1 Introduction

In recent years, there has been much progress in the development of integrable models in probability, particularly interacting particle systems related to representation theory and integrable systems. A well-known example is the coupled system of SDE’s

$$\begin{aligned} dx_1=d\beta _1,\qquad dx_i=d\beta _i+e^{x_{i-1}-x_i} dt, \ i=2,\ldots ,N, \end{aligned}$$
(1.1)

where \(\beta _i\) are independent one-dimensional Brownian motions. This process is closely related to the Toda lattice and has been extensively studied [5, 6, 23, 34, 39, 44, 45].

It is natural to consider non-commutative generalisations of such processes. In this paper, we consider some interacting systems of Brownian particles in the space \(\mathcal {P}\) of positive \(n\times n\) real symmetric matrices. One of the main examples we consider is a generalisation of the system (1.1), a diffusion process in \(\mathcal {P}^N\) with infinitesimal generator

$$\begin{aligned} T=\Delta _{X_1}+\sum _{i=2}^N [\Delta _{X_i} + 2 \,\text{ tr }\,(X_{i-1}\partial _{X_i})], \end{aligned}$$
(1.2)

where \(\Delta _X\) denotes the Laplacian, and \(\partial _X\) denotes the partial matrix derivative, on \(\mathcal {P}\). In the case \(n=1\) with \(x_i=\ln X_i\), it is equivalent to the system (1.1). We will show that this process is related to a quantisation of the non-Abelian Toda lattice in \(\mathcal {P}\).

In another direction, Matsumoto and Yor [30] obtained an analogue of Pitman’s \(2M-X\) theorem for exponential functions of Brownian motion, which is closely related to Dufresne’s identity [16]. These results were recently extended to the matrix setting by Rider and Valkó [42] (see also Bougerol [7] for related results in the complex case). We discuss this example in some detail, and give a new proof (and slight generalisation) of Rider and Valkó’s result. We also consider an example of a pair of interacting Brownian particles in \(\mathcal {P}\) with a ‘reflecting wall’.

The outline of the paper is as follows. In the next section we present some preliminary background material. This is followed, in Sects. 36, by a series of examples with small numbers of particles. In Sect. 7, we discuss the example (1.2) and its relation to a quantisation of the non-Abelian Toda lattice. In Sect. 8, we outline how this example is related to some Bäcklund transformations for the classical non-Abelian Toda lattice. In Sect. 9, we briefly discuss a related class of processes in which the underlying motion of particles is not governed by the Laplacian, but rather a related diffusion process which was introduced and studied by Norris, Rogers and Williams [33]. In Sect. 10, we briefly outline how the framework developed in this paper extends to the complex setting, with particular emphasis on some extensions of the (Hermitian) Matrix Dufresne identity of [42].

2 Preliminaries

We mostly follow the nomenclature of Terras [47], to which we refer the reader for more background. Let \(\mathcal {P}\) denote the space of positive \(n\times n\) real symmetric matrices. For \(a\in GL(n)\) and \(X\in \mathcal {P}\), write \(X[a]=a^t X a\). This defines an action of GL(n) on \(\mathcal {P}\).

For \(X\in \mathcal {P}\), we will use the notation

$$\begin{aligned} |X|=\det (X),\qquad \, \text{ etr }\,(X)=\exp (\,\text{ tr }\,(X)), \end{aligned}$$

and denote by \(X^{1/2}\) the unique positive square root of X.

2.1 Differential operators

The partial derivative on \(\mathcal {P}\) is defined, writing \(X=(x_{ij})\), by

$$\begin{aligned} \partial _X = \left( \frac{1}{2}(1+\delta _{ij})\frac{\partial }{\partial x_{ij}} \right) _{1\le i,j \le n}. \end{aligned}$$

We define the Laplacian on \(\mathcal {P}\) by \(\Delta _X = \,\text{ tr }\,\vartheta _X^2\), where \(\vartheta _X = X \partial _X\). The Laplacian is a GL(n)-invariant differential operator on \(\mathcal {P}\), meaning \((\Delta f)^a=\Delta f^a\) for all \(a\in GL(n)\), where \(f^a(X)=f(X[a])\). In fact, the differential operators \(L_k=\,\text{ tr }\,(\vartheta _X)^k,\ k=1,2,\ldots \) are all GL(n)-invariant [47, Exercise 1.1.27]. This follows from the fact that, if \(Y=X[a]\) for some fixed \(a\in GL(n)\), then

$$\begin{aligned} \partial _X=a\, \partial _Y \, a^t. \end{aligned}$$
(2.1)

If \(Y=X^{-1}\), then \(\vartheta _Y=-\vartheta _X'\), where \(\vartheta _X' f = (\partial _X f)X\). It follows that \(L_k^Y=(-1)^k L_k^X\). In particular, the Laplacian is invariant under this change of variable.

2.2 Chain rule for quadratic change of variables

It is known [11, 29] that for each \(X,Y\in \mathcal {P}\), the equation \(Y=AXA\) has a unique solution in \(\mathcal {P}\), namely

$$\begin{aligned} A=X^{-1/2}(X^{1/2}YX^{1/2})^{1/2} X^{-1/2}. \end{aligned}$$

This fact may also be deduced from [49, Theorem 4.1]. If X is fixed, then

$$\begin{aligned} \partial _A=XA\, \partial _Y+\partial _Y AX. \end{aligned}$$
(2.2)

2.3 Calculus

For \(A\in \mathcal {P}\) fixed:

$$\begin{aligned} \vartheta _X \,\text{ tr }\,(AX)= & {} XA,\qquad \vartheta _X \,\text{ tr }\,(AX^{-1}) = -AX^{-1} \end{aligned}$$
(2.3)
$$\begin{aligned} \Delta _X \,\text{ tr }\,(AX)= & {} \frac{n+1}{2} \,\text{ tr }\,(AX),\qquad \Delta _X \ \,\text{ tr }\,(AX^{-1})= \frac{n+1}{2} \,\text{ tr }\,(AX^{-1})\end{aligned}$$
(2.4)
$$\begin{aligned} \Delta _X \ F= & {} \left[ \Delta _X \ln F + \,\text{ tr }\,(\vartheta _X\ln F)^2\right] F \end{aligned}$$
(2.5)
$$\begin{aligned} \Delta _X \, \text{ etr }\,(AX)= & {} \left[ \frac{n+1}{2} \,\text{ tr }\,(AX) + \,\text{ tr }\,(AXAX)\right] \, \text{ etr }\,(AX) \end{aligned}$$
(2.6)
$$\begin{aligned} \Delta (fg)= & {} (\Delta f)g+f (\Delta g)+2\,\text{ tr }\,[(\vartheta _Xf)( \vartheta _X g)] \end{aligned}$$
(2.7)
$$\begin{aligned} \vartheta _X (AX)= & {} \frac{1}{2} (A+\,\text{ tr }\,A) X \end{aligned}$$
(2.8)

For positive integers k,

$$\begin{aligned} \vartheta _X \,\text{ tr }\,X^k= & {} k X^k,\qquad \vartheta _X^2 \,\text{ tr }\,X^k=k \vartheta _X(X^k) = \frac{k^2}{2} X^k+ \frac{k}{2} \sum _{j=1}^{k} X^j \,\text{ tr }\,X^{k-j} \nonumber \\\end{aligned}$$
(2.9)
$$\begin{aligned} \Delta _X \,\text{ tr }\,X^k= & {} \frac{k^2}{2} \,\text{ tr }\,X^k + \frac{k}{2} \sum _{j=1}^{k} \,\text{ tr }\,X^j \,\text{ tr }\,X^{k-j} \end{aligned}$$
(2.10)
$$\begin{aligned} \Delta _X \,\text{ tr }\,X^{-k}= & {} \frac{k^2}{2} \,\text{ tr }\,X^{-k}+ \frac{k}{2} \sum _{j=1}^{k} \,\text{ tr }\,X^{-j} \,\text{ tr }\,X^{j-k} \end{aligned}$$
(2.11)

2.4 Integration

Denote the GL(n)-invariant volume element on \(\mathcal {P}\) by

$$\begin{aligned} \mu (dX)=|X|^{-(n+1)/2}\prod _{1\le i\le j\le n} dx_{ij}, \end{aligned}$$

where \(dx_{ij}\) is the Lebesgue measure on \(\mathbb {R}\). If we write \(X=a[k]\), where a is diagonal with entries \(a_1,\ldots ,a_n>0\) and \(k\in O(n)\), then

$$\begin{aligned} \mu (dX)=c_n dk \prod _{i<j} |a_i-a_j| \prod _{i=1}^n a_i^{-(n-1)/2-1} da_i , \end{aligned}$$
(2.12)

where dk is the normalised Haar measure on O(n) and \(c_n\) is a normalisation constant.

2.5 Power and Gamma functions

For \(s\in \mathbb {C}^n\), define the power function

$$\begin{aligned} p_s(X)=\prod _{k=1}^n |X^{(k)}|^{s_k}, \end{aligned}$$

where \(X^{(k)}\) denotes the \(k\times k\) upper left hand corner of X. For \(s\in \mathbb {C}^n\) satisfying

$$\begin{aligned} 2\mathfrak {R}(s_k+\ldots +s_n)>k-1,\qquad k=1,\ldots ,n \end{aligned}$$
(2.13)

define

$$\begin{aligned} \Gamma _n(s)=\int _\mathcal {P}p_s(X) \, \text{ etr }\,(- X) \mu (dX) = \pi ^{n(n-1)/4} \prod _{k=1}^n \Gamma \left( s_k+\ldots +s_n-\frac{k-1}{2}\right) . \end{aligned}$$

For \(s=(0,\ldots ,0,\nu )\), we will write \(p_s(X)=e_\nu (X)=|X|^\nu \) and \(\Gamma _n(\nu )=\Gamma _n(0,\ldots ,0,\nu )\). The spherical functions on \(\mathcal {P}\) are defined, for \(s\in \mathbb {C}^n\), by

$$\begin{aligned} h_s(X)=\int _{O(n)} p_s(X[k])\; dk, \end{aligned}$$

where dk denotes the normalised Haar measure on O(n).

The power function \(p_s\) is an eigenfunction of the Laplacian on \(\mathcal {P}\), with eigenvalue

$$\begin{aligned} \lambda _2(s)=\sum _{i=1}^n r_i^2+\frac{n-n^3}{48},\qquad r_i=s_i+\cdots +s_n+\frac{n+1-2i}{4}. \end{aligned}$$

(See, for example, [47, Exercise 1.2.12 & Equation (1.93)].) The functions \(p_{s,k}(X)=p_s(X[k]),\ k\in O(n)\), and \(h_s(X)\), are also eigenfunctions of \(\Delta \) with eigenvalue \(\lambda _2(s)\).

For \(s=(0,\ldots ,0,\nu )\), so that \(p_s(X)=e_\nu (X)=|X|^\nu \), we note:

$$\begin{aligned} \vartheta _X e_\nu (X)= & {} \nu \, e_\nu (X)\, I_n,\qquad \,\text{ tr }\,\vartheta _X \, e_\nu (X) = n\nu \, e_\nu (X), \qquad \,\text{ tr }\,\vartheta _X \, \ln e_\nu (X) = n\nu ,\nonumber \\ \end{aligned}$$
(2.14)
$$\begin{aligned} \Delta _X \, e_\nu (X)= & {} n\nu ^2 \, e_\nu (X),\qquad \Delta _X \, \ln e_\nu (X) = 0. \end{aligned}$$
(2.15)

2.6 Bessel functions

Following Terras [47], for \(s\in \mathbb {C}^n\) and \(V,W\in \mathcal {P}\), define

$$\begin{aligned} K_n(s|V,W)=\int _\mathcal {P}p_s(Y) \, \text{ etr }\,(- VY - WY^{-1}) \mu (dY). \end{aligned}$$

As remarked in [47], following Exercise 1.2.16, we can always reduce one of the positive matrix arguments in \(K_n(s|V,W)\) to the identity: if \(W=I[g]\), where g is upper triangular with positive diagonal coefficients, then

$$\begin{aligned} K_n(s| V,W) = p_s(W) K_n(s|gVg^t,I). \end{aligned}$$
(2.16)

In the following, if \(s=(0,\ldots ,0,\nu )\) we will write \(K_n(s|V,W)=K_n(\nu |V,W)\).

For \(\nu \in \mathbb {C}\) and \(X\in \mathcal {P}\), define

$$\begin{aligned} B_\nu (X)=\int _\mathcal {P}e_\nu (Y) \, \text{ etr }\,(- XY - Y^{-1}) \mu (dY). \end{aligned}$$
(2.17)

This function was introduced by Herz [22], and is related to \(K_n\) by

$$\begin{aligned} B_\nu (X)=K_n(\nu | X,I). \end{aligned}$$

We note that \(B_{-\nu }(X)=e_\nu (X)B_\nu (X)\), which implies

$$\begin{aligned} e_\nu (V) K_n(\nu |V,W)=e_\nu (W) K_n(-\nu |V,W). \end{aligned}$$
(2.18)

The asymptotic behaviour of \(B_\nu (X)\) for large arguments has been studied via Laplace’s method in the paper [12], see also [20, Appendix B]. If we fix \(M\in \mathcal {P}\) and let \(X=z^2 M^2/2\), then it holds that, as \(z\rightarrow \infty \),

$$\begin{aligned} B_\nu (X)=C(\nu ,M) z^{-\nu -n(n+1)/4} e^{-z\,\text{ tr }\,M} (1+O(z^{-1})), \end{aligned}$$

where

$$\begin{aligned} C(\nu ,M)=2^{\nu -n(n+1)/4} \pi ^{n(n+1)/4} |M|^{-\nu -1/2} \prod _{i<j}(m_i+m_j)^{-1/2}, \end{aligned}$$

and \(m_i\) denote the eigenvalues of M. In particular, taking \(M=I\) and \(z^2/2=\alpha \), say, we deduce from the same application of Laplace’s method, the following lemma, which we record here for later reference.

Lemma 1

For \(\alpha >0\), let \(A(\alpha )\) be distributed according to the (matrix GIG) law

$$\begin{aligned} B_\nu (\alpha I)^{-1} \, \text{ etr }\,(-\alpha A-A^{-1}) \mu (dA). \end{aligned}$$

Then \(\alpha ^{1/2} A(\alpha )\rightarrow I\), in probability, as \(\alpha \rightarrow \infty \).

