Melissa Tacy
Department of Mathematics, The University of Auckland, Auckland, New Zealand
[email protected]
Abstract.
We investigate the small-scale equidistribution properties of random waves, , in . Numerical evidence suggests that such objects display a fine scale filament structure. We show that the X-ray of along any line segment is uniformly equidistributed. So any limiting behaviour must be weaker than scaring. On the other hand, we show that at Planck scale in phase space there are (with high probability) logarithmic fluctuations above what would be expected given equidistribution. Taken together these results suggest that the filament structure may be a configuration space echo of the phase-space concentrations.
Key words and phrases:
Eigenfunction equidistribution at small scales, random waves, filament structure, scaring, phase space
2020 Mathematics Subject Classification:
58J50, 35S05, 35P20, 60B10
In this paper we are concerned with the small-scale structure of random plane waves. Random plane waves are functions of the following form:
where is a set of equally spaced momenta. Exact choices for vary somewhat between papers in this field but is usually a neighbourhood of the unit sphere . The randomness is injected through the coefficients which are chosen according to a probability distribution. Such functions were conjectured, by Berry [3], to provide a good model for the behaviour of chaotic modes in billiard systems.
There are a number of interesting probability distributions from which to draw the . For example we may choose to look at cases where each is chosen as an independent random variable such as Gaussian or Rademacher (see for instance [3], [21],[8]). Alternatively in situations where it is preferable to be able to fix the norm of the coefficients may be chosen from a uniform probability density on the high dimensional sphere (see for example [20],[6],[14],[22] and [11]).
In this paper we treat a model where each is a (independent, identically distributed) Gaussian random variable and is a set of -separated momenta drawn from for . This range of corresponds to the range of random waves for which we have numerical experiments. The website of Alex Barnett [2] records a number of these experiments including videos generated at the AIM workshop Topological complexity of random sets showing the effect of varying . It is convenient to choose so that the momenta are equally spaced, this is however not strictly necessary. What is necessary is that the spacing between momenta is never less than (this is an uncertainty principle requirement) and that the number of momenta in a region scales with the volume of that region. To allow normalisation we restrict our attention to the behaviour of inside the ball of radius one about zero, . Then the variance for the Gaussian random variables is determined by adopting the convention that
This then means that the variance where .
The key questions of quantum chaos (see for example the survey [23]) are concerned with limits of where is a semiclassical pseudodifferential operator. A sequence of on a compact manifold with is said to equidistribute if
where is the unit co-tangent bundle equipped with Liouville measure and is the principal symbol of . This leads to a natural question, are random waves equidistributed?
It is reasonably easy to ascertain that if (the symbol of ) is independent of then random waves are indeed equidistributed. A more subtle question pertains to the small-scale structure. If for example is a symbol that is zero off a small (shrinking in ) region of space in does equidistribution (in an almost sure sense) still hold? If not does a weaker form of equidistribution hold where is only proportional to the phase-space volume of
Figure 1. A random plane waves in dimension two. Reproduced from [2] with permission of Alex Barnett.
