Uniform analysis on local fields and applications to orbital integrals

By Raf Cluckers, Julia Gordon, and Immanuel Halupczok

Abstract

We study upper bounds, approximations, and limits for functions of motivic exponential class uniformly in non-Archimedean local fields whose characteristic is or sufficiently large. Our results together form a flexible framework for doing analysis over local fields in a field-independent way. As corollaries, we obtain many new transfer principles, for example, for local constancy, continuity, and existence of various kinds of limits. Moreover, we show that the Fourier transform of an -function of motivic exponential class is again of motivic exponential class. As an application in representation theory, we prove uniform bounds for the Fourier transforms of orbital integrals on connected reductive -adic groups.

1. Introduction

One branch of analysis on non-Archimedean local fields deals with functions from (subsets of) to . In this paper, we develop a framework which makes it possible to carry out this kind of analysis uniformly in the local field in a similar sense as algebraic geometry works in a field-independent way. This extends the work pursued in Reference 7, Reference 26, Appendix B, Reference 10, and Reference 18, Section 4. One of our motivations is the program initiated by Hales to reformulate in a field-independent way the entire theory of complex admissible representations of reductive groups over local fields, Reference 20Reference 21. We note that in retrospect, one can see an inkling of the possibility of such a reformulation already in one of the earliest books on the subject; see Reference 17, p. 121. This paper develops the framework in which this program can be carried out. Another motivation is to enable further applications in the line of e.g. Reference 6Reference 8Reference 11Reference 18Reference 26Reference 27. In particular, in Section 4, we develop uniform in bounds for the Fourier transforms of orbital integrals, normalized by the discriminant, discussed in more detail below.

Using our framework has several direct implications. The most striking one is probably that for any property that can be expressed uniformly in the field using the formalism, one has a transfer principle like the one of Ax–Koshen/Ershov Reference 2Reference 16, i.e., whether the property holds can be transferred between local fields of positive and mixed characteristic (provided that the residue field characteristic is big enough). Another direct implication is that if some value (obtained using the formalism) can be bounded for each local field independently, then one obtains some precise information on how the bound depends on , for free. An example of this approach, applied to Fourier transforms of orbital integrals, appears in Section 4.

By “doing analysis uniformly in several fields”, we mean that the analytic operations are carried out on certain abstract objects which can be specialized to every specific field, yielding the familiar concrete objects. We take a very naive approach to this: we simply define an abstract object to be the collection of its specializations. We do however require that this collection is given in some uniform way (to be made precise below). As an example, one kind of object we work with is definable sets in the sense of model theory. According to the above approach, we consider a definable set as a collection of sets , all given by the same formula, where runs over the local fields we are interested in.

All our uniformity results are valid for all local fields whose characteristic is or sufficiently large. For this reason, it makes sense to think of two objects and as being the same if they differ only for of small positive characteristic. (Formally defining the objects this way would have been possible but unhandy.) Note that in particular, we obtain uniformity results even for local fields which are arbitrarily ramified extensions of , for every . This is new, compared to most previous papers, and it builds on Reference 12.

The central objects of study are -functions”, as in Reference 12, also called “functions of -class” or “of motivic exponential class”. (Those generalize the “motivic exponential functions” of Reference 7Reference 9Reference 10, based on the “motivic constructible exponential functions” of Reference 15.) A -function is a uniformly given collection of functions from, say, to for every local field . As an example, given a polynomial , we obtain a -function by setting , where is the norm on the local field (in fact, this function lies in a smaller class of -functions – a similar class without the exponentials, also defined below). Another example can be built by composing a nontrivial additive character on with a polynomial. In general, the class of -functions is defined as the -algebra generated by those and some other functions, which involve the valuation on , certain exponential sums over the residue field, and additive characters . (The word “exponential” refers to the presence of those additive characters which behave like exponentiation.)

The class of -functions has intentionally been defined in such a way that it is closed under integration in the following sense. Given any -function , say in variables, there exists a -function in the first variables such that for every local field of characteristic or , we have

(Here, the integral is a usual Lebesgue-integral with respect to a suitably normalized Haar measure on . For the moment, we assume that is integrable for each and each .) In this sense, we consider as a “field-independent integral of ”. In this paper, we prove that a whole zoo of analytic operations on -functions can be done in a similar, field-independent way.

For an operation which turns functions into other functions, being able to carry it out field-independently in this paper means that the class of -functions is closed under the operation, in the same sense as it is closed under integration as above.

Another type of question that we want to be able to answer field-independently is whether a function has a given (analytic) property (like being integrable, being continuous, being locally constant, existence of various kinds of limits). This really becomes meaningful only in a parametrized version: Given a -function in, say, variables, we want to understand how the set

depends on . Typically, the collection is not a definable set; however, for most of the properties we will consider, we will prove that is what we call a -locus”: for some -function . These loci have already appeared in Reference 7; in the present paper, they play a central role.

Even though -loci are not definable sets, they are almost as flexible; for example, the collection of -loci is closed under positive boolean combinations, under the -quantifier, and under various kinds of for-almost-all quantifiers. This flexibility is inherited by our formalism: It suffices to prove that -functions and -loci are closed under very few “fundamental” operations to then obtain a multitude of other operations essentially for free (i.e., by combining the fundamental ones in various ways). Accordingly, this paper contains a few long and deep proofs (of the fundamental operations), from which then many other results follow easily.

As stated at the beginning of this introduction, our formalism yields Ax–Kochen/ Ershov-like transfer principles. More precisely, one has a general transfer principle for arbitrary conditions that can be expressed in terms of -loci. (This follows directly from the transfer principle for equalities in Reference 15.) This means that we obtain a transfer principle for every analytic condition which can be expressed uniformly in the sense of this paper. For example, given a -function , it only depends on the residue field of whether is continuous/integrable/bounded/constant/locally constant/has a limit at /etc., provided that the residue field characteristic is large enough.

The outline of the paper is as follows. Section 1.2 contains the basic definitions that are necessary to understand all results; in particular, -functions and -loci are defined.

In Section 1.3, we recall some results from previous papers needed in this paper. Here, Proposition 1.3.1 plays a key role, listing the most important operations under which the collection of -loci is closed.

In the remainder of Section 1, we recall some more results from Reference 7Reference 12 (in particular about integrability), we introduce one new basic “operation” on -loci (namely eventual behavior, Section 1.4), and we give some first example applications of the locus formalism.

The remaining uniform analytic operations are grouped by topic in the next two sections, as follows.

Section 2 contains the results about bounds and suprema. Given a bounded -function (i.e., such that is bounded for each ), we obtain results about the dependence of the bound on and also, for families of functions, the dependence of the bound on the family parameter. A statement about a bound on is the same as a statement about , so it would be natural to first prove that the class of -functions is closed under taking suprema, a result which would also be useful for many other purposes. Unfortunately, given a real-valued -function , the supremum is, in general, not a -function in . However, what we obtain instead is that the supremum is “approximately” a -function. For all applications of the supremum in this paper, this turns out to be strong enough.

The main topic of Section 3 is limit behavior: pointwise limits, uniform limits, and limits with respect to the -norm for various . As applications, we prove results about continuity and about Fourier transforms.

Finally, in Section 4, we give an application of the results of Section 2 to Fourier transforms of orbital integrals. The Fourier transform of an orbital integral of a reductive Lie algebra is itself a locally integrable function on , and it is known that when normalized by the square root of the discriminant, this function is bounded on compact sets. We prove that for a definable family of definable compact sets , the bound is of the form , where is the cardinality of the residue field of and and are constants that depend only on and the formulas defining the sets .

This systematic study of bounds for -functions grew out of the similar study for motivic functions without the exponential that was initiated in order to answer a question that arose in an application of the Trace Formula Reference 26, Appendix B. Though we do not have such an application in mind for the Fourier transforms of orbital integrals, we use this opportunity to establish this bound, both for possible future applications and to illustrate applicability of this machinery. In particular, now that the orbital integrals are proved to be motivic exponential functions (see Lemma 4.3.2 and the preceding paragraph), all the results of Section 3 on the speed of convergence are applicable in particular to families of orbital integrals, giving for free the estimates by a negative power of for all kinds of sequences of orbital integrals that are known to converge to zero. We do not pursue any specific results of this kind in this paper, but note that they should follow from the results of Section 3 and Lemma 4.3.2 automatically. We hope that these results will be useful in applications of the Trace Formula in the spirit of Reference 26.

1.1. Summary of the results

Spread over the paper, we give many examples of results that can be deduced from the basic ones using the formalism. For the convenience of the reader, we here give a summary of all results (including those that were known before). Those marked with a are “fundamental” ones, i.e., those which need real work; those without a are the ones deduced using the formalism.

Operations on -functions:

Applying any of the following operations to a -function yields a -function again:

Integral Theorem 1.5.1
Bound Theorem 2.1.2
Approximate supremum Theorem 2.1.3
Pointwise limit Theorems 3.1.3, 3.1.8
Continuous extension Proposition 3.1.6
Uniform and -limit Theorem 3.2.1
Fourier transform Theorem 3.3.1

Properties of -functions:

The following properties of -functions are given by -loci (2nd column) and can be transferred in the Ax–Kochen/Ershov way (3rd column):

Constant Corollary 1.3.2 Corollary 1.3.4
Locally constant Corollary 1.4.3 Corollary 1.4.4
-integrable Theorem 1.5.1 Reference 7, Thm. 4.4.1
Bounded Theorem 2.1.2 Reference 7, Thm. 4.4.2
Limit is Theorems 3.1.1, 3.1.2 Corollary 3.1.7
(Pointwise) limit exists Thms. 3.1.3, 3.1.4, 3.1.8 Corollary 3.1.7
Continuous Corollary 3.1.5 Corollary 3.1.7
Limits exist: uniform, , Theorem 3.2.1 Corollary 3.2.4
-, -, -integrable Corollary 3.2.2 Corollary 3.2.5
-norm equal to Corollary 3.2.3 Corollary 3.2.5

Operations on -loci and properties of -loci:

Applying any of the following operations to a -locus yields a -locus again:

Boolean combinations, -quantification Proposition 1.3.1
Eventual behavior Proposition 1.4.1
Almost everywhere quantification Corollary 3.2.3

Consequences of being a -function / a -locus:

Transfer principle for -loci Theorem 1.3.3
Bounds are uniform in the field Theorems 2.1.1, 2.1.2
Speed of convergence of pointwise limits Thms. 3.1.1, 3.1.2, 3.1.8
-convergence pointwise convergence Lemma 3.2.6

Applications to Fourier transforms of orbital integrals:

Orbital integrals are of -class (up to a -constant) Lemma 4.3.2
Fourier transforms of orbital integrals are of -class Lemma 4.3.2
Uniform bound for Fourier transforms of orbital integrals Theorem 4.3.1

1.2. Basic definitions: Functions and loci of -class

The main object of study in this paper is the class of -functions”; by incorporating all finite field extensions of for all as in Reference 12, they generalize the “motivic exponential functions” of Reference 10, Section 2 and Reference 9, Section 2, all based on Reference 15. We note that this class of functions differs also in another way from the class of motivic constructible exponential functions originally defined in Reference 15: instead of the abstract motivic functions of Reference 15, we consider collections of functions, uniformly in local fields. This avoids issues related to what is called null-functions in Reference 3. We recall this framework and fix our terminology.

Definition 1.2.1 (Local fields).

Let be the collection of all non-Archimedean local fields equipped with a uniformizer for the valuation ring of . (So more formally, elements of are pairs .) Here, by a non-Archimedean local field we mean a finite extension of or of for any prime .

Given an integer , let be the collection of such that has characteristic either or at least .

We will use the notation “for all to mean “there exists an such that for all ”.

Note that in various previous papers, the characteristic of the residue field is required to be sufficiently large. In contrast, results in this paper only require , which allows the characteristic of to be arbitrary if the characteristic of is zero. The reason that this weaker hypothesis is enough is that we build on the results of Reference 12. On the other hand, it still seems far out of reach to treat local fields of small positive characteristic, given that the model theory of such fields is not understood.

Notation 1.2.2 (Sorts).

Given a local field , we sometimes write for the underlying set of , for the valuation ring of , for the maximal ideal, for the residue field, and for the number of elements of . The value group (even though always being ) will sometimes be denoted by . Moreover, we introduce a notation for the residue rings: For positive integers , set (where ). We write for the valuation map, for the residue map, and more generally for the canonical projections.

We use (or ”) as shorthand notation for being a subset of for some , and likewise (or ”) and ”, the latter meaning that is a subset of a product , for some and .

Note that is always a power of the ideal , but which power it is depends on and on . Namely, for the unique nonnegative integer such that the order of equals , one has . In particular, we have and hence whenever the residue field characteristic of does not divide . This somewhat technical definition of the sorts is necessary to obtain the desired uniformity in . In particular, since any formula uses only finitely many sorts, this implies that it suffices to exclude finitely many to achieve that the formula does not use any residue ring at all (except ). This fits with the papers that exclude small residue characteristic but do not need residue rings.

We use the same generalized Denef–Pas language as in Reference 12, Section 2.2 and consider each as a structure in that language. It is defined as follows.

Definition 1.2.3.

The language has sorts (the valued field), (the value group), and for each (the residue rings), with the ring language on and on each , the ordered abelian group language on , the valuation map , and generalized angular component maps sending to . See Reference 12, Section 2.2 for more details.

For some applications, it may be useful to have additional constants in the language (e.g. for the elements of a fixed subring of or for the uniformizer ). Using some standard techniques from model theory, all of our results can be suitably reformulated in such an enriched language; see Appendix A for details.

