20.07.2013 Views

NEAR OPTIMAL BOUNDS IN FREIMAN'S THEOREM

NEAR OPTIMAL BOUNDS IN FREIMAN'S THEOREM

NEAR OPTIMAL BOUNDS IN FREIMAN'S THEOREM

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S<br />

<strong>THEOREM</strong><br />

TOMASZ SCHOEN<br />

Abstract<br />

We prove that if for a finite set A of integers we have |A + A| ≤K|A|, thenAis 1+C(log K)−1/2<br />

contained in a generalized arithmetic progression of dimension at most K<br />

1+C(log K)−1/2<br />

and of size at most exp(K )|A| for some absolute constant C. We also<br />

discuss a number of applications of this result.<br />

1. Introduction<br />

Freiman’s theorem [8] is one of the most fundamental theorems in additive number<br />

theory. It asserts that if for a finite A ⊆ Z we have |A + A| ≤K|A|, thenA is<br />

a subset of a d-dimensional generalized arithmetic progression P1 +···+Pd with<br />

|P1|···|Pd| =|A|f , where d and f depend only on K (here Pi stands for usual<br />

arithmetic progressions). This result provides a complete description of sets for which<br />

the sumset is not much larger than the set itself. Freiman’s theorem has a plethora<br />

of deep applications to many important problems, such as the one recently described<br />

by Gowers in [9] and[11], who used it to get an effective version of Szemerédi’s<br />

theorem (for further works that make important use of Freiman’s result, see, e.g.,<br />

[6], [5], [25], [26]–[28]). However, most applications require a quantitative version of<br />

Freiman’s theorem. It is easy to see that one cannot do better than d(K) ≤ K − 1<br />

and f (K) = e O(K) . The estimates which followed from Freiman’s original proof<br />

were quite poor. The first useful estimates for d and f were provided by Ruzsa<br />

[19], whose ingenious argument combines techniques from different fields. Ruzsa’s<br />

bounds were good enough for many applications and his approach paved the way<br />

for further improvements. The next substantial progress was made by Chang [3],<br />

who showed that one can take d ≤ K 2+o(1) and f (K) ≤ exp(K 2+o(1) ). Her proof<br />

coupled Ruzsa’s main ideas with Chang’s so-called spectral lemma and covering<br />

lemma. Further improvements were due to Sanders [21, Theorem 1.3], who used a<br />

much-expanded version of Bogolyubov-Ruzsa’s lemma [20, Lemma 2.1] to obtain the<br />

best currently known bounds d(K) ≤ K 7/4+o(1) and f (K) ≤ exp(K 7/4+o(1) ). The main<br />

DUKE MATHEMATICAL JOURNAL<br />

Vol. 158, No. 1, c○ 2011 DOI 10.1215/00127094-1276283<br />

Received 16 February 2010. Revision received 13 July 2010.<br />

2010 Mathematics Subject Classification. Primary 11P70; Secondary 11B25.<br />

Author’s work supported in part by Ministerstwo Nauki i Szkolnictwa Wy˙zszego grant N N201 543538.<br />

1


2 TOMASZ SCHOEN<br />

result of this article is the following theorem, which gives nearly optimal estimates<br />

for d and f .<br />

<strong>THEOREM</strong> 1<br />

There exists an absolute constant C such that each set A ⊆ Z satisfying |A +<br />

A| ≤K|A| is contained in a generalized arithmetic progression of dimension at<br />

1+C(log K)−1/2<br />

most d(K) and size at most f (K)|A|, with d(K) ≤ K and f (K) ≤<br />

1+C(log K)−1/2<br />

exp(K ).<br />

Our approach is yet another follow-up of Ruzsa’s original argument, and our improvement<br />

relies on further refinement of Bogolyubov-Ruzsa’s lemma. A new addition<br />

to this setup is an elementary result, Lemma 3, which roughly says that one can<br />

find two sets “closely related” to A which have much better additive properties. As<br />

a consequence of our method, we infer that there is a proper arithmetic progression<br />

P of dimension Oε(Kε ) and size bounded from above by KO(1) |A| such that<br />

|A ∩ P |≥exp(−Oε(Kε ))|A| (for a precise statement, see Theorem 8 in Section 4).<br />

This fact has been already anticipated by Chang [3], Gowers [10], and Green and<br />

Tao [14, Conjecture 1.6], and is often required in applications. Remarkably, Green<br />

and Tao [14] proved that this result is equivalent to an inverse theorem for Gowers’s<br />

U 3-norm. It can also be considered a first step toward resolving Freiman-Ruzsa’s<br />

polynomial conjecture. We also note that, using Freiman-Bilu’s [1] argument (see also<br />

[3]), one cannot deduce from Theorem 1 that A is contained in a progression with<br />

1+C(log K)−1/2<br />

linear dimension and size at most exp(K )|A|. The reason is that we cannot<br />

guarantee our progression to be proper. Our proof can be modified to achieve it as<br />

well, but in this case the size of the progression goes up to exp(K2+o(1) )|A|.<br />

Finally, our method can be directly applied to Green-Ruzsa’s [13, Theorem 1.1]<br />

proof of Freiman’s theorem in the “torsion” case, giving the following statement, the<br />

proof of which we omit here (further improvement of this theorem will appear in [7]).<br />

<strong>THEOREM</strong> 2<br />

Let G be an arbitrary Abelian group, and let A ⊆ G satisfy |A + A| ≤K|A|. Then A<br />

is contained in a coset progression P +H of dimension d(K) ≤ (K+2) 3+C(log(K+2))−1/2<br />

and size f (K)|A| ≤exp((K + 2) 3+C(log(K+2))−1/2<br />

)|A| for some absolute constant C.<br />

This paper is organized as follows. Section 2 contains some basic definitions and facts<br />

that we use here. The main part of the paper is Section 3, where we prove Lemma 3,<br />

which is key to our argument. In order to keep the paper self-contained, we briefly<br />

outline the rest of Ruzsa’s argument (with Chang’s refinements) in Section 4. We<br />

conclude with some applications of our results and a few comments on our approach<br />

and its possible further developments.


<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S <strong>THEOREM</strong> 3<br />

2. Preliminaries<br />

This part of the article contains basic definitions and notation we will use later on.<br />

As usual, we set<br />

A + B ={a + b : a ∈ A, b ∈ B},<br />

and the k-fold sumset of A is denoted by kA.Byageneralized arithmetic progression<br />

of dimension d, we mean every set of the form P = P1+···+Pd, where P1,...,Pd are<br />

usual arithmetic progressions. The size of P is defined as the product |P1|···|Pd|.The<br />

dimension and size of progression P are denoted by dim(P ) and size(P ), respectively.<br />

If each x ∈ P has unique representation x = p1 +···+pd, pi ∈ Pi, thenwesay<br />

that P is proper. Then the cardinality of P is equal to its size.<br />

Let G, H be Abelian groups, and let A ⊆ G, B ⊆ H . We say that A is Fkisomorphic<br />

(i.e., Freiman isomorphic of order k) toB if there exists a bijective map<br />

ϕ : A → B such that<br />

if and only if<br />

x1 +···+xk = y1 +···+yk<br />

ϕ(x1) +···+ϕ(xk) = ϕ(y1) +···+ϕ(yk)<br />

for every x1,...,xk,y1,...,yk ∈ A.<br />

We call (x,x ′ ,y,y ′ ) an additive quadruple if x + y = x ′ + y ′ . The number of<br />

additive quadruples in X2 × Y 2 is denoted be E(X, Y ).<br />

In this paper, by Zn we always mean Z/nZ. The Fourier coefficients of the<br />

indicator function of a set X ⊆ Zn are defined by<br />

ˆX(s) = <br />

e −2πixs/n ,<br />

x∈X<br />

where s ∈ Zn. Parseval’s formula states that n−1<br />

s=0 | ˆX(s)| 2 =|X|n. We observe also<br />

that, for X, Y ⊆ Zn, wehave<br />

E(X, Y ) = 1 n−1<br />

| ˆX(s)|<br />

n<br />

2 | ˆY (s)| 2 . (1)<br />

s=0<br />

Our argument makes use of Plünnecke-Ruzsa’s inequality, which says, in particular<br />

that if |A + A| ≤K|A|, then for all k, l ∈ N, wehave<br />

|kA − lA| ≤K k+l |A|.


4 TOMASZ SCHOEN<br />

By α, we denote the distance from α to the nearest integer. For Ɣ ⊆ Zn and<br />

γ>0, the Bohr set B(Ɣ, γ ) is defined as<br />

B(Ɣ, γ ) ={x ∈ Zn : gx/n ≤γ for every g ∈ Ɣ}.<br />

3. The main lemma<br />

The key idea of our approach is to identify a relatively large subset of A that has<br />

better additive properties than A. One possible way would be to find A ′ ⊆ A such that<br />

|kA ′ |≤K εk |A ′ | for not too-big k. A result of this kind was proved in [17, Lemma<br />

1], but under a stronger assumption on A. Unfortunately, this method does not seem<br />

to work in our general setting. Let us remark, however, that Bogolyubov-Ruzsa’s<br />

lemma uses the number of additive quadruples rather than the size of the subset.<br />

This suggests the following approach: using the assumption |A + A| ≤K|A|, one<br />

should find for some small k a relatively dense set X, Y ⊆ kA − kA with E(X, Y )<br />

much bigger than one can expect from the size of A + A. As far as we know, the<br />

first such result was proved by Katz and Koester [15]. They showed that there is<br />

B ⊆ A ± A, |B| ≫|A|/K C with E(B,B) ≫|B| 3 /K 36/37 . Lemma 3 provides a<br />

much better estimate for E(X, Y ) in a “highly nonsymmetric” case, when one set is<br />

large and the other is small (a similar result for a diagonal case would give a much<br />

better result—see our concluding remarks below). Finally, let us also mention that<br />

Sanders in [22], and [23] (seealso[24] for related theorems), used a similar result<br />

for a different purpose—that of improving the bounds in many summed versions of<br />

Roth-Meshulam’s theorem.<br />

LEMMA 3<br />

Suppose that A is a subset of an Abelian group and that |A+A| ≤K|A|. Then for every<br />

ε>0, there are sets X ⊆ A and Y ⊆ A + A such that |X| ≥(2K2 ) −21/ε|A|,<br />

|Y |≥<br />

|A|, and E(X, Y ) ≥ K−2ε |X| 2 |Y |.<br />

Proof<br />

For a set B ⊆ A, denote by D = D(B) the set of all t ∈ B − B which has at least<br />

|B| 2 |B|2 |B|2<br />

≥ ≥<br />

2|B − B| 2|A − A| 2K2 |A|<br />

representations as b − b ′ , b,b ′ ∈ B (the last inequality follows from Plünnecke-<br />

Ruzsa’s inequality). Note that |D| ≥|B|/2.<br />

We first show that there exists a set B ⊆ A such that |B| ≥(2K 2 ) −21/ε +1 |A| and<br />

|A + Bt| ≥K −ε |A + B| (2)<br />

for all t ∈ D = D(B), where Bt = B ∩ (B + t). We construct B in the following<br />

iterative procedure. We start with A 1 = A. Now suppose that A l with


<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S <strong>THEOREM</strong> 5<br />

|Al |≥(2K2 ) −2l−1 +1 |A| has already been defined. If there exists t0 ∈ D(Al ) with<br />

|A + Al t0 |


6 TOMASZ SCHOEN<br />

Since |D ′ |≥|B|/2 ≥ (2K2 ) −21/ε|A|,<br />

the assertion follows for X = D ′ and Y =<br />

A + B. ✷<br />

Proof of Theorem 1<br />

In order to show Theorem 1, one needs to mimic Ruzsa-Chang’s proof using Lemma 3.<br />

In order to make the article self-contained, we briefly sketch their argument.<br />

Let n be a prime satisfying<br />

48|12A − 12A| >n>24|12A − 12A|.<br />

By Plünnecke-Ruzsa’s inequality, it follows that n ≤ 48K 24 |A|. By Ruzsa’s lemma<br />

(see [18, Theorem 2]), there exists a set A ′ ⊆ A such that |A ′ |≥|A|/12, whichis<br />

F12-isomorphic to T ⊆ Zn. Clearly, |T + T |≤12K|T |. Let X ⊆ T and Y ⊆ T + T<br />

be sets such that the assertion of Lemma 3 holds with<br />

Thus, we have<br />

ε = (log K) −1/2 .<br />

|X| ≥ 2(12K) 2 −2 1/ε<br />

|T |≥2 −10 (288K 2 ) −21/ε<br />

K −24 n.<br />

The next ingredient of our approach is Chang’s spectral lemma (see [3]and[13]).<br />

For Ɣ ⊆ Zn, define<br />

<br />

<br />

Span(Ɣ) = εgg : εg ∈{−1, 0, 1} .<br />

g∈Ɣ<br />

Then the spectral lemma can be stated as follows.<br />

LEMMA 4([3, Lemma 3.1])<br />

Let X ⊆ Zn, and let ={r ∈ Zn : | ˆX(r)| ≥λ|X|}. Then there is a set Ɣ ⊆ Zn such<br />

that |Ɣ| ≤2λ −2 log(n/|X|) and ⊆ Span(Ɣ).<br />

Let us emphasize at this point that Chang’s spectral lemma is absolutely crucial for<br />

our argument. While in Chang’s paper one can omit it and still get reasonable bounds<br />

d(K) ≤ K O(1) and f (K) ≤ exp(K o(1) ), without Lemma 4 our approach completely<br />

fails. In fact, as it is easy to see, Chang’s spectral lemma is especially useful if λ is<br />

much bigger than the density of X, and we apply it precisely in such a case using its<br />

full potential.<br />

Our goal is to get a version of Bogolyubov-Ruzsa’s lemma for sets X and Y .


<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S <strong>THEOREM</strong> 7<br />

LEMMA 5<br />

The set X + Y − X − Y contains a Bohr set B(Ɣ, γ ) such that |Ɣ| ≪K 3ε log K and<br />

γ ≫ K −3ε log −1 K.<br />

Proof<br />

Define<br />

= r ∈ Zn : | ˆX(r)| ≥K −ε |X|/2 .<br />

We will show that B = B(, 1/6) ⊆ X + Y − X − Y .Letr(x) be the number of<br />

representations of x in X + Y − X − Y .Ifx ∈ B, then, by (1) and Parseval’s formula,<br />

we have<br />

r(x) = 1 n−1<br />

| ˆX(s)|<br />

n<br />

2 | ˆY (s)| 2 e 2πixs/n<br />

≥ 1<br />

n<br />

≥ 1<br />

2n<br />

s=0<br />

<br />

s∈<br />

<br />

s∈<br />

s=0<br />

| ˆX(s)| 2 | ˆY (s)| 2 cos(2πxs/n) − 1<br />

n<br />

| ˆX(s)| 2 | ˆY (s)| 2 − 1<br />

n<br />

<br />

| ˆX(s)| 2 | ˆY (s)| 2<br />

s∈<br />

<br />

| ˆX(s)| 2 | ˆY (s)| 2<br />

s∈<br />

≥ 1 n−1<br />

| ˆX(s)|<br />

2n<br />

2 | ˆY (s)| 2 − 3 <br />

| ˆX(s)|<br />

2n<br />

2 | ˆY (s)| 2<br />

s∈<br />

> E(X, Y )/2 − 3<br />

8n K−2ε |X| 2<br />

n−1<br />

| ˆY (s)| 2<br />

s=0<br />

≥ E(X, Y )/2 − (3/8)K −2ε |X| 2 |Y | > 0.<br />

By Chang’s spectral lemma (Lemma 4), there is a set Ɣ such that |Ɣ| ≪<br />

K 2ε log(n/|X|) ≪ K 3ε log K and ⊆ Span(Ɣ), so that<br />

B Ɣ, 1/(6|Ɣ|) ⊆ B(, 1/6) ⊆ X + Y − X − Y. ✷<br />

Remark. We proved Bogolyubov-Ruzsa’s lemma in a standard way, but one can easily<br />

strengthen the assertion by removing one summand. Indeed, by (3), we have<br />

1 n−1<br />

ˆX(s)| ˆY (s)|<br />

n<br />

s=0<br />

2 e −2πib0s/n −ε<br />

≥ K |X||Y |.<br />

Therefore, using a similar argument as in the proof of Lemma 5, we infer that a shift<br />

of B(Ɣ, γ ) is contained in X + Y − Y.


8 TOMASZ SCHOEN<br />

Let us first recall Ruzsa’s lemma on the geometry of numbers (see [20, Lemma<br />

3.2]).<br />

LEMMA 6<br />

Let B(Ɣ, γ ) be a Bohr set in Zn. Then there is a proper progression P ⊆ B(Ɣ, γ ) of<br />

dimension |Ɣ| and size at least (γ/|Ɣ|) |Ɣ| n.<br />

Thus by Lemmas 5 and 6, one can find a proper progression P ⊆ X + Y − X − Y ⊆<br />

3T − 3T with dim(P ) ≤|Ɣ| ≪K 3ε log K and<br />

size(P ) ≥ exp −K 3ε (log K) 3/2+o(1) n.<br />

Let P ′ betheimageofP under a considered F12-isomorphism between A ′ and T .It<br />

induces an F2-isomorphism between P and P ′ , which implies that P ′ is also proper.<br />

For the last step of the argument we need Chang’s covering lemma [3,Step4].<br />

LEMMA 7<br />

If |A + A| ≤K|A| and |B + A| ≤L|B|, then there are sets S1,...,Sk each of size<br />

≤ 2K such that A ⊆ B − B + (S1 − S1) +···+(Sk − Sk) and k ≤ 1 + log(KL).<br />

Now, we have<br />

so that<br />

|P ′ + A| ≤|5A − 4A| ≤K 9 |A| ≤K 9 exp K 3ε (log K) 3/2+o(1) |P ′ |<br />

A ⊆ P ′ − P ′ + (S1 − S1) +···+(Sk − Sk)<br />

with k ≤ K 3ε (log K) 5/2+o(1) . Finally, A is contained in a progression Q with<br />

and<br />

1+C(log K)−1/2<br />

dim(Q) ≤ dim(P ) + 2kK ≤ K<br />

size(Q) ≤ 2 dim(P ′ 1+C(log K)−1/2<br />

) ′ 2kK K<br />

size(P )3 ≤ e |A|<br />

for some constant C. This completes the proof of Theorem 1. <br />

4. Some applications<br />

We focus on the consequences of a new version of Bogolyubov-Ruzsa’s lemma in<br />

Freiman’s theorem as one of the most important applications. However, as an immediate<br />

consequence of the results obtained in Section 3, we obtain the following step in<br />

the direction of Freiman-Ruzsa’s polynomial conjecture, which may also have other<br />

applications.


<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S <strong>THEOREM</strong> 9<br />

<strong>THEOREM</strong> 8<br />

There exists an absolute constant C such that, if A ⊆ Z satisfies |A+A| ≤K|A|, then<br />

C(log K)−1/2<br />

there is a proper generalized arithmetic progression P with dim(P ) ≤ K<br />

and size(P ) ≤ 48K24 C(log K)−1/2<br />

|A| such that |A ∩ P |≥exp(K )|A|.<br />

Using Theorem 8 and Theorem 1, one can directly improve many results based on<br />

Freiman’s theorem; we mention only some of them here. The most exciting case is<br />

Gowers’s theorem; however, here the effect is not so impressive. It is proved in [11,<br />

Theorem 18.6] that, for every k ∈ N, the density of a subset of {1,...,n} which does<br />

not contain any k-term arithmetic progression does not exceed<br />

(log log n) −ck ,<br />

where ck = 2−2k+9. Applying our version of Bogolyubov-Ruzsa’s lemma, one can at<br />

most increase the value of ck. On the other hand, one can deduce from Lemma 5 (and<br />

in view of the Remark) that, for each set A ⊆{1,...,n} of density δ =|A|/n, one<br />

can find an arithmetic progression of length<br />

1/δ)1/2<br />

e−c(log<br />

cδn<br />

in 5A for some absolute constant c>0, beating the previous best bound cδn cδ1/3<br />

obtained by Sanders [20]. Another interesting consequence is a result related to a<br />

problem considered by Konyagin and Łaba [16, Corollary 3.7]. Inserting our estimates<br />

in Sanders’s proof (see [21, Theorem 1.2]), we infer that there exists c>0 such that,<br />

for every finite set A ⊆ R and every transcendental α ∈ R, wehave<br />

|A + α · A| ≥(log |A|) c log log |A| |A|.<br />

Let us also formulate our version of Bogolyubov-Ruzsa’s lemma for F n 2<br />

follows.<br />

<strong>THEOREM</strong> 9<br />

There are absolute constants C1,C2 such that, if A ⊆ Fn 2<br />

to read as<br />

has density δ, then6A<br />

√<br />

contains a subspace H of codimension C1 exp(C2 log(1/δ)).<br />

We refer the reader to [12] for other consequences of Lemma 3 in F n 2 . The last<br />

application provides progress toward Ruzsa’s conjecture, which states that, for every<br />

finite set of squares A, wehave|A + A| ≫|A| 2−ε . Currently, the best estimate<br />

|A + A| ≥(log |A|) 1/12 |A| is due to Chang [4].


10 TOMASZ SCHOEN<br />

<strong>THEOREM</strong> 10<br />

There is a positive constant c such that, for every finite set A ⊆{1 2 , 2 2 ,...}, we have<br />

|A + A| ≥(log |A|) c log log |A| |A|.<br />

Proof<br />

Set |A + A| =K|A|. By Theorem 8, there is a proper generalized arithmetic pro-<br />

C(log K)−1/2<br />

gression P = P1 +···+Pd with d ≤ K and size(P ) ≤ 48K24 |A| such<br />

that<br />

K)−1/2<br />

−KC(log<br />

|A ∩ P |≥e |A|.<br />

Suppose that |P1| =maxi |Pi| ≥|P | 1/d , and write P = <br />

x∈P2+···+Pd (P1 + x). Now,<br />

we make use of a theorem of Bombieri and Zannier [2, Theorem 1] stating that an<br />

arithmetic progression of length k contains O(k2/3+ε ) squares. Therefore, we have<br />

|A ∩ P |≤ <br />

|(P1 + x) ∩ A| ≪ <br />

|P1| 3/4 ≤ (48K 24 |A|) 1−1/(4d) ,<br />

hence<br />

x∈P2+···+Pd<br />

x∈P2+···+Pd<br />

K ≥ (log |A|) c log log |A| ,<br />

and the assertion follows. ✷<br />

It would be interesting to strengthen Lemma 3 since, clearly, each such improvement<br />

would result in better bounds in Freiman’s theorem. The argument presented here has<br />

one major weakness—it works only for small sets X (one would prefer to have an<br />

analog of this result which is also valid for sets X of roughly the same order as Y ).<br />

Indeed, if one could prove a version of Lemma 3 with |X| ∼|Y | and, say ε = 1/20,<br />

then Balog-Szemerédi-Gowers’s theorem would give |X ′ +X ′ |≤K 1−c |X ′ |, for some<br />

X ′ ⊆ X and c>0. Then, by iterating this process, one would most probably get a<br />

result very close to Freiman-Ruzsa’s polynomial conjecture.<br />

References<br />

[1] Y. BILU, “Structure of sets with small sumset” in Structure Theory of Set Addition,<br />

Astérisque 258, Soc. Math. France, Paris, 1999, 77 – 108. MR 1701189<br />

[2] E. BOMBIERI and U. ZANNIER, A note on squares in arithmetic progressions, II, Atti<br />

Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl. 13<br />

(2002), 69 – 75. MR 1949479


<strong>NEAR</strong> <strong>OPTIMAL</strong> <strong>BOUNDS</strong> <strong>IN</strong> FREIMAN’S <strong>THEOREM</strong> 11<br />

[3] M.-C. CHANG, A polynomial bound in Freiman’s theorem, Duke Math. J. 113 (2002),<br />

399 – 419. MR 1909605<br />

[4] ———, On problems of Erdős and Rudin, J. Funct. Anal. 207 (2004), 444 – 460.<br />

MR 2032997<br />

[5] ———, On sum-product representations in Zq, J. Eur. Math. Soc. (JEM) 8 (2006),<br />

435 – 463. MR 2250167<br />

[6] ———, Some consequences of the polynomial Freiman-Ruzsa conjecture, C. R. Math.<br />

Acad. Sci. Paris 347 (2009), 583 – 588. MR 2532910 2101, 2, 6, 8911<br />

[7] K. CWAL<strong>IN</strong>A and T. SCHOEN, Linear bound on dimension in Green-Ruzsa’s theorem,<br />

in preparation.<br />

[8] G. A. FREĬMAN, Foundations of a Structural Theory of Set Addition, Trans. Math.<br />

Monog. 37, Amer. Math. Soc., Providence, 1973. MR 0360496<br />

[9] W. T. GOWERS, A new proof of Szemerédi’s theorem for arithmetic progressions of<br />

length four, Geom. Funct. Anal. 8 (1998), 529 – 551. MR 1631259<br />

[10] ———, “Rough structure and classification” in GAFA 2000 (Tel Aviv, 1999), Geom.<br />

Funct. Anal. 2000, 79 – 117. MR 1826250<br />

[11] ———, A new proof of Szemerédi’s theorem, Geom. Funct. Anal. 11 (2001),<br />

465 – 588. MR 1844079<br />

[12] B. GREEN, “Finite field models in additive combinatorics” in Surveys in Combinatorics<br />

2005, London Math. Soc. Lecture Note Ser. 327, Cambridge Univ. Press,<br />

Cambridge, 2005, 1 – 27. MR 2187732<br />

[13] B. GREEN and I. Z. RUZSA, Freiman’s theorem in an arbitrary abelian group, J. Lond.<br />

Math. Soc. (2) 75 (2007), 163 – 175. MR 2302736<br />

[14] B. GREEN and T. TAO, An equivalence between inverse sumset theorems and inverse<br />

conjectures for the U 3 norm, Math. Proc. Cambridge. Philos. Soc. 149 (2010),<br />

1 – 19. MR 2651575<br />

[15] N. H. KATZ and P. KOESTER, On additive doubling and energy, preprint,<br />

arXiv:0802.4371v1 [math.CO]<br />

[16] S. KONYAG<strong>IN</strong> and I. ŁABA, Distance sets of well-distributed planar sets for polygonal<br />

norms, Israel J. Math. 152 (2006), 157 – 175. MR 2214458<br />

[17] T. ŁUCZAK and T. SCHOEN, On a problem of Konyagin, Acta Arith. 134 (2008),<br />

101 – 109. MR 2429639<br />

[18] I. Z. RUZSA, Arithmetical progressions and the number of sums, Period. Math. Hungar.<br />

25 (1992), 105 – 111. MR 1200845<br />

[19] ———, Generalized arithmetical progressions and sumsets, Acta Math. Hungar. 65<br />

(1994), 379 – 388. MR 1281447<br />

[20] T. SANDERS, Additive structures in sumsets, Math. Proc. Cambridge. Philos. Soc. 144<br />

(2008), 289 – 316. MR 2405891<br />

[21] ———, Appendix to Roth’s theorem on progressions revisited by J. Bourgain, J. Anal.<br />

Math. 104 (2008), 193 – 206. MR 2403434<br />

[22] ———, On a nonabelian Balog-Szemerédi-type lemma, J. Aust. Math. Soc. 89 (2010),<br />

127 – 132. MR 2727067<br />

[23] ———, Structure in sets with logarithmic doubling, to appear in Canad. Math. Bull.,<br />

preprint, arXiv:1002.1552v1 [math.CA]


12 TOMASZ SCHOEN<br />

[24] I. SHKREDOV and S. YEKHAN<strong>IN</strong>, Sets with large additive energy and symmetric sets,<br />

J. Combin. Theory Ser. A 118 (2011), 1086 – 1093.<br />

[25] B. SUDAKOV, E. SZEMERÉDI,andV. H. VU, On a question of Erdős and Moser, Duke<br />

Math. J. 129 (2005), 129 – 155. MR 2155059<br />

[26] E. SZEMERÉDI and V. H. VU, Long arithmetic progressions in sum-sets and the number<br />

of x-sum-free sets, Proc. Lond. Math. Soc. (3) 90 (2005), 273 – 296. MR 2142128<br />

[27] ———, Finite and infinite arithmetic progressions in sumsets, Ann. of Math. (2) 163<br />

(2006), 1 – 35. MR 2195131<br />

[28] ———, Long arithmetic progressions in sumsets: Thresholds and bounds, J. Amer.<br />

Math. Soc. 19 (2006), 119 – 169. MR 2169044 21121, 992, 62494611,<br />

8, 91, 944411 1<br />

Faculty of Mathematics and Computer Science, Adam Mickiewicz University, 61-614 Poznań,<br />

Poland; schoen@amu.edu.pl


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS<br />

PASCAL AUTISSIER<br />

Résumé<br />

On donne des résultats de non-densité pour les points entiers sur des variétés affines,<br />

dans l’esprit de la conjecture de Lang-Vojta. En particulier, soit X une variété projective<br />

de dimension d ≥ 2 sur un corps de nombres K (resp., sur C). Soit H la somme de<br />

2d diviseurs amples sur X qui se coupent proprement. On montre que tout ensemble<br />

de points quasi-entiers (resp., toute courbe entière) sur X − H est non Zariski-dense.<br />

Abstract<br />

We give nondensity results for integral points on affine varieties, in the spirit of the<br />

Lang-Vojta conjecture. In particular, let X be a projective variety of dimension d ≥ 2<br />

over a number field K (resp., over C). Let H be the sum of 2d properly intersecting<br />

ample divisors on X. We show that any set of quasi-integral points (resp., any integral<br />

curve) on X − H is not Zariski-dense.<br />

1. Introduction<br />

On s’intéresse ici aux solutions à coordonnées (quasi-)entières de systèmes d’équations<br />

polynomiales à coefficients dans un corps de nombres K: on prouve qu’une variété<br />

projective sur K privée de suffisamment d’hypersurfaces n’a pas beaucoup de points<br />

(quasi-)entiers. On a également les résultats analogues en géométrie hyperbolique:<br />

une telle variété (sur C) n’a pas de courbe entière non dégénérée.<br />

Plus précisément, soit S un ensemble fini de places de K. On note OKS l’anneau<br />

des S-entiers de K. Cet article s’inscrit dans le cadre d’un problème important de la<br />

géométrie arithmétique.<br />

CONJECTURE 1.1 ([6, Conjecture 4.2], [13, Conjecture 3.4.3])<br />

Soit X une variété projective lisse sur K de diviseur canonique KX.SoitD un diviseur<br />

effectif sur X, à croisements normaux. Posons Y = X − D. On suppose KX + D big<br />

(par exemple, ample) sur X. Alors tout ensemble E ⊂ Y (K) S-entier sur Y est non<br />

Zariski-dense dans Y .<br />

DUKE MATHEMATICAL JOURNAL<br />

Vol. 158, No. 1, c○ 2011 DOI 10.1215/00127094-1276292<br />

Received 14 December 2009. Revision received 16 July 2010.<br />

2010 Mathematics Subject Classification. Primary 14G25; Secondary 11J97, 11G35.<br />

13


14 PASCAL AUTISSIER<br />

L’hypothèse sur KX + D est une condition de positivité de ce diviseur. Notons que<br />

cette conjecture est encore largement ouverte: le cas où X = P2 K n’est par exemple pas<br />

connu. Levin en a proposé des cas particuliers intéressants, lorsque D a suffisamment<br />

de composantes irréductibles.<br />

CONJECTURE 1.2 ([8, Conjecture 5.3A])<br />

Soit X une variété projective sur K de dimension d ≥ 1. Soient D1,...,Dr des<br />

diviseurs effectifs big sur X qui se coupent proprement. Posons Y = X−D1∪···∪Dr.<br />

On suppose r ≥ d + 2. Alors aucun ensemble S-entier sur Y n’est Zariski-dense<br />

dans Y .<br />

Pour une explication du lien entre 1.1 et 1.2, on pourra consulter [8, paragraphe 14.2].<br />

Un théorème classique de Siegel montre cette conjecture lorsque X est une courbe.<br />

Le cas des surfaces lisses est obtenu par Levin [8, théorème 6.2A.c].<br />

Par ailleurs, c’est une conséquence directe du théorème du sous-espace (voir<br />

Schmidt [11], Schlickewei [10]) lorsque X = P d K et les Di sont des hyperplans<br />

en position générale. Plus généralement, cet énoncé est connu de Vojta [15, corollaire<br />

0.3] pour X lisse et r ≥ d + ρ + 1, oùρ désigne le nombre de Picard de<br />

X K.<br />

Dans un article récent, Corvaja, Levin, et Zannier [2] démontrent la conjecture<br />

1.2 avec d 2 − d + 1 diviseurs Di amples sur X lisse (d ≥ 3). On se propose ici<br />

d’améliorer ce résultat en considérant un nombre de diviseurs linéaire en d (au lieu<br />

de quadratique).<br />

THÉORÈME 1.3<br />

Soit X une variété projective sur K de dimension d ≥ 2. Soient D1,...,D2d des<br />

diviseurs effectifs amples sur X qui se coupent proprement. Posons Y = X − D1 ∪<br />

···∪D2d. Alors tout ensemble E ⊂ Y (K) S-entier sur Y est non Zariski-dense dans<br />

Y .<br />

On l’obtient comme corollaire d’un critère (théorème 2.11) qui donne des conditions<br />

géométriques de non-Zariski-densité des points S-entiers.<br />

La démonstration s’inspire de la méthode introduite par Corvaja et Zannier dans<br />

le cas des courbes (voir [3]) et des surfaces (voir [4]). Cette méthode, dont l’ingrédient<br />

arithmétique principal est le théorème du sous-espace, a ensuite été étendue en dimension<br />

supérieure par Levin [8], l’auteur [1], et par Corvaja, Levin, et Zannier [2].<br />

L’idée nouvelle ici est de considérer un faisceau cohérent codant les ordres<br />

d’annulation optimaux en les Di et de prouver un théorème de convexité (théorème<br />

3.6) sur les sections globales de ce faisceau.<br />

À titre d’application du critère 2.11, on montre en outre l’énoncé suivant.


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 15<br />

THÉORÈME 1.4<br />

Soit X une variété projective sur K de dimension d ≥ 2. Soient D1,...,Dd+2 des<br />

diviseurs effectifs non nuls sur X qui se coupent proprement. Posons L = d+2 i=1 Di<br />

et Y = X − D1 ∪···∪Dd+2. On suppose que L − (d + 1)Di est ample pour tout<br />

i ∈{1,...,d+ 2}. Alors aucun ensemble S-entier sur Y n’est Zariski-dense dans Y .<br />

L’hypothèse sur les L − (d + 1)Di est vérifiée lorsque les Di vivent dans un cône<br />

“suffisamment étroit” du groupe de Néron-Severi de X.<br />

Par ailleurs, Vojta [13] adéveloppé un “dictionnaire” entre la géométrie diophantienne<br />

et la théorie de Nevanlinna: l’étude des points S-entiers sur les variétés sur<br />

K est mise en analogie avec l’étude des courbes entières sur les variétés complexes.<br />

À titre d’exemple, le théorème de Siegel sur les courbes est l’analogue du théorème<br />

de Picard.<br />

Pour l’étayer, on obtient aussi le critère (théorème 2.12) qui, dans ce dictionnaire,<br />

correspond au théorème 2.11.<br />

La section 2 décrit les critères de quasi-hyperbolicité, qui sont démontrés à la<br />

section 4. La section 3 donne le théorème de convexité 3.6 (ainsi qu’un lemme<br />

d’algèbre locale prouvé à la section 6). On en déduit les théorèmes 1.3 et 1.4 à la<br />

section 5.<br />

2. Définitions et énoncés<br />

Soit K un corps. Commençons par quelques définitions de géométrie.<br />

Conventions<br />

On appelle variété sur K tout schéma quasi-projectif et géométriquement intègre sur<br />

K. Le mot “diviseur” sous-entend “diviseur de Cartier”.<br />

Soit X une variété projective sur K de dimension d ≥ 1. Lorsque L est un<br />

diviseur sur X tel que h 0 (X, L) ≥ 1, ondésigne par BL le lieu de base de Ɣ(X, L) et<br />

par L : X − BL → P(Ɣ(X, L)) le morphisme défini par Ɣ(X, L). Pour tout diviseur<br />

effectif D sur X, on note 1D la section globale de OX(D) qu’il définit.<br />

Définition 2.1<br />

Un diviseur L sur X est dit grand lorsque h 0 (X, L) ≥ 1 et L est génériquement fini.<br />

Définition 2.2<br />

Soient D1,...,Dr des diviseurs effectifs sur X. On dit que D1,...,Dr se coupent<br />

proprement lorsque pour toute partie I non vide de {1,...,r}, la section globale<br />

(1Di)i∈I de <br />

i∈I OX(Di) est régulière (autrement dit, pour tout x ∈ <br />

i∈I Di, en<br />

notant ϕi une équation locale de Di en x, les(ϕi)i∈I forment une suite régulière de<br />

l’anneau local OX,x).


16 PASCAL AUTISSIER<br />

Remarque 2.3<br />

Supposons X de Cohen-Macaulay (par exemple, lisse sur K); alors d’après [5, lemme<br />

A.7.1], les diviseurs D1,...,Dr se coupent proprement si et seulement si pour toute<br />

partie I non vide de {1,...,r}, le fermé <br />

i∈I Di est purement de codimension #I<br />

dans X (éventuellement vide).<br />

Définition 2.4<br />

Lorsque L est un diviseur sur X tel que q = h 0 (X, L) ≥ 1 et E un diviseur effectif<br />

non nul sur X, on pose<br />

α(L, E) = 1<br />

q<br />

<br />

h 0 (X, L − kE).<br />

Introduisons ensuite les notions d’hyperbolicité étudiées dans cet article.<br />

k≥1<br />

Convention 2.5<br />

Lorsque K est un corps de nombres et v une place de K,ondésigne par Kv le complété<br />

de K en v, et on normalise la valeur absolue ||v de sorte que |x|v =|x| [Kv:R] si v est<br />

archimédienne et |p|v = p −[Kv:Qp] si v est p-adique.<br />

Lorsque K est un corps de nombres et S un ensemble fini de places de K, on note<br />

OK,S l’anneau des S-entiers de K (i.e., l’ensemble des x ∈ K tels que |x|v ≤ 1 pour<br />

toute place finie v/∈ S).<br />

Soit K un corps de nombres.<br />

Définition 2.6<br />

Soient Y une variété sur K, K ′ une extension finie de K,etS un ensemble fini de places<br />

de K ′ . Un ensemble E ⊂ Y (K ′ ) est dit S-entier sur Y lorsqu’il existe un OK ′ ,S-schéma<br />

intègre et quasi-projectif Y de fibre générique YK ′ tel que E ⊂ Y(OK ′ ,S).<br />

Définition 2.7<br />

Soient Y une variété sur K,etK ′ une extension finie de K. Un ensemble E ⊂ Y (K ′ )<br />

est dit quasi-entier sur Y lorsqu’il existe un ensemble fini S de places de K ′ tel que E<br />

soit S-entier sur Y .<br />

Définition 2.8<br />

Soit Y une variété sur K. On dit que Y est arithmétiquement quasi-hyperbolique<br />

lorsqu’il existe un fermé Z = Y tel que pour toute extension finie K ′ de K et tout<br />

ensemble quasi-entier E ⊂ Y (K ′ ) sur Y , l’ensemble E − Z(K ′ ) soit fini.


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 17<br />

Définition 2.9<br />

Soit U une variété complexe. Une courbe entière sur U est une application holomorphe<br />

f : C → U non constante.<br />

Définition 2.10<br />

Soit U une variété complexe. On dit que U est Brody quasi-hyperbolique lorsqu’il<br />

existe un fermé Z = U tel que pour toute courbe entière f sur U, onaitf (C) ⊂ Z.<br />

Terminons cette section avec les énoncés principaux de ce travail. Ces critères de<br />

quasi-hyperbolicité généralisent (et simplifient) des résultats de Levin [8, théorèmes<br />

8.3A et 8.3B] et de l’auteur (voir [1,théorèmes 3.3 et 3.5]).<br />

THÉORÈME 2.11<br />

Soit X une variété projective sur K de dimension d ≥ 1. Soient D1,...,Dr des<br />

diviseurs effectifs non nuls sur X qui se coupent proprement. Posons Y = X − D1 ∪<br />

···∪Dr. Soit m ≥ 1 un entier. On suppose que le diviseur L = m r i=1 Di est grand<br />

sur X et que α(L, Di) >mpour tout i ∈{1,...,r}. AlorsYest arithmétiquement<br />

quasi-hyperbolique.<br />

THÉORÈME 2.12<br />

Soit V une variété complexe projective de dimension d ≥ 1. Soient D1,...,Dr des<br />

diviseurs effectifs non nuls sur V qui se coupent proprement. Posons U = V − D1 ∪<br />

···∪Dr. Soit m ≥ 1 un entier. On suppose que le diviseur L = m r i=1 Di est<br />

grand sur V et que α(L, Di) >mpour tout i ∈{1,...,r}. AlorsUest Brody<br />

quasi-hyperbolique.<br />

3. Préliminaires<br />

Soient K un corps et W un K-espace vectoriel de dimension finie.<br />

Définitions<br />

Une filtration de W est une famille décroissante F = (Fx)x∈R+ de sous-espaces<br />

vectoriels de W telle que Fx ={0} pour tout x assez grand. Une base B de W est<br />

dite adaptée à la filtration F lorsque B ∩ Fx est une base de Fx pour tout réel x ≥ 0.<br />

LEMME 3.1<br />

Soient F et G deux filtrations de W . Il existe alors une base de W adaptée à F et à<br />

G .<br />

Démonstration<br />

C’est le lemme 3.2 de [4]. <br />

Soit r un entier ≥ 1. On note ici l’ensemble des t = (t1,...,tr) ∈ Rr + tels que<br />

t1 +···+tr = 1.