Numerical studies of random waves have suggested that at very small scales there are fluctuations that exceed equidistribution. As seen in Figure 1 (reproduced with permission of Alex Barnett) random waves appear to have enhancements along some straight lines. This so named filament structure was intensely discussed at the AIM workshop Topological complexity of random sets (see the report [1]). Some participants taking the opinion that the apparent structure was simply a numerical artefact other believing that it could be quantified. No firm conclusion was reached at the workshop or since. In this paper we take some steps toward understanding what precisely gives rise to this filament structure. As observed in the AIM report [1] the linear structure appears to be constructed of many smaller filaments roughly aligned along a straight lines. The simplest form of failure to equidistribute would then be to have some straight line segments where was unusually large. Due to VanderKam [19] and Zelditch [20] the maximum fluctuation we could expect would be logarithmic. In this paper we will see that this sort of concentration does not occur. In Theorems 2.1 and 2.2 we show that the X-ray transform of the modulus squared of a random wave is uniformly equidistributed with high probability. That is for a unit length line segment, properly contained in , from in direction the random variable obeys
and that there is a (details in Theorem 2.2) so that
(0.1)
In terms of what we might expect to see in a numerical simulation the uniformity (in ) of the tail bound (0.1) is crucial. If for example we were only able to show that for fixed
we could conclude that if we picked a line segment then considered all potential random waves the chance that we would see a failure to equidistribute on that particular segment is small. For an example of this kind of result see [10] (in the setting where the coefficients are chosen according to a uniform distribution on a high dimensional sphere). Such a statement would not preclude the possibility that most random waves have some for which is large. Therefore to rule out structure of this form we do indeed need the stronger bound (0.1). In Section 2 we establish Theorem 0.1 which guarantees uniform equidistribution of .
Theorem 0.1.
Let be the random variable given by
then
Further there is a (explicitly given by in Section 2 by (2.7)) such that if , as ,
and the exception set is given by
independent of . The error is introduced by approximating by . The validity of the approximation follows from a concentration of measure argument which in turn explicitly controls the quantity.
The explicit expressions for this control are found in Theorem 2.1 (equations (2.2), (2.3) and (2.4)).
It is instructive to compare Theorem 0.1 with general results about restriction of spectral clusters to curves. When the requirement that ensures that is a spectral cluster of width one. Therefore we can apply the results of Burq-Gérard-Tvetkov [5] and Hu [13] to obtain, that for any choice of coefficients,
The results of Theorem 0.1 should be interpreted as saying that choices of coefficients that saturate the general bounds are highly unusual.
In the setting of a Riemannian manifold with ergodic geodesic flow the quantum ergodic restriction results of [18] and [9] imply that for density one subsequences of exact eigenfunctions ,
where is the length of the curve . The agreement between this and Theorem 0.1 should be seen as evidence of the agreement between quantum ergodic systems and the predictions of the random wave model.
We then turn our attention to the phase-space picture and consider a random variable which measures the phase-space concentration of near the point given by
(0.3)
Here is a semiclassical pseudodifferential operator that localises in phase space near . The parameter controls the shape of the localisation, the parameter controls the distance from Planck scale. When the operator localises up to the limits allowed by the uncertainty principle. As increases the localisation becomes more relaxed. Normalisations are chosen so that if is phase-space equidistributed . The details of symbol of are described in Section 3. We discover that once is large, that is , we again obtain a uniform equidistribution result similar to that for the X-ray transform of . However when we not only fail to obtain equidistribution but are able to show that with high probability there are logarithmic enhancements near some points . The intermediate scales remain something of a mystery and will likely require very delicate analysis to fully resolve.
Theorem 0.2.
Let be given by (0.3). For and small enough (but independent of ) there exist constants so that
(0.4)
On the other hand if , then for ,
(0.5)
Further if and ,
(0.6)
Remark.
The proof of Theorem 0.2 depends on estimating by an norm where is finite but growing in . Then is directly computed. It is the lower bound on this norm that requires (see equation (3.6)). The key technical dependancies for are determined in Lemma 4.1, in particular the comparison between (4.1) and (4.2).
Therefore once we are at Planck scale we expect to see some regions of phase space that where displays concentration that is logarithmically larger than that predicted by equidistribution. This result should be seen as aligning with Burq-Lebeau’s result [6] that
(0.7)
Indeed when the operator localises to a small ball of radius around . Since is constituted of waves of frequency of size the value of cannot vary much on balls of radius . Taking into account the normalisation we therefore get that
Therefore the results of this paper for should be seen as morally equivalent to the Burq-Lebeau result.
