This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Matrix-valued Bratu equation associated with the symmetric domains of type BDI and CI

Hiroto Inoue Institute of Mathematics for Industry, Kyushu University
Abstract

A formulation of the exponential matrix solution of the matrix-valued Bratu equation is given, based on the structure of the symmetric domain of type BDI. Moreover, an analog for the symmetric domain of type CI is given.

Notation

Throughout this article, Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R}) is the set of nn-dimensional positive-symmetric matrices and Mn,r()\mathrm{M}_{n,r}(\mathbb{R}) is the set of (n×r)(n\times r) real matrices.

1 Introduction

We consider the matrix-valued Bratu equation

(h(s)h(s)1)=(aat)h(s)1,aMn,r()(h^{\prime}(s)h(s)^{-1})^{\prime}=(a{}^{t}\!a)h(s)^{-1},\quad a\in\mathrm{M}_{n,r}(\mathbb{R}) (1.1)

for a Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R})-valued function h(s)h(s) of ss\in\mathbb{R} introduced in [1]. The solution of the initial value problem of the Bratu equation (1.1) was given in [1] using a specific exponential matrix.

In this article, we give a geometrical formulation of this solution based on the geometrical structure of the symmetric domains G/KG/K. In detail, we express the exponential matrix solution of the equation (1.1) by the defining function of the symmetric domain of type BDI according to Cartan’s classification, ref. [2, 3]. Moreover, we give an analog associated with the symmetric domain of type CI, as Table 1.

Table 1: Matrix-valued differential equations associated to symmetric domains

domaincondition forΔ(w,esBw)equation ofΔ(esBu0,esBu0)1BDI(aat)s22+O(s3)(hh1)=(aat)h1aMn,r()CINone(hh1)=(ch1)2cSymn()\begin{array}[]{c||c|c}\hline\cr\hline\cr\text{domain}&\begin{array}[]{l}\text{condition for}\\ \text{$\Delta(w,e^{sB}w)$}\end{array}&\begin{array}[]{l}\text{equation of}\\ \Delta(e^{sB}u_{0},e^{sB}u_{0})^{-1}\end{array}\\ \hline\cr\text{BDI}&\displaystyle(a{}^{t}\!a)\frac{s^{2}}{2}+O(s^{3})&\begin{array}[]{r}(h^{\prime}h^{-1})^{\prime}=(a{}^{t}\!a)h^{-1}\\ a\in\mathrm{M}_{n,r}(\mathbb{R})\end{array}\\ \hline\cr\text{CI}&\text{None}&\begin{array}[]{r}(h^{\prime}h^{-1})^{\prime}=(ch^{-1})^{2}\\ c\in\operatorname{Sym}_{n}(\mathbb{R})\end{array}\\ \hline\cr\end{array}

2 Proof for exponential matrix solution

2.1 Block-Gauss decomposition

Manifold Ω(J)\Omega(J)

We define a submanifold Ω(J)\Omega(J) in the cone Sym2n+r+()\operatorname{Sym}^{+}_{2n+r}(\mathbb{R}) by

Ω(J)={GSym2n+r+();JGJ=G1},\Omega(J)=\left\{G\in\operatorname{Sym}^{+}_{2n+r}(\mathbb{R});\;JGJ=G^{-1}\right\},
J=(00In0Ir0In00).J=\begin{pmatrix}0&0&I_{n}\\ 0&I_{r}&0\\ I_{n}&0&0\end{pmatrix}.

Setting the set V(J)V(J) by

V(J)={BSym2n+r();JBJ=B},V(J)=\left\{B\in\operatorname{Sym}_{2n+r}(\mathbb{R});\;JBJ=-B\right\},

we can consider the following map:

V(J)BexpBΩ(J).V(J)\ni B\mapsto\exp B\in\Omega(J).

We see that any element BV(J)B\in V(J) has the following form:

B=B(b,a,c):=(bacat0atctab),bSymn(),aMn,r(),cAltn().B=B(b,a,c):=\begin{pmatrix}b&a&c\\ {}^{t}\!a&0&-{}^{t}\!a\\ {}^{t}\!c&-a&-b\end{pmatrix},\quad\begin{array}[]{l}b\in\operatorname{Sym}_{n}(\mathbb{R}),\\ a\in\mathrm{M}_{n,r}(\mathbb{R}),\\ c\in\mathrm{Alt}_{n}(\mathbb{R}).\end{array}

For an element BV(J)B\in V(J), we define an Ω(J)\Omega(J)-valued function G(s;B)G(s;B) by

G(s;B):=exp(sB(b,a,c))(s).G(s;B):=\exp(sB(b,a,c))\quad(s\in\mathbb{R}).

We define the following sets of matrices:

𝒢(J)={GGL2n+r();JGJ=G1t}O(n+r,n),\displaystyle\mathcal{G}(J)=\left\{G\in\operatorname{GL}_{2n+r}(\mathbb{R});\;JGJ={}^{t}\!G^{-1}\right\}\cong\mathrm{O}(n+r,n),
𝒩(J)={N𝒢(J);N=(In00Ir0In)},\displaystyle\mathcal{N}(J)=\left\{N\in\mathcal{G}(J);\;N=\begin{pmatrix}I_{n}&0&0\\ *&I_{r}&0\\ *&*&I_{n}\end{pmatrix}\right\},
={N=(In00n1tIr0n2tn1In),n2+n2t=n1n1t},\displaystyle\qquad=\left\{N=\begin{pmatrix}I_{n}&0&0\\ {}^{t}\!n_{1}&I_{r}&0\\ {}^{t}\!n_{2}&-n_{1}&I_{n}\end{pmatrix},n_{2}+{}^{t}\!n_{2}=-n_{1}{}^{t}\!n_{1}\right\},
Ω0(J)={AΩ(J);A=(000000)},\displaystyle\Omega_{0}(J)=\left\{A\in\Omega(J);\;A=\begin{pmatrix}*&0&0\\ 0&*&0\\ 0&0&*\end{pmatrix}\right\},
={A=(h000Ir000h1),ht=h}.\displaystyle\qquad=\left\{A=\begin{pmatrix}h&0&0\\ 0&I_{r}&0\\ 0&0&h^{-1}\end{pmatrix},{}^{t}\!h=h\right\}.
Proposition 2.1 (Block-Gauss decomposition for Ω(J)\Omega(J)).

The following map is bijective:

𝒩(J)×Ω0(J)Ω(J)(N,A)NANt.\begin{array}[]{ccc}\mathcal{N}(J)\times\Omega_{0}(J)&\rightarrow&\Omega(J)\\ (N,A)&\mapsto&NA{}^{t}\!N.\end{array}

We denote the variable change defined through this decomposition by

GN,Ah,n1,n2,G\rightarrow N,A\rightarrow h,n_{1},n_{2},

or simply by GhG\rightarrow h.

Now we state the exponential matrix solution in [I2020] as follows.

Proposition 2.2.

For B=B(b,a,0)V(J)B=B(b,a,0)\in V(J), define a Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R})-valued function h~(s)\tilde{h}(s) by the variable change

G(s;B)h~(s).G(s;B)\rightarrow\tilde{h}(s).

Then, h~(s)\tilde{h}(s) satisfies the matrix-valued Bratu equation;

(h~(s)h~(s)1)=(aat)h~(s)1.(\tilde{h}^{\prime}(s)\tilde{h}(s)^{-1})^{\prime}=(a{}^{t}\!a)\tilde{h}(s)^{-1}.

In the rest of this section, we give an alternative proof for this proposition based on the Lagrangian calculus.

2.2 Lagrangian calculus

Lemma 2.1.

Let G=G(s)G=G(s) be an Ω(J)\Omega(J)-valued function. Under the variable change

GN,Ah,n1,n2,G\rightarrow N,A\rightarrow h,n_{1},n_{2},

the following equation holds:

12tr((GG1)2)\displaystyle\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2})
=12tr((AA1)2)+tr(ANtN1tA1N1N)\displaystyle=\frac{1}{2}\operatorname{tr}((A^{\prime}A^{-1})^{2})+\operatorname{tr}(A{}^{t}\!N^{\prime}{}^{t}\!N^{-1}A^{-1}N^{-1}N^{\prime})
=tr((hh1)2)+2tr(hn1n1t)+tr(hmthm),\displaystyle=\operatorname{tr}((h^{\prime}h^{-1})^{2})+2\operatorname{tr}(hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime})+\operatorname{tr}(h{}^{t}\!mhm),
m=n1n1t+n2=mt.m=n_{1}{}^{t}\!n^{\prime}_{1}+n^{\prime}_{2}=-{}^{t}\!m.
Proof.

To prove this lemma, we use the following orthogonal relation.

Lemma 2.2.

Define a subset of M=M2n+r()M=\mathrm{M}_{2n+r}(\mathbb{R}) as

M0={XM;X=(000)},M_{\leq 0}=\left\{X\in M;\;X=\begin{pmatrix}*&0&0\\ *&*&0\\ *&*&*\end{pmatrix}\right\},
M<0={XM;X=(000000)}.M_{<0}=\left\{X\in M;\;X=\begin{pmatrix}0&0&0\\ *&0&0\\ *&*&0\end{pmatrix}\right\}.

Then, it holds that

tr(X1X2)=0\operatorname{tr}(X_{1}X_{2})=0

for any X1M0,X2M<0X_{1}\in M_{\leq 0},X_{2}\in M_{<0}.

