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.

The nodal line of a smooth function f:MR, defined on a smooth compact surface M, 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 M, in the high-energy limit.

Let (ϕj,λj)j1 be the Laplace eigenfunctions on M, with energies λj in increasing order counted with multiplicity, i.e.,

Δϕj+λjϕj=0,
(1.1)

endowed with the Dirichlet boundary conditions ϕ|M0 in the presence of nontrivial boundary. In this context, Yau’s conjecture asserts that the nodal length L(ϕj) of ϕj is commensurable with λj, in the sense that

cMλjL(ϕj)CMλj,

with some constants CM>cM>0. Yau’s conjecture was resolved for M analytic,9,10,14 and, more recently, a lower bound22 and a polynomial upper bound21,23 were asserted in full generality (i.e., for M smooth).

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 u:R2R prescribed uniquely by the covariance function

E[u(x)u(y)]=J0(xy),
(1.2)

with x,yR2 and J0(·) being the Bessel J function.

Let

K1u(x)=ϕu(x)(0)E[u(x)u(x)=0]
(1.3)

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

K1u(x)122
(1.4)

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 L(u;R) of u(·) restricted to a radius-R disk B(R)R2 to be precisely

E[L(u;R)]=B(R)K1u(x)dx=122Area(B(R)).
(1.5)

Berry6 found that as R, the variance (L(u;R)) satisfies the asymptotic law

Var(L(u;R))=1256R2logR+O(R2),
(1.6)

much smaller than the a priori heuristic prediction Var(L(u;R))R3 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 {(x1,x2):x2=0}R2 serving as a model for the boundary, he introduced a Gaussian random field v(x1,x2):R×R>0R 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

2Jj=1Jsin(x2sin(θj))cos(x1cos(θj)+ϕj)

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

rv(x,y)E[v(x)v(y)]=J0(xy)J0(xỹ),
(1.7)

where x = (x1, x2), y = (y1, y2), and ỹ=(y1,y2) is the mirror symmetry of y; the law of v is invariant with respect to horizontal shifts,

v(,)v(a+,),
(1.8)

aR, 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(‖xy‖) 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 K1v(x)=K1v(x2) 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,

K1v(x2)12π
(1.9)

and attributed this “nodal deficiency” 12π<122, 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,

K1v(x2)=1221+cos(2x2π/4)πx2132πx2+E(x2),
(1.10)

with some prescribed error term E(·). Here E(x2) is of order 1|x2|, so not smaller by magnitude than 132πx2, but of oscillatory nature, and will not contribute to the Kac-Rice integral along expanding domains, as neither the term cos(2x2π/4)πx2. In this situation, a natural choice for expanding domains is the rectangles DR[1,1]×[0,R], R (say). As an application of the Kac–Rice formula (1.5), in this case, it easily follows that

E[L(v;DR)]=122Area(DR)1322πlogR+O(1),
(1.11)

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.

The (unit) sphere M=S2 is one of but few surfaces, where the solutions to the Helmholtz equation (1.1) admit an explicit solution. For a number Z0, the space of solutions of (1.1) with λ = ( + 1) is the (2 + 1)-dimensional space of degree- spherical harmonics, and conversely, all solutions to (1.1) are spherical harmonics of some degree ≥ 0. Given ≥ 0, let E{η,1,η,2+1} be any L2-orthonormal basis of the space of spherical harmonics of degree . The random field

T̃(x)=4π2+1k=12+1akη,k(x),
(1.12)

with ak being i.i.d. standard Gaussian random variables, is the degree- random spherical harmonics.

The law of T̃ is invariant with respect to the chosen orthonormal basis E, uniquely defined via the covariance function

E[T̃(x)T̃(y)]=P(cosd(x,y)),
(1.13)

with P(·) being the Legendre polynomial of degree , and d(·, ·) is the spherical distance between x,yS2. The random fields {T̃} 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 L(T̃) be the total nodal length of T̃, of high interest for various pure and applied disciplines, including the above. Berard4 evaluated the expected nodal length to be precisely

E[L(T̃)]=2π(+1),
(1.14)

and as , its variance is asymptotic34 to

Var(L(T̃))132log,
(1.15)

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 L(T̃), asserting the asymptotic Gaussianity of

L(T̃)E[L(T̃)]132log,

is sufficiently robust to also yield the central limit theorem, as R for the nodal length L(u;R) of Berry’s random waves, as was recently demonstrated,31 also claimed by Ref. 27.

Our principal results concern the hemisphere H2S2, endowed with the Dirichlet boundary conditions along the equator. We will widely use the spherical coordinates

H2={(θ,ϕ):θ[0,π/2],ϕ[0,2π)},

with the equator identified with {θ=π/2}H2. Here, all the Laplace eigenfunctions are necessarily spherical harmonics restricted to H2, subject to some extra properties. Recall that a concrete (complex-valued) orthonormal basis of degree is the Laplace spherical harmonics {Y,m}m=, given in the spherical coordinates by

Y,m(θ,ϕ)=γmPm(cosθ)eimϕ,γm2+14π(m)!(+m)!

for m ≥ 0 and Y,m(θ,ϕ)=(1)mY,m*(θ,ϕ) for m < 0, with Pm() being the associated Legendre polynomials of degree on order m. For ≥ 0 and |m| ≤ , the spherical harmonic Y,m obeys the Dirichlet boundary condition on the equator if and only if m mod 2, spanning a subspace of dimension inside the (2 + 1)-dimensional space of spherical harmonics of degree (Ref. 17, Example 4). [Its ( + 1)-dimensional orthogonal complement is the subspace, satisfying the Neumann boundary condition.] Conversely, every Laplace eigenfunction on H2 is necessarily a spherical harmonic of some degree ≥ 0 that is a linear combination of Y,m with m mod 2.

