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

Integrability of pushforward measures by analytic maps

Itay Glazer Department of Mathematics, University of Oxford, Andrew Wiles Building, Radcliffe Observatory Quarter (550), Woodstock Road, Oxford, OX2 6GG, UK and Department of Mathematics, Technion – Israel Institute of Technology, Haifa 3200003, Israel [email protected] https://sites.google.com/view/itay-glazer Yotam I. Hendel KU Leuven, Department of Mathematics, B-3001 Leuven, Belgium and Ben Gurion University of the Negev, Department of Mathematics, Be’er Sheva 8410501, Israel [email protected] https://sites.google.com/view/yotam-hendel  and  Sasha Sodin School of Mathematical Sciences, Queen Mary University of London, Mile End Road, London E1 4NS, UK and Einstein Institute of Mathematics, The Hebrew University of Jerusalem, Givat Ram, Jerusalem 91904, Israel [email protected] https://webspace.maths.qmul.ac.uk/a.sodin/
Abstract.

Given a map ϕ:XY\phi:X\rightarrow Y between FF-analytic manifolds over a local field FF of characteristic 0, we introduce an invariant ϵ(ϕ)\epsilon_{\star}(\phi) which quantifies the integrability of pushforwards of smooth compactly supported measures by ϕ\phi. We further define a local version ϵ(ϕ,x)\epsilon_{\star}(\phi,x) near xXx\in X. These invariants have a strong connection to the singularities of ϕ\phi.

When YY is one-dimensional, we give an explicit formula for ϵ(ϕ,x)\epsilon_{\star}(\phi,x), and show it is asymptotically equivalent to other known singularity invariants such as the FF-log-canonical threshold lctF(ϕϕ(x);x)\operatorname{lct}_{F}(\phi-\phi(x);x) at xx.

In the general case, we show that ϵ(ϕ,x)\epsilon_{\star}(\phi,x) is bounded from below by the FF-log-canonical threshold λ=lctF(𝒥ϕ;x)\lambda=\operatorname{lct}_{F}(\mathcal{J}_{\phi};x) of the Jacobian ideal 𝒥ϕ\mathcal{J}_{\phi} near xx. If dimY=dimX\dim Y=\dim X, equality is attained. If dimY<dimX\dim Y<\dim X, the inequality can be strict; however, for F=F=\mathbb{C}, we establish the upper bound ϵ(ϕ,x)λ/(1λ)\epsilon_{\star}(\phi,x)\leq\lambda/(1-\lambda), whenever λ<1\lambda<1.

Finally, we specialize to polynomial maps φ:XY\varphi:X\rightarrow Y between smooth algebraic \mathbb{Q}-varieties XX and YY. We geometrically characterize the condition that ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty over a large family of local fields, by showing it is equivalent to φ\varphi being flat with fibers of semi-log-canonical singularities.

Key words and phrases:
LpL^{p}-spaces, log-canonical threshold, analytic maps, pushforward measures, constructible functions, Young’s convolution inequality, regularization by convolution
1991 Mathematics Subject Classification:
14B05, 03C98, 14E15, 32B20, 60B15

1. Introduction

The goal of this paper is to explore a singularity invariant ϵ(ϕ)\epsilon_{\star}(\phi) of a map ϕ\phi between two manifolds over a local field. This invariant quantifies the integrability of pushforward measures by ϕ\phi; 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 FF of characteristic 0, i.e., \mathbb{R}, \mathbb{C} or a finite extension of p\mathbb{Q}_{p}. If XX is an FF-analytic manifold of dimension nn, let (UαX,ψα:UαFn)α𝒜(U_{\alpha}\subset X,\psi_{\alpha}:U_{\alpha}\to F^{n})_{\alpha\in\mathcal{A}} be an atlas. We denote by C(X)C^{\infty}(X) the space of smooth functions on XX, i.e. functions f:Xf:X\to\mathbb{C} such that fψα1|ψα(Uα)f\circ\psi_{\alpha}^{-1}|_{\psi_{\alpha}(U_{\alpha})} is smooth for each α𝒜\alpha\in\mathcal{A}, and by Cc(X)C_{c}^{\infty}(X) the subspace of compactly supported smooth functions (if FF is non-Archimedean, smooth means locally constant). We similarly write (X)\mathcal{M}^{\infty}(X) for the space of smooth measures on XX, i.e. measures such that each (ψα)(μ|Uα)(\psi_{\alpha})_{*}(\mu|_{U_{\alpha}}) has a smooth density with respect to the Haar measure. We use c(X)\mathcal{M}_{c}^{\infty}(X) to denote the space of smooth compactly supported measures on XX. For 1q1\leq q\leq\infty, consider the class c,q(X)\mathcal{M}_{c,q}(X) of finite Radon measures μ\mu on XX that are compactly supported and such that for any α𝒜\alpha\in\mathcal{A} the measure (ψα)(μ|Uα)(\psi_{\alpha})_{*}(\mu|_{U_{\alpha}}) is absolutely continuous with density in Lq(Fn)L^{q}(F^{n}). All these classes do not depend on the choice of the atlas. For μc,1(X)\mu\in\mathcal{M}_{c,1}(X) we define

ϵ(μ):=sup{ϵ0|μc,1+ϵ(X)}.\epsilon_{\star}(\mu):=\sup\left\{\epsilon\geq 0\,|\,\mu\in\mathcal{M}_{c,1+\epsilon}(X)\right\}.

Note that by Jensen’s inequality μc,1+ϵ(X)\mu\in\mathcal{M}_{c,1+\epsilon}(X) for all 0ϵ<ϵ(μ)0\leq\epsilon<\epsilon_{\star}(\mu).

Now let ϕ:XY\phi:X\to Y be an analytic map between FF-analytic manifolds X,YX,Y. If ϕ\phi is locally dominant, i.e. ϕ(U)\phi(U) contains a non-empty open set for each open subset UXU\subseteq X, then ϕμc,1(Y)\phi_{*}\mu\in\mathcal{M}_{c,1}(Y) whenever μc,1(X)\mu\in\mathcal{M}_{c,1}(X). We now set for each xXx\in X,

(1.1) ϵ(ϕ;x):=supUxinfμc(U)ϵ(ϕμ)=supUxinfμc,(U)ϵ(ϕμ),\epsilon_{\star}(\phi;x):=\sup_{U\ni x}\inf_{\mu\in\mathcal{M}_{c}^{\infty}(U)}\epsilon_{\star}(\phi_{*}\mu)=\sup_{U\ni x}\inf_{\mu\in\mathcal{M}_{c,\infty}(U)}\epsilon_{\star}(\phi_{*}\mu),

where the supremum is over all open neighborhoods UU of xx. Finally, we can define

(1.2) ϵ(ϕ):=infμc(X)ϵ(ϕμ)=infxXϵ(ϕ;x).\epsilon_{\star}(\phi):=\inf_{\mu\in\mathcal{M}_{c}^{\infty}(X)}\epsilon_{\star}(\phi_{*}\mu)=\inf_{x\in X}\epsilon_{\star}(\phi;x).

Note that if there exists UxU\ni x such that ϕμ\phi_{*}\mu lies in c,q(Y)\mathcal{M}_{c,q}(Y) for all 1<q<1<q<\infty and all μc(U)\mu\in\mathcal{M}_{c}^{\infty}(U), then ϵ(ϕ;x)=\epsilon_{\star}(\phi;x)=\infty. The best case scenario in this setting is obtained when ϕμc,(Y)\phi_{*}\mu\in\mathcal{M}_{c,\infty}(Y) for every μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X). In this case we say that ϕ\phi is an LL^{\infty}-map.

The main motivation for ϵ(ϕ;x)\epsilon_{\star}(\phi;x) comes from singularity theory. In general, bad singularities of ϕ\phi should manifest themselves in poor analytic behavior of the pushforward ϕμ\phi_{*}\mu of μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X). This phenomenon has been extensively studied in the case Y=FmY=F^{m}, through the analysis of the Fourier transform (ϕμ)\mathcal{F}(\phi_{*}\mu), which takes the form of an oscillatory integral in the Archimedean case, and of an exponential sum in the pp-adic case. When the dimension of the target space is equal to m=1m=1, the rate of decay of (φμ)\mathcal{F}(\varphi_{*}\mu) 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 ϕ\phi still result in faster decay rates of (ϕμ)\mathcal{F}(\phi_{*}\mu). For more information on (ϕμ)\mathcal{F}(\phi_{*}\mu), 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 §\mathsection3.2 below.

The study of the integrability properties of ϕμ\phi_{*}\mu and their relation to the singularities of ϕ\phi has received less treatment. The invariant ϵ(ϕ;x)\epsilon_{\star}(\phi;x) is a natural step in this direction, and is more robust than Fourier-type invariants as it is also meaningful when YY is any smooth manifold, which is especially important for applications (see §\mathsection1.1).

The invariant ϵ(ϕ;x)\epsilon_{\star}(\phi;x) tends to be small whenever the singularities of ϕ\phi are bad near xx. When ϕ\phi is a submersion, the pushforward ϕμ\phi_{*}\mu of any μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X) is smooth and in particular lies in c,(Y)\mathcal{M}_{c,\infty}(Y). Moreover, Aizenbud and Avni [AA16] have shown that for algebraic maps φ:XY\varphi:X\rightarrow Y between smooth algebraic \mathbb{Q}-varieties, the condition that the corresponding map φp:X(p)Y(p)\varphi_{\mathbb{Q}_{p}}:X(\mathbb{Q}_{p})\rightarrow Y(\mathbb{Q}_{p}) of p\mathbb{Q}_{p}-analytic varieties is an LL^{\infty}-map is equivalent to a certain mild singularity property, namely that φ\varphi is flat with fibers of rational singularities (abbreviated (FRS), see Definition 1.8).

When analyzing ϵ(ϕ;x)\epsilon_{\star}(\phi;x), one can further restrict the infima in (1.1)–(1.2) to the class of compactly supported measures μ\mu 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 ϕμ\phi_{*}\mu 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 ϵ(ϕμ)>0\epsilon_{\star}(\phi_{*}\mu)>0. Positivity of ϵ(ϕμ)\epsilon_{\star}(\phi_{*}\mu) 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 ϵ(ϕ;x)\epsilon_{\star}(\phi;x), and in particular to obtain upper and lower bounds on ϵ(ϕ;x)\epsilon_{\star}(\phi;x) 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 §\mathsection4), we give a lower bound on ϵ(ϕ;x)\epsilon_{\star}(\phi;x). In particular, this provides a proof for the positivity of ϵ\epsilon_{\star}, without using the theory of motivic integration. In Theorem 1.5 (proved in §\mathsection5) an upper bound on ϵ(ϕ;x)\epsilon_{\star}(\phi;x) is given in the complex case. In Theorem 1.2 (proved in §\mathsection3) an explicit formula is given for ϵ(ϕ;x)\epsilon_{\star}(\phi;x) when YY is one-dimensional, over any local field. Finally, in Theorem 1.12 (proved in §\mathsection6), we specialize to polynomial maps between smooth algebraic varieties, and geometrically characterize the condition ϵ(ϕ;x)=\epsilon_{\star}(\phi;x)=\infty.

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 GG is an FF-analytic group, and take a finite measure ν\nu on GG. Can one find a number kk\in\mathbb{N} such that the kk-th convolution power νk\nu^{*k} lies in c,(G)\mathcal{M}_{c,\infty}(G), and if so, what is the smallest such number k(ν)k_{\star}(\nu)?

An important class of examples comes from the realm of word maps. Given a word ww in rr letters, by which we mean an element either of the free group FrF_{r}, or of the free Lie algebra r\mathcal{L}_{r}, one can consider the corresponding word maps wG:GrGw_{G}:G^{r}\rightarrow G or w𝔤:𝔤r𝔤w_{\mathfrak{g}}:\mathfrak{g}^{r}\rightarrow\mathfrak{g}, where 𝔤\mathfrak{g} is the Lie algebra of GG. When GG is a compact real or pp-adic Lie group, ww induces a natural measure ν:=(wG)πG\nu:=(w_{G})_{*}\pi_{G} where πG\pi_{G} is the normalized Haar measure on GG. Here, one may further ask what is the LL^{\infty}-mixing time of the word measure ν\nu, namely, how large should kk be to ensure that νkπG1\left\|\nu^{*k}-\pi_{G}\right\|_{\infty}\ll 1. 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(2,2,\mathbb{R}), or, more generally, in Sp(2W,2W,\mathbb{R})) 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: GG and XX are analytic, and ν=ϕμ\nu=\phi_{*}\mu is the pushforward of a measure μc,(X)\mu\in\mathcal{M}_{c,\infty}(X) under a locally dominant analytic map ϕ:XG\phi:X\to G. For xXx\in X, let

k(ϕ;x)=minUxmaxμc(U)k(ϕμ),k_{\star}(\phi;x)=\min_{U\ni x}\max_{\mu\in\mathcal{M}_{c}^{\infty}(U)}k_{\star}(\phi_{*}\mu),

be the smallest kk that works for any μ\mu supported in a sufficiently small neighborhood UU of xx.

According to Young’s convolution inequality for locally compact groups (see [KR78, Corollary 2.3] and Remark 1.10),

νc,1+ϵ(G)νkc,1+r(G),wherer={kϵ1(k1)ϵifk<1+ϵϵ,ifk1+ϵϵ.\nu\in\mathcal{M}_{c,1+\epsilon}(G)\Longrightarrow\nu^{*k}\in\mathcal{M}_{c,1+r}(G),\quad\text{where}\quad r=\begin{cases}\frac{k\epsilon}{1-(k-1)\epsilon}&\text{if}\,k<\frac{1+\epsilon}{\epsilon},\\ \infty&\text{if}\,k\geq\frac{1+\epsilon}{\epsilon}.\end{cases}

This implies

(1.3) k(ϕ;x)1+ϵ(ϕ;x)ϵ(ϕ;x)+1<.k_{\star}(\phi;x)\leq\left\lfloor\frac{1+\epsilon_{\star}(\phi;x)}{\epsilon_{\star}(\phi;x)}\right\rfloor+1<\infty.

As mentioned above, a more classical approach to bounding k(ϕ;x)k_{\star}(\phi;x) relies on the study of the decay of the Fourier transform of ϕμ\phi_{*}\mu. While the Fourier-analytic approach often provides sharper bounds (see [GH24, Proposition 5.7]), it is mainly applicable for abelian groups such as G=FnG=F^{n}, or mildly non-abelian groups such as the Heisenberg model. One can use non-commutative Fourier transform to analyze compact Lie groups such as SOn()\mathrm{SO}_{n}(\mathbb{R}), but such representation theoretic techniques are much less effective for compact pp-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 ϕ\phi. Thus one can argue that the approach to regularization via (1.3) is currently the most efficient one for treating pp-adic analytic groups and non-compact Lie groups such as SLn()SL_{n}(\mathbb{R}).

1.2. Main results

We now discuss the main results in this paper.

1.2.1. A lower bound on ϵ\epsilon_{\star}

While the mere positivity of ϵ\epsilon_{\star} (and the mere finiteness of kk_{\star}) 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 F=F=\mathbb{C}. For an analytic map ψ:XF\psi:X\to F, define the FF-log-canonical threshold

(1.4) lctF(ψ;x):=sup{s>0:Ux s.t.μc,(U),X|ψ(x)|Fs𝑑μ(x)<},\operatorname{lct}_{F}(\psi;x):=\sup\left\{s>0:\exists U\ni x\text{ s.t.}\,\forall\mu\in\mathcal{M}_{c,\infty}(U),\,\int_{X}\left|\psi(x)\right|_{F}^{-s}d\mu(x)<\infty\right\},

where UU runs over all open neighborhoods of xx, and ||F\left|\,\cdot\,\right|_{F} is the absolute value on FF, normalized such that μF(aA)=|a|FμF(A)\mu_{F}(aA)=\left|a\right|_{F}\cdot\mu_{F}(A), for all aF×a\in F^{\times}, AFA\subseteq F, and where μF\mu_{F} is a Haar measure on FF. In particular, ||=||2\left|\cdot\right|_{\mathbb{C}}=\left|\cdot\right|^{2} is the square of the usual absolute value ||\left|\cdot\right| on \mathbb{C}. More generally, if JJ is a non-zero ideal of analytic functions generated by ψ1,,ψ\psi_{1},\cdots,\psi_{\ell}, define

(1.5) lctF(J;x):=sup{s>0:Ux s.t. μc,(U),Xmin1il[|ψi(x)|Fs]𝑑μ(x)<}.\operatorname{lct}_{F}(J;x):=\sup\left\{s>0:\exists U\ni x\text{ \,s.t.\, }\forall\mu\in\mathcal{M}_{c,\infty}(U),\,\int_{X}\min_{1\leq i\leq l}\Big{[}\left|\psi_{i}(x)\right|_{F}^{-s}\Big{]}d\mu(x)<\infty\right\}.

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 §\mathsection2).

Given a locally dominant analytic map ϕ:XY\phi:X\to Y between two FF-analytic manifolds, we define the Jacobian ideal sheaf 𝒥ϕ\mathcal{J}_{\phi} as follows. If XFnX\subseteq F^{n} and YFmY\subseteq F^{m} are open subsets, we define 𝒥ϕ\mathcal{J}_{\phi} to be the ideal in the algebra of analytic functions on XX, generated by the m×mm\times m-minors of the differential dx(ϕ)d_{x}(\phi) of ϕ\phi. Note that if ψ1:XXFn\psi_{1}:X\rightarrow X^{\prime}\subseteq F^{n} and ψ2:YYFm\psi_{2}:Y\rightarrow Y^{\prime}\subseteq F^{m} are analytic diffeomorphisms, then ψ1(𝒥ψ2ϕψ11)=𝒥ϕ\psi_{1}^{*}\left(\mathcal{J}_{\psi_{2}\circ\phi\circ\psi_{1}^{-1}}\right)=\mathcal{J}_{\phi}. Hence, the definition of 𝒥ϕ\mathcal{J}_{\phi} can be generalized (or glued) to an ideal sheaf on XX, if XX and YY are FF-analytic manifolds.

We now describe the first main result, which provides a lower bound on ϵ(ϕ;x)\epsilon_{\star}(\phi;x) in terms of the Jacobian ideal 𝒥ϕ\mathcal{J}_{\phi} of ϕ\phi. The proof is given in §\mathsection4.

Theorem 1.1.

Let X,YX,Y be analytic FF-manifolds, dimX=ndimY=m\dim X=n\geq\dim Y=m, and let ϕ:XY\phi:X\to Y be a locally dominant analytic map. Then for every xXx\in X,

(1.6) ϵ(ϕ;x)lctF(𝒥ϕ;x);\epsilon_{\star}(\phi;x)\geq\operatorname{lct}_{F}(\mathcal{J}_{\phi};x);

if m=nm=n, equality is achieved.

As we will see later in Remark 1.6(2), for dimX>dimY\dim X>\dim Y the inequality (1.6) may be strict. In the next paragraph we discuss an additional case in which ϵ\epsilon_{\star} can be computed explicitly: dimY=1\dim Y=1.

1.2.2. A formula in the one-dimensional case and a reverse Young inequality

When the target space YY 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 §\mathsection3.

The first one is an exact formula relating ϵ\epsilon_{\star} to the log-canonical threshold.

Theorem 1.2.

Let XX be an analytic FF-manifold, and let ϕ:XF\phi:X\to F be a locally dominant analytic map. Then for each xXx\in X, we have:

(1.7) ϵ(ϕ;x)={if lctF(ϕϕ(x);x)1,lctF(ϕϕ(x);x)1lctF(ϕϕ(x);x)if lctF(ϕϕ(x);x)<1.\epsilon_{\star}(\phi;x)=\begin{cases}\infty&\text{if }\operatorname{lct}_{F}(\phi-\phi(x);x)\geq 1,\\ \frac{\operatorname{lct}_{F}(\phi-\phi(x);x)}{1-\operatorname{lct}_{F}(\phi-\phi(x);x)}&\text{if }\operatorname{lct}_{F}(\phi-\phi(x);x)<1.\end{cases}

By Theorem 1.2, by (1.3) and by a Thom–Sebastiani type result for lctF\operatorname{lct}_{F} (Proposition 3.11(1)), one can further show:

(1.8) 1lctF(ϕx;x)k(ϕ;x)1lctF(ϕx;x)+1.\left\lceil\frac{1}{\operatorname{lct}_{F}(\phi_{x};x)}\right\rceil\leq k_{\star}(\phi;x)\leq\left\lfloor\frac{1}{\operatorname{lct}_{F}(\phi_{x};x)}\right\rfloor+1.

We therefore see that that ϵ(ϕ;x),lctF(ϕx;x)\epsilon_{\star}(\phi;x),\operatorname{lct}_{F}(\phi_{x};x) and 1/k(ϕ;x)1/k_{\star}(\phi;x) are asymptotically equivalent as lctF(ϕx;x)0\operatorname{lct}_{F}(\phi_{x};x)\rightarrow 0. In §\mathsection3.2, we shall see that these invariants are also tightly related to an invariant δ\delta_{\star} quantifying Fourier decay. We will further see in §\mathsection3.4 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) ϵ11+ϵ1+ϵ21+ϵ2ϵ1+ϵc,1+ϵ1(F)c,1+ϵ2(F)c,1+ϵ(F).\frac{\epsilon_{1}}{1+\epsilon_{1}}+\frac{\epsilon_{2}}{1+\epsilon_{2}}\geq\frac{\epsilon}{1+\epsilon}\Longrightarrow\mathcal{M}_{c,1+\epsilon_{1}}(F)*\mathcal{M}_{c,1+\epsilon_{2}}(F)\subseteq\mathcal{M}_{c,1+\epsilon}(F).

Using the connection between ϵ(ϕ;x)\epsilon_{\star}(\phi;x) and k(ϕ;x)k_{\star}(\phi;x) as well as the structure of pushforward measures, we show the following converse to (1.9):

Theorem 1.3 (Reverse Young inequality).

Let ν1,ν2c,1(F)\nu_{1},\nu_{2}\in\mathcal{M}_{c,1}(F) be pushforward measures of the form νj=(ϕj)μj\nu_{j}=(\phi_{j})_{*}\mu_{j}, where μjc(Fnj)\mu_{j}\in\mathcal{M}_{c}^{\infty}(F^{n_{j}}) and ϕj:FnjF\phi_{j}:F^{n_{j}}\to F are analytic, locally dominant. If ν1ν2c,1+ϵ(F)\nu_{1}*\nu_{2}\in\mathcal{M}_{c,1+\epsilon}(F) for some ϵ>0\epsilon>0, then

(1.10) ϵ(ν1)1+ϵ(ν1)+ϵ(ν2)1+ϵ(ν2)>ϵ1+ϵ.\frac{\epsilon_{\star}(\nu_{1})}{1+\epsilon_{\star}(\nu_{1})}+\frac{\epsilon_{\star}(\nu_{2})}{1+\epsilon_{\star}(\nu_{2})}>\frac{\epsilon}{1+\epsilon}.

In particular, if ν1\nu_{1} is equal to ν2\nu_{2}, it lies in c,1+ϵ2+ϵ(F)\mathcal{M}_{c,1+\frac{\epsilon}{2+\epsilon}}(F).

Remark 1.4.

Under the assumptions of Theorem 1.3, if ν1ν2c,(F)\nu_{1}*\nu_{2}\in\mathcal{M}_{c,\infty}(F) then (1.10) (applied with ϵ+\epsilon\to+\infty) implies

(1.11) ϵ(ν1)1+ϵ(ν1)+ϵ(ν2)1+ϵ(ν2)1.\frac{\epsilon_{\star}(\nu_{1})}{1+\epsilon_{\star}(\nu_{1})}+\frac{\epsilon_{\star}(\nu_{2})}{1+\epsilon_{\star}(\nu_{2})}\geq 1.

In general, one cannot hope for a strict inequality in (1.11); indeed taking F=F=\mathbb{\mathbb{R}} or F=pF=\mathbb{Q}_{p} for p3mod4p\equiv 3\mod 4, ϕ1=ϕ2=x2\phi_{1}=\phi_{2}=x^{2} and μ1=μ2\mu_{1}=\mu_{2} a uniform measure on some ball around 0, one has ϵ(ν1)=ϵ(ν2)=1\epsilon_{\star}(\nu_{1})=\epsilon_{\star}(\nu_{2})=1, so that (1.11=1=1. On the other hand ν1ν2c,(F)\nu_{1}*\nu_{2}\in\mathcal{M}_{c,\infty}(F).

When F=F=\mathbb{C}, or more generally when the codimension of (ϕ1ϕ2)1(0)(\phi_{1}*\phi_{2})^{-1}(0) in Fn1+n2F^{n_{1}+n_{2}} is 11, we expect (1.11) to hold with a strict inequality.

1.2.3. Upper bounds on ϵ\epsilon_{\star}

For n>mn>m, the lower bound (1.6) may in general not be an equality. However, the next result shows that when F=F=\mathbb{C}, (1.6) is asymptotically sharp as lct(𝒥ϕ;x)0\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x)\rightarrow 0.

Theorem 1.5.

