Integrability of pushforward measures by analytic maps
Abstract.
Given a map between -analytic manifolds over a local field of characteristic , we introduce an invariant which quantifies the integrability of pushforwards of smooth compactly supported measures by . We further define a local version near . These invariants have a strong connection to the singularities of .
When is one-dimensional, we give an explicit formula for , and show it is asymptotically equivalent to other known singularity invariants such as the -log-canonical threshold at .
In the general case, we show that is bounded from below by the -log-canonical threshold of the Jacobian ideal near . If , equality is attained. If , the inequality can be strict; however, for , we establish the upper bound , whenever .
Finally, we specialize to polynomial maps between smooth algebraic -varieties and . We geometrically characterize the condition that over a large family of local fields, by showing it is equivalent to being flat with fibers of semi-log-canonical singularities.
Key words and phrases:
-spaces, log-canonical threshold, analytic maps, pushforward measures, constructible functions, Young’s convolution inequality, regularization by convolution1991 Mathematics Subject Classification:
14B05, 03C98, 14E15, 32B20, 60B151. Introduction
The goal of this paper is to explore a singularity invariant of a map between two manifolds over a local field. This invariant quantifies the integrability of pushforward measures by ; we define it in (1.2) below after introducing some notation that will also be used in the sequel.
Throughout this paper, we fix a local field of characteristic , i.e., , or a finite extension of . If is an -analytic manifold of dimension , let be an atlas. We denote by the space of smooth functions on , i.e. functions such that is smooth for each , and by the subspace of compactly supported smooth functions (if is non-Archimedean, smooth means locally constant). We similarly write for the space of smooth measures on , i.e. measures such that each has a smooth density with respect to the Haar measure. We use to denote the space of smooth compactly supported measures on . For , consider the class of finite Radon measures on that are compactly supported and such that for any the measure is absolutely continuous with density in . All these classes do not depend on the choice of the atlas. For we define
Note that by Jensen’s inequality for all .
Now let be an analytic map between -analytic manifolds . If is locally dominant, i.e. contains a non-empty open set for each open subset , then whenever . We now set for each ,
(1.1) |
where the supremum is over all open neighborhoods of . Finally, we can define
(1.2) |
Note that if there exists such that lies in for all and all , then . The best case scenario in this setting is obtained when for every . In this case we say that is an -map.
The main motivation for comes from singularity theory. In general, bad singularities of should manifest themselves in poor analytic behavior of the pushforward of . This phenomenon has been extensively studied in the case , through the analysis of the Fourier transform , which takes the form of an oscillatory integral in the Archimedean case, and of an exponential sum in the -adic case. When the dimension of the target space is equal to , the rate of decay of is closely related to singularity invariants such as the log-canonical threshold (see e.g. [Igu78], [Den91, Corollary 1.4.5] and [CMN19, Theorem 1.5] for the non-Archimedean case, and [AGZV88, Chapter 7] for the Archimedean case). In higher dimension, the connections are less explicit, however, milder singularities of still result in faster decay rates of . For more information on , we refer to Igusa’s work [Igu78], the surveys of Denef, Meuser and León-Cardenal [Den91, LC22, Meu16], as well as the book [AGZV88, Parts II, III] by Arnold, Guseĭn-Zade, and Varchenko, and also the discussion in 3.2 below.
The study of the integrability properties of and their relation to the singularities of has received less treatment. The invariant is a natural step in this direction, and is more robust than Fourier-type invariants as it is also meaningful when is any smooth manifold, which is especially important for applications (see 1.1).
The invariant tends to be small whenever the singularities of are bad near . When is a submersion, the pushforward of any is smooth and in particular lies in . Moreover, Aizenbud and Avni [AA16] have shown that for algebraic maps between smooth algebraic -varieties, the condition that the corresponding map of -analytic varieties is an -map is equivalent to a certain mild singularity property, namely that is flat with fibers of rational singularities (abbreviated (FRS), see Definition 1.8).
When analyzing , one can further restrict the infima in (1.1)–(1.2) to the class of compactly supported measures which are constructible, in the sense of [CGH14, Section 3] and [CM11, Definition 1.2] (see also [LR97]). This class is preserved under pushforward by analytic maps, and therefore is constructible as well. Moreover, constructible measures admit a well behaved structure theory and have tame analytic behavior (see e.g. [CL08, CL10, CM11, CM13, CGH14, CGH18]), and, in particular, it follows from [GH21, CM13] that . Positivity of in the real case can further be deduced from [RS88, Section 2].
Our goal in this paper is to explore in more detail the properties of , and in particular to obtain upper and lower bounds on in terms of other singularity invariants which may be easier to compute, such as the log-canonical threshold of certain ideals. In Theorem 1.1 (proved in 4), we give a lower bound on . In particular, this provides a proof for the positivity of , without using the theory of motivic integration. In Theorem 1.5 (proved in 5) an upper bound on is given in the complex case. In Theorem 1.2 (proved in 3) an explicit formula is given for when is one-dimensional, over any local field. Finally, in Theorem 1.12 (proved in 6), we specialize to polynomial maps between smooth algebraic varieties, and geometrically characterize the condition .
1.1. Application: regularization by convolution
Apart from the geometric motivation discussed above, an additional source of motivation comes from the study of random walks on groups. Assume that is an -analytic group, and take a finite measure on . Can one find a number such that the -th convolution power lies in , and if so, what is the smallest such number ?
An important class of examples comes from the realm of word maps. Given a word in letters, by which we mean an element either of the free group , or of the free Lie algebra , one can consider the corresponding word maps or , where is the Lie algebra of . When is a compact real or -adic Lie group, induces a natural measure where is the normalized Haar measure on . Here, one may further ask what is the -mixing time of the word measure , namely, how large should be to ensure that . Questions of this kind have been studied e.g. in [AA16, LST19, GH24, AG, AGL].
Similar problems have also appeared in a variety of other applications; we mention the work of Ricci and Stein on singular integrals on non-abelian groups (see the ICM survey of Stein [Ste87]). In the study of one-dimensional random operators, the regularity of the distribution of transfer matrices (lying in SL(), or, more generally, in Sp()) plays a role in the work of Shubin–Vakilian–Wolff [SVW98] as well as in the recent work [GS22].
Motivated by the above examples, we focus on the following setting: and are analytic, and is the pushforward of a measure under a locally dominant analytic map . For , let
be the smallest that works for any supported in a sufficiently small neighborhood of .
According to Young’s convolution inequality for locally compact groups (see [KR78, Corollary 2.3] and Remark 1.10),
This implies
(1.3) |
As mentioned above, a more classical approach to bounding relies on the study of the decay of the Fourier transform of . While the Fourier-analytic approach often provides sharper bounds (see [GH24, Proposition 5.7]), it is mainly applicable for abelian groups such as , or mildly non-abelian groups such as the Heisenberg model. One can use non-commutative Fourier transform to analyze compact Lie groups such as , but such representation theoretic techniques are much less effective for compact -adic groups and non-compact Lie groups. To treat the latter cases, one can use algebro-geometric techniques as in [AA16, GH24]; however, this method requires some assumptions on . Thus one can argue that the approach to regularization via (1.3) is currently the most efficient one for treating -adic analytic groups and non-compact Lie groups such as .
1.2. Main results
We now discuss the main results in this paper.
1.2.1. A lower bound on
While the mere positivity of (and the mere finiteness of ) are sufficient for some applications, other ones require explicit bounds. Our first result provides a bound in terms of an important exponent known as the log-canonical threshold, see e.g. [Mus12, Kol] for . For an analytic map , define the -log-canonical threshold
(1.4) |
where runs over all open neighborhoods of , and is the absolute value on , normalized such that , for all , , and where is a Haar measure on . In particular, is the square of the usual absolute value on . More generally, if is a non-zero ideal of analytic functions generated by , define
(1.5) |
This definition does not depend on the choice of the generators, and thus it extends in a straightforward way to sheaves of ideals. Furthermore, the log-canonical threshold is always strictly positive (see 2).
Given a locally dominant analytic map between two -analytic manifolds, we define the Jacobian ideal sheaf as follows. If and are open subsets, we define to be the ideal in the algebra of analytic functions on , generated by the -minors of the differential of . Note that if and are analytic diffeomorphisms, then . Hence, the definition of can be generalized (or glued) to an ideal sheaf on , if and are -analytic manifolds.
