1 Introduction

The theory of Campana points is of growing interest in arithmetic geometry due to its ability to interpolate between rational and integral points. Two competing notions of Campana points can be found in the literature, both extending a definition of “orbifold rational points” for curves within Campana’s theory of “orbifoldes géométriques” in [7,8,9,10]. They capture the idea of rational points which are integral with respect to a weighted boundary divisor. These two notions have been termed Campana points and weak Campana points in the recent paper [27] of Pieropan, Smeets, Tanimoto and Várilly-Alvarado, in which the authors initiate a systematic quantitative study of points of the former type on smooth Campana orbifolds and prove a logarithmic version of Manin’s conjecture for Campana points on vector group compactifications. The only other quantitative results in the literature are found in [5, 6, 26, 32, 33], and the former four of these indicate the close relationship between Campana points and m-full solutions of equations. We recall that, given \(m \in \mathbb {Z}_{\ge 2}\), we say that \(n \in \mathbb {Z} \setminus \{0\}\) is m-full if all primes in the prime decomposition of n have multiplicity at least m.

In this paper, we bring together the perspectives in the above papers and provide the first result for Campana points on singular orbifolds. As observed in [27, Sect. 1.1], the study of weak Campana points of bounded height is challenging and requires new ideas for the regularisation of certain Fourier transforms, and these ideas for the orbifolds in consideration are the main innovation of this paper. We adopt a height zeta function approach, similar to the one employed in [27] and modelled on the work of Loughran in [19] and Batyrev and Tschinkel in [3] on toric varieties, in order to prove log Manin conjecture-type results for both types of Campana points on \(\left( \mathbb {P}^{d-1}_K,\left( 1-\frac{1}{m}\right) Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \right) \), where \(Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \) is the zero locus of a norm form \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) associated to a K-basis \({{\,\mathrm{\varvec{\omega }}\,}}\) of a Galois extension of number fields E/K of degree \(d \ge 2\) coprime to \(m \in \mathbb {Z}_{\ge 2}\) if d is not prime. Although toric varieties also play a prominent role in [26], the tori there are split, whereas we shall work with anisotropic tori. When \(K = \mathbb {Q}\), we derive from the result for weak Campana points an asymptotic for the number of elements of E of bounded height with m-full norm over \(\mathbb {Q}\). We compare the result for Campana points to a conjecture of Pieropan, Smeets, Tanimoto and Várilly-Alvarado [27, Conj. 1.1, p. 3].

1.1 Results

Theorem 1.1

Let E/K be a Galois extension of number fields of degree \(d \ge 2\), and let \(m \ge 2\) be an integer which is coprime to d if d is not prime. Let \({{\,\mathrm{\varvec{\omega }}\,}}\) be a K-basis of E. Denote by \(\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\) the \(\mathbb {Q}\)-divisor \(\left( 1-\frac{1}{m}\right) Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \) of \(\mathbb {P}^{d-1}_K\) for \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) the norm form corresponding to \({{\,\mathrm{\varvec{\omega }}\,}}\). Let H denote the anticanonical height function on \(\mathbb {P}^{d-1}_K\) from Definition 4.5. Then there exists an explicit finite set \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \subset {{\,\mathrm{Val}\,}}\left( K\right) \) such that, for any finite set of places \(S \supset S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), the number \(N\left( \left( \mathbb {P}^{d-1}_{K},\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) ,H,B,S\right) \) of weak Campana \(\mathcal {O}_{K,S}\)-points of height at most \(B \in \mathbb {R}_{\ge 1}\) on the orbifold \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) with respect to the model \(\mathbb {P}^{d-1}_{\mathcal {O}_{K,S}}\) of \(\mathbb {P}^{d-1}_K\) has the asymptotic formula

$$\begin{aligned} N\left( \left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) ,H,B,S\right) \sim c\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\right) B^{\frac{1}{m}}\left( \log B\right) ^{b\left( d,m\right) -1} \end{aligned}$$

as \(B \rightarrow \infty \) for some explicit positive constant \(c\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\right) \), where

$$\begin{aligned} b\left( d,m\right) = \frac{1}{d}\left( {{d+m-1}\atopwithdelims (){d-1}} - {{m-1}\atopwithdelims (){d-1}}\right) . \end{aligned}$$

Note 1.2

If \({{\,\mathrm{\varvec{\omega }}\,}}\) is a relative integral basis of E/K, then \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) = S_{\infty }\), the set of archimedean places of K, in Theorem 1.1 (see Remark 4.4).

Each rational point \(P \in \mathbb {P}^{d-1}\left( \mathbb {Q}\right) \) possesses precisely two sets of coordinates in \(\mathbb {Z}^d_{\text {prim}} = \{\left( x_0,\dots ,x_{d-1}\right) \in \mathbb {Z}^d : \gcd \left( x_0,\dots ,x_{d-1}\right) = 1\}\). Interpreting H and \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) as functions on this set, we immediately obtain the following result.

Corollary 1.3

Taking \(K=\mathbb {Q}\) and letting \({{\,\mathrm{\varvec{\omega }}\,}}\) be an integral basis with the notation and hypotheses of Theorem 1.1, we have

$$\begin{aligned} \#\{x \in \mathbb {Z}^d_{\text {prim}} : H\left( x\right) \le B, \, N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) \text { is }m\text {-full}\} \sim 2c\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S_{\infty }\right) B^{\frac{1}{m}} \left( \log B\right) ^{b\left( d,m\right) -1}. \end{aligned}$$

Arithmetically special (e.g. prime, square-free) values of norm forms are a topic of long-standing interest in number theory (see e.g. [13, 20]).

Campana points are only defined and studied for smooth orbifolds (i.e. smooth varieties for which the orbifold divisor has strict normal crossings support) in [27]. In order to study the Campana points of \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \), which is smooth only when \(d = 2\), we must first generalise the definition of Campana points, which we do in Sect. 2.1. Using the same strategy employed in the proof of Theorem 1.1, we then derive an asymptotic for the number of Campana points on \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \).

Theorem 1.4

With the notation and hypotheses of Theorem 1.1, denote by \(\widetilde{N}\Big (\big (\mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\big ),H,B,S\Big )\) the number of Campana \(\mathcal {O}_{K,S}\)-points on \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) of height at most \(B \in \mathbb {R}_{\ge 1}\) with respect to H. Then there exists an explicit positive constant \(\widetilde{c}\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\right) \) such that, as \(B \rightarrow \infty \), we have

$$\begin{aligned} \widetilde{N}\left( \left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) ,H,B,S\right) \sim \widetilde{c}\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\right) B^{\frac{1}{m}}. \end{aligned}$$

Remark 1.5

It is not clear if the exponent of the logarithm in Theorem 1.1 admits a geometric interpretation as it does in Theorem 1.4 (cf. [27, Conj. 1.1, p. 3]).

1.2 Conventions

1.2.1 Algebra

We take \(\mathbb {N} = \mathbb {Z}_{\ge 1}\). We denote by \(R^*\) the group of units of a ring R. Given a group G, we denote by \(1_G\) the identity element of G, and for any \(n \in \mathbb {N}\), we set \(G[n] = \{g \in G: g^n = 1_G\}\). For any perfect field F, we fix an algebraic closure \(\overline{F}\) and set \(G_F = {{\,\mathrm{Gal}\,}}\left( \overline{F}/F\right) \). Given a topological group G, we denote by \(G^{\wedge } = {{\,\mathrm{Hom}\,}}\left( G,S^1\right) \) its group of continuous characters, where \(S^1 = \{z \in \mathbb {C}: z \overline{z} = 1\} \subset \mathbb {C}^*\) is the circle group. A monomial in the variables \(x_1,\dots ,x_n\) is a product \(x_1^{a_1}\cdots x_n^{a_n}\), \(\left( a_1,\dots ,a_n\right) \in \mathbb {Z}_{\ge 0}^n\). For any \(n \in \mathbb {N}\), we denote by \(\mu _n\) the group of nth roots of unity and by \(S_n\) the symmetric group of order n.

1.2.2 Geometry

We denote by \(\mathbb {P}_R^n\) the projective n-space over the ring R. We omit the subscript if the ring R is clear. Given a homogeneous polynomial \(f \in R[x_0,\dots ,x_n]\), we denote by \(Z\left( f\right) = {{\,\mathrm{Proj}\,}}R[x_0,\dots ,x_n]/\left( f\right) \) the zero locus of f viewed as a closed subscheme of \(\mathbb {P}_R^n\). A variety over a field F is a geometrically integral separated scheme of finite type over F. Given a variety X defined over F and an extension E/F, we denote by \(X_E= V \times _{{{\,\mathrm{Spec}\,}}F} {{\,\mathrm{Spec}\,}}E\) the base change of X over E, and we write \(\overline{X} = X \times _{{{\,\mathrm{Spec}\,}}F} {{\,\mathrm{Spec}\,}}\overline{F}\). When \(F = K\) and \(E = K_v\) for a number field K and a place v of K, we write \(X_{v} = X_{K_v}\). Given a field F, we define \(\mathbb {G}_{m,F} = {{\,\mathrm{Spec}\,}}F[x_0,x_1]/\left( x_0 x_1-1\right) \). We omit the subscript F if the field is clear.

1.2.3 Number theory

Given an extension of number fields L/K with K-basis \({{\,\mathrm{\varvec{\omega }}\,}}= \{\omega _0,\dots ,\omega _{d-1}\}\), we write \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x_0,\dots ,x_{d-1}\right) = N_{L/K}\left( x_0 \omega _0 + \cdots + x_{d-1}\omega _{d-1}\right) \) for the associated norm form. We denote by \({{\,\mathrm{Val}\,}}\left( K\right) \) the set of valuations of a number field K, and we denote by \(S_{\infty }\) the set of archimedean valuations. For \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), we denote by \(\mathcal {O}_v\) the maximal compact subgroup of \(K_v\). For a finite set of places S containing \(S_{\infty }\), we denote by \(\mathcal {O}_{K,S} = \{\alpha \in K : \alpha \in \mathcal {O}_v \text { for all }v \not \in S\}\) the ring of algebraic S-integers of K. We write \(\mathcal {O}_K = \mathcal {O}_{K,S_{\infty }}\). For \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) non-archimedean, we denote by \(\pi _v\) and \(q_v\) a uniformiser for the residue field of \(K_v\) and the size of the residue field of \(K_v\) respectively. If \(v \mid \infty \), then we set \(\log q_v = 1\). For each \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), we choose the absolute value \(|x|_v = |N_{K_v/\mathbb {Q}_p}\left( x\right) |_p\) for the unique \(p \in {{\,\mathrm{Val}\,}}\left( \mathbb {Q}\right) \) with \(v \mid p\) and the usual absolute value \(|\cdot |_p\) on \(\mathbb {Q}_p\). We denote by \(\mathbb {A}_K =\widehat{\prod }_{v \in {{\,\mathrm{Val}\,}}\left( K\right) }^{\mathcal {O}_v}K_v\) the adele ring of K with the restricted product topology.

2 Background

2.1 Campana points

In this section we define Campana orbifolds, Campana points and weak Campana points, generalising the definitions in [27, Sect. 3.2] in such a way that the exponents in Theorem 1.4 match those in [27, Conj. 1.1, p. 3].

Definition 2.1

A Campana orbifold over a field F is a pair \(\left( X,D_{\epsilon }\right) \) consisting of a proper, normal variety X over F and an effective Cartier \(\mathbb {Q}\)-divisor

$$\begin{aligned} D_{\epsilon } = \sum _{\alpha \in \mathcal {A}} \epsilon _{\alpha } D_{\alpha } \end{aligned}$$

on X, where the \(D_{\alpha }\) are prime divisors and \(\epsilon _{\alpha } = 1 - \frac{1}{m_{\alpha }}\) for some \(m_{\alpha } \in \mathbb {Z}_{\ge 2} \cup \{\infty \}\) (by convention, we take \(\frac{1}{\infty } = 0\)). We define the support of the \(\mathbb {Q}\)-divisor \(D_{\epsilon }\) to be

$$\begin{aligned} D_{\text {red}} = \sum _{\alpha \in \mathcal {A}} D_{\alpha }. \end{aligned}$$

We say that \(\left( X,D_{\epsilon }\right) \) is smooth if X is smooth and \(D_{\text {red}}\) has strict normal crossings (see [31, Sect. 41.21] for the definition of strict normal crossings divisors).

Let \(\left( X,D_{\epsilon }\right) \) be a Campana orbifold over a number field K. Let \(S \subset {{\,\mathrm{Val}\,}}\left( K\right) \) be a finite set containing \(S_{\infty }\).

Definition 2.2

A model of \(\left( X,D_{\epsilon }\right) \) over \(\mathcal {O}_{K,S}\) is a pair \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \), where \(\mathcal {X}\) is a flat proper model of X over \(\mathcal {O}_{K,S}\) (i.e. a flat proper \(\mathcal {O}_{K,S}\)-scheme with \(\mathcal {X}_{\left( 0\right) } \cong X\)) and \(\mathcal {D}_{\epsilon } = \sum _{\alpha \in \mathcal {A}}\epsilon _{\alpha } \mathcal {D}_{\alpha }\) for \(\mathcal {D}_{\alpha }\) the Zariski closure of \(D_{\alpha }\) in \(\mathcal {X}\).

Define \(\mathcal {D}_{\text {red}} = \sum _{\alpha \in \mathcal {A}}\mathcal {D}_{\alpha }\). Denote by \(\mathcal {D}_{\alpha _v}\), \(\alpha _v \in \mathcal {A}_v\) the irreducible components of \(\mathcal {D}_{\text {red}}\) over \({{\,\mathrm{Spec}\,}}\mathcal {O}_v\). We write \(\alpha _v \mid \alpha \) if \(\mathcal {D}_{\alpha _v} \subset \mathcal {D}_{\alpha }\).

Let \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) be a model for \(\left( X,D_{\epsilon }\right) \) over \(\mathcal {O}_{K,S}\). For \(v \not \in S\), any \(P \in X\left( K\right) \) induces some \(\mathcal {P}_v \in \mathcal {X}\left( \mathcal {O}_v\right) \) by the valuative criterion of properness [14, Thm. II.4.7, p. 101].

Definition 2.3

Let \(P \in X\left( K\right) \) and take a place \(v \not \in S\). For each \(\alpha _v \in \mathcal {A}_v\), we define the local intersection multiplicity \(n_v\left( \mathcal {D}_{\alpha _v},P\right) \) of \(\mathcal {D}_{\alpha _v}\) and P at v to be \(\infty \) if \(\mathcal {P}_v \subset \mathcal {D}_{\alpha _v}\), and the colength of the ideal \(\mathcal {P}_v^*\mathcal {D}_{\alpha _v} \subset \mathcal {O}_v\) otherwise. We then define the quantities

$$\begin{aligned} n_v\left( \mathcal {D}_{\alpha },P\right) = \sum _{\alpha _v \mid \alpha } n_v\left( \mathcal {D}_{\alpha _v},P\right) , \quad n_v\left( \mathcal {D}_{\epsilon },P\right) = \sum _{\alpha \in \mathcal {A}} \epsilon _{\alpha } n_{v}\left( \mathcal {D}_{\alpha },P\right) . \end{aligned}$$

We are now ready to define weak Campana points and Campana points. Both notions arise from [1], with the former appearing in its current form in [2, Sect. 1].

Definition 2.4

We say that \(P \in X\left( K\right) \) is a weak Campana \(\mathcal {O}_{K,S}\)-point of \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) if the following implications hold for all places \(v \not \in S\) of K and for all \(\alpha \in \mathcal {A}\):

  1. (1)

    If \(\epsilon _\alpha = 1\) (meaning \(m_{\alpha } = \infty \)), then \(n_v\left( \mathcal {D}_{\alpha },P\right) = 0\).

  2. (2)

    If \(n_v\left( \mathcal {D}_{\epsilon },P\right) > 0\), then

    $$\begin{aligned} \sum _{\alpha \in \mathcal {A}} \frac{1}{m_{\alpha }}n_v\left( \mathcal {D}_{\alpha },P\right) \ge 1. \end{aligned}$$

We denote the set of weak Campana \(\mathcal {O}_{K,S}\)-points of \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) by \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) _{\mathbf {w}}\left( \mathcal {O}_{K,S}\right) \).

Definition 2.5

We say that \(P \in X\left( K\right) \) is a Campana \(\mathcal {O}_{K,S}\)-point of \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) if the following implications hold for all places \(v \not \in S\) of K and for all \(\alpha \in \mathcal {A}\).

  1. (3)

    If \(\epsilon _{\alpha } = 1\) (meaning \(m_{\alpha } = \infty \)), then \(n_v\left( \mathcal {D}_{\alpha },P\right) = 0\).

  2. (4)

    If \(\epsilon _\alpha \ne 1\) and \(n_v\left( \mathcal {D}_{\alpha _v},P\right) > 0\) for some \(\alpha _v \mid \alpha \), then \( n_v\left( \mathcal {D}_{\alpha _v},P\right) \ge m_{\alpha }. \)

We denote the set of Campana \(\mathcal {O}_{K,S}\)-points of \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) by \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \left( \mathcal {O}_{K,S}\right) \).

Remark 2.6

Informally, weak Campana points are rational points \(P \in X\left( K\right) \) avoiding \(\cup _{\epsilon _\alpha = 1}\mathcal {D}_{\alpha }\) which, upon reduction modulo any place \(v \not \in S\), either do not lie on \(D_{\text {red}}\) or lie on \(D_\alpha \) with multiplicity at least \(m_\alpha \) on average over \(\alpha \). Similarly, Campana points are rational points \(P \in X\left( K\right) \) avoiding \(\cup _{\epsilon _\alpha = 1}\mathcal {D}_{\alpha }\) which, upon reduction modulo any place \(v \not \in S\), either do not lie on \(D_{\text {red}}\) or lie on each v-adic irreducible component of each \(D_{\alpha }\) with multiplicity either 0 or at least \(m_{\alpha }\).

Remark 2.7

Our definition of Campana points differs from the one in [27, Sect. 3.2], in which one requires simply that \(n_v\left( \mathcal {D}_\alpha ,P\right) \ge m_{\alpha }\) instead of \(n_v\left( \mathcal {D}_{\alpha _v},P\right) \ge m_{\alpha }\) in the second implication. If one were to apply this definition to the orbifold \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) of Theorem 1.1, which is singular for all \(d \ge 3\) as \(Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \) is not a strict normal crossings divisor in this case, then the weak Campana points and the Campana points would be the same, but the asymptotic of Theorem 1.1 differs to [27, Conj. 1.1, p. 3] for \(d \ge 3\) (at least if one takes the thin set there to be the empty set). Using the definitions above, we obtain the asymptotic for Campana points in Theorem 1.4, whose exponents match this conjecture.

Lemma 2.8

Let \(\left( X,D_{\epsilon }\right) \) be a smooth Campana orbifold over a number field K which is Kawamata log terminal (i.e. \(\epsilon _{\alpha } < 1\) for all \(\alpha \in \mathcal {A}\)), and let \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) be a model of \(\left( X,D_{\epsilon }\right) \) over \(\mathcal {O}_{K,S}\) with \(\mathcal {X}\) smooth over \(\mathcal {O}_{K,S}\) and \(\mathcal {D}_{\text {red}}\) a relative strict normal crossings divisor in \(\mathcal {X}/\mathcal {O}_{K,S}\) as defined in [17, Sect. 2]. Then the definition of Campana points on \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) above coincides with the one in [27, Sect. 3.2].

Proof

Since \(\mathcal {D}_{\text {red}}\) is a relative strict normal crossings divisor, each irreducible component \(\mathcal {D}_{\alpha }\) is smooth over \(\mathcal {O}_{K,S}\). In particular, its base change over \({{\,\mathrm{Spec}\,}}\mathcal {O}_v\) is smooth for any \(v \not \in S\), so the divisors \(\mathcal {D}_{\alpha _v}\), \(\alpha _v \mid \alpha \) are disjoint. Then, for any rational point \(P \in X\left( K\right) \), the reduction of P at the place v can lie on at most one of the divisors \(\mathcal {D}_{\alpha _v}\), \(\alpha _v \mid \alpha \), so \(n_v\left( \mathcal {D}_{\alpha },P\right) = \sum _{\alpha _v \mid \alpha } n_v\left( \mathcal {D}_{\alpha _v},P\right) \) is either 0 or at least \(m_{\alpha }\) if and only if each \(n_v\left( \mathcal {D}_{\alpha _v},P\right) \), \(\alpha _v \mid \alpha \), is either 0 or at least \(m_{\alpha }\). \(\square \)

2.2 Toric varieties

Definition 2.9

An (algebraic) torus over a field F is an algebraic group T over F such that \(\overline{T} \cong \mathbb {G}_m^n\) for some \(n \in \mathbb {N}\). The splitting field of a torus T over a field F is defined to be the smallest Galois field extension E of F for which \(T_E \cong \mathbb {G}_m^n\).

