Complex Analysis and CR Geometry
Complex Analysis and CR Geometry
Giuseppe Zampieri
To my children Caterina, Sebastiano and Emanuele Zampieri
Preface vii
Chapter 1. Several Complex Variables 1
1.1. Analytic functions in complex spaces 1
1.2. Cauchy formula in polydiscs 5
1.3. Analytic functions and power series 11
1.4. Subharmonic functions 14
1.5. Separate analyticity 22
1.6. Levi forms—continuity principle (I)—Hartogs extension
theorem 24
1.7. Logarithmic supermean of Taylor radius of holomorphic
functions—continuity principle (II)—propagation of
holomorphic extendibility 30
1.8. Domains of holomorphy and pseudoconvex domains 39
1.9. L2 -estimates for ∂¯ on q-pseudoconvex domains of Cn 51
1.10. Subelliptic estimates for ∂¯ 68
Chapter 2. Real Structures 77
2.1. Euclidean spaces and their diffeomorphisms 77
2.2. Integration of vector fields and vector bundles—Frobenius
theorem 83
2.3. Real symplectic spaces—Frobenius-Darboux theorem 90
2.4. Subelliptic estimates and hypoellipticity of systems of
vector fields 96
2.5. Miscellanea: foliations—orbits 102
Chapter 3. Real/Complex Structures 109
3.1. Complex structures—real underlying structures—
complexifications 109
3.2. CR manifolds 115
3.3. CR functions and CR mappings 125
3.4. The Levi form of a submanifold M ⊂ Cn and an abstract
CR structure 134
3.5. Real/complex symplectic spaces 138
3.6. Approximation of CR functions by polynomials 146
v
vi CONTENTS
vii
viii PREFACE
f : R2n → C, we have
(1.1.8) ∂x f dx + ∂y f dy = ∂z f dz + ∂z̄ f dz̄
that we also write as
(1.1.9) ¯
df = ∂f + ∂f.
Equation (1.1.9) serves as a definition ¯ . Note that in
of ∂f and ∂f
P
(1.1.8) ∂x f dx + ∂y f dy stands for j ∂xj f dxj + ∂yj f dyj and ∂z f dz +
P
∂z̄ f dz̄ stands for j ∂zj f dzj + ∂z̄j f dz̄j . Equation (1.1.8) states that
the system of derivatives ∂z and ∂z̄ is the dual system to the system of
differentials dz and dz̄. It is often used, in alternative to (1.1.4), (1.1.5),
to define ∂z and ∂z̄ from dz and dz̄.
√
Remark 1.1.1. The length 2 of dzj and dz̄j , which is imposed by
the notation, forces the dual system ∂zj and ∂z̄j to have the unnatural
length √12 ; this is why the chain rule is preferable to duality in their
definition.
We have to notice that in case f is real valued then (1.1.8) yields
(1.1.10) df = 2Re ∂f.
There is no doubt that (1.1.10) is great: it tells us that the “real” dif-
ferentials are the real parts of the “complex” differentials. It serves the
intrinsic version of the identification R2n ' Cn between the euclidean
structure and the real part of the hermitian structure: xx̃ + y ỹ = Re z z̃.
As for (1.1.9), it says that the complexified differentials, that is, the set
of the combinations of the dxj ’s and dyj ’s with complex coefficients, are
identified with the combinations of the dzj ’s and dz̄j ’s. Also, (1.1.10)
identifies real differentials with “diagonal” combinations, in which the
coefficients of the dz̄j ’s are the conjugates of those of the dzj ’s. We will
come back to this point in Chapter 3.
Differential forms which are combinations of the dzj ’s are said to
be of type (1, 0) and those which are combinations of the dz̄j ’s of type
(0, 1). Thus, ∂f and ∂f ¯ are the components of df of type (1, 0) and
(0, 1), respectively. Let Ω be a domain in Cn .
Definition 1.1.2. A function f ∈ C 1 (Ω) is said to be holomorphic
or complex analytic when df = ∂f , that is, when f satisfies the differ-
ential system ∂z̄j f = 0, j = 1, ..., n. The set of holomorphic functions
in Ω is denoted by hol(Ω).
The above system is called the Cauchy-Riemann system (CR system
for short). Thus, df , which is in general R-linear, is in this case C-linear,
4 1. SEVERAL COMPLEX VARIABLES
∂ζ̄ g(ζ)
We note that the last integrand in (1.2.3) is −dω = ζ−z
dζ ∧ dζ̄. Since
dω is integrable at its singular point z, then
ZZ ZZ
dω → dω.
∆\∆ ∆
But in the above integral we can make the substitution which we are
allowed by the 1-dimensional Cauchy formula:
Z
1 f (ζ1 , ..., ζn )
(1.2.5) f (ζ1 , ..., ζn−1 , zn ) = dζn .
2πi |ζn −zn |=rn ζn − zn
From (1.2.4), (1.2.5) we get (1.2.1) by Fubini’s theorem.
Remark 1.2.4. The separate C 1 regularity in each zj is needed
for Stokes’s formula, and the joint C 0 regularity is needed for Fubini’s
theorem.
For a function f ∈ C 0 (Ω) separately (C 1 and) holomorphic in each
zj , we wish to apply (1.2.1) to all possible P̄ ⊂⊂ Ω. Since the functions
of z produced by the integrals in ζ are C ∞ , as one can check by bringing
derivatives in the zj ’s inside the convergent integral, we get at once
Corollary 1.2.5. If f is C 0 on Ω and separately holomorphic in
each zj , then it is C ∞ and in particular jointly holomorphic.
We write the point ranging in ∂0 P (z, r) as ζ = (ζj ) for ζj = zj +
iθj
rj e . Then we have
Z
−n f (ζ)
(2πi) dζ1 ∧ ... ∧ dζn
∂0 P (z,r) (ζ1 − z1 ) · ... · (ζn − zn )
Z
= f (z + (r1 ei2πθ1 , ..., rn ei2πθn ))dθ1 ∧ ... ∧ dθn .
[0,1]n
Therefore, the domain of convergence D of a power series shows its
first two interesting features:
• It is invariant under rotations: if z ∈ D, then eiθ z ∈ D for any
θ ∈ [0, 2π]n .
• If w ∈ D, then, for any z which satisfies |zj | ≤ |wj | for all j,
we also have z ∈ D.
An open set which enjoys the first property is called a “Reinhardt do-
main”: it can be described through its representative in the space of
moduli |D| ⊂ (R+ )n defined by |D| := {(|z1 |, ..., |zn |) : z ∈ D}. A Rein-
hardt domain which enjoys the second property is called “complete”: it
is therefore characterized as a union of polydiscs centered at 0. There-
fore, for a domain of Cn to be the domain of convergence of a power
series, we have to require it to be Reinhardt and complete. But not all
complete Reinhardt domains are precisely the convergence domain of a
power series. Some additional geometric requirements must be fulfilled.
Theorem 1.3.4. D is “logarithmically convex” in the sense that the
set D∗ := {ξ ∈ Rn : ξj = log|zj |, j = 1, ..., n, for z ∈ D} is convex.
◦
Proof. Remember that D = B; hence, what we have to prove is
that
(
|aα z α | < c,
(1.3.3)
|aα wα | < c
implies for λ ∈ [0, 1]
(1.3.4) |aα z λα w(1−λ)α | < c.
But this is a straightforward calculation.
We compare now analytic functions and convergent power series.
We have already seen that they coincide over polydiscs, hence over the
union of those which are centered at the same point: this follows easily
from analytic continuation. In particular, over a Reinhardt complete
(non-empty) domain, a holomorphic function is always represented as
a sum of a power series centered at 0. (Of course for a general domain
1.3. ANALYTIC FUNCTIONS AND POWER SERIES 13
In the same way as for the function g previously discussed, we can see
(α)
that fα (z) = f α!(0) z α on Ω . But then f is a sum of a power series.
14 1. SEVERAL COMPLEX VARIABLES
(i) ϕ is upper semicontinuous, i.e., for any zo , ϕ(zo ) ≥ lim sup ϕ(z);
z→zo
(ii) for any K ⊂⊂ Ω and for any h continuous on K and har-
◦
monic on K,
(1.4.1) ϕ|∂K ≤ h|∂K implies ϕ|K ≤ h|K .
In particular, by taking as h the constant function on K with value
max ϕ, which exists and is finite by the upper semicontinuity of ϕ,
∂K
we have that subharmonic functions satisfy the maximum principle.
We denote by SH(Ω) the set of subharmonic functions on Ω. A func-
tion ϕ is said to be superharmonic when −ϕ is subharmonic. The
harmonic functions are precisely those which are at the same time
sub/superharmonic. This is always true once we know that any contin-
uous function on ∂∆ is endowed with a unique harmonic extension to
∆ (given by the Poisson integral as explained in Section 3.7). An easier
argument for C 2 functions is based on Theorem 1.4.9.
Subharmonicity does not appear to be a local property at first sight
but it turns out to be. Let ∆ be the standard disc in C and ∆zo r the
disc of center zo and radius r. Here is the main characterization.
Theorem 1.4.2. Let ϕ : Ω → [−∞, +∞) be an upper semicontin-
uous function. The following are equivalent,
(i) ϕ ∈ SH(Ω).
(ii) For any disc ∆zo r ⊂⊂ Ω and for any polynomial P = P (z),
ϕ|∂∆zo r ≤ Re P |∂∆zo r implies ϕ|∆¯ zo r ≤ Re P |∆¯ zo r .
(iii) (spherical submean) For any ∆ ¯ zo r ⊂⊂ Ω, we have
Z
ϕ(zo ) ≤ (2πr)−1 ϕds,
∂∆zo r
(iv)loc (Local solid submean) For any zo there is ro < dist(zo , ∂Ω)
such that for any r ≤ ro , (iv) holds.
Proof. (i) ⇒ (ii) and (iv) ⇒ (iv)loc are obvious, whereas (iii) ⇒
(iv) is immediate by integration on dr.
(ii) ⇒ (iii): We take a continuous function f (θ) ≥ ϕ(zo + reiθ ) and
+N
ak eikθ
P
approximate f by a real trigonometric polynomial g(θ) =
k=−N
16 1. SEVERAL COMPLEX VARIABLES
where the last equality follows from the fact that all non-constant har-
monics vanish when integrated over complete cycles. Finally, by the
third inequality of (1.4.2), we have
Z 2π Z 2π
1 1
(1.4.4) g(θ)dθ ≤ f (zo + reiθ )dθ + .
2π 0 2π 0
Combining (1.4.3) and (1.4.4) and then taking infimum in f and we
get (iii).
◦
(iv)loc ⇒ (i): Let K ⊂⊂ Ω and let h be harmonic on K, continuous
on K and such that
(1.4.5) ϕ|∂K ≤ h|∂K .
Let M := max (ϕ − h); we reason by contradiction and assume M > 0.
K
We define the function v := ϕ − h and the set F := v −1 (M ): v is upper
◦
semicontinuous and F is a compact subset of K because of (1.4.5).
Let zo be a point of ∂F . Then, there is some disc ∆(zo , r) such that
∆(zo , r) ∩ (K \ F ) 6= ∅. Also, we can suppose that r is taken so small
that (iv)loc holds for such a zo and r. It follows that
M = v(zo )
Z 2π Z r
2 −1
≤ (πr ) tvdt dθ.
0 0
But the mean value on the right hand side is taken over a disc which
has an open portion outside F , the set of maximum value M for v.
Hence this mean value must be < M , a contradiction.
1.4. SUBHARMONIC FUNCTIONS 17
Subharmonicity is preserved under composition with convex in-
creasing functions.
Theorem 1.4.7. Let χ be a function from R to R such that χ̇ ≥
··
0 and χ ≥ 0 (extended at −∞ by χ(−∞) := lim χ(t)). Let ϕ be
t→−∞
subharmonic; then χ(ϕ) is also subharmonic.
Proof. Since χ is convex, then for any to and for suitable c, we
have
(1.4.7) χ(to ) ≤ χ(t) − c(t − to ).
Replacing t with ϕ(z + reiθ ) in (1.4.7) and integrating over θ, we get
(1.4.8) Z 2π Z 2π
−1 iθ iθ
χ(to ) ≤ (2π) χ(ϕ)(z + re )dθ − c( ϕ(z + re )dθ − 2πto ) .
0 0
−1 2π
R
We choose now to = (2π) ϕ(z + reiθ )dθ and get
0
Z 2π
−1 iθ
χ(ϕ(z)) ≤ χ (2π) ϕ(z + re )dθ
0
Z 2π
−1
≤ (2π) χ(ϕ)(z + reiθ )dθ,
0
where the first inequality follows from the subharmonicity of ϕ and
the fact that χ is increasing and the second from (1.4.8) under the
already-mentioned choice of to .
We have said that subharmonicity is preserved under composition
with a convex increasing χ but it is indeed strengthened: χ(ϕ) is “more”
subharmonic than ϕ is and how much more depends on how convex χ
is. According to [43], ϕ is said to be “strongly subharmonic” when there
exists χ positive, increasing, strictly convex, such that χ−1 (ϕ) is still
subharmonic. Thus, for instance, if ψ is subharmonic and positive, then
any power ψ a for a > 1 is strongly subharmonic. It is another useful
exercise to prove that if ϕ is C 2 , strongly subharmonic and satisfies
¯ > 0.
∂ϕ 6= 0, then ∂∂ϕ
We recall now that it is proved in [43] that if ϕ is strongly sub-
harmonic and has a finite harmonic majorant in the unit ball, then its
boundary measure is absolutely continuous: just plain subharmonicity
would not suffice for this conclusion. A very intense case, corresponding
to χ = exp, occurs when ϕ is logarithmically subharmonic, that is, log ϕ
is subharmonic. Let us further discuss logarithmic subharmonicity. We
1.4. SUBHARMONIC FUNCTIONS 19
begin by the obvious remark that for f ∈ hol(Ω), the mean property
of f , and hence the submean of |f |, implies that |f | ∈ SH(Ω). We can
say more.
Lemma 1.4.8. If f ∈ hol(Ω), then log|f | ∈ SH(Ω).
Proof. Let K be a compact of Ω and P a polynomial such that
(1.4.9) log |f |∂K ≤ Re P ∂K .
We apply “exp” and observe that eRe P = |eP |; hence (1.4.9) yields
f
(1.4.10) ≤ 1.
eP
∂K
By the maximum principle, (1.4.10) remains true when passing from
∂K to K and next, the reverse of the implication from (1.4.9) to (1.4.10)
yields
log |f |K ≤ Re P K .
Thus, log |f | is subharmonic by (ii) of Theorem 1.4.2.
1 1
In particular, since |f | = exp ν1 log(|f |), we have that |f | is sub-
ν ν
harmonic for any ν. This will be the main tool in the problem of sepa-
rate analyticity that we will solve in Section 1.5. However, the subhar-
1
monicity of all the |f | ν ’s, that is, the strong subharmonicity of f in any
degree, does not suffice in proving such a delicate property as propaga-
tion of CR extendibility. We will see that logarithmic subharmonicity
of |f | is needed.
We want to give a differential description of subharmonicity.
Though the characterization applies to a more general situation,
we want to confine ourselves to the case of ϕ ∈ C 2 .
Theorem 1.4.9. Let ϕ ∈ C 2 (Ω); then ϕ is subharmonic if and only
if
(1.4.11) ∂z ∂z̄ ϕ ≥ 0.
Note that ∂z ∂z̄ coincides, up to a constant factor, with the cele-
brated Laplace operator
∂x2 + ∂y2 = 4∂z ∂z̄ .
Proof. We first prove that the submean property implies (1.4.11).
