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

On the number of nodal domains of homogeneous caloric polynomials

Matthew Badger  and  Cole Jeznach Department of Mathematics, University of Connecticut, Storrs, CT 06269-1009. matthew.badger@uconn.edu School of Mathematics, University of Minnesota, Minneapolis, MN, 55455. jezna001@umn.edu
(Date: January 13, 2024)
Abstract.

We investigate the minimum and maximum number of nodal domains across all time-dependent homogeneous caloric polynomials of degree dd in n×\mathbb{R}^{n}\times\mathbb{R} (space ×\times time), i.e., polynomial solutions of the heat equation satisfying tp0\partial_{t}p\not\equiv 0 and

p(λx,λ2t)=λdp(x,t)for all xnt, and λ>0.p(\lambda x,\lambda^{2}t)=\lambda^{d}p(x,t)\quad\text{for all $x\in\mathbb{R}^{n}$, $t\in\mathbb{R}$, and $\lambda>0$.}

When n=1n=1, it is classically known that the number of nodal domains is precisely 2d/22\lceil d/2\rceil. When n=2n=2, we prove that the minimum number of nodal domains is 2 if d0(mod4)d\not\equiv 0\pmod{4} and is 3 if d0(mod4)d\equiv 0\pmod{4}. When n3n\geq 3, we prove that the minimum number of nodal domains is 22 for all dd. Finally, we show that the maximum number of nodal domains is Θ(dn)\Theta(d^{n}) as dd\rightarrow\infty and lies between dnn\lfloor\frac{d}{n}\rfloor^{n} and (n+dn)\binom{n+d}{n} for all nn and dd. As an application and motivation for counting nodal domains, we confirm existence of the singular strata in Mourgoglou and Puliatti’s two-phase free boundary regularity theorem for caloric measure.

Key words and phrases:
heat equation, caloric polynomial, nodal domain, free boundary regularity
2020 Mathematics Subject Classification:
Primary: 35K05. Secondary: 26C05, 26C10, 35R35.
M. Badger was partially supported by NSF DMS grant 2154047. C. Jeznach was partially supported by Simons Collaborations in MPS grant 563916 and NSF DMS grant 2000288.

1. Introduction

A general motivation for studying caloric polynomials comes from their ubiquity in the theory of parabolic PDEs as finite order solutions of the heat equation. After showing that tangent functions of solutions to a parabolic PDE are homogeneous caloric polynomials (hereafter abbreviated hcps), one can deduce strong unique continuation principles and estimate the dimension of nodal and singular sets of solutions [Che98]. In a similar vein, it was recently shown that zero sets of hcps appear as the supports of tangent measures in non-variational free boundary problems for caloric measure [MP21]. Hcps also arise in geometric contexts such as understanding the dimension of ancient caloric functions on manifolds with polynomial growth [CM21].

In this paper, with a goal of confirming existence of singular strata in the aforementioned free boundary regularity problem for caloric measure (see §1.3), we investigate basic topology of nodal domains of hcps in n+1n×\mathbb{R}^{n+1}\equiv\mathbb{R}^{n}\times\mathbb{R} (space ×\times time). In particular, for each ambient dimension and for each parabolic degree (see §1.1), we would like to determine the minimum and maximum number of possible nodal domains. This type of question has been studied extensively for spherical harmonics (see §1.2 for a brief survey). Also, up to scaling by a constant, for each degree dd, there is a unique hcp in 1+1\mathbb{R}^{1+1} of degree dd and the number of nodal domains is precisely 2d/22\lceil d/2\rceil (see §2). Thus, the problem is to determine what happens for time-dependent hcps in at least two space variables. We fully determine the minimum number of domains (with detailed constructions) and establish asymptotic bounds for the maximum number of domains.

Theorem 1.1 (minimum number of nodal domains).

When n2n\geq 2, the minimum number mn,dm_{n,d} of nodal domains of time-dependent homogeneous caloric polynomials in n+1\mathbb{R}^{n+1} of degree d2d\geq 2 satisfies (see Figure 1.1)

(1.1) m2,d={2,when d0(mod4),3,when d0(mod4).m_{2,d}=\begin{cases}2,&\text{when }d\not\equiv 0\pmod{4},\\ 3,&\text{when }d\equiv 0\pmod{4}.\end{cases}

When n3n\geq 3, we have mn,d=2m_{n,d}=2 for all d2d\geq 2.

Refer to caption
Refer to caption
Refer to caption
Figure 1.1. Gallery of nodal sets of homogeneous caloric polynomials in 2+1\mathbb{R}^{2+1} achieving the minimum number m2,dm_{2,d} of nodal domains. For increased visibility, we show the intersection of the full nodal set with a spherical annulus. Left: (5.7) with d=4d=4, ε=0.2\varepsilon=0.2, α=π/10\alpha=\pi/10. Middle: (5.2) with d=5d=5, ε=0.3\varepsilon=0.3, α=π/10\alpha=\pi/10. Right: (5.1) with d=6d=6, ε=0.05\varepsilon=0.05.

We prove the case n3n\geq 3 of Theorem 1.1 in §3 by modifying a construction from [BET17] for homogeneous harmonic polynomials (hereafter abbreviated hhps). The case n=2n=2 is established in §§4 and 5, partly by employing the perturbation technique from [Lew77]. Notably, the proof that m2,4k=3m_{2,4k}=3 is the most difficult argument in the paper.

Theorem 1.2 (maximum number of nodal domains).

For all n2n\geq 2, the maximum number Mn,dM_{n,d} of nodal domains of time-dependent homogeneous caloric polynomials in n+1\mathbb{R}^{n+1} of degree dd is Θ(dn)\Theta(d^{n}) as dd\rightarrow\infty. More precisely, for all n2n\geq 2 and d2d\geq 2,

(1.2) dnnMn,d(n+dn).\left\lfloor\frac{d}{n}\right\rfloor^{n}\leq M_{n,d}\leq\binom{n+d}{n}.

We prove Theorem 1.2 in §3. The lower bound on Mn,dM_{n,d} in (1.2) is based on an elementary construction using products of hcps in 1+1\mathbb{R}^{1+1}. The upper bound follows from an indirect application of Courant’s nodal domain theorem, exploiting a connection between negative time-slices of caloric polynomials in 1+1\mathbb{R}^{1+1} and Hermite orthogonal polynomials.

Remark 1.3.

Classical theorems in algebraic geometry imply that the maximal number of nodal domains of polynomials p:n+1p:\mathbb{R}^{n+1}\rightarrow\mathbb{R} of degree dd is Θ(dn+1)\Theta(d^{n+1}) as dd\rightarrow\infty and lies between (d0)++(dn+1)\binom{d}{0}+\cdots+\binom{d}{n+1} and (d+1)n(d+2)(d+1)^{n}(d+2); see [Sta07, Proposition 2.4] for the lower bound (from hyperplane arrangements) and [Mil64, Theorem 3] for the upper bound. Thus, perhaps not unsurprisingly, the behavior of hcps is distinct from the behavior of general polynomials. Finding the exact value of Mn,dM_{n,d} appears to be a difficult problem. Except for some low degrees (d6d\leq 6), the corresponding problem for spherical harmonics in 3\mathbb{R}^{3} is also open [Ley96].

We discuss proof strategies for the main theorems and outline the paper in §1.4.

1.1. Definitions and examples

Let us decode the terminology in the statements of the main theorems. We study the heat equation (tΔ)u=0(\partial_{t}-\Delta)u=0 in n+1={(x,t):xn,t}\mathbb{R}^{n+1}=\{(x,t):x\in\mathbb{R}^{n},t\in\mathbb{R}\}, where ΔΔx=1nxixi\Delta\equiv\Delta_{x}=\sum_{1}^{n}\partial_{x_{i}}\partial_{x_{i}}. The natural notion of homogeneity in this setting is anisotropic.

Definition 1.4.

A function f(x,t)f(x,t) is parabolically homogeneous of degree dd\in\mathbb{R} if

(1.3) f(λx,λ2t)\displaystyle f(\lambda x,\lambda^{2}t) =λdf(x,t)for all λ>0 and (x,t).\displaystyle=\lambda^{d}f(x,t)\quad\text{for all $\lambda>0$ and $(x,t)$.}
Definition 1.5.

A polynomial p(x,t)p(x,t) is a homogeneous caloric polynomial (hcp) of degree dd if pp satisfies the heat equation and is parabolically homogeneous of degree d={0,1,2,}d\in\mathbb{N}=\{0,1,2,\dotsc\}. We say that pp is time-dependent if tp0\partial_{t}p\not\equiv 0.

Example 1.6.

For any exponent kk\in\mathbb{N} and multi-index αn\alpha\in\mathbb{N}^{n}, the monomial tkxαt^{k}x^{\alpha} is parabolically homogeneous of degree 2k+|α|=2k+α1++αn2k+\left\lvert\alpha\right\rvert=2k+\alpha_{1}+\cdots+\alpha_{n}. In particular, if p(x,t)p(x,t) is parabolically homogeneous of degree dd\in\mathbb{N}, then each monomial in p(x,t)p(x,t) has the form tkxαt^{k}x^{\alpha} where 2k+|α|=d2k+\left\lvert\alpha\right\rvert=d. Parabolic and algebraic homogeneity are distinct notions; e.g.,

(1.4) p(x,t)=t2+tx2+x4/12p(x,t)=t^{2}+tx^{2}+x^{4}/12

is an hcp of degree 44 in 1+1\mathbb{R}^{1+1}, but pp is not algebraically homogeneous. Nevertheless, the parabolic degree and the algebraic degree of an hcp always coincide.

Remark 1.7.

If pp is a time-dependent hcp, then the degree of pp is at least 2.

Definition 1.8.

The nodal domains of a continuous function u(x,t)u(x,t) on n+1\mathbb{R}^{n+1} are the connected components of the set {(x,t):u(x,t)0}\{(x,t):u(x,t)\neq 0\}. We let 𝒩(u){+}\mathcal{N}(u)\in\mathbb{N}\cup\{+\infty\} denote the number of nodal domains of uu. The nodal set of uu is {(x,t):u(x,t)=0}\{(x,t):u(x,t)=0\}.

Remark 1.9.

Let 𝕊n={(x,t)n+1:x12++xn2+t2=1}\mathbb{S}^{n}=\{(x,t)\in\mathbb{R}^{n+1}:x_{1}^{2}+\cdots+x_{n}^{2}+t^{2}=1\}. By parabolic homogeneity, if uu is an hcp of degree d1d\geq 1 in n+1\mathbb{R}^{n+1}, then 𝒩(u)\mathcal{N}(u) is the number of connected components of {(x,t)𝕊n:u(x,t)0}\{(x,t)\in\mathbb{S}^{n}:u(x,t)\neq 0\} in 𝕊n\mathbb{S}^{n}.

Remark 1.10.

For any n1n\geq 1, the mean value property (e.g., see [Eva10, p. 50]) implies that a non-constant solution u:n+1u:\mathbb{R}^{n+1}\rightarrow\mathbb{R} of the heat equation takes positive and negative values in any neighborhood of a zero of uu. In particular, mn,d2m_{n,d}\geq 2 for all n1n\geq 1 and d2d\geq 2.

Example 1.11.

For any 1jn1\leq j\leq n, the polynomial p(x,t)=2t+xj2p(x,t)=2t+x_{j}^{2} in n+1\mathbb{R}^{n+1} is an hcp of degree 22. Moreover, p(x,t)p(x,t) has exactly two nodal domains: {t>xj2}\{t>x_{j}^{2}\} and {t<xj2}\{t<x_{j}^{2}\}. Thus, for all n1n\geq 1, the minimum number of nodal domains of time-dependent hcps in n+1\mathbb{R}^{n+1} of degree 2 is mn,2=2m_{n,2}=2.

Example 1.12.

Up to scaling by a constant multiple, for every d1d\geq 1, there exists a unique hcp pd(x,t)p_{d}(x,t) of degree dd in 1+1\mathbb{R}^{1+1} and m1,d=M1,d=𝒩(pd)=2d/2m_{1,d}=M_{1,d}=\mathcal{N}(p_{d})=2\lceil d/2\rceil. See §2 for the details.

Example 1.13.

The polynomial

(1.5) p(x,y,t)=150t(3x+y)+27x3+267x2y+144xy264y3p(x,y,t)=150t(3x+y)+27x^{3}+267x^{2}y+144xy^{2}-64y^{3}

is an hcp of degree 3 in 2+1\mathbb{R}^{2+1} and 𝒩(p)=2\mathcal{N}(p)=2. This example can be found by evaluating (5.2) with d=3d=3, ε=1\varepsilon=1, and (cosα,sinα)=(3/5,4/5)(\cos\alpha,\sin\alpha)=(3/5,4/5) and multiplying by a constant to obtain a polynomial with integer coefficients. It can be checked that p(x,y,t)=0\nabla p(x,y,t)=0 if and only if (x,y,t)=(0,0,0)(x,y,t)=(0,0,0).

Example 1.14.

The polynomial

(1.6) p(x,y,t)=7500t2+150t(37x27xy+13y2)+192x4+176x3y+1623x2y2351xy3108y4\begin{split}p(x,y,t)=7500t^{2}&+150t(37x^{2}-7xy+13y^{2})\\ &+192x^{4}+176x^{3}y+1623x^{2}y^{2}-351xy^{3}-108y^{4}\end{split}

is an hcp of degree 4 in 2+1\mathbb{R}^{2+1} and 𝒩(p)=3\mathcal{N}(p)=3. This example can be found by evaluating (5.7) with d=4d=4, ε=1/2\varepsilon=1/2, and (cosα,sinα)=(3/5,4/5)(\cos\alpha,\sin\alpha)=(3/5,4/5) and multiplying by a constant to obtain a polynomial with integer coefficients.

Example 1.15.

The polynomial

(1.7) p(x,y,z,t)=12t2+12tx2+x4+y46y2z2+z4p(x,y,z,t)=12t^{2}+12tx^{2}+x^{4}+y^{4}-6y^{2}z^{2}+z^{4}

is an hcp of degree 4 in 3+1\mathbb{R}^{3+1} and 𝒩(p)=2\mathcal{N}(p)=2. See Proposition 3.1. It can be checked that p(x,y,z,t)=0\nabla p(x,y,z,t)=0 if and only if (x,y,z,t)=(0,0,0,0)(x,y,z,t)=(0,0,0,0).

1.2. Comparison with spherical harmonics and Grushin spherical harmonics

Steady-state solutions of the heat equation on n+1\mathbb{R}^{n+1} correspond to harmonic functions on n\mathbb{R}^{n}. Nodal geometry of homogeneous harmonic polynomials pp in n\mathbb{R}^{n} (also called solid harmonics) and of the so-called spherical harmonics p|𝕊n1p|_{\mathbb{S}^{n-1}} is well-studied. See [EJN07, NS09, Log18a, Log18b] for a short sample, including results for Laplace-Beltrami eigenfunctions on closed Riemannian manifolds beyond the sphere.

Parallel to the quantities mn,dm_{n,d} and Mn,dM_{n,d} defined in Theorems 1.1 and 1.2, we let m~n,d\tilde{m}_{n,d} and M~n,d\tilde{M}_{n,d} denote the minimum and maximum number of nodal domains of hhps in n\mathbb{R}^{n} of degree dd, respectively. In the line (n=1n=1), the only harmonic functions are affine and m~1,1=M~1,1=2\tilde{m}_{1,1}=\tilde{M}_{1,1}=2 trivially. In the plane (n=2n=2), since harmonic functions can be written as the real part of a complex-analytic function, the nodal sets of hhp degree dd are rotations of {(x,y):Re(x+iy)d=0}\{(x,y):\mathrm{Re}(x+iy)^{d}=0\} and m~2,d=M~2,d=2d\tilde{m}_{2,d}=\tilde{M}_{2,d}=2d for all d1d\geq 1. The situation becomes more interesting when n3n\geq 3. Lewy [Lew77] proved that m~3,d=2\tilde{m}_{3,d}=2 whenever d1d\geq 1 is odd and m~3,d=3\tilde{m}_{3,d}=3 whenever d2d\geq 2 is even. An explicit example of a degree 3 hhp with exactly two nodal domains,

(1.8) p(x,y,z)=x33xy2+z3(3/2)(x2+y2)z,p(x,y,z)=x^{3}-3xy^{2}+z^{3}-(3/2)(x^{2}+y^{2})z,

was found independently by Szulkin [Szu79]. Badger, Engelstein, and Toro [BET17] gave a simple construction (utilizing the explicit description of hhps in 2\mathbb{R}^{2}) that shows m~n,d=2\tilde{m}_{n,d}=2 for all n4n\geq 4 and d1d\geq 1, independent of the parity of dd.

When n3n\geq 3, any hhp pp in n\mathbb{R}^{n} of degree dd satisfies the equation

Δ𝕊n1p|𝕊n1=d(d+n2)p|𝕊n1,-\Delta_{\mathbb{S}^{n-1}}p|_{\mathbb{S}^{n-1}}=d(d+n-2)p|_{\mathbb{S}^{n-1}},

where Δ𝕊n1\Delta_{\mathbb{S}^{n-1}} denotes the Laplace-Beltrami operator on the sphere. Courant’s nodal domain theorem asserts that when listed with multiplicity, the mm-th eigenfunction of the Laplace-Beltrami operator on a closed C1C^{1} Riemannian manifold has at most mm nodal domains [CH53, CLMM20]. Since the dimension of the vector space of hhps of degree d2d\geq 2 in n\mathbb{R}^{n} is exactly (n+d1n1)(n+d3n1)\binom{n+d-1}{n-1}-\binom{n+d-3}{n-1} (see [ABR01, Proposition 5.8]), the maximal number of linearly independent hhps of degree at most dd is exactly (n+d1n1)+(n+d2n1)=O(dn1)\binom{n+d-1}{n-1}+\binom{n+d-2}{n-1}=O(d^{n-1}) as dd\rightarrow\infty. Thus, M~n,d=O(dn1)\tilde{M}_{n,d}=O(d^{n-1}) as dd\rightarrow\infty. It is known that the upper bound on Mn,dM_{n,d} provided by Courant’s theorem is not sharp. See [Ley96] for further discussion and the (still to this day) state-of-the-art bounds on M2,dM_{2,d}.

In [LTY15], Liu, Tian, and Yang study the minimum number m~2,dG\tilde{m}^{G}_{2,d} of nodal domains of Grushin spherical harmonics, i.e. parabolically homogenenous polynomial solutions p(x,y,t)p(x,y,t) of the operator LG=x2+y2+(x2+y2)t2L_{G}=\partial_{x}^{2}+\partial_{y}^{2}+(x^{2}+y^{2})\partial_{t}^{2} on 2+1\mathbb{R}^{2+1}. In particular, they prove that m~2,dG=2\tilde{m}^{G}_{2,d}=2 when d0(mod4)d\equiv 0\pmod{4}, whereas m~2,dG3\tilde{m}^{G}_{2,d}\geq 3 when d0(mod4)d\equiv 0\pmod{4}; moreover, they provide examples that show m2,4G=m2,8G=m2,12G=3m^{G}_{2,4}=m^{G}_{2,8}=m^{G}_{2,12}=3. The method of proof is the perturbation technique of Lewy op. cit. Other than the fact that the parabolic scaling is the natural scaling for solutions of the Grushin operator and the heat operator, there does not seem to be any immediate connection between Grushin spherical harmonics and hcps. To wit, when p(x,y,t)p(x,y,t) is a Grushin spherical harmonic, so is p(x,y,t)p(x,y,-t), whereas this strong symmetry property is not enjoyed by time-dependent solutions of the heat equation. Thus, the main results in [LTY15] cannot be used to establish Theorem 1.1 or vice-versa.

