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.
I. INTRODUCTION
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
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,
However, the rigorous bounds we are aware of are much weaker,
with some σ > 0. The upper bound with 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 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 .
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 = ℓ} and of the excursion set {F ⩾ ℓ} of the random plane wave (which is a scaling limit of the ensemble of spherical harmonics) for non-zero levels ℓ ≠ 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.
II. THE SETUP AND THE MAIN RESULT
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 of large radius L and is normalized by for all . 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
where dL is the spherical distance on . We call such an ensemble (fL) regular if the following two conditions hold:
-smoothness: with estimates uniform in L and with some ν > 0.
Power decay of correlations: , , with some γ > 0 and with the implicit constant independent of L.
Condition (1) yields that almost surely, with estimates uniform in L. We also note that condition (2) is equivalent to the estimate |kL(d)| ≲ (1 + d)−γ for 0 ⩽ d ⩽ πL (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 . By N(fL), we denote the number of these loops.
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 ⩽ (π − ɛ)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 instead of the sphere.
III. MAIN STEPS IN THE PROOF
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
which has the same distribution as fL. Let α be another small parameter, which is significantly bigger than α′, α′ ≪ α ≪ 1, and let
Then, as we will see, with high probability, given fL, the change in the topology of the zero set is determined by the signs of at Cr(α), that is, by the collection of the random values . To make the correlations between these random values negligible, the set Cr(α) should be well-separated on the sphere . At the same time, the set Cr(α) has to be relatively large; otherwise, the impact of fluctuations in signs of on the number will be negligible. In Lemma 17, we will show that
there exist positive ɛ0, c, and C such that, given 0 < ɛ ⩽ ɛ0 and L ⩾ L0(ɛ), for L−2+ɛ ⩽ α ⩽ L−2+2ɛ, with probability very close to 1, the set Cr(α) is L1−Cɛ-separated and |Cr(α)| ⩾ Lcɛ.
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 at Cr(α) affect the topology of the zero set , 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 depends only on the signs of the eigenvalues of the Hessian ∇2fL(p) and the signs of , p ∈ Cr(α). We will describe how these signs determine the structure of the zero set in small neighborhoods of the points p ∈ Cr(α). Outside these neighborhoods, the zero lines of 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, either consists of a simple loop encircling the point p when the sign of is opposite to that of the eigenvalues of ∇2fL(p) or is empty when these signs coincide. We call such connected components of 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 . 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 determines how the graph G(fL) is turned into a collection of loops in , which we will call the Bogomolny–Schmit loops. Figure 1 illustrates how the sign of at the saddle point p ∈ Cr(α) determines the structure of the zero set in a small neighborhood of the saddle point p ∈ Cr(α).
The sign of determines the structure of the zero set near the saddle point p.
We see that in both cases, the fluctuations in the number of connected components of are caused by fluctuations in the signs,
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 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.
IV. NOTATION
Throughout the paper, we will be using the following notation:
L is a large parameter that tends to +∞. We always assume that L ⩾ 1.
denotes the sphere in centered at the origin and of radius L, while, as usual, denotes the unit sphere in . By dL, we denote the spherical distance on . By , we denote the open spherical disk of radius ρ centered at x.
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(L−C) with every C > 0, and w.h.p. for “with high probability,” which means the property in question holds outside an event of probability O(L−c) with some c > 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 (-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.
A ≲ B means A ⩽ CB, A ≳ B means A ⩾ cB, and A ≃ B means that A ≲ B and A ≳ B simultaneously.
The sign ≪ means “sufficiently smaller than” and ≫ means “sufficiently larger than.” For instance, the assumption “given A and B such that A ≪ B” means that there exists c ∈ (0, 1) such that the corresponding conclusion holds for every positive A and B satisfying A ⩽ cB.
V. PRELIMINARIES
For convenience, we collect here standard facts that we will be using throughout this work.
