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

A New Two-component Sasa–Satsuma Equation: Large-time Asymptotics on the Line

Xiaodan Zhao School of Control and Computer Engineering, North China Electric Power University, Beijing 102206, P.R. China [email protected] (X. Zhao)  and  Lei Wang School of Mathematics and Physics, North China Electric Power University, Beijing 102206, P.R. China [email protected] (L. Wang)
Abstract.

We consider the initial value problem for a new two-component Sasa–Satsuma equation associated with 4×44\times 4 Lax pair with decaying initial data on the line. By utilizing the spectral analysis, the solution of the new two-component Sasa–Satsuma system is transformed into the solution of a 4×44\times 4 matrix Riemann–Hilbert problem. Then the long-time asymptotics of the solution is obtained by means of the nonlinear steepest descent method of Deift and Zhou for oscillatory Riemann–Hilbert problems. We show that there are three main regions in the half-plane <x<-\infty<x<\infty, t>0t>0, where the asymptotics has qualitatively different forms: a left fast decaying sector, a central Painlevé sector where the asymptotics is described in terms of the solution of a new coupled Painlevé II equation which is related to a 4×44\times 4 matrix Riemann–Hilbert problem, and a right slowly decaying oscillatory sector.

Key words and phrases:
Two-component Sasa–Satsuma equation; Cauchy problem; Long-time asymptotic behavior; Riemann–Hilbert problem; Nonlinear steepest descent method
Corresponding author.

1. Introduction

The one-dimensional cubic NLS (nonlinear Schrödinger) equation

(1.1) iqT+12qXX+|q|2q=0,\text{i}q_{T}+\frac{1}{2}q_{XX}+|q|^{2}q=0,

is a universal model for the evolution of quasi-monochromatic waves in weakly nonlinear dispersive media [6]. It has important applications in many different physical contexts, such as deep water waves, nonlinear fiber optics, acoustics, plasma physics, Bose–Einstein condensation (e.g., see [2, 14, 37, 40] and references therein). One of the most successful related applications is the description of optical solitons in fibres. However, in order to illustrate the propagation of the ultrashort (femtosecond) optical pulses, the NLS equation becomes less accurate [38], and thus some additional effects such as the third-order dispersion, self-steepening and stimulated Raman scattering should be added to meet this requirement. In this setting, Kodama and Hasegawa [23] proposed the higher-order NLS equation

(1.2) iqT+12qXX+|q|2q+iε{β1qXXX+β2|q|2qX+β3q(|q|2)X}=0,\text{i}q_{T}+\frac{1}{2}q_{XX}+|q|^{2}q+\text{i}\varepsilon\left\{\beta_{1}q_{XXX}+\beta_{2}|q|^{2}q_{X}+\beta_{3}q(|q|^{2})_{X}\right\}=0,

where ε\varepsilon is a small parameter and represents the integrable perturbation of the NLS equation, β1\beta_{1}, β2\beta_{2} and β3\beta_{3} are real parameters.

In general, Equation (1.2) is not completely integrable, unless certain restrictions are imposed on β1\beta_{1}, β2\beta_{2} and β3\beta_{3}. In particular, when the ratio of their coefficients satisfies β1:β2:β3=1:6:3\beta_{1}:\beta_{2}:\beta_{3}=1:6:3, the Equation (1.2) can be immediately reduced to the well-known integrable Sasa–Satsuma equation [39] in the form

(1.3) iqT+12qXX+|q|2q+iε{qXXX+6|q|2qX+3q(|q|2)X}=0.\text{i}q_{T}+\frac{1}{2}q_{XX}+|q|^{2}q+\text{i}\varepsilon\left\{q_{XXX}+6|q|^{2}q_{X}+3q(|q|^{2})_{X}\right\}=0.

For the convenience of analyzing the Sasa–Satsuma equation (1.3), according to [39], one can introduce variable transformations

(1.4) u(x,t)=q(X,T)exp{i6ε(XT18ε)},t=T,x=XT12ε,\displaystyle u(x,t)=q(X,T)\exp\left\{-\frac{\text{i}}{6\varepsilon}\left(X-\frac{T}{18\varepsilon}\right)\right\},\ t=T,\ x=X-\frac{T}{12\varepsilon},

then Equation (1.3) reduces to a complex modified KdV (Korteweg–de Vries)-type equation

(1.5) ut+ε{uxxx+6|u|2ux+3u(|u|2)x}=0.u_{t}+\varepsilon\{u_{xxx}+6|u|^{2}u_{x}+3u(|u|^{2})_{x}\}=0.

On account of its integrability and physical implications, the Sasa–Satsuma equation has attracted much attention and various works have been presented since it was discovered. For instance, the double hump soliton solutions of the Sasa–Satsuma equation have been obtained in [34, 39] by means of inverse scattering approach. While, its multi-soliton solutions have been constructed in [19] by the Kadomtsev–Petviashvili hierarchy reduction method. Besides, the Darboux transformation [26] and the RH (Riemann–Hilbert) problem approach [48] were also imposed separately on this equation to obtain the high-order soliton solutions. Moreover, breather and rogue wave solutions for Sasa–Satsuma equation were also derived [4, 13, 35, 44]. In addition to the initial-boundary value problem for Sasa–Satsuma equation on the half-line and a finite interval were also investigated via the unified transform method in [45] and [47], respectively. Beyond that, the long-time asymptotic behaviour of the solution to Sasa–Satsuma equation (1.5) with decaying initial data were analyzed in [27, 29] and [21] respectively in the sectors 0<c1<x/t<c20<c_{1}<x/t<c_{2} and |x|c3t1/3|x|\leq c_{3}t^{1/3} by using the nonlinear steepest descent method for oscillatory Riemann–Hilbert problems. Very recently, data-driven solutions and parameter discovery of the Sasa–Satsuma equation was studied via the physics-informed neural networks method in [32].

Since various complex systems such as multimode or wavelength-division multiplexing fibers usually involve more than one component, the studies should be extended to multi-component Sasa–Satsuma equation cases [16, 28, 36, 41, 50]. Based on this fact, in present paper, we will consider a new integrable two-component Sasa–Satsuma equation [16, 20, 41],

(1.6) ut\displaystyle u_{t} =uxxx+6|u|2ux+3(|u|2)xu+3w(uw)x,\displaystyle=u_{xxx}+6|u|^{2}u_{x}+3(|u|^{2})_{x}u+3w(uw)_{x},
wt\displaystyle w_{t} =wxxx+6|u|2wx+3(|u|2)xw+6w2wx,\displaystyle=w_{xxx}+6|u|^{2}w_{x}+3(|u|^{2})_{x}w+6w^{2}w_{x},

where uu is a complex-valued function, ww is a real-valued function. It is readily to see that when w=0w=0, the new two-component Sasa–Satsuma equation (1.6) can be reduced to the Sasa–Satsuma equation (1.5) with ε=1\varepsilon=-1. On the basis of spectral analysis of the 4×44\times 4 matrix Lax pair for the two-component Sasa–Satsuma equation, the NN-soliton formulas expressed by the ratios of determinants were discussed in [41] via the RH approach, moreover, the traveling soliton, breather soliton and rogue wave solutions have been constructed by Darboux transformation method in [16]. Recently, the initial-boundary value problem of Equation (1.6) on the half-line has been solved with the aid of the unified transformation method [20].

The inverse scattering transform based on the RH problem is a very powerful tool in the study of the nonlinear integrable equations. It can obtain explicit soliton solutions for the integrable systems under reflectionless potentials condition. However, as is well-known, one can not solve the RH problems in a closed form unless in the case of reflectionless potentials. As a consequence, the study on large-time asymptotic behavior of solutions becomes an attractive topic in integrable systems. There were a number of progresses in this formidable subject [3, 22, 49], nevertheless, the nonlinear steepest descent method for oscillatory Riemann–Hilbert problems proposed by Deift and Zhou [15] turned out to be a great achievement in the further development of analyzing the long-time asymptotics for the initial value problems of integrable nonlinear evolution equations. Up to now, with this method, numerous new significant long-time asymptotic results for various nonlinear completely integrable models associated with 2×22\times 2 matrix spectral problems were obtained in a rigorous and transparent form (see [5, 7, 8, 10, 11, 25, 30, 42, 46]). Recently, the long-time asymptotics for some integrable nonlinear evolution equations associated with higher-order matrix Lax pairs were studied in accordance with the procedures of the Deift–Zhou nonlinear steepest descent method, such as Degasperis–Procesi equation [9], coupled NLS equation [17], Sasa–Satsuma equation [21, 27, 29], Spin-1 Gross–Pitaevskii equation [18], matrix modified KdV equation [31], three-component coupled NLS system [33] and so on.

The main goal of the present paper is to extend the nonlinear steepest descent method to study the long-time asymptotic behavior for the Cauchy problem of the new two-component Sasa–Satsuma equation (1.6) associated with a 4×44\times 4 Lax pair on the line with the initial data

(1.7) u(x,0)=u0(x),w(x,0)=w0(x),u(x,0)=u_{0}(x),\quad w(x,0)=w_{0}(x),

where u0(x)u_{0}(x) and w0(x)w_{0}(x) belong to the Schwartz space 𝒮()\mathcal{S}({\mathbb{R}}). Our first step is to formulate the main matrix RH problem corresponding to Cauchy problem (1.6)-(1.7). The most outstanding structure of this system is that it admits a 4×44\times 4 matrix spectral problem, however, all the 4×44\times 4 matrices in this paper can be rewritten as 2×22\times 2 block ones. Thus we can directly formulate the 4×44\times 4 matrix RH problem by the combinations of the entries in matrix-valued eigenfunctions instead of using the Fredholm integral equation to construct another set of eigenfunctions [9, 12]. As a consequence, a RH representation of the solution of the Cauchy problem (1.6)-(1.7) is given (Theorem 2.1). Then, this representation obtained allows us to apply the nonlinear steepest descent method for the associated 4×44\times 4 matrix RH problem and to obtain a detailed description for the leading-order term of the asymptotics of the solution.

We will first consider the asymptotic behavior of the solution in oscillatory sector characterized by (3.1) (Theorem 3.2). It is noted that the phase function Φ(x,t;ξ)\Phi(x,t;\xi) of e±tΦ(x,t;ξ)\text{e}^{\pm t\Phi(x,t;\xi)} involved in the jump matrix has two stationary points in this region. This immediately leads us to introduce a 3×33\times 3 matrix-valued function δ(ξ)\delta(\xi) function to remove the middle matrix term when we split the jump matrix into an appropriate upper/lower triangular form. However, the function δ\delta cannot be solved explicitly since it satisfies a 3×33\times 3 matrix RH problem. Recalling that the topic of our paper is studying the asymptotic behavior of solution, we can replace function δ(ξ)\delta(\xi) with detδ(ξ)\det\delta(\xi) by adding an error term by following the idea first introduced in [17]. Then the exact solution of a class of model RH problem that is relevant near the critical points is derived (Theorem 3.1), which generalizes the model problem considered in [17, 27, 29, 33] and also can be used to analyze the long-time asymptotics of other integrable models. Next, we study the asymptotics of solution in Painlevé sector given in (4.1) (Theorem 4.2). Our main result shows that the leading-order asymptotics for Equation (1.6) is depicted in terms of the solution of a new coupled Painlevé II equation (B.5). Interestingly, we noticed that the functions up(y)u_{p}(y) and wp(y)w_{p}(y) of solution of (B.5) are complex-valued and real-valued functions, respectively, moreover, up(y)u_{p}(y) has constant phase, which is very different from the result obtained for the matrix modified KdV equation in same region [31]. This innovative work enriches the Painlevé asymptotic theory in the field of long-time dynamic analysis for integrable systems. Finally, the asymptotic behavior of the solution in the fast decay sector (5.1) is derived by performing a trivial contour deformation (Theorem 5.1).

The organization of this paper is as follows. In Section 2, a basic 4×44\times 4 matrix RH problem with the aid of the inverse scattering method is constructed, whose solution gives the solution of the initial value problem (1.6)-(1.7), where the Lax pair of the new two-component Sasa–Satsuma equation and the relevant matrices are written as block forms. Sections 3-5 perform the asymptotic analysis of this RH problem leading to asymptotic formulas for the solution. We mainly analyze three regions in the (x,t)(x,t)-half-plane where the asymptotic behavior of the solution is qualitatively different: (i) A slowly decaying oscillatory sector (Theorem 3.2), (ii) A Painlevé sector (Theorem 4.2), (iii) A fast decay sector (Theorem 5.1). Appendix A is devoted to give the proof of Theorem 3.1. The RH problem associated with the new coupled Painlevé II equation is discussed in Appendix B.

2. Basic Riemann–Hilbert problem

An essential ingredient in the following analysis is the 4×44\times 4 matrix Lax pair [41] of the new two-component Sasa–Satsuma equation (1.6), which reads

(2.1) ψx=Uψ=(iξσ+U1)ψ,\displaystyle\psi_{x}=U\psi=(-\text{i}\xi\sigma+U_{1})\psi,
(2.2) ψt=Vψ=(4iξ3σ+V1)ψ,\displaystyle\psi_{t}=V\psi=(4\text{i}\xi^{3}\sigma+V_{1})\psi,

where ψ(x,t;ξ)\psi(x,t;\xi) is a 4×44\times 4 matrix-valued function, ξ\xi is the spectral parameter, σ=diag(1,1,1,1)\sigma=\text{diag}(-1,1,1,1). The 4×44\times 4 matrix-valued functions

(2.3) U1(x,t)=\displaystyle U_{1}(x,t)= (0qq𝟎3×3),q(x,t)=(uuw),\displaystyle\begin{pmatrix}0&q\\ -q^{\dagger}&\mathbf{0}_{3\times 3}\end{pmatrix},\quad q(x,t)=\begin{pmatrix}-u&-u^{*}&-w\end{pmatrix},
(2.4) V1(x,t;ξ)=\displaystyle V_{1}(x,t;\xi)= 4ξ2U1+2iξ(U12+U1x)σ+[U1,U1x]+U1xx2U13.\displaystyle-4\xi^{2}U_{1}+2\text{i}\xi(U_{1}^{2}+U_{1x})\sigma+[U_{1},U_{1x}]+U_{1xx}-2U_{1}^{3}.

where ``"``*" and ``"``{\dagger}" denote complex conjugation of a complex number and Hermitian conjugation of a complex matrix or vector, respectively. A direct calculation shows that the zero-curvature equation UtVx+[U,V]=0U_{t}-V_{x}+[U,V]=0 is equivalent to the new two-component Sasa–Satsuma equation (1.6). On the other hand, it is noted the matrices UU and VV obey the symmetry conditions:

(2.5) U(x,t;ξ)\displaystyle U(x,t;\xi) =U(x,t;ξ),V(x,t;ξ)=V(x,t;ξ),\displaystyle=-U^{\dagger}(x,t;\xi^{*}),\quad V(x,t;\xi)=-V^{\dagger}(x,t;\xi^{*}),
(2.6) U(x,t;ξ)\displaystyle U(x,t;\xi) =𝒜U(x,t;ξ)𝒜,V(x,t;ξ)=𝒜V(x,t;ξ)𝒜,\displaystyle=\mathcal{A}U^{*}(x,t;-\xi^{*})\mathcal{A},\quad V(x,t;\xi)=\mathcal{A}V^{*}(x,t;-\xi^{*})\mathcal{A},

where

(2.7) 𝒜=(1𝟎1×3𝟎3×1σ1),σ1=(010100001).\mathcal{A}=\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \mathbf{0}_{3\times 1}&\sigma_{1}\end{pmatrix},\quad\sigma_{1}=\begin{pmatrix}0&1&0\\ 1&0&0\\ 0&0&1\end{pmatrix}.

Introducing a new eigenfunction μ(x,t;ξ)\mu(x,t;\xi) by

(2.8) μ(x,t;ξ)=ψ(x,t;ξ)eı(ξx4ξ3t)σ,\mu(x,t;\xi)=\psi(x,t;\xi)\text{e}^{\i(\xi x-4\xi^{3}t)\sigma},

we obtain

(2.9) μx=iξ[σ,μ]+U1μ,\displaystyle\mu_{x}=-\text{i}\xi[\sigma,\mu]+U_{1}\mu,
(2.10) μt=4iξ3[σ,μ]+V1μ,\displaystyle\mu_{t}=4\text{i}\xi^{3}[\sigma,\mu]+V_{1}\mu,

We now define two Jost solutions μ±(x,t;ξ)\mu_{\pm}(x,t;\xi) of (2.9) for ξ\xi\in{\mathbb{R}} with μ±(x,t;ξ)𝕀4×4\mu_{\pm}(x,t;\xi)\to\mathbb{I}_{4\times 4} as x±x\to\pm\infty by the following Volterra integral equations

(2.11) μ±(x,t;ξ)=𝕀4×4+±xeiξ(yx)σ[U1(y,t)μ±(y,t;ξ)]eiξ(yx)σdy.\mu_{\pm}(x,t;\xi)=\mathbb{I}_{4\times 4}+\int_{\pm\infty}^{x}\text{e}^{\text{i}\xi(y-x)\sigma}[U_{1}(y,t)\mu_{\pm}(y,t;\xi)]\text{e}^{-\text{i}\xi(y-x)\sigma}\text{d}y.

Denote μ±L(x,t;ξ)\mu_{\pm L}(x,t;\xi) and μ±R(x,t;ξ))\mu_{\pm R}(x,t;\xi)) be the first column and last three columns of the 4×44\times 4 matrices μ±(x,t;ξ)\mu_{\pm}(x,t;\xi). Then, the following are consequences of standard analysis of the iterates that:

(i)(i) μ+L\mu_{+L}, μR\mu_{-R} are analytic in +{\mathbb{C}}_{+} and can be continuously extended to +{\mathbb{C}}_{+}\cup{\mathbb{R}}, (μ+L,μR)𝕀4×4(\mu_{+L},\mu_{-R})\rightarrow\mathbb{I}_{4\times 4} as ξ\xi\rightarrow\infty;

(ii)(ii) μL\mu_{-L}, μ+R\mu_{+R} are analytic in {\mathbb{C}}_{-} and can be continuously extended to {\mathbb{C}}_{-}\cup{\mathbb{R}}, (μL,μ+R)𝕀4×4(\mu_{-L},\mu_{+R})\rightarrow\mathbb{I}_{4\times 4} as ξ\xi\rightarrow\infty.

Since μ±(x,t;ξ)\mu_{\pm}(x,t;\xi) are both fundamental matrices of solutions of (2.9), thus, they satisfy the scattering relation

(2.12) μ+(x,t;ξ)=μ(x,t;ξ)ei(ξx4ξ3t)σs(ξ)ei(ξx4ξ3t)σ,ξ.\mu_{+}(x,t;\xi)=\mu_{-}(x,t;\xi)\text{e}^{-\text{i}(\xi x-4\xi^{3}t)\sigma}s(\xi)\text{e}^{\text{i}(\xi x-4\xi^{3}t)\sigma},\quad\xi\in{\mathbb{R}}.

Evaluation at x,t=0x\rightarrow-\infty,\ t=0 gives

(2.13) s(ξ)=limxeiξxσμ+(x,0;ξ)eiξxσ,s(\xi)=\lim_{x\rightarrow-\infty}\text{e}^{\text{i}\xi x\sigma}\mu_{+}(x,0;\xi)\text{e}^{-\text{i}\xi x\sigma},

that is,

(2.14) s(ξ)=𝕀4×4+eiξxσ[U(x,0)μ+(x,0;ξ)]eiξxσdx.s(\xi)=\mathbb{I}_{4\times 4}-\int_{-\infty}^{+\infty}\text{e}^{\text{i}\xi x\sigma}[U(x,0)\mu_{+}(x,0;\xi)]\text{e}^{-\text{i}\xi x\sigma}\text{d}x.

This implies that the scattering matrix s(ξ)s(\xi) can be determined in terms of the initial values u0(x)u_{0}(x) and w0(x)w_{0}(x).

The symmetries in (2.5) and (2.6) implies that

(2.15) μ±1(x,t;ξ)=μ±(x,t;ξ),μ±(x,t;ξ)=𝒜μ±(x,t;ξ)𝒜.\mu_{\pm}^{-1}(x,t;\xi)=\mu_{\pm}^{\dagger}(x,t;\xi^{*}),\quad\mu_{\pm}(x,t;\xi)=\mathcal{A}\mu_{\pm}^{*}(x,t;-\xi^{*})\mathcal{A}.

Moreover, the tracelessness of U(x,t)U(x,t) shows that det[μ±(x,t;ξ)]=1.\det[\mu_{\pm}(x,t;\xi)]=1. Then (2.12) yields det[s(ξ)]=1.\det[s(\xi)]=1. By (2.15), the 4×44\times 4 matrix-valued spectral function s(k)s(k) obeys the symmetries

(2.16) s1(ξ)=s(ξ),s(ξ)=𝒜s(ξ)𝒜,ξ.s^{-1}(\xi)=s^{\dagger}(\xi^{*}),\quad s(\xi)=\mathcal{A}s^{*}(-\xi^{*})\mathcal{A},\quad\xi\in{\mathbb{R}}.