Inserting G=NANtG=NA{}^{t}\!N into GG1G^{\prime}G^{-1}, we get

GG1\displaystyle G^{\prime}G^{-1}
=NANtN1tA1N1+NAA1N1+NN1\displaystyle=NA{}^{t}\!N^{\prime}{}^{t}\!N^{-1}A^{-1}N^{-1}+NA^{\prime}A^{-1}N^{-1}+N^{\prime}N^{-1}
=:X1+X2+X3.\displaystyle=:X_{1}+X_{2}+X_{3}.

Here we see that

X1Ad(N)M<0t,X2Ad(N)M0tM0,X_{1}\in\operatorname{Ad}(N){}^{t}\!M_{<0},\quad X_{2}\in\operatorname{Ad}(N){}^{t}\!M_{\leq 0}\cap M_{\leq 0},
X3M<0.X_{3}\in M_{<0}.

Then, by Lemma 2.2, we get

12tr((GG1)2)\displaystyle\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2}) =tr(X1X3)+12tr(X2X2).\displaystyle=\operatorname{tr}(X_{1}X_{3})+\frac{1}{2}\operatorname{tr}(X_{2}X_{2}).

Inserting

AA1\displaystyle A^{\prime}A^{-1} =(hh10000000h1h),\displaystyle=\begin{pmatrix}h^{\prime}h^{-1}&0&0\\ 0&0&0\\ 0&0&-h^{-1}h^{\prime}\end{pmatrix},
N1N\displaystyle N^{-1}N^{\prime} =(000n1t00mn10),m=n1n1t+n2,\displaystyle=\begin{pmatrix}0&0&0\\ {}^{t}\!n^{\prime}_{1}&0&0\\ m&-n^{\prime}_{1}&0\end{pmatrix},\;m=n_{1}{}^{t}\!n^{\prime}_{1}+n^{\prime}_{2},

we calculate as

tr(X1X3)\displaystyle\operatorname{tr}(X_{1}X_{3}) =tr(ANtN1tA1N1N)\displaystyle=\operatorname{tr}(A{}^{t}\!N^{\prime}{}^{t}\!N^{-1}A^{-1}N^{-1}N^{\prime})
=2tr(hn1n1t)+tr(hmthm),\displaystyle=2\operatorname{tr}(hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime})+\operatorname{tr}(h{}^{t}\!mhm),
12tr(X2X2)\displaystyle\frac{1}{2}\operatorname{tr}(X_{2}X_{2}) =12tr((AA1)2)=tr((hh1)2).\displaystyle=\frac{1}{2}\operatorname{tr}((A^{\prime}A^{-1})^{2})=\operatorname{tr}((h^{\prime}h^{-1})^{2}).

Therefore, we get the Lagrangian as

12tr((GG1)2)=\displaystyle\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2})= tr((hh1)2)+2tr(hn1n1t)\displaystyle\operatorname{tr}((h^{\prime}h^{-1})^{2})+2\operatorname{tr}(hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime})
+tr(hmthm).\displaystyle+\operatorname{tr}(h{}^{t}\!mhm).

Next we proceed to derive the Euler-Lagrange equation. For the preparation, we deal with the constrain condition for G(s)G(s) as follows.

Lemma 2.3.

For the Lagrangian

=12tr((GG1)2)+tr((JGJGtI)λ)+tr((GGt)μ)\mathcal{L}=\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2})+\operatorname{tr}((JGJ{}^{t}\!G-I)\lambda)+\operatorname{tr}((G-{}^{t}\!G)\mu)

for GL(n,)\operatorname{GL}(n,\mathbb{R})-valued function G,λ,μG,\lambda,\mu, the Euler-Lagrange equation reads

{(GG1)=0,G(μμt)G+J(μμt)J=0.\displaystyle\begin{cases}(G^{\prime}G^{-1})^{\prime}=0,\\ G(\mu-{}^{t}\!\mu)G+J(\mu-{}^{t}\!\mu)J=0.\end{cases}
Proof.

By the property of the matrix trace, we have the equation

tr(JG1JG2tλ)\displaystyle\operatorname{tr}(JG_{1}J{}^{t}\!G_{2}\lambda) =tr(G1JG2tλJ)\displaystyle=\operatorname{tr}(G_{1}\cdot J{}^{t}\!G_{2}\lambda J)
=tr(G2JG1tJλt).\displaystyle=\operatorname{tr}(G_{2}\cdot J{}^{t}\!G_{1}J{}^{t}\!\lambda).

Then, we can derive the Euler-Lagrange equation as

G:\displaystyle G: G1(G′′GG1G)G1=JGtλJ+JGtJλt,\displaystyle G^{-1}(G^{\prime\prime}-G^{\prime}G^{-1}G^{\prime})G^{-1}=J{}^{t}\!G\lambda J+J{}^{t}\!GJ{}^{t}\!\lambda,
(GG1)=JλJ+λt+G(μμt),\displaystyle\Leftrightarrow(G^{\prime}G^{-1})^{\prime}=J\lambda J+{}^{t}\!\lambda+G(\mu-{}^{t}\!\mu),
λ:\displaystyle\lambda: GJGt=J,\displaystyle GJ{}^{t}\!G=J,
μ:\displaystyle\mu: GGt=0.\displaystyle G-{}^{t}\!G=0.

Applying the map XXJXtJX\mapsto X-J{}^{t}\!XJ to the both sides of the first equation, we get

2(GG1)=G(μμt)+J(μμt)JG1\displaystyle 2(G^{\prime}G^{-1})^{\prime}=G(\mu-{}^{t}\!\mu)+J(\mu-{}^{t}\!\mu)JG^{-1}
\displaystyle\Leftrightarrow 2(GG1)G=G(μμt)G+J(μμt)J.\displaystyle 2(G^{\prime}G^{-1})^{\prime}G=G(\mu-{}^{t}\!\mu)G+J(\mu-{}^{t}\!\mu)J.

We see that the left hand side is a symmetric matrix and the right hand side is a skew-symmetric matrix. Therefore, they turn to be 0. ∎

Derivation of Euler-Lagrange equation

Now we derive the Euler-Lagrange equation for our Lagrangian

\displaystyle\mathcal{L} =(h,n1,n2,h,n1,n2)\displaystyle=\mathcal{L}(h,n_{1},n_{2},h^{\prime},n_{1}^{\prime},n_{2}^{\prime})
=tr((hh1)2)+2tr(hn1n1t)+tr(hmthm),\displaystyle=\operatorname{tr}((h^{\prime}h^{-1})^{2})+2\operatorname{tr}(hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime})+\operatorname{tr}(h{}^{t}\!mhm),
m=n1n1t+n2.\displaystyle m=n_{1}{}^{t}\!n^{\prime}_{1}+n^{\prime}_{2}.

The result is written as follows:

variable x:\displaystyle\text{variable }x: E-L eq. ddsx=x\displaystyle\text{E-L eq. }\displaystyle\frac{d}{ds}\frac{\partial\mathcal{L}}{\partial x^{\prime}}=\frac{\partial\mathcal{L}}{\partial x}
h:\displaystyle h: (hh1)=hn1n1t+hmthm,\displaystyle(h^{\prime}h^{-1})^{\prime}=hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime}+h{}^{t}\!mhm,
n1:\displaystyle n_{1}: {2hn1+hmthn1}=hmhn1,\displaystyle\left\{2hn_{1}^{\prime}+h{}^{t}\!mhn_{1}\right\}^{\prime}=hmhn_{1}^{\prime},
n2:\displaystyle n_{2}: (hmth)=0.\displaystyle(h{}^{t}\!mh)^{\prime}=0.

From the third equation, we get

hmth=c~:constant matrix,c~Altn().h{}^{t}\!mh=\tilde{c}:\text{constant matrix},\quad\tilde{c}\in\mathrm{Alt}_{n}(\mathbb{R}).

Thus, the above equations become

h:\displaystyle h: (hh1)=hn1n1tc~h1c~h1,\displaystyle(h^{\prime}h^{-1})^{\prime}=hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime}-\tilde{c}h^{-1}\tilde{c}h^{-1},
n1:\displaystyle n_{1}: 2(hn1)+2(c~n1)=0.\displaystyle 2(hn_{1}^{\prime})^{\prime}+2(\tilde{c}n_{1})^{\prime}=0.

From the second equation, we get

hn1+cn1=a~:constant matrix.hn_{1}^{\prime}+{c}n_{1}=\tilde{a}:\text{constant matrix}.

Especially, in the case c~=0\tilde{c}=0, we get hn1=a~hn_{1}^{\prime}=\tilde{a}. Inserting this into the first equation, we obtain the matrix-valued Bratu equation

h:\displaystyle h: (hh1)=(a~a~t)h1.\displaystyle(h^{\prime}h^{-1})^{\prime}=(\tilde{a}{}^{t}\!\tilde{a})h^{-1}.

Next we identify the constant matrices c~,a~\tilde{c},\tilde{a}.

Defining the function G(s)G(s) by the relation Gh,n1,n2G\rightarrow h,n_{1},n_{2}, we have that

GG1=(hn1hmhn1hmth)=(a~c~).G^{\prime}G^{-1}=\begin{pmatrix}*&hn_{1}^{\prime}-hmhn_{1}&h{}^{t}\!mh\\ *&*&*\\ *&*&*\end{pmatrix}=\begin{pmatrix}*&\tilde{a}&\tilde{c}\\ *&*&*\\ *&*&*\end{pmatrix}.

On the other hand, G(s)G(s) satisfies the original E-L equation

GG1=(constant matrix).G^{\prime}G^{-1}=\text{(constant matrix)}.

This shows that G(s)G(s) is given as

G=G(s;B(b~,a~,c~)),b~Symn().G=G(s;B(\tilde{b},\tilde{a},\tilde{c})),\quad\tilde{b}\in\operatorname{Sym}_{n}(\mathbb{R}).

