This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Orbital integrals and normalizations of measures

Julia Gordon Department of Mathematics
University of British Columbia
121-1984 Mathematics Rd., Vancouver BC V6T 1Z2, Canada
[email protected]
Abstract.

This note provides an informal introduction, with examples, to some technical aspects of the re-normalization of measures on orbital integrals used in the work of Langlands, Frenkel-Langlands-Ngô, and Altug on Beyond Endoscopy. In particular, we survey different relevant measures on algebraic tori and explain the connection with the Tamagawa numbers. We work out the example of GL2\operatorname{GL}_{2} in complete detail. The Appendix by Matthew Koster illustrates, for the Lie algebras 𝔰𝔩2\mathfrak{sl}_{2} and 𝔰𝔬3\mathfrak{so}_{3}, the relation between the so-called geometric measure on the orbits and Kirillov’s measure on co-adjoint orbits in the linear dual of the Lie algebra.

1. Introduction

Haar measure on a locally compact topological group is unique up to a constant. In many situations the normalization matters (as we shall see below). In particular, it seems that some measures are more convenient than others in the approach to Beyond Endoscopy by Frenkel-Langlands-Ngô, Arthur, and Altug. The main goal of this note is to track some of the normalizations of the measures that arise in the literature on orbital integrals on reductive groups over non-Archimedean local fields, and provide an introduction to this aspect of Altug’s lectures. In particular, our first goal is to give an exposition of the formula (3.31) in [FLN10], which is the first technical step towards Poisson summation. The second goal is to summarize the relationship between measures on pp-adic manifolds, point-counts over the residue field, and local LL-functions. These relations are scattered over the literature, and the aim here is to collect the references in one place, and provide some examples.

Fundamentally, there are two approaches to choosing a normalization of a Haar measure on the set of FF-points of an algebraic group for a local field FF: one can consider a measure associated with a specified differential form; or one can choose a specific compact subgroup and prescribe the volume of that subgroup. As we shall see, both approaches have certain advantages, and converting between these two kinds of normalizations can be surprisingly tricky. All the objects we consider here will be affine algebraic varieties, and we will only consider algebraic differential forms. To avoid confusion, we will try to consistently denote varieties by bold letters, while various sets of rational points will be denotes by letters in the usual font.

In this context, we start, in §2, with a quick survey of A.Weil’s definition of the measure on the set 𝐕(F)\mathbf{V}(F) of FF-points of an affine variety 𝐕\mathbf{V} associated with an algebraic volume form on 𝐕\mathbf{V}. We discuss the relation between this measure and counting rational points of 𝐕\mathbf{V} over the residue field, and the results on the various measures on 𝐓(F)\mathbf{T}(F) for an algebraic torus 𝐓\mathbf{T} that follow. Next in §3, we compare two natural measures on the orbits of the adjoint action of an algebraic group 𝐆(F)\mathbf{G}(F) on itself. This comparison is the main reason for writing this note. More specifically, we introduce Steinberg map and derive the relationship between two measures on the orbits: the so-called geometric measure, obtained by considering the stable orbits as fibres of Steinberg map, and the measure obtained as a quotient of two natural measures coming from volume forms. In §4, we combine the outcome with the results of §2 to obtain the relationship between the geometric measure and the so-called canonical measure (which is the one most frequently used to define orbital integrals). We do the GL2\operatorname{GL}_{2} example in detail. Finally, in §5, we assemble the local results into a global calculation, first, in the context of the analytic class number formula, and then in the case of Eichler-Selberg Trace Formula.

Acknowledgment and disclaimer.

These notes would not have been possible without many conversations with W. Casselman over the years; in particular, among many other insights, I thank him for pointing out the key point of §2.2.3. I learned most of the material presented in these notes while working on a seemingly unrelated project with Jeff Achter, S. Ali Altug, and Luis Garcia. I am very grateful to these people. One of the things that surprised me during that work was the difficulty of doing calculations with Haar measure and tracking its normalizations in the literature. The goal of these notes is to illustrate practical ways of doing such calculations, and provide references (and emphasize the normalizations of the measures in these sources) to the best of my ability at the moment. I am not aiming at presenting general (or rigorous) proofs here.

My sincere gratitude also goes to the organizers of the Program “On the Langlands Program: Endoscopy and Beyond” at NUS for inviting me to give these lectures and for their patience and encouragement during the preparation of these notes; and to the referee for many helpful suggestions.

2. Volume forms and point-counting

Everywhere in this note, FF stands for a non-Archimedean local field (of characteristic zero or positive characteristic), with the ring of integers 𝒪F\mathcal{O}_{F} and residue field kFk_{F} of cardinality qq. We denote its uniformizing element (a uniformizer for short) by ϖ\varpi; by definition, ϖ\varpi is a generator of the maximal ideal of 𝒪F\mathcal{O}_{F}. We denote the normalized valuation on FF by ord\operatorname{ord}; thus, ord(ϖ)=1\operatorname{ord}(\varpi)=1.

2.1. The measure on the affine line

We start with choosing, once and for all, an additive Haar measure on the affine line. For a non-Archimedean local field FF, we normalize the additive Haar measure on FF so that vol(𝒪F)=1\operatorname{vol}(\mathcal{O}_{F})=1. For an affine line over FF, given a choice of the coordinate xx, there is an invariant differential form dxdx; we declare that the associated measure |dx||dx| on 𝔸1(F)\mathbb{A}^{1}(F) also gives volume 11 to the ring of integers (this choice is analogous to setting up the ‘unit interval’ on the xx-axis over the reals and declaring that the interval [0,1][0,1] has ‘volume’ (i.e., length) 11 with respect to the measure |dx||dx|). Now that this choice is made, any non-vanishing top degree differential form ω\omega (defined over FF or any finite extension of FF) on a dd-dimensional FF-variety 𝐕\mathbf{V} determines a measure |ω||\omega| on the set 𝐕(F)\mathbf{V}(F) of its FF-points, where |||\cdot| stands for the absolute value on FF (respectively, its unique extension to the field of definition of ω\omega). Thus, our definition of the measure associated with a volume form is such that for the additive group 𝔾a\mathbb{G}_{a} both approaches to normalizing the measure give the same natural measure: the measure that gives 𝔾a(𝒪F)\mathbb{G}_{a}(\mathcal{O}_{F}) volume 11. We shall see that such a measure is closely related to counting points over the residue field.

Remark 2.1.

Note that our choice of the normalization of the measure on the affine line differs from that of [FLN10]: if the local field under consideration arises as a completion of a global field at a finite place, we normalize the Haar measure on it so that the ring of integers has volume 11, whereas in [FLN10], the authors fix a choice of a character of the global field and normalize the measure on every completion so that it is self-dual with respect to that character. This choice is important for the Poisson summation formula; however, as the authors point out, this makes all the measures locally non-canonical. Since our exposition is local, we chose to omit this complication. However, this means that given any variety 𝐕\mathbf{V} defined over a global field KK, our calculations of measures on 𝐕(Kv)\mathbf{V}(K_{v}) at every place vv differ from those of [FLN10] by N𝔡vdim(𝐕)/2=|ΔK/|vdim(𝐕)/2N{\mathfrak{d}}_{v}^{\dim(\mathbf{V})/2}=|\Delta_{K/\mathbb{Q}}|_{v}^{\dim(\mathbf{V})/2}, where 𝔡\mathfrak{d} is the different, N:KN:K\to\mathbb{Q} is the norm map, and ΔK/\Delta_{K/\mathbb{Q}} is the discriminant of KK.111more precisely, we quote (approximately) from [FLN10]: ‘because of this there are no canonical local calculations. The ideal 𝔡v{\mathfrak{d}}_{v} is however equal to 𝒪v\mathcal{O}_{v} almost everywhere. So there are canonical local formulas almost everywhere.’ If K=K=\mathbb{Q}, this issue disappears.

There is a natural notion of integration on the set of pp-adic points of a variety with respect to a volume form, see [Wei82]. A key feature of this theory is that if 𝐗{\mathbf{X}} is a smooth scheme over 𝒪F\mathcal{O}_{F}, and ω\omega is a top degree non-vanishing differential form on 𝐗{\mathbf{X}}, defined over 𝒪F\mathcal{O}_{F}, then the volume of 𝐗(𝒪F){\mathbf{X}}(\mathcal{O}_{F}) with respect to the measure |ω||\omega| is given by the number of points on the closed fibre of 𝐗{\mathbf{X}}:

(1) vol|ω|(𝐗(𝒪F))=#𝐗(kF)qdim(𝐗).\operatorname{vol}_{|\omega|}({\mathbf{X}}(\mathcal{O}_{F}))=\frac{\#{\mathbf{X}}(k_{F})}{q^{\dim({\mathbf{X}})}}.

The relationship between volumes and point-counts for more general sets (e.g. not requiring smoothness) was further explored by Serre [Ser81, Chapter III], Oesterlé [Oes82], and in the greatest generality,222a far-reaching generalization of these ideas is the theory of motivic integration started by Batyrev [Bat99], Kontsevich, Denef-Loeser [DL01], and Cluckers-Loeser [CL08]. W. Veys [Vey92].

Here we will only need to consider the case of reductive algebraic groups. We start with algebraic tori, where the volumes already carry interesting arithmetic information.

2.2. Tori

Let 𝐓\mathbf{T} be an algebraic torus defined over FF. Let FsepF^{\operatorname{sep}} be the separable closure of FF. As discussed above, to define a measure on T:=𝐓(F)T:=\mathbf{T}(F) one can start with a differential form or with a compact subgroup. To define either, we first need a choice of coordinates on 𝐓\mathbf{T}. One natural choice, albeit not defined over FF unless 𝐓\mathbf{T} is FF-split, comes from any basis of the character lattice of 𝐓\mathbf{T}. Let χ1,,χr\chi_{1},\dots,\chi_{r}, where rr is the rank of 𝐓\mathbf{T}, be any set of generators of X(𝐓)X^{\ast}(\mathbf{T}) over \mathbb{Z} (the characters a priori are defined over FsepF^{\operatorname{sep}}). We note that this choice is equivalent to a choice of an isomorphism 𝐓𝔾mr\mathbf{T}\simeq\mathbb{G}_{m}^{r} over FsepF^{\operatorname{sep}}. Then we can define a volume form (defined over FsepF^{\operatorname{sep}} but not over FF, unless 𝐓\mathbf{T} is split over FF):

(2) ωT=dχ1χ1dχrχr.\omega_{T}=\frac{d\chi_{1}}{\chi_{1}}\wedge\ldots\wedge\frac{d{\chi_{r}}}{\chi_{r}}.

The group of FF-points 𝐓(F)\mathbf{T}(F) has a unique maximal compact subgroup in the pp-adic topology; we denote it by TcT^{c}, following the notation of [Shy77]. Ono gave the description of this subgroup in terms of characters:

Tc={t𝐓(F):|χ(t)|=1 for χX(𝐓)F},T^{c}=\{t\in\mathbf{T}(F):|\chi(t)|=1\text{ for }\chi\in X^{\ast}(\mathbf{T})_{F}\},

where X(𝐓)FX^{\ast}(\mathbf{T})_{F} is the sublattice of X(𝐓)X^{\ast}(\mathbf{T}) consisting of the characters defined over FF.

Question 1: What is the volume of TcT^{c} with respect to the measure |ωT||\omega_{T}|?

The question is well-defined because any other \mathbb{Z}-basis of X(𝐓)X^{\ast}(\mathbf{T}) would differ from {χi}\{\chi_{i}\} by a \mathbb{Z}-matrix of determinant ±1\pm 1, and hence the resulting volume form would give rise to the same measure.

The complete answer to this question is quite involved and requires machinery beyond the scope of this note. Here we show some basic examples illustrating the easy part and the difficulty, and provide further references in §2.2.3.

Example 2.2.

𝐓\mathbf{T} is FF-split. For 𝔾m\mathbb{G}_{m}, we have: the invariant form as above is dx/xdx/x, where x=χ1:𝔾m𝔾mx=\chi_{1}:\mathbb{G}_{m}\to\mathbb{G}_{m} is the identity character and the natural coordinate, and 𝔾mc=𝒪F×\mathbb{G}_{m}^{c}=\mathcal{O}_{F}^{\times}. The volume calculation gives:

vol|dxx|(𝒪F×)=𝒪F×1|x||dx|=𝒪F×|dx|=vol|dx|(𝒪F)vol|dx|(ϖ𝒪F)\displaystyle\operatorname{vol}_{\left|\frac{dx}{x}\right|}(\mathcal{O}_{F}^{\times})=\int_{\mathcal{O}_{F}^{\times}}\frac{1}{|x|}|dx|=\int_{\mathcal{O}_{F}^{\times}}|dx|=\operatorname{vol}_{|dx|}(\mathcal{O}_{F})-\operatorname{vol}_{|dx|}(\varpi\mathcal{O}_{F})
=11q=#kF×q,\displaystyle=1-\frac{1}{q}=\frac{\#k_{F}^{\times}}{q},

as predicted by (1). Then for an FF-split torus of rank rr,

vol|ωT|(Tc)=(11q)r.\operatorname{vol}_{|\omega_{T}|}(T^{c})=\left(1-\frac{1}{q}\right)^{r}.

The next easiest case is Weil restriction of scalars.

Example 2.3.

Let E/FE/F be a quadratic extension, and 𝐓:=ResE/F𝔾m\mathbf{T}:=\operatorname{Res}_{E/F}\mathbb{G}_{m}. Assume that the residue characteristic p2p\neq 2.

We write E=F(ϵ)E=F(\sqrt{\epsilon}), where ϵ\epsilon is any non-square in FF, and we can choose it to be in 𝒪F\mathcal{O}_{F} without loss of generality. By definition, 𝐓\mathbf{T} has two characters defined over EE, call them z1z_{1} and z2z_{2} and think of them as EE-coordinates on 𝐓(E)\mathbf{T}(E). Then the volume form ωT\omega_{T} is simply ωT=dz1z1dz2z2\omega_{T}=\frac{dz_{1}}{z_{1}}\wedge\frac{dz_{2}}{z_{2}}. Note that it is defined over EE but not over FF. We can try to rewrite it in FF-coordinates: we write z1=x+ϵy,z2=xϵyz_{1}=x+\sqrt{\epsilon}y,z_{2}=x-\sqrt{\epsilon}y, and get:

(3) ωT=d(x+ϵy)x+ϵyd(xϵy)xϵy\displaystyle\omega_{T}=\frac{d(x+\sqrt{\epsilon}y)}{x+\sqrt{\epsilon}y}\wedge\frac{d(x-\sqrt{\epsilon}y)}{x-\sqrt{\epsilon}y}
=(dx+ϵdy)(dxϵdy)x2ϵy2=2ϵNE/F(x+ϵy)dxdy,\displaystyle=\frac{(dx+\sqrt{\epsilon}dy)\wedge(dx-\sqrt{\epsilon}dy)}{x^{2}-\epsilon y^{2}}=\frac{-2\sqrt{\epsilon}}{N_{E/F}(x+\sqrt{\epsilon}y)}dx\wedge dy,

where NE/FN_{E/F} is the norm map.

We note that 𝐓(F)=E×\mathbf{T}(F)=E^{\times} as a set, and the norm map is the generator of the group of FF-characters of 𝐓\mathbf{T}. Thus the subgroup TcT^{c} of 𝐓(F)\mathbf{T}(F) is

Tc={x+ϵyE×:|x2ϵy2|F=1}.T^{c}=\{x+\sqrt{\epsilon}y\in E^{\times}:|x^{2}-{\epsilon}y^{2}|_{F}=1\}.

Its volume with respect to the measure |ωT||\omega_{T}| is:

(4) vol|ωT|(Tc)=|2ϵ|{(x,y)F2:|x2ϵy2|=1}𝑑x𝑑y.\operatorname{vol}_{|\omega_{T}|}(T^{c})=|2\sqrt{\epsilon}|\int_{\{(x,y)\in F^{2}:|x^{2}-\epsilon y^{2}|=1\}}dxdy.

Thus we have reduced the computation of the volume of TcT^{c} with respect to the volume form ωT\omega_{T} to the computation of the volume of the subset of 𝔸2(F)\mathbb{A}^{2}(F), which we denote by CC, defined by

C:={(x,y)F2:|x2ϵy2|=1}C:=\{(x,y)\in F^{2}:|x^{2}-\epsilon y^{2}|=1\}

with respect to the usual measure on the plane |dxdy||dx\wedge dy| (note that CC is open in 𝔸2(F)\mathbb{A}^{2}(F) in the pp-adic topology). The computation of this volume illustrates the way to use point-counting over the residue field, and for this reason we do it in detail. There are two cases: ϵ\epsilon is a unit (i.e., E/FE/F is unramified), and ϵ\epsilon is not a unit.

First, consider the unramified case. Note that ord(x2ϵy2)=min(ord(x2),ord(y2))\operatorname{ord}(x^{2}-\epsilon y^{2})=\min(\operatorname{ord}(x^{2}),\operatorname{ord}(y^{2})) since ϵ\epsilon is a non-square unit. Then our set can be decomposed as:

C={(x,y):x𝒪F×,y𝒪F}{(x,y):x𝒪F𝒪F×,y𝒪F×},C=\{(x,y):x\in\mathcal{O}_{F}^{\times},y\in\mathcal{O}_{F}\}\sqcup\{(x,y):x\in\mathcal{O}_{F}\setminus\mathcal{O}_{F}^{\times},y\in\mathcal{O}_{F}^{\times}\},

and its volume with respect to the affine plane measure |dxdy||dx\wedge dy| is, therefore:

vol|dxdy|(C)=(q1)qq2+1qq1q=q1q(1+1q).\operatorname{vol}_{|dx\wedge dy|}(C)=\frac{(q-1)q}{q^{2}}+\frac{1}{q}\frac{q-1}{q}=\frac{q-1}{q}\left(1+\frac{1}{q}\right).

There is an alternative (and more insightful) way to do this calculation: first, as above, observe that necessarily the set CC is contained in 𝒪F2\mathcal{O}_{F}^{2}. Then consider the reduction mod ϖ\varpi map (x,y)(x¯,y¯)(x,y)\mapsto(\bar{x},\bar{y}) from 𝒪F2\mathcal{O}_{F}^{2} to kF2k_{F}^{2}, where ϖ\varpi is the uniformizer of the valuation of FF. Each fibre of this map is a translate of the set (ϖ)×(ϖ)𝒪F2(\varpi)\times(\varpi)\subset\mathcal{O}_{F}^{2}, thus each fibre has volume q2q^{-2} with respect to the measure that we have denoted by |dxdy||dx\wedge dy|. Therefore we just need to compute the number of these fibres to complete the calculation. There are two ways to do it: one is to proceed by hand, which in this case is easy enough. Another is to appeal to a generalization of Hensel’s Lemma: the affine 𝒪F\mathcal{O}_{F}-scheme defined by x2ϵy20x^{2}-\epsilon y^{2}\neq 0 is smooth; this implies that the reduction map from the set of its 𝒪F\mathcal{O}_{F}-points is surjective onto the set of kFk_{F}-points of its special fibre (see, e.g., [Ser81, §3] and [Bou85, III.4.5, Corollaire 3, p.271]). The set CC can be written as:

C={(x,y)𝒪F2:(x2ϵy2)¯0}.C=\{(x,y)\in\mathcal{O}_{F}^{2}:\overline{(x^{2}-{\epsilon}y^{2})}\neq 0\}.

Since the reduction map is surjective in this case, we just need to find the number of points (x¯,y¯)kF2(\bar{x},\bar{y})\in k_{F}^{2} such that x¯2ϵ¯y¯20\bar{x}^{2}-\bar{\epsilon}\bar{y}^{2}\neq 0. The set of points satisfying this condition is in bijection with 𝔽q2×\mathbb{F}_{q^{2}}^{\times}, where 𝔽q2\mathbb{F}_{q^{2}} is the quadratic extension of our residue field k=𝔽qk=\mathbb{F}_{q}, and we get the same result as above for the volume of CC.

If the extension is ramified, the calculation changes. In this case ord(ϵ)=1\operatorname{ord}(\epsilon)=1, hence for x2ϵy2x^{2}-\epsilon y^{2} to be a unit, xx has to be a unit and there is no condition on yy other than it has to be an integer. Again consider the reduction modulo the uniformizer map. Its image in this case is 𝔽q××𝔽q\mathbb{F}_{q}^{\times}\times\mathbb{F}_{q} (again, this can be checked by hand in this specific case), and thus the volume of the set CC is (q1)qq2\frac{(q-1)q}{q^{2}}.

We summarize (cf. [Lan13]):

(5) vol|ωT|(Tc)={(q1q)2𝐓 split q1qq+1q𝐓 non-split unramifed 1qq1q𝐓 non-split ramified. \operatorname{vol}_{|\omega_{T}|}(T^{c})=\begin{cases}&(\frac{q-1}{q})^{2}\quad{\mathbf{T}}\text{ split }\\ &\frac{q-1}{q}\frac{q+1}{q}\quad{\mathbf{T}}\text{ non-split unramifed }\\ &\frac{1}{\sqrt{q}}\frac{q-1}{q}\quad{\mathbf{T}}\text{ non-split ramified. }\\ \end{cases}

Note the factor |2||2| from (4) disappears (i.e., |2|v=1|2|_{v}=1) since we are assuming that p2p\neq 2. The factor 1q\frac{1}{\sqrt{q}} in the ramified case comes from the factor |ϵ||\sqrt{\epsilon}| in (4) (in this case, ϵ=ϖ\epsilon=\varpi up to a unit since p2p\neq 2).

For p=2p=2, everything in the calculation is slightly different (and a lot longer) but the answers are similar. Not to interrupt the flow of the exposition, we postpone the discussion of p=2p=2 till §5.2 below.

2.2.1. Norm-1 torus of a quadratic extension

Finally, to illustrate general difficulties of this volume computation, we consider the example of the norm-1 torus of a quadratic extension.

There is an exact sequence of algebraic FF-tori:

(6) 1ResE/F(1)𝔾mResE/F𝔾m𝔾m1,1\to\operatorname{Res}_{E/F}^{(1)}\mathbb{G}_{m}\to\operatorname{Res}_{E/F}\mathbb{G}_{m}\to\mathbb{G}_{m}\to 1,

where the last map is the norm map; its kernel is an algebraic torus 𝐓1:=ResE/F(1)𝔾m\mathbf{T}_{1}:=\operatorname{Res}_{E/F}^{(1)}\mathbb{G}_{m} over FF, called the norm-1 torus.

It is tempting to try to use this exact sequence to compute the volume of T1cT_{1}^{c} with respect to ωT1\omega_{T_{1}}, but that is not the right way to proceed. The standard way to do this calculation is to consider an isogeny between 𝐓1×𝔾m\mathbf{T}_{1}\times\mathbb{G}_{m} and 𝐓\mathbf{T} and use the results of Ono on the behaviour of various invariants attached to tori under isogenies; see also [Shy77]. However, this would take us too far afield; instead we proceed with an elementary calculation. Before we do this calculation for a pp-adic field, consider for a moment the situation when F=F=\mathbb{R} and E=E=\mathbb{C}, in order to get some geometric intuition.

Example 2.4.

Let 𝐓:=Res/𝔾m()=×\mathbf{T}:=\operatorname{Res}_{\mathbb{C}/\mathbb{R}}\mathbb{G}_{m}(\mathbb{R})=\mathbb{C}^{\times}; then the norm-1 torus is the unit circle S1S^{1}.

The same calculation as in Example 2.3 shows that the volume form ωT\omega_{T} gives the measure |ωT|=|2idxdyx2+y2|=2x2+y2|dxdy||\omega_{T}|=|\frac{2idx\wedge dy}{x^{2}+y^{2}}|=\frac{2}{x^{2}+y^{2}}|dx\wedge dy| on ×\mathbb{C}^{\times}.

Similarly, if we go by the definition of the volume form ωS1\omega_{S^{1}} on S1S^{1}, we obtain the following. The generator of the character group X(S1)X^{\ast}(S^{1}) (over \mathbb{C}) is simply the identity character zz=x+iyz\mapsto z=x+iy. We get that ωS1=dzz\omega_{S^{1}}=\frac{dz}{z} by definition, but intuitively it is not yet clear what is the measure defined by this form. We write dz=dx+idydz=dx+idy, and note that 1z=z¯=xiy\frac{1}{z}=\bar{z}=x-iy when zS1z\in S^{1}. We obtain:

(7) ωS1=(xiy)(dx+idy)=(xiy)dx+(y+ix)dy.\omega_{S^{1}}=(x-iy)(dx+idy)=(x-iy)dx+(y+ix)dy.

It is still not obvious what measure this form gives; it would be convenient to rewrite it using the local coordinate of some chart on the circle. Here we can use the fact that we are working over \mathbb{R} and take θ\theta to be the arc length; then x=cosθx=\cos\theta, y=sinθy=\sin\theta, 0θ<2π0\leq\theta<2\pi is the familiar (transcendental) parametrization. We get:

(8) ωS1=(xiy)dx+(y+ix)dy\displaystyle\omega_{S^{1}}=(x-iy)dx+(y+ix)dy
=(cosθisinθ)d(cosθ)+(sinθ+icosθ)d(sinθ)\displaystyle=(\cos\theta-i\sin\theta)d(\cos\theta)+(\sin\theta+i\cos\theta)d(\sin\theta)
=cosθsinθdθ+sinθcosθdθ+i(cos2θdθ+sin2θdθ)=idθ.\displaystyle=-\cos\theta\sin\theta d\theta+\sin\theta\cos\theta d\theta+i(\cos^{2}\theta d\theta+\sin^{2}\theta d\theta)=id\theta.

Since |i|=1|i|=1, we see that the measure |ωS1||\omega_{S^{1}}| coincides with the arc length.

The exact sequence (6) gives a relation between this measure and the measures on S1S^{1} and 𝔾m\mathbb{G}_{m}, which is the same as rewriting the measure dxdydx\wedge dy in polar coordinates. Indeed, we have (from calculus) dxdy=rdrdθdx\wedge dy=rdr\wedge d\theta, so dxdyx2+y2=drrdθ\frac{dx\wedge dy}{x^{2}+y^{2}}=\frac{dr}{r}\wedge d\theta. We obtain the relation between ωS1\omega_{S^{1}} and the form ωT\omega_{T} on 𝐓=Res/𝔾m\mathbf{T}=\operatorname{Res}_{\mathbb{C}/\mathbb{R}}\mathbb{G}_{m}:

(9) ωT=2ωS1ω𝔾m.\omega_{T}=2\omega_{S^{1}}\wedge\omega_{\mathbb{G}_{m}}.

The appearance of the factor 22 in this relation, combined with the fact the norm map to 𝔾m\mathbb{G}_{m} is not surjective on \mathbb{R}-points and is 2:12:1 illustrates that the relation between the measures on 𝐓\mathbf{T} and 𝐓1\mathbf{T}_{1} is not straightforward (if one cares for a power of 22). Armed with this caution, we move on to the pp-adic fields.

Example 2.5.

Let 𝐓1:=ResE/F(1)𝔾m\mathbf{T}_{1}:=\operatorname{Res}_{E/F}^{(1)}\mathbb{G}_{m} be the norm-1 torus of a quadratic extension as above, but with E=F(ϵ)E=F(\sqrt{\epsilon}) an extension of non-Archimedean local fields as in Example 2.3. As before, we assume p2p\neq 2 (the case p=2p=2 is treated below in §5.2).

As above, we would like to understand the form ω𝐓1\omega_{\mathbf{T}_{1}}. We observe that the relation (7) can be easily adapted to this case (essentially, replacing ii with ϵ\sqrt{\epsilon}). What we need is an algebraic parametrization of the conic x2ϵy2=1x^{2}-\epsilon y^{2}=1. Such a parametrization is given in projective coordinates (x:y:z)(x:y:z) by:

(10) x=ϵt2+1,y=2t,z=1ϵt2,tF.x=\epsilon t^{2}+1,\quad y=2t,\quad z=1-\epsilon t^{2},\quad t\in F.

Then a calculation similar to (8) shows that in the affine chart z0z\neq 0,

ω𝐓1=2ϵ1ϵt2dt.\omega_{\mathbf{T}_{1}}=-\frac{2\sqrt{\epsilon}}{1-{\epsilon}t^{2}}dt.

We note here that we could have used the same rational parametrization for the unit circle in the example above; then at this point we would have obtained the same answer: if we plug in ϵ=1\epsilon=-1, we get that the “volume” of the circle with respect to |ωS1||\omega_{S^{1}}| is 21t2+1𝑑t=2π2\int_{\mathbb{R}}\frac{1}{t^{2}+1}\,dt=2\pi, as expected.

Continuing with the pp-adic calculation, we can discard |2||2| since p2p\neq 2. Next, we observe that T1c=𝐓1(F)T_{1}^{c}=\mathbf{T}_{1}(F) (our torus is anisotropic; it has no non-trivial FF-characters, and hence the condition defining T1cT_{1}^{c} is vacuous). Therefore, the volume of T1cT_{1}^{c} with respect to ω𝐓1\omega_{\mathbf{T}_{1}} is

(11) vol|ω𝐓1|(T1c)=F|ϵ1ϵt2||dt|.\operatorname{vol}_{|\omega_{\mathbf{T}_{1}}|}(T_{1}^{c})=\int_{F}\left|\frac{\sqrt{\epsilon}}{{1-\epsilon}t^{2}}\right|\,|dt|.

Now we need to consider two cases.
Case 1. The extension is unramified, i.e., ϵ\epsilon is a non-square unit. Then (11) becomes (using the fact that the volume of the pp-adic annulus {t:ord(t)=n}\{t:\operatorname{ord}(t)=n\} with respect to the measure |dt||dt| equals qnq(n+1)q^{-n}-q^{-(n+1)}) :

(12) vol|ω𝐓1|(T1c)=\displaystyle\operatorname{vol}_{|\omega_{\mathbf{T}_{1}}|}(T_{1}^{c})= F1|1ϵt2||dt|=𝒪F|dt|+n=11q2n(qnq(n1))\displaystyle\int_{F}\frac{1}{|1-\epsilon t^{2}|}\,|dt|=\int_{\mathcal{O}_{F}}|dt|+\sum_{n=1}^{\infty}\frac{1}{q^{2n}}(q^{n}-q^{(n-1)})
=1+(11q)n=11qn=1+1q.\displaystyle=1+\left(1-\frac{1}{q}\right)\sum_{n=1}^{\infty}\frac{1}{q^{n}}=1+\frac{1}{q}.

