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

Bergman kernel functions associated to measures supported on totally real submanifolds

George Marinescu and Duc-Viet Vu
Abstract

We prove that the Bergman kernel function associated to a smooth measure supported on a piecewise-smooth maximally totally real submanifold KK in n\mathbb{C}^{n} is of polynomial growth (e.g, in dimension one, KK is a finite union of transverse Jordan arcs in \mathbb{C}). Our bounds are sharp when KK is smooth. We give an application to equidistribution of zeros of random polynomials extending a result of Shiffman-Zelditch to the higher dimensional setting.

Keywords: Equilirium measure, Bernstein-Markov measure, totally real submanifolds, equidistribution, random polynomial.

Mathematics Subject Classification 2020: 32U15, 32Q15, 32V20, 60F10, 60B20, 41A10.

1 Introduction

Let KK be a non-pluripolar compact subset in n\mathbb{C}^{n}, i.e, KK is not contained in {φ=}\{\varphi=-\infty\} for some plurisubharmonic (psh) function φ\varphi on n\mathbb{C}^{n}. Let μ\mu be a probability measure whose support is non-pluripolar and is contained in KK, and QQ be a real continuous function on KK. Let 𝒫k\mathcal{P}_{k} be the space of restrictions to KK of complex polynomials of degree at most kk on n\mathbb{C}^{n}. The scalar product

s1,s2L2(μ,kQ):=Ks1s¯2e2kQ𝑑μ\langle s_{1},s_{2}\rangle_{L^{2}(\mu,kQ)}:=\int_{K}s_{1}\overline{s}_{2}e^{-2kQ}d\mu

induces the L2(μ,kQ)L^{2}(\mu,kQ)-norm on 𝒫k\mathcal{P}_{k}. The Bergman kernel function of order kk associated to μ\mu with weight QQ is defined by

Bk(x):=sups𝒫k|s(x)ekQ(x)|2/sL2(μ,kQ)2B_{k}(x):=\sup_{s\in\mathcal{P}_{k}}|s(x)e^{-kQ(x)}|^{2}/\|s\|^{2}_{L^{2}(\mu,kQ)}

for xKx\in K. Equivalently if (s1,,sdk)(s_{1},\ldots,s_{d_{k}}) (here dkd_{k} denotes the dimension of 𝒫k\mathcal{P}_{k}) is an orthonormal basis of 𝒫k\mathcal{P}_{k} with respect to the L2(μ,kQ)L^{2}(\mu,kQ)-norm, then

Bk(x)=j=1dk|sj(x)|2e2kQ(x).B_{k}(x)=\sum_{j=1}^{d_{k}}|s_{j}(x)|^{2}e^{-2kQ(x)}.

When Q0Q\equiv 0, we say that BkB_{k} is unweighted. In this case (Q0Q\equiv 0) the inverse of BkB_{k} is known as the Christoffel function in the literature on orthogonal polynomials. In practice we also use a modified version of the Bergman kernel function as follows:

B~k(x):=sups𝒫k|s(x)|2/sL2(μ,kQ)2\tilde{B}_{k}(x):=\sup_{s\in\mathcal{P}_{k}}|s(x)|^{2}/\|s\|^{2}_{L^{2}(\mu,kQ)}

for xnx\in\mathbb{C}^{n}. The advantage of B~k\tilde{B}_{k} is that it is well defined on n\mathbb{C}^{n}.

The asymptotics of the Bergman kernel function (or its inverse, the Christoffel function) is essential for many applications in (higher dimensional or not) real analysis including approximation theory, random matrix theory, etc. There is an immense literature on such asymptotics. We refer to [7, 9, 16, 24, 26, 35, 44, 47, 60, 65], to cite a just few, for an overview on this very active research field.

Most standard settings are measures supported on concrete domains on nn\mathbb{R}^{n}\subset\mathbb{C}^{n} (such as balls or simplexes in n\mathbb{R}^{n}) or in the unit ball in n\mathbb{C}^{n}. Considering measures on n\mathbb{C}^{n} whose support are not necessarily in n\mathbb{R}^{n} are also important in many applications; e.g., one can consult [7, 39, 61, 60] where the authors consider measures supported on finite unions of piecewise smooth Jordan curves \mathbb{C} or domains in \mathbb{C} bounded by Jordan curves. We refer to the end of this section for a concrete application to the equidistribution of zeros of random polynomials.

All of settings mentioned above are particular cases of a more natural situation where our measures are supported on piecewise-smooth domains in a generic Cauchy-Riemann submanifold KK in n\mathbb{C}^{n}. This is the context in which we will work on in this paper.

We underline that in view of potential applications, it is important to work with piecewise-smooth compact sets KK (rather than only smooth ones). In what follows, by a (convex) polyhedron in M\mathbb{R}^{M}, we mean a subset in M\mathbb{R}^{M} which is the intersection of a finite number of closed half-hyperplanes in M\mathbb{R}^{M}.

Definition 1.1.

A subset KK of a real MM-dimensional smooth manifold YY is called a nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold of dimension mm if for every point pK,p\in K, there exists a local chart (Wp,Ψ)(W_{p},\Psi) of YY such that Ψ\Psi is a 𝒞5\mathcal{C}^{5}-diffeomorphism from WpW_{p} to the unit ball of M\mathbb{R}^{M} and Ψ(KWp)\Psi(K\cap W_{p}) is the intersection with the unit ball of a finite union of convex polyhedra of the same dimension m.m.

A point pKp\in K is said to be a regular point of KK if the above local chart (Wp,Ψp)(W_{p},\Psi_{p}) can be chosen such that Ψp(KWp)\Psi_{p}(K\cap W_{p}) is the intersection of the unit ball with an mm-dimensional vector subspace in M\mathbb{R}^{M}, in other words, KK is a mm-dimensional submanifold locally near pp. The regular part of KK is the set of regular points of K.K. The singular part of KK is the complement of the regular part of KK in K.K. Hence if KK is a smooth manifold with boundary, then the boundary of KK is the singular part of KK and its complement in KK is the regular part of KK.

Now let KK be a nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold of X.X. Since XX is a complex manifold, its real tangent spaces have a natural complex structure JJ. We say that KK is (Cauchy-Riemann or CR for short) generic in the sense of Cauchy-Riemann geometry if for every pKp\in K and every sequence of regular points (pm)mK(p_{m})_{m}\subset K approaching to pp, any limit space of the sequence of tangent spaces of KK at pmp_{m} is not contained in a complex hyperplane of the (real) tangent space at pp of XX (equivalently, if EE is a limit space of the sequence (TpmK)m(T_{p_{m}}K)_{m\in\mathbb{N}} of tangent spaces at pmp_{m}, then we have E+JE=TpXE+JE=T_{p}X, where TpXT_{p}X is the real tangent space of XX at pp).

For a CR generic KK, note that the space TpKJTpKT_{p}K\cap JT_{p}K (pp is a regular point in KK) is invariant under JJ and hence has a complex structure induced by JJ. In this case, the complex dimension of TpKJTpKT_{p}K\cap JT_{p}K is the same for every pp and is called the CR dimension of KK. If rr denotes the CR dimension of KK, then r=dimKnr=\dim K-n. Thus the dimension of a generic KK is at least n.n.

If YY is CR generic and dimY=n\dim Y=n, then YY is said to be (maximally) totally real, and it is locally the graph of a smooth function over a small ball centered at 0n0\in\mathbb{R}^{n} which is tangent at 0 to n\mathbb{R}^{n}. Examples of piecewise-smooth totally real submanifolds are polygons in \mathbb{C} or boundaries of polygons in \mathbb{C}, and polyhedra of dimension nn in nn\mathbb{R}^{n}\subset\mathbb{C}^{n}.

A notion playing an important role in the study of Bergman kernel functions is the following extremal function

VK,Q:=sup{ψ(n):ψQ on K},V_{K,Q}:=\sup\{\psi\in\mathcal{L}(\mathbb{C}^{n}):\,\psi\leq Q\quad\text{ on }K\},

where (n)\mathcal{L}(\mathbb{C}^{n}) is the set of psh functions ψ\psi on n\mathbb{C}^{n} such that ψ(z)log|z|\psi(z)-\log|z| is bounded at infinity on n\mathbb{C}^{n}.

Since KK is non-pluripolar, the upper semi-continuous regularisation VK,QV_{K,Q}^{*} of VK,QV_{K,Q} belongs to (n)\mathcal{L}(\mathbb{C}^{n}). If VK,Q=VK,QV_{K,Q}=V^{*}_{K,Q}, then we say (K,Q)(K,Q) is regular. A stronger notion is the following: we say that KK is locally regular if for every zKz\in K there is an open neighborhood UU of zz in n\mathbb{C}^{n} such that for every increasing sequence of psh functions (uj)j(u_{j})_{j} on UU with uj0u_{j}\leq 0 on KUK\cap U, then

(supjuj)0(\sup_{j}u_{j})^{*}\leq 0

on KUK\cap U. One can see that if KK is a locally regular set, then (K,Q)(K,Q) is regular for every continuous function QQ on KK. The following result answers the question raised in [9, Remark 1.8].

Theorem 1.2.

Let KK be a compact generic Cauchy-Riemann nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n}. Then KK is locally regular.

Note that it was known that KK is locally regular if KK is smooth real analytic; see, e.g., [9, Corollary 1.7].

The measure μ\mu is said to be a Bernstein-Markov measure (with respect to (K,Q)(K,Q)) if for every ϵ>0\epsilon>0, there exists C>0C>0 such that

supK|s|2e2kQCeϵksL2(μ,kQ)2\displaystyle\sup_{K}|s|^{2}e^{-2kQ}\leq Ce^{\epsilon k}\|s\|^{2}_{L^{2}(\mu,kQ)}

for every s𝒫ks\in\mathcal{P}_{k}. In other words, the Bergman kernel function of order kk grows at most subexponentially, i.e., supKBk=O(eϵk)\sup_{K}B_{k}=O(e^{\epsilon k}) as kk\to\infty for every ϵ>0\epsilon>0.

For some examples of Bernstein-Markov measures and criteria checking this condition, we refer to [16]. However, apart from few explicit geometric situations, there are not many (geometric) examples of Bernstein-Markov measures in higher dimensions. This is the motivation for our next main result giving a large geometric class of Bernstein-Markov measures.

Theorem 1.3.

Let KK be a compact generic Cauchy-Riemann nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n}. Let μ\mu be a finite measure supported on KK such that there exist constants τ>0,r0>0\tau>0,r_{0}>0 satisfying μ(𝔹(z,r)K)rτ\mu(\mathbb{B}(z,r)\cap K)\geq r^{\tau} for every zKz\in K, rr0r\leq r_{0} (where 𝔹(z,r)\mathbb{B}(z,r) denotes the ball of radius rr centered at zz in n\mathbb{C}^{n}). Then for every continuous function QQ on KK, μ\mu is a Bernstein-Markov measure with respect to (K,Q)(K,Q).

To the best of our knowledge, the above result was only known when KK is real analytic. We are not aware of results of this type in previous literature for maximally totally real submanifolds. A measure μ0\mu_{0} on KK is said to be a smooth volume form on KK if μ0\mu_{0} is given by a smooth volume form locally at every regular point in KK, and if for every singular point pp in KK, there is a local chart (Ψ,Wp)(\Psi,W_{p}) as in Definition 1.1, such that Ψ(KWp)=j=1sPj\Psi(K\cap W_{p})=\bigcup_{j=1}^{s}P_{j}, where PjP_{j}’s are polyhedra in n\mathbb{C}^{n} of the same dimension for every 1js1\leq j\leq s, and the restriction of μ0\mu_{0} to PjP_{j} is a smooth volume form on PjP_{j} for 1js1\leq j\leq s.

Let LebK\mathop{\mathrm{Leb}}\nolimits_{K} be now a smooth volume form on KK. Then for any M>0M>0 the measure μ=|zz0|MLebK\mu=|z-z_{0}|^{M}\mathop{\mathrm{Leb}}\nolimits_{K} satisfies the hypothesis of Theorem 1.3. Here is our next main result.

Theorem 1.4.

Let KK be a compact generic Cauchy-Riemann nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n} of dimension nKn_{K}. Let QQ be a Hölder continuous function of Hölder exponent α(0,1)\alpha\in(0,1) on KK, and let LebK\mathop{\mathrm{Leb}}\nolimits_{K} be a smooth volume form on KK, and μ=ρLebK\mu=\rho\mathop{\mathrm{Leb}}\nolimits_{K}, where ρ0\rho\geq 0 and ρλL1(LebK)\rho^{-\lambda}\in L^{1}(\mathop{\mathrm{Leb}}\nolimits_{K}) for some constant λ>0\lambda>0. Then we have

supKBkCk2nK(λ+1)/(αλ),\sup_{K}B_{k}\leq Ck^{2n_{K}(\lambda+1)/(\alpha\lambda)}\,,

for some constant C>0C>0 independent of kk.

We would like to point out that [11, Remark 3.2] shows that the regularity of weights affects considerably the growth of the Bergman kernel function. One can also consult the last reference for upper bounds for K=nK=\mathbb{C}^{n} and μ=Lebn\mu=\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}^{n}} (the Lebesgue measure on n\mathbb{C}^{n}).

There have been very few works concerning polynomial growth of Bergman kernel functions associated to measures supported on real submanifolds in higher dimensions. With an exception of [12], known upper bounds on BkB_{k} were proved mostly based on special geometric structures of the compact KnK\subset\mathbb{R}^{n} (see, e.g.,[44, 47]). Such a method is not useful in dealing with general situations as in Theorem 1.4. In [12] it was supposed that KK is smooth real algebraic in m\mathbb{R}^{m} or the closure of a bounded convex open subset in m\mathbb{R}^{m}, and their arguments use this hypothesis in an essential way.

Note that μ=|zz0|MLebK\mu=|z-z_{0}|^{M}\mathop{\mathrm{Leb}}\nolimits_{K} for any constant M>0M>0 satisfies the hypothesis of Theorem 1.4. In general it is not possible to bound from below BkB_{k} by a polynomial in kk, see Remark 3.7.

By [12, Theorems 2 and 4], if KK is the closure of a bounded open convex subset in n\mathbb{R}^{n} and Q0Q\equiv 0 and μ\mu is the restriction of the Lebesgue measure on n\mathbb{R}^{n} to KK, then knBnk^{-n}B_{n} is bounded and bounded away from 0 on a fixed compact subset in the interior of KK (the behaviour of BkB_{k} at boundary points is more complicated). On the other hand, for general μ\mu on such KK, by [24], the upper bound for BkB_{k} on KK can not be O(kn)O(k^{n}) in general. To be precise, it was proved there that if KK is a smooth Jordan curve in \mathbb{C}, and μ0\mu_{0} is the arc measure on KK, and μ=(zz0)αμ0\mu=(z-z_{0})^{\alpha}\mu_{0} for some constant α>0\alpha>0, then Bk(z0)k1+αB_{k}(z_{0})\approx k^{1+\alpha} as kk\to\infty. One can see also [45] for a similar asymptotic in the case where KK is the closure of the unit ball in n\mathbb{R}^{n}.

We note that by [31, Corollary 2.13], if (K,μ,Q)(K,\mu,Q) is as in the hypothesis of Theorem 1.4), then the triple (K,μ,Q)(K,\mu,Q) is 11-Bernstein-Markov in the sense that for every constant 0<δ10<\delta\leq 1, there exists a constant C>0C>0 such that

supK|s|2e2kQCeCk1δsL2(μ,kQ)2\displaystyle\sup_{K}|s|^{2}e^{-2kQ}\leq Ce^{Ck^{1-\delta}}\|s\|^{2}_{L^{2}(\mu,kQ)}

for every s𝒫ks\in\mathcal{P}_{k}. This is much weaker than our bound. Nevertheless, [31, Corollary 2.13] is applicable to a broader class of KK.

When KK is smooth (no boundary) and Q𝒞1,δ(K)Q\in\mathcal{C}^{1,\delta}(K) for some constant δ>0\delta>0 (e.g, KK is the unit circle in \mathbb{C} as in a classical setting), we obtain sharp bounds which have potential applications in studying sampling or interpolation problems of multivariate polynomials on maximally totally real sets in n\mathbb{C}^{n}. The case where KK is compact smooth real algebraic was considered in [12]. Here is our next main result.

Theorem 1.5.

Let KK be a compact 𝒞5\mathcal{C}^{5} maximally totally real submanifold without boundary in n\mathbb{C}^{n}. Let μ\mu be a smooth volume form on KK regarded as a measure on n\mathbb{C}^{n}. Let Q𝒞1,δ(K)Q\in\mathcal{C}^{1,\delta}(K) for some constant δ>0\delta>0. Let BkB_{k} be the Bergman kernel function associated to μ\mu with weight QQ. Then there exists a constant C>0C>0 such that

supKBkCkn\displaystyle\sup_{K}B_{k}\leq Ck^{n} (1.1)

for every k0k\geq 0.

When KK is smooth compact real algebraic of dimension nn in m\mathbb{R}^{m}, it was proved in [12] that Bk/kn1B_{k}/k^{n}\approx 1. The proof of the upper bound for BkB_{k} in [12] relies crucially on the algebraicity of KK. Our approach to Theorems 1.4 and 1.5 is different and is based on constructions of analytic discs partly attached to KK, subharmonic functions on unit discs, and fine regularity of extremal plurisubharmonic envelopes associated to KK.

In dimension one, we refer to [39, Theorem 4.3] for a similar bound when the case where KK is analytic, and to [24, 61] and references therein for asymptotics of BkB_{k} (it behaves like kk at regular points, but the asymptotic of BkB_{k} at singular points is more complicated).

We are not aware of any estimates of flavor similar to Theorem 1.5 in the previous literature for general smooth maximally totally real submanifolds in higher dimensions (except the real algebraic case mentioned above in [12]). We refer to [1, 43, 55] for more precise bounds when KK is a convex subset in n\mathbb{R}^{n}.

Our next result is the following convergence which is a consequence of Theorem 1.4.

Theorem 1.6.

Let KK be a compact generic Cauchy-Riemann nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n}. Let QQ be a Hölder continuous function on KK and μ\mu be as in Theorem 1.4. Then there exists constant C>0C>0 such that

12klogB~kVK,Q𝒞0(n)Clogkk\bigg{\|}\frac{1}{2k}\log\tilde{B}_{k}-V_{K,Q}\bigg{\|}_{\mathcal{C}^{0}(\mathbb{C}^{n})}\leq C\,\frac{\log k}{k}\,\cdot

Finally we note that we also obtain a version of (Bernstein-)Markov inequality for maximally totally real submanifolds which might be useful elsewhere; see Theorem 3.13 below.

Zeros of random polynomials. We give an application of the above results to the study of equidistribution of zeros of random polynomials.

Let KK be a non-pluripolar set in n\mathbb{C}^{n} and let μ\mu be a probability measure on n\mathbb{C}^{n} such that the support of μ\mu is contained in KK and is non-pluripolar. Let QQ be a continuous weight on KK. Let 𝒫k(K)\mathcal{P}_{k}(K) be the space of restrictions of complex polynomials of degree at most kk in n\mathbb{C}^{n} to KK. Let dk:=dim𝒫k(K)d_{k}:=\dim\mathcal{P}_{k}(K), and let p1,,pdkp_{1},\ldots,p_{d_{k}} be an orthonormal basis of 𝒫k(K)\mathcal{P}_{k}(K) with respect to the L2(μ,kQ)L^{2}(\mu,kQ)-scalar product. Consider the random polynomial

p:=j=1dkαjpj,\displaystyle p:=\sum_{j=1}^{d_{k}}\alpha_{j}p_{j}, (1.2)

where αj\alpha_{j}’s are complex i.i.d. random variables. The study of zeros of random polynomials has a long history. The most classical example may be the Kac polynomial where n=1n=1, and pj=zjp_{j}=z^{j}.

The distribution of zeros of more general random polynomials associated to orthonormal polynomials (as in (1.2)) was considered in [58] by observing that 1,z,,zk1,z,\ldots,z^{k} are an orthonormal basis of the restriction of the space of polynomials in \mathbb{C} to 𝕊1\mathbb{S}^{1} with respect to the L2L^{2}-norm induced by the Haar measure μ0\mu_{0} on 𝕊1\mathbb{S}^{1}. In this setting, the necessary and sufficient conditions for the distribution of αj\alpha_{j}’s so that the zeros of pp is equidistributed almost surely or in probability with respect to the Lebesgue measure μ0\mu_{0} on the unit circle as kk\to\infty are known; see [14, 25, 40, 42].

