1. Introduction
To make the introduction as accessible as possible to the reader, we put the topic of the present paper in a historical context, thereby highlighting the results relevant for its content. The study of integrable or exactly solvable systems has a long history. Classical mathematicians such as Euler, Lagrange, Liouville, Riemann and Poincaré amongst many others, investigated nonlinear systems which could be integrated more or less explicitly. We start at the beginning of the area of soliton theory of which we give a short description based on [
1]. The first recorded observation of a solitary wave is the one made by Russell in 1834 [
2]. He observed along a narrow channel in Scotland that, when a barge was suddenly stopped, the water mass put in motion by the boat formed at the prow of the barge a solitary wave that moved forward for quite some time at a constant speed and maintaining its height. After returning from this trip, he started careful experiments with water waves to classify them and after some time he came empirically to a number of results: the existence of solitary waves, a formula for their speed and a nonlinear differential equation they had to satisfy. The astronomer Airy, based on his own work on water waves, denied the existence of solitary waves, and this led to bitter disputes. This controversy around the validity of Russell’s claims went on till 1895, when Korteweg and his student de Vries published the paper [
3] on propagation of shallow water waves in a narrow rectangular canal. It settled the controversy entirely in favor of Russell. In their paper, they argued that the height
of the waves in the canal should satisfy the differential equation
where
x is a space coordinate along the canal and
t a time coordinate. Moreover, they gave a set of solitary wave solutions of this equation, the so-called
cnoidal waves. Since then, Equation (
1) has been called the KdV equation.
Interest in the KdV equation reignited when Zabusky and Kruskal [
4] found remarkable interaction properties for the solitary wave solutions to KdV. They coined the name
solitons for these solutions. Soon after that, Gardner, Greene, Kruskal and Miura [
5] introduced
inverse scattering transform or IST to solve the Cauchy problem for KdV, which turned out to be applicable to many more equations. As a result, the research on both theoretical and practical aspects of these equations intensified. E.g., in 1995, at the centennial of the eponymous KdV paper, the practical applications [
6] of the KdV equation alone were recognized in various fields: fluid mechanics, optics, oceanography, plasma physics, astronomy and electrical transmission lines, among others.
An alternative formulation of the KdV equation was found by Lax [
7]. He considered the Schrödinger operator
and the third order operator
, where
u is a function depending on
x,
t and possibly other parameters and
∂ denotes the partial derivative
. Then, a direct calculation yields that their commutator is a zero-th order operator in
∂
and we see that the evolution equation for
is equivalent to
u satisfying the KdV equation. Equation (
2) was soon called the
Lax form of the KdV. This form suggests that
is obtained by conjugating a
t-independent operator like
with a
t-dependent one and we will see it coming back at all kinds of integrable equations. For example, the Lax form of the nonlinear Schrödinger equation was found by Zacharov and Shabat in [
8] and NLS was the starting point in [
9] to treat Hamiltonian methods for soliton equations. It laid the foundation for the quantum version of IST developed by the Fadeev school in St. Petersburg, see [
10]. This also brought solvable models from statistical mechanics into the picture and likewise quantum groups made their appearance in the integrable world. A rich collection of contributions to these topics and integrability can be found in [
11].
The third nonlinear equation that needs to be mentioned is a two-dimensional version of KdV that was used in plasma physics by Khadomtsev and Petviashvilii and is therefore called the KP equation. For a function
, it reads
Let
be another function,
the operator
and
the operator
. Consider the following linear system for
:
Then, the compatibility of this system is
. Eliminating
v from these equations gives you that
u satisfies the KP Equation (
3). Krichever searched for an analogue for the KP equation of the results in [
12,
13,
14,
15] for the KdV equation and succeeded [
16,
17]. The algebraic geometric data he needed were the following: a complete irreducible algebraic curve
X, a rank 1 torsion free coherent sheaf
over
X, a non-singular point
, a local coordinate
around
such that
z is an isomorphism between a closed neighbourhood
of
and the unit disc around infinity on the Riemann sphere and finally a trivialization
of
over
. If
X is nonsingular, then
X is a compact Riemann surface and
is a line bundle over
X.
