We show that the variance of the number of connected components of the zero set of the two-dimensional Gaussian ensemble of random spherical harmonics of degree n grows as a positive power of n. The proof uses no special properties of spherical harmonics and works for any sufficiently regular ensemble of Gaussian random functions on the two-dimensional sphere with distribution invariant with respect to isometries of the sphere. Our argument connects the fluctuations in the number of nodal lines with those in a random loop ensemble on planar graphs of degree four, which can be viewed as a step toward justification of the Bogomolny–Schmit heuristics.

Let (fn) be the ensemble of random Gaussian spherical harmonics of degree n on the two-dimensional sphere, and let N(fn) be the number of connected components of the zero set {fn = 0}. It is known that

E[N(fn)]=(c+o(1))n2,n,

with a positive numerical constant c and that the random variable N(fn) exponentially concentrates around its mean.4 A beautiful Bogomolny–Schmit heuristics2 suggests that, for any ɛ > 0 and n large enough,

n2ε<Var[N(fn)]<n2+ε.

However, the rigorous bounds we are aware of are much weaker,

nσVar[N(fn)]n4σ,

with some σ > 0. The upper bound with σ=215 follows from the exponential concentration of N(fn) around its mean (see Remark 1.2 of Ref. 4). The purpose of this paper is to prove the lower bound. The proof we give uses no special properties of spherical harmonics and shows that this lower bound holds for any smooth non-degenerate ensemble of Gaussian random functions on the two-dimensional sphere S2 with a distribution invariant with respect to isometries of the sphere and correlations decaying at least as a positive power of the appropriately scaled distance on S2.

It is worth mentioning that in a recent study,1 Beliaev, McAuley, and Muirhead found non-trivial lower bounds for fluctuations of the number of connected components in the disk of radius R ≫ 1 of the level sets {F = } and of the excursion set {F} of the random plane wave (which is a scaling limit of the ensemble of spherical harmonics) for non-zero levels ≠ 0. It is expected that in their case, the fluctuations are much larger than the ones we study. The techniques used in their work are quite different.

Let (fL) be an ensemble of Gaussian random functions on the two-dimensional sphere. It is convenient to assume that the function fL is defined on the sphere S2(L)={xR3:|x|=L} of large radius L and is normalized by E[fL(x)2]=1 for all xS2(L). We always assume that the distribution of fL is invariant with respect to the isometries of the sphere. Then, the covariance kernel of fL has the form

KL(x,y)=E[fL(x)fL(y)]=kL(dL(x,y)),x,yS2(L),

where dL is the spherical distance on S2(L). We call such an ensemble (fL) regular if the following two conditions hold:

  1. C3+-smoothness: KLC3+ν,3+ν(S2(L)) with estimates uniform in L and with some ν > 0.

  2. Power decay of correlations: KL(x,y)(1+dL(x,y))γ, x,yS2(L), with some γ > 0 and with the implicit constant independent of L.

Condition (1) yields that almost surely, fLC3(S2(L)) with estimates uniform in L. We also note that condition (2) is equivalent to the estimate |kL(d)| ≲ (1 + d)γ for 0 ⩽ dπL (with the implicit constant independent of L).

By Z(fL), we denote the random zero set of fL, which is, almost surely, a collection of disjoint simple smooth closed random curves (“loops”) on S2(L). By N(fL), we denote the number of these loops.

Theorem.
Let (fL) be a regular Gaussian ensemble. Then, there existsσ > 0 such that, forLL0,
Var[N(fL)]Lσ.

There are many natural regular Gaussian ensembles, but a nuisance is that the spherical harmonics ensemble is not among them. Spherical harmonics are symmetric with respect to the center of the sphere, so their values at the antipodal points on the sphere coincide up to the sign. The correlations for this ensemble still satisfy condition (2) but only in the range 0 ⩽ d ⩽ (πɛ)L with any ɛ > 0. For this reason, our theorem cannot be applied to this ensemble directly. Luckily, the case of the spherical harmonics requires only minor modifications in the proof of the theorem, which we will outline in the last section of this work. Essentially, we just need an analogue of our theorem for the projective plane RP2 instead of the sphere.

Heuristically, the fluctuations in the topology of the zero set are caused by fluctuations in the signs of the critical values. To exploit this heuristics, we fix a random function fL and slightly perturb it by a multiple of its independent copy gL, i.e., consider the random function

f̃L=1α2fL+αgL,0<α1,

which has the same distribution as fL. Let α be another small parameter, which is significantly bigger than α′, α′ ≪ α ≪ 1, and let

Cr(α)=pS2(L):fL(p)=0,|fL(p)|α.

Then, as we will see, with high probability, given fL, the change in the topology of the zero set Z(f̃L) is determined by the signs of f̃L at Cr(α), that is, by the collection of the random values gL(p):pCr(α). To make the correlations between these random values negligible, the set Cr(α) should be well-separated on the sphere S2(L). At the same time, the set Cr(α) has to be relatively large; otherwise, the impact of fluctuations in signs of f̃L on the number N(f̃L) will be negligible. In Lemma 17, we will show that

  • there exist positiveɛ0,c, andCsuch that, given 0 < ɛɛ0andLL0(ɛ), forL−2+ɛαL−2+2ɛ, with probability very close to 1, the set Cr(α) isL1−-separated and |Cr(α)| ⩾ L.

The proof of this lemma given in Secs. VII–IX is the longest and probably the most delicate part of our work.

To understand how the signs of f̃L at Cr(α) affect the topology of the zero set Z(f̃L), we develop in Sec. VI a little caricature of the quantitative Morse theory. This caricature is non-random—its applicability to the random function fL relies on the fact that with high probability, the Hessian ∇2fL cannot degenerate at the points where the function fL and its gradient ∇fL are simultaneously small. We show that if the parameter α′ is small enough, then with high probability, the topology of Z(f̃L) depends only on the signs of the eigenvalues of the Hessian ∇2fL(p) and the signs of f̃L(p), p ∈ Cr(α). We will describe how these signs determine the structure of the zero set Z(f̃L) in small neighborhoods of the points p ∈ Cr(α). Outside these neighborhoods, the zero lines of Z(f̃L) stay close to the ones of Z(fL).

First, we consider the critical points p ∈ Cr(α) for which both eigenvalues of the Hessian ∇2fL(p) have the same sign, i.e., the points that are local extrema of fL. In this case, we show that there exists a disk D(p, δ) centered at p of a small radius δ such that with high probability, Z(f̃L)D(p,δ) either consists of a simple loop encircling the point p when the sign of f̃L(p) is opposite to that of the eigenvalues of ∇2fL(p) or is empty when these signs coincide. We call such connected components of Z(f̃L)blinking circles.

Now, we turn to the case when the eigenvalues of the Hessian ∇2fL(p) have opposite signs, i.e., to the saddle points of fL. In this case, the situation is more intricate. We define a degree four graph G(fL) embedded in S2(L). Its vertices are small neighborhoods J(p, δ) of saddle points p ∈ Cr(α). The edges are arcs in the set Z(fL) that connect these neighborhoods. We will show that with high probability, the collection of signs {sgn(f̃L(p)):pCr(α)} determines how the graph G(fL) is turned into a collection of loops in Z(f̃L), which we will call the Bogomolny–Schmit loops. Figure 1 illustrates how the sign of f̃L at the saddle point p ∈ Cr(α) determines the structure of the zero set Z(f̃L) in a small neighborhood of the saddle point p ∈ Cr(α).

FIG. 1.

The sign of f̃L determines the structure of the zero set Z(f̃L) near the saddle point p.

FIG. 1.

The sign of f̃L determines the structure of the zero set Z(f̃L) near the saddle point p.

Close modal

We see that in both cases, the fluctuations in the number of connected components ofZ(f̃L)are caused by fluctuations in the signs,

sgn(f̃L(p))=sgn(1α2fL(p)+αgL(p)),pCr(α).

We show that, since the points of the set Cr(α) are well separated and the covariance kernel kL(d) decays at least as a power of d, with probability very close to one, we can replace the values gL(p), p ∈ Cr(α), by a collection of independent standard Gaussian random variables.

Thus, conditioning on fL, we may assume that the values of f̃L at Cr(α) are independent normal random variables (not necessarily mean zero). To conclude, we apply Lemma 21 on the variance of the number of loops generated by percolation-like processes on planar graphs of degree 4.

Note that this chain of arguments can be viewed as the first, although very modest, step toward justification of the Bogomolny–Schmit heuristics.

Throughout the paper, we will be using the following notation:

  • L is a large parameter that tends to +. We always assume that L ⩾ 1.

  • S2(L) denotes the sphere in R3 centered at the origin and of radius L, while, as usual, S2 denotes the unit sphere in R3. By dL, we denote the spherical distance on S2(L). By D(x,ρ)S2(L), we denote the open spherical disk of radius ρ centered at x.

  • G¯ is the closure of the set G.

  • We use the abbreviations a.s. for “almost surely,” w.o.p. for “with overwhelming probability,” which means that the property in question holds outside an event of probability O(LC) with everyC > 0, and w.h.p. for “with high probability,” which means the property in question holds outside an event of probability O(Lc) with somec > 0.

  • C and c (with or without indices) are positive constants that might only depend on the parameters in the definition of the regular Gaussian ensemble fL (C3+ν-smoothness and the power decay of correlations). One can think that the constant C is large (in particular, C ⩾ 1), while the constant c is small (in particular, c ⩽ 1). The values of these constants are irrelevant for our purposes and may vary from line to line.

  • AB means ACB, AB means AcB, and AB means that AB and AB simultaneously.

  • The sign ≪ means “sufficiently smaller than” and ≫ means “sufficiently larger than.” For instance, the assumption “given A and B such that AB” means that there exists c ∈ (0, 1) such that the corresponding conclusion holds for every positive A and B satisfying AcB.

For convenience, we collect here standard facts that we will be using throughout this work.

It will be convenient to associate with each point pS2(L) its own coordinate chart. For pS2(L), let Πp be a plane in R3 passing through the origin and orthogonal to p. The Euclidean structure on Πp is inherited from R3. By Sp2(L), we denote the hemisphere of S2(L) centered at p. By

Ψp:XΠp:|X|LSp2(L),

we denote the map inverse to the orthogonal projection. Note that

|XY|dL(Ψp(X),Ψp(Y))2|XY|

whenever |X|,|Y|12L.

Let f:S2(L)R be a smooth function. We put Fp = f◦Ψp and identify dkf(p) with dkFp(0), i.e., with a k-linear form on Πp. Then, the gradient ∇f(p) = ∇Fp(0) is a vector in Πp such that df(p) (v) = ⟨∇f(p), v⟩, v ∈ Πp, and the Hessian Hf(p) = ∇2f(p) = ∇2Fp(0) is a self-adjoint operator on Πp such that d2f(p) (u, v) = ⟨Hf(p)u, v⟩, u, v ∈ Πp.

The notation fCk(S2(L)) means that, for every pS2(L), fΨpCk{XΠp:|X|12L}, and

fCk=defj=0kmaxS2(L)djf.

Obviously, fCkmaxpS2(L)fΨpCk|X|12L. In the other direction, it is not difficult to see that if fCk(S2(L)), then fΨpCk|X|12LCkfCk.

Fix pS2(L) and the orthogonal coordinate system (X1, X2) on the plane Πp, and set i1ikkf(p)=Xi1XikkFp(0), where, as above, Fp = f◦Ψp.

1. Independence

To simplify the notation, next, we will deal with the case L = 1. The general case can be easily obtained by scaling.

Lemma 1.

Letfbe aC2+ν-smooth random Gaussian function on the sphereS2whose distribution is invariant with respect to the isometries of the sphere. Then, the following Gaussian random variables are independent:

  • f(p) andf(p), as well asf(p) and2f(p),

  • 1f(p) and2f(p),

  • f(p) and1,22f(p), and

  • 1,12f(p)and1,22f(p), as well as2,22f(p)and1,22f(p).

