Bergman kernel functions associated to measures supported on totally real submanifolds
Abstract
We prove that the Bergman kernel function associated to a smooth measure supported on a piecewise-smooth maximally totally real submanifold in is of polynomial growth (e.g, in dimension one, is a finite union of transverse Jordan arcs in ). Our bounds are sharp when 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 be a non-pluripolar compact subset in , i.e, is not contained in for some plurisubharmonic (psh) function on . Let be a probability measure whose support is non-pluripolar and is contained in , and be a real continuous function on . Let be the space of restrictions to of complex polynomials of degree at most on . The scalar product
induces the -norm on . The Bergman kernel function of order associated to with weight is defined by
for . Equivalently if (here denotes the dimension of ) is an orthonormal basis of with respect to the -norm, then
When , we say that is unweighted. In this case () the inverse of 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:
for . The advantage of is that it is well defined on .
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 (such as balls or simplexes in ) or in the unit ball in . Considering measures on whose support are not necessarily in 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 or domains in 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 in . 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 (rather than only smooth ones). In what follows, by a (convex) polyhedron in , we mean a subset in which is the intersection of a finite number of closed half-hyperplanes in .
Definition 1.1.
A subset of a real -dimensional smooth manifold is called a nondegenerate piecewise-smooth submanifold of dimension if for every point there exists a local chart of such that is a -diffeomorphism from to the unit ball of and is the intersection with the unit ball of a finite union of convex polyhedra of the same dimension
A point is said to be a regular point of if the above local chart can be chosen such that is the intersection of the unit ball with an -dimensional vector subspace in , in other words, is a -dimensional submanifold locally near . The regular part of is the set of regular points of The singular part of is the complement of the regular part of in Hence if is a smooth manifold with boundary, then the boundary of is the singular part of and its complement in is the regular part of .
Now let be a nondegenerate piecewise-smooth submanifold of Since is a complex manifold, its real tangent spaces have a natural complex structure . We say that is (Cauchy-Riemann or CR for short) generic in the sense of Cauchy-Riemann geometry if for every and every sequence of regular points approaching to , any limit space of the sequence of tangent spaces of at is not contained in a complex hyperplane of the (real) tangent space at of (equivalently, if is a limit space of the sequence of tangent spaces at , then we have , where is the real tangent space of at ).
For a CR generic , note that the space ( is a regular point in ) is invariant under and hence has a complex structure induced by . In this case, the complex dimension of is the same for every and is called the CR dimension of . If denotes the CR dimension of , then . Thus the dimension of a generic is at least
If is CR generic and , then is said to be (maximally) totally real, and it is locally the graph of a smooth function over a small ball centered at which is tangent at to . Examples of piecewise-smooth totally real submanifolds are polygons in or boundaries of polygons in , and polyhedra of dimension in .
A notion playing an important role in the study of Bergman kernel functions is the following extremal function
where is the set of psh functions on such that is bounded at infinity on .
Since is non-pluripolar, the upper semi-continuous regularisation of belongs to . If , then we say is regular. A stronger notion is the following: we say that is locally regular if for every there is an open neighborhood of in such that for every increasing sequence of psh functions on with on , then
on . One can see that if is a locally regular set, then is regular for every continuous function on . The following result answers the question raised in [9, Remark 1.8].
Theorem 1.2.
Let be a compact generic Cauchy-Riemann nondegenerate piecewise-smooth submanifold in . Then is locally regular.
Note that it was known that is locally regular if is smooth real analytic; see, e.g., [9, Corollary 1.7].
The measure is said to be a Bernstein-Markov measure (with respect to ) if for every , there exists such that
for every . In other words, the Bergman kernel function of order grows at most subexponentially, i.e., as for every .
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 be a compact generic Cauchy-Riemann nondegenerate piecewise-smooth submanifold in . Let be a finite measure supported on such that there exist constants satisfying for every , (where denotes the ball of radius centered at in ). Then for every continuous function on , is a Bernstein-Markov measure with respect to .
To the best of our knowledge, the above result was only known when is real analytic. We are not aware of results of this type in previous literature for maximally totally real submanifolds. A measure on is said to be a smooth volume form on if is given by a smooth volume form locally at every regular point in , and if for every singular point in , there is a local chart as in Definition 1.1, such that , where ’s are polyhedra in of the same dimension for every , and the restriction of to is a smooth volume form on for .
Let be now a smooth volume form on . Then for any the measure satisfies the hypothesis of Theorem 1.3. Here is our next main result.
Theorem 1.4.
