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

\MakePerPage

footnote

Filament structure of random waves

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, uu, in n\mathbb{R}^{n}. Numerical evidence suggests that such objects display a fine scale filament structure. We show that the X-ray of |u|2|u|^{2} along any line segment is uniformly equidistributed. So any limiting behaviour must be weaker than L2L^{2} 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 uu of the following form:

u=ξjΛcje1hx,ξju=\sum_{\xi_{j}\in\Lambda}c_{j}e^{\frac{1}{h}\langle x,\xi_{j}\rangle}

where Λn\Lambda\subset\mathbb{R}^{n} is a set of equally spaced momenta. Exact choices for Λ\Lambda vary somewhat between papers in this field but Λ\Lambda is usually a neighbourhood of the unit sphere 𝕊n1\mathbb{S}^{n-1}. The randomness is injected through the coefficients cjc_{j} which are chosen according to a probability distribution. Such functions uu 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 cjc_{j}. For example we may choose to look at cases where each cjc_{j} 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 2\ell^{2} norm of c=[c1,,c|Λ|]Tc=[c_{1},\dots,c_{|\Lambda|}]^{T} the coefficients may be chosen from a uniform probability density on the high dimensional sphere 𝕊|Λ|1\mathbb{S}^{|\Lambda|-1} (see for example [20],[6],[14],[22] and [11]).

In this paper we treat a model where each cjc_{j} is a (independent, identically distributed) Gaussian random variable and Λ=Λβ\Lambda=\Lambda_{\beta} is a set of hh-separated momenta drawn from 𝕊n1×[1hβ,1+hβ]\mathbb{S}^{n-1}\times[1-h^{\beta},1+h^{\beta}] for 0β10\leq\beta\leq 1. This range of β\beta 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 β\beta. It is convenient to choose Λβ\Lambda_{\beta} 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 hh (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 uu inside the ball of radius one about zero, B1(0)B_{1}(0). Then the variance for the Gaussian random variables is determined by adopting the convention that

𝔼[uL2(B1(0))2]=Vol(B1(0)).\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|^{2}_{L^{2}(B_{1}(0))}\right]=\mathrm{Vol}(B_{1}(0)).

This then means that the variance σ2=1N\sigma^{2}=\frac{1}{N} where N=|Λ|N=|\Lambda|.

The key questions of quantum chaos (see for example the survey [23]) are concerned with limits of Puh,uh\langle Pu_{h},u_{h}\rangle where PP is a semiclassical pseudodifferential operator. A sequence of uhu_{h} on a compact manifold MM with h0h\to 0 is said to equidistribute if

Puh,uh1Vol(SM)SMp𝑑L\langle Pu_{h},u_{h}\rangle\to\frac{1}{\mathrm{Vol}(S^{\star}M)}\int_{S^{\star}M}p\,d{L}

where SMS^{\star}M is the unit co-tangent bundle equipped with Liouville measure and pp is the principal symbol of PP. This leads to a natural question, are random waves equidistributed?

It is reasonably easy to ascertain that if p(x,ξ)p(x,\xi) (the symbol of PP) is independent of hh then random waves are indeed equidistributed. A more subtle question pertains to the small-scale structure. If for example p(x,ξ;h)p(x,\xi;h) is a symbol that is zero off a small (shrinking in hh) region of space in TMT^{\star}M does equidistribution (in an almost sure sense) still hold? If not does a weaker form of equidistribution hold where Pu,u\langle Pu,u\rangle is only proportional to the phase-space volume of p(x,ξ;h)?p(x,\xi;h)?

Refer to caption
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 γ\gamma where uL2\left|\!\left|{u}\right|\!\right|_{L^{2}} 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 γ(x,ξ)\gamma_{(x,\xi)} a unit length line segment, properly contained in B1(0)B_{1}(0), from xx in direction ξ\xi the random variable uL2(γ(x,ξ))=(γ(x,ξ)|u|2𝑑lγ)1/2\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}=\left(\int_{\gamma_{(x,\xi)}}|u|^{2}dl_{\gamma}\right)^{1/2} obeys

𝔼[uL2(γ(x,ξ))]=1+o(1)h0\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}\right]=1+o(1)\quad h\to 0

and that there is a κ>0\kappa>0 (details in Theorem 2.2) so that

Pr{c:(x,ξ),|uL2(γ(x,ξ))1|>m(h)}\displaystyle Pr\left\{c:\exists(x,\xi),\left|\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}-1\right|>m(h)\right\} exp(CN2κm2(h))\displaystyle\leq\exp\left(-CN^{2\kappa}m^{2}(h)\right) (0.1)
m(h)>Nκ+ϵ,m(h)0.\displaystyle m(h)>N^{-\kappa+\epsilon},m(h)\to 0.

In terms of what we might expect to see in a numerical simulation the uniformity (in (x,ξ)(x,\xi)) of the tail bound (0.1) is crucial. If for example we were only able to show that for fixed (x,ξ)(x,\xi)

Pr{c:|uL2(γ(x,ξ))1|>m(h)}exp(CN2κm2(h))m(h)>Nκ+ϵ,m(h)0Pr\left\{c:\left|\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}-1\right|>m(h)\right\}\leq\exp\left(-CN^{2\kappa}m^{2}(h)\right)\quad m(h)>N^{-\kappa+\epsilon},m(h)\to 0

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 (x,ξ)(x,\xi) for which uL2(γ(x,ξ))\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma(x,\xi))} 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 uL2(γ(x,ξ))\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma(x,\xi))}.

Theorem 0.1.

Let Fβ(x,ξ)F_{\beta}(x,\xi) be the random variable given by

Fβ(x,ξ)=Fβ(x,ξ;c)=uL2(γ(x,ξ))=λjΛβcjeih,ξjL2(γ(x,ξ))F_{\beta}(x,\xi)=F_{\beta}(x,\xi;c)=\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma(x,\xi))}=\left|\!\left|{\sum_{\lambda_{j}\in\Lambda_{\beta}}c_{j}e^{\frac{i}{h}\langle\cdot,\xi_{j}\rangle}}\right|\!\right|_{L^{2}(\gamma(x,\xi))}

then

𝔼[Fβ(x,ξ)]=1+o(1)h0.\mathbb{E}[F_{\beta}(x,\xi)]=1+o(1)\quad h\to 0.

Further there is a κ(n)\kappa(n) (explicitly given by in Section 2 by (2.7)) such that if m:++m:\mathbb{R}^{+}\to\mathbb{R}^{+}, m(h)0m(h)\to 0 as h0h\to 0, m(h)Nκ(n)+ϵm(h)\geq{}N^{-\kappa(n)+\epsilon} and the exception set S(m)S(m) is given by

S(m)={c:(x,ξ) so that |Fβ(x,ξ;c)1|m(h)}S(m)=\{c:\exists(x,\xi)\text{ so that }|F_{\beta}(x,\xi;c)-1|\geq{}m(h)\}

then

Pr(S(m))exp(N2κ(n)m2(h)).Pr(S(m))\leq\exp\left(-N^{2\kappa(n)}m^{2}(h)\right). (0.2)
Remark.

We will see, in Section 2, that in fact

𝔼[Fβ(x,ξ)2]=1\mathbb{E}[F_{\beta}(x,\xi)^{2}]=1

independent of hh. The error is introduced by approximating 𝔼[Fβ(x,ξ)]\mathbb{E}[F_{\beta}(x,\xi)] by (𝔼[Fβ(x,ξ)2])1/2\left(\mathbb{E}[F_{\beta}(x,\xi)^{2}]\right)^{1/2}. The validity of the approximation follows from a concentration of measure argument which in turn explicitly controls the quantity.

|𝔼[Fβ(x,ξ)](𝔼[Fβ(x,ξ)2])1/2|.\left|\mathbb{E}[F_{\beta}(x,\xi)]-\left(\mathbb{E}[F_{\beta}(x,\xi)^{2}]\right)^{1/2}\right|.

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 β=1\beta=1 the requirement that ξjΛβ\xi_{j}\in\Lambda_{\beta} ensures that uu 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,

uL2(γ(x,ξ))\displaystyle\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma(x,\xi))} h14uL2(B2(0))\displaystyle\lesssim h^{-\frac{1}{4}}\left|\!\left|{u}\right|\!\right|_{L^{2}(B_{2}(0))}
uL2(𝒞)\displaystyle\left|\!\left|{u}\right|\!\right|_{L^{2}(\mathcal{C})} h16uL2(B2(0))𝒞 geodesically curved.\displaystyle\lesssim h^{-\frac{1}{6}}\left|\!\left|{u}\right|\!\right|_{L^{2}(B_{2}(0))}\quad\text{$\mathcal{C}$ geodesically curved}.

The results of Theorem 0.1 should be interpreted as saying that choices of coefficients c=[c1,,cN]c=[c_{1},\dots,c_{N}] that saturate the general bounds are highly unusual.

In the setting of (M,g)(M,g) 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 ϕj\phi_{j},

ϕjL2(γ)L(γ)\left|\!\left|{\phi_{j}}\right|\!\right|_{L^{2}(\gamma)}\to L(\gamma)

where L(γ)L(\gamma) is the length of the curve γ\gamma. 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 Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) which measures the phase-space concentration of uu near the point (x,ξ)n×𝕊n1(x,\xi)\in\mathbb{R}^{n}\times\mathbb{S}^{n-1} given by

Gα,β,μ(x,ξ)=Gα,β,μ(x,ξ;c)=P(x,ξ)α,μu()L2.G_{\alpha,\beta,\mu}(x,\xi)=G_{\alpha,\beta,\mu}(x,\xi;c)=\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}}. (0.3)

Here P(x,ξ)α,μP^{\alpha,\mu}_{(x,\xi)} is a semiclassical pseudodifferential operator that localises uu in phase space near (x,ξ)(x,\xi). The parameter α\alpha controls the shape of the localisation, the parameter μ\mu controls the distance from Planck scale. When μ=1\mu=1 the operator P(x,ξ)α,μP^{\alpha,\mu}_{(x,\xi)} localises uu up to the limits allowed by the uncertainty principle. As μ\mu increases the localisation becomes more relaxed. Normalisations are chosen so that if uu is phase-space equidistributed |Gα,β,μ(x,ξ)|=1+O(μ1)|G_{\alpha,\beta,\mu}(x,\xi)|=1+O(\mu^{-1}). The details of symbol of Pα,μP^{\alpha,\mu} are described in Section 3. We discover that once μ\mu is large, that is μhϵ)\mu\approx h^{-\epsilon}), we again obtain a uniform equidistribution result similar to that for the X-ray transform of |u|2|u|^{2}. However when μC\mu\leq C we not only fail to obtain equidistribution but are able to show that with high probability there are logarithmic enhancements near some points (x,ξ)(x,\xi). The intermediate scales CμhϵC\leq\mu\leq h^{-\epsilon} remain something of a mystery and will likely require very delicate analysis to fully resolve.

Theorem 0.2.

Let Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) be given by (0.3). For μC\mu\leq C and small enough (but independent of hh) there exist constants C1,C2C_{1},C_{2} so that

C1log(1/h)𝔼[Gα,β,μL(B1(0)×𝕊n1)]C2log(1/h).C_{1}\sqrt{\log(1/h)}\leq\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}(B_{1}(0)\times\mathbb{S}^{n-1})}\right]\leq C_{2}\sqrt{\log(1/h)}. (0.4)

On the other hand if μhϵ\mu\geq h^{-\epsilon}, then for (x,ξ)B1(0)×𝕊n1(x,\xi)\in B_{1}(0)\times\mathbb{S}^{n-1},

𝔼[Gα,β,μ(x,ξ)]=1+O(hϵ2).\mathbb{E}\left[G_{\alpha,\beta,\mu}(x,\xi)\right]=1+O\left(h^{\frac{\epsilon}{2}}\right). (0.5)

Further if δ>0\delta>0 and m(h)max(hϵδ,h2ϵ(n+1)4(nβ)δ)m(h)\geq{}\max\left(h^{\epsilon-\delta},h^{\frac{2\epsilon(n+1)}{4(n-\beta)}-\delta}\right),

Pr{c:(x,ξ), so that |Gα,β,μ(x,ξ;c)1|>m(h)}exp(N2ϵ(n+1)4(nβ)m2(h)).Pr\{c:\exists(x,\xi),\text{ so that }\left|G_{\alpha,\beta,\mu}(x,\xi;c)-1\right|>m(h)\}\leq\exp(-N^{\frac{2\epsilon(n+1)}{4(n-\beta)}}m^{2}(h)). (0.6)
Remark.