There are many works (in one or higher dimension) following [58], to cite just a few, [13, 15, 5, 4, 6]. In all of these works, it seems to us that the question of large deviation type estimates for the equidistribution of zeros of random polynomials has not been investigated in details. As it will be clear in our proof below, the new ingredient needed for such an estimate is an quantitative rate of convergence between 1/(2k)logB~k1/(2k)\log\tilde{B}_{k} and the extremal function associated to KK. This is what we obtained in Theorem 1.6. To state our result we need some hypothesis on μ\mu and the distribution of αj\alpha_{j}’s.

Assume now that the distribution of αj\alpha_{j}’s is fLebf\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}}, where ff is a nonnegative bounded Borel function on \mathbb{C} satisfying the following mild regularities:

|z|>r|f|LebC/r2\displaystyle\int_{|z|>r}|f|\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}}\leq C/r^{2} (1.3)

for some constant C>0C>0 independent of rr. This condition was introduced in [13, 15]. We want to study the distribution of zeros of pp as kk\to\infty. We denote by [p=0][p=0] the current of integration along the zero divisor {p=0}\{p=0\} of pp. Note that if n=1n=1, then [p=0][p=0] is the sum of Dirac masses at zeros of pp.

If (K,Q,μ)(K,Q,\mu) is Bernstein-Markov, it was proved in [15, Theorem 4.2], that almost surely

k1[p=0]ddclog|VK,Q|\displaystyle k^{-1}[p=0]\to dd^{c}\log|V_{K,Q}^{*}| (1.4)

as kk\to\infty, where the convergence is the weak one between currents. In other words, for every smooth form Φ\Phi of degree (2n2)(2n-2) with compact support in n\mathbb{C}^{n}, one has

k1{p=0}Φn𝑑dclog|VK,Q|Φk^{-1}\int_{\{p=0\}}\Phi\to\int_{\mathbb{C}^{n}}dd^{c}\log|V_{K,Q}^{*}|\wedge\Phi

as kk\to\infty. Theorem 1.3 above thus provides us a large class of measures for which the equidistribution of zeros of pp holds.

Our goal now is to obtain a rate of convergence in (1.4). To this end, it is reasonable to ask for finer regularity on μ\mu and of the distribution of αj\alpha_{j}. We don’t try to make the most optimal condition. Here is our hypothesis:

(H1) |f(z)||z|3|f(z)|\leq|z|^{-3} for |z||z| sufficiently large.

(H2) let KK be a non-degenerate 𝒞5\mathcal{C}^{5} piecewise-smooth generic Cauchy-Riemann submanifold of n\mathbb{C}^{n}, and QQ be a Hölder continuous function on KK. Let μ=ρLebK\mu=\rho\mathop{\mathrm{Leb}}\nolimits_{K}, where ρλL1(LebK)\rho^{-\lambda}\in L^{1}(\mathop{\mathrm{Leb}}\nolimits_{K}) for some constant λ>0\lambda>0.

The condition (H1) ensures that (1.3) holds, and the joint-distribution of α1,,αdk\alpha_{1},\ldots,\alpha_{d_{k}} is dominated by the Fubini-Study volume form on dk\mathbb{C}^{d_{k}} (by definition the Fubini-Study volume form is equal to ωFSdk\omega_{FS}^{d_{k}}, where ωFS\omega_{FS} is the Fubini-Study form on dkdk\mathbb{P}^{d_{k}}\supset\mathbb{C}^{d_{k}}). Clearly the Gaussian random variables satisfy this condition.

The condition (H2) is a natural generalization of the classical setting with Kac polynomials where KK is the unit circle in \mathbb{C}. Indeed in [58] the authors considered the setting in which μ\mu is the surface area on a closed analytic curve in \mathbb{C} bounding a simply connected domain Ω\Omega in \mathbb{C} or μ\mu is the restriction of the Lebesgue measure on \mathbb{C} to Ω\Omega. This setting is relevant to the theory of random matrix theory as already pointed out in [58]. We refer to [14, 53, 54] for partial generalizations (without quantitative estimates) to domains with smooth boundary in \mathbb{C}. We would like to mention also that in some cases, certain large deviation type estimates for random polynomials in dimension one were known; see [37, Theorem 10] for polynomial error terms, and [28, Theorem 1.1], and [34, Theorem 3.10] for exponential error terms.

To our best knowledge there has been no quantitative generalization of results in [58] to higher dimension. It was commented in the last paper that their method seems to have no simple generalization to the case of higher dimension.

We now recall distances on the space of currents. For every constant β0\beta\geq 0, and T,ST,S closed positive currents of bi-degree (m,m)(m,m) on the complex projective space n\mathbb{P}^{n}, define

distβ(T,S):=supΦ:Φ𝒞[β],β[β]1|TS,Φ|,\mathop{\mathrm{dist}}\nolimits_{-\beta}(T,S):=\sup_{\Phi:\,\|\Phi\|_{\mathcal{C}^{[\beta],\beta-[\beta]}}\leq 1}|\langle T-S,\Phi\rangle|,

where [β][\beta] denotes the greatest integer less than or equal to β\beta, and Φ\Phi is a smooth form of degree (2nm)(2n-m) on n\mathbb{P}^{n}.

It is a standard fact that the distance distβ\mathop{\mathrm{dist}}\nolimits_{-\beta} for β>0\beta>0 induces the weak topology on the space of closed positive currents (see for example [33, Proposition 2.1.4]). We have the following interpolation inequality: for 0<β1β20<\beta_{1}\leq\beta_{2}, there is a constant cβ1,β2c_{\beta_{1},\beta_{2}} such that

distβ2distβ1cβ1,β2[distβ2]β1/β2;\displaystyle\mathop{\mathrm{dist}}\nolimits_{-\beta_{2}}\leq\mathop{\mathrm{dist}}\nolimits_{-\beta_{1}}\leq c_{\beta_{1},\beta_{2}}[\mathop{\mathrm{dist}}\nolimits_{-\beta_{2}}]^{\beta_{1}/\beta_{2}}; (1.5)

see [33, Lemma 2.1.2] or [48, 62].

Note that the currents [p=0][p=0] and ddcVK,Qdd^{c}V_{K,Q} extends trivially through the hyperplane at infinity n\n\mathbb{P}^{n}\backslash\mathbb{C}^{n} to be closed positive currents of bi-degree (1,1)(1,1) on n\mathbb{P}^{n} (this is due to the correspondence (2.1) above). Hence one can consider distβ\mathop{\mathrm{dist}}\nolimits_{-\beta} between k1[p=0]k^{-1}[p=0] and ddcVK,Qdd^{c}V_{K,Q} as closed positive currents on n\mathbb{P}^{n}.

Theorem 1.7.

(A large deviation type estimate) Let M1M\geq 1 be a constant. Assume that (H1) and (H2) are satisfied. Then there exists a constant CM>0C_{M}>0 so that

𝒫k{(α1,,αdk)dk:dist2(k1[p=0],ddcVK,Q)CMlogkk}CMkM,\displaystyle\mathscr{P}_{k}\bigg{\{}(\alpha_{1},\ldots,\alpha_{d_{k}})\in\mathbb{C}^{d_{k}}:\mathop{\mathrm{dist}}\nolimits_{-2}\big{(}k^{-1}[p=0],dd^{c}V_{K,Q}\big{)}\geq\frac{C_{M}\log k}{k}\bigg{\}}\leq C_{M}k^{-M}, (1.6)

for every kk, where 𝒫k\mathscr{P}_{k} denotes the joint-distribution of α1,,αdk\alpha_{1},\ldots,\alpha_{d_{k}}.

By (1.5), one obtains similar estimates for distβ\mathop{\mathrm{dist}}\nolimits_{-\beta} with 0<β20<\beta\leq 2 as in Theorem 1.7. We don’t know if the right-hand side of (1.6) can be improved. We state now a direct consequence of Theorem 1.7 which gives a higher dimensional generalization of [58, Theorems 1 and 2] (except that we only obtain the error term O(logkk)O(\frac{\log k}{k}) instead of O(k1)O(k^{-1})); see also Theorem 1.9 below. Denote by 𝔼k(k1[p=0]){\mathbb{E}}_{k}(k^{-1}[p=0]) the expectation of the random normalized currents of zeros k1[p=0]k^{-1}[p=0].

Corollary 1.8.

Assume that (H1) and (H2) are satisfied. Then we have

𝔼k(k1[p=0])=ddcVK,Q+O(logkk),\displaystyle{\mathbb{E}}_{k}\big{(}k^{-1}[p=0]\big{)}=dd^{c}V_{K,Q}+O\left(\frac{\log k}{k}\right), (1.7)

where O(logkk)O(\frac{\log k}{k}) denotes a current SkS_{k} of order 0 in n\mathbb{C}^{n} such that for every smooth form Φ\Phi of degree (2nm)(2n-m) with compact support in n\mathbb{C}^{n} such that Φ𝒞21\|\Phi\|_{\mathcal{C}^{2}}\leq 1, we have

|Sk,Φ|Clogkk,|\langle S_{k},\Phi\rangle|\leq C\frac{\log k}{k}\,,

for some constant CC independent of k,Φk,\Phi.

Even when the αj\alpha_{j}’s are Gaussian variables, we underline that the decay obtained in [58] is only O(k1)O(k^{-1}), this error term is optimal in dimension 1 (one can see it by a careful examination of computations in [58, Proposition 3.3]).

Since zeros of random polynomials in higher dimension form no longer a discrete set, one might be somehow not at ease to speak of questions like correlation of zeros. To remedy this problem, one can reformulate the equidistribution of zeros of random polynomials in the following way. Let LL be a complex line in n\mathbb{C}^{n} or an (complex) algebraic curve in n\mathbb{C}^{n}. Sine generic polynomials intersect transversely LL, almost surely the number of intersection points (without counting multiplicities) of the random hypersurface {p=0}\{p=0\} and LL is exactly kdegLk\deg L by Bezout’s theorem. Define

μk,L:=1kdegLj=1kdegLδzj,\mu_{k,L}:=\frac{1}{k\deg L}\sum_{j=1}^{k\deg L}\delta_{z_{j}},

where z1,,zkdegLz_{1},\ldots,z_{k\deg L} are zeros of pp on LL. Let [L][L] be the current of integration along LL. Since VK,QV_{K,Q} is bounded, the product

μL:=1degLddcVK,Q[L]\mu_{L}:=\frac{1}{\deg L}dd^{c}V_{K,Q}\wedge[L]

is a well-defined measure supported on LL (it is simply ddc(VK,Q|L)dd^{c}(V_{K,Q}|_{L}) if LL is smooth).

Theorem 1.9.

Let M1M\geq 1 be a constant. Assume that (H1) and (H2) are satisfied. Then there exists a constant CM>0C_{M}>0 so that

𝒫k{(α1,,αdk)dk:dist2(μk,L,μL)CMlogkk}CMkM,\mathscr{P}_{k}\bigg{\{}(\alpha_{1},\ldots,\alpha_{d_{k}})\in\mathbb{C}^{d_{k}}:\mathop{\mathrm{dist}}\nolimits_{-2}\big{(}\mu_{k,L},\mu_{L}\big{)}\geq\frac{C_{M}\log k}{k}\bigg{\}}\leq C_{M}k^{-M},

for every kk. In particular, the measure μk,L\mu_{k,L} converges weakly to μL\mu_{L} as kk\to\infty.

Now since zeros of pp on LL is discrete and is equidistributed as kk\to\infty, one can ask as in [58] how zeros of pp on LL (if scaled appropriately) are correlated. Nevertheless such questions seem to be still out of reach in the higher dimensional setting. Finally we note that one can even consider LL to be a transcendental curve in n\mathbb{C}^{n}. In this case generic polynomials pp still intersect LL transversely asymptotically (see [41]); the question of equidistribution is however more involving.

Acknowledgement. We thank Norman Levenberg for many fruitful discussions.

2 Bergman kernel functions associated to a line bundle

The results mentioned in the Introduction have their direct generalizations in the context of complex geometry where n\mathbb{C}^{n} is replaced by a compact Kähler manifold. Working in such a generality will make the presentation more clear and enlarge the range of applicability of the theory. We will now describe the setting.

Let XX be a projective manifold of dimension nn. Let (L,h0)(L,h_{0}) be an ample line bundle equipped with a Hermitian metric h0h_{0} whose Chern form ω\omega is positive. Let KK be a compact non-pluripolar subset in XX. Let μ\mu be a probability measure on XX such that the support of μ\mu is non-pluripolar and is contained in KK. Let hh be a Hermitian metric on L|KL|_{K} such that h=e2ϕh0h=e^{-2\phi}h_{0}, where ϕ\phi is a continuous function on KK. For s1,s2H0(X,L)s_{1},s_{2}\in H^{0}(X,L), we define

s1,s2:=Xs1,s2h𝑑μ.\langle s_{1},s_{2}\rangle:=\int_{X}\langle s_{1},s_{2}\rangle_{h}d\mu.

Since Suppμ{\rm Supp}\mu is non-pluripolar, the last scalar product defines a norm called L2(μ,h)L^{2}(\mu,h)-norm on H0(X,L)H^{0}(X,L). Let kk\in\mathbb{N}. We obtain induced Hermitian metric hkh^{k} on LkL^{k} and a similar norm L2(μ,hk)L^{2}(\mu,h^{k}) on H0(X,Lk)H^{0}(X,L^{k}). Put dk:=dimH0(X,Lk)d_{k}:=\dim H^{0}(X,L^{k}). Let {s1,,sdk}\{s_{1},\ldots,s_{d_{k}}\} be an orthonormal basis of H0(X,Lk)H^{0}(X,L^{k}) with respect to L2(μ,hk)L^{2}(\mu,h^{k}). The Bergman kernel function of order kk associated to (L,h,μ)(L,h,\mu) is

Bk(x):=j=1dk|sj(x)|hk2=sup{|s(x)|hk2:sH0(X,Lk),sL2(μ,hk)=1}B_{k}(x):=\sum_{j=1}^{d_{k}}|s_{j}(x)|_{h^{k}}^{2}=\sup\big{\{}|s(x)|_{h^{k}}^{2}:\quad s\in H^{0}(X,L^{k}),\quad\|s\|_{L^{2}(\mu,h^{k})}=1\big{\}}

for xKx\in K.

When μ\mu is a volume form on XX and h=h0h=h_{0}, the Bergman kernel function is an object of great importance in complex geometry, for example see [49] for a comprehensive study.

The setting considered in Introduction corresponds to the case where X=nX=\mathbb{P}^{n} and (L,h0)=(𝒪(1),hFS)(L,h_{0})=(\mathcal{O}(1),h_{FS}) is the hyperplane line bundle on n\mathbb{P}^{n} endowed with the Fubini-Study metric. We consider n\mathbb{C}^{n} as an open subset in n\mathbb{P}^{n} and the weight QQ corresponds to ϕ12log(1+|z|2)\phi-\frac{1}{2}\log(1+|z|^{2}). Recall that there is a natural identification between (n)\mathcal{L}(\mathbb{C}^{n}) and the set of ωFS\omega_{FS}-psh functions on n\mathbb{P}^{n} (where ωFS\omega_{FS} denotes the Fubini-Study form on n\mathbb{P}^{n}) given by

uu12log(1+|z|2)\displaystyle u\,\longleftrightarrow\,u-\frac{1}{2}\log(1+|z|^{2}) (2.1)

for u(n)u\in\mathcal{L}(\mathbb{C}^{n}).

Another well-known example is the case where KK is the unit sphere in n\mathbb{R}^{n} (here n2n\geq 2; see, e.g, [50]) and XX is the complexification of KK, i.e, K=𝕊n1nK=\mathbb{S}^{n-1}\subset\mathbb{R}^{n} which is considered as usual a compact subset of X:={z02+z12++zn2=1}nX:=\{z_{0}^{2}+z_{1}^{2}+\cdots+z_{n}^{2}=1\}\subset\mathbb{P}^{n}. The line bundle LL on XX is the restriction of the hyperplane bundle 𝒪(1)n\mathcal{O}(1)\to\mathbb{P}^{n} to XX. We remark that in this case H0(X,Lk)H^{0}(X,L^{k}) is equal to the restriction of the space of H0(n,𝒪(k))H^{0}(\mathbb{P}^{n},\mathcal{O}(k)) to XX. Hence the restriction of H0(X,Lk)H^{0}(X,L^{k}) to KK is that of the space of complex polynomials in n\mathbb{C}^{n} to KK. To see this, notice that XX is a smooth hypersurface in n\mathbb{P}^{n}. Consider the standard exact sequence of sheaves:

0𝒪(kdegX)𝒪(k)𝒪(k)|X0,0\to\mathcal{O}(k-\deg X)\to\mathcal{O}(k)\to\mathcal{O}(k)|_{X}\rightarrow 0,

where the second arrow is the multiplication by a section of 𝒪(degX)\mathcal{O}(\deg X) whose zero divisor is equal to XX. We thus obtain a long exact sequence of cohomology spaces:

0H0(n,𝒪(kdegX))H0(n,𝒪(1))H0(n,𝒪(k)|X)H1(n,𝒪(kdegX))0\to H^{0}(\mathbb{P}^{n},\mathcal{O}(k-\deg X))\to H^{0}(\mathbb{P}^{n},\mathcal{O}(1))\to H^{0}(\mathbb{P}^{n},\mathcal{O}(k)|_{X})\\ \to H^{1}(\mathbb{P}^{n},\mathcal{O}(k-\deg X))\to\cdots

In the last sequence, H0(n,𝒪(k)|X)H^{0}(\mathbb{P}^{n},\mathcal{O}(k)|_{X}) is exactly equal to H0(X,𝒪(k)|X)H^{0}(X,\mathcal{O}(k)|_{X}), and H1(n,𝒪(kdegX))=0H^{1}(\mathbb{P}^{n},\mathcal{O}(k-\deg X))=0 by Kodaira-Nakano vanishing theorem; see [38, p. 156]. As above, the weight QQ on KK in the spherical model corresponds to ϕ12log(1+|z|2)|X\phi-\frac{1}{2}\log(1+|z|^{2})|_{X} in the setting (K,X,𝒪(1)|X)(K,X,\mathcal{O}(1)|_{X}).

The measure μ\mu is said to be a Bernstein-Markov measure (with respect to (K,ϕ,L)(K,\phi,L)) if for every ϵ>0\epsilon>0 there exists C=C(ϵ)>0C=C(\epsilon)>0 such that

supK|s|hk2CeϵksL2(μ,hk)2\displaystyle\sup_{K}|s|_{h^{k}}^{2}\leq Ce^{\epsilon k}\|s\|^{2}_{L^{2}(\mu,h^{k})} (2.2)

for every sH0(X,Lk)s\in H^{0}(X,L^{k}). In other words, the Bergman kernel function of order kk grows at most subexponentially, i.e, supKBk=O(eϵk)\sup_{K}B_{k}=O(e^{\epsilon k}) as kk\to\infty for every ϵ>0\epsilon>0. Theorem 1.3 is a particular case of the following result.

Theorem 2.1.

Let KK be a compact nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth Cauchy-Riemann generic submanifold of XX. Then for every continuous function ϕ\phi on KK, if μ\mu is a finite measure whose support is equal to KK such that there exist constants τ>0,r0>0\tau>0,r_{0}>0 satisfying μ(𝔹(z,r)K)rτ\mu(\mathbb{B}(z,r)\cap K)\geq r^{\tau} for every zKz\in K, and every rr0r\leq r_{0} (where 𝔹(z,r)\mathbb{B}(z,r) denotes the ball of radius rr centered at zz induced by a fixed smooth Riemannian metric on XX), then μ\mu is a Bernstein-Markov measure with respect to (K,ϕ,L)(K,\phi,L).

Let

ϕK:=sup{ψPSH(X,ω):ψϕ on K}.\phi_{K}:=\sup\{\psi\in{\rm PSH}(X,\omega):\,\psi\leq\phi\>\text{ on }K\}. (2.3)

Since KK is non-pluripolar, the function ϕK\phi_{K}^{*} is a bounded ω\omega-psh function. If ϕK=ϕK\phi_{K}=\phi^{*}_{K}, then we say (K,ϕ)(K,\phi) is regular. A stronger notion is the following: we say that KK is locally regular if for every zKz\in K there is an open neighborhood UU of zz such that for every increasing sequence of psh functions (uj)j(u_{j})_{j} on UU with uj0u_{j}\leq 0 on KUK\cap U, then

(supjuj)0(\sup_{j}u_{j})^{*}\leq 0

on KUK\cap U. The following result answers the question raised in [9, Remark 1.8].

Theorem 2.2.

