We consider a random Gaussian model of Laplace eigenfunctions on the hemisphere, satisfying the Dirichlet boundary conditions along the equator. For this model, we find a precise asymptotic law for the corresponding zero density functions, in both short range (around the boundary) and long range (far away from the boundary) regimes. As a corollary, we were able to find a logarithmic negative bias for the total nodal length of this ensemble relative to the rotation invariant model of random spherical harmonics. Jean Bourgain’s research, and his enthusiastic approach to the nodal geometry of Laplace eigenfunctions, has made a crucial impact in the field and the current trends within. His works on the spectral correlations {Theorem 2.2 in the work of Krishnapur et al. [Ann. Math. 177(2), 699–737 (2013)]} and Bombieri and Bourgain [Int. Math. Res. Not. (IMRN) 11, 3343–3407 (2015)] have opened a door for an active ongoing research on the nodal length of functions defined on surfaces of arithmetic flavor, such as the torus or the square. Furthermore, Bourgain’s work [J. Bourgain, Isr. J. Math. 201(2), 611–630 (2014)] on toral Laplace eigenfunctions, also appealing to spectral correlations, allowed for inferring deterministic results from their random Gaussian counterparts.
I. INTRODUCTION
A. Nodal length of Laplace eigenfunctions
The nodal line of a smooth function , defined on a smooth compact surface , with or without a boundary, is its zero set f−1(0). If f is non-singular, i.e., f has no critical zeros, then its nodal line is a smooth curve with no self-intersections. An important descriptor of f is its nodal length, i.e., the length of f−1(0), receiving much attention in the last couple of decades, in particular, concerning the nodal length of the eigenfunctions of Laplacian Δ on , in the high-energy limit.
Let be the Laplace eigenfunctions on , with energies λj in increasing order counted with multiplicity, i.e.,
endowed with the Dirichlet boundary conditions in the presence of nontrivial boundary. In this context, Yau’s conjecture asserts that the nodal length of ϕj is commensurable with , in the sense that
B. (Boundary-adapted) random wave model
In his highly influential work,5 Berry proposed to compare the high-energy Laplace eigenfunctions on generic chaotic surfaces and their nodal lines to random monochromatic waves and their nodal lines, respectively. The random monochromatic waves (also called Berry’s “random wave model” or RWM) are a centered isotropic Gaussian random field prescribed uniquely by the covariance function
with and J0(·) being the Bessel J function.
Let
be the zero density, also called the “first intensity” function of u, with ϕu(x) being the probability density function of the random variable u(x). In this isotropic case, it is easy to directly evaluate
and then appeal to the Kac–Rice formula, valid under the easily verified non-degeneracy conditions on the random field u, to evaluate the expected nodal length of u(·) restricted to a radius-R disk to be precisely
Berry6 found that as R → ∞, the variance satisfies the asymptotic law
much smaller than the a priori heuristic prediction made based on the natural scaling of the problem, due to what is now known as “Berry’s cancellation”34 of the leading non-oscillatory term of the two-point correlation function (also known as the “second zero intensity”).
Furthermore, in the same work,6 Berry studied the effect induced on the nodal length of eigenfunctions, satisfying the Dirichlet condition on a nontrivial boundary, both in its vicinity and far away from it. With the (infinite) horizontal axis serving as a model for the boundary, he introduced a Gaussian random field of boundary-adapted (non-stationary) monochromatic random waves, forced to vanish at x2 = 0. Formally, v(x1, x2) is the limit, as J → ∞, of the superposition
of J plane waves of wavenumber 1 forced to vanish at x2 = 0. Alternatively, v is the centered Gaussian random field prescribed by the covariance function
where x = (x1, x2), y = (y1, y2), and is the mirror symmetry of y; the law of v is invariant with respect to horizontal shifts,
, but not the vertical shifts.
By comparing (1.2) to (1.7), we observe that far away from the boundary (i.e., x2, y2 → ∞), rv(x, y) ≈ J0(‖x − y‖) so that in that range, the (covariance of) boundary-adapted waves converge to the (covariance of) isotropic ones (1.2), although the decay of the error term in this approximation is slow and of oscillatory nature. Intuitively, it means that, at infinity, the boundary has a small impact on the random waves, although it takes its toll on the nodal bias, as was demonstrated by Berry, as follows.