The fourth type of system that exhibit soliton-like behavior was found by M.Toda [
18]. He was inspired by results of Fermi [
19] and Ford [
20] and his goal was to elucidate certain characteristic features of nonlinear waves, using nonlinear lattice models. He found lattice models that possess multi-soliton solutions, periodic waves and solvability of the initial value problem with IST. Since then, they carry his name. We mention one of them, the
finite Toda lattice. For more examples, we refer the reader to [
21]. The finite Toda lattice describes the motion of
n particles, all with mass equal to 1, on a line with only nearest neighbor interaction. Let
denote the position of the
j-th particle and
its momentum. The equations of motion of the
n particles are:
where we put
and
. Introduce the new coordinates
and recall that
. Then, the equations of motion (
5) transform into
It was H. Flaschka [
22] who showed that Equation (
6) can be written in Lax form. Consider the following
-matrices
Then, one verifies directly that Equation (
6) amounts to the Lax equation
With the help of this equation, one shows that the eigenvalues of
L are time-independent and polynomials in
and
. Moreover, the functions
are
n linear independent integrals for the vector field
for the Hamiltonian
and
are in involution with respect to the standard Poisson bracket
on
, i.e.,
. For proofs, we refer the reader to Chapter 3 in [
23]. Note that the matrix
B in the Lax form of the finite Toda lattice is the anti-symmetric part of
L in the splitting of the
-matrices in an anti-symmetric part and a lower triangular part. Hence, (
7) is a Lax equation coupled to this decomposition. An integrable hierarchy, coupled to an infinite dimensional variant of this decomposition, is described in [
24]. This decomposition is different from the one on which the discrete KP hierarchy [
25] is based.
Gelfand and Dickey realized that the third order operator
A occurring in the Lax form (
2) of KdV was nothing but the differential operator part of the third power of the square root
of
in the pseudo-differential operators Psd in
∂. Here, Psd is an extension of the differential operators with coefficients, functions on which
∂ acts via Leibniz. It consists of possibly infinite series
satisfying
and
and it has the property that each monic element of Psd of order
possesses an
m-th root in Psd. Then, it was clear how to generalize it. In [
26], the authors considered, instead of the Schrödinger operator
, a monic differential operator
of order
in
∂ of the form
and for all
the monic differential operator
consisting of the differential operator part of the
m-th power of the
n-th root
of
in the pseudo-differential operators. Take the infinite set of Lax equations for
Gelfand and Dickey showed that this system of equations is compatible and determined its Hamiltonian structure. For
, this system is usually called the
KdV hierarchy and for the generalization one uses the name
n-th Gelfand–Dickey hierarchy or
n-KdV hierarchy. The strict version of the n-th Gelfand–Dickey hierarchy is a wider deformation of
and is treated in [
27].
A next big step was the appearance of a long series of papers by members of the Sato school in Kyoto on transformation groups for soliton equations. They corresponded with various infinite dimensional groups and were full of all kinds of intriguing ideas and connections. We focus on the main example: the KP hierarchy, that is the one asssociated with the group
. Date, Jimbo, Kashiwara and Miwa introduced it in [
28] and considered elements
L in Psd of the form
and we see
as a deformation of the commutative algebra
. Note that any
is of this form. In that case, we are dealing with a deformation of
by conjugating with the dressing operator
K. We assume that the coefficients in the pseudo-differential operators are differentiable with respect to the parameters
and that each
commutes with
∂. Let
be the differential operator part of
, then the evolution equations for
L are
These Lax equations are equivalent to the zero curvature relations for
Now, the first two nontrivial powers of
L start as follows
Hence,
and
and from the calculation at the introduction of the KP equation it follows that
is a solution of the KP equation. This justifies the name of the hierarchy. Each of the Gelfand–Dickey hierarchies is a subsystem of the KP hierarchy. Any solution
L of the KP hierarchy for which
is a differential operator delivers the solution
of the
n-th Gelfand–Dickey and in terms of deformations this corresponds to deforming the commutative algebra
inside the differential operators. The geometric picture sketched in [
28] is that the group
acts in the unspecified space of
-functions, that the
-orbit of the vacuum is the Sato Grassmannian and this
-orbit is the solutions of the KP hierarchy. This could use some clarification. In a later paper, the authors specified the space of
-functions as the polynomials in
and the picture became better, but not good enough for, e.g., the solutions found by Krichever. Segal and Wilson followed a different path in [
29] to obtain solutions of the KP hierarchy. They used the linearization of KP. This is a set of relations in the Psd-module of oscillating functions
If you find for
L and all the
a suitable
in that module such that these relations hold, then
L is a solution of the KP hierarchy. Segal and Wilson introduced a Grasmann manifold based on a space of Hilbert–Schmidt operators from which they constructed solutions for the KP hierarchy and for each point of this variety they introduced a
-function that relates to the wave function in the way, the Sato school predicted. The solutions found by Krichever fall also within this framework. The strict KP hierarchy is a wider deformation of
introduced in [
30]. The geometric picture is a flag variety covering the Grassmann manifold of Segal and Wilson and can be found in [
31]. In this case, one needs two
-functions to describe the relevant wave function [
32]. It will not come as a surprise that the strict versions of the Gelfand–Dickey hierarchies are subsystems of the strict KP hierarchy.
