Overcrowding Estimates for zero count and nodal length of stationary Gaussian processes
Abstract.
Assuming certain conditions on the spectral measures of centered stationary Gaussian processes on (or ), we show that the probability of the event that their zero count in an interval (resp., nodal length in a square domain) is larger than , where is much larger than the expected value of the zero count in that interval (resp., nodal length in that square domain), is exponentially small in .
1. Introduction
1.1. Main Results
In this paper we consider centered stationary Gaussian processes on /, study the unlikely event that there is an excessive zero count/nodal length in a region and obtain probability estimates for them. Let be a function such that for every we have . Then its zero/nodal set is a smooth one-dimensional submanifold. Let be any function and , we define and to be the zero count of in and the nodal length of in respectively, that is
When there is no ambiguity of the function under consideration, we simply use and .
Let be a centered stationary Gaussian process on / whose spectral measure is a symmetric Borel probability measure. Stationarity of implies that and , under the assumption that has a finite second moment these expectations are also finite. Under some assumptions on , we get probability estimates for the unlikely events and . The following are the assumptions we make on and their significance.
- •
- •
Lemmas 2.1 and 2.7 are the quantitative versions of the above statements.
Notations.
Before stating the main results, we introduce some useful notations.
-
1.
Let be a positive Borel measure on . For , we define and by
(1) -
2.
For a positive Borel measure on , we denote by the push forward of by the map , where is defined by . For , we define the following quantities
(2) The relation between the moments and is discussed in Remark 3.10.
-
3.
Let be a positive Borel measure on /. Then is said to satisfy assumption / if:
is a symmetric Borel probability measure on which has a nontrivial absolutely continuous part w.r.t. the Lebesgue measure on . () is a symmetric Borel probability measure on such that there exist , satisfying and the marginals of on and satisfy (). ()
Theorem 1.1.
Theorem 1.2.
Remark 1.3.
If we fix and take in (3), we can conclude that there are constants such that for every we have
and this indicates short-range repulsion of the zeros of .
In Section 6, we apply Theorems 1.1 and 1.2 to specific classes of spectral measures and get tail estimates for and ; these results are summarized in the following tables.
Growth of | Example of | Constraints on and | ||
, for | Any with supp | , for | ||
, for | , with | , for | ||
, for | , for any | |||
, for |
Growth of | Example of or | Constraints on and | ||
, for | Random plane wave on | , for | ||
, for | , with | , for | ||
, for | , for any | |||
, for |
1.2. Prior work comparison with our results
Before we comment about how our results compare with what is already known, let us briefly review the results known about zero count and nodal length of stationary Gaussian processes. A more detailed and comprehensive account of results known so far can be found in [1, 5, 15].
As before, let be a centered stationary Gaussian processes on / whose spectral measure is a symmetric Borel probability measure. We denote its covariance kernel by . The exact value of the expectations of and are given by the Kac-Rice formulas. Under the assumption that the tails of are light enough and some integrablility assumptions on and , it was shown in [6] that Var and a central limit theorem was also established for . Finiteness of moments of and bounds for these were established in [3, 4, 14] under the assumption that the moments of are finite and more recently the asymptotics of the central moments of were obtained in [1] assuming certain decay of . Under the assumption that has very light tails and is integrable, it was shown in [5] that concentrates exponentially around its mean. Central limit theorems and variance asymptotics for in specific examples of stationary Gaussian processes were obtained in [8, 13]. Under the assumption that moments of are finite, finiteness of moments of were established in [2].
These results can be broadly classified into two categories based on the assumptions made on their spectral measure or covariance.
-
•
Decay of covariance (and light tails of ): The philosophy behind such an assumption is that decay of implies quasi-independence of events in well separated regions and hence for a large , is approximately a sum of identically distributed random variables which are -dependent. Although it is quite easy to state this idea, building on it and making it work is far from trivial. Hence the concentration results in [5, 6] can be viewed as stemming from some underlying independence.
-
•
Finiteness of moments of : In the other results like [3, 4, 14] which established finiteness of moments of , the only assumption made on is finiteness of its moments. The significance of this, as discussed earlier, is that a control on the moments imply a control on the oscillations and hence on the zero count also.
