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

CR Paneitz operator on non-embeddable CR manifolds

Yuya Takeuchi Division of Mathematics
Institute of Pure and Applied Sciences
University of Tsukuba
1- 1- 1 Tennodai, Tsukuba, Ibaraki 305-8571 Japan
[email protected], [email protected]
Abstract.

The CR Paneitz operator is closely related to some important problems in CR geometry. In this paper, we consider this operator on a non-embeddable CR manifold. This operator is essentially self-adjoint and its spectrum is discrete except zero. Moreover, the eigenspace corresponding to each non-zero eigenvalue is a finite dimensional subspace of the space of smooth functions. Furthermore, we show that the CR Paneitz operator on the Rossi sphere, an example of non-embeddable CR manifolds, has infinitely many negative eigenvalues, which is significantly different from the embeddable case.

Key words and phrases:
CR Paneitz operator, embeddability
2020 Mathematics Subject Classification:
32V20, 58J50
This work was supported by JSPS KAKENHI Grant Number JP21K13792.

1. Introduction

The CR Paneitz operator plays a crucial role in CR geometry of dimension theree. For example, this operator appears in the transformation law of the logarithmic singularity of the Szegő kernel [Hirachi1993], which is also known as the CR QQ-curvature [Fefferman-Hirachi2003]. Moreover, the non-negativity of the CR Paneitz operator is deeply connected to global embeddability [Chanillo-Chiu-Yang2012, Takeuchi2020-Paneitz] and the CR positive mass theorem [Cheng-Malchiodi-Yang2017], which has an application to the CR Yamabe problem.

Let (M,T1,0M,θ)(M,T^{1,0}M,\theta) be a closed pseudo-Hermitian manifold of dimension three. We consider the CR Paneitz operator PP on MM as an unbounded operator on L2(M)L^{2}(M) with domain C(M)C^{\infty}(M), which is closable. In the embeddable case, Hsiao [Hsiao2015] has studied analytic properties of this operator; see also [Takeuchi2023-GJMS] for some improvements and generalizations to higher dimensions. The CR Paneitz operator on an embeddable CR manifold is essentially self-adjoint and has closed range. Moreover, its spectrum is a discrete subset of \mathbb{R} and consists only of eigenvalues. Furthermore, the eigenspace associated with each non-zero eigenvalue is a finite-dimensional subspace of C(M)C^{\infty}(M).

It is natural to ask what happens in the non-embeddable case. We will first prove that PP is essentially self-adjoint even in the non-embeddable case.

{theorem}

The CR Paneitz operator PP is essentially self-adjoint; equivalently, the maximal closed extension of PP is self-adjoint.

We use the same letter PP for the maximal closed extension of the CR Paneitz operator by abuse of notation. Let EE be the resolution of the identity for PP, and set

(1.1) πλE([λ,λ]):L2(M)DomP\pi_{\lambda}\coloneqq E([-\lambda,\lambda])\colon L^{2}(M)\to\operatorname{Dom}P

for λ0\lambda\geq 0. Note that πλ\pi_{\lambda} is an orthogonal projection of L2(M)L^{2}(M). We will show that if λ>0\lambda>0, then PπλP\pi_{\lambda} is a smoothing operator (section 4). As an application, we study the spectrum SpecP\operatorname{Spec}P of PP. As opposed to the embeddable case, we can only show that SpecP\operatorname{Spec}P is discrete except zero.

{theorem}

The set SpecP{0}\operatorname{Spec}P\setminus\{0\} is a discrete subset of {0}\mathbb{R}\setminus\{0\} and consists only of eigenvalues of finite multiplicity. Moreover, any eigenfunction of PP corresponding to each non-zero eigenvalue is smooth.

We will also show that πλ\pi_{\lambda} is a Heisenberg pseudodifferential operator (section 4). In particular, the singular support of its Schwartz kernel is contained in the diagonal ΔMM×M\Delta_{M}\subset M\times M. Our proof is inspired by that of [Hsiao-Marinescu2017]. Remark that π0\pi_{0} is not necessarily a Heisenberg pseudodifferential operator; see section 5.3.

We also study analytic properties of the CR Paneitz operator on the Rossi sphere, a homogeneous non-embeddable strictly pseudoconvex CR manifold; see [Chen-Shaw2001] for example. The unit sphere

(1.2) S3{(z,w)2||z|2+|w|2=1}S^{3}\coloneqq\Set{(z,w)\in\mathbb{C}^{2}}{\lvert z\rvert^{2}+\lvert w\rvert^{2}=1}

has the canonical CR structure T1,0S3T^{1,0}S^{3}. This CR structure is spanned by

(1.3) Z1w¯zz¯w.Z_{1}\coloneqq\overline{w}\frac{\partial}{\partial z}-\overline{z}\frac{\partial}{\partial w}.

A canonical contact form θ\theta on S3S^{3} is given by

(1.4) θ12(zdz¯+wdw¯z¯dzw¯dw)|S3.\theta\coloneqq\frac{\sqrt{-1}}{2}(zd\overline{z}+wd\overline{w}-\overline{z}dz-\overline{w}dw)|_{S^{3}}.

For a real number 0<|t|<10<\lvert t\rvert<1, the Rossi sphere (St3,T1,0St3)(S^{3}_{t},T^{1,0}S^{3}_{t}) is defined by

(1.5) (St3,T1,0St3)(S3,(Z1+tZ1¯)).(S^{3}_{t},T^{1,0}S^{3}_{t})\coloneqq(S^{3},\mathbb{C}(Z_{1}+t\overline{Z_{1}})).

Denote by P(t)P(t) the CR Paneitz operator with respect to (St3,T1,0St3,θ)(S^{3}_{t},T^{1,0}S^{3}_{t},\theta). We can apply the theory of spherical harmonics since P(t)P(t) is invariant under U(2)U(2)-action. Similar to the method in [Abbas-Brown-Ramasami-Zeytuncu2019], we can prove the following

{theorem}

The CR Paneitz operator P(t)P(t) on the Rossi sphere (St3,T1,0St3,θ)(S^{3}_{t},T^{1,0}S^{3}_{t},\theta) has infinitely many negative eigenvalues counted without multiplicity.

This paper is organized as follows. In section 2, we recall basic facts on CR manifolds and the definition of the CR Paneitz operator. section 3 gives a brief exposition of the Heisenberg calculus, which is a main tool for studying the CR Paneitz operator. section 4 is devoted to proofs of sections 1 and 1 as applications of Heisenberg calculus. In section 5, we study the CR Paneitz operator on the Rossi sphere via spherical harmonics. In section 6, we propose some related problems and give some observations.

2. CR manifolds

Let MM be a smooth three-dimensional manifold without boundary. A CR structure is a complex line subbundle T1,0MT^{1,0}M of the complexified tangent bundle TMTM\otimes\mathbb{C} such that

(2.1) T1,0MT0,1M=0,T^{1,0}M\cap T^{0,1}M=0,

where T0,1MT^{0,1}M is the complex conjugate of T1,0MT^{1,0}M in TMTM\otimes\mathbb{C}. Introduce an operator ¯b:C(M)Γ((T0,1M))\overline{\partial}_{b}\colon C^{\infty}(M)\to\Gamma((T^{0,1}M)^{\ast}) by

(2.2) ¯bf=(df)|T0,1M.\overline{\partial}_{b}f=(df)|_{T^{0,1}M}.

A CR manifold (M,T1,0M)(M,T^{1,0}M) is said to be embeddable if there exists a smooth embedding FF from MM into some N\mathbb{C}^{N} such that FT1,0MT1,0NF_{\ast}T^{1,0}M\subset T^{1,0}\mathbb{C}^{N}.

A CR structure T1,0MT^{1,0}M is said to be strictly pseudoconvex if there exists a nowhere-vanishing real one-form θ\theta on MM such that θ\theta annihilates T1,0MT^{1,0}M and

(2.3) 1dθ(Z,Z¯)>0,0ZT1,0M;-\sqrt{-1}d\theta(Z,\overline{Z})>0,\qquad 0\neq Z\in T^{1,0}M;

we call such a one-form a contact form. The triple (M,T1,0M,θ)(M,T^{1,0}M,\theta) is called a pseudo-Hermitian manifold. Denote by TT the Reeb vector field with respect to θ\theta; that is, the unique vector field satisfying

(2.4) θ(T)=1,Tdθ=0.\theta(T)=1,\qquad T\lrcorner\,d\theta=0.

Let Z1Z_{1} be a local frame of T1,0MT^{1,0}M, and set Z1¯=Z1¯Z_{\overline{1}}=\overline{Z_{1}}. Then (T,Z1,Z1¯)(T,Z_{1},Z_{\overline{1}}) gives a local frame of TMTM\otimes\mathbb{C}, called an admissible frame. Its dual frame (θ,θ1,θ1¯)(\theta,\theta^{1},\theta^{\overline{1}}) is called an admissible coframe. The two-form dθd\theta is written as

(2.5) dθ=1l11¯θ1θ1¯,d\theta=\sqrt{-1}l_{1\overline{1}}\theta^{1}\wedge\theta^{\overline{1}},

where l11¯l_{1\overline{1}} is a positive function. We use l11¯l_{1\overline{1}} and its multiplicative inverse l11¯l^{1\overline{1}} to raise and lower indices.

