightarrow \infty }g(X)=0$
$f(X) =O_{p_1,\dots }(g(X))$
: there exists a constant C depending only on
$p_1,\dots $
such that for all values of all variables
$|f(X)|\leq Cg(X)$
$f(X) \ll _{p_1,\dots }g(X)$
:
$f(X) =O_{p_1,\dots }(g(X))$
$f(X) \gg _{p_1,\dots }g(X)$
: only used when both sides non-negative, and then means
$g(X) \ll _{p_1,\dots }f(X)$
2 Nontrivial bound for the relative class group
In this section, our main goal is to establish Lemma 2.1, which is a relative version of a widely used lemma of Ellenberg–Venkatesh [Reference Ellenberg and VenkateshEV07, Lemma 2.3]. Let
$L/K$
be an extension of number fields. The relative class group
$ {\mathrm {Cl}}_{L/K}\subset {\mathrm {Cl}}_L$
is defined to be the kernel of the map
$ {\mathrm {Nm}}_{L/K}: {\mathrm {Cl}}_L\to {\mathrm {Cl}}_K$
induced from the usual norm on fractional ideals of L.
Lemma 2.1 (Relative Ellenberg–Venkatesh).
Let
$L/K$
be an extension of number fields such that
$d:=[L:{\mathbb {Q}}]$
, let
$\ell \geq 1 $
be an integer, and let
$\theta $
be a real number such that
$0< \theta < \frac {1}{4\ell (d-1)}$
. Suppose there exist M pairs of distinct prime ideals
$(\wp _i, \bar {\wp }_i)$
in
$\mathcal {O}_L$
and integers
$a_i,b_i$
such that
-
1. For each i, we have
$\wp _i \cap \mathcal {O}_K = \bar {\wp }_i\cap \mathcal {O}_K$ and
$ {\mathrm {Nm}}_{L/K}(\wp _i)^{a_i} = {\mathrm {Nm}}_{L/K}(\bar {\wp }_i)^{b_i}$ ;
-
2. For each i, the norms satisfy
$ {\mathrm {Nm}}(\wp _i)^{a_i}= {\mathrm {Nm}}(\bar {\wp _i})^{b_i}< {\mathrm {Disc}}(L/K)^{\theta }$ ; and
-
3. For each i, there is no intermediate field H with
$K\subset H \subsetneq L$ such that powers of
$\wp _i$ and
$\bar {\wp _i}$ are both extended from a prime of H. (An ideal
$\mathfrak {a}\subset \mathcal {O}_L $ is extended from a prime of H if there exists a prime ideal
$ {\mathrm {\mathfrak {p}}}\subset \mathcal {O}_H$ such that
$\mathfrak {a} = {\mathrm {\mathfrak {p}}} \mathcal {O}_L$ .)
Then

Remark 2.2. In our application, we will only use prime ideals that are split completely in
$L/K$
; therefore the first and third conditions will be automatically satisfied, and the second condition will be equivalent to
$ {\mathrm {Nm}}(\wp _i\cap \mathcal {O}_K)< {\mathrm {Disc}}(L/K)^{\theta }$
.
The original lemma [Reference Ellenberg and VenkateshEV07] gives a nontrivial bound

for the absolute class group
$ {\mathrm {Cl}}_L[\ell ]$
under the same assumptions. Our improvement is in proving a saving from the trivial bound on the size of the relative class group
$ {\mathrm {Cl}}_{L/K}[\ell ]$
instead of the absolute class group. Such a refinement is crucial for our application since we will treat
$ {\mathrm {Cl}}_K$
and
$ {\mathrm {Cl}}_{L/K}$
separately.
By definition and basic properties of groups, we have an exact sequence

Let
$\iota $
be the natural map from
$ {\mathrm {Cl}}_K$
to
$ {\mathrm {Cl}}_L$
by extension of ideals. The composition
$ {\mathrm {Nm}}_{L/K}\circ \iota : {\mathrm {Cl}}_K[\ell ] \to {\mathrm {Cl}}_K[\ell ]$
is equal to multiplication by
$[L:K]$
and thus is an isomorphism when
$(\ell , [L:K]) =1$
. So when
$\ell $
is relatively prime to
$[L:K]$
, the map
$ {\mathrm {Nm}}_{L/K}: {\mathrm {Cl}}_L[\ell ] \rightarrow {\mathrm {Cl}}_K[\ell ]$
is surjective and we have a short exact sequence

Thus when
$(\ell , [L:K]) =1$
,

This leads to a strategy we use in this paper to bound
$h_{\ell }(L)$
, that is, to bound
$h_{\ell }(K)$
and
$h_{\ell }(L/K)$
separately.
In the following lemma we give the description of the cokernel of the map
$\mathrm {Nm}_{L/K}: {\mathrm {Cl}}_L \to {\mathrm {Cl}}_K$
.
Lemma 2.3. Given an extension
$L/K$
of number fields, the cokernel of
$\mathrm {Nm}_{L/K}: {\mathrm {Cl}}_L \to {\mathrm {Cl}}_K$
is isomorphic to
$ {\mathrm {Gal}}(M/K)$
, where
$M = H_K\cap L$
and
$H_K$
is the Hilbert class field of K.
Proof. By class field theory, the map
$ {\mathrm {Nm}}_{L/K}: {\mathrm {Cl}}_L \to {\mathrm {Cl}}_K$
agrees with the restriction map on the Galois groups
$ {\mathrm {Gal}}(H_L/L) \to {\mathrm {Gal}}(H_K/K)$
after identifying class groups and Galois groups with the reciprocity map. Since
$H_K/K$
is Galois, we have
$ {\mathrm {Gal}}(H_L/L) \twoheadrightarrow {\mathrm {Gal}}(H_KL/L)\simeq {\mathrm {Gal}}(H_K/M) = \mathrm {Im}( {\mathrm {Nm}}_{L/K}: {\mathrm {Cl}}_L \to {\mathrm {Cl}}_K)$
, and the lemma follows.
As a consequence, we have an upper bound
$|\mathrm {Coker}(\mathrm {Nm}_{L/K}: {\mathrm {Cl}}_L \to {\mathrm {Cl}}_K )|\leq [L:K]$
, and thus we can obtain an upper bound on
$h(L/K)$
, which then gives a “trivial bound” on the
$\ell $
-torsion in the relative class group.
Lemma 2.4. Given a relative extension
$L/K$
with
$[L:K]=d$
and an arbitrary integer
$\ell>0$
, we have

and

Proof. We have

where the first inequality follows from Lemma 2.3, and the second inequality comes from the theorem of Brauer-Siegel and the expression of the relative discriminant in terms of absolute discriminants. We have an absolute lower bound
$\frac { {\mathrm {Rg}}(L)}{ {\mathrm {Rg}}(K)} \gg _{[L:{\mathbb {Q}}]} 1$
on the ratio of regulators by [Reference Friedman and SkoruppaFS99], and hence
$ h(L/K)\ll _{[L:{\mathbb {Q}}]} \frac { {\mathrm {Rg}}(L)}{ {\mathrm {Rg}}(K)} h(L/K). $
Then the second statement of the lemma follows from the trivial inequality
$h_{\ell }(L/K) \le h(L/K)$
.
Now we are ready to prove Lemma 2.1. The idea of the proof is similar to the proof of Lemma
$2.3$
in [Reference Ellenberg and VenkateshEV07], where the major difference is we use a different set T of ideals that cut out the nontrivial saving in the proof. As a result, we can show a saving that is completely coming from the relative class group.
Proof of Lemma 2.1.
First, at the cost of replacing M by at worst
$M/d$
(which is harmless), we may assume that distinct prime pairs lie over distinct primes of K. After renumbering, let
$\wp _0,\dots \wp _{N}$
be a subset of our chosen primes such that, for
$0\leq i\leq N$
, the ideals
$\wp _i^{\ell a_i}\bar {\wp _i}^{-\ell b_i}$
all have the same image in
$ {\mathrm {Cl}}_{L/K}$
. We have M pairs of primes, and each
$\wp _i^{a_i} \bar {\wp _i}^{-b_i}$
is an element of
$ {\mathrm {Cl}}_{L/K}$
. Since the kernel of the
$\ell $
th power map on
$ {\mathrm {Cl}}_{L/K}$
is
$ {\mathrm {Cl}}_{L/K}[\ell ]$
, we have that the image of this map has size
$| {\mathrm {Cl}}_{L/K}|/| {\mathrm {Cl}}_{L/K}[\ell ]|$
. Thus by the pigeonhole principle, we can always consider such a subset with
$N+1\geq M/(| {\mathrm {Cl}}_{L/K}|/| {\mathrm {Cl}}_{L/K}[\ell ]|).$
Let
$\alpha _i\in L^*$
such that
$(\alpha _i)=\wp _i^{\ell a_i}\bar {\wp _i}^{-\ell b_i} \wp _{0}^{-\ell a_0}\bar {\wp }_{0}^{\ell b_0}$
for
$1\leq i\leq N$
. Let
$w_1,\dots ,w_m$
be the infinite places of K, and let
$v_{i,j}$
be the places of L above
$w_i$
, for
$1\leq j \leq k_i$
. Let
$k=\sum _i k_i$
be the number of infinite places of L. We consider the space
${\mathbb {R}}^k$
whose coordinates are indexed by the infinite places
$v_{i,j}$
of L. The coordinates then naturally occur in m blocks, with the ith block having
$k_i$
coordinates. Consider the map
$\mathcal {L}: L^* \rightarrow {\mathbb {R}}^{k}$
given by
$\mathcal {L}(\alpha )= (\log (|\alpha |_{v_{i,j}}^{\operatorname {deg}(v_{i,j})}))_{v_{i,j}}$
, where the degree of a complex place is
$2$
and of a real place is
$1$
.
Let
$U_i: {\mathbb {R}}^{k_i} \rightarrow {\mathbb {R}}^{k_i}$
be the map such that
$U_i(x_1,\dots ,x_{k_i})=(x_1+\cdots +x_{k_i},x_1-x_2,x_2-x_3,\dots , x_{{k_i}-1}-x_{k_i})$
. If, for all i, we apply
$U_i$
to the ith block coordinates of
${\mathbb {R}}^k$
corresponding to places above
$w_i$
, we obtain a map
$U:{\mathbb {R}}^k \rightarrow {\mathbb {R}}^k$
.
Let
$\mathcal {O}^*:=\ker ( {\mathrm {Nm}}_{L/K}: \mathcal {O}_L^*\rightarrow \mathcal {O}_K^*)$
. For
$\alpha \in \mathcal {O}^*$
, the first coordinate in each block of
$U(\mathcal {L}(\alpha ))$
is
$0$
. Let
$P: {\mathbb {R}}^k \rightarrow {\mathbb {R}}^{k-m}$
be the map that just omits the first coordinate in each block. Note that the image of
$ {\mathrm {Nm}}_{L/K}: \mathcal {O}_L^*\rightarrow \mathcal {O}_K^*$
has finite index in
$\mathcal {O}_K^*$
, and thus by Dirichlet’s unit theorem,
$U(\mathcal {O}^*)$
has rank
$k-m$
. Let V be the covolume of
$PU \mathcal {L}(\mathcal {O}^*)$
in
${\mathbb {R}}^{k-m}$
(using the standard volume form on
${\mathbb {R}}^{k-m}$
).
Consider the images of
$PU \mathcal {L}(\alpha _i)$
in
${\mathbb {R}}^{k-m}/PU \mathcal {L}(\mathcal {O}^*)$
. Take a box of volume
$2^{k-m}$
around each
$PU \mathcal {L}(\alpha _i)$
(of length
$2$
in each coordinate). Suppose two of these boxes overlap. Then that means there exists a unit
$u\in \mathcal {O}^*$
, and a choice of
$r,s$
such that
$PU \mathcal {L}(u\alpha _r/\alpha _s)$
has its coordinates all at most
$2$
in absolute value.
For this chosen
$r,s$
, let us consider the height of the element
$u\alpha _r/\alpha _s$
in the sense of [Reference Ellenberg and VenkateshEV07, Equation (2.2)]:

Let
$x_v:=\log (|u\alpha _r/\alpha _s|_v^{\deg (v)})$
, that is, these are the coordinates of
$\mathcal {L}(u\alpha _r/\alpha _s)$
. Let
$t_i:=\sum _{v\mid w_i} x_v$
. Then since
$|x_{v_{i,j}}-x_{v_{i,j'}}|\leq 2$
, we have
$x_{v_{i,j}}=t_i/k_i +O_k(1).$
We have
$\sum _i t_i=0$
, since
$u,\alpha _r,$
and
$\alpha _s$
all have norm
$1$
. Thus
$ \sum _{i,j} \max (x_{v_{i,j}},0)=\sum _i t_i +O_k(1)=O_k(1). $
Thus the archimedean contribution to
$H(u\alpha _r/\alpha _s)$
is
$O_k(1)$
. The nonarchmiedean contribution is at most
$ {\mathrm {Disc}}(L/K)^{2\theta \ell }$
, and so
$H(u\alpha _r/\alpha _s)=O_k( {\mathrm {Disc}}(L/K)^{2\theta \ell })$
.
Let
$d=[L:K].$
The result [Reference Ellenberg and VenkateshEV07, Lemma 2.2] tells us that elements of small height belong to proper subfields. More precisely, for
$ {\mathrm {Disc}}(L/K)$
sufficiently large given k,
$[K:{\mathbb {Q}}]$
, and
$1/(2d-2)-2\theta \ell $
, by [Reference Ellenberg and VenkateshEV07, Lemma 2.2], we have that
$K_1:=K(u\alpha _r/\alpha _s)$
is a proper subfield of L. We are using
$1/(2d-2)-2\theta \ell>0$
in the application of [Reference Ellenberg and VenkateshEV07, Lemma 2.2].
We consider the valuations of
$u\alpha _r/\alpha _s$
. We have that
$u\in \mathcal {O}_L^*$
, so has valuation
$0$
at every prime of L or
$K_1$
. Recall that in L we have
$(u\alpha _r/\alpha _s)=\wp _r^{\ell a_i}\bar {\wp _s}^{\ell b_j} \bar {\wp _r}^{-\ell a_i}{\wp _s}^{-\ell b_j}$
. However any element from
$K_1$
that, in L, has positive valuation at
$\wp _r$
must also have positive valuation at some other prime of L dividing
$\wp _r \cap K$
, by our assumption that no power of
$\wp _r$
is extended from a prime in any smaller field. Since
$\wp _r \cap K\ne \wp _s \cap K$
, this is a contradiction. Thus, we conclude that for
$ {\mathrm {Disc}}(L/K)$
sufficiently large given k and
$1/(2d-2)-2\theta \ell $
, no two of our boxes overlap, and in particular
$2^{k-m}N \leq V$
.
It remains to compute V. We have that V is the determinant of a
$(k-m)\times (k-m)$
matrix
$R_{L/K}$
whose rows correspond to basis elements of
$\mathcal {L}(\mathcal {O}^*)$
and whose entries the the coordinates of
$PU $
applied to those basis elements. Consider the
$k\times k$
matrix R whose first
$k-m$
rows correspond to basis elements of
$\mathcal {L}(\mathcal {O}^*)$
, and next
$m-1$
rows correspond to basis elements of
$\mathcal {L}(\mathcal {O}_K^*)$
, and the entries in each of those rows are U applied to those basis elements. Let the final row have a
$1$
in the first entry of the first block and a
$0$
in every other column. If we rearrange the columns of R so that the first columns of each block become the final columns of the matrix, the we obtain a block diagonal matrix, whose upper-left
$(k-m)\times (k-m)$
minor is
$R_{L/K}$
, and has determinant of absolute value V. The bottom-right
$m\times m$
minor has entries
$d\log (|\gamma _i|_{w_{j}}^{\operatorname {deg}(w_{j})})$
for a system of fundamental units
$\gamma _i$
of
$\mathcal {O}_K^*$
(and last row with a single
$1$
entry), and thus this minor has determinant of absolute value
$d^{m-1} {\mathrm {Rg}} K$
.
Since the composition of the inclusion
$\mathcal {O}_K^* \hookrightarrow \mathcal {O}_L^*$
with the norm map raises elements to the power d, the quotient
$\mathcal {O}_L^*/(\mathcal {O}^* \mathcal {O}_K^*)$
is annihilated by d, and we have that
$[\mathcal {O}_L^*:\mathcal {O}^* \mathcal {O}_K^*]\leq d^d$
. So the determinant of R is at most
$d^d$
times the determinant of the
$k\times k$
matrix
$R'$
whose rows are
$U \mathcal {L}$
applied to basis of
$\mathcal {O}_L^*$
(and last row as in R). Since
$\det (U)=O_d(1)$
, we have
$\det R=O_d( {\mathrm {Rg}} L)$
. Thus,
$V=O_d(\det R/ {\mathrm {Rg}} K)=O_d( {\mathrm {Rg}} L/ {\mathrm {Rg}} K)$
.
So, for
$ {\mathrm {Disc}}(L/K)$
sufficiently large given
$[L:{\mathbb {Q}}]$
and
$1/(2d-2)-2\theta \ell $
, we then have

where the final bound is by Lemma 2.4. Note that given d and
$\theta>0$
, there are only finitely many
$\ell $
such that
$1/(2d-2)-2\theta \ell>0$
, so we can reduce to requiring
$ {\mathrm {Disc}}(L/K)$
sufficiently large given
$[L:{\mathbb {Q}}]$
and
$\theta $
, which proves the lemma in this case.
Finally, if
$ {\mathrm {Disc}}(L/K)=O_{[L:{\mathbb {Q}}],\theta }(1)$
, then since there are
$O_{[L:{\mathbb {Q}}]}( {\mathrm {Disc}}(L/K)^\theta )$
primes of L of norm up to
$ {\mathrm {Disc}}(L/K)^\theta $
, we have that
$M=O_{[L:{\mathbb {Q}}],\theta }(1)$
and the lemma follows from Lemma 2.4 in this case.
3 Uniform bounds on cubic extensions and the average
$3$
-part of quadratic extensions
In this section we are going to bound the sum of
$h_3(K/F)$
for quadratic extensions
$K/F$
of a general number field F, and the closely related number of
$S_3$
cubic extensions of F. This proof will employ different techniques and ideas for different regimes of X versus
$ {\mathrm {Disc}}(F)$
. Let
$N_F(G,X)$
be the number of isomorphism classes of G-extensions
$L/F$
with
$ {\mathrm {Disc}}(L/F)\le X$
. The main theorem is the following, which we expect will be useful for other applications as well.
Theorem 3.1. Let F be a number field, let
$h = h_2(F)$
denote the size of the
$2$
-torsion subgroup of the class group of F, and let
$D_F = {\mathrm {Disc}}(F)$
. For any
$X \geq 1$
we have