An -formula uniformly yields sets for all . As stated in the introduction, it will be convenient to consider a definable set as the collection of the sets it actually defines in the fields .

Definition 1.2.4 (Definable sets and functions).

A collection of subsets for some is called a definable set if there is an -formula such that for each in .

For definable sets and , a collection of functions for some is called a definable function and is denoted by if the collection of graphs of the is a definable set. (For this to make sense, we assume the of to be at least as large as the ones of and .)

In reality, we are only interested in definable sets and functions for “big ”; i.e., we think of and as being equal if

However, for practical reasons, it is often easier to work with representatives instead of introducing this equivalence relation formally. Nevertheless, we will often implicitly enlarge , for example calling definable if is definable for large enough .

Our definition of definable sets is the one which will be most convenient for this paper, but note that there are various other possibilities. For example, if one were to define a definable set to be the collection indexed only by the local fields of characteristic , one could recover Equation 1.2.1 by the Ax–Kochen/Ershov transfer principle.

We apply the typical set-theoretical notation to definable sets , e.g. writing (if for each ), , and so on, which may increase if necessary. More generally, we will often omit the indices when the intended meaning is clear, writing e.g. to mean the corresponding collection of subsets of the ; also, by “for all , we mean “for all and all , ”.

Using definable functions and sets as building blocks, we introduce “functions of -class”; those will then be generalized to “functions of -class”.

Definition 1.2.5 (Functions of -class).

Let be a definable set. A collection of functions is called a function of -class (or simply a -function) on if there exist integers , , and , nonzero integers , definable functions and , and definable sets such that for all and all ,

where , for some and .

We write to denote the ring of -functions on .

In this definition, we already omitted lots of indices from the notation. Equation Equation 1.2.2 e.g. really means

(Note that each is a finite set, so makes sense.)

Definition 1.2.6 (Additive characters).

For any local field , let be the set of the additive characters on that are trivial on the maximal ideal of and nontrivial on .

Expressions involving additive characters of -adic fields often give rise to exponential sums; that’s what the “exp” stands for in the definition below.

Definition 1.2.7 (Functions of -class).

Let be a definable set. A collection of functions for and is called a function of -class (or a -function) on if there exist integers and , functions in , definable sets , and definable functions and for , such that for all , all , and all ,

where for and , by abuse of notation, is defined as with any element in such that , which is well defined since is constant on . We write to denote the ring of -functions on .

In the same way as we omit indices , we will also omit indices from the notation, for example writing to mean ”. Also, we will freely consider collections as collections indexed by and not depending on .

Loci of vanishing of -functions will play a key role in this paper. To some extent, we will be able to apply the same reasoning as we do to the definable sets. We develop some terminology for that.

Definition 1.2.8 (Loci).

A locus of -class (or a -locus) is a zero set of a function , i.e., , where . Similarly, a locus of -class is a zero set of a function . (The latter one is a collection indexed by and .)

Note that any definable set is also a -locus, since characteristic functions of definable sets are of -class. The converse, however, is not true; for example, is a -locus, but it is not definable. Moreover, definable sets never depend on .

Sometimes, it is notationally more convenient to refer to the condition describing a set instead of the set itself.

Notation 1.2.9 (Conditions).

Fix some integers and consider a collection of conditions on elements . We call a definable condition if the collection of sets is definable. Analogously, we call a condition of -class or of -class, respectively, if is a locus of the corresponding class. Again, we also say -condition or -condition for short.

Thus a -condition is a family of conditions given by a -function , namely: holds if and only if (for all appropriate , , ).

Note that in the case , each is a condition on elements of a one-point set, so it makes sense to ask whether holds without specifying any element. (Definable such are essentially just first order sentences.)

Remark 1.2.10.

In the remainder of this section and in all of Sections 2 and 3, literally every result that is stated for -functions and -loci is also valid if one replaces all occurrences of by ; i.e., if no input object of a result uses the additive character , then does not appear in the output objects either. (This is obvious from the proofs.)

1.3. The locus formalism, transfer, and constancy

By definition of first order formulas, the class of definable conditions is closed under finite boolean combinations and under quantification. Results from Reference 7Reference 12 state that this is partially also true for the new classes of conditions introduced in Notation 1.2.9: each of them is closed under finite positive boolean combinations and under universal quantification. Note however that unlike definable conditions, they are not closed under negation and neither under existential quantification. For -conditions this follows from Example 1.3.5, and for -conditions it follows from Example 1.3.6.

The following proposition summarizes the positive results for further reference.

Proposition 1.3.1 (Basic operations on loci).

In the following, runs over a definable set , and runs over a definable set .

(1)

Any definable condition is a -condition.

(2)

If and are -conditions on , then so are and .

(3)

If is a -condition on and , then is a -condition on .

(4)

A -condition stays a -condition when considered as a condition on and that is independent of .

Again, we are omitting indices. For example, is the family of conditions, where, for , holds if and only if holds for all .

Proof of Proposition 1.3.1.

(1) is just a reformulation of the fact that definable sets are loci. (2) is just a simple manipulation of the -functions corresponding to and ; see Corollary 3.5.4 of Reference 7. (3) is the (Iva)-part of Theorem 4.4.2 of Reference 12 (cf. also Theorem 4.3.2 of Reference 7). (4) follows from the corresponding statement for -functions, i.e., that such functions in can also be considered as functions in and .

Another way of stating Proposition 1.3.1 is that any (syntactically correct) mathematical expression built out of -formulas, (for a -function), , , and defines a -condition. (Note that (4) is implicitly used when we write something like , which is a condition on .) The -functions in turn can be given by any expression of the form as in Definition 1.2.5 or 1.2.7.

Such expressions can be considered as a generalization of first order formulas, and they provide a short and convenient way to prove that new conditions are of -class. Several of the main results of this paper (and also of Reference 7) just state that certain additional operations can be used in expressions defining -conditions. As an example of how this formalism works, we prove:

Corollary 1.3.2 (Constancy).

Let be in , for some definable sets and . Then the set of for which the function is constant is a -locus.

Proof.

That is constant can be written as a -condition as follows:

To illustrate our conventions and formalism once more, let us give the following extra explanation of the above proof: is a -function in , , , so this difference being zero defines a -condition on , , , namely the family , where holds if and only if . By Proposition 1.3.1(3),

is a -condition, and using Proposition 1.3.1(3) once more, so is

This corresponds to a -locus , which, by definition of , is the one desired by the corollary: For every , we have if and only if holds for all (which just means that is constant).

One motivation to introduce -conditions is that for them, one has a transfer principle analogous to the Ax–Kochen/Ershov Theorem for first order sentences: Whether a -condition holds only depends on the residue field, provided that the residue field characteristic is large enough. More precisely, we have the following.

Theorem 1.3.3 (Transfer).

Suppose that is a -condition on , for some definable set . Then there exists an such that for all with residue field characteristics at least , the following holds.

If holds for all and all , then for any whose residue field is isomorphic to the one of , holds for all and all .

Proof.

This is Proposition 9.2.1 of Reference 15.

Note that as a “basic transfer result”, it would be enough to have Theorem 1.3.3 for 0-ary -conditions. Indeed, Theorem 1.3.3 for general is obtained by applying transfer to the 0-ary condition

which is of -class by Proposition 1.3.1(3).

This transfer principle can be combined with many other results from this paper. As a first example, we obtain:

Corollary 1.3.4 (Transfer for constancy).

Let and be definable sets, and let be in . Then there exists such that, for any of residue field characteristic , the truth of the following statement depends only on (the isomorphism class of) the residue field of :

Proof.

By Corollary 1.3.2, being constant is a -condition on , so Theorem 1.3.3 applies.

It might be tempting to replace our -loci by a larger class which is additionally closed under existential quantification, and maybe even under negation (thus turning it into a class of definable sets in a first order language). However, the following example shows that this would destroy the transfer principle.

Example 1.3.5.

Transfer (in the style of Theorem 1.3.3) does not hold for the statement

Indeed: The image consists of the -th roots of unity, where is the residue characteristic of . If itself has characteristic , then the entire image consists only of the -th roots of unity, whereas if has characteristic , then contains all -th roots of unity for all . Thus Equation 1.3.1 holds if and only if has positive characteristic.

Example 1.3.6.

We give an example that, in general, -loci are not closed under existential quantification. If -loci would be closed under existential quantification, then in particular would be a -locus, i.e., with . However, for any and any , the zero set of is ultimately periodic; i.e., there exist such that for , whether is zero or not depends only on . Indeed, this follows from Proposition 1.4.2: For large enough and in a fixed congruence class modulo some suitable , we have

for finitely many nonnegative integers , rational , and complex , and such a function is either constant equal to zero (namely when all are zero) or has only finitely many zeros, e.g. by Reference 7, Lemma 2.1.8.

1.4. Eventual behavior and local constancy

The first new building block for -conditions provided by this paper is the following proposition about eventual behavior. (One should not confuse this result with Theorem 14 of Reference 18, where is not allowed to depend on .)

Proposition 1.4.1 (Eventual behavior).

If is a -condition on running over a definable set and on running over , then the condition given by

is a -condition on . Moreover, can be chosen to depend definably on ; i.e., there exists a definable function such that for each , each and each , if holds, then holds for all .

Recall that Equation 1.4.1 defines a condition depending not only on but implicitly also on and ; that is, for any , any , and any , the condition holds if and only if there is an integer (possibly depending on ) such that for all integers with one has .

Thus the first part of the proposition states that in expressions defining -conditions (as explained below Proposition 1.3.1), we are also allowed to use the quantifier (meaning for all sufficiently large ).

The proof of this proposition (below) essentially follows a reasoning often used in Reference 7, whose main ingredient is that the dependence of on is controlled by Presburger-definable data. However, given that we work in a context allowing arbitrary ramification, this becomes true only after introducing additional -variables. The following proposition summarizes several of the results from Reference 12 we shall need.

Proposition 1.4.2.

Let and be definable and let be a -function on . Then there exist a definable bijection commuting with the projection to and a finite definable partition of such that we have the following for each part . We set .

(1)

The set is a Presburger Cell over , i.e.,

where and are integers, is the projection of , is either a definable function or constant equal to , and is either a definable function or constant equal to .

(2)

If , there exists an integer and a definable function such that for all , and similarly for . (And without loss, can be taken the same for all and .)

(3)

There are finitely many functions in and distinct pairs with nonnegative integers and rational numbers such that for all , for each , and each ,

(The map is called a reparameterization.)

Proof.

Let be the definable functions into the value group appearing in the definition of . Apply Reference 12, Corollary 5.2.3 to and to those functions . This yields a reparameterization and a finite definable partition of such that each is linear over on each part ; i.e., is linear for each fixed , with the coefficient of independent of and . This already implies (3).

By refining the partition using Presburger cell decomposition, we obtain (1). Finally, we apply Reference 12, Corollary 5.2.4 to the functions and bounding the cells. This yields a partition of each such that, after using that partition to refine the partition , (2) holds.

Proof of Proposition 1.4.1.

It is enough to find a definable as in the “moreover” part. Indeed, using such a , can be expressed as

which is a -condition by Proposition 1.3.1(1), (2), and (3) (since it can be written as ).

Let in be a function such that holds iff . To obtain , we apply Proposition 1.4.2 (with ); we also use the notation from there.

We may reduce to the case where is nonzero only on a single part . Indeed, the function which is equal to iff and otherwise is of -class, and we can define to be the maximum of the corresponding to the .

If the upper bound defining is not , then using Proposition 1.4.2(2) we can take , so now assume . In that case, we claim that we can take . More precisely, we claim that for each , the set either is empty or has no upper bound.

From

(see Equation 1.4.2) and using that all pairs are distinct, we obtain that for each , the set

is empty if for each and unbounded otherwise. Now the claim follows, using that is a union of sets of the form .

Using Proposition 1.4.1, we can turn global -properties of functions into local ones. As an example, we prove several variants of: local constancy is a property of -class.

Corollary 1.4.3 (Local constancy).

Let be in , for some definable sets and for some . Then we have the following (where for each , on we put the topology induced by the one on ).

(1)

The set of for which the function is locally constant is a -locus.

(2)

The set of such that the function is constant on a neighborhood of is a -locus. Moreover, the radius of constancy can be chosen definably; i.e., there exists a definable function such that for all , for all , for all , and for all , if is constant on a neighborhood of , then it is constant on the intersection of with the ball of valuative radius around in .

Proof.

(1) follows from (2) and Proposition 1.3.1, since being locally constant can be expressed by

(2) That is constant on a neighborhood of can be expressed by

where is the ball around of valuative radius , which again is a condition of -class by Propositions 1.3.1 and 1.4.1 (and, if one wants, Corollary 1.3.2).

Moreover, the witness for the quantifier provided by Proposition 1.4.1 is the desired definable function .

In particular, we obtain a transfer principle for local constancy.

Corollary 1.4.4 (Transfer for local constancy).

Corollary 1.3.4 still holds if one replaces “constant” by “locally constant”.

Proof.

Apply Theorem 1.3.3 to the condition is locally constant”, which is a -property (about ) by Corollary 1.4.3.

1.5. Integration

We put the Haar measure on so that has measure ; on and on , we use the counting measure, and for we use the measure on induced by the product measure on . Likewise, we put the discrete topology on and on , the valuation topology on , the product topology on , and the subset topology on .

Maybe the most important aspect of -functions is that they have nice and natural properties related to integration; see e.g. the following theorem stating stability under integration, generalizing Theorem 9.1.4 of Reference 15, by including also -adic fields of small residue field characteristic.

Theorem 1.5.1 (Integration, Reference 12, Theorems 4.1.1, 4.4.2).

