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

Necessity of a logarithmic estimate for hypoellipticity of some degenerately elliptic operators

Timur Akhunov Department of Mathematics and Computer Science
Wabash College
301 W Wabash Ave, Crawfordsville, IN 47933, USA
 and  Lyudmila Korobenko Mathematics Department
Reed College
3203 Southeast Woodstock Boulevard
Portland, OR 97202-8199, USA
Abstract.

This paper extends a class of degenerate elliptic operators for which hypoellipticity requires more than a logarithmic gain of derivatives of a solution in every direction. Work of Hoshiro and Morimoto in late 80s characterized a necessity of a super-logarithmic gain of derivatives for hypoellipticity of a sum of a degenerate operator and some non-degenerate operators like Laplacian. The operators we consider are similar, but more general. We examine operators of the form L1(x)+g(x)L2(y)L_{1}(x)+g(x)L_{2}(y), where L1(x)L_{1}(x) is one-dimensional and g(x)g(x) may itself vanish. The argument of the paper is based on spectral projections, analysis of a spectral differential equation, and interpolation between standard and operator-adapted derivatives. Unlike prior results in the literature, our methods do not require explicit analytic construction in the non-degenerate direction. In fact, our result allows non-analytic and even non-smooth coefficients for the non-degenerate part.

Key words and phrases:
hypoellipticity, infinite vanishing, loss of derivatives
1991 Mathematics Subject Classification:
35H10, 35H20, 35S05, 35G05, 35B65, 35A18
The authors thank Cristian Rios, who introduced us to the problem of hypoellipticity and worked with us on [AKR19] that motivated results of this paper. We also thank the anonymous referee for numerous insightful suggestions

1. Introduction

Consider a degenerate elliptic operator

(1) L(x,y)=L1(x)+g(x)L2(y)\displaystyle L(x,y)=L_{1}(x)+g(x)L_{2}(y)

where

L1=j,k=1najk(x)xjxk+j=1naj(x)xj+a0(x)\displaystyle L_{1}=-\sum_{j,k=1}^{n}a_{jk}(x)\partial_{x_{j}}\partial_{x_{k}}+\sum_{j=1}^{n}a_{j}(x)\partial_{x_{j}}+a_{0}(x)

with xnx\in\operatorname{\mathbb{R}}^{n}, ajka_{jk}, aja_{j} smooth real valued coefficients with ajka_{jk} a non-negative matrix. Similarly, let L2L_{2} be denoted as

(2) L2=j,k=1mbjk(y)yjyk+j=1mbj(y)yj+b0(y)\displaystyle L_{2}=-\sum_{j,k=1}^{m}b_{jk}(y)\partial_{y_{j}}\partial_{y_{k}}+\sum_{j=1}^{m}b_{j}(y)\partial_{y_{j}}+b_{0}(y)

where ymy\in\operatorname{\mathbb{R}}^{m}, bjkb_{jk}, bjb_{j} are smooth real functions and bjkb_{jk} is a non-negative matrix:

(3) j,k=1mbj,k(y)ηjηk0\displaystyle\sum_{j,k=1}^{m}b_{j,k}(y)\eta_{j}\eta_{k}\geq 0

In this paper we investigate conditions under which local hypoellipticity of LL, or smoothness of solutions of Lu=fLu=f whenever ff is locally smooth (see Definition 2), requires the following superlogarithmic estimate introduced by Morimoto [Mor87]:

Definition 1 (Superlogarithmic estimate).

We say that an operator LL satisfies a superlogarithmic estimate near y0Rmy_{0}\in R^{m} if for any small enough compact set KRmK\Subset R^{m} containing y0y_{0} and for all ε>0\varepsilon>0, there exists a family of constants Cε,KC_{\varepsilon,K}, such that the following estimate holds.

(4) logξu^(ξ)2εRe(Lu,u)+Cε,Ku2,uC0(K).||\log\!\left<\xi\right>\hat{u}(\xi)||^{2}\leq\varepsilon\,\mathrm{Re}(Lu,u)+C_{\varepsilon,K}||u||^{2},\quad u\in C_{0}^{\infty}(K).

where

(5) ξ:=e2+|ξ|2\displaystyle\!\left<\xi\right>:=\sqrt{e^{2}+|\xi|^{2}}

This estimate is satisfied by certain operators L2L_{2} with degeneracy in (3) so that the vanishing is faster than any polynomial in yy in some directions. For example,

(6) L2=y12+e1|y1|1κy22\displaystyle L_{2}=\partial_{y_{1}}^{2}+e^{-\frac{1}{|y_{1}|^{1-\kappa}}}\partial_{y_{2}}^{2}

considered by [KS84] satisfies (4) for κ>0\kappa>0, see [Chr01] Lemma 5.2. The Hörmander bracket condition for L2L_{2} corresponds to a polynomial degeneracy for L2L_{2} and leads to subelliptic estimate, defined as

(7) ξδu^(ξ)2C(δ,K)(Re(L2u,u)+u2), for uC0(K).\displaystyle||\!\left<\xi\right>^{\delta}\hat{u}(\xi)||^{2}\leq C(\delta,K)\left(\mathrm{Re}(L_{2}u,u)+||u||^{2}\right),\text{ for }\quad u\in C_{0}^{\infty}(K).

for some δ>0\delta>0 and ξ\!\left<\xi\right> defined in (5) as before. Note that (7) implies (4). Superlogarithmic estimate (4) limits the degree of vanishing that an operator L2L_{2} can exhibit. If a stronger vanishing is considered, e.g. (6) for κ0\kappa\leq 0, then (4) fails and a weaker gain function should replace the logarithm in (4). See [Chr01] for a further discussion of degeneracy and gain function.

This is a companion paper to [AKR19], where we proved hypoellipticity of LL from (1) under the assumption that operators L1L_{1} and L2L_{2} satisfy Definition 1, see [AKR19, Theorem 1] and [Mor87]. It was also demonstrated that, at least for particular classes of operators, the superlogarithmic estimate (4) for L1L_{1} is necessary, see [AKR19, Example 1]. Here we address in greater generality the question of necessity of these superlogarithmic estimates.

We first set up relevant notation. In what follows, for any subsets EE, FF in n\operatorname{\mathbb{R}}^{n} we write EFE\Subset F to indicate that the closure of EE, E¯\overline{E}, is compact and E¯F\overline{E}\subset F.

For completeness, we first define local smooth hypoellipticity here.

Definition 2.

The operator L(x,y)L(x,y) is locally (CC^{\infty}) hypoelliptic at (x0,y0)(x_{0},y_{0}) if there exists a neighborhood Ω\Omega of (x0,y0)(x_{0},y_{0}), such that for every f(x,y)f(x,y) smooth in Ω\Omega and a distributional solution uu satisfying Lu=fLu=f in the weak sense in Ω\Omega, then uC(Ω)u\in C^{\infty}(\Omega^{\prime}) for all neighbohoods Ω\Omega^{\prime} of (x0,y0)(x_{0},y_{0}) such that ΩΩ\Omega^{\prime}\Subset\Omega.

In particular, the region of smoothness of a weak solution depends on the operator LL and the region Ω\Omega, but not on the specific function ff, so long as ff is smooth. Our method allows us to work with a limited regularity form of hypoellipticity as well. We model this form of hypoellipticity on the Definition 2.

Definition 3.

We say that LL is (locally) HsH^{s}-hypoelliptic at (x0,y0)(x_{0},y_{0}), if there exists a neighborhood Ω\Omega of (x0,y0)(x_{0},y_{0}), such that for every fHs(Ω)f\in H^{s}(\Omega) and uL2(Ω)u\in L^{2}(\Omega) satisfying Lu=fLu=f in the weak sense in Ω\Omega, then uHs(Ω)u\in H^{s}(\Omega^{\prime}) for all neighbohoods Ω\Omega^{\prime} of (x0,y0)(x_{0},y_{0}) such that ΩΩ\Omega^{\prime}\Subset\Omega.

Morimoto in [Mor87], motivated by the results in [KS84], provided a non-probabilistic proof that in order for a symmetric operator L(x,y)=x2+L2(y)L(x,y)=-\partial_{x}^{2}+L_{2}(y), x,ymx\in\operatorname{\mathbb{R}},\ y\in\operatorname{\mathbb{R}}^{m} to be hypoelliptic, L2L_{2} has to satisfy Definition 1. Independently at the same time, Hoshiro [Hos87] obtained a similar result as a corollary of an a priori estimate for a general self-adjoint partial differential operator. Our main theorem is a generalization of the Morimoto’s and Hoshiro’s necessity estimates to the operator of the form L=L1(x)+g(x)L2(y)L=L_{1}(x)+g(x)L_{2}(y), where the function gg is allowed to vanish, allowing the operator LL itself to be more degenerate than L2L_{2} and hence not satisfy (4) at some (x0,y0)(x_{0},y_{0}).

Theorem 4.

Consider a possibly degenerate elliptic operator

(8) L=a2(x)x2+a1(x)x+a0(x)+g(x)L2(y,y),\displaystyle L=-a_{2}(x)\partial_{x}^{2}+a_{1}(x)\partial_{x}+a_{0}(x)+g(x)L_{2}(y,\partial_{y}),

where L2L_{2}, defined in (2), is a self-adjoint operator, i.e.

(9) (L2u,v)=(u,L2v), for uC0\displaystyle(L_{2}u,v)=(u,L_{2}v)\text{, for }u\in C^{\infty}_{0}

Assume further a2(x0)>0a_{2}(x_{0})>0, g(x)0g(x)\geq 0 and the coefficients gg and aia_{i} for i=0,1,2i=0,1,2 are smooth.

Suppose that LL from (8) is HsH^{s}-hypoelliptic near (x0,y0)(x_{0},y_{0}) for s>12s>\frac{1}{2}. Then L2L_{2} must satisfy (4) near y0y_{0}.

Remark 5.

Proof of Theorem 4 for 12<s1\frac{1}{2}<s\leq 1 is valid with a2(x)Cx1a_{2}(x)\in C^{1}_{x} and other coefficients merely continuous.

Corollary 6.

Suppose LL from (8) satisfies all of the assumptions of Theorem 4, with the exception of HsH^{s} hypoellipticity. We replace that assumption with the CC^{\infty} hypoellipticity of LL at (x0,y0)(x_{0},y_{0}). Then L2L_{2} satisfies (4) near y0y_{0}. Note that to establish CC^{\infty} hypoellipticity, more smoothness of the coefficients than continuity in Theorem 4 may be needed.

