Periodic integrable systems with delta-potentials
E. Emsiz, E.M. Opdam and J.V. Stokman
KdV Institute for Mathematics, Universiteit van Amsterdam, Plantage Muidergracht 24, 1018 TV Amsterdam, The Netherlands. E-mail address : eemsiz@science.uva.nl, opdam@science.uva.nl, jstokman@science.uva.nl
-
Abstract.
In this paper we study root system generalizations of the quantum Bose-gas on the circle with pair-wise delta function interactions. The underlying symmetry structures are shown to be governed by the associated graded of Cherednik's (suitably filtered) degenerate double affine Hecke algebra, acting by Dunkl-type differential-reflection operators.
We use Gutkin's generalization of the equivalence between the impenetrable Bose-gas and the free Fermi-gas to derive the Bethe ansatz equations and the Bethe ansatz eigenfunctions.
1 Introduction
Given any affine root system
, Gutkin and Sutherland [10] , [29] defined a quantum integrable system whose Hamiltonian
has a potential
expressible as a weighted sum of delta functions at the affine root hyperplanes of
. For the affine root system of type
, the quantum integrable system essentially reduces to the quantum Bose-gas on the circle with pair-wise delta function interactions, which has been subject of intensive studies over the past 40 years.
The special case of the impenetrable Bose-gas on the circle was exactly solved by relating the model to the free Fermi-gas on the circle (see Girardeau [9] ). Soon afterwards fundamental progress was made for arbitrary pair-wise delta function interactions by Lieb & Liniger [21] , Yang [30] and Yang & Yang [31] , leading to the derivation of the associated Bethe ansatz equations and Bethe ansatz eigenfunctions. Yang & Yang [31] showed that the solutions of the Bethe ansatz equations are controlled by a strictly convex master function. One of the aims of the present paper is to generalize these results to Gutkin's and Sutherland's quantum integrable systems associated to affine root systems.
Quantum Calogero-Moser systems are root system generalizations of quantum Bose-gases on the line or circle with long range pair-wise interactions. In special cases quantum Calogero-Moser systems naturally arose from harmonic analysis on symmetric spaces. A decisive role in the studies of quantum Calogero-Moser systems has been played by certain non-bosonic analogs of systems, which are defined in terms of Dunkl-type commuting differential-reflection operators. Suitable degenerations of affine Hecke algebras naturally appear here as the fundamental objects governing the algebraic relations between the Dunkl-type operators and the natural Weyl group action.
In this paper we define Dunkl-type commuting differential-reflection operators associated to the root system generalizations of the quantum Bose-gas with delta-function interactions.
We furthermore show that the Dunkl-type operators, together with the natural affine Weyl group action, realize a faithful representation of the associated graded of Cherednik's [3] (suitably filtered) degenerate double affine Hecke algebra. There results show that these quantum integrable systems naturally fit into the class of quantum Calogero-Moser integrable systems, a point of view which also has been advertised from the perspective of harmonic analysis in [14,Sect.5] .
The quantum integrable systems under consideration for affine root systems
of classical type still have reasonable physical interpretations in terms of interaction one-dimensional quantum bosons. In these cases various results of the present paper can be found in the vast physics literature on this subject. We will give the precise connections to the literature in the main body of the text.
The knowledge on the quantum Bose-gas with pair-wise delta function interactions still far exceeds the knowledge on its root system generalizations. In fact, an important feature of the quantum Bose-gas with pair-wise delta function interactions is its realization as the restriction to a fixed particle sector of the quantum integrable field theory in
dimensions governed by the quantum nonlinear Schrödinger equation. This point of view has led to the study of this model by quantum inverse scattering methods. With these methods a proof of full orthogonality of the Bethe eigenfunctions on a period box(with respect to Lebesgue measure) is derived in [5] and the quadratic norms of the Bethe eigenfunctions are evaluated in terms of the determinant of the Hessian of the master function (conjectured by Gaudin [8,Sect.4.3.3] and proved by Korepin [19] ).
At this point we can only speculate on the generalizations of these results to arbitrary root systems. It seems that the quantum inverse scattering techniques are only in reach for classical root systems, in which case we have quantum field theories with (non)periodic integrable boundary conditions to our disposal, see [27] . It seems though reasonable to expect that the Bethe eigenfunctions are orthogonal on a fundamental domain for the reflection representation of the affine Weyl group (with respect to Lebesgue measure), and that their quadratic norms are expressible in terms of the determinant of the Hessian of the master function at the associated spectral point.
The contents of the paper is as follows. Sections 2 and 3 are meant to introduce the quantum integrable systems and to state and clarify the results on the associated spectral problem. We first introduce in Section 2 the relevant notations on affine root systems.
Following Gutkin [11] we formulate the spectral problem for the quantum integrable systems under consideration as an explicit boundary value problem. We state the main results on the boundary value problem (Bethe ansatz equations and Bethe ansatz eigenfunctions) and we introduce the associated master function. In Section 3 we formulate the analog of Girardeau's equivalence between the impenetrable Bose-gas and the free Fermi-gas on the circle for the quantum integrable systems under consideration.
In Section 4 we introduce Dunkl-type commuting differential-reflection operators and show that they realize, together with the natural affine Weyl group action, a faithful realization of the associated graded
of Cherednik's [3] (suitably filtered) degenerate double affine Hecke algebra. In Section 5 we show that Gutkin's [11] integral-reflection operators, together with the ordinary directional derivatives, yield an equivalent realization of
.
The equivalence is realized by Gutkin's [11] propagation operator. In Section 6 we show that the Dunkl operators naturally act on a space of functions with higher order normal derivative jumps over the affine root hyperplanes.
In Section 7 we return to the boundary value problem of Section 2 . Using the Hecke-type algebra
we refine an clarify Gutkin's [11] generalization of Girardeau's equivalence between the boundary value problem for the impenetrable Bose gas and the boundary value problem for the free Fermi-gas as formulated in Section 3 . The results in this section entail that the boundary value problem is equivalent to a boundary value problem with trivial boundary value conditions, on the cost of having to deal with a nontrivial (partially fermionic) affine Weyl group action. In Section 8 we study the reformulated boundary value problem, leading in Section 9 to the derivation of the Bethe ansatz equations. In Section 10 we study the master function and show how it leads to a natural parametrization of the solutions of the Bethe ansatz equations. In Sections 11 the solutions of the Bethe ansatz equations are further analyzed. In Section 12 it is proved that the boundary value problem has solutions if and only if the associated spectral value is a regular solution of the Bethe ansatz equations. In case of root system of type
, this is known as the Pauli principle for the interacting bosons!
2 The boundary value problem
In this section we recall Gutkin's [11] reformulation of the spectral problem for periodic integrable systems with delta-potentials in terms of a concrete boundary value problem.
We furthermore state the main results on the solutions of the boundary value problem and we detail the physical background.
In order to fix notations we start by recalling some well known facts on affine root systems, see e.g. [16] for a detailed exposition. Let
be an Euclidean space of dimension
. Let
be a finite, irreducible crystallographic root system in the dual Euclidean space
. We denote
for the inner product on
and
for the corresponding norm. The co-root of
is the unique vector
satisfying
We write
for the resulting co-root system in
. We fix a basis
for the root system
. Let
be the corresponding decomposition in positive and negative roots. We denote
for the half sum of positive roots and
for the highest root with respect to the basis
. The highest root
is a long root in
. We define the fundamental Weyl chamber in
by
|
(2.1)
|
Let
be the vector space of affine linear functionals on
. Then
as vector spaces, where the second component is identified with the constant functions on
. The gradient map
is the projection onto
along this decomposition.