Proof.
Without loss of generality (WLOG), we assume that p is the North Pole of the sphere S2 and suppress the dependence on p, letting Ψ = Ψp. Then,
Ψ(X1,X2)=X1,X2,1(X12+X22),
and F = f◦Ψ is a Gaussian function on the unit disk {|X1|2 + |X2|2 < 1} with the covariance
K(X,Y)=E[F(X)F(Y)]=kd(Ψ(X),Ψ(Y))(d is the spherical distance)=karccosΨ(X),Ψ(Y)=karccosj=12XjYj+1j=12Xj2121j=12Yj212.
Note that
K(X1,X2,Y1,Y2)=K(X1,X2,Y1,Y2)=K(X1,X2,Y1,Y2).
(5.1)
To prove properties (i)–(iv), we need to check the corresponding statements for F(0), XjF(0), and XiXj2F(0). The covariances of these random variables can be computed using the relations
EF(X)1X12X2mF(Y)m1Y1m2Y2=+mK(X,Y)1X12X2m1Y1m2Y2.
The rest follows by differentiation of relations (5.1). For instance,
2KY1Y2(X1,X2,Y1,Y2)=2KY1Y2(X1,X2,Y1,Y2),
whence
EF(0)Y1Y22F(0)=2KY1Y2(0,0)=0.
Similarly,
4K2X1Y1Y2(X1,X2,Y1,Y2)=4K2X1Y1Y2(X1,X2,Y1,Y2),
whence
EX12F(0)Y1Y22F(0)=4KX12Y1Y2(0,0)=0
and so on.□

2. Non-degeneracy

Lemma 2.
Let (fL) be a regular Gaussian ensemble, and letpS2(L). Then,
lim infLE(jfL(p))2>0,j{1,2},
and
lim infLE(i,j2fL(p))2>0,i,j{1,2}.

Proof.

Again, we assume that p is the North Pole of the sphere S2(L). Put ΨL(X1,X2)=X1,X2,L2(X12+X22), and consider the Gaussian functions FL = fL◦ΨL defined in the disks 12LD. The corresponding covariances E[FL(X)FL(Y)]=KL(X,Y) are C3+ν,3+ν-smooth on 12LD×12LD with some ν > 0. Their partial derivatives up to the third order are bounded locally uniformly in LL0. Hence, by a version of the Arzelá–Ascoli theorem, any sequence KLj contains a locally uniformly C2+ν,2+ν-convergent subsequence.

The limiting function K is a C2+ν,2+ν-smooth Hermitian-positive function on R2×R2, which depends only on the Euclidean distance |XY|. Hence, by Bochner’s theorem, it is a Fourier integral of a finite positive rotation-invariant measure ρ,
K(X,Y)=R2e2πiλ,XYdρ(λ),
where
R2|λ|4dρ(λ)<.
Furthermore, since
KL(X,Y)1+dL(ΨL(X),ΨL(Y))γ,
uniformly in LL0, the limiting function K satisfies
K(X,Y)1+|XY|γ.
Therefore, the measure ρ cannot degenerate to the point measure at the origin.
The rest is straightforward. Suppose, for instance, that for some sequence Lj,
limjE(1fLj(p))2=0.
Then,
limj2KLjX1Y1(0,0)=0.
Passing to a subsequence, we conclude that
(2πi)2R2λ12dρ(λ)=2X1Y1R2e2πiλ,XYdρ(λ)X=Y=0=0.
Since the measure ρ is positive and rotation-invariant, this is possible only when ρ is a point mass at the origin. This contradiction concludes the proof.□

3. Power decay of correlations

The power decay of the correlations between fL(p) and fL(q) when dL(p, q) is large and the a prioriC3,3-smoothness of the covariance yield the power decay of correlations between the Gaussian vectors

v(p)=(fL(p),fL(p))=(Fp(0),Fp(0))=(Fp(0),X1Fp(0),X2Fp(0))

and

v(q)=(fL(q),fL(q))=(Fq(0),Fq(0))=(Fq(0),Y1Fq(0),Y2Fq(0))

[we keep fixed the coordinate systems (X1, X2) and (Y1, Y2) in the planes Πp and Πq].

Lemma 3.
Let (fL) be a regular Gaussian ensemble. Then, for anyp,qS2(L),
max1i,j3E[vi(p)vj(q)]1+dL(p,q)γ/4.

Proof.

Put Kp,q(X,Y)=EFp(X)Fq(Y), where Fp = fL◦Ψp and (X,Y)Q¯×Q¯, where Q¯=[1,1]×[1,1]. We have E[Fp(0)YiFq(0)]=YiKp,q(0,0), and E[XiFp(0)YjFq(0)]=XiYj2Kp,q(0,0).

There is nothing to prove if dL(p, q) ⩽ 1, so we assume that dL(p, q) ⩾ 1. Then, Kp,qC(Q¯×Q¯)dγ and Kp,qC2,2(Q¯×Q¯)1. Now, we use the classical Landau–Hadamard inequality in the following form: if h:[0,1]R is a C2-smooth function and Mj = max[0,1]|h(j)|, 0 ⩽ j ⩽ 2, then M1max(M0,M0M2). Applying this to the functions XiKp,q(X, Y) and YjKp,q(X, Y), we get XiKp,qC(Q¯×Q¯)dγ/2 and YjKp,qC(Q¯×Q¯)dγ/2, i, j = 1, 2. Applying the Landau–Hadamard inequality again, this time to the functions YjXiKp,q(X,Y), we get XiYj2Kp,qC(Q¯×Q¯)dγ/4, i, j = 1, 2.

In particular, these estimates hold at X = Y = 0, which gives us what we needed.□

  • To simplify our notation, in the following, we assume that the parameterγ > 0 is chosen so that the correlations between the Gaussian vectors (fL(p), ∇fL(p)) and (fL(q), ∇fL(q)) decay as(1+dL(p,q))γ.

Let (fL) be a regular Gaussian ensemble. Quite often, we will be using the following a priori bound:

  • w.o.p,fLC3(S2(L))<logL.

This bound immediately follows from the classical estimate

PfLC3(S2(L))>tCL2ect2.

For a self-contained proof, see Secs. A9–A11 of Ref. 6.

Here, we introduce the (non-random) class C3(A, Δ, α, β) of smooth functions f on S2(L) such that the number of connected components of the zero set of a small perturbation f̃ of f can be recovered from the values of f̃ at the critical points of f with small critical values (provided that the values of f̃ at these points are not too small). Later, we will show that w.h.p., our random function fL belongs to this class.

Given α, β ⩽ 1 and Δ ⩾ 1, we let

Cr(α)={pS2(L):|f(p)|α,f(p)=0},
Cr(α,β)={pS2(L):|f(p)|α,|f(p)|β},

and

Cr(α,β,Δ)={pS2(L):|f(p)|α,|f(p)|β,(2f(p))1opΔ},

where ‖⋅‖6op stands for the operator norm.

By C3(A), we denote the class of C3-smooth functions f on S2(L) with fC3A. Given the parameters

αβ1AΔ,

by C3(A, Δ, α, β), we denote the class of functions fC3(A) for which Cr(α, β) = Cr(α, β, Δ), i.e., the Hessian of f does not degenerate [(2f)1opΔ] on the almost singular set Cr(α, β) where f and ∇f are simultaneously small.

Given fC3(A, Δ, α, β), α′ ≪ α, and gC3(A), we set

ft=f+tg,0tα.

Next, we develop a little caricature of the quantitative Morse theory, which shows that the collection of signs of ft at Cr(α) defines the topology of the zero set Z(ft), provided that minCr(α)|ft| is not too small, and gives “an explicit formula” that recovers the number of connected components of Z(ft) from this collection of signs and the structure of Z(f).

Lemma 4.
Suppose thatfC3(A) andp ∈ Cr(α, β, Δ) with
1AΔ,AΔ2β1.
Then,
  • the spherical diskD(p, 2Δβ)contains a unique critical pointzoff,

  • there are no other critical points offin the diskDp,c(AΔ)1, and

  • |f(z)| ⩽ 2α, provided thatAΔ2β2α.

Proof.
We will work on the plane Πp, and let F = f◦Ψp and HF=2F. To find the critical point z, we use a simplified Newton method,
Xn+1=XnHF(0)1F(Xn),X0=0.
Put Φ(X)=XHF(0)1F(X). First, we check that in the disk D0,c(AΔ)1, the map Φ is a 13-contraction.
Indeed,
Φ(X)Φ(Y)=(XY)HF(0)1F(X)F(Y)=(XY)HF(0)1HF(X)(XY)+O(A|XY|2),
whence
|Φ(X)Φ(Y)|IHF(0)1HF(X)op|XY|+ΔO(A|XY|2).
Furthermore,
IHF(0)1HF(X)opHF(0)1opHF(0)HF(X)opΔO(A|X|),
and finally,
|Φ(X)Φ(Y)|AΔO(|X|)+O(|XY|)|XY|13|XY|,
provided that |X|, |Y| ⩽ c(AΔ)−1 with sufficiently small positive c.
Next, we check that the map Φ preserves the disk D(0, c(AΔ)−1). We have
Φ(X)=XHF(0)1F(X)=HF(0)1F(X)HF(0)X=HF(0)1F(0)+O(A|X|2),
and then,
|Φ(X)|Δβ+AcAΔ2O(1)=βΔ+c2AΔO(1)<cAΔ,
provided that βAΔ2 ≪ 1 and that the positive constant c is sufficiently small.
Thus, Φ has a unique fixed point Z in the disk D(0, c(AΔ)−1), and
|Z|=|0Z|n0|XnXn+1|n03n|X0X1|=32|0Φ(0)|=32|Φ(0)|=32|HF(0)1F(0)|32Δβ.
At last,
|F(Z)||F(0)|+|F(0)||Z|+O(A|Z|2)α+β2βΔ+O(Aβ2Δ2)2α,
provided that 2Δ2α.□

Lemma 5.
Letf, gC3(A), whereft = f + tgwith 0 ⩽ tα. Suppose thatp ∈ Cr(α) withHf(p)1opΔand that
1AΔ,Aαα,AΔ2α1.
Then, there exists a unique critical pointptofftsuch thatdL(p, pt) ≪ Δαand |ft(p) − ft(pt)| ≪ Aα)2. Moreover, there are no other critical points offtat distancec(AΔ)−1fromp.

Proof of Lemma 5.
First, we note that ftC3(2A) and that |ft(p)| ⩽ |f(p)| + ′ and |∇ft(p)|≲ ′ ≪ α. Furthermore, Hft(p)1op2Δ. Indeed, we have
Hft(p)1op=Hf(p)1(I+(Hft(p)Hf(p))Hf(p)1)1opΔ(I+(Hft(p)Hf(p))Hf(p)1)1op.
Noting that Hft(p)Hf(p)opAα, we see that
(Hft(p)Hf(p))Hf(p)1opΔAαΔα1.
Therefore,
(I+(Hft(p)Hf(p))Hf(p)1)1op<2,
and finally, Hft(p)1op2Δ.
Then, by Lemma 4 (applied to the function ft with β = ′), there exists a unique critical point pt of ft with
dL(p,pt)2(Aα)2ΔΔα.
Besides, for dL(p, x) ≪ Δα, we have dL(pt, x) ≪ Δα and then |∇ft(x)| = |∇ft(x) − ∇ft(pt)| ≪ A ⋅ Δα, whence
|ft(p)ft(pt)|A(Δα)2.

At last, by part B of Lemma 4, there are no other critical points of ft at distance ⩽c(AΔ)−1 from p.□

Given fC3(A), p ∈ Cr(α), and Hf(p)1opΔ, we look at the behavior of ft in the δ-neighborhood of p. As above, ft = f + tg, with gC3(A) and 0 ⩽ tα′. Throughout this section, we assume that the parameters α′, α, δ, A, and Δ satisfy the set of conditions

α1AΔ,AααA2Δ3
(6.1)

and that

δ=c(AΔ)1,
(6.2)

with sufficiently small constant c. Note that these conditions are more restrictive than the ones used in Lemma 5, so we will be using freely that lemma.

1. Local extrema

First, we consider the case when the Hessian Hf(p) is positive or negative definite, that is, its eigenvalues have the same sign. With a little abuse of terminology, we say that the function ft is convex (concave) in D(p, δ) if the function ft◦Ψp is convex (correspondingly, concave) in Ψp1D(p,δ)Πp.

Lemma 6.

Suppose that the eigenvalues of the HessianHf(p)have the same sign and that conditions (6.1) and (6.2) hold. Then,

  • the functionftis either concave or convex function inD(p, δ), and

  • the functionftdoes not vanish on∂D(p, δ), and moreover, the sign offtD(p,δ)coincides with the sign of the eigenvalues ofHf(p).