A contact form θ\theta induces a canonical connection \nabla, called the Tanaka-Webster connection with respect to θ\theta. It is defined by

(2.6) T=0,Z1=ω11Z1,Z1¯=ω1¯1¯Z1¯(ω1¯1¯=ω11¯)\nabla T=0,\quad\nabla Z_{1}={\omega_{1}}^{1}Z_{1},\quad\nabla Z_{\overline{1}}={\omega_{\overline{1}}}^{\overline{1}}Z_{\overline{1}}\quad\left\lparen{\omega_{\overline{1}}}^{\overline{1}}=\overline{{\omega_{1}}^{1}}\right\rparen

with the following structure equations:

(2.7) dθ1=θ1ω11+A11¯θθ1¯,\displaystyle d\theta^{1}=\theta^{1}\wedge{\omega_{1}}^{1}+{A^{1}}_{\overline{1}}\theta\wedge\theta^{\overline{1}},
(2.8) dl11¯=ω11l11¯+l11¯ω1¯1¯.\displaystyle dl_{1\overline{1}}={\omega_{1}}^{1}l_{1\overline{1}}+l_{1\overline{1}}{\omega_{\overline{1}}}^{\overline{1}}.

The tensor A11=A1¯1¯¯A_{11}=\overline{A_{\overline{1}\overline{1}}} is called the Tanaka-Webster torsion. The curvature form Ω11=dω11{\Omega_{1}}^{1}=d{\omega_{1}}^{1} of the Tanaka-Webster connection satisfies

(2.9) Ω11Scall11¯θ1θ1¯modulo θ,{\Omega_{1}}^{1}\equiv\mathrm{Scal}\cdot l_{1\overline{1}}\theta^{1}\wedge\theta^{\overline{1}}\qquad\text{modulo }\theta,

where Scal\mathrm{Scal} is the Tanaka-Webster scalar curvature. We denote the components of a successive covariant derivative of a tensor by subscripts preceded by a comma, for example, K11¯,1K_{1\overline{1},1}; we omit the comma if the derivatives are applied to a function. We use the index 0 for the component TT or θ\theta in our index notation. In this notation, the operator ¯b\overline{\partial}_{b} is given by

(2.10) ¯bf=f1¯θ1¯.\overline{\partial}_{b}f=f_{\overline{1}}\theta^{\overline{1}}.

The commutators of the second derivatives for uC(M)u\in C^{\infty}(M) are given by

(2.11) u11¯u1¯1=1l11¯u0,u01u10=A11u1;u_{1\overline{1}}-u_{\overline{1}1}=\sqrt{-1}l_{1\overline{1}}u_{0},\qquad u_{01}-u_{10}=A_{11}u^{1};

see [Lee1988]*(2.14). Define the Kohn Laplacian b\Box_{b} and the sub-Laplacian Δb\Delta_{b} by

(2.12) buu1¯1¯,Δbu(b+¯b)u\Box_{b}u\coloneqq-{u_{\overline{1}}}^{\overline{1}},\qquad\Delta_{b}u\coloneqq(\Box_{b}+\overline{\Box}_{b})u

for uC(M)u\in C^{\infty}(M). It follows from eq. 2.11 that

(2.13) ¯b=b1T.\overline{\Box}_{b}=\Box_{b}-\sqrt{-1}T.

It is known that (M,T1,0M)(M,T^{1,0}M) is embeddable if and only if the Kohn Laplacian b\Box_{b} has closed range [Boutet_de_Monvel1975, Kohn1986].

The CR Paneitz operator PP is the fourth-order differential operator given by

(2.14) P¯bb+𝒬,P\coloneqq\overline{\Box}_{b}\Box_{b}+\mathcal{Q},

where

(2.15) 𝒬u1(A1¯1¯u1¯),1¯.\mathcal{Q}u\coloneqq\sqrt{-1}(A^{\overline{1}\overline{1}}u_{\overline{1}})_{,\overline{1}}.

This operator is real and formally self-adjoint; see [Gover-Graham2005]*Proposition 5.1 for example. Note that our PP is just the operator P0,0P_{0,0} in this paper. Define an operator 𝒞:Γ((T0,1M))C(M)\mathcal{C}\colon\Gamma((T^{0,1}M)^{\ast})\to C^{\infty}(M) by

(2.16) 𝒞(τ1¯θ1¯)¯b(τ1¯,1¯)+1(A1¯1¯τ1¯),1¯.\mathcal{C}(\tau_{\overline{1}}\theta^{\overline{1}})\coloneqq-\overline{\Box}_{b}({\tau_{\overline{1},}}^{\overline{1}})+\sqrt{-1}(A^{\overline{1}\overline{1}}\tau_{\overline{1}})_{,\overline{1}}.

It follows from eqs. 2.10 and 2.12 that P=𝒞¯bP=\mathcal{C}\overline{\partial}_{b}. It is known that the CR Paneitz operator is non-negative if (M,T1,0M)(M,T^{1,0}M) is embeddable [Takeuchi2020-Paneitz]*Theorem 1.1. Conversely, (M,T1,0M)(M,T^{1,0}M) is embeddable if the CR Paneitz operator is non-negative and the Tanaka-Webster scalar curvature is positive [Chanillo-Chiu-Yang2012]*Theorem 1.4(a).

3. Heisenberg calculus

In this section, we recall basic properties of Heisenberg pseudodifferential operators; see [Beals-Greiner1988, Ponge2008-Book] for a comprehensive introduction to the Heisenberg calculus.

Throughout this section, we fix a closed pseudo-Hermitian manifold (M,T1,0M,θ)(M,T^{1,0}M,\theta) of dimension three. Set

(3.1) 𝔤M(TM/HM)HM.\mathfrak{g}M\coloneqq(TM/HM)\oplus HM.

The Reeb vector field TT defines a nowhere-vanishing section [T][T] of TM/HMTM/HM. For sections X0X_{0} and Y0Y_{0} of TM/HMTM/HM and XX^{\prime} and YY^{\prime} of HMHM, the Lie bracket [X0+X,Y0+Y][X_{0}+X^{\prime},Y_{0}+Y^{\prime}] is defined by

(3.2) [X0+X,Y0+Y]dθ(X,Y)[T].[X_{0}+X^{\prime},Y_{0}+Y^{\prime}]\coloneqq-d\theta(X^{\prime},Y^{\prime})[T].

This bracket makes 𝔤M\mathfrak{g}M a bundle of two-step nilpotent Lie algebras. The dilation δr\delta_{r} on 𝔤M\mathfrak{g}M is defined by

(3.3) δr|TM/HMr2,δr|HMr.\delta_{r}|_{TM/HM}\coloneqq r^{2},\qquad\delta_{r}|_{HM}\coloneqq r.

For mm\in\mathbb{Z}, the space SHm(M)S_{H}^{m}(M) consists of functions in C((𝔤M){0})C^{\infty}((\mathfrak{g}M)^{\ast}\setminus\{0\}) that are homogeneous of degree mm on each fiber. This space has a bilinear product

(3.4) :SHm1(M)×SHm2(M)SHm1+m2(M).\ast\colon S_{H}^{m_{1}}(M)\times S_{H}^{m_{2}}(M)\to S_{H}^{m_{1}+m_{2}}(M).

For mm\in\mathbb{Z}, denote by ΨHm(M)\Psi_{H}^{m}(M) the space of Heisenberg pseudodifferential operators A:C(M)C(M)A\colon C^{\infty}(M)\to C^{\infty}(M) of order mm. This space is closed under sum, complex conjugate, transpose, and formal adjoint [Ponge2008-Book]*Proposition 3.1.23. In particular, any AΨHm(M)A\in\Psi_{H}^{m}(M) extends to a linear operator

(3.5) A:𝒟(M)𝒟(M),A\colon\mathscr{D}^{\prime}(M)\to\mathscr{D}^{\prime}(M),

where 𝒟(M)\mathscr{D}^{\prime}(M) is the space of distributions on MM. For example, ZΓ(T1,0M)Z\in\Gamma(T^{1,0}M) is an element of ΨH1(M)\Psi_{H}^{1}(M) and TΨH2(M)T\in\Psi_{H}^{2}(M). Note that ΨH(M)mΨHm(M)\Psi_{H}^{-\infty}(M)\coloneqq\bigcap_{m\in\mathbb{Z}}\Psi_{H}^{m}(M) coincides with the space of smoothing operators on MM. As in the usual pseudodifferential calculus, there exists the Heisenberg principal symbol, which has some good properties:

{proposition}

[[Ponge2008-Book]*Propositions 3.2.6 and 3.2.9] (i) The Heisenberg principal symbol σm\sigma_{m} gives the following exact sequence:

(3.6) 0ΨHm1(M)ΨHm(M)σmSHm(M)0.0\to\Psi_{H}^{m-1}(M)\hookrightarrow\Psi_{H}^{m}(M)\xrightarrow{\sigma_{m}}S_{H}^{m}(M)\to 0.

(ii) For A1ΨHm1(M)A_{1}\in\Psi_{H}^{m_{1}}(M) and A2ΨHm2(M)A_{2}\in\Psi_{H}^{m_{2}}(M), the operator A1A2A_{1}A_{2} is a Heisenberg pseudodifferential operator of order m1+m2m_{1}+m_{2}, and

(3.7) σm1+m2(A1A2)=σm1(A1)σm2(A2).\sigma_{m_{1}+m_{2}}(A_{1}A_{2})=\sigma_{m_{1}}(A_{1})\ast\sigma_{m_{2}}(A_{2}).