We now describe the first main result, which provides a lower bound on in terms of the Jacobian ideal of . The proof is given in 4.
Theorem 1.1.
Let be analytic -manifolds, , and let be a locally dominant analytic map. Then for every ,
(1.6) |
if , equality is achieved.
1.2.2. A formula in the one-dimensional case and a reverse Young inequality
When the target space is one-dimensional, Hironaka’s theorem on the embedded resolution of singularities [Hir64] provides a powerful tool to study the structural properties of algebraic and analytic maps. This theorem, as well the asymptotic expansion of pushforward measures about a critical value of the map, allows us to obtain the following much more detailed results, the proofs of which are given in 3.
The first one is an exact formula relating to the log-canonical threshold.
Theorem 1.2.
Let be an analytic -manifold, and let be a locally dominant analytic map. Then for each , we have:
(1.7) |
By Theorem 1.2, by (1.3) and by a Thom–Sebastiani type result for (Proposition 3.11(1)), one can further show:
(1.8) |
We therefore see that that and are asymptotically equivalent as . In 3.2, we shall see that these invariants are also tightly related to an invariant quantifying Fourier decay. We will further see in 3.4 that the close relation between all these quantities is a special feature of the one-dimensional case, and does not generalize to higher dimensions.
We next provide a reverse Young result for pushforward measures by analytic maps. Recall that Young’s convolution inequality (see e.g. [Wei40, pp. 54-55]) implies that
(1.9) |
Using the connection between and as well as the structure of pushforward measures, we show the following converse to (1.9):
Theorem 1.3 (Reverse Young inequality).
Let be pushforward measures of the form , where and are analytic, locally dominant. If for some , then
(1.10) |
In particular, if is equal to , it lies in .
Remark 1.4.
Under the assumptions of Theorem 1.3, if then (1.10) (applied with ) implies
(1.11) |
In general, one cannot hope for a strict inequality in (1.11); indeed taking or for , and a uniform measure on some ball around , one has , so that (1.11) . On the other hand .
When , or more generally when the codimension of in is , we expect (1.11) to hold with a strict inequality.
1.2.3. Upper bounds on
For , the lower bound (1.6) may in general not be an equality. However, the next result shows that when , (1.6) is asymptotically sharp as .
Theorem 1.5.
Let be analytic -manifolds, and let be a locally dominant analytic map. Then, whenever ,
(1.12) |
Remark 1.6.
- (1)
-
(2)
The upper bound (1.12) is asymptotically tight, in the sense that the value of can be arbitrarily close to the upper bound (1.12), as seen from the following family of examples. Let . Then , and thus by [How01, Main Theorem and Example 5], it follows that
so that the upper bound in (1.12) becomes , whereas the lower bound (1.6) is . We see that the true value is closer to the upper bound than to the lower bound.
One may wonder whether Theorem 1.5 can be extended to . In the current proof, the volumes of balls in complex manifolds are bounded from below using Lelong’s monotonicity theorem, and the latter fails for and for any . If is a polynomial map between smooth varieties, defined over , we expect the upper bound in Theorem 1.5 to hold for for infinitely many primes . This would follow from a positive answer to Question 1.14.
1.2.4. Applications to convolutions of algebraic morphisms
Throughout this and the next subsections we assume is a number field. In [GH19, GH21] and [GH24], the first two authors have studied the following convolution operation in algebraic geometry:
Definition 1.7.
Let and be morphisms from algebraic -varieties to an algebraic -group . We define their convolution by
We denote by the -th self convolution of .
We restrict ourselves to the setting where are smooth algebraic -varieties and is a connected algebraic -group. The main motto is that the algebraic convolution operation has a smoothing effect on morphisms, similarly to the usual convolution operation in analysis (see [GH19, Proposition 1.3] and [GH21, Proposition 3.1]). For example, starting from any dominant map , the -th self convolution is flat for every ([GH21, Theorem B]). To explain the connection to this work, we introduce the following property:
Definition 1.8 ([AA16, Definition II]).
-
(1)
A -variety has rational singularities if it is normal and for every resolution of singularities , the pushforward of the structure sheaf has no higher cohomologies.
-
(2)
A morphism between smooth -varieties is called (FRS) if it is flat and if every fiber of has rational singularities.
In [AA16, Theorem 3.4] (see Theorem 6.1 below), Aizenbud and Avni proved the following. A morphism between smooth -varieties, is (FRS) if and only if for every non-Archimedean local field , one has for every . A similar characterization can be given for , see Corollary 6.2.
This characterization of -maps allows one to study random walks on analytic groups as in 1.1 in an algebro-geometric way, via the above algebraic convolution operation. Starting from a pushforward of by an algebraic map , instead of showing that , it is enough to show that is an (FRS) morphism. This method was used in [AA16, GH24] to study word maps. Moreover, in [GH19, GH21] it was shown that any locally dominant morphism becomes (FRS) after sufficiently many self-convolutions.
Corollary 1.9.
Let be a smooth -variety, be a connected -algebraic group and let be a locally dominant morphism. Then is (FRS) for any .
Remark 1.10.
In the setting of locally compact groups, Young’s convolution inequality is commonly stated under the assumption that the group is unimodular. In [KR78, Lemma 2.1 and Corollary 2.3] a version for non-unimodular groups is given; if is a locally compact group, with modular character , and if satisfy , then we have whenever and . However, since the modular character is a continuous homomorphism, it bounded on the compact support of . Hence, for every and , we deduce that
for some constant depending on and . In particular, .
1.2.5. An algebraic characterization of
Let be a morphism between smooth -varieties. We would like to characterize the condition that for all in certain families of local fields, in terms of the singularities of . The singularity properties we consider play a central role in birational geometry (see [Kol97]).
Let be a normal -variety, and let be a rational top form on the smooth locus of . The zeros and poles of give rise to a divisor on . Let be a resolution of singularities, namely, a proper morphism from a smooth variety , which is an isomorphism over . Then defines a unique rational top form on . Moreover, when is nice enough (e.g. if is a local complete intersection), is -Cartier, and we can define its pullback . The divisor on is called the relative canonical divisor, and one can verify that it does not depend on the choice of . can be written as , for some prime divisors , . We say that has canonical singularities (resp. log-canonical singularities), if (resp. ) for all . When is a local complete intersection (e.g. a fiber of a flat morphism between smooth schemes), canonical singularities are equivalent to rational singularities. Let us give an example:
Example 1.11.
Let be the variety defined by , with . Then has canonical singularities if and only if , and log-canonical singularities if and only if .
As seen from Example 1.11, log-canonical singularities are very close to being canonical, so one could suspect being flat with fibers of log-canonical singularities is equivalent to , that is to being almost in . Unfortunately, the normality hypothesis required for log-canonical singularities turns out to be too strong. For example, the map satisfies for all , but the fiber over is not normal (see 6, after Corollary 6.2). This technical issue can be resolved by considering the slightly weaker notion of semi-log-canonical singularities, which is an analogue of log-canonical singularities for demi-normal schemes (see [KSB88, Section 4]). Indeed, the variety is demi-normal and has semi-log-canonical singularities. We can now state the main result of this section:
Theorem 1.12.
Let be a map between smooth -varieties. Then the following are equivalent:
-
(1)
is flat with fibers of semi-log-canonical singularities.
-
(2)
For every local field containing , we have , that is, for every , the measure lies in for all .
-
(3)
For every large enough prime , such that , we have .
-
(4)
We have .
We prove Theorem 1.12 by showing the implications and . The implications and are immediate. In the proof of and , we reduce to the case , and show that satisfies that is (FRS) for every dominant map . By analyzing the jets of , and using a jet-scheme interpretation of semi-log canonical singularities (Lemma 6.5), we deduce Item (1). The proof also uses the Archimedean counterpart of [AA16, Theorem 3.4], which is stated in Corollary 6.2.
1.3. Future directions and further applications
1.3.1. as an invariant of singularities
Similarly to the analytic definition of in (1.2), one can also define an algebro-geometric invariant. Let be the collection of all non-Archimedean local fields of characteristic .
Definition 1.13.