Let be in , for some definable sets and .

(1)

The set of such that is -integrable over is a -locus.

(2)

There exists a function in such that

whenever is -integrable over . In more detail, for every , every , and every , if is -integrable, then .

Note that Theorem 1.5.1 immediately implies a generalized version, where the set we are integrating over is allowed to depend on : If is a -function on (and denotes the fiber of at ), then , where is the characteristic function of . Thus by applying the theorem to , we obtain that -integrability of on is a -condition, and the value of that integral (where it exists) is equal to for some function of -class.

In Reference 7, we proved that the condition that is locally integrable is also a -condition of the corresponding class. This can be deduced from Theorem 1.5.1 in a similar way as for local constancy (Corollary 1.4.3). However note also that since integrability on all small balls is equivalent to integrability on all balls, one doesn’t even need Proposition 1.4.1.

2. Bounds and approximate suprema

2.1. Results about bounds and approximate suprema

One result of Reference 7 (generalized to the present context in Reference 12) is that a -function being bounded is a -condition (see Theorem 2.1.2(1)). In this section, we describe how bounds depend on the field and on additional parameters; this generalizes results from Appendix B of Reference 26. We give three main results on bounds, in increasing strength and complexity, with Theorem 2.1.3 containing the core result, with the hardest and longest proof of this paper.

Theorem 2.1.1 (Bounds).

Let be in , where is a definable set and . Then there exist an integer and a definable such that for all , all , and all , the following holds.

If the function

is bounded on , then one actually has

where and where is the norm on .

By a definable , we mean a definable set where is a singleton in for each , where we identify with the integer with . Since we do not use parameters in our language until the appendix, a definable can be bounded by for some integers and depending on but not on ; that is, for each (where is considered as an element of via the natural map ). Indeed, this follows from quantifier elimination Reference 12, Theorem 5.1.1. If one is moreover only interested in of sufficiently large residue characteristic, then one may assume to be zero, so that the exponent in Equation 2.1.1 becomes for some integers not depending on .

In the appendix, can be more general depending on the constants added to and on the requirements (axioms) put on the interpretations of the constant symbols.

We observe that in the case with , and still using that we do not use parameters, Theorem 2.1.1 yields that for any -function on , there exist integers such that for any (in ) and any for which is bounded, the bound can be taken to be . Note also that is relevant only when one is interested in uniformity in for small residue characteristic and arbitrary (but finite) ramification.

Part (2) of the following theorem generalizes Theorem 2.1.1 by allowing the domain of to be an arbitrary definable set (instead of just ). In part (1), we recall the result from Reference 7Reference 12 about boundedness, since it fits nicely with part (2).

Theorem 2.1.2 (Bounds).

Let be in , where and are definable sets.

(1)

The set of such that the function is bounded is a -locus.

(2)

There exists a definable function such that for every , every , and every , if the function

is bounded on , then one has

This theorem implies Theorem 2.1.1 as follows.

Proof of Theorem 2.1.1.

Applying Theorem 2.1.2(2) with yields a definable function . To obtain Theorem 2.1.1, we need to know that we can replace by for some definable and some integer .

Note that when allowing big ramification, a definable function is not necessarily piecewise linear. However, by Reference 12, Corollary 5.2.5, and up to working piecewise with finitely many definable pieces , it differs from a linear definable function by at most for some integer . More precisely, for every and every , , where does not depend on . Moreover, we may assume that the coefficients of (which are rational numbers) are independent of and the constant term is given by a definable constant .

Thus on each piece , we obtain a bound of the form . Finally we take maxima of the so-obtained finite collection of and , yielding a bound of the desired form.

In general, the supremum over of is not a -function in . However, it can be reasonably well approximated by the square root of a -function. The precise statement is the following.

Theorem 2.1.3 (Approximate suprema).

Let be in , where and are definable sets. Then there exist a definable and a function in taking nonnegative real values such that the following holds: For each , each , and each for which the function

is bounded on , one has

The remark about the definable just below Theorem 2.1.1 also applies to the of Theorem 2.1.3. We do not believe that Theorem 2.1.3 would hold without squaring, i.e., with instead of . Indeed, using a one-point set as , this would imply that the absolute value of a -function can be approximated by a -function, which already seems unlikely for the function , .

Theorem 2.1.3 implies Theorem 2.1.2 as follows.

Proof of Theorem 2.1.2.

(1) is the Bdd-part of Reference 12, Theorem 4.4.2.

(2) Apply Theorem 2.1.3 to to obtain a (real-valued) function in such that whenever is bounded. It remains to show that any can be bounded by for some definable (as functions depending on ). (We will more generally bound for arbitrary , i.e., not necessarily real valued.)

Using the definition of , one easily reduces to the case . Indeed, by that definition,

for some , some definable sets , and some definable functions and . The complex norm of the inner sum is bounded by a product of some for each , so it remains to bound the .

So now suppose . By definition of , we have

for some definable , , and some nonzero integers . Clearly, each of the factors of the terms of the sum can be bounded as desired (using again ), and hence so can the sum.

2.2. Proof of Theorem 2.1.3

The proof of Theorem 2.1.3 is based on two deep results, which we develop first: one which allows us to reduce to functions living on and , and another one which allows us to find approximate suprema of functions on .

A key result of Reference 7 is its Proposition 4.5.8; the following is the generalization of that result to the present context given in Reference 12. (Here, for simplicity, we state a slightly weakened version.)

Proposition 2.2.1 (Reference 12, Proposition 4.6.1).

Let be an integer, let and be definable, and let be in . Write for variables running over and for variables running over . Then there exist integers , , a definable surjection over , definable functions , and functions in for , such that the following conditions hold for each and each ; we omit the indices to keep the notation lighter.

(1)

One has

(2)

If one sets, for ,

and

then

where the volume is taken with respect to the Haar measure on .

Here, as usual, is the set of such that lies in , and for the map to “be over means that makes a commutative diagram with the natural maps to .

Note that (2) in particular implies that if has positive measure, then

By virtue of Proposition 2.2.1, one can simplify the domain of a -function while preserving most of its behavior on growth and size, as follows.

Corollary 2.2.2.

Let , , and be definable sets and let be in . Then there exist a nonnegative integer , a definable surjection

over , and a nonnegative real valued function in such that for all , , , and , one has

where is the fiber of above and where by the right hand inequality, we in particular mean that the supremum is finite.

Proof of Corollary 2.2.2.

Write for some and some . If there is nothing to prove. We proceed by induction on . Apply Proposition 2.2.1 to (where the from the proposition is ), yielding integers , a definable surjection , and in for . We may assume that each fiber of has positive measure by treating the fibers of measure zero separately, using induction on ; in particular, Equation 2.2.1 holds, i.e., . Now set

Then is as desired (with this ): The left hand inequality of Equation 2.2.2 follows from Equation 2.2.1, and the right hand inequality follows from

and .

We now come to the second ingredient in the proof of Theorem 2.1.3. Suppose that is a bounded -function with domain in the value group, say for some definable . To obtain an approximate maximum of , we will choose a finite subset such that takes its maximum on , up to some factor . We will need to be able to do this in families in such a way that and do not depend on the parameters, and the elements of depend definably on the parameters. After various simplifications, what we end up needing are the following two lemmas: Lemma 2.2.3, which is used in the case where is infinite, and Lemma 2.2.4, which is used in the case where is finite but growing in size (when varying the family members).

Lemma 2.2.3.

Suppose that for , we have and . Then there exist positive integers and such that the following holds for every tuple and every . Suppose that the function

is bounded on . Then we have

Lemma 2.2.4.

Suppose that for , we have and . Then there exist positive integers and and Presburger definable functions with for all and and such that the following holds for every tuple , every , and every . Consider the function

then we have

We will prove both lemmas together.

Proof of Lemmas 2.2.3 and 2.2.4.

We start working on Lemma 2.2.4. First note that we may impose lower bounds on by taking some more Presburger functions covering the whole interval when is smaller; we will do this whenever convenient.

We will treat three special cases: namely when all are negative, when all are positive, and when all are zero. This will then be put together to obtain the result for arbitrary . For the latter to work, we will prove slightly stronger statements in the special cases, namely the following three claims.

Claim ().

If all are negative, then there exists such that for every and every , there exists such that

Claim ().

If all are zero, then there exist such that for every , every , and every sufficiently large , there exists , where such that

The third claim follows from Claim by replacing by :

Claim ().

If all are positive, then there exists such that for every , every , and every , there exists such that

Before we prove these claims, here is how they imply Lemma 2.2.4.

Write as a sum , according to the sign of the , and apply the corresponding claims to , , and . All the possible values of appearing in the three claims are Presburger functions in . These are the functions we use to conclude the lemma, so we need to prove that for any tuple , there exists a such that for some constant . The idea for this is that using that the bound on is decreasing and the one on is increasing, we can deduce that for one of the , dominates the other two summands of at . We then can bound in terms of . Here are the details:

We may assume that the values obtained from Claims  and (1) are the same. By imposing a lower bound on , we can ensure that

Indeed, we have , so imposing implies the left hand inequality; the right hand inequality is obtained in a similar way using .

It is sufficient to bound , since all outside this interval are equal to one of the definable functions anyway.

Let be such that is maximal among

We will prove that for this choice of , we have

This then implies, for , that

for some suitable multiple of ( works independently of ); this implies the lemma. Let us now prove Equation 2.2.4.

Suppose first that . Then we have

where the last inequality follows from . This yields

and hence .

The case works in exactly the same way.

Finally, suppose . Then we get

using . Analogously, we get , and hence, by the same computation as in Equation 2.2.6, . This finishes the proof that the three claims imply Lemma 2.2.4.

Before proving the claims, we carry out similar (but simpler) arguments for Lemma 2.2.3. Again, we write as a sum , according to the sign of the . Since the conclusion of Lemma 2.2.3 only speaks about tuples for which is bounded, we may assume that is constant and vanishes entirely.

We apply Claim  to , we don’t need Claim , and instead of using Claim (0), we simply use (for the from Claim ). Then the same arguments as for Lemma 2.2.4 yield

(This time, we do a case distinction on which of and is larger.)

Thus for both lemmas, it remains to prove the three claims. More precisely, it suffices to prove Claims (0) and , since Claim (1) is equivalent to Claim .

Proof of Claim .

The function is a polynomial of degree . Let be the vector space of all polynomials of degree and consider the map

Set , i.e., the preimage of the unit sphere in with respect to the maximum norm. Since is injective, is compact, so the maximum

exists. From this, we deduce that Claim (0) holds for . Indeed, set , where

By imposing , we obtain , and hence for . Thus

which is what we had to show.

Proof of Claim .

Set , , and . We can write as

Let be the tuple of all coefficients.

Set and . We will find an depending only on and (but not on ) such that

where and are the usual norms on . This then implies, for all , that

where the last inequality is ensured by choosing large enough, namely, such that holds for all and . Thus Equation 2.2.7 implies Claim .

To obtain Equation 2.2.7, instead of bounding , we may as well bound (since they differ at most by a factor ). We have , where is the matrix with coefficients

Suppose for the moment that is invertible. Then, by definition of the operator norm , we have , so it suffices to find an upper bound on of the form , where only depends on and . (Note that the entire matrix only depends on , , and .)

Up to a constant factor (namely ), is bounded by the maximum of the absolute values of the entries of . These entries are of the form , where is a minor of . Since is a polynomial in the entries of and those entries are bounded (even independently of , since all exponents are nonpositive), we have an upper bound on , and it remains to find a lower bound on of the form . This at the same time will prove that is invertible.

Considering as an indeterminate, we have . It is enough to prove that when considered as such a Laurent polynomial. Indeed, this then implies that for large enough, we have for some fixed , where is the least negative power of appearing in .

We have

where runs over all bijections . (We fix an order on for and the sign of to be well defined.) Each summand of the sum Equation 2.2.8 is a monomial in , namely of degree

Let be the smallest (i.e., most negative) degree in appearing among the summands in Equation 2.2.8. To prove , we will prove that the sum of the monomials of degree in is nonzero.

Let us write as a disjoint union of many intervals of length :

for . The degree is minimal if and only if, for every with , we have . This is equivalent to sending to for every .

Write for the set of bijections . Then the sum of the monomials of minimal degree in Equation 2.2.8 is (maybe up to sign)

Each factor in the product over is a Vandermonde determinant (corresponding to a polynomial of degree evaluated at each element of ) and hence nonzero. Thus , which is what remained to be proven.

Proof of Theorem 2.1.3.

By Corollary 2.2.2, it is enough to consider a nonnegative real-valued for and find a nonnegative real-valued and a definable such that

It will be handy to consider a slightly more general situation, namely where lives on a definable subset and where the suprema run over .

By a recursive procedure, it is enough to consider only the situations where for some or . In the first case, we set

(which lies in by Theorem 1.5.1). This is obviously at least as big as , and it exceeds the supremum at most by a factor , so it is as desired.

In the case , the idea is to use a rectilinearization result as in Reference 4 to reduce to Lemmas 2.2.3 and 2.2.4, though to deal with our setting allowing highly ramified fields, we need rectilinearization in the form stated in Reference 12, Proposition 5.2.6. The details are as follows.

By Reference 12, Proposition 5.2.3, after possibly introducing new residue ring variables (which we can later get rid of again as explained above) and after a finite partition of , we may suppose that has linear ingredients”; i.e., all functions involved in the definition of depend linearly on . Moreover, by Reference 12, Proposition 5.2.6 we may suppose that either for all or for some definable function . This involves introducing more new residue ring variables, another finite partition of , and applying a bijection which is linear over . (The linearity of this bijection ensures that still has linear ingredients.)