We have seen how research in integrable systems has developed from single equations to infinite dimensional integrable hierarchies. They play an important role in both mathematics and theoretical physics and common grounds are a big stimulus for cooperation and success. We mention a few areas in physics where they play an important role. First of all, they often occur in matrix models at the application of large
N-techniques, see, e.g., Refs. [
33,
34,
35]. The next subject combines a wide range of disciplines: two-dimensional quantum field theory, intersection theory on moduli spaces of Riemann surfaces, integrable hierarchies, matrix integrals and random surfaces to name a few. At the beginning of the 1990s, Edward Witten [
36] started a program [
37] to search for a relationship between random surfaces and the algebraic topology of moduli space and formulated a series of conjectures based on his findings. The first of these conjectures was proved by Kontsevich [
38], who showed that the string partition function is a
-function of the KdV hierarchy. Another topic is topological strings on Calabi–Yau geometries. It provides a unifying picture connecting non-critical (super)strings, integrable hierarchies and various matrix models, see [
39,
40].
In [
41], we studied integrable hierarchies related to deforming various commutative subalgebras of the
-matrices with coefficients in
k,
or
. We also constructed solutions of these hierarchies by conjugating the original matrices with matrices from operators of the form identity plus a Hilbert–Schmidt operator. In operator terms, this is a mild perturbation as Hilbert–Schmidt operators are compact. In a later paper [
42], we took the main example of a commutative subalgebra and studied the scaling properties of the corresponding hierarchy. Here, we stick to the basic example and present a far wider class of dressing matrices for both hierarchies with matrices corresponding to bounded operators on suitable Banach spaces. To be more concrete, we first need some notation.
Let
R be a commutative
k-algebra over the field
k. We write
for the
-matrices with coefficients from
R and similarly
for the space of
-matrices with coefficients from
R. The transpose
of any matrix
or any
matrix
A with coefficients from
R is defined as in the finite dimensional case. Let
V be the
R-module of all
-matrices with coefficients from
R, i.e.,
Inside
V, we consider the
R-submodule
The space
is a free
R-module with
as the basis, where
is the vector with the
i-th coordinate equal to one and the remaining ones equal to zero. For each
and each
, the product
is well defined and determines a vector in
V. Hence, if we write
then
is an
R-linear map in
. Inside
, we have the basic matrices
and
, whose matrix entries, in Kronecker notation, are given by
It is convenient to use the notation
for an
. A central role in this paper is played by the shift matrix
S, its transpose
and their powers, where
S is the matrix
that corresponds to the shift operator
defined by
The way the two deformations of the commutative algebra are performed is by conjugating them with invertible lower triangular -matrices with coefficients from R. At one deformation, we conjugate with lower triangular matrices with the identity on the central diagonal. The evolution equations of the deformed generators are Lax equations and are determined by the decomposition of a matrix in in their upper triangular and strict lower triangular parts. The evolution equations of the deformed generators in this case are called the Lax equations of the -hierarchy.
For the second wider deformation, the central diagonal is merely assumed to be invertible. Also in this case, the evolution equations of the deformed generators are Lax equations. Only now they are based on the splitting of matrices in their strict upper triangular and lower triangular parts. The evolution equations of these deformations of are called the Lax equations of the strict -hierarchy.
In the case of the -hierarchy, the dressing matrix of the deformation turns out to be unique, and in the strict case it is determined up to a multiple of the identity. The uniqueness of the dressing operator enables one to prove directly the equivalence of the Lax form of the k[S]-hierarchy with a set of Sato–Wilson equations. There exists an analogue of the Sato–Wilson equations for the strict -hierarchy. It always implies the Lax equations of this hierarchy and it suffices for the geometric solutions that we produce further on. However, we show how they can fail.
