1 March, 2005
1991 Mathematics Subject Classification. 22E35, 22E50.
The local character expansion near a tame, semisimple element
  
 
Jeffrey D. Adler 
 Jonathan Korman
 E-mail address : adler@uakron.edu  The University of Akron, Akron, OH 44325-4002  E-mail address : jkorman@math.toronto.edu  The University of Toronto, Toronto, Ontario M5S 3G3 
- 
 
 Abstract.
 Consider the character of an irreducible admissible representation of a 
 
-adic reductive group. The Harish-Chandra-Howe local expansion expresses this character near a semisimple element as a linear combination of Fourier transforms of nilpotent orbital integrals. Under mild hypotheses, we describe an explicit region on which the local character expansion is valid. 
 0  Introduction
 Let 
 
 denote the group of 
 
-points of a reductive 
 
-group 
 
, where 
 
 is a nonarchimedean local field. To simplify the present discussion, assume that 
 
is connected and that 
 
 has characteristic zero. Let 
 
denote an irreducible admissible representation of 
 
. Let 
 
 denote a fixed Haar measure on 
 
. The distribution character  
 
 of 
 
 is the map 
 
 given by 
 
, where 
 
is the (finite-rank) operator on  
 
 given by 
 
. 
From Howe  [12] and Harish-Chandra  [9] , the distribution  
 
 is represented by a locally constant function on the set of regular semisimple elements in 
 
. We will denote this function also by  
 
. 
For any semisimple 
 
, the local character expansion about 
 
 (see  [11] and [10] ) is the identity  
 
 valid for all regular semisimple  
 
 in the Lie algebra 
 
 of the centralizer of 
 
 such that  
 
 is close enough to 
 
. Here, the sum is over the set of nilpotent orbits 
 
 in 
 
;  
 
 is the function that represents the distribution that is the Fourier transform of the orbital integral  
 
 associated to 
 
; 
 
; and  
 
 is the exponential map, or some suitable substitute. 
This is a qualitative result, in the sense that it gives no indication of how close  
 
 must be to 
 
in order for the identity to be valid. Many questions in harmonic analysis on 
 
 require more quantitative versions of such qualitative results. As an example of a quantitative result, DeBacker  [6] has determined (under some hypotheses on 
 
) a neighborhood of validity for the local character expansion near the identity, thus verifying a conjecture of Hales, Moy, and Prasad (see  [14] ). 
In this paper, we generalize DeBacker's result for any semisimple 
 
 satisfying mild tameness hypotheses. (See § 7 for the hypotheses, and Corollary   11.10 for a precise statement of the main result.) When 
 
 is regular, we recover a generalization of Theorem  19 of [13] . 
We have taken care not to assume that 
 
 is connected. For example, 
 
 could be a semidirect product 
 
 for some 
 
of finite order. This case is of particular interest, since understanding the 
 
-twisted characters of  
 
 near the identity is equivalent to understanding the characters of 
 
 near 
 
. 
Note that Theorem  2.1.5(3) of  [6] plays a key role in our proof. 
 Acknowledgements. We thank Stephen DeBacker, Fiona Murnaghan, and Ju-Lee Kim for helpful conversations. 
 1  Notation and conventions
 Let 
 
 denote nonarchimedean local field, and let 
 
 denote a discrete valuation on 
 
. For any algebraic extension field 
 
 of 
 
, 
 
 extends uniquely to a valuation (also denoted 
 
) of 
 
. Fix a complex-valued, additive character 
 
on 
 
 that is nontrivial on the ring 
 
 of integers in 
 
 and trivial on the prime ideal of 
 
. 
For a reductive 
 
-group 
 
, let  
 
 denote its connected part, and let  
 
 denote its Lie algebra. Let  
 
 denote the dual of  
 
. Let 
 
, the group of 
 
-rational points of 
 
; and let 
 
and  
 
. Let  
 
 denote the center of 
 
. 
We use similar notation and font conventions for other groups. That is, given a group 
 
, we have  
 
,  
 
, etc. 
Let  
 
(resp.   
 
) denote the adjoint or coadjoint representation of 
 
 (resp.   
 
) on  
 
 or  
 
. Let  
 
denote the conjugation action of 
 
 on itself. For an element or subset 
 
 in 
 
 and an element or subset 
 
 in 
 
 or  
 
 (resp.  
 
), we will sometimes write 
 
 instead of 
 
 (resp. 
 
). 
An element 
 
 is semisimple if 
 
is a semisimple linear transformation of 
 
. When  
 
, this is equivalent to 
 
 belonging to a torus. For a subset 
 
of 
 
, let  
 
 denote the set of semisimple elements in 
 
 (so  
 
). An element 
 
 is regular semisimple if the coefficient of  
 
 in 
 
is nonzero (where 
 
 is the rank of the component  
 
 in 
 
; see  [5] ). We denote the set of regular semisimple elements in 
 
 by  
 
. Similarly we say that an element 
 
 is regular semisimple if the coefficient of  
 
 in 
 
is nonzero. 
We denote the set of regular semisimple elements in 
 
 by  
 
. 
For a subset 
 
 of 
 
 (resp.  
 