Every compact nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth Cauchy-Riemann generic submanifold of XX is locally regular.

Note that Theorem 1.2 is a direct consequence of the above result. It was shown in [9, Corollary 1.7] that KK is locally regular if KK is smooth real analytic. Theorem 2.1 is actually a direct consequence of Theorem 2.2 and the criterion [16, Proposition 3.4] giving a sufficient condition for measures being Bernstein-Markov.

The Monge-Ampère current (ddcϕK+ω)n(dd^{c}\phi_{K}^{*}+\omega)^{n} is called the equilibrium measure associated to (K,ϕ)(K,\phi). It is well-known that the last measure is supported on KK. By [9, Theorem B] one has

dk1Bkμ(ddcϕK+ω)n,k,\displaystyle d_{k}^{-1}B_{k}\mu\to(dd^{c}\phi_{K}^{*}+\omega)^{n}\,,\quad k\to\infty, (2.4)

provided that μ\mu is a Bernstein-Markov measure associated to (K,ϕ,L)(K,\phi,L). The last property suggests that the Bergman kernel function BkB_{k} cannot behave too wildly at infinity.

Theorem 2.3.

Let KK be a compact nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth Cauchy-Riemann generic submanifold of XX. Let nKn_{K} be the dimension of KK. Let ϕ\phi be a Hölder continuous function of Hölder exponent α(0,1)\alpha\in(0,1) on KK, let LebK\mathop{\mathrm{Leb}}\nolimits_{K} be a smooth volume form on KK, and μ=ρLebK\mu=\rho\mathop{\mathrm{Leb}}\nolimits_{K}, where ρ0\rho\geq 0 and ρλL1(LebK)\rho^{-\lambda}\in L^{1}(\mathop{\mathrm{Leb}}\nolimits_{K}) for some constant λ>0\lambda>0. Then, there exists a constant C>0C>0 such that

supKBkCk2nK(λ+1)/(αλ)\sup_{K}B_{k}\leq Ck^{2n_{K}(\lambda+1)/(\alpha\lambda)}

for every kk.

Note that by the proof of [29, Theorem 1.3] or [30, Theorem 3.6], for every Hölder continuous function ϕ1\phi_{1} on XX, and μ1:=ωn\mu_{1}:=\omega^{n}, the Bergman kernel function of order kk associated to (X,μ1,ϕ1)(X,\mu_{1},\phi_{1}) grows at most polynomially on KK as kk\to\infty; see also [8, Theorem 3.1] for the case where ϕ1\phi_{1} is smooth.

Theorem 2.4.

Assume that the following two conditions hold:

(i) KK is maximally totally real and has no singularity (i.e, KK is smooth and without boundary),

(ii) ϕ𝒞1,δ(K)\phi\in\mathcal{C}^{1,\delta}(K) for some constant δ>0\delta>0.

Then there exists a constant C>0C>0 such that for every kk, the following holds

Bk(x)Ckn,B_{k}(x)\leq Ck^{n},

for every xKx\in K.

Consider the case when X=nX=\mathbb{P}^{n}, L:=𝒪(1)L:=\mathcal{O}(1), h0=hFSh_{0}=h_{FS} is be the Fubini-Study metric on 𝒪(1)\mathcal{O}(1), and KK is a smooth maximally totally real compact submanifold in nn\mathbb{C}^{n}\subset\mathbb{P}^{n}, and hh is the trivial line bundle on KK. It is clear that ϕ:=Q12log(1+|z|2)\phi:=Q-\frac{1}{2}\log(1+|z|^{2}) is in 𝒞1,δ(K)\mathcal{C}^{1,\delta}(K) if QQ is so. In this case the hypothesis of Theorem 2.4 are fulfilled. Thus Theorem 2.4 implies Theorem 1.5.

As a consequence of Theorem 2.3, we obtain the following estimate generalizing Theorem 1.6 in Introduction.

Theorem 2.5.

Let KK be a compact nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth generic submanifold KK of XX. Let ϕ\phi be a Hölder continuous function on KK. Let μ\mu be a smooth volume form on KK. Then we have

12klogB~kϕK𝒞0(X)=O(logkk),\left\|\frac{1}{2k}\log\tilde{B}_{k}-\phi_{K}\right\|_{\mathcal{C}^{0}(X)}=O\left(\frac{\log k}{k}\right),

as kk\to\infty, where

B~k:=e2kϕBk=sup{|s(x)|h0k2:sH0(X,Lk),sL2(μ,hk)=1}.\tilde{B}_{k}:=e^{2k\phi}B_{k}=\sup\big{\{}|s(x)|_{h_{0}^{k}}^{2}:\>s\in H^{0}(X,L^{k}),\>\|s\|_{L^{2}(\mu,h^{k})}=1\big{\}}.

3 Bernstein-Markov property for totally real submanifolds

In the first part of this section we prove Theorem 2.2, and hence Theorem 2.1 as commented in the paragraph after Theorem 2.2. In the second part of the section, assuming Theorem 2.3, we prove Theorem 2.5.

3.1 Local regularity

Let XX be a compact Kähler manifold of dimension nn with a Kähler form ω\omega. Let KK be a compact non-pluripolar subset on XX and ϕ\phi be a continuous function on KK. Recall

ϕK:=sup{ψ:ψω-psh,ψϕonK}.\phi_{K}:=\sup\big{\{}\psi:\quad\psi\quad\omega\text{-psh},\quad\psi\leq\phi\quad\text{on}\quad K\big{\}}.

By non-pluripolarity of KK, we have ϕK<\phi_{K}<\infty. Hence ϕK\phi^{*}_{K} is a bounded ω\omega-psh function on XX.

When K=XK=X and ϕ𝒞1,1\phi\in\mathcal{C}^{1,1}, it was proved in [10, 59, 22] that ϕK𝒞1,1\phi_{K}\in\mathcal{C}^{1,1}. In general, the best regularity for ϕK\phi_{K} is Hölder one, see Theorem 3.9 below and comment following it. One can check that if KK is locally regular, then (K,ϕ)(K,\phi) is regular for every ϕ\phi.

Let 𝔻\mathbb{D} be the open unit disc in .\mathbb{C}. An analytic disc ff in XX is a holomorphic mapping from 𝔻\mathbb{D} to XX which is continuous up to the boundary 𝔻\partial\mathbb{D} of 𝔻.\mathbb{D}. For an interval I𝔻,I\subset\partial\mathbb{D}, ff is said to be II-attached to a subset EXE\subset X if f(I)E.f(I)\subset E. Fix a Riemannian metric on XX and denote by dist(,)\mathop{\mathrm{dist}}\nolimits(\cdot,\cdot) the distance induced by it. For xXx\in X and r+r\in\mathbb{R}^{+}, let 𝔹(x,r)\mathbb{B}(x,r) be the ball of radius rr centered at xx with respect to the fixed metric. Here is the crucial property for us showing the existence of well-behaved analytic discs partly attached to a generic Cauchy-Riemann submanifold.

Proposition 3.1.

([64, Proposition 2.5]) Let KK be a compact generic nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold of XX. Then, there are positive constants c0,r0c_{0},r_{0} and θ0(0,π/2)\theta_{0}\in(0,\pi/2) such that for any a0Ka_{0}\in K and any a𝔹(a0,r0)\{a0},a\in\mathbb{B}(a_{0},r_{0})\backslash\{a_{0}\}, there exist a 𝒞2\mathcal{C}^{2} analytic disc f:𝔻¯Xf:\overline{\mathbb{D}}\rightarrow X such that ff is [eiθ0,eiθ0][e^{-i\theta_{0}},e^{i\theta_{0}}]-attached to K,K, dist(f(1),a0)c0δ\mathop{\mathrm{dist}}\nolimits(f(1),a_{0})\leq c_{0}\delta with δ=dist(a,a0),\delta=\mathop{\mathrm{dist}}\nolimits(a,a_{0}), f𝒞2c0\|f\|_{\mathcal{C}^{2}}\leq c_{0}, and there is z𝔻z^{*}\in\mathbb{D} so that |1z|c0δ|1-z^{*}|\leq\sqrt{c_{0}\delta} and f(z)=a.f(z^{*})=a. Moreover if a0a_{0} is in a fixed compact subset KK^{\prime} of the regular part of KK, then we have |1z|c0δ|1-z^{*}|\leq c_{0}\delta.

It was stated that f𝒞1(𝔻¯)f\in\mathcal{C}^{1}(\overline{\mathbb{D}}) instead of f𝒞2(𝔻¯)f\in\mathcal{C}^{2}(\overline{\mathbb{D}}) in [64, Proposition 2.5]. But the latter regularity is indeed clear from the construction in the proof of [64, Proposition 2.5]. Note that the compactness of XX is not necessary in the above result. In particular, if KUK\Subset U for some open subset UU of XX, then the analytic disc ff can be chosen to lie entirely in UU. Here is a slight improvement of Proposition 3.1.

Proposition 3.2.

Assume that one of the following assumptions hold:

(1) KK is a compact generic 𝒞5\mathcal{C}^{5} smooth submanifold with 𝒞5\mathcal{C}^{5} smooth boundary that is also generic,

(2) KK is a union of a finite number of compact sets as in (1).

Then there are positive constants c0,r0c_{0},r_{0} and θ0(0,π/2)\theta_{0}\in(0,\pi/2) such that for any a0Ka_{0}\in K and any a𝔹(a0,r0)\{a0},a\in\mathbb{B}(a_{0},r_{0})\backslash\{a_{0}\}, there exist a 𝒞2\mathcal{C}^{2} analytic disc f:𝔻¯Xf:\overline{\mathbb{D}}\rightarrow X such that ff is [eiθ0,eiθ0][e^{-i\theta_{0}},e^{i\theta_{0}}]-attached to K,K, dist(f(1),a0)c0δ\mathop{\mathrm{dist}}\nolimits(f(1),a_{0})\leq c_{0}\delta with δ=dist(a,a0),\delta=\mathop{\mathrm{dist}}\nolimits(a,a_{0}), f𝒞2c0\|f\|_{\mathcal{C}^{2}}\leq c_{0}, and there is z𝔻z^{*}\in\mathbb{D} so that |1z|c0δ|1-z^{*}|\leq c_{0}\delta.

Proof.

If KK fulfills one of the conditions (1) or (2) then KK can be covered by a finite number of sets KjK_{j} such that for every jj there exist an open subset UjU_{j} in XX and a smooth family (Kjs)sSj(K_{js})_{s\in S_{j}} of 𝒞5\mathcal{C}^{5} smooth generic CR submanifolds KjsK_{js} in UjU_{j} such that KjsK_{js} is 𝒞5\mathcal{C}^{5} smooth without boundary in UjU_{j}, for every ss, and satisfies Kj=sSjKjsK_{j}=\cup_{s\in S_{j}}K_{js}. Now the desired assertion follows directly from Proposition 3.1 applied to each KjsK_{js} and points in UjU_{j} correspondingly. We note that the constants c0,r0,θ0c_{0},r_{0},\theta_{0} can be chosen independent of sSjs\in S_{j} because as shown in the proof of [64, Proposition 2.5], they depend only on bounds on 𝒞3\mathcal{C}^{3}-norm of diffeomorphisms defining local charts in KjsK_{js} (see [64, Lemma 4.1]), these bounds are independent of sSjs\in S_{j} for the family (Kjs)sSj(K_{js})_{s\in S_{j}} is smooth. ∎

Examples for compact KK satisfying the hypothesis of Proposition 3.2 include a collection of finite number of smooth Jordan arcs in \mathbb{C} regardless of their configuration or the closure of an open subset with smooth boundary in XX.

Lemma 3.3.

Let θ0(0,π/6),β(0,1)\theta_{0}\in(0,\pi/6),\beta\in(0,1) and let c>0c>0 be a constant. Let ψ\psi be a subharmonic function on 𝔻\mathbb{D}. Assume that

lim supz𝔻eiθψ(z)c|θ|βforθ(θ0,θ0)andsup𝔻ψc.\displaystyle\limsup_{z\in\mathbb{D}\to e^{i\theta}}\psi(z)\leq c|\theta|^{\beta}\quad\text{for}\quad\theta\in(-\theta_{0},\theta_{0})\quad\text{and}\quad\sup_{\mathbb{D}}\psi\leq c. (3.1)

Then, there exists a constant CC depending only on (θ0,β,c)(\theta_{0},\beta,c) so that for any z𝔻,z\in\mathbb{D}, we have

ψ(z)C|1z|β.\displaystyle\psi(z)\leq C|1-z|^{\beta}. (3.2)

Moreover if lim supz𝔻eiθψ(z)g(eθ)\limsup_{z\in\mathbb{D}\to e^{i\theta}}\psi(z)\leq g(e^{\theta}) for some function g𝒞1,δg\in\mathcal{C}^{1,\delta} on [eiθ0,eiθ0][e^{-i\theta_{0}},e^{i\theta_{0}}] and for some δ>0\delta>0 so that g(1)=0g(1)=0, then

ψ(z)C|1z|,\displaystyle\psi(z)\leq C|1-z|, (3.3)

for some constant CC independent of z𝔻z\in\mathbb{D}.

Proof.

The desired inequality (3.2) is essentially contained in [64, Lemma 2.6]. The hypothesis of continuity up to boundary of ψ\psi in the last lemma is superfluous and the proof there still works in our current setting. Note that the proof of [64, Lemma 2.6] does not work for β=1\beta=1 because the harmonic extension of a Lipschitz function on 𝔻\partial\mathbb{D} is not necessarily Lipschitz on 𝔻¯\overline{\mathbb{D}}. However since the harmonic extension of a 𝒞1,δ\mathcal{C}^{1,\delta} function on 𝔻\partial\mathbb{D} to 𝔻\mathbb{D} is also 𝒞1,δ\mathcal{C}^{1,\delta} on 𝔻¯\overline{\mathbb{D}} (see, e.g, [36, Page 41]), we obtain (3.3). ∎

End of the proof of Theorem 2.2.

Let a0Ka_{0}\in K and a small ball 𝔹\mathbb{B} of XX around a0a_{0}. Consider an increasing sequence (uj)j(u_{j})_{j} of psh functions bounded uniformly from above on 𝔹\mathbb{B} such that uj0u_{j}\leq 0 on K𝔹K\cap\mathbb{B}. We need to check that (supjuj)0(\sup_{j}u_{j})^{*}\leq 0 on K𝔹K\cap\mathbb{B}. Now, we will essentially follow arguments from the proof of [64, Theorem 2.3]. Let 𝔹\mathbb{B}^{\prime} be a relatively compact subset of 𝔹\mathbb{B} containing a0a_{0}. We will check that there exists a constant C>0C>0 such that for every a𝔹a\in\mathbb{B}^{\prime}, we have

uj(a)Cdist(a,K)1/5.\displaystyle u_{j}(a)\leq C\mathop{\mathrm{dist}}\nolimits(a,K)^{1/5}. (3.4)

The desired assertion is deduced from the last inequality by taking dist(a,K)0\mathop{\mathrm{dist}}\nolimits(a,K)\to 0. It remains to check (3.4).

Let a0a^{\prime}_{0} be a point in KK such that dist(a,a0)=dist(a,K)\mathop{\mathrm{dist}}\nolimits(a,a^{\prime}_{0})=\mathop{\mathrm{dist}}\nolimits(a,K). Put δ:=dist(a,K)=dist(a,a0)\delta:=\mathop{\mathrm{dist}}\nolimits(a,K)=\mathop{\mathrm{dist}}\nolimits(a,a^{\prime}_{0}). By Proposition 3.1, there exists an analytic disc f:𝔻𝔹f:\mathbb{D}\to\mathbb{B} continuous up to boundary and za𝔻z_{a}\in\mathbb{D} with |za1|Cδ1/2|z_{a}-1|\leq C\delta^{1/2} such that f(za)=af(z_{a})=a and dist(f(1),a0)Cδ\mathop{\mathrm{dist}}\nolimits\big{(}f(1),a^{\prime}_{0}\big{)}\leq C\delta, and f([eiθ0,eiθ0])Kf([e^{-i\theta_{0}},e^{i\theta_{0}}])\subset K, for some constants CC and θ0\theta_{0} independent of aa.

Put vj:=ujfv_{j}:=u_{j}\circ f. Since uj0u_{j}\leq 0 on 𝔹K\mathbb{B}\cap K and f([eiθ0,eiθ0])Kf([e^{-i\theta_{0}},e^{i\theta_{0}}])\subset K, we get vj(eiθ)0v_{j}(e^{i\theta})\leq 0 for θ[θ0,θ0]\theta\in[-\theta_{0},\theta_{0}]. Moreover since uju_{j} is uniformly bounded from above, there is a constant MM such that vjMv_{j}\leq M for every jj. This allows us to apply Lemma 3.3 for β=1ϵ\beta=1-\epsilon (for some constant ϵ>0\epsilon>0 small) and cc big enough. We infer that vj(z)|1z|1/2v_{j}(z)\lesssim|1-z|^{1/2}. Substituting z=zaz=z_{a} in the last inequality gives

uj(a)=uj(f(za))=vj(za)|1za|(1ϵ)/2δ(1ϵ)/4.u_{j}(a)=u_{j}\big{(}f(z_{a})\big{)}=v_{j}(z_{a})\lesssim|1-z_{a}|^{(1-\epsilon)/2}\lesssim\delta^{(1-\epsilon)/4}.

Hence, (3.4) follows by choosing ϵ:=1/5\epsilon:=1/5. The proof is finished. ∎

Let h0,h,ϕh_{0},h,\phi be as in the previous section. Recall that the Chern form of h0h_{0} is equal to ω\omega. Define

ϕK,k:=sup{k1log|σ|h0:σH0(X,Lk),supK(|σ|h0kekϕ)1}.\phi_{K,k}:=\sup\big{\{}k^{-1}\log|\sigma|_{h_{0}}:\quad\sigma\in H^{0}(X,L^{k}),\,\sup_{K}(|\sigma|_{h_{0}^{k}}e^{-k\phi})\leq 1\big{\}}.

Clearly ϕK,kϕK\phi_{K,k}\leq\phi_{K}. We recall the following well-known fact.

Lemma 3.4.

The sequence (ϕK,k)k(\phi_{K,k})_{k} increases pointwise to ϕK\phi_{K} as kk\to\infty.

As a direct consequence of the above lemma, we see that ϕK\phi_{K} is lower semi-continuous.

Proof.

Since we couldn’t find a proper reference, we present detailed arguments here. We just need to use Demailly’s analytic approximation. Since ϕK\phi_{K} is bounded, without loss of generality, we can assume that ϕK<0\phi_{K}<0. Clearly, ϕK,kϕK\phi_{K,k}\leq\phi_{K}. Fix aXa\in X. Let δ\delta be a positive constant. Let ψ\psi be a negative ω\omega-psh function with ψϕ\psi\leq\phi on KK such that ψ(a)ϕK(a)δ\psi(a)\geq\phi_{K}(a)-\delta. Let ϵ(0,1)\epsilon\in(0,1). Observe ddc(1ϵ)ψ+ωϵωdd^{c}(1-\epsilon)\psi+\omega\geq\epsilon\omega. This allows us to apply [27, Theorem 14.21] to ψ\psi. Let (σj)j(\sigma_{j})_{j} is an orthonormal basis of H0(X,Lk)H^{0}(X,L^{k}) with respect to L2L^{2}-norm generated by the Hermitian metric hϵ,ψ,k:=ek(1ϵ)ψh0kh_{\epsilon,\psi,k}:=e^{-k(1-\epsilon)\psi}h_{0}^{k} and ωn\omega^{n}. Set

ψϵ,k:=1klogj=1dk|σj|h0k\psi_{\epsilon,k}:=\frac{1}{k}\log\sum_{j=1}^{d_{k}}|\sigma_{j}|_{h_{0}^{k}}

Then

ψϵ,k(1ϵ)ψ\psi_{\epsilon,k}\geq(1-\epsilon)\psi

and ψk\psi_{k} converges pointwise to (1ϵ)ψ(1-\epsilon)\psi as kk\to\infty. Note that

ψϵ,k=sup{1klog|σ|h0k:σH0(X,Lk):σL2(ωn,hϵ,ψ,k)=1}.\psi_{\epsilon,k}=\sup\left\{\frac{1}{k}\log|\sigma|_{h_{0}^{k}}:\sigma\in H^{0}(X,L^{k}):\|\sigma\|_{L^{2}(\omega^{n},h_{\epsilon,\psi,k})}=1\right\}.