Solutions of both hierarchies are constructed by producing wave matrices in the linearization module of each hierarchy. Therefore, we recall the essentials of this approach. We start the last section with the description of a family of Banach spaces B that all have a countable Schauder basis; they have the approximation property and the elements with only a finite number of coordinates with respect to the Schauder basis are dense in B. For these spaces, we show how LU factorizations in can be used to construct for a nonempty open subset of solutions of the -hierarchy and its strict version. We show that the dressing matrix coefficients of these solutions are quotients of analytic functions on the group of commuting flows of the hierarchies. Next, we introduce a subgroup of that is the analogue of the restricted linear group that was used in cases in which B is a Hilbert space to construct solutions of the KP hierarchy. We conclude by showing that is contained in this open subset and that leads to a description of each set of solutions in terms of a homogeneous space of .
The content of the various sections is as follows:
Section 2 describes the scene of the deformations, the algebra
, the maximality of
and the properties required later on. The next section is devoted to a description of the two deformations; it contains a discussion of the Lax equations they have to satisfy and we describe there the link with their Sato–Wilson form. The form of the relevant
-module, the equations of the linearization and a characterization of the special vectors, called wave matrices, can all be found in
Section 4. In the last section, we present the role of LU factorization in constructing solutions of both hierarchies, we discuss properties of the dressing matrix coefficients and prove that
is contained in the open set of
, where LU factorization is effective in producing solutions of the hierarchies.
3. The -Hierarchy and Its Strict Version
In this section, we discuss the two deformations of
that we consider and the evolution equations we want the deformations of
S to satisfy. At the first deformation, each
in
is deformed into
, where
is an element of the form
One directly checks that any element
, with
, has this form and we call
a
-
deformation of
S. We also call
U the
dressing matrix of
. At the second deformation, we transform each matrix
into
, where
is an element of the form
Also in this case, one can easily verify that any matrix
, with
, possesses the form (
18) and therefore it is called a
-
deformation of
S. Likewise, we also call
P the
dressing matrix of the deformation
. Moreover, we showed in [
42].
Lemma 2. Conversely, there holds for the deformations (17) and (18) - (a)
Any of the form (17) can uniquely be written in the form with , i.e., is a -deformation of S. - (b)
Any of the form (18) can be written in the form , where is unique and is determined up to a factor from . In particular, is a -deformation of S.
Next, we discuss the evolution equations that an -deformation of S has to satisfy and those for a -deformation of S. Thereby, each is seen as an infinitesimal generator of a flow. In that light, we assume in both cases that R is equipped with a set of commuting k-linear derivations , where each should be seen as an algebraic substitute for the derivative with respect to the flow parameter corresponding to the flow generated by each . By letting each act coefficient-wise on matrices in , we obtain a set of derivations of , also denoted by . The data we call a setting for both deformations.
For each
-deformation
of
S and all
, we consider the cut-offs
Note that, since all
commute, the
satisfy for all
where the right-hand side is of degree
or lower, like
. This shows that it makes sense to unite the following set of Lax equations for the
in one combined system, the so-called
-hierarchy:
It suffices to prove the equations just for
. For, since
and
are both
k-linear derivations of
, all basis elements
of the deformation
of
satisfy the same Lax equations. Equation (
20) itself is called the
Lax equations of the -hierarchy. Note that the Lax Equation (
20) shows that the action of each
on the coefficients of
expresses each of them in a polynomial expression of the coefficients of
. Any
-deformation
of
S in
that satisfies all Equation (
20),
, is called a
solution of the
-hierarchy in the setting
. Note that in each setting there is at least one solution of the
-hierarchy, namely
, the
trivial solution of the
-hierarchy. We can express the conditions when a
-deformation
is a solution of the
-hierarchy, in terms of
U. For, there holds
Lemma 3. Any with is a solution of the -hierarchy if and only if U satisfies the relations: for all Proof. Since
, we obtain for
that
If
U satisfies (
21), then
yields the Lax equations for
. Conversely, if
is a solution of the
-hierarchy, then
commutes with
and thus
commutes with
S. The element
only has strict negative diagonals and Lemma 1 implies that
and this proves the claim. □
Since the relations (
21) are the analogue of the Sato–Wilson equations for the KP hierarchy, we call them the
Sato–Wilson form of the -hierarchy. This is the form that comes out of the approach in
Section 4 and that we will use at the construction in
Section 5.