The same upper bound holds for
$N_F(S_3, X)$
. (The sum is over
$K\subset \bar {{\mathbb {Q}}}$
.)
We obtain the following convenient corollary for general number fields F.
Corollary 3.2. For any number field F and any
$X \geq 1$
, we have

The same upper bound holds for
$N_F(S_3,X)$
under the same hypotheses.
Using the trivial bound on
$h_2(F)$
in Corollary 3.2, the right-hand side becomes
$O_{[F:\mathbb {Q}],\epsilon }( {\mathrm {Disc}}(F)^{4/3+\epsilon } X)$
for any F, and when F is a
$2$
-extension, it follows from Lemma 5.6 that the right-hand side is
$O_{[F:\mathbb {Q}],\epsilon }( {\mathrm {Disc}}(F)^{1+\epsilon } X)$
.
Remark 3.3. Theorem 3.1 relies on an application of the Brauer–Siegel theorem to give a lower bound on the residue
$\mathrm {Res}_{s=1} \zeta _F(s)$
of the Dedekind zeta function of F. In particular, the implied constant depends on
$\epsilon $
ineffectively. However, the key results of Sections 3.1–3.3 from which Theorem 3.1 is obtained, namely Propositions 3.4, 3.6, and 3.12, have effective constants and an explicit dependence on the residue, and could be used to prove an effective analogue of Theorem 3.1 with dependence on the residue.
In order to prove Theorem 3.1, we will use different approaches for three regimes of X compared to
$ {\mathrm {Disc}}(F)$
: the small range, the large range, and the intermediate range. In Section 3.1, for X in the small range, we will use class field theory to give what we regard as a weak bound on the number of cubic extensions with a fixed discriminant. This will give a correspondingly weak bound on the number of cubic extensions with bounded discriminant. In Section 3.2, for X in the large range, by using the functional equation of the Shintani zeta function for cubic rings, we can use the bound on the coefficients we obtained from class field theory and get improved bounds on the number of cubic fields with discriminants quite large compared to
$ {\mathrm {Disc}}(F)$
. In Section 3.3, we will take advantage of the fact that the bound from the Shintani zeta function is actually also an upper bound for counting all cubic rings. Since the number of cubic orders associated to each cubic field is large, we show it is impossible to get too many cubic fields with X in the intermediate range. Finally, in Section 3.4, we prove Theorem 3.1 by combining the previous propositions for all ranges of X.
3.1 Small range: class field theory
In this section we will use class field theory to give the following upper bound on the number of small degree extensions of F.
Proposition 3.4. For any number field F and any
$X \geq 1$
, we have

To prove this proposition, we apply class field theory and the trivial bound on torsion in relative class groups to give the following pointwise bound on the number of relative
$S_3$
cubic extensions.
Lemma 3.5. For any positive integer n and any number field F, the number of extensions
$L/F$
in
$\bar {{\mathbb {Q}}}$
with
$\mathrm {Disc}(L/F)=n$
and
$[L:F]\leq 3$
is
$O_{[F:\mathbb {Q}],\epsilon }(n^{1/2+\epsilon } h_2(F) \mathrm {Disc}(F)^{1/2+\epsilon }).$
Proof. If
$L/F$
is an
$S_3$
cubic extension with
$\mathrm {Disc}(L/F)=n$
, then its associated quadratic resolvent
$K/F$
has
$ {\mathrm {Disc}}(K/F) = m$
and
$m| {\mathrm {Disc}}(L/F)$
. By class field theory, quadratic extensions of F are in bijection to surjective homomorphisms from the idèle class group
$C_F = \mathbb {A}^{\times }_F/F^{\times }$
to
$C_2$
, where
$\mathbb {A}^{\times }_F$
is the idèle group. We have the following exact sequence of
$C_F$
,

where the product is over places v of F and
$\mathcal {O}_v^\times $
is, at finite places, the group of elements of the completion
$F_v$
with valuation
$0$
, and, at infinite places, the group of nonzero elements of the completion
$F_v$
. Since
$ {\mathrm {Hom}}(\cdot , A)$
is left exact, we have

Here
$| {\mathrm {Hom}}( {\mathrm {Cl}}_F, C_2)| = h_2(F)$
. The global Artin map of class field theory maps
$O_v^\times \subset C_F$
to the inertia group of
$ {\mathrm {Gal}}(F^{ab}/F)$
, and the relative discriminant ideal of an extension only depends on the image of the inertia groups in the Galois group (e.g., since the discriminant is given by the Artin conductor of the permutation representation). Thus given a
$\rho \in {\mathrm {Hom}}(C_F, C_2)$
, its image in
$ {\mathrm {Hom}}( \prod _v O_{v}^{\times }, C_2)$
determines the relative discriminant ideal of the quadratic field corresponding to
$\rho $
. Therefore for each fixed m, the number of possible
$L/F$
with
$ {\mathrm {Disc}}(L/F)=m$
is bounded by
$h_2(F)$
. For each fixed n, there are at most
$O_{[F:{\mathbb {Q}}],\epsilon }(n^{\epsilon })$
possible m, therefore there are at most
$O_{[F:\mathbb {Q}],\epsilon }(n^\epsilon h_2(F))$
quadratic extensions
$K/F$
with
$\mathrm {Disc}(L/F)|n$
.
Similarly, for each fixed quadratic extension
$K/F$
, any associated
$S_3$
cubic extension
$L/F$
has Galois closure
$\tilde {L}/F$
such that
$\tilde {L}/K$
has
$ {\mathrm {Gal}}(\tilde {L}/K) = C_3$
. Thus, we can use class field theory again to determine that, for each fixed
$K/F$
, the number of associated
$S_3$
cubic extensions with
$ {\mathrm {Disc}}(L/F)|n$
is

by the trivial bound on
$h_3(K/F)$
in Lemma 2.4. Altogether, the total number of
$S_3$
-extensions is now
$O_{[F:\mathbb {Q]},\epsilon }(n^{1/2+\epsilon } h_2(F) \mathrm {Disc}(F)^{1/2+\epsilon })$
, as claimed. As above, we can bound the number of
$C_3$
-extensions
$K/F$
by
$O_{[F:\mathbb {Q}],\epsilon }(n^\epsilon h_3(F))$
and the lemma follows.
Summing Lemma 3.5 over integers n immediately yields Proposition 3.4.
3.2 Large range: the Shintani zeta function
In this section, we are going to give a better bound for
$N_F(S_3, X)$
when X is very large compared to
$ {\mathrm {Disc}}(F)$
. The main idea is to use the bound in Lemma 3.5 to get a better estimate on
$N_F(S_3, X)$
using Shintani zeta functions over a general number field F. Moreover, the upper bound we obtain in this section is also an upper bound on the number of cubic rings over
$\mathcal {O}_F$
, not just cubic fields. This will be important in Section 3.3.
Given a number field F, Shintani zeta functions over F count cubic rings over
$\mathcal {O}_F$
with a specified signature. More specifically, given a cubic ring R over
$\mathcal {O}_F$
(that is, a ring that is a locally free rank three
$\mathcal {O}_F$
-module), let
$ {\mathrm {disc}}(R/\mathcal {O}_F)$
denote the relative discriminant ideal of R and
$\mathrm {Disc}(R/\mathcal {O}_F) = {\mathrm {Nm}}_{F/{\mathbb {Q}}}( {\mathrm {disc}}(R/\mathcal {O}_F))$
. For each archimedean place v of F, let
$F_v$
be the completion of F at v, and let
$R_v:= R\otimes _{\mathcal {O}_F} F_v$
.Footnote
1
Each
$R_v$
is a cubic étale algebra over
$F_v$
. If
$F_v \simeq \mathbb {C}$
, then there is only one such cubic étale algebra
$\mathbb {C}^3$
, while if
$F_v \simeq \mathbb {R}$
, then either
$R_v \simeq \mathbb {R}^3$
or
$R_v \simeq \mathbb {R}\times \mathbb {C}$
. The signature of R is the tuple
$\mathrm {sgn}(R) = (R_v)_{v\mid \infty }$
that records the isomorphism class of each
$R_v$
.
We will use
$N_{\mathcal {O}_F,3}(X)$
to denote the number of isomorphism classes of cubic rings
$R/\mathcal {O}_F$
over the maximal order
$\mathcal {O}_F$
of a number field F with
$0< {\mathrm {Disc}}(R/\mathcal {O}_F)\le X$
, and
$N_{\mathcal {O}_F,\alpha }(X)$
to denote the number of isomorphism classes of cubic rings
$R/\mathcal {O}_F$
with
$0< {\mathrm {Disc}}(R/\mathcal {O}_F)\le X$
and with signature
$\alpha $
. Then the main proposition we are going to prove for this subsection is as follows.
Proposition 3.6. For any number field F and
$X \geq \mathrm {Disc}(F)^3 h_2(F)^{2/3}$
, we have

We start by introducing properties of Shintani zeta functions over a general number field, including all useful notations. For a fixed signature
$\alpha $
, let

where the sum runs over isomorphism classes of cubic rings R over
$\mathcal {O}_F$
.
Let
$\zeta _F(s)$
be the Dedekind zeta function of F. For convenience in what is to come, we define

Additionally, for each pair of signatures
$\alpha ,\beta $
and each infinite place
$v \mid \infty $
of F, we define a factor
$c_{v,\alpha \beta }(s)$
as follows. If
$F_v \simeq \mathbb {C}$
, set
$c_{v,\alpha \beta }(s) = \sin ^2(\pi s)\sin (\pi s - \pi /6) \sin (\pi s + \pi /6)$
. If
$F_v \simeq \mathbb {R}$
and
$\alpha _v \simeq \beta _v$
, set
$c_{v,\alpha \beta }(s) = \sin (2\pi s)/2$
. In the remaining cases, set

and define
$c_{\alpha \beta }(s) := \prod _{v\mid \infty } c_{v,\alpha \beta }(s)$
.
The basic properties of
$\xi _{F,\alpha }(s)$
are recorded in the following proposition that collects results due to Shintani [Reference ShintaniShi72], Wright [Reference WrightWri82, Reference WrightWri85], and Datskovsky and Wright [Reference Datskovsky and WrightDW86]. We also refer the interested reader to the work of Taniguchi [Reference TaniguchiTan06] for more information.
Proposition 3.7. We have the following properties for Shintani zeta functions over a general number field F:
-
1. For each number field F and each signature
$\alpha $ , the function
$\xi _{F,\alpha }(s)$ converges absolutely in the region
$\Re (s)>1$ .
-
2. The function
$\xi _{F,\alpha }(s)$ has a meromorphic continuation to
$\mathbb {C}$ with poles only at
$s=1$ and
$s=5/6$ . Each of these poles is simple, with residues
$$\begin{align*}\mathfrak{A}_F\cdot(1+3^{-r(\alpha)-r_2(F)}) \text{ and } \mathfrak{B}_F\cdot 3^{-r(\alpha)/2} \end{align*}$$
$r(\alpha ) = \#\{ v \mid \infty : \alpha _v \simeq \mathbb {R}^3\}$ .
-
3. The function
$\xi _{F,\alpha }(s)$ satisfies a functional equation,
$$\begin{align*}\xi_{F,\alpha}(1-s) = \left(\frac{3^{(6s-2)}}{\pi^{4s}}\Gamma(s)^{2}\Gamma(s-1/6)\Gamma(s+1/6)\right)^{[F:\mathbb{Q}]} \mathrm{Disc}(F)^{4s-2} \sum_{\beta} c_{\alpha\beta}(s) \hat\xi_{F,\beta}(s), \end{align*}$$
$\beta $ and where
$\hat \xi _{F,\beta }(s)$ is the dual Shintani zeta function defined via the Dirichlet series
$$\begin{align*}\hat\xi_{F,\beta}(s) = \sum_{\substack{R/\mathcal{O}_F: \\ \mathrm{Disc}(R/\mathcal{O}_F) \neq 0 \\ \mathrm{sgn}(R)=\beta \\ R \otimes \mathcal{O}_{F_v} \in \Sigma_v \, \forall v \mid 3}} \frac{|\mathrm{Aut}_{\mathcal{O}_F}(R)|^{-1}}{\mathrm{Disc}(R/\mathcal{O}_F)^s}, \end{align*}$$
$\mathcal {O}_F$ such that
$R \otimes \mathcal {O}_{F_v}$ lies in a particular subset
$\Sigma _v$ of cubic rings over
$\mathcal {O}_{F_v}$ for each place v dividing
$3$ .
-
4. For each
$\alpha $ , the function
$\hat \xi _{F,\alpha }(s)$ has a meromorphic continuation to all of
$\mathbb {C}$ with poles at
$s=1$ and
$s=5/6$ (both simple), and satisfies the inequality
$\hat \xi _{F,\alpha }(s) < \xi _{F,\alpha }(s)$ for real
$s>1$ .
-
5. For each
$\alpha $ , the functions
$(s-1)(s-5/6) \xi _{F,\alpha }(s)$ and
$(s-1)(s-5/6)\hat \xi _{F,\alpha }(s)$ are entire of order
$1$ .
Proof. We provide references for these claims without regard to necessarily providing the original reference. The first claim follows from [Reference WrightWri85, Theorem 4.1]. The second through fourth primarily follow from [Reference Datskovsky and WrightDW86, Theorem 6.2], with the claim that
$\hat \xi _{F,\alpha }(s) < \xi _{F,\alpha }(s)$
for real
$s>1$
following from the absolute convergence of
$\xi _{F,\alpha }(s)$
in this region. The fifth follows from a generalization of [Reference ShintaniShi72, Theorem 2.1.iii] as is explained on [Reference WrightWri82, pg. 96].
We next record one more useful property of
$\xi _{F,\alpha }(s)$
.
Lemma 3.8. If F has degree at least
$2$
, then
$\xi _{F,\alpha }(0)=0$
for each signature
$\alpha $
.
Proof. From the formulas given, it follows that each
$c_{\alpha \beta }(s)$
has a zero of order
$[F:\mathbb {Q}]$
at
$s=1$
. The claim then follows from the functional equation for
$\xi _{F,\alpha }(s)$
.
While Proposition 3.7 establishes that
$\xi _{F,\alpha }(s)$
converges absolutely in the region
$\Re (s)>1$
, it does not guarantee that this convergence is uniform in F. To control this uniformity, we must control the number of cubic rings over
$\mathcal {O}_F$
with small discriminant. This will essentially be afforded to us by Proposition 3.4 on the number of small discriminant cubic algebras over F, except that
$\xi _{F,\alpha }(s)$
counts all orders in cubic étale algebras over F, not just the maximal orders. We will later exploit this overcounting in the proof of Proposition 3.15, but for now, we simply record the following result of Datskovsky and Wright [Reference Datskovsky and WrightDW86, Theorem 6.1] in order to track the number of orders in a cubic étale algebra
$A/F$
.
Lemma 3.9 (Datskovsky–Wright).
Let A be a cubic étale algebra over F with maximal order
$\mathcal {O}_A$
. Thus, either
$A\simeq F^3$
,
$A \simeq K \times F$
where
$K/F$
is a quadratic extension, or
$A \simeq L$
where
$L/F$
is a relative cubic field extension. For each n, let
$a_n$
denote the number of orders
$\mathcal {O} \subset A$
for which
$[\mathcal {O}_A:\mathcal {O}]^2 = n$
, equivalently, the number of orders
$\mathcal {O}\subset A$
for which
$ {\mathrm {Disc}}(\mathcal {O}/\mathcal {O}_F) = {\mathrm {Disc}}(\mathcal {O}_A/\mathcal {O}_F) \cdot n$
. Then the Dirichlet series
$f_A(s):=\sum _{n} a_n n^{-s}$
satisfies