Case 2. The extension is ramified, i.e., ord(ϵ)=1\operatorname{ord}(\epsilon)=1. Then |1ϵt2|=1|1-\epsilon t^{2}|=1 if t𝒪Ft\in\mathcal{O}_{F} and |1ϵt2|=q2n1|1-\epsilon t^{2}|=q^{2n-1} if ord(t)=n<0\operatorname{ord}(t)=n<0. Thus, the integral computing the volume of T1cT_{1}^{c} again breaks down as a sum:

(13) vol|ω𝐓1|(T1c)=\displaystyle\operatorname{vol}_{|\omega_{\mathbf{T}_{1}}|}(T_{1}^{c})= 1qF1|1ϵt2||dt|=1q(𝒪F|dt|+n=11q2n1(qnq(n1)))\displaystyle\frac{1}{\sqrt{q}}\int_{F}\frac{1}{|1-\epsilon t^{2}|}\,|dt|=\frac{1}{\sqrt{q}}\left(\int_{\mathcal{O}_{F}}|dt|+\sum_{n=1}^{\infty}\frac{1}{q^{2n-1}}(q^{n}-q^{(n-1)})\right)
=1q(1+q(11q)n=11qn)=2q.\displaystyle=\frac{1}{\sqrt{q}}\left(1+q\left(1-\frac{1}{q}\right)\sum_{n=1}^{\infty}\frac{1}{q^{n}}\right)=\frac{2}{\sqrt{q}}.

Let us compare the results of this calculation with the approach to volumes via point-counting. If we take the equation x2ϵy2=1x^{2}-\epsilon y^{2}=1 and reduce it modulo the uniformizer, we get an equation of a conic over 𝔽q\mathbb{F}_{q}. In the unramified case, this conic is in bijection with 1(𝔽q){\mathbb{P}}^{1}(\mathbb{F}_{q}) via (10); thus we expect the volume to equal q+1q\frac{q+1}{q}, which agrees with (12).

In the ramified case, when ϵ\epsilon is not a unit, the reduction of the same equation modulo the uniformizer gives a disjoint union of two lines over the finite field: it is the subvariety of the affine plane defined by x2=1x^{2}=1. The point-count over 𝔽q\mathbb{F}_{q} gives us 2q2q, thus the volume we obtain is 2qq=2\frac{2q}{q}=2, which agrees with (13) once we make the correction for the fact that our volume form had a factor of ϵ\sqrt{\epsilon} and thus was not defined over FF (this again illustrates why in [Wei82] the disciminant factor appears in the definition of the volume form). We note that when p2p\neq 2, the affine scheme SpecF[x,y]/(x2ϵy21)\operatorname{Spec}F[x,y]/(x^{2}-\epsilon y^{2}-1) is smooth over 𝒪F\mathcal{O}_{F} (in both the ramified and unramified cases, which can be checked by the Jacobi criterion, [BLR80, §2.2]), and this justifies the fact that the point-count on the reduction modϖ\mod\varpi does give us the correct answer.

2.2.2. The Néron model

How do the above calculations generalize to an arbitrary algebraic torus? The issue is that for a torus that is not FF-split, it is not a priori obvious how to choose ‘coordinates’ defined over 𝒪F\mathcal{O}_{F}; more precisely, one first needs to define an integral model for 𝐓\mathbf{T}, i.e., a scheme over 𝒪F\mathcal{O}_{F} such that its generic fibre is 𝐓\mathbf{T}. In order to use the formula (1), this model would also need to be a smooth scheme over 𝒪F\mathcal{O}_{F}. There is a canonical way to define such a smooth integral model for 𝐓\mathbf{T}, namely, the weak Néron model, [BLR80, Chapter 4], which we shall denote by 𝒯\mathcal{T}. In general it is a scheme not of finite type (it can have infinitely many connected components). The 𝒪F\mathcal{O}_{F}-points of its identity component T0:=𝒯0(𝒪F)T^{0}:={\mathcal{T}}^{0}(\mathcal{O}_{F}) provide another canonical compact subgroup of 𝐓(F)\mathbf{T}(F). This subgroup is traditionally used in the literature to normalize the Haar measures on tori (and plays a role in normalization of measures on general reductive groups, as we shall see below), but it plays no explicit role in this note, hence we do not discuss any details of its definition.

Moreover, once we have the 𝒪F\mathcal{O}_{F}-model for 𝐓\mathbf{T}, we can use the local coordinates associated with this model to define a volume form on 𝐓(F)\mathbf{T}(F). Unlike the volume form defined above by using the characters, this form actually has coefficients in FF; it is called the canonical volume form in the literature, following the article [Gro97]; we call it ωcan\omega^{{\operatorname{can}}}, but we shall not use any explicit information about it in this note.

In general, T0T^{0} is a subgroup of finite index in TcT^{c} (this index is an interesting arithmetic invariant, see [Bit11] for a detailed study), and the relationship between the form ωT\omega_{T} defined above and ωcan\omega^{{\operatorname{can}}} is discussed in [GG99]. The subgroup TcT^{c} is the set of 𝒪F\mathcal{O}_{F}-points of the so-called standard integral model of 𝐓\mathbf{T}, which is not smooth in general; roughly speaking, the coordinates on the standard model come from the characters χi\chi_{i} as above in the examples (see [Bit11, §1.1]).

If 𝐓\mathbf{T} splits over an unramified extension of FF, the situation is simple (see [Bit11] and references therein for details):

Theorem 2.6.

Suppose that 𝐓\mathbf{T} splits over an unramified extension of FF. Then

  1. (1)

    T0=TcT^{0}=T^{c}, and the special fibre 𝒯κ0{\mathcal{T}}^{0}_{\kappa} of 𝒯0{\mathcal{T}}^{0} is an algebraic torus over 𝔽q\mathbb{F}_{q}.

  2. (2)

    vol|ωT|(Tc)=vol|ωcan|(Tc)=#𝒯κ0(𝔽q)qdim(𝐓)\operatorname{vol}_{|\omega_{T}|}(T^{c})=\operatorname{vol}_{|\omega^{{\operatorname{can}}}|}(T^{c})=\frac{\#{\mathcal{T}}^{0}_{\kappa}(\mathbb{F}_{q})}{q^{\dim(\mathbf{T})}}.

We give one illustrative example without any details, and summarize the known general results below in §2.2.3.

Example 2.7.

Consider again Example 2.5, where 𝐓\mathbf{T} is the norm-1 torus of a quadratic extension. It is anisotropic over FF, and consequently its Néron model is a scheme of finite type over FF. If E/FE/F is unramified, the ‘standard model’ (see [Bit11]) coincides with the Néron model, and is simply defined by the equation x2ϵy2=1x^{2}-\epsilon y^{2}=1 over 𝒪F\mathcal{O}_{F}. It is connected and its special fibre is an irreducible conic over 𝔽q\mathbb{F}_{q}, which has q+1q+1 rational points over 𝔽q\mathbb{F}_{q}, as discussed above.

If E/FE/F is ramified, e.g. E=F[ϖ]E=F[\sqrt{\varpi}], then |Tc/T0|=2|T^{c}/{T}^{0}|=2 (the special fibre of 𝒯\mathcal{T} has two connected components, each isomorphic to an affine line as we saw above – note that it is not an algebraic torus!) And in this case we have

vol|ωcan|(T0)=#𝒯κ0(𝔽q)qdim(𝐓)=qq=1,\operatorname{vol}_{|\omega^{{\operatorname{can}}}|}({T}^{0})=\frac{\#{\mathcal{T}}^{0}_{\kappa}(\mathbb{F}_{q})}{q^{\dim(\mathbf{T})}}=\frac{q}{q}=1,

and also ωT=|ϖ|Eωcan=1qωcan\omega_{T}=|\sqrt{\varpi}|_{E}\omega^{{\operatorname{can}}}=\frac{1}{\sqrt{q}}\omega^{{\operatorname{can}}}.

Example 2.8.

For an arbitrary finite extension E/FE/F and 𝐓=ResE/F𝔾m\mathbf{T}=\operatorname{Res}_{E/F}\mathbb{G}_{m}, we have 𝐓(F)=E×\mathbf{T}(F)=E^{\times}, and Tc=𝒪E×T^{c}=\mathcal{O}_{E}^{\times}. If α1,,αr\alpha_{1},\dots,\alpha_{r} are elements of 𝒪E\mathcal{O}_{E} that form a basis for 𝒪E\mathcal{O}_{E} over 𝒪F\mathcal{O}_{F}, then χi(x):=TrE/F(αix)\chi_{i}(x):=\operatorname{Tr}_{E/F}(\alpha_{i}x) form a basis of X(𝐓)X^{\ast}(\mathbf{T}) over \mathbb{Z}. Then by definition of the discriminant (as the norm of the different 𝔡E/F{\mathfrak{d}}_{E/F}), the measure |ωT|=|dχi||\omega_{T}|=|\wedge d\chi_{i}| equals |det(TrE/F(αi))|Fωcan|\det(\operatorname{Tr}_{E/F}(\alpha_{i}))|_{F}\omega^{{\operatorname{can}}}, i.e. the conversion factor is the square root of the FF-absolute value of the discriminant of EE:

ωT=|ΔE/F|ωcan.\omega_{T}=\sqrt{|\Delta_{E/F}|}\omega^{\operatorname{can}}.

This calculation is generalized to an arbitrary reductive group (not just an arbitrary torus) in [GG99].

2.2.3. References to the general results: a non-self-contained answer to Question 1

  1. (1)

    In general, the index [Tc:T0]=|(X)Itor|[T^{c}:T^{0}]=|(X_{\ast})_{I}^{tor}| can be computed by looking at the inertia co-invariants on the co-character lattice of 𝐓\mathbf{T}, see [Bit11, (3.1)] which follows [Kot97, §7].

  2. (2)

    The relation vol|ωcan|(T0)=#𝒯κ0(𝔽q)qdim(𝐓)\operatorname{vol}_{|\omega^{{\operatorname{can}}}|}(T^{0})=\frac{\#{\mathcal{T}}^{0}_{\kappa}(\mathbb{F}_{q})}{q^{\dim(\mathbf{T})}} holds for any algebraic torus, [Bit11, Proposition 2.14].

  3. (3)

    As noted above, the form ωT\omega_{T} is not generally defined over FF. It turns out (see [GG99]) that it only needs to be corrected by a factor that is a square root of an element FF to get to a form defined over FF. We saw this already in the case when 𝐓\mathbf{T} is of the form ResE/F𝔾m\operatorname{Res}_{E/F}\mathbb{G}_{m}; the general case follows from this calculation and a theorem of Ono that relates an arbitrary torus with a torus of the form ResE/F𝔾m\operatorname{Res}_{E/F}\mathbb{G}_{m}, see proof of Corollary 7.3 in [GG99]. Specifically, Corollary 7.3 in [GG99] states (using our notation) that if FF has characteristic zero or if 𝐓\mathbf{T} splits over a Galois extension of FF of degree relatively prime to the the characteristic of FF, then:

    (14) |ωTDM|=|ωcan|,\left|\frac{\omega_{T}}{\sqrt{D_{M}}}\right|=|\omega^{\operatorname{can}}|,

    where DMD_{M} is a refined Artin conductor of the motive MM associated with 𝐓\mathbf{T}, defined in [GG99, (4.5)]. We discuss this motive briefly in the next subsection, but do not discuss the definition of the Artin conductor.

  4. (4)

    Combining these results gives an answer to Question 1 above.

2.2.4. The local LL-functions

There is yet another way to express the number of 𝔽q\mathbb{F}_{q}-points of 𝒯κ0{\mathcal{T}}^{0}_{\kappa}, and hence the volume of T0T^{0}, entirely in terms of the representation of the Galois group on the character lattice, using Artin L-factor.

Indeed, an algebraic torus 𝐓\mathbf{T} over FF is uniquely determined by the action of the Galois group of its splitting field EE on X(𝐓)X^{\ast}(\mathbf{T}). Let Γ=Gal(E/F)\Gamma=\operatorname{Gal}(E/F) and let II be the inertia subgroup and let Fr\operatorname{Fr} be the Frobenius automorphism of kEk_{E} over kFk_{F}. Recall that we have the exact sequence of groups

1IΓFr1,1\to I\to\Gamma\to\langle\operatorname{Fr}\rangle\to 1,

and the cyclic group Fr\langle\operatorname{Fr}\rangle is isomorphic to the Galois group of the residue field Gal(kE/kF)\operatorname{Gal}(k_{E}/k_{F}). Thus we get a natural action of Gal(kE/kF)\operatorname{Gal}(k_{E}/k_{F}) on the set of inertia invariants X(𝐓)IX^{\ast}(\mathbf{T})^{I}. Let us denote this integral representation by

σT:Gal(kE/kF)Aut(X(𝐓)I)GLdI(),\sigma_{T}:\operatorname{Gal}(k_{E}/k_{F})\to{\operatorname{Aut}}_{\mathbb{Z}}(X^{\ast}(\mathbf{T})^{I})\simeq\operatorname{GL}_{d_{I}}(\mathbb{Z}),

where dI=rank(X(𝐓)I)d_{I}=\operatorname{rank}(X^{\ast}(\mathbf{T})^{I}). The Artin LL-factor associated with this representation is, by definition,

(15) L(s,σT)=det(IdIσT(Fr)qs)1 for s,L(s,\sigma_{T})=\det\left(I_{d_{I}}-\frac{\sigma_{T}(\operatorname{Fr})}{q^{s}}\right)^{-1}\quad\text{ for }s\in\mathbb{C},

where IdII_{d_{I}} is the identity matrix of size dId_{I}. Then the following relation holds (we are quoting it from [Bit11, Propostion 2.14]):

Theorem 2.9.

ωcan(T0)=#𝒯κ0(𝔽q)qdim(𝐓)=L(1,σT)1\omega^{{\operatorname{can}}}(T^{0})=\frac{\#{\mathcal{T}}^{0}_{\kappa}(\mathbb{F}_{q})}{q^{\dim(\mathbf{T})}}=L(1,\sigma_{T})^{-1}.

Note that if E/FE/F is unramified, the inertia is trivial, so dI=rank(𝐓)d_{I}=\operatorname{rank}(\mathbf{T}).

Thus, to summarize, we have defined a natural invariant form ωT\omega_{T} on 𝐓\mathbf{T} and described the maximal compact subgroup TcT^{c} of 𝐓(F)\mathbf{T}(F) in terms of the characters of 𝐓\mathbf{T}. If 𝐓\mathbf{T} splits over an unramified extension, the volume of TcT^{c} with respect to this differential form equals

(16) vol|ωT|(Tc)=#𝒯κ0(𝔽q)qdim(𝐓)=L(1,σT)1,\operatorname{vol}_{|\omega_{T}|}(T^{c})=\frac{\#{\mathcal{T}}^{0}_{\kappa}(\mathbb{F}_{q})}{q^{\dim(\mathbf{T})}}=L(1,\sigma_{T})^{-1},

where the second equality holds for any 𝐓\mathbf{T}, not necessarily unramified. In general, the volume of TcT^{c} with respect to this differential form contains two more factors – the index of T0T^{0} in TcT^{c} and the ratio between the differential forms ωT\omega_{T} and ωcan\omega^{{\operatorname{can}}}.

2.3. Reductive groups

Similarly to the case of tori discussed above, for a general reductive group 𝐆\mathbf{G} over a local field FF, the choice of a normalization of Haar measure is linked with a choice of a ‘canonical’ differential form or a ‘canonical’ compact subgroup of 𝐆(F)\mathbf{G}(F) (unlike an algebraic torus, the set of FF-points of a general reductive group can have more than one conjugacy class of maximal compact subgroups, and this choice matters for the normalization of measure). Luckily, for the questions studied in [FLN10] the choice of the normalization of measure on 𝐆(F)\mathbf{G}(F) does not matter – it only contributes some global constant.

However, for completeness, we record that a ‘canonical’ choice of a compact subgroup G0G^{0} and an associated volume form ωG\omega_{G} is described by B. Gross in [Gro97], using Bruhat-Tits theory. The group G0G^{0} is the set of 𝒪F\mathcal{O}_{F}-points of a smooth scheme 𝐆¯\underline{\mathbf{G}} over 𝒪F\mathcal{O}_{F} whose generic fibre is 𝐆\mathbf{G}. Hence, by Weil’s general argument (since 𝐆¯\underline{\mathbf{G}} is smooth over 𝒪F\mathcal{O}_{F}), the volume of G0G^{0} is obtained by counting points on the special fibre of 𝐆¯\underline{\mathbf{G}} (see ([Gro97, Proposition 4.7]):

(17) vol|ωG|(G0)=#𝐆¯κ(𝔽q)qdim𝐆.\operatorname{vol}_{|\omega_{G}|}(G^{0})=\frac{\#\underline{\mathbf{G}}_{\kappa}(\mathbb{F}_{q})}{q^{\dim\mathbf{G}}}.

Moreover, Gross defines an Artin-Tate motive MM associated with 𝐆\mathbf{G} such that the volume of the canonical compact subroup with respect to the canonical form is given by the value of the Artin LL-function associated with this motive at 11. (If 𝐆\mathbf{G} is an algebraic torus, the associated motive is precisely the representation of the Galois group on its character lattice as above, and Gross’ result amounts precisely to the statement of Theorem 2.9 above).

3. Orbital integrals: the geometric measure

3.1. The two normalizations

Let 𝐆\mathbf{G} be a connected reductive algebraic group over FF. We denote the sets of regular semisimple elements in G:=𝐆(F)G:=\mathbf{G}(F) by GrssG^{\operatorname{rss}} (respectively, 𝔤rss\mathfrak{g}^{\operatorname{rss}} for the Lie algebra 𝔤\mathfrak{g} of 𝐆\mathbf{G}).22footnotetext: The simplest way to characterize the set of regular semisimple elements is to use any faithful representation of GG to think of its elements as matrices; then an element γ𝐆(F)\gamma\in\mathbf{G}(F) (respectively, X𝔤X\in\mathfrak{g}) is regular and semisimple if and only if its eigenvalues (in an algebraic closure of FF) are distinct; we will also give a precise definition below in §3.3.1.

Let γG:=𝐆(F)rss\gamma\in G:=\mathbf{G}(F)^{\operatorname{rss}} (in this note we are only interested in this setting). The adjoint orbit (or simply, ‘orbit’, or sometimes, ‘rational orbit’ when we want to emphasize that it is the group of FF-points of 𝐆\mathbf{G} that is acting on it) of γ\gamma is the set

𝒪(γ):={gγg1g𝐆(F)}.\mathcal{O}(\gamma):=\{g\gamma g^{-1}\mid g\in\mathbf{G}(F)\}.

The centralizer of γ\gamma is by definition the group CG(γ)={g𝐆(F)gγg1=γ}C_{G}(\gamma)=\{g\in\mathbf{G}(F)\mid g\gamma g^{-1}=\gamma\}. We will also briefly refer to the notion of a stable orbit of γ\gamma. It is a finite union of rational orbits; as a first approximation, it can be thought of as the set

𝒪(γ)stable:={gγg1g𝐆(Fsep)}𝐆(F).\mathcal{O}(\gamma)^{\mathrm{stable}}:=\{g\gamma g^{-1}\mid g\in\mathbf{G}(F^{\operatorname{sep}})\}\cap\mathbf{G}(F).

However, this is not the correct definition in general; see [Kot82]. We will not need a precise definition in this note. If 𝐆=GLn\mathbf{G}=\operatorname{GL}_{n}, then for γGrss\gamma\in G^{\operatorname{rss}}, the stable orbit and rational orbit are the same.

When γGrss\gamma\in G^{\operatorname{rss}}, the identity component of the centralizer of γ\gamma is a maximal torus TGT\subset G; it can be thought of as a set of FF-points T=𝐓(F)T=\mathbf{T}(F) of an algebraic torus 𝐓\mathbf{T} defined over FF. This leads to two natural approaches to normalizing the measure on the orbit of γ\gamma:

  1. (1)

    Normalize the measures on GG and TT according to one of the methods discussed above and consider the quotient measure.

  2. (2)

    Describe the space of all (stable) orbits, and derive a measure on each orbit as a quotient measure with respect to the measure on the space of orbits. 333In fact, there is a third natural approach if we are working with the orbital integrals on the Lie algebra rather than the group: namely, to identify 𝔤\mathfrak{g} with 𝔤\mathfrak{g}^{\ast} and consider the differential form on the orbit itself which comes from Kirillov’s symplectic form on co-adjoint orbits, see [Kot05]. We will not discuss this approach here as it is not related to the main subject of the note. However, the example of 𝔰𝔩2\mathfrak{sl}_{2} where one can clearly see the relation of this measure to what one would expect from calculus is provided in the Appendix by Matthew Koster.

Since the GG-invariant measure on each orbit is unique up to a constant multiple, the two orbital integrals defined with respect to these measures will of course differ by a constant; however, this constant can, and does, depend on the orbit. The goal of this section is to give a detailed explanation for the formula that relates the two orbital integrals; this is equation (3.31) in [FLN10]. More specifically, we start with a review of the construction of Steinberg map 𝔠:𝐆𝔸G\mathfrak{c}:\mathbf{G}\to\mathbb{A}_{G}, in §3.2 below. The set 𝔸G(F)\mathbb{A}_{G}(F) has an open dense subset whose points parametrize the stable orbits of regular semisimple elements in GG – each fibre of the map 𝔠\mathfrak{c} over a point of this subset is such a stable orbit. The relation (3.31) in [FLN10] we aim to explain is:

(18) 𝔠1(a)f(g)d|ωa|=|Δ(t)|L(1,σT\G)Ostable(t,f),\int_{\mathfrak{c}^{-1}(a)}f(g)d|\omega_{a}|=|\Delta(t)|L(1,\sigma_{T\backslash G})O^{\mathrm{stable}}(t,f),

where a=𝔠(t)a=\mathfrak{c}(t). We start by defining all the ingredients of this formula (as we shall see, this formula is not really about orbital integrals; it is simply a statement about the relationship between two invariant measures on an orbit). We also simultaneously treat the orbital integrals on Lie algebras.

3.2. The space 𝔸G\mathbb{A}_{G}; Chevalley and Steinberg maps

We start with the Lie algebra, where the situation is simpler. We recall that 𝐆\mathbf{G} acts on 𝔤\mathfrak{g} via adjoint action, denoted by Ad:𝐆GL(𝔤)\operatorname{Ad}:\mathbf{G}\to\operatorname{GL}(\mathfrak{g}) (for the classical groups and their Lie algebras, Ad(g)\operatorname{Ad}(g) is simply matrix conjugation by an element g𝐆(F)g\in\mathbf{G}(F)). When we talk about orbits in 𝔤\mathfrak{g}, it is the orbits under the adjoint action. For X𝔤X\in\mathfrak{g}, its centralizer CG(X)C_{G}(X) is, by definition, its stabilizer (in 𝐆(F)\mathbf{G}(F)) under the adjoint action. If X𝔤rssX\in\mathfrak{g}^{\operatorname{rss}}, then CG(X)C_{G}(X) is a maximal torus in 𝐆(F)\mathbf{G}(F).

3.2.1. Reductive Lie algebra; algebraically closed field

444This section is entirely based on [Kot05, §14].

Let 𝔤\mathfrak{g} be the Lie algebra of 𝐆\mathbf{G}. For the moment let us work over an algebraically closed field kk of characteristic 0 (in fact, assuming sufficiently large characteristic is sufficient here but we will not pursue this direction). Let 𝔱=Lie(𝐓)\mathfrak{t}=\mathrm{Lie}(\mathbf{T}) be a maximal Cartan subalgebra. Then the ring of polynomial functions on 𝔱\mathfrak{t} is the symmetric algebra S=S(𝔱)S=S(\mathfrak{t}^{\ast}). The Weyl group acts on SS, and the ring of invariants SWS^{W} is the ring of regular functions on the quotient 𝔱/W\mathfrak{t}/W. In other words, 𝔱/W\mathfrak{t}/W is a variety over kk, isomorphic to SpecSW\operatorname{Spec}S^{W}. Furthermore, in fact SWS^{W} is itself a polynomial ring, and so 𝔱/W\mathfrak{t}/W is isomorphic to the affine space 𝔸r\mathbb{A}^{r}, where r=Rank(𝐓)r=\operatorname{Rank}(\mathbf{T}).

Example 3.1.

Let 𝐆=GLn\mathbf{G}=\operatorname{GL}_{n}, and let 𝔱𝔤=𝔤𝔩n\mathfrak{t}\subset\mathfrak{g}=\mathfrak{gl}_{n} be the Cartan subalgebra consisting of diagonal matrices. Then W=SnW=S_{n}, S=k[x1,,xn]S=k[x_{1},\dots,x_{n}], and SWS^{W} is the algebra of symmetric polynomials. As we know, it is generated by the elementary symmetric polynomials. Thus, the map 𝔱𝔸r\mathfrak{t}\to\mathbb{A}^{r} is given by: t=diag(t1,,tr)(σ1(t¯),,σr(t¯))t=\operatorname{diag}(t_{1},\dots,t_{r})\mapsto(\sigma_{1}(\bar{t}),\dots,\sigma_{r}(\bar{t})), where t¯=(t1,,tr)\bar{t}=(t_{1},\dots,t_{r}) and σk(t¯)={i1,,ik}{1,,r}ti1tik\sigma_{k}(\bar{t})=\sum_{\{i_{1},\dots,i_{k}\}\subset\{1,\dots,r\}}t_{i_{1}}\dots t_{i_{k}} is the kk-th elementary symmetric polynomial. Note that in particular, for n=2n=2, we get the map t(Tr(t),det(t))t\mapsto(\operatorname{Tr}(t),\det(t)).

Let k[𝔤]k[\mathfrak{g}] be the FF-algebra of polynomial functions on 𝔤\mathfrak{g}, and let k[𝔤]Gk[\mathfrak{g}]^{G} be the subalgebra of the polynomials invariant under the adjoint action of GG. We quote from [Kot05, §14.2]: Chevalley’s restriction theorem can be stated as:

k[𝔤]GSW,k[\mathfrak{g}]^{G}\cong S^{W},

where the isomorphism is given by restricting the polynomial functions from 𝔤\mathfrak{g} to 𝔱\mathfrak{t}. Dually to the inclusion k[𝔤]Gk[𝔤]k[\mathfrak{g}]^{G}\hookrightarrow k[\mathfrak{g}], we get the surjection (which we will refer to as Chevalley map)

𝔠𝔤:𝔤𝔸G=𝔱/W,\mathfrak{c}_{\mathfrak{g}}:\mathfrak{g}\to\mathbb{A}_{G}=\mathfrak{t}/W,

which maps X𝔤X\in\mathfrak{g} to the unique WW-orbit in 𝔱\mathfrak{t} consisting of elements conjugate to the semisimple part of XX. An important observation (which is not used in these notes but is very relevant for the subject) is that the nilpotent cone in 𝔤\mathfrak{g} is 𝔠𝔤1(0)\mathfrak{c}_{\mathfrak{g}}^{-1}(0).

In general, the role of ‘elementary symmetric polynomials’ is played by the traces of irreducible representations of 𝔤\mathfrak{g} determined by the fundamental weights. Namely, let {μi}i=1r\{\mu_{i}\}_{i=1}^{r} be the fundamental weights determined by a choice of simple roots for 𝔤\mathfrak{g} (i.e., the weights of 𝔤\mathfrak{g} defined by μi,αj=δij\langle\mu_{i},\alpha_{j}\rangle=\delta_{ij}, where Δ={αj}j=1r\Delta=\{\alpha_{j}\}_{j=1}^{r} is a base of the root system of 𝔤\mathfrak{g}). Let ρi\rho_{i} be the representation of 𝔤\mathfrak{g} of highest weight μi\mu_{i}. Then SWS^{W}, as an algebra, is generated by Tr(ρi)\operatorname{Tr}(\rho_{i}) (see, e.g., [Hum72, 23.1]).

For the type AnA_{n}, one recovers the elementary symmetric polynomials from this construction. Namely, for 𝔰𝔩n\mathfrak{sl}_{n}, it happens that the exterior powers of the standard representation are irreducible, and they give all the fundamental representations: ρi=iρ1\rho_{i}=\wedge^{i}\rho_{1}, for i=1,,n1i=1,\dots,{n-1}, where ρ1\rho_{1} is the standard representation, which has highest weight μ1\mu_{1}. Consequently, since the coefficients of the characteristic polynomial of a matrix are (up to sign) the traces of its exterior powers, we obtain:

Example 3.2.

For 𝔰𝔩n\mathfrak{sl}_{n}, Chevalley map can be realized explicitly as X(ai)X\mapsto(a_{i}), where aia_{i} are the coefficients of the characteristic polynomial of XX.

3.2.2. Reductive Lie algebra, non-algebraically closed field

When the field FF is not algebraically closed, the space 𝔸G\mathbb{A}_{G} can be defined as SpecF[𝔤]G\operatorname{Spec}F[\mathfrak{g}]^{G}, avoiding the need to choose a maximal torus; it turns out that the morphism 𝔠\mathfrak{c} is still defined over FF (see [Kot05, §14.3]). However, in this note we are only considering the case of 𝐆\mathbf{G} split over FF, and it is convenient for us to continue using an explicit definition of the map. Namely, if 𝐆\mathbf{G} is split over FF, we can choose the split maximal torus TsplT^{\operatorname{spl}} in GG, and define 𝔸G=𝔱spl/W\mathbb{A}_{G}=\mathfrak{t}^{\operatorname{spl}}/W exactly as above. The definition of the map 𝔠𝔤\mathfrak{c}_{\mathfrak{g}} stays the same. Consider explicitly what happens in the 𝐆=GL2\mathbf{G}=\operatorname{GL}_{2} example.

Example 3.3.

As above, Chevalley map is the map 𝔠𝔤𝔩2:𝔤𝔩2𝔸2\mathfrak{c}_{\mathfrak{gl}_{2}}:\mathfrak{gl}_{2}\to\mathbb{A}^{2}, X(Tr(X),det(X))X\mapsto(\operatorname{Tr}(X),\det(X)). All the split Cartan subalgebras are conjugate in 𝔤\mathfrak{g}. The image under Chevalley map of any split Cartan subalgebra 𝔱\mathfrak{t} of 𝔤\mathfrak{g} is the set

(a1,a2)𝔸2:a124a2 is a square in F.(a_{1},a_{2})\in\mathbb{A}^{2}:a_{1}^{2}-4a_{2}\text{ is a square in }F.

We observe that the FF-conjugacy classes of Cartan subalgebras in 𝔤𝔩2{\mathfrak{gl}_{2}} are in bijection with quadratic extensions of FF: as discussed above in Example 2.8, for each quadratic extension EE of FF we get the torus RE/F𝔾mR_{E/F}\mathbb{G}_{m} in GL2\operatorname{GL}_{2}. Its Lie algebra maps under Chevalley map onto the set

(a1,a2)𝔸2(F):a124a2 is a square in E.(a_{1},a_{2})\in\mathbb{A}^{2}(F):{a_{1}^{2}-4a_{2}}\text{ is a square in }E.

