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

An analogy of Jacobi’s formula and its applications

Chiba Jun NEC Corporation, 7-1, Shiba 5-chome Minato-ku, Tokyo 108-8001, Japan jun.chiba323@nec.com  and  Matsumoto Keiji Department of Mathematics, Faculty of Science, Hokkaido University, Sapporo 060-0810, Japan matsu@math.sci.hokudai.ac.jp
Abstract.

We give an analogy of Jacobi’s formula, which relates the hypergeometric function with parameters (1/4,1/4,1)(1/4,1/4,1) and theta constants. By using this analogy and twice formulas of theta constants, we obtain a transformation formula for this hypergeometric function. As its application, we express the limit of a pair of sequences defined by a mean iteration by this hypergeometric function.

Key words and phrases:
Hypergeometric Function, Theta constants, Mean iteration.
2010 Mathematics Subject Classification:
Primary 33C05; Secondary 14K25, 33C90.

1. Introduction

Jacobi’s formula in elliptic function theory is an equality

F(12,12,1;λ(τ))=ϑ00(τ)2,λ(τ)=ϑ10(τ(z))4ϑ00(τ(z))4,F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau))=\vartheta_{00}(\tau)^{2},\quad\lambda(\tau)=\frac{\vartheta_{10}(\tau(z))^{4}}{\vartheta_{00}(\tau(z))^{4}},

as holomorphic functions on the upper half plane ={τIm(τ)>0}\mathbb{H}=\{\tau\in\mathbb{C}\mid\mathrm{Im}(\tau)>0\}, where F(a,b,c;z)F(a,b,c;z) and ϑpq(τ)\vartheta_{pq}(\tau) denote the hypergeometric series on {z|z|<1}\{z\in\mathbb{C}\mid|z|<1\} and the theta constant with characteristics p,q{0,1}p,q\in\{0,1\} on \mathbb{H}, respectively; refer to (2.1) and (3.1) for their definitions. This formula is applied to the study of the arithmetic-geometric mean as shown in [BB], and generalized to Thomae’s formula in [Th]. In the present, there are several kinds of its analogies, which are applied to studies of mean iterations, refer to [BB], [MS], [MT], [Sh] and the references therein.

In this paper, we give an analogy of Jacobi’s formula, which is an equality

F(14,14,1;ζ(τ))=ϑ00(τ)2,ζ(τ)=4ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8,F(\frac{1}{4},\frac{1}{4},1;\zeta(\tau))=\vartheta_{00}(\tau)^{2},\quad\zeta(\tau)=\frac{4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}},

as holomorphic functions on \mathbb{H}. Strictly, we initially give this equality on a fundamental region 𝔻12\mathbb{D}_{12} of

Γ12={g=gijSL2()g11g12,g21g220mod2}\Gamma_{12}=\{g=g_{ij}\in\mathrm{SL}_{2}(\mathbb{Z})\mid g_{11}g_{12},g_{21}g_{22}\equiv 0\bmod 2\}

acting on \mathbb{H}, and extend it to the whole space \mathbb{H}. Though this equality is naturally extended to \mathbb{H} by the simply connectedness of \mathbb{H}, F(14,14,1;z)F(\frac{1}{4},\frac{1}{4},1;z) is not regarded as single valued when τ\tau runs over \mathbb{H}. To understand its behavior well, we investigate the monodromy representation of the hypergeometric differential equation (14,14,1)\mathcal{F}(\frac{1}{4},\frac{1}{4},1). It is realized in not SL2()\mathrm{SL}_{2}(\mathbb{Z}) but GL2([𝐢])\mathrm{GL}_{2}(\mathbb{Z}[\mathbf{i}]), and its projectivization becomes PΓ12=Γ12/{±I2}\mathrm{P}\Gamma_{12}=\Gamma_{12}/\{\pm I_{2}\}, where 𝐢=1\mathbf{i}=\sqrt{-1} and I2I_{2} is the unit matrix of size 22. By using circuit matrices in GL2([𝐢])\mathrm{GL}_{2}(\mathbb{Z}[\mathbf{i}]), we can simplify not only the formula for ϑ00(τ)\vartheta_{00}(\tau) with respect to the action PΓ12\mathrm{P}\Gamma_{12} in Fact 3.3 but also a description of the behavior of F(14,14,1;τ(z))F(\frac{1}{4},\frac{1}{4},1;\tau(z)); for details refer to Lemma 4.3 and Corollary 4.7. We also remark that the monodromy representation of (12,12,1)\mathcal{F}(\frac{1}{2},\frac{1}{2},1) is realized as

Γ(2,4)={g=(gij)SL2()g11,γ221mod4,g12,g210mod2},\Gamma(2,4)=\{g=(g_{ij})\in\mathrm{SL}_{2}(\mathbb{Z})\mid g_{11},\gamma_{22}\equiv 1\bmod 4,g_{12},g_{21}\equiv 0\bmod 2\},

which is a subgroup in Γ(2)={gSL2()gI2mod2}\Gamma(2)=\{g\in\mathrm{SL}_{2}(\mathbb{Z})\mid g\equiv I_{2}\bmod 2\} of index 22, and that there is a similar advantage in using not PΓ(2)=Γ(2)/{±I2}\mathrm{P}\Gamma(2)=\Gamma(2)/\{\pm I_{2}\} but Γ(2,4)\Gamma(2,4) for Jacobi’s formula, see Lemma 4.3 and Corollary 4.4.

By using our analogy of Jacobi’s formula together with twice formulas for theta constants in (3.3), we obtain a transformation formula for F(14,14,1;z)F(\frac{1}{4},\frac{1}{4},1;z) in Theorem 5.1. As studied in [Go], [HKM], [KS1], [KS2] and [Ma], some transformation formulas for hypergeometric functions are applied to expressing limits of mean iterations. By following the idea in [HKM], we define two functions on the set

𝕊(1,)={(x,rx)2x>0,1<r<}\mathbb{S}_{(-1,\infty)}=\{(x,rx)\in\mathbb{R}^{2}\mid x>0,\ -1<r<\infty\}

by

μ1(x,y)=mA(x,μ0(x,y)),μ2=ν(x,μ0(x,y)),\mu_{1}(x,y)=m_{A}(x,\mu_{0}(x,y)),\quad\mu_{2}=\nu(x,\mu_{0}(x,y)),

where mAm_{A}, mGm_{G} and mHm_{H} are the arithmetic, geometric and harmonic means, and

μ0=mG(x,mA(x,y)),ν(x,y)=2mH(x,y)mA(x,y).\mu_{0}=m_{G}(x,m_{A}(x,y)),\quad\nu(x,y)=2m_{H}(x,y)-m_{A}(x,y).

We can easily see that μ1(x,y)\mu_{1}(x,y) is a mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)}, however μ2(x,y)\mu_{2}(x,y) is not, since the inequality min(x,y)μ2(x,y)\min(x,y)\leq\mu_{2}(x,y) does not hold for any (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)}. In spite of this situation, we can define a pair of sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} by the recurrence relation

(xn+1,yn+1)=(μ1(xn,yn),μ2(xn,yn))(x_{n+1},y_{n+1})=(\mu_{1}(x_{n},y_{n}),\mu_{2}(x_{n},y_{n}))

for any given initial term (x0,y0)𝕊(1,)(x_{0},y_{0})\in\mathbb{S}_{(-1,\infty)}, and show in Lemma 6.12 that they converge as nn\to\infty and limnxn=limnyn>0\lim\limits_{n\to\infty}x_{n}=\lim\limits_{n\to\infty}y_{n}>0. We express this limit by F(14,14,1;z)F(\frac{1}{4},\frac{1}{4},1;z) in Theorem 6.13.

2. Fundamental properties of F(a,b,c;z)F(a,b,c;z)

We begin with defining the hypergeometric series.

Definition 2.1.

The hypergeometric series F(a,b,c;z)F(a,b,c;z) is defined by

F(a,b,c;z)=n=0(a,n)(b,n)(c,n)(1,n)zn,F(a,b,c;z)=\sum_{n=0}^{\infty}\frac{(a,n)(b,n)}{(c,n)(1,n)}z^{n}, (2.1)

where zz is a complex variable, a,b,ca,b,c are complex parameters with c0,1,2,c\neq 0,-1,-2,\cdots, and (a,n)=Γ(a+n)/Γ(a)=a(a+1)(a+n1)(a,n)=\mathit{\Gamma}(a+n)/\mathit{\Gamma}(a)=a(a+1)\cdots(a+n-1). It converges absolutely and uniformly on any compact set in the unit disk {z|z|<1}\{z\in\mathbb{C}\mid|z|<1\} for any fixed a,b,ca,b,c.

This series satisfies the hypergeometric differential equation

(a,b,c):[z(1z)d2dz2+{c(a+b+1)z}ddzab]f(z)=0,\mathcal{F}(a,b,c):\big{[}z(1-z)\frac{d^{2}}{dz^{2}}+\{c-(a+b+1)z\}\frac{d}{dz}-ab\big{]}\cdot f(z)=0,

which is a second order linear ordinary differential equation with regular singular points z=0,1,z=0,1,\infty. The space 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}) of local solutions to (a,b,c)\mathcal{F}(a,b,c) around a point z˙Z={0,1}\dot{z}\in Z=\mathbb{C}-\{0,1\} is a 22-dimensional complex vector space, its element admits the analytic continuation along any path in ZZ. In particular, a loop ρ\rho in ZZ with a base point z˙\dot{z} leads to a linear transformation of 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}). Thus, we have a homomorphism from the fundamental group π1(Z,z˙)\pi_{1}(Z,\dot{z}) to the general linear group GL(𝒮(a,b,c;z˙))\mathrm{GL}(\mathcal{S}(a,b,c;\dot{z})) of 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}), which is called the monodromy representation of (a,b,c)\mathcal{F}(a,b,c). We take z˙=12\dot{z}=\frac{1}{2} as a base point of ZZ, and loops ρ0\rho_{0} and ρ1\rho_{1} with terminal z˙\dot{z} as positively oriented circles with radius 12\frac{1}{2} and center 0 and 11, respectively. Since π1(Z,z˙)\pi_{1}(Z,\dot{z}) is a free group generated by ρ0\rho_{0} and ρ1\rho_{1}, the monodromy representation of (a,b,c)\mathcal{F}(a,b,c) is uniquely characterized by the images of ρ0\rho_{0} and ρ1\rho_{1}. We give their explicit forms with respect to a basis of 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}) given by Euler type integrals.

Fact 2.2 ([Ki, §17]).
  1. (ii)

    If Re(b),Re(cb),Re(1a)>0\mathrm{Re}(b),\mathrm{Re}(c-b),\mathrm{Re}(1-a)>0 then the integrals

    f1(z)=11/ztb1(1t)cb1(1tz)a𝑑t,f2(z)=01tb1(1t)cb1(1tz)a𝑑tf_{1}(z)=\int_{1}^{1/z}t^{b-1}(1-t)^{c-b-1}(1-tz)^{-a}dt,\quad f_{2}(z)=\int_{0}^{1}t^{b-1}(1-t)^{c-b-1}(1-tz)^{-a}dt

    converge and span the space 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}). Here we assign a branch of the integrand tb1(1t)cb1(1tz)at^{b-1}(1-t)^{c-b-1}(1-tz)^{-a} on the open interval (0,1)(0,1) in the tt-space by arg(t)=arg(1t)=arg(1tz)=0\arg(t)=\arg(1-t)=\arg(1-tz)=0 for real zz near to z˙=12\dot{z}=\frac{1}{2}, and that on the open interval (1,1z)(1,\frac{1}{z}) by its analytic continuation via the lower half plane of the tt-space, which transforms arg(1t)\arg(1-t) from 0 to π\pi. Under some generic conditions on a,b,ca,b,c, they admit expressions in terms of the hypergeometric series

    f1(z)\displaystyle f_{1}(z) =eπ𝐢(cb1)B(cb,1a)(1z)cabF(ca,cb,cab+1;1z),\displaystyle=e^{\pi\mathbf{i}(c-b-1)}B(c-b,1-a)(1-z)^{c-a-b}F(c-a,c-b,c-a-b+1;1-z),
    =e2π𝐢(cb)e2π𝐢ae2π𝐢a1(B(b,cb)F(a,b,c;z)B(b,cab)F(a,b,a+bc+1;1z)),\displaystyle=\frac{e^{2\pi\mathbf{i}(c-b)}\!-\!e^{2\pi\mathbf{i}a}}{e^{2\pi\mathbf{i}a}\!-\!1}\left(B(b,c\!-\!b)F(a,b,c;z)\!-\!B(b,c\!-\!a\!-\!b)F(a,b,a\!+\!b\!-\!c\!+\!1;1\!-\!z)\right),
    f2(z)\displaystyle f_{2}(z) =B(b,cb)F(a,b,c;z),\displaystyle=B(b,c-b)F(a,b,c;z),

    where BB denotes the beta function.