That has linear ingredients means that there exist and () such that for any , , and , the map is of the form

for some depending on , , and . Thus we can either apply Lemma 2.2.3 (if ) or Lemma 2.2.4 with . In both cases, we obtain an integer and finitely many definable functions with such that

for all . (In the case, we take to be constant equal to .) Now set

Then we have

So this is as desired, given that can easily be bounded by an integer power of .

3. Limits of -functions

This section contains various results about limits, namely: existence of limits in various contexts is given by -conditions, and the limit itself is a -function in the parameters. We consider both classical pointwise limits and -limits. Even though we obtain a whole variety of different results, it should be noted that certain other, subtly different statements do not seem to hold; see the remark after Theorem 3.1.3.

3.1. Pointwise limits and continuity

The first result states that limits being is a -condition and that convergence cannot be arbitrarily slow.

Theorem 3.1.1 (-limits).

Let be in , where is a definable set and where .

(1)

The set is a -locus.

(2)

There exists a rational number and a definable function , such that for each , for each , and for each , the following holds.

If one has

then one actually has

The above result also holds for more general kinds of limits.

Theorem 3.1.2 (-limits).

Let be in , where and are definable sets, and let be a surjective definable function.

(1)

The set is a -locus.

(2)

There exists a rational number and a definable function such that for all , for each , and for each , the following holds.

If one has

then one actually has

For some , the limits appearing in Theorems 3.1.1 and 3.1.2 might not even exist. (The theorems do not assume that they do.) The next result states that existence of limits is also a -condition and that if a limit exists, then its value is given by a -function.

Theorem 3.1.3 (Limits).

Let be in for some definable set .

(1)

The set is a -locus.

(2)

There exists in such that the following holds for all , for each and each .

then

Again, we can ask for a version of this for more general limits. We obtain:

Theorem 3.1.4 (Limits).

Let be in , where and are definable sets, and let be a surjective definable function. Then the set is a -locus.

Note that, surprisingly, in this generality it is not straightforward to prove that the value of the limit is of -class, and we even doubt that this is true. It seems that such a result would require some better understanding of the number of rational points of families of varieties over finite fields. On the other hand, it should be possible to obtain the generalization in a slightly altered context, namely after adding function symbols for Skolem functions to the sorts , or after adding more denominators as in the rational motivic functions of Reference 25.

From Theorem 3.1.4, one easily deduces that continuity is a -condition. In the following, given a definable set , we write for its topological closure (i.e., is the topological closure of in for each ), which is also definable.

Corollary 3.1.5 (Continuity).

Let be in , where and are definable sets and . Then

and

are -loci.

Proof.

The conditions at can be expressed respectively as follows:

Now use Theorem 3.1.2 and Proposition 1.3.1.

Note that if is continuous on and if exists for each , then the unique continuous extension of is not known by us to be of -class, for the same reasons as explained after Theorem 3.1.4. However, it is of -class in many slightly more restrictive cases. The most general setting in which this can be proven would probably be very technical, so in the following, we just prove it under reasonable assumptions.

Proposition 3.1.6 (Continuous extension).

Let be in , where and are definable sets and and suppose that is clopen (i.e., is open and closed for every ). Then can be extended to a -function on such that for every , if exists, then is continuous at .

Namely, with all indices, for every , for every , and for every , if exists, then is continuous at .

Proof.

We can simply define

where is the closed ball of valuative radius around (considered as a definable set). More precisely, we use Theorems 1.5.1 and 3.1.3 to find an such that for every and every , we have

whenever the integral and the limit exist, where is the concrete closed ball in given by and .

Since is clopen, for sufficiently large , we have , so that we are just averaging over this ball (where has measure zero by dimension considerations). If is continuous at , then the integral exists for large , and continuity also implies that the limit of those average values exists and is equal to , as desired.

As usual (namely, by transfer for -conditions) the above theorems imply corresponding transfer principles for limits, as follows.

Corollary 3.1.7 (Transfer principle for limits and continuity).

Let be in , where and are definable sets, and let be a definable function.

Then there exists such that, for any of residue characteristic , the truth of each of the following statements depends only on (the isomorphism class of) the residue field of ; here, in the third and fourth statements, we assume for some .

The proofs of Theorems 3.1.2 and 3.1.3 use Theorem 2.1.3 in a crucial way, to reduce to the case of a limit over a single -variable; other proofs, relying on Proposition 2.2.1 above and the rectilinearization result given by Proposition 5.2.6 of Reference 12, may also be thinkable. Here is the single -variable version of Theorems 3.1.2 and 3.1.3 (formulated in a more concise way):

Theorem 3.1.8 (Limits).

Let be in , where is a definable set. Then we have the following.

(1)

The condition exists” is a -condition on .

(2)

There exists a function , a definable function , and a rational number such that for every , every , and every , if the limit exists, then

for every . In particular, the limit is equal to .

This theorem implies Theorems 3.1.13.1.4 as follows.

Proof of Theorems 3.1.1 and 3.1.2.

Theorem 3.1.1 follows from Theorem 3.1.2 (using and ), so we now prove Theorem 3.1.2.

We may suppose that and that is the coordinate projection to the -variable. By Theorem 2.1.3, there exists a real-valued function in such that differs from by at most a factor of the form , for some definable . This implies that we can without loss replace by , considering the limit . In this version, the theorem follows pretty directly from Theorem 3.1.8. Indeed, that the limit is can be expressed by the condition exists and ”, which is of -class, where is the function obtained by applying Theorem 3.1.8 to , and claim Equation 3.1.3 follows from Equation 3.1.4.

Proof of Theorems 3.1.3 and 3.1.4.

Define in by

By applying Theorem 2.1.3 to , we find in such that this differs from by at most a factor not depending on , where and both run over . In particular we have that iff is a Cauchy sequence in the sense of the -limit. Thus the limit exists if and only if , which is a -condition by Theorem 3.1.8. This finishes the proof of Theorem 3.1.4 and of Theorem 3.1.3(1).

To obtain a function as desired in Theorem 3.1.3(2), we can simply take the limit of along any sequence of with , so we obtain it e.g. by applying Theorem 3.1.8(2) to

where runs over .

Proof of Theorem 3.1.8.

We apply Proposition 1.4.2 to , yielding (using the same notation as there) a reparameterization and a partition of into sets

(for some definable , , and some integers ) such that for , we have

for some , , and .

Denote the preimage of under by

For any fixed and , the fibers form a partition of . The limit exists if and only if, for each for which is unbounded, the corresponding restricted limit exists and moreover all these limits are equal. In particular, those for which are irrelevant. Indeed, recall that for each , either is constant or, by Proposition 1.4.2(2), is bounded for all . Therefore, in the remainder of the proof, we only consider those parts (and ) for which .

After shrinking to the union of those parts (and shrinking accordingly and restricting to ), we can entirely get rid of the reparameterization by replacing each by its preimage . To see this, what we need to check is that we have analogues of Equation 3.1.5 and Equation 3.1.6 for those . Those analogues can be obtained, provided that the restriction of to is of the form for some bijection , where is the projection of to . Indeed, such a is automatically definable (using ), and hence Equation 3.1.5 and Equation 3.1.6 for can be obtained by pre-composing and with .

Since is a reparameterization, the fiber is equal to a disjoint union of certain fibers , and what we need to check is that is actually equal to a single fiber . This follows from the fact that different , cannot be disjoint, by Equation 3.1.5: both fibers contain all large satisfying . Thus the reparameterization is indeed unnecessary, so from now on we assume that

for some and that, for , we have

We assume without loss that , and we let be the collection of those satisfying either or and .

Since for each fixed the pairs are distinct, given , the limit exists if and only if, jointly,

and

Each of these equations is a -condition on , so their conjunction is also a -condition (by Proposition 1.3.1), hence finishing the proof of (1).

When the limit exists, it is equal to the following -function:

for any with (e.g. the smallest for some linear ordering of the ).

It remains to find and . To this end, we may from now on without loss suppose that all are negative (by subtracting and by ignoring those for which the limit does not exist). In particular, the limit is if it exists.

We need to choose and in such a way that we have

Let and be the maximum of all and of all , respectively, and set . Then to obtain Equation 3.1.9, it suffices to prove

for all , where is the number of summands in Equation 3.1.9.

By the same argument as in the proof of Theorem 2.1.2 (or by actually applying Theorem 2.1.2 with being a singleton), we obtain a bound of the form for some definable . From this, one can easily obtain an as desired (e.g. choosing it such that and for ).

3.2. Limits of functions

In this section, we consider various types of limits of sequences of functions. For simplicity we restrict to functions on (though the notions would also make sense on other definable sets, using the counting measures on and on , and the proofs also go through in this generalized setting). We ask the usual two questions: Does the limit exist, and if yes, what is it? Those results will be used to study Fourier transforms of functions in Section 3.3.

By uniform convergence of a sequence of functions , , we mean convergence with respect to the sup norm, and by -convergence (for ), we mean convergence with respect to the usual -norm. (In particular, -convergence means uniform convergence except on a subset of measure zero.) Note that we do not require the individual functions to be -integrable. Instead, -convergence towards a limit function means that the -norm of the difference exists for sufficiently large and that it goes to .

Theorem 3.2.1 (Limits of sequences of functions).

Let be in

for some definable set and some . Given , , and , consider the sequence of functions (on )

indexed by integers .

(1)

For each of the notions of convergence listed below, the set of such that the sequence Equation 3.2.1 of functions converges in that sense is a -locus. The notions of convergence are: pointwise convergence, uniform convergence, -convergence, -convergence.

(2)

There exists a in such that the following holds for all , for each , and for each .

If the sequence Equation 3.2.1 of functions converges pointwise or uniformly or in -norm for any , then the function

is the limit of this sequence.

We believe that also part (1) of this theorem is true for -convergence for any . However, the proof would be more involved (even in the case ).

From the theorem, we deduce various related results (some of which would also have simple direct proofs).

Corollary 3.2.2 (-integrability).

Fix and let be in . Then the set of for which is -integrable is a -locus.

Proof.

The case is exactly Theorem 1.5.1(1). (We repeat the result for here only for completeness.) For , we obtain the corollary by applying Theorem 3.2.1 to . Indeed, for any fixed , the -limit exists (and is ) if and only if is -integrable.

Corollary 3.2.3 (The “almost everywhere” quantifier).

Suppose that is a -condition, where runs over a definable set and runs over . Then

is a -condition on . (Here, by “almost all”, we mean that the complement has measure .)

Proof.

Choose such that is if and only if holds, and set . Then for any , the sequence diverges for if and only if does not hold. Thus holds for almost all if and only if the -limit of exists, so that the corollary follows by applying Theorem 3.2.1 to .

As usual, we also obtain some transfer results.

Corollary 3.2.4 (Transfer principles for limits of functions).

Let be in for some definable set and some . Given , , and , consider the sequence of functions (on )

indexed by integers .

Fix a notion of convergence among: pointwise convergence / uniform convergence / convergence in -norm / in -norm.

Then there exists such that, for any of residue characteristic , the truth of each of the following statements depends only on (the isomorphism class of) the residue field of .

Concerning transfer of convergence to , let be obtained from as in Theorem 3.2.1(2). Then convergence to can be expressed by a -condition as follows (using Theorem 3.2.1(1) and Proposition 1.3.1):

Corollary 3.2.5 (Transfer principles for -norms).

Let be in for some definable set and some . Then there exists such that, for any of residue characteristic , the truth of each of the following four statements depends only on (the isomorphism class of) the residue field of .

Let us now prove Theorem 3.2.1. One ingredient to the second part is that existence of the -limit implies almost everywhere existence of the pointwise limit. This is not true in general, but it is true for sequences of functions of -class. Here is the precise statement.

Lemma 3.2.6.

Fix and let be in for some definable set . Also fix , , and . If the sequence of functions (on ) converges in -norm for , then converges for almost all .

(Without the assumption of being of -class, one can only find a subsequence which converges for almost all .)

Proof of Lemma 3.2.6.

We apply the same reasoning as in the proof of Theorem 3.1.8, using the same notation (except that our set plays the role of the set from the proof of Theorem 3.1.8), namely: We find a reparameterization and a partition of into finitely many , we can disregard those that have an upper bound , and after that, we can get rid of the reparameterization again in the same way as for Theorem 3.1.8.

Fix , , and for the entire remainder of the proof, and suppose that there exists a set of positive measure such that does not converge for any . For each such , either Equation 3.1.7 or Equation 3.1.8 fails. Thus, up to shrinking , either

(a) there exists an with such that for all in , where or

(b) there exist , with such that for all in .

If (a) holds, we fix and choose, among the different for which (a) holds, the one corresponding to the dominant term, i.e., such that is maximal, and among equal , the one such that is maximal. Then for large , the -norm of restricted to is eventually larger than half of the -norm of the dominant term on and hence diverges.

Now suppose that no as in (a) exists. Then we are in case (b), and and correspond to two different subsequences of , which, on , converge to two different functions, namely , resp. . Again, this contradicts -convergence of .

Proof of Theorem 3.2.1.

We first prove (2). Using Theorem 3.1.3, we find a function such that

for every , , and for which that limit exists. In particular, if the sequence converges pointwise (or uniformly), then the limit is . If converges in -norm, then converges for almost all by Lemma 3.2.6, and for those for which it does, the limit is . Therefore, is also the -limit.

We now consider (1) with the various notions of convergence. By replacing by (with obtained from part (2)), we may assume that if exists, then it is equal to .

Pointwise convergence: This can be expressed as

which is a -condition by Theorem 3.1.8.