Let be a morphism between smooth -varieties. We define , and , where varies over all Zariski open neighborhoods of .
It is a consequence of Theorem 1.1 that for any , and it essentially follows from [GH21] that . We further expect to have a purely algebro-geometric formula, and to have a good behavior in families, which means the following. Suppose that is a morphism over , where and are smooth morphisms. This gives a family of morphisms between smooth varieties. It follows from [Var83], that the function is lower semicontinuous. By Theorem 1.2, is lower semicontinuous as well, if has fibers of dimension . We expect to be lower semicontinuous in general. We further expect the following question to have a positive answer.
Question 1.14.
Let be as in Definition 1.13. Is it true that ?
Given a -morphism as above, one may further wonder whether the quantity in (1.1) stays the same when the supremum is taken over all Zariski open neighborhoods of instead of analytic ones.
1.3.2. as an invariant of words
As discussed in 1.1, one particularly interesting potential application is to the study of word map on semisimple algebraic groups. In [GH24, Theorem A] it was shown that Lie algebra word maps , where is a simple Lie algebra, become (FRS) after self convolutions, where is the degree of .
Question 1.15.
Can we find such that for any of length , and every simple algebraic group , the word map is (FRS)?
A potential way to tackle Question 1.15 is by studying in the sense of Definition 1.13; For each (resp. ), we define (resp. ), where runs over all simple, simply connected algebraic groups, and . We can now ask the following:
Question 1.16.
Can we find for each , a constant such that:
-
(1)
For every of length , we have ?
-
(2)
For every of degree , we have ?
1.3.3. as an invariant of representations
In a recent work [GGH] of the first two authors with Julia Gordon, we apply the results and point of view of this paper to the realm of representation theory, and define and study a new invariant of representations of reductive groups over local fields. Harish-Chandra’s regularity theorem says that every character of an irreducible representation of is given by a locally -function. Since characters are of motivic nature, a variant of [GH21, Theorem F] suggests that they should in fact be locally in , for some . This gives rise to a new invariant , which is not equivalent to previously known invariants of representations, such as the Gelfand–Kirillov dimension (see e.g. [Vog78]). We use a geometric construction and Theorem 1.1 to provide a formula for in terms of the nilpotent orbits appearing the local character expansion of (see [GGH, Theorems A and D]).
1.4. Conventions
-
(1)
By we mean the set .
-
(2)
We use (resp. ) to denote number (resp. local) fields and (resp. ) for their rings of integers.
-
(3)
Given an algebraic map between smooth -varieties, and a local field , we denote by the corresponding -analytic map. We sometimes write instead of if the local field is clear from the setting.
-
(4)
If are two functions, we write if there exists a number , possibly depending on , and , such that . We write if .
-
(5)
We write for the differential of an analytic map at , and we denote by the Jacobian of if . The generalization of the Jacobian to the case of unequal dimensions is defined in 1.2.
-
(6)
Throughout the paper, we write for the absolute value on , normalized so that , for all , , and where is a Haar measure on . Note that is the square of the usual absolute value on .
-
(7)
We write for the -dimensional Hausdorff measure. We recall that given a metric space and a subset , we define , where
where denotes the diameter of a set and . The normalization constant is chosen such that coincides with the Lebesgue measure in the case of .
-
(8)
All the measures we consider are non-negative, unless stated otherwise.
Acknowledgement.
We thank Nir Avni, Joseph Bernstein, Lev Buhovski, Raf Cluckers and Stephan Snegirov for useful conversations and correspondences. We thank David Kazhdan for suggesting that the condition of semi-log-canonical singularities should play a role in Theorem 1.12. We are particularly grateful to Rami Aizenbud for numerous discussions on this project, and for initiating the collaboration between the authors.
SS was supported in part by the European Research Council starting grant 639305 (SPECTRUM), a Royal Society Wolfson Research Merit Award (WM170012), and a Philip Leverhulme Prize of the Leverhulme Trust (PLP-2020-064).
2. Preliminaries: embedded resolution of singularities
Let be a local field of characteristic zero. We use the following analytic version of Hironaka’s theorem [Hir64] on embedded resolution of singularities. The map below is called a log-principalization (or uniformization) of .
Theorem 2.1 (See [VZnG08, Theorem 2.3], [DvdD88, Theorem 2.2], [BM89] and [Wlo09]).
Let be an open subset, and let be -analytic maps, generating a non-zero ideal in the algebra of -analytic functions on . Then there exist an -analytic manifold , a proper -analytic map and a collection of closed submanifolds of of codimension , equipped with pairs of non-negative integers , such that the following hold:
-
(1)
is locally a composition of a finite number of blow-ups at closed submanifolds, and is an isomorphism over the complement of the common zero set of in .
-
(2)
For every , there are local coordinates in a neighborhood , such that each containing is given by the equation . Moreover, if without loss of generality contain , then there exists an -analytic unit , such that the pullback of is the principal ideal
(2.1) and such that the Jacobian of (i.e. ) is given by:
(2.2)
Remark 2.2.
-
(1)
Condition (2.1) means that for each , one can write for some analytic functions , and for at least one . Note that the are the same for all the ’s.
-
(2)
If , so that , we may further assume that locally on each chart, for some constant . Indeed, . If is an -th power in then the same holds for in a small neighborhood of . In this case, we may apply the change of coordinates
If is not an -th power we may multiply it by such that is an -th power, and apply a similar change of coordinates.
The next lemma follows directly by changing coordinates using a log-principalization of and computing the integral with respect to the new coordinates.
3. The one-dimensional case
3.1. A formula for
In this section we provide a formula for in the one-dimensional case (Theorem 1.2). The formula will be phrased in terms of the -log-canonical threshold, where is any local field of characteristic zero.
When is non-Archimedean, we denote by its ring of integers, by its residue field, and by the number of elements in . Write for a fixed uniformizer (i.e., a generator of the maximal ideal of the ring of integers) of , and let be as in 1.4, so that . We write for the Haar measure on , normalized such that when is non-Archimedean, and such that is the Lebesgue measure when is Archimedean. We write . We write instead of when we integrate a function with respect to . We denote by the maximum norm on .
For an analytic map , and , we set .
To prove Theorem 1.2, we reduce to the monomial case using Hironaka’s resolution of singularities (2), and prove the monomial case in Lemma 3.2. However, we first note that the upper bound in (1.7) can be proved by elementary arguments, as follows:
Lemma 3.1.
Let be an analytic -manifold, and let be a locally dominant analytic map. Then for every with , one has:
Proof.
We need to show
Let . Then there exists a neighborhood of such that for every . Write . Let , and note that . By Jensen’s inequality we have:
i.e. we have the distributional estimate . Using Fubini’s theorem, we obtain:
whenever . This implies that . ∎
We now return to the main narrative. Specializing to the setting of Hironaka’s theorem, we consider the monomial case.
Lemma 3.2.
Let be a monomial map , let be a continuous function and let , for and . Then:
Furthermore, if , the second bound is in fact an equality.
Proof.
Without loss of generality we may assume . We first consider the special case when is the unique minimum. We write as a composition , where is given by and is the projection to the last coordinate. Write , where , and similarly . Note that and that if , then and . Hence the Radon–Nikodym density of is equal to
Since we get for every ,
and hence:
In particular, integrating over the first coordinates, we get that . If then as required. Thus we may assume that . Then , whenever , i.e. whenever
If we also have , whence the inequality in the statement of the lemma is in fact an equality.
It is left to deal with the case when is not uniquely achieved. For the lower bound, take
for an arbitrarily small . Since inside a small neighborhood of , we deduce that for . Since can be taken arbitrarily small we are done. Similarly, for the upper bound we take
with and deduce that . ∎
Proof of Theorem 1.2.
Let be a locally dominant analytic map, and let . Replacing with , we may assume that . We may further assume that is open, and apply Theorem 2.1, to get a log-principalization , such that, locally on a chart around a point in , , and
Let , with , and in a neighborhood of . Then , where
3.2. Relation to Fourier decay and other invariants
Let be an -analytic map between -analytic manifolds and . We have seen that each of the exponents and provides a different quantification for the singularities of near . When , one can further consider other invariants involving the Fourier transform of pushforward measures.