Proof of Lemma 6.
Put F = f◦Ψp and Ft = ft◦Ψp, and suppose that HF(0)=Hf(p) is positive definite (otherwise, replace f by −f), that is, HF(0)x,xΔ1|x|2. Then, for any XΨp1D(p,δ), we have
HF(X)x,xHF(0)x,xHF(X)HF(0)op|x|2(Δ1CA|X|)|x|2(2Δ)1|x|2
since HF(X)HF(0)opA|X| and |X| ⩽ δ = c(AΔ)−1 with sufficiently small c. Noting that HFt(X)HF(X)opAαΔ1, we get HFt(X)x,x(4Δ)1|x|2, which proves (i).
To prove (ii), we take XΨp1D(p,δ). Then,
Ft(X)=Ft(0)+Ft(0),X+12HFt(0)X,X+O(Aδ3)12HFt(0)X,X|Ft(0)||Ft(0),X|O(Aδ3).
Furthermore, using that |Ft(0)| = |ft(p)| ⩽ α + O(′) ≲ α and that |⟨∇Ft(0), X⟩| ⩽ |∇Ft(0)|⋅|X| = O(αA) ⋅ δα, we conclude that
Ft(X)|X|δ/212(4Δ)1(δ/2)2O(α+Aδ3)=(32Δ)1δ2O(Aδ3)=(32Δ)1O(Aδ)δ2δ=c(AΔ)1132CcΔ1δ2164Δ1δ2,
provided that the constant c in (6.2) (the definition of δ) was chosen so small that Cc164.□

Summary: Let p be a local extremum of f. Suppose that conditions (6.1) and (6.2) hold.

  • Then, Z(ft) ∩ ∂D(p, δ) = .

  • Z(ft) ∩ D(p, δ) is either empty, homeomorphic to S1, or a singleton.

  • Suppose that |ft(p)| ≳ Aα)2. Then, by Lemma 5, ft(pt) has the same sign as ft(p). Therefore, Z(ft) ∩ D(p, δ) = whenever ft(p) and the eigenvalues of Hf(p) have the same sign, and Z(ft) ∩ D(p, δ) is homeomorphic to S1 whenever ft(p) and the eigenvalues of Hf(p) have opposite signs.

2. Saddle points

Now, we turn to the case when p ∈ Cr(α) is a saddle point of f, that is, the eigenvalues of Hf(p) have opposite signs. We will work on the plane Πp and set F = f◦Ψp, G = g◦Ψp, and Ft = ft◦Ψp = F + tG. By H(X)=HF(0)X,X, we denote the quadratic form generated by the Hessian HF(0). WLOG, we assume that

H(X)=aX12bX22,Δ1abA.

We take δ = c(AΔ)−1 with a sufficiently small positive constant c, set

J(δ)=def|H|aδ2|X1|3δ,

and call this set a joint. By

*J(δ)=def|H|=aδ2|X1|3δ,

we denote the curvilinear part of the full boundary ∂J(δ) of the joint J(δ) (Fig. 2).

FIG. 2.

Joint J(δ) with four terminals.

FIG. 2.

Joint J(δ) with four terminals.

Close modal

Lemma 7.

Suppose that the eigenvalues ofHf(p)have the opposite signs and that conditions (6.1) and (6.2) hold. Then, the functionFtdoes not vanish on*J(δ). Moreover, the signs ofFtandHcoincide on*J(δ).

Proof of Lemma 7.
Everywhere in J(δ), we have
Ft=F(0)+12H+O(Aδ3)+αG=12H+Oα+(δ3+α)A=αδ312H+OAδ3.
Furthermore, on *J(δ), we have |H| = 2c2A−2Δ−3, while 3 = c3A−2Δ−3. This proves the lemma.□

The set |H|aδ22δ|X1|3δ consists of four disjoint curvilinear quadrangles. We call them terminals and denote them by Ti, 1 ⩽ i ⩽ 4.

Lemma 8.

Under the same assumptions as in Lemma 7, each of the setsZ(Ft) ∩ Ticonsists of one curve, which joins the vertical segments on the boundary ofTi.

Proof of Lemma 8.
Everywhere in J(δ), we have
(Ft)X2=12HX2+(FX212HX2)+tGX2.
Hence,
|(Ft)X2|b|X2|CA|X|2O(Aα).
In each of the terminals Ti,
b|X2|b3abδ>abδδΔ
(in the last estimate, we use that ba ⩾ Δ−1), and
|X|2(3δ)2+10abδ219δ2,
whence
CA|X|219CAδ2=19CAcAΔδδ2Δ,
provided that the constant c in the definition of δ is sufficiently small. Furthermore, ′ is also much smaller than Δ−1δ (since ′ ≪ A−1Δ−2). Thus, |(Ft)X2|>0 everywhere in Ti. It remains to recall that, by Lemma 7, the function Ft has at least one change in sign on each vertical section of Ti. Therefore, by the implicit function theorem, Z(Ft) ∩ Ti is a graph of a smooth function.□

Under the same assumptions as in Lemmas 7 and 8, by Lemma 5, the joint J(δ) contains only one critical point Xt=(X1t,X2t) of Ft and |Xt| ≪ Δα. Consider the sets

I1=X=(X1,X2t):X1RJ(δ),I2=X=(X1t,X2):X2RJ(δ).

Since

|Xt|Δα(AΔ)2=c1(AΔ)1δ(AΔ)1/2δa/bδ(Δ1abA),

it is easy to see that both sets are the segments.

Lemma 9.

Under the same assumptions as in Lemmas 7 and 8, the only extremum of the restriction of the functionFtto the segmentI1is a local minimum atX1=X1t, and the only extremum of the restriction of the functionFtto the segmentI2is a local maximum atX2=X2t.

Proof of Lemma 9.
Consider the function X1Ft(X1,X2t). It has a critical point at X1=X1t, the second derivative at this point is not less than
2aCAαCA|Xt|2ΔCA(α+Δα)1Δ
(since AΔ2α ≪ 1, and ′ ≪ α), and the C3-norm of Ft is bounded by CA. Therefore, for X1>X1t, we have
(Ft)X1(X1,X2t)2a(X1X1t)CA(X1X1t)2Δ1(X1X1t)CA|X1X1t|2.
Note that the RHS of the last expression is positive since
X1X1t<2δ=2c(AΔ)1
with sufficiently small constant c. Similarly, (Ft)X1(X1,X2t)<0 for X1<X1t.

The proof of the second statement is almost identical and we skip it.□

Lemma 10.

Suppose thatFt(Xt) ≠ 0 [i.e., zero is not a critical value of the restriction of the functionFtto the jointJ(δ)]. Then, under the same assumptions as in Lemmas 79, the setZ(Ft) ∩ J(δ) consists of two connected components, which enter and exit the jointJ(δ) through the terminalsTi.

Furthermore, the set {Ft ≠ 0} ∩ J(δ) consists of three connected components. One of them containsFt(Xt), while on the other two components,Fthas the sign opposite to the sign ofFt(Xt).

Proof of Lemma 10.

Since zero is not a critical value of the restriction FtJ(δ), the set Z(Ft) ∩ J(δ) consists of a finitely many disjoint smooth curves. By Lemma 8, this set has at least two connected components, the ones that enter and exit the joint J(δ) through the terminals. If there exists a third component, then, again by Lemma 8, it cannot intersect the terminals, while by Lemma 7, it also cannot intersect the rest of the boundary *J(δ). Hence, it stays inside the joint. Therefore, it is a closed curve, which bounds a domain G with ḠJ(δ). Since Ft vanishes on ∂G, G must contain the (unique) critical point Xt of Ft and ∂G separates Xt from ∂J(δ). On the other hand, Lemma 9 together with Lemma 7 yields that on one of the segments Ii, i = 1, 2, the function Ft does not change its sign. The resulting contradiction proves the first part of the lemma.

To prove the second part, first, we notice that, since the sets {Ft > 0} ∩ J(δ) and {Ft < 0} ∩ J(δ) cannot be simultaneously connected, the set {Ft ≠ 0} ∩ J(δ) has at least three connected components. One of them, we call it Ω0, contains the critical point Xt, and therefore, by Lemma 9, it contains one of the segments Ii. Since Ft does not vanish on *J(δ) (Lemma 7), the boundary of Ω0 contains two opposite sides of *J(δ), the ones on which the end points of the segment Ii lie. For the same reason, there are two more connected components of the set {Ft ≠ 0} ∩ J(δ), and each of these two components contains on its boundary one of two remaining opposite sides of the set *J(δ). At last, arguing as in the proof of the first part (also using again Lemmas 8 and 9), we see that the fourth connected component of the set {Ft ≠ 0} ∩ J(δ) cannot exist.□

Summary: Let p be a saddle point of f. Suppose that conditions (6.1) and (6.2) hold, and let J(p, δ) = ΨpJ(δ) be the corresponding joint. Suppose that 0 is not a critical value of ft.

  • Then, the set Z(ft) ∩ J(p, δ) consists of two connected components. Each of them enters and exits the joint through its own terminals ΨpTi.

  • We say that the joint J(p, δ) has positive type if the set J(p, δ) ∩{ft > 0} is connected [therefore, the set J(p, δ) ∩{ft < 0} is disconnected and consists of two connected components]. Otherwise, we say that the joint J(p, δ) has negative type. Suppose that |ft(p)|≳ Aα)2. Then, the type of the joint J(p, δ) coincides with the sign of ft(p).

Fix the functions fC3(A, Δ, α, β) and gC3(A). Let ft = f + tg and f̃=fα, and consider the gradient flow zt, 0 ⩽ tα′, defined by the ordinary differential equation (ODE),

dztdt=(tft)(zt)|ft(zt)|2ft(zt)
(6.3)

with the initial condition z0Z(f).

Lemma 11.

Suppose thatAΔ2β2α ≪ (AΔ)−2βand′ ≪ α. Letδ = c(AΔ)−1with a sufficiently small constantc > 0. Then, for any arcIZ(f) \⋃p∈Cr(α)D(p, 2δ2), the flowztprovides aC1-homotopy ofIonto an arcĨZ(f̃)\pCr(α)D(p,δ2).Vice versa, for any arcĨZ(f̃)\pCr(α)D(p,2δ2), the inverse flowzα′−tprovides aC1-homotopy ofĨonto an arcIZ(f) \⋃p∈Cr(α)D(p, δ2). Moreover, these homotopies move the points by at mostO(′/β).

Proof of Lemma 11.
Let X=def|f|<β and
Ω̄=Ω̄(ε,η)=def(t,x)[ε,α+ε]×(S2(L)\X):|ft(x)|η
(the choice of small positive parameters ɛ and η has no importance), and let Ω be the interior of Ω̄. Note that
|tft|=|g|Aeverywhere,
and
|ft||f|Aα12βon S2(L)\X.
Therefore, the flow moves the points with the speed
dztdt=(tft)(zt)|ft(zt)|2Aβ.
The RHS of the ODE (6.3) is a C1-function on Ω̄. Therefore, for any initial point z0Z(f)\X̄, the ODE has a unique C1-solution. The solution exists until it reaches the boundary of Ω. Note that along the trajectory zt, we have
ddtft(zt)=tft(zt)+ft(zt),dztdt=0,
whence ft(zt) = 0 [recall that z0Z(f)]. Hence, if the solution zt is not defined on [0, α′], then there exists τα′ such that dL(zt, X) → 0 as t ↑ τ. This means that the closure of the trajectory zt, 0 ⩽ t < τ, contains a point z̄ with |f(z̄)|β. Recalling that the point zt moves with the speed ⩽ 2A/β, we see that dL(z̄,z0)2Aα/β. Furthermore, by the continuity of ft, we have fτ(z̄)=0, whence |f(z̄)|α|g(z̄)|Aαα. Combining this with the gradient estimate |f(z̄)|β and applying Lemma 4, we conclude that there is a unique critical point p of f with dL(p,z̄)2βΔ, i.e., with
dL(p,z0)2βΔ+2Aαβ,
which is much less than δ2.
It remains to check that this critical point p belongs to Cr(α), which is straightforward,
|f(p)||f(z̄)|+dL(z̄,p)β+O(dL(z̄,p)2A)Aα+Δβ2+AΔ2β2α,
since ′ and AΔ2β2 are both much less than α.

Since the function f̃=f+αg belongs to the class C3(2A, 2Δ, 2α, 2β), the same arguments can be also applied to the inverse flow zα′−t.□

We start with functions fC3(A, Δ, α, β) and gC3(A) and consider the perturbation f̃=f+αg. We assume that the parameters

ααβ1AΔ

satisfy the following relations:

Aαα,AΔ2β2α(AΔ)2β,A2Δ3α1

[which particularly yield conditions (6.1)]. We also assume that the perturbation f̃ is not too small on Cr(α),

minCr(α)|f̃|AΔ2α2.