It follows from the first symmetry in (2.16) that we can write s(ξ)s(\xi) as

(2.17) s(ξ)=(det[a(ξ)]b(ξ)adj[a(ξ)]b(ξ)a(ξ)),ξ,s(\xi)=\begin{pmatrix}\det[a^{\dagger}(\xi^{*})]&b(\xi)\\ -\text{adj}[a^{\dagger}(\xi^{*})]b^{\dagger}(\xi^{*})&a(\xi)\end{pmatrix},\quad\xi\in{\mathbb{R}},

where adj(A)(A) denotes the adjoint matrix of AA in the context of linear algebra, a(ξ)a(\xi) is a 3×33\times 3 matrix-valued function and b(ξ)b(\xi) is a 1×31\times 3 row vector-valued function. On the other hand, it is easy to see from (2.14) that a(ξ)a(\xi) is analytic for ξ\xi\in{\mathbb{C}}_{-}, however, b(ξ)b(\xi) is only defined in {\mathbb{R}}. Furthermore, we also from the second symmetry in (2.16) have

(2.18) a(ξ)=σ1a(ξ)σ1,ξ;b(ξ)=b(ξ)σ1,ξ.a(\xi)=\sigma_{1}a^{*}(-\xi^{*})\sigma_{1},\quad\xi\in{\mathbb{C}}_{-};\quad b(\xi)=b^{*}(-\xi^{*})\sigma_{1},\quad\xi\in{\mathbb{R}}.

To exclude soliton-type phenomena, for convenience, we assume that det[a(ξ)]\det[a(\xi)] has no zeros in {\mathbb{C}}_{-}. Then, we have the following main result in this section, which shows how solutions of (1.6) can be constructed by a basic 4×44\times 4 matrix Riemann–Hilbert problem.

Theorem 2.1.

Define a piecewise meromorphic 4×44\times 4 matrix-valued function

(2.19) M(x,t;ξ)={(μ+L(x,t;ξ)det[a(ξ)]1,μR(x,t;ξ)),ξ+,(μL(x,t;ξ),μ+R(x,t;ξ)a1(ξ)),ξ,M(x,t;\xi)=\left\{\begin{aligned} &\left(\mu_{+L}(x,t;\xi)\det[a^{\dagger}(\xi^{*})]^{-1},\mu_{-R}(x,t;\xi)\right),\quad\xi\in{\mathbb{C}}_{+},\\ &\left(\mu_{-L}(x,t;\xi),\mu_{+R}(x,t;\xi)a^{-1}(\xi)\right),\qquad\qquad\xi\in{\mathbb{C}}_{-},\end{aligned}\right.

and the 4×44\times 4 matrix-valued jump matrix J(x,t;ξ)J(x,t;\xi) by

(2.20) J(x,t;ξ)=(1+γ(ξ)γ(ξ)γ(ξ)e2i(ξx4ξ3t)γ(ξ)e2i(ξx4ξ3t)𝕀3×3),J(x,t;\xi)=\begin{pmatrix}1+\gamma(\xi)\gamma^{\dagger}(\xi)&\gamma(\xi)\text{e}^{2\text{i}(\xi x-4\xi^{3}t)}\\ \gamma^{\dagger}(\xi)\text{e}^{-2\text{i}(\xi x-4\xi^{3}t)}&\mathbb{I}_{3\times 3}\end{pmatrix},

where

(2.21) γ(ξ)=b(ξ)a1(ξ),γ(ξ)=γ(ξ)σ1,ξ.\gamma(\xi)=-b(\xi)a^{-1}(\xi),\quad\gamma(\xi)=\gamma^{*}(-\xi^{*})\sigma_{1},\quad\xi\in{\mathbb{R}}.

Then the following 4×44\times 4 matrix RH problem:
\bullet M(x,t;ξ)M(x,t;\xi) is a sectionally meromorphic function with respect to {\mathbb{R}};
\bullet The limiting values M±(x,t;ξ)=limε0M(x,t;ξ±iε)M_{\pm}(x,t;\xi)=\lim_{\varepsilon\to 0}M(x,t;\xi\pm\text{i}\varepsilon) satisfy the jump condition M+(x,t;ξ)=M(x,t;ξ)J(x,t;ξ)M_{+}(x,t;\xi)=M_{-}(x,t;\xi)J(x,t;\xi) for ξ\xi\in{\mathbb{R}};
\bullet As ξ\xi\rightarrow\infty, M(x,t;ξ)M(x,t;\xi) has the asymptotics: M(x,t;ξ)=𝕀4×4+O(ξ1)M(x,t;\xi)=\mathbb{I}_{4\times 4}+O\left(\xi^{-1}\right);
has a unique solution for each (x,t)2(x,t)\in{\mathbb{R}}^{2}.

Moreover, define {u(x,t),w(x,t)}\{u(x,t),w(x,t)\} in terms of M(x,t;ξ)M(x,t;\xi) by

(2.22) (u(x,t)u(x,t)w(x,t))=2ilimξ(ξM(x,t;ξ))12,\displaystyle\begin{pmatrix}u(x,t)&u^{*}(x,t)&w(x,t)\end{pmatrix}=2\text{i}\lim_{\xi\rightarrow\infty}(\xi M(x,t;\xi))_{12},

which solves the Cauchy problem of new two-component Sasa–Satsuma equation (1.6).

Proof.

It is a simple consequence of Liouville’s theorem that if a solution exists, it is unique. The existence of solution of RH problem follows by means of Zhou’s vanishing lemma argument [51] since

(2.23) J(x,t;ξ)=J(x,t;ξ)=𝒜J(x,t;ξ)𝒜,ξ.J(x,t;\xi)=J^{\dagger}(x,t;\xi^{*})=\mathcal{A}J^{*}(x,t;-\xi^{*})\mathcal{A},\quad\xi\in{\mathbb{R}}.

Expanding this solution as ξ\xi\to\infty,

(2.24) M(x,t;ξ)=𝕀4×4+M1(x,t)ξ+M2(x,t)ξ2+O(ξ3)M(x,t;\xi)=\mathbb{I}_{4\times 4}+\frac{M_{1}(x,t)}{\xi}+\frac{M_{2}(x,t)}{\xi^{2}}+O(\xi^{-3})

and inserting this into equation (2.9) one finds that the solution of (1.6) is given by (2.22).

3. Asymptotic analysis in oscillating sector 𝒪\mathcal{O}

The representation of the solution of the Cauchy problem for a nonlinear integrable equation in terms of the solution of an associated RH problem makes it possible to analyze the long-time asymptotics via the Deift–Zhou steepest descent method [15]. The goal of this section is devoted to deriving the long-time asymptotic behavior of solution for the new two-component Sasa–Satsuma equation (1.6) in oscillating sector 𝒪\mathcal{O} defined by

(3.1) 𝒪={(x,t)2|t>1, 0<xNt,x3/t},Nconstant.\mathcal{O}=\{(x,t)\in{\mathbb{R}}^{2}|t>1,\ 0<x\leq Nt,\ x^{3}/t\to\infty\},\,N\,\text{constant}.

We first prove an important result (Theorem 3.1), which expresses the large zz behavior of solution of a model RH problem in terms of the solution of parabolic cylinder functions and will be very useful in the study of long-time asymptotics in the oscillating sector.

3.1. A model RH problem

Refer to caption
Figure 1. The oriented contour X=j=14XjX=\cup_{j=1}^{4}X_{j}.

Define the oriented contour X=j=14XjX=\cup_{j=1}^{4}X_{j} by

(3.2) X1\displaystyle X_{1} ={ϵeiπ4|0ϵ<},X2={ϵe3iπ4|0ϵ<},\displaystyle=\{\epsilon\text{e}^{-\frac{\text{i}\pi}{4}}|0\leq\epsilon<\infty\},\,\,\,\,\ X_{2}=\{\epsilon\text{e}^{\frac{3\text{i}\pi}{4}}|0\leq\epsilon<\infty\},
X3\displaystyle X_{3} ={ϵe3iπ4|0ϵ<},X4={ϵeiπ4|0ϵ<},\displaystyle=\{\epsilon\text{e}^{-\frac{3\text{i}\pi}{4}}|0\leq\epsilon<\infty\},\,\,\ X_{4}=\{\epsilon\text{e}^{\frac{\text{i}\pi}{4}}|0\leq\epsilon<\infty\},

see Figure 1. For a 1×31\times 3 complex-valued row vector ρ\rho, define the function ν\nu by ν(ρ)=12πln(1+ρρ)>0\nu(\rho)=\frac{1}{2\pi}\ln(1+\rho\rho^{\dagger})>0 and the 4×44\times 4 jump matrix by

(3.3) JX(ρ;z)={(1ρeiz22z2iν(ρ)03×1𝕀3×3),zX1,(1ρ1+ρρeiz22z2iν(ρ)03×1𝕀3×3),zX2,(101×3ρ1+ρρeiz22z2iν(ρ)𝕀3×3),zX3,(101×3ρeiz22z2iν(ρ)𝕀3×3),zX4.J^{X}(\rho;z)=\left\{\begin{aligned} &\begin{pmatrix}1&\rho\text{e}^{-\frac{\text{i}z^{2}}{2}}z^{-2\text{i}\nu(\rho)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad\quad\,\ z\in X_{1},\\ &\begin{pmatrix}1&-\frac{\rho}{1+\rho\rho^{\dagger}}\text{e}^{-\frac{\text{i}z^{2}}{2}}z^{-2\text{i}\nu(\rho)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\,\,\,\ z\in X_{2},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ -\frac{\rho^{\dagger}}{1+\rho\rho^{\dagger}}\text{e}^{\frac{\text{i}z^{2}}{2}}z^{2\text{i}\nu(\rho)}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad z\in X_{3},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \rho^{\dagger}\text{e}^{\frac{\text{i}z^{2}}{2}}z^{2\text{i}\nu(\rho)}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad\qquad z\in X_{4}.\end{aligned}\right.

We consider the following RH problem:
\bullet MX(ρ;z)M^{X}(\rho;z) is analytic for zXz\in{\mathbb{C}}\setminus X and extends continuously to XX;
\bullet Across XX, the boundary values M±XM^{X}_{\pm} satisfy the jump relation M+X(ρ;z)=MX(ρ;z)JX(ρ;z)M^{X}_{+}(\rho;z)=M^{X}_{-}(\rho;z)J^{X}(\rho;z);
\bullet MX(ρ;z)𝕀4×4M^{X}(\rho;z)\rightarrow\mathbb{I}_{4\times 4}, as zz\rightarrow\infty.

Theorem 3.1.

The RH problem has a unique solution MX(ρ;z)M^{X}(\rho;z) for each 1×31\times 3 row vector ρ\rho, and this solution satisfies

(3.4) MX(ρ;z)=𝕀4×4+M1X(ρ)z+O(z2),z,M^{X}(\rho;z)=\mathbb{I}_{4\times 4}+\frac{M^{X}_{1}(\rho)}{z}+O\left(z^{-2}\right),\quad z\rightarrow\infty,

where

(3.5) (M1X)12(ρ)=iβX,(M1X)21(ρ)=i(βX),βX=νΓ(iν)eiπ4πν22πρ,\left(M^{X}_{1}\right)_{12}(\rho)=-\text{i}\beta^{X},\,\left(M^{X}_{1}\right)_{21}(\rho)=-\text{i}\left(\beta^{X}\right)^{\dagger},\,\beta^{X}=\frac{\nu\Gamma(-\text{i}\nu)\text{e}^{\frac{\text{i}\pi}{4}-\frac{\pi\nu}{2}}}{\sqrt{2\pi}}\rho,

where Γ()\Gamma(\cdot) denotes the standard Gamma function. Moreover, for each compact subset 𝒟\mathcal{D} of {\mathbb{C}},

(3.6) supρ𝒟supzX|MX(ρ;z)|<.\sup_{\rho\in\mathcal{D}}\sup_{z\in{\mathbb{C}}\setminus X}|M^{X}(\rho;z)|<\infty.
Proof.

See Appendix A. ∎

3.2. Transformations

The Deift–Zhou nonlinear steepest descent method for RH problems consists of making a series of invertible transformations in order to arrive at a problem that can be approximated in the large-tt limit. For this purpose, we first note that the jump matrix J(x,t;ξ)J(x,t;\xi) defined in (2.20) involves the exponentials e±tΦ(x,t;ξ)\text{e}^{\pm t\Phi(x,t;\xi)}, where Φ(x,t;ξ)\Phi(x,t;\xi) is given by

(3.7) Φ(x,t;ξ)=8iξ32iξxt.\Phi(x,t;\xi)=8\text{i}\xi^{3}-2\text{i}\xi\frac{x}{t}.

Suppose (x,t)𝒪(x,t)\in\mathcal{O}. By solving the equation Φ(x,t;ξ)/ξ=0,\partial\Phi(x,t;\xi)/\partial\xi=0, we see that there are two real critical points located at

(3.8) ±ξ0=±x12t,\pm\xi_{0}=\pm\sqrt{\frac{x}{12t}},

moreover, the signature table for ReΦ(x,t;ξ)\Phi(x,t;\xi) is shown in Figure 2.

Refer to caption
Figure 2. The signature table for ReΦ(x,t;ξ)\Phi(x,t;\xi) and ±ξ0\pm\xi_{0}.

The jump matrix J(x,t;ξ)J(x,t;\xi) enjoys two distinct factorizations:

(3.9) J(x,t;ξ)={(1γetΦ𝟎3×1𝕀3×3)(1𝟎1×3γetΦ𝕀3×3),(1𝟎1×3γetΦ1+γγ𝕀3×3)(1+γγ𝟎1×3𝟎3×1(𝕀3×3+γγ)1)(1γetΦ1+γγ𝟎3×1𝕀3×3).J(x,t;\xi)=\left\{\begin{aligned} &\begin{pmatrix}1&\gamma\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \gamma^{\dagger}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\\ &\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \frac{\gamma^{\dagger}\text{e}^{t\Phi}}{1+\gamma\gamma^{\dagger}}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1+\gamma\gamma^{\dagger}&\mathbf{0}_{1\times 3}\\ \mathbf{0}_{3\times 1}&(\mathbb{I}_{3\times 3}+\gamma^{\dagger}\gamma)^{-1}\end{pmatrix}\begin{pmatrix}1&\frac{\gamma\text{e}^{-t\Phi}}{1+\gamma\gamma^{\dagger}}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}.\end{aligned}\right.

Thus, by signature table in Figure 2, we can see that the jump matrix has the wrong factorization for ξ(ξ0,ξ0)\xi\in(-\xi_{0},\xi_{0}). Hence the first transformation is to introduce M(1)M^{(1)} by

(3.10) M(1)(x,t;ξ)=M(x,t;ξ)([detδ(ξ)]1𝟎1×3𝟎3×1δ(ξ)),M^{(1)}(x,t;\xi)=M(x,t;\xi)\begin{pmatrix}[\det\delta(\xi)]^{-1}&\mathbf{0}_{1\times 3}\\ \mathbf{0}_{3\times 1}&\delta(\xi)\end{pmatrix},

where the 3×33\times 3 matrix-valued function satisfies the following RH problem:
\bullet δ(ξ)\delta(\xi) is analytic for ξ[ξ0,ξ0]\xi\in{\mathbb{C}}\setminus[-\xi_{0},\xi_{0}], and it takes continuous boundary values on (ξ0,ξ0)(-\xi_{0},\xi_{0}) from the upper and lower half-planes;
\bullet The boundary values on the jump contour (ξ0,ξ0)(-\xi_{0},\xi_{0}) (oriented to right) are related as

(3.11) δ+(ξ)=(𝕀3×3+γ(ξ)γ(ξ))δ(ξ),ξ(ξ0,ξ0);\delta_{+}(\xi)=(\mathbb{I}_{3\times 3}+\gamma^{\dagger}(\xi)\gamma(\xi))\delta_{-}(\xi),\quad\xi\in(-\xi_{0},\xi_{0});

\bullet δ(ξ)=𝕀3×3+O(ξ1)\delta(\xi)=\mathbb{I}_{3\times 3}+O(\xi^{-1}) as ξ\xi\to\infty.
Since the jump matrix 𝕀3×3+γ(ξ)γ(ξ)\mathbb{I}_{3\times 3}+\gamma^{\dagger}(\xi)\gamma(\xi) is positive definite, the vanishing lemma of Zhou [51] yields the existence and uniqueness of the function δ(ξ)\delta(\xi). Furthermore, it is easy to see that detδ(ξ)\det\delta(\xi) obeys a scalar RH problem:
\bullet detδ(ξ)\det\delta(\xi) is analytic for ξ[ξ0,ξ0]\xi\in{\mathbb{C}}\setminus[-\xi_{0},\xi_{0}];
\bullet detδ(ξ)\det\delta(\xi) takes continuous boundary values detδ±(ξ)\det\delta_{\pm}(\xi) on (ξ0,ξ0)(-\xi_{0},\xi_{0}), and they are related by the jump condition

(3.12) detδ+(ξ)=detδ(ξ)(1+γ(ξ)γ(ξ)),ξ(ξ0,ξ0);\det\delta_{+}(\xi)=\det\delta_{-}(\xi)(1+\gamma(\xi)\gamma^{\dagger}(\xi)),\quad\xi\in(-\xi_{0},\xi_{0});

\bullet limξdetδ(ξ)=1\lim_{\xi\to\infty}\det\delta(\xi)=1.

Proposition 3.1.

The functions δ(ξ)\delta(\xi) and detδ(ξ)\det\delta(\xi) have the following properties:

(3.13) δ(ξ)=[δ(ξ)]1=σ1δ(ξ)σ1,detδ(ξ)=[(detδ(ξ))]1=[detδ(ξ)],\delta(\xi)=[\delta^{\dagger}(\xi^{*})]^{-1}=\sigma_{1}\delta^{*}(-\xi^{*})\sigma_{1},\quad\det\delta(\xi)=[(\det\delta(\xi^{*}))^{*}]^{-1}=[\det\delta(-\xi^{*})]^{*},

and

(3.14) |δ(ξ)|,|detδ(ξ)|const<,|\delta(\xi)|,|\det\delta(\xi)|\leq\text{const}<\infty,

where |X|2=tr(XX)|X|^{2}=\text{tr}(X^{\dagger}X) for any matrix XX.

Then M(1)M^{(1)} satisfies the following matrix RH problem:
\bullet M(1)(x,t;ξ)M^{(1)}(x,t;\xi) is analytic for ξ\xi\in{\mathbb{C}}\setminus{\mathbb{R}}, and it takes continuous boundary values on {\mathbb{R}} from the upper and lower half-planes;
\bullet The boundary values M±(1)(x,t;ξ)M^{(1)}_{\pm}(x,t;\xi) on the jump contour {\mathbb{R}} are related as

(3.15) M+(1)(x,t;ξ)=M(1)(x,t;ξ)J(1)(x,t;ξ),ξ,M^{(1)}_{+}(x,t;\xi)=M^{(1)}_{-}(x,t;\xi)J^{(1)}(x,t;\xi),\quad\xi\in{\mathbb{R}},

where the jump matrix is given by

(3.16) J(1)(x,t;ξ)={(1γδdetδetΦ𝟎3×1𝕀3×3)(1𝟎1×3δ1γ(detδ)1etΦ𝕀3×3),|ξ|>ξ0,(1𝟎1×3δ1γ1+γγ(detδ)1etΦ𝕀3×3)(1δ+γ1+γγdetδ+etΦ𝟎3×1𝕀3×3),|ξ|<ξ0;\displaystyle J^{(1)}(x,t;\xi)=\left\{\begin{aligned} &\begin{pmatrix}1&\gamma\delta\det\delta\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta^{-1}\gamma^{\dagger}(\det\delta)^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad\qquad\quad\,\ |\xi|>\xi_{0},\\ &\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta_{-}^{-1}\frac{\gamma^{\dagger}}{1+\gamma\gamma^{\dagger}}(\det\delta_{-})^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\\ \end{pmatrix}\begin{pmatrix}1&\delta_{+}\frac{\gamma}{1+\gamma\gamma^{\dagger}}\det\delta_{+}\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\ |\xi|<\xi_{0};\end{aligned}\right.

\bullet M(1)(x,t;ξ)=𝕀4×4+O(ξ1)M^{(1)}(x,t;\xi)=\mathbb{I}_{4\times 4}+O(\xi^{-1}) as ξ\xi\to\infty.

Refer to caption
Figure 3. The jump contour Σ\Sigma (solid line) and the open sets {Dj}16\{D_{j}\}_{1}^{6}.

The next step is to deform the contour such that the jump matrix involves the exponential factor etΦ\text{e}^{t\Phi} on the parts of the contour where ReΦ\Phi is negative, the factor etΦ\text{e}^{-t\Phi} on the parts where ReΦ\Phi is positive and the jumps on the original contour {\mathbb{R}} are small remainders with respect to tt. To achieve this goal, we should introduce the analytic approximations of γ(ξ)\gamma(\xi) and γ(ξ)1+γ(ξ)γ(ξ)\frac{\gamma^{\dagger}(\xi)}{1+\gamma(\xi)\gamma^{\dagger}(\xi)}. The symmetry of γ(ξ)\gamma(\xi) in (2.21) implies that we can rewrite

(3.17) γ(ξ)=\displaystyle\gamma(\xi)= (γ1(ξ)γ1(ξ)γ2(ξ)),ξ,\displaystyle\begin{pmatrix}\gamma_{1}(\xi)&\gamma_{1}^{*}(-\xi^{*})&\gamma_{2}(\xi)\end{pmatrix},\ \xi\in{\mathbb{R}},
(3.18) γ~(ξ)\displaystyle\tilde{\gamma}^{\dagger}(\xi)\doteq γ(ξ)1+γ(ξ)γ(ξ)=(γ3(ξ)γ3(ξ)γ4(ξ)),ξ.\displaystyle\frac{\gamma^{\dagger}(\xi)}{1+\gamma(\xi)\gamma^{\dagger}(\xi)}=\begin{pmatrix}\gamma_{3}(\xi)&\gamma_{3}^{*}(-\xi^{*})&\gamma_{4}(\xi)\end{pmatrix},\ \xi\in{\mathbb{R}}.

Let DjDj(ξ)D_{j}\doteq D_{j}(\xi), j=1,,6j=1,\cdots,6 denote the open subsets of {\mathbb{C}} displayed in Figure 3, then we have the following lemma.

Lemma 3.1.