In 1.1, we have defined as the minimal number of self-convolutions after which the pushforward densities of smooth measures supported near become bounded. Note that by the Plancherel theorem, for each we have if and only if , whence
Thus, the exponent is, in general, roughly comparable to the -class of rather than to the -class of the pushforward measure itself. Instead of the -class, we now focus on an invariant quantifying the Fourier decay of on the power-law scale:
where
(3.1) |
The study of the invariant , and variations of it, goes back at least to the 1920’s, when the classical van der Corput lemma was introduced, relating lower bounds on the derivative of a smooth function , to bounds as in (3.1), see [Ste93, Proposition 2], and [CCW99]. This invariant was also studied extensively in Igusa’s work [Igu78] in the case ; it is much less understood in high dimensions.
Remark 3.3.
Note that and are preserved under analytic changes of coordinates around . On the other hand, and might depend on the choice of coordinate system. For example, the map satisfies (by Theorem 1.1), while and . By applying the change of coordinates , we get , and still have , while and .
We next discuss the relations between the different exponents. In the one-dimensional case, it turns out that all of the exponents above are essentially equivalent (whenever ). In order to explain this, we need to discuss the structure of pushforward measures by analytic maps.
3.2.1. Asymptotic expansions of pushforward measures and their Fourier transform
Let be a local field, be a locally dominant analytic map with an open set. Let , and consider the pushforward measure . Fix a non-trivial additive character of . We may identify between and by where . The Fourier transform can then be written as
To and , one can further associates Igusa’s local zeta function
Igusa’s local zeta function admits a meromorphic continuation to the complex plane (see [BG69, Ati70, Ber72] for the Archimedean case, and [Igu74, Igu78] for the non-Archimedean case). The poles of (as well as certain twisted versions of it), and the Laurent expansions around them, controls the asymptotic expansions for as and for when , via the theory of Mellin transform (see [Igu78, Theorems 4.2, 4.3 and 5.3]). We next describe the asymptotic expansions of and .
Definition 3.4 (Asymptotic expansion, see [Igu78, Section I.2]).
Let be or . A sequence of complex-valued functions on an open subset of , with in a punctured neighborhood of , is said to constitute an asymptotic scale, if for every ,
A function is said to have an asymptotic expansion at , if there exists a sequence of complex numbers such that for every , there exists such that for all close enough to :
In this case, we write
(3.2) |
Example 3.5.
Given a monotone increasing sequence of real numbers, with no finite accumulation points, and given a sequence of positive integers, set to be the sequence:
for . Then is an asymptotic scale at .
We now describe the asymptotic expansions of pushforward measures and their Fourier transforms. We fix a local field of characteristic and an analytic -manifold . If is non-Archimedean we further fix a uniformizer . We set and denote by the angular component map
Theorem 3.6 ([Jea70, Mal74, Igu78], see also [VZnG17, Section 4] and [Den91, Theorem 1.3.2 and Corollary 1.4.5]).
Let be a locally dominant analytic map, let , and write . Suppose that is the only critical value of . Then there exist:
-
•
a sequence , consisting of strictly increasing positive real numbers with , if , or a finite set of complex numbers , with , and for some , if is non-Archimedean;
-
•
a sequence of positive integers;
-
•
smooth functions on ,
such that:
-
(1)
For , admits an asymptotic expansion111The asymptotic expansion is also termwise differentiable, and uniform in the angular component; we refer to [Igu78, p. 19-24] for the precise meaning of those notions. of the form
(3.3) and admits an asymptotic expansion of the form,
(3.4) - (2)
-
(3)
For each , the functions are determined by the functions . If , the map taking the latter to the former is one-to-one, and, moreover, the leading function is not identically zero (provided that is defined so that is not identically ).
Remark 3.7.
- (1)
-
(2)
Note that Igusa’s theory (and in particular Theorem 3.6) was originally developed for polynomial maps but works for analytic maps as well. Indeed, the proof uses resolution of singularities to reduce to the case of pushforward of measures with monomial density by monomial maps. The same reduction can be made for analytic maps via an analytic version of resolution of singularities (as stated in Theorem 2.1). For a generalization of Theorem 3.6 to the case of meromorphic maps, see [VZnG17, Section 5].
Corollary 3.8.
Let be a locally dominant analytic map, let , and suppose that . Then , i.e. the supremum in the definition of is not achieved.
Theorem 3.6 also implies the following corollary, relating to :
Corollary 3.9.
Let be a locally dominant analytic map such that is the only critical value. Then for each
(3.5) |
Proof.
Write . If , then by Theorem 3.6, can be expanded near , so that the leading term is for some and . Similarly, can be expanded near . If , then by Item (3) of Theorem 3.6, is the leading term of , and thus . If , then Item (3) implies that .
Since is the only critical value of , is bounded outside any neighborhood of , so is integrable if and only if it is integrable in a small ball around , and this holds if and only if , i.e. either and then , or .
The case when is non-Archimedean should be done with care, since there might be multiple terms in (3.3) with the same real part, and some cancellations may occur (see Example 3.10 below). Let and suppose . Then by Theorem 3.6, can be written as , where
is equal to if and otherwise, and furthermore,
We first show there is an arithmetic progression for some , and , such that
(3.6) |
for all , for some constant independent of . It is enough to show (3.6) for the terms in where is maximal such that for some , is not identically zero. Note that
Let be such that for some . Choose such that each of is constant on the ball . Note that the functions , are the irreducible characters of and hence they are linearly independent. In particular, there exists , such that
for all . Taking for , we deduce (3.6) as required.
In the proof of Corollary 3.9 we have seen there might be some cancellations between the terms in (3.3) with the same real part, and that these cancellations are insignificant for infinitely many values of . Here is a simple example which illustrates this phenomenon.
Example 3.10.
Let and let be a prime. Let be the map . Write . Then for almost all we have
where . Note that by Schur orthogonality, if and is if . In particular
and therefore, the expansion of as in (3.3) is
3.2.2. Relations between the invariants
We now show that and are essentially determined by the log-canonical threshold whenever . In particular, we show that
are asymptotically equivalent as .
Theorem 1.2 already shows that if . We further have the following:
Proposition 3.11.
Let be a dominant -analytic map. Then:
-
(1)
If , then . In particular,
-
(2)
We have:
Proof.
Let us first prove Item (1). Let . Replacing with , we may assume that . We may choose an open neighborhood such that is the only critical value of , is compact and such that for each , and each , one has . Taking any which does not vanish at , we get by Theorem 1.2 and Lemma 3.2 that
Corollary 3.9 implies that
Since the equalities above hold for of arbitrarily small support around , we get as required. Since Fourier transform translates convolution into product, we have
which implies the second part of Item (1) (see also [Den91, Section 5.1]).
Remark 3.12.
Proof of Theorem 1.3.
Let , and suppose that
(3.7) |
where . Assume by contradiction that
For each , choose a finite cover of by open balls in , such that for each , has at most one critical value of . We can write with supported inside . Taking , we can find and such that
Now, Corollary 3.9 implies that
whence
Using Corollary 3.9 once again, we obtain that
3.3. Consistency of the various bounds in the one-dimensional case
The following proposition, which can be seen as a variant of the Łojasiewicz gradient inequality (see e.g. [Lo65, p.92], [BM88, Proposition 6.8]), shows that the formula for given in Theorem 1.2 in the one-dimensional case is consistent with the lower and upper bounds in Theorems 1.1 and 1.5. The proof of Proposition 3.13 is similar to [Fee19, Theorem 1], but applied to any local field.
Note that for , we have , which we denote by .
Proposition 3.13.
Let be an analytic map. Then for every , we have
where the middle (resp. left) inequality holds whenever (resp. ).
Proof.
The middle inequality follows from Theorems 1.1 and 1.2, and the left inequality follows from the right inequality, so it is left to show that for a fixed .
Recall that is the maximum norm on . We would like to relate between and , where is a small ball around , for small enough. Let be a resolution of singularities of . Then we have
(3.8) |
Since is compact, by working locally over finitely many pieces, and using Theorem 2.1, we may replace by a compact neighborhood of , , and further assume that for :
(3.9) |
for some analytic unit and a constant . Note that on all of the entries of are smaller, in absolute value, than some constant , so that the operator norm of is bounded by a constant. Since , for we get:
which concludes the proof. ∎
Remark 3.14.