We summarize the above calculation in the following table:

G(s;B(b,a,c))h(s),n1(s),n2(s)(GG1)=0{(hh1)=hn1n1tch1ch1hn1+cn1=ah(n1n1t+n2)th=c12tr((GG1)2)tr((hh1)2)+2tr(hn1n1t)+tr(hmthm)\begin{array}[]{ccc}\hline\cr\\ G(s;B(b,a,c))&\rightarrow&h(s),n_{1}(s),n_{2}(s)\\ \\ (G^{\prime}G^{-1})^{\prime}=0&&\begin{cases}(h^{\prime}h^{-1})^{\prime}=hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime}-{c}h^{-1}{c}h^{-1}\\ hn_{1}^{\prime}+{c}n_{1}={a}\\ h{}^{t}\!(n_{1}{}^{t}\!n^{\prime}_{1}+n^{\prime}_{2})h=c\end{cases}\\ &&\\ \displaystyle\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2})&&\operatorname{tr}((h^{\prime}h^{-1})^{2})+2\operatorname{tr}(hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime})+\operatorname{tr}(h{}^{t}\!mhm)\\ \\ \hline\cr\end{array}
Theorem 2.1.

The equation

G(s)G(s)1=B(b,a,c)G^{\prime}(s)G(s)^{-1}=B(b,a,c)

for an Ω(J)\Omega(J)-valued function G(s)G(s) is expressed as

{(hh1)=hn1n1tch1ch1,hn1+cn1=a,h(n1n1t+n2)th=c\begin{cases}(h^{\prime}h^{-1})^{\prime}=hn_{1}^{\prime}{}^{t}\!n_{1}^{\prime}-{c}h^{-1}{c}h^{-1},\\ hn_{1}^{\prime}+{c}n_{1}={a},\\ h{}^{t}\!(n_{1}{}^{t}\!n^{\prime}_{1}+n^{\prime}_{2})h=c\end{cases} (2.1)

by the variable change G(s)h(s),n1(s),n2(s)G(s)\rightarrow h(s),n_{1}(s),n_{2}(s). Especially, in the case c=0{c}=0, it is equivalent to the matrix-valued Bratu equation

(hh1)=(aat)h1.(h^{\prime}h^{-1})^{\prime}=({a}{}^{t}\!{a})h^{-1}.

The exponential matrix solution (Proposition 2.2) follows from this theorem.

3 Expression in symmetric domains

Considering a realization of Ω(J)\Omega(J) as a symmetric domain, we express

  • the condition c=0c=0 for B=B(b,a,c)B=B(b,a,c),

  • the variable change GhG\rightarrow h by the block-Gauss decomposition

in Proposition 2.2 in terms of the structure of symmetric domains.

3.1 Siegel domain of type BDI

Siegel domain D(J)D(J) of type BDI

For the set

M={U=(U1U2U3)M2n+r,n();rankU=n},\displaystyle M=\left\{U=\begin{pmatrix}U_{1}\\ U_{2}\\ U_{3}\end{pmatrix}\in\mathrm{M}_{2n+r,n}(\mathbb{R});\;\operatorname{rank}U=n\right\},

we consider the group actions

GL(2n+r,)MGL(n,)\operatorname{GL}(2n+r,\mathbb{R})\curvearrowright M\curvearrowleft\operatorname{GL}(n,\mathbb{R})

by the matrix multiplication. For the quatient

D:=M/GL(n,)={u=[U]=[U1U2U3];UM},D:=M/\operatorname{GL}(n,\mathbb{R})=\left\{u=[U]=\begin{bmatrix}U_{1}\\ U_{2}\\ U_{3}\end{bmatrix};\;U\in M\right\},

we equip it with the induced topology to consider DD as a domain. We let

J=(00In0Ir0In00)J=-\begin{pmatrix}0&0&I_{n}\\ 0&I_{r}&0\\ I_{n}&0&0\end{pmatrix}

and define a subdomain D(J)DD(J)\subset D by

M(J)={UM;UtJU>0},M(J)=\left\{U\in M;\;{}^{t}\!UJU>0\right\},
D(J):={u=[U]D;UM(J)}.D(J):=\left\{u=[U]\in D;\;U\in M(J)\right\}.
Lemma 3.1 (Realization as a Siegel domain).

The domain D(J)D(J) is the following set:

D(J)={u=[U1U2In]D;U1tU1U2tU2>0}.D(J)=\left\{u=\begin{bmatrix}U_{1}\\ U_{2}\\ I_{n}\end{bmatrix}\in D;\;-{}^{t}\!U_{1}-U_{1}-{}^{t}\!U_{2}U_{2}>0\right\}.

We define a subdomain ΣD\Sigma\subset D (the Shilov boundary of D(J)D(J)) by

Σ={u=[U1U2In]D;U1tU1U2tU2=0}.\Sigma=\left\{u=\begin{bmatrix}U_{1}\\ U_{2}\\ I_{n}\end{bmatrix}\in D;\;-{}^{t}\!U_{1}-U_{1}-{}^{t}\!U_{2}U_{2}=0\right\}.

We take two points u0,wDu_{0},w\in D as

u0=[U0]D(J),w=[W]Σ,\displaystyle u_{0}=[U_{0}]\in D(J),\quad w=[W]\in\Sigma,
U0=(In0In),W=(00In).\displaystyle U_{0}=\begin{pmatrix}-I_{n}\\ 0\\ I_{n}\end{pmatrix},\quad W=\begin{pmatrix}0\\ 0\\ I_{n}\end{pmatrix}.

For two points u1,u2D(J)u_{1},u_{2}\in D(J) expressed as

uj=[Uj],Uj=(Uj,1Uj,2Uj,3)(j=1,2),u_{j}=[U_{j}],\;U_{j}=\begin{pmatrix}U_{j,1}\\ U_{j,2}\\ U_{j,3}\end{pmatrix}\quad(j=1,2),

we define Δ(u1,u2)Mn()\Delta(u_{1},u_{2})\in\mathrm{M}_{n}(\mathbb{R}) by

Δ(u1,u2):=(U1Γ11)tJ(U2Γ21),\displaystyle\Delta(u_{1},u_{2}):={}^{t}\!(U_{1}\Gamma_{1}^{-1})J(U_{2}\Gamma_{2}^{-1}),
Γj=WtUj=Uj,3(j=1,2).\Gamma_{j}={}^{t}\!WU_{j}=U_{j,3}\quad(j=1,2).

Expression of D(J)D(J) as a symmetric space G/KG/K

Consider the action 𝒢(J)D(J)\mathcal{G}(J)\curvearrowright D(J) defined by

gu=[gU](g𝒢(J),u=[U]D(J)).g\cdot u=[gU]\quad(g\in\mathcal{G}(J),u=[U]\in D(J)).

We also define the following subgroups:

𝒫(J)={g𝒢(J);g=(000)},\displaystyle\mathcal{P}(J)=\left\{g\in\mathcal{G}(J);\;g=\begin{pmatrix}*&*&*\\ 0&*&*\\ 0&0&*\end{pmatrix}\right\},
𝒦(J)={g𝒢(J);gt=g1}O(n+r)×O(n),\displaystyle\mathcal{K}(J)=\left\{g\in\mathcal{G}(J);\;{}^{t}\!g=g^{-1}\right\}\cong\mathrm{O}(n+r)\times\mathrm{O}(n),
𝒢(J)u0={g𝒢(J);gu0=u0}.\displaystyle\mathcal{G}(J)_{u_{0}}=\left\{g\in\mathcal{G}(J);\;gu_{0}=u_{0}\right\}.
Lemma 3.2.

The following assertions hold:

  1. 1.

    𝒢(J)=𝒫(J)𝒦(J)\mathcal{G}(J)=\mathcal{P}(J)\mathcal{K}(J).

  2. 2.

    𝒢(J)u0=𝒦(J)\mathcal{G}(J)_{u_{0}}=\mathcal{K}(J).

  3. 3.

    The action 𝒢(J)D(J)\mathcal{G}(J)\curvearrowright D(J) is transitive, and it holds that

    D(J)SO0(n+r,n)/SO(n+r)×SO(n).D(J)\cong\operatorname{SO}_{0}(n+r,n)/\operatorname{SO}(n+r)\times\operatorname{SO}(n).
Proof.

2. Let

A=12(In0In02Ir0In0In),A=\frac{1}{\sqrt{2}}\begin{pmatrix}I_{n}&0&-I_{n}\\ 0&\sqrt{2}I_{r}&0\\ I_{n}&0&I_{n}\end{pmatrix},

and

v0:=Au0=[In00],K:=AJAt=(In00In+r),v_{0}:=Au_{0}=\begin{bmatrix}I_{n}\\ 0\\ 0\end{bmatrix},\quad K:=AJ{}^{t}\!A=\begin{pmatrix}I_{n}&0\\ 0&-I_{n+r}\end{pmatrix},
𝒢(K)=A𝒢(J)At.\mathcal{G}(K)=A\mathcal{G}(J){}^{t}\!A.

For gGL(2n+r,)g\in\operatorname{GL}(2n+r,\mathbb{R}), we see the equivalence

gv0=v0g=(0).\displaystyle gv_{0}=v_{0}\Leftrightarrow g=\begin{pmatrix}*&*\\ 0&*\end{pmatrix}.

Thus, for g𝒢(K)g\in\mathcal{G}(K), it holds the equivalence

gv0=v0g=g1t.\displaystyle gv_{0}=v_{0}\Leftrightarrow g={}^{t}\!g^{-1}.