Remark 1. Let be the zero-th diagonal of , then the zero curvature form of the -hierarchy, see [42], implies that the commuting diagonal matrices satisfy the compatibility conditions Next, we treat the evolution equations for the
-deformations
. In that case, we consider for each
the strict cut-off
Since all the
commute, there holds
which shows that the
have the common property that the commutator with
has degree
m or lower. The same holds for the matrix
, so it makes sense to unite the following set of Lax equations for the
in one combined system
Because of the form of the
and the similarity with the Lax Equation (
20), we call this system the
strict -hierarchy. Equation (
24) itself is called the
Lax equations of the strict -hierarchy. Note that also in the strict case the Lax Equation (
24) shows that the action of each
on the coefficients of
expresses each of them in a polynomial expression of the matrix coefficients of
. Any
-deformation
of
S in
that satisfies Equation (
24) is called a
solution of the strict
-hierarchy in the setting
. By the same argument as for the
case, it suffices to prove Equation (
24) for
, since all basis elements
of the wider deformation
of
satisfy the same Lax equations. Note that in each setting there is at least one solution of the strict
-hierarchy, namely
, the
trivial solution of the strict
-hierarchy.
Next, we discuss the Sato–Wilson equations for a
-deformation
, with
, where
and
. The analogues for
P of Equation (
21) are
These are called the
Sato–Wilson equations for
P. If
P satisfies Equation (
25), then
yields the Lax Equation (
24) for
. Hence, Equations (
25) are sufficient for
to be a solution of the strict
-hierarchy. They have a characterization in terms of the components
d and
UProposition 1. Take a -deformation , with , and let . Then, the Sato–Wilson equations for P are equivalent with being a solution of the -hierarchy and d is a solution of the systemwhich is a compatible systems of equations if is a solution of the -hierarchy, see (22). Proof. We express
in the components
d and
U
Since
has only strict negative diagonals, the zero-th diagonal of
is
and the sum of all strict negative diagonals of
is
. Similarly, we decompose
Hence, the zero-th diagonal parts of
and
are equal if and only if
d satisfies the system (
26) and the equality of the sum of all their strict negative diagonals is equivalent with the Sato–Wilson equations for
. □
Remark 2. Note that the choice of the dressing operator P of is crucial in (25). Assume that for Equation (25) holds. Then, is anyway a solution of the strict -hierarchy. We have seen that all are also dressing operators of . The right-hand side of Equation (25) is the same for all dressing operators . However, the left-hand side is equal toand that is only equal to the right-hand side if for all . In other words, r should be a constant for all the . Hence, if you would pick out a dressing operator with for some i, then the Sato–Wilson equations would not hold for that . Nevertheless, the Lax equations of the strict KP hierarchy hold for . 4. Wave Matrices
Let
be a setting where one looks for solutions to both hierarchies. The homogeneous spaces of solutions of both hierarchies are constructed by producing special vectors, called wave matrices, in a suitable
-module that we briefly recall. We start with the upper triangular matrix
Here,
t is the shorthand notation for
,
and each
is a homogeneous polynomial of degree
j in the
, where each
has degree
i. Note that
commutes with
S and satisfies for all
,
. Recall that each
was the algebraic substitute on
R for the partial derivative with respect to the flow parameter of
. Therefore, we write
. The
-module that we need consists of formal products of a perturbation factor from
and
. The products will be formal, for making sense out of the product of a matrix from
and the matrix
as a matrix requires convergence conditions and we want to give an algebraic description of the module. Consider therefore the space
consisting of the formal products
Addition (resp. scalar multiplication) is defined on
by adding up the perturbation factors of two elements (resp. by applying the scalar multiplication on the perturbation factor). Something similar is done with the
-module structure: for each
, define
Clearly, this makes
into a free
-module with generator
. Besides the
-action, each
also acts on
by the formula
Here, we impose a Leibniz rule on the formal product. Finally there is a right-hand action of
S on
. Since
S and
commute, we can define it by
Analogous to the terminology in the function case, see, e.g., Ref. [
28], we call the elements of
oscillating -matrices. Note that any
with
invertible is a generator of the free
-module
. Examples are the choices
resp.
in which case we call
an oscillating
-matrix of type
resp.
. With the three actions just defined, we can introduce inside
two systems of equations leading to solutions of the respective hierarchies.