Let X,YX,Y be analytic \mathbb{C}-manifolds, and let ϕ:XY\phi:X\to Y be a locally dominant analytic map. Then, whenever lct(𝒥ϕ;x)<1\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x)<1,

(1.12) ϵ(ϕ;x)lct(𝒥ϕ;x)1lct(𝒥ϕ;x).\epsilon_{\star}(\phi;x)\leq\frac{\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x)}{1-\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x)}.
Remark 1.6.
  1. (1)

    When Y=FY=F, we have lctF(ϕϕ(x);x)lctF(𝒥ϕ;x)\operatorname{lct}_{F}(\phi-\phi(x);x)\leq\operatorname{lct}_{F}(\mathcal{J}_{\phi};x) (see Proposition 3.13), so that (1.12) holds also for FF\neq\mathbb{C}, whenever lctF(𝒥ϕ;x)<1\operatorname{lct}_{F}(\mathcal{J}_{\phi};x)<1. In particular, Theorem 1.2 implies Theorem 1.5 in the case that Y=Y=\mathbb{C}.

  2. (2)

    The upper bound (1.12) is asymptotically tight, in the sense that the value of ϵ\epsilon_{\star} can be arbitrarily close to the upper bound (1.12), as seen from the following family of examples. Let ϕ:=x1mx2mxnm\phi:=x_{1}^{m}x_{2}^{m}...x_{n}^{m}. Then ϕ=x1m1x2mxnm,,x1mx2mxnm1\nabla\phi=\langle x_{1}^{m-1}x_{2}^{m}...x_{n}^{m},...,x_{1}^{m}x_{2}^{m}...x_{n}^{m-1}\rangle, and thus by [How01, Main Theorem and Example 5], it follows that

    lct(𝒥ϕ;0)=1m1n,\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};0)=\frac{1}{m-\frac{1}{n}},

    so that the upper bound in (1.12) becomes ϵ(ϕ;0)1m11n\epsilon_{\star}(\phi;0)\leq\frac{1}{m-1-\frac{1}{n}}, whereas the lower bound (1.6) is ϵ(ϕ;0)1m1n\epsilon_{\star}(\phi;0)\geq\frac{1}{m-\frac{1}{n}}. We see that the true value ϵ(ϕ;0)=1m1\epsilon_{\star}(\phi;0)=\frac{1}{m-1} is closer to the upper bound than to the lower bound.

One may wonder whether Theorem 1.5 can be extended to FF\neq\mathbb{C}. 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 F=F=\mathbb{\mathbb{R}} and for any p\mathbb{Q}_{p}. If φ:XY\varphi:X\to Y is a polynomial map between smooth varieties, defined over \mathbb{Q}, we expect the upper bound in Theorem 1.5 to hold for φp:X(p)Y(p)\varphi_{\mathbb{Q}_{p}}:X(\mathbb{Q}_{p})\to Y(\mathbb{Q}_{p}) for infinitely many primes pp. 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 KK 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 φ:XG\varphi:X\rightarrow G and ψ:YG\psi:Y\rightarrow G be morphisms from algebraic KK-varieties X,YX,Y to an algebraic KK-group GG. We define their convolution by

φψ:X×YG, (x,y)φ(x)Gψ(y).\varphi*\psi:X\times Y\rightarrow G,\text{ }(x,y)\mapsto\varphi(x)\cdot_{G}\psi(y).

We denote by φk:XkG\varphi^{*k}:X^{k}\rightarrow G the kk-th self convolution of φ\varphi.

We restrict ourselves to the setting where X,YX,Y are smooth algebraic KK-varieties and GG is a connected algebraic KK-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 φ:XG\varphi:X\rightarrow G, the kk-th self convolution φk:XkG\varphi^{*k}:X^{k}\rightarrow G is flat for every kdimGk\geq\dim G ([GH21, Theorem B]). To explain the connection to this work, we introduce the following property:

Definition 1.8 ([AA16, Definition II]).
  1. (1)

    A KK-variety ZZ has rational singularities if it is normal and for every resolution of singularities π:Z~Z\pi:\widetilde{Z}\rightarrow Z, the pushforward π(𝒪Z~)\pi_{*}(\mathcal{O}_{\widetilde{Z}}) of the structure sheaf has no higher cohomologies.

  2. (2)

    A morphism φ:XY\varphi:X\rightarrow Y between smooth KK-varieties is called (FRS) if it is flat and if every fiber of φ\varphi has rational singularities.

In [AA16, Theorem 3.4] (see Theorem 6.1 below), Aizenbud and Avni proved the following. A morphism φ:XY\varphi:X\rightarrow Y between smooth KK-varieties, is (FRS) if and only if for every non-Archimedean local field FKF\supseteq K, one has (φF)μc,(Y(F))\left(\varphi_{F}\right)_{*}\mu\in\mathcal{M}_{c,\infty}(Y(F)) for every μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)). A similar characterization can be given for F=F=\mathbb{C}, see Corollary 6.2.

This characterization of LL^{\infty}-maps allows one to study random walks on analytic groups as in §\mathsection1.1 in an algebro-geometric way, via the above algebraic convolution operation. Starting from a pushforward ν=(φF)μ\nu=\left(\varphi_{F}\right)_{*}\mu of μc,(X(F))\mu\in\mathcal{M}_{c,\infty}(X(F)) by an algebraic map φ:XG\varphi:X\to G, instead of showing that νkc,(G(F))\nu^{*k}\in\mathcal{M}_{c,\infty}(G(F)), it is enough to show that φk:XkG\varphi^{*k}:X^{k}\rightarrow G 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 φ:XG\varphi:X\rightarrow G becomes (FRS) after sufficiently many self-convolutions.

Using (1.3), Theorem 1.1 and Corollary 6.2, explicit bounds can be given on the required number of self convolutions, in terms of the Jacobian ideal of φ\varphi.

Corollary 1.9.

Let XX be a smooth KK-variety, GG be a connected KK-algebraic group and let φ:XG\varphi:X\to G be a locally dominant morphism. Then φk\varphi^{*k} is (FRS) for any k1+lct(𝒥φ)lct(𝒥φ)+1k\geq\left\lfloor\frac{1+\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\varphi})}{\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\varphi})}\right\rfloor+1.

Remark 1.10.

In the setting of locally compact groups, Young’s convolution inequality is commonly stated under the assumption that the group GG is unimodular. In [KR78, Lemma 2.1 and Corollary 2.3] a version for non-unimodular groups is given; if GG is a locally compact group, with modular character :G>0\triangle:G\rightarrow\mathbb{\mathbb{R}}_{>0}, and if 1p,q,r1\leq p,q,r\leq\infty satisfy 1p+1q=1+1r\frac{1}{p}+\frac{1}{q}=1+\frac{1}{r}, then we have μ(11pν)rμpνq\left\|\mu*(\triangle^{1-\frac{1}{p}}\nu^{\prime})\right\|_{r}\leq\left\|\mu\right\|_{p}\left\|\nu^{\prime}\right\|_{q} whenever μc,p(G)\mu\in\mathcal{M}_{c,p}(G) and νc,q(G)\nu^{\prime}\in\mathcal{M}_{c,q}(G). However, since the modular character :G>0\triangle:G\rightarrow\mathbb{\mathbb{R}}_{>0} is a continuous homomorphism, it bounded on the compact support of ν\nu^{\prime}. Hence, for every μc,p(G)\mu\in\mathcal{M}_{c,p}(G) and νc,q(G)\nu\in\mathcal{M}_{c,q}(G), we deduce that

μνr=μ(11p(1p1ν))rμp1p1νqCμpνq,\left\|\mu*\nu\right\|_{r}=\left\|\mu*\left(\triangle^{1-\frac{1}{p}}(\triangle^{\frac{1}{p}-1}\cdot\nu)\right)\right\|_{r}\leq\left\|\mu\right\|_{p}\cdot\left\|\triangle^{\frac{1}{p}-1}\cdot\nu\right\|_{q}\leq C\cdot\left\|\mu\right\|_{p}\cdot\left\|\nu\right\|_{q},

for some constant CC depending on GG and ν\nu. In particular, μνc,r(G)\mu*\nu\in\mathcal{M}_{c,r}(G).

1.2.5. An algebraic characterization of ϵ=\epsilon_{\star}=\infty

Let φ:XY\varphi:X\rightarrow Y be a morphism between smooth KK-varieties. We would like to characterize the condition that ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty for all FF in certain families of local fields, in terms of the singularities of φ\varphi. The singularity properties we consider play a central role in birational geometry (see [Kol97]).

Let XX be a normal KK-variety, and let ωΩtop(Xsm)\omega\in\Omega^{\mathrm{top}}(X_{\mathrm{sm}}) be a rational top form on the smooth locus XsmX_{\mathrm{sm}} of XX. The zeros and poles of ω\omega give rise to a divisor div(ω)\operatorname{div}(\omega) on XX. Let π:X~X\pi:\widetilde{X}\rightarrow X be a resolution of singularities, namely, a proper morphism from a smooth variety X~\widetilde{X}, which is an isomorphism over XsmX_{\mathrm{sm}}. Then πω\pi^{*}\omega defines a unique rational top form on X~\widetilde{X}. Moreover, when XX is nice enough (e.g. if XX is a local complete intersection), div(ω)\operatorname{div}(\omega) is \mathbb{Q}-Cartier, and we can define its pullback πdiv(ω)\pi^{*}\operatorname{div}(\omega). The divisor KX~/X:=div(πω)πdiv(ω)K_{\widetilde{X}/X}:=\operatorname{div}(\pi^{*}\omega)-\pi^{*}\operatorname{div}(\omega) on X~\widetilde{X} is called the relative canonical divisor, and one can verify that it does not depend on the choice of ω\omega. KX~/XK_{\widetilde{X}/X} can be written as KX~/X=i=1MaiEiK_{\widetilde{X}/X}=\sum_{i=1}^{M}a_{i}E_{i}, for some prime divisors EiE_{i}, aia_{i}\in\mathbb{Q}. We say that XX has canonical singularities (resp. log-canonical singularities), if ai0a_{i}\geq 0 (resp. ai1a_{i}\geq-1) for all 1iM1\leq i\leq M. When XX 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 X𝔸nX\subseteq\mathbb{A}_{\mathbb{C}}^{n} be the variety defined by i=1nxidi=0\sum_{i=1}^{n}x_{i}^{d_{i}}=0, with n3n\geq 3. Then XX has canonical singularities if and only if i=1n1di>1\sum_{i=1}^{n}\frac{1}{d_{i}}>1, and log-canonical singularities if and only if i=1n1di1\sum_{i=1}^{n}\frac{1}{d_{i}}\geq 1.

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 ϵ(φp)=\epsilon_{\star}(\varphi_{\mathbb{Q}_{p}})=\infty, that is to being almost in LL^{\infty}. Unfortunately, the normality hypothesis required for log-canonical singularities turns out to be too strong. For example, the map φ(x,y)=xy\varphi(x,y)=xy satisfies ϵ(φp)=\epsilon_{\star}(\varphi_{\mathbb{Q}_{p}})=\infty for all pp, but the fiber over 0 is not normal (see §\mathsection6, 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 {xy=0}\left\{xy=0\right\} is demi-normal and has semi-log-canonical singularities. We can now state the main result of this section:

Theorem 1.12.

Let φ:XY\varphi:X\rightarrow Y be a map between smooth KK-varieties. Then the following are equivalent:

  1. (1)

    φ\varphi is flat with fibers of semi-log-canonical singularities.

  2. (2)

    For every local field FF containing KK, we have ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty, that is, for every μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)), the measure φμ\varphi_{*}\mu lies in c,q(Y(F))\mathcal{M}_{c,q}(Y(F)) for all 1<q<1<q<\infty.

  3. (3)

    For every large enough prime pp, such that pK\mathbb{Q}_{p}\supseteq K, we have ϵ(φp)=\epsilon_{\star}(\varphi_{\mathbb{Q}_{p}})=\infty.

  4. (4)

    We have ϵ(φ)=\epsilon_{\star}(\varphi_{\mathbb{C}})=\infty.

We prove Theorem 1.12 by showing the implications (1)(2)(3)(1)(1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1) and (1)(2)(4)(1)(1)\Rightarrow(2)\Rightarrow(4)\Rightarrow(1). The implications (2)(3)(2)\Rightarrow(3) and (2)(4)(2)\Rightarrow(4) are immediate. In the proof of (3)(1)(3)\Rightarrow(1) and (4)(1)(4)\Rightarrow(1), we reduce to the case Y=𝔸nY=\mathbb{A}^{n}, and show that φ:XY\varphi:X\rightarrow Y satisfies that φψ\varphi*\psi is (FRS) for every dominant map ψ:YY\psi:Y^{\prime}\rightarrow Y. By analyzing the jets of φ\varphi, 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.

In the proof of (1)(2)(1)\Rightarrow(2), we reduce to showing Item (2) for constructible measures, and utilize their structure theory, namely, we use [CM11, CM12] for the Archimedean case, and [CGH18] for the non-Archimedean case.

1.3. Future directions and further applications

1.3.1. ϵ\epsilon_{\star} as an invariant of singularities

Similarly to the analytic definition of ϵ(ϕ)\epsilon_{\star}(\phi) in (1.2), one can also define an algebro-geometric invariant. Let Loc0\operatorname{Loc}_{0} be the collection of all non-Archimedean local fields FF of characteristic 0.

Definition 1.13.

Let φ:XY\varphi:X\rightarrow Y be a morphism between smooth \mathbb{Q}-varieties. We define ϵ(φ):=minFLoc0ϵ(φF)\epsilon_{\star}(\varphi):=\min\limits_{F\in\operatorname{Loc}_{0}}\epsilon_{\star}(\varphi_{F}), and ϵ(φ;x)=supUxϵ(φ|U)\epsilon_{\star}(\varphi;x)=\sup\limits_{U\ni x}\,\epsilon_{\star}(\varphi|_{U}), where UU varies over all Zariski open neighborhoods of xX(¯)x\in X(\overline{\mathbb{Q}}).

It is a consequence of Theorem 1.1 that ϵ(φ)>0\epsilon_{\star}(\varphi)>0 for any φ\varphi, and it essentially follows from [GH21] that ϵ(φ)\epsilon_{\star}(\varphi)\in\mathbb{Q}. We further expect ϵ(φ;x)\epsilon_{\star}(\varphi;x) to have a purely algebro-geometric formula, and to have a good behavior in families, which means the following. Suppose that φ:X~Y~\varphi:\widetilde{X}\rightarrow\widetilde{Y} is a morphism over 𝔸1\mathbb{A}_{\mathbb{C}}^{1}, where π1:X~𝔸1\pi_{1}:\widetilde{X}\rightarrow\mathbb{A}_{\mathbb{C}}^{1} and π2:Y~𝔸1\pi_{2}:\widetilde{Y}\rightarrow\mathbb{A}_{\mathbb{C}}^{1} are smooth morphisms. This gives a family {φt:X~tY~t}t𝔸1\left\{\varphi_{t}:\widetilde{X}_{t}\rightarrow\widetilde{Y}_{t}\right\}_{t\in\mathbb{A}_{\mathbb{C}}^{1}} of morphisms between smooth varieties. It follows from [Var83], that the function xlct(φπ1(x);x)x\mapsto\operatorname{lct}(\varphi_{\pi_{1}(x)};x) is lower semicontinuous. By Theorem 1.2, xϵ(φπ1(x);x)x\mapsto\epsilon_{\star}(\varphi_{\pi_{1}(x)};x) is lower semicontinuous as well, if π2:Y~𝔸1\pi_{2}:\widetilde{Y}\rightarrow\mathbb{A}^{1} has fibers of dimension 11. We expect xϵ(φπ1(x);x)x\mapsto\epsilon_{\star}(\varphi_{\pi_{1}(x)};x) to be lower semicontinuous in general. We further expect the following question to have a positive answer.

Question 1.14.

Let φ:XY\varphi:X\rightarrow Y be as in Definition 1.13. Is it true that ϵ(φ)=ϵ(φ)\epsilon_{\star}(\varphi_{\mathbb{C}})=\epsilon_{\star}(\varphi)?

Given a \mathbb{Q}-morphism φ:XY\varphi:X\rightarrow Y as above, one may further wonder whether the quantity ϵ(φ;x)=supUxinfμc,(U)ϵ((φ)μ)\epsilon_{\star}(\varphi_{\mathbb{C}};x)=\sup_{U\ni x}\inf_{\mu\in\mathcal{M}_{c,\infty}(U)}\epsilon_{\star}((\varphi_{\mathbb{C}})_{*}\mu) in (1.1) stays the same when the supremum is taken over all Zariski open neighborhoods UU of xx instead of analytic ones.

1.3.2. ϵ\epsilon_{\star} as an invariant of words

As discussed in §\mathsection1.1, 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 w𝔤:𝔤r𝔤w_{\mathfrak{g}}:\mathfrak{g}^{r}\rightarrow\mathfrak{g}, where 𝔤\mathfrak{g} is a simple Lie algebra, become (FRS) after deg(w)4\sim\operatorname{deg}(w)^{4} self convolutions, where deg(w)\operatorname{deg}(w) is the degree of ww.

Question 1.15.

Can we find α,C>0\alpha,C>0 such that for any wFrw\in F_{r} of length (w)\ell(w), and every simple algebraic group GG, the word map wGC(w)αw_{G}^{*C\ell(w)^{\alpha}} is (FRS)?

A potential way to tackle Question 1.15 is by studying ϵ(wG)\epsilon_{\star}(w_{G}) in the sense of Definition 1.13; For each wFrw\in F_{r} (resp. wrw\in\mathcal{L}_{r}), we define ϵ(w):=infGϵ(wG)\epsilon_{\star}(w):=\inf_{G}\,\epsilon_{\star}(w_{G}) (resp. ϵ(w):=inf𝔤ϵ(w𝔤)\epsilon_{\star}(w):=\inf_{\mathfrak{g}}\,\epsilon_{\star}(w_{\mathfrak{g}})), where GG runs over all simple, simply connected algebraic groups, and 𝔤=Lie(G)\mathfrak{g}=\operatorname{Lie}(G). We can now ask the following:

Question 1.16.

Can we find for each ll\in\mathbb{N}, a constant ϵ(l)>0\epsilon(l)>0 such that:

  1. (1)

    For every wFrw\in F_{r} of length (w)=l\ell(w)=l, we have ϵ(w)ϵ(l)\epsilon_{\star}(w)\geq\epsilon(l)?

  2. (2)

    For every wrw\in\mathcal{L}_{r} of degree deg(w)=l\operatorname{deg}(w)=l, we have ϵ(w)ϵ(l)\epsilon_{\star}(w)\geq\epsilon(l)?

1.3.3. ϵ\epsilon_{\star} 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 GG over local fields. Harish-Chandra’s regularity theorem says that every character Θ(π)\Theta(\pi) of an irreducible representation π\pi of GG is given by a locally L1L^{1}-function. Since characters are of motivic nature, a variant of [GH21, Theorem F] suggests that they should in fact be locally in L1+ϵL^{1+\epsilon}, for some ϵ>0\epsilon>0. This gives rise to a new invariant ϵ(π)\epsilon_{\star}(\pi), 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 ϵ(π)\epsilon_{\star}(\pi) in terms of the nilpotent orbits appearing the local character expansion of π\pi (see [GGH, Theorems A and D]).

1.4. Conventions

  1. (1)

    By \mathbb{N} we mean the set {0,1,2,}\{0,1,2,...\}.

  2. (2)

    We use KK (resp. FF) to denote number (resp. local) fields and 𝒪K\mathcal{O}_{K} (resp. 𝒪F\mathcal{O}_{F}) for their rings of integers.

  3. (3)

    Given an algebraic map φ:XY\varphi:X\rightarrow Y between smooth KK-varieties, and a local field FF, we denote by φF:X(F)Y(F)\varphi_{F}:X(F)\rightarrow Y(F) the corresponding FF-analytic map. We sometimes write φ\varphi instead of φF\varphi_{F} if the local field FF is clear from the setting.

  4. (4)

    If f,g:Xf,g:X\to\mathbb{R} are two functions, we write fgf\lesssim g if there exists a number C>0C>0, possibly depending on XX, ff and gg, such that fCgf\leq Cg. We write fgf\sim g if gfgg\lesssim f\lesssim g.

  5. (5)

    We write dxϕd_{x}\phi for the differential of an analytic map ϕ:XY\phi:X\rightarrow Y at xXx\in X, and we denote by Jacxϕ:=det(dxϕ)\operatorname{Jac}_{x}\phi:=\det(d_{x}\phi) the Jacobian of ϕ\phi if dimX=dimY\dim X=\dim Y. The generalization of the Jacobian to the case of unequal dimensions is defined in §\mathsection1.2.

  6. (6)

    Throughout the paper, we write ||F\left|\,\cdot\,\right|_{F} for the absolute value on FF, normalized so that μF(aA)=|a|FμF(A)\mu_{F}(aA)=\left|a\right|_{F}\cdot\mu_{F}(A), for all aF×a\in F^{\times}, AFA\subseteq F, and where μF\mu_{F} is a Haar measure on FF. Note that ||\left|\,\cdot\,\right|_{\mathbb{C}} is the square of the usual absolute value on \mathbb{C}.

  7. (7)

    We write HnH_{n} for the nn-dimensional Hausdorff measure. We recall that given a metric space XX and a subset ZXZ\subseteq X, we define Hn(Z):=supδ0Hnδ(Z)H_{n}(Z):=\underset{\delta\rightarrow 0}{\sup}H_{n}^{\delta}(Z), where

    Hnδ(Z)=inf{i=1α(n)(diam(Ui))n:i=1UiZ,diam(Ui)<δ},H_{n}^{\delta}(Z)=\inf\left\{\sum_{i=1}^{\infty}\alpha(n)\left(\mathrm{diam}(U_{i})\right)^{n}:\bigcup_{i=1}^{\infty}U_{i}\supseteq Z,\mathrm{diam}(U_{i})<\delta\right\},

    where diam(A)\mathrm{diam}(A) denotes the diameter of a set AXA\subseteq X and α(n):=(π/4)n/2Γ(n2+1)\alpha(n):=\frac{(\pi/4)^{n/2}}{\Gamma(\frac{n}{2}+1)}. The normalization constant α(n)\alpha(n) is chosen such that HnH_{n} coincides with the Lebesgue measure in the case of X=nX=\mathbb{\mathbb{R}}^{n}.

  8. (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 FF 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 π:X~U\pi:\widetilde{X}\rightarrow U below is called a log-principalization (or uniformization) of JJ.

Theorem 2.1 (See [VZnG08, Theorem 2.3], [DvdD88, Theorem 2.2], [BM89] and [Wlo09]).

Let UFnU\subseteq F^{n} be an open subset, and let f1,,fr:UFf_{1},...,f_{r}:U\rightarrow F be FF-analytic maps, generating a non-zero ideal JJ in the algebra of FF-analytic functions on UU. Then there exist an FF-analytic manifold X~\widetilde{X}, a proper FF-analytic map π:X~U\pi:\widetilde{X}\rightarrow U and a collection of closed submanifolds {Ei}iT\left\{E_{i}\right\}_{i\in T} of X~\widetilde{X} of codimension 11, equipped with pairs of non-negative integers {(ai,bi)}iT\left\{(a_{i},b_{i})\right\}_{i\in T}, such that the following hold:

  1. (1)

    π\pi 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 V(J)\mathrm{V}(J) of JJ in UU.

  2. (2)

    For every cX~c\in\widetilde{X}, there are local coordinates (x~1,,x~n)(\widetilde{x}_{1},...,\widetilde{x}_{n}) in a neighborhood VcV\ni c, such that each EiE_{i} containing cc is given by the equation x~i=0\widetilde{x}_{i}=0. Moreover, if without loss of generality E1,,EmE_{1},...,E_{m} contain cc, then there exists an FF-analytic unit v:VFv:V\rightarrow F, such that the pullback of JJ is the principal ideal

    (2.1) πJ=x~1a1x~mam\pi^{*}J=\langle\widetilde{x}_{1}^{a_{1}}\cdots\widetilde{x}_{m}^{a_{m}}\rangle

    and such that the Jacobian of π\pi (i.e. detdx~π\det\,d_{\widetilde{x}}\pi) is given by:

    (2.2) Jacx~(π)=v(x~)x~1b1x~mbm.\operatorname{Jac}_{\widetilde{x}}(\pi)=v(\widetilde{x})\cdot\widetilde{x}_{1}^{b_{1}}\cdots\widetilde{x}_{m}^{b_{m}}.
Remark 2.2.
  1. (1)

    Condition (2.1) means that for each fi:UFf_{i}:U\rightarrow F, one can write fiπ(x~)=ui(x~)x~1a1x~mamf_{i}\circ\pi(\widetilde{x})=u_{i}(\widetilde{x})\widetilde{x}_{1}^{a_{1}}\cdots\widetilde{x}_{m}^{a_{m}} for some analytic functions uiu_{i}, and uj(0)0u_{j}(0)\neq 0 for at least one j{1,,r}j\in\{1,...,r\}. Note that the a1,,ama_{1},...,a_{m} are the same for all the fif_{i}’s.

  2. (2)

    If r=1r=1, so that J=fJ=\langle f\rangle, we may further assume that fπ(x~1,,x~n)=Cx~1a1x~mamf\circ\pi(\widetilde{x}_{1},...,\widetilde{x}_{n})=C\cdot\widetilde{x}_{1}^{a_{1}}\cdots\widetilde{x}_{m}^{a_{m}} locally on each chart, for some constant C0C\neq 0. Indeed, u(0)0u(0)\neq 0. If u(0)u(0) is an a1a_{1}-th power in FF then the same holds for u(x~)u(\widetilde{x}) in a small neighborhood of 0. In this case, we may apply the change of coordinates

    (x~1,,x~n)(u(x~)1a1x~1,,x~n).(\widetilde{x}_{1},...,\widetilde{x}_{n})\mapsto(u(\widetilde{x})^{-\frac{1}{a_{1}}}\widetilde{x}_{1},...,\widetilde{x}_{n}).