Therefore, for g𝒢(J)g\in\mathcal{G}(J),

gu0=u0g=g1t.\displaystyle gu_{0}=u_{0}\Leftrightarrow g={}^{t}\!g^{-1}.

Isomorphism :Ω(J)D(J)\mathcal{F}:\Omega(J)\xrightarrow{\sim}D(J)

We define a map :Ω(J)D\mathcal{F}:\Omega(J)\xrightarrow{}D by

(G)=[G3InG2h],G=(hG2G3)Ω(J),\mathcal{F}(G)=\begin{bmatrix}G_{3}-I_{n}\\ G_{2}\\ h\end{bmatrix},\quad G=\begin{pmatrix}h&*&*\\ G_{2}&*&*\\ G_{3}&*&*\end{pmatrix}\in\Omega(J),

and a map π:Ω(J)Symn+()\pi:\Omega(J)\rightarrow\operatorname{Sym}^{+}_{n}(\mathbb{R}) by

π(G)=h,G=(hG2G3)Ω(J).\pi(G)=h,\quad G=\begin{pmatrix}h&*&*\\ G_{2}&*&*\\ G_{3}&*&*\end{pmatrix}\in\Omega(J).

We define an action 𝒢(J)Ω(J)\mathcal{G}(J)\curvearrowright\Omega(J) by

gG=g1tGg1(g𝒢(J),GΩ(J)).g\cdot G={}^{t}\!g^{-1}Gg^{-1}\quad(g\in\mathcal{G}(J),G\in\Omega(J)).
Lemma 3.3.
  1. 1.

    For any g𝒢(J),GΩ(J)g\in\mathcal{G}(J),G\in\Omega(J), it holds that

    (gG)=g(G).\mathcal{F}(g\cdot G)=g\mathcal{F}(G).
  2. 2.

    It holds that (Ω(J))D(J)\mathcal{F}(\Omega(J))\subset D(J), and :Ω(J)D(J)\mathcal{F}:\Omega(J)\rightarrow D(J) is bijective.

  3. 3.

    For any GΩ(J)G\in\Omega(J), it holds that

    12π(G)=Δ((G),(G))1.\frac{1}{2}\pi(G)=\Delta(\mathcal{F}(G),\mathcal{F}(G))^{-1}.

By this lemma, we obtain the following commutative diagram:

Ω(J)\textstyle{\Omega(J)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\mathcal{F}}\scriptstyle{\circlearrowright}12π\scriptstyle{\frac{1}{2}\pi}D(J)\textstyle{D(J)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ(,)1\scriptstyle{\Delta(\cdot,\cdot)^{-1}}Symn+()\textstyle{\operatorname{Sym}^{+}_{n}(\mathbb{R})}
Proof.

1. First, we prove for the case G=I2n+rG=I_{2n+r}. Take an arbitrary g𝒢(J)g\in\mathcal{G}(J) and express it as

g=pk,p𝒫(J),k𝒦(J)g=pk,\quad p\in\mathcal{P}(J),\;k\in\mathcal{K}(J)
p=(p1p2p3t0Irp2t00p11t).p=\begin{pmatrix}p_{1}&-p_{2}&{}^{t}\!p_{3}\\ 0&I_{r}&{}^{t}\!p_{2}\\ 0&0&{}^{t}\!p_{1}^{-1}\end{pmatrix}.

Then, from the relation p1t=JpJ{}^{t}\!p^{-1}=JpJ we get

pI2n+r\displaystyle p\cdot I_{2n+r} =p1tp1\displaystyle={}^{t}\!p^{-1}p^{-1}
=(p11t00p2tIr0p3tp2p1)(p11p2p30Irp2t00p1t)\displaystyle=\begin{pmatrix}{}^{t}\!p_{1}^{-1}&0&0\\ {}^{t}\!p_{2}&I_{r}&0\\ {}^{t}\!p_{3}&-p_{2}&p_{1}\end{pmatrix}\begin{pmatrix}p_{1}^{-1}&p_{2}&p_{3}\\ 0&I_{r}&-{}^{t}\!p_{2}\\ 0&0&{}^{t}\!p_{1}\end{pmatrix}
=(p11tp11p2tp11p3tp11).\displaystyle=\begin{pmatrix}{}^{t}\!p_{1}^{-1}p_{1}^{-1}&*&*\\ {}^{t}\!p_{2}p_{1}^{-1}&*&*\\ {}^{t}\!p_{3}p_{1}^{-1}&*&*\end{pmatrix}.

Therefore, its image under the map \mathcal{F} is

(pI2n+r)\displaystyle\mathcal{F}(p\cdot I_{2n+r}) =[p3tp11Inp2tp11p11tp11]=[p3tp1p2tp11t].\displaystyle=\begin{bmatrix}{}^{t}\!p_{3}p_{1}^{-1}-I_{n}\\ {}^{t}\!p_{2}p_{1}^{-1}\\ {}^{t}\!p_{1}^{-1}p_{1}^{-1}\end{bmatrix}=\begin{bmatrix}{}^{t}\!p_{3}-p_{1}\\ {}^{t}\!p_{2}\\ {}^{t}\!p_{1}^{-1}\end{bmatrix}.

On the other hand, we see that

p(I2n+r)=p[In0In]=[p1+p3tp2tp11t].\displaystyle p\mathcal{F}(I_{2n+r})=p\begin{bmatrix}-I_{n}\\ 0\\ I_{n}\end{bmatrix}=\begin{bmatrix}-p_{1}+{}^{t}\!p_{3}\\ {}^{t}\!p_{2}\\ {}^{t}\!p_{1}^{-1}\end{bmatrix}.

It shows that (pI2n+r)=p(I2n+r)\mathcal{F}(p\cdot I_{2n+r})=p\mathcal{F}(I_{2n+r}). Therefore,

(gI2n+r)=(pI2n+r)\displaystyle\mathcal{F}(g\cdot I_{2n+r})=\mathcal{F}(p\cdot I_{2n+r})
=pu0=pku0=g(I2n+r).\displaystyle=pu_{0}=pku_{0}=g\mathcal{F}(I_{2n+r}).

Next, we show the assertion for the general case g𝒢(J),GΩ(J)g\in\mathcal{G}(J),G\in\Omega(J). We express GG as

G=g0I2n+r,g0𝒢(J).G=g_{0}\cdot I_{2n+r},\quad g_{0}\in\mathcal{G}(J).

Then, we see that

(gG)=(gg0I2n+r)=gg0(I2n+r)\displaystyle\mathcal{F}(g\cdot G)=\mathcal{F}(gg_{0}\cdot I_{2n+r})=gg_{0}\mathcal{F}(I_{2n+r})
=g(g0I2n+r)=g(G).\displaystyle=g\mathcal{F}(g_{0}\cdot I_{2n+r})=g\mathcal{F}(G).

2. The equation (Ω(J))=D(J)\mathcal{F}(\Omega(J))=D(J) is follows from 1 and the transitivity of the action of 𝒢(J)\mathcal{G}(J). The injectivity of \mathcal{F} holds because the isotropy subgroups are coincident to be 𝒦(J)\mathcal{K}(J).

3. Take an arbitrary GΩ(J)G\in\Omega(J) and express it as

G=(hG2G3),hG3+G3th+G2tG2=0.G=\begin{pmatrix}h&*&*\\ G_{2}&*&*\\ G_{3}&*&*\end{pmatrix},\quad hG_{3}+{}^{t}\!G_{3}h+{}^{t}\!G_{2}G_{2}=0.

Thus, its image is

(G)=[U],U=(U1U2U3)=(G3InG2h).\displaystyle\mathcal{F}(G)=[U],\quad U=\begin{pmatrix}U_{1}\\ U_{2}\\ U_{3}\end{pmatrix}=\begin{pmatrix}G_{3}-I_{n}\\ G_{2}\\ h\end{pmatrix}.

Setting Γ:=h\Gamma:=h, we get

Δ((G),(G))=(UΓ1)tJ(UΓ1)\displaystyle\Delta(\mathcal{F}(G),\mathcal{F}(G))={}^{t}\!(U\Gamma^{-1})J(U\Gamma^{-1})
=(G3h1h1)t(G3h1h1)h1G2tG2h1\displaystyle=-{}^{t}\!(G_{3}h^{-1}-h^{-1})-(G_{3}h^{-1}-h^{-1})-h^{-1}{}^{t}\!G_{2}G_{2}h^{-1}
=2h1=2π(G)1.\displaystyle=2h^{-1}=2\pi(G)^{-1}.

Characterization of B=B(b,a,0)B=B(b,a,0)

𝔤(J)={X𝔤𝔩(2n+r,);XJ+JXt=0}𝔰𝔬(n+r,n),\displaystyle\mathfrak{g}(J)=\left\{X\in\mathfrak{gl}(2n+r,\mathbb{R});\;XJ+J{}^{t}\!X=0\right\}\cong\mathfrak{so}(n+r,n),
𝔭(J)={X𝔤(J);X=Xt}.\displaystyle\mathfrak{p}(J)=\left\{X\in\mathfrak{g}(J);\;X={}^{t}\!X\right\}.

An arbitrary element 𝔭(J)\mathfrak{p}(J) has the form

B=B(b,a,c):=(bacat0atctab),bSymn(),aMn,r(),cAltn().B=B(b,a,c):=\begin{pmatrix}b&a&c\\ {}^{t}\!a&0&-{}^{t}\!a\\ {}^{t}\!c&-a&-b\end{pmatrix},\quad\begin{array}[]{l}b\in\operatorname{Sym}_{n}(\mathbb{R}),\\ a\in\mathrm{M}_{n,r}(\mathbb{R}),\\ c\in\mathrm{Alt}_{n}(\mathbb{R})\end{array}.