The proof of Theorem 0.2 depends on estimating Gα,β,μ(,)L\left|\!\left|{G_{\alpha,\beta,\mu}(\cdot,\cdot)}\right|\!\right|_{L^{\infty}} by an LphL^{p_{h}} norm where php_{h} is finite but growing in hh. Then Gα,β,μ(,)phph\left|\!\left|{G_{\alpha,\beta,\mu}(\cdot,\cdot)}\right|\!\right|^{p_{h}}_{p_{h}} is directly computed. It is the lower bound on this norm that requires μC\mu\leq C (see equation (3.6)). The key technical dependancies for μ\mu 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 uu 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

C1log(1/h)𝔼[uL]C2log(1/h).C_{1}\sqrt{\log(1/h)}\leq\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{\infty}}\right]\leq C_{2}\sqrt{\log(1/h)}. (0.7)

Indeed when α=0\alpha=0 the operator P(x,ξ)α,μP^{\alpha,\mu}_{(x,\xi)} localises uu to a small ball of radius hh around xx. Since uu is constituted of waves of frequency of size 1/h1/h the value of |u(x)||u(x)| cannot vary much on balls of radius hh. Taking into account the normalisation we therefore get that

P(x,ξ)0,μL22|u(x)|2.\left|\!\left|{P^{0,\mu}_{(x,\xi)}}\right|\!\right|_{L^{2}}^{2}\approx|u(x)|^{2}.

Therefore the results of this paper for α=0\alpha=0 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 γ(x,ξ)\gamma(x,\xi) cannot be enough to register in the L2L^{2} norm. However it is quite possible to have large number of logarithmically bright points aligned along a ray without the L2L^{2} 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 L2L^{2} 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 ϕ\phi”. The idea is that if ϕ:[0,)[0,)\phi:[0,\infty)\to[0,\infty) is invertible, we will use measure concentration to find (see Section 1) sufficient conditions so that

C1ϕ1(𝔼[ϕ(X)])𝔼[X]C2ϕ1(𝔼[ϕ(X)])C1,C2+.C_{1}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)\leq\mathbb{E}\left[X\right]\leq C_{2}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)\quad C_{1},C_{2}\in\mathbb{R}^{+}.

Here XX is a random variable that takes values in [0,)[0,\infty). The rational for seeking this kind of estimate is to be able to chose ϕ\phi so that 𝔼[ϕ(X)]\mathbb{E}[\phi(X)] is easy (or at least tractable) to compute. Just by assuming ϕ\phi 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 |ϕ(𝔼[X])𝔼[ϕ(X)]|\left|\phi\left(\mathbb{E}[X]\right)-\mathbb{E}\left[\phi(X)\right]\right| 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 XX and ϕ\phi) so that

C1ϕ1(𝔼[ϕ(X)])𝔼[X]C2ϕ1(𝔼[ϕ(X)])C1,C2+C_{1}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)\leq\mathbb{E}\left[X\right]\leq C_{2}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)\quad C_{1},C_{2}\in\mathbb{R}^{+}

or

𝔼[X]=ϕ1(𝔼[ϕ(X)])+Er\mathbb{E}\left[X\right]=\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)+Er

where ErEr is small compared to ϕ1(𝔼[ϕ(X)])\phi^{-1}\left(\mathbb{E}[\phi(X)]\right). 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 c=[c1,,cN]Tc=[c_{1},\dots,c_{N}]^{T} be an NN-tuple of independent Gaussian variables each with variance σ2=N1\sigma^{2}=N^{-1}. Let μ\mu be the joint NN-dimensional Gaussian measure

μ=NN/2(2π)N/2eN|c|22λN\mu=\frac{N^{N/2}}{(2\pi)^{N/2}}e^{-\frac{N|c|^{2}}{2}}\lambda^{N} (1.1)

where λN\lambda^{N} is the standard NN-dimensional Lebesgue measure. Suppose G=G(c)G=G(c) is a random variable obeying the Lipschitz bounds

|G(c)G(d)|Lcd2|G(c)-G(d)|\leq L\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}} (1.2)

Then if MGM_{G} is a median for GG there is a constant CC so that

Pr{GMG+t}CeNt22L2Pr\{G\geq M_{G}+t\}\leq Ce^{-\frac{Nt^{2}}{2L^{2}}} (1.3)

and

Pr{GMGt}CeNt22L2.Pr\{G\leq M_{G}-t\}\leq Ce^{-\frac{Nt^{2}}{2L^{2}}}. (1.4)
Proof.

Let the set BNB\subset\mathbb{R}^{N} be given by

B={cN:G(c)MG}.B=\{c\in\mathbb{R}^{N}:G(c)\leq M_{G}\}.

Since GG obeys the Lipschitz estimate

|G(c)G(d)|Lcd2|G(c)-G(d)|\leq L\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}
Bt/L={cN:dist(c,B)<tL}{cNG(c)<MG+t}.B_{t/L}=\left\{c\in\mathbb{R}^{N}:\mathrm{dist}(c,B)<\frac{t}{L}\right\}\subset\{c\in\mathbb{R}^{N}\mid G(c)<M_{G}+t\}.

Then the Gaussian isoperimetric inequality (see [16] and [4]) tells us that

μ(Bt/L)μ(Ht/L)\mu(B_{t/L})\geq\mu(H_{t/L})

for any half-plane with measure μ(B)=μ(H)=12\mu(B)=\mu(H)=\frac{1}{2}. We calculate with the half-plane {yy1<0}\{y\mid y_{1}<0\}. So

μ(Ht/L)\displaystyle\mu(H_{t/L}) =N1/22πtLeNs22𝑑s\displaystyle=\frac{N^{1/2}}{\sqrt{2\pi}}\int_{-\infty}^{\frac{t}{L}}e^{-\frac{Ns^{2}}{2}}ds
=1N1/22πtLeNs22𝑑s\displaystyle=1-\frac{N^{1/2}}{\sqrt{2\pi}}\int_{\frac{t}{L}}^{\infty}e^{-\frac{Ns^{2}}{2}}ds
=1O(eNt22L2).\displaystyle=1-O\left(e^{-\frac{Nt^{2}}{2L^{2}}}\right).

That is,

μ{cN:G(c)<M+t}=1O(eNt22L2)\mu\{c\in\mathbb{R}^{N}:G(c)<M+t\}=1-O\left(e^{-\frac{Nt^{2}}{2L^{2}}}\right)

so

μ{cN:G(c)M+t}CeNt22L2\mu\{c\in\mathbb{R}^{N}:G(c)\geq M+t\}\leq Ce^{-\frac{Nt^{2}}{2L^{2}}}

that is (1.3). Equation (1.4) follows from a similar argument. ∎

We are now in a position to explore the relationship between ϕ1(𝔼[ϕ(X)])\phi^{-1}\left(\mathbb{E}[\phi(X)]\right) and 𝔼[X]\mathbb{E}[X]. We model our arguments on those of Burq-Lebeau [6] (Theorems 4 and 5) when ϕ(τ)=τp\phi(\tau)=\tau^{p}. We say that ϕ\phi is admissible if:

  1. (1)

    ϕC1\phi\in C^{1} and strictly increasing.

  2. (2)

    ϕ(0)=0\phi(0)=0.

  3. (3)

    ϕ(||)\phi(|\cdot|) obeys a generalised Minkowski inequality (with respect to μ\mu the joint Gaussian measure)

    ϕ1(ϕ(|X(c)+Y(c)|)𝑑μ)ϕ1(ϕ(|X(c)|)𝑑μ)+ϕ1(ϕ(|Y(c)|)𝑑μ).\phi^{-1}\left(\int\phi(|X(c)+Y(c)|)d\mu\right)\leq\phi^{-1}\left(\int\phi(|X(c)|)d\mu\right)+\phi^{-1}\left(\int\phi(|Y(c)|)d\mu\right). (1.5)

Notice that τp\tau^{p} clearly satisfies all requirements. There are a number of different conditions that could be placed on ϕ\phi to ensure (3). For example assuming (in addition to (1) and (2)) that both the functions ϕ(||)\phi(|\cdot|) and log|ϕ(eτ)|-\log|\phi^{\prime}(e^{\tau})| are convex is sufficient (see [15] for this and other possible conditions).

Theorem 1.2.

Suppose that ϕ\phi is admissible, |ϕ(τ)|C|τ|q1|\phi^{\prime}(\tau)|\leq C|\tau|^{q-1} for some q1q\geq{}1. Let c=[c1,,cN]Tc=[c_{1},\dots,c_{N}]^{T} where each cic_{i} is a Gaussian random variable with σ2=N1\sigma^{2}=N^{-1}. Now if X=X(c)X=X(c) is a non-negative, random variable that obeys the Lipschitz bound

|X(c)X(d)|Lcd2|X(c)-X(d)|\leq L\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}

with LN1/2κL\leq N^{1/2-\kappa} for some κ>0\kappa>0 then

𝔼[X]=ϕ1(𝔼[ϕ(X)])+O(ϕ1(CNκqΓ(q/2))).\mathbb{E}\left[X\right]=\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)+O\left(\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right)\right).

In particular if (for small enough ϵ\epsilon)

ϕ1(CNκqΓ(q/2))ϵϕ1(𝔼[ϕ(X)])\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right)\leq\epsilon\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)

then there are C1,C2+C_{1},C_{2}\in\mathbb{R}^{+} so that

C1ϕ1(𝔼[ϕ(X)])𝔼[X]C2ϕ1(𝔼[ϕ(X)]).C_{1}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)\leq\mathbb{E}\left[X\right]\leq C_{2}\phi^{-1}\left(\mathbb{E}[\phi(X)]\right). (1.6)
Remark.

The parameter κ\kappa controls the Lipschitz bound. Later (see 2.7 and 3.9) for specific random variables we will determine suitable κ\kappa.

Proof.

Let MXM_{X} be a median for the random variable XX. We will proceed by relating both ϕ1(𝔼[ϕ(X)])\phi^{-1}(\mathbb{E}[\phi(X)]) and 𝔼[X]\mathbb{E}[X] to MXM_{X}. Consider

ϕ(|ϕ1(𝔼[ϕ(X)])MX|)=ϕ(|ϕ1(ϕ(X)𝑑μ)ϕ1(ϕ(MX)𝑑μ)|).\phi\left(\left|\phi^{-1}(\mathbb{E}[\phi(X)])-M_{X}\right|\right)=\phi\left(\left|\phi^{-1}\left(\int\phi(X)d\mu\right)-\phi^{-1}\left(\int\phi(M_{X})d\mu\right)\right|\right). (1.7)

Note that by using the same argument that yields the reverse triangle inequality from the triangle inequality we have that

|ϕ1(ϕ(X)𝑑μ)ϕ1(ϕ(MX)𝑑μ)|ϕ1(ϕ(|XMX|)𝑑μ).\left|\phi^{-1}\left(\int\phi(X)d\mu\right)-\phi^{-1}\left(\int\phi(M_{X})d\mu\right)\right|\leq\phi^{-1}\left(\int\phi(|X-M_{X}|)d\mu\right).