The nature of the results from these two categories also differ: in the former, the assumptions made are strong and so are the conclusions (concentration) and in the latter, the assumptions are much weaker and the conclusions (non-explicit moment bounds) are also weak. In terms of the assumptions made on the spectral measure/covariance, our result falls in the second category, but surprisingly we get quite strong conclusions. To illustrate this claim, we present some consequences of Theorem 1.1; the following are two instances where we get the exact asymptotics for the deviation probabilities :
-
•
If is compactly supported, there is such that if and , then
-
•
If is such that for some , then for every and such that , we have
It is also interesting to compare the exponential concentration result ([5], Theorem 1.1) with Theorem 1.1; both the results give estimates for . The former gives estimates when the fluctuation is of order , while our result gives estimates when and is large (how large depends on the growth of moments ) enough. Hence only in certain situations can we actually compare these two results and one such instance is when is compactly supported and has a density which belongs to . There exists (depending on ) such that for and , the estimates for the deviation probability from [5] and Theorem 1.1 are , and , respectively. So this naturally raises a question about the true deviation probabilities: either the true deviations for are much smaller than or there is some such that the behaviour of the deviation probabilities are different for and which hints at a possible JLM law.
1.3. Plan of the paper
In Section 2, we present the main ideas in the proofs of Theorems 1.1 and 1.2 and also give a sketch of their proofs. The relevant details and calculations required to complete the proofs are presented in Sections 4 and 5. In Section 3, we recall some known results and present other preliminary results required in proving our main theorems. In Section 6, we deduce the tail bounds for and presented in Tables 1 and 2 from Theorems 1.1 and 1.2 respectively.
2. Idea of the proof
In this section, we present the key ideas in the proofs of Theorems 1.1 and 1.2; we also give a brief sketch of their proofs here. The calculations and detailed arguments will be presented in Sections 4 and 5.
2.1. Proof idea of Theorem 1.1
We now present a simple yet very useful idea regarding the zero count of a smooth function and this serves as the starting point for proving the upper bound in Theorem 1.1. This idea has previously been employed in [3, 4, 14] to study moments of the zero count of random functions. The following lemma is inspired by these earlier versions and is a slight modification of those.
Lemma 2.1.
Let and . Let be a smooth function which has at least distinct roots in and . Then for every , we have
Proof.
Let be a smooth function and let be two zeros of , then there exists a zero of in . This along with the fact that has at least distinct roots in implies that there are points such that for every , we have . For , we have
(5) |
Using the estimate for in (5), we have for every
Thus inductively we can establish that for every and every
and this establishes our claim. ∎
We get the following result as an immediate consequence of Lemma 2.1.
Lemma 2.2.
Let be a smooth random function, then for every and we have
(6) |
Upper bound in Theorem 1.1
We use Lemma 2.2 to prove the upper bound in Theorem 1.1 and hence we only need to estimate the two terms on the r.h.s. of (6). The second term is estimated using well known tail bounds for the supremum of Gaussian processes; we recall these standard results in Section 3.2. Since we are working with stationary Gaussian processes, calculating the metric entropy and the corresponding Dudley integral are quite straightforward.
The first term corresponds to a small ball event and the discussion in Section 3.1 is devoted to obtaining probability estimates for such small ball events. We briefly explain how this is done: say and we want to estimate . For and , we define which are equispaced points in . Consider the Gaussian vector defined by Let and be the covariance matrix and the density of respectively, then
where is the smallest eigenvalue of and hence the following is an estimate for the small ball probability
(7) |
Result 3.3 gives a lower bound for and hence an upper bound for the small ball probability in (7). We then use (7) with and an optimal choice of to get an esimate for the first term in (6).
For given values of and , we need to make an optimal choice of (it should be small enough so that the first event is indeed a small ball event and large enough so that the second event is unlikely) so that the r.h.s. of (6) is as small as possible.
Lower bound in Theorem 1.1
We use the same notations as above and take . If the sign of a continuous function defined on alternates at the points , then there is necessarily a 0-crossing of the function between every two successive points and hence the zero count in is at least . Thus we have
(8) |
and hence to get a lower bound for the above probability, we need a lower bound on the density and we get this below. being a Gram-matrix with all diagonal entries equal to 1, we have ([9], Lemma 3) and hence
(9) |
As before, we use the lower bound for from Result 3.3 to get a lower bound on the density in (9) and then use this to get a lower bound for the l.h.s. of (8).