2.7 Standard probability distributions

The Wishart distribution on \(\mathcal {P}\) with parameters \(\Sigma \in \mathcal {P}\) and \(p>n-1\) has density

$$\begin{aligned} \Gamma _n(p/2)^{-1} |\Sigma ^{-1}X/2|^{p/2} \, \text{ etr }\,(-\Sigma ^{-1}X/2) \mu (dX). \end{aligned}$$

If \(p\ge n\) is an integer and A is a \(p\times n\) random matrix with independent standard normal entries, then \(\Sigma ^{1/2}A^tA\Sigma ^{1/2}\) is Wishart distributed with parameters \(\Sigma \) and p. The inverse Wishart distribution on \(\mathcal {P}\), with parameters \(\Sigma \in \mathcal {P}\) and \(p>n-1\), is the law of the inverse of a Wishart matrix with parameters \(\Sigma \) and p, and has density

$$\begin{aligned} \Gamma _n(p/2)^{-1} |\Sigma ^{-1}X^{-1}/2|^{p/2} \, \text{ etr }\,(-\Sigma ^{-1}X^{-1}/2) \mu (dX). \end{aligned}$$

The matrix GIG (generalised inverse Gaussian) distribution on \(\mathcal {P}\) with parameters \(\nu \in \mathbb {R}\) and \(A,B\in \mathcal {P}\) is defined by

$$\begin{aligned} K_n(\nu | A,B)^{-1} |X|^{\nu } \, \text{ etr }\,(-AX-BX^{-1}) \mu (dX). \end{aligned}$$

2.8 Invariant kernels

A kernel k(XY) is invariant if, for all \(a\in GL(n)\),

$$\begin{aligned} k(X[a],Y[a])=k(X,Y). \end{aligned}$$

For k sufficiently smooth, this implies that

$$\begin{aligned} \Delta _X k(X,Y)=\Delta _Y k(X,Y). \end{aligned}$$

Examples include

$$\begin{aligned} k(X,Y)=\, \text{ etr }\,(-YX^{-1}) \end{aligned}$$
(2.19)

and, more generally, for \(\nu \in \mathbb {C}\),

$$\begin{aligned} k_\nu (X,Y)=e_{-\nu }(YX^{-1}) \, \text{ etr }\,(-YX^{-1}). \end{aligned}$$
(2.20)

We note that the kernel k defined by (2.19) satisfies

$$\begin{aligned} \vartheta _X \ln k= & {} - \vartheta _Y \ln k = YX^{-1} , \end{aligned}$$
(2.21)
$$\begin{aligned} \Delta _X k= & {} \Delta _Y k = \left[ - \frac{n+1}{2} \,\text{ tr }\,(YX^{-1}) + \,\text{ tr }\,(YX^{-1}YX^{-1})\right] k . \end{aligned}$$
(2.22)

2.9 Brownian motion and diffusion

We define Brownian motion in \(\mathcal {P}\) with drift \(\nu \in \mathbb {R}\) to be the diffusion process in \(\mathcal {P}\) with generator

$$\begin{aligned} \Delta _X^{(\nu )} = \Delta _X + 2\nu \, \,\text{ tr }\,\vartheta _{X}. \end{aligned}$$

More generally, if \(\varphi \) is a positive eigenfunction of the Laplacian on \(\mathcal {P}\) with eigenvalue \(\lambda \), then we may consider the corresponding Doob transform

$$\begin{aligned} \Delta _X^{(\varphi )} = \varphi (X)^{-1} \circ (\Delta _X-\lambda ) \circ \varphi (X) = \Delta _X+2\,\text{ tr }\,(\vartheta _X\ln \varphi (X) \ \vartheta _X). \end{aligned}$$

If \(\varphi (X)=|X|^\nu \) for some \(\nu \in \mathbb {R}\), then \(\lambda =\nu ^2 n\) and \(\Delta _X^{(\varphi )} \equiv \Delta _X^{(\nu )}\). We shall refer to the diffusion process with infinitesimal generator \(\Delta _X^{(\varphi )}\) as a Brownian motion in \(\mathcal {P}\) with drift \(\varphi \).

A Brownian motion in \(\mathcal {P}\), with drift \(\nu \), may be constructed as follows. Let \(b_t,\ t\ge 0\) be a standard Brownian motion in the Lie algebra \(\mathfrak {gl}(n,\mathbb {R})\) of real \(n\times n\) matrices, that is, each matrix entry evolves as a standard Brownian motion on the real line. Set \(\beta _t=b_t/\sqrt{2}+\nu t\). Define a Markov process \(G_t,\ t\ge 0\) in GL(n) via the Stratonovich SDE: \(\partial G_t = \partial \beta _t\, G_t\). When \(\nu =0\), this is called a right-invariant Brownian motion in GL(n); thus, we shall refer to G as a right-invariant Brownian motion in GL(n) with drift \(\nu \). Then \(Y=G^tG\) is a Brownian motion in \(\mathcal {P}\) with drift \(\nu \). Note that Y satisfies the Stratonovich SDE

$$\begin{aligned} \partial Y = G^t (\partial \beta +\partial \beta ^t) G. \end{aligned}$$

By orthogonal invariance of the underlying Brownian motion in \(\mathfrak {gl}(n,\mathbb {R})\), one may replace the G and \(G^t\) factors in this equation by \(Y^{1/2}\) to obtain a closed SDE for the evolution of Y.

We will also consider more general diffusions on \(\mathcal {P}^r\), with generators of the form

$$\begin{aligned} L=\sum _{i=1}^r [\Delta _{X_i} + \,\text{ tr }\,(a_i(X)\partial _{X_i})], \end{aligned}$$

where the \(a_i\) are locally Lipschitz functions on \(\mathcal {P}^r\). For such generators, we may take the domain to be \(C^2_c(\mathcal {P}^r)\), the set of continuously twice differentiable, compactly supported, functions on \(\mathcal {P}^r\). If \(\rho \) is a probability measure on \(\mathcal {P}^r\) and the martingale problem associated with \((L,\rho )\) is well posed, then we may construct a realisation of the corresponding Markov process by solving the (Stratonovich) SDE’s:

$$\begin{aligned} \partial X_i = X_i^{1/2} \partial S_i X_i^{1/2} + a_i(X) dt, \end{aligned}$$

where \(b_1,\ldots ,b_r\) are independent standard Brownian motions in \(\mathfrak {gl}(n,\mathbb {R})\), \(S_i=(b_i+b_i^t)/\sqrt{2}\), and X(0) is chosen according to \(\rho \).

3 Brownian particles with one-sided interaction

Consider the differential operator on \(\mathcal {P}^2\) defined by

$$\begin{aligned} T = \Delta _Y + \Delta _X + 2\,\text{ tr }\,(Y\partial _X). \end{aligned}$$
(3.1)

Let \(k(X,Y)=\, \text{ etr }\,(-YX^{-1})\) and consider the integral operator defined, for suitable test functions f on \(\mathcal {P}^2\), by

$$\begin{aligned} (K f)(X) = \int _\mathcal {P}k(X,Y) f(X,Y) \mu (dY),\qquad X\in \mathcal {P}. \end{aligned}$$

Then, on a suitable domain, the following intertwining relation holds:

$$\begin{aligned} \Delta \circ K = K \circ T. \end{aligned}$$
(3.2)

Indeed, let us write \(k=k(X,Y)\), \(f=f(X,Y)\) and note the following identities:

$$\begin{aligned} \Delta _X k= & {} \Delta _Y k,\qquad \vartheta _X k=YX^{-1}k,\\ \Delta _X(kf)= & {} f \Delta _X k+k\Delta _X f+2\,\text{ tr }\,(\vartheta _Xk\vartheta _Xf) = f \Delta _Y k+k\Delta _X f+2k\,\text{ tr }\,(Y\partial _X f). \end{aligned}$$

It follows, using the fact that \(\Delta \) is self-adjoint with respect to \(\mu \), that

$$\begin{aligned} \Delta (K f) (X)= & {} \Delta _X \int _\mathcal {P}k f \mu (dY) \\= & {} \int _\mathcal {P}(f\Delta _Y k+k\Delta _X f+2k\,\text{ tr }\,(Y\partial _X f)) \mu (dY)\\= & {} \int _\mathcal {P}k \; (Tf)\; \mu (dY), \end{aligned}$$

as required.

Now suppose \(\varphi \) is a positive eigenfunction of \(\Delta \) with eigenvalue \(\lambda \) such that

$$\begin{aligned} \tilde{\varphi }(X)=\int _\mathcal {P}\varphi (Y) k(X,Y) \mu (dY)<\infty . \end{aligned}$$

Then \(\tilde{\varphi }\) is also a positive eigenfunction of \(\Delta \) with eigenvalue \(\lambda \):

$$\begin{aligned} \Delta _X \tilde{\varphi }(X)= & {} \int _\mathcal {P}\varphi (Y) \Delta _X k(X,Y) \mu (dY) \\= & {} \int _\mathcal {P}\varphi (Y) \Delta _Y k(X,Y) \mu (dY) \\= & {} \int _\mathcal {P}[\Delta _Y \varphi (Y)] k(X,Y) \mu (dY) \\= & {} \lambda \int _\mathcal {P}\varphi (Y) k(X,Y) \mu (dY) = \lambda \tilde{\varphi }(X). \end{aligned}$$

For example, if \(\varphi =p_s\), for some \(s\in \mathbb {R}^n\) satisfying (2.13), then \(\tilde{\varphi }=\Gamma _n(s)\varphi \) (see, for example, [47, Exercise 1.2.4]). Similarly, if \(s\in \mathbb {R}^n\) satisfies (2.13), and \(\varphi =h_s\) or \(\varphi =p_{s,k}\) for some \(k\in O(n)\), then it also holds that \(\tilde{\varphi }=\Gamma _n(s)\varphi \). More generally, \(\tilde{\varphi }\) is a constant multiple of \(\varphi \) whenever \(\varphi \) is a simultaneous eigenfunction of the Laplacian and the integral operator with kernel k(XY); note that these two operators commute, since \(\Delta _X k = \Delta _Y k\).

Define

$$\begin{aligned} (K_\varphi f)(X) = \tilde{\varphi }(X)^{-1} \int _\mathcal {P}\varphi (Y) k(X,Y) f(X,Y) \mu (dY), \end{aligned}$$

and

$$\begin{aligned} T^{(\varphi )}=\varphi (Y)^{-1}\circ (T-\lambda ) \circ \varphi (Y)=\Delta ^{(\varphi )}_Y + \Delta _X + 2\,\text{ tr }\,(Y\partial _X). \end{aligned}$$

Then (3.2) extends to:

$$\begin{aligned} \Delta ^{(\tilde{\varphi })} \circ K_\varphi = K_\varphi \circ T^{(\varphi )}. \end{aligned}$$
(3.3)

The intertwining relation (3.3) has a probabilistic interpretation, as follows. Set

$$\begin{aligned} \pi (X,Y)=\tilde{\varphi }(X)^{-1} \varphi (Y) k(X,Y). \end{aligned}$$

Let \(\rho \) be a probability measure on \(\mathcal {P}\) and define a probability measure on \(\mathcal {P}\times \mathcal {P}\) by

$$\begin{aligned} \sigma (dX,dY)=\pi (X,Y) \rho (dX) \mu (dY). \end{aligned}$$

Suppose that \(\varphi \) is such that the martingale problems associated with \((\Delta ^{(\tilde{\varphi })},\rho )\) and \((T^{(\varphi )},\sigma )\) are well-posed, and that \((X_t,Y_t)\) is a diffusion process with infinitesimal generator \(T^{(\varphi )}\) and initial law \(\sigma \). Then it follows from the theory of Markov functions (see Appendix A) that, with respect to its own filtration, \(X_t\) is a Brownian motion with drift \(\tilde{\varphi }\) and initial distribution \(\rho \); moreover, the conditional law of \(Y_t\), given \(X_s,\ s\le t\), only depends on \(X_t\) and is given by \( \pi (X_t,Y) \mu (dY)\). This statement is analogous to the Burke output theorem for the M/M/1 queue, although in this context the ‘output’ (a Brownian motion with drift \(\tilde{\varphi }\)) need not have the same law as the ‘input’ (a Brownian motion with drift \(\varphi \)). Note however that these Brownian motions are equivalent whenever \(\tilde{\varphi }\) is a constant multiple of \(\varphi \), so whenever this holds the output does have the same law as the input. This is always the case when \(n=1\), as was observed in the paper [39].

We note that the intertwining relation (3.3) also implies that

$$\begin{aligned} (T^{(\varphi )})^*\pi =0, \end{aligned}$$

where \((T^{(\varphi )})^*\) is the formal adjoint of \(T^{(\varphi )}\).

One can replace k by any invariant kernel \(k'\) and the above remains valid with

$$\begin{aligned} T=\Delta _Y + \Delta _X + 2\,\text{ tr }\,(\vartheta _X\ln k'(X,Y)\ \vartheta _X). \end{aligned}$$

For example, if \(k'=k_{\nu }\), defined by (2.20), then

$$\begin{aligned} T = \Delta _Y + \Delta ^{(\nu )}_X + 2\,\text{ tr }\,(Y\partial _X). \end{aligned}$$

In this case, we require \(\tilde{\varphi }= k_\nu \varphi \) to be finite, where

$$\begin{aligned} k_\nu \varphi (X)=\int _\mathcal {P}\varphi (Y) k_\nu (X,Y) \mu (dY). \end{aligned}$$

For example, if \(\varphi (X)=|X|^\lambda \), then this holds provided \(2(\lambda -\nu )>n-1\), in which case \(k_\nu \varphi =\Gamma _n(\lambda -\nu )\varphi \). For this example, the associated martingale problems are well posed, as shown in Appendix B (Example 2), so we may state the following theorem.