Remark 7.

Note, that the focus in Theorem 4 is on the implications of hypoellipticity. In particular, if g(x)0g(x)\equiv 0, the operator LL fails to be hypoelliptic and the Theorem 4 does not apply.

When g(x)g(x) has an isolated zero at x0x_{0}, our earlier work with Cristian Rios in [AKR19] proved CC^{\infty} hypoellipticity of LL if it satisfies (4) extending the non-degenerate case of g(x)g(x) studied by Morimoto in [Mor87]. Thus, in the context of an isolated zero (or non-vanishing) gg, Theorem 4 is sharp for CC^{\infty} hypoellipticity of (8).

Finally, [MM97a] proved hypoellipticity of LL from (8), when L2=y2L_{2}=\partial_{y}^{2} assuming only that the stochastic average

gI:=Ig(x)𝑑x>0, for any interval I,\displaystyle g_{I}:=\int_{I}g(x)dx>0\text{, for any interval }I\subset\operatorname{\mathbb{R}},

including the case of non-isolated zeros. Morimoto and Morioka further connected quantitative estimates for gIg_{I} with the logarithmic estimate (4) for some operators in [MM97b]. Thus Theorem 4 is likely non-trivial in some cases of g(x)g(x) with non-isolated zeros.

The strength of Theorem 4 lies both in the fact that we replace x2-\partial^{2}_{x} by a general one-dimensional elliptic operator, and, more importantly, allow for broader conditions on the degeneracy of the non-negative weight function g(x)g(x) in (8). This is in line with the results by Fediĭ [Fed71], who considered L=x2+g(x)y2L=\partial^{2}_{x}+g(x)\partial^{2}_{y}, and Kohn [Koh98] who generalized Fediĭ’s result to L=L1(x)+g(x)L2(y)L=L_{1}(x)+g(x)L_{2}(y) where L1L_{1} and L2L_{2} are subelliptic in their variables. Note that the subelliptic estimate (7) is stronger that (4), so Theorem 4 together with [AKR19, Theorem 1] characterizes hypoellipticity in the case of g(x)g(x) with isolated vanishing. Furthermore, Theorem 4 extends a necessary condition for hypoellipticity to a wide class of operators, many of them not previously considered in the literature.

Theorem 4 can be generalized in several ways. First, one can ask if the result still holds for a more general operator L1(x)L_{1}(x). We believe that the ultimate goal would be the following conjecture.

Conjecture 8.

Suppose that the operator LL from (1) is symmetric, i.e. satisfies (9) and g(x)0g(x)\geq 0. The operator LL is locally CC^{\infty}-hypoelliptic, if and only if both L1L_{1} and L2L_{2} satisfy (4).

As discussed in Remark 7, results in [AKR19] imply the sufficiency part of this conjecture in the case of g(x)g(x) that has at most isolated zeros. Theorem 4 confirms the necessity direction in the case that L1L_{1} is one dimensional and elliptic.

The second direction to generalize Theorem 4 is to allow for a more general dependence on the variables, ultimately considering operators of the form L1(x,y)+L2(x,y)L_{1}(x,y)+L_{2}(x,y) with either L1L_{1} or L2L_{2} being degenerate elliptic. It is a nontrivial question, what happens to hypoellipticity when two (or more) operators are added. For example, Fediĭ’s operator x2+g(x)y2\partial^{2}_{x}+g(x)\partial^{2}_{y} is hypoelliptic whenever gg is positive away from zero [Fed71], while a very similar three-dimensional operator x2+y2+g(x)z2\partial^{2}_{x}+\partial^{2}_{y}+g(x)\partial^{2}_{z} considered by Kusuoka and Strook [KS84] is hypoelliptic if and only if gg vanishes slower than e1/x2e^{-1/x^{2}}. See also [Chr01], [Mor87], [Mor92], and references within for more discussion of these “intertwined cases”. It is possible that methods used in our paper may also extend to some complex valued elliptic operators satisfying Hörmander bracket condition [Koh05], similar to the note of Christ [Chr05].

To summarize, Theorem 4 together with [AKR19, Theorem 1], contribute to our understanding of how regularity properties of operators in lower dimensions play into the regularity of their sum in higher dimensions. On the technical level, prior works of [Hos87], [Mor86] and [Mor87] used an explicit construction of an analytic function of xx motivated by techniques of analytic hypoellipticity [M7́8] or complex analysis. In contrast, the argument of this paper is grounded in analysis of ODEs, which allows for non-analytic and even nonsmooth coefficients. Additionally, we have adapted Littlewood-Paley projections to the degenerate operator to interpolate between standard derivatives and functional calculus, what we hope improves exposition.

This paper is organized as follows. The proof of Theorem 4 is split into the next three sections. Section 2 uses an initial value problem for an ODE and spectral projections to construct a special solution to L~u=0\tilde{L}u=0. In Section 3 we use the Closed Graph Theorem and a version of Sobolev Embedding to obtain qualitative estimates for the solution of a hypoelliptic operator. In section 4 we apply the results of the previous two section to spectral projections of any L2L^{2} function. Section 5 concludes the proof by using interpolation to combine the estimates for the spectral projections into (4). Finally, the Appendix A, contains some technical results about spectral projections used throughout the paper.

2. Spectral solutions via an ODE

We begin the proof of Theorem 4 with a simple change of operator. It is convenient to reduce (8) to a case of a constant coefficient: a21a_{2}\equiv 1 first.

Lemma 9.

Consider the operator P~=1a2(x)L\tilde{P}=\frac{1}{a_{2}(x)}L in a neighborhood small enough, where a2(x)a_{2}(x) does not vanish. I.e.

P~=x2+a1a2(x)x+a0a2(x)+ga2(x)L2(y,y).\displaystyle\tilde{P}=-\partial_{x}^{2}+\frac{a_{1}}{a_{2}}(x)\partial_{x}+\frac{a_{0}}{a_{2}}(x)+\frac{g}{a_{2}}(x)L_{2}(y,\partial_{y}).

Then P~\tilde{P} is hypoelliptic if and only if LL from (8) is hypoelliptic. By relabeling a1a_{1}, a0a_{0} and g(x)g(x) with analogues in P~\tilde{P}, we can assume a2(x)=1a_{2}(x)=1 in LL without loss of generality.

Proof.

Suppose LL is hypoelliptic. Note that P~\tilde{P} has the same structure as LL, except for a coefficient of x2\partial_{x}^{2}.

If F=P~uF=\tilde{P}u is smooth near (x0,y0)(x_{0},y_{0}), so is f(x,y)=a2(x)F(x,y)=Luf(x,y)=a_{2}(x)F(x,y)=Lu. By hypoellipticity of LL, uu must be smooth near (x0,y0)(x_{0},y_{0}). Thus P~\tilde{P} is hypoelliptic. A similar argument works for the converse. ∎

The strategy of the proof of Theorem 4 is to convert the problem Lu=0Lu=0 into an ODE using spectral projections, formally discussed in appendix A. Informally, the operator L2L_{2} is replaced by a spectral parameter λ\lambda. The following ODE becomes relevant to our analysis in light of Lemma 9.

(10) {x2v(x,λ)+a1(x)xv(x,λ)+[a0(x)+g(x)λ]v(x,λ)=0v(x0,λ)=1,v˙(x0,λ)=0\displaystyle\begin{cases}-\partial_{x}^{2}v(x,\lambda)+a_{1}(x)\partial_{x}v(x,\lambda)+[a_{0}(x)+g(x)\lambda]v(x,\lambda)=0\\ v(x_{0},\lambda)=1,\,\dot{v}(x_{0},\lambda)=0\end{cases}

We analyze this ODE first before returning to the original operator LL from (8) in Lemma 13.

Proposition 10 (Proposition 1.2.4, [Hör97]).

Let x0x_{0}\in\operatorname{\mathbb{R}}, II a connected neighborhood of x0x_{0} and let Φ:IRm×m\Phi:I\to R^{m\times m} be a continuous function with values in m×mm\times m matrices. Then the initial value problem

(11) {y(x)=Φ(x)y for |xx0|rx0y(x0)=y0\displaystyle\begin{cases}y^{\prime}(x)=\Phi(x)y&\,\,\text{ for }|x-x_{0}|\leq r_{x_{0}}\\ y(x_{0})=y_{0}&\end{cases}

has a unique solution for all y0my_{0}\in\operatorname{\mathbb{R}}^{m}. Moreover, if ΦLIM\lVert\Phi\rVert_{L^{\infty}_{I}}\leq M for some M>1M>1, then |y(x)||y0|eM|xx0||y(x)|\leq|y_{0}|e^{M|x-x_{0}|} for all xIx\in I.

We need estimates for derivatives of the solutions to (11), primarily for y(x)y^{\prime}(x).

Lemma 11.

If in addition Φ(j)(x)M\lVert\Phi^{(j)}(x)\rVert\leq M for 0jk10\leq j\leq k-1, k1k\geq 1, then

|xky(x)|C(k)(1+M)keM|xx0||y0|\displaystyle|\partial_{x}^{k}y(x)|\leq C(k)(1+M)^{k}e^{M|x-x_{0}|}|y_{0}|

Moreover, estimates for the derivatives of Φ\Phi are not needed for k=1k=1.

Proof.

Note, that

|y(x)||Φ(x)||y(x)|MeM|xx0||y0|\displaystyle|y^{\prime}(x)|\leq\lvert\Phi(x)\rvert\lvert y(x)\rvert\leq Me^{M|x-x_{0}|}|y_{0}|

Differentiating (11) we get

y′′=Φy+Φy\displaystyle y^{\prime\prime}=\Phi y^{\prime}+\Phi^{\prime}y

Hence

|y′′(x)|(MeM|xx0|+MMeM|xx0|)|y0|\displaystyle\lvert y^{\prime\prime}(x)\rvert\leq(Me^{M|x-x_{0}|}+M\cdot Me^{M|x-x_{0}|})|y_{0}|

Proceeding inductively, we get the estimates for higher derivatives of yy up to kk. ∎

We now use these results to solve (10).

Lemma 12.

