December 18, 2004. Revised February 15, 2005
The author was supported by NSF grant DMS-040126
.
Power series solution of a nonlinear Schrödinger equation
Michael Christ
Michael Christ, Department of Mathematics, University of California, Berkeley, CA 94720-3840, USA E-mail address : mchrist@math.berkeley.edu
-
Abstract.
A slightly modified variant of the cubic periodic one-dimensional nonlinear Schrödinger equation is shown to be well-posed, in a relatively weak sense, in certain function spaces wider than
. Solutions are constructed as sums of infinite series of multilinear operators applied to initial data; no fixed point argument or energy inequality are used.
1 Introduction
1.1 The NLS Cauchy problem
The Cauchy problem for the one-dimensional periodic cubic nonlinear Schrödinger equation is
|
(NLS)
|
where
,
, and the parameter
equals
. Bourgain [2] has shown this problem to be wellposed in the Sobolev space
for all
, in the sense of uniformly continuous dependence on the initial datum. In
it is wellposed globally in time, and as is typical in this subject, the uniqueness aspect of wellposedness is formulated in a certain auxiliary space more restricted than
, in which existence is also established. For
it is illposed in the sense of uniformly continuous dependence [3] , and is illposed in stronger senses [5] as well.
The objectives of this paper are twofold. Firstly, we seek to establish the existence of solutions for wider classes of initial data than
. Secondly, we aim to develop an alternative method of solution.
The spaces of initial data considered here are the spaces
for
and
, defined as follows:
Definition 1.1.
.
Here
is the usual space of distributions, and
is equipped with the norm
.
We write
, and are mainly interested in these spaces since, for
, they are larger function spaces than the borderline Sobolev space
in which NLS is already known to be wellposed.
1.2 Motivations
At least four concrete considerations motivate analysis of the Cauchy problem in these particular function spaces. Firstly,
scales like
where
as
, thus spanning the gap between the optimal exponent
for Sobolev space wellposedness, and the scaling exponent
.
A second motivation is the work of Kappeler and Topalov [9] ,[10] , who showed via an inverse scattering analysis that the periodic KdV and mKdV equations are wellposed for wider ranges of Sobolev spaces
than had previously been known. It is reasonable to seek a corresponding improvement for NLS , but this problem has been shown to be illposed in strong senses in
for all
[5] . Christ and Erdogan have investigated in unpublished work the relatively simple “action variable” portion of the inverse scattering theory relevant to NLS , and have found that for any distribution in
with small norm, the sequence of gap lengths for the associated Dirac operator belongs to
and has comparable norm.1
Thus
for
may be a natural setting for inverse scattering theory for the Dirac operator relevant to the periodic cubic nonlinear Schrödinger equation.
A related third motivation is the goal of developing an alternative approach to the results of Kappeler and Topalov, independent of inverse scattering theory. NLS seems to be technically simpler than mKdV or KdV, so it may be a reasonable starting point.
Fourthly, Grünrock [7] has proved wellposedness for the cubic nonlinear Schrödinger equation in spaces analogous to
, with
replaced by
, and for other PDE in these function spaces, as well.
1.3 Modified equation
In order for the Cauchy problem to make any sense in
for
it seems to be essential to modify the differential equation. We consider
|
(NLS*)
|
where
|
(1.1)
|
equals the mean value of the absolute value squared of
. In NLS* ,
is shorthand for
, which is independent of
for all sufficiently smooth solutions; modifying the equation in this way merely introduces a unimodular scalar factor
, where
. For parameters
such that
is not embedded in
,
is not defined for typical
, but of course the same goes for the function
, and we will nonetheless prove that the equation makes reasonable sense for such initial data.
The coefficient
in front of
is the unique one for which solutions depend continuously on initial data in
for
.
1.4 Conclusions
Our main result is as follows. Recall that there exists a unique mapping
, defined for
, which for all sufficiently large
extends to a uniformly continuous mapping from
to
, such that
is a solution of the modified Cauchy problem NLS* .
is of course a dense subset of
for any
.
Theorem 1.1.
For any
, any
, and any
, there exists
for which the solution mapping
extends by continuity to a uniformly continuous mapping from the ball centered at
of radius
in
to
.
For the unmodified equation this has the following consequence. Denote by
the set of all
such that
. Denote by
the usual solution [2] of the unmodified Cauchy problem NLS with initial datum
, for
.