We set

CrS(α)={pCr(α):pis a saddle point of f},CrE(α)={pCr(α):pis a local extremum of f}.

We put δ = c(AΔ)−1 with sufficiently small positive constant c and consider the disks D(p, δ) and p ∈ CrE(α) and the joints J(p, δ) and p ∈ CrS(α). If the constant c in the definition of δ was chosen sufficiently small, then all these disks and joints are mutually disjoint [recall that by Lemma 5, the points from the set Cr(α) are c0(AΔ)−1-separated with a positive constant c0].

1. Stable loops

These are connected components of Z(f) and Z(f̃) that do not intersect the set

U=U(Cr(α),δ)=defpCrE(α)D(p,δ)pCrS(α)J(p,δ).

We denote by NI(f) the number of stable loops in Z(f) and by NI(f̃) the number of stable loops in Z(f̃).

Observe that Dp,abδJ(p,δ), p ∈ CrS(α), and that ab(AΔ)1/2. We see that, for each p ∈ Cr(α), we have D(p, 2δ2) ⊂ U. Therefore, Lemma 11 applies to stable loops in Z(f) as well as to stable loops in Z(f̃) and yields a one-to-one correspondence between the set of stable loops in Z(f) and the set of stable loops in Z(f̃). That is, NI(f̃)=NI(f).

2. Blinking circles

These are small connected components of f̃ that surround the points p ∈ CrE(α) and lie in the interiors of the corresponding disks D(p, δ). Recall that, by Lemma 6, Z(f̃) cannot intersect the boundary circle ∂D(p, δ) of such a disk.

By the summary in the end of the local extrema Sec. VI D 1, the number of such components is

NII(f̃)=defpCrE(α):f̃(p)and the eigenvalues of Hf(p)have opposite signs.

3. The Bogomolny–Schmit loops

This is the most interesting part of Z(f̃). Consider the graph G = G(f) embedded in S2(L). The vertices of G are the joints J(p, δ),p ∈ CrS(α). The edges are connected components of the set

Z(f)\pCrS(α)J(p,δ)
(6.4)

that touch the boundaries ∂J(p, δ) [these components are homeomorphic to intervals, while the other connected components of the set (6.4) are homeomorphic to circles]. Each vertex of this graph has degree 4. The signs of f̃(p), p ∈ CrS(α), determine the way the graph G is turned into a collection of loops [(Fig. 3) see the summary in the end of the saddle point, Sec. VI D 2]. By NIII(f̃), we denote the number of loops in this collection.

FIG. 3.

Creation of the Bogomolny–Schmit loops.

FIG. 3.

Creation of the Bogomolny–Schmit loops.

Close modal

4. The main lemma of Sec. VI

At last, we are able to state the main result of this section.

Lemma 12.
LetfC3(A, Δ, α, β),gC3(A), andf̃=f+αg. Suppose that the parameters
ααβ1AΔ
(6.5)
satisfy following relations
Aαα,AΔ2β2α(AΔ)2β,A2Δ3α1
(6.6)
and that
minCr(α)|f̃|AΔ2α2.
(6.7)
Then,
N(f̃)=NI(f̃)+NII(f̃)+NIII(f̃).

Now, we return to regular Gaussian ensembles (fL).

Lemma 13.
Given a sufficiently small positiveɛ, letαL−2+2ɛandβ2L3ɛα. Then, there existsL0 = L0(ɛ) such that, for eachLL0, w.h.p.,
maxCr(α,β)(2fL)1opL3ε,
where ‖⋅‖opdenotes the operator norm.

Proof.
Fix a small ɛ > 0 and consider the set Cr(5α, 4β). Let p be the probability that a given point xS2(L) belongs to the set Cr(5α, 4β). By the invariance of the ensemble (fL), this probability does not depend on x. The statistical independence of fL(x) and ∇fL(x) (Lemma 1), non-degeneracy of their distributions (Lemma 2), and uniform boundedness of their variances yield that
p=P|fL(x)|5αP|fL(x)|4βαβ2.
Next, note that, by Fubini’s theorem,
E[area(Cr(5α,4β))]=parea(S2(L))pL2,
whence, by Chebyshev’s inequality,
Parea(Cr(5α,4β))>pL2L12εL12ε.
Thus, w.h.p.,
area(Cr(5α,4β))pL2L12ε(αL2)β2L12εαL2L2εβ2L2.5ε.

Denote by μμ(x) the eigenvalue of the Hessian matrix ∇2fL with the minimal absolute value and by w = w(x) the corresponding normalized eigenvector. Assume that fLC3<logL (recall that this holds w.o.p.), and suppose that, for some x ∈ Cr(α, β) and LL0, |μ(x)| < Δ−1, where Δ = L3ɛ. We will show that then, the set Cr(5α, 4β) contains a subset G̃=G̃(x) with area(G̃)β2L3ε(logL)1. This will immediately imply the lemma.

Fix a point x ∈ Cr(α, β) with |μ(x)| < Δ−1, take the corresponding map Ψx, and let F = fL◦Ψx. Put τ = βΔ ≪ 1. For Y = tw(x), 0 ⩽ tτ, we have
|F(Y)||F(0)|+|2F(0)Y|+O(|Y|2FC3({|X|1}))β+Δ1τ+τ2O(logL)=2β+β2Δ2O(logL)3β.
Then, letting I = [0, tw(x)], we get
|F(Y)||F(0)|+maxI|F||Y|α+3βτ=α+3β2Δ4α,
since β2Δ ⩽ α.
Put ρ = (log L)−1 with a sufficiently small positive constant c, and denote by Ω the ρ-neighborhood of the segment I on the plane Πx. Then,
maxΩ¯|F|maxI|F|+O(ρFC3({|X|1}))3β+ρO(logL)<3.5β,
and
maxΩ¯|F|maxI|F|+ρmaxI|F|+O(ρ2FC3({|X|1}))4α+3βρ+ρ2O(logL)=4α+β2o(1)<5α.
Let Ω̃=Ψx(Ω). We see that Ω̃Cr(5α,4β).
At the same time,
area(Ω̃)τρ=cβ2Δ(logL)1=cβ2L3ε(logL)1,
completing the proof.□

The next lemma gives us a lower bound for the probability that a given point xS2(L) belongs to the set

Cr(α,β,Δ)={xCr(α,β):(2fL(x))1opΔ}.

Let p be the probability that a given point xS2(L) belongs to the set Cr(α, β). By the invariance of the ensemble (fL), this probability does not depend on x.

Lemma 14.
For anyα, β ⩽ 1 and any Δ ⩾ 2,
PxCr(α,β,Δ)(1C(logΔ)14Δ12)p.

Proof.
We need to show that, conditioned on x ∈ Cr(α, β), the probability that
(2fL(x))1opΔ
is (logΔ)14Δ12. Since fL(x) and its Hessian ∇2fL(x) are independent of the gradient ∇fL(x), it will suffice to show that
P(2fL(x))1op>Δ|f(x)|α(logΔ)14Δ12.

Fix xS2(L) and denote by μ1(x) and μ2(x) the eigenvalues of the Hessian. First, we show that, conditioned on the event {|f(x)| ⩽ α}, with large probability, |μ1(x) ⋅ μ2(x)| = |det ∇2fL(x)| cannot be too small, and then that, with large probability, max(|μ1(x)|, |μ2(x)|) = ‖∇2fL(x)‖op cannot be too big. Together, these two estimates will do the job.

Fix coordinates (X1, X2) in the plane Πx. Then,
det2fL(x)=1,12fL(x)2,22fL(x)(1,22fL(x))2.
Recalling that, by Lemma 1, 1,22fL(x) is independent of the vector
fL(x),1,12fL(x),2,22fL(x)t
and that by Lemma 2, the distribution of 1,22fL(x) does not degenerate, we conclude that, for any δ > 0,
P|μ1(x)μ2(x)|<δ|f(x)|αsupsRP|(1,22fL(x))2s|<δδ.
In the second step, taking into account that 2fL(x)op2max1i,j2|i,j2fL(x)|, we need to estimate from above the absolute values of the second order derivatives of fL at x conditioned on fL(x). The mixed derivative 1,22fL(x) has a bounded variance and is independent of fL(x). Furthermore, we use the normal correlation theorem, which says that if (θ, ξ) is a two-dimensional Gaussian vector, then the expectation and variance of θ conditioned on ξ equal
E[θ|ξ]=E[θξ]Var[ξ]ξ
and
Var[θ|ξ]=Var[θ](E[θξ])2Var[ξ]Var[θ].
By this theorem, the distribution of i,i2fL(x), i = 1, 2, conditioned on fL(x), is normal with bounded conditional mean
Ei,i2fL(x)|fL(x)=EfL(x)i,i2fL(x)fL(x)1
(recall that fL is normal and that we are interested only in the values |fL(x)| ⩽ α ⩽ 1) and with the bounded conditional variance
Vari,i2fL(x)|fL(x)Vari,i2fL(x)1.
Thus, Pmax(|μ1(x)|,|μ2(x)|)>λ|f(x)|αecλ2. Therefore, after conditioning on {|f(x)| ⩽ α}, with probability at least 1Cδ+ecλ2, we have
δ|μ1(x)||μ2(x)|=min(|μ1(x)|,|μ2(x)|)max(|μ1(x)|,|μ2(x)|)λmin(|μ1(x)|,|μ2(x)|)=λ(2fL(x))1op1.
Letting λ = δΔ and δ=12cΔ1logΔ and noting that δ+ecλ2Δ12(logΔ)14, we complete the proof.□

In this section, we will look at the set Cr(α, β) of almost singular points of fL,

Cr(α,β)=xS2(L):|fL(x)|α,|fL(x)|β,

with very small parameters α and β, and at its one- and two-point functions p and p(x, y). As above, p is the probability that a given point xS2(L) belongs to the set Cr(α, β). Recall that, by the invariance of the ensemble (fL), this probability does not depend on x and that the statistical independence of fL(x) and ∇fL(x) yields that

p=P|fL(x)|αP|fL(x)|βαβ2.

By p(x, y), we denote the probability that two given points x,yS2(L) belong to the set Cr(α, β). By the invariance of the distribution of fL with respect to isometries of the sphere, p(x, y) depends only on the spherical distance between the points x and y.

Lemma 15.
Let (fL) be a regular Gaussian ensemble. Then, the estimates hold uniformly inα, β ⩽ 1 and inLL0,
p(x,y)maxdL(x,y)Θ,1p2,
(8.1)
with some positive constant Θ, and fordL(x, y) ⩾ 1, we havep(x, y) = Wp2, with
|W1|dL(x,y)γ.
(8.2)

Note that the proof of the short-distance estimate (8.1) is quite lengthy, while the long-distance estimate (8.2) is a straightforward consequence of the power decay of correlations of (fL).

1. Beginning the proof

Let x,yS2(L). Fix the coordinate systems in the planes Πx and Πy, and let Γ(x, y) be the covariance matrix of the Gaussian six-dimensional vector,

v(x,y)=fL(x),fL(x),fL(y),fL(y)t.

Then,

p(x,y)=1(2π)3detΓ(x,y)Ωexp12ξtΓ(x,y)1ξd6ξ,

where

Ω=ξR6:|ξ1|,|ξ4|α,|ξ2|2+|ξ3|2,|ξ5|2+|ξ6|2β2.

Note that vol(Ω) ≲ α2β4p2. Hence, to prove estimate (8.1), we need to bound from below the minimal eigenvalue λ = λ(x, y) of the covariance matrix Γ = Γ(x, y),

λ=minξtΓ(x,y)ξ:ξR6,|ξ|=1.

Note that λ does not depend on the choice of the coordinate systems in the planes Πx and Πy.

First, we show that there exists a sufficiently large constant d0, independent of L, so that λ(x, y) is bounded from below by a positive constant whenever dL(x, y) ⩾ d0. Hence, proving estimate (8.1), we assume that dL(x, y) ⩽ d0 [while later, proving the long-distance estimate (8.2), we will assume that dL(x, y) ⩾ d0]. The value of sufficiently large constant d0 is inessential for our purposes.

Denote by C the covariance matrix of ∇fL(x) and put

Γ̃=10000C000010000C.

Since the matrix C is non-degenerate uniformly in LL0 (Lemma 2), the matrix Γ̃ is also non-degenerate uniformly in LL0.