We rewrite the submean contained in (iii) of Theorem 1.4.2 as
Z 2π
ϕ(zo + reiθ ) − ϕ(zo ) dθ ≥ 0.
(1.4.12)
0
20 1. SEVERAL COMPLEX VARIABLES
Theorem 1.4.11 is a variant of the famous Hartogs lemma.
Remark 1.4.12. The use of Fatou’s lemma inside the spherical
submean estimate would not suffice for the conclusion. In that case we
should apply Fatou’s lemma to the function θ 7→ sup ϕν (z + reiθ ) but
z
we do not know if its lim sup is still ≤ l. However, a proof which uses
1-dimensional integrals is also available: notice that it uses the Poisson
kernel and not just the submean (cf. [71]) as is hereby done.
But the first sum vanishes because ∂z̄k z̃i = 0. We have likewise
X X
∂ zh (∂z̃i r̃)∂zk z̃i = ∂z̃j ∂z̃i r̃∂zh z̃j ∂zk z̃i
i ji
X X
+ ∂z̃j ∂z̃i r̃∂zh z̃¯j ∂zk z̃i + ∂z̄i r̃∂z̄h ∂zk z̃i .
ji ji
Again, the sums in the second line cancel. Writing F (z) for z̃(z), we
get the identity for matrices
Lr = t J(F ) · Lr̃ · J(F ),
which is another way of stating (1.6.3).
The Levi form is called the “complex hessian” of r. The full hessian,
the quadratic term of the expansion of r, written in coordinates (z, z̄) ∈
n
Cn × C is
1X 1X X
hess r = ∂zi ∂zj rdzi ⊗dzj + ∂z̄i ∂z̄j rdz̄i ⊗dz̄j + ∂zi ∂z̄j rdzi ⊗dz̄j .
2 ij 2 ij ij
The sum of the first two terms on the right side is a real harmonic
polynomial
P called the “Levi polynomial” that we denote by LPr :=
Re ij ∂zi ∂zj rdzi ⊗dzj ; thus hess r = LPr +Lr . By a coordinate change
we can see that the Levi polynomial can be canceled from a graphing
function: the Levi form is therefore the holomorphic invariant of the
hessian. A first rough way of getting rid of the Levi polynomial consists
in taking the mean of the value of the hessian over a vector u and iu.
Proposition 1.6.3. We have
1
Lr (u, ū) = [hess r(u, u) + hess r(iu, iu)] .
2
Proof. We have
(1.6.4) LPr (iu, iu) = −LPr (u, u),
(1.6.5) Lr (iu, iu) = Lr (u, ū).
It follows that
hess r(u, u) + hess r(iu, iu)
= (LPr (iu, iu) + LPr (u, u)) + (Lr (iu, iu) + Lr (u, ū))
= 2Lr (u, ū).
26 1. SEVERAL COMPLEX VARIABLES
z 0 7→ z 0 .
By the implicit function theorem the equation of M becomes
y1 = Q(z 0 , z̄ 0 ) + o2 (x1 , z 0 ).
We now want to eliminate the Levi polynomial inside Q. We perform
the quadratic transform
n
z1 7→ z1 − i ij 2aij zi0 zj0 ,
P
change z̃ = z̃(z) under which the inequality which defines the domain
changes into
(1.6.7) ỹ1 < −|z̃2 |2 + ....
For this purpose, it is enough to consider the transformation
(
z̃1 = z1 + i 23 z22 ,
z̃2 = √z22 ,
On the other hand it is clear that this family {Az0 } covers a full neigh-
borhood B 0 of zo = 0. The conclusion is then a consequence of the
extension criterion herewith proved.
30 1. SEVERAL COMPLEX VARIABLES
dζ ∧ dζ̄
Z
ζ
ϕν (z) = ϕ(z − )χ(ζ) .
∆×...×∆ ν (−2i)n
fact,
Z 2π
1
ϕν (wo + reiθ )dθ
2π 0
Z Z Z 2π
1 iθ ζ dζ ∧ dζ̄
= ϕ(wo + re − )dθ χ(ζ)
∆ 2π 0 ν −2i
dζ ∧ dζ̄
ZZ
ζ
≥ ϕ(wo − )χ(ζ) = ϕν (wo ).
∆ ν −2i
We come back to general dimension n. Since the ϕν ’s are plurisub-
harmonic, then the (ϕν ) 1 ’s are decreasing with & 0 (by iterated 2-
dimensional submean over -discs). By letting ν → ∞, the ϕ 1 ’s them-
selves are decreasing: writing ν in place of 1 , we conclude ϕν & ϕ.
Last, we show that ϕ is plurisubharmonic. It is upper semicontinuous
for the same reason as in one variable (cf. the beginning of the proof of
Lemma 1.4.6). Moreover, by taking composition with the parametriza-
tion of a disc A = Φ(∆), we have
ϕν ◦ Φ & ϕ ◦ Φ.
Since each ϕν ◦ Φ is subharmonic, then also the monotonic limit ϕ ◦ Φ
is subharmonic on account of Lemma 1.4.6.
Let Ω be an open domain, v a unit vector in Cn , f a holomorphic
function on Ω. We consider the function
ν 1 !−1
∂ f (z) ν
rfv (z) = lim sup v
(1.7.1) , z ∈ Ω.
ν ν!
The function rfv (z) represents the radius of the maximal disc in the v
direction in which the Taylor series of f at z is separately convergent
and hence f is separately holomorphic. When we uniformize over z 0
ranging in a neighborhood B of z in the (n − 1)-plane orthogonal to v
passing through z, we get
Proposition 1.7.3. We have
(1.7.2) inf
0
rfv (z 0 ) = sup{r : f is holomorphic in v∆r × B}.
z ∈B
−z
zo − z is normal to ∂Ω at zo . We write v := |zzoo −z| and d(z) := |zo − z|;
thus, d(z) is the distance of z to ∂Ω. In this situation, for a holomorphic
function on Ω we have
if inf
0
rfv (z 0 ) > d(z) for some B, then f extends holomorphically across ∂Ω at zo .
z ∈B
≤ 1.
Proof. “≥” is obvious. As for “≤”, let r < inf rfv . We then have
K
Lemma 1.8.4. Let Ω be a domain of holomorphy. Then
d(z, ∂Ω) = inf rfv (z).
v∈S 2n−1 , f ∈hol(Ω)
42 1. SEVERAL COMPLEX VARIABLES
= d(K̂Ω , ∂Ω).
Definition 1.8.6. A domain Ω is said to be holomorphically convex
when for any K compact in Ω, the set K̂Ω is still compact in Ω.
In particular, Theorem 1.8.5 says that a domain of holomorphy is
holomorphically convex. Also, since K̂Ω is controlled by the convex
hull K̂, it follows that a convex domain is holomorphically convex.
Holomorphic convexity is stable under intersection.
Proposition 1.8.7. T If Ωj is a domain of holomorphy for any j,
then the interior Ω of j Ωj is holomorphically convex.
Proof. We have
d(K̂Ω , ∂Ωj ) ≥ d(K̂Ωj , ∂Ωj )
(1.8.3)
= d(K, ∂Ωj ),
where the first inequality follows from the inclusion K̂Ωj ⊃ K̂Ω and the
second equality is a consequence of Theorem 1.8.5. If we take “inf” in
j
(1.8.3) we conclude
(1.8.4) d(K̂Ω , ∂Ω) = d(K, ∂Ω).
1.8. DOMAINS OF HOLOMORPHY AND PSEUDOCONVEX DOMAINS 43
¯
defined on Ω and induces f on ω. It is not necessarily ∂-closed; to
obtain it, we have to use a “correction” term of type zn v for a solution
¯ ∗f
¯ = ∂χ∧π
v of ∂v and set
zn
f˜ := χπ ∗ f − zn v.
The existence of the solution v to the above equation is assured by the
vanishing hypothesis of the cohomology groups ( for degree k + 1 if f
has degree k). Thus, f can be “lifted” to f˜ ∂-closed.
¯ Also, ω inherits
from Ω the vanishing of the cohomology:
(1.8.7) Kk (ω) = 0, k ≥ 1.
¯ = 0 on ω, we first lift it to Ω, then solve ∂¯ũ = f˜
In fact, if f satisfies ∂f
on Ω and finally define u := j ∗ ũ. This is the solution of ∂u¯ = f on ω.
By the inductive assumption, (1.8.7) implies that ω is a domain of
holomorphy (which is trivially true at the first inductive step ω ⊂ C1 )
and therefore there is a function f ∈ hol(ω) which does not extend at
zo . Next, using again that K1 (Ω) = 0, we can find f˜ ∈ hol(Ω) such that
f˜|ω = f ; hence f˜ cannot extend holomorphically and Ω is a domain of
holomorphy.
(iv) ⇒ (i): In a domain of holomorphy Ω we have (Lemma 1.8.4)
d(z) = inf rfv (z), z ∈ Ω.
v, f
Thus − log d = sup(− log rfv ); now, the functions − log rfv ’s enjoy the
v, f
submean property along discs according to Theorem 1.7.5 and the same
is true for their “sup”. Since, in addition, − log d is continuous, then
it is plurisubharmonic. The proof of the theorem is complete.
Proposition 1.8.14. Let Ω be a domain of Cn and suppose that
for every point in ∂Ω there is a basis of neighborhoods {B} such that
B ∩ Ω satisfies (ii) of Theorem 1.8.13; then Ω also satisfies (ii).
Proof. We denote by ψB∩Ω the plurisubharmonic exhaustion func-
tion on each B ∩ Ω. By taking a locally finite covering of ∂Ω by the B’s
and gluing the various ψB∩Ω by taking their sup over overlappings, we
find ψ 1 plurisubharmonic on a neighborhood Ω1 of ∂Ω in Ω, such that
ψ 1 → +∞ as z → ∂Ω, and ψ 1 is locally bounded at ∂(Ω\Ω1 ). Let ψ 2 be
a plurisubharmonic function on Cn with ψ 2 → +∞ if |z| → +∞ such
that ψ 2 > ψ 1 over a neighborhood V of ∂(Ω \ Ω1 ). Define ψ to be ψ 2 on
Ω2 := (Ω \ Ω1 ) ∪ V and sup(ψ 1 , ψ 2 ) on Ω1 ; ψ is an exhaustion function
for Ω and it is plurisubharmonic because this property is stable under
“sup” over finite indices.
1.8. DOMAINS OF HOLOMORPHY AND PSEUDOCONVEX DOMAINS 47
Proposition 1.8.14 reveals that pseudoconvexity is a local property.
In combination with Theorem 1.8.13, it shows that the property of
being a domain of holomorphy is also local. Moreover, it implies a
version of this property in terms of germs.
Remark 1.8.15. If Ω is a domain of holomorphy, then for any
zo ∈ ∂Ω there does not exist a forced extension map for germs of
holomorphic functions
ext
(1.8.8) hol(Ω)zo 99K hol(Cn )zo .
This means that there is a neighborhood B of zo and a holomorphic
function f ∈ hol(Ω ∩ B) (indeed f ∈ hol(Ω)), such that f ∈ / hol(B 0 ) for
any B 0 ⊂ B. It is an open question whether the opposite implication
holds, that is, whether non-extendibility passes from germs to global
sections. Of course the question makes sense only if it is formulated for
any boundary point zo ∈ ∂Ω. In fact, for a single point zo we can well
have extension for global holomorphic functions but not for germs. For
instance, if Ω is defined in C2 by y2 > ϕ(y1 ) where ϕ ≡ 0 for |y1 | < 1
ext
and ϕ = (|y1 | − 1)2 for |y1 | ≥ 1, then hol(Ω) 99K hol(C2 ) holds but
ext
hol(Ω)0 99K hol(C2 )0 does not.
We come back to Theorem 1.8.13. We have proved, in conjunction
with the conclusions of Theorem 1.8.13, that if there exists an exhaus-
tion plurisubharmonic function ψ for Ω, then − log d is also plurisub-
harmonic. This is a deep point in which we use the full strength of the
characterization. A direct but not simple proof could be given along
the lines of [50, Theorem 2.3.7, p. 46]. Note that in the next geometric
characterization and for ∂Ω of class C 2 we deal in fact with − log(−r)
where r is a defining function for ∂Ω with r < 0 on Ω; but it is clear that
this function could replace − log d in Theorem 1.8.13. We also need an
extra amount of Levi positivity which is provided by an additional term
λ|z|2 for large λ.
Let Ω be a domain with C 2 boundary defined by r < 0 with ∂r 6= 0
at points where r = 0, set M = ∂Ω, let LM = Lr |T C M be the Levi
−
form of M and let s+ 0
M , sM , sM be the number of its eigenvalues which
are > 0, < 0, = 0, respectively. We describe M , in a neighborhood of
a boundary point zo , as a graph y1 = f (x1 , z 0 , z̄ 0 ), set r = −y1 + f ,
and denote by z 7→ z ∗ the projection on M along the y1 -axis. The
following relates, in a neighborhood of zo , the signature of the Levi
form of the exhaustion function ψ = − log(−r) + λ|z|2 for suitable λ
to the signature of the Levi form of the boundary LM .
48 1. SEVERAL COMPLEX VARIABLES
Zampieri [101] (cf. also Hörmander [50, Theorem 2.6.12] for another
proof).
(1.9.1) ¯ =f
∂u ¯ = 0,
for f satisfying ∂f
where the datum f and the solution u are (0, k)- and (0, k − 1)-forms,
respectively. We prove solvability of (1.9.1) in C ∞ (Ω̄) for all k ≥ q + 1
when Ω is bounded, has smooth boundary, and is “q-pseudoconvex”. If
Ω is unbounded or does not have C ∞ boundary, but can be exhausted
by bounded, C ∞ , q-pseudoconvex domains Ων % Ω, then (1.9.1) can
be solved in C ∞ (Ω) in corresponding degrees k. In particular, all pseu-
doconvex domains have this exhaustion and this implies the vanishing
of all cohomology groups of ∂¯ in degree k ≥ 1 (the implication (ii) ⇒
(iii) of Theorem 1.8.13, that is, the solution of the Levi problem).
The key point in solving (1.9.1) is to also call into play the adjoint
operator ∂¯∗ : in particular, we take full advantage of the ellipticity in
the interior of the system (∂,¯ ∂¯∗ ).
To study equation (1.9.1), the most effective tool is the Hilbert space
¯
techniques in the context of the ∂-Neumann problem. To exploit them,
we first establish the a priori estimates of Morrey, Kohn and Hörmander
in the weighted L2 spaces. By a suitable choice of the weight, these
¯
ensure the existence of the ∂-Neumann operator and its continuity in
the Sobolev spaces H . The C ∞ regularity up to the boundary of ∂¯
s
J
P0
ujK := sign jK
uJ . We define |u|2 := |uJ |2 ; this norm is inde-
|J|=k
pendent of the choice of the orthonormal basis. We consider a bounded
domain Ω ⊂⊂ Cn with boundary ∂Ω of class C ∞ defined by r(z, z̄) < 0
with ∂r 6= 0 over the points where r = 0. We choose our orthonormal
frame of (1, 0)-forms such that ∂ωn r = 1 (and for this we need possibly
to renormalize r) and ∂ωj r = 0 for any j ≤ n − 1; hence ∂ωj , j ≤ n − 1,
are a basis for T 1,0 ∂Ω. The coefficients of our forms are taken in vari-
ous spaces Λ such as C ∞ (Ω̄), C ∞ (Ω), Cc∞ (Ω), L2 (Ω), H s (Ω) and the
corresponding spaces of k-forms are denoted by Λk . Though our a pri-
ori estimates are proved over smooth forms, they are stated in Hilbert
norms. Thus, let ||u||H 0 be the H 0 = L2 -norm and, for a real function
ϕ, let the weighted L2 -norm be defined by
X0 Z
(1.9.2) 2
||u||Hϕ0 := e−ϕ |uJ |2 dV,
|J|=k Ω
where ϕj stands for ∂ωj ϕ and the dots denote a 0-order operator which
does not involve ϕ. We denote by ∂¯ϕ∗ the Hϕ0 -adjoint of ∂.