where

with notation as above, and where for a number field L,
$\zeta _L(s)$
denotes the Dedekind zeta function of L.
Combining Lemma 3.9 and Lemma 3.5, we derive the following lemma on the values of Shintani zeta function in a right half-plane.
Lemma 3.10. For any real
$\sigma>3/2$
and any
$t \in \mathbb {R}$
, we have
$\xi _{F,\alpha }(\sigma +it) \ll _{[F:\mathbb {Q}],\sigma ,\epsilon } \mathrm {Disc}(F)^{1/2+\epsilon }h_2(F)$
.
Proof. By the triangle inequality, it suffices to prove the lemma when
$t=0$
. Assume
$\sigma>3/2$
is given. With the notation of Lemma 3.9, for any cubic étale algebra A, we have
$\zeta _A(\sigma ) \leq \zeta _F(\sigma )^3$
whenever
$\sigma> 1$
. It follows that

where the sum runs over isomorphism classes of field extensions
$L/F$
of degree at most
$3$
. The second inequality comes from
$\zeta _A(4\sigma )\ge 1$
, valid for
$\sigma>1$
. The factors of
$\zeta _F(\sigma )$
may be bounded by suitable powers of the Riemann zeta function
$\zeta (\sigma )$
, depending only on the degree of
$F/\mathbb {Q}$
. So it suffices to understand the Dirichlet series of field discriminants. By Lemma 3.5, we then find that for any
$\epsilon < \sigma - 3/2$
,

yielding the lemma.
Next, we apply the functional equation to deduce a discriminant-aspect “convexity bound” for the Shintani zeta function.
Lemma 3.11. For any
$-1/2 \leq \sigma \leq 3/2$
and any
$t \in \mathbb {R}$
, the Shintani zeta function satisfies

Proof. Let
$\delta> 0$
be small. For
$\sigma = 3/2+\delta $
, it follows from Lemma 3.10 that

For
$s = \sigma + it$
with
$\sigma = 3/2 + \delta $
, we apply the functional equation and Stirling’s formula to extract cancellation between the gamma factors and the
$c_{\alpha \beta }(s)$
in order to find also

The general result follows upon using the Phragmen–Lindelöf principle [Reference Iwaniec and KowalskiIK04, Theorem 5.53] applied to the function
$\frac {(s-1)(s-5/6)}{(s+1)^2} \xi _{F,\alpha }(s)$
and letting
$\delta $
be sufficiently small. Specifically, we find for any
$-1/2- \delta < \sigma < 3/2 + \delta $
that

Upon choosing
$\delta = \epsilon $
and then replacing
$\epsilon $
in the equation above by
$\epsilon /(1+2[F:\mathbb {Q}])$
, we obtain the required statement.
Now we apply the functional equation in Proposition 3.7 and the above lemmas to bound the number of cubic rings
$N_{\mathcal {O}_F,3}(X)$
.
Proof of Proposition 3.6.
Let
$\phi (x)$
be a positive smooth function satisfying
$\phi (x)\geq 1$
whenever
${0\leq x \leq 1}$
and whose Mellin transform exhibits rapid decay away from the real axis; for example, we may take
$\phi (x)=e^{1-x}$
. Letting R run over cubic rings over
$\mathcal {O}_F$
, we have

This latter term may be evaluated via Perron’s formula and the full Shintani zeta function
$\xi _{F,\alpha }(s)$
,

where
$\Phi (s)$
is the Mellin transform of
$\phi (x)$
. By Lemma 3.11, the exponential decay of
$\Phi (s)$
off the real axis allows us to shift the contour to the line
$\Re (s)=-1/2-\epsilon $
for any
$\epsilon>0$
. Doing so, we must account for the poles of
$\xi _{F,\alpha }(s)$
at
$s=1$
and
$s=5/6$
and of
$\Phi (s)$
at
$s=0$
. From Proposition 3.7, it follows that the contribution from the pole at
$s=1$
subsumes that of the pole at
$s=5/6$
. Moreover, Lemma 3.8 shows that in fact the integrand has no pole at
$s=0$
owing to the zero of
$\xi _{F,\alpha }(s)$
at
$s=0$
. Altogether, the contribution from the poles is seen to be
$O_{[F:\mathbb {Q}]}(X\cdot \mathrm {Res}_{s=1}\zeta _F(s))$
.
Finally, appealing to the functional equation and Lemma 3.10, we find

where
$\hat \xi (s) = \sum _{\beta }c_{\alpha \beta }(s)\hat \xi _{F,\beta }(s)$
.
We conclude that for any number field F and any
$X \geq 1$
, we have

Using the well-known upper bound
$\mathrm {Res}_{s=1} \zeta _F(s) \ll _{[F:\mathbb {Q}],\epsilon } \mathrm {Disc}(F)^\epsilon $
due to Landau (see also [Reference BrauerBra47, Lemma 3]), we conclude the proposition.
3.3 Intermediate range: propagation of orders
In this section, our main proposition is the following bound on
$N_F(S_3, X)$
that is better than both Proposition 3.4 and 3.6 when X is a bit smaller than the region allowed by Proposition 3.6. Let
$\kappa _F:=\mathrm {Res}_{s=1}\zeta _F(s)$
.
Proposition 3.12. Let F be a number field and let
$X \geq 1$
. We have

The key idea behind Proposition 3.12 is that Proposition 3.6 actually gives an upper bound on the number of cubic rings over
$\mathcal {O}_F$
instead of cubic fields, and if there are too many cubic fields in this intermediate range, and too many orders inside such fields on average, then the orders inside these fields would overrun the bounds from Proposition 3.6. The bulk of the work in this section is therefore to understand the number of orders inside small discriminant fields, on average.
We will first apply this idea to bound the average
$3$
-class number of quadratic extensions of F, and then we apply class field theory to bound the number
$N_F(S_3, X)$
of general
$S_3$
cubic fields.
We begin with the following lemma.
Lemma 3.13. Let
$K/F$
be a quadratic extension of number fields, and let
$\mathcal {R}(K)$
denote the set of isomorphism classes of cubic rings R over
$\mathcal {O}_F$
whose underlying cubic étale algebra
$A:=R\otimes _{\mathbb {Q}} F$
has discriminant
$ {\mathrm {disc}}(K/F)$
and is either of the form
$A \simeq F \times K$
or
$A \simeq L$
for a non-Galois cubic extension
$L/F$
with quadratic resolvent
$K/F$
. Then we have

where the summation on the right-hand side runs over those squarefree ideals of
$\mathcal {O}_K$
whose classes are in the subgroup
$3 {\mathrm {Cl}}_K+ {\mathrm {Cl}}_F \subseteq {\mathrm {Cl}}_K$
.
Proof. If
$R \in \mathcal {R}(K)$
, then
$\mathrm {Disc}(R/\mathcal {O}_F) = \mathrm {Disc}(K/F) [\mathcal {O}_A:R]^2$
. Thus, with the notation of Lemma 3.9, we have

Note that the summation on the right is restricted to those cubic extensions
$L/F$
with quadratic resolvent K and which satisfy
$ {\mathrm {disc}}(L/F) = {\mathrm {disc}}(K/F)$
. For each such
$L/F$
, by Lemma 3.9 we have

where
$\widetilde {L}$
is the Galois closure of L over F, and
$\chi _{\widetilde {L}/K}$
is one of the nontrivial cyclic cubic characters of the class group
$\mathcal {C}= \mathrm {Cl}_K/\mathrm {Cl}_F$
that cuts out
$\widetilde {L}/K$
. The second equality comes from the relation
$\zeta _L(s)/\zeta _F(s)= L(s, \chi _{\widetilde {L}/K})$
. In fact, we have
$L(s,\chi _{\widetilde {L}/K}) = L(s,\chi _{\widetilde {L}/K}^2) = L(s, \overline {\chi }_{\widetilde {L}/K})$
, since all are equal to the irreducible degree
$2$
Artin L-function of
$L/F$
. Thus, we obtain also

Meanwhile, the Dirichlet series for orders inside
$K\times F$
is

Notice that
$\zeta _K(s)$
is the L-function attached to the trivial character of
$\mathcal {C}$
. Therefore,

by orthogonality of characters and the relation
$|\widehat {\mathcal {C}}[3]| = h_3(K/F)$
.
We next have the following simple lemma.
Lemma 3.14. Let F be a number field and let
$\delta>0$
. There is an effectively computable constant C, depending at most on
$\delta $
and
$[F:\mathbb {Q}]$
, such that for any
$X \geq C \mathrm {Disc}(F)^{1/2+\delta } \kappa _F^{-1}$
, we have

Proof. Let
$m \geq \frac {[F:\mathbb {Q}]}{2}+2$
be an integer. Then the number of ideals
$\mathfrak {a} \subseteq \mathcal {O}_L$
with
$ {\mathrm {Nm}}(\mathfrak {a}) \leq X$
satisfies

The convexity bound for
$\zeta _F(s)$
(see [Reference Iwaniec and KowalskiIK04, Equation (5.20)], for example) implies that when
$\sigma :=\Re (s)$
lies in the critical strip
$0 \leq \sigma \leq 1$
, we have

Thus, the integral above converges absolutely when the line of integration is moved anywhere inside the critical strip. Taking it arbitrarily close to the line
$\Re (s)=0$
, we conclude that

The result follows.
We can now apply our count of orders to get an upper bound on the average
$3$
-class number of quadratic extensions.
Proposition 3.15. Let F be a number field and let
$X \geq 1$
. We have

Proof. Let Y be given by

where C is the constant from Lemma 3.14. For each quadratic extension
$K/F$
with discriminant at most X, we consider the contribution to
$N_{\mathcal {O}_F,3}(Y)$
from the orders in the family
$\mathcal {R}(K)$
from Lemma 3.13. For any
$R \in \mathcal {R}(K)$
, any ring automorphism extends to an automorphism of the étale algebra
$A = R \otimes _{\mathbb {Q}} F$
. Thus,
$|\mathrm {Aut}_{\mathcal {O}_F} (R)| \leq |\mathrm {Aut}_F(A)|$
, and we obtain an inequality of generating functions,

which is the Dirichlet series from Lemma 3.13. Appealing to Lemma 3.13, this series is in turn bounded below by

since the summation over ideals
$\mathfrak {a}$
has non-negative coefficients. It then follows from our assumption
$Y\geq \kappa _F^{-2} {X C^2\mathrm {Disc}(F)^{1+\epsilon }} $
that we may appeal to Lemma 3.14 (with
$X = Y^{1/2}/\mathrm {Disc}(K/F)^{1/2}$
) to find that the contribution to
$N_{\mathcal {O}_F,3}(Y)$
of orders arising in
$\mathcal {R}(K)$
is

Thus,

and we conclude that

Finally, by Proposition 3.6, we have
$N_{\mathcal {O}_F,3}(Y) \ll _{[F:\mathbb {Q}],\epsilon } \mathrm {Disc}(F)^{\epsilon /2} Y$
whenever
$ Y\geq \mathrm {Disc}(F)^3 h_2(F)^{2/3},$
and the result follows using dyadic summation and the fact that
$\kappa _F \ll _{[F:\mathbb {Q}],\epsilon } \mathrm {Disc}(F)^{\epsilon /2}$
.
Proof of Proposition 3.12.
Following the proof of Lemma
$6.2$
in [Reference Datskovsky and WrightDW88], for a square-free ideal
$\mathfrak {q}$
of F of norm
$ {\mathrm {Nm}} \mathfrak {q} \leq X^{1/2}$
, the number of
$S_3$
cubic fields
$L/F$
with quadratic resolvent
$K/F$
and
$ {\mathrm {disc}}(L/F) = \mathfrak {q}^2 {\mathrm {disc}}(K/F)$
is bounded by
$O(4^{\omega (\mathfrak {q})} h_3(K/F))$
, where
$\omega (\mathfrak {q})$
denotes the number of prime ideals dividing
$\mathfrak {q}$
. Thus, for fixed
$\mathfrak {q}$
, the number of
$S_3$
-extensions
$L/F$
with
$ {\mathrm {disc}}(L/F)\leq X$
whose discriminant is
$\mathfrak {q}^2$
times that of their quadratic resolvent, may be bounded by a uniform constant times

the latter inequality provided by Proposition 3.15. Next we will sum these expressions over
$\mathfrak {q}$
with norm at most
$X^{1/2}$
in order to bound
$N_F(S_3, X)$
. The Dirichlet series
$\sum _{\mathfrak {q}} \frac {4^{\omega (q)}}{ {\mathrm {Nm}} (\mathfrak {q})^s}$
is bounded term by term by
$\zeta (s)^{4[F:{\mathbb {Q}}]}$
. Thus it follows that

If the second summand above is bigger than the first, then the proposition immediately follows. Otherwise,
$X^{1/2-\epsilon }\leq \mathrm {Disc}(F) h_2(F)^{1/3}$
and the first summand is bigger. If also
$\epsilon \leq 1/4$
, then
$X\ll _{[F:{\mathbb {Q}}]} \mathrm {Disc}(F)^8$
and so the
$X^\epsilon $
in the first term can be subsumed into the
$\mathrm {Disc}(F)^\epsilon $
, and the proposition follows. The
$\epsilon>1/4$
case of the proposition immediately follows from the
$\epsilon \leq 1/4$
case.
3.4 Proof of Theorem 3.1
We can now combine our results in the various ranges to prove Theorem 3.1. By the bijection between
$3$
-torsion elements in relative class groups of quadratic extensions and
$S_3$
relative cubic extensions with square-free discriminant ideal [Reference Datskovsky and WrightDW88, Section 5], we have