Let (ψ(N))N(\psi^{(N)})_{N} be a sequence of continuous functions decreasing to ψ\psi as NN\to\infty. Using Hartog’s lemma applied to (ψ1/N,k)k(\psi_{1/N,k})_{k\in\mathbb{N}}, we see that there is a sequence (kN)N(k_{N})_{N}\subset\mathbb{N} increasing to \infty such that

(11/N)ψψ1N,kN(11/N)ψ(N)+1/N.(1-1/N)\psi\leq\psi_{\frac{1}{N},k_{N}}\leq(1-1/N)\psi^{(N)}+1/N.

Consequently, ψ1N,kN\psi_{\frac{1}{N},k_{N}} converges pointwise to ψ\psi as NN\to\infty.

Recall that ψϕ\psi\leq\phi on KK. By Hartog’s lemma again and the continuity of ϕ\phi, for every constant δ>0\delta^{\prime}>0 and NN large enough, we have

ψ1N,kNϕ+δ\displaystyle\psi_{\frac{1}{N},k_{N}}\leq\phi+\delta^{\prime} (3.5)

on KK. It follows that

kN1log|σ|h0kNϕ+δ\displaystyle k_{N}^{-1}\log|\sigma|_{h_{0}^{k_{N}}}\leq\phi+\delta^{\prime} (3.6)

for every σH0(X,LkN)\sigma\in H^{0}(X,L^{k_{N}}) with σL2(ωn,h1N,ψ,kN)=1\|\sigma\|_{L^{2}(\omega^{n},h_{\frac{1}{N},\psi,k_{N}})}=1. We deduce that

ϕK,kNkN1log|σ|h0kNδ\phi_{K,k_{N}}\geq k_{N}^{-1}\log|\sigma|_{h_{0}^{k_{N}}}-\delta^{\prime}

for such σ\sigma. In other words, ϕK,kNψ1N,kNδ\phi_{K,k_{N}}\geq\psi_{\frac{1}{N},k_{N}}-\delta^{\prime} for NN big enough. Letting NN\to\infty gives

lim infNϕK,kN(a)limNψ1N,kN(a)δ=ψ(a)δϕK(a)δδ.\liminf_{N\to\infty}\phi_{K,k_{N}}(a)\geq\lim_{N\to\infty}\psi_{\frac{1}{N},k_{N}}(a)-\delta^{\prime}=\psi(a)-\delta^{\prime}\geq\phi_{K}(a)-\delta^{\prime}-\delta.

Letting δ,δ\delta,\delta^{\prime} tend to 0 yields that lim infNϕK,kN(a)=ϕK(a)\liminf_{N\to\infty}\phi_{K,k_{N}}(a)=\phi_{K}(a). Hence ϕK,kNϕK\phi_{K,k_{N}}\to\phi_{K} as NN\to\infty. We have actually shown that for every sequence (kN)N(k^{\prime}_{N})_{N}\subset\mathbb{N} converging to \infty, there is a subsequence (kN)N(k_{N})_{N} such that ϕK,kN\phi_{K,k_{N}} converges to ϕK\phi_{K}. Thus the desired assertion follows. ∎

Using arguments from [17, Lemma 3.2] and Lemma 3.4 gives the following.

Lemma 3.5.

Assume that (K,ϕ)(K,\phi) is regular. Then ϕK,k\phi_{K,k} converges uniformly to ϕK\phi_{K} as kk\to\infty.

Proof.

For readers’ convenience, we briefly recall the proof here. Since ϕK=ϕK\phi_{K}=\phi_{K}^{*}, we see that ϕK\phi_{K} is upper semi-continuous. This combined with the fact that ϕK\phi_{K} is already lower semi-continuous gives that ϕK\phi_{K} is continuous. By Lemma 3.4, we have the pointwise convergence of ϕK,k\phi_{K,k} to ϕK\phi_{K}. Using the envelop defining ϕK\phi_{K}, observe next that

kϕK,k+mϕK,m(k+m)ϕK,k+m\displaystyle k\phi_{K,k}+m\phi_{K,m}\leq(k+m)\phi_{K,k+m} (3.7)

for every k,mk,m.

We fix a Riemannian metric dd on XX. Let ϵ>0\epsilon>0. Since XX is compact, ϕK\phi_{K} is uniformly continuous on XX. Hence there exists a constant δ>0\delta>0 such that d(ϕK(x),ϕK(y))ϵd\big{(}\phi_{K}(x),\phi_{K}(y)\big{)}\leq\epsilon if d(x,y)δd(x,y)\leq\delta for every x,yXx,y\in X. Fix x0Xx_{0}\in X. Let k0>0k_{0}>0 be a natural number such that for kk0k\geq k_{0}, we have

d(ϕK(x0),ϕK,k(x0))ϵ.d\big{(}\phi_{K}(x_{0}),\phi_{K,k}(x_{0})\big{)}\leq\epsilon.

Since the line bundle LL is positive, ϕK,r\phi_{K,r} is continuous for rr big enough. Hence without loss of generality we can assume that ϕK,r\phi_{K,r} is continuous for every rr, for only big rr matters for us. By shrinking δ\delta if necessary, we obtain that for every 1rk01\leq r\leq k_{0}, one has

d(ϕK,r(x),ϕK,r(y))ϵd\big{(}\phi_{K,r}(x),\phi_{K,r}(y)\big{)}\leq\epsilon

if d(x,y)δd(x,y)\leq\delta. Write k=k0r+sk=k_{0}r+s for 0sk010\leq s\leq k_{0}-1. Using this and (3.7) yields

kϕK,krk0ϕK,k0+sϕK,s.k\phi_{K,k}\geq rk_{0}\phi_{K,k_{0}}+s\phi_{K,s}.

It follows that

ϕK,krk0kϕK,k0+skϕK,s.\phi_{K,k}\geq r\frac{k_{0}}{k}\,\phi_{K,k_{0}}+\frac{s}{k}\,\phi_{K,s}\,.

Thus

ϕK,k(x)ϕK(x)rk0k(ϕK,k0(x)ϕK(x))(1rk0k)ϕK(x)+skϕK,s(x).\phi_{K,k}(x)-\phi_{K}(x)\geq\frac{rk_{0}}{k}(\phi_{K,k_{0}}(x)-\phi_{K}(x))-\Big{(}1-r\frac{k_{0}}{k}\Big{)}\phi_{K}(x)+\frac{s}{k}\,\phi_{K,s}(x).

The choice of δ\delta now implies that there exists a constant C>0C>0 such that the right-hand side is bounded from below by 3ϵCk0/k-3\epsilon-Ck_{0}/k if d(x,x0)δd(x,x_{0})\leq\delta. Since ϕK,kϕK\phi_{K,k}\leq\phi_{K}, we obtain the uniform convergence of ϕK,k\phi_{K,k} to ϕK\phi_{K}. ∎

Put

ϕ~K,k:=12klogB~k=12klogj=1dk|sj|h0k2.\tilde{\phi}_{K,k}:=\frac{1}{2k}\log\tilde{B}_{k}=\frac{1}{2k}\log\sum_{j=1}^{d_{k}}|s_{j}|^{2}_{h_{0}^{k}}.
Proposition 3.6.

Assume that (K,ϕ)(K,\phi) is regular and (K,μ,ϕ)(K,\mu,\phi) satisfies the Bernstein-Markov property. Then we have

ϕ~K,kϕK𝒞0(X)0,as k.\big{\|}\tilde{\phi}_{K,k}-\phi_{K}\big{\|}_{\mathscr{C}^{0}(X)}\to 0,\>\>\text{as $k\to\infty$}. (3.8)

In particular,

limkB~k1/k=e2ϕK.\lim_{k\to\infty}\tilde{B}_{k}^{1/k}=e^{2\phi_{K}}. (3.9)

Note that the limit in (3.9) is independent of μ\mu. We refer to [7, Lemma 2.8] for more informations in the case KnX=nK\subset\mathbb{C}^{n}\subset X=\mathbb{P}^{n}.

Proof.

When X=nX=\mathbb{P}^{n} and L=𝒪(1)L=\mathcal{O}(1), this is Lemma 3.4 in [17]. The arguments there work for our setting. We reproduce here the proof for the readers’ convenience. It suffices to check the first desired property (3.8). Observe that

supK(|s|h0k2e2kϕ)=supK|s|hk2(supKBk)sL2(μ,kϕ)2.\displaystyle\sup_{K}(|s|^{2}_{h_{0}^{k}}e^{-2k\phi})=\sup_{K}|s|^{2}_{h^{k}}\leq\big{(}\sup_{K}B_{k}\big{)}\|s\|^{2}_{L^{2}(\mu,k\phi)}. (3.10)

Combining this with the Bernstein-Markov property, we see that for every ϵ>0\epsilon>0, there holds

supK(|s|h0k2e2kϕ)eϵksL2(μ,kϕ)2\displaystyle\sup_{K}(|s|^{2}_{h_{0}^{k}}e^{-2k\phi})\leq e^{\epsilon k}\|s\|^{2}_{L^{2}(\mu,k\phi)} (3.11)

for every sH0(X,Lk)s\in H^{0}(X,L^{k}). Observe also that

|s|h0ksupK(|s|h0kekϕ)ekϕK,k\displaystyle|s|_{h_{0}^{k}}\leq\sup_{K}(|s|_{h_{0}^{k}}e^{-k\phi})e^{k\phi_{K,k}} (3.12)

on XX. Applying the last inequality to s:=sjs:=s_{j} and using (3.11) we infer that

12klogB~kϵ+ϕK,k.\frac{1}{2k}\log\tilde{B}_{k}\leq\epsilon+\phi_{K,k}.

In other words, ϕ~K,kϵ+ϕK,k\tilde{\phi}_{K,k}\leq\epsilon+\phi_{K,k} on XX. On the other hand, if supK(|s|h0kekϕ)1\sup_{K}(|s|_{h_{0}^{k}}e^{-k\phi})\leq 1, then

sL2(μ,kϕ)C\|s\|_{L^{2}(\mu,k\phi)}\leq C

for some constant CC independent of kk. It follows that

B~kC1e2kϕK,k.\tilde{B}_{k}\geq C^{-1}e^{2k\phi_{K,k}}.

Consequently ϕ~K,kϕK,k+O(k1)\tilde{\phi}_{K,k}\geq\phi_{K,k}+O(k^{-1}). Thus using Lemma 3.5 we obtain the desired assertion. This finishes the proof. ∎

Remark 3.7.

Recall ϕKϕ\phi_{K}\leq\phi on KK. If xKx\in K is a point so that ϕK(x)<ϕ(x)\phi_{K}(x)<\phi(x), then by Proposition 3.6 we see that Bk1(x)B_{k}^{-1}(x) grows exponentially as kk\to\infty. Consider now the case where K=XK=X and ϕ\phi is not an ω\omega-psh function. In this case there exists xXx\in X with ϕX(x)<ϕ(x)\phi_{X}(x)<\phi(x), and hence BkB_{k} becomes exponentially small as kk\to\infty.

3.2 Hölder regularity of extremal plurisubharmonic envelopes

Let α(0,1]\alpha\in(0,1] and YY be a metric space. We denote by 𝒞0,α(Y)\mathcal{C}^{0,\alpha}(Y) the space of functions on YY of finite 𝒞0,α\mathcal{C}^{0,\alpha}-norm. If 0<α<10<\alpha<1, then we also write 𝒞α\mathcal{C}^{\alpha} for 𝒞0,α\mathcal{C}^{0,\alpha}. The following notion introduced in [30] will play a crucial role for us.

Definition 3.8.

For α(0,1]\alpha\in(0,1] and α(0,1],\alpha^{\prime}\in(0,1], a non-pluripolar compact KK is said to be (𝒞0,α,𝒞0,α)(\mathcal{C}^{0,\alpha},\mathcal{C}^{0,\alpha^{\prime}})-regular if for any positive constant C,C, the set {ϕK:ϕ𝒞0,α(K) and ϕ𝒞0,α(K)C}\{\phi_{K}:\phi\in\mathcal{C}^{0,\alpha}(K)\text{ and }\|\phi\|_{\mathcal{C}^{0,\alpha}(K)}\leq C\} is a bounded subset of 𝒞0,α(X).\mathcal{C}^{0,\alpha^{\prime}}(X).

The following provides examples for the last notion.

Theorem 3.9 ([64, Theorem 2.3]).

Let α\alpha be an arbitrary number in (0,1).(0,1). Then any compact generic nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold KK of XX is (𝒞0,α,𝒞0,α/2)(\mathcal{C}^{0,\alpha},\mathcal{C}^{0,\alpha/2})-regular. Moreover if KK has no singularity, then KK is (𝒞0,α,𝒞0,α)(\mathcal{C}^{0,\alpha},\mathcal{C}^{0,\alpha})-regular.

The following remark follows immediately from Proposition 3.2 and the proof of [64, Theorem 2.3].

Remark 3.10.

If KK is as in Proposition 3.2, then KK is also (𝒞0,α,𝒞0,α)(\mathcal{C}^{0,\alpha},\mathcal{C}^{0,\alpha})-regular for 0<α<10<\alpha<1. The union of a finite number of open subsets with smooth boundary in XX is an example of such KK.

If K=XK=X, and ϕ𝒞0,1\phi\in\mathcal{C}^{0,1}, then it was shown in [22] that ϕX𝒞0,1\phi_{X}\in\mathcal{C}^{0,1}, hence XX is (𝒞0,1,𝒞0,1)(\mathcal{C}^{0,1},\mathcal{C}^{0,1})-regular; see also [10, 22, 59] for more information. In the case where K=XK=X or KK is an open subset with smooth boundary in XX it was proved in [30] that KK is (𝒞0,α,𝒞0,α)(\mathcal{C}^{0,\alpha},\mathcal{C}^{0,\alpha})-regular for 0<α<10<\alpha<1. This was extended for KK as in the statement of Theorem 3.9 in [64, Theorem 2.3]; see also [46]. We don’t know if Theorem 3.9 holds for α=1\alpha=1. Here is a partial result whose proof is exactly as of [64, Theorem 2.3] by using (3.3) instead of (3.2) (and noting that the analytic disc in Proposition 3.1 is 𝒞2(𝔻¯)\mathcal{C}^{2}(\overline{\mathbb{D}}), hence in particular, is 𝒞1,δ\mathcal{C}^{1,\delta} for some δ(0,1]\delta\in(0,1]).

Theorem 3.11.

Let δ(0,1),C1>0\delta\in(0,1),C_{1}>0 be constants. Let KK be a compact generic 𝒞5\mathcal{C}^{5} smooth submanifold (without boundary) of XX. Then there exists a constant C2>0C_{2}>0 such that for every ϕ𝒞1,δ(K)\phi\in\mathcal{C}^{1,\delta}(K) with ϕ𝒞1,δC1\|\phi\|_{\mathcal{C}^{1,\delta}}\leq C_{1}, then ϕK𝒞0,1C2\|\phi_{K}\|_{\mathcal{C}^{0,1}}\leq C_{2}.

We mention at this point an example in [56] of a domain KK with 𝒞0\mathcal{C}^{0} boundary but (K,ϕ)(K,\phi) is not regular even for ϕ=0\phi=0. Applying Theorem 3.11 to X=nX=\mathbb{P}^{n} we obtain the following result that implies [57, Conjecture 6.2] as a special case.

Theorem 3.12.

Let α(0,1]\alpha\in(0,1]. Let KK be a compact generic nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n}. Let

VK:=sup{ψ(n):ψ0 on K}.V_{K}:=\sup\{\psi\in\mathcal{L}(\mathbb{C}^{n}):\psi\leq 0\,\text{ on }K\}.

Then VK𝒞1/2(n)V_{K}\in\mathcal{C}^{1/2}(\mathbb{C}^{n}). Additionally if KK has no singularity, then VK𝒞0,1(n)V_{K}\in\mathcal{C}^{0,1}(\mathbb{C}^{n}).

We note that the fact that VK𝒞0,1(n)V_{K}\in\mathcal{C}^{0,1}(\mathbb{C}^{n}) when KK has no singularity was proved in [57] (and as can be seen from the above discussion, this property also follows essentially from [64]). As a direct consequence of Theorem 3.12, we record here a Bernstein-Markov type inequality of independent interest. We will not use it anywhere in the paper.

Theorem 3.13.

Let KK be a compact generic nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold in n\mathbb{C}^{n}. Then there exists a constant C>0C>0 such that for every complex polynomial pp on n\mathbb{C}^{n} we have

pL(K)C(degp)2pL(K).\|\nabla p\|_{L^{\infty}(K)}\leq C(\deg p)^{2}\|p\|_{L^{\infty}(K)}. (3.13)

If additionally KK has no singularity (e.g, K=𝕊2n1K=\mathbb{S}^{2n-1}), then

pL(K)CdegppL(K).\|\nabla p\|_{L^{\infty}(K)}\leq C\deg p\,\|p\|_{L^{\infty}(K)}. (3.14)

Note that the exponent of degp\deg p is optimal as it is well-known for the classical Markov and Bernstein inequalities in dimension one. The above result was known when KK is algebraic in n\mathbb{R}^{n}, see [12, 19]. We underline that inequalities similar to those in Theorem 3.13 also hold for other situations (with the same proof), for example, K=𝕊n1nK=\mathbb{S}^{n-1}\subset\mathbb{R}^{n} and considering KK as a maximally totally real submanifold in the complexification of 𝕊n1\mathbb{S}^{n-1}.

Markov (or Bernstein) type inequalities are a subject of great interest in approximation theory. There is a large literature on this topic, e. g. [12, 19, 18, 21, 20, 23, 52, 66], to cite just a few.

Proof.

Let

ϕK:=sup{ψ(n):supKψ0}\phi_{K}:=\sup\{\psi\in\mathcal{L}(\mathbb{C}^{n}):\sup_{K}\psi\leq 0\}

which is 𝒞1\mathcal{C}^{1} if KK has no singularity or 𝒞1/2\mathcal{C}^{1/2} in general by Theorem 3.12. Let pp be a complex polynomial in n\mathbb{C}^{n}. Put k:=degpk:=\deg p. Since 1k(log|p|logmaxK|p|)\frac{1}{k}(\log|p|-\log\max_{K}|p|) is a candidate in the envelope defining ϕK\phi_{K}, we get

|p|ekϕKmaxK|p||p|\leq e^{k\phi_{K}}\max_{K}|p|

on n\mathbb{C}^{n}. We use the same notation CC to denote a constant depending only on K,nK,n. Let a=(a1,,an)Kna=(a_{1},\ldots,a_{n})\in K\subset\mathbb{C}^{n}. Let r>0r>0 be a small constant. Consider the analytic disc Da:=(a1+r𝔻,a2,,an)D_{a}:=(a_{1}+r\mathbb{D},a_{2},\ldots,a_{n}). Applying the Cauchy formula to the restriction of pp to DaD_{a} shows that

|z1p(a)|r1maxDa|p|r1(maxK|p|)maxDaekϕK.|\partial_{z_{1}}p(a)|\leq r^{-1}\max_{D_{a}}|p|\leq r^{-1}(\max_{K}|p|)\max_{D_{a}}e^{k\phi_{K}}.

Since ϕK=0\phi_{K}=0 on KK, using 𝒞1/2\mathcal{C}^{1/2} regularity of ϕK\phi_{K} gives

|z1p(a)|r1maxDa|p|r1(maxK|p|)eCkr1/2|\partial_{z_{1}}p(a)|\leq r^{-1}\max_{D_{a}}|p|\leq r^{-1}(\max_{K}|p|)e^{Ckr^{1/2}}

for some constant C>0C>0 independent of pp and aa. Choosing r=k2r=k^{-2} in the last inequality yields

|z1p(a)|Ck2maxDa|p|.|\partial_{z_{1}}p(a)|\leq Ck^{2}\max_{D_{a}}|p|.

Similarly we also get

|zjp(a)|Ck2maxDa|p||\partial_{z_{j}}p(a)|\leq Ck^{2}\max_{D_{a}}|p|

for every 1jn1\leq j\leq n. Hence the first desired inequality (3.13) for general KK. When KK has no singularity, the arguments are similar. This finishes the proof. ∎

Here is a quantitative version of Lemma 3.4.

Proposition 3.14.

Let KK be a compact generic nondegenerate 𝒞5\mathcal{C}^{5} piecewise-smooth submanifold of XX. Let ϕ\phi be a Hölder continuous function on KK. Then, we have

ϕK,kϕK𝒞0(X)=O(logkk).\big{\|}\phi_{K,k}-\phi_{K}\big{\|}_{\mathscr{C}^{0}(X)}=O\left(\frac{\log k}{k}\right).
Proof.