1.3. Free boundary regularity for caloric measure

The phrase caloric measure refers to a family of probability measures ωΩX,t\omega^{X,t}_{\Omega} that are supported on a subset of the boundary Ω\partial\Omega of a space-time domain Ωn+1=n×\Omega\subset\mathbb{R}^{n+1}=\mathbb{R}^{n}\times\mathbb{R} and indexed by the points (X,t)Ω(X,t)\in\Omega. They arise in connection with the Dirichlet problem for the heat equation. Stochastically, ωΩX,t(E)\omega^{X,t}_{\Omega}(E) is the probability that the trace (B(t+s),ts)s0(B(t+s),t-s)_{s\geq 0} of a Brownian traveler B(s)B(s) starting at B(t)=XB(t)=X and sent into the past first intersects Ω\partial\Omega inside the set En+1E\subset\mathbb{R}^{n+1}. For a consolidated introduction to caloric measure, see [BG23, §3], and for extensive background, see [Wat12] or [Doo01]. Recent progress on free boundary regularity for caloric measure was made by Mourgoglou and Puliatti [MP21], propelling the time-dependent theory for caloric measure closer to the better developed, time-independent theory for harmonic and elliptic measure (see e.g. [KPT09, AM19, BET22]). Among other results—and setting aside certain technical assumptions related to the heat potential theory—their work leads to the following description of the asymptotic shape of the free boundary in the two-phase setting.

Theorem 1.16 (Mourgoglou-Puliatti).

Assume that Ω+=n+1Ω¯\Omega^{+}=\mathbb{R}^{n+1}\setminus\overline{\Omega^{-}} and Ω=n+1Ω+¯\Omega^{-}=\mathbb{R}^{n+1}\setminus\overline{\Omega^{+}} are complementary domains in n+1\mathbb{R}^{n+1} with a sufficiently regular (for heat potential theory), common boundary Ω=Ω+=Ω\partial\Omega=\partial\Omega^{+}=\partial\Omega^{-}. Let ω±=ωΩ±X±,t0\omega^{\pm}=\omega^{X_{\pm},t_{0}}_{\Omega^{\pm}} be caloric measures for Ω±\Omega^{\pm} with poles at (X±,t0)Ω±(X_{\pm},t_{0})\in\Omega^{\pm} or poles at infinity. If ω±\omega^{\pm} are doubling measures, ω+ωω+\omega^{+}\ll\omega^{-}\ll\omega^{+}, and the Radon-Nikodym derivatives dω/dω+d\omega^{-}/d\omega^{+} and dω+/dωd\omega^{+}/d\omega^{-} are bounded continuous functions on Ω{tt0}\partial\Omega\cap\{t\leq t_{0}\}, then Ω=Γ1Γ2Γd0,\partial\Omega=\Gamma_{1}\cup\Gamma_{2}\cup\cdots\cup\Gamma_{d_{0}}, where geometric blow-ups (tangent sets) Σ=limiri1(Ωx)\Sigma=\lim_{i\rightarrow\infty}r_{i}^{-1}(\partial\Omega-x) of Ω\partial\Omega at xΓdx\in\Gamma_{d} along sequence of scales ri0r_{i}\rightarrow 0 are zero sets of homogeneous caloric polynomials pp of degree dd such that {p>0}\{p>0\} and {p<0}\{p<0\} are connected. Cf. [MP21, Theorems III, IV].

The main results in this paper validate Mourgoglou and Puliatti’s theory by confirming existence of time-dependent hcps with two nodal domains. Furthermore, we obtain a refined description of the free boundary in low dimensions.

Theorem 1.17.

When n=1n=1, Ω=Γ1\partial\Omega=\Gamma_{1}. When n=2n=2,

(1.9) Ω=k0Γ4k+1Γ4k+2Γ4k+3;\partial\Omega=\bigcup_{k\geq 0}\Gamma_{4k+1}\cup\Gamma_{4k+2}\cup\Gamma_{4k+3};

for every d0(mod4)d\not\equiv 0\pmod{4}, the stratum Γd\Gamma_{d} is nonempty for some pair of domains satisying the free boundary condition. When n=3n=3, the stratum Γd\Gamma_{d} can be nonempty for every d1d\geq 1.

Proof.

When n=1n=1, Γd=\Gamma_{d}=\emptyset for all d2d\geq 2 by Example 1.12. When n=2n=2, Γd=\Gamma_{d}=\emptyset for all d=4kd=4k by Theorem 1.1. For the remaining pairs of nn and dd, the examples of hcps pp in n+1\mathbb{R}^{n+1} of degree dd with 𝒩(p)=2\mathcal{N}(p)=2 constructed in the proof of Theorem 1.1 (see the proofs of Proposition 3.1 and Theorems 5.4 and 5.5 for details) have smooth zero sets outside any neighborhood of the origin. This fact is enough to ensure that the domains Ωp+={p>0}\Omega_{p}^{+}=\{p>0\} and Ωp={p<0}\Omega_{p}^{-}=\{p<0\} associated to pp satisfy the background regularity hypothesis in Theorem 1.16. If ωp±\omega_{p}^{\pm} denote the caloric measures on Ωp±\Omega_{p}^{\pm} with poles at infinity, then it is known that ωp+=ωp\omega_{p}^{+}=\omega_{p}^{-} and dωp/dωp+1d\omega_{p}^{-}/d\omega_{p}^{+}\equiv 1 (see [MP21, §6]). Finally, by parabolic homogeneity, {p=0}\{p=0\} is the unique blow-up of Ωp±={p=0}\partial\Omega_{p}^{\pm}=\{p=0\} at the origin. Therefore, Γd\Gamma_{d} is nonempty in Ωp±\partial\Omega^{\pm}_{p}. ∎

1.4. Proof strategies and outline of the paper

In Section 2, we recall classical facts about hcps, including their connection with Hermite polynomials. We also introduce a basis of hcps of degree dd in n+1\mathbb{R}^{n+1}, which we use in the constructions in Sections 3 and 5.

In Section 3, we first prove that mn,d=2m_{n,d}=2 for n3n\geq 3 and any d1d\geq 1 by modifying a construction in [BET17]. Next, we build time-dependent hcps in n+1\mathbb{R}^{n+1} with a large number of nodal domains by taking products of hcps in 1+1\mathbb{R}^{1+1}. Finally, after showing that any nodal domain of an hcp necessarily intersects {t=1}\{t=-1\}, we employ the proof of Courant’s nodal domain theorem on negative time slices of hcps to establish the upper bound on the number of nodal domains in Theorem 1.2. This leaves us to determine the value of m2,dm_{2,d} for d1d\geq 1.

In Section 4, we show that m2,d3m_{2,d}\geq 3 whenever d1d\geq 1 satisfies d0(mod4)d\equiv 0\pmod{4}. The main argument leverages the fact that if pp is an hcp of degree dd in 2+1\mathbb{R}^{2+1}, then the nodal set of p|𝕊2p|_{\mathbb{S}^{2}} near the north and south poles is asymptotic to the zero set of a homogeneous harmonic polynomial in two variables, which are easy to describe and are perfectly understood.

In Section 5, we construct examples of time-dependent hcps of degree d1d\geq 1 in 2+1\mathbb{R}^{2+1} with exactly m2,dm_{2,d} nodal domains. The basic strategy dates back to [Lew77]: starting with an hcp of degree dd in 2+1\mathbb{R}^{2+1} whose zero set we can explicitly describe, we perturb the polynomial to produce a time-dependent hcp with the desired number of nodal domains. This style of argument requires a careful analysis of how perturbation affects the topology of nodal sets; see Lemma 5.2 for a precise statement and Section 6 for the proof of the lemma.

2. Basic properties of homogeneous caloric polynomials

Given an hcp p(x,t)p(x,t) of degree dd, we typically shall choose to write p(x,t)p(x,t) in the form

(2.1) p(x,t)=tmpm(x)+tm1pm1(x)++p0(x)for all xnt.p(x,t)=t^{m}p_{m}(x)+t^{m-1}p_{m-1}(x)+\cdots+p_{0}(x)\quad\text{for all $x\in\mathbb{R}^{n}$, $t\in\mathbb{R}$}.

Each coefficient pmj=pmj(x)p_{m-j}=p_{m-j}(x) is necessarily an algebraically homogeneous polynomial of degree d2(mj)d-2(m-j) (see Example 1.6). Moreover, applying the heat operator to pp and collecting like powers of tt, we obtain the relations

(2.2) 0=Δpm,mpm=Δpm1,,(mj)pmj=Δpmj1,,p1=Δp0.0=\Delta p_{m},\quad mp_{m}=\Delta p_{m-1},\quad\cdots,\quad(m-j)p_{m-j}=\Delta p_{m-j-1},\quad\cdots,\quad p_{1}=\Delta p_{0}.

As such, we refer to pmp_{m} as the harmonic coefficient of pp and obtain that the other coefficients pmjp_{m-j} are polyharmonic: Δj+1pmj=0\Delta^{j+1}p_{m-j}=0. In fact, the relations (2.2) and the requirement that each coefficient pi(x)p_{i}(x) be homogeneous gives a characterization of p(x,t)p(x,t) being an hcp. Thus, we arrive at the following elementary method of generating hcps: starting with any choice of m0m\geq 0 and hhp pm(x)p_{m}(x), use (2.2) to inductively solve for homogeneous coefficients pm1(x),,p0(x)p_{m-1}(x),\dots,p_{0}(x); then p(x,t)p(x,t) defined by (2.1) is an hcp.

Now, given any homogeneous polynomial q(x)=bxdq(x)=bx^{d} with x1x\in\mathbb{R}^{1}, the polynomial r(x)=1(d+2)(d+1)cx2q(x)r(x)=\frac{1}{(d+2)(d+1)}cx^{2}q(x) is the unique homogeneous function such that r′′(x)=cq(x)r^{\prime\prime}(x)=cq(x). Since the only hhps in 1\mathbb{R}^{1} are of the form q(x)=bq(x)=b or q(x)=bxq(x)=bx, it follows that, up to scaling by a constant, there exists a unique hcp pd(x,t)p_{d}(x,t) in 1+1\mathbb{R}^{1+1} for each degree d0d\geq 0. We adopt the following normalization for the hcps pd(x,t)p_{d}(x,t), emphasizing the time variable.

Definition 2.1.

For each d0d\geq 0, define pd:1+1p_{d}:\mathbb{R}^{1+1}\rightarrow\mathbb{R} as follows. When d=2kd=2k is even,

pd(x,t):=tk+k2!tk1x2+k(k1)4!tk2x4++k!(2k)!x2k=j=0kk!(kj)!(2j)!tkjx2j.p_{d}(x,t):=t^{k}+\tfrac{k}{2!}t^{k-1}x^{2}+\tfrac{k(k-1)}{4!}t^{k-2}x^{4}+\cdots+\tfrac{k!}{(2k)!}x^{2k}=\sum_{j=0}^{k}\tfrac{k!}{(k-j)!(2j)!}t^{k-j}x^{2j}.

When d=2k+1d=2k+1 is odd,

pd(x,t):=tkx+k3!tk1x3+k(k1)5!tk2x5++k!(2k+1)!x2k+1=j=0kk!(kj)!(2j+1)!tkjx2j+1.p_{d}(x,t):=t^{k}x+\tfrac{k}{3!}t^{k-1}x^{3}+\tfrac{k(k-1)}{5!}t^{k-2}x^{5}+\cdots+\tfrac{k!}{(2k+1)!}x^{2k+1}=\sum_{j=0}^{k}\tfrac{k!}{(k-j)!(2j+1)!}t^{k-j}x^{2j+1}.
Definition 2.2 (see [Sze75, §5.5]).

The Hermite polynomials H0(x)=1H_{0}(x)=1, H1(x)=2xH_{1}(x)=2x, H2(x)=4x22H_{2}(x)=4x^{2}-2, H3(x)=8x312xH_{3}(x)=8x^{3}-12x, H4(x)=16x448x2+12H_{4}(x)=16x^{4}-48x^{2}+12, H5(x)=32x5160x3+120xH_{5}(x)=32x^{5}-160x^{3}+120x, etc. are the family of orthogonal polynomials for the weighted space L2(,ex2dx)L^{2}(\mathbb{R},e^{-x^{2}}dx) defined by requiring that degHd=d\deg H_{d}=d, the coefficient of xdx^{d} in Hd(x)H_{d}(x) is positive, and

(2.3) Hd(x)Hd(x)ex2𝑑x=π1/22dd!δddfor all d,d.\int_{\mathbb{R}}H_{d}(x)H_{d^{\prime}}(x)\,e^{-x^{2}}dx=\pi^{1/2}2^{d}d!\,\delta_{dd^{\prime}}\quad\text{for all }d,d^{\prime}\in\mathbb{N}.
Lemma 2.3 (see [Sze75, §5.5]).

Equivalently, for all d0d\geq 0,

(2.4) Hd(x)=j=0d/2d!j!(d2j)!(1)j(2x)d2j.H_{d}(x)=\sum_{j=0}^{\lfloor d/2\rfloor}\frac{d!}{j!(d-2j)!}(-1)^{j}(2x)^{d-2j}.

After establishing the connection between the “basic hcps” and the Hermite polynomials (see e.g. [RW59], [Che98], [PKS21]), one can use facts about the zeros of Hd(x)H_{d}(x) and parabolic scaling to derive the following description of pd(x,t)p_{d}(x,t) and its nodal set.

Theorem 2.4.

For all d2d\geq 2, the “basic hcp” pd(x,t)p_{d}(x,t) assumes the form

(2.5) pd(x,t)={(t+ad,1x2)(t+ad,kx2)when d=2k is even,x(t+ad,1x2)(t+ad,kx2)when d=2k+1 is odd,p_{d}(x,t)=\left\{\begin{array}[]{rl}(t+a_{d,1}x^{2})\cdots(t+a_{d,k}x^{2})&\text{when $d=2k$ is even},\\ x(t+a_{d,1}x^{2})\cdots(t+a_{d,k}x^{2})&\text{when $d=2k+1$ is odd},\end{array}\right.

for some distinct numbers 0<ad,1<<ad,k0<a_{d,1}<\dots<a_{d,k}. Moreover, if we write

p2k1(x,t)\displaystyle p_{2k-1}(x,t) =x(t+a1x2)(t+ak1x2),\displaystyle=x(t+a_{1}x^{2})\cdots(t+a_{k-1}x^{2}), p2k+1(x,t)\displaystyle p_{2k+1}(x,t) =x(t+c1x2)(t+ckx2),\displaystyle=x(t+c_{1}x^{2})\cdots(t+c_{k}x^{2}),
p2k(x,t)=(t+b1x2)(t+bkx2),p_{2k}(x,t)=(t+b_{1}x^{2})\cdots(t+b_{k}x^{2}),

with the aia_{i}​’s, bib_{i}​’s, and cic_{i}​’s each listed in increasing order, then the coefficients associated with consecutive polynomials are interlaced:

(2.6) {b1<a1<b2<a2<<ak1<bk,c1<b1<c2<b2<<bk1<ck<bk.\left\{\begin{array}[]{l}b_{1}<a_{1}<b_{2}<a_{2}<\cdots<a_{k-1}<b_{k},\\ \,c_{1}<b_{1}<c_{2}<b_{2}<\cdots<b_{k-1}<c_{k}<b_{k}.\end{array}\right.
Proof.

Replacing jj by kjk-j in the summations in Definition 2.1 yield that for all d0d\geq 0,

(2.7) pd(x,t)=j=0d/2d/2!j!(d2j)!tjxd2j.p_{d}(x,t)=\sum_{j=0}^{\lfloor d/2\rfloor}\frac{\lfloor d/2\rfloor!}{j!(d-2j)!}t^{j}x^{d-2j}.

Comparing (2.4) and (2.7), we see that

(2.8) pd(x,1)=d/2!d!Hd(x/2).p_{d}(x,-1)=\frac{\lfloor d/2\rfloor!}{d!}H_{d}(x/2).

Because the Hermite polynomial HdH_{d} is even, when dd is even, HdH_{d} is odd, when dd is odd, and orthogonal polynomials have a full number of distinct real roots (see [Sze75, Theorem 3.3.1]), we can factor Hd(x)=(x2r12)(xrk2)H_{d}(x)=(x^{2}-r_{1}^{2})\cdots(x-r_{k}^{2}) for some 0<rk<<r10<r_{k}<\cdots<r_{1}, when d=2kd=2k is even, and Hd(x)=x(x2r12)(x2rk2)H_{d}(x)=x(x^{2}-r_{1}^{2})\cdots(x^{2}-r_{k}^{2}), when d=2k+1d=2k+1 is odd. Hence

pd(x,1)=d/2!d!r12rk2xd2d/2(x24r121)(x24rk21).p_{d}(x,-1)=\frac{\lfloor d/2\rfloor!}{d!}r_{1}^{2}\cdots r_{k}^{2}x^{d-2\lfloor d/2\rfloor}\left(\frac{x^{2}}{4r_{1}^{2}}-1\right)\cdots\left(\frac{x^{2}}{4r_{k}^{2}}-1\right).

Together with parabolic homogeneity and the fact that pd(x,t)p_{d}(x,t) was normalized to have leading term tkt^{k} or tkxt^{k}x, this yields (2.5) with ad,i=1/(4ri2)a_{d,i}=1/(4r_{i}^{2}). Thus, (2.6) follows from the interlacing of roots of consecutive Hermite polynomials [Sze75, Theorem 3.3.2].∎

Refer to caption
Figure 2.1. The nodal set of a degree dd hcp in 1+1\mathbb{R}^{1+1} is a union of d/2\lfloor d/2\rfloor nested, downward-opening parabolas with a common turning point at the origin, and when dd is odd, an additional vertical line (the tt-axis). Thus, the number of nodal domains is precisely 2d/22\lceil d/2\rceil. From left to right, we illustrate the cases d=2d=2, …, d=5d=5. Inside the nodal set of pdpd+1p_{d}p_{d+1}, the “nodal parabolas” of consecutive hcps pdp_{d} and pd+1p_{d+1} are intertwined: the “widest” parabola of pd+1p_{d+1} sits above the “widest” parabola of pdp_{d}; the “widest” parabola of pdp_{d} sits above the “second widest” parabola of pd+1p_{d+1}; etc.
Corollary 2.5.

Any hcp p(x,t)p(x,t) in 1+1\mathbb{R}^{1+1} of degree d1d\geq 1 has exactly 2d/22\lceil d/2\rceil nodal domains. See Figure 2.1.

In contrast to the case n=1n=1, given a homogeneous polynomial q(x)q(x) with xnx\in\mathbb{R}^{n} for some n2n\geq 2, there is more than one way to produce a homogeneous polynomial r(x)r(x) such that Δr(x)=cq(x)\Delta r(x)=cq(x). We can build an explicit basis for the vector space d(n+1)\mathfrak{H}_{d}(\mathbb{R}^{n+1}) of all hcps of degree dd in n+1\mathbb{R}^{n+1} (and the zero function) using products of basic hcps in 1+1\mathbb{R}^{1+1}.

Definition 2.6.

Let n2n\geq 2. For each multi-index α=(α1,,αn)\alpha=(\alpha_{1},\dots,\alpha_{n}), define

(2.9) pα(x,t):=pα1(x1,t)pαn(xn,t).p_{\alpha}(x,t):=p_{\alpha_{1}}(x_{1},t)\cdots p_{\alpha_{n}}(x_{n},t).
Lemma 2.7.

For all t<0t<0, the set {pα(,t):α is a multi-index in n}\{p_{\alpha}(\cdot,t):\alpha\text{ is a multi-index in }\mathbb{N}^{n}\} is an orthogonal basis for the weighted space L2(n,e|x|2/4tdx)L^{2}(\mathbb{R}^{n},e^{-|x|^{2}/4t}\,dx).

Proof.

Let t<0t<0 and let α\alpha and β\beta be multi-indices in n\mathbb{N}^{n} with αβ\alpha\neq\beta. By Fubini’s theorem,

npα(x,t)pβ(x,t)e|x|2/4|t|𝑑x\displaystyle\int_{\mathbb{R}^{n}}p_{\alpha}(x,t)p_{\beta}(x,t)e^{-\left\lvert x\right\rvert^{2}/4\left\lvert t\right\rvert}\,dx =i=1npαi(xi,t)pβi(xi,t)exi2/4|t|𝑑xi.\displaystyle=\prod_{i=1}^{n}\int_{\mathbb{R}}p_{\alpha_{i}}(x_{i},t)p_{\beta_{i}}(x_{i},t)e^{-x_{i}^{2}/4\left\lvert t\right\rvert}\,dx_{i}.