Let be the zero density of v, defined analogously to (1.3), depending on the height x2 only, independent of x1 by the inherent invariance (1.8). Berry showed that as x2 → 0,
and attributed this “nodal deficiency” , relative to (1.4), to the a.s. orthogonality of the nodal lines touching the boundary (Ref. 13, Theorem 2.5). Though a significant proportion of the details of the computation were omitted, we validated Berry's assertions for ourselves.
Furthermore, as x2 → ∞,
with some prescribed error term E(·). Here E(x2) is of order , so not smaller by magnitude than , but of oscillatory nature, and will not contribute to the Kac-Rice integral along expanding domains, as neither the term . In this situation, a natural choice for expanding domains is the rectangles , R → ∞ (say). As an application of the Kac–Rice formula (1.5), in this case, it easily follows that
i.e., a logarithmic “nodal deficiency” relative to (1.5), impacted by the boundary infinitely many wavelengths away from it. The logarithmic fluctuations (1.6) in the isotropic case u, possibly also holding for v, give rise to a hope to be able to detect the said, also logarithmic, negative boundary impact (1.11) via a single sample of the nodal length or, at least, very few ones.
C. Random spherical harmonics
The (unit) sphere is one of but few surfaces, where the solutions to the Helmholtz equation (1.1) admit an explicit solution. For a number , the space of solutions of (1.1) with λ = ℓ(ℓ + 1) is the (2ℓ + 1)-dimensional space of degree-ℓ spherical harmonics, and conversely, all solutions to (1.1) are spherical harmonics of some degree ℓ ≥ 0. Given ℓ ≥ 0, let be any L2-orthonormal basis of the space of spherical harmonics of degree ℓ. The random field
with ak being i.i.d. standard Gaussian random variables, is the degree-ℓ random spherical harmonics.
The law of is invariant with respect to the chosen orthonormal basis , uniquely defined via the covariance function
with Pℓ(·) being the Legendre polynomial of degree ℓ, and d(·, ·) is the spherical distance between . The random fields are the Fourier components in the L2-expansion of every isotropic random field,24 of interest, for instance, in cosmology and the study of Cosmic Microwave Background (CMB) radiation.
Let be the total nodal length of , of high interest for various pure and applied disciplines, including the above. Berard4 evaluated the expected nodal length to be precisely
and as ℓ → ∞, its variance is asymptotic34 to
in accordance with Berry’s result (1.6), save for the scaling, and the invariance of the nodal lines with respect to the symmetry x ↦ −x of the sphere, resulting in a doubled leading constant in (1.15) relative to (1.6) suitably scaled. A more recent proof25 of the central limit theorem for , asserting the asymptotic Gaussianity of
D. Principal results: Nodal bias for the hemisphere, at the boundary, and far away
Our principal results concern the hemisphere , endowed with the Dirichlet boundary conditions along the equator. We will widely use the spherical coordinates
with the equator identified with . Here, all the Laplace eigenfunctions are necessarily spherical harmonics restricted to , subject to some extra properties. Recall that a concrete (complex-valued) orthonormal basis of degree ℓ is the Laplace spherical harmonics , given in the spherical coordinates by
for m ≥ 0 and for m < 0, with being the associated Legendre polynomials of degree ℓ on order m. For ℓ ≥ 0 and |m| ≤ ℓ, the spherical harmonic Yℓ,m obeys the Dirichlet boundary condition on the equator if and only if m ≢ ℓ mod 2, spanning a subspace of dimension ℓ inside the (2ℓ + 1)-dimensional space of spherical harmonics of degree ℓ (Ref. 17, Example 4). [Its (ℓ + 1)-dimensional orthogonal complement is the subspace, satisfying the Neumann boundary condition.] Conversely, every Laplace eigenfunction on is necessarily a spherical harmonic of some degree ℓ ≥ 0 that is a linear combination of Yℓ,m with m ≢ ℓ mod 2.
The principal results of this paper concern the following model of boundary-adapted random spherical harmonics:
where aℓ,m are the standard (complex-valued) Gaussian random variables subject to the constraint so that Tℓ(·) is real-valued. Our immediate concern is for the law of Tℓ, which, as for any centered Gaussian random field, is uniquely determined by its covariance function, claimed by the following proposition.