) let 
 
denote the characteristic function of 
 
 on 
 
(resp.  
 
). 
Call an element 
 
 nilpotent if there is some one-parameter subgroup 
 
 of 
 
 defined over 
 
 such that 
 
. Let  
 
 denote the set of nilpotent elements in 
 
, and 
 
the set of nilpotent orbits under the adjoint action of 
 
on  
 
. We will leave out the subscript when it is understood. One can similarly define a set  
 
 of nilpotent elements in  
 
. 
For any compact group 
 
, let  
 
 denote the set of equivalence classes of irreducible, continuous representations of 
 
. We will not always distinguish between a representation and its equivalence class. Recall that if 
 
 is abelian, then  
 
 is a group. 
Let  
 
 and extend the ordering on 
 
 to one on  
 
 as follows: 
for all 
 
, 
 |  |  | 
 |  |  | 
 |  |  | 
 
  
 If 
 
, define  
 
to be  
 
. There is a natural way to extend the additive structure on 
 
 to an additive structure on 
 
. 
 2  Apartments and buildings.
 For any extension 
 
 of finite ramification degree, let 
 
denote the extended Bruhat-Tits building of 
 
 over 
 
. Note that if 
 
 is Galois, then 
 
embeds naturally in the set of 
 
-fixed points of 
 
, with equality when 
 
 is tame (see  [16,(5.11)] ). 
Every maximal 
 
-split torus 
 
 in 
 
 has an associated apartment 
 
in 
 
. Let 
 
 be a maximal 
 
-torus in 
 
 containing 
 
. Then 
 
 splits over some Galois extension 
 
, so 
 
has an apartment 
 
in 
 
. The Galois fixed point set of the apartment of 
 
 in 
 
is the apartment of 
 
 in 
 
 [20,§2.6] . 
Suppose 
 
 is an 
 
-Levi 
 
-subgroup (that is, 
 
is a Levi subgroup of 
 
) for some finite Galois extension 
 
. There is a natural family of 
 
-equivariant embeddings of 
 
into 
 
. When 
 
 is tame, this in turn induces a family of embeddings of 
 
into 
 
. In general, there is no canonical way to pick a distinguished member of this family. However, all such embeddings share the same image, and no statement we make will depend on the choice of embedding. 
More generally, suppose that 
 
 is the centralizer of some  
 
. Then  
 
 for some 
 
. Let 
 
 denote the connected part of the centralizer of  
 
. Then 
 
 is an 
 
-Levi 
 
-subgroup of 
 
 for some 
 
, and 
 
 acts on 
 
 via an automorphism of order dividing 
 
. For every extension 
 
 of finite residue degree, 
 
is the direct product of the affine space 
 
and the reduced building 
 
. Since  
 
, 
 
 acts on 
 
via an automorphism of order dividing 
 
, and on 
 
via an affine transformation whose 
 
th power is a translation. Thus, for some translation 
 
 of 
 
, 
 
 acts on 
 
via an automorphism of order dividing 
 
. 
If 
 
 is a tame Galois action on 
 
 (see  [19] or [16] ), or the residue characteristic of 
 
 does not divide 
 
 (see  [17] ), then we may identify 
 
with the set of 
 
-fixed points in 
 
. If in addition 
 
 is tame, then (again) we have a family of embeddings of 
 
into 
 
. Again, no statement we make will depend on the choice of embedding. 
 3  Moy-Prasad filtrations.
 For any 
 
-torus 
 
 in 
 
, let 
 
denote the absolute root system of 
 
with respect to 
 
. We can also interpret 
 
as the set of nontrivial eigencharacters for the adjoint action of 
 
 on  
 
. 
When 
 
 is maximal, let 
 
denote the set of affine roots of 
 
 with respect to 
 
 and 
 
. If 
 
, let 
 
denote the gradient of 
 
, and let 
 
denote the root space corresponding to 
 
. We denote the root lattice in  
 
 corresponding to 
 
 by  
 
 [14,3.2] . 
Let 
 
denote the lattice of characters of 
 
, and let  
 
 denote the parahoric subgroup of 
 
. For  
 
, define 
 
 and for 
 
, 
 
 For each  
 
, Moy and Prasad define lattices  
 
 in 
 
and  
 
 in  
 
. When 
 
, they define a normal subgroup  
 
 of the parahoric subgroup  
 
 of 
 
. In particular, for all 
 
and  
 
,  
 |  | (3.1) | 
Similarly,  
 
 is defined in terms of the filtrations on 
 
and on root groups. 
These definitions depend on the normalization of the valuation 
 
; our normalization agrees with that of Yu  [21] . Thus, for example, for any  
 
,  
 
. 
However, the definitions do not depend on the choice of 
 
 containing 
 
 in its apartment. Note that for all  
 
,  
 
 and  
 
. 
Moy and Prasad also define  
 
 and  
 
 (irrespective of whether or not 
 
 is 
 
-split). The above normalization was chosen to have the following property  [1,1.4.1] : when 
 
 is tame and 
 
, we have 
 
 
 We will also use the following notation. For  
 
, let 
 
 
 It is proven in  [2] that  
 
 (resp.   
 