On the other hand, there exists a crucial difference between the usual pseudodifferential calculus and the Heisenberg one. Since the product of Heisenberg principal symbol is non-commutative, the commutator [A1,A2][A_{1},A_{2}] of A1ΨHm1(M)A_{1}\in\Psi_{H}^{m_{1}}(M) and A2ΨHm2(M)A_{2}\in\Psi_{H}^{m_{2}}(M) is not an element of ΨHm1+m21(M)\Psi_{H}^{m_{1}+m_{2}-1}(M) in general. However, we have the following

{lemma}

[[Takeuchi2023-GJMS]*Lemma 4.2] If AΨHm(M)A\in\Psi_{H}^{m}(M), then [T,A]ΨHm+1(M)[T,A]\in\Psi_{H}^{m+1}(M).

Next we consider approximate inverses of Heisenberg pseudodifferential operators. We write ABA\sim B if ABA-B is a smoothing operator. Let AΨHm(M)A\in\Psi_{H}^{m}(M). An operator BΨHm(M)B\in\Psi_{H}^{-m}(M) is called a parametrix of AA if ABIAB\sim I and BAIBA\sim I. The existence of a parametrix of a Heisenberg pseudodifferential operator is determined only by its Heisenberg principal symbol.

{proposition}

[[Ponge2008-Book]*Proposition 3.3.1] Let AΨHm(M)A\in\Psi_{H}^{m}(M) with Heisenberg principal symbol aSHm(M)a\in S_{H}^{m}(M). Then the following are equivalent:

  1. (1)

    AA has a parametrix;

  2. (2)

    there exists BΨHm(M)B\in\Psi_{H}^{-m}(M) such that ABI,BAIΨH1(M)AB-I,BA-I\in\Psi_{H}^{-1}(M);

  3. (3)

    there exists bSHm(M)b\in S_{H}^{-m}(M) such that ab=ba=1a\ast b=b\ast a=1.

Now consider the Heisenberg differential operator Δb+I\Delta_{b}+I of order 22. It is known that this operator has a parametrix; see the proof of [Ponge2008-Book]*Proposition 3.5.7 for example. Since Δb+I\Delta_{b}+I is positive and self-adjoint, the k/2k/2-th power (Δb+I)k/2(\Delta_{b}+I)^{k/2} of Δb+I\Delta_{b}+I, kk\in\mathbb{Z}, is a Heisenberg pseudodifferential operator of order kk [Ponge2008-Book]*Theorems 5.3.1 and 5.4.10. Using this operator, we define

(3.8) WHk(M):={u𝒟(M)(Δb+I)k/2uL2(M)}.W_{H}^{k}(M):=\Set{u\in\mathscr{D}^{\prime}(M)\mid(\Delta_{b}+I)^{k/2}u\in L^{2}(M)}.

This space is a Hilbert space with the inner product

(3.9) (u,v)k=((Δb+I)k/2u,(Δb+I)k/2v)L2.\lparen u,v\rparen_{k}=\lparen(\Delta_{b}+I)^{k/2}u,(\Delta_{b}+I)^{k/2}v\rparen_{L^{2}}.

The space C(M)C^{\infty}(M) is dense in WHk(M)W_{H}^{k}(M) for any kk\in\mathbb{Z}, and we have

(3.10) C(M)=kWHk(M),𝒟(M)=kWHk(M)C^{\infty}(M)=\bigcap_{k\in\mathbb{Z}}W_{H}^{k}(M),\qquad\mathscr{D}^{\prime}(M)=\bigcup_{k\in\mathbb{Z}}W_{H}^{k}(M)

as topological vector spaces [Ponge2008-Book]*Proposition 5.5.3. Remark that the Hilbert space WHk(M)W_{H}^{k}(M) coincides with the Folland-Stein space Sk,2(M)S^{k,2}(M) as a topological vector space [Ponge2008-Book]*Proposition 5.5.5. Heisenberg pseudodifferential operators act on these Hilbert spaces as follows:

{proposition}

[[Ponge2008-Book]*Propositions 5.5.8 and [Takeuchi2023-GJMS]*Proposition 4.6] Any AΨHm(M)A\in\Psi_{H}^{m}(M) extends to a continuous linear operator

(3.11) A:WHk+m(M)WHk(M)A\colon W_{H}^{k+m}(M)\to W_{H}^{k}(M)

for every kk\in\mathbb{Z}. In particular if m<0m<0, the operator A:L2(M)L2(M)A\colon L^{2}(M)\to L^{2}(M) is compact.

4. Proof of sections 1 and 1

Let (M,T1,0M,θ)(M,T^{1,0}M,\theta) be a closed strictly pseudoconvex CR manifold of dimension three. If (M,T1,0M)(M,T^{1,0}M) is embeddable, then there exist the orthogonal projection to Kerb\operatorname{Ker}\Box_{b}, known as the Szegő projection, and the partial inverse of b\Box_{b} [Beals-Greiner1988]*Theorem 25.20. If (M,T1,0M)(M,T^{1,0}M) is non-embeddable, then we can not use the partial inverse of b\Box_{b}. However, we have an approximate Szegő projection and partial inverse of b\Box_{b}.

{theorem}

[[Beals-Greiner1988]*Proposition 25.4 and Corollaries 25.64 and 25.67] There exist SΨH0(M)S\in\Psi_{H}^{0}(M) and NΨH2(M)N\in\Psi_{H}^{-2}(M) such that

(4.1) bN+SNb+SI,bSSb0,SSS2,¯bS0.\Box_{b}N+S\sim N\Box_{b}+S\sim I,\qquad\Box_{b}S\sim S\Box_{b}\sim 0,\qquad S\sim S^{\ast}\sim S^{2},\qquad\overline{\partial}_{b}S\sim 0.

We first show some commutation relations of SS and S¯\overline{S}.

{lemma}

One has [S,¯b],[S¯,b]ΨH1(M)[S,\overline{\Box}_{b}],[\overline{S},\Box_{b}]\in\Psi_{H}^{1}(M).

Proof.

It follows from section 4 that

(4.2) S¯b=S(b1T)1TS1[S,T]¯bS1[S,T].S\overline{\Box}_{b}=S(\Box_{b}-\sqrt{-1}T)\sim-\sqrt{-1}TS-\sqrt{-1}[S,T]\sim\overline{\Box}_{b}S-\sqrt{-1}[S,T].

We obtain from section 3 that

(4.3) [S,¯b]1[S,T]ΨH1(M).[S,\overline{\Box}_{b}]\sim-\sqrt{-1}[S,T]\in\Psi_{H}^{1}(M).

Taking the complex conjugate yields [S¯,b]ΨH1(M)[\overline{S},\Box_{b}]\in\Psi_{H}^{1}(M). ∎

Next we consider the composition SS and S¯\overline{S}.

{lemma}

One has SS¯,S¯SΨH1(M)S\overline{S},\overline{S}S\in\Psi_{H}^{-1}(M).

Proof.

It follows from section 4 that

(4.4) ΔbSS¯¯bSS¯=S¯bS¯+[¯b,S]S¯[¯b,S]S¯ΨH1(M)\Delta_{b}S\overline{S}\sim\overline{\Box}_{b}S\overline{S}=S\overline{\Box}_{b}\overline{S}+[\overline{\Box}_{b},S]\overline{S}\sim[\overline{\Box}_{b},S]\overline{S}\in\Psi_{H}^{1}(M)

Since Δb\Delta_{b} has a parametrix, we have SS¯ΨH1(M)S\overline{S}\in\Psi_{H}^{-1}(M). Taking the complex conjugate gives S¯SΨH1(M)\overline{S}S\in\Psi_{H}^{-1}(M). ∎

{remark}

If MM is embeddable and SS is the Szegő projection, then we have SS¯S¯S0S\overline{S}\sim\overline{S}S\sim 0; see [Hsiao2015]*Lemma 4.2.

Let PP be the CR Paneitz operator on (M,T1,0M,θ)(M,T^{1,0}M,\theta). Since P=𝒞¯bP=\mathcal{C}\overline{\partial}_{b}, we have PS0PS\sim 0. Taking the complex conjugate yields PS¯0P\overline{S}\sim 0. Hence Π0S+S¯ΨH0(M)\Pi_{0}\coloneqq S+\overline{S}\in\Psi_{H}^{0}(M) satisfies PΠ00P\Pi_{0}\sim 0. Set G0NN¯ΨH4(M)G_{0}\coloneqq N\overline{N}\in\Psi_{H}^{-4}(M). It follows from sections 4 and 4 that

(4.5) PG0¯b(IS)N¯(IS)¯bN¯(IS)(IS¯)IΠ0PG_{0}\equiv\overline{\Box}_{b}(I-S)\overline{N}\equiv(I-S)\overline{\Box}_{b}\overline{N}\equiv(I-S)(I-\overline{S})\equiv I-\Pi_{0}

modulo ΨH1(M)\Psi_{H}^{-1}(M); note that 𝒬ΨH2(M)\mathcal{Q}\in\Psi_{H}^{2}(M). Thus we have

(4.6) R0PG0+Π0IΨH1(M).R_{0}\coloneqq PG_{0}+\Pi_{0}-I\in\Psi_{H}^{-1}(M).
{lemma}

The operator I+R0ΨH0(M)I+R_{0}\in\Psi_{H}^{0}(M) has a parametrix A0ΨH0(M)A_{0}\in\Psi_{H}^{0}(M). Moreover, A0A_{0} satisfies A0IΨH1(M)A_{0}-I\in\Psi_{H}^{-1}(M).

Proof.