We note that the image of the set of semisimple elements of 𝔤\mathfrak{g} is the complement of the origin in 𝔸2(F)\mathbb{A}^{2}(F), and the image of the set of regular semisimple elements is the complement of the locus a124a2=0a_{1}^{2}-4a_{2}=0.

This situation is general: all Cartan subalgebras become conjugate to 𝔱\mathfrak{t} over the algebraic closure of FF; Chevalley map is defined over FF, and on FF-points, the images of (LieS)(F)(\operatorname{Lie}S)(F) under Chevalley map cover a Zariski open subset of 𝔸G(F)\mathbb{A}_{G}(F) as SS runs over a set of representatives of the FF-conjugacy classes of tori.

Now we return to the group itself; here the situation gets more complicated because of the central isogenies.

3.2.3. Semi-simple simply connected split group

Assume that 𝐆\mathbf{G} is split over FF, and let 𝐓\mathbf{T} be an FF-split maximal torus of 𝐆\mathbf{G}. We shall see that 𝐓/W𝔱/W=𝔸G\mathbf{T}/W\simeq\mathfrak{t}/W=\mathbb{A}_{G} in this case. To do this, we construct a basis for the coordinate ring of 𝐓/W\mathbf{T}/W (see [FLN10, §3.3]). Let α1,αr\alpha_{1},\dots\alpha_{r} be a set of simple roots for 𝐆\mathbf{G} relative to 𝐓\mathbf{T} (since 𝐆\mathbf{G} is assumed to be semi-simple, the root lattice spans the same vector space as the character lattice X(𝐓)X^{\ast}(\mathbf{T}), so there are rr simple roots). Let μi\mu_{i} be the fundamental weights, as above, defined by μi(αj)=δij\mu_{i}(\alpha_{j}^{\vee})=\delta_{ij} for 1i,jr1\leq i,j\leq r. We recall that for a semi-simple algebraic group, simply connected means that the character lattice X(𝐓)X^{\ast}(\mathbf{T}) coincides with the weight lattice, i.e. μi\mu_{i} with i=1,ri=1,\dots r constitute a \mathbb{Z}-basis of X(𝐓)X^{\ast}(\mathbf{T}).

Let ρi\rho_{i} be the algebraic representation of GG of the highest weight μi\mu_{i} for i=1,ri=1,\dots r, and let ai(t)=Trρi(t)a_{i}(t)=\operatorname{Tr}\rho_{i}(t). These functions are algebraically independent over FF and

𝐓/WSpecF[a1,,ar].\mathbf{T}/W\simeq\operatorname{Spec}F[a_{1},\dots,a_{r}].

As above, we get the map 𝔠:𝐆𝔸G\mathfrak{c}:\mathbf{G}\to\mathbb{A}_{G}, defined by g(Trρi(g))g\mapsto(\operatorname{Tr}\rho_{i}(g)). This map for the group is called Steinberg map.

Example 3.4.

As a baby example, take 𝐆=SL2\mathbf{G}=\operatorname{SL}_{2}, with 𝐓\mathbf{T} the torus of diagonal matrices, and let ρ1\rho_{1} be its standard representation on F2F^{2}. For x𝔽×x\in\mathbb{F}^{\times}, let t(x)𝐓(F)t(x)\in\mathbf{T}(F) be the one-parameter subgroup of diagonal matrices, t(x)=diag(x,x1)t(x)=\operatorname{diag}(x,x^{-1}). Then the weights of ρ1\rho_{1} are μ1:=(diag(x,x1)x)\mu_{1}:=\left(\operatorname{diag}(x,x^{-1})\mapsto x\right) and μ1=(diag(x,x1)x1)-\mu_{1}=\left(\operatorname{diag}(x,x^{-1})\mapsto x^{-1}\right) (which form a single Weyl orbit). We have a:=Tr(ρ1)(t(x))=x+x1a:=\operatorname{Tr}(\rho_{1})(t(x))=x+x^{-1}, and this is the coordinate on the affine line 𝔸1=𝔸SL2\mathbb{A}^{1}=\mathbb{A}_{\operatorname{SL}_{2}}.

More generally, for 𝐆=SLn\mathbf{G}=\operatorname{SL}_{n}, we have r=n1r=n-1, and with the standard choice of simple roots αi(diag(x1,xn))=xixi+11\alpha_{i}(\operatorname{diag}(x_{1},\dots x_{n}))=x_{i}x_{i+1}^{-1}, the above construction yields ρ1\rho_{1} - the standard representation of SLn(F)\operatorname{SL}_{n}(F) on FnF^{n}, and ρi=iρ1\rho_{i}=\wedge^{i}\rho_{1} (see [FH91, §15.2] for a detailed treatment over \mathbb{C}, which in fact works for algebraic representations over FF). We recover the same ‘characteristic polynomial’ map: the trace of the ii-th alternating power of the standard representation applied to a diagonal matrix is precisely the ii-th coefficient of its characteristic polynomial (which is, up to sign, the degree ii elementary symmetric polynomial of the eigenvalues).

(Note, however, that this is a coincidence that holds just for groups of type AnA_{n}: the isomorphism ρiiρ1\rho_{i}\simeq\wedge^{i}\rho_{1} does not hold for other types; we discuss this issue below in §3.7).

Caution: Note that unlike the typical situation when one has an algebraic homomorphism of Lie algebras which then is ‘integrated’ to obtain a homomorphism of simply connected Lie groups, Chevalley map on 𝔤\mathfrak{g} is not the differential of Steinberg map (e.g. for SL2\operatorname{SL}_{2}, the map on 𝔰𝔩2\mathfrak{sl}_{2} is Xdet(X)X\mapsto\det(X), while on SL2\operatorname{SL}_{2} the map is gTr(g)g\mapsto\operatorname{Tr}(g)).

3.2.4. Split reductive group with simply connected derived subgroup

Let 𝐆\mathbf{G} be a split, reductive group of rank rr, with simply connected derived group 𝐆der\mathbf{G}^{\operatorname{der}} (whose Lie algebra we will denote by 𝔤der\mathfrak{g}^{\operatorname{der}}). Let 𝐙\mathbf{Z} be the connected component of the centre of 𝐆\mathbf{G}. By our assumption that 𝐆\mathbf{G} is split, 𝐙\mathbf{Z} is a split torus. Let TZT\supset Z be a split maximal torus in GG, Tder=TGderT^{\operatorname{der}}=T\cap G^{\operatorname{der}} (note that TderT^{\operatorname{der}} is not the derived group of TT), and let WW be the Weyl group of GG relative to TT. Let 𝔸Gder=Tder/W\mathbb{A}_{G^{\operatorname{der}}}=T^{\operatorname{der}}/W be the Steinberg quotient for the semisimple group 𝐆der\mathbf{G}^{\operatorname{der}}. Let us denote Rank(𝐙)\operatorname{Rank}(\mathbf{Z}) by rZr_{Z}. (Naturally, the most common situation is rZ=1r_{Z}=1. ) We have 𝔸Gder𝔸rrZ\mathbb{A}_{G^{\operatorname{der}}}\simeq\mathbb{A}^{r-r_{Z}}. We have the exact sequence of algebraic groups [(3.1) in [FLN10]]:

(19) 1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐀\textstyle{{\mathbf{A}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐙×𝐆der\textstyle{\mathbf{Z}\times\mathbf{G}^{\operatorname{der}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐆\textstyle{\mathbf{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1,𝐀=𝐙𝐆der.\textstyle{1,\quad{\mathbf{A}}=\mathbf{Z}\cap\mathbf{G}^{\operatorname{der}}.}

For example, for 𝐆=GL2\mathbf{G}=\operatorname{GL}_{2}, the group 𝐀{\mathbf{A}} is the algebraic group μ2\mu_{2} of square roots of 11; it is defined by the equation x2=1x^{2}=1. 555It is important to think of 𝐀{\mathbf{A}} as a group scheme. As the authors point out, this group scheme presents an ‘annoying difficulty’ in characteristic 22 (by not being étale).

Naïvely, then, one would like to define Steinberg-Hitchin base as 𝔸Gder×𝐙\mathbb{A}_{G^{\operatorname{der}}}\times\mathbf{Z}, and establish a correspondence between the stable conjugacy classes in GG and the points of the base, as it was done for semi-simple simply connected groups. The obstacle is that we cannot really define a good map from 𝐆\mathbf{G} to 𝔸Gder×𝐙\mathbb{A}_{G^{\operatorname{der}}}\times\mathbf{Z} over FF by means of the exact sequence (19): first, the decomposition g=gzg=g^{\prime}z with gGderg^{\prime}\in G^{\operatorname{der}} and zZz\in Z is defined only up to replacing gg^{\prime} and zz with agag^{\prime}, azaz (a𝐀(F)a\in{\mathbf{A}}(F)), and second, the map (g,z)gz(g^{\prime},z)\to g^{\prime}z is in general not surjective on FF-points: for example for GL2\operatorname{GL}_{2}, its image only consists of elements whose determinant is a square in FF.

The way to deal with this issue is described in [FLN10, (3.15)]: the set of FF-points of the Steiberg-Hitchin base 𝔄G\mathfrak{A}_{G} is defined as the union over cocycles ηH1(F,A)\eta\in H^{1}(F,A) of the spaces (𝔅η(F)×𝐙η(F))/𝐀(F)(\mathfrak{B}_{\eta}(F)\times\mathbf{Z}_{\eta}(F))/{\mathbf{A}}(F), where 𝔅η\mathfrak{B}_{\eta} and 𝐙η\mathbf{Z}_{\eta} are torsors of, respectively, 𝔸Gder\mathbb{A}_{G^{\operatorname{der}}} and 𝐙\mathbf{Z}, defined by the cocycle η\eta.

Finally, note that for the Lie algebra there is no issue because the Lie algebra actually splits as a direct sum 𝔤=𝔤der𝔷\mathfrak{g}=\mathfrak{g}^{\operatorname{der}}\oplus\mathfrak{z}, and this is why we could treat all reductive Lie algebras above on equal footing.

The situation is more complicated if 𝐆der\mathbf{G}^{\operatorname{der}} is not simply connected, as Steinberg quotient in this case will no longer be an affine space. We will not address this case (as well as the non-split case) in this note.

3.3. Weyl discriminant

We recall the definition and the basic properties of the Weyl discriminant (for the Lie algebra, the main source is [Kot05, §§7, 14]).

3.3.1. Weyl disriminant on the Lie algebra

Definition 3.5.

Let 𝔤\mathfrak{g} be a reductive Lie algebra, let X𝔤X\in\mathfrak{g} be a regular semisimple element, and let T=CG(X)T=C_{G}(X) be its centralizer with the Lie algebra 𝔱=Lie(T)\mathfrak{t}=\operatorname{Lie}(T). Then

D(X)=det(ad(X)|𝔤/𝔱)D(X)=\det(\operatorname{ad}(X)|_{\mathfrak{g}/\mathfrak{t}})

is called the Weyl discriminant of XX.

The discriminant is, in fact, a polynomial function on 𝔤\mathfrak{g} (and thus extends to all of 𝔤\mathfrak{g} from the dense subset of regular semisimple elements): D(X)D(X) is the lowest non-vanishing coefficient of the characteristic polynomial of ad(X)\operatorname{ad}(X) (see [Kot05, 7.5]). This interpretation allows us to give an intrinsic characterization of the set of regular semisimple elements: in fact, X𝔤X\in\mathfrak{g} is regular semisimple if and only if D(X)0D(X)\neq 0; thus it can be taken as a definition of regular semisimple.

We also recall the expression for D(X)D(X) in terms of roots:

(20) D(X)=αΦα(X)=(1)dim𝔤rank𝔤2(αΦ+α(X))2,D(X)=\prod_{\alpha\in\Phi}\alpha(X)=(-1)^{\frac{\dim\mathfrak{g}-\operatorname{rank}\mathfrak{g}}{2}}\left(\prod_{\alpha\in\Phi^{+}}\alpha(X)\right)^{2},

where Φ\Phi is the set of all roots and Φ+\Phi^{+} is any set of positive roots.

Example 3.6.

We compute the explicit expressions for the Weyl discriminant in terms of the eigenvalues of XX, in the cases 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n} and 𝔤=𝔰𝔭2n\mathfrak{g}=\mathfrak{sp}_{2n}, for use in future examples.

Ler X𝔤X\in\mathfrak{g} have eigenvalues λiF¯\lambda_{i}\in\bar{F}.

For 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}, the roots are αij(X)=λiλj\alpha_{ij}(X)=\lambda_{i}-\lambda_{j}, 1i,jn and ij1\leq i,j\leq n\text{ and }i\neq j. Then the Weyl discriminant of XX coincides with the polynomial discriminant of the characteristic polynomial of XX:

D(X)=1i,jnij(λiλj).D(X)=\prod_{{1\leq i,j\leq n}\atop{i\neq j}}(\lambda_{i}-\lambda_{j}).

(We observe that the eigenvalues satisfy the relation i=1nλi=Tr(X)=0\sum_{i=1}^{n}\lambda_{i}=\operatorname{Tr}(X)=0).

For 𝔤=𝔰𝔭n\mathfrak{g}=\mathfrak{sp}_{n}, the explicit expression for the roots depends on the choice of the coordinates for the standard representation (though of course the answer does not). We define Sp2n\operatorname{Sp}_{2n} and 𝔰𝔭2n\mathfrak{sp}_{2n} explicitly as:

Sp2n(F)={gGL2n(F):gtJg=J},𝔰𝔭2n(F)={X𝔤𝔩2n(F):XtJ+JX=0},\operatorname{Sp}_{2n}(F)=\{g\in\operatorname{GL}_{2n}(F):g^{t}Jg=J\},\quad\mathfrak{sp}_{2n}(F)=\{X\in\mathfrak{gl}_{2n}(F):X^{t}J+JX=0\},

where J=[0InIn0]J=\left[\begin{smallmatrix}0&I_{n}\\ -I_{n}&0\end{smallmatrix}\right] and InI_{n} stands for the n×nn\times n-identity matrix. Then the eigenvalues of any element X𝔰𝔭2nX\in\mathfrak{sp}_{2n} satisfy λn+i=λi\lambda_{n+i}=-\lambda_{i}, 1in1\leq i\leq n, and the set of values of the roots at XX is (cf. [Hum72, §12.1]): {±(λi±λj),1i,jn,ij}{±2λi,1in}\{\pm(\lambda_{i}\pm\lambda_{j}),1\leq i,j\leq n,i\neq j\}\cup\{\pm 2\lambda_{i},1\leq i\leq n\}. Then we get:

D(X)=(1)n(n+1)22n1i,jnij(λi2λj2)i=1nλi.D(X)=(-1)^{\frac{n(n+1)}{2}}2^{n}\prod_{{1\leq i,j\leq n}\atop{i\neq j}}(\lambda_{i}^{2}-\lambda_{j}^{2})\prod_{i=1}^{n}\lambda_{i}.

We also observe that in a reductive Lie algebra, the Weyl discriminant of any element is computed entirely via the derived subalgebra 𝔤der\mathfrak{g}^{\operatorname{der}}, by definition (since 𝔤/𝔱=𝔤der/𝔱der\mathfrak{g}/\mathfrak{t}=\mathfrak{g}^{\operatorname{der}}/\mathfrak{t}^{\operatorname{der}}).

3.3.2. Weyl discriminant on the group

On the group, the definition is obtained essentially by reducing to the Lie algebra:

Definition 3.7.

Let γ𝐆(F)\gamma\in\mathbf{G}(F) be a regular semisimple element, and let T=CG(X)T=C_{G}(X) be its centralizer with the Lie algebra 𝔱=Lie(T)\mathfrak{t}=\operatorname{Lie}(T). Then the Weyl discriminant of γ\gamma is

D(γ)=det(1Ad(γ)|𝔤/𝔱).D(\gamma)=\det(1-\operatorname{Ad}(\gamma)|_{\mathfrak{g}/\mathfrak{t}}).

Similarly to the Lie algebra case, the Weyl discriminant has an expression in terms of the (multiplicative) roots:

(21) D(γ)=αΦ(1α(γ))=(1)dim𝔤rank𝔤2ρ2(γ)(αΦ+(1α(γ)))2,D(\gamma)=\prod_{\alpha\in\Phi}(1-\alpha(\gamma))=(-1)^{\frac{\dim\mathfrak{g}-\operatorname{rank}\mathfrak{g}}{2}}\rho^{2}(\gamma)\left(\prod_{\alpha\in\Phi^{+}}(1-\alpha(\gamma))\right)^{2},

where ρ\rho is half the sum of positive roots, so 2ρ2\rho is the sum of positive roots (in the above formula, ρ2(γ)\rho^{2}(\gamma) is the value of the character 2ρ2\rho at γ\gamma). Note that the second part of the formula expressing the Weyl discriminant as a product over positive roots now has an extra factor that did not arise in the Lie algebra case (the examples below illustrate this).

We again show the calculation for the general linear and symplectic groups. Note that the final expressions are a lot simpler when restricted to 𝐆der\mathbf{G}^{\operatorname{der}}.

Example 3.8.

In all examples, we give an explicit expression for the Weyl discriminant of a regular semisimple element γG(F)\gamma\in G(F) with eigenvalues {λi}F¯\{\lambda_{i}\}\subset\bar{F}. We observe that these expressions do not depend on the field (so one could even consider F=F=\mathbb{C}).

  1. (1)

    G=GL2G=\operatorname{GL}_{2}: D(γ)=(1λ1λ2)(1λ2λ1)=(λ1λ2)2det(γ)D(\gamma)=(1-\frac{\lambda_{1}}{\lambda_{2}})(1-\frac{\lambda_{2}}{\lambda_{1}})=-\frac{(\lambda_{1}-\lambda_{2})^{2}}{\det(\gamma)}.

  2. (2)

    G=GLnG=\operatorname{GL}_{n}: D(γ)=1i,jnij(1λiλj)D(\gamma)=\prod_{{1\leq i,j\leq n}\atop{i\neq j}}\left(1-\frac{\lambda_{i}}{\lambda_{j}}\right).

  3. (3)

    G=Sp2nG=\operatorname{Sp}_{2n}: D(γ)=1i<jndij1indiD(\gamma)=\prod_{1\leq i<j\leq n}d_{ij}\cdot\prod_{1\leq i\leq n}d_{i}, where

    dij\displaystyle d_{ij} =(1λiλj)(1λjλi)(1λiλj)(11λiλj)\displaystyle=\left(1-\frac{\lambda_{i}}{\lambda_{j}}\right)\left(1-\frac{\lambda_{j}}{\lambda_{i}}\right)\left(1-\lambda_{i}\lambda_{j}\right)\left(1-\frac{1}{\lambda_{i}\lambda_{j}}\right)
    di\displaystyle d_{i} =(1λi2)(11/λi2).\displaystyle=\left(1-\lambda_{i}^{2}\right)\left(1-1/\lambda_{i}^{2}\right).
  4. (4)

    G=GSp2nG={\mathrm{GSp}}_{2n}. By definition, GSp2n(F){\mathrm{GSp}}_{2n}(F) is the algebraic group whose functor of points is defined as, for any FF-algebra RR,

    GSp2n(R)={gGL2n(R):ν(g)R×,gtJg=ν(g)J},{\mathrm{GSp}}_{2n}(R)=\{g\in\operatorname{GL}_{2n}(R):\exists\nu(g)\in R^{\times},g^{t}Jg=\nu(g)J\},

    where JJ is the same matrix as the one used to define Sp2n\operatorname{Sp}_{2n}. It fits into the exact sequence of algebraic groups

    1Sp2nGSp2n𝔾m1,1\to\operatorname{Sp}_{2n}\to{\mathrm{GSp}}_{2n}\to\mathbb{G}_{m}\to 1,

    where the map to 𝔾m\mathbb{G}_{m} is the map gν(g)g\mapsto\nu(g), called the multiplier. We have GSp2nder=Sp2n{\mathrm{GSp}}_{2n}^{\operatorname{der}}=\operatorname{Sp}_{2n}, so GSp2n{\mathrm{GSp}}_{2n} is a good example (other than GLn\operatorname{GL}_{n}) of a reductive but not semi-simple algebraic group whose derived subgroup is simply connected.

    If the element γ\gamma has multiplier ν\nu, then as above for G=Sp2nG=\operatorname{Sp}_{2n}, D(γ)=1i<jndij1indiD(\gamma)=\prod_{1\leq i<j\leq n}d_{ij}\cdot\prod_{1\leq i\leq n}d_{i}, but now we have

    dij\displaystyle d_{ij} =(1λiλj)(1λjλi)(1λiλjν)(1νλiλj)\displaystyle=\left(1-\frac{\lambda_{i}}{\lambda_{j}}\right)\left(1-\frac{\lambda_{j}}{\lambda_{i}}\right)\left(1-\frac{\lambda_{i}\lambda_{j}}{\nu}\right)\left(1-\frac{\nu}{\lambda_{i}\lambda_{j}}\right)
    di\displaystyle d_{i} =(1λi2ν)(1νλi2).\displaystyle=\left(1-\frac{\lambda_{i}^{2}}{\nu}\right)\left(1-\frac{\nu}{\lambda_{i}^{2}}\right).

3.4. Orbital integrals: the Lie algebra case

We start with a prototype case of a Lie algebra.

3.4.1. Definitions: Lie algebra

Let 𝐆\mathbf{G} be a connected reductive group defined over a local field FF, as above. The orbital integrals (for regular semisimple elements) on the Lie algebra are distributions on the space Cc(𝔤)C_{c}^{\infty}(\mathfrak{g}) of the locally constant compactly supported functions on 𝔤\mathfrak{g}, defined as follows.

Let X𝔤X\in\mathfrak{g} be a regular semisimple element, and let fCc(𝔤)f\in C_{c}^{\infty}(\mathfrak{g}). Since XX is regular semisimple, its centralizer is a torus T=CG(X)=𝐓(F)T=C_{G}(X)=\mathbf{T}(F), as discussed above, and thus the adjoint orbit of XX can be identified with the quotient 𝐓(F)\𝐆(F)\mathbf{T}(F)\backslash\mathbf{G}(F). Both 𝐓(F)\mathbf{T}(F) and 𝐆(F)\mathbf{G}(F) can be endowed with any of the natural measures discussed above in §3.1. Once the measures on GG and TT are fixed, there is a unique quotient measure on T\GT\backslash G, which we will denote by μT\G\mu_{T\backslash G} (see e.g., [Kot05, §2.4] for the definition of the quotient measure in this context). The orbital integral with respect to this measure is

(22) OX(f):=T\Gf(Ad(g1)X)𝑑μT\G.O_{X}(f):=\int_{T\backslash G}f(\operatorname{Ad}(g^{-1})X)d\mu_{T\backslash G}.

We observe that there are finitely many FF-conjugacy classes of tori in GG; thus there are finitely many choices of measures that we need to make on the representatives of these conjugacy classes, and these choices endow the orbit of every regular semisimple element with a measure. If the canonical measures (in the sense of [Gro97], discussed above in 2.2.2) are chosen on the tori, the resulting orbital integrals are called canonical. This approach to the normalization of measures on the orbits is the one typically used in the literature.

On the other hand, one can use Chevalley map defined above to normalize the measures on orbits. For a general reductive group 𝐆\mathbf{G} and X𝔤X\in\mathfrak{g} regular semisimple, the fibre 𝔠1(𝔠(X))\mathfrak{c}^{-1}(\mathfrak{c}(X)) over the point a:=𝔠(X)𝔸G(F)a:=\mathfrak{c}(X)\in\mathbb{A}_{G}(F) is the stable orbit of XX, which is a finite union of FF-rational orbits. Thus, every FF-rational orbit is an open subset of 𝔠1(a)\mathfrak{c}^{-1}(a) for some a𝔸G(F)a\in\mathbb{A}_{G}(F), and if we define a measure on this fibre, we get a measure on every FF-rational orbit contained in it by restriction.

In the Introduction, we have fixed measures on affine spaces with a choice of a basis. The Lie algebra 𝔤\mathfrak{g} is an affine space; it does not come with a canonical choice of a basis, and this choice would not matter much in the discussion below; we can choose an arbitrary FF-basis {ei}i=1n\{e_{i}\}_{i=1}^{n} of 𝔤\mathfrak{g} for our purposes. This basis then gives rise to a differential form ω𝔤=i=1ndxi\omega_{\mathfrak{g}}=\wedge_{i=1}^{n}dx_{i} on 𝔤\mathfrak{g}, which gives a measure |ω𝔤||\omega_{\mathfrak{g}}| as in the Introduction. The space 𝔸G\mathbb{A}_{G} is also an affine space under our assumptions (since at the moment we are working with the Lie algebra); and in our construction it comes with a choice of basis {ρi}i=1r\{\rho_{i}\}_{i=1}^{r} as in §3.2.1. We let ω𝔸G\omega_{\mathbb{A}_{G}} be the differential form associated with this basis.

Thus we get the quotient measure on each fibre 𝔠1(𝔠(X))\mathfrak{c}^{-1}(\mathfrak{c}(X)): it is the measure associated with the differential form ω𝔠(X)geom\omega_{\mathfrak{c}(X)}^{\operatorname{geom}} such that

(23) ω𝔤=ω𝔠(X)geomω𝔸G.\omega_{\mathfrak{g}}=\omega^{{\operatorname{geom}}}_{\mathfrak{c}(X)}\wedge\omega_{\mathbb{A}_{G}}.

That is, by definition of ω𝔠(X)geom\omega_{\mathfrak{c}(X)}^{{\operatorname{geom}}}, for any fCc(𝔤)f\in C_{c}^{\infty}(\mathfrak{g}),

(24) 𝔤f(X)d|ω𝔤|=𝔸G𝔠1(𝔠(X))f(X)d|ω𝔠(X)geom|d|ω𝔸G|.\int_{\mathfrak{g}}f(X)\,d|\omega_{\mathfrak{g}}|=\int_{\mathbb{A}_{G}}\int_{\mathfrak{c}^{-1}(\mathfrak{c}(X))}f(X)\,d|\omega^{{\operatorname{geom}}}_{\mathfrak{c}(X)}|\,d|\omega_{\mathbb{A}_{G}}|.

Our immediate goal is to derive the relationship between these two measures on the orbit: μT\G\mu_{T\backslash G} and |dωageom||d\omega_{a}^{{\operatorname{geom}}}|, where a=𝔠(X)a=\mathfrak{c}(X). First we observe that since both measures are quotient measures of a chosen Haar measure on GG, their ratio does not depend on the choice of the measure on GG, as long as it is compatible with the choice of the measure on the Lie algebra; thus at this point, the choice of the measure on GG is determined by our choice of the form ω𝔤\omega_{\mathfrak{g}}. (Conversely, one often chooses a measure on GG first, and this determines ω𝔤\omega_{\mathfrak{g}}.) At the same time, the choice of the measures on the representatives of conjugacy classes of tori affects the measure μT\G\mu_{T\backslash G} but not the measure |ωageom||\omega_{a}^{{\operatorname{geom}}}|. Here we address two natural choices of such measures:

  1. (i)

    Let ωG\omega_{G} be a volume form on GG, and on each algebraic torus TT, define the form ωT\omega_{T} using the characters of the torus as in (2). Then we get the quotient measure ωT\G\omega_{T\backslash G} on each orbit. This is the measure discussed in [FLN10]. We discuss this measure in this section.

  2. (ii)

    Use the measure denoted above by |ωcan||\omega^{\operatorname{can}}|, associated with the Néron model, on each torus. This measure on the orbits is discussed in the next section.

Thus our first goal is to determine the ratio of ωT\G\omega_{T\backslash G} to ωageom\omega_{a}^{\operatorname{geom}}, for each X𝔤rssX\in\mathfrak{g}^{\operatorname{rss}} (which determines TT and aa). It turns out that the conversion between these two measures is based on exactly the same calculation as the Weyl integration formula, which we now review.

3.4.2. Weyl integration formula, revisited

We follow [Kot05, §7, §14.1.1], and use the same notation (except we continue to use boldface letters to denote varieties). For a torus TGT\subset G, let WT=𝐍G(T)(F)/𝐓(F)W_{T}={\mathbf{N}_{G}(T)}(F)/\mathbf{T}(F) be the relative Weyl group of TT (cf. [Kot05, §7.1]). Weyl integration formula (which we quote in this form from [Kot05, §7.7]), for an arbitrary Schwartz-Bruhat function fCc(𝔤)f\in C_{c}^{\infty}(\mathfrak{g}), asserts:

(25) 𝔤f(Y)d|ω𝔤(Y)|=T1|WT|𝔱|D(X)|T\Gf(Ad(g1)X)d|ωT\G|d|ω𝔱(X)|,\int_{\mathfrak{g}}f(Y)d|\omega_{\mathfrak{g}}(Y)|=\sum_{T}\frac{1}{|W_{T}|}\int_{\mathfrak{t}}|D(X)|\int_{T\backslash G}f(\operatorname{Ad}(g^{-1})X)d|\omega_{T\backslash G}|\,d|\omega_{\mathfrak{t}}(X)|,

where the sum on the right-hand side is over the representatives of the conjugacy classes of tori in GG.

The proof of this formula relies on a computation of the Jacobian of the map

(26) (T\G)×𝔤\displaystyle(T\backslash G)\times\mathfrak{g} 𝔤\displaystyle\to\mathfrak{g}
(g,X)\displaystyle(g,X) Ad(g1)X.\displaystyle\mapsto\operatorname{Ad}(g^{-1})X.

This map is |WT|:1|W_{T}|:1, and its Jacobian at XX is precisely |D(X)||D(X)| (see [Kot05, §7.2] for a beautiful exposition).

3.4.3. The relation between geometric and canonical orbital integrals for the Lie algebra

Let us just naïvely compare the right-hand side of the Weyl integration formula with the right-hand side of (24) above (since the left-hand sides are the same). First, note (as already discussed above) that our space 𝔸G\mathbb{A}_{G} is, up to a set of measure zero, a disjoint union of images of the representatives of the conjugacy classes of tori, and for each torus, Chevalley map 𝔠:T𝔸G\mathfrak{c}:T\to\mathbb{A}_{G} is |WT|:1|W_{T}|:1. Thus, the right-hand side of (23) would look exactly like the right-hand side of the Weyl integration formula (25) if we could replace integration over 𝔸G\mathbb{A}_{G} with the sum of integrals over the representatives of the conjugacy classes of Cartan subalgebras 𝔱=Lie(T)\mathfrak{t}=\operatorname{Lie}(T) (as TT ranges over the conjugacy classes of maximal tori).