Here is an alternative approach to Proposition 3.13, as suggested by the referee. For simplicity suppose . We may assume that . Recall from [HS06, Definition 1.1.1] that the integral closure of an ideal in a commutative ring , is the set of elements for which there are such that for some . We now take to be the ring of convergent power series, and . By [HS06, Corollary 7.1.4], we have and in particular (see [Mus12, Property 1.15]). However, by [Mus12, Property 1.12] (see also [dFM09, Proposition 2.4]) we have , which implies that as required. This argument likely generalizes to any local field of characteristic , i.e. the full generality of Proposition 3.13.
3.4. Some examples in higher dimension
We next discuss the higher dimensional case (i.e. ). Here we will see that the connection between the four invariants , , and is not as tight as in the one-dimensional case. We first provide a simpler description of .
Lemma 3.15.
Let be a local field of characteristic and be an analytic map. Then
where runs over all non-zero linear functionals .
Proof.
For each , we have
for each with and . Setting , we have,
which concludes the proof of the lemma. ∎
The following examples demonstrate that the connection between the four invariants can get loose as the dimension of grows.
Example 3.16.
Consider the map , defined by . Then:
-
(1)
.
-
(2)
, where .
-
(3)
.
-
(4)
.
Item (1) follows from Theorem 1.1. Young’s inequality (see 1.3) shows that , on the other hand we have which implies that at least convolutions are needed to obtain the (FRS) property (see [GH24, Lemmas 3.23 and 3.26]), hence . Note that for any with we have so , by Lemma 3.15.
Remark 3.17.
The following example shows that a reverse Young inequality (Theorem 1.3) does not hold in dimension .
Example 3.18.
Consider . Then , while , and consequently .
4. Lower bound: proof of Theorem 1.1
In this section we prove Theorem 1.1. We start with the equidimensional case, which we restate as Proposition 4.1 below.
Recall that given a locally dominant analytic map between two -analytic manifolds, the Jacobian ideal sheaf is defined locally as the ideal in the algebra of analytic functions on generated by the -minors of the differential of .
Proposition 4.1.
Let be -analytic manifolds, , and let be a locally dominant analytic map. Then for any ,
(4.1) |
Proof.
We may assume that are compact balls in , and . Since , we have , where . Since is analytic and is compact, there is an open dense set and such that and , for every . We choose a disjoint cover of by locally closed subsets222Recall that a set is locally closed if it is the difference of two open sets. such that is a diffeomorphism. Write , , and note that and , where
We have
(4.2) |
On the other hand, since for each , using Jensen’s inequality, we have:
(4.3) |
which implies the proposition. ∎
Proof of Theorem 1.1.
Let be a locally dominant analytic map, and let . Since the claim is local, we may assume that is an open subset, and , with . Let be a small ball around . For each subset of size , let be the corresponding -minor of , and set
If , the map is locally dominant, where , and . For each of positive measure, let , and write
Let be a large ball in which contains the projection of from to the last coordinates . Since where is a projection to the first coordinates, we have
By Jensen’s inequality, we have
for every . Since is of full measure in , and using Proposition 4.1 and (1.5), we have:
for every , as required. ∎
5. Proof of Theorem 1.5 – an upper bound over
In this section we prove Theorem 1.5. We use the following easy consequence of the coarea formula (see [Fed69]).
Lemma 5.1.
Let be a compact ball, and let be a dominant analytic map. Let and write . Then:
where the integral is taken with respect to the -dimensional Hausdorff measure (recall 1.4(7)).
Let be a locally dominant analytic map between complex analytic manifolds, and let . Since the claim is local, we may assume that and is a ball in . To show (1.12), it is enough to bound , where is of arbitrarily small radius around . Denote . By Lemma 5.1, we have:
whence
(5.1) |
We apply Theorem 2.1 to the ideal . Let be the corresponding resolution. Without loss of generality, we can assume that is an open subset, so that
and each of the minors of satisfies
where at least one of the functions does not vanish at .
We first perform the change of variables in the external integral in (5.1), yielding
and then perform the change of variables in the internal integral. Since the map is continuously differentiable, and is compact, the product of the singular values of the restriction of to any subspace is bounded from below in absolute value by a number times the absolute value of the determinant of . Thus,
(5.2) |
where . By the Cauchy–Binet formula, the right-hand side is equal to
We bound the internal integral from below as follows. For each , set
Then
(5.3) |
Claim 5.2.
For any ,
Proof of Claim 5.2.
Let be the affine map given by . Then , where is a ball of radius centered at the origin. By a theorem of Lelong [Lel57] (see also [Thi67, LG86]), for any analytic set of pure dimension , and any , one has
where is a linear subspace of the same dimension . The limit on the right-hand side is the Lelong number of , which is the algebraic multiplicity of at ; it is strictly positive; thus
Applying this estimate to and observing that
for any Borel set , we obtain the claimed assertion. ∎
We now proceed with the proof of the theorem. Using Claim 5.2, we deduce
whence,
This integral diverges whenever there is an index such that
i.e.
On the other hand,
and under the assumption that this quantity is strictly less than we have an index such that
For this index,
This concludes the proof of Theorem 1.5.
6. Geometric characterization of
From now on let be a number field and let be its ring of integers. Let be the collection of all non-Archimedean local fields which contain . We use the notation , for the collection of with large enough residual characteristic, depending on some given data. Throughout this section, we denote by a ball of radius centered at .
In this section we focus on algebraic morphisms between algebraic -varieties. We would like to characterize morphisms where for certain collections of local fields , in terms of the singularities of . In order to effectively do this, it is necessary to consider an “algebraically closed” collection of local fields, such as the following:
-
•
.
-
•
, or .
Aizenbud and Avni have shown the following characterization of the (FRS) property:
Theorem 6.1 ([AA16, Theorem 3.4]).
Let be a map between smooth -varieties. Then is (FRS) if and only if for each and every , one has .
The Archimedean counterpart of this theorem was studied in [Rei], where it was shown that given an (FRS) morphism , then and are -morphisms. For the other direction, the non-Archimedean proof of [AA16, Theorem 3.4] (see [AA16, Section 3.7]), can be easily adapted to the complex case, with less complications due to the fact that is algebraically closed. We arrive at the following characterization of -morphisms.
Corollary 6.2.
Let be a map between smooth -varieties. Then the following are equivalent:
-
(1)
is (FRS).
-
(2)
For every local field containing , the map is an -morphism.
-
(3)
For each , the map is an -map.
-
(4)
The map is an -map.
Our goal is to characterize the weaker property that over , or over all (Theorem 1.12, restated below as Theorem 6.6). Let us first present an example showing the (FRS) condition is too strong for this purpose. Let be the map . Then:
In particular, we have
Hence, the measure does not have bounded density. On the other hand, since , and by considering the asymptotic expansion of as in Theorem 3.6, one sees:
-
(1)
explodes logarithmically around , i.e. the density of behaves like , around .
-
(2)
for every prime .
By Corollary 6.2, cannot be (FRS), and indeed is not normal, so in particular it does not have rational singularities.
In order to prove Theorem 1.12, we recall the notion of jet schemes. Let be an affine -scheme whose coordinate ring is
Then the -th jet scheme of is the affine scheme with the following coordinate ring:
where is the -th formal derivative of .
Let be a morphism between affine spaces. Then the -th jet morphism of is given by formally deriving , . Similarly, the -th jet of a morphism of affine -schemes, is given by the formal derivative of . Both and can be generalized to arbitrary -schemes and -morphisms (see [CLNS18, Chapter 3] and [EM09] for more details).
Given a subscheme of a smooth variety , with defined by an ideal , we denote by the log-canonical threshold of the pair . Mustaţă showed that the log-canonical threshold can be characterized in terms of the growth rate of the dimensions of the jet schemes of :
Theorem 6.3 ([Mus02, Corollary 0.2], [CLNS18, Corollary 7.2.4.2]).
Let be a smooth, geometrically irreducible -variety, and let be a closed subscheme. Then
Furthermore, the supremum is achieved for divisible enough. In particular, is a rational number.
Note that depends on and , and neither on the ambient space not on the embedding of in .
We now introduce the following definitions from [GH24]. For a morphism between schemes, we denote by the scheme theoretic fiber of over .