Uniform convergence: Apply Theorem 3.1.2 to , where just sends to .

-convergence: The function

is of -class (since the square of the absolute value can be obtained by multiplying with the complex conjugate and then using Theorem 1.5.1), so this going to 0 is a -condition by Theorem 3.1.8.

-convergence: By Theorem 4.4.3 of Reference 12 -functions are locally constant almost everywhere. More precisely, there exists a definable set such that for every , , , and , the set has dimension less than and is locally constant on . Define to be equal to on and equal to outside . Then and have the same behavior concerning -convergence, but has the additional property that -convergence is equivalent to uniform convergence; this has already been dealt with above.

3.3. Stability under Fourier transformation for -functions of -class

That the Fourier transform of an -function of -class is again of -class follows directly from closedness under integration; see Reference 12, Corollary 4.3.1 for a version in the current context. For -functions of -class, this can be obtained using our formalism as follows. It seems that this result is new even for fixed . It compares to a real counterpart about stability under Fourier transformation of Theorem 8.10 of Reference 5.

Theorem 3.3.1 (Stability under Fourier transformation).

Let be in for some and some definable set . Then there exists in such that the following holds for all in , for all , and for each .

If the map is -integrable on , then is the Fourier transform of .

Proof.

The Fourier transform of an -function is defined as the -limit of the sequence of functions

indexed by (where runs over and is the scalar product). By Theorem 1.5.1 (and the comment below that theorem), is of -class, so the limit is of -class by Theorem 3.2.1.

Note that exactly the same proof also works for -functions for .

4. An application to orbital integrals

Some of the motivation for this paper came from harmonic analysis on reductive groups, and we conclude it by an illustration: we apply the powerful techniques developed above to Fourier transforms of orbital integrals.

4.1. Preliminaries

Let be a connected reductive algebraic group over a local field , with Lie algebra . We need to represent and as members of a family of definable sets, and we do it as in Reference 18, Section 2. Let us briefly review this setting and the notation.

4.1.1. Classification of reductive groups

We start with the fixed choices as in Reference 18, Section 2.1. The fixed choices are constructed so that for each Galois extension of local fields whose Galois group matches the data from the fixed choice (see item (1) below) (and with residue characteristic of not or ), a fixed choice determines (uniquely up to isomorphism) a split reductive group defined over and an action of on .

A fixed choice consists of:

(1)

an abstract finite group with an enumeration of its elements, a fixed subgroup , and a distinguished element . Later when we associate a split reductive algebraic -group with our fixed choice, the group is interpreted as the Galois group of a Galois extension of with being the inertia subgroup, and is interpreted as a generator of the Galois group of the maximal unramified subextension;

(2)

a Dynkin diagram with an action of (this is interpreted as the Dynkin diagram of );

(3)

two abstract lattices of the rank determined by the Dynkin diagram with action of (these are interpreted as the character and cocharacter lattices of a split maximal torus in ).

Each fixed choice yields a finite set of isomorphism classes of connected reductive groups defined over , as described in Reference 18, Section 6. Those isomorphism classes corresponding to are distinguished by the elements of a definable parameter set (which we will call the “cocycle space”). An element of encodes (a) an extension over which splits (which has to have with the inertia subgroup mapping to ) and (b) the Galois cocycle that determines as a twist of . We observe that if is unramified over , then itself is also determined uniquely by the fixed choices. The precise details of this construction will not be important here, and we refer the reader to Reference 18, Section 2 for the details.

Naturally, not all fixed choices of , , etc., give rise to a family of reductive groups (e.g. has to be solvable in order to be a Galois group of an extension of local fields, etc.). From here onwards, we are only interested in fixed choices that do satisfy the compatibility conditions that ensure they give rise to a family of nonempty reductive groups.

Here it is important for us to note that even though the set is infinite for all , there are only finitely many isomorphism classes of connected reductive groups corresponding to the same fixed choices, as apparent by inspection from Reference 18, Section 6, and these isomorphism classes are parameterized independently of . Indeed, in the case (1) of Reference 18, Section 6, there are possible invariants in the Brauer group with denominator ; in the cases (2) and (3), excluding type , there are two possible ramified quadratic extensions when . If contains any factors of type , there are still finitely many possibilities, corresponding to distinct degree ramified extensions with Galois group .

In summary, we have the following.

Proposition 4.1.1 (Reference 18, Section 2.2.2, Reference 19, Theorem 4).

For every fixed choice there exist , a definable set , and definable families and such that for every local field of residue characteristic greater than the following holds.

For , the set is the set of -points of a connected reductive group with absolute root datum determined by , and the set is the set of -points of the Lie algebra of .

In the following, we deal with the sets , , etc., while implicitly thinking of them as elements of the definable families and . In particular, by a “definable subset ”, we really mean a definable subset , where .

4.1.2. Orbital integrals

Our goal is to provide uniform estimates for Fourier transforms of the orbital integrals on the Lie algebra.

Let . An orbital integral at is the distribution on , defined by

where is the centralizer of , and is a -invariant measure on . Note that such a measure is unique up to a constant multiple and is uniquely determined by the normalization of the Haar measures on and on . The fact that orbital integrals converge and thus are well-defined distributions (which is a nontrivial question when is not regular semisimple) was proved by Deligne and Ranga Rao for the fields of characteristic zero, and by G. McNinch for of sufficiently large positive characteristic.

In this paper, however, we focus on regular semisimple elements . We will denote the set of regular semisimple elements in by . By definition, the centralizer of a regular semisimple element is a maximal torus. There are finitely many conjugacy classes of maximal tori in , with the upper bound on their number depending only on the fixed choices determining , so in order to pin down the normalization of the measures on all regular semisimple orbits it suffices to pick a Haar measure on and on the (finitely many) representatives of conjugacy classes of the maximal tori. The classical statements, quoted below, on the boundedness of the Fourier transforms of orbital integral are independent of these normalizations of the measures. However, the main point of this section is to discuss how these bounds depend on the field , and therefore we need a consistent normalization of measures. Following e.g. Kottwitz’s note on bounds for orbital integrals Reference 26, Appendix A, we choose on and on all its maximal tori the so-called canonical measures, which assign volume to the canonical parahoric subgroup in the sense of Gross. With this normalization of measures, for , let denote the orbital integral at , considered as a distribution on .

4.1.3. Fourier transform

Given an additive character of and a nondegenerate -invariant bilinear form on , the Fourier transform for a Schwartz-Bruhat function is defined as

Later we will allow the character to vary through the collection of level zero characters satisfying Definition 1.2.6. We will also need the bilinear form to be a definable function of and . Such a form exists for all of characteristic zero or sufficiently large positive characteristic by Reference 1; see also Reference 8, Section 4.1.

4.1.4. The bounds on Fourier transforms of orbital integrals

Let be the discriminant of ,

where is the root system of .

It is a well-known theorem of Harish-Chandra that for any , the orbital integral at (as a distribution on the space of locally constant compactly supported functions on ) is represented by a function , defined and locally constant on the set of regular semisimple elements ; this function, extended by zero to all of , is locally integrable on . Moreover, the function is locally bounded on . Following Reference 24, we will call such a function a “nice” function on .

R. Herb generalized the boundedness result, also making it uniform in in some sense Reference 23. To state Herb’s theorem precisely, we first need a definition. Let be a Cartan subalgebra of , let be the torus in such that , and let be the split component of . Fix an arbitrary open compact subgroup of , and define a function on by

where is the quotient measure obtained from the Haar measures on and on , and is the normalized Haar measure on . This definition is due to Harish-Chandra, and in his definition the normalization of the measure on does not matter. We note that is compact; we have already chosen a normalization of the measure on . Let us normalize the measure on so that the volume of is . This gives a collection of measures on , where are the split components of the representatives of the conjugacy classes of maximal tori in .

For any fixed collection of normalizations of measures on consistent with the normalization of measures on the maximal tori, Harish-Chandra proved the following.

Lemma 4.1.2 (Reference 22, Lemma 7.9).

Let and be two Cartan subalgebras in . For , ,

Herb proved that the function is uniformly bounded on compact sets.

Theorem 4.1.3 (Reference 23, Theorem 1.3).

Let be a Cartan subalgebra of , and let be a compact subset of the set of regular elements in . Then

Combining this theorem with Harish-Chandra’s lemma, one obtains that Fourier transforms of regular semisimple orbital integrals are uniformly bounded on compact sets in a given Lie algebra over a given field.

4.2. Uniformity: The questions

Our goal is to answer a series of related natural questions:

(1)

How does the bound in Theorem 4.1.3 vary for a family of compact sets ?

(2)

Given a definable set , how does the bound in Theorem 4.1.3 depend on the field ? More specifically, do the same bounds apply for the fields of positive characteristic and of characteristic zero with isomorphic residue fields?

(3)

Given a definable set , how does the bound depend on the cardinality of the residue field as varies through the family of all completions of a given global field?

4.3. Uniform bounds for Fourier transforms of regular semisimple orbital integrals

Here we prove two results, analogous to Theorems 1 and 2 of Reference 26, Appendix B, for the Fourier transforms of regular semisimple orbital integrals, which answer the above questions.

Theorem 4.3.1.

Let be a fixed choice as in Section 4.1.1. Let be the corresponding cocycle space, and let be the definable family of reductive groups as in Proposition 4.1.1, with the corresponding family of Lie algebras. Let be a definable family of subsets of , understood as in the remark after Proposition 4.1.1. Then:

(1)

There exist constants , , and that depend only on and on the formula defining , such that for each non-Archimedean local field with residue characteristic at least , the following holds. Let be any connected reductive algebraic group over corresponding to the fixed choice . Then for all with , for all ,

(2)

Moreover, if some constants and provide a bound for all fields of sufficiently large positive characteristic, then there exist constants and (as above, depending only on the data defining the group and the family of subsets ) such that for all local fields of characteristic zero and residue characteristic larger than , the bound

holds for all with , for all , and vice versa, i.e., with positive characteristic and characteristic zero swapped.

In order to prove this theorem, we need a lemma establishing that orbital integrals with respect to the canonical measure are motivic. The fact that orbital integrals are motivic functions was first proved for semisimple orbital integrals in Reference 11, but only for unramified reductive groups. This result was extended to all orbital integrals and all connected reductive groups in Reference 8, but the measure on the orbits used there was not necessarily the canonical measure. Further, this lemma was needed for all semisimple elements, not necessarily regular, in Reference 26, Appendix B, but at that time we could only prove it up to a uniformly bounded factor (denoted by in Reference 26, Appendix B). Now we include the proof in full generality for completeness, even though in this paper we need the statement only for regular semisimple elements.

First, we need one more notation. We observe that the connected component of a centralizer of a semisimple element in is the set of -points of a connected reductive algebraic group, and there is a finite list of possible root data associated with such groups; cf. Reference 26, Section A.2. In particular, there are finitely many fixed choices giving rise to split forms of such centralizers and for each , a corresponding cocycle space parameterizing all their forms. We let be the disjoint union of all these cocycle spaces (considered as a definable set); this yields a definable family which has the property that all centralizers of elements of arise as for some . We also note that the condition on the pair stating that the centralizer of is isomorphic to as reductive groups over is a definable condition.

Lemma 4.3.2.

Let be as in Theorem 4.3.1 above, and let be a family of subgroups of as above. Then there exists an integer and a function in such that for any definable family of test functions on of -class there exists a function in such that for all local fields of residue characteristic greater than , the following holds.

For every , let be the corresponding connected reductive group with Lie algebra . Let be a semisimple element with centralizer isomorphic to for some . Then

We recall from Section 4.1.2 that denotes the orbital integral at with respect to the canonical measure.

Remark 4.3.3.

We note that everywhere in this lemma, we could replace -class with -class; see Remark 1.2.10.

Proof.

Recall that we deal with the sets , , etc., while thinking of them as elements of a definable family: , for any . Since the centralizer of is a definable set (with its defining formulas using as a parameter), we see that the condition that the centralizer of is isomorphic to for some gives a definable subset of . In particular, there exists a definable family of subsets of such that

where denotes the centralizer of under the adjoint action of .

By Reference 19, Theorem 6, there exist a function and a family of motivic measures on the groups such that the canonical measure on satisfies

Now, if is the quotient measure on the orbit of , where is the canonical measure on , then we can represent it as

Consider the quotient measure . It is a quotient of two motivic measures, and thus it is motivic by exactly the same argument as the one of Reference 26, Lemma 14.15. The statement of the lemma follows.

4.3.1. Proof of Theorem 4.3.1

Let be an arbitrary regular element, and let be its centralizer (which is a torus independent of , with ). First, note that the statement of Theorem 4.1.3 is independent of the normalization of the measure on . More precisely, it is independent of the normalization of the measure on , the maximal split subtorus of . We can choose such normalization that the volume of the compact torus is equal to . Then we have

where is the resulting quotient measure. In particular, for the choice of measure that corresponds to the canonical measures (as above) on and on , we obtain, by Theorem 4.1.3, that is bounded on .

By Lemma 4.3.2, and since is a function of -class (generalized to allow for as in Reference 8, B.3.1), the right hand side of Equation 4.3.3 is “of -class up to -constant”; i.e., we can write Equation 4.3.3 in the form

where is a -class function (also generalized as in Reference 8, B.3.1), , and the parameter determines the isomorphism class of the centralizer of . Since here is assumed to be regular, its centralizer is a torus, and corresponds to the choice of the Cartan subalgebra that contains .

Let us consider the function

By Theorem 4.1.3, we know that for every and every pair , the function is bounded (as a function of and ) on .