    If u(0)u(0) is not an a1a_{1}-th power we may multiply it by bF×b\in F^{\times} such that u(x~)bu(\widetilde{x})b is an a1a_{1}-th power, and apply a similar change of coordinates.

The next lemma follows directly by changing coordinates using a log-principalization of JJ and computing the integral min1il[|fi(x)|s]𝑑μ(s)\int\min\limits_{1\leq i\leq l}\big{[}\left|f_{i}(x)\right|^{-s}\big{]}d\mu(s) with respect to the new coordinates.

Lemma 2.3 (See e.g. [Mus12, Theorem 1.2] and [VZnG08, Theorem 2.7]).

Let J=f1,,frJ=\langle f_{1},...,f_{r}\rangle be an ideal of FF-analytic functions on UFnU\subseteq F^{n}. Let π:X~U\pi:\widetilde{X}\rightarrow U be a log-principalization of JJ, with data {Ei}iT\left\{E_{i}\right\}_{i\in T} and {(ai,bi)}iT\left\{(a_{i},b_{i})\right\}_{i\in T} as in Theorem 2.1. Then the FF-log-canonical threshold of JJ at xUx\in U is equal to:

(2.3) lctF(J;x)=mini:xπ(Ei)bi+1ai.\operatorname{lct}_{F}(J;x)=\min_{i:\,x\in\pi(E_{i})}\frac{b_{i}+1}{a_{i}}.
Remark 2.4.

Note that while the data {Ei}iT\left\{E_{i}\right\}_{i\in T} and {(ai,bi)}iT\left\{(a_{i},b_{i})\right\}_{i\in T} depends on the log-principalization π\pi, Definition 1.5 is intrinsic, thus the right-hand side of (2.3) does not depend on π\pi.

3. The one-dimensional case

3.1. A formula for ϵ\epsilon_{\star}

In this section we provide a formula for ϵ\epsilon_{\star} in the one-dimensional case (Theorem 1.2). The formula will be phrased in terms of the FF-log-canonical threshold, where FF is any local field of characteristic zero.

When FF is non-Archimedean, we denote by 𝒪F\mathcal{O}_{F} its ring of integers, by kFk_{F} its residue field, and by qFq_{F} the number of elements in kFk_{F}. Write ϖF𝒪F\varpi_{F}\in\mathcal{O}_{F} for a fixed uniformizer (i.e., a generator of the maximal ideal of the ring of integers) of FF, and let ||F\left|\,\cdot\,\right|_{F} be as in §\mathsection1.4, so that |ϖF|F=qF1\left|\varpi_{F}\right|_{F}=q_{F}^{-1}. We write μF\mu_{F} for the Haar measure on FF, normalized such that μF(𝒪F)=1\mu_{F}(\mathcal{O}_{F})=1 when FF is non-Archimedean, and such that μF\mu_{F} is the Lebesgue measure when FF is Archimedean. We write μ𝒪F:=μF|𝒪F\mu_{\mathcal{O}_{F}}:=\mu_{F}|_{\mathcal{O}_{F}}. We write dxdx instead of μF\mu_{F} when we integrate a function g(x)g(x) with respect to μF\mu_{F}. We denote by (a1,,an)F:=maxi|ai|F\left\|(a_{1},...,a_{n})\right\|_{F}:=\max_{i}\left|a_{i}\right|_{F} the maximum norm on FnF^{n}.

For an analytic map ϕ:XF\phi:X\rightarrow F, and xXx\in X, we set ϕx(z):=ϕ(z)ϕ(x)\phi_{x}(z):=\phi(z)-\phi(x).

To prove Theorem 1.2, we reduce to the monomial case using Hironaka’s resolution of singularities (§\mathsection2), 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 XX be an analytic FF-manifold, and let ϕ:XF\phi:X\to F be a locally dominant analytic map. Then for every xXx\in X with lctF(ϕx;x)<1\operatorname{lct}_{F}(\phi_{x};x)<1, one has:

ϵ(ϕ;x)lctF(ϕx;x)1lctF(ϕx;x).\epsilon_{\star}(\phi;x)\leq\frac{\operatorname{lct}_{F}(\phi_{x};x)}{1-\operatorname{lct}_{F}(\phi_{x};x)}.
Proof.

We need to show

111+ϵ(ϕ;x)lctF(ϕx;x).1-\frac{1}{1+\epsilon_{\star}(\phi;x)}\leq\operatorname{lct}_{F}(\phi_{x};x).

Let ϵ<ϵ(ϕ;x)\epsilon<\epsilon_{\star}(\phi;x). Then there exists a neighborhood UU of xx such that ϕσL1+ϵ\phi_{*}\sigma\in L^{1+\epsilon} for every σc,(U)\sigma\in\mathcal{M}_{c,\infty}(U). Write ϕσ=g(t)μF\phi_{*}\sigma=g(t)\mu_{F}. Let B(a,δ):={t:|ta|Fδ}B(a,\delta):=\left\{t:\left|t-a\right|_{F}\leq\delta\right\}, and note that μF(B(a,δ))δ\mu_{F}(B(a,\delta))\sim\delta. By Jensen’s inequality we have:

(ϕσ)(B(a,δ))=1δB(a,δ)δg(t)𝑑t(1δB(a,δ)(δg(t))1+ϵ𝑑t)11+ϵδ111+ϵ,(\phi_{*}\sigma)(B(a,\delta))=\frac{1}{\delta}\int_{B(a,\delta)}\delta g(t)dt\leq\left(\frac{1}{\delta}\int_{B(a,\delta)}\left(\delta g(t)\right)^{1+\epsilon}dt\right)^{\frac{1}{1+\epsilon}}\lesssim\delta^{1-\frac{1}{1+\epsilon}},

i.e. we have the distributional estimate σ({z:|ϕ(z)ϕ(x)|Fδ})δ111+ϵ\sigma\left(\left\{z:\left|\phi(z)-\phi(x)\right|_{F}\leq\delta\right\}\right)\lesssim\delta^{1-\frac{1}{1+\epsilon}}. Using Fubini’s theorem, we obtain:

U|ϕx(z)|Fs𝑑σ\displaystyle\int_{U}\left|\phi_{x}(z)\right|_{F}^{-s}d\sigma =U(01{(t,z):|ϕx(z)|Fst}(t,z)𝑑t)𝑑σ\displaystyle=\int_{U}\left(\int_{0}^{\infty}1_{\left\{(t,z)\,:\left|\phi_{x}(z)\right|_{F}^{-s}\geq t\right\}}(t,z)dt\right)d\sigma
=0σ{z:|ϕ(z)ϕ(x)|Fst}𝑑t\displaystyle=\int_{0}^{\infty}\sigma\left\{z:\left|\phi(z)-\phi(x)\right|_{F}^{-s}\geq t\right\}dt
=0σ{z:|ϕ(z)ϕ(x)|Ft1/s}𝑑t\displaystyle=\int_{0}^{\infty}\sigma\left\{z:\left|\phi(z)-\phi(x)\right|_{F}\leq t^{-1/s}\right\}dt
1+1t1s(111+ϵ)𝑑t<,\displaystyle\lesssim 1+\int_{1}^{\infty}t^{-\frac{1}{s}\left(1-\frac{1}{1+\epsilon}\right)}dt<\infty,

whenever s<111+ϵs<1-\frac{1}{1+\epsilon}. This implies that lctF(ϕx;x)111+ϵ(ϕ;x)\operatorname{lct}_{F}(\phi_{x};x)\geq 1-\frac{1}{1+\epsilon_{\star}(\phi;x)}. ∎

We now return to the main narrative. Specializing to the setting of Hironaka’s theorem, we consider the monomial case.

Lemma 3.2.

Let f:FnFf:F^{n}\rightarrow F be a monomial map f(x1,,xn)=x1a1xnanf(x_{1},...,x_{n})=x_{1}^{a_{1}}\cdot...\cdot x_{n}^{a_{n}}, let g:Fn0g:F^{n}\rightarrow\mathbb{\mathbb{R}}_{\geq 0} be a continuous function and let μ=g(x)|x1|Fb1|xn|FbnμFn\mu=g(x)\left|x_{1}\right|_{F}^{b_{1}}\cdot...\cdot\left|x_{n}\right|_{F}^{b_{n}}\mu_{F}^{n}, for a1,,an1a_{1},...,a_{n}\in\mathbb{Z}_{\geq 1} and b1,,bn0b_{1},...,b_{n}\in\mathbb{Z}_{\geq 0}. Then:

ϵ(fμ){=if minibi+1ai1,minibi+1ai1minibi+1aiotherwise.\epsilon_{\star}(f_{*}\mu)\,\begin{cases}=\infty&\text{if\, }\min_{i}\frac{b_{i}+1}{a_{i}}\geq 1,\\ \geq\frac{\min_{i}\frac{b_{i}+1}{a_{i}}}{1-\min_{i}\frac{b_{i}+1}{a_{i}}}&\text{otherwise}.\end{cases}

Furthermore, if g(0)0g(0)\neq 0, the second bound is in fact an equality.

Proof.

Without loss of generality we may assume minibi+1ai=bn+1an\min_{i}\frac{b_{i}+1}{a_{i}}=\frac{b_{n}+1}{a_{n}}. We first consider the special case when bn+1an\frac{b_{n}+1}{a_{n}} is the unique minimum. We write ff as a composition f=prψf=\operatorname{pr}\circ\psi, where ψ:FnFn\psi:F^{n}\rightarrow F^{n} is given by ψ(x1,,xn)=(x1,,xn1,x1a1xnan)\psi(x_{1},...,x_{n})=(x_{1},...,x_{n-1},x_{1}^{a_{1}}\cdot...\cdot x_{n}^{a_{n}}) and pr\operatorname{pr} is the projection to the last coordinate. Write x=(x¯,xn)x=(\overline{x},x_{n}), where x¯=(x1,,xn1)\overline{x}=(x_{1},...,x_{n-1}), and similarly y=(y¯,yn)y=(\overline{y},y_{n}). Note that Jac(x1,,xn)(ψ)=anx1a1xn1an1xnan1\operatorname{Jac}_{(x_{1},...,x_{n})}(\psi)=a_{n}x_{1}^{a_{1}}\cdot...\cdot x_{n-1}^{a_{n-1}}x_{n}^{a_{n}-1} and that if ψ(x¯,xn)=(y¯,yn)\psi(\overline{x},x_{n})=(\overline{y},y_{n}), then x¯=y¯\overline{x}=\overline{y} and xnan=yny1a1yn1an1x_{n}^{a_{n}}=y_{n}y_{1}^{-a_{1}}\cdot...\cdot y_{n-1}^{-a_{n-1}}. Hence the Radon–Nikodym density of ψμ\psi_{*}\mu is equal to

d(ψμ)d(μFn)(y¯,yn)\displaystyle\frac{d(\psi_{*}\mu)}{d(\mu_{F}^{n})}(\overline{y},y_{n}) =t:ψ(y¯,t)=y(j=1n1|yj|Fbj)|t|Fbng(y¯,t)|Jac(y1,,yn1,t)(ψ)|F\displaystyle=\sum_{t:\psi(\overline{y},t)=y}\frac{\left(\prod_{j=1}^{n-1}\left|y_{j}\right|_{F}^{b_{j}}\right)\left|t\right|_{F}^{b_{n}}g(\overline{y},t)}{\left|\operatorname{Jac}_{(y_{1},...,y_{n-1},t)}(\psi)\right|_{F}}
=t:ψ(y¯,t)=yg(y¯,t)|an|F1(j=1n1|yj|Fbjaj)|t|Fbnan+1\displaystyle=\sum_{t:\psi(\overline{y},t)=y}g(\overline{y},t)\left|a_{n}\right|_{F}^{-1}\cdot\left(\prod_{j=1}^{n-1}\left|y_{j}\right|_{F}^{b_{j}-a_{j}}\right)\left|t\right|_{F}^{b_{n}-a_{n}+1}
=(t:ψ(y¯,t)=yg(y¯,t))|an|F1(j=1n1|yj|Fbjaj(bnan+1)ajan)|yn|Fbnan+1an.\displaystyle=\left(\sum_{t:\psi(\overline{y},t)=y}g(\overline{y},t)\right)\cdot\left|a_{n}\right|_{F}^{-1}\left(\prod_{j=1}^{n-1}\left|y_{j}\right|_{F}^{b_{j}-a_{j}-\frac{(b_{n}-a_{n}+1)a_{j}}{a_{n}}}\right)\left|y_{n}\right|_{F}^{\frac{b_{n}-a_{n}+1}{a_{n}}}.

Since bn+1an<bj+1aj\frac{b_{n}+1}{a_{n}}<\frac{b_{j}+1}{a_{j}} we get for every 1jn11\leq j\leq n-1,

an(bjaj+1)>(bnan+1)aj,a_{n}(b_{j}-a_{j}+1)>(b_{n}-a_{n}+1)a_{j},

and hence:

bjaj(bnan+1)ajan>1.b_{j}-a_{j}-\frac{(b_{n}-a_{n}+1)a_{j}}{a_{n}}>-1.

In particular, integrating over the first n1n-1 coordinates, we get that d(fμ)dμF(s)|s|Fbnan+1an\frac{d(f_{*}\mu)}{d\mu_{F}}(s)\lesssim\left|s\right|_{F}^{\frac{b_{n}-a_{n}+1}{a_{n}}}. If bn+1an1\frac{b_{n}+1}{a_{n}}\geq 1 then ϵ(fμ)=\epsilon_{\star}(f_{*}\mu)=\infty as required. Thus we may assume that bn+1an<1\frac{b_{n}+1}{a_{n}}<1. Then fμc,1+ϵ(F)f_{*}\mu\in\mathcal{M}_{c,1+\epsilon}(F), whenever bnan+1an(1+ϵ)>1\frac{b_{n}-a_{n}+1}{a_{n}}(1+\epsilon)>-1, i.e. whenever

ϵ<bn+1anbn1=bn+1an1bn+1an=minibi+1ai1minibi+1ai.\epsilon<\frac{b_{n}+1}{a_{n}-b_{n}-1}=\frac{\frac{b_{n}+1}{a_{n}}}{1-\frac{b_{n}+1}{a_{n}}}=\frac{\min_{i}\frac{b_{i}+1}{a_{i}}}{1-\min_{i}\frac{b_{i}+1}{a_{i}}}.

If g(0)0,g(0)\neq 0, we also have d(fμ)dμF(s)|s|Fbnan+1an\frac{d(f_{*}\mu)}{d\mu_{F}}(s)\gtrsim\left|s\right|_{F}^{\frac{b_{n}-a_{n}+1}{a_{n}}}, whence the inequality in the statement of the lemma is in fact an equality.

It is left to deal with the case when minibi+1ai\min_{i}\frac{b_{i}+1}{a_{i}} is not uniquely achieved. For the lower bound, take

μ~=|x1|Fb1|xn1|Fbn1|xn|Fbnδg(x)μFn,\widetilde{\mu}=\left|x_{1}\right|_{F}^{b_{1}}...\left|x_{n-1}\right|_{F}^{b_{n-1}}\left|x_{n}\right|_{F}^{b_{n}-\delta}g(x)\mu_{F}^{n},

for an arbitrarily small δ>0\delta>0. Since μ~μ\widetilde{\mu}\geq\mu inside a small neighborhood of 0, we deduce that fμc,1+ϵ(F)f_{*}\mu\in\mathcal{M}_{c,1+\epsilon}(F) for ϵ<bnδ+1an1bnδ+1an\epsilon<\frac{\frac{b_{n}-\delta+1}{a_{n}}}{1-\frac{b_{n}-\delta+1}{a_{n}}}. Since δ\delta can be taken arbitrarily small we are done. Similarly, for the upper bound we take

μ~=|x1|Fb1+δ|xn1|Fbn1+δ|xn|Fbng(x)μFnμ,\widetilde{\mu}=\left|x_{1}\right|_{F}^{b_{1}+\delta}\cdot...\cdot\left|x_{n-1}\right|_{F}^{b_{n-1}+\delta}\left|x_{n}\right|_{F}^{b_{n}}g(x)\mu_{F}^{n}\leq\mu,

with g(0)0g(0)\neq 0 and deduce that ϵ(fμ)bn+1an1bn+1an\epsilon_{\star}(f_{*}\mu)\leq\frac{\frac{b_{n}+1}{a_{n}}}{1-\frac{b_{n}+1}{a_{n}}}. ∎

Proof of Theorem 1.2.

Let ϕ:XF\phi:X\to F be a locally dominant analytic map, and let x0Xx_{0}\in X. Replacing ϕ\phi with ϕx0=ϕϕ(x0)\phi_{x_{0}}=\phi-\phi(x_{0}), we may assume that ϕ(x0)=0\phi(x_{0})=0. We may further assume that XFnX\subseteq F^{n} is open, and apply Theorem 2.1, to get a log-principalization π:X~X\pi:\widetilde{X}\rightarrow X, such that, locally on a chart around a point in π1(x0)\pi^{-1}(x_{0}), ϕx0π(x~1,,x~n)=Cx~1a1x~nan\phi_{x_{0}}\circ\pi(\widetilde{x}_{1},...,\widetilde{x}_{n})=C\widetilde{x}_{1}^{a_{1}}\cdot...\cdot\widetilde{x}_{n}^{a_{n}}, and

Jacx~(π)=v(x~)x~1bnx~nbn.\operatorname{Jac}_{\widetilde{x}}(\pi)=v(\widetilde{x})\cdot\widetilde{x}_{1}^{b_{n}}\cdot...\cdot\widetilde{x}_{n}^{b_{n}}.

Let σc,(X)\sigma\in\mathcal{M}_{c,\infty}(X), with σ=g(x)μFn\sigma=g(x)\mu_{F}^{n}, and g>0g>0 in a neighborhood of x0x_{0}. Then σ=πμ\sigma=\pi_{*}\mu, where

μ=(gπ)(x~1,,x~n)|v(x~)|F|x~1|Fb1|x~n|FbnμFn.\mu=(g\circ\pi)(\widetilde{x}_{1},...,\widetilde{x}_{n})\left|v(\widetilde{x})\right|_{F}\left|\widetilde{x}_{1}\right|_{F}^{b_{1}}...\left|\widetilde{x}_{n}\right|_{F}^{b_{n}}\mu_{F}^{n}.

Since ϕσ=(ϕπ)μ\phi_{*}\sigma=(\phi\circ\pi)_{*}\mu, the theorem now follows from Lemmas 3.2 and 2.3. ∎

3.2. Relation to Fourier decay and other invariants

Let ϕ:XY\phi:X\rightarrow Y be an FF-analytic map between FF-analytic manifolds XX and YY. We have seen that each of the exponents ϵ(ϕ;x)\epsilon_{\star}(\phi;x) and lctF(ϕx;x)\operatorname{lct}_{F}(\phi_{x};x) provides a different quantification for the singularities of ϕ\phi near xXx\in X. When Y=FmY=F^{m}, one can further consider other invariants involving the Fourier transform of pushforward measures.

In §\mathsection1.1, we have defined k(ϕ;x)k_{\star}(\phi;x) as the minimal number of self-convolutions after which the pushforward densities of smooth measures supported near xx become bounded. Note that by the Plancherel theorem, for each νc,1(Fm)\nu\in\mathcal{M}_{c,1}(F^{m}) we have νkc,2(Fm)\nu^{*k}\in\mathcal{M}_{c,2}(F^{m}) if and only if (ν)L2k(Fm)\mathcal{F}(\nu)\in L^{2k}(F^{m}), whence

νkc,(Fm)(ν)L2k(Fm)ν2kc,(Fm).\nu^{*k}\in\mathcal{M}_{c,\infty}(F^{m})\Longrightarrow\mathcal{F}(\nu)\in L^{2k}(F^{m})\Longrightarrow\nu^{*2k}\in\mathcal{M}_{c,\infty}(F^{m}).

Thus, the exponent k(ϕ;x)k_{\star}(\phi;x) is, in general, roughly comparable to the LpL^{p}-class of (ϕμ)\mathcal{F}(\phi_{*}\mu) rather than to the LpL^{p}-class of the pushforward measure ϕμ\phi_{*}\mu itself. Instead of the LpL^{p}-class, we now focus on an invariant quantifying the Fourier decay of ϕμ\phi_{*}\mu on the power-law scale:

δ(ϕ;x):=supUxinfμc,(U)δ(ϕμ),\delta_{\star}(\phi;x):=\sup_{U\ni x}\inf_{\mu\in\mathcal{M}_{c,\infty}(U)}\delta_{\star}(\phi_{*}\mu),

where

(3.1) δ(ν):=sup{δ0:|(ν)(y)|yFδ}.\delta_{\star}(\nu):=\sup\{\delta\geq 0\,:\,\left|\mathcal{F}(\nu)(y)\right|\lesssim\left\|y\right\|_{F}^{-\delta}\}.

The study of the invariant δ(μ)\delta_{\star}(\mu), 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 f:f:\mathbb{\mathbb{R}}\rightarrow\mathbb{\mathbb{R}}, 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 dimY=1\dim Y=1; it is much less understood in high dimensions.

Remark 3.3.

Note that ϵ(ϕ;x)\epsilon_{\star}(\phi;x) and lctF(ϕx;x)\operatorname{lct}_{F}(\phi_{x};x) are preserved under analytic changes of coordinates around xx. On the other hand, δ(ϕ;x)\delta_{\star}(\phi;x) and k(ϕ;x)k_{\star}(\phi;x) might depend on the choice of coordinate system. For example, the map ϕ(x,y)=(x,x2(1+y1000))\phi(x,y)=(x,x^{2}(1+y^{1000})) satisfies ϵ(ϕ)=1999\epsilon_{\star}(\phi)=\frac{1}{999} (by Theorem 1.1), while δ(ϕ)=12\delta_{\star}(\phi)=\frac{1}{2} and k(ϕ)4k_{\star}(\phi)\leq 4. By applying the change of coordinates ψ(x,y)=(x,yx2)\psi(x,y)=(x,y-x^{2}), we get ψϕ(x,y)=(x,x2y1000)\psi\circ\phi(x,y)=(x,x^{2}y^{1000}), and still have ϵ(ψϕ)=1999\epsilon_{\star}(\psi\circ\phi)=\frac{1}{999}, while δ(ψϕ)=11000\delta_{\star}(\psi\circ\phi)=\frac{1}{1000} and k(ψϕ)1000k_{\star}(\psi\circ\phi)\geq 1000.

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 lctF(ϕ;x)1\operatorname{lct}_{F}(\phi;x)\leq 1). 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 FF be a local field, f:UFf:U\rightarrow F be a locally dominant analytic map with UFnU\subseteq F^{n} an open set. Let μc(U)\mu\in\mathcal{M}_{c}^{\infty}(U), and consider the pushforward measure fμf_{*}\mu. Fix a non-trivial additive character Ψ\Psi of FF. We may identify between FF and FF^{\vee} by tΨtt\longleftrightarrow\Psi_{t} where Ψt(y)=Ψ(ty)\Psi_{t}(y)=\Psi(ty). The Fourier transform (fμ)\mathcal{F}(f_{*}\mu) can then be written as

(fμ)(t)=FΨ(ty)𝑑fμ(y)=UΨ(tf(x))𝑑μ(x).\mathcal{F}(f_{*}\mu)(t)=\int_{F}\Psi(ty)df_{*}\mu(y)=\int_{U}\Psi(t\cdot f(x))d\mu(x).

To μ\mu and ff, one can further associates Igusa’s local zeta function

Zμ,f(s):=U|f(x)|Fs𝑑μ(x),s,Re(s)>0.Z_{\mu,f}(s):=\int_{U}\left|f(x)\right|_{F}^{s}d\mu(x),\,\,\,\,\,\,s\in\mathbb{C},\,\,\,\,\operatorname{Re}(s)>0.

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 Zμ,f(s)Z_{\mu,f}(s) (as well as certain twisted versions of it), and the Laurent expansions around them, controls the asymptotic expansions for fμf_{*}\mu as |y|F0\left|y\right|_{F}\rightarrow 0 and for (fμ)\mathcal{F}(f_{*}\mu) when |t|F\left|t\right|_{F}\rightarrow\infty, via the theory of Mellin transform (see [Igu78, Theorems 4.2, 4.3 and 5.3]). We next describe the asymptotic expansions of fμf_{*}\mu and (fμ)\mathcal{F}(f_{*}\mu).

Definition 3.4 (Asymptotic expansion, see [Igu78, Section I.2]).

Let xx_{\infty} be 0 or \infty. A sequence {φk}k\{\varphi_{k}\}_{k\in\mathbb{N}} of complex-valued functions on an open subset UU of F{,}F\in\{\mathbb{\mathbb{R}},\mathbb{C}\}, with φk(x)0\varphi_{k}(x)\neq 0 in a punctured neighborhood of xx_{\infty}, is said to constitute an asymptotic scale, if for every kk,

limxxφk+1(x)φk(x)=0.{\displaystyle\lim\limits_{x\to x_{\infty}}\frac{\varphi_{k+1}(x)}{\varphi_{k}(x)}=0}.