So

ϕ(|ϕ1(𝔼[ϕ(X)])MX|)\displaystyle\phi\left(\left|\phi^{-1}(\mathbb{E}[\phi(X)])-M_{X}\right|\right) ϕ(|XMX|)𝑑μ\displaystyle\leq\int\phi(|X-M_{X}|)d\mu
=0ϕ(s)μ({|XMX|s})𝑑s\displaystyle=\int_{0}^{\infty}\phi^{\prime}(s)\mu(\{|X-M_{X}|\geq{}s\})ds
C0sq1eNs22L2𝑑s.\displaystyle\leq C\int_{0}^{\infty}s^{q-1}e^{-\frac{Ns^{2}}{2L^{2}}}ds.

If LN1/2κL\leq N^{1/2-\kappa} we then get

ϕ(|ϕ1(𝔼[ϕ(X)])MX|)C0sq1ecN2κs2𝑑s=CNκqΓ(q/2).\phi\left(\left|\phi^{-1}(\mathbb{E}[\phi(X)])-M_{X}\right|\right)\leq C\int_{0}^{\infty}s^{q-1}e^{-cN^{2\kappa}s^{2}}ds=\frac{C}{N^{\kappa q}}\Gamma(q/2).

On the other hand

|𝔼[X]MX|\displaystyle|\mathbb{E}[X]-M_{X}| |XMX|𝑑μ\displaystyle\leq\int|X-M_{X}|d\mu
=0μ(|XMX|s)𝑑s\displaystyle=\int_{0}^{\infty}\mu(|X-M_{X}|\geq{}s)ds
0ecN2κs2𝑑s\displaystyle\leq\int_{0}^{\infty}e^{-cN^{2\kappa}s^{2}}ds
CNκ0es2𝑑sCNκ.\displaystyle\leq\frac{C}{N^{\kappa}}\int_{0}^{\infty}e^{-s^{2}}ds\leq\frac{C}{N^{\kappa}}.

Now since CNκCNκqΓ(q/2)\frac{C}{N^{\kappa}}\leq\frac{C}{N^{\kappa q}}\Gamma(q/2) we have that

|E[X]MX|CNκqΓ(q/2).|E[X]-M_{X}|\leq\frac{C}{N^{\kappa q}}\Gamma(q/2).

Therefore

E[X]=ϕ1(E[ϕ(X)])+O(ϕ1(CNκqΓ(q/2))).E[X]=\phi^{-1}\left(E[\phi(X)]\right)+O\left(\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right)\right).

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.

Corollary 1.3.

Under the conditions of Theorem 1.2,

μ({c:|X(c)ϕ1(𝔼[ϕ(X(c))])|t})exp(CN2κt2)\mu\left(\{c:|X(c)-\phi^{-1}\left(\mathbb{E}[\phi(X(c))]\right)|\geq{}t\}\right)\leq\exp(-CN^{2\kappa}t^{2}) (1.8)

for all t2ϕ1(CNκqΓ(q/2))t\geq{}2\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right).

Proof.

We saw in the proof of Theorem 1.2 that

|ϕ1(𝔼[ϕ(X)])MX|ϕ1(CNκqΓ(q/2)).\left|\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)-M_{X}\right|\leq\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right).

So if

|Xϕ1(𝔼[ϕ(X)])|t|X-\phi^{-1}\left(\mathbb{E}[\phi(X)]\right)|\geq{}t

with tϕ1(CNκqΓ(q/2))t\geq{}\phi^{-1}\left(\frac{C}{N^{\kappa q}}\Gamma(q/2)\right),

|XM|t/2.|X-M|\geq{}t/2.

The corollary follows the from concentration of measure for Gaussian variables.

2. The X-ray transform of |u|2|u|^{2}

In this section we study the random variable obtained by taking a X-ray transform of |u|2|u|^{2} along a line segment γ\gamma. That is we define the random variable

Fβ(x,ξ)=Fβ(x,ξ;c)=(γ(x,ξ)|u|2𝑑lγ)1/2=uL2(γ(x,ξ))F_{\beta}(x,\xi)=F_{\beta}(x,\xi;c)=\left(\int_{\gamma_{(x,\xi)}}|u|^{2}dl_{\gamma}\right)^{1/2}=\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})} (2.1)

where γ(x,ξ)\gamma_{(x,\xi)} is the unit line segment, properly contained in B1(0)B_{1}(0), from xx in direction ξ\xi. We will show that this random variable is equidistributed both in the sense that for fixed (x,ξ)(x,\xi), Fβ(x,ξ)F_{\beta}(x,\xi) is equidistributed but also in the stronger sense that the probability that Fβ(x,ξ)F_{\beta}(x,\xi) fails equidistribution for any (x,ξ)(x,\xi) is exponentially small. Since γ(x,ξ)\gamma_{(x,\xi)} is of unit length a uniform equidistribution statement for FβF_{\beta} takes the following form. For every (x,ξ)(x,\xi),

𝔼[Fβ(x,ξ)]=1+o(1)\mathbb{E}[F_{\beta}(x,\xi)]=1+o(1)

and

Pr{c:(x,ξ) so that |Fβ(x,ξ;c)1|m(h)}Cexp(cN2κm2(h))κ>0Pr\{c:\exists(x,\xi)\text{ so that }|F_{\beta}(x,\xi;c)-1|\geq{}m(h)\}\leq C\exp\left(cN^{2\kappa}m^{2}(h)\right)\kappa>0

here m(h)0m(h)\to 0 as h0h\to 0 but m(h)>Nκ+ϵm(h)>N^{-\kappa+\epsilon} to ensure exponential decay. In Theorem 2.2 we will obtain explicit representations for valid κ\kappa, m(h)m(h) and an explicit decay for the error term in the expectation.

Theorem 2.1.

Suppose that for every (x,ξ)B1(0)×𝕊n1(x,\xi)\in B_{1}(0)\times\mathbb{S}^{n-1} the random variable Fβ(x,ξ)F_{\beta}(x,\xi) is given by (2.1). Then for n=2n=2

𝔼[Fβ(x,ξ)]=1+O(1N12β4),\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{1}{N^{\frac{1}{2}-\frac{\beta}{4}}}\right), (2.2)

for n=3n=3

𝔼[Fβ(x,ξ)]=1+O(log(N)N1/2)\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{\sqrt{\log(N)}}{N^{1/2}}\right) (2.3)

and for n3n\geq{}3,

𝔼[Fβ(x,ξ)]=1+O(1N1/2).\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{1}{N^{1/2}}\right). (2.4)
Proof.

We will use the results of Section 1 to compute 𝔼[uL2(γ(x,ξ))]\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}\right] by computing the 𝔼[uL2(γ(x,ξ))2]\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}^{2}\right]. In this case we are using ϕ(τ)=τ2\phi(\tau)=\tau^{2}. Now

𝔼[uL2(γ(x,ξ))2]=𝔼[j,lcjclγ(x,ξ)eihy,ξjξl𝑑y]\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}^{2}\right]=\mathbb{E}\left[\sum_{j,l}c_{j}c_{l}\int_{\gamma_{(x,\xi)}}e^{\frac{i}{h}\langle y,\xi_{j}-\xi_{l}\rangle}dy\right]

and since the Gaussian variables cjc_{j} are mean 0 and σ2=N1\sigma^{2}=N^{-1},

𝔼[uL2(γ(x,ξ))2]=1.\mathbb{E}\left[\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}^{2}\right]=1.

So we need only control the Lipschitz norm associated with Fβ(x,ξ)F_{\beta}(x,\xi).

|Fβ(x,ξ;c)Fβ(x,ξ;c)||u(c)L2(γ(x,ξ))udL2(γ(x,ξ))|u(c)u(d)L2(γ(x,ξ)).|F_{\beta}(x,\xi;c)-F_{\beta}(x,\xi;c)|\leq|\left|\!\left|{u(c)}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}-\left|\!\left|{u_{d}}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}|\leq\left|\!\left|{u(c)-u(d)}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}.

If we can estimate u(c)u(d)L2(γ(x,ξ))\left|\!\left|{u(c)-u(d)}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})} in terms of u(c)u(d)L2(B1(0))\left|\!\left|{u(c)-u(d)}\right|\!\right|_{L^{2}(B_{1}(0))} the almost orthogonality of the eihy,ξje^{\frac{i}{h}\langle y,\xi_{j}\rangle} induced by the order hh spacing will yield a suitable Lipschitz constant. Since uu is a linear combination of plane waves with frequencies close to 1h\frac{1}{h} it makes sense to consider uu as an approximate solution (or quasimode) to

(h2Δ1)u=0.(h^{2}\Delta-1)u=0.

In particular uu is a quasimode of order hβh^{\beta}, that is

(h2Δ1)uL2(B1(0))=OL2(hβ).\left|\!\left|{(h^{2}\Delta-1)u}\right|\!\right|_{L^{2}(B_{1}(0))}=O_{L^{2}}(h^{\beta}).

This characterisation allows us to use results on the LpL^{p} growth of quasimodes on submanifolds [17]. Note that, while we restrict our attention to 𝟙B1(0)u\mathds{1}_{B_{1}(0)}u for exact computation, in cases (such as this) where we are only concerned with bounds we may instead work with χ(x)u(x)\chi(x)u(x) where χ\chi is smooth, χ=1\chi=1 on B1(0)B_{1}(0) and is zero outside B2(0)B_{2}(0). The function χ(x)u(x)\chi(x)u(x) remains a quasimode of order hβh^{\beta} for any 0β10\leq\beta\leq 1. Similarly if we are only concerned with upper estimates we may extend the line segment γ(x,ξ)\gamma(x,\xi) so that the ends lie outside of the support of χ(x)u(x)\chi(x)u(x).

In the case where β=1\beta=1 we can immediately use the submanifold estimates of [17] (Theorem 1.7). We can retrieve estimates for other values of β\beta by scaling. Let

uβ=u(h1βx)u_{\beta}=u(h^{1-\beta}x)

and Er[u]=(h2Δ1)uEr[u]=(h^{2}\Delta-1)u. Then

(h2βΔ1)uβ=Er[u](h1βx).(h^{2\beta}\Delta-1)u_{\beta}=Er[u](h^{1-\beta}x).

Since uu is an OL2(hβ)O_{L^{2}}(h^{\beta}) quasimode of (h2Δ1)(h^{2}\Delta-1) then uβu_{\beta} is an OL2(hβ)O_{L^{2}}(h^{\beta}) quasimode of (h2βΔ1)(h^{2\beta}\Delta-1). So applying [17] at the semiclassical scale hβh^{\beta} we obtain

uβL2(γ(x,ξ))hβδ(n)uβL2\left|\!\left|{u_{\beta}}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}\lesssim h^{-\beta\delta(n)}\left|\!\left|{u_{\beta}}\right|\!\right|_{L^{2}}
δ(n)={14n=2n1212n>3.\delta(n)=\begin{cases}\frac{1}{4}&n=2\\ \frac{n-1}{2}-\frac{1}{2}&n>3.\end{cases} (2.5)

When n=3n=3

uβL2(γ(x,ξ))hβ2log(1/h)uβL2\left|\!\left|{u_{\beta}}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}\lesssim h^{-\frac{\beta}{2}}\sqrt{\log(1/h)}\left|\!\left|{u_{\beta}}\right|\!\right|_{L^{2}}

and so

uL2(γ(x,ξ))hβδ(n)+β12uL2\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma_{(x,\xi)})}\lesssim h^{-\beta\delta(n)+\frac{\beta-1}{2}}\left|\!\left|{u}\right|\!\right|_{L^{2}} (2.6)

in dimensions other than three (it holds with a log loss in dimension three). When the ξj\xi_{j} actually lie on the sphere we can use [7] to remove the log loss. Therefore the Lipschitz constant for Fβ(x,ξ)F_{\beta}(x,\xi) is hβδ(n)+β12h^{-\beta\delta(n)+\frac{\beta-1}{2}} (with the log loss for n=3n=3). Recall that NN grows as hβnh^{\beta-n}. So LN1/2κ(n)L\leq N^{1/2-\kappa(n)},

κ(n)={12β4n=212n>3\kappa(n)=\begin{cases}\frac{1}{2}-\frac{\beta}{4}&n=2\\ \frac{1}{2}&n>3\end{cases} (2.7)

and again with κ(3)=1/2\kappa(3)=1/2 the same bounds hold with a log loss in dimension three. Therefore using Theorem 1.2 we have that for fixed (x,ξ)(x,\xi) and n=2n=2

𝔼[Fβ(x,ξ)]=1+O(1N12β4),\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{1}{N^{\frac{1}{2}-\frac{\beta}{4}}}\right),

for n=3n=3

𝔼[Fβ(x,ξ)]=1+O(logNN1/2),\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{\sqrt{\log N}}{N^{1/2}}\right),

for n3n\geq{}3,

𝔼[Fβ(x,ξ)]=1+O(1N1/2).\mathbb{E}\left[F_{\beta}(x,\xi)\right]=1+O\left(\frac{1}{N^{1/2}}\right).