Let be a compact generic Cauchy-Riemann nondegenerate piecewise-smooth submanifold in of dimension . Let be a Hölder continuous function of Hölder exponent on , and let be a smooth volume form on , and , where and for some constant . Then we have
for some constant independent of .
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 and (the Lebesgue measure on ).
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 were proved mostly based on special geometric structures of the compact (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 is smooth real algebraic in or the closure of a bounded convex open subset in , and their arguments use this hypothesis in an essential way.
Note that for any constant satisfies the hypothesis of Theorem 1.4. In general it is not possible to bound from below by a polynomial in , see Remark 3.7.
By [12, Theorems 2 and 4], if is the closure of a bounded open convex subset in and and is the restriction of the Lebesgue measure on to , then is bounded and bounded away from on a fixed compact subset in the interior of (the behaviour of at boundary points is more complicated). On the other hand, for general on such , by [24], the upper bound for on can not be in general. To be precise, it was proved there that if is a smooth Jordan curve in , and is the arc measure on , and for some constant , then as . One can see also [45] for a similar asymptotic in the case where is the closure of the unit ball in .
We note that by [31, Corollary 2.13], if is as in the hypothesis of Theorem 1.4), then the triple is -Bernstein-Markov in the sense that for every constant , there exists a constant such that
for every . This is much weaker than our bound. Nevertheless, [31, Corollary 2.13] is applicable to a broader class of .
When is smooth (no boundary) and for some constant (e.g, is the unit circle in 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 . The case where is compact smooth real algebraic was considered in [12]. Here is our next main result.
Theorem 1.5.
Let be a compact maximally totally real submanifold without boundary in . Let be a smooth volume form on regarded as a measure on . Let for some constant . Let be the Bergman kernel function associated to with weight . Then there exists a constant such that
(1.1) |
for every .
When is smooth compact real algebraic of dimension in , it was proved in [12] that . The proof of the upper bound for in [12] relies crucially on the algebraicity of . Our approach to Theorems 1.4 and 1.5 is different and is based on constructions of analytic discs partly attached to , subharmonic functions on unit discs, and fine regularity of extremal plurisubharmonic envelopes associated to .
In dimension one, we refer to [39, Theorem 4.3] for a similar bound when the case where is analytic, and to [24, 61] and references therein for asymptotics of (it behaves like at regular points, but the asymptotic of 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 is a convex subset in .
Our next result is the following convergence which is a consequence of Theorem 1.4.
Theorem 1.6.
Let be a compact generic Cauchy-Riemann nondegenerate piecewise-smooth submanifold in . Let be a Hölder continuous function on and be as in Theorem 1.4. Then there exists constant such that
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 be a non-pluripolar set in and let be a probability measure on such that the support of is contained in and is non-pluripolar. Let be a continuous weight on . Let be the space of restrictions of complex polynomials of degree at most in to . Let , and let be an orthonormal basis of with respect to the -scalar product. Consider the random polynomial
(1.2) |
where ’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 , and .
The distribution of zeros of more general random polynomials associated to orthonormal polynomials (as in (1.2)) was considered in [58] by observing that are an orthonormal basis of the restriction of the space of polynomials in to with respect to the -norm induced by the Haar measure on . In this setting, the necessary and sufficient conditions for the distribution of ’s so that the zeros of is equidistributed almost surely or in probability with respect to the Lebesgue measure on the unit circle as 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 and the extremal function associated to . This is what we obtained in Theorem 1.6. To state our result we need some hypothesis on and the distribution of ’s.
Assume now that the distribution of ’s is , where is a nonnegative bounded Borel function on satisfying the following mild regularities:
(1.3) |
for some constant independent of . This condition was introduced in [13, 15]. We want to study the distribution of zeros of as . We denote by the current of integration along the zero divisor of . Note that if , then is the sum of Dirac masses at zeros of .
If is Bernstein-Markov, it was proved in [15, Theorem 4.2], that almost surely
(1.4) |
as , where the convergence is the weak one between currents. In other words, for every smooth form of degree with compact support in , one has
as . Theorem 1.3 above thus provides us a large class of measures for which the equidistribution of zeros of 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 and of the distribution of . We don’t try to make the most optimal condition. Here is our hypothesis:
(H1) for sufficiently large.
(H2) let be a non-degenerate piecewise-smooth generic Cauchy-Riemann submanifold of , and be a Hölder continuous function on . Let , where for some constant .
The condition (H1) ensures that (1.3) holds, and the joint-distribution of is dominated by the Fubini-Study volume form on (by definition the Fubini-Study volume form is equal to , where is the Fubini-Study form on ). Clearly the Gaussian random variables satisfy this condition.