There exist decompositions

(3.19) γ1(ξ)=\displaystyle\gamma_{1}(\xi)= γ1,a(x,t;ξ)+γ1,r(x,t;ξ),|ξ|>ξ0,ξ,\displaystyle\gamma_{1,a}(x,t;\xi)+\gamma_{1,r}(x,t;\xi),\quad|\xi|>\xi_{0},~{}\xi\in{\mathbb{R}},
(3.20) γ3(ξ)=\displaystyle\gamma_{3}(\xi)= γ3,a(x,t;ξ)+γ3,r(x,t;ξ),|ξ|<ξ0,ξ,\displaystyle\gamma_{3,a}(x,t;\xi)+\gamma_{3,r}(x,t;\xi),\quad|\xi|<\xi_{0},~{}\xi\in{\mathbb{R}},

where the functions γj,a\gamma_{j,a} and γj,r\gamma_{j,r}, j=1,3j=1,3 have the following properties:

(1) For each (x,t)𝒪(x,t)\in\mathcal{O} and j=1,3j=1,3, γj,a(x,t;ξ)\gamma_{j,a}(x,t;\xi)is defined and continuous for ξD¯j\xi\in\bar{D}_{j} and analytic in DjD_{j}.

(2) The function γj,a\gamma_{j,a} satisfies the following estimates

(3.21) |γj,a(x,t;ξ)γj(ξ0)|C|ξξ0|et4|ReΦ(x,t;ξ)|,ξD¯j,j=1,3,|\gamma_{j,a}(x,t;\xi)-\gamma_{j}(\xi_{0})|\leq C|\xi-\xi_{0}|\text{e}^{\frac{t}{4}|\text{Re}\Phi(x,t;\xi)|},~{}\xi\in\bar{D}_{j},\ j=1,3,

and

(3.22) |γ1,a(x,t;ξ)|C1+|ξ|2et4|ReΦ(x,t;ξ)|,ξD¯1.|\gamma_{1,a}(x,t;\xi)|\leq\frac{C}{1+|\xi|^{2}}\text{e}^{\frac{t}{4}|\text{Re}\Phi(x,t;\xi)|},~{}\xi\in\bar{D}_{1}.

(3) The L1,L2L^{1},L^{2} and LL^{\infty} norms of the function γ1,r(x,t;)\gamma_{1,r}(x,t;\cdot) on (,ξ0)(ξ0,)(-\infty,-\xi_{0})\cup(\xi_{0},\infty) are O(t3/2)O(t^{-3/2}) as tt\rightarrow\infty uniformly with respect to (x,t)𝒪(x,t)\in\mathcal{O}.

(4) The L1,L2L^{1},L^{2} and LL^{\infty} norms of the function γ3,r(x,t;)\gamma_{3,r}(x,t;\cdot) on (ξ0,ξ0)(-\xi_{0},\xi_{0}) are O(t3/2)O(t^{-3/2}) as tt\rightarrow\infty uniformly with respect to (x,t)𝒪(x,t)\in\mathcal{O}.

Proof.

See Lemma 4.8 in [24]. ∎

For l=2,4l=2,4, the decomposition of γl(ξ)=γl,a(x,t;ξ)+γl,r(x,t;ξ)\gamma_{l}(\xi)=\gamma_{l,a}(x,t;\xi)+\gamma_{l,r}(x,t;\xi) can be similarly found. Thus, we establish the decompositions γ(ξ)=γa(x,t;ξ)+γr(x,t;ξ)\gamma(\xi)=\gamma_{a}(x,t;\xi)+\gamma_{r}(x,t;\xi) of γ(ξ)\gamma(\xi), and γ~(ξ)=γ~a(x,t;ξ)+γ~r(x,t;ξ)\tilde{\gamma}^{\dagger}(\xi)=\tilde{\gamma}^{\dagger}_{a}(x,t;\xi)+\tilde{\gamma}^{\dagger}_{r}(x,t;\xi) of γ~(ξ)\tilde{\gamma}^{\dagger}(\xi) by setting

γa(x,t;ξ)=\displaystyle\gamma_{a}(x,t;\xi)= (γ1,a(x,t;ξ)γ1,a(x,t;ξ)γ2,a(x,t;ξ)),\displaystyle\begin{pmatrix}\gamma_{1,a}(x,t;\xi)&\gamma_{1,a}^{*}(x,t;-\xi^{*})&\gamma_{2,a}(x,t;\xi)\end{pmatrix},
γr(x,t;ξ)=\displaystyle\gamma_{r}(x,t;\xi)= (γ1,r(x,t;ξ)γ1,r(x,t;ξ)γ2,r(x,t;ξ)),\displaystyle\begin{pmatrix}\gamma_{1,r}(x,t;\xi)&\gamma_{1,r}^{*}(x,t;-\xi^{*})&\gamma_{2,r}(x,t;\xi)\end{pmatrix},
γ~a(x,t;ξ)=\displaystyle\tilde{\gamma}_{a}^{\dagger}(x,t;\xi)= (γ~3,a(x,t;ξ)γ~3,a(x,t;ξ)γ~4,a(x,t;ξ)),\displaystyle\begin{pmatrix}\tilde{\gamma}_{3,a}(x,t;\xi)&\tilde{\gamma}_{3,a}^{*}(x,t;-\xi^{*})&\tilde{\gamma}_{4,a}(x,t;\xi)\end{pmatrix},
γ~r(x,t;ξ)=\displaystyle\tilde{\gamma}^{\dagger}_{r}(x,t;\xi)= (γ~3,r(x,t;ξ)γ~3,r(x,t;ξ)γ~4,r(x,t;ξ)).\displaystyle\begin{pmatrix}\tilde{\gamma}_{3,r}(x,t;\xi)&\tilde{\gamma}_{3,r}^{*}(x,t;-\xi^{*})&\tilde{\gamma}_{4,r}(x,t;\xi)\end{pmatrix}.

Now, we can introduce M(2)(x,t;ξ)M^{(2)}(x,t;\xi) by

(3.23) M(2)(x,t;ξ)=M(1)(x,t;ξ)G(ξ),M^{(2)}(x,t;\xi)=M^{(1)}(x,t;\xi)G(\xi),

where the sectionally analytic function GG is defined by

(3.24) G(ξ)={(1γaδdetδetΦ𝟎3×1𝕀3×3),ξD1,(1δ+γ~adetδ+etΦ𝟎3×1𝕀3×3),ξD2,(1𝟎1×3δ1γ~a(detδ)1etΦ𝕀3×3),ξD3,(1𝟎1×3δ1γa(detδ)1etΦ𝕀3×3),ξD4,𝕀4×4,ξD5D6.G(\xi)=\left\{\begin{aligned} &\begin{pmatrix}1&\gamma_{a}\delta\det\delta\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad\quad\,\,\ \xi\in D_{1},\\ &\begin{pmatrix}1&-\delta_{+}\tilde{\gamma}_{a}\det\delta_{+}\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\,\,\,\ \xi\in D_{2},\\ &\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta_{-}^{-1}\tilde{\gamma}_{a}^{\dagger}(\det\delta_{-})^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\ \xi\in D_{3},\\ &\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ -\delta^{-1}\gamma^{\dagger}_{a}(\det\delta)^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in D_{4},\\ &\mathbb{I}_{4\times 4},\qquad\qquad\qquad\qquad\qquad\qquad\,\,\ \xi\in D_{5}\cup D_{6}.\end{aligned}\right.

Let Σ\Sigma\subset{\mathbb{C}} denote the contour displayed in Figure 3. It follows that M(2)(x,t;ξ)M^{(2)}(x,t;\xi) satisfies the following RH problem:
\bullet M(2)(x,t;ξ)M^{(2)}(x,t;\xi) is analytic off Σ\Sigma, and it takes continuous boundary values on Σ\Sigma;
\bullet Across the oriented contour Σ\Sigma, the boundary values M±(2)(x,t;ξ)M^{(2)}_{\pm}(x,t;\xi) are connected by the following formula:

(3.25) M+(2)(x,t;ξ)=M(2)(x,t;ξ)J(2)(x,t;ξ),ξΣ;M^{(2)}_{+}(x,t;\xi)=M^{(2)}_{-}(x,t;\xi)J^{(2)}(x,t;\xi),\quad\xi\in\Sigma;

\bullet M(2)(x,t;ξ)𝕀4×4M^{(2)}(x,t;\xi)\to\mathbb{I}_{4\times 4}, as ξ\xi\to\infty;
where the jump matrix J(2)(x,t;ξ)J^{(2)}(x,t;\xi) is given by

(3.26) J1(2)=\displaystyle J^{(2)}_{1}= (1γaδdetδetΦ𝟎3×1𝕀3×3),J2(2)=(1δ+γ~adetδ+etΦ𝟎3×1𝕀3×3),\displaystyle\begin{pmatrix}1&\gamma_{a}\delta\det\delta\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\ J^{(2)}_{2}=\begin{pmatrix}1&-\delta_{+}\tilde{\gamma}_{a}\det\delta_{+}\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},
J3(2)=\displaystyle J^{(2)}_{3}= (1𝟎1×3δ1γ~a(detδ)1etΦ𝕀3×3),J4(2)=(1𝟎1×3δ1γa(detδ)1etΦ𝕀3×3),\displaystyle\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ -\delta_{-}^{-1}\tilde{\gamma}_{a}^{\dagger}(\det\delta_{-})^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\ J^{(2)}_{4}=\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta^{-1}\gamma^{\dagger}_{a}(\det\delta)^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},
J5(2)=\displaystyle J^{(2)}_{5}= (1𝟎1×3δ1γ~r(detδ)1etΦ𝕀3×3)(1δ+γ~rdetδ+etΦ𝟎3×1𝕀3×3),\displaystyle\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta_{-}^{-1}\tilde{\gamma}_{r}^{\dagger}(\det\delta_{-})^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\\ \end{pmatrix}\begin{pmatrix}1&\delta_{+}\tilde{\gamma}_{r}\det\delta_{+}\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},
J6(2)=\displaystyle J^{(2)}_{6}= (1γrδdetδetΦ𝟎3×1𝕀3×3)(1𝟎1×3δ1γr(detδ)1etΦ𝕀3×3),\displaystyle\begin{pmatrix}1&\gamma_{r}\delta\det\delta\text{e}^{-t\Phi}\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\mathbf{0}_{1\times 3}\\ \delta^{-1}\gamma^{\dagger}_{r}(\det\delta)^{-1}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},

with Jj(2)J^{(2)}_{j} denoting the restriction of J(2)J^{(2)} to the contour labeled by jj in Figure 3.

3.3. Local models

Obviously, as tt\rightarrow\infty, the matrix J(2)𝕀4×4J^{(2)}-\mathbb{I}_{4\times 4} decays to zero everywhere except near the critical points ±ξ0\pm\xi_{0}. This implies that the main contribution to the long-time asymptotics of M(2)M^{(2)} should come from the neighborhoods of ±ξ0\pm\xi_{0}.

In order to relate M(2)M^{(2)} to the solution MXM^{X} of the model RH problem in Theorem 3.1, we introduce the following scaling transforms for ξ\xi near ±ξ0\pm\xi_{0}

(3.27) Tξ0:ξz48tξ0ξ0,\displaystyle T_{-\xi_{0}}:~{}\xi\mapsto\frac{z}{\sqrt{48t\xi_{0}}}-\xi_{0},
(3.28) Tξ0:ξz48tξ0+ξ0.\displaystyle T_{\xi_{0}}:~{}\xi\mapsto\frac{z}{\sqrt{48t\xi_{0}}}+\xi_{0}.

Observe that the 3×33\times 3 matrix-valued function δ(ξ)\delta(\xi) can not be expressed explicitly, in order to proceed to the next step, we write, for example,

(3.29) (γaδdetδetΦ)(ξ)=(γa(δdetδ𝕀3×3)detδetΦ)(ξ)+(γa(detδ)2etΦ)(ξ).\left(\gamma_{a}\delta\det\delta\text{e}^{-t\Phi}\right)(\xi)=\left(\gamma_{a}(\delta-\det\delta\mathbb{I}_{3\times 3})\det\delta\text{e}^{-t\Phi}\right)(\xi)+\left(\gamma_{a}(\det\delta)^{2}\text{e}^{-t\Phi}\right)(\xi).

For the second term in the right-hand side of (3.29), by Plemelj formula, we know that

(3.30) detδ(ξ)=exp{12πiξ0ξ0ln(1+γ(s)γ(s))sξds}=(ξξ0ξ+ξ0)iν(ξ0)eχ(ξ),\det\delta(\xi)=\exp\left\{\frac{1}{2\pi\text{i}}\int_{-\xi_{0}}^{\xi_{0}}\frac{\ln(1+\gamma(s)\gamma^{\dagger}(s))}{s-\xi}\text{d}s\right\}=\left(\frac{\xi-\xi_{0}}{\xi+\xi_{0}}\right)^{-\text{i}\nu(\xi_{0})}\text{e}^{\chi(\xi)},

where

(3.31) ν(ξ0)=\displaystyle\nu(\xi_{0})= 12πln(1+γ(ξ0)γ(ξ0))>0,\displaystyle\frac{1}{2\pi}\ln(1+\gamma(\xi_{0})\gamma^{\dagger}(\xi_{0}))>0,
(3.32) χ(ξ)=\displaystyle\chi(\xi)= 12πiξ0ξ0ln(1+γ(s)γ(s)1+γ(ξ0)γ(ξ0))dssξ,\displaystyle\frac{1}{2\pi\text{i}}\int_{-\xi_{0}}^{\xi_{0}}\ln\left(\frac{1+\gamma(s)\gamma^{\dagger}(s)}{1+\gamma(\xi_{0})\gamma^{\dagger}(\xi_{0})}\right)\frac{\text{d}s}{s-\xi},

thus, a direct computation yields that

(3.33) Tξ0(γa(detδ)2etΦ)(ξ)=ϑ2υ2γa(z48tξ0+ξ0),\displaystyle T_{\xi_{0}}\left(\gamma_{a}(\det\delta)^{2}\text{e}^{-t\Phi}\right)(\xi)=\vartheta^{2}\upsilon^{2}\gamma_{a}\left(\frac{z}{\sqrt{48t\xi_{0}}}+\xi_{0}\right),

where

(3.34) ϑ\displaystyle\vartheta =(192tξ03)iν2e8itξ03+χ(ξ0),\displaystyle=(192t\xi_{0}^{3})^{\frac{\text{i}\nu}{2}}\text{e}^{8\text{i}t\xi_{0}^{3}+\chi(\xi_{0})},
(3.35) υ\displaystyle\upsilon =ziνeiz24(1+z/432tξ03)(2ξ0z/48tξ0+2ξ0)iνeχ([z/48tξ0]+ξ0)χ(ξ0).\displaystyle=z^{-\text{i}\nu}\text{e}^{-\frac{\text{i}z^{2}}{4}(1+z/\sqrt{432t\xi_{0}^{3}})}\left(\frac{2\xi_{0}}{z/\sqrt{48t\xi_{0}}+2\xi_{0}}\right)^{-\text{i}\nu}\text{e}^{\chi([z/\sqrt{48t\xi_{0}}]+\xi_{0})-\chi(\xi_{0})}.

However, for the first term in the right-hand side of (3.29), if we denote

(3.36) δ~(ξ)=(γa(δdetδ𝕀3×3)etΦ)(ξ),\tilde{\delta}(\xi)=\left(\gamma_{a}(\delta-\det\delta\mathbb{I}_{3\times 3})\text{e}^{-t\Phi}\right)(\xi),

one can find that δ~\tilde{\delta} satisfies the following RH problem:
\bullet δ~(ξ)\tilde{\delta}(\xi) is analytic in ξ[ξ0,ξ0]\xi\in{\mathbb{C}}\setminus[-\xi_{0},\xi_{0}] with continuous boundary values on (ξ0,ξ0)(-\xi_{0},\xi_{0});
\bullet On the jump contour (ξ0,ξ0)(-\xi_{0},\xi_{0}), δ~(ξ)\tilde{\delta}(\xi) satisfies the jump condition

(3.37) δ~+(ξ)=(1+γ(ξ)γ(ξ))δ~(ξ)+f(ξ)etΦ(x,t;ξ),ξ(ξ0,ξ0),\tilde{\delta}_{+}(\xi)=(1+\gamma(\xi)\gamma^{\dagger}(\xi))\tilde{\delta}_{-}(\xi)+f(\xi)\text{e}^{-t\Phi(x,t;\xi)},\quad\xi\in(-\xi_{0},\xi_{0}),

where

(3.38) f(ξ)=γa(ξ)[γ(ξ)γ(ξ)γ(ξ)γ(ξ)𝕀3×3]δ(ξ);f(\xi)=\gamma_{a}(\xi)[\gamma^{\dagger}(\xi)\gamma(\xi)-\gamma(\xi)\gamma^{\dagger}(\xi)\mathbb{I}_{3\times 3}]\delta_{-}(\xi);

\bullet δ~(ξ)𝟎3×3,\tilde{\delta}(\xi)\to\mathbf{0}_{3\times 3}, as ξ\xi\to\infty.
By [1], the function δ~(ξ)\tilde{\delta}(\xi) can be expressed by

(3.39) δ~(ξ)\displaystyle\tilde{\delta}(\xi) =X(ξ)ξ0ξ0etΦ(x,t;s)f(s)X+(s)(sξ)ds,\displaystyle=X(\xi)\int_{-\xi_{0}}^{\xi_{0}}\frac{\text{e}^{-t\Phi(x,t;s)}f(s)}{X_{+}(s)(s-\xi)}\text{d}s,
X(ξ)\displaystyle X(\xi) =exp{12πiξ0ξ0ln(1+γ(s)γ(s))sξds}.\displaystyle=\exp\left\{\frac{1}{2\pi\text{i}}\int_{-\xi_{0}}^{\xi_{0}}\frac{\ln(1+\gamma(s)\gamma^{\dagger}(s))}{s-\xi}\text{d}s\right\}.

Define L^{z=κeiπ4:0<κ<+}\hat{L}\doteq\{z=\kappa\text{e}^{-\frac{\text{i}\pi}{4}}:0<\kappa<+\infty\}, then we have the following lemma.

Lemma 3.2.

As tt\rightarrow\infty, for zL^z\in\hat{L}, the estimate for δ~(ξ)\tilde{\delta}(\xi) holds:

(3.40) |(Tξ0δ~)(z)|Ct12.\left|(T_{\xi_{0}}\tilde{\delta})(z)\right|\leq Ct^{-\frac{1}{2}}.
Proof.

See Lemma 3.2 in [31]. ∎

Remark 3.1.

In a similar way, we have for zL^z\in\hat{L}^{*},

(3.41) |(Tξ0δ^)(z)|Ct12,t,\left|(T_{\xi_{0}}\hat{\delta})(z)\right|\leq Ct^{-\frac{1}{2}},\quad t\rightarrow\infty,

where δ^(ξ)=((δ1(detδ)1𝕀3×3)γaetΦ)(ξ)\hat{\delta}(\xi)=\left((\delta^{-1}-(\det\delta)^{-1}\mathbb{I}_{3\times 3})\gamma^{\dagger}_{a}\text{e}^{t\Phi}\right)(\xi). Analogous estimates also follow for Tξ0T_{-\xi_{0}}.

Let 𝒳±ξ0X±ξ0\mathcal{X}_{\pm\xi_{0}}\doteq X\pm\xi_{0} be the cross XX defined by (3.2) centered at ±ξ0\pm\xi_{0} and Dϵ(±ξ0)D_{\epsilon}(\pm\xi_{0}) denote the open disk of radius ϵ\epsilon centered at ±ξ0\pm\xi_{0} for a small ϵ>0\epsilon>0, 𝒳±ξ0ϵ=𝒳±ξ0Dϵ(±ξ0)\mathcal{X}_{\pm\xi_{0}}^{\epsilon}=\mathcal{X}_{\pm\xi_{0}}\cap D_{\epsilon}(\pm\xi_{0}). Then, combining above analysis, we find that as tt\to\infty, the jump matrix J(2)(x,t;ξ)J^{(2)}(x,t;\xi) in Dϵ(ξ0)D_{\epsilon}(\xi_{0}) tends to J(ξ0)(γ(ξ0);z)J^{(\xi_{0})}(\gamma(\xi_{0});z) for zXz\in X, where

(3.42) J(ξ0)(γ(ξ0);z)={(1ϑ2z2iνeiz22γ(ξ0)03×1𝕀3×3),ξ(𝒳ξ0ϵ)1,(1ϑ2z2iνeiz22γ(ξ0)1+γ(ξ0)γ(ξ0)03×1𝕀3×3),ξ(𝒳ξ0ϵ)2,(101×3ϑ2z2iνeiz22γ(ξ0)1+γ(ξ0)γ(ξ0)𝕀3×3),ξ(𝒳ξ0ϵ)3,(101×3ϑ2z2iνeiz22γ(ξ0)𝕀3×3),ξ(𝒳ξ0ϵ)4.J^{(\xi_{0})}(\gamma(\xi_{0});z)=\left\{\begin{aligned} &\begin{pmatrix}1&\vartheta^{2}z^{-2\text{i}\nu}\text{e}^{-\frac{\text{i}z^{2}}{2}}\gamma(\xi_{0})\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\left(\mathcal{X}_{\xi_{0}}^{\epsilon}\right)_{1},\\ &\begin{pmatrix}1&-\vartheta^{2}z^{-2\text{i}\nu}\text{e}^{-\frac{\text{i}z^{2}}{2}}\frac{\gamma(\xi_{0})}{1+\gamma(\xi_{0})\gamma^{\dagger}(\xi_{0})}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\,\ \xi\in\left(\mathcal{X}_{\xi_{0}}^{\epsilon}\right)_{2},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ -\vartheta^{-2}z^{2\text{i}\nu}\text{e}^{\frac{\text{i}z^{2}}{2}}\frac{\gamma^{\dagger}(\xi_{0})}{1+\gamma(\xi_{0})\gamma^{\dagger}(\xi_{0})}&\mathbb{I}_{3\times 3}\end{pmatrix},\qquad\xi\in\left(\mathcal{X}_{\xi_{0}}^{\epsilon}\right)_{3},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \vartheta^{-2}z^{2\text{i}\nu}\text{e}^{\frac{\text{i}z^{2}}{2}}\gamma^{\dagger}(\xi_{0})&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\left(\mathcal{X}_{\xi_{0}}^{\epsilon}\right)_{4}.\end{aligned}\right.