A function f:Uf:U\rightarrow\mathbb{C} is said to have an asymptotic expansion at xx_{\infty}, if there exists a sequence {an}n\{a_{n}\}_{n\in\mathbb{N}} of complex numbers such that for every k0k\geq 0, there exists C>0C>0 such that for all xx close enough to xx_{\infty}:

|f(x)i=0kaiφi(x)|Cφk+1(x).\left|f(x)-\sum\limits_{i=0}^{k}a_{i}\varphi_{i}(x)\right|\leq C\cdot\varphi_{k+1}(x).

In this case, we write

(3.2) f(x)k=0akφk(x) as xx.f(x)\approx\sum\limits_{k=0}^{\infty}a_{k}\varphi_{k}(x)\text{ as }x\rightarrow x_{\infty}.
Example 3.5.

Given a monotone increasing sequence Λ=(1<λ0<λ1<<λn<)\Lambda=(-1<\lambda_{0}<\lambda_{1}<...<\lambda_{n}<...) of real numbers, with no finite accumulation points, and given a sequence {mn}n\{m_{n}\}_{n\in\mathbb{N}} of positive integers, set φ0,φ1,φ2,\varphi_{0},\varphi_{1},\varphi_{2},... to be the sequence:

xλ0log(x)m01,xλ0log(x)m02,,xλ0,xλ1log(x)m11,,xλ1,x^{\lambda_{0}}\log(x)^{m_{0}-1},x^{\lambda_{0}}\log(x)^{m_{0}-2},...,x^{\lambda_{0}},x^{\lambda_{1}}\log(x)^{m_{1}-1},...,x^{\lambda_{1}},...

for x>0x>0. Then {φk}k\{\varphi_{k}\}_{k\in\mathbb{N}} is an asymptotic scale at x=0x_{\infty}=0.

We now describe the asymptotic expansions of pushforward measures and their Fourier transforms. We fix a local field FF of characteristic 0 and an analytic FF-manifold XX. If FF is non-Archimedean we further fix a uniformizer ϖF𝒪F\varpi_{F}\in\mathcal{O}_{F}. We set F1×:={xF:|x|F=1}F_{1}^{\times}:=\{x\in F:\left|x\right|_{F}=1\} and denote by ac:F×F1×\mathbb{\mathrm{ac}}:F^{\times}\rightarrow F_{1}^{\times} the angular component map

ac(t)={tϖFval(t)if F is non-Archimedeant|t|if F is Archimedean.\mathbb{\mathrm{ac}}(t)=\begin{cases}t\varpi_{F}^{-\mathbb{\mathrm{val}}(t)}&\text{if }F\text{ is non-Archimedean}\\ \frac{t}{\left|t\right|}&\text{if }F\text{ is Archimedean}.\end{cases}
Theorem 3.6 ([Jea70, Mal74, Igu78], see also [VZnG17, Section 4] and [Den91, Theorem 1.3.2 and Corollary 1.4.5]).

Let f:XFf:X\rightarrow F be a locally dominant analytic map, let μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X), and write fμ=g(y)dμFf_{*}\mu=g(y)d\mu_{F}. Suppose that 0 is the only critical value of ff. Then there exist:

  • a sequence Λ={λk}k0\Lambda=\{\lambda_{k}\}_{k\geq 0}, consisting of strictly increasing positive real numbers with limkλk=\lim\limits_{k\to\infty}\lambda_{k}=\infty, if F{,}F\in\{\mathbb{\mathbb{R}},\mathbb{C}\}, or a finite set of complex numbers λk\lambda_{k}, with Reλk>0\operatorname{Re}\lambda_{k}>0, and Imλk2πNln(qF)\operatorname{Im}\lambda_{k}\in\frac{2\pi}{N\ln(q_{F})}\mathbb{N} for some NN\in\mathbb{N}, if FF is non-Archimedean;

  • a sequence {mk}k0\{m_{k}\}_{k\geq 0} of positive integers;

  • smooth functions ak,m,μ,bk,m,μa_{k,m,\mu},b_{k,m,\mu} on F1×F_{1}^{\times},

such that:

  1. (1)

    For F{,}F\in\{\mathbb{\mathbb{R}},\mathbb{C}\}, g(y)g(y) 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) g(y)k0m=1mkak,m,μ(ac(y))|y|Fλk1(log|y|F)m1,|y|F0,g(y)\approx\sum_{k\geq 0}\sum\limits_{m=1}^{m_{k}}a_{k,m,\mu}(\mathbb{\mathrm{ac}}(y))\cdot\left|y\right|_{F}^{\lambda_{k}-1}(\log\left|y\right|_{F})^{m-1},\left|y\right|_{F}\rightarrow 0,

    and (fμ)\mathcal{F}(f_{*}\mu) admits an asymptotic expansion of the form,

    (3.4) (fμ)(t)k0m=1mkbk,m,μ(ac(t))|t|Fλk(log|t|F)m1,|t|F.\mathcal{F}(f_{*}\mu)(t)\approx\sum_{k\geq 0}\sum\limits_{m=1}^{m_{k}}b_{k,m,\mu}(\mathbb{\mathrm{ac}}(t))\cdot\left|t\right|_{F}^{-\lambda_{k}}(\log\left|t\right|_{F})^{m-1},\left|t\right|_{F}\rightarrow\infty.
  2. (2)

    For FF non-Archimedean, g(y)g(y) and (fμ)\mathcal{F}(f_{*}\mu) admits an expansion as in (3.3) and (3.4), where ((\approx)) is replaced with ((==)), and both sums are finite.

  3. (3)

    For each λk\lambda_{k}, the functions (bk,m,μ)m=1mk(b_{k,m,\mu})_{m=1}^{m_{k}} are determined by the functions (ak,m,μ)m=1mk(a_{k,m,\mu})_{m=1}^{m_{k}}. If Reλk<1\operatorname{Re}\lambda_{k}<1, the map taking the latter to the former is one-to-one, and, moreover, the leading function bk,mk,μb_{k,m_{k},\mu} is not identically zero (provided that mkm_{k} is defined so that ak,mk,μa_{k,m_{k},\mu} is not identically 0).

Remark 3.7.
  1. (1)

    The maps (ak,m,μ)m=1mk(bk,m,μ)m=1mk(a_{k,m,\mu})_{m=1}^{m_{k}}\mapsto(b_{k,m,\mu})_{m=1}^{m_{k}} of item (3) can be explicitly described; see [VZnG17, Proposition 4.6] and [Igu78, Section 2.2].

  2. (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].

We record the following immediate corollary of Equation (3.3) in Theorem 3.6:

Corollary 3.8.

Let f:XFf:X\to F be a locally dominant analytic map, let μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X), and suppose that ϵ(fμ)<\epsilon_{\star}(f_{*}\mu)<\infty. Then fμc,1+ϵ(fμ)(F)f_{*}\mu\notin\mathcal{M}_{c,1+\epsilon_{\star}(f_{*}\mu)}(F), i.e. the supremum in the definition of ϵ\epsilon_{\star} is not achieved.

Theorem 3.6 also implies the following corollary, relating ϵ\epsilon_{\star} to δ\delta_{\star}:

Corollary 3.9.

Let ϕ:XF\phi:X\rightarrow F be a locally dominant analytic map such that 0 is the only critical value. Then for each μc(X)\mu\in\mathcal{M}_{c}^{\infty}(X)

(3.5) ϵ(ϕμ)={if δ(ϕμ)1,δ(ϕμ)1δ(ϕμ)if δ(ϕμ)<1.\epsilon_{\star}(\phi_{*}\mu)=\begin{cases}\infty&\text{if }\delta_{\star}(\phi_{*}\mu)\geq 1,\\ \frac{\delta_{\star}(\phi_{*}\mu)}{1-\delta_{\star}(\phi_{*}\mu)}&\text{if }\delta_{\star}(\phi_{*}\mu)<1.\end{cases}
Proof.

Write ϕμ=g(y)dμF\phi_{*}\mu=g(y)d\mu_{F}. If F=F=\mathbb{\mathbb{R}}, then by Theorem 3.6, g(y)g(y) can be expanded near y=0y=0, so that the leading term is |y|Fλk01(log|y|F)mk01\left|y\right|_{F}^{\lambda_{k_{0}}-1}(\log\left|y\right|_{F})^{m_{k_{0}}-1} for some λk0>0\lambda_{k_{0}}\in\mathbb{\mathbb{R}}_{>0} and mk0m_{k_{0}}\in\mathbb{N}. Similarly, (ϕμ)\mathcal{F}(\phi_{*}\mu) can be expanded near t=t=\infty. If λk0<1\lambda_{k_{0}}<1, then by Item (3) of Theorem 3.6, |y|Fλk0(log|y|F)mk01\left|y\right|_{F}^{-\lambda_{k_{0}}}(\log\left|y\right|_{F})^{m_{k_{0}}-1} is the leading term of (ϕμ)\mathcal{F}(\phi_{*}\mu), and thus δ(ϕμ)=λk0\delta_{\star}(\phi_{*}\mu)=\lambda_{k_{0}}. If λk0=1\lambda_{k_{0}}=1, then Item (3) implies that δ(ϕμ)1\delta_{\star}(\phi_{*}\mu)\geq 1.

Since 0 is the only critical value of ϕ\phi, g(y)g(y) is bounded outside any neighborhood of 0, so g(y)1+ϵg(y)^{1+\epsilon} is integrable if and only if it is integrable in a small ball BB around 0, and this holds if and only if (λk01)(1+ϵ)>1(\lambda_{k_{0}}-1)(1+\epsilon)>-1, i.e. either λk01\lambda_{k_{0}}\geq 1 and then ϵ(ϕμ)=\epsilon_{\star}(\phi_{*}\mu)=\infty, or ϵ<λk01λk0=δ(ϕμ)1δ(ϕμ)\epsilon<\frac{\lambda_{k_{0}}}{1-\lambda_{k_{0}}}=\frac{\delta_{\star}(\phi_{*}\mu)}{1-\delta_{\star}(\phi_{*}\mu)}.

The case when FF 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 λ:=min𝑘Reλk\lambda:=\underset{k}{\min}\operatorname{Re}\lambda_{k} and suppose λ<1\lambda<1. Then by Theorem 3.6, g(y)g(y) can be written as g(y)=glead(y)+gerror(y)g(y)=g_{\mathrm{lead}}(y)+g_{\mathrm{error}}(y), where

glead(y):=mj=0N1a~j,m,μ(ac(y))|y|Fλ+2πijNln(qF)1(log|y|F)m1,g_{\mathrm{lead}}(y):=\sum_{m}\sum_{j=0}^{N-1}\widetilde{a}_{j,m,\mu}(\mathbb{\mathrm{ac}}(y))\cdot\left|y\right|_{F}^{\lambda+\frac{2\pi ij}{N\ln(q_{F})}-1}(\log\left|y\right|_{F})^{m-1},

a~j,m,μ(ac(y))\widetilde{a}_{j,m,\mu}(\mathbb{\mathrm{ac}}(y)) is equal to akj,m,μ(ac(y))a_{k_{j},m,\mu}(\mathbb{\mathrm{ac}}(y)) if λkj=λ+2πijNln(qF)Λ\lambda_{k_{j}}=\lambda+\frac{2\pi ij}{N\ln(q_{F})}\in\Lambda and 0 otherwise, and furthermore,

gerror(y):=k:Reλk>λm=1mkak,m,μ(ac(y))|y|Fλk1(log|y|F)m1.g_{\mathrm{error}}(y):=\sum_{k:\operatorname{Re}\lambda_{k}>\lambda}\sum\limits_{m=1}^{m_{k}}a_{k,m,\mu}(\mathbb{\mathrm{ac}}(y))\cdot\left|y\right|_{F}^{\lambda_{k}-1}(\log\left|y\right|_{F})^{m-1}.

We first show there is an arithmetic progression Ia,N,l:={a+bN}l<bI_{a,N,l}:=\{a+bN\}_{l<b\in\mathbb{N}} for some a,la,l\in\mathbb{N}, and y0F1×y_{0}\in F_{1}^{\times}, such that

(3.6) |glead(y)|>CF|y|Fλ1,\left|g_{\mathrm{lead}}(y)\right|>C_{F}\left|y\right|_{F}^{\lambda-1},

for all ySa,N,l,d,y0:={z𝒪F:val(ac(z)y0)d,val(z)Ia,N,l}y\in S_{a,N,l,d,y_{0}}:=\left\{z\in\mathcal{O}_{F}:\mathbb{\mathrm{val}}(\mathbb{\mathrm{ac}}(z)-y_{0})\geq d,\,\mathbb{\mathrm{val}}(z)\in I_{a,N,l}\right\}, for some constant CFC_{F} independent of yy. It is enough to show (3.6) for the terms in glead(y)g_{\mathrm{lead}}(y) where m=m0m=m_{0} is maximal such that for some 0jN10\leq j\leq N-1, a~j,m,μ(ac(y))\widetilde{a}_{j,m,\mu}(\mathbb{\mathrm{ac}}(y)) is not identically zero. Note that

|y|F2πijNln(qF)=qFval(y)2πijNln(qF)=eln(qF)val(y)2πijNln(qF)=eval(y)2πijN.\left|y\right|_{F}^{\frac{2\pi ij}{N\ln(q_{F})}}=q_{F}^{-\mathbb{\mathrm{val}}(y)\frac{2\pi ij}{N\ln(q_{F})}}=e^{-\ln(q_{F})\mathbb{\mathrm{val}}(y)\frac{2\pi ij}{N\ln(q_{F})}}=e^{-\mathbb{\mathrm{val}}(y)\frac{2\pi ij}{N}}.

Let y0F1×y_{0}\in F_{1}^{\times} be such that a~j,m0,μ(y0)0\widetilde{a}_{j,m_{0},\mu}(y_{0})\neq 0 for some 0jN10\leq j\leq N-1. Choose dd\in\mathbb{N} such that each of a~0,m0,μ(z),,a~0,m0,μ(z)\widetilde{a}_{0,m_{0},\mu}(z),\dots,\widetilde{a}_{0,m_{0},\mu}(z) is constant on the ball val(zy0)d\mathbb{\mathrm{val}}(z-y_{0})\geq d. Note that the functions f0,,fN1:/Nf_{0},...,f_{N-1}:\mathbb{Z}/N\mathbb{Z}\rightarrow\mathbb{C}, fj(t)=et2πijNf_{j}(t)=e^{-t\frac{2\pi ij}{N}} are the irreducible characters of /N\mathbb{Z}/N\mathbb{Z} and hence they are linearly independent. In particular, there exists aa\in\mathbb{N}, such that

|j=0N1a~j,m0,μ(y0)et2πijN|=CF>0,\left|\sum_{j=0}^{N-1}\widetilde{a}_{j,m_{0},\mu}(y_{0})e^{-t\frac{2\pi ij}{N}}\right|=C_{F}>0,

for all tIa,N,0t\in I_{a,N,0} . Taking ySa,N,l,d,y0y\in S_{a,N,l,d,y_{0}} for l1l\gg 1, we deduce (3.6) as required.

Finally, by Item (3) of Theorem 3.6 and by an argument similar to the one above, it follows that δ(ϕμ)=λ\delta_{\star}(\phi_{*}\mu)=\lambda. Hence, (3.6) implies that Sa,N,l,d,y0g(y)1+ϵ𝑑μF\int_{S_{a,N,l,d,y_{0}}}g(y)^{1+\epsilon}d\mu_{F} diverges for every ϵλ1λ=δ(ϕμ)1δ(ϕμ)\epsilon\geq\frac{\lambda}{1-\lambda}=\frac{\delta_{\star}(\phi_{*}\mu)}{1-\delta_{\star}(\phi_{*}\mu)}. On the other hand, a similar argument as in the case F=F=\mathbb{\mathbb{R}} shows that Fg(y)1+ϵ𝑑μF\int_{F}g(y)^{1+\epsilon}d\mu_{F} converges for ϵ<λ1λ\epsilon<\frac{\lambda}{1-\lambda}, so the corollary follows. ∎

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 val(y)\mathbb{\mathrm{val}}(y). Here is a simple example which illustrates this phenomenon.

Example 3.10.

Let dd\in\mathbb{N} and let p>dp>d be a prime. Let ϕ:pp\phi:\mathbb{Q}_{p}\rightarrow\mathbb{Q}_{p} be the map ϕ(x)=xd\phi(x)=x^{d}. Write ϕμp=g(y)μp\phi_{*}\mu_{\mathbb{\mathbb{Z}}_{p}}=g(y)\mu_{\mathbb{\mathbb{Z}}_{p}}. Then for almost all ypy\in\mathbb{\mathbb{Z}}_{p} we have

g(y)=#ϕ1(y)|y|p1+1d=a(ac(y))1{zp:d|val(z)}(y)|y|p1+1d.g(y)=\#\phi^{-1}(y)\cdot\left|y\right|_{p}^{-1+\frac{1}{d}}=a(\mathbb{\mathrm{ac}}(y))1_{\{z\in\mathbb{\mathbb{Z}}_{p}:d|\mathbb{\mathrm{val}}(z)\}}(y)\cdot\left|y\right|_{p}^{-1+\frac{1}{d}}.

where a(z):=#{xp:xd=z}a(z):=\#\{x\in\mathbb{\mathbb{Z}}_{p}:x^{d}=z\}. Note that by Schur orthogonality, j=0d1et2πijd=d\sum_{j=0}^{d-1}e^{-t\frac{2\pi ij}{d}}=d if d|td|t and is 0 if dtd\nmid t\in\mathbb{N}. In particular

1{zp:d|val(z)}(y)=1dj=0d1eval(y)2πijd=1dj=0d1|y|p2πijdln(qF),1_{\{z\in\mathbb{\mathbb{Z}}_{p}:d|\mathbb{\mathrm{val}}(z)\}}(y)=\frac{1}{d}\sum_{j=0}^{d-1}e^{-\mathbb{\mathrm{val}}(y)\frac{2\pi ij}{d}}=\frac{1}{d}\sum_{j=0}^{d-1}\left|y\right|_{p}^{\frac{2\pi ij}{d\ln(q_{F})}},

and therefore, the expansion of g(y)g(y) as in (3.3) is

g(y)=j=0d1a(ac(y))d|y|p1+1d+2πijdln(qF).g(y)=\sum_{j=0}^{d-1}\frac{a(\mathbb{\mathrm{ac}}(y))}{d}\left|y\right|_{p}^{-1+\frac{1}{d}+\frac{2\pi ij}{d\ln(q_{F})}}.

3.2.2. Relations between the invariants

We now show that ϵ(ϕ;x),δ(ϕ;x)\epsilon_{\star}(\phi;x),\delta_{\star}(\phi;x) and k(ϕ;x)k_{\star}(\phi;x) are essentially determined by the log-canonical threshold lctF(ϕx;x)\operatorname{lct}_{F}(\phi_{x};x) whenever lctF(ϕ;x)1\operatorname{lct}_{F}(\phi;x)\leq 1. In particular, we show that

ϵ(ϕ;x),δ(ϕ;x),1k(ϕ;x) and lctF(ϕx;x)\epsilon_{\star}(\phi;x),\delta_{\star}(\phi;x),\frac{1}{k_{\star}(\phi;x)}\text{ and }\operatorname{lct}_{F}(\phi_{x};x)

are asymptotically equivalent as lctF(ϕx;x)0\operatorname{lct}_{F}(\phi_{x};x)\rightarrow 0.

Theorem 1.2 already shows that ϵ(ϕ;x)=lctF(ϕx;x)1lctF(ϕx;x)\epsilon_{\star}(\phi;x)=\frac{\operatorname{lct}_{F}(\phi_{x};x)}{1-\operatorname{lct}_{F}(\phi_{x};x)} if lctF(ϕx;x)<1\operatorname{lct}_{F}(\phi_{x};x)<1. We further have the following:

Proposition 3.11.

Let ϕ:XF\phi:X\rightarrow F be a dominant FF-analytic map. Then:

  1. (1)

    If lctF(ϕx;x)<1\operatorname{lct}_{F}(\phi_{x};x)<1, then δ(ϕ;x)=lctF(ϕx;x)\delta_{\star}(\phi;x)=\operatorname{lct}_{F}(\phi_{x};x). In particular,

    lctF(ϕxϕy;(x,y)){=lctF(ϕx;x)+lctF(ϕy;y),lctF(ϕx;x)+lctF(ϕy;y)<1,1,otherwise.\operatorname{lct}_{F}(\phi_{x}*\phi_{y};(x,y))\,\begin{cases}=\operatorname{lct}_{F}(\phi_{x};x)+\operatorname{lct}_{F}(\phi_{y};y)~{},&\operatorname{lct}_{F}(\phi_{x};x)+\operatorname{lct}_{F}(\phi_{y};y)<1,\\ \geq 1~{},&\text{otherwise}.\end{cases}
  2. (2)

    We have:

    1lctF(ϕx;x)k(ϕ;x)1lctF(ϕx;x)+1.\left\lceil\frac{1}{\operatorname{lct}_{F}(\phi_{x};x)}\right\rceil\leq k_{\star}(\phi;x)\leq\left\lfloor\frac{1}{\operatorname{lct}_{F}(\phi_{x};x)}\right\rfloor+1.
Proof.

Let us first prove Item (1). Let xXx\in X. Replacing ϕ\phi with ϕx\phi_{x}, we may assume that ϕ(x)=0\phi(x)=0. We may choose an open neighborhood UU such that 0 is the only critical value of ϕ|U\phi|_{U}, U¯\overline{U} is compact and such that for each μc(U)\mu\in\mathcal{M}_{c}^{\infty}(U), and each s<lctF(ϕ;x)s<\operatorname{lct}_{F}(\phi;x), one has |ϕ(x)|Fs𝑑μ(s)<\int\left|\phi(x)\right|_{F}^{-s}d\mu(s)<\infty. Taking any μ\mu which does not vanish at xx, we get by Theorem 1.2 and Lemma 3.2 that

ϵ(ϕμ)=lctF(ϕx;x)1lctF(ϕx;x)=ϵ(ϕ;x)<.\epsilon_{\star}(\phi_{*}\mu)=\frac{\operatorname{lct}_{F}(\phi_{x};x)}{1-\operatorname{lct}_{F}(\phi_{x};x)}=\epsilon_{\star}(\phi;x)<\infty.

Corollary 3.9 implies that

δ(ϕμ)=ϵ(ϕμ)1+ϵ(ϕμ)=ϵ(ϕ;x)1+ϵ(ϕ;x)=lctF(ϕx;x).\delta_{\star}(\phi_{*}\mu)=\frac{\epsilon_{\star}(\phi_{*}\mu)}{1+\epsilon_{\star}(\phi_{*}\mu)}=\frac{\epsilon_{\star}(\phi;x)}{1+\epsilon_{\star}(\phi;x)}=\operatorname{lct}_{F}(\phi_{x};x).

Since the equalities above hold for μ\mu of arbitrarily small support around xx, we get δ(ϕ;x)=δ(ϕμ)=lctF(ϕx;x)\delta_{\star}(\phi;x)=\delta_{\star}(\phi_{*}\mu)=\operatorname{lct}_{F}(\phi_{x};x) as required. Since Fourier transform translates convolution into product, we have

δ(ϕϕ;(x,y))=δ(ϕ;x)+δ(ϕ;y),\delta_{\star}(\phi*\phi;(x,y))=\delta_{\star}(\phi;x)+\delta_{\star}(\phi;y),

which implies the second part of Item (1) (see also [Den91, Section 5.1]).

We now turn to Item (2). Set k0:=1lctF(ϕx;x)k_{0}:=\left\lceil\frac{1}{\operatorname{lct}_{F}(\phi_{x};x)}\right\rceil. For a positive integer k<k0k<k_{0} we get by Item (1), that lctF(ϕxk;(x,,x))<1\operatorname{lct}_{F}(\phi_{x}^{*k};(x,...,x))<1, so that ϵ(ϕxk;(x,,x))<\epsilon_{\star}(\phi_{x}^{*k};(x,...,x))<\infty. This implies the lower bound k(ϕ;x)k0k_{\star}(\phi;x)\geq k_{0}. The upper bound follows from (1.3) and Theorem 1.2. ∎

Remark 3.12.

Item (1) of Proposition 3.11 can be seen as an FF-analytic interpretation of a theorem by Thom–Sebastiani [ST71] (see e.g. [MSS18] for the case F=F=\mathbb{C}).

The combination of Corollaries 3.8 and 3.9 implies Theorem 1.3, as follows.

Proof of Theorem 1.3.

Let μjc(Fnj)\mu_{j}\in\mathcal{M}_{c}^{\infty}(F^{n_{j}}), and suppose that

(3.7) ν1ν21+ϵ(F),ϵ>0,\nu_{1}*\nu_{2}\in\mathcal{M}_{1+\epsilon}(F),\quad\epsilon>0,

where νj=(ϕj)μj\nu_{j}=(\phi_{j})_{*}\mu_{j}. Assume by contradiction that

ϵ(ν1)1+ϵ(ν1)+ϵ(ν2)1+ϵ(ν2)ϵ1+ϵ,\frac{\epsilon_{\star}(\nu_{1})}{1+\epsilon_{\star}(\nu_{1})}+\frac{\epsilon_{\star}(\nu_{2})}{1+\epsilon_{\star}(\nu_{2})}\leq\frac{\epsilon}{1+\epsilon},

For each j{1,2}j\in\{1,2\}, choose a finite cover iUi,j\bigcup_{i}U_{i,j} of Supp(μj)\operatorname{Supp}(\mu_{j}) by open balls in FnjF^{n_{j}}, such that for each ii, ϕj|Ui,j\phi_{j}|_{U_{i,j}} has at most one critical value zi,jz_{i,j} of ϕj\phi_{j}. We can write μj=iμi,j\mu_{j}=\sum_{i}\mu_{i,j} with μi,j\mu_{i,j} supported inside Ui,jU_{i,j}. Taking ϕi,j=ϕj|Ui,jzi,j\phi_{i,j}=\phi_{j}|_{U_{i,j}}-z_{i,j}, we can find i1i_{1} and i2i_{2} such that

ϵ(νj)=ϵ((ϕij,j)(μij,j)),j=1,2.\epsilon_{\star}(\nu_{j})=\epsilon_{\star}\left((\phi_{i_{j},j})_{*}(\mu_{i_{j},j})\right),\quad j=1,2.