Since

(4.7) I(I+R0)I=(I+R0)II=R0ΨH1(M),I(I+R_{0})-I=(I+R_{0})I-I=R_{0}\in\Psi_{H}^{-1}(M),

I+R0I+R_{0} has a parametrix A0A_{0} by section 3. We obtain from R0ΨH1(M)R_{0}\in\Psi_{H}^{-1}(M) and section 3 that

(4.8) σ0(A0)=σ0((I+R0)A0)=σ0(I),\sigma_{0}(A_{0})=\sigma_{0}((I+R_{0})A_{0})=\sigma_{0}(I),

which means A0IΨH1(M)A_{0}-I\in\Psi_{H}^{-1}(M). ∎

The proof of the following proposition is inspired by that of [Beals-Greiner1988]*Proposition 25.4.

{proposition}

There exist ΠΨH0(M)\Pi_{\infty}\in\Psi_{H}^{0}(M) and GΨH4(M)G_{\infty}\in\Psi_{H}^{-4}(M) such that

(4.9) GP+ΠPG+Π=I,\displaystyle G_{\infty}P+\Pi_{\infty}\sim PG_{\infty}+\Pi_{\infty}=I,
(4.10) GG,ΠΠ2Π,ΠΠ0ΨH1(M),\displaystyle G_{\infty}^{\ast}\sim G_{\infty},\qquad\Pi_{\infty}^{\ast}\sim\Pi_{\infty}^{2}\sim\Pi_{\infty},\qquad\Pi_{\infty}-\Pi_{0}\in\Psi_{H}^{-1}(M),
(4.11) ΠPPΠ0,ΠGGΠ0.\displaystyle\Pi_{\infty}P\sim P\Pi_{\infty}\sim 0,\qquad\Pi_{\infty}G_{\infty}\sim G_{\infty}\Pi_{\infty}\sim 0.
Proof.

Let A0ΨH0(M)A_{0}\in\Psi_{H}^{0}(M) be a parametrix of I+R0I+R_{0}, and set

(4.12) G(IΠ0A0)G0A0ΨH4(M),ΠIPGΨH0(M).G_{\infty}\coloneqq(I-\Pi_{0}A_{0})G_{0}A_{0}\in\Psi_{H}^{-4}(M),\qquad\Pi_{\infty}\coloneqq I-PG_{\infty}\in\Psi_{H}^{0}(M).

First note that

(4.13) Π=I(PPΠ0A0)G0A0(I+R0PG0)A0Π0A0.\Pi_{\infty}=I-(P-P\Pi_{0}A_{0})G_{0}A_{0}\sim(I+R_{0}-PG_{0})A_{0}\sim\Pi_{0}A_{0}.

In particular, PΠPΠ0A00P\Pi_{\infty}\sim P\Pi_{0}A_{0}\sim 0 and ΠΠ0ΨH1(M)\Pi_{\infty}-\Pi_{0}\in\Psi_{H}^{-1}(M). These yield

(4.14) Π=Π(PG+Π)ΠΠ(GP+Π)Π=Π.\displaystyle\Pi_{\infty}^{\ast}=\Pi_{\infty}^{\ast}(PG_{\infty}+\Pi_{\infty})\sim\Pi_{\infty}^{\ast}\Pi_{\infty}\sim(G_{\infty}^{\ast}P+\Pi_{\infty}^{\ast})\Pi_{\infty}=\Pi_{\infty}.

We also have

(4.15) ΠGΠ(IΠ)G0A0=(ΠΠ2)G0A00\Pi_{\infty}G_{\infty}\sim\Pi_{\infty}(I-\Pi_{\infty})G_{0}A_{0}=(\Pi_{\infty}-\Pi_{\infty}^{2})G_{0}A_{0}\sim 0

and

(4.16) GΠ=(GP+Π)GΠ\displaystyle G_{\infty}\Pi_{\infty}=(G_{\infty}^{\ast}P+\Pi_{\infty}^{\ast})G_{\infty}\Pi_{\infty} (GPG+ΠG)Π\displaystyle\sim(G_{\infty}^{\ast}PG_{\infty}+\Pi_{\infty}G_{\infty})\Pi_{\infty}
(4.17) G(IΠ)Π\displaystyle\sim G_{\infty}^{\ast}(I-\Pi_{\infty})\Pi_{\infty}
(4.18) =G(ΠΠ2)\displaystyle=G_{\infty}^{\ast}(\Pi_{\infty}-\Pi_{\infty}^{2})
(4.19) 0.\displaystyle\sim 0.

Therefore

(4.20) GG(IΠ)=GPG(GP+Π)G=G,G_{\infty}^{\ast}\sim G_{\infty}^{\ast}(I-\Pi_{\infty})=G_{\infty}^{\ast}PG_{\infty}\sim(G_{\infty}^{\ast}P+\Pi_{\infty})G_{\infty}=G_{\infty},

which completes the proof. ∎

Consider PP as an unbounded closed operator acting on L2(M)L^{2}(M) by the maximal closed extension. The domain DomP\operatorname{Dom}P contains WH4(M)W_{H}^{4}(M) by section 3. Conversely, any uDomPu\in\operatorname{Dom}P is an element of WH4(M)W_{H}^{4}(M) modulo RanΠ\operatorname{Ran}\Pi_{\infty} by the lemma below.

{lemma}

For uDomPu\in\operatorname{Dom}P, one has uΠuWH4(M)u-\Pi_{\infty}u\in W_{H}^{4}(M). In particular, DomP=RanΠ+WH4(M)\operatorname{Dom}P=\operatorname{Ran}\Pi_{\infty}+W_{H}^{4}(M).

Proof.

Set

(4.21) RGP+ΠIΨH(M).R_{\infty}\coloneqq G_{\infty}P+\Pi_{\infty}-I\in\Psi_{H}^{-\infty}(M).

If v=PuL2(M)v=Pu\in L^{2}(M), then

(4.22) uΠu=GvRuWH4(M).u-\Pi_{\infty}u=G_{\infty}v-R_{\infty}u\in W_{H}^{4}(M).

In particular, uRanΠ+WH4(M)u\in\operatorname{Ran}\Pi_{\infty}+W_{H}^{4}(M). Moreover, PΠ0P\Pi_{\infty}\sim 0 implies RanΠDomP\operatorname{Ran}\Pi_{\infty}\subset\operatorname{Dom}P. Therefore we have DomP=RanΠ+WH4(M)\operatorname{Dom}P=\operatorname{Ran}\Pi_{\infty}+W_{H}^{4}(M). ∎

Proof of section 1.

It suffices to show that PP is symmetric. Let u,vDomPu,v\in\operatorname{Dom}P. It follows from section 4 that vvΠvv^{\prime}\coloneqq v-\Pi_{\infty}v is in WH4(M)W_{H}^{4}(M). Since PΠ0P\Pi_{\infty}\sim 0,

(4.23) (Pu,Πv)0=(ΠPu,v)0=((PΠ)u,v)0=(u,PΠv)0.\lparen Pu,\Pi_{\infty}v\rparen_{0}=\lparen\Pi_{\infty}^{\ast}Pu,v\rparen_{0}=\lparen(P\Pi_{\infty})^{\ast}u,v\rparen_{0}=\lparen u,P\Pi_{\infty}v\rparen_{0}.

Take a sequence (vj)(v_{j}) in C(M)C^{\infty}(M) such that vjv_{j} converges to vv^{\prime} in WH4(M)W_{H}^{4}(M) as j+j\to+\infty. Then PvjPv_{j} converges to PvPv^{\prime} in L2(M)L^{2}(M) as j+j\to+\infty by the continuity of P:WH4(M)L2(M)P\colon W_{H}^{4}(M)\to L^{2}(M). Thus we have

(4.24) (Pu,v)0=limj(Pu,vj)0=limj(u,Pvj)0=(u,Pv)0.\lparen Pu,v^{\prime}\rparen_{0}=\lim_{j\to\infty}\lparen Pu,v_{j}\rparen_{0}=\lim_{j\to\infty}\lparen u,Pv_{j}\rparen_{0}=\lparen u,Pv^{\prime}\rparen_{0}.

Therefore (Pu,v)0=(u,Pv)0\lparen Pu,v\rparen_{0}=\lparen u,Pv\rparen_{0}, which means that PP is symmetric. ∎

Let EE be the resolution of the identity for PP and fix λ0\lambda\geq 0. Set

(4.25) πλE([λ,λ]):L2(M)DomP.\pi_{\lambda}\coloneqq E([-\lambda,\lambda])\colon L^{2}(M)\to\operatorname{Dom}P.

This is an orthogonal projection of L2(M)L^{2}(M) and satisfies πλP=Pπλ\pi_{\lambda}P=P\pi_{\lambda} on DomP\operatorname{Dom}P. If λ>0\lambda>0,

(4.26) Nλt1χ[λ,λ]c(t)𝑑E(t):L2(M)DomPN_{\lambda}\coloneqq\int_{\mathbb{R}}t^{-1}\chi_{[-\lambda,\lambda]^{c}}(t)\,dE(t)\colon L^{2}(M)\to\operatorname{Dom}P

is a continuous self-adjoint operator and satisfies

(4.27) PNλ+πλ=I on L2(M),\displaystyle PN_{\lambda}+\pi_{\lambda}=I\text{\ on \ }L^{2}(M),
(4.28) NλP+πλ=I on DomP.\displaystyle N_{\lambda}P+\pi_{\lambda}=I\text{\ on \ }\operatorname{Dom}P.
{theorem}

If λ>0\lambda>0, then Pπλ0P\pi_{\lambda}\sim 0.

Proof.

We first show that PπλP\pi_{\lambda} defines a continuous map from L2(M)L^{2}(M) to WH4k(M)W_{H}^{4k}(M) for any k0k\in\mathbb{Z}_{\geq 0}. It follows from PG+Π=IPG_{\infty}+\Pi_{\infty}=I that

(4.29) Pπλ=(GP+Π)Pπλ=G(Pπλ)2+ΠPπλ.P\pi_{\lambda}=(G_{\infty}^{\ast}P+\Pi_{\infty}^{\ast})P\pi_{\lambda}=G_{\infty}^{\ast}(P\pi_{\lambda})^{2}+\Pi_{\infty}^{\ast}P\pi_{\lambda}.