The condition (H2) is a natural generalization of the classical setting with Kac polynomials where is the unit circle in . Indeed in [58] the authors considered the setting in which is the surface area on a closed analytic curve in bounding a simply connected domain in or is the restriction of the Lebesgue measure on to . 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 . 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 , and closed positive currents of bi-degree on the complex projective space , define
where denotes the greatest integer less than or equal to , and is a smooth form of degree on .
It is a standard fact that the distance for 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 , there is a constant such that
(1.5) |
Note that the currents and extends trivially through the hyperplane at infinity to be closed positive currents of bi-degree on (this is due to the correspondence (2.1) above). Hence one can consider between and as closed positive currents on .
Theorem 1.7.
(A large deviation type estimate) Let be a constant. Assume that (H1) and (H2) are satisfied. Then there exists a constant so that
(1.6) |
for every , where denotes the joint-distribution of .
By (1.5), one obtains similar estimates for with 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 instead of ); see also Theorem 1.9 below. Denote by the expectation of the random normalized currents of zeros .
Corollary 1.8.
Assume that (H1) and (H2) are satisfied. Then we have
(1.7) |
where denotes a current of order in such that for every smooth form of degree with compact support in such that , we have
for some constant independent of .
Even when the ’s are Gaussian variables, we underline that the decay obtained in [58] is only , 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 be a complex line in or an (complex) algebraic curve in . Sine generic polynomials intersect transversely , almost surely the number of intersection points (without counting multiplicities) of the random hypersurface and is exactly by Bezout’s theorem. Define
where are zeros of on . Let be the current of integration along . Since is bounded, the product
is a well-defined measure supported on (it is simply if is smooth).
Theorem 1.9.
Let be a constant. Assume that (H1) and (H2) are satisfied. Then there exists a constant so that
for every . In particular, the measure converges weakly to as .
Now since zeros of on is discrete and is equidistributed
as , one can ask as in [58]
how zeros of on (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 to be a transcendental curve in .
In this case generic polynomials still intersect
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 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 be a projective manifold of dimension . Let be an ample line bundle equipped with a Hermitian metric whose Chern form is positive. Let be a compact non-pluripolar subset in . Let be a probability measure on such that the support of is non-pluripolar and is contained in . Let be a Hermitian metric on such that , where is a continuous function on . For , we define
Since is non-pluripolar, the last scalar product defines a norm called -norm on . Let . We obtain induced Hermitian metric on and a similar norm on . Put . Let be an orthonormal basis of with respect to . The Bergman kernel function of order associated to is
for .
When is a volume form on and , 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 and is the hyperplane line bundle on endowed with the Fubini-Study metric. We consider as an open subset in and the weight corresponds to . Recall that there is a natural identification between and the set of -psh functions on (where denotes the Fubini-Study form on ) given by
(2.1) |
for .
Another well-known example is the case where is the unit sphere in (here ; see, e.g, [50]) and is the complexification of , i.e, which is considered as usual a compact subset of . The line bundle on is the restriction of the hyperplane bundle to . We remark that in this case is equal to the restriction of the space of to . Hence the restriction of to is that of the space of complex polynomials in to . To see this, notice that is a smooth hypersurface in . Consider the standard exact sequence of sheaves:
where the second arrow is the multiplication by a section of whose zero divisor is equal to . We thus obtain a long exact sequence of cohomology spaces:
In the last sequence, is exactly equal to , and by Kodaira-Nakano vanishing theorem; see [38, p. 156]. As above, the weight on in the spherical model corresponds to in the setting .
The measure is said to be a Bernstein-Markov measure (with respect to ) if for every there exists such that
(2.2) |
for every . In other words, the Bergman kernel function of order grows at most subexponentially, i.e, as for every . Theorem 1.3 is a particular case of the following result.
Theorem 2.1.
Let be a compact nondegenerate piecewise-smooth Cauchy-Riemann generic submanifold of . Then for every continuous function on , if is a finite measure whose support is equal to such that there exist constants satisfying for every , and every (where denotes the ball of radius centered at induced by a fixed smooth Riemannian metric on ), then is a Bernstein-Markov measure with respect to .
Let
(2.3) |
Since is non-pluripolar, the function is a bounded -psh function. If , then we say is regular. A stronger notion is the following: we say that is locally regular if for every there is an open neighborhood of such that for every increasing sequence of psh functions on with on , then
on . The following result answers the question raised in [9, Remark 1.8].
Theorem 2.2.
Every compact nondegenerate piecewise-smooth Cauchy-Riemann generic submanifold of 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 is locally regular if 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 is called the equilibrium measure associated to . It is well-known that the last measure is supported on . By [9, Theorem B] one has
(2.4) |
provided that is a Bernstein-Markov measure associated to . The last property suggests that the Bergman kernel function cannot behave too wildly at infinity.