Definition 6.4.
Let be a morphism of smooth, geometrically irreducible -varieties, and let .
-
(1)
is called -flat if for every we have .
-
(2)
is called -jet flat if is -flat for every .
-
(3)
is called jet-flat if it is -jet flat.
In particular, note that is flat if and only if it is -flat.
Note that by Theorem 6.3, is -jet-flat if and only if for all . We will need the following lemma to give a jet scheme interpretation to rational and semi-log-canonical singularities (from Theorem 6.3).
Lemma 6.5.
Let be a morphism of smooth -varieties. Then:
-
(1)
is (FRS) if and only if is flat with locally integral fibers, for every (in particular, is jet-flat).
-
(2)
is jet-flat if and only if is flat with fibers of semi-log-canonical singularities.
Proof.
Item (1) is proved in [GH24, Corollary 3.12] and essentially follows from a characterization of rational singularities, by Mustaţă [Mus01]. For the proof of Item (2), note that by [GH24, Corollary 2.7], is jet-flat if and only if is flat over . Since a fiber of a morphism between smooth varieties is flat if and only if its fibers are local complete intersections, the latter condition is equivalent to the condition that every , and every , the scheme is a local complete intersection. By [Ish18, Corollary 10.2.9] and [EI15, Corollary 3.17], this is equivalent to the condition that has semi-log-canonical singularities, for every . ∎
6.1. Proof of Theorem 1.12
We are now in a position to prove Theorem 1.12. Let us recall its formulation, slightly restated using Lemma 6.5.
Theorem 6.6.
Let be a map between smooth -varieties. Then the following are equivalent:
-
(1)
is jet-flat.
-
(2)
For every local field containing , we have .
-
(3)
For every we have .
-
(4)
We have .
The proof of Theorem 6.6 is done by showing both implications and . The implications and are immediate. We first prove and . Then we will prove in the non-Archimedean case in 6.2, and the Archimedean case in 6.3.
Proposition 6.7.
Let be a map between smooth -varieties. Assume that either or for all . Then is jet-flat.
Proof.
Working locally, and composing with an étale map , we may assume that . Let be any dominant morphism, and let and . By Theorem 1.1, for all , we have for some . Taking large enough, by Young’s convolution inequality, one has:
where is as in Definition 1.7. By Corollary 6.2, we get that is (FRS). We now claim that since is a morphism whose convolution with any dominant morphism produces an (FRS) morphism, must be jet-flat.
Indeed, assume it is not the case. Then by Theorem 6.3 and Definition 6.4, there exist and such that the scheme theoretic fiber of over satisfies
Moreover, this supremum is achieved for divisible enough. Thus the map
is not -flat for divisible enough. But on the other hand, the map
satisfies that is not -flat for divisible enough (since ). Thus we may find such that is not -flat and is not -flat . But then is not flat (see [GH24, Lemma 3.26]), which is a contradiction by Fact 6.5, as is (FRS). ∎
6.2. : the non-Archimedean case
We now turn to the proof of , in the non-Archimedean case. We first prove the following variant of [CGH23, Theorem 4.12].
Proposition 6.8 ([CGH23, Theorem 4.12]).
Let be a jet-flat map between smooth -varieties. Then there exists , such that for each , each and each non-vanishing , one can find such that for each and one has,
Proof.
We may assume that and that . We may further assume that is affine, and thus embeds in . Let be the canonical measure on (see [Ser81, Section 3.3], and also [CCL12, Section 1.2]). It is enough to consider measures which are of the form .
Write for the density of with respect to , and set . Then the collections and are both motivic functions, in the sense of [CGH18, Section 1.2]. By [CGH18, Theorem 2.1.3], there exists a motivic function , which approximates the supremum of , that is:
(6.1) |
for all and , where depends only on the local field333In the statement of [CGH18, Theorem 2.1.3], the approximation (6.1) is stated for instead of . Since is a non-negative real-valued motivic function, their argument yields the current statement as well (see the first four lines of the proof on p.146)..
Since is motivic, and using [CGH18, Proposition 1.4.2], for each , we may divide into a finite disjoint union , with independent of , such that on each part , the following hold. There exist finitely many , independent of , a finite set (of size depending on ), and a finite partition of into subsets , such that for all :
for some constants depending on and . Moreover, for fixed and , the set is either finite or a fixed congruence class modulo some .
To prove the proposition, it is enough to show that for each and , we have on each , for some constant depending on . It is enough to prove this for infinite, as otherwise we have
Now suppose is infinite. By rearranging the constants , we may assume that the pairs are disjoint, and that in lexicographic order, that is, either or and . Note that if , then we are done, since for each and each :
(6.2) |
where . Assume towards contradiction that , and . Then for all large enough , one has
(6.3) |
Now let be the map , for . Then by [GH24, Corollary 3.18], is (FRS). By Corollary 6.2, we have
(6.4) |
On the other hand, note that
Further note that
Thus, we have
By (6.3), for each large enough, we may find , such that
We are now ready to prove of Theorem 6.6.
Proof of of Theorem 6.6, non-Archimedean case.
Let be a jet-flat morphism. We may assume . Let and write for the density of with respect to . Let be the set of such that is smooth over . For every , the map is smooth over , and therefore is locally constant on .
By [CGH18, Corollary 1.4.3], the constancy radius of can be taken to be definable, i.e. there exists a definable function such that is constant around every ball . In particular, for every we have . In addition, by Proposition 6.8, we have , for . We arrive at the following:
where the last inequality follows by [CGH18, Theorem 3.1.1], since and thus for some and every large enough. ∎
In [CGH23, Theorem 4.12], Cluckers and the first two authors showed that if is a jet-flat morphism, which is defined over , and one chooses and to be the canonical measures on and (see [CGH23, Lemma 4.2]), then the constant in Proposition 6.8 can be taken to be independent of (i.e. ). [CGH23, Theorem 4.12], together with the ideas of the proof of of Theorem 6.6, allows us to give bounds on the norms of , which are independent of :
Proposition 6.9.
Let be a dominant morphism between finite type -schemes and , with smooth and geometrically irreducible. For any prime , let be the density of with respect to . Then the following are equivalent:
-
(1)
is jet-flat.
-
(2)
For every , there exists such that for every prime ,
Proof.
The implication follows from Proposition 6.7. By of Theorem 6.6, it is enough to prove for . Suppose that is jet-flat. We may assume that . Indeed, working locally, and since is smooth, we may assume there exists a morphism , such that is an étale map. If is the density of with respect to , and is an upper bound on the size of the geometric fibers of , then
for large enough. In particular, we have:
As in the proof of of Theorem 6.6 above, by [CGH18, Corollary 1.4.3], there exists a definable function such that , where . Hence, applying [CGH23, Theorem 4.12] we can find such that for :
By [CGH18, Theorem 3.1.1], there exists and such that for every and every prime . We therefore get the desired claim as:
6.3. : the Archimedean case
In this subsection we prove in the cases and . Let be a jet-flat morphism between smooth algebraic varieties, defined over . Using restriction of scalars, we may assume that . We would like to show that for each , we have for all . We start with the following proposition.
Proposition 6.10.
Let be a jet-flat map between smooth -varieties. Then for every , every non-vanishing , and every , one can find and such that for each and one has
Remark 6.11.
Proposition 6.10 is weaker than its non-Archimedean counterpart (Proposition 6.8). The main obstacle is that we do not know of an Archimedean analogue to [CGH18, Theorem 2.1.3], that is, whether one can approximate the supremum of constructible functions by constructible functions as in (6.1). It is conjectured by Raf Cluckers that such a statement should be true under suitable assumptions (see [AM, Conjecture 6.9]). We thank the anonymous referee for bringing this to our attention. We further believe that the current proposition should hold for for a sufficiently large independent of .
Analogously to the non-Archimedean case, we introduce the following notion of constructible functions:
Definition 6.12 ([CM11, Section 1.1], see also [LR97]).
-
(1)
A restricted analytic function is a function such that is analytic and .
-
(2)
A subset is subanalytic if it is definable in – the extension of the ordered real field by all restricted analytic functions. A function is subanalytic if its graph is subanalytic.