Since ΠPΠP0\Pi_{\infty}^{\ast}P\sim\Pi_{\infty}P\sim 0, the second term maps L2(M)L^{2}(M) to WH4k(M)W_{H}^{4k}(M) continuously for any k0k\in\mathbb{Z}_{\geq 0}. If PπλP\pi_{\lambda} maps L2(M)L^{2}(M) to WH4k(M)W_{H}^{4k}(M) continuously, then G(Pπλ)2G_{\infty}^{\ast}(P\pi_{\lambda})^{2} maps L2(M)L^{2}(M) to WH4k+4(M)W_{H}^{4k+4}(M) continuously. Hence Pπλ:L2(M)WH4k+4(M)P\pi_{\lambda}\colon L^{2}(M)\to W_{H}^{4k+4}(M) is continuous. We obtain by induction that PπλP\pi_{\lambda} is a continuous operator from L2(M)L^{2}(M) to WH4k(M)W_{H}^{4k}(M) for any k0k\in\mathbb{Z}_{\geq 0}. Taking the adjoint yields that PπλP\pi_{\lambda} extends to a continuous operator from WH4l(M)W_{H}^{-4l}(M) to L2(M)L^{2}(M) for any l0l\in\mathbb{Z}_{\geq 0}.

We next show that πλ\pi_{\lambda} extends to a continuous operator acting on WH4l(M)W_{H}^{-4l}(M) for any l0l\in\mathbb{Z}_{\geq 0}. It follows from PNλ+πλ=IPN_{\lambda}+\pi_{\lambda}=I that

(4.30) Π=ΠPNλ+Ππλ.\Pi_{\infty}^{\ast}=\Pi_{\infty}^{\ast}PN_{\lambda}+\Pi_{\infty}^{\ast}\pi_{\lambda}.

Similarly, PG+Π=IPG_{\infty}+\Pi_{\infty}=I yields that

(4.31) πλ=GPπλ+Ππλ.\pi_{\lambda}=G_{\infty}^{\ast}P\pi_{\lambda}+\Pi_{\infty}^{\ast}\pi_{\lambda}.

Thus we have

(4.32) πλΠ=GPπλΠPNλ,\pi_{\lambda}-\Pi_{\infty}^{\ast}=G_{\infty}^{\ast}P\pi_{\lambda}-\Pi_{\infty}^{\ast}PN_{\lambda},

which maps L2(M)L^{2}(M) to WH4k(M)W_{H}^{4k}(M) continuously for any k0k\in\mathbb{Z}_{\geq 0}. Taking the adjoint gives that πλΠ\pi_{\lambda}-\Pi_{\infty} extends to a continuous operator from WH4l(M)W_{H}^{-4l}(M) to L2(M)L^{2}(M) for any l0l\in\mathbb{Z}_{\geq 0}. Since Π\Pi_{\infty} defines a continuous operator acting on WH4l(M)W_{H}^{-4l}(M), so does πλ\pi_{\lambda}. This and eq. 4.29 imply that PπλP\pi_{\lambda} extends to a continuous operator from WH4l(M)W_{H}^{-4l}(M) to WH4k(M)W_{H}^{4k}(M) for any k,l0k,l\in\mathbb{Z}_{\geq 0}, which means that it is a smoothing operator. ∎

Proof of section 1.

Fix λ>0\lambda>0. Then SpecP[λ,λ]=SpecPπλ\operatorname{Spec}P\cap[-\lambda,\lambda]=\operatorname{Spec}P\pi_{\lambda} and

(4.33) Ker(PμI)=Ker(PπλμI)\operatorname{Ker}(P-\mu I)=\operatorname{Ker}(P\pi_{\lambda}-\mu I)

for any 0<|μ|λ0<\lvert\mu\rvert\leq\lambda. On the other hand, it follows from section 4 that PπλP\pi_{\lambda} is a compact self-adjoint operator acting on L2(M)L^{2}(M), and so SpecPπλ\operatorname{Spec}P\pi_{\lambda} is discrete except 0. Moreover, Ker(PπλμI)\operatorname{Ker}(P\pi_{\lambda}-\mu I) is a finite dimensional subspace of C(M)C^{\infty}(M) for any μ0\mu\neq 0. These completes the proof. ∎

{theorem}

If λ>0\lambda>0, then πλ\pi_{\lambda} and NλN_{\lambda} are Heisenberg pseudodifferential operators of order 0 and 4-4 respectively. Moreover, πλ\pi_{\lambda} and NλN_{\lambda} coincide with Π\Pi_{\infty} and GG_{\infty} respectively modulo ΨH(M)\Psi_{H}^{-\infty}(M).

Proof.

It follows from NλP+πλ=IN_{\lambda}P+\pi_{\lambda}=I that

(4.34) NλPΠ+πλΠ=Π.N_{\lambda}P\Pi_{\infty}+\pi_{\lambda}\Pi_{\infty}=\Pi_{\infty}.

Taking the adjoint yields that

(4.35) ΠPNλ+Ππλ=Π.\Pi_{\infty}^{\ast}PN_{\lambda}+\Pi_{\infty}^{\ast}\pi_{\lambda}=\Pi_{\infty}^{\ast}.

Thus we have

(4.36) (ΠΠπλ)(ΠπλΠ)=ΠPNλ2PΠ,(\Pi_{\infty}^{\ast}-\Pi_{\infty}^{\ast}\pi_{\lambda})(\Pi_{\infty}-\pi_{\lambda}\Pi_{\infty})=\Pi_{\infty}^{\ast}PN_{\lambda}^{2}P\Pi_{\infty},

which is a smoothing operator. On the other hand,

(4.37) (ΠΠπλ)(ΠπλΠ)\displaystyle(\Pi_{\infty}^{\ast}-\Pi_{\infty}^{\ast}\pi_{\lambda})(\Pi_{\infty}-\pi_{\lambda}\Pi_{\infty})
(4.38) =(Π+GPπλπλ)(Π+PπλGπλ)\displaystyle=(\Pi_{\infty}^{\ast}+G_{\infty}^{\ast}P\pi_{\lambda}-\pi_{\lambda})(\Pi_{\infty}+P\pi_{\lambda}G_{\infty}-\pi_{\lambda})
(4.39) =ΠΠ+ΠPπλGΠπλ+GPπλΠ\displaystyle=\Pi_{\infty}^{\ast}\Pi_{\infty}+\Pi_{\infty}^{\ast}P\pi_{\lambda}G_{\infty}-\Pi_{\infty}^{\ast}\pi_{\lambda}+G_{\infty}^{\ast}P\pi_{\lambda}\Pi_{\infty}
(4.40) +G(Pπλ)2GGPπλ2πλΠ+PπλG+πλ2\displaystyle\quad+G_{\infty}^{\ast}(P\pi_{\lambda})^{2}G_{\infty}-G_{\infty}^{\ast}P\pi_{\lambda}^{2}-\pi_{\lambda}\Pi_{\infty}+P\pi_{\lambda}G_{\infty}+\pi_{\lambda}^{2}
(4.41) ΠΠπλπλΠ+πλ\displaystyle\sim\Pi_{\infty}-\Pi_{\infty}^{\ast}\pi_{\lambda}-\pi_{\lambda}\Pi_{\infty}+\pi_{\lambda}
(4.42) =Π+GPπλπλ+PπλGπλ+πλ\displaystyle=\Pi_{\infty}+G_{\infty}^{\ast}P\pi_{\lambda}-\pi_{\lambda}+P\pi_{\lambda}G_{\infty}-\pi_{\lambda}+\pi_{\lambda}
(4.43) Ππλ.\displaystyle\sim\Pi_{\infty}-\pi_{\lambda}.

In particular, πλ\pi_{\lambda} is a Heisenberg pseudodifferential operator of order 0.

Next consider NλN_{\lambda}. It follows from PG+Π=IPG_{\infty}+\Pi_{\infty}=I that

(4.44) (Iπλ)G+NλΠ=Nλ,G(Iπλ)+ΠNλ=Nλ.(I-\pi_{\lambda})G_{\infty}+N_{\lambda}\Pi_{\infty}=N_{\lambda},\qquad G_{\infty}^{\ast}(I-\pi_{\lambda})+\Pi_{\infty}^{\ast}N_{\lambda}=N_{\lambda}.

Hence

(4.45) NλG\displaystyle N_{\lambda}-G_{\infty} =Nλ(Ππλ)πλG\displaystyle=N_{\lambda}(\Pi_{\infty}-\pi_{\lambda})-\pi_{\lambda}G_{\infty}
(4.46) =(Ππλ)Nλ(Ππλ)+G(Iπλ)(Ππλ)\displaystyle=(\Pi_{\infty}^{\ast}-\pi_{\lambda})N_{\lambda}(\Pi_{\infty}-\pi_{\lambda})+G_{\infty}^{\ast}(I-\pi_{\lambda})(\Pi_{\infty}-\pi_{\lambda})
(4.47) +(Ππλ)GΠG,\displaystyle\quad+(\Pi_{\infty}-\pi_{\lambda})G_{\infty}-\Pi_{\infty}G_{\infty},

which is a smoothing operator. Therefore NλN_{\lambda} is a Heisenberg pseudodifferential operator of order 4-4. ∎

5. CR Paneitz operator on the Rossi sphere

In this section, we prove the existence of infinitely many negative eigenvalues of the CR Paneitz operator on the Rossi sphere. Our proof is inspired by that of [Abbas-Brown-Ramasami-Zeytuncu2019]*Theorem 5.7.