Corollary 1.2.
Let
and
. For any
there exists
such that for any finite constant
, the mapping
is uniformly continuous as a mapping from
intersected with the ball centered at
of radius
in
, equipped with the
norm, to
.
The unpublished result of the author and Erdogan shows that for initial data in
, for which the solution is known to exist globally in time,
uniformly for all
, provided that
is sufficiently small. This result, once published, will combine with Theorem 1.1 to yield global wellposedness for sufficiently small data.
The following result quantifies the relation between the nonlinear evolution NLS* ond the corresponding linear Cauchy problem
|
(1.2)
|
Proposition 1.3.
Let
and
. Let
also satisfy
.
Then there exist
and
such that for any initial datum
satisfying
, the solutions
of NLS* and
of 1.2 satisfy
|
(1.3)
|
Here
the solution defined by approximating
by elements of
and passing to the limit. Thus for
the nonlinear terms are in a sense smoother than the linear evolution.
Our next result indicates that the function
defined by the limiting procedure of Theorem 1.1 is a solution of the differential equation in a more natural sense than merely being a limit of smooth solutions. Define Fourier truncation operators
, acting on
, by
for all
, and
whenever
.
acts also on functions
by acting on
for each time
separately. We denote by
the limiting function whose existence, for nonsmooth
, is established by Theorem 1.1 .
Proposition 1.4.
Let
,
, and
. Write
. Then for any
there exists
such that whenever
,
exists in the sense that
|
(1.4)
|
Moreover if
is interpreted as this limit, then
satisfies NLS* in the sense of distributions in
.
More generally, the same holds for any sequence of Fourier multipliers of the form
where each sequence
is finitely supported,
, and
as
for each
; the limit is of course independent of the sequence
. Making sense of the nonlinearity via this limiting procedure is connected with general theories of multiplication of distributions [1] ,[6] , but the existence here of the limit over all sequences
gives
stronger claim to the title of solution than in the general theory.
Unlike the fixed point method, our proof yields no uniqueness statement corresponding to these existence results. But this failing is unavoidable; for all
, solutions of the Cauchy problem in the class
, in the sense of Proposition 1.4 , are not unique [4] .
1.5 Method
Define the partial Fourier transform
|
(1.5)
|
Our approach is to regard the partial differential equation as an infinite coupled nonlinear system of ordinary differential equations for these Fourier coefficients, to express the solution as a power series in the initial datum
|
(1.6)
|
where each
is a bounded multilinear operator2
from a product of
copies of
to
, to show that the individual terms
are well-defined, and to show that the formal series converges absolutely in
to a solution in the sense of 1.4 . The case
follows from a very small modification of the analysis for
, so we discuss primarily
, indicating the necessary modifications for
at the end of the paper.
The analysis is quite elementary, much of the paper being devoted to setting up the definitions and notation required to describe the operators
. A single number theoretic fact enters the discussion: the number of factorizations of an integer
as a product of two integer factors is
as
, for all
; this same fact was used in a more sophisticated way by Bourgain [2] .
The author is grateful to J. Bourgain, C. Kenig, H. Koch, and D. Tataru for invitations to conferences that stimulated this work, and to Betsy Stovall for thorough proofreading of the manuscript.
2 A system of coupled ordinary differential equations
2.1 General discussion
Define
|
(2.1)
|
It factors as
|
(2.2)
|
Written in terms of Fourier coefficients
, the equation
becomes
|
(2.3)
|
Here the first summation is taken over all
satisfying the indicated identity, and the second over all
. Substituting
|
(2.4)
|
2.3 becomes
|
(2.5)
|
where the notation
means that the sum is taken over all
for which neither
nor
. This notational convention will be used throughout the discussion.
The effect of the term
in the modified differential equation NLS* is to cancel out a term
, which would otherwise appear on the right-hand side of 2.5 .
Reformulated as an integral equation, 2.5 becomes
|
(2.6)
|
However, in deriving 2.6 from 2.5 , we have interchanged the integral over
with the summation over
without any justification. 2.6 is fully equivalent to
|
(2.7)
|
Substituting for
in the right-hand side of 2.6 by means of the equation itself yields
|
(2.8)
|
| |
We recognize
as a Taylor polynomial for
, but for our purposes it will not be necessary to exploit this by recombining terms, and in particular we will not exploit the coefficient
which makes this exponential unimodular.