Using this, the first case follows directly from Proposition 3.4. Propositions 3.12 and 3.15 and give the same upper bound for
$N_F(S_3,X)$
and
$\sum _{K} h_3(K/F)$
. This bound gives the second and third cases, depending on the size of X, using the fact that
$\mathrm {Res}_{s=1} \zeta _F(s) \gg _{[F:{\mathbb {Q}}],\epsilon } \mathrm {Disc}(F)^{-\epsilon }$
(though, this is only known to hold with an effectively computable implied constant when
$\zeta _F(s)$
does not have an exceptional Landau–Siegel zero; see [Reference StarkSta74]). The fifth case follows from (3.8) and Proposition 3.6, using that every cubic extension of F has a maximal order which is a cubic ring over
$\mathcal {O}_F$
, and every quadratic extension
$K/F$
has a maximal order
$\mathcal {O}_K$
such that
$\mathcal {O}_K\oplus \mathcal {O}_F$
is a cubic ring over
$\mathcal {O}_F$
. The fourth case then follows from with fifth case with the additional observation that
$N_{F}(S_3, X) \leq N_F(S_3, D_F^3h^{2/3})$
for any
$X \leq D_F^3h^{2/3}$
(and analogously for class numbers).
4 Zero-density estimates for ray class L-functions
In this section, we collect the results we need on zero-density estimates for ray class L-functions and use them to show that most quadratic extensions will have sufficiently many small split primes. This will allow us to apply our relative Ellenberg–Venkatesh Lemma 2.1 and get a nontrivial bound on
$h_3(K/F)$
for most quadratic
$K/F$
.
Let
$\chi $
be a nontrivial ray class character of a number field k with conductor
$\mathfrak {f}_\chi $
. The associated L-function
$L(s,\chi )$
is defined by

the series running over the integral ideals of k. Let
$r_1^+(\chi )$
denote the number of real places of k that are split in the cyclic extension of k cut out by
$\chi $
, let
$r_1^-(\chi )$
denote the number of real places that are ramified in the extension, and let
$r_2$
denote the number of complex places of k. If we define

then the completed L-function
$\Lambda (s,\chi ) := ( {\mathrm {Nm}}(\mathfrak {f}_\chi ) \mathrm {Disc}(k))^{s/2} \gamma _\chi (s) L(s,\chi )$
is entire and satisfies a functional equation,
$\Lambda (1-s,\chi ) = \varepsilon (\chi ) \Lambda (s,\bar \chi )$
. The poles of
$\gamma _\chi (s)$
thus lend
$L(s,\chi )$
its trivial zeros; these are the nonpositive even integers with multiplicity
$r_1^+(\chi )+r_2$
, and the negative odd integers with multiplicity
$r_1^-(\chi )+r_2$
. See, for example, [Reference Iwaniec and KowalskiIK04, §5.13] for more on the analytic properties of such L-functions.
The following lemma summarizes the analytic properties we shall need.
Lemma 4.1. Assume notation as above and set
$q_\chi = {\mathrm {Nm}}(\mathfrak {f}_\chi ) \mathrm {Disc}(k)$
. Then:
-
1. For any
$t\geq 0$ , the number of nontrivial zeros
$\rho =\beta +i\gamma $ of
$L(s,\chi )$ with
$|\gamma -t| \leq 1$ is
$O_{[k:\mathbb {Q}]} (\log ((t+1) q_\chi ))$ .
-
2. For any s with
$-3/2 \leq \Re (s) \leq 2$ , we have
$$\begin{align*}-\frac{L^\prime}{L}(s,\chi) = -\frac{r_1^+(\chi)+r_2}{s} - \frac{r_1^-(\chi)+r_2}{s+1}-\sum_{\substack{ \rho: \\ |s-\rho|<1}} \frac{1}{s-\rho} + O_{[k:\mathbb{Q}]}(\log((|s|+1) q_\chi)), \end{align*}$$
$\rho $ of
$L(s,\chi )$ .
Proof. These are standard properties of L-functions. For example, see Proposition 5.7 in Iwaniec and Kowalski [Reference Iwaniec and KowalskiIK04]. The second conclusion is stated there only for
$-1/2 \leq \Re (s) \leq 2$
, but in the region
$-3/2 \leq \Re (s) \leq -1/2$
, the functional equation relates
$-\frac {L^\prime }{L}(s,\chi )$
to
$-\frac {L^\prime }{L}(1-s,\bar \chi )$
, which is absolutely convergent. Thus, the conclusion is easier in this case, with the factors arising from the functional equation being dominant.
We next recall a consequence of Proposition A.2 in the appendix by Lemke Oliver and Thorner to work of Pasten [Reference PastenPas17], which follows upon taking
$\pi $
to be the trivial representation of k. See also [Reference Thorner and ZamanTZ21, Theorem 1.2] for a somewhat stronger statement, which would allow also
$\epsilon = 0$
below with an explicit value of c.
Theorem 4.2. Let k be a number field, and for any ray class character
$\chi $
of k, let

Then there is a constant
$c>0$
, depending only on the degree
$[k:\mathbb {Q}]$
, such that for any
$Q,T>1$
, any
$1/2 \leq \sigma <1$
, and any
$\epsilon> 0$
, we have

Here, the outer summation runs over ideals
$\mathfrak {q}$
of
$\mathcal {O}_k$
with bounded norm and the inner summation runs over primitive ray class characters of conductor
$\mathfrak {q}$
.
Theorem 4.2 implies that over a general number field k, most ray class L-functions do not have zeros near
$s=1$
. Therefore we will utilize Theorem 4.2 to show that for most quadratic extensions over k we can find enough prime ideals with a specified splitting behavior. Recall that
$\pi _k(Y)$
denotes the prime ideal counting function of k, and for an extension
$F/k$
,
$\pi _k(Y;F,e)$
denotes the number of primes in k of norm at most Y that are split in F.
Lemma 4.3. Let k be a number field and let
$\mathcal {F}_2(X)$
be the set of quadratic extensions of k with
$ {\mathrm {Disc}}(F/k)\le X$
. Let
$c = c_{[k:{\mathbb {Q}}]}$
be the constant as given in Theorem 4.2. For any
$\epsilon _1>0$
and
$X\geq 2$
, there exists a set
$\mathcal {E} = \mathcal {E}(k, X, \epsilon _1)\subset \mathcal {F}_2(X)$
of exceptional quadratic extensions
$F/k$
such that:
-
1. for all
$F \in \mathcal {F}_2(X)\setminus \mathcal {E}$ , for all
$4 \leq Y \le X$ there exists a constant
$C_{[k:{\mathbb {Q}}],\epsilon _1}$ depending only on
$[k:{\mathbb {Q}}],\epsilon _1$ such that
$$ \begin{align*}\pi_k(Y; F, e) \geq \frac{1}{8} \pi_k(Y/2) - C_{[k:{\mathbb{Q}}],\epsilon_1}Y^{\sigma_1} \log^2( X {\mathrm{Disc}}(k)), \end{align*} $$
$\sigma _1 = \max (1- \frac {\epsilon _1}{4c},1/2)$ ; and
-
2.
$|\mathcal {E}| \ll _{[k:{\mathbb {Q}}], \epsilon _1} {\mathrm {Disc}}(k)^{\epsilon _1}X^{\epsilon _1}$ .
Proof. For each quadratic extension
$F/k$
, there is a unique quadratic ray class character
$\chi _{F/k}$
with conductor
$ {\mathrm {disc}}(F/k)$
such that
$\frac {\zeta _F(s)}{\zeta _k(s)} = L(s, \chi _{F/k})$
. As in the statement of the lemma, we set
${\sigma _1 = 1-\epsilon _1/4c}$
and define

Then Theorem 4.2 implies
$|\mathcal {E}| \ll _{[k:{\mathbb {Q}}], \epsilon _1/4} ( {\mathrm {Disc}}(k) X^{3/2} )^{c(1-\sigma _1)+\epsilon _1/4}\ll _{[k:{\mathbb {Q}}],\epsilon _1} {\mathrm {Disc}}(k)^{\epsilon _1}X^{\epsilon _1}$
.
Suppose
$F \notin \mathcal {E}$
. In order to bound from below the number of primes of k with
$\chi _{F/k}(\mathfrak {p}) = 1$
, we first observe that, away from the
$O(\log {\mathrm {Disc}}(F))$
ramified primes, the function
$(1+\chi _{F/k}(\mathfrak {p}))/2$
is precisely
$1$
or
$0$
according to whether
$\mathfrak {p}$
is split or inert. As is standard, we first consider a version weighted by the von Mangoldt function
$\Lambda _k$
of k, and we observe

The first term is related to
$\pi _k(Y)$
by means of partial summation, so our goal is to show that the second is small.
Via Perron’s formula, we find

It follows from Lemma 4.1 and the functional equation that the integral converges absolutely over any vertical line. Thus, we may shift the contour all the way to the left, finding

where the summation runs over all zeros of
$L(s,\chi )$
, nontrivial or trivial, and where
$\mathrm {Res}_0$
and
$\mathrm {Res}_{-1}$
denote the residues of the integrand at
$s=0$
and
$s=-1$
, respectively. The trivial zeros of
$L(s,\chi _{F/k})$
occur at negative integers, and each has multiplicity at most
$[k:\mathbb {Q}]$
, so their total contribution to the sum may be bounded unifomly by
$O_{[k:\mathbb {Q}]}(Y^{-2})$
. Using Lemma 4.1, the residue
$\mathrm {Res}_{-1}$
is bounded by
$O_{[k:\mathbb {Q}]}((\log Y)(\log X\mathrm {Disc}(k))Y^{-1})$
, which is sufficient. Again using Lemma 4.1, we see that

since
$F \not \in \mathcal {E}$
and
$Y \leq X $
.
By Lemma 4.1, the summation over zeros of height greater than
$T = X^{1/2} $
is bounded by

Notice that by definition of
$\mathcal {E}$
, for
$F \notin \mathcal {E}$
there are no zeros for
$L(s, \chi _{F/k})$
with
$\Re (\rho )>\sigma _1$
and
$|\Im (\rho )| \leq T = X^{1/2} $
. Using the functional equation, there are also no zeros with
$\Re (\rho )\in (0,1-\sigma _1)$
and
$|\Im (\rho )| \leq T$
. The summation over low-lying zeroes of
$L(s, \chi _{F/k})$
may therefore be bounded by

Finally, by the standard translation from the von Mangoldt function to the prime counting function, we can conclude the lemma.
In order to apply Lemma 4.3 efficiently, we need the following effective lower bound on the number of prime ideals in a general number field k.
Lemma 4.4 [Reference ZamanZam17].
There exist absolute, effective constants
$\gamma $
,
$\beta $
, and
$D_0>0$
such that if k is a number field with
$ {\mathrm {Disc}}(k)\ge D_0$
, then for
$Y\ge {\mathrm {Disc}}(k)^{\beta }$
, we have

An effective lower bound, weaker than that above, can be derived from [Reference Lagarias, Montgomery and OdlyzkoLMO79], but Zaman [Reference ZamanZam17] was the first to prove a bound of this strength; he did so with
$\beta = 35$
and
$\gamma = 19$
. It follows from [Reference Thorner and ZamanTZ17] that
$\beta = 694$
and
$\gamma = 5$
are also admissible, and from [Reference Thorner and ZamanTZ19a] that
$\gamma = 1/[k:{\mathbb {Q}}]$
is admissible with an inexplicit value of
$\beta $
.
5 A uniform tail estimate
A
$2$
-extension is a G-extension for some
$2$
-group G. In this section, our main goal is to prove the following theorem on a tail estimate for the
$3$
-torsion in class groups of
$2$
-extensions with a uniform dependence on the base field k. It will be the critical input for the proof of our main theorems in Section 6.
Theorem 5.1. Let k be a number field. For each
$m\ge 2$
, there exist constants
$\delta _m>0$
and
$\alpha _m$
depending at most on m and the degree
$[k:{\mathbb {Q}}]$
such that for any
$X,Y>0$
, we have

where the summation is over all towers
$F_m/F_{m-1}/ \cdots F_1/F_0=k$
in
$\bar {{\mathbb {Q}}}$
with
$F_{i+1}/F_{i}$
being quadratic extensions for each
$ 0\le i\le m-1$
.
We emphasize the point that Theorem 5.1 is stated with respect to the absolute discriminant, while the results of Section 3 are with respect to the norm of the relative discriminant.
Remark 5.2. The statement of Theorem 5.1 for
$m=1$
is empty except when
$Y \le {\mathrm {Disc}}(k)$
. If
${Y \leq {\mathrm {Disc}}(k)}$
, the statement for
$m=1$
is nontrivial, but follows from Corollary 3.2, and noting that it is stated with respect to relative discriminant, we may take
$\alpha _1$
to be any constant greater than
$1/3$
.
Remark 5.3. For our purposes, the actual values of
$\delta _m$
and
$\alpha _m$
are irrelevant; it suffices to know only that the constants
$\delta _m$
and
$\alpha _m$
exist. However, in the course of the proof, we provide explicit but nonoptimal values. This will show that the value of
$\delta _2$
may be chosen without dependence on
$[k:\mathbb {Q}]$
. We choose specific nonoptimal values of many related exponents in the course of the proof, because we wish to avoid tracking further dependencies of implied constants that would come with avoiding such choices.
Note that the tail estimate we prove in Theorem 5.1 handles all
$2$
-extensions due to the following easy lemma.
Lemma 5.4 ([Reference Klüners and WangKW20], Lemma 2.3).
Let
$n=\ell ^r$
be a prime power and
$G\subset S_n$
be an
$\ell $
-group and
$E/F$
be an extension of number fields with
$ {\mathrm {Gal}}(E/F)\cong G$
. Then there exists a tower of fields

such that
$ {\mathrm {Gal}}(F_{i+1}/F_i)=C_\ell $
for all
$0\leq i \leq r-1$
.
We will prove Theorem 5.1 first for
$m = 2$
. Then treating
$m=2$
as the base case, we will apply an inductive proof to prove Theorem 5.1 for general m. For
$m=2$
, the theorem requires us to sum over all towers
$F_2/F_1/F_0=k$
of relative quadratic extensions. We will separate the discussion depending on how large
$ {\mathrm {Disc}}(F_1)$
is. In Section 5.1, we will consider the summation when
$ {\mathrm {Disc}}(F_1)$
is away from
$X^{1/3}$
using results established in Section 3. In Section 5.2, we will consider the summation when
$ {\mathrm {Disc}}(F_1)$
is near
$X^{1/3}$
using results established in Section 2 and 4.
Taking
$Y=1$
in Theorem 5.1, we get the following immediate corollary, essentially an analogue of Corollary 3.2, which will be useful in the induction argument to come.
Corollary 5.5. Let k be a number field. For each
$m \geq 2$
, let
$\alpha _m$
be the constant from Lemma 5.1. Then for any
$X \geq 1$
, we have

5.1 Base case: noncritical range
In this subsection, we will consider the following summation:

where
$\mathfrak {N}(X,Y) := [Y, X^{1/3-\delta _0}) \cup (X^{1/3+\delta _0}, X^{1/2}]$
for an arbitrary small number
$\delta _0$
satisfying
${0<\delta _0<1/6}$
. We regard this range for
$F_1$
as noncritical because the results of Section 3 together with the following lemma on the
$2$
-class number of relative
$2$
-extensions will give the desired bound (Lemma 5.7) in this range.
Lemma 5.6 ([Reference Klüners and WangKW20], Theorem
$2.7$
).
Let k be a number field and let
$F/k$
be a
$2$
-extension. Then for any
$\epsilon>0$
, we have
$h_2(F) \ll _{[F:\mathbb {Q}],\epsilon } \mathrm {Disc}(F)^\epsilon h_2(k)^{[F:k]}$
.
Lemma 5.7. Let k be a number field and let
$\delta _0 \in (0, 1/6)$
. There exists
$\delta>0$
, only depending on
$\delta _0$
, and an absolute constant
$\alpha>0$
such that for all
$X,Y>0$
, we have

where the summation is over all towers
$F_2/F_1/F_0=k$
with
$F_2/F_1$
and
$F_1/F_0$
being relative quadratic extensions and where
$\mathfrak {N}(X,Y) = [Y, X^{1/3-\delta _0}) \cup (X^{1/3+\delta _0}, X^{1/2}]$
.
Proof of Lemma 5.7.
Writing
$D := {\mathrm {Disc}}(F_1)$
and
$h:=h_2(F_1)$
, from Section 3 (and the translation from relative to absolute discriminant), we have

Specifically, the first line follows from Proposition 3.4 and (3.8), the second from Proposition 3.15, and the third from the final case of Theorem 3.1.
We will use the first line of (5.5) for
$ {\mathrm {Disc}}(F_1) \in (X^{1/3+\delta _0}, X^{1/2}]$
, the second line for
$ {\mathrm {Disc}}(F_1) \in [Y_0, X^{1/3-\delta _0})$
, and the third line for
$ {\mathrm {Disc}}(F_1) \in [Y, Y_0)$
, where we choose a value for
$Y_0$
so that we can apply the third line to the range
$[Y, Y_0)$
for every
$F_1/k$
. In particular, if
$C_{[K:{\mathbb {Q}}]}$
is such that we always have
$h(K)^{2/3}\leq C_{[K:{\mathbb {Q}}]} {\mathrm {Disc}}(K)$
, we can choose
$Y_0=C_{2[k:{\mathbb {Q}}]}^{-1/6}X^{1/6}$
. Applying (5.5) as described and applying Lemma 5.6 to bound h, we find

We will now explain the final inequality in more detail. From each of the four sums we can factor out
$ {\mathrm {Disc}}(k)^{-1+\epsilon '}h_2(k)^{8/3}\ll _{[k:{\mathbb {Q}}],\epsilon '} {\mathrm {Disc}}(k)^{1/3+2\epsilon '}$
, by the trivial bound on
$h_2(k)$
. Then the remaining factors are each of the form

for some
$\gamma , \beta , A, B$
. Corollary 3.2, translating to absolute discriminant, gives, for
$B>1$
,