-
(3)
A function is called constructible if there exist subanalytic functions and , such that:
(6.5) We denote the class of constructible functions on by .
-
(4)
Given an analytic manifold , a measure is called constructible if locally it is of the form , where is a regular top-form, and is constructible. We denote the class of constructible measures by . Similarly, we write , and .
Note that is defined by polynomials, so it is definable in the real field, and in particular subanalytic. Since contains indicators of balls, we may assume that when proving Proposition 6.10 and Theorem 6.6.
Proof of Proposition 6.10.
We may assume that , and . Write for the density of with respect to . Set
(6.6) |
For each , let
By [CM11, Theorem 1.3], the functions and are constructible. Writing as in (6.5), and using a preparation theorem for constructible functions [CM12, Corollary 3.5], there exist and , such that or , and for each one can write:
(6.7) |
where , , , and where are certain subanalytic units, called strong functions (see [CM11, Definition 2.3]). We may assume that as otherwise, is bounded on and we are done. In addition, [CM12, Corollary 3.5] also ensures that for each , either for each , or . This additional property is achieved by writing each strong function as a converging infinite sum for some , and split it into a finite sum and an infinite sum . By taking large enough, and rearranging the terms in (6.7), Cluckers and Miller ensured that whenever . Following the same argument, and taking even larger, one can guarantee that for some fixed , as large as we wish. Hence, we may assume that has the following form:
(6.8) |
where for , and large as we like. For , we may further assume that are mutually different and lexicographically ordered, i.e. either , or and . In particular, by taking small enough, we have for :
(6.9) |
We claim that . Assume not, then we have for small enough:
(6.10) |
We now use an argument analogous to the one in Proposition 6.8. Take to be the map
for and let be a bump function which is equal to one on the unit ball in . Then is (FRS), and thus by Corollary 6.2
(6.11) |
Repeating precisely the same argument as in Proposition 6.8, and using (6.10), we may find such that:
which leads to a contradiction. Hence , therefore on for small enough we have
For , we have and thus . This concludes the proof. ∎
In order to prove Theorem 6.6, we need to control the oscillations of constructible functions.
Definition 6.13.
Let be a subanalytic function. Define by
(6.12) |
Note that is subanalytic, and for any we have for almost every . The next lemma extends this construction to the ring of constructible functions.
Lemma 6.14.
Let . Then there exists a subanalytic function such that for any we have for almost all , and
(6.13) |
Proof.
If for , and suppose we already constructed , for each . Then we may set . Hence, by (6.5), we may assume that
(6.14) |
for some subanalytic . By setting for , we may assume .
There is an open subanalytic subset with complement of measure , such that and all are continuous (see e.g. [DvdD88, Theorem 3.2.11]). Set
and denote . Let be as in (6.12) and set
Note that , are subanalytic, and thus also is subanalytic. Moreover, by the continuity of we get that and thus also for all . Note that for any real numbers we have
(6.15) |
By (6.12), for every and we have , and
Similarly, we have:
and
By (6.15), for every we have:
We can now finish the proof of Theorem 6.6.
Proof of the Archimedean part of of Theorem 6.6.
Let and write for the density of with respect to . Let be as in Lemma 6.14, and set
Then for each define the following subanalytic set:
We fix large enough. Setting
(6.16) |
and using Lemma 6.14, Proposition 6.10 and Hölder’s inequality, we have:
(6.17) |
Since is a constructible function, using a similar argument as in the Proposition 6.10, and writing as in (6.8) and (6.9), we get , as , for and . But since , we must have . In particular, taking large enough, we get that for some . Thus, the following holds for any :
(6.18) |
References
- [AA16] A. Aizenbud and N. Avni. Representation growth and rational singularities of the moduli space of local systems. Invent. Math., 204(1):245–316, 2016.
- [AG] N. Avni and I. Glazer. On the Fourier coefficients of word maps on unitary groups. arXiv:2210.04164.
- [AGL] N. Avni, I. Glazer, and M. Larsen. Fourier and small ball estimates for word maps on unitary groups. arXiv:2402.11108.
- [AGZV88] V. I. Arnold, S. M. Guseĭn-Zade, and A. N. Varchenko. Singularities of differentiable maps. Vol. II, volume 83 of Monographs in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1988. Monodromy and asymptotics of integrals, Translated from the Russian by Hugh Porteous, Translation revised by the authors and James Montaldi.
- [AM] F. Adiceam and O. Marmon. Homogeneous Forms Inequalities. arXiv:2305.19782.
- [Ati70] M. F. Atiyah. Resolution of singularities and division of distributions. Comm. Pure Appl. Math., 23:145–150, 1970.
- [Ber72] I. N. Bernšteĭn. Analytic continuation of generalized functions with respect to a parameter. Funkcional. Anal. i Priložen., 6(4):26–40, 1972.
- [BG69] I. N. Bernšteĭn and S. I. Gefand. Meromorphy of the function . Funkcional. Anal. i Priložen., 3(1):84–85, 1969.
- [BM88] E. Bierstone and P. D. Milman. Semianalytic and subanalytic sets. Inst. Hautes Études Sci. Publ. Math., (67):5–42, 1988.
- [BM89] E. Bierstone and P. D. Milman. Uniformization of analytic spaces. J. Amer. Math. Soc., 2(4):801–836, 1989.
- [CCL12] R. Cluckers, G. Comte, and F. Loeser. Local metric properties and regular stratifications of -adic definable sets. Comment. Math. Helv., 87(4):963–1009, 2012.
- [CCW99] A. Carbery, M. Christ, and J. Wright. Multidimensional van der Corput and sublevel set estimates. J. Amer. Math. Soc., 12(4):981–1015, 1999.
- [CGH14] R. Cluckers, J. Gordon, and I. Halupczok. Integrability of oscillatory functions on local fields: transfer principles. Duke Math. J., 163(8):1549–1600, 2014.
- [CGH18] R. Cluckers, J. Gordon, and I. Halupczok. Uniform analysis on local fields and applications to orbital integrals. Trans. Amer. Math. Soc. Ser. B, 5:125–166, 2018.
- [CGH23] R. Cluckers, I. Glazer, and Y. I. Hendel. A number theoretic characterization of E-smooth and (FRS) morphisms: estimates on the number of -points. Algebra Number Theory, 17(12):2229–2260, 2023.
- [CL08] R. Cluckers and F. Loeser. Constructible motivic functions and motivic integration. Invent. Math., 173(1):23–121, 2008.
- [CL10] R. Cluckers and F. Loeser. Constructible exponential functions, motivic Fourier transform and transfer principle. Ann. of Math. (2), 171(2):1011–1065, 2010.
- [CLNS18] A. Chambert-Loir, J. Nicaise, and J. Sebag. Motivic integration, volume 325 of Progress in Mathematics. Birkhäuser/Springer, New York, 2018.
- [CM11] R. Cluckers and D. J. Miller. Stability under integration of sums of products of real globally subanalytic functions and their logarithms. Duke Math. J., 156(2):311–348, 2011.
- [CM12] R. Cluckers and D. J. Miller. Loci of integrability, zero loci, and stability under integration for constructible functions on Euclidean space with Lebesgue measure. Int. Math. Res. Not. IMRN, (14):3182–3191, 2012.
- [CM13] R. Cluckers and D. J. Miller. Lebesgue classes and preparation of real constructible functions. J. Funct. Anal., 264(7):1599–1642, 2013.
- [CMN19] R. Cluckers, M. Mustaţă, and K. H. Nguyen. Igusa’s conjecture for exponential sums: optimal estimates for nonrational singularities. Forum Math. Pi, 7:e3, 28, 2019.
- [Den91] J. Denef. Report on Igusa’s local zeta function. Number 201-203, pages Exp. No. 741, 359–386 (1992). 1991. Séminaire Bourbaki, Vol. 1990/91.
- [dFM09] T. de Fernex and M. Mustaţă. Limits of log canonical thresholds. Ann. Sci. Éc. Norm. Supér. (4), 42(3):491–515, 2009.
- [DvdD88] J. Denef and L. van den Dries. -adic and real subanalytic sets. Ann. of Math. (2), 128(1):79–138, 1988.