The situation is summarized by the commutative diagram:

(27) (T\G)×𝔱\textstyle{(T\backslash G)\times\mathfrak{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔤\textstyle{\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔱\textstyle{\mathfrak{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔸G=𝔱spl/W\textstyle{\mathbb{A}_{G}=\mathfrak{t}^{\operatorname{spl}}/W}

Here the horizontal map on the top is the map (26); this map is |WT|:1|W_{T}|:1 and its Jacobian at (g,X)(g,X) is |D(X)||D(X)| (see [Kot05, §7.2]). The vertical arrow on the left is projection onto 𝔱\mathfrak{t}; the vertical arrow on the right is Chevalley map 𝔠\mathfrak{c}; and the horizontal arrow at the bottom is 𝔠|𝔱\mathfrak{c}|_{\mathfrak{t}}, which is also |WT|:1|W_{T}|:1 .

We have the forms ω𝔤\omega_{\mathfrak{g}} on 𝔤\mathfrak{g} and ω𝔸G\omega_{\mathbb{A}_{G}} on 𝔸G\mathbb{A}_{G}; let us choose the invariant differential form ωT\omega_{T} on TT defined by the characters of TT as in (2); we also need an invariant top degree form ωG\omega_{G} on GG, which is required to be compatible with ω𝔤\omega_{\mathfrak{g}} under the exponential map, which determines it uniquely. As discussed above, given ωG\omega_{G} and ωT\omega_{T}, we get the quotient measure |ωT\G||\omega_{T\backslash G}| that corresponds to a differential form ωT\G\omega_{T\backslash G} satisfying ωTωT\G=ωG\omega_{T}\wedge\omega_{T\backslash G}=\omega_{G}, and a differential form ω𝔠1(a)geom\omega^{{\operatorname{geom}}}_{\mathfrak{c}^{-1}(a)} on each fibre of the map 𝔠:𝔤𝔸G\mathfrak{c}:\mathfrak{g}\to\mathbb{A}_{G}. Both ωT\G\omega_{T\backslash G} and ω𝔠1(a)geom\omega_{\mathfrak{c}^{-1}(a)}^{{\operatorname{geom}}} are generators of the top exterior power of the cotangent bundle of 𝐓\𝐆\mathbf{T}\backslash\mathbf{G}, hence they differ by a constant (which can depend on aa).

Looking at the top, right and bottom maps in the diagram (27), respectively, we see that these differential forms are related as follows (the first and third lines follow from the Jacobian formula and the fact that the horizontal maps are |WT|:1|W_{T}|:1; the second line is the definition of ω𝔠1(a)geom\omega^{{\operatorname{geom}}}_{\mathfrak{c}^{-1}(a)} with a=𝔠(X)a=\mathfrak{c}(X)):

(28) |WT|1|D(X)|ωT\Gω𝔱=ω𝔤\displaystyle|W_{T}|^{-1}|D(X)|\omega_{T\backslash G}\wedge\omega_{\mathfrak{t}}=\omega_{\mathfrak{g}}
ω𝔤(X)=ω𝔠1(𝔠(X))geomω𝔸G(𝔠(X))\displaystyle\omega_{\mathfrak{g}}(X)=\omega^{{\operatorname{geom}}}_{\mathfrak{c}^{-1}(\mathfrak{c}(X))}\wedge\omega_{\mathbb{A}_{G}}(\mathfrak{c}(X))
|WT|1|Jac(𝔠|𝔱)|ω𝔱=ω𝔸G.\displaystyle|W_{T}|^{-1}|{\operatorname{Jac}}(\mathfrak{c}|_{\mathfrak{t}})|\omega_{\mathfrak{t}}=\omega_{\mathbb{A}_{G}}.

We conclude that

(29) |D(X)|ωT\G(X)=|Jac(𝔠|𝔱)(X)|ω𝔠1(𝔠(X))geom,X𝔱.|D(X)|\omega_{T\backslash G}(X)=|{\operatorname{Jac}}(\mathfrak{c}|_{\mathfrak{t}})(X)|\omega^{{\operatorname{geom}}}_{\mathfrak{c}^{-1}(\mathfrak{c}(X))},\quad X\in\mathfrak{t}.

The Jacobian of the restriction of Chevalley map 𝔠|𝔱\mathfrak{c}|_{\mathfrak{t}} at XX is αΦ+α(X)\prod_{\alpha\in\Phi^{+}}\alpha(X), up to a constant in F×F^{\times} (see [Kot05, §14.1]). This constant depends on the choice of coordinates on 𝔱\mathfrak{t}. We use the basis of the character lattice {χi}i=1r\{\chi_{i}\}_{i=1}^{r}, as in (2), to define the coordinates on 𝔱\mathfrak{t}. With this choice of coordinates, the constant turns out to be ±1\pm 1; the sign depends on the ordering of the characters χi\chi_{i} and does not affect the resulting measure. The reason for this is that the constant is 11 for the split torus (this is not trivial; it follows from the argument in [Bou02, ch.5,§5], and the group version of this statement is also proved in [FLN10] over \mathbb{C} (see proof of Proposition 3.29, especially (3.33) and (3.34)); the argument holds for any split torus). If TT is not split, we can work over an extension EE where TT splits, and since our coordinate system is precisely the one used for the split torus over EE, the equality continues to hold. We observe that on the Lie algebra, we have

(30) |αΦ+α(X)|=|D(X)|1/2.|\prod_{\alpha\in\Phi^{+}}\alpha(X)|=|D(X)|^{1/2}.

Putting the relations (28), (29), and (30) together, we obtain the following Proposition.

Proposition 3.9.

(cf. [FLN10, Proposition 3.29].) Let 𝔠:𝔤𝔸G\mathfrak{c}:\mathfrak{g}\to\mathbb{A}_{G} be Chevalley map as above; let X𝔤X\in\mathfrak{g} be a regular semisimple element, let the algebraic torus TT be its centralizer, with the Lie algebra 𝔱\mathfrak{t}. Then with the measures defined as above, we have:

|ω𝔠1(𝔠(X))geom|=|D(X)|1/2|ωT\G|.|\omega^{{\operatorname{geom}}}_{\mathfrak{c}^{-1}(\mathfrak{c}(X))}|=|D(X)|^{1/2}|\omega_{T\backslash G}|.

We conclude this section with an example illustrating the proposition.

Example 3.10.

Let 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} and let 𝔱=𝔱spl\mathfrak{t}=\mathfrak{t}^{{\operatorname{spl}}} be the subalgebra of diagonal matrices. Then we have 𝔠:X(t):=[t00t]t2\mathfrak{c}:X(t):=\left[\begin{smallmatrix}t&0\\ 0&-t\end{smallmatrix}\right]\mapsto-t^{2}; here the Jacobian is just the derivative (since we are dealing with a function of one variable), so Jac(𝔠|𝔱)=2t=α(X(t)){\operatorname{Jac}}(\mathfrak{c}|_{\mathfrak{t}})=-2t=-\alpha(X(t)).

Now consider 𝔱E\mathfrak{t}_{E} – a non-split Cartan subalgebra corresponding to a quadratic extension E=F[ϵ]E=F[\sqrt{\epsilon}]: 𝔱E={X(t):=[0tϵt0],tF}\mathfrak{t}_{E}=\left\{X(t):=\left[\begin{smallmatrix}0&t\\ \epsilon t&0\end{smallmatrix}\right],\ t\in F\right\}. We have 𝔠|𝔱E=ϵt2\mathfrak{c}|_{\mathfrak{t}_{E}}=-\epsilon t^{2}, and its Jacobian is 2ϵt=ϵα(X(t))-2\epsilon t=-\sqrt{\epsilon}\alpha(X(t)) (note that the eigenvalues of our element are ±ϵt\pm\sqrt{\epsilon}t). At the same time, on 𝔱E\mathfrak{t}_{E}, the measure ωT\omega_{T} is ϵdt\sqrt{\epsilon}dt (recall that ωT\omega_{T} is defined by means of characters of TT over the algebraic closure). Hence, with this choice of the differential form, we obtain, again, with a=𝔠(X(t))a=\mathfrak{c}(X(t)):

da=ϵα(X(t))dt=α(X(t))ωT(X(t)).da=-\sqrt{\epsilon}\alpha(X(t))dt=-\alpha(X(t))\omega_{T}(X(t)).

3.5. The simplest group case

Let us assume that 𝐆\mathbf{G} is semi-simple, split, and simply connected. We are now almost ready to explain the relation (3.31) of [FLN10] (see equation (18) above). The definitions are essentially the same as in the Lie algebra case:

  • an orbit of a regular semisimple element γG:=𝐆(F)\gamma\in G:=\mathbf{G}(F), as a manifold over FF, can be identified with T\GT\backslash G, where TT is the centralizer of γ\gamma. As above, if ωG\omega_{G} is a volume form on GG and ωT\omega_{T} - a volume form on TT, we get the measure |ωT\G||\omega_{T\backslash G}| on the orbit of γ\gamma.

  • The regular fibres of the map 𝔠:G𝔸G\mathfrak{c}:G\to\mathbb{A}_{G} are stable orbits; each stable orbit of a regular semisimple element is a finite disjoint union of FF-rational orbits, and thus we get the geometric measure |ωageom||\omega_{a}^{{\operatorname{geom}}}| on each such orbit, by considering the quotient of the measures on GG and 𝔸G\mathbb{A}_{G}.

For γG\gamma\in G, let

Δ(γ):=ρ1(γ)α>0(1α(γ));thus |Δ(γ)|2=|D(γ)|.\Delta(\gamma):=\rho^{-1}(\gamma)\prod_{\alpha>0}(1-\alpha(\gamma));\quad\text{thus }|\Delta(\gamma)|^{2}=|D(\gamma)|.
Theorem 3.11.

([FLN10, Relation (3.31)].) Let 𝐆\mathbf{G} be a connected semi-simple simply connected group over a local field FF, and let γG\gamma\in G be a regular semisimple element. Then for any fCc(G)f\in C_{c}^{\infty}(G), the orbital integrals with respect to the geometric measure on the orbit of γ\gamma, and the measure ωT\G\omega_{T\backslash G} (which, by definition, is the quotient of the measures |ωG||\omega_{G}| on GG and |ωT||\omega_{T}| on TT, with ωT\omega_{T} defined by (2)) are related via:

𝒪(γ)f(g)d|ω𝔠(γ)geom(g)|=|Δ(γ)|T\Gf(Ad(g1)γ)ωT\G(g),\int_{\mathcal{O}(\gamma)}f(g)d|\omega_{\mathfrak{c}(\gamma)}^{{\operatorname{geom}}}(g)|=|\Delta(\gamma)|\int_{T\backslash G}f(\operatorname{Ad}(g^{-1})\gamma)\omega_{T\backslash G}(g),

where on the left, the orbit 𝒪(γ)\mathcal{O}(\gamma) is thought of as an open subset of the stable orbit 𝔠1(𝔠(γ))\mathfrak{c}^{-1}(\mathfrak{c}(\gamma)) and endowed with the geometric measure |ω𝔠(γ)geom(g)||\omega_{\mathfrak{c}(\gamma)}^{{\operatorname{geom}}}(g)| as above.

We first explain two differences with the statement in [FLN10].

Remark 3.12.

Our expression does not (yet) include the factor L(1,σT\G)L(1,\sigma_{T\backslash G}) that appears in (3.31) of [FLN10]. This factor appears simply by their definition of the measure dg¯v:=L(1,σT\G)ωT\Gd\bar{g}_{v}:=L(1,\sigma_{T\backslash G})\omega_{T\backslash G} which appears on the right-hand side of (3.31). As we shall see in the next section, using the measure dg¯vd\bar{g}_{v} ensures that the local orbital integral on the right-hand side is 11 for almost all places of a given number field (which is desirable for defining the orbital integral globally), and for almost all places this coincides with the orbital integral with respect to the canonical measure.

Remark 3.13.

Note that we stated the theorem as a relation between orbital integrals, whereas in [FLN10] it is stated as a relation between stable orbital integrals. Since the measure is a local notion, this is an equivalent statement: in fact, the assertion of the theorem is just that the two measures on the stable orbit (and hence, by restriction, on every rational orbit) are related via

|ω𝔠(γ)geom(g)|=|Δ(γ)||ωT\G(g)|.|\omega_{\mathfrak{c}(\gamma)}^{{\operatorname{geom}}}(g)|=|\Delta(\gamma)||\omega_{T\backslash G}(g)|.

3.5.1. Sketch of the proof.

As the measures are defined by differential forms, the calculation is carried out in the exterior power of the cotangent space, and hence it is essentially the same calculation as for the Lie algebra above. The only ingredients that needs to be treated slightly differently are the discriminant and the Jacobian of the map from TT to T/WT/W. Indeed, for γGrss(Fsep)\gamma\in G^{\operatorname{rss}}(F^{\operatorname{sep}}), we still have the exact sequence of tangent spaces (see [FLN10, Lemma 26])

0Tanγ(𝔠1(a))Tanγ𝐆Tana(𝔸G)0,0\to Tan_{\gamma}(\mathfrak{c}^{-1}(a))\to Tan_{\gamma}\mathbf{G}\to Tan_{a}(\mathbb{A}_{G})\to 0,

and by definition, ω𝔠(γ)geomω𝔸G=ωG\omega_{\mathfrak{c}(\gamma)}^{\operatorname{geom}}\wedge\omega_{\mathbb{A}_{G}}=\omega_{G}; ωG=ωTωT\G\omega_{G}=\omega_{T}\wedge\omega_{T\backslash G}. The proof proceeds exactly as for Lie algebras, except the map (26) needs to be replaced with the map

(31) (T\G)×G\displaystyle(T\backslash G)\times G G\displaystyle\to G
(g,γ)\displaystyle(g,\gamma) g1γg,\displaystyle\mapsto g^{-1}\gamma g,

and the map 𝔱𝔸G=T/W\mathfrak{t}\to\mathbb{A}_{G}=T/W is replaced with the map TT/WT\to T/W. The Jacobian of the first map is the group version of the Weyl discriminant (and fits into the group version of Weyl integration formula in the exact same way as it did for the Lie algebra):

|WT|1|D(γ)|ωT\GωT=ωG.|W_{T}|^{-1}|D(\gamma)|\omega_{T\backslash G}\wedge\omega_{T}=\omega_{G}.

Next, we need to relate ωT\omega_{T} with ω𝔸G\omega_{\mathbb{A}_{G}}.

Lemma 3.14.

([FLN10, Proposition 3.29].) Let 𝐆\mathbf{G} be a split, semi-simple, simply connected group, and let TGT\subset G be a maximal torus. Let ωT\omega_{T} be defined by (2). Then

|WT|1ωT(γ)=|Δ(γ)|ω𝔸G(𝔠(γ)).|W_{T}|^{-1}\omega_{T}(\gamma)=|\Delta(\gamma)|\omega_{\mathbb{A}_{G}}(\mathfrak{c}(\gamma)).
Proof.

For T=TsplT=T^{\operatorname{spl}}, this is proved in [FLN10], as well as in [Bou02] (where the field is assumed algebraically closed, but the proof works verbatim for the split torus). Now it remains to consider the restriction of 𝔠\mathfrak{c} to an arbitrary (not necessarily split) maximal torus. The map 𝔠\mathfrak{c} on TT can be defined as a composition

𝐓𝐓spl𝔸G=𝐓spl/W,\mathbf{T}\to\mathbf{T}^{\operatorname{spl}}\to\mathbb{A}_{G}=\mathbf{T}^{\operatorname{spl}}/W,

where the first map is an isomorphism over the algebraic closure of FF. The pullback of the form ωTspl\omega_{T^{\operatorname{spl}}} on TsplT^{\operatorname{spl}} is precisely the form ωT\omega_{T} on TT, and thus the equality remains true. ∎

The theorem follows, precisely as in the Lie algebra case. To conclude this section, we compute some examples illustrating the above Lemma (which show that it is substantially non-trivial even for the split torus).

3.5.2. Examples of Jacobians and discriminants on the group

Example 3.15.

We again start with 𝐆=SL2\mathbf{G}=\operatorname{SL}_{2}. Let γt=[t00t1]\gamma_{t}=\left[\begin{smallmatrix}t&0\\ 0&t^{-1}\end{smallmatrix}\right]. The map 𝔠\mathfrak{c} on the diagonal torus is given by: 𝔠:[t00t1]t+t1\mathfrak{c}:\left[\begin{smallmatrix}t&0\\ 0&t^{-1}\end{smallmatrix}\right]\mapsto t+t^{-1}. Its Jacobian (i.e., the derivative) is 1t21-t^{-2}; so we get:

Jac(𝔠|Tspl)(γt)=1t2=(1(α)(γt))=t2(1α(γt)).{\operatorname{Jac}}(\mathfrak{c}|_{T^{{\operatorname{spl}}}})(\gamma_{t})=1-t^{-2}=(1-(-\alpha)(\gamma_{t}))=t^{-2}(1-\alpha(\gamma_{t})).

We observe that for SL2\operatorname{SL}_{2}, the half-sum of positive roots is ρ=12α\rho=\frac{1}{2}\alpha, so ρ(γt)=t\rho(\gamma_{t})=t.

However, this is not yet the whole story. We are interested in the ratio between the measure ωT=dtt\omega_{T}=\frac{dt}{t} on TT and the measure dada on 𝔸1T/W\mathbb{A}^{1}\simeq T/W, and our map, as above, is given by a=t+t1a=t+t^{-1}. We just computed: da=(1t2)dtda=(1-t^{-2})dt. Then we have:

dtt=(1t2)1tda=ρ(γt)α>0(1α(γt))1da.\frac{dt}{t}=\frac{(1-t^{-2})^{-1}}{t}da=\rho(\gamma_{t})\prod_{\alpha>0}(1-\alpha(\gamma_{t}))^{-1}da.

It is instructive to do one more, higher rank, example.

Example 3.16.

Let 𝐆=Sp4\mathbf{G}=\operatorname{Sp}_{4} (defined explicitly as in Example 3.6 above), and consider the split torus T={diag(t1,t2,t11,t21)tiF×}T=\{\operatorname{diag}(t_{1},t_{2},t_{1}^{-1},t_{2}^{-1})\mid t_{i}\in F^{\times}\}.

Let γt1,t2=diag(t1,t2,t11,t21)\gamma_{t_{1},t_{2}}=\operatorname{diag}(t_{1},t_{2},t_{1}^{-1},t_{2}^{-1}). In these coordinates, Steinberg map is given explicitly by the elementary symmetric polynomials:

𝔠:γt1,t2(a,b),\displaystyle\mathfrak{c}:\gamma_{t_{1},t_{2}}\mapsto(a,b),
a=t1+t2+t11+t21,b=t1t2+t2t11+t1t21+t11t21+2.\displaystyle a=t_{1}+t_{2}+t_{1}^{-1}+t_{2}^{-1},\quad b=t_{1}t_{2}+t_{2}t_{1}^{-1}+t_{1}t_{2}^{-1}+t_{1}^{-1}t_{2}^{-1}+2.

The Jacobian of this map is (we are skipping the details of a painful calculation)

|at1bt1at2bt2|=(1t12)(1t22)(1t11t21)(t1t2),\left|\begin{smallmatrix}\frac{\partial a}{\partial t_{1}}&\frac{\partial b}{\partial t_{1}}\\ \frac{\partial a}{\partial t_{2}}&\frac{\partial b}{\partial t_{2}}\end{smallmatrix}\right|=(1-t_{1}^{-2})(1-t_{2}^{-2})(1-t_{1}^{-1}t_{2}^{-1})(t_{1}-t_{2}),

which we recognize as:

Jac(𝔠|T)(γt1,t2)=t1α<0(1α(γt1,t2)),{\operatorname{Jac}}(\mathfrak{c}|_{T})(\gamma_{t_{1},t_{2}})=t_{1}\prod_{\alpha<0}(1-\alpha(\gamma_{t_{1},t_{2}})),

Note the factor of t1t_{1} in front (which is not a root value). Thus we obtained:

dt1dt2=±α<0(1α(γt1,t2))1t11dadb\displaystyle dt_{1}\wedge dt_{2}=\pm\prod_{\alpha<0}(1-\alpha(\gamma_{t_{1},t_{2}}))^{-1}t_{1}^{-1}da\wedge db
=±(α>0(1α(γt1,t2))1)ρ2(γt1,t2)t11dadb.\displaystyle=\pm\left(\prod_{\alpha>0}(1-\alpha(\gamma_{t_{1},t_{2}}))^{-1}\right)\rho^{2}(\gamma_{t_{1},t_{2}})t_{1}^{-1}da\wedge db.

The plus-minus sign in front is not important and depends on the ordering of the coordinates.

Now, we are interested in the ratio between the invariant measure ωT=dt1dt2t1t2\omega_{T}=\frac{dt_{1}\wedge dt_{2}}{t_{1}t_{2}} on TT and the measure dω𝔸G=dadbd\omega_{\mathbb{A}_{G}}=da\wedge db on T/WT/W. We note that in this case ρ\rho, the half-sum of positive roots, evaluated at γt1,t2\gamma_{t_{1},t_{2}} is ρ(γt1,t2)=(t12t22t1t2(t1t2))1/2=t12t2\rho(\gamma_{t_{1},t_{2}})=\left(t_{1}^{2}t_{2}^{2}\frac{t_{1}}{t_{2}}(t_{1}t_{2})\right)^{1/2}=t_{1}^{2}t_{2}, and compute further (here we write γ:=γt1,t2\gamma:=\gamma_{t_{1},t_{2}} to avoid notational clutter):

ωT\displaystyle\omega_{T} =dt1dt2t1t2=±(α>0(1α(γ))1)ρ2(γ)t11t1t2dadb\displaystyle=\frac{dt_{1}\wedge dt_{2}}{t_{1}t_{2}}=\frac{\pm\left(\prod_{\alpha>0}(1-\alpha(\gamma))^{-1}\right)\rho^{2}(\gamma)t_{1}^{-1}}{t_{1}t_{2}}da\wedge db
=±(α>0(1α(γ)1)ρ(γ)dadb=Δ(γ)dadb.\displaystyle=\pm\left(\prod_{\alpha>0}(1-\alpha(\gamma)^{-1}\right)\rho(\gamma)da\wedge db=\Delta(\gamma)da\wedge db.

3.6. The general case

First, suppose that 𝐆\mathbf{G} is a connected split reductive group over FF, with 𝐆der\mathbf{G}^{\operatorname{der}} simply connected. Then if one uses the correct general notion of Steinberg-Hitchin base as defined in [FLN10], all measures are invariant under the action of the centre, and hence relation (3.31) of [FLN10] holds in this case as well, with no further proof needed.

If 𝐆der\mathbf{G}^{\operatorname{der}} is not simply connected, the space we denoted by 𝔸Gder\mathbb{A}_{G^{\operatorname{der}}} is no longer an affine space, and one needs to use zz-extensions. If 𝐆\mathbf{G} is not split, we need to consider Galois action on Steinberg-Hitchin base. Both topics are discussed in [FLN10] but are beyond the scope of these notes.

3.7. Aside: naïve approach for classical groups – what works and what doesn’t

Suppose for a moment that GGLn(F)G\hookrightarrow\operatorname{GL}_{n}(F) is a split classical reductive group. It is tempting (and often done in Number theory666For example, [Gek03], [AW15], [DKS], [AAG+19], etc. A reader not interested in this type of a calculation can safely skip this section.) to still try to use the coefficients of the characteristic polynomial to define the maps from 𝔤\mathfrak{g} and GG to 𝔸G\mathbb{A}_{G}. This works (with further caution discussed below) for the groups of type AnA_{n}, CnC_{n} and BnB_{n}, but does not quite work for type DnD_{n}.

First, consider Chevalley map on the Lie algebra.

If 𝔤=𝔰𝔭2n\mathfrak{g}=\mathfrak{sp}_{2n}, then the characteristic polynomial of any element X𝔤X\in\mathfrak{g} has the form: fX(t)=t2n+a1t2n2++anf_{X}(t)=t^{2n}+a_{1}t^{2n-2}+\dots+a_{n}. We can define 𝔠𝔰𝔭(X)\mathfrak{c}^{\prime}_{\mathfrak{sp}}(X) to be the tuple of coefficients (a1,,an)𝔸n(a_{1},\dots,a_{n})\in\mathbb{A}^{n}. The relationship between this map and Chevalley map is determined by the relation between the fundamental representation ρi\rho_{i} and the ii-th exterior power of the standard representation iρ1\wedge^{i}\rho_{1}, for i=1,,ni=1,\dots,n. For the symplectic Lie algebra, it turns out that iρ1\wedge^{i}\rho_{1} is a direct sum of ρi\rho_{i} and some representations of lower highest weights (see e.g. [FH91, Theorem 17.5]). Hence, the transition matrix between the characters of ρi\rho_{i} and the characters of iρ1\wedge^{i}\rho_{1} (i.e., the coefficients of the characteristic polynomial up to sign) is upper-triangular with 11s on the diagonal. Therefore, we get a measure-preserving isomorphism between the affine space 𝔸G\mathbb{A}_{G} with coordinates Tr(ρi)\operatorname{Tr}(\rho_{i}) and the affine space 𝔸G\mathbb{A}_{G}^{\prime} with the coordinates given by the coefficients of the characteristic polynomial. This implies that the map 𝔠𝔰𝔭\mathfrak{c}^{\prime}_{\mathfrak{sp}} could be used instead of 𝔠𝔰𝔭\mathfrak{c}_{\mathfrak{sp}} in all the calculations, without affecting the results.

For the odd orthogonal Lie algebras 𝔰𝔬2n+1\mathfrak{so}_{2n+1}, the exterior powers iρ1\wedge^{i}\rho_{1} are irreducible for 1=1,,n1=1,\dots,n, and for i=1,,n1i=1,\dots,n-1, coincide with the first n1n-1 fundamental representations; however, the last fundamental representation, the spin representation is not obtained this way (see [FH91, Theorem 19.14]). For the even orthogonal Lie algebra 𝔰𝔬2n\mathfrak{so}_{2n}, the representations iρ1\wedge^{i}\rho_{1} are irreducible for 1=1,,n11=1,\dots,n-1, and for i=1,,n2i=1,\dots,n-2, coincide with the first n2n-2 fundamental representations, and nρ1\wedge^{n}\rho_{1} decomposes as a direct sum of two irreducible representations whose weights are double the fundamental weights (see [FH91, Theorem 19.2]). Nevertheless, for the odd orthogonal Lie algebras, the coefficients of the characteristic polynomial still distinguish the stable conjugacy classes of regular semisimple elements; for the even orthogonal Lie algebras, one needs to add the pfaffian.

Passing to Steinberg map and algebraic groups: for the symplectic group, the coefficients of the characteristic polynomial can still be used without affecting any of the measure conversions, since this group is simply connected, and an argument similar to the above argument on the Lie algebra applies. For special orthogonal groups, 𝔸G\mathbb{A}_{G} is not an affine space since 𝐆\mathbf{G} is not simply connected; the coefficients of the characteristic polynomial give a map to an affine space. It seems to be a worthwhile exercise to work out the relationship between these two spaces and their measures; I have not done this calculation.

Finally, we discuss the cases 𝐆=GLn\mathbf{G}=\operatorname{GL}_{n} and 𝐆=GSp2n\mathbf{G}=\operatorname{GSp}_{2n} in some more detail since the latter calculation is needed in [AAG+19]. For GLn\operatorname{GL}_{n}, we just map gg to the coefficients of its characteristic polynomial. For GSp2n\operatorname{GSp}_{2n}, we can define 𝔠char.p.(g)=(a1,,an,ν(g))\mathfrak{c}^{\operatorname{char.p.}}(g)=(a_{1},\dots,a_{n},\nu(g)), where a1,ana_{1},\dots a_{n} are the first nn non-trivial coefficients of the characteristic polynomial, and ν(g)\nu(g) is the multiplier (this is ad hoc; one could have used the determinant instead to be consistent with GLn\operatorname{GL}_{n}); the superscript ‘char.p.{\operatorname{char.p.}}’ is to remind us that we are using the characteristic polynomial and distinguish this map from the standard map 𝔠\mathfrak{c}. The codomain of the map 𝔠char.p.\mathfrak{c}^{\operatorname{char.p.}} is the space we call 𝔸Gchar.p.\mathbb{A}_{G}^{\operatorname{char.p.}} which is 𝔸n1×𝔾m\mathbb{A}^{n-1}\times\mathbb{G}_{m} if 𝐆=GLn\mathbf{G}=\operatorname{GL}_{n}, and 𝔸n×𝔾m\mathbb{A}^{n}\times\mathbb{G}_{m} if 𝔾=GSp2n\mathbb{G}=\operatorname{GSp}_{2n}. The restriction of 𝔠char.p.\mathfrak{c}^{\operatorname{char.p.}} to 𝐆der\mathbf{G}^{\operatorname{der}} (if we forget the 𝔾m\mathbb{G}_{m}-component) coincides with the map 𝔠\mathfrak{c}^{\prime} constructed above for 𝐆der\mathbf{G}^{\operatorname{der}} (which coincides with 𝔠\mathfrak{c} if 𝐆der=SLn\mathbf{G}^{\operatorname{der}}=\operatorname{SL}_{n}).

The measure on the base 𝔸Gchar.p.\mathbb{A}_{G}^{\operatorname{char.p.}} in this case should be defined as the product of the measures associated with the form dada on the affine space, and ds/sds/s on 𝔾m\mathbb{G}_{m}, where we denote the coordinates on 𝔸k×𝔾m\mathbb{A}^{k}\times\mathbb{G}_{m} by (a,s)(a,s) (with k=n1k=n-1 or k=nk=n). With this definition, the resulting measure is, essentially, independent of the specific map used for the last coordinate (for example, in the case of GSp2n\operatorname{GSp}_{2n}, if the determinant instead of the multiplier were mapped to ss, the measure would just change by the factor |n|F|n|_{F}, which is 11 unless the residue characteristic of FF divides nn; but this caveat is the reason we prefer to work with the multiplier).

Let ωachar.p.\omega_{a}^{\operatorname{char.p.}} be the form on the fibre (𝔠char.p.)1(a)(\mathfrak{c}^{\operatorname{char.p.}})^{-1}(a) defined the same way as the form ωageom\omega_{a}^{{\operatorname{geom}}} in (23) and §3.5, but using the map 𝔠char.p.\mathfrak{c}^{\operatorname{char.p.}} instead:

(32) ωachar.p.(dadss)=ωG.\omega_{a}^{\operatorname{char.p.}}\wedge(da\wedge\frac{ds}{s})=\omega_{G}.

As before, the forms ωG\omega_{G}, ωT\omega_{T} and ωG/T\omega_{G/T} are, of course, invariant under the action of the centre of GG, but the centre does not even act on 𝔸Gchar.p.\mathbb{A}_{G}^{\operatorname{char.p.}} as a group. Nevertheless, multiplication by scalars still makes sense on this space.