18 PASCAL AUTISSIER<br />

Définition 3.2<br />

Une partie N de N r est dite saturée lorsque a + b ∈ N pour tout a ∈ N r et tout<br />

b ∈ N.<br />

Soit A un anneau local noethérien. Soit (ϕ1,...,ϕr) une suite régulière de A. Pour<br />

toute partie saturée N de Nr ,ondésigne par I(N) l’idéal de A engendré par l’ensemble<br />

{ϕ b1<br />

1 ,...,ϕbr r , b ∈ N}. On aura besoin d’un résultat d’algèbre locale (qui ne sera<br />

démontré qu’àlasection6).<br />

LEMME 3.3<br />

Soient N et M deux parties saturées de N r . On a alors l’égalité d’idéaux<br />

I(N) ∩ I(M) = I(N ∩ M).<br />

Remarque 3.4<br />

On utilisera dans la suite le cas particulier suivant du lemme 3.3: pour t ∈ et<br />

x ∈ R+, notons N(t,x) l’ensemble des b ∈ Nr tels que t1b1 +···+trbr ≥ x. Ona<br />

alors l’inclusion<br />

I N(t,x) ∩ I N(u,y) ⊂ I N(λt + (1 − λ)u,λx+ (1 − λ)y) <br />

pour tous t, x, u, y, et tout λ ∈ [0, 1], puisque N(t,x) ∩ N(u,y) ⊂ N(λt + (1 −<br />

λ)u,λx+ (1 − λ)y).<br />

Maintenant, soit X une variété projective sur K de dimension d ≥ 1. SoitLun diviseur sur X tel que q = h0 (X, L) ≥ 1. SoientD1,...,Dr des diviseurs effectifs<br />

sur X qui se coupent proprement, tels que r i=1 Di soit non vide.<br />

Introduisons quelques notations. Soit t ∈ . Pour x ∈ R+, ondéfinit l’idéal<br />

J(t,x) de OX par<br />

J(t,x) = <br />

b∈N(t,x)<br />

OX<br />

<br />

−<br />

r<br />

biDi<br />

i=1<br />

et on pose F (t)x = Ɣ(X, J(t,x) ⊗ L). Observons que (F (t)x)x∈R+ est une filtration<br />

de Ɣ(X, L). Pour tout s ∈ Ɣ(X, L) −{0}, on note µt(s) = sup{y ∈ R+ | s ∈ F (t)y}.<br />

On pose finalement<br />

F ( t ) = 1<br />

+∞<br />

h<br />

q<br />

0 X, J( t,x) ⊗ L dx.<br />

Remarque 3.5<br />

Soit B = (s1,...,sq) une base de Ɣ(X, L). On a alors l’inégalité<br />

F ( t ) ≥ 1<br />

q<br />

+∞<br />

0<br />

0<br />

# F ( t )x ∩ B dx = 1<br />

q<br />

<br />

,<br />

q<br />

µt(sk),<br />

k=1


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 19<br />

qui est même une égalité siB est adaptée à la filtration F ( t ).<br />

Voici le résultat principal de cette section.<br />

THÉORÈME 3.6<br />

L’application F : → R+ ainsi définie est concave. On a en particulier F ( t ) ≥<br />

mini α(L, Di) pour tout t ∈ .<br />

Démonstration<br />

Soient t ∈ , u ∈ et λ ∈ [0, 1]. Il s’agit de montrer que<br />

F λt + (1 − λ)u ≥ λF ( t ) + (1 − λ)F ( u ).<br />

Le lemme 3.1 construit une base B = (s1,...,sq) de W = Ɣ(X, L) adaptée aux<br />

filtrations F ( t ) et F ( u ).<br />

Soit (x,y) ∈ R2 + . Le lemme 3.3 (voir aussi remarque 3.4) fournit l’inclusion<br />

d’idéaux J(t,x) ∩ J(u,y) ⊂ J(λt + (1 − λ)u,λx+ (1 − λ)y), puisque les diviseurs<br />

D1,...,Dr se coupent proprement. On a en particulier l’inclusion d’espaces vectoriels<br />

F ( t )x ∩ F ( u )y ⊂ F λt + (1 − λ) u <br />

λx+(1−λ)y .<br />

Soit s ∈ W −{0}. Ce qui précède implique que s appartient à F (λt + (1 −<br />

λ)u)λx+(1−λ)y pour tout x


20 PASCAL AUTISSIER<br />

Choisissons une section rationnelle s de L définie et non nulle en P . On pose<br />

1<br />

h ˆL(P ) =−<br />

[K ′ <br />

ln s(P )v,<br />

: Q]<br />

où v parcourt l’ensemble des places de K ′ .Ceréel ne dépend pas du choix de s.<br />

On va utiliser la version suivante du théorème du sous-espace.<br />

PROPOSITION 4.2<br />

On suppose L grand sur X. Notons q = h0 (X, L). Soient s1,...,sl des sections non<br />

nulles engendrant Ɣ(X, L). Soit ε>0. Il existe alors un fermé Z = X tel que pour<br />

toute extension finie K ′ de K et tout ensemble fini S de places de K ′ , l’ensemble des<br />

points P ∈ (X − Z)(K ′ ) vérifiant<br />

<br />

v∈S<br />

max<br />

J ∈L<br />

<br />

j∈J<br />

v<br />

ln sj(P ) −1<br />

v ≥ (q + qε)[K′ : Q]h ˆL(P ) (1)<br />

est fini, où L désigne l’ensemble des parties J de {1,...,l} telles que (sj)j∈J soit une<br />

base de Ɣ(X, L).<br />

Démonstration<br />

Posons P = P(Ɣ(X, L)).Désignons par π : X ′ → X l’éclatement de X relativement<br />

à BL et par E = π −1 (BL) le diviseur exceptionnel. En notant L ′ = OX ′(−E) ⊗ π ∗L, on a un morphisme L ′ : X′ → P génériquement fini qui prolonge L. Il existe donc<br />

un fermé Z1 = X ′ tel que L ′ |X ′ −Z1 soit à fibres finies.<br />

On munit L0 = OP(1) d’une métrique adélique ( ′ v )v telle que la section globale<br />

1E de ∗ L ′L∨0 ⊗ π ∗L vérifie<br />

sup<br />

X ′ 1E<br />

v<br />

′<br />

v ≤ 1<br />

pour toute place v. On applique alors la version de Vojta [14, théorème 0.3, reformulation<br />

3.4] du théorème du sous-espace.<br />

Il existe une réunion finie H de K-hyperplans de P P q−1<br />

K telle que pour toute<br />

extension finie K ′ de K et tout ensemble fini S de places de K ′ , l’ensemble des points<br />

P ∈ (P − H )(K ′ ) vérifiant<br />

<br />

ln sj(P ) ′−1<br />

v<br />

v∈S<br />

max<br />

J ∈L<br />

j∈J<br />

≥ (q + qε)[K′ : Q]hL0 ˆ (P )<br />

est fini.<br />

Prenons Z = BL ∪ π(Z1 ∪ −1<br />

L ′ (H )). Alors pour toute extension finie K′ de K<br />

et tout ensemble fini S de places de K ′ , l’ensemble des points P ∈ (X − Z)(K ′ )<br />

vérifiant (1) est fini.


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 21<br />

Montrons donc les critères de quasi-hyperbolicité delasection2.<br />

Démonstration du théorème 2.11<br />

On reprend en l’améliorant la démonstration du [1,théorème 3.3]. On procède en deux<br />

étapes: dans la première, on construit un fermé Z = X candidat à contenir “presque<br />

tous les points entiers”; dans la seconde, on prouve que Y est arithmétiquement quasihyperbolique.<br />

Étape 1<br />

Posons ε = (1/4m)(min α(L, Di) − m) et q = h<br />

i 0 (X, L), et fixons un entier b ≥<br />

(1/ε + 3)r. On choisit aussi une base B0 de Ɣ(X, L).<br />

Désignons par P l’ensemble des parties I non vides de {1,...,r} telles que<br />

<br />

i∈I Di soit non vide. Soit I ∈ P . On note I l’ensemble des a = (ai)i ∈ NI tels<br />

que <br />

i∈I ai = b. Soita∈I . Pour x ∈ R+, ondéfinit l’idéal J(x) de OX par<br />

J(x) = <br />

− <br />

b<br />

OX<br />

où la somme porte sur les b ∈ NI tels que<br />

<br />

aibi ≥ bx,<br />

i∈I<br />

i∈I<br />

biDi<br />

et on pose F (I,a)x = Ɣ(X, J(x) ⊗ L). Le lemme 3.1 fournit une base BI,a de<br />

Ɣ(X, L) adaptée à la filtration F (I,a).<br />

On munit chaque faisceau OX(Di) d’une métrique adélique. Appliquons le<br />

théorème du sous-espace (proposition 4.2) avec {s1,...,sl} =B0 ∪ <br />

I,a BI,a (remarquons<br />

que cette réunion est finie puisque P et les I le sont).<br />

Il existe un fermé Z = X tel que pour toute extension finie K ′ de K et tout<br />

ensemble fini S de places de K ′ , l’ensemble des points P ∈ (X − Z)(K ′ ) vérifiant<br />

l’inégalité (1) est fini.<br />

Étape 2<br />

Soient K ′ une extension finie de K et S un ensemble fini de places de K ′ contenant<br />

les places archimédiennes. Soit E ⊂ Y (K ′ ) un ensemble S-entier sur Y . Raisonnons<br />

par l’absurde en supposant E − Z(K ′ ) infini. Choisissons une suite injective (Pn)n≥0<br />

d’éléments de E − Z(K ′ ).<br />

Quitte à extraire, on peut supposer (par compacité) que pour tout v ∈ S, la suite<br />

(Pn)n≥0 converge dans X(K ′ v ) vers un yv ∈ X(K ′ v ).<br />

Pour tout v ∈ S, on note Iv l’ensemble des i ∈{1,...,r} tels que yv ∈ Di.<br />

Quitte à extraire de nouveau, on peut supposer que pour tout v ∈ S tel que Iv soit non


22 PASCAL AUTISSIER<br />

vide et tout i ∈ Iv, la suite<br />

<br />

ln 1Di(Pn)v<br />

<br />

ln 1Dj (Pn)v<br />

j∈Iv<br />

converge vers un tvi ∈ [0, 1]. Observons que l’on a <br />

i∈Iv tvi = 1.<br />

Fait: Soitv∈S. Il existe une base (s1v,...,sqv) de Ɣ(X, L) contenue dans<br />

{s1,...,sl} telle que l’on ait la minoration suivante pour tout n ≥ 0:<br />

−<br />

q<br />

ln skv(Pn)v ≥−(q + 2qε)ln1L(Pn)v − O(1), (2)<br />

k=1<br />

où le O(1) est indépendant de n.<br />

Prouvons ce fait. Si Iv est vide, on prend {s1v,...,sqv} =B0 et on obtient la<br />

minoration (2) en remarquant que ln 1L(Pn)v = O(1) (puisque yv /∈ L).<br />

On suppose maintenant Iv non vide. On a donc Iv ∈ P . Choisissons un a v =<br />

(avi)i ∈ Iv tel que avi ≤ (b + r)tvi pour tout i ∈ Iv. On prend alors {s1v,...,sqv} =<br />

BIv,a v .Vérifions que ce choix convient.<br />

Soit s ∈ Ɣ(X, L) −{0}. Notons µ(s) le plus grand rationnel µ tel que s ∈<br />

F (Iv,av )µ. Enécrivant localement s = <br />

b fb<br />

<br />

i∈Iv 1bi<br />

Di au voisinage de yv, on<br />

trouve<br />

− ln s(Pn)v ≥−max<br />

b<br />

Par définition des tvi, ona<br />

−bi ln 1Di(Pn)v ≥−<br />

n≥0<br />

<br />

bi ln 1Di(Pn)v − O(1).<br />

i∈Iv<br />

avibi<br />

b + r<br />

− mε<br />

r<br />

<br />

ln 1Dj (Pn)v<br />

pour tout n assez grand et tout i ∈ Iv.<br />

On en déduit l’inégalité (pour tout n ≥ 0)<br />

<br />

µ(s)b<br />

<br />

− ln s(Pn)v ≥− − mε ln 1Dj (Pn)v − O(1).<br />

b + r<br />

On écrit cette inégalité pour s = skv, puis on somme sur k. En observant que l’on a<br />

q<br />

µ(skv) =<br />

k=1<br />

j∈Iv<br />

j∈Iv<br />

+∞<br />

dim F (Iv,av )x dx (remarque 3.5)<br />

0<br />

≥ min α(L, Di)q (théorème 3.6)<br />

i<br />

b + r<br />

= (1 + 4ε)qm ≥ (q + 3qε)m,<br />

b


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 23<br />

on obtient alors<br />

−<br />

q<br />

ln skv(Pn)v ≥−(q + 2qε)m <br />

ln 1Dj (Pn)v − O(1).<br />

k=1<br />

Le fait énoncé (2) s’en déduit en remarquant que ln 1Dj (Pn)v = O(1) pour tout<br />

j/∈ Iv.<br />

Maintenant, l’ensemble E est S-entier sur Y , donc pour tout n ≥ 0, ona<br />

[K ′ : Q]h ˆL(Pn) =− <br />

ln 1L(Pn)v + O(1).<br />

v∈S<br />

j∈Iv<br />

En utilisant la minoration (2), on trouve (pour tout n ≥ 0)<br />

− <br />

v∈S<br />

q<br />

ln skv(Pn)v ≥ (q + 2qε)[K ′ : Q]h ˆL(Pn) − O(1).<br />

k=1<br />

D’où une contradiction avec (1). <br />

Démonstration du théorème 2.12<br />

On reprend de la même manière la démonstration du [1, théorème 3.5], qui utilise<br />

la version de Vojta [16, théorème 2] du théorème de Cartan au lieu du théorème du<br />

sous-espace. Les détails sont laissés au lecteur. <br />

5. Démonstration des théorèmes 1.3 et 1.4<br />

Soient K un corps de nombres et X une variété projective sur K de dimension<br />

d ≥ 2. Pour tous diviseurs L1,...,Ld sur X,ondésigne par <br />

L1 ···Ld leur nombre<br />

d’intersection.<br />

Définition 5.1<br />

Un diviseur L sur X est dit big lorsque lim inf<br />

n→+∞ (1/nd )h 0 (X, nL) > 0.<br />

Remarque 5.2<br />

D’après le lemme de Kodaira [7, p. 141], le diviseur L est big si et seulement si nL<br />

est grand pour tout entier n assez grand.<br />

Définition 5.3<br />

Soit L un diviseur big sur X. On dit que L est presque ample lorsqu’il existe un entier<br />

n ≥ 1 tel que BnL soit vide.<br />

On montre ici de légères extensions des théorèmes 1.3 et 1.4 de l’introduction.


24 PASCAL AUTISSIER<br />

THÉORÈME 1.3BIS<br />

Soient D1,...,D2d des diviseurs effectifs presque amples sur X qui se coupent<br />

proprement. Alors X − D1 ∪···∪D2d est arithmétiquement quasi-hyperbolique.<br />

Démonstration<br />

D’après [1, théorème 4.4], il existe (m1,...,m2d) ∈ (N ∗ ) 2d tel qu’en posant L =<br />

2d<br />

i=1 miDi, onait<br />

lim inf<br />

n→+∞<br />

1<br />

n α(nL, miDi) > 1 pour tout i ∈{1,...,2d}.<br />

On en déduit le résultat en appliquant le théorème 2.11 (les diviseurs<br />

m1D1,...,m2dD2d se coupent encore proprement). <br />

THÉORÈME 1.4BIS<br />

Soient D1,...,Dr des diviseurs effectifs non nuls sur X qui se coupent proprement,<br />

avec r ≥ d + 2. Posons L = r i=1 Di et Y = X − D1 ∪···∪Dr. On suppose<br />

que L − (d + 1)Di est ample pour tout i ∈{1,...,r}.AlorsYest arithmétiquement<br />

quasi-hyperbolique.<br />

Démonstration<br />

D’après le lemme 5.4 ci-dessous, on a lim inf<br />

n→+∞ (1/n)α(nL, Di) > 1 pour tout i ∈<br />

{1,...,r}. On obtient le résultat en appliquant le théorème 2.11. <br />

LEMME 5.4<br />

Soient L un diviseur ample sur X et D un diviseur effectif non nul sur X. Soit t ≥ 1<br />

un entier tel que L − tD soit ample. On a alors les minorations<br />

lim inf<br />

n→+∞<br />

1<br />

α(nL, D) ≥<br />

n<br />

t<br />

(d + 1) L d<br />

d<br />

j=0<br />

L d−j (L − tD) j > t<br />

d + 1 .<br />

Démonstration<br />

La formule de Hirzebruch-Riemann-Roch donne que χ(X, nL−kD) est une fonction<br />

polynomiale en (n, k) dont on peut expliciter la composante homogène dominante.<br />

Pour tout (n, k) tel que 0 ≤ k ≤ tn,onaχ(X, nL − kD) = (1/d!)〈(nL −<br />

kD) d 〉+O(n d−1 ).<br />

Par ailleurs, pour tout n assez grand et tout k ∈{0,...,tn}, onah i (X, nL −<br />

kD) = 0 pour tout i ≥ 1, puisque L et L − tD sont amples. On a en particulier<br />

h 0 (X, nL − kD) = (1/d!)〈(nL − kD) d 〉+O(n d−1 ).


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 25<br />

On trouve ainsi les estimations suivantes:<br />

tn<br />

k=1<br />

h 0 (X, nL − kD) = 1<br />

d!<br />

= 1<br />

d!<br />

= 1<br />

d!<br />

tn<br />

k=1<br />

d<br />

j=0 k=1<br />

d<br />

j=0<br />

〈(nL − kD) d 〉+O(n d−1 ) <br />

tn<br />

Or un calcul montre (par multilinéarité) la formule<br />

d<br />

j=0<br />

C j (−1) j<br />

d−j j<br />

d L D<br />

j + 1 t j+1 = t<br />

d + 1<br />

C j<br />

d 〈Ld−j D j 〉n d−j (−k) j + O(n d )<br />

C j<br />

d 〈Ld−j D j 〉 (−1)j<br />

j + 1 t j+1 n d+1 + O(n d ).<br />

d<br />

〈L d−j (L − tD) j 〉.<br />

D’où la première inégalité del’énoncé.<br />

La deuxième inégalité s’en déduit facilement: les diviseurs L et L − tD sont<br />

amples, donc on a 〈L d−j (L − tD) j 〉 > 0 pour tout j ∈{1,...,d}. <br />

6. Démonstration du lemme 3.3<br />

Soit A un anneau local noethérien. On aura besoin d’un résultat classique dû à Krull.<br />

LEMME 6.1<br />

Soient I et J deux idéaux de A avec J = A. On a alors l’égalité<br />

<br />

(I + J k ) = I.<br />

k≥1<br />

Démonstration<br />

Il suffit d’appliquer [9, corollaire 11.D.2, p. 69], dans l’anneau A/I. <br />

Soit (ϕ1,...,ϕr) une suite régulière de A. Rappelons que pour toute partie saturée N<br />

de Nr ,ondésigne par I(N) l’idéal 〈ϕ b1<br />

1 ···ϕbr r ,b∈ N〉 de A. Le lemme 3.3 est une<br />

conséquence directe de l’énoncé suivant.<br />

LEMME 6.2<br />

Soit N une partie saturée de N r . On a alors l’égalité d’idéaux<br />

Démonstration<br />

Notons ici I1(N) = <br />

I(N) = <br />

c∈N r −N<br />

〈ϕ c1+1<br />

1<br />

c∈Nr −N 〈ϕc1+1 1 ,...,ϕcr +1<br />

r<br />

j=0<br />

,...,ϕ cr +1<br />

r<br />

〉.<br />

〉. Vérifions d’abord l’inclusion


26 PASCAL AUTISSIER<br />

I(N) ⊂ I1(N):sib ∈ N et c ∈ Nr − N, alors il existe un indice i tel que bi ≥ ci + 1,<br />

donc ϕ b1<br />

1 ,...,ϕbr r ∈ ϕci+1 i A ⊂〈ϕc1+1 1 ,...,ϕcr +1<br />

r 〉.<br />

On prouve l’inclusion I1(N) ⊂ I(N) par récurrence sur r. Le cas r = 1 est<br />

facile (avec le lemme 6.1 si N est vide). Supposons r ≥ 2 et le résultat au cran r − 1.<br />

Montrons par une (deuxième) récurrence sur k que I1(N) ⊂ I(N) + ϕk 1A pour tout<br />

k ∈ N. Sik = 0, cette inclusion est évidente. Supposons k ≥ 1 et le résultat au cran<br />

k − 1. Soitsun élément de I1(N). Ils’écrit s = x + ϕ k−1<br />

1 a avec un x ∈ I(N) et un<br />

a ∈ A.<br />

Posons Nk = {b = (b2,...,br) ∈ Nr−1 | (k − 1,b2,...,br) ∈ N}. Soit<br />

b ∈ N r−1 − Nk. Alorsϕ k−1<br />

1 a = s − x appartient à I1(N) donc en particulier à<br />

〈ϕk 1 ,ϕb2+1 2 ,...,ϕbr +1<br />

r 〉. Autrement dit, il existe a ′ ∈ A tel que ϕ k−1<br />

1 a − ϕk 1a′ ∈<br />

〈ϕ b2+1<br />

2 ,...,ϕbr +1<br />

r 〉. Puisque (ϕ b2+1<br />

2 ,...,ϕbr +1<br />

r ,ϕ1) est une suite régulière de A, on<br />

obtient a − ϕ1a ′ ∈〈ϕ b2+1<br />

2 ,...,ϕbr +1<br />

r 〉 en simplifiant par ϕ k−1<br />

1 .Réduisons modulo ϕ1<br />

on trouve dans A ′ = A/ϕ1A que ā ∈〈¯ ϕ2 b2+1 ,..., ϕr ¯ br +1 〉.<br />

On a ainsi que ā ∈ <br />

b∈Nr−1 〈 ¯ −Nk ϕ2 b2+1 ,..., ϕr ¯ br +1 〉.Or(¯ ϕ2,..., ϕr) ¯ est une suite<br />

régulière de A ′ , donc ā ∈〈¯ ϕ2 b2 ··· ϕr ¯ br , b ∈ Nk〉 par la première hypothèse de<br />

récurrence. On écrit a = y + ϕ1a1 avec y ∈〈ϕ b2<br />

2 ···ϕbr r ,b∈ Nk〉 et a1 ∈ A. Ilen<br />

découle l’égalité s = x + ϕ k−1<br />

1 y + ϕk 1a1 avec ϕ k−1<br />

1 y ∈〈ϕk−1 1 ϕb2 2 ···ϕbr r ,b∈ Nk〉 ⊂<br />

I(N), qui implique bien que s est dans I(N) + ϕk 1A. D’où l’inclusion I1(N) ⊂ <br />

k∈N (I(N) + ϕk 1A). On conclut par le lemme 6.1. <br />

Remerciements. Je remercie Yuri Bilu pour m’avoir fourni la référence [2]. Je remercie<br />

également les rapporteurs pour leurs suggestions pertinentes.<br />

References<br />

[1] P. AUTISSIER, Géométrie, points entiers et courbes entières, Ann. Sci. Éc. Norm.<br />

Supér. (4) 42 (2009), 221 – 239. MR 2518077 14, 17, 21, 23, 24<br />

[2] P. CORVAJA, A. LEV<strong>IN</strong> et U. ZANNIER, Integral points on threefolds and other<br />

varieties, Tohoku Math. J. 61 (2009), 589 – 601. MR 2598251 14, 26<br />

[3] P. CORVAJA et U. ZANNIER, A subspace theorem approach to integral points on curves,<br />

C. R. Math. Acad. Sci. Paris 334 (2002), 267 – 271. MR 1891001 14<br />

[4] ———, On integral points on surfaces, Ann. of. Math. (2) 160 (2004), 705 – 726.<br />

MR 2123936 14, 17<br />

[5] W. FULTON, Intersection Theory, 2nd ed., Ergeb. Math. Grenzgeb. 3, Springer, Berlin,<br />

1998. MR 1644323 16<br />

[6] S. LANG, Number Theory, III: Diophantine Geometry, Encyclopaedia Math. Sci. 60,<br />

Springer, Berlin, 1991. MR 1112552 13


SUR LA NON-DENSITÉ DES PO<strong>IN</strong>TS ENTIERS 27<br />

[7] R. LAZARSFELD, Positivity in Algebraic Geometry, I: Classical Setting—Line Bundles<br />

and Linear Series, Ergeb. Math. Grenzgeb. 48, Springer, Berlin, 2004.<br />

MR 2095471 23<br />

[8] A. LEV<strong>IN</strong>, Generalizations of Siegel’s and Picard’s theorems, Ann. of Math. (2) 170<br />

(2009), 609 – 655. MR 2552103 14, 17<br />

[9] H. MATSUMURA, Commutative Algebra, 2nd ed., Math. Lecture Note Ser. 56,<br />

Benjamin/Cummings, Reading, Mass., 1980. MR 0575344 25<br />

[10] H. P. SCHLICKEWEI, The p-adic Thue-Siegel-Roth-Schmidt theorem, Arch. Math.<br />

(Basel) 29 (1977), 267 – 270. MR 0491529 14<br />

[11] W. M. SCHMIDT, Diophantine approximation, Lecture Notes in Math. 785, Springer<br />

Berlin, 1980. MR 0568710 14<br />

[12] J.-P. SERRE, Lectures on the Mordell-Weil Theorem, 3rd ed., Aspects Math., Friedr.<br />

Vieweg and Sons, Braunschweig, 1997. MR 1757192<br />

[13] P. VOJTA, Diophantine Approximations and Value Distribution Theory, Lecture Notes<br />

in Math. 1239, Springer, Berlin, 1987. MR 0883451 13, 15<br />

[14] ———, A refinement of Schmidt’s subspace theorem, Amer. J. Math. 111 (1989),<br />

489 – 518. MR 1002010 20<br />

[15] ———, Integral points on subvarieties of semiabelian varieties, I, Invent. Math. 126<br />

(1996), 133 – 181. MR 1408559 14<br />

[16] ———, On Cartan’s theorem and Cartan’s conjecture, Amer. J. Math. 119 (1997),<br />

1 – 17. MR 1428056 23<br />

[17] S. ZHANG, Small points and adelic metrics,J.AlgebraicGeom.4 (1995), 281 – 300.<br />

MR 1311351 19<br />

Institut de Mathématiques de Bordeaux, Université Bordeaux I, 33405 Talence CEDEX,<br />

France; pascal.autissier@math.u-bordeaux1.fr


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS<br />

CASIM ABBAS<br />

Abstract<br />

Giroux showed that every contact structure on a closed 3-dimensional manifold is<br />

supported by an open book decomposition. We extend this result by showing that the<br />

open book decomposition can be chosen in such a way that the pages are solutions to<br />

a homological perturbed holomorphic curve equation.<br />

Contents<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2. Existence and local foliations . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3. From local foliations to global ones . . . . . . . . . . . . . . . . . . . 52<br />

4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

Appendix. Some local computations near the punctures . . . . . . . . . . . 78<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

1. Introduction<br />

This paper is the starting point of a larger program by the author, Hofer, and Lisi<br />

investigating a perturbed holomorphic curve equation in the symplectization of a 3dimensional<br />

contact manifold (see [6], [7]). One aim of this program is to provide<br />

an alternative proof of the Weinstein conjecture in dimension 3 as outlined in [4]<br />

complementing Taubes’s gauge theoretical proof (see [33], [34]). A special case of<br />

this paper’s main result has been used in the proof of the Weinstein conjecture for<br />

planar contact structures in [4]. Another reason for studying this equation is to construct<br />

foliations by surfaces of section with nontrivial genus. This is usually impossible to<br />

do with the unperturbed holomorphic curve equation since solutions generically do<br />

not exist.<br />

Consider a closed 3-dimensional manifold M equipped with a contact form λ.<br />

This is a 1-form which satisfies λ ∧ dλ = 0 at every point of M. We denote the<br />

associated contact structure by ξ = ker λ, and we denote the Reeb vector field by Xλ.<br />

DUKE MATHEMATICAL JOURNAL<br />

Vol. 158, No. 1, c○ 2011 DOI 10.1215/00127094-1276301<br />

Received 20 July 2009. Revision received 8 July 2010.<br />

2010 Mathematics Subject Classification. Primary 53D35; Secondary 35J62.<br />

29


30 CASIM ABBAS<br />

Recall that the Reeb vector field is defined by the two equations<br />

iXλdλ = 0 and iXλλ = 1.<br />

Definition 1.1 (Open book decomposition)<br />

Assume that K ⊂ M is a link in M and that τ : M\K → S 1 is a fibration so<br />

that the fibers Fϑ = τ −1 (ϑ) are interiors of compact embedded surfaces ¯Fϑ with<br />

∂ ¯Fϑ = K, where ϑ is the coordinate along K. We also assume that K has a tubular<br />

neighborhood K × D, D ⊂ R 2 being the open unit disk, such that τ restricted to<br />

K × (D\{0}) is given by τ(ϑ, r, φ) = φ, where (r, φ) are polar coordinates on D.<br />

Then we call τ an open book decomposition of M, the link K is called the binding of<br />

the open book decomposition, and the surfaces Fϑ are called the pages of the open<br />

book decomposition.<br />

It is a well-known result in 3-dimensional topology that every closed 3-dimensional<br />

orientable manifold admits an open book decomposition. Indeed, Alexander proved<br />

the following theorem in 1923 (see [10], [30]):<br />

<strong>THEOREM</strong> 1.2<br />

Every closed, orientable manifold M of dimension 3 is diffeomorphic to<br />

W (h) ∪Id (∂W × D 2 ),<br />

where D 2 is the closed unit disk in R 2 ,whereW is an orientable surface with boundary,<br />

and where h : W → W is an orientation-preserving diffeomorphism which restricts<br />

to the identity near ∂W.HereW (h) denotes the manifold obtained from W × [0, 2π]<br />

by identifying (x,0) with (h(x), 2π).<br />

The above decomposition is an open book decomposition, and the pages are given by<br />

Fϑ := (W ×{ϑ}) ∪Id (∂W × Iϑ), 0 ≤ ϑ


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 31<br />

dλ ′ induces an area form on each fiber Fϑ with K consisting of closed orbits of the<br />

Reeb vector field Xλ ′,andλ′ orients K as the boundary of (Fϑ,dλ ′ ).<br />

We refer to a contact form λ ′ above as a Giroux contact form. Note that λ ′ is not unique<br />

and that it is in general different from the original contact form λ. The following<br />

theorem by Giroux guarantees existence of such open book decompositions, and it<br />

contains a uniqueness statement as well (see also [16, Proposition 2]).<br />

<strong>THEOREM</strong> 1.4 ([16, Theorem 3])<br />

Every co-oriented contact structure ξ = ker λ on a closed 3-dimensional manifold is<br />

supported by some open book. Conversely, if two contact structures are supported by<br />

the same open book, then they are diffeomorphic.<br />

In the topological category, it is possible to modify an open book decomposition<br />

such that the pages of the new decomposition have lower genus at the expense of<br />

increasing the number of connected components of K. It was not known for some<br />

time whether a similar statement could also be made in the context of supporting open<br />

book decompositions. In particular, it was unclear whether every contact structure<br />

was supported by an open book decomposition whose pages were punctured spheres<br />

(planar pages). The author and his collaborators could resolve the Weinstein conjecture<br />

for contact forms inducing a planar contact structure in 2005 (see [4]). So the question<br />

of whether all contact structures are planar became a priority, which prompted Etnyre<br />

to address it in [14]. He showed that overtwisted contact structures always admit<br />

supporting open book decompositions with planar pages, but many contact structures<br />

do not. Since then, planar open book decompositions have become an important tool<br />

in contact geometry.<br />

In this paper, we will prove that every contact structure has a supporting open book<br />

decomposition such that the pages solve a homological perturbed Cauchy-Riemann<br />

type equation which we now describe after introducing some notation. We write<br />

πλ = π : TM → ξ for the projection along the Reeb vector field Xλ. Fix a complex<br />

multiplication J : ξ → ξ so that the map ξ ⊕ ξ → R, definedby<br />

(h, k) → dλ(h, J k),<br />

defines a positive definite metric on the fibers. We call such complex multiplications<br />

compatible (with dλ). The equation of interest here is the following nonlinear firstorder<br />

elliptic system. The solutions consist of 5-tuplets (S,j,Ɣ,ũ, γ ) where (S,j)<br />

is a closed Riemann surface with complex structure j, Ɣ ⊂ S is a finite subset,<br />

ũ = (a,u) : ˙S → R × M is a proper map with ˙S = S \ Ɣ,andγ is a 1-form on S so


32 CASIM ABBAS<br />

that<br />

Here the energy E(ũ) is defined by<br />

⎧<br />

⎪⎨<br />

π ◦ Tu◦ j = J ◦ π ◦ Tu on ˙S<br />

(u<br />

⎪⎩<br />

∗λ) ◦ j = da + γ on ˙S<br />

dγ = d(γ ◦ j) = 0 on S<br />

E(ũ) < ∞.<br />

E(ũ) = sup<br />

ϕ∈<br />

<br />

˙S<br />

ũ ∗ d(ϕλ),<br />

(1.1)<br />

where consists of all smooth maps ϕ : R → [0, 1] with ϕ ′ (s) ≥ 0 for all s ∈ R.<br />

Note that equation (1.1) reduces to the usual pseudoholomorphic curve equation<br />

in the symplectization R × M if we set γ = 0. The following proposition, which is a<br />

modification of a result by Hofer [17], shows that solutions to problem (1.1) approach<br />

cylinders over periodic orbits of the Reeb vector field.<br />

PROPOSITION 1.5<br />

Let (M,λ) be a closed 3-dimensional manifold equipped with a contact form λ.Then<br />

the associated Reeb vector field has periodic orbits if and only if the associated PDE<br />

problem (1.1) has a nonconstant solution.<br />

Proof<br />

Let (S,j,Ɣ,ũ, γ ) be a nonconstant solution of (1.1). If Ɣ = ∅, then the results in [17]<br />

imply that, near a puncture, the solution is asymptotic to a periodic orbit (see also [3]<br />

for a complete proof). Here we use the fact that γ is exact near the punctures. The<br />

aim now is to show that, in the absence of punctures, the map a is constant while the<br />

image of u lies on a periodic Reeb orbit. Assume that Ɣ =∅.Since<br />

we find, after applying d, that<br />

u ∗ λ =−da ◦ j − γ ◦ j,<br />

ja =−d(da ◦ j) = u ∗ dλ.<br />

In view of the equation π ◦ Tu◦ j = J ◦ π ◦ Tu, we see that u ∗ dλ is a nonnegative<br />

integrand. Applying Stokes’s theorem, we obtain <br />

S u∗ dλ = 0, implying that<br />

π ◦ Tu≡ 0.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 33<br />

Hence a is a harmonic function on S and therefore constant. So far, we also know that<br />

the image of u lies on a Reeb trajectory, and it remains to show that this trajectory is<br />

actually periodic.<br />

Let τ : ˜S → S be the universal covering map. The complex structure j lifts to a<br />

complex structure ˜j on ˜S. Now pick smooth functions f, g on ˜S such that<br />

dg = τ ∗ γ =: ˜γ, −df = τ ∗ (γ ◦ j) = ˜γ ◦ ˜j.<br />

Then the map u ◦ τ : ˜S → M satisfies<br />

(u ◦ τ) ∗ λ = df.<br />

Theimageofu◦ τ lies on a trajectory x of the Reeb vector field in view of<br />

<br />

D(u ◦ τ)(z)ζ = Df (z)ζ · Xλ (u ◦ τ)(z) ,<br />

hence (u ◦ τ)(z) = x(h(z)) for some smooth function h on ˜S, and it follows that, after<br />

maybe adding a constant to f ,wehave<br />

(u ◦ τ)(z) = x f (z) .<br />

The function f does not descend to S. If it did, it would have to be constant since it<br />

is harmonic. On the other hand, this would imply that u is constant in contradiction to<br />

our assumption that it is not. Therefore, there is a point q ∈ S and two lifts z0,z1 ∈ ˜S<br />

such that f (z0) >f(z1).Letℓ : S 1 → S be a loop which lifts to a path α :[0, 1] → ˜S<br />

with α(0) = z0 and α(1) = z1. Considering the map<br />

v := u ◦ ℓ : S 1 −→ M,<br />

we see that v(t) = (u ◦ τ ◦ α)(t) = x f (α(t)) and x(f (z0)) = x(f (z1)), that is, the<br />

trajectory x is a periodic orbit. Hence the image of u is a periodic orbit for the Reeb<br />

vector field. <br />

The following is the main result of this paper.<br />

<strong>THEOREM</strong> 1.6<br />

Let M be a closed 3-dimensional manifold, and let λ ′ be a contact form on M. Then<br />

the following holds for a suitable contact form λ = fλ ′ ,wheref is a positive function<br />

on M. There exists a smooth family (S,jτ ,Ɣτ , ũτ = (aτ ,uτ ),γτ )τ∈S 1 of solutions to<br />

(1.1) for a suitable compatible complex structure J :kerλ → ker λ such that<br />

• all maps uτ have the same asymptotic limit K at the punctures, where K is a<br />

finite union of periodic trajectories of the Reeb vector field Xλ;


34 CASIM ABBAS<br />

• uτ ( ˙S) ∩ uτ ′( ˙S) =∅ if τ = τ ′ ;<br />

• M\K = <br />

τ∈S 1 uτ ( ˙S);<br />

• the projection P onto S 1 defined by p ∈ uτ ( ˙S) ↦→ τ is a fibration;<br />

• the open book decomposition given by (P,K) supports the contact structure<br />

ker λ, and λ is a Giroux form.<br />

Here is a very brief outline of the argument. The reader is invited to skip forward to<br />

Section 4 to see in more detail how all the partial results of this paper are tied together<br />

to prove the main result. In Section 2, we find a Giroux contact form which has a<br />

certain normal form near the binding. Following an argument by Wendl ([39], [38]),<br />

we will then almost be able to turn the Giroux leaves into solutions of (1.1) without<br />

harmonic form except for the fact that we have to accept a confoliation form instead<br />

of a contact form. Pick one of these Giroux leaves as a starting point. The next step is<br />

to prove a result which permits us to perturb the Giroux leaf into a genuine solution<br />

of (1.1) while simultaneously perturbing the confoliation form slightly into a contact<br />

form. This is where the harmonic form in (1.1) is required. We actually obtain a local<br />

family of nearby solutions, not just one. In Section 3, we prove a compactness result<br />

which extends the local family of solutions into a global one. The remarkable fact is<br />

that there is a compactness result in the context of this paper, although there is none<br />

in general for the perturbed holomorphic curve equation. The special circumstances<br />

in this paper imply a crucial a priori bound which implies that a sequence of solutions<br />

has a pointwise convergent subsequence with a measurable limit. The objective is then<br />

to show that the regularity of this limit is much better, that it is actually smooth.<br />

We consider two solutions (S,j,Ɣ,ũ, γ ) and (S ′ ,j ′ ,Ɣ ′ , ũ ′ ,γ ′ ) equivalent if there<br />

exists a biholomorphic map φ :(S,j) → (S ′ ,j ′ ) mapping Ɣ to Ɣ ′ (preserving the<br />

enumeration) so that ũ ′ ◦ φ = ũ. We will often identify a solution (S,j,Ɣ,ũ, γ ) of<br />

(1.1) with its equivalence class [S,j,Ɣ,ũ, γ ].WenotethatwehaveanaturalR-action<br />

on the solution set by associating to c ∈ R and [S,j,Ɣ,ũ, γ ] the new solution<br />

c + [S,j,Ɣ,ũ, γ ] = [S,j,Ɣ,(a + c, u),γ], ũ = (a,u).<br />

A crucial concept for our discussion is the notion of a finite energy foliation F .<br />

Definition 1.7 (Finite energy foliation)<br />

A foliation F of R × M is called a finite energy foliation if every leaf F is the image<br />

of an embedded solution [S,j,Ɣ,ũ, γ ] of the equations (1.1), that is,<br />

F = ũ( ˙S),


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 35<br />

so that u( ˙S) ⊂ M is transverse to the Reeb vector field, and for every leaf F ∈ F we<br />

also have c + F ∈ F for every c ∈ R; that is, the foliation is R-invariant.<br />

We recall the concept of a global surface of section. Let M be a closed 3-manifold,<br />

and let X be a nowhere-vanishing smooth vector field.<br />

Definition 1.8 (Surface of section)<br />

(a) A local surface of section for (M,X) consists of an embedded compact surface<br />

⊂ M with boundary, so that ∂ consists of a finite union of periodic orbits<br />

(called the binding orbits). In addition, the interior ˙ = \ ∂ is transverse<br />

to the flow.<br />

(b) A local surface of section is called a global surface of section if, in addition,<br />

every orbit other than a binding orbit hits ˙ in forward and backward time.<br />

Furthermore, the globally defined return map : ˙ → ˙ has a bounded<br />

return time; that is, there exists a constant c>0 so that every x ∈ ˙ hits ˙<br />

again in forward time not exceeding c.<br />

Using Proposition 2.5 below, the existence part of Giroux’s theorem can be rephrased<br />

as follows.<br />

<strong>THEOREM</strong> 1.9<br />

Let M be a closed orientable 3-manifold, and let ˜λ be a contact form on M. Then<br />

there exists a smooth function f : M → (0, ∞) so that the contact form λ = f ˜λ has<br />

a Reeb vector field admitting a global surface of section.<br />

Existence results for finite energy foliations with a given contact form λ are hard to<br />

come by since they usually have striking consequences. In [20] for example, Hofer,<br />

Wysocki, and Zehnder show that every compact strictly convex energy hypersurface<br />

S in R 4 carries either two or infinitely many closed characteristics. The proof relies on<br />

constructing a special finite energy foliation. In special cases they were established by<br />

Hofer, Wysocki, and Zehnder [22] and by Wendl ([38], [40]). Proofs usually require<br />

a starting point, that is, a finite energy foliation for a slightly different situation as the<br />

given one. Then some kind of continuation argument is employed where all kinds of<br />

things can and do happen to the original foliation. In [22], the authors start with an<br />

explicit finite energy foliation for the round 3-dimensional sphere S 3 ⊂ R 4 which is<br />

then deformed. Wendl’s papers also use a rather special manifold as a starting point.<br />

The main result of this paper, Theorem 1.6, provides a starting finite energy foliation<br />

for any closed 3-dimensional contact manifold (M,ker λ) since it is obtained from<br />

deforming the leaves of Giroux’s open book decomposition. The pages are usually<br />

not punctured spheres, and generically there are no pseudoholomorphic curves on


36 CASIM ABBAS<br />

punctured surfaces with genus which are transverse to the Reeb vector field. This<br />

makes the introduction of the harmonic form in (1.1) a necessity. The price to be paid<br />

is that compactness issues are more complicated.<br />

Wendl [39] published a proof of Theorem 1.6 for the special case where ker λ is<br />

a planar contact structure, that is, where the surfaces ˙S are punctured spheres. This<br />

result was outlined in [4]. Regardless of whether the contact structure is planar or not,<br />

there are two main steps in the proof: existence of a solution and compactness of a<br />

family of solutions. While the author established the compactness part for Theorem<br />

1.6 long before [4] appeared, we will use the same argument described by Wendl in<br />

[39] for the existence part since it simplifies the proof considerably.<br />

The main theorem of this article was the first step in the proof of the Weinstein<br />

conjecture for the planar case in [4]. Recall that the Weinstein conjecture [38, p. 358]<br />

states the following: Every Reeb vector field X on a closed contact manifold M admits<br />

a periodic orbit.<br />

In fact, Weinstein added the additional hypothesis that the first cohomology group<br />

H 1 (M,R) with real coefficients vanishes, but there seems to be no indication that this<br />

additional hypothesis is needed.<br />

Moreover, Theorem 1.6 is also the starting point for the construction of global<br />

surfaces of section in the forthcoming paper [7]. Another application will be an<br />

alternative proof of the Weinstein conjecture in dimension 3 (see [7]) as outlined in<br />

[4]. This complements Taubes’s recent proof of the Weinstein conjecture in dimension<br />

3 using a perturbed version of the Seiberg-Witten equations (see [33], [34]). The<br />

main issue with the homological perturbed holomorphic curve equation (1.1) isthat<br />

there is no natural compactification of the space of solutions unless the harmonic<br />

forms are uniformly bounded. In the forthcoming papers [6] and[7] the lack of compactness<br />

is investigated, and bounds for the harmonic forms are derived in particular<br />

cases.<br />

2. Existence and local foliations<br />

2.1. Local model near the binding orbits<br />

We use the same approach as in [39] and[38] to prove existence of a solution to<br />

(1.1). Given a closed contact 3-manifold (M,ξ), Giroux’s theorem implies that there<br />

is an open book decomposition as in Theorem 1.2 supporting ξ. On the other hand,<br />

any other contact structure ξ ′ supported by the same open book is diffeomorphic to<br />

ξ. Starting with an open book decomposition for M, we construct a contact structure<br />

supported by Giroux contact form λ which has a certain normal form near the<br />

binding.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 37<br />

Definition 2.1<br />

Let θ ∈ S 1 = R/2πZ denote polar coordinates on the unit disk D ⊂ R 2 by (r, φ),<br />

and let γ1,γ2 :[0, +∞) → R be smooth functions. A 1-form<br />

λ = γ1(r) dθ + γ2(r) dφ<br />

is called a local model near the binding if the following conditions are satisfied:<br />

(1) the functions γ1,γ2,andγ2(r)/r 2 are smooth if considered as functions on the<br />

disk D (in particular, γ ′ 1 (0) = γ ′ 2 (0) = γ2(0) = 0);<br />

(2) µ(r) := γ1(r)γ ′ 2 (r) − γ ′ 1 (r)γ2(r) > 0 if r>0;<br />

(3) γ1(0) > 0 and γ ′ 1 (r) < 0 if r>0;<br />

(4) limr→0(µ(r)/r) = γ1(0)γ ′′<br />

2 (0) > 0;<br />

(5) κ := (γ ′′ ′′<br />

1 (0)/γ 2 (0)) /∈ Z and κ ≤−1/2;<br />

(6) A(r) = (1/µ 2 (r))(γ ′′<br />

2 (r)γ ′ ′′<br />

1 (r) − γ 1 (r)γ ′ 2<br />

We explain some of the conditions above. First, since<br />

(r)) is of order r for small r>0.<br />

λ ∧ dλ = µ(r)dθ ∧ dr ∧ dφ = µ(r)<br />

dθ ∧ dx ∧ dy,<br />

r<br />

the form λ is a contact form on S 1 × D. The Reeb vector field is given by<br />

X(θ,r,φ) = γ ′ 2 (r)<br />

µ(r)<br />

∂<br />

∂θ − γ ′ 1 (r)<br />

µ(r)<br />

∂<br />

∂φ<br />

The trajectories of X all lie on tori Tr = S 1 × ∂Dr:<br />

We compute<br />

and<br />

=: α(r) ∂<br />

∂θ<br />

+ β(r) ∂<br />

∂φ .<br />

θ(t) = θ0 + α(r) t, φ(t) = φ0 + β(r) t. (2.1)<br />

γ<br />

lim α(r) = lim<br />

r→0 r→0<br />

′′<br />

2 (r)<br />

µ ′ (r)<br />

γ<br />

lim β(r) =−lim<br />

r→0 r→0<br />

′′<br />

1 (r)<br />

µ ′ (r)<br />

= γ ′′<br />

2 (0)<br />

Recalling that ∂ ∂ ∂<br />

= x − y , we obtain for r = 0<br />

∂φ ∂y ∂x<br />

X = 1 ∂<br />

γ1(0) ∂θ ,<br />

γ1(0)γ ′′<br />

1<br />

=<br />

(0) γ1(0)<br />

2<br />

=− γ ′′<br />

1 (0)<br />

γ1(0)γ ′′<br />

2 (0).