Let x0x_{0} be given. Then (10) has a unique solution v(x,λ)v(x,\lambda) of (10) on [x01,x0+1][x_{0}-1,x_{0}+1]. Furthermore, the following estimate holds:

|xkv(x,λ)|C(k,x0)λmax{0,k12}exp(Cx0λ12|xx0|)\displaystyle|\partial_{x}^{k}v(x,\lambda)|\leq C({k,x_{0}})\!\left<\lambda\right>^{\max\{0,\frac{k-1}{2}\}}\exp(C_{x_{0}}\!\left<\lambda\right>^{\frac{1}{2}}|x-x_{0}|)

where Cx0C_{x_{0}} depends on x0x_{0} through the coefficients of (10). Moreover, the constant for k=0k=0 and k=1k=1, C(0,x0)=1C({0,x_{0}})=1. For k=0,1,2k=0,1,2 no derivatives of the coefficients a0a_{0}, a1a_{1} and gg are needed. Note, that we use (5) for λ\!\left<\lambda\right>.

Proof.

To apply Proposition 10 to (10), define a(x)=[a0(x)+g(x)λ]a(x)=[a_{0}(x)+g(x)\lambda] and y(x)=(vv)y(x)=\begin{pmatrix}v\\ v^{\prime}\end{pmatrix}. Then y(x0)=(10)y(x_{0})=\begin{pmatrix}1\\ 0\end{pmatrix} and y(x)=Φ(x)y(x)y^{\prime}(x)=\Phi(x)y(x) with

(12) Φ(x)=(01a(x)a1(x))\displaystyle\Phi(x)=\begin{pmatrix}0&1\\ a(x)&a_{1}(x)\end{pmatrix}

Looking at the eigenvalues of Φ(x)\Phi(x), we obtain

Φ(x)C(|a1(x)|+|a(x)|),\lVert\Phi(x)\rVert\leq C(|a_{1}(x)|+\sqrt{|a(x)|}),

where CC is a universal constant. Therefore, from the definition of a(x)a(x), we estimate

(13) Φ()LlocC(gLlocλ+a0Lloc+a1Lloc)C(x0)λ12\displaystyle\lVert\Phi(\cdot)\rVert_{L^{\infty}_{loc}}\leq C(\sqrt{\lVert g\rVert_{L^{\infty}_{loc}}\lambda+\lVert a_{0}\rVert_{L^{\infty}_{loc}}}+\lVert a_{1}\rVert_{L^{\infty}_{loc}})\leq C(x_{0})\!\left<\lambda\right>^{\frac{1}{2}}

Proposition 10 completes the proof for k=0k=0 and k=1k=1.

For higher derivatives of vv, we observe that differentiating (12), will give an estimate similar to (13) for xjΦ()Lloc\lVert\partial_{x}^{j}\Phi(\cdot)\rVert_{L^{\infty}_{loc}} . That is, xa(x)\partial_{x}a(x) and xa1(x)\partial_{x}a_{1}(x) have similar to dependence on λ\lambda, i.e.

xjΦ()LlocC(j,x0)λ12.\lVert\partial_{x}^{j}\Phi(\cdot)\rVert_{L^{\infty}_{loc}}\leq C(j,x_{0})\!\left<\lambda\right>^{\frac{1}{2}}.

As control of v(k)(x)v^{(k)}(x) requires estimates of y(k1)(x)y^{(k-1)}(x) for k1k\geq 1, Lemma 11 completes the proof. ∎

We are now ready to construct solutions to (8) via Proposition 23 and other results from the Appendix. We define an operator v(x,B)v(x,B) on a subset of L2L^{2} via v(x,λ)v(x,\lambda). The following Lemma summarizes the connection between the spectral construction and (8). We will later extend the domain of the vv operator, by pre-applying Littlewood-Paley projections.

Lemma 13.

Let u(y)u(y) be such that, eBεuL2e^{\sqrt{B}\varepsilon}u\in L^{2} for some ε>0\varepsilon>0. Define ww by

(14) w(x,y)=v(x,B)u(y)=1v(x,λ)𝑑Eλu(y).\displaystyle w(x,y)=v(x,B)u(y)=\int_{1}^{\infty}v(x,\lambda)dE_{\lambda}u(y).

where v(x,λ)v(x,\lambda) is the solution of (10). Let 0<ε<ε0<\varepsilon^{\prime}<\varepsilon. Then for |xx0|ε2Cx0|x-x_{0}|\leq\frac{\varepsilon^{\prime}}{2C_{x_{0}}} for Cx0C_{x_{0}} from Lemma 12, ww is well-defined. Likewise, Lw(x,y)L2Lw(x,y)\in L^{2} for LL from (8). Moreover, using Lemma 9 to replace a2a_{2} with 11, we establish that

Lw=0.Lw=0.
Proof.

By Proposition 23

Bw=1λv(x,λ)𝑑Eλu(y)Bw=\int_{1}^{\infty}\lambda v(x,\lambda)dE_{\lambda}u(y)

Hence by Lemma 12 for |xx0|ε2Cx0|x-x_{0}|\leq\frac{\varepsilon^{\prime}}{2C_{x_{0}}}

BwL22\displaystyle\lVert Bw\rVert_{L^{2}}^{2} 1λ2(exp(Cx0λ12|xx0|))2dEλu(y)2\displaystyle\leq\int_{1}^{\infty}\lambda^{2}(\exp(C_{x_{0}}\!\left<\lambda\right>^{\frac{1}{2}}|x-x_{0}|))^{2}d\lVert E_{\lambda}u(y)\rVert^{2}
C1λ2eελdEλu(y)2C1eελdEλu(y)<\displaystyle\leq C\int_{1}^{\infty}\lambda^{2}e^{\varepsilon^{\prime}\sqrt{\lambda}}d\lVert E_{\lambda}u(y)\rVert^{2}\leq C\int_{1}^{\infty}e^{\varepsilon\sqrt{\lambda}}d\lVert E_{\lambda}u(y)\rVert<\infty

for ε<ε\varepsilon^{\prime}<\varepsilon and xx within the interval specified above.
We now justify that xwL2\partial_{x}w\in L^{2} and can be expressed as

(15) xw=1xv(x,λ)dEλu(y).\displaystyle\partial_{x}w=\int_{1}^{\infty}\partial_{x}v(x,\lambda)dE_{\lambda}u(y).

To do that we use difference quotients to interchange derivatives and integration in λ\lambda. Indeed,

Dhw(x):=w(x+h,y)w(x,y)h=11h[v(x+h,λ)v(x,λ)]𝑑Eλu(y)\displaystyle D_{h}w(x):=\frac{w(x+h,y)-w(x,y)}{h}=\int_{1}^{\infty}\frac{1}{h}[v(x+h,\lambda)-v(x,\lambda)]dE_{\lambda}u(y)

Hence

Dhw(x)=101xv(x+sh,λ)dsdEλu(y)\displaystyle D_{h}w(x)=\int_{1}^{\infty}\int_{0}^{1}\partial_{x}v(x+sh,\lambda)dsdE_{\lambda}u(y)

Since |xx0|ε2Cx0|x-x_{0}|\leq\frac{\varepsilon^{\prime}}{2C_{x_{0}}} and for |h|<εε2sCx0|h|<\frac{\varepsilon-\varepsilon^{\prime}}{2sC_{x_{0}}} we use Lemma 12 to obtain

|xv(x+sh,λ)|λ12eCx0ελ|xx0+sh|eελ\displaystyle|\partial_{x}v(x+sh,\lambda)|\leq\!\left<\lambda\right>^{\frac{1}{2}}e^{C_{x_{0}}\varepsilon\sqrt{\lambda}|x-x_{0}+sh|}\leq e^{\varepsilon\sqrt{\lambda}}

Therefore, DhwL2D_{h}w\in L^{2} for hh sufficiently small. A similar argument shows that for any hn0h_{n}\to 0, DhnwD_{h_{n}}w is Cauchy and thus has a limit, that we call xv\partial_{x}v. By Lebesgue Dominated Convergence 01xv(x+sh,λ)dsv(x,λ)\int_{0}^{1}\partial_{x}v(x+sh,\lambda)ds\to v(x,\lambda) as h0h\to 0. This proves (15).

Analysis of x2wL2\partial_{x}^{2}w\in L^{2} and an analogue of (15) is similar. Combining the terms implies LwL2Lw\in L^{2}.

We now demonstrate that under the hypothesis of the Lemma Lw=0Lw=0. Using spectral projections, e.g. Proposition 23

Lw(x,y)=x2w+a1(x)xw+1[a0(x)+g(x)λ]v(x,λ)𝑑Eλu(y)\displaystyle Lw(x,y)=-\partial_{x}^{2}w+a_{1}(x)\partial_{x}w+\int_{1}^{\infty}[a_{0}(x)+g(x)\lambda]v(x,\lambda)dE_{\lambda}u(y)

We have shown above that we can pass xx derivatives into the projection for ww. Hence

Lw(x,y)=1[x2v(x,λ)+a1(x)xv(x,λ)+[a0(x)+g(x)λ]v(x,λ)dEλu(y)Lw(x,y)=\int_{1}^{\infty}[-\partial_{x}^{2}v(x,\lambda)+a_{1}(x)\partial_{x}v(x,\lambda)+[a_{0}(x)+g(x)\lambda]v(x,\lambda)dE_{\lambda}u(y)

As v(x,λ)v(x,\lambda) solves (10), Lw=0Lw=0. ∎

3. Hypoellipticity and closed graph theorem

For generic uL2u\in L^{2} or even C0C^{\infty}_{0}, we may not be able to guarantee the hypothesis of Lemma 13. To go around this concern, we first use the Closed Graph Theorem to control smoothness of solutions to Lw=0Lw=0 using hypoellipticity. We start with a bit of notation.

Definition 14.

For r>0r>0 define Ir:=(x0r,x0+r)I_{r}:=(x_{0}-r,x_{0}+r)

Lemma 15.

Let y0Ryny_{0}\in R^{n}_{y} with y0Ωyny_{0}\in\Omega\Subset\operatorname{\mathbb{R}}^{n}_{y} and let r>0r>0. For IrI_{r} from Definition 14, define

(16) X:={wL2(Ir×Ω)Lw=0 in the weak sense on Ir×Ω}\displaystyle X:=\{w\in L^{2}({I_{r}}\times\Omega)\mid Lw=0\text{ in the weak sense on }I_{r}\times\Omega\}

Then XX is a closed subspace of Lloc2(Ir×Ω)L^{2}_{loc}({I_{r}}\times\Omega).

Proof.