2.2 A sample term
One representative additional term is
(2.9)
(
i
ω
)
4
∑
j
1
−
j
2
+
j
3
=
n
*
∑
m
1
1
−
m
2
1
+
m
3
1
=
j
1
*
∑
m
1
2
−
m
2
2
+
m
3
2
=
j
2
*
∑
m
1
3
−
m
2
3
+
m
3
3
=
j
3
*
∫
0
≤
r
1
,
r
2
,
r
3
≤
s
≤
t
a
m
1
1
(
r
1
)
a
¯
m
2
1
(
r
1
)
a
m
3
1
(
r
1
)
a
¯
m
1
2
(
r
2
)
a
m
2
2
(
r
2
)
a
¯
m
3
2
(
r
2
)
a
m
1
3
(
r
3
)
a
¯
m
2
3
(
r
3
)
a
m
3
3
(
r
3
)
e
i
σ
(
j
1
,
j
2
,
j
3
,
n
)
s
e
i
σ
(
m
1
1
,
m
2
1
,
m
3
1
,
j
1
)
r
1
e
−
i
σ
(
m
1
2
,
m
2
2
,
m
3
2
,
j
2
)
r
2
e
i
σ
(
m
1
3
,
m
2
3
,
m
3
3
,
j
3
)
r
3
d
r
1
d
r
2
d
r
2
d
s
.
Substituting via 2.6 for each coefficient
yields a main term
|
(2.10)
|
plus higher-degree terms, where the superscript
indicates here that the sum is taken over only certain
(more precisely, over most of a copy of
affinely embedded in
),
if
is even and
if
is odd, and
|
(2.11)
|
with
|
(2.12)
|
here
are defined as functions of
by the equations governing the sums in 2.9 . Continuing in this way yields formally an infinite expansion for the sequence
in terms of multilinear expressions in the initial datum
. This expansion is doubly infinite; the single (and relatively simple) term 2.10 is for instance an infinite sum over most of a copy of
for each
.
The discussion up to this point has been purely formal, with no justification of convergence.
In the next section we will begin to describe the terms in this expansion systematically.
3 Trees
On a formal level
equals an infinite sum
where each
is a sum of finitely many multilinear operators, each of degree
. We now describe a class of trees which will be used both to name, and to analyze, these multilinear operators.
Definition 3.1.
A tree
is a finite partially ordered set with the following properties:
-
(1)
Whenever
and
and
, then either
or
.
-
(2)
There exists a unique element
satisfying
for all
.
-
(3)
Each
has either three children, or no children;
is said to be a child of
if
and if there exists no
satisfying
.
-
(4)
For each
there is given an element of
, denoted
.
Definition 3.2.
Elements of
are called nodes. A terminal node is one with zero children. The maximal element of
is called its root node. For any
,
is a tree, with root node
.
denotes the set of all terminal nodes of
, while
denotes the set
of all non-terminal nodes. The three children of any
are denoted by
.
The number
of nodes of a tree is of the form
for some nonnegative integer
.
The number of terminal nodes is then
|
(3.1)
|
Definition 3.3.
An ornamented tree is a tree
together with the following additional structure:
-
(1)
Associated to each node
is copy of
, indexed by the variable
.
-
(2)
There is given a partition of the set of all non-terminal nodes of
into two disjoint classes, called simple nodes and general nodes. Terminal nodes are neither simple nor general.
-
(3)
For each non-terminal node
, and each
such that the child
is non-terminal, there is given a coefficient
.
-
(4)
Associated to each node
is a
-valued function
of
, defined by
|
(3.2)
|
and
|
(3.3)
|
actually depends only on
. We will use the symbol
to denote both the ornamented tree and the underlying tree, and will often write
instead of
.
Definition 3.4.
Let
be a tree.
denotes the set of all
satisfying the restrictions
|
(3.4)
|
|
(3.5)
|
|
(3.6)
|
Definition 3.5.
Let
be any tree.
denotes the function
if
is terminal, and
if
is non-terminal.
Let
be sufficiently small positive numbers, to be chosen later. The following key definition involves these quantities.
Definition 3.6.
Given an ornamented tree
and
, we say that a node
is frozen if
is non-terminal and
|
(3.7)
|
If
is not frozen, then
is said to be alive. A non-terminal node
is said to be exceptional if
.