A. Local coordinates
It will be convenient to associate with each point its own coordinate chart. For , let Πp be a plane in passing through the origin and orthogonal to p. The Euclidean structure on Πp is inherited from . By , we denote the hemisphere of centered at p. By
we denote the map inverse to the orthogonal projection. Note that
whenever .
Let 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 means that, for every , , and
Obviously, . In the other direction, it is not difficult to see that if , then .
B. Statistical properties of the gradient and the Hessian
Fix and the orthogonal coordinate system (X1, X2) on the plane Πp, and set , 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.
Let f be a C2+ν-smooth random Gaussian function on the sphere whose distribution is invariant with respect to the isometries of the sphere. Then, the following Gaussian random variables are independent:
f(p) and ∇f(p), as well as ∇f(p) and ∇2f(p),
∂1f(p) and ∂2f(p),
f(p) and , and
and , as well as and .
2. Non-degeneracy
Again, we assume that p is the North Pole of the sphere . Put , and consider the Gaussian functions FL = fL◦ΨL defined in the disks . The corresponding covariances are C3+ν,3+ν-smooth on with some ν > 0. Their partial derivatives up to the third order are bounded locally uniformly in L ⩾ L0. Hence, by a version of the Arzelá–Ascoli theorem, any sequence contains a locally uniformly C2+ν,2+ν-convergent subsequence.
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 priori C3,3-smoothness of the covariance yield the power decay of correlations between the Gaussian vectors
and
[we keep fixed the coordinate systems (X1, X2) and (Y1, Y2) in the planes Πp and Πq].
Put , where Fp = fL◦Ψp and , where . We have , and .
There is nothing to prove if dL(p, q) ⩽ 1, so we assume that dL(p, q) ⩾ 1. Then, and . Now, we use the classical Landau–Hadamard inequality in the following form: if is a C2-smooth function and Mj = max[0,1]|h(j)|, 0 ⩽ j ⩽ 2, then . Applying this to the functions Xi↦Kp,q(X, Y) and Yj↦Kp,q(X, Y), we get and , i, j = 1, 2. Applying the Landau–Hadamard inequality again, this time to the functions , we get , 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 parameter γ > 0 is chosen so that the correlations between the Gaussian vectors (fL(p), ∇fL(p)) and (fL(q), ∇fL(q)) decay as .
C. A priori smoothness of fL
Let (fL) be a regular Gaussian ensemble. Quite often, we will be using the following a priori bound:
w.o.p, .
This bound immediately follows from the classical estimate
For a self-contained proof, see Secs. A9–A11 of Ref. 6.
VI. SMOOTH FUNCTIONS WITH CONTROLLED TOPOLOGY OF THE ZERO SET
Here, we introduce the (non-random) class C3(A, Δ, α, β) of smooth functions f on such that the number of connected components of the zero set of a small perturbation of f can be recovered from the values of at the critical points of f with small critical values (provided that the values of at these points are not too small). Later, we will show that w.h.p., our random function fL belongs to this class.
A. The sets Cr(α), Cr(α, β), and Cr(α, β, Δ) and the class C3(A, Δ, α, β)
Given α, β ⩽ 1 and Δ ⩾ 1, we let
and
where ‖⋅‖6op stands for the operator norm.
By C3(A), we denote the class of C3-smooth functions f on with . Given the parameters
by C3(A, Δ, α, β), we denote the class of functions f ∈ C3(A) for which Cr(α, β) = Cr(α, β, Δ), i.e., the Hessian of f does not degenerate [] on the almost singular set Cr(α, β) where f and ∇f are simultaneously small.
Given f ∈ C3(A, Δ, α, β), α′ ≪ α, and g ∈ C3(A), we set
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).
B. Near any almost singular point there is a unique critical point of f
the spherical disk D(p, 2Δβ) contains a unique critical point z of f,
there are no other critical points of f in the disk , and
|f(z)| ⩽ 2α, provided that AΔ2β2 ≪ α.