Theorem 2.3.
Let be a compact nondegenerate piecewise-smooth Cauchy-Riemann generic submanifold of . Let be the dimension of . Let be a Hölder continuous function of Hölder exponent on , let be a smooth volume form on , and , where and for some constant . Then, there exists a constant such that
for every .
Note that by the proof of [29, Theorem 1.3] or [30, Theorem 3.6], for every Hölder continuous function on , and , the Bergman kernel function of order associated to grows at most polynomially on as ; see also [8, Theorem 3.1] for the case where is smooth.
Theorem 2.4.
Assume that the following two conditions hold:
(i) is maximally totally real and has no singularity (i.e, is smooth and without boundary),
(ii) for some constant .
Then there exists a constant such that for every , the following holds
for every .
Consider the case when , , is be the Fubini-Study metric on , and is a smooth maximally totally real compact submanifold in , and is the trivial line bundle on . It is clear that is in if 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 be a compact nondegenerate piecewise-smooth generic submanifold of . Let be a Hölder continuous function on . Let be a smooth volume form on . Then we have
as , where
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 be a compact Kähler manifold of dimension with a Kähler form . Let be a compact non-pluripolar subset on and be a continuous function on . Recall
By non-pluripolarity of , we have . Hence is a bounded -psh function on .
When and , it was proved in [10, 59, 22] that . In general, the best regularity for is Hölder one, see Theorem 3.9 below and comment following it. One can check that if is locally regular, then is regular for every .
Let be the open unit disc in An analytic disc in is a holomorphic mapping from to which is continuous up to the boundary of For an interval is said to be -attached to a subset if Fix a Riemannian metric on and denote by the distance induced by it. For and , let be the ball of radius centered at 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 be a compact generic nondegenerate piecewise-smooth submanifold of . Then, there are positive constants and such that for any and any there exist a analytic disc such that is -attached to with , and there is so that and Moreover if is in a fixed compact subset of the regular part of , then we have .
It was stated that instead of 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 is not necessary in the above result. In particular, if for some open subset of , then the analytic disc can be chosen to lie entirely in . Here is a slight improvement of Proposition 3.1.
Proposition 3.2.
Assume that one of the following assumptions hold:
(1) is a compact generic smooth submanifold with smooth boundary that is also generic,
(2) is a union of a finite number of compact sets as in (1).
Then there are positive constants and such that for any and any there exist a analytic disc such that is -attached to with , and there is so that .
Proof.
If fulfills one of the conditions (1) or (2) then can be covered by a finite number of sets such that for every there exist an open subset in and a smooth family of smooth generic CR submanifolds in such that is smooth without boundary in , for every , and satisfies . Now the desired assertion follows directly from Proposition 3.1 applied to each and points in correspondingly. We note that the constants can be chosen independent of because as shown in the proof of [64, Proposition 2.5], they depend only on bounds on -norm of diffeomorphisms defining local charts in (see [64, Lemma 4.1]), these bounds are independent of for the family is smooth. ∎
Examples for compact satisfying the hypothesis of Proposition 3.2 include a collection of finite number of smooth Jordan arcs in regardless of their configuration or the closure of an open subset with smooth boundary in .
Lemma 3.3.
Let and let be a constant. Let be a subharmonic function on . Assume that
(3.1) |
Then, there exists a constant depending only on so that for any we have
(3.2) |
Moreover if for some function on and for some so that , then
(3.3) |
for some constant independent of .
Proof.
The desired inequality (3.2) is essentially contained in [64, Lemma 2.6]. The hypothesis of continuity up to boundary of 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 because the harmonic extension of a Lipschitz function on is not necessarily Lipschitz on . However since the harmonic extension of a function on to is also on (see, e.g, [36, Page 41]), we obtain (3.3). ∎
End of the proof of Theorem 2.2.
Let and a small ball of around . Consider an increasing sequence of psh functions bounded uniformly from above on such that on . We need to check that on . Now, we will essentially follow arguments from the proof of [64, Theorem 2.3]. Let be a relatively compact subset of containing . We will check that there exists a constant such that for every , we have
(3.4) |
The desired assertion is deduced from the last inequality by taking . It remains to check (3.4).
Let be a point in such that . Put . By Proposition 3.1, there exists an analytic disc continuous up to boundary and with such that and , and , for some constants and independent of .
Put . Since on and , we get for . Moreover since is uniformly bounded from above, there is a constant such that for every . This allows us to apply Lemma 3.3 for (for some constant small) and big enough. We infer that . Substituting in the last inequality gives
Hence, (3.4) follows by choosing . The proof is finished. ∎
Let be as in the previous section. Recall that the Chern form of is equal to . Define
Clearly . We recall the following well-known fact.