For the
-hierarchy, this system is as follows: consider a
-
deformation of
S in
with the set of projections
. The goal is now to find an oscillating
-matrix
of type
such that in
the following set of equations holds
Since
is a free
-module with generator
, the first equation
implies
and thus
. By Lemma 2, this determines
uniquely. The same argument allows one to translate each
into an identity in
:
Thus, we obtain that
satisfies the Sato–Wilson Equation (
21) and
is a solution of the
-hierarchy. The system (
28) is called the
linearization of the -hierarchy and
a
wave matrix of the -hierarchy. Note that
is the wave matrix corresponding to the trivial solution of the
-hierarchy,
.
In the case of the strict
-hierarchy, we start with a
-
deformation of
S together with the projections
. Now, we look for an oscillating
-matrix
of type
that satisfies in
the following set of equations:
Also,
is a generator of
and again we can translate Equation (
29) into identities in
. Thus, the first equation
becomes
and the second set of equations in (
29) yields the Sato–Wilson Equation (
25) of the strict
-hierarchy. In particular,
is a solution of that hierarchy. The system (
29) is called the
linearization of the strict -hierarchy and a
satisfying this system a
wave matrix of the strict -hierarchy.
For both hierarchies, we use in the sequel a milder property that oscillating
-matrices of a certain type have to satisfy in order to become a wave matrix of that hierarchy. For a proof, see [
41].
Proposition 2. Let be an oscillating -matrix of type in and the corresponding potential solution of the -hierarchy. Similarly, let be an oscillating -matrix of type in with potential solution of the strict version.
- (a)
Assume there exists for each an element such that Then, each and ψ is a wave matrix for the -hierarchy.
- (b)
Suppose there exists for each an element such that Then, each and φ is a wave matrix for the strict -hierarchy.
Since you do not meet formal products of lower triangular -matrices and upper triangular -matrices in real life, the only way to construct wave matrices of both hierarchies is to give an analytic framework, where you can produce well-defined products of such matrices. This is done in the next section.
5. LU Factorizations and Solutions of Both Hierarchies
All the relevant
-matrices that will be produced in the sequel correspond to bounded operators acting on a separable real or complex Banach space with a countable Schauder basis. Since
or
, we have on each
k the standard norm
The Banach spaces
B we will work with are different spaces of
matrices. In all cases the Schauder basis consists of the
introduced in
Section 2. The choice we make for
B is either one of the spaces
of
matrices, defined by
with the norm
or the space
defined by
with the supremum norm
. Any
, the space of all bounded
k-linear maps from
B to itself, corresponds to left multiplication
with an
-matrix
with respect to
, i.e.,
The invertible transformations in
are denoted by
. For example, the operator
as defined in
Section 2 by Formula (
12) maps each of these Banach spaces
to itself and defines an operator in
with operator norm equal to 1.
All the Banach spaces
have two properties in common that we need later on: first of all, they have the approximation property, i.e., the bounded finite dimensional operators in
are dense with respect to the operator norm in the compact operators
in
, see, e.g., Ref. [
43]. Secondly, each vector in
is the limit of a Cauchy sequence of vectors with only a finite number of nonzero coordinates with respect to
.
Now, we come to the description of the role of LU factorization in
for the construction of solutions of both hierarchies. Two decompositions of
play a role in this process. The first splits
as
, with the corresponding matrices
It gives rise to the decomposition of the Lie algebra
as
, where
The Lie groups corresponding to
and
are, respectively
Since the exponential map is a local diffeomorphism from an open neighbourhood of 0 in
resp.
to an open neighbourhood of the identity element in
resp.
, it follows that
is a neighbourhood of the identity in
. As any point
of
can be reached with left multiplication with elements of
and right multiplication with elements of
, the set
is open in
. It is called the
big cell in
with respect to the groups
and
. Since the groups
and
intersect only in the identity, the splitting of an
from
in the product of a unipotent lower triangular matrix and an upper triangular matrix is unique. We use a slightly twisted version of this
LU factorization of
: each
splits uniquely as
The component
we call the
unipotent component of the LU factorization (
32).
The second decomposition is a variation of the foregoing and consists of splitting a
as
, where the matrices of both components are given by
This leads to the decomposition
of
, where
The Lie groups corresponding to
and
are, respectively
Let
denote the diagonal matrices in
. Then, we have
and the big cell
with respect to
and
is also equal to
. This yields an alternative LU factorization for an element in
in the product of a lower triangular matrix and a unipotent upper triangular matrix. We will again use a twisted version of this factorization: each
in
splits uniquely as follows:
The component
we call the
parabolic component of the LU factorization (
35).