This suggests that in the neighborhood Dϵ(ξ0)D_{\epsilon}(\xi_{0}) of ξ0\xi_{0}, we can approximate M(2)(x,t;ξ)M^{(2)}(x,t;\xi) by a 4×44\times 4 matrix-valued function of the form

(3.43) M(ξ0)(x,t;ξ)=ϑσMX(γ(ξ0);z)ϑσ,M^{(\xi_{0})}(x,t;\xi)=\vartheta^{-\sigma}M^{X}(\gamma(\xi_{0});z)\vartheta^{\sigma},

where MX(γ(ξ0);z)M^{X}(\gamma(\xi_{0});z) is the solution of the model RH problem considered in Subsection 3.1 by setting ρ=γ(ξ0)\rho=\gamma(\xi_{0}).

Lemma 3.3.

The function M(ξ0)(x,t;ξ)M^{(\xi_{0})}(x,t;\xi) defined in (3.43) is analytic and bounded for ξDϵ(ξ0)𝒳ξ0ϵ\xi\in D_{\epsilon}(\xi_{0})\setminus\mathcal{X}_{\xi_{0}}^{\epsilon}. On the contour 𝒳ξ0ϵ\mathcal{X}_{\xi_{0}}^{\epsilon}, M(ξ0)M^{(\xi_{0})} satisfies the jump relation M+(ξ0)=M(ξ0)J(ξ0)M_{+}^{(\xi_{0})}=M_{-}^{(\xi_{0})}J^{(\xi_{0})}, and J(ξ0)J^{(\xi_{0})} obeys the estimate for 1n1\leq n\leq\infty:

(3.44) J(2)J(ξ0)Ln(𝒳ξ0ϵ)Ct1212nlnt.\left\|J^{(2)}-J^{(\xi_{0})}\right\|_{L^{n}\left(\mathcal{X}_{\xi_{0}}^{\epsilon}\right)}\leq Ct^{-\frac{1}{2}-\frac{1}{2n}}\ln t.

As tt\rightarrow\infty, we have

(3.45) [M(ξ0)(x,t;ξ)]1𝕀4×4L(Dϵ(ξ0))=O(t12),\left\|\left[M^{(\xi_{0})}(x,t;\xi)\right]^{-1}-\mathbb{I}_{4\times 4}\right\|_{L^{\infty}(\partial D_{\epsilon}(\xi_{0}))}=O(t^{-\frac{1}{2}}),

and

(3.46) 12πiDϵ(ξ0)([M(ξ0)(x,t;ξ)]1𝕀4×4)dξ=ϑσM1X(γ(ξ0))ϑσ48tξ0+O(t1).-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(\xi_{0})}\left(\left[M^{(\xi_{0})}(x,t;\xi)\right]^{-1}-\mathbb{I}_{4\times 4}\right)\text{d}\xi=\frac{\vartheta^{-\sigma}M^{X}_{1}(\gamma(\xi_{0}))\vartheta^{\sigma}}{\sqrt{48t\xi_{0}}}+O(t^{-1}).
Proof.

The analyticity and bound of M(ξ0)M^{(\xi_{0})} are a consequence of Theorem 3.1. Moreover, for ξ(𝒳ξ0ϵ)1\xi\in(\mathcal{X}_{\xi_{0}}^{\epsilon})_{1}, that is, z=48tξ0ςeiπ4z=\sqrt{48t\xi_{0}}\varsigma\text{e}^{-\frac{\text{i}\pi}{4}}, 0ςϵ0\leq\varsigma\leq\epsilon, it follows from the Lemma 3.35 in [15] that

(3.47) υ2γa(z48tξ0+ξ0)z2iνeiz22γ(ξ0)L((𝒳ξ0ϵ)1)C|eiκ2z2|t12lnt,0<κ<12.\left\|\upsilon^{2}\gamma_{a}\left(\frac{z}{\sqrt{48t\xi_{0}}}+\xi_{0}\right)-z^{-2\text{i}\nu}\text{e}^{-\frac{\text{i}z^{2}}{2}}\gamma(\xi_{0})\right\|_{L^{\infty}\left((\mathcal{X}_{\xi_{0}}^{\epsilon})_{1}\right)}\leq C\left|\text{e}^{-\frac{\text{i}\kappa}{2}z^{2}}\right|t^{-\frac{1}{2}}\ln t,\quad 0<\kappa<\frac{1}{2}.

Thus, together with (3.40) and (3.41), we have

(3.48) J(2)J(ξ0)L((𝒳ξ0ϵ)1)C|eiκ2z2|t12lnt.\left\|J^{(2)}-J^{(\xi_{0})}\right\|_{L^{\infty}\left((\mathcal{X}_{\xi_{0}}^{\epsilon})_{1}\right)}\leq C\left|\text{e}^{-\frac{\text{i}\kappa}{2}z^{2}}\right|t^{-\frac{1}{2}}\ln t.

Hence, we arrive at

(3.49) J(2)J(ξ0)L1((𝒳ξ0ϵ)1)Ct1lnt.\left\|J^{(2)}-J^{(\xi_{0})}\right\|_{L^{1}\left((\mathcal{X}_{\xi_{0}}^{\epsilon})_{1}\right)}\leq Ct^{-1}\ln t.

By the general inequality fLnfL11/nfL11/n\|f\|_{L^{n}}\leq\|f\|^{1-1/n}_{L^{\infty}}\|f\|_{L^{1}}^{1/n}, we find

(3.50) J(2)J(ξ0)Ln((𝒳ξ0ϵ)1)Ct1212nlnt.\left\|J^{(2)}-J^{(\xi_{0})}\right\|_{L^{n}\left((\mathcal{X}_{\xi_{0}}^{\epsilon})_{1}\right)}\leq Ct^{-\frac{1}{2}-\frac{1}{2n}}\ln t.

The norms on (𝒳ξ0ϵ)j(\mathcal{X}_{\xi_{0}}^{\epsilon})_{j}, j=2,3,4j=2,3,4 can be estimated in a similar way. Therefore, (3.44) follows.

If ξDϵ(ξ0)\xi\in\partial D_{\epsilon}(\xi_{0}), the variable z=48tξ0(ξξ0)z=\sqrt{48t\xi_{0}}(\xi-\xi_{0}) tends to infinity as tt\rightarrow\infty. It follows from (3.4) that

(3.51) MX(γ(ξ0);z)=𝕀4×4+M1X(γ(ξ0))48tξ0(ξξ0)+O(t1),t,ξDϵ(ξ0).M^{X}(\gamma(\xi_{0});z)=\mathbb{I}_{4\times 4}+\frac{M^{X}_{1}(\gamma(\xi_{0}))}{\sqrt{48t\xi_{0}}(\xi-\xi_{0})}+O(t^{-1}),\quad t\rightarrow\infty,~{}\xi\in\partial D_{\epsilon}(\xi_{0}).

Then the equation (3.43) yields

(3.52) [M(ξ0)(x,t;ξ)]1𝕀4×4=ϑσM1X(γ(ξ0))ϑσ48tξ0(ξξ0)+O(t1),t,ξDϵ(ξ0).\left[M^{(\xi_{0})}(x,t;\xi)\right]^{-1}-\mathbb{I}_{4\times 4}=-\frac{\vartheta^{-\sigma}M^{X}_{1}(\gamma(\xi_{0}))\vartheta^{\sigma}}{\sqrt{48t\xi_{0}}(\xi-\xi_{0})}+O(t^{-1}),\quad t\rightarrow\infty,~{}\xi\in\partial D_{\epsilon}(\xi_{0}).

The estimate (3.45) then immediately follows. By Cauchy’s formula and (3.52), we find (3.46). ∎

On the other hand, it is easy to check that as tt\to\infty, the jump matrix J(2)(x,t;ξ)J^{(2)}(x,t;\xi) in Dϵ(ξ0)D_{\epsilon}(-\xi_{0}) tends to 𝒜[J(ξ0)(γ(ξ0);z)]𝒜\mathcal{A}\left[J^{(\xi_{0})}(\gamma(\xi_{0});-z^{*})\right]^{*}\mathcal{A} for zXz\in X. Thus, we can approximate M(2)(x,t;ξ)M^{(2)}(x,t;\xi) in the neighborhood Dϵ(ξ0)D_{\epsilon}(-\xi_{0}) of ξ0-\xi_{0} by 𝒜[M(ξ0)(x,t;ξ)]𝒜\mathcal{A}\left[M^{(\xi_{0})}(x,t;-\xi^{*})\right]^{*}\mathcal{A}.

3.4. Find asymptotic formula

Define the approximate solution M(app)(x,t;ξ)M^{(app)}(x,t;\xi) by

(3.53) M(app)(x,t;ξ)={M(ξ0)(x,t;ξ),ξDϵ(ξ0),𝒜[M(ξ0)(x,t;ξ)]𝒜,ξDϵ(ξ0),𝕀4×4,elsewhere.M^{(app)}(x,t;\xi)=\left\{\begin{aligned} &M^{(\xi_{0})}(x,t;\xi),\qquad\qquad\quad\xi\in D_{\epsilon}(\xi_{0}),\\ &\mathcal{A}\left[M^{(\xi_{0})}(x,t;-\xi^{*})\right]^{*}\mathcal{A},\ \xi\in D_{\epsilon}(-\xi_{0}),\\ &\mathbb{I}_{4\times 4},\qquad\qquad\qquad\,\,\ \qquad{\text{e}lsewhere}.\end{aligned}\right.

Let M^(x,t;ξ)\hat{M}(x,t;\xi) be

(3.54) M^(x,t;ξ)=M(2)(x,t;ξ)[M(app)(x,t;ξ)]1.\hat{M}(x,t;\xi)=M^{(2)}(x,t;\xi)\left[M^{(app)}(x,t;\xi)\right]^{-1}.

Denote Σ^=ΣDϵ(ξ0)Dϵ(ξ0)\hat{\Sigma}=\Sigma\cup\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0}) and suppose that the boundaries of Dϵ(±ξ0)D_{\epsilon}(\pm\xi_{0}) are oriented counterclockwise, see Figure 4. Then M^(x,t;ξ)\hat{M}(x,t;\xi) satisfies the following 4×44\times 4 matrix RH problem:
\bullet M^(x,t;ξ)\hat{M}(x,t;\xi) is analytic for ξΣ^\xi\in{\mathbb{C}}\setminus\hat{\Sigma} with continuous boundary values on Σ^\hat{\Sigma};
\bullet Across the contour Σ^\hat{\Sigma}, the limiting values M^±(x,t;ξ)\hat{M}_{\pm}(x,t;\xi) obey the jump relation

(3.55) M^+(x,t;ξ)=M^(x,t;ξ)J^(x,t;ξ);\hat{M}_{+}(x,t;\xi)=\hat{M}_{-}(x,t;\xi)\hat{J}(x,t;\xi);

\bullet As ξ\xi\to\infty, M^(x,t;ξ)𝕀4×4\hat{M}(x,t;\xi)\to\mathbb{I}_{4\times 4};
where the jump matrix J^(x,t;ξ)\hat{J}(x,t;\xi) is given by

(3.56) J^={M(app)J(2)[M+(app)]1,ξΣ^(Dϵ(ξ0)Dϵ(ξ0)),[M(app)]1,ξDϵ(ξ0)Dϵ(ξ0),J(2),ξΣ^(Dϵ(ξ0)¯Dϵ(ξ0)¯).\hat{J}=\left\{\begin{aligned} &M^{(app)}_{-}J^{(2)}\left[M^{(app)}_{+}\right]^{-1},\,\,\ \xi\in\hat{\Sigma}\cap(D_{\epsilon}(\xi_{0})\cup D_{\epsilon}(-\xi_{0})),\\ &\left[M^{(app)}\right]^{-1},\qquad\qquad\quad\xi\in\partial D_{\epsilon}(\xi_{0})\cup\partial D_{\epsilon}(-\xi_{0}),\\ &J^{(2)},\qquad\qquad\qquad\qquad\quad\xi\in\hat{\Sigma}\setminus(\overline{D_{\epsilon}(\xi_{0})}\cup\overline{D_{\epsilon}(-\xi_{0})}).\end{aligned}\right.
Refer to caption
Figure 4. The oriented contour Σ^\hat{\Sigma}.

Now we rewrite Σ^\hat{\Sigma} as:

Σ^=(Dϵ(ξ0)Dϵ(ξ0))(𝒳ξ0ϵ𝒳ξ0ϵ)Σ~,\hat{\Sigma}=(\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0}))\cup(\mathcal{X}_{-\xi_{0}}^{\epsilon}\cup\mathcal{X}_{\xi_{0}}^{\epsilon})\cup\tilde{\Sigma}\cup{\mathbb{R}},

where

Σ~=14Σj(Dϵ(ξ0)Dϵ(ξ0)),\tilde{\Sigma}=\bigcup_{1}^{4}\Sigma_{j}\setminus(D_{\epsilon}(-\xi_{0})\cup D_{\epsilon}(\xi_{0})),

and denote w^=J^𝕀4×4\hat{w}=\hat{J}-\mathbb{I}_{4\times 4}, then we have the following lemma.

Lemma 3.4.

For 1n1\leq n\leq\infty, the following estimates hold:

(3.57) w^Ln(Dϵ(ξ0)Dϵ(ξ0))\displaystyle\|\hat{w}\|_{L^{n}(\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0}))} Ct12,\displaystyle\leq Ct^{-\frac{1}{2}},
(3.58) w^Ln(𝒳ξ0ϵ𝒳ξ0ϵ)\displaystyle\|\hat{w}\|_{L^{n}(\mathcal{X}_{-\xi_{0}}^{\epsilon}\cup\mathcal{X}_{\xi_{0}}^{\epsilon})} Ct1212nlnt,\displaystyle\leq Ct^{-\frac{1}{2}-\frac{1}{2n}}\ln t,
(3.59) w^Ln(Σ~)\displaystyle\|\hat{w}\|_{L^{n}(\tilde{\Sigma})} Cect,\displaystyle\leq C\text{e}^{-ct},
(3.60) w^Ln()\displaystyle\|\hat{w}\|_{L^{n}({\mathbb{R}})} Ct32.\displaystyle\leq Ct^{-\frac{3}{2}}.
Proof.

It follows the same line as Lemma 3.4 in [31]. ∎

The uniformly vanishing bound on w^\hat{w} in Lemma 3.4 establishes RH problem for M^(x,t;ξ)\hat{M}(x,t;\xi) as a small-norm Riemann–Hilbert problem, for which there is a well known existence and uniqueness theorem [15]. In fact, we may write

(3.61) M^(x,t;ξ)=𝕀4×4+12πiΣ^(μ^w^)(x,t;s)sξds,\hat{M}(x,t;\xi)=\mathbb{I}_{4\times 4}+\frac{1}{2\pi\text{i}}\int_{\hat{\Sigma}}\frac{(\hat{\mu}\hat{w})(x,t;s)}{s-\xi}\text{d}s,

where the 4×44\times 4 matrix-valued function μ^(x,t;ξ)\hat{\mu}(x,t;\xi) is the unique solution of

(3.62) μ^=𝕀4×4+C^w^μ^.\hat{\mu}=\mathbb{I}_{4\times 4}+\hat{C}_{\hat{w}}\hat{\mu}.

The singular integral operator C^w^\hat{C}_{\hat{w}}: L2(Σ^)L2(Σ^)L^{2}(\hat{\Sigma})\to L^{2}(\hat{\Sigma}) is defined for fL2(Σ^)f\in L^{2}(\hat{\Sigma}) by

(3.63) C^w^f\displaystyle\hat{C}_{\hat{w}}f C^(fw^),\displaystyle\doteq\hat{C}_{-}(f\hat{w}),
(3.64) (C^f)(ξ)\displaystyle(\hat{C}_{-}f)(\xi) limξΣ^Σ^f(s)sξds2πi,\displaystyle\doteq\lim_{\xi\to\hat{\Sigma}_{-}}\int_{\hat{\Sigma}}\frac{f(s)}{s-\xi}\frac{\text{d}s}{2\pi\text{i}},

where C^\hat{C}_{-} is the well known Cauchy operator. Then, by Lemma 3.4 and (3.63), we find

(3.65) C^w^B(L2(Σ^))Cw^L(Σ^)Ct12lnt,\|\hat{C}_{\hat{w}}\|_{B(L^{2}(\hat{\Sigma}))}\leq C\|\hat{w}\|_{L^{\infty}(\hat{\Sigma})}\leq Ct^{-\frac{1}{2}}\ln t,

where B(L2(Σ^))B(L^{2}(\hat{\Sigma})) denotes the Banach space of bounded linear operators L2(Σ^)L2(Σ^)L^{2}(\hat{\Sigma})\rightarrow L^{2}(\hat{\Sigma}). Therefore, there exists a T>0T>0 such that 1C^w^B(L2(Σ^))1-\hat{C}_{\hat{w}}\in B(L^{2}(\hat{\Sigma})) is invertible for all (x,t)𝒪,(x,t)\in\mathcal{O}, t>Tt>T. And hence the existence of both μ^\hat{\mu} and M^\hat{M} immediately follow.

Moreover, standard estimates using the Neumann series shows that μ^(x,t;ξ)\hat{\mu}(x,t;\xi) satisfies

(3.66) μ^(x,t;)𝕀4×4L2(Σ^)=O(t12),t.\|\hat{\mu}(x,t;\cdot)-\mathbb{I}_{4\times 4}\|_{L^{2}(\hat{\Sigma})}=O(t^{-\frac{1}{2}}),\quad t\rightarrow\infty.

In fact, Equation (3.62) is equivalent to μ^=𝕀4×4+(1C^w^)1C^w^𝕀4×4\hat{\mu}=\mathbb{I}_{4\times 4}+(1-\hat{C}_{\hat{w}})^{-1}\hat{C}_{\hat{w}}\mathbb{I}_{4\times 4}. Using the Neumann series, one can obtain

(1C^w^)1B(L2(Σ^))11C^w^B(L2(Σ^))\|(1-\hat{C}_{\hat{w}})^{-1}\|_{B(L^{2}(\hat{\Sigma}))}\leq\frac{1}{1-\|\hat{C}_{\hat{w}}\|_{B(L^{2}(\hat{\Sigma}))}}

whenever C^w^B(L2(Σ^))<1\|\hat{C}_{\hat{w}}\|_{B(L^{2}(\hat{\Sigma}))}<1. Then, we find

(3.67) μ^(x,t;)𝕀4×4L2(Σ^)=(1C^w^)1C^w^𝕀4×4L2(Σ^)(1C^w^)1B(L2(Σ^))C^(w^)L2(Σ^)Cw^L2(Σ^)1C^w^B(L2(Σ^))Cw^L2(Σ^)\displaystyle\begin{aligned} \|\hat{\mu}(x,t;\cdot)-\mathbb{I}_{4\times 4}\|_{L^{2}(\hat{\Sigma})}=&\|(1-\hat{C}_{\hat{w}})^{-1}\hat{C}_{\hat{w}}\mathbb{I}_{4\times 4}\|_{L^{2}(\hat{\Sigma})}\\ \leq&\|(1-\hat{C}_{\hat{w}})^{-1}\|_{B(L^{2}(\hat{\Sigma}))}\|\hat{C}_{-}(\hat{w})\|_{L^{2}(\hat{\Sigma})}\\ \leq&\frac{C\|\hat{w}\|_{L^{2}(\hat{\Sigma})}}{1-\|\hat{C}_{\hat{w}}\|_{B(L^{2}(\hat{\Sigma}))}}\leq C\|\hat{w}\|_{L^{2}(\hat{\Sigma})}\end{aligned}

for all tt large enough. In view of Lemma 3.4, this gives (3.66).

It then follows from (3.10), (3.23), (3.54), (3.53) and (3.61) that

(3.68) limξξ(M(x,t;ξ)𝕀4×4)=limξξ(M^(x,t;ξ)𝕀4×4)=12πiΣ^(μ^w^)(x,t;ξ)dξ.\lim_{\xi\rightarrow\infty}\xi(M(x,t;\xi)-\mathbb{I}_{4\times 4})=\lim_{\xi\rightarrow\infty}\xi(\hat{M}(x,t;\xi)-\mathbb{I}_{4\times 4})=-\frac{1}{2\pi\text{i}}\int_{\hat{\Sigma}}(\hat{\mu}\hat{w})(x,t;\xi)\text{d}\xi.