We invoke duality. Let wnwLloc2w_{n}\to w\in L^{2}_{loc} with Lwn=0Lw_{n}=0 weakly. Then for ϕC0(Ir×Ω)\phi\in C^{\infty}_{0}({I_{r}}\times\Omega)

(Lw,ϕ)=(w,Lϕ)=limn(wn,Lϕ)=0\displaystyle(Lw,\phi)=(w,L^{*}\phi)=\lim_{n}(w_{n},L^{*}\phi)=0

Thus Lw=0Lw=0 weakly and wXw\in X. ∎

Lemma 16.

[Closed Graph] Let ΩΩyn\Omega^{\prime}\Subset\Omega\subset\operatorname{\mathbb{R}}^{n}_{y} and 0<r<r0<r^{\prime}<r. Let s>0s>0. Define IrI_{r} and IrI_{r^{\prime}} by Definition 14.
Suppose the operator LL from (8) is HsH^{s} hypoelliptic at (x0,y0)(x_{0},y_{0}) in the sense of Definition 3 and XX is as in (16). Then the operator

T:XHs(Ir×Ω) defined by Tw=w\displaystyle T:X\to H^{s}(I_{r^{\prime}}\times\Omega^{\prime})\text{ defined by }Tw=w

has a closed graph. In particular, if wXw\in X, then

(17) wHs(Ir×Ω)C~(Ω,Ω,r,r)wL2(Ir×Ω)\displaystyle\lVert w\rVert_{H^{s}(I_{r^{\prime}}\times\Omega^{\prime})}\leq\tilde{C}(\Omega,\Omega^{\prime},r,r^{\prime})\lVert w\rVert_{L^{2}({I_{r}}\times\Omega)}
Remark 17.

As rrr^{\prime}\to r or d(Ω,Ω)0d(\partial\Omega^{\prime},\partial\Omega)\to 0 the constant in (17) may go to infinity.

Proof.

First, we demonstrate TT is well defined. Let ψC0(Ir×Ω)\psi\in C^{\infty}_{0}(I_{r}\times\Omega) be a bump function 0ψ10\leq\psi\leq 1 with ψ(x,y)1\psi(x,y)\equiv 1 on Ir×ΩI_{r^{\prime}}\times\Omega^{\prime}. Then, for all uXu\in X, L(ψu)(x,y)=0L(\psi u)(x,y)=0 for (x,y)Ir×Ω(x,y)\in I_{r^{\prime}}\times\Omega^{\prime}. By Definition 3 ψuHs\psi u\in H^{s}. Hence uHs(Ir×Ω)u\in H^{s}(I_{r^{\prime}}\times\Omega^{\prime}).

To demonstrate that TT has a closed graph, suppose wnXw_{n}\in X,

wnwX, and Twnw~Hs(Ir×Ω),w_{n}\to w\in X,\text{ and }Tw_{n}\to\tilde{w}\in H^{s}(I_{r^{\prime}}\times\Omega^{\prime}),

Passing to subsequences if necessary, we may assume that wnw~w_{n}\to\tilde{w} and wnww_{n}\to w for a.e. (x,y)Ir×Ω(x,y)\in I_{r^{\prime}}\times\Omega^{\prime}. Since Hs(Ir×Ω)L2(Ir×Ω)H^{s}(I_{r^{\prime}}\times\Omega^{\prime})\subset L^{2}(I_{r^{\prime}}\times\Omega^{\prime}), it is then clear that w=w~w=\tilde{w} and the graph of TT is closed. By the closed graph theorem, c.f. [Yos95] p.80, TT is continuous, and hence satisfies (17). ∎

We now need a version of Sobolev Embedding to apply Lemma 16 in HysH^{s}_{y} for 12<sn+12\frac{1}{2}<s\leq\frac{n+1}{2}, where yny\in\operatorname{\mathbb{R}}^{n}. If we use CC^{\infty} hypoellipticity or HsH^{s} hypoellipticity for ss bigger than this range, the standard Sobolev Embedding Hn+12+(n+1)C0(n+1)H^{\frac{n+1}{2}^{+}}(\operatorname{\mathbb{R}}^{n+1})\subset C^{0}(\operatorname{\mathbb{R}}^{n+1}) suffices.

Lemma 18.

[Sobolev Embedding] Let

(18) s=s1+s2 for s1>12 and s20\displaystyle s=s_{1}+s_{2}\text{ for }s_{1}>\frac{1}{2}\text{ and }s_{2}\geq 0

Suppose u(x,y)Hx,ys(x×yn)u(x,y)\in H^{s}_{x,y}(\operatorname{\mathbb{R}}_{x}\times\operatorname{\mathbb{R}}^{n}_{y}) . Then xu(x,)x\to u(x,\cdot) is a continuous uniformly bounded function from Hys2\operatorname{\mathbb{R}}\to H^{s_{2}}_{y} with

(19) uLxHys2Cs1,s2uHx,ys\displaystyle\lVert u\rVert_{L^{\infty}_{x}H^{s_{2}}_{y}}\leq C_{s_{1},s_{2}}\lVert u\rVert_{H^{s}_{x,y}}
Proof.

Define

(20) us:=ys2xs1u,\displaystyle u_{s}:=\!\left<\partial_{y}\right>^{s_{2}}\!\left<\partial_{x}\right>^{s_{1}}u,

where xs1\!\left<\partial_{x}\right>^{s_{1}} is a pseudodifferential operator with symbol ξs1\!\left<\xi\right>^{s_{1}} and similarly for ys2\!\left<\partial_{y}\right>^{s_{2}}. We can re-write this formula as

u^s(ξ,η)=(1+|ξ|2)s1(1+|η|2)s2(1+|ξ|2+|η|2))s(1+|ξ|2+|η|2))su^(η,ξ)\displaystyle\hat{u}_{s}(\xi,\eta)=\frac{(1+|\xi|^{2})^{s_{1}}(1+|\eta|^{2})^{s_{2}}}{(1+|\xi|^{2}+|\eta|^{2}))^{s}}(1+|\xi|^{2}+|\eta|^{2}))^{s}\hat{u}(\eta,\xi)

By hypothesis, uHsu\in H^{s}, and the term (1+|ξ|2)s1(1+|η|2)s2(1+|ξ|2+|η|2))s\frac{(1+|\xi|^{2})^{s_{1}}(1+|\eta|^{2})^{s_{2}}}{(1+|\xi|^{2}+|\eta|^{2}))^{s}} is a bounded Fourier multiplier for s1s_{1}, s20s_{2}\geq 0. Thus

(21) usLx,y2Cs1,s2uHx,ys\displaystyle\lVert u_{s}\rVert_{L^{2}_{x,y}}\leq C_{s_{1},s_{2}}\lVert u\rVert_{H^{s}_{x,y}}

We return to the analysis of uu using (20). Write formally

(22) u(x,y)=xs1ys2us(x,y),\displaystyle u(x,y)=\!\left<\partial_{x}\right>^{-s_{1}}\!\left<\partial_{y}\right>^{-s_{2}}u_{s}(x,y),

and define, g(x)g(x) as an inverse Fourier transform of xs1\!\left<\partial_{x}\right>^{-s_{1}}, i.e.

g^(ξ)=(1+|ξ|2)s1Lξ2 for s1>12.\displaystyle\hat{g}(\xi)=(1+|\xi|^{2})^{-s_{1}}\in L^{2}_{\xi}\text{ for }s_{1}>\frac{1}{2}.

Note that from the definition gLx2Cs1\lVert g\rVert_{L^{2}_{x}}\leq C_{s_{1}}. With this notation we rewrite (22) as

(23) u(x,y)=g(x)xys2us.\displaystyle u(x,y)=g(x)\ast_{x}\!\left<\partial_{y}\right>^{-s_{2}}u_{s}.

Applying Cauchy-Schwartz, (21) and the L2L^{2} estimate for gg implies

(24) uLxHys2Cs1ys2usLx2Hys2=Cs1usLx,y2Cs1,s2uHx,ys,\displaystyle\lVert u\rVert_{L^{\infty}_{x}H^{s_{2}}_{y}}\leq C_{s_{1}}\lVert\!\left<\partial_{y}\right>^{-s_{2}}u_{s}\rVert_{L^{2}_{x}H^{s_{2}}_{y}}=C_{s_{1}}\lVert u_{s}\rVert_{L^{2}_{x,y}}\leq C_{s_{1},s_{2}}\lVert u\rVert_{H^{s}_{x,y}},

which concludes (19). Furthermore, to demonstrate continuity we use (23)

u(x,)u(x,)Hys2|g(xz)g(xz)|ys2us(z,)Hys2𝑑z\displaystyle\lVert u(x,\cdot)-u(x^{\prime},\cdot)\rVert_{H^{s_{2}}_{y}}\leq\int\left|g(x-z)-g(x^{\prime}-z)\right|\lVert\!\left<\partial_{y}\right>^{-s_{2}}u_{s}(z,\cdot)\rVert_{H^{s_{2}}_{y}}dz

Then by Cauchy-Schwartz

u(x,)u(x,)Hys2(|g(xz)g(xz)|2𝑑z)1/2usLz,y2\displaystyle\lVert u(x,\cdot)-u(x^{\prime},\cdot)\rVert_{H^{s_{2}}_{y}}\leq\left(\int|g(x-z)-g(x^{\prime}-z)|^{2}dz\right)^{1/2}\lVert u_{s}\rVert_{L^{2}_{z,y}}

As xxx^{\prime}\to x, the first factor on the right converges to zero by properties of the translation operator in L2L^{2}.∎

We conclude the section by combining Lemmas 16 and 18.

Proposition 19.

Let s=s1+s2s=s_{1}+s_{2} with s1>12s_{1}>\frac{1}{2} and s2>0s_{2}>0. Suppose the operator LL from (8) is HsH^{s} hypoelliptic at (x0,y0)(x_{0},y_{0}). Then for some ΩΩyn\Omega^{\prime}\Subset\Omega\Subset\operatorname{\mathbb{R}}^{n}_{y} and r>0r>0 small enough,

(25) w(x0,)Hs2(Ω)C~(s1,s2,Ω,Ω,r)wL2(Ir×Ω)\displaystyle\lVert w(x_{0},\cdot)\rVert_{H^{s_{2}}(\Omega^{\prime})}\leq\tilde{C}(s_{1},s_{2},\Omega,\Omega^{\prime},r)\lVert w\rVert_{L^{2}({I_{r}}\times\Omega)}

4. Hypoellipticity with spectral projections

As mentioned above, for generic uLy2u\in L^{2}_{y} or even C0C^{\infty}_{0}, we may not be able to guarantee the hypothesis of Lemma 13. We show below that for hypoelliptic operators, Lemma 13 applies for any PjuP_{j}u. In particular, this will imply that for hypoelliptic operator LL the spectral projection PjP_{j} from (48) is smoothing. We state this in a form that is crucial for the proof of Theorem 4.