Thus, the left hand side vanishes if and only if at least one of the terms on the right hand side vanish. Since αβ\alpha\neq\beta, there exists 1in1\leq i\leq n such that αiβi\alpha_{i}\neq\beta_{i}. Write γi=(αi+βi)/2\gamma_{i}=(\alpha_{i}+\beta_{i})/2. By a simple change of variables, with xi=2|t|1/2yx_{i}=2|t|^{1/2}y, parabolic homogeneity, (2.8), and (2.3),

|t|1/2pαi(xi,t)pβi(xi,t)exi2/4|t|𝑑xi=2pαi(2|t|1/2y,t)pβi(2|t|1/2y,t)ey2𝑑y=2|t|γiRpαi(2y,1)pβi(2y,1)ey2𝑑y=2|t|γiC(α,β)Hαi(y)Hβi(y)ey2𝑑y=0.\begin{split}&|t|^{-1/2}\int_{\mathbb{R}}p_{\alpha_{i}}(x_{i},t)p_{\beta_{i}}(x_{i},t)e^{-x_{i}^{2}/4\left\lvert t\right\rvert}\,dx_{i}=2\int_{\mathbb{R}}p_{\alpha_{i}}(2|t|^{1/2}y,t)p_{\beta_{i}}(2|t|^{1/2}y,t)\,e^{-y^{2}}dy\\ &\quad=2|t|^{\gamma_{i}}\int_{R}p_{\alpha_{i}}(2y,-1)p_{\beta_{i}}(2y,-1)\,e^{-y^{2}}dy=2|t|^{\gamma_{i}}C(\alpha^{\prime},\beta^{\prime})\int_{\mathbb{R}}H_{\alpha_{i}}(y)H_{\beta_{i}}(y)\,e^{-y^{2}}dy=0.\end{split}

By a similar argument, {pα(,t):α is a multi-index in n}\{p_{\alpha}(\cdot,t):\alpha\text{ is a multi-index in }\mathbb{N}^{n}\} is a basis for L2(n,e|x|2/4tdx)L^{2}(\mathbb{R}^{n},e^{-|x|^{2}/4t}\,dx), because {Hd(x):d0}\{H_{d}(x):d\geq 0\} is a basis for L2(,ex2dx)L^{2}(\mathbb{R},e^{-x^{2}}dx). ∎

Lemma 2.8.

For all n1n\geq 1 and d0d\geq 0, the set 𝔓d(n+1)={pα:|α|=d}\mathfrak{P}_{d}(\mathbb{R}^{n+1})=\{p_{\alpha}:|\alpha|=d\} is a basis for d(n+1)\mathfrak{H}_{d}(\mathbb{R}^{n+1}) and dimd(n+1)=(n1+dn1)\dim\mathfrak{H}_{d}(\mathbb{R}^{n+1})=\binom{n-1+d}{n-1}.

Proof.

For a proof that dimd(n+1)=(n1+dn1)\dim\mathfrak{H}_{d}(\mathbb{R}^{n+1})=\binom{n-1+d}{n-1}, see [CM21, Lemma 2.2]. The fact that #𝔓d(n+1)=(n1+dn1)\#\mathfrak{P}_{d}(\mathbb{R}^{n+1})=\binom{n-1+d}{n-1} is a standard exercise; see e.g. [Eva10, p. 12]. Now, the set is 𝔓d(n+1)\mathfrak{P}_{d}(\mathbb{R}^{n+1}) is linearly independent, because {pα(,1):|α|=d}\{p_{\alpha}(\cdot,-1):|\alpha|=d\} is linearly independent by Lemma 2.7. Therefore, 𝔓d(n+1)\mathfrak{P}_{d}(\mathbb{R}^{n+1}) is a basis for d(n+1)\mathfrak{H}_{d}(\mathbb{R}^{n+1}). ∎

Example 2.9.

The hhp x2y2x^{2}-y^{2} in 2\mathbb{R}^{2} can be expressed as a linear combination of the basic hcps p1(x,t)p_{1}(x,t) and p1(y,t)p_{1}(y,t): x2y2=2(t+12x2)2(t+12y2)x^{2}-y^{2}=2(t+\frac{1}{2}x^{2})-2(t+\frac{1}{2}y^{2}).

We will also need the following fact in the next section.

Lemma 2.10.

If pp is an hcp of degree dd in n+1\mathbb{R}^{n+1}, then the function v(x):=p(x,1)v(x):=p(x,-1) satisfies

(2.10) div(e|x|2/4v(x))=(d/2)e|x|2/4v(x).\displaystyle-\mathrm{div}(e^{-\left\lvert x\right\rvert^{2}/4}\nabla v(x))=(d/2)e^{-\left\lvert x\right\rvert^{2}/4}v(x).
Proof.

Indeed, for t<0t<0, we have p(x,t)=(t)d/2p((t)1/2x,1)=(t)d/2v((t)1/2x)p(x,t)=(-t)^{d/2}p((-t)^{-1/2}x,-1)=(-t)^{d/2}v((-t)^{1/2}x). Applying the heat operator tΔ\partial_{t}-\Delta and simplifying yields

0\displaystyle 0 =(t)d21[(d/2)v((t)1/2x)+(1/2)v((t)1/2x)(t)1/2xΔv((t)1/2x)].\displaystyle=(-t)^{\frac{d}{2}-1}\left[-(d/2)v((-t)^{-1/2}x)+(1/2)\nabla v((-t)^{-1/2}x)\cdot(-t)^{-1/2}x-\Delta v((-t)^{-1/2}x)\right].

Thus, (1/2)v(x)xΔv(x)=(d/2)v(x)(1/2)\nabla v(x)\cdot x-\Delta v(x)=(d/2)v(x). Therefore,

div(e|x|2/4v(x))=e|x|2/4[(1/2)vxΔv(x)]=(d/2)e|x|2/4v(x).-\mathrm{div}(e^{-|x|^{2}/4}\nabla v(x))=-e^{-|x|^{2}/4}\left[(1/2)\nabla v\cdot x-\Delta v(x)\right]=(d/2)e^{-|x|^{2}/4}v(x).\qed

3. HCP in high spatial dimensions

In this section, we establish the case n3n\geq 3 of Theorem 1.1 and prove Theorem 1.2.

Proposition 3.1 (Theorem 1.1 when n3n\geq 3).

If ϕ(x,y)\phi(x,y) is an hhp of degree d1d\geq 1 in 2\mathbb{R}^{2} and ψ(z,t)\psi(z,t) is an hcp of degree dd in 1+1\mathbb{R}^{1+1}, then u(x,y,z,t):=ϕ(x,y)+ψ(z,t)u(x,y,z,t):=\phi(x,y)+\psi(z,t) has exactly two nodal domains. In particular, mn,d=2m_{n,d}=2 for all n3n\geq 3 and d2d\geq 2.

Proof.

We modify the proof of [BET17, Lemma 1.7], which gives a construction of hhps in n\mathbb{R}^{n} with two nodal domains when n4n\geq 4. The essential change is to show how to incorporate parabolic scaling. A trivial (but important!) observation is that for any (x,y)(x,y) and (z,t)(z,t), the expressions |ϕ(λx,λy)|=λd|ϕ(x,y)||\phi(\lambda x,\lambda y)|=\lambda^{d}|\phi(x,y)| and |ψ(λz,λ2t)|=λd|ψ(z,t)||\psi(\lambda z,\lambda^{2}t)|=\lambda^{d}|\psi(z,t)| are (weakly) increasing as functions of λ[0,1]\lambda\in[0,1]. For any (x,y)(0,0)(x,y)\neq(0,0), let “the line segment from (0,0)(0,0) to (x,y)(x,y)” have its usual meaning. For any (z,t)(0,0)(z,t)\neq(0,0), let “the line segment from (0,0)(0,0) to (z,t)(z,t)” mean the curve described by {(λz,λ2t):λ[0,1]}\{(\lambda z,\lambda^{2}t):\lambda\in[0,1]\}. Then |ϕ||\phi| and |ψ||\psi| are increasing along line segments started at the origin and |ϕ||\phi| and |ψ||\psi| are decreasing along line segments terminating at the origin. Also, the sets {ϕ>0}\{\phi>0\}, {ϕ<0}\{\phi<0\}, {ψ>0}\{\psi>0\}, and {ψ<0}\{\psi<0\} are each nonempty. Keeping these preliminaries in mind, we will now show that {u>0}\{u>0\} is path-connected. (Applying the same argument to u-u shows that {u<0}\{u<0\} is path-connected.)

Suppose that (x1,y1,z1,t1)(x_{1},y_{1},z_{1},t_{1}) and (x2,y2,z2,t2)(x_{2},y_{2},z_{2},t_{2}) are points at which ϕ(xi,yi)+ψ(zi,ti)>0\phi(x_{i},y_{i})+\psi(z_{i},t_{i})>0. Then for each i=1,2i=1,2, at least one of the terms ϕ(xi,yi)\phi(x_{i},y_{i}) and ψ(zi,ti)\psi(z_{i},t_{i}) is positive. Because the argument that follows only involves “line segments” with an endpoint at the origin and the monotonicity of |ϕ||\phi| and |ψ||\psi| along those line segments, we may suppose without loss of generality that ψ(z1,t1)>0\psi(z_{1},t_{1})>0 and ψ(z2,t2)>0\psi(z_{2},t_{2})>0. We emphasize that ϕ(x1,y1)\phi(x_{1},y_{1}) and ϕ(x2,y2)\phi(x_{2},y_{2}) may have any sign (positive, negative, zero). Choose any auxiliary point (x+,y+)(x_{+},y_{+}) at which ϕ(x+,y+)>0\phi(x_{+},y_{+})>0. We can build a “piecewise linear” path in {u>0}\{u>0\} from (x1,y1,z1,t1)(x_{1},y_{1},z_{1},t_{1}) to (x2,y2,z2,t2)(x_{2},y_{2},z_{2},t_{2}) as follows. The path is a concatenation of six “line segments” (see Figure 3.1):

Refer to caption
Figure 3.1. Connecting (v1,w1)=(x1,y1,z1,t1)(v_{1},w_{1})=(x_{1},y_{1},z_{1},t_{1}) to (v2,w2)=(x2,y2,z2,t2)(v_{2},w_{2})=(x_{2},y_{2},z_{2},t_{2}) in 3+1{u>0}\mathbb{R}^{3+1}\cap\{u>0\}, where ϕ(x,y)\phi(x,y) is a degree 4 hhp in 2\mathbb{R}^{2} (left) and ψ(z,t)\psi(z,t) is a degree 4 hcp in 1+1\mathbb{R}^{1+1} (right).
  1. (1)

    First follow the line segment from (x1,y1,z1,t1)(x_{1},y_{1},z_{1},t_{1}) to (0,0,z1,t1)(0,0,z_{1},t_{1}).

  2. (2)

    Second follow the line segment from (0,0,z1,t1)(0,0,z_{1},t_{1}) to (x+,y+,z1,t1)(x_{+},y_{+},z_{1},t_{1}).

  3. (3)

    Third follow the line segment from (x+,y+,z1,z2)(x_{+},y_{+},z_{1},z_{2}) to (x+,y+,0,0)(x_{+},y_{+},0,0).

  4. (4)

    Fourth follow the line segment from (x+,y+,0,0)(x_{+},y_{+},0,0) to (x+,y+,z2,t2)(x_{+},y_{+},z_{2},t_{2}).

  5. (5)

    Fifth follow the line segment from (x+,y+,z2,t2)(x_{+},y_{+},z_{2},t_{2}) to (0,0,z2,t2)(0,0,z_{2},t_{2}).

  6. (6)

    Sixth follow the line segment from (0,0,z2,t2)(0,0,z_{2},t_{2}) to (x2,y2,z2,t2)(x_{2},y_{2},z_{2},t_{2}).

It remains to confirm that the six line segments lie in {u>0}\{u>0\}. The first segment lies in the positivity set, because u(x1,y1,z1,t1)>0u(x_{1},y_{1},z_{1},t_{1})>0 and ψ(z1,t1)>0\psi(z_{1},t_{1})>0. (If ϕ(x1,y1)0\phi(x_{1},y_{1})\leq 0, then uu increases from the initial point to the terminal point. If ϕ(x1,y1)0\phi(x_{1},y_{1})\geq 0, then uu decreases from the initial point to the terminal point, but uu never falls below ψ(z1,t1)\psi(z_{1},t_{1}).) Along the second segment, ϕ0\phi\geq 0 and ψ>0\psi>0, so u>0u>0. Along the third segment, ϕ>0\phi>0 and ψ0\psi\geq 0, so u>0u>0. After reversing the orientation, identical arguments show that the fourth, fifth, and sixth segments lie in {u>0}\{u>0\}, as well. ∎

We now move on to the lower and upper bounds on Mn,dM_{n,d} for any spatial dimension.

Proposition 3.2.

For all n2n\geq 2 and d2d\geq 2, there exists a time-dependent hcp of degree dd in n+1\mathbb{R}^{n+1} with at least d/nn\lfloor d/n\rfloor^{n} nodal domains.

Proof.

Let c=d/nc=\lfloor d/n\rfloor. If c1c\leq 1, then the conclusion is trivial. Thus, we may assume that c2c\geq 2. Let pp be an hcp in 1+1\mathbb{R}^{1+1} of degree cc and let qq be an hcp in 1+1\mathbb{R}^{1+1} of degree b=d(n1)ccb=d-(n-1)c\geq c. A direct computation shows that u(x,t):=q(xn,t)i=1n1p(xi,t)u(x,t):=q(x_{n},t)\prod_{i=1}^{n-1}p(x_{i},t) is an hcp of degree dd. Moreover, uu is time-dependent, since c2c\geq 2 implies pp and qq are time-dependent. By Corollary 2.5,

(3.1) 𝒩(u)𝒩(p)n1𝒩(q)cn1bcn.\mathcal{N}(u)\geq\mathcal{N}(p)^{n-1}\mathcal{N}(q)\geq c^{n-1}b\geq c^{n}.

We can verify the first inequality in (3.1) as follows. Suppose that (x,t)(x,t) and (y,s)(y,s) are points in n+1\mathbb{R}^{n+1} such that u(x,t)0u(x,t)\neq 0, u(y,s)0u(y,s)\neq 0, and

  1. (a)

    (xi,t)(x_{i},t) and (yi,s)(y_{i},s) belong to different nodal domains of pp for some 1in11\leq i\leq n-1, or

  2. (b)

    (xn,t)(x_{n},t) and (yn,s)(y_{n},s) belong to different nodal domains of qq.

If (a) holds, assign j=ij=i and r=pr=p; otherwise, if (b) holds, assign j=nj=n and r=qr=q. Let γ:[0,1]n+1\gamma:[0,1]\rightarrow\mathbb{R}^{n+1} be any curve such that γ(0)=(x,t)\gamma(0)=(x,t) and γ(1)=(y,s)\gamma(1)=(y,s) and let π\pi be the orthogonal projection onto the ejen+1e_{j}e_{n+1}-plane. Then πγ:[0,1]1+1\pi\circ\gamma:[0,1]\rightarrow\mathbb{R}^{1+1} is a curve connecting (xj,t)(x_{j},t) to (yj,s)(y_{j},s). Since (xj,t)(x_{j},t) and (yj,s)(y_{j},s) lie in different nodal domains of rr, it follows that r(π(γ(c)))=0r(\pi(\gamma(c)))=0 for some c(0,1)c\in(0,1). Thus, u(γ(c))=0u(\gamma(c))=0, as well, by definition of uu. Since γ\gamma was an arbitrary curve connecting (x,t)(x,t) to (y,s)(y,s), we conclude that (x,t)(x,t) and (y,s)(y,s) lie in different nodal domains of uu. Therefore, 𝒩(u)𝒩(p)n1𝒩(q)\mathcal{N}(u)\geq\mathcal{N}(p)^{n-1}\mathcal{N}(q). ∎

Example 3.3.

Let u(x,y,t)=(2t+x2)(2t+y2)u(x,y,t)=(2t+x^{2})(2t+y^{2}). Then 𝒩(u)=6>4=𝒩(2t+x2)𝒩(2t+y2)\mathcal{N}(u)=6>4=\mathcal{N}(2t+x^{2})\mathcal{N}(2t+y^{2}). Thus, the first inequality in (3.1) can be strict.

The next lemma will help us bound the number of nodal domains of an hcp from above.

Lemma 3.4.

If uu is an hcp of degree d1d\geq 1 in n+1\mathbb{R}^{n+1}, then every nodal domain of uu intersects the hyperplane {t=1}\{t=-1\}. Thus, 𝒩(u)\mathcal{N}(u) is at most the number of connected components of {xn:u(x,1)0}\{x\in\mathbb{R}^{n}:u(x,-1)\neq 0\} in n\mathbb{R}^{n}.

Proof.

By parabolic scaling, it suffices to prove that every nodal domain of uu intersects the lower half-space {t<0}\{t<0\}. Suppose to get a contradiction that there exists a nodal domain UU of uu such that U{t0}U\subset\{t\geq 0\}. In fact, observe that U{t>0}U\subset\{t>0\}, since UU is open. Replacing uu by u-u, if necessary, we may further assume that U{u>0}U\subset\{u>0\}. Write

𝔼+n:={(x,t)n+1:x12++xn2+|t|=1 and t0}.\mathbb{E}^{n}_{+}:=\{(x,t)\in\mathbb{R}^{n+1}:x_{1}^{2}+\cdots+x_{n}^{2}+|t|=1\text{ and }t\geq 0\}.

Define an auxiliary hcp of degree 2d2d in n+1\mathbb{R}^{n+1} by f(x,t):=i=1np2d(xi,t)f(x,t):=\sum_{i=1}^{n}p_{2d}(x_{i},t). By Theorem 2.4, there exist constants a1,,ad>0a_{1},\dots,a_{d}>0 such that f(x,t)=i=1n(t+a1xi2)(t+adxi2).f(x,t)=\sum_{i=1}^{n}(t+a_{1}x_{i}^{2})\cdots(t+a_{d}x_{i}^{2}). Hence f(x,t)>0f(x,t)>0 for all (x,t)𝔼+n(x,t)\in\mathbb{E}^{n}_{+}. Thus, because ff is continuous and 𝔼+n\mathbb{E}^{n}_{+} is compact, we can find c>0c>0 such that f(x,t)>cf(x,t)>c for all (x,t)𝔼+n(x,t)\in\mathbb{E}^{n}_{+}.

To proceed, choose ε>0\varepsilon>0 small enough so that vε:=uεfv_{\varepsilon}:=u-\varepsilon f is positive at some point in UU. Let VV be any connected component of U{vε>0}U\cap\{v_{\varepsilon}>0\}. If (x,t)V(x,t)\in V and x12++xn2+|t|=λ2x_{1}^{2}+\cdots+x_{n}^{2}+|t|=\lambda^{2}, then (x/λ,t/λ2)𝔼+n(x/\lambda,t/\lambda^{2})\in\mathbb{E}^{n}_{+} and we see that

vε(x,t)\displaystyle v_{\varepsilon}(x,t) =λdu(x/λ,t/λ2)ελ2df(x/λ,t/λ2)λd(uL(𝔼+n)λdεc)<0\displaystyle=\lambda^{d}u(x/\lambda,t/\lambda^{2})-\varepsilon\lambda^{2d}f(x/\lambda,t/\lambda^{2})\leq\lambda^{d}(\left\lVert u\right\rVert_{L^{\infty}(\mathbb{E}^{n}_{+})}-\lambda^{d}\varepsilon c)<0

whenever λ1\lambda\gg 1 (depending only on dd, cc, ε\varepsilon, and uL(𝔼+n)\|u\|_{L^{\infty}(\mathbb{E}^{n}_{+})}). This shows that VV is bounded. Furthermore, since VU{vε=0}{u=0}{vε=0}\partial V\subset\partial U\cup\{v_{\varepsilon}=0\}\subset\{u=0\}\cup\{v_{\varepsilon}=0\}, we have vε0v_{\varepsilon}\leq 0 on V\partial V. Therefore, vε0v_{\varepsilon}\leq 0 throughout VV by the maximum principle for solutions of the heat equation. This contradicts our assertion that vεv_{\varepsilon} is positive at some point of VV. ∎

Example 3.5.