Definition 2.10

A toric variety is a smooth projective variety X equipped with a faithful action of an algebraic torus T such that there is an open dense orbit containing a rational point.

Definition 2.11

Let T be a torus over a field F. The character group of T is \(X^*\left( \overline{T}\right) = {{\,\mathrm{Hom}\,}}\left( \overline{T},\mathbb {G}_m\right) \), and we have \(X^*\left( T\right) = X^*\left( \overline{T}\right) ^{G_F}\). The cocharacter group of T is \(X_*\left( \overline{T}\right) = {{\,\mathrm{Hom}\,}}\left( X^*\left( \overline{T}\right) , \mathbb {Z}\right) \), and we have \(X_*\left( T\right) = X_*\left( \overline{T}\right) ^{G_F}\). We let \(X^*\left( T\right) _{\mathbb {R}} = X^*\left( T\right) \otimes _{\mathbb {Z}} \mathbb {R}\) and \(X_*\left( T\right) _{\mathbb {R}} = X_*\left( T\right) \otimes _{\mathbb {Z}} \mathbb {R}\).

Definition 2.12

An algebraic torus T over a field F is anisotropic if it has trivial character group over F, i.e. \(X^*\left( T\right) = 0.\)

Let T be a torus over a number field K with splitting field E. Set \(T_\infty = \prod _{v \mid \infty }T_v\). For \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), let \(T\left( \mathcal {O}_v\right) \) denote the maximal compact subgroup of \(T\left( K_v\right) \).

Definition 2.13

Let \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) and \(w \in {{\,\mathrm{Val}\,}}\left( E\right) \) with \(w \mid v\).

For \(v \not \mid \infty \) with ramification degree \(e_v\) in E/K, define the maps

$$\begin{aligned} \deg _{T,v}: T\left( K_v\right) \rightarrow X_*\left( T_v\right) , \quad t_v \mapsto [\chi _v \mapsto v\left( \chi _v\left( t_v\right) \right) ] \end{aligned}$$

and \(\deg _{T,E,v} = e_v \deg _{T,v}\).

For \(v \mid \infty \), define the maps

$$\begin{aligned} \deg _{T,v}: T\left( K_v\right) \rightarrow X_*\left( T_v\right) _{\mathbb {R}}, \quad t_v \mapsto [\chi _v \mapsto \log \left| \chi _v\left( t_v\right) \right| _v] \end{aligned}$$

and \(\deg _{T,E,v} = [E_w:K_v] \deg _{T,v}\).

Finally, define the maps

$$\begin{aligned} \deg _T = \sum _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \left( \log q_v\right) \deg _{T,v}, \quad \deg _{T,E} = \sum _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \left( \log q_w\right) \deg _{T,E,v}. \end{aligned}$$

Lemma 2.14

[4, Sect. 2.2], [19, Sect. 4.2] Let \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), and let f be either \(\deg _{T,v}\) or \(\deg _{T,E,v}\).

  1. (i)

    If v is non-archimedean, then we have the exact sequence

    $$\begin{aligned} 0 \rightarrow T\left( \mathcal {O}_v\right) \rightarrow T\left( K_v\right) \xrightarrow {f} X_*\left( T_v\right) . \end{aligned}$$

    The image of f is open and of finite index. Further, if v is unramified in E, then f is surjective.

  2. (ii)

    If v is archimedean, then we have the short exact sequence

    $$\begin{aligned} 0 \rightarrow T\left( \mathcal {O}_v\right) \rightarrow T\left( K_v\right) \xrightarrow {f} X_*\left( T_v\right) _{\mathbb {R}} \rightarrow 0. \end{aligned}$$

    Further, f admits a canonical section.

  3. (iii)

    Letting g be either \(\deg _T\) or \(\deg _{T,E}\) and denoting its kernel by \(T\left( \mathbb {A}_K\right) ^1\), we have the split short exact sequence

    $$\begin{aligned} 0 \rightarrow T\left( \mathbb {A}_K\right) ^1 \rightarrow T\left( \mathbb {A}_K\right) \xrightarrow {g} X_*\left( T\right) _{\mathbb {R}} \rightarrow 0, \end{aligned}$$

    hence we have an isomorphism

    $$\begin{aligned} T\left( \mathbb {A}_K\right) \cong T\left( \mathbb {A}_K\right) ^1 \times X_*\left( T\right) _{\mathbb {R}}. \end{aligned}$$
    (2.1)

Definition 2.15

Let \(\chi \) be a character of \(T\left( \mathbb {A}_K\right) \). We say that \(\chi \) is automorphic if it is trivial on \(T\left( K\right) \). We say that \(\chi \) is unramified at \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) if \(\chi _v\) is trivial on \(T\left( \mathcal {O}_v\right) \), and we say that it is unramified if it is unramified at every \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \).

The canonical sections of the maps \(T\left( K_v\right) \xrightarrow {\deg _{T,v}} X_*\left( T_v\right) _{\mathbb {R}}\) for each \(v \mid \infty \) from Lemma 2.14(ii) induce a canonical section of the composition

$$\begin{aligned} T\left( \mathbb {A}_K\right) \rightarrow \prod _{v \mid \infty } T\left( K_v\right) \rightarrow X_*\left( T_\infty \right) _\mathbb {R}, \end{aligned}$$

which in turn induces a “type at infinity map”

$$\begin{aligned} T\left( \mathbb {A}_K\right) ^\wedge \rightarrow X^*\left( T_\infty \right) _{\mathbb {R}}, \quad \chi \mapsto \chi _{\infty }. \end{aligned}$$
(2.2)

Defining \(\mathsf {K}_T = \prod _{v}T\left( \mathcal {O}_v\right) \), the splitting (2.1) for \(g = \deg _T\) induces a map

$$\begin{aligned} \left( T\left( \mathbb {A}_K\right) ^1/T\left( K\right) \mathsf {K}_T\right) \rightarrow X^* \left( T_{\infty }\right) _{\mathbb {R}} \end{aligned}$$

which has finite kernel and image a lattice of codimension \({{\,\mathrm{rank}\,}}X^*\left( T\right) \) (see [4, Lem. 4.52, p. 96]).

Note 2.16

When T is anisotropic, we have \(T\left( \mathbb {A}_K\right) ^1 = T\left( \mathbb {A}_K\right) \) by Lemma 2.14(iii), and then we see from the above that there is a map

$$\begin{aligned} \left( T\left( \mathbb {A}_K\right) /T\left( K\right) \mathsf {K}_T\right) \rightarrow X^* \left( T_{\infty }\right) _{\mathbb {R}} \end{aligned}$$

with finite kernel and image a lattice of full rank.

2.3 Hecke characters

Definition 2.17

A Hecke character for K is an automorphic character of \(\mathbb {G}_{m,K}\).

Each Hecke character \(\chi \) has a conductor \(q\left( \chi \right) \in \mathbb {N}\) (see [18, Sect. 5.10]), which measures the ramification of \(\chi \) at the non-archimedean places of K.

Definition 2.18

A Hecke character is principal if it is trivial on \(\mathbb {G}_{m,K}\left( \mathbb {A}_K\right) ^1\).

By Lemma 2.14(iii), \(\chi \) is principal if and only if \( \chi = \Vert \cdot \Vert ^{i\theta }\) for some \(\theta \in \mathbb {R}\), where \(\Vert \cdot \Vert \) denotes the adelic norm map, i.e.

$$\begin{aligned} \Vert \cdot \Vert : \mathbb {A}_K^* \rightarrow S^1, \quad \left( x_v\right) _v \mapsto \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) }|x_v|_v. \end{aligned}$$

Definition 2.19

The (Hecke) L-function \(L\left( \chi ,s\right) \) of a Hecke character \(\chi \) is

$$\begin{aligned} L\left( \chi ,s\right) = \prod _v \left( 1-\frac{\chi _v\left( \pi _v\right) }{q_v^s}\right) ^{-1}, \end{aligned}$$

where the product is taken over all places \(v \not \mid \infty \) at which \(\chi \) is unramified.

The Dedekind zeta function of K is

$$\begin{aligned} \zeta _K\left( s\right) = L\left( 1,s\right) . \end{aligned}$$

Given a Hecke character \(\chi \) for a number field L and \(w \in {{\,\mathrm{Val}\,}}\left( L\right) \), we denote by \(L_w\left( \chi ,s\right) \) the local factor at w for the Euler product of \(L\left( \chi ,s\right) \), i.e.

$$\begin{aligned} L_w\left( \chi ,s\right) = {\left\{ \begin{array}{ll} \left( 1 - \frac{\chi _w \left( \pi _w\right) }{q_w^s}\right) ^{-1} \text { if }w \not \mid \infty \text { and }\chi \text { is unramified at }w, \\ 1 \text { otherwise}. \end{array}\right. } \end{aligned}$$

When working over the field \(L \supset K\), we define \(L_v\left( \chi ,s\right) \) for each \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) by

$$\begin{aligned} L_v\left( \chi ,s\right) = \prod _{w \mid v} L_w\left( \chi ,s\right) . \end{aligned}$$

Theorem 2.20

[15, Sect. 6] The L-function of a Hecke character \(\chi \) admits a meromorphic continuation to \(\mathbb {C}\). If \(\chi = \Vert \cdot \Vert ^{i \theta }\) for some \(\theta \in \mathbb {R}\), then this continuation admits a single pole of order 1 at \(s = 1+ i\theta \). Otherwise, it is holomorphic.

Definition 2.21

Let \(\psi \) be a character of \(\prod _{v | \infty } K_v^*\). The restriction of \(\psi \) to each \(\mathbb {R}_{>0} \subset K_v^*\) is of the form \(x \mapsto |x|^{i\kappa _v}\) for some \(\kappa _v \in \mathbb {R}\). We define

$$\begin{aligned} \Vert \psi \Vert = \max _{v \mid \infty } |\kappa _v|. \end{aligned}$$

Lemma 2.22

[18, Exercise 3, §5.2, p. 100] Let \(\chi \) be a non-principal Hecke character of K, let C be a compact subset of \({\text {Re}}s \ge 1\) and let \(\varepsilon > 0\). Then

$$\begin{aligned} L\left( \chi ,s\right) \ll _{\varepsilon , C} q\left( \chi \right) ^{\varepsilon } \left( 1+ \Vert \chi _{\infty }\Vert \right) ^{\varepsilon }, \quad \left( s-1\right) \zeta _K\left( s\right) \ll _C 1, \quad s \in C. \end{aligned}$$

Definition 2.23

Let E/K be Galois, let \(\chi \) be a Hecke character for E and let \(g \in {{\,\mathrm{Gal}\,}}\left( E/K\right) \). We define the (Galois) twist of \(\chi \) by g to be the character

$$\begin{aligned} \chi ^g: \mathbb {A}_E^* \rightarrow S^1, \quad \left( t_w\right) _w \mapsto \chi \left( \left( g_w\left( t_w\right) \right) _{gw}\right) . \end{aligned}$$

Here, gw denotes the place of E obtained by the action of g on \({{\,\mathrm{Val}\,}}\left( E\right) \), and \(g_w: E_w \rightarrow E_{gw}\) is the induced map on completions. One may easily verify that \(\chi ^g\) is trivial on \(E^*\), hence it is also a Hecke character for E.

3 The norm torus

In this section, we fix an extension of number fields L/K of degree \(d \ge 2\) with Galois closure E and a K-basis \({{\,\mathrm{\varvec{\omega }}\,}}= \{\omega _0,\dots ,\omega _{d-1}\}\). We write \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x_0,\dots ,x_{d-1}\right) \) for the norm form corresponding to \({{\,\mathrm{\varvec{\omega }}\,}}\), and \(G = {{\,\mathrm{Gal}\,}}\left( E/K\right) \). From the equality

$$\begin{aligned} N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x_0, \dots , x_{d-1}\right) = \prod _{g \in G/{{\,\mathrm{Gal}\,}}\left( E/L\right) }\left( x_0 g\left( \omega _0\right) + \cdots + x_{d-1} g\left( \omega _{d-1}\right) \right) , \end{aligned}$$
(3.1)

we see that \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) is irreducible over K and has splitting field E. We denote by T the norm torus \(T_{{{\,\mathrm{\varvec{\omega }}\,}}} = \mathbb {P}^{d-1}_K\setminus Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \). As noted in [19, Sect. 1.2], \(\mathbb {P}^{d-1}_K\) is a toric variety with respect to T, and \(T \cong R_{L/K}\mathbb {G}_m/\mathbb {G}_m\) is anisotropic. Since its boundary is \(Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \), its splitting field is E. We have the short exact sequence

$$\begin{aligned} 0 \rightarrow \mathbb {G}_m \rightarrow R_{L/K}\mathbb {G}_m \rightarrow T \rightarrow 0. \end{aligned}$$
(3.2)

Note 3.1

The isomorphisms \(T\left( \mathbb {A}_K\right) \cong \mathbb {A}_L^*/\mathbb {A}_K^*\) and \(T\left( K\right) \cong L^*/K^*\) follow from Hilbert’s Theorem 90 [22, Prop. IV.3.8, p. 283] by applying Galois cohomology to (3.2). They allow us to interpret an automorphic character \(\chi \) of T as a Hecke character for L, and we will do so frequently. In fact, distinct automorphic characters of T correspond to distinct Hecke characters of L by [3, Cor. 1.4.16, p. 606] and [3, Thm. 3.1.1, p. 619]. Since \(T\left( K_v\right) \cong \left( \prod _{w \mid v}L_w^*\right) /K_v^*\) for each \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), we see that, if \(\chi \) is unramified at v, then it is unramified as a Hecke character at all \(w \mid v\). In particular, if \(\chi \) is unramified at v and v is unramified in L/K, then \(\prod _{w \mid v} \chi _w\left( \pi _w\right) = 1\), since \(\pi _v\) is a uniformiser for \(L_w\) for each \(w \mid v\).

3.1 Geometry

In this section we study fan-theoretic objects related to T. We begin by describing the fan \(\Sigma \subset X_*\left( \overline{T}\right) _{\mathbb {R}}\) associated to the equivariant compactification \(\mathbb {P}^{d-1}_K\) of T and the associated piecewise-linear function \(\varphi _{\Sigma }\) (see [3, Sect. 1.2]) used to define the Batyrev–Tschinkel height function.

Denoting by \(l_0\left( x\right) ,\dots ,l_{d-1}\left( x\right) \in E[x]\) the E-linear factors of \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) \), we have the E-isomorphism

$$\begin{aligned} \begin{aligned} \Phi : \overline{T} = \mathbb {P}^{d-1} \setminus \bigcup _{i=0}^{d-1}Z\left( l_i\right) \xrightarrow {\sim } \mathbb {G}_m^{d-1} = \mathbb {P}^{d-1} \setminus \bigcup _{j=0}^{d-1}Z\left( x_j\right) , \\ [x_0,\dots ,x_{d-1}] \mapsto [l_0\left( x\right) ,\dots ,l_{d-1}\left( x\right) ]. \end{aligned} \end{aligned}$$

By [16, Sect. 1.1], the fan associated to \(\mathbb {P}^{d-1}_E\) as a compactification of \(\mathbb {G}_{m,E}^{d-1}\) is the fan whose r-dimensional cones are generated by the r-fold subsets of \(\{e_0',\dots ,e_{d-1}'\}\) for \(0 \le r \le d-1\), where \(e_i' \in X_*\left( \mathbb {G}_m^{d-1}\right) \cong {{\,\mathrm{Hom}\,}}\left( \mathbb {G}_m,\mathbb {G}_m^{d-1}\right) \) is defined by

$$\begin{aligned} e'_i: \mathbb {G}_m \rightarrow \mathbb {G}_m^{d-1}, \quad t \mapsto [x_{0,i}\left( t\right) ,\dots ,x_{d-1,i}\left( t\right) ], \end{aligned}$$

where in turn

$$\begin{aligned} x_{j,i}\left( t\right) = {\left\{ \begin{array}{ll} t \text { if } i=j, \\ 1 \text { otherwise.} \end{array}\right. } \end{aligned}$$

Definition 3.2

Set \(e_i = \Phi ^{-1} \circ e'_i\) for \(i=0,\dots ,d-1\), and define \(\Sigma \) to be the fan whose r-dimensional cones are generated by the r-fold subsets of \(\{e_0,\dots ,e_{d-1}\}\) for \(0 \le r \le d-1\).

It follows that \(\Sigma \) is the fan associated to \(\mathbb {P}^{d-1}_E\) as a compactification of \(T_E\). Also, we see that \(\sum _{i=0}^{d-1}e_i = 0\) and that \(\{e_1,\dots ,e_{d-1}\}\) is the dual of the basis \(\{m_1,\dots ,m_{d-1}\}\) of \(X^*\left( \overline{T}\right) \), where \(m_i\left( x\right) = \frac{l_i\left( x\right) }{l_0\left( x\right) }\) for \(i=1,\dots ,d-1\).

We now show that the action of G on \(\Sigma \left( 1\right) \) is compatible with its action on the E-linear factors of \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\). Denote by \(*\) the action of G, and set \(l_{g\left( i\right) } = g * l_i\).

Lemma 3.3

For all \(g \in G\) and \(i=0,\dots ,d-1\), we have

$$\begin{aligned} g * e_i = e_{g\left( i\right) }. \end{aligned}$$

Proof

Let \(g \in G\). It suffices to show that

$$\begin{aligned} \left( g * e_i\right) \left( m_j\right) = e_{g\left( i\right) }\left( m_j\right) \end{aligned}$$
(3.3)

for all \(i \in \{0,\dots ,d-1\}\) and \(j \in \{1,\dots ,d-1\}\). Note that, for any \(i,j,k \in \{0,\dots ,d-1\}\), we have

$$\begin{aligned} e_i\left( \frac{l_j}{l_k}\right) = \delta _{ij} - \delta _{ik}, \end{aligned}$$
(3.4)

where \(\delta _{ij}\) is the Kronecker delta symbol, defined by

$$\begin{aligned} \delta _{ij} = {\left\{ \begin{array}{ll} 1 \quad \text { if } i=j, \\ 0 \quad \text { otherwise.} \end{array}\right. } \end{aligned}$$

Then (3.3) becomes

$$\begin{aligned} \delta _{ig^{-1}\left( j\right) } - \delta _{ig^{-1}\left( 0\right) } = \delta _{g\left( i\right) j} - \delta _{g\left( i\right) 0}, \end{aligned}$$

which clearly holds. \(\square \)

By [16, Thm. 1.22, p. 217], \(\Sigma \) is the fan associated to the compactification \(\mathbb {P}^{d-1}_K = \mathbb {P}^{d-1}_E / G\) of T over K.

By [3, Prop. 1.2.12, p. 597], the line bundle \(L\left( \varphi _{\Sigma }\right) \) associated to the piecewise-linear function \(\varphi _{\Sigma }: X_*\left( \overline{T }\right) _{\mathbb {R}} \rightarrow \mathbb {R}\) (see [3, Prop. 1.2.9, p. 597]) defined by \(\varphi _{\Sigma }\left( e_i\right) = 1\) for all \(i=0,\dots ,d-1\) is the anticanonical bundle \(-K_{\mathbb {P}^{d-1}}\).

It follows from the above that G acts transitively on \(\Sigma \left( 1\right) = \{\langle e_0 \rangle , \dots , \langle e_{d-1} \rangle \}\). For \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) non-archimedean, let \(G_v\) denote the associated decomposition subgroup of G. By the proof of [3, Thm. 3.1.3, p. 619], the \(G_v\)-orbits of \(\Sigma \left( 1\right) \) are in bijection with the places of L over v, and the length of the \(G_v\)-orbit corresponding to a place \(w \mid v\) is its inertia degree.