Proposition 20.

Let s=s1+s2s=s_{1}+s_{2} with s1>12s_{1}>\frac{1}{2} and s2>0s_{2}>0. Suppose LL from (8) is HsH^{s} hypoelliptic near (x0,y0)(x_{0},y_{0}) and let the operator PjP_{j} be defined by (48). Let ε>0\varepsilon>0 and y0ΩΩyny_{0}\in\Omega^{\prime}\Subset\Omega\Subset\operatorname{\mathbb{R}}^{n}_{y}, then

(26) PjuHs2(Ω)C~(ε,x0,s1,s2,Ω,Ω)exp(εej)PjuL2(Ω)\displaystyle\lVert P_{j}u\rVert_{H^{s_{2}}(\Omega^{\prime})}\leq\tilde{C}(\varepsilon,x_{0},s_{1},s_{2},\Omega,\Omega^{\prime})\exp(\varepsilon\sqrt{e^{j}})\lVert P_{j}u\rVert_{L^{2}(\Omega)}

for all u(y)Ly2(Ω)u(y)\in L^{2}_{y}(\Omega) and all j0j\geq 0. The constant C~\tilde{C}\to\infty as ε0\varepsilon\to 0, and similarly as d(Ω,Ω)0d(\partial\Omega^{\prime},\partial\Omega)\to 0, see Remark 17.

Proof.

Let uL2u\in L^{2} and define uj=Pjuu_{j}=P_{j}u and

wj=v(x,B)uj=ψj(B)v(x,B)u.w_{j}=v(x,B)u_{j}=\psi_{j}(B)v(x,B)u.

Note by (10) wj(x0)=ujw_{j}(x_{0})=u_{j}. By Lemma 26

(27) eεBuj(y)Ly2|jj|1eεejPjPju<\displaystyle\lVert e^{\varepsilon\sqrt{B}}u_{j}(y)\rVert_{L^{2}_{y}}\leq\sum_{|j^{\prime}-j|\leq 1}e^{\varepsilon\sqrt{e^{j^{\prime}}}}\lVert P_{j^{\prime}}P_{j}u\rVert<\infty

Therefore, Lemma 13 applies to uju_{j} and gives

Lwj(x,y)=0 for |xx0|r0.Lw_{j}(x,y)=0\text{ for }|x-x_{0}|\leq r_{0}.

Thus hypoellipticity of LL implies that wjClocw_{j}\in C^{\infty}_{loc}. Note, in particular, PjuP_{j}u is locally smooth. It remains to establish (26) - a quantitative control of this smoothness.

We now invoke Proposition 19. By (25)

(28) wj(x0,)Hs2(Ω)C~wjL2(Ir×Ω)\displaystyle\lVert w_{j}(x_{0},\cdot)\rVert_{H^{s_{2}}(\Omega^{\prime})}\leq\tilde{C}\lVert w_{j}\rVert_{L^{2}(I_{r}\times\Omega)}

where the constant C~\tilde{C} is independent of wjw_{j}. In particular, as r0r\to 0, C~\tilde{C} may go to infinity by Remark 17.

Recall that wj(x0,y)=Pjuw_{j}(x_{0},y)=P_{j}u. Therefore we can restate (28) as

(29) PjuHs2(Ω)C~wjL2(Ir×Ω)\displaystyle\lVert P_{j}u\rVert_{H^{s_{2}}(\Omega^{\prime})}\leq\tilde{C}\lVert w_{j}\rVert_{L^{2}(I_{r}\times\Omega)}

Proposition 23 and the support of ψj\psi_{j} imply

(30) wj(x,y)L2(Ir×Ω)2x0rx0+rej1ej+1|v(x,λ)|2ψj2(λ)dEλuLy22𝑑x\displaystyle\lVert w_{j}(x,y)\rVert_{L^{2}(I_{r}\times\Omega)}^{2}\leq\int_{x_{0}-r}^{x_{0}+r}\int_{e^{j-1}}^{e^{j+1}}|v(x,\lambda)|^{2}\psi_{j}^{2}(\lambda)d\lVert E_{\lambda}u\rVert_{L^{2}_{y}}^{2}dx

From Lemma 12 we get

|v(x,λ)|eCx0λ12|xx0|\displaystyle|v(x,\lambda)|\leq e^{C_{x_{0}}\!\left<\lambda\right>^{\frac{1}{2}}|x-x_{0}|}

Note, that Cx0C_{x_{0}}, unlike C~\tilde{C} is independent of ε\varepsilon. Hence for λ[ej1,ej+1]\lambda\in[{e^{j-1}},{e^{j+1}}] and x[x0r,x0+r]x\in[{x_{0}-r},{x_{0}+r}]:

|v(x,λ)|2e2Cx0ej+12r=eCx0ej2r, where Cx0 is replaced with a larger constant\displaystyle|v(x,\lambda)|^{2}\leq e^{2C_{x_{0}}e^{\frac{j+1}{2}}r}=e^{C_{x_{0}}e^{\frac{j}{2}}r},\text{ where }C_{x_{0}}\text{ is replaced with a larger constant}
that is still independent of jj and ε\varepsilon in the last equality.

For ε>0\varepsilon>0 in the hypothesis of our Proposition choose

r=Cx01ε.\displaystyle r=C_{x_{0}}^{-1}\varepsilon.

Then

|v(x,λ)|2exp(εej).\displaystyle|v(x,\lambda)|^{2}\leq\exp(\varepsilon\sqrt{e^{j}}).

We now return to (30) with this estimate of vv on our projected band:

wj(x,y)L2(Ir×Ω)2x0rx0+rexp(εej)ej1ej+1ψj2(λ)dEλuLy22𝑑x\displaystyle\lVert w_{j}(x,y)\rVert_{L^{2}(I_{r}\times\Omega)}^{2}\leq\int_{x_{0}-r}^{x_{0}+r}\exp(\varepsilon\sqrt{e^{j}})\int_{e^{j-1}}^{e^{j+1}}\psi_{j}^{2}(\lambda)d\lVert E_{\lambda}u\rVert_{L^{2}_{y}}^{2}dx

Since PjuL2(Ω)2=ej1ej+1ψj2(λ)dEλuL2(Ω)2\lVert P_{j}u\rVert_{L^{2}(\Omega)}^{2}=\int_{e^{j-1}}^{e^{j+1}}\psi_{j}^{2}(\lambda)d\lVert E_{\lambda}u\rVert_{L^{2}(\Omega)}^{2}, we rewrite this as

wjL2(Ir×Ω)2εexp(εej)PjuL2(Ω)2\displaystyle\lVert w_{j}\rVert_{L^{2}(I_{r}\times\Omega)}^{2}\leq\varepsilon\exp(\varepsilon\sqrt{e^{j}})\lVert P_{j}u\rVert^{2}_{L^{2}(\Omega)}

This estimate together with (29) gives

PjuHs2(Ω)2C~εexp(εej)PjuL2(Ω)2\displaystyle\lVert P_{j}u\rVert_{H^{s_{2}}(\Omega^{\prime})}^{2}\leq\tilde{C}\varepsilon\exp(\varepsilon\sqrt{e^{j}})\lVert P_{j}u\rVert^{2}_{L^{2}(\Omega)}

Note, that this estimate does not imply that Pju0P_{j}u\to 0, as ε0\varepsilon\to 0, since C~\tilde{C} depends on ε\varepsilon itself. Absorbing, ε\varepsilon into a new constant C~\tilde{C} concludes the proof. ∎

5. Interpolation and proof Theorem 4

Lemma 21.

Let ε>0\varepsilon>0, s2>0s_{2}>0. For each ξn\xi\in\operatorname{\mathbb{R}}^{n} let R=R(ξ)R=R(\xi) be defined by

(31) eR=(s2logξ2ε)2.\displaystyle e^{R}=\left(\frac{s_{2}\log\!\left<\xi\right>}{2\varepsilon}\right)^{2}.

Consider a sequence of functions αk(y)Hys2\alpha_{k}(y)\in H^{s_{2}}_{y} for k=0,1,2k=0,1,2\ldots Then

(32) kR(ξ)logξαk^(ξ)Lξ22C(ε,s2)k=0e2εekαkHys22\displaystyle\lVert\sum_{k\leq R(\xi)}\log\!\left<\xi\right>\widehat{\alpha_{k}}(\xi)\rVert_{L^{2}_{\xi}}^{2}\leq C(\varepsilon,s_{2})\sum_{k=0}^{\infty}e^{-2\varepsilon\sqrt{e^{k}}}\lVert\alpha_{k}\rVert_{H^{s_{2}}_{y}}^{2}
(33) kRlogξαk^(ξ)Lξ22C(εs2)2k=0ekαkLy22\displaystyle\lVert\sum_{k\geq R}\log\!\left<\xi\right>\widehat{\alpha_{k}}(\xi)\rVert_{L^{2}_{\xi}}^{2}\leq C\left(\frac{\varepsilon}{s_{2}}\right)^{2}\sum_{k=0}^{\infty}e^{k}\lVert\alpha_{k}\rVert_{L^{2}_{y}}^{2}
Proof.

Both of the inequalities are proved by a clever use of the Cauchy-Schwartz inequality.