For \(2a>n-1\), we define the Markov kernel

$$\begin{aligned} \Pi _a(X,dY)=\Gamma _n(a)^{-1} |YX^{-1}|^a \, \text{ etr }\,(-YX^{-1}) \mu (dY). \end{aligned}$$
(3.4)

Theorem 2

Suppose \(2(\lambda -\nu )>n-1\), and let \((X_t,Y_t)\) be a diffusion process in \(\mathcal {P}^2\) with infinitesimal generator

$$\begin{aligned} T_{\lambda ,\nu }=\Delta _Y^{(\lambda )} + \Delta ^{(\nu )}_X + 2\,\text{ tr }\,(Y\partial _X), \end{aligned}$$

and initial law \(\delta _{X_0}(dX) \Pi _{\lambda -\nu }(X,dY)\). Then, with respect to its own filtration, \(X_t\) is a Brownian motion with drift \(\lambda \) started at \(X_0\). Moreover, the conditional law of \(Y_t\), given \(X_s,\ s\le t\), only depends on \(X_t\) and is given by \( \Pi _{\lambda -\nu }(X_t,dY)\).

The above example extends naturally to a system of N particles with one-sided interactions, as follows. Let \(\nu _2,\ldots ,\nu _N\in \mathbb {R}\), and \(\varphi \) a positive eigenfunction of \(\Delta \) such that

$$\begin{aligned} \tilde{\varphi }(X)=(k_{\nu _N}\circ \cdots \circ k_{\nu _2})\varphi (X)<\infty . \end{aligned}$$

For example, if \(\varphi (X)=|X|^{\nu _1}\) then this condition is satisfied provided \(\nu _i<\nu _1\) for all \(1<i\le N\), in which case we have

$$\begin{aligned} \tilde{\varphi }(X)=\prod _{i=2}^N \Gamma _n(\nu _1-\nu _i) \ |X|^{\nu _1}. \end{aligned}$$

Define

$$\begin{aligned} T= & {} \Delta _{X_1}^{(\varphi )}+\sum _{i=2}^N [\Delta _{X_i}^{(\nu _i)} + 2 \,\text{ tr }\,(X_{i-1}\partial _{X_i})],\nonumber \\ \pi (X_1,\ldots ,X_N)= & {} \tilde{\varphi }(X_N)^{-1} \varphi (X_1) \prod _{i=2}^N k_{\nu _i}(X_i,X_{i-1}),\nonumber \\ (Kf)(X_N)= & {} \int _{\mathcal {P}^{N-1}} \pi (X_1,\ldots ,X_{N-1},X_N) \, f(X_1,\ldots , X_N) \,\mu _{N-1}(dX_1,\ldots , dX_{N-1}).\nonumber \\ \end{aligned}$$
(3.5)

Then

$$\begin{aligned} \Delta _{X_N}^{(\tilde{\varphi })} \circ K = K \circ T. \end{aligned}$$

This implies that \(T^*\pi =0\) and, moreover, if \(\varphi \) is such that the relevant martingale problems are well posed and the system is started in equilibrium, then \(X_N\) is a Brownian motion, in its own filtration, with drift \(\tilde{\varphi }\). This certainly holds in the case \(\varphi (X)=|X|^{\nu _1}\), with \(\nu _i<\nu _1\) for all \(1<i\le N\). Note that this can also be seen as a direct consequence of Theorem 2.

Finally we remark that, by a simple change of variables, one may also consider

$$\begin{aligned} T'=\Delta _{X_1}^{(\varphi )}+\sum _{i=2}^N [\Delta _{X_i}^{(\nu _i)} - 2 \,\text{ tr }\,(X_i X_{i-1}^{-1}\vartheta _{X_i})]. \end{aligned}$$
(3.6)

The operator \(T'\) is related to T as follows. Write \(T=T_X(\varphi ,\nu )\), \(T'=T'_X(\varphi ,\nu )\). Then, under the change of variables \(Y_i=X_i^{-1}\), \(T_X(\varphi ,\nu )=T'_Y(\bar{\varphi },\bar{\nu })\), where \(\bar{\varphi }(Y)=\varphi (X^{-1})\) and \(\bar{\nu }_i=-\nu _i\).

In this case, if we assume that \(\tilde{\varphi } = (k_{-\nu _N}\circ \cdots \circ k_{-\nu _2})\varphi \) is finite, and define

$$\begin{aligned} \pi '(X_1,\ldots ,X_N)= & {} \tilde{\varphi }(X_N)^{-1} \varphi (X_1) \prod _{i=2}^N k_{-\nu _i}(X_{i-1},X_{i}),\\ (K'f)(X_N)= & {} \int _{\mathcal {P}^{N-1}} \pi '(X_1,\ldots ,X_{N-1},X_N) \, f(X_1,\ldots , X_N) \,\mu _{N-1}(dX_1,\ldots , dX_{N-1}), \end{aligned}$$

then it holds that \(\Delta _{X_N}^{(\tilde{\varphi })} \circ K' = K' \circ T'\), with the analogous conclusions.

4 Connection with Bessel functions

The previous example, with two particles, extends naturally to

$$\begin{aligned} G=\Delta _Y + \Delta _{X_1} +2\,\text{ tr }\,(Y\partial _{X_1})+ \Delta _{X_2}-2\,\text{ tr }\,(X_2Y^{-1}X_2\partial _{X_2}). \end{aligned}$$

Note that this is a combination of the T and \(T'\) of the previous section.

Writing \(X=(X_1,X_2)\), define

$$\begin{aligned} H=\Delta _{X_1}+\Delta _{X_2}-2\,\text{ tr }\,(X_1^{-1}X_2),\qquad q(X,Y)= \, \text{ etr }\,(-YX_1^{-1}-X_2 Y^{-1}). \end{aligned}$$

Consider the integral operator, defined for suitable f on \(\mathcal {P}^2\times \mathcal {P}\) by

$$\begin{aligned} (Qf)(X)=\int _\mathcal {P}q(X,Y) f(X,Y) \mu (dY),\qquad X\in \mathcal {P}^2. \end{aligned}$$

Then the following intertwining relation holds:

$$\begin{aligned} H \circ Q = Q\circ G. \end{aligned}$$
(4.1)

Indeed, let us write \(q=q(X,Y)\), \(f=f(X,Y)\) and note that

$$\begin{aligned} H_Xq=\Delta _Y q,\qquad H_X(qf)=f \Delta _Y q+q(G-\Delta _Y)f. \end{aligned}$$

The claim follows, using the fact that \(\Delta \) is self-adjoint with respect to \(\mu \).

Suppose that \(\varphi \) is a positive eigenfunction of \(\Delta \) with eigenvalue \(\lambda \) such that

$$\begin{aligned} \psi (X)=\int _\mathcal {P}\varphi (Y) q(X,Y) \mu (dY)<\infty . \end{aligned}$$

Then \(\psi \) is a positive eigenfunction of H with eigenvalue \(\lambda \). We remark that, if \(\varphi =p_s\), then

$$\begin{aligned} \psi (X)= K_n(s|\ X_1^{-1},X_2). \end{aligned}$$

Let us define

$$\begin{aligned} (Q_\varphi f)(X)= & {} \psi (X)^{-1} \int _\mathcal {P}\varphi (Y) q(X,Y) f(X,Y) \mu (dY),\\ H^{(\psi )}= & {} \psi (X)^{-1}\circ (H-\lambda ) \circ \psi (X),\\ G^{(\varphi )}= & {} \varphi (Y)^{-1}\circ (G-\lambda ) \circ \varphi (Y)=\Delta ^{(\varphi )}_Y + \Delta _{X_1}\\&+2\,\text{ tr }\,(Y\partial _{X_1})+ \Delta _{X_2}-2\,\text{ tr }\,(X_2Y^{-1}X_2\partial _{X_2}). \end{aligned}$$

Then (4.1) extends to:

$$\begin{aligned} H^{(\psi )} \circ Q_\varphi = Q_\varphi \circ G^{(\varphi )}. \end{aligned}$$
(4.2)

This intertwining relation has a probabilistic interpretation, as follows. Let \(\rho \) be a probability measure on \(\mathcal {P}^2\) and define a probability measure on \(\mathcal {P}^2\times \mathcal {P}\) by

$$\begin{aligned} \sigma (dX,dY)=\psi (X)^{-1} \rho (dX) \varphi (Y) q(X,Y) \mu (dY). \end{aligned}$$

Suppose that \(\varphi \) is such that the martingale problems associated with \((H^{(\psi )},\rho )\) and \((G^{(\varphi )},\sigma )\) are well-posed, and that \((X_t,Y_t)\) is a diffusion process with infinitesimal generator \(G^{(\varphi )}\) and initial law \(\sigma \). Then, with respect to its own filtration, \(X_t\) is a diffusion with generator \(H^{(\psi )}\) and initial distribution \(\rho \); moreover, the conditional law of \(Y_t\), given \(X_s,\ s\le t\), only depends on \(X_t\) and is given by

$$\begin{aligned} \psi (X_t)^{-1} \varphi (Y) q(X_t,Y) \mu (dY). \end{aligned}$$

The above example is a special case of a more general construction which will be discussed in Sect. 7.

5 Matrix Dufresne identity and \(2M-X\) theorem

Let \(M=\Delta _Y +\,\text{ tr }\,(Y\partial _A)\). If \(Y=AXA\) then, in the variables (XA), we can write

$$\begin{aligned} M=\Delta _X -2\,\text{ tr }\,(XAX\partial _X)+\,\text{ tr }\,(AXA\partial _A). \end{aligned}$$

To see this, let \(f=f(X,A)=g(AXA,A)\) and first note that, by invariance,

$$\begin{aligned} \Delta _Xf=\Delta _Y g(Y,A)\ \Big |_{Y=AXA}. \end{aligned}$$

Let us write \(g_1(Y,A)=\partial _Y g(Y,A)\), \(g_2(Y,A)=\partial _A g(Y,A)\). By (2.1) and (2.2),

$$\begin{aligned} \partial _X f = A g_1(AXA,A) A,\qquad \partial _A f = XA g_1(AXA,A) + g_1(AXA,A) AX. \end{aligned}$$

It follows that

$$\begin{aligned} \,\text{ tr }\,(AXA\partial _A)f-2\,\text{ tr }\,(XAX\partial _X)f = \,\text{ tr }\,(AXAg_2(AXA,A)) = \,\text{ tr }\,(Y\partial _A) g(Y,A) \ \Big |_{Y=AXA}, \end{aligned}$$

as required.

Let us define

$$\begin{aligned} J=\Delta _X-\,\text{ tr }\,X,\qquad p(X,A)=\, \text{ etr }\,(-AX-A^{-1}) \end{aligned}$$

and the corresponding integral operator

$$\begin{aligned} (Pf)(X)=\int _\mathcal {P}p(X,A) f(X,A) \mu (dA). \end{aligned}$$

Then, on a suitable domain, the following intertwining relation holds:

$$\begin{aligned} J \circ P = P\circ M. \end{aligned}$$
(5.1)

Indeed, let us write \(p=p(X,A)\), \(f=f(X,A)\) and first note that

$$\begin{aligned} \Delta _X(pf)=f\Delta _Xp+p(\Delta _X-2\,\text{ tr }\,(XAX\partial _X))f. \end{aligned}$$

Now, using the fact that \(\Delta \) is self-adjoint with respect to \(\mu \), together with the identity

$$\begin{aligned} \Delta _X\, \text{ etr }\,(-AX)=\Delta _A\, \text{ etr }\,(-AX), \end{aligned}$$

we have

$$\begin{aligned} \int _\mathcal {P}f (\Delta _Xp) \mu (dA)= & {} \int _\mathcal {P}\, \text{ etr }\,(-A^{-1}) f\ [\Delta _A \, \text{ etr }\,(-AX)] \mu (dA)\\= & {} - \int _\mathcal {P}\,\text{ tr }\,(\vartheta _A\, \text{ etr }\,(-AX)\ \vartheta _A(\, \text{ etr }\,(-A^{-1}) f))\mu (dA) \\= & {} - \int _\mathcal {P}p \,\text{ tr }\,(-AX \vartheta _A-AXA^{-1})f \mu (dA) . \end{aligned}$$

It follows that \(J(Pf)(X)=P(Mf)(X)\), as required.

Note that the intertwining relation (5.1) implies

$$\begin{aligned} J \circ D = D \circ \Delta , \end{aligned}$$
(5.2)

where D is the linear operator defined, for suitable \(f:\mathcal {P}\rightarrow \mathbb {C}\) by

$$\begin{aligned} (Df)(X)=\int _\mathcal {P}f(AXA) \, \text{ etr }\,(-AX-A^{-1}) \mu (dA). \end{aligned}$$

The intertwining relation (5.2) is essentially equivalent to [42, Corollary 6].

Now suppose \(\varphi \) is a positive eigenfunction of \(\Delta \) with eigenvalue \(\lambda \) such that \(\beta =D\varphi <\infty \). Then it follows from (5.2) that \(\beta \) is a positive eigenfunction of J with eigenvalue \(\lambda \). Note that if we write \(\beta (X)=\varphi (X) B_\varphi (X)\), then this implies

$$\begin{aligned} (\Delta _X^{(\varphi )}-\,\text{ tr }\,X)B_\varphi (X)=0. \end{aligned}$$
(5.3)

This suggests that, for suitable \(\varphi \), the function \(B_\varphi \) admits a natural probabilistic interpretation, via the Feynman-Kac formula, and this is indeed the case.

Proposition 3

Let \(\varphi \) be a positive eigenfunction of \(\Delta \) such that \(D\varphi <\infty \), and the martingale problem associated with \(\Delta ^{(\varphi )}\) is well posed for any initial condition in \(\mathcal {P}\). Let Y be a Brownian motion in \(\mathcal {P}\) with drift \(\varphi \) started at X, and denote by \(\mathbb {E}_X\) the corresponding expectation. Assume that, for any \(X\in \mathcal {P}\),

$$\begin{aligned} Z=\int _0^\infty \,\text{ tr }\,Y_s\ ds<\infty \end{aligned}$$
(5.4)

almost surely, and define \(M_\varphi (X)=\mathbb {E}_X e^{-Z}\). Suppose that \(\lim _{X\rightarrow 0} M_\varphi (X) =1\) and

$$\begin{aligned} \qquad \lim _{X\rightarrow 0} B_\varphi (X) = C_\varphi , \end{aligned}$$
(5.5)

where \(C_\varphi >0\) is a constant. Then \(B_\varphi (X)=C_\varphi \ M_\varphi (X)\) and, moreover, \(B_\varphi \) is the unique bounded solution to (5.3) satisfying the boundary condition (5.5).