By our assumption on the power decay of correlations, we have

max1i,j6Γij(x,y)Γ̃ij(1+dL(x,y))γ.

Then, provided that dL(x, y) ⩾ d0 with d0 ≫ 1, we get

Γ(x,y)1Γ̃1opdL(x,y)γ,

and therefore,

Γ(x,y)1opΓ̃1opOdL(x,y)γ12Γ̃1op.

Thus, until the end of the proof of the short-distance estimate (8.1), we assume that dL(x, y) ⩽ d0 with some positive d0 independent of L.

Let v(x) denote the three-dimensional Gaussian vector v(x)=fL(x),fL(x)t, and let a=(ξ1,ξ2,ξ3)t and b=(ξ4,ξ5,ξ6)t. Then,

ξtΓ(x,y)ξ=Ev(x,y),ξ2=E(v(x),a+v(y),b)2,

whence

λ=minE(v(x),a+v(y),b)2:a,bR3,|a|2+|b|2=1.

By the compactness of the unit sphere in R3, there exist a,bR3, |a|2 + |b|2 = 1, such that

λ=E(v(x),a+v(y),b)2=v(x),a+v(y),b2.
(8.3)

Here and until the end of the Proof of Lemma 15, ‖⋅‖ stands for the L2-norm, i.e., η2=E[|η|2]=Var[η] for the Gaussian random variable η.

2. The normal and the tangential derivatives

Let C be the big circle on S2(L) that passes through the points x and y, and let IC be the shortest of the two arcs of C with the endpoints x and y. We orient C by moving from x to y along I and choose the coordinate systems in Πx and Πy so that one coordinate vector is parallel to the tangent to C (at x and y correspondingly), while the other one is orthogonal to C. We keep the same orientation for both coordinate systems. We denote by the derivative along C and by the derivative in the normal direction to C and decompose

v(x)=v(x)+v(x),

where v(x)=fL(x),fL(x),0t and v(x)=0,0,v(x)t. Then, by (8.3),

λ=v(x),a+v(x),a+v(y),b+v(y),b2,
(8.4)

where a=(a1,a2,0)t and a=(0,0,a3)t, similarly for b′ and b, and |a′|2 + |a|2 + |b′|2 + |b|2 = 1.

Now, consider another Gaussian six-dimensional vector

ṽ(x,y)=(fL(x),fL(x),fL(x),fL(y),fL(y),fL(y))t.

Since the distribution of fL is invariant with respect to orthogonal transformations, the Gaussian vectors v(x, y) and ṽ(x,y) have the same covariance matrix in the chosen coordinate systems in Πx and Πy. Therefore,

λ=v(x),av(x),a+v(y),bv(y),b2.
(8.5)

Juxtaposing (8.4) with (8.5), we conclude that

λ=v(x),a+v(y),b2+v(x),a+v(y),b2.

We split the rest of the proof of estimate (8.1) into two cases: (i) |a|2+|b|212 and (ii) |a|2+|b|212. In the first case, we shall use a straightforward argument, while in the second case, we shall prove an equivalent estimate on the Fourier side.

3. Case (i): |a|2+|b|212

In this case, we use the estimate λ ≳‖⟨v(x), a⟩ + ⟨v(y), b⟩‖2. By the invariance of the distribution of fL, the random pairs (fL(x), fL(y)) and (fL(y), fL(x)) have the same distribution. Therefore,

v(x),a+v(y),b=v(y),a+v(x),b.

Since |a|2+|b|212, at least one of the following holds:

  • either |a+b|12 or |ab|12.

We assume that |a+b|12 (the other case is similar and slightly simpler), let e = a + b, |e|12, and notice that

v(x)+v(y),e=v(x),a+v(y),b+v(x),b+v(y),av(x),a+v(y),b+v(x),b+v(y),aλ.

We take the point zC, zx, so that dL(y, z) = dL(x, y). Then, by the invariance of the distribution of fL with respect to the isometries of the sphere,

v(y)+v(z),e=v(x)+v(y),e,

whence

λv(x)v(z),e.

Put x0 = x and x1 = z, then take the point x2C, x2x0, so that dL(x2, x1) = dL(x1, x0), and continue this way until dL(x0, xN) ⩾ d0, where d0 is the correlation length defined above. Then,

v(x0)v(xN),eNλ.

On the other hand, since d0 is the correlation length and dL(x0, xN) ⩾ d0, we have

v(x0)v(xN),e2v(x0),e2+v(xN),e21

(at the last step, we use that the distribution of v does not degenerate uniformly in LL0 and that |e|12). Therefore, λN−2.

Recalling that by the definition of N, we have (N − 1)dL(x, y) < d0, and we get λdL(x,y)2, which concludes our consideration of the first case.

4. Case (ii): |a|2+|b|212

In this case, we restrict the function fL to the big circle C and treat it as a periodic random Gaussian function F:RR with translation-invariant distribution. To simplify the notation, we omit the index L. By ρ, we denote the spectral measure of F, that is,

E[F(X)F(Y)]=ρ^(XY)=Re2πiξ(XY)dρ(ξ),

where X,YR correspond to the points x,yC. Then, we have

λEv(X),a+v(Y),b2=R(a1+a2ξ)e2πiξX+(b1+b2ξ)e2πiξY2dρ(ξ)=R(a1+a2ξ)+(b1+b2ξ)eiδξ2dρ(ξ),

where δ=12π|XY|=12πdL(x,y). Furthermore,

ρ(R)=E[F(0)2]=1,
m=defRξ2dρ(ξ)=E[F(0)2],m1,

by the uniform non-degeneracy of ∇fL and

|ρ^(s)|=|E[F(0)F(s)]||s|γ,|s|12L,

by the power decay of correlations of fL. Note that since the function F is 2πL-periodic, the Fourier transform ρ^ of its spectral measure is also 2πL-periodic.

Recalling that 12|a|2+|b|21, we notice that |a1|+|a2|+|b1|+|b2|12 as well. These remarks reduce the lower bound for λ we are after to a question in harmonic analysis.

Given δ > 0, consider the exponential sum of degree 4,

pδ(ξ)=(a1+a2ξ)+(b1+b2ξ)eiδξ,a1,a2,b1,b2C,
(8.6)

and denote ‖pδW = |a1| + |a2| + |b1| + |b2| (recall that an exponential sum is the expression

S(ξ)=j=1nqj(ξ)eiλjξ,degS=defj=1n(degqj+1),

where λj are real numbers and qj are polynomials in ξ with complex coefficients). Then, the following lemma does the job.

Lemma 16.
Letpδbe the exponential sum (8.6) of degree 4. Letρbe a probability measure onRwith the 2πL-periodic Fourier transformρ^. Assume that
m=Rξ2dρ(ξ)1
and
|ρ^(s)|C|s|γ,|s|12L.
Then, givenδ0 > 0, there existc = c(C, γ, m, δ0) > 0 andL0 = L0(C, γ, m, δ0) such that, for every 0 < δδ0and everyLL0,
R|pδ|2dρcδ6pδW2.

1. Beginning the Proof of Lemma 16

Let A=2m. Then, by Chebyshev’s inequality,

ρ(R\[A,A])A2m=14.

Take B = δ0A and let 0 < δδ0. Then, [−A, A] ⊂ [−−1, −1]. Let κ > 0 be a sufficiently small parameter, which we will choose later, and consider the set

Ξ=ξ[Bδ1,Bδ1]:|pδ(ξ)|<pδWκ3δ3.

We claim that

  • the set Ξ is a union of at most B + 5 intervals Ij, 1 ⩽ jB + 5, and

  • the length of each interval Ij is ⩽ C(B, δ0) κ.

Having these claims, we will show that ρ[A,A]\Ξ14, whence

[A,A]\Ξ|pδ|2dρ14(κδ)6pδW2.

This will complete the Proof of Lemma 16.

2. Proof of Claim (a)

To show (a), we consider the exponential sum P = |pδ|2 of degree 9. By the classical Langer lemma (see Lemma 1.3 of Ref. 3), the number of zeroes of any exponential sum of degree N on any interval JR cannot exceed

(N1)+Δ2π|J|,

where Δ is the maximal distance between the exponents in the exponential sum. Hence, the number of solutions to equation P(ξ) = t on the interval [−−1, −1] does not exceed

8+4δ2π2Bδ1<8+2B.

Hence, the set Ξ consists of at most 12((8+2B)+2)=5+B intervals, proving (a).

3. Proof of Claim (b)

To show (b), we apply Turán’s lemma (see Theorem 1.5 of Ref. 3), which states that for any exponential sum S of degree N and any pair of closed intervals IJ,

maxJ|S|C|J||I|N1maxI|S|,

where C is a numerical constant. Applying this lemma to the exponential sum pδ of degree 4 and to each of the intervals Ij ⊂ [−−1, −1], we get

max[Bδ1,Bδ1]|pδ|C2Bδ1|Ij|3maxIj|pδ|C2Bδ1|Ij|3pδWκ3δ3(since IjΞ)=κ3C2B|Ij|3pδW.

On the other hand,

max[Bδ1,Bδ1]|pδ|=maxξ[B,B](a1+δ1a2ξ)+(b1+δ1b2ξ)eiξ(by scaling)c(B)|a1|+δ1|a2|+|b1|+δ1|b2|(by compactness)δδ0c(B)(1+δ0)1pδW.

Thus, |Ij| ⩽ C(B, δ0)κ, proving (b).

4. Completing the Proof of Lemma 16

Recall that ρ([A,A])34 and that given κ > 0, we defined the set

Ξ=ξ[Bδ1,Bδ1]:|pδ(ξ)|<pδWκ3δ3,

satisfying (a) and (b). Then, we have the following alternative:

  • either ρ([A,A]\Ξ)14 or ρ([A,A]Ξ)12.

In the first case,

R|pδ|2dρ[A,A]\Ξ|pδ|2dρ14pδW2κ6δ6,

and we are done (modulo the choice of the parameter κ, which will be made later). It remains to show that ifκis sufficiently small, then the second case cannot occur.

Suppose that ρ([A,A]Ξ)12. Then, ρ(Ξ)12. By claim (a), Ξ is a union of at most B + 5 intervals Ij, hence, for at least one of them, ρ(Ij) ⩾ c(B) > 0. We call this interval I and denote by ν the restriction of the measure ρ on I. We choose a large parameter S so that 1 ≪ Sκ−1 and estimate the integral

J=SS1|s|Sν^(s)2ds

from below and from above, obtaining the estimates that will contradict each other.

Denote by ξI the center of the interval I. Then, for |s| ⩽ S and ξI, we have

e2πiξse2πiξIs2π|s||ξξI|2πS12|I|2πSκ12C(B,δ0)(since, by claim (b),|I|C(B,δ0)κ)<12

provided that Sκ is sufficiently small. Therefore, for |s| ⩽ S, we have

ν^(s)2=Ieisξdρ(ξ)214|ρ(I)|2c1(B),

whence

Jc1(B)SS1|s|Sdsc2(B)S.

On the other hand, using the identity

Rφ(s)|ν^(s)|2ds=R×Rφ^(λη)dν(λ)dν(η),

with φ(s) = (1 −|s|/S)+ and noting that the Fourier transform of this function is non-negative, we get

SS1|s|Sν^(s)2dsSS1|s|Sρ^(s)2ds.

Then, recalling that the function ρ^ is 2πL-periodic and that |ρ^(s)|min(1,C|s|γ) for |s| ⩽ L/2 and assuming without loss of generality that γ<12, we get

JS(min(S,L))2γ.

Choosing κ sufficiently small and S sufficiently large, we arrive at a contradiction, which completes the Proof of Lemma 16, and therefore, that of estimate (8.1) in Lemma 15.□

As above, we denote by C the covariance matrix of ∇fL(x) and put

Γ̃=10000C000010000C.

Since the matrix C is non-degenerate uniformly in LL0, the matrix Γ̃ is also non-degenerate uniformly in LL0.

We assume that dL(x, y) ⩾ d0, where d0 is sufficiently large (and independent of L). Then, we have

max1i,j6Γij(x,y)Γ̃ijdL(x,y)γ.

Therefore,

detΓ(x,y)=detΓ̃+OdL(x,y)γ

and

Γ(x,y)1Γ̃1opdL(x,y)γ.

Recall that

p(x,y)=1(2π)3detΓ(x,y)Ωexp12ξtΓ(x,y)1ξd6ξ,

where

Ω=ξR6:|ξ1|,|ξ4|α,|ξ2|2+|ξ3|2,|ξ5|2+|ξ6|2β2,

and that

1(2π)3detΓ̃Ωexp12ξtΓ̃1ξd6ξ=p2.