Theorem 2.2.

Let Fβ(x,ξ)F_{\beta}(x,\xi) be the random variable as given in Theorem 2.1 and κ(n)\kappa(n) be given by (2.7). For m:++m:\mathbb{R}^{+}\to\mathbb{R}^{+}, m(h)0m(h)\to 0 as h0h\to 0 and

m(h)Nκ(n)+ϵ={h(2β)(12β4)ϵ(nβ)n=2hnβ2ϵ(nβ)n3.m(h)\geq{}N^{-\kappa(n)+\epsilon}=\begin{cases}h^{\left(2-\beta\right)\left(\frac{1}{2}-\frac{\beta}{4}\right)-\epsilon(n-\beta)}&n=2\\ h^{\frac{n-\beta}{2}-\epsilon(n-\beta)}&n\geq{}3.\end{cases}

define the the exception set S(m)S(m) by

S(m)={c:(x,ξ) so that |Fβ(x,ξ;c)1|m(h)}.S(m)=\{c:\exists(x,\xi)\text{ so that }|F_{\beta}(x,\xi;c)-1|\geq{}m(h)\}.

Let μ\mu be the joint Gaussian measure given by (1.1). Then the following estimate on the size of S(m)S(m) holds.

Pr(S(m))=μ(S(m))exp(N2κ(n)m2(h)).Pr(S(m))=\mu(S(m))\leq\exp\left(-N^{2\kappa(n)}m^{2}(h)\right). (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 X=X(y;c)X=X(y;c) depending on a parameter yy are obtained via a three step process.

  1. (1)

    A finite grid of yνy^{\nu} is produced so that X(y;c)X(y;c) is approximated sufficiently by X(yν;c)X(y^{\nu};c). This reduces the problem about the size of S(m)S(m) to the size of a finite union of sets (where S(m)S(m) is contained in the union).

  2. (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 h1h^{-1}.

  3. (3)

    The number of grid-points is estimated. Usually this number will grow only polynomially in h1h^{-1}. 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 |Fβ(x,ξ;c)|2|F_{\beta}(x,\xi;c)|^{2} as a random variable depending on the parameter (x,ξ)(x,\xi). Therefore we will form a grid of (xν,ξν)(x^{\nu},\xi^{\nu}) and approximate |Fβ(x,ξ;c)|2|F_{\beta}(x,\xi;c)|^{2} by |Fβ(xν,ξν;c)|2|F_{\beta}(x^{\nu},\xi^{\nu};c)|^{2}. We have then reduced the problem about the size of S(m)S(m) to the size of a finite union

ν{c:|Fβ(xν,ξν;c)1|m(h)/2}.\bigcup_{\nu}\{c:|F_{\beta}(x^{\nu},\xi^{\nu};c)-1|\geq{}m(h)/2\}.

We find that the number of grid points (xν,ξν)(x^{\nu},\xi^{\nu}) 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 {c:|Fβ(xν,ξν;c)1|m(h)/2}\{c:|F_{\beta}(x^{\nu},\xi^{\nu};c)-1|\geq{}m(h)/2\}, overwhelming the growth from the number of gridpoints. Therefore the union bound decays exponentially.

Suppose that cS(m)c\in S(m), then there is some (x,ξ)(x,\xi) for which either

Fβ2(x,ξ;c)(1+m(h))2orFβ2(x,ξ;c)(1m(h))2.F^{2}_{\beta}(x,\xi;c)\geq{}(1+m(h))^{2}\quad\text{or}\quad F^{2}_{\beta}(x,\xi;c)\leq(1-m(h))^{2}.

We approximate Fβ2(x,ξ;c)F^{2}_{\beta}(x,\xi;c) by Fβ2(xν,ξν;c)F^{2}_{\beta}(x^{\nu},\xi^{\nu};c) via a Taylor series. Writing Fβ2(x,ξ;c)F^{2}_{\beta}(x,\xi;c) parametrically we see that

|xjFβ2(x,ξ;c)|201|uxj(x+sξ)||u(x+sξ)|𝑑s\left|\frac{\partial}{\partial x_{j}}F^{2}_{\beta}(x,\xi;c)\right|\leq 2\int_{0}^{1}|u_{x_{j}}(x+s\xi)||u(x+s\xi)|ds
|ξjFβ2(x,ξ;c)|201|uxj(x+sξ)||u(x+sξ)|𝑑s.\left|\frac{\partial}{\partial\xi_{j}}F^{2}_{\beta}(x,\xi;c)\right|\leq 2\int_{0}^{1}|u_{x_{j}}(x+s\xi)||u(x+s\xi)|ds.

So applying Cauchy Schwartz

|xjFβ2(x,ξ;c)|+|ξjFβ2(x,ξ;c)|CuxjL2(γ(x,ξ))uL2(γ(x,ξ)).\left|\frac{\partial}{\partial x_{j}}F^{2}_{\beta}(x,\xi;c)\right|+\left|\frac{\partial}{\partial\xi_{j}}F^{2}_{\beta}(x,\xi;c)\right|\leq C\left|\!\left|{u_{x_{j}}}\right|\!\right|_{L^{2}(\gamma(x,\xi))}\left|\!\left|{u}\right|\!\right|_{L^{2}(\gamma(x,\xi))}.

Since uxju_{x_{j}} is also a O(hβ)O(h^{\beta}) quasimode of (h2Δ1)(h^{2}\Delta-1) we have that

Fβ2(x,ξ;c)=Fβ2(xν,ξν;c)+O(h2βδ(n)|(x,ξ)(xν,ξν)|)F^{2}_{\beta}(x,\xi;c)=F^{2}_{\beta}(x^{\nu},\xi^{\nu};c)+O\left(h^{-2\beta\delta(n)}|(x,\xi)-(x^{\nu},\xi^{\nu})|\right)

where δ(n)\delta(n) is given by (2.5). Therefore in dimension n=2n=2 if we place our grid points with spacing hβ2+(2β)(12β4)h^{\frac{\beta}{2}+\left(2-\beta\right)\left(\frac{1}{2}-\frac{\beta}{4}\right)} we can always write (for some (xν,ξν)(x^{\nu},\xi^{\nu}))

Fβ2(x,ξ;c)=Fβ2(xν,ξν;c)+O(Nκ(n))=Fβ2(xν,ξν;c)+O(h(nβ)(12β4)).F^{2}_{\beta}(x,\xi;c)=F^{2}_{\beta}(x^{\nu},\xi^{\nu};c)+O(N^{-\kappa(n)})=F^{2}_{\beta}(x^{\nu},\xi^{\nu};c)+O\left(h^{\left(n-\beta\right)\left(\frac{1}{2}-\frac{\beta}{4}\right)}\right).

Similarly for n3n\geq{}3 we set the spacing at hn2+nβ2h^{n-2+\frac{n-\beta}{2}} to have

Fβ2(x,ξ;c)=Fβ2(xν,ξν;c)+O(Nκ(n))=Fβ2(xν,ξν;c)+O(hnβ2).F^{2}_{\beta}(x,\xi;c)=F^{2}_{\beta}(x^{\nu},\xi^{\nu};c)+O(N^{-\kappa(n)})=F^{2}_{\beta}(x^{\nu},\xi^{\nu};c)+O\left(h^{\frac{n-\beta}{2}}\right).

Recall that m(h)Nκ(n)+ϵ=h(nβ)(κ(n)ϵ)m(h)\geq{}N^{-\kappa(n)+\epsilon}=h^{(n-\beta)(\kappa(n)-\epsilon)}. So (at least for small enough hh) in both cases there is a grid point (xν,ξν)(x^{\nu},\xi^{\nu}) where

|Fβ(xν,ξν;c)1|m(h)/2.|F_{\beta}(x^{\nu},\xi^{\nu};c)-1|\geq{}m(h)/2.

For ϕ(τ)=τ2\phi(\tau)=\tau^{2}, we already saw that ϕ1(E[ϕ(F(xν,ξν))])=1\phi^{-1}\left(E[\phi(F(x^{\nu},\xi^{\nu}))]\right)=1 so we can now apply Corollary 1.3 to obtain

μ({c:|Fβ(xν,ξν;c)1|m(h)/2})exp(CN2κm2(h)).\mu\left(\{c:|F_{\beta}(x^{\nu},\xi^{\nu};c)-1|\geq{}m(h)/2\}\right)\leq\exp(-CN^{2\kappa}m^{2}(h)). (2.9)

Putting together (2.9) with the results of Theorem 2.1, the probability that there there is any line γ(x,ξ)\gamma(x,\xi) 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 PP is a semiclassical localiser, is PuL2\left|\!\left|{Pu}\right|\!\right|_{L^{2}} equidistributed? We say that PP is a semiclassical localiser if

P=Op(p(y,η;h))=ph(y,hD)=1(2πh)neihyz,ηp(y,η;h)u(z)𝑑z𝑑ηP=Op(p(y,\eta;h))=p_{h}(y,hD)=\frac{1}{(2\pi h)^{n}}\int e^{\frac{i}{h}\langle y-z,\eta\rangle}p(y,\eta;h)u(z)dzd\eta

where p(y,η;h)p(y,\eta;h) is compactly supported in n×n\mathbb{R}^{n}\times\mathbb{R}^{n}. The support of p(y,η;h)p(y,\eta;h) may depend on hh but is constrained by the uncertainty principle,

|xpηp|h1.|\nabla_{x}p\cdot\nabla_{\eta}p|\leq h^{-1}.

For the purposes of this paper we will be assuming that p(y,η;h)p(y,\eta;h) is supported near a point (x,ξ)(x,\xi). We will introduce two parameters α\alpha and μ\mu to measure how near.

  1. (1)

    The hαh^{\alpha} scale controls the level of angular frequency localisation. The radial frequency is localised so that |η|4|\eta|\leq 4. We can ignore contributions from frequencies |η|4|\eta|\geq{}4 since uu is a sum of plane waves with frequencies in [1hβ,1+hβ][1-h^{\beta},1+h^{\beta}], β0\beta\geq{}0.

  2. (2)

    The parameter μ\mu controls how close/far we are from Planck scale. When μ=1\mu=1 we are exactly at Planck scale and therefore at the limits of the uncertainty principle. Large μ\mu (growing in h1h^{-1}) correspond to more relaxed localisation.

Let (x,ξ)n×𝕊n1(x,\xi)\in\mathbb{R}^{n}\times\mathbb{S}^{n-1}. Because of the curvature we obtain linearisation when α=β/2\alpha=\beta/2 so we are interested in the cases αβ/2\alpha\leq\beta/2. Figure 2 shows the support of p(x,ξ)α,μ(y,η)p^{\alpha,\mu}_{(x,\xi)}(y,\eta) in both configuration and Fourier space.

Refer to caption
Figure 2. The configuration an Fourier supports of the symbol p(x,ξ)α,μ(y,η)p^{\alpha,\mu}_{(x,\xi)}(y,\eta).