Proof

It follows from the Feynman-Kac formula that \(M_\varphi \) satisfies

$$\begin{aligned} (\Delta _X^{(\varphi )}-\,\text{ tr }\,X)M_\varphi (X)=0. \end{aligned}$$

To prove uniqueness, up to a constant factor, suppose U(X) is another bounded solution which vanishes as \(X\rightarrow 0\). Note that, by (5.4), it must hold that \(Y_t\rightarrow 0\) almost surely as \(t\rightarrow \infty \). Thus,

$$\begin{aligned} U(Y_t) \exp \left( -\int _0^t \,\text{ tr }\,Y_s\ ds\right) \end{aligned}$$

is a bounded martingale which converges to 0 almost surely, as \(t\rightarrow \infty \), hence must be identically zero almost surely, which implies \(U= 0\), as required. \(\square \)

If \(\varphi (X)=|X|^{-\nu /2}\), then \(B_\varphi =B_{-\nu }\) is the matrix K-Bessel function defined by (2.17). If we denote the eigenvalues of \(Y_t\) by \(\lambda _i(Y_t)\), arranged in decreasing order, then, as shown in [42], it holds almost surely that, for any initial condition \(X\in \mathcal {P}\),

$$\begin{aligned} \lim _{t\rightarrow \infty } \frac{1}{t} \log \lambda _i(Y_t) = - \nu + (n-2i+1)/2. \end{aligned}$$

In particular, if \(\nu >(n-1)/2\), then (5.4) holds. In this example, the process \(Y_t\) is \(GL(n,\mathbb {R})\)-invariant, so we may write

$$\begin{aligned} M_\varphi (X)=\mathbb {E}_X \exp \left( -\int _0^\infty \,\text{ tr }\,Y_s\ ds\right) = \mathbb {E}_I \exp \left( -\int _0^\infty \,\text{ tr }\,(XY_s)\ ds\right) , \end{aligned}$$

and it follows, using bounded convergence, that \(\lim _{X\rightarrow 0} M_\varphi (X) =1\). On the other hand, again using bounded convergence, we have

$$\begin{aligned} \lim _{X\rightarrow 0} B_{-\nu }(X) = \Gamma _n(\nu ). \end{aligned}$$

Putting this together and applying Proposition 3 yields the following conclusion, in agreement with [42, Theorem 2]. When \(n=1\), this is Dufresne’s identity [16].

Corollary 4

If Y is a Brownian motion in \(\mathcal {P}\) with drift \(-\nu /2\), started at the identity, then \(\int _0^\infty Y_s\ ds\) is inverse Wishart distributed with parameters I/2 and \(2\nu \).

More generally, suppose \(\varphi =h_s\), where \(s\in \mathbb {R}^n\). Define new variables \(r_i\) by

$$\begin{aligned} 2(s_i+\cdots +s_n)=2r_i+i-\frac{n+1}{2}. \end{aligned}$$

It is well known that the spherical function \(h_s\) is invariant under permutations of the \(r_i\), so we may assume that \(r_1>\cdots >r_n\). Then it may be shown [8], via a straightforward modification of the proof of the second part of Theorem 3.1 in [9], that

$$\begin{aligned} \lim _{t\rightarrow \infty } \frac{1}{2t} \log \lambda _i(t) = r_i, \end{aligned}$$

almost surely, for any initial condition. In particular, (5.4) holds if, and only if, \(r_1<0\). This condition also ensures that

$$\begin{aligned} \int _\mathcal {P}h_s(AXA)\, \text{ etr }\,(-A^{-1})\mu (dA)<\infty , \end{aligned}$$
(5.6)

for all \(X\in \mathcal {P}\). Then, by the uniqueness property of the spherical functions on \(\mathcal {P}\) with a given set of eigenvalues [47, Proposition 1.2.4],

$$\begin{aligned} \int _\mathcal {P}h_s(AXA)\, \text{ etr }\,(-A^{-1})\mu (dA)=c_s h_s(X), \end{aligned}$$
(5.7)

where

$$\begin{aligned} c_s=\int _\mathcal {P}h_s(A^2)\, \text{ etr }\,(-A^{-1})\mu (dA) . \end{aligned}$$
(5.8)

This implies that

$$\begin{aligned} B_s(X)=h_s(X)^{-1} \int _\mathcal {P}h_s(AXA) \, \text{ etr }\,(-AX-A^{-1}) \mu (dA) \end{aligned}$$
(5.9)

is bounded. Using the homogeneity property \(h_s(cX)=c^dh_s(X)\), where \(d=\sum _k ks_k\), we see that, for any fixed \(X\in \mathcal {P}\),

$$\begin{aligned} \lim _{c\rightarrow 0} \mathbb {E}_{cX} \exp \left( -\int _0^\infty \,\text{ tr }\,Y_s\ ds\right) = \lim _{c\rightarrow 0} \mathbb {E}_{X} \exp \left( - c\int _0^\infty \,\text{ tr }\,Y_s\ ds\right) = 1 \end{aligned}$$

and

$$\begin{aligned} \lim _{c\rightarrow 0} B_s(cX) = \lim _{c\rightarrow 0} h_s(X)^{-1} \int _\mathcal {P}h_s(AXA) \, \text{ etr }\,(-cAX-A^{-1}) \mu (dA) = c_s. \end{aligned}$$

Assuming that these limits extend to

$$\begin{aligned} \lim _{X\rightarrow 0} \mathbb {E}_X \exp \left( -\int _0^\infty \,\text{ tr }\,Y_s\ ds\right) =1,\qquad \lim _{X\rightarrow 0} B_s(X)=c_s, \end{aligned}$$
(5.10)

Proposition 3 would then imply that

$$\begin{aligned} \mathbb {E}_X \exp \left( -\int _0^\infty \,\text{ tr }\,Y_s\ ds\right) = c_s^{-1} B_s(X). \end{aligned}$$
(5.11)

Again using the homogeneity property of \(h_s\), this is equivalent to the identity:

$$\begin{aligned} \int _0^\infty \,\text{ tr }\,Y_s\ ds \,{\mathop {=}\limits ^{law}}\,\,\text{ tr }\,(AX), \end{aligned}$$
(5.12)

where A is distributed according to the probability measure

$$\begin{aligned} \nu _s(dA)=c_s^{-1} h_s(X)^{-1} h_s(AXA)\, \text{ etr }\,(-A^{-1})\mu (dA) . \end{aligned}$$

To make this claim rigorous, one would need to establish the existence of the limits in (5.10). We will not pursue this here.

We remark that, writing \(r=-\mu \), we may compute, for \(n=1,2,3\):

$$\begin{aligned} c_s=\prod _i\Gamma (2\mu _i)\prod _{i<j} B(1/2,\mu _i+\mu _j), \end{aligned}$$
(5.13)

where \(B(x,y)=\Gamma (x)\Gamma (y)/\Gamma (x+y)\) is the beta function. The analogue of this formula in the complex case is given by (10.10) below, which is valid for all n. It seems natural to expect (5.13) to be valid for all n also.

Returning to the general setting, let us define

$$\begin{aligned} (P_\varphi f)(X)= & {} \beta (X)^{-1} \int _\mathcal {P}\varphi (AXA) p(X,A) f(X,A) \mu (dA),\\ J^{(\beta )}= & {} \beta (X)^{-1}\circ (J-\lambda ) \circ \beta (X),\\ M^{(\varphi )}= & {} \varphi (Y)^{-1}\circ (M-\lambda ) \circ \varphi (Y)=\Delta ^{(\varphi )}_Y +\,\text{ tr }\,(Y\partial _A). \end{aligned}$$

As before, with the change of variables \(Y=AXA\), we can also write

$$\begin{aligned} M^{(\varphi )}=\Delta _X^{(\varphi )} -2\,\text{ tr }\,(XAX\partial _X)+\,\text{ tr }\,(AXA\partial _A). \end{aligned}$$

Then (5.1) extends to:

$$\begin{aligned} J^{(\beta )} \circ P_\varphi = P_\varphi \circ M^{(\varphi )}. \end{aligned}$$
(5.14)

This intertwining relation has a probabilistic interpretation, as follows. Let \(\rho \) be a probability measure on \(\mathcal {P}\) and define a probability measure on \(\mathcal {P}\times \mathcal {P}\) by

$$\begin{aligned} \sigma (dX,dA)= \rho (dX) \gamma _X(dA), \end{aligned}$$

where

$$\begin{aligned} \gamma _X(dA)=\beta (X)^{-1} \varphi (AXA) p(X,A) \mu (dA). \end{aligned}$$

Suppose that \(\varphi \) is such that the martingale problems associated with \((J^{(\beta )},\rho )\) and \((M^{(\varphi )},\sigma )\) are well-posed, and that \((X_t,A_t)\) is a diffusion process with infinitesimal generator \(M^{(\varphi )}\) and initial law \(\sigma \). Then we may apply Theorem 15 to conclude that, with respect to its own filtration, \(X_t\) is a diffusion with generator \(J^{(\beta )}\) and initial distribution \(\rho \); moreover, the conditional law of \(A_t\), given \(X_s,\ s\le t\), only depends on \(X_t\) and is given by \(\gamma _{X_t}(dA)\).

These conditions certainly hold when \(\varphi (X)=|X|^{\nu /2}\), for any \(\nu \in \mathbb {R}\), in which case we obtain the following generalisation of [42, Proposition 23]. Define \(\beta _\nu (X)=|X|^{\nu /2} B_\nu (X)\).

Theorem 5

Let \(Y_t,\ t\ge 0\) be a Brownian motion in \(\mathcal {P}\) with drift \(\nu /2\) started at I, and let \(A_t=\int _0^t Y_s ds\). Fix \(X_0\in \mathcal {P}\), choose \(\tilde{A}_0\) at random, independent of Y, according to the distribution \(\gamma _{X_0}(dA)\), and define

$$\begin{aligned} \tilde{Y}_t=\tilde{A}_0 X_0^{1/2} Y_t X_0^{1/2} \tilde{A}_0,\qquad \tilde{A}_t = \tilde{A}_0 + \int _0^t \tilde{Y}_s\ ds. \end{aligned}$$

Then \(X_t=\tilde{A}_t^{-1} \tilde{Y}_t \tilde{A}_t^{-1},\ t\ge 0\) is a diffusion in \(\mathcal {P}\) with infinitesimal generator

$$\begin{aligned} L_\nu =\Delta _X+2\,\text{ tr }\,(\vartheta _X\ln \beta _\nu (X)\ \vartheta _X), \end{aligned}$$

started at \(X_0\). In particular, as a degenerate case, the process \(A_t^{-1} Y_t A_t^{-1},\ t>0\) is a diffusion in \(\mathcal {P}\) with infinitesimal generator \(L_\nu \).

Proof

The relevant martingale problems are well posed, as shown in Appendix B (Example 5), so the first claim follows from Theorem 15, as outlined above. For the second, we can let \(X_0=mI\) and consider the limit as \(m\rightarrow \infty \). By Lemma 1, \(m^{1/2} \tilde{A}_0\rightarrow I\) in probability, as required. \(\square \)

The second statement was proved, under the condition \(2|\nu |>n-1\), by Rider and Valkó [42]. Related results in the complex setting have been obtained by Bougerol [7]. In the case \(n=1\), the above theorem is due to Matsumoto and Yor [30], see also Baudoin [2]. We note that, as observed in [42], the law of the process with generator \(L_\nu \) is invariant under a change of sign of the underlying drift \(\nu \), since \(\beta _\nu =\beta _{-\nu }\), cf. (2.18).

More generally, if \(\varphi \) is such that, as \(X^{-1}\rightarrow 0\), the measure \(\gamma _X(dA)\) is concentrated around \(AXA=I\), and the relevant martingale problems are well-posed, then the corresponding statement should hold: if \(Y_t\) is a Brownian motion in \(\mathcal {P}\) with drift \(\varphi \) and \(A_t=\int _0^t Y_s ds\), then \(A_t^{-1} Y_t A_t^{-1},\ t>0\) is a diffusion in \(\mathcal {P}\) with generator \(J^{(\beta )}\).

6 Two particles with one-sided interaction and a ‘reflecting wall’

Let \(\nu \in \mathbb {R}\), and define

$$\begin{aligned} R_Q=\Delta _Q^{(\nu /2)} +\,\text{ tr }\,\partial _Q,\qquad N=R_Q + \Delta _X^{(-\nu /2)} +2\,\text{ tr }\,(Q\partial _X). \end{aligned}$$

We first note that R is self-adjoint with respect to the measure

$$\begin{aligned} \pi (dQ)=|Q|^{\nu }\, \text{ etr }\,(-Q^{-1})\mu (dQ). \end{aligned}$$
(6.1)

Define

$$\begin{aligned} S_X = \Delta _X^{(-\nu /2)} -\,\text{ tr }\,X^{-1},\qquad C f(X)=\int _\mathcal {P}\, \text{ etr }\,(-QX^{-1}) f(X,Q) \pi (dQ). \end{aligned}$$

Then \(R_Q \, \text{ etr }\,(-QX^{-1}) = S_X \, \text{ etr }\,(-QX^{-1})\), which implies \(S \circ C = C \circ N\).

Now suppose \(\rho \) is a positive eigenfunction of R with eigenvalue \(\lambda \) such that

$$\begin{aligned} \gamma (X)=\int _\mathcal {P}\rho (Q) \, \text{ etr }\,(-QX^{-1}) \pi (dQ)<\infty . \end{aligned}$$

Then \(\gamma \) is a positive eigenfunction of S with eigenvalue \(\lambda \). Define

$$\begin{aligned} S^{(\gamma )}= & {} \gamma (X)^{-1}\circ (S-\lambda ) \circ \gamma (X),\\ R^{(\rho )}= & {} \rho (Q)^{-1}\circ (R-\lambda ) \circ \rho (Q),\qquad N^{(\rho )}=\rho (Q)^{-1}\circ (N-\lambda ) \circ \rho (Q),\\ (C_\rho f)(X)= & {} \gamma (X)^{-1} \int _\mathcal {P}\rho (Q) \, \text{ etr }\,(-QX^{-1}) f(X,Q) \pi (dQ). \end{aligned}$$

Then the above intertwining relation extends to

$$\begin{aligned} S^{(\gamma )} \circ C_\rho = C_\rho \circ N^{(\rho )}. \end{aligned}$$

For example, if \(\rho =1\), then \(\gamma (X)=B_{\nu }(X^{-1})\). We note that S is related to the J of the previous section, via \(S_X=J_Y^{(\varphi )}\), where \(Y=X^{-1}\) and \(\varphi (Y)=|Y|^{\nu /2}\).