C. Near any point p ∈ Cr(α) there is a unique critical point of ft
At last, by part B of Lemma 4, there are no other critical points of ft at distance ⩽c(AΔ)−1 from p.□
D. Local matters
Given f ∈ C3(A), p ∈ Cr(α), and , we look at the behavior of ft in the δ-neighborhood of p. As above, ft = f + tg, with g ∈ C3(A) and 0 ⩽ t ⩽ α′. Throughout this section, we assume that the parameters α′, α, δ, A, and Δ satisfy the set of conditions
and that
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 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 .
Suppose that the eigenvalues of the Hessian have the same sign and that conditions (6.1) and (6.2) hold. Then,
the function ft is either concave or convex function in D(p, δ), and
the function ft does not vanish on ∂D(p, δ), and moreover, the sign of coincides with the sign of the eigenvalues of .
Then, Z(ft) ∩ ∂D(p, δ) = ∅.
Z(ft) ∩ D(p, δ) is either empty, homeomorphic to , or a singleton.
Suppose that |ft(p)| ≳ A(Δα)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 have the same sign, and Z(ft) ∩ D(p, δ) is homeomorphic to whenever ft(p) and the eigenvalues of 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 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 , we denote the quadratic form generated by the Hessian . WLOG, we assume that
We take δ = c(AΔ)−1 with a sufficiently small positive constant c, set
and call this set a joint. By
we denote the curvilinear part of the full boundary ∂J(δ) of the joint J(δ) (Fig. 2).
The set consists of four disjoint curvilinear quadrangles. We call them terminals and denote them by Ti, 1 ⩽ i ⩽ 4.
Under the same assumptions as in Lemma 7, each of the sets Z(Ft) ∩ Ti consists of one curve, which joins the vertical segments on the boundary of Ti.
Under the same assumptions as in Lemmas 7 and 8, by Lemma 5, the joint J(δ) contains only one critical point of Ft and |Xt| ≪ Δα. Consider the sets
Since
it is easy to see that both sets are the segments.
Under the same assumptions as in Lemmas 7 and 8, the only extremum of the restriction of the function Ft to the segment I1 is a local minimum at , and the only extremum of the restriction of the function Ft to the segment I2 is a local maximum at .
The proof of the second statement is almost identical and we skip it.□
Suppose that Ft(Xt) ≠ 0 [i.e., zero is not a critical value of the restriction of the function Ft to the joint J(δ)]. Then, under the same assumptions as in Lemmas 7–9, the set Z(Ft) ∩ J(δ) consists of two connected components, which enter and exit the joint J(δ) through the terminals Ti.
Furthermore, the set {Ft ≠ 0} ∩ J(δ) consists of three connected components. One of them contains Ft(Xt), while on the other two components, Ft has the sign opposite to the sign of Ft(Xt).
Since zero is not a critical value of the restriction , 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 . 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)|≳ A(Δα)2. Then, the type of the joint J(p, δ) coincides with the sign of ft(p).
E. Global matters: The gradient flow
Fix the functions f ∈ C3(A, Δ, α, β) and g ∈ C3(A). Let ft = f + tg and , and consider the gradient flow zt, 0 ⩽ t ⩽ α′, defined by the ordinary differential equation (ODE),
with the initial condition z0 ∈ Z(f).
Suppose that AΔ2β2 ≪ α ≪ (AΔ)−2β and Aα′ ≪ α. Let δ = c(AΔ)−1 with a sufficiently small constant c > 0. Then, for any arc I ⊂ Z(f) \⋃p∈Cr(α)D(p, 2δ2), the flow zt provides a C1-homotopy of I onto an arc . Vice versa, for any arc , the inverse flow zα′−t provides a C1-homotopy of onto an arc I ⊂ Z(f) \⋃p∈Cr(α)D(p, δ2). Moreover, these homotopies move the points by at most O(Aα′/β).