The subset
is the affine root system associated to
. We extend the basis
of
to a basis
of the affine root system
. Observe that
maps
onto
.
For a root
,
defines the orthogonal reflection in the root hyperplane
. The affine Weyl group
associated to
is the sub-group of the affine linear isomorphisms of
generated by the orthogonal reflections
(
). The sub-group
generated by the orthogonal reflections
(
) is the Weyl group associated to
. We denote
for the longest Weyl group element in
. It is well known that
(respectively
) is a Coxeter group with Coxeter generators the simple reflections
for
(respectively
for
).
A second important presentation of
is given by
with
the co-root lattice of
, acting by translations on
. The gradient map
induces a surjective group homomorphism
by
for
.
Alternatively,
if
is the
-component of
in the semi-direct product decomposition 2.2 . The space
of affine linear functionals on
is a
-module by
(
). Observe that
is
-stable, and
for roots
. Furthermore,
for
and
. The length of
is defined by
.
Alternatively,
is the minimal positive integer
such that
can be written as product of
simple reflections. Such an expression
(
) is called reduced.
The weight lattice of
is defined by
Another convenient description is
|
(2.3)
|
which follows from the fact that
is already spanned over
by the short co-roots in
.
We denote
(respectively
) for the cone of dominant (respectively strictly dominant) weights with respect to the choice
of positive roots in
. Recall that
.
We write
for the irregular vectors in
with respect to the affine root hyperplane arrangement
. Its open, dense complement
is called the set of regular vectors in
.
We denote
for the collection of connected components of
. An element
is called an alcove. The affine Weyl group
acts simply transitively on
. Explicitly,
(disjoint union) with the fundamental alcove
defined by
We call a vector
(
) sub-regular if it does not lie on any other root hyperplane
(
).
The symmetric algebra
is canonically an
-module algebra. Using the standard identification
where
is the algebra of real-valued polynomial functions on
, the
-module structure takes the form
We denote
and
for the subalgebra of
-invariants in
and
, respectively.
Let
(
) be the derivative in direction
,
for
continuously differentiable at
. The assignment
uniquely extends to an algebra isomorphism of
onto the algebra of constant coefficient differential operators on
(say acting on
). We denote
for the constant coefficient differential operator corresponding to
. For example, the
-invariant constantcoefficient differential operator
associated to the polynomial
is the Laplacean
on
.
The boundary value problem, which we will define now in a moment, depends on certain coupling constants called multiplicity functions.
Definition 2.1.
A multiplicity function
is a
-invariant function
satisfying
for all
.
Unless stated explicitly otherwise, we fix a strictly positive multiplicity function
. To simplify notations we write
for the value of
at the root
.
Let
be the space of complex valued, continuous functions
on
which are smooth on the closures
for all alcoves
(in other words,
has a smooth extension to some open neighborhood
for all
), and which satisfy the derivative jump conditions
|
(2.4)
|
for sub-regular vectors
(
). The following boundary value problem was formulated by Gutkin and Sutherland, see [29] and [10] .
Definition 2.2.
Fix a spectral parameter
. We denote
for the space of functions
solving the system
|
(2.5)
|
of differential equations away from the root hyperplane configuration
.
The central theme of this paper is the study of the subspace
of
-invariant solutions, where
acts on
by the usual action
|
(2.6)
|
for
and
.
Example 2.3 (Free case
).
A function
is a weak eigenfunction of the Laplacean
on
with eigenvalue
, cf. the proof of [
11,Thm.2.7]
.
The regularity theorem for elliptic differential operators now implies that
is smooth.
Consequently
is the solution space to the spectral problem for the free bosonic quantum integrable system on
associated to the Laplacean
on
. It is easy to show that
is zero-dimensional unless
, in which case it is spanned by the plane wave
(cf. with the analysis in the impenetrable case
in Section 3 ).
The boundary value problem (see Definition 2.2 ) is a reformulation of the spectral problem associated to the quantum integrable system with Hamiltonian
|
(2.7)
|
where
is the Kronecker delta-function, see e.g. [29] , [10] , [11] as well as Proposition 6.4 and Remark 7.4 . Here we interpret
as a linear map
, with
the space of distributions on
, as
|
(2.8)
|
for a test function
, with
the Euclidean volume measure on
and
(
) the corresponding volume measure on the root hyperplane
. Our focus on
-invariant solutions amounts to studying the associated bosonic (=
-invariant) theory under
-periodicity restraints (or equivalently, we view the quantum system on the torus
).
The quantum Hamiltonian 2.7 for
of type
takes the explicit form
Here we have embedded
into
as the hyperplane defined by
. The study of
-invariant solutions to the boundary value problem then essentially amounts to analyzing the spectral problem for the system describing
quantum bosons on the circle with pair-wise repulsive delta function interactions. In this special case the quantum system has been extensively studied in the physics literature, see e.g. [9] , [21] , [30] , [31] , [8] and [17] . The upgrade to other classical root systems amounts to adding particular reflection terms to the physical model, see e.g. [27] , [2] , [8] , [15] , [18] and [23] .
We are now in a position to formulate the main results on the solution space of the boundary value problem. Define the regularized
-function by
|
(2.9)
|
cf. [8] , [14] . We call the spectral value
regular if its isotropy sub-group in
is trivial (equivalently,
for all
). We call
singular otherwise.
Furthermore,
is called real (respectively purely imaginary) if
(respectively
).
Theorem 2.4.
Let
. The space
of
-invariant solutions to the boundary value problem is one-dimensional or zero-dimensional. It is one-dimensional if and only if the spectral value
is a purely imaginary, regular solution of the Bethe ansatz equations
|
(2.10)
|
If
is one-dimensional, then there exists a unique
normalized by
. The solution
is the unique
-invariant function satisfying
|
(2.11)
|
We give a reformulation of Theorem 2.4 in Section 3 . The Bethe ansatz equations are derived Section 9 . The regularity restraint on
is proved in Section 12 .
A key role in the analysis of the Bethe ansatz equations 2.10 is played by the following master function.
Definition 2.6.
The master function
is defined by
|
(2.14)
|
The master function
enters into the description of the set
of solutions
of the Bethe ansatz equations 2.10 in the following way.
Proposition 2.7.
For
there exists a unique extremum
of the master function
. The assignment
defines a
-equivariant bijection
.
The proof of Proposition 2.7 , which hinges on the strict convexity of
(
), is given in Section 10 . The regularity condition on the spectrum in Theorem 2.4 also turns out to be a consequence of the strict convexity of the master function
(
), see Section 12 .
The following proposition yields precise information on the location of the deformed weight
.
Proposition 2.8.
For
and
we have
|
(2.15)
|
where
. Furthermore,
if and only if
.
Proposition 2.8 is proved in Section 11 . The lower bound in 2.15 shows how far away the spectral values
(
) are from being singular.