Whether
is frozen depends on the values of
for all nodes
, as well as on
for all non-terminal
, not merely on the structure of
; a non-terminal node will be frozen for some
, but alive for others. Thus it would be more felicitous to say that a pair
is frozen, rather than a node
.
Exceptional nodes are of course frozen. If
is a general node all three of whose children of
are terminal, then
cannot be exceptional, for
cannot vanish, by 3.5 . But if
has at least one non-terminal child, then nothing prevents
from vanishing, and if
is a simple node all of whose children are terminal, then
is certainly exceptional.
Definition 3.7.
A weathered ornamented tree
is an ornamented tree
together with a subset
and the collection
|
(3.8)
|
4 Multilinear operators associated to trees
Definition 4.1.
Let
be any tree, and let
. The associated tree coefficients are
|
(4.1)
|
where
|
(4.2)
|
The following upper bounds for the coefficients
are the only information concerning them that will be used in the analysis.
Lemma 4.1.
Let
be any tree, and let
. Then for all
,
|
(4.3)
|
|
(4.4)
|
The notation
means
. The sum here is taken over all of the
possible choices of
; these choices in turn determine the quantities
. This lemma will be proved in § 6 .
Definition 4.2.
Let
be any ornamented tree. The tree operator
associated to
is for each
the multilinear operator that maps
, where each
is a sequence of complex numbers, to the sequence of complex numbers
|
(4.5)
|
5 Formalities
With all these definitions and notations in place, we can finally formulate the conclusion of the discussion in § 2 ; proofs will be supplied later.
By this last statement we mean that
converges absolutely for each ornamented tree
, and that if its sum is denoted by
then the resulting series
likewise converges.S All but the last sentence follows from the discussion in § 2 and the definitions in §§ 3 ,section:treeops.
The operators
and coefficients
were defined so that the following holds automatically.
Lemma 5.2.
There exists
with the following property. Let
be any numerical sequence and define
. Suppose that the infinite series defining
converges absolutely and uniformly for all
and that its sum is
, uniformly for every ornamented tree
.
Define
to be the sequence
. Then
satisfies the integral equation 2.6 for
. Moreover the function
defined by
is a solution of the modified Cauchy problem NLS* in the corresponding sense 2.7 .
The main estimate in our analysis is as follows.
Proposition 5.3.
Let
. Then for any exponent
satisfying also
, there exist
and
such that for all ornamented trees
and all sequences
,
|
(5.3)
|
Propositions 5.1 and 5.3 and Lemma 5.2 will be proved in subsequent sections. Together, they give:
Corollary 5.4.
Let
. For any
there exists
such that the solution mapping
for the modified Cauchy problem NLS* , initially defined for all sufficiently smooth
, extends by uniform continuity to a real analytic mapping from
to
.
We emphasize that analytic dependence on
is not asserted.
6 Tree coefficient bound
-
Proof of Lemma 4.1 .
The first conclusion of the lemma holds simply because
. The proof of the main conclusion 4.4 proceeds recursively. In step 1 we integrate with respect to
for certain nodes
, holding fixed all other coordinates
in the integral defining
. Specifically, we hold fixed the coordinate
whenever at least one child of
is not terminal. We also fix
for every simple node
having only terminal children.
The former coordinates
, and underlying nodes
, are said to be temporarily fixed; the latter coordinates and nodes are said to be permanently fixed. No other coordinates are fixed at this step.
When
there is nothing to prove. Otherwise there must always exist at least one node, all of whose children are terminal If there exists such a node which is also general, then at least coordinate
is not fixed. The subset, or slice, of
defined by setting each of the fixed coordinates equal to some constant is either empty, or takes the product form
, where
denotes the parent of
. Integrating over this slice with respect to all of the non-fixed coordinates thus yields
where the first product is taken over all fixed
, and the second over all remaining non-fixed
.
None of the quantities
can vanish in step 1, since a general node having only terminal children can never be exceptional, by 3.5 . Therefore the preceding expression equals
This may be expanded as a sum of
terms, where
is the number of non-fixed nodes in
. Each of these terms has the form
for some numbers
.
The other possibility in step 1 is that
, but every nonterminal node that has only terminal children is simple. In that case all coordinates
are fixed at step 1, no integration is performed, and we move on to step 2.