Lemma 3.4.
The sequence increases pointwise to as .
As a direct consequence of the above lemma, we see that 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 is bounded, without loss of generality, we can assume that . Clearly, . Fix . Let be a positive constant. Let be a negative -psh function with on such that . Let . Observe . This allows us to apply [27, Theorem 14.21] to . Let is an orthonormal basis of with respect to -norm generated by the Hermitian metric and . Set
Then
and converges pointwise to as . Note that
Let be a sequence of continuous functions decreasing to as . Using Hartog’s lemma applied to , we see that there is a sequence increasing to such that
Consequently, converges pointwise to as .
Recall that on . By Hartog’s lemma again and the continuity of , for every constant and large enough, we have
(3.5) |
on . It follows that
(3.6) |
for every with . We deduce that
for such . In other words, for big enough. Letting gives
Letting tend to yields that . Hence as . We have actually shown that for every sequence converging to , there is a subsequence such that converges to . Thus the desired assertion follows. ∎
Lemma 3.5.
Assume that is regular. Then converges uniformly to as .
Proof.
For readers’ convenience, we briefly recall the proof here. Since , we see that is upper semi-continuous. This combined with the fact that is already lower semi-continuous gives that is continuous. By Lemma 3.4, we have the pointwise convergence of to . Using the envelop defining , observe next that
(3.7) |
for every .
We fix a Riemannian metric on . Let . Since is compact, is uniformly continuous on . Hence there exists a constant such that if for every . Fix . Let be a natural number such that for , we have
Since the line bundle is positive, is continuous for big enough. Hence without loss of generality we can assume that is continuous for every , for only big matters for us. By shrinking if necessary, we obtain that for every , one has
if . Write for . Using this and (3.7) yields
It follows that
Thus
The choice of now implies that there exists a constant such that the right-hand side is bounded from below by if . Since , we obtain the uniform convergence of to . ∎
Put
Proposition 3.6.
Assume that is regular and satisfies the Bernstein-Markov property. Then we have
(3.8) |
In particular,
(3.9) |
Note that the limit in (3.9) is independent of . We refer to [7, Lemma 2.8] for more informations in the case .
Proof.
When and , 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
(3.10) |
Combining this with the Bernstein-Markov property, we see that for every , there holds
(3.11) |
for every . Observe also that
(3.12) |
on . Applying the last inequality to and using (3.11) we infer that
In other words, on . On the other hand, if , then
for some constant independent of . It follows that
Consequently . Thus using Lemma 3.5 we obtain the desired assertion. This finishes the proof. ∎
Remark 3.7.
Recall on . If is a point so that , then by Proposition 3.6 we see that grows exponentially as . Consider now the case where and is not an -psh function. In this case there exists with , and hence becomes exponentially small as .
3.2 Hölder regularity of extremal plurisubharmonic envelopes
Let and be a metric space. We denote by the space of functions on of finite -norm. If , then we also write for . The following notion introduced in [30] will play a crucial role for us.
Definition 3.8.
For and a non-pluripolar compact is said to be -regular if for any positive constant the set is a bounded subset of
The following provides examples for the last notion.
Theorem 3.9 ([64, Theorem 2.3]).
Let be an arbitrary number in Then any compact generic nondegenerate piecewise-smooth submanifold of is -regular. Moreover if has no singularity, then is -regular.
Remark 3.10.
If is as in Proposition 3.2, then is also -regular for . The union of a finite number of open subsets with smooth boundary in is an example of such .
If , and , then it was shown in [22] that , hence is -regular; see also [10, 22, 59] for more information. In the case where or is an open subset with smooth boundary in it was proved in [30] that is -regular for . This was extended for 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 . 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 , hence in particular, is for some ).
Theorem 3.11.
Let be constants. Let be a compact generic smooth submanifold (without boundary) of . Then there exists a constant such that for every with , then .
We mention at this point an example in [56] of a domain with boundary but is not regular even for . Applying Theorem 3.11 to we obtain the following result that implies [57, Conjecture 6.2] as a special case.
Theorem 3.12.
Let . Let be a compact generic nondegenerate piecewise-smooth submanifold in . Let
Then . Additionally if has no singularity, then .
We note that the fact that when 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 be a compact generic nondegenerate piecewise-smooth submanifold in . Then there exists a constant such that for every complex polynomial on we have
(3.13) |
If additionally has no singularity (e.g, ), then
(3.14) |
Note that the exponent of is optimal as it is well-known for the classical Markov and Bernstein inequalities in dimension one. The above result was known when is algebraic in , 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, and considering as a maximally totally real submanifold in the complexification of .