The Bethe ansatz functions
and the necessity of the Bethe ansatz equations 2.10 on the allowed spectrum were obtained by Lieb and Liniger [21] for root system
of type
, and soon after generalized to root system
of type
by Gaudin [7] , [8] (see also [18] ). For
of type
, Yang and Yang [31] introduced the master function
(also known as the Yang-Yang action) and derived the special case of Proposition 2.7 using its strict convexity. In physics literature the regularity of the spectral parameter
(see Theorem 2.4 ) is usually imposed as additional requirement, since it automatically ensures that eigenstates admit a plane wave expansion within any alcove
. The regularity condition for root system
of type
can be viewed as a Pauli type principle for the interacting quantum bosons, since it implies that the momenta of the quantum bosons are pair-wise different.
An actual proof of the regularity of the spectrum was obtained by Izergin and Korepin [17] using quantum inverse scattering methods. In this derivation the regularity condition again follows from the strict convexity of the Yang-Yang action. Estimates for the momenta gaps of the quantum particles play a role in the study of the thermodynamical limit, see [21] and [31] . See e.g. [8,Sect4.3.2] for the exact analog of the estimates 2.15 for
of type
.
It is believed [17] that quantum integrable systems governed by a strictly convex master function always have a regularity restraint on the spectrum, although a conceptual understanding is not known as far as we know. We remark though that our derivation of the regularity restraint on the spectrum is in accordance to this point of view. In the following section we give though a conceptual understanding of the fermionic nature of the quantum integrable system at hand by generalizing Girardeau's [11] equivalence between the impenetrable quantum Bose-gas and the free quantum Fermi-gas.
3 Girardeau's isomorphism
To obtain a better (physical) understanding of the results described in the previous section it is instructive to reflect on the extremal case
(impenetrable case).
Let
be the space of real analytic functions on
, which we consider as an
-module with respect to the usual action 2.6 . Consider for
the space
|
(3.1)
|
which is an
-submodule of
(see [28] or Section 8 ). We observed already in Example 2.3 that
|
(3.2)
|
Denote
for the subspace of
-translation invariant functions in
.
Lemma 3.1.
For
, we have
unless
. For
the space
is spanned by
(
).
-
Proof.
By [28] , a function
can be uniquely written in the form
with
, see also Section 8 . Such a function
is
-translation invariant iff
|
(3.3)
|
for all
,
and
. This implies that
and that
is bounded on
for all
. The latter condition implies that
is constant for all
. Returning to 3.3 with
, the
-translation invariance of
is equivalent to
if
. Hence
unless
, in which case
is spanned by
(
). □
We denote
for the space of functions
satisfying
for all
and
. Since translations
have even length,
consists of
-translation invariant functions. In particular,
is the solution space to the spectral problem for free fermionic quantum integrable system on
associated to the Laplacean
on
.
Corollary 3.2.
Let
. The space
is zero-dimensional or one-dimensional. It is one-dimensional iff
is a regular element from
, in which case
is spanned by
|
(3.4)
|
-
Proof.
Let
and
with
, cf. Lemma 3.1 . Then we have
iff
for all
. For singular
this implies
for all
. For regular
we conclude that
is a constant multiple of
. □
Following the analogy with Girardeau's [9] analysis of the impenetrable quantum bosons on the circle, we define now a linear map
by
|
(3.5)
|
The map
is injective: for
in the image of
, the function
is the unique real analytic continuation of
to
.
For
we interpret the boundary conditions 2.4 as
for all
. The solution spaces
of the associated boundary value problem (see Definition 2.2 ) can now be analyzed as follows.
Proposition 3.3.
For
we have (i) The map
restricts to a linear isomorphism
.
(ii) The space
is zero-dimensional or one-dimensional. It is one-dimensional iff
is a regular element from
. In that case
is spanned by
, which is the unique
-invariant function satisfying
-
Proof.
(i) A function
vanishes on the root hyperplanes
(
), hence so does
. The function
furthermore satisfies the differential equations 2.5 , hence
.
For
we define
by
for
and
. This is well defined since
vanishes on the root hyperplanes
(
). Since
is
-alternating we have
. The function
satisfies the differential equations 2.5 , hence
, where the last equality follows from 3.2 . The proof is now completed by observing that
is the inverse of the map
.
(ii) This follows from (i) and Corollary 3.2 . □
For root system
of type
, Proposition 3.3 is due to Girardeau [9] .
The crucial step for establishing Theorem 2.4 is to generalize Proposition 3.3 (i) to arbitrary multiplicity function
. The main ingredients for this generalization have already been developed by Gutkin [11] . Gutkin considers a
-deformed
-action
on e.g.
, which is the usual (bosonic)
-action 2.6 with added fermionic component
. The multiplicity function
determines the relative strength of the fermionic component. Concretely, we consider the integral-reflection operators
|
(3.6)
|
with the integral operators
for
defined by
|
(3.7)
|
Then the assignment
(
) uniquely defines a left
-action on
, cf. [11] , [13] or Section 5 . In the terminology of [13] , the system of integral operators
is an operator calculus with respect to the affine Weyl group
.
It follows that the operators
and
(
) satisfy the braid relations of
with respect to the fixed basis
of
, as well as the quadratic relations
and
. In particular, given a reduced expression
for
, the operators
and
are well defined. The fermionic nature of the integral operators
is explained by the following lemma.
Lemma 3.4.
For
and
we have
if and only if
.
In particular,
for all
.
-
Proof.
It is immediate that
if
. Conversely, suppose
. It follows from [11,Lem.2.1(iii)] that
|
(3.8)
|
Applied to
we obtain
. □
The analog of Proposition 3.3 (i) is the statement that the map
restricts to a linear isomorphism
|
(3.9)
|
for arbitrary positive multiplicity function
, where
is the subspace of
-invariant functions in
. The proof of 3.9 is given in Section 7 .
With the isomorphism 3.9 at hand, Theorem 2.4 is equivalent to the following theorem.
Theorem 3.5.
Let
. The space
is one-dimensional or zero-dimensional. It is one-dimensional if and only if
is a purely imaginary, regular solution of the Bethe ansatz equations 2.10 . If
is one-dimensional then
|
(3.10)
|
is the unique function in
normalized by
.
The ”regular part” of Theorem 3.5 is proved in Section 9 . The exclusion of singular spectral parameters
(Pauli principle) is proved in Section 12 .
In order to reveal the full symmetry structures underlying the isomorphism 3.9 , we will consider the upgrade of the map
to an
-dependent linear isomorphism
of
which intertwines the
-action with the usual
-action 2.6 , and which acts as
when applied to
-invariant functions. The map which does the job is Gutkin's [11] propagation operator, defined by
for
and
(see Section 5 for details). The propagation operator
now restricts to an isomorphism
|
(3.11)
|
for all
(cf. [11] and Theorem 7.3 ), which implies 3.9 by restricting to the subspaces of
-invariant functions. We will show in Sections 4–7 that the
-action
on
together with the directional derivarives
(
) makes
into a module over the associated graded
of Cherednik's [3] (suitably filtered) degenerate double affine Hecke algebra. On the other hand,
is an
-module by adding to the usual
-action 2.6 certain explicit Dunkl-type differential-reflection operators. With these upgraded symmetry structures, 3.11 will be shown to be an isomorphism of
-modules (see Theorem 7.3 (iii)). In sections 4–7 the Hecke algebra structure is explored to (re-)prove and clarify crucial results on the boundary value problem (see Definition 2.2 ).