In the second sum, since
$\beta>0$
, we can bound
$ {\mathrm {Disc}}(F_1)^{1/2+\epsilon '}<X^{(1/3-\delta _0)(1/2+\epsilon ')}$
, and use the bound from (5.8) to find that the remaining factor is
$\ll _{\epsilon ', [k:{\mathbb {Q}}]}X^{1-3\delta _0/2+\epsilon '}$
. In the first, third, and fourth sums, where
$\beta <-1$
, we use partial summation and apply (5.8), to find that (5.7) is
$ \ll _{\epsilon ', [k:{\mathbb {Q}}]} X^\gamma A^{\beta +1}. $
In particular the remaining factor in the first sum is
$ \ll _{\epsilon ', [k:{\mathbb {Q}}]} X^{1-3\delta _0/2+\epsilon '}, $
the remaining factor in the third sum is
$ \ll _{\epsilon ', [k:{\mathbb {Q}}]} XY_0^{-1/2+\epsilon '} \ll _{\epsilon ',\epsilon , [k:{\mathbb {Q}}]}X ^{11/12+\epsilon '}, $
and the remaining factor in the fourth sum is
$ \ll _{\epsilon ',\epsilon , [k:{\mathbb {Q}}]} XY^{-1+\epsilon }. $
We now would like to choose
$\delta ,\alpha $
and
$\epsilon '$
such that the final bound in (5.6) is smaller than the bound we are trying to prove, that is, we want

We can choose, for example,
$\alpha =2/5$
and
$\delta =\min \{ \delta _0,1/15\}$
. Above if we take
$\epsilon '=\min \{\delta _0/2,1/60\}$
, then the inequality above and the lemma follows.
5.2 Base case: critical range
In this section, we will consider the following summation:

where
$\mathfrak {C}(X,Y) = \mathfrak {C}(X)= [X^{1/3-\delta _0}, X^{1/3+\delta _0}]$
, which we regard as the critical range of
$ {\mathrm {Disc}}(F_1)$
. In this range, summing the best bound for each
$F_1$
is not sufficient, and we need to extract an additional saving from the sum over
$F_1$
. We will apply the results developed in Sections 2 and 4 to obtain this saving.
It will be convenient to use the following simple bound on the number of relative quadratic extensions.
Lemma 5.8. Let k be a number field. Then

Proof. By class field theory, the number
$N_k(C_2, X)$
is bounded by the product of
$h_2(k)$
and the number of integral ideals with bounded norm. The latter can be counted by integrating
$\zeta _k(s)$
. If
$a_n$
denotes the number of integral ideals of norm n, then by Perron’s formula we have

Shifting the contour integral to
$\Re (s)=1-\epsilon $
and using the convexity bound (3.6) on
$\zeta _k(s)$
, we get the upper bound

where the upper bound on the residue of the Dedekind zeta function comes from Landau.
Lemma 5.9. Let k be a number field and let
$\delta _0 \in (0, 1/100]$
. There exists an absolute positive constant
$\delta $
, and a value
$\alpha $
, depending only on
$[k:{\mathbb {Q}}]$
, such that for all
$X\ge 1$
we have

where the summation is over all towers
$F_2/F_1/F_0=k$
with
$F_2/F_1$
and
$F_1/F_0$
being relative quadratic extensions and
$\mathfrak {C}(X) = [X^{1/3-\delta _0}, X^{1/3+\delta _0}]$
.
Proof. We first give the idea of the proof in broad terms to motivate the notation that is to follow. Exploiting the factorization
$h_3(F_2/k) = h_3(F_2/F_1) h_3(F_1/k)$
, Corollary 3.2 provides strong control on the average of
$h_3(F_1/k)$
, while Lemma 2.1 provides a nontrivial bound on
$h_3(F_2/F_1)$
if there are “enough” small primes that split in
$F_2/F_1$
. Lemma 4.4 ensures there are enough small primes in k. Applying Lemma 4.3 twice, we see first that “most” extensions
$F_1/k$
have roughly the same number of small primes as k, and then second that in “most” extensions
$F_2/F_1$
for such
$F_1$
, many of those small primes are split. This lets us apply Lemma 2.1, and yields the lemma.
To make this precise, and in preparation for applying Lemma 4.3, we let
$\delta _1=1/20$
and
$\epsilon _1=1/30$
. We set
$c = \max \{c_{[k:{\mathbb {Q}}]}, c_{2[k:{\mathbb {Q}}]},1\}$
to be at least as large as the two constants given in Theorem 4.2 and
$\sigma _1 = 1 - \epsilon _1/4c$
. Throughout the proof, we will make assumptions on X, including that X is at least any absolute constant necessary to apply the results we use, and then we will handle the remaining values of X at the end of the proof. Applying Lemma 4.3 the first time, we obtain an exceptional set
$\mathcal {E}_0 =\mathcal {E}(k, X^{1/3+\delta _0}, \epsilon _1)$
. Then for each
$F_1/k\notin \mathcal {E}_0$
, we apply Lemma 4.3 a second time to obtain an exceptional set
$\mathcal {E}(F_1) = \mathcal {E}(F_1, X/ {\mathrm {Disc}}(F_1)^2, \epsilon _1)$
. By Lemma 4.3, both exceptional sets contain few elements:

We now consider the consequences of Lemma 4.3 for fields outside the exceptional sets. For
$F_1/k \notin \mathcal {E}_0$
with
$ {\mathrm {Disc}}(F_1) \in \mathfrak {C}(X)$
and
$x =(X/(2 {\mathrm {Disc}}(F_1)^2))^{\delta _1} \leq X^{1/3+\delta _0}$
, by Lemma 4.3 we have, for X sufficiently large,

We next assume that
$X\ge 2^{\frac {1}{1/3-2/100}} 4^{\frac {1}{(1/3-2/100)\delta _1}} {\mathrm {Disc}}(k)^A$
where
$A =\frac { \max \{ \beta , 8c\gamma /\epsilon _1 \} }{(1/3-2/100) \delta _1} = A([k:{\mathbb {Q}}])$
, and
$\gamma $
and
$\beta $
are absolute constants for which Lemma 4.4 holds. This assumption ensures that
$x/4 \geq {\mathrm {Disc}}(k)^\beta $
and that
$ x^{\frac {1-\sigma _1}{2}}\geq {\mathrm {Disc}}(k)^\gamma $
. Let also
$D_0$
be an absolute constant such that Lemma 4.4 holds for those
$\gamma $
and
$\beta $
. If
$ {\mathrm {Disc}}(k) \geq D_0$
, we use Lemma 4.4, and for
$X\ge {\mathrm {Disc}}(k)^A$
and X sufficiently large given
$[k:{\mathbb {Q}}]$
, we obtain

Here we used
$x/4 \geq {\mathrm {Disc}}(k)^\beta $
to apply Lemma 4.4, and
$ x^{\frac {1-\sigma _1}{2}}\geq {\mathrm {Disc}}(k)^\gamma $
as a somewhat generous assumption to ensure that the lower bound from Lemma 4.4 is sufficiently large to subtract off the term above and still, up to constants, obtain the same lower bound. We require that X is large enough so that (5.15) holds also for the finitely many fields k with
$ {\mathrm {Disc}}(k) \leq D_0$
. Under these assumptions, we conclude for
$F_1 \notin \mathcal {E}_0$
that

Next, for
$F_2/F_1\notin \mathcal {E}(F_1)$
, by Lemma 4.3 and (5.16), and the assumptions above on X, we have

We now consider the contribution to the summation from the various kinds of exceptional and nonexceptional fields. First, we consider the summation
$F_2/F_1/k$
where
$F_1/k\in \mathcal {E}_0$
. Using the trivial bound on the class group and the sparsity of exceptional fields along with Lemmas 5.8 and 5.6, we find

Second, we consider the contribution from those
$F_2/F_1/k$
where
$F_1 \notin \mathcal {E}_0$
but
$F_2/F_1\in \mathcal {E}(F_1)$
. Using the trivial bound on relative class group
$h_3(F_2/F_1)$
and an average bound on
$h_3(F_1/k)$
, we find

Precisely, the first inequality follows from Lemma 2.4. The second inequality follows from the estimate
$|\mathcal {E}(F_1)|\ll _{[k:{\mathbb {Q}}], \epsilon _1} {\mathrm {Disc}}(k)^{\epsilon _1} X^{\epsilon _1}$
from (5.13). The third inequality follows from Corollary 3.2 and partial summation.
We now consider the contribution from those
$F_2/F_1/k$
where neither
$F_1/k \notin \mathcal {E}_0$
nor
$F_2/F_1 \notin \mathcal {E}(F_1)$
. Using the relative Ellenberg–Venkatesh method developed in Section 2 and existence of small split primes for nonexceptional fields developed in Section 4 and made explicit in (5.17), we find,

Precisely, the first inequality comes from Lemma 2.1 with
$\theta =\delta _1=1/20$
and and (5.17) with
${x= {\mathrm {Disc}}(F_2/F_1)^{\delta _1}}$
. The second inequality follows from Lemma 5.8 and Lemma 5.6. We apply Corollary 3.2 and partial summation for the last inequality.
Finally, in the complementary case that X is not sufficiently large, we may assume that
$X \ll _{[k:\mathbb {Q}]} {\mathrm {Disc}}(k)^A$
, so that by applying the trivial bound
$h_3(F_2/k) \ll _{[k:\mathbb {Q}],\epsilon } X^{1/2+\epsilon }$
, we find

Here the second inequality follows from Lemma 5.8 twice, and the last inequality follows from the bound on X.
So to deduce the statement of the lemma, we note we have four upper bounds of the form
$ {\mathrm {Disc}}(k)^{a_i} X^{b_i+\epsilon }$
. We can take, for example,
$\epsilon =1/1000$
, and note that each of the
$b_i$
is less than
$998/1000$
. Then we can take
$\delta =1/1000$
, and
$\alpha $
can be taken to be
$\max _i{a_i}$
, which depends on
$[k:{\mathbb {Q}}]$
.
5.3 Induction
Finally, in this section, we will prove Theorem 5.1 in general. We will use an induction argument with the initial case
$m=2$
.
Proof of Theorem 5.1.
The statement for
$m = 2$
follows from the combination of Lemmas 5.7 and 5.9. Now assume for some i that we can prove the lemma for every
$m\le i$
. Our goal is to prove the lemma for
$m = i+1$
.
Firstly, we treat the summation when
$ {\mathrm {Disc}}(F_i) \in \mathfrak {N}(X,Y)=[Y, X^{1/3-\delta _0}) \cup (X^{1/3+\delta _0}, X^{1/2}]$
, the noncritical range. We use an argument similar to that in Lemma 5.7 where the noncritical range is treated for
$m=2$
. For any
$0< \delta _0 < 1/6$
and
$Y \geq 1$
, using the results on counting cubic fields in Section 3, as collected in (5.5), and the induction hypothesis in the form of Corollary 5.5, we find

The inequality above follows as in (5.6), with
$F_i$
in place of
$F_1$
, using the cases of (5.5) in three ranges of
$ {\mathrm {Disc}}(F_i)$
followed by partial summation with Corollary 5.5 (in place of Corollary 3.2), and the
$h_2(k)^2$
factor from bounding
$h_2(F_1)$
is replaced by
$h_2(k)^{2^i}$
in the bound for
$h_2(F_i)$
, using Lemma 5.6.
We now treat the contribution from towers with
$ {\mathrm {Disc}}(F_i)\in \mathfrak {C}(X) = [X^{1/3-\delta _0},X^{1/3+\delta _0}]$
. We separate the discussion into two cases depending on the size of
$ {\mathrm {Disc}}(F_{i-1})$
. Fix
$1 \leq Y_2 \leq X^{\frac {1}{6}-\frac {\delta _0}{2}}$
. Then, we find

Precisely, the first inequality follows from (3.8), Proposition 3.4, and Lemma 5.6. The second inequality follows from partial summation, and the induction hypothesis with
$m=i$
applied to
$F_i/\cdots /F_1/k$
with
$Y_2$
being the lower bound for
$ {\mathrm {Disc}}(F_{i-1})$
. The third inequality follows from the trivial bound on
$h_2(k)$
.
Finally we consider those towers with
$ {\mathrm {Disc}}(F_{i-1}) \le Y_2$
. Let
$\delta ^{\prime }_2$
and
$\alpha ^{\prime }_2$
be the constants associated to
$[F_{i-1}:{\mathbb {Q}}]$
and
$m =2$
. We have

Here precisely, the first inequality follows from the induction hypothesis with
$m=2$
and base field
$k = F_{i-1}$
. The second inequality follows from the induction hypothesis with
$m=i-1$
and taking
$\epsilon = 1/12$
.
If we let
$Y_2=X^b$
, then to obtain the desired bound, (5.23) requires us to choose
$\delta _0$
small, given b and
$\delta _i$
, and (5.24) requires us to take b small given
$\delta _2'$
and
$\alpha _2'$
(as long as we have taken
$\delta _0$
absolutely sufficiently small). Since we can do this by choosing b and then
$\delta _0$
, we will be able to obtain the theorem. In particular, making the convenient but nonoptimal choices
$b=\min \{\delta ^{\prime }_2/(2(1+\alpha ^{\prime }_2)),1/(8(1+\alpha ^{\prime }_2))\}$
and
$Y_2 = X^{b}$
and
$\delta _0 = \min \{\delta _i/9, b/2,1/16\}$
, we conclude that the statement of the theorem holds for
$m=i+1$
with
$ \alpha _{i+1} = \max \{\alpha _i+2^{i}, \alpha _{i-1}\} \quad \text {and} \quad \delta _{i+1} = \min \{\frac {\delta ^{\prime }_2}{2}, \frac {b}{8},\frac {\delta _i}{36},\frac {1}{32}\}. $
6 Average of
$3$
-torsion in class groups of
$2$
-extensions
In this section, we will prove a refined version of Theorem 1.1, in Theorem 6.1 below, using crucially the tail estimates from Section 5. We use the same approach to prove Theorem 1.2.
It is well-known that the statistics of class groups of imaginary versus real quadratic fields are quite different, and of course the Cohen-Lenstra-Martinet heuristics reflect this. So for higher degree number fields we also expect that the class group statistics will differ based on the behavior at infinity of the fields, though perhaps it is less clear which aspects of the behavior at infinity need to be specified. For their heuristics, Cohen and Martinet separated number fields by a certain function of the Galois module structure of the units in the Galois closure (see [Reference Cohen and MartinetCM90, Définitions 6.4 and 2.2]). By [Reference Cohen and MartinetCM90, Théorème 6.7] their separation depends only on the structure of the induced representations of the trivial representation from the decomposition groups at infinity to the Galois group. For us, it is simpler to separate number fields by the conjugacy classes of the decomposition groups themselves. This data determines the data Cohen and Martinet use at infinity to separate number fields of the same Galois group, and slightly refines it. We define the group signature
$\Sigma $
of an extension of number fields
$K/k$
to be the ordered tuple
$(\Sigma _v)_v$
, where v ranges over the real places of k, and
$\Sigma _v$
is the conjugacy class of complex conjugation in
$ {\mathrm {Gal}}(K/k)$
over v. Let
$E_k(m,X)$
be the set of degree
$2^m 2$
-extensions
$K/k$
(in
$\bar {{\mathbb {Q}}}$
) with
$ {\mathrm {Disc}} K\leq X$
. Let
$E_k(G,X)$
be the set of G-extensions
$K/k$
(in
$\bar {{\mathbb {Q}}}$
) with
$ {\mathrm {Disc}} K\leq X$
and let
$E_k^\Sigma (G,X)$
be those with group signature
$\Sigma $
.
Theorem 6.1. Let k be a number field and m a positive integer. Then there exists
$C_m>0$
such that

Moreover, let
$G\subset S_{2^m}$
be a transitive permutation
$2$
-group containing a transposition and
$\Sigma $
a group signature that occurs for some G-extension of k. Then there exists
$C_{G,\Sigma },C_G>0$
such that