By (3.46), (3.53), (3.57) and (3.66), we can get

(3.69) 12πiDϵ(ξ0)Dϵ(ξ0)(μ^w^)(x,t;ξ)dξ\displaystyle-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0})}(\hat{\mu}\hat{w})(x,t;\xi)\text{d}\xi
=\displaystyle= 12πiDϵ(ξ0)Dϵ(ξ0)w^dξ12πiDϵ(ξ0)Dϵ(ξ0)(μ^𝕀4×4)w^dξ\displaystyle-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0})}\hat{w}\text{d}\xi-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0})}(\hat{\mu}-\mathbb{I}_{4\times 4})\hat{w}\text{d}\xi
=\displaystyle= 12πiDϵ(ξ0)((𝒜[M(ξ0)(x,t;ξ)]𝒜)1𝕀4×4)dξ\displaystyle-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(-\xi_{0})}\left(\left(\mathcal{A}\left[M^{(\xi_{0})}(x,t;-\xi^{*})\right]^{*}\mathcal{A}\right)^{-1}-\mathbb{I}_{4\times 4}\right)\text{d}\xi
12πiDϵ(ξ0)((M(ξ0)(x,t;ξ))1𝕀4×4)dξ\displaystyle-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(\xi_{0})}\left(\left(M^{(\xi_{0})}(x,t;\xi)\right)^{-1}-\mathbb{I}_{4\times 4}\right)\text{d}\xi
+O(μ^𝕀4×4L2(Dϵ(ξ0)Dϵ(ξ0))w^L2(Dϵ(k0)Dϵ(ξ0)))\displaystyle+O(\|\hat{\mu}-\mathbb{I}_{4\times 4}\|_{L^{2}(\partial D_{\epsilon}(-\xi_{0})\cup\partial D_{\epsilon}(\xi_{0}))}\|\hat{w}\|_{L^{2}(\partial D_{\epsilon}(-k_{0})\cup\partial D_{\epsilon}(\xi_{0}))})
=\displaystyle= 𝒜[ϑσM1X(γ(ξ0))ϑσ]𝒜48tξ0+ϑσM1X(γ(ξ0))ϑσ48tξ0+O(t1).\displaystyle-\frac{\mathcal{A}\left[\vartheta^{-\sigma}M^{X}_{1}(\gamma(\xi_{0}))\vartheta^{\sigma}\right]^{*}\mathcal{A}}{\sqrt{48t\xi_{0}}}+\frac{\vartheta^{-\sigma}M^{X}_{1}(\gamma(\xi_{0}))\vartheta^{\sigma}}{\sqrt{48t\xi_{0}}}+O(t^{-1}).

Similarly, the contributions from 𝒳k0ϵ𝒳k0ϵ\mathcal{X}_{-k_{0}}^{\epsilon}\cup\mathcal{X}_{k_{0}}^{\epsilon}, Σ~\tilde{\Sigma} and {\mathbb{R}} to the right-hand side of (3.68) are O(t1lnt)O(t^{-1}\ln t), O(ect)O(\text{e}^{-ct}) and O(t3/2)O(t^{-3/2}), respectively.

Thus, taking into account that the reconstructional formula (2.22) and (3.5), we get

(u(x,t)u(x,t)w(x,t))=\displaystyle\begin{pmatrix}u(x,t)&u^{*}(x,t)&w(x,t)\end{pmatrix}= 2ilimξ(ξM(x,t;ξ))12\displaystyle 2\text{i}\lim_{\xi\rightarrow\infty}(\xi M(x,t;\xi))_{12}
(3.70) =\displaystyle= 2ilimξξ(M^(x,t;ξ)𝕀4×4)12\displaystyle 2\text{i}\lim_{\xi\rightarrow\infty}\xi(\hat{M}(x,t;\xi)-\mathbb{I}_{4\times 4})_{12}
=\displaystyle= 2i(iϑ2βXi(ϑ2βX)σ148tξ0)+O(lntt).\displaystyle 2\text{i}\left(\frac{-\text{i}\vartheta^{2}\beta^{X}-\text{i}\left(\vartheta^{2}\beta^{X}\right)^{*}\sigma_{1}}{\sqrt{48t\xi_{0}}}\right)+O\left(\frac{\ln t}{t}\right).

Collecting above computations, we obtain the long-time asymptotic result of the solution in oscillating sector 𝒪\mathcal{O}.

Theorem 3.2.

Let u0(x)u_{0}(x) and w0(x)w_{0}(x) lie in the Schwartz space 𝒮()\mathcal{S}({\mathbb{R}}), and generate the scattering data in sense that: the determinant of the 3×33\times 3 matrix-valued spectral function a(ξ)a(\xi) defined in (2.17) has no zeros in {\mathbb{C}}_{-}. Then, in oscillating sector 𝒪\mathcal{O}, as tt\rightarrow\infty, the solution of the initial problem for the new two-component Sasa-Satsuma equation (1.6) on the line satisfies the following asymptotic formula

(3.71) (u(x,t)w(x,t))=(uas(x,t)was(x,t))t+O(lntt),\begin{pmatrix}u(x,t)&w(x,t)\end{pmatrix}=\frac{\begin{pmatrix}u_{as}(x,t)&w_{as}(x,t)\end{pmatrix}}{\sqrt{t}}+O\left(\frac{\ln t}{t}\right),

where the leading-order coefficient uas(x,t)u_{as}(x,t) and was(x,t)w_{as}(x,t) are given by

(3.72) uas(x,t)=\displaystyle u_{as}(x,t)= νeπν224πξ0((192tξ03)iνe16itξ03+2χ(ξ0)+πi4Γ(iν)γ1(ξ0)\displaystyle\frac{\nu\text{e}^{-\frac{\pi\nu}{2}}}{\sqrt{24\pi\xi_{0}}}\left((192t\xi_{0}^{3})^{\text{i}\nu}\text{e}^{16\text{i}t\xi_{0}^{3}+2\chi(\xi_{0})+\frac{\pi\text{i}}{4}}\Gamma(-\text{i}\nu)\gamma_{1}(\xi_{0})\right.
+(192tξ03)iνe16itξ032χ(ξ0)πi4Γ(iν)γ1(ξ0)),\displaystyle\left.+(192t\xi_{0}^{3})^{-\text{i}\nu}\text{e}^{-16\text{i}t\xi_{0}^{3}-2\chi(\xi_{0})-\frac{\pi\text{i}}{4}}\Gamma(\text{i}\nu)\gamma_{1}(-\xi_{0})\right),
(3.73) was(x,t)=\displaystyle w_{as}(x,t)= νeπν26πξ0|Γ(iν)γ2(ξ0)|cos(νln(192tξ03)+16tξ03+π4+argγ2(ξ0)\displaystyle\frac{\nu\text{e}^{-\frac{\pi\nu}{2}}}{\sqrt{6\pi\xi_{0}}}\left|\Gamma(-\text{i}\nu)\gamma_{2}(\xi_{0})\right|\cos\bigg{(}\nu\ln(192t\xi_{0}^{3})+16t\xi_{0}^{3}+\frac{\pi}{4}+\arg\gamma_{2}(\xi_{0})
+argΓ(iν)1πξ0ξ0ln(1+γ(s)γ(s)1+γ(ξ0)γ(ξ0))dssξ0),\displaystyle+\arg\Gamma(-\text{i}\nu)-\frac{1}{\pi}\int_{-\xi_{0}}^{\xi_{0}}\ln\left(\frac{1+\gamma(s)\gamma^{\dagger}(s)}{1+\gamma(\xi_{0})\gamma^{\dagger}(\xi_{0})}\right)\frac{\text{d}s}{s-\xi_{0}}\bigg{)},

where ξ0\xi_{0}, ν\nu, χ(ξ)\chi(\xi) and γ1(ξ)\gamma_{1}(\xi), γ2(ξ)\gamma_{2}(\xi) are given by (3.8), (3.31), (3.32) and (3.17), respectively.

4. Asymptotic analysis in Painlevé sector 𝒫\mathcal{P}

In this section, we study the long-time asymptotics of solution to the new two-component Sasa–Satsuma equation (1.6) in Painlevé region 𝒫\mathcal{P} defined by

(4.1) 𝒫={(x,t)2|t>1,|x|Nt1/3},Nconstant.\mathcal{P}=\{(x,t)\in{\mathbb{R}}^{2}|t>1,\ |x|\leq Nt^{1/3}\},\ N\ \text{constant}.

Let

𝒫𝒫{x0},𝒫𝒫{x0}\displaystyle\mathcal{P}_{\geq}\doteq\mathcal{P}\cap\{x\geq 0\},\quad\mathcal{P}_{\leq}\doteq\mathcal{P}\cap\{x\leq 0\}

denote the right and left halves of 𝒫\mathcal{P}. The long-time asymptotic formula of the solution for the case (x,t)𝒫(x,t)\in\mathcal{P}_{\geq} will be established, the case when (x,t)𝒫(x,t)\in\mathcal{P}_{\leq} can be handled in a similar but easy way.

At first, we will prove Theorem 4.1, which expresses the large zz behavior of the solution of a model RH problem in terms of the solution of a new coupled Painlevé II equation. This result will be important for analyzing the long-time asymptotics of the solution of system (1.6) in region 𝒫\mathcal{P}_{\geq}.

4.1. Another model RH problem

Refer to caption
Figure 5. The oriented contour ZZ.

Given z00z_{0}\geq 0, let ZZ denote the contour Z=j=15ZjZ=\cup_{j=1}^{5}Z_{j}, where the line segments

(4.2) Z1\displaystyle Z_{1} ={z0+ςeπi6|ς0},Z2={z0+ςe5πi6|ς0},\displaystyle=\{z_{0}+\varsigma\text{e}^{\frac{\pi\text{i}}{6}}|\varsigma\geq 0\},\qquad\,\,\ Z_{2}=\{-z_{0}+\varsigma\text{e}^{\frac{5\pi\text{i}}{6}}|\varsigma\geq 0\},
Z3\displaystyle Z_{3} ={z0+ςe5πi6|ς0},Z4={z0+ςeπi6|ς0},\displaystyle=\{-z_{0}+\varsigma\text{e}^{-\frac{5\pi\text{i}}{6}}|\varsigma\geq 0\},\quad Z_{4}=\{z_{0}+\varsigma\text{e}^{-\frac{\pi\text{i}}{6}}|\varsigma\geq 0\},
Z5\displaystyle Z_{5} ={ς|z0ςz0},\displaystyle=\{\varsigma|-z_{0}\leq\varsigma\leq z_{0}\},

are oriented as in Figure 5. We consider the following family of RH problems parameterized by y>0y>0 and a 1×31\times 3 complex-valued row vector pp with p=pσ1p=p^{*}\sigma_{1} for MZM^{Z}:
\bullet MZ(y,p;z)M^{Z}(y,p;z) is analytic in Z{\mathbb{C}}\setminus Z with continuous boundary values on ZZ;
\bullet For zZz\in Z, the boundary values M±ZM^{Z}_{\pm} satisfy the jump relation M+Z(y,p;z)=MZ(y,p;z)JZ(y,p;z);M^{Z}_{+}(y,p;z)=M^{Z}_{-}(y,p;z)J^{Z}(y,p;z);
\bullet MZ(y,p;z)𝕀4×4M^{Z}(y,p;z)\rightarrow\mathbb{I}_{4\times 4}, as zz\rightarrow\infty;
where the jump matrix JZJ^{Z} is defined by

(4.3) JZ(y,p;z)={(101×3pe2i(43z3yz)𝕀3×3),zZ1Z2,(1pe2i(43z3yz)03×1𝕀3×3),zZ3Z4,(1pe2i(43z3yz)03×1𝕀3×3)(101×3pe2i(43z3yz)𝕀3×3),zZ5.J^{Z}(y,p;z)=\left\{\begin{aligned} &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ p^{\dagger}\text{e}^{2\text{i}(\frac{4}{3}z^{3}-yz)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad z\in Z_{1}\cup Z_{2},\\ &\begin{pmatrix}1&p\text{e}^{-2\text{i}(\frac{4}{3}z^{3}-yz)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad z\in Z_{3}\cup Z_{4},\\ &\begin{pmatrix}1&p\text{e}^{-2\text{i}(\frac{4}{3}z^{3}-yz)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ p^{\dagger}\text{e}^{2\text{i}(\frac{4}{3}z^{3}-yz)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad z\in Z_{5}.\end{aligned}\right.
Theorem 4.1.

Define the parameter subset

(4.4) ={(y,t,z0)3|0yC1,t2,y/2z0C2},\mathbb{P}=\{(y,t,z_{0})\in{\mathbb{R}}^{3}|0\leq y\leq C_{1},\ t\geq 2,\ \sqrt{y}/2\leq z_{0}\leq C_{2}\},

where C1C_{1}, C2>0C_{2}>0 are constants. Then the RH problem for MZM^{Z} with jump matrix JZJ^{Z} has a unique solution MZ(y,p;z)M^{Z}(y,p;z) whenever (y,t,z0)(y,t,z_{0})\in\mathbb{P}. There are smooth functions {MjZ(y)}\{M_{j}^{Z}(y)\} such that

(4.5) MZ(y,p;z)=𝕀4×4+j=1NMjZ(y)zj+O(zN1),z,M^{Z}(y,p;z)=\mathbb{I}_{4\times 4}+\sum_{j=1}^{N}\frac{M_{j}^{Z}(y)}{z^{j}}+O(z^{-N-1}),\quad z\rightarrow\infty,

however, the (1,2)(1,2) entry and (1,4)(1,4) entry of leading coefficient M1ZM_{1}^{Z} are given by

(4.6) (M1Z(y))12=up(y),(M1Z(y))14=iwp(y),\left(M_{1}^{Z}(y)\right)_{12}=u_{p}(y),\quad\left(M_{1}^{Z}(y)\right)_{14}=\text{i}w_{p}(y),

where up(y)u_{p}(y) and wp(y)w_{p}(y) are complex-valued and real-valued functions, respectively, and are the smooth solution of a new coupled Painlevé II equation (B.5). Moreover, MZ(y,p;z)M^{Z}(y,p;z) is uniformly bounded for zZz\in{\mathbb{C}}\setminus Z, and satisfies the symmetries

(4.7) MZ(y,p;z)=[(MZ)(y,p;z)]1=𝒜[MZ(y,p;z)]𝒜.M^{Z}(y,p;z)=\left[\left(M^{Z}\right)^{\dagger}(y,p;z^{*})\right]^{-1}=\mathcal{A}\left[M^{Z}(y,p;-z^{*})\right]^{*}\mathcal{A}.
Proof.

It is easy to see that

Re(2i(43z3yz))=ς(83ς243z0ς4z02+y)83ς343z0ς2,\displaystyle\text{Re}\left(2\text{i}\left(\frac{4}{3}z^{3}-yz\right)\right)=\varsigma\left(-\frac{8}{3}\varsigma^{2}-4\sqrt{3}z_{0}\varsigma-4z_{0}^{2}+y\right)\leq-\frac{8}{3}\varsigma^{3}-4\sqrt{3}z_{0}\varsigma^{2},

for all z=z0+ςeπi6z=z_{0}+\varsigma\text{e}^{\frac{\pi\text{i}}{6}} and z=z0+ςe5πi6z=-z_{0}+\varsigma\text{e}^{\frac{5\pi\text{i}}{6}} with ς0,z00\varsigma\geq 0,z_{0}\geq 0 and 0<y4z020<y\leq 4z_{0}^{2}. Thus, we have

|e2i(43z3+yz)|Ce|z±z0|2(|z±z0|+z0),zZ1Z2.\displaystyle|\text{e}^{2\text{i}(\frac{4}{3}z^{3}+yz)}|\leq C\text{e}^{-|z\pm z_{0}|^{2}(|z\pm z_{0}|+z_{0})},\quad z\in Z_{1}\cup Z_{2}.

Analogous estimates hold for zZ3Z4z\in Z_{3}\cup Z_{4}. This shows that JZ𝕀4×4J^{Z}\to\mathbb{I}_{4\times 4} exponentially fast as zz\rightarrow\infty.

Note that, the jump matrix JZJ^{Z} satisfies the same symmetries (B.6) and (B.14) as JPJ^{P}. In other words, JZJ^{Z} is Hermitian and positive definite on ZZ\cap{\mathbb{R}} and satisfies JZ(y,p;z)=(JZ)(y,p;z)J^{Z}(y,p;z)=(J^{Z})^{\dagger}(y,p;z^{*}) on ZZ\setminus{\mathbb{R}}. This implies that the jump conditions and the jump matrices in the RH problem for MZM^{Z} satisfy the hypotheses of Zhou’s vanishing lemma [51], that is, the jump contour Z1Z2Z_{1}\cup Z_{2} has the necessary invariance under Schwarz reflection with orientation and if zz lies in the part of jump contour on the real axis, JZ+(JZ)J^{Z}+(J^{Z})^{\dagger} is positive definite. Thus we deduce the unique existence of the solution MZM^{Z}. The symmetries of JZJ^{Z} implies that (4.7) follows. Moreover, the RH problem for MZ(y,p;z)M^{Z}(y,p;z) can be transformed into the RH problem for MP(y;z)M^{P}(y;z) stated in Appendix B up to a trivial contour deformation. Therefore, we complete the proof of Theorem 4.1. ∎

4.2. Transformations

Suppose (x,t)𝒫(x,t)\in\mathcal{P}_{\geq}. Then, as tt\rightarrow\infty, the critical points ±ξ0\pm\xi_{0} given by (3.8) approach 0. In this case, we only need the triangular factorization of the jump matrix in the form as follows:

(4.8) J(x,t;ξ)=(1γ(ξ)etΦ(x,t;ξ)03×1𝕀3×3)(101×3γ(ξ)etΦ(x,t;ξ)𝕀3×3).J(x,t;\xi)=\begin{pmatrix}1&\gamma(\xi)\text{e}^{-t\Phi(x,t;\xi)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}(\xi)\text{e}^{t\Phi(x,t;\xi)}&\mathbb{I}_{3\times 3}\end{pmatrix}.
Refer to caption
Figure 6. The sets 𝒟\mathcal{D} and 𝒟\mathcal{D}^{*} and oriented contour Γ\Gamma.

Define the contour Γ\Gamma\subset{\mathbb{C}} by Γ=j=14Γj\Gamma=\cup_{j=1}^{4}\Gamma_{j}\cup{\mathbb{R}}, where

(4.9) Γ1\displaystyle\Gamma_{1} ={ξ0+ςeπi6|ς0},Γ2={ξ0+ςe5πi6|ς0},\displaystyle=\{\xi_{0}+\varsigma\text{e}^{\frac{\pi\text{i}}{6}}|\varsigma\geq 0\},\quad\quad\,\,\ \Gamma_{2}=\{-\xi_{0}+\varsigma\text{e}^{\frac{5\pi\text{i}}{6}}|\varsigma\geq 0\},
Γ3\displaystyle\Gamma_{3} ={ξ0+ςe5πi6|ς0},Γ4={ξ0+ςeπi6|ς0}.\displaystyle=\{-\xi_{0}+\varsigma\text{e}^{-\frac{5\pi\text{i}}{6}}|\varsigma\geq 0\},\quad\Gamma_{4}=\{\xi_{0}+\varsigma\text{e}^{-\frac{\pi\text{i}}{6}}|\varsigma\geq 0\}.

The orientation of Γ\Gamma and the triangular domains 𝒟\mathcal{D}, 𝒟\mathcal{D}^{*} are shown in Figure 6. Recalling (3.17), by Lemma 3.1, we also have the following analytic decomposition lemma for γ1(ξ)\gamma_{1}(\xi) and γ2(ξ)\gamma_{2}(\xi).

Lemma 4.1.

For j=1,2j=1,2, we have

(4.10) γj(ξ)=γj,a(x,t;ξ)+γj,r(x,t;ξ),ξ(,ξ0)(ξ0,),\gamma_{j}(\xi)=\gamma_{j,a}(x,t;\xi)+\gamma_{j,r}(x,t;\xi),\quad\xi\in(-\infty,-\xi_{0})\cup(\xi_{0},\infty),

where the functions γj,a\gamma_{j,a} and γj,r\gamma_{j,r} satisfy:
(i) For (x,t)𝒫(x,t)\in\mathcal{P}_{\geq}, γj,a(x,t;ξ)\gamma_{j,a}(x,t;\xi) is defined and continuous for ξ𝒟¯\xi\in\bar{\mathcal{D}} and analytic for ξ𝒟\xi\in\mathcal{D}.
(ii) The function γj,a\gamma_{j,a} satisfies

(4.11) |γj,a(x,t;ξ)|C1+|ξ|2et4|ReΦ(x,t;ξ)|,ξ𝒟¯,|\gamma_{j,a}(x,t;\xi)|\leq\frac{C}{1+|\xi|^{2}}\text{e}^{\frac{t}{4}|\text{Re}\Phi(x,t;\xi)|},~{}\xi\in\bar{\mathcal{D}},

and

(4.12) |γj,a(x,t;ξ)γ(ξ0)|C|ξξ0|et4|ReΦ(x,t;ξ)|,ξ𝒟¯.|\gamma_{j,a}(x,t;\xi)-\gamma(\xi_{0})|\leq C|\xi-\xi_{0}|\text{e}^{\frac{t}{4}|\text{Re}\Phi(x,t;\xi)|},~{}\xi\in\bar{\mathcal{D}}.

(iii) The L1,L2L^{1},L^{2} and LL^{\infty} norms of the function γj,r(x,t;)\gamma_{j,r}(x,t;\cdot) on (,ξ0)(ξ0,)(-\infty,-\xi_{0})\cup(\xi_{0},\infty) are O(t3/2)O(t^{-3/2}) as tt\rightarrow\infty uniformly with respect to (x,t)𝒫(x,t)\in\mathcal{P}_{\geq}.

Thus, we have obtained a decomposition γ(ξ)=γa(x,t;ξ)+γr(x,t;ξ)\gamma(\xi)=\gamma_{a}(x,t;\xi)+\gamma_{r}(x,t;\xi) by setting

γa(x,t;ξ)=\displaystyle\gamma_{a}(x,t;\xi)= (γ1,a(x,t;ξ),γ1,a(x,t;ξ),γ2,a(x,t;ξ)),ξ𝒟,\displaystyle(\gamma_{1,a}(x,t;\xi),\gamma^{*}_{1,a}(x,t;-\xi^{*}),\gamma_{2,a}(x,t;\xi)),\quad\xi\in\mathcal{D},
γr(x,t;ξ)=\displaystyle\gamma_{r}(x,t;\xi)= (γ1,r(x,t;ξ),γ1,r(x,t;ξ),γ2,r(x,t;ξ)),ξ.\displaystyle(\gamma_{1,r}(x,t;\xi),\gamma^{*}_{1,r}(x,t;-\xi),\gamma_{2,r}(x,t;\xi)),\quad\xi\in{\mathbb{R}}.