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
which is if has no singularity or in general by Theorem 3.12. Let be a complex polynomial in . Put . Since is a candidate in the envelope defining , we get
on . We use the same notation to denote a constant depending only on . Let . Let be a small constant. Consider the analytic disc . Applying the Cauchy formula to the restriction of to shows that
Since on , using regularity of gives
for some constant independent of and . Choosing in the last inequality yields
Similarly we also get
for every . Hence the first desired inequality (3.13) for general . When has no singularity, the arguments are similar. This finishes the proof. ∎
Here is a quantitative version of Lemma 3.4.
Proposition 3.14.
Let be a compact generic nondegenerate piecewise-smooth submanifold of . Let be a Hölder continuous function on . Then, we have
Proof.
The desired estimate was proved for in [29, Corollary 4.4]. For the general case, we use the proof of [64, Theorem 2.3] (and [30]). Let be the continuous extension of to as in the proof of [64, Theorem 2.3]. It was showed there that . On the other hand, one can check directly that . Hence, we get
for every . 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
It remains to bound from above . Combining the polynomial upper bound for in Theorem 2.3 and (3.10), one gets, for some constants independent of ,
for every . This coupled with (3.12) yields
on . It follows that
This finishes the proof. ∎
4 Polynomial growth of Bergman kernel functions
4.1 Families of analytic discs attached to
The goal of this part is to construct suitable families of analytic discs partly attached to . 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 in this section.
Here is our result giving the desired family of analytic discs.
Theorem 4.1.
Let be a constant. Then there exist constants and such that for every , the following properties are satisfied. Let be a regular point of of distance at least to the singularity of , and let be a local chart around in such that corresponds to the origin in . Then there is a map such that the following properties are fulfilled:
(i) is holomorphic for every , and
for every , .
(ii) for and , and
(iii) Let denote the restriction of to . Then is bijective onto its image, and the image of is contained in (here denotes the ball centred at of radius in ), and
for every and .
We proceed with the proof of the last theorem. Denote by the complex variable on and by the variable on For any and let be the Euclidean ball centered at of radius of , and for , we write for . Let be a compact submanifold with or without boundary of The Euclidean metric on induces a metric on For and , let be the space of real-valued functions on which are differentiable up to the order and whose derivatives are Hölder continuous of order For any tuple consisting of functions in , we define its -norm to be the maximum of the ones of its components.
Let be a continuous function on . Let
which is a holomorphic function on . Recall that the real part of is . Let denotes the imaginary part of . Put
We now go back to our current situation with . We endow with an arbitrary Riemannian metric. For , let denotes the arc of of arguments from to . Let be a regular point in and let denotes the distance of to the singular part of . Recall that we assume in this section that .
Lemma 4.2.
There exist a constant depending only on and a local chart around where is biholomorphic with such that the two following conditions hold:
we have
there is a map from to so that , and
where the canonical coordinates on are denoted by and
(4.1) |
Note that is indeed (because is so), but is sufficient for our purpose in what follows.
Proof.
The existence of local coordinates so that is standard, see [3] or [64, Lemma 4.1]. Perhaps one needs to explain a bit about the radius : the existence of comes from the fact that for every singular point in , there are an open neighorhood of in and sets for such that is smooth generic CR submanifold of dimension in , and
and is the closure in of a relatively compact open subset in , and (in ) is contained in the singularity of . Thus by applying the standard local coordinates to points in we obtain the existence of . This finishes the proof. ∎
From now on, we only use the local coordinates introduced in Lemma 4.2 and identify points in with those in via
Lemma 4.3 ([64, Lemma 3.1]).
There exist a function such that for and .
In what follows, we identify with Let be a function described in Lemma 4.3. Let Define and and Let be a positive number in which plays a role as a scaling parameter in the equation (4.2) below.
In order to construct an analytic disc partly attached to , it suffices to find a map
which is Hölder continuous, satisfying the following Bishop-type equation
(4.2) |
where the Hilbert transform 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 and a real number satisfying the following property: for any and any the equation (4.2) has a unique solution which is in and such that
(4.3) |
for any and .
From now on, we consider (hence the distance from to the singularity of is at least ). Let be the unique solution of (4.2). For simplicity, we use the same notation to denote the harmonic extension of to Let be the harmonic extension of to Recall that
Lemma 4.5.