Hence,

p(x,y)=(2π)3detΓ̃+O(dL(x,y)γ)12Ωexp12ξtΓ̃1ξ+O(dL(x,y)γ)d6ξ=(2π)31+O(dL(x,y)γ)1detΓ̃Ωexp12ξtΓ̃1ξd6ξ=1+O(dL(x,y)γ)p2,

completing the proof of estimate (8.2) in Lemma 15.□

Now, we are ready to prove our main lemma:

Lemma 17.

There exist positiveɛ0andcand positiveCsuch that, given 0 < ɛɛ0andLL0(ɛ), forL−2+ɛαL−2+2ɛ, w.h.p., the set Cr(α) isL1−-separated, and |Cr(α)| ⩾ L.

In this part, we assume that αL−2+2ɛ, choose β and ρ so that

β2L3εα,ρ=β(logL)1,

and fix a maximal ρ-separated set X(ρ) on S2(L). Then, |X(ρ)|(L/ρ)2.

First, we note that, w.h.p., the points of the set Cr(α) are L−4ɛ-separated. This is a straightforward consequence of part (B) in Lemma 4 combined with a priori w.o.p.-bound fLC3<logL and with the w.h.p.-estimate maxCr(α)(2fL)1opL3ε provided by Lemma 13. Hence, we need to estimate the probability of the event

E=z1,z2Cr(α):L4εdL(z1,z2)L1Cε

with an appropriately chosen constant C.

Suppose that the event E occurs. Denote by x1 and x2 the closest to z1 and z2 points in X(ρ). Then,

12L4εdL(x1,x2)2L1Cε,
(9.1)

and

|fL(xi)|α+O(ρ2fLC2)<w.o.p.α+O(β2(logL)1)<2α,|fL(xi)|O(ρfLC2)<w.o.p.β(logL)1logL=β,

i.e., x1, x2 ∈ Cr(2α, β). We claim that

  • the mean number of pairs of pointsx1,x2Cr(2α,β)X(ρ)satisfying(9.1)is bounded from above byLɛ.

By Chebyshev’s inequality, this yields that the probability that there exists at least one such pair is also bounded from above by Lɛ, which proves the L1−-separation.

The mean we need to estimate equals

x1,x2Cr(2α,β)X(ρ)(9.1) occursp(x1,x2),

where p(x1,x2)=P{x1,x2Cr(2α,β)} is the two-point function estimated in Lemma 15. By Lemma 15, p(x1, x2) ≲ L4ɛΘp2, so the whole sum is

x1X(ρ)L4εΘp2{x2X(r):dL(x1,x2)2L1Cϵ}L4εΘp2(L1Cε/ρ)2(L/ρ)2L(C4Θ)εα2β4(L4/β4)log4L(pαβ2,ρ=β/logL)L(C4Θ)εL4εlog4L(αL2+2ε)<Lε,

provided that the constant C is sufficiently large. This proves that the set Cr(α) is L1−-separated.□

In this part, we assume that αL−2+ɛ and introduce the parameters β, Δ, and r, satisfying

Δ=L14ε,β2L3ε=13α,r=βΔ.

We fix a maximal r-separated set X(r) on S2(L). Then, the disks D(x, r), xX(r), cover S2(L) with a bounded multiplicity of covering and |X(r)|(L/r)2. We set

Y=X(r)Cr(13α,β,12Δ).

W.o.p., for LL0, we have fLC3logL. Then, by Lemma 4, each disk D(y, r), yY, contains a unique critical point z ∈ Cr(α), and

(2fL(z))1op(2fL(y))1op1+O(r(2fL(y))1opfLC3)12Δ+O(rΔ2logL)<Δ,

provided that LL0. This yields two useful observations that hold w.o.p.:

  • |Cr(α)|≳|Y|, and

  • if y1, y2Y, then either dL(y1, y2) ⩽ 2r and the number of such pairs (y1, y2) is |X(r)| or d(y1, y2) ⩾ Lɛ. Indeed, if the points y1, y2Y generate the same critical point z, then dL(y1, y2) ⩽ 2r. If they generate different critical points z, we note that, by part (B) of Lemma 4, the set of critical points z of fL with (2fL(z))1opΔ is cΔ−1(log L)−1-separated, and thus, in this case, d(y1, y2) ⩾ Lɛ.

In the following, we will show that, for sufficiently large L,

E[|Y|]L12ε

and that

Var[|Y|]Lcε(E[|Y|])2.

These two estimates combined with the first observation readily yield what we need.

1. Estimating E[|Y|]

This estimate is straightforward,

E[|Y|]=PxCr(13α,β,12Δ)|X(r)|
PxCr(13α,β)Lr2(by Lemma14)
αβ2L2β2Δ2
(9.2)
L12ε(αL2+ε,Δ=L14ε).
(9.3)

2. Estimating Var[|Y|]

In this section, p and p(x, y) will denote the one- and two-point functions of the set Cr13α,β. Given x,yS2(L), we put

pΔ=PyCr13α,β,12Δ,pΔ(x,y)=Px,yCr13α,β,12Δ.

Then, E[|Y|]=pΔ|X(r)| and

Var[|Y|]=yX(r)(pΔpΔ2)+x,yX(r)xy(pΔ(x,y)pΔ2).

The first sum on the RHS is bounded by

pΔ|X(r)|=E[|Y|](9.3)L12εE[|Y|]2.

Hence, we need to estimate the double sum only.

In the double sum, we consider separately the terms with dL(x, y) ⩽ 2r, the terms with 2r < dL(x, y) < Lɛ, the terms with LɛdL(x, y) ⩽ Lɛ, and the terms with dL(x, y) ⩾ Lɛ.

a. The terms with dL(x, y) ⩽ 2r.

Taking into account that the number of such pairs is |X(r)|, we bound this sum by pΔ|X(r)|=E[|Y|](9.3)L12ε(E[|Y|])2.

b. The terms with 2r < dL(x, y) < Lɛ.

By the second observation, we conclude that w.o.p., this case cannot occur, that is, the probability that there exists a pair of almost-singular points x,yCr(13α,β,12Δ) with 2r < dL(x, y) < Lɛ is O(LC) with any positive C. Thus, in this range, pΔ(x, y) = O(LC), while the total number of pairs x, y is bounded by |X(r)|2(L/r)4(L/β)4L10, provided that ɛ in the definition of the parameters β and α is sufficiently small. That is, the sum is negligibly small.

c. The terms with LɛdL(x, y) ⩽ Lɛ.

In this case, we estimate each summand in the double sum by p(x, y), which, by the short-distance estimate in Lemma 15, is

max(dL(x,y)Θ,1)p2LΘεpΔ2.

Then, the whole double sum is

LΘεpΔ2|X(r)|(Lε/r)2(pΔ|X(r)|)2L2+(2+Θ)ε=L2+(2+Θ)εE[|Y|]2,

which gives what we needed with a large margin.

d. The terms with dL(x, y) ⩾ Lɛ.

In this case,

pΔ(x,y)pΔ2p(x,y)pΔ2=(p(x,y)p2)+(p2pΔ2).

By the long-distance estimate (8.2) of the two-point function p(x, y),

p(x,y)p2Lγεp2LγεpΔ2,

while, by Lemma 14,

p2pΔ2(logΔ)1/4Δ1/2p2Lε/10pΔ2.

Thus, pΔ(x,y)pΔ2LcεpΔ2, and the whole double sum is bounded by LcεpΔ2|X(r)|2=LcεE[|Y|]2.

This completes the proof of the estimate of Var[ |Y| ] and hence that of Lemma 17.□

Lemma 18
(asymptotic independence). Let (fL) be a regular Gaussian ensemble, and letZS2(L)be anL1−κ-separated set with sufficiently small positiveκ. Then, there exist a collection (ξ(z))zZof independent standard Gaussian random variables and positive constantsc1, c2so that, forLL0,
PmaxzZ|fL(z)ξ(z)|>Lc1<eLc2.

Proof.
We will follow rather closely the Proof of Theorem 3.1 in Ref. 5. We fix sufficiently large L and introduce the following notation:
  • HfL is a Gaussian Hilbert space generated by fL, i.e., the closure of finite linear combinations ∑cjfL(xj) with the scalar product generated by the covariance.

  • H is “a big Gaussian Hilbert space” that contains HfL and countably many mutually orthogonal one-dimensional subspaces that are orthogonal to HfL.

  • J(z) = fL(z), zZ, are unit vectors in HfL with

J(z),J(z)H=E[fL(z)fL(z)¯]Lγ(1κ),zz.
We claim that there exists a collection of orthonormal vectorsJ̃(z)zZHsuch that
maxzZJ̃(z)J(z)H2L0.45γ.
To prove this claim, we consider the Hermitian matrix Γ = Γ(z, z′)z,z′∈Z with the elements
Γ(z,z)=J(z),J(z)H,zz,L0.9γ,z=z.
For L > L0(γ, κ), the matrix Γ is positive-definite. Indeed, by the classical Gershgorin theorem, each eigenvalue of Γ lies in one of the intervals (Γ(z, z) − t(z), Γ(z, z) + t(z)) with
t(z)=zZ\{z}|Γ(z,z)|,
so we need to check that, for each zZ, t(z) < Γ(z, z), which is easy to see since
t(z)<|Z|Lγ(1κ)|Z|L2κL2κLγ(1κ)L0.9γ,
provided that κ is sufficiently small. Since the Hermitian matrix Γ is positive-definite, we can find a collection of vectors {I(z)}zZHspan{J(z)}zZ with the Gram matrix Γ. Note that I(z)H=Γ(z,z)=L0.45γ. Then, we let
J̃(z)=J(z)+I(z)J(z)+I(z)H,zZ.
By construction, the system of vectors {J̃(z)}zZ is orthonormal in H. Furthermore,
1J(z)+I(z)H1+L0.45γ,
whence
J(z)J̃(z)HJ(z)+I(z)H1J(z)H+I(z)H2L0.45γ,
proving the claim.
It remains to note that, for each zZ, we have
P|J̃(z)J(z)|>t=e12(tJ̃(z)J(z)H1)2e18t2L0.9γ,
whence, by the union bound,
PmaxzZ|J̃(z)J(z)|>t|Z|e18t2L0.9γL2κe18t2L0.9γ.
Letting t = L−0.4γ, we complete the Proof of Lemma 18.□

In this section, we present two simple and standard lemmas, which will be employed later. To keep this work relatively self-contained, we will include their proofs.

Lemma 19.
Given0<p012, let (ηj) be a collection ofNindependent random variables on the probability space Ω such thatηjattains the value 1 with probabilitypj,p0pj ⩽ 1 − p0, and the value 0 with probability 1 − pj. Let
SN=j=1Nηj.
Then, there existsɛ = ɛ(p0) > 0 such that for any measurable function,Q: Ω → [0, 1] withΩQdP1ε, and for anymR, we have
Ω|SNm|2QdPc(p0)N.

Proof of Lemma 19.
During the proof, the value of c(p0) may vary from line to line. Take λ=1/N. First, we claim that |E[eiλSN]|1c(p0). Indeed,
|E[eiληj]|=|eiλpj+1(1pj)|1c(p0)λ2,
whence
E[eiλSN](1c(p0)λ2)Nec(p0)Nλ21c(p0).
Therefore, for any mR,
E[eiλ(SNm)1]c(p0).
Next, using that |eit − 1| ⩽ |t|, we proceed as follows:
c(p0)E[(eiλ(SNm)1)(Q+(1Q))]Ωλ|SNm|QdP+Ω2(1Q)dP1NΩ|SNm|2QdP+2ε.
Taking ɛc(p0)/4, we complete the proof.□

Lemma 20.
Let(Ω1×Ω2,P1×P2)be a product probability space, and let 0 < p ⩽ 1 and0<ε12p. LetQ: Ω1 × Ω2 → [0, 1] be a measurable function withΩ1×Ω2QdP1ε, and letX ⊂ Ω1be an event withP1(X)p. Then,
P1ω1X:Ω2Q(ω1,ω2)dP2(ω2)12εp1p2.