Now we can deform the contour by introducing the new 4×44\times 4 matrix-valued function M(1)(x,t;ξ)M^{(1)}(x,t;\xi) as follows:

(4.13) M(1)(x,t;ξ)=M(x,t;ξ)×{(1γa(x,t;ξ)etΦ(x,t;ξ)03×1𝕀3×3),ξ𝒟,(101×3γa(x,t;ξ)etΦ(x,t;ξ)𝕀3×3),ξ𝒟,𝕀4×4,elsewhere.M^{(1)}(x,t;\xi)=M(x,t;\xi)\times\left\{\begin{aligned} &\begin{pmatrix}1&\gamma_{a}(x,t;\xi)\text{e}^{-t\Phi(x,t;\xi)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\,\,\ \xi\in\mathcal{D},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ -\gamma^{\dagger}_{a}(x,t;\xi^{*})\text{e}^{t\Phi(x,t;\xi)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\mathcal{D}^{*},\\ &\mathbb{I}_{4\times 4},\qquad\qquad\qquad\qquad\qquad\qquad\ \ \ \text{elsewhere}.\end{aligned}\right.

Then M(1)(x,t;ξ)M^{(1)}(x,t;\xi) satisfies the following RH problem:
\bullet M(1)(x,t;ξ)M^{(1)}(x,t;\xi) is analytic in Γ{\mathbb{C}}\setminus\Gamma with continuous boundary values on Γ\Gamma;
\bullet For ξΓ\xi\in\Gamma, the limiting values M±(1)M^{(1)}_{\pm} obey the jump condition M+(1)(x,t;ξ)=M(1)(x,t;ξ)J(1)(x,t;ξ),M^{(1)}_{+}(x,t;\xi)=M^{(1)}_{-}(x,t;\xi)J^{(1)}(x,t;\xi), where

(4.14) J(1)={(101×3γaetΦ𝕀3×3),ξΓ1Γ2,(1γaetΦ03×1𝕀3×3),ξΓ3Γ4,(1γretΦ03×1𝕀3×3)(101×3γretΦ𝕀3×3),ξ(,ξ0)(ξ0,),(1γetΦ03×1𝕀3×3)(101×3γetΦ𝕀3×3),ξ(ξ0,ξ0);J^{(1)}=\left\{\begin{aligned} &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}_{a}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\ \xi\in\Gamma_{1}\cup\Gamma_{2},\,\ \begin{pmatrix}1&\gamma_{a}\text{e}^{-t\Phi}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\ \xi\in\Gamma_{3}\cup\Gamma_{4},\\ &\begin{pmatrix}1&\gamma_{r}\text{e}^{-t\Phi}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}_{r}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\ \xi\in(-\infty,-\xi_{0})\cup(\xi_{0},\infty),\\ &\begin{pmatrix}1&\gamma\text{e}^{-t\Phi}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}\text{e}^{t\Phi}&\mathbb{I}_{3\times 3}\end{pmatrix},\,\,\ \xi\in(-\xi_{0},\xi_{0});\end{aligned}\right.

\bullet As ξ\xi\to\infty, M(1)(x,t;ξ)𝕀4×4M^{(1)}(x,t;\xi)\to\mathbb{I}_{4\times 4}.

4.3. Local model

Refer to caption
Figure 7. The oriented contour Γ^\hat{\Gamma} and Γϵ=j=15Γjϵ\Gamma^{\epsilon}=\cup_{j=1}^{5}\Gamma^{\epsilon}_{j}.

Select suitable ϵ>0\epsilon>0 and denote Dϵ(0)={ξ||ξ|<ϵ}D_{\epsilon}(0)=\{\xi\in{\mathbb{C}}||\xi|<\epsilon\}. Define contour Γϵ=(ΓDϵ(0))((,ξ0)(ξ0,))\Gamma^{\epsilon}=(\Gamma\cap D_{\epsilon}(0))\setminus((-\infty,-\xi_{0})\cup(\xi_{0},\infty)), see Figure 7. Let

(4.15) z(3t)13ξ,y(3t)13x,z\doteq(3t)^{\frac{1}{3}}\xi,\quad y\doteq(3t)^{-\frac{1}{3}}x,

such that

(4.16) tΦ(x,t;ξ)=2i(43z3yz).t\Phi(x,t;\xi)=2\text{i}\left(\frac{4}{3}z^{3}-yz\right).

Then for fixed zz and as tt\to\infty, the jump matrix J(1)(x,t;ξ)J^{(1)}(x,t;\xi) tends to

(4.17) J(1)(x,t;ξ){(101×3γ(0)e2i(43z3yz)𝕀3×3),ξΓ1ϵΓ2ϵ,(1γ(0)e2i(43z3yz)03×1𝕀3×3),ξΓ3ϵΓ4ϵ,(1γ(0)e2i(43z3yz)03×1𝕀3×3)(101×3γ(0)e2i(43z3yz)𝕀3×3),ξΓ5ϵ,J^{(1)}(x,t;\xi)\rightarrow\left\{\begin{aligned} &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}(0)\text{e}^{2\text{i}(\frac{4}{3}z^{3}-yz)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\quad\xi\in\Gamma^{\epsilon}_{1}\cup\Gamma^{\epsilon}_{2},\\ &\begin{pmatrix}1&\gamma(0)\text{e}^{-2\text{i}(\frac{4}{3}z^{3}-yz)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\\ \end{pmatrix},\qquad\xi\in\Gamma^{\epsilon}_{3}\cup\Gamma^{\epsilon}_{4},\\ &\begin{pmatrix}1&\gamma(0)\text{e}^{-2\text{i}(\frac{4}{3}z^{3}-yz)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\\ \end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}(0)\text{e}^{2\text{i}(\frac{4}{3}z^{3}-yz)}&\mathbb{I}_{3\times 3}\end{pmatrix},\,\,\xi\in\Gamma^{\epsilon}_{5},\end{aligned}\right.

which is nothing but the jump matrix JZJ^{Z} defined in (4.3) with p=γ(0)p=\gamma(0). Thus, as tt\rightarrow\infty, in Dϵ(0)D_{\epsilon}(0), the solution M(1)(x,t;ξ)M^{(1)}(x,t;\xi) can be approximated by a 4×44\times 4 matrix-valued function M(0)(x,t;ξ)M^{(0)}(x,t;\xi) defined by

(4.18) M(0)(x,t;ξ)MZ(y,γ(0);z),M^{(0)}(x,t;\xi)\doteq M^{Z}(y,\gamma(0);z),

where MZ(y,γ(0);z)M^{Z}(y,\gamma(0);z) is the solution of the model RH problem established in Subsection 4.1 with z0=y/2z_{0}=\sqrt{y}/2. Moreover, if (x,t)𝒫(x,t)\in\mathcal{P}_{\geq}, then (y,t,z0)(y,t,z_{0})\in\mathbb{P}, where \mathbb{P} is the parameter subset defined in (4.4). Thus, Theorem 4.1 implies that M(0)(x,t;ξ)M^{(0)}(x,t;\xi) is well-defined by (4.18).

Lemma 4.2.

For (x,t)𝒫(x,t)\in\mathcal{P}_{\geq}, the function M(0)(x,t;ξ)M^{(0)}(x,t;\xi) is analytic for ξDϵ(0)Γϵ\xi\in D_{\epsilon}(0)\setminus\Gamma^{\epsilon} such that |M(0)(x,t;ξ)|C|M^{(0)}(x,t;\xi)|\leq C. Across the contour Γϵ\Gamma^{\epsilon}, the continuous boundary values M±(0)M^{(0)}_{\pm} obey the jump relation M+(0)=M(0)J(0)M_{+}^{(0)}=M_{-}^{(0)}J^{(0)}, where jump matrix J(0)J^{(0)} satisfies, for 1n1\leq n\leq\infty,

(4.19) J(1)J(0)Ln(Γϵ)Ct13(1+1n).\left\|J^{(1)}-J^{(0)}\right\|_{L^{n}(\Gamma^{\epsilon})}\leq Ct^{-\frac{1}{3}(1+\frac{1}{n})}.

Moreover, as tt\rightarrow\infty, we have

(4.20) [M(0)(x,t;ξ)]1𝕀4×4L(Dϵ(0))=O(t13),\left\|\left[M^{(0)}(x,t;\xi)\right]^{-1}-\mathbb{I}_{4\times 4}\right\|_{L^{\infty}(\partial D_{\epsilon}(0))}=O\left(t^{-\frac{1}{3}}\right),

and

(4.21) 12πiDϵ(0)([M(0)(x,t;ξ)]1𝕀4×4)dξ=M1(0)(y)(3t)13+O(t23),-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(0)}\left(\left[M^{(0)}(x,t;\xi)\right]^{-1}-\mathbb{I}_{4\times 4}\right)\text{d}\xi=\frac{M_{1}^{(0)}(y)}{(3t)^{\frac{1}{3}}}+O(t^{-\frac{2}{3}}),

where

(4.22) (M1(0)(y))12=up(y),(M1(0)(y))14=iwp(y),\left(M_{1}^{(0)}(y)\right)_{12}=u_{p}(y),\quad\left(M_{1}^{(0)}(y)\right)_{14}=\text{i}w_{p}(y),

furthermore, the complex-valued function up(y)u_{p}(y) and real-valued function wp(y)w_{p}(y) are the smooth solution of the new coupled Painlevé II equation (B.5).

Proof.

The proof follows the similar lines as Lemma 4.1 in [31]. ∎

4.4. Find asymptotic formula

Define the contour Γ^=ΓDϵ(0)\hat{\Gamma}=\Gamma\cup\partial D_{\epsilon}(0) and let the boundary of Dϵ(0)D_{\epsilon}(0) is oriented counterclockwise as depicted in Figure 7. We now introduce M^(x,t;ξ)\hat{M}(x,t;\xi) by

(4.23) M^(x,t;ξ)={M(1)(x,t;ξ)[M(0)(x,t;ξ)]1,ξDϵ(0),M(1)(x,t;ξ),ξDϵ(0).\hat{M}(x,t;\xi)=\left\{\begin{aligned} &M^{(1)}(x,t;\xi)\left[M^{(0)}(x,t;\xi)\right]^{-1},\quad\xi\in D_{\epsilon}(0),\\ &M^{(1)}(x,t;\xi),\qquad\qquad\qquad\,\,\,\qquad\xi\in{\mathbb{C}}\setminus D_{\epsilon}(0).\end{aligned}\right.

It then can be shown that M^(x,t;ξ)\hat{M}(x,t;\xi) satisfies the following RH problem:
\bullet M^(x,t;ξ)\hat{M}(x,t;\xi) is analytic outside the contour Γ^\hat{\Gamma} with continuous boundary values on Γ^\hat{\Gamma};
\bullet For ξΓ^\xi\in\hat{\Gamma}, we have the jump relation M^+(x,t;ξ)=M^(x,t;ξ)J^(x,t;ξ);\hat{M}_{+}(x,t;\xi)=\hat{M}_{-}(x,t;\xi)\hat{J}(x,t;\xi);
\bullet M^(x,t;ξ)𝕀4×4\hat{M}(x,t;\xi)\to\mathbb{I}_{4\times 4}, as ξ\xi\to\infty;
where the jump matrix J^(x,t;ξ)\hat{J}(x,t;\xi) is described by

(4.24) J^={M(0)J(1)[M+(0)]1,ξΓ^Dϵ(0),[M(0)]1,ξDϵ(0),J(1),ξΓ^Dϵ(0)¯.\hat{J}=\left\{\begin{aligned} &M^{(0)}_{-}J^{(1)}\left[M^{(0)}_{+}\right]^{-1},\,\,\ \xi\in\hat{\Gamma}\cap D_{\epsilon}(0),\\ &\left[M^{(0)}\right]^{-1},\qquad\qquad\ \xi\in\partial D_{\epsilon}(0),\\ &J^{(1)},\qquad\qquad\qquad\quad\ \xi\in\hat{\Gamma}\setminus\overline{D_{\epsilon}(0)}.\end{aligned}\right.
Lemma 4.3.

Let ω^=J^𝕀4×4\hat{\omega}=\hat{J}-\mathbb{I}_{4\times 4}, Γ~=Γ(Dϵ(0)¯)\tilde{\Gamma}=\Gamma\setminus({\mathbb{R}}\cup\overline{D_{\epsilon}(0)}). For each 1n1\leq n\leq\infty, we have

(4.25) ω^Ln(Dϵ(0))Ct13,\displaystyle\|\hat{\omega}\|_{L^{n}(\partial D_{\epsilon}(0))}\leq Ct^{-\frac{1}{3}},
(4.26) ω^Ln(Γϵ)Ct13(1+1n),\displaystyle\|\hat{\omega}\|_{L^{n}(\Gamma^{\epsilon})}\leq Ct^{-\frac{1}{3}(1+\frac{1}{n})},
(4.27) ω^Ln([ξ0,ξ0])Ct32,\displaystyle\|\hat{\omega}\|_{L^{n}({\mathbb{R}}\setminus[-\xi_{0},\xi_{0}])}\leq Ct^{-\frac{3}{2}},
(4.28) ω^Ln(Γ~)Cect.\displaystyle\|\hat{\omega}\|_{L^{n}(\tilde{\Gamma})}\leq C\text{e}^{-ct}.
Proof.

See the proof of Lemma 4.3 in [11]. ∎

As the discussion in Subsection 3.4, the estimates in Lemma 4.3 show that the RH problem for M^\hat{M} has a unique solution given by

(4.29) M^(x,t;ξ)=𝕀4×4+12πiΓ^(ν^ω^)(x,t;s)dssξ,\hat{M}(x,t;\xi)=\mathbb{I}_{4\times 4}+\frac{1}{2\pi\text{i}}\int_{\hat{\Gamma}}(\hat{\nu}\hat{\omega})(x,t;s)\frac{\text{d}s}{s-\xi},

where ν^=𝕀4×4+(1C^ω^)1C^ω^𝕀4×4\hat{\nu}=\mathbb{I}_{4\times 4}+(1-\hat{C}_{\hat{\omega}})^{-1}\hat{C}_{\hat{\omega}}\mathbb{I}_{4\times 4} satisfies the estimate

(4.30) ν^(x,t;)𝕀4×4L2(Γ^)=O(t13),t.\|\hat{\nu}(x,t;\cdot)-\mathbb{I}_{4\times 4}\|_{L^{2}(\hat{\Gamma})}=O(t^{-\frac{1}{3}}),\quad t\rightarrow\infty.

As tt\rightarrow\infty, it then follows from (4.13), (4.23) and (4.29) that

(4.31) limξξ(M(x,t;ξ)𝕀4×4)\displaystyle\lim_{\xi\rightarrow\infty}\xi\left(M(x,t;\xi)-\mathbb{I}_{4\times 4}\right) =limξξ(M^(x,t;ξ)𝕀4×4)\displaystyle=\lim_{\xi\rightarrow\infty}\xi\left(\hat{M}(x,t;\xi)-\mathbb{I}_{4\times 4}\right)
=12πiΓ^(ν^ω^)(x,t;ξ)dξ\displaystyle=-\frac{1}{2\pi\text{i}}\int_{\hat{\Gamma}}(\hat{\nu}\hat{\omega})(x,t;\xi)\text{d}\xi
=12πiDϵ(0)ω^(x,t;ξ)dξ+O(t23)\displaystyle=-\frac{1}{2\pi\text{i}}\int_{\partial D_{\epsilon}(0)}\hat{\omega}(x,t;\xi)\text{d}\xi+O(t^{-\frac{2}{3}})
=M1(0)(y)(3t)13+O(t23).\displaystyle=\frac{M_{1}^{(0)}(y)}{(3t)^{\frac{1}{3}}}+O(t^{-\frac{2}{3}}).

Using the reconstruction formula (2.22), we hence obtain long-time asymptotics of the solution in Painlevé sector 𝒫\mathcal{P}.

Theorem 4.2.

Under the assumptions of Theorem 3.2, the solution of the new two-component Sasa–Satsuma equation (1.6) satisfies the following asymptotic formula in Painlevé region 𝒫\mathcal{P} as tt\to\infty

(4.32) u(x,t)=\displaystyle u(x,t)= 2i(3t)13up(x(3t)13)+O(t23),\displaystyle\frac{2\text{i}}{(3t)^{\frac{1}{3}}}u_{p}\left(\frac{x}{(3t)^{\frac{1}{3}}}\right)+O(t^{-\frac{2}{3}}),
(4.33) w(x,t)=\displaystyle w(x,t)= 2(3t)13wp(x(3t)13)+O(t23),\displaystyle-\frac{2}{(3t)^{\frac{1}{3}}}w_{p}\left(\frac{x}{(3t)^{\frac{1}{3}}}\right)+O(t^{-\frac{2}{3}}),

where complex-valued function up(y)u_{p}(y) and real-valued function wp(y)w_{p}(y) denote the smooth solution of the new coupled Painlevé II equation (B.5) corresponding to pγ(0)p\doteq\gamma(0) according to Lemma B.1. Particularly, the function up(y)u_{p}(y) has constant phase, namely, argup\arg u_{p} is independent of yy.

5. Asymptotic analysis in fast decay sector \mathcal{F}

Finally, we will study the long-time asymptotic behavior of solution to Equation (1.6) in the fast decay region \mathcal{F} defined by

(5.1) ={(x,t)2|t>1,Nt<x,t},Nconstant.\mathcal{F}=\{(x,t)\in{\mathbb{R}}^{2}|t>1,\ Nt<-x,\ t\to\infty\},\ N\ \text{constant}.

In this region, the signature table for real part of phase function Φ(x,t;ξ)\Phi(x,t;\xi) is shown in Figure 8.

Refer to caption
Figure 8. The contour Ξ1Ξ2\Xi_{1}\cup\Xi_{2} and open sets Ω\Omega, Ω\Omega^{*}.

5.1. Transformations

Define open sets

(5.2) Ω={ξ|Imξ(η,0)},Ω={ξ|Imξ(0,η)},\Omega=\{\xi\in{\mathbb{C}}|\text{Im}\xi\in(-\eta,0)\},\quad\Omega^{*}=\{\xi\in{\mathbb{C}}|\text{Im}\xi\in(0,\eta)\},

as shown in Figure 8. Then, according to the analysis of previous sections, we can similarly obtain the analytic decomposition of γ(ξ)\gamma(\xi): γ(ξ)=γa(x,t;ξ)+γr(x,t;ξ)\gamma(\xi)=\gamma_{a}(x,t;\xi)+\gamma_{r}(x,t;\xi). Moreover, the function γa(x,t;ξ)\gamma_{a}(x,t;\xi) is defined and continuous for ξΩ¯\xi\in\bar{\Omega} and analytic in Ω\Omega, the L1,L2L^{1},L^{2} and LL^{\infty} norms of the function γr(x,t;)\gamma_{r}(x,t;\cdot) on {\mathbb{R}} are O(t3/2)O(t^{-3/2}) as tt\rightarrow\infty. In order to perform the contour deformation, we would like to define the lines Ξ1={ξ|Imξ=η}\Xi_{1}=\{\xi\in{\mathbb{C}}|\text{Im}\xi=-\eta\}, Ξ2={ξ|Imξ=η}\Xi_{2}=\{\xi\in{\mathbb{C}}|\text{Im}\xi=\eta\}, and the orientation is depicted in Figure 8. Then, let us perform the transform

(5.3) M(1)(x,t;ξ)=M(x,t;ξ)×{(1γa(x,t;ξ)etΦ(x,t;ξ)03×1𝕀3×3),ξΩ,(101×3γa(x,t;ξ)etΦ(x,t;ξ)𝕀3×3),ξΩ,𝕀4×4,elsewhere.M^{(1)}(x,t;\xi)=M(x,t;\xi)\times\left\{\begin{aligned} &\begin{pmatrix}1&\gamma_{a}(x,t;\xi)\text{e}^{-t\Phi(x,t;\xi)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\ \ \xi\in\Omega,\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ -\gamma^{\dagger}_{a}(x,t;\xi^{*})\text{e}^{t\Phi(x,t;\xi)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\Omega^{*},\\ &\mathbb{I}_{4\times 4},\qquad\qquad\qquad\qquad\quad\qquad\,\,\ \ \ \text{elsewhere}.\end{aligned}\right.

Hence, the matrix RH problem for M(1)(x,t;ξ)M^{(1)}(x,t;\xi) is as follows:
\bullet M(1)(x,t;ξ)M^{(1)}(x,t;\xi) is analytic for ξ(Ξ1Ξ2)\xi\in{\mathbb{C}}\setminus({\mathbb{R}}\cup\Xi_{1}\cup\Xi_{2});
\bullet The continuous boundary values M±(x,t;ξ)M_{\pm}(x,t;\xi) at Ξ1Ξ2{\mathbb{R}}\cup\Xi_{1}\cup\Xi_{2} satisfy the jump condition M+(1)(x,t;ξ)=M(1)(x,t;ξ)J(1)(x,t;ξ)M^{(1)}_{+}(x,t;\xi)=M^{(1)}_{-}(x,t;\xi)J^{(1)}(x,t;\xi);
\bullet M(1)(x,t;ξ)=𝕀4×4+O(ξ1)M^{(1)}(x,t;\xi)=\mathbb{I}_{4\times 4}+O(\xi^{-1}), as ξ\xi\to\infty;
where the jump matrix

(5.4) J(1)(x,t;ξ)={(1γa(x,t;ξ)etΦ(x,t;ξ)03×1𝕀3×3),ξΞ1,(101×3γa(x,t;ξ)etΦ(x,t;ξ)𝕀3×3),ξΞ2,(1γr(x,t;ξ)etΦ(x,t;ξ)03×1𝕀3×3)(101×3γr(x,t;ξ)etΦ(x,t;ξ)𝕀3×3),ξ.J^{(1)}(x,t;\xi)=\left\{\begin{aligned} &\begin{pmatrix}1&\gamma_{a}(x,t;\xi)\text{e}^{-t\Phi(x,t;\xi)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\Xi_{1},\\ &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}_{a}(x,t;\xi^{*})\text{e}^{t\Phi(x,t;\xi)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in\Xi_{2},\\ &\begin{pmatrix}1&\gamma_{r}(x,t;\xi)\text{e}^{-t\Phi(x,t;\xi)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix}\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ \gamma^{\dagger}_{r}(x,t;\xi^{*})\text{e}^{t\Phi(x,t;\xi)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad\xi\in{\mathbb{R}}.\end{aligned}\right.