The desired estimate was proved for K=XK=X in [29, Corollary 4.4]. For the general case, we use the proof of [64, Theorem 2.3] (and [30]). Let ϕ~\tilde{\phi} be the continuous extension of ϕ\phi to XX as in the proof of [64, Theorem 2.3]. It was showed there that ϕK=ϕ~X\phi_{K}=\tilde{\phi}_{X}. On the other hand, one can check directly that ϕK,kϕ~X,k\phi_{K,k}\geq\tilde{\phi}_{X,k}. Hence, we get

|ϕK,kϕK|=ϕKϕK,kϕ~Xϕ~X,k=O(logkk),|\phi_{K,k}-\phi_{K}|=\phi_{K}-\phi_{K,k}\leq\tilde{\phi}_{X}-\tilde{\phi}_{X,k}=O\left(\frac{\log k}{k}\right),

for every kk. This implies the conclusion. ∎

End of the proof of Theorem 2.5.

The desired estimate is deduced directly by using Proposition 3.14, Theorem 2.3, and following the same arguments as in the proof of Proposition 3.6. We just briefly recall here how to do it. Firstly as in the proof of Proposition 3.6, we have

ϕ~K,kϕK,kO(k1).\tilde{\phi}_{K,k}-\phi_{K,k}\geq O(k^{-1}).

It remains to bound from above ϕ~K,kϕK,k\tilde{\phi}_{K,k}-\phi_{K,k}. Combining the polynomial upper bound for BkB_{k} in Theorem 2.3 and (3.10), one gets, for some constants C,N>0C,N>0 independent of kk,

supK(|s|h0kekϕ)CkNsL2(μ,kϕ)\displaystyle\sup_{K}(|s|_{h_{0}^{k}}e^{-k\phi})\leq Ck^{N}\|s\|_{L^{2}(\mu,k\phi)}

for every sH0(X,Lk)s\in H^{0}(X,L^{k}). This coupled with (3.12) yields

|s|h0kkNekϕK,k\displaystyle|s|_{h_{0}^{k}}\leq k^{N}e^{k\phi_{K,k}}

on XX. It follows that

ϕ~K,k=12klogB~kϕK,k+Nlogkk.\tilde{\phi}_{K,k}=\frac{1}{2k}\log\tilde{B}_{k}\leq\phi_{K,k}+N\,\frac{\log k}{k}.

This finishes the proof. ∎

4 Polynomial growth of Bergman kernel functions

This section is devoted to the proof of Theorems 2.4 and 2.3.

4.1 Families of analytic discs attached to KK

The goal of this part is to construct suitable families of analytic discs partly attached to KK. This is actually implicitly contained in [63]. We don’t need all of properties of the family of analytic discs given in [63]. For readers’ convenience we recall briefly the construction below. We will only consider the case where dimK=n\dim K=n in this section.

Here is our result giving the desired family of analytic discs.

Theorem 4.1.

Let C0>0C_{0}>0 be a constant. Then there exist constants C>0,r0>0C>0,r_{0}>0 and θ0(0,π/2)\theta_{0}\in(0,\pi/2) such that for every 0<t<r00<t<r_{0}, the following properties are satisfied. Let p0p_{0} be a regular point of KK of distance at least t/C0t/C_{0} to the singularity of KK, and let Wp0W_{p_{0}} be a local chart around p0p_{0} in XX such that p0p_{0} corresponds to the origin 0 in n\mathbb{C}^{n}. Then there is a 𝒞2\mathcal{C}^{2} map F:𝔻¯×𝔹n1Wp0F:\overline{\mathbb{D}}\times\mathbb{B}_{n-1}\to W_{p_{0}} such that the following properties are fulfilled:

(i) F(,y)F(\cdot,y) is holomorphic for every y𝔹n1y\in\mathbb{B}_{n-1}, and

DF(ξ,y)Ct,\|DF(\xi,y)\|\leq Ct,

for every ξ𝔻\xi\in\mathbb{D}, y𝔹n1y\in\mathbb{B}_{n-1}.

(ii) F(eiθ,y)KF(e^{i\theta},y)\in K for θ0θθ0-\theta_{0}\leq\theta\leq\theta_{0} and y𝔹n1y\in\mathbb{B}_{n-1}, and

|F(1,y)|t/C0,|F(1,y)|\leq t/C_{0},

(iii) Let GG denote the restriction of FF to [eiθ0,eiθ0]×𝔹n1[e^{-i\theta_{0}},e^{i\theta_{0}}]\times\mathbb{B}_{n-1}. Then GG is bijective onto its image, and the image of GG is contained in 𝔹(p0,t/C0)K\mathbb{B}(p_{0},t/C_{0})\cap K (here 𝔹(p0,t/C0)\mathbb{B}(p_{0},t/C_{0}) denotes the ball centred at p0p_{0} of radius t/C0t/C_{0} in XX), and

C1tn|detDG(eiθ,y)|Ctn,C^{-1}t^{n}\leq\big{|}\det DG(e^{i\theta},y)\big{|}\leq Ct^{n},

for every θ0θθ0-\theta_{0}\leq\theta\leq\theta_{0} and y𝔹n1y\in\mathbb{B}_{n-1}.

We proceed with the proof of the last theorem. Denote by z=x+iyz=x+iy the complex variable on \mathbb{C} and by ξ=eiθ\xi=e^{i\theta} the variable on 𝔻.\partial\mathbb{D}. For any mm\in\mathbb{N} and r>0,r>0, let 𝔹m(0,r)\mathbb{B}_{m}(0,r) be the Euclidean ball centered at 0 of radius rr of m\mathbb{R}^{m}, and for r=1r=1, we write 𝔹m\mathbb{B}_{m} for 𝔹m(0,1)\mathbb{B}_{m}(0,1). Let ZZ be a compact submanifold with or without boundary of m.\mathbb{R}^{m}. The Euclidean metric on m\mathbb{R}^{m} induces a metric on Z.Z. For β(0,1]\beta\in(0,1] and kk\in\mathbb{N}, let 𝒞k,β(Z)\mathcal{C}^{k,\beta}(Z) be the space of real-valued functions on ZZ which are differentiable up to the order kk and whose kthk^{th} derivatives are Hölder continuous of order β.\beta. For any tuple v=(v0,,vm)v=(v_{0},\cdots,v_{m}) consisting of functions in 𝒞k,β(Z)\mathcal{C}^{k,\beta}(Z), we define its 𝒞k,β\mathcal{C}^{k,\beta}-norm to be the maximum of the ones of its components.

Let u0u_{0} be a continuous function on 𝔻\partial\mathbb{D}. Let

𝒞u0(z):=12πππu0(eiθ)eiθ+zeiθz𝑑θ\mathcal{C}u_{0}(z):=\frac{1}{2\pi}\int_{-\pi}^{\pi}u_{0}(e^{i\theta})\frac{e^{i\theta}+z}{e^{i\theta}-z}d\theta

which is a holomorphic function on 𝔻\mathbb{D}. Recall that the real part of 𝒞u0\mathcal{C}u_{0} is u0u_{0}. Let 𝒯u0(z)\mathcal{T}u_{0}(z) denotes the imaginary part of 𝒞u0(z)\mathcal{C}u_{0}(z). Put

𝒯1u0:=𝒯u0𝒯u0(1).\mathcal{T}_{1}u_{0}:=\mathcal{T}u_{0}-\mathcal{T}u_{0}(1).

For basic properties of 𝒯1\mathcal{T}_{1}, one can consult [51] or [2].

We now go back to our current situation with XX. We endow XX with an arbitrary Riemannian metric. For θ0(0,π)\theta_{0}\in(0,\pi), let [eiθ0,eiθ0][e^{-i\theta_{0}},e^{i\theta_{0}}] denotes the arc of 𝔻\partial\mathbb{D} of arguments from θ0-\theta_{0} to θ0\theta_{0}. Let p0p_{0} be a regular point in KK and let rp0r_{p_{0}} denotes the distance of p0p_{0} to the singular part of KK. Recall that we assume in this section that dimK=n\dim K=n.

Lemma 4.2.

There exist a constant cK>1c_{K}>1 depending only on (K,X)(K,X) and a local chart (Wp0,Ψ)(W_{p_{0}},\Psi) around p0,p_{0}, where Ψ:Wp0𝔹2n\Psi:W_{p_{0}}\rightarrow\mathbb{B}_{2n} is biholomorphic with Ψ(p0)=0\Psi(p_{0})=0 such that the two following conditions hold:

(i)(i) we have

Ψ𝒞5cK,Ψ1𝒞5cK,\displaystyle\|\Psi\|_{\mathcal{C}^{5}}\leq c_{K},\quad\|\Psi^{-1}\|_{\mathcal{C}^{5}}\leq c_{K},

(ii)(ii) there is a 𝒞3\mathcal{C}^{3} map hh from 𝔹¯n\overline{\mathbb{B}}_{n} to n\mathbb{R}^{n} so that h(0)=Dh(0)=0h(0)=Dh(0)=0, and

Ψ(KWp0){(𝐱,h(𝐱)):𝐱𝔹¯n(0,rp0/cK)},\Psi(K\cap W_{p_{0}})\supset\big{\{}(\mathbf{x},h(\mathbf{x})):\mathbf{x}\in\overline{\mathbb{B}}_{n}(0,r_{p_{0}}/c_{K})\big{\}},

where the canonical coordinates on n=n+in\mathbb{C}^{n}=\mathbb{R}^{n}+i\mathbb{R}^{n} are denoted by 𝐳=𝐱+i𝐲,\mathbf{z}=\mathbf{x}+i\mathbf{y}, and

h𝒞3cK.\displaystyle\|h\|_{\mathcal{C}^{3}}\leq c_{K}. (4.1)

Note that hh is indeed 𝒞5\mathcal{C}^{5} (because KK is so), but 𝒞3\mathcal{C}^{3} is sufficient for our purpose in what follows.

Proof.

The existence of local coordinates so that h(0)=Dh(0)=0h(0)=Dh(0)=0 is standard, see [3] or [64, Lemma 4.1]. Perhaps one needs to explain a bit about the radius rp0/cKr_{p_{0}}/c_{K}: the existence of cKc_{K} comes from the fact that for every singular point aa in KK, there are an open neighorhood UU of aa in XX and sets AjBjUA_{j}\Subset B_{j}\subset U for 1jm1\leq j\leq m such that BjB_{j} is 𝒞5\mathcal{C}^{5} smooth generic CR submanifold of dimension dimK\dim K in UU, and

KU=j=1m(AjU),K\cap U=\cup_{j=1}^{m}(A_{j}\cap U),

and AjA_{j} is the closure in UU of a relatively compact open subset in BjB_{j}, and Aj\partial A_{j} (in UU) is contained in the singularity of KK. Thus by applying the standard local coordinates to points in BjB_{j} we obtain the existence of cKc_{K}. This finishes the proof. ∎

From now on, we only use the local coordinates introduced in Lemma 4.2 and identify points in Wp0W_{p_{0}} with those in 𝔹2n\mathbb{B}_{2n} via Ψ.\Psi.

Lemma 4.3 ([64, Lemma 3.1]).

There exist a function u0𝒞(𝔻)u_{0}\in\mathcal{C}^{\infty}(\partial\mathbb{D}) such that u0(eiθ)=0u_{0}(e^{i\theta})=0 for θ[π/2,π/2]\theta\in[-\pi/2,\pi/2] and xu0(1)=1\partial_{x}u_{0}(1)=-1.

In what follows, we identify n\mathbb{C}^{n} with n+in.\mathbb{R}^{n}+i\mathbb{R}^{n}. Let u0u_{0} be a function described in Lemma 4.3. Let 𝝉1,𝝉2𝔹¯n1n1.\boldsymbol{\tau}_{1},\boldsymbol{\tau}_{2}\in\overline{\mathbb{B}}_{n-1}\subset\mathbb{R}^{n-1}. Define 𝝉1:=(1,𝝉1)n\boldsymbol{\tau}^{*}_{1}:=(1,\boldsymbol{\tau}_{1})\in\mathbb{R}^{n} and 𝝉2:=(0,𝝉2)n\boldsymbol{\tau}^{*}_{2}:=(0,\boldsymbol{\tau}_{2})\in\mathbb{R}^{n} and 𝝉:=(𝝉1,𝝉2).\boldsymbol{\tau}:=(\boldsymbol{\tau}_{1},\boldsymbol{\tau}_{2}). Let tt be a positive number in (0,1](0,1] which plays a role as a scaling parameter in the equation (4.2) below.

In order to construct an analytic disc partly attached to KK, it suffices to find a map

U:𝔻𝔹nn,U:\partial\mathbb{D}\rightarrow\mathbb{B}_{n}\subset\mathbb{R}^{n},

which is Hölder continuous, satisfying the following Bishop-type equation

U𝝉,t(ξ)=t𝝉2𝒯1(h(U𝝉,t))(ξ)t𝒯1u0(ξ)𝝉1,\displaystyle U_{\boldsymbol{\tau},t}(\xi)=t\boldsymbol{\tau}^{*}_{2}-\mathcal{T}_{1}\big{(}h(U_{\boldsymbol{\tau},t})\big{)}(\xi)-t\mathcal{T}_{1}u_{0}(\xi)\,\boldsymbol{\tau}^{*}_{1}, (4.2)

where the Hilbert transform 𝒯1\mathcal{T}_{1} is extended to a vector valued function by acting on each component. The existence of solution of the last equation is a standard fact in the Cauchy-Riemann geometry.

Proposition 4.4 ([63, Proposition 3.3]).

There are a positive number t1(0,1)t_{1}\in(0,1) and a real number c1>0c_{1}>0 satisfying the following property: for any t(0,t1]t\in(0,t_{1}] and any 𝛕𝔹¯n12,\boldsymbol{\tau}\in\overline{\mathbb{B}}_{n-1}^{2}, the equation (4.2) has a unique solution U𝛕,tU_{\boldsymbol{\tau},t} which is 𝒞2,12\mathcal{C}^{2,\frac{1}{2}} in (ξ,𝛕)(\xi,\boldsymbol{\tau}) and such that

D(ξ,𝝉)jU𝝉,t𝒞12(𝔻)c1t,\displaystyle\|D^{j}_{(\xi,\boldsymbol{\tau})}U_{\boldsymbol{\tau},t}\|_{\mathcal{C}^{\frac{1}{2}}(\partial\mathbb{D})}\leq c_{1}t, (4.3)

for any 𝛕𝔹¯n12\boldsymbol{\tau}\in\overline{\mathbb{B}}_{n-1}^{2} and j=0,1j=0,1.

From now on, we consider t<min{t1,C0rp0}t<\min\{t_{1},C_{0}r_{p_{0}}\} (hence the distance from p0p_{0} to the singularity of KK is at least t/C0t/C_{0}). Let U𝝉,tU_{\boldsymbol{\tau},t} be the unique solution of (4.2). For simplicity, we use the same notation U𝝉,t(z)U_{\boldsymbol{\tau},t}(z) to denote the harmonic extension of U𝝉,t(ξ)U_{\boldsymbol{\tau},t}(\xi) to 𝔻.\mathbb{D}. Let P𝝉,t(z)P_{\boldsymbol{\tau},t}(z) be the harmonic extension of h(U𝝉,t(ξ))h\big{(}U_{\boldsymbol{\tau},t}(\xi)\big{)} to 𝔻.\mathbb{D}. Recall that

Lemma 4.5.

([63, Lemma 3.4]) There exists a constant c2c_{2} so that for every t(0,t1]t\in(0,t_{1}] and every (z,𝛕)𝔻¯×𝔹¯n12,(z,\boldsymbol{\tau})\in\overline{\mathbb{D}}\times\overline{\mathbb{B}}_{n-1}^{2}, we have

D(z,𝝉)jU𝝉,t(z)c2tandD(z,𝝉)jP𝝉,t(z)c2t2,\displaystyle\|D^{j}_{(z,\boldsymbol{\tau})}U_{\boldsymbol{\tau},t}(z)\|\leq c_{2}t\quad\text{and}\quad\|D^{j}_{(z,\boldsymbol{\tau})}P_{\boldsymbol{\tau},t}(z)\|\leq c_{2}t^{2}, (4.4)

for j=0,1.j=0,1.

We note that the hypothesis that D2h(0)=0D^{2}h(0)=0 was required in [63], but it is actually superfluous in the proof of Lemma 4.5. Define

F(z,𝝉,t):=U𝝉,t(z)+iP𝝉,t(z)+itu0(z)𝝉1F(z,\boldsymbol{\tau},t):=U_{\boldsymbol{\tau},t}(z)+iP_{\boldsymbol{\tau},t}(z)+it\,u_{0}(z)\,\boldsymbol{\tau}^{*}_{1}

which is a family of analytic discs parametrized by (𝝉,t).(\boldsymbol{\tau},t). Compute

F(1,𝝉,t)=U(𝝉,t)(1)+iP𝝉,t(1)=t𝝉2+h(t𝝉2).F(1,\boldsymbol{\tau},t)=U(\boldsymbol{\tau},t)(1)+iP_{\boldsymbol{\tau},t}(1)=t\boldsymbol{\tau}^{*}_{2}+h(t\boldsymbol{\tau}^{*}_{2}).

Hence if |𝝉2||\boldsymbol{\tau}_{2}| is small enough, we see that

|F(1,𝝉,t)|t|𝝉2|<rp0/(2cK).\displaystyle|F(1,\boldsymbol{\tau},t)|\leq t|\boldsymbol{\tau}_{2}|<r_{p_{0}}/(2c_{K}). (4.5)

This combined with Lemma 4.5 yields that

|U𝝉,t(eiθ)||θ||U𝝉,t(1)||θ|rp0/(2cK)<rp0/cK\displaystyle|U_{\boldsymbol{\tau},t}(e^{i\theta})|\leq|\theta||U_{\boldsymbol{\tau},t}(1)|\leq|\theta|r_{p_{0}}/(2c_{K})<r_{p_{0}}/c_{K} (4.6)

if θ\theta and 𝝉2\boldsymbol{\tau}_{2} are small enough. Now the defining formula of FF and the fact that u00u_{0}\equiv 0 on [eiπ/2,eiπ/2][e^{-i\pi/2},e^{i\pi/2}] imply that

F(ξ,𝝉,t)=U𝝉,t(ξ)+iP𝝉,t(ξ)=U𝝉,t(ξ)+ih(U𝝉,t(ξ))KF(\xi,\boldsymbol{\tau},t)=U_{\boldsymbol{\tau},t}(\xi)+iP_{\boldsymbol{\tau},t}(\xi)=U_{\boldsymbol{\tau},t}(\xi)+ih\big{(}U_{\boldsymbol{\tau},t}(\xi)\big{)}\in K

by (4.6) if θ\theta and 𝝉2\boldsymbol{\tau}_{2} are small enough. In other words, there is a small constant θ0\theta_{0} such that if |𝝉2|<θ0|\boldsymbol{\tau}_{2}|<\theta_{0}, then FF is [eiθ0,eiθ0][e^{-i\theta_{0}},e^{i\theta_{0}}]-attached to KK.

Proposition 4.6.

By decreasing θ0\theta_{0} and t1t_{1} if necessary, we obtain the following property: for every 𝛕1𝔹¯n1,\boldsymbol{\tau}_{1}\in\overline{\mathbb{B}}_{n-1}, the map F(,𝛕1,,t):[eiθ0,eiθ0]×𝔹¯n1(0,θ0)KF(\cdot,\boldsymbol{\tau}_{1},\cdot,t):[e^{-i\theta_{0}},e^{i\theta_{0}}]\times\overline{\mathbb{B}}_{n-1}(0,\theta_{0})\rightarrow K is a diffeomorphism onto its image, and

C1tndetDF(,𝝉1,,t)LCtn,C^{-1}t^{n}\leq\|\det DF(\cdot,\boldsymbol{\tau}_{1},\cdot,t)\|_{L^{\infty}}\leq Ct^{n},

for some constant C>0C>0 independent of t,𝛕1t,\boldsymbol{\tau}_{1}.

Proof.

The desired assertion was implicitly obtained in the proof of [63, Proposition 3.5]. We present here complete arguments for readers’ convenience. Recall

F(eiθ,𝝉1,t)=U𝝉,t(eiθ)+ih(U𝝉,t(eiθ)F(e^{i\theta},\boldsymbol{\tau}_{1},t)=U_{\boldsymbol{\tau},t}(e^{i\theta})+ih(U_{\boldsymbol{\tau},t}(e^{i\theta})

By the Cauchy-Riemann equations, we have

yU𝝉,t(1)=txu0(1)𝝉1xP𝝉,t(1)=t𝝉1xP𝝉,t(1).\partial_{y}U_{\boldsymbol{\tau},t}(1)=-t\partial_{x}u_{0}(1)\boldsymbol{\tau}^{*}_{1}-\partial_{x}P_{\boldsymbol{\tau},t}(1)=t\boldsymbol{\tau}^{*}_{1}-\partial_{x}P_{\boldsymbol{\tau},t}(1).