Since the function belongs to the class C3(2A, 2Δ, 2α, 2β), the same arguments can be also applied to the inverse flow zα′−t.□
F. The upshot
We start with functions f ∈ C3(A, Δ, α, β) and g ∈ C3(A) and consider the perturbation . We assume that the parameters
satisfy the following relations:
[which particularly yield conditions (6.1)]. We also assume that the perturbation is not too small on Cr(α),
We set
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 that do not intersect the set
We denote by NI(f) the number of stable loops in Z(f) and by the number of stable loops in .
Observe that , p ∈ CrS(α), and that . 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 and yields a one-to-one correspondence between the set of stable loops in Z(f) and the set of stable loops in . That is, .
2. Blinking circles
These are small connected components of that surround the points p ∈ CrE(α) and lie in the interiors of the corresponding disks D(p, δ). Recall that, by Lemma 6, 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
3. The Bogomolny–Schmit loops
This is the most interesting part of . Consider the graph G = G(f) embedded in . The vertices of G are the joints J(p, δ),p ∈ CrS(α). The edges are connected components of the set
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 , 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 , we denote the number of loops in this collection.
4. The main lemma of Sec. VI
At last, we are able to state the main result of this section.
VII. LOWER BOUNDS FOR THE HESSIAN OF fL ON THE ALMOST SINGULAR SET
Now, we return to regular Gaussian ensembles (fL).
Denote by μ = μ(x) the eigenvalue of the Hessian matrix ∇2fL with the minimal absolute value and by w = w(x) the corresponding normalized eigenvector. Assume that (recall that this holds w.o.p.), and suppose that, for some x ∈ Cr(α, β) and L ⩾ L0, |μ(x)| < Δ−1, where Δ = L3ɛ. We will show that then, the set Cr(5α, 4β) contains a subset with . This will immediately imply the lemma.
The next lemma gives us a lower bound for the probability that a given point belongs to the set
Let p be the probability that a given point belongs to the set Cr(α, β). By the invariance of the ensemble (fL), this probability does not depend on x.
Fix 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.
VIII. THE TWO-POINT FUNCTION OF THE SET Cr(α, β)
In this section, we will look at the set Cr(α, β) of almost singular points of fL,
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 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
By p(x, y), we denote the probability that two given points 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.
A. Estimates of the two-point function
B. Proof of the short-distance estimate (8.1)
1. Beginning the proof
Let . 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,
Then,
where
Note that vol(Ω) ≲ α2β4 ≃ p2. 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),
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 the covariance matrix of ∇fL(x) and put
Since the matrix is non-degenerate uniformly in L ⩾ L0 (Lemma 2), the matrix is also non-degenerate uniformly in L ⩾ L0.
By our assumption on the power decay of correlations, we have
Then, provided that dL(x, y) ⩾ d0 with d0 ≫ 1, we get
and therefore,
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 , and let and . Then,
whence
By the compactness of the unit sphere in , there exist , |a|2 + |b|2 = 1, such that
Here and until the end of the Proof of Lemma 15, ‖⋅‖ stands for the L2-norm, i.e., for the Gaussian random variable η.
2. The normal and the tangential derivatives
Let be the big circle on that passes through the points x and y, and let be the shortest of the two arcs of with the endpoints x and y. We orient by moving from x to y along and choose the coordinate systems in Πx and Πy so that one coordinate vector is parallel to the tangent to (at x and y correspondingly), while the other one is orthogonal to . We keep the same orientation for both coordinate systems. We denote by ∂‖ the derivative along and by ∂⊥ the derivative in the normal direction to and decompose
where and . Then, by (8.3),
where and , similarly for b′ and b″, and |a′|2 + |a″|2 + |b′|2 + |b″|2 = 1.
Now, consider another Gaussian six-dimensional vector
Since the distribution of fL is invariant with respect to orthogonal transformations, the Gaussian vectors v(x, y) and have the same covariance matrix in the chosen coordinate systems in Πx and Πy. Therefore,
We split the rest of the proof of estimate (8.1) into two cases: (i) and (ii) . 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):
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,
Since , at least one of the following holds:
either or .