To find the relation between the differential forms ωachar.p.\omega_{a}^{\operatorname{char.p.}} and ωT\G\omega_{T\backslash G} on a given orbit, let us work over the algebraic closure of FF for a moment. Over F¯\bar{F}, every element γ𝐆(F¯)\gamma\in\mathbf{G}(\bar{F}) can be written as ηzγ\eta_{z}\gamma^{\prime} with ηz𝐙(F¯)\eta_{z}\in\mathbf{Z}(\bar{F}) and γ𝐆der(F¯)\gamma^{\prime}\in\mathbf{G}^{\operatorname{der}}(\bar{F}) (defined up to an element of 𝐀(F¯){\mathbf{A}}(\bar{F}); we just pick one such pair). For ηz:=zIdZ\eta_{z}:=z{Id}\in Z with zF¯z\in\bar{F}, and g𝐆der(F¯)g\in\mathbf{G}^{\operatorname{der}}(\bar{F}), each coefficient ai(ηzg)a_{i}(\eta_{z}g) of the characteristic polynomial of ηzg\eta_{z}g differs from ai(g)a_{i}(g) by a power of zz. Then the form ωachar.p.\omega_{a}^{\operatorname{char.p.}} would have to scale by the power of zz as well, to preserve (32). We denote by Oγchar.p.O_{\gamma}^{\operatorname{char.p.}} the orbital integral of γ\gamma with respect to the form ω𝔠char.p.(γ)char.p.\omega_{\mathfrak{c}^{\operatorname{char.p.}}(\gamma)}^{\operatorname{char.p.}}, as a distribution on Cc(G)C_{c}^{\infty}(G). We compute explicitly the relation between this integral and the integral with respect to ωT\G\omega_{T\backslash G} for GSp2n\operatorname{GSp}_{2n}.

Example 3.17.

𝐆=GSp2n\mathbf{G}=\operatorname{GSp}_{2n}. In this case the scaling factor is |z|S|z|^{S}, where SS is the sum of the degrees (as homogeneous polynomials in the roots) of the first nn coefficients of the characteristic polynomial, i.e., 1+2++n=n(n+1)21+2+\dots+n=\frac{n(n+1)}{2}. If γ=ηzγ\gamma=\eta_{z}\gamma^{\prime} with det(γ)=1\det(\gamma^{\prime})=1 (and ηz𝐙(F¯)\eta_{z}\in\mathbf{Z}(\bar{F})), then |z|=|det(γ)|1/2n|z|=|\det(\gamma)|^{1/2n}. We obtain, for γGSp2n(F)rss\gamma\in\operatorname{GSp}_{2n}(F)^{\operatorname{rss}}:

Oγchar.p.(f)=|D(γ)|1/2|det(γ)|(n+1)4G/Tf(g1γg)ωG/T.O^{\operatorname{char.p.}}_{\gamma}(f)=|D(\gamma)|^{1/2}|\det(\gamma)|^{-\frac{(n+1)}{4}}\int_{G/T}f(g^{-1}\gamma g)\omega_{G/T}.

3.8. Summary

To summarize, so far the following choices have been made (we use the same notation as in [FLN10] whenever possible):

  • The measure |dx||dx| on FF, such that the volume of 𝒪F\mathcal{O}_{F} is 11. If we are working over a global field KK, and F=KvF=K_{v} is its completion at a finite place vv, this measure differs from [FLN10] for a finite number of places vv. For orbital integrals, this discrepancy gives rise to the factor |ΔK/|vdim(G)Rank(G)|\Delta_{K/\mathbb{Q}}|_{v}^{\dim(G)-\operatorname{Rank}(G)} (independent of the element) at each place vv.

  • An invariant differential form ωG\omega_{G} on 𝐆\mathbf{G} – appears on the both sides and does not affect the ratio between measures.

  • For an algebraic torus 𝐓\mathbf{T}, a choice of {χ1,,χn}\{\chi_{1},\dots,\chi_{n}\} - a \mathbb{Z}-basis of X(𝐓)X^{\ast}(\mathbf{T}). This choice does not affect anything.

Given γG\gamma\in G - a regular semisimple element, with T=CG(γ)T=C_{G}(\gamma), the following differential forms and measures have been constructed from these choices:

  • ωT=dχ1dχr\omega_{T}=d\chi_{1}\wedge\dots\wedge d\chi_{r};

  • ωT\G\omega_{T\backslash G} (which should be thought of as a measure on the orbit of γ\gamma, with T=CG(γ)T=C_{G}(\gamma)) satisfying ωG=ωTωT\G\omega_{G}=\omega_{T}\wedge\omega_{T\backslash G}. (Note that the centralizers of stably conjugate elements are isomorphic as algebraic tori over FF, so one can also think of ωT\G\omega_{T\backslash G} as a form on the stable orbit of γ\gamma.)

  • ωa\omega_{a}, also on the stable orbit of γ\gamma, with a=𝔠(γ)a=\mathfrak{c}(\gamma), satisfying ω𝔸Gωa=ωG\omega_{\mathbb{A}_{G}}\wedge\omega_{a}=\omega_{G}.

  • In [FLN10], there is a renormalized measure dg¯v:=L(1,σG)L(1,σT)ωT\Gd\bar{g}_{v}:=\frac{L(1,\sigma_{G})}{L(1,\sigma_{T})}\omega_{T\backslash G}.

Recall the notation D(γ)D(\gamma) for the Weyl discriminant of γGrss\gamma\in G^{\operatorname{rss}}, |Δ(γ)|=|D(γ)|1/2|\Delta(\gamma)|=|D(\gamma)|^{1/2}, as well as the definition of Artin LL-factor, (15). The following relations between these measures have been established:

  • For γGrss\gamma\in G^{\operatorname{rss}}, and a=𝔠(γ)a=\mathfrak{c}(\gamma), ωa=|Δ(γ)|ωT\G\omega_{a}=|\Delta(\gamma)|\omega_{T\backslash G}.

  • Consequently, for the measure dg¯vd\bar{g}_{v} defined in [FLN10, §3.4 below (3.17)] we have: ωa=|Δ(γ)|L(1,σG/T)dg¯v\omega_{a}=|\Delta(\gamma)|{L(1,\sigma_{G/T})}d\bar{g}_{v}, where the representation σT\G\sigma_{T\backslash G} of the Galois group is, by definition, the quotient of the representation σT\sigma_{T} on X(𝐓)X^{\ast}(\mathbf{T}) by the subrepresentation σG\sigma_{G} on the characters of GG 777 the rank of this subrepresentation is the same as the rank of the centre of GG; if GG is a semisimple group, we have σT\G=σT\sigma_{T\backslash G}=\sigma_{T}., and hence (L(1,σG)L(1,σT))1\left(\frac{L(1,\sigma_{G})}{L(1,\sigma_{T})}\right)^{-1} is precisely L(1,σT\G)L(1,\sigma_{T\backslash G}).

This establishes the identity (3.31) in [FLN10] (we note that Orb(f)\mathrm{Orb}(f) is defined in loc.cit. as the integral with respect to the measure dg¯vd\bar{g}_{v} on the orbit). Now we move on to the discussion of the factor L(1,σT\G)L(1,\sigma_{T\backslash G}) and the relationship with the canonical measures in the sense of Gross.

4. Orbital integrals: from differential forms to ‘canonical measures’

So far, we have worked with measures coming from differential forms, as summarized above. However, in many parts of the literature the so-called canonical measures are used. They are defined by means of defining a canonical subgroup, and then normalizing the measure so that the volume of this subgroup is 11. This introduces the following factors:

  • By definition of the canonical measure, for a torus 𝐓\mathbf{T},

    μTcan=1volωcan(𝒯0)|ωcan|,\mu_{T}^{{\operatorname{can}}}=\frac{1}{\operatorname{vol}_{\omega^{{\operatorname{can}}}}(\mathcal{T}^{0})}|\omega^{{\operatorname{can}}}|,

    where ωcan\omega^{{\operatorname{can}}} is the so-called canonical invariant volume form (discussed briefly in §2.2.2 above; the details of the definition are not important here).

    By Theorem 2.9 above,

    volωcan(𝒯0)=L(1,σT)1.\operatorname{vol}_{\omega^{{\operatorname{can}}}}(\mathcal{T}^{0})=L(1,\sigma_{T})^{-1}.

    Hence, on 𝐓\mathbf{T}, we have

    (33) ωcan=volωcan(𝒯0)volωT(𝒯0)ωT.\omega^{{\operatorname{can}}}=\frac{\operatorname{vol}_{\omega^{{\operatorname{can}}}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}\omega_{T}.
  • Recall that ωG=ωGcan\omega_{G}=\omega_{G}^{{\operatorname{can}}} since we are assuming 𝐆\mathbf{G} is split; this is also true more generally for 𝐆\mathbf{G} unramified, (and in any case, the choice of the form on 𝐆\mathbf{G} matters much less than the choice of the form on 𝐓\mathbf{T}, as discussed above). Therefore, on the orbit of γ\gamma, we have:

    ωT\Gcan=volωT(𝒯0)volωcan(𝒯0)ωT\G,\omega^{{\operatorname{can}}}_{T\backslash G}=\frac{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega^{{\operatorname{can}}}}(\mathcal{T}^{0})}\omega_{T\backslash G},

    and

    (34) μT\Gcan=μGcanμTcan=1volωG(G0)|ωG|1volωcan(𝒯0)|ωcan|=volωcan(𝒯0)volωG(G0)|ωG||ωcan|\displaystyle\mu^{{\operatorname{can}}}_{T\backslash G}=\frac{\mu_{G}^{\operatorname{can}}}{\mu_{T}^{\operatorname{can}}}=\frac{{\frac{1}{\operatorname{vol}_{\omega_{G}}(G^{0})}}{|\omega_{G}|}}{{\frac{1}{\operatorname{vol}_{\omega^{\operatorname{can}}}(\mathcal{T}^{0})}}{|\omega^{\operatorname{can}}|}}=\frac{\operatorname{vol}_{\omega^{\operatorname{can}}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{G}}(G^{0})}\frac{|\omega_{G}|}{|\omega^{\operatorname{can}}|}
    =volωcan(𝒯0)volωG(G0)|ωG|volωcan(𝒯0)volωT(𝒯0)|ωT|=volωT(𝒯0)volωG(G0)|ωG||ωT|=volωT(𝒯0)volωG(G0)|ωT\G|.\displaystyle=\frac{\operatorname{vol}_{\omega^{\operatorname{can}}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{G}}(G^{0})}\frac{|\omega_{G}|}{\frac{\operatorname{vol}_{\omega^{\operatorname{can}}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}|\omega_{T}|}=\frac{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{G}}(G^{0})}\frac{|\omega_{G}|}{|\omega_{T}|}=\frac{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}{\operatorname{vol}_{\omega_{G}}(G^{0})}|\omega_{T\backslash G}|.

    When 𝐓\mathbf{T} splits over an unramified extension, by Theorem 2.9 above, volωT(𝒯0)=L(1,σT)1\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})=L(1,\sigma_{T})^{-1}. Thus at almost all places vv, the measure dg¯vd\bar{g}_{v} is related to the canonical measure on the orbit by:

    (35) dg¯v=L(1,σG)volωG(G0)μT\Gcan.d\bar{g}_{v}=L(1,\sigma_{G})\operatorname{vol}_{\omega_{G}}(G^{0})\mu^{{\operatorname{can}}}_{T\backslash G}.

    Combining (34) with the relation ωa=±Δ(γ)ωT\G\omega_{a}=\pm\Delta(\gamma)\omega_{T\backslash G}, we also obtain the relation between the geometric measure and canonical measure (for all places):

    (36) |ωa|=|Δ(γ)|1volωT(𝒯0)μT\Gcan,|\omega_{a}|=|\Delta(\gamma)|\frac{1}{\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})}\mu^{{\operatorname{can}}}_{T\backslash G},

    where the factor volωG(G0)\operatorname{vol}_{\omega_{G}}(G^{0}) disappears since the same measure on 𝐆\mathbf{G} needs to be used on both sides when defining ωa\omega_{a}. We recall that the factor volωT(𝒯0){\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0})} is discussed above in §2.2.3.

4.1. Example: GL2\operatorname{GL}_{2}

For 𝐆=GL2\mathbf{G}=\operatorname{GL}_{2}, we can make everything completely explicit. The orbital integrals of spherical functions with respect to canonical measure are computed, for example, in [Kot05, §5]. We combine this computation with our conversion factors to obtain the integrals with respect to the geometric measure. We observe that the result agrees with the formula (3.6) of [Lan13].

In [Kot05, §5] the orbital integrals are computed using the reduced building (i.e. the tree) for GL2\operatorname{GL}_{2}, and expressed in terms of the integer dγd_{\gamma} (for γG(F)rss\gamma\in G(F)^{\operatorname{rss}}). The number dγd_{\gamma} is defined in terms of the valuations of the eigenvalues of γ\gamma, see the top of p.415 for the split case, p.417 for the unramified case, and (5.9.9) for the ramified case.

In fact, we have

|D(γ)|={q2dγ,γ split or unramified,q2dγ1,γ ramified.|D(\gamma)|=\begin{cases}q^{-2d_{\gamma}},\quad\gamma\text{ split or unramified},\\ q^{-2d_{\gamma}-1},\quad\gamma\text{ ramified}.\end{cases}

(This is the definition in the split and unramified cases and an easy exercise in the ramified case.)

Here we only look at the simplest orbital integral of f0f_{0}, the characteristic function of the maximal compact subgroup G0=GL2(𝒪F)G_{0}=\operatorname{GL}_{2}(\mathcal{O}_{F}).

  • If γ\gamma is split over FF, then, from formula (5.8.4) loc.cit.:

    (37) Oγ(f0)=qdγ=|D(γ)|1/2O_{\gamma}(f_{0})=q^{d_{\gamma}}=|D(\gamma)|^{-1/2}
  • If γ\gamma is elliptic (which in GL2rss\operatorname{GL}_{2}^{\operatorname{rss}}, is the same as not split), then Oγ(f0)=|Vγ|O_{\gamma}(f_{0})=|V^{\gamma}|, the cardinality of the set of fixed points of the action of γ\gamma on the building; see formula (5.9.3). Note that here the right-hand side does not depend on the choice of the measures on GG and on the centralizer of γ\gamma (which we denote by T=Gγ=CG(γ)T=G_{\gamma}=C_{G}(\gamma) to consolidate notation with this part of [Kot05]). Thus, there is a unique choice of measures for which this equality is true. This equality is explained in §3.4 of loc.cit.; see also the explanation just above (5.9.1). In fact, for elliptic γ\gamma, one has

    vol(Z\Gγ)Oγ(f0)=|Vγ|,\operatorname{vol}(Z\backslash G_{\gamma})O_{\gamma}(f_{0})=|V^{\gamma}|,

    where on the left the volume and the orbital integral are taken with respect to the same choice of the measure on GγG_{\gamma}, and the measure on GG that gives G0G_{0} volume 11 (note that in this formula both sides are independent of the choice of the measure on GγG_{\gamma}). Thus the measure on GγG_{\gamma} that makes (5.9.3) work is precisely the measure such that vol(Z\Gγ)=1\operatorname{vol}(Z\backslash G_{\gamma})=1.

    Plugging in the calculations of |Vγ||V^{\gamma}| from loc.cit., in the two remaining cases we obtain:

  • If γ\gamma is unramified (5.9.7):

    (38) Oγ(f0)=1+(q+1)qdγ1q1.O_{\gamma}(f_{0})=1+(q+1)\frac{q^{d_{\gamma}}-1}{q-1}.
  • If γ\gamma is ramified (5.9.10):

    (39) Oγ(f0)=2qdγ+11q1.O_{\gamma}(f_{0})=2\frac{q^{d_{\gamma}+1}-1}{q-1}.

Assume as usual that p2p\neq 2. Suppose we started with the measure on GγG_{\gamma} that gave volume 11 to its maximal compact subgroup, and the measure on ZZ such that the volume of ZG0Z\cap G_{0} is 11. In the unramified case, the map from TcT^{c} to Z\TZ\backslash T is surjective, and this choice of measures gives the quotient Z\TZ\backslash T volume 11. In the ramified case, the image of TcT^{c} in Z\TZ\backslash T has index 22, and thus the volume of Z\TZ\backslash T we get from this natural measure on TT is not 11 but 22. The upshot is that in the ramified case, the measure giving the volume 11 to Z\GγZ\backslash G_{\gamma} does not come from a natural measure on GγG_{\gamma}.

Thus, combining the relation (36) with (37), (38), and (39), respectively, and using (5), which applies since for all tori in GL2\operatorname{GL}_{2}, Tc=𝒯0T^{c}=\mathcal{T}^{0}, we obtain the integrals with respect to the geometric measure:

(40) Oγgeom(f0)=|D(γ)|1/2volωT(Tc)Oγ(f0)\displaystyle O_{\gamma}^{\operatorname{geom}}(f_{0})=\frac{|D(\gamma)|^{1/2}}{\operatorname{vol}_{\omega_{T}}(T^{c})}O_{\gamma}(f_{0})
={(11q)2γ is hyperbolicq2(q21)(1+(q+1)qdγ1q1)qdγγ is unramified ellipticqqq1(qdγ+11q1)qdγ1/2γ is ramified elliptic\displaystyle=\begin{cases}&(1-\frac{1}{q})^{-2}\quad\gamma\text{ is hyperbolic}\\ &\frac{q^{2}}{(q^{2}-1)}\left(1+(q+1)\frac{q^{d_{\gamma}}-1}{q-1}\right)q^{-d_{\gamma}}\quad\gamma\text{ is unramified elliptic}\\ &\frac{q\sqrt{q}}{q-1}\left(\frac{q^{d_{\gamma}+1}-1}{q-1}\right)q^{-d_{\gamma}-1/2}\quad\gamma\text{ is ramified elliptic}\end{cases}
=(11q)2{1γ is hyperbolic12q+1qdγγ is unramified elliptic1q(dγ+1)γ is ramified elliptic\displaystyle=\left(1-\frac{1}{q}\right)^{-2}\begin{cases}&1\quad\gamma\text{ is hyperbolic}\\ &1-\frac{2}{q+1}q^{-d_{\gamma}}\quad\gamma\text{ is unramified elliptic}\\ &1-q^{-(d_{\gamma}+1)}\quad\gamma\text{ is ramified elliptic}\end{cases}

These formulas agree with [Lan13, (2.2.10)], if one multiplies our results by volωG(G0)=(11q)2(1+1q)\operatorname{vol}_{\omega_{G}}(G_{0})=(1-\frac{1}{q})^{2}(1+\frac{1}{q}), as required by the relation (35).

4.2. The next step

In [Lan13], Langlands works out Poisson summation on the geometric side of the stable Trace Formula for SL2\operatorname{SL}_{2}. Roughly speaking, Poisson summation formula is a relation between the sum of the values of a smooth function over a lattice in a vector space, and the sum of the values of its Fourier transform over a dual lattice. Here the space is the set of adèlic points of the Steinberg-Hitchin base for SL2\operatorname{SL}_{2}, which is just the affine line. The lattice in it is the image of the diagonal embedding of the base field (we can take \mathbb{Q} for simplicity). The geometric side of the Trace Formula contains a sum over the conjugacy classes of elliptic elements γSL2()\gamma\in\operatorname{SL}_{2}(\mathbb{Q}), which corresponds to a sum over \mathbb{Q} in the Steinberg-Hitchin base. Thus at least for the elliptic part, it is tempting to take the function to be a stable orbital integral (i.e., the integral of some fixed test function over a fibre of Steinberg map 𝔠1(a)\mathfrak{c}^{-1}(a), as a function of aa), and apply Poisson summation. However, for that the function needs to satisfy some smoothness assumption. Now we can at least make some preliminary remarks about how far our function is from being smooth, at least at every finite place.

If we take γSL2(p)\gamma\in\operatorname{SL}_{2}(\mathbb{Q}_{p}), the GL2(p)\operatorname{GL}_{2}(\mathbb{Q}_{p})-orbital integral computed above is the stable orbital integral of γ\gamma. All along we have been assuming that γ\gamma is a regular semisimple element. It is well-known that the singularities of orbital integrals occur only at non-regular elements (and we will see this explicitly in a moment, in this example). More precisely, it is a result of Harish-Chandra that for a given test function ff, when a measure of the form |ωG/T||\omega_{G/T}| is used on each regular semisimple orbit, the orbital integral γOγ(f)\gamma\mapsto O_{\gamma}(f) is a smooth (i.e., locally constant) function on the open set GrssG^{\operatorname{rss}} of regular semisimple elements. This function is not bounded as γ\gamma approaches a non-regular element; however, its growth is controlled by |D(γ)|1/2|D(\gamma)|^{-1/2}. Specifically, Harish-Chandra proved that (still with ff fixed), the so-called normalized orbital integral, namely, the function γ|D(γ)|1/2Oγ(f)\gamma\mapsto|D(\gamma)|^{1/2}O_{\gamma}(f) is bounded on compact subsets of GG, and locally integrable on GG. We note that since D(γ)D(\gamma) vanishes at non-regular elements, this normalized orbital integral is also zero off the regular set. Thus, the normalized orbital integral, as a function of γ\gamma (for a given test function ff), is locally constant on GrssG^{\operatorname{rss}}, zero on non-regular semisimple elements, and locally bounded on GG. However, this does not imply that it is continuous on GG. Indeed, while it is locally constant on the set of regular semisimple elements, as γ\gamma approaches a non-regular element, the neighbourhoods of constancy get smaller; at a non-regular element γ0\gamma_{0} itself this function is zero since D(γ0)=0D(\gamma_{0})=0; by Harish-Chandra’s theorem this function is bounded on any compact neighbourhood of γ0\gamma_{0}; however, nothing says that it is continuous at γ0\gamma_{0}: without a careful choice of measures, it will have “jumps”. As we shall see in our SL2\operatorname{SL}_{2}-example, the extension of the normalized orbital integral by zero to non-regular elements does not actually give a continuous function on GG; however, when the geometric measure is chosen, one gets a function on the Steinberg-Hitchin base with just a removable discontinuity.

In SL2\operatorname{SL}_{2}, we have just two non-regular semisimple elements, namely, ±Id\pm{\mathrm{Id}}. Their images under Steinberg map 𝔠SL2:SL2𝔸1\mathfrak{c}_{SL_{2}}:\operatorname{SL}_{2}\to\mathbb{A}^{1} are ±2\pm 2. Fix pp (for now, p2p\neq 2) and consider, for example, a neighbourhood of the point 2𝔸SL2(p)=𝔸1(p)2\in\mathbb{A}_{\operatorname{SL}_{2}}(\mathbb{Q}_{p})=\mathbb{A}^{1}(\mathbb{Q}_{p}). It consists of the images of split, ramified, and unramified elements with sufficiently large dγd_{\gamma} (the split/ramified/unramifed is determined by the discriminant of the characteristic polynomial of a given element, as discussed above in Example 3.3). The formula (40) shows that as dγd_{\gamma}\to\infty (i.e, as γ\gamma approaches ±Id\pm\mathrm{Id}), the stable orbital integral of f0f_{0} on the orbit of γ\gamma with respect to the geometric measure gives a continuous function on 𝔸1\mathbb{A}^{1}, with value (1q1)2(1-q^{-1})^{-2} at a=2𝔸SL2(p)a=2\in\mathbb{A}_{\operatorname{SL}_{2}}(\mathbb{Q}_{p}). (This, of course, cannot be said about the orbital integrals with respect to the canonical measure, as they get large - of the size qdγq^{d_{\gamma}}; as |D(γ)|1/2qdγ|D(\gamma)|^{1/2}\asymp q^{-d_{\gamma}}, we see the confirmation of Harish-Chandra’s boundedness result; but still the function γ|D(γ)|1/2Ocan(γ)\gamma\mapsto|D(\gamma)|^{1/2}O^{\operatorname{can}}(\gamma) has “jumps” at ±Id\pm Id; once we make the adjustments by the volumes of the maximal compact subgroups of the corresponding tori, it becomes continuous). This continuity result is one of the insights of [Lan13]. However, as we see explicitly from (40), this function is continuous but not smooth (i.e. not constant on any neighbourhood of a=±2a=\pm 2); and so far this is just the description of the situation at a single place, whereas ultimately we will need a global Poisson summation formula. This causes some of the technical difficulties discussed in Altug’s lectures.

5. Global volumes

5.1. The analytic class number formula for an imaginary quadratic field

Here we recast the analytic class number formula for an imaginary quadratic field KK as a volume computation, using the above ideas. It was observed by Ono, [Ono63] (see also [Shy77]), that the analytic class number formula in this case amounts to the fact that the Tamagawa number τ(𝐓)\tau(\mathbf{T}) of the torus 𝐓=ResK/𝔾m\mathbf{T}=\operatorname{Res}_{K/\mathbb{Q}}\mathbb{G}_{m} equals 11. We will assume that τ(𝐓)=1\tau(\mathbf{T})=1 and derive the analytic class number formula for KK from this fact. This also serves as preparation for §5.3 where the same volume term combines with an orbital integral for an interesting result.

The analytic class number formula for a general number field is the relation

(41) lims1+(s1)ζK(s)=2r+tπtRKhKwK|ΔK|1/2,\lim_{s\to 1^{+}}(s-1)\zeta_{K}(s)=\frac{2^{r+t}\pi^{t}R_{K}h_{K}}{w_{K}|\Delta_{K}|^{1/2}},

where: ζK\zeta_{K} is the Dedekind zeta-function of KK, RKR_{K} is the regulator (we will not need it in this note so we skip the definition), ΔK\Delta_{K} is the discriminant of KK, hKh_{K} is the class number, wKw_{K} is the number of roots of 11 in KK, rr is the number of real embeddings, and 2t2t is the number of complex embeddings of KK.

Let us consider an imaginary quadratic field K=(D)K=\mathbb{Q}(\sqrt{-D}) (with D>0D>0); denote its ring of integers by 𝒪\mathcal{O}. In this case, we have r=0r=0, t=1t=1, the regulator RKR_{K} is automatically equal to 11, and the left-hand side equals the value at s=1s=1 of ζK(s)ζ(s)=L(1,χK)\frac{\zeta_{K}(s)}{\zeta(s)}=L(1,\chi_{K}) – the value (in the sense of a conditionally convergent product) of the Dirichlet LL-function L(1,χK)L(1,\chi_{K}) 888This equality is the simplest case of the correspondence between Artin and Hecke LL-functions.. Here ζ(s)\zeta(s) is the Riemann zeta-function, and χK\chi_{K} is the Dirichlet character associated with KK :

(42) χK(p)={1p splits in K1p is inert in K0p ramifies in K.\chi_{K}(p)=\begin{cases}1&\quad p\text{ splits in }K\\ -1&\quad p\text{ is inert in }K\\ 0&\quad p\text{ ramifies in }K\end{cases}.

Thus, for an imaginary quadratic field KK the analytic class number formula reduces to:

(43) L(1,χK)=2πhKwKΔK.L(1,\chi_{K})=\frac{2\pi h_{K}}{w_{K}\sqrt{\Delta_{K}}}.

Our goal is to prove this relation by using only the known facts about algebraic groups and the measure conversions discussed above. The algebraic group in question here is just the torus 𝐓=ResK/𝔾m\mathbf{T}=\operatorname{Res}_{K/\mathbb{Q}}\mathbb{G}_{m}.

Let 𝔸K\mathbb{A}_{K} be the adèles999There is an unfortunate clash of standard notation: we used 𝔸G\mathbb{A}_{G} to denote Steinberg quotient of 𝐆\mathbf{G}; hopefully this causes no confusion. Another notation clash is μK\mu_{K} for the group of roots of 11 in KK, as we have been using the letter μ\mu (with subscripts and superscripts) to denote various measures. of KK and let 𝔸Kfin\mathbb{A}_{K}^{\operatorname{fin}} be the finite adèles. In general K×K^{\times} embeds (diagonally) in 𝔸K×\mathbb{A}_{K}^{\times} with discrete image; for KK imaginary quadratic, the image of the embedding K×(𝔸Kfin)×K^{\times}\hookrightarrow(\mathbb{A}_{K}^{\operatorname{fin}})^{\times} is still discrete (in fact, this is true only when K=K=\mathbb{Q} or is imaginary quadratic, see e.g.,[Mil08]). We know (weak approximation, see e.g., [PR91]) that for \mathbb{Q},

×\(𝔸fin)×/pp×={1}.\mathbb{Q}^{\times}\big{\backslash}(\mathbb{A}_{\mathbb{Q}}^{\operatorname{fin}})^{\times}\big{/}\prod_{p}\mathbb{Z}_{p}^{\times}=\{1\}.

Since the image of K×K^{\times} in (𝔸Kfin)×(\mathbb{A}_{K}^{\operatorname{fin}})^{\times} is discrete, we can define a similar double quotient for 𝔸Kfin\mathbb{A}_{K}^{\operatorname{fin}}: K×\(𝔸Kfin)×/v(𝒪v)×K^{\times}\big{\backslash}(\mathbb{A}_{K}^{\operatorname{fin}})^{\times}\big{/}\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times}, which, roughly speaking, should measure the size of the class group of KK. The reason this is not exactly the class group is the intersection of the image of K×K^{\times} with the compact subgroup v(𝒪v)×\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times}. More precisely, we have the exact sequence:

(44) 1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(K×v(𝒪v)×)\v(𝒪v)×\textstyle{(K^{\times}\cap\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times})\backslash\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K×\(𝔸Kfin)×\textstyle{K^{\times}\big{\backslash}(\mathbb{A}_{K}^{\operatorname{fin}})^{\times}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cl(K)\textstyle{\mathrm{Cl}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1,\textstyle{1,}

where Cl(K)\mathrm{Cl}(K) is the class group of KK.

The group K×v(𝒪v)×K^{\times}\cap\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times} is precisely the group μK\mu_{K} of roots of 11 in KK (the elements of K×K^{\times} that are units at every finite place).

The key point is that if we normalize the volume of the group of units 𝒪v×\mathcal{O}_{v}^{\times} to be 11 at every place, and call this measure νT\nu_{T} 101010An important coincidence that happens for our torus TT, because it is obtained from 𝔾m\mathbb{G}_{m} by restriction of scalars, is that the measure νT\nu_{T} coincides with the canonical measure μTcan\mu_{T}^{\operatorname{can}} at every finite place, as discussed above in the point (1) of §2.2.3. See [Shy77] for the general situation., then we get from the above exact sequence:

(45) volνT(K×\(𝔸Kfin)×/v(𝒪v)×)=hKwK.\operatorname{vol}_{\nu_{T}}\left(K^{\times}\big{\backslash}(\mathbb{A}_{K}^{\operatorname{fin}})^{\times}\big{/}\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times}\right)=\frac{h_{K}}{w_{K}}.