¯ The operator
∂¯ϕ is still closed, densely defined but it is no longer true that smooth
∗
forms belong to D∂¯ϕ∗ . For this, they must satisfy certain boundary con-
ditions. Namely, there must exist g ∈ (Hϕ0 )k−1 such that
If this is the case, one sets ∂¯ϕ∗ u := g. But integration by parts shows
that (1.9.4) cannot hold unless
X Z
e−ϕ ∂ωj (r)ujK ψ̄K dS = 0 for any K and ψK ,
j=1,...,n ∂Ω
P
that is, ∂ωj (r)ujK |∂Ω ≡ 0 for any K. (Here dS is the element of
j=1,...,n
hypersurface in ∂Ω.) Since we have chosen our basis with the property
∂ωj r|∂Ω = κjn , we then conclude
(1.9.5) u belongs to D∂¯ϕ∗ iff uJ |∂Ω = 0 whenever n ∈ J.
As a set, Hϕ0 coincides with H 0 ; from (1.9.5) it also follows that D∂¯ϕ∗ =
D∂¯∗ . We call a form which belongs to D∂¯∗ “tangential”. Over such a
form the action of the Hilbert adjoint of ∂¯ coincides with that of its
“formal adjoint” and is therefore expressed by a “divergence operator”:
X0 X
(1.9.6) ∂¯ϕ∗ u = − δωj (ujK )ω̄K + ... for any u ∈ D∂¯∗ ,
|K|=k−1 j
(1.9.8)
¯ 2 0 + ||∂¯∗ u||2 0 + C||u||2 0
||∂u||Hϕ ϕ Hϕ Hϕ
X0 X Z X0 X Z
−ϕ
≥+ e ϕij uiK ūjK dV − e−ϕ ϕjj |uJ |2 dV
|K|=k−1 ij Ω |J|=k j≤qo Ω
X0 X Z X0 X Z
+ e−ϕ rij uiK ūjK dS − e−ϕ rjj |uJ |2 dS
|K|=k−1 ij ∂Ω |J|=k j≤qo ∂Ω
X0 X X0 X
+ (1 − ) ||δωj uJ ||2Hϕ0 + ||∂ω̄j uJ ||2Hϕ0 .
|J|=k j≤qo |J|=k j≥qo +1
Here rj stands for ∂ωj r. Recall that ∂ωj r = h∂ωj , ωn i and this is 0
or 1 according to j < n or j = n. On the other hand, for j = n,
it is the coefficient unK which vanishes when restricted to ∂Ω. We
thus conclude that the boundary integrals vanish in both equalities
of (1.9.9). Note also that by taking the sum of the two terms in the
right side of (1.9.9), after discarding the boundary integrals, we put
in evidence the commutators [δωi , ∂ω̄j ]. In the last sum of the second
equality of (1.9.7), we now want to interchange ∂ω̄j with δωj for indices
j ≤ qo . A double integration by parts yields
Z
||∂ω̄j uJ ||ϕ = ||δωj uJ ||ϕ − e−ϕ δωj , ∂ω̄j uJ ūJ dV + R,
2 2
(1.9.10)
Ω
X0 X Z
− e−ϕ [δωj , ∂ω̄j ](uJ )ūJ dV
|J|=k j≤qo Ω
!
X0 X X
+ ||δωj uJ ||2ϕ + ||∂ω̄j uJ ||2ϕ + R.
|J|=k j≤qo j≥qo +1
= rij .
Similarly
hω̄n , [∂ωi , ∂ω̄j ]i = −h∂ ω̄n , ∂ωi ∧ ∂ω̄j i
= −rij .
It follows that
(1.9.14) [∂ωi , ∂ω̄j ] = rij ∂ωn − rij ∂ω̄n + ...,
where the dots denote combinations of ∂ωh ’s and ∂ω̄h ’s for h ≤ n − 1.
We denote by chij = hωh , [∂ωi , ∂ω̄j ]i(= −h∂ω ¯ h , ∂ω ∧ ∂ω̄ i), resp. dh =
i j ij
hω̄h , [∂ωi , ∂ω̄j ](= −h∂ ω̄h , ∂ωi ∧∂ω̄j i), the coefficients with which the ∂ωh ’s
56 1. SEVERAL COMPLEX VARIABLES
and ∂ω̄h ’s, respectively, occur in the dotted terms of (1.9.14). (Inciden-
tally, they are related by the relation
dhij = hω̄h , [∂ωi , ∂ω̄j ]i
= hωh , [∂ω̄i , ∂ωj ]i
= −hωh , [∂ωj , ∂ω̄i ]i = −chji .)
Let (ϕij )ij be the matrix of Lϕ in the basis ω1 , ..., ωn . We have
n−1
X
ϕij = ∂ωi ∂ω̄j (ϕ) + (∂ ω̄n )ij ϕω̄n + (∂ ω̄h )ij ϕω̄h
h=1
(1.9.15) n−1
X
= ∂ωi ∂ω̄j (ϕ) + rij ϕω̄n − dhij ϕω̄h .
h=1
It follows that
[−ϕi , ∂ω̄j ] = ∂ω̄j ∂ωi (ϕ)
= −[∂ωi , ∂ω̄j ](ϕ) + ∂ωi ∂ω̄j (ϕ)
(1.9.16) n−1
X
= ϕij − rij ϕωn − chij ϕωh .
h=1
δωn (un K )ū. We also denote by S the sum in the last line in (1.9.11).
Thus, we can rewrite the right side of (1.9.11) as
!
X0 XZ XZ
e−ϕ ϕij · − e−ϕ ϕjj ·
|K|=k−1 ij Ω j≤qo Ω
!
X0 XZ XZ
+ e−ϕ rij · − e−ϕ rjj · + S + R,
|J|=k ij ∂Ω j≤qo ∂Ω
For the same reason, in the right side of (1.9.8), the boundary integrals
vanish. If we then apply the variant of (1.9.8) with, say, ϕ = 0, we get
an estimate which fully expresses the elliptic regularity in the interior
¯ ∂¯∗ ):
of the system (∂,
(1.9.20) ||∂¯∗ u||2 + ||∂u||
¯ 2 + ||u||2 > ||u||2 1
H for any u ∈ (Cc∞ (Ω))k ,
∼
q+1 qo
X X
(1.9.21) λj (z) − rjj (z) ≥ 0.
j=1 j=1
In some cases we can do better. For instance, assume that s− (z) is con-
stant; then λs− < 0 ≤ λs− +1 and hence the negative eigenvectors span
− −
a bundle V s that,
P identified withP the span of the first s coordinate
vectors, yields j≤s− +1 λj (z) − j≤s− rjj (z) ≥ 0. Note that a pseudo-
convex domain is characterized by s− (z) ≡ 0; thus, it is 0-pseudoconvex
in our sense.
1.9. L2 -ESTIMATES FOR ∂¯ 59
(It is meant here that the inequality is true for any T s−i and for suitable
T s .) Classically, this formula is referred to as a consequence of the
“noncharacteristicity” of (∂,¯ ∂¯∗ ). If i = 1 and s = 1 it immediately
follows from the splitting
(1.9.29) N = ∂ω̄n + T ,
¯ 2 +||∂¯∗ u||2 +||u||2 contains ||∂ω̄n u||2 .
and the fact that the “energy” ||∂u||
For general i and s, one replaces ∂ω̄n inside (1.9.29) (or ∂ωn inside the
twin decomposition N = ∂ωn + T for a new T ) according to the (non-
characteristicity) relations
N uj K = ∂ω̄n uj K + T uj K
¯ j n K + T uj K + P 0 u,
= ∂ω̄j un K − (∂u)
N un K = ∂ωn un K + T un K
(1.9.30)
n−1
X
= (∂¯∗
u)K − ∂ωj uj K + T un K + Pt0 u,
t
j=1
where P 0 and Pt0 are 0-order operators. If we then apply N i−1 T s−i
to the two terms of (1.9.30), remark that ∂ω̄j and ∂ωj j ≤ n − 1 are
derivatives of type T , and put into an s − 1 order term the error from
commutators for bringing N in head position, we get (1.9.28). Now,
(1.9.28), together with induction, shows that in order to estimate the
various ||N i T s−i u||’s, that is, the full norm||u||s , it is sufficient to es-
timate the ||T s u||’s. This is the central point of the proof. To see it,
we begin by remarking that T s u still belongs to D∂¯∗ . Next, we observe
that we can describe the commutator [∂¯t∗ , T s ] as As + Ats−1 where As
and Ats−1 are operators of order s and s − 1, respectively, the first being
independent of t and involving at most one non-tangential derivative.
¯ T s ] has order ≤ s and is independent of t; hence it can be
Also, [∂,
1.9. L2 -ESTIMATES FOR ∂¯ 61
¯
The ∂-Neumann operator. We give the outline of the construction
by Kohn. We begin by recalling from [50, Lemma 4.1.3] and, more
closely, [36, Lemma 4.3.2] that in a domain Ω with C ∞ boundary,
C ∞ (Ω̄)k ∩ D∂¯∗ is dense in H 0 (Ω)k ∩ D∂¯∗ ∩ D∂¯ for the triplet norm
¯ 2 0 + ||∂¯∗ u||2 0 .
||u||2 0 + ||∂u||
H H t H
For this reason, (1.9.25) is in fact true for any u ∈ H 0 (Ω)k ∩ D∂¯t∗ ∩ D∂¯
instead of C ∞ (Ω̄)k ∩ D∂¯t∗ . We define now the “complex Laplacian” by
t := ∂¯∂¯t∗ + ∂¯t∗ ∂.
¯
This is a linear closed densely defined operator on forms with coeffi-
cients in H 0 which leaves unchanged their degree. A form u belongs
62 1. SEVERAL COMPLEX VARIABLES
to Dt iff u ∈ D∂¯t∗ and ∂u ¯ ∈ D∂¯∗ , that is, u satisfies the celebrated
t
¯
“∂-Neumann conditions”. In our normalization ωn = ∂r, they are de-
scribed over a smooth form u by
un K |∂Ω ≡ 0,
(1.9.34) ∂ω̄n uj K |∂Ω = (∂ω̄n uj K − ∂ω̄j un K )|∂Ω
¯ n K |∂Ω ≡ 0.
= (∂u)
It is readily checked that
¯ 2 0 + ||∂¯∗ u||2 0 .
ht u, uiH 0 = ||∂u||H t H
Thus, (1.9.27) for the choice s = 0 (extended for all u ∈ H 0 ∩ Dt ) can
be rewritten as
||u||2H 0 < ht u, uiH 0
∼
≤ ||t u||H 0 ||u||H 0 ,
which yields
||u||H 0 < ||t u||H 0 for any u ∈ H 0 (Ω)k ∩ Dt .
∼
The proof is the same as above and consists in getting control of the
commutators [Λs , t ] by a suitable choice of t = ts . By the density of
C ∞ (Ω̄) into H s , (1.9.35) can be rewritten as
(1.9.36)
||Nt u||H s < ||u||H s for any u ∈ H s (Ω) such that Nt u ∈ H s (Ω)k .
∼
1.9. L2 -ESTIMATES FOR ∂¯ 63
= f − ∂¯t∗ Nt ∂f
¯
= f.
Remark 1.9.10. If under the choice t = ts , we have (1.9.27) for
H , we also have it for any H j , j ≤ s. In particular, the solution u
s
of Theorem 1.9.9 satisfies not only the estimate in the second line of
(1.9.39) but also ||u||2H j < ||f ||2H j for any j ≤ s.
∼
We did not just find a solution in Hϕs t , but the one which is orthog-
¯ in fact ∂¯∗ u = 0 because ∂¯∗ ◦ ∂¯∗ = 0. This serves
onal to the kernel of ∂; t t t
to define the “Bergman projection”
Bs : H s → ¯
ker ∂,
f 7→ f − ∂¯t∗ Nt ∂f. ¯
over forms with coefficients in C ∞ (Ω̄) (resp. C ∞ (Ω)) (cf. the lines after
(1.8.6)).
Theorem 1.9.11. Let Ω be bounded, smooth and q-pseudoconvex or
let it satisfy (1.9.24). Then
Kk (Ω̄) = 0 if k ≥ q + 1.
Proof. We follow [58]. We have to solve the equation ∂u ¯ = f for
∞ k ¯
f ∈ C (Ω̄) such that ∂f = 0. We claim that we can find a sequence
{uν }ν of H ν solutions such that ∂¯t∗ u = 0 for t = tν and
||uν+1 − uν ||H ν ≤ 2−ν .
Suppose we have carried out our construction up to uν and now look
for uν+1 . Take ũν+1 ∈ H ν+1 with (∂¯ũν+1 = f, ∂¯t∗ ũν+1 = 0) according
to Theorem 1.9.9. Put h := −ũν+1 + uν ; the problem that we have to
overcome is that h is only H ν and not yet H ν+1 . Thus, in each patch
of a finite covering {Bj }j≥1 of ∂Ω we “push points of Ω̄ inward to Ω”
by means of the mapping
Ω̄ ∩ Bj → Ω, z 7→ z + vj ,
where vS
j is an “inner normal” to Ω in Bj . Let B0 be a neighborhood
of Ω \ Bj , take a partition of the unity {ψj }j≥0 subordinate to the
j≥1
covering {Bj }j≥0 and put
X
h (z) = ψ0 (z)h(z) + ψj (z)h(z + vj ).
j≥1
¯ ∂¯∗ ) in the interior of Ω, the function
Note that by the ellipticity of (∂, tν
∞
ψj hj . We have
P
h is C (Ω̄). We also rewrite the above sum as
j≥0
X
¯ =
∂h ¯ j ∧ hj
∂ψ
j≥0
X
= ¯ j ∧ (hj − h),
∂ψ
j≥0
P¯ ¯ P ψj h) = 0. Thus,
where the last equality follows from ∂(ψj )h = ∂(
j≥0 j≥0
¯ ||H ν → 0 as → 0. At this point we recall Theorem 1.9.7 and find
||∂h
a solution v of
¯ = ∂h ¯ ,
(
∂v
¯ ||H ν .
||v ||H ν < ||∂h
∼
The solution v can be found in any Sobolev space, in particular in
¯ ∈ C ∞ (Ω̄); its norm is controlled by the norm of ∂h
H ν+1 , because ∂h ¯
1.9. L2 -ESTIMATES FOR ∂¯ 65
¯
Notes. The ∂-Neumann problem was suggested by Garabedian
and Spencer [44]. The basic a priori estimates were first proved for
1-forms by Morrey [69]. Next, Kohn [55] derived the general estimates
and proved the regularity up to the boundary on strongly pseudocon-
vex domains. The L2 estimates with weights which are 0 at the bound-
ary were used by Hörmander [49] to bypass the boundary regularity
problem. The regularity at the boundary by use of the weighted ∂- ¯
Neumann problem, that is, the conclusion of Theorem 1.9.11 for q = 0,
was proved by Kohn in [57]. A very clear and exhaustive explanation
¯
of the ∂-Neumann method by Kohn can be found in the book by Chen
and Shaw [36]. As for the condition of q-pseudoconvexity and its use
¯
for the ∂-Neumann weighted estimates, we refer to Zampieri [101] and
Ahn [1]. Another approach, through the integral method, is mainly due
to Henkin (cf. [4]) and is very well explained in the last six chapters of
the book by Boggess [27].