These intertwining relations (and their probabilistic interpretations) may be viewed as non-commutative, ‘positive temperature’ versions of the following well known relation between reflecting Brownian motion and the three-dimensional Bessel process: for appropriate initial conditions, a Brownian motion, reflected off a [Brownian motion reflected at zero], is a three-dimensional Bessel process ([13, 1, Prop. 3.5]). In the above, note that \(\partial _Q=Q^{-1}\vartheta _Q\) and \(Q\partial _X=QX^{-1}\vartheta _X\), so we may interpret the Q-process as a Brownian motion in \(\mathcal {P}\) with ‘soft reflection off the identity’, and the X-process as a second Brownian motion in \(\mathcal {P}\) with ‘soft reflection off Q’.

7 Whittaker functions and related processes

7.1 Whittaker functions of several matrix arguments

For \(X=(X_1,\ldots ,X_N)\in \mathcal {P}^N\) and \(\nu \in \mathbb {C}\), we define

$$\begin{aligned} e_\nu (X) = \prod _{i=1}^N e_\nu (X_i) \end{aligned}$$

and, for \(X=(X_1,\ldots ,X_N)\in \mathcal {P}^N\) and \(\lambda \in \mathbb {C}^N\),

$$\begin{aligned} e_\lambda (X)=\prod _{i=1}^N e_{\lambda _i}(X_i). \end{aligned}$$

For \(N\ge 1\), we define the product measure

$$\begin{aligned} \mu _N(dX)=\mu (dX_1)\ldots \mu (dX_N). \end{aligned}$$

Let \(\mathcal {T}=\mathcal {P}\times \mathcal {P}^2\times \cdots \times \mathcal {P}^N\) and, for \(X\in \mathcal {P}^N\), denote by \(\mathcal {T}(X)\) the set of \(Y=(Y^1,\ldots ,Y^N)\in \mathcal {T}\) such that \(Y^N=X\). For \(Y\in \mathcal {T}\), define

$$\begin{aligned} \mathcal {F}(Y)=\sum _{1\le i\le m<N} \,\text{ tr }\,[Y^m_i(Y^{m+1}_i)^{-1}]+\,\text{ tr }\,[Y^{m+1}_{i+1}(Y^m_i)^{-1}]. \end{aligned}$$
(7.1)

For \(Y\in \mathcal {T}\) and \(\lambda \in \mathbb {C}^N\), define

$$\begin{aligned} e_\lambda (Y)=e_{\lambda _1}(Y^1) \prod _{2\le m\le N} e_{\lambda _m} (Y^m) \,e_{-\lambda _m}(Y^{m-1}). \end{aligned}$$

Proposition 6

Let \(X\in \mathcal {P}^N\) and \(\lambda \in \mathbb {C}^N\).

  1. (i)

    The following integral converges:

    $$\begin{aligned} \psi _\lambda (X)=\int _{\mathcal {T}(X)} e_\lambda (Y) e^{-\mathcal {F}(Y)} \prod _{1\le m<N} \mu _m(dY^m). \end{aligned}$$
  2. (ii)

    The integrand \(e_\lambda (Y) e^{-\mathcal {F}(Y)}\) vanishes as \(Y\rightarrow \partial \mathcal {T}(X)\).

The proof is given in Appendix C.3. We note that, when \(N=2\),

$$\begin{aligned} \psi _\lambda (X)=e_{\lambda _2}(X) K_n(\lambda _1-\lambda _2|\ X_1^{-1},X_2). \end{aligned}$$

The following properties are straightforward to verify from the definition.

Proposition 7

Let \(X\in \mathcal {P}^N\), \(a\in GL(n)\), \(\lambda \in \mathbb {C}^N\), \(\nu \in \mathbb {C}\) and write \(\lambda '_i=\lambda _i+\nu \) and \(X'=(X_N^{-1},\ldots ,X_1^{-1})\). Then

$$\begin{aligned}&\psi _\lambda (X_1[a],\ldots ,X_N [a])=|a^ta|^{\sum _i\lambda _i}\psi _\lambda (X),\\&\psi _{\lambda '}(X)=e_\nu (X)\psi _\lambda (X),\qquad \psi _\lambda (X)=\psi _{-\lambda }(X'). \end{aligned}$$

We also anticipate that \(\psi _\lambda (X)\) is symmetric in the parameters \(\lambda _1,\ldots ,\lambda _N\); in the case \(N=2\), this symmetry holds and follows from (2.18).

7.2 Interpretation as eigenfunctions

Consider the differential operator

$$\begin{aligned} H^{(N)}=\sum _{i=1}^N \Delta _{X_i} - V(X),\qquad V(X)=2 \sum _{i=1}^{N-1} \,\text{ tr }\,(X_i^{-1}X_{i+1}). \end{aligned}$$
(7.2)

This is a quantisation of the N-particle non-Abelian Toda lattice on \(\mathcal {P}\).

For \(\nu \in \mathbb {C}\) and \((X,Y)\in \mathcal {P}^N\times \mathcal {P}^{N-1}\) define

$$\begin{aligned} Q^{(N)}_{\nu }(X,Y) = e_{\nu }(X)\, e_{-\nu }(Y)\, \, \text{ etr }\,\left( - \sum _{i=1}^{N-1} (Y_i X_i^{-1}+X_{i+1}Y_i^{-1}) \right) . \end{aligned}$$
(7.3)

We identify \(Q^{(N)}_\nu \) with the integral operator defined, for appropriate f, by

$$\begin{aligned} Q^{(N)}_\nu f(X)=\int _{\mathcal {P}^{N-1}} Q^{(N)}_\nu (X,Y) \,f(Y) \,\mu _{N-1}(dY). \end{aligned}$$

Note that, for \(\lambda \in \mathbb {C}^N\),

$$\begin{aligned} \psi ^{(N)}_\lambda =Q^{(N)}_{\lambda _N} \psi ^{(N-1)}_{\lambda _1,\ldots ,\lambda _{N-1}}, \qquad \psi ^{(1)}_\lambda (X)=|X|^\lambda . \end{aligned}$$
(7.4)

It is straightforward to show that

$$\begin{aligned} (H^{(N)}_X-n\nu ^2) \; Q^{(N)}_\nu (X,Y)=H^{(N-1)}_Y \; Q^{(N)}_\nu (X,Y), \end{aligned}$$

with the convention \(H^{(1)}=\Delta \). It follows that, on a suitable domain,

$$\begin{aligned} (H^{(N)}-n\nu ^2) \circ Q^{(N)}_\nu = Q^{(N)}_\nu \circ H^{(N-1)} . \end{aligned}$$
(7.5)

For \(\lambda \in \mathbb {C}^N\), set

$$\begin{aligned} H_\lambda = H^{(N)} - \sum _{i=1}^N n\lambda _i^2. \end{aligned}$$
(7.6)

The intertwining relation (7.5) implies that, for any \(\lambda \in \mathbb {C}^N\),

$$\begin{aligned} H_\lambda \psi _\lambda = 0. \end{aligned}$$
(7.7)

The integral representation of Proposition 6 is a generalisation of the Givental-type formula obtained in [21] for the eigenfunctions of the quantum Toda lattice, also known as GL(n)-Whittaker functions. We remark that a slightly richer family of eigenfunctions of \(H^{(N)}\) can be obtained by taking \(\psi _\lambda ^{(1)}\) to be an arbitrary eigenfunction of \(\Delta \) in the recursive definition (7.4), provided the corresponding integrals converge.

7.3 Feynman–Kac interpretation

Define Brownian motion in \(\mathcal {P}^N\) with drift \(\lambda \in \mathbb {R}^N\) to be the diffusion process in \(\mathcal {P}^N\) with infinitesimal generator

$$\begin{aligned} \Delta _\lambda = \sum _{i=1}^N \Delta _{X_i}^{(\lambda _i)}. \end{aligned}$$

We begin with a lemma.

Lemma 8

let Y be a Brownian motion in \(\mathcal {P}^2\) with drift \(\lambda \) started at X, with \(\nu =\lambda _1-\lambda _2>(n-1)/2\). Then

$$\begin{aligned} \int _0^\infty \,\text{ tr }\,[Y_1(t)^{-1} Y_2(t)] dt \,{\mathop {=}\limits ^{law}}\,\,\text{ tr }\,(AW^{-1}) , \end{aligned}$$

where \(A=X_1^{-1/2} X_2 X_1^{-1/2}\) and W is a Wishart random matrix with parameters I and \(2\nu \).

This follows from the matrix Dufresne identity [42, Theorem 1] (Corollary 4 in the present paper), together with the fact that the eigenvalue process of \(Y_1(t)^{-1/2} Y_2(t) Y_1(t)^{-1/2}\) has the same law as that of a Brownian motion in \(\mathcal {P}\) with generator \(2\Delta -2\nu \, \,\text{ tr }\,\vartheta _{X}\), started at A. A proof of the latter claim is given in Appendix C.4.

Now let \(Y(t),\ t\ge 0\) be a Brownian motion in \(\mathcal {P}^N\) with drift \(\lambda \in \mathbb {R}^N\) started at X. Suppose that \(\lambda _i-\lambda _j>(n-1)/2\) for all \(i<j\), and define

$$\begin{aligned} \varphi _\lambda (X)= \mathbb {E}\exp \left( -2\sum _{i=1}^{N-1} \int _0^\infty \,\text{ tr }\,[Y_i(t)^{-1} Y_{i+1}(t)] dt \right) . \end{aligned}$$

Proposition 9

Suppose that \(\lambda _i-\lambda _j>(n-1)/2\) for all \(i<j\). Then

$$\begin{aligned} \psi _\lambda (X)=\prod _{i<j} \Gamma _n(\lambda _i-\lambda _j) \ e_\lambda (X) \varphi _\lambda (X). \end{aligned}$$

Moreover, under these hypotheses, \(\psi _\lambda (X)\) is the unique solution to (7.7), up to a constant factor, such that \(e_{-\lambda }(X) \psi _\lambda (X)\) is bounded.

Proof

Define

$$\begin{aligned} \varphi _\lambda (X)= \mathbb {E}\exp \left( -2\sum _{i=1}^{N-1} \int _0^\infty \,\text{ tr }\,[Y_i(t)^{-1} Y_{i+1}(t)] dt \right) . \end{aligned}$$

It follows from Lemma 8 that

$$\begin{aligned} \lim _{V(X)\rightarrow 0} \varphi _\lambda (X)= 1. \end{aligned}$$

By Feynman-Kac, \((\Delta _\lambda -V)\varphi _\lambda =0\), hence \(f_\lambda =e_\lambda \varphi _\lambda \) satisfies the eigenvalue equation (7.7). A standard martingale argument (as in the proof of 3) then shows that \(f_\lambda \) is the unique solution to (7.7) such that \(e_{-\lambda } f_\lambda \) is bounded and

$$\begin{aligned} \lim _{X\rightarrow +\infty } e_{-\lambda } (X)f_\lambda (X)= 1. \end{aligned}$$

It therefore suffices to show that \(e_{-\lambda } \psi _\lambda \) is bounded and

$$\begin{aligned} \lim _{V(X)\rightarrow 0} e_{-\lambda }(X) \psi _\lambda (X) = \prod _{i<j} \Gamma _n(\lambda _i-\lambda _j). \end{aligned}$$

We prove these statements by induction on N, using the recursion (7.4).

For \(N=1\), the claim holds since \(\psi _\lambda (X)=e_\lambda (X)\) in this case. For \(N\ge 2\), we have

$$\begin{aligned} \psi ^{(N)}_\lambda (X)=\int _{\mathcal {P}^{N-1}} Q^{(N)}_{\lambda _N}(X,Y) \psi ^{(N-1)}_{\lambda _1,\ldots ,\lambda _{N-1}}(Y) \mu _{N-1}(dY), \end{aligned}$$

where

$$\begin{aligned} Q^{(N)}_{\nu }(X,Y) = e_{\nu }(X)\, e_{-\nu }(Y)\, \, \text{ etr }\,\left( - \sum _{i=1}^{N-1} (Y_i X_i^{-1}+X_{i+1}Y_i^{-1}) \right) . \end{aligned}$$

Let us write

$$\begin{aligned} g^{(N)}_\lambda (X)=e_{-\lambda }(X) \psi ^{(N)}_\lambda (X) \end{aligned}$$

and

$$\begin{aligned} R_\lambda (Y)=e_{\lambda _1-\lambda _N,\ldots ,\lambda _{N-1}-\lambda _N}(Y)\, \, \text{ etr }\,\left( -\sum _{i=1}^{N-1} \left( Y_i + A_i Y_i^{-1}\right) \right) , \end{aligned}$$

where \(A_i=X_i^{-1/2} X_{i+1} X_i^{-1/2}\). Changing variables from \(Y_i\) to \(X_i^{-1/2} Y_i X_i^{-1/2} \), we can write

$$\begin{aligned} g^{(N)}_\lambda (X) =\int _{\mathcal {P}^{N-1}} R_\lambda (Y) g^{(N-1)}_{\lambda _1,\ldots ,\lambda _{N-1}}(Y') \mu _{N-1}(dY), \end{aligned}$$

where \(Y'_i=X_i^{1/2} Y_i X_i^{1/2}\). By induction on N, we see immediately that

$$\begin{aligned} g^{(N)}_\lambda (X)\le \prod _{i<j} \Gamma _r(\lambda _i-\lambda _j). \end{aligned}$$

Here we are using

$$\begin{aligned} R_\lambda (Y) \le e_{\lambda _1-\lambda _N,\ldots ,\lambda _{N-1}-\lambda _N}(Y)\, \, \text{ etr }\,\left( -\sum _{i=1}^{N-1} Y_i \right) , \end{aligned}$$

which implies

$$\begin{aligned} \int _{\mathcal {P}^{N-1}} R_\lambda (Y) \mu _{N-1}(dY) \le \prod _{i=1}^{N-1} \Gamma _n(\lambda _i-\lambda _N). \end{aligned}$$

Now observe that, for each \(Y\in \mathcal {P}^{N-1}\), if \(V(X)\rightarrow 0\) then \(V^{(N-1)}(Y')\rightarrow 0\). Thus the claim follows, again by induction, using the dominated convergence theorem. \(\square \)

In the scalar case \(n=1\), the above proposition is a special case of [3, Proposition 5.1], see also [36, Corollary 3].