The principal results of this paper concern the following model of boundary-adapted random spherical harmonics:

T(x)=8π2+1m=m mod 2a,mY,m(x),
(1.16)

where a,m are the standard (complex-valued) Gaussian random variables subject to the constraint a,m=a,m̄ so that T(·) is real-valued. Our immediate concern is for the law of T, which, as for any centered Gaussian random field, is uniquely determined by its covariance function, claimed by the following proposition.

Proposition 1.1.
The covariance function ofTas in (1.16) is given by
r(x,y)E[T(x)T(y)]=P(cosd(x,y))P(cosd(x,ȳ)),
(1.17)
whereȳis the mirror symmetry ofyaround the equator, i.e.,y=(θ,ϕ)ȳ=(πθ,ϕ)in the spherical coordinates.

It is evident, either from the definition or the covariance, that the law of T is invariant with respect to rotations of H2 around the axis orthogonal to the equator, that is, in the spherical coordinates,

T(θ,ϕ)T(θ,ϕ+ϕ0),
(1.18)

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 yỹ relative to the x axis in the Euclidean situation is substituted by mirror symmetry yȳ 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

K1,(x)=12πVar(T(x))ET(x)T(x)=0
(1.19)

be the zero density of T, which, unlike the rotation invariant spherical harmonics (1.12), genuinely depends on xH2. More precisely, by the said invariance with respect to (1.18), the zero density K1,(x) depends on the polar angle θ only. We rescale by introducing the variable

ψ=(π2θ)
(1.20)

and, with a slight abuse of notation, write

K1,(ψ)=K1,(x).

Our principal result deals with the asymptotics of K1,(·), in two different regimes, in line with (1.9) and (1.10), respectively.

Theorem 1.2.

  1. ForC > 0 sufficiently large, as, one has
    K1,(ψ)=(+1)221+2π1ψcos{(+1/2)ψ/π/4}116πψ
    (1.21)
    +1516πψcos{(+1/2)2ψ/π/2}+O(ψ3/22)
    uniformly forC < ψ < πℓ, with the constant involved in the `O'-notation absolute.
  2. For ≥ 1, one has the uniform asymptotics
    K1,(ψ)=2π1+O(1)+O(ψ2),
    (1.22)
    with the constant involved in the `O'-notation absolute.

Clearly, the statement (1.22) is asymptotic for ψ small only, otherwise yielding the mere bound K1,(ψ) = O(), which is easy. As a corollary to Theorem 1.2, one may evaluate the asymptotic law of the total expected nodal length of T and detect the negative logarithmic bias relative to (1.14), in full accordance with Berry’s result (1.11).

Corollary 1.3.
As, the expected nodal length has the following asymptotics:
E[L(T)]=2π(+1)221322log()+O(1).

Another surface admitting explicit solution to the Helmholtz equation (1.1) is the standard torus T2=R2/Z2. 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

fn(x)=μ2=naμe(μ,x)
(2.1)

over all lattice points μ=(μ1,μ2)Z2 lying on the radius-n centered circle, where n is a sum of two squares with e(y)≔e2πiy, ⟨μ, x⟩ = μ1x1μ2x2, and x=(x1,x2)T2. 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 aμ=aμ̄, resulting in the ensemble of “arithmetic random waves.”

The expected nodal length of fn was computed29 to be

E[L(fn)]=2π2n,
(2.2)

and the useful upper bound

Var(L(fn))nr2(n)

was also asserted, with r2(n) being the number of lattice points lying on the radius-n circle or, equivalently, the dimension of the eigenspace {fn} as in (2.1). A precise asymptotic law for Var(L(fn)) 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 or2(n)r2(n)4 for the number of length-six spectral correlations, i.e., six-tuples of lattice points {μ: ‖μ2 = 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.

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.

In the analysis of K1,(x), we naturally encounter the distribution of T(x), determined by

Var(T(x))=1P(cosd(x,x̄)),

and the distribution of ∇T(x) conditioned on T(x) = 0, determined by its 2 × 2 covariance matrix

Ω(x)=E[T(x)tT(x)|T(x)=0].

Let x correspond to the spherical coordinates (θ, ϕ). An explicit computation shows that the covariance matrix Ω(x) depends only on θ, and below, we will often abuse a notation to write Ω(θ) instead, and also, when convenient, Ω(ψ) with ψ as in (1.20). A direct computation shows the following:

Lemma 3.1.
The 2 × 2 covariance matrix ofT(x) conditioned onT(x) = 0 is the following real symmetric matrix:
Ω(x)=(+1)2I2+S(x),
(3.1)
where
S(x)=S11,(x)00S22,(x),
and forx = (θ, ϕ),
S11,(x)=2(+1)cos(2θ)P(cos(π2θ))+sin2(2θ)P(cos(π2θ))+11P(cos(π2θ))sin2(2θ)[P(cos(π2θ))]2,S22,(x)=2(+1)P(cos(π2θ)).