38 CASIM ABBAS<br />

that is, the central orbit has minimal period 2πγ1(0). If the ratio α(r)/β(r) is irrational<br />

then the torus Tr carries no periodic trajectories. Otherwise, Tr is foliated with periodic<br />

trajectories of minimal period<br />

τ = 2πm<br />

α<br />

= 2πn<br />

β ,<br />

where α/β = m/n or β/α = n/m for suitable integers m, n (choose whatever makes<br />

sense if either α or β is zero). We calculate<br />

dα<br />

lim<br />

r→0 dr<br />

γ<br />

= lim<br />

r→0<br />

′′<br />

2 (r)µ(r) − γ ′ 2 (r)µ′ (r)<br />

µ 2 (r)<br />

γ<br />

= lim<br />

r→0<br />

′′′<br />

2 (r)<br />

2µ ′ − lim<br />

(r) r→0<br />

= γ ′′′<br />

= 0<br />

since µ ′′ (0) = γ1(0)γ ′′′<br />

coordinates on the disk, we get<br />

2 (0)<br />

2µ ′ (0)<br />

µ ′′ (r)<br />

2µ ′ γ<br />

(r)<br />

′ 2 (r)<br />

r<br />

′′′ ′′<br />

γ1(0)γ 2 (0)γ 2<br />

− (0)<br />

2(µ ′ (0)) 2<br />

2 (0) and µ′ (0) = γ1(0)γ ′′<br />

2<br />

X(θ,x,y) = α(x,y) ∂<br />

∂θ<br />

− β(x,y) y ∂<br />

∂x<br />

r<br />

µ(r)<br />

(0) > 0. Converting to Cartesian<br />

+ β(x,y) x ∂<br />

∂y ,<br />

and linearizing the Reeb vector field along the center orbit yields<br />

⎛<br />

0 0 0<br />

⎞<br />

DX(θ,0, 0) = ⎝ 0 0 −β(0) ⎠ .<br />

0 β(0) 0<br />

The linearization of the Reeb flow is given by<br />

⎛<br />

1 0 0<br />

⎞<br />

Dφt(θ,0, 0) = ⎝ 0 cos β(0)t − sin β(0)t ⎠ (2.2)<br />

0 sin β(0)t cos β(0)t<br />

with<br />

The spectrum of (t) is given by<br />

(t) = e β(0)tJ , J =<br />

σ (t) ={e ±iβ(0)t }.<br />

<br />

0 −1<br />

.<br />

1 0


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 39<br />

The binding orbit has period 2πγ1(0), since<br />

and it is nondegenerate and elliptic.<br />

′′ γ 1<br />

γ1(0)β(0) =− (0)<br />

γ ′′<br />

2<br />

(0) /∈ Z,<br />

Example 2.2<br />

For the contact form Tdθ+(1/k)(xdy−ydx) = Tdθ+(r 2 /k)dφ the central orbit<br />

S 1 ×{0} is degenerate, but<br />

is a local model near the binding if<br />

In this case,<br />

and we note that<br />

If<br />

λ = (1 − r 2 )(T dθ+ r2<br />

k dφ)<br />

k, T > 0, kT /∈ Z, and kT ≥ 1<br />

2 .<br />

µ(r) = 2rT<br />

k (1 − r2 ) 2 > 0 and<br />

A(r) = 1<br />

µ 2 (r)<br />

γ ′′<br />

2<br />

(r)γ ′<br />

1<br />

α(r)<br />

β(r) =−γ ′ 2 (r)<br />

γ ′ 1<br />

′′ ′<br />

(r) − γ 1 (r)γ 2 (r) =<br />

1 − 2r2<br />

=<br />

(r) kT<br />

γ ′′<br />

1 (0)<br />

γ ′′<br />

2<br />

m<br />

=<br />

n<br />

(0) =−kT,<br />

4kr<br />

T (1 − r2 .<br />

) 4<br />

for integers n, m, then the invariant torus Tr is foliated with periodic orbits. The<br />

case m = 0 is only possible if r = 1/ √ 2.Ifr is sufficiently small, then |m| ≥2.<br />

Indeed, we would otherwise be able to find sequences rl ↘ 0 and {nl} ⊂Z such that<br />

kT/(1 − 2r 2 l ) = nl, which is impossible. The binding orbit has period 2πT while<br />

the periodic orbits close to the binding orbit have much larger periods equal to<br />

τ = 2πTm (1 − r2 ) 2<br />

.<br />

1 − 2r2 Example 2.3<br />

Consider the contact form λ = T (1 − r 2 )dθ + (r 2 /k)dφ on S 1 × D. Itisalsoa<br />

local model near the binding if k, T > 0, kT ≥ 1/2, andkT is not an integer. We


40 CASIM ABBAS<br />

even have A(r) ≡ 0. In contrast to Example 2.2, if kT /∈ Q, then the invariant tori Tr<br />

carry no periodic orbits. If kT = n/m ∈ Q but not in Z, all invariant tori are foliated<br />

with periodic orbits of period 2πmT with |m| ≥2 while the binding orbit has period<br />

2πT. The function A(r) is identically zero. This is the contact form on the irrational<br />

ellipsoid in R 4 .<br />

The following proposition is essentially due to Wendl. The construction in the proof<br />

was used by Thurston and Winkelnkemper [35] to show existence of contact forms on<br />

closed 3-manifolds.<br />

PROPOSITION 2.4 ([39, Proposition 1])<br />

Let M be a 3-dimensional manifold given by an open book decomposition<br />

M = W (h) ∪Id (∂W × D 2 )<br />

as described in Theorem 1.2. We denote the pages by<br />

Fα := (W ×{α}) ∪Id (∂W × Iα), 0 ≤ α


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 41<br />

([0,ε] × ( ˙<br />

nS1 ), {0} ×( ˙<br />

nS1 )), where we take an n-fold disjoint union of circles<br />

S1 ≈ R/2πZ according to the number n of components of ∂W.<br />

We<br />

<br />

claim that there is an area form on W that satisfies<br />

• = 2πn,<br />

W<br />

• |C = dt ∧ dθ.<br />

Indeed, start with any area form ′ so that <br />

W ′ = 2πn.Thenwehave ′ |C =<br />

f ′ (t,θ)dt∧dθ with a positive smooth function f ′ (after switching signs if necessary).<br />

Now pick a new smooth positive function f which is equal to some constant c if<br />

t ≤ (1/3)ε and which agrees with f ′ if t ≥ (2/3)ε so that the resulting area form<br />

still satisfies <br />

= 2πn. Do one component of ∂W at a time. Rescaling the<br />

W<br />

t-coordinate, we may assume that c = 1.<br />

Let α1 be any 1-form on W which equals (1 + t) dθ near ∂W. Then by Stokes’s<br />

theorem we obtain<br />

<br />

<br />

<br />

( − dα1) = 2πn− α1 = 2πn+ dθ = 0.<br />

W<br />

∂W<br />

∂W<br />

The 2-form −dα1 on W is closed and vanishes near ∂W. Then there exists a 1-form<br />

β on W with<br />

dβ = − dα1<br />

and β ≡ 0 near ∂W.Nowdefineα2 := α1 + β.Thenα2 satisfies the following:<br />

• dα2 is an area form on W inducing the same orientation as , (2.3)<br />

• α2 = (1 + t) dθ near ∂W. (2.4)<br />

The set of 1-forms on W satisfying (2.3) and (2.4) is therefore nonempty and also<br />

convex. We define the following 1-form on W × [0, 2π], where α is any 1-form on<br />

W satisfying (2.3) and (2.4):<br />

˜α(x,τ):= τα(x) + (2π − τ)(h ∗ α)(x).<br />

This 1-form descends to the quotient W (h), and the restriction to each fiber of the<br />

fiber bundle W (h) π → S 1 satisfies condition (2.3). Moreover, since h ≡ Id near ∂W,<br />

we have ˜α(x,τ) = 2π(1 + t) dθ for all (x,τ) = ((t,θ),τ) near ∂W(h) = ∂W × S 1 .<br />

Let dτ be a volume form on S 1 . We claim that<br />

λ1 := −δ ˜α + π ∗ dτ


42 CASIM ABBAS<br />

are contact forms on W (h) whenever δ>0 is sufficiently small. Pick (x,τ) ∈ W(h),<br />

and let {u, v, w} be a basis of T(x,τ)W (h) with π∗u = π∗v = 0. Then<br />

(λ1 ∧ dλ1)(x,τ)(u, v, w)<br />

= δ 2 (˜α ∧ d ˜α)(x,τ)(u, v, w) − δ [dτ(π∗w) d ˜α(x,τ)(u, v)]<br />

= 0<br />

for sufficiently small δ>0,anddλ1 is a volume form on W . Now we have to continue<br />

the contact forms λ1 beyond ∂W(h) ≈ ∂W × S1 onto ∂W × D2 . At this point it is<br />

convenient to change coordinates. We identify C × S1 with ∂W × (D2 1+ε \D2 1 ), where<br />

D2 ρ is the 2-disk of radius ρ. Using polar coordinates (r, φ) on D2 1+ε with 0 ≤ φ ≤ 2π<br />

and 0 0 and γ ′<br />

1 (r) < 0 if r>0,<br />

hence the curves γ = γδ have to turn counterclockwise in the first quadrant starting<br />

at the point (γ1(0), 0) and later connecting with (−δ(1 − ε0), 1). In the case where<br />

δ>0, the Reeb vector fields are given by<br />

Xδ(θ,r,φ) = γ ′ 2 (r)<br />

µ(r)<br />

∂<br />

∂θ − γ ′ 1 (r)<br />

µ(r)<br />

∂<br />

∂φ ,


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 43<br />

and, in particular,<br />

Xδ(θ,r,φ) = ∂<br />

∂φ for r ≥ 1 − ε0, (2.5)<br />

which implies that the Reeb vector fields Xδ converge as δ ↘ 0. In addition to λδ being<br />

contact forms for δ>0, we also want the given open book decomposition to support<br />

ker λδ, hence Xδ needs to be transverse to the pages of the open book decomposition<br />

which is equivalent to γ ′ 1 (r) = 0. A curve γ (r) fulfilling these conditions can clearly<br />

be constructed. <br />

The following result shows that we can always assume that a Giroux contact form is<br />

equal to any of the forms provided by Proposition 2.4.<br />

PROPOSITION 2.5<br />

Let M be a closed 3-dimensional manifold with contact structure ξ. Then, for every<br />

δ>0, there is a diffeomorphism ϕδ : M → M such that ker λδ = ϕ∗ξ where λδ is<br />

given by Proposition 2.4.<br />

Proof<br />

Existence of an open book decomposition supporting ξ follows from the existence<br />

part of Giroux’s theorem. On the other hand, Proposition 2.4 yields contact forms λδ<br />

such that ker λδ is also supported by the same open book decomposition as ξ for any<br />

δ>0. By the uniqueness part of Giroux’s theorem, ξ and ker λδ are diffeomorphic. <br />

It follows from our previous construction of the forms λδ that λ0 satisfies λ0 ∧dλ0 > 0<br />

on ∂W ×D1−ε0 and that λ0 = dφ otherwise. For δ → 0, the Reeb vector fields Xδ will<br />

converge to some vector field X0, which is the Reeb vector field of λ0 if r


44 CASIM ABBAS<br />

the standard complex structure on the cylinder [0, +∞) × S 1 after introducing polar<br />

coordinates near the punctures. Moreover, all the punctures are positive ∗ the family<br />

(ũα)0≤α≤2π is a finite energy foliation, and the curves ũα are ˜Jδ-holomorphic near the<br />

punctures.<br />

Proof<br />

We parameterize the leaves of the open book decomposition uα : ˙S → M, 0 ≤ α<<br />

2π, and we assume that they look as follows near the binding:<br />

uα :[0, +∞) × S 1 −→ S 1 × D1,<br />

uα(s, t) = t,r(s)e iα ,<br />

(2.6)<br />

where r are smooth functions with lims→∞ r(s) = 0 to be determined shortly. We<br />

use the notation (r, φ) for polar coordinates on the disk D = D1. We identify some<br />

neighborhood U of the punctures of ˙S with a finite disjoint union of half-cylinders<br />

[0, +∞) × S1 . Recall that the binding orbit is given by<br />

<br />

t<br />

<br />

x(t) = , 0, 0 , 0 ≤ t ≤ 2πγ1(0)<br />

γ1(0)<br />

and that it has minimal period T = 2πγ1(0). We define smooth functions aα : ˙S → R<br />

by<br />

s<br />

aα(z) := 0 γ1<br />

<br />

′ ′ r(s ) ds if z = (s, t) ∈ [0, +∞) × S1 ⊂ U,<br />

0 if z/∈ U<br />

so that<br />

u ∗<br />

α λ0 ◦ j = daα,<br />

where j is a complex structure on ˙S which equals the standard structure i on [0, +∞)×<br />

S 1 (i.e., near the punctures). We want to turn the maps ũα = (aα,uα) : ˙S → R × M<br />

into ˜J0-holomorphic curves for a suitable almost-complex structure ˜J0 on R × M.<br />

Recall that the contact structure is given by<br />

<br />

∂ ∂<br />

ker λδ = Span{η1,η2} = Span , −γ2(r)<br />

∂r ∂θ<br />

We define complex structures Jδ :kerλδ → ker λδ by<br />

<br />

Jδ(θ,r,φ) − γ2(r) ∂<br />

∂θ<br />

∂<br />

<br />

+ γ1(r) := −<br />

∂φ<br />

1<br />

h(r)<br />

∗ A puncture pj is called positive for the curve (aα,uα) if limz→pj aα(z) =+∞.<br />

∂<br />

<br />

+ γ1(r) .<br />

∂φ<br />

∂<br />

∂r<br />

(2.7)


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 45<br />

and<br />

Jδ(θ,r,φ) ∂<br />

∂r<br />

<br />

:= h(r) − γ2(r) ∂<br />

∂θ<br />

∂<br />

<br />

+ γ1(r) ,<br />

∂φ<br />

where h :(0, 1] → R\{0} are suitable smooth functions. Also recall that γ1,γ2<br />

depend on δ away from the binding orbit. We want Jδ to be compatible with dλδ,that<br />

is, we want<br />

dλδ(η1,Jη1) = h(r) µ(r) > 0 and dλδ(η2,Jη2) = µ(r)<br />

> 0<br />

h(r)<br />

so that h(r) > 0. We also demand that Jδ extends smoothly over the binding {r = 0}.<br />

Expressing the vectors η1 and η2 in Cartesian coordinates, we have<br />

and<br />

η1 = 1<br />

r<br />

<br />

x ∂ ∂<br />

+ y<br />

∂x ∂y<br />

η2 =−γ2(r) ∂<br />

∂θ + γ1(r) x ∂<br />

∂y − γ1(r) y ∂<br />

∂x .<br />

We introduce the following generators of the contact structure:<br />

and<br />

We compute from this<br />

Now<br />

ε1 := γ1(r) ∂<br />

∂y<br />

= yγ1(r)<br />

r<br />

ε2 := γ1(r) ∂<br />

∂x<br />

= xγ1(r)<br />

r<br />

− xγ2(r)<br />

r 2<br />

η1 + x<br />

η2<br />

r2 <br />

∂<br />

∂θ<br />

yγ2(r)<br />

+<br />

r2 ∂<br />

∂θ<br />

η1 − y<br />

η2.<br />

r2 η1 = 1<br />

rγ1(r) (yε1 + xε2), η2 = xε1 − yε2.<br />

Jδε1 = yγ1(r)h(r)<br />

η2 −<br />

r<br />

x<br />

r2h(r) η1<br />

<br />

1<br />

=<br />

r xyγ1(r)h(r)<br />

xy<br />

−<br />

r3 <br />

ε1<br />

h(r)γ1(r)<br />

<br />

1<br />

−<br />

r y2 x<br />

γ1(r)h(r) +<br />

2<br />

r3 <br />

γ1(r)h(r)<br />

ε2


46 CASIM ABBAS<br />

and<br />

Jδε2 = xγ1(r)h(r)<br />

η2 +<br />

r<br />

y<br />

r2h(r) η1<br />

<br />

= − 1<br />

r xyγ1(r)h(r)<br />

xy<br />

+<br />

r3 <br />

ε2<br />

h(r)γ1(r)<br />

<br />

1<br />

+<br />

r x2 y<br />

γ1(r)h(r) +<br />

2<br />

r3 <br />

ε1.<br />

γ1(r)h(r)<br />

Inserting x = r cos φ, y = r sin φ, and demanding that the limit is φ-independent as<br />

r → 0, we arrive at the condition that rh(r) γ1(r) ≡±1 for small r. Recalling that<br />

we need h>0, we obtain<br />

h(r) = 1<br />

rγ1(r)<br />

for small r.<br />

As usual, we continue Jδ to an almost-complex structure ˜Jδ on R × M by setting<br />

˜Jδ(θ,r,φ) ∂<br />

∂τ<br />

:= Xδ(θ,r,φ),<br />

where τ denotes the coordinate in the R-direction. We emphasize that ˜Jδ also makes<br />

sense for δ = 0. We now arrange r(s) in (2.6) such that the Giroux leaves ũα = (aα,uα)<br />

become ˜J0-holomorphic curves. ∗ We compute for r ≤ 1 − ε0<br />

∂sũα + ˜J0(uα)∂tũα = γ1(r) ∂ ∂<br />

+ r′<br />

∂τ ∂r + <br />

˜J0(uα)<br />

∂<br />

<br />

∂θ<br />

= γ1(r) ∂ ∂<br />

+ r′<br />

∂τ ∂r + ˜J0(uα) γ1(r) Xδ(uα) <br />

<br />

+ ˜J0(uα)<br />

∂<br />

<br />

− γ1(r)Xδ(uα)<br />

∂θ<br />

′ ∂<br />

= r<br />

∂r + ′<br />

˜J0(uα)<br />

γ 1 (r)<br />

<br />

γ1(r)<br />

µ(r)<br />

∂ ∂<br />

<br />

− γ2(r)<br />

∂φ ∂θ<br />

<br />

= r ′ − γ ′ 1 (r)<br />

<br />

∂<br />

µ(r)h(r) ∂r ,<br />

hence the Giroux leaves satisfy the equation if we choose r to be a solution of the<br />

ordinary differential equation<br />

r ′ (s) =<br />

γ ′ 1 (r(s))<br />

µ(r(s)) h(r(s)) .<br />

∗ The calculation shows that we can make them Jδ-holomorphic for all δ ≥ 0 near the binding.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 47<br />

Note that r ′ (s) < 0. We also choose h(r) ≡ 1 for r ≥ 1 − ε0. We continue the<br />

almost-complex structures Jδ :kerλδ → ker λδ (which were only defined near the<br />

binding) smoothly to all of M. Away from the binding we have Xδ = ∂/∂φ, andwe<br />

extend Jδ as before to T (R × M). Away from the binding, if δ = 0, wehavethat<br />

ker λ0 coincides with the tangent spaces of the pages of the open book decomposition.<br />

Because aα is constant away from the binding, the solutions ũα which we constructed<br />

near the binding fit together smoothly with the pages of the open book decomposition<br />

and solve the holomorphic curve equation for the almost-complex structure J0. <br />

Remark 2.7<br />

Near the binding orbit the function r(s) satisfies a differential equation of the form<br />

and<br />

r ′ (s) = r(s) r(s) := γ ′ 1 (r(s))γ1(r(s))<br />

µ(r(s))<br />

lim<br />

r→0<br />

(r) = γ ′′<br />

1 (0)<br />

γ ′′<br />

2<br />

(0) =: κ.<br />

r(s),<br />

Writing r(s) = c(s)e κs , the function c(s) satisfies c ′ (s) = ((r(s)) − κ)c(s), and<br />

hence it is a decreasing function which converges to a constant as s →+∞.<br />

We return to Examples 2.2 and 2.3, and we compute r(s) for large s. The differential<br />

equation in the case of Example 2.3 for large s is<br />

so that<br />

r ′ (s) = γ ′ 1 (r(s))γ1(r(s))<br />

r(s) =−kT<br />

µ(r(s))<br />

1 − r 2 (s) r(s),<br />

r(s) =<br />

1<br />

√ 1 + ce 2kT s ,<br />

where c is a constant. In Example 2.2, the differential equation reads<br />

and solutions satisfy<br />

r ′ (s) = γ ′ 1 (r(s))γ1(r(s))<br />

r(s) =−<br />

µ(r(s))<br />

kT<br />

1 − r2 (s) r(s),<br />

r(s) = ce −kT s e (1/2)r2 (s) .<br />

2.2. Functional analytic setup and the implicit function theorem<br />

In the following theorem we prove the existence of a smooth family of solutions near a<br />

given solution. In Proposition 2.6, we constructed a finite energy foliation for the data


48 CASIM ABBAS<br />

(λ0,J0) with vanishing harmonic form. The form λ0, however, is only a confoliation<br />

form. We produce solutions for the perturbed data (λδ,Jδ), and harmonic forms appear<br />

if the surface S is not a sphere. The key result is an application of the implicit function<br />

theorem in a suitable setting.<br />

<strong>THEOREM</strong> 2.8<br />

Assume one of the following.<br />

(1) Let (a0,u0) : ˙S → R × M be one of the ˜J0-holomorphic curves described in<br />

Proposition 2.6 with complex structure j0 on S (we refer to such u0 as a Giroux<br />

leaf) and confoliation form λ0.<br />

(2) Let ( ˙S,j0,a0,u0,γ0) be a solution of the differential equation (1.1) for some<br />

dλ0-compatible complex structure J0 :kerλ0 → ker λ0 which, near the binding<br />

orbit, agrees with (2.7), and where λ0 is a contact form which is a local<br />

model near the binding. Assume that u0 is an embedding and that it is of the<br />

form u0 = φg(v0), whereg : S → R is a smooth function, φ is the flow of the<br />

Reeb vector field, and where v0 : ˙S → M is a Giroux leaf as in Proposition<br />

2.6.<br />

(3) Let Jδ be a smooth family of dλδ-compatible complex structures also agreeing<br />

with (2.7) near the binding orbit, where (λδ)−ε


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 49<br />

Note that this is well defined because u is transverse to the Reeb vector field so that<br />

πTu(z) :TzS → ker λ(u(z)) is an isomorphism. By the second equation of (1.1), we<br />

then have to solve the equation df ◦ jf + u ∗ 0 λ ◦ jf = da + γ for a,f,γ on ˙S which<br />

is equivalent to the equation<br />

¯∂jf (a + if ) = u ∗<br />

0 λ ◦ jf − i(u ∗<br />

0 λ) − γ − i(γ ◦ jf ). (2.9)<br />

Recall that we are looking for a of the form a = a0 + b, where b is a suitable realvalued<br />

function defined on the whole surface S. We obtain the differential equation<br />

¯∂jf (b + if ) = u ∗<br />

0 λ ◦ jf − i(u ∗<br />

0 λ) − ¯∂jf a0 − γ − i(γ ◦ jf ), (2.10)<br />

and it follows from a straightforward calculation (see the appendix) that all expressions<br />

on the right-hand side of (2.10) are bounded near the punctures; in particular, they<br />

are contained in the spaces L p (T ∗ S ⊗ C) for any p. This is what the assumption<br />

κ ≤−(1/2) from Definition 2.1 is needed for. We work in the function space b +if ∈<br />

W 1,p (S,C), where p>2. For any complex structure j on S, the space L p (T ∗ S ⊗ C)<br />

of complex-valued 1-forms of class L p decomposes into complex linear and complex<br />

antilinear forms (with respect to j). We use the notation<br />

L p (T ∗ S ⊗ C) = L p (T ∗ S ⊗ C) 1,0<br />

j ⊕ Lp (T ∗ S ⊗ C) 0,1<br />

j .<br />

The operator b + if ↦→ ¯∂jf (b + if ) is then a section in the vector bundle<br />

L p (T ∗ S ⊗ C) 0,1 :=<br />

<br />

b+if ∈W 1,p (S,C)<br />

{b + if }×L p (T ∗ S ⊗ C) 0,1<br />

jf → W 1,p (S,C).<br />

This vector bundle is of course trivial, but here are some explicit local trivializations<br />

for f, g ∈ W 1,p (S,R) sufficiently close to each other:<br />

If we write<br />

we see that fg is invertible with<br />

fg : Lp (T ∗S ⊗ C) 0,1<br />

jf −→Lp (T ∗S ⊗ C) 0,1<br />

jg<br />

τ ↦−→ τ + i(τ ◦ jg).<br />

τ + i(τ ◦ jg) = τ ◦ (IdTS − jf ◦ jg),<br />

−1<br />

fg τ = τ ◦ (IdTS − jf ◦ jg) −1 .<br />

(2.11)<br />

It follows from the Hodge decomposition theorem that every cohomology class [σ ] ∈<br />

H 1 (S,R) has a unique harmonic representative ψj(σ ) ∈ H 1 j (S), where H 1 j (S) is


50 CASIM ABBAS<br />

defined as<br />

H 1<br />

j (S) := γ ∈ E 1 (S) dγ = 0, d(γ ◦ j) = 0 <br />

(2.12)<br />

and where E 1 (S) denotes the space of all (smooth) real-valued 1-forms on S, and<br />

we write E 0,1 (S) = E 0,1<br />

j (S) for the space of complex antilinear 1-forms on S with<br />

respect to j, that is, complex-valued 1-forms σ such that iσ + σj = 0. Note that<br />

our definition coincides with the set of closed and co-closed 1-forms on S. Moreover,<br />

by elliptic regularity, we may also consider Sobolev forms. We identify H 1 (S,R)<br />

with R 2g , and we consider the following parameter-dependent section in the bundle<br />

L p (T ∗ S ⊗ C) 0,1 → W 1,p (S,C)<br />

F : W 1,p (S,C) × R 2g −→ L p (T ∗ S ⊗ C) 0,1<br />

F (b + if, σ ):= ¯∂jf (b + if ) − u ∗<br />

0λ ◦ jf + i(u ∗<br />

0λ) + ¯∂jf a0 + ψjf (σ ) + i <br />

ψjf (σ ) ◦ jf<br />

(2.13)<br />

with jf as in (2.8). Recalling that z ↦→ jf (z) may not be differentiable, we interpret<br />

the equation d(γ ◦ jf ) = 0 in the sense of weak derivatives. The solution set of (2.10)<br />

is then the zero set of F . We consider the real parameter δ which we dropped from<br />

the notation, fixed at the moment. For g ≡ 0 and b + if small in the W 1,p -norm,<br />

we consider the composition ˆF (b + if, σ ) = fg(F (b + if, σ )). Its linearization<br />

in the point (b + if, σ ) = (0,σ0), where σ0 is defined by ψj0(σ0) = γ0 and where<br />

F (0,σ0) = 0,is<br />

where<br />

D ˆF (0,σ0) :W 1,p (S,C) × R 2g −→ L p (T ∗ S ⊗ C) 0,1<br />

j0<br />

D ˆF (0,σ0)(ζ,σ) = ¯∂j0ζ + ψj0(σ ) + i(ψj0(σ ) ◦ j0) + Lζ,<br />

L : W 1,p (S,C) → W 1,p (T ∗ S ⊗ C) 0,1<br />

j0 ↩→ L p (T ∗ S ⊗ C) 0,1<br />

j0<br />

is the compact linear map<br />

where<br />

Lζ =− 1<br />

2 u∗<br />

0 λ ◦ (Aζ + j0Aζj0) + i<br />

2 u∗<br />

0 λ ◦ (j0Aζ − Aζj0) + Bζ + iBζj0,<br />

Bζ = d<br />

<br />

<br />

<br />

dτ<br />

τ=0<br />

ψjτ k(σ0), ζ = h + ik,


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 51<br />

and where<br />

Aζ = d<br />

<br />

<br />

jτ<br />

dτ k = h (πλTu0)<br />

τ=0<br />

−1 [J (u0)DXλ(u0)<br />

− DXλ(u0)J (u0) + DJ(u0)Xλ(u0)](πλTu0).<br />

The linear term L therefore does not contribute to the Fredholm index of D ˆF (0,σ0).<br />

We note that the linear map ζ ↦→ Lζ only depends on the imaginary part of ζ .We<br />

claim that the operator<br />

W 1,p (S,C) × R 2g −→ L p (T ∗ S ⊗ C) 0,1<br />

j0<br />

(ζ,σ) ↦−→ ¯∂j0ζ + ψj0(σ ) + i(ψj0(σ ) ◦ j0)<br />

is a surjective Fredholm operator of index 2. Then we would have ind(D ˆF (0,σ0)) = 2<br />

as well. Here is the argument: The Riemann-Roch theorem asserts that the kernel and<br />

the cokernel of the Cauchy-Riemann operator ¯∂j (acting on smooth complex-valued<br />

functions on S) are both finite-dimensional and that<br />

<br />

dimR ker ¯∂j<br />

0,1<br />

− dimR E (S)/Im ¯∂j = 2 − 2g,<br />

where g is the genus of the surface S. The only holomorphic functions on S are the<br />

constant functions, hence E0,1 (S)/Im ¯∂j has dimension 2g.<br />

On the other hand, the vector space H 1 j (S) of all (real-valued) harmonic 1-forms<br />

on S also has dimension 2g (see [15]). We now consider the linear map<br />

: H 1<br />

j (S) −→ E0,1 (S)/Im ¯∂j<br />

(γ ):= [γ + i(γ ◦ j)],<br />

where [ . ] denotes the equivalence classes of (0, 1)-forms. Assume that (γ ) = [0],<br />

that is, that there is a complex-valued smooth function f = u+iv on S such that ¯∂jf =<br />

γ +i(γ ◦j).Sinceγ is a harmonic 1-form, we conclude that d(dv◦j) = d(du◦j) = 0<br />

(i.e., both u and v are harmonic). Since there are only constant harmonic functions<br />

on S we obtain γ = 0 (i.e., is injective and also bijective). Hence, (0, 1)-forms<br />

γ + i(γ ◦ j) with γ ∈ H 1 j (S) make up the cokernel of ¯∂j : C ∞ (S,C) → E 0,1 (S).<br />

This proves the claim that the operator D ˆF (0,σ0) is Fredholm of index 2. We<br />

now show that the operator D ˆF (0,σ0) is surjective. Using the decomposition<br />

L p (T ∗ S ⊗ C) 0,1<br />

j0 = R(¯∂j0) ⊕ H 1<br />

j0 (S)


52 CASIM ABBAS<br />

and denoting the corresponding projections by π1,π2, we see that it suffices to prove<br />

the surjectivity of the operator<br />

T : W 1,p (S,C) → R(¯∂j0)<br />

Tζ := ¯∂j0ζ + π1(Lζ ),<br />

which is a Fredholm operator of index 2. Assume that ζ ∈ ker T .Unlessζ ≡ 0, the<br />

set {z ∈ S | ζ (z) = 0} consists of finitely many points by the similarity principle (see<br />

[23]), and the local degree of each zero is positive. On the other hand, the sum of all<br />

the local degrees has to be zero, hence elements in the kernel of T are nowhere zero.<br />

Actually, if h + ik ∈ ker T ,thenevenk is nowhere zero because h + c + ik ∈ ker T<br />

for any real constant c since the zero-order term L only depends on the imaginary part<br />

of ζ . Therefore, ∗ dim ker T ≤ 2, and since the Fredholm index of T equals 2, we<br />

actually have dim ker T = 2. This proves the surjectivity of T and also of D ˆF (0,σ0)<br />

so that the set M of all pairs (b + if, γ ) solving the differential equation (2.10) isa<br />

2-dimensional manifold with T(0,γ0)M = ker D ˆF (0,γ0). If we add a real constant to<br />

b + if , then we obtain again a solution of (2.10). If we divide M by this R-action,<br />

then we obtain a 1-dimensional family of solutions (ũτ )−ε


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 53<br />

where the map r satisfies for every (i, j) ∈ N 2 a decay estimate of the form<br />

with Mij and d positive constants.<br />

|∇ i<br />

s ∇j<br />

t r(s, t)| ≤Mij e −ds<br />

Our situation is less general than in [31], so we will explain the notation in the context<br />

of this paper. The setup is a manifold M with contact form λ and contact structure<br />

ξ = ker λ. Consider a periodic orbit ¯P of the Reeb vector field Xλ with period T ,and<br />

we may assume here that T is its minimal period. We introduce P (t) := ¯P (Tt/2π)<br />

such that P (0) = P (2π). IfJ : ξ → ξ is a dλ-compatible complex structure, then<br />

the set of all ˜J -holomorphic half-cylinders<br />

ũ = (a,u) :[R,∞) × S 1 → R × M, S 1 = R/2πZ<br />

for which |a(s, t) − Ts/2π| and |u(s, t) − P (t)| decay at some exponential rate (in<br />

local coordinates near the orbit P (S 1 )) is denoted by M(P,J). Note that it is assumed<br />

here that the domain [R,∞) × S 1 is endowed with the standard complex structure. A<br />

smooth map U :[R,∞)×S 1 → P ∗ ξ for which U(s, t) ∈ ξP (t) is called an asymptotic<br />

representative of ũ if there is a proper embedding ψ :[R,∞) × S 1 → R × S 1<br />

asymptotic to the identity so that<br />

ũ ψ(s, t) = Ts/2π, exp P (t) U(s, t) , ∀ (s, t) ∈ [R,∞) × S 1<br />

(exp is the exponential map corresponding to some metric on M, e.g., the one induced<br />

by λ and J ). Every ũ ∈ M(P,J) has an asymptotic representative (see [31]). The<br />

asymptotic operator AP,J is defined as follows:<br />

(AP,Jh)(t) :=− T<br />

2π J P (t) d<br />

<br />

<br />

ds<br />

s=0<br />

<br />

Dφ−s(φs(P (t)))h(φs(P (t))) ,<br />

where φs is the flow of the Reeb vector field and where h is a section in P ∗ ξ →<br />

S 1 . Because the Reeb flow preserves the splitting TM = R Xλ ⊕ ξ we have also<br />

(AP,Jh)(t) ∈ ξP (t).<br />

We compute the asymptotic operator AP,J for the binding orbit<br />

¯P (t) =<br />

<br />

t<br />

<br />

, 0, 0 ∈ S<br />

γ1(0) 1 × R2 .<br />

Recall that the above periodic orbit has minimal period T = 2πγ1(0). Using<br />

<br />

φs P (t) = φs(t,0, 0) = t + s<br />

<br />

, 0, 0 ,<br />

γ1(0)


54 CASIM ABBAS<br />

formula (2.2) for the linearization of the Reeb flow with h(t) = (0,ζ(t),η(t)), and<br />

the fact that J (t,0, 0) ∂ ∂<br />

∂<br />

= and J (t,0, 0) ∂x ∂y ∂y =−∂ , we compute<br />

∂x<br />

′ h (t)<br />

(AP,Jh)(t) =−γ1(0)J (t,0, 0)<br />

γ1(0) +<br />

<br />

0 β(0) ζ (t)<br />

−β(0) 0 η(t)<br />

where J0 =<br />

if<br />

0 −1<br />

1 0<br />

=−J0 h ′ (t) − γ1(0)β(0) h(t)<br />

=−J0h ′ (t) + κh(t),<br />

<br />

and κ = γ ′′ ′′<br />

1 (0)/γ 2 (0) ∈ (−1, 0). Hence λ ∈ σ (AP,J) precisely<br />

h ′ (t) = (λ − κ) J0 h(t) and h(2π) = h(0)<br />

(i.e., σ (AP,J) ={κ + l | l ∈ Z}), and the largest negative eigenvalue is given by κ.<br />

The corresponding eigenspace consists of all constant vectors h(t) ≡ const ∈ R 2 .<br />

The eigenspace for the eigenvalues κ + l consists of all<br />

h(t) = e J0lt h0, with h0 ∈ R 2 .<br />

<strong>THEOREM</strong> 3.2 (Compactness)<br />

Let λ be a contact form on M which is a local model near the binding (of the Giroux<br />

leaf v0), and let J :kerλ→ker λ be a dλ-compatible complex structure. Consider a<br />

smooth family of solutions (S,jτ ,aτ ,uτ ,γτ ,J)0≤τ


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 55<br />

Remark 3.3<br />

We may assume without loss of generality that v0 ≡ u0 and f0 ≡ 0.Ifz ∈ ˙S,thenwe<br />

denote by T (z) > 0 the positive return time of the point u0(z); that is, we have<br />

T (z) := inf T>0 φT (u0(z)) ∈ u0( ˙S) < +∞.<br />

We claim that the return time z ↦→ T (z) extends continuously over the punctures of<br />

the surface, and that therefore there is an upper bound<br />

T := sup T (z) < ∞.<br />

z∈ ˙S<br />

Using (2.1)and(2.6), we note that, asymptotically near the punctures, φT (u0(s, t)) ∈<br />

S 1 × R 2 has the following structure:<br />

φT (u0(s, t)) = t + α(r(s))T, r(s) exp[i(α0 + β(r(s))T )] ,<br />

where r(s) is a strictly decreasing function, α0 is some constant, and α(r),β(r) are<br />

suitable functions for which the limits limr→0 β(r) and limr→0 α(r) exist and are not<br />

zero. Hence, if T = T (u0(s, t)) is the positive return time at the point u0(s, t), then<br />

T u0(s, t) =<br />

and therefore the limit for s →+∞exists.<br />

2π<br />

|β(r(s))| ,<br />

The remainder of this section is devoted to the proof of Theorem 3.2. We recall<br />

that the functions aτ and fτ satisfy the Cauchy-Riemann type equation (2.9)whichis<br />