7.4 Whittaker measures on \(\mathcal {P}^N\)

The following generalises an integral identity due to Stade [46]. The proof is straightforward, by induction, using (7.4). Denote by \(e_n\) the unit vector \((0,0,\ldots ,0,1)\) in \(\mathbb {C}^n\).

Proposition 10

Let \(s\in \mathbb {C}^n\), \(A\in \mathcal {P}\) and \(\lambda ,\nu \in \mathbb {C}^N\). Set \(a=\sum _i(\lambda _i+\nu _i)\). Assume that \(\mathfrak {R}(\lambda _i+\nu _j)>(n-1)/2\) for all i and j, and \(\mathfrak {R}(s_k+\cdots +s_n+a)>(k-1)/2\) for all k. Then

$$\begin{aligned}&\int _{\mathcal {P}^N} p_s(X_1) \, \text{ etr }\,\left( -A^{-1}X_1\right) \psi _\lambda (X)\psi _\nu (X) \mu _N(dX) \\&\quad = \frac{\Gamma _n(s+ae_n)}{\Gamma _n(a)} p_{s+ae_n}(A) \prod _{i,j} \Gamma _n(\lambda _i+\nu _j). \end{aligned}$$

In particular, if \(\lambda ,\nu \in \mathbb {R}^N\) satisfy \(\lambda _i+\nu _j>(n-1)/2\) for all ij, then

$$\begin{aligned} W_{\lambda ,\nu }(dX) =\prod _{i,j} \Gamma _n(\lambda _i+\nu _j)^{-1} \, \text{ etr }\,(-X^{-1}_N) \psi _{-\lambda }(X)\psi _{-\nu }(X) \mu _N(dX) \end{aligned}$$

is a probability measure on \(\mathcal {P}^N\) which generalises the Whittaker measures of [14, 38]. Here we have made the change of variables \(X'=(X_N^{-1},\ldots ,X_1^{-1})\), as in Proposition 7. We remark that, by Proposition 10, for all B such that \(I+B\in \mathcal {P}\), we have

$$\begin{aligned} \int _{\mathcal {P}^N} \, \text{ etr }\,(-BX_N^{-1}) W_{\lambda ,\nu }(dX) = |I+B|^{-a}. \end{aligned}$$

This implies that the \(N^{th}\) marginal of \(W_{\lambda ,\nu }\) is the inverse Wishart distribution with parameters \(\Sigma =I/2\) and \(p=2a\), as defined in Sect. 2.7.

7.5 Triangular processes

Consider the differential operator on \(\mathcal {T}\) defined for \(\lambda \in \mathbb {C}^N\) by

$$\begin{aligned} G_\lambda = \sum _{1\le i\le m\le N} \Delta _{Y^m_i}^{(\lambda _i)} + 2\,\text{ tr }\,[( \epsilon _{im} Y^{m-1}_i (Y^m_i)^{-1}- \epsilon _{1i} Y^m_i (Y^{m-1}_{i-1})^{-1}) \vartheta _{Y^m_i}] , \end{aligned}$$
(7.8)

where \(\epsilon _{ij}=1-\delta _{ij}\).

Define a kernel from \(\mathcal {P}^N\) to \(\mathcal {T}\) by

$$\begin{aligned} \Sigma _\lambda (X,dY)= \delta _X(dY^N) e_\lambda (Y) e^{-\mathcal {F}(Y)} \prod _{1\le m<N} \mu _m(dY^m) , \end{aligned}$$

and note that

$$\begin{aligned} \psi _\lambda (X) = \int _\mathcal {T}\Sigma _\lambda (X,dY) . \end{aligned}$$

Then the following intertwining relation holds:

$$\begin{aligned} H_\lambda \circ \Sigma _\lambda = \Sigma _\lambda \circ G_\lambda . \end{aligned}$$
(7.9)

This extends (7.5) and is readily verified by induction. Define

$$\begin{aligned} L_\lambda = \psi _\lambda (X)^{-1} \circ H_\lambda \circ \psi _\lambda (X). \end{aligned}$$
(7.10)

As shown in Appendix B (Example 6), for any \(\lambda \in \mathbb {R}^N\), the martingale problems associated with \(G_\lambda \) and \(L_\lambda \) are well posed, for any initial conditions. We thus deduce from the intertwining relation (7.9), and Theorem 15, the following theorem.

Theorem 11

Let \(\lambda \in \mathbb {R}^N\) and \(X\in \mathcal {P}^N\), and suppose that \(Y(t),t\ge 0\) is a diffusion in \(\mathcal {T}\) with generator \(G_\lambda \) and initial law

$$\begin{aligned} \sigma _\lambda (X,dY) = \psi _\lambda (X)^{-1} \Sigma _\lambda (X,dY). \end{aligned}$$

Then \(Y^N(t),\ t\ge 0\) is a diffusion in \(\mathcal {P}^N\) with generator \(L_\lambda \). Moreover, for each \(t\ge 0\), the conditional law of Y(t), given \(Y^N(s),\ s\le t\), is \(\sigma _\lambda (Y^N(t),dY)\).

Remark 12

Theorem 11 is a matrix generalisation of Proposition 9.1 in [34]. It is not clear whether Theorem 3.1 in [34] admits a similar generalisation, other than in the case \(N=2\), where it is given by the statement of Theorem 5.

We note that \(G_\lambda \) contains the autonomous

$$\begin{aligned} T_\lambda = \Delta _\lambda + 2\sum _{i=2}^N \,\text{ tr }\,[X_{i-1} \partial _{X_i}],\qquad T'_\lambda =\Delta _\lambda -2\sum _{i=2}^N \,\text{ tr }\,[X_i X_{i-1}^{-1} \vartheta _{X_i}]. \end{aligned}$$

These are special cases of the diffusions with one-sided interactions discussed in Sect. 3, see (3.5) and (3.6). It follows from Theorem 11 that, if \(X(t),\ t\ge 0\) is a diffusion in \(\mathcal {P}^N\) with generator \(T_\lambda \) (resp. \(T'_\lambda \)), with appropriate initial conditions, then \(X_N(t),\ t\ge 0\) is distributed as the first (resp. last) coordinate of a diffusion in \(\mathcal {P}^N\) with generator \(L_\lambda \). In the scalar case, this yields very precise information about the law of \(X_N(t)\), for special initial conditions, see for example [34, Corollary 4.1]. We note however that this application in the scalar case relies on the Plancherel theory for the quantum Toda lattice, currently unavailable for its non-Abelian generalisation (7.2). In the scalar case, for a particular (singular) initial condition, the random variable \(X_N(t)\) may be interpreted as the logarithmic partition function of the semi-discrete directed polymer in a Brownian environment introduced in [39]. Unfortunately this interpretation does not extend to the non-Abelian setting, other than in an intuitive sense, because the SDE’s do not admit an explicit solution as an integral in this case. Nevertheless, to understand the law of \(X_N(t)\) when \(n>1\), and also its asymptotic behaviour as \(N,t\rightarrow \infty \), seems to be an interesting topic for future research.

8 The non-Abelian Toda lattice

The non-Abelian Toda lattice is a Hamiltonian system which describes the evolution of a system of particles \(X_1,\ldots ,X_N\) in the space of invertible \(n\times n\) matrices. There is a standard version, introduced by A.M. Polyakov [10], which generalises the classical Toda lattice. There is also an indefinite version, in which the potential has the opposite sign, as considered by Popowicz [40, 41]. As in the scalar case [35, 37], it is the indefinite version which is relevant to our setting.

Writing \(A_i=X_{i+1} X_i^{-1} \) and \(B_i=\dot{X}_i X_i^{-1}\), the Hamiltonian is given by

$$\begin{aligned} H = \,\text{ tr }\,\left( \frac{1}{2}\sum _{i=1}^N B_i^2 - \sum _{i=1}^{N-1} A_i\right) , \end{aligned}$$

and the equations of motion are

$$\begin{aligned} \dot{B}_i = A_{i-1}-A_i ,\qquad \dot{A}_i=B_{i+1} A_i - A_i B_i . \end{aligned}$$
(8.1)

This system admits the Lax representation \(\dot{L}=[L,M]\), where

$$\begin{aligned} L=\begin{pmatrix} B_1 &{} \quad -1 &{} \quad 0 &{} \quad \cdots &{} \quad 0\\ A_1 &{} \quad B_2 &{} \quad -1 &{} \quad \cdots &{} \quad 0\\ 0 &{} \quad A_2 &{} \quad \smash {\ddots } &{} \quad \smash {\ddots } &{} \quad \smash {\vdots }\\ \smash {\vdots }&{} \quad &{} \quad &{} \quad B_{N-1}&{} \quad -1\\ 0&{} \quad \dots &{} \quad 0&{} \quad A_{N-1}&{} \quad B_N \end{pmatrix}, \qquad M=\begin{pmatrix} 0 &{} \quad 0 &{} \quad \dots &{} \quad 0\\ A_1 &{} \quad 0 &{} \quad \dots &{} \quad 0\\ \vdots &{} \quad \ddots &{} \quad \ddots &{} \quad \vdots \\ 0&{} \quad \dots &{} \quad A_{N-1}&{} \quad 0 \end{pmatrix}. \end{aligned}$$
(8.2)

Here we are using the notation \([L,M]=LM-ML\), where multiplication of matrices with matrix-valued entries is ordered, that is,

$$\begin{aligned} (LM)_{ij}=\sum _k L_{ik}M_{kj}. \end{aligned}$$

The Lax representation implies that, for each positive integer k, \(C_k=\sum _i (L^k)_{ii}\) is a constant of motion for the system, that is \(\dot{C}_k=0\). Note that these constants of motion are matrix-valued. In the following we will use the notation \(L=L^{(N)}(A,B)\) for the above Lax matrix, and \(C^{(N)}_k\) for the corresponding constants of motion.

The equations of motion (8.1) can be written, equivalently, as

$$\begin{aligned} \dot{X}_i=P_i,\qquad \dot{P}_i=P_iX_i^{-1}P_i+X_iX_{i-1}^{-1}X_i-X_{i+1}. \end{aligned}$$
(8.3)

In the scalar case, writing \(X_i=e^{x_i}\), these reduce to the (indefinite) Toda equations

$$\begin{aligned} \ddot{x}_i = e^{x_i-x_{i-1}}-e^{x_{i+1}-x_i}. \end{aligned}$$

Observe that the space \(\mathcal {P}^N\times \mathcal {S}^N\) is invariant under the evolution (8.3), where \(\mathcal {S}\) denotes the set of real symmetric \(n\times n\) matrices. The system therefore admits a natural quantisation in this space, with Hamiltonian given by (7.2).

The diffusion with generator \(G_\lambda \), defined by (7.8) with \(\lambda \in \mathbb {R}^N\), is in fact a stochastic version of a series of Bäcklund transformations between non-Abelian Toda systems with different numbers of particles, as we shall now explain.

Recall the kernel function \(Q^{(N)}_{\nu }(X,Y)\) defined by (7.3), and consider the following evolution in \(\mathcal {P}^N\times \mathcal {P}^{N-1}\):

$$\begin{aligned} B_i= & {} \dot{X}_i X_i^{-1} = \vartheta _{X_i} \ln Q^{(N)}_{\nu }(X,Y)=\nu +Y_i X_i^{-1}-X_i Y_{i-1}^{-1},\ 1\le i\le N,\nonumber \\ \end{aligned}$$
(8.4)
$$\begin{aligned} B_i'= & {} \dot{Y}_i Y_i^{-1} = - \vartheta _{Y_i} \ln Q^{(N)}_{\nu }(X,Y)=\nu +Y_i X_i^{-1}- X_{i+1}Y_i^{-1},\ 1\le i\le N-1.\nonumber \\ \end{aligned}$$
(8.5)

Here we are using the conventions \(X_{N+1}=Y_N=Y_0^{-1}=0\). If (8.4) and (8.5) hold, then it is straightforward to compute

$$\begin{aligned} \dot{B}_i=A_{i-1}-A_i,\ 1\le i\le N,\qquad \dot{B}_i'=A'_{i-1}-A'_i,\ 1\le i\le N-1, \end{aligned}$$
(8.6)

where \(A_i=X_{i+1} X_i^{-1} \), \(A_i'=Y_{i+1} Y_i^{-1}\), with conventions \(A_0=A_N=0\) and \(A'_0=A'_{N-1}=0\). As such, this defines a Bäcklund transformation between the N- and \((N-1)\)-particle systems. In the case \(\nu =0\), it can be seen as a degeneration of the auto-Bäcklund transformation for the semi-infinite non-Abelian Toda lattice described in the paper [41]. Moreover, the constants of motion are related by \(C^{(N)}_k=C^{(N-1)}_k+\nu ^k I\). This follows from the relation, readily verified from (8.4) and (8.5):

$$\begin{aligned} L^{(N)}(A,B)\, D(X,Y) = D(X,Y)\, \hat{L}^{(N-1)}(A',B',\nu ), \end{aligned}$$
(8.7)

where

$$\begin{aligned} \hat{L}^{(N-1)}(A',B',\nu )=\begin{pmatrix} &{}&{}&{}0\\ &{} L^{(N-1)}(A',B')&{}&{} \vdots \\ &{}&{}&{} 0\\ &{}&{}&{}-1\\ 0&{}\cdots &{}0&{}\nu I \end{pmatrix}, \end{aligned}$$

and

$$\begin{aligned} D(X,Y)=\begin{pmatrix} 1 &{} 0 &{} \cdots &{} 0&{}0\\ X_2 Y_1^{-1} &{} 1 &{} \cdots &{} 0&{}0\\ 0 &{} X_3 Y_2^{-1} &{} \smash {\ddots } &{} \smash {\vdots }&{} \smash {\vdots }\\ \smash {\vdots }&{}&{}&{}1&{}0\\ 0&{}\dots &{}0&{}X_N Y_{N-1}^{-1}&{}1 \end{pmatrix}. \end{aligned}$$