It is evident, either from the definition or the covariance, that the law of Tℓ is invariant with respect to rotations of around the axis orthogonal to the equator, that is, in the spherical coordinates,
ϕ ∈ [0, 2π). The boundary impact of (1.17) relative to (1.13) is in perfect harmony with the boundary impact of the covariance (1.7) of Berry’s boundary-adapted model relative to the isotropic case (1.2), except that the mirror symmetry relative to the x axis in the Euclidean situation is substituted by mirror symmetry relative to the equator for the spherical geometry. These generalize to two dimensions the boundary impact on the ensemble of stationary random trigonometric polynomials on the circle,16,32 resulting in the ensemble of non-stationary random trigonometric polynomials vanishing at the end points.1,15
Let
be the zero density of Tℓ, which, unlike the rotation invariant spherical harmonics (1.12), genuinely depends on . More precisely, by the said invariance with respect to (1.18), the zero density K1,ℓ(x) depends on the polar angle θ only. We rescale by introducing the variable
and, with a slight abuse of notation, write
Our principal result deals with the asymptotics of K1,ℓ(·), in two different regimes, in line with (1.9) and (1.10), respectively.
- For C > 0 sufficiently large, as ℓ → ∞, one has(1.21)uniformly for C < ψ < πℓ, with the constant involved in the `O'-notation absolute.
- For ℓ ≥ 1, one has the uniform asymptoticswith the constant involved in the `O'-notation absolute.(1.22)
Clearly, the statement (1.22) is asymptotic for ψ small only, otherwise yielding the mere bound K1,ℓ(ψ) = O(ℓ), which is easy. As a corollary to Theorem 1.2, one may evaluate the asymptotic law of the total expected nodal length of Tℓ and detect the negative logarithmic bias relative to (1.14), in full accordance with Berry’s result (1.11).
II. DISCUSSION
A. Toral eigenfunctions and spectral correlations
Another surface admitting explicit solution to the Helmholtz equation (1.1) is the standard torus . Here, the Laplace eigenfunctions with eigenvalue 4π2n all correspond to an integer n expressible as a sum of two squares and are given by a sum
over all lattice points lying on the radius- centered circle, where n is a sum of two squares with e(y)≔e2πiy, ⟨μ, x⟩ = μ1x1 + μ2x2, and . Following Oravecz, Rudnick, and Wigman,28 one endows the eigenspace of {fn} with a Gaussian probability measure with the coefficient aμ standard (complex-valued) i.i.d. Gaussian, save for , resulting in the ensemble of “arithmetic random waves.”
The expected nodal length of fn was computed29 to be
and the useful upper bound
was also asserted, with r2(n) being the number of lattice points lying on the radius- circle or, equivalently, the dimension of the eigenspace {fn} as in (2.1). A precise asymptotic law for was subsequently established,19 shown to fluctuate, depending on the angular distribution of the lattice points. A non-central non-universal limit theorem was asserted26 also depending on the angular distribution of the lattice points.
An instrumental key input to both the said asymptotic variance and the limit law was Bourgain’s first nontrivial upper bound (Ref. 19, Theorem 2.2) of for the number of length-six spectral correlations, i.e., six-tuples of lattice points {μ: ‖μ‖2 = n} summing up to 0. Bourgain’s bound was subsequently improved and generalized to higher order correlations,7 in various degrees of generality, conditionally or unconditionally. These results are still actively used within the subsequent and ongoing research, in particular,8 and its followers.
B. Boundary impact
It makes sense to compare the torus to the square with Dirichlet boundary and test what kind of impact it would have relative to (2.2) on the expected nodal length, as the “boundary-adapted arithmetic random waves,” which were addressed in Ref. 11. It was concluded, building on the work of Bourgain and Bombieri,7 and by appealing to a different notion of spectral correlation, namely, the spectral semi-correlations, that, even at the level of expectation, the total nodal bias is fluctuating from nodal deficiency (negative bias) to nodal surplus (positive bias) depending on the angular distribution of the lattice points and its interaction with the direction of the square boundary, at least, for generic energy levels. A similar experiment conducted by Gnutzmann and Lois for cuboids of arbitrary dimensions, averaging for eigenfunctions admitting separation of variables belonging to different eigenspaces, revealed consistency with Berry’s nodal deficiency ansatz stemming from (1.11).