The last term is O(t2)O(t^{2}) by Lemma 4.5. Thus the first component of yU𝝉,t(1)\partial_{y}U_{\boldsymbol{\tau},t}(1) is greater than t/2t/2 provided that tt2t\leq t_{2} small enough. A direct computation gives yU𝝉,t(1)=θU𝝉,t(1).\partial_{y}U_{\boldsymbol{\tau},t}(1)=\partial_{\theta}U_{\boldsymbol{\tau},t}(1). Consequently, the first component of

θF(eiθ,𝝉1,t)=θU𝝉,t(1)+iθh(U𝝉,t(1))\partial_{\theta}F(e^{i\theta},\boldsymbol{\tau}_{1},t)=\partial_{\theta}U_{\boldsymbol{\tau},t}(1)+i\partial_{\theta}h\big{(}U_{\boldsymbol{\tau},t}(1)\big{)}

is greater than t/2t/2 for tt2t\leq t_{2} (note that Dh(0)=0Dh(0)=0). Moreover, as computed above, we have

F(1,𝝉1,t)=t𝝉2+h(t𝝉2).F(1,\boldsymbol{\tau}_{1},t)=t\boldsymbol{\tau}_{2}^{*}+h(t\boldsymbol{\tau}_{2}^{*}).

Thus D𝝉2,θF(1,𝝉,t)D_{\boldsymbol{\tau}_{2},\theta}F(1,\boldsymbol{\tau},t) is a nondegenerate matrix whose determinant satisfies the desired inequalities if |θ|<θ0|\theta|<\theta_{0} is small enough. The proof is finished. ∎

End of proof of Theorem 4.1.

Let θ0\theta_{0} be as above and smaller than θ1\theta_{1} and M>|θ0|1M>|\theta_{0}|^{-1} be a big constant. Fix a parameter 𝝉1\boldsymbol{\tau}_{1} and define

Ft(ξ,𝝉2):=F(ξ,𝝉1,𝝉2/M,t).F_{t}(\xi,\boldsymbol{\tau}_{2}):=F(\xi,\boldsymbol{\tau}_{1},\boldsymbol{\tau}_{2}/M,t).

By the above results, we see that the family FtF_{t} satisfies all of required properties (because |𝝉2/M|<θ0|\boldsymbol{\tau}_{2}/M|<\theta_{0}). This finishes the proof. ∎

4.2 Upper bound of Bergman kernel functions

We start with the following useful estimate in one dimension.

Lemma 4.7.

Let β(0,1)\beta\in(0,1). Then there exists a constant Cβ>0C_{\beta}>0 such that for every θ0(0,π]\theta_{0}\in(0,\pi] and every constant M>0M>0, and every subharmonic function gg on 𝔻\mathbb{D} such that gg is continuous up to 𝔻\partial\mathbb{D} and |g(ξ)|M|g(\xi)|\leq M for ξ𝔻\{eiθ:θ0θθ0}\xi\in\partial\mathbb{D}\backslash\{e^{i\theta}:-\theta_{0}\leq\theta\leq\theta_{0}\}, we have

g(z)Cβ[|1z|βθ01M+(1|z|)1θ0θ0g(eiθ)𝑑θ].g(z)\leq C_{\beta}\bigg{[}|1-z|^{\beta}\theta_{0}^{-1}M+(1-|z|)^{-1}\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta})d\theta\bigg{]}.
Proof.

Let θ1[0,2π)\theta_{1}\in[0,2\pi). Put

I:={eiθ:θ0θθ0},I:={eiθ:θ0/4θθ0/4}.I:=\{e^{i\theta}:-\theta_{0}\leq\theta\leq\theta_{0}\},\quad I^{\prime}:=\{e^{i\theta}:-\theta_{0}/4\leq\theta\leq\theta_{0}/4\}.

Let g1g_{1} be the harmonic function on 𝔻\mathbb{D} such that g1𝒞0,1(𝔻)g_{1}\in\mathcal{C}^{0,1}(\partial\mathbb{D}), and g1(ξ)=Mg_{1}(\xi)=M for ξ𝔻\I\xi\in\partial\mathbb{D}\backslash I and g10g_{1}\equiv 0 on II^{\prime}. Observe that g1𝒞0,1(𝔻)Cθ01M\|g_{1}\|_{\mathcal{C}^{0,1}(\partial\mathbb{D})}\leq C\theta_{0}^{-1}M for some constant C>0C>0 independent of M,θ1,θ0M,\theta_{1},\theta_{0}.

By a classical result on harmonic functions on the unit disc (see [36, Page 41] or (3.4) in [64]), we have

g1𝒞0,β(𝔻)g1𝒞0,β(𝔻)θ01M\|g_{1}\|_{\mathcal{C}^{0,\beta}(\mathbb{D})}\lesssim\|g_{1}\|_{\mathcal{C}^{0,\beta}(\partial\mathbb{D})}\lesssim\theta_{0}^{-1}M

for every β(0,1)\beta\in(0,1). As a result, we get

g1(z)=g1(z)g1(1)|1z|βg1𝒞0,β(𝔻)|1z|βθ01M.\displaystyle g_{1}(z)=g_{1}(z)-g_{1}(1)\lesssim|1-z|^{\beta}\|g_{1}\|_{\mathcal{C}^{0,\beta}(\partial\mathbb{D})}\lesssim|1-z|^{\beta}\theta_{0}^{-1}M. (4.7)

Let g2g_{2} be the harmonic function on 𝔻\mathbb{D} such that g2(eiθ)=0g_{2}(e^{i\theta})=0 for |θ|>θ0|\theta|>\theta_{0}, and g2(eiθ)=g(eiθ)g_{2}(e^{i\theta})=g(e^{i\theta}) for θ[θ0,θ0]\theta\in[-\theta_{0},\theta_{0}]. Observe that gg1+g2g\leq g_{1}+g_{2} because the latter function is harmonic and greater than or equal to gg on the boundary of 𝔻\mathbb{D}. Using Poisson’s formula, we see that

g2(z)(1|z|)1ππg2(eiθ)𝑑θ=(1|z|)1θ0θ0g(eiθ)𝑑θ.g_{2}(z)\leq(1-|z|)^{-1}\int_{-\pi}^{\pi}g_{2}(e^{i\theta})d\theta=(1-|z|)^{-1}\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta})d\theta.

Summing this and (4.7) gives the desired assertion. The proof is finished. ∎

Let KK be a 𝒞5\mathcal{C}^{5} smooth (without boundary) maximally totally real submanifold in XX. Let sH0(X,Lk)s\in H^{0}(X,L^{k}) with sL2(μ,hk)=1\|s\|_{L^{2}(\mu,h^{k})}=1 and M:=supK|s|hk2M:=\sup_{K}|s|^{2}_{h^{k}}. Let p0Kp_{0}\in K. Consider a local chart (U,𝐳)(U,\mathbf{z}) around p0p_{0} with coordinates 𝐳\mathbf{z}, and p0p_{0} corresponds to the origin 0 in n\mathbb{C}^{n}. Shrinking UU if necessary we can assume also that LL is trivial on UU.

We trivialize (L,h0)(L,h_{0}) over UU such that h0=eψh_{0}=e^{-\psi} for some psh function ψ\psi on UU with ψ(0)=ϕ(0)\psi(0)=-\phi(0) (we implicitly fix a local holomorphic frame on L|UL|_{U} so that one can identify Hermitian metrics on L|UL|_{U} with functions on UU), and identify ss with a holomorphic function gsg_{s} on UU. Thus

h=eϕh0=eϕψh=e^{-\phi}h_{0}=e^{-\phi-\psi}

on UU. In particular h(0)=1h(0)=1 on UU.

Lemma 4.8.

There exists a constant C1>0C_{1}>0 independent of k,s,p0k,s,p_{0} such that

sup{|𝐳|1/k}|s(𝐳)|hk2C1M,\displaystyle\sup_{\{|\mathbf{z}|\leq 1/k\}}|s(\mathbf{z})|^{2}_{h^{k}}\leq C_{1}M, (4.8)

and

C11|s(𝐳)|hk2|gs(𝐳)|2C1|s(𝐳)|hkC12M,\displaystyle C_{1}^{-1}|s(\mathbf{z})|_{h^{k}}^{2}\leq|g_{s}(\mathbf{z})|^{2}\leq C_{1}|s(\mathbf{z})|_{h^{k}}\leq C_{1}^{2}M, (4.9)

for 𝐳𝔹(0,k1)\mathbf{z}\in\mathbb{B}(0,k^{-1}). Moreover for every constant ϵ>0\epsilon>0, there exist a constant cϵ>0c_{\epsilon}>0 independent of k,s,p0k,s,p_{0} such that

|gs(0)|2|gs(𝐳)|2+ϵM,|g_{s}(0)|^{2}\leq|g_{s}(\mathbf{z})|^{2}+\epsilon M,

for |𝐳|1/(cϵk)|\mathbf{z}|\leq 1/(c_{\epsilon}k).

Proof.

The desired inequality (4.9) follows immediately from (4.8) and the equalities

|gs(𝐳)|2=|s(𝐳)|hk2ek(ψ(𝐳)+ϕ(𝐳)),ψ(0)+ϕ(0)=0.|g_{s}(\mathbf{z})|^{2}=|s(\mathbf{z})|_{h^{k}}^{2}e^{k(\psi(\mathbf{z})+\phi(\mathbf{z}))},\quad\psi(0)+\phi(0)=0.

Recall that

ϕK:=sup{ψω-psh:ψϕon K}.\phi_{K}:=\sup\{\psi\quad\text{$\omega$-psh}:\psi\leq\phi\quad\text{on }\quad K\}.

Note that ϕKϕ\phi_{K}\leq\phi on KK. By hypothesis ϕ𝒞1,δ(K)\phi\in\mathcal{C}^{1,\delta}(K) for some constant δ>0\delta>0. This combined with Theorem 3.11 and the fact that KK has no singularity yields that ϕK\phi_{K} is Lipschitz. Using the fact that k1log|s|h0kk^{-1}\log|s|_{h^{k}_{0}} is ω\omega-psh, we get the Bernstein-Walsh inequality

|s|h0k2(supK|s|hk2)e2kϕK=Me2kϕK.|s|_{h_{0}^{k}}^{2}\leq\big{(}\sup_{K}|s|^{2}_{h^{k}}\big{)}e^{2k\phi_{K}}=Me^{2k\phi_{K}}.

Hence

|s|hk2Me2k(ϕKϕ).|s|^{2}_{h^{k}}\leq Me^{2k(\phi_{K}-\phi)}.

This combined with the Lipschitz property of ϕK\phi_{K} and ϕ\phi (and also the property that ϕK(0)ϕ(0)0\phi_{K}(0)-\phi(0)\leq 0 on KK) yields

sup{𝐳:|𝐳|1/k}|s(𝐳)|hk2C1M,\displaystyle\sup_{\{\mathbf{z}:|\mathbf{z}|\leq 1/k\}}|s(\mathbf{z})|^{2}_{h^{k}}\leq C_{1}M, (4.10)

for some constant C1C_{1} independent of k,s,p0k,s,p_{0}. Hence (4.8) also follows.

By arguing as in the proof of Theorem 3.13, one obtains the following version of Bernstein-Markov inequality: for 𝐳𝔹(0,k1/2)\mathbf{z}\in\mathbb{B}(0,k^{-1}/2), there holds

|gs(𝐳)|ksup𝔹(0,k1)|gs|kM1/2.\displaystyle|\nabla g_{s}(\mathbf{z})|\lesssim k\sup_{\mathbb{B}(0,k^{-1})}|g_{s}|\lesssim kM^{1/2}. (4.11)

Consequently, for 𝐳𝔹(0,k1/cϵ)\mathbf{z}\in\mathbb{B}(0,k^{-1}/c_{\epsilon}) with cϵc_{\epsilon} big enough, we get

|gs(𝐳)gs(0)||𝐳|sup𝔹(0,k1/C0)|gs|ϵ1/2M1/2.|g_{s}(\mathbf{z})-g_{s}(0)|\leq|\mathbf{z}|\sup_{\mathbb{B}(0,k^{-1}/C_{0})}|\nabla g_{s}|\leq\epsilon^{1/2}M^{1/2}.

This finishes the proof. ∎

Proof of the upper bound in Theorem 2.4.

Let sH0(X,Lk)s\in H^{0}(X,L^{k}) with sL2(μ,hk)=1\|s\|_{L^{2}(\mu,h^{k})}=1. Let M:=sL(K,hk)2M:=\|s\|^{2}_{L^{\infty}(K,h^{k})}. We need to prove that MknM\lesssim k^{n}.

Let p0Kp_{0}\in K and consider a local chart (U,𝐳)(U,\mathbf{z}) around p0p_{0} with coordinates 𝐳\mathbf{z}, the point p0p_{0} corresponds to 0 in the local chart (U,𝐳)(U,\mathbf{z}). Let ϵ>0\epsilon>0 be a small constant to be chosen later. Let C1,cϵC_{1},c_{\epsilon} be the constants in Lemma 4.8. Let AC12ϵ2A\geq C_{1}^{2}\epsilon^{-2} be a big constant. Using the Lipschitz continuity of ϕ\phi yields

{|𝐳|k1}|gs|2𝑑μ={|𝐳|k1}|s|hk2ek(ψ+ϕ)𝑑μ1\displaystyle\int_{\{|\mathbf{z}|\leq k^{-1}\}}|g_{s}|^{2}d\mu=\int_{\{|\mathbf{z}|\leq k^{-1}\}}|s|_{h^{k}}^{2}e^{k(\psi+\phi)}d\mu\lesssim 1 (4.12)

(uniformly in ss). Now let F:𝔻×YXF:\mathbb{D}\times Y\to X be the family of analytic discs in Theorem 4.1 associated to C0:=2cϵC_{0}:=2c_{\epsilon} for p0,p_{0}, and t:=k1t:=k^{-1}, where Y:=𝔹n1Y:=\mathbb{B}_{n-1}. Let θ0(0,π/2)\theta_{0}\in(0,\pi/2) be the constant in the last theorem.

Let g:=|gsF|2g:=|g_{s}\circ F|^{2}. Put 𝐳y:=F(11/A,y)\mathbf{z}_{y}:=F(1-1/A,y). By expressing

𝐳y=F(11/A,y)F(1,y)+F(1,y),\mathbf{z}_{y}=F(1-1/A,y)-F(1,y)+F(1,y),

one gets

|𝐳y|1/(cϵk)|\mathbf{z}_{y}|\leq 1/(c_{\epsilon}k)

if AA is big enough. Applying Lemma 4.7 to g(,y)g(\cdot,y) and β=1/2\beta=1/2 and using (4.9) yield

|gs(𝐳y)|2=g(ξy,y)C(A1/2M+Aθ0θ0g(eiθ,y)𝑑θ),|g_{s}(\mathbf{z}_{y})|^{2}=g(\xi_{y},y)\leq C^{\prime}\big{(}A^{-1/2}M+A\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta\big{)},

for some constant C>0C^{\prime}>0 independent of s,p0,ks,p_{0},k. This combined with Lemma 4.8 yields

|gs(0)|2ϵM+CA1/2M+CAθ0θ0g(eiθ,y)𝑑θ.|g_{s}(0)|^{2}\leq\epsilon M+C^{\prime}A^{-1/2}M+C^{\prime}A\int_{\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta.

Integrating the last inequality over yYy\in Y gives

|gs(0)|2\displaystyle|g_{s}(0)|^{2} ϵM+CA1/2M+CA(vol(Y))1Yvolyθ0θ0g(eiθ,y)𝑑θ\displaystyle\leq\epsilon M+C^{\prime}A^{-1/2}M+C^{\prime}A(\mathop{\mathrm{vol}}(Y))^{-1}\int_{Y}vol_{y}\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta
2ϵM+CA(vol(Y))1Yvolyθ0θ0g(eiθ,y)𝑑θ\displaystyle\leq 2\epsilon M+C^{\prime}A(\mathop{\mathrm{vol}}(Y))^{-1}\int_{Y}vol_{y}\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta

if Aϵ2C2A\geq\epsilon^{-2}C^{\prime 2}. By Properties of FF and (4.12), the second term in the right-hand side of the last inequality is kn\lesssim k^{n}. We infer that

|gs(0)|22ϵM+AC2kn|g_{s}(0)|^{2}\leq 2\epsilon M+AC_{2}k^{n}

where C2C_{2} is a constant independent of k,s,p0k,s,p_{0}. Consequently by (4.9), one gets

|s(p0)|hk22C1ϵM+AC2C1kn|s(p_{0})|^{2}_{h^{k}}\leq 2C_{1}\epsilon M+AC_{2}C_{1}k^{n}

for every p0Kp_{0}\in K. By choosing ϵ:=1/(4C1)\epsilon:=1/(4C_{1}) and AA big enough as required, one gets

MM/2+AC1C2kn.M\leq M/2+AC_{1}C_{2}k^{n}.

Thus the desired upper bound for MM follows. The proof is finished. ∎

We now proceed with the proof of Theorem 2.3.

End of proof of Theorem 2.3 for dimK=n\dim K=n.

We assume dimK=n\dim K=n. We will explain how to treat the case dimKn\dim K\geq n later. Let sH0(X,Lk)s\in H^{0}(X,L^{k}) with sL2(μ,hk)=1\|s\|_{L^{2}(\mu,h^{k})}=1. Put M:=supK|s|hkM:=\sup_{K}|s|_{h^{k}}. Let α(0,1)\alpha\in(0,1) be a Hölder exponent of ϕ\phi. We have ϕK𝒞α/2\phi_{K}\in\mathcal{C}^{\alpha/2} (note α<1\alpha<1). We follow essentially the scheme of the proof for the upper bound of BkB_{k} in Theorem 2.4.

Denote by Kk,αK_{k,\alpha} the set of points in KK of distance at least k2/αk^{-2/\alpha} to the singular part of KK, and Uk,αU_{k,\alpha} the set of points in XX of distance at most k2/αk^{-2/\alpha} to KK. As in the proof of Lemma 4.8, one has

suppUk,α|s(p)|hk2CM,MCsuppKk,α|s(p)|hk2,\displaystyle\sup_{p\in U_{k,\alpha}}|s(p)|^{2}_{h^{k}}\leq CM,\quad M\leq C\sup_{p\in K_{k,\alpha}}|s(p)|^{2}_{h^{k}}, (4.13)

for some constant C4C\geq 4 independent of k,sk,s. In particular it is sufficient to estimate |s(p)|hk2|s(p)|^{2}_{h^{k}} for pKk,αp\in K_{k,\alpha}.

Let p0Kk,αp_{0}\in K_{k,\alpha} (hence p0p_{0} is a regular point of KK, and the ball 𝔹(p0,k2/α)K\mathbb{B}(p_{0},k^{-2/\alpha})\cap K lies entirely in the regular part of KK) and a local chart (U,𝐳)(U,\mathbf{z}) around p0p_{0} with coordinates 𝐳\mathbf{z}, the point p0p_{0} corresponds to 0 in the local chart (U,𝐳)(U,\mathbf{z}).