The proof of Theorem 6.1 gives the constants explicitly. We let
$r_1(F)$
denote the number of real places of F, and
$r_2(F)$
the number of pairs of complex places of F. We have

where the summation runs over those
$F \in E_k(m-1,\infty )$
. When the group G has a transposition, Lemma 6.6 below shows that
$G=C_2\wr H$
for some transitive subgroup
$H \subseteq S_{2^{m-1}}$
. The image of a group signature
$\Sigma $
for G gives a group signature
$\bar {\Sigma }$
for H. So a G-extension
$K/k$
of signature
$\Sigma $
has an index
$2$
subfield
$F/k$
which is an H-extension of signature
$\bar {\Sigma }$
, and there is a number
$u(\Sigma )$
, depending only on
$\Sigma $
, which gives the number of infinite places of F that are split in K, or equivalently,
$ {\mathrm {rk}}\mathcal {O}_K^*- {\mathrm {rk}} \mathcal {O}_F^*$
, the relative unit rank of
$K/F$
. Then we have

Before proving Theorem 6.1, we first give a result on the asymptotic number of
$2$
-extensions, that is, the denominator in Theorem 6.1.
Theorem 6.2. Let k be a number field and let m be a positive integer. There exists
$D_m>0$
such that

Moreover, let
$G\subset S_{2^m}$
be a transitive permutation
$2$
-group containing a transposition and
$\Sigma $
a group signature. Then there exists
$D_{G,\Sigma }\geq 0$
and
$D_G>0$
such that

If there exists an H-extension of k with group signature
$\bar {\Sigma }$
, then
$D_{G,\Sigma }>0$
. Further, if
$G\subset S_{2^m}$
is a transitive permutation
$2$
-group not containing a transposition, then
$|E_k(G,X)| = O_{k,G,\epsilon }(X^{1/2+\epsilon })$
.
Proof. Malle [Reference MalleMal02, Reference MalleMal04] conjectured that for
$2$
-groups
$|E_k(G,X)|$
has linear growth if and only if G contains a transposition. When G has no transpositions, it is proved by [Reference Klüners and MalleKM04, Corollary 7.3] that
$|E_k(G,X)| = O_{k,G,\epsilon }(X^{1/2+\epsilon })$
. When G contains a transposition, by [Reference KlünersKlü12, Lemma
$5.5$
] G must be isomorphic to
$C_2\wr H$
for some permutation
$2$
-group H. It is proved in [Reference KlünersKlü12, Theorem
$5.8$
] that when there exists a H-extension and
$|E_k(H,X)|$
is not growing too fast, Malle’s conjecture
$|E_k(G,X)| \sim D_G X$
holds for
$G = C_2 \wr H$
. The existence of
$2$
-group extensions is shown by a celebrated theorem of Shafarevich [Reference ŠafarevičŠ54], and the upper bound for
$|E_k(H,X)|$
is known for all permutation
$2$
-groups H, again by [Reference Klüners and MalleKM04, Corollary 7.3]. This gives the first statement of the theorem and a version of the second statement without signature conditions. Following Klüners’ proof, or a simpler version of the proof of Theorem 6.1, one can obtain a version with the signature condition and also with the precise constants


where
$\bar {\Sigma }$
is the group signature on H given from
$\Sigma $
via the quotient map
$G\rightarrow H$
, and
$p_\Sigma $
is the proportion of the
$2^{r_1(F)}$
possible behaviors at infinity for a quadratic extension
$K/F$
such that when K is a G-extension with such behavior then it has group signature
$\Sigma $
. (We note that
$p_\Sigma $
is determined group theoretically from
$\Sigma $
and is nonzero – see the proof of Theorem 6.1.)
Theorem 6.2 shows that, among
$2$
-extensions of a given degree, the fields with Galois groups without transpositions are thin families. Thus, the restriction in Theorem 6.1 to groups G with transpositions is natural. In fact, while Theorem 6.1 does not give class number averages for the thin families of extensions with Galois group G for G without transpositions, a key step in its proof is to show that class group elements arising from such extensions give a negligible contribution to the average in (6.1). This is a result of independent interest that complements Theorem 6.1. Thus, we begin by considering thin families in Section 6.1, which culminates in Theorem 6.5 that makes this discussion precise. Then, in Section 6.2, we turn to the proof of Theorem 6.1.
6.1 Thin families of
$2$
-extensions
In this section, we focus on studying
$3$
-torsion in class groups of
$2$
-extensions with a permutation Galois group G without a transposition. We begin by proving an upper bound on the count of these extensions, but with explicit base field discriminant dependence. We first give the following lemma, which is a uniform version of Theorem
$1.6$
in [Reference Klüners and WangKW20] for
$\ell =2$
.
Lemma 6.3. Let k be a number field. Then there exists
$\alpha _m>0$
depending at most on
$[k:{\mathbb {Q}}]$
and m such that the number of degree
$2^m 2$
-extensions
$K/k$
with
$ {\mathrm {Disc}}(K/k) = D$
is bounded by
$O_{[k:{\mathbb {Q}}], m, \epsilon }(D^{\epsilon } {\mathrm {Disc}}(k)^{\alpha _m})$
.
Proof. We proceed by induction over m. For
$m = 1$
, we know from class field theory (see the proof of Lemma 3.5) that the number of relative quadratic extensions
$K/k$
with
$ {\mathrm {Disc}}(K/k) = D$
is at most
$O_{[k:{\mathbb {Q}}], \epsilon }(h_2(k)\cdot D^{\epsilon }) = O_{[k:{\mathbb {Q}}], \epsilon }( D^{\epsilon } {\mathrm {Disc}}(k)^{1/2+\epsilon })$
. Now assuming the statement is true for
$m =i$
, we will prove that it also holds for
$m = i+1$
. By Lemma 5.4, it suffices to count quadratic extensions of degree
$2^i 2$
-extensions
$F/k$
with
$ {\mathrm {Disc}}(F/k)^2| D$
. The number of such F is bounded by
$O_{[k:{\mathbb {Q}}], \epsilon , i}( D^{\epsilon } {\mathrm {Disc}}(k)^{\alpha _i})$
by the induction hypothesis. For each such F, the number of quadratic relative extensions
$K/F$
with
$ {\mathrm {Disc}}(K/F) = D/ {\mathrm {Disc}}(F/k)^2$
is bounded by
$O_{[F:{\mathbb {Q}}], \epsilon }(h_2(F) \cdot D^{\epsilon })$
from class field theory. Applying Lemma 5.6 and summation over all divisors of D, we obtain the upper bound
$O_{[k:{\mathbb {Q}}], i+1, \epsilon }( D^{\epsilon } {\mathrm {Disc}}(k)^{\alpha _{i+1}})$
, where
$\alpha _{i+1}$
can be taken as
$\alpha _{i}+ 2^{i-1}$
.
Theorem 6.4. Let k be a number field k and let
$G \subseteq S_{2^m}$
be a transitive permutation
$2$
-group without a transposition. Then there exists
$\alpha>0$
depending at most on
$[k:{\mathbb {Q}}]$
and m such that

Proof. Notice that when G does not contain a transposition, the discriminant ideal
$ {\mathrm {disc}}(K/k)$
must have exponent at least
$2$
at every ramified prime, that is, it is powerful. (The exponent of a prime in
$ {\mathrm {disc}}(K/k)$
can be computed from the exponent in the different of a prime above in the Galois closure, which is the Artin conductor of the permutation representation [Reference NeukirchNeu99, VII.11.8]. In particular, the exponent of a ramified prime in
$ {\mathrm {disc}}(K/k)$
is at least
$2^m$
minus the number of orbits of the inertia subgroup, and the number of orbits is
$2^m - 1$
if and only if the inertia subgroup is cyclic and generated by a transposition.) Thus, its norm,
$ {\mathrm {Disc}}(K/k)$
, must be a powerful integer, and there are
$O(X^{1/2})$
powerful integers below X. The result now follows from Lemma 6.3.
The
$\ell $
-torsion conjecture then would imply that the summation of
$| {\mathrm {Cl}}_K[3]|$
for G-extensions K without a transposition should be also thin, that is,
$O(X^{1/2+\epsilon })$
. By applying the trivial bound for the class group
$h_3(K)= O_{[K:{\mathbb {Q}}], \epsilon }( {\mathrm {Disc}}(K)^{1/2+\epsilon })$
, we can conclude immediately that
$\sum _{K\in E_k(G,X)} h_3(K)= O_{k, \epsilon }(X^{1+\epsilon })$
. However, it is crucial for our main theorem that we prove something slightly better than this.
Theorem 6.5. Let k be a number field and let
$G\subset S_{2^m}$
be a transitive permutation
$2$
-group without a transposition. Then there exist
$\delta>0$
and
$\alpha $
depending on G and
$[k:{\mathbb {Q}}]$
such that

Proof. A G-extension K can be constructed as a relative quadratic extension over a degree
$2^{m-1} 2$
-extension F with
$ {\mathrm {Gal}}(F/k) =: H\subset S_{2^{m-1}}$
by Lemma 5.4. We fix
$\delta _1=1/24$
and
$\epsilon _1=1/4$
. Let
$\gamma $
,
$\beta $
and
$D_0$
be absolute constants allowable in Lemma 4.4. Let
$Y = X^{\delta '}$
for some
$\delta '$
that we will choose sufficiently small in terms of m and
$[k:{\mathbb {Q}}]$
. Then by Theorem 5.1 and dyadic summation, taking
$\delta '\leq \delta _m$
, we have

Therefore it suffices to consider the summation over G-extensions where the associated H-extension
$F/k$
has
$ {\mathrm {Disc}}(F)\le Y = X^{\delta '}$
.
We will apply an argument similar to Section 5.2 to complete the proof. For each F, denote the set
$\mathcal {E}(F) = \mathcal {E}(F, X/ {\mathrm {Disc}}(F)^2, \epsilon _1)$
of exceptional quadratic extensions as given by Lemma 4.3, for
$X \gg 1$
. The set has size
$|\mathcal {E}(F)| \ll _{[F:{\mathbb {Q}}],\epsilon _1 } X^{\epsilon _1}$
. Let
$x = (X/(2Y^2))^{\delta _1}$
. We assume X is sufficiently large in terms of
$[k:\mathbb {Q}]$
and
$\delta '$
sufficiently small in terms of our absolute constants, so that for
$\sigma _1 = \max \{1- \epsilon _1/4c,1/2\}$
with
$c = c_{[F:{\mathbb {Q}}]}$
in Theorem 4.2 and F with
$ {\mathrm {Disc}}(F) \geq D_0$
, we may apply Lemmas 4.3 and 4.4 in concert to conclude for
$K\not \in \mathcal {E}(F)$
,

We also assume that X is sufficiently large that this holds for the finite number of fields F with
$ {\mathrm {Disc}}(F) \leq D_0$
.
Using the trivial bound for
$h_3(K/k)$
, the summation over
$K/F/k$
with
$K/F\in \mathcal {E}(F)$
is

where the last inequality follows from combining the bound on
$|\mathcal {E}(F)|$
with an upper bound on the number of such F, for example as is provided by Corollary 5.5 and dyadic summation.
For fields
$K/F \notin \mathcal {E}(F)$
, we use Lemma 2.1 with
$\theta =\delta _1$
and (6.5) with
$x= {\mathrm {Disc}}(K/F)^{\delta _1}$
to show
$h_3(K/F) \ll _{[k:\mathbb {Q}],m,\epsilon } X^{1/2-\delta _1+\epsilon } {\mathrm {Disc}}(F)^{-1/2+2\delta _1+\gamma }$
. Combining with the trivial bound on
$h_3(F/k)$
, we have
$h_3(K/k) \ll _{[k:\mathbb {Q}],m,\epsilon } X^{1/2-\delta _1+\epsilon }Y^{2\delta _1+\gamma }$
. Then we find

by Theorem 6.4, where
$\alpha $
is associated to
$[k:{\mathbb {Q}}]$
in Theorem 6.4. We can use dyadic summation to remove the lower bound on
$ {\mathrm {Disc}}(K)$
and obtain the same bound (with a different implied constant).
Finally, for
$X\ll _{[k:{\mathbb {Q}}],m,\delta '} 1$
, we have

by Corollary 5.5. Choosing
$\delta '$
and sufficiently small in terms of
$[k:{\mathbb {Q}}]$
and m, we conclude the theorem.
6.2 Proof of the main theorem
Now we are ready to prove Theorem 6.1. We start with the following lemmas that allow us to move from summing over G-extensions to summing over H-extensions, where
$G\simeq C_2\wr H$
.
Lemma 6.6. A transitive permutation
$2$
-group
$G\subset S_{2d}$
contains a transposition if and only if
$G\simeq C_2\wr H$
for some
$H\subset S_{d}$
. For such a G, if
$K/k$
is a G-extension, then there exists a unique subfield
$F/k$
with
$[K:F] = 2$
. Moreover,
$F/k$
is an H-extension. For such a G, there is only one permutation group H up to isomorphism such that
$G\simeq C_2\wr H$
.
Proof. The first statement is Lemma
$5.5$
in [Reference KlünersKlü12]. By Galois theory, an F as in the second statement corresponds to a subgroup M of G that contains
$ {\mathrm {Stab}}_G(1)$
as an index
$2$
subgroup. We claim that if M is a subgroup of G and contains
$ {\mathrm {Stab}}_G(1)$
as a proper subgroup, then
$C_2^d\rtimes {\mathrm {Stab}}_H(1)\subset M$
. To show the claim, we write
$ {\mathrm {Stab}}_G(1) = C_2^{d-1}\rtimes {\mathrm {Stab}}_H(1)$
where
$C_2^{d-1}$
is the subspace of
$C_2^d$
with first entry
$0$
. Suppose
$g=v\rtimes h\in M$
and
$g\notin {\mathrm {Stab}}_G(1)$
. If
$h\in {\mathrm {Stab}}_H(1)$
, then
$v_1\neq 0$
, and multiplication by all
$u\rtimes h^{-1}\in {\mathrm {Stab}}_G(1)$
shows that
$C_2^d \subset M$
. If
$h\notin {\mathrm {Stab}}_H(1)$
, then for
$u\in C_2^{d-1}$
with a
$1$
in position
$h^{-1}(1)$
and
$0$
’s elsewhere, we have that
$(v\rtimes h) u (v\rtimes h)^{-1}\in C_2^d$
with a
$1$
in position
$1$
and
$0$
’s elsewhere, so we conclude the claim. Since
$ {\mathrm {Stab}}_G(1) = C_2^{d-1}\rtimes {\mathrm {Stab}}_H(1)$
is index
$2$
in
$C_2^d\rtimes {\mathrm {Stab}}_H(1)$
, there is a unique subgroup of G that contains
$ {\mathrm {Stab}}_G(1)$
as an index
$2$
subgroup, showing the second statement. The Galois closure of F, via Galois theory, corresponds to the intersection of the conjugates of
$C_2^d\rtimes {\mathrm {Stab}}_H(1)$
, which is
$C_2^d$
. Thus is follows that
$ {\mathrm {Gal}}(F/k)$
is given by the action of
$G/C_2^d$
on
$(C_2^d\rtimes {\mathrm {Stab}}_H(1))/C_2^d$
, which is the permutation group H, giving the third statement. If
$G \simeq C_2\wr H$
, then H acts on
$2$
-element blocks of the elements that G acts on, and all elements of G preserve the block structure. From this, it follows that two elements are together in a block if and only if G contains a transposition that interchanges them. From this it follows that H is determined uniquely as a permutation group.
We now want to be precise about exactly what we mean by permutation group isomorphisms.
Definition 6.7. A permutation group G is a group G, a set
$B_G$
, and a faithful action of G on
$B_G$
. Given a set X, we write
$\operatorname {Sym} X$
for the group of permutations of the elements of X. Thus a permutation group G is naturally a subgroup of
$\operatorname {Sym} B_G$
. The degree of G is
$|B_G|$
. An isomorphism of permutation groups
$(G,B_G)$
and
$(H,B_H)$
is a pair
$(\phi ,\pi )$
where
$\phi : G \rightarrow H$
is a group isomorphism and
$\pi : B_G \rightarrow B_H$
is a bijection such that for all
$a\in B_G$
and
$g\in G$
we have
$\phi (g)(\pi (a))=\pi (g(a))$
. We write
$ {\mathrm {Aut}}_{\operatorname {perm}}(G)$
for the group of permutation automorphisms of G, which is also the normalizer
$N_{\operatorname {Sym} {B_G}}(G)$
of G in the symmetric group
$\operatorname {Sym} {B_G}$
.
In particular, a G-extension
$K/k$
includes the data of a bijection between the embeddings
$K\rightarrow \bar {k}$
and
$B_G$
. Note that
$B_{C_2\wr H}=\{1,2\}\times B_H$
.
Lemma 6.8. Let G be a transitive permutation
$2$
-group and
$G=C_2\wr H$
, with H a permutation group of degree d. Given an H-extension
$F/k$
, and a quadratic extension
$K/F$
(in
$\bar {{\mathbb {Q}}}$
), then either (1)
$ {\mathrm {Gal}}(K/k)$
does not contain a transposition or (2)
$K/k$
can be realized as a G-extension with
$2^d$
choices of
$ {\mathrm {Gal}}(K/k)\simeq G$
(i.e., choices of
$\{K \rightarrow \tilde {K} \}\rightarrow B_G$
) that are compatible with the permutation group isomorphism
$ {\mathrm {Gal}}(F/k)\simeq H$
(i.e., choice of
$\{F \rightarrow \tilde {F} \}\rightarrow B_H$
) in the quotient. Moreover, any G-extension
$K/k$
arises in this construction from a unique F, one quadratic extension
$K/F$
, and one of the
$2^d$
choices of
$ {\mathrm {Gal}}(K/k)\simeq G$
.
Proof. Given an H-extension
$F/k$
and a quadratic extension
$K/F$
, we have
$2^d$
choices of identification of the embeddings
$K\rightarrow \tilde {K}$
with
$\{1,2\}\times B_H$
that are compatible with the map from the embeddings of F to
$B_H$
. Any of these gives an injection of groups
$ {\mathrm {Gal}}(K/k)\subset C_2\wr H$
compatible with the permutation actions. Any transposition in
$C_2\wr H$
is an element of
$C_2^d$
that is nontrivial in exactly one coordinate. Since H is transitive, this implies that any subgroup of
$C_2\wr H$
containing a transposition and with image in H all of H must contain all of
$C_2^d$
and thus be all of
$C_2\wr H$
. So if
$ {\mathrm {Gal}}(K/k)$
contains a transposition, we see that
$ {\mathrm {Gal}}(K/k)\simeq G$
, and there are
$2^d$
ways of choosing such a permutation group isomorphism that are compatible with the map from the embeddings of F to
$B_H$
.
Given a G-extension
$K/k$
, from Lemma 6.6 we see that it arises in this way from a unique F, and the lemma follows.
We use the following lemma to translate group signature data between our G-extensions and our quadratic extensions of H-extensions.
Lemma 6.9. Let H be a permutation group of degree
$2^{m-1}$
for some positive integer m, and
$G = C_2 \wr H$
the permutation group of degree
$2^m$
. If
$K/k$
is a G-extension, and F is the unique subfield with
$[K:F]=2$
, then the group signature of
$K/k$
is determined by F and the group signature of
$K/F$
.
Let
$M_F$
denote the set of group signatures
$\Sigma _2$
for quadratic extensions of F such that
$K/k$
has group signature
$\Sigma $
when
$K/F$
has group signature
$\Sigma _2$
. Then
$|M_F|$
does not depend on F.
Let
$M_\Sigma :=|M_F|$
. For a G-extension
$K/k$
of group signature
$\Sigma $
, we let
$u(\Sigma )$
be the difference in unit ranks between K and F, which only depends on
$\Sigma $
. Let
$\Sigma _0$
be a group signature for an H-extension
$F/k$
. Then