We note that the hypothesis that was required in [63], but it is actually superfluous in the proof of Lemma 4.5. Define
which is a family of analytic discs parametrized by Compute
Hence if is small enough, we see that
(4.5) |
This combined with Lemma 4.5 yields that
(4.6) |
if and are small enough. Now the defining formula of and the fact that on imply that
by (4.6) if and are small enough. In other words, there is a small constant such that if , then is -attached to .
Proposition 4.6.
By decreasing and if necessary, we obtain the following property: for every the map is a diffeomorphism onto its image, and
for some constant independent of .
Proof.
The desired assertion was implicitly obtained in the proof of [63, Proposition 3.5]. We present here complete arguments for readers’ convenience. Recall
By the Cauchy-Riemann equations, we have
The last term is by Lemma 4.5. Thus the first component of is greater than provided that small enough. A direct computation gives Consequently, the first component of
is greater than for (note that ). Moreover, as computed above, we have
Thus is a nondegenerate matrix whose determinant satisfies the desired inequalities if is small enough. The proof is finished. ∎
End of proof of Theorem 4.1.
Let be as above and smaller than and be a big constant. Fix a parameter and define
By the above results, we see that the family satisfies all of required properties (because ). 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 . Then there exists a constant such that for every and every constant , and every subharmonic function on such that is continuous up to and for , we have
Proof.
Let . Put
Let be the harmonic function on such that , and for and on . Observe that for some constant independent of .
By a classical result on harmonic functions on the unit disc (see [36, Page 41] or (3.4) in [64]), we have
for every . As a result, we get
(4.7) |
Let be the harmonic function on such that for , and for . Observe that because the latter function is harmonic and greater than or equal to on the boundary of . Using Poisson’s formula, we see that
Summing this and (4.7) gives the desired assertion. The proof is finished. ∎
Let be a smooth (without boundary) maximally totally real submanifold in . Let with and . Let . Consider a local chart around with coordinates , and corresponds to the origin in . Shrinking if necessary we can assume also that is trivial on .
We trivialize over such that for some psh function on with (we implicitly fix a local holomorphic frame on so that one can identify Hermitian metrics on with functions on ), and identify with a holomorphic function on . Thus
on . In particular on .
Lemma 4.8.
There exists a constant independent of such that
(4.8) |
and
(4.9) |
for . Moreover for every constant , there exist a constant independent of such that
for .
Proof.
The desired inequality (4.9) follows immediately from (4.8) and the equalities
Recall that
Note that on . By hypothesis for some constant . This combined with Theorem 3.11 and the fact that has no singularity yields that is Lipschitz. Using the fact that is -psh, we get the Bernstein-Walsh inequality
Hence
This combined with the Lipschitz property of and (and also the property that on ) yields
(4.10) |
for some constant independent of . 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 , there holds
(4.11) |
Consequently, for with big enough, we get
This finishes the proof. ∎
Proof of the upper bound in Theorem 2.4.
Let with . Let . We need to prove that .
Let and consider a local chart around with coordinates , the point corresponds to in the local chart . Let be a small constant to be chosen later. Let be the constants in Lemma 4.8. Let be a big constant. Using the Lipschitz continuity of yields
(4.12) |
(uniformly in ). Now let be the family of analytic discs in Theorem 4.1 associated to for and , where . Let be the constant in the last theorem.
Let . Put . By expressing
one gets
if is big enough. Applying Lemma 4.7 to and and using (4.9) yield
for some constant independent of . This combined with Lemma 4.8 yields
Integrating the last inequality over gives
if . By Properties of and (4.12), the second term in the right-hand side of the last inequality is . We infer that
where is a constant independent of . Consequently by (4.9), one gets
for every . By choosing and big enough as required, one gets
Thus the desired upper bound for follows. The proof is finished. ∎
We now proceed with the proof of Theorem 2.3.
End of proof of Theorem 2.3 for .
We assume . We will explain how to treat the case later. Let with . Put . Let be a Hölder exponent of . We have (note ). We follow essentially the scheme of the proof for the upper bound of in Theorem 2.4.
Denote by the set of points in of distance at least to the singular part of , and the set of points in of distance at most to . As in the proof of Lemma 4.8, one has
(4.13) |
for some constant independent of . In particular it is sufficient to estimate for .
Let (hence is a regular point of , and the ball lies entirely in the regular part of ) and a local chart around with coordinates , the point corresponds to in the local chart .
We trivialize over such that for some smooth psh function on with (we implicitly fix a local holomorphic frame on so that one can identify Hermitian metrics on with functions on ), and identify with a holomorphic function on . Thus
on . In particular on . Using the Hölder continuity of yields
(uniformly in ). As in the proof of Lemma 4.8, a Berstein-Walsh type inequality implies that by increasing if necessary (independent of ) there holds
(4.14) |
for . Let . One also sees that there is a constant independent of such that
(4.15) |
if . Since , where , applying Hölder inequality to gives
(4.16) |
Let be a constant. Now let be the family of analytic discs in Theorem 4.1 associated to for and .