Proposition 3.4

Let \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) be non-archimedean with ramification degree \(e_v\) in E/K, and let

$$\begin{aligned} \Sigma \left( 1\right) = \bigcup _{w \mid v}\Sigma _w\left( 1\right) \end{aligned}$$

denote the decomposition of \(\Sigma \left( 1\right) = \{\langle e_0 \rangle , \dots , \langle e_{d-1} \rangle \}\) into \(G_v\)-orbits. For each \(w \mid v\), let \(n_w\) be the sum of the elements of \(\Sigma _w\left( 1\right) \) and let \(f_w\left( x\right) \) be the product of the linear factors in the \(G_v\)-orbit of \(\{l_0,\dots ,l_{d-1}\}\) corresponding to \(\Sigma _w\left( 1\right) \) by Lemma 3.3. Then the map \(\deg _{T,E,v}: T\left( K_v\right) \rightarrow X_*\left( T_v\right) \) is given by

$$\begin{aligned} t_v \mapsto e_v \sum _{w \mid v} \frac{v\left( f_w\left( t_v\right) \right) }{\deg f_w}n_w. \end{aligned}$$

Proof

The image of \(t_v\) in \(X_*\left( T_v\right) \cong X_*\left( \overline{T}\right) ^{G_v}\) under \(\deg _{T,E,v}\) is the cocharacter

$$\begin{aligned} \varphi _{t_v}: X^*\left( T_v\right) \rightarrow \mathbb {Z}, \quad \lambda \mapsto e_v v\left( \lambda \left( t_v\right) \right) . \end{aligned}$$

We first show that \(\{n_w : w \mid v\}\) spans \(X_*\left( \overline{T}\right) ^{G_v}\). Given \(g \in G\) and \(\sigma = \sum _{i=0}^{d-1} a_i e_i\), we have \(g * \sigma = \sum _{i=0}^{d-1} a_{g^{-1}\left( i\right) }e_i\), so \(g * \sigma = \sigma \) if and only if there exists \(r_g \in \mathbb {Z}\) such that \(a_i = a_{g^{-1}\left( i\right) } + r_g\) for all \(i \in \{0,\dots ,d-1\}\). Setting \(s = \#G\), we have

$$\begin{aligned} a_i = a_{g^s\left( i\right) } = a_{g^{s-1}\left( i\right) } + r_g = \cdots = a_i + s r_g, \end{aligned}$$

hence \(r_g = 0\). We deduce that \(\sigma \in \Sigma ^{G_v}\) if and only if \(a_i = a_j\) for all \(e_i, e_j\) in the same \(G_v\)-orbit, so the result follows. Moreover, we observe that \(\sum _{w \mid v}a_w n_w = \sum _{w \mid v}b_w n_w\) if and only if there exists \(r \in \mathbb {Z}\) such that \(b_w = a_w + r\) for all \(w \mid v\), since there is a unique expression for \(\sigma \in X_*\left( \overline{T}\right) \) in the form \(\sigma = \sum _{i=0}^{d-1} c_i e_i\) where \(c_d = 0\).

Now, write

$$\begin{aligned} \varphi _{t_v} = \sum _{w \mid v}{\alpha _w n_w}. \end{aligned}$$

Define \(\mu _i \in X^*\left( \overline{T}\right) \) and \(\lambda _w \in X^*\left( \overline{T}\right) ^{G_v}\) for all \(i \in \{0,\dots ,{d-1}\}\) and all \(w \mid v\) by

$$\begin{aligned} \mu _i\left( x\right) = \frac{l_i\left( x\right) ^d}{N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) }, \quad \lambda _w\left( x\right) = \prod _{e_i \in \Sigma _w} \mu _i\left( x\right) = \frac{f_w\left( x\right) ^d}{N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) ^{\deg f_w}}. \end{aligned}$$

By (3.4), we have

$$\begin{aligned} e_i\left( \mu _j\right) = {\left\{ \begin{array}{ll} d -1 \quad \text { if } i = j, \\ -1 \qquad \text { otherwise}. \end{array}\right. } \end{aligned}$$

Then, setting \(d_w = \deg f_w\), we see that

$$\begin{aligned} n_w\left( \lambda _{w'}\right) = {\left\{ \begin{array}{ll} d d_w - d_w^2 \quad \text { if } w = w', \\ -d_w d_{w'} \qquad \text { otherwise}, \end{array}\right. } \end{aligned}$$

so we deduce that

$$\begin{aligned} e_v v\left( \lambda _w\left( t_v\right) \right) = dd_w \alpha _w - d_w\sum _{w' \mid v} d_{w'} \alpha _{w'} \end{aligned}$$
(3.5)

for all \(w \mid v\). On the other hand, we have

$$\begin{aligned} e_v v\left( \lambda _w\left( t_v\right) \right) = e_v d v\left( f_w\left( t_v\right) \right) - e_v d_w\sum _{w' \mid v} v\left( f_{w'}\left( t_v\right) \right) . \end{aligned}$$
(3.6)

Set \(\beta _w = d_w \alpha _w - e_v v\left( f_w\left( t_v\right) \right) \). Combining (3.5) and (3.6), we obtain

$$\begin{aligned} d\beta _w = d_w \sum _{w' \mid v}\beta _{w'}, \end{aligned}$$

hence \(\beta _{w'} = \frac{d_{w'}}{d_w}\beta _w\) for all \(w \mid v\), \(w' \mid v\). Since \(K_v \cong E_w^{G_v}\) for any \(w \mid v\), it follows that \(d_w \mid v\left( f_w\left( t_v\right) \right) \), so \(\beta _w \in d_w\mathbb {Z}\) for all \(w \mid v\). We deduce that there exists an integer \(n \in \mathbb {Z}\) such that, for all \(w \mid v\), we have \(\beta _w = d_w n\), hence

$$\begin{aligned} \alpha _w = e_v \frac{v \left( f_w\left( t_v\right) \right) }{\deg f_w} + n. \end{aligned}$$

Since \(\sum _{w \mid v} n_w = \sum _{i}e_i = 0\), we conclude that

$$\begin{aligned} \varphi _{t_v} = e_v \sum _{w \mid v} \frac{v\left( f_w\left( t_v\right) \right) }{\deg f_w}n_w. \end{aligned}$$

\(\square \)

We now study polynomials introduced by Batyrev and Tschinkel in [3, Sect. 2.2], which play a key role in the analysis of local Fourier transforms in Sect. 5.

Definition 3.5

Let \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) be non-archimedean, and let \(\Sigma \left( 1\right) = \bigcup _{i=1}^l\Sigma _i\left( 1\right) \) be the decomposition of \(\Sigma \left( 1\right) \) into \(G_v\)-orbits. Let \(d_i\) be the cardinality of \(\Sigma _i\left( 1\right) \). For each \(\Sigma _i\left( 1\right) \), define an independent variable \(u_i\). Let \(\sigma \in \Sigma ^{G_v}\), and let \(\Sigma _{i_1}\left( 1\right) \cup \cdots \cup \Sigma _{i_k}\left( 1\right) \) be the set of 1-dimensional faces of \(\sigma \). We define the rational function

$$\begin{aligned} R_{\sigma ,v}\left( u_1,\dots ,u_l\right) = \frac{u_{i_1}^{d_{i_1}}\cdots u_{i_k}^{d_{i_k}}}{\left( 1-u_{i_1}^{d_{i_1}}\right) \cdots \left( 1-u_{i_k}^{d_{i_k}}\right) }, \end{aligned}$$

and we define the polynomial \(Q_{\Sigma ,v}\left( u_1,\dots ,u_l\right) \) by

$$\begin{aligned} \frac{Q_{\Sigma ,v}\left( u_1,\dots ,u_l\right) }{\left( 1-u_1^{d_1}\right) \cdots \left( 1-u_l^{d_l}\right) } = \sum _{\sigma \in \Sigma ^{G_v}}R_{\sigma ,v}\left( u_1,\dots ,u_l\right) . \end{aligned}$$

Proposition 3.6

For all non-archimedean valuations \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), we have

$$\begin{aligned} Q_{\Sigma ,v}\left( u_1,\dots ,u_l\right) = 1 - u_1^{d_1}\cdots u_l^{d_l}. \end{aligned}$$

Proof

Observe that the \(G_v\)-invariant cones in \(\Sigma \) are precisely those cones generated by a set of 1-dimensional cones of the form \(\Sigma _{i_1}\left( 1\right) \cup \cdots \cup \Sigma _{i_k}\left( 1\right) \) for some \(i_1,\dots ,i_k \in \{1,\dots ,l\}\) pairwise distinct with \(k < l\). From this observation, we deduce that

$$\begin{aligned} \frac{Q_{\Sigma ,v}\left( u_1,\dots ,u_l\right) }{\left( 1-u_1^{d_1}\right) \cdots \left( 1-u_l^{d_l}\right) } = \sum _{k=1}^{l-1} \sum _{\begin{array}{c} i_1,\dots ,i_k \in \{1,\dots ,l\} \\ \text { pairwise distinct} \end{array}}\frac{u_{i_1}^{d_{i_1}}\cdots u_{i_k}^{d_{i_k}}}{\left( 1-u_{i_1}^{d_{i_1}}\right) \cdots \left( 1-u_{i_k}^{d_{i_k}}\right) }. \end{aligned}$$

In particular, we see that

$$\begin{aligned} Q_{\Sigma ,v}\left( u_1,\dots ,u_l\right) = \sum _{\left( t_1,\dots ,t_l\right) \in \{0,1\}^l} \prod _{i=1}^l\left( t_i + \left( 1-2t_i\right) u_i^{d_i}\right) - u_1^{d_1} \cdots u_l^{d_l}, \end{aligned}$$

so it suffices to prove that

$$\begin{aligned} \sum _{\left( t_1,\dots ,t_l\right) \in \{0,1\}^l} \prod _{i=1}^l\left( t_i + \left( 1-2t_i\right) u_i^{d_i}\right) = 1. \end{aligned}$$
(3.7)

Splitting the sum into two smaller sums for \(t_1 = 0\) and \(t_1 = 1\), we obtain

$$\begin{aligned} \begin{aligned}&\sum _{\left( t_1,\dots ,t_l\right) \in \{0,1\}^l} \prod _{i=1}^l\left( t_i + \left( 1-2t_i\right) u_i^{d_i}\right) \\&\quad = \left( u_1^{d_1} + \left( 1-u_1^{d_1}\right) \right) \sum _{\left( t_2,\dots ,t_n\right) \in \{0,1\}^{l-1}} \prod _{i=2}^l \left( t_i + \left( 1-2t_i\right) u_i^{d_i}\right) \\&\quad = \sum _{\left( t_2,\dots ,t_n\right) \in \{0,1\}^{l-1}} \prod _{i=2}^l \left( t_i + \left( 1-2t_i\right) u_i^{d_i}\right) . \end{aligned} \end{aligned}$$

Repeating this process for each variable \(t_2,\dots ,t_l\), we deduce (3.7). \(\square \)

3.2 Haar measures and volume

Let \(\omega \) be an invariant d-form on T. By a classical construction (see [12, Sect. 2.1.7]), \(\omega \) gives rise to a Haar measure \(|\omega |_v\) on \(T\left( K_v\right) \) for each \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \). In [23, Sect. 3.3], Ono constructs the convergence factors

$$\begin{aligned} c_v = {\left\{ \begin{array}{ll} L_v\left( X^*\left( \overline{T}\right) ,1\right) ^{-1} \quad \text { if } v \not \mid \infty , \\ 1\qquad \qquad \qquad \qquad \qquad \text { if } v \mid \infty . \end{array}\right. } \end{aligned}$$

Here, \(L_v\left( X^*\left( \overline{T}\right) ,s\right) \) is the local factor at v of the Artin L-function \(L\left( X^*\left( \overline{T}\right) ,s\right) \). Defining \(\mu _v = c_v^{-1}|\omega |_v\), the product of the \(\mu _v\) converges to give a Haar measure \(\mu \) on \(T\left( \mathbb {A}_K\right) \), which is independent of \(\omega \) by the product formula.

Note 3.7

From the short exact sequence (3.2), we obtain

$$\begin{aligned} L\left( X^*\left( \overline{T}\right) ,s\right) = \frac{\zeta _L \left( s\right) }{\zeta _K\left( s\right) }. \end{aligned}$$

Lemma 3.8

With respect to the Haar measure \(\mu \), we have

$$\begin{aligned} {{\,\mathrm{vol}\,}}\left( T\left( \mathbb {A}_K\right) /T\left( K\right) \right) = d \frac{{{\,\mathrm{Res}\,}}_{s=1}\zeta _L\left( s\right) }{{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }. \end{aligned}$$

Proof

By [23, Sect. 3.5] and [24, Main Thm., p. 68], we have

where is the In this proof, two equations appear in overly large font in the eProof, although they are the correct size in the PDF version. of T, i.e.

By [29, Prop. 8.3, p. 58] and [19, Cor. 4.6, p. 2568], the rationality of T implies that is trivial. Further, we have \({{\,\mathrm{Pic}\,}}T \cong \mathbb {Z}/d\mathbb {Z}\) (see [14, Prop. II.6.5(c), p. 133]). Since \(\zeta _K\left( s\right) \) and \(\zeta _L\left( s\right) \) both have a simple pole at \(s=1\), we have \(L\left( X^*\left( \overline{T}\right) ,1\right) = \frac{{{\,\mathrm{Res}\,}}_{s=1}\zeta _L\left( s\right) }{{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }\). Finally, as T is anisotropic, we have \(T\left( \mathbb {A}_K\right) ^1 = T\left( \mathbb {A}_K\right) \). \(\square \)

4 Heights and indicator functions

In this section we define functions which allow us to use harmonic analysis to study weak Campana points. Let L/K be an extension of number fields with K-basis \({{\,\mathrm{\varvec{\omega }}\,}}= \{\omega _0,\dots ,\omega _{d-1}\}\) and Galois closure E/K. When \(L=E\), for any \(i,j \in \{0,\dots ,d-1\}\) and \(g \in G = {{\,\mathrm{Gal}\,}}\left( E/K\right) \), write

$$\begin{aligned} \omega _i \cdot \omega _j = \sum _{k=0}^{d-1}a^{ij}_{k}\omega _k, \quad g\left( \omega _i\right) = \sum _{k=0}^{d-1}b^g_{k}\omega _k, \quad 1 = \sum _{k=0}^{d-1}c_k \omega _k. \end{aligned}$$

Definition 4.1

When \(L=E\), we define \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) to be the minimal subset of \({{\,\mathrm{Val}\,}}\left( K\right) \) containing \(S_{\infty }\) such that \(a^{ij}_k, b^g_k,c_k \in \mathcal {O}_v\) for all \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), \(i,j,k \in \{0,\dots ,d-1\}\) and \(g \in G\). Otherwise, we define \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) to be the minimal subset of \({{\,\mathrm{Val}\,}}\left( K\right) \) containing \(S_{\infty }\) such that \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) is an irreducible polynomial over \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\).

Remark 4.2

When \(L=E\), by (3.1), the \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \)-integrality of the \(a^{ij}_k\) and \(b^g_k\) implies that \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) is defined over \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\), while the \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \)-integrality of the \(c_k\) implies that the coefficients of \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) are not all divisible by some \(\alpha \in \mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) } \setminus \mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }^*\). Since \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) is irreducible over K, we deduce that \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) is irreducible over \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\), hence, for any L, the Zariski closure of \(Z\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \) in \(\mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }}\) is \({{\,\mathrm{Proj}\,}}\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }[x_0,\dots ,x_{d-1}]/\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \).

From now on, we fix the model \(\left( \mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }},\mathcal {D}^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) for \(\left( \mathbb {P}^{d-1}_K,\Delta _m^{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \), where \(\mathcal {D}^{{{\,\mathrm{\varvec{\omega }}\,}}}_m = \left( 1-\frac{1}{m}\right) {{\,\mathrm{Proj}\,}}\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }[x_0,\dots ,x_{d-1}]/\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \). We denote by \(\mathcal {D}^{{{\,\mathrm{\varvec{\omega }}\,}}}_{\text {red}}\) the support of \(\mathcal {D}^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\).

Note 4.3

In both the Galois and non-Galois cases, the conditions on \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) ensure that we may take the “obvious” model above. The potentially stronger conditions in the Galois case (in which we obtain our results) ensure compatibility between intersection multiplicity and toric multiplication, as we shall see in Sect. 4.2.

Remark 4.4

When \(L=E\) and \({{\,\mathrm{\varvec{\omega }}\,}}\) is a relative integral basis, we get \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) = S_{\infty }\), since every algebraic integer is expressible as an \(\mathcal {O}_K\)-linear combination of elements of a relative integral basis, and \(\mathcal {O}_L\) is closed under multiplication and conjugation.

4.1 Definitions

Definition 4.5

[3, Sect. 2.1] For each place v of K, we define the local height function

$$\begin{aligned} H_v : T\left( K_v\right) \rightarrow \mathbb {R}_{>0}, \quad t_v \mapsto e^{\varphi _{\Sigma }\left( \deg _{T,E,v}\left( t_v\right) \right) \log q_v}. \end{aligned}$$

We then define the global height function

$$\begin{aligned} H: T\left( \mathbb {A}_K\right) \rightarrow \mathbb {R}_{>0}, \quad \left( t_v\right) _v \mapsto \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) }H_v\left( t_v\right) . \end{aligned}$$

Definition 4.6

For each place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), define the function

$$\begin{aligned} H'_v: T\left( K_v\right) \rightarrow \mathbb {R}_{>0}, \quad x \mapsto \frac{\max \{|x_i|_v^d\}}{|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) |_v}. \end{aligned}$$

Remark 4.7

Note that \(H'_v\left( x\right) \ge 1\) for all \(x \in T\left( K_v\right) \). Indeed, one may always select v-adic coordinates \(x_i\) such that \(\max \{|x_i|_v\} = 1\), and \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) has coefficients in \(\mathcal {O}_v\) by Remark 4.2, so, by the strong triangle inequality, we have \(|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) |_v \le 1\).

Lemma 4.8

For all but finitely many places \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have \(H'_v = H_v\).

Proof

Note that \(H_v'\) is the local Weil function associated to the basis of global sections of \(-K_{\mathbb {P}^{d-1}}\) consisting of all monomials of degree d in [3, Def. 2.1.1, p. 606]. It is well-known (see [12, Sect. 2.2.3]) that two height functions corresponding to adelic metrisations of the same line bundle are equal over all but finitely many places. \(\square \)

Definition 4.9

We define the finite set

$$\begin{aligned} S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) = S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \cup \{v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) : H'_v \ne H_v\} \cup \{v \in {{\,\mathrm{Val}\,}}\left( K\right) : E/K \text { is ramified at }v\}. \end{aligned}$$

Definition 4.10

For each place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), define the local indicator function

$$\begin{aligned} \phi _{m,v} : T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto {\left\{ \begin{array}{ll} 1 \quad \text { if } H'_v\left( t_v\right) = 1 \text { or } H'_v\left( t_v\right) \ge q_v^m, \\ 0 \quad \text { otherwise}. \end{array}\right. } \end{aligned}$$

Setting \(\phi _{m,v} = 1\) for \(v \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we then define the global indicator function

$$\begin{aligned} \phi _m : T\left( \mathbb {A}_K\right) \rightarrow \{0,1\}, \quad \left( t_v\right) _v \mapsto \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \phi _{m,v}\left( t_v\right) . \end{aligned}$$

Remark 4.11

Let \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) be a non-archimedean place of K. Since \(H'_v\) is continuous with discrete image in \(\mathbb {R}_{>0}\), its level sets are clopen. It follows that \(\phi _{m,v}\) is continuous for all \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \). Also, since \(\phi _{m,v}\left( T\left( \mathcal {O}_v\right) \right) = 1\) for all \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) by Lemma 2.14(i), we see that \(\phi _m\) is well-defined and continuous on \(T\left( \mathbb {A}_K\right) \).

Lemma 4.12

The weak Campana \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\)-points of \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) are precisely the rational points \(t \in T\left( K\right) \) such that \(\phi _m\left( t\right) = 1\).

Proof

Take \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), and let \(t_0,\dots ,t_{d-1}\) be a set of \(\mathcal {O}_v\)-coordinates for \(t \in T\left( K\right) \) with at least one \(t_i \in \mathcal {O}_v^*\). Then we have

$$\begin{aligned} H'_v\left( t\right) = \frac{1}{|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( t_0,\dots ,t_{d-1}\right) |_v} = q_v^{v\left( N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( t_0,\dots ,t_{d-1}\right) \right) } = q_v^{n_v\left( \mathcal {D}^{{{\,\mathrm{\varvec{\omega }}\,}}}_{\text {red}},t\right) }. \end{aligned}$$

\(\square \)

4.2 Invariant subgroups

For this section, let \(L = E\) be Galois over K.