By the Cauchy-Schwartz inequality

kRlogξ|αk^(ξ)|(kRlog2ξξ2s2e2εek)12(kRξ2s2e2εek|αk^|2)12\displaystyle\sum_{k\leq R}\log\!\left<\xi\right>|\widehat{\alpha_{k}}(\xi)|\leq\left(\sum_{k\leq R}\frac{\log^{2}\!\left<\xi\right>}{\!\left<\xi\right>^{2s_{2}}}e^{2\varepsilon\sqrt{e^{k}}}\right)^{\frac{1}{2}}\cdot\left(\sum_{k\leq R}\frac{\!\left<\xi\right>^{2s_{2}}}{e^{2\varepsilon\sqrt{e^{k}}}}|\widehat{\alpha_{k}}|^{2}\right)^{\frac{1}{2}}

The terms of the first sum are increasing in kk, so the sum can be estimated

kRlog2ξξ2s2e2εeklog2ξξ2s2Re2εeR\displaystyle\sum_{k\leq R}\frac{\log^{2}\!\left<\xi\right>}{\!\left<\xi\right>^{2s_{2}}}e^{2\varepsilon\sqrt{e^{k}}}\leq\frac{\log^{2}\!\left<\xi\right>}{\!\left<\xi\right>^{2s_{2}}}R\cdot e^{2\varepsilon\sqrt{e^{R}}}

Substitution of the identity (31) (which is equivalent to ξs2=e2εeR\!\left<\xi\right>^{s_{2}}=e^{2\varepsilon\sqrt{e^{R}}}) implies

kRlog2ξξ2s2e2εeklog2ξ(2logs2+2loglogξ2log2ε)ξs2C(ε,s2)\displaystyle\sum_{k\leq R}\frac{\log^{2}\!\left<\xi\right>}{\!\left<\xi\right>^{2s_{2}}}e^{2\varepsilon\sqrt{e^{k}}}\leq\frac{\log^{2}\!\left<\xi\right>(2\log s_{2}+2\log\log\!\left<\xi\right>-2\log 2\varepsilon)}{\!\left<\xi\right>^{s_{2}}}\leq C(\varepsilon,s_{2})

Combining the last three estimates we obtain

kRlogξαk^(ξ)Lξ22C(ε,s2)kRξ2s2e2εek|αk^|2dξ\displaystyle\lVert\sum_{k\leq R}\log\!\left<\xi\right>\widehat{\alpha_{k}}(\xi)\rVert_{L^{2}_{\xi}}^{2}\leq C(\varepsilon,s_{2})\int\sum_{k\leq R}\frac{\!\left<\xi\right>^{2s_{2}}}{e^{2\varepsilon\sqrt{e^{k}}}}|\widehat{\alpha_{k}}|^{2}d\xi

Observe

kRξ2s2e2εek|αk^|2k=0ξ2s2e2εek|αk^|2\displaystyle\sum_{k\leq R}\frac{\!\left<\xi\right>^{2s_{2}}}{e^{2\varepsilon\sqrt{e^{k}}}}|\widehat{\alpha_{k}}|^{2}\leq\sum_{k=0}^{\infty}\frac{\!\left<\xi\right>^{2s_{2}}}{e^{2\varepsilon\sqrt{e^{k}}}}|\widehat{\alpha_{k}}|^{2}

Interchanging the sum and integration using Fubini concludes the proof of (32).

Similarly for (33), by Cauchy-Schwartz

(kRlogξ|αk^(ξ)|)2(kRlog2ξek)(kRek|αk^|2)\displaystyle\left(\sum_{k\geq R}\log\!\left<\xi\right>|\widehat{\alpha_{k}}(\xi)|\right)^{2}\leq\left(\sum_{k\geq R}\frac{\log^{2}\!\left<\xi\right>}{e^{k}}\right)\cdot\left(\sum_{k\geq R}e^{k}|\widehat{\alpha_{k}}|^{2}\right)

The first sum is a geometric series in kk, so

kRlog2ξek=log2ξeRee13(4εs2)2\displaystyle\sum_{k\geq R}\frac{\log^{2}\!\left<\xi\right>}{e^{k}}=\frac{\log^{2}\!\left<\xi\right>}{e^{R}}\frac{e}{e-1}\leq 3\left(4\frac{\varepsilon}{s_{2}}\right)^{2}

Integrating in ξ\xi gives

kRlogξαk^L22C(εs2)2kek|α^k|2dξ\displaystyle\lVert\sum_{k\geq R}\log\!\left<\xi\right>\widehat{\alpha_{k}}\rVert_{L^{2}}^{2}\leq C\left(\frac{\varepsilon}{s_{2}}\right)^{2}\int\sum_{k}e^{k}|\hat{\alpha}_{k}|^{2}d\xi

Applying Fubini concludes (33) ∎

Proof of Theorem 4.

Suppose LL from (8) is HsH^{s} hypoelliptic for s>12s>\frac{1}{2}, then Proposition 20 applies for s2=ss1>0s_{2}=s-s_{1}>0 with some s1>12s_{1}>\frac{1}{2}.

Let u(y)C0(Ω)u(y)\in C^{\infty}_{0}(\Omega^{\prime}), and define αj=Pju\alpha_{j}=P_{j}u. Since PjP_{j}’s are a resolution of identity by Lemma 26, we can decompose

u=j0Pju=j0αj\displaystyle u=\sum_{j\geq 0}P_{j}u=\sum_{j\geq 0}\alpha_{j}

By the triangle inequality

logξu^L2logξjRα^jL2+logξjRα^jL2\lVert\log\!\left<\xi\right>\hat{u}\rVert_{L^{2}}\leq\lVert\log\!\left<\xi\right>\sum_{j\leq R}\hat{\alpha}_{j}\rVert_{L^{2}}+\lVert\log\!\left<\xi\right>\sum_{j\geq R}\hat{\alpha}_{j}\rVert_{L^{2}}

By Cauchy-Schwartz

logξu^L222(logξjRα^jL22+logξjRα^jL22)\displaystyle\lVert\log\!\left<\xi\right>\hat{u}\rVert_{L^{2}}^{2}\leq 2\left(\lVert\log\!\left<\xi\right>\sum_{j\leq R}\hat{\alpha}_{j}\rVert_{L^{2}}^{2}+\lVert\log\!\left<\xi\right>\sum_{j\geq R}\hat{\alpha}_{j}\rVert_{L^{2}}^{2}\right)

We are now set to use Lemma 21 for our choice of αj\alpha_{j} (replacing kk with jj):

(34) logξuL22Cεs22j=0ejPjuL22+Cεj=0e2εejPjuHs22:=I+II\displaystyle\lVert\log\!\left<\xi\right>u\rVert_{L^{2}}^{2}\leq C\frac{\varepsilon}{s_{2}}^{2}\sum_{j=0}^{\infty}e^{j}\lVert P_{j}u\rVert_{L_{2}}^{2}+C_{\varepsilon}\sum_{j=0}^{\infty}e^{-2\varepsilon\sqrt{e^{j}}}\lVert P_{j}u\rVert_{H^{s_{2}}}^{2}:=I+II

We estimate II and IIII separately. For II we estimate the first sum via (49) for f(λ)=λf(\lambda)=\sqrt{\lambda}. As f(λ)>0f(\lambda)>0 and increasing for λ1\lambda\geq 1

I=Cε2jRe|ej1|2PjuL22Ceε2Bu2\displaystyle I=C\varepsilon^{2}\sum_{j\geq R}e\cdot|\sqrt{e^{j-1}}|^{2}\lVert P_{j}u\rVert_{L_{2}}^{2}\leq Ce\varepsilon^{2}\lVert\sqrt{B}u\rVert^{2}

Choose ε>0\varepsilon>0 small enough, so that Ceε1Ce\varepsilon\leq 1. Then

IεBu2\displaystyle I\leq\varepsilon\lVert\sqrt{B}u\rVert^{2}

Using (Bu,u)=(L2u,u)(Bu,u)=(L_{2}u,u) and the fact that BB is an extension of the operator L2L_{2} and uC0u\in C^{\infty}_{0} we obtain

Iε(L2u,u)\displaystyle I\leq\varepsilon(L_{2}u,u)

For IIII we observe from Proposition 20

PjuHs22C(ε,s)e2εejPjuL22\displaystyle\lVert P_{j}u\rVert_{H^{s_{2}}}^{2}\leq C(\varepsilon,s)e^{2\varepsilon\sqrt{e^{j}}}\lVert P_{j}u\rVert_{L^{2}}^{2}

Hence

IIC(ε,s)jPjuL22Cεu2\displaystyle II\leq C(\varepsilon,s)\sum_{j}\lVert P_{j}u\rVert_{L^{2}}^{2}\leq C_{\varepsilon}\lVert u\rVert^{2}

Combining II and IIII into (34) we obtain

logξuL2(Ω)2ε(L2u,u)+C(ε,s)uL2(Ω)2\displaystyle\lVert\log\!\left<\xi\right>u\rVert_{L^{2}(\Omega^{\prime})}^{2}\leq\varepsilon(L_{2}u,u)+C(\varepsilon,s)\lVert u\rVert^{2}_{L^{2}(\Omega)}\hskip 20.0pt\qed

Appendix A Spectral projections for the operator

We start with recalling the relevant facts about the spectral projection and constructing Littlewood-Paley projections adapted to the operator.

Note that by (2), L2L_{2} is an operator bounded below. That is, given Ωn\Omega\Subset\operatorname{\mathbb{R}}^{n} there is a constant cΩc_{\Omega}\in\operatorname{\mathbb{R}} so that

(35) (L2u,u)cΩu2 for all uC0(Ω)\displaystyle(L_{2}u,u)\geq c_{\Omega}\lVert u\rVert^{2}\text{ for all }u\in C_{0}^{\infty}(\Omega)

Fix a neighborhood of (x0,y0)(x_{0},y_{0}) with compact closure. Since Theorem 4 is a local result, we may change the operator LL from (8) away from this neighborhood to assume (35) holds with a constant c=cΩ(x0,y0)c=c_{\Omega_{(x_{0},y_{0})}}.

Remark 22.

We may assume without loss of generality that (35) holds for c=1c=1 and hence L2L_{2} is a positive operator:

(36) (L2u,u)u2\displaystyle(L_{2}u,u)\geq\lVert u\rVert^{2}
Proof.

If c<1c<1 in (35), we replace L2L_{2} with L2:=L2c+1L_{2}:=L_{2}-c+1 for cc from (35) and a0(x)a_{0}(x) with a0(x)+g(x)[c1]a_{0}(x)+g(x)[c-1]. Note that with this change the operator LL in (8) remains unchanged. Such a replacement makes L2L_{2} a positive operator, which we would still call L2L_{2}. ∎

We now extend the domain of L2L_{2} from C0(Ω)C^{\infty}_{0}(\Omega) to a unique maximal domain inside L2(Ω)L^{2}(\Omega) to make it self adjoint. More precisely, let BB be the Friedreich extension of L2L_{2} (c.f. Theorem 2 on p. 317 of [Yos95]). We then have that BB satisfies

(37) (Bu,u)u2,\displaystyle(Bu,u)\geq\lVert u\rVert^{2},

since Bu=L2uBu=L_{2}u for uC0u\in C^{\infty}_{0}. Moreover, as BB is a self-adjoint operator, the spectral theorem/calculus applies to it. We summarize the relevant parts from [Yos95] ch XI sec. 5 and 10 (p.300-343)]

Proposition 23.

