Abstract

We consider the initial-boundary value problem for Benjamin-Ono equation on a half-line. We study traditionally important problems of the theory of nonlinear partial differential equations, such as global in time existence of solutions to the initial-boundary value problem and the asymptotic behavior of solutions for large time.

1. Introduction

In this paper we study the large time asymptotic behavior of solutions to the initial-boundary value problem for the Benjamin-Ono equation on a half-line:

𝑢𝑡+𝑢𝑢𝑥+𝑢𝑥𝑥𝑢=0,𝑥>0,𝑡>0,(𝑥,0)=𝑢0(𝑥),𝑥>0,𝑢(0,𝑡)=0,𝑡>0,(1.1) where 𝑢=PV0+𝑢((𝑦,𝑡)/(𝑦𝑥))𝑑𝑦 is the Hilbert transformation, and PV means the principal value of the singular integral. We note that in the case of the whole line we have the relations 𝜕2𝑥=𝜕𝑥(𝜕2𝑥)1/2 since the operator can be written as follows: =1(𝑖𝜉/|𝜉|)=(𝜕2𝑥)1/2𝜕𝑥, where (𝜑)(𝜉)=(1/2𝜋)𝜑(𝑥)e𝑖𝑥𝜉𝑑𝑥 is the usual Fourier transform, and 1 denotes the inverse Fourier transform. This equation is of great interest in many areas of Physics (see [1, 2]). The Cauchy problem (1.1) was studied by many authors. The existence of solutions in the usual Sobolev spaces 𝐇𝑠,0 was proved in [39] and the smoothing properties of solutions were studied in [1014]. In paper [15] it was proved that for small initial data in 𝐇2,0𝐇1,1 solutions decay as 𝑡 in 𝐋 norm at the same rate 1/𝑡 as for the case of the linear Benjamin-Ono equation, where

𝐇𝑚,𝑠=𝜙𝐋2𝜙𝑚,𝑠=1+𝑥2𝑠/21𝜕2𝑥𝑚/2𝜙𝐋2<.(1.2)

The initial-boundary value problem (1.1) plays an important role in the contemporary mathematical physics. For the general theory of nonlinear equations on a half-line we refer to the book [16], where it was developed systematically a general theory of the initial-boundary value problems for nonlinear evolution equations with pseudodifferential operators on a half-line, where pseudodifferential operator 𝕂 on a half-line was introduced by virtue of the inverse Laplace transformation of the product of the symbol 𝐾(𝑝)=𝑂(𝑝𝛽) which is analytic in the right complex half-plane, and the Laplace transform of the derivative 𝜕𝑥[𝛽]𝑢. Thus, for example, in the case of 𝐾(𝑝)=𝑝3/2 we get the following definition of the fractional derivative 𝜕𝑥3/2:

𝜕𝑥3/2𝜙=1𝑝3/2𝜙𝜙(0)𝑝.(1.3) Here and below 𝑝𝛽 is the main branch of the complex analytic function in the complex half-plane Re𝑝0, so that 1𝛽=1 (we make a cut along the negative real axis (,0)). Note that due to the analyticity of 𝑝𝛽 for all Re𝑝>0 the inverse Laplace transform gives us the function which is equal to 0 for all 𝑥<0. In spite of the importance and actuality there are few results about the initial-boundary value problem for pseudodifferential equations with nonanalytic symbols. For example, in paper [17] there was considered the case of rational symbol 𝐾(𝑝) which have some poles in the right complex half-plane. There was proposed a new method for constructing the Green operator based on the introduction of some necessary condition at the singularity points of the symbol 𝐾(𝑝). In the paper [18] one of the authors considered the initial-boundary value problem for a pseudodifferential equation with symbol 𝐾(𝑝)=|𝑝|1/2 and nonlinearity |𝑢|𝜎𝑢.