Lemma 4.13

For all \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) and \(x,y \in T\left( K_v\right) \), we have

$$\begin{aligned} H'_v \left( x \cdot y\right) \le H'_v \left( x\right) H'_v \left( y\right) . \end{aligned}$$

Proof

Choose sets of projective coordinates \(\{x_0,\dots ,x_{d-1}\}\) and \(\{y_0,\dots ,y_{d-1}\}\) for x and y respectively. Note that

$$\begin{aligned} \left( x_0 \omega _0 + \cdots + x_{d-1} \omega _{d-1}\right) \cdot \left( y_0 \omega _0 + \cdots + y_{d-1} \omega _{d-1}\right) = \left( z_0 \omega _0 + \cdots + z_{d-1} \omega _{d-1}\right) , \end{aligned}$$

where, for \(a^{ij}_k \in \mathcal {O}_v\) as in Definition 4.1, we have

$$\begin{aligned} z_k = \sum _{i=0}^{d-1}\sum _{j=0}^{d-1} a^{ij}_{k} x_i y_j. \end{aligned}$$

Using \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x \cdot y\right) = N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( y\right) \) and the strong triangle inequality, we deduce that

$$\begin{aligned} H'_v\left( x \cdot y\right)&= \frac{\max \{|\sum _{i=0}^{d-1}\sum _{j=0}^{d-1}a^{ij}_k x_i y_j|_v^{d}\}}{|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x \cdot y\right) |_v} \\&\le \max \{|a^{ij}_{k}|_v^d\} \frac{\max \{|x_i|_v^d\}}{|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) |_v} \frac{\max \{|y_j|_v^{d}\}}{|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( y\right) |_v} \le H'_v\left( x\right) H'_v\left( y\right) . \end{aligned}$$

\(\square \)

Lemma 4.14

For any place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), the level set

$$\begin{aligned} \mathcal {K}_v = \{t_v \in T\left( K_v\right) : H'_v\left( t_v\right) = 1\} \end{aligned}$$

is a subgroup of \(T\left( \mathcal {O}_v\right) \).

Proof

From Proposition 3.4 and Lemma 2.14(i), it is clear that \(H_v'\left( t_v\right) = 1\) implies \(t_v \in T\left( \mathcal {O}_v\right) \), so \(\mathcal {K}_v \subset T\left( \mathcal {O}_v\right) \). It is also clear that \(H'_v\left( 1\right) = 1\), and closure under multiplication follows from Lemma 4.13 and Remark 4.7. It only remains to verify that \(x \in \mathcal {K}_v\) implies \(x^{-1} \in \mathcal {K}_v\). Let \(x \in \mathcal {K}_v\), and choose coordinates \(x_0,\dots ,x_{d-1}\) with \(\max \{|x_i|_v^d\}\) = 1. Since \(H'_v\left( x\right) = 1\), we must have \(|N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) |_v = 1\). Note that

$$\begin{aligned} \left( x_0 \omega _0 + \cdots + x_{d-1} \omega _{d-1}\right) ^{-1} = \frac{1}{N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x_0,\dots ,x_{d-1}\right) }\prod _{\begin{array}{c} g \in G \\ g \ne 1_G \end{array}}\left( x_0 g\left( \omega _0\right) + \cdots + x_0 g\left( \omega _0\right) \right) . \end{aligned}$$

Recursively applying Lemma 4.13, we obtain

$$\begin{aligned} H_v'\left( x^{-1}\right) \le \prod _{\begin{array}{c} g \in G \\ g \ne 1_G \end{array}} H_v'\left( g\left( x\right) \right) . \end{aligned}$$

By Remark 4.7, it suffices to show that, for any \(g \in G\), we have \(H_v'\left( g\left( x\right) \right) = 1\). Since \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( g\left( x\right) \right) = N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) \), it suffices by Remark 4.7 to show that \(\max \{|g\left( x\right) _i|_v\} \le 1\). This follows from the fact that \(b^g_k \in \mathcal {O}_v\) since \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), see Definition 4.1. \(\square \)

Corollary 4.15

For every place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), the function \(H'_v\) is \(\mathcal {K}_v\)-invariant.

Proof

Take \(x \in \mathcal {K}_v\), and let \(y \in T\left( K_v\right) \). Then by Lemma 4.13, we have

$$\begin{aligned} H'_v\left( x \cdot y\right) \le H'_v\left( x\right) H'_v \left( y\right) = H'_v\left( y\right) , \end{aligned}$$

while on the other hand, since \(x^{-1} \in \mathcal {K}_v\) by Lemma 4.14, we have

$$\begin{aligned} H'_v\left( y\right) = H'_v \left( x^{-1} \cdot \left( x \cdot y\right) \right) \le H'_v \left( x^{-1}\right) H'_v \left( x \cdot y\right) = H'_v \left( x \cdot y\right) , \end{aligned}$$

so we conclude that \(H'_v \left( x \cdot y\right) = H'_v\left( y\right) \). \(\square \)

Lemma 4.16

For each place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), the functions \(H_v\) and \(\phi _{m,v}\) are both \(\mathcal {K}_v\)-invariant and 1 on \(\mathcal {K}_v\). Further, \(\mathcal {K}_v\) is compact, open and of finite index in \(T\left( \mathcal {O}_v\right) \). Moreover, when \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have \(\mathcal {K}_v = T\left( \mathcal {O}_v\right) \).

Proof

Let \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \). By [3, Thm. 2.1.6(i), p. 608], \(H_v\) is \(T\left( \mathcal {O}_v\right) \)-invariant, hence trivial and invariant on all of \(T\left( \mathcal {O}_v\right) \). By Corollary 4.15, the function \(\phi _{m,v}\) is \(\mathcal {K}_v\)-invariant; since \(\phi _{m,v}\left( 1\right) = 1\), it is also trivial on \(\mathcal {K}_v\).

Now, since \(\mathcal {K}_v = \left( H_v'|_{T\left( \mathcal {O}_v\right) }\right) ^{-1}\left( \{1\}\right) \), it is open. Since the cosets of an open subgroup form an open cover of a topological group, any open subgroup of a compact topological group is closed and of finite index. Then \(\mathcal {K}_v \subset T\left( \mathcal {O}_v\right) \) is closed, hence compact, and of finite index. Finally, we note that, when \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have \(H'_v = H_v\), and \(H_v^{-1}\left( \{1\}\right) = T\left( \mathcal {O}_v\right) \) by Lemma 2.14(i) and the equality \(\varphi ^{-1}_{\Sigma }\left( \{0\}\right) = \{0\}\), so \(\mathcal {K}_v = T\left( \mathcal {O}_v\right) \). \(\square \)

Definition 4.17

For each \(v \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), set \(\mathcal {K}_v = T\left( \mathcal {O}_v\right) \). Let \(\mathcal {K} = \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \mathcal {K}_v\), and let \(\mathcal {U}\) be the group of automorphic characters of T which are trivial on \(\mathcal {K}\).

4.3 Height zeta function and Fourier transforms

Definition 4.18

For \({\text {Re}}s \gg 0\), we define the height zeta function

$$\begin{aligned} Z_m : \mathbb {C} \rightarrow \mathbb {C}, \quad s \mapsto \sum _{x \in \mathbb {P}^{d-1}\left( K\right) } \frac{\phi _{m}\left( x\right) }{H\left( x\right) ^s}. \end{aligned}$$

Definition 4.19

Let \(\mu _v\) and \(\mu \) be the Haar measures introduced in Sect. 3.2. Let \(f: T\left( \mathbb {A}_K\right) \rightarrow \mathbb {C}\) be a continuous function given as a product of local factors \(f_v: T\left( K_v\right) \rightarrow \mathbb {C}\) such that \(f_v\left( T\left( \mathcal {O}_v\right) \right) = 1\) for all but finitely many places \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \). For each place \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \) and each character \(\chi \) of \(T\left( \mathbb {A}_K\right) \), we define the local Fourier transform of \(\chi _v\) with respect to \(f_v\) to be

$$\begin{aligned} \widehat{H}_v\left( f_v,\chi _v;-s\right) = \int _{T\left( K_v\right) }\frac{f_v\left( t_v\right) \chi _v\left( t_v\right) }{H_v\left( t_v\right) ^s}d\mu _v \end{aligned}$$

for all \(s \in \mathbb {C}\) for which the integral exists. We then define the global Fourier transform of \(\chi \) with respect to f to be

$$\begin{aligned} \widehat{H}\left( f,\chi ;-s\right) = \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \widehat{H}_v\left( f_v,\chi _v;-s\right) = \int _{T\left( \mathbb {A}_K\right) }\frac{f\left( t\right) \chi \left( t\right) }{H\left( t\right) ^s}d\mu . \end{aligned}$$

5 Weak Campana points

In this section we prove Theorem 1.1. Fix an extension of number fields L/K of degree d with K-basis \({{\,\mathrm{\varvec{\omega }}\,}}\), set \(T = T_{{{\,\mathrm{\varvec{\omega }}\,}}}\) as in Sect. 3 and let \(m \in \mathbb {Z}_{\ge 2}\).

5.1 Strategy

Following [3, 19], we will apply a Tauberian theorem [3, Thm. 3.3.2, p. 624] to our height zeta function \(Z_m\left( s\right) \) in order to find an asymptotic for the number of weak Campana points of bounded height. By loc. cit., it suffices to show that \(Z_m\left( s\right) \) is absolutely convergent for \({\text {Re}}s > \frac{1}{m}\) and that \(Z_m\left( s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\) admits an extension to a holomorphic function on \({\text {Re}}s \ge \frac{1}{m}\) which is not zero at \(s = \frac{1}{m}\). In order to do this, we will apply the version of the Poisson summation formula given by Bourqui [4, Thm. 3.35, p. 64]. Formally applying this version with \(\mathcal {G} = T\left( \mathbb {A}_K\right) \), \(\mathcal {H} = T\left( K\right) \), \(dg = d\mu \), dh the discrete measure on \(T\left( K\right) \) and \(F\left( t\right) = \frac{\phi _m\left( t\right) }{H\left( t\right) ^s}\) for some \(s \in \mathbb {C}\) with \({\text {Re}}s > \frac{1}{m}\) gives

$$\begin{aligned} Z_m\left( s\right) = \frac{1}{{{\,\mathrm{vol}\,}}\left( T\left( \mathbb {A}_K\right) /T\left( K\right) \right) } \sum _{\chi \in \left( T\left( \mathbb {A}_K\right) /T\left( K\right) \right) ^{\wedge } } \widehat{H}\left( \phi _m, \chi ; -s\right) . \end{aligned}$$
(5.1)

5.2 Analytic properties of Fourier transforms

Lemma 5.1

For any place \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), any character \(\chi _v\) of \(T\left( K_v\right) \) and any \(\epsilon > 0\), the local Fourier transform \(\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \) is absolutely convergent and is bounded uniformly (in terms of \(\epsilon \) and v) on \({\text {Re}}s \ge \epsilon \).

Proof

Let \({\text {Re}}s \ge \epsilon \). Since

$$\begin{aligned} |\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) | \le \int _{T\left( K_v\right) }\left| \frac{\phi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) }{H_v\left( t_v\right) ^s}\right| d\mu _v \le \widehat{H}_v\left( 1,1;-\epsilon \right) , \end{aligned}$$

it suffices to prove that \(\widehat{H}_v\left( 1,1;-\epsilon \right) \) is convergent. For \(v \mid \infty \), this follows from [3, Prop. 2.3.2, p. 614], so assume that \(v \not \mid \infty \). The following argument is essentially the one in [3, Rem. 2.2.8, p. 613], but we fill in the details for the sake of clarity.

Since \(H_v\) and \(d\mu _v\) are \(T\left( \mathcal {O}_v\right) \)-invariant and \(\int _{T\left( \mathcal {O}_v\right) }d\mu _v = 1\), we have

$$\begin{aligned} \begin{aligned} \widehat{H}_v\left( 1,1;-\epsilon \right) = \int _{T\left( K_v\right) }\frac{1}{H_v\left( t_v\right) ^\epsilon }d\mu _v = \sum _{\overline{t_v} \in T\left( K_v\right) /T\left( \mathcal {O}_v\right) }\frac{1}{H_v\left( \overline{t_v}\right) ^{\epsilon }}. \end{aligned} \end{aligned}$$

Now, by Lemma 2.14(i), \(T\left( K_v\right) /T\left( \mathcal {O}_v\right) \) can be identified with a sublattice of finite index in \(X_*\left( T_v\right) \), and this sublattice coincides with \(X_*\left( T_v\right) \) when v is unramified in L/K. Then we see that, interpreting \(H_v\) as a function on \(X_*\left( T_v\right) \), we have

$$\begin{aligned} \sum _{\overline{t_v} \in T\left( K_v\right) /T\left( \mathcal {O}_v\right) }\frac{1}{H_v\left( \overline{t_v}\right) ^{\epsilon }} \le \sum _{n_v \in X_*\left( T_v\right) }\frac{1}{H_v\left( n_v\right) ^{\epsilon }}, \end{aligned}$$

and the proof of [3, Thm. 2.2.6, p. 611] and Proposition 3.6 give

$$\begin{aligned} \sum _{n_v \in X_*\left( T_v\right) }\frac{1}{H_v\left( n_v\right) ^{\epsilon }} = \left( 1-\frac{1}{q_v^{d\epsilon }}\right) \prod _{w \mid v}\left( 1-\frac{1}{q_w^{\epsilon }}\right) ^{-1}, \end{aligned}$$

so we deduce that \(\widehat{H}_v\left( 1,1;-\epsilon \right) \) is convergent, and this concludes the proof. \(\square \)

Lemma 5.2

For any \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), the local Fourier transform \(\widehat{H}_v\left( \phi _{m,v},1;-s\right) \) is non-zero for all \(s \in \mathbb {R}_{> 0}\).

Proof

The proof is analogous to the proof of [19, Lem. 5.1, p. 2575]. \(\square \)

Lemma 5.3

Let \(L = E\) be a Galois extension of K. For any place \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), let \(\chi _v\) be a character of \(T\left( K_v\right) \) which is non-trivial on \(\mathcal {K}_v\). Then

$$\begin{aligned} \widehat{H}_v \left( \phi _{m,v},\chi _v;-s\right) = 0. \end{aligned}$$

Proof

Since \(\phi _{m,v}\) and \(H_v\) are \(\mathcal {K}_v\)-invariant, the result follows by character orthogonality. \(\square \)

Corollary 5.4

Let \(L = E\) be a Galois extension of K, and let \(\chi \) be an automorphic character of T. If \(\chi \not \in \mathcal {U}\) for \(\mathcal {U}\) as in Definition 4.17, then

$$\begin{aligned} \widehat{H}\left( \phi _m,\chi ;-s\right) = 0. \end{aligned}$$

Lemma 5.5

Let \(v \not \mid \infty \) be a non-archimedean place of K unramified in L/K, and let \(\chi \) be an automorphic character of T which is unramified at v. Then we have

$$\begin{aligned} \widehat{H}_v\left( 1,\chi _v;-s\right) = \left( 1-\frac{1}{q_v^{ds}}\right) \prod _{w \mid v}\left( 1-\frac{\chi _w\left( \pi _w\right) }{q_w^s}\right) ^{-1} = L_v \left( \chi ,s\right) \zeta _{K,v}\left( ds\right) ^{-1}. \end{aligned}$$

Proof

The result follows from [3, Thm. 2.2.6, p. 611] and Proposition 3.6. \(\square \)

Definition 5.6

Given a vector \(u = \left( u_1,\dots ,u_r\right) \in \mathbb {N}^r\), define \(f_{r,n,u}\left( x_1\dots ,x_r\right) \) to be the sum of all degree-n monomials in \(x_1,\dots ,x_r\) weighted via u, i.e.

$$\begin{aligned} f_{r,n,u}\left( x_1,\dots ,x_r\right) = \sum _{\begin{array}{c} \sum _{i=1}^r u_i a_i = n \\ \forall i \, a_i \in \mathbb {Z}_{\ge 0} \end{array}} x_1^{a_1}\dots x_r^{a_r}. \end{aligned}$$

Set

$$\begin{aligned} f_{r,n}\left( x_1,\dots ,x_r\right) = f_{r,n,\left( 1,\dots ,1\right) }\left( x_1,\dots ,x_r\right) . \end{aligned}$$

Proposition 5.7

Let \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) be a non-archimedean place of K, and let \(\chi \in \mathcal {U}\). Let \(w_1,\dots ,w_r \in {{\,\mathrm{Val}\,}}\left( L\right) \) be the places of L over v. Let \(u_i\) be the inertia degree of \(w_i\) over v for each \(i=1,\dots ,r\). Set

$$\begin{aligned} c_{\chi ,v,n} = f_{r,n,u}\left( \chi _{w_1}\left( \pi _{w_1}\right) ,\dots ,\chi _{w_r}\left( \pi _{w_r}\right) \right) . \end{aligned}$$

Then, for \({\text {Re}}s > 0\), we have

$$\begin{aligned} \widehat{H}_v\left( \phi _{m,v};\chi _v;-s\right) = 1 + \sum _{n=m}^{\infty } \frac{c_{\chi ,v,n} - c_{\chi ,v,n-d}}{q_v^{ns}}. \end{aligned}$$

Proof

Let \(s \in \mathbb {C}\) with \({\text {Re}}s > 0\). As \(\chi \in \mathcal {U}\) and \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), it follows that \(\chi \) is unramified at v. Then, expanding geometric series, we have

$$\begin{aligned} L_v\left( \chi ,s\right) = \prod _{i=1}^r\left( 1-\frac{\chi _{w_i}\left( \pi _{w_i}\right) }{q_v^{u_i s}}\right) ^{-1} = 1+ \sum _{n=1}^{\infty }\frac{c_{\chi ,v,n}}{q_v^{ns}}, \end{aligned}$$

so, by Lemma 5.5, we obtain

$$\begin{aligned} \widehat{H}_v\left( 1,\chi _v;-s\right) = 1 + \sum _{n=1}^{\infty } \frac{c_{\chi ,v,n} - c_{\chi ,v,n-d}}{q_v^{ns}}. \end{aligned}$$

On the other hand, we may write

$$\begin{aligned} \widehat{H}_v\left( 1,\chi _v;-s\right) = \int _{T\left( K_v\right) }\frac{\chi _v\left( t_v\right) }{H_v\left( t_v\right) ^s}d\mu _v = \sum _{n=0}^{\infty } \frac{1}{q_v^{ns}}\int _{H_v\left( t_v\right) = q_v^n}\chi _v\left( t_v\right) d\mu _v, \end{aligned}$$

so, comparing these expressions, we see for \(n \ge 1\) that

$$\begin{aligned} c_{\chi ,v,n} - c_{\chi ,v,n-d} = \int _{H_v\left( t_v\right) = q_v^n}\chi _v\left( t_v\right) d\mu _v. \end{aligned}$$

Since \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have \(\phi _{m,v}\left( t_v\right) = 1\) if and only if \(H_v\left( t_v\right) = 1\) or \(H_v\left( t_v\right) \ge q_v^m\), so the result follows.

5.3 Regularisation

Now that we have expressions for the local Fourier transforms at all but finitely many places, our goal is to find “regularisations” for the global Fourier transforms, i.e. functions expressible as Euler products whose convergence is well-understood and whose local factors approximate the local Fourier transforms well (as expansions in \(q_v\)) at all but finitely many places. As in [3, 19, 27], we will construct our regularisations from L-functions.

Proposition 5.8

Let G be a subgroup of \(S_d\) acting freely and transitively on \(\{1,\dots ,d\}\), and let \(m \ge 2\) be a positive integer. Let \(S_m\) act upon \(G^m\) by permutation of coordinates, and let G act on \(G^m/S_m\) by right multiplication of every element of a representative m-tuple. Set \(S\left( G,m\right) = \left( G^m/S_m\right) /G\).

  1. (i)

    If \(\overline{\left( g_1,\dots ,g_m\right) } = \overline{\left( h_1,\dots ,h_m\right) }\) in \(S\left( G,m\right) \), then

    $$\begin{aligned} \sum _{i=1}^d x_{g_1\left( i\right) }\cdots x_{g_m\left( i\right) } = \sum _{i=1}^d x_{h_1\left( i\right) }\cdots x_{h_m\left( i\right) }, \end{aligned}$$
    (5.2)

    hence, for \(\overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) \), we may define the sum

    $$\begin{aligned} \phi _{\overline{\left( g_1,\dots ,g_m\right) }}\left( x_1,\dots ,x_d\right) = \sum _{i=1}^d x_{g_1\left( i\right) }\cdots x_{g_m\left( i\right) }. \end{aligned}$$
  2. (ii)

    If \(\overline{\left( g_1,\dots ,g_m\right) } \ne \overline{\left( h_1,\dots ,h_m\right) }\), then the sums \(\phi _{\overline{\left( g_1,\dots ,g_m\right) }}\left( x_1,\dots ,x_d\right) \) and \(\phi _{\overline{\left( h_1,\dots ,h_m\right) }}\left( x_1,\dots ,x_d\right) \) share no summands. Further, a monomial appears twice in \(\phi _{\overline{\left( g_1,\dots ,g_m\right) }}\left( x_1,\dots ,x_d\right) \) if and only if \(\left( g_1,\dots ,g_m\right) = \left( g_1 g_r,\dots ,g_m g_r\right) \) for some \(r \in \{1,\dots ,m\}\) with \(g_r \ne 1_G\).