Proof of Lemma 20.
Put
X=ω1X:Ω2Q(ω1,ω2)dP2(ω2)12εp1.
Then,
1εΩ1×Ω2QdP=Ω1\X+X+X\XΩ2Q(ω1,ω2)dP2(ω2)dP1(ω1)<1P1(X)+P1(X)+(12εp1)(P1(X)P1(X))12ε+2εp1P1(X),
which yields the lemma.□

In the following, we will apply this lemma, mostly, with X = Ω1 and p = 1.

Let G = G(V, E) be a finite graph embedded in the sphere S2 with each vertex having degree four. We allow G to have multiple edges as well as “circular edges,” which connect a vertex with itself. The vertices of the graph are the joints (see Sec. VI D 2), the edges are curves on S2 connecting the vertices, and the faces are the connected components of the open set S2\(VE).

Each vertex vV can be replaced by one of two possible “avoided crossings” at v (Fig. 4).

FIG. 4.

The vertex v and its states σv.

FIG. 4.

The vertex v and its states σv.

Close modal

We call the choice of the avoided crossing at vthe state of the vertexv and denote it by σv. When the states are assigned to all vertices in V, the collection of states σV = {σv: vV} turns the graphs G into a collection of loops Γ = Γ(σV).

We will deal with a random loop model when the states σv are independent random variables taking their values with probabilities p(v) and 1 − p(v). By (Ω,P), we denote the probability space on which the random states are defined. Then, N(Γ) = N(Γ(σV)) is a random variable on (Ω,P).

Given 0<p012, we put V(p0)=vV:p0p(v)1p0 and denote by |V(p0)| the cardinality of the set of vertices V(p0).

Lemma 21.
For any0<p012, there exist positivec(p0),C(p0), andɛ = ɛ(p0) such that for any functionQ(σV) defined on the set of all possible states and taking the values in the interval [0, 1] withΩQ(σV)dP1εand for anymR,
Ω(N(Γ(σV))m)2Q(σV)dPc(p0)|V(p0)|,
provided that |V(p0)| ⩾ C(p0).

We fix a function Q as above. In several steps, we will reduce the statement of the lemma to the anti-concentration bound for the sum of independent Bernoulli random variables provided by Lemma 19. In each of these steps, we will be using the following decoupling argument.

1. Decoupling

Suppose that the vertices are split into two disjoint parts V = V′⊔V and decompose correspondingly σV=(σV,σV). A collection of states σV assigned to the vertices from V′ generates (Fig. 5)

  • a collection Γ′ = Γ(σV) of disjoint loops

  • and a graph G(σV) with vertices at V, all of them having degree 4.

FIG. 5.

Left: the graph G. Right: the loops Γ′ and the graph G(σV).

FIG. 5.

Left: the graph G. Right: the loops Γ′ and the graph G(σV).

Close modal

By N(Γ′), we denote the number of loops in the collection Γ′. The collection of states σV turns the graph G(σV) into a collection of loops Γ=Γ(σV,σV). Then, Γ = Γ′ ⊔Γ, and N(Γ) = N(Γ′) + N).

Let φ(σV) be any non-negative bounded measurable function. Since the random variables σV and σV are independent, we have

ΩφσVωdPω=Ω×ΩφσVω,σVωdPωdPω.=ΩΩφσVω,σVωdPωdPω.

Hence, for any event X′ ⊂Ω,

Ωφ(σV)dPP(X)infσVΣΩφ(σV,σV(ω))dP(ω),

where Σ′ = {σV(ω′): ω′ ∈ X′}. Letting φ = (N(Γ) − m)2Q and taking into account that N(Γ) − m = N) − (mN(Γ′)), we get

ΩNΓσVm2QσVdPPXinfσVΣinfRΩNΓσVω2QσV,σVωdPω.

The choice of the event X′ ⊂ Ω, or what is the same, of the set of states Σ′ is in our hands. Choosing it, we need to keep the value of the integral,

infσVΣΩQ(σV,σV(ω))dP(ω),

close to 1, while P(X) should stay bounded away from zero. This will be done with the help of Lemma 20. After that, it will suffice to prove Lemma 21 for the graph G(σV) with vertices at V.

In the following, we will apply this decoupling argument several times. To simplify the notation, after each step, we treat the function Q as depending only on the states σV of the remaining set of vertices V, ignoring its dependence on the fixed states σV ∈ Σ′.

As above, we let V(p0)=vV:p0p(v)1p0. We put V′ = V\V(p0) and V = V(p0) and consider

X=ωΩ:ΩQ(σV(ω),σV(ω))dP(ω)12ε.

Then, by Lemma 20 (applied with p = 1), P(X)12.

Hence, from now on, we assume that for each vertex v of the graph G, we have p0p(v) ⩽ 1 − p0, while ΩQ(σV)dP12ε.

Denote by F the set of the faces of the graph G. Given a face fF, we denote by v(f) the set of vertices in V that lie on the boundary f. The Euler formula gives us

|V||E|+|F|=1+ν,

where ν is the number of connected components of the graph G. Since the degree of each vertex in G equals 4, we have |E| = 2|V|, whence |F| ⩾ |V| + 2. Furthermore, since each vertex in G lies on the boundary of at most four different faces, we have

|F|>|V|14fF|v(f)|14fF|v(f)|5|v(f)|54{fF:|v(f)|5}.

We let F*={fF:|v(f)|4} be the set of all faces having at most 4 vertices on the boundary and conclude that F*>15|F|>15|V|.

We call faces f,fF*separated if they do not have common vertices on their boundaries: v(f)v(f)=. We fix a maximal collection FF* of separated faces. Since for any face fF, there are at most 12 other faces fF* with v(f)v(f), we conclude from the maximality of F that it also contains sufficiently many faces,

|F|113F*>165|V|.

Next, we choose a simple cycle c on the boundary of each face fF and mark a vertex vc on that cycle according to the following rule: We take a vertex v′ in v(f) and, starting at v′, walk along f turning left at each vertex so that the face f always remains on the left-hand side. We stop when we return for the first time to the vertex that we already have passed and mark that vertex and the corresponding cycle (Fig. 6).

FIG. 6.

Marking a cycle and a vertex on the boundary of the face f.

FIG. 6.

Marking a cycle and a vertex on the boundary of the face f.

Close modal

We will be using the following property of the marked cycles: if the cycle exits in a vertex along an edgee, then it returns to this vertex along an edge, which is one of two edges adjacent toe (this follows from the fact that the same face cannot lie on both sides of some edge). This property yields that there exist states of vertices on c, which turn c into a separate loop.

We denote by VMV the set of all marked vertices and by VUM = V\VM the set of all unmarked vertices.

FIG. 7.

Good and bad cycles.

FIG. 7.

Good and bad cycles.

Close modal
FIG. 8.

Separation vs merging.

FIG. 8.

Separation vs merging.

Close modal

For any particular assignment of the states of the unmarked vertices, the marked cycle c is called good if the edges of c merge into an edge of the graph G(σ(VUM)) that connects the vertex vc with itself (Fig. 7). Note that this event depends only on the states of at most 3 unmarked vertices lying on c\{vc}. Therefore, for any marked cycle c, we have Pcis good p03. Hence, denoting by NG the number of good cycles, we obtain

ENGp03|{marked cycles}|=p03|F|165p03|V|.

Using first the Chebyshev inequality and then the independence of the random states, we get

PNG12E[NG]4Var[NG]E[NG]21302p06|{marked cycles}||V|21302p06|V|1.

To simplify the notation, we let d(p0)=1130p03 and D(p0)=d(p0)2. Then, letting X=ω:NG>d(p0)|V|, we see that P(X)p=def1D(p0)|V|1. Put

X=ωX:ΩQ(σVUM(ω),σVM(ω))dP(ω)14εp1

(recall that at this moment ΩQdP12ε). Then, by Lemma 20, P(X)12p, which is 14, provided that |V| > 2D(p0). From now on, we fix the states of the unmarked vertices corresponding to the event X′ and consider the remaining graph with vertices in VM. At this step, ΩQdP14εp1.

All marked cycles c are split into two classes: bad cycles and good cycles. Correspondingly, we decompose the set of all marked vertices VM into the disjoint union VM = VM.B.VM.G. and consider

X=ωΩ:ΩQ(σVM.B.(ω),σVM.G.(ω))dP(ω)18εp1

By Lemma 20 (applied with X = Ω), P(X)12. We fix the states of marked bad vertices corresponding to the event X′.

We are left with the graph G with vertices at VM.G.. For each vertex vVM.G., there is “a circular edge” ev with the endpoints at v, which came from the corresponding cycle c. Each state σv of the vertex v either creates from this circular edge a separate loop or merges it with other edges (Fig. 8).

We fix a collection of states σVM.G.* of good marked vertices such that that none of the corresponding good cycles turns into a separate loop and denote by Γ(σVM.G.*) the loop ensemble obtained from the graph G after the assignment of the states σVM.G.*. Introduce a collection of independent Bernoulli random variables ηvvVM.G., letting ηv = 0 if σv=σVM.G.*(v) and ηv = 1 otherwise. Then,

N(Γ(σVM.G.))=N(Γ(σVM.G.*))+vVM.G.ηv.

Applying Lemma 19, we get the uniform in m lower bound,

ΩN(Γ(σVM.G.))m2Q(σVM.G.)dPc(p0)|VM.G.|.

To finish off the Proof of Lemma 21, it remains to recall that good marked vertices are in the one-to-one correspondence with good cycles, that is, |VM.G.| = NG, and that the states of unmarked vertices were fixed so that NGd(p0)|V|.□

We choose a sufficiently small ɛ > 0 and take α′ = L−2+ɛ and α = L−2+2ɛ. Then, we take the function fL and its independent copy gL and put

f̃L=1α2fL+αgL.

This is a random Gaussian function equidistributed with fL. We will show that

infmRE(N(f̃L)m)2Lcε,

which immediately yields the lower bound for Var[N(fL)] we are after. Note that

E(N(f̃L)m)2=EfLEgLN(1α2fL+αgL)m2.

That is, it suffices to show that with probability at least 12 in fL, we have

EgL(N(f̃L)m)2Lcε.

We will prove a somewhat stronger statement that this inequality holds if the function fL satisfies the following conditions:

  • fLC3(A, Δ, α, β) (introduced in Sec. VI A) with A = log L, Δ = L3ɛ, and with β chosen so that β2L7ɛ = α (i.e., β=L152ε),

  • the set Cr(α) is L1−-separated, and

  • |Cr(α′)| ⩾ L.

By Lemmas 13 and 17, these three conditions hold w.h.p. in fL. From now on, we fix fL so that these conditions hold and omit the index gL, meaning P=PgL, E=EgL, etc.

The rest will essentially follow from our little Morse caricature summarized in Lemma 12 combined with Lemma 21 on the fluctuations in the number of random loops. In order to apply Lemma 12, first, we observe that the relations

Aαα,AΔ2β2α(AΔ)2β,A2Δ3α1,

required in Lemma 12 readily follow from our choice of the parameters α, α′, β, A, and Δ made few lines above. Lemma 12 also needs the lower bound

minCr(α)|f̃L|AΔ2α2,

which holds w.h.p. with a large margin since AΔ2α2 = L−4+10ɛ log L, while, as we will momentarily see, w.h.p. in gL, we have

minCr(α)|f̃L|>αLc1,
(13.1)

where c1 ⩽ 1 is a constant from Lemma 18 (recall that α′ = L−2+ɛ). Indeed, since gL(p) is a standard Gaussian random variable, the probability that

|1α2fL(p)+αgL(p)|αLc1

at a given point p is Lc1. By the union bound, the probability that this happens somewhere on Cr(α) is

Lc1|Cr(α)|Lc1+2CεLc1/2,

provided that ɛ is sufficiently small.

Thus, Lemma 12 applied to the functions fL and f̃L yields that

N(f̃L)=NI(f̃L)+NII(f̃L)+NIII(f̃L)

on the major part of the probability space where gLC3(log L) and where estimate (13.1) holds. The first term on the RHS, NI(f̃L), comes from the stable connected components of Z(fL). Hence, on the large part of the probability space, the fluctuations in N(f̃L) come only from the blinking circles NII(f̃L) and from the Bogomolny–Schmit loops NIII(f̃L), and after we have fixed the function fL, both these quantities depend only on the configuration of (random) signs of f̃L(p) and p ∈ Cr(α).

To make these random signs independent, using Lemma 18, we choose a collection of independent standard Gaussian random variables ξ(p), p ∈ Cr(α), so that

maxpCr(α)|gL(p)ξ(p)|Lc1.
(13.2)

Denote by Ω′ the event that gLC3logL, and both estimates (13.1) and (13.2) hold. Then,

τ(L)=defP(Ω\Ω)=o(1),L,

while on Ω′, we have

sgnf̃L(p)=s(p)=defsgn1α2fL(p)+αξ(p),pCr(α).