) is a 
 
-domain: a 
 
-invariant, open and closed subset of 
 
 (resp.  
 
). 
For any 
 
and any 
 
, the group 
 
is abelian. 
Under many conditions (for example, if 
 
 contains a tamely ramified maximal torus, or if 
 
 is simply connected), there exists a (  
 
-equivariant) isomorphism (see  [14] or  [21] )  
 |  | (3.2) | 
and thus an isomorphism  
 |  | (3.3) | 
Yu  [22] has defined a more complicated filtration on 
 
 than the one above. Using this filtration to define  
 
, he shows that  3.3 is valid for all 
 
. However, for the groups that we will consider, Yu's filtration is equivalent to the one above. 
 4  Singular depth
 From now on, 
 
 is a reductive 
 
-group,  
 
, and 
 
 is the centralizer of 
 
in 
 
.
 Definition 4.1. 
For  
 
, let 
 
where  
 
 is the set of generalized eigenvalues of the action of 
 
 on 
 
. 
Remark 4.2. 
If 
 
is connected and 
 
 is regular, then the definition of 
 
 given above agrees with the definition in  [
13,§1] 
.  
Remark 4.3. 
 Note that 
 
 for all  
 
 and that 
 
 for all  
 
. 
Remark 4.4. 
 Suppose  
 
, and 
 
 is an extension that contains all of the generalized eigenvalues of both 
 
 and 
 
 acting on 
 
. Since 
 
 and 
 
 commute, we can write 
 
 as a direct sum of subspaces  
 
, where  
 
 is simultaneously an  
 
-eigenspace for 
 
 and a generalized  
 
-eigenspace for 
 
. 
 Lemma 4.5. 
 If  
 
, then 
 
. 
- 
 
Proof.
Pick a maximal 
 
-torus 
 
 in 
 
 with  
 
, and a splitting field 
 
 for 
 
. 
Then  
 
 where the last equality follows from Theorem  4.1.5(1) of  [7] . Thus, 
 
for all 
 
. In particular, this is true for all 
 
. □ 
 
 Lemma 4.6. 
 If  
 
, then 
 
. If 
 
 is compact mod  
 
, then so is 
 
, and conversely. If 
 
 is semisimple, then so is 
 
. 
 
- 
 
Proof.
Let  
 
,  
 
, and  
 
 be as in Remark   4.4 . For all 
 
, Lemma   4.5 implies that 
 
, and thus  
 
 Thus, 
 
. The second statement of the lemma follows from the fact that  
 
 is a unit if and if so is  
 
. If 
 
is diagonalizable over some field extension, then 
 
and 
 
are simultaneously diagonalizable, so the last statement follows. □ 
 
 For  
 
, following  [10,§18] , define  
 
 (When  
 
, define 
 
.) For 
 
let 
 |  |  | 
 |  |  | 
 
  
 Note that  
 
, and these are open, dense subsets of  
 
. 
 Corollary 4.7. 
   
 
. 
 
- 
 
Proof.
Let  
 
. Let  
 
,  
 
, and  
 
 be as in Remark   4.4 . As in the proof of Lemma   4.6 , 
 
for all 
 
. Thus,  
 
 Therefore,  
 
. □ 
 
 
 5  Intertwining
 Definition 5.1. 
Let 
 
 be a compact open subgroup of 
 
 and let  
 
. For 
 
, recall that 
 
 is the representation of  
 
 given as 
 
. 
Definition 5.2. 
If 
 
 and  
 
 are continuous representations of compact subgroups 
 
 and  
 
 (respectively) of 
 
, then let 
 
. 
 
 Lemma 5.3. 
 Let 
 
 and 
 
 be compact subgroups of 
 
, and let  
 
 be a compact subgroup of 
 
. Let  
 
 and let 
 
 denote a one-dimensional representation of 
 
. Let  
 
 be a decomposition of 
 
 into representations of  
 
. If 
 
 then for some  
 
, 
 
. 
- 
 
Proof.
We have 
 
. Therefore, 
 
for some 
 
. □ 
 
 
 6  Partial traces
 From now on, let 
 
denote an irreducible admissible representation of 
 
. 
Let  
 
 denote the distribution character of 
 
. This distribution is represented by a locally constant function (also denoted  
 
) on  
 
. Let 
 
denote the depth of 
 
 [14,§5] . 
For any irreducible representation 
 
 of a compact open subgroup 
 
, let  
 
 denote the 
 
-isotypic subspace of  
 
. Let  
 
 denote the 
 
-equivariant projection from  
 
 to  
 
. Define the distribution  
 
 by 
 
for all 
 
. 
Then  
 
, which can be thought of as the `partial trace of 
 
 with respect to 
 
', is represented by the locally constant function 
 
on 
 
. It follows from the definitions that 
 
 |  | (6.1) | 
Note that for each fixed  
 
, 
 
has finite rank, so all but finitely many terms in this sum vanish.
 Lemma 6.2. 
  
- 
 
(1)
If 
 
 and 
 
, then 
 
. 
-  
(2)
If 
 
, then 
 
. 
 
