Home Mathematics On the role of embeddability in conformal pseudo-hermitian geometry
Article Open Access

On the role of embeddability in conformal pseudo-hermitian geometry

  • Andrea Malchiodi EMAIL logo
Published/Copyright: November 5, 2024

Abstract

In this article, we review some recent results about the role of embeddability in conformal CR (Cauchy-Riemann) geometry. We will show how this condition enters in the second variation of the pseudo-hermitian counterpart of the Einstein-Hilbert action, in the positivity of the pseudohermitian mass and in the extremality properties of the CR (Cauchy-Riemann) Yamabe quotient. Given also some explicit examples at hand, this aspect shows sharp differences compared to the Riemannian case.

MSC 2010: 32V20; 53C17; 32V30

1 Introduction

The Yamabe problem consists in finding on a Riemannian manifold a metric conformal to a background one such that its scalar curvature is constant. This represents a higher-dimensional counterpart of the classical uniformization problem in two dimensions, and it has been introduced in [45].

Let ( M , g ) be a Riemannian manifold of dimension n 3 , and that for simplicity, we assume closed, i.e. compact and without boundary. If one denotes the scalar curvature by R g and performs the conformal change of metric g ˜ = w 4 n 2 g with w positive on M , then the scalar curvature R g ˜ of g ˜ is given by

(1.1) L g w = R g ˜ w n + 2 n 2 ,

where L g stands for the conformal Laplacian: L g w = 4 ( n 1 ) n 2 Δ g w + R g w . This operator enjoys the covariance law

(1.2) L g ˜ ϕ = w n + 2 n 2 L g ( w ϕ ) ; ϕ C ( M ) .

By this formula, the Yamabe problem amounts to finding a positive solution of

(Y) L g u = R ¯ u n + 2 n 2 ,

for some R ¯ R , which can be interpreted as a Lagrange multiplier: noting that for u smooth one has

M 4 ( n 1 ) n 2 g u 2 + R g u 2 d μ g = M u L g u d μ g ,

one can search for critical points of the Yamabe quotient on W 1 , 2 ( M ) { u : M R u , u L 2 ( M ) } .

(1.3) Q g ( u ) = M 4 ( n 1 ) n 2 g u 2 + R g u 2 d μ g M u 2 * d μ g 2 2 * ; u W 1 , 2 ( M ) ,

where 2 * = 2 n n 2 = n + 2 n 2 + 1 . Such critical points, that one could also look for constrained to the unit sphere of L 2 * ( M ) and taken non-negative, are solutions of (Y). Given the conformal covariance of L g , Q g satisfies the invariance property Q g ˜ ( u ) = Q g ( w u ) if g ˜ = w 4 n 2 g .

The quantity Q g has a clear geometric interpretation: define the normalized Einstein-Hilbert action as the Riemannian functional given by

(1.4) ( g ) = M ( g ) Vol g ( M ) n 2 n M R g d μ g .

The factor Vol g ( M ) n 2 n makes ( g ) scaling invariant. Its critical points are Einstein metrics, but its second variation at those is strongly indefinite, which makes its variational analysis difficult in general. On the round sphere ( S n , g S n ) , in particular, it is known that conformal deformations of the metric give rise to non-negative second variations, while tt-variations of the metric, for which one takes out the conformal part and the invariance by diffeomorphisms, have negative second variation [42]. The study by Koiso [28] is one of the first in which a more systematic study of the second variation properties of ( g ) at Einstein metrics was initiated.

The relation of Q g to ( g ) is that if g ˜ = u 4 n 2 g , then Q g ( u ) = ( g ˜ ) , so while Einstein metrics are free critical points of ( g ) , those constrained to a fixed conformal class are Yamabe metrics. Even extremizing within a given conformal class is not easy, since the lack of compactness of the embedding W 1 , 2 ( M ) L 2 * ( M ) makes the extremization of Q g challenging, because minimizing sequences for Q g might a priori not converge. This is due to the presence of bubbling, which means concentration of the H 1 - or of the L 2 * -norms at arbitrarily small scales.

Problem (Y) was solved by considering the Yamabe quotient, which is defined as follows:

Y ( M , g ) inf u C ( M ; R + ) Q g ( u ) = inf g ˜ [ g ] ( g ˜ ) .

The aforementioned covariance of Q g implies that indeed Y ( M , g ) depends only on the conformal class [ g ] of g , and will be therefore denoted by Y ( M , [ g ] ) . Conformal classes [ g ] for which Y ( M , [ g ] ) > 0 (respectively, = 0 or < 0 ) are said to be of positive Yamabe class (respectively, of null or negative class). In the study by Trudinger [41], it was proven that for every n 3 , there exists ε n > 0 such that Y ( M , [ g ] ) is attained when Y ( M , [ g ] ) ε n , which holds in particular when the Yamabe class is negative or null.

This result was sharpened by Aubin [5], where it was proven that the infimum of Q g is attained whenever

(1.5) Y ( M , [ g ] ) < Y ( S n , [ g S n ] ) ,

where g S n denotes the round metric of the sphere. As ( S n , g S n ) is conformally equivalent to R n via stereographic projection, Y ( S n , [ g S n ] ) coincides up to a dimensional constant with the Sobolev constant of R n , whose extremals (classified also in [39]) decay at infinity like the fundamental solution of the Laplacian. Such profiles, after a proper dilation, can be glued using normal coordinates to any point p of M : the geometry of ( M , g ) affects the expansion of the Yamabe quotient depending on the scaling parameter. Being the decay of the Green’s function of R n faster in higher dimensions, the role of the geometry is more localized, and it is ruled by the Weyl tensor W g at p when n 6 . Instead, for n 5 (or when g is locally conformally flat), it is determined by the mass of ( M , g ) once it is conformally deformed by the Green’s function of L g with pole at p . Using the positive mass theorem from [36,37] and [34], the aforementioned inequality (1.5) was proven also for the latter cases in the study by Schoen [35].

The purpose of this article is to explore the counterparts of the aforementioned facts in Cauchy-Riemann (CR) and pseudohermitian geometry, which sometimes are surprisingly in sharp contrast to the Riemannian case. All the basic definitions will be recalled in Section 2, while here we will only list some main concepts. Pseudohermitian manifolds are ( 2 n + 1 ) -dimensional manifolds with an n -dimensional complex sub-bundle H of the complexified tangent bundle T C M of M , verifying H H ¯ = { 0 } , [ H , H ] H , and on which a complex rotation J acts. Letting H ( M ) = Re ( H H ¯ ) , there exists a one-form θ , the contact form, such that H ( M ) = ker θ . Classical examples are hypersurfaces of C n or the Heisenberg group H n . If the Levi form L θ ( W , Z ¯ ) i d θ ( W , Z ¯ ) is non-degenerate, pseudohermitian manifolds carry a natural connection along H ( M ) , the Tanaka-Webster connection [18,40,43]. The trace W J , θ of the corresponding curvature tensor is known as the Webster curvature, and represents a pseudohermitian counterpart of the scalar curvature. Furthermore, θ ( d θ ) n acts naturally as a volume form.

In the study of properties of CR geometry, we will stress a crucial difference between the cases n = 1 and higher due to the embeddable or non-embeddable character of the structures. By a well-known theorem by Boutet de Monvel [9], if a closed CR manifold M with n 2 is strictly pseudoconvex, i.e., when the Levi form is positive-definite), then it can be CR embedded in C N for some natural integer N . This fact is not always true when n = 1 [13], and in fact generically false for perturbations of the standard S 3 [10], causing differences in all counterparts of the above-mentioned results concerning the Riemannian case.

We begin considering the second variation of the counterpart of the normalized pseudohermitian Einstein-Hilbert action, which we define by

W ˜ ( J , θ ) = M θ ( d θ ) n Q 2 Q M W J , θ θ ( d θ ) n .

Here, Q = 2 n + 2 stands for the homogeneous dimension of the manifold M , which is equal to the topological dimension plus 1. We also note that since a local deformation of the contact structure can be achieved via contactomorphisms to the original one by a result in [21], it is sufficient for us to vary only J and θ .

Embedding S 2 n + 1 into C n , the sphere naturally inherits a pseudohermitian structure ( J 0 , θ 0 ) from the ambient, see Section 2 for the explicit definitions. It can be shown that first variations of W ˜ vanishes at ( S 2 n + 1 , J 0 , θ 0 ) , while for second-order ones (see Section 2 for explicit formulas), we have then the following result.

Theorem 1.1

[2] Consider the standard pseudohermitian structure ( J 0 , θ 0 ) on S 2 n + 1 . Then the second variation d θ 2 W ˜ ( J 0 , θ 0 ) with respect to volume-preserving deformations of θ is non-negative definite. Concerning the variation with respect to J, we also have that

  1. if n > 1 then d J 2 W ˜ ( J 0 , θ 0 ) is positive-definite;

  2. if n = 1 the second variation d J 2 W ˜ ( J 0 , θ 0 ) is non-negative definite along deformations that infinitesimally preserve embeddabillity and is negative-definite along their orthogonal ones.