Now, Corollary 3.9 implies that

δ((ϕij,j)(μij,j))=ϵ((ϕij,j)(μij,j))1+ϵ((ϕij,j)(μij,j)),\delta_{\star}\left((\phi_{i_{j},j})_{*}(\mu_{i_{j},j})\right)=\frac{\epsilon_{\star}\left((\phi_{i_{j},j})_{*}(\mu_{i_{j},j})\right)}{1+\epsilon_{\star}\left((\phi_{i_{j},j})_{*}(\mu_{i_{j},j})\right)},

whence

δ((ϕi1,1)(μi1,1)(ϕi2,2)(μi2,2))ϵ1+ϵ.\delta_{\star}\left((\phi_{i_{1},1})_{*}(\mu_{i_{1},1})*(\phi_{i_{2},2})_{*}(\mu_{i_{2},2})\right)\leq\frac{\epsilon}{1+\epsilon}.

Using Corollary 3.9 once again, we obtain that

ϵ(ν1ν2)ϵ((ϕi1,1)(μi1,1)(ϕi2,2)(μi2,2))ϵ.\epsilon_{\star}(\nu_{1}*\nu_{2})\leq\epsilon_{\star}\left((\phi_{i_{1},1})_{*}(\mu_{i_{1},1})*(\phi_{i_{2},2})_{*}(\mu_{i_{2},2})\right)\leq\epsilon.

In view of Corollary 3.8, this contradicts (3.7). ∎

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 ϵ(ϕ;x)\epsilon_{\star}(\phi;x) 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 ϕ:FnF\phi:F^{n}\rightarrow F, we have 𝒥ϕ=ϕx1,,ϕxn\mathcal{J}_{\phi}=\langle\frac{\partial\phi}{\partial x_{1}},...,\frac{\partial\phi}{\partial x_{n}}\rangle, which we denote by ϕ\langle\nabla\phi\rangle.

Proposition 3.13.

Let ϕ:FnF\phi:F^{n}\rightarrow F be an analytic map. Then for every xFnx\in F^{n}, we have

lctF(ϕ;x)1lctF(ϕ;x)lctF(ϕx;x)1lctF(ϕx;x)lctF(ϕ;x)lctF(ϕx;x),\frac{\operatorname{lct}_{F}(\langle\nabla\phi\rangle;x)}{1-\operatorname{lct}_{F}(\langle\nabla\phi\rangle;x)}\geq\frac{\operatorname{lct}_{F}(\phi_{x};x)}{1-\operatorname{lct}_{F}(\phi_{x};x)}\geq\operatorname{lct}_{F}(\langle\nabla\phi\rangle;x)\geq\operatorname{lct}_{F}(\phi_{x};x),

where the middle (resp. left) inequality holds whenever lctF(ϕx;x)<1\operatorname{lct}_{F}(\phi_{x};x)<1 (resp. lctF(ϕ;x)<1\operatorname{lct}_{F}(\langle\nabla\phi\rangle;x)<1).

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 lctF(ϕ;x0)lctF(ϕx0;x0)\operatorname{lct}_{F}(\langle\nabla\phi\rangle;x_{0})\geq\operatorname{lct}_{F}(\phi_{x_{0}};x_{0}) for a fixed x0Fnx_{0}\in F^{n}.

Recall that (a1,,an)F:=maxi|ai|F\left\|(a_{1},...,a_{n})\right\|_{F}:=\max_{i}\left|a_{i}\right|_{F} is the maximum norm on FnF^{n}. We would like to relate between BϕFs𝑑x\int_{B}\left\|\nabla\phi\right\|_{F}^{-s}dx and B|ϕx0|Fs𝑑x\int_{B}\left|\phi_{x_{0}}\right|_{F}^{-s}dx, where BB is a small ball around x0x_{0}, for s>0s>0 small enough. Let π:X~Fn\pi:\widetilde{X}\rightarrow F^{n} be a resolution of singularities of ϕx0\phi_{x_{0}}. Then we have

(3.8) BϕFs𝑑x=π1(B)ϕ|π(x~)Fs|Jacx~(π)|Fdx~.\int_{B}\left\|\nabla\phi\right\|_{F}^{-s}dx=\int_{\pi^{-1}(B)}\left\|\nabla\phi|_{\pi(\widetilde{x})}\right\|_{F}^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}d\widetilde{x}.

Since π1(B)\pi^{-1}(B) is compact, by working locally over finitely many pieces, and using Theorem 2.1, we may replace π1(B)\pi^{-1}(B) by a compact neighborhood LFnL\subseteq F^{n} of 0, π(0)=x0\pi(0)=x_{0}, and further assume that for x~=(x~1,,x~n)\widetilde{x}=(\widetilde{x}_{1},...,\widetilde{x}_{n}):

(3.9) ϕ~(x~):=ϕπ(x~)=Cx~1a1x~nan and Jacx~(π)=v(x~)x~1b1x~nbn,\widetilde{\phi}(\widetilde{x}):=\phi\circ\pi(\widetilde{x})=C\widetilde{x}_{1}^{a_{1}}\cdots\widetilde{x}_{n}^{a_{n}}\text{ and }\operatorname{Jac}_{\widetilde{x}}(\pi)=v(\widetilde{x})\widetilde{x}_{1}^{b_{1}}\cdots\widetilde{x}_{n}^{b_{n}},

for some analytic unit v(x~)v(\widetilde{x}) and a constant CFC\in F. Note that on LL all of the entries of dx~πd_{\widetilde{x}}\pi are smaller, in absolute value, than some constant C~\widetilde{C}, so that the operator norm dx~πop:=supvFn:vF=1dx~πvF\left\|d_{\widetilde{x}}\pi\right\|_{\mathrm{op}}:=\underset{v\in F^{n}:\left\|v\right\|_{F}=1}{\sup}\left\|d_{\widetilde{x}}\pi\cdot v\right\|_{F} of dx~πd_{\widetilde{x}}\pi is bounded by a constant. Since ϕ~|x~=ϕ|π(x~)dx~π\nabla\widetilde{\phi}|_{\widetilde{x}}=\nabla\phi|_{\pi(\widetilde{x})}\cdot d_{\widetilde{x}}\pi, for s>0s>0 we get:

Lϕ|π(x~)Fs|Jacx~(π)|Fdx~\displaystyle\text{$\int_{L}\left\|\nabla\phi|_{\pi(\widetilde{x})}\right\|_{F}^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}$}d\widetilde{x} Ldx~πopsϕ~|x~Fs|Jacx~(π)|Fdx~Lϕ~|x~Fs|Jacx~(π)|Fdx~\displaystyle\leq\int_{L}\left\|d_{\widetilde{x}}\pi\right\|_{\mathrm{op}}^{s}\left\|\nabla\widetilde{\phi}|_{\widetilde{x}}\right\|_{F}^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}d\widetilde{x}\lesssim\int_{L}\left\|\nabla\widetilde{\phi}|_{\widetilde{x}}\right\|_{F}^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}d\widetilde{x}
L(maxj|x~1|Fa1|x~j|Faj1|x~n|Fan)s|Jacx~(π)|F𝑑x~\displaystyle\lesssim\int_{L}\left(\max_{j}\left|\widetilde{x}_{1}\right|_{F}^{a_{1}}\cdot...\cdot\left|\widetilde{x}_{j}\right|_{F}^{a_{j}-1}\cdot...\cdot\left|\widetilde{x}_{n}\right|_{F}^{a_{n}}\right)^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}d\widetilde{x}
L(|x~1|Fa1|x~n|Fan)s|Jacx~(π)|F𝑑x~π(L)|ϕx0|Fs𝑑x,\displaystyle\lesssim\int_{L}\left(\left|\widetilde{x}_{1}\right|_{F}^{a_{1}}...\left|\widetilde{x}_{n}\right|_{F}^{a_{n}}\right)^{-s}\left|\operatorname{Jac}_{\widetilde{x}}(\pi)\right|_{F}d\widetilde{x}\lesssim\int_{\pi(L)}\left|\phi_{x_{0}}\right|_{F}^{-s}dx,

which concludes the proof. ∎

Remark 3.14.

Here is an alternative approach to Proposition 3.13, as suggested by the referee. For simplicity suppose F=F=\mathbb{C}. We may assume that x=0nx=0\in\mathbb{C}^{n}. Recall from [HS06, Definition 1.1.1] that the integral closure I¯\overline{I} of an ideal II in a commutative ring RR, is the set of elements rRr\in R for which there are ajIja_{j}\in I^{j} such that j=0najrnj=0\sum_{j=0}^{n}a_{j}r^{n-j}=0 for some nn\in\mathbb{N}. We now take R={x1,,xn}R=\mathbb{C}\{x_{1},...,x_{n}\} to be the ring of convergent power series, I:=ϕI:=\langle\nabla\phi\rangle and J:=ϕ0J:=\langle\phi_{0}\rangle. By [HS06, Corollary 7.1.4], we have JI¯J\subseteq\overline{I} and in particular lct(J;0)lct(I¯;0)\operatorname{lct}_{\mathbb{C}}(J;0)\leq\operatorname{lct}_{\mathbb{C}}(\overline{I};0) (see [Mus12, Property 1.15]). However, by [Mus12, Property 1.12] (see also [dFM09, Proposition 2.4]) we have lct(I;0)=lct(I¯;0)\operatorname{lct}_{\mathbb{C}}(I;0)=\operatorname{lct}_{\mathbb{C}}(\overline{I};0), which implies that lct(J;0)lct(I;0)\operatorname{lct}_{\mathbb{C}}(J;0)\leq\operatorname{lct}_{\mathbb{C}}(I;0) as required. This argument likely generalizes to any local field FF of characteristic 0, 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. dimY>1\dim Y>1). Here we will see that the connection between the four invariants ϵ(ϕ;x)\epsilon_{\star}(\phi;x), δ(ϕ;x)\delta_{\star}(\phi;x), k(ϕ;x)k_{\star}(\phi;x) and lctF(ϕx;x)\operatorname{lct}_{F}(\phi_{x};x) is not as tight as in the one-dimensional case. We first provide a simpler description of δ(ϕ;x)\delta_{\star}(\phi;x).

Lemma 3.15.

Let FF be a local field of characteristic 0 and ϕ:FnFm\phi:F^{n}\rightarrow F^{m} be an analytic map. Then

δ(ϕ;x)=inf{δ(ϕ;x)},\delta_{\star}(\phi;x)=\inf_{\ell}\{\delta_{\star}(\ell\circ\phi;x)\},

where \ell runs over all non-zero linear functionals :FmF\ell:F^{m}\rightarrow F.

Proof.

For each μc(Fn)\mu\in\mathcal{M}_{c}^{\infty}(F^{n}), we have

|(ϕμ)(z)|zFδ|(ϕμ)(ta1,,tam)||t|Fδ,\left|\mathcal{F}(\phi_{*}\mu)(z)\right|\lesssim\left\|z\right\|_{F}^{-\delta}\Longleftrightarrow\left|\mathcal{F}(\phi_{*}\mu)(ta_{1},...,ta_{m})\right|\lesssim\left|t\right|_{F}^{-\delta},

for each a=(a1,,am)a=(a_{1},...,a_{m}) with aF=1\left\|a\right\|_{F}=1 and tFt\in F. Setting a(y1,,ym)=i=1maiyi\ell_{a}(y_{1},...,y_{m})=\sum_{i=1}^{m}a_{i}y_{i}, we have,

(ϕμ)(ta1,,tam)=FmΨ(ta(y))𝑑ϕμ=FΨ(tz)d(aϕ)μ=((aϕ)μ)(t),\mathcal{F}(\phi_{*}\mu)(ta_{1},...,ta_{m})=\int_{F^{m}}\Psi(t\ell_{a}(y))\cdot d\phi_{*}\mu=\int_{F}\Psi(tz)\cdot d(\ell_{a}\circ\phi)_{*}\mu=\mathcal{F}\left((\ell_{a}\circ\phi)_{*}\mu\right)(t),

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 YY grows.

Example 3.16.

Consider the map ϕ:mm\phi:\mathbb{C}^{m}\rightarrow\mathbb{C}^{m}, defined by ϕ(x1,,xm)=(x1d,x1dx2,,x1dxm)\phi(x_{1},...,x_{m})=(x_{1}^{d},x_{1}^{d}x_{2},...,x_{1}^{d}x_{m}). Then:

  1. (1)

    ϵ(ϕ)=1dm1\epsilon_{\star}(\phi)=\frac{1}{dm-1}.

  2. (2)

    lct(J)=1d\operatorname{lct}_{\mathbb{C}}(J)=\frac{1}{d}, where J=x1d,x1dx2,,x1dxm=x1dJ=\langle x_{1}^{d},x_{1}^{d}x_{2},...,x_{1}^{d}x_{m}\rangle=\langle x_{1}^{d}\rangle.

  3. (3)

    dmk(ϕ)dm+1dm\leq k_{\star}(\phi)\leq dm+1.

  4. (4)

    δ(ϕ)=1d\delta_{\star}(\phi)=\frac{1}{d}.

Item (1) follows from Theorem 1.1. Young’s inequality (see 1.3) shows that k(ϕ)dm+1k_{\star}(\phi)\leq dm+1, on the other hand we have lct(J)=1d\operatorname{lct}_{\mathbb{C}}(J)=\frac{1}{d} which implies that at least dmdm convolutions are needed to obtain the (FRS) property (see [GH24, Lemmas 3.23 and 3.26]), hence k(ϕ)dmk_{\star}(\phi)\geq dm. Note that for any a1,..,ama_{1},..,a_{m}\in\mathbb{C} with a=1\left\|a\right\|_{\mathbb{C}}=1 we have lct(a1x1d+a2x1dx2++amx1dxm)=1d\operatorname{lct}_{\mathbb{C}}(a_{1}x_{1}^{d}+a_{2}x_{1}^{d}x_{2}+...+a_{m}x_{1}^{d}x_{m})=\frac{1}{d} so δ(ϕ)=1d\delta_{\star}(\phi)=\frac{1}{d}, by Lemma 3.15.

Remark 3.17.

One can replace ϕ\phi in Example 3.16, with

ϕ(x1,,xm)=(x1,x1dx2,,x1dxm).\phi(x_{1},...,x_{m})=(x_{1},x_{1}^{d}x_{2},...,x_{1}^{d}x_{m}).

Here, we get ϵ(ϕ)=1dmd\epsilon_{\star}(\phi)=\frac{1}{dm-d}, which is similar to Example 3.16 when mm is large, while lct(J)=1\operatorname{lct}_{\mathbb{C}}(J)=1, where J=x1,x1dx2,,x1dxm=x1J=\langle x_{1},x_{1}^{d}x_{2},...,x_{1}^{d}x_{m}\rangle=\langle x_{1}\rangle, which is much larger than in Example 3.16.

The following example shows that a reverse Young inequality (Theorem 1.3) does not hold in dimension m2m\geq 2.

Example 3.18.

Consider ϕ(x,y)=(x,x2(1+y1000))\phi(x,y)=(x,x^{2}(1+y^{1000})). Then ϵ(ϕ)=1999\epsilon_{\star}(\phi)=\frac{1}{999}, while δ(ϕ)=12\delta_{\star}(\phi)=\frac{1}{2}, and consequently k(ϕ)4k_{\star}(\phi)\leq 4.

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 ϕ:XY\phi:X\to Y between two FF-analytic manifolds, the Jacobian ideal sheaf 𝒥ϕ\mathcal{J}_{\phi} is defined locally as the ideal in the algebra of analytic functions on XX generated by the m×mm\times m-minors of the differential dx(ϕ)d_{x}(\phi) of ϕ\phi.

Proposition 4.1.

Let X,YX,Y be FF-analytic manifolds, dimX=dimY=n\dim X=\dim Y=n, and let ϕ:XY\phi:X\to Y be a locally dominant analytic map. Then for any x0Xx_{0}\in X,

(4.1) ϵ(ϕ;x0)=lctF(𝒥ϕ;x0).\epsilon_{\star}(\phi;x_{0})=\operatorname{lct}_{F}(\mathcal{J}_{\phi};x_{0}).
Proof.

We may assume that X,YX,Y are compact balls in FnF^{n}, and x0=0x_{0}=0. Since dimX=dimY=n\dim X=\dim Y=n, we have 𝒥ϕ=Jacx(ϕ)\mathcal{J}_{\phi}=\langle\operatorname{Jac}_{x}(\phi)\rangle, where Jacx(ϕ)=det(dxϕ)\operatorname{Jac}_{x}(\phi)=\det\,(d_{x}\phi). Since ϕ\phi is analytic and XX is compact, there is an open dense set UXU\subseteq X and MM\in\mathbb{N} such that #{ϕ1ϕ(x)}M\#\left\{\phi^{-1}\phi(x)\right\}\leq M and Jacx(ϕ)0\operatorname{Jac}_{x}(\phi)\neq 0, for every xUx\in U. We choose a disjoint cover iUi\bigcup_{i\in\mathbb{N}}U_{i} of UU by locally closed subsets222Recall that a set is locally closed if it is the difference of two open sets. UiU_{i} such that ϕ|Ui\phi|_{U_{i}} is a diffeomorphism. Write μ:=μFn|X\mu:=\mu_{F}^{n}|_{X}, μi:=μFn|Ui\mu_{i}:=\mu_{F}^{n}|_{U_{i}}, and note that ϕμ=g(y)μFn\phi_{*}\mu=g(y)\cdot\mu_{F}^{n} and ϕμi=gi(y)μFn\phi_{*}\mu_{i}=g_{i}(y)\cdot\mu_{F}^{n}, where

g(y)=xϕ1(y)|Jacx(ϕ)|F1 and gi(y)=|Jacϕ1(y)(ϕ)|F1.g(y)=\sum_{x\in\phi^{-1}(y)}\left|\operatorname{Jac}_{x}(\phi)\right|_{F}^{-1}\text{ and }g_{i}(y)=\left|\operatorname{Jac}_{\phi^{-1}(y)}(\phi)\right|_{F}^{-1}.

We have

(4.2) Yg(y)1+s𝑑y=ϕ(U)g(y)s𝑑ϕμ=Ugϕ(x)s𝑑μU1|Jacx(ϕ)|Fs𝑑μ.\int_{Y}g(y)^{1+s}dy=\int_{\phi(U)}g(y)^{s}d\phi_{*}\mu=\int_{U}g\circ\phi(x)^{s}d\mu\geq\int_{U}\frac{1}{\left|\operatorname{Jac}_{x}(\phi)\right|_{F}^{s}}d\mu.

On the other hand, since #{iI:yϕ(Ui)}M\#\left\{i\in I:y\in\phi(U_{i})\right\}\leq M for each yYy\in Y, using Jensen’s inequality, we have:

Yg(y)1+s𝑑y\displaystyle\int_{Y}g(y)^{1+s}dy Y(iI:yϕ(Ui)gi(y))1+s𝑑yMsYiI:yϕ(Ui)gi(y)1+sdy\displaystyle\leq\int_{Y}\left(\sum_{i\in I:y\in\phi(U_{i})}g_{i}(y)\right)^{1+s}dy\leq M^{s}\int_{Y}\sum_{i\in I:y\in\phi(U_{i})}g_{i}(y)^{1+s}dy
(4.3) =MsiIYgi(y)1+s𝑑y=MsiIUi1|Jacx(ϕ)|Fs𝑑μ=MsU1|Jacx(ϕ)|Fs𝑑μ,\displaystyle=M^{s}\sum_{i\in I}\int_{Y}g_{i}(y)^{1+s}dy=M^{s}\sum_{i\in I}\int_{U_{i}}\frac{1}{\left|\operatorname{Jac}_{x}(\phi)\right|_{F}^{s}}d\mu=M^{s}\int_{U}\frac{1}{\left|\operatorname{Jac}_{x}(\phi)\right|_{F}^{s}}d\mu,

which implies the proposition. ∎

Proof of Theorem 1.1.

Let ϕ:XY\phi:X\to Y be a locally dominant analytic map, and let x0Xx_{0}\in X. Since the claim is local, we may assume that XFnX\subseteq F^{n} is an open subset, and Y=FmY=F^{m}, with nmn\geq m. Let BB be a small ball around x0x_{0}. For each subset I{1,,n}I\subseteq\{1,...,n\} of size mm, let MIM_{I} be the corresponding m×mm\times m-minor of dxϕd_{x}\phi, and set

VI:={xB:maxI{1,,n}|MI(x)|F=|MI(x)|F}.V_{I}:=\left\{x\in B:\max_{I^{\prime}\subseteq\{1,...,n\}}\left|M_{I^{\prime}}(x)\right|_{F}=\left|M_{I}(x)\right|_{F}\right\}.

If VIV_{I}\neq\varnothing, the map ϕI(x):=(ϕ(x),xj1,,xjnm)\phi_{I}(x):=(\phi(x),x_{j_{1}},...,x_{j_{n-m}}) is locally dominant, where Jacx(ϕI)=MI(x)\operatorname{Jac}_{x}(\phi_{I})=M_{I}(x), and {j1,,jnm}={1,,n}\I\{j_{1},...,j_{n-m}\}=\{1,...,n\}\backslash I. For each VIV_{I} of positive measure, let μB:=μFn|B\mu_{B}:=\mu_{F}^{n}|_{B}, μI:=μFn|BVI\mu_{I}:=\mu_{F}^{n}|_{B\cap V_{I}} and write

ϕμB=g(y)μFm,ϕμI=gI(y)μFm and (ϕI)μI=g~I(z)μFn.\phi_{*}\mu_{B}=g(y)\cdot\mu_{F}^{m},\,\,\,\phi_{*}\mu_{I}=g_{I}(y)\cdot\mu_{F}^{m}\text{ and }\left(\phi_{I}\right)_{*}\mu_{I}=\widetilde{g}_{I}(z)\cdot\mu_{F}^{n}.

Let B~\widetilde{B} be a large ball in FnmF^{n-m} which contains the projection of ϕI(BVI)\phi_{I}(B\cap V_{I}) from FnF^{n} to the last nmn-m coordinates FnmF^{n-m}. Since ϕ=πIϕI\phi=\pi_{I}\circ\phi_{I} where πI:FnFm\pi_{I}:F^{n}\rightarrow F^{m} is a projection to the first mm coordinates, we have

gI(y)=Fnmg~I(y,zm+1,,zn)𝑑zm+1𝑑zn=B~g~I(y,zm+1,,zn)𝑑zm+1𝑑zn.g_{I}(y)=\int_{F^{n-m}}\widetilde{g}_{I}(y,z_{m+1},...,z_{n})dz_{m+1}...dz_{n}=\int_{\widetilde{B}}\widetilde{g}_{I}(y,z_{m+1},...,z_{n})dz_{m+1}...dz_{n}.

By Jensen’s inequality, we have

FmgI(y)1+s𝑑y=\displaystyle\int_{F^{m}}g_{I}(y)^{1+s}dy= FmμFnm(B~)1+s(1μFnm(B~)B~g~I(y,zm+1,,zn)𝑑zm+1𝑑zn)1+s𝑑y\displaystyle\int_{F^{m}}\mu_{F}^{n-m}(\widetilde{B})^{1+s}\left(\frac{1}{\mu_{F}^{n-m}(\widetilde{B})}\int_{\widetilde{B}}\widetilde{g}_{I}(y,z_{m+1},...,z_{n})dz_{m+1}...dz_{n}\right)^{1+s}dy
\displaystyle\leq μFnm(B~)sFm×B~g~I(y,zm+1,,zn)1+s𝑑y𝑑zm+1𝑑znFng~I(z)1+s𝑑z,\displaystyle\mu_{F}^{n-m}(\widetilde{B})^{s}\cdot\int_{F^{m}\times\widetilde{B}}\widetilde{g}_{I}(y,z_{m+1},...,z_{n})^{1+s}dydz_{m+1}...dz_{n}\lesssim\int_{F^{n}}\widetilde{g}_{I}(z)^{1+s}dz,

for every s>0s>0. Since IVI\bigcup_{I}V_{I} is of full measure in BB, and using Proposition 4.1 and (1.5), we have:

Fmg(y)1+s𝑑y\displaystyle\int_{F^{m}}g(y)^{1+s}dy IFmgI(y)1+s𝑑yIFng~I(z)1+s𝑑z\displaystyle\lesssim\sum_{I}\int_{F^{m}}g_{I}(y)^{1+s}dy\lesssim\sum_{I}\int_{F^{n}}\widetilde{g}_{I}(z)^{1+s}dz
IVI|MI(x)|Fs𝑑xBminI{1,,n}[|MI(x)|Fs]𝑑x<,\displaystyle\lesssim\sum_{I}\int_{V_{I}}\left|M_{I}(x)\right|_{F}^{-s}dx\lesssim\int_{B}\min_{I\subseteq\{1,...,n\}}[\left|M_{I}(x)\right|_{F}^{-s}]dx<\infty,

for every s<lctF(𝒥ϕ;x0)s<\operatorname{lct}_{F}(\mathcal{J}_{\phi};x_{0}), as required. ∎

5. Proof of Theorem 1.5 – an upper bound over \mathbb{C}

In this section we prove Theorem 1.5. We use the following easy consequence of the coarea formula (see [Fed69]).