We conclude this section by considering the limit
to the impenetrable case. This limit is interpreted as
where
(
) and
is an arbitrary positive multiplicity function (the limits turn out to be independent of the choice of
).
In the limit
the Bethe ansatz equations 2.10 reduce to
which has
as purely imaginary solutions
(see 2.3 ). Furthermore we have
|
(3.12)
|
for
, which follows by taking the limit
in 2.15 . For
(
) a regular solution to the Bethe ansatz equation,
(see 3.10 ) can alternatively be written as
see [14] or Section 8 . It follows that
for
, uniformly on compacta. Pulling the limits through the map
, we obtain
for
, uniformly on compacta.
4 Dunkl operators and Hecke algebras
The conserved integrals for quantum integrable systems of Calogero-Moser type can be conveniently expressed in terms of Dunkl-type operators, which are explicit commuting first-order differential-reflection operators, see e.g. [6] and [1] . The Dunkl operators, together with the usual Weyl group action 2.6 , form a faithful representation of suitable degenerations of affine Hecke algebras, cf. [2] . We derive the Dunkl operators and the underlying Hecke algebra structures for the periodic quantum integrable systems with delta potentials as introduced in Section 2 . We define the Dunkl-type operators as differential-reflection operators on the space
of smooth functions on
, which is a module over the affine Weyl group
with respect to the usual action 2.6 . The connection to the boundary value problem (see Definition 2.2 ) will be clarified in Section 7 .
We denote
for the characteristic function of the interval
, so
if
and
if
. For
the function
(
) defines a smooth function on
, which is constant on the alcoves
of
.
In fact, for
and
we have
|
(4.1)
|
hence
is nonzero on a given alcove
(
) for only finitely many positive roots
. The Dunkl-type operators
|
(4.2)
|
thus are well defined linear operators on
, depending linearly on
. In fact, for
and
we have by 4.1
|
(4.3)
|
In particular, for the fundamental alcove
we simply have
|
(4.4)
|
The Dunkl operators
(
) and the
-action 2.6 on
satisfy the following fundamental commutation relations.
Theorem 4.1.
(i) We have the cross relation
(ii) The Dunkl operators
(
) pair-wise commute.
-
Proof.
(i) Fix
and
. By a direct computation we have
Since
we obtain
| |
which is the desired cross relation.
(ii) We derive the commutativity of the Dunkl operators
(
) as a direct consequence of 4.4 and the cross relation. Let
and
. We show by induction on the length
of
that
|
(4.5)
|
By 4.4 , equation 4.5 is obviously valid for
the unit element of
. To prove the induction step, it suffices to show that
|
(4.6)
|
for all
. For the proof of 4.6 , first observe that
|
(4.7)
|
for all
, which follows from applying the cross relation twice. Now 4.6 follows from the fact that the right hand side of 4.7 is symmetric in
and
. □
By Theorem 4.1 (ii), the assignment
uniquely extends to an algebra morphism
. We denote
for the differential-reflection operator on
associated to
.
Let
be the subspace of functions
for which
has a smooth extension to some open neighborhood
for all alcoves
. Clearly
is a left
-module by 2.6 . Observe that a function
has a natural multi-valued extension to
defined by
|
(4.8)
|
where
is the finite collection of alcoves
such that
. It follows from 4.3 that the Dunkl operators
(
) restrict to well-defined linear operators on
.
We formulate the second part of the following theorem in terms of the space
. It obviously also holds true for the larger space
.
Theorem 4.2.
(i) There exists a unique complex unital associative algebra
satisfying
-
(a)
as vector space, with
the group algebra of
.
-
(b)
The maps
and
, with
the unit element of
, are algebra embeddings of
and
into
.
-
(c)
The cross relations
holds in
for
and
. Here we have identified
and
with their images in
through the algebra embeddings of (b).
(ii) The assignment
, together with the
-action 2.6 , defines a faithful representation
.
-
Proof.
Suppose that
as endomorphism of
with only finitely many
's non zero. We show that all
's are zero. Equation 4.4 implies
|
(4.9)
|
Applying 4.9 to functions
of the form
with
and with
having support in the fundamental alcove
, we conclude that
as constant coefficient differential operator on smooth functions in some open ball
, hence
.
The proof of the theorem is now standard: let
be the complex unital associative algebra generated by
and
(
) with defining relations as in (b) and (c) (so the vectors
pair-wise commute, the
(
) are involutions satisfying the Coxeter relations associated to
and
, and the generators satisfy the cross relations from (c)). By Theorem 4.1 and by the paragraph preceding this theorem, the assignment
, together with the
-action 2.6 , uniquely defines an algebra morphism
.
By the previous paragraph and by the cross relations in
it follows that
is injective and that
as vector spaces (the Poincaré-Birkhoff-Witt Theorem for
). Both statements of the theorem are now immediately clear. □
We will use the notation
to indicate that the subspace
is a
-submodule or
-submodule of
with respect to the
-action.
Remark 4.3.
If the values
of the multiplicity function
are considered to be independent central variables in the definition of
, then
is graded by imposing the degree of
to be zero and the degrees of
and
to be one. As graded algebra,
is the associated graded of Cherednik's [
3]
degenerate double affine Hecke algebra
, considered as filtered algebra by the same degree function (the only difference in the definition of
is the cross relation (see Theorem 4.2 (c)), which now is of the form
for
, where
is considered as
-module algebra with the action of
defined by
).
Lemma 4.4.
The center
of
contains
.
-
Proof.
Observe that the cross relations in
(see Theorem 4.2 (c)) imply
|
(4.10)
|
for
and
. It follows from 4.10 that
. □
For trivial multiplicity parameters
, the operator
(
) on
is the constant-coefficient differential operator
on
. We have the following striking fact when
is
-invariant.
Corollary 4.6.
For
we have
as operators on
.
-
Proof.
Let
and
. By 4.4 we have
. Let
and
. By Lemma 4.4 applied twice (once with multiplicity function
, once with
), we have
| |
hence
. □
Remark 4.7.
The Dunkl operators
, Theorem 4.1 , Theorem 4.2 and Corollary 4.6 have their obvious analogs in the context of finite root systems. In that case, the Dunkl-type operators are
realizing, together with the
-action 2.6 , an action of the degenerate affine Hecke algebra
on the space of smooth functions on
(or, on the associated
-space). For classical root systems these operators were constructed using solutions of classical Yang-Baxter equations and reflection equations in [
26]
, [
23]
(type A) and [
18]
.
This construction fits into Cherednik's [
2]
general framework relating root system analogs of
-matrices to (degenerate) affine Hecke algebras and Dunkl operators.
5 Integral-reflection operators
Heckman and Opdam [14] clarified the role of the degenerate affine Hecke algebra
in Gutkin's [11] work when the underlying root system is finite. It led to an explicit action of
as directional derivatives and integral-reflection operators. In this section we extend these results to the present affine set-up, which yields a second
-action on
. We show that Gutkin's [11] notion of propagation operator leads to an explicit isomorphism between this
-module structure on
and
(see Theorem 4.2 (ii)).