We now want to make some complementary comments. The solution
¯ = f that we have constructed in Theorem 1.9.11 is obtained by
u of ∂u
approximation and is not necessarily the “canonical” solution. This is
the one in the form ∂¯∗ N f and hence, it satisfies ∂¯∗ u = 0; in particular,
68 1. SEVERAL COMPLEX VARIABLES
1
Thus we get control of “(TR ) 2m−1 ” by means of ∂ω and ∂ω̄ . Estimate
(1.10.4) follows from Theorem 2.4.6 below which is in turn a conse-
quence of Lemma 2.4.7 applied to the pair Re ∂ω and Im ∂ω . We call the
attention to this point. Finite type of systems of complex vector fields
implies subellipticity but under the strict assumption that the system
is closed under conjugation (cf. the counterexample by Kohn [60]). We
postpone to Remark 2.4.9 below the analysis of this point.
We state now 12 -subelliptic estimates. In these estimates, the
Sobolev spaces are unweighted and the forms u have compact support
in Ω̄. They apply to a domain Ω which is strongly q-pseudoconvex in
the sense that (1.9.21) holds with the inequality “≥ 0” replaced by
“≥ c” for c > 0. In particular, LM is not the 0-form and M has type
2: we have 12 -subelliptic estimates in this case:
Theorem 1.10.2. Let Ω be smooth and strongly q-pseudoconvex in
a neighborhood B of zo . Then, for any u ∈ Cc∞ (Ω̄ ∩ B)k ∩ D∂¯∗ , we have
¯ 2 + ||∂¯∗ u||2 + ||u||2
(1.10.5) |||u|||21 + |||∂r u|||− 1 < ||∂u|| if ≥ q + 1.
0 0 0
2 2 ∼
X0 X Z
− (1 − ) rjj |uJ |2 dS
|J|=k j≤qo ∂Ω
X0 X Z
− rjj |uJ |2 dS,
|J|=k qo +1≤j≤n−1 ∂Ω
Then, for k ≥ q + 1 and for any u ∈ Cc∞ (Ω̄ ∩ B)k ∩ D∂¯∗ , we have
(1.10.11)
¯ 2 +||∂¯∗ u||2 if < 1 where m = max mjo .
|||u|||2 +|||∂r u|||−1+ < ||∂u||0 0
∼ 2m−1 jo
In both cases, the term which involves boundary integrals can be dis-
carded since it is positive. What we get is
(1.10.13)
X 0 X X 0 X
2
||∂ω̄j uJ ||20
||∂ω j
uJ ||0 +
|J|=k j≤qo |J|=k qo +1≤j≤n
< ||∂u||¯ 2 + ||∂¯∗ u||2 + ||u||2 ,
0 0 0
∼
X0 X X0 X X0
2 2
||∂ωjo ujo K ||20
||∂ u
ωj J 0|| + ||∂ u
ω̄j J 0|| +
|J|=k j≤qo |J|=k qo +1≤j≤n |K|=k−1
j6=jo or j ∈J
/
¯ ¯
2 ∗ 2 2
< ||∂u|| 0 + ||∂ u||0 + ||u||0 ,
∼
Let TR = √i2 (∂ωn − ∂ω̄n ) be the totally real vector field tangential to M .
Since ∂ωjo has finite type mjo , we know that TR is obtained by mjo − 1
iterated commutators of ∂ωjo and ∂ω̄jo . We then have (cf. Section 2.4)
|||TR ujo K |||2−1+ 1 < ||∂ωjo ujo K ||20 + ||∂ω̄jo ujo K ||20 + ||u||20
2
mjo −1 ∼
(1.10.15)
¯ 2 + ||∂¯∗ u||2 + ||u||2 .
< ||∂u||0 0 0
∼
Recall what has already been said: subellipticity follows from finite type
for systems of real vector fields. Therefore, the same is true for systems
of complex vector fields when these are closed under conjugation such
as the system ∂ωjo and ∂ω̄jo (but it is not necessarily true for general
complex systems). We carry out our proof. We reason in the same
way for any index j of the system {∂ωj } ⊂ V ⊥ for which (i) and (ii)
of the statement of Theorem 1.10.4 hold. By changing the metric in
¯
M (which does not affect the ∂-Neumann conditions) we can assume
that this system is orthonormal. We remark now that, u having degree
k ≥ q + 1, any coefficient uJ can be written as uJ = ujK for some index
j of the system or else uJ = unK . We start from the first; they satisfy
(1.10.16)
¯ 2 + ||∂¯∗ u||2 + ||u||2
|||TR uJ |||2−1+ < ||∂u|| for = inf 2−(mj −1) .
0 0 0
∼ j
1.10. SUBELLIPTIC ESTIMATES FOR ∂¯ 75
In this situation, we say that the energy, that is, the quantity in the
right of (1.10.16) contains derivative TR of these coefficients. The
energy also contains a full ∂ωj (resp. ∂ω̄j ) derivative for 1 ≤ j ≤ qo
(resp. qo + 1 ≤ j ≤ n). The missing barred and unbarred derivatives,
except from ∂ωn , can be “-interchanged” with the complemantary ones
on account of
(1.10.17)
|||∂ω uJ |||2−1+ − |||∂ω̄ uJ |||2−1+ ≤ [∂ω , ∂ω̄ ]uJ , uJ + ||uJ ||20
j j j j −1+
This concludes the proof of (1.10.11) for any coefficient uJ such that
n∈ / J. As for the coefficients unK , we have that barred and unbarred
¯
derivatives are fully interchangeable on account of the ∂-Neumann con-
dition unK |M ≡ 0. In other terms, the energy contains the full ||unK ||1 -
norm.
We have a result of subellipticity for an important class of domains,
the so-called “finite type decoupled” domains. We choose q = qo and
write z 0 = (z1 , ...zq ) and z 00 = (zq+1 , ..., zn−1 ).
Theorem 1.10.8. Let Ω be defined by r < 0 for r = 2Re zn −
n−1
h(z 0 ) + |fj (zj00 )|2 where hij (z 0 ) ≥ 0 in Cqo and the fj (zj00 )’s are
P
j=q+1
holomorphic functions with vanishing order mj at zo = 0. Then -
1
subelliptic estimates hold in degree k ≥ q + 1 for = 22m−1 for m =
maxj mj .
Proof. Let π : Cn → Cn−1 be the projection over the first n − 1 coor-
dinates; since h and the fj ’s do not depend on zn , that is Ω is “rigid”,
then we have an identification
π0
∼ ¯ n−1
Lr |T C M → ∂ ∂r|Cz ,...,z .
1 n−1
We take ∂ωj = ∂zj − rzj ∂zn , j = 1, ..., n − 1 and ∂ωn = ∂zn as a basis of
(1, 0) vector fields and its dual system ω1 , ..., ωn = ∂r as a basis of (1, 0)
forms; we choose a Hermitian metric in which this is orthonormal. In
this basis we have
−(hij ) 0
(rij )i,j≤n−1 = .
0 m2j |zj00 |2(mj −1) κij
It readily follows that Ω is q-pseudoconvex and that (1.10.10) is fulfilled.
76 1. SEVERAL COMPLEX VARIABLES
Real Structures
But in the basis ∂x̃i the coefficients on the right side of (2.1.4) are
exactly those of ṽ := Jx x̃ · v. We have thus obtained
∂v → ∂ṽ .
Hence, under coordinate change, derivations and tangent vectors trans-
form accordingly. This allows
P us to identify, as we intensively do, vectors
v and derivations ∂v = j vj ∂xj . As for the basis of differential dxj ,
dual to ∂xj in the inner product h·, ·i, the chain rule applied to the
inverse transformation xj = xj (x̃) yields
X
dxj → ∂x̃i (xj )dx̃i .
i
P
For a general differential, that is, a combination w = j wj dxj , we get
!
X X
(2.1.5) w→ ∂x̃i (xj )wj dx̃i .
i j
But the coefficients on the right are those of (t Jx̃ x)w = t (Jx x̃)−1 w
which we also denote by w̃, and therefore (2.1.5) reads
w → w̃.
We are ready to enlarge the setting of our analysis and deal with the
diffeomorphic images of the euclidean space which is the local structure
of an immersed manifold. In greater generality, a manifold M is the
∼
union of the homeomorphic images Uj → Mj of open sets Uj ⊂ Rn such
Φj
that if Φjk (resp. Φkj ) is the homeomorphism induced by Φj (resp. Φk )
on overlappings Mj ∩ Mk 6= ∅, then
Φjk ◦ Φ−1
kj is a diffeomorphism.
bj = L(x̃j )
X
= ak ∂xk (x̃j ).)
k
P
In the same way, if θ is written in the two systems as θ = j αj dxj and
θ = j βj dx̃j , then the coefficients are related by (βj ) = t (Jx̃ x)(αj ).
P
Lj = F∗ ∂tj , j = 1, ..., n − l,
resp.
θj = (F −1 )∗ dtj , j = 1, ..., n − l.
2.2. INTEGRATION OF VECTOR BUNDLES 83
The last two terms in the second equality belong to E. Thus, we have
the fact that [L, W ] ∈ E is reduced to [Li , Lj ] ∈ E for any i, j.
Definition 2.2.8. A k-dimensional manifold Γ is said to be an
integral leaf of the bundle E when Tx Γ = Ex for any x ∈ Γ.
Our problem is to state under which condition on E there is the
unique integral leaf of E through each point. If E has rank k = 1,
we have already seen that the problem has a positive answer. In the
general situation, the answer is contained in the following
Theorem 2.2.9 (Frobenius theorem). Let E be a bundle on Ω of
rank k and class C m , m ≥ 2. The following are equivalent:
(i) E is involutive.
(ii) For any xo ∈ Ω there is a unique local integral leaf Γ of E
passing through xo .
(iii) For any xo ∈ Ω there are local coordinates x = (x0 , x00 ), x0 =
(x01 , ..., x0k ), such that E = Span{∂x01 , ..., ∂x0k } in a neighborhood
of xo .
Remark 2.2.10. A system of manifolds {Γ} as in (ii), that is, a local
covering of Ω by disjoint manifolds, is said to be a “local foliation” of
Ω. According to (iii), it is possible to reduce a foliation, under a local
coordinate change, to the system of k-planes defined by x00 = const. In
fact, reducing the bundle to Span{∂x0j }j=1,...,k is equivalent to reducing
its integral leaves to x00 = const. Thus, under a C 1 diffeomorphism, the
leaves of a foliation are always “straightenable”. We remark now that
the changes of coordinates which preserve the leaves x00 = const, or the
bundle Span{∂x0j }j=1,...,k , are those in the form
(
x̃0 = x̃0 (x0 , x00 ),
(2.2.4)
x̃00 = x̃00 (x00 ).
We also remark that these changes do not need to preserve each single
∂x0j but, as has already been said, the set of their linear combinations.
This freedom is the key point in the proof of Theorem 2.2.9, which
consists of two steps: linear algebra to adjust the basis {Lj } so that
86 2. REAL STRUCTURES
the terms commute, that is, [Li , Lj ] = 0 and a coordinate change of type
(2.2.4) provided by the solution of Cauchy’s problem, which transforms
Lj for j ≤ k into ∂x0j .
[Li , Lj ] = 0, i, j = 1, ..., k.
As for the second point, the reduction of Lk+1 to ∂x̃k+1 , we are asking
that
X X
aj ∂ x j = ∂x̃k+1 (xj )∂xj ,
j j
that is,
∂x̃k+1 (xj ) = (
aj , j = 1, ..., n,
(2.2.6) 6 k + 1,
x̃j , j =
xj |x̃k+1 =0 =
0, j = k + 1.
Since aj = aj (x00 ), then the solutions of (2.2.6) satisfy x00 = x00 (x̃00 )
which is the crucial point of the proof. These solutions yield the explicit
description of the change of coordinates
(
xj = x̃j + aj (x00 (ξ˜j , x̃00k+2 , ..., x̃00n ))x̃k+1 for j 6= k + 1,
(2.2.7)
xk+1 = ak+1 (x00 (ξ˜k+1 , x̃00k+2 , ..., x̃00n ))x̃k+1 ,
where ξ˜j is the point in the interval (0, x̃00k+1 ) of mean value for
∂x̃k+1 xj (·, x̃00k+1 , ..., x̃00n ); thus (2.2.7) is certainly of type (2.2.5).
We can then solve in the unknown (zih )i=1,...,k the linear systems
k
X
aij zih = κjh , j = 1, ..., k,
i=1
88 2. REAL STRUCTURES
where κjh is the Kronecker symbol. These are a set of k systems with
k
index h = 1, ..., k. We then put L̃h = zih Li . We have
P
i=1
k
X n
X
h
L̃h = zi aij (x)∂xj
i=1 j=1
k k n
!
X X X
= zih aij (x)∂xj + aij (x)∂xj
i=1 j=1 j=k+1
(2.2.8) ! !
k
X k
X n
X k
X
= aij (x)zih ∂xj + aij (x)zih ∂xj
j=1 i=1 j=k+1 i=1
n
X
= ∂xh + chj ∂xj ,
j=k+1
k
aij (x)zih for any j ≥ k +
P
where we have used the notation chj :=
i=1
1. Hence, [L̃i , L̃j ] belongs to Span{∂xk+1 , ..., ∂xn } but it also belongs
to Span{L̃1 , ..., L̃k } which is transversal to the former. It follows that
[L̃i , L̃j ] = 0; we can thus apply Proposition 2.2.11, which yields the
conclusion of the proof of Theorem 2.2.9.
For further application to symplectic geometry we need the follow-
ing variant.
Corollary 2.2.12. Let E = Span{L1 , ..., Ln } be an involutive
bundle and let the constants cijr be defined by the relations [Li , Lj ] =
∞
P
r cijr Lr . Let h1 , ..., hk ∈ C satisfy
X
(2.2.9) Li hj − Lj hi = cijr hr .
r
δ the embedding (x, ξ) 7→ ((x, ξ), (x, ξ)) and hence t π 0 ◦ δ(x, ξ) = ξdx.)
Thus we learned from Exercise 2.3.1 that the forms ω and σ can be
intrinsically defined without use of coordinates. A change of coordinates
˜ that is, a C 1 diffeomorphism of T ∗ M , is said to be
F : (x, ξ) → (x̃, ξ),
“symplectic” when it preserves σ, that is, (F −1 )∗ σ = σ̃.
Note that σ is non-degenerate, that is, if v is a vector of T T ∗ M such
that σ(v, ·) ≡ 0 on T T ∗ M , then v = 0. Let h·, ·i be the duality between
T T ∗ M and T ∗ T ∗ M . Associated to σ, there is therefore a “Hamiltonian”
isomorphism which interchanges the duality with the inner product σ:
(2.3.1) H : T ∗ T ∗ M → T T ∗ M, u 7→ Hu,
where Hu is defined by σ(v, Hu) = hv, ui, for any v ∈ T T ∗ M .
In coordinates (x, ξ) of T ∗ M and for u = u0 dx + u00 dξ, the isomorphism
H is written as
H(u0 dx + u00 dξ) = u00 ∂x − u0 ∂ξ .
For a real smooth function f = f (x, ξ) on T ∗ M , we define its Hamil-
tonian vector field by
Hf = H(df )
= ∂ξ f ∂x − ∂x f ∂ξ .
We define the “Poisson bracket” of two functions on T ∗ M by
{f, g} = Hf (g)
= ∂ξ f ∂x g − ∂x f ∂ξ g.