  3. (iii)

    If d is coprime to m, then we have

    $$\begin{aligned} f_{d,m}\left( x_1,\dots ,x_d\right) = \sum _{\overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) } \phi _{\left( g_1,\dots ,g_m\right) }\left( x_1,\dots ,x_d\right) . \end{aligned}$$
  4. (iv)

    If d is prime and \(m = kd\), then we have

    $$\begin{aligned} f_{d,m}\left( x_1,\dots ,x_d\right) + \left( d-1\right) x_1^k \cdots x_d^k = \sum _{\overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) } \phi _{\left( g_1,\dots ,g_m\right) }\left( x_1,\dots ,x_d\right) . \end{aligned}$$

Proof

First, we prove (i). Note that \(\overline{\left( g_1,\dots ,g_m\right) } = \overline{\left( h_1,\dots ,h_m\right) }\) if and only if \(\{h_1,\dots ,h_m\} = \{g_1g,\dots ,g_m g\}\) as multisets for some \(g \in G\). If the coordinates of \(\left( h_1,\dots ,h_m\right) \in G^m\) are a permutation of those of \(\left( g_1,\dots ,g_m\right) \in G^m\), then

$$\begin{aligned} x_{h_1\left( i\right) }\cdots x_{h_m\left( i\right) } = x_{g_1\left( i\right) }\cdots x_{g_m\left( i\right) } \end{aligned}$$

for any \(i \in \{1,\dots ,d\}\), while if \(\left( h_1,\dots ,h_m\right) = \left( g_1g,\dots ,g_m g\right) \) for some \(g \in G\), then for any \(i \in \{1,\dots ,d\}\), we have

$$\begin{aligned} x_{g_1 g\left( i\right) }\cdots x_{g_m g\left( i\right) } = x_{g_1\left( j\right) }\cdots x_{g_m\left( j\right) } \end{aligned}$$

for the unique j with \(g\left( i\right) = j\). In either case, we obtain (5.2). The claim follows.

Next we prove (ii). Note that for \(\overline{\left( g_1,g_2,\dots ,g_m\right) } \in S\left( G,m\right) \), we may take \(g_1 = 1_G\) without loss of generality. Suppose that

$$\begin{aligned} x_{i}x_{g_2\left( i\right) }\cdots x_{g_m\left( i\right) } = x_{j} x_{h_2\left( j\right) }\cdots x_{h_m\left( j\right) } \end{aligned}$$

for some \(i,j \in \{1,\dots ,d\}\). This is equivalent to the equality of multisets

$$\begin{aligned} \{i,g_2\left( i\right) ,\dots ,g_m\left( i\right) \} = \{j,h_2\left( j\right) ,\dots ,h_m\left( j\right) \}. \end{aligned}$$

If \(i=j\), we have

$$\begin{aligned} \{g_2\left( i\right) ,\dots ,g_m\left( i\right) \} = \{h_2\left( i\right) ,\dots ,h_m\left( i\right) \}, \end{aligned}$$

and by the freeness of the action of G on \(\{1,\dots , d\}\), we have \(\left( g_2,\dots ,g_m\right) = \left( h_2,\dots ,h_m\right) \) up to reordering, i.e. the m-tuple \(\left( 1_G,h_2,\dots ,h_m\right) \) is a permutation of \(\left( 1_G,g_2,\dots ,g_m\right) \). If \(i \ne j\), we may take \(g_2\left( i\right) = j\) without loss of generality (note that \(g_2 \ne 1_G\) in this case), and we obtain the equality of multisets

$$\begin{aligned} \{j, h_2\left( j\right) ,\dots ,h_m\left( j\right) \} = \{g_2\left( i\right) , h_2 g_2 \left( i\right) ,\dots ,h_m g_2\left( i\right) \}. \end{aligned}$$

Once again, by freeness of the action of G on \(\{1,\dots ,d\}\), we get that

$$\begin{aligned} \{1_G,g_2,\dots ,g_m\} = \{g_2,h_2 g_2, \dots , h_m g_2\} \end{aligned}$$

as multisets, from which both claims follow.

Before proceeding to the proofs of (iii) and (iv), we make the following observation: if a multiset \(M = \{g_1,\dots ,g_m\}\) of elements of G is closed under right multiplication by some \(g \in G\) of order \(n \ge 2\), then n divides m. Indeed, M must contain the n distinct elements \(g_1, g_1 g, g_1 g^2,\dots ,g_1 g^{n-1}\). Without loss of generality, we may take \(g_l = g_1 g^{l-1}\) for \(l =1,\dots ,n\). Since \(\{g_1,\dots ,g_n\} = \{g_1,g_1 g,\dots , g_1 g^{n-1}\}\) is clearly closed under right multiplication by g, so is the sub-multiset \(\{g_{n+1},\dots ,g_m\}\). Iterating this process, we deduce that n divides the cardinality of M, as otherwise we would eventually obtain a non-empty sub-multiset closed under right multiplication by g with fewer than n elements, which is clearly impossible.

We now prove (iii) and (iv). By (i) and (ii), every degree-m monomial in \(x_1,\dots ,x_d\) appears in a unique summand of \(\sum _{\overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) } \phi _{\left( g_1,\dots ,g_m\right) }\left( x_1,\dots ,x_d\right) \), and a monomial appears twice in \(\phi _{\overline{\left( g_1,\dots ,g_m\right) }}\left( x_1,\dots ,x_d\right) \) if and only if the multiset \(\{g_1,\dots ,g_m\}\) is closed under right multiplication by some \(g_r \ne 1_G\). By the above, the latter is possible only if the order of \(g_r\) divides m, while by Lagrange’s theorem, the order of \(g_r\) divides d. If d is coprime to m, then we deduce that no such \(g_r\) exists, thus we obtain (iii). If \(d=p\) is prime and \(m=kp\), then \(g_r\) is necessarily a generator of the cyclic group G. It then follows as in the above observation that

$$\begin{aligned} \overline{\left( g_1,\dots ,g_{kp}\right) } = \overline{\left( 1_G,g_r,\dots ,g_r^{p-1},\dots ,1_G,g_r,\dots ,g_r^{p-1}\right) }. \end{aligned}$$

Letting g be a generator of G, we see that \(\phi _{\overline{\left( 1_G,g,\dots ,g^{p-1},\dots ,1_G,g,\dots ,g^{p-1}\right) }}\left( x_1,\dots ,x_p\right) = px_1^k \cdots x_p^k\) is the only one of the polynomials \(\phi _{\overline{\left( g_1,\dots ,g_{kp}\right) }}\left( x_1,\dots ,x_p\right) \), \(\overline{\left( g_1,\dots ,g_{kp}\right) } \in S\left( G,kp\right) \) in which a monomial appears more than once, and so we obtain (iv). \(\square \)

Remark 5.9

It follows immediately from Proposition 5.8 that

$$\begin{aligned} \#S\left( G,m\right) = {\left\{ \begin{array}{ll} \frac{1}{d}{{d+m-1}\atopwithdelims (){d-1}} \text { if }d\text { and }m\text { are coprime}, \\ \frac{1}{d}\left( {{d+m-1}\atopwithdelims (){d-1}} - 1\right) + 1 \text { if } d\text { is prime and} d\text { divides }m, \end{array}\right. } \end{aligned}$$

since the number of monomials of degree m in d variables is \({{d+m-1}\atopwithdelims (){d-1}}\).

Remark 5.10

As we shall shortly see, it is the ability to partition (or in the case where d is a prime dividing m, nearly partition) the sum of all degree-m monomials in d variables as in Proposition 5.8 that allows us to construct well-behaved regularisations.

For the rest of this section, let \(L = E\) be Galois over K with Galois group G, and assume that m is coprime to d if d is not prime.

Lemma 5.11

Let \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) be a non-archimedean place which is totally split in E/K and let \(\chi \in \mathcal {U}\). Then

$$\begin{aligned} \frac{\prod _{\overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) } L_v\left( \chi ^{g_1}\cdots \chi ^{g_m},ms\right) }{\prod _{\overline{\left( h_1,\dots ,h_{m-d}\right) } \in S\left( G,m-d\right) } L_v\left( \chi ^{h_1}\cdots \chi ^{h_{m-d}},ms\right) } = \prod _{\overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) } L_v\left( \chi ^{g_1}\cdots \chi ^{g_m},ms\right) , \end{aligned}$$

where

$$\begin{aligned} S'\left( G,m\right) = \left\{ \overline{\left( g_1,\dots ,g_m\right) } \in S\left( G,m\right) : \#\{g_1,\dots ,g_m\} \le d-1\right\} . \end{aligned}$$

Proof

Let \(G=\{g_1,\dots ,g_d\}\). First, we show that every factor of the denominator of the left-hand side appears on the numerator. Let \(\overline{\left( h_1,\dots ,h_{m-d}\right) } \in S\left( G,m-d\right) \). Then we claim that

$$\begin{aligned} L_v\left( \chi ^{h_1} \cdots \chi ^{h_{m-d}},ms\right) = L_v\left( \chi ^{h_1} \cdots \chi ^{h_{m-d}} \chi ^{g_1} \cdots \chi ^{g_d},ms\right) . \end{aligned}$$

Since G acts freely and transitively on the places \(w_1,\dots ,w_d\) of E over v, we have

$$\begin{aligned} \{\chi _{w_i}^{g_1}\left( \pi _{w_i}\right) ,\dots ,\chi _{w_i}^{g_d}\left( \pi _{w_i}\right) \} = \{\chi _{w_1}\left( \pi _{w_1}\right) ,\dots ,\chi _{w_d}\left( \pi _{w_d}\right) \} \end{aligned}$$

for any \(i=1,\dots ,d\). Since \(\chi _v\) is trivial on \(\mathcal {K}_v = T\left( \mathcal {O}_v\right) \), we have from Note 3.1 that \(\prod _{i=1}^d\chi _{w_i}\left( \pi _{w_i}\right) = \chi _v\left( 1\right) = 1\), so \(\left( \chi ^{g_1}\cdots \chi ^{g_d}\right) _w\left( \pi _w\right) =1\) for all \(w \mid v\). Then the equality follows.

It now suffices to show that, for \(\overline{\left( h_1,\dots ,h_{m-d}\right) } \ne \overline{\left( h'_1,\dots ,h'_{m-d}\right) }\), we have

$$\begin{aligned} \overline{\left( h_1,\dots ,h_{m-d},g_1,\dots ,g_d\right) } \ne \overline{\left( h'_1,\dots ,h'_{m-d},g_1,\dots ,g_d\right) }. \end{aligned}$$

If not, then \(\{h'_1,\dots ,h'_{m-d},g_1,\dots ,g_d\} = \{gh_1,\dots ,gh_{m-d},gg_1,\dots ,gg_d\}\) as multisets for some \(g \in G\). Since we have \(\{gg_1,\dots ,gg_d\} = \{g_1,\dots ,g_d\}\), this implies that \(\{h'_1,\dots ,h'_{m-d}\} = \{gh_1,\dots ,gh_{m-d}\}\) as multisets, but then we have \(\overline{\left( h_1,\dots ,h_{m-d}\right) } = \overline{\left( h'_1,\dots ,h'_{m-d}\right) }\), which is false. \(\square \)

Remark 5.12

It follows from the proof of Lemma 5.11 that \(\#S'\left( G,m\right) = S\left( G,m\right) - S\left( G,m-d\right) \). Combining this with Remark 5.9, we obtain

$$\begin{aligned} \#S'\left( G,m\right) = \frac{1}{d}\left( {{d+m-1}\atopwithdelims (){d-1}} - {{m-1}\atopwithdelims (){d-1}}\right) = b\left( d,m\right) . \end{aligned}$$

Note that the term \(\frac{1}{d}{{m-1}\atopwithdelims (){d-1}}\) only appears when \(d \le m\).

Definition 5.13

For all \(\chi \in \mathcal {U}\), \({\text {Re}}s > 0\) and non-archimedean places \(v \not \mid \infty \), set

$$\begin{aligned} F_{m,\chi ,v}\left( s\right) = \prod _{\overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) } L_v\left( \chi ^{g_1}\cdots \chi ^{g_m},ms\right) , \quad G_{m,\chi ,v}\left( s\right) = \frac{\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) }{F_{m,\chi ,v}\left( s\right) }, \end{aligned}$$

and define

$$\begin{aligned}&F_{m,\chi }\left( s\right) = \prod _{v \not \mid \infty } F_{m,\chi ,v}\left( s\right) = \prod _{\overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) } L\left( \chi ^{g_1}\cdots \chi ^{g_m},ms\right) , \nonumber \\&G_{m,\chi }\left( s\right) = \prod _{v \not \mid \infty } G_{m,\chi ,v}\left( s\right) . \end{aligned}$$
(5.3)

For any non-archimedean place \(v \not \mid \infty \), write

$$\begin{aligned} \widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) = \sum _{n=0}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}}, \end{aligned}$$

where \(a_{\chi ,v,n} = \int _{H_v\left( t_v\right) = q_v^n}\phi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) d\mu _v\), and write

$$\begin{aligned} F_{m,\chi ,v}\left( s\right) = 1 + \sum _{n=1}^{\infty }\frac{b_{\chi ,v,mn}}{q_v^{mns}} \end{aligned}$$

for the expansion of \(F_{m,\chi ,v}\left( s\right) \) as a multidimensional geometric series in \(q_v^{ms}\), so

$$\begin{aligned} G_{m,\chi ,v}\left( s\right) = \frac{\sum _{n=0}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}}}{ 1 + \sum _{n=1}^{\infty }\frac{b_{\chi ,v,mn}}{q_v^{mns}}} = \sum _{n=0}^{\infty } \frac{d_{\chi ,v,n}}{q_v^{ns}}, \end{aligned}$$

where \(d_{\chi ,v,n}\) is defined for all \(n \ge 0\) by the iterative formula

$$\begin{aligned} d_{\chi ,v,n} = a_{\chi ,v,n} - \sum _{r=1}^{\lfloor \frac{n}{m} \rfloor } b_{\chi ,v,mr}d_{\chi ,v,n-mr}. \end{aligned}$$
(5.4)

In particular, we have \(d_{\chi ,v,n} = a_{\chi ,v,n}\) for \(0 \le n \le m-1\).

Corollary 5.14

For \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) a non-archimedean place, we have \(d_{\chi ,v,0} = 1\) and \(d_{\chi ,v,n} = 0\) for all \(n \in \{1,\dots ,m\}\).

Proof

Let \(c_{\chi ,v,n}\) be as defined in Proposition 5.7. Since \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have from loc. cit. that \(a_{\chi ,v,0} =1\), \(a_{\chi ,v,n} = 0\) for \(n \in \{1,\dots ,m-1\}\) and \(a_{\chi ,v,m} = c_{\chi ,v,m} - c_{\chi ,v,m-d}\). Then, by (5.4), we see that \(d_{\chi ,v,0} = a_{\chi ,v,0} = 1\) and \(d_{\chi ,v,n} = a_{\chi ,v,n} = 0\) for \(1 \le n \le m-1\). Further, we obtain

$$\begin{aligned} d_{\chi ,v,m} = a_{\chi ,v,m} - b_{\chi ,v,m}d_{\chi ,v,0} = c_{\chi ,v,m} - c_{\chi ,v,m-d} - b_{\chi ,v,m}, \end{aligned}$$

so, to complete the proof, it suffices to show that \(b_{\chi ,v,m} = c_{\chi ,v,m} - c_{\chi ,v,m-d}\).

Since E/K is Galois, all of the places \(w_1,\dots ,w_r\) of E over v share a common inertia degree \(d_v\). Since \(\chi _v\left( T\left( \mathcal {O}_v\right) \right) = 1\), it is unramified as a Hecke character at all of the \(w_i\) (see Note 3.1), and for any \(g_1,\dots ,g_m \in G\), so is \(\chi ^{g_1}\cdots \chi ^{g_m}\). Then

$$\begin{aligned} \begin{aligned} L_v\left( \chi ^{g_1} \cdots \chi ^{g_m},ms\right)&= \prod _{i=1}^r\left( 1-\frac{\left( \chi ^{g_1}\cdots \chi ^{g_m}\right) _{w_i}\left( \pi _{w_i}\right) }{q_v^{d_v ms}}\right) ^{-1} \\&= 1 + \frac{1}{q_v^{d_v ms}} \sum _{i=1}^r \left( \chi ^{g_1}\cdots \chi ^{g_m}\right) _{w_i} \left( \pi _{w_i}\right) + O\left( \frac{1}{q_v^{\left( d_v m+1\right) s}}\right) . \end{aligned} \end{aligned}$$
(5.5)

First, suppose that v is totally split in E/K. Then (5.5) gives

$$\begin{aligned} L_v\left( \chi ^{g_1} \cdots \chi ^{g_m},ms\right) = 1 + \frac{ \phi _{\left( g_1,\dots ,g_m\right) } \left( \chi _{w_1}\left( \pi _{w_1}\right) ,\dots ,\chi _{w_d}\left( \pi _{w_d}\right) \right) }{q_v^{ms}} + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) . \end{aligned}$$

Since G acts freely and transitively on the \(w_i\), it follows from Proposition 5.8 and Lemma 5.11 that \(b_{\chi ,v,m} = c_{\chi ,v,m} - c_{\chi ,v,m-d}\), and so \(d_{\chi ,v,m} = 0\).

Now assume that v is not totally split in E/K. If \(\gcd \left( d,m\right) = 1\), then \(c_{\chi ,v,m} = c_{\chi ,v,m-d} = 0\), as \(c_{\chi ,v,n} = 0\) whenever \(d_v \not \mid n\). If d is prime, then v is inert and we have \(\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) = 1\) since \(T\left( K_v\right) = T\left( \mathcal {O}_v\right) \). Then, in either case, we have \(c_{\chi ,v,m} - c_{\chi ,v,m-d} = 0\), and (5.5) implies that \(b_{\chi ,v,m} = 0\), hence \(d_{\chi ,v,m} = 0\). \(\square \)

Corollary 5.15

For any \(\chi \in \mathcal {U}\), we have

$$\begin{aligned} \widehat{H}\left( \phi _{m};\chi ;-s\right) = \prod _{v \mid \infty }\widehat{H}_v\left( 1,\chi _v;-s\right) F_{m,\chi }\left( s\right) G_{m,\chi }\left( s\right) , \end{aligned}$$

where \(G_{m,\chi }\left( s\right) \) is holomorphic and uniformly bounded with respect to \(\chi \) for \({\text {Re}}s \ge \frac{1}{m}\) and \(G_{m,1}\left( \frac{1}{m}\right) \ne 0\). In particular, \(\widehat{H} \left( \phi _m,\chi ;-s\right) \) possesses a holomorphic continuation to the line \({\text {Re}}s = \frac{1}{m}\), apart from possibly at \(s = \frac{1}{m}\). When \(\chi =1\), the right-hand side has a pole of order \(b\left( d,m\right) \) at \(s=\frac{1}{m}\).

Proof

By construction, \(\widehat{H}\left( \phi _{m};\chi ;-s\right) = \prod _{v \mid \infty }\widehat{H}_v\left( 1,\chi _v;-s\right) F_{m,\chi }\left( s\right) G_{m,\chi }\left( s\right) \). We will prove the stronger result that \(G_{m,\chi }\left( s\right) \) is holomorphic on \({\text {Re}}s > \frac{1}{m+1}\) and uniformly bounded with respect to both \(\chi \) and \(\epsilon \) on \({\text {Re}}s \ge \frac{1}{m+1}+\epsilon \) for all \(\epsilon > 0\).