 
- 
 
Proof.
The first statement is  [13,Lemma14] . Since  
 
, we have  
 
 
 Let  
 
 be a compact open subgroup of 
 
, and let  
 
. Considered as a representation of  
 
, 
 
 decomposes into a finite sum of distinct irreducible representations  
 
 with multiplicities  
 
: 
 
 For each 
 
, let  
 
 denote the 
 
-isotypic subspace of  
 
. Let  
 
 denote the  
 
-equivariant projection from  
 
 to  
 
. For 
 
, define 
 
 Remark 6.3. 
 Note that 
 
. Moreover,  
 
 has invariance properties analogous to those given for  
 
 in Lemma   6.2 .  
 
 Proposition 6.4. 
 Let 
 
. If 
 
, then 
 
. If 
 
, then 
 
. 
 
- 
 
Proof.
Define a pairing 
 
 on  
 
 with respect to which 
 
 is unitary. Let 
 
 be a basis for  
 
. For 
 
, 
 |  |  |  
 |  |  |  
 |  |  |  
 
 Fixing  
 
 and letting  
 
 vary, we have a sum of matrix coefficients of 
 
. Note that by Lemma   6.2 (1) we have,
 |  |  |  
 |  |  |  
 
 Fixing  
 
 and letting  
 
 vary, we again have a sum of matrix coefficients of 
 
. 
Therefore, our first statement follows from Corollary  14.3 of [10] . 
To prove the second statement, note that since  
 
 is a projection onto a subspace of  
 
, if 
 
, then 
 
. □
 
 Lemma 6.5. 
 Fix 
 
. Let 
 
 be a character of a closed subgroup  
 
 of  
 
. Let 
 
. Suppose  
 
 Then  
 
. 
 
- 
 
Proof.
For all  
 
,  
 
 acts on  
 
 via the scalar 
 
. Let 
 
 be the multi-set of diagonal entries of a matrix that represents, with respect to some basis, the action of 
 
on  
 
. Then for some 
 
,  
 
 The conclusion then follows from Lemma  14.2 of  [10] . □ 
 
 
 7  Hypotheses
 From now on, we will make certain assumptions: Let 
 
, 
 
, and 
 
 be as in § 4 . 
That is, 
 
 is a reductive linear algebraic 
 
-group,  
 
, and 
 
 is the centralizer of 
 
 in 
 
. 
 Hypothesis 7.1. 
 Assume that the eigenvalues of 
 
 belong to a tamely ramified extension of 
 
. 
 When  
 
, this hypothesis implies that 
 
 is an 
 
-Levi subgroup for some tame 
 
. 
 Hypothesis 7.2. 
 Assume that 
 
 embeds in 
 
, as in § 2 , and 
 |  |  | 
 |  |  | 
 
  
  
 We will pursue elsewhere the question of when this hypothesis holds. For now, we note that, given Hypothesis   7.1 , it holds when  
 
. Additionally, it holds when 
 
 is a finite, tamely ramified Galois extension, 
 
for some 
 
-group 
 
 (where  
 
 denotes restriction of scalars), and 
 
. 
(See Lemmas  2.2.3 and  2.2.9 of  [3] .) 
 Hypothesis 7.3. 
 Suppose the order of  
 
 is prime to the residue characteristic of 
 
. 
 We will need this hypothesis in order to prove Lemma   8.1 . However, there are other conditions that would imply this result. 
 Hypothesis 7.4. 
 There is a nondegenerate 
 
-invariant symmetric bilinear form 
 
 on 
 
 such that we can identify  
 
 with  
 
 via the map  
 
 defined by 
 
. 
 (Groups satisfying this hypothesis are discussed in  [4,§4] .) Thus, we can (and eventually will) identify  
 
 with  
 
 and  
 
 with  
 
 The following hypothesis concerns the existence of a “mock exponential” map. 
 Hypothesis 7.5. 
 Let 
 
. There exists a bijection  
 
 such that for all 
 
 and all 
 
, we have that  
 
 induces the group isomorphism  
 
 of  3.2 . 
Moreover, for all 
 
,  
 
, and 
- 
 
(1)
for 
 
, all  
 
, and all  
 
, we have 
 
 modulo  
 
; 
-  
(2)
for all  
 
 we have 
 
; 
-  
(3)
for all 
 
, all  
 
 with 
 
, all  
 
, and all   
 
, we have  
 
. 
 
 
 Note that item ( 3 ) in the hypothesis asserts, for all 
 
, a weaker version of Proposition  1.6.3 of  [1] , and the remainder of the hypothesis is a weaker version of Hypothesis  3.2.1 in [6] . Item ( 1 ) implies that  
 
 carries a Haar measure on 
 
 into a Haar measure on  
 
. 
Mock exponential maps are known to exist in several situations. For example, for 
 
, the map 
 
works. For a classical group that splits over a tame extension of 
 
, with odd residual characteristic, the Cayley transform works. If 
 
has characteristic zero and 
 
 is sufficiently large, then the exponential map works. We will need the next two hypotheses in order to apply Theorem   10.14 . 
 Hypothesis 7.6. 
 Assume that 
 
satisfies Hypothesis  3.4.3 of [
6] 
, concerning the convergence of nilpotent orbital integrals.  
 This is automatically satisfied when 
 
 has characteristic zero.
 Hypothesis 7.7. 
 Assume that 
 
 and 
 
satisfy the hypotheses of Theorem  2.1.5(3) of [
6] 
.  
 