We now carry out step 2. If a node
was permanently fixed at step 1 then it remains fixed for all subsequent steps; we never integrate with respect to
. More generally, any node that is permanently fixed at any step remains fixed through all subsequent steps.
On the other hand, once we've integrated with respect to some
, then the node
is also removed from further consideration. All other nodes remain active, including those temporarily fixed in step 1. Denote by
the set of all nodes that are active after the completion of step 1.
is itself a tree. There is an associated subset
of
, defined by the inequalities
whenever
, and also by
if
and
was permanently fixed in step 1. To each node
is associated a modified phase
, defined to be
, where the sum is taken over all
such that we integrated with respect to
in the first step.
A node
is permanently fixed at the second step if
is terminal in
and satisfies
. A node
is temporarily fixed at the second step if
is not terminal in
. We now integrate
over
with respect to
for all
that are neither temporarily nor permanently fixed. As in step 1, this integral has a product structure, and
terms are obtained, where
is the number of variables with respect to which we integrate.
In step 3 we consider the tree
consisting of all
that were temporarily fixed in step 2. Associated to
is a set
, and associated to each node
is a modified phase
, the sum being taken over all
such that
was not fixed in step 2. A node
is then permanently fixed if it is terminal in
and
.
is temporarily fixed if it is not terminal in
. We then integrate with respect to
for all
that are neither temporarily nor permanently fixed.
This procedure terminates after finitely many steps, when for each node
, either
has become permanently fixed, or we have integrated with respect to
. This yields a sum of at most
terms. Each term arises from some particular choice of the parameters
, and is expressed as an integral with respect to
for all nodes
that were permanently fixed at some step; the vector
indexed by all such
varies over a subset of
where
is the number of such
. At step
, each integration with respect to some
yields a factor of
, multiplied by some unimodular factor; recall that
, since otherwise
would have been permanently fixed.
Thus for each term we obtain an upper bound of
, where the product is taken over all nonexceptional nodes
; this bound must still be integrated with respect to all
where
ranges over all the exceptional nodes. Each such coordinate
is restricted to
. Thus we obtain a total bound
|
(6.1)
|
where for each
,
is the number of exceptional nodes encountered in this procedure, that is, the number of permanently fixed nodes, and where for each
,
denotes the product over all nonexceptional nodes
that are nonexceptional with respect to
. □
7 A simple
bound
This section is devoted to a preliminary bound for simplified multilinear operators. For any tree
and any sequences
, define
|
(7.1)
|
The notation
indicates that the sum is taken over all indices
that satisfy 3.4 as well as
; the restrictions 3.5 and 3.6 are not imposed here.
Lemma 7.1.
For any tree
and any sequences
|
(7.2)
|
with equality when all
are nonnegative.
-
Proof.
Recall that for some nonnegative integer
,
,
, and
.
Consider the set
whose elements are
together with all
such that
and
. Thus
. Define
|
(7.3)
|
Consider the
-linear mapping
from
to
defined so that
has coordinates
and all
.
and
are well-defined linear functionals of
, because given the quantities
for all
,
can be recovered for all other
via the relations 3.4 , by ascending induction on
. We claim that
is invertible. Indeed, from the quantities
and all
with
and
,
can be recovered for all
by descending induction on
, using again 3.4 at each stage. For instance, at the initial step,
for
, and then
can be recovered via 3.4 . Thus
is injective, hence invertible.
By descending induction on nodes it follows in the same way from 3.4 that
satisfies a certain linear relation of the form
|
(7.4)
|
where each coefficient
equals
. By the conclusion of the preceding paragraph, this can be the only relation to which
is subject; the sum defining
is taken over all
satisfying this relation. Therefore
equals the summation over all
and all
, without restriction, of
. The lemma follows. □
Corollary 7.2.
For any ornamented tree, the sum defining
converges absolutely for all
whenever all
, and the resulting sequence satisfies
|
(7.5)
|
There is no bound for
in terms of the quantities
for
. It is the additional factors
in the second tree coefficient bound 4.4 , reflecting the dispersive character of the partial differential equation, which make possible estimates in terms of weaker
norms.
-
Proof of Proposition 5.1 .
Consider any tree
and associated function
|
(7.6)
|
for
, with
, under the assumption that for each
,
belongs to
. Here for each
,
is the sequence whose components are
.