For a place \(v \not \mid \infty \) and \(s \in \mathbb {C}\) with \({\text {Re}}s = \sigma \ge \epsilon \) for some \(\epsilon > 0\), we have

$$\begin{aligned} \begin{aligned} \sum _{n=0}^{\infty }\left| \frac{a_{\chi ,v,n}}{q_v^{ns}}\right|&= \sum _{n=0}^{\infty }\frac{1}{q_v^{n\sigma }}\left| \int _{H_v\left( t_v\right) =q_v^n}\phi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) d\mu _v\right| \\&\le \sum _{n=0}^{\infty }\frac{1}{q_v^{n\sigma }}\int _{H_v\left( t_v\right) =q_v^n}\left| \phi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) \right| d\mu _v \\&= \int _{T\left( K_v\right) }\left| \frac{\phi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) }{H_v\left( t_v\right) ^s}\right| d\mu _v,\\ \end{aligned} \end{aligned}$$

so, by Lemma 5.1, the series \(\sum _{n=0}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}}\) is absolutely convergent and bounded by a constant depending only on \(\epsilon \) and v. Now, for any \(N \in \mathbb {N}\), we have

$$\begin{aligned} \left| \sum _{n=0}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}} - \sum _{n=0}^{N}\frac{a_{\chi ,v,n}}{q_v^{ns}}\right| = \left| \sum _{n=N+1}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}}\right| \le \sum _{n=0}^{\infty }\frac{|a_{\chi ,v,n}|}{q_v^{n\epsilon }}, \end{aligned}$$

from which it follows that \(\sum _{n=0}^{\infty }\frac{a_{\chi ,v,n}}{q_v^{ns}}\) is also uniformly convergent, hence the function \(\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \) is holomorphic on \({\text {Re}}s > 0\). Then, we note that \(F_{m,\chi ,v}\left( s\right) \) is clearly holomorphic on \({\text {Re}}s > 0\), and we have

$$\begin{aligned} \frac{1}{|F_{m,\chi ,v}\left( s\right) |} = \prod _{\overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) } \left| L_v\left( \chi ^{g_1}\cdots \chi ^{g_m},ms\right) ^{-1}\right| \le \left( 1 + \frac{1}{q_v^{m\epsilon }}\right) ^{db\left( d,m\right) }, \end{aligned}$$

hence \(G_{m,\chi ,v}\left( s\right) \) is holomorphic on \({\text {Re}}s > 0\) and is bounded uniformly in terms of \(\epsilon \) and v on \({\text {Re}}s \ge \epsilon \).

To conclude the result, it suffices to prove that there exists \(N \in \mathbb {N}\) such that

$$\begin{aligned} \prod _{q_v > N }G_{m,\chi ,v}\left( s\right) \end{aligned}$$

is holomorphic and uniformly bounded with respect to \(\chi \) on \({\text {Re}}s \ge \frac{1}{m+1} + \epsilon \) for all \(\epsilon > 0\). Let \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) be non-archimedean, and let \({\text {Re}}s = \sigma \ge \frac{1}{m+1} + \epsilon \). From

$$\begin{aligned} \widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) = \left( 1-\frac{1}{q_v^{ds}}\right) L_v\left( \chi ,s\right) , \end{aligned}$$

and the definition of \(F_{m,\chi ,v}\left( s\right) \), we have

$$\begin{aligned} |a_{\chi ,v,n}| \le 2d^n, \quad |b_{\chi ,v,n}| \le \left( b\left( d,m\right) d\right) ^n. \end{aligned}$$

Then, by (5.4), it follows inductively that we have

$$\begin{aligned} |d_{\chi ,v,n}| \le 2^n \left( b\left( d,m\right) d\right) ^n = \left( 2b\left( d,m\right) d\right) ^n. \end{aligned}$$
(5.6)

Choose \(N > \left( 2b\left( d,m\right) d\right) ^{\frac{1}{\sigma }}\) so that, for all places \(v \not \mid \infty \) with \(q_v > N\), we have \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \). Now, any normally convergent infinite product is holomorphic (see [28, Sect. 2]), and \(\prod _{q_v > N}G_{m,\chi ,v}\left( s\right) \) converges normally if and only if

$$\begin{aligned} \sum _{q_v > N} \sum _{n=m+1}^{\infty }\frac{|d_{\chi ,v,n}|}{q_v^{n\sigma }} \end{aligned}$$

converges. By (5.6) and our condition on N, we need only check convergence of

$$\begin{aligned} \sum _{q_v > N}\frac{1}{q_v^{\left( m+1\right) \sigma }}, \end{aligned}$$

which is clear. Then \(G_{m,\chi }\left( s\right) \) is holomorphic on \({\text {Re}}s > \frac{1}{m+1}\). Further, for \({\text {Re}}s \ge \frac{1}{m+1}+\epsilon \), we have the bound

$$\begin{aligned} \left| \prod _{q_v> N}G_{m,\chi ,v}\left( s\right) \right| \le \prod _{q_v > N}\left( 1+\sum _{n=m+1}^{\infty }\left( \frac{2b\left( d,m\right) d}{q_v^{\frac{1}{m+1}+\epsilon }}\right) ^n\right) , \end{aligned}$$

which is uniform with respect to \(\chi \). Now, as a convergent infinite product, \(G_{m,1}\left( \frac{1}{m}\right) \) is zero if and only if \(G_{m,1,v}\left( \frac{1}{m}\right) = \frac{\widehat{H}_v\left( \phi _{m,v},1;-\frac{1}{m}\right) }{F_{m,1,v}\left( \frac{1}{m}\right) } = 0\) for some place \(v \not \mid \infty \). However, \(\widehat{H}_v\left( \phi _{m,v},1;-\frac{1}{m}\right) \ne 0\) by Lemma 5.2, so we conclude that \(G_{m,1}\left( \frac{1}{m}\right) \ne 0\). The order of the pole of the right-hand side being \(b\left( d,m\right) \) follows from Theorem 2.20, since

$$\begin{aligned} F_{m,1}\left( s\right) = \zeta _E\left( ms\right) ^{b\left( d,m\right) }. \end{aligned}$$

\(\square \)

Note 5.16

In constructing the regularisation \(F_{m,\chi }\left( s\right) \), one must ensure that

$$\begin{aligned} \frac{\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) }{F_{m,\chi ,v}\left( s\right) } = 1 + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) \end{aligned}$$

for all non-archimedean places v with \(q_v\) is sufficiently large. As seen above, the restrictions on d, m and E ensure that this is automatic for all such places which are not totally split, i.e. we only need to approximate the local Fourier transform at totally split places not in \(S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \). Without these restrictions, one might have to approximate the local Fourier transform at places of more than one splitting type simultaneously, and to do this would require a new approach.

Before applying our key theorems, we give one more result, which will be used in order to move from the Poisson summation formula to the Tauberian theorem.

Lemma 5.17

[19, Lem. 5.9, p. 2577] Choose an \(\mathbb {R}\)-vector space norm \(\Vert \cdot \Vert \) on \(X^*\left( T_{\infty }\right) _{\mathbb {R}}\) and let \(\mathcal {L} \subset X^*\left( T_{\infty }\right) _{\mathbb {R}}\) be a lattice. Let C be a compact subset of \({\text {Re}}s \ge \frac{1}{m}\) and let \(g: X^*\left( T_{\infty }\right) _{\mathbb {R}} \times C \rightarrow \mathbb {C}\) be a function. If there exists \(0 \le \delta < \frac{1}{d-1}\) such that

$$\begin{aligned} |g\left( \psi ,s\right) | \ll _C \left( 1+ \Vert \psi \Vert \right) ^\delta \end{aligned}$$

for all \(\psi \in X^*\left( T_{\infty }\right) _{\mathbb {R}}\) and \(s \in C\), then the sum

$$\begin{aligned} \sum _{\psi \in \mathcal {L}}g\left( \psi ,s\right) \prod _{v \mid \infty }\widehat{H}_{v}\left( 1,\psi ;-s\right) \end{aligned}$$

is absolutely and uniformly convergent on C.

Theorem 5.18

Let

$$\begin{aligned} \Omega _{m}\left( s\right) = Z_m \left( s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }. \end{aligned}$$

Then \(\Omega _{m}\left( s\right) \) admits an extension to a holomorphic function on \({\text {Re}}s \ge \frac{1}{m}\).

Proof

Let \(s \in \mathbb {C}\) with \({\text {Re}}s > \frac{1}{m}\). Combining the formal application (5.1) of the Poisson summation formula with Lemma 3.8 and Corollary 5.4 gives

$$\begin{aligned} \begin{aligned} Z_m \left( s\right) = \frac{{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{d {{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) } \sum _{\chi \in \mathcal {U}} \widehat{H}\left( \phi _m,\chi ;-s\right) . \end{aligned} \end{aligned}$$
(5.7)

By Corollary 5.15, the function \(s \mapsto \frac{\phi _m\left( t\right) }{H\left( t\right) ^s}\) is \(L^1\) for \({\text {Re}}s > \frac{1}{m}\). To show that this application is valid, we apply Bourqui’s criterion [4, Cor. 3.36, p. 64], by which it suffices to show that the right-hand side of (5.7) is absolutely convergent, \(s \mapsto \frac{\phi _m\left( t\right) }{H\left( t\right) ^s}\) is continuous and there exists an open neighbourhood \(U \subset T\left( \mathbb {A}_K\right) \) of the origin and strictly positive constants \(C_1\) and \(C_2\) such that for all \(u \in U\) and all \(t \in T\left( \mathbb {A}_K\right) \), we have

$$\begin{aligned} C_1 \left| \frac{\phi _{m}\left( t\right) }{H\left( t\right) ^s}\right| \le \left| \frac{\phi _{m}\left( u t\right) }{H\left( u t\right) ^s}\right| \le C_2 \left| \frac{\phi _{m}\left( t\right) }{H\left( t\right) ^s}\right| . \end{aligned}$$

We may take \(U = \mathcal {K}\) by Lemma 4.16, and continuity is clear. It only remains to prove the absolute convergence. We will prove the stronger result that

$$\begin{aligned} \sum _{\chi \in \mathcal {U}} \widehat{H}\left( \phi _m,\chi ;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) } \end{aligned}$$

is absolutely and uniformly convergent on any compact subset C of the half-plane \({\text {Re}}s \ge \frac{1}{m}\), which will both verify validity of the application and prove the theorem.

Since \(\mathcal {K} \subset \mathsf {K}_T\) is of finite index, the map (2.2) yields a homomorphism

$$\begin{aligned} \mathcal {U} \rightarrow X^*\left( T_{\infty }\right) _{\mathbb {R}}, \quad \chi \mapsto \chi _{\infty }, \end{aligned}$$

with finite kernel \(\mathcal {N}\) and image \(\mathcal {L}\) a lattice of full rank. We obtain

$$\begin{aligned} \begin{aligned}&\sum _{\chi \in \mathcal {U}} \widehat{H}\left( \phi _m,\chi ;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) } \\&\quad = \sum _{\psi \in \mathcal {L}} \prod _{v \mid \infty } \widehat{H}_{v}\left( 1,\psi ;-s\right) \sum _{\begin{array}{c} \chi \in \mathcal {U} \\ \chi _{\infty } = \psi \end{array}} \prod _{v \not \mid \infty } \widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }, \end{aligned} \end{aligned}$$

where the inner sum is finite. Then, for \(s \in C\), we have

$$\begin{aligned} \begin{aligned}&\sum _{\chi \in \mathcal {U}} \widehat{H}\left( \phi _m,\chi ;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) } \\&\quad \ll \sum _{\psi \in \mathcal {L}} \prod _{v \mid \infty } \left| \widehat{H}_v\left( 1,\psi ;-s\right) \right| \sum _{\begin{array}{c} \chi \in \mathcal {U} \\ \chi _{\infty } = \psi \end{array}} \prod _{v \not \mid \infty } \left| \widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\right| . \end{aligned} \end{aligned}$$

Now, for \(\chi \in \mathcal {U}\), we deduce from the proof of Corollary 5.15 that

$$\begin{aligned} \left| \prod _{v \not \mid \infty } \widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\right| \ll _C \left| F_{m,\chi }\left( s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\right| . \end{aligned}$$

In order to deduce the result from Lemma 5.17, it suffices to prove that, for each \(\psi \in \mathcal {L}\) and some constant \(0 \le \delta < \frac{1}{d-1}\), we have

$$\begin{aligned} \left| \sum _{\begin{array}{c} \chi \in \mathcal {U} \\ \chi _{\infty } = \psi \end{array}}F_{m,\chi }\left( s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\right| \ll _C \left( 1+ \Vert \psi \Vert \right) ^{\delta } \end{aligned}$$

for \(\Vert \cdot \Vert \) as in Definition 2.21. As \(\mathcal {K} \subset \mathsf {K}_T\) is of finite index, there exists a constant \(Q>0\) such that \(q\left( \chi \right) < Q\) for all \(\chi \in \mathcal {U}\) (cf. [19, Proof of Thm. 5.12, p. 2579]). Since \(F_{m,\chi }\left( s\right) \) is a product of \(b\left( d,m\right) \) L-functions of Hecke characters evaluated at ms, it follows from Lemma 2.22 that

$$\begin{aligned} \left| \sum _{\begin{array}{c} \chi \in \mathcal {U} \\ \chi _{\infty } = \psi \end{array}}F_{m,\chi }\left( s\right) \left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\right| \ll _{\varepsilon ,C}|\mathcal {N}| \cdot Q^{\varepsilon }\left( 1+\Vert \psi \Vert \right) ^{\varepsilon } \end{aligned}$$

for all for all \(\varepsilon > 0\) and \(s \in C\). The result now follows from Lemma 5.17. \(\square \)

5.4 The leading constant

In order to apply [3, Thm. 3.3.2, p. 624] and deduce Theorem 1.1 from Theorem 5.18, it only remains to show that \(\Omega _m\left( \frac{1}{m}\right) \ne 0\).

Definition 5.19

Let \(\mathcal {U}^G\) be the subgroup of G-invariant elements of \(\mathcal {U}\), and set

$$\begin{aligned} \mathcal {U}_0 = {\left\{ \begin{array}{ll} \mathcal {U}[m] \text { if } d=2, \\ \mathcal {U}^G \cap \mathcal {U}[m] \text { otherwise.} \end{array}\right. } \end{aligned}$$

Lemma 5.20

For any Galois extension of number fields E/K, the subgroup \(\mathcal {U}[m] \le \left( T\left( \mathbb {A}_K\right) /T\left( K\right) \right) ^\wedge \) is finite. In particular, \(\mathcal {U}_0\) is a finite subgroup of \(\mathcal {U}\).

Proof

By class field theory [22, Ch. VI, Sect. 6; Ch. VII, Sect. 6], \(\mathcal {U}\) may be interpreted as a subset of \({{\,\mathrm{Gal}\,}}\left( E^{{{\,\mathrm{ab}\,}}}_{S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }/E\right) ^\wedge \) for \(E^{{{\,\mathrm{ab}\,}}}_{S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\) the maximal \(S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \)-unramified abelian extension of E, hence \(\mathcal {U}[m]\) is in bijection with a subset of \({{\,\mathrm{Hom}\,}}\left( {{\,\mathrm{Gal}\,}}\left( E^{\text {ab}}_{S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }/E\right) ,\mu _m\right) \), a finite set. \(\square \)

Lemma 5.21

  1. (i)

    The characters \(\chi \in \mathcal {U}\) contributing to the pole of \(Z_m\left( s\right) \) of order \(b\left( d,m\right) \) at \(s=\frac{1}{m}\) are precisely those \(\chi \in \mathcal {U}_0\) such that \(\prod _{v \mid \infty }\widehat{H}_v\left( 1,\chi _v;-\frac{1}{m}\right) G_{m,\chi }\left( \frac{1}{m}\right) \ne 0\).

  2. (i)

    Suppose that \(d \ne 2\). If d and m are coprime, then \(\mathcal {U}_0 = \{1\}\). If d is prime and m is a multiple of d, then \(\mathcal {U}_0 = \{\chi \in \mathcal {U}: \chi ^d = 1\}\).

Proof

From Theorem 2.20 and Corollary 5.15, \(\chi \in \mathcal {U}\) contributes to the pole of \(Z_m\left( s\right) \) at \(s = \frac{1}{m}\) if and only if each factor of \(F_{m,\chi }\left( s\right) \) in (5.3) equals \(\zeta _E\left( ms\right) \) and \(\prod _{v \mid \infty }\widehat{H}_v\left( 1,\chi _v;-\frac{1}{m}\right) G_{m,\chi }\left( \frac{1}{m}\right) \ne 0\). Denoting by \(\psi \) the Hecke character associated to \(\chi \), this means precisely that \(\prod _{v \mid \infty }\widehat{H}_v\left( 1,\chi _v;-\frac{1}{m}\right) G_{m,\chi }\left( \frac{1}{m}\right) \ne 0\) and, for each \(\overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) \), we have \(\left( \psi ^{g_1}\dots \psi ^{g_m}\right) _v = 1\) for all \(v \not \mid \infty \), which is equivalent by strong approximation [11, Thm., p. 67] to \(\psi ^{g_1}\dots \psi ^{g_m}= 1\). By Note 3.1, this holds if and only if

$$\begin{aligned} \chi ^{g_1}\dots \chi ^{g_m} = 1 \text { for all } \overline{\left( g_1,\dots ,g_m\right) } \in S'\left( G,m\right) . \end{aligned}$$
(5.8)

To conclude the first part, it only remains to show that (5.8) holds if and only if \(\chi \in \mathcal {U}_0\). Taking \(\overline{\left( g_1,\dots ,g_m\right) } = \overline{\left( 1,\dots ,1\right) }\) in (5.8), we obtain \(\chi ^m =1\). If \(d=2\), then \(S'\left( G,m\right) = \left\{ \overline{\left( 1,\dots ,1\right) }\right\} \), and we are done. Otherwise, taking \(\overline{\left( g_1,\dots ,g_m\right) } = \overline{\left( g,1,\dots ,1\right) }\) for any \(g \in G\), we obtain \(\chi ^{m-1}\chi ^g = 1\), so \(\chi ^m = 1\) and \(\chi =\chi ^g\) for all \(g \in G\). Conversely, if \(\chi ^m = 1\) and \(\chi = \chi ^g\) for all \(g \in G\), then (5.8) holds.

Let now \(d \ne 2\), \(\chi \in \mathcal {U}_0\), \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) and \(w \mid v\). We have \(\psi _w\left( \pi _w\right) = \psi _w^g\left( \pi _w\right) = \psi _{gw}\left( \pi _{gw}\right) \) for all \(g \in G\). Since \(\prod _{w \mid v} \psi _w\left( \pi _w\right) = 1\) (see Note 3.1) and G acts transitively on the places of E over v, we obtain \(\psi _w^d\left( \pi _w\right) = 1\), hence \(\chi ^d = 1\) by strong approximation. On the other hand, \(\chi ^m = 1\). For d and m coprime, we conclude that \(\chi = 1\). \(\square \)

Proposition 5.22

The limit

$$\begin{aligned} \Omega _m\left( \frac{1}{m}\right) = \lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\sum _{\chi \in \mathcal {U}_0}\widehat{H}\left( \phi _m,\chi ;-s\right) \end{aligned}$$

is non-zero.

Proof

We have

$$\begin{aligned} \sum _{\chi \in \mathcal {U}_0}\widehat{H}\left( \phi _m,\chi ;-s\right) = \sum _{\chi \in \mathcal {U}_0}\int _{T\left( \mathbb {A}_K\right) }\frac{\phi _m\left( t\right) \chi \left( t\right) }{H\left( t\right) ^s}d\mu = \int _{T\left( \mathbb {A}_K\right) }\frac{\phi _m\left( t\right) }{H\left( t\right) ^s} \sum _{\chi \in \mathcal {U}_0} \chi \left( t\right) d\mu . \end{aligned}$$

Let \(t \in T\left( \mathbb {A}_K\right) \). Note that, if there exists \(\chi ' \in \mathcal {U}_0\) with \(\chi '\left( t\right) \ne 1\), then

$$\begin{aligned} \sum _{\chi \in \mathcal {U}_0}\chi \left( t\right) = \sum _{\chi \in \mathcal {U}_0} \chi \chi '\left( t\right) = \chi '\left( t\right) \sum _{\chi \in \mathcal {U}_0}\chi \left( t\right) , \end{aligned}$$

so \(\sum _{\chi \in \mathcal {U}_0}\chi \left( t\right) = 0\). Then we have

$$\begin{aligned} \begin{aligned} \sum _{\chi \in \mathcal {U}_0}\widehat{H}\left( \phi _m,\chi ;-s\right) = |\mathcal {U}_0|\int _{T\left( \mathbb {A}_K\right) ^{\mathcal {U}_0,\phi _m}}\frac{1}{H\left( t\right) ^s}d\mu , \end{aligned} \end{aligned}$$

where

$$\begin{aligned} T\left( \mathbb {A}_K\right) ^{\mathcal {U}_0, \phi _m} = \{t \in T\left( \mathbb {A}_K\right) : \phi _m\left( t\right) = \chi \left( t\right) = 1 \text { for all } \chi \in \mathcal {U}_0\}. \end{aligned}$$

For any \(\chi \in \mathcal {U}_0\) and non-archimedean place \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), comparing the series expressions of \(\widehat{H}_v\left( \phi _{m,v},\chi _v;-s\right) \) and \(F_{m,\chi ,v}\left( s\right) = F_{m,1,v}\left( s\right) \) in Definition 5.13, we see that

$$\begin{aligned} \int _{H_v\left( t_v\right) = 1}\chi _v\left( t_v\right) d\mu _v = \int _{H_v\left( t_v\right) = 1} d\mu _v, \quad \int _{H_v\left( t_v\right) = q_v^m}\chi _v\left( t_v\right) d\mu _v = \int _{H_v\left( t_v\right) = q_v^m} d\mu _v, \end{aligned}$$

so \(\chi _v\left( t_v\right) = 1\) for all \(\chi \in \mathcal {U}_0\) whenever \(H_v\left( t_v\right) = 1\) or \(H_v\left( t_v\right) = q_v^m\).