In Secs. III A and III B, we prove Lemma 3.1, that is, we evaluate the 2 × 2 covariance matrix of ∇T(x) conditioned upon T(x) = 0. First, in Sec. III A, we evaluate the unconditional 3 × 3 covariance matrix Σ(x) of (T(x), ∇T(x)), and then, in Sec. III B, we apply the standard procedure for conditioning multivariate Gaussian random variables.

The covariance matrix of

(T(x),T(x))

could be expressed as

Σ(x)=A(x)B(x)Bt(x)C(x),

where

A(x)=Var(T(x)),B(x)=E[T(x)yT(y)]x=y,C(x)=E[xT(x)yT(y)]x=y.

The 1 × 2 matrix B(x) is

B(x)=B,1(x)B,2(x),

where B(x) depends only on θ, and by an abuse of notation, we write

B,1(x)=θyr(x,y)x=y=sin(2θ)P(cos(π2θ)),B,2(x)=1sinθyϕyr(x,y)x=y=0.

The entries of the 2 × 2 matrix C(x) are

C(x)=C,11(x)C,12(x)C,21(x)C,22(x),

where again recalling that x = (θ, ϕ), we write

C,11(x)=θxθyr(x,y)x=y=P(1)cos(2θ)P(cos(π2θ))sin2(2θ)P(cos(π2θ)),C,12(x)=C,21(x)=1sinθyϕyθxr(x,y)x=y=0,C,22(x)=1sinθyϕy1sinθxϕxr(x,y)x=y=P(1)P(cos(π2θ)).

The conditional covariance matrix of the Gaussian vector (∇T(x)|T(x) = 0) is given by the following standard Gaussian transition formula:

Ω(x)=C(x)1Var(T(x))Bt(x)B(x).
(3.2)

Again taking x = (θ, ϕ) and observing that

Bt(x)B(x)Var(T(x))=11P(cos(π2θ))sin2(2θ)[P(cos(π2θ))]2000

and

P(1)=(+1)2,

we have

Ω(x)=(+1)2I2cos(2θ)P(cos(π2θ))+sin2(2θ)P(cos(π2θ))00P(cos(π2θ))11P(cos(π2θ))sin2(2θ)[P(cos(π2θ))]2000,

which is the statement of Lemma 3.1.

The asymptotic analysis (1.21) is in two steps. First, we evaluate the variance Var(T(x)) and each entry in S(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,(x) where both Var(T(x)) − 1 and the entries of S(x) are assumed to be small.

Lemma 4.1
(Hilb’s asymptotics).
P(cosφ)=φsinφ1/2J0((+1/2)φ)+δ(φ)
uniformly for 0 ≤ φπɛ, whereJ0is the Bessel function of the first kind. For the error term, we have the bounds
δ(φ)φ2O(1),0<φC/φ1/2O(3/2),C/φπε,
whereCis a fixed positive constant and the constants involved in theO-notation depend onConly.

Lemma 4.2.
The following asymptotic representation for the Bessel functions of the first kind holds:
J0(x)=2πx1/2cos(xπ/4)k=0(1)kg(2k)(2x)2k+2πx1/2cos(x+π/4)k=0(1)kg(2k+1)(2x)2k1,
whereɛ > 0, |arg x| ≤ πɛ,g(0) = 1, andg(k)=(1)(32)((2k1)2)22kk!=(1)k[(2k)!!]222kk!.

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 ≥ 1 and C < ψ < ℓπ,

P(cos(ψ/))=2π1/2sin1/2(ψ/)cos((+1/2)ψ/π/4)12ψ/cos((+1/2)ψ/+π/4)+O((ψ/)1/23/2).

Observing that

1/2sin1/2(ψ/)=1/21ψ/+O((ψ/)32)=1ψ+O(ψ3/22),1/2sin1/2(ψ/)12ψ=O(ψ3/2),

we write

P(cos(ψ/))=2π1ψcos((+1/2)ψ/π/4)+O(ψ3/2)+O(ψ3/22).
(4.1)

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),

P(cos(ψ/))=2π11/2sin1+1/2(ψ/)sin((+1/2)ψ/π/4)18ψ/sin((+1/2)ψ/+π/4)+O(12(ψ/)52)

and

P(cos(ψ/))=2π21/2sin2+1/2(ψ/)cos((+1/2)ψ/π/4)+18ψ/cos((+1/2)ψ/+π/4)2π11/2sin3+1/2(ψ/)cos((1+1/2)ψ/+π/4)+18ψ/cos((1+1/2)ψ/π/4)+O(ψ7/24).