The question then remains, just what are we seeing in numerical simulation that display filament structure? We know from Theorem 0.1 any scaring along a line segment cannot be enough to register in the norm. However it is quite possible to have large number of logarithmically bright points aligned along a ray without the norm along that ray becoming large. To the human eye examining a numerical experiment, such a situation may appear as a scar. This suggests we should look at weaker forms of structure rather than scaring. One possibility, suggested by the phase-space failure to equidistribute, is that the bright points tend to align along rays that have a Planck-scale enhancement in the phase-space sense.
Throughout this paper we will use concentration of measure arguments to allow us to “commute the expectation operator with a function ”. The idea is that if is invertible, we will use measure concentration to find (see Section 1) sufficient conditions so that
Here is a random variable that takes values in . The rational for seeking this kind of estimate is to be able to chose so that is easy (or at least tractable) to compute. Just by assuming is convex and strictly increasing we can get the upper bound (by Jensen). However, typically, we need the full inequality. In some cases we will in fact be able to find lower order control on therefore obtaining control on the “Jensen gap”.
1. Measure concentration
In this section we set up those measure concentration results relevant to this paper. We use those results to find sufficient conditions (on and ) so that
or
where is small compared to . The major measure concentration result we use here is analogous to the Levy concentration of measure for coefficients chosen uniformly on a high dimensional sphere. In that case the key driver of the proof was the isoperimetric inequality on spheres, which states that of all sets of fixed measure the ones with the “smallest perimeter” are hemispheres. To deal with Gaussian variables we simply replace this inequality with the Gaussian isoperimetric inequality, see [16] (Theorem 1) and [4] (Theorem 3.1), which states that of all sets of fixed measure the ones with “smallest perimeter” are half-planes.
Lemma 1.1.
Let be an -tuple of independent Gaussian variables each with variance . Let be the joint -dimensional Gaussian measure
(1.1)
where is the standard -dimensional Lebesgue measure. Suppose is a random variable obeying the Lipschitz bounds
(1.2)
Then if is a median for there is a constant so that
(1.3)
and
(1.4)
Proof.
Let the set be given by
Since obeys the Lipschitz estimate
Then the Gaussian isoperimetric inequality (see [16] and [4]) tells us that
for any half-plane with measure . We calculate with the half-plane . So
That is,
so
that is (1.3). Equation (1.4) follows from a similar argument.
∎
We are now in a position to explore the relationship between and . We model our arguments on those of Burq-Lebeau [6] (Theorems 4 and 5) when . We say that is admissible if:
(1)
and strictly increasing.
(2)
.
(3)
obeys a generalised Minkowski inequality (with respect to the joint Gaussian measure)
(1.5)
Notice that clearly satisfies all requirements. There are a number of different conditions that could be placed on to ensure (3). For example assuming (in addition to (1) and (2)) that both the functions and are convex is sufficient (see [15] for this and other possible conditions).
Theorem 1.2.
Suppose that is admissible, for some . Let where each is a Gaussian random variable with . Now if is a non-negative, random variable that obeys the Lipschitz bound
with for some then
In particular if (for small enough )
then there are so that
(1.6)
Remark.
The parameter controls the Lipschitz bound. Later (see 2.7 and 3.9) for specific random variables we will determine suitable .
Proof.
Let be a median for the random variable . We will proceed by relating both and to . Consider
(1.7)
Note that by using the same argument that yields the reverse triangle inequality from the triangle inequality we have that
So
If we then get
On the other hand
Now since we have that
Therefore
∎
Since this results ensures that the expectation and median are very close together we can also state the concentration of measure results in terms of the expectation.
The corollary follows the from concentration of measure for Gaussian variables.
∎
2. The X-ray transform of
In this section we study the random variable obtained by taking a X-ray transform of along a line segment . That is we define the random variable
(2.1)
where is the unit line segment, properly contained in , from in direction . We will show that this random variable is equidistributed both in the sense that for fixed , is equidistributed but also in the stronger sense that the probability that fails equidistribution for any is exponentially small. Since is of unit length a uniform equidistribution statement for takes the following form. For every ,
and
here as but to ensure exponential decay. In Theorem 2.2 we will obtain explicit representations for valid , and an explicit decay for the error term in the expectation.