We have the equality
H{f,g} = [Hf , Hg ],
whose proof is straightforward. In particular, the product {·, ·} enjoys
the same properties as [·, ·]: alternancy, additivity, Jacobi identity. For
˜ we can check whether it is
a change of coordinates F : (x, ξ) 7→ (x̃, ξ),
symplectic just by inspecting if it preserves the Poisson relations
{x̃i , x̃j } = 0, {ξ˜i , ξ˜j } = 0, {x̃i , ξ˜j } = κij .
Note that these relations are trivial for the initial system (x, ξ).
We use the superscript ⊥ σ to denote the orthogonal with respect to
σ.
Definition 2.3.2. A submanifold V of T ∗ M is said to be involutive
if and only if T V ⊥ σ ⊂ T V .
92 2. REAL STRUCTURES
only along the fibers. Thus, the homogeneous map Φ−1 is expressed, in
terms of usual coordinates, by (x, ξ) 7→ (x|ξ|, ξ). In a conic manifold
there is intrinsically defined a “radial” vector field ρ. Its action on a
C 1 function f : Y → R is expressed by
d
(2.3.2) ρ(f ) = Mt∗ f |t=1 .
dt
In particular, in the coordinates of T ∗ M , we have ρ =
P
j ξj ∂ξj . It
is readily seen that a function is homogeneous of degree µ, that is, it
satisfies
Mt∗ f = tµ f,
if and only if it satisfies the Euler relation
ρf = µf.
In particular, in T ∗ M , x and ξ are homogeneous of degree 0 and 1, re-
spectively. When Y is symplectic, in the sense that it is endowed with
a non-degenerate closed 2-form σ, we ask for σ to satisfy the homo-
geneity condition Mt∗ σ = tσ and call Y a conic symplectic manifold.
One can then define a homogeneous 1-form ω by putting
(2.3.3) ω(v) = σ(ρ, v) for any vector field v.
Note that (2.3.3) is equivalent to ω = H −1 (ρ). In other words, in a
symplectic manifold, a homogeneous structure is equivalently defined
by a radial vector field ρ or a canonical 1-form ω. There naturally arises
the question if, σ being closed, it is in fact exact and namely satisfies
σ = dω. Since this is true for T ∗ M , one expects that it is also true for
a general Y . The question is hard and its answer stated positively is
equivalent to saying that Y is symplectically diffeomorphic to T ∗ M .
Theorem 2.3.5 (Darboux-Frobenius theorem). Let Y be a symplec-
tic conic manifold and V a conic involutive submanifold of Y of codi-
mension l in a neighborhood of po = (xo , ξo ). We also assume that V is
regular, that is, ω|V 6= 0. Then we can find local homogeneous symplec-
tic coordinates (x, ξ) in a neighborhood of po , that is, a local symplectic
∼
homogeneous diffeomorphism Y → T ∗ Rn , such that xo = 0, ξo = eio
(the io -th coordinate vector) for io ≤ n − l, and
(2.3.4) V = {(x, ξ) : ξn−l+1 = 0, ..., ξn = 0}.
Cf. Theorem 2.3.7 for the proof.
Remark 2.3.6. (a) In particular, if we choose Y = V , the theorem
says that any symplectic homogeneous manifold Y is locally symplec-
tically diffeomorphic to T ∗ Rn (Darboux theorem).
94 2. REAL STRUCTURES
where the first and second equalities follow from the first and second
lines of (2.3.7), respectively.
We can do the same to find fio for io ≤ n − (l − k) which satisfies
the symplectic relations with the coordinates previously found and such
that dfio = dξio . The proof is complete.
Theorem 2.3.7 says that when rank σ|V ≡ const, then ker σ|V is
integrable. It also provides the explicit straightening of its integral
leaves: under a suitable choice of local symplectic coordinates, they
are reduced to the (l − 2k)-dimensional planes whose tangent bun-
dle is Span{∂xi }i=n−(l−k)+1,...,n−k , that is, the planes defined by xi =
const for i ≤ n − (l − k), xi = 0 for i ≥ n − k + 1, ξj = const for j ≤
n − (l − k), and ξj = 0 for j ≥ n − (l − k) + 1.
(2.4.1) Lj = Span Li , [Li1 , Li2 ], ..., [Li1 , [Li2 , ...[Lij −1 , Lij ]...]]
| {z }
j
for any i, i1 , ..., ij ∈ {1, ..., k} .
2.4. SUBELLIPTIC ESTIMATES FOR SYSTEMS OF VECTOR FIELDS 97
and
Lm
xo
h
lh (xo ) := dim
m .
Lxoh−1
This gives rise to a chain of spaces of iterated commutators which
becomes stationary after, say, r steps:
(2.4.2) L1xo ⊂ L2xo ⊂ ... ⊂ Lm
xo .
r
Notice that the entries of (2.4.2) are not necessarily bundles (constant
rank) but just distributions of tangent vectors.
Definition 2.4.1. The system L1 , ..., Lk is said to be of “finite
type” or to satisfy the “Hörmander condition” at xo when the chain
ends after finitely many, say r, steps with Lm
xo = Txo M .
r
XZ
= (aj (x) − aj (x − t))∂xα g(x − t)∂tj χν (t)
j
+ ∂tj aj (x − t)∂xα g(x − t)χν (t) dt1 ∧ ... ∧ dtn
+ [∂xα , L](g ∗ χν ) − [∂xα , L](g) ∗ χν .
2.4. SUBELLIPTIC ESTIMATES FOR SYSTEMS OF VECTOR FIELDS 99
where c is independent
R of ν. (This
independence Rfollows from the fact
that kχν k = 1 and |t| ∂tj χν dt1 ∧ ... ∧ dtn ≤ |∂tj χ|dt1 ∧ ... ∧ dtn
for any ν.) On the other hand, [∂xα , L] is of order s and hence the H 0 -
norm of the two terms on the last line of (2.4.7) is controlled by kgks
uniformly in ν.
By taking all choices of |α| ≤ s, we get (2.4.6) for s integer. As it
has already been said, this implies (2.4.5), that is, the continuity from
H s to H s of the mapping
(2.4.9) g 7→ L(g ∗ χν ) − L(g) ∗ χν .
Next, by interpolation lemma, the mapping (2.4.9) is continuous also
for s fractionary. This concludes the proof of the Lemma.
Theorem 2.4.5. If L1 , ..., Lk is subelliptic, then it is also hypoel-
liptic.
Proof. (i): Let ψ be a smooth cut-off function such that ψ ≡ 1
at x. We show that ψu ∈ H 0 and Lj (ψu) ∈ H 0 for any j implies
ψu ∈ H . In fact, let us approximate ψu by uν := χν ∗ (ψu) ∈ Cc∞ ; by
Lemma 2.4.4 applied with s = 0 and g = χu, we have that uν → ψu
and Lj uν → Lj (ψu) in H 0 -norm. From
X
||uν − uµ ||2H < ||Lj uν − Lj uµ ||2H 0 + ||uν − uµ ||2H 0 ,
∼
j
We notice now that if the system L1 , ..., Lr is not of finite type but
the Lie span L has locally constant rank a < n, then in new variables
L = Span{∂x0j }1≤j≤a ; this follows from the Frobenius theorem. Any
function u of x00j , j ≥ a + 1 only, is then a solution of the system
Lj u = 0 but it is not necessarily regular. Thus the sufficient condition
in next theorem is also essentially necessary.
1
Theorem 2.4.6. Let L1 , ..., Lk be of finite type m. Then it is 2m−1
subelliptic (and hence in particular hypoelliptic).
≤ |hL2 u, [L1 , Λ−2+2 [L1 , L2 ]]uiH 0 | + |hL2 u, Λ−2+2 [L1 , L2 ]L1 ui|
| {z } | {z }
order −1 + 2 order −1 + 2
(
||L1 u||2H −1+2 + ||L2 u||2H 0 + ||u||2H 0
< ||L2 u||2H −1+2 +
∼ ||L2 u||2H −1+2 + ||L1 u||2H 0 + ||u||2H 0 ,
where we can choose either of the two alternatives in the third line
of (2.4.11). When m = 2, we let = 21 , choose either of the alterna-
tives and get (2.4.10). Otherwise, we reason by induction and suppose
(2.4.10) true for commutators of m − 1 vector fields. When we go on to
m, we have to use the second alternative in (2.4.11) with L2 replaced
by [L2 , [L3 , [..., Lm ]]]. Notice that, since we get a factor 2 for each com-
mutator, the choice = 2−(m−1) is the one needed in the induction. We
then conclude
||[L1 , [L2 , [..., Lm ]]]u||2H −1+
< ||[L2 , [..., Lm ]]u||2H −1+2 + ||L1 u||2H 0 + ||u||2H 0
∼
X
< ||Lj u||2H 0 + ||L1 u||2H 0 + ||u||2H 0 .
∼
2≤j≤k
End of Proof of Theorem 2.4.6. We are assuming that L1 , ..., Lk
have finite type, say, m. Hence all derivatives of T M can be expressed
through commutators involving at most m vector fields Lj ’s. Hence,
(2.4.10) yields the subelliptic estimate for = 2−(m−1) .
Remark 2.4.8. If L1 , ..., Lk has finite type m, then u “gains” =
2−(m−1) derivatives with respect to (Lj u, j = 1, ..., k). A more efficient
proof shows that we have, in fact, the better gain of any with < m1
(cf. [75]). However, for application to hypoellipticity, this improvement
is irrelevant.
Remark 2.4.9. We spend a few words about the “complex” ver-
sion of Theorem 2.4.6 and Lemma 2.4.7. If a system of vector fields
Lj ∈ T 1,0 Cn ⊕ T 0,1 Cn is closed under conjugation then, identified to
a system in T R2n , it can enjoy the conclusions of Lemma 2.4.7 and
Theorem 2.4.6. Otherwise, Lemma 2.4.7 holds only if an amount of
102 2. REAL STRUCTURES
kL̄j uk2H 0 norm is added on the right. On the other hand, on account
of the estimate kL̄j uk2H 0 < kLj uk2H 0 + kuk2 1 , the norm kuk2 1 is ab-
∼ H2 H2
sorbed in the left side provided that m = 2 (and taking summation
over all bracket of order ≤ 2). This explains why Theorem 2.4.6 stays
true in complex version when m = 2 and, instead, it fails for m > 2
(cf. Kohn [60]).
It follows that, if L is the Lie span of the Ljµ ’s, then L0 (x0 ) ⊂ L(0, x0 ).
0
Thus, let Γ0 be the integral leaf of L0 in the x0 -space Rn−1 which exists
by the inductive hypothesis. Then (−, +) × Γ0 is the integral leaf of
L in Rn .
When M or X is not C ω , we cannot expect the existence of leaves
such that T Γ = L(X)|Γ . Instead, we can look for manifolds Γ of minimal
dimension which satisfy
(2.5.1) Tx Γ ⊃ Xx for any x ∈ Γ.
As was already noticed, since T Γ is stable by linear combinations and
commutators, (2.5.1) implies
(2.5.2) Tx Γ ⊃ Lx (X) for any x ∈ Γ.
It turns out that, if M and X are C ∞ , there are no more Frobenius
or Nagano leaves but, nevertheless, there exist locally the manifolds of
minimal dimension which satisfy (2.5.2). To describe them, we need
new terminology.
Definition 2.5.2. A “piecewise smooth integral curve”, or “polyg-
onal integral path” of X in M , is the junction of finitely many integral
curves γ1 , ..., γs of smooth sections of X with each γj being issued from
the ending point of the former γj−1 .
Definition 2.5.3. We call the set of all ending points of piecewise
smooth integral curves of X issued from xo the “orbit” of X through
xo ∈ M and denote it by O(xo , X, M ). We call the set of the ending
points of the piecewise smooth integral curves contained in B the orbit
of X in a neighborhood B of xo ∈ M and denote it by O(xo , X, B).
Some examples are in order.
Example 2.5.4. Let X = Span{L1 , L2 } in R3 where L1 = ∂x1 and
L2 = ∂x2 + x1 x3 ∂x3 . Note that [L1 , L2 ] = x3 ∂x3 . It follows that
(2.5.3) rank X = 2,
3 for x3 6= 0,
(2.5.4) rank L(X) =
2 for x3 = 0.
104 2. REAL STRUCTURES
↑ p1 p1 ↑
C
P ' ∼2
↑ p2 p2 ↑
C
Π ' ∼3
.
We explain p1 and p−1 n
1 . A projection h : R → S to a submanifold
S induces a foliation defined by its fibers {h−1 (x)}x∈S ; a foliation and
a choice of a transversal manifold of dimension complementary to the
fibers give a projection which assigns to each x the point h(x) in which
the leaf through x meets S. The mapping p2 is clear by itself.
Example 2.5.12. We take two local foliations R = {R} and S =
{S} with leaves of dimension l + k and l, respectively. We claim that
2.5. MISCELLANEA: FOLIATIONS—ORBITS 107
Real/Complex Structures
and the dual identification for cotangent vectors, that is, forms,
(3.1.4)
a − ib a + ib
T ∗ R2n → T ∗ Cn ×T ∗ Cn T ∗ C̄n , adx + bdy 7→ dz + dz̄.
2 2
Under (3.1.3) and (3.1.4), the duality between T R2n and T ∗ R2n in the
left hand side is interchanged with the real part of the duality between
T Cn and T ∗ Cn on the right hand side
ha∂x + b∂y , a0 dx + b0 dyi = Re h(a + ib)∂z , (a0 − ib0 )dzi,
that is,
aa0 + bb0 = Re ((a + ib) · (a0 − ib0 )) .
This identifies the euclidean structure of R2n with the real part of the
hermitian structure of Cn . After we have extracted the real structure
R2n underlying Cn , we want to complexify it, that is, to go on to C2n .
We first revisit the above construction in an intrinsic way and prefer to
reason on a complex manifold X , a union of local holomorphic patches
∼
{Xj } parametrized by means of homeomorphisms Uj → Xj over open
Φj
domains Uj ⊂ Cn which are holomorphically compatible over over-
lappings Ui ∩ Uj . We denote by X̄ the conjugate, the manifold with
local holomorphic parametrizations Φ̄j : Ūj → X̄j , and define the real
underlying manifold X R through the diagonal identification
∼
(3.1.5) X R → X ×X X̄ .
δX
fields and forms or to use the mapping j 0 and t j 0−1 . The two ways are
related by the diagonal correspondences
δ 0C
(3.1.10) C ⊗R T X R → T X ×X T X̄
and
t δ 0−1C
(3.1.11) C ⊗R T ∗ X R → T ∗ X ×X T ∗ X̄ .
If we wish to express (3.1.10) and (3.1.11) in coordinates, we have just
to use (3.1.3), (3.1.4) and there replace real coefficients a and b by
complex ones. We note that the “diagonal” relations a − ib = a + ib
correspond precisely to the choices of a and b real.
Definition 3.1.1. A complex structure on X is the morphism J
induced on T X R by the multiplication by i on T X :
i
TX → TX
|| ||
J
TX R
→ T X R.
Note that we have the identity J 2 = −id as the real counterpart
of i2 = −1. Also, in T X R , the complex structure is given in local
holomorphic coordinates z = x + iy of X by J ∂x = ∂y , J ∂y = −∂x .
With the aid of J we can write (3.1.10) and (3.1.11) as
(3.1.12)
ξ + i t J −1 ξ ξ − i t J −1 ξ
X − iJ X X + iJ X
X 7→ , , ξ 7→ , ,
2 2 2 2
with inverses
(3.1.13) (Z, W̄ ) 7→ Z + W̄ , (ζ, µ̄) 7→ ζ + µ̄.