Since we have that

11/2sin1+1/2(ψ/)=11/21(ψ/)3/2+O((ψ/)1/2)=2ψ3/2+O(ψ1/2),11/2sin1+1/2(ψ/)1ψ=O(ψ5/22),

we have

P(cos(ψ/))=2π11/2sin1+1/2(ψ/)sin((+1/2)ψ/π/4)+O(ψ5/22),
(4.2)

and observing that

21/2sin2+1/2(ψ/)=21/21(ψ/)5/2+O((ψ/)1/2)=4ψ5/2+(ψ1/22),21/2sin2+1/2(ψ/)1ψ=O(ψ7/24),11/2sin3+1/2(ψ/)=11/21(ψ/)7/2+O((ψ/)3/2)=4ψ7/2+O(ψ3/22),

we obtain

P(cos(ψ/))=2π21/2sin2+1/2(ψ/)cos((+1/2)ψ/π/4)+O(ψ3/22)+O(ψ7/24).
(4.3)

The estimates in (4.1), (4.2), and (4.3) imply that for ≥ 1 and uniformly for C < ψ < ℓπ, with C > 0, we have

P(cos(ψ/))=2π1ψcos((+1/2)ψ/π/4)+O(ψ3/2)+O(ψ3/22),{P(cos(ψ/))}2=2π1ψcos2((+1/2)ψ/π/4)+O(ψ2)+O(ψ2).
(4.4)

With the same abuse of notation as above, we write S(ψ) ≔ S(x) as in Lemma 3.1 and in analogous manner for its individual entries Sij;(ψ) ≔ S11;(x). We have

S11;(ψ)=22π1ψcos((+1/2)ψ/π/4)4π1ψsin2((+1/2)ψ/π/4)
(4.5)
+O(ψ3/2)+O(ψ3/22),S22;(ψ)=22π1ψ3/2sin((+1/2)ψ/π/4)+O(ψ1/22)+O(ψ5/2).
(4.6)

The next proposition prescribes a precise asymptotic expression for the density function K1,(·) via a Taylor expansion of the relevant Gaussian expectation as a function of the associated covariance matrix entries.

Proposition 4.3.
ForC > 0 sufficiently large, we have the following expansion onC < ψ < ℓπ:
K1,(ψ)=(+1)22+L(ψ)+E(ψ),
(4.7)
with the leading term
L(ψ)=(+1)42s(ψ)+12trS(ψ)+34s2(ψ)+14s(ψ)trS(ψ)116trS2(ψ)132(trS(ψ))2,
wheres(ψ) = P(cos(ψ/)) and the error termE(ψ) is bounded by
|E(ψ)|=O((|s(ψ)|3+|S(ψ)|3)),
with constant involved in theO-notation absolute.

Proof.
To prove Proposition 4.3, we perform a precise Taylor analysis for the density function K1,(ψ), assuming that both s(ψ) and the entries of S(ψ) are small. We introduce the scaled covariance matrix [see (3.1)],
Δ(ψ)=2(+1)Ω(ψ)=I2+S(ψ).
The density function K1,(·) could be expressed as
K1,(ψ)=12π11s(ψ)12πdetΔ(ψ)(+1)2R2zexp12zΔ1(ψ)ztdz.
On (C, πℓ), with C sufficiently large, we Taylor expand
11s(ψ)=1+12s(ψ)+38s2(ψ)+O(s3(ψ))
since, using the high degree asymptotics of the Legendre polynomials (Hilb’s asymptotics), we see that |P(cos(ψ/))| is bounded away from 1. Next, we consider the Gaussian integral
I(S(ψ))=R2zexp12z(I2+S(ψ))1ztdz,
observing that on (C, πℓ), for C sufficiently large, we can Taylor expand
(I2+S(ψ))1=I2S(ψ)+S2(ψ)+O(S3(ψ))
and the exponential as follow:
exp12z(I2+S(ψ))1zt=expzzt21+12zS(ψ)S2(ψ)+O(S3(ψ))zt+1212z(S(ψ)S2(ψ)+O(S3(ψ)))zt2+Oz(S(ψ)S2(ψ)+O(S3(ψ)))zt3
so that
I(S(ψ))=R2zexpzzt21+12zS(ψ)zt12zS2(ψ)zt+18zS(ψ)zt2dz+O(S3(ψ)).
We introduce the following notations:
I0(S(ψ))=R2zexpzzt2dz=2π0ρexp12ρ2ρdρ=2ππ2=2π3/2,
I1(S(ψ))=12R2zexpzzt2zS(ψ)ztdz=323/2π3/2trS(ψ),
and
I2(S(ψ))=12R2zexpzzt2zS2(ψ)ztdz=12R2zexpzzt2S11;2(ψ)z12+S22;2(ψ)z22dz=323/2π3/2trS(ψ)2.
We also define
I3(S(ψ))=18R2zexpzzt2zS(ψ)zt2dz=18R2zexpzzt2S11;2(ψ)z14+S22;2(ψ)z24+2S11;(ψ)S22;(ψ)z12z22dz
(4.8)
and note that
R2zexpzzt2(z12+z22)2dz=2π0ρexpρ22ρ4ρdρ=2152π3/2,R2zexpzzt2z14dz=15234π3/2
(4.9)
and that
R2zexpzzt2z12z22dz=12R2zexp12zzt(z12+z22)2dzR2zexp12zztz14dz=15214π3/2.
(4.10)
Substituting (4.9) and (4.10) into (4.8), we obtain
I3(S(ψ))=181542π3/23S11;2(ψ)+3S22;2(ψ)+2S11;(ψ)S22;(ψ)=15264π3/2{2trS2(ψ)+[trS(ψ)]2}.
Write
I(S(ψ))=I0(S(ψ))+I1(S(ψ))+I2(S(ψ))+I3(S(ψ))+O(S3(ψ))=2π3/2+323/2π3/2trS(ψ)9162π3/2trS2(ψ)+15264π3/2[trS(ψ)]2+O(S3(ψ)).
We finally expand
1detΔ(ψ)=1det(I2+S(ψ))
and note that
det(I2+S(ψ))=[1+S11;(ψ)][1+S22;(ψ)]=1+trS(ψ)+detS(ψ),
and so,
1detΔ(ψ)=112trS(ψ)+detS(ψ)+38trS(ψ)+detS(ψ)2+O(S3(ψ))=112trS(ψ)12detS(ψ)+38[trS(ψ)]2+O(S3(ψ))=112trS(ψ)+14trS2(ψ)+18[trS(ψ)]2+O(S3(ψ)),
where we have used the fact that S11;2(ψ) and S22;2(ψ) are the eigenvalues of S2(ψ), and we have written det S(ψ) as follows:
detS(ψ)=12[S11;(ψ)+S22;(ψ)]2[S11;2(ψ)+S22;2(ψ)]=12trS(ψ)2trS2(ψ).
In conclusion, we have
K1,(ψ)=(+1)22ππ1+12s(ψ)+38s2(ψ)+O(s3(ψ))×2π3/2+323/2π3/2trS(ψ)9162π3/2trS2(ψ)+15264π3/2[trS(ψ)]2+O(S3(ψ))×112trS(ψ)+14trS2(ψ)+18[trS(ψ)]2+O(S3(ψ))=(+1)2222+s(ψ)+12trS(ψ)+34s2(ψ)+14s(ψ)trS(ψ)116trS2(ψ)132[trS(ψ)]2+O(s3(ψ))+O(S3(ψ)).