For all k1k\geq 1, the basic hcp p2k(x,t)p_{2k}(x,t) of degree 2k2k in 1+1\mathbb{R}^{1+1} has 𝒩(p2k)=2k\mathcal{N}(p_{2k})=2k, but 𝒩(p2k|{t=1})=2k+1\mathcal{N}(p_{2k}|\{t=-1\})=2k+1. See Figure 2.1.

We are ready to present an analogue of Courant’s nodal domain theorem for negative time-slices of hcps, which gives the upper bound on Mn,dM_{n,d}. For the original version of Courant’s theorem, see [CH53, Chapter VI, §6].

Theorem 3.6.

If uu is an hcp of degree dd in n+1\mathbb{R}^{n+1}, then 𝒩(u)(n+dn)\mathcal{N}(u)\leq\binom{n+d}{n}.

Proof.

Let HH denote the Hilbert space L2(n,ex2/4dx)L^{2}(\mathbb{R}^{n},e^{-x^{2}/4}dx). By Lemma 2.7, as α\alpha ranges over all multi-indices in n\mathbb{N}^{n}, the polynomials pα(x):=pα(x,1)p_{\alpha}(x):=p_{\alpha}(x,-1) form an orthogonal basis for HH. Assign q(x):=u(x,1)q(x):=u(x,-1). By Lemma 3.4, 𝒩(u)𝒩(q)\mathcal{N}(u)\leq\mathcal{N}(q). Thus, to establish the theorem, it suffices to prove that k:=𝒩(q)(n+dn)=#{pα(x):|α|d}k:=\mathcal{N}(q)\leq\binom{n+d}{n}=\#\{p_{\alpha}(x):|\alpha|\leq d\}. Enumerate the nodal domains of qq by V1,,VknV_{1},\dotsc,V_{k}\subset\mathbb{R}^{n}. We suppose for the sake of contradiction that k>(n+dn)k>\binom{n+d}{n}. Consider the homogeneous system of (n+dn)\binom{n+d}{n} linear equations i=1kcα,iai=0\sum_{i=1}^{k}c_{\alpha,i}a_{i}=0, where

cα,i:=nq(x)χVi(x)pα(x)ex2/4𝑑xfor all |α|d and 1ik.c_{\alpha,i}:=\int_{\mathbb{R}^{n}}q(x)\chi_{V_{i}(x)}p_{\alpha}(x)\,e^{-x^{2}/4}dx\quad\text{for all $|\alpha|\leq d$ and $1\leq i\leq k$}.

Since k>(n+dn)k>\binom{n+d}{n}, we can find a non-zero solution vector (a1,,ak)(a_{1},\dots,a_{k}). By definition of the linear system, the function gg defined by g(x):=i=1kaiq(x)χVi(x)g(x):=\sum_{i=1}^{k}a_{i}q(x)\chi_{V_{i}}(x) is orthogonal to pαp_{\alpha} in HH for all |α|d|\alpha|\leq d. On the one hand, by (2.10) and integration by parts,

(3.2) n|g(x)|2e|x|2/4𝑑x=i=1kai2Vi|q(x)|2e|x|2/4𝑑x=i=1kai2Vid2q(x)2e|x|2/4𝑑x=d2ng(x)2e|x|2/4𝑑x.\begin{split}&\int_{\mathbb{R}^{n}}\left\lvert\nabla g(x)\right\rvert^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx=\sum_{i=1}^{k}a_{i}^{2}\int_{V_{i}}\left\lvert\nabla q(x)\right\rvert^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx\\ &\qquad=\sum_{i=1}^{k}a_{i}^{2}\int_{V_{i}}\frac{d}{2}q(x)^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx=\frac{d}{2}\int_{\mathbb{R}^{n}}g(x)^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx.\end{split}

On the other hand, we can expand gg with respect to the orthogonal basis {pα}\{p_{\alpha}\} for HH, say g(x)=|α|>dbαpα(x)g(x)=\sum_{\left\lvert\alpha\right\rvert>d}b_{\alpha}p_{\alpha}(x) for some coefficients bαb_{\alpha}. Now, by (2.10) and integration by parts

nxpα(x)xpβ(x)e|x|2/4𝑑x\displaystyle\int_{\mathbb{R}^{n}}\nabla_{x}p_{\alpha}(x)\cdot\nabla_{x}p_{\beta}(x)e^{-\left\lvert x\right\rvert^{2}/4}\,dx =n|β|2pα(x)pβ(x)e|x|2/4𝑑x=0for all αβ.\displaystyle=\int_{\mathbb{R}^{n}}\dfrac{\left\lvert\beta\right\rvert}{2}p_{\alpha}(x)p_{\beta}(x)e^{-\left\lvert x\right\rvert^{2}/4}\,dx=0\quad\text{for all }\alpha\neq\beta.

Thus, using (2.10) and integration by parts once more,

(3.3) n|g(x)|2e|x|2/4𝑑x=|α|>dbα2n|xpα(x)|2e|x|2/4𝑑x=|α|>dbα2n|α|2pα(x)2e|x|2/4𝑑xd+12ng(x)2e|x|2/4𝑑x.\begin{split}&\int_{\mathbb{R}^{n}}\left\lvert\nabla g(x)\right\rvert^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx=\sum_{\left\lvert\alpha\right\rvert>d}b_{\alpha}^{2}\int_{\mathbb{R}^{n}}\left\lvert\nabla_{x}p_{\alpha}(x)\right\rvert^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx\\ &\qquad=\sum_{\left\lvert\alpha\right\rvert>d}b_{\alpha}^{2}\int_{\mathbb{R}^{n}}\dfrac{\left\lvert\alpha\right\rvert}{2}p_{\alpha}(x)^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx\geq\dfrac{d+1}{2}\int_{\mathbb{R}^{n}}g(x)^{2}e^{-\left\lvert x\right\rvert^{2}/4}\;dx.\end{split}

Since g0g\not\equiv 0, (3.2) and (LABEL:eqn:eigp2) are incompatible. Therefore, k(n+dn)k\leq\binom{n+d}{n}. ∎

Note that Proposition 3.2 and Theorem 3.6 yield Theorem 1.2.

4. HCP in 2+1\mathbb{R}^{2+1}, Part I: lower bounds

To complete the proof of Theorem 1.1, it remains to determine the minimum possible number of nodal domains for hcps in 2+1\mathbb{R}^{2+1}, where the story is more complicated than in 1+1\mathbb{R}^{1+1} (see Corollary 2.5) and in high enough spatial dimensions (see Proposition 3.1). In this section, we aim to show that m2,4k3m_{2,4k}\geq 3 for all k1k\geq 1, i.e. the number of nodal domains of an hcp in 2+1\mathbb{R}^{2+1} of degree d4d\geq 4 with d0(mod4)d\equiv 0\pmod{4} is at least 3.

Towards our goal, we first lower bound the number of nodal domains of a continuous function in 2\mathbb{R}^{2} with “alternating nodal structure” at the origin and at infinity. A chamber of uu in VV is a connected component of V{u0}V\cap\{u\neq 0\} relative to VV. A positive chamber is a chamber on which u>0u>0; a negative chamber is a chamber on which u<0u<0.

Lemma 4.1.

Suppose that u:2u:\mathbb{R}^{2}\rightarrow\mathbb{R} is continuous, u(0)=0u(0)=0, and uu has the following nodal structure near the origin and near infinity for some integers nin+,nout+1n^{+}_{\mathrm{in}},n^{+}_{\mathrm{out}}\geq 1 with nin++nout+3n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}\geq 3:

  • There exists ε>0\varepsilon>0 (small) such that in Bε(0)B_{\varepsilon}(0), the chambers of uu are sectors based at the origin, uu alternates signs on adjacent chambers, and uu is positive on nin+n^{+}_{\mathrm{in}} of the chambers. (When nin+=1n^{+}_{\mathrm{in}}=1, we allow either 0 or 11 negative chambers.)

  • There exists M>εM>\varepsilon (large) such that in BM(0)cB_{M}(0)^{c}, the chambers of uu are sectors extending off to infinity, uu alternates signs on adjacent chambers, and uu is positive on nout+n^{+}_{\mathrm{out}} of the chambers. (When nout+=1n^{+}_{\mathrm{out}}=1, we allow either 0 or 11 negative chambers.)

The number of nodal domains of uu is at least nin++nout+n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}.

Refer to caption
Figure 4.1. Parameters nin+n^{+}_{\mathrm{in}} and nout+n^{+}_{\mathrm{out}} count the number of positive chambers of uu near the origin and near infinity, respectively. On the right, we illustrate the base case nin+=2,nout+=1n^{+}_{\mathrm{in}}=2,n^{+}_{\mathrm{out}}=1 (with zero negative chambers at infinity).
Proof.

We adopt the phrase chamber at the origin for chambers relative to Bε(0)B_{\varepsilon}(0) and the phrase chamber at infinity for chambers relative to 2BM(0)\mathbb{R}^{2}\setminus B_{M}(0). The proof is by induction on nin++nout+n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}. For any fixed value of nin++nout+n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}, it suffices to establish either the case nin+nout+n^{+}_{\mathrm{in}}\geq n^{+}_{\mathrm{out}} or the case nout+nin+n^{+}_{\mathrm{out}}\geq n^{+}_{\mathrm{in}}, as each case follows from the other case and inversion. By the alternating hypothesis, if nin+2n^{+}_{\mathrm{in}}\geq 2, then uu also has nin+n^{+}_{\mathrm{in}} negative chambers at the origin; if nout+2n^{+}_{\mathrm{out}}\geq 2, then uu also has nout+n^{+}_{\mathrm{out}} negative chambers at infinity.

For the base case, suppose that nin++nout+=3n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}=3, say nin+=2n^{+}_{\mathrm{in}}=2 and nout+=1n^{+}_{\mathrm{out}}=1. There are two alternatives. First alternative. If the two positive chambers of uu at the origin belong to different nodal domains, then uu has at least three nodal domains, since {u>0}\{u>0\} has at least two connected components and {u<0}\{u<0\} is nonempty. Second alternative. Suppose that the two positive chambers of uu at the origin belong to the same nodal domain. Then we can find a simple, closed curve γ{u>0}{(0,0)}\gamma\subset\{u>0\}\cup\{(0,0)\} that connects two points in distinct positive chambers of Bε(0)B_{\varepsilon}(0); see Figure 4.1. By the Jordan curve theorem, γ\gamma disconnects the two negative chambers at the origin. Hence uu has at least three nodal domains, since {u<0}\{u<0\} has at least two connected components and {u>0}\{u>0\} is nonempty. The base case holds.

For the induction step, suppose that there exists mm\in\mathbb{N} such that the lemma holds whenever 3nin++nout+m3\leq n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}\leq m. Assume that nin++nout+=m+1n^{+}_{\mathrm{in}}+n^{+}_{\mathrm{out}}=m+1 and nin+nout+n^{+}_{\mathrm{in}}\geq n^{+}_{\mathrm{out}}. We remark that this implies nin+12(m+1)2n^{+}_{\mathrm{in}}\geq\lceil\frac{1}{2}(m+1)\rceil\geq 2. We must prove that uu has at least m+1m+1 nodal domains. There are two cases, one easy, one harder.

Easy case. Suppose that no two chambers at the origin belong to the same nodal domain of uu. Then 𝒩(u)2nin+212(m+1)m+1\mathcal{N}(u)\geq 2n^{+}_{\mathrm{in}}\geq 2\lceil\frac{1}{2}(m+1)\rceil\geq m+1 (since nin+2n^{+}_{\mathrm{in}}\geq 2).

Harder case. Suppose that two distinct chambers CC and DD at the origin belong to the same nodal domain. Among all candidates, we choose CC and DD that have

(4.1) the least number of (positive and negative) chambers between them.\text{the least number of (positive and negative) chambers between them}.

Replacing uu by u-u, if needed, we may assume without loss of generality that CC and DD are positive chambers. By the alternating condition, the minimum number of chambers strictly between CC and DD is at least one and is odd. Suppose there are 2k+12k+1 such chambers and enumerate them and CC and DD in order:

(4.2) C,N1,P1,,Pk,Nk+1,D,C,\ N_{1},\ P_{1},\dots,\ P_{k},\ N_{k+1},\ D,

where N1,,Nk+1N_{1},\dots,N_{k+1} are negative and P1,,PkP_{1},\dots,P_{k} are positive. (When k=0k=0, the enumeration is C,N1,DC,N_{1},D.) Pause and note that the 2k+22k+2 chambers CC, N1N_{1}, P1P_{1}, …, PkP_{k}, Nk+1N_{k+1} lie in 2k+22k+2 disjoint nodal domains of uu, otherwise we would violate (4.1). Also note that k12nin+1k\leq\frac{1}{2}n^{+}_{\mathrm{in}}-1, when nin+n^{+}_{\mathrm{in}} is even, and k12nin+32k\leq\frac{1}{2}n^{+}_{\mathrm{in}}-\frac{3}{2} when nin+n^{+}_{\mathrm{in}} is odd. (To see this, draw some examples for small values of nin+n^{+}_{\mathrm{in}}.) Either way, nin+2k+2n^{+}_{\mathrm{in}}\geq 2k+2.

To proceed, let σ{u>0}{(0,0)}\sigma\subset\{u>0\}\cup\{(0,0)\} be a simple closed curve that connects a point in CC to a point in DD, passes through the origin in C¯\overline{C} and D¯\overline{D}, and encloses the intermediate chambers N1,P1,,Pk,Nk+1N_{1},P_{1},\cdots,P_{k},N_{k+1} in the bounded component Ω\Omega of 2σ\mathbb{R}^{2}\setminus\sigma. Next, let u~:2\tilde{u}:\mathbb{R}^{2}\rightarrow\mathbb{R} be any continuous function such that u~|2Ω=u\tilde{u}|_{\mathbb{R}^{2}\setminus\Omega}=u and u~>0\tilde{u}>0 throughout Ω\Omega. Effectively, this collapses the chambers in (4.2) into a single positive chamber. Since there are k+2k+2 positive chambers in (4.2), it follows that nin+(u~)=nin+(u)(k+1)n^{+}_{\mathrm{in}}(\tilde{u})=n^{+}_{\mathrm{in}}(u)-(k+1) and nout+(u~)=nout+(u)n^{+}_{\mathrm{out}}(\tilde{u})=n^{+}_{\mathrm{out}}(u). Suppose to get a contradiction that nin+(u~)+nout+(u~)2n^{+}_{\mathrm{in}}(\tilde{u})+n^{+}_{\mathrm{out}}(\tilde{u})\leq 2. Then

k+2=(2k+2)(k+1)+1nin+(u)(k+1)+nout+(u)=nin+(u~)+nout+(u~)2.k+2=(2k+2)-(k+1)+1\leq n^{+}_{\mathrm{in}}(u)-(k+1)+n^{+}_{\mathrm{out}}(u)=n^{+}_{\mathrm{in}}(\tilde{u})+n^{+}_{\mathrm{out}}(\tilde{u})\leq 2.

Hence k=0k=0 and 4m+1=nin+(u)+nout+(u)=nin+(u~)+nout+(u~)+134\leq m+1=n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)=n^{+}_{\mathrm{in}}(\tilde{u})+n^{+}_{\mathrm{out}}(\tilde{u})+1\leq 3, which is absurd. Therefore, 3nin+(u~)+nout+(u~)nin+(u)+nout+(u)1m3\leq n^{+}_{\mathrm{in}}(\tilde{u})+n^{+}_{\mathrm{out}}(\tilde{u})\leq n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)-1\leq m.

By the induction hypothesis, 𝒩(u~)nin+(u~)+nout+(u~)=nin+(u)+nout+(u)(k+1)\mathcal{N}(\tilde{u})\geq n^{+}_{\mathrm{in}}(\tilde{u})+n^{+}_{\mathrm{out}}(\tilde{u})=n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)-(k+1). Now, the nodal domains of u~\tilde{u} are in one-to-one correspondence with the nodal domains of uu that intersect nΩ\mathbb{R}^{n}\setminus\Omega. Together with the additional 2k+12k+1 nodal domains of uu inside Ω\Omega, corresponding to the chambers N1N_{1}, P1P_{1}, …, PkP_{k}, Nk+1N_{k+1}, we conclude that in total 𝒩(u)nin+(u)+nout+(u)+knin+(u)+nout+(u)=m+1\mathcal{N}(u)\geq n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)+k\geq n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)=m+1. This completes the induction step. ∎

Lemma 4.2.

If u=ud+fu=u_{d}+f near the origin in 2\mathbb{R}^{2}, where udu_{d} is an hhp in 2\mathbb{R}^{2} of degree d1d\geq 1, fC1f\in C^{1}, and in polar coordinates, |f(r,θ)|+|θf(r,θ)|=o(rd)|f(r,\theta)|+\left\lvert\partial_{\theta}f(r,\theta)\right\rvert=o(r^{d}) as r0r\rightarrow 0, then uu has 2d2d chambers with alternating signs in all sufficiently small neighborhoods of the origin and nin+(u)=dn^{+}_{\mathrm{in}}(u)=d.

Proof.

Up to a rotation and renormalization, udu_{d} is given in polar coordinates by ud(r,θ)=rdsin(dθ)u_{d}(r,\theta)=r^{d}\sin(d\theta). Thus the nodal domains of udu_{d} consist of 2d2d sectors at the origin with opening π/d\pi/d with alternating signs. If we consider the function gr(θ):𝕊1g_{r}(\theta):\mathbb{S}^{1}\rightarrow\mathbb{R} defined by gr(θ)=rdu(r,θ)g_{r}(\theta)=r^{-d}u(r,\theta), then zero set of grg_{r} is the same as that of uu intersected with Br(0)\partial B_{r}(0). Moreover, the assumptions on ff near the origin give us that

(4.3) gr(θ)=sin(dθ)+e1(r,θ),θgr(θ)=dcos(dθ)+e2(r,θ),\begin{split}g_{r}(\theta)=\sin(d\theta)+e_{1}(r,\theta),\\ \partial_{\theta}g_{r}(\theta)=d\cos(d\theta)+e_{2}(r,\theta),\end{split}

where |ei(r,θ)|0\left\lvert e_{i}(r,\theta)\right\rvert\rightarrow 0 as r0r\rightarrow 0 for i=1,2i=1,2. Since |cos(dθ)|=1\left\lvert\cos(d\theta)\right\rvert=1 on whenever sin(dθ)=0\sin(d\theta)=0, it is straight-forward to check from (4.3) that the following holds. For each ε>0\varepsilon>0, there is some r0>0r_{0}>0 small so that for all 0<r<r00<r<r_{0}, gr(θ)g_{r}(\theta) has exactly 2d2d zeros in 𝕊1\mathbb{S}^{1} (one in each of the sectors {|θjπ/d|<ε}\{\left\lvert\theta-j\pi/d\right\rvert<\varepsilon\} for 0j2d10\leq j\leq 2d-1), at which grg_{r} changes sign. If we order such zeros θ0(r),θ1(r),,θ2d1(r)\theta_{0}(r),\theta_{1}(r),\dotsc,\theta_{2d-1}(r), then the θj(r)\theta_{j}(r) are continuous in rr since uu is a continuous function, and θj(r)jπ/d\theta_{j}(r)\rightarrow j\pi/d as r0r\downarrow 0. Recalling that the nodal set of grg_{r} coincides with 𝒩(u)Br(0)\mathcal{N}(u)\cap\partial B_{r}(0), one obtains the desired conclusion. ∎

As an application of Lemmas 4.1 and 4.2, we obtain a lower bound on the number of nodal domains of an hcp in 2+1\mathbb{R}^{2+1} whose harmonic coefficient has degree at least 2.

Proposition 4.3.

If p(x,y,t)p(x,y,t) is an hcp in 2+1\mathbb{R}^{2+1} of the form (2.1) and degpm2\deg p_{m}\geq 2, then pp has at least 2degpm2\deg p_{m} nodal domains.

Proof.