5.1. Definition of the Rossi sphere

The unit sphere

(5.1) S3{(z,w)2||z|2+|w|2=1}S^{3}\coloneqq\Set{(z,w)\in\mathbb{C}^{2}}{\lvert z\rvert^{2}+\lvert w\rvert^{2}=1}

has the canonical CR structure T1,0S3T^{1,0}S^{3}. This CR structure is spanned by

(5.2) Z1w¯zz¯w.Z_{1}\coloneqq\overline{w}\frac{\partial}{\partial z}-\overline{z}\frac{\partial}{\partial w}.

A canonical contact form θ\theta on S3S^{3} is given by

(5.3) θ12(zdz¯+wdw¯z¯dzw¯dw)|S3.\theta\coloneqq\frac{\sqrt{-1}}{2}(zd\overline{z}+wd\overline{w}-\overline{z}dz-\overline{w}dw)|_{S^{3}}.

The Reeb vector field TT with respect to θ\theta is written as

(5.4) T=1(zz+wwz¯z¯w¯w¯).T=\sqrt{-1}\left\lparen z\frac{\partial}{\partial z}+w\frac{\partial}{\partial w}-\overline{z}\frac{\partial}{\partial\overline{z}}-\overline{w}\frac{\partial}{\partial\overline{w}}\right\rparen.

The admissible coframe corresponding to (T,Z1,Z1¯)(T,Z_{1},Z_{\overline{1}}) is given by

(5.5) (θ,θ1(wdzzdw)|S3,θ1¯θ1¯).(\theta,\theta^{1}\coloneqq(wdz-zdw)|_{S^{3}},\theta^{\overline{1}}\coloneqq\overline{\theta^{1}}).

For this coframe, we have

(5.6) dθ=1θ1θ1¯,dθ1=21θθ1.d\theta=\sqrt{-1}\theta^{1}\wedge\theta^{\overline{1}},\qquad d\theta^{1}=2\sqrt{-1}\theta\wedge\theta^{1}.

These imply

(5.7) l11¯=1,A11=0,ω11=21θ,Scal=2.l_{1\overline{1}}=1,\qquad A_{11}=0,\qquad{\omega_{1}}^{1}=-2\sqrt{-1}\theta,\qquad\mathrm{Scal}=2.

In particular,

(5.8) b=l11¯(Z1Z1¯ω1¯1¯(Z1)Z1¯)=Z1Z1¯.\Box_{b}=-l^{1\overline{1}}(Z_{1}Z_{\overline{1}}-{\omega_{\overline{1}}}^{\overline{1}}(Z_{1})Z_{\overline{1}})=-Z_{1}Z_{\overline{1}}.

For a real number 0<|t|<10<\lvert t\rvert<1, the Rossi sphere (St3,T1,0St3)(S^{3}_{t},T^{1,0}S^{3}_{t}) is defined by

(5.9) (St3,T1,0St3)(S3,(Z1+tZ1¯)).(S^{3}_{t},T^{1,0}S^{3}_{t})\coloneqq(S^{3},\mathbb{C}(Z_{1}+tZ_{\overline{1}})).

The complex vector field

(5.10) Z1(t)Z1+tZ1¯Z_{1}(t)\coloneqq Z_{1}+tZ_{\overline{1}}

gives a global frame of T1,0St3T^{1,0}S^{3}_{t}. The admissible coframe with respect to (T,Z1(t),Z1¯(t)Z1(t)¯)(T,Z_{1}(t),Z_{\overline{1}}(t)\coloneqq\overline{Z_{1}(t)}) is of the form

(5.11) (θ,θ1(t)(1t2)1(θ1tθ1¯),θ1¯(t)θ1(t)¯).(\theta,\theta^{1}(t)\coloneqq(1-t^{2})^{-1}(\theta^{1}-t\theta^{\overline{1}}),\theta^{\overline{1}}(t)\coloneqq\overline{\theta^{1}(t)}).

For this coframe, we have

(5.12) dθ=1(1t2)θ1(t)θ1¯(t),\displaystyle d\theta=\sqrt{-1}(1-t^{2})\theta^{1}(t)\wedge\theta^{\overline{1}}(t),
(5.13) dθ1(t)=12(1+t2)1t2θθ1(t)+14t1t2θθ1¯(t).\displaystyle d\theta^{1}(t)=\sqrt{-1}\frac{2(1+t^{2})}{1-t^{2}}\theta\wedge\theta^{1}(t)+\sqrt{-1}\frac{4t}{1-t^{2}}\theta\wedge\theta^{\overline{1}}(t).

This implies

(5.14) l11¯(t)=1t2,A11(t)=14t(1t2)2,ω11(t)=12(1+t2)1t2θ;l_{1\overline{1}}(t)=1-t^{2},\qquad A^{11}(t)=\sqrt{-1}\frac{4t}{(1-t^{2})^{2}},\qquad{\omega_{1}}^{1}(t)=-\sqrt{-1}\frac{2(1+t^{2})}{1-t^{2}}\theta;

see [Chanillo-Chiu-Yang2012]*Proposition 2.5 for example.

{lemma}

The Kohn Laplacian b(t)\Box_{b}(t) and 𝒬(t)\mathcal{Q}(t) with respect to (St3,T1,0St3,θ)(S^{3}_{t},T^{1,0}S^{3}_{t},\theta) satisfy

(5.15) (1t2)b(t)=bt(Z1)2t(Z1¯)2+t2¯b,\displaystyle(1-t^{2})\Box_{b}(t)=\Box_{b}-t(Z_{1})^{2}-t(Z_{\overline{1}})^{2}+t^{2}\overline{\Box}_{b},
(5.16) (1t2)2𝒬(t)=4t(Z1¯)24t2(b+¯b)+4t3(Z1)2.\displaystyle(1-t^{2})^{2}\mathcal{Q}(t)=4t(Z_{\overline{1}})^{2}-4t^{2}(\Box_{b}+\overline{\Box}_{b})+4t^{3}(Z_{1})^{2}.
Proof.

It follows from the definition of the Kohn Laplacian and eq. 5.8 that

(5.17) (1t2)b(t)\displaystyle(1-t^{2})\Box_{b}(t) =(1t2)l11¯(t)(Z1(t)Z1¯(t)ω1¯1¯(t)(Z1(t))Z1¯(t))\displaystyle=-(1-t^{2})l^{1\overline{1}}(t)\left\lparen Z_{1}(t)Z_{\overline{1}}(t)-{\omega_{\overline{1}}}^{\overline{1}}(t)(Z_{1}(t))Z_{\overline{1}}(t)\right\rparen
(5.18) =Z1(t)Z1¯(t)\displaystyle=-Z_{1}(t)Z_{\overline{1}}(t)
(5.19) =bt(Z1)2t(Z1¯)2+t2¯b.\displaystyle=\Box_{b}-t(Z_{1})^{2}-t(Z_{\overline{1}})^{2}+t^{2}\overline{\Box}_{b}.

Similarly, (1t2)2𝒬(t)(1-t^{2})^{2}\mathcal{Q}(t) is given by

(5.20) (1t2)2𝒬(t)\displaystyle(1-t^{2})^{2}\mathcal{Q}(t) =1(1t2)2(Z1¯(t)+ω1¯1¯(t)(Z1¯(t)))(A1¯1¯(t)Z1¯(t))\displaystyle=\sqrt{-1}(1-t^{2})^{2}\left\lparen Z_{\overline{1}}(t)+{\omega_{\overline{1}}}^{\overline{1}}(t)(Z_{\overline{1}}(t))\right\rparen\left\lparen A^{\overline{1}\overline{1}}(t)Z_{\overline{1}}(t)\right\rparen
(5.21) =4t(Z1¯(t))2\displaystyle=4t(Z_{\overline{1}}(t))^{2}
(5.22) =4t(Z1¯)24t2(b+¯b)+4t3(Z1)2,\displaystyle=4t(Z_{\overline{1}})^{2}-4t^{2}(\Box_{b}+\overline{\Box}_{b})+4t^{3}(Z_{1})^{2},

which completes the proof. ∎

5.2. Spherical harmonics

In this subsection, we recall some facts on spherical harmonics, which give a good orthogonal decomposition of L2(S3)L^{2}(S^{3}). We denote by 𝒫p,q(2)\mathscr{P}_{p,q}(\mathbb{C}^{2}) the space of complex homogeneous polynomials of bidegree (p,q)(p,q) and by p,q(2)\mathscr{H}_{p,q}(\mathbb{C}^{2}) the space of harmonic f𝒫p,q(2)f\in\mathscr{P}_{p,q}(\mathbb{C}^{2}). The restriction map

(5.23) |S3:p,q(2)C(S3)|_{S^{3}}\colon\mathscr{H}_{p,q}(\mathbb{C}^{2})\to C^{\infty}(S^{3})

is injective since any fp,qf\in\mathscr{H}_{p,q} is harmonic. The image of p,q(2)\mathscr{H}_{p,q}(\mathbb{C}^{2}) under this map is written as p,q(S3)\mathscr{H}_{p,q}(S^{3}).

{lemma}

[c.f. [Abbas-Brown-Ramasami-Zeytuncu2019]*Propositions 2.1 and 2.2] The Hilbert space L2(S3)L^{2}(S^{3}) has the following orthogonal decomposition:

(5.24) L2(S3)=p,qp,q(S3).L^{2}(S^{3})=\bigoplus_{p,q}\mathscr{H}_{p,q}(S^{3}).