Proof.
Substituting the estimates (4.4), (4.5), and (4.6) into (4.7), we obtain
K1,(ψ)=(+1)2222+22π1ψcos((+1/2)ψ/π/4)+74π1ψcos2((+1/2)ψ/π/4)2π1ψsin2((+1/2)ψ/π/4)+O(ψ3/22),
and since cos2(x)=12[1+cos(2x)] and sin2(x)=12[1cos(2x)], we can write
74πψcos2((+1/2)ψ/π/4)2πψsin2((+1/2)ψ/π/4)=74πψ12[1+cos((+1/2)2ψ/π/2)]2πψ12[1cos((+1/2)2ψ/π/2)]=74πψ122πψ12+74πψ12+2πψ12cos((+1/2)2ψ/π/2)=18πψ+158πψcos((+1/2)2ψ/π/2).
The above equation implies
K1,(ψ)=(+1)2222+22π1ψcos((+1/2)ψ/π/4)18πψ+158πψcos((+1/2)2ψ/π/2)+O(ψ3/22)=(+1)221+2π1ψcos((+1/2)ψ/π/4)116πψ+1516πψcos((+1/2)2ψ/π/2)+O(ψ3/22),
the statement (1.21) of Theorem 1.2(1).□

The aim of this section is to study the asymptotic behavior of the density function K1,(ψ) for 0 < ψ < ϵ0, with ϵ0 > 0 sufficiently small. We have

K1,(ψ)=12π1P(cos(ψ/))12πdetΔ(ψ)(+1)R2zexp12ztΔ1(ψ)zdz,

where Δ(ψ) is the scaled conditional covariance matrix

Δ(ψ)=C(ψ)Bt(ψ)B(ψ)1P(cos(ψ/)).

We have that

1P(cos(ψ/))=(+1)2ψ222(1)(+1)(+2)44ψ424+136(2)(1)(+1)(+2)(+3)6ψ626+O(ψ8),
(5.1)

with constant involved in the `O'-notation absolute. We also have

Bt(ψ)=sin(ψ/)P(cos(ψ/))10=(+1)2ψ2(1)(+1)(+2)24ψ323+112(2)(1)(+1)(+2)(+3)l6ψ525+O(ψ7)0,

and C(ψ) is the 2 × 2 symmetric matrix with entries

C,11(ψ)=P(1)+cos(ψ/)P(cos(ψ/))sin2(ψ/)P(cos(ψ/))12=134(1)(+1)(+2)4ψ222+524(2)(1)(+1)(+2)(+3)6ψ424+O(ψ6),C,12(ψ)=0,C,22(ψ)=P(1)P(cos(ψ/))12=(1)(+1)(+2)44ψ222(2)(1)(+1)(+2)(+3)246ψ424+O(ψ6).

We obtain that

Δ(ψ)=δ11,(ψ)00δ22,(ψ),

with

δ11,(ψ)=12832(2)(1)(+1)(+2)(+3)6ψ4+O(ψ6)=12832ψ4+O(1ψ4)+O(ψ6)
(5.2)

and

δ22,(ψ)=(1)(+1)(+2)44ψ222+O(ψ4)=ψ216+O(1ψ2)+O(ψ4).
(5.3)

We introduce the change in variable ξ=Δ1/2(ψ)z, and we write

K1,(ψ)=12π1P(cos(ψ/))12π(+1)R2δ11,(ψ)ξ12+δ22,(ψ)ξ22expξtξ2dξ.

Using the expansions in (5.1), (5.2), and (5.3), we write

K1,(ψ)=12πψ2/4+O(1ψ2)+O(ψ4)(+1)ψ4+O(1ψ)+O(ψ3)2π=(+1)12π+O(1)+O(ψ2),

which is (1.22).