where the sums are over group signatures
$\Sigma $
for G-extensions.
Proof. We first consider the structure of order
$2$
elements of G, to understand the possible group signatures. For
$g\in G=C_2 \wr H$
, let
$\bar {g}$
denote the image of g in H. The group H acts on a set of
$2^{m-1}$
elements
$B_H$
. The element
$g\in G$
has order dividing
$2$
if and only if
$\bar {g}$
has order dividing
$2$
and the
$C_2$
coordinates of g (of which there is one for each element of
$B_H$
) are constant on
$\bar {g}$
orbits of
$B_H$
. Moreover, for any
$g\in G$
, if
$\bar {g}$
transposes two elements
$a,b$
of
$B_H$
, and
$g'$
is obtained from g by changing the
$C_2$
coordinates at positions
$a,b$
each to the other element of
$C_2$
, then
$g'$
is conjugate to g by the element that is nontrivial in the a coordinate and trivial in all other coordinates. In particular, a conjugacy class
$[g]$
of an order
$2$
element
$g\in G$
is determined by the image of the conjugacy class in H, and for any element h of that conjugacy class in H, the
$C_2$
coordinates of the h fixed points of
$B_H$
.
Next, we need to relate the infinite places of F above a place v of k to the permutation group H. To do this, at each real place v of k we choose an embedding
$i_v: \bar {{\mathbb {Q}}}\rightarrow {\mathbb {C}}$
that restricts to v on k. If F is an H-extension, the elements
$B_H$
correspond to embeddings
$\tau : F\rightarrow \tilde {F}$
(as part of the data of the H-extension), and via
$i_v:\bar {{\mathbb {Q}}}\rightarrow {\mathbb {C}}$
these correspond to embeddings
$i_v\circ \tau : F\rightarrow {\mathbb {C}}$
. So for each real place v of k, we now have a correspondence between
$B_H$
and the complex embeddings of F that restrict to
$i_v$
on k. With these choices of
$i_v$
, we can now specify an actual element of H corresponding to complex conjugation over v (as opposed to a conjugacy class). Precisely, we define
$\sigma _v(F)$
to be the element of H that acts on the embeddings
$\tau : F\rightarrow \tilde {F}$
so that
$\sigma _v(F)(\tau )=i_v^{-1} \circ \overline {i_v\circ \tau }$
, where the bar denotes complex conjugation, that is,
$\sigma _v(F)$
is the pullback of complex conjugation via
$i_v$
. So pairs of complex embeddings corresponding to a single place of F over v correspond to elements of
$B_H$
transposed by
$\sigma _v(F)$
, and real places of F over V correspond to fixed points of
$\sigma _v(F)$
. Similarly, for a G-extension
$K/k$
of group signature
$\Sigma $
, we have
$\sigma _v(K) \in {\Sigma }_v$
and the analogous correspondence between places of K above v and
$\sigma _v(K)$
orbits of
$B_G$
. If
$F/k$
is the index two subfield of K, then at the fixed points of
$\sigma _v(F)$
on
$B_H$
, the
$C_2$
coordinate of
$\sigma _v(K)$
is trivial when
$K/F$
is split at the corresponding real place of F (i.e., K has two real places above) and nontrivial when
$K/F$
is ramified at that place (i.e., K has one complex place above). Using the conclusion of the previous paragraph, we conclude that if
$K/k$
is a G-extension with index two subfield
$F/k$
, the conjugacy class of
$\sigma _v(K)$
, and hence the group signature of
$K/k$
, is determined by F and the group signature of
$K/F$
.
Now we consider how many group signatures of the quadratic extension
$K/F$
will lead to group signature
$\Sigma $
for
$K/k$
. If we choose
$\sigma _v\in \Sigma _v$
for all real places v of k, then
$|M_F|$
is the product over v of the number of elements
$w\in C_2^{(B_H^{\sigma _v})}$
such that
$w\sigma _v$
is conjugate to
$\sigma _v$
(where
$B_H^{\sigma _v}$
denotes the elements of
$B_H$
fixed by
$\sigma _v$
). In particular,
$|M_F|$
does not depend on F.
Any number field F has
$2^{r_1(F)}$
possible group signatures for quadratic extensions (choices for a conjugacy class in
$S_2$
over each real place of F). Given an H-extension
$F/k$
with group signature
$\Sigma _0$
, each of these group signatures of quadratic extensions
$K/F$
then determines some group signature
$\Sigma $
of
$K/k$
such that
$\bar {\Sigma }=\Sigma _0$
. Thus the first equation of (6.7) follows.
Given some group signature
$\Sigma $
of a G-extension
$K/k$
, with index
$2$
subfield F, for any
$\Sigma _2\in M_F$
, we have that
$u(\Sigma )$
is the number of split places in
$\Sigma _2$
. Also, note that if we use
$G=S_2$
, we have a definition of
$u(\Sigma _2)$
which is also the number of split places in
$\Sigma _2$
. Thus, for any
$\Sigma _2\in M_F$
, we have
$u(\Sigma _2)=u(\Sigma )$
.
Given an H-extension
$F/k$
with group signature
$\Sigma _0$
, we are interested in summing
$3^{-u(\Sigma _2)}$
over all the over all the group signatures
$\Sigma _2$
of quadratic extensions of F. There are
$\binom {r_1(F)}{i}$
choices of group signatures
$\Sigma _2$
of quadratic extensions of F with i real places of F split, and these have
$u(\Sigma _2)=i+r_2(F).$
Thus

Next we recall the following important result by Datskovsky and Wright [Reference Datskovsky and WrightDW88]. Note that for a quadratic extension, a group signature is just the data of which real places become complex in the extension, which is equivalent to “the
$M_\infty $
-signature” of Datskovsky and Wright [Reference Datskovsky and WrightDW88, Section 1].
Theorem 6.10 (Datskovsky–Wright).
Let F be a number field. For a group signature
$\Sigma $
of a quadratic extension
$K/F$
, let
$u(\Sigma )$
be the number of infinite places of F where
$\Sigma $
is trivial (which is also the relative unit rank of
$K/F$
). Then

and

Proof. This follows from combining Theorems 4.2 and 5.1 in [Reference Datskovsky and WrightDW88].
We now turn to the proof of our main theorem.
Proof of Theorem 6.1.
Suppose
$G \subseteq S_{2^m}$
is a
$2$
-group with a transposition, so by Lemma 6.6,
$G = C_2 \wr H$
for a unique
$H \subseteq S_{2^{m-1}}$
. Let
$\bar {\Sigma }$
be the group signature for H that is the image of
$\Sigma $
. So a G-extension
$K/k$
with group signature
$\Sigma $
has an index two subfield
$F/k$
with the group signature
$\bar {\Sigma }$
. Given an H-extension F, let
$\mathcal {Q}_F^{\Sigma }$
be the set of quadratic extensions
$K/F$
(in
$\bar {{\mathbb {Q}}}$
) of group signature in
$M_F$
, as in Lemma 6.9.
Let
$Y \geq 1$
, and suppose
$X_0 = X_0(Y)$
is sufficiently large so that: 1)
$X_0\geq Y^2$
; 2) and for any
$X\geq X_0$
and each
$F \in E_k(H,Y)$
, we have

Since there are only finitely many fields
$F \in E_k(H,Y)$
, such an
$X_0$
formally exists by Theorem 6.10. By Theorem 5.1 and dyadic summation, we find for any
$X \geq X_0$

where
$F \in E_k^{\bar {\Sigma }}(H,Y)$
is the unique index 2 subfield of K containing k. Then by Lemma 6.8 and Theorem 6.5, we find, for
$X\geq X_0$
,

Then by (6.8), for
$X\geq X_0$
,

and thus also

since we assumed
$X \geq X_0 \geq Y^2$
. Letting
$Y \to \infty $
, the sum over
$F \in E_k^{\bar {\Sigma }}(H,Y)$
converges by virtue of Theorem 6.2, the trivial bounds
$h_3(F) \ll _{[F:\mathbb {Q}],\epsilon } {\mathrm {Disc}}(F)^{1/2+\epsilon }$
and
$ {\mathrm {Res}}_{s=1}\zeta _F(s) \ll _{[F:\mathbb {Q}],\epsilon } {\mathrm {Disc}}(F)^{\epsilon }$
, and partial summation. The error term on the right-hand side also converges, so the left-hand side must converge as well. It follows that

We can then sum over all group signatures and use (6.7) to obtain

To count each field K exactly once (instead of once for each G extension structure), we can divide the above sum by
$| {\mathrm {Aut}}_{\operatorname {perm}}(G)|$
. We then sum over (isomorphism classes of) G of the form
$C_2\wr H$
, and by Lemma 6.6 this will give a sum over
$T_m$
, the set of all (isomorphism classes of) transitive permutation
$2$
-groups H of degree
$ 2^{m-1}$
. Using Theorem 6.5, we obtain

Above, each field F appears in the sum
$| {\mathrm {Aut}}_{\operatorname {perm}}(H)|$
times. Next, we claim that
$| {\mathrm {Aut}}_{\operatorname {perm}}(H)|/ | {\mathrm {Aut}}_{\operatorname {perm}}(C_2\wr H)|=2^{-2^{m-1}}.$
The blocks of
$B_G$
that
$C_2\wr H$
acts on are the pairs that appear in G as transpositions, so these are preserved by any permutation isomorphism of
$C_2\wr H$
, and we have a map
$ {\mathrm {Aut}}_{\operatorname {perm}}(C_2\wr H)\rightarrow {\mathrm {Aut}}_{\operatorname {perm}}(H)$
. It is easy to see this is a surjection and the kernel is given by all the permutations of
$\{1, 2\}\times B_H$
that fix the
$B_H$
coordinates, showing the claim. Thus we have

Combined with Theorem 6.2, this yields the theorem. (Note we may obtain the constants in Theorem 6.2 by the same argument as above, using the count of quadratic extensions of F in Theorem 6.10 in place of the average
$3$
-torsion result.)
6.3 Relative class group averages
For the families in Theorem 6.1, the approach of the proof of Theorem 6.1 also permits us to determine the average size of the
$3$
-torsion subgroup of the relative class group
$ {\mathrm {Cl}}_{K/F}$
, where F is the index two subfield of K, proving Theorem 1.2 (when combined with Proposition 7.1). These averages are particularly nice when the Galois closure group and
$u := r_1(K)+r_2(K)-r_1(F)-r_2(F)$
is fixed. For G a transitive permutation
$2$
-group with a transposition, let
$E_k^u(G,X)$
denote the set of G-extensions
$K/k$
with
$ {\mathrm {Disc}} K\leq X$
and
$K/F$
of relative unit rank u, where
$F/k$
is the unique index
$2$
subfield of K.
Theorem 6.11. Let k be a number field, m a positive integer, and
$G \subseteq S_{2^m}$
a transtive
$2$
-group with a transposition. If u is such that
$E_k^u(G,\infty )$
is nonempty, then for all

Proof. The proof uses the main input we have built for the proof of Theorem 6.1 and is nearly identical to that proof, so we will be brief. Let
$F/k$
be the index
$2$
subfield of K. In particular, using the trivial inequality
$h_3(K/F) \leq h_3(K/k)$
, it follows from Theorem 5.1 that for any
$X,Y>0$
,

Again using the trivial inequality
$h_3(K/F) \leq h_3(K/k)$
, we see from Lemma 6.8 and Theorem 6.5 that

for some
$\delta>0$
depending on m and k. Let r be the number of group signatures of quadratic extensions of F such that
$K/F$
of that signature have
$\mathrm {rk}(\mathcal {O}_K^\times / \mathcal {O}_F^\times ) = u$
. We take
$X_0(Y)$
as in the proof of Theorem 6.1, and have, for
$X\geq X_0$