Put d=degpmd=\deg p_{m}. By Remark 1.9, it suffices to prove that the nodal set of p|𝕊2p|_{\mathbb{S}^{2}} has at least 2d2d nodal domains. Since d1d\geq 1, p(N)=p(S)=0p(N)=p(S)=0, where N=(0,0,1)N=(0,0,1) and S=(0,0,1)S=(0,0,-1) are the north and south poles of 𝕊2\mathbb{S}^{2}, respectively. Let πN\pi_{N} denote the stereographic projection from the north pole of 𝕊2\mathbb{S}^{2} onto ^2=2{}\widehat{\mathbb{R}}^{2}=\mathbb{R}^{2}\cup\{\infty\}. Then πN(S)=0\pi_{N}(S)=0 and πN(N)=\pi_{N}(N)=\infty. Thus, if we can show that the nodal domains of u:=pπN1u:=p\circ\pi_{N}^{-1} near the origin and near infinity are homeomorphic to 2d2d sectors with alternating signs, then nin+(u)=nout+(u)=dn^{+}_{\mathrm{in}}(u)=n^{+}_{\mathrm{out}}(u)=d and 𝒩(p)=𝒩(p|𝕊2)=𝒩(u)nin+(u)+nout+(u)=2d4\mathcal{N}(p)=\mathcal{N}(p|_{\mathbb{S}^{2}})=\mathcal{N}(u)\geq n^{+}_{\mathrm{in}}(u)+n^{+}_{\mathrm{out}}(u)=2d\geq 4 by Lemma 4.1.

The stereographic projection πN:𝕊2{N}2\pi_{N}:\mathbb{S}^{2}\setminus\{N\}\rightarrow\mathbb{R}^{2} is given by

πN(x,y,t)\displaystyle\pi_{N}(x,y,t) =(x1t,y1t),πN1(X,Y)=(2XR2+1,2YR2+1,R21R2+1),\displaystyle=\left(\dfrac{x}{1-t},\dfrac{y}{1-t}\right),\quad\pi_{N}^{-1}(X,Y)=\left(\dfrac{2X}{R^{2}+1},\dfrac{2Y}{R^{2}+1},\dfrac{R^{2}-1}{R^{2}+1}\right),

where R=R(X,Y)=(X2+Y2)1/2R=R(X,Y)=(X^{2}+Y^{2})^{1/2}. Recalling (2.1), we have

(4.4) u(X,Y)=j=0m(R21R2+1)jpj(2XR2+1,2YR2+1)=j=0m(R21R2+1)j(2R2+1)d+2(mj)pj(X,Y)=:j=0maj(R)pj(X,Y),\begin{split}u(X,Y)&=\sum_{j=0}^{m}\left(\dfrac{R^{2}-1}{R^{2}+1}\right)^{j}p_{j}\left(\dfrac{2X}{R^{2}+1},\dfrac{2Y}{R^{2}+1}\right)\\ &=\sum_{j=0}^{m}\left(\dfrac{R^{2}-1}{R^{2}+1}\right)^{j}\left(\dfrac{2}{R^{2}+1}\right)^{d+2(m-j)}p_{j}(X,Y)=:\sum_{j=0}^{m}a_{j}(R)p_{j}(X,Y),\end{split}

where each term pjp_{j} is algebraically homogeneous of order d+2(mj)d+2(m-j), the lowest order term pmp_{m} is an hhp of degree dd (see e.g. Figure 3.1 for the case d=4d=4), and the coefficients aj(R)a_{j}(R) are radial, bounded real-analytic functions of XX and YY with limR0aj(R)=(1)j2d+2(mj)\lim_{R\rightarrow 0}a_{j}(R)=(-1)^{j}2^{d+2(m-j)}. In particular, ama_{m} does not vanish in {0<R1/2}\{0<R\leq 1/2\}. It follows that f:=u/ampmf:=u/a_{m}-p_{m} is real-analytic near 0 and in polar coordinates, |f(R,θ)|+|θf(R,θ)|=o(Rd)|f(R,\theta)|+\left\lvert\partial_{\theta}f(R,\theta)\right\rvert=o(R^{d}) as R0+R\rightarrow 0^{+}, since each of the terms p0p_{0}, …, pm1p_{m-1} is homogeneous with degree at least d+2d+2. By Lemma 4.2, we conclude that u/am=pm+fu/a_{m}=p_{m}+f has 2d2d chambers at the origin with alternating signs. Thus, since ama_{m} does not vanish near the origin, u=am(u/am)u=a_{m}(u/a_{m}) also has 2d2d chambers at the origin with alternating signs and nin+(u)=nin+(u/am)=dn^{+}_{\mathrm{in}}(u)=n^{+}_{\mathrm{in}}(u/a_{m})=d.

Finally, let v:=pπS1v:=p\circ\pi_{S}^{-1}, where πS\pi_{S} denotes the stereographic projection from the south pole of 𝕊2\mathbb{S}^{2} onto ^2\widehat{\mathbb{R}}^{2}. Repeating the argument for uu shows that nin+(v)=dn^{+}_{\mathrm{in}}(v)=d. Therefore, because uu and vv are related by inversion, we have nout+(u)=nin+(v)=dn^{+}_{\mathrm{out}}(u)=n^{+}_{\mathrm{in}}(v)=d.∎

Next, we present a lower bound on the number of nodal domains for a certain class of hcps of degree 4k4k.

Proposition 4.4.

If p(x,y,t)p(x,y,t) is an an hcp in 2+1\mathbb{R}^{2+1} of the form

(4.5) p(x,y,t)\displaystyle p(x,y,t) =t2k+t2k1p2k1(x,y)++p0(x,y)\displaystyle=t^{2k}+t^{2k-1}p_{2k-1}(x,y)+\cdots+p_{0}(x,y)

for some k1k\geq 1, then pp has at least 33 nodal domains.

Refer to caption
Figure 4.2. Proof of Proposition 4.4. If the union γ~\tilde{\gamma} of the path γ\gamma and its reflection RγR\circ\gamma lie in the positivity set of p|𝕊2p|_{\mathbb{S}^{2}}, then the negativity set of p|𝕊2p|_{\mathbb{S}^{2}} is
disconnected.
Proof.

By Remark 1.9, it suffices to prove that the positivity set of p|𝕊2p|_{\mathbb{S}^{2}} is disconnected or the negativity set of p|𝕊2p|_{\mathbb{S}^{2}} is disconnected. Thus, if {p|𝕊2>0}\{p|_{\mathbb{S}^{2}}>0\} is disconnected, we are done.

Suppose that the positivity set of p|𝕊2p|_{\mathbb{S}^{2}} is connected. By (4.5), p(0,0,1)=p(0,0,1)=1p(0,0,1)=p(0,0,-1)=1. Thus, the north pole N=(0,0,1)N=(0,0,1) and south pole S=(0,0,1)S=(0,0,-1) belong to the positivity set of p|𝕊2p|_{\mathbb{S}^{2}}. Hence there exists a simple path γ:[0,1]{p|𝕊2>0}\gamma:[0,1]\rightarrow\{p|_{\mathbb{S}^{2}}>0\} such that γ(0)=N\gamma(0)=N and γ(1)=S\gamma(1)=S. Now, when tt is fixed, p(x,y,t)p(x,y,t) is an even polynomial in xx and yy, since each coefficient pi(x,y)p_{i}(x,y) in (4.5) has even degree. In particular, writing R(x,y,t)=(x,y,t)R(x,y,t)=(-x,-y,t), we have RγR\circ\gamma is a mirrored path from NN to SS contained in {p|𝕊2>0}\{p|_{\mathbb{S}^{2}}>0\}. Let γ~\tilde{\gamma} be the union of the traces of γ\gamma and its reflection RγR\circ\gamma. See Figure 4.2. By the Jordan curve theorem, γ~\tilde{\gamma} separates 𝕊2γ~\mathbb{S}^{2}\setminus\tilde{\gamma} into two connected components. Let Q𝕊2Q\in\mathbb{S}^{2} be any point such that p(Q)<0p(Q)<0. Then p(R(Q))=p(Q)<0p(R(Q))=p(Q)<0, but QQ and R(Q)R(Q) lie in different connected components of 𝕊2γ~\mathbb{S}^{2}\setminus\tilde{\gamma}. Therefore, the negativity set of p|𝕊2p|_{\mathbb{S}^{2}} is disconnected. ∎

Combining the previous two results, we see that m2,d3m_{2,d}\geq 3 when d0(mod4)d\equiv 0\pmod{4}.

Corollary 4.5 (A lower bound for d0(mod4)d\equiv 0\pmod{4}).

If d0(mod4)d\equiv 0\pmod{4} and d4d\geq 4, then any time-dependent hcp pp of degree dd in 2+1\mathbb{R}^{2+1} has at least 33 nodal domains. Hence m2,d3m_{2,d}\geq 3.

Proof.

Since pp has degree d=4k4d=4k\geq 4, its leading tt term is either t2kt^{2k} or of the form tmpm(x,y)t^{m}p_{m}(x,y) with degpm2\deg p_{m}\geq 2. In the first case, 𝒩(p)3\mathcal{N}(p)\geq 3 by Proposition 4.4. In the second case, 𝒩(p)2degpm4\mathcal{N}(p)\geq 2\deg p_{m}\geq 4 by Proposition 4.3. ∎

5. HCP in 2+1\mathbb{R}^{2+1}, Part II: constructions

In this section, we construct examples of time-dependent hcps in 2+1\mathbb{R}^{2+1} of degree d2d\geq 2 with two nodal domains when d0(mod4)d\not\equiv 0\pmod{4} and with three nodal domains when d0(mod4)d\equiv 0\pmod{4}. By Remark 1.9, counting the nodal domains of an hcp pp is equivalent to counting the nodal domains of p|𝕊2p|_{\mathbb{S}^{2}}. With this reduction, the general strategy in constructing examples is the same in all cases (and is the strategy introduced by [Lew77] and used by [EJN07], [LTY15], and [BH16] in related contexts):

  1. (a)

    Begin with an hcp ϕ1\phi_{1} of degree dd whose nodal set can be described explicitly.

  2. (b)

    Find another hcp ϕ2\phi_{2} of degree dd so that the nodal set of the perturbation u=ϕ1εϕ2u=\phi_{1}-\varepsilon\phi_{2} in 𝕊2\mathbb{S}^{2} is either one Jordan curve (uu has two nodal domains) or the nodal set of uu is the union of two disjoint Jordan curves (uu has three nodal domains).

The key difficulty in this strategy is finding certain compatibility conditions between ϕ1,ϕ2\phi_{1},\phi_{2}. In general, the nodal set {ϕ1|𝕊2=0}\{\phi_{1}|_{\mathbb{S}^{2}}=0\} is the union of a relatively open smooth set, where |ϕ1|0|\nabla\phi_{1}|\neq 0, and isolated singular points, where |ϕ1|=0|\nabla\phi_{1}|=0. Understanding the picture near smooth portion of the nodal set is straightforward; see e.g. [Lew77, Lemma 2]. However, understanding how the nodal domains of ϕ1\phi_{1} change under perturbation near a singular point is quite delicate, requiring knowledge both of the local structure of ϕ1\phi_{1} and of the sign of ϕ2\phi_{2}. This makes finding ϕ1\phi_{1} and ϕ2\phi_{2} challenging. After reviewing the main perturbation lemmas, we present our examples in order of increasing difficulty: d2(mod4)d\equiv 2\pmod{4} in §5.1, dd odd in §5.2, and d0(mod4)d\equiv 0\pmod{4} in §5.3.

Refer to caption
Figure 5.1. Zero set of ψ(x,y)=Im((x+iy)6)\psi(x,y)=\mathrm{Im}((x+iy)^{6}) (in black) and its perturbation ψ(x,y)εf(x,y)\psi(x,y)-\varepsilon f(x,y) near the origin when ff is C1C^{1}, ε\varepsilon is small, and f(0,0)>0f(0,0)>0 (left, in green) or f(0,0)<0f(0,0)<0 (right, in red). The green lines lie in the positivity set for ψ\psi, whereas the red lines lie in the negativity set for ψ\psi.

The first of two perturbation lemmas that we need, Lemma 5.1, is the consequence of [Lew77, Lemma 4] recorded on the second paragraph on p. 1239 of Lewy’s paper. In the original source, it is stated that ff should be real-analytic, but inspecting the proof shows that it suffices to assume ff is C1C^{1}.

Lemma 5.1.

Let ψ(x,y)=Im((x+iy)d)\psi(x,y)=\mathrm{Im}((x+iy)^{d}) for some d2d\geq 2. If f:Br(0)f:B_{r}(0)\rightarrow\mathbb{R} is C1C^{1} and f(0,0)>0f(0,0)>0, then there exists τ(0,r)\tau\in(0,r) and ε0>0\varepsilon_{0}>0 such that for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the nodal set of ψεf\psi-\varepsilon f in Bτ(0)B_{\tau}(0) consists of dd pairwise disjoint simple curves, with one curve inside each of the connected components of {ψ>0}\{\psi>0\}, and

limε0HD({ψεf=0}Bτ(0),{ψ=0}Bτ(0))=0.\lim_{\varepsilon\rightarrow 0}\mathrm{HD}\left(\{\psi-\varepsilon f=0\}\cap B_{\tau}(0),\{\psi=0\}\cap B_{\tau}(0)\right)=0.

The same conclusion holds when f(0,0)<0f(0,0)<0 except that then the nodal set of the perturbation ψεf\psi-\varepsilon f lies in {ψ<0}\{\psi<0\}. See Figure 5.1.

We also need a variant of Lemma 5.1, in which ψ(x,y)\psi(x,y) is replaced by a function G(x,y)G(x,y) whose nodal set is given locally by the union of mm graphs with simple intersection at a common point. We emphasize that the following lemma is inspired by [Lew77].

Lemma 5.2.

Suppose that G:Br(0)2G:B_{r}(0)\subset\mathbb{R}^{2}\rightarrow\mathbb{R} takes the form G(x,y)=i=1mgi(x,y)G(x,y)=\prod_{i=1}^{m}g_{i}(x,y) for some m2m\geq 2, where g1,,gm:Br(0)g_{1},\dots,g_{m}:B_{r}(0)\rightarrow\mathbb{R} are real-analytic functions satisfying

  • gi(0,0)=0g_{i}(0,0)=0 and ygi(0,0)0\partial_{y}g_{i}(0,0)\neq 0 for all ii,

  • {gi=0}{gj=0}={(0,0)}\{g_{i}=0\}\cap\{g_{j}=0\}=\{(0,0)\} for all iji\neq j.

If F:Br(0)F:B_{r}(0)\rightarrow\mathbb{R} is C1C^{1} and F(0,0)>0F(0,0)>0, then there exists τ(0,r)\tau\in(0,r) and ε0>0\varepsilon_{0}>0 such that for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the nodal set of GεFG-\varepsilon F in Bτ(0)B_{\tau}(0) consists of mm pairwise disjoint simple curves, one inside each of the mm connected components of {G>0}\{G>0\}, and

limε0HD({GεF=0}Bτ(0),{G=0}Bτ(0))=0,\lim_{\varepsilon\rightarrow 0}\mathrm{HD}\left(\{G-\varepsilon F=0\}\cap B_{\tau}(0),\{G=0\}\cap B_{\tau}(0)\right)=0,

where HD\mathrm{HD} is the Hausdorff distance. The same conclusion holds when F(0,0)<0F(0,0)<0 except that then the nodal set of the perturbation GεFG-\varepsilon F lies in {G<0}\{G<0\}. See Figure 5.2.

Refer to caption
Refer to caption
Refer to caption
Figure 5.2. Zero set of G(x,y)=(x4yy2)(x2(x21)+12y)(3x3y)G(x,y)=(x^{4}-y-y^{2})(x^{2}(x^{2}-1)+\frac{1}{2}y)(3x^{3}-y) and its perturbation GεFG-\varepsilon F: ε=105\varepsilon=10^{-5}, F(x,y)=1F(x,y)=1 (left), F(x,y)=1F(x,y)=-1 (right).
Remark 5.3.

When F(0,0)>0F(0,0)>0, the negativity set of GεFG-\varepsilon F near the origin is connected. When F(0,0)<0F(0,0)<0, the positivity set of GεFG-\varepsilon F near the origin is connected.

We defer the proof of Lemma 5.2, which may be considered a (somewhat long) exercise with the implicit function theorem, to §6. Note that by rotating coordinates, Lemma 5.1 follows from Lemma 5.2.

5.1. Two nodal domains when d2(mod4)d\equiv 2\pmod{4}

When d=4k+2d=4k+2 for some k0k\geq 0, the basic hcp pd(x,t)=t2k+1+c2kt2kx2++c0x4k+2p_{d}(x,t)=t^{2k+1}+c_{2k}t^{2k}x^{2}+\cdots+c_{0}x^{4k+2} in 1+1\mathbb{R}^{1+1} (see Definition 2.1) satisfies

pd(0,1)>0andpd(0,1)<0.p_{d}(0,1)>0\quad\text{and}\quad p_{d}(0,-1)<0.

This simple observation will let us build hcps in 2+1\mathbb{R}^{2+1} of degree dd with two nodal domains by essentially copying Lewy’s construction of odd degree hhps in 3\mathbb{R}^{3} with two nodal domains.

Theorem 5.4 (cf. [Lew77, Theorem 1]).

Assume d=4k+2d=4k+2 for some k0k\geq 0. Let ψ(x,y)=Im((x+iy)d)\psi(x,y)=\mathrm{Im}((x+iy)^{d}) and let pd(x,t)p_{d}(x,t) be the basic hcp in 1+1\mathbb{R}^{1+1}. For all sufficiently small ε>0\varepsilon>0,

(5.1) uε(x,y,t)ψ(x,y)εpd(x,t)u_{\varepsilon}(x,y,t)\coloneqq\psi(x,y)-\varepsilon p_{d}(x,t)

is a time-dependent hcp in 2+1\mathbb{R}^{2+1} of degree dd and uεu_{\varepsilon} has two nodal domains.

Proof.

It is clear that uεu_{\varepsilon} is a time-dependent hcp of degree dd whenever ε0\varepsilon\neq 0. To proceed, we argue that the nodal set of uεu_{\varepsilon} looks like the one in Figure 1.1 (right) when ε>0\varepsilon>0 is small. Note that the nodal set of ψ|𝕊2\psi|_{\mathbb{S}^{2}} has only two singular points: the north and south pole.

We start with a description of the nodal set of uε|𝕊2u_{\varepsilon}|_{\mathbb{S}^{2}} near the north pole. Define vε(x,y):=ψ(x,y)εpd(x,1)v_{\varepsilon}(x,y):=\psi(x,y)-\varepsilon p_{d}(x,1) so that by parabolic homogeneity,

uε(x,y,t)=td/2uε(x/t,y/t,1)=td/2vε(x/t,y/t)for all t>0.u_{\varepsilon}(x,y,t)=t^{d/2}u_{\varepsilon}(x/\sqrt{t},y/\sqrt{t},1)=t^{d/2}v_{\varepsilon}(x/\sqrt{t},y/\sqrt{t})\quad\text{for all $t>0$}.

Hence the nodal set of uε|𝕊2u_{\varepsilon}|_{\mathbb{S}^{2}} in each time-slice {t=t0}\{t=t_{0}\} (with 0<t0<10<t_{0}<1) agrees with the zeros of vεv_{\varepsilon} on the circle {x2+y2=(1t02)/t0}\{x^{2}+y^{2}=(1-t_{0}^{2})/t_{0}\}. Thus, the nodal set uεu_{\varepsilon} on a small spherical cap at the north pole is homeomorphic to the nodal set of vεv_{\varepsilon} in a small disk at the origin (see Figure 5.1). Recall that pd(0,1)>0p_{d}(0,1)>0. Therefore, we can use Lemma 5.1 to conclude that when ε>0\varepsilon>0 is sufficiently small, the nodal set of uεu_{\varepsilon} in a small spherical cap at the north pole consists of dd “southward-opening U-shaped” curves lying in every other longitudinal sector in 𝕊2\mathbb{S}^{2} of angle ϑ:=πd\vartheta:=\frac{\pi}{d}, starting with {0<θ<ϑ}\{0<\theta<\vartheta\} and ending with {(2d2)ϑ<θ<(2d1)ϑ}\{(2d-2)\vartheta<\theta<(2d-1)\vartheta\}.