Remark 2.3.
Result 3.3 gives a lower bound for and the only assumption on required to establish this result is (). Hence it is the same assumption, namely (), which gives both the following bounds:
-
•
an upper bound for the small ball estimate in (7),
-
•
a lower bound for the probability of the event in (8).
This is not very surprising and we can perceive this as: the propensity of to oscillate makes it difficult for it to be confined to a small ball.
2.2. Proof idea of Theorem 1.2
We now present some deterministic results about nodal length of a smooth function which will be useful in proving Theorem 1.2. Let us first introduce some notations. For a smooth function, and , we define and as follows
The following result in an easy consequence of Lemma 2.1.
Corollary 2.4.
Let and be such that . Suppose is a smooth function such that and , then has no more than zeros in the interval .
The following integral geometric result gives a bound for the nodal length of a smooth function and it appears in the proof of Lemma 5.11 from [7].
Result 2.5.
Let be a smooth function and let , then we have the following bound for the nodal length of
The following lemma is an immediate consequence of Corollary 2.4 and is inspired by Lemma 5.11 in [7].
Lemma 2.6.
Let , be such that . Let be a smooth function satisfying the following
Then for every , we have .
Proof.
Let be such that , then for every we have
Thus we have and . Corollary 2.4 now gives the desired result. ∎
The following lemma which is a direct and easy consequence of Lemma 2.6 is the main ingredient in the proof of Theorem 1.2.
Lemma 2.7.
Let , be such that and define . Suppose that is a smooth function, then we have the following.
-
•
If for every , the following bounds hold for
(10) then for every we have .
-
•
If for every , the following bounds hold for
(11) then for every we have .
Thus if both (10) and (11) hold for , we can conclude from Result 2.5 that
The following result is an immediate consequence of Lemma 2.7.
Lemma 2.8.
Let , be such that and define . Suppose that is a smooth random function such that almost surely has no singular zeros, then we have
(12) |
where the events , , and are defined as follows
Proof sketch of Theorem 1.2
The spectral measure of satisfies (), we assume without loss of generality that and in (). Hence both the marginals and are symmetric probability measures on which satisfy (). Let and be centered stationary Gaussian processes on whose spectral measures are and respectively.
We further assume that is not supported on any line , the degenerate case supp (for some ) is much easier and will be analysed in Section 5. In this case, it follows from Bulinskaya’s lemma (Result 3.9) that almost surely does not have singular zeros. Hence we can use Lemma 2.8 and the stationarity of to conclude that
(13) |
and we now see how to get bounds for the terms on the r.h.s. of (13). Since is a stationary Gaussian process, so are all its derivatives. Like in the one dimensional case, estimating and using the well known tail bounds for supremum of Gaussian processes is quite straightforward and we do this in Section 5.
We now observe that and . Hence and correspond to small ball events for and respectively. By our assumption, both their spectral measures satisfy () and a way to estimate these small ball probabilities was already discussed in the few lines leading up to (7).
Similar to the one dimensional case, for fixed we need to make an optimal choice of so that the r.h.s. of (13) is minimized.
3. Preliminaries
In this section we present some preliminary results and recall other known results which will be used in establishing Theorems 1.1 and 1.2.
3.1. Small Ball Probability
We now get small ball probability estimates for stationary Gaussian processes on whose spectral measures satisfy ().
Let be a Gaussian process on an interval and let be distinct points in such that the Gaussian vector is non-degenerate and let be its density. Then for every , we have
(14) |
For a stationary Gaussian process whose spectral measure satisfies (), we can get an estimate for which appears in (14) and this is what is done below.
Result 3.1 ([12], Turan’s lemma).
Let , where and . Then there is a constant such that for every interval , every measurable set and every , we have
where denotes the Lebesgue measure.
Fact 3.2.