 8  Lie algebra decompositions
 Lemma 8.1. 
 Let 
 
 be a tame extension containing the eigenvalues of 
 
. For all 
 
, and  
 
, we have  
 
 where the sum is over the set of eigenvalues of the action of 
 
 on 
 
, and each  
 
 is the associated eigenspace. 
 
- 
 
Proof.
Let 
 
 be the order of  
 
. Then 
 
for some maximal torus 
 
 such that  
 
 is 
 
-split. Since  
 
 acts trivially on 
 
and acts on each 
 
-root space in 
 
, we have from  3.1 that  
 
 where the sum runs over the eigenvalues 
 
 of 
 
, and each  
 
 denotes the corresponding eigenspace. Thus it is enough to show that for each 
 
,  
 
 This will follow from Lemma   8.3 and Hypothesis   7.3 . □ 
 
 We may identify  
 
 with the 
 
-eigenspace of 
 
in  
 
. Define  
 
 to be the sum of all of the other eigenspaces of 
 
. Identify  
 
 with  
 
and define  
 
. These objects are all defined over 
 
, and  
 
 The following result is well known when 
 
 is connected. See  [1,Prop. 1.9.2] . 
 Corollary 8.2. 
 Suppose 
 
 and  
 
. Then  
 
 
 Lemma 8.3. 
 Let 
 
 denote the ring of integers in 
 
, and let 
 
 be a uniformizer. Let  
 
 be a finite-dimensional vector space over 
 
, and 
 
 a lattice in  
 
. Let  
 
 denote a diagonalizable linear map such that 
 
 for some 
 
. Let  
 
 be the eigenvalues of  
 
, and  
 
 the corresponding eigenspaces. Suppose that for 
 
, we have that 
 
. Then  
 
 
 
- 
 
Proof.
Without loss of generality, we may replace  
 
 by  
 
, and thus assume that 
 
, so that  
 
 for all 
 
. Let 
 
, and write  
 
, where  
 
. 
We must show that 
 
 for all 
 
. Let  
 
, and suppose that 
 
is nonempty. Pick a minimal nonempty subset 
 
 such that 
 
 for some coefficients  
 
. Pick such coefficients, and let 
 
 be the resulting sum. 
We must have 
 
. Let  
 
. Then 
 
. But we may rewrite this element as  
 
 contradicting the minimality of  
 
. □ 
 
 
 9  Some lemmas
 Some results in this section are stated in terms of the Lie algebras 
 
 and 
 
. 
However, by Hypothesis   7.4 , the analogous results for  
 
 and  
 
 are also valid (with the same proofs). 
 Lemma 9.1. 
 Let 
 
, 
 
, and  
 
. If 
 
 then  
 
. 
 
- 
 
Proof.
Write 
 
, with  
 
. Let 
 
. 
From Hypothesis   7.1 , there is a tame extension 
 
 of 
 
 containing the eigenvalues of 
 
. Write  
 
, and  
 
, where each sum is over the set of eigenvalues for the action of 
 
on 
 
, and each  
 
 and  
 
 belongs to the corresponding eigenspace. Then  
 
, where  
 
. 
From our hypothesis on 
 
, there is some 
 
 so that  
 
. From Lemma   4.6 , 
 
. Therefore, from Lemma   8.1 it will be enough to show that  
 
. Note that from Hypothesis   7.5 ( 3 ),  
 
. 
Therefore, Lemma   8.1 implies that  
 
. 
Suppose 
 
. Then  
 
. By the definition of 
 
,  
 
, and our conclusion follows. 
Now suppose that 
 
. Let 
 
. Then  
 
and  
 
, so  
 
. □ 
 
 Proposition 9.2. 
 Let 
 
, 
 
, 
 
, and  
 
. 
Write 
 
 according to the decomposition in Corollary   8.2 . If  
 
, then  
 
.  
 
- 
 
Proof.
For some 
 
, 
 
. By Lemma   9.1 ,  
 
. On the other hand, since  
 
 decomposes as 
 
, we have  
 
. Thus 
 
, which implies that  
 
. Finally, recall from Lemma   4.6 that 
 
. □ 
 
 Proposition 9.3. 
 If 
 
, 
 
, and  
 
, then  
 
 
- 
 
Proof.
This follows from Corollary  3.2.6 of  [2] and Hypothesis   7.2 . □ 
 
 Definition 9.4. 
Let 
 
. A character  
 
 is called degenerate if the coset that corresponds to 
 
 under the isomorphism  3.3 contains nilpotent elements. One can similarly define what it means for a character of 
 
 to be degenerate.  
 Lemma 9.5. 
 Let 
 
,  
 
, and suppose 
 
. 
Let  
 
 be a character such that 
 
 for some  
 
. Suppose that for some 
 
, 
 
 is trivial on  
 
. Then 
 
 is trivial on  
 
.  
- 
 
 
Proof.
By  3.3 , 
 
 corresponds to some coset  
 
. By Lemma  1.8.1 of  [1] , 
 
. Pick 
 
such that 
 
. Then  
 
. Write 
 
 with respect to the decomposition in Corollary   8.2 . By Proposition   9.2 ,  
 
. Hence,  
 
. 
Since 
 
 is trivial on  
 
, we have that  
 
. Therefore  
 
, implying that 
 
 is trivial on  
 
. □
 
 Remark 9.6. 
 In fact, the proof shows that 
 
 is trivial on a slightly larger subgroup that corresponds to the lattice 
 
 (for some 
 
 such that  
 
) via  3.2 .  
 