When the mappings in (3.1.13) are restricted to the diagonal, they
take values in T X R and T ∗ X R , respectively. We denote by “Re 00 the
composition of (3.1.13) restricted to the diagonal with (π1 |T X ×T X T X̄ )−1
and (π1 |T ∗ X ×T X T ∗ X̄ )−1 , respectively:
(3.1.14)
Re : T X → T X R , Z 7→ Z + Z̄, Re : T ∗ X → T ∗ X R , ξ 7→ ξ + ξ. ¯
We remark that in C ⊗R T X R there are two “distinguished” bundles
T 1,0 X = { X−iJ
2
X
: X ∈ T X R } and T 0,1 X = { X+iJ
2
X
: X ∈ T X R },
the bundles of holomorphic and antiholomorphic vector fields or of type
(1, 0) and (0, 1), respectively. They satisfy
(3.1.15) T 1,0 X ∩ T 0,1 X = {0}, T 1,0 X ⊕ T 0,1 X = (T X R )C .
They can be characterized as the eigenspaces of the eigenvalues +i and
−i for J C , respectively. In coordinates, they are described as T 1,0 X =
112 3. REAL/COMPLEX STRUCTURES
(J L, J h∗ L) = (W, h∗ W ),
h∗ J L = J h∗ L.
3.2. CR manifolds
F ∩ J F ⊂ F ⊂ F + J F,
116 3. REAL/COMPLEX STRUCTURES
where the term on the left (resp. the right) is the maximal C-space
contained in F (resp. the minimal C-space containing F ). We set
dimCR (F ) = dimC (F ∩ J F ).
It is clear that
(3.2.1) dimR F = dimC (F ∩ J F ) + dimC (F + J F ),
which is obtained by dividing by 2 the two sides of
(3.2.2) dimR F + dimR J F = dimR (F ∩ J F ) + dimR (F + J F ).
On account of (3.2.1) we have
dimR F
(3.2.3) dimR F − n ≤ dimCR F ≤ .
2
We set l = codRCn F, s = codCCn (F + JF ); one can check that if rkR and
rkC are the real and complex rank of a system of (linear) equations for
F , respectively, then l = rkR and s = l − rkC , which yields
dimCR F = n − l + s.
There are two extremal cases:
(
s = 2l , dimCR F = n − 2l = dimF
2
,
(3.2.4)
s = 0, dimCR F = n − l.
The first corresponds to F being complex. The second means that
F + J F = Cn and F is said to be “generic” in that case. In the sequel,
we refer to the number s as the “defect of genericity” of F . Note that
when F is generic, then dimR F ≥ n.
Lemma 3.2.2. For a pair of linear subspaces F, G ⊂ E the following
are equivalent:
(i) dimR F = dimR G and dimCR F = dimCR G.
∼
(ii) There is an isomorphism ϕ : F → G which is C-complex, that
is, ϕ ◦ J = J ◦ ϕ.
Proof. It is not restrictive to assume
(
F + J F = G + J G = E,
(3.2.5)
F ∩ J F = G ∩ J G.
We choose a basis e1 , . . . , en−l , J e1 , . . . , J en−l of F ∩ J F = G ∩ J G
and two completions f1 , . . . , fl and g1 , . . . gl for F and G, respectively.
Thus, we can explicitly define our morphism ϕ by
ϕ(ej ) = ej , ϕ(J ej ) = J ej , ϕ(fj ) = gj .
Such a ϕ fulfills all requirements of the statement.
3.2. CR MANIFOLDS 117
We note that (3.2.6) and (3.2.7) are stable: if they hold at zo , they
hold also in a neighborhood of zo . This is clear for (3.2.6), which is a
condition of transversality T M ∩ J T M = {0}. As for (3.2.7), we have
to remark that this is the condition of C-independence of a system of
equations of M : ∂r1 ∧ · · · ∧ ∂rl 6= 0. Another distinguished situation is
given by
= hk (x0k , z̃ 00 , 0) − hk x0k + ok , z̃ 00 , 0 + ok
= ok .
Remark 3.2.8. There is a C ω version of the above statement. In
that case the equations yj0 = hj can be found such that (3.2.12) holds
for any I, J, K and not just for those whose sum of lengths is ≤ k.
122 3. REAL/COMPLEX STRUCTURES
equations
z2 = |z1 |2 + γ(z12 + z̄12 ) + O3 (x3 , ..., xn , z1 ),
y = O3 (x , ..., x , z ),
3 3 n 1
(3.2.23)
...,
yn = O3 (x3 , ..., xn , z1 ).
Note that the xj , j = 3, ..., enter into the error term of the first equa-
tion because by a holomorphic transformation of type z̃1 = z1 + zj , j =
3, ..., terms like zj z̄1 are absorbed by |z̃1 |2 . Also, |z1 |2 enters into the
errors of the remaining equations because, by the first,
|z1 |2 = z2 − 2Re (γz12 ) + O3
= Re (z2 − 2γz12 ) + O3 .
In (3.2.23), the first error term O3 is complex and the others are real.
We write the first line as z2 = h and assume γ < 21 . We note that
the set of the complex tangencies forms an (n − 2)-dimensional real
submanifold. In fact, they are defined as the 0-set of the determinant
of the system
(3.2.24)
∂z (z2 − h) −(z̄1 + 2γz1 ) 1 ∗ ∗ · ∗
∂z (z̄2 − h̄) −(z̄1 + 2γ z̄1 ) 0 ∗ ∗ · ∗
i
∂z (y3 − O3 ) = ∗ 0 −2 ∗ · ∗
+ error.
... ... ... ... ... ... ...
∂z (yn − O )3
∗ 0 ∗ ∗ · − 2i
This determinant is
i
(− )n−2 (z̄1 + 2γz1 ) + error,
2
and therefore it has only one zero near z1 = 0.
We pass to the general case and suppose that M is generic of CR
dimension m apart from some CR singularities where the CR dimension
grows to m + 1. We divide the variables as (z1 , z2 , z 0 , z 00 ); by (3.2.23)
and Theorem 3.2.7, we can put the equations of M in the form
(
z2 = q(z1 , z 00 ) + O3 ,
y 0 = Q(z1 , z 00 ) + O3 .
We assume that zo = 0 is a non-degenerate isolated CR singularity in
the (z1 , z2 )-plane. By the discussion of the C2 case, we have q(z1 , 0) =
|z1 |2 + γ(z12 + z̄12 ). We point out the following.
• The points of higher-dimensional complex tangencies, that is,
the points where dimCR M = m + 1, do not necessarily form
124 3. REAL/COMPLEX STRUCTURES
We define
Y := k(M C ).
We wish to prove that
T Y |M = k 0 (T M C )|M = T M + iT M ;
in particular, T Y having constant (complex) rank, Y is a (complex)
manifold and fulfills the requirements in the statement. In fact, we
have
(3.3.7) δ C0 T M C = {α∂z + β̄∂z̄ : α∂z r + β̄∂z̄ r = 0},
We then have
i z1
∂¯b = − ∂z̄1 − ∂x2 ,
2 2
as we can check from
i z1 z1
Φ∗ (∂¯b ) = − ∂z̄1 + (∂z2 − ∂z̄2 ) − (∂z2 + ∂z̄2 )
2 2 2
= L̄.
We refer to the manifold Y of the preceding theorem as an “intrinsic
complexification” of M . It can be characterized by
Theorem 3.3.9. Let M be C ω and Y be a complex manifold con-
taining M and such that T Y |M = T M + iT M . Then, any f which is
CR and C ω , extends uniquely as holomorphic to Y in a neighborhood
of M .
Proof. Let f C : M C → C be the complexification of f , defined,
e.g., through a parametrization Φ of M and its complexification ΦC .
Our hypothesis that f is CR ensures that L̄f C = 0 for any L̄ ∈ T 0,1 Cn ∩
T M C . Thus, f C is constant on the fibers of k defined by (3.3.6). Hence,
there is a well-defined holomorphic function f˜ on Y = kM C given by
f˜(z) = f C (z̃) for z̃ ∈ M C with k(z̃) = z.
In general, if M is not C ω , we cannot expect that there is a complex
submanifold Y ⊂ Cn in which M is embedded as generic. However,
this is true if we allow a projection instead of an immersion. Before
discussing this, we recall that we have already treated the inversion
problem for holomorphic mappings: the inverse is holomorphic, if it ex-
ists. The same is not true for CR mappings. For instance, the mapping
R × R → C, (x, y) → z = x + iy is one-to-one and CR but its inverse
Re × Im is not CR. However, it is an instructive exercise to verify that
if dimCR M = dimCR M 0 , then any f : M → M 0 which is CR and is a
C 1 -diffeomeorphism is in fact a CR diffeomorphism, that is, f −1 is also
CR. With this preliminary we can state
Theorem 3.3.10. Let M ⊂ Cn be CR. Then, there is a complex
manifold Y and a complex projection j : Cn Y such that
(
j is a CR diffeomorphism between M and M 0 := jM ,
T Y |M 0 = T M 0 + iT M 0 .
Proof. We choose Y as the complex plane Tzo M + iTzo M , identi-
fied in a choice of coordinates to a plane of Cn , and choose a transversal
3.3. CR FUNCTIONS AND CR MAPPINGS 131
where each aji is C ω and C-valued. Since rkC {L̄j }j=1,...,m is m, we can
assume, by a linear change of coordinates, that the m × m minor A :=
(aji )j,i≤m is non-degenerate. We decompose the coordinates as
x = (x0 , x00 ), x0 = (x1 , ..., xm ), x00 = (xm+1 , ..., x2m+l ).
A linear change of coordinates x̃0 = A−1 x0 has the effect of multiplying
the coefficients of the ∂x0 ’s in L̄j by A−1 , thus giving
2m+l
X
L̄j = ∂xj + ãji ∂x00i .
i=m+1
It follows that
[L̄j , L̄i ] ∈ Span{∂x00i }i=m+1,...,2m+l .
On the other hand, we have
[L̄j , L̄i ] ∈ Span{L̄j }j=1,...,m ,
by the assumption of involutivity. Since the two above “Span”’s are
transversal one to another, we conclude that
[L̄j , L̄i ] = 0 for all i, j,
that is, we have found a basis of vector fields for L̄ which commute.
We now complexify R2m+l , with variable x, to C2m+l , with variable z,
3.3. CR FUNCTIONS AND CR MAPPINGS 133
and extend holomorphically the coefficients from aji (x) to aji (z). We
define
X
L̄Cj = ∂zj + aji (z)∂zi .
i
M̃ := Ψ(Φ−1 (M )).
We first consider the Levi form through the so-called “intrinsic ap-
proach” which applies to any abstract CR structure (M, L) as well. By
definition, L and L̄ are involutive, that is, [L1 , L2 ] ∈ L and [L̄1 , L̄2 ] ∈ L̄
for any pair of vector fields L1 , L2 ∈ L. Also, [L1 , L̄2 ] belongs to C⊗T M
but it does not necessarily belong to L ⊕ L̄. The Levi form precisely
measures how much L ⊕ L̄ fails to be involutive. Let
C ⊗ TM
p : C ⊗ TM →
L ⊕ L̄
1
LM (zo )(u) := p[L, L̄]zo ,
i
= p[N, N̄ ]|zo .
Proposition 3.4.3. L ⊕ L̄ is involutive if and only if LM ≡ 0.
Proof. The “only if” part is obvious. “If”: we have to show that if
[L, L̄] ∈ L ⊕ L̄ for any L ∈ L, then [L1 , L̄2 ] ∈ L ⊕ L̄ for any L1 , L2 ∈ L.
Now, we have
[L1 + L2 , L1 + L2 ] = [L1 , L̄1 ] + [L2 , L̄2 ]
(3.4.1)
+ [L1 , L̄2 ] − [L1 , L̄2 ] .
We also have
[L1 + iL2 , L1 + iL2 ] = [L1 , L̄1 ] + [L2 , L̄2 ]
(3.4.2)
− i([L1 , L̄2 ] + [L1 , L̄2 ]).
By (3.4.1) and (3.4.2), LM ≡ 0 implies [L1 , L̄2 ] − [L1 , L̄2 ] ∈ L ⊕ L̄
and [L1 , L̄2 ] + [L1 , L̄2 ] ∈ L ⊕ L̄, respectively, and hence in conclusion
[L1 , L̄2 ] ∈ L ⊕ L̄.
We wish to discuss in further detail LM when M is an embedded
CR generic manifold M ⊂ Cn . In this case L = T 1,0 M, L̄ = T 0,1 M are
M
both involutive bundles and C⊗T L⊕L̄
= C ⊗ TTCMM is the complexification
of the totally real part of the tangent bundle. The Levi form is the map
Tzo M
LM (zo ) : Tz1,0
o
M → TzCo M
,
1
u 7→ i
p[L, L̄]|zo ,
for L(zo ) = u. We want to revisit Proposition 3.4.3 in this situation.
What must be said is that T 1,0 M and T 0,1 M are involutive whereas
T C M is not, in general. Though these bundles can all be identified
∼ ∼
by the isomorphisms Re : T 1,0 M → T C M and Re : T 0,1 M → T C M ,
nevertheless Re [·, ·] 6= [Re ·, Re ·]. In detail, let M be defined by r1 =
0, ..., rl = 0 with ∂r1 ∧...∧∂rl 6= 0 and write r = t (r1 , ..., rl ). Take a pair
of vector fields L1 , L2 in T 1,0 M ; these are characterized by Lj (ri ) = 0
136 3. REAL/COMPLEX STRUCTURES
Also, we have R(r) = 0 for any such R if and only if (∂zi ∂z̄j r)i j |T C M ≡ 0,
that is, if and only if M is “Levi flat”. On the other hand, the condition
R(r) = 0 for any R characterizes the involutivity of T C M . We have
thus obtained the proof of the following statement which rephrases
Proposition 3.4.3.
Theorem 3.4.4. T C M is involutive if and only if M is Levi flat.
Indeed, what we proved is more than we stated. Whether M is
Levi flat or not, the vector field R + R̄ obtained by commutation of
Re L1 ∈ ker LM and Re L2 ∈ ker LM still belongs to ker LM .
Theorem 3.4.5. ker LM is involutive.
Thus, Theorem 3.4.4 is a corollary of Theorem 3.4.5 applied for
ker LM = T C M . For an embedded submanifold M ⊂ Cn , we want to
develop further the relation between LM and the Levi form that we
already introduced in Section 1.6 through a system of equations of M .
We note that, in case M is generic, the natural projection combined
with J , i.e.,
TM
→ TM Cn , [u] mod T C M 7→ [J u] mod T M,
T CM
3.4. LEVI FORM OF EMBEDDED AND ABSTRACT STRUCTURES 137
T M J∼ K ∗
→ TM Cn → TM Cn .
T CM
∗
(Here, as always, ∂rj are identified to sections of TM Cn by the aid of the
∼
correspondence 2Re ∂rj → drj .) We define, in analogy to Section 1.6
where the hypersurface case was treated, the “extrinsic Levi” form LM
by
X
(3.4.3) LM (u, ū) = ¯ j (u, ū)∂rj .
∂ ∂r
j
¯ j , L ∧ L̄i = ∂ ∂r
h∂rj , [L, L̄]i = h∂ ∂r ¯ j (u, ū).
Remark 3.4.7. We have already encountered this circumstance in
¯ is alternant on T 1,0 ⊕T 0,1 , but, when restricted
Section 1.9: the form ∂ ∂r
to (T ⊕ {0}) × ({0} ⊕ T 0,1 ), it becomes hermitian.