Let wΣw\in\Sigma as above,

Lemma 3.4.

For an element B(b,a,c)𝔭(J)B(b,a,c)\in\mathfrak{p}(J), it holds the following Taylor expansion:

Δ(w,esBw)=cs+(aatbccb)s22+O(s3).\Delta(w,e^{sB}w)=-cs+\left(a{}^{t}\!a-bc-cb\right)\frac{s^{2}}{2}+O(s^{3}).
Proof.

Let

Γ1=WtW=In,Γ(s)=WtesBW,\Gamma_{1}={}^{t}\!WW=I_{n},\quad\Gamma(s)={}^{t}\!We^{sB}W,
Φ(s)=WtJesBW.\Phi(s)={}^{t}\!WJe^{sB}W.

Then, we have

Δ(w,esBw)\displaystyle\Delta(w,e^{sB}w) =(WΓ11)tJ(esBWΓ(s)1)\displaystyle={}^{t}\!(W\Gamma_{1}^{-1})J(e^{sB}W\Gamma(s)^{-1})
=Φ(s)Γ(s)1.\displaystyle=\Phi(s)\Gamma(s)^{-1}.

By the definition, it is easily seen that

Φ(s)=cs+(aatbc+cb)s22+O(s3),\displaystyle\Phi(s)=-cs+(a{}^{t}\!a-bc+cb)\frac{s^{2}}{2}+O(s^{3}),
Γ(s)1=In+bs+O(s2).\displaystyle\Gamma(s)^{-1}=I_{n}+bs+O(s^{2}).

Then, we obtain the expansion

Δ(w,esBw)=cs+(aatbccb)s22+O(s3).\Delta(w,e^{sB}w)=-cs+\left(a{}^{t}\!a-bc-cb\right)\frac{s^{2}}{2}+O(s^{3}).

Proposition 3.1.

For B𝔭(J)B\in\mathfrak{p}(J), assume the Taylor expansion

Δ(w,esBw)=(aat)s22+O(s3),aMn,r()\Delta(w,e^{sB}w)=(a{}^{t}\!a)\frac{s^{2}}{2}+O(s^{3}),\quad a\in\mathrm{M}_{n,r}(\mathbb{R})

at s=0s=0. Define a Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R})-valued function h(s)h(s) by

h(s)=Δ(esBu0,esBu0)1.h(s)=\Delta(e^{sB}u_{0},e^{sB}u_{0})^{-1}.

Then, h(s)h(s) satisfies the matrix-valued Bratu equation;

(hh1)=2(aat)h1.(h^{\prime}h^{-1})^{\prime}=2(a{}^{t}\!a)h^{-1}.
Proof.

From the assumption, the element BB has the form B=B(b,a,0)B=B(b,a,0). Since

1(esBu0)=e2sB=:G(s),\mathcal{F}^{-1}(e^{sB}u_{0})=e^{-2sB}=:G(s),

the function h~(s):=π(G(s))\tilde{h}(s):=\pi(G(s)) satisfies the matrix-valued Bratu equation

(h~h~1)=4(aat)h~1.(\tilde{h}^{\prime}\tilde{h}^{-1})^{\prime}=4(a{}^{t}\!a)\tilde{h}^{-1}.

Therefore, the function

h(s):=Δ(esBu0,esBu0)1=12h~(s){h}(s):=\Delta(e^{sB}u_{0},e^{sB}u_{0})^{-1}=\frac{1}{2}\tilde{h}(s)

satisfies the equation

(hh1)=2(aat)h1.({h}^{\prime}{h}^{-1})^{\prime}=2(a{}^{t}\!a){h}^{-1}.

3.2 Bounded domain of type BDI and power series solution

Bounded domain D(L)D(L) of type BDI

Let

L=(In000Ir000In).\displaystyle L=\begin{pmatrix}I_{n}&0&0\\ 0&-I_{r}&0\\ 0&0&-I_{n}\end{pmatrix}.

We define a subdomain D(L)DD(L)\subset D by

M(L)={UM;UtLU>0},\displaystyle M(L)=\left\{U\in M;\;{}^{t}\!ULU>0\right\},
D(L):={u=[U]D;UM(L)}.\displaystyle D(L):=\left\{u=[U]\in D;\;U\in M(L)\right\}.

We notice the relation L=AJAtL=AJ{}^{t}\!A with a matrix

A=12(In0In02Ir0In0In).A=\frac{1}{\sqrt{2}}\begin{pmatrix}I_{n}&0&-I_{n}\\ 0&\sqrt{2}I_{r}&0\\ I_{n}&0&I_{n}\end{pmatrix}.
Lemma 3.5.
  1. 1.

    It holds the following bijection:

    D(J)D(L):uAu\displaystyle D(J)\xrightarrow{\sim}D(L):u\mapsto Au
  2. 2.

    (Realization as a bounded domain) D(L)D(L) is described as

    D(L)={v=[InV2V3]D;InV2tV2V3tV3>0}.D(L)=\left\{v=\begin{bmatrix}I_{n}\\ V_{2}\\ V_{3}\end{bmatrix}\in D;\;I_{n}-{}^{t}\!V_{2}V_{2}-{}^{t}\!V_{3}V_{3}>0\right\}.

Next, we take two points v0,zv_{0},z as

v0=[V0]D(L),z=[Z]AΣ,\displaystyle v_{0}=[V_{0}]\in D(L),\quad z=[Z]\in A\Sigma,
V0=2(In00)=AU0,Z=12(In0In)=AW.\displaystyle V_{0}=-\sqrt{2}\begin{pmatrix}I_{n}\\ 0\\ 0\end{pmatrix}=AU_{0},\quad Z=\frac{1}{\sqrt{2}}\begin{pmatrix}-I_{n}\\ 0\\ I_{n}\end{pmatrix}=AW.

For two points v1,v2D(L)v_{1},v_{2}\in D(L) expressed as

vj=[Vj],Vj=(Vj,1Vj,2Vj,3)(j=1,2),v_{j}=[V_{j}],\quad V_{j}=\begin{pmatrix}V_{j,1}\\ V_{j,2}\\ V_{j,3}\end{pmatrix}\quad(j=1,2),

we define ψ(v1,v2)Mn()\psi(v_{1},v_{2})\in\mathrm{M}_{n}(\mathbb{R}) by

ψ(v1,v2)\displaystyle\psi(v_{1},v_{2}) =(V1Γ11)tL(V2Γ21),\displaystyle={}^{t}\!(V_{1}\Gamma_{1}^{-1})L(V_{2}\Gamma_{2}^{-1}),
Γj=ZtVj=(Vj,3Vj,1)/2(j=1,2).\Gamma_{j}={}^{t}\!ZV_{j}=(V_{j,3}-V_{j,1})/{\sqrt{2}}\quad(j=1,2).
Lemma 3.6.

For any u1,u2D(J)u_{1},u_{2}\in D(J), it holds that

ψ(Au1,Au2)=Δ(u1,u2).\psi(Au_{1},Au_{2})=\Delta(u_{1},u_{2}).

By above lemma, we have the following commutative diagram:

D(J)×D(J)\textstyle{D(J)\times D(J)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A×A\scriptstyle{A\times A}\scriptstyle{\circlearrowright}Δ(,)\scriptstyle{\Delta(\cdot,\cdot)}D(L)×D(L)\textstyle{D(L)\times D(L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ(,)\scriptstyle{\psi(\cdot,\cdot)}Symn+()\textstyle{\operatorname{Sym}^{+}_{n}(\mathbb{R})}
Proof.

Take arbitrary u1,u2D(J)u_{1},u_{2}\in D(J) and express them as

uj=[Uj](j=1,2).u_{j}=[U_{j}]\quad(j=1,2).

With the relation

Γj=Zt(AUj)=WtUj,\Gamma_{j}={}^{t}\!Z(AU_{j})={}^{t}\!WU_{j},

we calculate as

ψ(Au1,Au2)\displaystyle\psi(Au_{1},Au_{2}) =(AU1Γ11)tL(AU2Γ21)\displaystyle={}^{t}\!(AU_{1}\Gamma_{1}^{-1})L(AU_{2}\Gamma_{2}^{-1})
=(U1Γ11)tJ(U2Γ21)\displaystyle={}^{t}\!(U_{1}\Gamma_{1}^{-1})J(U_{2}\Gamma_{2}^{-1})
=Δ(u,u).\displaystyle=\Delta(u,u).

Put 𝔭(L)={X𝔤(L);X=Xt}\mathfrak{p}(L)=\left\{X\in\mathfrak{g}(L);\;X={}^{t}\!X\right\}.

Proposition 3.2.

For B𝔭(L)B\in\mathfrak{p}(L), assume the following Taylor expansion

ψ(z,esBz)=(aat)s22+O(s3),aMn,r()\psi(z,e^{sB}z)=(a{}^{t}\!a)\frac{s^{2}}{2}+O(s^{3}),\quad a\in\mathrm{M}_{n,r}(\mathbb{R})

at s=0s=0. Define a Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R})-valued function h(s)h(s) by

h(s)=ψ(esBv0,esBv0)1,h(s)=\psi(e^{sB}v_{0},e^{sB}v_{0})^{-1},

Then, h(s)h(s) satisfies the matrix-valued Bratu equation;

(hh1)=2(aat)h1.(h^{\prime}h^{-1})^{\prime}=2(a{}^{t}\!a)h^{-1}.

Derivation of the power series solution

We take an arbitrary B𝔭(L)B\in\mathfrak{p}(L) and express it

B=(0CtC0),CMn+r,n().B=\begin{pmatrix}0&{}^{t}\!C\\ C&0\end{pmatrix},\quad C\in\mathrm{M}_{n+r,n}(\mathbb{R}).