Theorem 2.1.
Suppose that for every the random variable is given by (2.1). Then for
(2.2)
for
(2.3)
and for ,
(2.4)
Proof.
We will use the results of Section 1 to compute by computing the . In this case we are using . Now
and since the Gaussian variables are mean and ,
So we need only control the Lipschitz norm associated with .
If we can estimate in terms of the almost orthogonality of the induced by the order spacing will yield a suitable Lipschitz constant. Since is a linear combination of plane waves with frequencies close to it makes sense to consider as an approximate solution (or quasimode) to
In particular is a quasimode of order , that is
This characterisation allows us to use results on the growth of quasimodes on submanifolds [17]. Note that, while we restrict our attention to for exact computation, in cases (such as this) where we are only concerned with bounds we may instead work with where is smooth, on and is zero outside . The function remains a quasimode of order for any . Similarly if we are only concerned with upper estimates we may extend the line segment so that the ends lie outside of the support of .
In the case where we can immediately use the submanifold estimates of [17] (Theorem 1.7). We can retrieve estimates for other values of by scaling. Let
and . Then
Since is an quasimode of then is an quasimode of . So applying [17] at the semiclassical scale we obtain
(2.5)
When
and so
(2.6)
in dimensions other than three (it holds with a log loss in dimension three). When the actually lie on the sphere we can use [7] to remove the log loss.
Therefore the Lipschitz constant for is (with the log loss for ). Recall that grows as . So ,
(2.7)
and again with the same bounds hold with a log loss in dimension three. Therefore using Theorem 1.2 we have that for fixed and
for
for ,
∎
Theorem 2.2.
Let be the random variable as given in Theorem 2.1 and be given by (2.7). For , as and
define the the exception set by
Let be the joint Gaussian measure given by (1.1). Then the following estimate on the size of holds.
(2.8)
Proof.
We use similar argument to that used in [12] and [8] to obtain uniform equidistribution statements on small balls. In those works uniform equidistribution results for a random variable depending on a parameter are obtained via a three step process.
(1)
A finite grid of is produced so that is approximated sufficiently by . This reduces the problem about the size of to the size of a finite union of sets (where is contained in the union).
(2)
The size of the exception sets are estimated at each grid-point. Usually this is achieved via a concentration of measure argument so that the decay is exponential in .
(3)
The number of grid-points is estimated. Usually this number will grow only polynomially in . Since the growth rate of the grid-points is overwhelmed by the exponential decay in each term a good union bound can be obtained.
In this case we are considering as a random variable depending on the parameter . Therefore we will form a grid of and approximate by . We have then reduced the problem about the size of to the size of a finite union
We find that the number of grid points necessary to obtain a suitable estimate does indeed grow polynomially. On the other hand concentration of measure arguments will allow us to prove exponential decay on the size of each of the , overwhelming the growth from the number of gridpoints. Therefore the union bound decays exponentially.
Suppose that , then there is some for which either
We approximate by via a Taylor series.
Writing parametrically we see that
So applying Cauchy Schwartz
Since is also a quasimode of we have that
where is given by (2.5). Therefore in dimension if we place our grid points with spacing we can always write (for some )
Similarly for we set the spacing at to have
Recall that . So (at least for small enough ) in both cases there is a grid point where
For , we already saw that so we can now apply Corollary 1.3 to obtain
(2.9)
∎
Putting together (2.9) with the results of Theorem 2.1, the probability that there there is any line over which equidistribution fails is so small that we would not expect to be able to see strong scars in numerical simulations. In the next section we consider the question of phase-space equidistribution. We obtain two distinct results. The first says that when we are far away from the Planck-scale equidistribution holds in the uniform sense. The more delicate result tells us that at the Planck scale phase-space equidistribution fails with high probability.