We trivialize (L,h0)(L,h_{0}) over UU such that h0=eψh_{0}=e^{-\psi} for some smooth psh function ψ\psi on UU with ψ(0)=ϕ(0)\psi(0)=-\phi(0) (we implicitly fix a local holomorphic frame on L|UL|_{U} so that one can identify Hermitian metrics on L|UL|_{U} with functions on UU), and identify ss with a holomorphic function gsg_{s} on UU. Thus

h=eϕh0=eϕψh=e^{-\phi}h_{0}=e^{-\phi-\psi}

on UU. In particular h(0)=1h(0)=1 on UU. Using the Hölder continuity of ϕ\phi yields

{|𝐳|k2/α}|gs|2𝑑μ={|𝐳|k2/α}|s|hk2ek(ψ+ϕ)𝑑μ1\displaystyle\int_{\{|\mathbf{z}|\leq k^{-2/\alpha}\}}|g_{s}|^{2}d\mu=\int_{\{|\mathbf{z}|\leq k^{-2/\alpha}\}}|s|_{h^{k}}^{2}e^{k(\psi+\phi)}d\mu\lesssim 1

(uniformly in s,ks,k). As in the proof of Lemma 4.8, a Berstein-Walsh type inequality implies that by increasing CC if necessary (independent of p0,k,sp_{0},k,s) there holds

C1|s(𝐳)|hk2|gs(𝐳)|2=|s(𝐳)|hk2ek(ψ(𝐳)+ϕ(𝐳))CM,\displaystyle C^{-1}|s(\mathbf{z})|_{h^{k}}^{2}\leq|g_{s}(\mathbf{z})|^{2}=|s(\mathbf{z})|_{h^{k}}^{2}e^{k(\psi(\mathbf{z})+\phi(\mathbf{z}))}\leq C\,M, (4.14)

for 𝐳𝔹(0,k2/α)\mathbf{z}\in\mathbb{B}(0,k^{-2/\alpha}). Let λ:=λ/(1+λ)\lambda^{\prime}:=\lambda/(1+\lambda). One also sees that there is a constant C0>0C_{0}>0 independent of k,s,p0k,s,p_{0} such that

|gs(0)|2|gs(𝐳)|2+M/C2/λ\displaystyle|g_{s}(0)|^{2}\leq|g_{s}(\mathbf{z})|^{2}+M/C^{2/\lambda^{\prime}} (4.15)

if |𝐳|2k2/α/C0|\mathbf{z}|\leq 2k^{-2/\alpha}/C_{0}. Since μ=ρLebK\mu=\rho\mathop{\mathrm{Leb}}\nolimits_{K}, where ρλL1(LebK)\rho^{-\lambda}\in L^{1}(\mathop{\mathrm{Leb}}\nolimits_{K}), applying Hölder inequality to |gs|2λ=(|gs|2λρλ)(ρλ)|g_{s}|^{2\lambda^{\prime}}=(|g_{s}|^{2\lambda^{\prime}}\rho^{\lambda^{\prime}})(\rho^{-\lambda^{\prime}}) gives

{|𝐳|k2/α}|gs|2λdLebK({|𝐳|k2/α}|gs|2𝑑μ)λ1.\displaystyle\int_{\{|\mathbf{z}|\leq k^{-2/\alpha}\}}|g_{s}|^{2\lambda^{\prime}}d\mathop{\mathrm{Leb}}\nolimits_{K}\lesssim\bigg{(}\int_{\{|\mathbf{z}|\leq k^{-2/\alpha}\}}|g_{s}|^{2}d\mu\bigg{)}^{\lambda^{\prime}}\lesssim 1. (4.16)

Let AC6A\geq C^{6} be a constant. Now let F:𝔻×YXF:\mathbb{D}\times Y\to X be the family of analytic discs in Theorem 4.1 associated to C0C_{0} for p0=0,p_{0}=0, and t:=k2/αt:=k^{-2/\alpha}.

Let g:=|gsf|2λg:=|g_{s}\circ f|^{2\lambda^{\prime}}. Put 𝐳y:=F(11/A,y)\mathbf{z}_{y}:=F(1-1/A,y) which is 2k2/α/C0\leq 2k^{-2/\alpha}/C_{0} by properties of FF if AA is big enough. Applying Lemma 4.7 to g(,y)g(\cdot,y) and β=1/2\beta=1/2 yield

|gs(𝐳y)|2λ=g(ξy,y)CMλ/A1/2+ACθ0θ0g(eiθ,y)𝑑θ|g_{s}(\mathbf{z}_{y})|^{2\lambda^{\prime}}=g(\xi_{y},y)\leq CM^{\lambda^{\prime}}/A^{1/2}+AC\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta

(again we increase CC if necessary independently of s,p0,ks,p_{0},k). By this and (4.15), we gets

|gs(0)|2λ\displaystyle|g_{s}(0)|^{2\lambda^{\prime}} Mλ/C2+CMλ/A1/2+ACθ0θ0g(eiθ,y)𝑑θ\displaystyle\leq M^{\lambda^{\prime}}/C^{2}+CM^{\lambda^{\prime}}/A^{1/2}+AC\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta
2Mλ/C2+ACθ0θ0g(eiθ,y)𝑑θ.\displaystyle\leq 2M^{\lambda^{\prime}}/C^{2}+AC\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta.

Integrating the last inequality over yYy\in Y gives

|gs(0)|λ2Mλ/C2+AC(vol(Y))1Yvolyθ0θ0g(eiθ,y)𝑑θ.\displaystyle|g_{s}(0)|^{\lambda^{\prime}}\leq 2M^{\lambda^{\prime}}/C^{2}+AC(\mathop{\mathrm{vol}}(Y))^{-1}\int_{Y}vol_{y}\int_{-\theta_{0}}^{\theta_{0}}g(e^{i\theta},y)d\theta. (4.17)

The second term in the right-hand side of the last inequality is bounded by

Ik:=𝔹(0,k2/α)K|gs(x)|2λ|detDG|1dLebKk2n/α𝔹(0,k2/α)K|gs(x)|2λdLebK.I_{k}:=\int_{\mathbb{B}(0,k^{-2/\alpha})\cap K}|g_{s}(x)|^{2\lambda^{\prime}}|\det DG|^{-1}d\mathop{\mathrm{Leb}}\nolimits_{K}\leq k^{2n/\alpha}\int_{\mathbb{B}(0,k^{-2/\alpha})\cap K}|g_{s}(x)|^{2\lambda^{\prime}}d\mathop{\mathrm{Leb}}\nolimits_{K}.

Using this, (4.17) and (4.16) gives

(gs(0))2λ2Mλ/C2+AC2k2n/λ\big{(}g_{s}(0)\big{)}^{2\lambda^{\prime}}\leq 2M^{\lambda^{\prime}}/C^{2}+AC_{2}k^{2n/\lambda^{\prime}}

for some constant C2C_{2} big enough (independent of k,s,p0k,s,p_{0}). This combined with (4.14) gives

|s(p0)|hk2λ2Mλ/C+ACC2k2n/λ|s(p_{0})|_{h^{k}}^{2\lambda^{\prime}}\leq 2M^{\lambda^{\prime}}/C+ACC_{2}k^{2n/\lambda^{\prime}}

for every p0Kkp_{0}\in K_{k}. By this and (4.13) we obtain

Mλk2n/αM^{\lambda^{\prime}}\lesssim k^{2n/\alpha}

by choosing CC big enough. Hence Mk2n/(αλ).M\lesssim k^{2n/(\alpha\lambda^{\prime})}. This finishes the proof. ∎

We now explain how to prove Theorem 2.3 when KK is not necessarily totally real.

End of proof of Theorem 2.3 for dimKn\dim K\geq n.

As in the case of dimK=n\dim K=n, it suffices to work with points in KK of distance at least k2/α/C0k^{-2/\alpha}/C_{0} to the singularity of KK for some C0>0C_{0}>0 big enough. Let p0p_{0} be such a point, and let sH0(X,Lk)s\in H^{0}(X,L^{k}) with sL2(μ,hk)=1\|s\|_{L^{2}(\mu,h^{k})}=1. Our goal is to show that |s(p0)|hkCk2nK(λ+1)/(αλ)|s(p_{0})|_{h^{k}}\leq Ck^{2n_{K}(\lambda+1)/(\alpha\lambda)} for some constant CC independent of ss and kk. We identify ss with a holomorphic function gsg_{s} on a small open neighborhood UU of p0p_{0} as usual. Hence as above one get

𝔹(p0,k2/α)K|gs|2λdLebK1,\displaystyle\int_{\mathbb{B}(p_{0},k^{-2/\alpha})\cap K}|g_{s}|^{2\lambda^{\prime}}d\mathop{\mathrm{Leb}}\nolimits_{K}\lesssim 1, (4.18)

where λ\lambda^{\prime} is as in the case of dimK=n\dim K=n.

Let rr be the CR dimension of KK. Recall that dimK=n+r\dim K=n+r. Using standard local coordinates near p0p_{0} on KK, one sees that by shrinking UU if necessary, there are holomorphic local coordinates (𝐳1,𝐳2)Unr×r(\mathbf{z}_{1},\mathbf{z}_{2})\in U\subset\mathbb{C}^{n-r}\times\mathbb{C}^{r} such that KK is locally given by the graph Im𝐳1=h(Re𝐳1,𝐳2)\mathop{\mathrm{Im}}\nolimits\mathbf{z}_{1}=h(\mathop{\mathrm{Re}}\nolimits\mathbf{z}_{1},\mathbf{z}_{2}), for h𝒞5h\in\mathcal{C}^{5}. In particular for every vector vrv\in\mathbb{R}^{r} (small enough) the real linear subspace Imz2=v\mathop{\mathrm{Im}}\nolimits z_{2}=v intersects KK at a generic CR 𝒞5\mathcal{C}^{5} smooth submanifold KvK_{v} in UU. Put gs,v:=gs|Kvg_{s,v}:=g_{s}|_{K_{v}}. Let C0>0C_{0}>0 be a big constant to be chosen later. By Fubini’s theorem and (4.18), we obtain

|v|k2/αdLebrKv|gs,v|2λdLebKv1.\int_{|v|\leq k^{-2/\alpha}}d\mathop{\mathrm{Leb}}\nolimits_{\mathbb{R}^{r}}\int_{K_{v}}|g_{s,v}|^{2\lambda^{\prime}}d\mathop{\mathrm{Leb}}\nolimits_{K_{v}}\lesssim 1.

It follows that there exists vv with |v|k2/α/C0|v|\leq k^{-2/\alpha}/C_{0} so that

Kv|gs,v|2λdLebKvk2r/α.\int_{K_{v}}|g_{s,v}|^{2\lambda^{\prime}}d\mathop{\mathrm{Leb}}\nolimits_{K_{v}}\lesssim k^{-2r/\alpha}.

Applying the proof of the case where dimK=n\dim K=n to KvK_{v}, we see that

|s(0,h(0,0,v),0,v)|hk2k2nK/(αλ),\displaystyle|s(0,h(0,0,v),0,v)|^{2}_{h^{k}}\lesssim k^{2n_{K}/(\alpha\lambda^{\prime})}, (4.19)

where we write (𝐳1,𝐳2)=(Re𝐳1,Im𝐳1,Re𝐳2,Im𝐳2)(\mathbf{z}_{1},\mathbf{z}_{2})=(\mathop{\mathrm{Re}}\nolimits\mathbf{z}_{1},\mathop{\mathrm{Im}}\nolimits\mathbf{z}_{1},\mathop{\mathrm{Re}}\nolimits\mathbf{z}_{2},\mathop{\mathrm{Im}}\nolimits\mathbf{z}_{2}), and p0p_{0} is identified with 0 in these local coordinates. On the other hand since |v|k2/α/C0|v|\leq k^{-2/\alpha}/C_{0}, using a version of Bernstein-Markov inequality (similar to (4.11)) yields

|s(0)|hk2|s(0,h(0,0,v),0,v)|hk2+12supK|s|hk2|s(0)|^{2}_{h^{k}}\leq|s(0,h(0,0,v),0,v)|^{2}_{h^{k}}+\frac{1}{2}\sup_{K}|s|^{2}_{h^{k}}

if C0C_{0} is big enough. This combined with (4.19) gives the desired upper bound. The proof is finished. ∎

We end the subsection with a remark.

Remark 4.9.

Let the hypothesis be as in Theorem 2.3. In the proof of Theorem 2.3, we actually proved the following local estimate. Let UU be an open subset in XX, and μU\mu_{U} be the restriction of μ\mu to UU, define

Bk,U:=sup{sH0(U,Lk)}|s|hk2sL2(μU,hk)2B_{k,U}:=\sup_{\{s\in H^{0}(U,L^{k})\}}\frac{|s|^{2}_{h^{k}}}{\|s\|^{2}_{L^{2}(\mu_{U},h^{k})}}\cdot

Then for every UUU^{\prime}\Subset U, there exists C>0C^{\prime}>0 such that

supKUBk,UCk2nK(λ+1)/(αλ).\sup_{K\cap U^{\prime}}B_{k,U}\leq C^{\prime}k^{2n_{K}(\lambda+1)/(\alpha\lambda)}.

5 Zeros of random polynomials

In this section we prove Theorem 1.7. Let Lebm\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}^{m}} be the Lebesgue measure on m\mathbb{C}^{m} for m1m\geq 1, and we denote by \|\cdot\| the standard Euclidean norm on m\mathbb{C}^{m}. Let ωFS,m\omega_{FS,m} be the Fubini-Study form on m\mathbb{P}^{m}, and let ΩFS,m:=ωFS,mm\Omega_{FS,m}:=\omega_{FS,m}^{m} be the Fubini-Study volume form on m\mathbb{P}^{m}. We always embed m\mathbb{C}^{m} in m\mathbb{P}^{m}. We recall the following key lemma.

Lemma 5.1.

([32, Proposition A.3 and Corollary A.5]) There exist constants C,λ>0C,\lambda>0 such that for every k0k\geq 0, and every ωFS,k\omega_{FS,k}-psh function uu on k\mathbb{P}^{k} with XuΩFS,k=0\int_{X}u\Omega_{FS,k}=0, then

uC(1+logk),{u<t}ΩFS,kCkeλtu\leq C(1+\log k),\quad\int_{\{u<-t\}}\Omega_{FS,k}\leq Cke^{-\lambda t}

for every t0t\geq 0.

The essential point is that the constants in the above result is uniformly in the dimension kk of k\mathbb{P}^{k}. In our applications, the dimension kk will tend to \infty. Let dkd_{k} be the dimension of the space of polynomials of degree at most kk on n\mathbb{C}^{n}. Note that dkknd_{k}\approx k^{n}.

Let ff be a bounded Borel function on \mathbb{C} such that there is a constant C0>0C_{0}>0 for which for every r>1r>1 we have

|z|rfdLebC0/r2.\int_{|z|\geq r}fd\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}}\leq C_{0}/r^{2}.

Let a1,a2,,adka_{1},a_{2},\ldots,a_{d_{k}} be complex-valued i.i.d random variable whose distribution is fLebf\mathop{\mathrm{Leb}}\nolimits_{\mathbb{C}}. Assume furthermore that the joint-distribution of (a1,,adk)(a_{1},\ldots,a_{d_{k}}) satisfies

𝒫k:=f(z1)f(zdk)Leb(z1)Leb(zdk)C0ΩFS,dk\mathscr{P}_{k}:=f(z_{1})\ldots f(z_{d_{k}})\mathop{\mathrm{Leb}}\nolimits(z_{1})\otimes\cdots\otimes\mathop{\mathrm{Leb}}\nolimits(z_{d_{k}})\leq C_{0}\Omega_{FS,d_{k}}

on dk\mathbb{C}^{d_{k}}.

Let p(dk):=(p1,,pdk)p^{(d_{k})}:=(p_{1},\ldots,p_{d_{k}}) be an orthonormal basis of 𝒫k(n)\mathcal{P}_{k}(\mathbb{C}^{n}). Let LL be a complex algebraic subvariety of dimension m1m\geq 1 in n\mathbb{C}^{n}. Note that the tolological closure of LL in n\mathbb{P}^{n} is an algebraic subvariety in n\mathbb{P}^{n}. Observe that

ωFS,nmCωm,\omega_{FS,n}^{m}\leq C\omega^{m},

where ω\omega is the standard Kähler form on n\mathbb{C}^{n}, and C>0C>0 is a constant. Fix a compact AA of volume vol(A):=AωFS,nm>0\mathop{\mathrm{vol}}(A):=\int_{A}\omega_{FS,n}^{m}>0 in n\mathbb{C}^{n}. We start with a version of [15, Lemma 2.4] with more or less the same proof. We use the Euclidean norm on dk\mathbb{C}^{d_{k}}.

Lemma 5.2.

Let M1M\geq 1 be a constant. Let EkE_{k} be the set of (a1,,adk)dk(a_{1},\ldots,a_{d_{k}})\in\mathbb{C}^{d_{k}} such that

A(log|j=1dkajpj(z)|1/2logj=1dk|pj(z)|2)ωFS,nm2Mvol(A)logdk.\int_{A}\bigg{(}\log\bigg{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\bigg{|}-1/2\log\sum_{j=1}^{d_{k}}|p_{j}(z)|^{2}\bigg{)}\omega_{FS,n}^{m}\geq 2M\mathop{\mathrm{vol}}(A)\log d_{k}.

Let EkE^{\prime}_{k} be the set of a(dk)a^{(d_{k})} so that a(dk)dk2M\|a^{(d_{k})}\|\geq d_{k}^{2M}. Then we have

𝒫k(EkEk)Cdk3M\mathscr{P}_{k}(E_{k}\cup E_{k}^{\prime})\leq Cd_{k}^{-3M}

for some constant C>0C>0 independent of kk and MM.

Proof.

Put a(dk):=(a1,,adk)a^{(d_{k})}:=(a_{1},\ldots,a_{d_{k}}), and

Ik(a1,,adk):=A(log|j=1dkajpj(z)|1/2logj=1dk|pj(z)|2)ωFS,nmI_{k}(a_{1},\ldots,a_{d_{k}}):=\int_{A}\bigg{(}\log\bigg{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\bigg{|}-1/2\log\sum_{j=1}^{d_{k}}|p_{j}(z)|^{2}\bigg{)}\omega_{FS,n}^{m}

Observe

|j=1dkajpj(z)|a(dk)(j=1dk|pj(z)|2)1/2.\bigg{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\bigg{|}\leq\|a^{(d_{k})}\|(\sum_{j=1}^{d_{k}}|p_{j}(z)|^{2})^{1/2}.

It follows that for a(dk)Eka^{(d_{k})}\in E_{k}, one has

2Mvol(A)logdkIk(a(dk))loga(dk)vol(A).2M\mathop{\mathrm{vol}}(A)\log d_{k}\leq I_{k}(a^{(d_{k})})\leq\log\|a^{(d_{k})}\|\mathop{\mathrm{vol}}(A).

This implies that for each kk there exists 1jkdk1\leq j_{k}\leq d_{k} such that |ajk|dk(4M1)/2|a_{j_{k}}|\geq d_{k}^{(4M-1)/2}. We infer that

𝒫k(Ek)\displaystyle\mathscr{P}_{k}(E_{k}) 𝒫k(a(dk):|aj|dk(4M1)/2for some 1jdk)\displaystyle\leq\mathscr{P}_{k}\bigg{(}a^{(d_{k})}:|a_{j}|\geq d_{k}^{(4M-1)/2}\,\text{for some }1\leq j\leq d_{k}\bigg{)}
dk|a1|dk(4M1)/2fLebCC0dk4M+1C0dk3M.\displaystyle\leq d_{k}\int_{|a_{1}|\geq d_{k}^{(4M-1)/2}}f\mathop{\mathrm{Leb}}\nolimits_{C}\leq C_{0}d_{k}^{-4M+1}\leq C_{0}d_{k}^{-3M}.

Similarly, we also get 𝒫k(Ek)dk3M\mathscr{P}_{k}(E^{\prime}_{k})\lesssim d_{k}^{-3M}. This finishes the proof. ∎

Put a(dk):=(a1,,adk)a^{(d_{k})}:=(a_{1},\ldots,a_{d_{k}}), and p(dk):=(p1,,pdk)p^{(d_{k})}:=(p_{1},\ldots,p_{d_{k}}). Define

φ(a(dk)):=zLlog|j=1dkajpj(z)|(a(dk)2+1)1/2p(dk)(z)ωFS,nm(z).\varphi(a^{(d_{k})}):=\int_{z\in L}\log\frac{\big{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\big{|}}{(\|a^{(d_{k})}\|^{2}+1)^{1/2}\|p^{(d_{k})}(z)\|}\omega_{FS,n}^{m}(z).

Observe that φ0\varphi\leq 0 on dk\mathbb{C}^{d_{k}}. We put

Iφ:=dkφ(a(dk))ΩFS,dkI_{\varphi}:=\int_{\mathbb{C}^{d_{k}}}\varphi(a^{(d_{k})})\Omega_{FS,d_{k}}
Lemma 5.3.

There exists a constant C>0C>0 such that for every k1k\geq 1 we have

IφClogdk.I_{\varphi}\geq-C\log d_{k}.
Proof.

Let II be the right-hand side of the desired inequality. By Funibi’s theorem and the transitivity of the unitary group on dk\mathbb{C}^{d_{k}}, one has

Iφ\displaystyle I_{\varphi} =zLωFS,nmdkφ(a(dk))ΩFS,dk\displaystyle=\int_{z\in L}\omega_{FS,n}^{m}\int_{\mathbb{C}^{d_{k}}}\varphi(a^{(d_{k})})\Omega_{FS,d_{k}}
=zLωFS,nmdklog|a1|(a2+1)1/2ΩFS,dk.\displaystyle=\int_{z\in L}\omega_{FS,n}^{m}\int_{\mathbb{C}^{d_{k}}}\log\frac{|a_{1}|}{(\|a\|^{2}+1)^{1/2}}\Omega_{FS,d_{k}}.