For reasons of brevity, in the aforementioned statement, we did not characterize the deformations that infinitesimally preserve embeddability. These were described first in [8] and depend on a Fourier expansion of a suitable complex-valued function with respect to the standard coordinates ( z 1 , z 2 ) of C 2 , see Section 2 for more details.

We consider next the Yamabe problem on closed CR manifolds, whose study was initiated in previous studies [25,27]. Similarly as mentioned earlier, fixing the CR structure J , one wants to find a conformal change of contact form so that the Webster curvature becomes constant. As we will see, also in this case, a Yamabe-type quotient is fundamental to attack this problem.

Referring to (2.4) for the notation of the sub-Laplacian Δ b , it is well-known that the Webster curvature trasforms under a change of contact form: by the law

(1.6) L b u 2 n + 2 n Δ b u + W J , θ u = W J , θ ˜ u Q + 2 Q 2 ; θ ˜ = u 4 Q 2 θ ,

in complete analogy to (1.1). The operator on the left-hand side is the CR sublaplacian and and has the following covariance property under a conformal change of contact form

L ˜ b ( φ ) = u Q + 2 Q 2 L b ( u φ ) ; θ ˜ = u 4 Q 2 θ ,

see (1.2), where L ˜ b stands for the conformal sublaplacian relative to ( J , θ ˜ ) . The CR Yamabe equation then becomes

(W) L b u = W ¯ u Q + 2 Q 2 ; W ¯ R .

Similarly to the Riemannian case, solutions can be found by considering the CR Yamabe quotient, or Webster quotient, defined by

(1.7) Y ( M , J ) inf θ ˜ M R J , θ ˜ θ ˜ d θ ˜ M θ ˜ d θ ˜ 1 2 = 1 2 inf u > 0 M 2 n + 2 n b u 2 + W J , θ u 2 θ ( d θ ) n u L 2 Q 2 2 ,

where θ ˜ is any contact form which annihilates ξ = H ( M ) (the contact bundle) and where b u 2 stands for the squared sub-gradient of u , see Section 2.

Jerison and Lee [25] proved that Y ( M , J ) Y ( S 2 n + 1 , J 0 ) , and that if strict inequality holds, then the infimum is attained. They also showed that if n 2 and ( M , J ) is not locally spherical, which means that the Chern tensor does not vanish identically, then indeed one has strict inequality. This was done using an argument in the spirit of [5] and a classification result from [26] about extremal functions for the Sobolev quotient in the Heisenberg group, the flat model of CR manifold.

In lower dimension, it is again expected that a positive mass theorem might play a role since the extremal functions still behave like the fundamental solution of Δ b , and have a slower decay when n = 1 . A concept of pseudohermitian mass can be defined for a suitable class of aymptotically flat pseudohermitian manifolds manifolds. Such manifolds can be obtained in particular from conformal blow-ups of compact ones in positive Webster class. This means that the infimum in (1.7) is positive, which coincides with the positivity of the operator L b , which for n = 1 has the expression

4 Δ b u + W J , θ u .

Under the assumption Y ( J ) > 0 , we have that L b is invertible, so for any p M , there exists a Green’s function G p for which

( 4 Δ b + W J , θ ) G p = 16 δ p θ d θ .

One can show that in CR normal coordinates ( z , t ) (which are pseudo-hermitian coordinates for a proper conformal change θ θ ˆ , see Section 3) G p admits the following expansion:

G p = 1 2 π ρ 2 + A + O ( ρ ) ,

where A is some real constant and where we have set ρ 4 ( z , t ) = z 4 + t 2 , z C , t R , with ρ being the homogeneous norm on H 1 . We consider then the new pseudohermitian manifold with a blow-up of contact form

(1.8) N = ( M \ { p } , J , θ = G p 2 θ ˆ ) ,

where θ ˆ is a contact form used in the definition of the above coordinates. The asymptotics of G p near p determines the geomerty of N at infinity, which becomes the Heisengberg group H 1 in the limit. Using a divergence structure appearing in the variation of the integral of W J , θ on N (similarly to the Riemannian case), one can define a notion of mass for such manifolds, see Definition 3.4. It also turns out that this mass is proportional to the constant A , see Lemma 3.5. One then has the following positive mass theorem for the three-dimensional case, under the assumption of embeddability of the CR structure.

Theorem 1.2

[16,38] Let M be a smooth, strictly pseudoconvex three-dimensional compact and embeddable CR manifold such that Y ( J ) > 0 . Let p M and let θ be a blow-up of contact form as in (1.8). Then

  1. m ( J , θ ) 0 ;

  2. if m ( J , θ ) = 0 , M is CR equivalent (or, together with θ ˆ , isomorphic as pseudohermitian manifold) to S 3 , endowed with its standard CR structure (and its standard contact form).

The aforementioned theorem was proved by Cheng et al. [16] for manifolds with non-negative CR Paneitz operator, see Proposition 3.6, using crucially a result from the study by Hsiao and Yung [24]. The fact that embeddability implies non-negativity of the Paneitz operator was shown by Takeuchi [38]. For more relations between these two conditions, see the studies by Case et al. [11] and Chanillo et al. [12]. In higher dimension, we are only aware of the results in Cheng et al. [15] and Cheng and Chiu [14], and it would be desirable to have more general ones.

The positivity of the mass allows to prove the attainment of Y ( M , J ) .

Corollary 1.3

[16] Suppose we are under the aforementioned assumptions and that the manifold is not conformally equivalent to the standard S 3 . Then Y ( M , J ) < Y ( S 3 , J 0 ) is attained.

Even though the resolution of the CR Yamabe problem was known before from [19] using the theory of critical points at infinity from [6] (see also [20] for the locally spherical case), there was no previous result about minimal solutions in three dimensions. More recently, a compactness result for solutions of the CR Yamabe equation was derived by Afeltra [1] for n = 1 concerning embeddable manifolds, and it is an open problem to have higher-dimensional counterparts.

We discuss next the sharpness of the above assumptions by exhibiting examples in striking contrast to the Riemannian case. One is given by Rossi spheres [33]: these are a one-parameter-family of CR structures on the 3-sphere of the form S s 3 ( S 3 , J ( s ) , θ 0 ) , where θ 0 and J ( 0 ) = J 0 are the standard contact form and CR structure of S 3 and J ( s ) is characterized by

(1.9) J ( s ) Z 1 ( s ) = i Z 1 ( s ) ; Z 1 ( s ) = Z 1 + s 1 + s 2 Z 1 ¯ , Z 1 ¯ ( s ) = Z 1 ¯ + s 1 + s 2 Z 1 ,

with Z 1 being the vector field Z 1 = z ¯ 2 z 1 z ¯ 1 z 2 restricted to S 3 . Rossi spheres are interesting prototype example because they are homogeneous and non-embeddable CR structures on the three-sphere. One has indeed the following theorem.

Theorem 1.4

[17] For s small, s 0 , the pseudo-hermitian mass of the Rossi spheres S s 3 is negative. More precisely, one has the expansion

m s = 18 π s 2 + o ( s 2 ) f o r s 0 .

We saw before (in both low-dimensional Riemannian and CR cases) that positivity of the mass implies attainment of the Sobolev quotient. We also strengthen the relation by means of the following result, again in sharp contrast with the Riemannian case.

Theorem 1.5

[17] For s small, s 0 , the infimum of the CR-Sobolev quotient of S s 3 coincides with Y ( S 3 , J S 3 ) and is not attained.

It is an open problem to prove counterparts of the last two theorems for more general perturbations of the standard S 3 or to provide other classes of such examples.

The plan of the article is the following. In Section 2, we discuss the variation of Webster curvature with respect to variations of pseudo-hermitian structure and give an idea for the proof of Theorem 1.1. In Section 3, we introduce the notion of pseudo-hermitian mass and sketch the proof of Theorem 1.4. In Section 4, we discuss instead the proofs of Theorems 1.4 and 1.5.

2 Second variation of the normalized Einstein-Hilbert action

In this section, after recalling some useful basic material on pseudohermitian manifolds, as well as some calculation concerning the variation of the contact form or of the CR structure, we display first- and second-variation formulas for the Webster curvature. About the forthcoming review material, we refer the reader to [18] and [30], and for the rest of the calculations to [2].

A CR manifold is a real smooth manifold M endowed with a complex sub-bundle H = T 1 , 0 M of the complexified tangent bundle of M , T C M , such that H H ¯ = { 0 } and [ H , H ] H . We take M to be of hypersurface type, that is, dim M = 2 n + 1 and dim C H = n . Let H ( M ) denote Re ( H H ¯ ) (we use sometimes the notation ξ ). Then there exists a natural complex structure on H ( M ) given by

J ( Z + Z ¯ ) = i ( Z Z ¯ ) .