  2. (iiii)

    The loops ρ0\rho_{0} and ρ1\rho_{1} lead to linear transformations sending the basis 𝐅(z)=t(f1(z),f2(z))\mathbf{F}(z)=\;^{t}(f_{1}(z),f_{2}(z)) of 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}) to M0𝐅(z)M_{0}\mathbf{F}(z) and M1𝐅(z)M_{1}\mathbf{F}(z), where

    M0=(e2π𝐢ce2π𝐢b101),M1=(e2π𝐢(cab)01e2π𝐢a1).M_{0}=\begin{pmatrix}e^{-2\pi\mathbf{i}c}&e^{-2\pi\mathbf{i}b}-1\\ 0&1\end{pmatrix},\quad M_{1}=\begin{pmatrix}e^{2\pi\mathbf{i}(c-a-b)}&0\\ 1-e^{-2\pi\mathbf{i}a}&1\end{pmatrix}. (2.2)

    Here we exchange the role of f1(z)f_{1}(z) and f2(z)f_{2}(z) in [Ki, §17], our matrices M0M_{0} and M1M_{1} are the conjugates of A0A_{0} and A1A_{1} in [Ki, p.123] by U=(0110)U=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}, respectively.

Remark 2.3.

We can make the analytic continuation of the hypergeometric series F(a,b,c;z)F(a,b,c;z) to the simply connected domain [1,)\mathbb{C}-[1,\infty) as a solution to (a,b,c)\mathcal{F}(a,b,c). We use the same symbol F(a,b,c;z)F(a,b,c;z) for this continuation, which is a single-valued holomorphic function on [1,)\mathbb{C}-[1,\infty). Moreover, by assigning F(a,b,c;t)F(a,b,c;t) on t(1,)t\in(1,\infty) to the limit of F(a,b,c;z)F(a,b,c;z) as ztz\to t with Im(z)>0\mathrm{Im}(z)>0, we extend F(a,b,c;z)F(a,b,c;z) to a single-valued function on {1}\mathbb{C}-\{1\}, which is discontinuous along (1,)(1,\infty) for general a,b,ca,b,c. For a path ρ\rho in Z={0,1}Z=\mathbb{C}-\{0,1\} ending at ww, Fρ(a,b,c;w)F_{\rho}(a,b,c;w) denotes the continuation of F(a,b,c;z)F(a,b,c;z) along ρ\rho.

Though we remove the point z=1z=1 from \mathbb{C} for F(a,b,c;z)F(a,b,c;z) in Remark 2.3, there is a formula for the limit of F(a,b,c;z)F(a,b,c;z) as z1z\to 1 with zz in the unit disk.

Fact 2.4 (Gauss-Kummer’s identity [Ki, Theorem 3.3]).

If Re(cab)>0\mathrm{Re}(c-a-b)>0 then

limz1,|z|<1F(a,b,c;z)=Γ(c)Γ(cab)Γ(ca)Γ(cb).\lim_{z\to 1,|z|<1}F(a,b,c;z)=\frac{\mathit{\Gamma}(c)\mathit{\Gamma}(c-a-b)}{\mathit{\Gamma}(c-a)\mathit{\Gamma}(c-b)}. (2.3)

For the basis 𝐅(z)=t(f1(z),f2(z))\mathbf{F}(z)=\;^{t}(f_{1}(z),f_{2}(z)) of 𝒮(a,b,c;z˙)\mathcal{S}(a,b,c;\dot{z}), we have a map from a neighborhood of z˙\dot{z} to the complex projective line 1\mathbb{P}^{1} by the ratio f1(z)/f2(z)f_{1}(z)/f_{2}(z) of f1(z)f_{1}(z) and f2(z)f_{2}(z). Schwarz’s map is define by its analytic continuation to {0,1}\mathbb{C}-\{0,1\}, which is multi valued in general.

Fact 2.5.

If the parameters a,b,ca,b,c are real and satisfy

1|1c|,1|cab|,1|ab|{2,3,4}{},|1c|+|cab|+|ab|<1,\frac{1}{|1-c|},\frac{1}{|c-a-b|},\frac{1}{|a-b|}\in\{2,3,4\dots\}\cup\{\infty\},\quad|1-c|+|c-a-b|+|a-b|<1,

then the image of Schwarz’s map is isomorphic to an open dense subset of the upper half plane ={τIm(τ)>0}\mathbb{H}=\{\tau\in\mathbb{C}\mid\mathrm{Im}(\tau)>0\} and its inverse is single valued. The projectivization of the image of its monodromy representation is conjugate to a discrete subgroup of PSL2()=SL2()/{±I2}\mathrm{PSL}_{2}(\mathbb{R})=\mathrm{SL}_{2}(\mathbb{R})/\{\pm I_{2}\} generated by two elements g0g_{0} and g1g_{1} with

ord(g0)=1|1c|,ord(g1)=1|cab|,ord(g01g11)=1|ab|,\mathrm{ord}(g_{0})=\frac{1}{|1-c|},\quad\mathrm{ord}(g_{1})=\frac{1}{|c-a-b|},\quad\mathrm{ord}(g_{0}^{-1}g_{1}^{-1})=\frac{1}{|a-b|},

where I2I_{2} is the unit matrix of size 22, and ord(g0)\mathrm{ord}(g_{0}) denotes the order of g0g_{0}.

Example 2.6.

If (a,b,c)=(12,12,1)(a,b,c)=(\frac{1}{2},\frac{1}{2},1) then we have (1|1c|,1|cab|,1|ab|)=(,,)\big{(}\frac{1}{|1-c|},\frac{1}{|c-a-b|},\frac{1}{|a-b|}\big{)}=(\infty,\infty,\infty),

M0=(1201),M1=(1021),M01M11=(3221).M_{0}=\begin{pmatrix}1&-2\\ 0&1\end{pmatrix},\quad M_{1}=\begin{pmatrix}1&0\\ 2&1\end{pmatrix},\quad M_{0}^{-1}M_{1}^{-1}=\begin{pmatrix}-3&2\\ -2&1\end{pmatrix}.

The group generated by these matrices is not the principal congruence subgroup Γ(2)\Gamma(2) of level 22 in SL2()\mathrm{SL}_{2}(\mathbb{Z}) but

Γ(2,4)={g=(gij)SL2()g11,g221mod4,g12,g210mod2}.\Gamma(2,4)=\{g=(g_{ij})\in\mathrm{SL}_{2}(\mathbb{Z})\mid g_{11},g_{22}\equiv 1\bmod 4,g_{12},g_{21}\equiv 0\bmod 2\}.

Note that Γ(2,4)\Gamma(2,4) is a subgroup in Γ(2)\Gamma(2) of index 22, and that ±I2\pm I_{2} are representatives of the quotient Γ(2)/Γ(2,4)\Gamma(2)/\Gamma(2,4).

Since the image of the monodromy representation is in PSL2()\mathrm{PSL}_{2}(\mathbb{R}) and f1(z)/f2(z)f_{1}(z)/f_{2}(z) is in \mathbb{H} for z(0,1)z\in(0,1), the image of the analytic continuation of f1(z)/f2(z)f_{1}(z)/f_{2}(z) to Z={0,1}Z=\mathbb{C}-\{0,1\} is the whole space \mathbb{H}.

Example 2.7.

If (a,b,c)=(14,14,1)(a,b,c)=(\frac{1}{4},\frac{1}{4},1) then we have (1|1c|,1|cab|,1|ab|)=(,2,)\big{(}\frac{1}{|1-c|},\frac{1}{|c-a-b|},\frac{1}{|a-b|}\big{)}=(\infty,2,\infty),

M0=(11𝐢01),M1=(101+𝐢1),M01M11=(1+2𝐢1+𝐢1+𝐢1).M_{0}=\begin{pmatrix}1&-1-\mathbf{i}\\ 0&1\end{pmatrix},\quad M_{1}=\begin{pmatrix}-1&0\\ 1+\mathbf{i}&1\end{pmatrix},\quad M_{0}^{-1}M_{1}^{-1}=\begin{pmatrix}-1+2\mathbf{i}&1+\mathbf{i}\\ 1+\mathbf{i}&1\end{pmatrix}.

By using

P=(1+𝐢𝐢01),P=\begin{pmatrix}-1+\mathbf{i}&\mathbf{i}\\ 0&1\end{pmatrix},

we change the basis 𝐅(z)\mathbf{F}(z) into P𝐅(z)P\mathbf{F}(z). Then

PM0P1=(1201),PM1P1=(0𝐢𝐢0)=𝐢J,PM01M11P1=(2𝐢𝐢𝐢0).PM_{0}P^{-1}=\begin{pmatrix}1&2\\ 0&1\end{pmatrix},\quad PM_{1}P^{-1}=\begin{pmatrix}0&\mathbf{i}\\ -\mathbf{i}&0\end{pmatrix}=\mathbf{i}J,\quad PM_{0}^{-1}M_{1}^{-1}P^{-1}=\begin{pmatrix}2\mathbf{i}&\mathbf{i}\\ -\mathbf{i}&0\end{pmatrix}.

The group generated by these matrices is

Γ(2,4)𝐢J={gGL2([𝐢])gΓ(2,4) or (𝐢J)gΓ(2,4)}=Γ(2,4)(𝐢J)Γ(2,4)\Gamma(2,4)\langle\mathbf{i}J\rangle=\{g\in\mathrm{GL}_{2}(\mathbb{Z}[\mathbf{i}])\mid g\in\Gamma(2,4)\textrm{ or }(\mathbf{i}J)g\in\Gamma(2,4)\}=\Gamma(2,4)\cup(\mathbf{i}J)\cdot\Gamma(2,4)

since

(0𝐢𝐢0)(1201)(0𝐢𝐢0)=(1021).\begin{pmatrix}0&\mathbf{i}\\ -\mathbf{i}&0\end{pmatrix}\begin{pmatrix}1&2\\ 0&1\end{pmatrix}\begin{pmatrix}0&\mathbf{i}\\ -\mathbf{i}&0\end{pmatrix}=\begin{pmatrix}1&0\\ -2&1\end{pmatrix}.

Note that the projectivization of Γ(2,4)𝐢J\Gamma(2,4)\langle\mathbf{i}J\rangle is isomorphic to that of

Γ12={g=(gij)SL2()g11g12g21g220mod2}.\Gamma_{12}=\{g=(g_{ij})\in\mathrm{SL}_{2}(\mathbb{Z})\mid g_{11}g_{12}\equiv g_{21}g_{22}\equiv 0\bmod 2\}.

Since the image of the monodromy representation is in PSL2()\mathrm{PSL}_{2}(\mathbb{R}) and ((1+𝐢)f1(z)+𝐢f2(z))/f2(z)\big{(}(-1+\mathbf{i})f_{1}(z)+\mathbf{i}f_{2}(z)\big{)}/f_{2}(z) is in \mathbb{H} for z(0,1)z\in(0,1), the image of the analytic continuation of f2(z)/f1(z)f_{2}(z)/f_{1}(z) to Z={0,1}Z=\mathbb{C}-\{0,1\} is a open dense subset of \mathbb{H}.

It is known that F(a,b,c;z)F(a,b,c;z) satisfies several kinds of transformation formulas with respect to variable changes. In this paper, we use the following.

Fact 2.8 ([Er, (24) in p.64, (18),(19) in p.112 of Vol.I]).
F(a,b,2b;4z(1+z)2)\displaystyle F(a,b,2b;\frac{4z}{(1+z)^{2}}) =(1+z)2aF(a,ab+12,b+12;z2),\displaystyle=(1+z)^{2a}F(a,a-b+\frac{1}{2},b+\frac{1}{2};z^{2}), (2.4)
F(a,b,a+b+12;z)\displaystyle F(a,b,\frac{a+b+1}{2};z) =F(a2,b2,a+b+12;4z(1z))\displaystyle=F(\frac{a}{2},\frac{b}{2},\frac{a+b+1}{2};4z(1-z)) (2.5)
=(12z)F(a+12,b+12,a+b+12;4z(1z)),\displaystyle=(1-2z)F(\frac{a+1}{2},\frac{b+1}{2},\frac{a+b+1}{2};4z(1-z)), (2.6)

where zz is sufficiently near to 0 and (1+z)2a(1+z)^{2a} takes the value 11 at z=0z=0.

3. Fundamental properties of ϑpq(τ)\vartheta_{pq}(\tau)

We next introduce theta constants and their properties by referring to [Er] and [Mu].

Definition 3.1.

The theta constant ϑpq(τ)\vartheta_{pq}(\tau) with characteristics p,qp,q is defined by

ϑpq(τ)=n=exp(π𝐢(n+p2)2τ+2π𝐢(n+p2)q2),\vartheta_{pq}(\tau)=\sum_{n=-\infty}^{\infty}\exp\big{(}\pi\mathbf{i}(n+\frac{p}{2})^{2}\tau+2\pi\mathbf{i}(n+\frac{p}{2})\frac{q}{2}\big{)}, (3.1)

where τ\tau is a variable in the upper half plane \mathbb{H}, and p,qp,q are parameters taking the value 0 or 11. It converges absolutely and uniformly on any compact set in \mathbb{H} for any fixed p,qp,q.

There are four functions ϑ00(τ)\vartheta_{00}(\tau), ϑ01(τ)\vartheta_{01}(\tau), ϑ10(τ)\vartheta_{10}(\tau), ϑ11(τ)\vartheta_{11}(\tau); the last one vanishes identically on \mathbb{H}, and the rests satisfy Jacobi’s identity

ϑ00(τ)4=ϑ01(τ)4+ϑ10(τ)4\vartheta_{00}(\tau)^{4}=\vartheta_{01}(\tau)^{4}+\vartheta_{10}(\tau)^{4} (3.2)

for any τ\tau\in\mathbb{H}. We have twice formulas of them.

Fact 3.2 ([Er, (15) in p.373 of Vol.II]).