¯∂jτ (aτ + ifτ ) = u ∗<br />

0 λ ◦ jτ − i(u ∗<br />

0 λ) − γτ − i(γτ ◦ jτ ),<br />

where the complex structure jτ is given by (2.8)or<br />

jτ (z) = πλTu0(z) −1 Tφfτ (z)(u0(z)) −1 · J φfτ (z)(u0(z)) <br />

Tφfτ (z) u0(z) πλTu0(z),<br />

and that γτ is a closed 1-form on S with d(γτ ◦ jτ ) = 0.<br />

The following L ∞ -bound is the crucial ingredient for the compactness result. We<br />

claim that<br />

sup<br />

0≤τ


56 CASIM ABBAS<br />

Restricting any of the solutions to a simply connected subset U ⊂ ˙S, we can<br />

write γτ = dhτ for a suitable function hτ : U → R, and the maps<br />

ũτ : U → R × M, ũτ = (aτ + hτ ,uτ )<br />

are ˜J -holomorphic curves. If two such curves ũτ and ũτ ′ have an isolated intersection,<br />

then the corresponding intersection number is positive (see [28], [5], or [27] for<br />

positivity of (self-)intersections for holomorphic curves). We claim that<br />

u0( ˙S) ∩ uτ ( ˙S) =∅, ∀ 0 0 is different from u0( ˙S), we conclude that if uτ and u0 intersect,<br />

then the intersection point of the corresponding holomorphic curves ũτ and ũ0 must<br />

be isolated. But on the other hand, this implies that uτ ′ and u0 would also intersect for<br />

all τ ′ sufficiently close to τ by positivity of the intersection number showing that the<br />

set O is open.<br />

We conclude from the above that we have a sequence τk ↘ ˜τ and points pk,qk ∈ ˙S<br />

such that uτk(pk) = u0(qk). Passing to a suitable subsequence, we may assume<br />

convergence of the sequences (pk)k∈N and (qk)k∈N to points p, q ∈ S. Because of<br />

u˜τ ( ˙S) ∩ u0( ˙S) =∅the points p, q must be punctures, and they have to be equal<br />

z0 = p = q ∈ S\ ˙S. The reason for this is the following. The maps uτk,u0 are<br />

asymptotic near the punctures to a disjoint union of finitely many periodic Reeb<br />

orbits which are not iterates of other periodic orbits. Also, different punctures always<br />

correspond to different periodic orbits. This follows from Giroux’s result and our<br />

constructions in Section 2 of this paper.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 57<br />

We now derive a contradiction using Siefring’s result. The harmonic forms γτk in<br />

equation (1.1) are defined on all of S. Hence they are exact on some open neighborhood<br />

U of the puncture z0,andγτk = dhτk for suitable functions hτk on U and similarly γ˜τ =<br />

dh˜τ . We may also assume that j˜τ |U = jτk|U = j0 after changing local coordinates<br />

near z0. Then on the set U, the maps ũτk = (aτk + hτk,uτk) and ũ0 = (a0 + h0,u0) are<br />

holomorphic curves with ũτk(pk) = ũ0(qk) while the images of ũ˜τ and ũ0 have empty<br />

intersection. Now let<br />

U˜τ ,Uτk, U0 :[R,∞) × S 1 → R 2<br />

be asymptotic representatives of the holomorphic curves ũ˜τ , ũτk, ũ0, respectively.<br />

Invoking Theorem 3.1 and our subsequent computation of the asymptotic operator<br />

and its spectrum, we obtain the following asymptotic formulas<br />

Uτ (s, t) − U0(s, t) = e λτ s eτ (t) + rτ (s, t) , τ = ˜τ,τk, s ≥ Rτ , (3.2)<br />

where Rτ > 0 is some constant and where λτ < 0 is some negative eigenvalue of<br />

the asymptotic operator AP,J. It is of the form λτ = κ + lτ , where lτ is an integer,<br />

κ = γ ′′<br />

1<br />

(0)/γ ′′<br />

2 (0) is not an integer, and where eτ (t) = e J0lτ t hτ , hτ ∈ R 2 \{0} is an<br />

eigenvector corresponding to the eigenvalue λτ = κ + lτ . Note that the above formula<br />

applies since Uτ − U0 cannot vanish identically. We will actually show that lτ ≡ 0.<br />

The asymptotic representative U0 is given by<br />

u0(s, t) = t,r(s)e iα0<br />

<br />

= t,U0(s, t) ,<br />

using equation (2.6), and we recall that r(s) = c(s)eκs , where c(s) → c∞ > 0 as<br />

s →+∞. An asymptotic representative of ũτ , however, is given by an expression<br />

such as<br />

<br />

uτ ψ(s, t) = t,Uτ (s, t) ,<br />

where ψ :[R,∞) × S 1 → R × S 1 is a proper embedding converging to the identity<br />

map as s →+∞. Writing (s ′ ,t ′ ) = ψ(s, t), we get using equations (2.1) forthe<br />

Reeb flow<br />

Uτ (s, t) = c(s ′ )e κs′<br />

e i(α0+β(r(s ′ ))fτ (s ′ ,t ′ ))<br />

= e κs eτ + rτ (s, t) .<br />

The asymptotic formula for Uτ a priori allows for other decay rates, but κ is the only<br />

possible one. Dividing by e κs and passing to the limit s →+∞, we obtain<br />

eτ = c∞e iα0 e iβ(0)fτ (∞) ,


58 CASIM ABBAS<br />

where fτ (∞) = lims→+∞ fτ (s, t) which is independent of t since fτ extends continuously<br />

over the punctures. Hence the difference Uτ − U0 has decay rate λτ ≡ κ as<br />

claimed unless the two eigenvectors eτ and e0 agree, which is equivalent to<br />

fτ (∞) ∈ 2π<br />

β(0) Z<br />

or τ = ˜τ in our case. The maps Uτ − U0 satisfy a Cauchy-Riemann type equation<br />

to which the similarity principle applies so that, for every zero (s, t) of Uτ − U0, the<br />

map σ ↦→ (Uτ − U0)(s + ɛ cos σ, t + ɛ sin σ ) has positive degree for small ɛ>0.<br />

The Cauchy-Riemann type equation mentioned above is derived in [31, Section 5.3]<br />

as well as in [1, Section 3] in a slightly different context, and also in [21]. If R is<br />

sufficiently large, then the map<br />

S 1 → S 1 , t ↦→ Wτ (R,t) := Uτ − U0<br />

|Uτ − U0| (R,t)<br />

is well defined, and it has degree lτ because the remainder term rτ (s, t) decays<br />

exponentially in s. Zeros of Uτ − U0 contribute in the following way: if R ′ 0 so large that<br />

degW˜τ (R ′ , ·) = l˜τ < 0.Forτ sufficiently close to ˜τ,wealsohavedegWτ (R ′ , ·) = l˜τ .<br />

On the other hand, we have degWτ (R,·) = 0 for R>R ′ sufficiently large. Equation<br />

(3.3) implies that the map Uτ − U0 must have zeros in [R ′ ,R] × S 1 to account<br />

for the difference in degrees, but we know that there are none for τ < ˜τ. This<br />

contradiction shows that l˜τ = 0 is impossible. Choose again R ′ > 0 so large that<br />

degW˜τ (R ′ , ·) = 0. The degree does not change if we slightly alter τ. In particular,<br />

we have degWτ (R ′ , ·) = 0 for τ > ˜τ close to ˜τ as well. For R>>R ′ ,wehave<br />

degW˜τ (R,·) = 0, and we recall that<br />

(Uτk − U0)(sk,tk) = 0, τk ↘ ˜τ<br />

for a suitable sequence (sk,tk) with sk →+∞and that the set of zeros of Uτk − U0 is<br />

discrete. This, however, contradicts equation (3.3) since the zeros have positive orders.<br />

Summarizing, we have shown that the assumption O = ∅leads to a contradiction<br />

which implies the a priori bound (3.1).<br />

The monotonicity of the functions fτ in τ and the bound (3.1) imply that the<br />

functions fτ converge pointwise to a measurable function fτ0 as τ ↗ τ0. Wealso


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 59<br />

know that fτ0 L ∞ ( ˙S) ≤ T . We then obtain a complex structure jτ0 on ˙S by<br />

jτ0(z) = πλTu0(z) −1 Tφfτ (z)(u0(z)) 0 −1 × J φfτ (z)(u0(z)) 0 <br />

Tφfτ (z) u0(z) 0 πλTu0(z).<br />

By definition, the complex structure jτ0 is also of class L∞ and jτ (z) → j1(z)<br />

pointwise. Our task is to improve the regularity of the limit fτ0 and the character of<br />

the convergence fτ → fτ0 . We also have to establish convergence of the functions aτ<br />

for τ ↗ τ0. The complex structures jτ are of course all smooth, but the limit jτ0 might<br />

only be measurable.<br />

3.1. The Beltrami equation<br />

For the reader’s convenience, we briefly summarize a few classical facts from the<br />

theory of quasiconformal mappings (see [8], [9]). The punctured surface ˙S carries<br />

metrics gτ , also of class L ∞ for τ = τ0 and smooth otherwise, so that<br />

In fact, gτ is given by<br />

gτ (z) jτ (z)v, jτ (z)w = gτ (z)(v, w), for all v, w ∈ Tz ˙S.<br />

gτ (z)(v, w) = dλ uτ (z) πλTuτ (z)v, J(uτ (z))πλTuτ (z)w .<br />

In the case τ = τ0, we replace πλTuτ (z) by Tφfτ 0 (z)(u0(z))πλTu0(z). Wehave<br />

sup τ gτ L ∞ ( ˙S) < ∞ and gτ → gτ0 pointwise as τ ↗ τ0. Our considerations about<br />

the regularity of the limit are of local nature, so we may replace ˙S with a ball B ⊂ C<br />

centered at the origin. Denoting the metric tensor of gτ by (g τ kl )1≤k,l≤2, wedefinethe<br />

following complex-valued smooth functions:<br />

and we note that<br />

µτ (z) :=<br />

1<br />

2 (gτ 11 (z) − gτ 22 (z)) + igτ 12 (z)<br />

1<br />

2 (gτ 11 (z) + gτ 22 (z)) + gτ 11 (z)gτ 22 (z) − (gτ ,<br />

12 (z))2<br />

sup µτ L∞ ( ˙S) < 1<br />

τ<br />

and that µτ → µτ0 pointwise. We view the functions µτ as functions on the whole<br />

complex plane by trivially extending them beyond B. Then they are also τ-uniformly<br />

bounded in L p (C) for all 1 ≤ p ≤∞and µτ → µτ0 in Lp (C) for 1 ≤ p


60 CASIM ABBAS<br />

for τ


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 61<br />

We used here the notation ∂ = (1/2)(∂s + i∂t) and ∂ = (1/2)(∂s − i∂t). Properties<br />

(1) and (2) above should be understood in the sense of distributions. The proof of (4)<br />

involves the Calderón-Zygmund inequality and the Riesz-Thorin convexity theorem<br />

(see [25] and[32]). Following [9], we define Bp with p>2 to be the space of all<br />

locally integrable functions on the plane which have weak derivatives in Lp (C),vanish<br />

in the origin, and which satisfy a global Hölder condition with exponent 1 − 2<br />

p .For<br />

u ∈ Bp, we then define a norm by<br />

|u(z2) − u(z1)|<br />

uBp := sup<br />

z1=z2 |z2 − z1| 1−(2/p) +∂uLp (C) +∂uLp (C)<br />

so that Bp becomes a Banach space. We usually choose p > 2 such that<br />

cp sup τ µτ L ∞ (C) < 1, where cp is the constant from item (4) above.<br />

<strong>THEOREM</strong> 3.4 ([9, Theorem 1])<br />

Assume that p>2 such that cp sup τ µτ L ∞ (C) < 1.Ifσ ∈ L p (C), then the equation<br />

has a unique solution u = uµ,σ ∈ Bp.<br />

∂u = µ∂u+ σ<br />

For the existence part of the theorem, one first solves the following fixed-point problem<br />

in L p (C):<br />

This is possible because the map<br />

q = Ɣ(µq) + Ɣσ.<br />

L p (C) −→ L p (C)<br />

q ↦−→ Ɣ(µq + σ )<br />

is a contraction in view of cpµL ∞ (C) < 1. Then<br />

u := A(µq + σ )<br />

is the desired solution. The following estimate is also derived in [9]:<br />

qL p (C) ≤ c ′<br />

p σ L p (C)<br />

with c ′ p = cp/(1 − cpµL ∞ (C)), which follows from<br />

qL p (C) ≤Ɣ(µq)L p (C) +ƔσL p (C)<br />

≤ cpµL ∞ (C)qL p (C) + cpσ L p (C).<br />

(3.4)


62 CASIM ABBAS<br />

Recalling our original situation, we have the following result which shows that<br />

there is some sort of conformal mapping for j1 on the ball B.<br />

<strong>THEOREM</strong> 3.5 ([9, Theorem 4])<br />

Let µ : C → C be an essentially bounded measurable function with µ|C\B ≡ 0 and<br />

p>2 such that cpµL ∞ (C) < 1. Then there is a unique map α : C → C with<br />

α(0) = 0 such that<br />

∂α = µ∂α<br />

in the sense of distributions with ∂α − 1 ∈ L p (C).<br />

The desired map α is given by<br />

α(z) = z + u(z),<br />

where u ∈ Bp solves the equation ∂u = µ∂u + µ. In particular, α ∈ W 1,p (B).<br />

Lemma 8 in [9] states that α : C → C is a homeomorphism. We can apply the<br />

theorem to all the µτ , 0 2 such that cp sup n µnL ∞ (C) < 1 and any compact set<br />

B ⊂ C.<br />

Proof<br />

We first estimate with g ∈ Lp (C) and z = 0<br />

|Ag(z)| = 1<br />

<br />

z<br />

<br />

<br />

g(ξ) dξ d¯ξ<br />

2π C ξ(ξ − z)<br />

≤ |z|<br />

2π gLp <br />

1<br />

<br />

<br />

(C) <br />

ξ(ξ − z)<br />

≤ CpgLp 2<br />

1−<br />

(C) |z| p ,<br />

L p<br />

p−1 (C)<br />

(3.5)


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 63<br />

where the last estimate holds in view of<br />

<br />

|ξ(ξ − z)| −p/(p−1) dξ d¯ξ ζ =z−1 <br />

ξ<br />

=<br />

C<br />

If q solves q = Ɣ(µq + µ),then<br />

C<br />

=|z| 2−2p/(p−1)<br />

¯∂(αn − α) = µn ∂(αn − α) − µ∂α+ µn ∂α<br />

|z 2 ζ 2 − z 2 ζ | −p/(p−1) |z| 2 dζ d¯ζ<br />

<br />

|ζ (ζ − 1)| −p/(p−1) dζ d¯ζ .<br />

C <br />

2πCp<br />

<br />

= µn ∂(αn − α) + µn − µ + (µn − µ)Ɣ(µq + µ),<br />

that is, the difference αn − α again satisfies an inhomogeneous Beltrami equation. By<br />

Theorem 3.4, wehave<br />

αn − α = A(µnqn + λn),<br />

where λn = µn − µ + (µn − µ)Ɣ(µq + µ) and where qn ∈ L p (C) solves qn =<br />

Ɣ(µnqn + λn). Combining this with (3.5)and(3.4), we obtain<br />

|αn(z) − α(z)| ≤Cp µnqn + λnL p (C) |z| 1−2/p<br />

≤ (Cp sup µnL<br />

n<br />

∞ (C) · c ′<br />

pλnLp (C) + Cp λnLp (C)) |z| 1−2/p . (3.6)<br />

Since µn − µLp (C) → 0 and (µn − µ)Ɣ(µq + µ)Lp (C) → 0 by Lebesgue’s<br />

theorem we also have λnLp (C) → 0 and therefore αn → α uniformly on compact<br />

sets. Since ¯∂(αn − α) = µn ∂(αn − α) + λn and αn − α = A(µnqn + λn), we verify<br />

that<br />

and<br />

∂(αn − α) = Ɣ(µnqn + λn) = qn<br />

¯∂(αn − α) = µnqn + λn.<br />

Invoking (3.4) once again, we see that both ∂(αn − α)L p (C) and ¯∂(αn − α)L p (C)<br />

can be bounded from above by a constant times λnL p (C) which converges to<br />

zero. <br />

We also need some facts concerning the classical case where µ ∈ C k,α (BR(0)),<br />

BR(0) ={z ∈ C ||z|


64 CASIM ABBAS<br />

<strong>THEOREM</strong> 3.7<br />

Let µ, γ, δ ∈ C k,α (BR ′(0)) with 0


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 65<br />

for a suitable constant c>0 depending on α and R. This is only implicitly proved in<br />

[11], so we sketch the proof of this inequality. We have<br />

and<br />

(T2 − T1)h(z) = A (µ2 − µ1)∂h + (γ2 − γ1)h (z)<br />

− zƔ (µ2 − µ1)∂h + (γ2 − γ1)h (0),<br />

∂ (T2 − T1)h (z) = Ɣ (µ2 − µ1)∂h + (γ2 − γ1)h (z)<br />

− Ɣ (µ2 − µ1)∂h + (γ2 − γ1)h (0),<br />

¯∂ (T2 − T1)h (z) = (µ2 − µ1)(z)∂h(z) + (γ2 − γ1)(z)h(z).<br />

We need [13, (21)–(24)]. Adapted to our notation, they look as follows with z, z1,z2 ∈<br />

BR(0):<br />

Recalling that<br />

and<br />

|(Ah)(z2) − (Ah)(z1)|<br />

|z2 − z1| α<br />

|(Ah)(z)| ≤4RhC 0 (BR(0))<br />

|(Ɣh)(z)| ≤ 2α+1<br />

α Rα hC 0,α (BR(0))<br />

|(Ɣh)(z2) − (Ɣh)(z1)|<br />

|z2 − z1| α<br />

≤ Cα hC0,α (BR(0)).<br />

≤ 2hC 0 (BR(0)) + 2α+2<br />

α Rα hC 0,α (BR(0))<br />

hC 1,α (BR(0)) := hC 0 (BR(0)) +∂hC 0,α (BR(0)) +¯∂hC 0,α (BR(0))<br />

and that the Hölder norm satisfies<br />

kC 0,α (BR(0)) := kC0 |k(z2) − k(z1)|<br />

(BR(0)) + sup<br />

z1=z2 |z2 − z1| α<br />

hkC 0,α (BR(0)) ≤ C hC 0,α (BR(0)) kC 0,α (BR(0))<br />

for a suitable constant C depending only on α and R, the asserted inequality for the<br />

operator norm of T2 − T1 follows. In the same way, we obtain<br />

g2 − g1C 1,α (BR(0)) ≤ c δ2 − δ1C 0,α (BR(0)).


66 CASIM ABBAS<br />

Since<br />

w2 − w1C 1,α (BR(0)) ≤(T2 − T1)w2C 1,α (BR(0))<br />

+ θ w2 − w1C 1,α (BR(0)) +g2 − g1C 1,α (BR(0))<br />

and since θ


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 67<br />

The integral <br />

d(fγ) vanishes by Stokes’s theorem since fγ is a smooth 1-form on<br />

S<br />

the closed surface S. The form da∧γ ◦jf is not smooth on S, but the integral vanishes<br />

anyway for the following reason. As we have proved in the appendix, the form γ ◦ jf<br />

is bounded near the punctures, and hence in local coordinates near a puncture it is of<br />

the form<br />

σ = F (w1,w2)dw1 + F2(w1,w2)dw2, w1 + iw2 ∈ C,<br />

where F1,F2 are smooth (except possibly at the origin) but bounded. Passing to polar<br />

coordinates via<br />

φ :[0, ∞) × S 1 −→ C\{0}<br />

φ(s, t) = e −(s+it) = w1 + iw2,<br />

we see that φ ∗ σ has to decay at the rate e −s for large s. The form da has γ1(r(s)) ds<br />

as its leading term. Computing the integral <br />

Ɣ a(γ ◦ jf ) over small loops Ɣ around<br />

the punctures and using Stokes’s theorem, we conclude that the contribution from<br />

neighborhoods<br />

<br />

of the punctures can be made arbitrarily small. Therefore, the integral<br />

˙S da ∧ γ ◦ jf must vanish.<br />

If is a volume form on S, then we may write u∗ 0λ ∧ γ = g · for a suitable<br />

smooth function g. Defining<br />

<br />

|u<br />

˙S<br />

∗<br />

<br />

0λ ∧ γ | := |g| ,<br />

˙S<br />

we have<br />

γ 2<br />

L2 ,jf =<br />

<br />

<br />

u<br />

˙S<br />

∗<br />

<br />

<br />

0λ ∧ γ <br />

<br />

≤ |u ∗<br />

0λ ∧ γ |<br />

˙S<br />

≤u ∗<br />

0 λL 2 ,jf γ L 2 ,jf ,<br />

which implies the assertion. <br />

We resume the proof of the compactness result, Theorem 3.2. All the considerations<br />

which follow are local. The task is to improve the regularity of the limit fτ0 and the<br />

nature of the convergence fτ → fτ0 . Because the proof is somewhat lengthy, we<br />

organize it in several steps. For τ


68 CASIM ABBAS<br />

be the conformal transformations as in Section 2, that is,<br />

Tατ (z) ◦ i = jτ (ατ (z)) ◦ Tατ (z), z ∈ B.<br />

The L ∞ -bound (3.1) on the family of functions (fτ ) and the above L 2 -bound imply<br />

convergence of the harmonic forms α ∗ τ γτ after maybe passing to a subsequence.<br />

PROPOSITION 3.9<br />

Let τ ′ k be a sequence converging to τ0, and let B ′ = Bε ′(0) with B′ ⊂ B. Then there is<br />

a subsequence (τk) ⊂ (τ ′ k ) such that the harmonic 1-forms α∗ τk γτk converge in C∞ (B ′ ).<br />

Proof<br />

First, the harmonic 1-forms α ∗ τ γτ satisfy the same L 2 -bound as in Proposition 3.8:<br />

α ∗<br />

τ γτ 2<br />

L2 (B) =<br />

<br />

<br />

=<br />

<br />

=<br />

B<br />

B<br />

Uτ<br />

α ∗<br />

τ γτ ◦ i ∧ α ∗<br />

τ γτ<br />

α ∗<br />

τ (γτ ◦ jτ ) ∧ α ∗<br />

τ γτ<br />

γτ ◦ jτ ∧ γτ<br />

≤u ∗<br />

0 λL 2 ,jτ<br />

≤ C,<br />

where C is a constant depending only on λ and u0 since<br />

We write<br />

sup jτ L∞ ( ˙S) < ∞.<br />

τ<br />

α ∗<br />

τ γτ = h 1<br />

τ ds + h2<br />

τ dt,<br />

where hk τ , k = 1, 2 are harmonic and bounded in L2 (B) independent of τ. Ify ∈ B<br />

and BR(y) ⊂ BR(y) ⊂ B, then the classical mean-value theorem<br />

h k 1<br />

τ (y) =<br />

πR2 <br />

h k<br />

τ (x)dx<br />

implies that, for any ball Bδ = Bδ(y) with Bδ ⊂ Bδ ⊂ B, we have the rather generous<br />

estimate<br />

h k<br />

τ C0 (Bδ(y)) ≤ 1<br />

√ h<br />

πδ k<br />

τ L2 √<br />

C<br />

(B) ≤ √ .<br />

πδ<br />

BR(y)


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 69<br />

With y ∈ B and with ν = (ν1,ν2) being the unit outer normal to ∂Bδ(y), weget<br />

∂sh k 1<br />

τ (y) =<br />

πδ2 <br />

∂sh<br />

Bδ(y)<br />

k<br />

τ (x)dx<br />

= 1<br />

πδ2 <br />

div(h<br />

Bδ(y)<br />

k<br />

τ , 0) dx<br />

= 1<br />

πδ2 <br />

h k<br />

τ ν1 ds<br />

and<br />

|∇h k 1<br />

τ (y)| =<br />

πδ 2<br />

∂Bδ(y)<br />

<br />

<br />

<br />

h<br />

∂Bδ(y)<br />

k<br />

τ νds<br />

≤ 2<br />

δ hk<br />

τ C 0 (Bδ(y))<br />

so that, for B ′ = Bε ′, r being the radius of B, andδ = r − ε′ ,wehave<br />

∇h k<br />

τ C 0 (B ′ ) ≤ 2√ C<br />

√ πδ 2 .<br />

By iterating this procedure on nested balls we obtain τ-uniform C 0 (B ′ )-bounds on all<br />

derivatives. Convergence as stated in the proposition then follows from the Ascoli-<br />

Arzela theorem. <br />

3.3. A uniform L p -bound for the gradient<br />

The first step of the regularity story is showing that the gradients of aτ + ifτ are<br />

uniformly bounded in L p (B ′ ) for some p>2 and for any ball B ′ with B ′ ⊂ B. It will<br />

soon become apparent why this gradient bound is necessary. Since we do not have a<br />

lot to start with, the proof will be indirect. Recall the differential equation (2.9)<br />

where<br />

We set<br />

¯∂jτ (aτ + ifτ ) = u ∗<br />

0 λ ◦ jτ − i(u ∗<br />

0 λ) − γτ − i(γτ ◦ jτ ),<br />

jτ (z) = πλTu0(z) −1 Tφfτ (z)(u0(z)) −1 ×J φfτ (z)(u0(z)) <br />

Tφfτ (z) u0(z) πλTu0(z).<br />

<br />

<br />

<br />

φτ (z) := aτ (z) + ifτ (z), z ∈ Uτ


70 CASIM ABBAS<br />

so that, for z ∈ B, wehave<br />

<br />

∂(φτ ◦ ατ )(z) = ∂jτ φτ ατ (z) ◦ ∂sατ (z)<br />

= u ∗<br />

0λ ◦ jτ − i(u ∗<br />

0λ)ατ (z) ◦ ∂sατ (z)<br />

<br />

− (α ∗<br />

τ γτ )(z) · ∂<br />

+ i(α∗ τ<br />

∂s γτ )(z) · ∂<br />

<br />

∂t<br />

and<br />

=: ˆFτ (z) + ˆGτ (z)<br />

=: ˆHτ (z),<br />

(3.8)<br />

sup ˆFτ L<br />

τ<br />

p (B) < ∞ for some p>2 (3.9)<br />

since ατ → ατ0 in W 1,p (B) and supτ jτ L∞ < ∞. Wealsohave<br />

sup ˆGτ C<br />

τ<br />

k (B ′ ) < ∞ (3.10)<br />

for any ball B ′ ⊂ B ′ ⊂ B and any integer k ≥ 0 in view of Proposition 3.9 (the<br />

proposition asserts uniform convergence after passing to a suitable subsequence, but<br />

uniform bounds on all derivatives are established in the proof). We claim now that,<br />

for every ball B ′ ⊂ B ′ ⊂ B, there is a constant CB ′ > 0 such that<br />

∇(φτ ◦ ατ )L p (B ′ ) ≤ CB ′, ∀ τ ∈ [0,τ0). (3.11)<br />

Arguing indirectly, we may assume that there is a sequence τk ↗ τ0 such that<br />

Now define<br />

∇(φτk ◦ ατk)L p (B ′ ) →∞ for some ball B ′ ⊂ B ′ ⊂ B. (3.12)<br />

εk := inf ε>0 ∃ x ∈ B ′ : ∇(φτk ◦ ατk)L p (Bε(x)) ≥ ε 2/p−1 ,<br />

which are positive numbers since ε (2/p)−1 →+∞. Because we assumed (3.12) we<br />

must have infk εk = 0, hence we will assume that εk → 0. Otherwise, if ε0 =<br />

(1/2) infk εk > 0, then we cover B ′ with finitely many balls of radius ε0, andwe<br />

would get a k-uniform L p -bound on each of them, contradicting (3.12). We claim that<br />

∇(φτk ◦ ατk)L p (Bε k (x)) ≤ ε (2/p)−1<br />

k , ∀ x ∈ B ′ . (3.13)<br />

Otherwise, we could find y ∈ B ′ so that<br />

∇(φτk ◦ ατk)Lp (Bε (y)) >ε k (2/p)−1<br />

k ,


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 71<br />

and we would still have the same inequality for a slightly smaller ε ′ k 0, and we define for z ∈ BR(0) the<br />

functions<br />

ξk(z) := (φτk ◦ ατk) xk + εk(z − xk) ,<br />

which makes sense if k is sufficiently large. The transformation<br />

: x ↦−→ xk + εk(x − xk)<br />

satisfies (B1(xk)) = Bεk(xk) and (B1(y)) ⊂ Bεk(xk + εk(y − xk)) so that<br />

<br />

|∇(φτk ◦ ατk)(x)|<br />

Bε (xk) k p dx = ε 2<br />

<br />

|∇(φτk k<br />

◦ ατk)<br />

B1(xk)<br />

xk + εk(z − xk) | p dz<br />

= ε 2<br />

<br />

k ε −p<br />

k |∇ξk(z)| p dz<br />

and<br />

B1(xk)<br />

∇ξkL p (B1(xk)) = ε 1−(2/p)<br />

k ∇(φτk ◦ ατk)L p (Bε k (xk)) (3.15)<br />

= 1<br />

by (3.14) and, for any y for which ξk|B1(y) is defined and for large enough k, wehave<br />

∇ξkL p (B1(y)) ≤ ε 1−(2/p)<br />

k ∇(φτk ◦ ατk)L p (Bε k (xk+εk(y−xk))) ≤ 1 (3.16)<br />

by (3.13). The functions ξk satisfy the equation<br />

¯∂ξk(z) = εk ˆFτk<br />

<br />

xk + εk(z − xk) + εk ˆGτk<br />

<br />

xk + εk(z − xk) =: Hτk(z) (3.17)


72 CASIM ABBAS<br />

and, for every R>0, wehave<br />

sup ∇ξkL<br />

k<br />

p (BR(0)) < ∞, ∇ξkL p (B2(0)) ≥ 1 (3.18)<br />

because of (3.16) and(3.15) sinceB1(xk) ⊂ B2(0) for large k. The upper bound<br />

on ∇ξkL p (BR(0)) depends on how many balls B1(y) are needed to cover BR(0). We<br />

compute for ρ>0<br />

HτkLp (Bρ(xk)) = ε 1−(2/p)<br />

k ˆHτkL p (Bρε (xk)),<br />

k<br />

with ˆHτk as in (3.8). We conclude from p>2, (3.9), and (3.10) that<br />

for any R>0 as k →∞. Defining<br />

HτkL p (BR(0)) −→ 0<br />

X l,p :={ψ ∈ W l,p (B,C 2 ) | ψ(0) = 0, ψ(∂B) ⊂ R 2 },l≥ 1, B⊂ C aball,<br />

the Cauchy-Riemann operator<br />

¯∂ : X l,p −→ W l−1,p (B,C 2 )<br />

is a bounded bijective linear map. By the open mapping principle, we have the following<br />

estimate:<br />

ψl,p,B ≤ C ∂ψl−1,p,B, ∀ ψ ∈ X l,p . (3.19)<br />

Let R ′ ∈ (0,R). Now pick a smooth function β : R 2 → [0, 1] with β|B R ′ (0) ≡ 1 and<br />

supp(β) ⊂ BR(0). Define<br />

We note that<br />

ζk(z) := Re ξk(z) − ξk(0) + iβ(z) Im ξk(z) − ξk(0) .<br />

sup Im(ξk)L<br />

k<br />

p (BR(0)) ≤ CR<br />

with a suitable constant CR > 0 because of the uniform bound<br />

sup Im(ξk)L<br />

k,R<br />

∞ (BR(0)) < ∞.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 73<br />

Using (3.19), we then obtain<br />

because of<br />

ξk − ξk(0)1,p,B R ′ (0) ≤ζk1,p,BR(0)<br />

≤ C ∂ζkL p (BR(0))<br />

≤ CR ′(HτkL p (BR(0)) +∇ξkL p (BR(0))<br />

+Im(ξk) − Im(ξk)(0)L p (BR(0)))<br />

∂ζk = Hτk + i(β − 1)∂ Im(ξk) + i ∂β Im(ξk) − Im(ξk)(0) .<br />

(3.20)<br />

Hence the sequence (ξk − ξk(0)) is uniformly bounded in W 1,p (BR ′(0)), and in particular,<br />

it has a subsequence which converges in Cα 2<br />

(BR ′(0)) for 0


74 CASIM ABBAS<br />

W 1,p<br />

loc (C) since the sequence (ξk − ξk) depends on the choice of the ball BR(0). Our<br />

sequence ξk − ξk(0) has a convergent subsequence on any ball.<br />

3.4. Convergence in W 1,p (B ′ )<br />

Pick a sequence τk ↗ τ0. We claim that the sequence ( ˆFτk) converges in L p (B) maybe<br />

after passing to a suitable subsequence (recall that, so far, we only have the uniform<br />

bound (3.9)). The functions ˆFτk converge pointwise almost everywhere after passing<br />

to some subsequence. Indeed, the sequence {(u∗ 0λ ◦ jτk − i(u∗ 0λ))ατ (z)} converges<br />

k<br />

already pointwise since jτk and ατk do (recall that the sequence (ατk) converges in<br />

W 1,p (B) and therefore uniformly). The sequence (∂sατk) converges in Lp (B) and<br />

therefore pointwise almost everywhere after passing to a suitable subsequence. Then<br />

by Egorov’s theorem, for any δ>0, there is a subset Eδ ⊂ B with |B\Eδ| ≤δ so that<br />

the sequence ˆFτk converges uniformly on Eδ. Letαbe the Lp-limit of the sequence<br />

(∂sατk), andletε>0. We introduce<br />

<br />

C := 2 sup (u<br />

0≤τ0 sufficiently small such that<br />

αL p (B\Eδ) ≤ ε<br />

3 C .<br />

Now choose k0 ≥ 0 so large that, for all k ≥ k0, wehave<br />

∂sατk − αL p (B) ≤ ε<br />

3 C<br />

Then, if k, l ≥ k0, wehave<br />

and ˆFτk − ˆFτlL ∞ (Eδ) ≤ ε<br />

3 |B| .<br />

ˆFτk − ˆFτlL p (B) ≤ˆFτk − ˆFτlL p (Eδ) +ˆFτk − ˆFτlL p (B\Eδ)<br />

proving the claim.<br />

≤|Eδ|ˆFτk − ˆFτlL ∞ (Eδ) + 2 sup ˆFτkL<br />

k≥k0<br />

p (B\Eδ)<br />

≤|B|ˆFτk − ˆFτl p<br />

L ∞ (Eδ)<br />

≤ ε<br />

+ C · sup ∂sατkL<br />

k≥k0<br />

p (B\Eδ)<br />

Recalling that φτ = aτ + ifτ and that the family fτ satisfies a uniform L ∞ -bound, we<br />

have<br />

sup Im(φτ ◦ ατ )L<br />

τ<br />

∞ (B) < ∞.


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 75<br />

Now pick three balls B ′′′ ⊂ B ′′ ⊂ B ′ ⊂ B such that the closure of one is contained<br />

in the next. Our aim is to establish W 1,p (B ′′′ )-convergence of a subsequence of the<br />

sequence (φτk ◦ ατk). By Proposition 3.10, we have a uniform Lp (B ′ )-bound on the<br />

gradient. If β : R2 → [0, 1] is a smooth function with supp (β) ⊂ B ′ and β|B ′′ ≡ 1<br />

and if<br />

ζτ = Re φτ ◦ ατ − φτ (0) + iβ Im φτ ◦ ατ − φτ (0) ,<br />

then we proceed in the same way as in (3.20), and we obtain<br />

ϕk1,p,B ′′ ≤ C ( ˆHτkL p (B ′ ) +∇(φτk ◦ ατk)L p (B ′ ) +Im(ϕk)L p (B ′ )),<br />

where we wrote<br />

ϕk := φτk ◦ ατk − (φτk ◦ ατk)(0),<br />

and where C>0 is a constant depending only on p, B ′ ,andB ′′ . The sequence (ϕk) is<br />

then uniformly bounded in W 1,p (B ′′ ) by Proposition 3.10, and it converges in L p (B ′′ )<br />

after passing to a suitable subsequence. We now use the regularity estimate<br />

ϕl − ϕk1,p,B ′′′ ≤ C ( ˆFτl − ˆFτkL p (B ′′ ) (3.22)<br />

+ˆGτl − ˆGτkL p (B ′′ ) +ϕl − ϕkL p (B ′′ )).<br />

Since the right-hand side converges to zero as k, l →∞, we obtain the following.<br />

PROPOSITION 3.12<br />

For every ball B ′ ⊂ B ′ ⊂ B, the sequence (φτk ◦ ατk − φτk(0)) has a subsequence<br />

which converges in W 1,p (B ′ ).<br />

3.5. Improving regularity using both the Beltrami and Cauchy-Riemann equations<br />

In order to improve the convergence of the conformal transformations ατk , we need<br />

to improve the convergence of the maps µτk → µτ0 and the regularity of its limit.<br />

It is known that the inverses α−1 of the conformal transformations ατk also satisfy a<br />

τk<br />

Beltrami equation (see [9])<br />

where<br />

∂α −1<br />

τ = ντ ∂α −1<br />

τ ,<br />

ντ (z) =− ∂ατ (α −1<br />

τ (z))<br />

∂ατ (α −1<br />

τ (z))µτ<br />

α −1<br />

τ (z)


76 CASIM ABBAS<br />

(follows from 0 = ∂(α−1 τ ◦ ατ ) = ∂α−1 τ (ατ )∂ατ + ∂α−1 τ (ατ )∂ατ ). After passing to a<br />

suitable subsequence, we may assume that ∂ατk and ∂ατk converge pointwise almost<br />

everywhere since they converge in Lp (B). Hence we may assume that the sequence<br />

(ντk) also converges pointwise almost everywhere. We also have<br />

|ντ (z)| ≤ <br />

µτ<br />

−1<br />

α (z) ,<br />

and hence ντ satisfies the same L ∞ -bound as µτ . By Lemma 3.6, we conclude that<br />

α −1<br />

τk<br />

−→ α−1<br />

1 in W 1,p (B)<br />

with the same p>2 as in Lemma 3.6 applied to the functions µτ . After passing to<br />

some subsequence, the sequence<br />

(ϕk) := φτk − aτk(0) ◦ ατk<br />

converges in W 1,p (B ′ ) for any ball B ′ ⊂ B by Proposition 3.12. Indeed, the expression<br />

φτk ◦ ατk − φτk(0) and ϕk differ by a constant term ifτk(0), but the sequence (ifτk(0))<br />

has a convergent subsequence.<br />

We would like to derive a decent notion of convergence for the sequence (ϕτk◦α −1<br />

τk ),<br />

but the space W 1,p is not well behaved under compositions. The composition of two<br />

functions of class W 1,p is only in W 1,p/2 . Since we cannot choose p>2 freely, we<br />

rather carry out the argument in Hölder spaces. By the Sobolev embedding theorem<br />

and Rellich compactness, we may assume that the sequences (ϕk) and (α−1) converge<br />

τk<br />

in C0,α (B ′ ) for any ball B ′ ⊂ B ′ ⊂ B and 0


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 77<br />

spaces, that is,<br />

ϕl − ϕkC k+1,γ (B ′′′ ) ≤ C ( ˆFτl − ˆFτkC k,γ (B ′′ )<br />

+ˆGτl − ˆGτkC k,γ (B ′′ ) +ϕl − ϕkC k,γ (B ′′ )).<br />

The sequence ( ˆFτk) now converges in the C0,α2-norm, and the sequence ( ˆGτk) converges<br />

in any Hölder norm. We obtain with the above regularity estimate C1,α2-convergence of<br />

the sequence (ϕk), and composing with α−1 yields C1,α4-convergence<br />

of (fτk) and (µτl).<br />

τk<br />

Invoking Theorem 3.7 again then improves the convergence of the transformations<br />

ατk, α−1 to C τk 2,α4.<br />

We now iterate the procedure using the regularity estimate for the<br />

Cauchy-Riemann operator in Hölder space and the estimate for the Beltrami equation<br />

in Theorem 3.7.<br />

Theorem 3.2 follows if we apply the implicit function theorem to the limit solution<br />

(S,jτ0, ũτ0 = (aτ0,uτ0),γτ0), hence we obtain the same limit for every sequence {τk},<br />

and we obtain convergence in C∞ .<br />

4. Conclusion<br />

The following remarks tie together the loose ends and prove the main result, Theorem<br />

1.6. We start with a closed 3-dimensional manifold with contact form λ ′ . Giroux’s<br />

theorem, Theorem 1.4, then permits us to change the contact form λ ′ to another<br />

contact form λ such that ker λ = ker λ ′ and such that there is a supporting open book<br />

decomposition with binding K consisting of periodic orbits of the Reeb vector field<br />

of λ. Invoking Proposition 2.4, we construct a family of 1-forms (λδ)0≤δ


78 CASIM ABBAS<br />

is finite, the images of uτ and u0 must agree for some sufficiently large τ, concluding<br />

the proof. <br />

Appendix. Some local computations near the punctures<br />

In this appendix, we present some local computations needed for the proof of Theorem<br />

2.8. The issue is to show that the 1-forms<br />

u ∗<br />

0 λ ◦ jf − da0 and u ∗<br />

0 λ + da0 ◦ jf<br />

are bounded on ˙S. We obtain in the second case of the theorem<br />

u ∗<br />

0 λ ◦ jf − da0 = u ∗<br />

0 λ ◦ (jf − jg) + γ0<br />

= dg ◦ (jf − jg) + γ0 + v ∗<br />

0 λ ◦ (jf − jg).<br />

The first case can be treated as a special case: here the objective is to show that the<br />

1-form v ∗ 0 λ ◦ (jf − i) = v ∗ 0 λ ◦ (jf − j0) is bounded near the punctures. We again drop<br />

the subscript δ in the notation since we are only concerned with a local analysis near<br />

the binding, and all the forms λδ are identical there. We use coordinates (θ,r,φ) near<br />

the binding. The contact structure is then generated by<br />

η1 = ∂<br />

∂r = (0, 1, 0), η2<br />

∂<br />

=−γ2<br />

∂θ<br />

+ γ1<br />

∂<br />

∂φ = (−γ2, 0,γ1).<br />

The projection onto the contact planes along the Reeb vector field is then given by<br />

πλ(v1,v2,v3) = 1<br />

µ (v1γ ′ ′<br />

1 + v3γ 2 ) η2 + v2 η1, with µ = γ1γ ′ ′<br />

2 − γ 1γ2, and the flow of the Reeb vector field is given by<br />

where<br />

φt(θ,r,φ) = θ + α(r)t,r,φ + β(r)t ,<br />

α(r) = γ ′ 2 (r)<br />

µ(r)<br />

and β(r) =− γ ′ 1 (r)<br />

µ(r) .<br />

The linearization of the flow Tφτ (θ,r,φ) preserves the contact structure. In the basis<br />

{η1,η2} it is given by<br />

<br />

1 0<br />

Tφτ (θ,r,φ) =<br />

with A(r) =<br />

τA(r) 1<br />

1<br />

µ 2 ′′ ′ ′′ ′<br />

γ 2 (r)γ 1 (r) − γ 1 (r)γ 2<br />

(r)<br />

(r) .