Next we use the observation from Section 4.1.1 that for every fixed choice there are, in fact, finitely many possible values of that give rise to distinct Lie algebras. Hence, for every , the function is bounded. Then by Theorem 2.1.1, we have

where the integers and depend only on the fixed choices, which completes the proof of part (1). Indeed, since we are working in sufficiently large residue field characteristic, the remark just below Theorem 2.1.1 allows us to use an integer instead of a definable in the exponent.

Part (2) follows immediately from Reference 10, Theorem 3.2, with a single function from part (1) on the left, and .

Finally, in the spirit of Reference 26, Appendix B, we state an easy corollary that might be useful.

Theorem 4.3.4.

Let be a connected reductive algebraic group over a number field . Let be a definable family of definable subsets of . Then there exist constants and that depend only on the global model of and the formulas defining the subsets such that for all with , for all but finitely many places , for all with , for all ,

where is the cardinality of the residue field of .

Proof.

Since there are finitely many possibilities for the root data of the groups as varies over the finite places of , this theorem immediately follows from Theorem 4.3.1, exactly in the same way as Theorem 14.1 of Reference 26 follows from Theorem 14.2.

Appendix A. Adding constants to the language

For simplicity, all results in this paper are stated using a language containing only basic constant symbols. However, in all our results, more constants can be added to the language for free, by usual model theoretic arguments and using that all our results are stated “in families”; the idea is that any constant can be replaced by a family parameter. (Note however that the transfer principles need some extra care; see below.) For example, one can fix a ring , assume that each local field is equipped with an injection , and assume that the language contains constant symbols for the elements of , or we can assume that each comes with a uniformizer for which the language contains a constant symbol. Since not all readers might be familiar with this, we recall how this works.

Fix a set of (new) constant symbols (in any of the sorts) and set (where is the language introduced in Definition 1.2.3). We write for the set of fields which come equipped with (arbitrary) interpretations of the constant symbols in . Fix an arbitrary subset , and define and analogously to Definition 1.2.1. (Usually, in model theory, one would require the subset to be defined by a set of first order sentences, but here, we do not even need that.)

Proposition A.1.

Each of the results listed in Section 1.1, with the exception of the transfer results, also holds if one replaces by and by everywhere.

Proof.

The results from Section 4 build on the ones from the previous sections, so it suffices to verify that the proposition applies to the results from those previous sections. First note that in each of those, only appears in the form “for each (or variants of that), so in particular, we may as well assume that .

Next, since each theorem involves only a finite number of “input objects” (definable sets, definable functions, -function, etc.) and each input object involves only finitely many formulas, only finitely many of the constants from are used, so we may as well assume that we only added a finite tuple of constants to the language.

Finally, note that each result involves a parameter set (always denoted by ). We introduce a new tuple of variables of the same sorts as , we let be the corresponding cartesian product of sorts, we replace the parameter set by , and in all input objects, we replace all occurrences of by . Then we apply the -version of the theorem. If the theorem has some “output objects” (this is the case for all results but Lemma 3.2.6), then in that output object, we replace the by again. The -objects obtained in this way have the desired properties.

For the transfer principles to hold with additional constants in the language, one needs extra control on these constants via suitable axioms specifying the subset . Instead of trying to formulate this in the most general possible way, we just formulate it for collections of constants that are commonly used (e.g. in Reference 7): Let be the ring of integers of some fixed number field, set , and let be the set of fields together with a ring homomorphism . Recall that an comes with a uniformizer . We consider an as an -structure by interpreting elements of via the continuous ring homomorphism which is equal to on and which sends to . Given an element , we write for its interpretation in .

Proposition A.2.

Let and be as described right above this proposition. Then each transfer result listed in Section 1.1 still holds if one replaces by and by everywhere.

Proof.

Theorem 1.3.3, which we cited from Reference 15, is actually stated in the language and for fields in in Reference 15 and follows from Denef–Pas quantifier elimination. Since all other transfer results are deduced from Theorem 1.3.3 (and from results to which Proposition A.1 applies), they hold accordingly.

Finally, let us make precise the comments on below Theorems 2.1.1 and 2.1.3. If and are as described right above Proposition A.2, then any -definable can still be bounded by for some integers and depending on . However, for more general and such integers and may no longer yield bounds for uniformly in .

Remark A.3.

All results in this paper also stay true if one consequently expands the language by some analytic structure as in Reference 13Reference 14 or by arbitrary additional structure on the residue ring sorts . Indeed, by Reference 12, Section 4.7, this is true for the results from Reference 12, on which the present paper is based.

Acknowledgment

The authors thank the referee for valuable comments.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. Summary of the results
    2. 1.2. Basic definitions: Functions and loci of -class
    3. Definition 1.2.1 (Local fields).
    4. Definition 1.2.3.
    5. Definition 1.2.4 (Definable sets and functions).
    6. Definition 1.2.5 (Functions of -class).
    7. Definition 1.2.6 (Additive characters).
    8. Definition 1.2.7 (Functions of -class).
    9. Definition 1.2.8 (Loci).
    10. 1.3. The locus formalism, transfer, and constancy
    11. Proposition 1.3.1 (Basic operations on loci).
    12. Corollary 1.3.2 (Constancy).
    13. Theorem 1.3.3 (Transfer).
    14. Corollary 1.3.4 (Transfer for constancy).
    15. Example 1.3.5.
    16. Example 1.3.6.
    17. 1.4. Eventual behavior and local constancy
    18. Proposition 1.4.1 (Eventual behavior).
    19. Proposition 1.4.2.
    20. Corollary 1.4.3 (Local constancy).
    21. Corollary 1.4.4 (Transfer for local constancy).
    22. 1.5. Integration
    23. Theorem 1.5.1 (Integration, 12, Theorems 4.1.1, 4.4.2).
  3. 2. Bounds and approximate suprema
    1. 2.1. Results about bounds and approximate suprema
    2. Theorem 2.1.1 (Bounds).
    3. Theorem 2.1.2 (Bounds).
    4. Theorem 2.1.3 (Approximate suprema).
    5. 2.2. Proof of Theorem 2.1.3
    6. Proposition 2.2.1 (12, Proposition 4.6.1).
    7. Corollary 2.2.2.
    8. Lemma 2.2.3.
    9. Lemma 2.2.4.
  4. 3. Limits of -functions
    1. 3.1. Pointwise limits and continuity
    2. Theorem 3.1.1 (-limits).
    3. Theorem 3.1.2 (-limits).
    4. Theorem 3.1.3 (Limits).
    5. Theorem 3.1.4 (Limits).
    6. Corollary 3.1.5 (Continuity).
    7. Proposition 3.1.6 (Continuous extension).
    8. Corollary 3.1.7 (Transfer principle for limits and continuity).
    9. Theorem 3.1.8 (Limits).
    10. 3.2. Limits of functions
    11. Theorem 3.2.1 (Limits of sequences of functions).
    12. Corollary 3.2.2 (-integrability).
    13. Corollary 3.2.3 (The “almost everywhere” quantifier).
    14. Corollary 3.2.4 (Transfer principles for limits of functions).
    15. Corollary 3.2.5 (Transfer principles for -norms).
    16. Lemma 3.2.6.
    17. 3.3. Stability under Fourier transformation for -functions of -class
    18. Theorem 3.3.1 (Stability under Fourier transformation).
  5. 4. An application to orbital integrals
    1. 4.1. Preliminaries
    2. Proposition 4.1.1 (18, Section 2.2.2, 19, Theorem 4).
    3. Lemma 4.1.2 (22, Lemma 7.9).
    4. Theorem 4.1.3 (23, Theorem 1.3).
    5. 4.2. Uniformity: The questions
    6. 4.3. Uniform bounds for Fourier transforms of regular semisimple orbital integrals
    7. Theorem 4.3.1.
    8. Lemma 4.3.2.
    9. Theorem 4.3.4.
  6. Appendix A. Adding constants to the language
    1. Proposition A.1.
    2. Proposition A.2.
  7. Acknowledgment

Mathematical Fragments

Definition 1.2.1 (Local fields).

Let be the collection of all non-Archimedean local fields equipped with a uniformizer for the valuation ring of . (So more formally, elements of are pairs .) Here, by a non-Archimedean local field we mean a finite extension of or of for any prime .

Given an integer , let be the collection of such that has characteristic either or at least .

We will use the notation “for all to mean “there exists an such that for all ”.

Definition 1.2.3.

The language has sorts (the valued field), (the value group), and for each (the residue rings), with the ring language on and on each , the ordered abelian group language on , the valuation map , and generalized angular component maps sending to . See Reference 12, Section 2.2 for more details.

Equation (1.2.1)
Definition 1.2.5 (Functions of -class).

Let be a definable set. A collection of functions is called a function of -class (or simply a -function) on if there exist integers , , and , nonzero integers , definable functions and , and definable sets such that for all and all ,

where , for some and .

We write to denote the ring of -functions on .

Definition 1.2.6 (Additive characters).

For any local field , let be the set of the additive characters on that are trivial on the maximal ideal of and nontrivial on .

Definition 1.2.7 (Functions of -class).

Let be a definable set. A collection of functions for and is called a function of -class (or a -function) on if there exist integers and , functions in , definable sets , and definable functions and for , such that for all , all , and all ,

where for and , by abuse of notation, is defined as with any element in such that , which is well defined since is constant on . We write to denote the ring of -functions on .

Notation 1.2.9 (Conditions).

Fix some integers and consider a collection of conditions on elements . We call a definable condition if the collection of sets is definable. Analogously, we call a condition of -class or of -class, respectively, if is a locus of the corresponding class. Again, we also say -condition or -condition for short.

Remark 1.2.10.

In the remainder of this section and in all of Sections 2 and 3, literally every result that is stated for -functions and -loci is also valid if one replaces all occurrences of by ; i.e., if no input object of a result uses the additive character , then does not appear in the output objects either. (This is obvious from the proofs.)

Proposition 1.3.1 (Basic operations on loci).

In the following, runs over a definable set , and runs over a definable set .

(1)

Any definable condition is a -condition.

(2)

If and are -conditions on , then so are and .

(3)

If is a -condition on and , then is a -condition on .

(4)

A -condition stays a -condition when considered as a condition on and that is independent of .

Corollary 1.3.2 (Constancy).

Let be in , for some definable sets and . Then the set of for which the function is constant is a -locus.

Theorem 1.3.3 (Transfer).

Suppose that is a -condition on , for some definable set . Then there exists an such that for all with residue field characteristics at least , the following holds.

If holds for all and all , then for any whose residue field is isomorphic to the one of , holds for all and all .

Corollary 1.3.4 (Transfer for constancy).

Let and be definable sets, and let be in . Then there exists such that, for any of residue field characteristic , the truth of the following statement depends only on (the isomorphism class of) the residue field of :

Example 1.3.5.

Transfer (in the style of Theorem 1.3.3) does not hold for the statement

Indeed: The image consists of the -th roots of unity, where is the residue characteristic of . If itself has characteristic , then the entire image consists only of the -th roots of unity, whereas if has characteristic , then contains all -th roots of unity for all . Thus 1.3.1 holds if and only if has positive characteristic.

Example 1.3.6.

We give an example that, in general, -loci are not closed under existential quantification. If -loci would be closed under existential quantification, then in particular would be a -locus, i.e., with . However, for any and any , the zero set of is ultimately periodic; i.e., there exist such that for , whether is zero or not depends only on . Indeed, this follows from Proposition 1.4.2: For large enough and in a fixed congruence class modulo some suitable , we have

for finitely many nonnegative integers , rational , and complex , and such a function is either constant equal to zero (namely when all are zero) or has only finitely many zeros, e.g. by Reference 7, Lemma 2.1.8.

Proposition 1.4.1 (Eventual behavior).

If is a -condition on running over a definable set and on running over , then the condition given by

is a -condition on . Moreover, can be chosen to depend definably on ; i.e., there exists a definable function such that for each , each and each , if holds, then holds for all .

Proposition 1.4.2.

Let and be definable and let be a -function on . Then there exist a definable bijection commuting with the projection to and a finite definable partition of such that we have the following for each part . We set .

(1)

The set is a Presburger Cell over , i.e.,

where and are integers, is the projection of , is either a definable function or constant equal to , and is either a definable function or constant equal to .

(2)

If , there exists an integer and a definable function such that for all , and similarly for . (And without loss, can be taken the same for all and .)

(3)

There are finitely many functions in and distinct pairs with nonnegative integers and rational numbers such that for all , for each , and each ,

(The map is called a reparameterization.)

Corollary 1.4.3 (Local constancy).

Let be in , for some definable sets and for some . Then we have the following (where for each , on we put the topology induced by the one on ).

(1)

The set of for which the function is locally constant is a -locus.

(2)

The set of such that the function is constant on a neighborhood of is a -locus. Moreover, the radius of constancy can be chosen definably; i.e., there exists a definable function such that for all , for all , for all , and for all , if is constant on a neighborhood of , then it is constant on the intersection of with the ball of valuative radius around in .

Corollary 1.4.4 (Transfer for local constancy).

Corollary 1.3.4 still holds if one replaces “constant” by “locally constant”.

Theorem 1.5.1 (Integration, Reference 12, Theorems 4.1.1, 4.4.2).

Let be in , for some definable sets and .

(1)

The set of such that is -integrable over is a -locus.

(2)

There exists a function in such that

whenever is -integrable over . In more detail, for every , every , and every , if is -integrable, then .

Theorem 2.1.1 (Bounds).

Let be in , where is a definable set and . Then there exist an integer and a definable such that for all , all , and all , the following holds.

If the function

is bounded on , then one actually has

where and where is the norm on .

Theorem 2.1.2 (Bounds).