The theta constants satisfy twice formulas

ϑ00(2τ)2=ϑ00(τ)2+ϑ01(τ)22,ϑ01(2τ)2=ϑ00(τ)ϑ01(τ),ϑ10(2τ)2=ϑ00(τ)2ϑ01(τ)22.\begin{array}[]{l}\vartheta_{00}(2\tau)^{2}=\dfrac{\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2}}{2},\\[5.69054pt] \vartheta_{01}(2\tau)^{2}=\vartheta_{00}(\tau)\vartheta_{01}(\tau),\\[5.69054pt] \vartheta_{10}(2\tau)^{2}=\dfrac{\vartheta_{00}(\tau)^{2}-\vartheta_{01}(\tau)^{2}}{2}.\end{array} (3.3)
Fact 3.3 ([Mu, Theorem 7.1]).

Let g=(gij)g=(g_{ij}) be an element in the projectivization PΓ12\mathrm{P}\Gamma_{12} of Γ12\Gamma_{12}. By multiplying 1-1 to gg if necessary, we may assume that gg satisfies either g21>0g_{21}>0, or g21=0g_{21}=0 and g22>0g_{22}>0. Then we have

ϑ00(gτ)2=χ(g)(g21τ+g22)ϑ00(τ)2\vartheta_{00}(g\cdot\tau)^{2}=\chi(g)\cdot(g_{21}\tau+g_{22})\vartheta_{00}(\tau)^{2}

for any τ𝔻12\tau\in\mathbb{D}_{12}, where

χ(g)={𝐢g221ifg212,g222,𝐢g21ifg212,g222.\chi(g)=\left\{\begin{array}[]{lcc}\mathbf{i}^{g_{22}-1}&\textrm{if}&g_{21}\in 2\mathbb{Z},g_{22}\notin 2\mathbb{Z},\\ \mathbf{i}^{-g_{21}}&\textrm{if}&g_{21}\notin 2\mathbb{Z},g_{22}\in 2\mathbb{Z}.\end{array}\right.
Fact 3.4 ([Mu, Table V in p.36]).

We have

ϑ00(1τ)2=(𝐢τ)ϑ00(τ)2,ϑ01(1τ)2=(𝐢τ)ϑ10(τ)2,ϑ10(1τ)2=(𝐢τ)ϑ01(τ)2.\vartheta_{00}(\frac{-1}{\tau})^{2}=(-\mathbf{i}\tau)\vartheta_{00}(\tau)^{2},\quad\vartheta_{01}(\frac{-1}{\tau})^{2}=(-\mathbf{i}\tau)\vartheta_{10}(\tau)^{2},\quad\vartheta_{10}(\frac{-1}{\tau})^{2}=(-\mathbf{i}\tau)\vartheta_{01}(\tau)^{2}.
Fact 3.5 (The inverse of Schwarz’s map).
  1. (ii)

    [Yo, §5.6,5.7,5.8 in Chap. II, Proposition 8.1 in Chap. III] The inverse of Schwarz’s map

    zτ=f1(z)f2(z)=𝐢F(12,12,1,1z)F(12,12,1,z)z\mapsto\tau=\frac{f_{1}(z)}{f_{2}(z)}=\mathbf{i}\cdot\frac{F(\frac{1}{2},\frac{1}{2},1,1-z)}{F(\frac{1}{2},\frac{1}{2},1,z)} (3.4)

    for (a,b,c)=(12,12,1)(a,b,c)=(\frac{1}{2},\frac{1}{2},1) is

    τλ(τ)=ϑ10(τ)4ϑ00(τ)4Z.\mathbb{H}\ni\tau\mapsto\lambda(\tau)=\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\in Z.

    The function λ(τ)\lambda(\tau) is invariant under the action τ(g11τ+g12)/(g21τ+g22)\tau\mapsto(g_{11}\tau+g_{12})/(g_{21}\tau+g_{22}) of g=(gij)Γ(2)g=(g_{ij})\in\Gamma(2) on τ\tau\in\mathbb{H}.

  2. (iiii)

    [MOT, Theorem 8.4] The inverse of Schwarz’s map

    zτ=\displaystyle z\mapsto\tau= (1+𝐢)f1(z)+𝐢f2(z)f2(z)\displaystyle\frac{(-1+\mathbf{i})f_{1}(z)+\mathbf{i}f_{2}(z)}{f_{2}(z)}
    =\displaystyle= 𝐢πF(14,14,1;z)+B(34,34)1zF(34,34,32,1z)πF(14,14,1,z)\displaystyle\mathbf{i}\cdot\frac{\pi F(\frac{1}{4},\frac{1}{4},1;z)+B(\frac{3}{4},\frac{3}{4})\sqrt{1-z}F(\frac{3}{4},\frac{3}{4},\frac{3}{2},1-z)}{\pi F(\frac{1}{4},\frac{1}{4},1,z)} (3.5)
    =\displaystyle= 𝐢πF(14,14,1;z)+B(14,14)F(14,14,12,1z)πF(14,14,1,z),\displaystyle\mathbf{i}\cdot\frac{-\pi F(\frac{1}{4},\frac{1}{4},1;z)+B(\frac{1}{4},\frac{1}{4})F(\frac{1}{4},\frac{1}{4},\frac{1}{2},1-z)}{\pi F(\frac{1}{4},\frac{1}{4},1,z)},

    for (a,b,c)=(14,14,1)(a,b,c)=(\frac{1}{4},\frac{1}{4},1) is

    τζ(τ)=4ϑ014(τ)ϑ104(τ)ϑ00(τ)8Z{1}={0}.\mathbb{H}\ni\tau\mapsto\zeta(\tau)=\frac{4\vartheta_{01}^{4}(\tau)\vartheta_{10}^{4}(\tau)}{\vartheta_{00}(\tau)^{8}}\in Z\cup\{1\}=\mathbb{C}-\{0\}.

    The function ζ(τ)\zeta(\tau) is invariant under the action τ(g11τ+g12)/(g21τ+g22)\tau\mapsto(g_{11}\tau+g_{12})/(g_{21}\tau+g_{22}) of g=(gij)Γ1,2g=(g_{ij})\in\Gamma_{1,2} on τ\tau\in\mathbb{H}.

Here we give a fundamental region of Γ(2)\Gamma(2) and that of Γ12\Gamma_{12} as

𝔻(2)\displaystyle\mathbb{D}(2) ={τ1<Re(τ)1,|τ12|12,|τ+12|>12},\displaystyle=\{\tau\in\mathbb{H}\mid-1<\mathrm{Re}(\tau)\leq 1,\ |\tau-\dfrac{1}{2}|\geq\dfrac{1}{2},\ |\tau+\dfrac{1}{2}|>\dfrac{1}{2}\}, (3.6)
𝔻12\displaystyle\mathbb{D}_{12} ={τ1<Re(τ)1,|τ|>1}{τ|τ|=1,0Re(τ)1}.\displaystyle=\{\tau\in\mathbb{H}\mid-1<\mathrm{Re}(\tau)\leq 1,\ |\tau|>1\}\cup\{\tau\in\mathbb{H}\mid|\tau|=1,0\leq\mathrm{Re}(\tau)\leq 1\}. (3.7)
Refer to caption
Figure 1. Fundamental regions for Γ(2)\Gamma(2) and Γ12\Gamma_{12}

4. Jacobi’s formula and its analogy

As given in [BB, Theorem 2.1], the hypergeometric series F(12,12,1;x)F(\frac{1}{2},\frac{1}{2},1;x) and the theta constants are related by Jacobi’s formula:

Fact 4.1 (Jacobi’s formula).

Set

λ(τ)=ϑ10(τ)4ϑ00(τ)4\lambda(\tau)=\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}

for any element τ\tau in 𝔻(2)\mathbb{D}(2). Then Jacobi’s formula

F(12,12,1;λ(τ))=ϑ00(τ)2F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau))=\vartheta_{00}(\tau)^{2} (4.1)

holds, where F(12,12,1;z)F(\frac{1}{2},\frac{1}{2},1;z) denotes the single-valued function on {1}\mathbb{C}-\{1\} given in Remark 2.3. By the restriction of Schwarz’s map

zτ=τ(z)=𝐢F(12,12,1;1z)F(12,12,1;z)z\mapsto\tau=\tau(z)=\mathbf{i}\frac{F(\frac{1}{2},\frac{1}{2},1;1-z)}{F(\frac{1}{2},\frac{1}{2},1;z)}

in (3.4) to the domain {z|z|<1,|z1|<1}\{z\in\mathbb{C}\mid|z|<1,\ |z-1|<1\}, the equality (4.1) is pulled back to

F(12,12,1;z)=θ00(τ(z))2.F(\frac{1}{2},\frac{1}{2},1;z)=\theta_{00}(\tau(z))^{2}. (4.2)
Remark 4.2.

Suppose that τ\tau is in the interior 𝔻(2)\mathbb{D}(2)^{\circ} of 𝔻(2)\mathbb{D}(2).

  1. (ii)

    If Re(τ)0\mathrm{Re}(\tau)\geq 0 then Im(λ(τ))0\mathrm{Im}(\lambda(\tau))\geq 0, otherwise Im(λ(τ))<0\mathrm{Im}(\lambda(\tau))<0.

  2. (iiii)

    If τ\tau approaches to a point in the boundary components {τRe(τ)=±1}\{\tau\in\mathbb{H}\mid\mathrm{Re}(\tau)=\pm 1\} of 𝔻(2)\mathbb{D}(2), then λ(τ)\lambda(\tau) converges to a negative real value, at which the hypergeometric series is naturally extended.

  3. (iiiiii)

    If τ\tau approaches to a point in the boundary component {τ|τ12|=12}\{\tau\in\mathbb{H}\mid|\tau-\frac{1}{2}|=\frac{1}{2}\} of 𝔻(2)\mathbb{D}(2), then λ(τ)\lambda(\tau) approaches to a real value greater than 11 via the upper half plane. Note that the addition of this component to 𝔻(2)\mathbb{D}(2) is compatible with the definition of F(a,b,c;z)F(a,b,c;z) on {1}\mathbb{C}-\{1\} in Remark 2.3.

To extend (4.1) to the formula on the whole space \mathbb{H}, we need to consider the continuation of F(12,12,1;λ(τ))F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau)). We prepare a lemma.

Lemma 4.3.

For g=(gij)Γ(2,4)𝐢Jg=(g_{ij})\in\Gamma(2,4)\langle\mathbf{i}J\rangle, let g=(gij)PΓ12g^{\prime}=(g^{\prime}_{ij})\in\mathrm{P}\Gamma_{12} be its normalization of gg in the formula of Fact 3.3. Then we have

g21τ+g22=χ(g)(g21τ+g22).g_{21}\tau+g_{22}=\chi(g^{\prime})\cdot(g^{\prime}_{21}\tau+g^{\prime}_{22}).
Proof..

At first, we consider the case g=(gij)Γ(2,4)g=(g_{ij})\in\Gamma(2,4). In this case, we have g221mod4g_{22}\equiv 1\bmod 4. If g21>0g_{21}>0 then g=gg^{\prime}=g and

χ(g)(g21τ+g22)=𝐢g221(g21τ+g22)=g21τ+g22.\chi(g^{\prime})\cdot(g^{\prime}_{21}\tau+g^{\prime}_{22})=\mathbf{i}^{g_{22}-1}\cdot(g_{21}\tau+g_{22})=g_{21}\tau+g_{22}.

If g21<0g_{21}<0 then g=gg^{\prime}=-g and

χ(g)(g21τ+g22)=𝐢g221(g21τg22)=g21τ+g22.\chi(g^{\prime})\cdot(g^{\prime}_{21}\tau+g^{\prime}_{22})=\mathbf{i}^{-g_{22}-1}\cdot(-g_{21}\tau-g_{22})=g_{21}\tau+g_{22}.

If g21=0g_{21}=0 then g22=|g22|g^{\prime}_{22}=|g_{22}| and

χ(g)(g21τ+g22)=𝐢|g22|1|g22|={𝐢g221g22=g22ifg22>0,𝐢g221(g22)=g22ifg22<0.\chi(g^{\prime})\cdot(g^{\prime}_{21}\tau+g^{\prime}_{22})=\mathbf{i}^{|g_{22}|-1}\cdot|g_{22}|=\left\{\begin{array}[]{ccc}\mathbf{i}^{g_{22}-1}\cdot g_{22}=g_{22}&\textrm{if}&g_{22}>0,\\ \mathbf{i}^{-g_{22}-1}\cdot(-g_{22})=g_{22}&\textrm{if}&g_{22}<0.\end{array}\right.

Next we consider the case g=(gij)(𝐢J)Γ(2,4)g=(g_{ij})\in(\mathbf{i}J)\cdot\Gamma(2,4). In this case, g21g_{21} takes the form of (𝐢)(4m+1)(-\mathbf{i})\cdot(4m+1) (mm\in\mathbb{Z}), and gg^{\prime} is given by a scalar multiplication of ±𝐢\pm\mathbf{i} to gg with g21=|4m+1|g_{21}^{\prime}=|4m+1|. Since they do not vanish, it is sufficient to show χ(g)g21=g12\chi(g^{\prime})\cdot g^{\prime}_{21}=g_{12}. In fact, we have

χ(g)g21=|4m+1|𝐢|4m+1|={(4m+1)𝐢(4m+1)=(𝐢)(4m+1)if4m+1>0,(4m+1)𝐢4m+1=(𝐢)(4m+1)if4m+1<0.\chi(g^{\prime})\cdot g^{\prime}_{21}=|4m+1|\cdot\mathbf{i}^{-|4m+1|}=\left\{\begin{array}[]{rll}(4m+1)\cdot\mathbf{i}^{-(4m+1)}=(-\mathbf{i})\cdot(4m+1)&\textrm{if}&4m+1>0,\\ -(4m+1)\cdot\mathbf{i}^{4m+1}=(-\mathbf{i})\cdot(4m+1)&\textrm{if}&4m+1<0.\end{array}\right.