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 79<br />

The complex structure(s) we chose earlier in (2.7) had the following form near the<br />

binding with respect to the basis {η1,η2}:<br />

<br />

<br />

J (θ,r,φ) =<br />

0 −rγ1(r)<br />

0<br />

.<br />

1<br />

rγ1(r)<br />

The induced complex structure jτ on the surface is then given by<br />

jτ (z) = [πλTv0(z)] −1 <br />

[Tφτ v0(z) ] −1 J φτ (v0(z)) <br />

Tφτ v0(z) πλTv0(z).<br />

With v0(s, t) = (t,r(s),α),wefindthat<br />

so that<br />

and<br />

jτ =<br />

<br />

πλTv0(s, t) =<br />

r ′ (s) 0<br />

0<br />

γ ′ 1 (s)<br />

µ(r(s))<br />

−τA(r)rγ1(r) − rγ1(r)γ ′ 1 (r)<br />

r ′ µ(r)<br />

r ′ µ(r)<br />

rγ1(r)γ ′ 1 (r)(1 + τ 2 A 2 (r)r 2 γ 2 1<br />

(r)) τA(r)rγ1(r)<br />

<br />

−τA(r)rγ1(r) −1<br />

=<br />

1 + τ 2A2 (r)r2γ 2 <br />

1 (r) τA(r)rγ1(r)<br />

<br />

<br />

−1 0<br />

= j0 + τA(r) γ1(r)<br />

τA(r)γ1(r) 1<br />

<br />

jτ − jσ = A(r)rγ1(r)(τ − σ )<br />

using the fact that r(s) satisfies the differential equation<br />

With v ∗ 0 λ = γ1(r) dt, we obtain<br />

r ′ (s) = γ ′ 1 (r(s))γ1(r(s))r(s)<br />

.<br />

µ(r(s))<br />

<br />

<br />

−1 0<br />

(τ + σ )A(r)rγ1(r) 1<br />

v ∗<br />

0λ ◦ (jτ − jσ )|(s,t) = (τ − σ )A r(s) r(s)γ 2<br />

<br />

1 r(s)<br />

· (τ + σ )A r(s) <br />

r(s)γ1 r(s) ds + dt .<br />

<br />

,


80 CASIM ABBAS<br />

Converting from coordinates (s, t) on the half-cylinder to Cartesian coordinates x +<br />

iy = e −(s+it) in the complex plane, we get, with ρ = x 2 + y 2 ,<br />

ds =− 1<br />

(xdx+ ydy) and dt =−1 (xdy− ydx).<br />

ρ2 ρ2 Recall that r(s) = c(s)e κs , where c(s) is a smooth function which converges to a<br />

constant as s →+∞, and we assumed that κ ≤−1/2 and that κ /∈ Z. Another<br />

assumption was that A(r) = O(r). Hence r(s) is close to ρ −κ if s is large (and ρ is<br />

small) and A(r(s)) = O(ρ −κ ). Also recall that γ1(r(s)) = O(1). Summarizing, we<br />

need the expression<br />

A r(s) r(s)ρ −2 ρ = O(ρ −2κ−1 )<br />

to be bounded, which amounts to κ ≤−1/2. The same argument applied to the form<br />

dg ◦ (jτ − jσ ) leads to the same conclusion.<br />

Acknowledgments. I am very grateful to Richard Siefring and Chris Wendl for explaining<br />

some of their work to me. Their results are indispensable for the arguments<br />

in this article. I would also like to thank Samuel Lisi for the numerous discussions we<br />

had about the subject of this article.<br />

References<br />

[1] C. ABBAS, Pseudoholomorphic strips in symplectizations, II: Fredholm theory and<br />

transversality, Comm. Pure Appl. Math. 57 (2004), 1 – 58. MR 2007355 58<br />

[2] ———, Pseudoholomorphic strips in symplectizations, III: Embedding properties and<br />

compactness, J. Symplectic Geom. 2 (2004), 219 – 260. MR 2108375<br />

[3] ———, Introduction to compactness results in symplectic field theory, in preparation.<br />

32<br />

[4] C. ABBAS, K. CIELIEBAK,andH. HOFER, The Weinstein conjecture for planar contact<br />

structures in dimension three, Comment. Math. Helv. 80 (2005), 771 – 793.<br />

MR 2182700 29, 31, 36<br />

[5] C. ABBAS and H. HOFER, Holomorphic curves and global questions in contact<br />

geometry, in preparation. 56<br />

[6] C. ABBAS, H. HOFER,andS. LISI, Renormalization and energy quantization in Reeb<br />

dynamics, in preparation. 29, 36<br />

[7] ———, Some applications of a homological perturbed Cauchy-Riemann equation,<br />

in preparation. 29, 36<br />

[8] L. V. AHLFORS, Lectures on Quasiconformal Mappings, Van Nostrand Math. Stud. 10,<br />

D. Van Nostrand, Toronto, 1966. 59, 60<br />

[9] L. V. AHLFORS and L. BERS, Riemann’s mapping theorem for variable metrics, Ann. of<br />

Math. (2) 72 (1960), 385 – 404. MR 0115006 59, 60, 61, 62, 75


HOLOMORPHIC OPEN BOOK DECOMPOSITIONS 81<br />

[10] J. W. ALEXANDER, A lemma on systems of knotted curves, Proc. Natl. Acad. Sci. USA<br />

9 (1923), 93 – 95. 30<br />

[11] L. BERS, Riemann Surfaces, lectures, New York University, 1957 – 1958.<br />

[12] L. BERS and L. NIRENBERG, “On a representation theorem for linear elliptic systems<br />

with discontinuous coefficients and its applications” in Convegno internazionale<br />

sulle equazioni lineari alle derivate parziali, Trieste (1954), 111 – 140,<br />

Cremonese, Rome, 1955. MR 0076981 60, 63, 64, 65, 66 60<br />

[13] S.-S. CHERN, An elementary proof of the existence of isothermal parameters on a<br />

surface, Proc. Amer. Math. Soc. 6 (1955), 771 – 782. MR 0074856 60, 63, 64,<br />

65<br />

[14] J. B. ETNYRE, Planar open book decompositions and contact structures,Int.Math.<br />

Res. Not. IMRN 2004, no. 79, 4255 – 4267. MR 2126827 31<br />

[15] O. FORSTER, Lectures on Riemann Surfaces, Grad. Texts in Math. 81, Springer, New<br />

York, 1981. MR 0648106 51<br />

[16] E. GIROUX,“Géométrie de contact: De la dimension trois vers les dimensions<br />

supérieures” in Proceedings of the International Congress of Mathematicians,<br />

Vol. II (Beijing, 2002), Higher Ed. Press, Beijing, 2002, 405 – 414. MR 1957051<br />

30, 31<br />

[17] H. HOFER, Pseudoholomorphic curves in symplectizations with applications to the<br />

Weinstein conjecture in dimension three, Invent. Math. 114 (1993), 515 – 563.<br />

MR 1244912 32<br />

[18] ———, “Holomorphic curves and real three-dimensional dynamics” in GAFA 2000<br />

(Tel Aviv, 1999), Geom. Funct. Anal. 2000, Special Volume, Part II, 674 – 704.<br />

MR 1826268<br />

[19] H. HOFER, K. WYSOCKI,andE. ZEHNDER, Properties of pseudoholomorphic curves in<br />

symplectizations, II: Embedding controls and algebraic invariants, Geom. Funct.<br />

Anal. 5 (1995), 270 – 328. MR 1334869<br />

[20] ———, The dynamics on three-dimensional strictly convex energy surfaces,<br />

Ann. of Math. (2) 148 (1998), 197 – 289. MR 1652928 35<br />

[21] ———, Properties of pseudoholomorphic curves in symplectizations,. III: Fredholm<br />

theory, Progr. Nonlinear Differential Equations Appl. 35,Birkhäuser, Basel,<br />

1999, 381 – 475. MR 1725579 58<br />

[22] ———, Finite energy foliations of tight three-spheres and Hamiltonian dynamics,<br />

Ann. of Math. (2) 157 (2003), 125 – 255. MR 1954266 35<br />

[23] H. HOFER and E. ZEHNDER, Symplectic Invariants and Hamiltonian Dynamics,<br />

Birkhäuser Advanced Texts: Basler Lehrbücher, Birkhäuser, Basel, 1994.<br />

MR 1306732 52<br />

[24] C. HUMMEL, Gromov’s Compactness Theorem for Pseudoholomorphic Curves, Prog.<br />

Math. 151,Birkhäuser, Basel, 1997. MR 1451624<br />

[25] P. D. LAX, Functional Analysis, John Wiley, New York, 2001. MR 1892228 61<br />

[26] L. LICHTENSTE<strong>IN</strong>, Zur Theorie der konformen Abbildung nichtanalytischer,<br />

singularitätenfreier Flächenstücke auf ebene Gebiete, J. Krak. Auz. (1916),<br />

192 – 217. 60


82 CASIM ABBAS<br />

[27] D. MCDUFF and D. SALAMON, J-holomorphic Curves and Symplectic Topology,Amer.<br />

Math. Soc. Colloq. Publ. 52, Amer. Math. Soc., Providence, 2004. MR 2045629<br />

56<br />

[28] M. J. MICALLEF and B. WHITE, The structure of branch points in minimal surfaces and<br />

in pseudoholomorphic curves, Ann. of Math. (2) 141 (1995), 35 – 85.<br />

MR 1314031 56<br />

[29] C. B. MORREY, On the solutions of quasi-linear elliptic partial differential equations,<br />

Trans. Amer. Math. Soc. 43, no. 1 (1938), 126 – 166. MR 1501936 60<br />

[30] D. ROLFSEN, Knots and Links, Mathematics Lecture Series 7, Publish or Perish,<br />

Berkeley, Cali., 1976. MR 0515288 30<br />

[31] R. SIEFR<strong>IN</strong>G, Relative asymptotic behavior of pseudoholomorphic half-cylinders,<br />

Comm. Pure Appl. Math. 61 (2008), 1631 – 1684. MR 2456182 52, 53, 58<br />

[32] E. M. STE<strong>IN</strong>, Singular Integrals and Differentiability Properties of Functions, Princeton<br />

Math. Ser. 30, Princeton Univ. Press, Princeton, 1970. MR 0290095 61<br />

[33] C. H. TAUBES, The Seiberg-Witten equations and the Weinstein conjecture, Geom.<br />

Topol. 11 (2007), 2117 – 2202. MR 2350473 29, 36<br />

[34] ———, The Seiberg-Witten equations and the Weinstein conjecture, II: More closed<br />

integral curves of the Reeb vector field, Geom. Topol. 13 (2009), 1337 – 1417.<br />

MR 2496048 29, 36<br />

[35] W. P. THURSTON and H. E. W<strong>IN</strong>KELNKEMPER, On the existence of contact forms,Proc.<br />

Amer. Math. Soc. 52 (1975), 345 – 347. MR 0375366 40<br />

[36] A. J. TROMBA, Teichmüller Theory in Riemannian Geometry, Lectures Math. ETH<br />

Zürich, Birkhäuser, Basel, 1992. MR 1164870<br />

[37] A. WE<strong>IN</strong>STE<strong>IN</strong>, On the hypotheses of Rabinowitz’ periodic orbit theorems,J.<br />

Differential Equations 33 (1979), 353 – 358. MR 0543704<br />

[38] C. WENDL, Finite energy foliations on overtwisted contact manifolds, Geom. Topol. 12<br />

(2008) 531 – 616. MR 2390353 34, 35, 36<br />

[39] ———, Open book decompositions and stable Hamiltonian structures, Expo. Math.<br />

28 (2010), 187 – 197. MR 2671115 34, 36, 40<br />

[40] ———, Strongly fillable contact manifolds and J-holomorphic foliations, Duke Math.<br />

J. 151 (2010), 337 – 387. MR 2605865 35<br />

Department of Mathematics, Michigan State University, East Lansing, Michigan 48824, USA;<br />

abbas@math.msu.edu


CALDERÓN <strong>IN</strong>VERSE PROBLEM WITH PARTIAL<br />

DATA ON RIEMANN SURFACES<br />

COL<strong>IN</strong> GUILLARMOU and LEO TZOU<br />

Abstract<br />

On a fixed smooth compact Riemann surface with boundary (M0,g), we show that,<br />

for the Schrödinger operator g + V with potential V ∈ C 1,α (M0) for some α>0,<br />

the Dirichlet-to-Neumann map N |Ɣ measured on an open set Ɣ ⊂ ∂M0 determines<br />

uniquely the potential V . We also discuss briefly the corresponding consequences for<br />

potential scattering at zero frequency on Riemann surfaces with either asymptotically<br />

Euclidean or asymptotically hyperbolic ends.<br />

Contents<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

2. Harmonic and holomorphic Morse functions on a Riemann surface . . . 86<br />

3. Carleman estimate for harmonic weights with critical points . . . . . . . 95<br />

4. Complex geometric optics on a Riemann surface . . . . . . . . . . . . . 99<br />

5. Identifying the potential . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

6. Inverse scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

Appendix. Boundary determination . . . . . . . . . . . . . . . . . . . . . . 114<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

1. Introduction<br />

The problem of determining the potential in the Schrödinger operator by boundary<br />

measurement goes back to Calderón [8]andGelfand[12]. Mathematically, it amounts<br />

to asking if one can detect some data from boundary measurement in a domain (or<br />

manifold) with boundary. The typical model to have in mind is the Schrödinger<br />

operator P := g + V , where g is a metric and V is a potential; then we define the<br />

Cauchy data space by<br />

C := (u|∂,∂νu|∂); u ∈ H 1 (), u∈ ker P ,<br />

DUKE MATHEMATICAL JOURNAL<br />

Vol. 158, No. 1, c○ 2011 DOI 10.1215/00127094-1276310<br />

Received 17 September 2009. Revision received 22 July 2010.<br />

2010 Mathematics Subject Classification. Primary 35R30; Secondary 58J32.<br />

Guillarmou’s work partially supported by National Science Foundation grant DMS-0635607.<br />

Tzou’s work partially supported by National Science Foundation grant DMS-0807502.<br />

83


84 GUILLARMOU and TZOU<br />

where ∂ν is the interior pointing normal vector field to ∂ and where H 1 () =<br />

W 1,2 () is the space of L2 functions with one derivative in L2 .<br />

The first natural question is the following full data inverse problem: does the<br />

Cauchy data space determine uniquely the metric g and/or the potential V ?Ina<br />

sense, the most satisfying known results are those obtained when the domain ⊂ Rn is already known and g is the Euclidean metric, for which the identification of V<br />

has been proved in dimension n>2 by Sylvester and Uhlmann [32] and Novikov<br />

[30] (where reconstruction is also obtained), and very recently in dimension 2 by<br />

Bukgheim [6] when the domain is simply connected. In a recent work [16], we<br />

proved the identification of V on a Riemann surface with boundary from full data<br />

measurement. A related question is the conductivity problem which consists in taking<br />

V = 0 and replacing g by −divσ ∇, where σ is a positive definite symmetric tensor.<br />

An elementary observation shows that the problem of recovering a sufficiently smooth<br />

isotropic conductivity (i.e., σ = σ0Id for a function σ0) is contained in the abovementioned<br />

problem of recovering a potential V . For domain of R2 , Nachman [29]<br />

used the ¯∂ techniques to show that the Cauchy data space determines the conductivity.<br />

Recently a new approach developed by Astala and Päivärinta in [2] improved this<br />

result to assuming that the conductivity is only an L∞ scalar function on a simply<br />

connected domain. This was later generalized to L∞ anisotropic conductivities by<br />

Astala, Lassas, and Päivärinta in [3]. We notice that there still are rather few results<br />

in the direction of recovering the Riemannian manifold (,g) when V = 0, for<br />

instance the surface case by Lassas and Uhlmann [24] (seealso[4], [17]), the real<br />

analytic manifold case by Lassas, Taylor, and Uhlmann [23] (seealso[15] forthe<br />

Einstein case), the case of manifolds admitting limiting Carleman weights and in a<br />

same conformal class by Dos Santos Ferreira, Kenig, Salo, and Uhlmann [9]. The<br />

second natural, but harder, problem is the partial data inverse problem: if Ɣ1 and Ɣ2<br />

are open subsets of ∂, does the partial Cauchy data space for P<br />

CƔ1,Ɣ2 := (u|Ɣ1,∂νu|Ɣ2); u ∈ H 1 <br />

(∂M0), Pu= 0, u= 0in∂ \ Ɣ1<br />

determine the domain , the metric, the potential? For a fixed domain of R n ,the<br />

recovery of the potential if n>2 with partial data measurements was initiated by<br />

Bukhgeim and Uhlmann [7] and later improved by Kenig, Sjöstrand, and Uhlmann<br />

[21] to the case where Ɣ1 and Ɣ2 are respectively open subsets of the “front” and<br />

“back” ends of the domain. We refer the reader to the references for a more precise<br />

formulation of the problem. In dimension 2, the recent works of Imanuvilov, Uhlmann,<br />

and Yamamoto [19] solves the problem for fixed domains of R 2 in the case when<br />

Ɣ1 = Ɣ2 and when the potential are in C 2,α () for some α>0. In this article, we<br />

address the same question when the background domain is a fixed Riemann surface<br />

with boundary. We prove the following recovery result.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 85<br />

<strong>THEOREM</strong> 1.1<br />

Let (M0,g) be a smooth compact Riemann surface with boundary, let Ɣ ⊂ ∂M0 be<br />

an open subset of the boundary, and let g be the nonnegative Laplacian on M0.For<br />

α ∈ (0, 1),letV1,V2 ∈ C1,α (M0) be two potentials and, for i = 1, 2,let<br />

C Ɣ<br />

i =: (u|Ɣ,∂νu|Ɣ); u ∈ H 1 (M0), (g + Vi)u = 0, u = 0on∂M0 \ Ɣ <br />

(1)<br />

be the respective Cauchy partial data spaces. If C Ɣ 1 = C Ɣ 2 ,thenV1 = V2.<br />

Here the space C1,α (M0) is the usual Hölder space for α ∈ (0, 1). Notice that when<br />

g + Vi do not have L2 eigenvalues for the Dirichlet condition, the statement above<br />

can be given in terms of Dirichlet-to-Neumann operators. Since ˆg = e−2ϕg when<br />

ˆg = e2ϕg for some function ϕ, it is clear that in the statement in Theorem 1.1, we<br />

need only to fix the conformal class of g instead of the metric g (or equivalently,<br />

we need only to fix the complex structure on M0). In particular, the smoothness<br />

assumption of the Riemann surface with boundary is not really essential since we<br />

can change it conformally to make it smooth; for the Cauchy data space, this just<br />

has the effect of changing the potential conformally (all we need is that this new<br />

potential be C1,α ). Observe also that Theorem 1.1 implies that, for a fixed Riemann<br />

surface with boundary (M0,g), the Dirichlet-to-Neumann map on Ɣ for the operator<br />

u →−divg(γ ∇gu) determines the isotropic conductivity γ if γ ∈ C3,α (M0) in the<br />

sense that two conductivities giving rise to the same Dirichlet-to-Neumann map are<br />

equal. This is a standard observation by transforming the conductivity problem to a<br />

potential problem with potential V := (gγ 1/2 )/γ 1/2 . So our result also extends that<br />

of Henkin and Michel [17] in the case of isotropic conductivities. Notice also that<br />

reconstruction methods for isotropic conductivities are obtained in a recent work of<br />

Henkin and Novikov [18] for full boundary data measurements on Riemann surfaces.<br />

The method to identify the potential follows [6] and[19] and is based on the<br />

construction of a large set of special complex geometric optic solutions of (g +<br />

V )u = 0, also called Faddeev-type solutions and introduced first by Faddeev in [10].<br />

More precisely, if Ɣ0 = ∂M0 \ Ɣ is the set where we do not know the Dirichletto-Neumann<br />

operator, then we construct solutions of the form u = Re e/h (a +<br />

r(h)) + eRe()/hs(h) with u|Ɣ0 = 0, where h>0 is a small parameter, and a<br />

are holomorphic functions on (M0,g) independent of h,and||r(h)||L2 = O(h) while<br />

||s(h)||L2 = O(h3/2 | log h|) as h → 0. The idea of [6] to reconstruct V (p) for p ∈ M0<br />

is to take with a nondegenerate critical point at p and then use stationary phase as<br />

h → 0. In our setting, the function needs to be purely real on Ɣ0 and Morse with a<br />

prescribed critical point at p. One of our main contributions is a geometric construction<br />

of the holomorphic Carleman weights satisfying such conditions. We should point<br />

out that we use a method quite different from that in [19] to construct this weight,<br />

and we believe that our method simplifies their construction even in their case. A


86 GUILLARMOU and TZOU<br />

Carleman estimate on the surface for this degenerate weight needs to be proved, and<br />

we follow the ideas of [19]. We manage to improve the regularity of the potential to<br />

C1,α instead of C2,α in [19]. Since it did not seem to be available in the literature in<br />

our geometric setting, we also provide a short proof in the appendix of the fact that the<br />

partial Cauchy data space C Ɣ determines a potential V ∈ C0,α (M0) on Ɣ (for α>0).<br />

An interesting fact about this proof is that it uses our Carleman estimate, and it allows<br />

rather weak assumptions on the regularity of V .<br />

In Section 6, we obtain two inverse scattering results as a corollary of Theorem<br />

1.1, first for partial data scattering at zero frequency for + V on asymptotically<br />

hyperbolic surfaces with potential decaying at the conformal infinity, and second for<br />

full data scattering at zero frequency for + V with V compactly supported on an<br />

asymptotically Euclidean surface.<br />

Another straightforward corollary in the asymptotically Euclidean case full data<br />

setting is the recovery of a compactly supported potential from the scattering operator<br />

at a positive frequency. The proof is essentially the same as for the operator Rn + V<br />

once we know Theorem 1.1, so we omit it.<br />

2. Harmonic and holomorphic Morse functions on a Riemann surface<br />

2.1. Riemann surfaces<br />

We start by recalling a few elementary definitions and results about Riemann surfaces<br />

(see, e.g., [11] for more details). Let (M0,g0) be a compact, oriented, connected,<br />

smooth Riemannian surface with boundary ∂M0. The surface M0 can be considered<br />

as a subset of a larger oriented Riemannian surface with boundary (M,g), extending<br />

smoothly the metric g0 to M. The conformal class of g and the orientation on<br />

the surface M induce a structure of Riemann surface with boundary (i.e., a surface<br />

equipped with a complex structure via holomorphic charts zα : Uα → C). The Hodge<br />

star operator ⋆ (well defined since M is oriented) acts on the cotangent bundle T ∗M, its eigenvalues are ±i, and the respective eigenspaces T ∗<br />

1,0<br />

M := ker(⋆ + iId) and<br />

T ∗<br />

0,1 M := ker(⋆ − iId) are subbundles of the complexified cotangent bundle CT ∗ M,<br />

and the splitting CT ∗M = T ∗<br />

∗<br />

1,0M ⊕ T0,1M holds as complex vector spaces. Since ⋆<br />

is conformally invariant on 1-forms on M, the complex structure depends only on the<br />

conformal class of g. In holomorphic coordinates z = x + iy in a chart Uα, one has<br />

⋆(udx + vdy) =−vdx + udy and<br />

T ∗<br />

1,0 M|Uα C dz, T ∗<br />

0,1 M|Uα C d ¯z,<br />

where dz = dx + idy and d ¯z = dx − idy. We define the natural projections induced<br />

by the splitting of CT ∗ M as<br />

π1,0 : CT ∗ M → T ∗<br />

1,0 M, π0,1 : CT ∗ M → T ∗<br />

0,1 M.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 87<br />

The exterior derivative d defines the de Rham complex 0 → 0 → 1 → 2 → 0,<br />

where k := kT ∗M denotes the real bundle of k-forms on M. Let us denote Ck the complexification of k . Then the ∂ and ¯∂ operators can be defined as differential<br />

operators ∂ : C 0 → T ∗<br />

1,0 M and ¯∂ : C0 → T ∗<br />

0,1<br />

M by<br />

∂f := π1,0 df, ¯∂ := π0,1 df, (2)<br />

where they satisfy d = ∂ + ¯∂ and where they are expressed in holomorphic coordinates<br />

by<br />

∂f = ∂zf dz,<br />

¯∂f = ∂¯zf d¯z,<br />

with ∂z := (∂x − i∂y)/2 and ∂¯z := (∂x + i∂y)/2. Similarly, one can define the ∂ and<br />

¯∂ operators from C 1 to C 2 by setting<br />

∂(ω1,0 + ω0,1) := dω0,1,<br />

¯∂(ω1,0 + ω0,1) := dω1,0<br />

if ω0,1 ∈ T ∗<br />

0,1M and ω1,0 ∈ T ∗<br />

1,0M. In coordinates, this is simply<br />

∂(udz+ vd¯z) = ∂v ∧ d ¯z,<br />

¯∂(udz+ vd¯z) = ¯∂u ∧ dz.<br />

There is a natural operator, the Laplacian acting on functions and defined by<br />

f := −2i ⋆¯∂∂f = d ∗ d,<br />

where d ∗ is the adjoint of d through the metric g and where ⋆ is the Hodge star operator<br />

mapping 2 to 0 and induced by g as well.<br />

2.2. Maslov index and boundary-value problem for the ∂ operator<br />

In this section, we consider the setting where M0 is an oriented Riemann surface with<br />

boundary ∂M0 and Ɣ ⊂ ∂M0 is an open subset, and we let Ɣ0 = ∂M0 \ Ɣ be its<br />

complement in ∂M0. We can identify the boundary with a disjoint union of circles<br />

∂M0 = m i=1 ∂iM0, where each ∂iM0 S1 .SinceƔwill be the piece of the boundary<br />

where we know the Cauchy data space, it is sufficient to assume that Ɣ is a connected<br />

nonempty open segment of ∂1M0 = S1 , which we now do. Following [26], we adopt<br />

the following notation: let E → M0 be a complex line bundle with complex structure<br />

J : E → E, andletD : C∞ (M0,E) → C∞ (M0,T∗ 0,1 ⊗ E) be a Cauchy-Riemann<br />

operator with smooth coefficients on M0 acting on sections of the bundle E. Observe<br />

that in the case when E = M0 × C is the trivial line bundle with the natural complex<br />

structure on M0,thenDcan be taken to be the operator ∂ introduced in (2). For q>1,<br />

we define<br />

DF : W ℓ,q<br />

F (M0,E) → W ℓ−1,q (M0,T ∗<br />

0,1 M0 ⊗ E),


88 GUILLARMOU and TZOU<br />

where F ⊂ E |∂M0 is a totally real subbundle (i.e., a subbundle such that JF ∩ F is<br />

the zero section) and where DF is the restriction of D to the Lq-based Sobolev space<br />

with ℓ derivatives and boundary condition F<br />

W ℓ,q<br />

F (M0,E):= ξ ∈ W ℓ,q (M0,E) ξ(∂M0) ⊂ F .<br />

When q = 2, we will use the standard notation H ℓ (M0) := W ℓ,2 (M0) for L 2 -based<br />

Sobolev spaces. The boundary Maslov index for a totally real subbundle F ⊂ E∂M0<br />

of a complex vector bundle is defined in generality in [26, Appendix C.3]; we only<br />

recall the definition for our setting.<br />

Definition 2.2.1<br />

Let E = M0 × C and ∂M0 = m j=1 ∂iM0 be a disjoint union of m circles. The<br />

boundary Maslov index µ(E,F) is the degree of the map ρ ◦ : ∂M0 → ∂M0,<br />

where<br />

|∂iM0 : S 1 ∂iM0 → GL(1, C)/GL(1, R)<br />

is the natural map assigning to z ∈ S 1 the totally real subspace Fz ⊂ C, where<br />

GL(1, C)/GL(1, R) is the space of totally real subbundles of C, and where ρ :<br />

GL(1, C)/GL(1, R) → S 1 is defined by ρ(A.GL(1, R)) := A 2 /|A| 2 .<br />

In this setting, we have the following boundary-value Riemann-Roch theorem stated<br />

in [26, Theorem C.1.10].<br />

<strong>THEOREM</strong> 2.2.2<br />

Let E → M0 be a complex line bundle over an oriented compact Riemann surface<br />

with boundary, and let F ⊂ E |∂M0 be a totally real subbundle. Let D be a smooth<br />

Cauchy-Riemann operator on E acting on W ℓ,q (M0,E) for some q ∈ (1, ∞) and<br />

ℓ ∈ N. Then we have the following.<br />

(1) The following operators are Fredholm:<br />

DF : W ℓ,q<br />

F (M0,E) → W ℓ−1,q (M0,T ∗<br />

0,1 M0 ⊗ E),<br />

D ∗<br />

F<br />

ℓ,q ∗<br />

: WF (M0,T0,1M0 ⊗ E) → W ℓ−1,q (M0,E).<br />

(2) The real Fredholm index of DF is given by<br />

Ind(DF ) = χ(M0) + µ(E,F),<br />

where χ(M0) is the Euler characteristic of M0 and where µ(E,F) is the<br />

boundary Maslov index of the subbundle F .


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 89<br />

(3) If µ(E,F) < 0, thenDF is injective, while if µ(E,F) + 2χ(M0) > 0, the<br />

operator DF is surjective.<br />

As an application, we obtain the following (here and in what follows, H m (M0) :=<br />

W m,2 (M0)).<br />

COROLLARY 2.2.3<br />

(i) For q > 1 and k ∈ N0, letω∈ W k,q (M0,T∗ 0,1M0). Then there exists u ∈<br />

W k+1,q (M0) real-valued on Ɣ0 such that ¯∂u = ω.<br />

(ii) For m>1/2, letf∈ H m (∂M0) be a real-valued function. Then there exists<br />

a holomorphic function v ∈ H m+(1/2) (M0) such that Re(v)|Ɣ0 = f .<br />

(iii) For k ∈ N and q>1, the space of W k,q (M0) holomorphic functions on M0<br />

which are real-valued on Ɣ0 is infinite-dimensional.<br />

Proof<br />

(i) Let L ∈ N be arbitrarily large, and let Ɣ be a connected nonempty open segment<br />

of one connected component ∂1M0 = S 1 of ∂M0. ThenƔ can be defined in a<br />

coordinate θ (respecting the orientation of the boundary) by Ɣ ={θ ∈ S 1 |<br />

0 0.SinceL can be taken as<br />

large as we want, this achieves the proof of (i).<br />

(ii) Let w ∈ H m+(1/2) (M0) be a real function with boundary value f on ∂M0.Then<br />

by (i), there exists R ∈ H m+(1/2) (M0) such that i ¯∂R =−¯∂w and R purely real<br />

on Ɣ0; thus v := iR + w is holomorphic such that Re(v) = f on Ɣ0.<br />

(iii) Taking the subbundle F as in the proof of (i), we have that dim ker DF =<br />

χ(M0) + 2L if L satisfies 2χ(M0) + 2L >0, and since L can be taken as large<br />

as we like, this concludes the proof.


90 GUILLARMOU and TZOU<br />

LEMMA 2.2.4<br />

Let {p0,p1,...,pn} ⊂M0 be a set of n + 1 disjoint points. Let c1,...,cK ∈ C, let<br />

N ∈ N, and let z be a complex coordinate near p0 such that p0 ={z = 0}. Thenif<br />

p0 ∈ int(M0), there exists a holomorphic function f on M0 with zeros of order at least<br />

N at each pj such that f is real on Ɣ0 and f (z) = c0 +c1z+···+cKz K +O(|z| K+1 )<br />

in the coordinate z.Ifp0 ∈ ∂M0, then the same is true except that f is not necessarily<br />

real on Ɣ0.<br />

Proof<br />

First, using linear combinations and induction on K, it suffices to prove the lemma for<br />

any K and c0 =···=cK−1 = 0, which we now show. Consider the subbundle F as<br />

in the proof of (i) in Corollary 2.2.3. The Maslov index µ(E,F) is given by 2L, and so<br />

for each N ∈ N, one can take L large enough to have µ(F,E)+2χ(M0) ≥ 2N(1+n).<br />

Therefore, by Theorem 2.2.2, the dimension of the kernel of ∂F will be greater than<br />

2(n + 1)N. Now, since for each pj and complex coordinate zj near pj the map<br />

u → (u(pj),∂zju(pj),...,∂N−1 zj u(pj)) ∈ CN is linear, this implies that there exists<br />

a nonzero element u ∈ ker DF which has zeros of order at least N at all pj.<br />

First, assume that p0 ∈ int(M0) and that we want the desired Taylor expansion<br />

at p0 in the coordinate z. In the coordinate z, one has u(z) = αzM + O(|z| M+1 ) for<br />

some α = 0 and M ≥ N. Define the function rK(z) = χ(z) cK<br />

α z−M+K , where χ(z) is<br />

a smooth cutoff function supported near p0 and which is 1 near p0 ={z = 0}. Since<br />

M ≥ N>1, this function has a pole at p0 and trivially extends smoothly to M0\{p0},<br />

which we still call rK. Observe that the function is holomorphic in a neighborhood<br />

of p0 but not at p0 where it is only meromorphic, so that in M0 \{p0}, ∂rK is a<br />

smooth and compactly supported section of T ∗<br />

0,1 M0, and therefore trivially extends<br />

smoothly to M0 (by setting its value to be zero p0) to a 1-form denoted ωK. Bythe<br />

surjectivity assertion in Corollary 2.2.3, there exists a smooth function RK satisfying<br />

∂RK =−ωK,andRK|Ɣ0 ∈ R.WenowhavethatRK + rK is a holomorphic function<br />

on M\{p0} meromorphic with a pole of order M − K at p0, and in coordinate z one<br />

has z M−K (RK(z) + rK(z)) = cK + O(|z|). Setting f = u(RK + rK), wehavethe<br />

desired holomorphic function. Note that f also vanishes to order N at all p1,...,pn<br />

since u vanishes. This achieves the proof. <br />

Now, if p0 ∈ ∂M0, we can consider a slightly larger smooth domain of M containing<br />

M0, and we apply the result above.<br />

2.3. Morse holomorphic functions with prescribed critical points<br />

The main result of this section is the following.<br />

PROPOSITION 2.3.1<br />

Let p be an interior point of M0, and let ɛ > 0 be small. Then there exists a<br />

holomorphic function on M0 which is Morse on M0 (up to the boundary) and


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 91<br />

real-valued on Ɣ0, which has a critical point p ′ at distance less than ɛ from p and<br />

such that Im((p ′ )) = 0.<br />

Let O be a connected open set of M such that Ō is a smooth surface with boundary,<br />

with M0 ⊂ Ō ⊂ M and Ɣ0 ⊂ ∂Ō.Fixk>2 a large integer; we denote by Ck ( Ō) the<br />

Banach space of Ck real-valued functions on Ō. Then the set of harmonic functions<br />

on Ō which are in the Banach space Ck ( Ō) (and smooth in O by elliptic regularity)<br />

is the kernel of the continuous map : Ck ( Ō) → Ck−2 ( Ō), and so it is a Banach<br />

subspace of Ck ( Ō). The set H ⊂ Ck ( Ō) of harmonic functions u in Ck ( Ō) such that<br />

there exists v ∈ Ck ( Ō) harmonic with u + iv holomorphic on O is a Banach subspace<br />

of Ck ( Ō) of finite codimension. Indeed, let {γ1,...,γN} be a homology basis for O;<br />

then we have<br />

defined by<br />

H = ker L with L :ker∩C k ( Ō) → CN<br />

L(u) :=<br />

<br />

1<br />

<br />

∂u .<br />

πi γj<br />

j=1,...,N<br />

For all Ɣ ′ 0 ⊂ ∂M0 such that the complement of Ɣ ′ 0<br />

We now show the following.<br />

LEMMA 2.3.2<br />

HƔ ′ 0 :={u ∈ H; u|Ɣ ′ 0<br />

contains an open subset, we define<br />

= 0}.<br />

The set of functions u ∈ HƔ ′ 0<br />

intersection of open dense sets) in HƔ ′ 0 with respect to the Ck (<br />

which are Morse in O is residual (i.e., a countable<br />

Ō) topology.<br />

Proof<br />

We use an argument very similar to that used by Uhlenbeck [34]. We start by defining<br />

m : O × HƔ ′ 0 → T ∗O by (p, u) ↦→ (p, du(p)) ∈ T ∗ p O. This is clearly a smooth map,<br />

linear in the second variable, and moreover mu := m(., u) = (·,du(·)) is Fredholm<br />

since O is finite-dimensional. The map u is a Morse function if and only if mu is<br />

transverse to the zero section, denoted T ∗<br />

0 O,ofT∗O, that is, if<br />

Image(Dpmu) + Tmu(p)(T ∗<br />

0 O) = Tmu(p)(T ∗ O) ∀p ∈ O such that mu(p) = (p, 0),<br />

which is equivalent to the fact that the Hessian of u at critical points is nondegenerate<br />

(see, e.g., [34, Lemma 2.8]). We recall the following transversality theorem.