We assume that (the other case is similar and slightly simpler), let e = a″ + b″, , and notice that
We take the point , z ≠ x, 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,
whence
Put x0 = x and x1 = z, then take the point , x2≠x0, 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,
On the other hand, since d0 is the correlation length and dL(x0, xN) ⩾ d0, we have
(at the last step, we use that the distribution of v⊥ does not degenerate uniformly in L ⩾ L0 and that ). Therefore, λ ≳ N−2.
Recalling that by the definition of N, we have (N − 1)dL(x, y) < d0, and we get λ ≳ dL(x,y)2, which concludes our consideration of the first case.
4. Case (ii):
In this case, we restrict the function fL to the big circle and treat it as a periodic random Gaussian function with translation-invariant distribution. To simplify the notation, we omit the index L. By ρ, we denote the spectral measure of F, that is,
where correspond to the points . Then, we have
where . Furthermore,
by the uniform non-degeneracy of ∇fL and
by the power decay of correlations of fL. Note that since the function F is 2πL-periodic, the Fourier transform of its spectral measure is also 2πL-periodic.
Recalling that , we notice that 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,
and denote ‖pδ‖W = |a1| + |a2| + |b1| + |b2| (recall that an exponential sum is the expression
where λj are real numbers and qj are polynomials in ξ with complex coefficients). Then, the following lemma does the job.
C. Proof of Lemma 16
1. Beginning the Proof of Lemma 16
Let . Then, by Chebyshev’s inequality,
Take B = δ0A and let 0 < δ ⩽ δ0. Then, [−A, A] ⊂ [−Bδ−1, Bδ−1]. Let κ > 0 be a sufficiently small parameter, which we will choose later, and consider the set
We claim that
the set Ξ is a union of at most B + 5 intervals Ij, 1 ⩽ j ⩽ B + 5, and
the length of each interval Ij is ⩽ C(B, δ0) κ.
Having these claims, we will show that , whence
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 cannot exceed
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 [−Bδ−1, Bδ−1] does not exceed
Hence, the set Ξ consists of at most 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 I ⊂ J,
where C is a numerical constant. Applying this lemma to the exponential sum pδ of degree 4 and to each of the intervals Ij ⊂ [−Bδ−1, Bδ−1], we get
On the other hand,
Thus, |Ij| ⩽ C(B, δ0)κ, proving (b).
4. Completing the Proof of Lemma 16
Recall that and that given κ > 0, we defined the set
satisfying (a) and (b). Then, we have the following alternative:
either or .
In the first case,
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 . Then, . 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 ≪ S ≪ κ−1 and estimate the integral
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 ξ ∈ I, we have
provided that S ⋅ κ is sufficiently small. Therefore, for |s| ⩽ S, we have
whence
On the other hand, using the identity
with φ(s) = (1 −|s|/S)+ and noting that the Fourier transform of this function is non-negative, we get
Then, recalling that the function is 2πL-periodic and that for |s| ⩽ L/2 and assuming without loss of generality that , we get
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.□
D. Proof of the long-distance estimate (8.2)
As above, we denote by the covariance matrix of ∇fL(x) and put
Since the matrix is non-degenerate uniformly in L ⩾ L0, the matrix is also non-degenerate uniformly in L ⩾ L0.
We assume that dL(x, y) ⩾ d0, where d0 is sufficiently large (and independent of L). Then, we have
Therefore,
and
Recall that
where
and that
Hence,
completing the proof of estimate (8.2) in Lemma 15.□
IX. STRUCTURE OF THE SET Cr(α) WITH L−2+ɛ ⩽ α ⩽ L−2+2ɛ
Now, we are ready to prove our main lemma:
There exist positive ɛ0 and c and positive C such that, given 0 < ɛ ⩽ ɛ0 and L ⩾ L0(ɛ), for L−2+ɛ ⩽ α ⩽ L−2+2ɛ, w.h.p., the set Cr(α) is L1−Cɛ-separated, and |Cr(α)| ⩾ Lcɛ.