 10  Harmonic analysis
 
 From distributions on 
 
 to distributions on  
 
.
 For the distribution  
 
 (or any other distribution) on 
 
, we follow Harish-Chandra  [10,§18] (see also  [18] ) to define a distribution 
 
 on  
 
 that captures the behavior of  
 
near 
 
. 
Fix 
 
. The proof of Proposition  1 of  [18] is valid for general reductive groups, so we may apply it to see that the following (surjective) map is everywhere submersive:  
 |  |  | 
 |  |  | 
 
 Theorem 10.1. 
 There exists a unique, surjective, linear map  
 |  |  | 
 
 such that for all 
 
 
 
 and 
 
.  
 
- 
 
Proof.
This is Theorem 11, p.  49, of  [8] applied to the map 
 
 above □ 
 
 Remark 10.2. 
The set 
 
 is an open (since 
 
 is submersive), 
 
-invariant neighborhood of 
 
 in 
 
, so 
 
. Note that the set 
 
 is not necessarily open. 
 Fix an open compact subgroup 
 
 of 
 
; let  
 
 denote its characteristic function. 
We have the following diagram:  
 |  |  | 
 
 
 where the first arrow is the restriction map; the second arrow is given by 
 
; and the third arrow is the map of Theorem   10.1 . Note that the support of  
 
 is contained in 
 
, an open, 
 
-invariant neighborhood of 
 
 in 
 
. 
Let  
 
 denote the distribution character of 
 
. 
 Definition 10.3. 
 Define the distribution 
 
 on  
 
 by 
 
. 
 Lemma 10.4. 
 Normalize the measure on 
 
 so that 
 
 has total measure 
 
. Then for each 
 
 
 
 where the sum is over a finite set (which depends on  
 
) of representations. 
 
 Note that a similar statement appears on p.  78 of [10] . 
- 
 
 
Proof.
Combine equation  6.1 with 
 |  |  |  
 |  |  |  
 |  |  |  
 |  |  |  
 |  |  |  
 
 
 
 For the following Lemma, note that 
 
is a 
 
-invariant neighborhood of 
 
 in  
 
, and that 
 
is dense in 
 
. 
 Lemma 10.5. 
 The distribution 
 
 is represented on  
 
 by the function 
 
. 
 
- 
 
Proof.
For all 
 
, 
 |  |  |  
 |  |  |  
 |  |  |  
 |  |  |  
 
 
 
 
 Review of the Fourier transform
 We recall the following from, for example, §4.1 of [2] . Let  
 
 be a Haar measure on 
 
. For any 
 
, we define the Fourier transform 
 
of  
 
 by  
 
 for  
 
. Let 
 
 be a Haar measure on  
 
. For 
 
we use the natural identification of  
 
 with 
 
 and define the Fourier transform 
 
by  
 
 for 
 
. We normalize the measures  
 
 and 
 
 so that for 
 
 and 
 
 
 |  | (10.6) | 
Recall that, from Hypothesis   7.4 , we can (and eventually will) identify 
 
 with its linear dual  
 
. With this identification, the Fourier transform becomes a map from 
 
to itself. Given our normalization of measures, we have that for 
 
and  
 
, 
 
. 
 From distributions on  
 
 to distributions on  
 
.
 Using 
 
 and  
 
 as in Hypothesis   7.5 , we now define a distribution  
 
 on  
 
(compare  [6,§2.1] ). 
 Definition 10.7. 
 For 
 
, define 
 
. 
Remark 10.8. 
For 
 
, 
 
 if and only if 
 
. Hence in this case 
 
. 
 
 Some interesting spaces
 The following were defined in  [6,§2.1] . 
 Definition 10.9. 
For any  
 
-invariant subset 
 
 of 
 
 or  
 
, let 
 
denote the space of  
 
-invariant distributions supported on 
 
. 
Definition 10.10. 
 Suppose 
 
. Define the spaces of distributions 
 
 and 
 
 
 Definition 10.11. 
Define the space of functions 
 
 
 Remark 10.12. 
One can define  
 
,  
 
, and  
 
similarly. 
 
Remark 10.13. 
  
 
 if and only if  
 
 (see Definition 4.2.1 and Lemma 4.2.3 of  [
2] 
).  
 If  
 
 is a distribution on 
 
, then we let  
 
 denote the restriction of  
 
 to  
 
. The following homogeneity result was proved by DeBacker  [6,Theorem2.1.5] . From a remark in the introduction to loc. cit., the result does not require that 
 
 be connected. 
 Theorem 10.14. 
 