For each place \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), define the continuous function

$$\begin{aligned} \theta _{m,v} : T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto {\left\{ \begin{array}{ll} 1\quad \text { if } H'_v\left( t_v\right) = 1 \text { or } H'_v\left( t_v\right) = q_v^m, \\ 0\quad \text { otherwise}. \end{array}\right. } \end{aligned}$$

Letting \(\theta _{m,v}\) be the indicator function of \(\mathcal {K}_v\) for \(v \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we define the function

$$\begin{aligned} \theta _m : T\left( \mathbb {A}_K\right) \rightarrow \{0,1\}, \quad \theta _m\left( \left( t_v\right) _v\right) = \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \theta _{m,v}\left( t_v\right) . \end{aligned}$$

By the above, we deduce that \(T\left( \mathbb {A}_K\right) ^{\theta _m} \subset T\left( \mathbb {A}_K\right) ^{\mathcal {U}_0, \phi _m}\), where

$$\begin{aligned} T\left( \mathbb {A}_K\right) ^{\theta _m} = \{t \in T\left( \mathbb {A}_K\right) : \theta _m\left( t\right) = 1\}. \end{aligned}$$

Then, by comparing limits along the real line, we see that it suffices to prove that

$$\begin{aligned} \lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\widehat{H}\left( \theta _m,1;-s\right) \ne 0. \end{aligned}$$

It is easily seen that for any non-archimedean place \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have

$$\begin{aligned} \widehat{H}_v\left( \theta _{m,v},1;-s\right) = 1 + \frac{a_{1,v,m}}{q_v^{ms}} \end{aligned}$$

for \(a_{\chi ,v,n}\) as in Definition 5.13, so, as in Corollary 5.15, we may deduce that

$$\begin{aligned} \widehat{H}\left( \theta _m,1;-s\right) = \zeta _E\left( ms\right) ^{b\left( d,m\right) }G_m\left( s\right) \end{aligned}$$

for \(G_m\left( s\right) \) a function holomorphic on \({\text {Re}}s \ge \frac{1}{m}\). It also follows that \(G_m\left( \frac{1}{m}\right) \ne 0\), since \(\widehat{H}_v\left( \theta _{m,v},1;-\frac{1}{m}\right) \ne 0\) analogously to Lemma 5.2. Then the result follows. \(\square \)

Corollary 5.23

We have

$$\begin{aligned} \Omega _m\left( \frac{1}{m}\right) = \frac{{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{d{{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) } \lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\sum _{\chi \in \mathcal {U}_0}\widehat{H}\left( \phi _m,\chi ;-s\right) \ne 0. \end{aligned}$$

Proof of Theorem 1.1

Since \(\Omega _m\left( \frac{1}{m}\right) \ne 0\), the result for \(S = S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) follows from [3, Thm. 3.3.2, p. 624] and Theorem 5.18, taking \(c\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \right) \) to be

$$\begin{aligned} \frac{m{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{\left( b\left( d,m\right) -1\right) !d {{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) }\lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) ^{b\left( d,m\right) }\sum _{\chi \in \mathcal {U}_0}\widehat{H}\left( \phi _m,\chi ;-s\right) . \end{aligned}$$

The result for \(S \supset S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) follows analogously upon redefining \(\phi _{m,v}\) to be identically 1 for each \(v \in S \setminus S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) in Definition 4.10. \(\square \)

6 Campana points

In this section we prove Theorem 1.4. We will be brief when the argument is largely similar to the case of weak Campana points, emphasising only the key differences. Fix a Galois extension E/K of number fields with K-basis \({{\,\mathrm{\varvec{\omega }}\,}}= \{\omega _0,\dots ,\omega _{d-1}\}\), let \(m \in \mathbb {Z}_{\ge 2}\) and set \(T = T_{{{\,\mathrm{\varvec{\omega }}\,}}}\) as in Sect. 3.

Definition 6.1

For each non-archimedean place \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), let

$$\begin{aligned} N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x\right) = \prod _{w \mid v} f_w\left( x\right) \end{aligned}$$

denote the v-adic decomposition of the norm form \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\) associated to \({{\,\mathrm{\varvec{\omega }}\,}}\) into irreducible polynomials \(f_w\left( x\right) \in \mathcal {O}_v[x]\). For each \(w \mid v\), we define the functions

$$\begin{aligned}&\widetilde{H}_w: T\left( K_v\right) \rightarrow \mathbb {R}_{>0}, \quad x \mapsto \frac{\max \{|x_i|_v^{{{\,\mathrm{deg}\,}}f_w}\}}{|f_w\left( x\right) |_v}, \\&\psi _{m,w}: T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto {\left\{ \begin{array}{ll} 1 \quad \text { if } \widetilde{H}_w\left( t_v\right) = 1 \text { or } \widetilde{H}_w\left( t_v\right) \ge q_v^m, \\ 0 \quad \text { otherwise}. \end{array}\right. } \end{aligned}$$

We then define the Campana local indicator function

$$\begin{aligned} \psi _{m,v}: T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto \prod _{w \mid v}\psi _{m,w}\left( t_v\right) . \end{aligned}$$

Setting \(\psi _{m,v} = 1\) for \(v \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we then define the Campana indicator function

$$\begin{aligned} \psi _m : T\left( \mathbb {A}_K\right) \rightarrow \{0,1\}, \quad \left( t_v\right) _v \mapsto \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \psi _{m,v}\left( t_v\right) . \end{aligned}$$

If \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), then for each \(w \mid v\), we also define the function

$$\begin{aligned} \sigma _{m,w}: T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto {\left\{ \begin{array}{ll} 1 \text { if } \widetilde{H}_w\left( t_v\right) = 1 \text { or } \widetilde{H}_w\left( t_v\right) = q_v^m, \\ 0 \text { otherwise} \end{array}\right. } \end{aligned}$$

and we define

$$\begin{aligned} \sigma _{m,v}: T\left( K_v\right) \rightarrow \{0,1\}, \quad t_v \mapsto \prod _{w \mid v}\sigma _{m,w}\left( t_v\right) . \end{aligned}$$

Letting \(\sigma _{m,v}\) be the indicator function for \(\mathcal {K}_v\) for \(v \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we define the function

$$\begin{aligned} \sigma _m : T\left( \mathbb {A}_K\right) \rightarrow \{0,1\}, \quad \left( t_v\right) _v \mapsto \prod _{v \in {{\,\mathrm{Val}\,}}\left( K\right) } \sigma _{m,v}\left( t_v\right) . \end{aligned}$$

Lemma 6.2

The Campana \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\)-points of \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) are precisely the rational points \(t \in T\left( K\right) \) such that \(\psi _m\left( t\right) = 1\).

Proof

Taking coordinates \(t_0,\dots ,t_{d-1}\) as in the proof of Lemma 4.12, we have

$$\begin{aligned} \widetilde{H}_w\left( t\right) = \frac{1}{|f_w\left( t_0,\dots ,t_{d-1}\right) |_v}= q_v^{v\left( f_w\left( t_0,\dots ,t_{d-1}\right) \right) } = q_v^{n_{\alpha _w}\left( \mathcal {Z}\left( f_w\right) ,t\right) } \end{aligned}$$

for all non-archimedean places \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) and places \(w \mid v\), where \(\mathcal {Z}\left( f_w\right) \) denotes the Zariski closure of \(Z\left( f_w\right) \) in \(\mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }}\). \(\square \)

Lemma 6.3

For all \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \), the function \(\psi _{m,v}\) is \(\mathcal {K}_v\)-invariant and 1 on \(\mathcal {K}_v\).

Proof

For \(v \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) the result is trivial, so let \(v \not \in S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) and \(w \mid v\). Since \(f_w\left( x \cdot y\right) = f_w\left( x\right) f_w\left( y\right) \) for \(x,y \in L\), it follows as in the proof of Lemma 4.13 that \(\widetilde{H}_w\left( x \cdot y\right) \le \widetilde{H}_w\left( x\right) \widetilde{H}_w\left( y\right) \) for all \(x,y \in T\left( K_v\right) \). Since \(H_v' = \prod _{w \mid v}\widetilde{H}_w\), we have \(\widetilde{H}_w\left( \mathcal {K}_v\right) = 1\) for all \(w \mid v\), hence it follows as in the proof of Corollary 4.15 that \(\widetilde{H}_w\) and \(\psi _{m,w}\) are \(\mathcal {K}_v\)-invariant. Since \(\psi _{m,v}\left( 1\right) = 1\), we deduce that \(\psi _{m,v}\left( \mathcal {K}_v\right) = 1\). \(\square \)

Proposition 6.4

Given \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) and \(t_v \in T\left( K_v\right) \), the image of \(t_v\) in \(X_*\left( T_v\right) \) is

$$\begin{aligned} \sum _{w \mid v} \frac{\log _{q_v}\left( \widetilde{H}_w\left( t_v\right) \right) }{\deg f_w}n_w, \end{aligned}$$

with \(n_w\) defined as in Proposition 3.4.

Proof

Follows from Proposition 3.4. \(\square \)

Corollary 6.5

For \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) non-archimedean with \(q_v\) sufficiently large, \(\chi \) an automorphic character of T unramified at v and \(s \in \mathbb {C}\) with \({\text {Re}}s > 0\), we have

$$\begin{aligned} \widehat{H}_v \left( \psi _{m,v},\chi _v;-s\right) = 1 + \frac{1}{q_v^{ms}}\sum _{\begin{array}{c} w \mid v \\ \deg f_w \mid m \end{array}} \chi _w\left( \pi _w\right) ^m + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) . \end{aligned}$$

Proof

Since \(\chi _v\), \(H_v\) and \(\psi _{m,v}\) are \(T\left( \mathcal {O}_v\right) \)-invariant and \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \), we have

$$\begin{aligned} \begin{aligned} \widehat{H}_v \left( \psi _{m,v},\chi _v;-s\right)&= \int _{T\left( K_v\right) }\frac{\psi _{m,v}\left( t_v\right) \chi _v\left( t_v\right) }{H_v\left( t_v\right) ^s}d\mu _v = \sum _{\overline{t_v} \in T\left( K_v\right) /T\left( \mathcal {O}_v\right) }\frac{\psi _{m,v}\left( \overline{t_v}\right) \chi _v\left( \overline{t_v}\right) }{H_v\left( \overline{t_v}\right) ^s} \\&= \sum _{n_v \in X_*\left( T_v\right) } \frac{\psi _{m,v}\left( n_v\right) \chi _v\left( n_v\right) }{e^{\varphi _{\Sigma }\left( n_v\right) s\log q_v}} = \sum _{r=0}^{\infty } \frac{\gamma _{\chi ,v,r}}{q_v^{rs}} , \end{aligned} \end{aligned}$$

where

$$\begin{aligned} \gamma _{\chi ,v,r} = \sum _{\begin{array}{c} n_v \in X_*\left( T_v\right) \\ H_v\left( n_v\right) = q_v^r \end{array}} \psi _{m,v}\left( n_v\right) \chi _v\left( n_v\right) . \end{aligned}$$

Put \(d_w = {{\,\mathrm{deg}\,}}f_w\) and let \(n_v = \sum _{w \mid v} \alpha _w n_w \in X_*\left( T_v\right) \) with \(\min _w\{\alpha _w\} = 0\). By Proposition 3.4 and Note 3.1, we have

$$\begin{aligned} \begin{aligned} \log _{q_v} H_v\left( n_v\right)&= \sum _{w \mid v} d_w \alpha _w, \quad \chi _v\left( n_v\right) = \prod _{w \mid v}\chi _w\left( \pi _w\right) ^{d_w \alpha _w}, \\ \psi _{m,v}\left( n_v\right)&= {\left\{ \begin{array}{ll} 1 \quad \text { if } \alpha _w = 0 \text { or } \alpha _w \ge \frac{m}{d_w} \text { for all } w \mid v, \\ 0 \quad \text { otherwise.} \end{array}\right. } \end{aligned} \end{aligned}$$

In particular, \(\psi _{m,v}\left( n_v\right) = 0\) whenever \(q_v \le H_v\left( n_v\right) \le q_v^{m-1}\), hence \(\gamma _{\chi ,v,r} = 0\) for \(1 \le r \le m-1\). Further, we see that \(\psi _{m,v}\left( n_v\right) =1\) and \(H_v\left( n_v\right) = q_v^m\) if and only if there is exactly one place \(w_0 \mid v\) such that \(\alpha _{w_0} = \frac{m}{d_{w_0}}\) and \(\alpha _w = 0\) for \(w \ne w_0\), so

$$\begin{aligned} \gamma _{\chi ,v,m} = \sum _{\begin{array}{c} w \mid v \\ \deg f_w \mid m \end{array}} \chi _w\left( \pi _w\right) ^m. \end{aligned}$$

Since \(|\psi _{m,v}\left( n_v\right) \chi _v\left( n_v\right) | \le 1\), we deduce that

$$\begin{aligned} |\gamma _{\chi ,v,r}| \le \#\left\{ \beta _1, \dots , \beta _d \in \mathbb {Z}_{\ge 0} : \sum _{i=1}^d \beta _i = r\right\} \le d^r. \end{aligned}$$

Analogously to the proof of Corollary 5.15, we deduce for \(q_v\) sufficiently large that

$$\begin{aligned} \sum _{r=m+1}^\infty \frac{\gamma _{\chi ,v,r}}{q_v^{rs}} = O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) . \end{aligned}$$

\(\square \)

Proposition 6.6

For all places \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) with \(q_v\) sufficiently large, we have

$$\begin{aligned} \widehat{H}_v \left( \psi _{m,v},\chi _v;-s\right) = L_v\left( \chi ^m,ms\right) \left( 1+ O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) \right) , \quad {\text {Re}}s > 0. \end{aligned}$$

Proof

Let \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) with \(q_v\) sufficiently large as in Corollary 6.5. If v is totally split in E/K, then \(\deg f_w = 1\) for all \(w \mid v\), so Corollary 6.5 gives

$$\begin{aligned} \widehat{H}_v\left( \psi _{m,v},\chi _v;-s\right) = 1 + \frac{1}{q_v^{ms}}\sum _{w \mid v} \chi _w^m\left( \pi _w\right) + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) , \end{aligned}$$

and so we deduce the equality, since

$$\begin{aligned} L_v\left( \chi ^m,ms\right) = \prod _{w \mid v}\left( 1-\frac{\chi _w^m\left( \pi _w\right) }{q_v^{ms}}\right) ^{-1} = 1 + \frac{1}{q_v^{ms}}\sum _{w \mid v} \chi _w^m\left( \pi _w\right) + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) . \end{aligned}$$

Now let v have inertia degree \(d_v > 1\) in E/K. Then \({{\,\mathrm{deg}\,}}f_w = d_v \mid d\) for all \(w \mid v\). If d and m are coprime, then \(d_v \not \mid m\), hence \(\gamma _{\chi ,v,m} = 0\) and the result follows from

$$\begin{aligned} L_v\left( \chi ^m,ms\right) = \prod _{w \mid v}\left( 1-\frac{\chi ^m_w\left( \pi _w\right) }{q_v^{d_v ms}}\right) ^{-1} = 1 + O\left( \frac{1}{q_v^{d_v ms}}\right) . \end{aligned}$$

If d is prime, then v is inert, so \(T\left( \mathcal {O}_v\right) = T\left( K_v\right) \), \(\widehat{H}_v\left( \psi _{m,v},\chi _v;-s\right) = 1\), and

$$\begin{aligned} L_v\left( \chi ^m,ms\right) = 1-\frac{1}{q_v^{dms}} = 1 + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) . \end{aligned}$$

\(\square \)

Proposition 6.7

For any \(\chi \in \mathcal {U}\), we have

$$\begin{aligned} \widehat{H} \left( \psi _m,\chi ;-s\right) = \prod _{v \mid \infty } \widehat{H}_v\left( 1,\chi _v;-s\right) L\left( \chi ^m,ms\right) \widetilde{G}_{m,\chi }\left( s\right) , \end{aligned}$$

where \(\widetilde{G}_{m,\chi }\left( s\right) \) is a function which is holomorphic on \({\text {Re}}s \ge \frac{1}{m}\), \(\widetilde{G}_{m,1}\left( \frac{1}{m}\right) \ne 0\) and \(\widehat{H} \left( \psi _m,1;-s\right) \) has a simple pole at \(s = \frac{1}{m}\).

Proof

Defining \(\widetilde{G}_{m,\chi ,v}\left( s\right) = \frac{\widehat{H}_v \left( \psi _{m,v},\chi _v;-s\right) }{L_v\left( \chi ^m,ms\right) }\) for each place \(v \not \mid \infty \), it follows as in the proof of Corollary 5.15 that \(\widetilde{G}_{m,\chi ,v}\left( s\right) \) is holomorphic and bounded uniformly in terms of \(\epsilon \) and v on \({\text {Re}}s \ge \epsilon \) for all \(\epsilon > 0\). Since Proposition 6.6 gives

$$\begin{aligned} \widetilde{G}_{m,\chi ,v}\left( s\right) = 1 + O\left( \frac{1}{q_v^{\left( m+1\right) s}}\right) , \end{aligned}$$

it follows as in the proof of Corollary 5.15 that \(\widetilde{G}_{m,\chi }\left( s\right) \) is holomorphic and uniformly bounded with respect to \(\chi \) for \({\text {Re}}s \ge \frac{1}{m}\) with \(\widetilde{G}_{m,1}\left( \frac{1}{m}\right) \ne 0\). Then, since

$$\begin{aligned} L\left( 1,ms\right) = \zeta _E\left( ms\right) , \end{aligned}$$

we conclude from Theorem 2.20 that \(\widehat{H} \left( \psi _m,1;-s\right) \) has a simple pole at \(s = \frac{1}{m}\). \(\square \)

Definition 6.8

For \({\text {Re}}s \gg 0 \), define the functions

$$\begin{aligned} \widetilde{Z}_m : \mathbb {C} \rightarrow \mathbb {C}, \quad s \mapsto \sum _{x \in \mathbb {P}^{d-1}\left( K\right) }\frac{\psi _m\left( x\right) }{H\left( x\right) ^s}, \quad \widetilde{\Omega }_m = \widetilde{Z}_m\left( s\right) \left( s-\frac{1}{m}\right) . \end{aligned}$$

The proofs of the following two results are analogous to those of their weak Campana counterparts, namely Theorem 5.18 and Lemma 5.21 respectively.

Theorem 6.9

The function \(\widetilde{\Omega }_m\left( s\right) \) admits a holomorphic extension to \({\text {Re}}s \ge \frac{1}{m}\).

Lemma 6.10

The characters \(\chi \in \mathcal {U}\) contributing to the simple pole of \(\widetilde{Z}_m\left( s\right) \) at \(s = \frac{1}{m}\) are precisely the characters \(\chi \in \mathcal {U}[m]\) such that \(\widetilde{G}_{m,\chi }\left( \frac{1}{m}\right) \ne 0\).

Proposition 6.11

The limit

$$\begin{aligned} \lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) \sum _{\chi \in \mathcal {U}[m]}\widehat{H}\left( \psi _m,\chi ;-s\right) \end{aligned}$$

is non-zero.