The function

ψ:=log|a1|(a(dk)2+1)1/2\psi:=\log\frac{|a_{1}|}{(\|a^{(d_{k})}\|^{2}+1)^{1/2}}

is ωFS,dk\omega_{FS,d_{k}}-psh on dk\mathbb{P}^{d_{k}} (where ωFS,dk\omega_{FS,d_{k}} is the Fubini-Study form on dk\mathbb{P}^{d_{k}}). Let ψ:=ψdkψΩFS,dk\psi^{\prime}:=\psi-\int_{\mathbb{C}^{d_{k}}}\psi\Omega_{FS,d_{k}}. Thus dkψΩFS,dk=0\int_{\mathbb{P}^{d_{k}}}\psi^{\prime}\Omega_{FS,d_{k}}=0 and ψ\psi^{\prime} is ωFS,dk\omega_{FS,d_{k}}-psh. Applying Lemma 5.1 to ψ\psi^{\prime} gives

ψc(1+logdk)\psi^{\prime}\leq c(1+\log d_{k})

for some constant c>0c>0 independent of k,ψk,\psi^{\prime}. Consequently,

ψ(a(dk))c(1+logdk)+dkψΩFS,dk\psi(a^{(d_{k})})\leq c(1+\log d_{k})+\int_{\mathbb{C}^{d_{k}}}\psi\Omega_{FS,d_{k}}

for every a(dk)ma^{(d_{k})}\in\mathbb{C}^{m}. In particular for a(dk)=(1,0,,0)a^{(d_{k})}=(1,0,\ldots,0), we obtain

dkψΩFS,dkc(1+logdk)1.\int_{\mathbb{C}^{d_{k}}}\psi\Omega_{FS,d_{k}}\geq-c(1+\log d_{k})-1.

Thus the desired inequality follows. ∎

End of the proof of Theorems 1.7 and 1.9.

Let φ:=φIφ\varphi^{\prime}:=\varphi-I_{\varphi}. We have dkφΩFS,dk=0\int_{\mathbb{C}^{d_{k}}}\varphi^{\prime}\Omega_{FS,d_{k}}=0. By Lemma 5.1 again, there are constants c,α>0c,\alpha>0 independent of kk such that

{φt}ΩFS,dkcdkeαt.\int_{\{\varphi^{\prime}\leq-t\}}\Omega_{FS,d_{k}}\leq cd_{k}e^{-\alpha t}.

Combining this with Lemma 5.3 yields

{φClogdkt}ΩFS,dkcdkeαt.\int_{\{\varphi\leq-C\log d_{k}-t\}}\Omega_{FS,d_{k}}\leq cd_{k}e^{-\alpha t}.

Let M1M\geq 1 be a constant. Choosing t:=4Mα1logdkt:=4M\alpha^{-1}\log d_{k}, where C>0C>0 big enough gives

{φ(C+3/α)logdk}ΩFS,dkcdk3M.\displaystyle\int_{\{\varphi\leq-(C+3/\alpha)\log d_{k}\}}\Omega_{FS,d_{k}}\leq cd_{k}^{-3M}. (5.1)

Let EkE^{\prime}_{k} be the set of a(dk)a^{(d_{k})} such that a(dk)dk2M\|a^{(d_{k})}\|\leq d_{k}^{2M} and φ(C+4M/α)logdk\varphi\geq-(C+4M/\alpha)\log d_{k}. Combining (5.1) and Lemma 5.2, we obtain that

𝒫k(dk\Ek)2cdk3M\mathscr{P}_{k}(\mathbb{C}^{d_{k}}\backslash E^{\prime}_{k})\lesssim 2cd_{k}^{-3M}

(we increase cc if necessary). On the other hand, by the definition of EkE^{\prime}_{k} and φ\varphi, we see that the set of a(dk)a^{(d_{k})} such that

zLlog|j=1dkajpj(z)|p(dk)(z)ωFS,nm(z)CMlogdk\int_{z\in L}\log\frac{\big{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\big{|}}{\|p^{(d_{k})}(z)\|}\omega_{FS,n}^{m}(z)\geq-C^{\prime}_{M}\log d_{k}

(for some constant CM>0C^{\prime}_{M}>0 big enough independent of kk) contains EkE^{\prime}_{k}. It follows that

𝒫k{a(dk):zLlog|j=1dkajpj(z)|p(dk)(z)ΩFS,n(z)CMlogdk}2cdk3M.\mathscr{P}_{k}\bigg{\{}a^{(d_{k})}:\int_{z\in L}\log\frac{\big{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(z)\big{|}}{\|p^{(d_{k})}(z)\|}\Omega_{FS,n}(z)\leq-C^{\prime}_{M}\log d_{k}\bigg{\}}\leq 2cd_{k}^{-3M}.

This together with Lemma 5.2 implies that there exists a Borel set FkF_{k} such that 𝒫k(Fk)3cdk3M\mathscr{P}_{k}(F_{k})\leq 3cd_{k}^{-3M}, and for a(dk)Fka^{(d_{k})}\not\in F_{k}, one has a(dk)dk2M\|a^{(d_{k})}\|\leq d_{k}^{2M}, and

zLlog|j=1dkajpj(x)|p(dk)(z)ωFS,nm(z)CMlogdk.\int_{z\in L}\log\frac{\big{|}\sum_{j=1}^{d_{k}}a_{j}p_{j}(x)\big{|}}{\|p^{(d_{k})}(z)\|}\omega_{FS,n}^{m}(z)\geq-C^{\prime}_{M}\log d_{k}.

Consequently for p=j=1dkajpjp=\sum_{j=1}^{d_{k}}a_{j}p_{j}, ψk:=1/(2k)logB~k\psi_{k}:=1/(2k)\log\tilde{B}_{k} and a(dk)Fka^{(d_{k})}\not\in F_{k}, there holds

CMlogk/kL(k1log|p(dk)|ψk)ωFS,nm.-C^{\prime}_{M}\log k/k\leq\int_{L}(k^{-1}\log|p^{(d_{k})}|-\psi_{k})\omega_{FS,n}^{m}.

Moreover a(dk)dk2M\|a^{(d_{k})}\|\leq d_{k}^{2M}, one has

|k1log|p|ψk|\displaystyle|k^{-1}\log|p|-\psi_{k}| =2max{k1log|p|,ψk}ψkk1log|p|\displaystyle=2\max\{k^{-1}\log|p|,\psi_{k}\}-\psi_{k}-k^{-1}\log|p|
ψkk1log|p|+2Mk1logdk.\displaystyle\leq\psi_{k}-k^{-1}\log|p|+2Mk^{-1}\log d_{k}.

It follows that

L|k1log|p|ψk|ωFS,nmCM′′logdk/k,\int_{L}|k^{-1}\log|p|-\psi_{k}|\omega_{FS,n}^{m}\leq C^{\prime\prime}_{M}\log d_{k}/k,

for some constant CM′′>0C^{\prime\prime}_{M}>0 independent of kk. This together with Theorem 1.6 implies

L|k1log|p|VK,Q|ωFS,nmCM′′logkk\int_{L}|k^{-1}\log|p|-V_{K,Q}|\omega_{FS,n}^{m}\leq C^{\prime\prime}_{M}\frac{\log k}{k}

by increasing CM′′C^{\prime\prime}_{M} if necessary. It follows that

dist2(k1[p=0][L],ddcVK,Q[L])L|k1log|p|ψk|ωFS,nmCM′′logkk,\mathop{\mathrm{dist}}\nolimits_{-2}\big{(}k^{-1}[p=0]\wedge[L],dd^{c}V_{K,Q}\wedge[L]\big{)}\lesssim\int_{L}|k^{-1}\log|p|-\psi_{k}|\omega_{FS,n}^{m}\leq\frac{C^{\prime\prime}_{M}\log k}{k},

for a(dk)Fka^{(d_{k})}\not\in F_{k}. Here recall that we define dist2\mathop{\mathrm{dist}}\nolimits_{-2} by considering [p=0][L][p=0]\wedge[L] and ddcVK,Q[L]dd^{c}V_{K,Q}\wedge[L] as currents on m\mathbb{P}^{m}. This finishes the proof. ∎

References

  • [1] J. Antezana, J. Marzo, and J. Ortega-Cerdà, Necessary conditions for interpolation by multivariate polynomials, Comput. Methods Funct. Theory, 21 (2021), pp. 831–849.
  • [2] M. S. Baouendi, P. Ebenfelt, and L. P. Rothschild, Real submanifolds in complex space and their mappings, vol. 47 of Princeton Mathematical Series, Princeton University Press, Princeton, NJ, 1999.
  • [3]  , Real submanifolds in complex space and their mappings, vol. 47 of Princeton Mathematical Series, Princeton University Press, Princeton, NJ, 1999.
  • [4] T. Bayraktar, Equidistribution of zeros of random holomorphic sections, Indiana Univ. Math. J., 65 (2016), pp. 1759–1793.
  • [5]  , Mass equidistribution for random polynomials, Potential Anal., 53 (2020), pp. 1403–1421.
  • [6] T. Bayraktar, D. Coman, and G. Marinescu, Universality results for zeros of random holomorphic sections, Trans. Amer. Math. Soc., 373 (2020), pp. 3765–3791.
  • [7] B. Beckermann, M. Putinar, E. B. Saff, and N. Stylianopoulos, Perturbations of Christoffel-Darboux kernels: detection of outliers, Found. Comput. Math., 21 (2021), pp. 71–124.
  • [8] R. Berman and S. Boucksom, Growth of balls of holomorphic sections and energy at equilibrium, Invent. Math., 181 (2010), pp. 337–394.
  • [9] R. Berman, S. Boucksom, and D. Witt Nyström, Fekete points and convergence towards equilibrium measures on complex manifolds, Acta Math., 207 (2011), pp. 1–27.
  • [10] R. J. Berman, Bergman kernels and equilibrium measures for line bundles over projective manifolds, Amer. J. Math., 131 (2009), pp. 1485–1524.
  • [11]  , Bergman kernels for weighted polynomials and weighted equilibrium measures of n\mathbb{C}^{n}, Indiana Univ. Math. J., 58 (2009), pp. 1921–1946.
  • [12] R. J. Berman and J. Ortega-Cerdà, Sampling of real multivariate polynomials and pluripotential theory, Amer. J. Math., 140 (2018), pp. 789–820.
  • [13] T. Bloom, Random polynomials and (pluri)potential theory, Ann. Polon. Math., 91 (2007), pp. 131–141.
  • [14] T. Bloom and D. Dauvergne, Asymptotic zero distribution of random orthogonal polynomials, Ann. Probab., 47 (2019), pp. 3202–3230.
  • [15] T. Bloom and N. Levenberg, Random polynomials and pluripotential-theoretic extremal functions, Potential Anal., 42 (2015), pp. 311–334.
  • [16] T. Bloom, N. Levenberg, F. Piazzon, and F. Wielonsky, Bernstein-Markov: a survey, Dolomites Res. Notes Approx., 8 (2015), pp. 75–91.
  • [17] T. Bloom and B. Shiffman, Zeros of random polynomials on m\mathbb{C}^{m}, Math. Res. Lett., 14 (2007), pp. 469–479.
  • [18] L. Bos, N. Levenberg, P. Milman, and B. A. Taylor, Tangential Markov inequalities characterize algebraic submanifolds of 𝐑N{\bf R}^{N}, Indiana Univ. Math. J., 44 (1995), pp. 115–138.
  • [19] L. P. Bos, N. Levenberg, P. D. Milman, and B. A. Taylor, Tangential Markov inequalities on real algebraic varieties, Indiana Univ. Math. J., 47 (1998), pp. 1257–1272.
  • [20] A. Brudnyi, On local behavior of holomorphic functions along complex submanifolds of N\mathbb{C}^{N}, Invent. Math., 173 (2008), pp. 315–363.
  • [21]  , Bernstein type inequalities for restrictions of polynomials to complex submanifolds of N\mathbb{C}^{N}, J. Approx. Theory, 225 (2018), pp. 106–147.
  • [22] J. Chu and B. Zhou, Optimal regularity of plurisubharmonic envelopes on compact Hermitian manifolds, Sci. China Math., 62 (2019), pp. 371–380.
  • [23] D. Coman and E. A. Poletsky, Transcendence measures and algebraic growth of entire functions, Invent. Math., 170 (2007), pp. 103–145.
  • [24] T. Danka and V. Totik, Christoffel functions with power type weights, J. Eur. Math. Soc. (JEMS), 20 (2018), pp. 747–796.
  • [25] D. Dauvergne, A necessary and sufficient condition for convergence of the zeros of random polynomials, Adv. Math., 384 (2021).
  • [26] P. A. Deift, Orthogonal polynomials and random matrices: a Riemann-Hilbert approach, vol. 3 of Courant Lecture Notes in Mathematics, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 1999.
  • [27] J.-P. Demailly, Analytic methods in algebraic geometry, vol. 1 of Surveys of Modern Mathematics, International Press, Somerville, MA; Higher Education Press, Beijing, 2012.
  • [28] T.-C. Dinh, Large deviation theorem for zeros of polynomials and Hermitian random matrices, J. Geom. Anal., 30 (2020), pp. 2558–2580.
  • [29] T.-C. Dinh, X. Ma, and G. Marinescu, Equidistribution and convergence speed for zeros of holomorphic sections of singular Hermitian line bundles, J. Funct. Anal., 271 (2016), pp. 3082–3110.
  • [30] T.-C. Dinh, X. Ma, and V.-A. Nguyên, Equidistribution speed for Fekete points associated with an ample line bundle, Ann. Sci. Éc. Norm. Supér. (4), 50 (2017), pp. 545–578.
  • [31] T.-C. Dinh and V.-A. Nguyên, Large deviation principle for some beta ensembles, Trans. Amer. Math. Soc., 370 (2018), pp. 6565–6584.
  • [32] T.-C. Dinh and N. Sibony, Distribution des valeurs de transformations méromorphes et applications, Comment. Math. Helv., 81 (2006), pp. 221–258.
  • [33]  , Super-potentials of positive closed currents, intersection theory and dynamics, Acta Math., 203 (2009), pp. 1–82.
  • [34] T.-C. Dinh and D.-V. Vu, Estimation of Deviation for Random Covariance Matrices, Michigan Math. J., 68 (2019), pp. 597–620.
  • [35] C. F. Dunkl and Y. Xu, Orthogonal polynomials of several variables, vol. 81 of Encyclopedia of Mathematics and its Applications, Cambridge University Press, Cambridge, 2001.
  • [36] F. D. Gakhov, Boundary value problems, Pergamon Press, Oxford-New York-Paris; Addison-Wesley Publishing Co., Inc., Reading, Mass.-London, 1966. Translation edited by I. N. Sneddon.
  • [37] F. Götze and J. Jalowy, Rate of convergence to the Circular Law via smoothing inequalities for log-potentials, Random Matrices Theory Appl., 10 (2021).
  • [38] P. Griffiths and J. Harris, Principles of algebraic geometry, Pure and Applied Mathematics, Wiley-Interscience [John Wiley & Sons], New York, 1978.
  • [39] B. Gustafsson, M. Putinar, E. B. Saff, and N. Stylianopoulos, Bergman polynomials on an archipelago: estimates, zeros and shape reconstruction, Adv. Math., 222 (2009), pp. 1405–1460.
  • [40] J. M. Hammersley, The zeros of a random polynomial, in Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, 1954–1955, vol. II, University of California Press, Berkeley-Los Angeles, Calif., 1956, pp. 89–111.
  • [41] D. T. Huynh and D.-V. Vu, On the set of divisors with zero geometric defect, J. Reine Angew. Math., 771 (2021), pp. 193–213.
  • [42] I. Ibragimov and D. Zaporozhets, On distribution of zeros of random polynomials in complex plane, in Prokhorov and contemporary probability theory, vol. 33 of Springer Proc. Math. Stat., Springer, Heidelberg, 2013, pp. 303–323.
  • [43] A. Kroó, Christoffel functions on convex and starlike domains in d\mathbb{R}^{d}, J. Math. Anal. Appl., 421 (2015), pp. 718–729.
  • [44] A. Kroó and D. S. Lubinsky, Christoffel functions and universality in the bulk for multivariate orthogonal polynomials, Canad. J. Math., 65 (2013), pp. 600–620.
  • [45]  , Christoffel functions and universality on the boundary of the ball, Acta Math. Hungar., 140 (2013), pp. 117–133.
  • [46] H.-C. Lu, T.-T. Phung, and T.-D. Tô, Stability and Hölder regularity of solutions to complex Monge-Ampère equations on compact Hermitian manifolds. arXiv:2003.08417, 2020. to appear in Annales de l’Institut Fourier.
  • [47] D. S. Lubinsky, A new approach to universality limits involving orthogonal polynomials, Ann. of Math. (2), 170 (2009), pp. 915–939.
  • [48] A. Lunardi, Interpolation theory, Appunti. Scuola Normale Superiore di Pisa (Nuova Serie). [Lecture Notes. Scuola Normale Superiore di Pisa (New Series)], Edizioni della Normale, Pisa, second ed., 2009.
  • [49] X. Ma and G. Marinescu, Holomorphic Morse inequalities and Bergman kernels, vol. 254 of Progress in Mathematics, Birkhäuser Verlag, Basel, 2007.
  • [50] J. Marzo and J. Ortega-Cerdà, Equidistribution of Fekete points on the sphere, Constr. Approx., 32 (2010), pp. 513–521.
  • [51] J. Merker and E. Porten, Holomorphic extension of CR functions, envelopes of holomorphy, and removable singularities, IMRS Int. Math. Res. Surv., (2006), pp. 1–287.
  • [52] R. Pierzchał a, Geometry of holomorphic mappings and Hölder continuity of the pluricomplex Green function, Math. Ann., 379 (2021), pp. 1363–1393.
  • [53] I. Pritsker and K. Ramachandran, Equidistribution of zeros of random polynomials, J. Approx. Theory, 215 (2017), pp. 106–117.
  • [54]  , Natural boundary and zero distribution of random polynomials in smooth domains, Comput. Methods Funct. Theory, 19 (2019), pp. 401–410.
  • [55] A. Prymak, Upper estimates of Christoffel function on convex domains, J. Math. Anal. Appl., 455 (2017), pp. 1984–2000.
  • [56] A. Sadullaev, PP-regularity of sets in 𝐂n{\bf C}^{n}, in Analytic functions, Kozubnik 1979 (Proc. Seventh Conf., Kozubnik, 1979), vol. 798 of Lecture Notes in Math., pp. 402–408.
  • [57] A. Sadullaev and A. Zeriahi, Hölder regularity of generic manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 16 (2016), pp. 369–382.
  • [58] B. Shiffman and S. Zelditch, Equilibrium distribution of zeros of random polynomials, Int. Math. Res. Not., (2003), pp. 25–49.
  • [59] V. Tosatti, Regularity of envelopes in Kähler classes, Math. Res. Lett., 25 (2018), pp. 281–289.
  • [60] V. Totik, Christoffel functions on curves and domains, Trans. Amer. Math. Soc., 362 (2010), pp. 2053–2087.
  • [61]  , Christoffel functions on curves and domains, Trans. Amer. Math. Soc., 362 (2010), pp. 2053–2087.
  • [62] H. Triebel, Interpolation theory, function spaces, differential operators, Johann Ambrosius Barth, Heidelberg, second ed., 1995.
  • [63] D.-V. Vu, Complex Monge–Ampère equation for measures supported on real submanifolds, Math. Ann., 372 (2018), pp. 321–367.
  • [64]  , Equidistribution rate for Fekete points on some real manifolds, Amer. J. Math., 140 (2018), pp. 1311–1355.
  • [65] Y. Xu, Asymptotics of the Christoffel functions on a simplex in 𝐑d{\bf R}^{d}, J. Approx. Theory, 99 (1999), pp. 122–133.
  • [66] Y. Yomdin, Smooth parametrizations in dynamics, analysis, diophantine and computational geometry, Jpn. J. Ind. Appl. Math., 32 (2015), pp. 411–435.

George Marinescu, University of Cologne, Department of Mathematics and Computer Science, Division of Mathematics, Weyertal 86-90, 50931 Köln, Germany.

E-mail address: [email protected]

Duc-Viet Vu, University of Cologne, Department of Mathematics and Computer Science, Weyertal 86-90, 50931 Köln, Germany

E-mail address: [email protected]