Moreover, dimp,q(S3)=p+q+1\dim\mathscr{H}_{p,q}(S^{3})=p+q+1; in particular, p,q(S3)\mathscr{H}_{p,q}(S^{3}) contains a non-zero element.

Moreover, any fp,q(S3)f\in\mathscr{H}_{p,q}(S^{3}) is a simultaneous eigenfunction of b\Box_{b} and ¯b\overline{\Box}_{b}.

{lemma}

[c.f. [Abbas-Brown-Ramasami-Zeytuncu2019]*Theorem 2.6] For any fp,q(S3)f\in\mathscr{H}_{p,q}(S^{3}), one has

(5.25) bf=(p+1)qf,¯bf=p(q+1)f.\Box_{b}f=(p+1)qf,\qquad\overline{\Box}_{b}f=p(q+1)f.

Furthermore, Z1Z_{1} and Z1¯Z_{\overline{1}} change the bidegree.

{lemma}

The vector fields Z1Z_{1} and Z1¯Z_{\overline{1}} map p,q(S3)\mathscr{H}_{p,q}(S^{3}) to p1,q+1(S3)\mathscr{H}_{p-1,q+1}(S^{3}) and p+1,q1(S3)\mathscr{H}_{p+1,q-1}(S^{3}) respectively.

Proof.

It follows from the definition of Z1Z_{1} and Z1¯Z_{\overline{1}} that

(5.26) Z1:𝒫p,q(2)𝒫p1,q+1(2),Z1¯:𝒫p,q(2)𝒫p+1,q1(2).Z_{1}\colon\mathscr{P}_{p,q}(\mathbb{C}^{2})\to\mathscr{P}_{p-1,q+1}(\mathbb{C}^{2}),\qquad Z_{\overline{1}}\colon\mathscr{P}_{p,q}(\mathbb{C}^{2})\to\mathscr{P}_{p+1,q-1}(\mathbb{C}^{2}).

Moreover, one can check that [Δ,Z1]=[Δ,Z1¯]=0[\Delta,Z_{1}]=[\Delta,Z_{\overline{1}}]=0, where Δ\Delta is the Euclidean Laplacian on 2\mathbb{C}^{2}. Hence Z1Z_{1} and Z1¯Z_{\overline{1}} map harmonic functions to those. ∎

5.3. Negative eigenvalues of the CR Paneitz operator

Set 𝒫(t)(1t2)2P(t)\mathcal{P}(t)\coloneqq(1-t^{2})^{2}P(t), where P(t)P(t) is the CR Paneitz operator with respect to (St3,T1,0St3,θ)(S^{3}_{t},T^{1,0}S^{3}_{t},\theta). It suffices to show that 𝒫(t)\mathcal{P}(t) has infinitely many negative eigenvalues.

Fix a positive integer kk and take v12k1,0(S3)v_{1}\in\mathscr{H}_{2k-1,0}(S^{3}) with v1L2=1\lVert v_{1}\rVert_{L^{2}}=1; the existence of such v1v_{1} follows from section 5.2. We set

(5.27) ck(l)(l2)(2kl+2)c_{k}(l)\coloneqq(l-2)(2k-l+2)

and

(5.28) vi(l=1i11ck(2l+1)ck(2l+2))(Z1)2i2v12k2i+1,2i2v_{i}\coloneqq\left\lparen\prod_{l=1}^{i-1}\frac{1}{\sqrt{c_{k}(2l+1)c_{k}(2l+2)}}\right\rparen(Z_{1})^{2i-2}v_{1}\in\mathscr{H}_{2k-2i+1,2i-2}

for each integer 2ik2\leq i\leq k. Note that

(5.29) ck(l)+ck(l+3)=ck(l+1)+ck(l+2)4.c_{k}(l)+c_{k}(l+3)=c_{k}(l+1)+c_{k}(l+2)-4.

In what follows, we use the convention vi=0v_{i}=0 for i0i\leq 0 or ik+1i\geq k+1.

{lemma}

The functions (vi)(v_{i}) satisfy

(5.30) bvi=ck(2i)vi,¯bvi=ck(2i+1)vi,\displaystyle\Box_{b}v_{i}=c_{k}(2i)v_{i},\qquad\overline{\Box}_{b}v_{i}=c_{k}(2i+1)v_{i},
(5.31) (Z1)2vi=ck(2i+1)ck(2i+2)vi+1,(Z1¯)2vi=ck(2i1)ck(2i)vi1.\displaystyle(Z_{1})^{2}v_{i}=\sqrt{c_{k}(2i+1)c_{k}(2i+2)}v_{i+1},\qquad(Z_{\overline{1}})^{2}v_{i}=\sqrt{c_{k}(2i-1)c_{k}(2i)}v_{i-1}.
Proof.

Since vi2k2i+1,2i2v_{i}\in\mathscr{H}_{2k-2i+1,2i-2}, section 5.2 implies the first and second equalities. The third one easily follows from the definition of viv_{i}. Moreover, if i2i\geq 2, then

(5.32) (Z1¯)2vi\displaystyle(Z_{\overline{1}})^{2}v_{i} =1ck(2i1)ck(2i)(Z1¯)2(Z1)2vi1\displaystyle=\frac{1}{\sqrt{c_{k}(2i-1)c_{k}(2i)}}(Z_{\overline{1}})^{2}(Z_{1})^{2}v_{i-1}
(5.33) =1ck(2i1)ck(2i)Z1¯¯b(Z1vi1)\displaystyle=-\frac{1}{\sqrt{c_{k}(2i-1)c_{k}(2i)}}Z_{\overline{1}}\overline{\Box}_{b}(Z_{1}v_{i-1})
(5.34) =ck(2i)ck(2i1)¯bvi1( Z1vi12k2i+2,2i3)\displaystyle=\frac{\sqrt{c_{k}(2i)}}{\sqrt{c_{k}(2i-1)}}\overline{\Box}_{b}v_{i-1}\qquad\text{($\because$ $Z_{1}v_{i-1}\in\mathscr{H}_{2k-2i+2,2i-3}$)}
(5.35) =ck(2i1)ck(2i)vi1,\displaystyle=\sqrt{c_{k}(2i-1)c_{k}(2i)}v_{i-1},

which completes the proof. ∎

Let VkV_{k} be the subspace of L2(S3)L^{2}(S^{3}) spanned by (vi)i=1k(v_{i})_{i=1}^{k}. This space is closed under b\Box_{b}, ¯b\overline{\Box}_{b}, (Z1)2(Z_{1})^{2}, and (Z1¯)2(Z_{\overline{1}})^{2}.

{lemma}

The family (vi)i=1k(v_{i})_{i=1}^{k} is an orthonormal basis of VkV_{k}.

Proof.

Since vi2k2i+1,2i2(S3)v_{i}\in\mathscr{H}_{2k-2i+1,2i-2}(S^{3}), it is sufficient to show viL2=1\lVert v_{i}\rVert_{L^{2}}=1 for 1ik1\leq i\leq k. We prove this by induction. It follows from the choice of v1v_{1} that v1L2=1\lVert v_{1}\rVert_{L^{2}}=1. Assume that viL2=1\lVert v_{i}\rVert_{L^{2}}=1 for some 1ik11\leq i\leq k-1. Then the integration by parts gives

(5.36) vi+1L22\displaystyle\lVert v_{i+1}\rVert_{L^{2}}^{2} =1ck(2i+1)ck(2i+2)S3vi+1((Z1)2vi¯)θdθ\displaystyle=\frac{1}{\sqrt{c_{k}(2i+1)c_{k}(2i+2)}}\int_{S^{3}}v_{i+1}\left\lparen\overline{(Z_{1})^{2}v_{i}}\right\rparen\,\theta\wedge d\theta
(5.37) =1ck(2i+1)ck(2i+2)S3((Z1¯)2vi+1)vi¯θdθ\displaystyle=\frac{1}{\sqrt{c_{k}(2i+1)c_{k}(2i+2)}}\int_{S^{3}}\left\lparen(Z_{\overline{1}})^{2}v_{i+1}\right\rparen\overline{v_{i}}\,\theta\wedge d\theta
(5.38) =S3|vi|2θdθ\displaystyle=\int_{S^{3}}\lvert v_{i}\rvert^{2}\,\theta\wedge d\theta
(5.39) =1,\displaystyle=1,

which completes the proof. ∎

It follows from the definition of the CR Paneitz operator and section 5.1 that

(5.40) 𝒫(t)vi\displaystyle\mathcal{P}(t)v_{i}
(5.41) =t2ck(2i3)ck(2i2)ck(2i1)ck(2i)vi2\displaystyle=t^{2}\sqrt{c_{k}(2i-3)c_{k}(2i-2)c_{k}(2i-1)c_{k}(2i)}v_{i-2}
(5.42) t(1+t2)(ck(2i2)+ck(2i+1))ck(2i1)ck(2i)vi1\displaystyle\quad-t(1+t^{2})(c_{k}(2i-2)+c_{k}(2i+1))\sqrt{c_{k}(2i-1)c_{k}(2i)}v_{i-1}
(5.43) +[(1+t2)2ck(2i)ck(2i+1)+t2(ck(2i2)ck(2i)+ck(2i+1)ck(2i+3))]vi\displaystyle\quad+\left[(1+t^{2})^{2}c_{k}(2i)c_{k}(2i+1)+t^{2}(c_{k}(2i-2)c_{k}(2i)+c_{k}(2i+1)c_{k}(2i+3))\right]v_{i}
(5.44) t(1+t2)(ck(2i)+ck(2i+3))ck(2i+1)ck(2i+2)vi+1\displaystyle\quad-t(1+t^{2})(c_{k}(2i)+c_{k}(2i+3))\sqrt{c_{k}(2i+1)c_{k}(2i+2)}v_{i+1}
(5.45) +t2ck(2i+1)ck(2i+2)ck(2i+3)ck(2i+4)vi+2;\displaystyle\quad+t^{2}\sqrt{c_{k}(2i+1)c_{k}(2i+2)c_{k}(2i+3)c_{k}(2i+4)}v_{i+2};

here we use the equality eq. 5.29. In particular, this yields that 𝒫(t)\mathcal{P}(t) maps VkV_{k} to itself. We would like to show 𝒫(t)|Vk\mathcal{P}(t)|_{V_{k}} has exactly one negative eigenvalue. We denote by 𝒫k(t)\mathcal{P}_{k}(t) the matrix representation of 𝒫(t)|Vk\mathcal{P}(t)|_{V_{k}} with respect to (vi)i=1k(v_{i})_{i=1}^{k}. Note that 𝒫k(t)\mathcal{P}_{k}(t) is a k×kk\times k Hermitian matrix since 𝒫(t)\mathcal{P}(t) is self-adjoint and (vi)i=1k(v_{i})_{i=1}^{k} is an orthonormal basis of VkV_{k}.