The CR structure is uniquely determined by H ( M ) and J . For H ( M ) and J to generate a CR structure, it is necessary that H is closed under the Lie bracket operation. In three dimension, H is one dimensional, so the condition [ H , H ] H automatically holds.

There also exists a non-zero real differential form θ whose kernel at every point coincides with H ( M ) ; this is unique up to scalar multiplication by a non-zero function. A triple ( M , J , θ ) as above is called a pseudohermitian structure. On a pseudohermitian manifold, the Levi form is defined as follows:

L θ ( V , W ¯ ) = i d θ ( V , W ¯ ) = i θ ( [ V , W ¯ ] ) .

A CR manifold is said to be strictly pseudoconvex (respectively, non-degenerate) if it admits a positive definite (respectively, non-degenerate) Levi form. Non-degeneracy is equivalent to the fact that θ is a contact form (see Proposition 1.9 and formula (1.66) in [18]). In this case, we have a unique vector field T such that i T d θ = 0 and θ ( T ) = 1 . For example, if z 1 , , z n + 1 are complex coordinates on C n + 1 , then on the unit sphere S 2 n + 1 standard choices are given by the formulas:

θ = θ 0 = i 2 α = 1 n + 1 ( z α d z ¯ α z ¯ α d z α ) ; T = T 0 = i 2 α = 1 n + 1 z α z α z ¯ α z ¯ α ,

with J = J 0 being the restriction of the ambient complex rotation to H ( S 2 n + 1 ) . As mentioned in Section 1, classical examples of pseudohermitian manifolds are the Heisenberg group or boundaries of pseudoconvex domains in complex spaces.

On a non-degenerate pseudohermitian manifold, there is a connection, known as the Tanaka-Webster connection. To define it, we recall some useful facts, mostly from [30]. If { T , Z α , Z α ¯ } is a frame dual to { θ , θ α , θ α ¯ } , we express the Levi form as follows:

L θ ( f α Z α , g β ¯ Z β ¯ ) = h α β ¯ f α g β ¯ .

The matrix h α β ¯ = δ α β will be used in a standard way to raise and lower indices. The Webster connection forms ω α β and the torsion forms τ β = A β α θ α are defined by the following equations:

(2.1) d θ β = θ α ω α β + θ τ β ; ω α β ¯ + ω β ¯ α = d h α β ¯ ; A α β = A β α .

Also, the curvature forms

Π α β = d ω α β ω α γ ω γ β

satisfy the structure equations

Π α β = R α ρ σ ¯ β θ ρ θ σ ¯ + W α γ β θ γ θ W α γ ¯ β θ γ ¯ θ + i θ α τ β i τ α θ β .

The Ricci tensor and the pseudohermitian scalar (or Webster) curvature are defined by the contractions:

R ρ σ ¯ = R α ρ σ ¯ α ; W = R α α .

The covariant differentiation is characterized by the following equations:

(2.2) Z α = ω α β Z β ; Z α ¯ = ω α ¯ β ¯ Z β ¯ ; T = 0 .

For a tensor S with components S , we will use the notation

S , α ( Z α S ) ; S , α ¯ ( Z α ¯ S ) .

We also have the following commutation rules for second-order covariant derivatives of functions u and ( 1 , 0 ) -forms σ = σ α θ α , see Lemma 2.3 in [30]:

u , α β ¯ u , β ¯ α = i h α β ¯ u , 0 ; u , α β = u , β α ; u , 0 α u , α 0 = A α β u , β ; σ α , β γ σ α , γ β = i A α γ σ β i A α β σ γ ; σ α , β ¯ γ ¯ σ α , γ ¯ β ¯ = i h α β ¯ A γ ¯ ρ ¯ σ ρ ¯ i h α γ ¯ A β ¯ ρ ¯ σ ρ ¯ ; σ α , β γ ¯ σ α , γ ¯ β = i h β γ ¯ σ α , 0 + R α β γ ¯ ρ σ ρ ; σ α , 0 β σ α , β 0 = σ α , γ A γ β σ γ A α β , γ ; σ α , 0 β σ α , β 0 = σ α , γ A β γ + σ γ A β , α γ .

As a consequence of Bianchi’s identities (see Lemma 2.2 in [30]), we have in particular that

(2.3) A α β , γ = A α γ , β for all indices α , β , γ .

The sub-Laplacian of a scalar function u C ( M ) is defined as (see formula (4.10) in [29])

(2.4) Δ b u = u , α α + u , α ¯ α ¯ ,

while the squared subgradient is given by

b u 2 = 2 u , α u , α .

Given the contact bundle ξ , consider a smooth family t J ( t ) of CR structures on ξ . Then, for all values of t , J ( t ) : ξ ξ satisfies J ( t ) 2 = I d . Take a basis of eigenvectors ( Z α ( t ) ) α such that J ( t ) Z α ( t ) = i Z α ( t ) , which implies for the conjugate vector fields J ( t ) Z ¯ α ( t ) = i Z ¯ α ( t ) . In this way, J ( t ) writes as

J ( t ) = i θ ( t ) α Z α ( t ) i θ ( t ) α ¯ Z α ¯ ( t ) .

Differentiating with respect to t the relation J ( t ) 2 = I d and using the integrability conditions

(2.5) θ ( [ Z α , Z β ] ) = 0 ; θ γ ¯ ( [ Z α , Z β ] ) = 0 ,

which hold along all the deformation, one finds the following result.

Lemma 2.1

Setting J ˙ J ˙ ( t ) = d d t J ( t ) , the following expression holds true

J ˙ = 2 E = 2 E α β ¯ θ α Z β ¯ + c o n j .

Moreover, for all indices α , β , γ , we have that

E α β = E β α ; E α , β γ ¯ = E β , α γ ¯ .

Differentiating the structure equations, one also finds:

Proposition 2.2

For all t, the variation of the torsion is given by

(2.6) A ˙ γ ¯ α = i E γ ¯ , 0 α + A l ¯ α F γ ¯ l ¯ F l α A γ ¯ l ,

where F α β is determined by

Z ˙ α = F α β Z β i E α β ¯ Z β ¯ ,

while for the variation of the connection, we have

ω ˙ β α = [ i ( A γ ¯ α E β γ ¯ + E γ ¯ α A β γ ¯ ) + F β , 0 α ] θ + ( i E γ , α ¯ β ¯ F α ¯ , γ β ¯ ) θ γ + ( i E γ ¯ , β α + F β , γ ¯ α ) θ γ ¯ .

For the Webster curvature, we have the variation formula:

(2.7) W ˙ = R ˙ α α ¯ = i E l , γ ¯ l ¯ γ ¯ i E l ¯ , γ l γ ( A l γ ¯ E γ ¯ l + A γ ¯ l E l γ ¯ ) n + R l γ ¯ F γ l + R r γ ¯ F r ¯ γ ¯ .

Taking one more derivative in t of the latter formula, we also obtain also the following result.

Proposition 2.3

For the second variation of W = W ( t ) along the deformation J ( t ) , we have the following formula at t = 0

(2.8) W ¨ = i E ˙ l , γ ¯ l ¯ γ ¯ A l γ ¯ E ˙ γ ¯ l n + R l γ ¯ F ˙ γ l n A ˙ l γ ¯ E γ ¯ l E ρ ¯ l E ρ , γ ¯ l γ ¯ E γ ¯ l E ρ , l ρ ¯ γ ¯ E l γ ¯ E γ ¯ , ρ ρ ¯ l E ρ l ¯ E γ ¯ , γ ρ ¯ l E γ ¯ , ρ ¯ l E ρ , l γ ¯ E l , ρ ¯ γ ¯ E γ ¯ , ρ l E ρ , ρ ¯ l ¯ E γ ¯ , γ l E ρ ¯ , ρ l E l , γ ¯ γ ¯ + conj.

With this formula at hand, one is able to compute the second variation of W ˜ with respect to J . We refer to Proposition 3.3 in [18] for a general pointwise second variation formula regarding Webster’s curvature.

Lemma 2.4

For the standard structure ( S 2 n + 1 , J 0 , θ 0 ) , we have that

(2.9) d d t t = 0 W ˜ ( J ( t ) , θ 0 ) = 0 ; d 2 d t 2 t = 0 W ˜ ( J ( t ) , θ 0 ) = i n S 2 n + 1 E α γ ¯ , 0 E γ ¯ α θ 0 ( d θ 0 ) n + conj. ,

where E = 2 d d t t = 0 J ( t ) .

Proof

Since θ 0 and the volume remain fixed, we just need to integrate W ˙ and W ¨ with respect to the volume form θ 0 ( d θ 0 ) n . The vanishing of the first variation follows from (2.7) and an integration by parts, together with the fact that at t = 0 one can assume F = 0 , see Lemma 2.4 in [2].