Let . Put which is by properties of if is big enough. Applying Lemma 4.7 to and yield
(again we increase if necessary independently of ). By this and (4.15), we gets
Integrating the last inequality over gives
(4.17) |
The second term in the right-hand side of the last inequality is bounded by
Using this, (4.17) and (4.16) gives
for some constant big enough (independent of ). This combined with (4.14) gives
for every . By this and (4.13) we obtain
by choosing big enough. Hence This finishes the proof. ∎
We now explain how to prove Theorem 2.3 when is not necessarily totally real.
End of proof of Theorem 2.3 for .
As in the case of , it suffices to work with points in of distance at least to the singularity of for some big enough. Let be such a point, and let with . Our goal is to show that for some constant independent of and . We identify with a holomorphic function on a small open neighborhood of as usual. Hence as above one get
(4.18) |
where is as in the case of .
Let be the CR dimension of . Recall that . Using standard local coordinates near on , one sees that by shrinking if necessary, there are holomorphic local coordinates such that is locally given by the graph , for . In particular for every vector (small enough) the real linear subspace intersects at a generic CR smooth submanifold in . Put . Let be a big constant to be chosen later. By Fubini’s theorem and (4.18), we obtain
It follows that there exists with so that
Applying the proof of the case where to , we see that
(4.19) |
where we write , and is identified with in these local coordinates. On the other hand since , using a version of Bernstein-Markov inequality (similar to (4.11)) yields
if is big enough. This combined with (4.19) gives the desired upper bound. The proof is finished. ∎
We end the subsection with a remark.
5 Zeros of random polynomials
In this section we prove Theorem 1.7. Let be the Lebesgue measure on for , and we denote by the standard Euclidean norm on . Let be the Fubini-Study form on , and let be the Fubini-Study volume form on . We always embed in . We recall the following key lemma.
Lemma 5.1.
([32, Proposition A.3 and Corollary A.5]) There exist constants such that for every , and every -psh function on with , then
for every .
The essential point is that the constants in the above result is uniformly in the dimension of . In our applications, the dimension will tend to . Let be the dimension of the space of polynomials of degree at most on . Note that .
Let be a bounded Borel function on such that there is a constant for which for every we have
Let be complex-valued i.i.d random variable whose distribution is . Assume furthermore that the joint-distribution of satisfies
on .
Let be an orthonormal basis of . Let be a complex algebraic subvariety of dimension in . Note that the tolological closure of in is an algebraic subvariety in . Observe that
where is the standard Kähler form on , and is a constant. Fix a compact of volume in . We start with a version of [15, Lemma 2.4] with more or less the same proof. We use the Euclidean norm on .
Lemma 5.2.
Let be a constant. Let be the set of such that
Let be the set of so that . Then we have
for some constant independent of and .
Proof.
Put , and
Observe
It follows that for , one has
This implies that for each there exists such that . We infer that
Similarly, we also get . This finishes the proof. ∎
Put , and . Define
Observe that on . We put
Lemma 5.3.
There exists a constant such that for every we have
Proof.
Let be the right-hand side of the desired inequality. By Funibi’s theorem and the transitivity of the unitary group on , one has
The function
is -psh on (where is the Fubini-Study form on ). Let . Thus and is -psh. Applying Lemma 5.1 to gives
for some constant independent of . Consequently,
for every . In particular for , we obtain
Thus the desired inequality follows. ∎
End of the proof of Theorems 1.7 and 1.9.
Let . We have . By Lemma 5.1 again, there are constants independent of such that
Combining this with Lemma 5.3 yields
Let be a constant. Choosing , where big enough gives
(5.1) |
Let be the set of such that and . Combining (5.1) and Lemma 5.2, we obtain that
(we increase if necessary). On the other hand, by the definition of and , we see that the set of such that
(for some constant big enough independent of ) contains . It follows that
This together with Lemma 5.2 implies that there exists a Borel set such that , and for , one has , and
Consequently for , and , there holds
Moreover , one has
It follows that
for some constant independent of . This together with Theorem 1.6 implies
by increasing if necessary. It follows that
for . Here recall that we define by considering and as currents on . 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 , 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 , 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 , 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 , Invent. Math., 173 (2008), pp. 315–363.
- [21] , Bernstein type inequalities for restrictions of polynomials to complex submanifolds of , 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 , 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, -regularity of sets in , 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 , 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]