Thus g21τ+g22=χ(g)(g21τ+g22)g_{21}\tau+g_{22}=\chi(g^{\prime})\cdot(g^{\prime}_{21}\tau+g^{\prime}_{22}) holds in any cases. \square

Corollary 4.4.

We extend Jacobi’s formula (4.1) to the equality on the whole space \mathbb{H} as

ϑ00(τ)2=\displaystyle\vartheta_{00}(\tau)^{2}= Fρ(12,12,1;λ(τ))\displaystyle F_{\rho}(\frac{1}{2},\frac{1}{2},1;\lambda(\tau)) (4.3)
=\displaystyle= (g21τ0+g22)F(12,12,1;λ(τ0))=F(12,12,1;λ(τ))g21τ+g11,\displaystyle(g_{21}\tau_{0}+g_{22})F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau_{0}))=\frac{F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau))}{-g_{21}\tau+g_{11}}, (4.4)

where τ\tau is any element in \mathbb{H}, τ0𝔻(2)\tau_{0}\in\mathbb{D}(2) and g=(gij)Γ(2,4)g=(g_{ij})\in\Gamma(2,4) satisfy

τ=gτ0=g11τ0+g12g21τ0+g22,\tau=g\cdot\tau_{0}=\frac{g_{11}\tau_{0}+g_{12}}{g_{21}\tau_{0}+g_{22}},

ρ\rho is the image of a path from τ0\tau_{0} to τ\tau in \mathbb{H} under the map τλ(τ){0,1}\mathbb{H}\ni\tau\mapsto\lambda(\tau)\in\mathbb{C}-\{0,1\}, F(12,12,1;z)F(\frac{1}{2},\frac{1}{2},1;z) denotes the single-valued function on {1}\mathbb{C}-\{1\} given in Remark 2.3, and Fρ(12,12,1;w)F_{\rho}(\frac{1}{2},\frac{1}{2},1;w) is its analytic continuation along ρ\rho.

Proof..

We obtain the identity (4.3) by the analytic continuation of (4.1) to \mathbb{H}. We also have

ϑ00(τ)2=\displaystyle\vartheta_{00}(\tau)^{2}= ϑ00(gτ0)2=(g21τ0+g22)ϑ00(τ0)2=(g21τ0+g22)F(12,12,1;λ(τ0))\displaystyle\vartheta_{00}(g\cdot\tau_{0})^{2}=(g_{21}\tau_{0}+g_{22})\cdot\vartheta_{00}(\tau_{0})^{2}=(g_{21}\tau_{0}+g_{22})\cdot F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau_{0}))
=\displaystyle= (g21τ0+g22)F(12,12,1;λ(τ))=F(12,12,1;λ(τ))g21τ+g11\displaystyle(g_{21}\tau_{0}+g_{22})\cdot F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau))=\frac{F(\frac{1}{2},\frac{1}{2},1;\lambda(\tau))}{-g_{21}\tau+g_{11}}

by (4.1) and Lemma 4.3. Here note that

τ0=g22τg12g21τ+g11,g21τ0+g22=g21g22τg12g21τ+g11+g22=1g21τ+g11,\tau_{0}=\frac{g_{22}\tau-g_{12}}{-g_{21}\tau+g_{11}},\quad g_{21}\tau_{0}+g_{22}=g_{21}\frac{g_{22}\tau-g_{12}}{-g_{21}\tau+g_{11}}+g_{22}=\frac{1}{-g_{21}\tau+g_{11}},

and that λ(τ)=λ(τ0)\lambda(\tau)=\lambda(\tau_{0}). \square

We give an analogy of Jacobi’s formula (4.1).

Theorem 4.5.

We set

ζ(τ)=4ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8\zeta(\tau)=\frac{4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}}

for any τ\tau in 𝔻12\mathbb{D}_{12}. Then we have

F(14,14,1;ζ(τ))=\displaystyle F\Big{(}\frac{1}{4},\frac{1}{4},1;\zeta(\tau)\Big{)}= ϑ00(τ)2,\displaystyle\vartheta_{00}(\tau)^{2}, (4.5)
F(34,34,1;ζ(τ))=\displaystyle F\Big{(}\frac{3}{4},\frac{3}{4},1;\zeta(\tau)\Big{)}= ϑ00(τ)6ϑ01(τ)4ϑ10(τ)4,\displaystyle\frac{\vartheta_{00}(\tau)^{6}}{\vartheta_{01}(\tau)^{4}-\vartheta_{10}(\tau)^{4}}, (4.6)

where the hypergeometric series is extended to the single-valued function on {1}\mathbb{C}-\{1\} given in Remark 2.3.

Proof..

By substituting a=b=1/2a=b=1/2 into (2.5), we have

F(12,12,1;λ)=F(14,14,1;4λ(1λ)),F(\frac{1}{2},\frac{1}{2},1;\lambda)=F(\frac{1}{4},\frac{1}{4},1;4\lambda(1-\lambda)),

where λ\lambda is sufficiently near to 0. Jacobi’s formula (4.1) together with Jacobi’s identity (3.2) yields that

ϑ00(τ)2\displaystyle\vartheta_{00}(\tau)^{2} =F(12,12,1;ϑ10(τ)4ϑ00(τ)4)=F(14,14,1;4ϑ10(τ)4ϑ00(τ)4(1ϑ10(τ)4ϑ00(τ)4))\displaystyle=F\Big{(}\frac{1}{2},\frac{1}{2},1;\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\Big{)}=F\Big{(}\frac{1}{4},\frac{1}{4},1;4\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\big{(}1-\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}})\Big{)}
=F(14,14,1;4ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8)\displaystyle=F\Big{(}\frac{1}{4},\frac{1}{4},1;\frac{4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}}\Big{)}

for τ\tau with sufficiently large imaginary part. We can extend this identity to any τ𝔻12\tau\in\mathbb{D}_{12} since if τ\tau belongs to the interior of 𝔻12\mathbb{D}_{12} then ζ(τ)\zeta(\tau) is in [1,)\mathbb{C}-[1,\infty) on which F(1/4,1/4,1;z)F(1/4,1/4,1;z) is single valued.

To show the second formula, use (2.6) with the substitution a=b=1/2a=b=1/2. Then we have

ϑ00(τ)2=(12ϑ10(τ)4ϑ00(τ)4)F(34,34,1;ζ(τ)),\vartheta_{00}(\tau)^{2}=\big{(}1-2\frac{\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\big{)}F\Big{(}\frac{3}{4},\frac{3}{4},1;\zeta(\tau)\Big{)},

which is equivalent to (4.6). \square

Remark 4.6.
  1. (ii)

    Substitute τ=𝐢\tau=\mathbf{i} into (4.5). Fact 3.4 implies ϑ01(𝐢)2=ϑ10(𝐢)2\vartheta_{01}(\mathbf{i})^{2}=\vartheta_{10}(\mathbf{i})^{2}, and Jacobi’s identity (3.2) becomes ϑ00(𝐢)4=ϑ01(𝐢)4+ϑ10(𝐢)4=2ϑ10(𝐢)4\vartheta_{00}(\mathbf{i})^{4}=\vartheta_{01}(\mathbf{i})^{4}+\vartheta_{10}(\mathbf{i})^{4}=2\vartheta_{10}(\mathbf{i})^{4}. By Gauss-Kummer’s identity (2.3), we have

    ϑ00(𝐢)2=F(14,14,1;4ϑ01(𝐢)4ϑ10(𝐢)4ϑ00(𝐢)8)=F(14,14,1;1)=Γ(1)Γ(12)Γ(34)Γ(34)=(π1/4Γ(34))2,\vartheta_{00}(\mathbf{i})^{2}=F(\frac{1}{4},\frac{1}{4},1;\frac{4\vartheta_{01}(\mathbf{i})^{4}\vartheta_{10}(\mathbf{i})^{4}}{\vartheta_{00}(\mathbf{i})^{8}})=F(\frac{1}{4},\frac{1}{4},1;1)=\frac{\mathit{\Gamma}(1)\mathit{\Gamma}(\frac{1}{2})}{\mathit{\Gamma}(\frac{3}{4})\mathit{\Gamma}(\frac{3}{4})}=\Big{(}\frac{\pi^{1/4}}{\mathit{\Gamma}(\frac{3}{4})}\Big{)}^{2},\\

    which yields

    ϑ00(𝐢)=π1/4Γ(34)\vartheta_{00}(\mathbf{i})=\frac{\pi^{1/4}}{\mathit{\Gamma}(\frac{3}{4})}

    since ϑ00(𝐢)\vartheta_{00}(\mathbf{i}) is positive real.

  2. (iiii)

    The right hand side of (4.6) has a simple pole at τ=𝐢\tau=\mathbf{i}. Here note that we cannot apply Gauss-Kummer’s identity (2.3) to the left hand side of (4.6) for τ=𝐢\tau=\mathbf{i} since cabc-a-b for (a,b,c)=(34,34,1)(a,b,c)=(\frac{3}{4},\frac{3}{4},1) is 13434=12<01-\frac{3}{4}-\frac{3}{4}=-\frac{1}{2}<0.

We extend Theorem 4.5 to formulas for the whole space \mathbb{H}.

Corollary 4.7.

The equalities (4.5) and (4.6) are extended to

ϑ00(τ)2=Fρ(14,14,1;ζ(τ))=(g21τ0+g22)F(14,14,1;ζ(τ0))=det(g)F(14,14,1;ζ(τ))g21τ+g11,\displaystyle\hskip 51.21504pt\begin{array}[]{ll}\vartheta_{00}(\tau)^{2}&=F_{\rho}(\frac{1}{4},\frac{1}{4},1;\zeta(\tau))\\ &=(g_{21}\tau_{0}+g_{22})F(\frac{1}{4},\frac{1}{4},1;\zeta(\tau_{0}))=\dfrac{\det(g)F(\frac{1}{4},\frac{1}{4},1;\zeta(\tau))}{-g_{21}\tau+g_{11}},\\ \end{array} (4.9)
ϑ00(τ)6ϑ01(τ)4ϑ10(τ)4=Fρ(34,34,1;ζ(τ))=g21τ0+g22det(g)F(34,34,1;ζ(τ0))=F(34,34,1;ζ(τ))g21τ+g11,\displaystyle\begin{array}[]{ll}\dfrac{\vartheta_{00}(\tau)^{6}}{\vartheta_{01}(\tau)^{4}-\vartheta_{10}(\tau)^{4}}&=F_{\rho}(\frac{3}{4},\frac{3}{4},1;\zeta(\tau))\\ &=\dfrac{g_{21}\tau_{0}+g_{22}}{\det(g)}F(\frac{3}{4},\frac{3}{4},1;\zeta(\tau_{0}))=\dfrac{F(\frac{3}{4},\frac{3}{4},1;\zeta(\tau))}{-g_{21}\tau+g_{11}},\end{array} (4.12)

where τ\tau is any element in \mathbb{H}, τ0𝔻12\tau_{0}\in\mathbb{D}_{12} and g=(gij)Γ(2,4)𝐢Jg=(g_{ij})\in\Gamma(2,4)\langle\mathbf{i}J\rangle satisfy

τ=gτ0=g11τ0+g12g21τ0+g22,\tau=g\cdot\tau_{0}=\frac{g_{11}\tau_{0}+g_{12}}{g_{21}\tau_{0}+g_{22}},

ρ\rho is the image of a path τ0\tau_{0} to τ\tau in \mathbb{H} under the map τζ(τ){1}\mathbb{H}\ni\tau\mapsto\zeta(\tau)\in\mathbb{C}-\{1\}, F(a,b,c;z)F(a,b,c;z) is extended to the single-valued function on {1}\mathbb{C}-\{1\} given in Remark 2.3, and Fρ(a,b,c;w)F_{\rho}(a,b,c;w) is its analytic continuation along ρ\rho.

Proof..

We can show

ϑ00(τ)2=Fρ(14,14,1;ζ(τ))=(g21τ0+g22)F(14,14,1;ζ(τ0))\vartheta_{00}(\tau)^{2}=F_{\rho}(\frac{1}{4},\frac{1}{4},1;\zeta(\tau))=(g_{21}\tau_{0}+g_{22})F(\frac{1}{4},\frac{1}{4},1;\zeta(\tau_{0}))

quite similarly to Proof of Corollary 4.4. Since

det(g)={1ifgΓ(2,4),1ifg(𝐢J)Γ(2,4),\det(g)=\left\{\begin{array}[]{rcl}1&\textrm{if}&g\in\Gamma(2,4),\\ -1&\textrm{if}&g\in(\mathbf{i}J)\cdot\Gamma(2,4),\end{array}\right.

we have

g21τ0+g22=1det(g)(g21τ+g11)g_{21}\tau_{0}+g_{22}=\frac{1}{\det(g)(-g_{21}\tau+g_{11})}

for any gΓ(2,4)𝐢Jg\in\Gamma(2,4)\langle\mathbf{i}J\rangle, which leads to the last equality in (4.9).