Recalling (2.8), we first notice that the terms involving E ˙ and F ˙ vanish since they correspond to the first variation of W ˜ in the direction E ˙ , but we are at a stationary point of W ˜ .

Concerning the quadratic terms in E , we observe that after integrating and using Lemma 2.1, we obtain cancellation in (2.8) of the first with the seventh, of the second with the fifth, of the third with the sixth, and of the fourth with the eighth. We are then left with

W ˜ ¨ = n S 2 n + 1 ( A ˙ l γ ¯ E γ ¯ l + A ˙ γ ¯ l E l γ ¯ ) θ 0 ( d θ 0 ) n .

Recalling formula (2.6) and the fact that we can take F β α = 0 at t = 0 , we obtain the desired conclusion.□

We are now in position to discuss the proof of Theorem 1.1.

Proof of Theorem 1.1

We begin considering the second variation in θ . By a direct calculation, one finds on a general CR manifold that

(2.10) d 2 d t 2 t = 0 W ˜ ( J , ( 1 + t v ) 4 Q 2 θ ) = 2 V ol θ , u ( M ) Q 2 Q M b n b v 2 4 Q 2 W J , θ v 2 θ ( d θ ) n .

For the standard sphere ( S 2 n + 1 , J 0 , θ 0 ) , recalling from [26] and [43] that

(2.11) b n = 2 + 2 n ; W J 0 , θ 0 = n ( n + 1 ) ,

we obtain the first statement in the theorem [32].

We consider next the second variation of W ˜ with respect to J . For z 1 , , z n + 1 , the complex coordinates of C 2 n S 2 n + 1 , we define the subspace

(2.12) Γ m = { u C ( S 2 n + 1 ) u ( e i θ z 1 , e i θ z n ) = e i m θ u ( z 1 , , z n ) } .

It is possible to prove that E , 0 = i m 2 + 2 for E Γ m , which by Lemma 2.4 gives

(2.13) W ¨ = n m Z ( m + 4 ) S 2 n + 1 E ( m ) 2 θ 0 d ( θ 0 ) n .

where we have decomposed E in Fourier modes as

(2.14) E = m Z E ( m ) .

For n = 1 , it was proved by Bland [8] via a normal form that the perturbed structures which infinitesimally preserve embeddability are precisely characterized by having vanishing Fourier components E ( m ) for m 4 . Therefore, we obtain the conclusion for the case of S 3 .

For n > 1 , it is possible to prove that by Theorem 4.1 by Bland and Duchamp [7], the deformation E from Lemma 2.1 must satisfy

ϕ α γ ¯ = i E α γ ¯ ,

with ϕ of the form

(2.15) ϕ = ¯ b ( ¯ b f ) + h σ ( h ) .

Here f , h are a complex-valued function and a two-form of type ( 0 , 2 ) whose negative Fourier components are zero (with h determined by f ). Here, ¯ b denotes the holomorphic differential of f and the musical isomorphism from Ω 1 , 0 ( S 2 n + 1 ) to T 0 , 1 ( S 2 n + 1 ) . In analogy with (2.12), the mth Fourier eigenspace Γ m for a tensor on S 2 n + 1 is defined by the action of the flow generated by the vector field T .

Both the operators ¯ b and ( ¯ b ) commute with the Lie derivative by T , and it is noticed on page 102 of the study by Bland and Duchamp [7] that h σ preserves the Fourier decomposition. Therefore, the tensor ϕ , and hence E as well, only consist of non-negative Fourier modes. Since then E ( m ) = 0 for m < 0 in the decomposition (2.14), the conclusion follows from (2.13).□

Related results might be obtained for manifolds with special structure, like e.g. Sasakian. In view of the almost spherical embedding results in [22] (see Theorem 1.8), it might also be interesting to find extensions for more general CR manifolds.

3 Pseudo-hermitian mass

We consider here the case of three-dimensional pseudo-hermitian manifolds. We recall the notion of pseudohermitian normal coordinates, see [27] (Theorem 2.1), which are defined as follows. Given a vector V + c T R 2 × R H 1 T p M , consider the curve γ V , c solving the ordinary differential equation

γ ˙ γ ˙ = 2 c T ; γ ( 0 ) = p M , γ ˙ ( 0 ) = V .

The parabolic exponential map

Ψ ( V + c T ) = γ V , c ( 1 )

realizes a diffeomorphism from a neighbourhood of 0 in T p M to one of p in M . We also denote by

θ = d t + i z d z ¯ i z ¯ d z

the standard contact form on H 1 . We recall the following result [27, Proposition 2.5]. For a differential form η , let us denote by η ( m ) the part of its Taylor series that is homogeneous of degree m in terms of the parabolic dilations (see the latter reference for more details).

Proposition 3.1

Let Z ˜ 1 be a special frame dual to θ ˜ 1 (such that h ˜ 1 1 ¯ = 2 ). Let θ 1 = 2 θ ˜ 1 be a unitary coframe (i.e., h 1 1 ¯ = 1 ). Then in pseudohermitian normal coordinates ( z , t ) with respect to Z ˜ 1 , θ ˜ 1 , we have

  1. θ ( 2 ) = θ ; θ ( 3 ) = 0 ; θ ( m ) = 1 m 2 ( i z θ 1 ¯ i z ¯ θ 1 ) ( m ) , m 4 ;

  2. θ ( 1 ) 1 = 2 d z ; θ ( 2 ) 1 = 0 ; θ ( m ) 1 = 1 m ( 2 z ω 1 1 + 2 t A 1 ¯ 1 ¯ θ 1 ¯ 2 z ¯ A 1 ¯ 1 ¯ θ ) ( m ) , m 3 ;

  3. ( ω 1 1 ) ( 1 ) = 0 ; ( ω 1 1 ) ( m ) = 1 m ( 2 R ( z θ 1 ¯ z ¯ θ 1 ) + A 11 , 1 ¯ ( 2 z 2 t θ 1 ) A 1 ¯ 1 ¯ , 1 ( 2 z ¯ θ 2 t θ 1 ¯ ) ) ( m ) , m 2 .

By the results in Section 3 of the study by Jerison and Lee [27], we can find a contact form θ ˆ near p and local coordinates ( z , t ) such that

θ ˆ = ( 1 + O ( ρ 4 ) ) θ + O ( ρ 5 ) d z + O ( ρ 5 ) d z ¯ ; θ ˆ 1 = 2 ( 1 + O ( ρ 4 ) ) d z + O ( ρ 4 ) d z ¯ + O ( ρ 3 ) θ .

Consider next a CR manifold of positive Webster class: in this case, the operator L b is invertible and satisfies the maximum principle. For any p M , we have therefore a distributional solution of

( 4 Δ b + W J , θ ) G p = 16 δ p θ d θ ,

which is smooth and positive away from p . It can be proved that (see Proposition 5.2 in [16])

(3.1) G p = 1 2 π ρ 2 + A + O ( ρ ) near p ,

where ρ 4 = z 4 + t 2 and where 1 ρ 2 is (up to a multiple) the fundamental solution of Δ b in H 1 . By the aforementioned results, if we make the conformal change

(3.2) ( M \ { p } , J , θ G p 2 θ ˆ ) ,

then we obtain

(3.3) θ = G p 2 θ ˆ = 1 2 π ρ 2 + A + O ( ρ ) 2 [ ( 1 + O ( ρ 4 ) ) θ + O ( ρ 5 ) d z + O ( ρ 5 ) d z ¯ ] = 1 ( 2 π ) 2 ρ 4 + 2 1 2 π A ρ 2 + O ( ρ 1 ) [ ( 1 + O ( ρ 4 ) ) θ + O ( ρ 5 ) d z + O ( ρ 5 ) d z ¯ ] = 1 ( 2 π ) 2 ρ 4 + 2 1 2 π A ρ 2 + O ( ρ 1 ) θ + O ( ρ ) d z + O ( ρ ) d z ¯ ;

(3.4) θ 1 = G p ( θ ˆ 1 + 2 i ( log G p ) , 1 ¯ θ ˆ ) .

After a CR inversion of the type

(3.5) z * = z v ; t * = t v 2 ; on U \ { p } ,

where we have set v = t + i z 2 , and after a proper change of frame, we obtain a pseudo-hermitian structure as in the following definition.

Definition 3.2

A three-dimensional pseudohermitian manifold ( N , J , θ ) is said to be asymptotically flat pseudohermitian if N = N 0 N , with N 0 compact and N diffeomorphic to H 1 \ B ρ 0 in which ( J , θ ) is close to ( J 0 , θ ) in the sense that

θ = ( 1 + 4 π A ρ 2 + O ( ρ 3 ) ) θ + O ( ρ 3 ) d z + O ( ρ 3 ) d z ¯ ; θ 1 = O ( ρ 3 ) θ + O ( ρ 4 ) d z ¯ + ( 1 + 2 π A ρ 2 + O ( ρ 3 ) ) 2 d z

for some unitary coframe θ 1 and some A R in some system of coordinates (called asymptotic coordinates). We also require that W L 1 ( N ) .