A. w.h.p., the set Cr(α) is L1−Cɛ-separated
In this part, we assume that α ⩽ L−2+2ɛ, choose β and ρ so that
and fix a maximal ρ-separated set on . Then, .
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 and with the w.h.p.-estimate provided by Lemma 13. Hence, we need to estimate the probability of the event
with an appropriately chosen constant C.
Suppose that the event occurs. Denote by x1 and x2 the closest to z1 and z2 points in . Then,
and
i.e., x1, x2 ∈ Cr(2α, β). We claim that
the mean number of pairs of points satisfying (9.1) is bounded from above by L−ɛ.
By Chebyshev’s inequality, this yields that the probability that there exists at least one such pair is also bounded from above by L−ɛ, which proves the L1−Cɛ-separation.
The mean we need to estimate equals
where is the two-point function estimated in Lemma 15. By Lemma 15, p(x1, x2) ≲ L4ɛΘp2, so the whole sum is
provided that the constant C is sufficiently large. This proves that the set Cr(α) is L1−Cɛ-separated.□
B. w.h.p., |Cr(α)| ⩾ Lcɛ
In this part, we assume that α ⩾ L−2+ɛ and introduce the parameters β, Δ, and r, satisfying
We fix a maximal r-separated set on . Then, the disks D(x, r), , cover with a bounded multiplicity of covering and . We set
W.o.p., for L ⩾ L0, we have . Then, by Lemma 4, each disk D(y, r), y ∈ Y, contains a unique critical point z ∈ Cr(α), and
provided that L ⩾ L0. This yields two useful observations that hold w.o.p.:
|Cr(α)|≳|Y|, and
if y1, y2 ∈ Y, then either dL(y1, y2) ⩽ 2r and the number of such pairs (y1, y2) is or d(y1, y2) ⩾ L−ɛ. Indeed, if the points y1, y2 ∈ Y 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 is cΔ−1(log L)−1-separated, and thus, in this case, d(y1, y2) ⩾ L−ɛ.
In the following, we will show that, for sufficiently large L,
and that
These two estimates combined with the first observation readily yield what we need.
1. Estimating
This estimate is straightforward,
2. Estimating Var[|Y|]
In this section, p and p(x, y) will denote the one- and two-point functions of the set . Given , we put
Then, and
The first sum on the RHS is bounded by
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) < L−ɛ, the terms with L−ɛ ⩽ dL(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 , we bound this sum by .
b. The terms with 2r < dL(x, y) < L−ɛ.
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 with 2r < dL(x, y) < L−ɛ is O(L−C) with any positive C. Thus, in this range, pΔ(x, y) = O(L−C), while the total number of pairs x, y is bounded by , 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 L−ɛ ⩽ dL(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
Then, the whole double sum is
which gives what we needed with a large margin.
d. The terms with dL(x, y) ⩾ Lɛ.
In this case,
By the long-distance estimate (8.2) of the two-point function p(x, y),
while, by Lemma 14,
Thus, , and the whole double sum is bounded by .
This completes the proof of the estimate of Var[ |Y| ] and hence that of Lemma 17.□
X. ASYMPTOTIC INDEPENDENCE
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.
is “a big Gaussian Hilbert space” that contains and countably many mutually orthogonal one-dimensional subspaces that are orthogonal to .
J(z) = fL(z), z ∈ Z, are unit vectors in with
XI. TWO SIMPLE LEMMAS
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.
A. Anticoncentration of the sums of Bernoulli random variables
B. Large sections lemma
In the following, we will apply this lemma, mostly, with X = Ω1 and p = 1.
XII. VARIANCE OF THE NUMBER OF LOOPS
Let G = G(V, E) be a finite graph embedded in the sphere 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 connecting the vertices, and the faces are the connected components of the open set .
Each vertex v ∈ V can be replaced by one of two possible “avoided crossings” at v (Fig. 4).
We call the choice of the avoided crossing at v the state of the vertex v and denote it by σv. When the states are assigned to all vertices in V, the collection of states σV = {σv: v ∈ V} 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 , we denote the probability space on which the random states are defined. Then, N(Γ) = N(Γ(σV)) is a random variable on .