Let be a centered stationary Gaussian process on whose spectral measure satisfies (). Hence admits the following decomposition , where . Then there is a constant such that and hence there is a large enough such that , where . For and consider the stationary Gaussian process on defined by
(15) |
Then the spectral measure of is a symmetric probability measure on which is the push forward of by the map given by
Hence considered as a measure on has a nontrivial absolutely continuous part w.r.t. the Lebesgue measure on given by , where
Define , if and are such that , then and hence
(16) |
The following result taken from the proofs of Lemma 3 and Theorem 2 of [9] is the main tool in obtaining the small ball probability estimates.
Result 3.3.
Let be a centered stationary Gaussian process on whose spectral measure satisfies (). Let , and be the Gaussian process on defined in (15). Let and denote the covariance matrix and density of the Gaussian vector respectively. Let be the smallest eigenvalue of . Then there exists and such that whenever , we have
Proof.
Let , , , and be as in Fact 3.2. The density is given by
(17) |
We now get a lower bound on and use this to get an upper bound on . Let and define . Then we have
Hence it follows from (16) that
(18) | ||||
We now use Turan’s lemma (Result 3.1) to get
(19) |
and hence we conclude the following from (16), (18) and (19)
(20) |
for some . Using the lower bound we get for from (20) in (17), we get the desired result
∎
Lemma 3.4 (Small ball probability).
In the following Lemma, we get small ball probability estimates by using Lemma 3.4 with specific values of , and . This will be used in the proofs of Theorems 1.1 and 1.2.
Lemma 3.5.
Proof.
Let denote the expression in the r.h.s. of (21) for , then we have
(22) |
Since and , we have
Also because , and are all greater than , we have the following upper bound for the expression on the r.h.s. of (22)
(23) |
where the inequality in (23) follows from our assumption that and this establishes our claim. ∎
3.2. Metric Entropy and Supremum of Gaussian Processes
We now recall well known results which give tail bounds for the supremum of Gaussian processes and we shall use them to get tail bounds for higher derivatives of stationary Gaussian processes. The definitions and results in this section are taken from Section 3.4 of [10].
Let be an index set and let be a centered Gaussian process on . Then induces a pseudo-metric on given by
Suppose is totally bounded and let . The -covering number denoted by is defined to be the minimal number of -balls required to cover . The -packing number denoted by is defined to be the maximal possible cardinality of subset which is such that , for distinct . The quantities and are related by
(24) |
The -entropy number is defined as . The following results are stated in terms of , but because of (24) similar results also hold with replaced by .
Result 3.6 ([10], Theorem 3.18).
Let be a centered Gaussian process and assume that is totally bounded. If is integrable at , then admits a version which is almost surely uniformly continuous on and for such a version we have
where .
Fact 3.7.
For a centered Gaussian process on an index set , we have
(25) |
The following result gives tail bounds for the supremum of a Gaussian process.
Result 3.8 ([10], Proposition 3.19).
Let be an almost surely continuous centered Gaussian process on the totally bounded set . Let denote either or and let . Then for every , we have
3.3. Derivatives of a stationary Gaussian process
We now present some basic facts about stationary Gaussian processes on /. Let be a centered stationary Gaussian process on / with spectral measure and covariance function . Further assume that is a probability measure all of whose moments / are finite. Then we have the following.
-
1.
For on and , we have
-
•
is smooth and for every , is a centered stationary Gaussian process with covariance function given by
(26) -
•
The derivatives of are given by
(27)
-
•
-
2.
For on and and , we have
-
•
is smooth and for every , is a centered stationary Gaussian process with covariance function given by
-
•
The derivatives of are given by
-
•
Bulinskaya’s lemma ([11], Lemma 6) for a stationary Gaussian process reads as follows.
Result 3.9.
Let be a stationary Gaussian process on with spectral measure which is a symmetric probability measure on . Then almost surely does not have any singular zero (a singular zero is a point such that ) if the following conditions holds.
-
•
.
-
•
The Gaussian vector is non-degenerate, which holds iff there is no for which supp.
Remark 3.10.
We now see how the moments and defined in (2) are related. We first note that
and hence . We also have
and hence we conclude that
(28) |
4. Proof of Theorem 1.1
Upper bound
As discussed in Section 2.1, we use Lemma 2.2 to prove the upper bound in Theorem 1.1 and hence we only need to estimate the two terms on the r.h.s. of (6). Let , be as in Lemma 3.4 and define , where . Let , and be such that . We also assume that , where is defined by
With as in (1), we choose , where .