Remark 3.3

The integrability of W is assumed because of the characterization of the pseudohermitian mass from the variation of the integral of W on N, similarly to [4]. In fact, with the conformal choice as in (3.2), by the transformation law (1.6) and by the definition of G p , we have that the Webster curvature of N vanishes identically.

After this definition, we are ready to introduce the notion of p-mass using a variational characterization, in the same spirit of [4]. Considering a one-parameter family of CR structures J ( s ) and using the notation of the previous section, we have that

J ˙ = 2 E = 2 E 11 θ 1 Z 1 ¯ + 2 E 1 ¯ 1 ¯ θ 1 ¯ Z 1 .

Denoting by W ( s ) the corresponding Tanaka-Webster curvature and using Proposition 2.2, we then find

d d s s = 0 N W ( s ) θ d θ = N W ˙ θ d θ = N [ i ( E 11 , 1 ¯ 1 ¯ E 1 ¯ 1 ¯ , 11 ) ( A 11 E 1 ¯ 1 ¯ + A 1 ¯ 1 ¯ E 11 ) ] d θ = N d ( E 11 , 1 θ θ 1 ) + conj. N ( A 11 E 1 ¯ 1 ¯ + conj. ) θ d θ = E 11 , 1 ¯ θ θ 1 + conj. N ( A 11 E 1 ¯ 1 ¯ + conj. ) θ d θ = i ø ˙ 1 1 θ N ( A 11 E 1 ¯ 1 ¯ + conj. ) θ d θ .

This formula leads us to the following definition of mass, requiring the condition

d d s s = 0 N W ( s ) θ d θ = m ˙ N ( A 11 E 1 ¯ 1 ¯ + conj. ) θ d θ ,

in analogy with the Riemannian case, see [31].

Definition 3.4

Let N be an asymptotically flat manifold. Then we define the p-mass of ( N , J , θ ) as follows:

m ( J , θ ) i ω 1 1 θ lim Λ + i S Λ ω 1 1 θ ,

where we have set S Λ = { ρ = Λ } in inverted CR coordinates.

First, we notice that m ( J , θ ) can be expressed in terms of the constant A appearing in Definition 3.2.

Lemma 3.5

If m ( J , θ ) is as in Definition 3.4 and if A is as in (3.1), then

m ( J , θ ) = 48 π 2 A .

Proof

It is possible to show that in inverted coordinates the connection form expands as follows:

(3.6) 6 π A z * ¯ ( z * 2 + i t * ) ρ * 6 + O ( ρ * 4 ) d z * + 6 π A z * ( z * 2 + i t * ) ρ * 6 + O ( ρ * 4 ) d z ¯ * + O ( ρ * 5 ) ( θ ) * .

By some elementary estimates, one finds

m ( J , θ ) = i ø 1 1 θ = 3 i 2 π A ρ 6 [ ( z 2 z ¯ + i z ¯ t ) d z ( z 2 z i z t ) d z ¯ ] θ = 6 i π A S z 2 ( z ¯ d z z d z ¯ ) θ 6 i π A S ( i z ¯ t d z + i z t d z ¯ ) θ ,

where we have set

S = { ρ = 1 } .

Using the relations

(3.7) z = r e i φ ; d z = e i φ ( d r + i r d φ ) ; z ¯ d z z d z ¯ = 2 i r 2 d φ ,

we obtain

m ( J , θ ) = 6 i π A S [ 2 i r 4 d φ + 2 i t r d r ] θ = 12 π A S [ r 4 d φ d t + t r ( d r d t + 2 r 2 d r d φ ) ] = 12 π A S ( r 4 d φ d t + t r d r d t + 2 t r 3 d r d φ ) .

On S we have that 4 r 3 d r + 2 t d t = 0 , so the last formula becomes

m ( J , θ ) = 12 π A S ( r 4 d φ d t t 2 d t d φ ) = 12 π A S d φ d t = 48 π 2 A .

Therefore, we obtain the conclusion.□

We derive next an integral formula for the p-mass by an approach somewhat inspired by [44], where a spinorial proof of the positive mass theorem was given. Here though the quantities involved are of higher order. To state the next result, with the notation for covariant differentiation introduced before, we let the operator Box-b be

b β = 2 β , 1 ¯ 1 .

Proposition 3.6

Let ( N , J , θ ) be an asymptotically flat pseudohermitian 3-manifold. Let β : N C be a C -smooth complex-valued function such that

(3.8) β = z ¯ + β 1 + O ( ρ 2 + ε ) n e a r ,

and

(3.9) b β = O ( ρ 4 ) ,

where β 1 is a term with the homogeneity of ρ 1 satisfying

(3.10) ( β 1 ) , 1 ¯ = 2 2 π A 1 ρ 2 2 A z 2 + i t

near infinity, for some ε ( 0 , 1 ) . Then one has

(3.11) 2 3 m ( J , θ ) = N b β 2 θ d θ + 2 N β , 1 ¯ 1 ¯ 2 θ d θ + 2 N W β , 1 ¯ 2 θ d θ + 1 2 N β ¯ P β θ d θ ,

where

P β ¯ b b β + 4 i ( A 11 β , 1 ¯ ) , 1 ¯

is the CR Paneitz operator.

The CR Paneitz operator satisfies the covariance property [23]

(3.12) P ( J , θ ˆ ) = u 4 P ( J , θ ) ; θ ˆ = u 2 .

Furthermore, one has the following result.

Theorem 3.7

[38] Let ( M , J ) be a closed embeddable strictly pseudoconvex CR manifold of dimension three. Then the CR Paneitz operator is non-negative.

It can be proved via some integration by parts that the positivity of P on the compact manifold M implies positivity of the last term in (3.11) as well, up to a multiple of m ( J , θ ) (which can be reabsorbed into the left-hand side). We have the following corollary.

Corollary 3.8

Let N be the blow-up of a three-dimensional embeddable CR manifold with positive CR Yamabe invariant. Let β be as in Proposition 3.6: then one has

(3.13) 4 3 m ( J , θ ) N b β 2 θ d + 2 N β , 1 ¯ 1 ¯ 2 θ d θ + 2 N R β , 1 ¯ 2 θ d θ .

Moreover, if θ is chosen as in (1.8) and β satisfies b β = 0 , then R = 0 and (3.13) is reduced to

(3.14) 4 3 m ( J , θ ) 2 N β , 1 ¯ 1 ¯ 2 θ d θ 0 .

By a deep theorem in [24], it is indeed possible to choose β as above so that b β = 0 , and therefore, from the above corollary, one deduces the non-negativity of the mass. It is an open problem to find solutions of b = 0 for general asymptotically flat pseudohermitian 3-manifolds.

The rigidity statement is obtained from the aforementioned integral formula for β , which implies that both the Webster curvature and the torsion vanish identically. By setting z = β ¯ , it is possible to prove the existence of a function t ˜ such that

d t ˜ = θ i z d z ¯ + i z ¯ d z .

In this way, we obtain a pseudohermitian isomorphism between a neighbourhood of infinity U in M and its image in H 1 , V (a neighbourhood of infinity in H 1 ), if we send q N into

q ( z ( q ) , t ( q ) ) = β ¯ ( q ) , q 0 q d t ˜ ,

where we are taking curves connecting q 0 to q inside U .

We call Ψ : V U (sets which we can assume to be connected by arcs) the inverse of this map, which cab ne extended to a covering map Ψ ˜ : H 1 N . Observe that V is contained in a fundamental domain. If Ψ ˜ is not 1-1, then there are at least two fundamental domains. But one of them has infinity volume, while any other one has finite volume. The contradiction shows Ψ ˜ is 1-1 and a pseudohermitian isomorphism between H 1 and N .

Corollary 1.3 is obtained by using a suitable test function whose construction follows the argument in the study by Schoen [35], with a proper gluing of the extremal function from [26] to the Green’s function G p .

4 Rossi spheres

We recall here some properties of Rossi spheres, introduced by Rossi et al. [33] as an example of non-embeddable CR manifold. These are a one-dimensional family of CR structures on S 3 , containing the standard one, obtained in the following way.

Considering the complex vector field Z 1 given by

Z 1 = z ¯ 2 z 1 z ¯ 1 z 2

and its conjugate Z 1 ¯ , one defines the CR structure J ( s ) so that satisfies J ( s ) Z 1 ( s ) = i Z 1 ( s ) , where

Z 1 ( s ) = Z 1 + s 1 + s 2 Z 1 ¯ , Z 1 ¯ ( s ) = Z 1 ¯ + s 1 + s 2 Z 1 .

