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

Kapaev’s global asymptotics of the fourth Painlevé transcendents. Elliptic asymptotics

Shun Shimomura Department of Mathematics, Keio University, 3-14-1, Hiyoshi, Kohoku-ku, Yokohama 223-8522 Japan  [email protected]
Abstract.

For the fourth Painlevé transcendents we derive elliptic asymptotic representations, which were announced by late Professor Kapaev without proofs. Then we newly obtain related results including the correction function.

2010 Mathematics Subject Classification. 34M55, 34M56, 34M40, 34M60, 33E05.

Key words and phrases. elliptic asymptotic representation; fourth Painlevé transcendents; WKB analysis; isomonodromy deformation; monodromy data.

To the memory of Professor Kapaev

1. Introduction

The fourth Painlevé equation

PIV y′′=(y)22y+32y3+4xy2+(4α+β+2x2)yβ22yy^{\prime\prime}=\frac{(y^{\prime})^{2}}{2y}+\frac{3}{2}y^{3}+4xy^{2}+(-4\alpha+\beta+2x^{2})y-\frac{\beta^{2}}{2y}

(y=dy/dx)(y^{\prime}=dy/dx) with α,β\alpha,\beta\in\mathbb{C} governs the isomonodromy deformation of the linear system

(1.1) dΨdξ=((ξ32+ξ(x+uv)+αξ)σ3+i(0ξ2u+2xu+uξ2v+2xvv0))Ψ,\displaystyle\frac{d\Psi}{d\xi}=\Biggl{(}\Bigl{(}\frac{\xi^{3}}{2}+\xi(x+\mathrm{u}\mathrm{v})+\frac{\alpha}{\xi}\Bigr{)}\sigma_{3}+i\begin{pmatrix}0&\xi^{2}\mathrm{u}+2x\mathrm{u}+\mathrm{u}^{\prime}\\ \xi^{2}\mathrm{v}+2x\mathrm{v}-\mathrm{v}^{\prime}&0\end{pmatrix}\Biggr{)}\Psi,
β=uvuv+2xuv(uv)2,y=uv\displaystyle\beta=\mathrm{u}^{\prime}\mathrm{v}-\mathrm{u}\mathrm{v}^{\prime}+2x\mathrm{u}\mathrm{v}-(\mathrm{u}\mathrm{v})^{2},\quad y=\mathrm{u}\mathrm{v}

by Kitaev [16] (for another isomonodromy system see [10]). Kapaev [13] announced global asymptotic results on solutions of PIV including elliptic asymptotic representations along generic directions in the complex plane, giving notice of publishing the proofs in [4] written by him in collaboration with Fokas, Its and Novokshenov. The monograph [4], however, contains none of these proofs, and Kapaev passed away without publishing the asymptotics on PIV except [9] for solutions on the real line and [14]. In Kapaev’s announcement, [13, Theorems 2 and 3] are on elliptic asymptotics described in terms of an elliptic integral as the inverse function; [13, Theorems 4 and 5] on trigonometric asymptotics, and [13, Theorems 6, 7, 8 and 9] on truncated solutions. Asymptotics of PIV along Stokes rays are also studied by Kitaev [16], [17], [18]. By isomonodromy technique, elliptic asymptotics of general solutions are known for PI ([15], [20], [21]), PII ([23], [24], [11], [12], [21]), P(D6)𝐈𝐈𝐈{}_{\mathbf{III}}(D_{6}) ([25], [26], [28]), P(D7)𝐈𝐈𝐈{}_{\mathbf{III}}(D_{7}) ([27]) and PV ([29]), and in each asymptotic formula the phase shift is expressed in terms of the monodromy data. For the complete PIV the proofs of elliptic asymptotics have not been published, though PIV with α=0\alpha=0 was treated by Vereshchagin [31] for 0<argx<π/40<\arg x<\pi/4. The present author believes that it is significant to present the proofs of elliptic representations for PIV in Kapaev’s announcement [13] for the following reasons: (i) in the study of a general solution of PIV, elliptic asymptotics are crucial information; (ii) as a practical matter, the process of deriving the elliptic representation needs some devices peculiar to PIV; and (iii) related important materials including the correction function, which appear in the proofs, are not referred to in the announcement.

In this paper we derive the elliptic asymptotics for the complete PIV by the isomonodromy deformation method along the lines of the arguments in [12], [20], [21] with discussions on the Boutroux equations and on the justification of the asymptotics as a solution of PIV. Then we newly obtain related results including the correction function Bϕ(t)B_{\phi}(t) given by (5.1), which contains information on the asymptotics and is essential in the justification procedure as in [21, Section 3].

The main results are stated in Section 2. Theorems 2.1 and 2.2 present elliptic representations of a general solution of PIV, which correspond to the announced [13, Theorem 2]. These results are also described by an alternative elliptic expression as of Corollary 2.3, which has an advantage in treating in general sectors (cf. Theorem 2.4). For elliptic expressions of Theorems 2.1 and 2.2 in directions neighbouring the positive real axis, degeneration to trigonometric asymptotics may be considered under certain suppositions, and is shown to be consistent with the result of [13, Theorem 4]. This fact supports the validity of signs in Theorems 2.1 and 2.2 contradicting those of [13, Theorem 2] (cf. Remark 2.3). Section 3 summarises necessary facts on the isomonodromy linear system (3.1) and on its monodromy data consisting of Stokes coefficients. Section 4 explains turning points, Stokes graphs and WKB solutions, which are necessary in the WKB analysis. In Section 5 we solve a direct monodromy problem for system (3.1) by the WKB analysis to obtain the key relations consisting of monodromy data and certain integrals (cf. Propositions 5.1 and 5.2). Asymptotics of these key relations are examined in Section 6 by the use of the ϑ\vartheta-function. In Section 7 from the formulas thus obtained asymptotic forms of the main theorems are derived by solving an inverse monodromy problem for the prescribed monodromy data. In this process we make technical devices to find necessary special properties of the elliptic function related to our case (Propositions 7.5 and 7.6). The justification as a solution of PIV is performed along the lines of Kitaev [21] with [19]. The final section is devoted to the proofs of necessary facts on the Boutroux equations summarised in Proposition 8.15, which determine AϕA_{\phi} parametrising the related elliptic function. Furthermore we clarify local structure of Stokes curves near coalescing turning points, which is used in drawing Stokes graphs in Section 4.

Throughout this paper we use the following symbols:

(1) The coefficient A(φ0)A(\varphi_{0}) defined by [13, (20)] is denoted by e3iϕAϕe^{3i\phi}A_{\phi} (ϕ=φ0)(\phi=\varphi_{0});

(2) σ1,\sigma_{1}, σ2\sigma_{2}, σ3\sigma_{3} denote the Pauli matrices

σ1=(0110),σ2=(0ii0),σ3=(1001);\sigma_{1}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix},\quad\sigma_{2}=\begin{pmatrix}0&-i\\ i&0\end{pmatrix},\quad\sigma_{3}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix};

(3) for complex-valued functions ff and gg, we write fgf\ll g or gfg\gg f if f=O(|g|)f=O(|g|), and write fgf\asymp g if gfgg\ll f\ll g.

2. Results

To state the results we explain the monodromy data [13, Section 2], [9, Section 2], and the Boutroux equations [13, Section 3]. For kk\in\mathbb{Z} system (1.1) admits the matrix solutions

(2.1) Ψk(ξ)=(I+O(ξ1))exp((18ξ4+12xξ2+(αβ)lnξ)σ3)\Psi_{k}^{\infty}(\xi)=(I+O(\xi^{-1}))\exp((\tfrac{1}{8}\xi^{4}+\tfrac{1}{2}x\xi^{2}+(\alpha-\beta)\ln\xi)\sigma_{3})

as ξ\xi\to\infty through the sector |argξ+π8π4k|<π4,|\arg\xi+\frac{\pi}{8}-\frac{\pi}{4}k|<\frac{\pi}{4}, and

Ψ0(ξ)=T0(ξ)ξασ3(I+J0lnξ)eint(x)σ3,J0=(0j+j0),int(x)=xuv𝑑x\Psi^{0}(\xi)=T_{0}(\xi)\xi^{\alpha\sigma_{3}}(I+J_{0}\ln\xi)e^{\mathrm{int}(x)\sigma_{3}},\quad J_{0}=\begin{pmatrix}0&j_{+}\\ j_{-}&0\end{pmatrix},\quad\mathrm{int}(x)=\int^{x}\mathrm{u}\mathrm{v}\,dx

as ξ0\xi\to 0, where T0(ξ)T_{0}(\xi) is invertible around ξ=0\xi=0, and j+=0j_{+}=0 if α120,1,2,,\alpha-\frac{1}{2}\not=0,1,2,\ldots, j=0j_{-}=0 if α121,2,3,.\alpha-\frac{1}{2}\not=-1,-2,-3,\ldots. The Stokes matrices

S2l1=(1s2l101),S2l=(10s2l1)(l)S_{2l-1}=\begin{pmatrix}1&s_{2l-1}\\ 0&1\end{pmatrix},\quad S_{2l}=\begin{pmatrix}1&0\\ s_{2l}&1\end{pmatrix}\quad(l\in\mathbb{Z})

are defined by Ψk+1(ξ)=Ψk(ξ)Sk\Psi_{k+1}^{\infty}(\xi)=\Psi_{k}^{\infty}(\xi)S_{k}, and satisfy Sk+4=eiπ(αβ)σ3σ3Skσ3eiπ(αβ)σ3,S_{k+4}=e^{-i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}S_{k}\sigma_{3}e^{i\pi(\alpha-\beta)\sigma_{3}}, sk+4=ske(1)k2πi(αβ).s_{k+4}=-s_{k}e^{(-1)^{k}2\pi i(\alpha-\beta)}. For the matrices M=eint(x)σ3(I+iπJ0)eint(x)σ3eiπασ3M=e^{-\mathrm{int}(x)\sigma_{3}}(I+i\pi J_{0})e^{\mathrm{int}(x)\sigma_{3}}e^{i\pi\alpha\sigma_{3}} and EE such that σ3Ψ0(eiπξ)σ3=Ψ0(ξ)M\sigma_{3}\Psi^{0}(e^{i\pi}\xi)\sigma_{3}=\Psi^{0}(\xi)M and Ψ1(ξ)=Ψ0(ξ)E\Psi_{1}^{\infty}(\xi)=\Psi^{0}(\xi)E, the semi-cyclic relation

S1S2S3S4=E1σ3M1Eeiπ(αβ)σ3σ3S_{1}S_{2}S_{3}S_{4}=E^{-1}\sigma_{3}M^{-1}Ee^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}

holds, and the traces of both sides lead to the surface of the monodromy data

0(α,β):\mathcal{M}_{0}(\alpha,\beta): ((1+s1s2)(1+s3s4)+s1s4)eiπ(αβ)(1+s2s3)eiπ(αβ)=2isinπα.((1+s_{1}s_{2})(1+s_{3}s_{4})+s_{1}s_{4})e^{-i\pi(\alpha-\beta)}-(1+s_{2}s_{3})e^{i\pi(\alpha-\beta)}=-2i\sin\pi\alpha.

For each mm\in\mathbb{Z} the semi-cyclic relation for Sj+mS_{j+m} (1j4)(1\leq j\leq 4) in Proposition 3.1 yields

m(α,β):\mathcal{M}_{m}(\alpha,\beta): ((1+s1+ms2+m)\displaystyle((1+s_{1+m}s_{2+m}) (1+s3+ms4+m)+s1+ms4+m)eiπ(1)m(αβ)\displaystyle(1+s_{3+m}s_{4+m})+s_{1+m}s_{4+m})e^{-i\pi(-1)^{m}(\alpha-\beta)}
(1+s2+ms3+m)eiπ(1)m(αβ)=2i(1)msinπα.\displaystyle-(1+s_{2+m}s_{3+m})e^{i\pi(-1)^{m}(\alpha-\beta)}=-2i(-1)^{m}\sin\pi\alpha.

Around a nonsingular point on m(α,β)\mathcal{M}_{m}(\alpha,\beta) suitable three of sj+ms_{j+m} (1j4)(1\leq j\leq 4) are independent. For any c{0}c\in\mathbb{C}\setminus\{0\} the gauge transformation Ψ=cσ3Ψ~\Psi=c^{\sigma_{3}}\tilde{\Psi} induces the action

[c]:(S1,S2,S3,S4)(cσ3S1cσ3,cσ3S2cσ3,cσ3S3cσ3,cσ3S4cσ3)[c]:\,\,\,(S_{1},S_{2},S_{3},S_{4})\mapsto(c^{-\sigma_{3}}S_{1}c^{\sigma_{3}},c^{-\sigma_{3}}S_{2}c^{\sigma_{3}},c^{-\sigma_{3}}S_{3}c^{\sigma_{3}},c^{-\sigma_{3}}S_{4}c^{\sigma_{3}})

on 0(α,β)\mathcal{M}_{0}(\alpha,\beta) consistent with the isomonodromy structure of (1.1), and each solution of PIV corresponds to an orbit, or equivalence class yielded by dividing 0(α,β)\mathcal{M}_{0}(\alpha,\beta) by [c][c]. Thus an orbit passing through a point (s1,s2,s3,s4)0(α,β)(s_{1},s_{2},s_{3},s_{4})\in\mathcal{M}_{0}(\alpha,\beta) parametrises a solution of PIV. Let us call it a solution labelled by (s1,s2,s3,s4)(s_{1},s_{2},s_{3},s_{4}). (In [13, Section 2] the gauge symmetry on (u,v)(\mathrm{u},\mathrm{v}) is considered.)

For 0<A<8270<A<\frac{8}{27} let z=0,z=0, x1,x_{1}, x3,x_{3}, x5x_{5} with x5<x3<x1<0x_{5}<x_{3}<x_{1}<0 be the zeros of the polynomial PA(z)=z4+4z3+4z2+4AzP_{A}(z)=z^{4}+4z^{3}+4z^{2}+4Az such that x10,x_{1}\to 0, x3,x52x_{3},x_{5}\to-2 as A0A\to 0 and that x1,x323,x_{1},x_{3}\to-\frac{2}{3}, x583x_{5}\to-\frac{8}{3} as A827A\to\frac{8}{27}. Let Π±\Pi_{\pm} be two copies of P1()([x5,x3][x1,0])P^{1}(\mathbb{C})\setminus([x_{5},x_{3}]\cup[x_{1},0]). The elliptic curve w2=PA(z)w^{2}=P_{A}(z) is the two sheeted Riemann surface ΠA,0=Π+Π\Pi_{A,0}=\Pi_{+}\cup\Pi_{-} glued along the cuts [x5,x3][x_{5},x_{3}], [x1,0][x_{1},0]. As long as (ϕ,A)×𝒟0(\phi,A)\in\mathbb{R}\times\mathcal{D}_{0} with 𝒟0={c0}{c827}\mathcal{D}_{0}=\mathbb{C}\setminus\{c\leq 0\}\cup\{c\geq\tfrac{8}{27}\}, the polynomial z4+4eiϕz3+4e2iϕz2+4e3iϕAzz^{4}+4e^{i\phi}z^{3}+4e^{2i\phi}z^{2}+4e^{3i\phi}Az admits the roots 0,0, zjz_{j} (j=1,3,5)(j=1,3,5) such that Reeiϕz5<Reeiϕz3<Reeiϕz1\mathrm{Re\,}e^{-i\phi}z_{5}<\mathrm{Re\,}e^{-i\phi}z_{3}<\mathrm{Re\,}e^{-i\phi}z_{1} and that zj=xjz_{j}=x_{j} if ϕ=0,\phi=0, 0<A<8270<A<\frac{8}{27} (cf. Corollary 8.2), and then the elliptic curve

w2=w(A,z)2=z4+4eiϕz3+4e2iϕz2+4e3iϕAz=z(zz1)(zz3)(zz5)w^{2}=w(A,z)^{2}=z^{4}+4e^{i\phi}z^{3}+4e^{2i\phi}z^{2}+4e^{3i\phi}Az=z(z-z_{1})(z-z_{3})(z-z_{5})

may be considered to be the two sheeted Riemann surface ΠA,ϕ=Π+Π\Pi_{A,\phi}=\Pi_{+}\cup\Pi_{-} that is a continuous modification of ΠA,0\Pi_{A,0} with Π±\Pi_{\pm} glued along cuts [z1,0],[z_{1},0], [z5,z3][z_{5},z_{3}]. Here the branch of w=z(zz1)(zz3)(zz5)=z21z1z11z3z11z5z1w=\sqrt{z(z-z_{1})(z-z_{3})(z-z_{5})}=z^{2}\sqrt{1-z_{1}z^{-1}}\sqrt{1-z_{3}z^{-1}}\sqrt{1-z_{5}z^{-1}} is such that 1zjz11\sqrt{1-z_{j}z^{-1}}\to 1 as zz\to\infty (j=1,3,5)(j=1,3,5) on the upper sheet Π+\Pi_{+}.

As will be shown in Corollary 8.16 with Remark 8.3, for each ϕ\phi\in\mathbb{R}, there exists AϕA_{\phi}\in\mathbb{C} such that, for any cycle 𝐜\mathbf{c} on ΠAϕ,ϕ\Pi_{A_{\phi},\phi}

Re𝐜w(Aϕ,z)z𝑑z=0\mathrm{Re\,}\int_{\mathbf{c}}\frac{w(A_{\phi},z)}{z}dz=0

and that AϕA_{\phi} has the following properties:

(1) for each ϕ\phi\in\mathbb{R}, AϕA_{\phi} is uniquely determined;

(2) Aϕ+π/2=Aϕ,{A}_{\phi+\pi/2}={A}_{\phi}, Aϕ=Aϕ¯{A}_{-\phi}=\overline{{A}_{\phi}};

(3) A0=827,A_{0}=\frac{8}{27}, A±π/4=0A_{\pm\pi/4}=0 and 0ReAϕ827;0\leq\mathrm{Re\,}A_{\phi}\leq\tfrac{8}{27};

(4) AϕA_{\phi} is continuous in ϕ\phi\in\mathbb{R} and is smooth in ϕ{πk/4|k}.\phi\in\mathbb{R}\setminus\{\pi k/4\,|\,k\in\mathbb{Z}\}.

The elliptic curve ΠAϕ,ϕ\Pi_{A_{\phi},\phi} degenerates if and only if ϕ=πk/4\phi=\pi k/4 with k.k\in\mathbb{Z}.

2.1. Solutions for 0<|ϕ|<π/40<|\phi|<\pi/4 in [13, Theorem 2]

For 0<|ϕ|<π/40<|\phi|<\pi/4 and A𝒟0A\in\mathcal{D}_{0}, let the primitive cycles 𝐚\mathbf{a} and 𝐛\mathbf{b} on ΠA,ϕ\Pi_{A,\phi} be as described on the upper sheet Π+\Pi_{+} in Figure 2.1. (The cycles 𝐚\mathbf{a} and 𝐛\mathbf{b} are consistent with those defined in [13, Section 3].) Then the Boutroux equations

(2.2) Re𝐚w(A,z)z𝑑z=Re𝐛w(A,z)z𝑑z=0\mathrm{Re\,}\int_{\mathbf{a}}\frac{w(A,z)}{z}dz=\mathrm{Re\,}\int_{\mathbf{b}}\frac{w(A,z)}{z}dz=0

admit a unique solution A=AϕA=A_{\phi}, which means [13, Theorem 1]. For 0<|ϕ|<π/40<|\phi|<\pi/4 the periods of ΠAϕ,ϕ\Pi_{A_{\phi},\phi} along 𝐚\mathbf{a} and 𝐛\mathbf{b} are given by

Ω𝐚=Ω𝐚ϕ=𝐚dzw(Aϕ,z),Ω𝐛=Ω𝐛ϕ=𝐛dzw(Aϕ,z),\Omega_{\mathbf{a}}=\Omega_{\mathbf{a}}^{\phi}=\int_{\mathbf{a}}\frac{dz}{w(A_{\phi},z)},\quad\Omega_{\mathbf{b}}=\Omega_{\mathbf{b}}^{\phi}=\int_{\mathbf{b}}\frac{dz}{w(A_{\phi},z)},

which satisfy ImΩ𝐛/Ω𝐚>0.\mathrm{Im\,}\Omega_{\mathbf{b}}/\Omega_{\mathbf{a}}>0.

z5z_{5}z3z_{3}z1z_{1}0𝐚\mathbf{a}𝐛\mathbf{b}Π+\Pi_{+}
Figure 2.1. Cycles 𝐚,\mathbf{a}, 𝐛\mathbf{b} on ΠA,ϕ=Π+Π\Pi_{A,\phi}=\Pi_{+}\cup\Pi_{-}

Let P(u;A)\mathrm{P}(u;A) denote the elliptic function defined by

Pu2=P4+4eiϕP3+4e2iϕP2+4e3iϕAP,P(0;A)=0,i.e.0P(u;A)dzw(A,z)=u.\mathrm{P}_{u}^{2}=\mathrm{P}^{4}+4e^{i\phi}\mathrm{P}^{3}+4e^{2i\phi}\mathrm{P}^{2}+4e^{3i\phi}A\mathrm{P},\quad\mathrm{P}(0;A)=0,\quad\text{i.e.}\,\,\,\int^{\mathrm{P}(u;A)}_{0}\frac{dz}{w(A,z)}=u.

Note that P(u;Aϕ)\mathrm{P}(u;A_{\phi}) does not degenerate as long as ϕkπ/4,\phi\not=k\pi/4, k.k\in\mathbb{Z}. Then [13, Theorem 2] with n=0,n=0, m=0,1m=0,1 may be described as follows.

Let y=y(𝐬,x)y=y(\mathbf{s},x) denote the solution of PIV labelled by the monodromy data 𝐬=(s1,s2,s3,s4)0(α,β).\mathbf{s}=(s_{1},s_{2},s_{3},s_{4})\in\mathcal{M}_{0}(\alpha,\beta).

Theorem 2.1.

Suppose that π/4<ϕ<0-\pi/4<\phi<0 and that (1+s1s2)(1+s2s3)10,(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, 1+s1s20.1+s_{1}s_{2}\not=0. Then

y(𝐬,x)=eiϕxP(eiϕt+χ0+O(tδ);Aϕ),2t=(eiϕx)2,\displaystyle y(\mathbf{s},x)=e^{-i\phi}x\mathrm{P}(e^{i\phi}t+\chi_{0-}+O(t^{-\delta});A_{\phi}),\quad 2t=(e^{-i\phi}x)^{2},
χ0Ω𝐚2πiln((1+s1s2)(1+s2s3)1)\displaystyle\chi_{0-}\equiv\frac{\Omega_{\mathbf{a}}}{2\pi i}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)
Ω𝐛2πiln(1+s1s2)+Ω𝐛3(αβ)modΩ𝐚+Ω𝐛\displaystyle\phantom{----------}-\frac{\Omega_{\mathbf{b}}}{2\pi i}\ln(1+s_{1}s_{2})+\frac{\Omega_{\mathbf{b}}}{3}(\alpha-\beta)\quad\mod\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}

as 2t=(eiϕx)22t=(e^{-i\phi}x)^{2}\to\infty through the cheese-like strip

S0(ϕ,t,κ0,δ0)={x=eiϕ2t|Ret>t,|Imt|<κ0}σ𝒫0{|xσ|<δ0}S_{0}(\phi,t_{\infty},\kappa_{0},\delta_{0})=\{x=e^{i\phi}\sqrt{2t}\,|\,\mathrm{Re\,}t>t_{\infty},\,\,\,|\mathrm{Im\,}t|<\kappa_{0}\}\setminus\bigcup_{\sigma\in\mathcal{P}_{0-}}\{|x-\sigma|<\delta_{0}\}

with 𝒫0={eiϕ2t0|eiϕt0+χ0=±13Ω𝐛+Ω𝐚+Ω𝐛}.\mathcal{P}_{0-}=\{e^{i\phi}\sqrt{2t_{0-}}\,|\,e^{i\phi}t_{0-}+\chi_{0-}=\pm\tfrac{1}{3}\Omega_{\mathbf{b}}+\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}\}. Here δ\delta is some positive number, κ0\kappa_{0} a given positive number, δ0\delta_{0} a given small positive number, and t=t(κ0,δ0)t_{\infty}=t_{\infty}(\kappa_{0},\delta_{0}) a sufficiently large positive number depending on (κ0,δ0).(\kappa_{0},\delta_{0}).

Theorem 2.2.

Suppose that 0<ϕ<π/40<\phi<\pi/4 and that (1+s1s2)(1+s2s3)10,(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, 1+s2s30.1+s_{2}s_{3}\not=0. Then

y(𝐬,x)=eiϕxP(eiϕt+χ0++O(tδ);Aϕ),2t=(eiϕx)2,\displaystyle y(\mathbf{s},x)=e^{-i\phi}x\mathrm{P}(e^{i\phi}t+\chi_{0+}+O(t^{-\delta});A_{\phi}),\quad 2t=(e^{-i\phi}x)^{2},
χ0+Ω𝐚2πiln((1+s1s2)(1+s2s3)1)\displaystyle\chi_{0+}\equiv\frac{\Omega_{\mathbf{a}}}{2\pi i}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)
+Ω𝐛2πiln(1+s2s3)+Ω𝐛3(αβ)modΩ𝐚+Ω𝐛\displaystyle\phantom{----------}+\frac{\Omega_{\mathbf{b}}}{2\pi i}\ln(1+s_{2}s_{3})+\frac{\Omega_{\mathbf{b}}}{3}(\alpha-\beta)\quad\mod\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}

as 2t=(eiϕx)22t=(e^{-i\phi}x)^{2}\to\infty through the cheese-like strip

S1(ϕ,t,κ0,δ0)={x=eiϕ2t|Ret>t,|Imt|<κ0}σ𝒫0+{|xσ|<δ0}S_{1}(\phi,t_{\infty},\kappa_{0},\delta_{0})=\{x=e^{i\phi}\sqrt{2t}\,|\,\mathrm{Re\,}t>t_{\infty},\,\,\,|\mathrm{Im\,}t|<\kappa_{0}\}\setminus\bigcup_{\sigma\in\mathcal{P}_{0+}}\{|x-\sigma|<\delta_{0}\}

with 𝒫0+={eiϕ2t0+|eiϕt0++χ0+=±13Ω𝐛+Ω𝐚+Ω𝐛}.\mathcal{P}_{0+}=\{e^{i\phi}\sqrt{2t_{0+}}\,|\,e^{i\phi}t_{0+}+\chi_{0+}=\pm\tfrac{1}{3}\Omega_{\mathbf{b}}+\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}\}.

Remark 2.1.

If Aϕ0,827A_{\phi}\not=0,\tfrac{8}{27}, then P(τ;Aϕ)=e3iϕAϕ((τ;g2,g3)13e2iϕ)1\mathrm{P}(\tau;A_{\phi})=e^{3i\phi}A_{\phi}(\wp(\tau;g_{2},g_{3})-\frac{1}{3}e^{2i\phi})^{-1} with g2=4e4iϕ(Aϕ13)g_{2}=-4e^{4i\phi}(A_{\phi}-\frac{1}{3}), g3=e6iϕ(Aϕ2+43Aϕ827),g_{3}=e^{6i\phi}(-A_{\phi}^{2}+\frac{4}{3}A_{\phi}-\frac{8}{27}), where (τ;g2,g3)\wp(\tau;g_{2},g_{3}) is the Weierstrass pe-function satisfying τ2=43g2g3.\wp_{\tau}^{2}=4\wp^{3}-g_{2}\wp-g_{3}. Furthermore, τ=e3iϕAϕP2w(Aϕ,P).\wp_{\tau}=-e^{3i\phi}A_{\phi}\mathrm{P}^{-2}w(A_{\phi},\mathrm{P}).

Remark 2.2.

The solutions y(𝐬,x)y(\mathbf{s},x) in Theorems 2.1 and 2.2, and the correction function Bϕ(t)B_{\phi}(t) in Proposition 7.4 are parametrised by (𝔰12,𝔰23)=(s1s2,s2s3)(\mathfrak{s}_{12},\mathfrak{s}_{23})=(s_{1}s_{2},s_{2}s_{3}). The variables 𝔰12,\mathfrak{s}_{12}, 𝔰23\mathfrak{s}_{23}, and 𝔰34=s3s4,\mathfrak{s}_{34}=s_{3}s_{4}, 𝔰41=s4s1\mathfrak{s}_{41}=s_{4}s_{1} are invariant under the action [c][c] on 0(α,β)\mathcal{M}_{0}(\alpha,\beta), and hence the family of the orbits generated by [c][c] on 0(α,β)\mathcal{M}_{0}(\alpha,\beta) may be identified with the two-dimensional surface:

0(α,β):\mathcal{M}^{*}_{0}(\alpha,\beta): ((1+𝔰12)(1+𝔰34)+𝔰41)eiπ(βα)(1+𝔰23)eiπ(αβ)+2isinπα=0,\displaystyle\bigl{(}(1+\mathfrak{s}_{12})(1+\mathfrak{s}_{34})+\mathfrak{s}_{41}\bigr{)}e^{i\pi(\beta-\alpha)}-(1+\mathfrak{s}_{23})e^{i\pi(\alpha-\beta)}+2i\sin\pi\alpha=0,
𝔰12𝔰34𝔰23𝔰41=0,\displaystyle\mathfrak{s}_{12}\mathfrak{s}_{34}-\mathfrak{s}_{23}\mathfrak{s}_{41}=0,

whose points parametrise solutions of PIV given in Theorems 2.1, 2.2, and [13, Theorems 2, 4, 5 and 6] with n=0n=0. Singular points on this surface is given by Proposition 3.2. It is easy to see that a point (𝔰12,𝔰23,𝔰34)0(α,β)(\mathfrak{s}_{12},\mathfrak{s}_{23},\mathfrak{s}_{34})\in\mathcal{M}^{*}_{0}(\alpha,\beta) satisfying the condition of Theorem 2.1 or 2.2 is nonsingular.

Remark 2.3.

An inconsistency appears between the signs of ln((1+s1s2)(1+s2s3)1)\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1) in [13, Theorem 2] and in our Theorems 2.1 and 2.2. The agreements of the trigonometric asymptotics with [13, Theorem 4] and [9, Theorem 4.1] discussed in Section 2.3 support the correctness of the sign ++ in our theorems.

2.2. Alternative expression of solutions

Let the elliptic curve ΠAϕ=Π+Π\Pi_{A_{\phi}}^{*}=\Pi^{*}_{+}\cup\Pi_{-}^{*}: v2=v(Aϕ,ζ)2=e4iϕw(Aϕ,eiϕζ)2v^{2}=v(A_{\phi},\zeta)^{2}=e^{-4i\phi}w(A_{\phi},e^{i\phi}\zeta)^{2} be as defined in Section 8, and the cycles 𝐚\mathbf{a}^{*} and 𝐛\mathbf{b}^{*} on ΠAϕ\Pi^{*}_{A_{\phi}} as drawn in Figure 8.1. These elliptic curve and cycles, which depend on AϕA_{\phi} only, are also the images of ΠAϕ,ϕ=Π+Π\Pi_{A_{\phi},\phi}=\Pi_{+}\cup\Pi_{-} and 𝐚\mathbf{a} and 𝐛\mathbf{b} under the mapping z=eiϕζz=e^{i\phi}\zeta. Then the Boutroux equation (2.2) is written in the form

(2.3) Ree2iϕ𝐚v(Aϕ,ζ)ζ𝑑ζ=Ree2iϕ𝐛v(Aϕ,ζ)ζ𝑑ζ=0,\mathrm{Re\,}\,e^{2i\phi}\int_{\mathbf{a}_{*}}\frac{v(A_{\phi},\zeta)}{\zeta}d\zeta=\mathrm{Re\,}\,e^{2i\phi}\int_{\mathbf{b}_{*}}\frac{v(A_{\phi},\zeta)}{\zeta}d\zeta=0,

in which

v(Aϕ,ζ):=e2iϕw(Aϕ,eiϕζ)=ζ4+4ζ3+4ζ2+4Aϕζ,v(A_{\phi},\zeta):=e^{-2i\phi}w(A_{\phi},e^{i\phi}\zeta)=\sqrt{\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A_{\phi}\zeta},

and for each ϕ,\phi\in\mathbb{R}, AϕA_{\phi} is a unique solution of (2.3) as in Proposition 8.15. Let 𝔓(u)=𝔓(u,Aϕ)\mathfrak{P}(u)=\mathfrak{P}(u,A_{\phi}) be the elliptic function defined by

0𝔓(u,Aϕ)dζv(Aϕ,ζ)=u.\int^{\mathfrak{P}(u,A_{\phi})}_{0}\frac{d\zeta}{v(A_{\phi},\zeta)}=u.

Then it is easy to see that P(eiϕu;Aϕ)=eiϕ𝔓(u;Aϕ)\mathrm{P}(e^{-i\phi}u;A_{\phi})=e^{i\phi}\mathfrak{P}(u;A_{\phi}), and that

Ω𝐚=𝐚dζv(Aϕ,ζ)=eiϕΩ𝐚,Ω𝐛=𝐛dζv(Aϕ,ζ)=eiϕΩ𝐛\Omega_{\mathbf{a}_{*}}^{*}=\int_{\mathbf{a}_{*}}\frac{d\zeta}{v(A_{\phi},\zeta)}=e^{i\phi}\Omega_{\mathbf{a}},\quad\Omega_{\mathbf{b}_{*}}^{*}=\int_{\mathbf{b}_{*}}\frac{d\zeta}{v(A_{\phi},\zeta)}=e^{i\phi}\Omega_{\mathbf{b}}

are the periods of ΠAϕ\Pi_{A_{\phi}}^{*}. Then the solutions given above are also written as follows.

Corollary 2.3.

Suppose that 0<|ϕ|<π/40<|\phi|<\pi/4, and that (1+s1s2)(1+s2s3)10(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, ηϕ(𝐬)0,\eta_{\phi}(\mathbf{s})\not=0, where ηϕ(𝐬)=1+s1s2\eta_{\phi}(\mathbf{s})=1+s_{1}s_{2} if π/4<ϕ<0-\pi/4<\phi<0, and =(1+s2s3)1=(1+s_{2}s_{3})^{-1} if 0<ϕ<π/4.0<\phi<\pi/4. Then

y(𝐬,x)=x𝔓(12x2+χ0+O(xδ);Aϕ),\displaystyle y(\mathbf{s},x)=x\mathfrak{P}(\tfrac{1}{2}x^{2}+\chi_{0}^{*}+O(x^{-\delta});A_{\phi}),
χ0Ω𝐚2πiln((1+s1s2)(1+s2s3)1)\displaystyle\chi_{0}^{*}\equiv\frac{\Omega_{\mathbf{a}_{*}}^{*}}{2\pi i}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)
Ω𝐛2πilnηϕ(𝐬)+Ω𝐛3(αβ)modΩ𝐚+Ω𝐛\displaystyle\phantom{----------}-\frac{\Omega_{\mathbf{b}_{*}}^{*}}{2\pi i}\ln\eta_{\phi}(\mathbf{s})+\frac{\Omega_{\mathbf{b}_{*}}^{*}}{3}(\alpha-\beta)\mod\Omega_{\mathbf{a}_{*}}^{*}\mathbb{Z}+\Omega_{\mathbf{b}_{*}}^{*}\mathbb{Z}

as xx\to\infty through the cheese-like strip

S0(ϕ,t,κ0,δ0)={x=eiϕ2t|Ret>t,|Imt|<κ0}σ𝒫0{|xσ|<δ0}S_{0}^{*}(\phi,t_{\infty},\kappa_{0},\delta_{0})=\{x=e^{i\phi}\sqrt{2t}\,|\,\mathrm{Re\,}t>t_{\infty},\,\,\,|\mathrm{Im\,}t|<\kappa_{0}\}\setminus\bigcup_{\sigma\in\mathcal{P}_{0}^{*}}\{|x-\sigma|<\delta_{0}\}

with 𝒫0={eiϕ2t0|e2iϕt0+χ0=±13Ω𝐛+Ω𝐚+Ω𝐛}.\mathcal{P}_{0}^{*}=\{e^{i\phi}\sqrt{2t_{0}^{*}}\,|\,e^{2i\phi}t^{*}_{0}+\chi^{*}_{0}=\pm\tfrac{1}{3}\Omega^{*}_{\mathbf{b}_{*}}+\Omega^{*}_{\mathbf{a}_{*}}\mathbb{Z}+\Omega^{*}_{\mathbf{b}_{*}}\mathbb{Z}\}.

The dependence on ϕ\phi of v(Aϕ,ζ)v(A_{\phi},\zeta) is on AϕA_{\phi} only such that Aϕ+π/2=AϕA_{\phi+\pi/2}=A_{\phi}, and hence so is that of Ω𝐚\Omega_{\mathbf{a}_{*}}^{*}, Ω𝐛\Omega_{\mathbf{b}_{*}}^{*} and 𝔓(u,Aϕ)\mathfrak{P}(u,A_{\phi}). Solutions in general directions are given as follows.

Theorem 2.4.

For any n𝐙n\in\mathbf{Z}, let 𝐬n=(s1+n,s2+n,s3+n,s4+n)\mathbf{s}_{n}=(s_{1+n},s_{2+n},s_{3+n},s_{4+n}). Suppose that 0<|ϕπn/2|<π/40<|\phi-\pi n/2|<\pi/4, and that (1+s1+ns2+n)(1+s2+ns3+n)10,(1+s_{1+n}s_{2+n})(1+s_{2+n}s_{3+n})-1\not=0, ηϕ(𝐬n)0,\eta_{\phi}(\mathbf{s}_{n})\not=0, where ηϕ(𝐬n)=1+s1+ns2+n\eta_{\phi}(\mathbf{s}_{n})=1+s_{1+n}s_{2+n} if π/4<ϕπn/2<0-\pi/4<\phi-\pi n/2<0, and =(1+s2+ns3+n)1=(1+s_{2+n}s_{3+n})^{-1} if 0<ϕπn/2<π/4.0<\phi-\pi n/2<\pi/4. Then PIV\mathrm{P}_{\mathrm{IV}} admits a solution represented as follows::

y(𝐬n,x)=x𝔓(12x2+χ0n+O(xδ);Aϕ),\displaystyle y(\mathbf{s}_{n},x)=x\mathfrak{P}(\tfrac{1}{2}x^{2}+\chi_{0}^{n}+O(x^{-\delta});A_{\phi}),
χ0nΩ𝐚2πiln((1+s1+ns2+n)(1+s2+ns3+n)1)\displaystyle\chi_{0}^{n}\equiv\frac{\Omega_{\mathbf{a}_{*}}^{*}}{2\pi i}\ln((1+s_{1+n}s_{2+n})(1+s_{2+n}s_{3+n})-1)
Ω𝐛2πilnηϕ(𝐬n)+Ω𝐛3(αβ)modΩ𝐚+Ω𝐛\displaystyle\phantom{----------}-\frac{\Omega_{\mathbf{b}_{*}}^{*}}{2\pi i}\ln\eta_{\phi}(\mathbf{s}_{n})+\frac{\Omega_{\mathbf{b}_{*}}^{*}}{3}(\alpha-\beta)\mod\Omega_{\mathbf{a}_{*}}^{*}\mathbb{Z}+\Omega_{\mathbf{b}_{*}}^{*}\mathbb{Z}

as xx\to\infty through the cheese-like strip

S0n(ϕ,t,κ0,δ0)={x=eiϕ2t|Ret>t,|Imt|<κ0}σ𝒫0n{|xσ|<δ0}S_{0}^{n}(\phi,t_{\infty},\kappa_{0},\delta_{0})=\{x=e^{i\phi}\sqrt{2t}\,|\,\mathrm{Re\,}t>t_{\infty},\,\,\,|\mathrm{Im\,}t|<\kappa_{0}\}\setminus\bigcup_{\sigma\in\mathcal{P}_{0}^{n}}\{|x-\sigma|<\delta_{0}\}

with 𝒫0n={eiϕ2t0n|e2iϕt0n+χ0n=±13Ω𝐛+Ω𝐚+Ω𝐛}.\mathcal{P}_{0}^{n}=\{e^{i\phi}\sqrt{2t_{0}^{n}}\,|\,e^{2i\phi}t^{n}_{0}+\chi^{n}_{0}=\pm\tfrac{1}{3}\Omega^{*}_{\mathbf{b}_{*}}+\Omega^{*}_{\mathbf{a}_{*}}\mathbb{Z}+\Omega^{*}_{\mathbf{b}_{*}}\mathbb{Z}\}.