Proceeding as in the proof of Theorem 6.1, the result follows.
7 Comparison to the Cohen–Lenstra–Martinet heuristics
7.1 Cohen–Lenstra–Martinet prediction for the average of
$h_3(K/F)$
Theorem 6.11 verifies new cases of the Cohen–Lenstra–Martinet heuristics, as we now show.
Proposition 7.1. The Cohen–Lenstra–Martinet heuristics [Reference Cohen and MartinetCM90, Hypothése 6.6] predict that the average of Theorem 6.11 is as proved in the theorem.
Proof. Let G be a transitive permutation
$2$
-group with a transposition and
$S_K$
the stabilizer of an element
$1\in B_G$
. Our
$K\in \mathcal {F}^u(G)$
is then the
$S_K$
fixed field of a Galois G extension
$\tilde {K}/k$
with Galois group G. By Lemma 6.6, we have
$G=C_2 \wr H$
, where H is a permutation group of degree d. In this notation, we have that
$S_K=(1\times C_2^{d-1}) \rtimes {\mathrm {Stab}}_H(1)$
, where
$1\in B_H$
is the image of
$1\in B_G$
. Then we define
$S_F=(C_2^d) \rtimes {\mathrm {Stab}}_H(1)$
(the fixed field of
$S_F$
will be the field we call F, the index
$2$
subfield of K). Let U be the sign representation of
$S_F/S_K$
over
${\mathbb {F}}_3$
.
We claim
$h_3(K/F)=| {\mathrm {Hom}}_{S_F} ( {\mathrm {Cl}}_{\tilde {K}/k},U )|$
, where
$ {\mathrm {Hom}}_{S_F}$
denotes morphisms of
$S_F$
modules. To see this, let
$e\in {\mathbb {F}}_3[G]$
be the element
$e=|S_K|^{-1}{\sum _{g\in S_K} g}$
. We have a map
$m_e: {\mathrm {Cl}}_{\tilde {K}/k}/3 \rightarrow {\mathrm {Cl}}_{K/k}/3$
given by multiplication by e, which is a surjection (as inclusion of ideals gives a section). Since e acts as the identity on U, we have that any element of
$ {\mathrm {Hom}}_{S_F} ( {\mathrm {Cl}}_{\tilde {K}/k},U )$
factors through
$m_e$
, so is determined by a map
$ {\mathrm {Hom}}_{S_F} ( {\mathrm {Cl}}_{K/k}/3,U ).$
Now
$ {\mathrm {Cl}}_{K/k}/3$
is a product of
$\pm 1$
eigenspaces for
$S_F/S_K$
, and
$ {\mathrm {Cl}}_{K/F}/3$
is exactly the
$-1$
eigenspace, which proves the claim.
Let
$V:= \operatorname {Ind}_{{S_F}}^{G} U$
. We claim that V is an irreducible representation of G. If
$h_i$
are coset representatives of
$ {\mathrm {Stab}}_H(1)$
in H, then
$1\rtimes h_i$
are coset representatives for
$S_F$
in G. Then V has a basis
$e_i:= (1\rtimes h_i) e_1$
. The action of G on V is
$(x\rtimes s) e_i = x_{s(i)} e_{s(i)}$
. So on the one hand, it is permuting the
$e_i$
from the permutation action of s, and on the other hand, it can change the sign for any particular basis independently. Thus if
$\sum c_i e_i$
is a nonzero element of V, we can act by some
$g\in G$
to obtain
$gx=e_1$
, and thus we can generate all of V from G actions, which shows that V is irreducible.
So,

In [Reference Wang and WoodWW21, Conjecture 3.5] a conjecture is given for averages over
$ {\mathrm {Cl}}_{\tilde {K}/k}$
in terms of expectation of a certain random variable X, and [Reference Wang and WoodWW21, Proposition 6.6] shows that this conjecture agrees with the Cohen-Lenstra-Martinet heuristics. Then using [Reference Wang and WoodWW21, Theorem 6.2 and Theorem 4.1], we have that the Cohen–Lenstra–Martinet prediction for the average of
$| {\mathrm {Sur}}_G(X ,V)|$
is
$\prod _{v}=|V^{\sigma _v}|^{-1}$
, where the product is over infinite places of the base field k, and
$\sigma _v$
is a decomposition group for v. (Note that we can subdivide our family, which only has a fixed relative unit rank of
$K/F$
, based on their decomposition groups at each of the infinite places of k. We will see that the prediction for each of these subfamilies is the same, and only depends on the relative unit rank of
$K/F$
.) The action of
$\sigma _v$
on the elements that H acts on, or equivalently the cosets
$h_iS_F$
, has fixed points corresponding to the infinite places of F (over v) that are split over k and
$2$
-cycles corresponding to infinite places of F (over v) that are ramified over k. We can write
$\sigma _v$
as generated by
$x_v \rtimes r_v$
. If
$h_i$
corresponds to a fixed point of
$r_v$
, then
$\sigma _v {\mathbb {F}}_3 e_i={\mathbb {F}}_3 e_i$
, and the corresponding infinite place of F is split in K if and only if the ith coordinate of
$x_v$
is trivial, which happens if and only if
${\mathbb {F}}_3 e_i$
is a trivial representation of
$\sigma _v$
. If
$h_i$
and
$h_j$
correspond to a
$2$
-cycle of
$r_v$
, then
${\mathbb {F}}_3 e_i +{\mathbb {F}}_3 e_j$
is the regular representation of
$\sigma _v$
and has one dimension of trivial representation. Thus
$\sum _v \dim _{{\mathbb {F}}_3} V^{\sigma _v}$
is the number of infinite places of F that are not ramified in K, which is exactly the relative unit rank as defined above.
7.2 Cohen–Lenstra–Martinet prediction for the average of
$h_3(K/k)$
Using a similar approach, we can determine the Cohen–Lenstra–Martinet prediction for the average of
$h_3(K/k)$
for G-extensions
$K/k$
of a particular enriched or group signature. In the case when
$G=D_4$
, we give below a numerical computation of the proven average for comparison.
Let G be a transitive permutation
$2$
-group with a transposition and
$S_K$
the stabilizer of an element. We consider the representation
$\operatorname {Ind}_{S_K}^G {\mathbb {F}}_3$
of G over
${\mathbb {F}}_3$
(which is the permutation representation of G). Define
$S_K$
,
$S_F$
,
$U,$
and V as in the proof of Proposition 7.1. Then
$\operatorname {Ind}_{S_K}^G {\mathbb {F}}_3=\operatorname {Ind}_{S_F}^{G} \operatorname {Ind}_{S_K}^{S_F} {\mathbb {F}}_3=\operatorname {Ind}_{S_F}^{G} {\mathbb {F}}_3 \times V.$
Let
$W=\operatorname {Ind}_{S_F}^{G} {\mathbb {F}}_3$
, which is the permutation representation of H. We see that W contains no copies of the irreducible representation V, since the action G on W factors through H and the action of G on V does not.
Suppose that
$W=\prod _i V_i^{a_i},$
where the
$V_i$
are irreducible representations of G (and note they all have G-action that factors through H). We note that by Frobenius reciprocity W contains exactly one copy of the trivial representation, and we let
$W'$
be the quotient of of W by this trivial representation.
We have
$h_3(K/k)=| {\mathrm {Hom}} ( {\mathrm {Cl}}_{K/k},{\mathbb {F}}_3 )|$
. Let
$e,m_e$
be as in the proof of Proposition 7.1. Since any element of
$ {\mathrm {Hom}}_{S_K} ( {\mathrm {Cl}}_{\tilde {K}},{\mathbb {F}}_3 )$
factors through
$m_e {\mathrm {Cl}}_{\tilde {K}/k}= {\mathrm {Cl}}_{K/k}$
, we have a natural bijection between
$ {\mathrm {Hom}} ( {\mathrm {Cl}}_{K/k},{\mathbb {F}}_3 )$
and
$ {\mathrm {Hom}}_{S_K} ( {\mathrm {Cl}}_{\tilde {K}/k},{\mathbb {F}}_3 )|$
. Also
$| {\mathrm {Hom}}_{S_K} ( {\mathrm {Cl}}_{\tilde {K}/k},{\mathbb {F}}_3 )|=| {\mathrm {Hom}}_{G} ( {\mathrm {Cl}}_{\tilde {K}/k},W \times V )|$
. If we let
$e'=|G|^{-1}\sum _{g\in G} g$
, we have that the map
$m_e: {\mathrm {Cl}}_{\tilde {K}} \rightarrow {\mathrm {Cl}}_k$
is a map whose kernel is exactly the relative class group
$ {\mathrm {Cl}}_{\tilde {K}/k},$
and it is also the map that gives the maximal trivial representation quotient of any G representation. Thus
$ {\mathrm {Cl}}_{\tilde {K}/k}$
has no trivial representation part, and
$h_3(K/k)=| {\mathrm {Hom}}_{G} ( {\mathrm {Cl}}_{\tilde {K}/k},W '\times V )|.$
From [Reference Wang and WoodWW21, Theorem 6.2, Theorem 4.1], for a G-representation Z with no trivial component, we have that the Cohen–Lenstra–Martinet prediction for the average of
$| {\mathrm {Sur}}_G( {\mathrm {Cl}}_{\tilde {K}} ,Z)|$
is
$\prod _{v}=|Z^{\sigma _v}|^{-1}$
, where the product is over infinite places v of the base field k, in a family of G-extensions where
$\sigma _v\in G$
is an element of the conjugacy class of complex conjugation over v. Thus the predicted average of
$h_3(K/k)$
is

where the left product is the predicted average of
$h_3(K/F)$
and the right product is the predicted average of
$h_3(F/k)$
(as can be worked out similarly to the above) for families with the corresponding behavior at infinite places.
Let
$F/k$
be the index two subfield of K. The action of
$\sigma _v$
on the permutation basis elements of W has fixed points corresponding to the split places of F (over v) and
$2$
-cycles corresponding to the ramified places of F (over v). Thus
$\prod _v |W^{\sigma _v}|=3^{r_1(F)+r_2(F)}$
and
$\prod _v |(W')^{\sigma _v}|=3^{r_1(F)+r_2(F)-r_1(k)-r_2(k)}$
. If
$W'$
is irreducible, this gives a nice formula for the predicted average of
$h_3(K/k)$
, which is
$(1+3^{-u(K/F)} ) (1+3^{-u(F/k)} )$
, where
$u(K/F)$
is the difference in unit ranks of K and F in the given family (and similarly for
$u(F/k)$
). However, when W is not irreducible, the formula can be more complicated.
Example 7.2. Let
$H=C_2$
, so
$G=D_4=\langle (1234), (24)\rangle $
and
$S_K=\langle (24)\rangle $
and
$S_F=\langle (13),(24)\rangle $
. Here
$W'$
is irreducible, the sign representation of H, and thus the average conjectured the Cohen–Lenstra–Martinet heuristics is
$(1+3^{-u(K/F)} )(1+3^{-u(F/k)} )$
. For example, if
$k={\mathbb {Q}}$
, in the following table we give the averages of
$h_3(K)$
predicted by the Cohen–Lenstra–Martinet heuristics, and a numerical computation of the proven averages from Theorem 6.1, in families with the given group signatures.

Essentially, in the proof of Theorem 6.1, we see that the averages of the
$h_3(K/F)$
and
$h_3(F/k)$
factors are independent, and the
$h_3(K/F)$
has average as predicted by the Cohen–Lenstra–Martinet heuristics. When we order fields in a family up to discriminant X, and take a uniform average and then let
$X\rightarrow \infty $
, we will call this a discriminant-uniform average. A consequence of Cohen and Martinet’s conjecture is that the discriminant-uniform average over G extensions K of
$h_3(F_K/k)$
is the same as the discriminant-uniform average over H extensions F of
$h_3(F/k)$
when the G and H extensions have corresponding behavior at infinite places (see [Reference Wang and WoodWW21, Theorem 9.2]).
In contrast, what we see in the proof of Theorem 6.1 is that discriminant-uniform average over G extensions K of
$h_3(F_K/k)$
is provably a weighted average of
$h_3(F/k)$
over F, against the measure on H-extensions in which a field F has measure proportional to
$ {\mathrm {Res}}_{s=1}\zeta _F(s)/(\zeta _F(2) {\mathrm {Disc}}(F)^2)$
. There is no particular reason to think that this weighted average of
$h_3(F/k)$
over F will give the same result as the discriminant-uniform average over F. Indeed, the weighted average is heavily influenced by the F of small discriminant. When
$H=C_2$
, the discriminant uniform average over
$C_2$
-extensions F of
$h_3(F/k)$
is known by Datskovsky–Wright, and is as predicted by Cohen–Lenstra–Martinet.
Since the quadratic fields of small discriminant have very little
$3$
-torsion in their class groups, one expects the weighted average of
$h_3(F/k)$
to be smaller than the discriminant-uniform average over F, and indeed that is what we see in the chart above.
This perspective also explains the counterexample to the Cohen–Lenstra–Martinet heuristics of Bartel and Lenstra [Reference Bartel and LenstraBL20]. Let f be the indicator function of whether the
$3$
-torsion in the class group of a quadratic field is trivial. For a cyclic quartic field K, let
$F_K$
be the quadratic subfield. Bartel and Lenstra prove that the discriminant-uniform average over cyclic quartic K of
$f(F_K)$
is a weighted average of
$f(F)$
over quadratic F. On the other hand, the Cohen–Lenstra–Martinet heuristics predict that the discriminant-uniform average over cyclic quartic K of
$f(F_K)$
is the discriminant-uniform average of
$f(F)$
over quadratic F.
8 Averages for other groups
Our methods apply to more groups than just
$2$
-groups. We focus here on results that may be obtained purely from the results of Section 3, and whose proofs in particular do not rely on arguments from Sections 2 and 4. This permits a proof of the following general result.
Theorem 8.1. Let
$H \subseteq S_n$
be transitive and set
$G = C_2 \wr H$
. Then G-extensions K of a number field k have a unique index two subfield
$F_K/k$
. Moreover
$F_K$
is an H-extension of k. For
$u \in \mathbb {Z}$
, let
$E_k^u(G,X) \subseteq E_k(G,X)$
be the subset of those K for which
$ {\mathrm {rk}} \mathcal {O}_K^* - {\mathrm {rk}} \mathcal {O}_{F_K}^* = u$
.
1) If
$E_k^u(G,\infty )$
is nonempty and
$|E_k(H,X)| \ll _{k,H,\epsilon } X^{2/3+\epsilon }$
for every
$X\geq 1$
, then

2) If
$E_k(H,\infty )$
is nonempty and

for every
$X \geq 1$
, then there is an explicit constant
$c_{k,G,3}$
such that

Before we discuss the proof of Theorem 8.1, we note that the hypotheses of the first case are satisfied, for example, by any group
$H \neq C_2$
in its regular representation that occurs as a Galois group over k [Reference Ellenberg and VenkateshEV06, Proposition 1.3], and by any nilpotent group H without a transposition [Reference AlbertsAlb20, Corollary 1.8]. The hypotheses of the second case are satisfied for any p-group with
$p \neq 5$
odd, as follows from [Reference Klüners and MalleKM04, Corollary 7.3] and the trivial bound on
$h_3(F/k)$
for
$p \geq 7$
, and from [Reference Klüners and WangKW20] for
$p=3$
.
Proof. The claims about the subfield
$F_K$
follow as in Lemma 6.6. Suppose now that
$E_k^u(G,\infty )$
is nonempty and
$|E_k(H,X)| \ll _{k,H,\epsilon } X^{2/3+\epsilon }$
for every
$X\geq 1$
. We mimic the proof of Theorem 6.1, indicating the necessary modifications. Let
$X \geq 1$
. From Corollary 3.2, it follows for any
$Y\geq 1$
that

where the second inequality is by partial summation and [BST+20, Theorem 1.1]. This provides an analogue of Theorem 5.1. We now establish a soft analogue of Theorem 6.5, with the remainder of the proof then proceeding as in those of Theorems 6.1 and 6.11. Note that if
$G^\prime \subseteq G$
surjects onto H and contains an element conjugate to
$\sigma = (1,0,\dots ,0) \rtimes 1_H$
in G, then in fact
$G^\prime = G$
.
For any fixed H-extension F with discriminant at most Y, let
$\mathcal {P}$
be a finite set of primes of k that split completely in F. If some prime of F above a prime
$\mathfrak {p} \in \mathcal {P}$
is inert in a quadratic extension
$K/F$
, while all the other primes of F above
$\mathfrak {p}$
are split in K, then by the note above we have that
$K/k$
is a
$C_2\wr H$
extension. If this happens, then we say that
$\mathfrak {p}$
is bad for
$K/F$
. It follows from [Reference Bhargava, Shankar and WangBSW15, Theorem 2] that

The expressions in the product above are uniformly (in
$\mathfrak {p}$
) bounded away from
$1$
, and in particular the product is
$O(\beta _n^{\#\mathcal {P}})$
for some constant
$\beta _n <1$
depending at most on n. Since there are infinitely many primes that split completely in F, this product may be made arbitrarily small, and thus

Proceeding now as in the proof of Theorem 6.1, the first claim follows. The second follows analogously.
Acknowledgements
We thank Rachel Newton, Arul Shankar, Takashi Taniguchi, Frank Thorne, and Jesse Thorner for helpful conversations related to this project. We also thank Brandon Alberts, Alex Bartel, Jordan Ellenberg, Jürgen Klüners, Lillian Pierce, Frank Thorne, Jesse Thorner, and Asif Zaman for comments on an earlier version of this paper. We thank the referees for many useful suggestions on the exposition.
Competing interest
The authors have no competing interest to declare.
Financial support
RJLO was partially supported by National Science Foundation grants DMS-1601398 and DMS-2200760 and by a Simons Foundation Fellowship in Mathematics. JW was partially supported by a Foerster-Bernstein Fellowship at Duke University and by National Science Foundation grant DMS-2201346. MMW was partially supported by a Packard Fellowship for Science and Engineering, National Science Foundation grant DMS-2052036, an NSF Waterman award, a MacArthur Fellowship, and the Radcliffe Institute for Advanced Study at Harvard University.