5.2. Find asymptotic formula

Now J(1)(x,t;ξ)J^{(1)}(x,t;\xi) decays exponentially fast to the identity matrix 𝕀4×4\mathbb{I}_{4\times 4} as tt\to\infty on the contours Ξ1Ξ2\Xi_{1}\cup\Xi_{2}. Set ω=J(1)𝕀4×4\omega=J^{(1)}-\mathbb{I}_{4\times 4}. Therefore, one can find that for 1n1\leq n\leq\infty

(5.5) ωLn()Ct32,\displaystyle\|\omega\|_{L^{n}({\mathbb{R}})}\leq Ct^{-\frac{3}{2}},
(5.6) ωLn(Ξ1Ξ2)Cect,\displaystyle\|\omega\|_{L^{n}(\Xi_{1}\cup\Xi_{2})}\leq C\text{e}^{-ct},

which immediately yields that

(5.7) limξξ(M(x,t;ξ)𝕀4×4)\displaystyle\lim_{\xi\rightarrow\infty}\xi\left(M(x,t;\xi)-\mathbb{I}_{4\times 4}\right) =limξξ(M(1)(x,t;ξ)𝕀4×4)\displaystyle=\lim_{\xi\rightarrow\infty}\xi\left(M^{(1)}(x,t;\xi)-\mathbb{I}_{4\times 4}\right)
=12πiΞ1Ξ2(μω)(x,t;ξ)dξ=O(t32),\displaystyle=-\frac{1}{2\pi\text{i}}\int_{{\mathbb{R}}\cup\Xi_{1}\cup\Xi_{2}}(\mu\omega)(x,t;\xi)\text{d}\xi=O\left(t^{-\frac{3}{2}}\right),

where μ(x,t;ξ)\mu(x,t;\xi) is defined by μ=𝕀4×4+(1C^ω)1C^ω𝕀4×4\mu=\mathbb{I}_{4\times 4}+(1-\hat{C}_{\omega})^{-1}\hat{C}_{\omega}\mathbb{I}_{4\times 4} and obeys

(5.8) μ(x,t;)𝕀4×4L2(Ξ1Ξ2)=O(t32),t.\|\mu(x,t;\cdot)-\mathbb{I}_{4\times 4}\|_{L^{2}({\mathbb{R}}\cup\Xi_{1}\cup\Xi_{2})}=O\left(t^{-\frac{3}{2}}\right),\quad t\rightarrow\infty.

Hence, by (2.22), we get the following theorem.

Theorem 5.1.

In the fast decay region \mathcal{F} described in (5.1) and under the conditions of Theorem 3.2, the solution of the new two-component Sasa–Satsuma equation (1.6) has the following long-time asymptotic behavior as tt\to\infty

(5.9) u(x,t)=O(t32),w(x,t)=O(t32).u(x,t)=O\left(t^{-\frac{3}{2}}\right),\quad w(x,t)=O\left(t^{-\frac{3}{2}}\right).

Appendix A Proof of Theorem 3.1

The proof of Theorem 3.1 relies on deriving an explicit formula for the solution MXM^{X} in terms of parabolic cylinder functions. First, note that the jump matrix JXJ^{X} obeys the symmetry

(A.1) JX(ρ;z)=(JX)(ρ;z).J^{X}(\rho;z)=\left(J^{X}\right)^{\dagger}(\rho;z^{*}).

It then follows that the RH problem for MX(ρ;z)M^{X}(\rho;z) admits a Zhou’s vanishing lemma [51], as a result, there exists a unique solution MX(ρ;z)M^{X}(\rho;z) which admits an expansion of the form (3.4) with respect to zz.

To find the leading-order coefficient of large zz asymptotic behavior for the solution MXM^{X}, we let

(A.2) Ψ(z)=MX(z)ziνσeiz24σ,\Psi(z)=M^{X}(z)z^{\text{i}\nu\sigma}\text{e}^{\frac{\text{i}z^{2}}{4}\sigma},

where we suppress the ρ\rho dependence for clarity. It follows that dΨ(z)dzΨ1(z)\frac{\text{d}\Psi(z)}{\text{d}z}\Psi^{-1}(z) has no jump discontinuity along any of the rays. On the other hand, one find as zz\to\infty,

(A.3) dΨ(z)dzΨ1(z)=i2σzi2[σ,M1X]+O(z1).\frac{\text{d}\Psi(z)}{\text{d}z}\Psi^{-1}(z)=\frac{\text{i}}{2}\sigma z-\frac{\text{i}}{2}[\sigma,M_{1}^{X}]+O\left(z^{-1}\right).

Therefore, Liouville’s argument implies that

(A.4) dΨ(z)dziz2σΨ(z)=βΨ(z),\frac{\text{d}\Psi(z)}{\text{d}z}-\frac{\text{i}z}{2}\sigma\Psi(z)=\beta\Psi(z),

where

(A.5) β=i2[σ,M1X]=(0β12β21𝟎3×3).\beta=-\frac{\text{i}}{2}[\sigma,M_{1}^{X}]=\begin{pmatrix}0~{}&\beta_{12}\\ \beta_{21}~{}&\mathbf{0}_{3\times 3}\\ \end{pmatrix}.

Here we write a 4×44\times 4 matrix AA as a block form

(A.6) A=(A11A12A21A22)A=\begin{pmatrix}A_{11}&A_{12}\\ A_{21}&A_{22}\end{pmatrix}

with A11A_{11} is scalar. Particularly, we have

(A.7) (M1X)12=iβ12,(M1X)21=iβ21.(M_{1}^{X})_{12}=-\text{i}\beta_{12},\quad(M_{1}^{X})_{21}=\text{i}\beta_{21}.

The symmetries (A.1) of JXJ^{X} together with the uniqueness of the solution of the RH problem imply the following symmetry for MXM^{X}:

(A.8) [MX(z)]1=[MX(z)],\left[M^{X}(z)\right]^{-1}=\left[M^{X}(z^{*})\right]^{\dagger},

which further yields that

(A.9) β21=β12.\beta_{21}=-\beta^{\dagger}_{12}.

Next, considering (A.4) and (A.5), we can obtain

(A.10) d2Ψ11(z)dz2+(i2+z24β12β21)Ψ11(z)=0,\displaystyle\frac{\text{d}^{2}\Psi_{11}(z)}{\text{d}z^{2}}+\left(\frac{\text{i}}{2}+\frac{z^{2}}{4}-\beta_{12}\beta_{21}\right)\Psi_{11}(z)=0,
(A.11) β12Ψ21(z)=dΨ11(z)dz+i2zΨ11(z),\displaystyle\beta_{12}\Psi_{21}(z)=\frac{\text{d}\Psi_{11}(z)}{\text{d}z}+\frac{\text{i}}{2}z\Psi_{11}(z),
(A.12) d2β12Ψ22(z)dz2(i2z24+β12β21)β12Ψ22(z)=0,\displaystyle\frac{\text{d}^{2}\beta_{12}\Psi_{22}(z)}{\text{d}z^{2}}-\left(\frac{\text{i}}{2}-\frac{z^{2}}{4}+\beta_{12}\beta_{21}\right)\beta_{12}\Psi_{22}(z)=0,
(A.13) Ψ12(z)=1β12β21(dβ12Ψ22(z)dzi2zβ12Ψ22).\displaystyle\Psi_{12}(z)=\frac{1}{\beta_{12}\beta_{21}}\left(\frac{\text{d}\beta_{12}\Psi_{22}(z)}{\text{d}z}-\frac{\text{i}}{2}z\beta_{12}\Psi_{22}\right).

Then, by simple change of variables, it can be shown that Equations (A.10) and (A.12) can be transformed into the parabolic cylinder equation

(A.14) d2g(k)dk2+(12k24+a)g(k)=0.\frac{\text{d}^{2}g(k)}{\text{d}k^{2}}+\left(\frac{1}{2}-\frac{k^{2}}{4}+a\right)g(k)=0.

However, it is known that the solution of Equation (A.14) can be expressed as

(A.15) g(k)=ν1Da(k)+ν2Da(k),g(k)=\nu_{1}D_{a}(k)+\nu_{2}D_{a}(-k),

where ν1\nu_{1}, ν2\nu_{2} are constants and Da()D_{a}(\cdot) is the standard parabolic cylinder function [43]. Then, denoting a=iβ12β21a=\text{i}\beta_{12}\beta_{21}, we have

(A.16) Ψ11(z)=\displaystyle\Psi_{11}(z)= ν1Da(e3iπ4z)+ν2Da(eπi4z),\displaystyle\nu_{1}D_{a}(\text{e}^{-\frac{3\text{i}\pi}{4}}z)+\nu_{2}D_{a}(\text{e}^{\frac{\pi\text{i}}{4}}z),
(A.17) β12Ψ22(z)=\displaystyle\beta_{12}\Psi_{22}(z)= ν3Da(e3iπ4z)+ν4Da(eπi4z).\displaystyle\nu_{3}D_{-a}(\text{e}^{\frac{3\text{i}\pi}{4}}z)+\nu_{4}D_{-a}(\text{e}^{-\frac{\pi\text{i}}{4}}z).

On the other hand, it follows from [43] that as kk\rightarrow\infty,

(A.18) Da(k)={kaek24(1+O(k2)),|argk|<3π4,kaek24(1+O(k2))2πΓ(a)eaπi+k24ka1(1+O(k2)),π4<argk<5π4,kaek24(1+O(k2))2πΓ(a)eaπi+k24ka1(1+O(k2)),5π4<argk<π4,D_{a}(k)=\left\{\begin{aligned} &k^{a}\text{e}^{-\frac{k^{2}}{4}}(1+O(k^{-2})),\qquad\qquad\qquad\qquad\quad\qquad\qquad\qquad\qquad|\arg k|<\frac{3\pi}{4},\\ &k^{a}\text{e}^{-\frac{k^{2}}{4}}(1+O(k^{-2}))-\frac{\sqrt{2\pi}}{\Gamma(-a)}\text{e}^{a\pi\text{i}+\frac{k^{2}}{4}}k^{-a-1}(1+O(k^{-2})),\qquad\frac{\pi}{4}<\arg k<\frac{5\pi}{4},\\ &k^{a}\text{e}^{-\frac{k^{2}}{4}}(1+O(k^{-2}))-\frac{\sqrt{2\pi}}{\Gamma(-a)}\text{e}^{-a\pi\text{i}+\frac{k^{2}}{4}}k^{-a-1}(1+O(k^{-2})),\ -\frac{5\pi}{4}<\arg k<-\frac{\pi}{4},\end{aligned}\right.

Hence, as argz(3π4,π4)\arg z\in(-\frac{3\pi}{4},-\frac{\pi}{4}), we find that

(A.19) Ψ11(z)\displaystyle\Psi_{11}(z) =eπν4Da(eπi4z),a=iν,\displaystyle=\text{e}^{-\frac{\pi\nu}{4}}D_{a}(\text{e}^{\frac{\pi\text{i}}{4}}z),~{}a=-\text{i}\nu,
(A.20) β12Ψ22(z)\displaystyle\beta_{12}\Psi_{22}(z) =β12e3πν4Da(e3πi4z),\displaystyle=\beta_{12}\text{e}^{\frac{3\pi\nu}{4}}D_{-a}(\text{e}^{\frac{3\pi\text{i}}{4}}z),

because as zz\rightarrow\infty,

(A.21) Ψ11ziνeiz24,Ψ22ziνeiz24𝕀3×3.\Psi_{11}\rightarrow z^{-\text{i}\nu}\text{e}^{-\frac{\text{i}z^{2}}{4}},\quad\Psi_{22}\rightarrow z^{\text{i}\nu}\text{e}^{\frac{\text{i}z^{2}}{4}}\mathbb{I}_{3\times 3}.

It then follows from the property of Da()D_{a}(\cdot)

(A.22) dDa(k)dk+k2Da(k)aDa1(k)=0,\displaystyle\frac{\text{d}D_{a}(k)}{\text{d}k}+\frac{k}{2}D_{a}(k)-aD_{a-1}(k)=0,

and (A.11), (A.13) that

(A.23) β12Ψ21(z)\displaystyle\beta_{12}\Psi_{21}(z) =aeπ(iν)4Da1(eπi4z),\displaystyle=a\text{e}^{\frac{\pi(\text{i}-\nu)}{4}}D_{a-1}(\text{e}^{\frac{\pi\text{i}}{4}}z),
Ψ12(z)\displaystyle\Psi_{12}(z) =β12eπ(i+3ν)4Da1(e3πi4z).\displaystyle=\beta_{12}\text{e}^{\frac{\pi(\text{i}+3\nu)}{4}}D_{-a-1}(\text{e}^{\frac{3\pi\text{i}}{4}}z).

Accordingly, for argz(π4,π4)\arg z\in(-\frac{\pi}{4},\frac{\pi}{4}), we can get

(A.24) Ψ11(z)\displaystyle\Psi_{11}(z) =eπν4Da(eπi4z),a=iν,\displaystyle=\text{e}^{-\frac{\pi\nu}{4}}D_{a}(\text{e}^{\frac{\pi\text{i}}{4}}z),~{}a=-\text{i}\nu,
(A.25) β12Ψ22(z)\displaystyle\beta_{12}\Psi_{22}(z) =β12eπν4Da(eπi4z),\displaystyle=\beta_{12}\text{e}^{-\frac{\pi\nu}{4}}D_{-a}(\text{e}^{-\frac{\pi\text{i}}{4}}z),
(A.26) β12Ψ21(z)\displaystyle\beta_{12}\Psi_{21}(z) =aeπ(iν)4Da1(eπi4z),\displaystyle=a\text{e}^{\frac{\pi(\text{i}-\nu)}{4}}D_{a-1}(\text{e}^{\frac{\pi\text{i}}{4}}z),
(A.27) Ψ12(z)\displaystyle\Psi_{12}(z) =β12eπ(3i+ν)4Da1(eπi4z).\displaystyle=\beta_{12}\text{e}^{-\frac{\pi(3\text{i}+\nu)}{4}}D_{-a-1}(\text{e}^{-\frac{\pi\text{i}}{4}}z).

Now, since across the ray argz=π4\arg z=-\frac{\pi}{4},

(A.28) Ψ+(z)=\displaystyle\Psi_{+}(z)= M+X(z)ziνσeiz24σ=MX(z)JX(z)ziνσeiz24σ\displaystyle M^{X}_{+}(z)z^{\text{i}\nu\sigma}\text{e}^{\frac{\text{i}z^{2}}{4}\sigma}=M^{X}_{-}(z)J^{X}(z)z^{\text{i}\nu\sigma}\text{e}^{\frac{\text{i}z^{2}}{4}\sigma}
=\displaystyle= Ψ(z)ziνσeiz24σJX(z)ziνσeiz24σ=Ψ(z)(1ρ𝟎3×1𝕀3×3),\displaystyle\Psi_{-}(z)z^{-\text{i}\nu\sigma}\text{e}^{-\frac{\text{i}z^{2}}{4}\sigma}J^{X}(z)z^{\text{i}\nu\sigma}\text{e}^{\frac{\text{i}z^{2}}{4}\sigma}=\Psi_{-}(z)\begin{pmatrix}1&\rho\\ \mathbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},

thus,

(A.29) β12eπ(3i+ν)4Da1(eπi4z)=eπν4Da(eπi4z)ρ+β12eπ(i+3ν)4Da1(e3πi4z).\beta_{12}\text{e}^{-\frac{\pi(3\text{i}+\nu)}{4}}D_{-a-1}(\text{e}^{-\frac{\pi\text{i}}{4}}z)=\text{e}^{-\frac{\pi\nu}{4}}D_{a}(\text{e}^{\frac{\pi\text{i}}{4}}z)\rho+\beta_{12}\text{e}^{\frac{\pi(\text{i}+3\nu)}{4}}D_{-a-1}(\text{e}^{\frac{3\pi\text{i}}{4}}z).

However, note from [43] that

(A.30) Da(eπi4z)=Γ(a+1)2π(eiπa2Da1(e3πi4z)+eiπa2Da1(eπi4z)),D_{a}(\text{e}^{\frac{\pi\text{i}}{4}}z)=\frac{\Gamma(a+1)}{\sqrt{2\pi}}\left(\text{e}^{\frac{\text{i}\pi a}{2}}D_{-a-1}(\text{e}^{\frac{3\pi\text{i}}{4}}z)+\text{e}^{-\frac{\text{i}\pi a}{2}}D_{-a-1}(\text{e}^{-\frac{\pi\text{i}}{4}}z)\right),

therefore, we can find that

(A.31) β12=Γ(iν)2πeπi4πν2νρ.\beta_{12}=\frac{\Gamma(-\text{i}\nu)}{\sqrt{2\pi}}\text{e}^{\frac{\pi\text{i}}{4}-\frac{\pi\nu}{2}}\nu\rho.

The estimate (3.6) is an consequence of the explicit solution MX(ρ;z)M^{X}(\rho;z) which is expressed in terms of Da~(z)D_{\tilde{a}}(z).

Appendix B A new coupled Painlevé II RH problem

Refer to caption
Figure 9. The oriented contour P=j=14PjP=\cup_{j=1}^{4}P_{j}.

Let PP denote the contour P=j=14PjP=\cup_{j=1}^{4}P_{j} oriented to the right as in Figure 9, where

(B.1) P1=\displaystyle P_{1}= {leπi6|0l<},P2={le5πi6|0l<},\displaystyle\{l\text{e}^{\frac{\pi\text{i}}{6}}|0\leq l<\infty\},\qquad P_{2}=\{l\text{e}^{\frac{5\pi\text{i}}{6}}|0\leq l<\infty\},
P3=\displaystyle P_{3}= {le5πi6|0l<},P4={leπi6|0l<}.\displaystyle\{l\text{e}^{-\frac{5\pi\text{i}}{6}}|0\leq l<\infty\},\quad P_{4}=\{l\text{e}^{-\frac{\pi\text{i}}{6}}|0\leq l<\infty\}.

Let pp be a 1×31\times 3 complex-valued row vector with p=pσ1p=p^{*}\sigma_{1} and define the jump matrix by

(B.2) JP(y;z)={(101×3pe2i(43z3yz)𝕀3×3),zP1P2,(1pe2i(43z3yz)03×1𝕀3×3),zP3P4.J^{P}(y;z)=\left\{\begin{aligned} &\begin{pmatrix}1&\textbf{0}_{1\times 3}\\ p^{\dagger}\text{e}^{2\text{i}(\frac{4}{3}z^{3}-yz)}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad z\in P_{1}\cup P_{2},\\ &\begin{pmatrix}1&p\text{e}^{-2\text{i}(\frac{4}{3}z^{3}-yz)}\\ \textbf{0}_{3\times 1}&\mathbb{I}_{3\times 3}\end{pmatrix},\quad z\in P_{3}\cup P_{4}.\end{aligned}\right.

Then we consider the following model RH problem:
\bullet MP(y;z)M^{P}(y;z) is analytic in P{\mathbb{C}}\setminus P with continuous boundary values on PP;
\bullet M+P(y;z)=MP(y;z)JP(y;z)M^{P}_{+}(y;z)=M^{P}_{-}(y;z)J^{P}(y;z), for zPz\in P;
\bullet MP(y;z)𝕀4×4M^{P}(y;z)\rightarrow\mathbb{I}_{4\times 4}, as zz\rightarrow\infty.

Lemma B.1.

The RH problem for MP(y;z)M^{P}(y;z) has a unique solution for each yy\in{\mathbb{R}}. Moreover, there are smooth functions {MjP(y)}\{M_{j}^{P}(y)\} of yy\in{\mathbb{R}} with decay as yy\rightarrow-\infty such that

(B.3) MP(y;z)=𝕀4×4+j=1NMjP(y)zj+O(zN1),z,M^{P}(y;z)=\mathbb{I}_{4\times 4}+\sum_{j=1}^{N}\frac{M_{j}^{P}(y)}{z^{j}}+O(z^{-N-1}),\quad z\rightarrow\infty,

and the (1,2)(1,2), (1,4)(1,4) entries of leading coefficient M1PM_{1}^{P} can be expressed by

(B.4) (M1P(y))12=up(y),(M1P(y))14=iwp(y),\left(M_{1}^{P}(y)\right)_{12}=u_{p}(y),\quad\left(M_{1}^{P}(y)\right)_{14}=\text{i}w_{p}(y),

where up(y)u_{p}(y) and wp(y)w_{p}(y) are complex-valued and real-valued functions, respectively, and satisfy a new coupled Painlevé II equation

(B.5) up′′(y)+8(2|up(y)|2+wp2(y))up(y)+yup(y)=0,\displaystyle u_{p}^{{}^{\prime\prime}}(y)+8\left(2|u_{p}(y)|^{2}+w_{p}^{2}(y)\right)u_{p}(y)+yu_{p}(y)=0,
wp′′(y)+8(2|up(y)|2+wp2(y))wp(y)+ywp(y)=0.\displaystyle w_{p}^{{}^{\prime\prime}}(y)+8\left(2|u_{p}(y)|^{2}+w_{p}^{2}(y)\right)w_{p}(y)+yw_{p}(y)=0.