3. Phase-space concentrations
We now turn our attention to the question of equidistribtion in phase space. That is if is a semiclassical localiser, is equidistributed? We say that is a semiclassical localiser if
where is compactly supported in . The support of may depend on but is constrained by the uncertainty principle,
For the purposes of this paper we will be assuming that is supported near a point . We will introduce two parameters and to measure how near.
(1)
The scale controls the level of angular frequency localisation. The radial frequency is localised so that . We can ignore contributions from frequencies since is a sum of plane waves with frequencies in , .
(2)
The parameter controls how close/far we are from Planck scale. When we are exactly at Planck scale and therefore at the limits of the uncertainty principle. Large (growing in ) correspond to more relaxed localisation.
Let . Because of the curvature we obtain linearisation when so we are interested in the cases . Figure 2 shows the support of in both configuration and Fourier space.
Figure 2. The configuration an Fourier supports of the symbol .
Define
where
(3.1)
The prefactor is chosen so that
The angular cut off will ensure that is of order one. Note that
So if the random wave is assumed to equidistribute down to small scale (in terms of semiclassical defect measures) then this normalisation ensures that
Consider the random variable given by
(3.2)
If we can show that
we establish that, apart from a set whose measure decays to zero in , equidistribution fails at small scales. We will show this is true for small , that is . We will then see that for large , that in fact uniform equidistribution results hold. So it is only at Planck scale that we will see fluctuations.
As in Burq-Lebeau [6] we will first control by for . To compute we again use Theorem 1.2. In this case we set and compute . To facilitate that calculation we will assume that is an even integer (to allow us to expand the expression for ).
Proposition 3.1.
Suppose is given by (3.2). Then for (for some ), there are constants so that,
(3.3)
Proof.
Since is compact, the first half of (3.3) follows from Hölder. To get the second half we will in fact control the norm of in terms of its norm. The requisite control comes directly from the regularity properties of the symbol . The semiclassical Sobolev estimates tell us that if is semiclassically localised,
for some with compact support then
Saying that is semiclassically localised is the same as saying that is localised in space and that (up to Schwartz type error) its Fourier transform is localised in the ball of radius . Notice that
Therefore indeed outside the ball of radius the Fourier transform of decays in a Schwartz fashion and we can apply the semiclassical Sobolev estimates. In this case and if we set we obtain
∎
Theorem 3.2.
Let be given by (3.2). For and small enough (but independent of ) there exist constants so that
(3.4)
Proof.
From the results of Proposition 3.1 it is enough to find upper and lower bounds for for . Therefore let be an even integer obeying .
Let , we will compute
and apply the measure concentration results of Theorem 1.2.
Since we have assumed is even we can write
Since
is quadratic in the column vector . We will, similar to [8], scale and perform transformations so that can be written as where is a diagonal matrix and is a standard -dimensional Gaussian variable (with variance one).
Let we can write
as
where is the matrix with entries
(3.5)
with
If we diagonalise to write where is unitary and is diagonal with entries we can write
where . Since is unitary is also an -dimensional Gaussian random variable with variance one.
Recall that
so must have only non-negative eigenvalues. Let be a large integer (eventually we will want so that ). Then
Let and let be the number of distinct in .
Then
with . Since the are independent Gaussian variables we can compute
where the are the number of distinct occurrences in . Obviously . So
We obtain uniform upper and lower bounds for which, alongside the compact nature of carry over to give upper and lower bounds on .