Basically, the group of commuting flows that is relevant for both hierarchies is described in the generator
of the module
of oscillating matrices. What needs to be done still is to make a proper choice for the
such that for those parameters
is the matrix of an element in
. We have seen that the operator norm
is equal to 1. Hence, the following choice defines a subgroup
in
:
For each
there holds
, with the polynomials
as in (
27), and
is thus a subgroup of
.
Now, we construct for each
g in the open set
a solution of the
-hierarchy and the strict
-hierarchy. For each
, consider the open subset
of
defined by
The subset
is nonempty if and only if
, which we assume from now on. The appropriate setting in both cases is the algebra
with the derivations
We start with the construction of the solutions of the
-hierarchy. Then, we have by definition for all
that
and thus on the matrix level
Note that all matrix coefficients of
and
belong to
, since the map
is a diffeomorphism between
and
. Equation (
37) leads to the following identity
Clearly,
is an oscillating matrix in
for which the products between the different factors are real. To show that
is a wave matrix for the
-hierarchy, it suffices to prove the property in Proposition 2. Thus, we compute for all
, the matrix
using both the left- and the right-hand sides of expression (
38). We start with the right-hand side. Since for all
,
, we obtain
Now, the matrix
is of the form
with all
. Next, we use the left-hand side of (
38) to compute
. This yields
In this formula, expression
possesses only negative diagonals and
has the form
with all
and
. Combining this with the expression found for the right-hand side gives for all
Thus,
satisfies the conditions in part (a) of Proposition 2 and hence it is a wave matrix of the
-hierarchy. In other words,
is a solution of the linearization of the
-hierarchy. The corresponding solution
of the
-hierarchy is
and
. Note that, since the factor
plays no role in the construction of
, multiplying
g from the right with an element of
does not affect the solution
.
Secondly, we present for a
the construction of the solution of the strict
-hierarchy. We proceed similarly, but now we use the LU factorization (
35) of
. By definition, we have for all
that
and thus on the matrix level
Note that all matrix coefficients of
and
belong to
, since the map
is a diffeomorphism between
and
. Equation (
40) leads to the following identity
Clearly,
is an oscillating matrix in
for which the product between the different factors is real. The idea is again to show that
is a wave matrix for the strict
-hierarchy and that is done by proving property (b) in Proposition 2. Thus, we compute for all
, the matrix
using both the left- and the right-hand sides of expression (
41). We start with the right-hand side. Again all the
are zero; hence,
The matrix
only has strict positive diagonals. Thus, the expression
is equal to a matrix of the form
with all
. Next, we use the left-hand side of (
41) to compute
once more. This yields
In this formula, the expression
does not possess any strict positive diagonals and the matrix
has the form
with all
and
. Combining this with the first expression found yields for all
Thus,
satisfies the conditions in part (b) of Proposition 2 and hence is a wave matrix of the strict
-hierarchy. The corresponding solution
of the strict
-hierarchy is
and
. Also, here the factor
plays no role at the construction of
. Hence, multiplying
g from the right with an element of
does not affect the solution
. For completeness, we resume the foregoing results.
Theorem 1. Let g be an element in the open set .
- (a)
For any γ in , let be the unipotent component of in the LU factorization (32) of . Let the oscillating matrix be defined by Formula (38). Then, is the wave matrix of the -hierarchy with respect to the matrix defined by Formula (39). The solution of the -hierarchy satisfies for all and all that . - (b)
For any γ in , let be the parabolic component of in the LU factorization (35) of . Let the oscillating matrix be defined by Formula (41). Then, is the wave matrix of the strict -hierarchy with respect to the matrix defined by Formula (42). The solution of the strict -hierarchy satisfies for all and all that .