To show (4.12), we have only to note that

ϑ00(τ)6ϑ01(τ)4ϑ10(τ)4\displaystyle\frac{\vartheta_{00}(\tau)^{6}}{\vartheta_{01}(\tau)^{4}-\vartheta_{10}(\tau)^{4}} =ϑ00(gτ0)2ϑ00(gτ0)4ϑ01(gτ0)4ϑ10(gτ0)4,\displaystyle=\vartheta_{00}(g\cdot\tau_{0})^{2}\frac{\vartheta_{00}(g\cdot\tau_{0})^{4}}{\vartheta_{01}(g\cdot\tau_{0})^{4}-\vartheta_{10}(g\cdot\tau_{0})^{4}},
ϑ00(gτ0)2\displaystyle\vartheta_{00}(g\cdot\tau_{0})^{2} =(g21τ0+g22)ϑ00(τ0)2,\displaystyle=(g_{21}\tau_{0}+g_{22})\vartheta_{00}(\tau_{0})^{2},
ϑ00(gτ0)4ϑ01(gτ0)4ϑ10(gτ0)4\displaystyle\frac{\vartheta_{00}(g\cdot\tau_{0})^{4}}{\vartheta_{01}(g\cdot\tau_{0})^{4}-\vartheta_{10}(g\cdot\tau_{0})^{4}} ={ϑ00(τ0)4ϑ01(τ0)4ϑ10(τ0)4ifgΓ(2,4),ϑ00(τ0)4ϑ10(τ0)4ϑ01(τ0)4ifg(𝐢J)Γ(2,4),\displaystyle=\left\{\begin{array}[]{ccl}\dfrac{\vartheta_{00}(\tau_{0})^{4}}{\vartheta_{01}(\tau_{0})^{4}-\vartheta_{10}(\tau_{0})^{4}}&\textrm{if}&g\in\Gamma(2,4),\\[11.38109pt] \dfrac{\vartheta_{00}(\tau_{0})^{4}}{\vartheta_{10}(\tau_{0})^{4}-\vartheta_{01}(\tau_{0})^{4}}&\textrm{if}&g\in(\mathbf{i}J)\cdot\Gamma(2,4),\\ \end{array}\right.

by (3.2) and Facts 3.3, 3.4, 3.5 (ii). \square

By regarding τ\tau as a dependent variable of zz in the domain {z|z|<1,|z1|<1}\{z\in\mathbb{C}\mid|z|<1,\ |z-1|<1\} by Schwarz’s map (3.5), we rewrite Theorem 4.5 into the following.

Corollary 4.8.

By the restriction of Schwarz’s map zτ(z)z\mapsto\tau(z) in (3.5) to the domain {z|z|<1,|z1|<1}\{z\in\mathbb{C}\mid|z|<1,\ |z-1|<1\}, the equalities (4.5), (4.6) are pulled back to

F(14,14,1;z)=ϑ00(τ(z))2,F(34,34,1;z)=ϑ00(τ(z))6ϑ01(τ(z))4ϑ10(τ(z))4.F(\frac{1}{4},\frac{1}{4},1;z)=\vartheta_{00}(\tau(z))^{2},\quad F(\frac{3}{4},\frac{3}{4},1;z)=\frac{\vartheta_{00}(\tau(z))^{6}}{\vartheta_{01}(\tau(z))^{4}-\vartheta_{10}(\tau(z))^{4}}.

5. A transformation formula for F(14,14,1;z)F(\frac{1}{4},\frac{1}{4},1;z).

Theorem 5.1.

We have a transformation formula

2+2+2w4F(14,14,1;1w2)=F(14,14,1;1(62w+2w3)2(22w+2+w+3)2),\frac{2+\sqrt{2+2w}}{4}F(\frac{1}{4},\frac{1}{4},1;1-w^{2})=F\Big{(}\frac{1}{4},\frac{1}{4},1;1-\frac{(6\sqrt{2w+2}-w-3)^{2}}{(2\sqrt{2w+2}+w+3)^{2}}\Big{)}, (5.1)

where ww is in a neighborhood of 11, and 2+2w\sqrt{2+2w} takes 22 at w=1w=1.

Proof..

By substituting 2τ2\tau into τ\tau for (4.5), and using (3.3), we have

F(14,14,1;ζ(2τ))=ϑ00(2τ)2=ϑ00(τ)2+ϑ01(τ)22=ϑ00(τ)2+ϑ01(τ)22ϑ00(τ)2ϑ00(τ)2,F(\frac{1}{4},\frac{1}{4},1;\zeta(2\tau))=\vartheta_{00}(2\tau)^{2}=\frac{\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2}}{2}=\frac{\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2}}{2\vartheta_{00}(\tau)^{2}}\cdot\vartheta_{00}(\tau)^{2}, (5.2)

where we restrict τ\tau\in\mathbb{H} to pure imaginary numbers with sufficiently large imaginary part. Note that

ζ(τ)\displaystyle\zeta(\tau) =4ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8=1(ϑ01(τ)4+ϑ10(τ)4)24ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8\displaystyle=\frac{4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}}=1-\frac{(\vartheta_{01}(\tau)^{4}+\vartheta_{10}(\tau)^{4})^{2}-4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}}
=1(ϑ01(τ)4ϑ10(τ)4)2ϑ00(τ)8=1(2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)4)2.\displaystyle=1-\frac{(\vartheta_{01}(\tau)^{4}-\vartheta_{10}(\tau)^{4})^{2}}{\vartheta_{00}(\tau)^{8}}=1-\Big{(}\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\Big{)}^{2}.

Put

w=2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)4,w=\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}},

then we have

ϑ01(τ)4=w+12ϑ00(τ)4,ϑ01(τ)2=2w+22ϑ00(τ)2,\vartheta_{01}(\tau)^{4}=\frac{w+1}{2}\cdot\vartheta_{00}(\tau)^{4},\quad\vartheta_{01}(\tau)^{2}=\frac{\sqrt{2w+2}}{2}\cdot\vartheta_{00}(\tau)^{2},

the last term of (5.2) is

ϑ00(τ)2+ϑ01(τ)22ϑ00(τ)2ϑ00(τ)2=2+2w+24F(14,14,1;1w2),\frac{\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2}}{2\vartheta_{00}(\tau)^{2}}\cdot\vartheta_{00}(\tau)^{2}=\frac{2+\sqrt{2w+2}}{4}\cdot F(\frac{1}{4},\frac{1}{4},1;1-w^{2}),

which is the left hand side of (5.1). Here note that ϑ01(τ)>0\vartheta_{01}(\tau)>0 for any pure imaginary number τ\tau.

Let us express ζ(2τ)\zeta(2\tau) in terms of ww. We have

ζ(2τ)\displaystyle\zeta(2\tau) =1(2ϑ01(2τ)4ϑ00(2τ)4ϑ00(2τ)4)2=1(8ϑ00(τ)2ϑ01(τ)2(ϑ00(τ)2+ϑ01(τ)2)2ϑ00(τ)2+ϑ01(τ)2)2\displaystyle=1-\Big{(}\frac{2\vartheta_{01}(2\tau)^{4}-\vartheta_{00}(2\tau)^{4}}{\vartheta_{00}(2\tau)^{4}}\Big{)}^{2}=1-\Big{(}\frac{8\vartheta_{00}(\tau)^{2}\vartheta_{01}(\tau)^{2}-(\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2})^{2}}{\vartheta_{00}(\tau)^{2}+\vartheta_{01}(\tau)^{2}}\Big{)}^{2}
=1(162w+2(2+2w+2)2(2+2w+2)2)2=1(62w+2w322w+2+w+3)2.\displaystyle=1-\Big{(}\frac{16\sqrt{2w+2}-(2+\sqrt{2w+2})^{2}}{(2+\sqrt{2w+2})^{2}}\Big{)}^{2}=1-\Big{(}\frac{6\sqrt{2w+2}-w-3}{2\sqrt{2w+2}+w+3}\Big{)}^{2}.

Hence F(14,14,1;ζ(2τ))F(\frac{1}{4},\frac{1}{4},1;\zeta(2\tau)) in (5.2) is equal to the right hand side of (5.1). \square

6. Mean iterations

By referring to [BB, Chapter 8], we give fundamental properties of mean iterations.

Definition 6.1 (Means).

A mean on a subset 𝕊\mathbb{S} in 2\mathbb{R}^{2} is a continuous function mm defined on 𝕊\mathbb{S} satisfying

min(x,y)m(x,y)max(x,y)\min(x,y)\leq m(x,y)\leq\max(x,y) (6.1)

for any (x,y)𝕊(x,y)\in\mathbb{S}.
A mean m(x,y)m(x,y) is strict if it satisfies

m(x,y)=x or m(x,y)=yx=y.m(x,y)=x\textrm{ or }m(x,y)=y\quad\Leftrightarrow\quad x=y. (6.2)

A mean m(x,y)m(x,y) is homogeneous if 𝕊\mathbb{S} and mm satisfy

(rx,ry)𝕊,m(rx,ry)=rm(x,y)(r\cdot x,r\cdot y)\in\mathbb{S},\quad m(r\cdot x,r\cdot y)=r\cdot m(x,y) (6.3)

for any (x,y)𝕊(x,y)\in\mathbb{S} and any r+={xx>0}r\in\mathbb{R}_{+}=\{x\in\mathbb{R}\mid x>0\}.
A mean m(x,y)m(x,y) is symmetric if 𝕊\mathbb{S} and mm satisfy

(y,x)𝕊,m(y,x)=m(x,y)(y,x)\in\mathbb{S},\quad m(y,x)=m(x,y) (6.4)

for any (x,y)𝕊(x,y)\in\mathbb{S}.

We set the arithmetic mean on 2\mathbb{R}^{2}, the geometric mean on +2\mathbb{R}_{+}^{2}, and the harmonic mean on +2\mathbb{R}_{+}^{2} by

mA(x,y)=x+y2,mG(x,y)=xy,mH(x,y)=2xyx+y,m_{A}(x,y)=\frac{x+y}{2},\quad m_{G}(x,y)=\sqrt{xy},\quad m_{H}(x,y)=\frac{2xy}{x+y},

respectively. They are strict, homogeneous and symmetric means, and they satisfy

mA(x,y)mG(x,y)mH(x,y)m_{A}(x,y)\geq m_{G}(x,y)\geq m_{H}(x,y)

on +2\mathbb{R}_{+}^{2}.

We modify [BB, Proposition 8.4] as follows.

Lemma 6.2.

Let m0m_{0}, m1m_{1} and m2m_{2} be means on 𝕊0\mathbb{S}_{0}, 𝕊1\mathbb{S}_{1} and 𝕊2\mathbb{S}_{2} satisfying

(m1(x,y),m2(x,y))𝕊0(m_{1}(x,y),m_{2}(x,y))\in\mathbb{S}_{0}

for any (x,y)𝕊1𝕊2.(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}. Then the function

m(x,y)=m0(m1(x,y),m2(x,y))m(x,y)=m_{0}(m_{1}(x,y),m_{2}(x,y))

is a mean on 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2}. If two of m0,m1,m2m_{0},m_{1},m_{2} are strict, then so is mm. If m1m_{1} and m2m_{2} are symmetric, then so is mm. If all three are homogeneous, then so is mm.

Proof..

Note that the function mm is defined on 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2}, and continuous on it. Since m0,m1,m2m_{0},m_{1},m_{2} are means, they satisfy

min(m1(x,y),m2(x,y))m0(m1(x,y),m2(x,y))max(m1(x,y),m2(x,y)),\min(m_{1}(x,y),m_{2}(x,y))\leq m_{0}(m_{1}(x,y),m_{2}(x,y))\leq\max(m_{1}(x,y),m_{2}(x,y)), (6.5)
min(x,y)m1(x,y)max(x,y),min(x,y)m2(x,y)max(x,y).\min(x,y)\leq m_{1}(x,y)\leq\max(x,y),\quad\min(x,y)\leq m_{2}(x,y)\leq\max(x,y). (6.6)

Hence we have

min(x,y)m(x,y)max(x,y),\min(x,y)\leq m(x,y)\leq\max(x,y), (6.7)

which show that mm is a mean on 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2}.

Suppose that m1m_{1} and m2m_{2} are strict. If xyx\neq y then the inequalities in (6.6) are strict, which imply that the inequalities in (6.7) become strict. Suppose that m0m_{0} and mim_{i} (i=1i=1 or i=2i=2) are strict. If (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2} satisfies xyx\neq y and m1(x,y)m2(x,y)m_{1}(x,y)\neq m_{2}(x,y), then the inequalities in (6.5) become strict, since m0m_{0} is strict. Thus in this case, the inequalities in (6.7) become strict. If (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2} satisfies xyx\neq y and m1(x,y)=m2(x,y)m_{1}(x,y)=m_{2}(x,y) then

m(x,y)=m0(m1(x,y),m2(x,y))=m0(mi(x,y),mi(x,y))=mi(x,y).m(x,y)=m_{0}(m_{1}(x,y),m_{2}(x,y))=m_{0}(m_{i}(x,y),m_{i}(x,y))=m_{i}(x,y).

In this case, the inequalities in (6.7) become strict, since mim_{i} is strict. Hence it turns out that mm is strict if two of m0,m1,m2m_{0},m_{1},m_{2} are strict.

If each of mim_{i} (i=1,2i=1,2) is a symmetric mean on 𝕊i\mathbb{S}_{i} then we have

(x,y)𝕊1𝕊2(y,x)𝕊1𝕊2,(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}\Rightarrow(y,x)\in\mathbb{S}_{1}\cap\mathbb{S}_{2},
m(y,x)=m0(m1(y,x),m2(y,x))=m0(m1(x,y),m2(x,y))=m(x,y),m(y,x)=m_{0}(m_{1}(y,x),m_{2}(y,x))=m_{0}(m_{1}(x,y),m_{2}(x,y))=m(x,y),

for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}, which show that mm is symmetric.