Remark 2.4.

By the single-valuedness on \mathbb{C}, y(𝐬n,x)y(\mathbf{s}_{n},x) coincides with y(𝐬n+4,x)y(\mathbf{s}_{n+4},x). Indeed sksk+1=sk+4sk+5s_{k}s_{k+1}=s_{k+4}s_{k+5} for every k.k\in\mathbb{Z}.

2.3. Observation related to trigonometric asymptotics

Let us observe trigonometric asymptotics as degeneration of our elliptic solutions. For elliptic expressions of solutions in Theorems 2.1, 2.2 or Corollary 2.3 we may calculate at least formal leading terms of the analytic continuations to strips lying along the positive real axis, which is expected to behave trigonometrically as in [13, Theorem 4]. Such problems were discussed in [21, Section 4] for the first and the second Painlevé transcendents.

Write the strip S0(ϕ,t,κ0,δ0)S_{0}(\phi,t_{\infty},\kappa_{0},\delta_{0}) in Theorem 2.1 in the form

S0(ϕ,t,κ0,δ0)\displaystyle S_{0}(\phi,t_{\infty},\kappa_{0},\delta_{0}) =eiϕ2T(t,κ0))Pϕ=2e2iϕT(t,κ0))Pϕ,\displaystyle=e^{i\phi}\sqrt{2T(t_{\infty},\kappa_{0}))}\setminus P_{\phi}=\sqrt{2e^{2i\phi}T(t_{\infty},\kappa_{0}))}\setminus P_{\phi},
T(t,κ0)\displaystyle T(t_{\infty},\kappa_{0}) ={t|Ret>t,|Imt|<κ0},\displaystyle=\{t\,|\,\mathrm{Re\,}t>t_{\infty},\,\,|\mathrm{Im\,}t|<\kappa_{0}\},

where 2T={x=2t|tT}\sqrt{2T}=\{x=\sqrt{2t}\,|\,t\in T\} and Pϕ=σ𝒫0{|xσ|<δ0}P_{\phi}=\bigcup_{\sigma\in\mathcal{P}_{0-}}\{|x-\sigma|<\delta_{0}\}. Let us suppose that there exists a strip 𝒮0\mathcal{S}_{0} such that

𝒮0={t|Ret>t0,K0<Imt<K1}ϕ0<ϕ<0e2iϕT(t,κ0),\mathcal{S}_{0}=\{t\,|\,\mathrm{Re\,}t>t_{\infty}^{0},\,\,-K_{0}<\mathrm{Im\,}t<-K_{1}\}\subset\bigcup_{-\phi_{0}<\phi<0}e^{2i\phi}T(t_{\infty},\kappa_{0}),

where K0K_{0}, K1K_{1} are positive constants and ϕ0\phi_{0} a small positive constant. Then 2𝒮0\sqrt{2\mathcal{S}_{0}} is a strip lying along the line 2t0<x<+\sqrt{2t^{0}_{\infty}}<x<+\infty, and has the properties: for each ϕ\phi,

(1) every t𝒮0e2iϕT(t,κ0)t\in\mathcal{S}_{0}\cap e^{2i\phi}T(t_{\infty},\kappa_{0}) fulfils tϕ1t\phi\asymp 1 with implied constants independent of ϕ\phi and tt;

(2) 2(𝒮0e2iϕT(t,κ0))PϕS0(ϕ,t,κ0,δ0).\sqrt{2(\mathcal{S}_{0}\cap e^{2i\phi}T(t_{\infty},\kappa_{0}))}\setminus P_{\phi}\subset S_{0}(\phi,t_{\infty},\kappa_{0},\delta_{0}).

By the property (2) the expression of y(𝐬,x)y(\mathbf{s},x) in Theorem 2.1 is valid in each region 𝒮0e2iϕT(t,κ0)\mathcal{S}_{0}\cap e^{2i\phi}T(t_{\infty},\kappa_{0}). By Remark 2.1 this is written in the form

(2.4) (2t)1/2y(𝐬,x)=e3iϕAϕ(τ;g2,g3)13e2iϕ,τ=eiϕt+χ0+O(tδ).(2t)^{-1/2}y(\mathbf{s},x)=\frac{e^{3i\phi}A_{\phi}}{\wp(\tau;g_{2},g_{3})-\tfrac{1}{3}e^{2i\phi}},\quad\tau=e^{i\phi}t+\chi_{0-}+O(t^{-\delta}).

Note that, by Corollary 8.20 and the property (1),

Ω𝐚=eiϕΩ𝐚=123ilnt(1+o(1)),Ω𝐛=eiϕΩ𝐛=3π(1+o(1)),\Omega_{\mathbf{a}}=e^{-i\phi}\Omega^{*}_{\mathbf{a}_{*}}=\tfrac{1}{2}\sqrt{3}i\ln t\,(1+o(1)),\quad\Omega_{\mathbf{b}}=e^{-i\phi}\Omega^{*}_{\mathbf{b}_{*}}=-\sqrt{3}\pi(1+o(1)),

and χ0=143π1ln((1+s1s2)(1+s2s3)1)lnti123(ln(1+s1s2)23πi(αβ))\chi_{0-}=\tfrac{1}{4}\sqrt{3}\pi^{-1}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)\ln t-i\tfrac{1}{2}\sqrt{3}(\ln(1+s_{1}s_{2})-\tfrac{2}{3}\pi i(\alpha-\beta)). The periods of (τ)\wp(\tau) are given by (2ω,2ω)=(3π(1+o(1)),123ilnt(1+o(1))(2\omega,2\omega^{\prime})=(\sqrt{3}\pi(1+o(1)),\tfrac{1}{2}\sqrt{3}i\ln t\,(1+o(1)) such that Im(ω/ω)=(2π)1lnt(1+o(1)).\mathrm{Im\,}(\omega^{\prime}/\omega)=(2\pi)^{-1}\ln t\,(1+o(1)). In (2.4), we have ϕ=O(t1)\phi=O(t^{-1}) and Aϕ=827+O(t1)A_{\phi}=\tfrac{8}{27}+O(t^{-1}) (cf. Proposition 8.19) as tt\to\infty, and then (u)\wp(u) degenerates to

(2.5) (u)=13ω^2+ω^2sin2(ω^u)+O(hcos2(2ω^u))or\displaystyle\wp(u)=-\tfrac{1}{3}\hat{\omega}^{-2}+\hat{\omega}^{-2}\sin^{-2}(\hat{\omega}u)+O(h\cos^{2}(2\hat{\omega}u))\quad\text{or}
(2.6) (u)=13ω^28ω^2hcos(2ω^(uω))+O(h2cos2(2ω^(uω)))\displaystyle\wp(u)=-\tfrac{1}{3}\hat{\omega}^{2}-8\hat{\omega}^{2}h\cos(2\hat{\omega}(u-\omega^{\prime}))+O(h^{2}\cos^{2}(2\hat{\omega}(u-\omega^{\prime})))

with ω^=π/(2ω)\hat{\omega}=\pi/(2\omega) and h=eiπω/ωh=e^{i\pi{\omega^{\prime}/\omega}} [8], [32]. In 𝒮0(ϕ0<ϕ<0e2iϕT(t,κ0))\mathcal{S}_{0}\cap(\bigcup_{-\phi_{0}<\phi<0}e^{2i\phi}T(t_{\infty},\kappa_{0})), from (2.5) we have the trigonometric expression

1(2t)1/2y(𝐬,x)+23\displaystyle\frac{1}{(2t)^{-1/2}y(\mathbf{s},x)+\tfrac{2}{3}} =12+12(ei2τ/3+ei2τ/3)(1+O(tε)),\displaystyle=\frac{1}{2}+\frac{1}{2}(e^{i2\tau/\sqrt{3}}+e^{-i2\tau/\sqrt{3}})(1+O(t^{-\varepsilon})),
2τ/3=2t/3\displaystyle{2\tau}/{\sqrt{3}}={2t}/{\sqrt{3}} +(2π)1ln((1+s1s2)(1+s2s3)1)lnt23π(αβ)+O(1),\displaystyle+({2\pi})^{-1}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)\ln t-\tfrac{2}{3}\pi(\alpha-\beta)+O(1),

which agrees with [13, Theorem 4, (30)] up to constants. From (2.6) we have

y(𝐬,x)+23\displaystyle y(\mathbf{s},x)+\tfrac{2}{3} =2t1/2(ei2τ^/3+ei2τ^/3)(1+O(tε)),\displaystyle=2t^{-1/2}(e^{i2\hat{\tau}/\sqrt{3}}+e^{-i2\hat{\tau}/\sqrt{3}})(1+O(t^{-\varepsilon})),
2τ^/3\displaystyle{2\hat{\tau}}/{\sqrt{3}} =2t/3+(2π)1ln(1(1+s1s2)(1+s2s3))lnt23π(αβ)+O(1),\displaystyle={2t}/{\sqrt{3}}+({2\pi})^{-1}\ln(1-(1+s_{1}s_{2})(1+s_{2}s_{3}))\ln t-\tfrac{2}{3}\pi(\alpha-\beta)+O(1),

which agrees with [13, Theorem 4, (28)] and also with [9, Theorem 1.1, (1.11)] up to constants. In the strip 0<K0<Imt<K10<K_{0}<\mathrm{Im\,}t<K_{1} a similar argument is possible for the solution in Theorem 2.2. The argument above, though not justified, suggests information about degeneration to the trigonometric asymptotics.

3. Basic facts

3.1. Monodromy data

The monodromy data and the monodromy manifold 0(α,β)\mathcal{M}^{*}_{0}(\alpha,\beta) described in Section 2 have the following properties.

Proposition 3.1.

For each mm\in\mathbb{Z},

S1+mS2+mS3+mS4+meiπ(αβ)σ3σ3\displaystyle S_{1+m}S_{2+m}S_{3+m}S_{4+m}e^{-i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}
={(ES1Sm)1σ3M1(ES1Sm)if m1;(ES01Sm+11)1σ3M1(ES01Sm+11)if m1.\displaystyle\phantom{---}=\begin{cases}(ES_{1}\cdots S_{m})^{-1}\sigma_{3}M^{-1}(ES_{1}\cdots S_{m})\quad&\text{if $m\geq 1;$}\\[11.38092pt] (ES_{0}^{-1}\cdots S_{m+1}^{-1})^{-1}\sigma_{3}M^{-1}(ES_{0}^{-1}\cdots S_{m+1}^{-1})\quad&\text{if $m\leq-1.$}\end{cases}
Proof..

The semi-cyclic condition S1S2S3S4=E1σ3M1Eeiπ(αβ)σ3σ3S_{1}S_{2}S_{3}S_{4}=E^{-1}\sigma_{3}M^{-1}Ee^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3} with Sk+4=eiπ(αβ)σ3σ3Skσ3eiπ(αβ)σ3S_{k+4}=e^{-i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}S_{k}\sigma_{3}e^{i\pi(\alpha-\beta)\sigma_{3}} yields

S2S3S4S1+4=\displaystyle S_{2}S_{3}S_{4}S_{1+4}= S11E1σ3M1Eeiπ(αβ)σ3σ3eiπ(αβ)σ3σ3S1σ3eiπ(αβ)σ3\displaystyle S_{1}^{-1}E^{-1}\sigma_{3}M^{-1}Ee^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}e^{-i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}S_{1}\sigma_{3}e^{i\pi(\alpha-\beta)\sigma_{3}}
=\displaystyle= (ES1)1σ3M1(ES1)eiπ(αβ)σ3σ3,\displaystyle(ES_{1})^{-1}\sigma_{3}M^{-1}(ES_{1})e^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3},
S0S1S2S3=\displaystyle S_{0}S_{1}S_{2}S_{3}= S0E1σ3M1Eeiπ(αβ)σ3σ3S41\displaystyle S_{0}E^{-1}\sigma_{3}M^{-1}Ee^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}S_{4}^{-1}
=\displaystyle= (ES01)1σ3M1ES01S0eiπ(αβ)σ3σ3S41\displaystyle(ES_{0}^{-1})^{-1}\sigma_{3}M^{-1}ES_{0}^{-1}S_{0}e^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}S_{4}^{-1}
=\displaystyle= (ES01)1σ3M1(ES01)eiπ(αβ)σ3σ3.\displaystyle(ES_{0}^{-1})^{-1}\sigma_{3}M^{-1}(ES_{0}^{-1})e^{i\pi(\alpha-\beta)\sigma_{3}}\sigma_{3}.

Repeating this procedure, we obtain the proposition. ∎

Proposition 3.2.

The surface 0(α,β)\mathcal{M}^{*}_{0}(\alpha,\beta) has a singular point 𝐬sing\mathbf{s}_{\mathrm{sing}} if and only if α12\alpha-\tfrac{1}{2}\in\mathbb{Z}, and then 𝐬sing=(𝔰12,𝔰23,𝔰34,𝔰41)=(eiπβ1,eiπβ1,eiπβ1,eiβπe2iβπ).\mathbf{s}_{\mathrm{sing}}=(\mathfrak{s}_{12},\mathfrak{s}_{23},\mathfrak{s}_{34},\mathfrak{s}_{41})=(e^{-i\pi\beta}-1,e^{i\pi\beta}-1,e^{-i\pi\beta}-1,e^{-i\beta\pi}-e^{-2i\beta\pi}).

Proof..

Write the surface in the form

f=eiπ(βα)(xz+u1)eiπ(αβ)y+2isinπα=0,(x1)(z1)=(y1)(u1)f=e^{i\pi(\beta-\alpha)}(xz+u-1)-e^{i\pi(\alpha-\beta)}y+2i\sin\pi\alpha=0,\quad(x-1)(z-1)=(y-1)(u-1)

with (x,y,z,u)=(1+𝔰12,1+𝔰23,1+𝔰34,1+𝔰41),(x,y,z,u)=(1+\mathfrak{s}_{12},1+\mathfrak{s}_{23},1+\mathfrak{s}_{34},1+\mathfrak{s}_{41}), and examine when fx,f_{x}, fy,f_{y}, fzf_{z}, fuf_{u} and ff have a common zero. If, say y10y-1\not=0, by using u1=(x1)(z1)/(y1)u-1=(x-1)(z-1)/(y-1), we have α12\alpha-\tfrac{1}{2}\in\mathbb{Z} and x=z=1/y=eiπβ.x=z=1/y=e^{-i\pi\beta}.

Remark 3.1.

Equation PIV admits a one-parameter family of classical solutions (respectively, a rational solution) if and only if {α12,12β,α12β12}\{\alpha-\tfrac{1}{2},\,\tfrac{1}{2}\beta,\,\alpha-\tfrac{1}{2}\beta-\tfrac{1}{2}\}\cap\mathbb{Z}\not=\emptyset (respectively, {(α12,12β),(α±16,2α12β)}2\{(\alpha-\tfrac{1}{2},\tfrac{1}{2}\beta),(\alpha\pm\tfrac{1}{6},2\alpha-\tfrac{1}{2}\beta)\}\cap\mathbb{Z}^{2}\not=\emptyset) [5], [6], [7, Chap. 6], [22], [30].

3.2. Isomonodromy linear system

Let us transform system (1.1) into a form suitable to our calculation. By

Ψ=u(1/2)σ3x(1/4)σ3Φ,uv=y,uxu=xz,\Psi=\mathrm{u}^{(1/2)\sigma_{3}}x^{-(1/4)\sigma_{3}}\Phi,\quad\mathrm{u}\mathrm{v}=y,\quad\frac{\mathrm{u}_{x}}{\mathrm{u}}=x\mathrm{z},

system (1.1) is changed into

dΦdξ=\displaystyle\frac{d\Phi}{d\xi}= ((ξ32+(x+y)ξ+αξ)σ3+ix1/2(0ξ2+xz+2xx1(yξ2xyz+y2+β)0))Φ,\displaystyle\Biggl{(}\Bigl{(}\frac{\xi^{3}}{2}+(x+y)\xi+\frac{\alpha}{\xi}\Bigr{)}\sigma_{3}+ix^{1/2}\begin{pmatrix}0&\xi^{2}+x\mathrm{z}+2x\\ x^{-1}(y\xi^{2}-xy\mathrm{z}+y^{2}+\beta)&0\end{pmatrix}\Biggr{)}\Phi,
y=\displaystyle y^{\prime}= 2xyz+2xyy2β.\displaystyle 2xy\mathrm{z}+2xy-y^{2}-\beta.

The change of variables

τ=x2,y=xη,ξ=x1/2ξ~\tau=x^{2},\quad y=x\eta,\quad\xi=x^{1/2}\tilde{\xi}

takes this system to

dΦdξ~=\displaystyle\frac{d\Phi}{d\tilde{\xi}}= τ((ξ~32+(1+η)ξ~+ατ1ξ~)σ3+i(0ξ~2+z+2ηξ~2ηz+η2+βτ10))Φ,\displaystyle\tau\Biggl{(}\Bigl{(}\frac{\tilde{\xi}^{3}}{2}+(1+\eta)\tilde{\xi}+\frac{\alpha\tau^{-1}}{\tilde{\xi}}\Bigr{)}\sigma_{3}+i\begin{pmatrix}0&\tilde{\xi}^{2}+\mathrm{z}+2\\ \eta\tilde{\xi}^{2}-\eta\mathrm{z}+\eta^{2}+\beta\tau^{-1}&0\end{pmatrix}\Biggr{)}\Phi,
τdηdτ=\displaystyle\tau\frac{d\eta}{d\tau}= τη(z+1)12(τη2+η+β).\displaystyle\tau\eta(\mathrm{z}+1)-\frac{1}{2}({\tau\eta^{2}}+\eta+\beta).

The further substitution

τ=2e2iϕt,η=eiϕψ,eiϕz=𝔷,λ=eiϕ/2ξ~\tau=2e^{2i\phi}t,\quad\eta=e^{-i\phi}\psi,\quad e^{i\phi}\mathrm{z}=\mathfrak{z},\quad\lambda=e^{i\phi/2}\tilde{\xi}

leads to

(3.1) dΦdλ=t(t,λ)Φ,(t,λ)=b3σ3+b1σ1+b2σ2,\displaystyle\frac{d\Phi}{d\lambda}=t\mathcal{B}(t,\lambda)\Phi,\qquad\mathcal{B}(t,\lambda)=b_{3}\sigma_{3}+b_{1}\sigma_{1}+b_{2}\sigma_{2},
(3.2) 2(𝔷+eiϕ)=eiϕψtψ+ψ+12(eiϕ+βψ1)t1\displaystyle 2(\mathfrak{z}+e^{i\phi})=e^{-i\phi}\frac{\psi_{t}}{\psi}+\psi+\tfrac{1}{2}(e^{-i\phi}+\beta\psi^{-1})t^{-1}

with

b1=ieiϕ/2((ψ+eiϕ)λ2(ψeiϕ)𝔷+ψ2+2e2iϕ+12βt1),\displaystyle b_{1}=ie^{-i\phi/2}\left((\psi+e^{i\phi})\lambda^{2}-(\psi-e^{i\phi})\mathfrak{z}+\psi^{2}+2e^{2i\phi}+\tfrac{1}{2}\beta t^{-1}\right),
b2=eiϕ/2((ψeiϕ)λ2(ψ+eiϕ)𝔷+ψ22e2iϕ+12βt1),\displaystyle b_{2}=e^{-i\phi/2}\left((\psi-e^{i\phi})\lambda^{2}-(\psi+e^{i\phi})\mathfrak{z}+\psi^{2}-2e^{2i\phi}+\tfrac{1}{2}\beta t^{-1}\right),
b3=λ3+2(eiϕ+ψ)λ+αt1λ1\displaystyle b_{3}=\lambda^{3}+2(e^{i\phi}+\psi)\lambda+\alpha t^{-1}\lambda^{-1}

In (3.2) as a resulting equation, ψt\psi_{t} denotes the derivative (d/dt)ψ.(d/dt)\psi. Let us now change the meaning of ψt\psi_{t} in such a way that ψt\psi_{t} is an arbitrary function not necessarily the derivative, and in what follows suppose that system (3.1) is equipped with 𝔷\mathfrak{z} containing such ψt.\psi_{t}. Then the isomonodromy property of (1.1) is converted to that of (3.1).

Proposition 3.3.

The monodromy data of (3.1) is invariant under a small change of tt if and only if ψt=(d/dt)ψ\psi_{t}=(d/dt)\psi holds in (3.2) and y(x)=eiϕxψy(x)=e^{-i\phi}x\psi with 2e2iϕt=x22e^{2i\phi}t=x^{2} solves PIV\mathrm{P}_{\mathrm{IV}}.

For kk\in\mathbb{Z} system (3.1) admits canonical solutions

(3.3) Φk(λ)=(I+O(λ1))exp((14tλ4+eiϕtλ2+(αβ)lnλ)σ3)\Phi_{k}^{\infty}(\lambda)=(I+O(\lambda^{-1}))\exp((\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3})

as λ\lambda\to\infty through the sector |arg(t1/4λ)+π8π4k|<π4|\arg(t^{1/4}\lambda)+\tfrac{\pi}{8}-\tfrac{\pi}{4}k|<\tfrac{\pi}{4}. The Stokes matrices are defined by Φk+1(λ)=Φk(λ)Sk\Phi_{k+1}^{\infty}(\lambda)=\Phi_{k}^{\infty}(\lambda)S_{k}^{*}. Recalling the Stokes matrices SkS_{k} with respect to Ψk(ξ)\Psi^{\infty}_{k}(\xi) solving (1.1), we have the following relation.

Proposition 3.4.

For every kk\in\mathbb{Z},

Sk=u(1/2)σ3x(1/4)σ3(eiϕ/2x1/2)(αβ)σ3Sk(eiϕ/2x1/2)(αβ)σ3u(1/2)σ3x(1/4)σ3.S_{k}=\mathrm{u}^{(1/2)\sigma_{3}}x^{-(1/4)\sigma_{3}}(e^{-i\phi/2}x^{1/2})^{-(\alpha-\beta)\sigma_{3}}S_{k}^{*}(e^{-i\phi/2}x^{1/2})^{(\alpha-\beta)\sigma_{3}}\mathrm{u}^{-(1/2)\sigma_{3}}x^{(1/4)\sigma_{3}}.
Proof..

The relation Ψk(ξ)=u(1/2)σ3x(1/4)σ3Φk(λ)(eiϕ/2x1/2)(αβ)σ3u(1/2)σ3x(1/4)σ3\Psi^{\infty}_{k}(\xi)=\mathrm{u}^{(1/2)\sigma_{3}}x^{-(1/4)\sigma_{3}}\Phi^{\infty}_{k}(\lambda)(e^{-i\phi/2}x^{1/2})^{(\alpha-\beta)\sigma_{3}}\mathrm{u}^{-(1/2)\sigma_{3}}x^{(1/4)\sigma_{3}} yields the conclusion. ∎

The Stokes coefficients sks^{*}_{k} for Φk(λ)\Phi^{\infty}_{k}(\lambda) are given by

S2l1=(1s2l101),S2l=(10s2l1)(l).S_{2l-1}^{*}=\begin{pmatrix}1&s^{*}_{2l-1}\\ 0&1\end{pmatrix},\quad S_{2l}^{*}=\begin{pmatrix}1&0\\ s^{*}_{2l}&1\end{pmatrix}\quad(l\in\mathbb{Z}).
Corollary 3.5.

For any k,jk,j\in\mathbb{Z}, sksk+2j+1=sksk+2j+1s_{k}^{*}s_{k+2j+1}^{*}=s_{k}s_{k+2j+1}.

4. Turning points, Stokes graph, WKB analysis

For system (3.1) we will treat the direct monodromy problem by WKB analysis. Let us start with the characteristic roots ±μ(t,λ)\pm\mu(t,\lambda) of (t,λ)\mathcal{B}(t,\lambda) constituting the essential part of the WKB solution (cf. Proposition 4.1). By μ(t,λ)2=b12+b22+b32\mu(t,\lambda)^{2}=b_{1}^{2}+b_{2}^{2}+b_{3}^{2}, we have

(4.1) μ(t,λ)2=λ6+4eiϕλ4+(4e2iϕ+2(αβ)t1)λ2+4e3iϕaϕ+α2t2λ2\mu(t,\lambda)^{2}=\lambda^{6}+4e^{i\phi}\lambda^{4}+(4e^{2i\phi}+2(\alpha-\beta)t^{-1})\lambda^{2}+4e^{3i\phi}a_{\phi}+\alpha^{2}t^{-2}\lambda^{-2}

with

(4.2) 4e3iϕaϕ=\displaystyle 4e^{3i\phi}a_{\phi}= 4e3iϕaϕ(t)=e2iϕ(ψt)2ψ1(ψ+2eiϕ)2ψ\displaystyle 4e^{3i\phi}a_{\phi}(t)=e^{-2i\phi}{(\psi_{t})^{2}}{\psi^{-1}}-(\psi+2e^{i\phi})^{2}\psi
+(e2iϕψt+(4αβ)ψ+2(2αβ)eiϕ)t1+14(e2iϕψβ2ψ1)t2.\displaystyle+(e^{-2i\phi}\psi_{t}+(4\alpha-\beta)\psi+2(2\alpha-\beta)e^{i\phi})t^{-1}+\tfrac{1}{4}(e^{-2i\phi}\psi-\beta^{2}\psi^{-1})t^{-2}.

To draw Stokes graphs it is necessary to know the location of the turning points. Now we note the following facts on the solution AϕA_{\phi} of the Boutroux equation (2.2) for |ϕ|<π/4|\phi|<\pi/4 (Proposition 8.15):

(i) A0=827,A_{0}=\frac{8}{27}, and then w(A0,z)=z(z+23)2(z+83);w(A_{0},z)=z(z+\frac{2}{3})^{2}(z+\frac{8}{3});

(ii) A±π/4=0,A_{\pm\pi/4}=0, and then w(A±π/4,z)=z2(z+2e±iπ/4)2;w(A_{\pm\pi/4},z)=z^{2}(z+2e^{\pm i\pi/4})^{2};

(iii) for 0<|ϕ|<π/4,0<|\phi|<\pi/4, 0<ReAϕ<8270<\mathrm{Re\,}A_{\phi}<\frac{8}{27} and w(Aϕ,z)w(A_{\phi},z) does not degenerate.

Then by Corollary 8.2, for |ϕ|<π/4|\phi|<\pi/4, the zeros of w(Aϕ,z)w(A_{\phi},z) may be numbered in such a way that Reeiϕz5Reeiϕz3Reeiϕz1\mathrm{Re\,}e^{-i\phi}z_{5}\leq\mathrm{Re\,}e^{-i\phi}z_{3}\leq\mathrm{Re\,}e^{-i\phi}z_{1} and that (z1,z3,z5)(23,23,83)(z_{1},z_{3},z_{5})\to(-\frac{2}{3},-\frac{2}{3},-\frac{8}{3}) as ϕ0\phi\to 0, and (0,2e±iπ/4,2e±iπ/4)\to(0,-2e^{\pm i\pi/4},-2e^{\pm i\pi/4}) as ϕ±π/4\phi\to\pm\pi/4 (cf. Section 2).

Our WKB analysis is carried out under the supposition aϕ(t)=Aϕ+O(t1)a_{\phi}(t)=A_{\phi}+O(t^{-1}) as tt\to\infty, that is, (5.1). The characteristic root satisfies λμ(t,λ)w(Aϕ,λ2)\lambda\mu(t,\lambda)\to w(A_{\phi},\lambda^{2}) as tt\to\infty. Let λj\lambda_{j} (1j6)(1\leq j\leq 6) be turning points of μ(t,λ)\mu(t,\lambda) such that

λ1z11/2,λ3z31/2,λ5z51/2ast,\displaystyle\lambda_{1}\to z_{1}^{1/2},\,\,\,\lambda_{3}\to z_{3}^{1/2},\,\,\,\lambda_{5}\to z_{5}^{1/2}\,\,\,\text{as}\,\,t\to\infty,
λ2=λ1,λ4=λ3,λ6=λ5,\displaystyle\lambda_{2}=-\lambda_{1},\,\,\,\lambda_{4}=-\lambda_{3},\,\,\,\lambda_{6}=-\lambda_{5},
|argλjπ2|<π4(j=1,3,5)for sufficiently large t.\displaystyle|\arg\lambda_{j}-\tfrac{\pi}{2}|<\tfrac{\pi}{4}\,\,\,(j=1,3,5)\,\,\,\text{for sufficiently large $t$}.

The algebraic function μ(t,λ)\mu(t,\lambda) is written in the form

μ(t,λ)\displaystyle\mu(t,\lambda) =λ6+4eiϕλ4+(4e2iϕ+2(αβ)t1)λ2+4e3iϕaϕ+α2t2λ2\displaystyle=\sqrt{\smash{\lambda^{6}}+4e^{i\phi}\lambda^{4}+(4e^{2i\phi}+2(\alpha-\beta)t^{-1})\lambda^{2}+4e^{3i\phi}\smash{a_{\phi}}+\alpha^{2}t^{-2}\lambda^{-2}}
=λ1(λ2λ12)(λ2λ32)(λ2λ52)(λ2λ02)\displaystyle=\lambda^{-1}\sqrt{(\lambda^{2}-\lambda_{1}^{2})(\lambda^{2}-\lambda_{3}^{2})(\lambda^{2}-\lambda_{5}^{2})(\lambda^{2}-\lambda_{0}^{2})}
=λ1(λλ1)(λλ2)(λλ3)(λλ4)(λλ5)(λλ6)(λλ0)(λ+λ0),\displaystyle=\lambda^{-1}\sqrt{(\lambda-\lambda_{1})(\lambda-\lambda_{2})(\lambda-\lambda_{3})(\lambda-\lambda_{4})(\lambda-\lambda_{5})(\lambda-\lambda_{6})(\lambda-\lambda_{0})(\lambda+\lambda_{0})},
λ0=O(t1),\displaystyle\lambda_{0}=O(t^{-1}),

which is considered on the two sheeted Riemann surface t=t+t\mathcal{R}_{t}=\mathcal{R}^{+}_{t}\cup\mathcal{R}^{-}_{t} glued along the cuts [λ5,λ3][\lambda_{5},\lambda_{3}], [λ1,λ0],[\lambda_{1},\lambda_{0}], [λ0,λ2][-\lambda_{0},\lambda_{2}], [λ4,λ6][\lambda_{4},\lambda_{6}] with t±=([λ5,λ3][λ1,λ0][λ0,λ2][λ4,λ6])\mathcal{R}^{\pm}_{t}=\mathbb{C}\setminus([\lambda_{5},\lambda_{3}]\cup[\lambda_{1},\lambda_{0}]\cup[-\lambda_{0},\lambda_{2}]\cup[\lambda_{4},\lambda_{6}]). The branches of μ(,λ)\mu(\infty,\lambda) with aϕ()=Aϕa_{\phi}(\infty)=A_{\phi} (by (5.1)) is chosen in such a way that μ(,λ)=λ3(1+O(λ2)\mu(\infty,\lambda)=\lambda^{3}(1+O(\lambda^{-2}) as λ\lambda\to\infty on the upper sheet t+.\mathcal{R}^{+}_{t}. For 0<|ϕ|<π/40<|\phi|<\pi/4, w(Aϕ,z)w(A_{\phi},z) does not degenerate and neither does t\mathcal{R}_{t}. Then each turning point satisfies λj(t)λj()=O(t1)\lambda_{j}(t)-\lambda_{j}(\infty)=O(t^{-1}) as t.t\to\infty.

In what follows we treat a Stokes graph with t=t=\infty, and the limit turning point λj()\lambda_{j}(\infty) is simply denoted by λj\lambda_{j}. By z=λ2z=\lambda^{2} the algebraic function w(Aϕ,z)w(A_{\phi},z) on ΠAϕ,ϕ\Pi_{A_{\phi},\phi} is mapped to λμ(,λ)\lambda\mu(\infty,\lambda) on \mathcal{R}_{\infty}, in which

μ(,λ)=\displaystyle\mu(\infty,\lambda)= λ6+4eiϕλ4+4e2iϕλ2+4e3iϕAϕ=(λ2λ12)(λ2λ32)(λ2λ52)\displaystyle\sqrt{\lambda^{6}+4e^{i\phi}\lambda^{4}+4e^{2i\phi}\lambda^{2}+4e^{3i\phi}A_{\phi}}=\sqrt{(\lambda^{2}-\lambda_{1}^{2})(\lambda^{2}-\lambda_{3}^{2})(\lambda^{2}-\lambda_{5}^{2})}
=\displaystyle= (λ2z1)(λ2z3)(λ2z5).\displaystyle\sqrt{(\lambda^{2}-z_{1})(\lambda^{2}-z_{3})(\lambda^{2}-z_{5})}.

The Stokes graph on \mathcal{R}_{\infty} consists of vertices and Stokes curves, where the Stokes curve is defined by Reλjλμ(,λ)𝑑λ=0,\mathrm{Re\,}\int^{\lambda}_{\lambda_{j}}\mu(\infty,\lambda)d\lambda=0, and the vertices are turning points and singular points. In our case the turning points and the Stokes curves have the following properties:

(i) if ϕ=0\phi=0, then λ1=λ3=136i,\lambda_{1}=\lambda_{3}=\tfrac{1}{3}\sqrt{6}i, λ5=236i,\lambda_{5}=\frac{2}{3}\sqrt{6}i, that is, λ1,3=136i\lambda_{1,3}=\tfrac{1}{3}\sqrt{6}i is a double turning point, and if ϕ=±π/4\phi=\pm\pi/4, then λ1=0,\lambda_{1}=0, λ3=λ5=2ie±iπ/8\lambda_{3}=\lambda_{5}=\sqrt{2}ie^{\pm i\pi/8}, that is, λ3,5=2ie±iπ/8\lambda_{3,5}=\sqrt{2}ie^{\pm i\pi/8} is a double turning point;

(ii) if ϕ\phi is close to 0, then the double turning point λ1,3\lambda_{1,3} resolves into

λ2±1=136i±ϕei3π/4+O(ϕ2)for ϕ>0,\displaystyle\lambda_{2\pm 1}=\tfrac{1}{3}\sqrt{6}i\pm\phi_{*}e^{i3\pi/4}+O(\phi_{*}^{2})\quad\text{for $\phi>0$},
λ2±1=136i±ϕeiπ/4+O(ϕ2)for ϕ<0,\displaystyle\lambda_{2\pm 1}=\tfrac{1}{3}\sqrt{6}i\pm\phi_{*}e^{i\pi/4}+O(\phi_{*}^{2})\quad\text{for $\phi<0$},

where ϕ=ϕ(ϕ)\phi_{*}=\phi_{*}(\phi) is such that ϕ0\phi_{*}\geq 0 and ϕ=o(1)\phi_{*}=o(1) as ϕ0\phi\to 0 (cf. Propositions 8.17 and 8.18);

(iii) by the Boutroux equations (2.2) with z=λ2z=\lambda^{2},

Reλ1λ3μ(,λ)𝑑λ=0,Reλ3λ5μ(,λ)𝑑λ=0,\mathrm{Re\,}\int_{\lambda_{1}}^{\lambda_{3}}\mu(\infty,\lambda)d\lambda=0,\quad\mathrm{Re\,}\int_{\lambda_{3}}^{\lambda_{5}}\mu(\infty,\lambda)d\lambda=0,

implying the existence of Stokes curves joining λ1\lambda_{1} to λ3\lambda_{3}, and λ3\lambda_{3} to λ5\lambda_{5};

(iv) the Stokes curves tending to \infty are asymptotic to the rays argλ=π8+π4k\arg\lambda=\frac{\pi}{8}+\frac{\pi}{4}k (1k8)(1\leq k\leq 8).

Taking these facts into account, we may draw the limit Stokes graphs for |ϕ|<π/4|\phi|<\pi/4 as in Figure 4.1, in which the cuts [λ5,λ3][\lambda_{5},\lambda_{3}], [λ1,λ2][\lambda_{1},\lambda_{2}], [λ4,λ6][\lambda_{4},\lambda_{6}] are omitted.

λ5\lambda_{5}λ6\lambda_{6}λ3\lambda_{3}λ1\lambda_{1}λ4\lambda_{4}λ2\lambda_{2}0π/4<ϕ<0-\pi/4<\phi<0
λ5\lambda_{5}λ6\lambda_{6}λ3\lambda_{3}λ1\lambda_{1}λ4\lambda_{4}λ2\lambda_{2}00<ϕ<π/40<\phi<\pi/4
λ3,5\lambda_{3,5}λ1,2\lambda_{1,2}λ4,6\lambda_{4,6}0ϕ=π/4\phi=-\pi/4
λ5\lambda_{5}λ6\lambda_{6}λ1,3\lambda_{1,3}λ2,4\lambda_{2,4}0ϕ=0\phi=0
λ3,5\lambda_{3,5}λ1,2\lambda_{1,2}λ4,6\lambda_{4,6}0ϕ=π/4\phi=\pi/4
Figure 4.1. Limit Stokes graphs on +\mathcal{R}_{\infty}^{+}

An unbounded domain DD\subset\mathcal{R}_{\infty} is said to be canonical if, for every λD\lambda\in D, there exist contours C±(λ)D{C}_{\pm}(\lambda)\subset D terminating in λ\lambda such that

Reλ0λμ(λ)𝑑λ,Reλ0+λμ(λ)𝑑λ+\mathrm{Re\,}\int_{\lambda^{-}_{0}}^{\lambda}\mu(\lambda)d\lambda\to-\infty,\quad\mathrm{Re\,}\int_{\lambda^{+}_{0}}^{\lambda}\mu(\lambda)d\lambda\to+\infty

as λ0\lambda_{0}^{-}\to\infty along C(λ)C_{-}(\lambda) and as λ0+\lambda_{0}^{+}\to\infty along C+(λ)C_{+}(\lambda), respectively (see [3], [4]). The interior of a canonical domain contains exactly one Stokes curve, and the boundary consists of Stokes curves. System (3.1) admits the following WKB solution ([3], [4, Theorem 7.2], [29, Proposition 3.8]).