The integral-reflection operators
(see 3.6 ) for
is a well-defined endomorphism of
. They satisfy
|
(5.1)
|
with respect to the
-action 2.6 on
. Recalling the interpretation of
as multi-valued function on
(see 4.8 ), we furthermore have
|
(5.2)
|
Observe that the directional derivatives
(
, viewed as operators on
, restrict to endomorphisms of
. The following result is the affine analog of [14,Thm.2.1] and [14,Cor.2.3] .
Theorem 5.1.
(i) The assignment
|
(5.3)
|
defines a representation of
on
.
(ii) The assignment
, together with the
-action on
from (i), uniquely defines a representation
.
-
Proof.
(i) This follows from [11,Thm.2.3] , see also [13] .
(ii) It suffices to verify the cross relations (see Theorem 4.2 (c)), which follow directly from [11,Lem.2.1] . □
By Theorem 5.1 (ii) we have for
and any choice of decomposition
as product of simple reflections (
),
|
(5.4)
|
We will use the notation
to indicate that the subspace
is a
-submodule or
-submodule of
with respect to the
-action.
Definition 5.2.
Gutkin's [
11]
propagation operator
is the endomorphism of
defined by
|
(5.5)
|
for
. In particular,
is the identity operator on
.
Theorem 5.3.
The propagation operator
is an isomorphism
of left
-modules. In particular, for
we have
as endomorphisms of
, where
is the restriction to
of the directional derivative
acting on
.
-
Proof.
It follows from [12,Thm.2.7(1)] that
is a linear isomorphism of
. The
-equivariance of
is immediately clear from equation 5.5 defining
. Hence it remains to show that
|
(5.6)
|
for
and
. To prove 5.6 we use the commutation relation
|
(5.7)
|
in
, which can be easily proved by induction on the length
of
using the cross relations in
(see Theorem 4.2 (c)). Fix
and
. By 5.7 and Theorem 5.1 (ii) we have
| |
where the last equality follows from 4.3 . □
Remark 5.4.
Theorem 5.3 has an obvious analog in the context of finite root systems (compare with Remark 4.7 ). In the case of finite root system of type A, the intertwining properties of the propagation operator with respect to the degenerate affine Hecke algebra actions were considered in [
15]
.
6 Derivative jump conditions
For
we denote
for the space of functions
satisfying
|
(6.1)
|
for
,
sub-regular and
(where the right hand side of 6.1 should be read as zero for
). We define
as the space of functions
satisfying the
-dependent normal derivative jump conditions 6.1 over the root hyperplanes of any order. Note that the space
was used to formulate the boundary value problem in Definition 2.2 .
A function
automatically also satisfies the jump conditions 6.1 for negative roots
,
sub-regular and
. Hence the space
does not dependent on the choice of positive roots
in
, and
is an
-submodule of
.
Lemma 6.1.
For
we have
. In particular we have
.
-
Proof.
For a function
and any alcove
we write
for the smooth extension of
to some open neighborhood
of
.
We first show that
. Fix
. It suffices to show that its multi-valued extension 4.8 to
is single-valued. Equation 6.1 for
implies that
is single-valued on
, where
is the set of sub-regular vectors in
.
Fix a vector
. For arbitrary alcoves
there exists a gallery
of alcoves in
such that
is adjacent to
for all
.
This follows easily from e.g. [13,Sect.2] and the fact that the finite isotropy sub-group
of
in
acts simply transitively on
. For adjacent alcoves
we write
for the set of sub-regular vectors contained in the wall
. Then
hence
is single-valued at
.
As a second step we show that
. The statement for
(
arbitrary) then follows by an easy induction argument.
Fix
. We first show that
for
. Let
and
be adjacent alcoves, with common wall (say)
(
). If
is parallel to the affine hyperplane
(i.e.
), then
exists for all sub-regular
and is equal to both
and
. In particular,
is continuous at sub-regular
. For
normal to
, say pointing towards
, we have
for all sub-regular
, where
By the zero normal derivative jumps of
we conclude that
exists and is continuous at sub-regular
. Hence
for
, and the continuous extension
of
to
is given by
where
is any alcove from
.
It remains to show that
equals
on
for all
. Let
. There exist alcoves
and an
such that
for
. Then
Hence
exists and
. We conclude that
, as desired. □
Observe that
is an
-submodule of
for
, while
is an
-submodule of
.
Theorem 6.2.
(i) For
the propagation operator
restricts to an isomorphism
of left
-modules.
(ii)
is an
-submodule of
.
(iii) The propagation operator
restricts to an isomorphism
of left
-modules.
-
Proof.
In view of Theorem 5.3 it suffices to show that
|
(6.2)
|
We generalize the proofs of [11,Thm.2.6&Thm.2.7] , which deal with the proof of 6.2 for
and
.
We use the commutation relations
|
(6.3)
|
in
, which follows from 4.10 applied to
(for
, the right hand side of 6.3 should be read as zero). By Theorem 5.1 (ii) and 5.1 , the commutation relations 6.3 in
imply
|
(6.4)
|
as endomorphisms of
. Note the reminiscence of 6.3 and 6.4 to the jump condition 6.1 .
Let
and write
. We have to show that
satisfies the derivative jumps 6.1 over sub-regular
(
) for
. In view of the
-equivariance of the propagation operator
(see Theorem 5.3 ), it suffices to derive the derivative jumps for
over sub-regular vectors
(
). Fix
,
sub-regular and
. For
small we have
for
. Hence
|
(6.5)
|
where the second equality follows from 5.2 . On the other hand,
|
(6.6)
|
Combining 6.5 and 6.6 now yields
| |
where the second equality follows from 6.4 . Thus
.
Conversely, suppose that
and write
. Suppose that
for some
(for
, this is interpreted as
). We show that
. In this case the argument is more subtle since we cannot reduce the analysis to the walls of the fundamental alcove
. Let
and
sub-regular. By Lemma 6.1 it suffices to show that
satisfies the derivative jump conditions 6.1 over
with
. Let
and
such that
and
(so
points towards
), and define operators
by
then we can write
|
(6.7)
|
where we use 5.1 and Theorem 5.1 (i) for the second equality. From the definition 3.7 of the integral operator
on
it easily follows that
for all
(cf. [11,Lem.2.1(ii)] ). Combined with 3.6 we obtain
|
(6.8)
|
It follows from 6.8 and 5.1 that
|
(6.9)
|
and that both operators
and
have polynomial expressions in the integral operators
without constant terms. In particular,
and
map
into
.
Consequently
| |
with the obvious interpretation for
, where the second equality follows from 6.9 and the third equality follows from 3.8 . Using 6.7 and the fact that
it now easily follows that
This completes the proof of 6.2 . □
Remark 6.3.
Theorem 6.2 has an obvious analog in the context of finite root systems (compare with Remark 4.7 ). In the case of finite root system of type A, the space
was considered in [
12]
.
In the following proposition we relate the Dunkl operators
to the quantum Hamiltonian
with delta function interactions (see 2.7 and 2.8 ). For a continuous function
we write
for the corresponding distribution:
for a test function
. For a distribution
and an open subset
, we write
for the restriction of
to
. Recall that
for the
-invariant polynomial
on
.