1,0
138 3. REAL/COMPLEX STRUCTURES
∗
Cn )zo ∼ Rl and set rξ := j ξj rj ;
P
Definition 3.4.8. Let ξ ∈ (TM
the “microlocal” Levi form is the 1-component hermitian form defined
by
and
n
X
σ = dω = dζj ∧ dzj
j=1
3.5. REAL/COMPLEX SYMPLECTIC SPACES 139
¯ = i∂∂r
purely imaginary, that is, ∂∂r ¯ I on vectors (ż + ż, ż 1 + ż 1 ). More
precisely
¯ ż + ż, ż 1 + ż 1 ) = 2iLr (ż, ż 1 ).
∂∂r(
Also, in the first term in the right of (3.5.3), we have ∂rR |T M = 0.
Thus (3.5.3) shows once more that σ R |T TM∗ Cn = 0. Moreover, if we
restrict (3.5.3) to vectors which are not only in T M but also in T C M ,
then the first term on the right is 0 and we get
(3.5.4) σ|T(zo ,ξo ) TM∗ Cn ∩π−1 (T C M ) = 2iLrξo |TzCo M .
In particular, this restriction applies to ker σ|T TM∗ Cn which is contained
in π −1 T C M . Thus (3.5.4) shows that the projection π induces an iden-
tification
∼
(3.5.5) ker σ → ker LM .
Before going ahead with our discussion, we want to relax and have an
alternative proof of what we have seen so far.
Guided Exercise 3.5.1. Prove (3.5.2) and (3.5.5) in coordinates.
(a) Start by proving
∗
(3.5.6) T(zo ,ξo ) TM Cn = {(ż, ζ̇) : ż ∈ Tzo M, ζ̇ = ξ∂r(zo )
+ ∂ż ∂rξo (zo ) + ∂ż ∂rξo (zo ) for ξ ∈ Rl }.
(Here (3.5.6) is obtained by differentiating the parametrization M ×
∗
Rl → TM Cn , (z, ξ) 7→ (z, ξ∂r(z)).)
(b) Describe the relation between σ|T TM∗ Cn and LM . First prove that
σ(zo , ξo ) = d(ξ∂r)(zo , ξo )
¯ ξo (zo ).
= dξ ∧ ∂r(zo ) + ∂∂r
1
Next, observe that for any ξ, ξ 1 ∈ Rl , ż+ż ∈ Tzo M and ż 1 +ż ∈ Tzo M ,
we have
1 1
hd(ξ∂r)(zo , ξo ) , (ξ, ż+ż)∧(ξ 1 , ż 1 +ż )i = 2i(∂rξI ż 1 −∂rξ I ż+Lrξo (ż, ż 1 )).
1
Conclude that if ż + ż ∈ TzCo M and ż 1 + ż ∈ TzCo M , then
1
hd(ξ∂r)(zo , ξo ) , (ż + ż) ∧ (ż 1 + ż )i = 2iLrξo (ż, ż 1 ).
This proves once more (3.5.4).
(c) Carry out the proof. Show that if
∗
(ż, ζ̇) = i(ż 1 , ζ̇ 1 ) ∈ T TM Cn ,
3.5. REAL/COMPLEX SYMPLECTIC SPACES 141
∗
Proof. Let V1 be a partial or intrinsic complexification of TM Cn
∗ n
in T C (cf. Section 3.3) and let f1 = 0, ..., fso = 0 be a system of
holomorphic equations which define V1 in T ∗ Cn . We choose ζ̃n−so +1 =
f1 , ..., ζ̃n = fs0 and supplement to a full system of coordinates according
to the obvious holomorphic version of Theorem 2.3.5. In the new coor-
0 0 0
dinates, V1 is interchanged with T ∗ Cn × Cs and TM ∗
Cn with Λ0 × Cs
where Λ0 is an R-lagrangian submanifold. But it is in fact I-symplectic
0 0 0
because (Λ0 ×Cs )⊥ = {0}×Cs . By choosing new coordinates in T ∗ Cn
such that Span{∂ζj0 }j=1,...,n−s0 ∩T(zo ,ζo0 ) Λ = ζo0 R, the real line engendered
n0
by the Euler vector field, we achieve that Λ0 = TM ∗
0C for a Levi non-
0 n0
degenerate hypersurface M ⊂ C .
Theorem 3.5.3 is the main tool in solving the problem that we de-
scribe now. Let ξo be a conormal to M at zo and assume rk LM (z, ξ) ≡
const in a neighborhood of (zo , ξo ): is there a way to find a hypersur-
face S ⊃ M with conormal ξo at zo such that s0S = s0M ? If M is Levi
non-degenerate, defined, say, by r1 = 0, ..., rl = 0, we may define our
“extension” S of M for instance by
Notes. The relation between the symplectic and the Levi forms
is classical and very fruitful in the geometric study of PDE’s. We fol-
lowed here the guidelines of Sato, Kashiwara and Kawai [78] and Kashi-
wara and Schapira [53]. The result on the straightening of a Levi foli-
ation in the symplectic space (Theorem 3.5.5) follows the presentation
by Zampieri in [102]. The non-straightening in the base space is dis-
cussed, among others, by Ebenfelt in [41]. The deformation argument
contained in Theorem 3.5.3 is developed by Zampieri in [102] as well
as the non-vanishing of the tangential cohomology contained in Theo-
rem 3.5.5. The latter was already obtained by Andreotti, Fredricks and
Nacinovich [5] in case the Levi form is non-degenerate; note that, in
their setting, M need not be C ω .
role in Sections 3.7, 3.8 and 3.10 where the holomorphic extension of
CR functions is studied.
We have used here the symbol “∨” to mean that the corresponding item
is omitted. Since every differential operator P on M is a combination
148 3. REAL/COMPLEX STRUCTURES
of the Xj ’s and the L̄j ’s, then we can define t P starting from (3.6.2).
With this preliminary, we are in a position to define “weak solutions”
of the tangential CR equations.
Definition 3.6.1. We say that u is a CR distribution, and we write
0
u ∈ DCR , when L̄u = 0, for all L̄ ∈ T 0,1 M , in the sense that
(3.6.3) hu, t L̄(v)i = 0 for any v ∈ Cc∞ ,
where h·, ·i denotes the duality between D0 and Cc∞ .
The adjunction being made under a choice of the basis Xj , L̄j , the
1
R tdefinition is not invariant. If u ∈ Lloc , thenV(3.6.3) is equivalent
above V
to u L̄(v)α = 0 for any L̄ and v where α = dξj ∧ dtj
j=1,...,n j=1,...,n−l
which declares the dependence of the condition on the choice
R of the
0
density α. In particular, we can characterize u ∈ CCR by udβ = 0
for any β with compact support of type (n, n − l − 1), that is, of type
P ∨
β = j βj ...dtj ....
n2
λ 2
(3.6.4) Tλ u(z) := hu(ξ)|N , e−λ(z−ξ) i, z ∈ N.
π
Z Z
Tλ uvdξ = hu, Kλ ivdξ
Z
= hu, Kλ vdξi = hu, Tλ vi.
150 3. REAL/COMPLEX STRUCTURES
Thus, what we have to show is that for any totally real maximal
submanifold N ⊂ M and for a suitable N 0 ⊂ M , we have Tλ u ⇒
u uniformly on N 0 . We first consider the case N = Rn (and next de-
form it by a diffeomorphism). We start from the identity
n2 Z
λ 2
(3.6.5) e−λ(x−ξ) dξ = 1.
π Rn
√
By using (3.6.5) and by means of the coordinate change s := λ(x − ξ)
for x ∈ Rn , we get
n2 Z
λ 2
Tλ u(x) − u(x) = e−λ(x−ξ) [u(ξ) − u(x)] dξ
π Rn
(3.6.6) n2 Z
1 −s2 s
= e u(x − √ ) − u(x) ds.
π Rn λ
The term between bracket, which we denote by uλ , satisfies uλ → 0
2 2
and e−s |uλ | < e−s ; thus, by the Lebesgue dominated convergence
∼
theorem, the integral converges to 0 for fixed x. We have to show that
this convergence is indeed uniform. In fact,
Z Z
|Tλ u − u| < ·+ ·
∼ |s|≥R |s|≤R
−1
= (R ) + R (λ−1 ),
where the first is infinitesimal uniformly in λ. The uniformity follows
and the case N = Rn is proved.
We take now a general, totally real maximal, submanifold N with
a local, orientation-preserving parametrization over Rn :
Rn → N, ξ 7→ Φ(ξ).
We decompose the n × n complex Jacobian matrix Φ0 as Φ0 = B + iC
and suppose C(0) = 0 in our normalization. We need
3.6. APPROXIMATION OF CR FUNCTIONS BY POLYNOMIALS 151
We denote by [∗] the term between brackets in the second to last term
of (3.6.8). We can assume that
(
Re (Φ(x) − Φ(ξ))2 ≥ c|ξ − x|2 − c1 ,
(3.6.9)
Re (Φ0 (x) · (ξ − x))2 ≥ c|ξ − x|2 − c1 .
It follows that
e−λ(Φ(ξ)−Φ(ξ))2 [∗] < e−λ|ξ−x|2 E(|ξ − x|),
∼
(3.6.10)
e−λ(Φ(ξ)−Φ(x))2 − e−λ(Φ0 (x)·(ξ−x))2 < e−λ|ξ−x|2 E(|ξ − x|),
∼
152 3. REAL/COMPLEX STRUCTURES
where E(|ξ − x|) is infinitesimal (in fact we could replace E(|ξ − x|) by
O(|ξ − x|) in the second line). By a similar argument to that which
follows (3.6.6), we see that
n2 Z
λ 2
(3.6.11) lim e−λ|ξ−x| Edξ = 0.
λ→∞ π Rn
√
To prove it, we just have to set s = λ|ξ − x| and apply the domi-
nated convergence theorem. By (3.6.8), (3.6.10) and (3.6.11) we have
Tλ u(z) → u(z) as λ → ∞. Next, by the same argument as for N = Rn ,
we conclude that Tλ u ⇒ u uniformly on a suitable N 0 ⊂ N . Now,
we want to show that this convergence also takes place out of N 0 in
a controlled neighborhood M 0 of N 0 in M . For this, we remark that
any point z in M 00 ⊂⊂ M 0 can be reached by a deformation Nz0 of N 0 ,
parametrized over Rn by Φz such that
(
N 0 and Nz0 coincide near ∂M 0 ,
|Im Φ0z | < 21 .
Now,
Z
1
· + (ũ(z1 ) − ũ(z2 ))
2πi ∂∆\B2 (z1 )
Z
1
= (ũ(τ ) − ũ(z1 ))
2πi ∂∆\B2 (z1 )
1 1
× − dτ
τ − z1 τ − z2
Z
1 (−ũ(z1 ) + ũ(z2 ))
+ dτ + (ũ(z1 ) − ũ(z2 )).
2πi ∂∆\B2 (z1 ) τ − z2
1 1
R
Because of the identity 2πi ∂∆ τ −z2
dτ = 1, the sum of the second and
third terms in the right side is B2 (z1 )∩∂∆ ũ(z1τ)−ũ(z 2)
R
−z2
dτ , and therefore
α
its absolute value is estimated by ||ũ||α |z1 − z2 | . The absolute value of
the first term on the right side can be estimated by
Z
<
||ũ||α |z1 − z2 | 2|θ − θ1 |−2+α dθ ∼ ||ũ||α |z1 − z2 |α .
∂∆\B2 (z1 )
As for the remaining integral over ∂∆ ∩ B2 (z1 ), we have the estimate
Z Z
1 <
· ∼ ||ũ||α (|θ − θ1 |α−1 + |θ − θ2 |α−1 )dθ
2πi ∂∆∩B2 (z1 ) ∂∆∩B2 (z1 )
<
∼ ||ũ||α |z1 − z2 |α .
We are ready for the construction of discs. Let M be defined by the
system of equations
yj0 = hj (x0 , z 00 ), hj (0) = 0, ∂hj (0) = 0.
We fix a point zo = (zo0 , zo00 ) ∈ M and prescribe the “z 00 ” component
of our discs as a holomorphic function w(τ ), τ ∈ ∆, whose boundary
value is small in C k,α -norm. We look for a disc A, with A(1) = zo , in
the form A = zo + (u(τ ) + iv(τ ), w(τ )).
Proposition 3.7.2 (Solution of Bishop’s equation). Let M be de-
fined by yj0 = hj with h ∈ C k+m+2 . Then for any w ∈ C k,α (∂∆, Cn−l )
small and normalized by w(1) = 0 and for any zo close to 0, there is a
unique u ∈ C k,α (∂∆) which solves the equation
(3.7.2) u = −T1 h(u, zo00 + w(τ )) + x0o .
Moreover, if w depends on some parameter η ∈ Rd so that Rd →
C k,α , η 7→ wη is C m , then also (η, x0o , zo00 ) 7→ uη,x0o ,zo00 , Rd × Cn−l →
C k,α (∂∆) is C m .
156 3. REAL/COMPLEX STRUCTURES
version due to Tumanov [95]). The edge of the wedge theorem is due to
Ajrapetyan and Henkin [3] and the Levi extension theorem was estab-
lished by Lewy in codimension 1 and next by Ajrapetyan and Henkin
and Boggess and Polking [29], in general codimension. For the presen-
tation of the material that we propose in this section, we followed the
book by Boggess [27], the book by Baouendi, Ebenfelt and Rothschild
[8] and, more closely, the survey by Tumanov [91].
6=
m
LC p1o
Let dim 1
LC po
= l1 ; in this situation we refer to m1 as the first Hörmander
number of M at po and to l1 as its multiplicity. In case LjC = L1C for
any j, we set m1 = +∞ with multiplicity l1 = l. Next, we look for
m2 > m1 such that
(3.8.3) LjC po = LmC po , j < m2 ,
1
Lm 2 m1
C po 6= LC po ,
Lm2
C po
and set l2 = dim Lm 1 ; again, m2 is possibly +∞. We continue the
C po
above process. We say M is of finite type when commutators span
the whole of C ⊗R TpCo M . Thus, the above chain ends with a number
mr < +∞ or mr = +∞ according to whether the type is finite or not.
There is a “real” construction of chains of spaces of iterated commu-
tators underlying the complex ones. We recall from Section 1.10 the
identifications Re : T 1,0 M → T C M and Re : T 0,1 MP→ T C M . If ∂ωk =
Pk − iYk , k = 1, ..., n − l, these
X P are defined by Re P : k (ak + ibk )∂ωk 7→
k ak Xk + bk Yk and Re : k (ak − ibk )∂ω̄k 7→ k ak Xk + bk Yk respec-
tively. Let us define LjR by means of iterated commutators of the vector
fields Xk , Yk that we denote by a common symbol Lk , i = 1, ..., 2(n−l):
LjR = Span {Lk , [Lk1 , Lk2 ], [Lk1 , [Lk2 , [..., Lkj ]]]}.
| {z }
j
(cf. [18, §4, Proposition 4.2 and Corollary 4.3]). In particular, for these
M , there always exist CR functions f ∈ CRM which do not extend to
any extra direction v such that hv, d(Im F )i > 0. This shows that the
statement in Theorem 3.8.2 is sharp.
(ii) When j = 1, the first condition of (3.8.10) is always fulfilled.
Also, since PI1 is not harmonic, then it is divisible by |w1 |2 and therefore
it has at most 2m1 − 2 zeroes on the unit circle |w1 | = 1. In particular,
for either of ±P , the second condition of (3.8.10) is satisfied.