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 F:RdR, a sufficiently smooth centered Gaussian random field, we define

K1,F(x)12πVar(F(x))E[|F(x)|F(x)=0],

the zero density (first intensity) of F. Then, the Kac–Rice formula asserts that for some suitable class of random fields F and D̄Rd, a compact closed subdomain of Rd, one has the equality

E[Vold1(F1(0)D̄)]=D̄K1,F(x)dx.
(6.1)

We would like to apply (6.1) to the boundary-adapted random spherical harmonics T to evaluate the asymptotic law of the total expected nodal length of T. Unfortunately, the aforementioned non-degeneracy conditions fail at the equator,

E={(θ,ϕ):θ=π/2}H2.

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 T, by the Dirichlet boundary condition.

Proposition 6.1.
The expected nodal length ofTsatisfies
E[L(T)]=H2K1,(x)dx+2π,
(6.2)
whereK1,(·) is the zero density ofT.

Proof.

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.

We construct a small neighborhood of the equator E, i.e., the set
Eε=(θ,ϕ):θπ2,π2ε,
and we denote
HεH2\Eε.
Since the Kac–Rice formula holds for T restricted to Hε, the expected nodal length for T restricted to Hε is
E[L(T|Hε)]=HεK1,(x)dx.
Since the restricted nodal length {L(T|Hε)}ε>0 is an increasing sequence of non-negative random variables with an a.s. limit
limε0L(T|Hε)=L(T)2π,
the monotone convergence theorem yields
limε0E[L(T|Hε)]=E[L(T)]2π.
(6.3)
Moreover, by the definition,
limε0HεK1,(x)dx=H2K1,(x)dx.
(6.4)
The equality of the limits in (6.3) and (6.4) shows that Proposition 6.1 holds.□

Proof of Corollary 1.3.
To analyze asymptotic behavior of the expected nodal length, we separate the contribution of the following three subregions of the hemisphere H2 in the Kac–Rice integral on the right-hand side of (6.2):
HC={(ψ,ϕ):0<ψ<ϵ0},HI={(ψ,ϕ):ϵ0<ψ<C},HF={(ψ,ϕ):C<ψ<π}.
Note that we express the three subregions of H2 in terms of the scaled variable ψ. In what follows, we argue that HF gives the main contribution.
In the (scaled) spherical coordinates, we may rewrite the Kac–Rice integral (6.2) as
E[L(T)]2π=π0πK1,(ψ)sinπ2ψ2dψ
and the contribution of the third range HF as
E[L(T|HF)]=πCπK1,(ψ)sinπ2ψ2dψ.
(6.5)
We are now going to invoke the asymptotics of K1,(ψ), prescribed by (1.21) for this range. The first term in (1.21) contributes
π(+1)22Cπsinπ2ψ2dψ=π(+1)2221sinC2=2π(+1)221C2+OC
(6.6)
to the integral (6.5).
The second term in (1.21) gives
π(+1)22Cπ2π1ψcos{(+1/2)ψ/π/4}sinπ2ψ2dψ
(6.7)
=π(+1)22Cπ2π1ψcos{(+1/2)ψ/π/4}cosψ2dψ=O(1),
as we argue. To this end, we rewrite
cos{(+1/2)ψ/π/4}cosψ2=12cos((1+1/)ψπ/4)+cos(ψπ/4),
whence both relevant integrals are bounded,
Cπcos((1+1/)ψπ/4)ψdψ,Cπcos(ψπ/4)ψdψ=O(1),
by integration by parts.
The logarithmic bias is an outcome of
π(+1)22Cπ116πψsinπ2ψ2dψ=116(+1)22logCπ+O(1)=116(+1)22log()+O(1).
(6.8)
Consolidating all the above estimates (6.6), (6.7), and (6.8), and the contribution of the error term in (1.21), we finally obtain
E[L(T|HF)]=2π(+1)22116(+1)22log()+O(1).
The contribution to the Kac–Rice integral on the right-hand side of (6.2) of the set HC is bounded by the straightforward inequality,
E[L(T|HC)]=π0ε0K1,(ψ)sinπ2ψ2dψ=O(1),
on recalling the uniform estimate (1.22). Finally, we may bound the contribution of the intermediate range HI as follows. We first write
E[L(T|HI)]=12πHI11PcosψETψ/|T(ψ/)=0dψ.
Then, we observe that on the intermediate range,
HI=(ψ/,ϕ):ε0<ψ<C,
and the variance at the denominator, i.e., 1 − P(cos(ψ/)), is bounded away from 0, and moreover, the diagonal entries of the unconditional covariance matrix C of the Gaussian vector ∇T are O(2), and so are the diagonal entries of the conditional matrix Ω, since they are bounded by the unconditional ones, as it follows directly from (3.2) or, alternatively, from the vastly general Gaussian correlation inequality.33 This easily gives the following upper bound:
E[T(ψ/)T(ψ/)=0]E[T(ψ/)2T(ψ/)=0]1/2E[T(ψ/)2]1/2=O().
Since the area of HC is O(−1), it follows that the total contribution of this range to the expected nodal length is O(1).□

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.

The data that support the findings of this study are available within the article.