Since all the eigenvalues are non-negative we can obtain lower bounds by obtain a lower bound for any fixed . We claim that if (the Planck scale case) then there is some so that
(3.6)
In which case the term associated with has the lower bound
To justify this claim we (in Lemma 4.1) compute and compare and . For we find that,
and therefore we can conclude (3.6). On the other hand
for any combination of so
In Lemma 4.1 we see that there are constants and such that
so if we have that
Now all we need is the concentration of measure that allows us to relate to . In Lemma 4.2 we find that for and
We will follow much the same process as we did in Theorems 2.1 and 2.2 where we controlled the behaviour of . That is we will set and use Theorem 1.2 to compute in terms of . Note that
Since the Gaussian variables have mean zero all non-diagonal terms fall out when we compute expectation so we only need compute
Since we can write
where has the same support an normalisation properties as . The contribution from the top terms is
Rescaling and translating/rotating so that and we see that the normalisation on ensures that this term is one. Since has the same normalisation properties as we can conclude that
Therefore
Now we can use the measure concentration to find . In Lemma 4.2 we find that if
Note that the cut offs we used to define have power (in ) type regularity. So by the same arguments that we used in the proof of uniform equidistribution of we can find a polynomial grid so that failure to equidistribute at some point implies a failure to equidistribute at a grid point . Then applying the concentration of measure at these points we obtain
Since the number of grid points only growth polynomially in (compared to the exponential decay of measure) we obtain (3.8).
∎
4. Technical Lemmata
This section is devoted to the proofs of the two technical Lemmata used a number of times throughout the paper.
The first lemma controls the growth of and dependent on (the parameter that controls how far we are from the Planck scale).
Lemma 4.1.
Suppose is given by (3.5). Then there exist constants so that
Immediately this tells us that there can only be of each where is nonzero. Conversely there is a smaller constant so that there are more than points for which
If we have
Therefore there are constants and so that
Since there are of order that appear in the sum we obtain constants (and here we abuse notation by allowing constants to vary line by line)
Now let’s turn our attention to . The lower bounds on the diagonal terms only give that
Therefore we need to use the off diagonal terms. What matters here is, how far apart do and have to be for
to be small? The additional smallness (apart from that which comes entirely from the support) is due to the non-stationary phase integral. If is large enough so that the oscillation of overwhelms the regularity of the contribution from this pair will be small. By rotations and translation it is enough to consider the case when and . In that case is supported in a tube (the long direction lies along the axis. First let’s consider the Planck scale case (that is ). Let . In this coordinate system if
(4.3)
the factor
does not complete a full oscillation over the support of . Therefore in these cases there are constants so that
Suppose that is one of the points such that
and similarly is one of the points such that
that is both and lie in the intersection between the cone with angle from the axis and the annulus (see Figure 3).
Figure 3. If and are too close to each other the oscillations of are not enough to overwhelm the regularity of the symbol
Therefore indeed, since , satisfies the conditions 4.3. So there are at least points for which
which yields
In this case the upper bound follows directly from the maximum number of so the integrand of is nonzero.
Now consider what happens as we move away from Planck scale . In this case the factor does not significantly oscillate if
(4.4)
but otherwise we are able to obtain some extra decay. Suppose first that
then integrating by parts in we find that
where has the same support and regularity properties as . Repeated applications of this argument show that for any
(4.5)
If the same argument using integration by parts in variables gives (4.5). By picking large enough we can ensure that the major contribution to the sum comes when (4.4) are satisfied. For any fixed there are ) suitable . Therefore we arrive at the estimate that
∎
The second lemma obtains Lipschitz bounds on . These are key to using the measure concentration arguments.
Lemma 4.2.
Suppose the following Lipschitz bound holds,
(4.6)
Proof.
Let
In this notation
So if we can find estimates for and we can interpolate all the others. First let’s see the case.
where are as in the proof of Lemma 4.1. Applying Cauchy-Schwartz
In the proof of Lemma 4.1 we have already estimated the sum by
So this leads to a Lipschitz bound
Now for the case. In that case
To get the last line we have integrated the variables first, integrating out the cut-off functions. Therefore since the are spaced at order (and are therefore almost orthogonal) we have
Interpolating with
which establishes the Lipschitz bounds.
∎
Acknowledgments
The author would like to thank Alex Barnett and Xiaolong Han for many interesting discussions on the small-scale structure of random waves. Particular thanks to Alex Barnett for allowing the reproduction of his numerical studies in this paper.