Proposition 6.4.
For
we have
as distributions on
.
-
Proof.
Note that
, so both sides are well defined distributions on
. By Corollary 4.6 we have
. The proposition now follows from a standard argument using Green's identity, cf. e.g. the proof of [11,Thm.2.7] . □
By Proposition 6.4 it is justified to interpret the restriction of the quantum Hamiltonian
to
as the endomorphism
on
. The complete integrability of the corresponding quantum system is then directly reflected by the commutativity of the Dunkl operators
(
) as endomorphisms of
. In particular, for any
the operator
on
commutes with
and
for
, which is in accordance with the distributional analysis in [11,Sect.3] .
7 Reformulation of the boundary value problem
For a left
-module
and
we define
|
(7.1)
|
which is an
-submodule of
in view of Lemma 4.4 . By Remark 4.5 , the module
consists of the vectors
transforming according to the central character
for the action of the center
of the degenerate affine Hecke algebra
. Set
Proposition 7.1.
For
we have
|
(7.2)
|
-
Proof.
Let
and
. The multi-valued function
is uniquely determined by its restriction to
, in which case it equals
. This implies
. The equality
follows now directly from Corollary 4.6 . □
Recall from the previous section that
is an
-submodule of
.
Corollary 7.2.
For
, the space
(see 3.1 ) is the
-submodule
of
.
-
Proof.
By the definition 3.1 of
we have
. By Proposition 7.1 it now follows that
. □
Observe that the solution space
of the boundary value problem (see Definition 2.2 ) can be written as
|
(7.3)
|
In the following theorem we relate
to the
-module
(see Theorem 6.2 (ii)), and we use the propagation operator
to relate
to
.
Theorem 7.3.
For
we have (i) The solution space
coincides with the
-submodule
of
.
(ii) The propagation operator
restricts to an isomorphism
of left
-modules.
(iii) The map
3.5 restricts to an isomorphism
.
-
Proof.
By Theorem 5.3 and Proposition 7.1 , the propagation operator
restricts to a linear isomorphism
|
(7.4)
|
Combined with 7.3 , Theorem 6.2 (i) and Lemma 6.1 it follows that
restricts to a linear isomorphism
|
(7.5)
|
cf. [11,Sect.3] . By Corollary 7.2 , 7.4 and Theorem 6.2 (iii) it furthermore follows that
restricts to an isomorphism
|
(7.6)
|
of left
-modules. Part (i) and part (ii) of the theorem now follows from combining 7.5 and 7.6 with 3.2 . Part (iii) of the theorem follows from (ii) and the fact that the propagation map
acts on
-invariant functions in the same way as the map
3.5 . □
Remark 7.4.
It follows from Proposition 6.4 and Theorem 7.3 (i) that functions
are weak eigenfunctions of the quantum Hamiltonian
(see 2.7 and 2.8 ) with eigenvalue
.
As observed in Section 3 , Theorem 7.3 (iii) can be used to reformulate the main results on the solution space
(see Theorem 2.4 ) to the boundary value problem in terms of the space of invariants
, where
now is the solution space to the boundary value problem with zero normal derivative jumps over sub-regular vectors. Theorem 3.5 is the resulting reformulation of Theorem 2.4 . In order to prove Theorem 3.5 we analyze the space
in detail in the following sections.
8 Invariants in
In this section we analyze the sub-space
of
-invariants of
. First we recall some well known properties of the space
from [28] and [14] . For technical purposes it is convenient to introduce the following terminology.
Definition 8.1.
Let
be a subset of the simple roots
. The spectral parameter
is called
-standard if
and if the isotropic sub-group of
in
is the standard parabolic sub-group
generated by the simple reflections
(
).
Lemma 8.2.
Let
. The
-orbit of
contains a
-standard spectral parameter for some subset
.
-
Proof.
Taking a
-translate of
we may assume that
with
and
. The isotropy group of
in
is a standard parabolic sub-group
for some subset
. Write
with
and
its orthocomplement in
. Set
which we view as the fundamental chamber for the action of the standard parabolic sub-group
on
. Taking a
-translate of
we may assume that
with
,
, and
as before. The isotropy sub-group of
in
then equals the isotropy sub-group of
in
, which is a standard parabolic sub-group
for some subset
since
. □
Observe that a
-standard spectral parameter
is regular if and only if
. Note furthermore that the module
(
) only depends on the orbit
. When analyzing the module
, we thus may assume without loss of generality that
is
-standard for some subset
. In particular, we will now assume this condition for the remainder of this section.
For
we denote
(respectively
) for the homogeneous polynomials
of degree
(respectively the polynomials
of degree
). The
-action 2.6 on
respects the natural grading
. Furthermore,
is a graded
-submodule of
, isomorphic to the regular representation of
(see e.g. [28,Thm.1.2] and references therein). We write
and
.
Denote by
the minimal coset representatives of
. Steinberg [28] established the decomposition
|
(8.1)
|
It follows from 8.1 that
, viewed as
-module by the action 2.6 , is isomorphic to the regular representation of
. Furthermore, we have
with
the
-submodule
We denote
.
Representations of the finite group
do not admit nontrivial continuous deformations, hence
is isomorphic to the regular representation of
for arbitrary multiplicity function
. In particular,
is one-dimensional for all spectral values
. In fact, by 5.2 the function
|
(8.2)
|
satisfies
and spans
. On the other hand, by 8.1 there exist unique polynomials
(
) such that
|
(8.3)
|
By 8.1 we have
|
(8.4)
|
so the polynomials
(
) are constants for regular
. In fact, from e.g. [8] and [14,Sect.2] we have
|
(8.5)
|
where the
-function
is given by 2.9 , hence
For singular
an explicit expression for
(
) is not known. For our purposes it suffices to have explicit expressions for the highest and the next to highest homogeneous components of
, which we will now proceed to derive.
We denote
for the parabolic root sub-system associated to the simple roots
. We write
for the cardinality of the corresponding set
of positive roots in
and
Recall that the minimal coset representatives
of
can be characterized by
The following lemma now gives a derivational expression for
(
).
-
Proof.
By 8.2 ,
(
) depends analytically on the spectral parameter
. In particular,
with
depends analytically on
, and we have the (point-wise) limit
|
(8.6)
|
For
we write
There exists an
such that
is regular for
, hence
by 8.5 . Splitting the sum into a double sum
with
and
and using
| |
we obtain
|
(8.7)
|
as analytic functions in
(note that
is analytic at
). By 8.6 ,
is the
th term in the power series expansion of 8.7 at
, which yields the desired result. □
Define the strictly positive constant
by
The highest and next to highest homogeneous terms of
(
) can now be explicitly computed as follows.
Proposition 8.4.
Let
be
-standard and
.
(i) The highest homogeneous term
of
is of degree
and is explicitly given by
(ii) Suppose that
is singular (i.e.
). The next to highest homogeneous term
of
is
with
|
(8.8)
|
Remark 8.5.
The formula for
should be read as identity between real analytic functions in
(the possible singularities are easily seen to be removable).
-
Proof.
(i) Observe that
is independent of
, and
Combined with Lemma 8.3 we conclude that the highest homogeneous term
of
is given by
|
(8.9)
|
On the other hand, by the Weyl denominator formula for
we have
Combined with the first equality in 8.9 we obtain the desired expression for
.