This Bäcklund transformation may be used to construct solutions for the N-particle system, recursively, as follows. Let \(\mathcal {T}\), \(\mathcal {T}(X)\), \(\mathcal {F}(Y)\) and \(e_\lambda (Y)\) be defined as in Sect.  7.1. Let \(\lambda \in \mathbb {R}^n\) and set \(\mathcal {F}_\lambda (Y)=\mathcal {F}(Y)-\ln e_\lambda (Y)\). By Proposition 6 (ii), \(\mathcal {F}_\lambda (Y)\rightarrow +\infty \) as \(Y\rightarrow \partial \mathcal {T}(X)\), hence there exists \(Y^*(\lambda ,X)\in \mathcal {T}(X)\) at which \(\mathcal {F}_\lambda (Y)\) achieves its minimum value on \(\mathcal {T}(X)\). Moreover, this minimiser must satisfy

$$\begin{aligned} \vartheta _{Y^m_i} \mathcal {F}_\lambda (Y) = 0,\qquad 1\le i\le m <N. \end{aligned}$$
(8.8)

Equivalently,

$$\begin{aligned} \lambda _m + Y^{m-1}_i (Y^m_i)^{-1} - Y^m_i (Y^{m-1}_{i-1})^{-1} = \lambda _{m+1} + Y^m_i (Y^{m+1}_i)^{-1} - Y^{m+1}_{i+1} (Y^{m}_{i})^{-1}, \end{aligned}$$

with the conventions \(Y^{m-1}_m=0\) and \((Y^{m-1}_{0})^{-1} =0\). The equations with \(2\le m\le N-1\) are equivalent to

$$\begin{aligned} \vartheta _{Y^m_i} \ln Q^{(m)}_{\lambda _m}(Y^m,Y^{m-1})=- \vartheta _{Y^m_i} \ln Q^{(m+1)}_{\lambda _{m+1}}(Y^{m+1},Y^m). \end{aligned}$$

Denote by \(\mathcal {T}_\lambda \) the set of \(Y\in \mathcal {T}\) which satisfy the critical point Eq. (8.8), and consider the evolution on \(\mathcal {T}\) defined by \(B^1_1=\dot{Y}^1_1 (Y^1_1)^{-1}=\lambda _1\) and, for \(2\le m\le N\),

$$\begin{aligned} B^m_i = \dot{Y}^m_i (Y^m_i)^{-1} = \vartheta _{Y^m_i} \ln Q^{(m)}_{\lambda _m}(Y^m,Y^{m-1}),\qquad 1\le i\le m\le N. \end{aligned}$$
(8.9)

Note that this is equivalent to

$$\begin{aligned} B^m_i = \lambda _m + Y^{m-1}_i (Y^m_i)^{-1} - Y^m_i (Y^{m-1}_{i-1})^{-1} ,\qquad 1\le i\le m\le N, \end{aligned}$$
(8.10)

with the conventions, as above, \(Y^{m-1}_m=0\) and \((Y^{m-1}_{0})^{-1} =0\). It corresponds precisely to the drift term of the diffusion in \(\mathcal {T}\) with infinitesimal generator \(G_\lambda /2\).

It follows from (8.6) that \(\mathcal {T}_\lambda \) is invariant under the evolution (8.9), as in the scalar (\(n=1\)) case [35, Proposition 8.1]. Thus, if \(Y(0)=Y^*(\lambda ,X)\) and we let Y(t) evolve according to (8.9), then \(Y^N(t),\ t\ge 0\) is a realisation of the N-particle non-Abelian Toda flow on \(\mathcal {P}^N\), with \(Y^N(0)=X\) and \(C^{(N)}_k=\sum _i \lambda _i^k I\). In the scalar case, this agrees with the statement of [35, Theorem 8.4].

To conclude, Theorem 11 says that if we add noise to the evolution (8.9), and choose the initial law on \(\mathcal {T}\) appropriately, then \(Y^N(t),\ t\ge 0\) evolves as a diffusion in \(\mathcal {P}^N\) with generator given by a Doob transform of the quantised Hamiltonian (7.2).

9 A related class of processes

Let G be a right-invariant Brownian motion in GL(n), satisfying \(\partial G = \partial \beta \, G\), where \(\beta \) is a Brownian motion in \(\mathfrak {gl}(n,\mathbb {R})\) with each entry having infinitesimal variance 1/2. Then \(Y=G^tG\) is a Brownian motion in \(\mathcal {P}\). As the evolution of Y is governed by \(\Delta \), its law is invariant under the action of GL(n) on \(\mathcal {P}\).

Norris, Rogers and Williams [33] consider the closely related Markov process \(X=GG^t\). This process has the same eigenvalues as Y, but its eigenvectors behave quite differently. It satisfies \(\partial X = \partial \beta \, X+X\, \partial \beta ^t\), from which the Markov property is evident, and one may compute its infinitesimal generator:

$$\begin{aligned} \Omega _X=\frac{1}{2}\,\text{ tr }\,(\vartheta _X\vartheta _X'+\vartheta _X'\vartheta _X) =\,\text{ tr }\,(X^2\partial _X^2)+\frac{1}{2}\,\text{ tr }\,\vartheta _X + \frac{1}{2} \,\text{ tr }\,(X)\,\text{ tr }\,\partial _X. \end{aligned}$$

Here, \(\vartheta _X' f = (\partial _X f)X\). The differential operator \(\Omega \) is easily seen to be O(n)-invariant, but not GL(n)-invariant. Nevertheless, it bears many similarities to the Laplacian. For example, it agrees with the Laplacian when applied to radial functions on \(\mathcal {P}\), as can be seen by noting that \(\vartheta _X' \,\text{ tr }\,X^k = \vartheta _X \,\text{ tr }\,X^k\), for all positive integers k (cf. (2.9)). It is self-adjoint with respect to \(\mu \), and invariant under the change of variables \(Y=X^{-1}\). If \(k(X,Y)=\, \text{ etr }\,(-YX^{-1})\), then \(\Omega _Xk=\Omega _Y k\). An important difference is the associated product rule, cf. (2.72.8):

$$\begin{aligned} \Omega (fg)=(\Omega f)g+f (\Omega g)+\,\text{ tr }\,[X^2(\partial _Xf )(\partial _X g)]+\,\text{ tr }\,[X^2(\partial _Xg)( \partial _X f)]. \end{aligned}$$
(9.1)

Many of the diffusions we have considered have natural analogues in which the underlying motion of particles is governed by \(\Omega \) rather than \(\Delta \). Let G be a right-invariant Brownian motion in GL(n) with drift \(\nu /2\), started at I. Then \(Y=G^t G\) is a Brownian motion in \(\mathcal {P}\) with drift \(\nu /2\), and \(X=GG^t\) is a diffusion in \(\mathcal {P}\) with generator

$$\begin{aligned} \Omega _X^{(\nu /2)}=\Omega _X+\nu \,\text{ tr }\,\vartheta _X. \end{aligned}$$

We note that this is the Doob transform of \(\Omega _X\) via the positive eigenfunction \(|X|^{\nu /2}\).

Consider the process in \(\mathcal {P}\) defined by \(Q = G (Q_0+A) G^{t}\), where \(A_t=\int _0^t Y_s^{-1} ds\) and \(Q_0\) is independent of G. This is a diffusion in \(\mathcal {P}\) with infinitesimal generator

$$\begin{aligned} \mathcal {R}=\Omega _X^{(\nu /2)}+\,\text{ tr }\,\partial _Q. \end{aligned}$$

This process (with a different normalisation) was studied in [42], where it was observed that \(\mathcal {R}\) is self-adjoint with respect to the measure (6.1), as in the case of a Brownian particle.

For several particles with one-sided interactions, the interactions need to be modified on account of the product rule (9.1). For example, we may consider

$$\begin{aligned} \mathcal {T}= \Omega ^{(\lambda )}_Y+\Omega _X +\,\text{ tr }\,[(XYX^{-1}+X^{-1}YX)\partial _X]. \end{aligned}$$

Assume \(2\lambda >n-1\). Then \(\Omega ^{(\lambda )} \circ K_\lambda = K_\lambda \circ \mathcal {T}\), where \((K_\lambda f)(X)= \int _\mathcal {P}f(X,Y) \Pi _\lambda (X,dY)\) and \(\Pi _\lambda \) is defined by (3.4). Assuming the associated martingale problem is well posed, this yields the following analogue of the ‘Burke’ Theorem 2: if \((X_t,Y_t)\) be a diffusion in \(\mathcal {P}^2\) with generator \(\mathcal {T}\) and initial law \(\delta _{X_0}(dX) \Pi _{\lambda }(X,dY)\) then, with respect to its own filtration, \(X_t\) is a diffusion in \(\mathcal {P}\) with generator \(\Omega ^{(\lambda )}_X\), started at \(X_0\); moreover, the conditional law of \(Y_t\), given \(X_s,\ s\le t\), only depends on \(X_t\) and is given by \( \Pi _{\lambda }(X_t,dY)\).

10 The complex case

Most of the discussion in this paper carries over naturally to the complex setting. We remark in particular that the matrix Dufresne identity and related \(2M-X\) theorem are studied in some detail in the complex setting by Rider and Valkó [42] and Bougerol [7]. In this section, we briefly outline how the framework developed in this paper may be extended to the complex setting, with particular emphasis on a complex version of the intertwining relation (5.2) and some of its consequences. We will also briefly indicate how this relates to a remarkable identity of Fitzgerald and Warren [19] concerning Brownian motion in a Weyl chamber, and closely related work of Nguyen and Remenik [31] and Liechty, Nguyen and Remenik [28] on non-intersecting Brownian bridges.

In this section, \(\mathcal {P}\) will denote the space of positive \(n\times n\) Hermitian matrices. For \(a\in GL(n,\mathbb {C})\) and \(X\in \mathcal {P}\), write \(X[a]=a^\dagger X a\). This defines an action of \(GL(n,\mathbb {C})\) on \(\mathcal {P}\). The \(GL(n,\mathbb {C})\)-invariant volume element on \(\mathcal {P}\) is given by \(\mu (dX)=|X|^{-n} dX\), where dX denotes the Lebesgue measure on \(\mathcal {P}\). We note that, writing \(A=a[k]\), where \(k\in U(n)\) and a is the diagonal matrix with entries given by the eigenvalues \(z_1,\ldots ,z_n\) of A, with \(z\in C_+=\{z_1>\cdots>z_n>0\}\), we have the decomposition, on \(U(n)\times \mathbb {R}_+^n\),

$$\begin{aligned} \mu (dA) = \pi ^{n(n-1)/2} \left( \prod _{j=1}^{n-1} j!\right) ^{-1} dk \prod _{i<j}(z_i-z_j)^2 \prod _i z_i^{-n} dz_i. \end{aligned}$$
(10.1)

The Laplacian on \(\mathcal {P}\) may be characterised (see, for example, [32]) as the unique invariant differential operator on \(\mathcal {P}\) which satisfies

$$\begin{aligned} \Delta _X \, \text{ etr }\,(X) = (n\,\text{ tr }\,X + \,\text{ tr }\,X^2) \, \text{ etr }\,(X). \end{aligned}$$
(10.2)

Let us define a ‘partial matrix derivative’ \(\partial _X\) on \(\mathcal {P}\), writing \(X=(x_{ij})\), by

$$\begin{aligned} (\partial _X)_{ij} = {\left\{ \begin{array}{ll} \frac{\partial }{\partial x_{ii}} &{} i=j\\ \frac{\partial }{\partial \bar{x}_{ij}} &{} i\ne j\end{array}\right. } \end{aligned}$$

where

$$\begin{aligned} \frac{\partial }{\partial \bar{z}} = \frac{1}{2}\left( \frac{\partial }{\partial \mathfrak {R}z}+\sqrt{-1} \frac{\partial }{\partial \mathfrak {I}z}\right) . \end{aligned}$$

If \(Y=X[a]\) for some fixed \(a\in GL(n)\), then \( \partial _X=a\, \partial _Y \, a^\dagger \). This implies that \(\,\text{ tr }\,\vartheta _X^2\) is \(GL(n,\mathbb {C})\)-invariant, where \(\vartheta _X = X \partial _X\). One can check that \( \partial _X \,\text{ tr }\,(X)=I\) and \(\partial _X X=n I\). This implies that \(\,\text{ tr }\,\vartheta _X^2\) satisfies (10.2) and hence \(\Delta _X = \,\text{ tr }\,\vartheta _X^2\), as in the real case.

In this setting, the spherical functions are defined as follows. For \(\lambda ,x\in \mathbb {C}^n\), define \(a_\lambda (x)=\det (x_i^{\lambda _j})\) and \(s_\lambda (x)=a_{\delta +\lambda }(x)/a_\delta (x)\), where \(\delta _i=n-i\). For \(X\in \mathcal {P}\), define \(s_\lambda (X)\) to be the function \(s_\lambda \) evaluated at the eigenvalues of X.

We define Brownian motion in \(\mathcal {P}\) to be the diffusion process with generator \(\Delta /2\). If \(\varphi \) is a positive eigenfunction of the Laplacian with eigenvalue \(\gamma \), we may consider the corresponding Doob transform \(\Delta _X^{(\varphi )}/2\), and call the diffusion process with this generator a Brownian motion in \(\mathcal {P}\) with drift \(\varphi \). In particular, for \(\lambda \in \mathbb {R}^n\), we may consider the positive eigenfunction \(s_\lambda \), where \(\lambda \) is such that \(\rho +\lambda \in \bar{C}\), where \(\rho _i=(n-2i+1)/2\) and \(C=\{x\in \mathbb {R}^n:\ x_1>\cdots >x_n\}\). Then, as is well known, the logarithmic eigenvalues of a Brownian motion on \(\mathcal {P}\) with drift \(s_\lambda \) evolve as a standard Brownian motion in \(\mathbb {R}^n\) with drift \(\rho +\lambda \) conditioned never to exit C.