Define

P(x,ξ)α,μ=p(x,ξ)α,μ(y,hD)=Op(p(x,ξ)α,μ(y,η))P_{(x,\xi)}^{\alpha,\mu}=p_{(x,\xi)}^{\alpha,\mu}(y,hD)=Op(p_{(x,\xi)}^{\alpha,\mu}(y,\eta))

where

p(x,ξ)α,μ(y,η)=Ahn2+αμn+12χ(μ2h1+2α|xy,ξ|)×χ(μ1h1+α|(xy)xy,ξξ|)χ(|η|4)χ(hα|η|η|ξ|).p_{(x,\xi)}^{\alpha,\mu}(y,\eta)=Ah^{-\frac{n}{2}+\alpha}\mu^{-\frac{n+1}{2}}\chi\left(\mu^{-2}h^{-1+2\alpha}|\langle x-y,\xi\rangle|\right)\\ \times\chi\left(\mu^{-1}h^{-1+\alpha}|(x-y)-\langle x-y,\xi\rangle\xi|\right)\chi\left(\frac{|\eta|}{4}\right)\chi\left(h^{-\alpha}\left|\frac{\eta}{|\eta|}-\xi\right|\right). (3.1)

The prefactor is chosen so that

A2h1β+(n1)(1α)jχ2(hα|ξjξ|)χ2(|y,ξ|)χ2(|yy,ξ|)𝑑y=1.A^{2}h^{1-\beta+(n-1)(1-\alpha)}\sum_{j}\chi^{2}\left(h^{-\alpha}\left|\xi_{j}-\xi\right|\right)\int\chi^{2}\left(|\langle y,\xi\rangle|\right)\chi^{2}\left(|y-\langle y,\xi\rangle|\right)dy=1.

The angular cut off will ensure that AA is of order one. Note that

P(x,ξ)α,μu()L22=P(x,ξ)α,μu(),P(x,ξ)α,μu()=u(),(P(x,ξ)α,μ)P(x,ξ)α,μu().\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}}^{2}=\langle P^{\alpha,\mu}_{(x,\xi)}u(\cdot),P^{\alpha,\mu}_{(x,\xi)}u(\cdot)\rangle=\langle u(\cdot),(P^{\alpha,\mu}_{(x,\xi)})^{\star}P^{\alpha,\mu}_{(x,\xi)}u(\cdot)\rangle.

So if the random wave is assumed to equidistribute down to small scale (in terms of semiclassical defect measures) then this normalisation ensures that

P(x,ξ)α,μu()L22=1+O(μ1).\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}}^{2}=1+O(\mu^{-1}).

Consider the random variable Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) given by

Gα,β,μ=P(x,ξ)α,μu()L2.G_{\alpha,\beta,\mu}=\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}}. (3.2)

If we can show that

𝔼[Gα,β,μ(x,ξ)L(B1(0)×𝕊n1)]clog(1/h)\mathbb{E}[\left|\!\left|{G_{\alpha,\beta,\mu}(x,\xi)}\right|\!\right|_{L^{\infty}(B_{1}(0)\times\mathbb{S}^{n-1})}]\geq{}c\sqrt{\log(1/h)}

we establish that, apart from a set whose measure decays to zero in hh, equidistribution fails at small scales. We will show this is true for small μ\mu, that is μC\mu\leq C. We will then see that for large μ\mu, μhϵ\mu\geq{}h^{-\epsilon} 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 Gα,β,μL\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}} by Gα,β,μLph\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}} for ph=δlog(h)p_{h}=\delta\log(h). To compute Gα,β,μLph\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}} we again use Theorem 1.2. In this case we set ϕ(τ)=τph\phi(\tau)=\tau^{p_{h}} and compute Gα,β,μLphph\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}}^{p_{h}}. To facilitate that calculation we will assume that php_{h} is an even integer (to allow us to expand the expression for Gα,β,μG_{\alpha,\beta,\mu}).

Proposition 3.1.

Suppose Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) is given by (3.2). Then for ph=δlog(1/h)p_{h}=\delta\log(1/h) (for some δ>0\delta>0), there are constants C1,C2C_{1},C_{2} so that,

C1Gα,β,μLph(B1(0)×Sn1)Gα,β,μL(B1(0)×Sn1)C2Gα,β,μLph(B1(0)×Sn1).C_{1}\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times S^{n-1})}\leq\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}(B_{1}(0)\times S^{n-1})}\leq C_{2}\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times S^{n-1})}. (3.3)
Proof.

Since B1(0)×𝕊n1B_{1}(0)\times\mathbb{S}^{n-1} is compact, the first half of (3.3) follows from Hölder. To get the second half we will in fact control the LL^{\infty} norm of Gα,β,μ2G^{2}_{\alpha,\beta,\mu} in terms of its Lph/2L^{p_{h}/2} norm. The requisite control comes directly from the regularity properties of the symbol p(x,ξ)α,μ(y,η)p^{\alpha,\mu}_{(x,\xi)}(y,\eta). The semiclassical Sobolev estimates tell us that if W:dW:\mathbb{R}^{d}\to\mathbb{R} is semiclassically localised,

W=χ(w,hD)W+OL2(h)W=\chi(w,hD)W+O_{L^{2}}(h)

for some χ(w,ω)\chi(w,\omega) with compact support then

WLChdpWL2.\left|\!\left|{W}\right|\!\right|_{L^{\infty}}\leq Ch^{-\frac{d}{p}}\left|\!\left|{W}\right|\!\right|_{L^{2}}.

Saying that WW is semiclassically localised is the same as saying that WW is localised in space and that (up to Schwartz type error) its Fourier transform is localised in the ball of radius 1/h1/h. Notice that

|Dxγ1Dξγ2p(x,ξ)α,μ|Ch(|γ1|+|γ2|).|D^{\gamma_{1}}_{x}D^{\gamma_{2}}_{\xi}p^{\alpha,\mu}_{(x,\xi)}|\leq Ch^{-(|\gamma_{1}|+|\gamma_{2}|)}.

Therefore indeed outside the ball of radius 1/h1/h the Fourier transform of p(x,ξ)α,μp^{\alpha,\mu}_{(x,\xi)} decays in a Schwartz fashion and we can apply the semiclassical Sobolev estimates. In this case d=2n1d=2n-1 and if we set ph=δlog(1/h)p_{h}=\delta\log(1/h) we obtain

Gα,β,μ2LCe4n2cGα,β,μ2Lph/2.\left|\!\left|{G^{2}_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}}\leq Ce^{\frac{4n-2}{c}}\left|\!\left|{G^{2}_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}/2}}.

Theorem 3.2.

Let Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) be given by (3.2). For μC\mu\leq C and small enough (but independent of hh) there exist constants C1,C2C_{1},C_{2} so that

C1log(1/h)𝔼[Gα,β,μL(B1(0)×𝕊n1)]C2log(1/h).C_{1}\sqrt{\log(1/h)}\leq\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}(B_{1}(0)\times\mathbb{S}^{n-1})}\right]\leq C_{2}\sqrt{\log(1/h)}. (3.4)
Proof.

From the results of Proposition 3.1 it is enough to find upper and lower bounds for 𝔼[Gα,β,μLph]\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}}\right] for phδlog(1/h)p_{h}\approx\delta\log(1/h). Therefore let php_{h} be an even integer obeying δ2log(1/h)phδlog(1/h)\frac{\delta}{2}\log(1/h)\leq p_{h}\leq\delta\log(1/h). Let ϕ(τ)=τph\phi(\tau)=\tau^{p_{h}}, we will compute

ϕ1(𝔼[ϕ(Gα,β,μLph(B1(0)×𝕊n1))])\phi^{-1}\left(\mathbb{E}\left[\phi\left(\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}\right)\right]\right)

and apply the measure concentration results of Theorem 1.2.

Since we have assumed php_{h} is even we can write

|Gα,β,μ(x,ξ)|ph=(|Gα,β,μ(x,ξ)|2)ph/2.|G_{\alpha,\beta,\mu}(x,\xi)|^{p_{h}}=\left(|G_{\alpha,\beta,\mu}(x,\xi)|^{2}\right)^{p_{h}/2}.

Since

u=λjΛβcjeihy,ξju=\sum_{\lambda_{j}\in\Lambda_{\beta}}c_{j}e^{\frac{i}{h}\langle y,\xi_{j}\rangle}

Gα,β,μ2(x,ξ)G^{2}_{\alpha,\beta,\mu}(x,\xi) is quadratic in the column vector c=[c1,,cN]Tc=[c_{1},\dots,c_{N}]^{T}. We will, similar to [8], scale and perform transformations so that Gα,β,μ2(x,ξ)G^{2}_{\alpha,\beta,\mu}(x,\xi) can be written as yTDyy^{T}Dy where DD is a diagonal matrix and yTy^{T} is a standard NN-dimensional Gaussian variable (with variance one).

Let w=σc=N1/2cw=\sigma c=N^{-1/2}c we can write |Gα,β,μ(x,ξ)|2|G_{\alpha,\beta,\mu}(x,\xi)|^{2} as

|Gα,β,μ(x,ξ)|2=wT[Aα,β,μ]w|G_{\alpha,\beta,\mu}(x,\xi)|^{2}=w^{T}\left[A_{\alpha,\beta,\mu}\right]\,w

where Aα,β,μA_{\alpha,\beta,\mu} is the matrix with entries

(Aα,β,μ)j,k=N1C1/2ψj(y)ψ¯k(y)𝑑y(A_{\alpha,\beta,\mu})_{j,k}=N^{-1}\int_{C_{1/2}}\psi_{j}(y)\overline{\psi}_{k}(y)dy (3.5)

with

ψj(y)=pα,μ(y,hD)eihy,ξj.\psi_{j}(y)=p^{\alpha,\mu}(y,hD)e^{\frac{i}{h}\langle y,\xi_{j}\rangle}.

If we diagonalise Aα,β,μA_{\alpha,\beta,\mu} to write Aα,β,μ=UTDUA_{\alpha,\beta,\mu}=U^{T}DU where UU is unitary and DD is diagonal with entries λj\lambda_{j} we can write

|Gα,β,μ(x,ξ)|2=yTDy=jλjyj2|G_{\alpha,\beta,\mu}(x,\xi)|^{2}=y^{T}Dy=\sum_{j}\lambda_{j}y_{j}^{2}

where y=Uwy=Uw. Since UU is unitary yy is also an NN-dimensional Gaussian random variable with variance one. Recall that

|Gα,β,μ(x,ξ)|2=p(x,ξ)α,μ(y,hD)uL22|G_{\alpha,\beta,\mu}(x,\xi)|^{2}=\left|\!\left|{p^{\alpha,\mu}_{(x,\xi)}(y,hD)u}\right|\!\right|_{L^{2}}^{2}

so Aα,β,μA_{\alpha,\beta,\mu} must have only non-negative eigenvalues. Let MM be a large integer (eventually we will want Mlog(1/h)M\sim\log(1/h) so that ph=2Mp_{h}=2M). Then

|Gα,β,μ(x,ξ)|2M\displaystyle|G_{\alpha,\beta,\mu}(x,\xi)|^{2M} =(jλjyj2)M\displaystyle=\left(\sum_{j}\lambda_{j}y_{j}^{2}\right)^{M}
=[j1,,jM]k=1Mλjkyjk2.\displaystyle=\sum_{[j_{1},\dots,j_{M}]}\prod_{k=1}^{M}\lambda_{j_{k}}y_{j_{k}}^{2}.