(ii) The next to highest homogeneous term
of
is
in view of Lemma 8.3 , where the prime denotes the
-derivative. The first
-sum in this expression is identically zero since it is an
-alternating polynomial of degree
. By a direct calculation the remaining expression can be rewritten as
The desired expression for
now follows from 8.9 . □
9 The Bethe ansatz equations
In this section we show that
implies that the spectral parameter
is a purely imaginary solution of the Bethe ansatz equations 2.10 .
From the results of the previous section it is clear that
is one-dimensional or zero-dimensional. In fact it is one-dimensional if and only if
, in which case we have
It is convenient to reformulate these observations in terms of
|
(9.1)
|
(viewed as operator on e.g.
or
), which satisfies the elementary commutation relations
(the operator
can be defined on the level of the algebra
as the element
, in which case it is the analog of the affine intertwiner from [4] and [25,Sect.4] ). The equality
clearly implies
.
Lemma 9.1.
If
is regular, then
implies
.
-
Proof.
By 8.4 we have a unique expansion
with
. We conclude from the equality
that
for all
. Since
is regular, this implies
for all
. □
For
we write
for the associated constant coefficient differential operator acting on smooth functions in
.
Lemma 9.2.
Let
. For
we have
| |
where we view the left hand sides as functions in
and the right hand sides as functions in
. In particular,
for
and
.
-
Proof.
Observe that
and
(acting on
) clearly commutes with the endomorphisms
and
(which act on
). Thus it suffices to prove the lemma for
, in which case the second formula is trivial. To prove the first formula for
we may assume without loss of generality that
is the unit element of
. Suppose that
is regular. Adirect computation using the definition 3.6 of
as integral-reflection operator yields
hence
In the latter formula the regularity restraint on
can be removed by continuity. □
We denote
for the projection onto
along the decomposition
. Observe that
|
(9.2)
|
if
is
-standard in view of Proposition 8.4 (i). In this section we consider the restraint on
such that
|
(9.3)
|
for the highest degree component
.
The map
, where
is obtained from the unique decomposition
|
(9.4)
|
defines an involution on
. Observe that
|
(9.5)
|
Lemma 9.3.
Suppose that
is
-standard.
(i) The equation 9.3 for
holds if and only if
satisfies the equations
|
(9.6)
|
(ii) For
and for multiplicity functions
such that
, we have
-
Proof.
(i) By 8.3 , Lemma 9.2 and Proposition 8.4 (i) we have
|
(9.7)
|
The proof now follows by equating the coefficients of
(
) in 9.7 using 9.4 .(ii) We first compare the denominators of
and
. If
is regular then
| |
Taking the limit
we obtain
A similar (and easier) computation leads to the comparative formula
for the numerators of
and
. Combining both formulas leads to the desired result. □
Recall from Section 2 that
is the set of purely imaginary solutions of the Bethe ansatz equations 2.10 .
Proposition 9.4.
Suppose that
is
-standard. The equation 9.3 for
holds if and only if
.
-
Proof.
We first show that
is purely imaginary if
satisfies the equation 9.6 . Let
(
) be the element in the
-orbit of
having its real part in
. Then
since the multiplicity function
is strictly positive, hence 9.6 and Lemma 9.3 (ii) imply
|
(9.8)
|
The modulus of the left hand (respectively right hand side) of 9.8 is
(respectively
) since the real part of
is in
and the multiplicity function
is strictly positive.
Thus
, implying that
is purely imaginary. Since
with
strictly positive integers and since the real part of
lies in
, we conclude that
is purely imaginary for all co-roots
(
). This implies
, hence
.
Combined with Lemma 9.3 (i) it follows that
satisfies 9.3 for
if and only if
is a purely imaginary solution of the equations 9.6 . For purely imaginary
we have
for all
due to the strict positivity of the multiplicity function
. The proof now follows from Lemma 9.3 (ii) and Remark 2.5 . □
As an immediate result we obtain the following “regular part” of Theorem 3.5 .
Corollary 9.5.
Suppose that
is regular. The space
is zero-dimensional or one-dimensional. It is one-dimensional if and only if
. In that case
is spanned by
3.10 .
-
Proof.
By the observations at the beginning of the section it suffices to show that
iff
.
Since
is a
-invariant subset and
only depends on the
-orbit of
, we may assume without loss of generality that
is
-standard. If
then 9.3 holds, hence
by Proposition 9.4 . Conversely, suppose that
. Since
is regular we have
by 9.2 , hence
by Proposition 9.4 .
By Lemma 9.1 this implies
, hence
. □
10 The master function
In this section we prove Proposition 2.7 , which yields a parametrization of the set
of purely imaginary solutions of the Bethe ansatz equations 2.10 by the weight lattice
.
We first rewrite the Bethe ansatz equations 2.10 in logarithmic form. By a direct computation using the elementary identity
the Bethe ansatz equations 2.10 for
can be rewritten as
|
(10.1)
|
for all
. On the other hand, for
the gradient of the master function
(see 2.14 ) is determined by
|
(10.2)
|
Comparing 10.1 and 10.2 yields the following result.
Lemma 10.1.
We have
if and only if
with
an extremal vector of the master function
for some
.
-
Proof.
is an irreducible root system in
, hence
spans
. Thus
is an extremal vector of
if and only if
for all
, which by 10.2 is equivalent to
for all
. Comparing to 10.1 , the proof now follows from 2.3 . □
We thus need to analyze the extrema of the master function
at a given weight
. Observe that the Hessian
of
at
is independent of
, and is given explicitly by
|
(10.3)
|
By the strict positivity of the multiplicity function
, it follows from 10.3 that the Hessian
is positive definite for all
, hence
is strictly convex. Furthermore, for all
,
hence
has a unique extremum
, which is a global minimum. It now follows from 10.2 that
(
) is uniquely determined by the equation
|
(10.4)
|
in
, where
(
) is defined by
Combined with Lemma 10.1 it now follows that the map
is a bijection from the weight lattice
onto
. The
-equivariance of this map is immediate from the equivariance property
for
and
. This completes the proof of Proposition 2.7 .
11 Moment gaps
In this section we prove Proposition 2.8 , which yields estimates for the location of the deformed weight
compared to the parametrizing weight
. In view of 10.2 and Lemma 10.1 , the deformed weight
(
) is the unique solution of The following lemma establishes the necessary bounds for
.
Lemma 11.1.
For
,
with
.
-
Proof.
Fix
and
. Let
be the set of roots
satisfying
, then
|
(11.1)
|
Each term in this sum is positive, hence
.
For the second inequality, we use the estimate for
,
leading to
|
(11.2)
|
in view of 11.1 . Now note that
defines a
-equivariant linear map
. By Schur's lemma it equals
for some constant
. To determine
explicitly we fix a basis
of
and we denote
for the corresponding dual basis of
. Then
with
. Combined with 11.2 we obtain
. □
Corollary 11.2.
Let
. We have
if and only if
.
-
Proof.
Let
and suppose that
. Then for all
,
by Lemma 11.1 , hence
.