References
[1]
Robert Adler, Jonathan Taylor, Shmuel Weinberger, and Steve Zelditch.
Topological complexity of random sets, workshop report.
Technical report, American Institute of Mathematics, 2009.
[3]
M. V. Berry.
Regular and irregular semiclassical wavefunctions.
J. Phys. A, 10(12):2083–2091, 1977.
[4]
Christer Borell.
The Brunn-Minkowski inequality in Gauss space.
Invent. Math., 30(2):207–216, 1975.
[5]
N. Burq, P. Gérard, and N. Tzvetkov.
Restrictions of the Laplace-Beltrami eigenfunctions to
submanifolds.
Duke Math. J., 138(3):445–486, 2007.
[6]
Nicolas Burq and Gilles Lebeau.
Injections de Sobolev probabilistes et applications.
Ann. Sci. Éc. Norm. Supér. (4), 46(6):917–962, 2013.
[7]
Xuehua Chen and Christopher D. Sogge.
A few endpoint geodesic restriction estimates for eigenfunctions.
Comm. Math. Phys., 329(2):435–459, 2014.
[8]
Matthew de Courcy-Ireland.
Shrinking scale equidistribution for monochromatic random waves on
compact manifolds.
International Mathematics Research Notices, 2020.
[9]
Semyon Dyatlov and Maciej Zworski.
Quantum ergodicity for restrictions to hypersurfaces.
Nonlinearity, 26(1):35–52, 2013.
[10]
Suresh Eswarathasan.
Expected values of eigenfunction periods.
J. Geom. Anal., 26(1):360–377, 2016.
[11]
Xiaolong Han.
Small scale equidistribution of random eigenbases.
Comm. Math. Phys., 349(1):425–440, 2017.
[12]
Xiaolong Han and Melissa Tacy.
Equidistribution of random waves on small balls.
Comm. Math. Phys., 376(3):2351–2377, 2020.
[13]
Rui Hu.
norm estimates of eigenfunctions restricted to submanifolds.
Forum Math., 21(6):1021–1052, 2009.
[14]
Kenneth Maples.
Quantum unique ergodicity for random bases of spectral projections.
Math. Res. Lett., 20(6):1115–1124, 2013.
[15]
H. P. Mulholland.
On generalizations of Minkowski’s inequality in the form of a
triangle inequality.
Proc. London Math. Soc. (2), 51:294–307, 1950.
[16]
V. N. Sudakov and B. S. Cirel’son.
Extremal properties of half-spaces for spherically invariant
measures.
Zap. Naucn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI),
41:14–24, 165, 1974.
Problems in the theory of probability distributions, II.
[17]
Melissa Tacy.
Semiclassical estimates of quasimodes on submanifolds.
Comm. Partial Differential Equations, 35(8):1538–1562, 2010.
[18]
John A. Toth and Steve Zelditch.
Quantum ergodic restriction theorems: manifolds without boundary.
Geom. Funct. Anal., 23(2):715–775, 2013.
[19]
Jeffrey M. VanderKam.
norms and quantum ergodicity on the sphere.
Internat. Math. Res. Notices, (7):329–347, 1997.
[20]
Steve Zelditch.
A random matrix model for quantum mixing.
Internat. Math. Res. Notices, (3):115–137, 1996.
[21]
Steve Zelditch.
Real and complex zeros of Riemannian random waves.
In Spectral analysis in geometry and number theory, volume 484
of Contemp. Math., pages 321–342. Amer. Math. Soc., Providence, RI,
2009.
[22]
Steve Zelditch.
Quantum ergodicity of random orthonormal bases of spaces of high
dimension.
Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.,
372(2007):20120511, 16, 2014.
[23]
Steve Zelditch.
Mathematics of quantum chaos in 2019.
Notices Amer. Math. Soc., 66(9):1412–1422, 2019.