Let J=[j1,,jM]J=[j_{1},\dots,j_{M}] and let |J|d|J|_{d} be the number of distinct jij_{i} in JJ. Then

|Gα,β,μ(x,ξ)|M=r=0R2r1|J|d2rk=1Mλjkyjk2|G_{\alpha,\beta,\mu}(x,\xi)|^{M}=\sum_{r=0}^{R}\sum_{2^{r-1}\leq|J|_{d}\leq 2^{r}}\prod_{k=1}^{M}\lambda_{j_{k}}y_{j_{k}}^{2}

with 2R=M2^{R}=M. Since the yjk2y_{j_{k}}^{2} are independent Gaussian variables we can compute

𝔼[k=1Myjk2]=(2π)MΓ(p1+1/2)Γ(p2+1/2)Γ(p|J|d+1/2)\mathbb{E}\left[\prod_{k=1}^{M}y_{j_{k}}^{2}\right]=\left(\frac{2}{\sqrt{\pi}}\right)^{M}\Gamma(p_{1}+1/2)\Gamma(p_{2}+1/2)\cdots\Gamma(p_{|J|_{d}}+1/2)

where the pip_{i} are the number of distinct occurrences in JJ. Obviously p1+p2++p|J|d=Mp_{1}+p_{2}+\cdots+p_{|J|_{d}}=M. So

𝔼[|Gα,β,μ(x,ξ)|M]=r=0R2r1<|J|d2rk=1|J|dλjkpkΓ(pk+1/2).\mathbb{E}\left[|G_{\alpha,\beta,\mu}(x,\xi)|^{M}\right]=\sum_{r=0}^{R}\sum_{2^{r-1}<|J|_{d}\leq{}2^{r}}\prod_{k=1}^{|J|_{d}}\lambda_{j_{k}}^{p_{k}}\Gamma(p_{k}+1/2).

We obtain uniform upper and lower bounds for 𝔼[|Gα,β,μ(x,ξ)|M]\mathbb{E}\left[|G_{\alpha,\beta,\mu}(x,\xi)|^{M}\right] which, alongside the compact nature of B1(0)×𝕊n1B_{1}(0)\times\mathbb{S}^{n-1} carry over to give upper and lower bounds on 𝔼[Gα,β,μ(x,ξ)Lphph]\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}(x,\xi)}\right|\!\right|_{L^{p_{h}}}^{p_{h}}\right]. Since all the eigenvalues λj\lambda_{j} are non-negative we can obtain lower bounds by obtain a lower bound for any fixed rr. We claim that if μC\mu\leq C (the Planck scale case) then there is some 𝒋\boldsymbol{j} so that

λ𝒋>cTr(Aα,β,μ).\lambda_{\boldsymbol{j}}>c\text{Tr}(A_{\alpha,\beta,\mu}). (3.6)

In which case the term associated with J=[𝒋,𝒋,,𝒋]J=[\boldsymbol{j},\boldsymbol{j},\dots,\boldsymbol{j}] has the lower bound

k=1|J|dλjkpkΓ(pk+1/2)>cN(Tr(Aα,β))MΓ(M+1/2)>CM(Tr(Aα,β))MMM.\prod_{k=1}^{|J|_{d}}\lambda_{j_{k}}^{p_{k}}\Gamma(p_{k}+1/2)>c^{N}(\text{Tr}(A_{\alpha,\beta}))^{M}\Gamma(M+1/2)>C^{M}(\text{Tr}(A_{\alpha,\beta}))^{M}M^{M}.

To justify this claim we (in Lemma 4.1) compute and compare Tr(Aα,β,μ)\mathrm{Tr}\,(A_{\alpha,\beta,\mu}) and Tr(Aα,β,μ2)\mathrm{Tr}\,(A^{2}_{\alpha,\beta,\mu}). For μC\mu\leq C we find that,

(iλi)2iλi2C\frac{\left(\sum_{i}\lambda_{i}\right)^{2}}{\sum_{i}\lambda_{i}^{2}}\leq{}C

and therefore we can conclude (3.6). On the other hand

(2π)MΓ(p1+1/2)Γ(p2+1/2)Γ(p|J|d+1/2)CMMM\left(\frac{2}{\sqrt{\pi}}\right)^{M}\Gamma(p_{1}+1/2)\Gamma(p_{2}+1/2)\cdots\Gamma(p_{|J|_{d}}+1/2)\leq C^{M}M^{M}

for any combination of pip_{i} so

𝔼[|Gα,β,μ(x,ξ)|M]CMMM(jλj)M=CMMM(Tr(Aα,β,μ))M.\mathbb{E}\left[|G_{\alpha,\beta,\mu}(x,\xi)|^{M}\right]\leq C^{M}M^{M}\left(\sum_{j}\lambda_{j}\right)^{M}=C^{M}M^{M}(\text{Tr}(A_{\alpha,\beta,\mu}))^{M}.

In Lemma 4.1 we see that there are constants a1a_{1} and a2a_{2} such that

a1Tr(Aα,β,μ)a2a_{1}\leq\text{Tr}(A_{\alpha,\beta,\mu})\leq a_{2}

so if ph=2Mp_{h}=2M we have that

C1ph1/2(𝔼[Gα,β,μ(x,ξ)Lph(B1(0)×𝕊n1)ph])1phC2ph1/2.C_{1}p_{h}^{1/2}\leq\left(\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}(x,\xi)}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}^{p_{h}}\right]\right)^{\frac{1}{p_{h}}}\leq C_{2}p_{h}^{1/2}.

Now all we need is the concentration of measure that allows us to relate ϕ1(𝔼[ϕ(Gα,β,μLph)])\phi^{-1}\left(\mathbb{E}\left[\phi(\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}})\right]\right) to 𝔼[Gα,β,μLph]\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}}\right]. In Lemma 4.2 we find that for μC\mu\leq C and

Gp=Gα,β,μLp(B1(0)×𝕊n1)G_{p}=\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p}(B_{1}(0)\times\mathbb{S}^{n-1})}
|Gp(c)Gp(d)|N121pcd2.|G_{p}(c)-G_{p}(d)|\leq N^{\frac{1}{2}-\frac{1}{p}}\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}.

So applying the results of Theorem 1.2

𝔼[Gα,β,μLph(B1(0)×𝕊n1)]=(𝔼[Gα,β,μLph(B1(0)×𝕊n1)ph])1/ph+O(ph12N1ph).\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}\right]=\left(\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}^{p_{h}}\right]\right)^{1/p_{h}}+O\left(\frac{p_{h}^{\frac{1}{2}}}{N^{\frac{1}{p_{h}}}}\right).

Since δ2log(1/h)phδlog(1/h)\frac{\delta}{2}\log(1/h)\leq p_{h}\leq\delta\log(1/h),

𝔼[Gα,β,μLph(B1(0)×𝕊n1)]=(𝔼[Gα,β,μLph(B1(0)×𝕊n1)ph])1/ph+O(ph12enβδ).\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}\right]=\left(\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p_{h}}(B_{1}(0)\times\mathbb{S}^{n-1})}^{p_{h}}\right]\right)^{1/p_{h}}+O\left(p_{h}^{\frac{1}{2}}e^{-\frac{n-\beta}{\delta}}\right).

Therefore by making δ\delta small enough we can obtain both upper and lower bounds, that is

C1log(1/h)𝔼[Gα,β,μL(B1(0)×𝕊n1)]C2log(1/h).C_{1}\sqrt{\log(1/h)}\leq\mathbb{E}\left[\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{\infty}(B_{1}(0)\times\mathbb{S}^{n-1})}\right]\leq C_{2}\sqrt{\log(1/h)}.

We now consider the case where we are far from Planck scale. That is where μ>hϵ\mu>h^{-\epsilon}. In this case we find that the uniform version of equidistribution holds.

Theorem 3.3.

Let Gα,β,μ(x,ξ)G_{\alpha,\beta,\mu}(x,\xi) be given by (3.2) and suppose μhϵ\mu\geq h^{-\epsilon}, then

𝔼[Gα,β,μ(x,ξ)]=1+O(hϵ2).\mathbb{E}\left[G_{\alpha,\beta,\mu}(x,\xi)\right]=1+O\left(h^{\frac{\epsilon}{2}}\right). (3.7)

Further if δ>0\delta>0 and m(h)max(hϵδ,h2ϵ(n+1)4(nβ)δ)m(h)\geq{}\max\left(h^{\epsilon-\delta},h^{\frac{2\epsilon(n+1)}{4(n-\beta)}-\delta}\right),

Pr{c:(x,ξ), so that |Gα,β,μ(x,ξ)1|>m(h)}exp(N2ϵ(n+1)4(nβ)m2(h)).Pr\{c:\exists(x,\xi),\text{ so that }\left|G_{\alpha,\beta,\mu}(x,\xi)-1\right|>m(h)\}\leq\exp(-N^{\frac{2\epsilon(n+1)}{4(n-\beta)}}m^{2}(h)). (3.8)
Proof.

We will follow much the same process as we did in Theorems 2.1 and 2.2 where we controlled the behaviour of F(x,ξ)F(x,\xi). That is we will set ϕ(τ)=τ2\phi(\tau)=\tau^{2} and use Theorem 1.2 to compute E[Gα,β,μ(x,ξ)]E\left[G_{\alpha,\beta,\mu}(x,\xi)\right] in terms of 𝔼[Gα,β,μ2(x,ξ)]\mathbb{E}\left[G^{2}_{\alpha,\beta,\mu}(x,\xi)\right]. Note that

Gα,β,μ2(x,ξ)=j,lcjclP(x,ξ)α,μeihy,ξj,P(x,ξ)α,μeihy,ξl.G^{2}_{\alpha,\beta,\mu}(x,\xi)=\sum_{j,l}c_{j}c_{l}\langle P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle},P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{l}\rangle}\rangle.

Since the Gaussian variables have mean zero all non-diagonal terms fall out when we compute expectation so we only need compute

jP(x,ξ)α,μeihy,ξj,P(x,ξ)α,μeihy,ξj.\sum_{j}\langle P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle},P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle}\rangle.

Since μhϵ\mu\geq{}h^{-\epsilon} we can write

P(x,ξ)α,μeihy,ξj=p(x,ξ)α,μ(y,ξj)eihy,ξj+hϵr(y,ξj)eihy,ξP^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle}=p^{\alpha,\mu}_{(x,\xi)}(y,\xi_{j})e^{\frac{i}{h}\langle y,\xi_{j}\rangle}+h^{\epsilon}r(y,\xi_{j})e^{\frac{i}{h}\langle y,\xi\rangle}

where r(y,ξj)r(y,\xi_{j}) has the same support an normalisation properties as p(x,ξ)α,μp^{\alpha,\mu}_{(x,\xi)}. The contribution from the top terms is

h2αβμ(n+1)A2jχ2(hα|ξj|ξj|ξ|)χ2(μ2h1+2α|xy,ξ|)×χ2(μ1h1+α|(xy)xy,ξ|)dy.h^{2\alpha-\beta}\mu^{-(n+1)}A^{2}\sum_{j}\chi^{2}\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\int\chi^{2}\left(\mu^{-2}h^{-1+2\alpha}|\langle x-y,\xi\rangle|\right)\\ \times\chi^{2}\left(\mu^{-1}h^{-1+\alpha}|(x-y)-\langle x-y,\xi\rangle|\right)dy.

Rescaling and translating/rotating so that x=0x=0 and ξ=e1\xi=e_{1} we see that the normalisation on AA ensures that this term is one. Since r(y,ξj)r(y,\xi_{j}) has the same normalisation properties as pα,μ(y,ξj)p^{\alpha,\mu}(y,\xi_{j}) we can conclude that

jP(x,ξ)α,μeihy,ξj,P(x,ξ)α,μeihy,ξj=1+O(hϵ).\sum_{j}\langle P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle},P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle y,\xi_{j}\rangle}\rangle=1+O(h^{\epsilon}).

Therefore

𝔼[Gα,β,μ2(x,ξ)]=1+O(hϵ).\mathbb{E}\left[G^{2}_{\alpha,\beta,\mu}(x,\xi)\right]=1+O(h^{\epsilon}).