Let each node
be designated as either finished or unfinished. Assume that for each finished node,
is independent of
, while for each unfinished node, either the sequence-valued function
satisfies the integral equation
(7.7)
y
u
(
t
,
n
)
=
y
u
(
0
,
n
)
−
i
ω
∫
0
t
|
y
u
(
s
,
n
)
|
2
y
u
(
s
,
n
)
d
s
+
i
ω
∑
j
−
k
+
l
=
n
*
∫
0
t
y
u
(
s
,
j
)
y
¯
u
(
s
,
k
)
y
u
(
s
,
l
)
e
i
σ
(
j
,
k
,
l
,
n
)
s
d
s
,
or its complex conjugate satisfies this same equation.
The
hypothesis guarantees that if we substitute the right-hand side of 7.7 for
in 7.6 for each unfinished node, then an absolutely convergent integral and sum are obtained. Thus we may interchange the outer integral with the sums. What results is a finite linear combination of expressions of the same character as 7.6 , each associated to a larger tree
. At most
such expressions are obtained, and each is multiplied by a unimodular numerical coefficient. Each non-terminal node of
is a non-terminal node of
, each finished node of
remains a terminal node of
, and each unfinished node of
becomes a non-terminal node of
, each of whose three children may independently be either finished or unfinished. An unfinished node
of
gives rise either to a simple node or a general node of
, depending on which of the two trilinear terms of 7.7 is substituted for
in 7.6 .
This discussion justifies the formal derivation of the expansion in § 2 . It follows by recursion that for any solution
of the modified Cauchy problem NLS* in
for sufficiently large
, the associated coefficients
are given by the absolutely convergent infinite power series 5.1 , 5.2 for all sufficiently small
. □
8 Tree sum majorants
Definition 8.1.
Let
be an ornamented tree. The tree sum majorant associated to
is the multilinear operator
|
(8.1)
|
Here
and
is initially defined when all
, in order to ensure absolute convergence of the sum.
Definition 8.2.
Let
be a weathered ornamented tree. The associated tree sum majorant is the multilinear operator
|
(8.2)
|
Thus
|
(8.3)
|
the sum being taken over all subsets
. The total number of such subsets is
.
Let
be a weathered ornamented tree. We seek an upper bound for the associated tree sum operator
. The factors
in the definition of
are favorable when
is large; frozen nodes are those for which
is relatively small, and hence these require special attention.
Denote by
any element of
. Let
|
(8.4)
|
is the set of all frozen nodes, so by its definition we have
|
(8.5)
|
with the shorthand notation
introduced earlier. In the remainder of the discussion, we always assume tacitly that
satisfies 8.5 .
This leads to a further decomposition
(8.6)
S
(
T
,
T
′
)
(
y
v
)
v
∈
T
∞
(
n
)
=
∑
Γ
∑
j
∈
J
(
T
,
T
′
,
Γ
)
:
j
v
0
=
n
∏
u
∈
T
0
〈
ρ
u
(
j
)
〉
−
1
∏
w
∈
T
∞
y
w
(
j
w
)
≤
C
|
T
|
∑
N
∏
v
∈
T
′
2
−
N
v
∑
M
∏
u
∈
T
0
\
T
′
2
−
(
1
−
δ
)
M
u
∑
Γ
∑
j
∈
J
(
T
,
T
′
,
Γ
)
:
j
v
0
=
n
∏
w
∈
T
∞
y
w
(
j
w
)
where
and
. The notation in the last line means that the first two sums are taken over all nonnegative integers
as
ranges over
and
over
; the third sum is taken over all
such that
|
(8.7)
|
and the sum with respect to
is taken over all
satisfying the additional restrictions
|
(8.8)
|
|
(8.9)
|
Thus there is an upper bound
for all
.
For any
and any parameter
, for any
,
where
depends only on
. Since the quantity
on the left-hand side equals
, for any
and any
there are at most
ordered pairs
satisfying 8.9 . Here
is an arbitrarily small constant, to be chosen later.
For any frozen node
,
is small relative to
, provided that
is taken to be small in the definition of a frozen node. Therefore we can choose for each combination of parameters
a family
of vector-valued functions
of cardinality at most
such that for any
satisfying 8.7 and any
, there exists
such that for each
and each
,
|
(8.10)
|
Thus for all nonnegative sequences
and all
,
|
(8.11)
|
where
,
, and
|
(8.12)
|
In 8.11 , the second summation is taken over all
satisfying both 8.7 and 8.5 . In 8.12 , the sum is taken over all
satisfying
, 8.8 , 8.9 , and the additional restriction 8.10 . We have finally arrived at our basic building blocks, the multilinear operators
.