92 GUILLARMOU and TZOU<br />

<strong>THEOREM</strong> 2.3.3 ([34, Theorem 2])<br />

Let m : X × HƔ ′ 0 → W be a Ck map, where X, HƔ ′ , and W are separable Banach<br />

0<br />

manifolds with W and X of finite dimension. Let W ′ ⊂ W be a submanifold such<br />

that k>max(1, dim X − dim W + dim W ′ ).Ifmis transverse to W ′ , then the set<br />

{u ∈ HƔ ′ 0 ; mu is transverse to W ′ } is dense in HƔ ′ , and more precisely, it is a residual<br />

0<br />

set.<br />

We want to apply it with X := O, W := T ∗O,andW ′ := T ∗<br />

0 O, and the map m is<br />

defined above. We have thus proved Lemma 2.3.2 if one can show that m is transverse<br />

to W ′ .Let(p, u) be such that m(p, u) = (p, 0) ∈ W ′ . Then, identifying T(p,0)(T ∗O) with TpO ⊕ T ∗ p O, one has<br />

D(p,u)m(z, v) = z, dv(p) + Hessp(u)z ,<br />

where Hesspu is the Hessian of u at the point p, viewed as a linear map from<br />

TpO to T ∗ p O. To prove that m is transverse to W ′ , we need to show that (z, v) →<br />

(z, dv(p) + Hessp(u)z) is surjective from TpO ⊕ HƔ ′ 0 to TpO ⊕ T ∗ p O,whichis<br />

realized, for instance, if the map v → dv(p) from HƔ ′ 0 to T ∗ p O is surjective. But from<br />

Lemma 2.2.4, we know that there exist holomorphic functions v and ˜v on O such that<br />

v and ˜v are purely real on Ɣ ′ 0 . Clearly, the imaginary parts of v and ˜v belong to HƔ ′ 0 .<br />

Furthermore, for a given complex coordinate z near p ={z = 0}, we can arrange<br />

them to have series expansion v(z) = z + O(|z| 2 ) and ˜v(z) = iz + O(|z| 2 ) around<br />

the point p. We see, by coordinate computation of the exterior derivative of Im(v) and<br />

Im(v ), thatdIm(v)(p) and d Im(˜v)(p) are linearly independent at the point p. This<br />

shows our claim and ends the proof of Lemma 2.3.2 by using Theorem 2.3.3. <br />

We now proceed to show that the set of all functions u ∈ HƔ ′ 0<br />

degenerate critical points on Ɣ ′ 0 is also residual.<br />

LEMMA 2.3.4<br />

For all p ∈ Ɣ ′ 0 and k ∈ N, there exists a holomorphic function u ∈ Ck (<br />

Im(u)|Ɣ ′ 0<br />

= 0 and ∂u(p) = 0.<br />

such that u has no<br />

Ō) such that<br />

Proof<br />

The proof is quite similar to that of Lemma 2.2.4. By Lemma 2.2.4, we can choose a<br />

holomorphic function v ∈ Ck ( Ō) such that v(p) = 0 and Im(v)|Ɣ ′ = 0; then either<br />

0<br />

∂v(p) = 0 and we are done, or ∂v(p) = 0. Assume now the second case, and let<br />

M ∈ N be the order of p as a zero of v. By the Riemann mapping theorem, there is<br />

a conformal mapping from a neighborhood Up of p in Ō to a neighborhood {|z| <<br />

ɛ, Im(z) ≥ 0} of the real line Im(z) = 0 in C, and one can assume that p ={z = 0}<br />

in these complex coordinates. Take r(z) = χ(z)z−M+1 , where χ ∈ C∞ 0 (|z| ≤ɛ) is a


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 93<br />

real-valued function with χ(z) = 1 in {|z|


94 GUILLARMOU and TZOU<br />

At a point (p, u) such that b(p, u) = 0, a simple computation yields that the differential<br />

D(p,u)b : TpƔ ′ 0 × HƔ ′ 0 → T(p,∂νu(p))(N ∗ Ɣ ′ 0 ) is given by (w, u′ ) ↦→ (w, ∂τ ∂νu(p)w +<br />

∂νu ′ (p)). This observation combined with Lemma 2.3.4 shows that, for all (p, u) ∈<br />

Ɣ ′ 0 × HƔ ′ 0 such that b(p, u) = (p, 0), b is transverse to N ∗ 0 Ɣ′ 0 at (p, 0). Now we can<br />

apply Theorem 2.3.3; with X = Ɣ ′ 0 , W = N ∗Ɣ ′ 0 ,andW ′ = N ∗ 0 Ɣ′ 0 , we see that the set<br />

{u ∈ HƔ ′ 0 ; bu is transverse to N ∗ 0 Ɣ′ 0 } is residual in HƔ ′ . In view of Lemma 2.3.5, we<br />

0<br />

deduce the following.<br />

LEMMA 2.3.7<br />

The set of functions u ∈ HƔ ′ 0 such that u has no degenerate critical point on Ɣ′ 0 is<br />

residual in HƔ ′ 0 .<br />

Observing the general fact that finite intersection of residual sets remains residual, the<br />

combination of Lemmas 2.3.7 and 2.3.2 yields this corollary.<br />

COROLLARY 2.3.8<br />

The set of functions u ∈ HƔ ′ 0<br />

points on Ɣ ′ 0 is residual in HƔ ′ 0 with respect to the Ck (<br />

dense.<br />

which are Morse in O and have no degenerate critical<br />

Ō) topology. In particular, it is<br />

We are now in a position to give a proof of the main proposition of this section.<br />

Proof of Proposition 2.1<br />

As explained above, choose O in such a way that Ō is a smooth surface with boundary,<br />

containing M0, sothatƔ0⊂∂O andsothatOcontains ∂M0\Ɣ0. LetƔ ′ 0 be an open<br />

and such<br />

subset of the boundary of Ō such that the closure of Ɣ0 is contained in Ɣ ′ 0<br />

that ∂Ō\Ɣ′ 0 =∅.Letpbe an interior point of M0. By Lemma 2.2.4, there exists a<br />

holomorphic function f = u + iv on Ō such that f is purely real on Ɣ′ 0 , v(p) = 1,<br />

and df (p) = 0 (thus v ∈ HƔ ′ 0 ).<br />

By Corollary 2.3.8, there exists a sequence (vj)j of Morse functions vj ∈ HƔ ′ 0<br />

such that vj → v in C k (M0) for any fixed k large. By the Cauchy integral formula,<br />

there exist harmonic conjugates uj of vj such that fj := uj + ivj → f in C k (M0).<br />

Let ɛ>0 be small, and let U ⊂ O be a neighborhood containing p andnoother<br />

critical points of f , and with boundary a smooth circle of radius ɛ. In complex local<br />

coordinates near p, we can identify ∂f and ∂fj to holomorphic functions on an open<br />

set of C. Then by Rouche’s theorem, it is clear that ∂fj has precisely one zero in U<br />

and that vj never vanishes in U if j is large enough. Fix to be one of the fj for j<br />

large enough. By construction, is Morse in O and has no degenerate critical points<br />

on Ɣ0 ⊂ Ɣ ′ 0 . We notice that, since the imaginary part of vanishes on all of Ɣ′ 0 ,itis<br />

clear from the reflection principle applied after using the Riemann mapping theorem


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 95<br />

(as in the proof of Lemma 2.3.5) that no point on Ɣ0 ⊂ Ɣ ′ 0 can be an accumulation<br />

point for critical points. Now ∂M0\Ɣ0 is contained in the interior of O, and therefore<br />

no points on ∂M0\Ɣ0 can be an accumulation point of critical points. Since is Morse<br />

in the interior of O, there are no degenerate critical points on ∂M0\Ɣ0. This ends the<br />

proof. <br />

3. Carleman estimate for harmonic weights with critical points<br />

In this section, we prove a Carleman estimate using harmonic weight with nondegenerate<br />

critical points in a way similar to [19]. Let = ϕ + iψ be a holomorphic Morse<br />

function with no critical points on Ɣ0, and which is purely real on Ɣ0. In particular,<br />

this implies that ∂νϕ|Ɣ0 = 0.<br />

PROPOSITION 3.1<br />

Let (M0,g) be a smooth Riemann surface with boundary, and let ϕ : M0 → R be a<br />

C k (M0) harmonic Morse function for k large. Then for all V ∈ L ∞ (M0) there exist<br />

C>0 and h0 > 0 such that, for all h ∈ (0,h0) and u ∈ C ∞ (M0) with u|∂M0 = 0,we<br />

have<br />

1<br />

h u2<br />

L 2 (M0)<br />

1<br />

+ u|dϕ|2<br />

h2 L2 (M0) +du2 L2 2<br />

(M0) +∂νuL2 (Ɣ0)<br />

<br />

≤ C e −ϕ/h (g + V )e ϕ/h u 2<br />

L2 1 2<br />

(M0) + ∂νuL h 2 <br />

(Ɣ) ,<br />

where ∂ν is the exterior unit normal vector field to ∂M0.<br />

We start by modifying the weight as follows. If ϕ0 := ϕ : M0 → R is a real-valued<br />

harmonic Morse function with critical points {p1,...,pN} in the interior of M0,then<br />

we let ϕj : M0 → R be harmonic functions such that pj is not a critical point of<br />

ϕj for j = 1,...,N, their existence is ensured by Lemma 2.2.4. Forallɛ>0, we<br />

define the convexified weight ϕɛ := ϕ − (h/2ɛ)( N j=0 |ϕj| 2 ). To prove the estimate,<br />

we will localize in charts j covering the surfaces. These charts will be taken so that<br />

if j ∩ ∂M0 = ∅,thenj∩∂M0 S1 is a connected component of ∂M0. Moreover,<br />

by Riemann’s mapping theorem (e.g., [25, Lemma 3.2]), this chart can be taken to be<br />

a neighborhood of |z| =1 in {z ∈ C; |z| ≤1} and such that the metric g is conformal<br />

to the Euclidean metric |dz| 2 .<br />

LEMMA 3.2<br />

Let be a chart of M0 as above, and let ϕɛ : → R be as above. Then there are<br />

constants C, C ′ > 0 such that, for all u ∈ C ∞ (M0) supported in and h>0 small<br />

(3)


96 GUILLARMOU and TZOU<br />

enough, we have the estimate<br />

C<br />

ɛ u2 L2 <br />

(M0) + C′ − Im(〈∂τ u, u〉L2 (∂M0)) + 1<br />

h<br />

<br />

∂M0<br />

|u| 2 ∂νϕɛdvg<br />

≤e −ϕɛ/h¯∂e ϕɛ/h 2<br />

uL2 (M0) , (4)<br />

where ∂ν and ∂τ denote, respectively, the exterior pointing normal vector fields and<br />

its rotation by an angle +π/2.<br />

Proof<br />

We use complex coordinates z = x +iy in the chart , where u is supported. Since the<br />

Lebesgue measure dxdy is bounded below and above by dvg, g is conformal to |dz| 2 ,<br />

and the boundary terms in (4) depend only on the conformal class, it then suffices to<br />

prove the estimates with respect to dxdy and the Euclidean metric. We thus integrate<br />

by parts with respect to dxdy and we have<br />

4e −ϕɛ/h¯∂e ϕɛ/h<br />

<br />

2 <br />

u = ∂x + i∂yϕɛ<br />

<br />

u + i∂y +<br />

h<br />

∂xϕɛ<br />

<br />

<br />

u<br />

h<br />

2<br />

<br />

<br />

= ∂x + i∂yϕɛ<br />

<br />

<br />

u<br />

h<br />

2 <br />

<br />

+ i∂y + ∂xϕɛ<br />

<br />

<br />

u<br />

h<br />

2<br />

+ 2<br />

<br />

ϕɛ|u|<br />

h <br />

2 − 1<br />

2 ∂xϕɛ.∂x|u| 2 − 1<br />

<br />

2<br />

∂yϕɛ.∂y|u|<br />

2<br />

+ 2<br />

<br />

∂νϕɛ|u|<br />

h ∂M0<br />

2 <br />

<br />

− 2 ∂xRe(u).∂yIm(u)<br />

M0<br />

− ∂xIm(u).∂yRe(u) <br />

(5)<br />

<br />

<br />

= ∂x + i∂yϕɛ<br />

<br />

<br />

u<br />

h<br />

2 <br />

<br />

+ i∂y + ∂xϕɛ<br />

<br />

<br />

u<br />

h<br />

2<br />

+ 1<br />

<br />

ϕɛ|u|<br />

h <br />

2<br />

+ 1<br />

<br />

∂νϕɛ|u|<br />

h ∂M0<br />

2 <br />

+ 2 ∂τ Re(u).Im(u),<br />

∂M0<br />

where := −(∂ 2 x + ∂2 y ), where ∂ν is the exterior pointing normal vector field to the<br />

boundary, and where ∂τ is the tangent vector field to the boundary (i.e., ∂ν rotated<br />

with an angle π/2) for the Euclidean metric |dz| 2 .Then〈uϕɛ,u〉=(h/ɛ)(|dϕ0| 2 +<br />

|dϕ1| 2 +···+|dϕN| 2 )|u| 2 ,sinceϕj are harmonic, so the proof follows from the fact<br />

that |dϕ0| 2 +|dϕ1| 2 +···+|dϕN| 2 is uniformly bounded away from zero. <br />

Themainsteptogofrom(4) to(3) is the following lemma, whose proof is a slight<br />

modification of the one in [19, Proposition 5.3].


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 97<br />

LEMMA 3.3<br />

With the same assumptions as in Proposition 3.1, and if is either an interior chart<br />

of (M0,g0) or a chart containing a whole boundary-connected component, then there<br />

are positive constants C, C ′ such that for all ɛ>0 small, for all 0 0,<br />

e −ϕɛ/h ϕɛ/h 2 C<br />

<br />

e u ≥<br />

ɛ<br />

du 2 + 1<br />

2<br />

u|dϕɛ|<br />

h2 + 2<br />

h 〈∂xu, u∂xϕɛ〉+ 2<br />

h 〈∂yu,<br />

<br />

u∂yϕɛ〉<br />

+ C ′<br />

<br />

(B∂τA − A∂τ B)|∂νu| 2 + 1<br />

h<br />

∂M0<br />

<br />

∂M0<br />

|∂νu| 2 ∂νϕɛ<br />

<br />

.<br />

(7)


98 GUILLARMOU and TZOU<br />

Using the fact that u is real-valued, that ϕ is harmonic, and that N<br />

j=0 |dϕj| 2 is<br />

uniformly bounded away from zero, we see that<br />

2<br />

h 〈∂xu, u∂xϕɛ〉+ 2<br />

h 〈∂yu, u∂yϕɛ〉 = 1<br />

h 〈u, uϕɛ〉 ≥ C<br />

ɛ u2<br />

for some C>0, and therefore, by possibly modifying C>0, wehave<br />

e −ϕɛ/h ϕɛ/h 2 C<br />

<br />

e u ≥ du<br />

ɛ<br />

2 + 1<br />

h2 u|dϕɛ| 2 + C<br />

ɛ u2<br />

<br />

+ boundary terms. (8)<br />

Now if the diameter of the support of u is chosen small (with size depending only on<br />

|Hessϕ0|(p)) with a unique critical point p of ϕ0 inside, one can use integration by<br />

parts and the fact that the critical point is nondegenerate to obtain<br />

¯∂u 2 + 1<br />

h2 u|∂ϕ0| 2 ≥ 1<br />

<br />

<br />

∂¯z(u<br />

h<br />

2 <br />

<br />

)∂zϕ0dxdy<br />

≥ 1<br />

<br />

<br />

u<br />

h<br />

2 ∂2 z ϕ0<br />

<br />

<br />

dxdy<br />

≥ C′′<br />

h u2<br />

(9)<br />

for some C ′′ > 0. Clearly, the same estimate holds trivially if does not contain a<br />

critical point of ϕ0. Using a partition of unity (θj)j in and absorbing terms of the<br />

form ||u¯∂θj|| 2 into the right-hand side, one obtains (9) for any function u supported<br />

in and vanishing at the boundary. Thus, combining with (8), there are positive<br />

constants C, C0,C1 such that, for h small enough, we have<br />

C<br />

<br />

du<br />

ɛ<br />

2 + 1<br />

≥ C1<br />

ɛ<br />

h 2 u|dϕɛ| 2 + C<br />

ɛ u2<br />

<br />

<br />

du 2 + 1<br />

h 2 u|dϕ0| 2 + 1<br />

h u2<br />

Combining now with (8)gives<br />

≥ C<br />

<br />

du<br />

ɛ<br />

2 + 1<br />

<br />

.<br />

h2 u|dϕ0| 2 − C0<br />

ɛ<br />

<br />

u2<br />

2<br />

e −ϕɛ/h<br />

<br />

ϕɛ/h 2<br />

C1<br />

e u ≥ du<br />

ɛ<br />

2 + 1<br />

h2 u|dϕ|2 + 1<br />

h u2<br />

<br />

+ boundary terms.<br />

Let us now discuss the boundary terms in (7). If ϕj aretakensothat∂νϕj = 0 on Ɣ0,<br />

then ∂νϕɛ = 0 on Ɣ0 and ∂νϕɛ = ∂νϕ + O(h/ɛ) on Ɣ, and thus<br />

<br />

1<br />

|∂νu|<br />

h<br />

2 |∂νϕɛ| ≤ C2<br />

<br />

|∂νu|<br />

h<br />

2<br />

∂M0<br />

for some constant C2 > 0. We finally claim that B∂τ A − A∂τ B = 1 on ∂M0 ∩ .<br />

Indeed, since the chart near a connected component can be taken to be an interior<br />

neighborhood of the circle |z| =1 in C, one has A + iB = e −it , where t ∈ S 1<br />

parameterize the boundary component, so that B∂τ A − A∂τ B = 1 since ∂τ = ∂t<br />

Ɣ


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 99<br />

for the Euclidean metric. Combining all these estimates and possibly modifying the<br />

constants achieves the proof. <br />

Proof of Proposition 3.1<br />

Using triangular inequality and absorbing the term ||Vu|| 2 into the left-hand side of<br />

(3), it suffices to prove (3) with g instead of g + V .Letv ∈ C ∞ 0 (M0); by Lemma<br />

3.3, we have that there exist constants C, C ′ ,C1,C2 > 0 such that<br />

C<br />

<br />

1<br />

ɛ h e−ϕɛ/h 2 1<br />

v +<br />

h2 e−ϕɛ/h 2 1<br />

v|dϕ| +<br />

h2 e−ϕɛ/hv|dϕɛ| 2 +e −ϕɛ/h<br />

<br />

2<br />

dv<br />

≤ <br />

j<br />

+e −ϕɛ/h ∂νv 2<br />

Ɣ0<br />

C ′<br />

ɛ<br />

≤ C1<br />

≤ C2<br />

1<br />

h e−ϕɛ/h χjv 2 + 1<br />

h 2 e−ϕɛ/h χjv|dϕ| 2 + 1<br />

h 2 e−ϕɛ/h χjv|dϕɛ| 2<br />

+e −ϕɛ/h<br />

d(χjv) 2<br />

<br />

+e −ϕɛ/h<br />

∂νvƔ0<br />

<br />

e −ϕɛ/h<br />

g(χjv) 2 +e −ϕɛ/h<br />

∂νv 2<br />

<br />

Ɣ<br />

j<br />

−ϕɛ/h<br />

e gv 2 +e −ϕɛ/h 2 −ϕɛ/h 2 −ϕɛ/h<br />

v +e dv +e ∂νv 2 <br />

Ɣ ,<br />

where (χj)j is a partition of unity associated to the complex charts j on M0. Since<br />

constants on both sides are independent of ɛ and h, we can take ɛ small enough so that<br />

C2e −ϕɛ/h v 2 + C2e −ϕɛ/h dv 2 can be absorbed to the left side. Now set v = e ϕɛ/h w<br />

with w|∂M0 = 0. Then we have (for some new constant C>0)<br />

1<br />

h w2 + 1<br />

h2 w|dϕ|2 + 1<br />

h2 w|dϕɛ| 2 +dw 2 +∂νω 2<br />

Ɣ0<br />

≤ C e −ϕɛ/h<br />

ge ϕɛ/h 2<br />

w +∂νω 2 <br />

Ɣ<br />

Finally, fix ɛ>0,setu := e 1 N ɛ j=0 |ϕj | 2<br />

w, and use the fact that e 1 N ɛ j=0 |ϕj | 2<br />

is independent<br />

of h and is bounded uniformly away from zero and above; we then obtain the desired<br />

estimate for 0 0, with phase a Carleman weight (here a Morse holomorphic function), and<br />

such that the phase has a nondegenerate critical point at p, in order to apply the stationary<br />

phase method. In general, complex geometric optic solutions are 1-parameter<br />

families of solutions of the equation (g + V )u = 0 which are perturbations of


100 GUILLARMOU and TZOU<br />

certain exponentially growing solutions (with respect to the parameter) of the free<br />

equation gu = 0. They were introduced by Faddeev in [10] and are also called<br />

Faddeev-type solutions. The oscillating part of the solutions is what will allow us to<br />

recover information on the perturbation V .<br />

In this section, the potential V has the regularity V ∈ C 1,α (M0) for some α>0.<br />

Choose p ∈ int(M0) such that there exists a holomorphic function = ϕ +iψ which<br />

is Morse on M0, C k in M0 for large k ∈ N, and such that ∂(p) = 0 and has<br />

only finitely many critical points in M0. Furthermore, we ask that is purely real<br />

on Ɣ0. By Proposition 2.3.1, such points p form a dense subset of M0. Given such a<br />

holomorphic function, the purpose of this section is to construct solutions u on M0 of<br />

( + V )u = 0 of the form<br />

u = e /h (a + ha0 + r1) + e /h (a + ha0 + r1) + e ϕ/h r2 with u|Ɣ0 = 0 (10)<br />

for h>0 small, where a is holomorphic and u ∈ C k (M0) for large k ∈ N, a0 ∈<br />

H 2 (M0) is holomorphic, and moreover where a(p) = 0 and a vanishes to high order<br />

at all other critical points p ′ ∈ M0 of . Furthermore, we ask that the holomorphic<br />

function a is purely imaginary on Ɣ0. The existence of such a holomorphic function<br />

is a consequence of Lemma 2.2.4. Given such a holomorphic function on M0, we<br />

consider a compactly supported extension to M, still denoted a.<br />

The remainder terms r1,r2 will be controlled as h → 0 and have particular<br />

properties near the critical points of . More precisely, r2 will be a OL 2(h3/2 | log h|)<br />

and r1 will be of the form hr12 + oL 2(h), where r12 is independent of h, which can<br />

be used to obtain sufficient information from the stationary phase method in the<br />

identification process.<br />

4.1. Construction of r1<br />

We will construct r1 to satisfy<br />

e −/h (g + V )e /h (a + r1) = OL2(h| log h|)<br />

and r1 = r11 + hr12. WeletGbe the Green operator of the Laplacian on the smooth<br />

surface with boundary M0 with Dirichlet condition, so that gG = Id on L2 (M0).In<br />

particular, this implies that ¯∂∂G = (i/2)⋆−1 , where ⋆−1 is the inverse of ⋆ mapping<br />

functions to 2-forms. We extend a to be a compactly supported Ck function on M0,<br />

and we will search for r1 ∈ H 2 (M0) satisfying ||r1||L2 = O(h) and<br />

e −2iψ/h ∂e 2iψ/h r1 =−∂G(aV ) + ω + OH 1(h| log h|), (11)<br />

where ω is a smooth holomorphic 1-form on M0. Indeed, using the fact that is<br />

holomorphic, we have<br />

e −/h ge /h =−2i ⋆¯∂e −/h ∂e /h =−2i ⋆¯∂e −(1/h)(− ¯) ∂e (1/h)(− ¯)<br />

=−2i ⋆¯∂e −2iψ/h ∂e 2iψ/h ;


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 101<br />

applying −2i ⋆¯∂ to (11), we obtain (note that ∂G(aV ) ∈ C 2,α (M0) by elliptic<br />

regularity)<br />

e −/h (g + V )e /h r1 =−aV + OL2(h| log h|).<br />

We choose ω to be a smooth holomorphic 1-form on M0 such that at all critical<br />

points p ′ of in M0, the form b := ∂G(aV ) − ω with value in T ∗<br />

1,0M0 vanishes<br />

to the highest possible order. Writing b = b(z)dz in local complex coordinates, b(z)<br />

is C2+α by elliptic regularity, and we have −2i∂¯zb(z) = aV ; therefore, ∂z∂¯zb(p ′ ) =<br />

∂2 ¯z b(p′ ) = 0 at each critical point p ′ = p by construction of the function a. Therefore,<br />

we deduce that at each critical point p ′ = p, ∂G(aV ) has Taylor series expansion<br />

2 j=0 cjzj + O(|z| 2+α ). That is, all the lower-order terms of the Taylor expansion of<br />

∂G(aV ) around p ′ are polynomials of z only.<br />

LEMMA 4.1.1<br />

Let {p0,...,pN} be finitely many points on M0, and let θ be a C2,α section of T ∗<br />

1,0M0. Then there exists a Ck holomorphic function f on M0 with k ∈ N large, such that f<br />

vanishes to high order at the points {p1,...,pN} and ω = ∂f satisfies the following:<br />

in complex local coordinates z near p0 , one has ∂ℓ z θ(p0) = ∂ℓ z ω(p0) for ℓ = 0, 1, 2,<br />

where θ = θ(z) dz and ω = ω(z) dz.<br />

Proof<br />

This is a direct consequence of Lemma 2.2.4. <br />

Applying this to the form ∂G(aV ) and using the observation made above, we can<br />

construct a C k holomorphic form ω such that in local coordinates z centered at a critical<br />

point p ′ of (i.e., p ′ ={z = 0} in this coordinate), for b =−∂G(aV )+ω = b(z) dz<br />

we have<br />

|∂ m<br />

¯z ∂ℓ<br />

z b(z)| =O(|z|2+α−ℓ−m ) for ℓ + m ≤ 2 if p ′ = p,<br />

|b(z)| =O(|z|) if p ′ = p.<br />

Now, we let χ1 ∈ C ∞ 0 (M0) be a cutoff function supported in a small neighborhood<br />

Up of the critical point p and identically 1 near p, while χ ∈ C ∞ 0 (M0) is defined<br />

similarly with χ = 1 on the support of χ1. We construct r1 = r11 + hr12 in two steps:<br />

first, we construct r11 to solve (11) locally near the critical point p of , and second,<br />

we construct the global correction term r12 away from p by using the extra vanishing<br />

of b in (12) at the other critical points.<br />

We define locally in complex coordinates centered at p and containing the support<br />

of χ<br />

r11 := χe −2iψ/h R(e 2iψ/h χ1b),<br />

(12)


102 GUILLARMOU and TZOU<br />

<br />

−1 where Rf (z) :=−(2πi) R2(1/¯z − ¯ξ)fd¯ξ ∧ dξ for f ∈ L∞ compactly supported<br />

is the classical Cauchy-Riemann operator inverting locally ∂z (r11 is extended by zero<br />

outside the neighborhood of p). The function r11 is in C3+α (M0), and we have<br />

e −2iψ/h ∂(e 2iψ/h r11) = χ1<br />

−∂G(aV ) + ω + η<br />

with η := e −2iψ/h R(e 2iψ/h χ1b)∂χ. (13)<br />

We then construct r12 by observing that b vanishes to order 2 + α at critical points of<br />

other than p (from (12)), and ∂χ = 0 in a neighborhood of any critical point of ψ,<br />

so we can find r12 satisfying<br />

2ir12∂ψ = (1 − χ1)b.<br />

This is possible since both ∂ψ and the right-hand side are valued in T ∗<br />

1,0 M0, and∂ψ<br />

has finitely many isolated zeroes on M0; indeed, r12 is then a function in C 2,α (M0 \ P )<br />

where P := {p1,...,pN} is the set of critical points other than p, and it extends to a<br />

C 1,α (M0) function which satisfies in local complex coordinates z near each pj<br />

|∂ β<br />

¯z ∂ γ<br />

z r12(z)| ≤C|z − pj| 1+α−β−γ , β + γ ≤ 2,<br />

where we used also the fact that ∂ψ can locally be considered as a holomorphic<br />

function with a zero of order 1 at each pj. This implies that r1 ∈ H 2 (M0), andwe<br />

have<br />

e −2iψ/h ∂(e 2iψ/h r1) = b + h∂r12 + η =−∂G(aV ) − ω + h∂r12 + η.<br />

Now the first error term ||∂r12||H 1 (M0) is bounded by<br />

||∂r12||H 1 (1<br />

− χ1)b(z)<br />

<br />

<br />

(M0) ≤ C <br />

∂zψ(z)<br />

H 2 (Up)<br />

<br />

≤ C<br />

for some constant C, where we used the fact that (1 − χ1)b(z)/∂zψ(z) is in H 2 (Up)<br />

and is independent of h. To deal with the η term, we need the following.<br />

LEMMA 4.1.2<br />

The estimates<br />

hold true wherer12 solves 2ir12∂ψ = b.<br />

||η||H 2 = O(| log h|),<br />

||η||H 1 ≤ O(h| log h|),<br />

||r1||L2 = O(h),<br />

||r1 − hr12||L2 = o(h)


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 103<br />

Proof<br />

We start by observing that<br />

||r1 − hr12||L2 =<br />

||r1||L2 ≤<br />

<br />

<br />

χe<br />

−2iψ/h R(e 2iψ/h χ1b) − h χ1b<br />

<br />

<br />

<br />

2i∂zψ<br />

<br />

<br />

χe<br />

−2iψ/h R(e 2iψ/h χ1b) − h χ1b<br />

<br />

<br />

<br />

2i∂zψ<br />

L 2 (Up)<br />

L 2 (Up)<br />

,<br />

+ h||r12||L 2 (M0).<br />

The first term is estimated in [19, Proposition 2.7] and it is a o(h), while ||r12||L 2<br />

is independent of h. Now we are going to estimate the H 2 norms of η. Locally in<br />

complex coordinates z centered at p (i.e., p ={z = 0}), we have<br />

η(z) =−∂zχ(z)e −2iψ(z)/h<br />

<br />

e 2iψ(ξ)/h 1 dξ1dξ2<br />

χ1(ξ)b(ξ) ,<br />

¯z − ¯ξ π<br />

ξ = ξ1 +iξ2. (15)<br />

C<br />

Since b is C2,α in U, we decompose b(ξ) =〈∇b(0),ξ〉+b(ξ) using Taylor’s formula,<br />

so we haveb(0) = ∂ξb(0) = 0 and we split the integral (15) with 〈∇b(0),ξ〉 andb(ξ).<br />

Since the integrand with the 〈∇b(0),ξ〉 term is smooth and compactly supported in ξ<br />

(recall that χ1 = 0 on the support of ∂zχ), we can apply stationary phase to get<br />

<br />

<br />

∂zχ(z)e −2iψ(z)/h<br />

<br />

e<br />

C<br />

2iψ(ξ)/h<br />

1<br />

¯z − ¯ξ<br />

χ1(ξ)〈∇b(0),ξ〉 dξ1dξ2<br />

π<br />

<br />

<br />

≤ Ch 2<br />

uniformly in z. Nowsetbz(ξ) = ∂zχ(z)χ1(ξ)b(ξ)/(z − ξ) which is C2,α in ξ and<br />

smooth in z. Letθ∈ C∞ 0 ([0, 1)) be a cutoff function which is equal to 1 near zero,<br />

and set θh(ξ) := θ(|ξ|/h). Then integrating by parts, we have<br />

<br />

C<br />

e 2iψ(ξ)/hbz(ξ)dξ1dξ2 = h 2<br />

<br />

<br />

− h<br />

e<br />

supp(χ1)<br />

2iψ(ξ)/h ∂¯ξ<br />

supp(χ1)<br />

e 2iψ(ξ)/h θh(ξ)∂ξ<br />

(14)<br />

<br />

1 − θh(ξ)<br />

2i∂¯ξψ ∂ξ<br />

bz(ξ)<br />

<br />

dξ1dξ2<br />

2i∂ξψ<br />

bz(ξ)<br />

<br />

dξ1dξ2. (16)<br />

2i∂ξψ<br />

Using polar coordinates with the fact that bz(0) = 0, it is easy to check that the<br />

second term in (16) is bounded uniformly in z by Ch 2 . To deal with the first term,<br />

we use bz(0) = ∂ξ bz(0) = ∂¯ξ bz(0) = 0 and a straightforward computation in polar<br />

coordinates shows that the first term of (16) is bounded uniformly in z by Ch 2 | log(h)|.<br />

We conclude that<br />

||η||L 2 ≤ C||η||L ∞ ≤ Ch2 | log h|.<br />

It is also direct to see that the same estimates hold with a loss of h −2 for any derivatives<br />

in z, ¯z of order less or equal to 2, since they only hit the χ(z) factor, the (¯z − ¯ξ) −1


104 GUILLARMOU and TZOU<br />

factor, or the oscillating term e −2iψ(z)/h . So we deduce that<br />

||η||H 2 = O(| log h|)<br />

and this ends the proof. <br />

We summarize the result of this section with the following.<br />

LEMMA 4.1.3<br />

Let k ∈ N be large, and let ∈ C k (M0) be a holomorphic function on M0 which is<br />

Morse in M0 with a critical point at p ∈ int(M0). Leta ∈ C k (M0) be a holomorphic<br />

function on M0 vanishing to high order at every critical point of other than p.Then<br />

there exists r1 ∈ H 2 (M0) such that ||r1||L2 = O(h) and<br />

e −/h ( + V )e /h (a + r1) = OL2(h| log h|).<br />

4.2. Construction of a0<br />

We have constructed the correction term r1 which solves the Schrödinger equation to<br />

order h as stated in Lemma 4.1.3. In this section, we construct a holomorphic function<br />

a0 which annihilates the boundary value of the solution on Ɣ0. In particular, we have<br />

the following.<br />

LEMMA 4.2.1<br />

There exists a holomorphic function a0 ∈ H 2 (M0) independent of h such that<br />

and<br />

e −/h ( + V )e /h (a + r1 + ha0) = OL2(h| log h|)<br />

[e /h (a + r1 + ha0) + e /h (a + r1 + ha0)]|Ɣ0 = 0.<br />

Proof<br />

First, notice that h −1 r1|∂M0 = r12|∂M0 ∈ H 3/2 (∂M0) is independent of h. Since is<br />

purely real on Ɣ0 and a is purely imaginary on Ɣ0, we see that this lemma amounts to<br />

constructing a holomorphic function a0 ∈ H 2 (M0) with the boundary condition<br />

Re(r12) + Re(a0) = 0 on Ɣ0.<br />

To construct a0, it suffices to use item (ii) in Corollary 2.2.3. <br />

4.3. Construction of r2<br />

The goal of this section is to complete the construction of the complex geometric optic<br />

solutions by the following proposition.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 105<br />

PROPOSITION 4.3.1<br />

There exist solutions to ( + V )u = 0 with boundary condition u|Ɣ0 = 0 of the<br />

form (10) with r1, a0 constructed in the previous sections and r2 satisfying r2L2 =<br />

O(h3/2 | log h|).<br />

This is a consequence of the following lemma (which follows from the Carleman<br />

estimate obtained above).<br />

LEMMA 4.3.2<br />

If V ∈ L ∞ (M0) and f ∈ L 2 (M0), then for all h>0 small enough, there exists a<br />

solution v ∈ L 2 to the boundary-value problem<br />

satisfying the estimate<br />

e ϕ/h (g + V )e −ϕ/h v = f, v|Ɣ0 = 0,<br />

vL 2 ≤ Ch1/2 f L 2.<br />

Proof<br />

The proof is identical to the proof of [19, Proposition 2.2]; we repeat the argument<br />

here for the convenience of the reader. Define for all h>0 the vector space A :=<br />

{u ∈ H 1 0 (M0); (g + V )u ∈ L2 (M0), ∂νu |Ɣ= 0} equipped with the scalar product<br />

<br />

(u, w)A := e −2ϕ/h (gu + Vu)(gw + Vw)dvg.<br />

M0<br />

Since ψ is constant along Ɣ0 and since ϕ + iψ is holomorphic, then ∂νϕ = 0 on Ɣ0.<br />

Therefore, we may apply the Carleman estimate of Proposition 3.1 to the weight ϕ<br />

to assert that the space A is a Hilbert space equipped with the scalar product above.<br />

By using the same estimate, the linear functional L : w → <br />

M0 e−ϕ/hfwdvg on A<br />

is continuous and its norm is bounded by h1/2 ||f ||L2. By the Riesz theorem, there is<br />

an element u ∈ A such that (u, .)A = L and with norm bounded by the norm of L.<br />

It remains to take v := e−ϕ/h (gu + Vu), which solves (g + V )e−ϕ/hv = e−ϕ/hf and, in addition, satisfies the desired norm estimate. Furthermore, since<br />

<br />

e −ϕ/h <br />

v(g + V )w dvg = e −ϕ/h fwdvg<br />

M0<br />

for all w ∈ A, by Green’s theorem we have<br />

<br />

e −ϕ/h <br />

v∂νw dvg = 0 =<br />

∂M0<br />

Ɣ0<br />

M0<br />

e −ϕ/h v∂νw dvg<br />

for all w ∈ A. Thisimpliesthatv = 0 on Ɣ0.


106 GUILLARMOU and TZOU<br />

Proof of Proposition 4.1<br />

We note that<br />

( + V ) e /h (a + r1 + ha0) + e/h (a + r1 + ha0) + e ϕ/h <br />

r2 = 0<br />

if and only if<br />

e −ϕ/h ( + V )e ϕ/h r2 =−e −ϕ/h ( + V ) e /h (a + r1 + ha0) + e /h (a + r1 + ha0) .<br />

By Lemma 4.2.1, the right-hand side of the above equation is OL2(h| log h|). Therefore,<br />

using Lemma 4.3.2, one can find such r2 which satisfies<br />

r2L 2 ≤ Ch3/2 | log h|, r2 |Ɣ0= 0.<br />

Since the ansatz e /h (a + r1 + ha0) + e /h (a + r1 + ha0) is arranged to vanish on<br />

Ɣ0, the solution<br />

u = e /h (a + r1 + ha0) + e /h (a + r1 + ha0) + e ϕ/h r2<br />

vanishes on Ɣ0 as well. <br />

5. Identifying the potential<br />

We now assume that V1,V2 ∈ C 1,α (M0) are two potentials, with α>0, such that<br />

the respective Cauchy data spaces C Ɣ 1 , C Ɣ 2 for the operators g + V1 and g + V2<br />

on Ɣ ⊂ ∂M0 are equal. We may also assume that V1 = V2 on Ɣ by boundary<br />

identification, a fact which is proved below in the appendix. Let Ɣ0 = ∂M0 \ Ɣ be the<br />

complement of Ɣ in ∂M0, and possibly by taking Ɣ slightly smaller, we may assume<br />

that Ɣ0 contains an open set. Let p ∈ M0 be an interior point of M0 such that, using<br />

Proposition 2.3.1, we can choose a holomorphic Morse function = ϕ + iψ on M0<br />

with purely real on Ɣ0, C k in M0 for some large k ∈ N, with a critical point at p.<br />

Note that Proposition 2.3.1 states that we can choose such that none of its critical<br />

points on the boundary are degenerate and such that critical points do not accumulate<br />

on the boundary.<br />

PROPOSITION 5.1<br />

If the Cauchy data spaces agree, that is, if C Ɣ 1 = C Ɣ 2 ,thenV1(p) = V2(p).<br />

Proof<br />

Let a be a holomorphic function on M0 which is purely imaginary on Ɣ0 with a(p) = 0<br />

and a(p ′ ) = 0 to large order for all other critical points p ′ of . The existence of a is


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 107<br />

ensured by Lemma 2.2.4.Letu1 and u2 be H 2 solutions on M0 to<br />

(g + Vj)uj = 0<br />

constructed in Section 4 with = φ + iψ for Carleman weight for u1 and − for<br />

u2, and thus of the form<br />

u1 = e /h (a + ha0 + r1) + e /h (a + ha0 + r1) + e ϕ/h r2,<br />

u2 = e −/h (a + hb0 + s1) + e −/h (a + hb0 + s1) + e −ϕ/h s2,<br />

and with boundary value uj|∂M0 = fj, where fj vanishes on Ɣ0. By Green’s formula,<br />

we can write<br />

<br />

M0<br />

<br />

u1(V1 − V2)u2 dvg =−<br />

<br />

=−<br />

M0<br />

∂M0<br />

(gu1.u2 − u1.gu2)dvg<br />

(∂νu1.f2 − f1.∂νu2)dvg.<br />

Since the Cauchy data for g + V1 agrees on Ɣ with that of g + V2, there exists a<br />

solution v of the boundary-value problem<br />

(g + V2)v = 0, v|∂M0 = f1,<br />

satisfying ∂νv = ∂νu1 on Ɣ. Sincefj = 0 on Ɣ0, thisimpliesthat<br />

<br />

<br />

u1(V1 − V2)u2 dvg =− (gu1.u2 − u1.gu2)dvg<br />

M0<br />

<br />

=−<br />

<br />

=−<br />

<br />

=−<br />

M0<br />

∂M0<br />

∂M0<br />

M0<br />

(∂νu1.f2 − f1.∂νu2)dvg<br />

(∂νv.f2 − v.∂νu2)dvg<br />

(gv.u2 − v.gu2)dvg = 0 (17)<br />

since g + V2 annihilates both v and u2. We substitute in the full expansion for u1<br />

and u2, and setting V := V1 − V2 and using the estimates in Lemmas 4.1.2, 4.3.1,and<br />

4.2.1, we obtain<br />

0 = I1 + I2 + o(h), (18)


108 GUILLARMOU and TZOU<br />

where<br />

I1 =<br />

<br />

M0<br />

<br />

I2 = 2h<br />

V (a 2 + a 2 <br />

)dvg + 2<br />

M0<br />

<br />

V Re ae 2iψ/h<br />

<br />

s1<br />

h<br />

+ e −2iψ/h<br />

<br />

a0 + r1<br />

<br />

h<br />

M0<br />

V |a| 2 Re(e 2iψ/h )dvg, (19)<br />

+ b0<br />

<br />

+ b0 + a0 + s1 + r1<br />

h<br />

<br />

dvg. (20)<br />

Remark 5.2<br />

We observe from the last identity in Lemma 4.1.2 that r1/h in the expression I2 can<br />

be replaced by the term r12 satisfying 2ir12∂ψ = b up to an error which can go in<br />

the o(h) in (18), and similarly for the term s1/h, which can be replaced by a terms12<br />

independent of h.<br />

We apply the stationary phase to these two terms in Lemmas 5.3 and 5.4.<br />

LEMMA 5.3<br />

The estimate<br />

<br />

I2 = 2h<br />

M0<br />

V Re <br />

a(b0 + a0 +s12 +r12)dvg + o(h)<br />

holds true where r12, s12,r12 ands12 are independent of h.<br />

Proof<br />

We start with the following.<br />

LEMMA 5.4<br />

Let f ∈ L 1 (M0). Thenash → 0,<br />

<br />

M0<br />

e 2iψ/h f dvg = o(1).<br />

Proof<br />

Since C k (M0) is dense in L 1 (M0) for all k ∈ N, it suffices to prove the lemma for<br />

f ∈ C k (M0).Letɛ>0 be small, and choose a cutoff function χ which is identically<br />

equal to 1 on the boundary such that<br />

<br />

M0<br />

χ|f | dvg ≤ ɛ.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 109<br />

Then, splitting the integral and using stationary phase for the 1 − χ term, we obtain<br />

<br />

<br />

e 2iψ/h <br />

<br />

f dvg<br />

≤ (1 − χ)e 2iψ/h <br />

<br />

f dvg<br />

+ χe 2iψ/h <br />

<br />

f dvg<br />

≤ ɛ + Oɛ(h),<br />

M0<br />

M0<br />

which concludes the proof by taking h small enough depending on ɛ. <br />

The proof of Lemma 5.3 is a direct consequence of Lemma 5.4 and Remark<br />

5.2.1. <br />

The second lemma will be proved at the end of this section.<br />

LEMMA 5.5<br />

The estimate<br />

I1 =<br />

<br />

M0<br />

V (a 2 + a 2 )dvg + hCpV (p)|a(p)| 2 Re(e 2iψ(p)/h ) + o(h)<br />

holds true with Cp = 0 and is independent of h.<br />

With Lemmas 5.3 and 5.4, we can write (18) as<br />

<br />

0 = V (a 2 + a 2 )dvg + O(h)<br />

M0<br />

and thus we can conclude that<br />

<br />

0 =<br />

M0<br />

M0<br />

V (a 2 + a 2 )dvg.<br />

Therefore, (18) becomes, with Lemma 5.3,<br />

0 = CpV (p)|a(p)| 2 Re(e 2iψ(p)/h <br />

) + 2 V Re a(b0 + ˜s12 + a0 + ˜r12) dvg + o(1).<br />

M0<br />

Since ψ(p) = 0, we may choose a sequence of hj → 0 such that Re(e 2iψ(p)/hj ) = 1<br />

and another sequence ˜hj → 0 such that Re(e 2iψ(p)/˜hj ) =−1 for all j. Adding the<br />

expansion with h = hj and h = ˜hj, we deduce that<br />

0 = 2CpV (p)|a(p)| 2 + o(1)<br />

as j →∞, and since Cp = 0, a(p) = 0, we conclude that V (p) = 0. The set<br />

of p ∈ M0 for which we can conclude this is dense in M0, by Proposition 2.3.1.<br />

Therefore, we can conclude that V (p) = 0 for all p ∈ M0. <br />

We now prove Lemma 5.5.


110 GUILLARMOU and TZOU<br />

Proof of Lemma 5.5<br />

Let χ be a smooth cutoff function on M0 which is identically 1 everywhere except<br />

outside a small ball containing p and no other critical point of , andletχ = 0 near<br />

p. We split the oscillatory integral into two parts:<br />

<br />

(e 2iψ/h + e −2iψ/h )V |a| 2 <br />

dvg = χ(e 2iψ/h + e −2iψ/h )V |a| 2 dvg<br />

M0<br />

M0<br />

<br />

+<br />

M0<br />

(1 − χ)(e 2iψ/h + e −2iψ/h )V |a| 2 dvg.<br />

The phase ψ has nondegenerate critical points. Therefore, a standard application of<br />

the stationary phase at p gives<br />

<br />

(1 − χ)(e 2iψ/h + e −2iψ/h )V (p)|a| 2 dvg = hCp|a(p)| 2 V (p)Re(e 2iψ(p)/h ) + o(h),<br />

M0<br />

where Cp is a nonzero number which depends on the Hessian of ψ at the point p.<br />

Define the potential V (·) := V (·) − V (p) ∈ C1,α (M0). Then we show that<br />

<br />

(1 − χ)(e 2iψ/h + e −2iψ/h )V |a| 2 dvg = o(h). (21)<br />

M0<br />

Indeed, first by integration by parts and using gψ = 0, one has<br />

<br />

(1 − χ)(e 2iψ/h + e −2iψ/h )V |a| 2 dvg<br />

M0<br />

= h<br />

2i<br />

= h<br />

2i<br />

<br />

<br />

M0<br />

M0<br />

〈d(e 2iψ/h − e −2iψ/h (1 − χ)|a|2<br />

),dψ〉V<br />

|dψ| 2<br />

dvg<br />

(e 2iψ/h − e −2iψ/h <br />

(1 − χ)|a| 2V<br />

) d<br />

|dψ| 2<br />

<br />

,dψ dvg.<br />

But we can see that 〈d((1 − χ)|a| 2V/| dψ| 2 ),dψ〉∈L1 (M0): this follows directly<br />

from the fact that V is in the Hölder space C1,α (M0) and V (p) = 0, and from the<br />

nondegeneracy of Hess(ψ). It then suffices to use Lemma 5.4 to conclude that (21)<br />

holds. Using a similar argument, we now show that<br />

<br />

χ(e 2iψ/h + e −2iψ/h )V |a| 2 dvg = o(h).<br />

M0


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 111<br />

Indeed, since a vanishes to large order at all boundary critical points of ψ, we may<br />

write<br />

<br />

χ(e 2iψ/h + e −2iψ/h )V |a| 2 dvg<br />

M0<br />

= h<br />

<br />

〈d(e<br />

2i M0<br />

2iψ/h − e −2iψ/h ),dψ〉V χ|a|2<br />

dvg<br />

|dψ| 2<br />

=− h<br />

<br />

(e<br />

2i M0<br />

2iψ/h − e −2iψ/h <br />

)divg V χ|a|2<br />

|dψ| 2 ∇g <br />

ψ dvg<br />

+ h<br />

<br />

(e<br />

2i<br />

2iψ/h − e −2iψ/h )V |a|2<br />

|dψ| 2 ∂νψ dvg.<br />

∂M0<br />

For the interior integral, we use Lemma 5.4 to conclude that<br />

− h<br />

<br />

(e<br />

2i M0<br />

2iψ/h − e −2iψ/h <br />

)divg V χ|a|2<br />

|dψ| 2 ∇g <br />

ψ dvg = o(h),<br />

and for the boundary integral we write ∂M0 = Ɣ0 ∪ Ɣ and observe that on Ɣ0, ψ = 0,<br />

so (e2iψ/h−e−2iψ/h ) = 0, while on Ɣ we have V = 0 from the boundary determination<br />

proved in Proposition A.1 of the appendix. Therefore,<br />

<br />

χ(e 2iψ/h + e −2iψ/h )V |a| 2 dvg = o(h),<br />

M0<br />

and the proof is complete. <br />

6. Inverse scattering<br />

We first obtain, as a trivial consequence of Theorem 1.1, a result about inverse<br />

scattering for asymptotically hyperbolic (AH) surfaces. Recall that an AH surface<br />

is an open complete Riemannian surface (X, g) such that X is the interior of a<br />

smooth compact surface with boundary ¯X, and for any smooth boundary-defining<br />

function x of ∂ ¯X, ¯g := x 2 g extends as a smooth metric to ¯X, with curvature tending<br />

to −1 at ∂ ¯X. IfV ∈ C ∞ ( ¯X) and if V = O(x 2 ), then we can define a scattering<br />

map as follows (see [20], [13] or[14]). First, the L 2 kernel kerL 2(g + V ) is<br />

a finite-dimensional subspace of xC ∞ ( ¯X) and in one-to-one correspondence with<br />

E := {(∂xψ)| ∂ ¯X; ψ ∈ kerL 2(g + V )}, where ∂x := ∇ ¯g x is the normal vector field to<br />

∂ ¯X for ¯g. Then, for f ∈ C ∞ (∂ ¯X), there exists a function u ∈ C ∞ ( ¯X), unique modulo<br />

kerL 2(g + V ), such that (g + V )u = 0 and u| ∂ ¯X = f . Then one can see that the<br />

scattering map S : C ∞ (∂ ¯X) → C ∞ (∂ ¯X)/E is defined by Sf := ∂xu| ∂ ¯X. We thus<br />

obtain the following.