Corresponding to these vector fields, we have the dual forms

θ ( s ) 1 = ( 1 + s 2 ) θ 1 s 1 + s 2 θ 1 ¯ , θ ( s ) 1 ¯ = ( 1 + s 2 ) θ 1 ¯ s 1 + s 2 θ 1 ,

where θ 1 = z 2 d z 1 z 1 d z 2 . By direct computation, one obtain

(4.1) i θ ( s ) 1 θ ( s ) 1 ¯ = ( 1 + s 2 ) i θ 1 θ 1 ¯ = ( 1 + s 2 ) d θ 0 ,

where d θ 0 = i θ 1 θ 1 ¯ , i.e., h 1 1 ¯ = 1 . Hence, from (4.1), we obtain

h 1 1 ¯ ( s ) = 1 1 + s 2 and h ( s ) 1 1 ¯ ( h 1 1 ¯ ( s ) ) 1 = 1 + s 2 .

By taking

θ ˜ ( s ) 1 = 1 2 ( 1 + s 2 ) θ ( s ) 1 ,

we have h ˜ 1 1 ¯ ( s ) = 2 . The Webster curvature W of ( J , θ 0 ) is identically equal to 2. We can then determine, from the structure equation for ( J ( s ) , θ ˆ ) , that

ω 1 ( s ) 1 = 2 i ( 1 + 2 s 2 ) θ ˆ , h ( s ) 1 1 ¯ A 1 ¯ 1 ¯ ( s ) = 4 i s 1 + s 2 , W ( s ) = 2 ( 1 + 2 s 2 ) .

The expression of θ 1 is

θ 1 = z 2 d z 1 z 1 d z 2 .

The sub-Laplacian associated to ( J ( s ) , θ ˆ ) reads as follows:

(4.2) b ( s ) = h ( s ) 1 1 ¯ ( Z 1 ( s ) Z 1 ¯ ( s ) + Z 1 ¯ ( s ) Z 1 ( s ) ) = ( 1 + 2 s 2 ) b ( 0 ) + 2 s 1 + s 2 ( Z 1 2 + Z 1 ¯ 2 ) .

It follows that, at s = 0 , the first- and second-order derivatives of b ( s ) with respect to s are given by

(4.3) Δ ˙ b = 2 Z 1 ¯ Z 1 ¯ + conj. ; Δ ¨ b = 4 Δ b .

Moreover, since W s = 2 ( 1 + 2 s 2 ) it follows that, still at s = 0

(4.4) W ˙ = 0 ; W ¨ = 8 .

We next analyze a symmetry property of Rossi spheres, which will imply in particular the symmetry of their mass in s (which we notice is independent of the blow-up point, by homogeneity of the structures). Consider the diffeomorphism ι : S 3 S 3 defined by

(4.5) ι ( z 1 , z 2 ) = ( i z 1 , z 2 ) ,

which fixes the point (0,1). A direct computation shows that ι Z 1 S 3 = i Z 1 S 3 and hence ι Z 1 ¯ S 3 = ( i ) Z 1 ¯ S 3 . By (4.1), we have

(4.6) ι Z 1 ( s ) = ι Z 1 + s 1 + s 2 ι Z 1 ¯ = i Z 1 + s 1 + s 2 ( i ) Z 1 ¯ = i Z 1 ( s ) .

It follows that

( ι J ( s ) ) Z 1 ( s ) = ι 1 J ( s ) ( ι Z 1 ( s ) ) = ι 1 J ( s ) ( i Z 1 ( s ) ) (by (4.6)) = ι 1 ( Z 1 ( s ) ) = ( 1 ) ( i ) Z 1 ( s ) ; (by the inverse of (4.6)) = i Z 1 ( s ) = J ( s ) Z 1 ( s ) .

Hence, we have shown

(4.7) J ( s ) = ι J ( s ) .

Let v ( s ) denote the conformal factor in θ ˇ ( s ) = e 2 v ( s ) θ ˆ , yielding CR normal coordinates with respect to J ( s ) , see Proposition 2.3 in [17]. It then follows that

v ( s ) = ι v ( s ) , θ ˇ ( s ) = ι θ ˇ ( s ) ,

and hence for the Green’s function G s with pole at p = ( 0 , 1 ) S 3 C 2 , we have that G ˇ s = ι G ˇ s by observing

(4.8) ι θ ˆ = θ ˆ .

Write

G ˇ s = 2 ρ s 2 + A s + O ( ρ s )

in s -CR normal coordinates near ( 0 , 1 ) . Then ρ s = ι ρ s = ρ s ι and

A s = ι A s = A s ι = A s

near the point ( 0 , 1 ) . So, we obtained from Lemma 3.5 that

m ( J ( s ) , θ ( s ) ) = m ( J ( s ) , θ ( s ) ) ,

where θ ( s ) = G ˇ ( s ) 2 θ ˇ ( s ) . This property is crucial in the role of the mass for the expansion of the Webster quotient uniformly in s sufficiently small.

Our next goal is to expand the conformal factor used to define CR normal coordinates on Rossi spheres. Fix p = ( 0 , 1 ) and consider a contact form θ ˇ ( s ) = e 2 v ( s ) θ ˆ , where θ ˆ = 2 θ 0 = i ( ¯ ) ( z 1 2 + z 2 2 ) yielding CR normal coordinates with respect to J ( s ) for N = 4 . Write (see again Proposition 2.3 in [17])

(4.9) v ( s ) = v 2 + v 3 + ,

where v 2 2 P 2 , v 3 P 3 . Recall that, in the notation of [27], P m denotes the vector space of polynomials in ( z , t ) that are homogeneous of degree m in terms of parabolic dilations (for which t has homogeneity 2), and m P m denotes the subspace of polynomials independent of t . One has the following expansion from Lemma 3.1 in Cheng et al. [17].

Lemma 4.1

In pseudo-hermitian coordinates, the conformal factor defining CR normal coordinates expands in homogeneous powers as as follows:

(4.10) v ( s ) = s 1 + s 2 ( z 2 + z ¯ 2 ) + 1 4 ( 1 + 2 s 2 ) z 2 + v 4 + .

From these results, one obtains an expansion of CR normal coordinates in terms of the ambient ones as follows:

(4.11) z ˜ CR = ( 1 + s 2 ) z 1 s 1 + s 2 z ¯ 1 + ( 1 + s 2 ) 3 2 s 2 1 4 z 1 w + s 1 + s 2 3 2 s 2 + 5 4 z ¯ 1 w + s ( 1 + s 2 ) 3 2 1 2 s 2 + 5 4 z 1 3 s 2 ( 1 + s 2 ) 1 2 s 2 3 4 z ¯ 1 3 + s 1 + s 2 3 2 s 4 + 3 4 s 2 1 z ¯ 1 2 z 1 + ( 1 + s 2 ) 3 2 s 4 9 4 s 2 + 1 4 z 1 2 z ¯ 1 + h.o.t.

One can also expand the contact from to find that

(4.12) t CR = i w ( 1 + 1 2 z 1 2 ) + i s z 1 2 ( z 1 2 z ¯ 1 2 ) + i s 2 ( z ¯ 1 4 z 1 4 ) + h.o.t.

Proposition 4.2

The CR normal coordinates on Rossi spheres with respect to θ ˇ = e 2 v θ ˆ are given by the formulas z CR = z ˜ CR 1 + s 2 , with z ˜ CR as in (4.11) and t CR as in (4.12).

We next perform a formal expansion of the Green’s function in powers of s with respect to the ambient coordinates of C 2 . Let L s denote the conformal sub-Laplacian for the J ( s ) -structure on S 3 . For s = 0 , the fundamental solution of L 0 G 0 = 64 π δ p with pole at p = ( 0 , 1 ) is given by

(4.13) G 0 = 2 ( ( 1 z 2 ) ( 1 z ¯ 2 ) ) 1 2 .

We aim to solve formally, up to an error O ( s 3 ) , L s G s = 0 away from p as a power series of s in the form

(4.14) G s = G 0 + s G 1 + 1 2 s 2 ( G 2 + α G 0 G 3 ) + o ( s 2 ) ,

where G 1 , G 2 are suitable explicit singular functions near p , α R and G 3 is a Hölder continuous function near p for which we would need to determine only G 3 ( p ) . We chose to expand the second-order term including separately α G 0 for practical convenience.

By using the first formula in (4.3), the first-order correction G 1 to G s can be chosen as follows:

(4.15) G 1 = 1 2 ( z 1 2 + z ¯ 1 2 ) 1 1 z 2 + 1 1 z ¯ 2 + 2 1 ( ( 1 z 2 ) ( 1 z ¯ 2 ) ) 1 2 .

We can invert L 0 explicitly for the terms with factors z 1 4 and z ¯ 1 4 . The solution is given by

G 2 , 1 ( z 1 4 + z ¯ 1 4 ) g 2 , 1 ( ( 1 z 2 ) ( 1 z ¯ 2 ) ) 5 2 ; G 2 , 2 3 4 g 2 , 2 ( ( 1 z 2 ) ( 1 z ¯ 2 ) ) 5 2 .