Proposition 4.1.

In a canonical domain system (3.1) with (t,λ)=b1σ1+b2σ2+b3σ3\mathcal{B}(t,\lambda)=b_{1}\sigma_{1}+b_{2}\sigma_{2}+b_{3}\sigma_{3} admits an asymptotic solution expressed as

ΨWKB(λ)=T(I+O(tδ))exp(λλΛ(τ)𝑑τ),T=(1b3μb1+ib2μb3b1ib21),\Psi_{\mathrm{WKB}}(\lambda)=T(I+O(t^{-\delta}))\exp\Bigl{(}\int^{\lambda}_{\lambda_{*}}\Lambda(\tau)d\tau\Bigr{)},\quad T=\begin{pmatrix}1&\frac{b_{3}-\mu}{b_{1}+ib_{2}}\\ \frac{\mu-b_{3}}{b_{1}-ib_{2}}&1\end{pmatrix},

as long as |b1±ib2|11,|b_{1}\pm ib_{2}|^{-1}\ll 1, |ψ|+|𝔷|1,|\psi|+|\mathfrak{z}|\ll 1, |λλj|t(2/3)(1δ)|\lambda-\lambda_{j}|\gg t^{-(2/3)(1-\delta)} (λj:(\lambda_{j}: a simple turning point)) with given implied constants. Here δ\delta is an arbitrary number such that 0<δ<10<\delta<1, λ\lambda_{*} is a fixed base point, and

Λ(λ)=tμ(t,λ)σ3diagT1Tλ,\displaystyle\Lambda(\lambda)=t\mu(t,\lambda)\sigma_{3}-\mathrm{diag}T^{-1}T_{\lambda},
diagT1Tλ=14(1b3μ)λlnb1+ib2b1ib2σ3+12λlnμμ+b3I.\displaystyle\mathrm{diag}T^{-1}T_{\lambda}=\frac{1}{4}\Bigl{(}1-\frac{b_{3}}{\mu}\Bigr{)}\frac{\partial}{\partial\lambda}\ln\frac{b_{1}+ib_{2}}{b_{1}-ib_{2}}\sigma_{3}+\frac{1}{2}\frac{\partial}{\partial\lambda}\ln\frac{\mu}{\mu+b_{3}}I.
Remark 4.1.

In the WKB solution we write Λ(λ)\Lambda(\lambda) in the component-wise form Λ(λ)=Λ3(λ)+ΛI(λ)\Lambda(\lambda)=\Lambda_{3}(\lambda)+\Lambda_{I}(\lambda), Λ3(λ)σ3,\Lambda_{3}(\lambda)\in\mathbb{C}\sigma_{3}, ΛI(λ)I\Lambda_{I}(\lambda)\in\mathbb{C}I with

Λ3(λ)=tμ(t,λ)σ3diagT1Tλ|σ3σ3,ΛI(λ)=diagT1Tλ|II.\Lambda_{3}(\lambda)=t\mu(t,\lambda)\sigma_{3}-\mathrm{diag}T^{-1}T_{\lambda}|_{\sigma_{3}}\sigma_{3},\quad\Lambda_{I}(\lambda)=-\mathrm{diag}T^{-1}T_{\lambda}|_{I}I.

The WKB solution fails in expressing asymptotics in a neighbourhood of a turning point. Around a simple turning point equation (3.1) is reduced to the system

(4.3) dWdζ=(01ζ0)W\frac{dW}{d\zeta}=\begin{pmatrix}0&1\\ \zeta&0\end{pmatrix}W

having solutions (Ai(ζ),Aiζ(ζ))T,{}^{T}(\mathrm{Ai}(\zeta),\mathrm{Ai}_{\zeta}(\zeta)), (Bi(ζ),Biζ(ζ))T{}^{T}(\mathrm{Bi}(\zeta),\mathrm{Bi}_{\zeta}(\zeta)) [4, Theorem 7.3], [29, Proposition 3.9], where Ai(ζ)\mathrm{Ai}(\zeta) is the Airy function and Bi(ζ)=eπi/6Ai(e2πi/3ζ)\mathrm{Bi}(\zeta)=e^{-\pi i/6}\mathrm{Ai}(e^{-2\pi i/3}\zeta) [1], [2].

Proposition 4.2.

Let λj\lambda_{j} be a simple turning point, and write ck=bk(λj),c_{k}=b_{k}(\lambda_{j}), ck=(bk)λ(λj)c^{\prime}_{k}=(b_{k})_{\lambda}(\lambda_{j}) (k=1,2,3)(k=1,2,3) and κc=c1c1+c2c2+c3c3.\kappa_{c}=c_{1}c_{1}^{\prime}+c_{2}c_{2}^{\prime}+c_{3}c_{3}^{\prime}. Suppose that ck,c_{k}, ckc_{k}^{\prime} are bounded and c1±ic20.c_{1}\pm ic_{2}\not=0. Then (3.1) admits a matrix solution of the form

Φj(λ)=Tj(I+O(tδ))(100t^1)W(ζ),Tj=(1c3c1+ic2c3c1ic21),\Phi_{j}(\lambda)=T_{j}(I+O(t^{-\delta^{\prime}}))\begin{pmatrix}1&0\\ 0&\hat{t}^{-1}\end{pmatrix}W(\zeta),\quad T_{j}=\begin{pmatrix}1&-\frac{c_{3}}{c_{1}+ic_{2}}\\ -\frac{c_{3}}{c_{1}-ic_{2}}&1\end{pmatrix},

in which t^=2(2κc)1/3(c1ic2)(t/4)1/3\hat{t}=2(2\kappa_{c})^{-1/3}(c_{1}-ic_{2})(t/4)^{1/3} and λλj=(2κc)1/3(t/4)2/3(ζ+ζ0)\lambda-\lambda_{j}=(2\kappa_{c})^{-1/3}(t/4)^{-2/3}(\zeta+\zeta_{0}) with |ζ0|t1/3,|\zeta_{0}|\ll t^{-1/3}, as long as |ζ|t(2/3δ)/3,|\zeta|\ll t^{(2/3-\delta^{\prime})/3}, that is, |λλj|t2/3+(2/3δ)/3.|\lambda-\lambda_{j}|\ll t^{-2/3+(2/3-\delta^{\prime})/3}. Here δ\delta^{\prime} is a given number such that 0<δ<2/3,0<\delta^{\prime}<2/3, and W(ζ)W(\zeta) solves (4.3), which admits canonical solutions Wν(ζ)W_{\nu}(\zeta) (ν)(\nu\in\mathbb{Z}) such that

Wν(ζ)=ζ(1/4)σ3(σ3+σ1)(I+O(ζ3/2))exp(23ζ3/2σ3)W_{\nu}(\zeta)=\zeta^{-(1/4)\sigma_{3}}(\sigma_{3}+\sigma_{1})(I+O(\zeta^{-3/2}))\exp(\tfrac{2}{3}\zeta^{3/2}\sigma_{3})

as ζ\zeta\to\infty through the sector |argζ(2ν1)π3|<2π3,|\arg\zeta-(2\nu-1)\frac{\pi}{3}|<\frac{2\pi}{3}, and that Wν+1(ζ)=Wν(ζ)GνW_{\nu+1}(\zeta)=W_{\nu}(\zeta)G_{\nu} with

G1=(1i01),G2=(10i1),Gν+1=σ1Gνσ1.G_{1}=\begin{pmatrix}1&-i\\ 0&1\end{pmatrix},\quad G_{2}=\begin{pmatrix}1&0\\ -i&1\end{pmatrix},\quad G_{\nu+1}=\sigma_{1}G_{\nu}\sigma_{1}.
Remark 4.2.

The matrix solutions ΨWKB(λ)\Psi_{\mathrm{WKB}}(\lambda) and Φj(λ)\Phi_{j}(\lambda) in Propositions 4.1 and 4.2 are simultaneously valid in the annulus 𝒜j:\mathcal{A}_{j}: t2/3+ε|λλj|t2/3+2εt^{-2/3+\varepsilon}\ll|\lambda-\lambda_{j}|\ll t^{-2/3+2\varepsilon} with, say, ε=23δ=16(23δ)=110\varepsilon=\frac{2}{3}\delta=\frac{1}{6}(\frac{2}{3}-\delta^{\prime})=\frac{1}{10}, in which the matching is possible.

5. Direct monodromy problem

By WKB analysis we calculate Stokes matrices of linear system (3.1) as a solution of the direct monodromy problem. Recall aϕ(t)a_{\phi}(t) given by (4.2) and a unique solution AϕA_{\phi} of the Boutroux equations (2.2) for 0<|ϕ|<π/40<|\phi|<\pi/4. In our calculation suppose that ψ\psi and ψt\psi_{t} are arbitrary functions and that

(5.1) aϕ(t)=Aϕ+Bϕ(t)t,Bϕ(t)1a_{\phi}(t)=A_{\phi}+\frac{B_{\phi}(t)}{t},\quad B_{\phi}(t)\ll 1

as tt\to\infty in the strip

Sϕ(t,κ1,δ1)={t|Ret>t,|Imt|<κ1,|ψ(t)|+|ψ(t)|1+|ψt(t)|<δ11},S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1})=\{t\,|\,\mathrm{Re\,}t>t^{\prime}_{\infty},\,\,\,|\mathrm{Im\,}t|<\kappa_{1},\,\,\,|\psi(t)|+|\psi(t)|^{-1}+|\psi_{t}(t)|<\delta_{1}^{-1}\},

where κ1\kappa_{1} is a given number, δ1\delta_{1} a given small number and tt_{\infty}^{\prime} a sufficiently large number.

𝐜13\mathbf{c}^{3}_{\infty_{1}}𝐜35\mathbf{c}^{5}_{3}𝐜53\mathbf{c}_{5}^{\infty_{3}}λ5\lambda_{5}λ3\lambda_{3}λ1\lambda_{1}0𝔠π/85π/8=𝐜13𝐜35𝐜53\mathfrak{c}^{5\pi/8}_{\pi/8}=\mathbf{c}_{\infty_{1}}^{3}\cup\mathbf{c}_{3}^{5}\cup\mathbf{c}_{5}^{\infty_{3}}
𝐜13\mathbf{c}_{\infty_{1}}^{3}𝐜31\mathbf{c}_{3}^{1}𝐜14\mathbf{c}^{\infty_{4}}_{1}λ5\lambda_{5}λ3\lambda_{3}λ1\lambda_{1}0𝔠π/87π/8=𝐜13𝐜31𝐜14\mathfrak{c}^{7\pi/8}_{\pi/8}=\mathbf{c}_{\infty_{1}}^{3}\cup\mathbf{c}_{3}^{1}\cup\mathbf{c}_{1}^{\infty_{4}}
Figure 5.1. Fragments of the Stokes graph for π/4<ϕ<0-\pi/4<\phi<0

Let π/4<ϕ<0.-\pi/4<\phi<0. Let us calculate the analytic continuations of the matrix solution Φ1(λ)\Phi_{1}^{\infty}(\lambda) given by (3.3) in the sector |arg(tλ)π8|<π4|\arg(t\lambda)-\tfrac{\pi}{8}|<\tfrac{\pi}{4} along the fragments of the Stokes graph

𝔠π/85π/8=𝐜13𝐜35𝐜53and𝔠π/87π/8=𝐜13𝐜31𝐜14\mathfrak{c}_{\pi/8}^{5\pi/8}=\mathbf{c}_{\infty_{1}}^{3}\cup\mathbf{c}_{3}^{5}\cup\mathbf{c}_{5}^{\infty_{3}}\,\,\,\text{and}\,\,\,\mathfrak{c}_{\pi/8}^{7\pi/8}=\mathbf{c}_{\infty_{1}}^{3}\cup\mathbf{c}_{3}^{1}\cup\mathbf{c}_{1}^{\infty_{4}}

with 𝐜13=(eiπ/8,λ3),\mathbf{c}_{\infty_{1}}^{3}=(e^{i\pi/8}\infty,\lambda_{3})^{\sim}, 𝐜35=(λ3,λ5),\mathbf{c}_{3}^{5}=(\lambda_{3},\lambda_{5})^{\sim}, 𝐜53=(λ5,e5π/8),\mathbf{c}_{5}^{\infty_{3}}=(\lambda_{5},e^{5\pi/8})^{\sim}, 𝐜31=(λ3,λ1),\mathbf{c}_{3}^{1}=(\lambda_{3},\lambda_{1})^{\sim}, 𝐜14=(λ1,e7π/8),\mathbf{c}_{1}^{\infty_{4}}=(\lambda_{1},e^{7\pi/8})^{\sim}, where (p,q)(p,q)^{\sim} denotes a curve joining pp to qq, and 𝐜35\mathbf{c}_{3}^{5} lies on the right shore of the cut [λ3,λ5][\lambda_{3},\lambda_{5}] (cf. Figure 5.1).

The Stokes matrix S1S2=Φ1(λ)1Φ3(λ)S^{*}_{1}S^{*}_{2}=\Phi_{1}^{\infty}(\lambda)^{-1}\Phi_{3}^{\infty}(\lambda) follows from the analytic continuation of Φ1(λ)\Phi_{1}^{\infty}(\lambda) along 𝔠π/85π/8\mathfrak{c}^{5\pi/8}_{\pi/8}, which is calculated by the matching procedure as below. In the steps (2), (3), (6), (7), analytic continuations are considered in the annulus 𝒜3\mathcal{A}_{3} and 𝒜5\mathcal{A}_{5}, which may be given by δ=320,\delta=\tfrac{3}{20}, δ=115\delta^{\prime}=\tfrac{1}{15} as in Remark 4.2. Thus, in what follows we set δ=320.\delta=\tfrac{3}{20}.

(1) For a WKB solution Ψ3()(λ)\Psi^{(\infty)}_{3}(\lambda) along 𝐜13\mathbf{c}_{\infty_{1}}^{3} with a base point λ~3\tilde{\lambda}_{3} such that λ~3λ3t1\tilde{\lambda}_{3}-\lambda_{3}\asymp t^{-1}, set Φ1(λ)=Ψ3()(λ)Γ,3\Phi_{1}^{\infty}(\lambda)=\Psi^{(\infty)}_{3}(\lambda)\Gamma_{\infty,3}. Then

Γ,3=\displaystyle\Gamma_{\infty,3}= Ψ3()(λ)1Φ1(λ)\displaystyle\Psi^{(\infty)}_{3}(\lambda)^{-1}\Phi_{1}^{\infty}(\lambda)
=\displaystyle= C3(λ3~)cI(λ3~)(I+O(tδ))\displaystyle C_{3}(\tilde{\lambda_{3}})c_{I}(\tilde{\lambda_{3}})(I+O(t^{-\delta}))
×exp(limλλ𝐜13(λ3λΛ3(τ)𝑑τ(14tλ4+eiϕtλ2+(αβ)lnλ)σ3)),\displaystyle\times\exp\biggl{(}-\lim_{\begin{subarray}{c}\lambda\to\infty\\[1.42271pt] \lambda\in\mathbf{c}_{\infty_{1}}^{3}\end{subarray}}\biggl{(}\int^{\lambda}_{\lambda_{3}}\Lambda_{3}(\tau)d\tau-(\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3}\biggr{)}\biggr{)},

where C3(λ~3)=exp(λ3λ~3Λ3(τ)𝑑τ),C_{3}(\tilde{\lambda}_{3})=\exp(\int^{\tilde{\lambda}_{3}}_{\lambda_{3}}\Lambda_{3}(\tau)d\tau), cI(λ~3)=exp(λ~3ΛI(τ)𝑑τ).c_{I}(\tilde{\lambda}_{3})=\exp(-\int^{\infty}_{\tilde{\lambda}_{3}}\Lambda_{I}(\tau)d\tau).

(2) For a canonical solution Φ3+(λ)\Phi_{3}^{+}(\lambda) along 𝐜13𝒜3\mathbf{c}^{3}_{\infty_{1}}\cap\mathcal{A}_{3} (cf. Remark 4.2), set Ψ3()(λ)=Φ3+(λ)Γ3+\Psi^{(\infty)}_{3}(\lambda)=\Phi_{3}^{+}(\lambda)\Gamma^{+}_{3}. Then

Γ3+=Φ3+(λ)1Ψ3()(λ)=(ζ~3)1/4(I+O(tδ))C3(λ~3)1(100c1ic2c3),\Gamma^{+}_{3}=\Phi_{3}^{+}(\lambda)^{-1}\Psi^{(\infty)}_{3}(\lambda)=(\tilde{\zeta}_{3})^{1/4}(I+O(t^{-\delta}))C_{3}(\tilde{\lambda}_{3})^{-1}\begin{pmatrix}1&0\\ 0&-\frac{c_{1}-ic_{2}}{c_{3}}\end{pmatrix},

where ck=bk(λ3)c_{k}=b_{k}(\lambda_{3}) and ζ~3λ~3λ3\tilde{\zeta}_{3}\asymp\tilde{\lambda}_{3}-\lambda_{3} is suitably chosen.

(3) For a canonical solution Φ3(λ)\Phi_{3}^{-}(\lambda) along 𝐜35𝒜3\mathbf{c}^{5}_{3}\cap\mathcal{A}_{3}, set Φ3+(λ)=Φ3(λ)Γ(3)\Phi_{3}^{+}(\lambda)=\Phi_{3}^{-}(\lambda)\Gamma_{(3)}. Then

Γ(3)=Φ3(λ)1Φ3+(λ)=G11=(1i01).\Gamma_{(3)}=\Phi_{3}^{-}(\lambda)^{-1}\Phi_{3}^{+}(\lambda)=G_{1}^{-1}=\begin{pmatrix}1&i\\ 0&1\end{pmatrix}.

(4) For a WKB solution Ψ3(5)(λ)\Psi_{3}^{(5)}(\lambda) along 𝐜35\mathbf{c}_{3}^{5} with a base point λ3\lambda_{3}^{\prime} such that λ3λ3t1\lambda_{3}^{\prime}-\lambda_{3}\asymp t^{-1}, set Φ3(λ)=Ψ3(5)(λ)Γ3\Phi_{3}^{-}(\lambda)=\Psi_{3}^{(5)}(\lambda)\Gamma_{3}^{-}. Then

Γ3=Ψ3(5)(λ)1Φ3(λ)=(ζ3)1/4(I+O(tδ))C3(λ3)(100c3c1ic2),\Gamma_{3}^{-}=\Psi_{3}^{(5)}(\lambda)^{-1}\Phi_{3}^{-}(\lambda)=(\zeta_{3}^{\prime})^{-1/4}(I+O(t^{-\delta}))C^{\prime}_{3}(\lambda_{3}^{\prime})\begin{pmatrix}1&0\\ 0&-\frac{c_{3}}{c_{1}-ic_{2}}\end{pmatrix},

where C3(λ3)=exp(λ3λ3Λ3(τ)𝑑τ)C_{3}^{\prime}(\lambda_{3}^{\prime})=\exp(\int^{\lambda_{3}^{\prime}}_{\lambda_{3}}\Lambda_{3}(\tau)d\tau) and ζ3λ3λ3\zeta_{3}^{\prime}\asymp\lambda_{3}^{\prime}-\lambda_{3} is suitably chosen.

(5) For a WKB solution Ψ5(3)(λ)\Psi_{5}^{(3)}(\lambda) along 𝐜35\mathbf{c}_{3}^{5} with a base point λ5\lambda_{5}^{\prime} such that λ5λ5t1\lambda_{5}^{\prime}-\lambda_{5}\asymp t^{-1}, set Ψ3(5)(λ)=Ψ5(3)(λ)Γ3,5\Psi_{3}^{(5)}(\lambda)=\Psi_{5}^{(3)}(\lambda)\Gamma_{3,5}. Then

Γ3,5=\displaystyle\Gamma_{3,5}= Ψ5(3)(λ)1Ψ3(5)(λ)\displaystyle\Psi_{5}^{(3)}(\lambda)^{-1}\Psi_{3}^{(5)}(\lambda)
=\displaystyle= C3(λ3)1C3(λ5)cI(λ3,λ5)(I+O(tδ))exp(λ3λ5Λ3(τ)𝑑τ),\displaystyle C_{3}^{\prime}(\lambda^{\prime}_{3})^{-1}C_{3}^{\prime}(\lambda^{\prime}_{5})c_{I}(\lambda_{3}^{\prime},\lambda_{5}^{\prime})(I+O(t^{-\delta}))\exp\biggl{(}\int_{\lambda_{3}}^{\lambda_{5}}\Lambda_{3}(\tau)d\tau\biggr{)},

where C3(λ5)=exp(λ5λ5Λ3(τ)𝑑τ)C_{3}(\lambda_{5}^{\prime})=\exp(\int^{\lambda_{5}^{\prime}}_{\lambda_{5}}\Lambda_{3}(\tau)d\tau) and cI(λ3,λ5)=exp(λ3λ5ΛI(τ)𝑑τ).c_{I}(\lambda^{\prime}_{3},\lambda^{\prime}_{5})=\exp(\int^{\lambda^{\prime}_{5}}_{\lambda^{\prime}_{3}}\Lambda_{I}(\tau)d\tau).

(6) For a canonical solution Φ5+(λ)\Phi_{5}^{+}(\lambda) along 𝐜35𝒜5\mathbf{c}_{3}^{5}\cap\mathcal{A}_{5}, set Ψ5(3)(λ)=Φ5+(λ)Γ5+.\Psi_{5}^{(3)}(\lambda)=\Phi_{5}^{+}(\lambda)\Gamma_{5}^{+}. Then

Γ5+=Φ5+(λ)1Ψ5(3)(λ)=(ζ5)1/4(I+O(tδ))C3(λ5)1(100d1id2d3),\Gamma_{5}^{+}=\Phi_{5}^{+}(\lambda)^{-1}\Psi_{5}^{(3)}(\lambda)=(\zeta_{5}^{\prime})^{1/4}(I+O(t^{-\delta}))C_{3}(\lambda^{\prime}_{5})^{-1}\begin{pmatrix}1&0\\ 0&-\frac{d_{1}-id_{2}}{d_{3}}\end{pmatrix},

where dk=bk(λ5)d_{k}=b_{k}(\lambda_{5}) and ζ5λ5λ5\zeta^{\prime}_{5}\asymp\lambda^{\prime}_{5}-\lambda_{5} is suitably chosen.

(7) For a canonical solution Φ5(λ)\Phi_{5}^{-}(\lambda) along 𝐜53𝒜5\mathbf{c}_{5}^{\infty_{3}}\cap\mathcal{A}_{5}, set Φ5+(λ)=Φ5(λ)Γ(5)\Phi_{5}^{+}(\lambda)=\Phi_{5}^{-}(\lambda)\Gamma_{(5)}. Then

Γ(5)=Φ5(λ)1Φ5+(λ)=(G1G2)1=(1ii0).\Gamma_{(5)}=\Phi_{5}^{-}(\lambda)^{-1}\Phi_{5}^{+}(\lambda)=(G_{1}G_{2})^{-1}=\begin{pmatrix}1&i\\ i&0\end{pmatrix}.

(8) For a WKB solution Ψ5()(λ)\Psi_{5}^{(\infty)}(\lambda) along 𝐜53\mathbf{c}_{5}^{\infty_{3}} with a base point λ~5\tilde{\lambda}_{5} such that λ~5λ5t1,\tilde{\lambda}_{5}-\lambda_{5}\asymp t^{-1}, set Φ5(λ)=Ψ5()(λ)Γ5\Phi_{5}^{-}(\lambda)=\Psi_{5}^{(\infty)}(\lambda)\Gamma_{5}^{-}. Then

Γ5=Ψ5()(λ)1Φ5(λ)=(ζ~5)1/4(I+O(tδ))C3(λ~5)(100d3d1id2),\Gamma_{5}^{-}=\Psi_{5}^{(\infty)}(\lambda)^{-1}\Phi_{5}^{-}(\lambda)=(\tilde{\zeta}_{5})^{-1/4}(I+O(t^{-\delta}))C_{3}(\tilde{\lambda}_{5})\begin{pmatrix}1&0\\ 0&-\frac{d_{3}}{d_{1}-id_{2}}\end{pmatrix},

where C3(λ~5)=exp(λ5λ~5Λ3(τ)𝑑τ)C_{3}(\tilde{\lambda}_{5})=\exp(\int^{\tilde{\lambda}_{5}}_{\lambda_{5}}\Lambda_{3}(\tau)d\tau) and ζ~5λ~5λ5\tilde{\zeta}_{5}\asymp\tilde{\lambda}_{5}-\lambda_{5} is suitably chosen.

(9) For Φ3(λ)\Phi_{3}^{\infty}(\lambda) solving (3.1) in the sector |argλ5π8|<π4|\arg\lambda-\frac{5\pi}{8}|<\frac{\pi}{4}, set Ψ5()(λ)=Φ3(λ)Γ5,.\Psi_{5}^{(\infty)}(\lambda)=\Phi_{3}^{\infty}(\lambda)\Gamma_{5,\infty}. Then

Γ5,=\displaystyle\Gamma_{5,\infty}= Φ3(λ)1Ψ5()(λ)\displaystyle\Phi_{3}^{\infty}(\lambda)^{-1}\Psi_{5}^{(\infty)}(\lambda)
=\displaystyle= C3(λ~5)1cI(λ~5)1(I+O(tδ))\displaystyle C_{3}(\tilde{\lambda}_{5})^{-1}c_{I}(\tilde{\lambda}_{5})^{-1}(I+O(t^{-\delta}))
×exp(limλλ𝐜53(λ5λΛ3(τ)𝑑τ(14tλ4+eiϕtλ2+(αβ)lnλ)σ3)),\displaystyle\times\exp\biggl{(}\lim_{\begin{subarray}{c}\lambda\to\infty\\[1.42271pt] \lambda\in\mathbf{c}_{5}^{\infty_{3}}\end{subarray}}\biggl{(}\int_{\lambda_{5}}^{\lambda}\Lambda_{3}(\tau)d\tau-(\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3}\biggr{)}\biggr{)},

where cI(λ~5)=exp(λ~5ΛI(τ)𝑑τ).c_{I}(\tilde{\lambda}_{5})=\exp(-\int^{\infty}_{\tilde{\lambda}_{5}}\Lambda_{I}(\tau)d\tau).

Product of the matrices above along 𝔠π/85π/8\mathfrak{c}^{5\pi/8}_{\pi/8} yields the Stokes matrix

(S1S2)1=\displaystyle(S_{1}^{*}S_{2}^{*})^{-1}= (1s1s21+s1s2)\displaystyle\begin{pmatrix}1&-s_{1}^{*}\\ -s_{2}^{*}&1+s_{1}^{*}s_{2}^{*}\end{pmatrix}
=\displaystyle= Φ3(λ)1Φ1(λ)\displaystyle\Phi_{3}^{\infty}(\lambda)^{-1}\Phi_{1}^{\infty}(\lambda)
=\displaystyle= Γ5,Γ5Γ(5)Γ5+Γ3,5Γ3Γ(3)Γ3+Γ,3\displaystyle\Gamma_{5,\infty}\Gamma_{5}^{-}\Gamma_{(5)}\Gamma_{5}^{+}\Gamma_{3,5}\Gamma_{3}^{-}\Gamma_{(3)}\Gamma_{3}^{+}\Gamma_{\infty,3}
=\displaystyle= ϵ1eJ53σ3(100d01)(1ii0)(100d0)\displaystyle\epsilon_{1}e^{J_{5}^{\infty 3}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-d_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&i\\ i&0\end{pmatrix}\begin{pmatrix}1&0\\ 0&-d_{0}\end{pmatrix}
×eJ3,5σ3(100c01)(1i01)(100c0)eJ31σ3\displaystyle\times e^{J_{3,5}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-c_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&i\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ 0&-c_{0}\end{pmatrix}e^{-J_{3}^{\infty 1}\sigma_{3}}
=\displaystyle= ϵ1(eJ3,5+J53J31i(c0eJ3,5+d0eJ3,5)eJ53+J31id01eJ3,5J53J31c0d01eJ3,5J53+J31)\displaystyle\epsilon_{1}\begin{pmatrix}e^{J_{3,5}+J_{5}^{\infty 3}-J_{3}^{\infty 1}}&-i(c_{0}e^{J_{3,5}}+d_{0}e^{-J_{3,5}})e^{J_{5}^{\infty 3}+J_{3}^{\infty 1}}\\ -id_{0}^{-1}e^{J_{3,5}-J_{5}^{\infty 3}-J_{3}^{\infty 1}}&-c_{0}d_{0}^{-1}e^{J_{3,5}-J_{5}^{\infty 3}+J_{3}^{\infty 1}}\end{pmatrix}

(up to the multiplier 1+O(tδ)1+O(t^{-\delta}) to each entry), in which ϵ12=1,\epsilon_{1}^{2}=1,

c0=(c1ic2)/c3,d0=(d1id2)/d3,ck=bk(λ3),dk=bk(λ5),\displaystyle c_{0}=(c_{1}-ic_{2})/c_{3},\quad d_{0}=(d_{1}-id_{2})/d_{3},\quad c_{k}=b_{k}(\lambda_{3}),\,\,\,d_{k}=b_{k}(\lambda_{5}),
(5.2) J3,5σ3=λ3λ5Λ3(τ)𝑑τ,𝐜35=(λ3,λ5): on the right shore of the cut [λ3,λ5],\displaystyle J_{3,5}\sigma_{3}=\int^{\lambda_{5}}_{\lambda_{3}}\Lambda_{3}(\tau)d\tau,\quad\text{$\mathbf{c}_{3}^{5}=(\lambda_{3},\lambda_{5})^{\sim}$: on the right shore of the cut $[\lambda_{3},\lambda_{5}]$},
J53σ3=limλλ𝐜53(λ5λΛ3(τ)𝑑τ(14tλ4+eiϕtλ2+(αβ)lnλ)σ3),\displaystyle J_{5}^{\infty 3}\sigma_{3}=\lim_{\begin{subarray}{c}\lambda\to\infty\\[1.42271pt] \lambda\in\mathbf{c}_{5}^{\infty_{3}}\end{subarray}}\biggl{(}\int_{\lambda_{5}}^{\lambda}\Lambda_{3}(\tau)d\tau-(\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3}\biggr{)},
J31σ3=limλλ𝐜13(λ3λΛ3(τ)𝑑τ(14tλ4+eiϕtλ2+(αβ)lnλ)σ3).\displaystyle J_{3}^{\infty 1}\sigma_{3}=\lim_{\begin{subarray}{c}\lambda\to\infty\\[1.42271pt] \lambda\in\mathbf{c}^{3}_{\infty_{1}}\end{subarray}}\biggl{(}\int_{\lambda_{3}}^{\lambda}\Lambda_{3}(\tau)d\tau-(\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3}\biggr{)}.

Then the diagonal entries of (S1S2)1(S_{1}^{*}S_{2}^{*})^{-1} give 1=ϵ1eJ3,5+J53J31(1+O(tδ)),1=\epsilon_{1}e^{J_{3,5}+J_{5}^{\infty 3}-J_{3}^{\infty 1}}(1+O(t^{-\delta})), 1+s1s2=ϵ1c0d01eJ3,5J53+J31(1+O(tδ)),1+s_{1}^{*}s_{2}^{*}=-\epsilon_{1}c_{0}d_{0}^{-1}e^{J_{3,5}-J_{5}^{\infty 3}+J_{3}^{\infty 1}}(1+O(t^{-\delta})), which imply

(5.3) 1+s1s2=c0d01e2J3,5(1+O(tδ)).1+s_{1}^{*}s_{2}^{*}=-c_{0}d_{0}^{-1}e^{2J_{3,5}}(1+O(t^{-\delta})).

Similarly the analytic continuation of Φ1(λ)\Phi_{1}^{\infty}(\lambda) along 𝔠π/87π/8\mathfrak{c}_{\pi/8}^{7\pi/8} yields

(S1S2S3)1=\displaystyle(S^{*}_{1}S^{*}_{2}S^{*}_{3})^{-1}= (1+s2s3s1s3s1s2s3s21+s1s2)\displaystyle\begin{pmatrix}1+s_{2}^{*}s_{3}^{*}&-s_{1}^{*}-s_{3}^{*}-s_{1}^{*}s_{2}^{*}s_{3}^{*}\\ -s_{2}^{*}&1+s_{1}^{*}s_{2}^{*}\end{pmatrix}
=\displaystyle= Φ4(λ)1Φ1(λ)\displaystyle\Phi_{4}^{\infty}(\lambda)^{-1}\Phi_{1}^{\infty}(\lambda)
=\displaystyle= ϵ2eJ14σ3(100e01)(1i01)(100e0)\displaystyle\epsilon_{2}e^{J_{1}^{\infty 4}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-e_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&i\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ 0&-e_{0}\end{pmatrix}
×eJ3,1σ3(100c01)(10i1)(100c0)eJ31σ3\displaystyle\times e^{J_{3,1}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-c_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&0\\ -i&1\end{pmatrix}\begin{pmatrix}1&0\\ 0&-c_{0}\end{pmatrix}e^{-J_{3}^{\infty 1}\sigma_{3}}
=\displaystyle= ϵ2((eJ3,1+c01e0eJ3,1)eJ14J31ie0eJ3,1+J14+J31ic01eJ3,1J14J31eJ3,1J14+J31),\displaystyle\epsilon_{2}\begin{pmatrix}(e^{J_{3,1}}+c_{0}^{-1}e_{0}e^{-J_{3,1}})e^{J_{1}^{\infty 4}-J_{3}^{\infty 1}}&-ie_{0}e^{-J_{3,1}+J_{1}^{\infty 4}+J_{3}^{\infty 1}}\\ ic_{0}^{-1}e^{-J_{3,1}-J_{1}^{\infty 4}-J_{3}^{\infty 1}}&e^{-J_{3,1}-J_{1}^{\infty 4}+J_{3}^{\infty 1}}\end{pmatrix},

in which ϵ22=1,\epsilon_{2}^{2}=1,

e0=(e1ie2)/e3,ek=bk(λ1),\displaystyle e_{0}=(e_{1}-ie_{2})/e_{3},\quad e_{k}=b_{k}(\lambda_{1}),
(5.4) J3,1σ3=λ3λ1Λ3(τ)𝑑τ,\displaystyle J_{3,1}\sigma_{3}=\int^{\lambda_{1}}_{\lambda_{3}}\Lambda_{3}(\tau)d\tau,
J14σ3=limλλ𝐜14(λ1λΛ3(τ)𝑑τ(14tλ4+eiϕtλ2+(αβ)lnλ)σ3).\displaystyle J_{1}^{\infty 4}\sigma_{3}=\lim_{\begin{subarray}{c}\lambda\to\infty\\[1.42271pt] \lambda\in\mathbf{c}_{1}^{\infty_{4}}\end{subarray}}\biggl{(}\int_{\lambda_{1}}^{\lambda}\Lambda_{3}(\tau)d\tau-(\tfrac{1}{4}t\lambda^{4}+e^{i\phi}t\lambda^{2}+(\alpha-\beta)\ln\lambda)\sigma_{3}\biggr{)}.

Observing the off-diagonal entries, we have (1+s1s2)(1+s2s3)1=c01e0e2J3,1(1+O(tδ))(1+s_{1}^{*}s_{2}^{*})(1+s_{2}^{*}s_{3}^{*})-1=c_{0}^{-1}e_{0}e^{-2J_{3,1}}(1+O(t^{-\delta})). Combining this with (5.3) and using Corollary 3.5, we have the following.

Proposition 5.1.

Suppose that π/4<ϕ<0.-\pi/4<\phi<0. Then

1+s1s2=c0d01e2J3,5(1+O(tδ)),\displaystyle 1+s_{1}s_{2}=-c_{0}d_{0}^{-1}e^{2J_{3,5}}(1+O(t^{-\delta})),
(1+s1s2)(1+s2s3)1=c01e0e2J3,1(1+O(tδ)),\displaystyle(1+s_{1}s_{2})(1+s_{2}s_{3})-1=c_{0}^{-1}e_{0}e^{-2J_{3,1}}(1+O(t^{-\delta})),

where J3,5J_{3,5} and J3,1J_{3,1} are integrals given by (5.2) and (5.4).

In the case where 0<ϕ<π/40<\phi<\pi/4, we calculate the analytic continuations of

Φ2(λ)along𝔠^3π/87π/8=𝐜25𝐜53𝐜34,Φ1(λ)along𝔠^π/87π/8=𝐜11𝐜13𝐜34\Phi_{2}^{\infty}(\lambda)\,\,\,\text{along}\,\,\hat{\mathfrak{c}}_{3\pi/8}^{7\pi/8}=\mathbf{c}_{\infty_{2}}^{5}\cup{\mathbf{c}}_{5}^{3-}\cup\mathbf{c}_{3}^{\infty_{4}},\quad\Phi_{1}^{\infty}(\lambda)\,\,\,\text{along}\,\,\hat{\mathfrak{c}}_{\pi/8}^{7\pi/8}=\mathbf{c}_{\infty_{1}}^{1}\cup\mathbf{c}_{1}^{3}\cup\mathbf{c}_{3}^{\infty_{4}}

with 𝐜11=(eπi/8,λ1)\mathbf{c}_{\infty_{1}}^{1}=(e^{\pi i/8}\infty,\lambda_{1})^{\sim}, 𝐜13=(λ1,λ3)\mathbf{c}_{1}^{3}=(\lambda_{1},\lambda_{3})^{\sim}, 𝐜34=(λ3,e7πi/8)\mathbf{c}_{3}^{\infty_{4}}=(\lambda_{3},e^{7\pi i/8}\infty)^{\sim}, 𝐜25=(e3πi/8,λ5)\mathbf{c}^{5}_{\infty_{2}}=(e^{3\pi i/8}\infty,\lambda_{5})^{\sim}, 𝐜53=(λ5,λ3){\mathbf{c}}_{5}^{3-}=(\lambda_{5},\lambda_{3})^{\sim}, where 𝐜53{\mathbf{c}}_{5}^{3-} lies on the left shore of the cut [λ3,λ5][\lambda_{3},\lambda_{5}] (cf. Figure 5.2).