As far as we know the case of nonanalytic conservative symbols 𝐾(𝑝) was not studied previously. In the present paper we fill this gap, considering as example the Benjamin-Ono equation (1.1) with a symbol 𝐾(𝑝)=𝑝|𝑝|. There are many natural open questions which we need to study. First we consider the following question: how many boundary data we should pose on problem (1.1) for its correct solvability? Also we study traditionally important problems of a theory of nonlinear partial differential equations, such as global in time existence of solutions to the initial-boundary value problem and the asymptotic behavior of solutions for large time. We adopt here the approach of book [16] based on the estimates of the Green function. The main difficulty for nonlocal equation (1.1) on a half-line is that the symbol 𝐾(𝑝)=𝑝|𝑝| is non analytic in the complex plane. Therefore we cannot apply the Laplace theory directly. To construct Green operator we proposed a new method based on the integral representation for sectionally analytic function and theory of singular integrodifferential equations with Hilbert kernel and the discontinues coefficients (see [18, 19]).

To state precisely the results of the present paper we give some notations. We denote 𝑡=1+𝑡2,{𝑡}=𝑡/𝑡. Direct Laplace transformation 𝑥𝜉 is

̂𝑢(𝜉)𝑥𝜉𝑢=0+𝑒𝜉𝑥𝑢(𝑥)𝑑𝑥,(1.4) and the inverse Laplace transformation 1𝜉𝑥 is defined by

𝑢(𝑥)1𝜉𝑥̂𝑢=(2𝜋𝑖)1𝑖𝑖𝑒𝜉𝑥̂𝑢(𝜉)𝑑𝜉.(1.5) Weighted Lebesgue space is 𝐋𝑞,𝑎(𝐑+)={𝜑𝒮;𝜑𝐋𝑞,𝑎<}, where

𝜑𝐋𝑞,𝑎=0+𝑥𝑎𝑞||||𝜑(𝑥)𝑞𝑑𝑥1/𝑞(1.6) for 𝑎>0, 1𝑞< and

𝜑𝐋=esssup𝑥𝐑+||||.𝜑(𝑥)(1.7) Sobolev space is

𝐇1𝐑+=𝜑𝒮;𝜕𝑥𝜑𝐋2.<(1.8) We define a linear functional 𝑓:

𝑓(𝜙)=0+𝑦𝜙(𝑦)𝑑𝑦.(1.9) Now we state the main results.

Theorem 1.1. Suppose that the initial data 𝑢0𝐙𝐇1(𝐑+)𝐋1,𝑎+1(𝐑+) with 𝑎(0,1) are such that the norm 𝑢0𝐙𝜀(1.10) is sufficiently small. Then there exists a unique global solution [𝑢𝐂0,);𝐇1𝐑+(1.11) to the initial-boundary value problem (1.1). Moreover the following asymptotic is valid in 𝐋(𝐑+)1𝑢=𝑡𝐴Λ𝑥𝑡1/2𝑥+min1,𝑡𝑂𝑡1(𝑎/2)(1.12) for 𝑡, where Λ(𝑥𝑡1/2)𝐋(𝐑+),Λ(0)=0 is defined below by the formula (2.191), and the constant 𝑢𝐴=𝑓00+𝑓(𝒩(𝑢))𝑑𝜏,𝒩(𝑢)=𝑢𝑥𝑢.(1.13)

Remark 1.2. Note that the time decay rate of the solution is faster comparing with the case of the corresponding Cauchy problem. So the nonlinearity 𝑢𝑢𝑥 in (1.1) is not the super critical case for our problem.

Remark 1.3. In the case of the negative half line 𝑥<0 we expect that the solutions have an oscillation character, and the time decay rate of the solution is the same as the case of the corresponding Cauchy problem. so the nonlinearity 𝑢𝑢𝑥 in (1.1) will be the super critical case.

2. Preliminaries

In subsequent consideration we will have frequently to use certain theorems of the theory of functions of complex variable, the statements of which we now quote. The proofs can be found in [19].