We consider the singular value decomposition of the submatrix CC in the following form:

C=(p1p3p2p)(0σ)qt,C=\begin{pmatrix}p_{1}&p_{3}\\ p_{2}&p\end{pmatrix}\begin{pmatrix}0\\ \sigma\end{pmatrix}{}^{t}\!q,
(p1p3p2p)O(n+r),qO(n),\begin{pmatrix}p_{1}&p_{3}\\ p_{2}&p\end{pmatrix}\in\mathrm{O}(n+r),\quad q\in\mathrm{O}(n),
σ=diag(σi)i=1nMn().\sigma=\operatorname{diag}(\sigma_{i})_{i=1}^{n}\in\mathrm{M}_{n}(\mathbb{R}).

We define the following functions of ss\in\mathbb{R}:

ch(sσ):=12(esσ+esσ),sh(sσ):=12(esσesσ).\mathrm{ch}(s\sigma):=\frac{1}{2}(e^{s\sigma}+e^{-s\sigma}),\quad\mathrm{sh}(s\sigma):=\frac{1}{2}(e^{s\sigma}-e^{-s\sigma}).
Proposition 3.3.

For B𝔭(L)B\in\mathfrak{p}(L), define a function

h(s):=ψ(esBv0,esBv0)1.h(s):=\psi(e^{sB}v_{0},e^{sB}v_{0})^{-1}.

Then, h(s)h(s) has the following expression:

h(s)\displaystyle h(s) =12𝒆(s)𝒆t(s),𝒆(s)=psh(sσ)qch(sσ).\displaystyle=\frac{1}{2}\bm{e}(s){}^{t}\!\bm{e}(s),\quad\bm{e}(s)=p\mathrm{sh}(s\sigma)-q\mathrm{ch}(s\sigma).
Proof.

From the singular value decomposition, we get the equation

B=P(00σ000σ00)Pt,P:=(q000p1p30p2p).B=P\begin{pmatrix}0&0&\sigma\\ 0&0&0\\ \sigma&0&0\end{pmatrix}{}^{t}\!P,\quad P:=\begin{pmatrix}q&0&0\\ 0&p_{1}&p_{3}\\ 0&p_{2}&p\end{pmatrix}.

Then, we can calculate its exponential as

esBV0\displaystyle e^{sB}V_{0} =P(ch(sσ)0sh(sσ)0Ir0sh(sσ)0ch(sσ))PtV0\displaystyle=P\begin{pmatrix}\mathrm{ch}(s\sigma)&0&\mathrm{sh}(s\sigma)\\ 0&I_{r}&0\\ \mathrm{sh}(s\sigma)&0&\mathrm{ch}(s\sigma)\end{pmatrix}{}^{t}\!PV_{0}
=2(qch(sσ)qtp3sh(sσ)qtpsh(sσ)qt).\displaystyle=\sqrt{2}\begin{pmatrix}q\mathrm{ch}(s\sigma){}^{t}\!q\\ p_{3}\mathrm{sh}(s\sigma){}^{t}\!q\\ p\mathrm{sh}(s\sigma){}^{t}\!q\end{pmatrix}.

Putting

Γ:=psh(sσ)qtqch(sσ)qt,𝒆(s)=Γq,\Gamma:=p\mathrm{sh}(s\sigma){}^{t}\!q-q\mathrm{ch}(s\sigma){}^{t}\!q,\quad\bm{e}(s)=\Gamma q,

we obtain that

ψ(esBv0,esBv0)\displaystyle\psi(e^{sB}v_{0},e^{sB}v_{0}) =(esBV0Γ1)tL(esBV0Γ1)\displaystyle={}^{t}\!(e^{sB}V_{0}\Gamma^{-1})L(e^{sB}V_{0}\Gamma^{-1})
=2Γ1tΓ1=2𝒆t(s)1𝒆(s)1.\displaystyle=2{}^{t}\!\Gamma^{-1}\Gamma^{-1}=2{}^{t}\!\bm{e}(s)^{-1}\bm{e}(s)^{-1}.

Corollary 3.1.

h(s)h(s) has the following power series expansion:

h(s)\displaystyle h(s) =12In+12n=1(qσ2nqt+pσ2npt)s2n(2n)!\displaystyle=\frac{1}{2}I_{n}+\frac{1}{2}\sum_{n=1}^{\infty}\left(q\sigma^{2n}\;{}^{t}\!q+p\sigma^{2n}\;{}^{t}\!p\right)\frac{s^{2n}}{(2n)!}
12n=1(pσ2n1qt+qσ2n1pt)s2n1(2n1)!.\displaystyle-\frac{1}{2}\sum_{n=1}^{\infty}\left(p\sigma^{2n-1}\;{}^{t}\!q+q\sigma^{2n-1}\;{}^{t}\!p\right)\frac{s^{2n-1}}{(2n-1)!}.
Proof.

It is shown by the direct calculation:

12𝒆(s)𝒆t(s)\displaystyle\frac{1}{2}\bm{e}(s){}^{t}\!\bm{e}(s)
=12(qp)(ch(sσ)2ch(sσ)sh(sσ)sh(sσ)ch(sσ)sh(sσ)2)(qtpt)\displaystyle=\frac{1}{2}\begin{pmatrix}-q&p\end{pmatrix}\begin{pmatrix}\mathrm{ch}(s\sigma)^{2}&\mathrm{ch}(s\sigma)\mathrm{sh}(s\sigma)\\ \mathrm{sh}(s\sigma)\mathrm{ch}(s\sigma)&\mathrm{sh}(s\sigma)^{2}\end{pmatrix}\begin{pmatrix}-{}^{t}\!q\\ {}^{t}\!p\end{pmatrix}
=14(qp)(ch(2sσ)+Insh(2sσ)sh(2sσ)ch(2sσ)In)(qtpt)\displaystyle=\frac{1}{4}\begin{pmatrix}-q&p\end{pmatrix}\begin{pmatrix}\mathrm{ch}(2s\sigma)+I_{n}&\mathrm{sh}(2s\sigma)\\ \mathrm{sh}(2s\sigma)&\mathrm{ch}(2s\sigma)-I_{n}\end{pmatrix}\begin{pmatrix}-{}^{t}\!q\\ {}^{t}\!p\end{pmatrix}
=14{2In+q(ch(2sσ)In)qt+p(ch(2sσ)In)pt\displaystyle=\frac{1}{4}\{2I_{n}+q(\mathrm{ch}(2s\sigma)-I_{n}){}^{t}\!q+p(\mathrm{ch}(2s\sigma)-I_{n}){}^{t}\!p
qsh(2sσ)ptpsh(2sσ)qt}.\displaystyle\quad-q\mathrm{sh}(2s\sigma){}^{t}\!p-p\mathrm{sh}(2s\sigma){}^{t}\!q\}.

4 Analog to symmetric domain of type CI

Manifold Ω(J)\Omega(J)

We define a submanifold Ω(J)\Omega(J) in the cone Sym2n+()\operatorname{Sym}^{+}_{2n}(\mathbb{R}) by

Ω(J)={GSym2n+();JGJ=G1},\Omega(J)=\left\{G\in\operatorname{Sym}^{+}_{2n}(\mathbb{R});\;JGJ=G^{-1}\right\},
J=(0iIniIn0).J=\begin{pmatrix}0&iI_{n}\\ -iI_{n}&0\end{pmatrix}.

Setting the set V(J)V(J) by

V(J)={BSym2n();JBJ=B},V(J)=\left\{B\in\operatorname{Sym}_{2n}(\mathbb{R});\;JBJ=-B\right\},

we can consider the following map:

V(J)BexpBΩ(J).V(J)\ni B\mapsto\exp B\in\Omega(J).

We see that any element BV(J)B\in V(J) has the following form:

B=B(b,c):=(bccb),b,cSymn().B=B(b,c):=\begin{pmatrix}b&c\\ c&-b\end{pmatrix},\quad\begin{array}[]{l}b,c\in\operatorname{Sym}_{n}(\mathbb{R}).\end{array}

For an element BV(J)B\in V(J), we define an Ω(J)\Omega(J)-valued function G(s;B)G(s;B) by

G(s;B):=exp(sB(b,c))(s).G(s;B):=\exp(sB(b,c))\quad(s\in\mathbb{R}).

We define the following sets of matrices:

𝒢(J)={GGL2n();JGJ=G1t}Sp(n,),\displaystyle\mathcal{G}(J)=\left\{G\in\operatorname{GL}_{2n}(\mathbb{R});\;JGJ={}^{t}\!G^{-1}\right\}\cong\mathrm{Sp}(n,\mathbb{R}),
𝒩(J)={N𝒢(J);N=(In0In)},\displaystyle\mathcal{N}(J)=\left\{N\in\mathcal{G}(J);\;N=\begin{pmatrix}I_{n}&0\\ *&I_{n}\end{pmatrix}\right\},
={N=(In0n1In),n1=n1t},\displaystyle\qquad=\left\{N=\begin{pmatrix}I_{n}&0\\ n_{1}&I_{n}\end{pmatrix},\;n_{1}={}^{t}\!n_{1}\right\},
Ω0(J)={AΩ(J);A=(00)},\displaystyle\Omega_{0}(J)=\left\{A\in\Omega(J);\;A=\begin{pmatrix}*&0\\ 0&*\end{pmatrix}\right\},
={A=(h00h1),ht=h}.\displaystyle\qquad=\left\{A=\begin{pmatrix}h&0\\ 0&h^{-1}\end{pmatrix},{}^{t}\!h=h\right\}.
Proposition 4.1 (Block-Gauss decomposition for Ω(J)\Omega(J)).