𝐜34\mathbf{c}_{3}^{\infty_{4}}𝐜53\mathbf{c}_{5}^{3-}𝐜25\mathbf{c}^{5}_{\infty_{2}}λ5\lambda_{5}λ3\lambda_{3}λ1\lambda_{1}0𝔠^3π/87π/8=𝐜25𝐜53𝐜34\hat{\mathfrak{c}}^{7\pi/8}_{3\pi/8}=\mathbf{c}_{\infty_{2}}^{5}\cup\mathbf{c}_{5}^{3-}\cup\mathbf{c}_{3}^{\infty_{4}}
𝐜34\mathbf{c}^{\infty_{4}}_{3}𝐜13\mathbf{c}^{3}_{1}𝐜11\mathbf{c}_{\infty_{1}}^{1}λ5\lambda_{5}λ3\lambda_{3}λ1\lambda_{1}0𝔠^π/87π/8=𝐜11𝐜13𝐜34\hat{\mathfrak{c}}^{7\pi/8}_{\pi/8}=\mathbf{c}_{\infty_{1}}^{1}\cup\mathbf{c}_{1}^{3}\cup\mathbf{c}_{3}^{\infty_{4}}
Figure 5.2. Fragments of the Stokes graph for 0<ϕ<π/40<\phi<\pi/4

This procedure results in

(S2S3)1=\displaystyle(S_{2}^{*}S_{3}^{*})^{-1}= (1+s2s3s3s21)\displaystyle\begin{pmatrix}1+s_{2}^{*}s_{3}^{*}&-s_{3}^{*}\\ -s_{2}^{*}&1\end{pmatrix}
=\displaystyle= Φ4(λ)1Φ2(λ)\displaystyle\Phi_{4}^{\infty}(\lambda)^{-1}\Phi_{2}^{\infty}(\lambda)
=\displaystyle= ϵ~1eJ34σ3(100c01)(1i01)(100c0)\displaystyle\tilde{\epsilon}_{1}e^{J_{3}^{\infty 4}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-c_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&i\\ 0&{1}\end{pmatrix}\begin{pmatrix}1&0\\ 0&-c_{0}\end{pmatrix}
×eJ5,3σ3(100d01)(0ii1)(100d0)eJ52σ3\displaystyle\times e^{J_{5,3}^{-}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-d_{0}^{-1}\end{pmatrix}\begin{pmatrix}0&i\\ i&{1}\end{pmatrix}\begin{pmatrix}1&0\\ 0&-d_{0}\end{pmatrix}e^{-J_{5}^{\infty 2}\sigma_{3}}
=\displaystyle= ϵ~1(c0d01eJ5,3+J34J52i(d0eJ5,3+c0eJ5,3)eJ34+J52id01eJ5,3J34J52eJ5,3J34+J52)\displaystyle\tilde{\epsilon}_{1}\begin{pmatrix}-c_{0}d_{0}^{-1}e^{-J^{-}_{5,3}+J_{3}^{\infty 4}-J_{5}^{\infty 2}}&-i(d_{0}e^{J^{-}_{5,3}}+c_{0}e^{-J^{-}_{5,3}})e^{J_{3}^{\infty 4}+J_{5}^{\infty 2}}\\ -id_{0}^{-1}e^{-J^{-}_{5,3}-J_{3}^{\infty 4}-J_{5}^{\infty 2}}&e^{-J^{-}_{5,3}-J_{3}^{\infty 4}+J_{5}^{\infty 2}}\end{pmatrix}

with ϵ~12=1,\tilde{\epsilon}_{1}^{2}=1, J34σ3=J14σ3|𝐜14𝐜34,λ1λ3J_{3}^{\infty 4}\sigma_{3}=J_{1}^{\infty 4}\sigma_{3}|_{\mathbf{c}_{1}^{\infty_{4}}\mapsto\mathbf{c}_{3}^{\infty_{4}},\,\,\lambda_{1}\mapsto\lambda_{3}}, J52σ3=J31σ3|𝐜13𝐜25,λ3λ5J_{5}^{\infty 2}\sigma_{3}=J_{3}^{\infty 1}\sigma_{3}|_{\mathbf{c}^{3}_{\infty_{1}}\mapsto\mathbf{c}^{5}_{\infty_{2}},\,\,\lambda_{3}\mapsto\lambda_{5}} and

(5.5) J5,3σ3=λ5λ3,Λ3(τ)𝑑τ,𝐜53=(λ5,λ3): on the left shore of the cut [λ3,λ5],J^{-}_{5,3}\sigma_{3}=\int^{\lambda_{3},-}_{\lambda_{5}}\Lambda_{3}(\tau)d\tau,\quad\text{$\mathbf{c}_{5}^{3-}=(\lambda_{5},\lambda_{3})^{\sim}$: on the left shore of the cut $[\lambda_{3},\lambda_{5}]$},

and

(S1S2S3)1=\displaystyle(S_{1}^{*}S_{2}^{*}S_{3}^{*})^{-1}= Φ4(λ)1Φ1(λ)\displaystyle\Phi_{4}^{\infty}(\lambda)^{-1}\Phi_{1}^{\infty}(\lambda)
=\displaystyle= ϵ~2eJ34σ3(100c01)(10i1)(100c0)\displaystyle\tilde{\epsilon}_{2}e^{J_{3}^{\infty 4}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-c_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&0\\ -i&{1}\end{pmatrix}\begin{pmatrix}1&0\\ 0&-c_{0}\end{pmatrix}
×eJ1,3σ3(100e01)(1i01)(100e0)eJ11σ3\displaystyle\times e^{J_{1,3}\sigma_{3}}\begin{pmatrix}1&0\\ 0&-e_{0}^{-1}\end{pmatrix}\begin{pmatrix}1&i\\ 0&{1}\end{pmatrix}\begin{pmatrix}1&0\\ 0&-e_{0}\end{pmatrix}e^{-J_{1}^{\infty 1}\sigma_{3}}
=\displaystyle= ϵ~2(eJ1,3+J34J11ie0eJ1,3+J34+J11ic01eJ1,3J34J11(c01e0eJ1,3+eJ1,3)eJ34+J11)\displaystyle\tilde{\epsilon}_{2}\begin{pmatrix}e^{J_{1,3}+J_{3}^{\infty 4}-J_{1}^{\infty 1}}&-ie_{0}e^{J_{1,3}+J_{3}^{\infty 4}+J_{1}^{\infty 1}}\\ ic_{0}^{-1}e^{J_{1,3}-J_{3}^{\infty 4}-J_{1}^{\infty 1}}&(c_{0}^{-1}e_{0}e^{J_{1,3}}+e^{-J_{1,3}})e^{-J_{3}^{\infty 4}+J_{1}^{\infty 1}}\end{pmatrix}

with ϵ~22=1,\tilde{\epsilon}_{2}^{2}=1, J11σ3=J31σ3|𝐜13𝐜11,λ3λ1J_{1}^{\infty 1}\sigma_{3}=J_{3}^{\infty 1}\sigma_{3}|_{\mathbf{c}^{3}_{\infty_{1}}\mapsto\mathbf{c}^{1}_{\infty_{1}},\,\,\lambda_{3}\mapsto\lambda_{1}} and J1,3σ3=J3,1σ3.J_{1,3}\sigma_{3}=-J_{3,1}\sigma_{3}. Thus we have the following.

Proposition 5.2.

Suppose that 0<ϕ<π/4.0<\phi<\pi/4. Then

1+s2s3=c0d01e2J5,3(1+O(tδ)),\displaystyle 1+s_{2}s_{3}=-c_{0}d_{0}^{-1}e^{-2J^{-}_{5,3}}(1+O(t^{-\delta})),
(1+s1s2)(1+s2s3)1=c01e0e2J1,3(1+O(tδ)),\displaystyle(1+s_{1}s_{2})(1+s_{2}s_{3})-1=c_{0}^{-1}e_{0}e^{2J_{1,3}}(1+O(t^{-\delta})),

where J5,3J^{-}_{5,3} is given by (5.5) and J1,3=J3,1J_{1,3}=-J_{3,1}.

6. Asymptotics of monodromy data

Recall that λμ(,λ)=w(Aϕ,λ2)\lambda\mu(\infty,\lambda)=w(A_{\phi},\lambda^{2}) is considered on the Riemann surface \mathcal{R}_{\infty} and that w(Aϕ,z)w(A_{\phi},z) defines the elliptic curve ΠAϕ,ϕ\Pi_{A_{\phi},\phi} with the primitive cycles 𝐚\mathbf{a} and 𝐛\mathbf{b} described as in Figure 2.1. Note that, by z=λ2z=\lambda^{2}, \mathcal{R}_{\infty} is mapped to ΠAϕ,ϕ\Pi_{A_{\phi},\phi}. Let 𝐚^,\hat{\mathbf{a}}, 𝐛^\hat{\mathbf{b}}\subset\mathcal{R}_{\infty} denote the inverse images of 𝐚,\mathbf{a}, 𝐛\mathbf{b}, respectively, such that 𝐚^+\hat{\mathbf{a}}\subset\mathcal{R}^{+}_{\infty} surrounds the cut [λ5,λ3][\lambda_{5},\lambda_{3}] and that 𝐛^\hat{\mathbf{b}} links with 𝐚^\hat{\mathbf{a}}. For λj(t)t\lambda_{j}(t)\in\mathcal{R}_{t} and λj=λj()\lambda_{j}=\lambda_{j}(\infty)\in\mathcal{R}_{\infty}, we have λj(t)=λj()+O(t1)\lambda_{j}(t)=\lambda_{j}(\infty)+O(t^{-1}) if 0<|ϕ|<π/4,0<|\phi|<\pi/4, and hence the cycles 𝐚^\hat{\mathbf{a}} and 𝐛^\hat{\mathbf{b}} may be regarded to be those on t\mathcal{R}_{t} as well for sufficiently large tt. Furthermore 𝐚,\mathbf{a}, 𝐛ΠAϕ,ϕ\mathbf{b}\subset\Pi_{A_{\phi},\phi} may be regarded to be the primitive cycles on Πaϕ(t),ϕ\Pi_{a_{\phi}(t),\phi} by (5.1).

6.1. Integrals

We would like to calculate the asymptotics of 𝐚^,𝐛^Λ3(λ)λ\int_{\hat{\mathbf{a}},\,\,\hat{\mathbf{b}}}\Lambda_{3}(\lambda)\lambda. By (4.1)

μ(t,λ)\displaystyle\mu(t,\lambda) =λ1(λ8+4eiϕλ6+4e2iϕλ4+4e3iϕaϕλ2+2(αβ)t1λ4+α2t2)1/2\displaystyle=\lambda^{-1}\left(\lambda^{8}+4e^{i\phi}\lambda^{6}+4e^{2i\phi}\lambda^{4}+4e^{3i\phi}a_{\phi}\lambda^{2}+2(\alpha-\beta)t^{-1}\lambda^{4}+\alpha^{2}t^{-2}\right)^{1/2}
=λ1𝐰(λ)(1+2(αβ)t1λ4𝐰(λ)2+O(t2𝐰(λ)2))1/2\displaystyle=\lambda^{-1}{\mathbf{w}(\lambda)}\Bigl{(}1+2(\alpha-\beta)t^{-1}\frac{\lambda^{4}}{\mathbf{w}(\lambda)^{2}}+O(t^{-2}\mathbf{w}(\lambda)^{-2})\Bigr{)}^{1/2}
=𝐰(λ)λ+(αβ)t1λ3𝐰(λ)+O(t2)\displaystyle=\frac{\mathbf{w}(\lambda)}{\lambda}+(\alpha-\beta)t^{-1}\frac{\lambda^{3}}{\mathbf{w}(\lambda)}+O(t^{-2})

along 𝐚^\hat{\mathbf{a}} and 𝐛^\hat{\mathbf{b}}, where

𝐰(λ)=w(aϕ,λ2),w(aϕ,z)=z4+4eiϕz3+4e2iϕz2+4e3iϕaϕz.\mathbf{w}(\lambda)=w(a_{\phi},\lambda^{2}),\quad w(a_{\phi},z)=\sqrt{z^{4}+4e^{i\phi}z^{3}+4e^{2i\phi}z^{2}+4e^{3i\phi}a_{\phi}z}.

Substitution λ2=z\lambda^{2}=z yields

(6.1) 𝐚^,𝐛^μ(t,λ)𝑑λ=\displaystyle\int_{\hat{\mathbf{a}},\,\,\hat{\mathbf{b}}}\mu(t,\lambda)d\lambda= 12𝐚,𝐛w(aϕ,z)z𝑑z+12(αβ)t1𝐚,𝐛zw(aϕ,z)𝑑z+O(t2)\displaystyle\frac{1}{2}\int_{\mathbf{a},\,\,\mathbf{b}}\frac{w(a_{\phi},z)}{z}dz+\frac{1}{2}(\alpha-\beta)t^{-1}\int_{\mathbf{a},\,\,\mathbf{b}}\frac{z}{w(a_{\phi},z)}dz+O(t^{-2})
=\displaystyle= e4iϕaϕ𝐚,𝐛dzzw(aϕ,z)+12(αβ)t1𝐚,𝐛zw(aϕ,z)𝑑z\displaystyle e^{4i\phi}a_{\phi}\int_{{\mathbf{a}},\,\,{\mathbf{b}}}\frac{dz}{zw(a_{\phi},z)}+\frac{1}{2}(\alpha-\beta)t^{-1}\int_{\mathbf{a},\,\,\mathbf{b}}\frac{z}{w(a_{\phi},z)}dz
+3e3iϕaϕ2ω𝐚,𝐛+O(t2),ω𝐚,𝐛=𝐚,𝐛dzw(aϕ,z),\displaystyle+\frac{3e^{3i\phi}a_{\phi}}{2}\omega_{\mathbf{a},\,\,\mathbf{b}}+O(t^{-2}),\quad\,\,\omega_{\mathbf{a},\,\,\mathbf{b}}=\int_{\mathbf{a},\,\,\mathbf{b}}\frac{dz}{w(a_{\phi},z)},

in which the second equality is obtained by using

w(aϕ,z)dzz=w(aϕ,z)2+eiϕw(aϕ,z)z+3e3iϕaϕdzw(aϕ,z)+2e4iϕaϕdzzw(aϕ,z).\int w(a_{\phi},z)\frac{dz}{z}=\frac{w(a_{\phi},z)}{2}+e^{i\phi}\frac{w(a_{\phi},z)}{z}+3e^{3i\phi}a_{\phi}\int\frac{dz}{w(a_{\phi},z)}+2e^{4i\phi}a_{\phi}\int\frac{dz}{zw(a_{\phi},z)}.

Let us calculate

diagT1Tλ|σ3dλ=14(1b3μ)ddλlnb1+ib2b1ib2dλ.\mathrm{diag}\,T^{-1}T_{\lambda}|_{\sigma_{3}}d\lambda=\frac{1}{4}\Bigl{(}1-\frac{b_{3}}{\mu}\Bigl{)}\frac{d}{d\lambda}\ln\frac{b_{1}+ib_{2}}{b_{1}-ib_{2}}d\lambda.

Recalling (3.1) and (3.2), and setting λ2=z\lambda^{2}=z, we have

(6.2) b1ib2=2ieiϕ/2(zz+),b1+ib2=2ieiϕ/2(zz),\displaystyle b_{1}-ib_{2}=2ie^{i\phi/2}(z-z_{+}),\quad b_{1}+ib_{2}=2ie^{-i\phi/2}(z-z_{-}),
z+:=eiϕ2ψtψψ2eiϕ+O(t1),z:=eiϕ2ψtψψ2eiϕ+O(t1),\displaystyle z_{+}:=-\frac{e^{-i\phi}}{2}\frac{\psi_{t}}{\psi}-\frac{\psi}{2}-e^{i\phi}+O(t^{-1}),\quad z_{-}:=\frac{e^{-i\phi}}{2}\frac{\psi_{t}}{\psi}-\frac{\psi}{2}-e^{i\phi}+O(t^{-1}),

with (b1ib2)(z+)=0,(b_{1}-ib_{2})(z_{+})=0, (b1+ib2)(z)=0.(b_{1}+ib_{2})(z_{-})=0. Furthermore

b3μ(t,λ)=z(z+2(eiϕ+ψ))(1w(aϕ,z)+O(t1)),\frac{b_{3}}{\mu(t,\lambda)}=z(z+2(e^{i\phi}+\psi))\Bigl{(}\frac{1}{w(a_{\phi},z)}+O(t^{-1})\Bigr{)},

which satisfies (b3/μ)(z±)=1(b_{3}/\mu)(z_{\pm})=1 on the upper sheet Π+\Pi_{+} of Πaϕ,ϕ\Pi_{a_{\phi},\phi}. Then it follows that

(6.3) diagT1Tλ|σ3dλ\displaystyle\mathrm{diag}\,T^{-1}T_{\lambda}|_{\sigma_{3}}d\lambda =14(1zz1zz+)dz\displaystyle=\frac{1}{4}\Bigl{(}\frac{1}{z-z_{-}}-\frac{1}{z-z_{+}}\Bigr{)}dz
+14(z+z+w(aϕ,z+)zz+w(aϕ,z)zz)dzw(aϕ,z)+O(t1)dz,\displaystyle+\frac{1}{4}\Bigl{(}z_{+}-z_{-}+\frac{w(a_{\phi},z_{+})}{z-z_{+}}-\frac{w(a_{\phi},z_{-})}{z-z_{-}}\Bigr{)}\frac{dz}{w(a_{\phi},z)}+O(t^{-1})dz,

and that

(6.4) 12z5z3(1zz+1zz)𝑑z=12ln(λ32z+)(λ52z)(λ32z)(λ52z+)=ln(c0d01),12z3z1(1zz+1zz)𝑑z=12ln(λ12z+)(λ32z)(λ12z)(λ32z+)=ln(c01e0).\begin{split}&\frac{1}{2}\int^{z_{3}}_{z_{5}}\Bigl{(}\frac{1}{z-z_{+}}-\frac{1}{z-z_{-}}\Bigr{)}dz=\frac{1}{2}\ln\frac{(\lambda_{3}^{2}-z_{+})(\lambda_{5}^{2}-z_{-})}{(\lambda_{3}^{2}-z_{-})(\lambda_{5}^{2}-z_{+})}=\ln(c_{0}d_{0}^{-1}),\\ &\frac{1}{2}\int^{z_{1}}_{z_{3}}\Bigl{(}\frac{1}{z-z_{+}}-\frac{1}{z-z_{-}}\Bigr{)}dz=\frac{1}{2}\ln\frac{(\lambda_{1}^{2}-z_{+})(\lambda_{3}^{2}-z_{-})}{(\lambda_{1}^{2}-z_{-})(\lambda_{3}^{2}-z_{+})}=\ln(c_{0}^{-1}e_{0}).\end{split}

Suppose that π/4<ϕ<0-\pi/4<\phi<0, and recall Proposition 5.1. Note that

J3,5=λ3λ5(tμ(t,λ)diagT1Tλ|σ3)𝑑λ,J_{3,5}=\int_{\lambda_{3}}^{\lambda_{5}}\Bigl{(}t\mu(t,\lambda)-\mathrm{diag}T^{-1}T_{\lambda}|_{\sigma_{3}}\Bigr{)}d\lambda,

along 𝐜35=(λ3,λ5)\mathbf{c}_{3}^{5}=(\lambda_{3},\lambda_{5})^{\sim} in (5.2), and that 𝐜35\mathbf{c}_{3}^{5} is 12(𝐚^)\frac{1}{2}(-\hat{\mathbf{a}}), which is the image of 12(𝐚)\frac{1}{2}(-\mathbf{a}) of Figure 2.1 under the map λ2=z.\lambda^{2}=z. The integral J3,1J_{3,1} of (5.4) is along 𝐜31\mathbf{c}_{3}^{1}, which is the image of 12𝐛\tfrac{1}{2}\mathbf{b}. Then by (6.1), (6.3) and (6.4), we have the following proposition, in which

(6.5) W(z)=(z+z+w(aϕ,z+)zz+w(aϕ,z)zz)1w(aϕ,z),\displaystyle W(z)=\Bigl{(}z_{+}-z_{-}+\frac{w(a_{\phi},z_{+})}{z-z_{+}}-\frac{w(a_{\phi},z_{-})}{z-z_{-}}\Bigr{)}\frac{1}{w(a_{\phi},z)},
(6.6) ω𝐚,𝐛=𝐚,𝐛dzw(aϕ,z),Cα,β(𝐚,𝐛)=12(αβ)𝐚,𝐛zdzw(aϕ,z).\displaystyle\omega_{\mathbf{a},\,\,\mathbf{b}}=\int_{\mathbf{a},\,\,\mathbf{b}}\frac{dz}{w(a_{\phi},z)},\qquad C_{\alpha,\beta}(\mathbf{a},\,\,\mathbf{b})=\frac{1}{2}(\alpha-\beta)\int_{\mathbf{a},\,\,\mathbf{b}}\frac{zdz}{w(a_{\phi},z)}.
Proposition 6.1.

Suppose that π/4<ϕ<0-\pi/4<\phi<0. Then

ln(1+s1s2)=\displaystyle\ln(1+s_{1}s_{2})= t2𝐚w(aϕ,z)z𝑑z+14𝐚W(z)𝑑zCα,β(𝐚)+πi+O(tδ),\displaystyle-\frac{t}{2}\int_{\mathbf{a}}\frac{w(a_{\phi},z)}{z}dz+\frac{1}{4}\int_{\mathbf{a}}W(z)dz-C_{\alpha,\beta}({\mathbf{a}})+\pi i+O(t^{-\delta}),
ln((1+s1s2)(1\displaystyle\ln((1+s_{1}s_{2})(1 +s2s3)1)\displaystyle+s_{2}s_{3})-1)
=\displaystyle= t2𝐛w(aϕ,z)z𝑑z+14𝐛W(z)𝑑zCα,β(𝐛)+O(tδ).\displaystyle-\frac{t}{2}\int_{\mathbf{b}}\frac{w(a_{\phi},z)}{z}dz+\frac{1}{4}\int_{\mathbf{b}}W(z)dz-C_{\alpha,\beta}({\mathbf{b}})+O(t^{-\delta}).

From Proposition 5.2 we have the following.

Proposition 6.2.

Suppose that 0<ϕ<π/4.0<\phi<\pi/4. Then

ln(1+s2s3)=\displaystyle\ln(1+s_{2}s_{3})= t2𝐚w(aϕ,z)z𝑑z14𝐚W(z)𝑑z+Cα,β(𝐚)+πi+O(tδ),\displaystyle\frac{t}{2}\int_{\mathbf{a}}\frac{w(a_{\phi},z)}{z}dz-\frac{1}{4}\int_{\mathbf{a}}W(z)dz+C_{\alpha,\beta}(\mathbf{a})+\pi i+O(t^{-\delta}),
ln((1+s1s2)(\displaystyle\ln((1+s_{1}s_{2})( 1+s2s3)1)\displaystyle 1+s_{2}s_{3})-1)
=\displaystyle= t2𝐛w(aϕ,z)z𝑑z+14𝐛W(z)𝑑zCα,β(𝐛)+O(tδ).\displaystyle-\frac{t}{2}\int_{\mathbf{b}}\frac{w(a_{\phi},z)}{z}dz+\frac{1}{4}\int_{\mathbf{b}}W(z)dz-C_{\alpha,\beta}(\mathbf{b})+O(t^{-\delta}).
Remark 6.1.

In the propositions above,

12𝐚,𝐛w(aϕ,z)z𝑑z=e4iϕaϕ𝐚,𝐛dzzw(aϕ,z)+32e3iϕaϕω𝐚,𝐛.\frac{1}{2}\int_{\mathbf{a},\,\,\mathbf{b}}\frac{w(a_{\phi},z)}{z}dz=e^{4i\phi}a_{\phi}\int_{\mathbf{a},\,\,\mathbf{b}}\frac{dz}{zw(a_{\phi},z)}+\frac{3}{2}e^{3i\phi}a_{\phi}\omega_{\mathbf{a},\,\,\mathbf{b}}.

6.2. Theta-function

Further calculation needs the theta-function

ϑ(z,τ)=n=eπiτn2+2πizn,τ=ω𝐛ω𝐚,Imτ>0,ν=1+τ2.\vartheta(z,\tau)=\sum_{n=-\infty}^{\infty}e^{\pi i\tau n^{2}+2\pi izn},\quad\tau=\frac{\omega_{\mathbf{b}}}{\omega_{\mathbf{a}}},\quad\mathrm{Im\,}\tau>0,\quad\nu=\frac{1+\tau}{2}.

For z,z, z~Πaϕ,ϕ=Π+Π\tilde{z}\in\Pi_{a_{\phi},\phi}=\Pi_{+}\cup\Pi_{-}, set

F(z~,z)=1ω𝐚z~zdzw(aϕ,z).F(\tilde{z},z)=\frac{1}{\omega_{\mathbf{a}}}\int^{z}_{\tilde{z}}\frac{dz}{w(a_{\phi},z)}.

For z0Πaϕ,ϕ=Π+Πz_{0}\in\Pi_{a_{\phi},\phi}=\Pi_{+}\cup\Pi_{-} write z0+=(z0,w(aϕ,z0+)),z_{0}^{+}=(z_{0},w(a_{\phi},z_{0}^{+})), z0=(z0,w(aϕ,z0+)z_{0}^{-}=(z_{0},-w(a_{\phi},z_{0}^{+}).

Proposition 6.3.

For any z0Πaϕ,ϕ,z_{0}\in\Pi_{a_{\phi},\phi}, w(z)=w(aϕ,z)w(z)=w(a_{\phi},z) fulfils

dz(zz0)w(z)\displaystyle\frac{dz}{(z-z_{0})w(z)} =1w(z0+)dlnϑ(F(z0+,z)+ν,τ)ϑ(F(z0,z)+ν,τ)g0(z0)dzw(z),\displaystyle=\frac{1}{w(z_{0}^{+})}d\ln\frac{\vartheta(F(z_{0}^{+},z)+\nu,\tau)}{\vartheta(F(z_{0}^{-},z)+\nu,\tau)}-g_{0}(z_{0})\frac{dz}{w(z)},
g0(z0)\displaystyle g_{0}(z_{0}) =w(z0+)2w(z0+)1ω𝐚w(z0+)(πi+ϑϑ(F(z0,z0+)+ν,τ)).\displaystyle=\frac{w^{\prime}(z^{+}_{0})}{2w(z^{+}_{0})}-\frac{1}{\omega_{\mathbf{a}}w(z_{0}^{+})}\Bigl{(}\pi i+\frac{\vartheta^{\prime}}{\vartheta}(F(z^{-}_{0},z^{+}_{0})+\nu,\tau)\Bigr{)}.
Proof..

Let z=φ(u)z=\varphi(u) be an elliptic function such that φu2=w(φ)2\varphi_{u}^{2}=w(\varphi)^{2} and z0=φ(u0)z_{0}=\varphi(u_{0}). Then (zz0)1w(z)1dz=(φ(u)φ(u0))1du.(z-z_{0})^{-1}w(z)^{-1}dz=(\varphi(u)-\varphi(u_{0}))^{-1}du. Write φ(u0±)=z0±,\varphi(u_{0}^{\pm})=z_{0}^{\pm}, φu(u0±)=±w(z0+).\varphi_{u}(u^{\pm}_{0})=\pm w(z_{0}^{+}). Around each u=u0±u=u_{0}^{\pm},

(φ(u)φ(u0±))1=±(uu0±)1w(z0+)w(z0+)2w(z0+)+O(uu0±),(\varphi(u)-\varphi(u_{0}^{\pm}))^{-1}=\pm\frac{(u-u_{0}^{\pm})^{-1}}{w(z_{0}^{+})}-\frac{w^{\prime}(z_{0}^{+})}{2w(z_{0}^{+})}+O(u-u_{0}^{\pm}),

and hence we have

(φ(u)φ(u0±))1=\displaystyle(\varphi(u)-\varphi(u_{0}^{\pm}))^{-1}= 1ω𝐚w(z0)(ϑϑ(ω𝐚1(uu0+)+ν,τ)ϑϑ(ω𝐚1(uu0)+ν,τ))+C0\displaystyle\frac{1}{\omega_{\mathbf{a}}w(z_{0})}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}(\omega_{\mathbf{a}}^{-1}(u-u_{0}^{+})+\nu,\tau)-\frac{\vartheta^{\prime}}{\vartheta}(\omega_{\mathbf{a}}^{-1}(u-u_{0}^{-})+\nu,\tau)\Bigr{)}+C_{0}
=\displaystyle= 1ω𝐚w(z0)(ϑϑ(F(z0+,z)+ν,τ)ϑϑ(F(z0,z))+ν,τ))+C0\displaystyle\frac{1}{\omega_{\mathbf{a}}w(z_{0})}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}(F(z_{0}^{+},z)+\nu,\tau)-\frac{\vartheta^{\prime}}{\vartheta}(F(z_{0}^{-},z))+\nu,\tau)\Bigr{)}+C_{0}

for some constant C0C_{0}. Passage to the limit zz0+z\to z^{+}_{0}, i.e. uu0+u\to u_{0}^{+} leads to

C0=1ω𝐚w(z0+)(ϑϑ(F(z0,z0+)+ν,τ)ϑ′′(ν)2ϑ(ν))w(z0+)2w(z0+),C_{0}=\frac{1}{\omega_{\mathbf{a}}w(z_{0}^{+})}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}(F(z_{0}^{-},z_{0}^{+})+\nu,\tau)-\frac{\vartheta^{\prime\prime}(\nu)}{2\vartheta^{\prime}(\nu)}\Bigr{)}-\frac{w^{\prime}(z^{+}_{0})}{2w(z^{+}_{0})},

which implies the required formula. ∎

Corollary 6.4.

For any z0Πaϕ,ϕz_{0}\in\Pi_{a_{\phi},\phi}, and for w(z)=w(aϕ,z)w(z)=w(a_{\phi},z),

(6.7) 𝐚dz(zz0)w(z)=g0(z0)ω𝐚,\displaystyle\int_{\mathbf{a}}\frac{dz}{(z-z_{0})w(z)}=-g_{0}(z_{0})\omega_{\mathbf{a}},
(6.8) (𝐛τ𝐚)dz(zz0)w(z)=2πiw(z0+)F(z0,z0+),\displaystyle\Bigl{(}\int_{\mathbf{b}}-\tau\int_{\mathbf{a}}\Bigr{)}\frac{dz}{(z-z_{0})w(z)}=\frac{2\pi i}{w(z_{0}^{+})}F(z_{0}^{-},z_{0}^{+}),
(6.9) (𝐛τ𝐚)dzzw(z)=2πiaϕω𝐚e3iϕ.\displaystyle\Bigl{(}\int_{\mathbf{b}}-\tau\int_{\mathbf{a}}\Bigr{)}\frac{dz}{zw(z)}=\frac{2\pi i}{a_{\phi}\omega_{\mathbf{a}}}e^{-3i\phi}.
Corollary 6.5.

For z+,z_{+}, zz_{-} and W(z)W(z) given by (6.2) and (6.5), and for w(z)=w(aϕ,z)w(z)=w(a_{\phi},z), we have, on Πaϕ,ϕ\Pi_{a_{\phi},\phi},

𝐚W(z)𝑑z=\displaystyle\int_{\mathbf{a}}W(z)dz= (z+w(z++)2)ω𝐚+ϑϑ(F(z+,z++)+ν,τ)\displaystyle\Bigl{(}z_{+}-\frac{w^{\prime}(z_{+}^{+})}{2}\Bigr{)}\omega_{\mathbf{a}}+\frac{\vartheta^{\prime}}{\vartheta}(F(z_{+}^{-},z_{+}^{+})+\nu,\tau)
(zw(z+)2)ω𝐚ϑϑ(F(z,z+)+ν,τ),\displaystyle\phantom{--------}-\Bigl{(}z_{-}-\frac{w^{\prime}(z_{-}^{+})}{2}\Bigr{)}\omega_{\mathbf{a}}-\frac{\vartheta^{\prime}}{\vartheta}(F(z_{-}^{-},z_{-}^{+})+\nu,\tau),
(𝐛τ𝐚)\displaystyle\Bigl{(}\int_{\mathbf{b}}-\tau\int_{\mathbf{a}}\Bigr{)} W(z)dz=2πi(F(z+,z++)F(z,z+)).\displaystyle W(z)dz=2\pi i(F(z_{+}^{-},z_{+}^{+})-F(z^{-}_{-},z_{-}^{+})).

Another expression of 𝐚W(λ)𝑑λ\int_{\mathbf{a}}W(\lambda)d\lambda is derived by using more information on the poles of P(u;aϕ)\mathrm{P}(u;a_{\phi}) (cf. Proposition 7.6).

Proposition 6.6.

Under the same condition as above

𝐚W(z)𝑑z=2(ϑϑ(12F(z+,z++)+12+τ6,τ)ϑϑ(12F(z,z+)+12+τ6,τ)).\int_{\mathbf{a}}W(z)dz=2\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}(\tfrac{1}{2}F(z_{+}^{-},z_{+}^{+})+\tfrac{1}{2}+\tfrac{\tau}{6},\tau)-\frac{\vartheta^{\prime}}{\vartheta}(\tfrac{1}{2}F(z_{-}^{-},z_{-}^{+})+\tfrac{1}{2}+\tfrac{\tau}{6},\tau)\Bigr{)}.
Proof..

Let z0Πaϕ,ϕ=Π+Π.z_{0}\in\Pi_{a_{\phi},\phi}=\Pi_{+}\cup\Pi_{-}. Note that z012w(aϕ,z0)z_{0}-\tfrac{1}{2}w^{\prime}(a_{\phi},z_{0}) is holomorphic around z0+=+Π+z_{0}^{+}=\infty^{+}\in\Pi_{+} and admits a pole at z0+=Πz_{0}^{+}=\infty^{-}\in\Pi_{-} with the residue 22. Then, we have

(z012w(aϕ,z0+))ω𝐚+(ϑ/ϑ)(F(z0,z0+)+ν,τ)=2(ϑ/ϑ)(12F(z0,z0+)+12+τ6,τ)+c0(z_{0}-\tfrac{1}{2}w^{\prime}(a_{\phi},z^{+}_{0}))\omega_{\mathbf{a}}+(\vartheta^{\prime}/\vartheta)(F(z_{0}^{-},z_{0}^{+})+\nu,\tau)=2(\vartheta^{\prime}/\vartheta)(\tfrac{1}{2}F(z_{0}^{-},z_{0}^{+})+\tfrac{1}{2}+\tfrac{\tau}{6},\tau)+c_{0}

for some constant c0.c_{0}. Indeed, as shown later by Proposition 7.6, the elliptic function φ0(u)\varphi_{0}(u) defined by u=0φ0w(aϕ,z)1𝑑zu=\int_{0}^{\varphi_{0}}w(a_{\phi},z)^{-1}dz has poles with the residue 1\mp 1 at ±13ω𝐛+ω𝐚+ω𝐛\pm\tfrac{1}{3}\omega_{\mathbf{b}}+\omega_{\mathbf{a}}\mathbb{Z}+\omega_{\mathbf{b}}\mathbb{Z}, which implies ω𝐚10+w(aϕ,z)1𝑑z+12+τ6=12τ6\omega_{\mathbf{a}}^{-1}\int^{\infty^{+}}_{0}w(a_{\phi},z)^{-1}dz+\tfrac{1}{2}+\tfrac{\tau}{6}=\tfrac{1}{2}-\tfrac{\tau}{6} for z0+=+Π+,z_{0}^{+}=\infty^{+}\in\Pi_{+}, and ω𝐚10w(aϕ,z)1𝑑z+12+τ6=12+τ2=ν\omega_{\mathbf{a}}^{-1}\int^{\infty^{-}}_{0}w(a_{\phi},z)^{-1}dz+\tfrac{1}{2}+\tfrac{\tau}{6}=\tfrac{1}{2}+\tfrac{\tau}{2}=\nu for z0=Πz_{0}^{-}=\infty^{-}\in\Pi_{-}; and furthermore, at the branch points z0=0,z1,z3,z5z_{0}=0,z_{1},z_{3},z_{5}, the leading terms of 12w(aϕ,z0+)ω𝐚-\tfrac{1}{2}w^{\prime}(a_{\phi},z^{+}_{0})\omega_{\mathbf{a}} and (ϑ/ϑ)(F(z0,z0+)+ν,τ)(\vartheta^{\prime}/\vartheta)(F(z_{0}^{-},z_{0}^{+})+\nu,\tau) are cancelled out (cf. Subsection 6.3). Putting z0=z±z_{0}=z_{\pm} and using Corollary 6.5, we obtain the proposition. ∎

6.3. Expression of Bϕ(t)B_{\phi}(t)

Recall that our calculations are carried out under the supposition (5.1) in the strip Sϕ(t,κ1,δ1).S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}). Let

Ω𝐚,𝐛:=𝐚,𝐛dzw(Aϕ,z),𝒥𝐚,𝐛:=𝐚,𝐛w(Aϕ,z)z𝑑z\Omega_{\mathbf{a},\,\,\mathbf{b}}:=\int_{\mathbf{a},\,\,\mathbf{b}}\frac{dz}{w(A_{\phi},z)},\qquad\mathcal{J}_{\mathbf{a},\,\,\mathbf{b}}:=\int_{\mathbf{a},\,\,\mathbf{b}}\frac{w(A_{\phi},z)}{z}dz

for 𝐚,\mathbf{a}, 𝐛\mathbf{b} on ΠAϕ,ϕ=Π+Π=limaϕAϕΠaϕ,ϕ,\Pi_{A_{\phi},\phi}=\Pi_{+}\cup\Pi_{-}=\lim_{a_{\phi}\to A_{\phi}}\Pi_{a_{\phi},\phi}, i.e. ΠAϕ,ϕ=limtΠaϕ(t),ϕ\Pi_{A_{\phi},\phi}=\lim_{t\to\infty}\Pi_{a_{\phi}(t),\phi}. We would like to express Bϕ(t)B_{\phi}(t) defined by (5.1) in terms of these quantities.