Lemma 5.1.

Let BnB\subseteq\mathbb{C}^{n} be a compact ball, and let ϕ:Bm\phi:B\to\mathbb{C}^{m} be a dominant analytic map. Let μB:=μn|B\mu_{B}:=\mu_{\mathbb{C}}^{n}|_{B} and write ϕμB=g(y)μm\phi_{*}\mu_{B}=g(y)\cdot\mu_{\mathbb{C}}^{m}. Then:

g(y)=ϕ1(y)dH2(nm)(x)det((dxϕ)(dxϕ)),g(y)=\int_{\phi^{-1}(y)}\frac{dH_{2(n-m)}(x)}{\det\left((d_{x}\phi)^{*}(d_{x}\phi)\right)},

where the integral is taken with respect to the 2(nm)2(n-m)-dimensional Hausdorff measure (recall §\mathsection1.4(7)).

Let ϕ:XY\phi:X\rightarrow Y be a locally dominant analytic map between complex analytic manifolds, and let x0Xx_{0}\in X. Since the claim is local, we may assume that Y=mY=\mathbb{C}^{m} and X=BX=B is a ball in n\mathbb{C}^{n}. To show (1.12), it is enough to bound ϵ(ϕμB)\epsilon_{\star}(\phi_{*}\mu_{B}), where BB is of arbitrarily small radius around x0x_{0}. Denote G(x):=det((dxϕ)(dxϕ))G(x):=\det\left((d_{x}\phi)^{*}(d_{x}\phi)\right). By Lemma 5.1, we have:

g(y)=ϕ1(y)dH2(mn)(x)G(x)g(y)=\int_{\phi^{-1}(y)}\frac{dH_{2(m-n)}(x)}{G(x)}

whence

(5.1) Yg(y)1+ϵ𝑑y=X[ϕ1(ϕ(x))dH2(mn)(x)G(x)]ϵ𝑑x.\int_{Y}g(y)^{1+\epsilon}dy=\int_{X}\left[\int_{\phi^{-1}(\phi(x))}\frac{dH_{2(m-n)}(x^{\prime})}{G(x^{\prime})}\right]^{\epsilon}dx.

We apply Theorem 2.1 to the ideal 𝒥ϕ\mathcal{J}_{\phi}. Let π:X~B\pi:\widetilde{X}\to B be the corresponding resolution. Without loss of generality, we can assume that X~m\widetilde{X}\subseteq\mathbb{C}^{m} is an open subset, so that

detdx~π=v(x~)x~1b1x~nbn,v(0)0,\det d_{\widetilde{x}}\pi=v(\widetilde{x})\widetilde{x}_{1}^{b_{1}}\cdots\widetilde{x}_{n}^{b_{n}}~{},\quad v(0)\neq 0,

and each of the m×mm\times m minors MIM_{I} of dx~ϕd_{\widetilde{x}}\phi satisfies

MI(π(x~))=uI(x~)x~1a1x~nan,M_{I}(\pi(\widetilde{x}))=u_{I}(\widetilde{x})\widetilde{x}_{1}^{a_{1}}\cdots\widetilde{x}_{n}^{a_{n}},

where at least one of the functions uIu_{I} does not vanish at 0.

We first perform the change of variables x=π(x~)x=\pi(\widetilde{x}) in the external integral in (5.1), yielding

Yg(y)1+ϵ𝑑y=X~|detdx~π|[ϕ1(ϕ(π(x~)))dH2(mn)(x)G(x)]ϵ𝑑x~,\int_{Y}g(y)^{1+\epsilon}dy=\int_{\widetilde{X}}\left|\det d_{\widetilde{x}}\pi\right|_{\mathbb{C}}\left[\int_{\phi^{-1}(\phi(\pi(\widetilde{x})))}\frac{dH_{2(m-n)}(x^{\prime})}{G(x^{\prime})}\right]^{\epsilon}d\widetilde{x},

and then perform the change of variables x=π(t)x^{\prime}=\pi(t) in the internal integral. Since the map π\pi is continuously differentiable, and BB is compact, the product of the singular values of the restriction of dtπd_{t}\pi to any subspace is bounded from below in absolute value by a number times the absolute value of the determinant of dtπd_{t}\pi. Thus,

(5.2) Yg(y)1+ϵ𝑑yX~|detdx~π|[ϕ~1(ϕ~(x~))|detdtπ|dH2(mn)(t)G(π(t))]ϵ𝑑x~,\int_{Y}g(y)^{1+\epsilon}dy\gtrsim\int_{\widetilde{X}}\left|\det d_{\widetilde{x}}\pi\right|_{\mathbb{C}}\left[\int_{\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))}\frac{\left|\det d_{t}\pi\right|_{\mathbb{C}}\,dH_{2(m-n)}(t)}{G(\pi(t))}\right]^{\epsilon}d\widetilde{x},

where ϕ~:=ϕπ\widetilde{\phi}:=\phi\circ\pi. By the Cauchy–Binet formula, the right-hand side is equal to

X~|detdx~π|[ϕ~1(ϕ~(x~))|v(t)||t1|b1|tn|bndH2(mn)(t)I|uI(t)||t1|a1|tn|an]ϵ𝑑x~.\int_{\widetilde{X}}\left|\det d_{\widetilde{x}}\pi\right|_{\mathbb{C}}\left[\int_{\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))}\frac{\left|v(t)\right|_{\mathbb{C}}\left|t_{1}\right|_{\mathbb{C}}^{b_{1}}\cdot...\cdot\left|t_{n}\right|_{\mathbb{C}}^{b_{n}}\,dH_{2(m-n)}(t)}{\sum_{I}\left|u_{I}(t)\right|_{\mathbb{C}}\left|t_{1}\right|_{\mathbb{C}}^{a_{1}}\cdot...\cdot\left|t_{n}\right|_{\mathbb{C}}^{a_{n}}}\right]^{\epsilon}d\widetilde{x}.

We bound the internal integral from below as follows. For each x~X~\widetilde{x}\in\widetilde{X}, set

Q(x~):={tX~:j=1n|tjx~j||x~j|14}Q(\widetilde{x}):=\left\{t\in\widetilde{X}\,\,:\,\,\sum_{j=1}^{n}\frac{\left|t_{j}-\widetilde{x}_{j}\right|_{\mathbb{C}}}{\left|\widetilde{x}_{j}\right|_{\mathbb{C}}}\leq\frac{1}{4}\right\}

Then

ϕ~1(ϕ~(x~))|v(t)||t1|b1|tn|bndH2(mn)(t)I|uI(t)||t1|a1|tn|an\displaystyle\int_{\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))}\frac{\left|v(t)\right|_{\mathbb{C}}\left|t_{1}\right|_{\mathbb{C}}^{b_{1}}\cdot...\cdot\left|t_{n}\right|_{\mathbb{C}}^{b_{n}}\,dH_{2(m-n)}(t)}{\sum_{I}\left|u_{I}(t)\right|_{\mathbb{C}}\left|t_{1}\right|_{\mathbb{C}}^{a_{1}}\cdot...\cdot\left|t_{n}\right|_{\mathbb{C}}^{a_{n}}}
\displaystyle\gtrsim ϕ~1(ϕ~(x~))Q(x~)|t1|b1a1|tn|bnan𝑑H2(mn)(t)\displaystyle\int_{\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))\cap Q(\widetilde{x})}\left|t_{1}\right|_{\mathbb{C}}^{b_{1}-a_{1}}\cdot...\cdot\left|t_{n}\right|_{\mathbb{C}}^{b_{n}-a_{n}}\,dH_{2(m-n)}(t)
(5.3) \displaystyle\gtrsim |x~1|b1a1|x~n|bnanH2(nm)(ϕ~1(ϕ~(x~))Q(x~)).\displaystyle\left|\widetilde{x}_{1}\right|_{\mathbb{C}}^{b_{1}-a_{1}}\cdot...\cdot\left|\widetilde{x}_{n}\right|_{\mathbb{C}}^{b_{n}-a_{n}}\,H_{2(n-m)}(\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))\cap Q(\widetilde{x})).
Claim 5.2.

For any x~X~\widetilde{x}\in\widetilde{X},

H2(nm)(ϕ~1(ϕ~(x~))Q(x~))min|I|=nmiI|x~i|.H_{2(n-m)}(\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))\cap Q(\widetilde{x}))\gtrsim\min_{|I|=n-m}\prod_{i\in I}\left|\widetilde{x}_{i}\right|_{\mathbb{C}}.
Proof of Claim 5.2.

Let T:X~X~T:\widetilde{X}\to\widetilde{X} be the affine map given by T(x~)j=x~j(1+x~j)T(\widetilde{x}^{\prime})_{j}=\widetilde{x}_{j}(1+\widetilde{x}^{\prime}_{j}). Then Q(x~)=TB12Q(\widetilde{x})=TB_{\frac{1}{2}}, where B12B_{\frac{1}{2}} is a ball of radius 12\frac{1}{2} centered at the origin. By a theorem of Lelong [Lel57] (see also [Thi67, LG86]), for any analytic set MM of pure dimension nmn-m, and any r>0r>0, one has

H2(nm)(MBr)H2(nm)(M0Br)limρ+0H2(nm)(MBρ)H2(nm)(M0Bρ),\frac{H_{2(n-m)}(M\cap B_{r})}{H_{2(n-m)}(M_{0}\cap B_{r})}\geq\lim_{\rho\to+0}\frac{H_{2(n-m)}(M\cap B_{\rho})}{H_{2(n-m)}(M_{0}\cap B_{\rho})},

where M0M_{0} is a linear subspace of the same dimension nmn-m. The limit on the right-hand side is the Lelong number of MM, which is the algebraic multiplicity of MM at x~\widetilde{x}; it is strictly positive; thus

H2(nm)(MBr)r2(nm).H_{2(n-m)}(M\cap B_{r})\gtrsim r^{2(n-m)}.

Applying this estimate to M=T1(ϕ~1(ϕ~(x~)))M=T^{-1}(\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))) and observing that

H2(nm)(TA)min|I|=nmiI|x~i|H2(nm)(A),H_{2(n-m)}(TA)\geq\min_{|I|=n-m}\prod_{i\in I}\left|\widetilde{x}_{i}\right|_{\mathbb{C}}\cdot H_{2(n-m)}(A),

for any Borel set AX~A\subset\widetilde{X}, we obtain the claimed assertion. ∎

We now proceed with the proof of the theorem. Using Claim 5.2, we deduce

ϕ~1(ϕ~(x~))|detdtπ|dH2(mn)(t)G(π(t))i=1n|x~i|biai+1,\int_{\widetilde{\phi}^{-1}(\widetilde{\phi}(\widetilde{x}))}\frac{\left|\det d_{t}\pi\right|_{\mathbb{C}}\,dH_{2(m-n)}(t)}{G(\pi(t))}\gtrsim\prod_{i=1}^{n}\left|\widetilde{x}_{i}\right|_{\mathbb{C}}^{b_{i}-a_{i}+1},

whence,

Yg(y)1+ϵ𝑑yX~i=1n|x~i|bi+ϵ(biai+1)dx~.\int_{Y}g(y)^{1+\epsilon}dy\gtrsim\int_{\widetilde{X}}\prod_{i=1}^{n}\left|\widetilde{x}_{i}\right|_{\mathbb{C}}^{b_{i}+\epsilon(b_{i}-a_{i}+1)}d\widetilde{x}.

This integral diverges whenever there is an index ii such that

bi+ϵ(biai+1)1,b_{i}+\epsilon(b_{i}-a_{i}+1)\leq-1,

i.e.

ϵ(ϕ;x)mini{bi+1aibi1:bi+1<ai}.\epsilon_{\star}(\phi;x)\leq\min_{i}\left\{\frac{b_{i}+1}{a_{i}-b_{i}-1}\,\,:\,\,b_{i}+1<a_{i}\right\}.

On the other hand,

lct(𝒥ϕ;x0)=mini{bi+1ai}\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x_{0})=\min_{i}\left\{\frac{b_{i}+1}{a_{i}}\right\}

and under the assumption that this quantity is strictly less than 11 we have an index ii such that

lct(𝒥ϕ;x0)=bi+1ai<1.\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x_{0})=\frac{b_{i}+1}{a_{i}}<1.

For this index,

bi+1aibi1=lct(𝒥ϕ;x0)1lct(𝒥ϕ;x0).\frac{b_{i}+1}{a_{i}-b_{i}-1}=\frac{\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x_{0})}{1-\operatorname{lct}_{\mathbb{C}}(\mathcal{J}_{\phi};x_{0})}.

This concludes the proof of Theorem 1.5.

6. Geometric characterization of ϵ=\epsilon_{\star}=\infty

From now on let KK be a number field and let 𝒪K\mathcal{O}_{K} be its ring of integers. Let Loc0\operatorname{Loc}_{0} be the collection of all non-Archimedean local fields FF which contain KK. We use the notation Loc0,\operatorname{Loc}_{0,\gg}, for the collection of FLoc0F\in\operatorname{Loc}_{0} with large enough residual characteristic, depending on some given data. Throughout this section, we denote by B(y,r)B(y,r) a ball of radius rr centered at yY(F)y\in Y(F).

In this section we focus on algebraic morphisms φ:XY\varphi:X\rightarrow Y between algebraic KK-varieties. We would like to characterize morphisms φ\varphi where ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty for certain collections of local fields FF, in terms of the singularities of φ\varphi. In order to effectively do this, it is necessary to consider an “algebraically closed” collection of local fields, such as the following:

  • {}\{\mathbb{C}\}.

  • {F}FLoc0,\{F\}_{F\in\operatorname{Loc}_{0,\gg}}, or {F}FLoc0\{F\}_{F\in\operatorname{Loc}_{0}}.

Aizenbud and Avni have shown the following characterization of the (FRS) property:

Theorem 6.1 ([AA16, Theorem 3.4]).

Let φ:XY\varphi:X\rightarrow Y be a map between smooth KK-varieties. Then φ\varphi is (FRS) if and only if for each FLoc0F\in\operatorname{Loc}_{0} and every μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)), one has (φF)(μ)c,(Y(F))\left(\varphi_{F}\right)_{*}(\mu)\in\mathcal{M}_{c,\infty}(Y(F)).

The Archimedean counterpart of this theorem was studied in [Rei], where it was shown that given an (FRS) morphism φ\varphi, then φ\varphi_{\mathbb{\mathbb{R}}} and φ\varphi_{\mathbb{C}} are LL^{\infty}-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 \mathbb{C} is algebraically closed. We arrive at the following characterization of LL^{\infty}-morphisms.

Corollary 6.2.

Let φ:XY\varphi:X\rightarrow Y be a map between smooth KK-varieties. Then the following are equivalent:

  1. (1)

    φ\varphi is (FRS).

  2. (2)

    For every local field FF containing KK, the map φF\varphi_{F} is an LL^{\infty}-morphism.

  3. (3)

    For each FLoc0,F\in\operatorname{Loc}_{0,\gg}, the map φF\varphi_{F} is an LL^{\infty}-map.

  4. (4)

    The map φ\varphi_{\mathbb{C}} is an LL^{\infty}-map.

Our goal is to characterize the weaker property that ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty over F=F=\mathbb{C}, or over all FLoc0F\in\operatorname{Loc}_{0} (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 φ:𝔸2𝔸1\varphi:\mathbb{A}_{\mathbb{Q}}^{2}\rightarrow\mathbb{A}_{\mathbb{Q}}^{1} be the map φ(x,y)=xy\varphi(x,y)=xy. Then:

φp1(B(0,pk))p2\displaystyle\varphi_{\mathbb{Q}_{p}}^{-1}(B(0,p^{-k}))\cap\mathbb{Z}_{p}^{2} ={(x,y)p2:val(x)+val(y)k}\displaystyle=\left\{(x,y)\in\mathbb{Z}_{p}^{2}:\mathbb{\mathrm{val}}(x)+\mathbb{\mathrm{val}}(y)\geq k\right\}
=rk,0lr{(x,y)p2:val(x)=l,val(y)=rl}.\displaystyle=\bigcup_{r\geq k,0\leq l\leq r}\left\{(x,y)\in\mathbb{Z}_{p}^{2}:\mathbb{\mathrm{val}}(x)=l,\mathbb{\mathrm{val}}(y)=r-l\right\}.

In particular, we have

φp(μp2)(B(0,pk))μ𝒪F(B(0,pk))\displaystyle\frac{\varphi_{\mathbb{Q}_{p}*}(\mu_{\mathbb{Z}_{p}}^{2})(B(0,p^{-k}))}{\mu_{\mathcal{O}_{F}}(B(0,p^{-k}))} =pkr=kl=0r(p1p)2plp(rl)\displaystyle=p^{k}\cdot\sum_{r=k}^{\infty}\sum_{l=0}^{r}\left(\frac{p-1}{p}\right)^{2}p^{-l}p^{-(r-l)}
=pk(p1p)2r=k(r+1)pr(p1p)2(k+1).\displaystyle=p^{k}\cdot\left(\frac{p-1}{p}\right)^{2}\sum_{r=k}^{\infty}(r+1)p^{-r}\geq\left(\frac{p-1}{p}\right)^{2}(k+1).

Hence, the measure φp(μp2)\varphi_{\mathbb{Q}_{p}*}(\mu_{\mathbb{Z}_{p}}^{2}) does not have bounded density. On the other hand, since lctp(φp;0)=1\operatorname{lct}_{\mathbb{Q}_{p}}(\varphi_{\mathbb{Q}_{p}};0)=1, and by considering the asymptotic expansion of φp(μp2)\varphi_{\mathbb{Q}_{p}*}(\mu_{\mathbb{Z}_{p}}^{2}) as in Theorem 3.6, one sees:

  1. (1)

    φp(μp2)\varphi_{\mathbb{Q}_{p}*}(\mu_{\mathbb{Z}_{p}}^{2}) explodes logarithmically around 0, i.e. the density of φp(μp2)\varphi_{\mathbb{Q}_{p}*}(\mu_{\mathbb{Z}_{p}}^{2}) behaves like val(t)\mathbb{\mathrm{val}}(t), around 0.

  2. (2)

    ϵ(φp)=\epsilon_{\star}(\varphi_{\mathbb{Q}_{p}})=\infty for every prime pp.

By Corollary 6.2, φ\varphi cannot be (FRS), and indeed {xy=0}\left\{xy=0\right\} 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 X𝔸KnX\subseteq\mathbb{A}_{K}^{n} be an affine KK-scheme whose coordinate ring is

K[x1,,xn]/(f1,,fk).K[x_{1},\dots,x_{n}]/(f_{1},\dots,f_{k}).

Then the mm-th jet scheme Jm(X)J_{m}(X) of XX is the affine scheme with the following coordinate ring:

K[x1,,xn,x1(1),,xn(1),,x1(m),,xn(m)]/({fj(u)}j=1,u=1k,m),K[x_{1},\dots,x_{n},x_{1}^{(1)},\dots,x_{n}^{(1)},\dots,x_{1}^{(m)},\dots,x_{n}^{(m)}]/(\{f_{j}^{(u)}\}_{j=1,u=1}^{k,m}),

where fi(u)f_{i}^{(u)} is the uu-th formal derivative of fif_{i}.

Let φ:𝔸n1𝔸n2\varphi:\mathbb{A}^{n_{1}}\rightarrow\mathbb{A}^{n_{2}} be a morphism between affine spaces. Then the mm-th jet morphism Jm(φ):𝔸n1(m+1)𝔸n2(m+1)J_{m}(\varphi):\mathbb{A}^{n_{1}(m+1)}\rightarrow\mathbb{A}^{n_{2}(m+1)} of φ\varphi is given by formally deriving φ\varphi, Jm(φ)=(φ,φ(1),,φ(m))J_{m}(\varphi)=(\varphi,\varphi^{(1)},\dots,\varphi^{(m)}). Similarly, the mm-th jet Jm(φ)J_{m}(\varphi) of a morphism φ:XY\varphi:X\rightarrow Y of affine KK-schemes, is given by the formal derivative of φ\varphi. Both Jm(X)J_{m}(X) and Jm(φ)J_{m}(\varphi) can be generalized to arbitrary KK-schemes and KK-morphisms (see [CLNS18, Chapter 3] and [EM09] for more details).

Given a subscheme ZXZ\subseteq X of a smooth variety XX, with ZZ defined by an ideal JJ, we denote by lct(X,Z):=lct(J)\operatorname{lct}(X,Z):=\operatorname{lct}(J) the log-canonical threshold of the pair (X,Z)(X,Z). Mustaţă showed that the log-canonical threshold lct(X,Z)\operatorname{lct}(X,Z) can be characterized in terms of the growth rate of the dimensions of the jet schemes of ZZ:

Theorem 6.3 ([Mus02, Corollary 0.2], [CLNS18, Corollary 7.2.4.2]).

Let XX be a smooth, geometrically irreducible KK-variety, and let ZXZ\subsetneq X be a closed subscheme. Then

lct(X,Z)=dimXsupm0dimJm(Z)m+1.\operatorname{lct}(X,Z)=\dim X-\sup_{m\geq 0}\frac{\dim J_{m}(Z)}{m+1}.

Furthermore, the supremum is achieved for mm divisible enough. In particular, lct(X,Z)\operatorname{lct}(X,Z) is a rational number.

Note that lct(X,Z)\operatorname{lct}(X,Z) depends on ZZ and dimX\dim X, and neither on the ambient space XX not on the embedding of ZZ in XX.

We now introduce the following definitions from [GH24]. For a morphism φ:XY\varphi:X\rightarrow Y between schemes, we denote by Xy,φX_{y,\varphi} the scheme theoretic fiber of φ\varphi over yYy\in Y.

Definition 6.4.

Let φ:XY\varphi:X\rightarrow Y be a morphism of smooth, geometrically irreducible KK-varieties, and let ϵ>0\epsilon>0.

  1. (1)

    φ\varphi is called ϵ\epsilon-flat if for every xXx\in X we have dimXφ(x),φdimXϵdimY\dim X_{\varphi(x),\varphi}\leq\dim X-\epsilon\dim Y.

  2. (2)

    φ\varphi is called ϵ\epsilon-jet flat if Jm(φ):Jm(X)Jm(Y)J_{m}(\varphi):J_{m}(X)\rightarrow J_{m}(Y) is ϵ\epsilon-flat for every mm\in\mathbb{N}.

  3. (3)

    φ\varphi is called jet-flat if it is 11-jet flat.

In particular, note that φ\varphi is flat if and only if it is 11-flat.

Note that by Theorem 6.3, φ\varphi is ϵ\epsilon-jet-flat if and only if lct(X,Xφ(x),φ)ϵdimY\operatorname{lct}(X,X_{\varphi(x),\varphi})\geq\epsilon\dim Y for all xXx\in X. 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 φ:XY\varphi:X\rightarrow Y be a morphism of smooth KK-varieties. Then:

  1. (1)

    φ\varphi is (FRS) if and only if Jm(φ)J_{m}(\varphi) is flat with locally integral fibers, for every m0m\geq 0 (in particular, φ\varphi is jet-flat).

  2. (2)

    φ\varphi is jet-flat if and only if φ\varphi 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], φ\varphi is jet-flat if and only if Jm(φ)J_{m}(\varphi) is flat over YJm(Y)Y\subseteq J_{m}(Y). 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 xXx\in X, and every mm\in\mathbb{N}, the scheme Jm(Xφ(x),φ)J_{m}(X_{\varphi(x),\varphi}) is a local complete intersection. By [Ish18, Corollary 10.2.9] and [EI15, Corollary 3.17], this is equivalent to the condition that Xφ(x),φX_{\varphi(x),\varphi} has semi-log-canonical singularities, for every xXx\in X. ∎

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 φ:XY\varphi:X\rightarrow Y be a map between smooth KK-varieties. Then the following are equivalent:

  1. (1)

    φ\varphi is jet-flat.

  2. (2)

    For every local field FF containing KK, we have ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty.

  3. (3)

    For every FLoc0,F\in\operatorname{Loc}_{0,\gg} we have ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty.

  4. (4)

    We have ϵ(φ)=\epsilon_{\star}(\varphi_{\mathbb{C}})=\infty.

The proof of Theorem 6.6 is done by showing both implications (1)(2)(3)(1)(1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1) and (1)(2)(4)(1)(1)\Rightarrow(2)\Rightarrow(4)\Rightarrow(1). The implications (2)(3)(2)\Rightarrow(3) and (2)(4)(2)\Rightarrow(4) are immediate. We first prove (3)(1)(3)\Rightarrow(1) and (4)(1)(4)\Rightarrow(1). Then we will prove (1)(2)(1)\Rightarrow(2) in the non-Archimedean case in §\mathsection6.2, and the Archimedean case in §\mathsection6.3.

Proposition 6.7.

Let φ:XY\varphi:X\rightarrow Y be a map between smooth KK-varieties. Assume that either ϵ(φ)=\epsilon_{\star}(\varphi_{\mathbb{C}})=\infty or ϵ(φF)=\epsilon_{\star}(\varphi_{F})=\infty for all FLoc0,F\in\operatorname{Loc}_{0,\gg}. Then φ\varphi is jet-flat.

Proof.