For all t<0t<0, parabolic homogeneity instead yields

uε(x,y,t)=|t|d/2uε(x/|t|,y/|t|,1)=|t|d/2(ψ(x/|t|,y/|t|)εpd(x/|t|,1)).u_{\varepsilon}(x,y,t)=|t|^{d/2}u_{\varepsilon}(x/\sqrt{|t|},y/\sqrt{|t|},-1)=|t|^{d/2}\big{(}\psi(x/\sqrt{|t|},y/\sqrt{|t|})-\varepsilon p_{d}(x/\sqrt{|t|},-1)\big{)}.

Since pd(0,1)<0p_{d}(0,-1)<0, we can use Lemma 5.1 to conclude that when ε>0\varepsilon>0 is sufficiently small, the nodal set of uεu_{\varepsilon} in a small spherical cap at the south pole consists of dd “northward-opening U-shaped” curves lying in every other longitudinal sector in 𝕊2\mathbb{S}^{2} of angle ϑ=πd\vartheta=\frac{\pi}{d}, starting with {ϑ<θ<2ϑ}\{\vartheta<\theta<2\vartheta\} and ending with {(2d1)ϑ<θ<2dϑ}\{(2d-1)\vartheta<\theta<2d\vartheta\}.

Outside of the polar regions, i.e. in the complement of the union of fixed spherical caps at the north and south pole, the nodal set of ψ|𝕊2(x,y)\psi|_{\mathbb{S}^{2}}(x,y) consists of dd disjoint smooth arcs, along which |ψ(x,y)|c>0|\nabla\psi(x,y)|\geq c>0 for some constant cc (depending on the size of the caps). Hence the same is true for uε|𝕊2u_{\varepsilon}|_{\mathbb{S}^{2}} when ε\varepsilon is sufficiently small by the implicit function theorem.

In the end, since the chambers occupied at the north and south pole alternate, we see that when ε\varepsilon is sufficiently small, the nodal set of uε|𝕊2u_{\varepsilon}|_{\mathbb{S}^{2}} is a single closed, smooth, Jordan curve. Therefore, uεu_{\varepsilon} has two nodal domains.∎

5.2. Two nodal domains when dd is odd

See Figure 1.1 (middle) for an illustration of the example in Theorem 5.5 when d=5d=5.

Theorem 5.5.

Assume d3d\geq 3 is odd. For all sufficiently small ε>0\varepsilon>0 and α>0\alpha>0,

(5.2) uε,α(x,y,t):=ypd1(x,t)εpd(xcosαysinα,t)\displaystyle u_{\varepsilon,\alpha}(x,y,t):=yp_{d-1}(x,t)-\varepsilon p_{d}(x\cos\alpha-y\sin\alpha,t)

is a time-dependent hcp in 2+1\mathbb{R}^{2+1} of degree dd and uε,αu_{\varepsilon,\alpha} has two nodal domains.

Proof.

Let d=2k+1d=2k+1 for some k1k\geq 1. By Theorem 2.4, we can write the basic hcps pd1p_{d-1} and pdp_{d} in 1+1\mathbb{R}^{1+1} from Definition 2.1 as

pd1(x,t)=(t+b1x2)(t+bkx2)andpd(x,t)=x(t+c1x2)(t+ckx2)p_{d-1}(x,t)=(t+b_{1}x^{2})\cdots(t+b_{k}x^{2})\quad\text{and}\quad p_{d}(x,t)=x(t+c_{1}x^{2})\cdots(t+c_{k}x^{2})

for some numbers 0<c1<b1<c2<b2<<bk1<ck<bk0<c_{1}<b_{1}<c_{2}<b_{2}<\cdots<b_{k-1}<c_{k}<b_{k}. Note that the expression pd(xcosαysinα,t)p_{d}(x\cos\alpha-y\sin\alpha,t) is just the composition of pd(x,t)p_{d}(x,t) with a rotation in the xx and yy coordinates and the Laplacian is rotationally-invariant. Thus, pd(xcosαysinα,t)p_{d}(x\cos\alpha-y\sin\alpha,t) and uε,α(x,y,t)u_{\varepsilon,\alpha}(x,y,t) are time-dependent hcps in 2+1\mathbb{R}^{2+1} for all ε\varepsilon and α\alpha. To proceed, fix ε>0\varepsilon>0 and α>0\alpha>0 (small) and write u=uε,αu=u_{\varepsilon,\alpha}, p=pd1p=p_{d-1}, q=pdq=p_{d}, and qα(x,y,t)=pd(xcosαysinα,t)q_{\alpha}(x,y,t)=p_{d}(x\cos\alpha-y\sin\alpha,t). Our goal is to show that when ε\varepsilon and α\alpha are small enough that {u=0}𝕊2\{u=0\}\cap\mathbb{S}^{2} is a Jordan curve, whence 𝒩(u)=𝒩(u|𝕊2)=2\mathcal{N}(u)=\mathcal{N}(u|_{\mathbb{S}^{2}})=2.

Refer to caption
Refer to caption
Refer to caption
Figure 5.3. Proof of Theorem 5.5 (1/2): Nodal set of p¯\overline{p} (top), q¯\overline{q} (middle), and u¯\overline{u} (bottom) when k=3k=3 and ε\varepsilon and α\alpha are sufficiently small.

Consider the standard spherical coordinates on 𝕊2\mathbb{S}^{2} given by

(5.3) x=cosθcosϕ,y=sinθcosϕ,t=sinϕ,π<θπ,π/2ϕπ/2x=\cos\theta\cos\phi,\;y=\sin\theta\cos\phi,\;t=\sin\phi,\qquad-\pi<\theta\leq\pi,\;-\pi/2\leq\phi\leq\pi/2

and write p¯\overline{p}, q¯\overline{q}, q¯α\overline{q}_{\alpha}, and u¯\overline{u} for the functions corresponding to ypd(x,t)yp_{d}(x,t), q(x,t)q(x,t), qα(x,y,t)q_{\alpha}(x,y,t), and uε,α(x,y,t)u_{\varepsilon,\alpha}(x,y,t) on 𝕊2\mathbb{S}^{2} written in spherical coordinates. Hence

(5.4) p¯(θ,ϕ)=sinθcosϕi=1k(sinϕ+bicos2θcos2ϕ),q¯(θ,ϕ)=cosθcosϕi=1k(sinϕ+cicos2θcos2ϕ),q¯α(θ,ϕ)=q¯(θ+α,ϕ),u¯(θ,ϕ)=p¯(θ,ϕ)εq¯α(θ,ϕ).\begin{split}\overline{p}(\theta,\phi)&=\sin\theta\cos\phi\prod_{i=1}^{k}\left(\sin\phi+b_{i}\cos^{2}\theta\cos^{2}\phi\right),\\ \overline{q}(\theta,\phi)&=\cos\theta\cos\phi\prod_{i=1}^{k}\left(\sin\phi+c_{i}\cos^{2}\theta\cos^{2}\phi\right),\\ \overline{q}_{\alpha}(\theta,\phi)&=\overline{q}(\theta+\alpha,\phi),\quad\overline{u}(\theta,\phi)=\overline{p}(\theta,\phi)-\varepsilon\overline{q}_{\alpha}(\theta,\phi).\end{split}

As an aid for the reader, in Figure 5.3, we depict nodal domains of p¯\overline{p}, q¯\overline{q}, and u¯\overline{u} in the θϕ\theta\phi-plane when k=3k=3. When α>0\alpha>0 is small, the picture for q¯α\overline{q}_{\alpha} is obtained by translating the picture for q¯\overline{q} slightly to the left. This is crucial for the construction. Observe that the positivity and negativity sets for u¯\overline{u} in the figure are connected. We must explain why this is so.

The nodal set of p¯\overline{p} is comprised of the great circle {θ=0 or π}\{\theta=0\text{ or }\pi\} (including the north and south poles {ϕ=±π2}\{\phi=\pm\frac{\pi}{2}\}) and the graphs {(θ,hi(θ)):π<θπ}\{(\theta,h_{i}(\theta)):-\pi<\theta\leq\pi\} of the kk functions

(5.5) hi(θ):=sin1(11+4bi2cos4θ2bicos2θ)(i=1,,k).h_{i}(\theta):=\sin^{-1}\left(\frac{1-\sqrt{1+4b_{i}^{2}\cos^{4}\theta}}{2b_{i}\cos^{2}\theta}\right)\qquad(i=1,\dots,k).

To find the formula for hih_{i}, expand cos2ϕ=1sin2ϕ\cos^{2}\phi=1-\sin^{2}\phi in the equation sinϕ+bicos2θcos2ϕ=0\sin\phi+b_{i}\cos^{2}\theta\cos^{2}\phi=0 and use the quadratic formula to solve for sinϕ\sin\phi (recalling the restriction that |sinϕ|1|\sin\phi|\leq 1). The functions hih_{i} are π\pi-periodic, hi(±π/2)=0h_{i}(\pm\pi/2)=0, and hi(θ)>hi+1(θ)h_{i}(\theta)>h_{i+1}(\theta) for all 1ik11\leq i\leq k-1 and θ±π/2\theta\neq\pm\pi/2, since b1<<bkb_{1}<\cdots<b_{k}. Moreover, from the definition of p¯\overline{p}, we observe that p¯\overline{p} takes opposite signs in adjacent nodal domains as indicated in Figure 5.3.

To understand the nodal structure of u¯\overline{u} when ε\varepsilon is small using Lemma 5.2 and Remark 5.3, we must determine the signs of q¯α\overline{q}_{\alpha} at the singular points in the nodal set of p¯\overline{p} (i.e. the points where two or more nodal lines of p¯\overline{p} intersect). See Figure 5.4.

Refer to caption
Figure 5.4. Proof of Theorem 5.5 (2/2): Nodal sets of p¯\overline{p} (top) and u¯\overline{u} (bottom) near θ=0\theta=0 (left) and θ=π/2\theta=\pi/2 (right) when k=3k=3. The sign of q¯α\overline{q}_{\alpha} at singular points in the nodal set of p¯\overline{p} determines the local configuration of nodal domains of u¯\overline{u} (see Lemma 5.2 and Remark 5.3).

The nodal set of q¯\overline{q} is the union of the great circle {θ=±π/2}\{\theta=\pm\pi/2\} and the graphs {(θ,gi(θ)):π<θπ}\{(\theta,g_{i}(\theta)):-\pi<\theta\leq\pi\} of the kk functions

(5.6) gi(θ):=sin1(11+4ci2cos4θ2cicos2θ)(i=1,,k).g_{i}(\theta):=\sin^{-1}\left(\frac{1-\sqrt{1+4c_{i}^{2}\cos^{4}\theta}}{2c_{i}\cos^{2}\theta}\right)\qquad(i=1,\dots,k).

Because of the interlacing property c1<b1<<ck<bkc_{1}<b_{1}<\cdots<c_{k}<b_{k}, the order of the graphs of the functions gig_{i} and hih_{i} alternate: g1h1gkhkg_{1}\geq h_{1}\geq\cdots\geq g_{k}\geq h_{k} with strict inequality when θ±π/2\theta\neq\pm\pi/2. In particular, along the lines θ=0\theta=0 and θ=π\theta=\pi, the sign of the function q¯\overline{q} at the singular point in the nodal set of p¯\overline{p} corresponding to hih_{i} is (1)i(-1)^{i} when θ=0\theta=0 and (1)i+1(-1)^{i+1} when θ=π\theta=\pi. By continuity, the same alternating sign pattern persists for q¯α\overline{q}_{\alpha} if α\alpha is sufficiently small. At the two remaining singular points (π/2,0)(-\pi/2,0) and (π/2,0)(\pi/2,0) in the nodal set of p¯\overline{p}, the function q¯\overline{q} is zero.111Exceptionally, when d=3d=3 and k=1k=1, the nodal set for p¯\overline{p} is regular at (±π/2,0)(\pm\pi/2,0). This is why we introduce the parameter α\alpha. Taking α>0\alpha>0 (and small), we get q¯α(π/2,0)=q¯((π/2)+α,0)>0\overline{q}_{\alpha}(-\pi/2,0)=\overline{q}((-\pi/2)+\alpha,0)>0 and q¯α(π/2,0)=q¯((π/2)+α)<0\overline{q}_{\alpha}(\pi/2,0)=\overline{q}((\pi/2)+\alpha)<0.

By the perturbation lemma, the nodal lines for p¯\overline{p} transform into the nodal lines for u¯\overline{u} as indicated in the figures provided that ε\varepsilon is sufficiently small (with α\alpha fixed before ε\varepsilon). Outside of a neighborhood of the singular points, the nodal lines for u¯\overline{u} are homeomorphic to the nodal lines for p¯\overline{p} by the implicit function theorem; cf. proof of Theorem 5.4. (We have purposely ignored the singularities of p¯\overline{p} on the lines ϕ=±π/2\phi=\pm\pi/2, because the nodal lines of p|𝕊2p|_{\mathbb{S}^{2}} are regular at the north and south poles.) A careful stitching of the local structure of {u¯=0}\{\overline{u}=0\} yields that {u=0}𝕊2\{u=0\}\cap\mathbb{S}^{2} is a single Jordan curve. ∎

5.3. Three nodal domains when d0(mod4)d\equiv 0\pmod{4}

See Figure 1.1 (left) for an illustration of the example in Theorem 5.6 when d=4d=4.

Theorem 5.6.

Assume d=4kd=4k for some k1k\geq 1. For all sufficiently small ε>0\varepsilon>0 and α>0\alpha>0,

(5.7) uε,α(x,y,t):=p2k(x,t)p2k(y,t)+εp2k+1(xcosαysinα,t)p2k1(xsinα+ycosα,t)\displaystyle u_{\varepsilon,\alpha}(x,y,t):=p_{2k}(x,t)p_{2k}(y,t)+\varepsilon p_{2k+1}(x\cos\alpha-y\sin\alpha,t)p_{2k-1}(x\sin\alpha+y\cos\alpha,t)

is a time-dependent hcp in 2+1\mathbb{R}^{2+1} of degree dd and uε,αu_{\varepsilon,\alpha} has three nodal domains.

Proof.

The shell of the proof is the same as for proof of Theorem 5.5, but the construction is more intricate. Assign p(x,y,t):=p2k(x,t)p2k(y,t)p(x,y,t):=p_{2k}(x,t)p_{2k}(y,t) and q(x,y,t):=p2k+1(x,t)p2k1(y,t)q(x,y,t):=p_{2k+1}(x,t)p_{2k-1}(y,t). By Theorem 2.4, we may express

p(x,y,t)\displaystyle p(x,y,t) =(t+b1x2)(t+bkx2)(t+b1y2)(t+bky2),\displaystyle=(t+b_{1}x^{2})\cdots(t+b_{k}x^{2})(t+b_{1}y^{2})\cdots(t+b_{k}y^{2}),
q(x,y,t)\displaystyle q(x,y,t) =xy(t+c1x2)(t+ckx2)(t+a1y2)(t+ak1y2),\displaystyle=xy(t+c_{1}x^{2})\cdots(t+c_{k}x^{2})(t+a_{1}y^{2})\cdots(t+a_{k-1}y^{2}),

where a1,,ak1a_{1},\dots,a_{k-1}, b1,,bkb_{1},\dots,b_{k}, and c1,,ckc_{1},\dots,c_{k} are positive constants satisfying (2.6). Fix small parameters ε>0\varepsilon>0 and α>0\alpha>0 and assign qα(x,y,t):=q(xcosαysinα,xsinα+ycosα,t)q_{\alpha}(x,y,t):=q(x\cos\alpha-y\sin\alpha,x\sin\alpha+y\cos\alpha,t) and u:=uε,α=p+εqαu:=u_{\varepsilon,\alpha}=p+\varepsilon q_{\alpha}. Since the Laplacian is rotationally invariant, we conclude that qαq_{\alpha} and uu are time-dependent hcps of degree dd. Our goal is to prove that the nodal set of u|𝕊2u|_{\mathbb{S}^{2}} is the union of two disjoint, closed Jordan curves, which implies that uu has 3 nodal domains.

Refer to caption
Refer to caption
Refer to caption
Figure 5.5. Proof of Theorem 5.6 (1/2): The nodal set of p¯\overline{p} (top), q¯\overline{q} (middle), and u¯\overline{u} when k=3k=3 and ε\varepsilon and α\alpha are sufficiently small. The two Jordan curves which form {u¯=0}\{\overline{u}=0\} are depicted in blue and orange.

Once again, we write p¯\overline{p}, q¯\overline{q}, q¯α\overline{q}_{\alpha}, and u¯\overline{u} to denote the functions pp, qq, qαq_{\alpha}, and uu expressed in the standard spherical coordinates (5.3). Thus,

(5.8) p¯(θ,ϕ)=i=1k(sinϕ+bicos2θcos2ϕ)i=1k(sinϕ+bisin2θcos2ϕ),q¯(θ,ϕ)=cosθsinθcos2ϕi=1k(sinϕ+cicos2θcos2ϕ)i=1k1(sinϕ+aisin2θcos2ϕ).\begin{split}\overline{p}(\theta,\phi)&=\prod_{i=1}^{k}(\sin\phi+b_{i}\cos^{2}\theta\cos^{2}\phi)\prod_{i=1}^{k}(\sin\phi+b_{i}\sin^{2}\theta\cos^{2}\phi),\\ \overline{q}(\theta,\phi)&=\cos\theta\sin\theta\cos^{2}\phi\prod_{i=1}^{k}(\sin\phi+c_{i}\cos^{2}\theta\cos^{2}\phi)\prod_{i=1}^{k-1}(\sin\phi+a_{i}\sin^{2}\theta\cos^{2}\phi).\end{split}

See Figure 5.5 for an illustration of the nodal domains of p¯\overline{p}, q¯\overline{q}, and u¯\overline{u} when k=3k=3. The remainder of the proof is devoted to showing that u¯\overline{u} has two disjoint, closed nodal curves, bounding one positive component and two negative components.

The nodal set of p¯\overline{p} is the union of the 2k2k graphs of the functions hip(θ)h^{p}_{i}(\theta) defined by (5.5) and the functions gip(θ):=hip(θπ/2)g^{p}_{i}(\theta):=h^{p}_{i}(\theta-\pi/2) for all 1ik1\leq i\leq k. (Here we used the elementary fact that the positive yy-axis is the rotation of the positive xx-axis by π/2\pi/2 radians.) Recall that the functions hiph^{p}_{i} are π\pi-periodic, hip(±π/2)=0h^{p}_{i}(\pm\pi/2)=0, and hip(θ)>hi+1p(θ)h^{p}_{i}(\theta)>h^{p}_{i+1}(\theta) for all 1ik11\leq i\leq k-1 and θ±π/2\theta\neq\pm\pi/2. Similarly, the functions gipg^{p}_{i} are π\pi-periodic, gip(0)=gip(π)=0g^{p}_{i}(0)=g^{p}_{i}(\pi)=0, and gip(θ)>gi+1p(θ)g^{p}_{i}(\theta)>g^{p}_{i+1}(\theta) for all 1ik11\leq i\leq k-1 and θ0,π\theta\neq 0,\pi. The sign of p¯\overline{p} changes across adjacent nodal domains.

Refer to caption
Figure 5.6. Proof of Theorem 5.6 (2/2): Sign pattern for q¯α\overline{q}_{\alpha} at singular points in the nodal set of p¯\overline{p} near θ=π/4\theta=\pi/4 when k=3k=3.