If each of mim_{i} (i=0,1,2i=0,1,2) is a homogeneous mean on 𝕊i\mathbb{S}_{i} then we have

(x,y)\displaystyle(x,y) 𝕊1𝕊2(rx,ry)𝕊1𝕊2,\displaystyle\in\mathbb{S}_{1}\cap\mathbb{S}_{2}\Rightarrow(r\cdot x,r\cdot y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2},
m(rx,ry)\displaystyle m(r\cdot x,r\cdot y) =m0(m1(rx,ry),m2(rx,ry))=m0(rm1(x,y),rm2(x,y))\displaystyle=m_{0}(m_{1}(r\cdot x,r\cdot y),m_{2}(r\cdot x,r\cdot y))=m_{0}(r\cdot m_{1}(x,y),r\cdot m_{2}(x,y))
=rm0(m1(x,y),m2(x,y))=rm(x,y),\displaystyle=r\cdot m_{0}(m_{1}(x,y),m_{2}(x,y))=r\cdot m(x,y),

for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2} and any r+r\in\mathbb{R}_{+}, which show that mm is homogeneous. \square

Definition 6.3 (A mean iteration).

Let m1m_{1} and m2m_{2} be means on 𝕊1\mathbb{S}_{1} and 𝕊2\mathbb{S}_{2} satisfying (m1(x,y),m2(x,y))𝕊1𝕊2(m_{1}(x,y),m_{2}(x,y))\in\mathbb{S}_{1}\cap\mathbb{S}_{2} for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}. For any fixed (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}, we have a pair of sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} by setting (x0,y0)=(x,y)(x_{0},y_{0})=(x,y) and iterating the application of two means

(xn+1,yn+1)=(m1(xn,yn),m2(xn,yn))(n0={0})(x_{n+1},y_{n+1})=(m_{1}(x_{n},y_{n}),m_{2}(x_{n},y_{n}))\quad(n\in\mathbb{N}_{0}=\mathbb{N}\cup\{0\}) (6.8)

to the previous terms. This construction of a pair of sequences is called a mean iteration.

Lemma 6.4 ([BB, Theorem 8.2]).

Let m1m_{1} and m2m_{2} be means on 𝕊1\mathbb{S}_{1} and 𝕊2\mathbb{S}_{2} satisfying (m1(x,y),m2(x,y))𝕊1𝕊2(m_{1}(x,y),m_{2}(x,y))\in\mathbb{S}_{1}\cap\mathbb{S}_{2} for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}. Suppose that the restrictions of two means m1m_{1} and m2m_{2} to 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2} are strict, and satisfy one of the following

  1. (1)(1)

    m1(x,y)m2(x,y)m_{1}(x,y)\geq m_{2}(x,y) for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2};

  2. (2)(2)

    m1(x,y)m2(x,y)m_{1}(x,y)\leq m_{2}(x,y) for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2};

  3. (3)(3)

    m1(x,y)m2(x,y)m_{1}(x,y)\leq m_{2}(x,y) for xyx\leq y, (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}, and m2(x,y)m1(x,y)m_{2}(x,y)\leq m_{1}(x,y) for yxy\leq x, (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}.

Then each of sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} defined by (6.8) converges monotonously and uniformly on any compact set in 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2}, and satisfies

limnxn=limnyn.\lim_{n\to\infty}x_{n}=\lim_{n\to\infty}y_{n}.

The function (x,y)limnxn(=limnyn)\displaystyle{(x,y)\mapsto\lim_{n\to\infty}x_{n}(=\lim_{n\to\infty}y_{n})} is a strict mean on 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2}, which is called the compound mean of m1m_{1} and m2m_{2} and denoted by m1m2m_{1}\diamond m_{2}. Moreover, m1m2m_{1}\diamond m_{2} is homogeneous or symmetric if each of m1m_{1} and m2m_{2} is so.

Lemma 6.5 (Invariant Principle [BB, Theorem 8.3]).

Suppose that the compound mean m1m2m_{1}\diamond m_{2} on 𝕊1𝕊2\mathbb{S}_{1}\cap\mathbb{S}_{2} exists for given two means m1m_{1} on 𝕊1\mathbb{S}_{1} and m2m_{2} on 𝕊2\mathbb{S}_{2}, which satisfy (m1(x,y),m2(x,y))𝕊1𝕊2(m_{1}(x,y),m_{2}(x,y))\in\mathbb{S}_{1}\cap\mathbb{S}_{2} for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}. Then it is uniquely characterized as a continuous function ψ:𝕊1𝕊2\psi:\mathbb{S}_{1}\cap\mathbb{S}_{2}\to\mathbb{R} such that

ψ(m1(x,y),m2(x,y))=ψ(x,y)\psi(m_{1}(x,y),m_{2}(x,y))=\psi(x,y)

for any (x,y)𝕊1𝕊2(x,y)\in\mathbb{S}_{1}\cap\mathbb{S}_{2}.

Example 6.6 (The arithmetic-geometric mean).

The arithmetic-geometric mean mAGm_{AG} is defined by the compound mean mAmGm_{A}\diamond m_{G} of the arithmetic mean m1=mAm_{1}=m_{A} on 2\mathbb{R}^{2} and the geometric mean m2=mGm_{2}=m_{G} on +2\mathbb{R}_{+}^{2}; that is, set (x0,y0)=(x,y)+2=2+2(x_{0},y_{0})=(x,y)\in\mathbb{R}_{+}^{2}=\mathbb{R}^{2}\cap\mathbb{R}_{+}^{2}, define a pair of sequences by

xn+1=mA(xn,yn)=xn+yn2,yn+1=mG(xn,yn)=xnyn,x_{n+1}=m_{A}(x_{n},y_{n})=\frac{x_{n}+y_{n}}{2},\quad y_{n+1}=m_{G}(x_{n},y_{n})=\sqrt{x_{n}y_{n}},

and take the limit mAG(x,y)=limnxn(=limnyn)\displaystyle{m_{AG}(x,y)=\lim_{n\to\infty}x_{n}(=\lim_{n\to\infty}y_{n})}.

By putting a=b=12a=b=\frac{1}{2}, z=w1w+1z=\frac{w-1}{w+1} for (2.4), we have

F(12,12,1;14w(1+w)2)=1+w2F(12,12,1;1w2),F(\frac{1}{2},\frac{1}{2},1;1-\frac{4w}{(1+w)^{2}})=\frac{1+w}{2}F(\frac{1}{2},\frac{1}{2},1;1-w^{2}), (6.9)

where ww is in a small neighborhood UU of 11\in\mathbb{C}. Substitute w=yxw=\frac{y}{x} into this formula, then we have

xF(12,12,1;1y2x2)=mA(x,y)F(12,12,1;1mG(x,y)2mA(x,y)2),\frac{x}{F(\frac{1}{2},\frac{1}{2},1;1-\frac{y^{2}}{x^{2}})}=\frac{m_{A}(x,y)}{F(\frac{1}{2},\frac{1}{2},1;1-\frac{m_{G}(x,y)^{2}}{m_{A}(x,y)^{2}})},

which implies

mAG(x,y)=xF(12,12,1;1y2x2)m_{AG}(x,y)=\frac{x}{F(\frac{1}{2},\frac{1}{2},1;1-\frac{y^{2}}{x^{2}})} (6.10)

by Lemma 6.5. Since (6.9) is valid on UU, (6.10) holds initially for x,yx,y close to each other. Note that the set {1y2/x2(x,y)+2}\{1-y^{2}/x^{2}\mid(x,y)\in\mathbb{R}_{+}^{2}\} coincides with the open interval (,1)(-\infty,1), on which F(12,12,1;z)F(\frac{1}{2},\frac{1}{2},1;z) admits the unique analytic continuation satisfying (6.9) as in Remark 2.3. By regarding the right hand side of (6.10) as its analytic continuation, we can extend (6.10) to a formula on +2\mathbb{R}_{+}^{2}.

Recall the formula (5.1):

2+2+2w4F(14,14,1;1w2)=F(14,14,1;1(62w+2w3)2(22w+2+w+3)2).\frac{2+\sqrt{2+2w}}{4}F(\frac{1}{4},\frac{1}{4},1;1-w^{2})=F\Big{(}\frac{1}{4},\frac{1}{4},1;1-\frac{(6\sqrt{2w+2}-w-3)^{2}}{(2\sqrt{2w+2}+w+3)^{2}}\Big{)}.

By taking the factor in the left hand side of (5.1), we set

μ1(w)=2+2+2w4.\mu_{1}(w)=\frac{2+\sqrt{2+2w}}{4}.

We also set

μ2(w)=62+2ww32(2+2+2w)\mu_{2}(w)=\frac{6\sqrt{2+2w}-w-3}{2(2+\sqrt{2+2w})}

so that

μ2(w)2μ1(w)2=(62w+2w3)2(22w+2+w+3)2,\frac{\mu_{2}(w)^{2}}{\mu_{1}(w)^{2}}=\frac{(6\sqrt{2w+2}-w-3)^{2}}{(2\sqrt{2w+2}+w+3)^{2}},

which is in the right hand side of (5.1). We extend them to homogeneous functions μi(x,y)\mu_{i}(x,y) (i=1,2)(i=1,2) of degree 11 defined on +2\mathbb{R}_{+}^{2} by setting xμi(y/x)x\cdot\mu_{i}(y/x) (i=1,2)(i=1,2). They are

μ1(x,y)=2x+2x(x+y)4,μ2(x,y)=62x(x+y)y3x2(2x+2x(x+y))x.\mu_{1}(x,y)=\frac{2x+\sqrt{2x(x+y)}}{4},\quad\mu_{2}(x,y)=\frac{6\sqrt{2x(x+y)}-y-3x}{2(2x+\sqrt{2x(x+y)})}\cdot x. (6.11)

To understand μ1\mu_{1} and μ2\mu_{2} well, we prepare functions

μ0(x,y)\displaystyle\mu_{0}(x,y) =mG(x,mA(x,y))=xx+y2,\displaystyle=m_{G}(x,m_{A}(x,y))=\sqrt{x\cdot\frac{x+y}{2}},
ν(x,y)\displaystyle\nu(x,y) =2mH(x,y)mA(x,y)=4xyx+yx+y2=6xyx2y22(x+y),\displaystyle=2m_{H}(x,y)-m_{A}(x,y)=\frac{4xy}{x+y}-\frac{x+y}{2}=\frac{6xy-x^{2}-y^{2}}{2(x+y)},

defined on 𝕊(1,)={(x,y)2x>0,x+y>0}\mathbb{S}_{(-1,\infty)}=\{(x,y)\in\mathbb{R}^{2}\mid x>0,\ x+y>0\} and on +2\mathbb{R}_{+}^{2}. Here we introduce a notation

𝕊I={(x,rx)2x+,rI}\mathbb{S}_{I}=\{(x,rx)\in\mathbb{R}^{2}\mid x\in\mathbb{R}_{+},\ r\in I\}

for an interval II in \mathbb{R}. For examples, 𝕊(0,)=+2\mathbb{S}_{(0,\infty)}=\mathbb{R}_{+}^{2} and

𝕊(1,1]={(x,rx)2x+,r(1,1]}={(x,y)2x+,x<yx}.\mathbb{S}_{(-1,1]}=\{(x,rx)\in\mathbb{R}^{2}\mid x\in\mathbb{R}_{+},\ r\in(-1,1]\}=\{(x,y)\in\mathbb{R}^{2}\mid x\in\mathbb{R}_{+},\ -x<y\leq x\}.
Lemma 6.7.

The function μ0(x,y)\mu_{0}(x,y) is a positive-valued strict homogeneous mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)}. It satisfies

0<y<x\displaystyle 0<y<x\quad x2<μ0(x,y)<x,\displaystyle\Rightarrow\quad\frac{x}{\sqrt{2}}<\mu_{0}(x,y)<x, (6.12)
y<x\displaystyle y<x\quad 0<xμ0(x,y)<μ0(x,y)y.\displaystyle\Rightarrow\quad 0<x-\mu_{0}(x,y)<\mu_{0}(x,y)-y. (6.13)
Proof..

Regard the projection prx:(x,y)xpr_{x}:(x,y)\mapsto x as a homogeneous mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)} and restrict the arithmetic mean mAm_{A} to 𝕊(1,)\mathbb{S}_{(-1,\infty)}. Then they satisfy (prx(x,y),mA(x,y))+2(pr_{x}(x,y),m_{A}(x,y))\in\mathbb{R}_{+}^{2} for any (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)}. Since mAm_{A} and mGm_{G} are strict and homogeneous on +2\mathbb{R}_{+}^{2}, μ0(x,y)=mG(prx(x,y),mA(x,y))\mu_{0}(x,y)=m_{G}(pr_{x}(x,y),m_{A}(x,y)) is a positive-valued strict homogeneous mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)} by Lemma 6.2.

If 0<y<x0<y<x then

x>μ0(x,y)=mG(x,mA(x,y))>mG(x,mA(x,0))=mG(x,x2)=x2.x>\mu_{0}(x,y)=m_{G}(x,m_{A}(x,y))>m_{G}(x,m_{A}(x,0))=m_{G}(x,\frac{x}{2})=\frac{x}{\sqrt{2}}.