We have that

E[T(x)T(y)]=8π2+1m=m mod 2Y,m(x)Ȳ,m(y)=128π2+1m=Y,m(x)Ȳ,m(y)+m=(1)m++1Y,m(x)Ȳ,m(y)=128π2+1m=Y,m(x)Ȳ,m(y)m=Y,m(x̄)Ȳ,m(y),

where we have used the fact that Y,m(θ, ϕ) = (−1)+mY,m(πθ, ϕ). We apply now the addition theorem for spherical harmonics,

P(cosd(x,y))=4π2+1m=Y,m(x)Ȳ,m(y)

so that

E[T(x)T(y)]=P(cosd(x,y))P(cosd(x̄,y)).

Remark A.1.
In particular, we note that
E[T2(x)]=P(x,x)P(x̄,x)=1P(cos(π2θ)),
and this implies that
Var(T(x))=1P(cos(π))=1(1)ifθ=01P(1)=0ifθ=π/21 as ifθ0,π/2.
Moreover, as , for θ ≠ 0, π/2,
E[T(x)T(y)]P(cosd(x,y))1.

Given the block-diagonal form of the covariance matrix Σ(x), it is sufficient to show that the vector (T(x),θT(x)) is non-degenerate for every xH2, and we assume with no loss of generality that x is of the form x=(θ,0)H2. Using the definition of the spherical harmonics basis {Y,m}, we have

T(θ,0)=8π2+1m=0m mod 22Re(a,m)γmPm(cosθ)=8π2+1m=0m mod 2ξmγmPm(cosθ),

where {ξm}m=0,2, is an array standard Gaussian variable N(0, 1). Assume, by contradiction, that the covariance matrix is singular for some θ̄ so that a.s. T(θ̄,0)=cθT(θ̄,0), with the (deterministic) constant c ≠ 0. Assume at first that is odd so that m is even, and recall that the associated Legendre polynomials are defined by

Pm(t)=(1)m2!(1t2)m/2d+mdt+m(t21),

i.e., for m, even they are polynomials in t ∈ [−1, 1] of degree .

Note that

(P,0(t),P,2(t),P,1(t))t

is a 2×1 vector of linearly independent polynomials of degree , with only odd monomials; hence there exist a full rank matrix Q of dimension 2×2 such that

(P,0(t),P,2(t),P,1(t))t=Q(t,t3,t)t.

Now, take

(ξ̃1,ξ̃3,ξ̃)(ξ,0,ξ,2,ξ,1)Q ,

which is a centered Gaussian vector with the covariance matrix QtQ, hence of full rank. After this transformation of basis, we may write

8π2+1m=0,2,,1ξmγmPm(cosθ)=8π2+1k=1,3,,ξ̃kγm{cosθ}k.

Hence, after the change in variable cos θ = t, we have

θT(θ,0)=1t2ddt8π2+1m=0,2,,1ξmγmPm(t)=8π2+11t2ddtk=1,3,,ξ̃kγmtk=8π2+11t2k=1,3,,ξ̃kγmktk1.

Then, for θT(θ̄,0) and T(θ̄,0) to be linearly dependent, there should exist a value θ̄ such that for t̄=cosθ̄,

ξ̃kkt̄k1=cξ̃kt̄k.

More explicitly, it follows that almost surely

ξ̃kk=cξ̃kt̄

for all k = 1, 3, …, , which is clearly absurd, since the random variables {ξ̃k} are the components of a full rank Gaussian vector (which belongs to a proper linear subspace with probability zero). The case where is symmetric, the only difference being that the functions Pm are polynomials after multiplication by sin θ. In light of all the above, the variance matrix of the vector (T, ∇T) is nonsingular for all θ(0,π2), as claimed.