It would be useful to test whether different Gaussian random fields on the square would result in different limiting nodal biases around the boundary corresponding to (1.22), which is likely to bring in a different notion of spectral correlation, not unlikely “quasi-semi-correlation.”3,18 Another question of interest is “de-randomize” any of these results, i.e., infer the corresponding results on deterministic eigenfunctions following the work of Bourgain.8 We leave all these to be addressed elsewhere.
III. JOINT DISTRIBUTION OF (fn(x), ∇fn(x))
In the analysis of K1,ℓ(x), we naturally encounter the distribution of Tℓ(x), determined by
and the distribution of ∇Tℓ(x) conditioned on Tℓ(x) = 0, determined by its 2 × 2 covariance matrix
Let x correspond to the spherical coordinates (θ, ϕ). An explicit computation shows that the covariance matrix Ωℓ(x) depends only on θ, and below, we will often abuse a notation to write Ωℓ(θ) instead, and also, when convenient, Ωℓ(ψ) with ψ as in (1.20). A direct computation shows the following:
In Secs. III A and III B, we prove Lemma 3.1, that is, we evaluate the 2 × 2 covariance matrix of ∇Tℓ(x) conditioned upon Tℓ(x) = 0. First, in Sec. III A, we evaluate the unconditional 3 × 3 covariance matrix Σℓ(x) of (Tℓ(x), ∇Tℓ(x)), and then, in Sec. III B, we apply the standard procedure for conditioning multivariate Gaussian random variables.
A. The unconditional covariance matrix
The covariance matrix of
could be expressed as
where
The 1 × 2 matrix Bℓ(x) is
where Bℓ(x) depends only on θ, and by an abuse of notation, we write
The entries of the 2 × 2 matrix Cℓ(x) are
where again recalling that x = (θ, ϕ), we write
B. Conditional covariance matrix
The conditional covariance matrix of the Gaussian vector (∇Tℓ(x)|Tℓ(x) = 0) is given by the following standard Gaussian transition formula:
Again taking x = (θ, ϕ) and observing that
and
we have
which is the statement of Lemma 3.1.
IV. PROOF OF THEOREM 1.2(1): PERTURBATIVE ANALYSIS AWAY FROM THE BOUNDARY
A. Perturbative analysis
The asymptotic analysis (1.21) is in two steps. First, we evaluate the variance Var(Tℓ(x)) and each entry in Sℓ(x) using the high degree asymptotics of the Legendre polynomials and its derivatives (Hilb’s asymptotics). In the second step, performed within Proposition 4.3, we exploit the analyticity of the Gaussian expectation (1.19) as a function of the entries of the corresponding non-singular covariance matrix, to Taylor expand K1,ℓ(x) where both Var(Tℓ(x)) − 1 and the entries of Sℓ(x) are assumed to be small.
For a Proof of Lemmas 4.1 and 4.2, we refer to Ref. 30, Theorem 8.21.6 and Ref. 20, Sec. 5.11, respectively.
Recall the scaled variable ψ related to θ via (1.20) so that an application of Lemmas 4.1 and 4.2 yields that for ℓ ≥ 1 and C < ψ < ℓπ,
Observing that
we write
A repeated application of Lemmas 4.1 and 4.2 also yields an asymptotic estimate for the first couple of derivatives of the Legendre polynomials (Ref. 12, Lemma 9.3),
and
Since we have that
we have
and observing that
we obtain
The estimates in (4.1), (4.2), and (4.3) imply that for ℓ ≥ 1 and uniformly for C < ψ < ℓπ, with C > 0, we have
With the same abuse of notation as above, we write Sℓ(ψ) ≔ Sℓ(x) as in Lemma 3.1 and in analogous manner for its individual entries Sij;ℓ(ψ) ≔ S11;ℓ(x). We have
The next proposition prescribes a precise asymptotic expression for the density function K1,ℓ(·) via a Taylor expansion of the relevant Gaussian expectation as a function of the associated covariance matrix entries.