Let be in , where and are definable sets.

(1)

The set of such that the function is bounded is a -locus.

(2)

There exists a definable function such that for every , every , and every , if the function

is bounded on , then one has

Theorem 2.1.3 (Approximate suprema).

Let be in , where and are definable sets. Then there exist a definable and a function in taking nonnegative real values such that the following holds: For each , each , and each for which the function

is bounded on , one has

Proposition 2.2.1 (Reference 12, Proposition 4.6.1).

Let be an integer, let and be definable, and let be in . Write for variables running over and for variables running over . Then there exist integers , , a definable surjection over , definable functions , and functions in for , such that the following conditions hold for each and each ; we omit the indices to keep the notation lighter.

(1)

One has

(2)

If one sets, for ,

and

then

where the volume is taken with respect to the Haar measure on .

Here, as usual, is the set of such that lies in , and for the map to “be over means that makes a commutative diagram with the natural maps to .

Equation (2.2.1)
Corollary 2.2.2.

Let , , and be definable sets and let be in . Then there exist a nonnegative integer , a definable surjection

over , and a nonnegative real valued function in such that for all , , , and , one has

where is the fiber of above and where by the right hand inequality, we in particular mean that the supremum is finite.

Lemma 2.2.3.

Suppose that for , we have and . Then there exist positive integers and such that the following holds for every tuple and every . Suppose that the function

is bounded on . Then we have

Lemma 2.2.4.

Suppose that for , we have and . Then there exist positive integers and and Presburger definable functions with for all and and such that the following holds for every tuple , every , and every . Consider the function

then we have

Equation (2.2.3)
Equation (2.2.4)
Equation (2.2.5)
Equation (2.2.6)
Equation (2.2.7)
Equation (2.2.8)
Equation (2.2.10)
Theorem 3.1.1 (-limits).

Let be in , where is a definable set and where .

(1)

The set is a -locus.

(2)

There exists a rational number and a definable function , such that for each , for each , and for each , the following holds.

If one has

then one actually has

Theorem 3.1.2 (-limits).

Let be in , where and are definable sets, and let be a surjective definable function.

(1)

The set is a -locus.

(2)

There exists a rational number and a definable function such that for all , for each , and for each , the following holds.

If one has

then one actually has

Theorem 3.1.3 (Limits).

Let be in for some definable set .

(1)

The set is a -locus.

(2)

There exists in such that the following holds for all , for each and each .

then

Theorem 3.1.4 (Limits).

Let be in , where and are definable sets, and let be a surjective definable function. Then the set is a -locus.

Corollary 3.1.5 (Continuity).

Let be in , where and are definable sets and . Then

and

are -loci.

Proposition 3.1.6 (Continuous extension).

Let be in , where and are definable sets and and suppose that is clopen (i.e., is open and closed for every ). Then can be extended to a -function on such that for every , if exists, then is continuous at .

Namely, with all indices, for every , for every , and for every , if exists, then is continuous at .

Corollary 3.1.7 (Transfer principle for limits and continuity).

Let be in , where and are definable sets, and let be a definable function.

Then there exists such that, for any of residue characteristic , the truth of each of the following statements depends only on (the isomorphism class of) the residue field of ; here, in the third and fourth statements, we assume for some .

Theorem 3.1.8 (Limits).

Let be in , where is a definable set. Then we have the following.

(1)

The condition exists” is a -condition on .

(2)

There exists a function , a definable function , and a rational number such that for every , every , and every , if the limit exists, then

for every . In particular, the limit is equal to .

Equation (3.1.5)
Equation (3.1.6)
Equation (3.1.7)
Equation (3.1.8)
Equation (3.1.9)
Theorem 3.2.1 (Limits of sequences of functions).

Let be in

for some definable set and some . Given , , and , consider the sequence of functions (on )

indexed by integers .

(1)

For each of the notions of convergence listed below, the set of such that the sequence 3.2.1 of functions converges in that sense is a -locus. The notions of convergence are: pointwise convergence, uniform convergence, -convergence, -convergence.

(2)

There exists a in such that the following holds for all , for each , and for each .

If the sequence 3.2.1 of functions converges pointwise or uniformly or in -norm for any , then the function

is the limit of this sequence.

Corollary 3.2.2 (-integrability).

Fix and let be in . Then the set of for which is -integrable is a -locus.

Corollary 3.2.3 (The “almost everywhere” quantifier).

Suppose that is a -condition, where runs over a definable set and runs over . Then

is a -condition on . (Here, by “almost all”, we mean that the complement has measure .)

Corollary 3.2.4 (Transfer principles for limits of functions).

Let be in for some definable set and some . Given , , and , consider the sequence of functions (on )

indexed by integers .

Fix a notion of convergence among: pointwise convergence / uniform convergence / convergence in -norm / in -norm.

Then there exists such that, for any of residue characteristic , the truth of each of the following statements depends only on (the isomorphism class of) the residue field of .

Corollary 3.2.5 (Transfer principles for -norms).

Let be in for some definable set and some . Then there exists such that, for any of residue characteristic , the truth of each of the following four statements depends only on (the isomorphism class of) the residue field of .

Lemma 3.2.6.

Fix and let be in for some definable set . Also fix , , and . If the sequence of functions (on ) converges in -norm for , then converges for almost all .

Theorem 3.3.1 (Stability under Fourier transformation).

Let be in for some and some definable set . Then there exists in such that the following holds for all in , for all , and for each .

If the map is -integrable on , then is the Fourier transform of .

Proposition 4.1.1 (Reference 18, Section 2.2.2, Reference 19, Theorem 4).

For every fixed choice there exist , a definable set , and definable families and such that for every local field of residue characteristic greater than the following holds.

For , the set is the set of -points of a connected reductive group with absolute root datum determined by , and the set is the set of -points of the Lie algebra of .

Theorem 4.1.3 (Reference 23, Theorem 1.3).

Let be a Cartan subalgebra of , and let be a compact subset of the set of regular elements in . Then

Theorem 4.3.1.

Let be a fixed choice as in Section 4.1.1. Let be the corresponding cocycle space, and let be the definable family of reductive groups as in Proposition 4.1.1, with the corresponding family of Lie algebras. Let be a definable family of subsets of , understood as in the remark after Proposition 4.1.1. Then:

(1)

There exist constants , , and that depend only on and on the formula defining , such that for each non-Archimedean local field with residue characteristic at least , the following holds. Let be any connected reductive algebraic group over corresponding to the fixed choice . Then for all with , for all ,

(2)

Moreover, if some constants and provide a bound for all fields of sufficiently large positive characteristic, then there exist constants and (as above, depending only on the data defining the group and the family of subsets ) such that for all local fields of characteristic zero and residue characteristic larger than , the bound

holds for all with , for all , and vice versa, i.e., with positive characteristic and characteristic zero swapped.

Lemma 4.3.2.

Let be as in Theorem 4.3.1 above, and let be a family of subgroups of as above. Then there exists an integer and a function in such that for any definable family of test functions on of -class there exists a function in such that for all local fields of residue characteristic greater than , the following holds.

For every , let be the corresponding connected reductive group with Lie algebra . Let be a semisimple element with centralizer isomorphic to for some . Then

We recall from Section 4.1.2 that denotes the orbital integral at with respect to the canonical measure.

Equation (4.3.3)
Proposition A.1.

Each of the results listed in Section 1.1, with the exception of the transfer results, also holds if one replaces by and by everywhere.

Proposition A.2.

Let and be as described right above this proposition. Then each transfer result listed in Section 1.1 still holds if one replaces by and by everywhere.