Given , we put and denote by |V(p0)| the cardinality of the set of vertices V(p0).
A. Beginning the Proof of Lemma 21
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 . 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.
By N(Γ′), we denote the number of loops in the collection Γ′. The collection of states turns the graph G(σV′) into a collection of loops . Then, Γ = Γ′ ⊔Γ″, and N(Γ) = N(Γ′) + N(Γ″).
Let φ(σV) be any non-negative bounded measurable function. Since the random variables σV′ and are independent, we have
Hence, for any event X′ ⊂Ω,
where Σ′ = {σV′(ω′): ω′ ∈ X′}. Letting φ = (N(Γ) − m)2Q and taking into account that N(Γ) − m = N(Γ″) − (m − N(Γ′)), we get
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,
close to 1, while 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 of the remaining set of vertices V″, ignoring its dependence on the fixed states σV′ ∈ Σ′.
B. Discarding the vertices v with p(v) < p0 or p(v) > 1 − p0
As above, we let . We put V′ = V\V(p0) and V″ = V(p0) and consider
Then, by Lemma 20 (applied with p = 1), .
Hence, from now on, we assume that for each vertex v of the graph G, we have p0 ⩽ p(v) ⩽ 1 − p0, while .
C. Many faces have at most 4 vertices on the boundary
Denote by F the set of the faces of the graph G. Given a face , we denote by the set of vertices in V that lie on the boundary . The Euler formula gives us
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
We let be the set of all faces having at most 4 vertices on the boundary and conclude that .
D. Choosing a maximal collection of separated faces
We call faces separated if they do not have common vertices on their boundaries: . We fix a maximal collection of separated faces. Since for any face , there are at most 12 other faces with , we conclude from the maximality of that it also contains sufficiently many faces,
E. Marking vertices and cycles
Next, we choose a simple cycle on the boundary of each face and mark a vertex on that cycle according to the following rule: We take a vertex v′ in and, starting at v′, walk along turning left at each vertex so that the face 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).
We will be using the following property of the marked cycles: if the cycle exits in a vertex along an edge e, then it returns to this vertex along an edge, which is one of two edges adjacent to e (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 , which turn into a separate loop.
We denote by VM ⊂ V the set of all marked vertices and by VUM = V\VM the set of all unmarked vertices.
F. Good marked cycles
For any particular assignment of the states of the unmarked vertices, the marked cycle is called good if the edges of merge into an edge of the graph G(σ(VUM)) that connects the vertex with itself (Fig. 7). Note that this event depends only on the states of at most 3 unmarked vertices lying on . Therefore, for any marked cycle , we have . Hence, denoting by NG the number of good cycles, we obtain
Using first the Chebyshev inequality and then the independence of the random states, we get
To simplify the notation, we let and . Then, letting , we see that . Put
(recall that at this moment ). Then, by Lemma 20, , which is , 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, .
G. Discarding bad cycles
All marked cycles 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
By Lemma 20 (applied with X = Ω), . We fix the states of marked bad vertices corresponding to the event X′.
H. Completing the Proof of Lemma 21
We are left with the graph G with vertices at VM.G.. For each vertex v ∈ VM.G., there is “a circular edge” ev with the endpoints at v, which came from the corresponding cycle . 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 of good marked vertices such that that none of the corresponding good cycles turns into a separate loop and denote by the loop ensemble obtained from the graph G after the assignment of the states . Introduce a collection of independent Bernoulli random variables , letting ηv = 0 if and ηv = 1 otherwise. Then,
Applying Lemma 19, we get the uniform in m lower bound,
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 NG ⩾ d(p0)|V|.□
XIII. TYING LOOSE ENDS TOGETHER: PROOF OF THE THEOREM
A. Perturbing fL
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
This is a random Gaussian function equidistributed with fL. We will show that
which immediately yields the lower bound for Var[N(fL)] we are after. Note that
That is, it suffices to show that with probability at least in fL, we have
B. Freezing fL
We will prove a somewhat stronger statement that this inequality holds if the function fL satisfies the following conditions:
fL ∈ C3(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., ),
the set Cr(α) is L1−Cɛ-separated, and
|Cr(α′)| ⩾ Lcɛ.