1.
J.-M.
Azaïs
,
F.
Dalmao
, and
J. R.
León
, “
CLT for the zeros of classical random trigonometric polynomials
,”
Ann. l’Inst. Henri Poincaré, Probab. Stat.
52
(
2
),
804
820
(
2016
).
2.
J.-M.
Azaïs
and
W.
Wschebor
,
Level Sets and Extrema of Random Processes and Fields
(
John Wiley & Sons, Inc.
,
Hoboken, NJ
,
2009
).
3.
J.
Benatar
,
D.
Marinucci
, and
I.
Wigman
, “
Planck-scale distribution of nodal length of arithmetic random waves
,”
J. Anal. Math.
141
(
2
),
707
749
(
2020
).
4.
P.
Bérard
, “
Volume des ensembles nodaux des fonctions propres du Laplacien
,”
Sém. Thèor. Spect. Géom.
3
,
1
9
(
1985
).
5.
M. V.
Berry
, “
Regular and irregular semiclassical wavefunctions
,”
J. Phys. A: Math. Gen.
10
(
12
),
2083
2091
(
1977
).
6.
M. V.
Berry
, “
Statistics of nodal lines and points in chaotic quantum billiards: Perimeter corrections, fluctuations, curvature
,”
J. Phys. A: Math. Gen.
35
,
3025
3038
(
2002
).
7.
E.
Bombieri
and
J.
Bourgain
, “
A problem on sums of two squares
,”
Int. Math. Res. Not. (IMRN)
11
,
3343
3407
(
2015
).
8.
J.
Bourgain
, “
On toral eigenfunctions and the random wave model
,”
Isr. J. Math.
201
(
2
),
611
630
(
2014
).
9.
J.
Brüning
Über
, “
Knoten Eigenfunktionen des Laplace-Beltrami operators
,”
Math. Z.
158
,
15
21
(
1978
).
10.
J.
Brüning
and
D.
Gromes
, “
Über die Länge der Knotenlinien schwingender Membranen
,”
Math. Z.
124
,
79
82
(
1972
).
11.
V.
Cammarota
,
O.
Klurman
, and
I.
Wigman
, “
Boundary effect on the nodal length for arithmetic random waves, and spectral semi-correlations
,”
Commun. Math. Phys.
376
,
1261
(
2020
).
12.
V.
Cammarota
,
D.
Marinucci
, and
I.
Wigman
, “
On the distribution of the critical values of random spherical harmonics
,”
J. Geom. Anal.
26
(
4
),
3252
3324
(
2016
).
13.
S.-Y.
Cheng
, “
Eigenfunctions and nodal sets
,”
Commun. Math. Helv.
51
,
43
55
(
1976
).
14.
H.
Donnelly
and
C.
Fefferman
, “
Nodal sets of eigenfunctions on Reimannian manifolds
,”
Invent. Math.
93
(
1
),
161
183
(
1988
).
15.
J. E. A.
Dunnage
, “
The number of real zeros of a random trigonometric polynomial
,”
Proc. London Math. Soc.
s3-16
(
1
),
53
84
(
1966
).
16.
A.
Granville
and
I.
Wigman
, “
The distribution of the zeros of random trigonometric polynomials
,”
Am. J. Math.
133
(
2
),
295
357
(
2011
).
17.
A.
Hassell
and
T.
Tao
, “
Upper and lower bounds for normal derivatives of Dirichlet eigenfunctions
,”
Math. Res. Lett.
9
,
289
305
(
2002
).
18.
O.
Klurman
and
A.
Sartori
,
Res. prog.
On the behaviour of nodallines near the boundary for Laplace eigenfunctions on the square
” (unpublished).
19.
M.
Krishnapur
,
P.
Kurlberg
, and
I.
Wigman
, “
Nodal length fluctuations for arithmetic random waves
,”
Ann. Math.
177
(
2
),
699
737
(
2013
).
20.
N. N.
Lebedev
,
Special Functions and Their Applications
(
Dover Publications, Inc.
,
New York
,
1972
).
21.
A.
Logunov
and
E.
Malinnikova
, “
Nodal sets of Laplace eigenfunctions: Estimates of the Hausdorff measure in dimensions two and three. 50 Years with Hardy spaces
,”
Oper. Theory Adv. Appl.
261
,
333
344
(
2018
).
22.
A.
Logunov
, “
Nodal sets of Laplace eigenfunctions: Proof of Nadirashvili’s conjecture and of the lower bound in Yau’s conjecture
,”
Ann. Math.
187
(
1
),
241
262
(
2018
).
23.
A.
Logunov
, “
Nodal sets of Laplace eigenfunctions: Polynomial upper estimates of the Hausdorff measure
,”
Ann. Math.
187
(
1
),
221
239
(
2018
).
24.
D.
Marinucci
and
G.
Peccati
,
Random Fields on the Sphere: Representation, Limit Theorems and Cosmological Applications
(
Cambridge University Press
,
2011
).
25.
D.
Marinucci
,
M.
Rossi
, and
I.
Wigman
, “
The asymptotic equivalence of the sample trispectrum and the nodal length for random spherical harmonics
,”
Ann. l’Inst. Henri Poincaré, Probab. Stat.
56
(
1
),
374
390
(
2020
).
26.
D.
Marinucci
,
G.
Peccati
,
M.
Rossi
, and
I.
Wigman
, “
Non-universality of nodal length distribution for arithmetic random waves
,”
Geom. Funct. Anal.
26
(
3
),
926
960
(
2016
).
27.
I.
Nourdin
,
G.
Peccati
, and
M.
Rossi
, “
Nodal statistics of planar random waves
,”
Commun. Math. Phys.
369
(
1
),
99
151
(
2019
).
28.
F.
Oravecz
,
Z.
Rudnick
, and
I.
Wigman
, “
The Leray measure of nodal sets for random eigenfunctions on the torus
,”
Ann. l’Inst. Fourier
58
(
1
),
299
335
(
2008
).
29.
Z.
Rudnick
and
I.
Wigman
, “
On the volume of nodal sets for Eigenfunctions of the Laplacian on the torus
,”
Ann. Henri Poincaré
9
(
1
),
109
130
(
2008
).
30.
G.
Szëgo
,
Orthogonal Polynomials
, 4th ed. (
American Mathematical Society; Colloquium Publications
,
Providence, RI
,
1975
), Vol. XXIII.
31.
A.
Vidotto
, “
Note on the Reduction principle for the nodal length of planar random waves
,” arXiv:2007.04228 (
2020
).
32.
C.
Qualls
, “
On the number of zeros of a stationary Gaussian random trigonometric polynomial
,”
J. London Math. Soc.
s2-2
(
2
),
216
220
(
1970
).
33.
T.
Royen
, “
A simple proof of the Gaussian correlation conjecture extended to multivariate gamma distributions
,” arXiv:1408.1028 (
2014
).
34.
I.
Wigman
, “
Fluctuations of the nodal length of random spherical harmonics
,”
Commun. Math. Phys.
298
(
3
),
787
831
(
2010
);
Erratum
309
(
1
),
293
294
(
2012
).