Conversely, suppose that
and let
such that
. By Proposition 2.7 this implies
. By the previous paragraph we conclude that
. On the other hand
, hence
and
. □
Proposition 2.8 is now a direct consequence of Corollary 11.2 and Lemma 11.1 .
12 The Pauli principle
In this section we complete the proof of Theorem 3.5 (and hence also of Theorem 2.4 ).
In view of Proposition 9.4 and Corollary 9.5 it suffices to show the following root system analog of the Pauli principle.
Proposition 12.1.
If
is singular then
.
For the proof of Proposition 12.1 we may assume without loss of generality that
is
-standard (in particular,
). We write
for the real sub-space spanned by the subset
of simple roots. Its complement in
is defined by
Observe that
iff
iff
is regular.
Consider the linear map
defined by
-
Proof.
Fix a singular
-standard solution
of the Bethe ansatz equations 2.10 (in particular
). By a similar computation as in the proof of Proposition 9.4 we obtain from 8.3 , Lemma 9.2 and Proposition 8.4 ,
| |
with vectors
(
) given by
| |
where we have used the involution on
defined by 9.4 , as well as 9.5 . For
we have
| |
Consequently 12.1 is equivalent to
Since
is a solution of the Bethe ansatz equations (see 9.6 for the convenient equivalent form of the Bethe ansatz equations) this is equivalent to
|
(12.2)
|
Note that 12.2 only depends on the coset
(
). Using the explicit expressions for
and
we can rewrite 12.2 as
|
(12.3)
|
We match 12.3 to the desired condition
as follows. Since
is an irreducible root system in
, the condition
is equivalent to
for all
, which in turn is equivalent to 12.3 if
|
(12.4)
|
for all
. To prove 12.4 we first observe that
by the explicit expression 8.8 for
. Using 2.13 this can be rewritten as
The second term can be rewritten as
| |
Combining the latter two formulas yields 12.4 . □
It follows from 10.3 that
with
and
the Hessian of the master function
at
. Since
is positive definite,
is a linear isomorphism. Proposition 12.1 thus is an immediate consequence of Lemma 12.2 .
References
-
V.M. Buchstaber, G. Felder, A.P. Veselov, Elliptic Dunkl operators, root systems, and functional equations, Duke Math. J. 76 (1994), 885–911.
-
I. Cherednik, A unification of Knizhnik-Zamolodchikov and Dunkl operators via affine Hecke algebras, Invent. Math. 106 (1991), 411–431.
-
I. Cherednik, Inverse Harish-Chandra transform and difference operators, Internat. Math. Res. Notices 1997, no. 15, 733–750.
-
I. Cherednik, Intertwining operators of double affine Hecke algebras, Selecta Math. (N.S.) 3, no. 4 (1997), 459–495.
-
T.C. Dorlas, Orthogonality and completeness of the Bethe ansatz eigenstates of the nonlinear Schroedinger model, Commun. Math. Phys. 154 (1993), 347–376.
-
C.F. Dunkl, Differential-difference operators associated to reflection groups, Trans. Amer. Math. Soc. 311, no. 1 (1989), 167–183.
-
M. Gaudin, Boundary energy of a Bose gas in one dimension, Phys. Rev. A 4, no. 1 (1971), 386–394.
-
M. Gaudin, La fonction d'Onde de Bethe, Collection du Commissariat à l'Énergie Atomique: Série Scientifique. Masson, Paris, 1983.
-
M. Girardeau, Relationship between systems of impenetrable bosons and fermions in one dimension, J. Math. Phys. 1, no. 6 (1960), 516–523.
-
E. Gutkin, B. Sutherland, Completely integrable systems and groups generated by reflections, Proc. Natl. Acad. Sci. USA 76, no. 12 (1979), 6057–6059.
-
E. Gutkin, Integrable systems with delta-potential, Duke Math. J. 49, no. 1 (1982), 1–21.
-
E. Gutkin, Conservation laws for the nonlinear Schrödinger equation, Ann. Inst. Henri Poincaré 2, no. 1 (1985), 76–74.
-
E. Gutkin, Operator calculi associated with reflection groups, Duke Math. J. 55, no. 1 (1987), 1–18.
-
G.J. Heckman, E.M. Opdam, Yang's system of particles and Hecke algebras, Ann. of Math. (2) 145, no. 1 (1997), 139–173.
-
K. Hikami, Notes on the
-function interacting gas. Intertwining operator in the degenerate affine Hecke algebra, J. Phys. A: Math. Gen. 31 (1998), L85–L91.
-
J.E. Humphreys, Reflection groups and Coxeter groups, Cambridge studies in Adv. Math. 29, Cambridge Univ. Press (1990).
-
A.G. Izergin, V.E. Korepin, The Pauli principle for one-dimensional bosons and the algebraic Bethe ansatz, Lett. Math. Phys. 6 (1982), 283–289.
-
Y. Komori, K. Hikami, Nonlinear Schrödinger model with boundary, integrability and scattering matrix based on the degenerate affine Hecke algebra, Int. J. Mod. Phys. A 12, no. 3 (1997), 5397–5410.
-
V.E. Korepin, Calculation of Bethe wave functions, Commun. Math. Phys. 86 (1982), 391–418.
-
V.E. Korepin, L.D. Faddeev, Quantization of solitons, Theor. Math. Phys. 25 (1975), 1039–1049.
-
E.H. Lieb, W. Liniger, Exact analysis of an interacting Bose gas. I. The general solution and the ground state, Phys. Rev. (2), 130 (1963), 1605–1616.
-
G. Lusztig, Affine Hecke algebras and their graded version, J. Amer. Math. Soc. 2, no. 3 (1989), 599–635.
-
S. Murakami, M. Wadati, Connection between Yangian symmetry and the quantum inverse scattering method, J. Phys. A: Math. Gen. 29 (1996), 7903–7915.
-
M.A. Olshanetsky, A.M. Perelomov, Quantum integrable systems related to Lie algebras, Physics Reports 94, no. 6 (1983), 313–404.
-
E.M. Opdam, Lecture notes on Dunkl operators for real and complex reflection groups, MSJ Memoirs, 8. Mathematical Society of Japan, Tokyo, 2000.
-
A.P. Polychronakos, Exchange operator formalism for integrable systems of particles, Phys. Rev. Lett. 69 (1992), 703–705.
-
E.K. Sklyanin, Boundary conditions for integrable quantum systems, J. Phys. A 21, no. 10 (1988), 2375–2389.
-
R. Steinberg, Differential equations invariant under finite reflection groups, Trans. Amer. Math. Soc. 112 (1964), 392–400.
-
B. Sutherland, Nondiffractive scattering: Scattering from kaleidoscopes, J. Math. Phys. 21, no. 7 (1980), 1770–1775.
-
C.N. Yang, Some exact results for the many-body problem in one dimension with repulsive delta-function interaction, Phys. Rev. Lett. 19 (1967), 1312–1315.
-
C.N. Yang, C.P. Yang, Thermodynamics of a one-dimensional system of bosons with repulsive delta-function interaction, J. Math. Phys. 10 (1969), 1115–1122.
KdV Institute for Mathematics, Universiteit van Amsterdam, Plantage Muidergracht 24, 1018 TV Amsterdam, The Netherlands. E-mail address : eemsiz@science.uva.nl, opdam@science.uva.nl, jstokman@science.uva.nl