For this choice of , we get the following estimate for the small ball probability in (6) from Lemma 3.5
(29) |
We now estimate , which is the other term in (6), using Results 3.6 and 3.8. Let be the covariance function of . The following calculation is to estimate the entropy numbers for the process . For this we first get an estimate of the -covering number. We denote by the pseudo-metric on induced by .
(30) |
where the last inequality follows from mean value theorem and (27). Let denote the ball in with center and radius . Then it follows from (30) that
and hence we get the following bound for the -covering number of
(31) |
We now get a bound on the -packing number using (24) and (31)
The above inequality holds for and for , we have . Hence we have
(32) |
We now conclude from Result 3.6 that
(33) |
It now follows from (25) and (33) that with , we have
(34) | ||||
It follows from Result 3.8 that for every and every , we have
and since , we conclude that
(35) |
Hence we conclude from (6), (29) and (35) that
Lower bound
Let be as in Result 3.3. For and such that , we define the dimensional Gaussian vector by
where . Let denote the covariance matrix of . The density of is
Let be the smallest eigenvalue of , then we get the following lower bound for from Result 3.3. There is such that , hence we have
We have from Lemma 3 in [9] that and hence we have the following lower bound for the density
It follows from (8) that
5. Proof of Theorem 1.2
We continue to work with the assumptions and the notations of Section 2.2. The starting point of our analysis here is (13), which is
By our assumption, both , satisfy the assumptions of Lemma 3.4 and let , be such that the conclusion of Lemma 3.4 holds for both , with these constants. Define , where . Let , and be such that and . We further assume that and since we have from (28) that , it follows that , where is defined by
We choose , where and estimate the terms on the r.h.s. of (13). We first estimate the small ball probabilities using Lemma 3.5 with and
and since , we have
(36) |
where the last inequality follows from our assumption that .
Like in the one dimensional case, Results 3.6 and 3.8 will be used to estimate the probability of the events and . We now calculate the expected supremum of the process on . Let be the pseudo-metric on induced by the Gaussian process , then for we have
For , we have
Hence we have , where and are defined in (2). We let and denote the balls centered at with radius in the Euclidean metric and respectively, then and hence , the -covering number of , has the following upper bound and hence the -packing number satisfies . Similar to (32), we can get the following estimate for the Dudley integral in this case also
We thus conclude from Result 3.6 that
and by a calculation similar to (34), we conclude that with we have
thus we get the following tail bound from Result 3.8 and the fact that
For and , , similar calculations as above yield the following tail bounds. For every , we have
(37) |
Since , it follows from (37) that
(38) |
It now follows from (13), (5) and (38) that
(39) |
We have from (28) that and now the desired result follows from (39).
Now suppose that there exists such that supp, without loss of generality we may assume that . In this case considered as a measure on satisfies () and almost surely for every , we have and hence the zero set of is . If we let denote the zero count of in , we have iff . The result now follows by using the estimates for obtained in Theorem 1.1.
6. Consequences of Theorems 1.1 1.2
6.1. One dimension
In this section, we deduce from Theorem 1.1 the tail estimates for the zero count given in Table 1.
- 1.
-
2.
Let be such that for some , then . We consider the cases and separately and make the following conclusions from (3).
-
•
Let , then for every there is such that whenever , we have
(42) Justification for (42). Let be such , then we can choose such that and hence for this choice of and large enough , we have
and hence
And if we choose large enough, the fact that will imply that and hence all the conditions of Theorem 1.1 are satisfied. Using chosen above in (3) and letting , we get (42). Similar arguments can also be used to justify the tail bounds we get for in the other cases, namely (43) and (46). Moment bounds. By a calculation similar to (41), we conclude using (42) that
-
•
Let , then for every there exists such that whenever is large enough and , we have
(43) We write as a union of many subintervals , each of length . Then if , at least one of the intervals must contain more than zeros and hence a union bound gives
(44) and hence there exists such that
(45) Moment bounds. Using (45) we get the following moment bounds for . There exists such that for every , we have
-
•
-
3.