As usual, we must next identify the singular points in the nodal set of p¯\overline{p}. The points (π/2,0)(-\pi/2,0) and (π/2,0)(\pi/2,0) common to the graph of each hiph^{p}_{i} and the points (0,0)(0,0) and (π,0)(\pi,0) common to the graph of each gipg^{p}_{i} are singular points.222When d=4d=4 and k=1k=1, the nodal set for p¯\overline{p} is regular at (π/2,0)(-\pi/2,0), (0,0)(0,0), (π/2,0)(\pi/2,0), and (π,0)(\pi,0). In addition, the nodal set of p¯\overline{p} has 4k24k^{2} singular points where the graph of an hiph^{p}_{i} intersects the graph of a gjpg^{p}_{j}. For each 1i,jk1\leq i,j\leq k, there is a unique angle θi,j(0,π/2)\theta_{i,j}\in(0,\pi/2) so that the graphs of hiph^{p}_{i} and gjpg^{p}_{j} intersect precisely at

(5.9) (θi,j,hip(θi,j)),(θi,jπ,hip(θi,j)),(θi,j,hip(θi,j)),(θi,j+π,hip(θi,j)).(\theta_{i,j},h_{i}^{p}(\theta_{i,j})),\quad(\theta_{i,j}-\pi,h_{i}^{p}(\theta_{i,j})),\quad(-\theta_{i,j},h_{i}^{p}(\theta_{i,j})),\quad(-\theta_{i,j}+\pi,h_{i}^{p}(\theta_{i,j})).

The nodal set of q¯\overline{q} includes two great circles {(θ,ϕ):θ=π/2,0,π/2,π}\{(\theta,\phi)\;:\;\theta=-\pi/2,0,\pi/2,\pi\} and the 2k12k-1 graphs of functions hiqh_{i}^{q} defined by

(5.10) hiq(θ):=sin1(11+4ci2cos4θ2cicos2θ)(i=1,,k)h^{q}_{i}(\theta):=\sin^{-1}\left(\frac{1-\sqrt{1+4c_{i}^{2}\cos^{4}\theta}}{2c_{i}\cos^{2}\theta}\right)\qquad(i=1,\dots,k)

and the functions giqg^{q}_{i} defined by

(5.11) giq(θ)sin1(11+4ai2sin4θ2aisin2θ)(i=1,,k1).g_{i}^{q}(\theta)\coloneqq\sin^{-1}\left(\frac{1-\sqrt{1+4a_{i}^{2}\sin^{4}\theta}}{2a_{i}\sin^{2}\theta}\right)\qquad(i=1,\dots,k-1).

The sign of q¯\overline{q} alternates on adjacent nodal domains. In view of (2.6), we have

(5.12) h1q\displaystyle h_{1}^{q} h1ph2qh2phkqhkp,\displaystyle\geq h_{1}^{p}\geq h_{2}^{q}\geq h_{2}^{p}\geq\cdots\geq h_{k}^{q}\geq h_{k}^{p},
(5.13) g1p\displaystyle g_{1}^{p} g1qg2pg2qgk1qgkp,\displaystyle\geq g_{1}^{q}\geq g_{2}^{p}\geq g_{2}^{q}\geq\cdots\geq g_{k-1}^{q}\geq g_{k}^{p},

with strict inequalities in (5.12) unless θ=±π/2\theta=\pm\pi/2 and strict inequalities in (5.13) unless θ=0,π\theta=0,\pi. From the strict inequalities, periodicity, and evenness, one can deduce the following sign pattern for q¯\overline{q} at the singular points in the nodal set of p¯\overline{p}:

(5.14) sgn(q¯(θi,j,hip(θi,j))=sgn(q¯(θi,jπ,hip(θi,j))=(1)i+j+1,sgn(q¯(θi,j,hip(θi,j))=sgn(q¯(θi,j+π,hip(θi,j))=(1)i+j.\begin{split}\mathrm{sgn}(\overline{q}(\theta_{i,j},h_{i}^{p}(\theta_{i,j}))&=\mathrm{sgn}(\overline{q}(\theta_{i,j}-\pi,h_{i}^{p}(\theta_{i,j}))=(-1)^{i+j+1},\\ \mathrm{sgn}(\overline{q}(-\theta_{i,j},h_{i}^{p}(\theta_{i,j}))&=\mathrm{sgn}(\overline{q}(-\theta_{i,j}+\pi,h_{i}^{p}(\theta_{i,j}))=(-1)^{i+j}.\end{split}

See Figure 5.6. By continuity, (5.14) persists for q¯α\overline{q}_{\alpha} if α>0\alpha>0 is small enough. In addition,

(5.15) q¯α(0,0)=q¯α(π,0)>0,q¯α(π/2,0)=q¯α(π/2,0)<0.\overline{q}_{\alpha}(0,0)=\overline{q}_{\alpha}(\pi,0)>0,\quad\overline{q}_{\alpha}(-\pi/2,0)=\overline{q}_{\alpha}(\pi/2,0)<0.

Near the singular points of p¯\overline{p}, we again apply Lemma 5.2 / Remark 5.3 (in a rotated coordinate system, as necessary) to see that locally the nodal set of u¯\overline{u} is given by finitely many simple curves, which are contained either in the positive or negative components of {p¯0}\{\overline{p}\neq 0\}, if q¯α\overline{q}_{\alpha} is negative or positive there, respectively. Moreover, the curves approach the nodal set of p¯\overline{p} when ε0\varepsilon\downarrow 0. (Note that u¯=p¯+εq¯α=p¯ε(q¯α)\overline{u}=\overline{p}+\varepsilon\overline{q}_{\alpha}=\overline{p}-\varepsilon(-\overline{q}_{\alpha}).) Careful piecing of all of this information together, one deduces that for α>0\alpha>0 and ε>0\varepsilon>0 small, the nodal domain of u¯\overline{u} consists of two disjoint simple curves that separate 2\mathbb{R}^{2} into three connected components (one positive connected component and two negative connected components).

Indeed away from the singular points of p¯\overline{p}, the nodal set of u¯\overline{u} consists of smooth curves tending to the nodal set of p¯\overline{p} as ε0\varepsilon\downarrow 0. The key observation is then that at the singular points corresponding to the θi,j\theta_{i,j} and θi,jπ\theta_{i,j}-\pi, connectivity is gained in the horizontal direction. That is to say, at the points (θi,j,hip(θi,j))(\theta_{i,j},h_{i}^{p}(\theta_{i,j})) and (θi+1,j+1,hi+1p(θi+1,j+1))(\theta_{i+1,j+1},h_{i+1}^{p}(\theta_{i+1,j+1})), q¯α\overline{q}_{\alpha} has the same sign, which is equal to the sign of p¯\overline{p} at (θ,hip(θi,j))(\theta,h_{i}^{p}(\theta_{i,j})) for θ(θi,jδ,θi,j+δ){θi,j}\theta\in(\theta_{i,j}-\delta,\theta_{i,j}+\delta)\setminus\{\theta_{i,j}\} and δ>0\delta>0 sufficiently small. Thus for the singular points (θi,j,hip(θi,j))(\theta_{i,j},h_{i}^{p}(\theta_{i,j})), horizontally-adjacent chambers of p¯\overline{p} become connected in the nodal domain of u¯\overline{u} by Lemma 5.2. On the other hand, connectivity is gained in the vertical direction at the singular points corresponding to the θi,j,θi,j+π-\theta_{i,j},-\theta_{i,j}+\pi in similar manner. Along with the fact that all positive chambers of p¯\overline{p} meeting (0,0)(0,0) and (0,π)(0,\pi) become connected in the nodal domain of u¯\overline{u}, and all the negative chambers of p¯\overline{p} meeting (±π/2,0)(\pm\pi/2,0) become connected in the nodal domains of q¯\overline{q}, one can conclude that the positivity set of u¯\overline{u} is connected, and u¯\overline{u} has only 22 negative chambers. See Figure 5.6. ∎

6. Proof of Lemma 5.2

Towards the proof of Lemma 5.2, we start with an easy variation of [Lew77, Lemma 3], which assumed that g(x)=xkg(x)=x^{k} for some integer k2k\geq 2 and ff is real-analytic.

Lemma 6.1.

Suppose that f,g:[0,t]f,g:[0,t]\rightarrow\mathbb{R} are Lipschitz functions such that f(0)=1f(0)=1 and for some numbers k>1k>1, a>0a>0, and C>0C>0,

(6.1) |g(x)axk|Cxk+1for all x[0,1],|g(x)-ax^{k}|\leq Cx^{k+1}\quad\text{for all $x\in[0,1]$,}
(6.2) |g(x)akxk1|Cxkand|f(x)|Cfor a.e. x[0,1].|g^{\prime}(x)-akx^{k-1}|\leq Cx^{k}\quad\text{and}\quad|f^{\prime}(x)|\leq C\quad\text{for a.e.\leavevmode\nobreak\ $x\in[0,1]$.}

There exist ε0=ε0(C,a,k)>0\varepsilon_{0}=\varepsilon_{0}(C,a,k)>0 such that for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the function Fε:[0,1]F_{\varepsilon}:[0,1]\rightarrow\mathbb{R},

(6.3) Fε(x)=g(x)εf(x)for all x[0,1],F_{\varepsilon}(x)=g(x)-\varepsilon f(x)\quad\text{for all $x\in[0,1]$},

has a unique root x0(0,τ)x_{0}\in(0,\tau), where τ=min{1,12C1ak}\tau=\min\{1,\frac{1}{2}C^{-1}ak\}. Moreover, x0(ε1/(k12),ε1/(k+1))x_{0}\in(\varepsilon^{1/(k-\frac{1}{2})},\varepsilon^{1/(k+1)}), Fε(x)<0F_{\varepsilon}(x)<0 for all 0x<x00\leq x<x_{0}, Fε(x)>0F_{\varepsilon}(x)>0 for all x0<xτx_{0}<x\leq\tau, and FεF_{\varepsilon} is strictly increasing on [ε1/(k12),τ][\varepsilon^{1/(k-\frac{1}{2})},\tau].

Proof.

Let 0<εε0<10<\varepsilon\leq\varepsilon_{0}<1. Write j=k12j=k-\frac{1}{2} and =k+1\ell=k+1, so that k1<j<k<k-1<j<k<\ell. For all 0xε1/j0\leq x\leq\varepsilon^{1/j}, we have

(6.4) Fε(x)axk+Cxε(1Cx)aεk/j+Cε/j+ε(Cε1/j1)=(aε1/(2j)+Cε3/(2j)+Cε1/j1)ε12ε\begin{split}F_{\varepsilon}(x)&\leq ax^{k}+Cx^{\ell}-\varepsilon(1-Cx)\\ &\leq a\varepsilon^{k/j}+C\varepsilon^{\ell/j}+\varepsilon(C\varepsilon^{1/j}-1)=(a\varepsilon^{1/(2j)}+C\varepsilon^{3/(2j)}+C\varepsilon^{1/j}-1)\varepsilon\leq-\frac{1}{2}\varepsilon\end{split}

provided that ε0\varepsilon_{0} (hence ε\varepsilon) is sufficiently small depending only on CC, kk, and aa. Similarly, for all ε1/xmin{1,12C1a}\varepsilon^{1/\ell}\leq x\leq\min\{1,\frac{1}{2}C^{-1}a\},

(6.5) Fε(x)axkCxk+1ε(1+C)=(aCx)xkε(1+C)12aεk/ε(1+C)(12aε1/1C)εε\begin{split}F_{\varepsilon}(x)\geq ax^{k}-Cx^{k+1}-\varepsilon(1+C)&=(a-Cx)x^{k}-\varepsilon(1+C)\\ &\geq\tfrac{1}{2}a\varepsilon^{k/\ell}-\varepsilon(1+C)\geq(\tfrac{1}{2}a\varepsilon^{-1/\ell}-1-C)\varepsilon\geq\varepsilon\end{split}

provided that ε0\varepsilon_{0} is sufficiently small depending only on CC, kk, and aa. Since FF is continuous, it follows that FF has at least one root x0x_{0} in the interval (ε1/j,ε1/)(\varepsilon^{1/j},\varepsilon^{1/\ell}). Now, for any ε1/jxmin{1,12C1ak}=:τ\varepsilon^{1/j}\leq x\leq\min\{1,\frac{1}{2}C^{-1}ak\}=:\tau at which FF is differentiable,

(6.6) Fε(x)akxk1CxkCε=(akCx)xk1Cε12akε(k1)/jCε(12akε1/(2j)C)εε\begin{split}F_{\varepsilon}^{\prime}(x)\geq akx^{k-1}-Cx^{k}-C\varepsilon&=(ak-Cx)x^{k-1}-C\varepsilon\\ &\geq\tfrac{1}{2}ak\varepsilon^{(k-1)/j}-C\varepsilon\geq(\tfrac{1}{2}ak\varepsilon^{-1/(2j)}-C)\varepsilon\geq\varepsilon\end{split}

provided that ε0\varepsilon_{0} is sufficiently small depending only on CC, kk, and aa. It follows that FF is strictly increasing on [ε1/j,τ][\varepsilon^{1/j},\tau]. Choosing ε0\varepsilon_{0} to be sufficiently small guarantees that ε1/j<x0<ε1/τ\varepsilon^{1/j}<x_{0}<\varepsilon^{1/\ell}\leq\tau. In conjunction with (6.4), this shows that Fε(x)<0F_{\varepsilon}(x)<0 for all 0x<x00\leq x<x_{0} and Fε(x)>0F_{\varepsilon}(x)>0 for all x0<xτx_{0}<x\leq\tau. Therefore, x0x_{0} is the unique root of FεF_{\varepsilon} in (0,τ)(0,\tau).∎

Remark 6.2.

The proof of Lemma 6.1 shows that in fact, ε0\varepsilon_{0} and τ\tau can be chosen to depend only on CC, kk and a0>1a_{0}>1 provided that a[a01,a0]a\in[a_{0}^{-1},a_{0}].

Using Lemma 6.1, we prove Lemma 5.2.

Proof of Lemma 5.2.

It suffices to prove the result for GεFG-\varepsilon F, since the nodal set of G+εFG+\varepsilon F is the same as GεF-G-\varepsilon F, and then we may apply the result for G-G in place of GG. Since ygi(0,0)0\partial_{y}g_{i}(0,0)\neq 0, the (real-analytic) implicit function theorem implies that in a neighborhood of the origin, the zero set of gig_{i} is given by the graph of real analytic function of one variable, hi(x)h_{i}(x), so gi(x,hi(x))=0g_{i}(x,h_{i}(x))=0. Each hih_{i} satisfies hi(0)=0h_{i}(0)=0 since gig_{i} vanishes at the origin. Let us continue by proving the result for the nodal set of GεFG-\varepsilon F in {x0}\{x\geq 0\}, since similar reasoning applies to the nodal set in {x0}\{x\leq 0\}.

Note that since the gig_{i} only share a common zero at the origin, the functions hi(x)h_{i}(x) are distinct real-analytic functions. In particular, for τ\tau chosen small enough, then the hih_{i} must be ordered, and thus we may as well assume that h1(x)<h2(x)<<hm(x)h_{1}(x)<h_{2}(x)<\cdots<h_{m}(x) for 0<x<τ0<x<\tau. For each ii, we write hi(x)=k=kiai,kxkh_{i}(x)=\sum_{k=k_{i}}^{\infty}a_{i,k}x^{k}, where ai,ki0a_{i,k_{i}}\neq 0 is the first non-zero in the expansion of hih_{i}. Of course ki1k_{i}\geq 1 since hi(0)=0h_{i}(0)=0. It is straightforward to see from the fact that the nodal set of gig_{i} is given by the graph of hih_{i}, that the Taylor series of gig_{i} takes the form

(6.7) gi(x,y)\displaystyle g_{i}(x,y) =y(bi+α=(α1,α2),|α|1bαxα1yα2)+k=kibi,kxk,\displaystyle=y\left(b_{i}^{*}+\sum_{\alpha=(\alpha_{1},\alpha_{2}),\left\lvert\alpha\right\rvert\geq 1}b_{\alpha}x^{\alpha_{1}}y^{\alpha_{2}}\right)+\sum_{k=k_{i}}^{\infty}b_{i,k}x^{k},

where bi=ygi(0,0)0b_{i}^{*}=\partial_{y}g_{i}(0,0)\neq 0, and bi,ki0b_{i,k_{i}}\neq 0. In particular, the first nonzero pure xkx^{k} term in the expansion for gig_{i} has the same power as that of hih_{i} (and in fact, ai,ki=bi,ki/bia_{i,k_{i}}=-b_{i,k_{i}}/b_{i}^{*}).

We remark that since yg(0,0)0\partial_{y}g(0,0)\neq 0, gig_{i} changes sign about its nodal set and GG changes sign about the graphs hih_{i}. Note that for ε0\varepsilon_{0} and τ\tau small, the zero set of GεFG-\varepsilon F in Bτ(0)B_{\tau}(0) is contained in the positivity set of GG, since F(0,0)>0F(0,0)>0. Hence we consider a chamber

Ui0Bτ(0){(x,y):x>0,hi0(x)<y<hi0+1(x)}.U_{i_{0}}\coloneqq B_{\tau}(0)\cap\{(x,y)\;:\;x>0,\;h_{i_{0}}(x)<y<h_{i_{0}+1}(x)\}.

(the case when one of these chambers is just {y>hm}\{y>h_{m}\} or {y<h1}\{y<h_{1}\} is similar) and prove the conclusion of the lemma inside this chamber. First, we need some estimates on GG and G\nabla G, especially near the graphs of hi0h_{i_{0}} and hi0+1h_{i_{0}+1}, so define the region Ui0δU_{i_{0}}^{\delta} for δ>0\delta>0 small by

Ui0δ{(x,y)Ui0:y=thi0(x)+(1t)hi0+1(x) for some t[0,δ)(1δ,1]}.\displaystyle U_{i_{0}}^{\delta}\coloneqq\left\{(x,y)\in U_{i_{0}}\;:\;y=th_{i_{0}}(x)+(1-t)h_{i_{0}+1}(x)\text{ for some }t\in[0,\delta)\cup(1-\delta,1]\right\}.

See Figure 6.1.

Refer to caption
Figure 6.1. The region Ui0δU_{i_{0}}^{\delta} when the boundary curves hi0(x)h_{i_{0}}(x) and hi0+1(x)h_{i_{0}+1}(x) are quadratic polynomials.

Note that for i0min{ki0,ki0+1}\ell_{i_{0}}\coloneqq\min\{k_{i_{0}},k_{i_{0}+1}\} and (x,y)Ui0(x,y)\in U_{i_{0}}, we have |y||x|i0\left\lvert y\right\rvert\lesssim\left\lvert x\right\rvert^{\ell_{i_{0}}}, i.e. |y|C|x|i0\left\lvert y\right\rvert\leq C\left\lvert x\right\rvert^{\ell_{i_{0}}} for some constant C>0C>0. In addition, from (6.7), we see that for ii0i\neq i_{0} and such (x,y)(x,y),

(6.8) |gi(x,y)||x|min{ki,i0}\displaystyle\left\lvert g_{i}(x,y)\right\rvert\lesssim\left\lvert x\right\rvert^{\min\{k_{i},\ell_{i_{0}}\}}

and moreover,

(6.9) |gi0(x,y)||x|i0,|G(x,y)||x|i=1mmin{ki,i0}.\left\lvert g_{i_{0}}(x,y)\right\rvert\lesssim\left\lvert x\right\rvert^{\ell_{i_{0}}},\;\left\lvert G(x,y)\right\rvert\lesssim\left\lvert x\right\rvert^{\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\}}.

We can compute directly

yG\displaystyle\partial_{y}G =i=1mygijigj,y2G=i=1my2gijigj+i=1mjiygiygjki,jgj.\displaystyle=\sum_{i=1}^{m}\partial_{y}g_{i}\prod_{j\neq i}g_{j},\quad\partial_{y}^{2}G=\sum_{i=1}^{m}\partial_{y}^{2}g_{i}\prod_{j\neq i}g_{j}+\sum_{i=1}^{m}\sum_{j\neq i}\partial_{y}g_{i}\partial_{y}g_{j}\prod_{k\neq i,j}g_{j}.