By Corollary 6.5 the integral 𝐚W(z)𝑑z\int_{\mathbf{a}}W(z)dz is a linear combination of z±z_{\pm}, w(z±+)w^{\prime}(z_{\pm}^{+}) and (ϑ/ϑ)(F(z±,z±+)+ν,τ)(\vartheta^{\prime}/\vartheta)(F(z_{\pm}^{-},z_{\pm}^{+})+\nu,\tau). In Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}), the functions ψ\psi, 1/ψ1/\psi and ψ\psi^{\prime} has no poles, and hence z±=z±(t)z_{\pm}=z_{\pm}(t) are bounded in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}) and so are w(z±+)w^{\prime}(z^{+}_{\pm}) except for neighbourhoods of the zeros of w(z)w(z), i.e., 0,z1,z3,z50,z_{1},z_{3},z_{5}. Furthermore, around these points, the leading terms of 12w(z±+)ω𝐚-\tfrac{1}{2}w^{\prime}(z^{+}_{\pm})\omega_{\mathbf{a}} and (ϑ/ϑ)(F(z±,z±+)+ν,τ)(\vartheta^{\prime}/\vartheta)(F(z_{\pm}^{-},z_{\pm}^{+})+\nu,\tau) are cancelled out. Indeed we have, say, around z±+=0z^{+}_{\pm}=0, 12w(z±+)ω𝐚=12e3iϕ/2Aϕ1/2ω𝐚(z±+)1/2+O(1)-\tfrac{1}{2}w^{\prime}(z_{\pm}^{+})\omega_{\mathbf{a}}=-\tfrac{1}{2}e^{3i\phi/2}A_{\phi}^{1/2}\omega_{\mathbf{a}}(z_{\pm}^{+})^{-1/2}+O(1) and F(z±,z±+)1=12e3iϕ/2Aϕ1/2ω𝐚(z±+)1/2+O(1)F(z_{\pm}^{-},z_{\pm}^{+})^{-1}=\tfrac{1}{2}e^{3i\phi/2}A_{\phi}^{1/2}\omega_{\mathbf{a}}(z_{\pm}^{+})^{-1/2}+O(1). Since z±=z±(t)z_{\pm}=z_{\pm}(t) moves on Πaϕ,ϕ\Pi_{a_{\phi},\phi} crossing 𝐚\mathbf{a}- and 𝐛\mathbf{b}-cycles, ω𝐚F(z±,z±+)=2p±(t)ω𝐚+2q±(t)ω𝐛+O(1)\omega_{\mathbf{a}}F(z_{\pm}^{-},z_{\pm}^{+})=2p_{\pm}(t)\omega_{\mathbf{a}}+2q_{\pm}(t)\omega_{\mathbf{b}}+O(1) with p±(t),p_{\pm}(t), q±(t),q_{\pm}(t)\in\mathbb{Z}, which implies the boundedness of Re(ϑ/ϑ)(F(z±,z±+)+ν,τ)\mathrm{Re\,}(\vartheta^{\prime}/\vartheta)(F(z^{-}_{\pm},z_{\pm}^{+})+\nu,\tau). Thus we have verified that Re𝐚W(z)𝑑z\mathrm{Re\,}\int_{\mathbf{a}}W(z)dz is bounded in Sϕ(t,κ1,δ1).S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}).

Suppose that π/4<ϕ<0-\pi/4<\phi<0. By (5.1)

1z(w(aϕ,z)w(Aϕ,z))=2e3iϕt1Bϕ(t)w(Aϕ,z)(1+O(t1Bϕ(t))).\frac{1}{z}(w(a_{\phi},z)-w(A_{\phi},z))=\frac{2e^{3i\phi}t^{-1}B_{\phi}(t)}{w(A_{\phi},z)}(1+O(t^{-1}B_{\phi}(t))).

By using this with Bϕ(t)1B_{\phi}(t)\ll 1 the first formula in Proposition 6.1 is written in the form

ln(1+s1s2)=t2𝐚(w(Aϕ,z)z+2e3iϕt1Bϕ(t)w(Aϕ,z))𝑑z+14𝐚W(z)𝑑zCα,β(𝐚)+πi+O(tδ),\ln(1+s_{1}s_{2})=-\frac{t}{2}\int_{\mathbf{a}}\Bigl{(}\frac{w(A_{\phi},z)}{z}+\frac{2e^{3i\phi}t^{-1}B_{\phi}(t)}{w(A_{\phi},z)}\Bigr{)}dz+\frac{1}{4}\int_{\mathbf{a}}W(z)dz-C_{\alpha,\beta}(\mathbf{a})+\pi i+O(t^{-\delta}),

that is,

t𝒥𝐚+2e3iϕΩ𝐚Bϕ(t)=12𝐚W(z)𝑑z2Cα,β(𝐚)+2πi2ln(1+s1s2)+O(tδ).t\mathcal{J}_{\mathbf{a}}+2e^{3i\phi}\Omega_{\mathbf{a}}B_{\phi}(t)=\frac{1}{2}\int_{\mathbf{a}}W(z)dz-2C_{\alpha,\beta}(\mathbf{a})+2\pi i-2\ln(1+s_{1}s_{2})+O(t^{-\delta}).

In Proposition 6.1, (ln(1+s1s2),ln((1+s1s2)(1+s2s3)1))(\ln(1+s_{1}s_{2}),\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)), generally depending on tt, is a solution of the direct monodromy problem. Suppose that

(6.10) ln(1+s1s2)1,ln((1+s1s2)(1+s2s3)1)1in Sϕ(t,κ1,δ1).\ln(1+s_{1}s_{2})\ll 1,\quad\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)\ll 1\,\,\,\text{in $S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1})$.}

By the Boutroux equations (2.2), we have Ree3iϕΩ𝐚Bϕ(t)1\mathrm{Re\,}e^{3i\phi}\Omega_{\mathbf{a}}B_{\phi}(t)\ll 1 as tt\to\infty in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}). From the second formula of Proposition 6.1 we similarly derive

t𝒥𝐛+2e3iϕΩ𝐛Bϕ(t)=12𝐛W(z)𝑑z2Cα,β(𝐛)2ln((1+s1s2)(1+s2s3)1)+O(tδ),t\mathcal{J}_{\mathbf{b}}+2e^{3i\phi}\Omega_{\mathbf{b}}B_{\phi}(t)=\frac{1}{2}\int_{\mathbf{b}}W(z)dz-2C_{\alpha,\beta}(\mathbf{b})-2\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)+O(t^{-\delta}),

in which 𝐛W(z)𝑑z\int_{\mathbf{b}}W(z)dz is expressed by ϑ(z,τ^)\vartheta(z,\hat{\tau}) with τ^=(ω𝐚)/ω𝐛,\hat{\tau}=(-\omega_{\mathbf{a}})/\omega_{\mathbf{b}}, and this formula yields Ree3iϕΩ𝐛Bϕ(t)1\mathrm{Re\,}e^{3i\phi}\Omega_{\mathbf{b}}B_{\phi}(t)\ll 1. These two estimates leads to the inequality |Bϕ(t)|C0|B_{\phi}(t)|\leq C_{0} in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}) for some C0>0,C_{0}>0, while the implied constant of (5.1) may be supposed to be 2C02C_{0} if tt^{\prime}_{\infty} is taken sufficiently large. Hence the boundedness of Bϕ(t)B_{\phi}(t) may be derived under (6.10) independently of (5.1). The case 0<ϕ<π/40<\phi<\pi/4 is discussed in the same way by using the Boutroux equations (2.2).

Remark 6.2.

Under a relaxed condition, say Bϕ(t)ln|t|B_{\phi}(t)\ll\ln|t| instead of 1\ll 1 of (5.1), each turning point is located within the distance O(t1ln|t|)O(t^{-1}\ln|t|) from the limit one, and Propositions 6.1 and 6.2 are obtained by choosing a slightly smaller δ\delta. Then the equivalence between (6.10) and the boundedness of Bϕ(t)B_{\phi}(t) may be proved as above.

Proposition 6.7.

Suppose that 0<|ϕ|<π/40<|\phi|<\pi/4 and let 𝔩(𝐬,ϕ)=ln(1+s1s2)\mathfrak{l}(\mathbf{s},\phi)=\ln(1+s_{1}s_{2}) if π/4<ϕ<0,-\pi/4<\phi<0, and =ln(1+s2s3)=-\ln(1+s_{2}s_{3}) if 0<ϕ<π/4.0<\phi<\pi/4. In Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}), we have

𝔩(𝐬,ϕ)1,ln((1+s1s2)(1+s2s3)1)1\mathfrak{l}(\mathbf{s},\phi)\ll 1,\quad\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)\ll 1

if and only if Bϕ(t)1B_{\phi}(t)\ll 1, and then

t𝒥𝐚+2e3iϕΩ𝐚Bϕ(t)=12𝐚W(z)𝑑z2Cα,β(𝐚)+2πi2𝔩(𝐬,ϕ)+O(tδ).t\mathcal{J}_{\mathbf{a}}+2e^{3i\phi}\Omega_{\mathbf{a}}B_{\phi}(t)=\frac{1}{2}\int_{\mathbf{a}}W(z)dz-2C_{\alpha,\beta}(\mathbf{a})+2\pi i-2\mathfrak{l}(\mathbf{s},\phi)+O(t^{-\delta}).

The following fact guarantees the possibility of the limit aϕAϕa_{\phi}\to A_{\phi} under integration.

Proposition 6.8.

Suppose that 0<|ϕ|<π/4.0<|\phi|<\pi/4. Then, in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}),

(z+z++zz+)dzw(aϕ,z)=(z+z++zz+)dzw(Aϕ,z)+O(t1).\biggl{(}\int^{z^{+}_{+}}_{z^{-}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\frac{dz}{w(a_{\phi},z)}=\biggl{(}\int^{z^{+}_{+}}_{z^{-}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\frac{dz}{w(A_{\phi},z)}+O(t^{-1}).
Proof..

By (5.1) it is easy to see ω𝐚,𝐛=Ω𝐚,𝐛+O(t1).\omega_{\mathbf{a},\,\,\mathbf{b}}=\Omega_{\mathbf{a},\,\,\mathbf{b}}+O(t^{-1}). Suppose that π/4<ϕ<0-\pi/4<\phi<0. By Proposition 6.1, Remark 6.1, Corollary 6.5 and (6.9),

ln((1+s1s2)\displaystyle\ln((1+s_{1}s_{2}) (1+s2s3)1)τln(1+s1s2)\displaystyle(1+s_{2}s_{3})-1)-\tau\ln(1+s_{1}s_{2})
=\displaystyle= aϕe4iϕt(𝐛τ𝐚)dzzw(aϕ,z)+14(𝐛τ𝐚)W(z)dz\displaystyle-a_{\phi}e^{4i\phi}t\Bigl{(}\int_{\mathbf{b}}-\tau\int_{\mathbf{a}}\Bigr{)}\frac{dz}{zw(a_{\phi},z)}+\frac{1}{4}\Bigl{(}\int_{\mathbf{b}}-\tau\int_{\mathbf{a}}\Bigr{)}W(z)dz
(Cα,β(𝐛)τCα,β(𝐚))πiτ+O(tδ)\displaystyle-(C_{\alpha,\beta}(\mathbf{b})-\tau C_{\alpha,\beta}(\mathbf{a}))-\pi i\tau+O(t^{-\delta})
=\displaystyle= 2πieiϕtω𝐚+πi2(F(z+,z++)F(z,z+))Cα,βπiτ+O(tδ)\displaystyle-\frac{2\pi ie^{i\phi}t}{\omega_{\mathbf{a}}}+\frac{\pi i}{2}(F(z^{-}_{+},z^{+}_{+})-F(z^{-}_{-},z^{+}_{-}))-C_{\alpha,\beta}^{*}-\pi i\tau+O(t^{-\delta})

with Cα,β=Cα,β(𝐛)τCα,β(𝐚).C_{\alpha,\beta}^{*}=C_{\alpha,\beta}(\mathbf{b})-\tau C_{\alpha,\beta}(\mathbf{a}). This implies

2eiϕt+12(z+z++zz+)dzw(aϕ,z)=O(1).-2e^{i\phi}t+\frac{1}{2}\biggl{(}\int_{z^{-}_{+}}^{z^{+}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\frac{dz}{w(a_{\phi},z)}=O(1).

Write ω𝐚F(z±,z±+)=2p±(t)ω𝐚+2q±(t)ω𝐛+O(1)\omega_{\mathbf{a}}F(z_{\pm}^{-},z_{\pm}^{+})=2p_{\pm}(t)\omega_{\mathbf{a}}+2q_{\pm}(t)\omega_{\mathbf{b}}+O(1) with p±(t),p_{\pm}(t), q±(t)q_{\pm}(t)\in\mathbb{Z}, and set i𝒥𝐚t/2+πq(t)=X,-i\mathcal{J}_{\mathbf{a}}t/2+\pi q(t)=X, i𝒥𝐛t/2πp(t)=Y,-i\mathcal{J}_{\mathbf{b}}t/2-\pi p(t)=Y, where p(t)=p+(t)p(t),p(t)=p_{+}(t)-p_{-}(t), q(t)=q+(t)q(t).q(t)=q_{+}(t)-q_{-}(t). By the Boutroux equations (2.2), ImX\mathrm{Im\,}X and ImY\mathrm{Im\,}Y are bounded. Note that ω𝐛𝒥𝐚ω𝐚𝒥𝐛=Ω𝐛𝒥𝐚Ω𝐚𝒥𝐛+O(t1)=4πieiϕ+O(t1)\omega_{\mathbf{b}}\mathcal{J}_{\mathbf{a}}-\omega_{\mathbf{a}}\mathcal{J}_{\mathbf{b}}=\Omega_{\mathbf{b}}\mathcal{J}_{\mathbf{a}}-\Omega_{\mathbf{a}}\mathcal{J}_{\mathbf{b}}+O(t^{-1})=-4\pi ie^{i\phi}+O(t^{-1}) by Corollary 8.10. Then the estimate above is

2eiϕt+ω𝐚p(t)+ω𝐛q(t)+O(1)\displaystyle-2e^{i\phi}t+\omega_{\mathbf{a}}p(t)+\omega_{\mathbf{b}}q(t)+O(1)
=\displaystyle= 2eiϕt+π1ω𝐚(Yi𝒥𝐛t/2)+π1ω𝐛(X+i𝒥𝐚t/2)+O(1)\displaystyle-2e^{i\phi}t+\pi^{-1}\omega_{\mathbf{a}}(-Y-i\mathcal{J}_{\mathbf{b}}t/2)+\pi^{-1}\omega_{\mathbf{b}}(X+i\mathcal{J}_{\mathbf{a}}t/2)+O(1)
=\displaystyle= 2eiϕt+π1it(ω𝐛𝒥𝐚ω𝐚𝒥𝐛)/2+π1(ω𝐛Xω𝐚Y)+O(1)\displaystyle-2e^{i\phi}t+\pi^{-1}it(\omega_{\mathbf{b}}\mathcal{J}_{\mathbf{a}}-\omega_{\mathbf{a}}\mathcal{J}_{\mathbf{b}})/2+\pi^{-1}(\omega_{\mathbf{b}}X-\omega_{\mathbf{a}}Y)+O(1)
=\displaystyle= π1(ω𝐛Xω𝐚Y)+O(1)1\displaystyle\pi^{-1}(\omega_{\mathbf{b}}X-\omega_{\mathbf{a}}Y)+O(1)\ll 1

with Im(ω𝐛/ω𝐚)>0\mathrm{Im\,}(\omega_{\mathbf{b}}/\omega_{\mathbf{a}})>0 uniformly. This implies XX, Y1Y\ll 1, and hence

πp(t)=i𝒥𝐛t/2+O(1),πq(t)=i𝒥𝐚t/2+O(1).\pi p(t)=-i\mathcal{J}_{\mathbf{b}}t/2+O(1),\quad\pi q(t)=i\mathcal{J}_{\mathbf{a}}t/2+O(1).

Observing that w(aϕ,z)1w(Aϕ,z)1=2ze3iϕBϕ(t)t1w(Aϕ,z)3+O(t2),w(a_{\phi},z)^{-1}-w(A_{\phi},z)^{-1}=-2ze^{3i\phi}B_{\phi}(t)t^{-1}w(A_{\phi},z)^{-3}+O(t^{-2}), we have

|(z+z++zz+)(1w(aϕ,z)1w(Aϕ,z))dz||(z+z++zz+)zBϕ(t)t1w(Aϕ,z)3dz|+O(t1)\displaystyle\biggl{|}\biggl{(}\int^{z^{+}_{+}}_{z^{-}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\Bigl{(}\frac{1}{w(a_{\phi},z)}-\frac{1}{w(A_{\phi},z)}\Bigr{)}dz\biggr{|}\ll\biggl{|}\biggl{(}\int^{z^{+}_{+}}_{z^{-}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\frac{zB_{\phi}(t)t^{-1}}{w(A_{\phi},z)^{3}}dz\biggr{|}+O(t^{-1})
|t1(z+z++zz+)zdzw(Aϕ,z)3|+O(t1)|t1(p(t)j𝐚+q(t)j𝐛)|+O(t1)\displaystyle\ll\biggl{|}t^{-1}\biggl{(}\int^{z^{+}_{+}}_{z^{-}_{+}}-\int^{z^{+}_{-}}_{z^{-}_{-}}\biggr{)}\frac{zdz}{w(A_{\phi},z)^{3}}\biggr{|}+O(t^{-1})\ll|t^{-1}(p(t)j_{\mathbf{a}}+q(t)j_{\mathbf{b}})|+O(t^{-1})
=|𝒥𝐛j𝐚𝒥𝐚j𝐛|+O(t1)=12|(/Aϕ)(𝒥𝐛Ω𝐚𝒥𝐚Ω𝐛)|+O(t1)t1\displaystyle=|\mathcal{J}_{\mathbf{b}}j_{\mathbf{a}}-\mathcal{J}_{\mathbf{a}}j_{\mathbf{b}}|+O(t^{-1})=\tfrac{1}{2}|(\partial/\partial A_{\phi})(\mathcal{J}_{\mathbf{b}}\Omega_{\mathbf{a}}-\mathcal{J}_{\mathbf{a}}\Omega_{\mathbf{b}})|+O(t^{-1})\ll t^{-1}

with j𝐚,𝐛=𝐚,𝐛zw(Aϕ,z)3𝑑zj_{\mathbf{a},\,\,\mathbf{b}}=\int_{\mathbf{a},\,\,\mathbf{b}}zw(A_{\phi},z)^{-3}dz. This completes the proof. ∎

7. Proofs of the main results

7.1. Derivation of ψ(t)\psi(t)

Suppose that π/4<ϕ<0.-\pi/4<\phi<0. Let 𝐬=(s1,s2,s3,s4)\mathbf{s}=(s_{1},s_{2},s_{3},s_{4}) with (1+s1s2)(1+s2s3)10,(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, 1+s1s201+s_{1}s_{2}\not=0 be a solution of the direct monodromy problem discussed above in which (ψ,ψt)(\psi,\psi_{t}) be such that (5.1) is valid in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}). As in the proof of Proposition 6.8 we set

ln((1+s1s2)\displaystyle\ln((1+s_{1}s_{2}) (1+s2s3)1)τln(1+s1s2)\displaystyle(1+s_{2}s_{3})-1)-\tau\ln(1+s_{1}s_{2})
=\displaystyle= 2πieiϕtω𝐚+πi2(F(z+,z++)F(z,z+))Cα,βπiτ+O(tδ)\displaystyle-\frac{2\pi ie^{i\phi}t}{\omega_{\mathbf{a}}}+\frac{\pi i}{2}(F(z^{-}_{+},z^{+}_{+})-F(z^{-}_{-},z^{+}_{-}))-C_{\alpha,\beta}^{*}-\pi i\tau+O(t^{-\delta})

with Cα,β=Cα,β(𝐛)τCα,β(𝐚).C_{\alpha,\beta}^{*}=C_{\alpha,\beta}(\mathbf{b})-\tau C_{\alpha,\beta}(\mathbf{a}). Then by Proposition 6.8 we have

(7.1) (0z++0z+)dzw(Aϕ,z)=2eiϕt+2χ~0+Ω𝐛+O(tδ),\biggl{(}\int^{z^{+}_{+}}_{0}-\int^{z^{+}_{-}}_{0}\biggr{)}\frac{dz}{w(A_{\phi},z)}=2e^{i\phi}t+2\tilde{\chi}_{0}+\Omega_{\mathbf{b}}+O(t^{-\delta}),

in which

2χ~0=1πi(Ω𝐚ln((1+s1s2)(1+s2s3)1)Ω𝐛ln(1+s1s2))+Γα,β,\displaystyle 2\tilde{\chi}_{0}=\frac{1}{\pi i}(\Omega_{\mathbf{a}}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)-\Omega_{\mathbf{b}}\ln(1+s_{1}s_{2}))+\Gamma_{\alpha,\beta},
Γα,β=(αβ)2πi(Ω𝐚𝐛zdzw(Aϕ,z)Ω𝐛𝐚zdzw(Aϕ,z)).\displaystyle\Gamma_{\alpha,\beta}=\frac{(\alpha-\beta)}{2\pi i}\Bigl{(}\Omega_{\mathbf{a}}\int_{\mathbf{b}}\frac{zdz}{w(A_{\phi},z)}-\Omega_{\mathbf{b}}\int_{\mathbf{a}}\frac{zdz}{w(A_{\phi},z)}\Bigr{)}.

Let us derive the asymptotic form as in Theorem 2.1 from (7.1). Set

u=0z++dzw(Aϕ,z),v=0z+dzw(Aϕ,z).u=-\int^{z^{+}_{+}}_{0}\frac{dz}{w(A_{\phi},z)},\quad v=-\int^{z^{+}_{-}}_{0}\frac{dz}{w(A_{\phi},z)}.

By the change of variables z=e3iϕAϕ(13e2iϕ)1z=e^{3i\phi}A_{\phi}(\wp-\tfrac{1}{3}e^{2i\phi})^{-1} (cf. Remark 2.1),

0z±+dzw(Aϕ,z)=±d43g2g3,\displaystyle-\int^{z^{+}_{\pm}}_{0}\frac{dz}{w(A_{\phi},z)}=\int^{\wp_{\pm}}_{\infty}\frac{d\wp}{\sqrt{4\wp^{3}-g_{2}\wp-g_{3}}},
±=13e2iϕ+e3iϕAϕz±1,g2=e4iϕ(434Aϕ),g3=e6iϕ(827+43AϕAϕ2),\displaystyle\wp_{\pm}=\tfrac{1}{3}{e^{2i\phi}}+e^{3i\phi}{A_{\phi}}z_{\pm}^{-1},\quad g_{2}=e^{4i\phi}(\tfrac{4}{3}-4A_{\phi}),\quad g_{3}=e^{6i\phi}(-\tfrac{8}{27}+\tfrac{4}{3}A_{\phi}-A_{\phi}^{2}),

and hence (u)=+,\wp(u)=\wp_{+}, (v)=\wp(v)=\wp_{-}. Observing that, by (4.2), (5.1) and (6.2),

z++z=ψ2eiϕ+O(t1),z+z=e3iϕAϕψ1+O(t1),\displaystyle z_{+}+z_{-}=-\psi-2e^{i\phi}+O(t^{-1}),\quad z_{+}z_{-}=-e^{3i\phi}A_{\phi}\psi^{-1}+O(t^{-1}),
w(Aϕ,z±)=λb3(λ)|λ2=z±=z±2+2(eiϕ+ψ)z±+O(t1),\displaystyle w(A_{\phi},z_{\pm})=\lambda b_{3}(\lambda)|_{\lambda^{2}=z_{\pm}}=z_{\pm}^{2}+2(e^{i\phi}+\psi)z_{\pm}+O(t^{-1}),

we have, up to the error term +O(t1)+O(t^{-1}),

(u)+(v)=\displaystyle\wp(u)+\wp(v)= ++=23e2iϕ+e3iϕAϕ(z++z)(z+z)1=23e2iϕ+ψ2+2eiϕψ,\displaystyle\wp_{+}+\wp_{-}=\tfrac{2}{3}e^{2i\phi}+e^{3i\phi}A_{\phi}(z_{+}+z_{-})(z_{+}z_{-})^{-1}=\tfrac{2}{3}e^{2i\phi}+\psi^{2}+2e^{i\phi}\psi,
((u)(v))=\displaystyle-(\wp^{\prime}(u)-\wp^{\prime}(v))= e3iϕAϕ(z+2w(Aϕ,z+)z2w(Aϕ,z))=2(eiϕ+ψ)e3iϕAϕ(z+1z1)\displaystyle e^{3i\phi}A_{\phi}(z_{+}^{-2}w(A_{\phi},z_{+})-z_{-}^{-2}w(A_{\phi},z_{-}))=2(e^{i\phi}+\psi)e^{3i\phi}A_{\phi}(z_{+}^{-1}-z_{-}^{-1})
=\displaystyle= 2(eiϕ+ψ)(+)=2(eiϕ+ψ)((u)(v)).\displaystyle 2(e^{i\phi}+\psi)(\wp_{+}-\wp_{-})=2(e^{i\phi}+\psi)(\wp(u)-\wp(v)).

Then by the addition theorem

(u+v)=(u)(v)+14((u)(v)(u)(v))2=e2iϕ3+O(t1),\wp(u+v)=-\wp(u)-\wp(v)+\frac{1}{4}\Bigl{(}\frac{\wp^{\prime}(u)-\wp^{\prime}(v)}{\wp(u)-\wp(v)}\Bigr{)}^{2}=\frac{e^{2i\phi}}{3}+O(t^{-1}),

which implies, by Proposition 7.6,

(0z+++0z+)dzw(Aϕ,z)=\displaystyle\biggl{(}\int^{z_{+}^{+}}_{0}+\int^{z_{-}^{+}}_{0}\biggr{)}\frac{dz}{w(A_{\phi},z)}= (u+v)=e2iϕ/3d43g2g3\displaystyle-(u+v)=-\int_{\infty}^{e^{2i\phi}/3}\frac{d\wp}{\sqrt{4\wp^{3}-g_{2}\wp-g_{3}}}
=\displaystyle= 0dzw(Aϕ,z)=Ω0{±13Ω𝐛}.\displaystyle\int^{\infty}_{0}\frac{dz}{w(A_{\phi},z)}=\Omega_{0}\in\{\pm\tfrac{1}{3}\Omega_{\mathbf{b}}\}.

This combined with (7.1) leads to

(7.2) 0z++dzw(Aϕ,z)=eiϕt+χ~0+12(Ω𝐛+Ω0)+O(tδ),0z+dzw(Aϕ,z)=eiϕtχ~012(Ω𝐛Ω0)+O(tδ).\begin{split}&\int^{z^{+}_{+}}_{0}\frac{dz}{w(A_{\phi},z)}=e^{i\phi}t+\tilde{\chi}_{0}+\frac{1}{2}(\Omega_{\mathbf{b}}+\Omega_{0})+O(t^{-\delta}),\\ &\int^{z^{+}_{-}}_{0}\frac{dz}{w(A_{\phi},z)}=-e^{i\phi}t-\tilde{\chi}_{0}-\frac{1}{2}(\Omega_{\mathbf{b}}-\Omega_{0})+O(t^{-\delta}).\end{split}
Remark 7.1.

A similar argument with use of the addition theorem leads to

(0z+++0z+)dzw(aϕ,z)=e2iϕ/3d43g~2g~3+O(t1)\displaystyle\biggl{(}\int^{z_{+}^{+}}_{0}+\int^{z_{-}^{+}}_{0}\biggr{)}\frac{dz}{w(a_{\phi},z)}=-\int_{\infty}^{e^{2i\phi}/3}\frac{d\wp}{\sqrt{4\wp^{3}-\tilde{g}_{2}\wp-\tilde{g}_{3}}}+O(t^{-1})
=e2iϕ/3d43g2g3+O(t1)=(0z+++0z+)dzw(Aϕ,z)+O(t1),\displaystyle=-\int_{\infty}^{e^{2i\phi}/3}\frac{d\wp}{\sqrt{4\wp^{3}-{g}_{2}\wp-{g}_{3}}}+O(t^{-1})=\biggl{(}\int^{z_{+}^{+}}_{0}+\int^{z_{-}^{+}}_{0}\biggr{)}\frac{dz}{w(A_{\phi},z)}+O(t^{-1}),

where g~2=e4iϕ(434aϕ),\tilde{g}_{2}=e^{4i\phi}(\tfrac{4}{3}-4a_{\phi}), g~3=e6iϕ(278+43aϕaϕ2).\tilde{g}_{3}=e^{6i\phi}(-\tfrac{27}{8}+\tfrac{4}{3}a_{\phi}-a_{\phi}^{2}). Combining this with Proposition 6.8 we have

0z±+dzw(aϕ,z)=0z±+dzw(Aϕ,z)+O(t1).\int^{z^{+}_{\pm}}_{0}\frac{dz}{w(a_{\phi},z)}=\int^{z^{+}_{\pm}}_{0}\frac{dz}{w(A_{\phi},z)}+O(t^{-1}).

Recall that P(u;Aϕ)\mathrm{P}(u;A_{\phi}) solves Pu2=w(Aϕ,P)2.\mathrm{P}_{u}^{2}=w(A_{\phi},\mathrm{P})^{2}. By Proposition 7.6,

(7.3) P(u;Aϕ)=1Ω𝐚(ϑϑ(uΩ𝐚+τ3+ν,τ)ϑϑ(uΩ𝐚τ3+ν,τ)+CP)\mathrm{P}(u;A_{\phi})=\frac{1}{\Omega_{\mathbf{a}}}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{3}+\nu,\tau^{*}\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{3}+\nu,\tau^{*}\Bigr{)}+C_{\mathrm{P}}\Bigr{)}

with CP=(ϑ/ϑ)(13τ+ν,τ)+(ϑ/ϑ)(13τ+ν,τ)C_{\mathrm{P}}=-(\vartheta^{\prime}/\vartheta)(\tfrac{1}{3}\tau^{*}+\nu,\tau^{*})+(\vartheta^{\prime}/\vartheta)(-\tfrac{1}{3}\tau^{*}+\nu,\tau^{*}) and τ=Ω𝐛/Ω𝐚\tau^{*}=\Omega_{\mathbf{b}}/\Omega_{\mathbf{a}}. Then we have

ψ(t)=\displaystyle\psi(t)= z+z2eiϕ+O(t1)\displaystyle-z_{+}-z_{-}-2e^{i\phi}+O(t^{-1})
=\displaystyle= P(t~ϕ+12(Ω𝐛+Ω0);Aϕ)P(t~ϕ+12(Ω𝐛Ω0);Aϕ)2eiϕ+O(tδ)\displaystyle-\mathrm{P}(\tilde{t}_{\phi}+\tfrac{1}{2}(\Omega_{\mathbf{b}}+\Omega_{0});A_{\phi})-\mathrm{P}(\tilde{t}_{\phi}+\tfrac{1}{2}(\Omega_{\mathbf{b}}-\Omega_{0});A_{\phi})-2e^{i\phi}+O(t^{-\delta})
=\displaystyle= 1Ω𝐚(ϑϑ(t~ϕΩ𝐚τ6±τ6+ν)ϑϑ(t~ϕΩ𝐚+τ6±τ6+ν)\displaystyle-\frac{1}{\Omega_{\mathbf{a}}}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{6}\pm\frac{\tau^{*}}{6}+\nu\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{6}\pm\frac{\tau^{*}}{6}+\nu\Bigr{)}
+ϑϑ(t~ϕΩ𝐚τ6τ6+ν)ϑϑ(t~ϕΩ𝐚+τ6τ6+ν))+Cϕ+O(tδ)\displaystyle+\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{6}\mp\frac{\tau^{*}}{6}+\nu\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{6}\mp\frac{\tau^{*}}{6}+\nu\Bigr{)}\Bigr{)}+C_{\phi}+O(t^{-\delta})
=\displaystyle= 1Ω𝐚(ϑϑ(t~ϕΩ𝐚+τ3+ν)ϑϑ(t~ϕΩ𝐚τ3+ν))+Cϕ+O(tδ)\displaystyle\frac{1}{\Omega_{\mathbf{a}}}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{3}+\nu\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{\tilde{t}_{\phi}}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{3}+\nu\Bigr{)}\Bigr{)}+C_{\phi}+O(t^{-\delta})
=\displaystyle= P(t~ϕ;Aϕ)+Cϕ,0+O(tδ),\displaystyle\mathrm{P}(\tilde{t}_{\phi};A_{\phi})+C_{\phi,0}+O(t^{-\delta}),

where Cϕ,Cϕ,0,C_{\phi},C_{\phi,0}\in\mathbb{C}, and t~ϕ=eiϕt+χ~0\tilde{t}_{\phi}=e^{i\phi}t+\tilde{\chi}_{0}. If ψt=(d/dt)ψ+O(t1)\psi_{t}=(d/dt)\psi+O(t^{-1}), then by (4.2) e2iϕ((d/dt)ψ)2=w(Aϕ,ψ)2+O(t1)e^{-2i\phi}((d/dt)\psi)^{2}=w(A_{\phi},\psi)^{2}+O(t^{-1}), which implies Cϕ,0=0.C_{\phi,0}=0. In the case 0<ϕ<π/40<\phi<\pi/4, for 𝐬\mathbf{s} such that (1+s1s2)(1+s2s3)10,(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, 1+s2s301+s_{2}s_{3}\not=0, the same argument is possible. Thus we have the following.

Proposition 7.1.

If ψt=dψ/dt+O(t1),\psi_{t}=d\psi/dt+O(t^{-1}), then equation (7.1) in Sϕ(t,κ1,δ1)S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}) implies

ψ(t)=P(eiϕt+χ~0;Aϕ)+O(tδ),\psi(t)=\mathrm{P}(e^{i\phi}t+\tilde{\chi}_{0};A_{\phi})+O(t^{-\delta}),

in which

2χ~0=1πi(Ω𝐚ln((1+s1s2)(1+s2s3)1)Ω𝐛𝔩(𝐬,ϕ))+Γα,β,2\tilde{\chi}_{0}=\frac{1}{\pi i}(\Omega_{\mathbf{a}}\ln((1+s_{1}s_{2})(1+s_{2}s_{3})-1)-\Omega_{\mathbf{b}}\mathfrak{l}(\mathbf{s},\phi))+\Gamma_{\alpha,\beta},

with 𝔩(𝐬,ϕ)=ln(1+s1s2)\mathfrak{l}(\mathbf{s},\phi)=\ln(1+s_{1}s_{2}) if π/4<ϕ<0,-\pi/4<\phi<0, and =ln(1+s2s3)=-\ln(1+s_{2}s_{3}) if 0<ϕ<π/40<\phi<\pi/4.

Let us calculate the value Γα,β\Gamma_{\alpha,\beta} in χ~0.\tilde{\chi}_{0}. The substitution z=P(u;Aϕ)z=\mathrm{P}(u;A_{\phi}) with (7.3) leads to

𝐛zdzw(Aϕ,z)=\displaystyle\int_{\mathbf{b}}\frac{zdz}{w(A_{\phi},z)}= 0Ω𝐛P(u;Aϕ)𝑑u\displaystyle\int^{\Omega_{\mathbf{b}}}_{0}\mathrm{P}(u;A_{\phi})du
=\displaystyle= 0Ω𝐛1Ω𝐚(ϑϑ(uΩ𝐚+τ3+ν)ϑϑ(uΩ𝐚τ3+ν)+CP)𝑑u\displaystyle\int^{\Omega_{\mathbf{b}}}_{0}\frac{1}{\Omega_{\mathbf{a}}}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{3}+\nu\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{3}+\nu\Bigr{)}+C_{\mathrm{P}}\Bigr{)}du
=\displaystyle= [lnϑ(Ωa1u+13τ+ν)lnϑ(Ωa1u13τ+ν)]0Ω𝐛+CPτ\displaystyle\Bigl{[}\ln\vartheta(\Omega_{\mathrm{a}}^{-1}u+\tfrac{1}{3}\tau^{*}+\nu)-\ln\vartheta(\Omega_{\mathrm{a}}^{-1}u-\tfrac{1}{3}\tau^{*}+\nu)\Bigr{]}_{0}^{\Omega_{\mathbf{b}}}+C_{\mathrm{P}}\tau^{*}
=\displaystyle= πi(τ+2(13τ+ν))+πi(τ+2(13τ+ν))+CPτ\displaystyle-\pi i(\tau^{*}+2(\tfrac{1}{3}\tau^{*}+\nu))+\pi i(\tau^{*}+2(-\tfrac{1}{3}\tau^{*}+\nu))+C_{\mathrm{P}}\tau^{*}
=\displaystyle= (43πi+CP)τ,\displaystyle(-\tfrac{4}{3}\pi i+C_{\mathrm{P}})\tau^{*},

and

𝐚zdzw(Aϕ,z)=0Ω𝐚1Ω𝐚(ϑϑ(uΩ𝐚+τ3+ν)ϑϑ(uΩ𝐚τ3+ν)+CP)𝑑u=CP.\int_{\mathbf{a}}\frac{zdz}{w(A_{\phi},z)}=\int^{\Omega_{\mathbf{a}}}_{0}\frac{1}{\Omega_{\mathbf{a}}}\Bigl{(}\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}+\frac{\tau^{*}}{3}+\nu\Bigr{)}-\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{u}{\Omega_{\mathbf{a}}}-\frac{\tau^{*}}{3}+\nu\Bigr{)}+C_{\mathrm{P}}\Bigr{)}du=C_{\mathrm{P}}.

From these quantities, the required constant follows.

Proposition 7.2.

We have

Γα,β=Γα,β,𝐛Γα,β,𝐚=23(αβ)Ω𝐛,\displaystyle\Gamma_{\alpha,\beta}=\Gamma_{\alpha,\beta,\mathbf{b}}-\Gamma_{\alpha,\beta,\mathbf{a}}=-\frac{2}{3}(\alpha-\beta)\Omega_{\mathbf{b}},
Γα,β,𝐛=Γα,β+12πi(αβ)CPΩ𝐛,Γα,β,𝐚=12πi(αβ)CPΩ𝐛\displaystyle\Gamma_{\alpha,\beta,\mathbf{b}}=\Gamma_{\alpha,\beta}+\frac{1}{2\pi i}(\alpha-\beta)C_{\mathrm{P}}\Omega_{\mathbf{b}},\quad\Gamma_{\alpha,\beta,\mathbf{a}}=\frac{1}{2\pi i}(\alpha-\beta)C_{\mathrm{P}}\Omega_{\mathbf{b}}

with CP=(ϑ/ϑ)(13τ+ν,τ)+(ϑ/ϑ)(13τ+ν,τ)C_{\mathrm{P}}=-(\vartheta^{\prime}/\vartheta)(\tfrac{1}{3}\tau^{*}+\nu,\tau^{*})+(\vartheta^{\prime}/\vartheta)(-\tfrac{1}{3}\tau^{*}+\nu,\tau^{*}).

By Propositions 7.1 and 7.2 we may derive from (7.1) asymptotic forms as in Theorems 2.1 and 2.2.

Proposition 7.3.

In (7.2), Ω0=13Ω𝐛\Omega_{0}=-\tfrac{1}{3}\Omega_{\mathbf{b}}.

Proof..