Define then

(4.16) G 2 = G 2 , 1 + G 2 , 2 ,

To obtain G 3 ( p ) , we use the Green’s representation formula, convoluting Ξ ( z 2 , z ¯ 2 ) with G 0 :

G 3 ( p ) = 1 64 π 2 S 3 24 ( z 2 1 ) 2 + ( z ¯ 2 1 ) 2 3 ( z 2 1 ) ( z ¯ 2 1 ) ( ( 1 z 2 ) ( 1 z ¯ 2 ) ) θ ˆ d θ ˆ .

The Taylor expansion of the integrand in z 2 , z ¯ 2 is

( 24 24 z ¯ 2 5 24 z ¯ 2 4 24 z ¯ 2 3 24 z ¯ 2 2 ) + ( 24 z ¯ 2 5 + 24 z ¯ 2 4 + 24 z ¯ 2 3 + 24 z ¯ 2 2 + 48 z ¯ 2 ) z 2 + ( 24 z ¯ 2 24 ) z 2 2 + ( 24 z ¯ 2 24 ) z 2 3 + ( 24 z ¯ 2 24 ) z 2 4 + ( 24 z ¯ 2 24 ) z 2 5 + .

Integrated over S 3 , this gives

S 3 ( 24 + 48 z 2 2 ) θ ˆ d θ ˆ = 482 π 2 + 96 π 2 = 192 π 2 ,

which implies that

(4.17) G 3 ( p ) = 3 .

Proposition 4.3

For every compact set K in S 3 \ { p } , p = ( 0 , 1 ) , there exists a constant C K > 0 such that the function G s G 0 + s G 1 + 1 2 s 2 ( G 2 + α G 0 G 3 ) in (4.14) satisfies

L s G s C K s 3 o n K .

Comparing the two expressions of the Green’s function from Proposition 4.3 and Lemma 4.2, one then obtains the expansion of the mass m s as in Theorem 1.4.

We next briefly discuss the proof of Theorem 1.5. In the study by Jerison and Lee [26], it was proven that Y ( S 3 , J S 3 ) is precisely attained by the following functions, up to composing ( z 1 , z 2 ) with elements of S U ( 2 )

(4.18) φ λ = λ ( z 1 2 + z 2 + 1 2 ) 2 ( z 2 z ¯ 2 ) 2 ( λ 2 z 1 2 + z 2 + 1 2 ) 2 λ 4 ( z 2 z ¯ 2 ) 2 1 2 ; λ > 0 .

Define also the family of minimizers of the Webster quotient on the standard S 3 :

(4.19) = { φ λ ( U ( ) ) : λ > 0 , U S U ( 2 ) } .

A first step in the proof is the following lemma, which can be obtained e.g. using a standard concentration-compactness principle.

Lemma 4.4

Fix s R , s small. Assume u s > 0 attains Y ( S 3 , J ( s ) ) . Then, if u s is normalized so that S 3 u s 4 θ 0 d θ 0 = 4 π 2 , up to a homogeneous action on S 3 there exists λ > 0 such that

u s φ λ S 1 , 2 ( S 3 ) = o s ( 1 ) ,

where o s ( 1 ) 0 as s 0 .

One first shows that the CR-Yamabe equation is always solvable, in a fixed neighbourhood of , up to a Lagrange multiplier: see [3] for a general reference on this method.

Proposition 4.5

For φ λ as in (4.18) there exists a unique w λ S 1 , 2 ( S 3 ) , depending smoothly on λ , such that w λ S 1 , 2 ( S 3 ) C s and which satisfies

(4.20) S 3 φ λ 2 φ λ λ w λ θ 0 d θ 0 = 0 ; L s ( φ λ + w λ ) 2 ( φ λ + w λ ) 3 = φ λ 2 φ λ λ

for some R , where C is a fixed constant. Moreover, there exists δ > 0 with the following property: if there exists a critical point of Q ( s ) in a δ -neighbourhood of (in S 1 , 2 norm), then it must be of the form φ λ + w λ up to a homogeneous action on S 3 and up to a scalar multiple, with w λ as above.

Recalling the latter statement in Proposition 4.5, we analyze the CR Sobolev quotient on functions of the form φ λ + w λ , showing that it is strictly higher than the standard spherical one. We first notice the following consequence of the discussion after (4.5).

Lemma 4.6

Let s > 0 be small, and let w λ ( s ) and w λ ( s ) denote the counterparts of w λ in Proposition 4.5 for s and s , respectively. Then one has that

Q ( s ) ( φ λ + w λ ( s ) ) = Q ( s ) ( φ λ + w λ ( s ) ) .

One can consider next two situations. The first is when the parameter λ in the previous lemma tends to infinity or to zero, and the second when log λ remains bounded. In the latter case, we will show that the CR Sobolev quotient would be strictly higher than Y ( S 3 , J S 3 ) , which would give a contradiction. On the other hand, we can also rule out the former case using the negativity of the mass of ( S 3 , J ( s ) ) for s small and non-zero. The proofs of the next two results can be found in the two appendices of [17].

Proposition 4.7

Let Λ > 1 be a fixed number. Then there exist C Λ > 0 such that, for λ [ 1 Λ , Λ ] and for s small one has Q ( s ) ( φ λ + w λ ) = 4 π + s 2 A λ + λ , s , where

A λ = 16 π λ 2 ( 3 + 12 λ 2 + 2 λ 4 + 12 λ 6 + 3 λ 8 ) ( 1 + λ 2 ) 6 ,

and where λ , s C Λ s 3 .

Notice that the minimizers in [26] stay unchanged when we compose with the antipodal map on S 3 and replace λ by 1 λ : this symmetry implies that A λ = A 1 λ . Therefore, in the next proposition, it is sufficient to consider large values of λ .

Proposition 4.8

The following expansion holds true, uniformly in s (small)

Q ( s ) ( φ λ + w λ ) = 4 π 8 3 m s λ 2 + O s 2 λ 3 = 4 π + 48 π s 2 λ 2 ( 1 + o s ( 1 ) ) + O s 2 λ 3 ,

for λ large.

Theorem 1.5 can then be proved as follows: assume by contradiction that u is a minimizer of the CR-Sobolev quotient Q ( s ) for s 0 small. By Lemma 4.4, u must then lie in a δ -neighbourhood of the manifold defined in (4.19). From the second part of Proposition 4.5, we have also that u = φ λ + w λ up to a homogeneous action on S 3 , where w λ is as in the first part of the Proposition. The conclusion then follows from Propositions 4.7 and 4.8, which cover all ranges of λ for s small enough.


Dedicated to Ermanno Lanconelli with admiration


  1. Funding information: The author was supported by the project Geometric problems with loss of compactness from Scuola Normale Superiore and by the PRIN Project 2022AKNSE4 “Variational and Analytical aspects of Geometric PDE.” He is also a member of GNAMPA as part of INdAM.

  2. Author contributions: The author confirms the sole responsibility for the conception of the study, presented results and manuscript preparation.

  3. Conflict of interest: The author states no conflict of interest.

References

[1] C. Afeltra, A Compactness Result for the CR Yamabe Problem in Three Dimensions, arXiv:2401.00906, 2023. 10.1142/S0219199725500038Search in Google Scholar

[2] C. Afeltra, J.-H. Cheng, A. Malchiodi, P. Yang, and X. Wang, On the variation of the Einstein-Hilbert action in pseudohermitian geometry, J. Reine Angew. Math. 813 (2024), 81–102. 10.1515/crelle-2024-0031Search in Google Scholar

[3] A. Ambrosetti and A. Malchiodi, Perturbation methods and semilinear elliptic problems on Rn, Progress in Mathematics, vol. 240, Birkhäuser Verlag, Basel, 2006. 10.1007/3-7643-7396-2Search in Google Scholar

[4] R. Arnowitt, S. Deser, and C. W. Misner, Dynamical structure and definition of energy in general relativity, Phys. Rev. 116 (1959), no. 2, 1322–1330. 10.1103/PhysRev.116.1322Search in Google Scholar

[5] T. Aubin, Équations différentielles non linéaires et problème de Yamabe concernant la courbure scalaire, J. Math. Pures Appl. 55 (1976), no. 3, 269–296. Search in Google Scholar

[6] A. Bahri and J.-M. Coron, On a nonlinear elliptic equation involving the critical Sobolev exponent: the effect of the topology of the domain, Comm. Pure Appl. Math. 41 (1988), no. 3, 253–294. 10.1002/cpa.3160410302Search in Google Scholar

[7] J. Bland and T. Duchamp, Moduli for pointed convex domains, Invent. Math. 104 (1991), no. 1, 61–112. 10.1007/BF01245067Search in Google Scholar

[8] J. S. Bland, Contact geometry and CR structures on S3, Acta Math. 172 (1994), no. 1, 1–49. 10.1007/BF02392789Search in Google Scholar