We will assume as fact that the Tamagawa number of 𝐓\mathbf{T} is 11 (this is so because 𝐓\mathbf{T} is obtained from 𝔾m\mathbb{G}_{m} by Weil restriction of scalars, as briefly discussed below). The analytic class number formula will follow as soon as we relate the volume on the left-hand side of (45) to the Tamagawa number τ(𝐓)\tau(\mathbf{T}).

5.1.1. Tamagawa measure

We briefly recall the definition of the Tamagawa measure, just for the special case of our torus 𝐓\mathbf{T}. We follow the definition of Ono, [Ono66], [Ono61], which has become standard. 111111We note that superficially, it differs from the definition that A. Weil uses in [Wei82], in the sense that Ono uses a specific set of convergence factors, and incorporates a global factor that makes his definition independent of the choice of the convergence factors at finitely many places. The resulting global measure, of course, is the same in the both sources.

1. Let (𝔸×)1(\mathbb{A}_{\mathbb{Q}}^{\times})^{1} be the set of norm-1 adèles (also referred to as special ideles):

(𝔸×)1:={(xv)𝔸:v|(xv)|v=1},(\mathbb{A}_{\mathbb{Q}}^{\times})^{1}:=\{(x_{v})\in\mathbb{A}_{\mathbb{Q}}:\prod_{v}|(x_{v})|_{v}=1\},

where the product is over all places of \mathbb{Q}.

We have the exact sequence

(46) 1(𝔸×)1𝔸×>0×1,1\to(\mathbb{A}_{\mathbb{Q}}^{\times})^{1}\to\mathbb{A}_{\mathbb{Q}}^{\times}\to\mathbb{R}_{>0}^{\times}\to 1,

where the first map is the inclusion and the second map is the product of absolute values over all places, x=(xv)v|xv|vx=(x_{v})\mapsto\prod_{v}|x_{v}|_{v}. Moreover, the exact sequence splits and we have a canonical decomposition

(47) 𝔸×(𝔸×)1×(×)0,\mathbb{A}_{\mathbb{Q}}^{\times}\simeq(\mathbb{A}_{\mathbb{Q}}^{\times})^{1}\times(\mathbb{R}^{\times})^{0},

as a direct product of topological groups, where (×)0(\mathbb{R}^{\times})^{0} stands for the connected component R>0×R_{>0}^{\times} (in the sense of the metric topology) of the group ×\mathbb{R}^{\times}. We note that the image of the diagonal embedding of ×\mathbb{Q}^{\times} into 𝔸×\mathbb{A}_{\mathbb{Q}}^{\times} is contained in (𝔸×)1(\mathbb{A}_{\mathbb{Q}}^{\times})^{1}, and it follows from (47) that the quotient ×\(𝔸×)1\mathbb{Q}^{\times}\backslash(\mathbb{A}_{\mathbb{Q}}^{\times})^{1} is compact.

2. To define the Tamagawa measure on 𝐓()\𝐓(𝔸)\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}), one needs to start with a volume form ω\omega on 𝐓\mathbf{T} defined over \mathbb{Q}. We note that even writing down such a form concretely is not trivial: the natural form ωT\omega_{T} defined in §2.2 is not defined over \mathbb{Q}. Fortunately, in our special case, the differential form ω:=1ΔKωT\omega:=\frac{1}{\sqrt{\Delta_{K}}}\omega_{T} is defined over \mathbb{Q}, see Example 2.8 (This easy case can also be verified directly by a calculation similar to that of Example 2.3). 121212See [GG99], Corollary 3.7, for a way to define such a form in general. In our special case of the quadratic field, it is an easy case of the discriminant-conductor formula that the Artin conductor of the motive constructed in [GG99] coincides with the discriminant ΔK\Delta_{K}, so our definition is a special case of the construction in [GG99].

Recall the local Artin LL-factors attached to the representation σT\sigma_{T} of Gal(K/)\operatorname{Gal}(K/\mathbb{Q}) on X(𝐓)X^{\ast}(\mathbf{T}), see (15), and let L(s,σT):=pLp(s,σT)L(s,\sigma_{T}):=\prod_{p}L_{p}(s,\sigma_{T}). Let rTr_{T} be the multiplicity of the trivial representation as a sub-representation of σT\sigma_{T}. In our case, σT\sigma_{T} is 22-dimensional, and rT=1r_{T}=1; a copy of the trivial representation in σT\sigma_{T} is generated by the norm character, which is stable under the action of the Galois group /2\mathbb{Z}/2\mathbb{Z}. Let

ρT:=lims1+(s1)rTL(s,σT).\rho_{T}:=\lim_{s\to 1+}(s-1)^{r_{T}}L(s,\sigma_{T}).

We see that in our case, ρT\rho_{T} coincides with the left-hand-side of (41). The measure μTama\mu^{Tama} on 𝐓(𝔸)\mathbf{T}(\mathbb{A}) is defined as:

(48) μTama:=ρT1|ω|pLp(1,σT)|ω|p,\mu^{Tama}:=\rho_{T}^{-1}|\omega_{\infty}|\prod_{p}L_{p}(1,\sigma_{T})|\omega|_{p},

where ω\omega_{\infty} is the form induced by ω\omega on 𝐓()\mathbf{T}(\mathbb{R}), in our case.

We make a note of some subtle features of this definition:

  1. (1)

    The definition does not depend on the choice of a volume form (as long as ω\omega is defined over \mathbb{Q}), since any two choices differ by a constant in \mathbb{Q}, which does not matter globally thanks to the product formula.

  2. (2)

    Without the convergence factors Lp(1,σT)L_{p}(1,\sigma_{T}), the product p|ω|p\prod_{p}|\omega|_{p} does not define a measure on 𝐓(𝔸)\mathbf{T}(\mathbb{A}), since (as one can easily see in our example) the maximal compact subgroup v(𝒪v)×\prod_{v\nmid\infty}(\mathcal{O}_{v})^{\times} of 𝐓(𝔸fin)\mathbf{T}(\mathbb{A}^{\operatorname{fin}}) would have infinite volume with respect to such a ‘measure’, since by (5), it contains the Euler product for the Riemann zeta function at 11. There is some choice involved in the definition of the convergence factors (for example, in Weil’s definition in [Wei82] the convergence factors in this case would be simply (11/p)(1-1/p), which would be sufficient to achieve convergence of the product measure). As Ono explains in §3.5 of [Ono61], if one modifies the individual convergence factors by any multipliers whose product converges, it does not affect the final result thanks to the global factor ρT1\rho_{T}^{-1}.

For future use, we define by μp\mu_{p} the measure Lp(1,σT)|ω|pL_{p}(1,\sigma_{T})|\omega|_{p} on 𝐓(p)\mathbf{T}(\mathbb{Q}_{p}).

3. The Tamagawa number of 𝐓\mathbf{T} is, by definition, the volume (with respect to the Tamagawa measure on the quotient, discussed below), of 𝐓()\𝐓(𝔸)1\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}, where

𝐓(𝔸)1={(xv)𝐓(𝔸):v|χ(xv)|v=1 for all χX(𝐓) that are defined over }.\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}=\{(x_{v})\in\mathbf{T}(\mathbb{A}_{\mathbb{Q}}):\prod_{v}|\chi(x_{v})|_{v}=1\text{ for all }\chi\in X^{\ast}(\mathbf{T})\text{ that are defined over }\mathbb{Q}\}.

The group of \mathbb{Q}-characters of 𝐓\mathbf{T} has rank 1, and is generated by the norm map. Thus,

𝐓(𝔸)1={(xv)𝐓(𝔸):v|NKv/v(xv)|v=1},\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}=\{(x_{v})\in\mathbf{T}(\mathbb{A}_{\mathbb{Q}}):\prod_{v}|N_{K_{v}/\mathbb{Q}_{v}}(x_{v})|_{v}=1\},

where the product is over the places of \mathbb{Q}. We note that 𝐓(𝔸)=𝔸K×\mathbf{T}(\mathbb{A}_{\mathbb{Q}})=\mathbb{A}_{K}^{\times}, and we have the exact sequence

(49) 1𝐓(𝔸)1𝔸K×>0×1,1\to\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}\to\mathbb{A}_{K}^{\times}\to\mathbb{R}_{>0}^{\times}\to 1,

where the map to >0×\mathbb{R}_{>0}^{\times} is the product of the local norm maps.

Let dmdm (using the notation and terminology of [Shy77]) be the measure on 𝐓(𝔸)1\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1} that ‘matches topologically’ in this exact sequence with the measure μTama\mu^{Tama} on 𝐓(𝔸)=𝔸K×\mathbf{T}(\mathbb{A}_{\mathbb{Q}})=\mathbb{A}_{K}^{\times} defined above, and the measure dtt\frac{dt}{t} on >0×\mathbb{R}_{>0}^{\times}. That is, dmdm is the measure on 𝐓(𝔸)1\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1} such that

μTama=dmdtt.\mu^{Tama}=dm\wedge\frac{dt}{t}.

Since 𝐓()=K×\mathbf{T}(\mathbb{Q})=K^{\times} is a discrete subgroup of 𝐓(𝔸)1\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}, the measure dmdm descends to the quotient by this subgroup, and the volume of 𝐓()\𝐓(𝔸)1\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1} with respect to this measure is, by definition, the Tamagawa number τ(𝐓)\tau(\mathbf{T}).

We note that the exact sequence (49) splits, and we have a group isomorphism

K×\𝔸K×𝐓()\𝐓(𝔸)1×>0×.K^{\times}\backslash\mathbb{A}_{K}^{\times}\simeq\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}\times\mathbb{R}_{>0}^{\times}.

5.1.2. The proof of the analytic class number formula

Now we are ready to go from (45) to the analytic class number formula for KK; we do it in two steps.

Step1. The finite places. Rewriting the relation (5) of Example 2.3 using the notation of this section (and noting that for p2p\neq 2, |ΔK|p=p1|\Delta_{K}|_{p}=p^{-1} if pp ramifies in KK and |ΔK|p=1|\Delta_{K}|_{p}=1 otherwise), we obtain, for v|pv|p:

(50) vol|ωT|(𝒪v×)=(11p)Lv(1,χK)1|ΔK|p1/2.\operatorname{vol}_{|\omega_{T}|}(\mathcal{O}_{v}^{\times})=\left(1-\frac{1}{p}\right)L_{v}(1,\chi_{K})^{-1}|\Delta_{K}|_{p}^{1/2}.

We will see below in §5.2 that this relation holds at p=2p=2 as well. Thus, volμp(𝒪v×)=1\operatorname{vol}_{\mu_{p}}(\mathcal{O}_{v}^{\times})=1, and we see explicitly in this example, that our measure μp\mu_{p} coincides with the pp-component of the measure νT\nu_{T} used in (45) (this is a very special case of [GG99], Corollary 7.3).

Step 2. The infinite places and putting it together. We also have the exact sequence (where S1×S^{1}\subset\mathbb{C}^{\times} is the unit circle)

(51) 1S1𝐓(𝔸)1𝔸Kfin1,1\to S^{1}\to\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}\to\mathbb{A}_{K}^{\operatorname{fin}}\to 1,

where the first map is the inclusion into 𝔸K\mathbb{A}_{K} that maps zS1z\in S^{1} to the adèle (1,z)(1,z) trivial at all the finite places, and the second map is the projection onto the finite adèles. Since the image of the diagonal embedding of K×K^{\times} intersects the image of S1S^{1} in 𝐓(𝔸)1\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1} trivially, (51) yields the exact sequence for the quotients by K×K^{\times}:

(52) 1S1K×\𝐓(𝔸)1K×\𝔸Kfin1.1\to S^{1}\to K^{\times}\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1}\to K^{\times}\backslash\mathbb{A}_{K}^{\operatorname{fin}}\to 1.

Now we need to carefully find the component at infinity dmdm_{\infty} (which is a measure on S1S^{1}) of the measure dmdm defined via the exact sequence (49). First, we choose a convenient basis for the character lattice of 𝐓\mathbf{T}, in view of this exact sequence: we use the characters χ1(z,z¯)=z\chi_{1}(z,\bar{z})=z and η(z):=zz¯=|z|2\eta(z):=z\bar{z}=|z|^{2}. These two characters, which can be thought of as the vectors (1,0)(1,0) and (1,1)(1,1) with respect to the ‘standard’ basis of X(𝐓)X^{\ast}(\mathbf{T}) in Example 2.3, still form a \mathbb{Z}-basis of X(𝐓)X^{\ast}(\mathbf{T}), and hence we can write ωT=dzzdηη\omega_{T}=\frac{dz}{z}\wedge\frac{d\eta}{\eta}. We will also use the coordinates (z,η)(z,\eta) on 𝐓\mathbf{T} for the rest of this calculation. We write every element of 𝐓(𝔸)\mathbf{T}(\mathbb{A}) as a=afaa=a_{f}a_{\infty}, where afa_{f} has the infinity component 11 and a=(1,(z,η))a_{\infty}=(1,(z,\eta)) has all the components at the finite places equal to 11. In this notation, 𝐓(𝔸)1\mathbf{T}(\mathbb{A})^{1} is defined by the condition |z|2=|η|=af1|z|^{2}=|\eta|=\|a_{f}\|^{-1}. We write

μTama=μfinTamaμTama,\mu^{Tama}=\mu_{{\operatorname{fin}}}^{Tama}\mu_{\infty}^{Tama},

where μfinTama=ρT1pμp\mu_{{\operatorname{fin}}}^{Tama}=\rho_{T}^{-1}\prod_{p}\mu_{p}, and μTama=|ω|\mu_{\infty}^{Tama}=|\omega|. We recall that by definition, ω=1|ΔK|ωT\omega=\frac{1}{\sqrt{|\Delta_{K}|}}\omega_{T}. Then by the definition of dmdm, we have

dmdηη=μTama=1|ΔK|daadηη,dm_{\infty}\wedge\frac{d\eta}{\eta}=\mu^{Tama}_{\infty}=\frac{1}{\sqrt{|\Delta_{K}|}}\frac{da_{\infty}}{a_{\infty}}\wedge\frac{d{\eta}}{\eta},

and thus

dm=1|ΔK|daa=1|ΔK|dzz.dm_{\infty}=\frac{1}{\sqrt{|\Delta_{K}|}}\frac{da_{\infty}}{a_{\infty}}=\frac{1}{\sqrt{|\Delta_{K}|}}\frac{dz}{z}.

We have computed above in (8) that the form dz/zdz/z gives precisely the arc length dθd\theta on the unit circle.

Thus, the volume of S1S^{1} with respect to dmdm_{\infty} is 2π|ΔK|\frac{2\pi}{\sqrt{|\Delta_{K}|}}.

Finally, we get from (50) and (45):

(53) 1=volTama(𝐓()\𝐓(𝔸)1)=2π|ΔK|volTama(𝐓()\𝐓(𝔸fin))\displaystyle 1=\operatorname{vol}^{Tama}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}})^{1})=\frac{2\pi}{\sqrt{|\Delta_{K}|}}\operatorname{vol}^{Tama}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}_{\mathbb{Q}}^{\operatorname{fin}}))
=2π|ΔK|ρT1volνT(K×\(𝔸Kfin)×)=2π|ΔK|L(1,χK)1hKwK,\displaystyle=\frac{2\pi}{\sqrt{|\Delta_{K}|}}\rho_{T}^{-1}\operatorname{vol}_{\nu_{T}}(K^{\times}\big{\backslash}\left(\mathbb{A}_{K}^{\operatorname{fin}})^{\times}\right)=\frac{2\pi}{\sqrt{|\Delta_{K}|}}L(1,\chi_{K})^{-1}\frac{h_{K}}{w_{K}},

recovering the analytic class number formula for KK. We note that the product L(1,χK)L(1,\chi_{K}) converges only conditionally.

5.2. What happens at p=2p=2

We do the detailed (and elementary but tedious) analysis of the changes one needs to make at p=2p=2 to all of the above calculations as they apply to an imaginary quadratic extension KK of \mathbb{Q} as above in order to completely justify the global calculation above, and also point out interesting geometric differences relevant for a norm-one torus of a quadratic extension131313This section can be skipped if the reader is willing to believe that (50) holds at p=2p=2 as well, and is not interested in the norm-1 torus (which is not used in the sequel)..

We write K=(d)K=\mathbb{Q}(\sqrt{d}), with dd a square-free (and in our case, negative) integer d=Dd=-D. Our main reference for the number theory information is e.g., [FT91, §VI.3].

5.2.1. Quadratic extension at p=2p=2

First of all, recall that the ring of integers of KK is

𝒪K={[d+12],d1mod4,[d],d2,3mod4.\mathcal{O}_{K}=\begin{cases}\mathbb{Z}\left[\frac{\sqrt{d}+1}{2}\right],&\quad d\equiv 1\mod 4,\\ \mathbb{Z}[\sqrt{d}],&\quad d\equiv 2,3\mod 4.\end{cases}

We will need the fact that the ring 𝒪K\mathcal{O}_{K} is generated over \mathbb{Z} by a root of the monic polynomial X2X+1d2X^{2}-X+\frac{1-d}{2} in the first case, and of X2dX^{2}-d in the second case. Therefore the behaviour of the prime 22 in KK depends on the residue of dd modulo 88. Indeed, if d1mod4d\equiv 1\mod 4, then we need to look at the reduction of the polynomial X2X+1d2X^{2}-X+\frac{1-d}{2} modulo 22; we see that it is irreducible over 𝔽2\mathbb{F}_{2} if 1d2\frac{1-d}{2} is odd, and it factors as x(x1)x(x-1) if 1d2\frac{1-d}{2} is even. Thus if d5mod8d\equiv 5\mod 8, KvK_{v} (where v|2v|2) is the unramified quadratic extension of 2\mathbb{Q}_{2}, while if d1mod8d\equiv 1\mod 8, the prime 22 splits in KK. In the remaining cases, 22 ramifies: if d3,7mod8d\equiv 3,7\mod 8, the relevant polynomial is X2dX^{2}-d, and its reduction mod2\mod 2 factors as (x1)2(x-1)^{2}; if d2,6mod8d\equiv 2,6\mod 8, then the reduction of our polynomial mod2\mod 2 is just x2x^{2}.

5.2.2. From Dedekind zeta-factor to Dirichlet LL-factor

While the local LL-factor at p=2p=2 itself looks a bit different from the other primes, the argument relating the local factor of L(1,χ)L(1,\chi) to the local factor of the Dedekind zeta-function is the same for all primes (including 22), when the Dirichlet character associated with KK is defined by (42). We observe that above, we have just computed the Dirichlet character of KK at 22:

(54) χK(2)={1d1mod81d5mod80 otherwise .\chi_{K}(2)=\begin{cases}1&\quad d\equiv 1\mod 8\\ -1&\quad d\equiv 5\mod 8\\ 0&\quad\text{ otherwise }\end{cases}.

By definition, the local factor at p=2p=2 (as at any other prime) of the Dedekind zeta-function is

ζ2(s)=𝔭(2)11NK/(𝔭)s,\zeta_{2}(s)=\prod_{\mathfrak{p}\supset(2)}\frac{1}{1-{\mathrm{N}}_{K/\mathbb{Q}}(\mathfrak{p})^{-s}},

where the product is over the prime ideals of 𝒪K\mathcal{O}_{K} lying over 22. It remains to recall that in all cases, for 𝔭\mathfrak{p} lying over (p)(p), NK/(𝔭)=pfN_{K/\mathbb{Q}}(\mathfrak{p})=p^{f}, where ff is the residue degree, see e.g. [FT91, II.4].

5.2.3. Quadratic extensions and norm-11 tori

If FF is a local field with residue characteristic 22, then |F×/(F×)2|=8|F^{\times}/(F^{\times})^{2}|=8, and FF has one unramified and 6 ramified quadratic extensions. (Indeed, we recall that F××𝒪F×F^{\times}\simeq\mathbb{Z}\times\mathcal{O}_{F}^{\times}, and for a 22-adic field, |𝒪F×/(𝒪F×)2|=4|\mathcal{O}_{F}^{\times}/(\mathcal{O}_{F}^{\times})^{2}|=4). In the case F=2F=\mathbb{Q}_{2} everything can again be computed in an elementary way, and this is what we do in this section. We can list the extensions explicitly using the discussion from §5.2.1: 2(5)\mathbb{Q}_{2}(\sqrt{5}) is the unramified extension; 2(3)\mathbb{Q}_{2}(\sqrt{3}) and 2(7)\mathbb{Q}_{2}(\sqrt{7}) are the ramified extensions coming from the non-square units, and 2(2)\mathbb{Q}_{2}(\sqrt{2}), 2(10)\mathbb{Q}_{2}(\sqrt{10}), and 2(6)\mathbb{Q}_{2}(\sqrt{6}) and 2(14)\mathbb{Q}_{2}(\sqrt{14}) are the ramified extensions corresponding to the elements ϖϵ\varpi\epsilon where ϖ=2\varpi=2 and ϵ\epsilon is a non-square unit. We compute the volumes of 𝐓(2)c\mathbf{T}(\mathbb{Q}_{2})^{c} with respect to the form ωT\omega_{T} as in §2.2, for the tori 𝐓=ResK/𝔾m\mathbf{T}=\operatorname{Res}_{K/\mathbb{Q}}\mathbb{G}_{m} as well as for the norm-1 tori of the corresponding extensions, for K=(d)K=\mathbb{Q}(\sqrt{d}), with d5,3,2mod8d\equiv 5,3,2\mod 8, to compare the calculation with Examples 2.3 and 2.5.

We recall that the identification E×=(ResE/F𝔾m)(F)E^{\times}=(\operatorname{Res}_{E/F}\mathbb{G}_{m})(F) is completely general; and for a quadratic extension EE, when we think of (ResE/F𝔾m)(F)(\operatorname{Res}_{E/F}\mathbb{G}_{m})(F) as a torus in GL2(F)\operatorname{GL}_{2}(F), the determinant in GL2\operatorname{GL}_{2} corresponds to the norm map NE/FN_{E/F} on E×E^{\times} under this identification. Till the end of the section, we keep the notation 𝐓=ResE/2𝔾m\mathbf{T}=\operatorname{Res}_{E/\mathbb{Q}_{2}}\mathbb{G}_{m} (we are now looking locally at p=2p=2, so to match the notation of §2.2, we have F=2F=\mathbb{Q}_{2} and EE is the completion of KK at p=2p=2).

1. The unramified case, E=2(5)E=\mathbb{Q}_{2}(\sqrt{5}). We write the elements of EE as x+φyx+\varphi y where φ=1+52E\varphi=\frac{1+\sqrt{5}}{2}\in E is a root of X2X1X^{2}-X-1 and x,y2x,y\in\mathbb{Q}_{2}; then 𝒪E={x+φyx,y2}\mathcal{O}_{E}=\{x+\varphi y\mid x,y\in\mathbb{Z}_{2}\} (of course, for elements of EE, the first representation is equivalent to simply writing x+y5x^{\prime}+y^{\prime}\sqrt{5}, but for the ring of integers this would not give the whole ring). The norm map in these coordinates is

NE/2(x+φy)=(x+y2+52y)(x+y252y)=x2+xyy2.N_{E/\mathbb{Q}_{2}}(x+\varphi y)=\left(x+\frac{y}{2}+\frac{\sqrt{5}}{2}y\right)\left(x+\frac{y}{2}-\frac{\sqrt{5}}{2}y\right)=x^{2}+xy-y^{2}.

This causes small changes to our naïve calculations of Example 2.3. In particular, the pullback to 𝐓\mathbf{T} of the differential form dxx\frac{dx}{x} on 𝔾m\mathbb{G}_{m} is now (exactly as in (3), with φ=152\varphi^{\prime}=\frac{1-\sqrt{5}}{2} denoting the Galois conjugate of φ\varphi)

(55) ωT=d(x+φy)x+φyd(x+φy)x+φy=(dx+φdy)(dx+φdy)x2+xyy2\displaystyle\omega_{T}=\frac{d(x+\varphi y)}{x+\varphi y}\wedge\frac{d(x+\varphi^{\prime}y)}{x+\varphi^{\prime}y}=\frac{(dx+\varphi dy)\wedge(dx+\varphi^{\prime}dy)}{x^{2}+xy-y^{2}}
=φφNE/F(x+φy)dxdy=5NE/F(x+φy)dxdy.\displaystyle=\frac{\varphi^{\prime}-\varphi}{N_{E/F}(x+\varphi y)}dx\wedge dy=\frac{-\sqrt{5}}{N_{E/F}(x+\varphi y)}dx\wedge dy.

Note the absence of the factor 22, compared with (3), which would have caused trouble here.

From the volume to the point-count: since the extension is unramified, this part does not change. In fancy terms, we can say that the so-called standard model over 2\mathbb{Z}_{2} for 𝐓\mathbf{T}, defined by the coordinates xx and yy that we chose, is smooth. The set of 𝔽2\mathbb{F}_{2}-points of its special fibre (in simple terms, the reduction mod 22, which is the uniformizer of our unramified extension) is still (𝔽4)×(\mathbb{F}_{4})^{\times}.

Thus, the volume formula (5) still holds in this case. If this extension was obtained as the completion at 22 of a quadratic extension K=(d)K=\mathbb{Q}(\sqrt{d}) of \mathbb{Q}, then d5mod8d\equiv 5\mod 8, and therefore in this case the discriminant of KK is just d\sqrt{d}. Hence, the relation (50) holds without any modification.

The norm-1 torus of this extension is treated very similarly to the unramified case with p2p\neq 2; and the final answer for its volume is given by the same formula as in the case p2p\neq 2.

2. Ramified case 1, E=2(2)E=\mathbb{Q}_{2}(\sqrt{2}). We have 𝒪E={x+y2x,y2}\mathcal{O}_{E}=\{x+y\sqrt{2}\mid x,y\in\mathbb{Z}_{2}\}, and the norm map is NE/2(x+2y)=x22y2N_{E/\mathbb{Q}_{2}}(x+\sqrt{2}y)=x^{2}-2y^{2} similarly to the p2p\neq 2 case. In this case the calculation of the form ωT\omega_{T} applies verbatim, so we get the extra factor 12\frac{1}{2} in (4). Specifically, (4) now becomes:

vol|ωT|(Tc)=122{(x,y)22:|x22y2|=1}𝑑x𝑑y.\operatorname{vol}_{|\omega_{T}|}(T^{c})=\frac{1}{2\sqrt{2}}\int_{\{(x,y)\in\mathbb{Z}_{2}^{2}:|x^{2}-2y^{2}|=1\}}dxdy.

The condition x22y22×x^{2}-2y^{2}\in\mathbb{Z}_{2}^{\times} is still equivalent to x2×x\in\mathbb{Z}_{2}^{\times}, and we obtain that vol|ωT|(Tc)=12qq1q\operatorname{vol}_{|\omega_{T}|}(T^{c})=\frac{1}{2\sqrt{q}}\frac{q-1}{q}; here q=2q=2, we are just writing it this way for the ease of comparison with (5). However, the relation (50) again holds without modification, since if the completion of K=(d)K=\mathbb{Q}(\sqrt{d}) at 22 is ramified, then ΔK=4d\Delta_{K}=4\sqrt{d}, and |ΔK|21/2|\Delta_{K}|_{2}^{1/2} has the extra factor 22 as well.

Example 5.1.

The norm-1 torus of E=2(2)E=\mathbb{Q}_{2}(\sqrt{2}): Unlike the full torus obtained by the restriction of scalars, for the norm-1 subtorus the reduction of the volume computation to counting residue-field points looks very different from Example 2.5. We include this point-count exercise without a full discussion of its implications for the computation of the volume with respect to ωT\omega_{T} or ωcan\omega^{\operatorname{can}}, to illustrate the difficulties that arise when the reduction modϖ\mod\varpi is not smooth.

We write the 22-adic expansions x=x0+2x1+4x2+,x=x_{0}+2x_{1}+4x_{2}+\ldots, y=y0+2y1+4y2+,y=y_{0}+2y_{1}+4y_{2}+\ldots, with xi,yi{0,1}x_{i},y_{i}\in\{0,1\}. Then

(56) x2\displaystyle x^{2} =(x0+2x1+4x2+8x3+16x4+)2\displaystyle=(x_{0}+2x_{1}+4x_{2}+8x_{3}+16x_{4}+\ldots)^{2}
=x02+4(x0x1+x12)+8(x0x2)+16(x22+x1x2+x0x3)\displaystyle=x_{0}^{2}+4(x_{0}x_{1}+x_{1}^{2})+8(x_{0}x_{2})+16(x_{2}^{2}+x_{1}x_{2}+x_{0}x_{3})
+25(x0x4+x1x3)+26(x32+x0x5+x1x4+x2x3)+.\displaystyle+2^{5}(x_{0}x_{4}+x_{1}x_{3})+2^{6}(x_{3}^{2}+x_{0}x_{5}+x_{1}x_{4}+x_{2}x_{3})+\ldots.

Thus the condition x22y2=1x^{2}-2y^{2}=1 becomes (each line comes from the congruence modulo the next power of 22 (indicated in the left column), and each congruence is congruence mod2\mod 2):

mod2:\displaystyle\mod 2: x0=1;\displaystyle\quad x_{0}=1;
mod22:\displaystyle\mod 2^{2}: y0=0;\displaystyle\quad y_{0}=0;
mod23:\displaystyle\mod 2^{3}: x0x1+x120;\displaystyle\quad x_{0}x_{1}+x_{1}^{2}\equiv 0;
mod24:\displaystyle\mod 2^{4}: x0x2(y0y1+y12)0;\displaystyle\quad x_{0}x_{2}-(y_{0}y_{1}+y_{1}^{2})\equiv 0;
mod25:\displaystyle\mod 2^{5}: x22+x1x2+x0x3(y0y2)0;\displaystyle\quad x_{2}^{2}+x_{1}x_{2}+x_{0}x_{3}-(y_{0}y_{2})\equiv 0;
mod26:\displaystyle\mod 2^{6}: x0x4+x1x3(y22+y1y2+y0y3)0;\displaystyle\quad x_{0}x_{4}+x_{1}x_{3}-(y_{2}^{2}+y_{1}y_{2}+y_{0}y_{3})\equiv 0;
mod27:\displaystyle\mod 2^{7}: x32+x0x5+x1x4+x2x3(y0y4+y1y3)0;\displaystyle\quad x_{3}^{2}+x_{0}x_{5}+x_{1}x_{4}+x_{2}x_{3}-(y_{0}y_{4}+y_{1}y_{3})\equiv 0;
\displaystyle\ldots .\displaystyle\quad\ldots.