Lemma 8.1.
Let
and
. Then for every exponent
satisfying
and
, there exists
such that for every
and for every sequence
,
|
(8.13)
|
-
Proof.
As was shown in the proof of Lemma 7.1 , each quantity
in the summation defining
can be expressed as a function, depending on
, of
together with all
, where
varies over the set
of all nodes that are neither frozen nor terminal, and
varies over
. More precisely,
equals
, where
is some function of all these
.
can thus be rewritten as
. If every
is held fixed, then as a function of
, this product belongs to
for
with bound
, by Hölder's inequality.
The total number of terms in the sum defining
is the total possible number of vectors
where
ranges over
and
over
. The number of such pairs for a given
is
, since
. Thus in all there are at most
terms. Minkowski's inequality thus gives the stated bound. □
8.13 is a satisfactory bound, but it must be summed over all
. An upper bound for the number of such functions
is, roughly speaking,
times the product over all
of
. However, this does not quite make sense since
is a function of
, which we wish to allow to vary while
remains fixed. Thus a correct upper bound is
|
(8.14)
|
where
for
and
for
, and the maximum is taken over
.
A difficulty now appears. For each
we have a compensating factor of
, but there is no upper bound whatsoever for the ratio
. Thus the factor gained for a given
cannot compensate for the factor lost for that same node.
However in aggregate the factors gained compensate for those lost, as will now be shown.
Lemma 8.2.
For any
there exists
such that uniformly for all
,
|
(8.15)
|
-
Proof.
If the constant
in the definition 3.7 of a frozen node is chosen to be sufficiently small, then any frozen node
has a child
such that
. Consider any chain
of nodes such that
is the parent of
for each
(
is called the
-th generation ancestor of
),
is frozen for all
,
is either not frozen or is terminal, and
. Then
; hence
.
If
is terminal then
by definition, whence the inequality
forces
for all
, as well. This means that
. In particular, this holds for
, so the factor
will be harmless in our estimates. We say that a node
is negligible if there exists such a chain, with
for some
.
For each nonnegligible frozen node
, choose one such chain with
, thus uniquely specifying
and
as functions of
; we then write
. Given
and
, there can be at most one
such that
and
is the
-th generation ancestor of
, simply because any node has at most one
-th generation ancestor. Now taking the product only over nonnegligible nodes
on the left-hand side,
|
(8.16)
|
since3
each factor
in the first product is majorized by
in the second product, where
and
is the
-th generation ancestor of
. This is not so for negligible nodes, but they contribute at most
to the left-hand side so the conclusion remains valid for the full product. Thus by choosing
so that
, since
, we obtain
. □
-
Conclusion of proof of Proposition 5.3 .
Combining the preceding two lemmas gives
|
(8.17)
|
for arbitrarily small
, provided
. Since
, it follows that
|
(8.18)
|
On the other hand, Lemma 7.1 gives a uniform
norm bound of
for the summation over all
. Thus if
and
we may interpolate to find that there exists
depending on
but not on
such that
|
(8.19)
|
Taking into account the factors
in 8.11 , summing over
as well as over all subsets
yields a convergent series and completes the proof of Proposition 5.3 . □
9 Loose ends
We may reinterpret the sum of our power series 5.1 , 5.2 as a function via the relation
with
defined by
, and will do so consistently without further comment, abusing notation mildly by writing
.
Lemma 9.1.
Let
. For any
there exists
such that for any
with norm
, the element
defined by 5.1 , 5.2 is a limit, in
norm, of smooth solutions of NLS* .
-
Proof.
All of our estimates apply also in the spaces
defined by the condition that
, provided that
and
. This follows from the proof given for
above, for the effect of working in
is to introduce a factor of
in the definition of the tree operator. The relation 3.4 ensures that
, whence
, so the estimates for
apply directly to all
. More generally, if
is equipped with the norm
then all estimates hold uniformly in
and
.