(iii) There is a sort of “hierarchy” between the Hörmander numbers
mj whose geometric meaning will be fully clear from the proof of The-
orem 3.8.2. According to it, (3.8.10) gives the control not of v o itself,
but of its projection on Rlj +...+lr . The higher j becomes, the weaker the
control of the extension gets.
(iv) When M is of finite type and semirigid (in all its arguments
w), then our proof provides an alternative proof of the extension of
f to a wedge W . The first conclusion in this direction is due to [10]
where a description of W is also given. We improve this description
by specifying the vanishing order in a precise direction w1 . Also, the
semirigidity in the first condition of (3.8.10) can be released, as well as
the hypothesis that the equations are in canonical form as in (3.8.10).
What is indeed essential is the weighted vanishing order.
Proof of Theorem 3.8.2. (a) Preliminaries on F α spaces. Let
0 < α < 1 and denote by τ = reiθ the variable in the standard disc
∆. Let us recall from [93], [18] and [15] some basics about attaching
analytic discs to M in the subspace F α of the Hölder space C α . This
is the space of real continuous functions σ(θ), θ ∈ [−π, π], which are
C 1,α out of 0 and for which the following norm is finite:
||σ||F α := ||σ||C 0 + ||θσ (1) ||C α .
(Here ·(1) denotes the first derivative.) We remark that for σ ∈ F α
we must have θσ (1) |θ=0 = 0 for otherwise θσ (1) → c 6= 0 which implies
|σ| ≥ log |c|
2
−log |θ| which contradicts the boundedness of σ. This shows
that F α is continuously embedded into C α . It is easy to check that F α
is a Banach algebra. Also, if σi ∈ F αi , i = 1, 2, then σ1 · σ2 ∈ F α1 +α2
for α1 + α2 < 1, resp. σ1 · σ2 ∈ C 1,β with β := (α1 + α2 ) − 1 for
α1 + α2 > 1. In both cases the multiplication is continuous with values
in the respective spaces.
Let T1 denote the Hilbert transform normalized by the condition
T1 (·)(1) = 0; it is easy to see that T1 is a bounded operator in F α as has
already been proved to be true in C α . We come back to our manifold M .
We write coordinates in Cn ' Cl ×Cn−l as (z, w) with z = x+iy, choose
166 3. REAL/COMPLEX STRUCTURES
where the last inequality follows from β < α. Since we are assuming
||w̃1 ||F α small, the conclusion follows.
(b) Construction of a singular disc attached to M with con-
trolled normal component. Let us suppose that (3.8.10) is fulfilled.
It is not restrictive to assume that the sector of the plane Cw1 where
g ≥ 0 contains (1 − τ )α iel+1 , τ ∈ ∆. (Here el+1 is the unit vector of
the w1 -plane.) Let α satisfy αmj > 1, α(mj − 1) < 1. We define, for a
small real parameter η > 0,
(3.8.14) w1 (τ ) = (w1 )η (τ ) := η(1 − τ )α iel+1 .
We attach to M̃ a family of F α -discs A(τ ) = Aη (τ ) whose “w1 -
component” is w1 (τ ). We recall from (a) that for i ≥ j we have that
η 7→ (zIi )η (τ ), R → C 1,β is C mi . We also write zIi (τ ) instead of
(zIi )η (τ ), zIi (τ ) = uIi (τ ) + iT1 uIi (τ ), and finally A(τ ) = (z(τ ), w(τ )).
We note that we have
(3.8.15) ∂ηs vIi |η=0 ≡ 0, ∂ηs uIi |η=0 ≡ 0, s ≤ mi − 1.
This is clear for s = 0, 1. If it is true for any s ≤ mi − 2, then it is also
true for s = mi − 1 because of hIi = Omi by a “feedback” procedure.
If we then take the Taylor expansion of ∂r vIi at η = 0, we get
∂ηmi ∂r vIi |η=0 mi
(3.8.16) ∂r vIi = η + o(η mi ).
mi !
By a feedback argument we can also prove that
(3.8.17) |vIi | ≤ c|w1 |mi , |uIi | ≤ c|w1 |mi .
In fact, in the classes F kα , regularity and vanishing order are coincident;
thus, the equation vIi = hIi gives control from below of the vanishing
order of vIi which is transferred as regularity to uIi through Hilbert
transform and again as vanishing order to vIi . In this way, we can
prove that each vIi and uIi belongs to F mi α (and also to C [mi α],{mi α}
where [mi α], resp. {mi α}, is the integer, resp. fractional, part of mi α).
Recall that if ξo is, say, the unit vector in the l0 := (l1 + ... + lj−1 + 1)-
m
direction, we have hξo , hi ≥ 0 if w1 is in a sector S of width > πj and
if |xIi | ≤ c|w1 |mi . We first observe that this latter condition, |xIi | ≤
c|w1 |mi , is automatically fulfilled by the components xIi = uIi of our
discs A(τ ) because of (3.8.17). We show now that ∂r vl0 < 0. In fact, in
our situation we have
(3.8.18) ¯
hξo , ∂ηmj vl0 i|η=0 ≥ 0 for any τ ∈ ∆.
3.8. ITERATED COMMUTATORS AND FINITE TYPE 169
||vl0 − vl0 N ||C 1,γ = ||hl0 (uη , ηw) − hl0 (uN η , SN η )||C 1,γ
m
= η mj ||h(mj ) (0)wmj − h(mj ) (0)SN j ||C 1,γ
m
< η mj ||wmj − SN j ||F α0 .
∼
This proves (3.8.22). Also, once N has been fixed so that (3.8.22) holds,
one chooses η such that ||ηwN ||C 1,γ << 1 so that the hypotheses of
170 3. REAL/COMPLEX STRUCTURES
Proposition 3.8.5 are fulfilled. We call à = (z̃, w̃) one of these discs for
the choices of the parameters N and η which have been pointed out,
and use the notation ṽ o := ∂r (ṽIj ,...,Ir ). With a smooth transversal disc
in our hands, we are in the same situation as for the Levi extension dis-
cussed in Section 3.7 (where the discs were smooth from the beginning
since we had α = 1 therein). By a deformation of the disc Ã, we are
ready to construct a manifold M1 with boundary M which gains one
more direction and has the property that CR functions extend from M
to M1 . For this, we consider Bishop’s equation
In the same way as in Section 3.7 we can see that M1 is a manifold with
boundary M which satisfies T M1 = T M +R+ ṽ o . By the approximation
Theorem 3.6.3 and by the maximum principle, we can extend f from
the boundaries ∂Ap ⊂ M to the “radii” Ip which fill the manifold M1 .
∼
v1 at z1 ? In the identifications (TM Cn )z → Rl through the dual basis to
∂r1 (z), ..., ∂rl (z), what is the l×l matrix B associated to Φ, that is, such
that v1 = Bvo ? In practice, there is no natural connection in TM Cn .
∗
But there is in fact in TM Cn and it turns out that its dual-inverse is the
n
connection in TM C we are looking for (cf. Section 3.11). Let us start
the construction of t Φ−1 and t B −1 . Let z and (z; ζ) be coordinates in Cn
and T ∗ Cn , respectively. Let M be described in a neighborhood of 0 as
a graph yj0 = hj (x0 , z 00 ) with hj (0) = 0 and ∂hj (0) = 0 for j = 1, ..., l.
Let prCn−l be the projection in the fibers of T ∗ Cn modulo Cn−l , given
by Cn → Cl , ζ → ζ 0 , and define
∗
(3.9.1) (TM Cn )0 := prCn−l (TM
∗
Cn ).
∗
We denote by T C M := T M ∩ iT M and T C (TM Cn )0 := T (TM∗
Cn )0 ∩
∗ n 0 ∗ n 0
iT (TM C ) the complex tangent bundles to M and (TM C ) , respec-
tively.
Theorem 3.9.1. The natural projection π : T ∗ Cn → Cn induces
an identification
∗ ∼
(3.9.2) T C (TM Cn )0 →0 T C M.
π
It follows that
(3.9.4)
∗
TpC0 (TM Cn )0 ={(u; t∂ 0 r + ∂u ∂ 0 ro + ∂¯u ∂ 0 ro ) : u ∈ T C M,
1
∂¯u ∂ 0 ro = (−t + is)∂ 0 r for some t and s in Rl }.
2
Now, since ∂ 0 r is non-degenerate, for any u ∈ T C M it is possible to find
a unique λ = 21 (−t + is) ∈ Cl which solves the equation ∂¯u ∂ 0 ro = λ∂ 0 r;
∗
hence, over any vector u ∈ T C M , there is a vector of T C (TM Cn )0 . Also,
as already mentioned, this vector is unique.
∗ ∗
For a CR curve γ in M , we call a CR curve in TM Cn , resp. (TM Cn )0 ,
which projects over γ a “lift”, resp. a “partial lift”, of γ. We denote
by γ ∗ and γ ∗0 a lift and a partial lift of γ, respectively.
3.9. PARTIAL LIFTS OF ANALYTIC DISCS AND CR CURVES 173
∗
Let z1 = A(1) and take ξ1 ∈ (TM Cn )z1 . Then, for any A attached to
M with A(1) = z1 , there is a unique partial lift A∗ 0 with A∗ 0 (1) = ξ10 ;
in fact, the disc
A∗ 0 = A; t ∂ 0 r(A)t Gξ10
(3.9.7)
clearly serves the purpose. Also, if ζ 0 (τ ) = t ∂ 0 rξ 0 (τ ) is another lift, then
if we define g(τ ) := (t ∂ 0 r(A)t G)−1 ζ(τ ), we have
(
g(τ ) ∈ R for any τ ∈ ∂∆,
g(τ ) extends holomorphically.
It follows that g(τ ) ≡ ξ10 (constant) and therefore ζ 0 = t ∂ 0 rt Gξ10 ; thus,
ζ 0 (τ ) is uniquely determined by ζ 0 (1) = ξ10 . Also, through the partial
lifts of A, we can define a connection Φ over ∂A similar to that obtained
over γ. Let zo = A(−1) and z1 = A(1), take ξo0 ∈ (TM ∗
Cn )0zo , and let
A∗0 ∗0 0
(zo ,ξo0 ) be the unique partial lift such that A(zo ,ξo ) (−1) = ξo . We define
∗
a morphism (TM ∗
Cn )0zo → (TM Cn )0z1 by
(3.9.8) t
Φ−1 : ξo0 7→ ξ10 where ξ10 = pr−1 (A∗0
(zo ,ξo ) (1)).
Since A∗0 t 0 t 0
(zo ,ξo ) is in the form ∂ r(A) Gξ1 , then if we want it to pass
through ξo0 at −1, we must have ξ10 = t G−1 (−1)ξo0 . In other words,
we have t B −1 = t G−1 (−1). Again, the definition contains a great deal
of geometric foundation. We have chosen a distinguished section A∗0
∗
of (TM Cn )0 over ∂A: the one uniquely endowed with holomorphic ex-
tension to Ā. Since CR curves γ ∗0 can be approximated by chains of
analytic discs A∗0 , then the matrix B associated to Φ is approximable
by a composition of matrices G(−1)’s.
Guided Exercise 3.9.7 (Twin of Exercise 3.9.3). Let χ be the
transformation of Exercise 3.9.3 which interchanges M 99K M̃ . Prove
that it interchanges attached analytic discs A 99K Ã.
Hint: take a partial lift A∗0 (τ ) = (A, ζ 0 (τ )) for ζ 0 (τ ) = t ∂z0 r(A)· t Gξo ,
and define Ã∗0 = χ(A∗0 ) and à = π(Ã∗0 ). Now, in contrast to full
lifts, we have that the partial lifts A∗0 and Ã∗0 are holomorphic. Thus
à = π(Ã∗0 ) is holomorphic. Also, ∂ à ⊂ M̃ since ∂ Ã∗0 ⊂ (TM ∗
Cn )0 . The
conclusion is illustrated by the diagram
χ
A∗0 → Ã∗0
(3.9.9) π↓ ↓π
A 99K Ã.
We end this section with a geometric application of the construction
of lifts. This part is not essential for what follows.
176 3. REAL/COMPLEX STRUCTURES
Φ : Ezo → Ez1 ,
the submean property, we have for any τ in ∆δ (τo ), the disc of center
τo and radius δ,
Z 2π Z r
2 −1
ϕν (τ ) ≤ (πr ) tϕν (τ + teiθ )dt dθ
0 0
(3.12.5) Z 2π Z r+δ
≤ (π(r + δ)2 )−1 tϕν (τo + teiθ )dt dθ + O(δ),
0 0
in x00 when the other variables are fixed. CR extendibility to all the di-
rections of the y 00 -plane Rn2 means holomorphic extendibility to |y 00 | <
with locally uniform in x. CR extendibility to a cone Γ2 ⊂ Rn2 means
that f extends holomorphically for y 00 ∈ Γ2 , |y 00 | < with locally
uniform in x; we also assume that, for fixed x0 , f is locally uniformly
continuous up to y 00 = 0. Clearly real analyticity or CR extendibility
in a group of variables x00 is more restrictive than the combination of
real analyticity or CR extendibility in any single x00j . We now state the
main result of the section; the proof will follow next.
¯
[1] H. Ahn—Global boundary regularity of the ∂-equation on q-pseudoconvex
domains, Math. Reich. 280 (2007), 343–350
[2] H. Ahn, L. Baracco and G. Zampieri—Subelliptic estimates and reg-
ularity of ∂¯ at the boundary of a Q-pseudoconvex domain of finite type,
Math. Reich. (2009)
[3] R. Ajrapetyan and G. Henkin—Analytic continuation of CR functions
through the “edge of the wedge”, Dokl. Acad. Nauk. SSSR 259 (1981),
777–781
[4] R. Ajrapetyan and G. Henkin—Integral representation of differential
forms on Cauchy-Riemann manifolds and the theory of CR functions, II,
Math. USSR-Sb 55 (1986), 91–111
[5] A. Andreotti, G. Fredricks and M. Nacinovich—On the absence
of Poincaré lemma in tangential Cauchy-Riemann complexes, Ann. S.N.S.
Pisa 27 (1981), 365–404
[6] A. Andreotti and C.D. Hill—Complex characteristic coordinates and
tangential Cauchy-Riemann equations, Ann. S.N.S. Pisa, 26 (1972), 299–
324
[7] S. Baouendi, C.H. Chang and and F. Treves—Microlocal analyticity
and extension of CR functions, J. Diff. Geom. 18(1983), 331–391
[8] M.S. Baouendi, P. Ebenfelt and L.P. Rothschild—Real Subman-
ifolds in Complex Space and Their Mappings, Princeton Math. Series,
Princeton Univ. Press (1999)
[9] M.S. Baouendi, H. Jacobowitz and F. Treves—On the analyticity of
CR mappings, Ann. Math. 122 (1985), 365–400
[10] M.S. Baouendi and L.P. Rothschild—Normal forms for generic mani-
folds and holomorphic extension of CR functions, J. Diff. Geom. 25 (1987),
431–467
[11] M.S. Baouendi and L.P. Rothschild—Cauchy Riemann functions on
manifolds of higher codimension, Invent. Math. 101 (1990), 45–56
[12] M.S. Baouendi, L.P. Rothschild and J.M. Trépreau—On the geom-
etry of analytic discs attached to real manifolds, J. Diff. Geometry (1994),
379–405
[13] M.S. Baouendi and F. Treves—A property of the functions and distri-
butions annihilated by a locally integrable system of complex vector fields.
Ann. Math. (2) 114 (1981), 387–421
[14] M.S. Baouendi and F. Treves—About holomorphic extension of CR
functions on real hypersurfaces in complex space, Duke Math. J. 51 (1984),
77–107
193
194 BIBLIOGRAPHY