If y<xy<x then

y<mA(x,y)=x+y2<μ0(x,y)<x,y<m_{A}(x,y)=\frac{x+y}{2}<\mu_{0}(x,y)<x,

which yield that

0<xμ0(x,y)<xmA(x,y)=mA(x,y)y<μ0(x,y)y0<x-\mu_{0}(x,y)<x-m_{A}(x,y)=m_{A}(x,y)-y<\mu_{0}(x,y)-y

even in the case y<0y<0. \square

Lemma 6.8.

The restriction of ν\nu to 𝕊[1/3,3]={(x,y)+2x3y3x}\mathbb{S}_{[1/3,3]}=\{(x,y)\in\mathbb{R}_{+}^{2}\mid\dfrac{x}{3}\leq y\leq 3x\} is a homogeneous symmetric mean, and that to 𝕊(1/3,3)={(x,y)+2x3<y<3x}\mathbb{S}_{(1/3,3)}=\{(x,y)\in\mathbb{R}_{+}^{2}\mid\dfrac{x}{3}<y<3x\} becomes strict.

Proof..

Since mAm_{A} and mHm_{H} are homogeneous and symmetric, ν\nu satisfies

ν(rx,ry)=rν(x,y),ν(y,x)=ν(x,y)\nu(r\cdot x,r\cdot y)=r\cdot\nu(x,y),\quad\nu(y,x)=\nu(x,y)

for any (x,y)+2(x,y)\in\mathbb{R}_{+}^{2} and any r+r\in\mathbb{R}_{+}. Suppose that (x,y)𝕊[1/3,3](x,y)\in\mathbb{S}_{[1/3,3]}. If x<yx<y then

ν(x,y)x\displaystyle\nu(x,y)-x =4xyx+yx+y2x=(3xy)(yx)2(x+y)0,\displaystyle=\frac{4xy}{x+y}-\frac{x+y}{2}-x=\frac{(3x-y)(y-x)}{2(x+y)}\geq 0,
yν(x,y)\displaystyle y-\nu(x,y) =y4xyx+y+x+y2=(3yx)(yx)2(x+y)0,\displaystyle=y-\frac{4xy}{x+y}+\frac{x+y}{2}=\frac{(3y-x)(y-x)}{2(x+y)}\geq 0,

and if x>yx>y then xν(x,y)0x-\nu(x,y)\geq 0 and ν(x,y)y0\nu(x,y)-y\geq 0, which show that the restriction of ν(x,y)\nu(x,y) to 𝕊[1/3,3]\mathbb{S}_{[1/3,3]} is a homogeneous symmetric mean. By these inequalities, we can also see that the restriction of ν\nu to 𝕊(1/3,3)\mathbb{S}_{(1/3,3)} becomes strict. \square

Remark 6.9.

The function ν(x,y)=2mH(x,y)mA(x,y)\nu(x,y)=2m_{H}(x,y)-m_{A}(x,y) on +2\mathbb{R}_{+}^{2} takes negative values. In fact, we have

y>(3+22)x or y<(322)xν(x,y)<0.y>(3+2\sqrt{2})x\textrm{ or }y<(3-2\sqrt{2})x\quad\Rightarrow\quad\nu(x,y)<0.
Proposition 6.10.
  1. ((i)\,)

    We can express the functions μ1\mu_{1} and μ2\mu_{2} by μ0\mu_{0} and ν\nu as

    μ1(x,y)=mA(x,μ0(x,y)),μ2(x,y)=ν(x,μ0(x,y));\mu_{1}(x,y)=m_{A}(x,\mu_{0}(x,y)),\quad\mu_{2}(x,y)=\nu(x,\mu_{0}(x,y)); (6.14)

    we can regard the functions μ1(x,y)\mu_{1}(x,y) and μ2(x,y)\mu_{2}(x,y) as defined on

    𝕊(1,)={(x,y)2x>0,x+y>0}.\mathbb{S}_{(-1,\infty)}=\{(x,y)\in\mathbb{R}^{2}\mid x>0,\ x+y>0\}.
  2. ((ii)\,)

    The function μ1\mu_{1} is a strict homogeneous mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)}. The function μ2\mu_{2} satisfies an inequality

    μ2(x,y)y\mu_{2}(x,y)\geq y (6.15)

    for any (x,y)𝕊(1,1]={(x,y)2x>0,x<yx}(x,y)\in\mathbb{S}_{(-1,1]}=\{(x,y)\in\mathbb{R}^{2}\mid x>0,\ -x<y\leq x\}. The restriction of μ2\mu_{2} to the domain

    𝕊(0,17)={(x,y)+2y<17x}\mathbb{S}_{(0,17)}=\{(x,y)\in\mathbb{R}_{+}^{2}\mid y<17x\}

    is a strict homogeneous mean.

  3. ((iii)\,)

    The functions μ1\mu_{1} and μ2\mu_{2} on 𝕊(1,)\mathbb{S}_{(-1,\infty)} satisfy inequalities

    μ1(x,y)μ2(x,y)0,\displaystyle\mu_{1}(x,y)-\mu_{2}(x,y)\geq 0, (6.16)
    μ1(x,y)+μ2(x,y)>0,\displaystyle\mu_{1}(x,y)+\mu_{2}(x,y)>0, (6.17)

    for any (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)}. If xyx\neq y then (6.16) holds strictly.
    If (x,y)𝕊(1,1](x,y)\in\mathbb{S}_{(-1,1]} then

    μ1(x,y)+μ2(x,y)x+y(>0).\mu_{1}(x,y)+\mu_{2}(x,y)\geq x+y\ (>0). (6.18)
Proof..

(i)(\ref{item:expr}) By straightforward calculations, we have

mA(x,μ0(x,y))=x+x(x+y)/22=2x+2x(x+y)4=μ1(x,y),\displaystyle m_{A}(x,\mu_{0}(x,y))=\frac{x+\sqrt{x\cdot(x+y)/2}}{2}=\frac{2x+\sqrt{2x(x+y)}}{4}=\mu_{1}(x,y),
ν(x,μ0(x,y))=2mH(x,μ0(x,y))mA(x,μ0(x,y))\displaystyle\nu(x,\mu_{0}(x,y))=2m_{H}(x,\mu_{0}(x,y))-m_{A}(x,\mu_{0}(x,y))
=\displaystyle= 4xx(x+y)/2x+x(x+y)/22x+2x(x+y)4=16x2x(x+y)(2x+2x(x+y))24(2x+2x(x+y))\displaystyle\frac{4x\sqrt{x\cdot(x+y)/2}}{x+\sqrt{x\cdot(x+y)/2}}-\frac{2x+\sqrt{2x(x+y)}}{4}=\frac{16x\sqrt{2x(x+y)}-(2x+\sqrt{2x(x+y)})^{2}}{4(2x+\sqrt{2x(x+y)})}
=\displaystyle= 3x2+6x2x(x+y)xy2(2x+2x(x+y))=μ2(x,y).\displaystyle\frac{-3x^{2}+6x\sqrt{2x(x+y)}-xy}{2(2x+\sqrt{2x(x+y)})}=\mu_{2}(x,y).

We can extend them to functions on 𝕊(1,)\mathbb{S}_{(-1,\infty)} since μ0(x,y)\mu_{0}(x,y) is defined on it, and takes positive values.

(ii)(\ref{item:means}) Since mAm_{A} and μ0\mu_{0} are strict homogeneous means on 2\mathbb{R}^{2} and 𝕊(1,)\mathbb{S}_{(-1,\infty)}, and prx:(x,y)xpr_{x}:(x,y)\to x is a homogeneous mean on 2\mathbb{R}^{2}, μ1(x,y)=mA(prx(x,y),μ0(x,y))\mu_{1}(x,y)=m_{A}(pr_{x}(x,y),\mu_{0}(x,y)) is a strict homogeneous mean on 𝕊(1,)=2𝕊(1,)\mathbb{S}_{(-1,\infty)}=\mathbb{R}^{2}\cap\mathbb{S}_{(-1,\infty)} by Lemma 6.2.

Under the assumption (x,y)𝕊(1,1](x,y)\in\mathbb{S}_{(-1,1]}, i.e., x>0x>0 and x<yx-x<y\leq x, we show the inequality (6.15). Since

μ2(x,y)y=\displaystyle\mu_{2}(x,y)-y= 2[mH(x,μ0)mA(x,μ0)]+mA(x,μ0)y,\displaystyle 2[m_{H}(x,\mu_{0})-m_{A}(x,\mu_{0})]+m_{A}(x,\mu_{0})-y,
2[mA(x,μ0)mH(x,μ0)]=(xμ0)2x+μ0,\displaystyle 2[m_{A}(x,\mu_{0})-m_{H}(x,\mu_{0})]=\frac{(x-\mu_{0})^{2}}{x+\mu_{0}},

we have

μ2(x,y)y0mA(x,μ0)y(xμ0)2x+μ0,\mu_{2}(x,y)-y\geq 0\quad\Leftrightarrow\quad m_{A}(x,\mu_{0})-y\geq\frac{(x-\mu_{0})^{2}}{x+\mu_{0}},

where μ0\mu_{0} denotes μ0(x,y)\mu_{0}(x,y). Note that

y<mA(x,y)<μ0<mA(x,μ0)<x,0xμ0x+μ0<1.y<m_{A}(x,y)<\mu_{0}<m_{A}(x,\mu_{0})<x,\quad 0\leq\frac{x-\mu_{0}}{x+\mu_{0}}<1.

Then we have

mA(x,μ0)ymA(x,y)y=xmA(x,y)xμ0xμ0x+μ0(xμ0)=(xμ0)2x+μ0.m_{A}(x,\mu_{0})-y\geq m_{A}(x,y)-y=x-m_{A}(x,y)\geq x-\mu_{0}\geq\frac{x-\mu_{0}}{x+\mu_{0}}(x-\mu_{0})=\frac{(x-\mu_{0})^{2}}{x+\mu_{0}}.

To show the restriction of μ2\mu_{2} to 𝕊(0,17)\mathbb{S}_{(0,17)} is a strict homogeneous mean, we check that 𝕊(1/3,3)\mathbb{S}_{(1/3,3)} contains the image of 𝕊(0,17)\mathbb{S}_{(0,17)} under the map

(x,y)(x,μ0(x,y)).(x,y)\mapsto(x,\mu_{0}(x,y)).

In fact, if 0<y<x0<y<x then μ0(x,y)\mu_{0}(x,y) satisfies inequalities

x3<x2<μ0(x,y)<x\frac{x}{3}<\frac{x}{\sqrt{2}}<\mu_{0}(x,y)<x

by (6.12). If xy<17xx\leq y<17x then

xμ0(x,y)=mG(x,mA(x,y))<mG(x,mA(x,17x))=mG(x,9x)=3x.x\leq\mu_{0}(x,y)=m_{G}(x,m_{A}(x,y))<m_{G}(x,m_{A}(x,17x))=m_{G}(x,9x)=3x.

Thus we have (x,μ0(x,y))𝕊(1/3,3)(x,\mu_{0}(x,y))\in\mathbb{S}_{(1/3,3)} for any (x,y)𝕊(0,17)(x,y)\in\mathbb{S}_{(0,17)}. Since ν\nu is a strict homogeneous mean on 𝕊(1/3,3)\mathbb{S}_{(1/3,3)} as shown in Lemma 6.8, and the restriction of μ0\mu_{0} to 𝕊(0,17)\mathbb{S}_{(0,17)} is strict homogeneous and that of prx\mathrm{pr}_{x} to 𝕊(0,17)\mathbb{S}_{(0,17)} is homogeneous, μ2(x,y)=ν(prx(x,y),μ0(x,y))\mu_{2}(x,y)=\nu(pr_{x}(x,y),\mu_{0}(x,y)) is a strict homogeneous mean on 𝕊(0,17)\mathbb{S}_{(0,17)} by Lemma 6.2.

(iii)(\ref{item:ineq}) We have

μ1(x,y)μ2(x,y)\displaystyle\mu_{1}(x,y)-\mu_{2}(x,y) =2(mA(x,μ0(x,y))mH(x,μ0(x,y)))0,\displaystyle=2(m_{A}(x,\mu_{0}(x,y))-m_{H}(x,\mu_{0}(x,y)))\geq 0,
μ1(x,y)+μ2(x,y)\displaystyle\mu_{1}(x,y)+\mu_{2}(x,y) =2mH(x,μ0(x,y))>0,\displaystyle=2m_{H}(x,\mu_{0}(x,y))>0,

by using (6.14) together with the positivity of μ0(x,y)\mu_{0}(x,y) on 𝕊(1,)\mathbb{S}_{(-1,\infty)}. Since

mA(x,μ0(x,y))=mH(x,μ0(x,y))x=μ0(x,y)x=y,m_{A}(x,\mu_{0}(x,y))=m_{H}(x,\mu_{0}(x,y))\Leftrightarrow x=\mu_{0}(x,y)\Leftrightarrow x=y,

if xyx\neq y then (6.16) holds strictly.

If (x,y)𝕊(1,1](x,y)\in\mathbb{S}_{(-1,1]} then x>0x>0 and x<yx-x<y\leq x, we have

μ1(x,y)+μ2(x,y)=2mH(x,μ0(x,y))2mH(mA(x,y),mA(x,y))=x+y(>0)\mu_{1}(x,y)+\mu_{2}(x,y)=2m_{H}(x,\mu_{0}(x,y))\geq 2m_{H}(m_{A}(x,y),m_{A}(x,y))=x+y\ (>0)

since xx and μ0(x,y)\mu_{0}(x,y) are greater than or equal to mA(x,y)m_{A}(x,y). \square

Remark 6.11.