[9] L. Boutet de Monvel, Intégration des équations de Cauchy-Riemann induites formelles, Séminaire Goulaouic-Lions-Schwartz 1974-1975: Équations aux dérivées partielles linéaires et non linéaires, École Polytech., Paris, 1975, Exp. No. 9, 149. Search in Google Scholar

[10] D. M. Burns and C. L. Epstein, Embeddability for three-dimensional CR-manifolds, J. Amer. Math. Soc. 3 (1990), no. 4, 809–841. 10.1090/S0894-0347-1990-1071115-4Search in Google Scholar

[11] J. S. Case, S. Chanillo, and P. Yang, The CR Paneitz operator and the stability of CR pluriharmonic functions, Adv. Math. 287 (2016), 109–122. 10.1016/j.aim.2015.10.002Search in Google Scholar

[12] S. Chanillo, H.-L. Chiu, and P. Yang, Embeddability for 3-dimensional Cauchy-Riemann manifolds and CR Yamabe invariants, Duke Math. J. 161 (2012), no. 15, 2909–2921. 10.1215/00127094-1902154Search in Google Scholar

[13] S.-C. Chen and M.-C. Shaw, Partial differential equations in several complex variables, AMS/IP Studies in Advanced Mathematics, vol. 19, American Mathematical Society, Providence, RI; International Press, Boston, MA, 2001. 10.1090/amsip/019Search in Google Scholar

[14] J.-H. Cheng and H.-L. Chiu, Positive mass theorem and the CR Yamabe equation on 5-dimensional contact spin manifolds, Adv. Math. 404 (2022), Paper No. 108446, 50. 10.1016/j.aim.2022.108446Search in Google Scholar

[15] J.-H. Cheng, H.-L. Chiu, and P. Yang, Uniformization of spherical CR manifolds, Adv. Math. 255 (2014), 182–216. 10.1016/j.aim.2014.01.002Search in Google Scholar

[16] J.-H. Cheng, A. Malchiodi, and P. Yang, A positive mass theorem in three dimensional Cauchy-Riemann geometry, Adv. Math. 308 (2017), 276–347. 10.1016/j.aim.2016.12.012Search in Google Scholar

[17] J.-H. Cheng, A. Malchiodi, and P. Yang, On the Sobolev quotient of three-dimensional CR manifolds, Rev. Mat. Iberoam. 39 (2023), no. 6, 2017–2066. 10.4171/rmi/1412Search in Google Scholar

[18] S. Dragomir and G. Tomassini, Differential geometry and analysis on CR manifolds, Progress in Mathematics, vol. 246, Birkhäuser Boston, Inc., Boston, MA, 2006. Search in Google Scholar

[19] N. Gamara, The CR Yamabe conjecture—the case n=1, J. Eur. Math. Soc. (JEMS) 3 (2001), no. 2, 105–137. 10.1007/pl00011303Search in Google Scholar

[20] N. Gamara and R. Yacoub, CR Yamabe conjecture–the conformally flat case, Pacific J. Math. 201 (2001), no. 1, 121–175. 10.2140/pjm.2001.201.121Search in Google Scholar

[21] J. W. Gray, Some global properties of contact structures, Ann. Math. 69 (1959), no. 2, 421–450. 10.2307/1970192Search in Google Scholar

[22] H. Herrmann, C. Y. Hsiao, G. Marinescu, and W. C. Shen, Semi-classical spectral asymptotics of Toeplitz operators on CR manifolds, arXiv 2303.17319, 2023.Search in Google Scholar

[23] K. Hirachi, Scalar pseudo-hermitian invariants and the szego kernel on three-dimensional CR manifolds, Complex geometry (Osaka, 1990), Lecture Notes in Pure and Appl. Math., vol. 143, Dekker, New York, 1993, pp. 67–76. Search in Google Scholar

[24] C.-Y. Hsiao and P.-L. Yung, Solving the Kohn Laplacian on asymptotically flat CR manifolds of dimension 3, Adv. Math. 281 (2015), 734–822. 10.1016/j.aim.2015.04.028Search in Google Scholar

[25] D. Jerison and J. M. Lee, The Yamabe problem on CR manifolds, J. Differential Geom. 25 (1987), no. 2, 167–197. 10.4310/jdg/1214440849Search in Google Scholar

[26] D. Jerison and J. M. Lee, Extremals for the Sobolev inequality on the Heisenberg group and the CR Yamabe problem, J. Amer. Math. Soc. 1 (1988), no. 1, 1–13. 10.1090/S0894-0347-1988-0924699-9Search in Google Scholar

[27] D. Jerison and J. M. Lee, Intrinsic CR normal coordinates and the CR Yamabe problem, J. Differential Geom. 29 (1989), no. 2, 303–343. 10.4310/jdg/1214442877Search in Google Scholar

[28] N. Koiso, On the second derivative of the total scalar curvature, Osaka Math. J. 16 (1979), no. 2, 413–421. Search in Google Scholar

[29] J. M. Lee, The Fefferman metric and pseudo-Hermitian invariants, Trans. Amer. Math. Soc. 296 (1986), no. 1, 411–429. 10.1090/S0002-9947-1986-0837820-2Search in Google Scholar

[30] J. M. Lee, Pseudo-Einstein structures on CR manifolds, Amer. J. Math. 110 (1988), no. 1, 157–178. 10.2307/2374543Search in Google Scholar

[31] J. M. Lee and T. H. Parker, The Yamabe problem, Bull. Amer. Math. Soc. (N.S.) 17 (1987), no. 1, 37–91. 10.1090/S0273-0979-1987-15514-5Search in Google Scholar

[32] A. Malchiodi and F. Uguzzoni, A perturbation result for the Webster scalar curvature problem on the CR sphere, J. Math. Pures Appl. (9) 81 (2002), no. 10, 983–997. 10.1016/S0021-7824(01)01249-1Search in Google Scholar

[33] H. Rossi, Attaching analytic spaces to an analytic space along a pseudoconcave boundary, Proc. Conf. Complex Analysis (Minneapolis, 1964), Springer, Berlin-Heidelberg-New York, 1965, pp. 242–256. 10.1007/978-3-642-48016-4_21Search in Google Scholar

[34] R. Schoen and S.-T. Yau, Conformally flat manifolds, Kleinian groups and scalar curvature, Invent. Math. 92 (1988), no. 1, 47–71. 10.1007/BF01393992Search in Google Scholar

[35] R. Schoen, Conformal deformation of a Riemannian metric to constant scalar curvature, J. Differential Geom. 20 (1984), no. 2, 479–495. 10.4310/jdg/1214439291Search in Google Scholar

[36] R. Schoen and S. T. Yau, On the proof of the positive mass conjecture in general relativity, Comm. Math. Phys. 65 (1979), no. 1, 45–76. 10.1007/BF01940959Search in Google Scholar

[37] R. Schoen and S. T. Yau, Proof of the positive mass theorem. II, Comm. Math. Phys. 79 (1981), no. 2, 231–260. 10.1007/BF01942062Search in Google Scholar

[38] Y. Takeuchi, Nonnegativity of the CR Paneitz operator for embeddable CR manifolds, Duke Math. J. 169 (2020), no. 18, 3417–3438. 10.1215/00127094-2020-0051Search in Google Scholar

[39] G. Talenti, Best constant in Sobolev inequality, Ann. Mat. Pura Appl. 110 (1976), no. 4, 353–372. 10.1007/BF02418013Search in Google Scholar

[40] N. Tanaka, A differential geometric study on strongly pseudo-convex manifolds, Lectures in Mathematics, Department of Mathematics, Kyoto University, vol. 9, Kinokuniya Book Store Co., Ltd., Tokyo, 1975. Search in Google Scholar

[41] N. S. Trudinger, Remarks concerning the conformal deformation of Riemannian structures on compact manifolds, Ann. Scuola Norm. Sup. Pisa Cl. Sci. 22 (1968), no. 3, 265–274. Search in Google Scholar

[42] J. A. Viaclovsky, Critical metrics for Riemannian curvature functionals, Geometric analysis, IAS/Park City Math. Series, vol. 22, American Mathematical Society, Providence, RI, 2016, pp. 197–274. 10.1090/pcms/022/05Search in Google Scholar

[43] S. M. Webster, Pseudo-Hermitian structures on a real hypersurface, J. Differential Geometry 13 (1978), no. 1, 25–41. 10.4310/jdg/1214434345Search in Google Scholar

[44] E. Witten, A new proof of the positive energy theorem, Comm. Math. Phys. 80 (1981), no. 3, 381–402. 10.1007/BF01208277Search in Google Scholar

[45] H. Yamabe, On a deformation of Riemannian structures on compact manifolds, Osaka Math. J. 12 (1960), 21–37. Search in Google Scholar

Received: 2024-03-23
Accepted: 2024-09-18
Published Online: 2024-11-05

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 5.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/agms-2024-0011/html
Scroll to top button