 
 
 
 11  The Main Theorem
 Recall that we are assuming the hypotheses in § 7 . Let  
 
 be as in definition   10.7 . 
 Theorem 11.1. 
 Let 
 
. Then  
 
. 
 
- 
 
 
Proof.
It is enough to show that  
 
 for all 
 
and all 
 
 such that 
 
. Fix 
 
and 
 
, and take 
 
. Suppose 
 
. We need to show that 
 
. By the linearity of  
 
, it suffices to show this for 
 
, where  
 
. In other words, it suffices to show that the character  
 
 of  
 
 is degenerate, where  
 
. We have  
 
 Thus 
 
implies that for some  
 
, 
 
, where 
 
is defined by 
 
. Using Lemma   10.4 with 
 
, we get  
 
 Thus for some  
 
, 
 |  | (11.2) |  
 where the last equality holds because 
 
 is supported on  
 
. 
Pick minimal 
 
 such that 
 
 is trivial on  
 
. Let 
 
. 
Since  
 
, one can restrict 
 
 to  
 
. Since the group  
 
 is abelian, this restriction decomposes into a finite sum of irreducible, one-dimensional representations  
 
. From equation  11.2 and Remark   6.3 , we see that for some 
 
, and for all 
 
,
 |  |  |  
 |  |  |  
 |  |  |  
 
 where we think of each sum as running over a set of coset representatives for  
 
, and in the last line above we use the fact that 
 
 is constant on  
 
. 
Therefore, for some 
 
,|  | (11.3) |  
 If we assume that 
 
, then we may apply Lemma   6.5 to see that  
 
 is trivial on  
 
. Moreover,  11.3 implies that 
 
for some  
 
, which by Proposition   6.4 means that 
 
. If we further assume that  
 
, then we may apply Lemma   9.5 to see that 
 
. Since 
 
 is an irreducible representation of  
 
, it follows that  
 
 permutes the  
 
's transitively. Therefore, 
 
 is trivial on  
 
, so 
 
. Since this is true for all 
 
 satisfying 
 
 and 
 
, we see that|  | (11.4) |  
 We now use this inequality to prove four others: 
The first two of these follow trivially from the last. However, we prove them independently because we want to isolate the one part of this paper, the proof of   11.8 , that relies on the hypothesis that 
 
. 
Note that  11.5 is obvious in the case where 
 
. So assume that 
 
. 
Recall that 
 
 is either  
 
 or 
 
. If  
 
, then 
 
. If  
 
, then 
 
. Since 
 
, 
 
, so 
 
. 
Since 
 
, we have  11.5 . 
To prove  11.6 , note that  11.5 implies that 
 
, so 
 
. Since 
 
, we have 
 
. Since  
 
, we have 
 
. 
To prove  11.7 , note that  
 
. From  11.5 , we conclude that 
 
. 
To prove  11.8 , we use  11.4 to reduce to the case where 
 
. If  
 
, then 
 
. If  
 
, then since 
 
, we have  
 
. Since 
 
,  
 
. Therefore (using the hypothesis that  
 
), we have  
 
 We have
 |  |  |  
 |  |  |  
 |  |  |  
 
 Now let 
 
. From  11.6 ,  
 
 is trivial on  
 
. Apply Hypothesis   7.5 ( 1 ) together with Remark   6.3 to obtain:  
 
 From  11.7 and Lemma   6.5 , 
 
. From  11.8 ,  
 
 is trivial on  
 
. By Remark   9.6 , the restriction  
 
 of  
 
 to  
 
 is represented by a coset  
 
 where  
 
. Since 
 
, Theorem  3.5 of  [15] implies that  
 
 is degenerate. 
Thus, 
 
. Use Proposition   9.3 to conclude that 
 
, and hence that  
 
 is degenerate. Thus, 
 
, is degenerate. 
 □
 
 From now on, use Hypothesis   7.4 to identify  
 
 with  
 
 for all 
 
and all  
 
. For 
 
, let  
 
 denote the corresponding nilpotent orbital integral and let  
 
 denote its Fourier transform (both are distributions). 
From Hypothesis   7.6 and work of Huntsinger (see Theorem A.1.2 of [3] ), it is known that  
 
 is represented by a locally constant function (which we will also denote  
 
) on  
 
. (When 
 
 has characteristic zero, this is a result of Harish-Chandra  [10,Theorem4.4] .) 
 Corollary 11.9. 
 Let 
 
. Then 
 
 
 where 
 
 are complex constants that depend on 
 
, 
 
 and 
 
. 
 
- 
 
Proof.
Let 
 
. Then by Remark   10.13 ,  
 
. Let 
 
. For all 
 
, let 
 
 denote  
 
, and note that  
 
. Then for some coefficients  
 
, we have 
 
 |  |  |  
 |  |  |  
 |  |  |  
 |  |  |  
 
 
 
 Corollary 11.10. 
 Let 
 
. Then  
 
 for all 
 
. 
 