Note that ψ(t)=P(eiϕt+χ~0;Aϕ)+O(tδ)\psi(t)=\mathrm{P}(e^{i\phi}t+\tilde{\chi}_{0};A_{\phi})+O(t^{-\delta}). By Proposition 7.6, if ψ(t)=e3iϕAϕ(eiϕtt0)2(1+o(1))\psi(t)=e^{3i\phi}A_{\phi}(e^{i\phi}t-t_{0})^{2}(1+o(1)) around t=t0t=t_{0}, then ψ(t)(eiϕtt0)1eiϕ+o(1)\psi(t)\sim(e^{i\phi}t-t_{0}^{-})^{-1}-e^{i\phi}+o(1) and ψ(t)(eiϕtt0+)1eiϕ+o(1)\psi(t)\sim-(e^{i\phi}t-t_{0}^{+})^{-1}-e^{i\phi}+o(1) around each of the points t0±:=t0±13Ω𝐛t_{0}^{\pm}:=t_{0}\pm\tfrac{1}{3}\Omega_{\mathbf{b}} up to O(tδ).O(t^{-\delta}). Then, by z±(t)=12(eiϕψtψ1ψ2eiϕ)+O(t1)z_{\pm}(t)=\tfrac{1}{2}(\mp e^{-i\phi}\psi_{t}\psi^{-1}-\psi-2e^{i\phi})+O(t^{-1}), it follows that

z+(t)=O(eiϕtt0),z+(t)=(eiϕtt0)1+O(1),z+(t)=(eiϕtt0+)1+O(1),\displaystyle z_{+}(t)=O(e^{i\phi}t-t_{0}^{-}),\quad z_{+}(t)=-(e^{i\phi}t-t_{0})^{-1}+O(1),\quad z_{+}(t)=(e^{i\phi}t-t_{0}^{+})^{-1}+O(1),
z(t)=(eiϕtt0)1+O(1),z(t)=(eiϕtt0)1+O(1),z(t)=O(eiϕtt0+)\displaystyle z_{-}(t)=-(e^{i\phi}t-t_{0}^{-})^{-1}+O(1),\quad z_{-}(t)=(e^{i\phi}t-t_{0})^{-1}+O(1),\quad z_{-}(t)=O(e^{i\phi}t-t_{0}^{+})

around each point. Since z±(t)=P(eiϕt+χ~0+12(Ω𝐛±Ω0);Aϕ)+O(tδ)z_{\pm}(t)=\mathrm{P}(e^{i\phi}t+\tilde{\chi}_{0}+\tfrac{1}{2}(\Omega_{\mathbf{b}}\pm\Omega_{0});A_{\phi})+O(t^{-\delta}) we conclude that Ω0=13Ω𝐛.\Omega_{0}=-\tfrac{1}{3}\Omega_{\mathbf{b}}.

7.2. Asymptotic representation of Bϕ(t)B_{\phi}(t)

Recalling Propositions 6.6 and 6.7, and applying Remark 7.1, we have

t𝒥𝐚+2e3iϕΩ𝐚Bϕ(t)=\displaystyle t\mathcal{J}_{\mathbf{a}}+2e^{3i\phi}\Omega_{\mathbf{a}}B_{\phi}(t)= ϑϑ(12F(z+,z++)+12+τ6,τ)ϑϑ(12F(z,z+)+12+τ6,τ)\displaystyle\frac{\vartheta^{\prime}}{\vartheta}(\tfrac{1}{2}F_{*}(z^{-}_{+},z^{+}_{+})+\tfrac{1}{2}+\tfrac{\tau^{*}}{6},\tau^{*})-\frac{\vartheta^{\prime}}{\vartheta}(\tfrac{1}{2}F_{*}(z^{-}_{-},z^{+}_{-})+\tfrac{1}{2}+\tfrac{\tau^{*}}{6},\tau^{*})
(αβ)CP+2πi2𝔩(𝐬,ϕ)+O(tδ),\displaystyle-(\alpha-\beta)C_{\mathrm{P}}+2\pi i-2\mathfrak{l}(\mathbf{s},\phi)+O(t^{-\delta}),

where

F(z±,z±+)=1Ω𝐚z±z±+dzw(Aϕ,z)=2Ω𝐚0z±+dzw(Aϕ,z).F_{*}(z^{-}_{\pm},z^{+}_{\pm})=\frac{1}{\Omega_{\mathbf{a}}}\int_{z^{-}_{\pm}}^{z^{+}_{\pm}}\frac{dz}{w(A_{\phi},z)}=\frac{2}{\Omega_{\mathbf{a}}}\int_{0}^{z^{+}_{\pm}}\frac{dz}{w(A_{\phi},z)}.

By Proposition 7.3, insertion of (7.2) with Ω0=13Ω𝐛\Omega_{0}=-\tfrac{1}{3}\Omega_{\mathbf{b}} yields the asymptotic representation of the correction function Bϕ(t),B_{\phi}(t), which plays an essential role in the justification of our asymptotic solution.

Proposition 7.4.

In Sϕ(t,κ1,δ1),S_{\phi}(t^{\prime}_{\infty},\kappa_{1},\delta_{1}),

t𝒥𝐚+2e3iϕΩ𝐚Bϕ(t)=2ϑϑ(eiϕt+χ0~Ω𝐚+ν,τ)(αβ)CP+2πi2𝔩(𝐬,ϕ)+O(tδ),t\mathcal{J}_{\mathbf{a}}+2e^{3i\phi}\Omega_{\mathbf{a}}B_{\phi}(t)=2\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{e^{i\phi}t+\tilde{\chi_{0}}}{\Omega_{\mathbf{a}}}+\nu,\tau^{*}\Bigr{)}-(\alpha-\beta)C_{\mathrm{P}}+2\pi i-2\mathfrak{l}(\mathbf{s},\phi)+O(t^{-\delta}),

where 𝔩(𝐬,ϕ)\mathfrak{l}(\mathbf{s},\phi) and CPC_{\mathrm{P}} are constants given in Propositions 7.1 and 7.2.

7.3. Proofs of Theorems 2.1 and 2.2

For a prescribed monodromy data 𝐬\mathbf{s}, the asymptotic expression of ψ(t)\psi(t) given in Proposition 7.1 is, at least formally, a solution of the inverse monodromy problem. To prove Theorems 2.1 and 2.2 let us make the justification for y=eiϕxψ(t)y=e^{-i\phi}x\psi(t) as a solution of PIV\mathrm{P}_{\mathrm{IV}} along the lines in [21, pp. 105–106, pp. 120–121]. Suppose that π/4<ϕ<0.-\pi/4<\phi<0. Let 𝐬=(s1,s2,s3,s4)\mathbf{s}=(s_{1},s_{2},s_{3},s_{4}) with (1+s1s2)(1+s2s3)10,(1+s_{1}s_{2})(1+s_{2}s_{3})-1\not=0, 1+s1s201+s_{1}s_{2}\not=0 be a given point on the monodromy manifold 0(α,β)\mathcal{M}_{0}(\alpha,\beta) for isomonodromy system (3.1). Set

ψas=ψas(𝐬,t):=\displaystyle\psi_{\mathrm{as}}=\psi_{\mathrm{as}}(\mathbf{s},t):= P(eiϕt+χ~0;Aϕ),\displaystyle\mathrm{P}(e^{i\phi}t+\tilde{\chi}_{0};A_{\phi}),
(Bϕ)as=(Bϕ)as(𝐬,t):=\displaystyle(B_{\phi})_{\mathrm{as}}=(B_{\phi})_{\mathrm{as}}(\mathbf{s},t):= e3iϕ2Ω𝐚(2ϑϑ(eiϕt+χ0~Ω𝐚+ν,τ)t𝒥𝐚+Cα,β,𝐬,ϕ),\displaystyle\frac{e^{-3i\phi}}{2\Omega_{\mathbf{a}}}\Bigl{(}2\frac{\vartheta^{\prime}}{\vartheta}\Bigl{(}\frac{e^{i\phi}t+\tilde{\chi_{0}}}{\Omega_{\mathbf{a}}}+\nu,\tau^{*}\Bigr{)}-t\mathcal{J}_{\mathbf{a}}+C_{\alpha,\beta,\mathbf{s},\phi}\Bigr{)},
Cα,β,𝐬,ϕ\displaystyle C_{\alpha,\beta,\mathbf{s},\phi} =2πi(αβ)CP2𝔩(𝐬,ϕ),\displaystyle=2\pi i-(\alpha-\beta)C_{\mathrm{P}}-2\mathfrak{l}(\mathbf{s},\phi),

which are leading term expressions of ψ(t)\psi(t) and Bϕ(t)B_{\phi}(t) without O(tδ)O(t^{-\delta}) in Propositions 7.1 and 7.4. Taking (4.2) and (5.1) into account, we set

eiϕψas=\displaystyle e^{-i\phi}\psi^{*}_{\mathrm{as}}= 12eiϕψast1+ψas((ψas+2eiϕ)2ψas+4e3iϕAϕ)+Δ(t,ψas,(Bϕ)as)t1,\displaystyle-\tfrac{1}{2}e^{-i\phi}\psi_{\mathrm{as}}t^{-1}+\sqrt{\psi_{\mathrm{as}}((\psi_{\mathrm{as}}+2e^{i\phi})^{2}\psi_{\mathrm{as}}+4e^{3i\phi}A_{\phi})+\Delta(t,\psi_{\mathrm{as}},(B_{\phi})_{\mathrm{as}})t^{-1}},
Δ(t,ψas,\displaystyle\Delta(t,\psi_{\mathrm{as}}, (Bϕ)as)=ψas(4(Bϕ)as(4αβ)ψas2(2αβ)eiϕ)+14β2t1,\displaystyle(B_{\phi})_{\mathrm{as}})=\psi_{\mathrm{as}}(4(B_{\phi})_{\mathrm{as}}-(4\alpha-\beta)\psi_{\mathrm{as}}-2(2\alpha-\beta)e^{i\phi})+\tfrac{1}{4}\beta^{2}t^{-1},

where the branch of the square root is chosen in such a way that ψas\psi^{*}_{\mathrm{as}} is compatible with (d/dt)ψas(d/dt)\psi_{\mathrm{as}}, that is, ψas=(d/dt)ψas+O(t1).\psi^{*}_{\mathrm{as}}=(d/dt)\psi_{\mathrm{as}}+O(t^{-1}). Then for aϕ(t,ψas,ψas)a_{\phi}(t,\psi_{\mathrm{as}},\psi^{*}_{\mathrm{as}}) and for (Bϕ)as(B_{\phi})_{\mathrm{as}} inequality (5.1) is valid in the domain

S~(ϕ,t,κ0,δ2)={t|Ret>t,|Imt|<κ0}tZϕ{|tt|<δ2},Zϕ=ZϕZϕ0,\tilde{S}(\phi,t_{\infty},\kappa_{0},\delta_{2})=\{t\,|\,\mathrm{Re\,}t>t_{\infty},\,|\mathrm{Im\,}t|<\kappa_{0}\}\setminus\bigcup_{t_{*}\in Z_{\phi}}\{|t-t_{*}|<\delta_{2}\},\quad Z_{\phi}=Z_{\phi}^{\infty}\cup Z_{\phi}^{0},

where Zϕ={t|eiϕt+χ~0=±13Ω𝐛+Ω𝐚+Ω𝐛},Z_{\phi}^{\infty}=\{t_{*}\,|\,e^{i\phi}t_{*}+\tilde{\chi}_{0}=\pm\tfrac{1}{3}\Omega_{\mathbf{b}}+\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}\}, Zϕ0={t|eiϕt+χ~0Ω𝐚+Ω𝐛}.Z_{\phi}^{0}=\{t_{*}\,|\,e^{i\phi}t_{*}+\tilde{\chi}_{0}\in\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}\}. Consider system (3.1) with (t,λ)\mathcal{B}(t,\lambda) containing (ψas,ψas)=(ψas(𝐬,t),ψas(𝐬,t))(\psi_{\mathrm{as}},\psi_{\mathrm{as}}^{*})=(\psi_{\mathrm{as}}(\mathbf{s},t),\psi_{\mathrm{as}}^{*}(\mathbf{s},t)). Then the direct monodromy problem for this system by the WKB analysis results in the monodromy data (𝐬)as(t)(\mathbf{s})_{\mathrm{as}}(t) such that (𝐬)as(t)𝐬C|t|δ\|(\mathbf{s})_{\mathrm{as}}(t)-\mathbf{s}\|\leq C|t|^{-\delta} for |t|>t(𝐬)|t|>t_{\infty}(\mathbf{s}), in which CC and δ\delta are some constant independent of 𝐬\mathbf{s}. Then the justification scheme of Kitaev [19] applies to our case. By the maximal modulus principle the excluded disc around each point in Zϕ0Z_{\phi}^{0} is removed. Thus we obtain Theorems 2.1 and 2.2.

7.4. Proof of Theorem 2.4

For given nn\in\mathbb{Z} the substitution

ϕ=12πn+ϕ~,x=eiπn/2x~,ξ=eiπn/4ξ~,u=eiπn/4u~(n),v=eiπn/4v~(n),\displaystyle\phi=\tfrac{1}{2}\pi n+\tilde{\phi},\quad x=e^{i\pi n/2}\tilde{x},\quad\xi=e^{i\pi n/4}\tilde{\xi},\quad\mathrm{u}=e^{i\pi n/4}\tilde{\mathrm{u}}_{(n)},\quad\mathrm{v}=e^{i\pi n/4}\tilde{\mathrm{v}}_{(n)},
α=eiπnα~,β=eiπnβ~,ψ=eiπn/2ψ~,y=eiπn/2y~,\displaystyle\alpha=e^{i\pi n}\tilde{\alpha},\quad\beta=e^{i\pi n}\tilde{\beta},\quad\psi=e^{i\pi n/2}\tilde{\psi},\quad y=e^{i\pi n/2}\tilde{y},

where (u~(n),v~(n))=(v~,u~)(\tilde{\mathrm{u}}_{(n)},\tilde{\mathrm{v}}_{(n)})=(\tilde{\mathrm{v}},\tilde{\mathrm{u}}) if nn is odd, and =(u~,v~)=(\tilde{\mathrm{u}},\tilde{\mathrm{v}}) if nn is even, changes isomonodromy system (1.1) to

(7.4) dΨ~dξ~=\displaystyle\frac{d\tilde{\Psi}}{d\tilde{\xi}}= σ2n((ξ~32+ξ~(x~+u~(n)v~(n))+α~ξ~)σ3\displaystyle\sigma_{2}^{n}\Biggl{(}\Bigl{(}\frac{\tilde{\xi}^{3}}{2}+\tilde{\xi}(\tilde{x}+\tilde{\mathrm{u}}_{(n)}\tilde{\mathrm{v}}_{(n)})+\frac{\tilde{\alpha}}{\tilde{\xi}}\Bigr{)}\sigma_{3}
+i(0ξ~2u~(n)+2x~u~(n)+(u~(n))x~ξ~2v~(n)+2x~v~(n)(v~(n))x~0))σ2nΨ~,\displaystyle\phantom{--}+i\begin{pmatrix}0&\tilde{\xi}^{2}\tilde{\mathrm{u}}_{(n)}+2\tilde{x}\tilde{\mathrm{u}}_{(n)}+(\tilde{u}_{(n)})_{\tilde{x}}\\ \tilde{\xi}^{2}\tilde{\mathrm{v}}_{(n)}+2\tilde{x}\tilde{\mathrm{v}}_{(n)}-(\tilde{v}_{(n)})_{\tilde{x}}&0\end{pmatrix}\Biggr{)}\sigma_{2}^{n}\tilde{\Psi},
β~=\displaystyle\tilde{\beta}= (u~(n))x~v~(n)u~(n)(v~(n))x~+2x~u~(n)v~(n)(u~(n)v~(n))2,y~=u~(n)v~(n),\displaystyle(\tilde{\mathrm{u}}_{(n)})_{\tilde{x}}\tilde{\mathrm{v}}_{(n)}-\tilde{\mathrm{u}}_{(n)}(\tilde{\mathrm{v}}_{(n)})_{\tilde{x}}+2\tilde{x}\tilde{\mathrm{u}}_{(n)}\tilde{\mathrm{v}}_{(n)}-(\tilde{\mathrm{u}}_{(n)}\tilde{\mathrm{v}}_{(n)})^{2},\quad\tilde{y}=\tilde{\mathrm{u}}_{(n)}\tilde{\mathrm{v}}_{(n)},

and y=eiϕxψy=e^{-i\phi}x\psi to y~=eiϕ~x~ψ~.\tilde{y}=e^{-i\tilde{\phi}}\tilde{x}\tilde{\psi}. For kk\in\mathbb{Z} system (7.4) admits the canonical solutions

σ2nΨ~k(ξ~)=Ψk(ξ~)=(I+O(ξ~1))exp((18ξ~4+12x~ξ~2+(α~β~)lnξ~)σ3)\sigma_{2}^{n}\tilde{\Psi}^{\infty}_{k}(\tilde{\xi})=\Psi_{k}^{\infty}(\tilde{\xi})=(I+O(\tilde{\xi}^{-1}))\exp((\tfrac{1}{8}\tilde{\xi}^{4}+\tfrac{1}{2}\tilde{x}\tilde{\xi}^{2}+(\tilde{\alpha}-\tilde{\beta})\ln\tilde{\xi})\sigma_{3})

as ξ~\tilde{\xi}\to\infty through the sector |argξ~+π8π4k|<π4|\arg\tilde{\xi}+\tfrac{\pi}{8}-\tfrac{\pi}{4}k|<\tfrac{\pi}{4}, where Ψk(ξ)\Psi^{\infty}_{k}(\xi) are solutions of (1.1) given by (2.1). The Stokes matrices S~k\tilde{S}_{k} (k)(k\in\mathbb{Z}) with

S~2l1=(1s~2l101),S~2l=(10s~2l1)(l)\tilde{S}_{2l-1}=\begin{pmatrix}1&\tilde{s}_{2l-1}\\ 0&1\end{pmatrix},\quad\tilde{S}_{2l}=\begin{pmatrix}1&0\\ \tilde{s}_{2l}&1\end{pmatrix}\quad(l\in\mathbb{Z})

for system (7.4) are defined by Ψk+1(ξ~)=Ψk(ξ~)S~k.\Psi_{k+1}^{\infty}(\tilde{\xi})=\Psi_{k}^{\infty}(\tilde{\xi})\tilde{S}_{k}. Observing that

Ψk(ξ~)=\displaystyle\Psi^{\infty}_{k}(\tilde{\xi})= (I+O(ξ1))exp((1)n(18ξ4+12xξ2+(αβ)lnξ)σ3)e(1)n(αβ)(iπn/4)σ3\displaystyle(I+O(\xi^{-1}))\exp((-1)^{n}(\tfrac{1}{8}\xi^{4}+\tfrac{1}{2}x\xi^{2}+(\alpha-\beta)\ln\xi)\sigma_{3})e^{-(-1)^{n}(\alpha-\beta)(i\pi n/4)\sigma_{3}}
=\displaystyle= σ2n((I+O(ξ1))exp((18ξ4+12xξ2+(αβ)lnξ)σ3)e(αβ)(iπn/4)σ3)σ2n\displaystyle\sigma_{2}^{n}\Bigl{(}(I+O(\xi^{-1}))\exp((\tfrac{1}{8}\xi^{4}+\tfrac{1}{2}x\xi^{2}+(\alpha-\beta)\ln\xi)\sigma_{3})e^{-(\alpha-\beta)(i\pi n/4)\sigma_{3}}\Bigr{)}\sigma_{2}^{n}

in |argξ+π8π4(k+n)|<π4|\arg\xi+\tfrac{\pi}{8}-\tfrac{\pi}{4}(k+n)|<\tfrac{\pi}{4}, we have Ψk(ξ~)=σ2nΨk+n(ξ)e(αβ)(iπn/4)σ3σ2n.\Psi_{k}^{\infty}(\tilde{\xi})=\sigma_{2}^{n}\Psi^{\infty}_{k+n}(\xi)e^{-(\alpha-\beta)(i\pi n/4)\sigma_{3}}\sigma_{2}^{n}. This relation immediately leads to

S~k=σ2ne(αβ)(iπn/4)σ3Sk+ne(αβ)(iπn/4)σ3σ2n\tilde{S}_{k}=\sigma_{2}^{n}e^{(\alpha-\beta)(i\pi n/4)\sigma_{3}}S_{k+n}e^{-(\alpha-\beta)(i\pi n/4)\sigma_{3}}\sigma_{2}^{n}

(cf. [13, (13)]), which implies s~ks~k+1=sk+nsk+1+n\tilde{s}_{k}\tilde{s}_{k+1}=s_{k+n}s_{k+1+n}. Thus system (1.1) for 0<|ϕπn/2|<π/40<|\phi-\pi n/2|<\pi/4 is converted to (7.4) for 0<|ϕ~|<π/40<|\tilde{\phi}|<\pi/4 with the monodromy data 𝐬~=(s~1,s~2,s~3,s~4)=𝐬n=(s1+n,s2+n,s3+n,s4+n)\tilde{\mathbf{s}}=(\tilde{s}_{1},\tilde{s}_{2},\tilde{s}_{3},\tilde{s}_{4})=\mathbf{s}_{n}=(s_{1+n},s_{2+n},s_{3+n},s_{4+n}). Then application of Corollary 2.3 yields the theorem.

7.5. Properties of the elliptic function P(u;A)\mathrm{P}(u;A)

The elliptic function p=P(u;A)p=\mathrm{P}(u;A) is a solution of

(7.5) 4e3iϕA=(p)2pp(p+2eiϕ)2(p=dp/du).4e^{3i\phi}A=\frac{(p^{\prime})^{2}}{p}-p(p+2e^{i\phi})^{2}\qquad(p^{\prime}=dp/du).

About this equation relations in (7.2) suggest the following interesting fact.

Proposition 7.5.

Let p=η=η(u)p=\eta=\eta(u) be a given solution of (7.5). Then the functions χ±=12ηη112ηeiϕ\chi_{\pm}=\mp\tfrac{1}{2}\eta^{\prime}\eta^{-1}-\tfrac{1}{2}\eta-e^{i\phi} also solves (7.5).

Proof..

Note that

(χ±+12η+eiϕ)2=14(η)2η2=14η1(4e3iϕA+η(η+2eiϕ)2)=e3iϕAη1+14(η+2eiϕ)2,(\chi_{\pm}+\tfrac{1}{2}\eta+e^{i\phi})^{2}=\tfrac{1}{4}(\eta^{\prime})^{2}\eta^{-2}=\tfrac{1}{4}\eta^{-1}(4e^{3i\phi}A+\eta(\eta+2e^{i\phi})^{2})=e^{3i\phi}A\eta^{-1}+\tfrac{1}{4}(\eta+2e^{i\phi})^{2},

which implies that χ±\chi_{\pm} and η\eta satisfy the relation

(7.6) η2χ±+ηχ±2+2eiϕηχ±e3iϕA=0.\eta^{2}\chi_{\pm}+\eta\chi_{\pm}^{2}+2e^{i\phi}\eta\chi_{\pm}-e^{3i\phi}A=0.

Since η\eta solves (7.5), using (7.6) we have

χ±\displaystyle\mp\chi^{\prime}_{\pm} =12(η′′η1(η)2η2)±12η\displaystyle=\tfrac{1}{2}(\eta^{\prime\prime}\eta^{-1}-(\eta^{\prime})^{2}\eta^{-2})\pm\tfrac{1}{2}\eta^{\prime}
=12η1(η2(η+2eiϕ)2e3iϕA)±12η\displaystyle=\tfrac{1}{2}\eta^{-1}(\eta^{2}(\eta+2e^{i\phi})-2e^{3i\phi}A)\pm\tfrac{1}{2}\eta^{\prime}
=η(±12ηη1+12η+eiϕ)χ±2(η+2eiϕ)χ±\displaystyle=\eta(\pm\tfrac{1}{2}\eta^{\prime}\eta^{-1}+\tfrac{1}{2}\eta+e^{i\phi})-\chi_{\pm}^{2}-(\eta+2e^{i\phi})\chi_{\pm}
=2ηχ±χ±(χ±+2eiϕ),\displaystyle=-2\eta\chi_{\pm}-\chi_{\pm}(\chi_{\pm}+2e^{i\phi}),

and hence η=±12χ±χ±112χ±eiϕ\eta=\pm\tfrac{1}{2}\chi^{\prime}_{\pm}\chi_{\pm}^{-1}-\tfrac{1}{2}\chi_{\pm}-e^{i\phi}. Then by (7.6)

(χ±)2χ±1=\displaystyle(\chi^{\prime}_{\pm})^{2}\chi_{\pm}^{-1}= χ±(2η+(χ±+2eiϕ))2\displaystyle\chi_{\pm}(2\eta+(\chi_{\pm}+2e^{i\phi}))^{2}
=\displaystyle= χ±(χ±+2eiϕ)2+4(η2χ±+ηχ±(χ±+2eiϕ))\displaystyle\chi_{\pm}(\chi_{\pm}+2e^{i\phi})^{2}+4(\eta^{2}\chi_{\pm}+\eta\chi_{\pm}(\chi_{\pm}+2e^{i\phi}))
=\displaystyle= χ±(χ±+2eiϕ)2+4e3iϕA,\displaystyle\chi_{\pm}(\chi_{\pm}+2e^{i\phi})^{2}+4e^{3i\phi}A,

which implies χ±\chi_{\pm} solves (7.5). ∎

Recall that P(u;A)\mathrm{P}(u;A) is a solution of (7.5) such that P(0;A)=0.\mathrm{P}(0;A)=0. Using Proposition 7.5 we have the following.

Proposition 7.6.

The elliptic function P(u;A)\mathrm{P}(u;A) has simple poles with residue 1-1 at u=13Ω𝐛+Ω𝐚+Ω𝐛u=\tfrac{1}{3}\Omega_{\mathbf{b}}+\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z} and simple poles with residue 11 at u=13Ω𝐛+Ω𝐚+Ω𝐛u=-\tfrac{1}{3}\Omega_{\mathbf{b}}+\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}.

𝐥1\mathbf{l}_{-1}𝐥1\mathbf{l}_{1}𝐥2\mathbf{l}_{-2}𝐥2\mathbf{l}_{2}0z1z_{1}z3z_{3}z5z_{5}+\infty^{+}\infty^{-}Π+\Pi_{+}Π\Pi_{-}
Figure 7.1. Cycle 𝐛=𝐥1𝐥2𝐥2𝐥1\mathbf{b}=\mathbf{l}_{-1}\cup\mathbf{l}_{-2}\cup\mathbf{l}_{2}\cup\mathbf{l}_{1}
Proof..

Consider the loop on ΠA,ϕ=Π+Π\Pi_{A,\phi}=\Pi_{+}\cup\Pi_{-} defined by 𝐥1𝐥2𝐥2𝐥1\mathbf{l}_{-1}\cup\mathbf{l}_{-2}\cup\mathbf{l}_{2}\cup\mathbf{l}_{1} homotopic to the cycle 𝐛\mathbf{b}, in which 𝐥1=[0,],\mathbf{l}_{-1}=[0,\infty^{-}], 𝐥2=[,z3]Π\mathbf{l}_{-2}=[\infty^{-},z_{3}]\subset\Pi_{-}, and 𝐥2=[z3,+],\mathbf{l}_{2}=[z_{3},\infty^{+}], 𝐥1=[+,0]Π+\mathbf{l}_{1}=[\infty^{+},0]\subset\Pi_{+} as in Figure 7.1. Let

γ:=𝐥1dzw(A,z),γ+:=Ω𝐛γ=𝐥1𝐥2𝐥2dzw(A,z)=(𝐛𝐥1)dzw(A,z),\gamma_{-}:=\int_{\mathbf{l}_{-1}}\frac{dz}{w(A,z)},\quad\gamma_{+}:=\Omega_{\mathbf{b}}-\gamma_{-}=\int_{\mathbf{l}_{-1}\cup\mathbf{l}_{-2}\cup\mathbf{l}_{2}}\frac{dz}{w(A,z)}=\Bigl{(}\int_{\mathbf{b}}-\int_{\mathbf{l}_{1}}\Bigr{)}\frac{dz}{w(A,z)},

with 𝐥1w(A,z)1𝑑z=𝐥1w(A,z)1𝑑z.\int_{\mathbf{l}_{1}}w(A,z)^{-1}dz=\int_{\mathbf{l}_{-1}}w(A,z)^{-1}dz. Then P(u)=P(u;A)\mathrm{P}(u)=\mathrm{P}(u;A) has simple poles at u=u±γ±modΩ𝐚+Ω𝐛u=u_{\pm}\equiv\gamma_{\pm}\mod\Omega_{\mathbf{a}}\mathbb{Z}+\Omega_{\mathbf{b}}\mathbb{Z}, and double zeros at u=u00,u=u_{0}\equiv 0, and around these poles and zeros, P(u)=±(uu±)1eiϕ+o(1)\mathrm{P}(u)=\pm(u-u_{\pm})^{-1}-e^{i\phi}+o(1) and P(u)=e3iϕA(uu0)2(1+o(1)).\mathrm{P}(u)=e^{3i\phi}A(u-u_{0})^{2}(1+o(1)). Let

P(u)=12P(u)P(u)112P(u)eiϕ.\mathrm{P}_{-}(u)=\tfrac{1}{2}\mathrm{P}^{\prime}(u)\mathrm{P}(u)^{-1}-\tfrac{1}{2}\mathrm{P}(u)-e^{i\phi}.

It is easy to see that P(u)\mathrm{P}_{-}(u) has simple poles with residue 11 at u=u00u=u_{0}\equiv 0, with residue 1-1 at u=u+γ+u=u_{+}\equiv\gamma_{+}, and at least vanishes at u=uγu=u_{-}\equiv\gamma_{-}. By Proposition 7.5, P(u)\mathrm{P}_{-}(u) also solves (7.5), and hence P(u+γ)=P(u)\mathrm{P}_{-}(u+\gamma_{-})=\mathrm{P}(u). Then the string of poles and zeros of P(u)\mathrm{P}(u) given by 𝒮:=Ω𝐛(γ+Ω𝐛)(γ++Ω𝐛)\mathcal{S}:=\Omega_{\mathbf{b}}\mathbb{Z}\cup(\gamma_{-}+\Omega_{\mathbf{b}}\mathbb{Z})\cup(\gamma_{+}+\Omega_{\mathbf{b}}\mathbb{Z}) coincides with the shifted one 𝒮+γ\mathcal{S}+\gamma_{-}. From this fact we derive 2γ=γ+=Ω𝐛γ,2\gamma_{-}=\gamma_{+}=\Omega_{\mathbf{b}}-\gamma_{-}, which implies γ=13Ω𝐛.\gamma_{-}=\tfrac{1}{3}\Omega_{\mathbf{b}}. Thus the proposition is obtained. ∎

8. Boutroux equations

We show basic facts on the Boutroux equations used in the preceding sections.

8.1. Basic facts

The definitions of the cycles 𝐚,\mathbf{a}, 𝐛\mathbf{b}, 𝐚,\mathbf{a}_{*}, 𝐛\mathbf{b}_{*} are based on the following stable configuration of zeros.

Lemma 8.1.

As long as A𝒟0:=({c0}{c827})A\in\mathcal{D}_{0}:=\mathbb{C}\setminus(\{c\leq 0\}\cup\{c\geq\frac{8}{27}\}), the polynomial ζ3+4ζ2+4ζ+4A\zeta^{3}+4\zeta^{2}+4\zeta+4A has zeros ζ1,\zeta_{1}, ζ3\zeta_{3}, ζ5\zeta_{5} with the properties::

(1)(1) ζj=ζj(A)\zeta_{j}=\zeta_{j}(A) (j=1,3,5)(j=1,3,5) are continuous in A;A;

(2)(2) Reζ5<Reζ3<Reζ1;\mathrm{Re\,}\zeta_{5}<\mathrm{Re\,}\zeta_{3}<\mathrm{Re\,}\zeta_{1};

(3)(3) (ζ5,ζ3,ζ1)(2,2,0)(\zeta_{5},\zeta_{3},\zeta_{1})\to(-2,-2,0) as A0A\to 0, and (ζ5,ζ3,ζ1)(83,23,23)(\zeta_{5},\zeta_{3},\zeta_{1})\to(-\frac{8}{3},-\frac{2}{3},-\frac{2}{3}) as A827.A\to\frac{8}{27}.

Proof..

Suppose that A𝒟0A\in\mathcal{D}_{0}. The substitution ζ=43(ξ1)\zeta=\frac{4}{3}(\xi-1) leads to the polynomial

ϖ(ξ)=4ξ33ξc0,c0=1274A𝒟1=({c1}{c1}),\varpi(\xi)=4\xi^{3}-3\xi-c_{0},\quad c_{0}=1-\tfrac{27}{4}A\in\mathcal{D}_{1}=\mathbb{C}\setminus(\{c\leq-1\}\cup\{c\geq 1\}),

which has multiple zeros if and only if c0=±1.c_{0}=\pm 1. The zeros of ϖ(ξ)\varpi(\xi) may be numbered in such a way that 1<ξ5<ξ3<ξ1<1-1<\xi_{5}<\xi_{3}<\xi_{1}<1 for 1<c0<1-1<c_{0}<1, and clearly ξj\xi_{j} (j=1,3,5)(j=1,3,5) move continuously in c0𝒟1.c_{0}\in\mathcal{D}_{1}. Then to verify (2) for ζj=43(ξj1)\zeta_{j}=\frac{4}{3}(\xi_{j}-1) (j=1,3,5)(j=1,3,5), it is sufficient to show that ξjξki\xi_{j}-\xi_{k}\not\in i\mathbb{R} as long as c0𝒟1c_{0}\in\mathcal{D}_{1} if jkj\not=k. By Cardano’s formula ξj,\xi_{j}, ξk{γ0,γ±1},\xi_{k}\in\{\gamma_{0},\gamma_{\pm 1}\}, in which 2γ0=u1/3+u1/3,2\gamma_{0}=u^{1/3}+u^{-1/3}, 2γ±1=ω±1u1/3+ω1u1/32\gamma_{\pm 1}=\omega^{\pm 1}u^{1/3}+\omega^{\mp 1}u^{-1/3} with ω=e2πi/3\omega=e^{2\pi i/3} and u=c0+c021u=c_{0}+\sqrt{c_{0}^{2}-1}. Suppose that 2(γ0γ1)=(1ω)u1/3(1ω1)u1/3=ir2(\gamma_{0}-\gamma_{1})=(1-\omega)u^{1/3}-(1-\omega^{-1})u^{-1/3}=ir with r.r\in\mathbb{R}. Then v=(1ω)u1/3v=(1-\omega)u^{1/3} fulfils v2irv+3=0v^{2}-irv+3=0, which implies (1ω)u1/3i(1-\omega)u^{1/3}\in i\mathbb{R} with 1ω=i3ω2.1-\omega=i\sqrt{3}\omega^{2}. Hence u=c0+c021=r0u=c_{0}+\sqrt{c_{0}^{2}-1}=r_{0}\in\mathbb{R}, that is, c0=(r0+r01)/2{c1}{c1}c_{0}=(r_{0}+r_{0}^{-1})/2\in\{c\leq-1\}\cup\{c\geq 1\}, contradicting c0𝒟1.c_{0}\in\mathcal{D}_{1}. Supposition γ1γ1i\gamma_{1}-\gamma_{-1}\in i\mathbb{R} implies u1/3u^{1/3}\in\mathbb{R} leading to the same contradiction. Thus the property (2) is verified. It is easy to see that (1) and (3) are fulfilled. ∎

The following corollary is obtained by setting ζ=eiϕz.\zeta=e^{-i\phi}z.

Corollary 8.2.

As long as (ϕ,A)×𝒟0(\phi,A)\in\mathbb{R}\times\mathcal{D}_{0}, the polynomial z3+4eiϕz2+4e2iϕz+4e3iϕAz^{3}+4e^{i\phi}z^{2}+4e^{2i\phi}z+4e^{3i\phi}A has zeros z1,z_{1}, z3z_{3}, z5z_{5} with the properties::

(1)(1) zj=zj(ϕ,A)z_{j}=z_{j}(\phi,A) (j=1,3,5)(j=1,3,5) are continuous in (ϕ,A);(\phi,A);

(2)(2) Reeiϕz5<Reeiϕz3<Reeiϕz1;\mathrm{Re\,}e^{-i\phi}z_{5}<\mathrm{Re\,}e^{-i\phi}z_{3}<\mathrm{Re\,}e^{-i\phi}z_{1};

(3)(3) (z5,z3,z1)(2eiϕ,2eiϕ,0)(z_{5},z_{3},z_{1})\to(-2e^{i\phi},-2e^{i\phi},0) as A0A\to 0, and (z5,z3,z1)(83eiϕ,23eiϕ,23eiϕ)(z_{5},z_{3},z_{1})\to(-\frac{8}{3}e^{i\phi},-\frac{2}{3}e^{i\phi},-\frac{2}{3}e^{i\phi}) as A827.A\to\frac{8}{27}.

Let us define the elliptic curve ΠA=Π+Π\Pi_{A}^{*}=\Pi_{+}^{*}\cup\Pi_{-}^{*} and the cycles 𝐚\mathbf{a}_{*} and 𝐛\mathbf{b}_{*} on ΠA\Pi^{*}_{A}. As long as A𝒟0A\in\mathcal{D}_{0}, by Lemma 8.1, the elliptic curve

v(A,ζ)2=ζ4+4ζ3+4ζ2+4Aζ=ζ(ζζ1)(ζζ3)(ζζ5),v(A,\zeta)^{2}=\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A\zeta=\zeta(\zeta-\zeta_{1})(\zeta-\zeta_{3})(\zeta-\zeta_{5}),

is the two sheeted Riemann surface ΠA=Π+Π\Pi^{*}_{A}=\Pi^{*}_{+}\cup\Pi^{*}_{-} glued along the cuts [ζ5,ζ3],[\zeta_{5},\zeta_{3}], [ζ1,0][\zeta_{1},0], where Π±\Pi^{*}_{\pm} are two copies of P1()([ζ5,ζ3][ζ1,0])P^{1}(\mathbb{C})\setminus([\zeta_{5},\zeta_{3}]\cup[\zeta_{1},0]). Then on ΠA\Pi^{*}_{A} the cycles 𝐚\mathbf{a}_{*} and 𝐛\mathbf{b}_{*} are defined as in Figure 8.1. Note that ΠA\Pi^{*}_{A} and (𝐚,𝐛)(\mathbf{a}_{*},\mathbf{b}_{*}) are parametrised by A𝒟0A\in\mathcal{D}_{0}. Let the elliptic curve ΠA,ϕ=Π+Π\Pi_{A,\phi}=\Pi_{+}\cup\Pi_{-} and the pair of cycles (𝐚,𝐛)(\mathbf{a},\mathbf{b}) be the images of ΠA\Pi^{*}_{A} and (𝐚,𝐛)(\mathbf{a}_{*},\mathbf{b}_{*}) under the map z=eiϕζz=e^{i\phi}\zeta. Then ΠA,ϕ\Pi_{A,\phi} is given by w(A,z)2=z4+4eiϕz3+4e2iϕz2+4e3iϕAzw(A,z)^{2}=z^{4}+4e^{i\phi}z^{3}+4e^{2i\phi}z^{2}+4e^{3i\phi}Az as in Section 2, and the cycles 𝐚\mathbf{a} and 𝐛\mathbf{b} are as in Figure 2.1 for 0<|ϕ|<π/40<|\phi|<\pi/4.