Now we can use the measure concentration to find 𝔼[Gα,β,μ(x,ξ)]\mathbb{E}\left[G_{\alpha,\beta,\mu}(x,\xi)\right]. In Lemma 4.2 we find that if

G=Gα,β,μ(,)L(B1(0)×𝕊n1)G_{\infty}=\left|\!\left|{G_{\alpha,\beta,\mu}(\cdot,\cdot)}\right|\!\right|_{L^{\infty}(B_{1}(0)\times\mathbb{S}^{n-1})}

then

|G(c)G(d)|N1/2μ(n+14)N1/2hϵ(n+14)=N12κ|G_{\infty}(c)-G_{\infty}(d)|\leq N^{1/2}\mu^{-\left(\frac{n+1}{4}\right)}\leq N^{1/2}h^{\epsilon\left(\frac{n+1}{4}\right)}=N^{\frac{1}{2}-\kappa}

for

κ=ϵnβ(n+14).\kappa=\frac{\epsilon}{n-\beta}\left(\frac{n+1}{4}\right). (3.9)

So by Theorem 1.2

𝔼[Gα,β,μ(x,ξ)]=1+O(hϵ+Nκ)=1+O(hϵ+h2ϵ(n+1)4(nβ)).\mathbb{E}\left[G_{\alpha,\beta,\mu}(x,\xi)\right]=1+O(h^{\epsilon}+N^{-\kappa})=1+O\left(h^{\epsilon}+h^{\frac{2\epsilon(n+1)}{4(n-\beta)}}\right).

Note that the cut offs we used to define p(x,ξ)α,μp^{\alpha,\mu}_{(x,\xi)} have power (in h1h^{-1}) type regularity. So by the same arguments that we used in the proof of uniform equidistribution of F(x,ξ)F(x,\xi) we can find a polynomial grid (xν,ξν)(x^{\nu},\xi^{\nu}) so that failure to equidistribute at some point (x,ξ)n×𝕊n1(x,\xi)\in\mathbb{R}^{n}\times\mathbb{S}^{n-1} implies a failure to equidistribute at a grid point (xν,ξν)(x^{\nu},\xi^{\nu}). Then applying the concentration of measure at these points we obtain

Pr{c:|Gα,β,μ(xν,ξν)1|m(h)}exp(N2κm2(h)).Pr\{c:|G_{\alpha,\beta,\mu}(x^{\nu},\xi^{\nu})-1|\geq{}m(h)\}\leq\exp(-N^{2\kappa}m^{2}(h)).

Since the number of grid points only growth polynomially in h1h^{-1} (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 Tr(Aα,β,μ)\mathrm{Tr}\,(A_{\alpha,\beta,\mu}) and Tr(Aα,β,μ2)\mathrm{Tr}\,(A_{\alpha,\beta,\mu}^{2}) dependent on μ\mu (the parameter that controls how far we are from the Planck scale).

Lemma 4.1.

Suppose Aα,β,μA_{\alpha,\beta,\mu} is given by (3.5). Then there exist constants a1,a2,a3,a4>0a_{1},a_{2},a_{3},a_{4}>0 so that

a1Tr(Aα,β,μ)a2a_{1}\leq\mathrm{Tr}\,(A_{\alpha,\beta,\mu})\leq a_{2} (4.1)

and

a3μn1Tr(Aα,β,μ2)a4μn1.a_{3}\mu^{-n-1}\leq{}\mathrm{Tr}\,(A_{\alpha,\beta,\mu}^{2})\leq a_{4}\mu^{-n-1}. (4.2)
Proof.

From (3.5) we have that,

Tr(Aα,β,μ)=hnβj|ψj(y)|2𝑑y\mathrm{Tr}\,(A_{\alpha,\beta,\mu})=h^{n-\beta}\sum_{j}\int|\psi_{j}(y)|^{2}dy

and

Tr(Aα,β,μ2)\displaystyle\mathrm{Tr}\,(A_{\alpha,\beta,\mu}^{2}) =h2(nβ)j,mψj(y)ψ¯j(y)ψm¯(y)ψm(y)𝑑y𝑑y\displaystyle=h^{2(n-\beta)}\sum_{j,m}\iint\psi_{j}(y)\overline{\psi}_{j}(y^{\prime})\overline{\psi_{m}}(y)\psi_{m}(y^{\prime})dydy^{\prime}
=h2(nβ)j,m|Ij,m|2\displaystyle=h^{2(n-\beta)}\sum_{j,m}|I_{j,m}|^{2}

where

Ij,m=ψj(y)ψ¯m(y)𝑑y.I_{j,m}=\int\psi_{j}(y)\overline{\psi}_{m}(y)dy.

Recall that

ψj(y)\displaystyle\psi_{j}(y) =p(x,ξ)α,μ(y,hD)eihy,ζj\displaystyle=p^{\alpha,\mu}_{(x,\xi)}(y,hD)e^{\frac{i}{h}\langle y,\zeta_{j}\rangle}
=hn2+αμn+12w(x,ξ)α,μ(y)qξα(hD)eihy,ζj\displaystyle=h^{-\frac{n}{2}+\alpha}\mu^{-\frac{n+1}{2}}w^{\alpha,\mu}_{(x,\xi)}(y)q^{\alpha}_{\xi}(hD)e^{\frac{i}{h}\langle y,\zeta_{j}\rangle}

where

w(x,ξ)α,μ(y)=χ(μ2h1+2α|xy,ξ|)χ(μ1h1α|(xy)xy,ξξ|)w^{\alpha,\mu}_{(x,\xi)}(y)=\chi\left(\mu^{-2}h^{-1+2\alpha}|\langle x-y,\xi\rangle|\right)\chi\left(\mu^{-1}h^{1-\alpha}|(x-y)-\langle x-y,\xi\rangle\xi|\right)

and

qξα(η)=χ(hα|η|η|ξ|)χ(|η|4).q^{\alpha}_{\xi}(\eta)=\chi\left(h^{-\alpha}\left|\frac{\eta}{|\eta|}-\xi\right|\right)\chi\left(\frac{|\eta|}{4}\right).

Therefore we can write

ψj(y)\displaystyle\psi_{j}(y) =hn2+αμn+12w(x,ξ)α,μ(y)h1(qξα()h[eih,ζj])\displaystyle=h^{-\frac{n}{2}+\alpha}\mu^{-\frac{n+1}{2}}w^{\alpha,\mu}_{(x,\xi)}(y)\mathcal{F}_{h}^{-1}\left(q^{\alpha}_{\xi}(\cdot)\mathcal{F}_{h}[e^{\frac{i}{h}\langle\cdot,\zeta_{j}\rangle}]\right)
=hn2+αμn+12eihy,ξjw(x,ξ)α,μ(y)χ(hα|ξj|ξj|ξ|).\displaystyle=h^{-\frac{n}{2}+\alpha}\mu^{-\frac{n+1}{2}}e^{\frac{i}{h}\langle y,\xi_{j}\rangle}w^{\alpha,\mu}_{(x,\xi)}(y)\chi\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right).

So we compute

Ij,m(x,ξ)\displaystyle I_{j,m}(x,\xi) =ψj(y)ψm(y)𝑑y\displaystyle=\int\psi_{j}(y)\psi_{m}(y)dy
=hn+2αμn1χ(hα|ξj|ξj|ξ|)χ(hα|ξm|ξm|ξ|)eihy,ξjξm(w(x,ξ)α,μ(y))2𝑑y.\displaystyle=h^{-n+2\alpha}\mu^{-n-1}\chi\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\chi\left(h^{-\alpha}\left|\frac{\xi_{m}}{|\xi_{m}|}-\xi\right|\right)\int e^{\frac{i}{h}\langle y,\xi_{j}-\xi_{m}\rangle}(w^{\alpha,\mu}_{(x,\xi)}(y))^{2}dy.

First note that Ij,m(x,ξ)=0I_{j,m}(x,\xi)=0 if either

|ξj|ξj|ξ|2hαor|ξm|ξm|ξ|>2hα.\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\geq{}2h^{\alpha}\quad\text{or}\quad\left|\frac{\xi_{m}}{|\xi_{m}|}-\xi\right|>2h^{\alpha}.

Immediately this tells us that there can only be Chα(n1)+βnCh^{\alpha(n-1)+\beta-n} of each where Ij,mI_{j,m} is nonzero. Conversely there is a smaller constant aa so that there are more than ahα(n1)+βnah^{\alpha(n-1)+\beta-n} points ξj\xi_{j} for which

χ(hα|ξj|ξj|ξ|)=1.\chi\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)=1.

If ξj=ξk\xi_{j}=\xi_{k} we have

Ij,j(x,ξ)=hn+2αμn1χ2(hα|ξj|ξj|ξ|)(wx,ξα,μ(y))2𝑑y.I_{j,j}(x,\xi)=h^{-n+2\alpha}\mu^{-n-1}\chi^{2}\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\int(w^{\alpha,\mu}_{x,\xi}(y))^{2}dy.

Therefore there are constants a1a_{1} and a2a_{2} so that

a1hα(n1)χ2(hα|ξj|ξj|ξ|)Ij,j(x,ξ)a2hα(n1)χ2(hα|ξj|ξj|ξ|).a_{1}h^{-\alpha(n-1)}\chi^{2}\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\leq I_{j,j}(x,\xi)\leq a_{2}h^{-\alpha(n-1)}\chi^{2}\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right).

Since there are of order hβnhα(n1)h^{\beta-n}h^{\alpha(n-1)} jj that appear in the sum we obtain constants (and here we abuse notation by allowing constants to vary line by line)

a1Tr(Aα,β)a2.a_{1}\leq{}\mathrm{Tr}\,(A_{\alpha,\beta})\leq{}a_{2}.

Now let’s turn our attention to Tr(Aα,β,μ2)\mathrm{Tr}\,(A^{2}_{\alpha,\beta,\mu}). The lower bounds on the diagonal terms only give that

a1hα(n1)+nβTr(Aα,μ2).a_{1}h^{-\alpha(n-1)+n-\beta}\leq{}\mathrm{Tr}\,(A^{2}_{\alpha,\mu}).

Therefore we need to use the off diagonal terms. What matters here is, how far apart do ξj\xi_{j} and ξm\xi_{m} have to be for

eihy,ξjξm(w(x,ξ)α,μ(y))2𝑑y\int e^{\frac{i}{h}\langle y,\xi_{j}-\xi_{m}\rangle}(w^{\alpha,\mu}_{(x,\xi)}(y))^{2}dy

to be small? The additional smallness (apart from that which comes entirely from the support) is due to the non-stationary phase integral. If ξjξm\xi_{j}-\xi_{m} is large enough so that the oscillation of eihy,ξjξme^{\frac{i}{h}\langle y,\xi_{j}-\xi_{m}\rangle} overwhelms the regularity of w(x,ξ)α,μ(y)w^{\alpha,\mu}_{(x,\xi)}(y) the contribution from this pair will be small. By rotations and translation it is enough to consider the case when x=0x=0 and ξ=e1\xi=e_{1}. In that case w(0,e1)α,μw^{\alpha,\mu}_{(0,e_{1})} is supported in a h12α×(h1α)n1h^{1-2\alpha}\times(h^{1-\alpha})^{n-1} tube (the long direction lies along the e1e_{1} axis. First let’s consider the Planck scale case (that is μ=1\mu=1). Let ηj,m=ξjξm\eta_{j,m}=\xi_{j}-\xi_{m}. In this coordinate system if

{|(ηj,m)1|ϵh2α|(ηj,m)|ϵhα\begin{cases}|(\eta_{j,m})_{1}|\leq\epsilon h^{2\alpha}\\ |(\eta_{j,m})^{\prime}|\leq{}\epsilon h^{\alpha}\end{cases} (4.3)

the factor

eihy,ηj,me^{\frac{i}{h}\langle y,\eta_{j,m}\rangle}

does not complete a full oscillation over the support of w(x,ξ)α,1w^{\alpha,1}_{(x,\xi)}. Therefore in these cases there are constants so that

a1hα(n1)χ(hα|ξj|ξj|ξ|)(hα|ξm|ξm|ξ|)|Ij,m|a2hα(n1)χ(hα|ξj|ξj|ξ|)(hα|ξm|ξm|ξ|).a_{1}h^{-\alpha(n-1)}\chi\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\left(h^{-\alpha}\left|\frac{\xi_{m}}{|\xi_{m}|}-\xi\right|\right)\leq|I_{j,m}|\\ \leq a_{2}h^{-\alpha(n-1)}\chi\left(h^{-\alpha}\left|\frac{\xi_{j}}{|\xi_{j}|}-\xi\right|\right)\left(h^{-\alpha}\left|\frac{\xi_{m}}{|\xi_{m}|}-\xi\right|\right).