The following map is bijective:

𝒩(J)×Ω0(J)Ω(J)(N,A)NANt.\begin{array}[]{ccc}\mathcal{N}(J)\times\Omega_{0}(J)&\rightarrow&\Omega(J)\\ (N,A)&\mapsto&NA{}^{t}\!N.\end{array}

We denote the variable change defined through this decomposition by

GN,Ah,n1,G\rightarrow N,A\rightarrow h,n_{1},

or simply by GhG\rightarrow h.

Lemma 4.1.

Under the variable change

GN,Ah,n1,G\rightarrow N,A\rightarrow h,n_{1},

it holds that

12tr((GG1)2)\displaystyle\frac{1}{2}\operatorname{tr}((G^{\prime}G^{-1})^{2})
=12tr((AA1)2)+tr(ANtN1tA1N1N)\displaystyle=\frac{1}{2}\operatorname{tr}((A^{\prime}A^{-1})^{2})+\operatorname{tr}(A{}^{t}\!N^{\prime}{}^{t}\!N^{-1}A^{-1}N^{-1}N^{\prime})
=tr((hh1)2)+tr(hn1thn1).\displaystyle=\operatorname{tr}((h^{\prime}h^{-1})^{2})+\operatorname{tr}(h{}^{t}\!n_{1}^{\prime}hn_{1}^{\prime}).

By the direct derivation of the Euler-Lagrange equation, we obtain the following analog of the matrix-valued Bratu equation.

Theorem 4.1.

The equation

G(s)G(s)1=B(b,c)G^{\prime}(s)G(s)^{-1}=B(b,c)

for an Ω(J)\Omega(J)-valued function G(s)G(s) is expressed as

(h(s)h(s)1)=(ch(s)1)2({h}^{\prime}(s){h}(s)^{-1})^{\prime}=(c{h}(s)^{-1})^{2}

by the variable change G(s)h(s),n1(s)G(s)\rightarrow h(s),n_{1}(s).

4.1 Siegel domain D(J;K)D(J;K) of type CI

Similarly as the type of BDI, we put

M={U=(U1U2)M2n,n();rankU=n}\displaystyle M=\left\{U=\begin{pmatrix}U_{1}\\ U_{2}\end{pmatrix}\in\mathrm{M}_{2n,n}(\mathbb{C});\;\operatorname{rank}U=n\right\}

and consider the actions

GL(2n,)MGL(n,)\displaystyle\operatorname{GL}(2n,\mathbb{C})\curvearrowright M\curvearrowleft\operatorname{GL}(n,\mathbb{C})

by the matrix multiplication. Put the quotient space

D:=M/GL(n,)={u=[U]=[U1U2];UM}D:=M/\operatorname{GL}(n,\mathbb{C})=\left\{u=[U]=\begin{bmatrix}U_{1}\\ U_{2}\end{bmatrix};\;U\in M\right\}

and regard it as a domain. We put

J=(0iIniIn0),K=(0InIn0)J=\begin{pmatrix}0&iI_{n}\\ -iI_{n}&0\end{pmatrix},\quad K=\begin{pmatrix}0&I_{n}\\ -I_{n}&0\end{pmatrix}

and define a subdomain D(J;K)D(J;K) by

M(J;K)={UM;UJU>0,UtKU=0},M(J;K)=\left\{U\in M;\;U^{*}JU>0,{}^{t}\!UKU=0\right\},
D(J;K):={u=[U]D;UM(J;K)}.D(J;K):=\left\{u=[U]\in D;\;U\in M(J;K)\right\}.
Lemma 4.2 (Realization as a Siegel upper half space).

D(J;K)D(J;K) is the following set:

D(J;K)\displaystyle D(J;K) ={u=[U1In];1i(U1U1)>0,U1t=U1}\displaystyle=\left\{u=\begin{bmatrix}U_{1}\\ I_{n}\end{bmatrix};\;\frac{1}{i}(U_{1}-U_{1}^{*})>0,{}^{t}\!U_{1}=U_{1}\right\}
Sp(n,)/U(n).\displaystyle\cong\operatorname{Sp}(n,\mathbb{R})/\mathrm{U}(n).

We define a subdomain ΣD\Sigma\subset D (Shilov boundary of D(J;K)D(J;K)) by

Σ={u=[U1In];1i(U1U1)=0,U1t=U1}.\Sigma=\left\{u=\begin{bmatrix}U_{1}\\ I_{n}\end{bmatrix};\;\frac{1}{i}(U_{1}-U_{1}^{*})=0,{}^{t}\!U_{1}=U_{1}\right\}.

We take two points u0,wu_{0},w as

u0=[U0]D(J;K),w=[W]Σ,\displaystyle u_{0}=[U_{0}]\in D(J;K),\quad w=[W]\in\Sigma,
U0=(iInIn),W=(0In).\displaystyle U_{0}=\begin{pmatrix}iI_{n}\\ I_{n}\end{pmatrix},\quad W=\begin{pmatrix}0\\ I_{n}\end{pmatrix}.

For two points u1,u2D(J;K)u_{1},u_{2}\in D(J;K) expressed as

uj=[Uj],Uj=(Uj,1Uj,2)(j=1,2),u_{j}=[U_{j}],\quad U_{j}=\begin{pmatrix}U_{j,1}\\ U_{j,2}\end{pmatrix}\quad(j=1,2),

we define Δ(u1,u2)Mn()\Delta(u_{1},u_{2})\in\mathrm{M}_{n}(\mathbb{C}) by

Δ(u1,u2)\displaystyle\Delta(u_{1},u_{2}) =(U1Γ11)J(U2Γ21),\displaystyle=(U_{1}\Gamma_{1}^{-1})^{*}J(U_{2}\Gamma_{2}^{-1}),
Γj=WtUj=Uj,2(j=1,2).\Gamma_{j}={}^{t}\!WU_{j}=U_{j,2}\quad(j=1,2).

Put

𝔤(J)={X𝔤𝔩(2n+r,);XJ+JXt=0}\displaystyle\mathfrak{g}(J)=\left\{X\in\mathfrak{gl}(2n+r,\mathbb{R});\;XJ+J{}^{t}\!X=0\right\}
𝔰𝔭(n,),\displaystyle\qquad\cong\mathfrak{sp}(n,\mathbb{R}),
𝔭(J)={X𝔤(J);X=Xt}\displaystyle\mathfrak{p}(J)=\left\{X\in\mathfrak{g}(J);\;X={}^{t}\!X\right\}
={B𝔤(J);B=(bccb),b,cSymn()}.\displaystyle=\left\{B\in\mathfrak{g}(J);\;B=\begin{pmatrix}b&c\\ c&-b\end{pmatrix},\begin{array}[]{l}b,c\in\operatorname{Sym}_{n}(\mathbb{R})\end{array}\right\}.
Proposition 4.2 (Analog to type CI).

For B𝔭(J)B\in\mathfrak{p}(J), define a Symn+()\operatorname{Sym}^{+}_{n}(\mathbb{R})-valued function h(s)h(s) by

h(s)=Δ(esBu0,esBu0)1.h(s)=\Delta(e^{sB}u_{0},e^{sB}u_{0})^{-1}.

Then, h(s)h(s) satisfies the equation

(hh1)=(ch1)2.(h^{\prime}h^{-1})^{\prime}=(ch^{-1})^{2}.

We can define an action 𝒢(J)D(J:K)\mathcal{G}(J)\curvearrowright D(J:K) by

gu=[gU](g𝒢(J),u=[U]D(J;K)).g\cdot u=[gU]\quad(g\in\mathcal{G}(J),u=[U]\in D(J;K)).

We also define subgroups:

𝒫(J)={g𝒢(J);g=(0)},\displaystyle\mathcal{P}(J)=\left\{g\in\mathcal{G}(J);\;g=\begin{pmatrix}*&*\\ 0&*\end{pmatrix}\right\},
𝒦(J)={g𝒢(J);gt=g1}U(n),\displaystyle\mathcal{K}(J)=\left\{g\in\mathcal{G}(J);\;{}^{t}\!g=g^{-1}\right\}\cong\mathrm{U}(n),
𝒢(J)u0={g𝒢(J);gu0=u0}.\displaystyle\mathcal{G}(J)_{u_{0}}=\left\{g\in\mathcal{G}(J);\;gu_{0}=u_{0}\right\}.
Lemma 4.3.

The following assertions hold.

  1. 1.

    𝒢(J)=𝒫(J)𝒦(J)\mathcal{G}(J)=\mathcal{P}(J)\mathcal{K}(J).

  2. 2.

    𝒢(J)u0=𝒦(J)\mathcal{G}(J)_{u_{0}}=\mathcal{K}(J).

  3. 3.

    The action 𝒢D(J;K)\mathcal{G}\curvearrowright D(J;K) is transitive and the following expression holds:

    D(J;K)Sp(n,)/U(n).D(J;K)\cong\operatorname{Sp}(n,\mathbb{R})/\mathrm{U}(n).