ζ5\zeta_{5}ζ3\zeta_{3}ζ1\zeta_{1}0𝐚\mathbf{a}_{*}𝐛\mathbf{b}_{*}Π+\Pi_{+}^{*}
Figure 8.1. Cycles 𝐚,\mathbf{a}_{*}, 𝐛\mathbf{b}_{*} on ΠA=Π+Π\Pi^{*}_{A}=\Pi^{*}_{+}\cup\Pi^{*}_{-}

8.2. Uniqueness

Let us treat the integrals

J𝐚(A)=𝐚v(A,ζ)ζ𝑑ζ,J𝐛(A)=𝐛v(A,ζ)ζ𝑑ζ,\displaystyle J_{\mathbf{a}_{*}}(A)=\int_{\mathbf{a}_{*}}\frac{v(A,\zeta)}{\zeta}d\zeta,\qquad J_{\mathbf{b}_{*}}(A)=\int_{\mathbf{b}_{*}}\frac{v(A,\zeta)}{\zeta}d\zeta,
v(A,ζ)=ζ4+4ζ3+4ζ2+4Aζ=e2iϕw(A,z).\displaystyle v(A,\zeta)=\sqrt{\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A\zeta}=e^{-2i\phi}w(A,z).

Then the Boutroux equations (2.2) become (2.3), that is,

(BE)ϕ Ree2iϕ𝐚v(Aϕ,ζ)ζ𝑑ζ=Ree2iϕ𝐛v(Aϕ,ζ)ζ𝑑ζ=0.\mathrm{Re\,}e^{2i\phi}\int_{\mathbf{a}_{*}}\frac{v(A_{\phi},\zeta)}{\zeta}d\zeta=\mathrm{Re\,}e^{2i\phi}\int_{\mathbf{b}_{*}}\frac{v(A_{\phi},\zeta)}{\zeta}d\zeta=0.
Example 8.1.

Let ϕ=0,\phi=0, A=827.A=\tfrac{8}{27}. Then ζ1=ζ3=23,\zeta_{1}=\zeta_{3}=-\tfrac{2}{3}, ζ5=83,\zeta_{5}=-\tfrac{8}{3}, and hence

J𝐚(827)=28/32/3(ζ+23)ζ(ζ+83)dζζ=2eiπ/28/32/3(23t)83tt𝑑t=4i3J_{\mathbf{a}_{*}}(\tfrac{8}{27})=2\int^{-2/3}_{-8/3}(\zeta+\tfrac{2}{3})\sqrt{\zeta(\zeta+\smash{\tfrac{8}{3})}}\frac{d\zeta}{\zeta}=2e^{i\pi/2}\int^{2/3}_{8/3}(\tfrac{2}{3}-t)\sqrt{\frac{\smash{\tfrac{8}{3}}-t}{t}}dt=\frac{4i}{\sqrt{3}}

and J𝐛(827)=0.J_{\mathbf{b}_{*}}(\tfrac{8}{27})=0. This implies A0=827A_{0}=\tfrac{8}{27} is a solution of the Boutroux equations (BE)ϕ=0.

In accordance with Kitaev [20, Section 7] we would like to show the uniqueness of a solution of (BE)ϕ=0. To do so we begin with the following.

Proposition 8.3.

If ReJ𝐚(A)=ReJ𝐛(A)=0,\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)=\mathrm{Re\,}J_{\mathbf{b}_{*}}(A)=0, then A.A\in\mathbb{R}.

Proof..

The condition ReJ𝐚(A)=0\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)=0 implies

(8.1) 0=\displaystyle 0= J𝐚(A)+J𝐚(A)¯=J𝐚(A)+J𝐚¯(A¯)=J𝐚(A)J𝐚(A¯)=4(AA¯)I0,\displaystyle J_{\mathbf{a}_{*}}(A)+\overline{J_{\mathbf{a}_{*}}(A)}=J_{\mathbf{a}_{*}}(A)+J_{\overline{\mathbf{a}_{*}}}(\overline{A})=J_{\mathbf{a}_{*}}(A)-J_{\mathbf{a}_{*}}(\overline{A})=4(A-\overline{A})I_{0},
I0:=\displaystyle I_{0}:= 𝐚dζv(A,ζ)+v(A¯,ζ).\displaystyle\int_{\mathbf{a}_{*}}\frac{d\zeta}{v(A,\zeta)+v(\overline{A},\zeta)}.

Let us derive AA\in\mathbb{R} by supposing the contrary AA¯0.A-\overline{A}\not=0. Then, by Lemma 8.1, it may be supposed that Reζ5<Reζ3<Reζ1\mathrm{Re\,}\zeta_{5}<\mathrm{Re\,}\zeta_{3}<\mathrm{Re\,}\zeta_{1}, which is divided into two cases according to the location of 0.0.

(i) Case Reζ30\mathrm{Re\,}\zeta_{3}\leq 0. It is sufficient to show that I00I_{0}\not=0 in (8.1). The algebraic function v(A,ζ)+v(A¯,ζ)v(A,\zeta)+v(\overline{A},\zeta) is considered on the two sheeted Riemann surface ΠA0\Pi_{A}^{*0} glued along the cuts [ζ1,0][\zeta_{1},0], [ζ1¯,0][\overline{\zeta_{1}},0], [ζ5,ζ3][\zeta_{5},\zeta_{3}], [ζ5¯,ζ3¯],[\overline{\zeta_{5}},\overline{\zeta_{3}}], which is constructed by adding to ΠA\Pi^{*}_{A} the new two cuts [ζ1¯,0][\overline{\zeta_{1}},0], [ζ5¯,ζ3¯][\overline{\zeta_{5}},\overline{\zeta_{3}}] and gluing along them. Choose the cycle 𝐚\mathbf{a}_{*} on ΠA0\Pi_{A}^{*0} in such a way that 𝐚\mathbf{a}_{*} surrounds the cuts [ζ5,ζ3][\zeta_{5},\zeta_{3}] and [ζ5¯,ζ3¯][\overline{\zeta_{5}},\overline{\zeta_{3}}] as in Figure 8.2, (a), and modify 𝐚\mathbf{a}_{*} as in Figure 8.2, (b), where ζ5=a+ib\zeta_{5}=a+ib, ζ3=c+ib\zeta_{3}=c+ib^{\prime}, a<c0a<c\leq 0, b,b.b,b^{\prime}\in\mathbb{R}.

ζ5\zeta_{5}ζ5¯\overline{\zeta_{5}}ζ3\zeta_{3}ζ3¯\overline{\zeta_{3}}ζ1\zeta_{1}ζ1¯\overline{\zeta_{1}}0𝐚\mathbf{a}_{*}(a) Cycle 𝐚\mathbf{a}_{*} on ΠA0\Pi_{A}^{*0}
ζ5\zeta_{5}ζ5¯\overline{\zeta_{5}}0ζ3\zeta_{3}ζ3¯\overline{\zeta_{3}}ζ1{\zeta_{1}}ζ1¯\overline{\zeta_{1}}a+iba+ibc+ibc+ib^{\prime}𝐚\mathbf{a}_{*}(b) Modification of 𝐚\mathbf{a}_{*}
Figure 8.2. Cycle 𝐚\mathbf{a}_{*} and modification in the case Reζ30\mathrm{Re\,}\zeta_{3}\leq 0

To simplify the description we write vA(ζ)=v(A,ζ),v_{A}(\zeta)=v(A,\zeta), vA¯(ζ)=v(A¯,ζ)v_{\overline{A}}(\zeta)=v(\overline{A},\zeta) and (vA±vA¯)(ζ)=v(A,ζ)±v(A¯,ζ)(v_{A}\pm v_{\overline{A}})(\zeta)=v(A,\zeta)\pm v(\overline{A},\zeta). Let us take the contour of I0I_{0} to be the modified 𝐚\mathbf{a}_{*} starting from and terminating in the point ζ=Reζ5+=a+\zeta=\mathrm{Re\,}\zeta_{5}+=a+ on the upper shore of the cut [a,c][a,c]. Then I0I_{0} is decomposed into the left- and right-vertical, and the horizontal parts:

I0=\displaystyle I_{0}= Ihor+Iright-v+Ileft-v,\displaystyle I_{\mathrm{hor}}+I_{\mathrm{right\text{-}v}}+I_{\mathrm{left\text{-}v}},
Ihor=\displaystyle I_{\mathrm{hor}}= 2acds(vA+vA¯)(s),\displaystyle 2\int^{c}_{a}\frac{ds}{(v_{A}+v_{\overline{A}})(s)},
Iright-v=\displaystyle I_{\mathrm{right\text{-}v}}= 0bidt(vA+vA¯)(c+it)+bbidt(vA+vA¯)(c+it)+b0idt(vAvA¯)(c+it),\displaystyle\int_{0}^{b^{\prime}}\frac{idt}{(v_{A}+v_{\overline{A}})(c+it)}+\int_{b^{\prime}}^{-b^{\prime}}\frac{idt}{(-v_{A}+v_{\overline{A}})(c+it)}+\int^{0}_{-b^{\prime}}\frac{idt}{(-v_{A}-v_{\overline{A}})(c+it)},
Ileft-v=\displaystyle-I_{\mathrm{left\text{-}v}}= 0bidt(vA+vA¯)(a+it)+bbidt(vAvA¯)(a+it)+b0idt(vAvA¯)(a+it).\displaystyle\int_{0}^{-b}\frac{idt}{(v_{A}+v_{\overline{A}})(a+it)}+\int_{-b}^{b}\frac{idt}{(v_{A}-v_{\overline{A}})(a+it)}+\int_{b}^{0}\frac{idt}{(-v_{A}-v_{\overline{A}})(a+it)}.

Here

Iright-v=\displaystyle I_{\mathrm{right\text{-}v}}= i4(AA¯)(0b(vAvA¯)(c+it)c+it𝑑t+0b(vA+vA¯)(c+it)c+it𝑑t)\displaystyle\frac{i}{4(A-\overline{A})}\biggl{(}\int_{0}^{b^{\prime}}\frac{(v_{A}-v_{\overline{A}})(c+it)}{c+it}dt+\int_{0}^{b^{\prime}}\frac{(v_{A}+v_{\overline{A}})(c+it)}{c+it}dt\biggr{)}
i4(AA¯)(0b(vA+vA¯)(c+it)c+it𝑑t0b(vAvA¯)(c+it)c+it𝑑t)\displaystyle-\frac{i}{4(A-\overline{A})}\biggl{(}\int_{0}^{-b^{\prime}}\frac{(v_{A}+v_{\overline{A}})(c+it)}{c+it}dt-\int_{0}^{-b^{\prime}}\frac{(v_{A}-v_{\overline{A}})(c+it)}{c+it}dt\biggr{)}
=\displaystyle= i2(AA¯)(0bvA(c+it)c+it𝑑t+0bvA¯(cit)cit𝑑t),\displaystyle\frac{i}{2(A-\overline{A})}\biggl{(}\int_{0}^{b^{\prime}}\frac{v_{A}(c+it)}{c+it}dt+\int_{0}^{b^{\prime}}\frac{v_{\overline{A}}(c-it)}{c-it}dt\biggr{)}\in\mathbb{R},

and similarly Ileft-v.I_{\mathrm{left\text{-}v}}\in\mathbb{R}. Let A=κ+iκA=\kappa+i\kappa^{\prime} with κ0.\kappa^{\prime}\not=0. If, say Imζ5>0\mathrm{Im\,}\zeta_{5}>0, along the upper shore of the cut [a,c][a,c], vA(s)=eπi/2φ(s)4iκsv_{A}(s)=e^{-\pi i/2}\sqrt{-\varphi(s)-4i\kappa^{\prime}s}, vA¯(s)=eπi/2φ(s)+4iκsv_{\overline{A}}(s)=-e^{\pi i/2}\sqrt{-\varphi(s)+4i\kappa^{\prime}s}, where φ(s)=s4+4s3+4s2+4κs<0\varphi(s)=s^{4}+4s^{3}+4s^{2}+4\kappa s<0, i.e. φ(s)>0-\varphi(s)>0 for a<s<c0.a<s<c\leq 0. Then the horizontal integral along the upper shore is

acds(vA+vA¯)(s)=i2acdsφ(s)+φ(s)2+16(κ)2s2=iγ0\int^{c}_{a}\frac{ds}{(v_{A}+v_{\overline{A}})(s)}=\frac{i}{\sqrt{2}}\int^{c}_{a}\frac{ds}{\sqrt{-\varphi(s)+\sqrt{\varphi(s)^{2}+16(\kappa^{\prime})^{2}s^{2}}}}=i\gamma_{0}

with some γ0>0\gamma_{0}>0. Thus we have Ihor,I_{\mathrm{hor}}\in\mathbb{C}\setminus\mathbb{R}, and hence I00,I_{0}\not=0, which contradicts (8.1). This implies AA\in\mathbb{R} in this case.

(ii) Case Reζ3>0\mathrm{Re\,}\zeta_{3}>0. Let ΠA=Π+Π\Pi^{**}_{A}=\Pi^{**}_{+}\cup\Pi^{**}_{-} be a two sheeted Riemann surface glued along the cuts [ζ3,ζ1][\zeta_{3},\zeta_{1}], [ζ5,0][\zeta_{5},0] with Π±=P1()([ζ3,ζ1][ζ5,0])\Pi^{**}_{\pm}=P^{1}(\mathbb{C})\setminus([\zeta_{3},\zeta_{1}]\cup[\zeta_{5},0]). Instead of ΠA\Pi^{*}_{A} and 𝐚\mathbf{a}_{*} of the case (i) we treat ΠA\Pi^{**}_{A} and a cycle on it. Draw a closed curve 𝐛0\mathbf{b}_{0} on the upper sheet Π+ΠA\Pi^{**}_{+}\subset\Pi^{**}_{A} surrounding the cut [ζ3,ζ1][\zeta_{3},\zeta_{1}] clockwise as in Figure 8.3.

ζ5\zeta_{5}ζ3\zeta_{3}ζ1\zeta_{1}0𝐛0\mathbf{b}_{0}ΠA\Pi_{A}^{**}
Figure 8.3. Cycle 𝐛0\mathbf{b}_{0} on ΠA\Pi_{A}^{**} in the case Reζ3>0\mathrm{Re\,}\zeta_{3}>0

Consider the algebraic function v(A,ζ)=ζ4+4ζ3+4ζ2+4Aζv^{*}(A,\zeta)=\sqrt{\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A\zeta} on ΠA\Pi^{**}_{A}, where the branch is chosen in such a way that v(A,ζ)v^{*}(A,\zeta) coincides with v(A,ζ)v(A,\zeta) along the upper shore of the cut [ζ3,ζ1][\zeta_{3},\zeta_{1}]. The cycle 𝐛0\mathbf{b}_{0} on ΠA\Pi^{**}_{A} is a substitute for 𝐛\mathbf{b}_{*} on ΠA.\Pi_{A}^{*}. Indeed

J𝐛0(A):=𝐛0v(A,ζ)ζ𝑑ζ=2ζ3ζ1v(A,ζ)ζ𝑑ζ=J𝐛(A).J_{\mathbf{b}_{0}}^{*}(A):=\int_{\mathbf{b}_{0}}\frac{v^{*}(A,\zeta)}{\zeta}d\zeta=2\int^{\zeta_{1}}_{\zeta_{3}}\frac{v(A,\zeta)}{\zeta}d\zeta=J_{\mathbf{b}_{*}}(A).

Hence ReJ𝐛(A)=0\mathrm{Re\,}J_{\mathbf{b}_{*}}(A)=0 is equivalent to ReJ𝐛0(A)=0.\mathrm{Re\,}J^{*}_{\mathbf{b}_{0}}(A)=0. Under this supposition,

0=\displaystyle 0= J𝐛0(A)+J𝐛0(A)¯=J𝐛0(A)+J𝐛0¯(A¯)=J𝐛0(A)J𝐛0(A¯)=4(AA¯)I0,\displaystyle J^{*}_{\mathbf{b}_{0}}(A)+\overline{J^{*}_{\mathbf{b}_{0}}(A)}=J^{*}_{\mathbf{b}_{0}}(A)+J^{*}_{\overline{\mathbf{b}_{0}}}(\overline{A})=J^{*}_{\mathbf{b}_{0}}(A)-J^{*}_{\mathbf{b}_{0}}(\overline{A})=4(A-\overline{A})I_{0}^{*},
I0:=\displaystyle I_{0}^{*}:= 𝐛0dζv(A,ζ)+v(A¯,ζ)\displaystyle\int_{\mathbf{b}_{0}}\frac{d\zeta}{v^{*}(A,\zeta)+v^{*}(\overline{A},\zeta)}

instead of (8.1). As in the case (i), to prove AA\in\mathbb{R} it is sufficient to show I00.I_{0}^{*}\not=0. The algebraic function v(A,ζ)+v(A¯,ζ)v^{*}(A,\zeta)+v^{*}(\overline{A},\zeta) is considered on the two sheeted Riemann surface ΠA0\Pi_{A}^{**0} glued along the cuts [ζ3,ζ1][\zeta_{3},\zeta_{1}], [ζ3¯,ζ1¯][\overline{\zeta_{3}},\overline{\zeta_{1}}], [ζ5,0][\zeta_{5},0], [ζ5¯,0][\overline{\zeta_{5}},0], and the cycle 𝐛0\mathbf{b}_{0} may be taken in such a way that 𝐛0\mathbf{b}_{0} surrounds the cuts [ζ3,ζ1][\zeta_{3},\zeta_{1}] and [ζ3¯,ζ1¯][\overline{\zeta_{3}},\overline{\zeta_{1}}] as in Figure 8.4, (a).

ζ5\zeta_{5}ζ5¯\overline{\zeta_{5}}ζ3\zeta_{3}ζ3¯\overline{\zeta_{3}}ζ1\zeta_{1}ζ1¯\overline{\zeta_{1}}0𝐛0\mathbf{b}_{0}ΠA0\Pi_{A}^{**0}(a) Cycle 𝐛0\mathbf{b}_{0} on ΠA0\Pi^{**0}_{A}
0ζ5\zeta_{5}ζ5¯\overline{\zeta_{5}}ζ3\zeta_{3}ζ3¯\overline{\zeta_{3}}ζ1\zeta_{1}ζ1¯\overline{\zeta_{1}}ΠA0\Pi_{A}^{**0}(b) Modification of 𝐛0\mathbf{b}_{0}
Figure 8.4. Modification of 𝐛0\mathbf{b}_{0} in the case Reζ3>0\mathrm{Re}\,\zeta_{3}>0

The cycle 𝐛0\mathbf{b}_{0} is modified as in Figure 8.4, (b), in which ζ1=a+ib,\zeta_{1}=a+ib, ζ3=c+ib\zeta_{3}=c+ib^{\prime}, a>c>0,a>c>0, b,b.b,b^{\prime}\in\mathbb{R}. The contour of the integral I0I^{*}_{0} is taken to be the modified 𝐛0\mathbf{b}_{0} that starts from and terminates in the point ζ=Reζ3+=c+\zeta=\mathrm{Re\,}\zeta_{3}+=c+ on the upper shore of the cut [c,a].[c,a]. Dividing I0I^{*}_{0} into the horizontal, and the left- and right-vertical parts, we may show I00I^{*}_{0}\not=0 by the same arguments as in the case (i). Thus the proposition is proved. ∎

As in Example 8.1, ReJ𝐚(827)=ReJ𝐛(827)=0\mathrm{Re\,}J_{\mathbf{a}_{*}}(\tfrac{8}{27})=\mathrm{Re\,}J_{\mathbf{b}_{*}}(\tfrac{8}{27})=0. Let us examine ReJ𝐚(A)\mathrm{Re\,}J_{\mathbf{a}_{*}}(A) or ReJ𝐛(A)\mathrm{Re\,}J_{\mathbf{b}_{*}}(A) for A{827}.A\in\mathbb{R}\setminus\{\tfrac{8}{27}\}.

Consider the polynomial f(ζ)=v(A,ζ)2=ζ4+4ζ3+4ζ2+4Aζf(\zeta)=v(A,\zeta)^{2}=\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A\zeta for AA\in\mathbb{R} on the real line (,+)Π+.(-\infty,+\infty)\subset\Pi_{+}^{*}. The zeros of f(ζ)f(\zeta) are located as follows, in which (z.3) is given by Lemma 8.1 and (z.4) is treated in Example 8.1:

(z.1) if A<0A<0, then 0<ζ1,0<\zeta_{1}, and ζ5,ζ3;\zeta_{5},\zeta_{3}\not\in\mathbb{R};

(z.2) if A=0A=0, then ζ5=ζ3<ζ1=0;\zeta_{5}=\zeta_{3}<\zeta_{1}=0;

(z.3) if 0<A<8270<A<\tfrac{8}{27}, then ζ5<ζ3<ζ1<0;\zeta_{5}<\zeta_{3}<\zeta_{1}<0;

(z.4) if A=827A=\tfrac{8}{27}, then ζ5<ζ3=ζ1<0;\zeta_{5}<\zeta_{3}=\zeta_{1}<0;

(z.5) if A>827A>\tfrac{8}{27}, then ζ5<0\zeta_{5}<0, and ζ3,ζ1.\zeta_{3},\zeta_{1}\not\in\mathbb{R}.

(z.5a)827<A<4\tfrac{8}{27}<A<4:0ccζ5\zeta_{5}ζ3\zeta_{3}ζ1\zeta_{1}(z.5b)A4A\geq 4:0ζ3\zeta_{3}ζ1\zeta_{1}ccζ5\zeta_{5}(z.1)A<0A<0:0ζ1\zeta_{1}ζ3\zeta_{3}ζ5\zeta_{5}(z.2)A=0A=0:ζ1,0\zeta_{1,0}ζ5,3\zeta_{5,3}(z.3)0<A<8270<A<\tfrac{8}{27}:0ζ1\zeta_{1}ζ3\zeta_{3}ζ5\zeta_{5}(z.4)A=827A=\tfrac{8}{27}:0ζ3,1\zeta_{3,1}ζ5\zeta_{5}(z.5)A>827A>\tfrac{8}{27}:0ζ5\zeta_{5}ζ3\zeta_{3}ζ1\zeta_{1}::f(ζ)>0f(\zeta)>0;::f(ζ)<0f(\zeta)<0;:: cut, f(ζ)<0f(\zeta)<0
Figure 8.5. f(ζ)f(\zeta) on (,+)(-\infty,+\infty)

Let us show that ReJ𝐚(A)0\mathrm{Re\,}J_{\mathbf{a}^{*}}(A)\not=0 or ReJ𝐛(A)0\mathrm{Re\,}J_{\mathbf{b}^{*}}(A)\not=0 for A{827}.A\in\mathbb{R}\setminus\{\tfrac{8}{27}\}. In each case, f(ζ)f(\zeta) behaves as in Figure 8.5: say, in the case (z.3), f(ζ)>0f(\zeta)>0 on (,ζ5)(ζ3,ζ1)(0,+)(-\infty,\zeta_{5})\cup(\zeta_{3},\zeta_{1})\cup(0,+\infty), and f(ζ)<0f(\zeta)<0 on (ζ5,ζ3)(ζ1,0)(\zeta_{5},\zeta_{3})\cup(\zeta_{1},0). It is easy to see that, in (z.2) or (z.3),

J𝐛(A)=2ζ3ζ1v(A,ζ)ζ𝑑ζ=2ζ3ζ1f(ζ)ζ𝑑ζ<0,J_{\mathbf{b}_{*}}(A)=2\int_{\zeta_{3}}^{\zeta_{1}}\frac{v(A,\zeta)}{\zeta}d\zeta=2\int_{\zeta_{3}}^{\zeta_{1}}\frac{-\sqrt{f(\zeta)}}{\zeta}d\zeta<0,

and that, in (z.2), J𝐚(0)=0.J_{\mathbf{a}_{*}}(0)=0. In (z.1), we write ζ5=c+ib,\zeta_{5}=c+ib, ζ3=cib\zeta_{3}=c-ib with c<0,c<0, b>0b>0 such that 2c+ζ1=4.2c+\zeta_{1}=-4. Then

J𝐛(A)=2ζ3ζ1v(A,ζ)ζ𝑑ζ=2c0f(ζ)ζ𝑑ζ<0.J_{\mathbf{b}_{*}}(A)=2\int_{\zeta_{3}}^{\zeta_{1}}\frac{v(A,\zeta)}{\zeta}d\zeta=2\int_{c}^{0}\frac{-\sqrt{f(\zeta)}}{\zeta}d\zeta<0.

It remains to discuss the case (z.5). Set ζ1=cib\zeta_{1}=c-ib and ζ3=c+ib\zeta_{3}=c+ib with b>0.b>0. The relation 2c+ζ5=42c+\zeta_{5}=-4 yields A=2(c+2)(c+1)2A=2(c+2)(c+1)^{2} and b=(3c+2)(c+2).b=\sqrt{(3c+2)(c+2)}. Then, according to the location of cc we consider two cases: (z.5a) 23<c<0-\tfrac{2}{3}<c<0 if 827<A<4\tfrac{8}{27}<A<4; and (z.5b) c0c\geq 0 if A4A\geq 4. First consider the case (z.5a) as in Figure 8.5. Note that

A0ζ1v(A,ζ)ζ𝑑ζ=20ζ1dζv(A,ζ)+(ζ1)Av(A,ζ1)ζ1=20ζ1dζv(A,ζ),\frac{\partial}{\partial A}\int^{\zeta_{1}}_{0}\frac{v(A,\zeta)}{\zeta}d\zeta=2\int^{\zeta_{1}}_{0}\frac{d\zeta}{v(A,\zeta)}+(\zeta_{1})_{A}\frac{v(A,\zeta_{1})}{\zeta_{1}}=2\int^{\zeta_{1}}_{0}\frac{d\zeta}{v(A,\zeta)},

since v(A,ζ1)=0.v(A,\zeta_{1})=0. Then, by f(ζ)<0f(\zeta)<0 on (c,0)(c,0),

AReJ𝐚(A)=2ARe0ζ1v(A,ζ)ζ𝑑ζ=4Re0ζ1dζv(A,ζ)=4Recζ1dζv(A,ζ),\frac{\partial}{\partial A}\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)=2\frac{\partial}{\partial A}\mathrm{Re\,}\int^{\zeta_{1}}_{0}\frac{v(A,\zeta)}{\zeta}d\zeta=4\mathrm{Re\,}\int^{\zeta_{1}}_{0}\frac{d\zeta}{v(A,\zeta)}=4\mathrm{Re\,}\int^{\zeta_{1}}_{c}\frac{d\zeta}{v(A,\zeta)},

in which the contour is taken along the upper shore of the cut [0,ζ1].[0,\zeta_{1}]. Using

f(c+it)=g(t)+ih(t),\displaystyle f(c+it)=g(t)+ih(t),
g(t)=t42(3c2+6c+2)t2+f(c),\displaystyle g(t)=t^{4}-2(3c^{2}+6c+2)t^{2}+f(c),
h(t)=4(c+1)t3+4(c+1)(c+2)(3c+2)t\displaystyle h(t)=-4(c+1)t^{3}+4(c+1)(c+2)(3c+2)t

fulfilling g(t)<0g(t)<0 and h(t)<0h(t)<0 for (3c+2)(c+2)<t<0-\sqrt{(3c+2)(c+2)}<t<0, and v(A,c+it)=eπi/2p(t)+iq(t)v(A,c+it)=e^{\pi i/2}\sqrt{{p}(t)+i{q}(t)} with p(t)=g(t){p}(t)=-g(t), q(t)=h(t){q}(t)=-h(t), we have, for 827<A<4,\tfrac{8}{27}<A<4,

AReJ𝐚(A)=4Re0bdtp(t)+iq(t)=22b0p(t)2+q(t)2+p(t)p(t)2+q(t)2𝑑t<0.\frac{\partial}{\partial A}\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)=4\mathrm{Re\,}\int^{-b}_{0}\frac{dt}{\sqrt{{p}(t)+i{q}(t)}}=-2\sqrt{2}\int^{0}_{-b}\frac{\sqrt{\sqrt{{p}(t)^{2}+{q}(t)^{2}}+{p}(t)}}{\sqrt{{p}(t)^{2}+{q}(t)^{2}}}dt<0.

From this combined with ReJ𝐚(827)=0\mathrm{Re\,}J_{\mathbf{a}_{*}}(\tfrac{8}{27})=0, it follows that ReJ𝐚(A)<0\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)<0 in the case (z.5a).

In the case (z.5b) in Figure 8.5, f(ζ)>0f(\zeta)>0 on (0,c)(0,c), and hence

AReJ𝐚(A)=\displaystyle-\frac{\partial}{\partial A}\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)= 2ARe0ζ1v(A,ζ)ζ𝑑ζ=4Re0ζ1dζv(A,ζ)=4Recζ1dζv(A,ζ)+I(c),\displaystyle 2\frac{\partial}{\partial A}\mathrm{Re\,}\int^{\zeta_{1}}_{0}\frac{v(A,\zeta)}{\zeta}d\zeta=4\mathrm{Re\,}\int^{\zeta_{1}}_{0}\frac{d\zeta}{v(A,\zeta)}=4\mathrm{Re\,}\int^{\zeta_{1}}_{c}\frac{d\zeta}{v(A,\zeta)}+I(c),
I(c)=\displaystyle I(c)= 40cdζv(A,ζ)0,\displaystyle 4\int_{0}^{c}\frac{d\zeta}{v(A,\zeta)}\geq 0,

in which the contour is taken along the upper shore of the cut [ζ1,0][\zeta_{1},0]. Note that g(t)>0g(t)>0 for c(3c+4)<t<0-\sqrt{c(3c+4)}<t<0 and h(t)<0h(t)<0 for (3c+2)(c+2)<t<0-\sqrt{(3c+2)(c+2)}<t<0, and set q(t)=h(t).q(t)=-h(t). Then

Recζ1dζv(A,ζ)\displaystyle\mathrm{Re\,}\int^{\zeta_{1}}_{c}\frac{d\zeta}{v(A,\zeta)} =Re0bidtv(A,c+it)=Imb0dtg(t)iq(t)\displaystyle=\mathrm{Re\,}\int^{-b}_{0}\frac{idt}{v(A,c+it)}=\mathrm{Im\,}\int^{0}_{-b}\frac{dt}{\sqrt{g(t)-iq(t)}}
=12b0g(t)2+q(t)2g(t)g(t)2+q(t)2𝑑t>0,\displaystyle=\frac{1}{\sqrt{2}}\int_{-b}^{0}\frac{\sqrt{\sqrt{g(t)^{2}+q(t)^{2}}-g(t)}}{\sqrt{g(t)^{2}+q(t)^{2}}}dt>0,

which implies (/A)ReJ𝐚(A)<0(\partial/\partial A)\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)<0 for A4.A\geq 4. This fact combined with ReJ𝐚(4)<0\mathrm{Re\,}J_{\mathbf{a}_{*}}(4)<0 leads to ReJ𝐚(A)<0\mathrm{Re\,}J_{\mathbf{a}_{*}}(A)<0 for A4.A\geq 4.

Thus we have shown the following uniqueness.

Proposition 8.4.

The Boutroux equations (BE)ϕ=0(\mathrm{BE})_{\phi=0} admit a unique solution A0=827A_{0}=\tfrac{8}{27}.

Corollary 8.5.

For every AA\in\mathbb{C}, (J𝐚(A),J𝐛(A))(0,0).(J_{\mathbf{a}_{*}}(A),J_{\mathbf{b}_{*}}(A))\not=(0,0).

Proposition 8.6.

Suppose that 0<|ϕ|π/4.0<|\phi|\leq\pi/4. Let AϕA_{\phi} solve (BE)ϕ(\mathrm{BE})_{\phi}, and let the elliptic curve v2=ζ4+4ζ3+4ζ2+4Aϕζv^{2}=\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A_{\phi}\zeta be degenerate. Then ϕ=±π/4\phi=\pm\pi/4 and A±π/4=0.A_{\pm\pi/4}=0.

Proof..

The degeneration occurs only when Aϕ=0,A_{\phi}=0, 827\tfrac{8}{27}. By Example 8.1 A=827A=\tfrac{8}{27} does not solve (BE)ϕ(\mathrm{BE})_{\phi} for 0<|ϕ|π/4.0<|\phi|\leq\pi/4. As shown in the case (z.2) above J𝐚(0)=0J_{\mathbf{a}_{*}}(0)=0 and J𝐛(0)<0J_{\mathbf{b}_{*}}(0)<0, which implies A=0A=0 solves (BE)ϕ(\mathrm{BE})_{\phi} only when ϕ=±π/4.\phi=\pm\pi/4. This completes the proof. ∎

8.3. Trajectory

The ratio I(A)=J𝐛(A)/J𝐚(A)I(A)={J_{\mathbf{b}_{*}}(A)}/{J_{\mathbf{a}_{*}}(A)} by Novokshenov [24, Appendix I] is useful in examining the Boutroux equations. The following is easily verified.

Proposition 8.7.

(1)(1) If I(A)I(A)\in\mathbb{R}, then AA solves (BE)ϕ(\mathrm{BE})_{\phi} for some ϕ.\phi\in\mathbb{R}.

(2)(2) If AA solves (BE)ϕ(\mathrm{BE})_{\phi} for some ϕ\phi\in\mathbb{R} and J𝐚(A)0,J_{\mathbf{a}_{*}}(A)\not=0, then I(A).I(A)\in\mathbb{R}.

Remark 8.1.

By Corollary 8.5 if J𝐚(A)=0J_{\mathbf{a}_{*}}(A)=0 then I(A)1=J𝐚(A)/J𝐛(A)=0I(A)^{-1}={J_{\mathbf{a}_{*}}(A)}/{J_{\mathbf{b}_{*}}(A)}=0, and around such a point I(A)1I(A)^{-1} has the same property as in Proposition 8.7.

Let D0={A|A solves (BE)ϕ for some ϕ}D_{0}=\{A\in\mathbb{C}\,|\,\text{$A$ solves $(\mathrm{BE})_{\phi}$ for some $\phi$}\}.

Proposition 8.8.

The set D0D_{0} is bounded.

Proof..

Let us calculate limAI(A)\lim_{A\to\infty}I(A). The zeros of ζ3+4ζ2+4ζ+4A=0\zeta^{3}+4\zeta^{2}+4\zeta+4A=0 are asymptotically expressed as ζje(12j)πi/3(4A)1/3\zeta_{j}\sim e^{(1-2j)\pi i/3}(4A)^{1/3} (j=1,3,5)(j=1,3,5) as AA\to\infty. Then, by ζ=eπi/3(4A)1/3z\zeta=e^{\pi i/3}(4A)^{1/3}z,

J𝐚(A)=\displaystyle J_{\mathbf{a}_{*}}(A)= 2ζ5ζ3v(A,ζ)ζ𝑑ζ2e2πi/3(4A)2/3e2πi/31z4zz𝑑z,\displaystyle 2\int^{\zeta_{3}}_{\zeta_{5}}\frac{v(A,\zeta)}{\zeta}d\zeta\sim 2e^{2\pi i/3}(4A)^{2/3}\int^{1}_{e^{2\pi i/3}}\frac{\sqrt{z^{4}-z}}{z}dz,
J𝐛(A)=\displaystyle J_{\mathbf{b}_{*}}(A)= 2ζ3ζ1v(A,ζ)ζ𝑑ζ2e2πi/3(4A)2/31e2πi/3z4zz𝑑z.\displaystyle 2\int^{\zeta_{1}}_{\zeta_{3}}\frac{v(A,\zeta)}{\zeta}d\zeta\sim 2e^{2\pi i/3}(4A)^{2/3}\int_{1}^{e^{-2\pi i/3}}\frac{\sqrt{z^{4}-z}}{z}dz.

Since

z4zz𝑑z=12z4z34dzz4z,\int\frac{\sqrt{z^{4}-z}}{z}dz=\frac{1}{2}\sqrt{z^{4}-z}-\frac{3}{4}\int\frac{dz}{\sqrt{z^{4}-z}},

limAI(A)\lim_{A\to\infty}I(A) is a ratio of periods of the elliptic curve v2=z4z,v^{2}=z^{4}-z, which implies ImlimAI(A)0.\mathrm{Im\,}\lim_{A\to\infty}I(A)\not=0. By Proposition 8.7 the set D0D_{0} is bounded. ∎

Recall the periods Ω𝐚=𝐚v(A,ζ)1𝑑ζ\Omega^{*}_{\mathbf{a}_{*}}=\int_{\mathbf{a}_{*}}v(A,\zeta)^{-1}d\zeta and Ω𝐛=𝐛v(A,ζ)1𝑑ζ\Omega^{*}_{\mathbf{b}_{*}}=\int_{\mathbf{b}_{*}}v(A,\zeta)^{-1}d\zeta of ΠA\Pi^{*}_{A} given in Section 2.2. To examine the conformality of I(A)I(A) we need the following (see also [31, Lemma 6.1]).

Lemma 8.9.

Ω𝐛J𝐚(A)Ω𝐚J𝐛(A)=4πi.\Omega^{*}_{\mathbf{b}_{*}}J_{\mathbf{a}_{*}}(A)-\Omega^{*}_{\mathbf{a}_{*}}J_{\mathbf{b}_{*}}(A)=-4\pi i.

Proof..

Let ζ(t)\zeta(t) be the elliptic function given by (ζ)2=v(A,ζ)2=ζ4+4ζ3+4ζ2+4Aζ,(\zeta^{\prime})^{2}=v(A,\zeta)^{2}=\zeta^{4}+4\zeta^{3}+4\zeta^{2}+4A\zeta, and let P,Q,R,SP,Q,R,S be the vertices of its period parallelogram such that Q=P+Ω𝐚Q=P+\Omega^{*}_{\mathbf{a}_{*}} and S=P+Ω𝐛S=P+\Omega^{*}_{\mathbf{b}_{*}}. Then the function g(u)=Puζ(t)2ζ(t)1𝑑t{g}(u)=\int^{u}_{P}{\zeta^{\prime}(t)^{2}}{\zeta(t)^{-1}}dt fulfils

g(u+Ω𝐛)g(u)=uu+Ω𝐛ζ(t)2ζ(t)𝑑t=𝐛v(A,ζ)ζ𝑑ζ=J𝐛(A),{g}(u+\Omega^{*}_{\mathbf{b}_{*}})-g(u)=\int^{u+\Omega^{*}_{\mathbf{b}_{*}}}_{u}\frac{\zeta^{\prime}(t)^{2}}{\zeta(t)}dt=\int_{\mathbf{b}_{*}}\frac{v(A,\zeta)}{\zeta}d\zeta=J_{\mathbf{b}_{*}}(A),

and hence

(PQ+RS)g(u)du=PQ(g(u)g(u+Ω𝐛))𝑑u=PQ(J𝐛(A))𝑑u=Ω𝐚J𝐛(A).\biggl{(}\int_{P}^{Q}+\int_{R}^{S}\biggr{)}g(u)du=\int_{P}^{Q}(g(u)-g(u+\Omega^{*}_{\mathbf{b}_{*}}))du=\int_{P}^{Q}(-J_{\mathbf{b}_{*}}(A))du=-\Omega^{*}_{\mathbf{a}_{*}}J_{\mathbf{b}_{*}}(A).