If (x,y)𝕊(7/9,17)(x,y)\in\mathbb{S}_{(-7/9,17)} then (x,μ0(x,y))𝕊(1/3,3)(x,\mu_{0}(x,y))\in\mathbb{S}_{(1/3,3)}.

By using functions μ1\mu_{1} and μ2\mu_{2} on 𝕊(1,)\mathbb{S}_{(-1,\infty)}, we define a pair of sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} for (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)} as

(x0,y0)=(x,y),(xn+1,yn+1)=(μ1(xn,yn),μ2(xn,yn)).(x_{0},y_{0})=(x,y),\quad(x_{n+1},y_{n+1})=(\mu_{1}(x_{n},y_{n}),\mu_{2}(x_{n},y_{n})). (6.19)
Lemma 6.12.

The sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} converge as nn\to\infty, and satisfy

limnxn=limnyn>0.\lim_{n\to\infty}x_{n}=\lim_{n\to\infty}y_{n}>0.
Proof..

For any (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)}, x1=μ1(x,y)x_{1}=\mu_{1}(x,y) and y1=μ2(x,y)y_{1}=\mu_{2}(x,y) satisfy x1>0x_{1}>0, x1<y1x1-x_{1}<y_{1}\leq x_{1} by (6.16) and (6.17) in Proposition 6.10. Since (x1,y1)𝕊(1,1](x_{1},y_{1})\in\mathbb{S}_{(-1,1]}, and μ1\mu_{1} is a strict mean on 𝕊(1,)\mathbb{S}_{(-1,\infty)}, x2=μ1(x1,y1)x_{2}=\mu_{1}(x_{1},y_{1}) satisfies

0<x2,y1x2x1.0<x_{2},\quad y_{1}\leq x_{2}\leq x_{1}.

Since (x1,y1)𝕊(1,1](x_{1},y_{1})\in\mathbb{S}_{(-1,1]} and (6.15), (6.16), (6.17) in Proposition 6.10, y2=μ2(x1,y1)y_{2}=\mu_{2}(x_{1},y_{1}) satisfies

y1y2,x2<y2x2.y_{1}\leq y_{2},\quad-x_{2}<y_{2}\leq x_{2}.

Thus we have

(x2,y2)𝕊(1,1],y1y2x2x1.(x_{2},y_{2})\in\mathbb{S}_{(-1,1]},\quad y_{1}\leq y_{2}\leq x_{2}\leq x_{1}.

Inductively, we have

(xn,yn)𝕊(1,1],y1ynxnx1(x_{n},y_{n})\in\mathbb{S}_{(-1,1]},\quad y_{1}\leq\cdots\leq y_{n}\leq x_{n}\leq\cdots\leq x_{1}

for any nn\in\mathbb{N} and any (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)}.

Since the sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} (n1)(n\geq 1) are monotonous and bounded, they converge. We set

limnxn=ξ,limnyn=η,\lim_{n\to\infty}x_{n}=\xi,\quad\lim_{n\to\infty}y_{n}=\eta,

which satisfy ξη\xi\geq\eta. Since xnynx_{n}\geq y_{n} and mA(xn,yn)μ0(xn,yn)xnm_{A}(x_{n},y_{n})\leq\mu_{0}(x_{n},y_{n})\leq x_{n}, we have

mA(xn+1,yn+1)=mH(xn,μ0(xn,yn))mH(mA(xn,yn),mA(xn,yn))=mA(xn,yn)>0m_{A}(x_{n+1},y_{n+1})=m_{H}(x_{n},\mu_{0}(x_{n},y_{n}))\geq m_{H}(m_{A}(x_{n},y_{n}),m_{A}(x_{n},y_{n}))=m_{A}(x_{n},y_{n})>0

which yields ξ+η>0\xi+\eta>0. By considering the limit nn\to\infty for

xn+1=μ1(xn,yn),x_{n+1}=\mu_{1}(x_{n},y_{n}),

we have

ξ=ξ+μ0(ξ,η)2ξ=μ0(ξ,η)ξ2=ξξ+η2.\xi=\frac{\xi+\mu_{0}(\xi,\eta)}{2}\quad\Leftrightarrow\quad\xi=\mu_{0}(\xi,\eta)\quad\Leftrightarrow\quad\xi^{2}=\xi\cdot\frac{\xi+\eta}{2}.

The last equality is equivalent to ξ=η\xi=\eta or ξ=0\xi=0. If ξ=0\xi=0 then 0=ξη0=\xi\geq\eta, which contradicts ξ+η>0\xi+\eta>0. Hence we have ξ=η>0\xi=\eta>0. \square

Theorem 6.13.

For the sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} given in (6.19), there exists N0N\in\mathbb{N}_{0} such that 0<ynxn0<y_{n}\leq x_{n} for any nNn\geq N, and that

limnxn=limnyn=xNF(14,14,1;1yN2xN2).\lim_{n\to\infty}x_{n}=\lim_{n\to\infty}y_{n}=\frac{x_{N}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N}^{2}}{x_{N}^{2}})}.
Proof..

Since the sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} satisfy ynxny_{n}\leq x_{n} for n1n\geq 1, and converge to a positive real number by Lemma 6.12, the number N0N\in\mathbb{N}_{0} exists. We have

xNF(14,14,1;1yN2xN2)\displaystyle\frac{x_{N}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N}^{2}}{x_{N}^{2}})} =m1(xN,yN)F(14,14,1;1m2(xN,yN)2m1(xN,yN)2)=xN+1F(14,14,1;1yN+12xN+12)\displaystyle=\frac{m_{1}(x_{N},y_{N})}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{m_{2}(x_{N},y_{N})^{2}}{m_{1}(x_{N},y_{N})^{2}})}=\frac{x_{N+1}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N+1}^{2}}{x_{N+1}^{2}})}
=xN+2F(14,14,1;1yN+22xN+22)==xN+kF(14,14,1;1yN+k2xN+k2)\displaystyle=\frac{x_{N+2}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N+2}^{2}}{x_{N+2}^{2}})}=\cdots=\frac{x_{N+k}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N+k}^{2}}{x_{N+k}^{2}})}

by Theorem 5.1, our definition of μ1,μ2\mu_{1},\mu_{2} in (6.11) and 0<ynxnyn+1xn+110<\frac{y_{n}}{x_{n}}\leq\frac{y_{n+1}}{x_{n+1}}\leq 1 for nNn\geq N. Let kk tend to \infty in the last term of the previous equality, then

xNF(14,14,1;1yN2xN2)=limnxn,\frac{x_{N}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N}^{2}}{x_{N}^{2}})}=\lim_{n\to\infty}x_{n},

holds, since limnxn=limnyn\displaystyle{\lim_{n\to\infty}x_{n}=\lim_{n\to\infty}y_{n}} and limz1F(14,14,1;1z2)=1\lim\limits_{z\to 1}F(\frac{1}{4},\frac{1}{4},1;1-z^{2})=1. \square

Alternative Proof.

For the real numbers xN,yNx_{N},y_{N}, there exists a pure imaginary number τ𝔻12\tau\in\mathbb{D}_{12} such that

1yN2xN2=ζ(τ)=4ϑ01(τ)4ϑ10(τ)4ϑ00(τ)8=1(2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)4)21-\frac{y_{N}^{2}}{x_{N}^{2}}=\zeta(\tau)=\frac{4\vartheta_{01}(\tau)^{4}\vartheta_{10}(\tau)^{4}}{\vartheta_{00}(\tau)^{8}}=1-\Big{(}\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{4}}\Big{)}^{2}

by Fact 3.5 (iiii). Thus we have a positive real number ξ\xi such that

xN=ξϑ00(τ)2,yN=ξ2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)2.x_{N}=\xi\cdot\vartheta_{00}(\tau)^{2},\quad y_{N}=\xi\cdot\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{2}}.

Since

μ1(xN,yN)\displaystyle\mu_{1}(x_{N},y_{N}) =ξμ1(ϑ00(τ)2,2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)2)=ξϑ00(2τ)2,\displaystyle=\xi\cdot\mu_{1}\big{(}\vartheta_{00}(\tau)^{2},\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{2}}\big{)}=\xi\cdot\vartheta_{00}(2\tau)^{2},
μ2(xN,yN)\displaystyle\mu_{2}(x_{N},y_{N}) =ξμ2(ϑ00(τ)2,2ϑ01(τ)4ϑ00(τ)4ϑ00(τ)2)=ξ2ϑ01(2τ)4ϑ00(2τ)4ϑ00(2τ)2\displaystyle=\xi\cdot\mu_{2}\big{(}\vartheta_{00}(\tau)^{2},\frac{2\vartheta_{01}(\tau)^{4}-\vartheta_{00}(\tau)^{4}}{\vartheta_{00}(\tau)^{2}}\big{)}=\xi\cdot\frac{2\vartheta_{01}(2\tau)^{4}-\vartheta_{00}(2\tau)^{4}}{\vartheta_{00}(2\tau)^{2}}

by our definition of μ1\mu_{1} and μ2\mu_{2}, xnx_{n} and yny_{n} admit expressions

xN+k=ξϑ00(2kτ)2,yN+k=ξ2ϑ01(2kτ)4ϑ00(2kτ)4ϑ00(2kτ)2.x_{N+k}=\xi\cdot\vartheta_{00}(2^{k}\tau)^{2},\quad y_{N+k}=\xi\cdot\frac{2\vartheta_{01}(2^{k}\tau)^{4}-\vartheta_{00}(2^{k}\tau)^{4}}{\vartheta_{00}(2^{k}\tau)^{2}}.

Hence we have

limnxn=ξlimkϑ00(2kτ)2=ξ=xNϑ00(τ)2=xNF(14,14,1;1yN2xN2)\lim_{n\to\infty}x_{n}=\xi\cdot\lim_{k\to\infty}\vartheta_{00}(2^{k}\tau)^{2}=\xi=\frac{x_{N}}{\vartheta_{00}(\tau)^{2}}=\frac{x_{N}}{F(\frac{1}{4},\frac{1}{4},1;1-\frac{y_{N}^{2}}{x_{N}^{2}})}

by Theorem 4.5. \square

Remark 6.14.

For (x,y)𝕊(1,)(x,y)\in\mathbb{S}_{(-1,\infty)} with y<0y<0, we have (x,y)𝕊(1,)(x,-y)\in\mathbb{S}_{(-1,\infty)} and

F(14,14,1;1y2x2)=F(14,14,1;1(y)2x2).F(\frac{1}{4},\frac{1}{4},1;1-\frac{y^{2}}{x^{2}})=F(\frac{1}{4},\frac{1}{4},1;1-\frac{(-y)^{2}}{x^{2}}).

However, the pair of sequences in (6.19) with (x0,y0)=(x,y)(x_{0},y_{0})=(x,y) is different from that with (x0,y0)=(x,y)(x_{0},y_{0})=(x,-y), and their limits as nn\to\infty do not coincide in general.

References

  • [BB] Borwein J.M. and Borwein P.B., Pi and the AGM, John Wiley & Sons, Inc., New York, 1998.
  • [Er] Erdélyi, A., Magnus, W., Oberhettinger, F., and Tricomi, F.G., Higher transcendental functions, Vol. I, II, McGraw-Hill, 1953.
  • [Go] Goto Y., Functional equations for Appell’s F1F_{1} arising from transformations of elliptic curves, Funkcial. Ekvac. 56 (2013), 379–396.
  • [Ki] Kimura T., Hypergeometric functions of two variables, Seminar Note in Math. Univ. of Tokyo, 1973.
  • [HKM] Hattori R., Kato T. and Matsumoto K., Mean iterations derived from transformation formulas for the hypergeometric function, Hokkaido Math. J. 38 (2009), 563–586.
  • [KS1] Koike K. and Shiga H., Isogeny formulas for the Picard modular form and a three terms arithmetic geometric mean, J. Number Theory 124 (2007), 123–141.
  • [KS2] Koike K. and Shiga H., An extended Gauss AGM and corresponding Picard modular forms, J. Number Theory 128 (2008), 2097–2126.
  • [Ma] Matsumoto K., A transformation formula for Appell’s hypergeometric function F1F_{1} and common limits of triple sequences by mean iterations, Tohoku Math. J. (2) 62 (2010), 263–268.
  • [MOT] Matsumoto K., Osafune S. and Terasoma T., Schwarz’s map for Appell’s second hypergeometric system with quarter integer parameters, to appear in Tohoku Math. J. (2).
  • [MS] Matsumoto K. and Shiga H., A variant of Jacobi type formula for Picard curves, J. Math. Soc. Japan 62 (2010), 305–319.
  • [MT] Matsumoto K. and Terasoma T., Thomae type formula for K3K3 surfaces given by double covers of the projective plane branching along six lines, J. Reine Angew. Math. 669 (2012), 121–149.
  • [Mu] Mumford D., Tata Lectures on Theta I, Birkhäuser, Boston, 2007.
  • [Sh] Shiga H., A Jacobi-type formula for a family of hyperelliptic curves of genus 33 with automorphism of order 44. Kyushu J. Math. 65 (2011), 169–177.
  • [Th] Thomae C.J., Beitrag zur Bestimmung von θ(0,0,0)\theta(0,0,\dots 0) durch die Klassenmoduln algebraischer Functionen, J. Reine Angew. Math. 71 (1870), 201–222.
  • [Yo] Yoshida, M., Hypergeometric functions, my love, Friedr. Vieweg & Sohn, Braunschweig, 1997.