These equations yield:

(57) x0=1,\displaystyle x_{0}=1, y0=0,\displaystyle\quad y_{0}=0,
x1 is arbitrary,\displaystyle x_{1}\text{ is arbitrary}, y1 is arbitrary,\displaystyle\quad y_{1}\text{ is arbitrary},
x2=y1,\displaystyle x_{2}=y_{1}, y2 is arbitrary,\displaystyle\quad y_{2}\text{ is arbitrary},
x3=x22+x1x2,\displaystyle x_{3}=x_{2}^{2}+x_{1}x_{2}, y3 is arbitrary,\displaystyle\quad y_{3}\text{ is arbitrary},
\displaystyle\ldots

This illustrates that Hensel’s Lemma, as expected, starts working once we have a solution mod8\mod 8, but not for solutions mod2\mod 2. In fancier terms, we have the reduction mod8\mod 8 map defined on the set of (2×2)(\mathbb{Z}_{2}\times\mathbb{Z}_{2})-solutions of the norm equation x22y2=1x^{2}-2y^{2}=1. The image of this map is the set {(x0,y0,x1,y1,x2,y2)}\{(x_{0},y_{0},x_{1},y_{1},x_{2},y_{2})\} defined by the first three lines of (57) inside 𝔸6(𝔽2)\mathbb{A}^{6}(\mathbb{F}_{2}); we see that it is a 33-dimensional hyperplane in 𝔸6(𝔽2)\mathbb{A}^{6}(\mathbb{F}_{2}). The fibre of the reduction mod8\mod 8 map over each point in its image is a translate of (23)Z2(2^{3})\subset Z_{2}, so the volume of each fibre is 123\frac{1}{2^{3}}. We obtain that the volume of our set of solutions is #𝔸3(𝔽2)23\frac{\#\mathbb{A}^{3}(\mathbb{F}_{2})}{2^{3}}. Note that this is geometrically quite different from the answer in the ramified case in Example 2.5, where the image of the reduction modp\mod p map was two copies of an affine line, and therefore the image of the reduction modp2\mod p^{2} map then was two copies of an affine plane over 𝔽p\mathbb{F}_{p}; and the image of reduction modp3\mod p^{3} map, inside 𝔸6(𝔽p)\mathbb{A}^{6}(\mathbb{F}_{p}), was two copies of 𝔸3(𝔽p)\mathbb{A}^{3}(\mathbb{F}_{p}).

3. Ramified case 2, E=2(3)E=\mathbb{Q}_{2}(\sqrt{3}). Exactly as in the case E=2(2)E=\mathbb{Q}_{2}(\sqrt{2}) considered above, for 𝐓=ResE/2𝔾m\mathbf{T}=\operatorname{Res}_{E/\mathbb{Q}_{2}}\mathbb{G}_{m}, the volume vol|ωT|(Tc)\operatorname{vol}_{|\omega_{T}|}(T^{c}) has the extra factor of 12\frac{1}{2} compared with the p2p\neq 2, and so does the discriminant (basically, the calculation of the volume of TcT^{c} is not sensitive to which ramified extension of 2\mathbb{Q}_{2} we are considering).

Example 5.2.

The norm-1 torus of E=2(3)E=\mathbb{Q}_{2}(\sqrt{3}). The calculation is also similar to 5.1 above: we use (56) to count solutions of the equation x23y2=1x^{2}-3y^{2}=1, but the geometry looks slightly different. As above, from (56) we get:

mod2:\displaystyle\mod 2: x02+y02=1;\displaystyle\quad x_{0}^{2}+y_{0}^{2}=1;
mod22:\displaystyle\mod 2^{2}: x023y021mod4;\displaystyle\quad x_{0}^{2}-3y_{0}^{2}\equiv 1\mod 4;
mod23:\displaystyle\mod 2^{3}: x0x1+x123(y0y1+y12)0mod2;\displaystyle\quad x_{0}x_{1}+x_{1}^{2}-3(y_{0}y_{1}+y_{1}^{2})\equiv 0\mod 2;

From the first two equations, we get: x0=1x_{0}=1, y0=0y_{0}=0 (note that the second congruence mod4\mod 4 does not allow for the option x0=0x_{0}=0, y0=1y_{0}=1, in contrast to the ramified case with p2p\neq 2; this is an example of a solution mod2\mod 2 that does not lift to a solution mod4\mod 4). Then the third equation becomes x1+x123y120mod2x_{1}+x_{1}^{2}-3y_{1}^{2}\equiv 0\mod 2, which allows for arbitrary x1x_{1} and makes y1=0y_{1}=0. Next, from the truncation mod24\mod 2^{4}, we get (plugging all this information in):

4(x1+x12)8(x230y2)0mod16.\quad 4(x_{1}+x_{1}^{2})-8(x_{2}-3\cdot 0\cdot y_{2})\equiv 0\mod 16.

If x1=0x_{1}=0, we get x2=0x_{2}=0; if x1=1x_{1}=1, then x2=1x_{2}=1. Summing it up, so far we have:

(58) x0=1,\displaystyle x_{0}=1, y0=0,\displaystyle\quad y_{0}=0,
x1 is arbitrary,\displaystyle x_{1}\text{ is arbitrary}, y1=0,\displaystyle\quad y_{1}=0,
x2=x1,\displaystyle x_{2}=x_{1}, y2 is arbitrary,\displaystyle\quad y_{2}\text{ is arbitrary},

Note that the pattern so far has been something we have not seen before: the congruences modulo an odd power of 22 might have some ‘carry-over’ to the next power; then the congruence modulo the next even power forces the truncated expression to become literally zero with no carry-over.

Continuing with one more step:

mod25:\displaystyle\mod 2^{5}:
x22+x1x2+x0x33(y22+y1y2+y0y3)0mod2;\displaystyle{}\quad x_{2}^{2}+x_{1}x_{2}+x_{0}x_{3}-3(y_{2}^{2}+y_{1}y_{2}+y_{0}y_{3})\equiv 0\mod 2;
mod26:\displaystyle\mod 2^{6}:
(x22+x1x2+x0x33(y22+y1y2+y0y3))+(x0x4+x1x33(y0y4+y1y3))0,\displaystyle{}\quad(x_{2}^{2}+x_{1}x_{2}+x_{0}x_{3}-3(y_{2}^{2}+y_{1}y_{2}+y_{0}y_{3}))+(x_{0}x_{4}+x_{1}x_{3}-3(y_{0}y_{4}+y_{1}y_{3}))\equiv 0,

where the last congruence is mod4\mod 4. As we plug in what we already know about the first terms, the first equation becomes x3+y220mod2x_{3}+y_{2}^{2}\equiv 0\mod 2, so x3=y2x_{3}=y_{2}. Plugging this into the second equation (and ignoring the squares), we get (2x22x3)+(x4+x3)0mod4(2x_{2}-2x_{3})+(x_{4}+x_{3})\equiv 0\mod 4, which determines x4x_{4} uniquely. Again we notice that by now Hensel’s Lemma started working as expected: at every step we get one linear relation in two unknown parameters, so the fibre over each truncated solution is an affine line over 2\mathbb{Q}_{2}.

To summarize, (58) says that the image of the reduction mod8\mod 8 map is a plane in 𝔸6(𝔽2)\mathbb{A}^{6}(\mathbb{F}_{2}), and thus for the volume of TcT^{c}, we get #𝔸2(𝔽2)23\frac{\#\mathbb{A}^{2}(\mathbb{F}_{2})}{2^{3}}.

Now we are ready to return to the global calculations.

5.3. Global orbital integrals in GL2\operatorname{GL}_{2}

Let us put together the information we have so far about the orbital integrals with respect to the canonical vs. geometric measures, and the information about the volume of K×\(𝔸Kfin)×K^{\times}\big{\backslash}\left(\mathbb{A}_{K}^{\operatorname{fin}}\right)^{\times} that we just obtained.

Let γGL2()\gamma\in\operatorname{GL}_{2}(\mathbb{Q}) be a regular semisimple element, such that its centralizer TT is non-split over \mathbb{R}. Let KK be the quadratic extension of \mathbb{Q} generated by the eigenvalues of γ\gamma. Then T=𝐓()T=\mathbf{T}(\mathbb{Q}) with 𝐓=ResK/𝔾m\mathbf{T}=\operatorname{Res}_{K/\mathbb{Q}}\mathbb{G}_{m}, as in §5.1. Let ffin=pfpf^{\operatorname{fin}}=\otimes_{p}f_{p}, with fpf_{p} equal to 𝟏GL2(p){\bf 1}_{\operatorname{GL}_{2}(\mathbb{Z}_{p})}, the characteristic function of GL2(p)\operatorname{GL}_{2}(\mathbb{Z}_{p}), for almost all pp, be a test function on G(𝔸fin)G(\mathbb{A}^{\operatorname{fin}}). Taking the product of the relations (36) at every prime pp, and applying the product formula to the absolute values p|D(γ)|p\prod_{p}|D(\gamma)|_{p} and p|ΔK|p\prod_{p}|\Delta_{K}|_{p}, we obtain:

(59) Oγgeom(ffin)=|D(γ)|1/2|ΔK|1/2L(1,σG/T)Oγcan(ffin).O_{\gamma}^{{\operatorname{geom}}}(f^{\operatorname{fin}})=|D(\gamma)|^{-1/2}|\Delta_{K}|^{1/2}L(1,\sigma_{G/T})O_{\gamma}^{{\operatorname{can}}}(f^{\operatorname{fin}}).

We observe that the canonical measure on the centralizer of γ\gamma (or any measure coinciding with it at almost all places) is convenient for defining a global orbital integral because with such a measure, all but finitely many factors Oγp(fp)O_{\gamma_{p}}(f_{p}) are equal to 11, and thus the global orbital integral is a finite product, namely, the product over the primes that divide D(γ)D(\gamma) and the primes where fp𝟏G(p)f_{p}\neq{\bf 1}_{G(\mathbb{Z}_{p})}, and there is no question of convergence.

In the Trace Formula, the orbital integrals are weighted by volumes; and the volume has to be taken with respect to the same measure on the centralizer that was used to define the orbital integral. Comparing (59) with (41), we see explicitly that for GL2\operatorname{GL}_{2}, the volume term contains some of the factors that also appear when we pass from the canonical measure to the geometric measure on the orbit. Now we are ready to explicate an observation (that is implicit in the work of Langlands) that switching to the geometric measure makes the volume term disappear in the case 𝐆=GL2\mathbf{G}=\operatorname{GL}_{2}, in addition to making the orbital integrals at finite places better-behaved. This comes at the cost of now having the orbital integral expressed as only a conditionally convergent infinite product. This theorem can be thought of as the main point of this note.

Theorem 5.3.

Let γ\gamma be a regular elliptic element of GL2\operatorname{GL}_{2} as above, that splits over a quadratic extension KK. Let 𝐓=ResK/𝔾m\mathbf{T}=\operatorname{Res}_{K/\mathbb{Q}}\mathbb{G}_{m}. Then

volcan(𝐓()\𝐓(𝔸fin))Oγcan(ffin)=|D(γ)|1/22πOγgeom(ffin).\operatorname{vol}^{\operatorname{can}}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}^{\operatorname{fin}}))O_{\gamma}^{\operatorname{can}}(f^{\operatorname{fin}})=\frac{|D(\gamma)|^{1/2}}{2\pi}O_{\gamma}^{\operatorname{geom}}(f^{\operatorname{fin}}).
Proof.

Combining the relation (59) with the calculation in (53), we see that the factors L(1,χK)L(1,\chi_{K}) and |ΔK|1/2|\Delta_{K}|^{1/2} cancel out, and we obtain:

volcan(𝐓()\𝐓(𝔸fin))Oγcan(ffin)=|ΔK|2πL(1,χT)volTama(𝐓()\𝐓(𝔸)1)Oγcan(ffin)\displaystyle\operatorname{vol}^{\operatorname{can}}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A}^{\operatorname{fin}}))O_{\gamma}^{\operatorname{can}}(f^{\operatorname{fin}})=\frac{\sqrt{|\Delta_{K}|}}{2\pi}L(1,\chi_{T})\operatorname{vol}^{Tama}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A})^{1})O_{\gamma}^{\operatorname{can}}(f^{\operatorname{fin}})
=|D(γ)|1/22πτ(𝐓)Oγgeom(ffin).\displaystyle=\frac{|D(\gamma)|^{1/2}}{2\pi}\tau(\mathbf{T})O_{\gamma}^{\operatorname{geom}}(f^{\operatorname{fin}}).

Since 𝐓\mathbf{T} is obtained from 𝔾m\mathbb{G}_{m} by restriction of scalars, we have τ(𝐓)=1\tau(\mathbf{T})=1, as discussed above in §5.1.1. ∎

We make a few remarks:

Remark 5.4.

1. In our statement, the left-hand side of the equation is actually independent of the choice of the measure on 𝐓(p)\mathbf{T}(\mathbb{Q}_{p}) at every finite place pp, as long as the volume and the orbital integral are taken with respect to the same measure; this is consistent with the right-hand side, which does not involve any measure on 𝐓\mathbf{T} at all.

2. The volume that appears in the Trace Formula is vol(𝐓()\𝐓(𝔸)1)\operatorname{vol}(\mathbf{T}(\mathbb{Q})\backslash\mathbf{T}(\mathbb{A})^{1}); when passing from the volume appearing on the left-hand side of Theorem 5.3 to this volume, the ratio between them will depend on the precise choice of the normalization of the component of the measure at infinity. Calculations of this sort (for a general reductive group, with various specific choices of the measure at infinity) appear in [GG99] and [Gro97].

3. The right-hand side of the relation in Theorem 5.3 might be preferable in two ways: there is no complicated volume term, and the orbital integral has the local components that are continuous as functions on the Steinberg-Hitchin base (however, now the orbital integral on the right is an infinite product that converges conditionally).

4. The proof of the theorem does not use the analytic class number formula. Moreover, the proof is general, except for three pieces:

  1. (1)

    The specific knowledge that at every finite place vv, volωT(𝒯0)\operatorname{vol}_{\omega_{T}}(\mathcal{T}^{0}) is |ΔK|v1/2Lv(1,χT)1|\Delta_{K}|_{v}^{1/2}L_{v}(1,\chi_{T})^{-1}. See [GG99] for a generalization of such a relation.

  2. (2)

    For general tori Tamagawa numbers can be difficult to compute explicitly (see [Rüd20] for some partial results), but for the maximal tori in GLn\operatorname{GL}_{n} they are 11.

  3. (3)

    The factor 2π2\pi in the denominator is specific to GL2\operatorname{GL}_{2}; in general, it needs to be replaced with the factor determined by the component of the chosen measure at infinity.

We conclude with a brief discussion of Eichler-Selberg Trace Formula, since it is the starting point for Altug’s lectures. This discussion is entirely based on [KL06].

5.4. Eichler-Selberg Trace Formula for GL2\operatorname{GL}_{2}

The Eichler-Selberg Trace Formula expresses the trace of a Hecke operator TnT_{n} on the space Sk(N)S_{k}(N) of cusp forms of weight kk and level NN. For simplicity of exposition, we set N=1N=1 in this note; in this case the central character is also trivial. In this setting, the Eichler-Selberg Trace Formula states:

(60) n1k/2TrTn\displaystyle n^{1-k/2}\operatorname{Tr}T_{n} =k112{1,n is a perfect square0 otherwise\displaystyle=\frac{k-1}{12}\begin{cases}&1,n\text{ is a perfect square}\\ &0\text{ otherwise }\end{cases}
12t2<4nρk1ρ¯k1ρρ¯mhw(t24nm2)\displaystyle-\frac{1}{2}\sum_{t^{2}<4n}\frac{\rho^{k-1}-\bar{\rho}^{k-1}}{\rho-\bar{\rho}}\sum_{m}h_{w}\left(\frac{t^{2}-4n}{m^{2}}\right)
12dnmin(d,nd)k1,\displaystyle-\frac{1}{2}\sum_{d\mid n}\min(d,\frac{n}{d})^{k-1},

where in the middle line ρ\rho and ρ¯\bar{\rho} are roots of the polynomial X2tX+nX^{2}-tX+n, and hw(t24nm2)h_{w}\left(\frac{t^{2}-4n}{m^{2}}\right) is the weighted class number of the order in [ρ]\mathbb{Q}[\rho] that has discriminant t24nm2\frac{t^{2}-4n}{m^{2}}. The sum over mm runs over the integers m1m\geq 1 such that m2m^{2} divides t24nt^{2}-4n, and t24nm2\frac{t^{2}-4n}{m^{2}} is 0 or 11 mod4\mod 4.

The goal of this section is to sketch, without any detail, a connection between this formula and the (geometric side of) Arthur-Selberg Trace formula.

5.4.1. The test function

We start with a very brief recall of the connection between modular forms and automorphic forms on GL2\operatorname{GL}_{2}. We refer to e.g., Knightly and Li [KL06] for all the details.

Let 𝐆=GL2\mathbf{G}=\operatorname{GL}_{2} for the rest of this section. A cusp form of weight kk generates (as a representation of 𝐆(𝔸)\mathbf{G}(\mathbb{A}_{\mathbb{Q}}) under the action by right translations) a closed subspace (π,V)(\pi,V) of L02(𝐆()\𝐆(𝔸))L^{2}_{0}(\mathbf{G}(\mathbb{Q})\backslash\mathbf{G}(\mathbb{A})); for a level 11 cusp form (which is our assumption here, so in particular we should assume k>2k>2), the central character of π\pi is trivial. By Flath’s theorem, the representation π\pi factors as a restricted tensor product

π=πfinπ=pπpπ.\pi=\pi_{\operatorname{fin}}\otimes\pi_{\infty}=\otimes^{\prime}_{p}\pi_{p}\otimes\pi_{\infty}.

Since (π,V)(\pi,V) came from a cusp form of weight kk, we have π=πk\pi_{\infty}=\pi_{k} – the discrete series representation of highest weight kk.

Every function fCc(𝐆(𝔸))f\in C_{c}(\mathbf{G}(\mathbb{A})) gives rise to a linear operator R(f)R(f) on L2(𝐙(𝔸)\𝐆(𝔸))L^{2}(\mathbf{Z}(\mathbb{A})\backslash\mathbf{G}(\mathbb{A})) defined by

(R(f)ϕ)(g):=(𝐙\𝐆)(𝔸)f(x)ϕ(gx)𝑑x.(R(f)\phi)(g):=\int_{(\mathbf{Z}\backslash\mathbf{G})(\mathbb{A})}f(x)\phi(gx)dx.

In this language, the Hecke operator TnT_{n} on SkS_{k} is precisely nk/21R(fn,k)n^{k/2-1}R(f_{n,k}), where fn,kf_{n,k} is a specific test function in the space L2(𝐆()\𝐆(𝔸)1)L^{2}(\mathbf{G}(\mathbb{Q})\backslash\mathbf{G}(\mathbb{A})^{1}). We quote the definition of this test function from [KL06].

  • Let ff_{\infty} be a matrix coefficient of the representation πk\pi_{k}. By orthogonality of matrix coefficients, this ensures that the image of R(f)R(f) is contained in (π,V)(\pi,V). (Since (π,V)(\pi,V) is irreducible, this means R(f)R(f) projects onto VV).

  • For pnp\nmid n, let fpf_{p} be the characteristic function of 𝐙(p)𝐆(p)\mathbf{Z}(\mathbb{Q}_{p})\mathbf{G}(\mathbb{Z}_{p}),

  • For pnp\mid n, let fpf_{p} be the characteristic function of 𝐙(p)Mn,p\mathbf{Z}(\mathbb{Q}_{p})M_{n,p}, where Mn,pM_{n,p} is the set of matrices of determinant nn in M2(p)M_{2}(\mathbb{Q}_{p}) (we prefer to think of it as the characteristic function of the union of the double cosets of the Cartan decomposition for GL2(p)\operatorname{GL}_{2}(\mathbb{Q}_{p}) of determinant nn).

5.4.2. The transition to Arthur Trace Formula

We plug f:=fn,kf:=f_{n,k} into Arthur’s Trace Formula for GL2\operatorname{GL}_{2}, and examine the geometric side. Since for our test function ff, the continuous and residual parts of the spectral side vanish, the geometric side in fact equals TrR(f)\operatorname{Tr}R(f) (see [KL06, §22]). Finally, Knightly and Li show that:

  • The first line of (60) matches the contribution of the trivial conjugacy class;

  • The last line matches the contribution of the unipotent and hyperbolic conjugacy classes.

  • The middle line matches the contribution of the elliptic conjugacy classes.

We discuss why the last claim is plausible. By definition of the test function ff, its orbital integrals vanish on all elements γ\gamma such that det(γ)n\det(\gamma)\neq n; hence, in the geometric side of Arthur’s Trace Formula, we are left with the sum over γ𝐆()\gamma\in\mathbf{G}(\mathbb{Q}) satisfying det(γ)=n\det(\gamma)=n. The conjugacy classes in GL2\operatorname{GL}_{2} are parametrized by characteristic polynomials, and the elliptic ones correspond to the polynomials with negative discriminants, so at least superficially, we recognize the sum over the integers tt such that t2<4nt^{2}<4n as a sum over the rational elliptic conjugacy classes.

Next, note that the expression ρk1ρ¯k1ρρ¯\frac{\rho^{k-1}-\bar{\rho}^{k-1}}{\rho-\bar{\rho}} is the value of the character of πk\pi_{k} on the corresponding conjugacy class; thus, we recognize it as the orbital integral of ff_{\infty} (see e.g. [Kot05, §1.11] for the discussion of characters as orbital integrals of matrix coefficients; see also [GGPS16, Ch.I, §5.2]).

Knightly and Li show (in our notation):

(61) Oγcan(ffin)=mhw(t24nm2),O_{\gamma}^{\operatorname{can}}(f^{\operatorname{fin}})=\sum_{m}h_{w}\left(\frac{t^{2}-4n}{m^{2}}\right),

where the sum over mm is as in (60). While the appearance of class numbers in our earlier calculations is suggestive, and proves this relation in the trivial case when t24nt^{2}-4n is square-free, it appears that our arguments are insufficient for getting a simpler proof of this claim in general (other than by essentially direct computation of the both sides, or matching the computation of the right-hand side in [KL06] with the calculation on the building in [Kot05], the results of which we already quoted above). A similar statement for GLn\operatorname{GL}_{n}, relating orbital integrals to sums of class numbers of orders, is proved by Zhiwei Yun, [Yun13].

6. Appendix A. Kirillov’s form on co-adjoint orbits: two examples

by Matthew Koster

In this appendix we illustrate Kirillov’s construction of a volume form on co-adjoint orbits in a Lie algebra. Here we work over \mathbb{R} in order to be able to use the intuition from calculus. In these examples, we also relate this form to the geometric measure discussed in the article. 141414This work was part of an NSERC summer USRA project in the summer of 2019; we acknowledge the support of NSERC.

6.1. The coadjoint orbits

A large part of this section is quoted from [Kot05, §17.3] for the reader’s convenience and to set up notation. Let GG denote a semisimple Lie group, 𝔤\mathfrak{g} its Lie algebra, and 𝔤\mathfrak{g}^{*} the linear dual space of 𝔤\mathfrak{g}. We denote elements of 𝔤\mathfrak{g} by capital letters, e.g. XX, and use starts to denote elements of 𝔤\mathfrak{g}^{\ast}; unless explicitly stated there is no a priori relationship between XX and XX^{*}.

GG acts on 𝔤\mathfrak{g} by Ad and acts on 𝔤\mathfrak{g}^{*} by Ad, where

Ad(g)(X),X=X,Ad(g1)(X).\langle\operatorname{Ad}^{*}(g)(X^{*}),X\rangle=\langle X^{*},\operatorname{Ad}(g^{-1})(X)\rangle.

Let 𝒪(X)𝔤\mathcal{O}(X^{*})\subset\mathfrak{g}^{*} denote the orbit of XX^{*} under this action (called a co-adjoint orbit).
We recall that the differential of the adjoint action Ad\operatorname{Ad} of GG is the action of 𝔤\mathfrak{g} on itself by ad\operatorname{ad} where adX(Z)=[X,Z]\operatorname{ad}_{X}(Z)=[X,Z]. The co-adjoint action of 𝔤\mathfrak{g} on 𝔤\mathfrak{g}^{*} is the differential of Ad\operatorname{Ad}^{*}; we denote it by ad\operatorname{ad}^{*}; explicitly, this action is defined by adX(Y),Z=Y,[Z,X]\langle\operatorname{ad}_{X}^{*}(Y^{\ast}),Z\rangle=\langle Y^{\ast},[Z,X]\rangle.

A choice of an element X𝔤X^{\ast}\in\mathfrak{g}^{\ast} defines a map φX:G𝔤\varphi_{X^{\ast}}:G\to\mathfrak{g}^{*} by φX(g)=Adg(X)\varphi_{X^{\ast}}(g)=\operatorname{Ad}^{*}_{g}(X^{*}). The differential of this map at the identity eGe\in G gives an identification of 𝔤/𝔠(X)\mathfrak{g}/\mathfrak{c}(X^{*}) with the tangent space TX𝒪(X)T_{X^{*}}\mathcal{O}(X^{*}) at XX^{\ast}, defined by XadX(X)X\mapsto\operatorname{ad}^{*}_{X}(X^{*}). Here 𝔠(X)\mathfrak{c}(X^{\ast}) is the stabilizer of XX^{\ast} under ad\operatorname{ad}^{\ast}, and we are viewing TX𝒪(X)T_{X^{*}}\mathcal{O}(X^{*}) as a subspace of TX𝔤𝔤T_{X^{*}}\mathfrak{g}^{*}\cong\mathfrak{g}^{*}. We denote this identification by ΦX:𝔤/𝔠(X)TX𝒪(X)𝔤\Phi_{X^{\ast}}:\mathfrak{g}/\mathfrak{c}(X^{\ast})\to T_{X^{\ast}}\mathcal{O}(X^{\ast})\hookrightarrow\mathfrak{g}^{*}. The element XX^{\ast} gives an alternating form ωX\omega_{X^{\ast}} on 𝔤\mathfrak{g}, defined by

(62) ωX(X,Y):=X,[X,Y]=ad(X)X,Y.\omega^{\prime}_{X^{\ast}}(X,Y):=\langle X^{\ast},[X,Y]\rangle=-\langle\operatorname{ad}^{\ast}(X)X^{\ast},Y\rangle.

This form clearly vanishes on 𝔠(X)\mathfrak{c}(X^{\ast}), and gives a non-degenerate bilinear form on 𝔤/𝔠(X)\mathfrak{g}/\mathfrak{c}(X^{\ast}), which we have just identified with TX𝒪(X)T_{X^{\ast}}\mathcal{O}(X^{\ast}). Thus, given a co-adjoint orbit 𝒪\mathcal{O}, we get a symplectic 22-form ω\omega^{\prime} on it by letting the value of ω\omega^{\prime} at X𝒪X^{\ast}\in\mathcal{O} equal ωX\omega^{\prime}_{X^{\ast}}. In particular, as a manifold, 𝒪\mathcal{O} has to have even dimension; if its dimension is 2k2k, then the kk-fold wedge product of the form ω\omega^{\prime} gives a volume form on 𝒪\mathcal{O}.

Over a field of characteristic zero, we can identify a semisimple Lie algebra with its dual; we will use the Killing form for this. Then the adjoint orbits in 𝔤\mathfrak{g} get identified with the co-adjoint orbits in 𝔤\mathfrak{g}^{\ast}, and thus we get a very natural algebraic volume form on each adjoint orbit in 𝔤\mathfrak{g}. Here our goal is to compute this form explicitly in two examples: the regular nilpotent orbit in 𝔰𝔩2()\mathfrak{sl_{2}}(\mathbb{R}) and a semisimple SO3()\mathrm{SO}_{3}(\mathbb{R})-orbit in 𝔰𝔬3()\mathfrak{so}_{3}(\mathbb{R}) (we do not use the accidental isomorphism in this calculation). In both cases the orbit will be two-dimensional, so we are just computing the form denoted by ω\omega^{\prime} above.

6.2. Rewriting the form as a form on an orbit in 𝔤\mathfrak{g}

Given X0𝔤X_{0}^{\ast}\in\mathfrak{g}^{\ast}, we compute the form ω\omega on the orbit of an element X0𝔤X_{0}\in\mathfrak{g} that corresponds to X0𝔤X_{0}^{\ast}\in\mathfrak{g}^{\ast} under the isomorphism defined by Killing form, in three steps:

  1. (1)

    Compute the map ΦX0\Phi_{X_{0}^{\ast}}.

  2. (2)

    For v1~,v2~TX0𝒪(X0)\widetilde{v_{1}},\widetilde{v_{2}}\in T_{X_{0}^{*}}\mathcal{O}(X_{0}^{*}) find v1,v2𝔤v_{1},v_{2}\in\mathfrak{g} with ΦX0(vi)=vi~\Phi_{X_{0}^{\ast}}(v_{i})=\widetilde{v_{i}} for i=1,2i=1,2, and then evaluate ωX0(v1~,v2~)=X0,[v1,v2]\omega_{X_{0}^{*}}(\widetilde{v_{1}},\widetilde{v_{2}})=\langle X_{0}^{*},[v_{1},v_{2}]\rangle.

  3. (3)

    Using Killing form, identify 𝔤\mathfrak{g} with 𝔤\mathfrak{g}^{\ast}, which identifies a co-adjoint orbit of X0X_{0}^{\ast} in 𝔤\mathfrak{g}^{\ast} with an adjoint orbit of an element X0𝔤X_{0}\in\mathfrak{g}. Then use the adjoint action to explicitly define the volume form ωX\omega_{X} on TX𝒪T_{X}\mathcal{O} at a point XX in this orbit by pulling back the form ωX0\omega_{X_{0}} .

6.3. Example I: a regular nilpotent orbit in 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R})

Let G=SL(2;)G=\operatorname{SL}(2;\mathbb{R}), 𝔤=𝔰𝔩(2;)\mathfrak{g}=\mathfrak{sl}(2;\mathbb{R}), and consider the standard basis {𝐞,𝐟,𝐡}\{\mathbf{e},\mathbf{f},\mathbf{h}\} for 𝔤\mathfrak{g} given by:

𝐞=[0100],𝐟=[0010],𝐡=[1001].\displaystyle{\bf e}=\begin{bmatrix}0&1\\ 0&0\end{bmatrix},\ \ \ \ {\bf f}=\begin{bmatrix}0&0\\ 1&0\end{bmatrix},\ \ \ \ {\bf h}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}.