Let be such that for some . We conclude from (3) that there exist constants such that whenever is large enough and , we have
(46) and hence by partitioning into subintervals of length , using the above estimate in each of them and a union bound as in (44) implies there is such that
(47) Moment bounds. We use (47) to conclude that there is such that for every , we have
6.2. Two dimensions
We now illustrate how the overcrowding estimates for nodal length given in Table 2 are deduced from Theorem 1.2. Since this analysis is similar to the one dimensional case considered above, we do this only for the first example in Table 2.
Let be compactly supported, say supp for some . Then and hence by taking in (4) and by letting , we conclude that there are constants such that whenever and is large enough, we have
Moment bounds. We get moment bounds using the above tail estimates. For ,
Acknowledgements
This work was carried out during my Ph.D. under the guidance of Manjunath Krishnapur. I thank him for suggesting me the questions considered in this paper, his patience and encouragement throughout this study and for being very generous with his time and insights on the subject of this paper. I also thank Riddhipratim Basu for helpful discussions.
References
- [1] Michele Ancona and Thomas Letendre. Zeros of smooth stationary Gaussian processes. arXiv e-prints, page arXiv:2007.03240, July 2020.
- [2] D. Armentano, J.-M. Azaïs, D. Ginsbourger, and J. R. León. Conditions for the finiteness of the moments of the volume of level sets. Electron. Commun. Probab., 24:Paper No. 17, 8, 2019.
- [3] Diego Armentano, Jean-Marc Azaïs, Federico Dalmao, Jose R. Léon, and Ernesto Mordecki. On the finiteness of the moments of the measure of level sets of random fields. arXiv e-prints, page arXiv:1909.10243, September 2019.
- [4] Jean-Marc Azaïs and Mario Wschebor. Level sets and extrema of random processes and fields. John Wiley & Sons, Inc., Hoboken, NJ, 2009.
- [5] Riddhipratim Basu, Amir Dembo, Naomi Feldheim, and Ofer Zeitouni. Exponential Concentration for Zeroes of Stationary Gaussian Processes. International Mathematics Research Notices, 2020(23):9769–9796, 12 2018.
- [6] Jack Cuzick. A central limit theorem for the number of zeros of a stationary Gaussian process. Ann. Probability, 4(4):547–556, 1976.
- [7] Harold Donnelly and Charles Fefferman. Nodal sets for eigenfunctions of the Laplacian on surfaces. J. Amer. Math. Soc., 3(2):333–353, 1990.
- [8] Marie F. Kratz and José R. León. Central limit theorems for level functionals of stationary Gaussian processes and fields. J. Theoret. Probab., 14(3):639–672, 2001.
- [9] M. Krishna and Manjunath Krishnapur. Persistence probabilities in centered, stationary, Gaussian processes in discrete time. Indian J. Pure Appl. Math., 47(2):183–194, 2016.
- [10] Pascal Massart. Concentration inequalities and model selection, volume 1896 of Lecture Notes in Mathematics. Springer, Berlin, 2007. Lectures from the 33rd Summer School on Probability Theory held in Saint-Flour, July 6–23, 2003, With a foreword by Jean Picard.
- [11] Fedor Nazarov and Mikhail Sodin. On the number of nodal domains of random spherical harmonics. Amer. J. Math., 131(5):1337–1357, 2009.
- [12] Fedor L. Nazarov. Complete version of Turan’s lemma for trigonometric polynomials on the unit circumference. In Complex analysis, operators, and related topics, volume 113 of Oper. Theory Adv. Appl., pages 239–246. Birkhäuser, Basel, 2000.
- [13] Ivan Nourdin, Giovanni Peccati, and Maurizia Rossi. Nodal statistics of planar random waves. Comm. Math. Phys., 369(1):99–151, 2019.
- [14] David Nualart and Mario Wschebor. Intégration par parties dans l’espace de Wiener et approximation du temps local. Probab. Theory Related Fields, 90(1):83–109, 1991.
- [15] Hugo Vanneuville. Overview and concentration results for nodal lines. https://metaphor.ethz.ch/x/2019/hs/401-4600-69L/Reading_group_nodal_1_2.pdf, 2019. Accessed 14/Dec/2020.