Working locally, and composing φ\varphi with an étale map Φ:Y𝔸m\Phi:Y\rightarrow\mathbb{A}^{m}, we may assume that Y=𝔸mY=\mathbb{A}^{m}. Let ψ:𝔸m𝔸m\psi:\mathbb{A}^{m}\rightarrow\mathbb{A}^{m} be any dominant morphism, and let μ1c(X(F))\mu_{1}\in\mathcal{M}_{c}^{\infty}(X(F)) and μ2c(Fm)\mu_{2}\in\mathcal{M}_{c}^{\infty}(F^{m}). By Theorem 1.1, for all FLoc{}F\in\operatorname{Loc}_{\gg}\cup\left\{\mathbb{C}\right\}, we have ψ(μ2)L1+ϵ\psi_{*}(\mu_{2})\in L^{1+\epsilon} for some ϵ>0\epsilon>0. Taking qq large enough, by Young’s convolution inequality, one has:

(φψ)(μ1×μ2)=φ(μ1)ψ(μ2)LqL1+ϵL,(\varphi*\psi)_{*}(\mu_{1}\times\mu_{2})=\varphi_{*}(\mu_{1})*\psi_{*}(\mu_{2})\subseteq L^{q}*L^{1+\epsilon}\subseteq L^{\infty},

where φψ\varphi*\psi is as in Definition 1.7. By Corollary 6.2, we get that φψ\varphi*\psi is (FRS). We now claim that since φ\varphi is a morphism whose convolution with any dominant morphism produces an (FRS) morphism, φ\varphi must be jet-flat.

Indeed, assume it is not the case. Then by Theorem 6.3 and Definition 6.4, there exist yK¯my\in\overline{K}^{m} and N>0N>0 such that the scheme theoretic fiber Xy,φX_{y,\varphi} of φ\varphi over yy satisfies

lct(X,Xy,φ)=dimXsupk0dimJkXy,φk+1m(12N).\operatorname{lct}(X,X_{y,\varphi})=\dim X-\sup_{k\geq 0}\frac{\dim J_{k}X_{y,\varphi}}{k+1}\leq m(1-\frac{2}{N}).

Moreover, this supremum is achieved for kk divisible enough. Thus the map

Jk(φ):Jk(X)Jk(Y)J_{k}(\varphi):J_{k}(X)\rightarrow J_{k}(Y)

is not (11N)\left(1-\frac{1}{N}\right)-flat for kk divisible enough. But on the other hand, the map

ψN(y1,,ym)=(y12N,,ym2N)\psi_{N}(y_{1},...,y_{m})=(y_{1}^{2N},...,y_{m}^{2N})

satisfies that Jl(ψN)J_{l}(\psi_{N}) is not 1N\frac{1}{N}-flat for divisible enough ll (since lct(yi2N)=12N\operatorname{lct}(y_{i}^{2N})=\frac{1}{2N}). Thus we may find k0k_{0}\in\mathbb{N} such that Jk0(φ)J_{k_{0}}(\varphi) is not (11N)\left(1-\frac{1}{N}\right)-flat and Jk0(ψN)J_{k_{0}}(\psi_{N}) is not 1N\frac{1}{N}-flat . But then Jk0(φψ)=Jk0(φ)Jk0(ψ)J_{k_{0}}(\varphi*\psi)=J_{k_{0}}(\varphi)*J_{k_{0}}(\psi) is not flat (see [GH24, Lemma 3.26]), which is a contradiction by Fact 6.5, as φψ\varphi*\psi is (FRS). ∎

6.2. (1)(2)(1)\Rightarrow(2): the non-Archimedean case

We now turn to the proof of (1)(2)(1)\Rightarrow(2), in the non-Archimedean case. We first prove the following variant of [CGH23, Theorem 4.12].

Proposition 6.8 ([CGH23, Theorem 4.12]).

Let φ:XY\varphi:X\rightarrow Y be a jet-flat map between smooth KK-varieties. Then there exists MM\in\mathbb{N}, such that for each FLoc0F\in\operatorname{Loc}_{0}, each μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)) and each non-vanishing τ(Y(F))\tau\in\mathcal{M}^{\infty}(Y(F)), one can find CF,μ,τ>0C_{F,\mu,\tau}>0 such that for each yY(F)y\in Y(F) and kk\in\mathbb{N} one has,

GF,μ(y,k):=(φμ)(B(y,qFk))τ(B(y,qFk))CF,μ,τkM.G_{F,\mu}(y,k):=\frac{(\varphi_{*}\mu)(B(y,q_{F}^{-k}))}{\tau(B(y,q_{F}^{-k}))}\leq C_{F,\mu,\tau}k^{M}.
Proof.

We may assume that Y=𝔸KmY=\mathbb{A}_{K}^{m} and that τ=μFm\tau=\mu_{F}^{m}. We may further assume that XX is affine, and thus embeds in 𝔸Kn\mathbb{A}_{K}^{n}. Let μ~\widetilde{\mu} be the canonical measure on X(F)X(F) (see [Ser81, Section 3.3], and also [CCL12, Section 1.2]). It is enough to consider measures μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)) which are of the form μl:=μ~|B(0,qFl)X(F)\mu_{l}:=\widetilde{\mu}|_{B(0,q_{F}^{l})\cap X(F)}.

Write gF(l,y)g_{F}(l,y) for the density of φμl\varphi_{*}\mu_{l} with respect to μFm\mu_{F}^{m}, and set GF(y,l,k):=GF,μl(y,k)G_{F}(y,l,k):=G_{F,\mu_{l}}(y,k). Then the collections {gF(l,y)}FLoc0\{g_{F}(l,y)\}_{F\in\operatorname{Loc}_{0}} and {GF(y,l,k)}FLoc0\{G_{F}(y,l,k)\}_{F\in\operatorname{Loc}_{0}} are both motivic functions, in the sense of [CGH18, Section 1.2]. By [CGH18, Theorem 2.1.3], there exists a motivic function H(l,k)={HF:×}FLoc0H(l,k)=\left\{H_{F}:\mathbb{Z}\times\mathbb{N}\rightarrow\mathbb{\mathbb{R}}\right\}_{F\in\operatorname{Loc}_{0}}, which approximates the supremum of G(y,l,k)G(y,l,k), that is:

(6.1) 1CFHF(l,k)supyY(F)GF(y,l,k)HF(l,k),\frac{1}{C_{F}}H_{F}(l,k)\leq\sup_{y\in Y(F)}G_{F}(y,l,k)\leq H_{F}(l,k),

for all FLoc0F\in\operatorname{Loc}_{0} and (l,k)×(l,k)\in\mathbb{Z}\times\mathbb{N}, where CFC_{F} depends only on the local field333In the statement of [CGH18, Theorem 2.1.3], the approximation (6.1) is stated for |GF(y,l,k)|\left|G_{F}(y,l,k)\right|_{\mathbb{C}} instead of GF(y,l,k)G_{F}(y,l,k). Since GF(y,l,k)G_{F}(y,l,k) 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 HH is motivic, and using [CGH18, Proposition 1.4.2], for each FLoc0F\in\operatorname{Loc}_{0}, we may divide ×\mathbb{Z}\times\mathbb{N} into a finite disjoint union ×=AF𝒜AF\mathbb{Z}\times\mathbb{N}=\bigsqcup_{A_{F}\in\mathcal{A}}A_{F}, with |𝒜|<\left|\mathcal{A}\right|<\infty independent of FF, such that on each part AF×A_{F}\subseteq\mathbb{Z}\times\mathbb{N}, the following hold. There exist finitely many aia_{i}\in\mathbb{N}, bib_{i}\in\mathbb{Q} independent of FF, a finite set ΛF\Lambda_{F} (of size depending on FF), and a finite partition of AFA_{F} into subsets AF=ξΛFAF,ξA_{F}=\bigsqcup_{\xi\in\Lambda_{F}}A_{F,\xi}, such that for all (k,l)AF,ξ(k,l)\in A_{F,\xi}:

HF(l,k)=i=1Lci(ξ,l,F)kaiqFbik,H_{F}(l,k)=\sum_{i=1}^{L}c_{i}(\xi,l,F)\cdot k^{a_{i}}q_{F}^{b_{i}k},

for some constants ci(ξ,l,F)c_{i}(\xi,l,F) depending on l,Fl,F and ξ\xi. Moreover, for fixed FLoc0F\in\operatorname{Loc}_{0} and ll\in\mathbb{N}, the set AF,ξ,l:={k:(k,l)AF,ξ}A_{F,\xi,l}:=\left\{k\in\mathbb{N}:(k,l)\in A_{F,\xi}\right\} is either finite or a fixed congruence class modulo some e1e\in\mathbb{Z}_{\geq 1}.

To prove the proposition, it is enough to show that for each FLoc0F\in\operatorname{Loc}_{0} and ll\in\mathbb{Z}, we have HF(l,k)CF,lkMH_{F}(l,k)\leq C_{F,l}k^{M} on each AF,ξ,lA_{F,\xi,l}, for some constant CF,lC_{F,l} depending on F,lF,l. It is enough to prove this for AF,ξ,lA_{F,\xi,l} infinite, as otherwise we have

HF(l,k)CF,l:=kAF,ξ,li=1L|ci(ξ,l,F)|kaiqFbik.H_{F}(l,k)\leq C_{F,l}:=\sum_{k\in A_{F,\xi,l}}\sum_{i=1}^{L}\left|c_{i}(\xi,l,F)\right|\cdot k^{a_{i}}q_{F}^{b_{i}k}.

Now suppose AF,ξ,lA_{F,\xi,l} is infinite. By rearranging the constants ci(ξ,l,F)c_{i}(\xi,l,F), we may assume that the pairs (ai,bi)(a_{i},b_{i}) are disjoint, and that (bi,ai)>(bi+1,ai+1)(b_{i},a_{i})>(b_{i+1},a_{i+1}) in lexicographic order, that is, either bi>bi+1b_{i}>b_{i+1} or bi=bi+1b_{i}=b_{i+1} and ai>ai+1a_{i}>a_{i+1}. Note that if b10b_{1}\leq 0, then we are done, since for each FLoc0F\in\operatorname{Loc}_{0} and each kAF,ξ,lk\in A_{F,\xi,l}:

(6.2) supyY(F)GF(y,l,k)HF(l,k)(i=1L|ci(ξ,l,F)|)kM,\sup_{y\in Y(F)}G_{F}(y,l,k)\leq H_{F}(l,k)\leq\left(\sum_{i=1}^{L}\left|c_{i}(\xi,l,F)\right|\right)k^{M},

where M=max{ai}M=\max\{a_{i}\}. Assume towards contradiction that b1>0b_{1}>0, and c1(ξ,l,F)0c_{1}(\xi,l,F)\neq 0. Then for all large enough kAF,ξ,lk\in A_{F,\xi,l}, one has

(6.3) supyY(F)GF(y,l,k)CF1HF(l,k)qF12b1k.\sup_{y\in Y(F)}G_{F}(y,l,k)\geq C_{F}^{-1}H_{F}(l,k)\geq q_{F}^{\frac{1}{2}b_{1}k}.

Now let ψR:𝔸Km𝔸Km\psi_{R}:\mathbb{A}_{K}^{m}\rightarrow\mathbb{A}_{K}^{m} be the map ψR(x1,,xm)=(x1R,,xmR)\psi_{R}(x_{1},...,x_{m})=(x_{1}^{R},...,x_{m}^{R}), for R:=4m/b1R:=\left\lceil 4m/b_{1}\right\rceil. Then by [GH24, Corollary 3.18], φψR:X×𝔸Km𝔸Km\varphi*\psi_{R}:X\times\mathbb{A}_{K}^{m}\rightarrow\mathbb{A}_{K}^{m} is (FRS). By Corollary 6.2, we have

(6.4) (φψR)(μl×μ𝒪Fm)c,(Fm).(\varphi*\psi_{R})_{*}(\mu_{l}\times\mu_{\mathcal{O}_{F}}^{m})\in\mathcal{M}_{c,\infty}(F^{m}).

On the other hand, note that

(φψR)1(B(y,qFk))φ1(B(y,qFk))×ψR1(B(0,qFk)).(\varphi*\psi_{R})^{-1}(B(y,q_{F}^{-k}))\supseteq\varphi^{-1}(B(y,q_{F}^{-k}))\times\psi_{R}^{-1}(B(0,q_{F}^{-k})).

Further note that

ψR1(B(0,qFk))=B(0,qFkR).\psi_{R}^{-1}(B(0,q_{F}^{-k}))=B\left(0,q_{F}^{-\left\lceil\frac{k}{R}\right\rceil}\right).

Thus, we have

(φψR)(μl×μ𝒪Fm)(B(y,qFk))μFm(B(y,qFk))\displaystyle\frac{(\varphi*\psi_{R})_{*}(\mu_{l}\times\mu_{\mathcal{O}_{F}}^{m})(B(y,q_{F}^{-k}))}{\mu_{F}^{m}(B(y,q_{F}^{-k}))} =μl×μ𝒪Fm((φψR)1(B(y,qFk))μFm(B(y,qFk))\displaystyle=\frac{\mu_{l}\times\mu_{\mathcal{O}_{F}}^{m}((\varphi*\psi_{R})^{-1}(B(y,q_{F}^{-k}))}{\mu_{F}^{m}(B(y,q_{F}^{-k}))}
μl(φ1(B(y,qFk)))μFm(B(y,qFk))μ𝒪Fm(ψR1(B(0,qFk)))\displaystyle\geq\frac{\mu_{l}\left(\varphi^{-1}(B(y,q_{F}^{-k}))\right)}{\mu_{F}^{m}(B(y,q_{F}^{-k}))}\cdot\mu_{\mathcal{O}_{F}}^{m}\left(\psi_{R}^{-1}(B(0,q_{F}^{-k}))\right)
=φμl(B(y,qFk))μFm(B(y,qFk))μ𝒪Fm(B(0,qFkR))\displaystyle=\frac{\varphi_{*}\mu_{l}(B(y,q_{F}^{-k}))}{\mu_{F}^{m}(B(y,q_{F}^{-k}))}\cdot\mu_{\mathcal{O}_{F}}^{m}\left(B\left(0,q_{F}^{-\left\lceil\frac{k}{R}\right\rceil}\right)\right)
GF(y,l,k)qFkRm.\displaystyle\geq G_{F}(y,l,k)\cdot q_{F}^{-\left\lceil\frac{k}{R}\right\rceil m}.

By (6.3), for each kk large enough, we may find y0Y(F)y_{0}\in Y(F), such that

(φψR)(μl×μ𝒪Fm)(B(y0,qFk))μFm(B(y0,qFk))qFkRmqF12b1kqFb18k,\frac{(\varphi*\psi_{R})_{*}(\mu_{l}\times\mu_{\mathcal{O}_{F}}^{m})(B(y_{0},q_{F}^{-k}))}{\mu_{F}^{m}(B(y_{0},q_{F}^{-k}))}\geq q_{F}^{-\left\lceil\frac{k}{R}\right\rceil m}q_{F}^{\frac{1}{2}b_{1}k}\geq q_{F}^{\frac{b_{1}}{8}k},

which contradicts (6.4). Thus b10b_{1}\leq 0 and we are done by (6.2). ∎

We are now ready to prove (1)(2)(1)\Rightarrow(2) of Theorem 6.6.

Proof of (1)(2)(1)\Rightarrow(2) of Theorem 6.6, non-Archimedean case.

Let φ:XY\varphi:X\to Y be a jet-flat morphism. We may assume Y=𝔸KmY=\mathbb{A}_{K}^{m}. Let μc(X(F))\mu\in\mathcal{M}_{c}^{\infty}(X(F)) and write gFg_{F} for the density of φμ\varphi_{*}\mu with respect to μFm\mu_{F}^{m}. Let Ysm,φY^{\mathrm{sm},\varphi} be the set of yYy\in Y such that φ\varphi is smooth over yy. For every FLoc0F\in\operatorname{Loc}_{0}, the map φF\varphi_{F} is smooth over Ysm,φ(F)Y^{\mathrm{sm},\varphi}(F), and therefore gF(y)g_{F}(y) is locally constant on Ysm,φ(F)Y^{\mathrm{sm},\varphi}(F).

By [CGH18, Corollary 1.4.3], the constancy radius of gF(y)g_{F}(y) can be taken to be definable, i.e. there exists a definable function α:Ysm,φ\alpha:Y^{\mathrm{sm},\varphi}\to\mathbb{N} such that gF(y)g_{F}(y) is constant around every ball B(y,qFαF(y))B(y,q_{F}^{-\alpha_{F}(y)}). In particular, for every yYsm,φ(F)y\in Y^{\mathrm{sm},\varphi}(F) we have gF(y)=GF,μ(y,αF(y))g_{F}(y)=G_{F,\mu}(y,\alpha_{F}(y)). In addition, by Proposition 6.8, we have GF,μ(y,k)CF,μkMG_{F,\mu}(y,k)\leq C_{F,\mu}k^{M}, for FLoc0F\in\operatorname{Loc}_{0}. We arrive at the following:

Fm|gF(y)|s𝑑y\displaystyle\int_{F^{m}}\left|g_{F}(y)\right|^{s}dy =Ysm,φ(F)|GF,μ(y,αF(y))|s𝑑yCF,μsFmαF(y)Ms𝑑y\displaystyle=\int_{Y^{\mathrm{sm},\varphi}(F)}\left|G_{F,\mu}(y,\alpha_{F}(y))\right|^{s}dy\leq C_{F,\mu}^{s}\int_{F^{m}}\alpha_{F}(y)^{Ms}dy
=CF,μsttMsμFm({yFm:αF(y)=t})\displaystyle=C_{F,\mu}^{s}\sum_{t\in\mathbb{N}}t^{Ms}\cdot\mu_{F}^{m}(\{y\in F^{m}:\alpha_{F}(y)=t\})
C(F)+CttMsqFλt<,\displaystyle\leq C^{\prime}(F)+C\sum_{t\in\mathbb{N}}t^{Ms}q_{F}^{-\lambda t}<\infty,

where the last inequality follows by [CGH18, Theorem 3.1.1], since limtμFm({yFm:αF(y)=t})=0\lim\limits_{t\to\infty}\mu_{F}^{m}(\{y\in F^{m}:\alpha_{F}(y)=t\})=0 and thus μFm({yFm:αF(y)=t})qFλt\mu_{F}^{m}(\{y\in F^{m}:\alpha_{F}(y)=t\})\leq q_{F}^{-\lambda t} for some λ>0\lambda>0 and every tt large enough. ∎

In [CGH23, Theorem 4.12], Cluckers and the first two authors showed that if φ:XY\varphi:X\to Y is a jet-flat morphism, which is defined over \mathbb{Z}, and one chooses μ=μX(p)\mu=\mu_{X(\mathbb{\mathbb{Z}}_{p})} and τ=μY(p)\tau=\mu_{Y(\mathbb{\mathbb{Z}}_{p})} to be the canonical measures on X(p)X(\mathbb{\mathbb{Z}}_{p}) and Y(p)Y(\mathbb{\mathbb{Z}}_{p}) (see [CGH23, Lemma 4.2]), then the constant Cp,μ,τC_{\mathbb{Q}_{p},\mu,\tau} in Proposition 6.8 can be taken to be independent of pp (i.e. Cp,μ,τ=CC_{\mathbb{Q}_{p},\mu,\tau}=C). [CGH23, Theorem 4.12], together with the ideas of the proof of (1)(2)(1)\Rightarrow(2) of Theorem 6.6, allows us to give bounds on the LsL^{s} norms of φμX(p)μY(p)\frac{\varphi_{*}\mu_{X(\mathbb{\mathbb{Z}}_{p})}}{\mu_{Y(\mathbb{\mathbb{Z}}_{p})}}, which are independent of pp:

Proposition 6.9.

Let φ:XY\varphi:X\to Y be a dominant morphism between finite type \mathbb{Z}-schemes XX and YY, with X,YX_{\mathbb{Q}},Y_{\mathbb{Q}} smooth and geometrically irreducible. For any prime pp, let gpg_{p} be the density of φμX(p)\varphi_{*}\mu_{X(\mathbb{\mathbb{Z}}_{p})} with respect to μY(p)\mu_{Y(\mathbb{\mathbb{Z}}_{p})}. Then the following are equivalent:

  1. (1)

    φ:XY\varphi_{\mathbb{Q}}:X_{\mathbb{Q}}\to Y_{\mathbb{Q}} is jet-flat.

  2. (2)

    For every s>1s>1, there exists C(s)>0C(s)>0 such that for every prime pp,

    gps:=Y(p)|gp(y)|s𝑑μY(p)<C(s).\left\|g_{p}\right\|_{s}:=\int_{Y(\mathbb{\mathbb{Z}}_{p})}\left|g_{p}(y)\right|^{s}d\mu_{Y(\mathbb{\mathbb{Z}}_{p})}<C(s).
Proof.

The implication (2)(1)(2)\Rightarrow(1) follows from Proposition 6.7. By (1)(2)(1)\Rightarrow(2) of Theorem 6.6, it is enough to prove (1)(2)(1)\Rightarrow(2) for ps1p\gg_{s}1. Suppose that φ\varphi_{\mathbb{Q}} is jet-flat. We may assume that Y=𝔸mY=\mathbb{A}^{m}. Indeed, working locally, and since YY_{\mathbb{Q}} is smooth, we may assume there exists a morphism ψ:Y𝔸m\psi:Y\rightarrow\mathbb{A}^{m}, such that ψ\psi_{\mathbb{Q}} is an étale map. If g~p\widetilde{g}_{p} is the density of ψφμX(p)\psi_{*}\varphi_{*}\mu_{X(\mathbb{\mathbb{Z}}_{p})} with respect to μpm\mu_{p}^{m}, and NN is an upper bound on the size of the geometric fibers of ψ\psi, then

|ψμY(p)μpm|<N,\left|\frac{\psi_{*}\mu_{Y(\mathbb{\mathbb{Z}}_{p})}}{\mu_{p}^{m}}\right|<N,

for pp large enough. In particular, we have:

Y(p)|gp(y)|sdμY(p)Y(p)|g~pψ(y))|sdμY(p)=pm|g~p(y~)|sd(ψμY(p))Npm|g~p(y~)|sdμpm.\int_{Y(\mathbb{\mathbb{Z}}_{p})}\left|g_{p}(y)\right|^{s}d\mu_{Y(\mathbb{\mathbb{Z}}_{p})}\leq\int_{Y(\mathbb{\mathbb{Z}}_{p})}\left|\widetilde{g}_{p}\circ\psi(y))\right|^{s}d\mu_{Y(\mathbb{\mathbb{Z}}_{p})}=\int_{\mathbb{Z}_{p}^{m}}\left|\widetilde{g}_{p}(\widetilde{y})\right|^{s}d(\psi_{*}\mu_{Y(\mathbb{\mathbb{Z}}_{p})})\leq N\int_{\mathbb{Z}_{p}^{m}}\left|\widetilde{g}_{p}(\widetilde{y})\right|^{s}d\mu_{p}^{m}.

As in the proof of (1)(2)(1)\Rightarrow(2) of Theorem 6.6 above, by [CGH18, Corollary 1.4.3], there exists a definable function α:(𝔸m)sm,φ\alpha:(\mathbb{A}^{m})^{\mathrm{sm},\varphi}\to\mathbb{N} such that gp(y)=Gp(y,αp(y))g_{p}(y)=G_{p}(y,\alpha_{\mathbb{Q}_{p}}(y)), where Gp(y,k):=Gp,μX(p)(y,k)G_{p}(y,k):=G_{\mathbb{Q}_{p},\mu_{X(\mathbb{\mathbb{Z}}_{p})}}(y,k). Hence, applying [CGH23, Theorem 4.12] we can find C,MC,M\in\mathbb{N} such that for p1p\gg 1:

pm|gp(y)|s𝑑y\displaystyle\int_{\mathbb{Z}_{p}^{m}}\left|g_{p}(y)\right|^{s}dy =(𝔸m)sm,φ(p)pm|Gp(y,αp(y))|s𝑑yCs(𝔸m)sm,φ(p)pmαp(y)Ms𝑑y\displaystyle=\int_{(\mathbb{A}^{m})^{\mathrm{sm},\varphi}(\mathbb{Q}_{p})\cap\mathbb{Z}_{p}^{m}}\left|G_{p}(y,\alpha_{\mathbb{Q}_{p}}(y))\right|^{s}dy\leq C^{s}\int_{(\mathbb{A}^{m})^{\mathrm{sm},\varphi}(\mathbb{Q}_{p})\cap\mathbb{Z}_{p}^{m}}\alpha_{\mathbb{Q}_{p}}(y)^{Ms}dy
=CsttMsμpm({ypm:αp(y)=t}).\displaystyle=C^{s}\sum_{t\in\mathbb{N}}t^{Ms}\cdot\mu_{p}^{m}(\{y\in\mathbb{Z}_{p}^{m}:\alpha_{\mathbb{Q}_{p}}(y)=t\}).