References

Reference [1]
Jeffrey D. Adler and Alan Roche, An intertwining result for -adic groups, Canad. J. Math. 52 (2000), no. 3, 449–467, DOI 10.4153/CJM-2000-021-8. MR1758228,
Show rawAMSref \bib{adler-roche}{article}{ author={Adler, Jeffrey D.}, author={Roche, Alan}, title={An intertwining result for $p$-adic groups}, journal={Canad. J. Math.}, volume={52}, date={2000}, number={3}, pages={449--467}, issn={0008-414X}, review={\MR {1758228}}, doi={10.4153/CJM-2000-021-8}, }
Reference [2]
James Ax and Simon Kochen, Diophantine problems over local fields. I, Amer. J. Math. 87 (1965), 605–630, DOI 10.2307/2373065. MR0184930,
Show rawAMSref \bib{AK1}{article}{ author={Ax, James}, author={Kochen, Simon}, title={Diophantine problems over local fields. I}, journal={Amer. J. Math.}, volume={87}, date={1965}, pages={605--630}, issn={0002-9327}, review={\MR {0184930}}, doi={10.2307/2373065}, }
Reference [3]
W. Casselman, J. Cely, and T. Hales, The spherical Hecke algebra, partition functions, and motivic integration, arXiv:1611.05773, to appear in Trans. Amer. Math. Soc., DOI: https://doi.org/10.1090/tran/7465.,
Show rawAMSref \bib{Casselman-Cely-Hales}{article}{ author={Casselman, W.}, author={Cely, J.}, author={Hales, T.}, title={The spherical {H}ecke algebra, partition functions, and motivic integration}, journal={arXiv:1611.05773, to appear in Trans. Amer. Math. Soc., DOI: https://doi.org/10.1090/tran/7465}, }
Reference [4]
Raf Cluckers, Presburger sets and -minimal fields, J. Symbolic Logic 68 (2003), no. 1, 153–162, DOI 10.2178/jsl/1045861509. MR1959315,
Show rawAMSref \bib{CPres}{article}{ author={Cluckers, Raf}, title={Presburger sets and $p$-minimal fields}, journal={J. Symbolic Logic}, volume={68}, date={2003}, number={1}, pages={153--162}, issn={0022-4812}, review={\MR {1959315}}, doi={10.2178/jsl/1045861509}, }
Reference [5]
Raf Cluckers, Georges Comte, Daniel J. Miller, Jean-Philippe Rolin, and Tamara Servi, Integration of oscillatory and subanalytic functions, Duke Math. J. 167 (2018), no. 7, 1239–1309, DOI 10.1215/00127094-2017-0056. MR3799699,
Show rawAMSref \bib{CCMRS}{article}{ author={Cluckers, Raf}, author={Comte, Georges}, author={Miller, Daniel J.}, author={Rolin, Jean-Philippe}, author={Servi, Tamara}, title={Integration of oscillatory and subanalytic functions}, journal={Duke Math. J.}, volume={167}, date={2018}, number={7}, pages={1239--1309}, issn={0012-7094}, review={\MR {3799699}}, doi={10.1215/00127094-2017-0056}, }
Reference [6]
Raf Cluckers, Clifton Cunningham, Julia Gordon, and Loren Spice, On the computability of some positive-depth supercuspidal characters near the identity, Represent. Theory 15 (2011), 531–567, DOI 10.1090/S1088-4165-2011-00403-9. MR2833466,
Show rawAMSref \bib{CCGordonS}{article}{ author={Cluckers, Raf}, author={Cunningham, Clifton}, author={Gordon, Julia}, author={Spice, Loren}, title={On the computability of some positive-depth supercuspidal characters near the identity}, journal={Represent. Theory}, volume={15}, date={2011}, pages={531--567}, issn={1088-4165}, review={\MR {2833466}}, doi={10.1090/S1088-4165-2011-00403-9}, }
Reference [7]
Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Integrability of oscillatory functions on local fields: transfer principles, Duke Math. J. 163 (2014), no. 8, 1549–1600, DOI 10.1215/00127094-2713482. MR3210968,
Show rawAMSref \bib{CGH}{article}{ author={Cluckers, Raf}, author={Gordon, Julia}, author={Halupczok, Immanuel}, title={Integrability of oscillatory functions on local fields: transfer principles}, journal={Duke Math. J.}, volume={163}, date={2014}, number={8}, pages={1549--1600}, issn={0012-7094}, review={\MR {3210968}}, doi={10.1215/00127094-2713482}, }
Reference [8]
Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Local integrability results in harmonic analysis on reductive groups in large positive characteristic (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 47 (2014), no. 6, 1163–1195, DOI 10.24033/asens.2236. MR3297157,
Show rawAMSref \bib{CGH2}{article}{ author={Cluckers, Raf}, author={Gordon, Julia}, author={Halupczok, Immanuel}, title={Local integrability results in harmonic analysis on reductive groups in large positive characteristic}, language={English, with English and French summaries}, journal={Ann. Sci. \'Ec. Norm. Sup\'er. (4)}, volume={47}, date={2014}, number={6}, pages={1163--1195}, issn={0012-9593}, review={\MR {3297157}}, doi={10.24033/asens.2236}, }
Reference [9]
Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Motivic functions, integrability, and applications to harmonic analysis on -adic groups, Electron. Res. Announc. Math. Sci. 21 (2014), 137–152, DOI 10.3934/era.2014.21.137. MR3356593,
Show rawAMSref \bib{CGH3}{article}{ author={Cluckers, Raf}, author={Gordon, Julia}, author={Halupczok, Immanuel}, title={Motivic functions, integrability, and applications to harmonic analysis on $p$-adic groups}, journal={Electron. Res. Announc. Math. Sci.}, volume={21}, date={2014}, pages={137--152}, issn={1935-9179}, review={\MR {3356593}}, doi={10.3934/era.2014.21.137}, }
Reference [10]
Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Transfer principles for bounds of motivic exponential functions, Families of automorphic forms and the trace formula, Simons Symp., Springer, [Cham], 2016, pp. 111–127. MR3675165,
Show rawAMSref \bib{CGH4}{article}{ author={Cluckers, Raf}, author={Gordon, Julia}, author={Halupczok, Immanuel}, title={Transfer principles for bounds of motivic exponential functions}, conference={ title={Families of automorphic forms and the trace formula}, }, book={ series={Simons Symp.}, publisher={Springer, [Cham]}, }, date={2016}, pages={111--127}, review={\MR {3675165}}, }
Reference [11]
Raf Cluckers, Thomas Hales, and François Loeser, Transfer principle for the fundamental lemma, On the stabilization of the trace formula, Stab. Trace Formula Shimura Var. Arith. Appl., vol. 1, Int. Press, Somerville, MA, 2011, pp. 309–347. MR2856374,
Show rawAMSref \bib{CHL}{article}{ author={Cluckers, Raf}, author={Hales, Thomas}, author={Loeser, Fran\c {c}ois}, title={Transfer principle for the fundamental lemma}, conference={ title={On the stabilization of the trace formula}, }, book={ series={Stab. Trace Formula Shimura Var. Arith. Appl.}, volume={1}, publisher={Int. Press, Somerville, MA}, }, date={2011}, pages={309--347}, review={\MR {2856374}}, }
Reference [12]
Raf Cluckers and Immanuel Halupczok, Integration of functions of motivic exponential class, uniform in all non-Archimedean local fields of characteristic zero (English, with English and French summaries), J. Éc. Polytech. Math. 5 (2018), 45–78. MR3732692,
Show rawAMSref \bib{CHallp}{article}{ author={Cluckers, Raf}, author={Halupczok, Immanuel}, title={Integration of functions of motivic exponential class, uniform in all non-Archimedean local fields of characteristic zero}, language={English, with English and French summaries}, journal={J. \'Ec. Polytech. Math.}, volume={5}, date={2018}, pages={45--78}, issn={2429-7100}, review={\MR {3732692}}, }
Reference [13]
R. Cluckers and L. Lipshitz, Fields with analytic structure, J. Eur. Math. Soc. (JEMS) 13 (2011), no. 4, 1147–1223, DOI 10.4171/JEMS/278. MR2800487,
Show rawAMSref \bib{CLip}{article}{ author={Cluckers, R.}, author={Lipshitz, L.}, title={Fields with analytic structure}, journal={J. Eur. Math. Soc. (JEMS)}, volume={13}, date={2011}, number={4}, pages={1147--1223}, issn={1435-9855}, review={\MR {2800487}}, doi={10.4171/JEMS/278}, }
Reference [14]
Raf Cluckers and Leonard Lipshitz, Strictly convergent analytic structures, J. Eur. Math. Soc. (JEMS) 19 (2017), no. 1, 107–149, DOI 10.4171/JEMS/662. MR3584560,
Show rawAMSref \bib{CLips}{article}{ author={Cluckers, Raf}, author={Lipshitz, Leonard}, title={Strictly convergent analytic structures}, journal={J. Eur. Math. Soc. (JEMS)}, volume={19}, date={2017}, number={1}, pages={107--149}, issn={1435-9855}, review={\MR {3584560}}, doi={10.4171/JEMS/662}, }
Reference [15]
Raf Cluckers and François Loeser, Constructible exponential functions, motivic Fourier transform and transfer principle, Ann. of Math. (2) 171 (2010), no. 2, 1011–1065, DOI 10.4007/annals.2010.171.1011. MR2630060,
Show rawAMSref \bib{CLexp}{article}{ author={Cluckers, Raf}, author={Loeser, Fran\c {c}ois}, title={Constructible exponential functions, motivic Fourier transform and transfer principle}, journal={Ann. of Math. (2)}, volume={171}, date={2010}, number={2}, pages={1011--1065}, issn={0003-486X}, review={\MR {2630060}}, doi={10.4007/annals.2010.171.1011}, }
Reference [16]
Ju. L. Eršov, On the elementary theory of maximal normed fields (Russian), Dokl. Akad. Nauk SSSR 165 (1965), 21–23. MR0190140,
Show rawAMSref \bib{Ersov}{article}{ author={Er\v sov, Ju. L.}, title={On the elementary theory of maximal normed fields}, language={Russian}, journal={Dokl. Akad. Nauk SSSR}, volume={165}, date={1965}, pages={21--23}, issn={0002-3264}, review={\MR {0190140}}, }
Reference [17]
I. M. Gel′fand, M. I. Graev, and I. I. Pyatetskii-Shapiro, Representation theory and automorphic functions, translated from the Russian by K. A. Hirsch, reprint of the 1969 edition, Generalized Functions, vol. 6, Academic Press, Inc., Boston, MA, 1990. MR1071179,
Show rawAMSref \bib{GelfandGraevPyat}{book}{ author={Gel\cprime fand, I. M.}, author={Graev, M. I.}, author={Pyatetskii-Shapiro, I. I.}, title={Representation theory and automorphic functions}, series={Generalized Functions}, volume={6}, edition={translated from the Russian by K. A. Hirsch, reprint of the 1969 edition}, publisher={Academic Press, Inc., Boston, MA}, date={1990}, pages={xviii+426}, isbn={0-12-279506-7}, review={\MR {1071179}}, }
Reference [18]
Julia Gordon and Thomas Hales, Endoscopic transfer of orbital integrals in large residual characteristic, Amer. J. Math. 138 (2016), no. 1, 109–148, DOI 10.1353/ajm.2016.0005. MR3462882,
Show rawAMSref \bib{GordonHales}{article}{ author={Gordon, Julia}, author={Hales, Thomas}, title={Endoscopic transfer of orbital integrals in large residual characteristic}, journal={Amer. J. Math.}, volume={138}, date={2016}, number={1}, pages={109--148}, issn={0002-9327}, review={\MR {3462882}}, doi={10.1353/ajm.2016.0005}, }
Reference [19]
Julia Gordon and David Roe, The canonical measure on a reductive -adic group is motivic (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 50 (2017), no. 2, 345–355, DOI 10.24033/asens.2322. MR3621433,
Show rawAMSref \bib{gordon-roe}{article}{ author={Gordon, Julia}, author={Roe, David}, title={The canonical measure on a reductive $p$-adic group is motivic}, language={English, with English and French summaries}, journal={Ann. Sci. \'Ec. Norm. Sup\'er. (4)}, volume={50}, date={2017}, number={2}, pages={345--355}, issn={0012-9593}, review={\MR {3621433}}, doi={10.24033/asens.2322}, }
Reference [20]
Thomas C. Hales, Can -adic integrals be computed?, Contributions to automorphic forms, geometry, and number theory, Johns Hopkins Univ. Press, Baltimore, MD, 2004, pp. 313–329. MR2058612,
Show rawAMSref \bib{Hales1}{article}{ author={Hales, Thomas C.}, title={Can $p$-adic integrals be computed?}, conference={ title={Contributions to automorphic forms, geometry, and number theory}, }, book={ publisher={Johns Hopkins Univ. Press, Baltimore, MD}, }, date={2004}, pages={313--329}, review={\MR {2058612}}, }
Reference [21]
Thomas C. Hales, Orbital integrals are motivic, Proc. Amer. Math. Soc. 133 (2005), no. 5, 1515–1525, DOI 10.1090/S0002-9939-04-07740-8. MR2111953,
Show rawAMSref \bib{Hales}{article}{ author={Hales, Thomas C.}, title={Orbital integrals are motivic}, journal={Proc. Amer. Math. Soc.}, volume={133}, date={2005}, number={5}, pages={1515--1525}, issn={0002-9939}, review={\MR {2111953}}, doi={10.1090/S0002-9939-04-07740-8}, }
Reference [22]
Harish-Chandra, Admissible invariant distributions on reductive -adic groups, with a preface and notes by Stephen DeBacker and Paul J. Sally, Jr., University Lecture Series, vol. 16, American Mathematical Society, Providence, RI, 1999, DOI 10.1090/ulect/016. MR1702257,
Show rawAMSref \bib{hc:queens}{book}{ author={Harish-Chandra}, title={Admissible invariant distributions on reductive $p$-adic groups}, series={University Lecture Series}, volume={16}, edition={with a preface and notes by Stephen DeBacker and Paul J. Sally, Jr.}, publisher={American Mathematical Society, Providence, RI}, date={1999}, pages={xiv+97}, isbn={0-8218-2025-7}, review={\MR {1702257}}, doi={10.1090/ulect/016}, }
Reference [23]
Rebecca A. Herb, Bounds for Fourier transforms of regular orbital integrals on -adic Lie algebras, Represent. Theory 5 (2001), 504–523, DOI 10.1090/S1088-4165-01-00125-X. MR1870601,
Show rawAMSref \bib{herb}{article}{ author={Herb, Rebecca A.}, title={Bounds for Fourier transforms of regular orbital integrals on $p$-adic Lie algebras}, journal={Represent. Theory}, volume={5}, date={2001}, pages={504--523}, issn={1088-4165}, review={\MR {1870601}}, doi={10.1090/S1088-4165-01-00125-X}, }
Reference [24]
Robert E. Kottwitz, Harmonic analysis on reductive -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. MR2192014,
Show rawAMSref \bib{kottwitz:clay}{article}{ author={Kottwitz, Robert E.}, title={Harmonic analysis on reductive $p$-adic groups and Lie algebras}, conference={ title={Harmonic analysis, the trace formula, and Shimura varieties}, }, book={ series={Clay Math. Proc.}, volume={4}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2005}, pages={393--522}, review={\MR {2192014}}, }
Reference [25]
Kien Huu Nguyen, Uniform rationality of the Poincaré series of definable, analytic equivalence relations on local fields, arXiv:1610.07952, to appear in Trans. Amer. Math. Soc., DOI: https://doi.org/10.1090/tran/7357.,
Show rawAMSref \bib{Kien:rational}{article}{ author={Nguyen, Kien~Huu}, title={Uniform rationality of the {P}oincar{\' e} series of definable, analytic equivalence relations on local fields}, journal={arXiv:1610.07952, to appear in Trans. Amer. Math. Soc., DOI: https://doi.org/10.1090/tran/7357}, }
Reference [26]
Sug Woo Shin and Nicolas Templier, Sato-Tate theorem for families and low-lying zeros of automorphic -functions, Appendix A by Robert Kottwitz, and Appendix B by Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Invent. Math. 203 (2016), no. 1, 1–177, DOI 10.1007/s00222-015-0583-y. MR3437869,
Show rawAMSref \bib{ShinTemp}{article}{ author={Shin, Sug Woo}, author={Templier, Nicolas}, title={Sato-Tate theorem for families and low-lying zeros of automorphic $L$-functions}, journal={Appendix A by Robert Kottwitz, and Appendix B by Raf Cluckers, Julia Gordon, and Immanuel Halupczok, Invent. Math.}, volume={203}, date={2016}, number={1}, pages={1--177}, issn={0020-9910}, review={\MR {3437869}}, doi={10.1007/s00222-015-0583-y}, }
Reference [27]
Zhiwei Yun, The fundamental lemma of Jacquet and Rallis, with an appendix by Julia Gordon, Duke Math. J. 156 (2011), no. 2, 167–227, DOI 10.1215/00127094-2010-210. MR2769216,
Show rawAMSref \bib{YGordon}{article}{ author={Yun, Zhiwei}, title={The fundamental lemma of Jacquet and Rallis, {\rm with an appendix by Julia Gordon}}, journal={Duke Math. J.}, volume={156}, date={2011}, number={2}, pages={167--227}, issn={0012-7094}, review={\MR {2769216}}, doi={10.1215/00127094-2010-210}, }

Article Information

MSC 2010
Primary: 14E18 (Arcs and motivic integration)
Secondary: 22E50 (Representations of Lie and linear algebraic groups over local fields), 40J99 (None of the above, but in this section)
Keywords
  • Transfer principles for motivic integrals
  • uniform bounds
  • motivic integration
  • motivic constructible exponential functions
  • loci of motivic exponential class
  • orbital integrals
  • admissible representations of reductive groups
  • Harish-Chandra characters
Author Information
Raf Cluckers
Université de Lille, Laboratoire Painlevé, CNRS - UMR 8524, Cité Scientifique, 59655 Villeneuve d’Ascq Cedex, France —and— Department of Mathematics, KU Leuven, Celestijnenlaan 200B, B-3001 Leuven, Belgium
Raf.Cluckers@univ-lille.fr
Homepage
Julia Gordon
Department of Mathematics, University of British Columbia, Vancouver, British Columbia V6T 1Z2, Canada
gor@math.ubc.ca
Homepage
Immanuel Halupczok
Mathematisches Institut, HHU Düsseldorf, Universitätsstrasse 1, 40225 Düsseldorf, Germany
math@karimmi.de
Homepage
Additional Notes

The first author was supported by the European Research Council under the European Community’s Seventh Framework Programme (FP7/2007-2013) with ERC Grant Agreement no. 615722 MOTMELSUM, by the Labex CEMPI (ANR-11-LABX-0007-01), and would like to thank both the Forschungsinstitut für Mathematik (FIM) at ETH Zürich and the IHÉS for the hospitality during part of the writing of this paper.

The second author was supported by NSERC.

The third author was partially supported by the SFB 878 of the Deutsche Forschungsgemeinschaft. Part of the work was done while he was affiliated with the University of Leeds.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 5, Issue 6, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2018 by the authors under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/25
  • MathSciNet Review: 3859937
  • Show rawAMSref \bib{3859937}{article}{ author={Cluckers, Raf}, author={Gordon, Julia}, author={Halupczok, Immanuel}, title={Uniform analysis on local fields and applications to orbital integrals}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={5}, number={6}, date={2018}, pages={125-166}, issn={2330-0000}, review={3859937}, doi={10.1090/btran/25}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.