Suppose that jj is one of the ϵn1hα(n1)+γn\epsilon^{n-1}h^{\alpha(n-1)+\gamma-n} points such that

|ξj|ξj|e1|ϵhα\left|\frac{\xi_{j}}{|\xi_{j}|}-e_{1}\right|\leq\epsilon h^{\alpha}

and similarly mm is one of the points such that

|ξm|ξm|ξ|ϵhα\left|\frac{\xi_{m}}{|\xi_{m}|}-\xi\right|\leq{}\epsilon h^{\alpha}

that is both ξj\xi_{j} and ξm\xi_{m} lie in the intersection between the cone with angle ϵhα2\frac{\epsilon h^{\alpha}}{2} from the e1e_{1} axis and the annulus [1hβ,1+hβ][1-h^{\beta},1+h^{\beta}] (see Figure 3).

Refer to caption
Figure 3. If ξj\xi_{j} and ξm\xi_{m} are too close to each other the oscillations of eihy,ξjξme^{\frac{i}{h}\langle y,\xi_{j}-\xi_{m}\rangle} are not enough to overwhelm the regularity of the symbol

Therefore indeed, since αβ/2\alpha\leq{}\beta/2, ηj,m\eta_{j,m} satisfies the conditions 4.3. So there are at least ϵ2(n1)h2α(n1)+2γ2n\epsilon^{2(n-1)}h^{2\alpha(n-1)+2\gamma-2n} points for which

|Ij,m|>aϵhα(n1)|I_{j,m}|>a_{\epsilon}h^{-\alpha(n-1)}

which yields

Tr(Aα,β,1)aϵ.\mathrm{Tr}\,(A_{\alpha,\beta,1})\geq{}a_{\epsilon}.

In this case the upper bound follows directly from the maximum number of (ξj,ξm)(\xi_{j},\xi_{m}) so the integrand of Ij,mI_{j,m} is nonzero.

Now consider what happens as we move away from Planck scale μ1\mu\gg 1. In this case the factor eihy,ηj,me^{\frac{i}{h}\langle y,\eta_{j,m}\rangle} does not significantly oscillate if

{|(ej,m)1|μ2h2α|(ηj,m)|μ1hα\begin{cases}|(e_{j,m})_{1}|\leq\mu^{-2}h^{2\alpha}\\ |(\eta_{j,m})^{\prime}|\leq{}\mu^{-1}h^{\alpha}\end{cases} (4.4)

but otherwise we are able to obtain some extra decay. Suppose first that

|(ηj,m)1|>2lμ2hα|(\eta_{j,m})_{1}|>2^{l}\mu^{-2}h^{\alpha}

then integrating by parts in y1y_{1} we find that

eihy,ηj,m(wα,μ(y))2𝑑y=heihy,ηj,m(ηj,m)1μ2h1+2αw~α,μ(y)𝑑y\int e^{\frac{i}{h}\langle y,\eta_{j,m}\rangle}(w^{\alpha,\mu}(y))^{2}dy=\int\frac{he^{\frac{i}{h}\langle y,\eta_{j,m}\rangle}}{(\eta_{j,m})_{1}}\mu^{-2}h^{-1+2\alpha}\tilde{w}^{\alpha,\mu}(y)dy

where w~α,μ(y)\tilde{w}^{\alpha,\mu}(y) has the same support and regularity properties as wα,μw^{\alpha,\mu}. Repeated applications of this argument show that for any LL

|Ij,m|2lLhα(n1).|I_{j,m}|\leq{}2^{-lL}h^{\alpha(n-1)}. (4.5)

If |(ηj,m)|>μ1hα|(\eta_{j,m})^{\prime}|>\mu^{-1}h^{\alpha} the same argument using integration by parts in yy^{\prime} variables gives (4.5). By picking LL large enough we can ensure that the major contribution to the sum j,m|Ij,m|2\sum_{j,m}|I_{j,m}|^{2} comes when (4.4) are satisfied. For any fixed jj there are O(hβnhα(n1)μn1O(h^{\beta-n}h^{\alpha(n-1)}\mu^{-n-1}) suitable mm. Therefore we arrive at the estimate that

a3μn1Tr(Aα,β,μ2)a4μn1.a_{3}\mu^{-n-1}\leq\mathrm{Tr}\,(A^{2}_{\alpha,\beta,\mu})\leq{}a_{4}\mu^{-n-1}.

The second lemma obtains Lipschitz bounds on G=Gα,β,μLp(B1(0)×𝕊n1)G=\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p}(B_{1}(0)\times\mathbb{S}^{n-1})}. These are key to using the measure concentration arguments.

Lemma 4.2.

Suppose Gp=Gα,β,μLp(B1(0)×𝕊n1)G_{p}=\left|\!\left|{G_{\alpha,\beta,\mu}}\right|\!\right|_{L^{p}(B_{1}(0)\times\mathbb{S}^{n-1})} the following Lipschitz bound holds,

|Gp(c)Gp(d)|N121pμn+14+n+12pcd2=μn+14+n+12phβn2(11p)cd2.|G_{p}(c)-G_{p}(d)|\leq N^{\frac{1}{2}-\frac{1}{p}}\mu^{-\frac{n+1}{4}+\frac{n+1}{2p}}\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}=\mu^{-\frac{n+1}{4}+\frac{n+1}{2p}}h^{\frac{\beta-n}{2}\left(1-\frac{1}{p}\right)}\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}. (4.6)
Proof.

Let

u\displaystyle u =ξjΛβcjeihx,ξj\displaystyle=\sum_{\xi_{j}\in\Lambda_{\beta}}c_{j}e^{\frac{i}{h}\langle x,\xi_{j}\rangle} v=ξjΛβdjeihx,ξj\displaystyle v=\sum_{\xi_{j}\in\Lambda_{\beta}}d_{j}e^{\frac{i}{h}\langle x,\xi_{j}\rangle}
U(x,ξ)\displaystyle U(x,\xi) =P(x,ξ)α,μu()L2\displaystyle=\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}} V(x,ξ)=P(x,ξ)α,μv()L2.\displaystyle V(x,\xi)=\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}v(\cdot)}\right|\!\right|_{L^{2}}.

In this notation

|Gp(c)Gp(d)|=|ULp(B1(0)×Sn1)VLp(B1(0)×Sn1)|UVLp(B1(0)×𝕊n1).|G_{p}(c)-G_{p}(d)|=\left|\left|\!\left|{U}\right|\!\right|_{L^{p}(B_{1}(0)\times S^{n-1})}-\left|\!\left|{V}\right|\!\right|_{L^{p}(B_{1}(0)\times S^{n-1})}\right|\leq\left|\!\left|{U-V}\right|\!\right|_{L^{p}(B_{1}(0)\times\mathbb{S}^{n-1})}.

So if we can find estimates for p=p=\infty and p=2p=2 we can interpolate all the others. First let’s see the p=p=\infty case.

U(x,ξ)V(x,ξ)=P(x,ξ)α,μu()L2P(x,ξ)α,μv()L2P(x,ξ)α,μ(uv)()L2.U(x,\xi)-V(x,\xi)=\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}u(\cdot)}\right|\!\right|_{L^{2}}-\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}v(\cdot)}\right|\!\right|_{L^{2}}\leq\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}(u-v)(\cdot)}\right|\!\right|_{L^{2}}.
P(x,ξ)α,μ(uv)()L2=(j,m(cjdj)(cmdm)Ij,m)1/2,\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}(u-v)(\cdot)}\right|\!\right|_{L^{2}}=\left(\sum_{j,m}(c_{j}-d_{j})(c_{m}-d_{m})I_{j,m}\right)^{1/2},

where Ij,mI_{j,m} are as in the proof of Lemma 4.1. Applying Cauchy-Schwartz

P(x,ξ)α,μ(uv)()L2cd2(j,mIj,m2)1/4.\left|\!\left|{P^{\alpha,\mu}_{(x,\xi)}(u-v)(\cdot)}\right|\!\right|_{L^{2}}\leq\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}\left(\sum_{j,m}I_{j,m}^{2}\right)^{1/4}.

In the proof of Lemma 4.1 we have already estimated the sum j,m|Ij,m|2\sum_{j,m}|I_{j,m}|^{2} by

j,m|Ij,m|2Cμn1N2.\sum_{j,m}|I_{j,m}|^{2}\leq C\mu^{-n-1}N^{2}.

So this leads to a Lipschitz bound

|G(c)G(d)|Cμn+14N1/2=Cμn+14hβn2.|G_{\infty}(c)-G_{\infty}(d)|\leq C\mu^{-\frac{n+1}{4}}N^{1/2}=C\mu^{-\frac{n+1}{4}}h^{\frac{\beta-n}{2}}.

Now for the p=2p=2 case. In that case

UVL2\displaystyle\left|\!\left|{U-V}\right|\!\right|_{L^{2}} =B1(0)×𝕊n1P(x,ξ)α,μ(uv),P(x,ξ)α,μ(uv)y𝑑x𝑑ω(ξ)\displaystyle=\int_{B_{1}(0)\times\mathbb{S}^{n-1}}\langle P^{\alpha,\mu}_{(x,\xi)}(u-v),P^{\alpha,\mu}_{(x,\xi)}(u-v)\rangle_{y}dxd\omega(\xi)
=j,m(cjdj)(cmdm)yB1(0)×𝕊n1(P(x,ξ)α,μeih,ξjP(x,ξ)α,μeih,ξm¯)|ydxdω(ξ)dy\displaystyle=\sum_{j,m}(c_{j}-d_{j})(c_{m}-d_{m})\int_{y}\int_{B_{1}(0)\times\mathbb{S}^{n-1}}\left(P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle\cdot,\xi_{j}\rangle}\overline{P^{\alpha,\mu}_{(x,\xi)}e^{\frac{i}{h}\langle\cdot,\xi_{m}\rangle}}\right)\Big{|}_{y}dxd\omega(\xi)dy
=j,m(cjcm)(cmdm)eihy,ξjξma(y)𝑑y.\displaystyle=\sum_{j,m}(c_{j}-c_{m})(c_{m}-d_{m})\int e^{\frac{i}{h}\langle y,\xi_{j}-\xi_{m}\rangle}a(y)dy.

To get the last line we have integrated the (x,ξ)(x,\xi) variables first, integrating out the cut-off functions. Therefore since the ξj\xi_{j} are spaced at order hh (and are therefore almost orthogonal) we have

UVL2Ccd2.\left|\!\left|{U-V}\right|\!\right|_{L^{2}}\leq C\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}.

Interpolating with LL^{\infty}

UVLpCN121pμn+14+n+12pcd2\left|\!\left|{U-V}\right|\!\right|_{L^{p}}\leq CN^{\frac{1}{2}-\frac{1}{p}}\mu^{-\frac{n+1}{4}+\frac{n+1}{2p}}\left|\!\left|{c-d}\right|\!\right|_{\ell^{2}}

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.
  • [2] Alex Barnett. Random plane waves, https://users.flatironinstitute.org/~ahb/rpws/. Accessed: 17-10-2022.
  • [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. LpL^{p} 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 LpL^{p} 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. LL^{\infty} 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.