Theorem 2.1. Let 𝜙(𝑞) be a complex function, which obeys the Hölder condition for all finite 𝑞 and tends to a definite limit 𝜙 as |𝑞|, such that for large 𝑞 the following inequality holds: ||𝜙(𝑞)𝜙||||𝑞||𝐶𝜇,𝜇>0.(2.1) Then Cauchy type integral 1𝐹(𝑧)=2𝜋𝑖𝑖𝑖𝜙(𝑞)𝑞𝑧𝑑𝑞(2.2) constitutes a function analytic in the left and right semiplanes. Here and below these functions will be denoted 𝐹+(𝑧) and 𝐹(𝑧), respectively. These functions have the limiting values 𝐹+(𝑝) and 𝐹(𝑝) at all points of imaginary axis Re𝑝=0, on approaching the contour from the left and from the right, respectively. These limiting values are expressed by Sokhotzki-Plemelj formula: 𝐹+(𝑝)=lim𝑧𝑝,Re𝑧<012𝜋𝑖𝑖𝑖𝜙(𝑞)1𝑞𝑧𝑑𝑞=2𝜋𝑖𝑃𝑉𝑖𝑖𝜙(𝑞)1𝑞𝑝𝑑𝑞+2𝐹𝜙(𝑝),(𝑝)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖𝜙(𝑞)1𝑞𝑧𝑑𝑞=2𝜋𝑖𝑃𝑉𝑖𝑖𝜙(𝑞)1𝑞𝑝𝑑𝑞2𝜙(𝑝).(2.3)

Subtracting and adding the formula (2.3) we obtain the following two equivalent formulas:

𝐹+(𝑝)𝐹𝐹(𝑝)=𝜙(𝑝),+(𝑝)+𝐹1(𝑝)=𝜋𝑖𝑃𝑉𝑖𝑖𝜙(𝑞)𝑞𝑝𝑑𝑞,(2.4) which will be frequently employed hereafter.

Theorem 2.2. An arbitrary function 𝜙(𝑝) given on the contour Re𝑝=0, satisfying the Hölder condition, can be uniquely represented in the form 𝜙(𝑝)=𝑈+(𝑝)𝑈(𝑝),(2.5) where 𝑈±(𝑝) are the boundary values of the analytic functions 𝑈±(𝑧) and the condition 𝑈±=0 holds. These functions are determined by formula 1𝑈(𝑧)=2𝜋𝑖𝑖𝑖𝜙(𝑞)𝑞𝑧𝑑𝑞.(2.6)

Theorem 2.3. An arbitrary function 𝜑(𝑝) given on the contour Re𝑝=0, satisfying the Hölder condition, and having zero index, 1ind𝜑(𝑡)=2𝜋𝑖𝑖𝑖𝑑ln𝜑(𝑝)=0,(2.7) is uniquely representable as the ratio of the functions 𝑋+(𝑝) and 𝑋(𝑝), constituting the boundary values of functions, 𝑋+(𝑧) and 𝑋(𝑧), analytic in the left and right complex semiplane and having in these domains no zero. These functions are determined to within an arbitrary constant factor and given by formula 𝑋±(𝑧)=𝑒Γ±(𝑧)1,Γ(𝑧)=2𝜋𝑖𝑖𝑖1𝑞𝑧ln𝜑(𝑞)𝑑𝑞.(2.8)

We consider the following linear initial-boundary value problem on half-line 𝑢𝑡𝑃𝑉0+𝑢𝑦𝑦(𝑦,𝑡)𝑥𝑦𝑑𝑦=0,𝑡>0,𝑥>0,𝑢(𝑥,0)=𝑢0(𝑥),𝑥>0,𝑢(0,𝑡)=0,𝑡>0.(2.9)