There exists a spectral projection EB(λ):L2(Ω)L2(Ω)E_{B}(\lambda):L^{2}(\Omega)\to L^{2}(\Omega), that is a resolution of identity and allows to form functions of the operator BB as follows:

  1. (1)

    EBE_{B} is a projection, i.e.

    (38) EB(λ)EB(λ)=EB[min(λ,λ)]\displaystyle E_{B}(\lambda)E_{B}(\lambda^{\prime})=E_{B}\left[\min(\lambda,\lambda^{\prime})\right]

    And a resolution of identity: EB()=IE_{B}(\infty)=I, EB()=0E_{B}(-\infty)=0; Moreover EE is right-continuous.

  2. (2)

    For a measurable function f(x)f(x) we can define the function f(B)f(B) of the operator BB as follows:

    (39) f(B)u=f(λ)𝑑EB(λ)u\displaystyle f(B)u=\int_{-\infty}^{\infty}f(\lambda)dE_{B}(\lambda)u

    Where domf(B)={uL2(Ω):f(B)u2:=|f(λ)|2dEB(λ)u2<}.\operatorname{dom}f(B)=\{u\in L^{2}(\Omega):\lVert f(B)u\rVert^{2}:=\int_{-\infty}^{\infty}|f(\lambda)|^{2}d||E_{B}(\lambda)u||^{2}<\infty\}.

  3. (3)

    In particular, for f(λ)=1f(\lambda)=1 and λ\lambda gives the following identities

    (40) u=𝑑EB(λ)u\displaystyle u=\int_{-\infty}^{\infty}dE_{B}(\lambda)u Bu=λ𝑑EB(λ)u\displaystyle\hskip 10.0ptBu=\int_{-\infty}^{\infty}\lambda dE_{B}(\lambda)u

    are defined on L2(Ω)L^{2}(\Omega) and dom(B)\operatorname{dom}(B) respectively.

  4. (4)

    For any measurable functions ff and g:g:\operatorname{\mathbb{R}}\to\operatorname{\mathbb{R}}, f(B)f(B) and g(B)g(B) commute on the dom(f(B)g(B))dom\left(f(B)g(B)\right) defined via (39). Moreover,

    (41) (f(B)g(B)u,v)=(f(B)u,g(B)v)=f(λ)g(λ)¯d(EB(λ)u,v)\displaystyle(f(B)g(B)u,v)=(f(B)u,g(B)v)=\int f(\lambda)\overline{g(\lambda)}d(E_{B}(\lambda)u,v)
Proof.

Existence of spectral projection is given Theorem XI.6.1 on p. 313 of [Yos95] via unitary operators. Part 2 and Part 3 are proved in Theorem XI.5.2. on p.311 using Part 1 and orthogonality. Part 4 is proved in Corollary XI.5.2 and for measurable functions in Theorem XI.12.3 on p. 343 of [Yos95]. ∎

We follow the notation of [Yos95] and denote

(42) E(λ,μ):=E(λ)E(μ) for λμ\displaystyle E(\lambda,\mu):=E(\lambda)-E(\mu)\text{ for }\lambda\geq\mu

One of the important consequences of part 4 of Proposition 23 is the following equality

Bu2=(Bu,u).\lVert\sqrt{B}u\rVert^{2}=(Bu,u).
Lemma 24.

Suppose (37) holds for BB. Then the projection EBE_{B} vanishes below λ=1\lambda=1, i.e. EB(1)=0E_{B}(1^{-})=0,

Proof.

It suffices to show that u:=E(μ)v=0u:=E(\mu)v=0 for all μ<1\mu<1 and all vL2(Ω)v\in L^{2}(\Omega). To achieve that we substitute uu into (37) and (40) to obtain

0(λ1)d(EB(λ)u,u)0\leq\int_{-\infty}^{\infty}(\lambda-1)d(E_{B}(\lambda)u,u)

We first consider the large frequencies and decompose the integral into a Riemann sum as follows

μ(λ1)d(E(λ)u,u)\displaystyle\int^{\infty}_{\mu}(\lambda-1)d(E(\lambda)u,u)
=supμ=λ0<λ1<<λN=Mj=0N(λj1)(E(λj,λj1)E(μ)v,E(μ)v),\displaystyle=\sup_{\mu=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=M}\sum_{j=0}^{N}(\lambda_{j}-1)(E(\lambda_{j},\lambda_{j-1})E(\mu)v,E(\mu)v),

Here the supremum is taken over all partitions μ=λ0<λ1<<λN=M\mu=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=M of all intervals [μ,M][μ,)[\mu,M]\subset[\mu,\infty). By (42) and (38) all terms in the sum above vanish. We conclude that

μ(λ1)d(E(λ)u,u)=0\displaystyle\int^{\infty}_{\mu}(\lambda-1)d(E(\lambda)u,u)=0

Hence the integral over the low frequencies must be non-negative. We proceed in the low frequency case as before

μ(λ1)d(E(λ)u,u)=supM=λ0<λ1<<λN=μj=1N(λj1)(E(λj,λj1)E(μ)v,E(μ)v)\displaystyle\int_{-\infty}^{\mu}(\lambda-1)d(E(\lambda)u,u)=\sup_{-M=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=\mu}\sum_{j=1}^{N}(\lambda_{j}-1)(E(\lambda_{j},\lambda_{j-1})E(\mu)v,E(\mu)v)

with the supremum taken over all partitions M=λ0<λ1<<λN=μ-M=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=\mu of all intervals [M,μ](,μ][-M,\mu]\subset(-\infty,\mu]. Observe that (38) implies orthogonality of projections localized at different frequencies, i.e. E(λj,λj1)E(λk,λk1)u=0E(\lambda_{j},\lambda_{j-1})E(\lambda_{k},\lambda_{k-1})u=0 for different jj and kk. Moreover, using (41) with f=χ(0,λj)χ(0,λj1)f=\chi_{(0,\lambda_{j})}-\chi_{(0,\lambda_{j-1})} and g=χ(0,λk)χ(0,λk1)g=\chi_{(0,\lambda_{k})}-\chi_{(0,\lambda_{k-1})} we have

μ(λ1)d(E(λ)u,u)=supM=λ0<λ1<<λN=μj=0N(λj1)E(λj,λj1)v2\displaystyle\int_{-\infty}^{\mu}(\lambda-1)d(E(\lambda)u,u)=\sup_{-M=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=\mu}\sum_{j=0}^{N}(\lambda_{j}-1)\lVert E(\lambda_{j},\lambda_{j-1})v\rVert^{2}
(μ1)supM=λ0<λ1<<λN=μj=0NE(λj,λj1)v2\displaystyle\leq(\mu-1)\sup_{-M=\lambda_{0}<\lambda_{1}<\ldots<\lambda_{N}=\mu}\sum_{j=0}^{N}\lVert E(\lambda_{j},\lambda_{j-1})v\rVert^{2}

Using orthogonality of distinct frequency interval again we obtain

μ(λ1)d(E(λ)u,u)(μ1)E(μ)v2\displaystyle\int_{-\infty}^{\mu}(\lambda-1)d(E(\lambda)u,u)\leq(\mu-1)\lVert E(\mu)v\rVert^{2}

Combining all the estimates established in this proof we obtain

0(λ1)d(EB(λ)u,u)(μ1)E(μ)v20,sinceμ<1.\displaystyle 0\leq\int_{-\infty}^{\infty}(\lambda-1)d(E_{B}(\lambda)u,u)\leq(\mu-1)\lVert E(\mu)v\rVert^{2}\leq 0,\quad\text{since}\ \ \mu<1.

We must conclude that E(μ)v=0E(\mu)v=0. ∎

In light of Lemma 24, we only need to consider λ1\lambda\geq 1. For the analysis below it is convenient to localize solutions to specific frequencies. Littlewood-Paley decomposition is ideal for that. Let ϕ\phi be a fixed cut-off function satisfying

ϕ(λ)C0([e,e]),ϕ(λ)1 for |λ|1 and 0ϕ1.\displaystyle\phi(\lambda)\in C^{\infty}_{0}([-e,e]),\,\,\phi(\lambda)\equiv 1\text{ for }|\lambda|\leq 1\text{ and }0\leq\phi\leq 1.

Moreover, we assume that ϕ\phi is even and nonincreasing in [0,e][0,e]. Now, for any integer jj let

ϕj(λ)ϕ(λej).\displaystyle\phi_{j}(\lambda)\equiv\phi(\lambda\cdot e^{-j}).

Then ϕj(λ)=1\phi_{j}(\lambda)=1 for |λ|ej|\lambda|\leq e^{j} and ϕj(λ)=0\phi_{j}(\lambda)=0 for |λ|ej+1|\lambda|\geq e^{j+1}. We further define cut-offs localized to |λ|ej|\lambda|\approx e^{j} as follows:

(43) ψj(λ)=ϕj(λ)ϕj1(λ) for j0,\displaystyle\psi_{j}(\lambda)=\phi_{j}(\lambda)-\phi_{j-1}(\lambda)\text{ for }j\geq 0,

so that suppψj(|λ|)=[ej1,ej+1]\operatorname{supp}{\psi_{j}(|\lambda|)}=[e^{j-1},e^{j+1}] and 0ψj10\leq\psi_{j}\leq 1.
We summarize properties of such cutoffs in this range of λ1\lambda\geq 1 as follows.

Lemma 25.

Let ψj\psi_{j} be as defined in (43) and consider λ1\lambda\geq 1. We then have

(44) suppψj(|λ|)=[ej1,ej+1] for j0;j=0ψj(λ)=1\displaystyle\operatorname{supp}\psi_{j}(|\lambda|)=[e^{j-1},e^{j+1}]\text{ for }j\geq 0;\hskip 10.0pt\sum_{j=0}^{\infty}\psi_{j}(\lambda)=1
(45) ψj(λ)ψk(λ)>0 if and only if |jk|1;12j=0ψj21\displaystyle\psi_{j}(\lambda)\cdot\psi_{k}(\lambda)>0\text{ if and only if }|j-k|\leq 1;\hskip 10.0pt\frac{1}{2}\leq\sum_{j=0}^{\infty}\psi_{j}^{2}\leq 1\

Furthermore, for any positive increasing function f:[1,)+f:[1,\infty)\to\operatorname{\mathbb{R}}^{+}

(46) j=0f(ej1)ψj2(λ)f(λ)2j=0f(ej+1)ψj2(λ)\displaystyle\sum_{j=0}^{\infty}f(e^{j-1})\psi_{j}^{2}(\lambda)\leq f(\lambda)\leq 2\sum_{j=0}^{\infty}f(e^{j+1})\psi_{j}^{2}(\lambda)
Proof.