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 , , etc.
C. Recalling a little Morse caricature
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
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
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
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
at a given point p is . By the union bound, the probability that this happens somewhere on Cr(α) is
provided that ɛ is sufficiently small.
Thus, Lemma 12 applied to the functions fL and yields that
on the major part of the probability space where gL ∈ C3(log L) and where estimate (13.1) holds. The first term on the RHS, , comes from the stable connected components of Z(fL). Hence, on the large part of the probability space, the fluctuations in come only from the blinking circles and from the Bogomolny–Schmit loops , and after we have fixed the function fL, both these quantities depend only on the configuration of (random) signs of and p ∈ Cr(α).
D. Fight for independence
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
while on Ω′, we have
Note that the random signs s(p) are independent and that there exists 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 . As before, we use the notation
Since, on Ω′, depends only on and on , there exist functions and such that, on Ω′, we have and . The function stays constant on Ω′, and by , we denote the value of that constant. Denoting by χΩ′ the indicator function of the event Ω′, we get
The conditional expectation 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,
Note that .
E. Fluctuations generated by the Bogomolny–Schmit loops
First, we consider the case when 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 and let
Then, by Lemma 20, , and therefore,
We fix ω1 ∈ X′ and the corresponding signs 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
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
This finishes the proof of our theorem in the case when .
F. Fluctuations generated by blinking circles
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
Then, by Lemma 20, . We fix ω1 ∈ X′ and the corresponding . The value of is the number of p ∈ CrE(α′) such that the sign s(p) is opposite to the sign of the eigenvalues of . To each p ∈ CrE(α′), we associate a Bernoulli random variable
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
Then, Lemma 19 does the job. This finishes off the proof of our theorem in the second case when .□
XIV. THE CASE OF SPHERICAL HARMONICS
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 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 .
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−Cɛ-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 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 , 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 remain positive on ∂D(p, δ) and convex in D(p, δ). By Lemma 5, has a critical point pt ∈ D(p, δ), which is its local minimum. If the blinking circle occurs, when it disappears, 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
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 ℓ passing through antipodal vertices, i.e., having common vertices with the symmetric cycle −ℓ. Fortunately, this does not happen often: there are at most eight faces such that and have a common vertex v (the case in which −v will be also a common vertex of and ). This follows from the following lemma.
Let be two closed symmetric (with respect to the inversion x ↦ −x) non-empty symmetric sets. Then, X ∩ X′ ≠ ∅.
First, we conclude our argument and then will prove Lemma 22. Let and 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 and by the path −γ with . Put X = γ ∪ −γ. This is a symmetric closed connected subset of . If is another such face (different from ), then we have another symmetric closed connected set X′, and by Lemma 22, X ∩ X′ ≠ ∅. Since (γ ∪ −γ) \{v, −v} is contained inside , while , we see that v and −v must be vertices on as well. Recalling that each vertex on our graph has degree 4, we see that there are at most eight such “bad faces” .
Assume that X1 ∩ X2 = ∅. 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 z ∈ X1. Then, −z ∈ X1 as well. Since X1 is connected, there exists a finite chain of points in X1 z = 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.
Consider the universal covering map , where is a horizontal infinite strip and . Note that, for , . By the path lifting lemma, is lifted to some curve Γ2 on S joining the top and the bottom boundary lines. The curve is lifted to some curve Γ1(τ), τ ∈ [0, 1], joining (t0, s0) with (t0 + ξ, −s0), where . Then, the curve extends Γ1 and projects to . 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(w − w2), 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}.
XV. DEDICATION
In memory of Jean Bourgain.
ACKNOWLEDGMENTS
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.)