B. Proof of Theorem 1.2(1)
V. PROOF OF THEOREM 1.2(2): PERTURBATIVE ANALYSIS AT THE BOUNDARY
The aim of this section is to study the asymptotic behavior of the density function K1,ℓ(ψ) for 0 < ψ < ϵ0, with ϵ0 > 0 sufficiently small. We have
where Δℓ(ψ) is the scaled conditional covariance matrix
We have that
with constant involved in the `O'-notation absolute. We also have
and Cℓ(ψ) is the 2 × 2 symmetric matrix with entries
We obtain that
with
and
We introduce the change in variable , and we write
which is (1.22).
VI. PROOF OF COROLLARY 1.3: EXPECTED NODAL LENGTH
A. Kac–Rice formula for expected nodal length
The Kac–Rice formula is a meta-theorem, allowing one to evaluate the moments of the zero set of a random field satisfying some smoothness and non-degeneracy conditions. For , a sufficiently smooth centered Gaussian random field, we define
the zero density (first intensity) of F. Then, the Kac–Rice formula asserts that for some suitable class of random fields F and , a compact closed subdomain of , one has the equality
We would like to apply (6.1) to the boundary-adapted random spherical harmonics Tℓ to evaluate the asymptotic law of the total expected nodal length of Tℓ. Unfortunately, the aforementioned non-degeneracy conditions fail at the equator,
Nevertheless, in a manner inspired by Ref. 11, Proposition 2.1, we excise a small neighborhood of this degenerate set and apply the monotone convergence theorem so to be able to prove that (6.1) holds precisely, save for the length of the equator that is bound to be contained in the nodal set of Tℓ, by the Dirichlet boundary condition.
To justify the Kac–Rice formula outside the equator, we use (Ref. 2, Theorem 6.8), which assumes the non-degeneracy of the 3 × 3 covariance matrix at all these points, a condition we were able to verify via an explicit computation, presented within Appendix B.
B. Expected nodal length
ACKNOWLEDGMENTS
We are grateful to Zeév Rudnick for raising the question addressed within this manuscript and to Mikhail Sodin for pointing out the necessity of validating the non-singularity of the covariance matrix to justify the use of Kac–Rice formula. V.C. has received funding from the Istituto Nazionale di Alta Matematica (INdAM) through the GNAMPA Research Project 2020 “Geometria stocastica e campi aleatori.” D.M. is supported by the MIUR Departments of Excellence Program Math@Tov. This paper is dedicated to the memory of Jean Bourgain.
DATA AVAILABILITY
The data that support the findings of this study are available within the article.
APPENDIX A: PROOF OF PROPOSITION 1.1
We have that
where we have used the fact that Yℓ,m(θ, ϕ) = (−1)ℓ+mYℓ,m(π − θ, ϕ). We apply now the addition theorem for spherical harmonics,
so that
APPENDIX B: NON-SINGULARITY OF THE COVARIANCE MATRIX
Given the block-diagonal form of the covariance matrix Σℓ(x), it is sufficient to show that the vector is non-degenerate for every , and we assume with no loss of generality that x is of the form . Using the definition of the spherical harmonics basis {Yℓ,m}, we have
where is an array standard Gaussian variable N(0, 1). Assume, by contradiction, that the covariance matrix is singular for some so that a.s. , with the (deterministic) constant c ≠ 0. Assume at first that ℓ is odd so that m is even, and recall that the associated Legendre polynomials are defined by
i.e., for m, even they are polynomials in t ∈ [−1, 1] of degree ℓ.
Note that
is a vector of linearly independent polynomials of degree ℓ, with only odd monomials; hence there exist a full rank matrix Q of dimension such that
Now, take
which is a centered Gaussian vector with the covariance matrix QtQ, hence of full rank. After this transformation of basis, we may write
Hence, after the change in variable cos θ = t, we have
Then, for and to be linearly dependent, there should exist a value such that for ,
More explicitly, it follows that almost surely
for all k = 1, 3, …, ℓ, which is clearly absurd, since the random variables are the components of a full rank Gaussian vector (which belongs to a proper linear subspace with probability zero). The case where ℓ is symmetric, the only difference being that the functions are polynomials after multiplication by sin θ. In light of all the above, the variance matrix of the vector (Tℓ, ∇Tℓ) is nonsingular for all , as claimed.