Proof

By the same reasoning as in the proof of Proposition 5.22, we have

$$\begin{aligned} \sum _{\chi \in \mathcal {U}[m]}\widehat{H}\left( \psi _m,\chi ;-s\right) = \sum _{\chi \in \mathcal {U}[m]}\int _{T\left( \mathbb {A}_K\right) }\frac{\psi _m\left( t\right) \chi \left( t\right) }{H\left( t\right) ^s}d\mu = |\mathcal {U}[m]|\int _{T\left( \mathbb {A}_K\right) ^{\mathcal {U}[m],\psi _m}}\frac{1}{H\left( t\right) ^s}d\mu , \end{aligned}$$

where

$$\begin{aligned} T\left( \mathbb {A}_K\right) ^{\mathcal {U}[m], \psi _m} = \{t \in T\left( \mathbb {A}_K\right) : \psi _m\left( t\right) = \chi \left( t\right) = 1 \text { for all } \chi \in \mathcal {U}[m]\}. \end{aligned}$$

Now, take \(\chi \in \mathcal {U}[m]\), \(v \not \in S'\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) non-archimedean. If \(\sigma _v\left( t_v\right) = 1\) for some \(t_v \in T\left( K_v\right) \), then the image of \(t_v\) in \(X_*\left( T_v\right) \) is of the form \(\sum _{w \mid v} \alpha _w n_w\), where each \(\alpha _w\) is either 0 or \(\frac{m}{d_v}\) for \(d_v\) the common inertia degree of the places of E over v, so

$$\begin{aligned} \chi _v\left( t_v\right) = \prod _{w \mid v}\chi _w\left( \pi _w\right) ^{d_v\alpha _w} = 1, \end{aligned}$$

since each \(d_v \alpha _w\) is 0 or m and \(\chi ^0 = \chi ^m = 1\). Then \(\chi _v\left( t_v\right) = 1\) for all \(\chi \in \mathcal {U}[m]\). In particular, we deduce that \(T\left( \mathbb {A}_K\right) ^{\sigma _m} \subset T\left( \mathbb {A}_K\right) ^{\mathcal {U}[m], \psi _m}\), where

$$\begin{aligned} T\left( \mathbb {A}_K\right) ^{\sigma _m} = \{t \in T\left( \mathbb {A}_K\right) : \sigma _m\left( t\right) = 1\}. \end{aligned}$$

Then it suffices to prove that

$$\begin{aligned} \lim _{s \rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) \widehat{H}\left( \sigma _m,1;-s\right) \ne 0. \end{aligned}$$

Analogously to the proof of Proposition 6.7, we may deduce that

$$\begin{aligned} \widehat{H}\left( \sigma _m,1;-s\right) = \zeta _E\left( ms\right) \widetilde{G}_m\left( s\right) \end{aligned}$$

for \(\widetilde{G}_m\left( s\right) \) a function holomorphic on \({\text {Re}}s \ge \frac{1}{m}\) with \(\widetilde{G}_m\left( \frac{1}{m}\right) \ne 0\), so the result follows. \(\square \)

Proof of Theorem 1.4

Since \(\widetilde{\Omega }_m\left( \frac{1}{m}\right) \ne 0\) by Proposition 6.11, the result for \(S = S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) now follows from [3, Thm. 3.3.2, p. 624] and Theorem 6.9, taking

$$\begin{aligned} \widetilde{c}\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \right) = \frac{m{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{d{{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) } \lim _{s \rightarrow \frac{1}{m}} \left( s-\frac{1}{m}\right) \sum _{\chi \in \mathcal {U}[m]}\widehat{H}\left( \psi _m,\chi ;-s\right) . \end{aligned}$$

The result for \(S \supset S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \) follows analogously upon redefining \(\psi _{m,v}\) to be identically 1 for each \(v \in S \setminus S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \). \(\square \)

7 Comparison to Manin-type conjecture

In this section we compare the leading constant in Theorem 1.4 with the Manin–Peyre constant in the conjecture of Pieropan, Smeets, Tanimoto and Várilly-Alvarado.

7.1 Statement of the conjecture

Let \(\left( X,D_{\epsilon }\right) \) be a smooth Campana orbifold over a number field K which is klt (i.e. \(\epsilon _{\alpha } < 1\) for all \(\alpha \in \mathcal {A}\)) and Fano (i.e. \(-\left( K_X + D_{\epsilon }\right) \) is ample). Let \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \) be a regular \(\mathcal {O}_{K,S}\)-model of \(\left( X,D_{\epsilon }\right) \) for some finite set \(S \subset {{\,\mathrm{Val}\,}}\left( K\right) \) containing \(S_{\infty }\) (i.e. \(\mathcal {X}\) regular over \(\mathcal {O}_{K,S}\)). Let \(\mathcal {L} = \left( L,\Vert \cdot \Vert \right) \) be an adelically metrised big line bundle with associated height function \(H_{\mathcal {L}}: X\left( K\right) \rightarrow \mathbb {R}_{>0}\) (see [25, Sect. 1.3]). For any subset \(U \subset X\left( K\right) \) and any \(B \in \mathbb {R}_{>0}\), we define

$$\begin{aligned} N\left( U,\mathcal {L},B\right) = \#\{P \in U: H_{\mathcal {L}}\left( P\right) \le B\}. \end{aligned}$$

Definition 7.1

Let V be a variety over a field k of characteristic zero, and let \(A \subset V\left( k\right) \). We say that A is of type I if there is a proper Zariski closed subset \(W \subset V\) with \(A \subset W\left( k\right) \). We say that A is of type II if there is a normal geometrically irreducible variety \(V'\) with \(\text {dim}V' = \text {dim}V\) and a finite surjective morphism \(\phi : V' \rightarrow V\) of degree \(\ge 2\) with \(A \subset \phi \left( V'\left( k\right) \right) \). We say that A is thin if it is contained in a finite union of subsets of \(V\left( k\right) \) of types I and II.

We are now ready to give the statement of the conjecture.

Conjecture 7.2

[27, Conj. 1.1, p. 3] Suppose that L is nef and \(\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \left( \mathcal {O}_{K,S}\right) \) is not thin. Then there exists a thin set \(Z \subset \left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \left( \mathcal {O}_{K,S}\right) \) and explicit positive constants \(a = a\left( \left( X,D_{\epsilon }\right) ,L\right) \), \(b = b\left( K,\left( X,D_{\epsilon }\right) ,L\right) \) and \(c = c\left( K,S,\left( \mathcal {X},\mathcal {D}_{\epsilon }\right) ,\mathcal {L},Z\right) \) such that, as \(B \rightarrow \infty \), we have

$$\begin{aligned} N\left( \left( \mathcal {X},\mathcal {D}_{\epsilon }\right) \left( \mathcal {O}_{K,S}\right) \setminus Z,\mathcal {L},B\right) \sim c B^a\left( \log B\right) ^{b-1}. \end{aligned}$$

7.2 Interpretation for norm orbifolds

The orbifold \(\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) \) in Theorem 1.4 is klt and Fano. It is smooth precisely when \(d=2\). The \(\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }\)-model \(\left( \mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }},\mathcal {D}_m^{{{\,\mathrm{\varvec{\omega }}\,}}}\right) \) is regular. The Batyrev–Tschinkel height arises from an adelic metrisation \(\mathcal {L}\) of \(-K_{\mathbb {P}^{d-1}} = \mathcal {O}\left( d\right) \). According to [27, Sect. 3.3], we have

$$\begin{aligned} c\left( K,S,\left( \mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }},\mathcal {D}_m^{{{\,\mathrm{\varvec{\omega }}\,}}}\right) ,\mathcal {L},Z\right) = \frac{1}{d}\tau \left( K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) ,\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) ,\mathcal {L},Z\right) , \end{aligned}$$

where

$$\begin{aligned} \tau \left( K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) ,\left( \mathbb {P}^{d-1}_K,\Delta ^{{{\,\mathrm{\varvec{\omega }}\,}}}_m\right) ,\mathcal {L},Z\right) = \int _{\overline{\mathbb {P}^{d-1}\left( K\right) }_\epsilon }H\left( t\right) ^{1-\frac{1}{m}}d\tau _{\mathbb {P}^{d-1}}. \end{aligned}$$

Here, \(\tau _{\mathbb {P}^{d-1}}\) is the Tamagawa measure defined in [19, Def. 2.8, p. 372], and \(\overline{\mathbb {P}^{d-1}\left( K\right) }_\epsilon \) denotes the topological closure of the Campana points inside \(\mathbb {P}^{d-1}\left( \mathbb {A}_K\right) \). If one assumes that weak approximation for Campana points holds for this orbifold (see [27, Question 3.9, p. 13]), it follows from the definition of \(\tau _{\mathbb {P}^{d-1}}\) (cf. [27, Sect. 9]) that

$$\begin{aligned} \tau = m\frac{{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{{{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) }\lim _{s\rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) \widehat{H}\left( \psi _m,1;-s\right) . \end{aligned}$$

Given the assumption on weak approximation, the conjectural leading constant is

$$\begin{aligned} c\left( K,S,\left( \mathbb {P}^{d-1}_{\mathcal {O}_{K,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) }},\mathcal {D}_m^{{{\,\mathrm{\varvec{\omega }}\,}}}\right) ,\mathcal {L},Z\right) = \frac{m{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{d{{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) }\lim _{s\rightarrow \frac{1}{m}}\left( s-\frac{1}{m}\right) \widehat{H}\left( \psi _m,1;-s\right) . \end{aligned}$$

On the other hand, the leading constant given by Theorem 1.4 in this case is

$$\begin{aligned} \widetilde{c}\left( {{\,\mathrm{\varvec{\omega }}\,}},m,S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) \right) = \frac{m{{\,\mathrm{Res}\,}}_{s=1}\zeta _K\left( s\right) }{d{{\,\mathrm{Res}\,}}_{s=1}\zeta _E\left( s\right) } \lim _{s \rightarrow \frac{1}{m}} \left( s-\frac{1}{m}\right) \sum _{\chi \in \mathcal {U}[m]}\widehat{H}\left( \psi _m,\chi ;-s\right) . \end{aligned}$$

We observe that our constant differs from the conjectural one in the potential inclusion of non-trivial characters in the limit.

7.3 The quadratic case

We now consider the case \(d=2\), in which the orbifold in Theorem 1.4 is smooth. Here, work of Nakahara and the author [21] shows that weak approximation for Campana points holds for any \(m \in \mathbb {Z}_{\ge 2}\).

In Theorem 1.4, it is not clear that there are non-trivial characters contributing to the leading constant and whether their contribution is positive if so. However, we now exhibit an extension for which a non-trivial character contributes positively to the leading constant, and all contributing characters do so positively.

Proposition 7.3

Let \(K = \mathbb {Q}\left( \sqrt{-39}\right) \), \(E = \mathbb {Q}\left( \sqrt{-3},\sqrt{13}\right) \) and \(m=2\). Choose the K-basis \({{\,\mathrm{\varvec{\omega }}\,}}= \{1,\frac{1+\sqrt{-3}}{2}\}\) of E. Then \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) = S_{\infty }\), \(\#\mathcal {U}[2] > 1\), and for every \(\chi \in \mathcal {U}[2]\), we have \(\lim _{s\rightarrow \frac{1}{2}}\left( s-\frac{1}{2}\right) \widehat{H}\left( \psi _2,\chi ;-s\right) > 0\).

Proof

Writing \(a \cdot 1 + b \cdot \frac{1 + \sqrt{-3}}{2}\) as \(\left( a,b\right) \) and \(G = {{\,\mathrm{Gal}\,}}\left( E/K\right) \) as \(\{1_G,g\}\), we have

$$\begin{aligned}&\left( 1,0\right) ^2 = \left( 1,0\right) , \, \left( 1,0\right) \cdot \left( 0,1\right) = \left( 0,1\right) , \, \left( 0,1\right) ^2 = \left( -1,1\right) , \\&1_G\left( \left( 1,0\right) \right) = \left( 1,0\right) , \, 1_G\left( \left( 0,1\right) \right) = \left( 0,1\right) , \, g\left( \left( 1,0\right) \right) = \left( 1,0\right) , \, g\left( \left( 0,1\right) \right) = \left( 1,-1\right) , \end{aligned}$$

and clearly \(1 = \left( 1,0\right) \), hence \(S\left( {{\,\mathrm{\varvec{\omega }}\,}}\right) = S_{\infty }\). Note that \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x,y\right) = x^2 + xy + y^2\).

Since \({{\,\mathrm{Cl}\,}}\left( E\right) \cong \mathbb {Z}/2\mathbb {Z}\), the Hilbert class field M of E is quadratic over E. We obtain the unramified Hecke character \(\chi _M\) of E, which is defined for all split \(w \in {{\,\mathrm{Val}\,}}\left( E\right) \) by \(\chi _{M,w}\left( \pi _w\right) = -1\) and is trivial at all other places. Since \(\chi _M\) is trivial on \(\mathbb {A}_K^*\), it may be viewed inside \(\mathcal {U}[2]\), hence \(\#\mathcal {U}[2] > 1\).

Let \(\chi \in \mathcal {U}[2]\). To show that \(\lim _{s\rightarrow \frac{1}{2}}\left( s-\frac{1}{2}\right) \widehat{H}\left( \psi _2,\chi ;-s\right) > 0\), it suffices to show that \(\widehat{H}_v\left( \psi _{2,v},\chi _v;-\frac{1}{2}\right) > 0\) for all \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \). If \(v \mid \infty \), then \(\widehat{H}_v\left( \psi _{2,v},\chi _v;-\frac{1}{2}\right) = \widehat{H}_v\left( 1,1;-\frac{1}{2}\right) > 0\), as \(\chi _v\) gives a continuous homomorphism from \(T\left( K_v\right) /T\left( \mathcal {O}_v\right) \cong \mathbb {R}_{>0}\) to \(\mu _2\), and \(\mathbb {R}_{>0}\) has no proper open subgroups. If v is inert, then we have \(\widehat{H}_v\left( \psi _{2,v},\chi _v;-\frac{1}{2}\right) = 1\). If v is not inert, then \(N_{{{\,\mathrm{\varvec{\omega }}\,}}}\left( x,y\right) = \left( x+\theta _1 y\right) \left( x+\theta _2 y\right) \) for \(\theta _1,\theta _2 \in \mathcal {O}_v\) roots of \(z^2 - z + 1\). By Proposition 6.4, we have \(H_v = H_v'\) if and only if there are no \(\left( a,b\right) \in \left( K_v^*\right) ^2\) with \(\min \{v\left( a\right) ,v\left( b\right) \} = 0\) such that \(v\left( a+\theta _1 b\right) , v\left( a + \theta _2 b\right) \ge 1\). If \(v\left( a+\theta _1 b\right) , v\left( a + \theta _2 b\right) \ge 1\), then we deduce from the equalities

$$\begin{aligned} \theta _1\left( a + \theta _2 b\right) - \theta _2 \left( a + \theta _1 b\right) = \left( \theta _1 - \theta _2\right) a, \quad \left( a + \theta _1 b\right) - \left( a + \theta _2 b\right) = \left( \theta _1 - \theta _2\right) b, \end{aligned}$$

that \(v\left( a\right) , v\left( b\right) \ge 1 - v\left( \theta _1 - \theta _2\right) \). Since \(\min \{v\left( a\right) ,v\left( b\right) \} = 0\), we have \(H_v' \ne H_v\) if and only if \(v\left( \theta _1 - \theta _2\right) \ge 1\). Since \(\left( \theta _1 - \theta _2\right) ^2 = -3\), the only such place is the unique place \(v_0\) of K above 3, and \(v_0\left( \theta _1 - \theta _2\right) = 1\).

For any split place \(v \ne v_0\), we have \(\mathcal {K}_v = T\left( \mathcal {O}_v\right) \) and \(\psi _{2,v} = \phi _{2,v}\), so

$$\begin{aligned} \widehat{H}_v\left( \psi _{2,v},\chi _v;-\frac{1}{2}\right) = 1 + \sum _{n=2}^{\infty }\frac{c_{\chi ,v,n} - c_{\chi ,v,n-2}}{q_v^{\frac{n}{2}}} \end{aligned}$$

by Proposition 5.7. In fact, for \(w_1\) and \(w_2\) the places of E over v, we have \(\chi _{w_2}\left( \pi _{w_2}\right) = \chi _{w_1}\left( \pi _{w_1}\right) ^{-1} \in \{1,-1\}\), hence \(c_{\chi ,v,n} - c_{\chi ,v,n-2} = 2\chi _{w_1}\left( \pi _{w_1}\right) ^n\), so

$$\begin{aligned} \widehat{H}_v\left( \psi _{2,v},\chi _v;-\frac{1}{2}\right) = 1 + \sum _{n=2}^{\infty }\frac{2\chi _{w_1}\left( \pi _{w_1}\right) ^n}{q_v^{\frac{n}{2}}} = 1 + \frac{2}{q_v} \left( \frac{1}{1-q_v^{-\frac{1}{2}}\chi _{w_1}\left( \pi _{w_1}\right) }\right) > 0. \end{aligned}$$

It only remains to check that \(\widehat{H}_{v_0}\left( \psi _{2,{v_0}},\chi _{v_0};-\frac{1}{2}\right) > 0\). We will make use of the following property of valuations:

$$\begin{aligned} v_0\left( \alpha + \beta \right) \ge \min \{v_0\left( \alpha \right) ,v_0\left( \beta \right) \}, \, \text {with equality if }v_0\left( \alpha \right) \ne v_0\left( \beta \right) . \end{aligned}$$
(7.1)

Assume that, for \(a,b \in \left( K_{v_0}^*\right) ^2\) as above, we have \(v_0\left( a+\theta _2 b\right) \ge 2\). We claim that \(v_0\left( a+\theta _1 b\right) = 1\). First, we deduce from (7.1) that

$$\begin{aligned} \begin{aligned} 2 \le v_0\left( a + \theta _2 b\right)&= v_0\left( \left( a + \theta _1 b\right) + \left( \theta _1 - \theta _2\right) b\right) \\&\ge \min \{v_0\left( a + \theta _1 b\right) , v_0\left( \left( \theta _1 - \theta _2\right) b\right) \}, \end{aligned} \end{aligned}$$

with equality if \(v_0\left( a + \theta _1 b\right) \ne v_0\left( \left( \theta _1 - \theta _2\right) b\right) \). Since \(v_0\left( \left( 2\theta _i - 1\right) ^2\right) = v_0\left( -3\right) = 2\), we have \(v_0\left( 2\theta _i - 1 \right) = 1\), so \(v_0 \left( \theta _i\right) = v_0\left( 2\theta _i\right) = v_0\left( 1\right) =0\) by (7.1), hence \(\min \{v_0\left( a\right) ,v_0\left( \theta _2 b\right) \} = \min \{v_0\left( a\right) ,v_0\left( b\right) \} = 0\). Then, since \(v_0\left( a+\theta _2 b\right) \ge 2\), it follows that \(v_0\left( a\right) = v_0\left( \theta _2 b\right) \). We deduce that \(v_0\left( a\right) = v_0\left( \theta _2 b\right) = v_0\left( b\right) \), so \(v_0\left( a\right) = v_0\left( b\right) = \min \{v_0\left( a\right) ,v_0\left( b\right) \} = 0\). Since \(v_0\left( \theta _1 - \theta _2\right) = 1\), we have \(v_0\left( \left( \theta _1 - \theta _2\right) b\right) = 1\). It follows that \(v_0\left( a + \theta _1 b\right) = 1\) by (7.1).

We deduce that \(\psi _{2,v_0}\left( t_{v_0}\right) = 1\) if and only if \(t_{v_0} \in \mathcal {K}_{v_0}\), hence

$$\begin{aligned} \widehat{H}_{v_0}\left( \psi _{2,{v_0}},\chi _{v_0};-\frac{1}{2}\right) = \int _{\mathcal {K}_{v_0}}d\mu _{v_0} > 0; \end{aligned}$$

positivity follows since \(\mathcal {K}_v \subset T\left( \mathcal {O}_v\right) \) is of finite index for all \(v \in {{\,\mathrm{Val}\,}}\left( K\right) \). \(\square \)

7.4 Possible thin sets

Assuming the truth of Conjecture 7.2, the question arises of which thin set Z should be removed in the setting of Proposition 7.3. Informally, its removal should remove the contribution of all non-trivial characters \(\chi \in \mathcal {U}[2]\). One might therefore postulate that, for each non-trivial character \(\chi \in \mathcal {U}[2]\), there is a finite morphism \(\varphi _{\chi }: C_{\chi } \rightarrow \mathbb {P}^1_K\), where \(C_{\chi }\) is a smooth projective curve, and \(Z = \bigcup _{\chi \in \mathcal {U}[2]}\varphi _{\chi }\left( C_\chi \left( K\right) \right) \). By the height bounds in [30, Sect. 9.7], we would have \(C_\chi \cong \mathbb {P}^1_K\) and \({{\,\mathrm{deg}\,}}\left( \varphi _\chi \right) =2\), making the morphisms \(\varphi _{\chi }\) degree-two endomorphisms of \(\mathbb {P}^1_K\). However, it is not clear how one should construct such endomorphisms. This may be an interesting direction to pursue in future work.