With these ingredients in place, all of the basic calculus and intertwining relations discussed previously carry over, with the obvious modifications. We will briefly illustrate this here in the context of the intertwining relation (5.2) and consider some of its consequences.

10.1 Matrix Dufresne identity

As in the real case, it is known [11, 29] that for each \(X,Y\in \mathcal {P}\), the equation \(Y=AXA\) has a unique solution in \(\mathcal {P}\), namely

$$\begin{aligned} A=X^{-1/2}(X^{1/2}YX^{1/2})^{1/2} X^{-1/2}. \end{aligned}$$

Moreover, if X is fixed, then

$$\begin{aligned} \partial _A=XA\, \partial _Y+\partial _Y AX. \end{aligned}$$
(10.3)

Let \(M=\Delta _Y +\,\text{ tr }\,(Y\partial _A)\). If \(Y=AXA\) then, in the variables (XA), we can write

$$\begin{aligned} M=\Delta _X -2\,\text{ tr }\,(XAX\partial _X)+\,\text{ tr }\,(AXA\partial _A). \end{aligned}$$

This follows from (10.3), as in the real case.

Let us define

$$\begin{aligned} J=\Delta _X-\,\text{ tr }\,X,\qquad p(X,A)=\, \text{ etr }\,(-AX-A^{-1}) \end{aligned}$$

and the corresponding integral operator

$$\begin{aligned} (Pf)(X)=\int _\mathcal {P}p(X,A) f(X,A) \mu (dA). \end{aligned}$$

Then, on a suitable domain, the following intertwining relation holds:

$$\begin{aligned} J \circ P = P\circ M. \end{aligned}$$
(10.4)

The proof is identical to the real case.

Note that the intertwining relation (10.4) implies

$$\begin{aligned} J \circ D = D \circ \Delta , \end{aligned}$$
(10.5)

where D is the linear operator defined, for suitable \(f:\mathcal {P}\rightarrow \mathbb {C}\) by

$$\begin{aligned} (Df)(X)=\int _\mathcal {P}f(AXA) \, \text{ etr }\,(-AX-A^{-1}) \mu (dA). \end{aligned}$$

Now suppose \(\varphi \) is a positive eigenfunction of \(\Delta \) with eigenvalue \(\gamma \) such that \(\beta =D\varphi <\infty \). Then it follows from (10.5) that \(\beta \) is a positive eigenfunction of J with eigenvalue \(\gamma \). Note that if we write \(\beta (X)=\varphi (X) B_\varphi (X)\), then this implies

$$\begin{aligned} (\Delta _X^{(\varphi )}-\,\text{ tr }\,X)B_\varphi (X)=0. \end{aligned}$$
(10.6)

As the real case, for suitable \(\varphi \), the function \(B_\varphi \) admits a natural probabilistic interpretation, via the Feynman-Kac formula.

Proposition 13

Let \(\varphi \) be a positive eigenfunction of \(\Delta \) such that \(D\varphi <\infty \), and the martingale problem associated with \(\Delta ^{(\varphi )}\) is well posed for any initial condition in \(\mathcal {P}\). Let Y be a Brownian motion in \(\mathcal {P}\) with drift \(\varphi \) started at X, and denote by \(\mathbb {E}_X\) the corresponding expectation. Assume that, for any \(X\in \mathcal {P}\),

$$\begin{aligned} Z=\frac{1}{2} \int _0^\infty \,\text{ tr }\,Y_s\ ds<\infty \end{aligned}$$
(10.7)

almost surely, and define \(M_\varphi (X)=\mathbb {E}_X e^{-Z}\). Suppose also that \(\lim _{X\rightarrow 0} M_\varphi (X) =1\) and

$$\begin{aligned} \qquad \lim _{X\rightarrow 0} B_\varphi (X) = C_\varphi , \end{aligned}$$
(10.8)

where \(C_\varphi >0\) is a constant. Then \(B_\varphi (X)=C_\varphi \ M_\varphi (X)\) and, moreover, \(B_\varphi \) is the unique bounded solution to (10.6) satisfying the boundary condition (10.8).

Again the proof is identical to the real case. We obtain the following corollary, in agreement with [42, Corollary 10].

Corollary 14

If \(\varphi (X)=|X|^{-\nu /2}\), where \(\nu >n-1\), then Z is inverse complex Wishart distributed with density proportional to

$$\begin{aligned} |A|^{-\nu } \, \text{ etr }\,(-A^{-1}) \mu (dA). \end{aligned}$$

Let \(\lambda \in \mathbb {R}^n\). It follows from the above that \(\beta _\lambda :=Ds_\lambda \), which is easily seen to be finite, is a eigenfunction of J. Thus, setting \(B_\lambda (X)=s_\lambda (X)^{-1}\beta _\lambda (X)\), we have

$$\begin{aligned} (\Delta _X^{(s_\lambda )}-\,\text{ tr }\,X) B_\lambda (X)=0. \end{aligned}$$
(10.9)

Let us now assume that \(\lambda \in C\), with \(2\lambda _1<1-n\). Then

$$\begin{aligned} c_\lambda = s_\lambda (I)^{-1} \int _\mathcal {P}s_\lambda (A^2) \, \text{ etr }\,(-A^{-1}) \mu (dA)<\infty , \end{aligned}$$

and, by uniqueness of the spherical functions,

$$\begin{aligned} \int _\mathcal {P}s_\lambda (AXA) \, \text{ etr }\,(-A^{-1}) \mu (dA)=c_\lambda s_\lambda (X). \end{aligned}$$

Recall the representation (10.1) and note that, writing \(\mu =-\lambda -\rho \),

$$\begin{aligned} s_{\lambda }(A^2) = a_{-\mu }(A^2)/a_\rho (A^2) = \det \left( z_i^{-2\mu _j} \right) \prod _{i<j}(z_i^2-z_j^2)^{-1} \prod _i z_i^{n-1}, \end{aligned}$$

where \(z_1,\ldots ,z_n\) denote the eigenvalues of A. Thus, using Schur’s Pfaffian identity and de Bruijn’s formula [15], we may compute

$$\begin{aligned} c_\lambda = \prod _i \Gamma (2\mu _i) \, \prod _{i<j}\frac{\pi }{\mu _i+\mu _j}. \end{aligned}$$
(10.10)

Let Y be a Brownian motion in \(\mathcal {P}\) with drift \(s_\lambda \), started at X, and denote by \(\mathbb {E}_X\) the corresponding expectation. The condition \(2\lambda _1<1-n\) ensures that \(Z<\infty \), almost surely. As in the real case, using the fact that the law of Y is invariant under multiplication by scalars, it is easy to see that \(\mathbb {E}_X e^{-Z} \rightarrow 1\) and \(B_\lambda (X)\rightarrow c_\lambda \) as \(X\rightarrow 0\) along any ray in \(\mathcal {P}\). To prove that these limits exist as \(X\rightarrow 0\) in \(\mathcal {P}\) is somewhat technical, and we will not pursue it here, but rather state it as a hypothesis:

Hypothesis 1

As \(X\rightarrow 0\) in \(\mathcal {P}\), \(\mathbb {E}_X e^{-Z} \rightarrow 1\) and \(B_\lambda (X)\rightarrow c_\lambda \).

Assuming Hypothesis 1, it follows from Proposition 13 that

$$\begin{aligned} B_\lambda (X) = c_\lambda \ \mathbb {E}e^{-Z}. \end{aligned}$$
(10.11)

Again using the fact that the law of Y is invariant under multiplication by scalars, the identity (10.11) is equivalent to the statement that Z has the same law as \(\,\text{ tr }\,(AX)\), where A is distributed according to the probability measure

$$\begin{aligned} c_\lambda ^{-1} \, s_\lambda (X)^{-1} \, s_\lambda (AXA)\, \, \text{ etr }\,(-A^{-1})\, \mu (dA). \end{aligned}$$

As this is a statement about the eigenvalues of Y, we may rephrase it in terms of a Brownian motion in C, as follows. Let \(-\mu ,x\in C\) and assume that \(\mu _1>0\). Let \(\xi (t)\) be a Brownian motion with drift \(-\mu \), started at x and conditioned never to exit C. Let

$$\begin{aligned} Z=\frac{1}{2} \sum _{i=1}^n \int _0^\infty e^{\xi _i(s)} ds. \end{aligned}$$

Then, assuming Hypothesis 1, Z has the same law as \(\,\text{ tr }\,(AX)\), where X has eigenvalues \(e^{x_1},\ldots ,e^{x_n}\) and A is distributed according to

$$\begin{aligned} c_{-\mu -\rho }^{-1}\, s_{-\mu -\rho }(X)^{-1} \, s_{-\mu -\rho }(AXA)\, \, \text{ etr }\,(-A^{-1})\, \mu (dA) . \end{aligned}$$

If \(\xi (t)\) is started at the origin, then Z has the same law as \(\sum _i z_i\), where z has density

$$\begin{aligned} b_\mu ^{-1} \det \left( z_i^{-2\mu _j} \right) \, \prod _{i<j}\frac{z_i-z_j}{z_i+z_j} \,\prod _i e^{-1/z_i} \frac{dz_i}{z_i} , \end{aligned}$$
(10.12)

on \(C_+\). The normalisation constant \(b_\mu \) is given by

$$\begin{aligned} b_\mu = \prod _i \Gamma (2\mu _i) \, \prod _{i<j} \frac{\mu _j-\mu _i}{\mu _i+\mu _j}. \end{aligned}$$
(10.13)

The measure (10.12) is a generalised Bures measure and defines a Pfaffian point process related to the BKP hierarchy [48, Proposition 4.4].

To conclude, consider the function

$$\begin{aligned} \tilde{B}_\lambda (X) = \prod _{i<j}\frac{\pi }{\mu _i+\mu _j} \, \det (e^{-\mu _i x_j})^{-1}\, \det \left( 2 K_{2\mu _i}(2e^{x_j/2})\right) , \end{aligned}$$

where \(K_\nu \) is the Macdonald function, and \(e^{x_1},\ldots ,e^{x_n}\) are the eigenvalues of X. It can be shown that \(\tilde{B}_\lambda (X)\) satisfies (10.9) and \(\lim _{X\rightarrow 0}B_\lambda (X)=c_\lambda \), so if Hypothesis 1 holds then \(B_\lambda (X) =\tilde{B}_\lambda (X)\).

10.2 Maximum of Dyson Brownian motion with negative drift

Let us write \(2\mu =\epsilon \alpha \), where \(-\alpha \in C\), \(\alpha _1>0\) and \(\epsilon >0\). Changing variables to \(y_i=\epsilon \ln z_i\), the density (10.12) becomes, up to a constant factor,

$$\begin{aligned} \det (e^{-\alpha _i y_j}) \prod _{i<j}\tanh \left( \frac{y_i-y_j}{2\epsilon }\right) \prod _i \exp (-e^{-y_i/\epsilon }) dy_i . \end{aligned}$$

In the limit as \(\epsilon \rightarrow 0\), this reduces to

$$\begin{aligned} \det (e^{-\alpha _i y_j}) \prod _i dy_i , \end{aligned}$$
(10.14)

on the set \(C_+=\{y_1>\cdots>y_n>0\}\). Note that, in this scaling limit,

$$\begin{aligned} \epsilon \ln \sum _i e^{y_i/\epsilon } \sim y_1. \end{aligned}$$

If \(\alpha _i=\alpha \), for all i, the measure (10.14) reduces to a special case of the Laguerre Orthogonal Ensemble (LOE) of random matrix theory.

On the other hand, by Brownian scaling and the method of Laplace, as \(\epsilon \rightarrow 0\) the random variable \(\epsilon \ln Z\) converges in law to \(\sup _t \eta _1(t)\), where \(\eta (t)\) is a Brownian motion in \(\mathbb {R}^n\) with drift \(-\alpha /2\), conditioned to stay in C, and started at the origin. Putting these observations together we conclude (modulo technicalities) that \(\sup _t \eta _1(t)\) has the same law as the first coordinate of the ensemble in \(C_+\) with density given by (10.14), in agreement with [19, Theorem 1].

We note that one may compute the law of \(\sup _t \eta _1(t)\) directly, for a general initial condition, as follows. Consider the process \(\xi (t)\), a BM with drift \(\nu \in C_+\), started at \(y\in C_+\) and conditioned never to exit C. Let T be the first exit time of \(\xi \) from \(C_+\). Then, using formulas from [4] for the exit probabilities of a Brownian motion with drift from C and \(C_+\),

$$\begin{aligned} f(\nu ,y):=P(T=\infty )=\det (e^{\nu _i y_j}-e^{-\nu _i y_j})/\det (e^{\nu _i y_j}). \end{aligned}$$

This implies that, if \(\eta (t)\) is a Brownian motion in \(\mathbb {R}^n\) with drift \(-\alpha /2\), conditioned to stay in C, and started at \(x\in C\), and \(M=\sup _t \eta _1(t)\), then, for \(\alpha _1>0\) and \(z\ge x_1\), we have the formula \(P(M\le z)=f(\nu ,y)\), where \(\nu _i=\alpha _{n-i+1}/2\) and \(y_i=z-x_{n-i+1}\).

We remark that f(vy) may also be interpreted as the probability that a collection of non-intersecting Brownian bridges of unit length, started at positions \(\nu _1,\ldots ,\nu _n\) and ending at positions \(y_1,\ldots ,y_n\), do not exit the domain \(C_+\). This follows from the reflection principles associated with C and \(C_+\), and gives a formula for the distribution function of the maximum of the ‘top’ bridge, namely \(F(r)=f(\nu ',y')\), where \(\nu _i'=r-\nu _i\) and \(y_i'=r-y_i\). In this context, the connection to the LOE was first discovered by Nguyen and Remenik [31] for bridges starting and ending at the origin, and for general starting and ending positions by Liechty, Nguyen and Remenik [28], where further asymptotic results are obtained in terms of Painlevé II and the KdV equation.