Note that the random signs s(p) are independent and that there exists p0(0,12] such that on Cr(α′) each of the two possible values of s(p) is attained with probability at least p0.

For any subset Z ⊂ Cr(α), we let sZ=s(p)pZ. As before, we use the notation

CrS(α)={pCr(α):pis a saddle point of fL},CrE(α)={pCr(α):pis a local extremum of fL}.

Since, on Ω′, NII(f̃L) depends only on sCrE(α) and NIII(f̃L) on sCrS(α), there exist functions ÑII(sCrE(α)) and ÑIII(sCrS(α)) such that, on Ω′, we have NII(f̃L)=ÑII(sCrE(α)) and NIII(f̃L)=ÑIII(sCrS(α)). The function NI(f̃L) stays constant on Ω′, and by ÑI, we denote the value of that constant. Denoting by χΩ′ the indicator function of the event Ω′, we get

E(N(f̃L)m)2Ω[NII(f̃L)+NIII(f̃L)(mNI(f̃L))]2χΩdP=Ω[ÑII+ÑIII(mÑI)]2E[χΩsCr(α)]dP.

The conditional expectation E[χΩsCr(α)] can be written as Q(sCr(α)), where Q is a function on a finite set SCr(α) of all possible collections of signs sCr(α). Thus,

E(N(f̃L)m)2Ω[ÑII+ÑIII(mÑI)]2QdP.

Note that E[Q]=P(Ω)=1τ(L).

First, we consider the case when |CrS(α)|12|Cr(α)| and look at the fluctuations in the number of the Bogomolny–Schmit loops. We use a decoupling argument similar to the one introduced in Sec. XII A 1. We decompose sCr(α)=(sCrE(α),sCrS(α)) and let

X=ω1Ω:ΩQ(sCrE(α)(ω1),sCrS(α)(ω2))dP(ω2)12τ(L).

Then, by Lemma 20, P(X)12, and therefore,

E(N(f̃L)m)212ess infω1XΩÑIII(sCrS(ω2))(mÑIÑII(sCrE(ω1)))2Q(sCrE(ω1),sCrS(ω2))dP(ω2).

We fix ω1X′ and the corresponding signs sCrE(α)(ω1) and consider the graph G(V, E) introduced in (6.6). The vertices of this graph are the joints J(p, δ) with p ∈ CrS(α) and δ = c(AΔ)−1 with sufficiently small positive constant c. The edges are connected components of the set

Z(fL)\pCrS(α)J(p,δ)

that touch the boundaries ∂J(p, δ). The random states σv are defined by the signs s(p), and, by construction, are independent. Furthermore, for the vertices v corresponding to the set CrS(α′), the probabilities of the states σv lie in the range [p0, 1 − p0]. Then, Lemma 21 yields that

infmRE(N(f̃L)m)2|CrS(α)|12Lcε.

This finishes the proof of our theorem in the case when |CrS(α)|12|Cr(α)|.

It remains to consider the case when at least half of the critical points in Cr(α′) are local extrema. In this case, we use the decomposition Cr(α) = (Cr(α) \CrE(α′)) ⊔ CrE(α′) and once again combine Lemma 20 with Lemma 19. Let

X=ω1Ω:ΩQ(sCr(α)\CrE(α)(ω1),sCrE(α)(ω2))dP(ω2)12τ(L).

Then, by Lemma 20, P(X)12. We fix ω1X′ and the corresponding sCr(α)\CrE(α)(ω1). The value of ÑII is the number of p ∈ CrE(α′) such that the sign s(p) is opposite to the sign of the eigenvalues of HfL(p). To each p ∈ CrE(α′), we associate a Bernoulli random variable

ηp=1,s(p)is opposite to the sign of the eigenvalues of HfL(p),0,otherwise.

Since the signs s(p) are independent, the variables ηp are independent as well. Recall that everywhere on Cr(α′), each of two possible values of s(p) is attained with probability ⩾p0, and note that

ÑII=pCrE(α)ηp.

Then, Lemma 19 does the job. This finishes off the proof of our theorem in the second case when |CrE(α)|12|Cr(α)|.□

As we have already mentioned in the Introduction, our theorem does not straightforwardly apply to the ensemble of Gaussian spherical harmonics. Here, we will outline minor modifications needed in this case.

The spherical harmonic fn is an even (when its degree n is even) or odd (when n is odd) function. Hence, its zero set Z(fn) is symmetric with respect to the origin. Hence, for the advanced readers, we are just working on the projective space RP2 instead of the sphere. For the rest of the readers, the critical points of fn come in symmetric pairs.

Instead of distances between critical points, we now have to talk about distances between symmetric pairs of points on S2(n).

The values fn(p) and fn(−p) are equal up to the sign (+ if fn is even and − if fn is odd), while for n1−-separated pairs (p1, −p1) and (p2, −p2), the random variables fn(p1) and fn(p2) are almost independent.

When applying Lemma 1 to replace gn(p) by ξ(p), we keep the relation between ξ(p) and ξ(−p) the same as between gn(p) and gn(−p), i.e., they coincide up to a sign, and make ξ(p) and ξ(p′) independent for p ≠ ± p′.

The rest of the argument goes as before with one simplification and two minor caveats.

The simplification is that for the spherical harmonics ensemble, the blinking circles cannot occur: by the classical Faber–Krahn inequality, the area of any nodal domain of a spherical harmonic of degree n on the sphere S2(n) cannot be less than a positive numerical constant, while the blinking circles are contained in the spherical disks D(p, δ), p ∈ Cr(α), of radius δ = c(AΔ)−1 = o(1) as n. Another way to see that there are no blinking circles is to recall that all local minima of a spherical harmonic are negative, and all local maxima are positive. Let the signs of the eigenvalues of the Hessian Hfn(p), p ∈ Cr(α), coincide, suppose that they are positive, that is, p is a local minimum of fn, and therefore, fn(p) < 0. By Lemma 6, fn and f̃n remain positive on ∂D(p, δ) and convex in D(p, δ). By Lemma 5, f̃n has a critical point ptD(p, δ), which is its local minimum. If the blinking circle occurs, when it disappears, f̃n should stay positive everywhere in D(p, δ), in particular, at pt, yielding the contradiction. Hence, we need to treat only the Bogomolny–Schmit loops.

Both caveats pertain to the Proof of Lemma 21, which estimates from below the fluctuations in the number of random loops. First of all, we note that since all steps of our construction were symmetric with respect to the mapping x ↦ −x, the results it produces are also symmetric. In particular, the joints J(p, δ) and J(−p, δ) are symmetric, and the set of connected components of

Z(fL)\pCrS(α)J(p,δ)

that touch the boundaries ∂J(p, δ) is also symmetric. Hence, the graph G(V, E), to which Lemma 21 was applied, is symmetric as well.

When defining marked cycles, we cannot choose a cycle passing through antipodal vertices, i.e., having common vertices with the symmetric cycle −. Fortunately, this does not happen often: there are at most eight faces f such that f and f have a common vertex v (the case in which −v will be also a common vertex of f and f). This follows from the following lemma.

Lemma 22.

LetX,XS2be two closed symmetric (with respect to the inversionx ↦ −x) non-empty symmetric sets. Then,XX′ ≠ .

First, we conclude our argument and then will prove Lemma 22. Let f and f have a pair of common vertices v and −v on their boundaries, and then, we can join v and −v by a path γ with γ\{v,v}f and by the path −γ with γ\{v,v}f. Put X = γ ∪ −γ. This is a symmetric closed connected subset of S2. If f is another such face (different from ±f), then we have another symmetric closed connected set X′, and by Lemma 22, XX′ ≠ . Since (γ ∪ −γ) \{v, −v} is contained inside ff, while (ff)(ff)=, we see that v and −v must be vertices on f(f) as well. Recalling that each vertex on our graph has degree 4, we see that there are at most eight such “bad faces” f.

Proof of Lemma 22.

Assume that X1X2 = . Without loss of generality, we assume that the set X2 contains the North and the South Poles. Since X1 and X2 are compact, there exists ɛ > 0 such that dist(X1, X2) > 6ɛ.

Take any point zX1. Then, −zX1 as well. Since X1 is connected, there exists a finite chain of points in X1z = z0, z1, …, zn = −z, with d(zj, zj−1) < ɛ (j = 1, …, n). Connecting zj−1 to zj by the shortest arc, we get a curve γ1 joining z to −z and staying in the ɛ-neighborhood of X1. Let γ = γ1 ∪ (−γ1). Then, γ is a symmetric curve going from z to −z and back and staying in the ɛ-neighborhood of X1.

In a similar way, we can construct a curve from the North Pole to the South Pole staying in the ɛ-neighborhood of X2. Let γ2 be the piece of that curve from the last intersection with the circle of radius ɛ around the North Pole to the first intersection with the circle of radius ɛ around the South Pole. Note that γ2 and both these circles stay in the ɛ-neighborhood of X2 and, thereby, are disjoint with γ.

Now, take the projection
(x1,x2,x3)x1x12+x22,x2x12+x22,x3
of S2{|x3|cosε} onto the cylinder C={x12+x22=1,|x3|cosε}. Note that it preserves the symmetry with respect to the origin, so we get two disjoint curves γ̃ and γ̃2 on this cylinder such that γ̃=γ1̃(γ1̃), where γ1̃ goes from some point z̃C to z̃ and stays at positive distance from the edge circles {x3 = ±cos ɛ} of C, while γ̃2 joins those edge circles.

Consider the universal covering map p:SC, where S={(t,s)R2:|s|cosε} is a horizontal infinite strip and p((t,s))=(cos2πt,sin2πt,s). Note that, for Z, p((t+12+,s))=p((t,s)). By the path lifting lemma, γ̃2 is lifted to some curve Γ2 on S joining the top and the bottom boundary lines. The curve γ̃1 is lifted to some curve Γ1(τ), τ ∈ [0, 1], joining (t0, s0) with (t0 + ξ, −s0), where ξZ+12. Then, the curve Γ1*(τ)=(t(τ)+ξ,s(τ)) extends Γ1 and projects to γ̃1. This extension process can now be repeated and done in both directions, so we get a curve Γ in S staying away from the boundary and such that the first coordinate of Γ goes from − to + (if ξ < 0, we reorient Γ). We still have Γ ∩Γ2 = , so the increment of arg(ww2), as w runs over Γ and w2 ∈ Γ2 stays fixed, should not depend on w2. However, this increment is +π when w2 is on the top boundary line of S and −π when w2 is on the bottom line. This contradiction proves the lemma.□

The second caveat is caused by the fact that good cycles now come in symmetric pairs, and the cycles in each pair simultaneously either merge other cycles or remain separate. Hence, our Bernoulli random variables ηp are now valued in {0, 2} instead of {0, 1}.

In memory of Jean Bourgain.

We are grateful to Dmitry Belyaev, Ron Peled, Evgenii Shustin, and Boris Tsirelson for helpful discussions and suggestions.

This work was partially supported by the U.S. NSF (Grant No. DMS-1900008, F.N.) and ERC Advanced Grant No. 692616 (M.S.)

1.
D.
Beliaev
,
M.
McAuley
, and
S.
Muirhead
, “
Fluctuations of the number of excursion sets of planar Gaussian fields
,” arXiv:1908.10708.
2.
E.
Bogomolny
and
C.
Schmit
, “
Percolation model for nodal domains of chaotic wave functions
,”
Phys. Rev. Lett.
88
,
114102
(
2002
).
3.
F.
Nazarov
, “
Local estimates for exponential polynomials and their applications to inequalities of the uncertainty principle type
,”
St. Petersburg Math. J.
5
,
663
717
(
1994
).
4.
F.
Nazarov
and
M.
Sodin
, “
On the number of nodal domains of random spherical harmonics
,”
Am. J. Math.
131
,
1337
1357
(
2009
).
5.
F.
Nazarov
and
M.
Sodin
, “
Fluctuations in random complex zeroes: Asymptotic normality revisited
,”
Int. Math. Res. Not. IMRN
2011
,
5720
5759
.
6.
F.
Nazarov
and
M.
Sodin
, “
Asymptotic laws for the spatial distribution and the number of connected components of zero sets of Gaussian random functions
,”
Z. Mat. Fiz. Anal. Geom.
12
,
205
278
(
2016
).