112 GUILLARMOU and TZOU<br />

COROLLARY 6.1<br />

Let (X, g) be an asymptotically hyperbolic manifold, and let V1,V2 ∈ x 2 C ∞ ( ¯X) be<br />

two potentials and Ɣ ⊂ ∂ ¯X an open subset of the conformal boundary. Assume that<br />

∂xu ∂ ¯X ; u ∈ kerL 2(g + V1) = ∂xu ∂ ¯X ; u ∈ kerL 2(g + V2) ,<br />

and let Sj be the scattering map for the operator g + Vj for j = 1, 2.IfS1f = S2f<br />

on Ɣ for all f ∈ C ∞ 0 (Ɣ), thenV1 = V2.<br />

Proof<br />

Let x be a smooth boundary-defining function of ∂ ¯X,let¯g = x 2 g be the compactified<br />

metric, and define ¯Vj := Vj/x 2 ∈ C ∞ ( ¯X). By conformal invariance of the Laplacian<br />

in dimension 2, one has<br />

g + Vj = x 2 (¯g + ¯Vj),<br />

and so if kerL 2(g + V1) = kerL 2(g + V2) and if S1 = S2 on Ɣ, then the Cauchy<br />

data spaces C Ɣ i for the operator ¯g + ¯Vj are the same. Then it suffices to apply the<br />

result in Theorem 1.1. <br />

Next we consider the asymptotically Euclidean scattering at zero frequency. An<br />

asymptotically Euclidean surface is a noncompact Riemann surface (X, g) which<br />

compactifies into a smooth surface ¯X with boundary such that the metric, in a collar<br />

neighborhood (0,ɛ)x × ∂ ¯X of the boundary, is of the form<br />

g = dx2 h(x)<br />

+<br />

x4 x<br />

where h(x) is a smooth 1-parameter family of metrics on ∂ ¯X such that h(0) = dθ2 S1 is<br />

the metric. Notice that using the coordinates r := 1/x, g is asymptotic to dr2 +r 2dθ2 S1 near r →∞. A particular case is given by the surfaces with Euclidean ends, that<br />

is, ends isometric to R2 \ B(0,R), where B(0,R) ={z ∈ R2 ; |z| ≥R}. Note that<br />

g is conformal to an asymptotically cylindrical metric (or “b-metric” in the sense of<br />

Melrose [27])<br />

2 ,<br />

gb := x 2 g = dx2<br />

+ h(x)<br />

x2 and that the Laplacian satisfies g = x2gb . Each end of X is of the form (0,ɛ)x ×S 1 θ ,<br />

and the operator gb has the expression in the ends<br />

gb =−(x∂x) 2 + ∂ ¯X + xP(x,θ; x∂x,∂θ)


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 113<br />

for some smooth differential operator P (x,θ; x∂x,∂θ) in the vector fields x∂x,∂θ<br />

down to x = 0. Let us define Vb := x −2 V , which is compactly supported, and<br />

H 2m<br />

b := u ∈ L 2 (X, dvolgb); m<br />

gb u ∈ L2 (X, dvolgb) , m ∈ N0.<br />

We also define the following spaces for α ∈ R:<br />

Fα := ker x α H 2 b (gb + Vb).<br />

Since the eigenvalues of S 1 are {j 2 ; j ∈ N0}, the relative index theorem of Melrose<br />

[27, Section 6.2] shows that gb + Vb is Fredholm from xαH 2 b to xαH 0 b if α /∈ Z.<br />

Moreover, from [27, Section 2.2.4], we have that any solution of (gb + Vb)u = 0 in<br />

xαH 2 b has an asymptotic expansion of the form<br />

u ∼ <br />

ℓj <br />

j>α,j∈Z ℓ=0<br />

x j (log x) ℓ uj,ℓ(θ) as x → 0<br />

for some sequence (ℓj)j of nonnegative integers and some smooth function uj,ℓ on<br />

S 1 . In particular, it is easy to check that kerL 2 (X,dvolg)(g + V ) = F1+ɛ for ɛ ∈ (0, 1).<br />

<strong>THEOREM</strong> 6.2<br />

Let (X, g) be an asymptotically Euclidean surface, let V1,V2 be two compactly supported<br />

smooth potentials, and let x be a boundary-defining function. Let ɛ ∈ (0, 1),<br />

and assume that, for any j ∈ Z and any function ψ ∈ ker x j−ɛ H 2 b (g + V1), thereisa<br />

ϕ ∈ ker x j−ɛ H 2 b (g + V2) such that ψ − ϕ = O(x ∞ ), and conversely. Then V1 = V2.<br />

Proof<br />

The idea is to reduce the problem to the compact case. First, we notice that by unique<br />

continuation, ψ = ϕ where V1 = V2 = 0. Now it remains to prove that, if Rη denotes<br />

the restriction of smooth functions on X to {x ≥ η} and if V is a smooth compactly<br />

supported potential in {x ≥ η}, then the set ∞ j=0 Rη(F−j−ɛ) is dense in the set NV<br />

of H 2 ({x ≥ η}) solutions of (g + V )u = 0. The proof is well known for positive<br />

frequency scattering (see, e.g., [28, Lemma 3.2]). Here it is very similar, so we do not<br />

give details. The main argument is to show that it converges in the L2 sense and then<br />

uses elliptic regularity; the L2 convergence can be shown as follows. Let f ∈ NV<br />

such that<br />

<br />

∞<br />

fψdvolg = 0, ∀ ψ ∈ F−j−ɛ.<br />

x≥η<br />

Then we want to show that f = 0. By[27, Proposition 5.64], there exists k ∈ N<br />

and a generalized right inverse Gb for Pb = gb + Vb (here, as before, x 2 Vb = V )<br />

j=0


114 GUILLARMOU and TZOU<br />

in x−k−ɛ H 2 b such that PbGb = Id. This holds in x−k−ɛ H 2 b for k large enough since<br />

the cokernel of Pb on this space becomes zero for k large. Let ω = Gbf so that<br />

(gb + Vb)ω = f ; in particular, this function is zero in {x n,<br />

while Brown [5] studied the case of Lipschitz domains with a continuous conductivity.<br />

Since the result in our setting is not explicitly written down but is certainly known<br />

from specialists, we provide a short proof with few details using the approach of [5].<br />

We prove the following.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 115<br />

PROPOSITION A.1<br />

Let Ɣ ⊂ ∂M0 be a nonempty open subset of the boundary. If V1,V2 ∈ C0,α (M0) for<br />

some α>0 and if their associated Cauchy data spaces C Ɣ 1 , C Ɣ 2 defined in (1) are<br />

equal, then V1|Ɣ = V2|Ɣ.<br />

The key to proving this proposition is the existence of solutions to (g + Vi)u = 0,<br />

which concentrate near a point p ∈ Ɣ. First, we need a solvability result for the<br />

equation (g + Vi)u = f , which is an easy consequence of the Carleman estimate<br />

of Proposition 3.1, and follows the method of Salo and Tzou [31, Section 6]. If we<br />

fix h>0 small and take ϕ = 1 in the Carleman estimate of Proposition 3.1, we<br />

easily obtain that there is a constant C such that for all functions in H 2 (M0) satisfying<br />

u|∂M0 = 0,<br />

u 2<br />

2<br />

H 2 +∂νuL2 (Ɣ0)<br />

2<br />

2<br />

≤ C(( + Vi)uL2 +∂νuL2 (Ɣ) ). (A.1)<br />

As a consequence, we deduce the following solvability result. Let<br />

B := w ∈ H 2 (M0) ∩ H 1<br />

0 (M0) | ∂νw|Ɣ = 0 <br />

be the closed subspace of H 2 (M0) under the H 2 norm, and let B ∗ be its dual space.<br />

Then we have the following.<br />

COROLLARY A.2<br />

Let i = 1, 2. Then for all f ∈ L 2 (M0) there exists u ∈ H 2 (M0) solving the equation<br />

(g + Vi)u = f<br />

with boundary condition u|Ɣ0 = 0, and uL2 ≤ Cf B∗. Proof<br />

Set A := {w ∈ H 1 0 (M0) | (g + Vi)w ∈ L2 ,∂νw|Ɣ = 0} equipped with the inner<br />

product<br />

<br />

(v, w)A := (g + Vi)v(g + Vi)w dvg.<br />

M0<br />

Thanks to (A.1), A is a Hilbert space and A = B. For each f ∈ L 2 (M0), let us define<br />

the linear functional on B:<br />

By (A.1), we have, for all w ∈ B,<br />

<br />

Lf : w ↦→<br />

M0<br />

wf dvg.<br />

|Lf (w)| ≤f B ∗wB ≤f B ∗wA.


116 GUILLARMOU and TZOU<br />

Therefore, by the Riesz theorem, there exists vf ∈ A such that<br />

<br />

<br />

(g + V )vf (g + Vi)w dvg = wf dvg<br />

M0<br />

for all w ∈ B. Furthermore, (g + Vi)vf L2 ≤fB∗. Setting u := (g + Vi)vf ,<br />

we have (g + Vi)u = f and uL2 ≤fB∗. To obtain the boundary condition for<br />

u, observe that since<br />

<br />

M0<br />

<br />

u(g + Vi)wdvg =<br />

for all w ∈ B, then by Green’s theorem,<br />

<br />

<br />

u∂νw dvg = 0 =<br />

M0<br />

Ɣ0<br />

M0<br />

M0<br />

fwdvg<br />

u∂νw dvg<br />

for all w ∈ B. Thisimpliesthatu = 0 on Ɣ0. <br />

Clearly, it suffices to assume that Ɣ is a small piece of the boundary which is contained<br />

in a single coordinate chart with complex coordinates z = x + iy, where |z| ≤1,<br />

Im(z) > 0, and the boundary is given by {y = 0}. Moreover, the metric is of the<br />

form e 2ρ |dz| 2 for some smooth function ρ.Letp ∈ Ɣ, and possibly by translating the<br />

coordinates, we can assume that p ={z = 0}. Letη ∈ C ∞ (M0) be a cutoff function<br />

supported in a small neighborhood of p. Forh>0 small, we define the smooth<br />

function vh ∈ C ∞ (M0) supported near p via the coordinate chart Z = (x,y) ∈ R 2 by<br />

vh(Z) := η(Z/ √ h)e (1/h)α·Z , (A.2)<br />

where α := (i, −1) ∈ C 2 is chosen such that α · α = 0 and the dot product · means<br />

the canonical symmetric C-bilinear form on C 2 . Thus we get (∂ 2 x + ∂2 y )eα·Z = 0 and<br />

thus ge α·Z = 0 by conformal covariance of the Laplacian. Therefore, we have in<br />

local coordinates<br />

gvh(Z) = 1<br />

h e(1/h)α·Z (gη)<br />

<br />

Z√h<br />

<br />

+ 2<br />

<br />

Z√h<br />

<br />

e(1/h)α·Z dη ,α· dZ . (A.3)<br />

h3/2 g<br />

LEMMA A.3<br />

If V ∈ C 0,α (M0) for q>2, then there exists a solution uh ∈ H 2 to (g + V )u = 0<br />

of the form<br />

uh = vh + Rh,<br />

with vh defined in (A.2) and RhL 2 ≤ Ch5/4 , satisfying supp(Rh|∂M0) ⊂ Ɣ.


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 117<br />

Proof of Lemma A.3<br />

We need to find Rh satisfying RhL 2 ≤ Ch5/4 and solving<br />

(g + V )Rh =−(g + V )vh =: Mh.<br />

Thanks to Corollary A.2, it suffices to show that MhB ∗ ≤ Ch5/4 . Thus, let w ∈ B;<br />

then we have, by (A.3),<br />

<br />

wMh dvg = I1 + I2 + I3,<br />

where<br />

<br />

I1 : =<br />

I2 : = 1<br />

h<br />

M0<br />

|Z|≤ √ h<br />

<br />

I3 : = 2<br />

h 3/2<br />

wVivhe 2ρ dZ,<br />

|Z|≤ √ we<br />

h<br />

(1/h)α·Z χ1<br />

<br />

|Z|≤ √ wχ2<br />

h<br />

<br />

Z√h<br />

<br />

e 2ρ dZ,<br />

<br />

Z√h<br />

<br />

e (1/h)α·Z dZ,<br />

and χ1 = gη, χ2 = i∂xη − ∂yη. In the above equation, the term I3 has the worst<br />

growth when h → 0. We analyze its behavior, and the preceding terms can be treated<br />

in similar fashion. One has<br />

I3 =−h 1/2<br />

=−h 1/2<br />

<br />

<br />

|Z|≤ √ wχ2<br />

h<br />

|Z|≤ √ (∂x − i∂y)<br />

h<br />

2<br />

<br />

Z√h<br />

<br />

(∂x − i∂y) 2 e (1/h)α·Z dZ<br />

<br />

Z√h<br />

<br />

wχ2 e (1/h)α·x dZ.<br />

Notice that the boundary term in the integration by parts vanishes because w ∈ H 1 0<br />

and ∂νw|∂M0 vanishes on the support of η. The term (∂x − i∂y) 2 (wχ2(Z/ √ h)) has<br />

derivatives hitting both χ2(Z/ √ h) and w. The worst growth in h would occur when<br />

both derivatives hit χ2(Z/ √ h), in which case a h −1 factor would come out. Combined<br />

with the h 1/2 term in front of the integral this gives a total of a h −1/2 in front. By this<br />

observation we have improved the growth from h −3/2 to h −1/2 . Repeating this line of argument<br />

and using the Cauchy-Schwarz inequality, we can see that |I3| ≤Ch 5/4 wH 2<br />

(an elementary computation shows that functions of the form χ(Z/ √ h)e (1/h)α·Z have<br />

L 2 norm bounded by Ch 3/4 ). Therefore, (1/h 3/2 )χ2(Z/ √ h)e (1/h)α·Z B ′ ≤ Ch5/4 ,<br />

and we are done.


118 GUILLARMOU and TZOU<br />

Proof of Proposition A.1<br />

It suffices to plug the solutions u1 h ,u2 h from Proposition A.3 into the boundary integral<br />

identity (17). A simple calculation using the fact that V1 − V2 is in C0,γ (M0) yields<br />

<br />

0 = u 1<br />

h (V1 − V2)u 2<br />

h dvg = Ch 3/2 V1(p) − V2(p) + o(h 3/2 ),<br />

M0<br />

and we are done. <br />

Acknowledgments. This work started during a summer evening in Pisa thanks to the<br />

hospitality of M. Mazzucchelli and A. G. Lecuona. We thank Sam Lisi, Rafe Mazzeo,<br />

Mikko Salo, and Eleny Ionel for pointing out very helpful references. The first author<br />

thanks the Mathematical Sciences Research Institute and the organizers of the 2008<br />

program “Analysis on Singular Spaces” for support during part of this project. Part of<br />

this work was also done while the first author was at the Institute for Advanced Study<br />

in Princeton, and he acknowledges the hospitality extended to him.<br />

References<br />

[1] G. ALESSANDR<strong>IN</strong>I, Singular solutions of elliptic equations and the determination of<br />

conductivity by boundary measurements, J. Differential Equations 84 (1990),<br />

252 – 272. MR 1047569 114<br />

[2] K. ASTALA and L. PÄIVÄR<strong>IN</strong>TA, Calderón’s inverse conductivity problem in the plane,<br />

Ann. of Math. (2) 163 (2006), 265 – 299. MR 2195135 84<br />

[3] K. ASTALA, L. PÄIVÄR<strong>IN</strong>TA, and M. LASSAS, Calderón’s inverse problem for<br />

anisotropic conductivity in the plane, Comm. Partial Differential Equations 30<br />

(2005), 207 – 224. MR 2131051 84<br />

[4] M. I. BELISHEV, The Calderon problem for two-dimensional manifolds by the<br />

BC-method, SIAM J. Math. Anal. 35 (2003), 172 – 182. MR 2001471 84<br />

[5] R. M. BROWN, Recovering the conductivity at the boundary from the Dirichlet to<br />

Neumann map: A pointwise result, J. Inverse Ill-Posed Probl. 9 (2001), 567 – 574.<br />

MR 1881563 114<br />

[6] A. L. BUKHGEIM, Recovering a potential from Cauchy data in the two-dimensional<br />

case, J. Inverse Ill-Posed Probl. 16 (2008), 19 – 33. MR 2387648 84, 85<br />

[7] A. L BUKHGEIM and G. UHLMANN, Recovering a potential from partial Cauchy data,<br />

Comm. Partial Differential Equations 27 (2002), 653 – 668. MR 1900557 84<br />

[8] A. P. CALDERÓN, “On an inverse boundary value problem” in Seminar on Numerical<br />

Analysis and its Applications to Continuum Physics, (Rio de Janeiro, 1980), Soc.<br />

Brasileira de Matemática 12, Soc. Brasileira de Matemática, Rio de Janeiro, 1980,<br />

65 – 73. MR 0590275 83<br />

[9] D. DOS SANTOS FERREIRA, C. E. KENIG, M. SALO,andG. UHLMANN, Limiting<br />

Carleman weights and anisotropic inverse problems, Invent. Math. 178 (2009),<br />

119 – 171. MR 2534094 84


CALDERÓN <strong>IN</strong>VERSE PROBLEM ON RIEMANN SURFACES 119<br />

[10] L. D. FADDEEV, Increasing solutions of the Schrödinger equation (in Russian), Dokl.<br />

Akad. Nauk SSSR 165 (1965), 514 – 517; English translation in Sov. Phys. Dokl.<br />

10 (1966), 1033 – 1035. 85, 100<br />

[11] H. M FARKAS and I. KRA, Riemann Surfaces, 2nd ed., Grad. Texts in Math. 71,<br />

Springer, New York, 1992. MR 1139765 86<br />

[12] I. M. GELFAND, “Some aspects of functional analysis and algebra” in Proceedings of<br />

the International Congress of Mathematicians (Amsterdam, 1954),<br />

North-Holland, Amsterdam, 253 – 276. MR 0095423 83<br />

[13] C. R. GRAHAM and M. ZWORSKI, Scattering matrix in conformal geometry, Invent.<br />

Math. 152 (2003), 89 – 118. MR 1965361 111<br />

[14] C. GUILLARMOU and L. GUILLOPÉ, The determinant of the Dirichlet-to-Neumann map<br />

for surfaces with boundary, Int. Math. Res. Not. IMRN 2007, no. 22, art. ID<br />

rnm099. MR 2376211 111<br />

[15] C. GUILLARMOU and A. SÁ BARRETO, Inverse problems for Einstein manifolds,<br />

Inverse Probl. Imaging 3 (2009), 1 – 15. MR 2558301 84<br />

[16] C. GUILLARMOU and L. TZOU,“Calderón inverse problem for the Schrödinger<br />

operator on Riemann surfaces” in The AMSI-ANU Workshop on Spectral Theory<br />

and Harmonic Analysis (Canberra, 2009), Proc. Centre Math. Appl. Austral. Nat.<br />

Univ. 44, Austral. Nat. Univ., Canberra, 2010, 129 – 141. MR 2655386 84<br />

[17] G. HENK<strong>IN</strong> and V. MICHEL, Inverse conductivity problem on Riemann surfaces, J.<br />

Geom. Anal. 18 (2008), 1033 – 1052. MR 2438910 84, 85<br />

[18] G. HENK<strong>IN</strong> and R. G. NOVIKOV, On the reconstruction of conductivity of bordered<br />

two-dimensional surface in R 3 from electrical currents measurements on its<br />

boundary, to appear in J. Geom. Anal., preprint, arXiv:1003.4897v1 [math.AP]<br />

85<br />

[19] O. Y. IMANUVILOV, G. UHLMANN, andM. YAMAMOTO, The Calderón problem with<br />

partial data in two dimensions, J.Amer.Math.Soc.23 (2010), 655 – 691. 84, 85,<br />

86, 95, 96, 103, 105<br />

[20] M. S. JOSHI and A. SÁ BARRETO, Inverse scattering on asymptotically hyperbolic<br />

manifolds, Acta Math. 184 (2000), 41 – 86. MR 1756569<br />

[21] C. E. KENIG, J. SJÖSTRAND,andG. UHLMANN, The Calderón problem with partial<br />

data, Ann. of Math. (2) 165 (2007), 567 – 591. MR 2299741 111 84<br />

[22] R. KOHN and M. VOGELIUS, Determining conductivity by boundary measurements,<br />

Comm. Pure Appl. Math. 37 (1984), 289 – 298. MR 0739921 114<br />

[23] M. LASSAS, M. TAYLOR,andG. UHLMANN, The Dirichlet-to-Neumann map for<br />

complete Riemannian manifolds with boundary, Comm. Anal. Geom. 11 (2003),<br />

207 – 222. MR 2014876 84<br />

[24] M. LASSAS and G. UHLMANN, On determining a Riemannian manifold from the<br />

Dirichlet-to-Neumann map. Ann. Sci. Éc. Norm. Supér. (4) 34 (2001), 771 – 787.<br />

MR 1862026 84<br />

[25] R. MAZZEO and M. TAYLOR, Curvature and uniformization, Israel J. Math. 130 (2002),<br />

323 – 346. MR 1919383 95<br />

[26] D. MCDUFF and D. SALAMON, J -Holomorphic curves and symplectic topology, Amer.<br />

Math. Soc. Colloq. Publ. 52, Amer. Math. Soc., Providence, 2004. MR 2045629<br />

87, 88


120 GUILLARMOU and TZOU<br />

[27] R. B. MELROSE, The Atiyah-Patodi-Singer Index Theorem, Res. Notes Math. 4,A.K.<br />

Peters, Wellesley, 1993. MR 1348401 112, 113, 114<br />

[28] ———, Geometric Scattering Theory, Cambridge Univ. Press, Cambridge, 1995.<br />

MR 1350074 113<br />

[29] A. NACHMAN, Reconstructions from boundary measurements, Ann. of Math. (2) 128<br />

(1988), 531 – 576. MR 0970610 84<br />

[30] R. G. NOVIKOV, A multidimensional inverse spectral problem for the equation<br />

−ψ + (v(x) − Eu(x))ψ = 0 (in Russian), Funktsional. Anal. i Prilozhen. 22<br />

(1988), no. 4, 11 – 22, 96; English translation in Funct. Anal. Appl. 22 (1988), no.<br />

4, 263 – 272 (1989). MR 0976992 84<br />

[31] M. SALO and L. TZOU, Carleman estimates and inverse problems for Dirac operators,<br />

Math. Ann. 344 (2009), 161 – 184. MR 2481057 115<br />

[32] J. SYLVESTER and G. UHLMANN, A global uniqueness theorem for an inverse boundary<br />

value problem, Ann. of Math. (2) 125 (1987), 153 – 169. MR 0873380 84<br />

[33] ———, Inverse boundary value problems at the boundary—continuous dependence,<br />

Comm. Pure Appl. Math. 41 (1988), 197 – 219. MR 0924684 114<br />

[34] K. UHLENBECK, Generic properties of eigenfunctions, Amer. J. Math. 98 (1976),<br />

1059 – 1078. MR 0464332 91, 92<br />

Guillarmou<br />

École Normale Supérieure, UMR CNRS 8553, F 75230 Paris CEDEX 05, France;<br />

cguillar@dma.ens.fr<br />

Tzou<br />

Department of Mathematics, Stanford University, Stanford, California 94305, USA;<br />

ltzou@math.stanford.edu


QUANTITATIVE VERSION OF THE OPPENHEIM<br />

CONJECTURE FOR <strong>IN</strong>HOMOGENEOUS<br />

QUADRATIC FORMS<br />

GREGORY MARGULIS and AMIR MOHAMMADI<br />

Abstract<br />

We prove a quantitative version of the Oppenheim conjecture for inhomogeneous<br />

quadratic forms. We also give an application to eigenvalue spacing on flat 2-tori with<br />

Aharonov-Bohm flux.<br />

Contents<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

2. Passage to space of inhomogeneous lattices . . . . . . . . . . . . . . . 128<br />

3. The case where p ≥ 3 . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

4. The case of signature (2, 2) . . . . . . . . . . . . . . . . . . . . . . . . 132<br />

5. Contribution from quasi-null subspaces . . . . . . . . . . . . . . . . . 136<br />

6. Proof of Theorem 4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

7. Proof of Theorem 1.10 . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

A. Equidistribution of spherical averages . . . . . . . . . . . . . . . . . . . 153<br />

B. Number of quasi-null subspaces . . . . . . . . . . . . . . . . . . . . . . 157<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

1. Introduction<br />

Let Q be a nondegenerate indefinite quadratic form on R n . Let ξ ∈ R n beavector,<br />

and define the (inhomogeneous) quadratic form Qξ by<br />

Qξ(x) = Q(x + ξ) for all x ∈ R n . (1)<br />

We will refer to Q = Q0 as the homogeneous part of Qξ. We say that Qξ has signature<br />

(p, q) if Q does. Recall that a quadratic form Qξ is called irrational if it is not a scalar<br />

DUKE MATHEMATICAL JOURNAL<br />

Vol. 158, No. 1, c○ 2011 DOI 10.1215/00127094-1276319<br />

Received 4 March 2010. Revision received 22 July 2010.<br />

2010 Mathematics Subject Classification. Primary 11E20; Secondary 58J50.<br />

Margulis’s work partially supported by National Science Foundation grant DMS-0801195.<br />

121


122 MARGULIS and MOHAMMADI<br />

multiple of a form with rational coefficients. In other words, Qξ is irrational if either Q<br />

is irrational as a homogeneous form, or, if Q is a rational form, then ξ is an irrational<br />

vector.<br />

Let ν be a continuous function on the sphere {v ∈ R n : v =1}. Define<br />

={v ∈ R n : v


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 123<br />

Recall from [EMM2]thatifQ is irrational of signature (2, 2), then it has at most four<br />

rational null subspaces. Let<br />

ÑQ,(a,b,T ) = #<br />

<br />

x ∈ Z n :<br />

Eskin, Margulis, and Mozes proved the following.<br />

x is not in a null subspace of Q<br />

<br />

. (6)<br />

x ∈ T and a


124 MARGULIS and MOHAMMADI<br />

As in [EMM1], we also have the following uniform version of Theorem 1.5. Let<br />

I(p, q) denote the space of inhomogeneous quadratic forms whose homogeneous<br />

parts are quadratic forms of signature (p, q) and discriminant ±1.<br />

<strong>THEOREM</strong> 1.6<br />

Let D be a compact subset of I(p, q) with p ≥ 3 and q ≥ 1, and let n = p + q.<br />

Then for every interval (a,b) and every θ>0, there exists a finite subset P of D<br />

such that each Qξ ∈ P is a rational form and, for every compact subset F ⊂ D \ P ,<br />

there exists T0 such that, for all Qξ ∈ F and T ≥ T0, we have<br />

(1 − θ)λQ,(b − a)T n−2 ≤ NQ,ξ,(a,b,T ) ≤ (1 + θ)λQ,(b − a)T n−2 , (10)<br />

where λQ, is as in (3).<br />

The proofs of the above theorems are straightforward inhomogeneous versions of the<br />

arguments and ideas developed in [DM] and[EMM1].<br />

As we mentioned before, Theorem 1.5 fails in signature (2, 2) and (2, 1). In this,<br />

paper, we prove an inhomogeneous version of Theorem 1.3. Indeed,asin[EMM2],<br />

one needs to assume certain Diophantine conditions on the quadratic form. Using<br />

similar ideas, we also give a partial result in the (2, 1) case (see Theorem 1.10 below).<br />

We start with the following definitions.<br />

Definition 1.7<br />

Let κ>0. A vector ξ = (ξ1,...,ξn) ∈ R n is called κ-Diophantine if there exist C =<br />

C(ξ) > 0 such that, for all 0


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 125<br />

then there are at most four subspaces L for which the above can hold. Let<br />

ÑQ,ξ,(a,b,T ) = #<br />

<br />

x ∈ Z n :<br />

x is not in an exceptional subspace of Qξ<br />

x ∈ T and a


126 MARGULIS and MOHAMMADI<br />

planes” interfere with each other. This fact prevents us from successfully using the<br />

ideas in [EMM2] when the signature is (2, 1). The absence of such an asymptotic<br />

in the homogeneous case with explicit Diophantine conditions is responsible for the<br />

“extra” assumption (i) in Theorem 1.10. Indeed, using this assumption, we are able to<br />

reduce the study to an investigation of the contribution from null vectors. We are then<br />

able to impose and utilize an explicit Diophantine condition on ξ (see Section 7 for<br />

details).<br />

Let us also mention that the Diophantine conditions in Theorems 1.10 and 1.9<br />

are necessary. Similar to [EMM1, Theorem 2.2], [Mark, Theorem 1.3], and [Sar,<br />

Theorem 2], it is possible to construct sets of second Baire category of quadratic<br />

forms for which the above fail. To be more explicit, it is shown in [EMM1], using<br />

category arguments, that there exists a dense set of second Baire category γ such<br />

that NQ,(1/8, 2,Ti) >Ti(log Ti) 1−ε for an infinite sequence Ti, where Qγ (x) =<br />

x2 1 + x2 2 − γ 2 (x2 3 + x2 4 ). Nowifwetakeanyξ∈ Z4 , then the same holds for Q γ<br />

ξ .<br />

This already gives the examples we wanted. However, these examples are in a sense<br />

“degenerate" since ξ ∈ Z4 . We will reproduce the argument in [EMM1] to get more<br />

“generic" counterexamples. First, note that an argument like that in [Mark, Appendix<br />

10] shows that, for any (α, ς) ∈ Q × Q4 ,wehave<br />

# v ∈ Z 4 : Q α<br />

ς (v) = 0 and v ≤T ∼ Cς,αT 2 log T. (15)<br />

Now the same argument as in [EMM1, Lemma 3.15] (see also [Mark, Section 9])<br />

gives: for every ε>0 and any S>0, the set (β,ζ) ∈ R × R 4 for which there exists<br />

T>Ssuch that NQ β ,ζ,(1/4, 1,T) ≥ T 2 (log T ) 1−ε is dense.<br />

Given that S>0, letUS be the set of ordered pairs (γ,ξ) ∈ R × R 4 for which<br />

there exists (β,ζ) ∈ R × R 4 and T>Ssuch that<br />

NQ β ,ζ,(1/4, 1,T) ≥ T 2 (log T ) 1−ε , |β − γ |


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 127<br />

Let the Hamiltonian be the geodesic flow for flat 2-tori. Sarnak [Sar] proved that,<br />

for almost all 2-dimensional flat tori (with respect to Lebesgue measure on the moduli<br />

space of 2-dimensional flat tori), the pair correlation density function converges to the<br />

pair correlation density of a Poisson process. He used averaging arguments to reduce<br />

the pair correlation problem to one of spacing between the values at integers of binary<br />

quadratic forms. This is related to the quantitative Oppenheim problem in the case of<br />

signature (2, 2). One corollary of Theorem 1.3 is that the Berry-Tabor conjecture holds<br />

for the pair correlation of 2-dimensional flat tori under certain explicit Diophantine<br />

conditions.<br />

Similarly, Theorem 1.9 has a corollary in this direction. Let h be a lattice in R 2 .<br />

Also, let α = (α1,α2) ∈ R 2 . Now the eigenvalues of the Laplacian<br />

− =− ∂2 ∂2<br />

−<br />

∂y2 ∂y2 with quasi-periodicity conditions<br />

(17)<br />

φ(x + v) = e 2πi〈α,v〉 φ(x) for all x ∈ R 2 and all v ∈ h (18)<br />

are of the form 4π 2 w + α 2 , where w ∈ h ∗ and where h ∗ is the dual lattice to h. Let<br />

0 ≤ λ0


128 MARGULIS and MOHAMMADI<br />

α ∈ R 2 be given, and define β as above. Suppose that at least one of the following<br />

holds.<br />

(i) The vector β = (β1,β2) is Diophantine.<br />

(ii) There exists N,C > 0 such that, for all triples of integers (p1,p2,q) with<br />

q ≥ 2, we have<br />

<br />

<br />

max Ai −<br />

i=1,2<br />

pi<br />

<br />

<br />

><br />

q<br />

C<br />

q<br />

Then for any interval (a,b) with 0 /∈ (a,b), we have<br />

lim<br />

T →∞ Rh,α(a,b,T ) = c 2<br />

h (b − a). (22)<br />

Hence the spectrum satisfies the Berry-Tabor conjecture for the pair correlation<br />

function.<br />

The case of h = Z 2 was proved by Marklof [Mark]. His approach utilizes results from<br />

the theory of unipotent flows combined with an application of theta sums. We also use<br />

the theory of unipotent flows in our proof; however, our strategy to control the integral<br />

of unbounded functions over certain orbits is dynamical and rests heavily on [EMM1]<br />

and [EMM2].<br />

Outline of the proof. Let Qξ be a quadratic form of signature (p, q),andletn = p+q.<br />

Fix an interval (a,b), andletU⊂Rnbe a “suitably chosen" compact set such that<br />

a


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 129<br />

Quadratic forms. Let n ≥ 3, andletn = p + q, where p ≥ 2. Let {e1,...,en} be<br />

the standard basis for R n . If p ≥ 3, let B be the “standard" form<br />

i<br />

B<br />

i=1<br />

xiei<br />

<br />

= 2x1xn +<br />

p<br />

i=2<br />

x 2<br />

i −<br />

n−1<br />

i=p+1<br />

x 2<br />

i . (23)<br />

Let H = SO(B),andlet{at} be the 1-parameter subgroup of H given by ate1 = e −t e1,<br />

atei = ei for 2 ≤ i ≤ n − 1, andaten = e t en. And let K = H ∩ ˆK, where ˆK is the<br />

group of orthogonal matrices with determinant 1. We let dk denote the Haar measure<br />

on K normalized so that K is a probability space.<br />

If (p, q) = (2, 2), welet<br />

B(x1,x2,x3,x4) = x1x4 − x2x3<br />

be the standard form on R4 . This is the determinant on M2(R) if we identify R4 with M2(R). Note that this identification shows that SO(2, 2) is locally isomorphic<br />

to SL2(R) × SL2(R) with the action v → g1vg −1<br />

2 , which leaves the determinant<br />

invariant. We let H = SL2(R) × SL2(R), K = SO(2) × SO(2), andat = (bt,bt),<br />

where bt = diag(e −t/2 ,e t/2 ). We let dk denote the Haar measure on K normalized so<br />

that K is a probability space. We often work with the standard lattice Z 4 and the form<br />

Q, in which case we continue to denote by {at} and K the corresponding 1-parameter<br />

and maximal compact subgroup of SO(Q).<br />

If (p, q) = (2, 1), welet<br />

B(x1,x2,x3) = x1x3 − x 2<br />

2<br />

be the standard form on R3 . This is the determinant on Sym2(R), the space of 2 × 2<br />

symmetric matrices, if we identify R3 with Sym2(R). This identification shows that<br />

SO(2, 1) is locally isomorphic to SL2(R) with the action v → gvtg, where tg is<br />

the transpose matrix. We let H = SL2(R). We let at = diag(e−t/2 ,et/2 ),andwe<br />

let K = SO(2) be the maximal compact subgroup of H. As before, dk denotes the<br />

normalized Haar measure on K.<br />

Let f be a continuous function with compact support on Rn . We define the theta<br />

transform of f by<br />

f ˆ(<br />

+ ξ) = <br />

f (v), (26)<br />

v∈+ξ<br />

where +ξ is any unimodular inhomogeneous lattice in R n . Note that ˆ<br />

f is a function<br />

on the space of inhomogeneous lattices.<br />

(24)<br />

(25)


130 MARGULIS and MOHAMMADI<br />

We fix some more notations. Let n = p + q, andletG = SLn(R) ⋉R n . Let<br />

Ɣ = SLn(Z) ⋉Z n , which is a lattice in G. We have the following, which is similar to<br />

Siegel’s integral formula.<br />

LEMMA 2.1<br />

Let f and fˆ be as above. Let µ be a probability measure on G/ Ɣ which is invariant<br />

under Rn . Then<br />

<br />

<br />

f ˆ(g)<br />

dµ(g) =<br />

G/ Ɣ<br />

Rn f (x) dx. (27)<br />

Proof<br />

Note that Rn is the unipotent radical of G. Now the lemma follows from Fubini’s<br />

theorem and the fact that µ is Rn-invariant. <br />

We end this section by recalling the definition of the α functions defined on the space<br />

of lattices. Let be a lattice in R n . A subspace L of R n is called -rational if L ∩ <br />

is a lattice in L. For a -rational subspace L, letd(L) be the volume of L/(L ∩ ).<br />

For 0 ≤ i ≤ n,define<br />

<br />

1<br />

αi() = sup<br />

d(L)<br />

<br />

: L is a -rational subspace of dimension i , (28)<br />

and let α() = maxi αi(). Now if ξ = + ξ is an inhomogeneous lattice, let<br />

α(ξ) = α(). There is a constant c = c(f ) depending on f such that, for any<br />

inhomogeneous lattice ξ = + ξ, wehave<br />

ˆ<br />

f (ξ)


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 131<br />

Theorem 1.6 is proved using the following, which is a result of combining Theorems<br />

3.1, A.3, andA.4. We have the following.<br />

<strong>THEOREM</strong> 3.2 ([EMM1, Theorem 3.5])<br />

Suppose that p ≥ 3 and q ≥ 1. Let f and fˆ be as above. Let ν be any continuous<br />

function on K. Then, for every compact subset D of G/ Ɣ, there exist finitely many<br />

points x1,...,xℓ ∈ G/ Ɣ such that<br />

(i)<br />

(ii)<br />

the orbit Hxi is closed and has finite H -invariant measure for all i;<br />

for any compact set F ⊂ D \ <br />

i Hxi, there exists t0 > 0 such that, for all<br />

x ∈ F and t>t0, we have<br />

<br />

<br />

<br />

<br />

f ˆ(atkx)<br />

ν(k) dk −<br />

<br />

fdµ ˆ<br />

<br />

<br />

νdk<br />

≤ ε, (31)<br />

K<br />

G/ Ɣ K<br />

where either µ is the G-invariant measure on G/ Ɣ or H ⋉R n x is closed and has<br />

H ⋉R n -invariant probability measure and µ is this measure.<br />

Proof<br />

We may and will assume that φ is nonnegative. Now define<br />

A(r) = ∈ G/ Ɣ : α1() >r . (32)<br />

Let gr be a continuous function on G/ Ɣ such that gr() = 0 if /∈ A(r),gr() = 1<br />

for all ∈ A(r + 1), and0≤ gr() ≤ 1 if r ≤ α1() ≤ r + 1. We have<br />

f ˆ = ( fˆ − fgr) ˆ + fgr. ˆ Note that fˆ − fgr ˆ is a continuous function with compact<br />

support on G/ Ɣ.<br />

Note that H 0 ⋉Rn is a maximal connected subgroup of G. Hence for every<br />

δ>0, there exists r0 such that if H ⋉Rny is a closed orbit of H ⋉Rn in G/ Ɣ with<br />

an H ⋉Rn-invariant probability measure σ ,thenσ (A(r) ∩ H ⋉Rny) r0. Hence for r sufficiently large, we get<br />

<br />

<br />

<br />

fdµ− ˆ ( fˆ − fgr) ˆ <br />


132 MARGULIS and MOHAMMADI<br />

Hence, if we apply Theorem 3.1, then there is a constant B depending on B1 such that<br />

we get<br />

<br />

( fgr)(atkx) ˆ ν(k) dk ≤ B sup |ν(k)| r −β/2<br />

(35)<br />

K<br />

for all x ∈ D.<br />

Now choose r>r0sufficiently large so that B(supk∈K |ν(k)|)r−β/2


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 133<br />

above. Define<br />

f ˜(g<br />

: qξ) = <br />

v∈X(qξ )<br />

f (gv). (36)<br />

The following is an analogue of [EMM2, Theorem 2.3] and will provide us with<br />

the upper bound required for the proof of Theorem 1.9. The proof of this theorem is<br />

the main technical part of this paper and will occupy the rest of this paper.<br />

<strong>THEOREM</strong> 4.1<br />

Let G, H, K, and {at} be as in Section 2 for the signature (2, 2) case. Let Qξ be a<br />

quadratic form of signature (2, 2) which is Diophantine. Let q ∈ SL4(R) and qξ be<br />

as above. Let ν be a continuous function on K. Then we have<br />

<br />

<br />

<br />

lim sup<br />

t→∞<br />

˜f (atk : qξ)ν(k) dk ≤ f ˆ(g)<br />

dµ(g)<br />

K<br />

G/ Ɣ<br />

K<br />

ν(k) dk, (37)<br />

where µ is the G-invariant probability measure on G/ Ɣ if the homogenous part, Q,<br />

is irrational and where the H ⋉R 4 -invariant probability measure on the closed orbit<br />

is H ⋉R 4 · qξ if Q is a rational form.<br />

Proof of Theorem 1.9<br />

Suppose that Qξ is as in the statement of Theorem 1.9. An argument like that<br />

of [EMM1, Sections 3.4, 3.5] combined with Theorem 4.1 gives: if 0 /∈ (a,b),<br />

then<br />

lim sup<br />

T →∞<br />

NQ,ξ,(a,b,T ) = lim sup ÑQ,ξ,(a,b,T ) ≤ λQ,(b − a)T<br />

T →∞<br />

2 . (38)<br />

This upper bound, combined with the lower bound obtained by Theorem 1.4, proves<br />

Theorem 1.9. <br />

The proof of Theorem 4.1 extensively utilizes results and ideas from [EMM1]<br />

and [EMM2], and we will try to use their terminology and notation for the convenience<br />

of the reader. We recall these theorems and terminology when we need them. Let us<br />

start with the following.<br />

<strong>THEOREM</strong> 4.2<br />

Let {at} and K be as in Theorem 4.1. Let be any lattice in R 4 . Then for i = 1, 3<br />

and any ε>0, we have<br />

<br />

sup αi(atk)<br />

t>0 K<br />

2−ε dk < ∞. (39)


134 MARGULIS and MOHAMMADI<br />

Hence there exists a constant c depending on ε and such that, for all t>0 and<br />

0


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 135<br />

If L is a µ1-quasi-null subspace and if T/2 ≤v L ≤T, then there exists C>0<br />

depending on µ1 such that either π2(v L )


136 MARGULIS and MOHAMMADI<br />

Indeed, an estimate like (43) would actually hold for f ˆ (by virtue of [Sch, Lemma<br />

2]) if A was defined by taking the union over all quasi-null subspaces. But we have<br />

replaced fˆ by f ˜,<br />

and this implies that we can take the union over Q instead. To see<br />

this, consider one of these subspaces, say L1. Assume that ξ ∈ L1 + w1ξ, where<br />

w1ξ ∈ Z4 . Now if there is k ∈ K such that (44) is satisfied with L = L1 and v ∈ Z4 ,<br />

then there is a constant cξ ≥ 1 depending on ξ such that d(atkH) 1/cξδ}. Hence there is c such that (43)<br />

holds.<br />

Now the proof of Theorem 4.1 will be completed if we can show that there is<br />

some η>0 depending on Qξ such that |AL t (δ)|


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 137<br />

Furthermore, if L is a null subspace (of the first type), then<br />

k ∈ K : d(atkL) 0 be small, and let<br />

A L,1<br />

t (δ, ε)<br />

<br />

=<br />

k ∈ K : δ1+ε


138 MARGULIS and MOHAMMADI<br />

For simplicity, let ζ = qξ, where q ∈ SL4(R) was chosen such that Q(v) =<br />

B(qv) for all v ∈ R 4 . Fix t > 0 and 0


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 139<br />

(i) there exists λL ∈ such that, if for some λ ∈ there is kφ ∈ EL t (δ, kθ,ε) for<br />

which the plane (L + λ + ζ )φ intersects B(r), then L + λ = L + λL;<br />

(ii) for all λ ∈ , we have maxkφ (at(kθ,kφ)[λ + ζ ]) Lφ ⊥ >c/(δ(1−ε)/2 ).<br />

Furthermore, if (ii) holds, then there exists a computable constant η1 > 0 such that,<br />

for 0 0 such that the function hλ is (C1,β1) good. It follows from the<br />

definition of (C, α)-good functions (see [KM]) that fλ(φ) =(at(kθ,kφ)(v L ∧ (λ +<br />

ζ )) is (C2,β2)-good for some C2,β2 > 0.<br />

Recall now that<br />

c<br />

maxkφ∈E (at(kθ,kφ)[λ + ζ ]) ⊥ Lφ > δ (1−ε)/2 ,<br />

δ1+ε


140 MARGULIS and MOHAMMADI<br />

Note that, for φ ∈ Eλ,wehavefλ(φ)


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 141<br />

let aφ =(at(kθ,kφ)[λ0 + ζ ]) Lφ ⊥. Recall that T/2 ≤vL ≤T, hence aT/2 ≤<br />

w ≤aT. As was mentioned above, K2 acts on the plane spanned by {e123,e124}<br />

by rotation on R 2 , and e123 is the contracting direction of {at}. Also note that we are<br />

assuming that |E L t (λ0)| >δ η2 . Hence, there exist kφ ∈ E L t (λ0) such that<br />

aT δ η2<br />

2 ≤ at(kθ,kφ)[v L ∧ (λ0 + ζ )] = aφat(kθ,kφ)v L . (57)<br />

Note now that we have aφ ≤ r and at(kθ,kφ)v L ≤δ. So we get a ≤ (2rδ 1−η2 )/T,<br />

as we wanted to show. <br />

Before proceeding, let us draw the following corollary. This will be used in the proof<br />

of Theorem 4.1 to control the contribution of “small” subspaces.<br />

COROLLARY 5.5<br />

Let η1 be as in Proposition 5.3, and let η2 1, there<br />

is a δ0 = δ0(M,) such that if δ2, one of the following holds.<br />

(i) The number of quasi-null subspaces of Q of norm between T/2 and T is<br />

O(T 1−τ2 ).