Setting

||𝑞||𝐾(𝑞)=𝑞,𝐾1(𝑞)=𝑞2||𝜉||,𝑘(𝜉)=1/2𝑒(1/2)𝑖arg𝜉,(2.10)

where Re𝑘(𝜉)>0 for Re𝜉>0, we define

𝒢(𝑡)𝜙=0+𝐺(𝑥,𝑦,𝑡)𝜙(𝑦)𝑑𝑦,(2.11)

where the function 𝐺(𝑥,𝑦,𝑡) is given by formula

1𝐺(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑑𝑝𝑒𝑝𝑥1𝐾1(𝑝)+𝜉(𝑒𝑝𝑦1+Ψ(𝜉,𝑦))12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑑𝑝𝑒𝑝𝑥1𝐾1𝑌(𝑝)+𝜉(𝑝,𝜉)×lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+(𝐾𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉(𝑒𝑞𝑦+Ψ(𝜉,𝑦))𝑑𝑞(2.12) for 𝜀>0,𝑥>0,𝑦>0,𝑡>0. Here and below

𝑌±=𝑒Γ±𝑤±.(2.13)Γ+(𝑝,𝜉) and Γ(𝑝,𝜉) are a left and right limiting values of sectionally analytic function Γ(𝑧,𝜉) given by

1Γ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝑞𝑧ln𝐾(𝑞)+𝜉𝐾1𝑤(𝑞)+𝜉(𝑞)𝑤+(𝑞)𝑑𝑞,(2.14) where

𝑤𝑧(𝑧)=𝑧+𝑘(𝜉)1/2,𝑤+𝑧(𝑧)=𝑧𝑘(𝜉)1/2,𝑒Ψ(𝜉,𝑦)=𝑘(𝜉)𝑦𝑌(+1𝑘(𝜉),𝜉)2𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌+(𝐾𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑒𝑞)+𝜉𝑞𝑦𝑑𝑞.(2.15)

All the integrals are understood in the sense of the principal values.

Proposition 2.4. Let the initial data be 𝑢0𝐋1(𝐑+). Then there exists a unique solution 𝑢(𝑥,𝑡) of the initial-boundary value problem (2.9), which has integral representation 𝑢(𝑥,𝑡)=𝒢(𝑡)𝑢0.(2.16)