From the definition of ψj\psi_{j} and supports of ϕj(λ)\phi_{j}(\lambda), we immediately get the support property (44) from (43). A telescoping sum argument establishes the series identity in (44). Indeed

(47) j=0nψj(λ)=ϕn(λ)ϕ1(λ)\displaystyle\sum_{j=0}^{n}\psi_{j}(\lambda)=\phi_{n}(\lambda)-\phi_{-1}(\lambda)

Since ϕn(λ)=1\phi_{n}(\lambda)=1 for |λ|en|\lambda|\leq e^{n} and ϕ1(λ)=0\phi_{-1}(\lambda)=0 for λ1\lambda\geq 1 , (44) follows.

For (45), the orthonormal property is immediate from the fact that ψj(λ)\psi_{j}(\lambda) is nonzero only when |λ|[ej1,ej+1]|\lambda|\in[e^{j-1},e^{j+1}]. From the size of ψj\psi_{j}, j=0ψj2j=0ψj=1\sum_{j=0}^{\infty}\psi_{j}^{2}\leq\sum_{j=0}^{\infty}\psi_{j}=1. By (45) and the Cauchy-Schwartz inequality,

1=(j=0ψj(λ))2j=0k=j1j+1ψjψkj=0(ψj2+12ψj12+12ψj+12)1=(\sum_{j=0}^{\infty}\psi_{j}(\lambda))^{2}\leq\sum_{j=0}^{\infty}\sum_{k=j-1}^{j+1}\psi_{j}\psi_{k}\leq\sum_{j=0}^{\infty}(\psi_{j}^{2}+\frac{1}{2}\psi_{j-1}^{2}+\frac{1}{2}\psi_{j+1}^{2})

which concludes (45).

On the support of each ψj\psi_{j} for j1j\geq 1, monotonicity of ff implies

f(ej1)f(λ)f(ej+1).f(e^{j-1})\leq f(\lambda)\leq f(e^{j+1}).

Multiplying by ψj(λ)2\psi_{j}(\lambda)^{2} makes the inequality valid for all λ\lambda:

f(ej1)ψj2(λ)f(λ)ψj2f(ej+1)ψj2(λ).f(e^{j-1})\psi_{j}^{2}(\lambda)\leq f(\lambda)\psi_{j}^{2}\leq f(e^{j+1})\psi_{j}^{2}(\lambda).

Summing up gives

j=0f(ej1)ψj2(λ)j=0f(λ)ψj2j=0f(ej+1)ψj2(λ)\displaystyle\sum_{j=0}^{\infty}f(e^{j-1})\psi_{j}^{2}(\lambda)\leq\sum_{j=0}^{\infty}f(\lambda)\psi_{j}^{2}\leq\sum_{j=0}^{\infty}f(e^{j+1})\psi_{j}^{2}(\lambda)

Finally, from (45)

j=0f(ej1)ψj2(λ)f(λ)2j=0f(ej+1)ψj2(λ)\displaystyle\sum_{j=0}^{\infty}f(e^{j-1})\psi_{j}^{2}(\lambda)\leq f(\lambda)\leq 2\sum_{j=0}^{\infty}f(e^{j+1})\psi_{j}^{2}(\lambda)

With ψj\psi_{j} as above, we define a frequency localization adapted to the operator via Proposition 23

(48) Pju:=ψj(B)u=ψj(λ)𝑑EB(λ)u\displaystyle P_{j}u:=\psi_{j}(B)u=\int\psi_{j}(\lambda)dE_{B}(\lambda)u

for jj\in\mathbb{N}. Combining Proposition 23, Lemma 24 and Lemma 25 we get

Lemma 26.

Let PjP_{j} be defined by (48). Then for any uL2(Ω)u\in L^{2}(\Omega) there holds

u=j=0Pjuu2j=0Pju2;\displaystyle u=\sum_{j=0}^{\infty}P_{j}u\hskip 15.0pt\lVert u\rVert^{2}\approx\sum_{j=0}^{\infty}\lVert P_{j}u\rVert^{2};
PjuPjv𝑑x=0 for |jj|>1 and vL2\displaystyle\int P_{j}uP_{j^{\prime}}vdx=0\text{ for }|j-j^{\prime}|>1\text{ and }v\in L^{2}

Finally, for f:[1,)+f:[1,\infty)\to\mathbb{R}_{+} nondecreasing

(49) j=0|f(ej1)|2Pju2f(B)u22j=0|f(ej+1)|2Pju2\displaystyle\sum_{j=0}^{\infty}|f\left(e^{j-1}\right)|^{2}\lVert P_{j}u\rVert^{2}\leq\lVert f(B)u\rVert^{2}\leq 2\sum_{j=0}^{\infty}|f\left(e^{j+1}\right)|^{2}\lVert P_{j}u\rVert^{2}
Proof.

By Proposition 23 Parts 2 and 4, Lemma 24 and (44)

u=𝑑E(λ)u=1j=0ψj(λ)dE(λ)u=j=0Pju\displaystyle u=\int_{-\infty}^{\infty}dE(\lambda)u=\int_{1}^{\infty}\sum_{j=0}^{\infty}\psi_{j}(\lambda)dE(\lambda)u=\sum_{j=0}^{\infty}P_{j}u

Similarly, using (45)

u2=(dE(λ)u,u)1j=0ψj2(λ)(dE(λ)u,u)=j=0Pju2\displaystyle\lVert u\rVert^{2}=\int_{-\infty}^{\infty}(dE(\lambda)u,u)\approx\int_{1}^{\infty}\sum_{j=0}^{\infty}\psi_{j}^{2}(\lambda)(dE(\lambda)u,u)=\sum_{j=0}^{\infty}\lVert P_{j}u\rVert^{2}

Orthogonality part of (45) for |jj|>1|j-j^{\prime}|>1 and (41) establish

PjuPjv𝑑x=0\int P_{j}uP_{j^{\prime}}vdx=0

Finally, from (46) and Fubini we conclude

f(B)u2=1|f(λ)|2d(Eλu,u)2j=01f(ej+1)2ψj2(λ)d(Eλu,u)\displaystyle\lVert f({B})u\rVert^{2}=\int_{1}^{\infty}|f(\lambda)|^{2}d(E_{\lambda}u,u)\leq 2\sum_{j=0}^{\infty}\int_{1}^{\infty}f(e^{j+1})^{2}\psi_{j}^{2}(\lambda)d(E_{\lambda}u,u)

The lower bound is established similarly, completing the proof. ∎

2. Acknowledgments

This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

References

  • [AKR19] Timur Akhunov, Lyudmila Korobenko, and Cristian Rios, Hypoellipticity of Fediĭ’s type operators under Morimoto’s logarithmic condition, Journal of Pseudo-Differential Operators and Applications 10 (2019), no. 3, 649–688.
  • [Chr01] Michael Christ, Hypoellipticity in the infinitely degenerate regime, Complex analysis and geometry (Columbus, OH, 1999), Ohio State Univ. Math. Res. Inst. Publ., vol. 9, de Gruyter, Berlin, 2001, pp. 59–84. MR 1912731 (2003e:35044)
  • [Chr05] by same author, A remark on sums of squares of complex vector fields, 2005.
  • [Fed71] V. S. Fediĭ, A certain criterion for hypoellipticity, Mat. Sb. (N.S.) 85 (127) (1971), 18–48. MR 0287160 (44 #4367)
  • [Hör97] Lars Hörmander, Lectures on nonlinear hyperbolic differential equations, Mathématiques & Applications (Berlin) [Mathematics & Applications], vol. 26, Springer-Verlag, Berlin, 1997. MR 1466700 (98e:35103)
  • [Hos87] Toshiko Hoshiro, A property of operators characterized by iteration and a necessary condition for hypoellipticity, J. Math. Kyoto Univ. 27 (1987), no. 3, 401–416.
  • [Koh98] Joseph J. Kohn, Hypoellipticity of some degenerate subelliptic operators, J. Funct. Anal. 159 (1998), no. 1, 203–216. MR 1654190 (99m:35035)
  • [Koh05] J. J. Kohn, Hypoellipticity and loss of derivatives, Ann. of Math. (2) 162 (2005), no. 2, 943–986, With an appendix by Makhlouf Derridj and David S. Tartakoff. MR 2183286
  • [KS84] Shigeo Kusuoka and Daniel Stroock, Applications of the Malliavin calculus. I, Stochastic analysis (Katata/Kyoto, 1982), North-Holland Math. Library, vol. 32, North-Holland, Amsterdam, 1984, pp. 271–306. MR 780762 (86k:60100a)
  • [M7́8] Guy Métivier, Propriété des itérés et ellipticité, Comm. Partial Differential Equations 3 (1978), no. 9, 827–876. MR 504629
  • [MM97a] Yoshinori Morimoto and Tatsushi Morioka, Hypoellipticity for a class of infinitely degenerate elliptic operators, Geometrical optics and related topics (Cortona, 1996), Progr. Nonlinear Differential Equations Appl., vol. 32, Birkhäuser Boston, Boston, MA, 1997, pp. 295–317. MR 2033500
  • [MM97b] by same author, The positivity of Schrödinger operators and the hypoellipticity of second order degenerate elliptic operators, Bull. Sci. Math. 121 (1997), no. 7, 507–547. MR 1485327
  • [Mor86] Yoshinori Morimoto, Nonhypoellipticity for degenerate elliptic operators, Publ. Res. Inst. Math. Sci. 22 (1986), no. 1, 25–30. MR 834345
  • [Mor87] by same author, A criterion for hypoellipticity of second order differential operators, Osaka J. Math. 24 (1987), no. 3, 651–675. MR 923880 (89b:35023)
  • [Mor92] Tatsushi Morioka, Hypoellipticity for some infinitely degenerate elliptic operators of second order, J. Math. Kyoto Univ. 32 (1992), no. 2, 373–386. MR 1173970 (93i:35035)
  • [Yos95] K. Yosida, Functional analysis, Classics in Mathematics, Cambridge University Press, 1995.