142 MARGULIS and MOHAMMADI<br />

(ii) There exists a split integral form Q ′ with coefficients bounded by a fixed power<br />

of T τ2 and 1 ≤ λ ∈ R satisfying Q − (1/λ)Q ′ ≤T −ρ such that the number<br />

of quasi-null subspaces of Q with norm between T/2 and T which are not null<br />

subspaces of Q ′ is O(T 1−τ2 ).<br />

We refer to [EMM2, Section 10] and also Section 7 of this paper for a more careful<br />

analysis of this theorem. The main ingredient in the proof is the system of inequalities<br />

from [EMM1]. The main difference is that these inequalities are applied to a certain<br />

dilated lattice in 2 R 4 = R 6 .<br />

Let us now outline the rest of the proof. Let Qξ be the inhomogeneous quadratic<br />

form as in the statement of Theorem 1.9. We apply the above theorem with appropriate<br />

parameters ρ,τ2 to be determined later. Let T ≥ 2 be given. Now if (i) above holds<br />

(which is always the case if Q is not EWAS), then we already have a good control on<br />

the number of quasi-null subspaces in question, and we will get the desired control<br />

on the measure of the set <br />

L AL t (δ, ε). Hence, we may assume that (ii) above holds.<br />

Again, if L is not a null subspace of the appropriate approximation of Q, we proceed<br />

as in case (i). So we need to consider the contribution from quasi-null subspaces which<br />

are null subspaces of some rational approximation. In this case, using Lemma 5.3,we<br />

are reduced to the case where only one translate of L has contribution. We then use<br />

the Diophantine property of ξ, and we get a control on the number of such subspaces.<br />

This completes the proof.<br />

We need to fix some more notations before proceeding with the above outline.<br />

If Q is a rational form, we may choose µ1 small enough such that all µ1-quasi-null<br />

subspaces are null subspaces. Also in this case, by replacing Q with a scalar multiple,<br />

we may and will assume that Q is a primitive integral form.<br />

Let T ≥ 2 be a fixed number. Recall from Theorem 5.6 that there are two<br />

possibilities for quasi-null subspaces. The case which requires more study is case<br />

(ii), so let us assume that we are in this case. Let QT = Q ′ (resp., QT = Q) ifQ<br />

is an irrational form (resp., if Q is split integral form), where Q ′ is given as in (ii)<br />

of Theorem 5.6. LetQt(δ, ε, τ2) be the set of all such quasi-null subspaces (resp.,<br />

null subspaces if Q is an rational) which are not exceptional subspaces and such that<br />

A2 δ (•,ε) is nonempty for them. Let L ∈ Qt(δ, ε, τ2) be such a subspace. We further<br />

assume that L is of the first type and that T/2 ≤vL≤T. Since QT is a split integral form, after possibly multiplying by a scalar bounded<br />

by a fixed power of T τ2 , there exists a nonsingular integral matrix such that QT (v) =<br />

B(pv). Furthermore, Theorem 5.6 guarantees that we may choose p such that its<br />

entries are bounded by a fixed power of T τ2 . Recall that the null subspaces of B are<br />

of two types based on the fact that the corresponding vector vL is in V1 or V2. From<br />

now on by a null space of first kind for B, we mean the following.


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 143<br />

Null subspaces of the first type. These are subspaces which are orbits of x11<br />

0<br />

x12 under<br />

0<br />

SL2 × SL2.<br />

Throughout the rest of the section, by a null subspace we mean a rational null<br />

subspace of the first type. Now let M be a rational null subspace of the first type for<br />

B. Such subspaces are characterized as annihilators of primitive row vectors. Hence<br />

an integral basis for M is m 0 0 m<br />

, , where gcd(m, n) = 1. We refer to this basis<br />

n 0 0 n<br />

as the standard integral basis for M.<br />

Recall that L ∈ Qt(δ, ε, τ2) with T/2 ≤vL≤T.Since L is a null subspace of<br />

QT , the subspace M = pL is a null subspace of B. Now let {v1,v2} be the standard<br />

basis for M, and let wi be the unique primitive integral multiple of p−1vi. Then<br />

{w1,w2} is a basis for L. Furthermore, since p is an integral matrix whose entries are<br />

bounded by a fixed power of T τ2 1/2−τ3 , we have T ≤wi ≤T 1/2+τ3 , where τ3 is a<br />

fixed multiple of τ2. We refer to the basis constructed in this way as a τ3-round basis<br />

for L.<br />

Fix 0 δ η2 ,<br />

in particular, that (i) of Proposition 5.3 holds for some λ ∈ . Fix{w1,w2} as a<br />

τ3-round. Note that τ3 is fixed when τ2 is chosen.<br />

Let q −1 λ = v ∈ Z 4 . Now, using Lemma 5.4, there is a constant c = c(q) such<br />

that (v + ξ)L ⊥ < (c(q)δ1−η2 )/T. This and the fact that L is a null subspace of QT<br />

give<br />

|〈wi ,v+ ξ〉QT |≤T 1/2+τ3 · cδ 1−η2<br />

T<br />

cδ1−η2<br />

=<br />

T 1/2−τ3<br />

, (58)<br />

where c is an absolute constant depending on Q. So {〈wi ,ξ〉QT }≤(cδ 1−η2 )/(T 1/2−τ3 ),<br />

where {}denotes the distance to the closest integer. Let us now collect the result of<br />

the above discussion in the following.<br />

LEMMA 5.7<br />

Let ε and τ2 be small, and let L ∈ Qt(δ, ε, τ2). Letkθ ∈ p1(A qL<br />

t (δ, ε)), and suppose<br />

that |E qL<br />

t (δ, kθ,ε,ξ)| >δ η2 with η2 as in Lemma 5.4. In particular, (i) in Proposition<br />

5.3 holds for qL and some λ ∈ . Then {〈wi ,ξ〉QT }≤(cδ 1−η2 )/(T 1/2−τ3 ),<br />

where τ3 is a fixed multiple of τ2 as above. Furthermore, since {w1,w2} is the image of<br />

standard basis {v1,v2} of M = pL, then we have {〈vi , pξ〉B} ≤(cδ 1−η2 )/(T 1/2−τ3 ).<br />

This lemma brings us to the situation where we can now use the Diophantine property<br />

of the vector ξ.<br />

As we mentioned in the brief outline following it, Theorem 5.6 deals with the<br />

Diophantine properties of Q. If Q fails to have the desired Diophantine condition, ξ


144 MARGULIS and MOHAMMADI<br />

needs to be a Diophantine vector thanks to our assumption on Qξ. In what follows,<br />

we make use of this assumption, and we control the number of quasi-null subspaces<br />

for which Lemma 5.7 can hold.<br />

The following is a simple consequence of Definition 1.7 (i.e., the Diophantine<br />

condition). The proof is easy, and we include it for the sake of completeness (and also<br />

for later reference).<br />

LEMMA 5.8<br />

Let x = (x1,...,xn) ∈ R n be a κ-Diophantine vector. Then for any κ ′ > (n+1)κ +1,<br />

we have that, for all β > 0 and all 0


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 145<br />

Proof<br />

We may and will assume that L is of the first type. We continue to use the notations<br />

as in Lemma 5.7. In particular, let M = pL, andlet{v1,v2} be the standard basis<br />

for M. Then Lemma 5.7 implies that {〈vi, pξ〉B} ≤(cδ 1−η2 )/(T 1/2−τ3 ). Recall that<br />

v1 = (m, 0,n,0) and v2 = (0,m,0,n), where gcd(m, n) = 1, and we have<br />

T 1−τ3 ≤v M =m 2 + n 2 ≤ T 1+τ3 , (60)<br />

where τ3 is a fixed multiple of the constant τ2 appearing in Theorem 5.6. We also note<br />

that 〈v1 , pξ〉B = m(pξ)4 − n(pξ)2 and that 〈v2 , pξ〉B(pξ)1 − m(pξ)3.<br />

Let τ ′ 1 ≪ 1/32κ be chosen. Then choose τ2 in Theorem 5.6 such that τ2 T τ ′ 1. Taking τ2 even smaller, we may and will assume that<br />

τ3


146 MARGULIS and MOHAMMADI<br />

Proof<br />

Let η1 be as in Proposition 5.3, andletη2 δ η2 for some kθ ∈ p1(A qL<br />

t (δ, ε)) is O(T 1−τ1 ). On the<br />

other hand, using Theorem 5.6, we see that the number of quasi-null subspaces with<br />

T/2 ≤v L ≤T and A L t (δ, ε) = ∅which are not in Qt(δ, ε, τ2) is O(T 1−τ2 ). The<br />

corollary follows with η = η1/2 and τ = min{τ1,τ2}. <br />

6. Proof of Theorem 4.1<br />

We complete the proof, of Theorem 4.1 in this section. Before proceeding to the proof,<br />

we need the following two statements.<br />

LEMMA 6.1<br />

If Qξ is an irrational (2, 2) form, then the number of 2-dimensional null subspaces,<br />

say L,ofQ such that ξ ∈ v + L for some v ∈ Z 4 is at most four.<br />

Proof<br />

First, note that using [EMM2, Lemma 10.3], we may and will assume that Q is a<br />

rational form. In this case, we actually show that there are at most two such subspaces.<br />

In order to see this, suppose that there are two null rational subspaces of the same<br />

type, say Li, for i = 1, 2 such that ξ ∈ vi + Li, where vi ∈ Z 4 . Now since the Li are<br />

of the same type, they are transversal. Let {w i 1 ,wi 2 } be an integral basis for Li. Then<br />

〈ξ,w i j 〉B ∈ Z for i, j = 1, 2. This (thanks to the transversality of the Li) implies that<br />

ξ is a rational vector, which is a contradiction. <br />

The following is essential to the proof of Theorem 4.1 and proved in Section 7.<br />

PROPOSITION 6.2<br />

The number of quasi-null subspaces with norm between T/2 and T is O(T ).<br />

As we mentioned, this is proved in Section 7. However, for the time being, let us<br />

remark that in the case where Q is a split integral form, this is immediate. Indeed,<br />

in that case we are dealing with null subspaces, and hence we need to show this for<br />

Q = B. As we observed, however, the null subspaces of B are classified by primitive<br />

integral vectors (m, n). Now if L is a null subspace of either type which corresponds<br />

to (m, n), then v L =m 2 + n 2 . Thus the result is obvious.<br />

We now turn to the proof of Theorem 4.1.


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 147<br />

Proof of Theorem 4.1<br />

We may and will assume that ˜f is nonnegative. Now define<br />

A(r) = ∈ X : α1() >r . (63)<br />

Let gr be a continuous function on X such that gr() = 0 if /∈ A(r), gr() = 1<br />

for all ∈ A(r + 1), and0 ≤ gr() ≤ 1 if r ≤ α1() ≤ r + 1. Let the constant c<br />

(depending on f and ξ) beasin(43), and let s ∈ N be a large number. Define<br />

( fgr) ˜ ≤<br />

s = ( fgr)χ ˜<br />

{ : fgr ˜ ()≤c2s } and ( fgr) ˜ ><br />

s = ( fgr)χ ˜<br />

{ : fgr ˜ ()>c2s }. (64)<br />

Now let µ be as in the statement of Theorem 4.1.Sincef˜−fgr ˜ is bounded and<br />

continuous, Theorems A.3 and A.4 imply that<br />

<br />

<br />

lim sup ( f˜ − fgr)(atkqξ)ν(k) ˜<br />

dk ≤ fdµ ˆ νdk. (65)<br />

t→∞ K<br />

G/ Ɣ K<br />

Hence Theorem 4.1 will be proved if we show that: given ɛ>0, we can choose r0<br />

<br />

such that, if r>r0, then lim supt K ˜ fgr dk < ɛ.<br />

Now let ɛ>0 be an arbitrary small number. Let us first control the contribution<br />

of ( fgr) ˜ > s when s is large enough. Let M>0 beanumberfixed(fornow)andlarge<br />

enough such that 1/M < ɛ/6C,where C is a universal constant appearing in (70) and,<br />

in particular, is independent of µ1 in the definition of quasi-null subspaces. Further<br />

assume that M is large enough such that all exceptional subspaces have norm less<br />

than M. Let µ1 in the definition of quasi-null subspace be small enough such that,<br />

for all L ∈ Qt(δ, ε) with vL 0<br />

be large enough such that the conclusions of Theorems 4.2 and 4.5, for this µ1 which<br />

we chose, as well as of Corollaries 5.5 and 5.11, hold for δ = 1/2j whenever j>s.<br />

Recall now that we have<br />

<br />

k ∈ K : ˜f (atkξ) >c2 j ⊂ k ∈ K : α13(atkξ) > 2 j <br />

1<br />

∪ Bt<br />

2j <br />

1<br />

∪ At<br />

2j <br />

,<br />

where At(1/2 j ) and Bt(1/2 j ) are as in (43). Let Ct(1/2 j ) ={k ∈ K : α13(atkξ) ><br />

2 j }.Wehave<br />

<br />

K<br />

h ><br />

s (atKqξ) dk ≤ <br />

s


148 MARGULIS and MOHAMMADI<br />

all j>s,we have<br />

<br />

1<br />

Ct<br />

2j <br />

<br />

+<br />

<br />

Bt<br />

<br />

1<br />

2j <br />

<br />

+<br />

<br />

At<br />

1<br />

2 j<br />

<br />

\ <br />

L∈Qt ( 1<br />

2j ,ε)<br />

AL <br />

1<br />

<br />

<br />

t ,ε <<br />

2j 1<br />

. (68)<br />

2 (1+ε/4)j<br />

Let η1 < 1/4 be as in Proposition 5.3, andletη = η1/2. The Conclusions of<br />

Corollaries 5.5 and 5.11 hold with this η and the corresponding τ.Now let Q < t (1/2j ,ε)<br />

(resp., Q ≥ t (1/2 j ,ε)) be the set of quasi-null subspaces in Qt(1/2 j ,ε) with norm less<br />

than M (resp., greater than or equal to M.)Wehave<br />

| <br />

L∈Q < t (1/2j ,ε) AL 1<br />

t (<br />

| <br />

L∈Q ≥ t (1/2 j ,ε) AL t<br />

( 1<br />

2 j ,ε)|≤ <br />

i<br />

t i+j e /(2 )<br />

i et /(2i+j+1 )<br />

t i+j e /(2 )<br />

et /(2i+j+1 )<br />

2 j ,ε)| ≤ <br />

<br />

L |It (δ)| ≤C1 i<br />

L |It (δ)|≤C2<br />

<br />

<br />

i<br />

1<br />

2 (1+η)j+i/2 +<br />

1<br />

2 (1+η)j+i/2 ,<br />

e −τt<br />

2 (1/2−τ)i+(1−τ)j<br />

where C1 and C2 are absolute constants independent of µ1. The inequality in the<br />

first line above follows from Corollary 5.5 and from the fact that the definition of<br />

Qt(1/2 j ,ε) excludes exceptional subspaces. The inequalities in the second line follow<br />

from Corollary 5.11. Wenowhave<br />

<br />

2 j<br />

<br />

<br />

<br />

j>s<br />

L∈Qt<br />

<br />

1<br />

2j ,ε<br />

<br />

,<br />

(69)<br />

AL <br />

1<br />

<br />

1<br />

t ,ε ≤ C<br />

2j 2s 1<br />

<br />

+ . (70)<br />

η M<br />

Here C ′ is an absolute constant which depends on C in (70). This inequality together<br />

with (68) gives<br />

<br />

K<br />

( fgr) ˜ ><br />

s (atKqξ) dk ≤ C′<br />

+ C<br />

2εs/4 <br />

1 1<br />

+<br />

2ηs M<br />

<br />

. (71)<br />

Recall that M was chosen such that C/M < ɛ/6. Now we choose s large enough<br />

such that the right-hand side of (71)islessthanɛ/2. The above estimate holds for all<br />

r.<br />

As we mentioned in the proof of Theorem 3.2, There exists r0 = r0(s, ɛ) such<br />

that if r>r0, then µ(A(r)) r0(s, ɛ). We have<br />

lim sup<br />

t→∞<br />

<br />

K<br />

( fgr) ˜ ≤<br />

s (atkqξ)ν(k) dk ≤ 2 s lim sup<br />

t→∞<br />

<br />

K<br />

gr(atkqξ)ν(k) dk ≤ ɛ/2.<br />

(72)<br />

Thus (72) and(71) give:ifr>r0(s, ε), then<br />

<br />

lim sup<br />

t→∞<br />

fgr(atkqξ) ˜ dk < ɛ. (73)<br />

K


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 149<br />

This finishes the proof of Theorem 4.1. <br />

7. Proof of Theorem 1.10<br />

The proof of Theorem 1.10 is very similar to that of Theorem 1.9. Indeed, our study<br />

in this case is simpler as we are dealing with the case where the homogeneous part Q<br />

is rational. Hence, we only need to consider the contribution of null subspaces to the<br />

counting function.<br />

As before, let q ∈ SL3(R) be such that Q(v) = B(qv) for all v ∈ R 3 . Since Q is<br />

rational, we may assume that q is in PGL3(Q). Let p ∈ GL3(Q) be a representative<br />

for q. Let = qZ 3 , and define qξ = q(Z 3 + ξ). As in Section 4, letX(qξ)<br />

be the set of vectors in qξ not contained in qL, where L ⊂ Z 3 is an exceptional<br />

(1-dimensional) subspace for Q. An argument like that in Lemma 6.1 shows that there<br />

are at most three such subspaces when Q is rational and that ξ is an irrational vector.<br />

For any continuous compactly supported function f on R 3 , define<br />

f ˜(g<br />

: qξ) = <br />

v∈X(qξ )<br />

f (gv). (74)<br />

As in previous discussions, we reduce the proof of Theorem 1.10 to the following<br />

theorem.<br />

<strong>THEOREM</strong> 7.1<br />

Let G, H, K, and {at} be as in Section 2 for the signature (2, 1) case. Let Qξ be a<br />

quadratic form of signature (2, 1) as in the statement of Theorem 1.10.Letq∈ SL3(R)<br />

and qξ be as above. Let ν be a continuous function on K. Then we have<br />

lim sup<br />

t→∞<br />

<br />

K<br />

<br />

˜f (atk : qξ)ν(k) dk ≤<br />

G/ Ɣ<br />

<br />

f ˆ(g)<br />

dµ(g)<br />

K<br />

ν(k) dk, (75)<br />

where µ is the H ⋉R 3 -invariant probability measure on the closed orbit H ⋉R 3 ·qξ.<br />

The proof of this theorem is very similar to that of Theorem 4.1. We will use the<br />

same notations as those in the previous sections for the sake of simplicity. The main<br />

notational difference to bear in mind is that in previous sections L denotes a 2dimensional<br />

subspace, where in this section L is a 1-dimensional (null) subspace.<br />

As before, we need to study the subsets of K where the function ˜f is “large”.<br />

We start by recalling the following. There is c>0 such that, for all large t and small


150 MARGULIS and MOHAMMADI<br />

0 c<br />

<br />

⊂ k ∈ K : α1(atkqξ)<br />

δ<br />

> 1<br />

<br />

∪ k ∈ K : α2(atkqξ) ><br />

δ<br />

1<br />

<br />

. (76)<br />

δ<br />

We fix t and δ as above. Let us make two important remarks before we continue.<br />

Remarks 7.2<br />

(i) Using reduction theory of the orthogonal group, we see that if α2(atkqξ) ><br />

1/δ, then actually α1(atkqξ) > 1/δ. Hence we only need to study the contribution<br />

coming from α1.<br />

(ii) The fact that Q is a rational form implies that there exists δ0 depending only on<br />

Q such that if 0


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 151<br />

For any null subspace L of Q, the subspace pL is a (rational) null subspace of<br />

B. We let vL denote the standard basis for pL. For any such L, let AL t (δ) be the<br />

corresponding set in (77). We have the following.<br />

LEMMA 7.3<br />

There exists 0


152 MARGULIS and MOHAMMADI<br />

Some important properties of these intervals were proved in [EMM2] and recalled in<br />

Lemma 5.1. What is important for us in Section 7 is property (ii) in Lemma 5.1.This<br />

property, tailored to our current assumptions, gives<br />

|A L<br />

t<br />

−t<br />

L e δ<br />

1/2 (δ)| ≤|It (δ)| ≈ . (81)<br />

T<br />

Proof of Theorem 7.1<br />

Let M be a large number which is fixed for now. Let t>0 be a large number. Assume<br />

also that j is a large number. We assume throughout that η = η1, where η1 is as in<br />

Lemma 7.3. LetL0,L1 be nonexceptional null subspaces such that {〈v Lk , pξ〉B} <<br />

1/2 j for k = 0, 1 and such that v L0 ≤v L1 are minimal with these properties.<br />

Indeed, we fix any two such subspaces if there are more than two subspaces satisfying<br />

these conditions. Assume that e t /2 j+i0+1 ≤v L 2 jη2 , where η2 is as in Lemma 7.3. Now using this<br />

lemma, we see that, for all i ≤ i0, the number of null subspaces L with e t /2 j+i+1 ≤<br />

v L i0, the subspaces of norm at most e t /2 i+j have<br />

no contribution to the set At(1/2 j ), that is, A L t (1/2j ) =∅for all such subspaces L.<br />

Hence if we use this fact and (81), we have: if j is such that Case 1 holds, then<br />

we have<br />

<br />

<br />

<br />

<br />

Lnull<br />

AL <br />

1<br />

t<br />

2j <br />

<br />

<br />

≤<br />

i<br />

≤ 1<br />

2 j(1+η)<br />

<br />

<br />

<br />

et /(2i+j+1 )≤vL


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 153<br />

Now, if we argue just as in Case 1, we get<br />

<br />

<br />

<br />

L null<br />

AL <br />

1<br />

t<br />

2j <br />

<br />

≤<br />

i<br />

<br />

<br />

<br />

et /(2i+j+1 )≤vL


154 MARGULIS and MOHAMMADI<br />

of G, and let K be the maximal compact subgroup of H . The question which is<br />

addressed in this section is that of the equidistribution of sets of the form atKx,<br />

where x ∈ G/ Ɣ and where A ={as : s ∈ R} is a suitable 1-parameter subgroup<br />

of H. This is a well-studied question (see [Sh2]). The main tools in the analysis are<br />

indeed Ratner’s theorem on classification of unipotent flow-invariant measures on<br />

G/ Ɣ, and linearization techniques for the action of unipotent groups on G/ Ɣ which<br />

were developed by Dani and Margulis [DM].<br />

Let H and W be closed subgroups of G. Following Dani and Margulis [DM],<br />

define X(H,W) ={g ∈ G : Wg ⊂ gH}. We recall the following.<br />

<strong>THEOREM</strong> A.1 ([DM, Theorem 3])<br />

Let G be a connected Lie group, and let Ɣ be a lattice in G. Let U ={ut} be<br />

an Ad-unipotent 1-parameter subgroup of G. Let φ be a bounded continuous function<br />

on G/ Ɣ. Let D be a compact subset of G/ Ɣ, and let ε>0 be given. Then<br />

there exist finitely many proper closed subgroups H1 = H1(φ,D,ε),...,Hk =<br />

Hk(φ,D,ε) such that Hi ∩ Ɣ is a lattice in Hi for all i, and compact subsets<br />

C1 = C1(φ,D,ε),...,Ck = Ck(φ,D,ε) of X(H1,U),...,X(Hk,U), respectively,<br />

for which the following holds. For any compact subset F of D \ <br />

i CiƔ/Ɣ<br />

there exists T0 > 0 such that, for all x ∈ F and T>T0, we have<br />

<br />

<br />

1<br />

T<br />

<br />

<br />

φ(utx) dt − φdg<br />

0.<br />

Note that the standard representation of H 0 on Rn is irreducible, hence H 0 is a<br />

maximal connected subgroup of H ⋉Rn . Note also that as H 0 is a maximal connected<br />

subgroup of SLn(R), we have that H 0 ⋉Rn is a maximal connected subgroup of G.


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 155<br />

Now let Ɣ be a lattice in G. Then Ɣ ∩ R n is a lattice in R n . We let = ϑ(Ɣ) =<br />

Ɣ/ Ɣ ∩ R n (this is a lattice in SLn(R)). We abuse the notation and let ϑ also denote<br />

the projection from G/ Ɣ onto SLn(R)/. We have the following.<br />

LEMMA A.2<br />

Let x ∈ G/ Ɣ. Then the orbit Hϑ(x) is closed in SLn(R)/ if and only if H ⋉R n x<br />

is closed in G/ Ɣ.<br />

Proof<br />

Note that closed orbits of H ⋉R n (resp. H ) have a finite H ⋉R n -invariant (resp.,<br />

H -invariant) measure by [Mar86, Section 3]. Let x = (g, v)Ɣ. Note that H ⋉R n x =<br />

ϑ −1 (Hϑ(x)), hence if the Hϑ(x) is closed, then so is H ⋉R n x. Suppose now that<br />

H ⋉R n x is closed. Then Ɣ1 = H ⋉R n ∩ (g, v)Ɣ(g, v) −1 is a lattice in H ⋉R n ,<br />

and hence Ɣ1 ∩ R n is a lattice in R n and Ɣ1/Ɣ1 ∩ R n is a lattice in H. Hence Hϑ(x)<br />

is closed, as we wanted. <br />

The following is special case of [EMM1, Theorem 4.4].<br />

<strong>THEOREM</strong> A.3<br />

Let the notation be as above. Further assume that is a lattice in H 0 ⋉R n . Let φ<br />

be a compactly supported continuous function on H 0 ⋉R n /. Then for every ε>0<br />

and any bounded measurable function ν on K, and for every compact subset D of<br />

H 0 ⋉R n /, there exist finitely many points x1,...,xℓ ∈ H 0 ⋉R n / such that<br />

(i) the orbit H 0 xi is closed and has finite H 0 -invariant measure for all i;<br />

(ii) for any compact set F ⊂ D \ <br />

i Hxi, there exists s0 > 0 such that, for all<br />

x ∈ F and s>s0, we have<br />

<br />

<br />

<br />

φ(askx) ν(k) dk −<br />

K<br />

H 0 ⋉R n /<br />

<br />

φdg<br />

K<br />

<br />

<br />

νdk<br />

≤ ε. (89)<br />

We also need a slight variant of [EMM1, Theorem 4.4]. This is the content of the<br />

following.<br />

<strong>THEOREM</strong> A.4<br />

Let G, H, K, and Ɣ be as above. Let A ={as : s ∈ R} also be as above. Let φ be<br />

a compactly supported continuous function on G/ Ɣ. Then for every ε>0 and any<br />

bounded measurable function ν on K, and for every compact subset D of G/ Ɣ,there<br />

exists finitely many points x1,...,xℓ ∈ G/ Ɣ such that<br />

(i) the orbit H ⋉R n xi is closed and has finite H ⋉R n -invariant measure for all<br />

i;


156 MARGULIS and MOHAMMADI<br />

(ii) for any compact set F ⊂ D \ <br />

i H ⋉Rn xi, there exists s0 > 0 such that, for<br />

all x ∈ F and s>s0, we have<br />

<br />

<br />

<br />

φ(askx) ν(k) dk −<br />

K<br />

G/ Ɣ<br />

<br />

φdg<br />

K<br />

<br />

<br />

νdk<br />

≤ ε. (90)<br />

Proof<br />

The proof of this theorem goes along the same lines as in [EMM1, Section 4]. Let U be<br />

as defined above. Let Hi = Hi(φ,KD,ε) and Ci = Ci(φ,KD,ε) for i = 1,...,k<br />

be given as in Theorem A.1 corresponding to U.For1 ≤ i ≤ k, define<br />

Gi = g ∈ G : Kg ⊂ X(Hi,U) . (91)<br />

Note that the group generated by <br />

k∈K k−1Uk is H 0 , as U is not contained in any<br />

proper normal subgroup of H and K is the maximal compact subgroup of H. Now let<br />

g ∈ Gi. Then k−1Uk ⊂ gHig−1 for all k ∈ K. Hence, H 0 ⊂ gH 0<br />

i g−1 . Now as H 0<br />

acts irreducibly on Rn , the only possibilities for gH 0<br />

i g−1 are H 0 and H 0 ⋉Rn . Thus<br />

we have<br />

gH 0<br />

i g−1 ⊂ H 0 ⋉R n , for any g ∈ Gi, 1 ≤ i ≤ k. (92)<br />

We also note that if g ∈ G is such that gH 0 g −1 ⊂ H 0 ⋉R n , then g ∈ NSLn(R)(H 0 ) ⋉<br />

R n . This fact and (92) say that if g1,g2 ∈ Gi for some i, then g −1<br />

1 g2 ∈ NSLn(R)(H 0 ) ⋉<br />

R n . Since H 0 ⋉R n is of finite index in NSLn(R)(H ) ⋉R n , we get that Gi can be<br />

covered by finitely many cosets of H 0 ⋉R n . Hence there are finitely many points<br />

x1,...,xℓ ∈ G/ Ɣ such that the H ⋉R n xi are closed and have a finite H ⋉R n -<br />

invariant measure for all 1 ≤ i ≤ ℓ and such that<br />

<br />

GiƔ/ Ɣ ⊂ <br />

H ⋉Rn xi. (93)<br />

i<br />

1≤i≤ℓ<br />

Note that since the X(Hi,U) are analytic submanifolds of G and since K is connected,<br />

we have that, for any g ∈ G \ <br />

i GiƔ/ Ɣ,<br />

<br />

<br />

k ∈ K : kg ∈ <br />

<br />

X(Hi,U)Ɣ = 0. (94)<br />

i<br />

Now (93), (94), and the fact that Ci ⊂ X(Hi,U), give<br />

|{k ∈ K : kx ∈ <br />

CiƔ/ Ɣ}| = 0 (95)<br />

i


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 157<br />

for any x ∈ F ⊂ D \ <br />

i H ⋉Rnxi. Now if we apply [EMM1, Lemma 4.2], then<br />

there exists an open subset W ⊂ G/ Ɣ such that <br />

i CiƔ/ Ɣ ⊂ W and such that<br />

|{k ∈ K : kx ∈ W }| 0 such that<br />

<br />

<br />

1<br />

T <br />

<br />

φ(uty) − φdg<br />

0 be any sufficiently small number, and let T > 2. There exists a rational<br />

3-dimensional subspace U of 2 R 4 with a reduced integral basis of norm at most<br />

T τ whose projections into V2 have norm less than T τ−1 such that one of the following<br />

holds.<br />

(a) The restriction of Q (6) to U is anisotropic over Q, in which case (i) in Theorem<br />

5.6 holds.<br />

(b) The restriction of Q (6) to U splits over Q, in which case (ii) in Theorem 5.6<br />

holds. Furthermore, in this case Q ′ (as in loc. cit.) is proportional to f (U,U ⊥ ),<br />

where U ⊥ is the orthogonal complement of U with respect to Q (6) and where f<br />

is as in Lemma 4.3. Moreover, the number of quasi-null subspaces with norm<br />

between T/2 and T which are not in U or U ⊥ is O(T 1−τ ).<br />

Let us denote by Q (3) (v) the restriction of Q (6) to U. As we mentioned before, this is<br />

a form with signature (2, 1) and U is a rational subspace. If L is a quasi-null subspace<br />

with T/2 ≤v L ≤T, then we have Q (6) (v L ) = 0. Hence if L is a quasi-null<br />

subspace in U, then Q (3) (v L ) = 0. In other words, Proposition 6.2 follows from the<br />

following.


158 MARGULIS and MOHAMMADI<br />

PROPOSITION B.2<br />

Let Q0(x,y,z) = 2xz − y 2 be the standard quadratic form of signature (2, 1) on<br />

R 3 . Let = gZ 3 ,whereg ∈ GL3(Q). Let T > 0 be a large parameter, and let τ<br />

be a sufficiently small parameter. Assume that the entries of g are rational numbers<br />

whose nominator and denominators are bounded by a fixed power of T τ , and also<br />

suppose that |det g| ≤T τ . Further assume that the length of the shortest vector in <br />

is c = O(1). Then if T is large enough, we have<br />

# w ∈ P () : Q0(w) = 0 and w ≤T 0. Note that C is dilation-invariant. Let λ = disc() 1/3 , and define<br />

1 = 1/λ. The lattice 1 is unimodular, and the length of the shortest vector in<br />

1 is at least c/λ. Let¯g ∈ PGL3(Q) denote the image of g. Indeed, 1 = ¯gZ 3 . The<br />

counting problem in (98) follows if we show that<br />

<br />

# w ∈ P (1) : Q0(w) = 0 and w ≤ T<br />

<br />


<strong>IN</strong>HOMOGENEOUS QUADRATIC FORMS 159<br />

Hence we see that C ∩P (1) ⊂ Ɣmv0. As is a finite set, we only need to consider<br />

Ɣmv0.<br />

Let e1 = (1, 0, 0), and let N0 ⊂ H be the stabilizer of e1. Let BR ={h ∈ H :<br />

h · e1 ∈ CR}, and let BR = BR/N0. Now let χR denote the characteristic function of<br />

CR. Define the following:<br />

FR(h) = <br />

χR(hγ v), for h ∈ H. (100)<br />

Ɣm/Ɣm∩N<br />

This is a function on H/Ɣ.Now using the well-developed machinery for counting<br />

problems using the mixing property (see in particular [BO] and[MS]), there exists<br />

ε>0 depending only on the spectral gap of H and τ such that<br />

FR(e) = |N0/N0 ∩ Ɣ|<br />

vol(BR)<br />

|H/Ɣ|<br />

1 + O(vol(BR) −ε ) , (101)<br />

where vol denotes the H -invariant Haar measure on H/N. Hence we have<br />

#(Ɣmv0 ∩ CR) ≤ FR/v0(e) = O(R/v0). (102)<br />

Since v0 ≥c/λ, this finishes the proof. <br />

Acknowledgments. We would like to thank J. Marklof for reading the first draft and<br />

for many helpful comments. We also thank the anonymous referees for their remarks<br />

and suggestions. The second author would like to thank A. Eskin for conversations<br />

regarding these remarks.<br />

References<br />

[BO] Y. BENOIST and H. OH, Effective equidistribution of S-integral points on<br />

symmetric varieties, preprint, arXiv:0706.1621v1 [math.NT] 159<br />

[BT] M. V. BERRY and M. TABOR, Level clustering in the regular spectrum, Proc.Royal<br />

Soc. A 356 (1977), 375 – 394. 126<br />

[DM] S. G. DANI and G. A. MARGULIS, “Limit distributions of orbits of unipotent flows<br />

and values of quadratic forms” in I. M. Gelfand Seminar, Adv.SovietMath.<br />

16, Amer. Math. Soc., Providence, 1993, 91 – 137. MR 1237827 123, 124,<br />

128, 154<br />

[EMM1] A. ESK<strong>IN</strong>, G. A. MARGULIS,andS. MOZES, Upper bounds and asymptotics in a<br />

quantitative version of the Oppenheim conjecture, Ann. of Math. (2) 147<br />

(1998), 93 – 141. MR 1609447 122, 123, 124, 125, 126, 128, 130, 131, 132,<br />

133, 134, 142, 155, 156, 157


160 MARGULIS and MOHAMMADI<br />

[EMM2] ———, Quadratic forms of signature (2,2) and eigenvalue spacings on<br />

rectangular 2-tori, Ann.of Math. (2) 161 (2005), 679 – 725. MR 2153398<br />

122, 123, 124, 125, 126, 128, 133, 134, 135, 136, 137, 141, 142, 146, 152, 157<br />

[GT] B. GREEN and T. TAO, The quantitative behavior of polynomial orbits on<br />

nilmanifolds, preprint. 151<br />

[KM] D. Y. KLE<strong>IN</strong>BOCK and G. A. MARGULIS, Flows on homogeneous spaces and<br />

Diophantine approximation on manifolds, Ann. of Math. (2) 148 (1998),<br />

339 – 360. MR 1652916 139<br />

[Mar] G. A. MARGULIS, “Lie groups and ergodic theory” in Algebra—Some Current<br />

Trends (Varna, 1986), Lecture Notes in Math. 1352, Springer, Berlin, 1988,<br />

130 – 146. MR 0981823<br />

[Mark] J. MARKLOF, Pair correlation densities of inhomogeneous quadratic forms, Ann.<br />

of Math. (2) 158 (2003), 419 – 471. MR 2018926 126, 128<br />

[MS] A. MOHAMMADI and A. S. GOLSEFIDY, Translates of horospherical measures and<br />

counting problems, preprint. 159<br />

[Sar] P. SARNAK, “Values at integers of binary quadratic forms” in Harmonic Analysis<br />

and Number Theory (Montreal, 1996),CMSConf.Proc.21,<br />

Amer. Math. Soc., Providence, 1997, 181 – 203. MR 1472786 126, 127<br />

[Sch] W. M. SCHMIDT, Asymptotic formulae for point lattices of bounded determinant<br />

and subspaces of bounded height, Duke Math. J. 35 (1968), 327 – 339.<br />

MR 0224562 130, 136<br />

[Sh2] N. SHAH, Limit distributions of expanding translates of certain orbits on<br />

homogeneous spaces, Proc. Indian Acad. Sci. Math. Sci. 106 (1996),<br />

105 – 125. MR 1403756 154<br />

Margulis<br />

Department of Mathematics, Yale University, New Haven, Connecticut 06520, USA;<br />

margulis@math.yale.edu<br />

Mohammadi<br />

Department of Mathematics, University of Chicago, Chicago, Illinois 60637, USA;<br />

amirmo@math.uchicago.edu

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!