Given
and any initial datum
satisfying
with the additional property that
for all
, we may choose
so that
;
depends on
but not on
. Thus the infinite series converges absolutely and uniformly in
if
and in
if
, where
depends only on
, not on
. By Lemma 5.2 , the series sums to a solution of NLS* in the sense 2.7 ; but since the sum is very smooth as a function of
(that is, its Fourier coefficients decay rapidly) this implies that it is a solution in the classical sense. Given an arbitrary
satisfying
, we can thus approximate it by such special initial data to conclude that
is indeed a limit, in
, of smooth solutions. □
-
Proof of Proposition 1.4 .
Let
be given, let
. We aim to prove that the nonlinear term
has an intrinsic meaning as
in the sense of distributions in
. Forming
is of course not the same thing as forming
.
Define
. Denote also by
the operator that maps a sequence-valued function
to
where
if
, and
otherwise. It suffices to prove that
|
(9.1)
|
converges in
norm as
, uniformly for all
, to
Convergence in the distribution sense follows easily from this by expressing any sufficiently smooth function of the time
as a superposition of characteristic functions of intervals
.
Now in the term
, the integral may be interchanged with the sum since the truncation operators restrict the summation to finitely many terms. Expanding
out as infinite series of tree operators applied to
, we obtain finally an infinite series of the general form
where
is a finite linear combination of
tree sum operators, with coefficients
, applied to
just as before, with the sole change that the extra restriction
for
for indices corresponding to children of the root node is placed on
in the summation defining
for each tree
.
Since we have shown that all bounds hold for the sums of the absolute values of the terms in the tree sum, it follows immediately that this trilinear term converges as
.
Convergence for the other nonlinear term is of course trivial. Likewise it is trivial that
and
, by linearity. □
This reasoning shows that the limit of each term equals the sum of a convergent power series, taking values in
, in
.
Given
, there exists
for which we have shown that for any
satisfying
, our power series expansion defines
, as an
-valued analytic function of
. Moreover for any
, both cubic terms in the integral equation 2.6 are well-defined as limits obtained by replacing
by
, evaluating the resulting cubic expressions, and passing to the limit
.
Lemma 9.2.
Whenever
, the function
defined as the sum of the power series expansion 5.1 satisfies the integral equation 2.7 when the nonlinear terms in 2.6 are defined by the limiting procedure described in the preceding paragraph.
-
Proof.
This follows by combining Lemma 5.2 with the result just proved. □
-
Proof of Proposition 1.3 .
Let
. If
, and if
is the solution of the Cauchy problem NLS* for the modified linear Schrödinger equation with initial datum
, then
is expressed as
where the
-th Fourier coefficient of
equals
with
. According to Proposition 5.3 ,
whenever
and
. Summing over
yields the conclusion. □
References
-
H. A. Biagioni, A nonlinear theory of generalized functions, Second edition. Lecture Notes in Mathematics, 1421. Springer-Verlag, Berlin, 1990.
-
J. Bourgain, Fourier transform restriction phenomena for certain lattice subsets and applications to nonlinear evolution equations. I. Schrödinger equations, Geom. Funct. Anal. 3 (1993), no. 2, 107–156.
-
N. Burq, P. Gérard, and N. Tzvetkov, An instability property of the nonlinear Schrödinger equation on
, Math. Res. Lett. 9 (2002), no. 2-3, 323–335.
-
M. Christ, Nonuniqueness of weak solutions of the nonlinear Schrödinger equation, preprint February 2005.
-
M. Christ, J. Colliander, and T. Tao, Instability of the periodic nonlinear Schrödinger equation, preprint, math.AP/0311227.
-
J.-F. Colombeau, Multiplication of distributions. A tool in mathematics, numerical engineering and theoretical physics, Lecture Notes in Mathematics, 1532. Springer-Verlag, Berlin, 1992.
-
A. Grünrock, abstract in Mathematisches Forschungsinstitut Oberwohlfach report 50/2004.
-
, An improved local wellposedness result for the modified KdV equation, IMRN 2004, no. 61, 3287-3308.
-
T. Kappeler and P. Topalov, Global Well-Posedness of KdV in
, preprint.
-
, Global Well-Posedness of mKDV in
, preprint.
-
Y. Tsutsumi,
-solutions for nonlinear Schrödinger equations and nonlinear groups, Funkcial. Ekvac. 30 (1987), no. 1, 115–125.
Michael Christ, Department of Mathematics, University of California, Berkeley, CA 94720-3840, USA E-mail address : mchrist@math.berkeley.edu