Moreover, the function up(y)u_{p}(y) has constant phase, that is, argup\arg u_{p} is independent of yy.

Proof.

The symmetry

(B.6) JP(y;z)=[JP(y;z)],J^{P}(y;z)=\left[J^{P}(y;z^{*})\right]^{\dagger},

implies that the jump condition and jump matrix in RH problem for MP(y;z)M^{P}(y;z) satisfy the hypotheses of Zhou’s vanishing lemma [51]. Therefore, the existence and uniqueness of MP(y;z)M^{P}(y;z) immediately follow, and hence the expansion (B.3). On the other hand, by the standard nonlinear steepest descent analysis, the coefficients MjP(y)M_{j}^{P}(y) have exponential decay as yy\rightarrow-\infty.

Let ψ(y;z)=MP(y;z)ei(43z3yz)σ\psi(y;z)=M^{P}(y;z)\text{e}^{\text{i}(\frac{4}{3}z^{3}-yz)\sigma}. Then, the function (y;z)=ψy(y;z)ψ1(y;z)\mathcal{B}(y;z)=\psi_{y}(y;z)\psi^{-1}(y;z) is an entire function of zz by Liouville’s theorem. Thus, one can find

(B.7) =(MyPizMPσ)(MP)1=iσz+i[σ,M1P]1z+0.\mathcal{B}=(M_{y}^{P}-\text{i}zM^{P}\sigma)\left(M^{P}\right)^{-1}=-\text{i}\sigma z+\text{i}[\sigma,M_{1}^{P}]\doteq\mathcal{B}_{1}z+\mathcal{B}_{0}.

Similarly, the function 𝒞(y;z)=ψz(y;z)ψ1(y;z)\mathcal{C}(y;z)=\psi_{z}(y;z)\psi^{-1}(y;z) is entire, and hence,

(B.8) 𝒞=\displaystyle\mathcal{C}= (MzP+i(4z2y)MPσ)(MP)1\displaystyle\left(M_{z}^{P}+\text{i}(4z^{2}-y)M^{P}\sigma\right)\left(M^{P}\right)^{-1}
=\displaystyle= 4iσz24i[σ,M1P]z4i[σ,M2P]+4i[σ,M1P]M1Piyσ\displaystyle 4\text{i}\sigma z^{2}-4\text{i}[\sigma,M_{1}^{P}]z-4\text{i}[\sigma,M_{2}^{P}]+4\text{i}[\sigma,M_{1}^{P}]M_{1}^{P}-\text{i}y\sigma
\displaystyle\doteq 𝒞2z2+𝒞1z+𝒞0.\displaystyle\mathcal{C}_{2}z^{2}+\mathcal{C}_{1}z+\mathcal{C}_{0}.

On the other hand, by (B.7), we also have

(B.9) MyPizMPσ=MP,M_{y}^{P}-\text{i}zM^{P}\sigma=\mathcal{B}M^{P},

and then inserting the expansion (B.3) into (B.9), one can obtain

(B.10) i[σ,M2P]=i[σ,M1P]M1PM1yP,\text{i}[\sigma,M_{2}^{P}]=\text{i}[\sigma,M_{1}^{P}]M_{1}^{P}-M_{1y}^{P},

that is,

(B.11) 𝒞0=4M1yPiyσ.\mathcal{C}_{0}=4M_{1y}^{P}-\text{i}y\sigma.

The definitions of \mathcal{B} and 𝒞\mathcal{C} yields that the function ψ\psi admits the Lax equations

(B.12) {ψy=ψ,ψz=𝒞ψ.\left\{\begin{aligned} \psi_{y}=&\mathcal{B}\psi,\\ \psi_{z}=&\mathcal{C}\psi.\end{aligned}\right.

Then the compatibility condition of (B.12) implies that

(B.13) 4M1yyP4i[σ,M1P]M1yP+4iM1yP[σ,M1P]y[σ,M1P]σ+yσ[σ,M1P]=0.4M^{P}_{1yy}-4\text{i}[\sigma,M_{1}^{P}]M_{1y}^{P}+4\text{i}M_{1y}^{P}[\sigma,M_{1}^{P}]-y[\sigma,M_{1}^{P}]\sigma+y\sigma[\sigma,M_{1}^{P}]=0.

However, the symmetric relation (B.6) and

(B.14) JP(y;z)=𝒜[JP(y;z)]𝒜J^{P}(y;z)=\mathcal{A}\left[J^{P}(y;-z^{*})\right]^{*}\mathcal{A}

implies that the solution of RH problem for MP(y;z)M^{P}(y;z) satisfies

(B.15) MP(y;z)=[(MP)(y;z)]1=𝒜[MP(y;z)]𝒜.M^{P}(y;z)=\left[\left(M^{P}\right)^{\dagger}(y;z^{*})\right]^{-1}=\mathcal{A}\left[M^{P}(y;-z^{*})\right]^{*}\mathcal{A}.

Hence, the leading order coefficient M1P(y)M_{1}^{P}(y) obeys

(B.16) [M1P(y)]=M1P(y)=𝒜[M1P(y)]𝒜.-\left[M_{1}^{P}(y)\right]^{\dagger}=M_{1}^{P}(y)=-\mathcal{A}\left[M_{1}^{P}(y)\right]^{*}\mathcal{A}.

For convenience, we write

(B.17) M1P(y)=((M1P(y))11(M1P(y))12(M1P(y))21(M1P(y))22),M_{1}^{P}(y)=\begin{pmatrix}\left(M_{1}^{P}(y)\right)_{11}&\left(M_{1}^{P}(y)\right)_{12}\\[4.0pt] \left(M_{1}^{P}(y)\right)_{21}&\left(M_{1}^{P}(y)\right)_{22}\end{pmatrix},

where (M1P(y))11\left(M_{1}^{P}(y)\right)_{11} is scalar. Now, substituting (B.17) into (B.13), using the first relation in (B.16), we have

(B.18) (M1P(y))11′′2i(M1P(y))12[(M1P(y))12]2i(M1P(y))12(M1P(y))12=0,\displaystyle\left(M_{1}^{P}(y)\right)^{\prime\prime}_{11}-2\text{i}\left(M_{1}^{P}(y)\right)_{12}\left[\left(M_{1}^{P}(y)\right)^{\prime}_{12}\right]^{\dagger}-2\text{i}\left(M_{1}^{P}(y)\right)^{\prime}_{12}\left(M_{1}^{P}(y)\right)^{\dagger}_{12}=0,
(B.19) (M1P(y))12′′+2i(M1P(y))12(M1P(y))222i(M1P(y))11(M1P(y))12+y(M1P(y))12=0,\displaystyle\left(M_{1}^{P}(y)\right)^{\prime\prime}_{12}+2\text{i}\left(M_{1}^{P}(y)\right)_{12}\left(M_{1}^{P}(y)\right)^{\prime}_{22}-2\text{i}\left(M_{1}^{P}(y)\right)^{\prime}_{11}\left(M_{1}^{P}(y)\right)_{12}+y\left(M_{1}^{P}(y)\right)_{12}=0,
(B.20) (M1P(y))22′′+2i(M1P(y))12(M1P(y))12+2i[(M1P(y))12](M1P(y))12=0.\displaystyle\left(M_{1}^{P}(y)\right)^{\prime\prime}_{22}+2\text{i}\left(M_{1}^{P}(y)\right)_{12}^{\dagger}\left(M_{1}^{P}(y)\right)^{\prime}_{12}+2\text{i}\left[\left(M_{1}^{P}(y)\right)^{\prime}_{12}\right]^{\dagger}\left(M_{1}^{P}(y)\right)_{12}=0.

Since M1P(y)M^{P}_{1}(y) and its derivatives decay as yy\to-\infty, it follows from (B.18) and (B.20) that

(B.21) (M1P(y))11=\displaystyle\left(M_{1}^{P}(y)\right)^{\prime}_{11}= 2i[(M1P(y))12(M1P(y))12],\displaystyle 2\text{i}\left[\left(M_{1}^{P}(y)\right)_{12}\left(M_{1}^{P}(y)\right)^{\dagger}_{12}\right],
(M1P(y))22=\displaystyle\left(M_{1}^{P}(y)\right)^{\prime}_{22}= 2i[(M1P(y))12(M1P(y))12].\displaystyle-2\text{i}\left[\left(M_{1}^{P}(y)\right)^{\dagger}_{12}\left(M_{1}^{P}(y)\right)_{12}\right].

Inserting (B.21) into (B.19), we get

(B.22) (M1P(y))12′′+8(M1P(y))12[(M1P(y))12(M1P(y))12]+y(M1P(y))12=0.\left(M_{1}^{P}(y)\right)^{\prime\prime}_{12}+8\left(M_{1}^{P}(y)\right)_{12}\left[\left(M_{1}^{P}(y)\right)^{\dagger}_{12}\left(M_{1}^{P}(y)\right)_{12}\right]+y\left(M_{1}^{P}(y)\right)_{12}=0.

Finally, according to the second symmetry in (B.16), we find (M1P(y))12=[(M1P(y))12]σ1\left(M_{1}^{P}(y)\right)_{12}=-\left[\left(M_{1}^{P}(y)\right)_{12}\right]^{*}\sigma_{1}, thus, we can write

(B.23) (M1P(y))12=(up(y)up(y)iwp(y)),\left(M_{1}^{P}(y)\right)_{12}=\begin{pmatrix}u_{p}(y)&-u^{*}_{p}(y)&\text{i}w_{p}(y)\end{pmatrix},

where up(y)u_{p}(y) and wp(y)w_{p}(y) are complex-valued and real-valued functions, respectively. Substituting (B.23) into (B.22), one immediately obtain

(B.24) up′′(y)+8(2|up(y)|2+wp2(y))up(y)+yup(y)=0,\displaystyle u_{p}^{{}^{\prime\prime}}(y)+8\left(2|u_{p}(y)|^{2}+w_{p}^{2}(y)\right)u_{p}(y)+yu_{p}(y)=0,
wp′′(y)+8(2|up(y)|2+wp2(y))wp(y)+ywp(y)=0.\displaystyle w_{p}^{{}^{\prime\prime}}(y)+8\left(2|u_{p}(y)|^{2}+w_{p}^{2}(y)\right)w_{p}(y)+yw_{p}(y)=0.

Writing up(y)=f(y)eiα(y)u_{p}(y)=f(y)\text{e}^{\text{i}\alpha(y)} with f(y)f(y), α(y)\alpha(y) being real functions, then the first equation in (B.24) reduces to the following system

(B.25) f′′+8(2f2+wp2)f+yff(α)2=\displaystyle f^{\prime\prime}+8(2f^{2}+w_{p}^{2})f+yf-f(\alpha^{\prime})^{2}= 0,\displaystyle 0,
(B.26) 2fα+fα′′=\displaystyle 2f^{\prime}\alpha^{\prime}+f\alpha^{\prime\prime}= 0.\displaystyle 0.

It then follows from (B.26) that

(B.27) f2α=C,f^{2}\alpha^{\prime}=C,

where CC is a real constant. Using this relation to eliminate α\alpha^{\prime} from (B.25), we find

(B.28) f′′+8(2f2+wp2)f+yfC2f3=0.\displaystyle f^{\prime\prime}+8(2f^{2}+w_{p}^{2})f+yf-C^{2}f^{-3}=0.

The decay of upu_{p}, wpw_{p} and their derivatives as yy\to-\infty imply that C=0C=0. Therefore, α(y)=argup(y)\alpha(y)=\arg u_{p}(y) is independent of yy.

The proof of lemma is now completed. ∎

References

  • [1] M.J. Ablowitz, A.S. Fokas, Complex Variables: Introduction and Applications, Cambridge University Press, Cambridge (2003).
  • [2] M.J. Ablowitz, B. Prinari, A.D. Trubatch, Discrete and Continuous Nonlinear Schrödinger Systems, Cambridge University Press, Cambridge (2003).
  • [3] M.J. Ablowitz, H. Segur, Asymptotic solutions of the Korteweg–de Vries equation, Stud. Appl. Math. 57 (1977) 13–44.
  • [4] N. Akhmediev, J.M. Soto-Crespo, N. Devine, N.P. Hoffmann, Rogue wave spectra of the Sasa–Satsuma equation, Phys. D 294 (2015) 37–42.
  • [5] L.K. Arruda, J. Lenells, Long-time asymptotics for the derivative nonlinear Schrödinger equation on the half-line, Nonlinearity 30 (2017) 4141–4172.
  • [6] D.J. Benney, A.C. Newell, Propagation of nonlinear wave envelopes, J. Math. Phys. 46 (1967) 133–139.
  • [7] M. Borghese, R. Jenkins, K.T.-R. McLaughlin, Long time asymptotic behavior of the focusing nonlinear Schrödinger equation, Ann. Inst. Henri Poincaré, Anal. Non Linéaire 35(4) (2018) 887–920.
  • [8] A. Boutet de Monvel, A. Its, V. Kotlyarov, Long-time asymptotics for the focusing NLS equation with time-periodic boundary condition on the half-line, Comm. Math. Phys. 290 (2009) 479–522.
  • [9] A. Boutet de Monvel, D. Shepelsky, A Riemann–Hilbert approach for the Degasperis–Procesi equation, Nonlinearity 26 (2013) 2081–2107.
  • [10] R. Buckingham, S. Venakides, Long-time asymptotics of the nonlinear Schrödinger equation shock problem, Comm. Pure Appl. Math. 60 (2007) 1349–1414.
  • [11] C. Charlier, J. Lenells, Airy and Painlevé asymptotics for the mKdV equation, J. London Math. Soc. 101(1) (2020) 194–225.
  • [12] C. Charlier, J. Lenells, Long-time asymptotics for an integrable evolution equation with a 3×33\times 3 Lax pair, Phys. D 426 (2021) 132987.
  • [13] S. Chen, Twisted rogue-wave pairs in the Sasa–Satsuma equation, Phys. Rev. E 88 (2013) 023202.
  • [14] R. Chiao, E. Garmire, C. Townes, Self-trapping of optical beams, Phys. Rev. Lett. 13 (1964) 479–482.
  • [15] P. Deift, X. Zhou, A steepest descent method for oscillatory Riemann–Hilbert problems. Asymptotics for the MKdV equation, Ann. Math. 137 (1993) 295–368.
  • [16] X. Geng, Y. Li, J. Wei, Y. Zhai, Darboux transformation of a two-component generalized Sasa–Satsuma equation and explicit solutions, Math. Meth. Appl. Sci. 44 (2021) 12727–12745.
  • [17] X. Geng, H. Liu, The nonlinear steepest descent method to long-time asymptotics of the coupled nonlinear Schrödinger equation, J. Nonlinear Sci. 28(2) (2018) 739–763.
  • [18] X. Geng, K. Wang, M. Chen, Long-time asymptotics for the Spin-1 Gross–Pitaevskii equation, Commun. Math. Phys. 382(1) (2021) 585–611.
  • [19] C. Gilson, J. Hietarinta, J. Nimmo, Y. Ohta, Sasa–Satsuma higher-order nonlinear Schrödinger equation and its bilinearization and multisoliton solutions, Phys. Rev. E 68 (2003) 016614.
  • [20] B. Hu, L. Zhang, J. Lin, The initial-boundary value problems of the new two-component generalized Sasa–Satsuma equation with a 4×44\times 4 matrix Lax pair, Anal. Math. Phys. 12 (2022) 109.
  • [21] L. Huang, J. Lenells, Asymptotics for the Sasa–Satsuma equation in terms of a modified Painlevé II transcendent, J. Differential Equations 268 (2020) 7480–7504.
  • [22] A. Its, Asymptotic behavior of the solutions to the nonlinear Schrödinger equation, and isomonodromic deformations of systems of linear differential equations, Dokl. Akad. Nauk SSSR 261(1) (1981) 14–18.
  • [23] Y. Kodama, A. Hasegawa, Nonlinear pulse propagation in a monomode dielectric guide, IEEE J. Quantum Elect. 23 (1987) 510–524.
  • [24] J. Lenells, The nonlinear steepest descent method for Riemann–Hilbert problems of low regularity, Indiana Univ. Math. J. 66 (2017) 1287–1332.
  • [25] Z. Li, S. Tian, J. Yang, E. Fan, Soliton resolution for the complex short pulse equation with weighted Sobolev initial data in space-time solitonic regions, J. Differential Equations 329 (2022) 31–88.
  • [26] L. Ling, The algebraic representation for high order solution of Sasa–Satsuma equation, Discrete Contin. Dyn. Syst. Ser. S 9 (2016) 1975–2010.
  • [27] H. Liu, X. Geng, B. Xue, The Deift–Zhou steepest descent method to long-time asymptotics for the Sasa–Satsuma equation, J. Differential Equations 265 (2018) 5984–6008.
  • [28] L. Liu, B. Tian, Y. Yuan, Z. Du, Dark-bright solitons and semirational rogue waves for the coupled Sasa–Satsuma equations, Phys. Rev. E 97 (2018) 052217.
  • [29] N. Liu, B. Guo, Long-time asymptotics for the Sasa–Satsuma equation via nonlinear steepest descent method, J. Math. Phys. 60(1) (2019) 011504.
  • [30] N. Liu, B. Guo, Painlevé-type asymptotics of an extended modified KdV equation in transition regions, J. Differential Equations 280 (2021) 203–235.
  • [31] N. Liu, X. Zhao, B. Guo, Long-time asymptotic behavior for the matrix modified Korteweg–de Vries equation, Phys. D 443 (2023) 133560.
  • [32] H. Luo, L. Wang, Y. Zhang, G. Lu, J. Su, Y. Zhao, Data-driven solutions and parameter discovery of the Sasa–Satsuma equation via the physics-informed neural networks method, Phys. D 440 (2022) 133489.
  • [33] W. Ma, Long-time asymptotics of a three-component coupled nonlinear Schrödinger system, J. Geom. Phys. 153 (2020) 103669.
  • [34] D. Mihalache, L. Torner, F. Moldoveanu, N. Panoiu, N. Truta, Inverse-scattering approach to femtosecond solitons in monomode optical fibers, Phys. Rev. E 48 (1993) 4699–4709.
  • [35] G. Mu, Z. Qin, Dynamic patterns of high-order rogue waves for Sasa–Satsuma equation, Nonlinear Anal. Real World Appl. 31 (2016) 179–209.
  • [36] K. Nakkeeran, K. Porsezian, P. Shanmugha Sundaram, A. Mahalingam, Optical solitons in NN-coupled higher order nonlinear Schrödinger equations, Phys. Rev. Lett. 80 (1998) 1425–1428.
  • [37] D.H. Peregrine, Water waves, nonlinear Schrödinger equations and their solutions, J. Aust. Math. Soc. B 25 (1983) 16–43.
  • [38] J.E. Rothenberg, Space-time focusing: breakdown of the slowly varying envelope approximation in the self-focusing of femtosecond pulses, Optim. Lett. 17 (1992) 1340–1342.
  • [39] N. Sasa, J. Satsuma, New-type of soliton solutions for a higher-order nonlinear Schrödinger equation, J. Phys. Soc. Jpn. 60 (1991) 409–417.
  • [40] C. Sulem, P.-L. Sulem, The Nonlinear Schrödinger Equation: Self-focusing and Wave Collapse, Springer, New York, 1999.
  • [41] J. Wang, T. Su, X. Geng, R. Li, Riemann–Hilbert approach and NN-soliton solutions for a new two-component Sasa–Satsuma equation, Nonlinear Dyn. 101 (2020) 597–609.
  • [42] K. Wang, X. Geng, M. Chen, Riemann–Hilbert approach and long-time asymptotics of the positive flow short-pulse equation, Phys. D 439 (2022) 133383.
  • [43] E.T. Whittaker, G.N. Watson, A Course of Modern Analysis, 4th ed. Cambridge University Press, Cambridge (1927).
  • [44] C. Wu, B. Wei, C. Shi, B. Feng, Multi-breather solutions to the Sasa–Satsuma equation, Proc. R. Soc. A 478 (2022) 20210711.
  • [45] J. Xu, E. Fan, The unified transform method for the Sasa–Satsuma equation on the half-line, Proc. R. Soc. A. 469 (2013) 20130068.
  • [46] J. Xu, E. Fan, Long-time asymptotics for the Fokas–Lenells equation with decaying initial value problem: Without solitons, J. Differential Equations 259(3) (2015) 1098–1148.
  • [47] J. Xu, Q. Zhu, E. Fan, The initial-boundary value problem for the Sasa–Satsuma equation on a finite interval via the Fokas method, J. Math. Phys. 59 (2018) 073508.
  • [48] B. Yang, Y. Chen, High-order soliton matrices for Sasa–Satsuma equation via local Riemann–Hilbert problem, Nonlinear Anal. Real World Appl. 45 (2019) 918–941.
  • [49] V.E. Zakharov, S.V. Manakov, Asymptotic behavior of nonlinear wave systems integrated by the inverse scattering method, Zh. Eksp. Teor. Fiz. 71 (1976) 203–215, Sov. Phys. JETP 44(1) (1976) 106–112.
  • [50] H. Zhang, Y. Wang, W. Ma, Binary Darboux transformation for the coupled Sasa–Satsuma equations, Chaos 27 (2017) 073102.
  • [51] X. Zhou, The Riemann–Hilbert problem and inverse scattering, SIAM J. Math. Anal. 20(4) (1989) 966–986.