- 
 
 
Proof.
For  
 
, we have  
 
. Let  
 
. From Lemmas  1 and 2 of  [5] , 
 
. Since 
 
on 
 
, 
 
, and so  
 
. Thus,  
 
, and thus  
 
, and so the right-hand side of the equation makes sense. Since  
 
, the left-hand side makes sense. The result now follows from Corollary   11.9 and Lemma   10.5 . □ 
 
 Corollary 11.11. 
Let 
 
. Then  
 
 
 
- 
 
Proof.
This follows from Corollary   11.10 and the 
 
-invariance of  
 
. □ 
 
 Remark 11.12. 
When 
 
 is regular, we have that  
 
 is a torus, and so the only nilpotent orbit in 
 
 is the 
 
 orbit. Thus in this case there is only one orbital integral in the character expansion and its Fourier transform is the identity function. This means that the domain of validity of the local character expansion near a regular semisimple element is a domain on which  
 
 is constant. Thus we recover a generalization of the main result of  [
13] 
.  
Remark 11.13. 
It would be desirable, for applications of motivic integration to character theory, to have a version of Theorem   11.1 (and thus of its corollaries) that is valid under the weaker hypothesis that 
 
. In order to obtain such a theorem, one would have to replace the one part of the proof of Theorem   11.1 that assumes 
 
: the proof of inequality  11.8 . This inequality allows us to apply Remark   9.6 , a slight strengthening of Lemma   9.5 , to the character  
 
. However, if we had a version of Lemma   9.5 strong enough to apply directly to the representation 
 
 (which is not necessarily one dimensional), then  11.8 would be unnecessary. We will pursue this matter elsewhere.  
 References
- 
 
J. D. Adler, Refined anisotropic 
 
-types and supercuspidal representations, Pacific J. Math  185 (1998), 1–32. 
- 
J. D. Adler and S. DeBacker, Some applications of Bruhat-Tits theory to harmonic analysis on the Lie algebra of a reductive 
 
-adic group, Michigan Math. J., 50 (2002), no.  2, 263–286. 
- 
 , Murnaghan-Kirillov theory for supercuspidal representations of tame general linear groups, J. Reine Angew. Math. 575 (2004). 
- 
J. D. Adler and A. Roche, An intertwining result for 
 
-adic groups, Canad. J. Math. 52  (2000), no.  3, 449–467. 
- 
L. Clozel, Characters of non-connected, reductive 
 
-adic groups, Canad. J. Math. 24  (1987), no.  1, 149–167. 
- 
S. DeBacker, Homogeneity results for invariant distributions of a reductive 
 
-adic group, Ann. Sci. École Norm. Sup. (4) 35 (2002), 391–422. 
- 
 , Some applications of Bruhat-Tits theory to harmonic analysis on a reductive  
 
-adic group, Michigan Math. J. 50 (2002), no.  2, 241–261. 
- 
Harish-Chandra (notes by G. van Dijk), Harmonic analysis on reductive 
 
-adic groups, Lecture Notes in Mathematics, vol.  162, Springer, 1970. 
- 
 , A submersion principle and its applications, Geometry and Analysis: Papers Dedicated to the Memory of V. K. Patodi, Indian Academy of Sciences, Bangalore, 1980, pp.  95–102. 
-  
 , Admissible invariant distributions on reductive 
 
-adic groups, Notes by Stephen DeBacker and Paul J. Sally, Jr., University Lecture Series, 16, Amer. Math. Soc., 1999. 
- 
R. Howe, The Fourier transform and germs of characters (case of  
 
 over a 
 
-adic field), Math. Ann. 208 (1974), 305–322. 
- 
 , Some qualitative results on the representation theory of  
 
 over a 
 
-adic field, Pacific J. Math. 73 (1977), 479–538. 
- 
J. Korman, On the local constancy of characters, preprint, 2004. 
- 
A. Moy and G. Prasad, Unrefined minimal 
 
-types for 
 
-adic groups, Invent. Math. 116  (1994), 393–408. 
- 
 , Jacquet functors and unrefined minimal 
 
-types, Comment. Math. Helvetici 71  (1996), 98–121. 
- 
G. Prasad, Galois fixed points in the Bruhat-Tits buildings of a reductive group, Bull. Soc. Math. France 129 (2001), 169–174. 
- 
G. Prasad and J.-K. Yu, On finite group actions on reductive groups and buildings, Invent. Math. 147 (2002), 545–560. 
- 
F. Rodier, Intégrabilité locale des caractères du groupe 
 
 où 
 
 est un corps local de caractéristique positive, Duke Math. J. 52 (1985), no.  3, 771–792. 
- 
G. Rousseau, Immeubles des groupes réductifs sur les corps locaux, Thèse, Université de Paris-sud, 1977. 
- 
J. Tits, Reductive groups over 
 
-adic fields, Automorphic forms, representations, and  
 
-functions (A. Borel and W. Casselman, eds.), Proc. Symp. Pure Math., vol.  33, part 
 
, AMS (1979), pp.  29–69. 
- 
J.-K. Yu, Construction of tame supercuspidal representations, J. Amer. Math. Soc. 14  (2001), no.  3, 579–622. 
- 
 , Smooth models associated to concave functions in Bruhat-Tits theory, preprint, September, 2002. 
E-mail address : adler@uakron.edu  The University of Akron, Akron, OH 44325-4002  E-mail address : jkorman@math.toronto.edu  The University of Toronto, Toronto, Ontario M5S 3G3