Let {𝐞,𝐟,𝐡}\{\mathbf{e}^{*},\mathbf{f}^{*},\mathbf{h}^{*}\} be the basis for 𝔤\mathfrak{g}^{*} dual to {𝐞,𝐟,𝐡}\{\mathbf{e},\mathbf{f},\mathbf{h}\} under the Killing form. Explicitly this means that 𝐞(𝐟)=4\mathbf{e}^{*}(\mathbf{f})=4, 𝐟(𝐞)=4\mathbf{f}^{*}(\mathbf{e})=4, and 𝐡(𝐡)=8\mathbf{h}^{*}(\mathbf{h})=8. We will compute Kirillov form on the co-adjoint orbit 𝒪𝐟\mathcal{O}_{\mathbf{f}^{\ast}} of 𝐟\mathbf{f}^{*}, which we identify with the adjoint orbit of 𝐞\mathbf{e} in 𝔤\mathfrak{g}.

When we refer to coordinates x,y,zx,y,z on 𝔤\mathfrak{g}, it is with respect to our chosen basis {𝐞,𝐟,𝐡}\{\mathbf{e},\mathbf{f},\mathbf{h}\}. Given this choice of coordinates, we have the basis of the space of 11-forms on 𝔤\mathfrak{g} given by dxdx, dydy, dzdz.

Under the isomorphism 𝔤𝔤\mathfrak{g}^{\ast}\simeq\mathfrak{g} defined by Killing form, a point (x,y,z)𝔤(x,y,z)\in\mathfrak{g}^{\ast} is mapped to [z/2xyz/2]𝔤\left[\begin{smallmatrix}z/2&x\\ y&-z/2\end{smallmatrix}\right]\in\mathfrak{g}. We can describe the orbit of 𝐞\mathbf{e} very explicitly in these coordinates.

6.3.1. The nilpotent cone

If we are working over \mathbb{R}, then the set of nilpotent elements in 𝔤\mathfrak{g} forms a cone: indeed, for a nilpotent matrix we have det[z/2xyz/2]=0\det\left[\begin{smallmatrix}z/2&x\\ y&-z/2\end{smallmatrix}\right]=0, i.e., z2+4xy=0z^{2}+4xy=0. (One can easily see that it is, indeed, a cone by the change of coordinates u=x+yu=x+y, v=xyv=x-y: in these coordinates, the equation of the orbit becomes z2+u2=v2z^{2}+u^{2}=v^{2}). See e.g. [DeB05, §2.3] for more detail of this picture.

The nilpotent cone consists of 3 orbits of SL2()\operatorname{SL}_{2}(\mathbb{R}): {0}\{0\}, the half-cone with v>0v>0 (which is the orbit of 𝐞\mathbf{e}), and the half-cone with v<0v<0 (the orbit of 𝐟\mathbf{f}). One can explicitly compute that given a matrix X0:=[z0/2x0y0z0/2]𝔤X_{0}:=\left[\begin{smallmatrix}z_{0}/2&x_{0}\\ y_{0}&-z_{0}/2\end{smallmatrix}\right]\in\mathfrak{g} satisfying z02=4x0y0z_{0}^{2}=-4x_{0}y_{0} (which forces x0y0<0x_{0}y_{0}<0 if we are working over \mathbb{R}), the element g0g_{0} below provides the conjugation so that X0=Adg0(𝐞)X_{0}=Ad^{*}_{g_{0}}(\mathbf{e}) (it is convenient for us to write g0g_{0} as a product of a diagonal and a unipotent matrix with a view toward further calculations):

(63) g0=[x00y01x0]=[x0001x0][10x0y01].g_{0}=\begin{bmatrix}\sqrt{x_{0}}&0\\ \sqrt{-y_{0}}&\frac{1}{\sqrt{x_{0}}}\end{bmatrix}=\begin{bmatrix}\sqrt{x_{0}}&0\\ 0&\frac{1}{\sqrt{x_{0}}}\\ \end{bmatrix}\cdot\begin{bmatrix}1&0\\ \sqrt{-{x_{0}}{y_{0}}}&1\\ \end{bmatrix}.

Note that since v=x0y0>0v=x_{0}-y_{0}>0, and x0y0<0x_{0}y_{0}<0, we have x0>0,y0<0x_{0}>0,y_{0}<0, which explains our choice of signs inside the square roots.

6.3.2. A measure from calculus

Given that our orbit is an open half-cone, we can write down a natural measure on it as a parametrized surface. In fact, as we think of a parametrization for this cone, we can be guided by the fact that we are looking for an SL2()\operatorname{SL}_{2}(\mathbb{R})-invariant measure. We recall Cartan decompositon: SL2()=UK\operatorname{SL}_{2}(\mathbb{R})=UK, where UU is the group of lower-triangular unipotent matrices and K=SO2()S1K=\mathrm{SO}_{2}(\mathbb{R})\simeq S^{1}. The adjoint action of SO2()\mathrm{SO}_{2}(\mathbb{R}) is given by a fairly complicated formula (see [DeB05, §2.3]), but at the same time one has the obvious action of S1S^{1} on the cone by rotations; thus it is reasonable to make a rotation-invariant measure on our cone. Therefore, we use cylindrical coordinates to parametrize the cone z2+u2=v2z^{2}+u^{2}=v^{2} and arrive at z=tcos(θ)z=t\cos(\theta), u=tsin(θ)u=t\sin(\theta), v=tv=t, which translates to the parametrization ρ:(0,)×[0,2π)𝔤\rho:(0,\infty)\times[0,2\pi)\to\mathfrak{g} given by:

(64) ρ(t,θ)=(t(cosθ+1),t(cosθ1),tsinθ).\rho(t,\theta)=(t(\cos\theta+1),t(\cos\theta-1),t\sin\theta).

The natural volume form on the cone is then dtdθdt\wedge d\theta; below we see how it compares to Kirillov’s volume form. Note: now that we made this guess at a form, we could just express the actions of UU and K=SO2()K=\mathrm{SO}_{2}(\mathbb{R}) on the cone in the (t,θ)(t,\theta) coordinates, and check if this form is invariant. However, we prefer to compute the Kirillov’s form directly and derive the comparison this way.

6.3.3. Computing Kirillov’s form

Step 1. The calculation at 𝐟\mathbf{f}^{\ast}. We compute the map Φ𝐟\Phi_{\mathbf{f}^{\ast}} defined in §6.1. It is a map from 𝔤\mathfrak{g} to 𝔤\mathfrak{g}^{\ast}, so given (x,y,z)=x𝐞+y𝐟+z𝐡𝔤(x,y,z)=x\mathbf{e}+y\mathbf{f}+z\mathbf{h}\in\mathfrak{g}, its image under Φ𝐟\Phi_{\mathbf{f}^{\ast}} is a linear functional on 𝔤\mathfrak{g}. Thus it makes sense to write Φ𝐟(x,y,z),(x,y,z)\langle\Phi_{\mathbf{f}^{\ast}}(x,y,z),(x^{\prime},y^{\prime},z^{\prime})\rangle, where (x,y,z)𝔤(x^{\prime},y^{\prime},z^{\prime})\in\mathfrak{g}. We evaluate:

Φ𝐟(x𝐞+y𝐟+z𝐡),(x,y,z)\displaystyle\langle\Phi_{\mathbf{f}^{\ast}}(x\mathbf{e}+y\mathbf{f}+z\mathbf{h}),(x^{\prime},y^{\prime},z^{\prime})\rangle =𝐟,[x𝐞+y𝐟+z𝐡,x𝐞+y𝐟+z𝐡]\displaystyle=\langle\mathbf{f}^{*},[x^{\prime}\mathbf{e}+y^{\prime}\mathbf{f}+z^{\prime}\mathbf{h},x\mathbf{e}+y\mathbf{f}+z\mathbf{h}]\rangle
=𝐟,2(zxxz)𝐞+2(yzzy)𝐟+(xyyx)𝐡\displaystyle=\langle\mathbf{f}^{*},2(z^{\prime}x-x^{\prime}z)\mathbf{e}+2(y^{\prime}z-z^{\prime}y)\mathbf{f}+(x^{\prime}y-y^{\prime}x)\mathbf{h}\rangle
=8(zxxz)\displaystyle=8(z^{\prime}x-x^{\prime}z)
=(2z𝐟x𝐡),(x,y,z).\displaystyle=\langle-(2z\mathbf{f}^{*}-x\mathbf{h}^{*}),(x^{\prime},y^{\prime},z^{\prime})\rangle.

Thus

(65) Φ𝐟(x𝐞+y𝐟+z𝐡)=(2z𝐟x𝐡).\Phi_{\mathbf{f}^{\ast}}(x\mathbf{e}+y\mathbf{f}+z\mathbf{h})=-(2z\mathbf{f}^{*}-x\mathbf{h}^{*}).

Step 2. We need to find a preimage under Φ𝐟\Phi_{\mathbf{f}^{\ast}} for a vector v~T𝐟𝒪𝐟\tilde{v}\in T_{\mathbf{f}^{\ast}}\mathcal{O}_{\mathbf{f}^{\ast}}. We recall that this tangent space is identified with a subspace of 𝔤\mathfrak{g}^{\ast}, which we later plan to identify with a subspace of 𝔤\mathfrak{g}. Because of this latter anticipated identification, we write v~=x𝐟+y𝐞+z𝐡\tilde{v}=x\mathbf{f}^{\ast}+y\mathbf{e}^{\ast}+z\mathbf{h}^{\ast}. We see directly from (65) that for v~\tilde{v} to be in the image of Φ𝐟\Phi_{\mathbf{f}^{\ast}} it has to satisfy y=0y=0, and then v=z𝐞x2𝐡v=z\mathbf{e}-\frac{x}{2}\mathbf{h} satisfies Φ𝐟(v)=v~\Phi_{\mathbf{f}^{\ast}}(v)=\tilde{v}.

We are now ready to compute the form ω𝐟\omega_{\mathbf{f}^{\ast}}. Let vi~=xi𝐟+yi𝐞+zi𝐡\widetilde{v_{i}}=x_{i}\mathbf{f}^{*}+y_{i}\mathbf{e}^{*}+z_{i}\mathbf{h}^{*} for i=1,2i=1,2; then as discussed above, we can take vi=zi𝐞xi2𝐡v_{i}=z_{i}\mathbf{e}-\frac{x_{i}}{2}\mathbf{h}. Then [v1,v2]=(z1x2x1z2)𝐞[v_{1},v_{2}]=(z_{1}x_{2}-x_{1}z_{2})\mathbf{e}, and finally we have that:

(66) ω𝐟(v~1,v~2)=𝐟,[v1,v2]=4(z1x2x1z2).\omega_{\mathbf{f}^{*}}(\tilde{v}_{1},\tilde{v}_{2})=\langle\mathbf{f}^{*},[v_{1},v_{2}]\rangle=4(z_{1}x_{2}-x_{1}z_{2}).

Under the identification of the differential 22-forms on a vector space with alternating 22-tensors, we recognize this form as 4dxdz-4dx^{*}\wedge dz^{*}, which we identify with the form ω𝐞:=4dxdz\omega_{\mathbf{e}}:=-4dx\wedge dz on the adjoint orbit of 𝐞𝔤\mathbf{e}\in\mathfrak{g}.

Step 3. Pullback of ω𝐞\omega_{\mathbf{e}} under the adjoint action. We compute the operator Adg0\operatorname{Ad}_{g_{0}} for the element g0g_{0} from (63) in our coordinates, in order to use it to pull back the form ω𝐞\omega_{\mathbf{e}}. By the right-hand side of (63), the matrix of Adg0\operatorname{Ad}_{g_{0}} in the basis {𝐞,𝐟,𝐡}\{\mathbf{e},\mathbf{f},\mathbf{h}\} is

(67) Adg0=[x00001x00001][100x0y012x0y0x0y001]=[x000y01x02y0x0x0y001].\operatorname{Ad}_{g_{0}}=\begin{bmatrix}x_{0}&0&0\\ 0&\frac{1}{x_{0}}&0\\ 0&0&1\end{bmatrix}\cdot\begin{bmatrix}1&0&0\\ -{x_{0}}{y_{0}}&1&2\sqrt{-{x_{0}}{y_{0}}}\\ \sqrt{-{x_{0}}{y_{0}}}&0&1\end{bmatrix}=\begin{bmatrix}x_{0}&0&0\\ -{y_{0}}&\frac{1}{x_{0}}&2\sqrt{-\frac{y_{0}}{x_{0}}}\\ \sqrt{-{x_{0}}{y_{0}}}&0&1\end{bmatrix}.

Thus,

(68) (Adg0)(dxdz)=|x00x0y00|dxdy|x00x0y01|dxdz+|0001|dydz\displaystyle(\operatorname{Ad}_{g_{0}})^{\ast}(dx\wedge dz)=\left|\begin{matrix}x_{0}&0\\ \sqrt{-x_{0}y_{0}}&0\end{matrix}\right|dx\wedge dy-\left|\begin{matrix}x_{0}&0\\ \sqrt{-x_{0}y_{0}}&1\end{matrix}\right|dx\wedge dz+\left|\begin{matrix}0&0\\ 0&1\end{matrix}\right|dy\wedge dz
=x0dxdz.\displaystyle=x_{0}dx\wedge dz.

By definition, the volume form at X0𝔤X_{0}\in\mathfrak{g} is (Adg01)(ω𝐞)(\operatorname{Ad}_{g_{0}^{-1}})^{\ast}(\omega_{\mathbf{e}}), and thus we get

ωX0=1x0dxdz.\omega_{X_{0}}=\frac{1}{x_{0}}dx\wedge dz.

Coverting to (t,θ)(t,\theta)-coordinates, we get:

ρ(dxdz)\displaystyle\rho^{*}(dx\wedge dz) =((cosθ+1)dttsinθdθ)(sinθdt+tcosθdθ)\displaystyle=((\cos\theta+1)dt-t\sin\theta\ d\theta)\wedge(\sin\theta\ dt+t\cos\theta\ d\theta)
=tcosθ(cosθ+1)dtdθtsin2θdθdt\displaystyle=t\cos\theta(\cos\theta+1)dt\wedge d\theta-t\sin^{2}\theta d\theta\wedge dt
=(tcos2θ+tcosθ+tsin2θ)dtdθ\displaystyle=(t\cos^{2}\theta+t\cos\theta+t\sin^{2}\theta)dt\wedge d\theta
=t(1+cosθ)dtdθ\displaystyle=t(1+\cos\theta)dt\wedge d\theta

and therefore

ρω\displaystyle\rho^{*}\omega =ρ(4dxdzx)=4t(1+cosθ)t(1+cosθ)dtdθ\displaystyle=\rho^{*}\left(\frac{4dx\wedge dz}{x}\right)=\frac{4t(1+\cos\theta)}{t(1+\cos\theta)}dt\wedge d\theta
=4dtdθ.\displaystyle=4\ dt\wedge d\theta.

6.3.4. Semisimple orbits in 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R})

The orbits of split semisimple elements in 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R}) are hyperboloids of one sheet asymptotically approaching the nilpotent cone on the outside; the orbits of elliptic elements are the individual sheets of hyperboloids of two sheets that lie inside the same asymptotic cone (see e.g., [DeB05, §2.3.3] for detail). Measures on them can be computed in a similar way (we return to this calculation below). For now we compute another example, a semisimple orbit in 𝔰𝔬3()\mathfrak{so}_{3}(\mathbb{R}).

6.4. Another example: a semisimple (elliptic) element in 𝔰𝔬3()\mathfrak{so}_{3}(\mathbb{R})

Let G=SO(3)G=\mathrm{SO}(3), 𝔤=𝔰𝔬(3)\mathfrak{g}=\mathfrak{so}(3), and let {X,H,Y}\{X,H,Y\} be the basis for 𝔤\mathfrak{g} given by:

X=(000001010)H=(001000100)Y=(010100000)X=\begin{pmatrix}0&0&0\\ 0&0&-1\\ 0&1&0\end{pmatrix}\ \ H=\begin{pmatrix}0&0&1\\ 0&0&0\\ -1&0&0\end{pmatrix}\ \ Y=\begin{pmatrix}0&-1&0\\ 1&0&0\\ 0&0&0\end{pmatrix}

Let {X,H,Y}\{X^{\ast},H^{\ast},Y^{\ast}\} be dual basis for 𝔤\mathfrak{g}^{*} under the Killing form. Explicitly this means X(X)=2X^{*}(X)=-2, Y(Y)=2Y^{*}(Y)=-2, and H(H)=2H^{\ast}(H)=-2. Denote by 𝒪H\mathcal{O}_{H^{\ast}} the co-adoint orbit of HH^{\ast}. Then a brief calculation shows that 𝒪H\mathcal{O}_{H^{\ast}} is the unit sphere:

𝒪H={(x,y,z)𝔤|x2+y2+z2=1}\mathcal{O}_{H^{\ast}}=\{(x,y,z)\in\mathfrak{g}^{*}\ |\ x^{2}+y^{2}+z^{2}=1\}

We compute Kirillov’s form on this sphere, using a slightly different method from the above (to illustrate various approaches to such computations). Namely, rather than computing the form at one fixed base point on the orbit and then using the group action to compute it at all points, we do the computation directly for each point X0X_{0} of our orbit 𝒪H\mathcal{O}_{H^{\ast}}.

As above, given X0𝒪HX_{0}\in\mathcal{O}_{H^{\ast}}, we have φX0:G𝔤\varphi_{X_{0}}:G\to\mathfrak{g}^{*} given by φX0(g)=Adg(X0)\varphi_{X_{0}}(g)=Ad^{*}_{g}(X_{0}). We write X0=(x0,y0,z0)X_{0}=(x_{0},y_{0},z_{0}) in {X,H,Y}\{X^{\ast},H^{\ast},Y^{\ast}\}-coordinates. As above, the differential of φX0\varphi_{X_{0}} at eGe\in G gives an identification ΦX0:𝔤/𝔠(X)TX0𝒪HTX0𝔤𝔤\Phi_{X_{0}}:\mathfrak{g}/\mathfrak{c}(X)\to T_{X_{0}}\mathcal{O}_{H^{\ast}}\hookrightarrow{}T_{X_{0}}\mathfrak{g}^{*}\cong\mathfrak{g}^{*}. This can be computed either by recalling that ΦX0(X),Y=X0,[Y,X]\langle\Phi_{X_{0}}(X),Y\rangle=\langle X_{0},[Y,X]\rangle or via the exponential map:

ΦX0(X)=ddt|t=0(φX0exp)(tX).\displaystyle\Phi_{X_{0}}(X)=\frac{d}{dt}\big{|}_{t=0}(\varphi_{X_{0}}\circ\exp)(tX).

The result of this computation is that with respect to our coordinates, the matrix representation for ΦX0\Phi_{X_{0}} is given by:

(0z0y0z00x0y0x00).\begin{pmatrix}0&z_{0}&-y_{0}\\ -z_{0}&0&x_{0}\\ y_{0}&-x_{0}&0\end{pmatrix}.

Write ω=f1dXdH+f2dXdY+f3dHdY\omega=f_{1}dX^{*}\wedge dH^{*}+f_{2}dX^{*}\wedge dY^{*}+f_{3}dH^{*}\wedge dY^{*}. We have:

f1(X0)=ωX0(X,H)=X0,[(0,z0,y0),(z0,0,x0)]=2(x02z0+y02z0+z03)=2z0\displaystyle f_{1}(X_{0})=\omega_{X_{0}}(\frac{\partial}{\partial X^{*}},\frac{\partial}{\partial H^{*}})=\langle X_{0},[(0,z_{0},-y_{0}),(-z_{0},0,x_{0})]\rangle=-2(x_{0}^{2}z_{0}+y_{0}^{2}z_{0}+z_{0}^{3})=-2z_{0}
f2(X0)=ωX0(X,Y)=X0,[(0,z0,y0),(y0,x0,0)]=2(x02y0+y03+y0z02)=2y0\displaystyle f_{2}(X_{0})=\omega_{X_{0}}(\frac{\partial}{\partial X^{*}},\frac{\partial}{\partial Y^{*}})=\langle X_{0},[(0,z_{0},-y_{0}),(y_{0},-x_{0},0)]\rangle=2(x_{0}^{2}y_{0}+y_{0}^{3}+y_{0}z_{0}^{2})=2y_{0}
f3(X0)=ωX0(H,Y)=X0,[(z0,0,x0),(y0,x0,0)]=2(x03+x0y02+x0z02)=2x0.\displaystyle f_{3}(X_{0})=\omega_{X_{0}}(\frac{\partial}{\partial H^{*}},\frac{\partial}{\partial Y^{*}})=\langle X_{0},[(-z_{0},0,x_{0}),(y_{0},-x_{0},0)]\rangle=-2(x_{0}^{3}+x_{0}y_{0}^{2}+x_{0}z_{0}^{2})=-2x_{0}.

Therefore,

(69) ω(x,y,z)\displaystyle\omega(x,y,z) =2xdHdY+2ydXdY2zdXdH\displaystyle=-2x\ dH^{*}\wedge dY^{*}+2y\ dX^{*}\wedge dY^{*}-2z\ dX^{*}\wedge dH^{*}
=2(xdHdYydXdY+zdXdH).\displaystyle=-2(x\ dH^{*}\wedge dY^{*}-y\ dX^{*}\wedge dY^{*}+z\ dX^{*}\wedge dH^{*}).

It is easy to check that if we parametrize the sphere using the spherical coordinates, this form is rewritten as twice the usual surface area element: ω(φ,θ)=2sinφ\omega(\varphi,\theta)=2\sin\varphi. We leave this check as an exercise.

6.4.1. General semi-simple orbits in 𝔰𝔬3(){\mathfrak{so}}_{3}(\mathbb{R}) and 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R})

Since the group SO3()\mathrm{SO}_{3}(\mathbb{R}) is compact, all its maximal tori are conjugate; consequently, every semisimple element in 𝔰𝔬3(){\mathfrak{so}}_{3}(\mathbb{R}) is conjugate to rHrH for some rr\in\mathbb{R} (and all semisimple orbits are spheres). It is clear that if we replace HH^{\ast} with rHrH^{\ast}, the form in (69) gets scaled by rr: ωrH=2rdxdz\omega_{rH^{\ast}}=2rdx\wedge dz, so it is again the natural area element on a sphere of radius rr.

We also note that all our calculations for these algebraic volume forms are valid over any field of characteristic different from 22 (the only reason we were working over the reals is the nice geometric picture and the intuitive parametric equations for the orbits as surfaces; note that despite our use of these transcendental parametrizations, in the end all the differential forms are algebraic).

Returning to semi-simple orbits in 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{R}), we can use the accidental isomorphism: over \mathbb{C}, 𝔰𝔩2\mathfrak{sl}_{2} and 𝔰03\mathfrak{s0}_{3} are isomorphic. Thus, the same calculation as above shows also that the value at the element t𝐡t\mathbf{h} of the Kirillov form on the orbit of t𝐡SL2()t{\mathbf{h}}\in\operatorname{SL}_{2}(\mathbb{R}) is 2tdxdy2tdx\wedge dy (note that the yy- and zz-coordinates are swapped in §6.3 and §6.4); this uniquely determines the invariant form on the orbit.

How does this form relate to the volume form ωcgeom\omega_{c}^{\operatorname{geom}} defined by (23) in §3.4? Using the coordinates of 6.3, Chevalley map is given by

[z/2xyz/2]z24xy.\left[\begin{smallmatrix}z/2&x\\ y&-z/2\end{smallmatrix}\right]\mapsto-\frac{z^{2}}{4}-xy.

The geometric measure is defined as a quotient: dxdy12dz=ωcgeomdcdx\wedge dy\wedge{\frac{1}{2}dz}=\omega_{c}^{\operatorname{geom}}\wedge dc, where c=z24xyc=-\frac{z^{2}}{4}-xy. Evaluating all the forms at the point t𝐡t\mathbf{h} (which corresponds to x=y=0x=y=0, z=2tz=2t, and thus c=t2c=-t^{2}), we see that ωcgeom\omega_{c}^{\operatorname{geom}} must satisfy

(ωcgeom)t𝐡(2tdt)=dxdydt,(\omega_{c}^{\operatorname{geom}})_{t\mathbf{h}}\wedge(-2tdt)=dx\wedge dy\wedge dt,

and therefore, (ωcgeom)t𝐡=12tdxdy(\omega_{c}^{\operatorname{geom}})_{t\mathbf{h}}=-\frac{1}{2t}dx\wedge dy. We obtain the conversion coefficient between Kirillov’s form and the geometric form: on the orbit of a split semi-simple element t𝐡t\mathbf{h} it is 14t2=D(t𝐡)1-\frac{1}{4t^{2}}=-D(t\mathbf{h})^{-1}. It would be interesting to find this coefficient for a general reductive Lie algebra.

References

  • [AAG+19] Jeffrey D. Achter, Salim Ali Altug, Luis Garcia, Julia Gordon, and with Appendix by Thomas Rüd and Wen-Wei Li, Counting Abelian varieties over finite fields via Frobenius densities, https://arxiv.org/abs/1905.11603 (2019).
  • [AW15] Jeffrey Achter and Cassandra Williams, Local heuristics and an exact formula for abelian surfaces over finite fields, Canad. Math. Bull. 58 (2015), no. 4, 673–691. MR 3415659
  • [Bat99] Victor V. Batyrev, Birational Calabi-Yau nn-folds have equal Betti numbers, New trends in algebraic geometry (Warwick, 1996), London Math. Soc. Lecture Note Ser., vol. 264, Cambridge Univ. Press, Cambridge, 1999, pp. 1–11. MR 1714818 (2000i:14059)
  • [Bit11] Rony A. Bitan, The discriminant of an algebraic torus, J. Number Theory 131 (2011), no. 9, 1657–1671.
  • [BLR80] Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud, Néron models, Springer-Verlag, Berlin, 1980.
  • [Bou85] N. Bourbaki, Eléménts de mathématique. algèbre commutative, Masson, 1985.
  • [Bou02] by same author, Lie groups and lie algebras. chapters 4-6., Springer-Verlag, 2002.
  • [CL08] Raf Cluckers and François Loeser, Constructible motivic functions and motivic integration, Invent. Math. 173 (2008), no. 1, 23–121.
  • [DeB05] Stephen DeBacker, Homogeneity for reductive pp-adic groups: an introduction, Harmonic analysis, the trace formula, and Shimura varieties, Clay Math. Proc., vol. 4, Amer. Math. Soc., Providence, RI, 2005, pp. 393–522. MR 2192014 (2006m:22016)
  • [DKS] Chantal David, Dimitris Koukoulopoulos, and Ethan Smith, Sums of Euler products and statistics of elliptic curves.
  • [DL01] Jan Denef and François Loeser, Definable sets, motives, and pp-adic integrals, J. Amer. Math. Soc. 14 (2001), no. 2, 429–469.
  • [FH91] William Fulton and Joe Harris, Representation theory: A first course, Graduate Texts in Mathematics, vol. 129, Springer, New York, 1991.
  • [FLN10] Edward Frenkel, Robert Langlands, and Báo Châu Ngô, Formule des traces et fonctorialité: le début d’un programme, Ann. Sci. Math. Québec 34 (2010), no. 2, 199–243. MR 2779866 (2012c:11240)
  • [FT91] A. Fröhlich and M.J. Taylor, Algebraic number theory, Cambridge studies in advanced mathematics, vol. 27, Cambridge University Press, 1991.
  • [Gek03] Ernst-Ulrich Gekeler, Frobenius distributions of elliptic curves over finite prime fields, Int. Math. Res. Not. (2003), no. 37, 1999–2018.
  • [GG99] Wee Teck Gan and Benedict H. Gross, Haar measure and the Artin conductor, Transactions of the AMS 351 (1999), no. 4, 1691–1704.
  • [GGPS16] I.M. Gelfand, M.I. Graev, and I. Pyatetskii-Shapiro, Generalized functions: volume 6. representation theory and automorphic functions, AMS Chelsea publishing, 2016, Originaly published in Russian, 1958.
  • [Gro97] Benedict H. Gross, On the motive of a reductive group, Invent. Math. 130 (1997), 287–313.
  • [Hum72] James E. Humphreys, Introduction to Lie algebras and representation theory, Graduate Texts in Mathematics, Springer-Verlag, 1972.
  • [KL06] Andrew Knightly and Charles Li, Traces of Hecke operators, Mathematica surveys and monographs, vol. 133, American Math. Society, 2006.
  • [Kot82] Robert E. Kottwitz, Rational conjugacy classes in reductive groups, Duke Math. J. 49 (1982), no. 4, 785–806. MR 683003 (84k:20020)
  • [Kot97] Robert Kottwitz, Isocrystals with additional structure. II, Compositio Math. 109 (1997), no. 3, 255–339.
  • [Kot05] Robert E. Kottwitz, Harmonic analysis on reductive pp-adic groups and Lie algebras, Harmonic analysis, the trace formula, and Shimura varieties, Clay Math. Proc., vol. 4, Amer. Math. Soc., Providence, RI, 2005, pp. 393–522. MR 2192014 (2006m:22016)
  • [Lan13] Robert P. Langlands, Singularités et transfert, Ann. Math. Qué. 37 (2013), no. 2, 173–253. MR 3117742
  • [Mil08] James S. Milne, Class field theory, Available at http://jmilne.org, 2008.
  • [Oes82] Joseph Oesterlé, Réduction modulo pnp^{n} des sous-ensembles analytiques fermés de ZpN{{Z}}^{N}_{p}, Invent. Math. 66 (1982), no. 2, 325–341.
  • [Ono61] Takashi Ono, Arithmetic of algebraic tori, Ann. of Math. (2) 74 (1961), 101–139.
  • [Ono63] by same author, On the Tamagawa number of algebraic tori, Ann. of Math. (2) 78 (1963), 47–73. MR 0156851
  • [Ono66] by same author, On Tamagawa numbers, Arithmetic properties of algebraic groups. Adèle groups., Proc. Symp. Pure Math., vol. 9, II, Amer. Math. Soc., Providence, RI, 1966, pp. 122–132.
  • [PR91] Vladimir Platonov and Andrei Rapinchuk, Algebraic groups and number theory, Academic Press, Inc., San Diego, 1991.
  • [Rüd20] Thomas Rüd, Explicit Tamagawa numbers for certain tori over number fields, https://arxiv.org/abs/2009.04431 (2020).
  • [Ser81] Jean-Pierre Serre, Quelques applications du théorème de densité de Chebotarev, Inst. Hautes Études Sci. Publ. Math. (1981), no. 54, 323–401.
  • [Shy77] Jih Min Shyr, On some class number relations of algebraic tori, Michigan Math. J. 24 (1977), no. 3, 365–377. MR 0491596
  • [Vey92] W. Veys, Reduction modulo pnp^{n} of pp-adic subanalytic sets, Math.Proc. Cambridge Philos. Soc. 112 (1992), 483–486.
  • [Wei82] André Weil, Adeles and algebraic groups, Progress in mathematics, vol. 23, Birkhäuser, 1982.
  • [Yun13] Zhiwei Yun, Orbital integrals and Dedekind zeta funcions, The Legacy of Srinivasa Ramanujan, RMS-Lecture notes series (2-13), no. 20, 399–420.