- [EI15] L. Ein and S. Ishii. Singularities with respect to Mather-Jacobian discrepancies. In Commutative algebra and noncommutative algebraic geometry. Vol. II, volume 68 of Math. Sci. Res. Inst. Publ., pages 125–168. Cambridge Univ. Press, New York, 2015.
- [EM09] L. Ein and M. Mustaţă. Jet schemes and singularities. In Algebraic geometry—Seattle 2005. Part 2, volume 80 of Proc. Sympos. Pure Math., pages 505–546. Amer. Math. Soc., Providence, RI, 2009.
- [Fed69] H. Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.
- [Fee19] P. M. N. Feehan. Resolution of singularities and geometric proofs of the łojasiewicz inequalities. Geom. Topol., 23(7):3273–3313, 2019.
- [GGH] I. Glazer, J. Gordon, and Y. I. Hendel. Integrability and singularities of Harish-Chandra characters. arXiv:2312.01591.
- [GH19] I. Glazer and Y. I. Hendel. On singularity properties of convolutions of algebraic morphisms. Selecta Math. (N.S.), 25(1):Art. 15, 41, 2019.
- [GH21] I. Glazer and Y. I. Hendel. On singularity properties of convolutions of algebraic morphisms—the general case. J. Lond. Math. Soc. (2), 103(4):1453–1479, 2021. With an appendix by Glazer, Hendel and Gady Kozma.
- [GH24] I. Glazer and Y. I. Hendel. On singularity properties of word maps and applications to probabilistic Waring-type problems. Mem. Amer. Math. Soc., to appear, 2024.
- [GS22] I. Goldsheid and S. Sodin. Lower bounds on Anderson-localised eigenfunctions on a strip. Comm. Math. Phys., 392(1):125–144, 2022.
- [Hir64] H. Hironaka. Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II. Ann. of Math. (2) (1964), 109–203; ibid. (2), 79:205–326, 1964.
- [How01] J. A. Howald. Multiplier ideals of monomial ideals. Trans. Amer. Math. Soc., 353(7):2665–2671, 2001.
- [HS06] C. Huneke and I. Swanson. Integral closure of ideals, rings, and modules, volume 336 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006.
- [Igu74] J. Igusa. Complex powers and asymptotic expansions. I. Functions of certain types. J. Reine Angew. Math., 268/269:110–130, 1974. Collection of articles dedicated to Helmut Hasse on his seventy-fifth birthday, II.
- [Igu78] J. Igusa. Forms of higher degree, volume 59 of Tata Institute of Fundamental Research Lectures on Mathematics and Physics. Tata Institute of Fundamental Research, Bombay; by the Narosa Publishing House, New Delhi, 1978.
- [Ish18] S. Ishii. Introduction to singularities. Springer, Tokyo, 2018. Second edition of [ MR3288750].
- [Jea70] P. Jeanquartier. Développement asymptotique de la distribution de Dirac attachée à une fonction analytique. C. R. Acad. Sci. Paris Sér. A-B, 201:A1159–A1161, 1970.
- [Kol] J. Kollár. Which powers of holomorphic functions are integrable? arXiv:0805.0756.
- [Kol97] J. Kollár. Singularities of pairs. In Algebraic geometry—Santa Cruz 1995, volume 62 of Proc. Sympos. Pure Math., pages 221–287. Amer. Math. Soc., Providence, RI, 1997.
- [KR78] A. Klein and B. Russo. Sharp inequalities for Weyl operators and Heisenberg groups. Math. Ann., 235(2):175–194, 1978.
- [KSB88] J. Kollár and N. I. Shepherd-Barron. Threefolds and deformations of surface singularities. Invent. Math., 91(2):299–338, 1988.
- [LC22] E. León-Cardenal. Archimedean zeta functions and oscillatory integrals. In -adic analysis, arithmetic and singularities, volume 778 of Contemp. Math., pages 3–24. Amer. Math. Soc., [Providence], RI, [2022] ©2022.
- [Lel57] Pierre Lelong. Intégration sur un ensemble analytique complexe. Bull. Soc. Math. France, 85:239–262, 1957.
- [LG86] Pierre Lelong and Lawrence Gruman. Entire functions of several complex variables, volume 282 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1986.
- [Lo65] S. Ł ojasiewicz. Ensembles semi-analytiques. In IHES notes. 1965.
- [LR97] J.-M. Lion and J.-P. Rolin. Théorème de préparation pour les fonctions logarithmico-exponentielles. Ann. Inst. Fourier (Grenoble), 47(3):859–884, 1997.
- [LST19] M. Larsen, A. Shalev, and P. H. Tiep. Probabilistic Waring problems for finite simple groups. Ann. of Math. (2), 190(2):561–608, 2019.
- [Mal74] Bernard Malgrange. Intégrales asymptotiques et monodromie. Ann. Sci. École Norm. Sup. (4), 7:405–430 (1975), 1974.
- [Meu16] D. Meuser. A survey of Igusa’s local zeta function. Amer. J. Math., 138(1):149–179, 2016.
- [MSS18] L. Maxim, M. Saito, and J. Schurmann. Thom-sebastiani theorems for filtered d-modules and for multiplier ideals. International Mathematics Research Notices, 2018.
- [Mus01] M. Mustaţă. Jet schemes of locally complete intersection canonical singularities. Invent. Math., 145(3):397–424, 2001. With an appendix by David Eisenbud and Edward Frenkel.
- [Mus02] M. Mustaţă. Singularities of pairs via jet schemes. J. Amer. Math. Soc., 15(3):599–615, 2002.
- [Mus12] M. Mustaţă. IMPANGA lecture notes on log canonical thresholds. In Contributions to algebraic geometry, EMS Ser. Congr. Rep., pages 407–442. Eur. Math. Soc., Zürich, 2012. Notes by Tomasz Szemberg.
- [Rei] A. Reiser. Pushforwards of measures on real varieties under maps with rational singularities. arXiv:1807.00079.
- [RS88] F. Ricci and E. M. Stein. Harmonic analysis on nilpotent groups and singular integrals. II. Singular kernels supported on submanifolds. J. Funct. Anal., 78(1):56–84, 1988.
- [Ser81] J. P. Serre. Quelques applications du théorème de densité de Chebotarev. Inst. Hautes Études Sci. Publ. Math., (54):323–401, 1981.
- [ST71] M. Sebastiani and R. Thom. Un résultat sur la monodromie. Invent. Math., 13:90–96, 1971.
- [Ste87] E. M. Stein. Problems in harmonic analysis related to curvature and oscillatory integrals. In Proceedings of the International Congress of Mathematicians, Vol. 1, 2 (Berkeley, Calif., 1986), pages 196–221. Amer. Math. Soc., Providence, RI, 1987.
- [Ste93] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
- [SVW98] C. Shubin, R. Vakilian, and T. Wolff. Some harmonic analysis questions suggested by Anderson-Bernoulli models. Geom. Funct. Anal., 8(5):932–964, 1998.
- [Thi67] Paul R. Thie. The Lelong number of a point of a complex analytic set. Math. Ann., 172:269–312, 1967.
- [Var83] A. N. Varchenko. Semicontinuity of the complex singularity exponent. Funktsional. Anal. i Prilozhen., 17(4):77–78, 1983.
- [Vog78] D. A. Vogan, Jr. Gelfand-Kirillov dimension for Harish-Chandra modules. Invent. Math., 48(1):75–98, 1978.
- [VZnG08] W. Veys and W. A. Zúñiga Galindo. Zeta functions for analytic mappings, log-principalization of ideals, and Newton polyhedra. Trans. Amer. Math. Soc., 360(4):2205–2227, 2008.
- [VZnG17] W. Veys and W. A. Zúñiga Galindo. Zeta functions and oscillatory integrals for meromorphic functions. Adv. Math., 311:295–337, 2017.
- [Wei40] A. Weil. L’intégration dans les groupes topologiques et ses applications, volume No. 869 of Actualités Scientifiques et Industrielles [Current Scientific and Industrial Topics]. Hermann & Cie, Paris, 1940. [This book has been republished by the author at Princeton, N. J., 1941.].
- [Wlo09] J. Wlodarczyk. Resolution of singularities of analytic spaces. In Proceedings of Gökova Geometry-Topology Conference 2008, pages 31–63. Gökova Geometry/Topology Conference (GGT), Gökova, 2009.