In the sequel, we need the decomposition
, where the subspaces
of
B and their complements
are defined by
Each
splits as follows with respect to
:
Next, we have a more detailed look at the matrix coefficients of the dressing matrices constructed in Theorem 1. Consider
, with
, and a
. Then,
, with
Hence, all the matrix coefficients
belong to the analytic functions
on
. If
, then
and
with
and
. In particular, we have
Using the decomposition (
43) for
and
, we obtain for all
the finite dimensional LU factorization
, which enables you to express each
and
as the quotient of two polynomial expressions in
. This is the case for a number of matrix coefficients that can be seen directly: for example, the first column of
has to equal that of
. So,
The LU factorization of
shows that
In particular, we obtain that for
,
. With these data, one can prove by induction on the size of the matrix that all other matrix coefficients of
and
are determined uniquely and have the mentioned form. This can be seen as follows: consider the matrix
Now, all
are nonzero and we apply the induction hypothesis to the submatrix
to obtain the unique coefficients
and
with
i and
j varying between 0 and
n. Now, we know already
and
so that we merely have to find
and
. For the first set, we have
and since
is invertible, this determines the first set completely and the whole matrix
. The product of the last row of
with the second column of
yields the equation
and that determines
uniquely. From the product of the last row of
with the third column of
one obtains the equation
, which fixes
and so on. Thus, we obtain the remaining coefficients from the second set. This procedure can best be carried out by a computer if you want to see how the coefficients evolve. Note that the matrix
also has the property that all its matrix coefficients are quotients of polynomial expressions in
. Now, we go back to the dressing matrices
and
from Theorem 1. Then,
and
and taking the inverse in both cases maintains the property that we want. So there holds
Theorem 2. The matrix coefficients of the dressing matrices and from Theorem 1 are quotients of polynomial expressions in the analytic functions .
Remark 3. The result in Theorem 2 is for the -hierarchy and its strict version the analogue of the property that the dressing operators for KP and strict KP as constructed in [29,31] have coefficients that depend meromorphically on the group of commuting flows. Let
denote the subspace of compact operators in
. In the group
, we consider
Clearly,
is contained in
. Note that for each
the operator
is a Fredholm operator of index zero. Using the fact that
is a two-sided ideal in
and some properties of Fredholm operators [
44], one shows that
is a group. One can see
as the analogue for the
-hierarchy and its strict version of the restricted linear group
used to construct solutions of the KP hierarchy and its strict version in the Hilbert setting, see [
29,
31]. For there holds
Proposition 3. The group lies in the open set .
Proof. We prove the statement in the proposition by using the geometry of
. Since
B has the approximation property, the finite-dimensional operators
in
are dense in
. So it suffices to prove the statement for a
with
Let
be the notation for the topological dual of
B. Recall that the elements of
are the image of the map
defined by:
As vectors in
B can be approximated arbitrarily close by vectors with a finite number of nonzero coordinates, we may assume that all the
have that property and then we have reduced the claim to proving the statement for elements of
that decompose for some
with respect to the splitting
as
The matrix of each element
we split similarly:
Then, the matrix of the operator
has the form
where
Let
be the matrix of the identity map in
. Note that the map from
defined by
is a continuous surjection. Hence, if we find a nonzero open subset of
such that for each vector
in that open set, it holds that
with
as an invertible upper triangular
-matrix and
a unipotent lower triangular matrix of the same size, then we have on an open subset of
and this is the decomposition we are looking for. Clearly, in order that we have the desired splitting of
, the condition
for the vector
is also necessary. We will show by induction on
N that there are
N nonzero polynomials
such that for all
in the complement of the union of the zero sets of all the
one has the decomposition
. For
, the matrix
is upper triangular and the desired decomposition holds for all
. Now, we take
; we split off the first row of
in
as follows:
where
is the matrix of the identity map in
We focus for the moment on the product of the last two matrices in (
45). Since the first column of
is nonzero, the polynomial
is nonzero. Now, we work on the complement of the zero set of
, so
is invertible. Define for all
the polynomials
. Note that
Then, the product of the last two matrices in (
45) is equal to
where each
and all
with
and
defined by
. Next, we push the top row of the right matrix to the right
The matrix
at the right has determinant
and will be part of
. Next, we move the matrix
in product (
45) to the right by using
where
is the column of length
N equal to
with
. The matrix
will be part of
. Thus, we reduce the case to the product
where the matrix at the right has determinant
. The induction hypothesis gives us then the sequence of nonzero polynomials
, so that on the complement of the union of all their zeros we have the desired decomposition. This proves the claim in the proposition. □
From Theorem 1 and Proposition 3, we may conclude
Corollary 2. The manifold describes solutions of the -hierarchy and the manifold describes solutions of the strict -hierarchy. As such, these manifolds are in the present context the analogues of the Grassmann manifold Gr(B) and its cover, the flag variety , used in [29,31] to construct solutions of the KP hierarchy and the strict KP hierarchy.