By [CGH18, Theorem 3.1.1], there exists LL\in\mathbb{N} and λ>0\lambda>0 such that μpm({ypm:αp(y)=t})pλt\mu_{p}^{m}(\{y\in\mathbb{Z}_{p}^{m}:\alpha_{\mathbb{Q}_{p}}(y)=t\})\leq p^{-\lambda t} for every t>Lt>L and every prime pp. We therefore get the desired claim as:

pm|gp(y)|s𝑑yCst=0LtMs+Cst>LtMspλt<Cs((L+1)LMs+t>LtMs2λt)<C(s).\int_{\mathbb{Z}_{p}^{m}}\left|g_{p}(y)\right|^{s}dy\leq C^{s}\sum_{t=0}^{L}t^{Ms}+C^{s}\sum_{t>L}t^{Ms}p^{-\lambda t}<C^{s}\left((L+1)\cdot L^{Ms}+\sum_{t>L}t^{Ms}2^{-\lambda t}\right)<C(s).\qed

6.3. (1)(2)(1)\Rightarrow(2): the Archimedean case

In this subsection we prove (1)(2)(1)\Rightarrow(2) in the cases F=F=\mathbb{\mathbb{R}} and F=F=\mathbb{C}. Let φ:XY\varphi:X\rightarrow Y be a jet-flat morphism between smooth algebraic varieties, defined over FF. Using restriction of scalars, we may assume that F=F=\mathbb{\mathbb{R}}. We would like to show that for each μc(X())\mu\in\mathcal{M}_{c}^{\infty}(X(\mathbb{\mathbb{R}})), we have φμq(Y())\varphi_{*}\mu\in\mathcal{M}^{q}(Y(\mathbb{\mathbb{R}})) for all 1q<1\leq q<\infty. We start with the following proposition.

Proposition 6.10.

Let φ:XY\varphi:X\rightarrow Y be a jet-flat map between smooth \mathbb{\mathbb{R}}-varieties. Then for every μc(X())\mu\in\mathcal{M}_{c}^{\infty}(X(\mathbb{\mathbb{R}})), every non-vanishing τ(Y())\tau\in\mathcal{M}^{\infty}(Y(\mathbb{\mathbb{R}})), and every pp\in\mathbb{N}, one can find Cμ,τ,p>0C_{\mu,\tau,p}>0 and Mμ,p>0M_{\mu,p}>0 such that for each yY()y\in Y(\mathbb{\mathbb{R}}) and 0<r<120<r<\frac{1}{2} one has

Y()((φμ)(B(y,r))τ(B(y,r)))p𝑑τ(y)<Cμ,τ,p|log(r)|Mμ,p.\int_{Y(\mathbb{\mathbb{R}})}\left(\frac{\left(\varphi_{*}\mu\right)(B(y,r))}{\tau(B(y,r))}\right)^{p}d\tau(y)<C_{\mu,\tau,p}\left|\log(r)\right|^{M_{\mu,p}}.
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 Mμ,p=CpM_{\mu,p}=C\cdot p for a sufficiently large CC independent of μ\mu.

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. (1)

    A restricted analytic function is a function f:nf:\mathbb{\mathbb{R}}^{n}\rightarrow\mathbb{\mathbb{R}} such that f|[1,1]nf|_{[-1,1]^{n}} is analytic and f|n\[1,1]n=0f|_{\mathbb{\mathbb{R}}^{n}\backslash[-1,1]^{n}}=0.

  2. (2)

    A subset AnA\subseteq\mathbb{\mathbb{R}}^{n} is subanalytic if it is definable in an\mathbb{R}_{\mathrm{an}} – the extension of the ordered real field by all restricted analytic functions. A function f:ABf:A\rightarrow B is subanalytic if its graph Γfn×m\Gamma_{f}\subseteq\mathbb{\mathbb{R}}^{n}\times\mathbb{\mathbb{R}}^{m} is subanalytic.

  3. (3)

    A function h:Ah:A\rightarrow\mathbb{\mathbb{R}} is called constructible if there exist subanalytic functions fi:Af_{i}:A\rightarrow\mathbb{\mathbb{R}} and fij:A>0f_{ij}:A\rightarrow\mathbb{\mathbb{R}}_{>0}, such that:

    (6.5) h(x)=i=1Nfi(x)j=1Nilog(fij(x)).h(x)=\sum_{i=1}^{N}f_{i}(x)\cdot\prod_{j=1}^{N_{i}}\log(f_{ij}(x)).

    We denote the class of constructible functions on AA by 𝒞(A)\mathcal{C}(A).

  4. (4)

    Given an analytic manifold ZZ, a measure μ(Z)\mu\in\mathcal{M}(Z) is called constructible if locally it is of the form f|ω|f\cdot\left|\omega\right|, where ω\omega is a regular top-form, and ff is constructible. We denote the class of constructible measures by 𝒞(Z)\mathcal{CM}(Z). Similarly, we write 𝒞c,q(Z)\mathcal{CM}_{c,q}(Z), 𝒞(Z)\mathcal{CM}^{\infty}(Z) and 𝒞c(Z)\mathcal{CM}_{c}^{\infty}(Z).

Note that X()X(\mathbb{\mathbb{R}}) is defined by polynomials, so it is definable in the real field, and in particular subanalytic. Since 𝒞(X())\mathcal{C}(X(\mathbb{\mathbb{R}})) contains indicators of balls, we may assume that μ𝒞c,(X())\mu\in\mathcal{CM}_{c,\infty}(X(\mathbb{\mathbb{R}})) when proving Proposition 6.10 and Theorem 6.6.

Proof of Proposition 6.10.

We may assume that Y=𝔸mY=\mathbb{A}_{\mathbb{\mathbb{R}}}^{m}, τ=μm\tau=\mu_{\mathbb{\mathbb{R}}}^{m} and μ𝒞c,(X())\mu\in\mathcal{CM}_{c,\infty}(X(\mathbb{\mathbb{R}})). Write g(y)g(y) for the density of φμ\varphi_{*}\mu with respect to μm\mu_{\mathbb{\mathbb{R}}}^{m}. Set

(6.6) G(y,r)=(ϕμ)(B(y,r))rm=1rmB(y,r)g(y)𝑑y.G(y,r)=\frac{(\phi_{*}\mu)(B(y,r))}{r^{m}}=\frac{1}{r^{m}}\int_{B(y,r)}g(y^{\prime})dy^{\prime}.

For each pp\in\mathbb{N}, let

Gp(r):=mG(y,r)p𝑑y.G_{p}(r):=\int_{\mathbb{\mathbb{R}}^{m}}G(y,r)^{p}dy.

By [CM11, Theorem 1.3], the functions G:m×>0G:\mathbb{\mathbb{R}}^{m}\times\mathbb{\mathbb{R}}_{>0}\rightarrow\mathbb{\mathbb{R}} and Gp(r):>0G_{p}(r):\mathbb{\mathbb{R}}_{>0}\rightarrow\mathbb{\mathbb{R}} are constructible. Writing GpG_{p} as in (6.5), and using a preparation theorem for constructible functions [CM12, Corollary 3.5], there exist δ>0\delta>0 and θ\theta\in\mathbb{\mathbb{R}}, such that θ=0\theta=0 or θ[0,δ]\theta\notin[0,\delta], and for each r(0,δ)r\in(0,\delta) one can write:

(6.7) Gp(r)=i=1MdiSi(r)|rθ|αilog(|rθ|)li,G_{p}(r)=\sum_{i=1}^{M}d_{i}\cdot S_{i}(r)\cdot\left|r-\theta\right|^{\alpha_{i}}\log(\left|r-\theta\right|)^{l_{i}},

where did_{i}\in\mathbb{\mathbb{R}}, li,Ml_{i},M\in\mathbb{N}, αi\alpha_{i}\in\mathbb{Q}, and where SiS_{i} are certain subanalytic units, called strong functions (see [CM11, Definition 2.3]). We may assume that θ=0\theta=0 as otherwise, Gp(r)G_{p}(r) is bounded on (0,δ)(0,\delta) and we are done. In addition, [CM12, Corollary 3.5] also ensures that for each 1iM1\leq i\leq M, either Si(r)=1S_{i}(r)=1 for each r(0,δ)r\in(0,\delta), or αi>1\alpha_{i}>-1. This additional property is achieved by writing each strong function Si(r)S_{i}(r) as a converging infinite sum Si(r)=j=0cijrj/pS_{i}(r)=\sum_{j=0}^{\infty}c_{ij}\cdot r^{j/p} for some p0p\in\mathbb{Q}_{\geq 0}, and split it into a finite sum j=0scijrj/p\sum_{j=0}^{s}c_{ij}\cdot r^{j/p} and an infinite sum Si~(r):=j>scijrj/p\widetilde{S_{i}}(r):=\sum_{j>s}c_{ij}\cdot r^{j/p}. By taking ss large enough, and rearranging the terms in (6.7), Cluckers and Miller ensured that αi>1\alpha_{i}>-1 whenever Si(r)1S_{i}(r)\neq 1. Following the same argument, and taking ss even larger, one can guarantee that αi>N\alpha_{i}>N for some fixed NN\in\mathbb{N}, as large as we wish. Hence, we may assume that Gp(r)G_{p}(r) has the following form:

(6.8) Gp(r)=i=1Mdirαilog(r)li+i=M+1MdiSi(r)rαilog(r)li,G_{p}(r)=\sum_{i=1}^{M^{\prime}}d_{i}r^{\alpha_{i}}\log(r)^{l_{i}}+\sum_{i=M^{\prime}+1}^{M}d_{i}S_{i}(r)r^{\alpha_{i}}\log(r)^{l_{i}},

where αi>N\alpha_{i}>N for M+1iMM^{\prime}+1\leq i\leq M, and NN\in\mathbb{N} large as we like. For 1iM1\leq i\leq M^{\prime}, we may further assume that (αi,li)(\alpha_{i},l_{i}) are mutually different and lexicographically ordered, i.e. either αi<αi+1\alpha_{i}<\alpha_{i+1}, or αi=αi+1\alpha_{i}=\alpha_{i+1} and li>li+1l_{i}>l_{i+1}. In particular, by taking δ\delta small enough, we have for 0<r<δ0<r<\delta:

(6.9) 12d1rα1|log(r)|l1<|Gp(r)|<2d1rα1|log(r)|l1.\frac{1}{2}d_{1}r^{\alpha_{1}}\left|\log(r)\right|^{l_{1}}<\left|G_{p}(r)\right|<2d_{1}r^{\alpha_{1}}\left|\log(r)\right|^{l_{1}}.

We claim that α10\alpha_{1}\geq 0. Assume not, then we have for rr small enough:

(6.10) G(,r)G(,r)p=Gp(r)1prα1p|log(r)|l1prα1p.\left\|G(\cdot\,,r)\right\|_{\infty}\geq\left\|G(\cdot\,,r)\right\|_{p}=G_{p}(r)^{\frac{1}{p}}\gtrsim r^{\frac{\alpha_{1}}{p}}\left|\log(r)\right|^{\frac{l_{1}}{p}}\gtrsim r^{\frac{\alpha_{1}}{p}}.

We now use an argument analogous to the one in Proposition 6.8. Take ψR:𝔸m𝔸m\psi_{R}:\mathbb{A}^{m}\rightarrow\mathbb{A}^{m} to be the map

ψR(x1,,xm)=(x1R,,xmR),\psi_{R}(x_{1},...,x_{m})=(x_{1}^{R},...,x_{m}^{R}),

for R:=2mp/|α1|R:=\left\lceil 2mp/\left|\alpha_{1}\right|\right\rceil and let ηCc(m)\eta\in C_{c}^{\infty}(\mathbb{\mathbb{R}}^{m}) be a bump function which is equal to one on the unit ball in m\mathbb{\mathbb{R}}^{m}. Then φψR:X×𝔸m𝔸m\varphi*\psi_{R}:X\times\mathbb{A}^{m}\rightarrow\mathbb{A}^{m} is (FRS), and thus by Corollary 6.2

(6.11) (φψR)(μ×η)c,(m).(\varphi*\psi_{R})_{*}(\mu\times\eta)\in\mathcal{M}_{c,\infty}(\mathbb{\mathbb{R}}^{m}).

Repeating precisely the same argument as in Proposition 6.8, and using (6.10), we may find {yr}r\{y_{r}\}_{r} such that:

(φψR)(μ×η)(B(yr,2r))μm(B(yr,2r))G(yr,r)ψRη(B(0,r))G(yr,r)rmRrα14p,\frac{(\varphi*\psi_{R})_{*}(\mu\times\eta)(B(y_{r},2r))}{\mu_{\mathbb{\mathbb{R}}}^{m}(B(y_{r},2r))}\gtrsim G(y_{r},r)\cdot\psi_{R*}\eta(B(0,r))\gtrsim G(y_{r},r)r^{\frac{m}{R}}\gtrsim r^{\frac{\alpha_{1}}{4p}},

which leads to a contradiction. Hence α10\alpha_{1}\geq 0, therefore on (0,δ)(0,\delta) for δ\delta small enough we have

Gp(r)|log(r)|l1.G_{p}(r)\lesssim\left|\log(r)\right|^{l_{1}}.

For r>δr>\delta, we have G(y,r)δmG(y,r)\lesssim\delta^{-m} and thus Gp(r)δpm<G_{p}(r)\lesssim\delta^{-pm}<\infty. This concludes the proof. ∎

In order to prove Theorem 6.6, we need to control the oscillations of constructible functions.

Definition 6.13.

Let f:nf:\mathbb{\mathbb{R}}^{n}\rightarrow\mathbb{\mathbb{R}} be a subanalytic function. Define αf:n×>00\alpha_{f}:\mathbb{\mathbb{R}}^{n}\times\mathbb{\mathbb{R}}_{>0}\rightarrow\mathbb{\mathbb{R}}_{\geq 0} by

(6.12) αf(y,r):=min(1,sup{t0:yB(y,t),|f(y)f(y)|<r}).\alpha_{f}(y,r):=\min\left(1,\sup\left\{t\in\mathbb{\mathbb{R}}_{\geq 0}:\forall y^{\prime}\in B(y,t),\,\left|f(y)-f(y^{\prime})\right|<r\right\}\right).

Note that αf\alpha_{f} is subanalytic, and for any r>0r>0 we have αf(y,r)>0\alpha_{f}(y,r)>0 for almost every yy. The next lemma extends this construction to the ring of constructible functions.

Lemma 6.14.

Let g𝒞(n)g\in\mathcal{C}(\mathbb{\mathbb{R}}^{n}). Then there exists a subanalytic function αg:n×>00\alpha_{g}:\mathbb{\mathbb{R}}^{n}\times\mathbb{\mathbb{R}}_{>0}\rightarrow\mathbb{\mathbb{R}}_{\geq 0} such that for any r>0r>0 we have αg(y,r)>0\alpha_{g}(y,r)>0 for almost all yny\in\mathbb{\mathbb{R}}^{n}, and

(6.13) |g(y)g(y)|r for all yB(y,αg(y,r)).\left|g(y)-g(y^{\prime})\right|\leq r\text{ for all }y^{\prime}\in B(y,\alpha_{g}(y,r)).
Proof.

If g=i=1Mgig=\sum_{i=1}^{M}g_{i} for gi𝒞(n)g_{i}\in\mathcal{C}(\mathbb{\mathbb{R}}^{n}), and suppose we already constructed αgi\alpha_{g_{i}}, for each ii. Then we may set αg(y,r)=miniαgi(y,rM)\alpha_{g}(y,r)=\min_{i}\,\alpha_{g_{i}}(y,\frac{r}{M}). Hence, by (6.5), we may assume that

(6.14) g(y)=f(y)j=1Nlog(fj(y)),g(y)=f(y)\cdot\prod_{j=1}^{N}\log(f_{j}(y)),

for some subanalytic f,f1,,fN𝒞(n)f,f_{1},...,f_{N}\in\mathcal{C}(\mathbb{\mathbb{R}}^{n}). By setting αg(y,r)=αg(y,1)\alpha_{g}(y,r)=\alpha_{g}(y,1) for r>1r>1, we may assume r1r\leq 1.

There is an open subanalytic subset UnU\subseteq\mathbb{\mathbb{R}}^{n} with complement UcU^{c} of measure 0, such that f|Uf|_{U} and all fj|Uf_{j}|_{U} are continuous (see e.g. [DvdD88, Theorem 3.2.11]). Set

M(y):=maxj(max{|f(y)|+1, 2fj(y), 2/fj(y)}),M(y):=\max_{j}\left(\max\left\{\left|f(y)\right|+1,\,2f_{j}(y),\,2/f_{j}(y)\right\}\right),

and denote h(y):=M(y)N2Nh(y):=\frac{M(y)^{-N}}{2N}. Let αf,αfj\alpha_{f},\alpha{}_{f_{j}} be as in (6.12) and set

αg(y,r)=12min{αf(y,rh(y)),αf1(y,rf1(y)h(y)),,αfN(y,rfN(y)h(y))}.\alpha_{g}(y,r)=\frac{1}{2}\min\left\{\alpha_{f}\left(y,rh(y)\right),\alpha_{f_{1}}\left(y,rf_{1}(y)h(y)\right),\dots,\alpha_{f_{N}}\left(y,rf_{N}(y)h(y)\right)\right\}.

Note that hh,αfi,αfij\alpha_{f_{i}},\alpha_{f_{ij}} are subanalytic, and thus also αg\alpha_{g} is subanalytic. Moreover, by the continuity of f|U,fj|Uf|_{U},f_{j}|_{U} we get that αf(y,),α(y,)fj>0\alpha_{f}(y,\,\cdot\,),\alpha{}_{f_{j}}(y,\,\cdot\,)>0 and thus also αg(y,)>0\alpha_{g}(y,\,\cdot\,)>0 for all yUy\in U. Note that for any real numbers a1,,aN,b1,,bN[L,L]a_{1},...,a_{N},b_{1},...,b_{N}\in[-L,L] we have

|i=1Naii=1Nbi|\displaystyle\left|\prod_{i=1}^{N}a_{i}-\prod_{i=1}^{N}b_{i}\right| =|j=1N(i=1Nj+1aii=Nj+2Nbii=1Njaii=Nj+1Nbi)|\displaystyle=\left|\sum_{j=1}^{N}\left(\prod_{i=1}^{N-j+1}a_{i}\prod_{i=N-j+2}^{N}b_{i}-\prod_{i=1}^{N-j}a_{i}\prod_{i=N-j+1}^{N}b_{i}\right)\right|
(6.15) j=1N|i=1Njaii=Nj+2Nbi||aNj+1bNj+1|LN1j=1N|aNj+1bNj+1|.\displaystyle\leq\sum_{j=1}^{N}\left|\prod_{i=1}^{N-j}a_{i}\prod_{i=N-j+2}^{N}b_{i}\right|\left|a_{N-j+1}-b_{N-j+1}\right|\leq L^{N-1}\sum_{j=1}^{N}\left|a_{N-j+1}-b_{N-j+1}\right|.

By (6.12), for every yB(y,αg(y,r))y^{\prime}\in B(y,\alpha_{g}(y,r)) and r<1r<1 we have |f(y)f(y)|rh(y)12\left|f(y^{\prime})-f(y)\right|\leq rh(y)\leq\frac{1}{2}, and

|f(y)||f(y)f(y)|+|f(y)|M(y).\left|f(y^{\prime})\right|\leq\left|f(y^{\prime})-f(y)\right|+\left|f(y)\right|\leq M(y).

Similarly, we have:

|log(fj(y))log(fj(y))|log(1+rh(y))rh(y)12,\left|\log(f_{j}(y^{\prime}))-\log(f_{j}(y))\right|\leq\log(1+rh(y))\leq rh(y)\leq\frac{1}{2},

and

|log(fj(y))|12+|log(fj(y))|12+max(fj(y),1fj(y))M(y).\left|\log(f_{j}(y^{\prime}))\right|\leq\frac{1}{2}+\left|\log(f_{j}(y))\right|\leq\frac{1}{2}+\max\left(f_{j}(y),\frac{1}{f_{j}(y)}\right)\leq M(y).

By (6.15), for every yB(y,αg(y,r))y^{\prime}\in B(y,\alpha_{g}(y,r)) we have:

|g(y)g(y)|\displaystyle\left|g(y)-g(y^{\prime})\right| =|f(y)j=1Nlog(fj(y))f(y)j=1Nlog(fj(y))|\displaystyle=\left|f(y)\cdot\prod_{j=1}^{N}\log(f_{j}(y))-f(y^{\prime})\cdot\prod_{j=1}^{N}\log(f_{j}(y^{\prime}))\right|
M(y)N(|f(y)f(y)|+j=1N|log(fj(y)fj(y))|)rh(y)(N+1)M(y)Nr.\displaystyle\leq M(y)^{N}\left(\left|f(y)-f(y^{\prime})\right|+\sum_{j=1}^{N}\left|\log\left(\frac{f_{j}(y)}{f_{j}(y^{\prime})}\right)\right|\right)\leq rh(y)(N+1)M(y)^{N}\leq r.\qed

We can now finish the proof of Theorem 6.6.

Proof of the Archimedean part of (1)(2)(1)\Rightarrow(2) of Theorem 6.6.

Let μ𝒞c,(X())\mu\in\mathcal{CM}_{c,\infty}(X(\mathbb{\mathbb{R}})) and write g(y)𝒞(m)g(y)\in\mathcal{C}(\mathbb{\mathbb{R}}^{m}) for the density of φμ\varphi_{*}\mu with respect to μm\mu_{\mathbb{\mathbb{R}}}^{m}. Let αg\alpha_{g} be as in Lemma 6.14, and set

S(2):={ym:12αg(y,12) and g(y)0}.S(2):=\left\{y\in\mathbb{\mathbb{R}}^{m}:\frac{1}{2}\leq\alpha_{g}(y,\frac{1}{2})\text{ and }g(y)\neq 0\right\}.

Then for each r3r\in\mathbb{\mathbb{R}}_{\geq 3} define the following subanalytic set:

S(r):={ym:1rαg(y,12)<1r1 and g(y)0}.S(r):=\left\{y\in\mathbb{\mathbb{R}}^{m}:\frac{1}{r}\leq\alpha_{g}(y,\frac{1}{2})<\frac{1}{r-1}\text{ and }g(y)\neq 0\right\}.

We fix LL\in\mathbb{N} large enough. Setting

(6.16) G(y,r)=(ϕμ)(B(y,r))rm=1rmB(y,r)g(y)𝑑yG(y,r)=\frac{(\phi_{*}\mu)(B(y,r))}{r^{m}}=\frac{1}{r^{m}}\int_{B(y,r)}g(y^{\prime})dy^{\prime}

and using Lemma 6.14, Proposition 6.10 and Hölder’s inequality, we have:

S(r)g(y)p𝑑y\displaystyle\int_{S(r)}g(y)^{p}dy μm(S(r))+S(r){g(y)>1}g(y)p𝑑y1+m1S(r)(y)G(y,1r)p𝑑y\displaystyle\leq\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))+\int_{S(r)\cap\{g(y)>1\}}g(y)^{p}dy\lesssim 1+\int_{\mathbb{\mathbb{R}}^{m}}1_{S(r)}(y)\cdot G(y,\frac{1}{r})^{p}dy
1+μm(S(r))11+1L(mG(y,1r)(L+1)p𝑑y)1L+1\displaystyle\lesssim 1+\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\left(\int_{\mathbb{\mathbb{R}}^{m}}G(y,\frac{1}{r})^{(L+1)p}dy\right)^{\frac{1}{L+1}}
(6.17) 1+μm(S(r))11+1L|log(1r)|Mμ,(L+1)pL+1=1+μm(S(r))11+1L|log(r)|Mμ,(L+1)pL+1.\displaystyle\lesssim 1+\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\left|\log\left(\frac{1}{r}\right)\right|^{\frac{M_{\mu,(L+1)p}}{L+1}}=1+\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\left|\log(r)\right|^{\frac{M_{\mu,(L+1)p}}{L+1}}.

Since μm(S(r))\mu_{\mathbb{\mathbb{R}}}^{m}(S(r)) is a constructible function, using a similar argument as in the Proposition 6.10, and writing μm(S(1/r))\mu_{\mathbb{\mathbb{R}}}^{m}(S(1/r)) as in (6.8) and (6.9), we get μm(S(r))rβ|log(r)|γ\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))\sim r^{\beta}\left|\log(r)\right|^{\gamma}, as rr\rightarrow\infty, for β\beta\in\mathbb{Q} and γ\gamma\in\mathbb{N}. But since r=2μm(S(r))<\sum_{r=2}^{\infty}\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))<\infty, we must have β<1\beta<-1. In particular, taking LL large enough, we get that μm(S(r))11+1Lr1δ|log(r)|γ\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\sim r^{-1-\delta}\left|\log(r)\right|^{\gamma^{\prime}} for some δ>0\delta>0. Thus, the following holds for any γ′′\gamma^{\prime\prime}\in\mathbb{\mathbb{R}}:

(6.18) r=2μm(S(r))11+1L|log(r)|γ′′<.\sum_{r=2}^{\infty}\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\left|\log(r)\right|^{\gamma^{\prime\prime}}<\infty.

By (6.17) and (6.18), we get

mg(y)p𝑑y=r=2S(r)g(y)p𝑑y1+r=3μm(S(r))11+1L|log(r)|Mμ,(L+1)pL+1<.\int_{\mathbb{\mathbb{R}}^{m}}g(y)^{p}dy=\sum_{r=2}^{\infty}\int_{S(r)}g(y)^{p}dy\lesssim 1+\sum_{r=3}^{\infty}\mu_{\mathbb{\mathbb{R}}}^{m}(S(r))^{\frac{1}{1+\frac{1}{L}}}\left|\log(r)\right|^{\frac{M_{\mu,(L+1)p}}{L+1}}<\infty.\qed

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 PλP^{\lambda}. 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 pp-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 /pk\mathbb{Z}/p^{k}\mathbb{Z}-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. pp-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 pp-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.