Combining this with (QR+SP)g(u)du=Ω𝐛J𝐚(A)(\int_{Q}^{R}+\int_{S}^{P})g(u)du=\Omega^{*}_{\mathbf{b}_{*}}J_{\mathbf{a}_{*}}(A) similarly obtained, we have

Ω𝐛J𝐚(A)Ω𝐚J𝐛(A)=(PQRSP)g(u)𝑑u=2πi(Res(u)+Res(u+))=4πi,\Omega^{*}_{\mathbf{b}_{*}}J_{\mathbf{a}_{*}}(A)-\Omega^{*}_{\mathbf{a}_{*}}J_{\mathbf{b}_{*}}(A)=\int_{(PQRSP)}g(u)du=2\pi i(\mathrm{Res}(u_{\infty}^{-})+\mathrm{Res}(u_{\infty}^{+}))=-4\pi i,

where u±u_{\infty}^{\pm} are poles of g(u)g(u) in the periodic parallelogram. ∎

As an immediate corollary for Ω𝐚,𝐛\Omega_{\mathbf{a},\,\mathbf{b}} and 𝒥𝐚,𝐛\mathcal{J}_{\mathbf{a},\,\mathbf{b}} (cf. Section 6.3) we have the following.

Corollary 8.10.

Ω𝐛𝒥𝐚Ω𝐚𝒥𝐛=4πieiϕ.\Omega_{\mathbf{b}}\mathcal{J}_{\mathbf{a}}-\Omega_{\mathbf{a}}\mathcal{J}_{\mathbf{b}}=-4\pi ie^{i\phi}.

Observing that (/A)J𝐚,𝐛(A)=2Ω𝐚,𝐛,(\partial/\partial A)J_{\mathbf{a}_{*},\,\mathbf{b}_{*}}(A)=2\Omega^{*}_{\mathbf{a}_{*},\,\mathbf{b}_{*}}, we have the following.

Proposition 8.11.

I(A)=8πiJ𝐚(A)2I^{\prime}(A)=-8\pi iJ_{\mathbf{a}_{*}}(A)^{-2} and (1/I)(A)=8πiJ𝐛(A)2(1/I)^{\prime}(A)=8\pi iJ_{\mathbf{b}_{*}}(A)^{-2}.

Remark 8.2.

By Example 8.1 and the proposition above, I(A)I(A) is conformal around A=A0=827A=A_{0}=\tfrac{8}{27}, and given by

I(A)=32πi(A827)(1+o(1)).I(A)=\tfrac{3}{2}\pi i(A-\tfrac{8}{27})(1+o(1)).

By Proposition 8.7 the inverse image of the interval (ε,ϵ)(-\varepsilon,\epsilon)\subset\mathbb{R} under I(A)I(A) is a local trajectory consisting of AϕA_{\phi} for |ϕ|<ε0|\phi|<\varepsilon_{0}, each AϕA_{\phi} solving (BE)ϕ, where ε\varepsilon and ε0\varepsilon_{0} are sufficiently small. This local trajectory is expressed as

Aϕ=827+iρ(ϕ),32πReρ(ϕ)(ε,ε),Imρ(ϕ)=o(Reρ(ϕ)) as ϕ0.A_{\phi}=\tfrac{8}{27}+i\rho(\phi),\quad\tfrac{3}{2}\pi\mathrm{Re\,}\rho(\phi)\in(-\varepsilon,\varepsilon),\quad\text{$\mathrm{Im\,}\rho(\phi)=o(\mathrm{Re\,}\rho(\phi))$ as $\phi\to 0$.}

Similarly, there exists a local trajectory for |ϕ±π/4|<ε0|\phi\pm\pi/4|<\varepsilon_{0} (cf. Proof of Proposition 8.6).

Suppose that 0<|ϕ|<π/40<|\phi|<\pi/4. Write

J𝐚(A)=u(A)+iv(A),J𝐛(A)=U(A)+iV(A),A=x+iy.J_{\mathbf{a}_{*}}(A)=u(A)+iv(A),\quad J_{\mathbf{b}_{*}}(A)=U(A)+iV(A),\quad A=x+iy.

Then AA solves (BE)ϕ if and only if

(8.2) u(A)v(A)tan2ϕ=0,U(A)V(A)tan2ϕ=0,u(A)-v(A)\tan 2\phi=0,\quad U(A)-V(A)\tan 2\phi=0,

which define the trajectory of AϕA_{\phi}. The Jacobian for (8.2) around A=AϕA=A_{\phi} is

detJ(x,y)=\displaystyle\det J(x,y)= det(uxvxtan2ϕuyvytan2ϕUxVxtan2ϕUyVytan2ϕ)=(1+tan22ϕ)(vxVyvyVx)\displaystyle\det\begin{pmatrix}u_{x}-v_{x}\tan 2\phi&u_{y}-v_{y}\tan 2\phi\\ U_{x}-V_{x}\tan 2\phi&U_{y}-V_{y}\tan 2\phi\end{pmatrix}=(1+\tan^{2}2\phi)(v_{x}V_{y}-v_{y}V_{x})
=\displaystyle= 4(1+tan22ϕ)ImΩ𝐚¯Ω𝐛=4(1+tan22ϕ)|Ω𝐚|2ImΩ𝐚Ω𝐛0\displaystyle-4(1+\tan^{2}2\phi)\mathrm{Im\,}\overline{\Omega^{*}_{\mathbf{a}_{*}}}\Omega^{*}_{\mathbf{b}_{*}}=-4(1+\tan^{2}2\phi)|\Omega^{*}_{\mathbf{a}_{*}}|^{2}\mathrm{Im\,}\frac{\Omega^{*}_{\mathbf{a}_{*}}}{\Omega^{*}_{\mathbf{b}_{*}}}\not=0

by Proposition 8.6. This fact implies that a given local trajectory for |ϕϕ0|<ε0|\phi-\phi_{0}|<\varepsilon_{0} with |ϕ0|<π/4|\phi_{0}|<\pi/4 is extended smoothly for 0<|ϕ|<π/40<|\phi|<\pi/4 and continuously for |ϕ|π/4|\phi|\leq\pi/4, and so are the trajectories described in Remark 8.2. For such an extended trajectory, which is bounded by Proposition 8.8, let AϕnA_{\phi_{n}} with ϕn0\phi_{n}\to 0 be a given sequence. By the boundedness there exists a subsequence convergent to some A0A_{0}^{*}\in\mathbb{C} solving (BE)ϕ=0, and then A0=827,A_{0}^{*}=\tfrac{8}{27}, which implies the uniqueness of the trajectory for |ϕ|π/4.|\phi|\leq\pi/4. Thus we have the following.

Proposition 8.12.

There exists a trajectory A=AϕA=A_{\phi} for |ϕ|π/4|\phi|\leq\pi/4 with the properties::

(1)(1) for each ϕ\phi, AϕA_{\phi} is a unique solution of (BE)ϕ;(\mathrm{BE})_{\phi};

(2)(2) AϕA_{\phi} is smooth in ϕ\phi for 0<|ϕ|<π/4,0<|\phi|<\pi/4, and continuous in ϕ\phi for |ϕ|π/4;|\phi|\leq\pi/4;

(3)(3) A0=827,A_{0}=\tfrac{8}{27}, A±π/4=0.A_{\pm\pi/4}=0.

For any ϕ\phi\in\mathbb{R} we have

e2iϕJ𝐚,𝐛(Aϕ)=e2i(ϕ±π/2)J𝐚,𝐛(Aϕ),e2iϕJ𝐚,𝐛(Aϕ)¯=e2iϕJ𝐚¯,𝐛¯(Aϕ¯),e^{2i\phi}J_{\mathbf{a}_{*},\,\mathbf{b}_{*}}(A_{\phi})=-e^{2i(\phi\pm\pi/2)}J_{\mathbf{a}_{*},\,\mathbf{b}_{*}}(A_{\phi}),\quad\overline{e^{2i\phi}J_{\mathbf{a}_{*},\,\mathbf{b}_{*}}(A_{\phi})}=e^{-2i\phi}J_{\overline{\mathbf{a}_{*}},\,\overline{\mathbf{b}_{*}}}(\overline{A_{\phi}}),

which leads to the following.

Proposition 8.13.

Aϕ±π/2=AϕA_{\phi\pm\pi/2}=A_{\phi} and Aϕ=Aϕ¯.A_{-\phi}=\overline{A_{\phi}}.

To know the shape of the trajectory A=AϕA=A_{\phi} it is sufficient to examine it for |ϕ|π/4|\phi|\leq\pi/4. For 0<|ϕ|<π/40<|\phi|<\pi/4 the derivative of (8.2) along A=AϕA=A_{\phi} with respect to t=tan2ϕt=\tan 2\phi is written in the form J(x,y)(x(t),y(t))T(v(Aϕ),V(Aϕ))T=𝐨J(x,y){}^{\mathrm{T}}\!(x^{\prime}(t),y^{\prime}(t))-{}^{\mathrm{T}}\!(v(A_{\phi}),V(A_{\phi}))=\mathbf{o} with Aϕ=x(t)+iy(t),A_{\phi}=x(t)+iy(t), where J(x,y)J(x,y) is the Jacobi matrix above. Then, for 0<|ϕ|<π/40<|\phi|<\pi/4

(8.3) (x(t),y(t))(0,0),(d/dϕ)Aϕ=2(x(t)+iy(t))cos22ϕ0.(x^{\prime}(t),y^{\prime}(t))\not=(0,0),\quad(d/d\phi)A_{\phi}=2(x^{\prime}(t)+iy^{\prime}(t))\cos^{-2}2\phi\not=0.

By Propositions 8.7 and 8.11, for every ϕ\phi such that |ϕ|π/4|\phi|\leq\pi/4,

ddtI(Aϕ)=8πi(x(t)+iy(t))J𝐚(Aϕ)2orddt(1/I)(Aϕ)=8πi(x(t)+iy(t))J𝐛(Aϕ)2,\frac{d}{dt}I(A_{\phi})=-\frac{8\pi i(x^{\prime}(t)+iy^{\prime}(t))}{J_{\mathbf{a}_{*}}(A_{\phi})^{2}}\,\,\,\text{or}\,\,\,\frac{d}{dt}(1/I)(A_{\phi})=\frac{8\pi i(x^{\prime}(t)+iy^{\prime}(t))}{J_{\mathbf{b}_{*}}(A_{\phi})^{2}}\in\mathbb{R},

where (J𝐚(Aϕ),J𝐛(Aϕ))(0,0)(J_{\mathbf{a}_{*}}(A_{\phi}),J_{\mathbf{b}_{*}}(A_{\phi}))\not=(0,0) by Corollary 8.5. Setting J𝐚(Aϕ)1J_{\mathbf{a}_{*}}(A_{\phi})^{-1} or J𝐛(Aϕ)1=P+iQ=i(QiP)J_{\mathbf{b}_{*}}(A_{\phi})^{-1}=P+iQ=i(Q-iP) with P2+Q2>0P^{2}+Q^{2}>0, and observing Im(d/dt)I(Aϕ)±1=0\mathrm{Im\,}(d/dt)I(A_{\phi})^{\pm 1}=0, we have

x(t)(P2Q2)2y(t)PQ=0.x^{\prime}(t)(P^{2}-Q^{2})-2y^{\prime}(t)PQ=0.

For 0<|ϕ|<π/4,0<|\phi|<\pi/4, by (8.3), x(t)0x^{\prime}(t)\not=0, and if y(t)=0,y^{\prime}(t)=0, then P=±QP=\pm Q, implying 2ϕ=π/4.2\phi=\mp\pi/4. For π/4<ϕ<0-\pi/4<\phi<0, x(t)>0x^{\prime}(t)>0, and for 0<ϕ<π/40<\phi<\pi/4, x(t)<0x^{\prime}(t)<0, since A±π/4=0A_{\pm\pi/4}=0 and A0=827A_{0}=\tfrac{8}{27}. If π/8<ϕ<0-\pi/8<\phi<0 (respectively, 0<ϕ<π/80<\phi<\pi/8) then y(t)<0y^{\prime}(t)<0, since |P|<|Q||P|<|Q|, PQ>0PQ>0 (respectively, PQ<0PQ<0).

Proposition 8.14.

Let Aϕ=x(t)+iy(t)A_{\phi}=x(t)+iy(t) with t=tan2ϕ.t=\tan 2\phi. Then, for |ϕ|π/4,|\phi|\leq\pi/4,

(1)(1) x(t)>0x^{\prime}(t)>0 for π/4<ϕ<0,-\pi/4<\phi<0, x(t)<0x^{\prime}(t)<0 for 0<ϕ<π/4;0<\phi<\pi/4;

(2)(2) y(t)<0y^{\prime}(t)<0 for 0<|ϕ|<π/4,0<|\phi|<\pi/4, y(t)>0y^{\prime}(t)>0 for π/4<|ϕ|<π/2;\pi/4<|\phi|<\pi/2;

(3)(3) x(0)=x(±tan(π/2))=0,x^{\prime}(0)=x^{\prime}(\pm\tan(\pi/2))=0, y(±tan(π/4))=0.y^{\prime}(\pm\tan(\pi/4))=0.

By this proposition and Remark 8.2 the trajectory AϕA_{\phi} for |ϕ|π/4|\phi|\leq\pi/4 is roughly drawn as in Figure 8.6.

A0=827A_{0}=\tfrac{8}{27}A±π/4=0A_{\pm\pi/4}=0
Figure 8.6. Trajectory AϕA_{\phi} for |ϕ|π/4|\phi|\leq\pi/4
Proposition 8.15.

There exists a trajectory A=AϕA=A_{\phi} for ϕ\phi\in\mathbb{R} with the properties::

(1)(1) for each ϕ\phi, AϕA_{\phi} is a unique solution of (BE)ϕ;(\mathrm{BE})_{\phi};

(2)(2) Aϕ+π/2=Aϕ,A_{\phi+\pi/2}=A_{\phi}, Aϕ=Aϕ¯;A_{-\phi}=\overline{A_{\phi}};

(3)(3) A0=827,A_{0}=\tfrac{8}{27}, A±π/4=0A_{\pm\pi/4}=0 and 0ReAϕ827;0\leq\mathrm{Re\,}A_{\phi}\leq\tfrac{8}{27};

(4)(4) AϕA_{\phi} is continuous in ϕ\phi\in\mathbb{R}, and smooth in ϕ{mπ/4|m}.\phi\in\mathbb{R}\setminus\{m\pi/4\,|\,m\in\mathbb{Z}\}.

As in the derivation of Corollary 8.2 by the change of variables ζ=eϕz\zeta=e^{-\phi}z Proposition 8.15 is converted to the results on the trajectory for the Boutroux equations (2.2) with the cycles 𝐚\mathbf{a} and 𝐛\mathbf{b} on the elliptic curve ΠA,ϕ:\Pi_{A,\phi}: w(A,z)2=z4+4eiϕz3+4e2iϕz2+4e3iϕAz.w(A,z)^{2}=z^{4}+4e^{i\phi}z^{3}+4e^{2i\phi}z^{2}+4e^{3i\phi}Az.

Corollary 8.16.

The trajectory A=AϕA=A_{\phi} in Proposition 8.15 fulfils

(1)(1^{\prime}) for each ϕ\phi\in\mathbb{R}, AϕA_{\phi} is a unique solution of (2.2).

Remark 8.3.

In the corollary above (1)(1^{\prime}) may be replaced with

(1′′)(1^{\prime\prime}) for each ϕ\phi, there exists AϕA_{\phi} uniquely such that, for every cycle 𝐜\mathbf{c} on ΠAϕ,ϕ\Pi_{A_{\phi},\phi},

Re𝐜w(Aϕ,z)z𝑑z=0.\mathrm{Re\,}\int_{\mathbf{c}}\frac{w(A_{\phi},z)}{z}dz=0.

8.4. Coalescing turning points

Let us observe coalescing turning points as an application of Proposition 8.14. On the trajectory AϕA_{\phi}, at Aϕ=0=827A_{\phi=0}=\tfrac{8}{27} and Aϕ=±π/4=0A_{\phi=\pm\pi/4}=0, coalescences of, respectively, ζ3\zeta_{3} and ζ1,\zeta_{1}, and ζ5\zeta_{5} and ζ3\zeta_{3} occur, and then the elliptic curve ΠA\Pi^{*}_{A} degenerates.

For Aϕ=827+εA_{\phi}=\tfrac{8}{27}+\varepsilon the zeros of v(Aϕ,ζ)2v(A_{\phi},\zeta)^{2} are denoted by ζ5=83+δ5,\zeta_{5}=-\tfrac{8}{3}+\delta_{5}, ζ3,1=23+δ3,1\zeta_{3,1}=-\tfrac{2}{3}+\delta_{3,1}, where δ\delta and δj\delta_{j} (j=1,3,5)(j=1,3,5) are small. Then δ3,1=±i(2ε)1/2+O(ε),\delta_{3,1}=\pm i(2\varepsilon)^{1/2}+O(\varepsilon), δ5=ε+O(ε2).\delta_{5}=-\varepsilon+O(\varepsilon^{2}). By Proposition 8.14, ε\varepsilon may be written in the form 2ε=eiθ(ϕ)ρ(ϕ)(1+o(1))2\varepsilon=e^{-i\theta(\phi)}\rho(\phi)(1+o(1)), where θ(ϕ)=π/2\theta(\phi)=\pi/2 if ϕ>0\phi>0 and =π/2=-\pi/2 if ϕ<0\phi<0, and ρ(0)=0\rho(0)=0 and ρ(ϕ)>0\rho(\phi)>0 for |ϕ|>0|\phi|>0 around ϕ=0\phi=0. Hence ζ3,1=23(132ieiθ(ϕ)/2ρ(ϕ)1/2(1+o(1)))\zeta_{3,1}=-\tfrac{2}{3}(1\mp\tfrac{3}{2}ie^{-i\theta(\phi)/2}\rho(\phi)^{1/2}(1+o(1))), ζ5=83(1+316eiθ(ϕ)ρ(ϕ)(1+o(1)))\zeta_{5}=-\tfrac{8}{3}(1+\tfrac{3}{16}e^{-i\theta(\phi)}\rho(\phi)(1+o(1))) as ϕ0,\phi\to 0, where each of \mp is chosen in such a way that Reζ3<Reζ1.\mathrm{Re\,}\zeta_{3}<\mathrm{Re\,}\zeta_{1}. By λ2=z=eiϕζ\lambda^{2}=z=e^{i\phi}\zeta, the turning points λj\lambda_{j} (j=1,3,5)(j=1,3,5) in Section 4 have the following property.

Proposition 8.17.

Around ϕ=0\phi=0,

λ5\displaystyle\lambda_{5} =i83+ϱ^(ϕ)(1+o(1)),\displaystyle=i\sqrt{\tfrac{8}{3}}+\hat{\varrho}(\phi)(1+o(1)),\quad
λ3\displaystyle\lambda_{3} =i23+e3πi/4ϱ(ϕ)(1+o(1)),λ1=i23+eπi/4ϱ(ϕ)(1+o(1))\displaystyle=i\sqrt{\tfrac{2}{3}}+e^{3\pi i/4}\varrho(\phi)(1+o(1)),\quad\lambda_{1}=i\sqrt{\tfrac{2}{3}}+e^{-\pi i/4}\varrho(\phi)(1+o(1))

as ϕ0+,\phi\to 0+, and

λ5\displaystyle\lambda_{5} =i83ϱ^(ϕ)(1+o(1)),\displaystyle=i\sqrt{\tfrac{8}{3}}-\hat{\varrho}(\phi)(1+o(1)),\quad
λ3\displaystyle\lambda_{3} =i23+eπi/4ϱ(ϕ)(1+o(1)),λ1=i23+e3πi/4ϱ(ϕ)(1+o(1))\displaystyle=i\sqrt{\tfrac{2}{3}}+e^{\pi i/4}\varrho(\phi)(1+o(1)),\quad\lambda_{1}=i\sqrt{\tfrac{2}{3}}+e^{-3\pi i/4}\varrho(\phi)(1+o(1))

as ϕ0,\phi\to 0-, where ϱ(0)=0\varrho(0)=0, ϱ(ϕ)>0\varrho(\phi)>0 for |ϕ|>0|\phi|>0, and ϱ^(ϕ)ϱ(ϕ)2.\hat{\varrho}(\phi)\asymp\varrho(\phi)^{2}.

For Aϕ=εA_{\phi}=\varepsilon near ϕ=±π/4\phi=\pm\pi/4, the zeros of v(A,ζ)2v(A,\zeta)^{2} are ζ5,3=2+δ5,3,\zeta_{5,3}=-2+\delta_{5,3}, ζ1=δ1\zeta_{1}=\delta_{1} with δ5,3=±(2ε)1/2+O(ε),\delta_{5,3}=\pm(2\varepsilon)^{1/2}+O(\varepsilon), δ1=ε+O(ε2).\delta_{1}=-\varepsilon+O(\varepsilon^{2}). Writing 2ε=eiθ(ϕ)ρ~(θ)2\varepsilon=e^{i\theta(\phi^{\prime})}\tilde{\rho}(\theta^{\prime}), where ϕ=±π/4ϕ\phi=\pm\pi/4\mp\phi^{\prime}, ϕ0\phi^{\prime}\geq 0, and ρ~(0)=0,\tilde{\rho}(0)=0, ρ~(ϕ)0\tilde{\rho}(\phi^{\prime})\geq 0, we have ζ5,3=2±eiθ(ϕ)/2ρ~(ϕ)1/2(1+o(1))\zeta_{5,3}=-2\pm e^{i\theta(\phi^{\prime})/2}\tilde{\rho}(\phi^{\prime})^{1/2}(1+o(1)), ζ1=12eiθ(ϕ)ρ~(ϕ)(1+o(1))\zeta_{1}=-\tfrac{1}{2}e^{i\theta(\phi^{\prime})}\tilde{\rho}(\phi^{\prime})(1+o(1)).

Proposition 8.18.

Around ϕ=±π/4\phi=\pm\pi/4,

λ5\displaystyle\lambda_{5} =eπi/8(i2+e3πi/4ϱ(ϕ)(1+o(1))),λ3=eπi/8(i2+eπi/4ϱ(ϕ)(1+o(1))),\displaystyle=e^{-\pi i/8}(i\sqrt{2}+e^{3\pi i/4}\varrho(\phi^{\prime})(1+o(1))),\quad\lambda_{3}=e^{-\pi i/8}(i\sqrt{2}+e^{-\pi i/4}\varrho(\phi^{\prime})(1+o(1))),
λ1\displaystyle\lambda_{1} =12eπi/8e3πi/4ϱ(ϕ)(1+o(1))),\displaystyle=\tfrac{1}{2}e^{-\pi i/8}e^{3\pi i/4}{\varrho}(\phi^{\prime})(1+o(1))),\quad

as ϕ=ϕ+π/40+,\phi^{\prime}=\phi+\pi/4\to 0+, and

λ5\displaystyle\lambda_{5} =eπi/8(i2+eπi/4ϱ(ϕ)(1+o(1))),λ3=eπi/8(i2+e3πi/4ϱ(ϕ)(1+o(1))),\displaystyle=e^{\pi i/8}(i\sqrt{2}+e^{\pi i/4}\varrho(\phi^{\prime})(1+o(1))),\quad\lambda_{3}=e^{\pi i/8}(i\sqrt{2}+e^{-3\pi i/4}\varrho(\phi^{\prime})(1+o(1))),
λ1\displaystyle\lambda_{1} =12eπi/8eπi/4ϱ(ϕ)(1+o(1))),\displaystyle=\tfrac{1}{2}e^{\pi i/8}e^{\pi i/4}{\varrho}(\phi^{\prime})(1+o(1))),

as ϕ=ϕπ/40,\phi^{\prime}=\phi-\pi/4\to 0-, where ϱ(0)=0{\varrho}(0)=0, and ϱ(ϕ)>0{\varrho}(\phi^{\prime})>0 for |ϕ|>0|\phi^{\prime}|>0.

The quantities above are also written as functions of ϕ.\phi. We have the following, say, around ϕ=0.\phi=0.

Proposition 8.19.

As ϕ0\phi\to 0- ((respectively, ϕ0+\phi\to 0+)), ζp=23ip2εϕ+O(εϕ)\zeta_{p}=-\tfrac{2}{3}-i^{p}\sqrt{2\varepsilon_{\phi}}+O(\varepsilon_{\phi}) ((respectively, =23+ip2εϕ+O(εϕ)=-\tfrac{2}{3}+i^{p}\sqrt{2\varepsilon_{\phi}}+O(\varepsilon_{\phi}))) for p=1,3p=1,3, and ζ5=83εϕ+O(εϕ2),\zeta_{5}=-\tfrac{8}{3}-\varepsilon_{\phi}+O(\varepsilon_{\phi}^{2}), where Aϕ=827+εϕA_{\phi}=\tfrac{8}{27}+\varepsilon_{\phi} with

εϕ=8i3ϕlnϕ(1+O(lnlnϕlnϕ)).\varepsilon_{\phi}=\frac{8i}{3}\frac{\phi}{\ln\phi}\Bigl{(}1+O\Bigl{(}\frac{\ln\ln\phi}{\ln\phi}\Bigr{)}\Bigr{)}.
Proof..

In the case ϕ0,\phi\to 0-, set Aϕ=827+εA_{\phi}=\tfrac{8}{27}+\varepsilon, and ζ1=23i2ε+O(ε),\zeta_{1}=-\tfrac{2}{3}-i\sqrt{2\varepsilon}+O(\varepsilon), ζ3=23+i2ε+O(ε),\zeta_{3}=-\tfrac{2}{3}+i\sqrt{2\varepsilon}+O(\varepsilon), ζ5=83ε+O(ε2),\zeta_{5}=-\tfrac{8}{3}-\varepsilon+O(\varepsilon^{2}), where ε=i|ε|(1+o(1))\varepsilon=i|\varepsilon|(1+o(1)). Then

12J𝐚(Aϕ)\displaystyle\tfrac{1}{2}J_{\mathbf{a}_{*}}(A_{\phi}) =i3ζ5ζ3(ζζ5)(ζ3ζ)(ζ1ζ)(ζ)dζζ\displaystyle=i^{3}\int^{\zeta_{3}}_{\zeta_{5}}\sqrt{(\zeta-\zeta_{5})(\zeta_{3}-\zeta)(\zeta_{1}-\zeta)(-\zeta)}\frac{d\zeta}{\zeta}
=i8/32/3+i2ε(ζ+83)(23+i2εζ)(23i2εζ)(ζ)dζζ+O(ε)\displaystyle=-i\int^{-2/3+i\sqrt{2\varepsilon}}_{-8/3}\sqrt{(\zeta+\tfrac{8}{3})(-\tfrac{2}{3}+i\sqrt{2\varepsilon}-\zeta)(-\tfrac{2}{3}-i\sqrt{2\varepsilon}-\zeta)(-\zeta)}\frac{d\zeta}{\zeta}+O(\varepsilon)
=i02+i2εt83t(2t)2+2ε𝑑t+O(ε)(ζ+83=t)\displaystyle=i\int^{2+i\sqrt{2\varepsilon}}_{0}\sqrt{\frac{t}{\frac{8}{3}-t}}\sqrt{(2-t)^{2}+2\varepsilon}\,dt+O(\varepsilon)\qquad(\zeta+\tfrac{8}{3}=t)
=i02+i2εt83t(2t+ε(2t)1+O(ε2(2t)3))𝑑t+O(ε)\displaystyle=i\int^{2+i\sqrt{2\varepsilon}}_{0}\sqrt{\frac{t}{\frac{8}{3}-t}}\Bigl{(}2-t+{\varepsilon}(2-t)^{-1}+O(\varepsilon^{2}(2-t)^{-3})\Bigr{)}dt+O(\varepsilon)
=i(233123εlnε+O(ε)),\displaystyle=i(\tfrac{2}{3}\sqrt{3}-\tfrac{1}{2}\sqrt{3}\varepsilon\ln\varepsilon+O(\varepsilon)),

and

12J𝐛(Aϕ)\displaystyle\tfrac{1}{2}J_{\mathbf{b}_{*}}(A_{\phi}) =i2ζ3ζ1(ζζ5)(ζζ3)(ζ1ζ)(ζ)dζζ\displaystyle=i^{2}\int^{\zeta_{1}}_{\zeta_{3}}\sqrt{(\zeta-\zeta_{5})(\zeta-\zeta_{3})(\zeta_{1}-\zeta)(-\zeta)}\frac{d\zeta}{\zeta}
=2/3+i2ε2/3i2ε(ζ+83)(ζ+23i2ε)(23i2εζ)(ζ)dζζ+O(ε3/2)\displaystyle=-\int^{-2/3-i\sqrt{2\varepsilon}}_{-2/3+i\sqrt{2\varepsilon}}\sqrt{(\zeta+\tfrac{8}{3})(\zeta+\tfrac{2}{3}-i\sqrt{2\varepsilon})(-\tfrac{2}{3}-i\sqrt{2\varepsilon}-\zeta)(-\zeta)}\frac{d\zeta}{\zeta}+O(\varepsilon^{3/2})
=3i2εi2ε2εt2𝑑t+O(ε3/2)(ζ+23=t)\displaystyle=\sqrt{3}\int^{-i\sqrt{2\varepsilon}}_{i\sqrt{2\varepsilon}}\sqrt{-2\varepsilon-t^{2}}\,dt+O(\varepsilon^{3/2})\qquad(\zeta+\tfrac{2}{3}=t)
=3π+O(ε3/2).\displaystyle=-\sqrt{3}\pi+O(\varepsilon^{3/2}).

Thus we have J𝐚(Aϕ)=i3(43εlnε+O(ε))J_{\mathbf{a}_{*}}(A_{\phi})=i\sqrt{3}(\tfrac{4}{3}-\varepsilon\ln\varepsilon+O(\varepsilon)) and J𝐛(Aϕ)=23πε+O(ε3/2)J_{\mathbf{b}_{*}}(A_{\phi})=-2\sqrt{3}\pi\varepsilon+O(\varepsilon^{3/2}). For small ϕ\phi, the Boutroux equations Ree2iϕJ𝐚(Aϕ)=Rer2iϕJ𝐛(Aϕ)=0\mathrm{Re\,}e^{2i\phi}J_{\mathbf{a}_{*}}(A_{\phi})=\mathrm{Re\,}r^{2i\phi}J_{\mathbf{b}_{*}}(A_{\phi})=0 yield ε=εϕ\varepsilon=\varepsilon_{\phi} as in the proposition. ∎

By 2Ω𝐚,𝐛=(/εϕ)J𝐚,𝐛(Aϕ)2\Omega^{*}_{\mathbf{a}_{*},\,\mathbf{b}_{*}}=(\partial/\partial\varepsilon_{\phi})J_{\mathbf{a}_{*},\,\mathbf{b}_{*}}(A_{\phi}) we have the following corollary.

Corollary 8.20.

Around ϕ=0\phi=0,

J𝐚(Aϕ)=43i3(12iϕ(1+O(δϕ))),J𝐛(Aϕ)=163πi3ϕ(lnϕ)1(1+O(δϕ)),\displaystyle J_{\mathbf{a}_{*}}(A_{\phi})=\tfrac{4}{3}i{\sqrt{3}}(1-2i\phi(1+O(\delta_{\phi}))),\quad J_{\mathbf{b}_{*}}(A_{\phi})=-\tfrac{16}{3}\pi i{\sqrt{3}}\,{\phi}\,({\ln\phi})^{-1}(1+O(\delta_{\phi})),
Ω𝐚=12i3lnϕ(1+O(δϕ)),Ω𝐛=3π+O(ϕ1/2),\displaystyle\Omega_{\mathbf{a}_{*}}^{*}=-\tfrac{1}{2}i\sqrt{3}\ln\phi\,(1+O(\delta_{\phi})),\quad\Omega_{\mathbf{b}_{*}}^{*}=-{\sqrt{3}\pi}+O({\phi}^{1/2}),

where δϕ=lnlnϕ(lnϕ)1\delta_{\phi}=\ln\ln\phi\,(\ln\phi)^{-1} as ϕ0.\phi\to 0.

Acknowledgements. The author is grateful to Professor Yousuke Ohyama for a stimulating conversation informing circumstances of Kapaev’s announcement [13] and inspiring the author to tackle this work.

References

  • [1] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, Dover, New York, 1972.
  • [2] A. Erdelyi, W. Magnus, F. Oberhettinger and F. G. Tricomi, Higher Transcendental Functions Vols I, II, III (Bateman Manuscript Project), McGraw-Hill, New York, 1953.
  • [3] M. V. Fedoryuk, Asymptotic Analysis, Springer-Verlag, New York, 1993.
  • [4] A. S. Fokas, A. R. Its, A. A. Kapaev and V. Yu. Novokshenov, Painlevé Transcendents, The Riemann-Hilbert Approach, Math. Surveys and Monographs 128, AMS Providence, 2006.
  • [5] V. I. Gromak, Single-parameter families of solutions of Painlevé equations, (Russian) Differentsial’nye Uravneniya 14 (1978), no. 12, 2131–2135, 2298.
  • [6] V. I. Gromak, On the theory of the fourth Painleve equation, (Russian) Differentsial’nye Uravneniya 23 (1987), no. 5, 760–768, 914; translation in Differential Equtions 23 (1987), 506–513.
  • [7] V. I. Gromak, I. Laine and S. Shimomura, Painlevé Differential Equations in the Complex Plane, De Gruyter Studies in Mathematics, 28. Walter de Gruyter & Co., Berlin, 2002.
  • [8] A. Hurwitz and R. Courant, Vorlesungen über allgemeine Funktionentheorie und elliptische Funktionen, Berlin, Springer, 1922.
  • [9] A. R. Its and A. A. Kapaev, Connection formulae for the fourth Painlevé transcendent; Clarkson-McLeod solution, J. Phys. A: Math. Gen. 31 (1998), 4073–4113.
  • [10] M. Jimbo and T. Miwa, Monodromy preserving deformation of linear ordinary differential equations with rational coefficients. II, Phys. D 2 (1981), 407–448.
  • [11] A. A. Kapaev, Global asymptotics of the second Painleé transcendent, Phys. Lett. A 167 (1992), 356–362.
  • [12] A. A. Kapaev, Essential singularity of the Painlevé function of the second kind and the nonlinear Stokes phenomenon. (Russian) Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 187 (1991), Differentsial’naya Geom. Gruppy Li i Mekh. 12, 139–170, 173, 176; translation in J. Math. Sci. 73 (1995), 500–517.
  • [13] A. A. Kapaev, Global asymptotics of the fourth Painleé transcendent, Steklov Mat. Inst. and IUPUI, Preprint # 96-06, 1996; available at http://www.pdmi.ras.ru/ preprint/1996/index.html.
  • [14] A. A. Kapaev, Connection formulae for the degenerated asymptotic solutions of the fourth Painlevé equation, arXiv:solv-int/9805011v1, 1998.
  • [15] A. A. Kapaev and A. V. Kitaev, Connection formulae for the first Painlevé transcendent in the complex domain, Lett. Math. Phys. 27 (1993), 243–252.
  • [16] A. V. Kitaev, Self-similar solutions of the modified nonlinear Schrödinger equation, Teor. Mat. Fiz. 64 (1985), 347–369; translation in Theor. Math. Phys. 64 (1985) 878–894.
  • [17] A. V. Kitaev, Method of isomonodromy deformations for complete third and fourth Painlevé equations, (Russian) PhD Thesis, Leningrad State University (1988).
  • [18] A. V. Kitaev, Asymptotic description of solutions of the fourth Painlevé equation on analogues of Stokes’s rays, (Russian) Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 169 (1988), Voprosy Kvant. Teor. Polya i Statist. Fiz. 8, 84–90, 187; translation in J. Soviet Math. 54 (1991), no. 3, 916–920.
  • [19] A. V. Kitaev, The justification of asymptotic formulas that can be obtained by the method of isomonodromic deformations, (Russian) Zap. Nauchn. Sem. Leningrad. Otdel.  Mat. Inst. Steklov. (LOMI) 179 (1989), Mat. Vopr. Teor. Rasprostr. Voln. 19, 101–109, 189–190; translation in J. Soviet Math. 57 (1991), no. 3, 3131–3135.
  • [20] A. V. Kitaev, The isomonodromy technique and the elliptic asymptotics of the first Painlevé transcendent, (Russian) Algebra i Analiz 5 (1993), 179–211; translation in St. Petersburg Math. J. 5 (1994), 577–605.
  • [21] A. V. Kitaev, Elliptic asymptotics of the first and second Painlevé transcendents, (Russian) Uspekhi Mat. Nauk 49 (1994), 77–140; translation in Russian Math. Surveys 49 (1994), 81–150.
  • [22] Y. Murata, Rational solutions of the second and the fourth Painlevé equations, Funkcial. Ekvac. 28 (1985), 1–32.
  • [23] V. Yu. Novokshenov, A modulated elliptic function as a solution of the second Painlevé equation in the complex plane, (Russian) Dokl. Akad. Nauk SSSR 311 (1990), 288–291; translation in Soviet Math. Dokl. 41 (1990), 252–255.
  • [24] V. Yu. Novokshenov, The Boutroux ansatz for the second Painlevé equation in the complex domain, (Russian) Izv. Akad. Nauk SSSR Ser. Mat. 54 (1990), 1229–1251; translation in Math. USSR-Izv. 37 (1991), 587–609.
  • [25] V. Yu. Novokshenov, Asymptotics in the complex plane of the third Painlevé transcendent, Difference equations, special functions and orthogonal polynomials, 432–451, World Sci. Publ., Hackensack, NJ, 2007.
  • [26] V. Yu. Novokshenov, Connection formulas for the third Painlevé transcendent in the complex plane, Integrable systems and random matrices, 55–69, Contemp. Math., 458, Amer. Math. Soc., Providence, RI, 2008.
  • [27] S. Shimomura, Boutroux ansatz for the degenerate third Painleve transcendents, Publ. Res. Inst. Math. Sci. (to appear) arXiv:2207.11495 math.CA
  • [28] S. Shimomura, Elliptic asymptotics for the complete third Painleve transcendents, Funkcial. Ekvac. (to appear) arXiv:2211.00886 math.CA
  • [29] S. Shimomura, Elliptic asymptotic representation of the fifth Painlevé transcendents, Kyushu J. Math. 76 (2022), 43–99. Corrigendum to ‘Elliptic asymptotic representation of the fifth Painlevé transcendents’, Kyushu J. Math. 77 (2023), 191–202. arXiv:2012.07321 math.CA
  • [30] H. Umemura and H. Watanabe, Solutions of the second and fourth Painlevé equations, I, Nagoya Math. J. 148 (1997), 151–198.
  • [31] V. L. Vereshchagin, Global asymptotics for the fourth Painlevé transcendent, (Russian) Mat. Sb. 188 (1997), 11–32; translation in Sb. Math. 188 (1997), 1739–1760.
  • [32] E. T. Whittaker and G. N. Watson, A course of modern analysis, Reprint of the fourth (1927) edition. Cambridge Mathematical Library. Cambridge University Press, Cambridge, 1996.