Since G\nabla G and 2G\nabla^{2}G are bounded in a neighborhood of (0,0)(0,0), it follows that for all (x,y)Ui0(x,y)\in U_{i_{0}},

(6.10) |yG(x,y)||x|i=1mmin{ki,i0}i0and|y2G(x,y)|\displaystyle\left\lvert\partial_{y}G(x,y)\right\rvert\lesssim\left\lvert x\right\rvert^{\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\}-\ell_{i_{0}}}\quad\text{and}\quad\left\lvert\partial_{y}^{2}G(x,y)\right\rvert |x|i=1mmin{ki,i0}2i0\displaystyle\lesssim\left\lvert x\right\rvert^{\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\}-2\ell_{i_{0}}}

by (6.8) and the fact that min{ki,i0}i0\min\{k_{i},\ell_{i_{0}}\}\leq\ell_{i_{0}}. Of course, similar estimates hold for xyG\partial_{x}\partial_{y}G and x2G\partial_{x}^{2}G, so we obtain |2G(x,y)||x|i=1mmin{ki,i0}2i0\left\lvert\nabla^{2}G(x,y)\right\rvert\lesssim\left\lvert x\right\rvert^{\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\}-2\ell_{i_{0}}} in Ui0U_{i_{0}}.

Next we estimate |G(x,y)|\left\lvert\nabla G(x,y)\right\rvert from below in Ui0δU_{i_{0}}^{\delta}. From our computation of yG\partial_{y}G,

yG(x,hi0(x))\displaystyle\partial_{y}G(x,h_{i_{0}}(x)) =ygi0(x,hi0(x))ii0gi(x,hi0(x)).\displaystyle=\partial_{y}g_{i_{0}}(x,h_{i_{0}}(x))\prod_{i\neq i_{0}}g_{i}(x,h_{i_{0}}(x)).

Recalling that ygi(0,0)0\partial_{y}g_{i}(0,0)\neq 0, then for τ\tau chosen sufficiently small, we have

(6.11) |yG(x,hi0(x))|ii0|gi(x,hi0(x))|.\displaystyle\left\lvert\partial_{y}G(x,h_{i_{0}}(x))\right\rvert\gtrsim\prod_{i\neq i_{0}}\left\lvert g_{i}(x,h_{i_{0}}(x))\right\rvert.

We may estimate from below for ii0i\neq i_{0},

(6.12) |gi(x,hi0(x))|=|01dds(gi(x,shi0(x)+(1s)hi(x)))𝑑s|=|(hi0(x)hi(x))01ygi(x,shi0(x)+(1s)hi(x))ds||hi0(x)hi(x)||x|min{ki,i0},\begin{split}\left\lvert g_{i}(x,h_{i_{0}}(x))\right\rvert&=\left\lvert\int_{0}^{1}\dfrac{d}{ds}\left(g_{i}(x,sh_{i_{0}}(x)+(1-s)h_{i}(x))\right)\;ds\right\rvert\\ &=\left\lvert\left(h_{i_{0}}(x)-h_{i}(x)\right)\int_{0}^{1}\partial_{y}g_{i}(x,sh_{i_{0}}(x)+(1-s)h_{i}(x))\;ds\right\rvert\\ &\gtrsim\left\lvert h_{i_{0}}(x)-h_{i}(x)\right\rvert\gtrsim\left\lvert x\right\rvert^{\min\{k_{i},\ell_{i_{0}}\}},\end{split}

where in the second to last inequality, we again used that ygi(0,0)0\partial_{y}g_{i}(0,0)\neq 0 (and take τ\tau small), and in the last, we are simply using the definition of the kik_{i} and the expansions of the hih_{i}. In conjunction with (6.11), we see that

|yG(x,hi0(x))||x|ii0min{ki,i0},\displaystyle\left\lvert\partial_{y}G(x,h_{i_{0}}(x))\right\rvert\gtrsim\left\lvert x\right\rvert^{\sum_{i\neq i_{0}}\min\{k_{i},\ell_{i_{0}}\}},

as long as τ\tau is taken sufficiently small, and similarly for |yG(x,hi0+1(x))|\left\lvert\partial_{y}G(x,h_{i_{0}+1}(x))\right\rvert. Along with (6.10), this gives by the mean value theorem that

(6.13) |G(x,y)||yG(x,y)||x|i=1mmin{ki,i0}i0=|x|ii0mmin{ki,i0}\begin{split}\left\lvert\nabla G(x,y)\right\rvert\geq\left\lvert\partial_{y}G(x,y)\right\rvert&\gtrsim\left\lvert x\right\rvert^{\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\}-\ell_{i_{0}}}=\left\lvert x\right\rvert^{\sum_{i\neq i_{0}}^{m}\min\{k_{i},\ell_{i_{0}}\}}\end{split}

when (x,y)Ui0δ(x,y)\in U_{i_{0}}^{\delta} provided that δ\delta is chosen small enough.

Set =i=1mmin{ki,i0}\ell^{*}=\sum_{i=1}^{m}\min\{k_{i},\ell_{i_{0}}\} for convenience. By (6.9), we have that, as long as λ>0\lambda>0 is chosen sufficiently small (but not depending on ε\varepsilon), GεF0G-\varepsilon F\neq 0 in B(λε)1/(0)B_{(\lambda\varepsilon)^{1/\ell^{*}}}(0), since F1F\gtrsim 1 there while |G|λε\left\lvert G\right\rvert\lesssim\lambda\varepsilon. Now if (x,y)Ui0δ(x,y)\in U_{i_{0}}^{\delta}, and (x,y)B(λε)1/(0)(x,y)\not\in B_{(\lambda\varepsilon)^{1/\ell^{*}}}(0), then by estimate (6.13),

|yG(x,y)||x|i0(λε)1i0/,\displaystyle\left\lvert\partial_{y}G(x,y)\right\rvert\gtrsim\left\lvert x\right\rvert^{\ell^{*}-\ell_{i_{0}}}\gtrsim(\lambda\varepsilon)^{1-\ell_{i_{0}}/\ell^{*}},

since (x,y)Ui0(x,y)\in U_{i_{0}} implies |y||x|i0\left\lvert y\right\rvert\lesssim\left\lvert x\right\rvert^{\ell_{i_{0}}} with i01\ell_{i_{0}}\geq 1, so |x||(x,y)|\left\lvert x\right\rvert\simeq\left\lvert(x,y)\right\rvert there. Since FF is C1C^{1} in a neighborhood of the origin, then as long as τ\tau and ε0\varepsilon_{0} are chosen sufficiently small, then εFL(Br/2(0))|yG(x,y)|/8\varepsilon\left\lVert\nabla F\right\rVert_{L^{\infty}(B_{r/2}(0))}\leq\left\lvert\partial_{y}G(x,y)\right\rvert/8 for such points, and thus in Ui0δB(λε)1/(0)U_{i_{0}}^{\delta}\setminus B_{(\lambda\varepsilon)^{1/\ell^{*}}}(0),

(6.14) |y(GεF)||yG|ε|F|(7/8)|yG|>0.\displaystyle\left\lvert\partial_{y}(G-\varepsilon F)\right\rvert\geq\left\lvert\partial_{y}G\right\rvert-\varepsilon\left\lvert\nabla F\right\rvert\geq(7/8)\left\lvert\partial_{y}G\right\rvert>0.

In particular, the implicit function theorem implies that locally near any zero (x,y)(x,y) of GεFG-\varepsilon F in Ui0δU_{i_{0}}^{\delta}, the set {GεF=0}\{G-\varepsilon F=0\} is the graph of a C1C^{1} function over the xx-axis.

For (x,y)Ui0Ui0δ/2(x,y)\in U_{i_{0}}\setminus U_{i_{0}}^{\delta/2}, we apply Lemma 6.1. In particular, for t[δ/2,1δ/2]t\in[\delta/2,1-\delta/2], write yt(x)thi0(x)+(1t)hi0+1(x)y_{t}(x)\coloneqq th_{i_{0}}(x)+(1-t)h_{i_{0}+1}(x),

gt(x)G(x,yt(x))andft(x)F(x,yt(x)).\displaystyle g_{t}(x)\coloneqq G(x,y_{t}(x))\quad\text{and}\quad f_{t}(x)\coloneqq F(x,y_{t}(x)).

Since hi0h_{i_{0}}, hi0+1h_{i_{0}+1}, and GG are real-analytic in a neighborhood of the origin, so is gtg_{t}, and similarly, ftf_{t} is Lipschitz since FF is (with Lipschitz constant independent of tt). The fact that G(0,0)=hi0(0)=hi0+1(0)=0G(0,0)=h_{i_{0}}(0)=h_{i_{0}+1}(0)=0 and G(0,0)=0\nabla G(0,0)=0 forces the Taylor expansion of gtg_{t} at the origin to have leading term a(t)xk(t)a(t)x^{k(t)} for some k(t)2k(t)\geq 2 and a(t)>0a(t)>0, since gtg_{t} is a non-trivial, non-negative analytic function that is positive in {x>0}\{x>0\}. (Recall that Ui0U_{i_{0}} is a positive chamber for GG.) In fact, k(t)k(t)\equiv\ell^{*} for all such tt, as can be seen from the estimates (6.9) and an estimate similar to (6.12), from which one deduces that |G(x,yt(x))|tx\left\lvert G(x,y_{t}(x))\right\rvert\simeq_{t}x^{\ell^{*}} for all xx sufficiently small, which forces k(t)=k(t)=\ell^{*}. Thus, Lemma 6.1 gives for each t[δ/2,1δ/2]t\in[\delta/2,1-\delta/2], some ε0(t)\varepsilon_{0}(t) and τ(t)\tau(t) so that the function gtεftg_{t}-\varepsilon f_{t} has exactly one root in (0,τ(t))(0,\tau(t)) for all ε<ε0(t)\varepsilon<\varepsilon_{0}(t). Since tt lives in a compact interval and a(t)a(t) is continuous and strictly positive, then it attains its positive minimum and finite maximum in [δ/2,1δ/2][\delta/2,1-\delta/2]. In particular, ε(t)\varepsilon(t) and τ(t)\tau(t) can be chosen to depend only on the maximum and minimum of a(t)a(t). By Remark 6.2, we can choose ε0\varepsilon_{0} and τ\tau small independently of tt so that satisfies the conclusion of Lemma 6.1 for all t[δ/2,1δ/2]t\in[\delta/2,1-\delta/2]. Of course, ε0\varepsilon_{0} and τ\tau will depend on δ\delta, but this is a small, fixed quantity depending on FF and GG.

We are in a position to conclude the proof. Our work so far shows that at any zero z𝒩(GεF)Ui0z\in\mathcal{N}(G-\varepsilon F)\cap U_{i_{0}}, there is a neighborhood VzV_{z} of zz such that 𝒩(GεF)Vz\mathcal{N}(G-\varepsilon F)\cap V_{z} coincides with a curve passing through zz: in Ui0δU_{i_{0}}^{\delta} this is by the implicit function theorem and in Ui0Ui0δ/2U_{i_{0}}\setminus U_{i_{0}}^{\delta/2} from the fact that the (unique) zeros of GεFG-\varepsilon F along the trajectories x(x,yt(x))Ui0x\rightarrow(x,y_{t}(x))\in U_{i_{0}} indexed by t[δ/2,1δ/2]t\in[\delta/2,1-\delta/2] form a continuous curve in tt. Let γ\gamma denote some connected component of 𝒩(GεF)Ui0¯\mathcal{N}(G-\varepsilon F)\cap\overline{U_{i_{0}}}. If γ\gamma meets Ui0δU_{i_{0}}^{\delta}, define

xmin=inf{x>0:(x,y)Ui0δγ for some y},\displaystyle x_{\mathrm{min}}=\inf\{x>0\;:\;(x,y)\in U_{i_{0}}^{\delta}\cap\gamma\text{ for some }y\},

which exists since GεF0G-\varepsilon F\neq 0 in B(λε)1/(0)B_{(\lambda\varepsilon)^{1/\ell^{*}}}(0). Choose a minimizing sequence xkxminx_{k}\rightarrow x_{\mathrm{min}} with corresponding points yky_{k} such that (xk,yk)Ui0δγ(x_{k},y_{k})\in U_{i_{0}}^{\delta}\cap\gamma, and note that we may assume (up to a subsequence) that (xk,yk)(xmin,ymin)(x_{k},y_{k})\rightarrow(x_{\mathrm{min}},y_{\mathrm{min}}) for some yminy_{\mathrm{min}}. Note that (xmin,ymin)(x_{\min},y_{\min}) is a zero of GεFG-\varepsilon F, and moreover, ymin=yt(xmin)y_{\min}=y_{t}(x_{\min}) for either t=δt=\delta or t=1δt=1-\delta. Indeed GεFG-\varepsilon F is nonzero on the graphs of hi0h_{i_{0}} and hi0+1h_{i_{0}+1}, and (6.14) says that if (xmin,ymin)int(Ui0δ)(x_{\min},y_{\min})\in\mathrm{int}(U_{i_{0}}^{\delta}), then locally near this point, 𝒩(GεF)\mathcal{N}(G-\varepsilon F) coincides with the graph of a C1C^{1} function over the xx-axis, which would contradict the definition of xminx_{\min}. By the remarks above, there is a neighborhood VV of (xmin,ymin)(x_{\min},y_{\min}) in which 𝒩(GεF)\mathcal{N}{(G-\varepsilon F)} coincides with a curve passing through (xmin,ymin)(x_{\min},y_{\min}), which therefore must coincide with γV\gamma\cap V. Hence (xmin,ymin)γ(Ui0Ui0δ/2)(x_{\min},y_{\min})\in\gamma\,\cap(U_{i_{0}}\setminus U_{i_{0}}^{\delta/2}). We have proved that every connected component of 𝒩(GεF)Ui0¯\mathcal{N}(G-\varepsilon F)\cap\overline{U_{i_{0}}} meets Ui0Ui0δ/2U_{i_{0}}\setminus U_{i_{0}}^{\delta/2}.

Recall that 𝒩(GεF)(Ui0Ui0δ/2)\mathcal{N}(G-\varepsilon F)\cap(U_{i_{0}}\setminus U_{i_{0}}^{\delta/2}) has exactly one connected component, given by the curve of zeros of GεFG-\varepsilon F along the trajectories x(x,yt(x))x\rightarrow(x,y_{t}(x)). Since any connected component of 𝒩(GεF)Ui0¯\mathcal{N}(G-\varepsilon F)\cap\overline{U_{i_{0}}} meets this region, it follows that 𝒩(GεF)Ui0¯\mathcal{N}(G-\varepsilon F)\cap\overline{U_{i_{0}}} has exactly one connected component and we know that this component is a simple curve. We leave it to the reader to verify that the curve tends to {G=0}\{G=0\} near the origin as ε0\varepsilon\downarrow 0. ∎

References

  • [ABR01] Sheldon Axler, Paul Bourdon, and Wade Ramey, Harmonic function theory, second ed., Graduate Texts in Mathematics, vol. 137, Springer-Verlag, New York, 2001. MR 1805196
  • [AM19] Jonas Azzam and Mihalis Mourgoglou, Tangent measures of elliptic measure and applications, Anal. PDE 12 (2019), no. 8, 1891–1941. MR 4023971
  • [BET17] Matthew Badger, Max Engelstein, and Tatiana Toro, Structure of sets which are well approximated by zero sets of harmonic polynomials, Anal. PDE 10 (2017), no. 6, 1455–1495. MR 3678494
  • [BET22] Matthew Badger, Max Engelstein, and Tatiana Toro, Slowly vanishing mean oscillations: non-uniqueness of blow-ups in a two-phase free boundary problem, preprint, arXiv:2210.17531, to appear in Vietnam J. Math., 2022.
  • [BG23] Matthew Badger and Alyssa Genschaw, Hausdorff dimension of caloric measure, preprint, arXiv:2108.12340v2, to appear in Amer. J. Math., 2023.
  • [BH16] P. Bérard and B. Helffer, A. Stern’s analysis of the nodal sets of some families of spherical harmonics revisited, Monatsh. Math. 180 (2016), no. 3, 435–468. MR 3513215
  • [CH53] R. Courant and D. Hilbert, Methods of mathematical physics. Vol. I, Interscience Publishers, Inc., New York, 1953. MR 65391
  • [Che98] Xu-Yan Chen, A strong unique continuation theorem for parabolic equations, Math. Ann. 311 (1998), no. 4, 603–630. MR 1637972
  • [CLMM20] S. Chanillo, A. Logunov, E. Malinnikova, and D. Mangoubi, Local version of Courant’s nodal domain theorem, preprint, arXiv:2008.00677, to appear in J. Differential Geom., 2020.
  • [CM21] Tobias Holck Colding and William P. Minicozzi, II, Optimal bounds for ancient caloric functions, Duke Math. J. 170 (2021), no. 18, 4171–4182. MR 4348235
  • [Doo01] Joseph L. Doob, Classical potential theory and its probabilistic counterpart, Classics in Mathematics, Springer-Verlag, Berlin, 2001, Reprint of the 1984 edition. MR 1814344
  • [EJN07] Alexandre Eremenko, Dmitry Jakobson, and Nikolai Nadirashvili, On nodal sets and nodal domains on S2S^{2} and 2\mathbb{R}^{2}, Ann. Inst. Fourier (Grenoble) 57 (2007), no. 7, 2345–2360. MR 2394544
  • [Eva10] Lawrence C. Evans, Partial differential equations, second ed., Graduate Studies in Mathematics, vol. 19, American Mathematical Society, Providence, RI, 2010. MR 2597943
  • [KPT09] C. Kenig, D. Preiss, and T. Toro, Boundary structure and size in terms of interior and exterior harmonic measures in higher dimensions, J. Amer. Math. Soc. 22 (2009), no. 3, 771–796. MR 2505300
  • [Lew77] Hans Lewy, On the minimum number of domains in which the nodal lines of spherical harmonics divide the sphere, Comm. Partial Differential Equations 2 (1977), no. 12, 1233–1244. MR 477199
  • [Ley96] Josef Leydold, On the number of nodal domains of spherical harmonics, Topology 35 (1996), no. 2, 301–321. MR 1380499
  • [Log18a] Alexander Logunov, Nodal sets of Laplace eigenfunctions: polynomial upper estimates of the Hausdorff measure, Ann. of Math. (2) 187 (2018), no. 1, 221–239. MR 3739231
  • [Log18b] Alexander Logunov, Nodal sets of Laplace eigenfunctions: proof of Nadirashvili’s conjecture and of the lower bound in Yau’s conjecture, Ann. of Math. (2) 187 (2018), no. 1, 241–262. MR 3739232
  • [LTY15] Hairong Liu, Long Tian, and Xiaoping Yang, The minimum numbers of nodal domains of spherical Grushin-harmonics, Nonlinear Anal. 126 (2015), 143–152. MR 3388875
  • [Mil64] J. Milnor, On the Betti numbers of real varieties, Proc. Amer. Math. Soc. 15 (1964), 275–280. MR 161339
  • [MP21] Mihalis Mourgoglou and Carmelo Puliatti, Blow-ups of caloric measure in time varying domains and applications to two-phase problems, J. Math. Pures Appl. (9) 152 (2021), 1–68. MR 4280831
  • [NS09] Fedor Nazarov and Mikhail Sodin, On the number of nodal domains of random spherical harmonics, Amer. J. Math. 131 (2009), no. 5, 1337–1357. MR 2555843
  • [PKS21] Vassilis G. Papanicolaou, Eva Kallitsi, and George Smyrlis, Entire solutions for the heat equation, Electron. J. Differential Equations (2021), Paper No. 44, 25. MR 4269120
  • [RW59] P. C. Rosenbloom and D. V. Widder, Expansions in terms of heat polynomials and associated functions, Trans. Amer. Math. Soc. 92 (1959), 220–266. MR 107118
  • [Sta07] Richard P. Stanley, An introduction to hyperplane arrangements, Geometric combinatorics, IAS/Park City Math. Ser., vol. 13, Amer. Math. Soc., Providence, RI, 2007, pp. 389–496. MR 2383131
  • [Sze75] Gábor Szegő, Orthogonal polynomials, fourth ed., American Mathematical Society Colloquium Publications, Vol. XXIII, American Mathematical Society, Providence, R.I., 1975. MR 0372517
  • [Szu79] Andrzej Szulkin, An example concerning the topological character of the zero-set of a harmonic function, Math. Scand. 43 (1978/79), no. 1, 60–62. MR 523825
  • [Wat12] Neil A. Watson, Introduction to heat potential theory, Mathematical Surveys and Monographs, vol. 182, American Mathematical Society, Providence, RI, 2012. MR 2907452