Proof of the proposition

Define a manifold

Ω(J)\displaystyle\Omega(J) ={gHerm2n+();gJg=J,gKgt=K}\displaystyle=\left\{g\in\operatorname{Herm}^{+}_{2n}(\mathbb{C});\;gJg^{*}=J,gK{}^{t}\!g=K\right\}
={gSym2n+();gJgt=J}\displaystyle=\left\{g\in\operatorname{Sym}^{+}_{2n}(\mathbb{R});\;gJ{}^{t}\!g=J\right\}

and a map :Ω(J)D\mathcal{F}:\Omega(J)\xrightarrow{}D by

(G)=[iInG2h],G=(hG2)Ω(J;K),\mathcal{F}(G)=\begin{bmatrix}iI_{n}-G_{2}\\ h\end{bmatrix},\quad G=\begin{pmatrix}h&*\\ G_{2}&*\end{pmatrix}\in\Omega(J;K),

and a map π:Ω(J)Symn+()\pi:\Omega(J)\rightarrow\operatorname{Sym}^{+}_{n}(\mathbb{R}) by

π(G)=h,G=(hG2)Ω(J;K).\pi(G)=h,\quad G=\begin{pmatrix}h&*\\ G_{2}&*\end{pmatrix}\in\Omega(J;K).

We define an action 𝒢(J)Ω(J)\mathcal{G}(J)\curvearrowright\Omega(J) by

gG=g1tGg1(g𝒢(J),GΩ(J)).g\cdot G={}^{t}\!g^{-1}Gg^{-1}\quad(g\in\mathcal{G}(J),G\in\Omega(J)).
Lemma 4.4.

The following assertions hold.

  1. 1.

    (gG)=g(G)\mathcal{F}(g\cdot G)=g\mathcal{F}(G) for any g𝒢(J),GΩ(J)g\in\mathcal{G}(J),G\in\Omega(J).

  2. 2.

    (Ω(J))D(J)\mathcal{F}(\Omega(J))\subset D(J) and the map :Ω(J)D(J)\mathcal{F}:\Omega(J)\rightarrow D(J) is bijective.

  3. 3.

    For any GΩ(J)G\in\Omega(J), it holds that

    12π(G)=Δ((G),(G))1.\frac{1}{2}\pi(G)=\Delta(\mathcal{F}(G),\mathcal{F}(G))^{-1}.

From the above lemma, we have the following commutative diagram:

Ω(J)\textstyle{\Omega(J)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\mathcal{F}}\scriptstyle{\circlearrowright}12π\scriptstyle{\frac{1}{2}\pi}D(J;K)\textstyle{D(J;K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ(,)1\scriptstyle{\Delta(\cdot,\cdot)^{-1}}Symn+()\textstyle{\operatorname{Sym}^{+}_{n}(\mathbb{R})}
Proof.

Each GΩ(J)G\in\Omega(J) is expressed as

G=g1tg1=(hhn1n1thIn+n1thn1),G={}^{t}\!g^{-1}g^{-1}=\begin{pmatrix}h&hn_{1}\\ {}^{t}\!n_{1}h&I_{n}+{}^{t}\!n_{1}hn_{1}\end{pmatrix},
g1t=(In0n1tIr)(h1200h12)𝒢(J).{}^{t}\!g^{-1}=\begin{pmatrix}I_{n}&0\\ {}^{t}\!n_{1}&I_{r}\end{pmatrix}\begin{pmatrix}h^{\frac{1}{2}}&0\\ 0&h^{-\frac{1}{2}}\end{pmatrix}\in\mathcal{G}(J).

We also see that

gu0\displaystyle gu_{0} =(In10I)(h1200h12)[iInIn]\displaystyle=\begin{pmatrix}I&-n_{1}\\ 0&I\end{pmatrix}\begin{pmatrix}h^{\frac{1}{2}}&0\\ 0&h^{-\frac{1}{2}}\end{pmatrix}\begin{bmatrix}iI_{n}\\ I_{n}\end{bmatrix}
=[ih1n1In].\displaystyle=\begin{bmatrix}ih^{-1}-n_{1}\\ I_{n}\end{bmatrix}.

1. First, we show in the case G=I2nG=I_{2n}. Take an arbitrary g𝒢(J)g\in\mathcal{G}(J) and express it as

g=pk,p𝒫(J),k𝒦(J),g=pk,\quad p\in\mathcal{P}(J),\;k\in\mathcal{K}(J),
p=(p1p20p11t).p=\begin{pmatrix}p_{1}&-p_{2}\\ 0&{}^{t}\!p_{1}^{-1}\end{pmatrix}.

Then, using the relation p1t=J1pJ{}^{t}\!p^{-1}=J^{-1}pJ, we get

pI2n\displaystyle p\cdot I_{2n} =p1tGp1\displaystyle={}^{t}\!p^{-1}Gp^{-1}
=(p11t0p2p1)(p11p2t0p1t)\displaystyle=\begin{pmatrix}{}^{t}\!p_{1}^{-1}&0\\ p_{2}&p_{1}\end{pmatrix}\begin{pmatrix}p_{1}^{-1}&{}^{t}\!p_{2}\\ 0&{}^{t}\!p_{1}\end{pmatrix}
=(p11tp11p2p11).\displaystyle=\begin{pmatrix}{}^{t}\!p_{1}^{-1}p_{1}^{-1}&*\\ p_{2}p_{1}^{-1}&*\end{pmatrix}.

Therefore,

(pI2n)\displaystyle\mathcal{F}(p\cdot I_{2n}) =[iInp2p11p11tp11]=[ip1p2p11t].\displaystyle=\begin{bmatrix}iI_{n}-p_{2}p_{1}^{-1}\\ {}^{t}\!p_{1}^{-1}p_{1}^{-1}\end{bmatrix}=\begin{bmatrix}ip_{1}-p_{2}\\ {}^{t}\!p_{1}^{-1}\end{bmatrix}.

On the other hand, we see that

p(I2n)=p[iInIn]=[ip1p2p11t]\displaystyle p\mathcal{F}(I_{2n})=p\begin{bmatrix}iI_{n}\\ I_{n}\end{bmatrix}=\begin{bmatrix}ip_{1}-p_{2}\\ {}^{t}\!p_{1}^{-1}\end{bmatrix}

and then (pI2n)=p(I2n)\mathcal{F}(p\cdot I_{2n})=p\mathcal{F}(I_{2n}). Therefore,

(gI2n)=(pI2n)\displaystyle\mathcal{F}(g\cdot I_{2n})=\mathcal{F}(p\cdot I_{2n})
=pu0=pku0=g(I2n).\displaystyle=pu_{0}=pku_{0}=g\mathcal{F}(I_{2n}).

Next, we show in the general case g𝒢(J),GΩ(J)g\in\mathcal{G}(J),G\in\Omega(J). Express an arbitrary GΩ(J)G\in\Omega(J) as

G=g0I2n,g0𝒢(J).G=g_{0}\cdot I_{2n},\quad g_{0}\in\mathcal{G}(J).

Then we see that

(gG)=(gg0I2n)=gg0(I2n)\displaystyle\mathcal{F}(g\cdot G)=\mathcal{F}(gg_{0}\cdot I_{2n})=gg_{0}\mathcal{F}(I_{2n})
=g(g0I2n)=g(G).\displaystyle=g\mathcal{F}(g_{0}\cdot I_{2n})=g\mathcal{F}(G).

3. Take an arbitrary GΩ(J)G\in\Omega(J) and express it as

G=(hG2tG2),G2thhG2=0.\displaystyle G=\begin{pmatrix}h&{}^{t}\!G_{2}\\ G_{2}&*\end{pmatrix},\quad{}^{t}\!G_{2}h-hG_{2}=0.

Then we see that

(G)=[U],U=(iInG2h).\mathcal{F}(G)=[U],\quad U=\begin{pmatrix}iI_{n}-G_{2}\\ h\end{pmatrix}.

Putting

Γ=WtU=h,\Gamma={}^{t}\!WU=h,

we see that

Δ((G),(G))=(UΓ1)J(UΓ1)\displaystyle\Delta(\mathcal{F}(G),\mathcal{F}(G))=(U\Gamma^{-1})^{*}J(U\Gamma^{-1})
=1i{(ih1G2h1)(ih1G2h1)}\displaystyle=\frac{1}{i}\left\{(ih^{-1}-G_{2}h^{-1})-(ih^{-1}-G_{2}h^{-1})^{*}\right\}
=2h1>0.\displaystyle=2h^{-1}>0.

Proof of Proposition 4.2. From the definition, we have

1(esBu0)=e2sB=:G(s).\mathcal{F}^{-1}(e^{sB}u_{0})=e^{-2sB}=:G(s).

Then, the function h~(s):=π(G(s))\tilde{h}(s):=\pi(G(s)) satisfies

(h~h~1)=4(ch1)2.(\tilde{h}^{\prime}\tilde{h}^{-1})^{\prime}=4(ch^{-1})^{2}.

Therefore, the function

h(s):=Δ(esBu0,esBu0)1=12h~(s){h}(s):=\Delta(e^{sB}u_{0},e^{sB}u_{0})^{-1}=\frac{1}{2}\tilde{h}(s)

satisfies

(hh1)=(ch1)2.({h}^{\prime}{h}^{-1})^{\prime}=(ch^{-1})^{2}.

5 Remarks

We gave individual calculations for each symmetric domain of type BDI and CI. The same calculation, which consists of the Lagrangian calculation and the realization of symmetric domains, is likely applicable to other symmetric domains classified by Cartan, ref [2, 3]. For this purpose, it is natural to expect a unified way to treat general symmetric domains.

References

  • [1] Inoue, H., Matrix-valued Bratu equation and the exact solution of its initial value problem, Int. Jour. Math. Ind., published on 12 Mar. 2020.
  • [2] Helgason, S., Differential geometry and symmetric spaces, Academic Press, 1962.
  • [3] Pyatetskii-Shapiro, I. I., Automorphic functions and the geometry of classical domains, Gordon and Breach Science Publishers, 1969.
Hiroto Inoue
Institute of Mathematics for Industry
Kyushu University
744 Motooka, Nishi-ku
Fukuoka, 819-0395, Japan
(E-mail: [email protected])