{proposition}

The Hermitian matrix 𝒫k(t)\mathcal{P}_{k}(t) has exactly one negative eigenvalue.

Proof.

We define an l×ll\times l submatrix 𝒫k,l(t)\mathcal{P}_{k,l}(t) of 𝒫k(t)\mathcal{P}_{k}(t) for each 1lk1\leq l\leq k by

(5.46) (𝒫k,l(t))i,j(𝒫k(t))i,j,(i,j=1,,l),(\mathcal{P}_{k,l}(t))_{i,j}\coloneqq(\mathcal{P}_{k}(t))_{i,j},\qquad(i,j=1,\dots,l),

and set ηk,l(t)det𝒫k,l(t)\eta_{k,l}(t)\coloneqq\det\mathcal{P}_{k,l}(t). In order to compute ηk,l(t)\eta_{k,l}(t), we apply some elementary row operations. Add inductively the ii-th row multiplied by (1+t2)ck(2i+2)/ck(2i+1)/t(1+t^{2})\sqrt{c_{k}(2i+2)/c_{k}(2i+1)}/t (resp. (ck(2i+2)ck(2i+4))/(ck(2i+1)ck(2i+3))-\sqrt{(c_{k}(2i+2)c_{k}(2i+4))/(c_{k}(2i+1)c_{k}(2i+3))}) to the (i+1)(i+1)-st row (resp. (i+2)(i+2)-nd row), which does not change the determinant. The resulting matrix is the upper triangular matrix given by

(5.47) (t2ck(3)ck(5)*t2ck(5)ck(7)0t2ck(2l+1)ck(2l+3))\begin{pmatrix}t^{2}c_{k}(3)c_{k}(5)&&&\text{\Huge{*}}\\ &t^{2}c_{k}(5)c_{k}(7)&&\\ &&\ddots&&\\ \text{\Huge{0}}&&&t^{2}c_{k}(2l+1)c_{k}(2l+3)\end{pmatrix}

Hence

(5.48) ηk,l(t)=t2lck(3)ck(2l+3)i=1l1ck(2i+3)2.\eta_{k,l}(t)=t^{2l}c_{k}(3)c_{k}(2l+3)\prod_{i=1}^{l-1}c_{k}(2i+3)^{2}.

In particular, ηk,l(t)\eta_{k,l}(t) is positive (resp. negative) if 1lk11\leq l\leq k-1 (resp. l=kl=k). Applying Cauchy’s interlace theorem inductively yields that 𝒫k,l(t)\mathcal{P}_{k,l}(t) has only positive eigenvalues for 1lk11\leq l\leq k-1 and 𝒫k,k(t)=𝒫k(t)\mathcal{P}_{k,k}(t)=\mathcal{P}_{k}(t) has exactly one negative eigenvalue. ∎

Proof of section 1.

As we proved above, there exists an eigenfunction 0fkVk0\neq f_{k}\in V_{k} of P(t)P(t) with negative eigenvalue for each positive integer kk. Since

(5.49) Vkp+q=2k1p,q(S3),L2(S3)=p,qp,q(S3),V_{k}\subset\bigoplus_{p+q=2k-1}\mathscr{H}_{p,q}(S^{3}),\qquad L^{2}(S^{3})=\bigoplus_{p,q}\mathscr{H}_{p,q}(S^{3}),

the family (fk)k=1(f_{k})_{k=1}^{\infty} is linearly independent. This yields that P(t)P(t) has infinitely many negative eigenvalues with multiplicity. On the other hand, it follows from section 1 that SpecP(t){0}\operatorname{Spec}P(t)\setminus\{0\} consists only of eigenvalues of finite multiplicity. Therefore P(t)P(t) has infinitely many negative eigenvalues without multiplicity. ∎

{remark}

The proof of section 5.3 implies that the kernel of the operator

(5.50) P(t):p+q=2k1p,qp+q=2k1p,qP(t)\colon\bigoplus_{p+q=2k-1}\mathscr{H}_{p,q}\to\bigoplus_{p+q=2k-1}\mathscr{H}_{p,q}

is equal to zero for any positive integer kk. In particular, any function annihilated by P(t)P(t) must be even. Hence the Schwartz kernel of the orthogonal projection π0(t)\pi_{0}(t) to KerP(t)\operatorname{Ker}P(t) has the singularity on

(5.51) {((z,w),±(z,w))S3×S3},\Set{((z,w),\pm(z,w))\in S^{3}\times S^{3}},

and so π0(t)\pi_{0}(t) can not be a Heisenberg pseudodifferential operator.

6. Concluding remarks

The author [Takeuchi2020-Paneitz]*Theorem 1.1 has proved that the CR Paneitz operator on any embeddable CR manifold is non-negative. On the other hand, we found that the CR Paneitz operator on the Rossi sphere has infinitely many negative eigenvalues in the previous section. It is natural to ask whether this phenomenon occurs in general non-embeddable CR manifolds.

{problem}

Does the CR Paneitz operator on any non-embeddable CR manifold have necessarily infinitely many negative eigenvalues?

Moreover, Hsiao [Hsiao2015]*Theorem 4.7 has shown that the CR Paneitz operator on any embeddable CR manifold has closed range; see also [Takeuchi2023-GJMS] for another proof via Heisenberg calculus. On the other hand, section 1 only asserts that there are infinitely many negative eigenvalues of P(t)P(t), and it does not make any claims about the distribution of those eigenvalues.

{problem}

Does the CR Paneitz operator on the Rossi sphere have closed range?

Using Mathematica for calculations, we find that the following holds for small values of kk:

(6.1) det(𝒫1(t)+3t2I)\displaystyle\det(\mathcal{P}_{1}(t)+3t^{2}I) =0,\displaystyle=0,
(6.2) det(𝒫2(t)+3t2I)\displaystyle\det(\mathcal{P}_{2}(t)+3t^{2}I) =36t2(1t2)2,\displaystyle=36t^{2}(1-t^{2})^{2},
(6.3) det(𝒫3(t)+3t2I)\displaystyle\det(\mathcal{P}_{3}(t)+3t^{2}I) =576t2(1t2)2(15+58t2+15t4),\displaystyle=576t^{2}(1-t^{2})^{2}(15+58t^{2}+15t^{4}),
(6.4) det(𝒫4(t)+3t2I)\displaystyle\det(\mathcal{P}_{4}(t)+3t^{2}I) =6480t2(1t2)2(1680+6549t2+15926t4+6549t6+1680t8),\displaystyle=6480t^{2}(1-t^{2})^{2}(1680+6549t^{2}+15926t^{4}+6549t^{6}+1680t^{8}),
(6.5) det(𝒫5(t)+3t2I)\displaystyle\det(\mathcal{P}_{5}(t)+3t^{2}I) =995328t2(1t2)2(44100+172683t2+422712t4+825970t6\displaystyle=995328t^{2}(1-t^{2})^{2}(44100+172683t^{2}+422712t^{4}+825970t^{6}
(6.6) +422712t8+172683t10+44100t12),\displaystyle\quad+422712t^{8}+172683t^{10}+44100t^{12}),
(6.7) det(𝒫6(t)+3t2I)\displaystyle\det(\mathcal{P}_{6}(t)+3t^{2}I) =4536000t2(1t2)2(95800320+376277184t2+924268539t4\displaystyle=4536000t^{2}(1-t^{2})^{2}(95800320+376277184t^{2}+924268539t^{4}
(6.8) +1815582548t6+3114137570t8+1815582548t10\displaystyle\quad+1815582548t^{6}+3114137570t^{8}+1815582548t^{10}
(6.9) +924268539t12+376277184t14+95800320t16).\displaystyle\quad+924268539t^{12}+376277184t^{14}+95800320t^{16}).

These calculations suggest that the determinant of 𝒫k(t)+3t2I\mathcal{P}_{k}(t)+3t^{2}I is non-negative for any 1<t<1-1<t<1. If this is true, the unique negative eigenvalue of 𝒫k(t)\mathcal{P}_{k}(t) (section 5.3) is bigger than or equal to 3t2-3t^{2}, and so the spectrum of P(t)P(t) has 0 as an accumulation point; in particular, the CR Paneitz operator on the Rossi sphere does not have closed range.

References