Proof. To derive an integral representation for the solutions of the problem (2.9) we suppose that there exists a solution 𝑢(𝑥,𝑡) of problem (2.9), which is continued by zero outside of 𝑥>0:
𝑢(𝑥,𝑡)=0,𝑥<0.(2.17)
Let 𝜙(𝑝) be a function of the complex variable 𝑝, which obeys the Hölder condition for all finite 𝑝 and tends to 0 as 𝑝±𝑖. We define the operator 1𝜙(𝑧)=2𝜋𝑖𝑖𝑖1𝑞𝑧𝜙(𝑞)𝑑𝑞.(2.18)
Since the operator is defined by a Cauchy type integral, it is readily observed that 𝜙(𝑧) constitutes a function analytic in the entire complex plane, except for points of the contour of integration Re𝑧=0. Also by Sokhotzki-Plemelj formula we have for Re𝑝=0+1𝜙=2𝜋𝑖𝑃𝑉𝑖𝑖11𝑞𝑝𝜙(𝑞)𝑑𝑞+2𝜙(𝑝),1𝜙=2𝜋𝑖𝑃𝑉𝑖𝑖11𝑞𝑝𝜙(𝑞)𝑑𝑞2𝜙(𝑝).(2.19) Here +𝜙 and 𝜙 are limits of 𝜙 as 𝑧 tends to 𝑝 from the left and right semi-plane, respectively.
We have for the Laplace transform 𝑢𝑥𝑥||𝑝||𝑝={𝑢}𝑢(0,𝑡)𝑝𝑢𝑥(0,𝑡)𝑝2.(2.20) Since {𝑢} is analytic for all Re𝑞>0, we have ̂𝑢(𝑞,𝑡)={𝑢}=̂𝑢(𝑝,𝑡).(2.21) Therefore applying the Laplace transform with respect to 𝑥 to problem (2.9) we obtain for 𝑡>0̂𝑢𝑡+𝐾(𝑝)̂𝑢(𝑝,𝑡)𝐾(𝑝)𝑝𝑢(0,𝑡)𝐾(𝑝)𝑝2𝑢𝑥(0,𝑡)=0,̂𝑢(𝑝,0)=̂𝑢0(𝑝),(2.22) where ||𝑝||𝐾(𝑝)=𝑝.(2.23) We rewrite (2.22) in the form ̂𝑢𝑡+𝐾(𝑝)̂𝑢(𝑝,𝑡)𝐾(𝑝)𝑝𝑢(0,𝑡)𝐾(𝑝)𝑝2𝑢𝑥(0,𝑡)=Φ(𝑝,𝑡),̂𝑢(𝑝,0)=̂𝑢0(𝑝),(2.24) with some function Φ(𝑝,𝑡) such that for all Re𝑝>0{Φ(𝑝,𝑡)}=0(2.25) and for |𝑝|>1||||1Φ(𝑝,𝑡)𝐶||𝑝||.(2.26) Applying the Laplace transformation with respect to time variable to problem (2.24) we find for Re𝑝>0̂1̂𝑢(𝑝,𝜉)=𝐾(𝑝)+𝜉̂𝑢0(𝑝)+𝐾(𝑝)𝑝̂𝑢(0,𝜉)+𝐾(𝑝)𝑝2̂𝑢𝑥(0,𝜉)+Φ(𝑝,𝜉).(2.27) Here the functions ̂̂𝑢(𝑝,𝜉),Φ(𝑝,𝜉),̂𝑢(0,𝜉), and ̂𝑢𝑥(0,𝜉) are the Laplace transforms for ̂𝑢(𝑝,𝑡),Φ(𝑝,𝑡),𝑢(0,𝑡), and 𝑢𝑥(0,𝑡) with respect to time, respectively. We will find the function Φ(𝑝,𝜉) using the analytic properties of function ̂̂𝑢 in the right-half complex planes Re𝑝>0 and Re𝜉>0. We have for Re𝑝=0̂1̂𝑢(𝑝,𝜉)=𝜋𝑖𝑃𝑉𝑖𝑖1̂𝑞𝑝̂𝑢(𝑞,𝜉)𝑑𝑞.(2.28) In view of Sokhotzki-Plemelj formula via (2.27) the condition (2.28) can be written as Θ+(𝑝,𝜉)=Λ+(𝑝,𝜉),(2.29) where the sectionally analytic functions Θ(𝑧,𝜉) and Λ(𝑧,𝜉) are given by Cauchy type integrals: 1Θ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑧1𝐾(𝑞)+𝜉Φ(𝑞,𝜉)𝑑𝑞,(2.30)Λ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝐾(𝑞)+𝜉̂𝑢0(𝑞)+𝐾(𝑞)𝑞̂𝑢(0,𝜉)+𝐾(𝑞)𝑞2̂𝑢𝑥(0,𝜉)𝑑𝑞.(2.31) To perform the condition (2.29) in the form of nonhomogeneous Riemann problem we introduce the sectionally analytic function: 1Ω(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝑞𝑧Ψ(𝑞,𝜉)𝑑𝑞,(2.32) where Ψ(𝑝,𝜉)=𝐾(𝑝)𝐾(𝑝)+𝜉Φ(𝑝,𝜉).(2.33) Taking into account the assumed condition (2.25) and making use of Sokhotzki-Plemelj formula (2.3) we get for limiting values of the functions Ω(𝑧,𝜉) and Θ(𝑧,𝜉)Ω(𝑝,𝜉)=𝜉Θ(𝑝,𝜉).(2.34) Also observe that from (2.30) and (2.32) by formula (2.4) 𝐾Θ(𝑝)+(𝑝,𝜉)Θ(𝑝,𝜉)=Ψ(𝑝,𝜉)=Ω+(𝑝,𝜉)Ω(𝑝,𝜉).(2.35) Substituting (2.29) and (2.34) into this equation we obtain nonhomogeneous Riemann problem Ω+(𝑝,𝜉)=𝐾(𝑝)+𝜉𝜉Ω(𝑝,𝜉)𝐾(𝑝)Λ+(𝑝,𝜉).(2.36)
It is required to find two functions for some fixed point 𝜉, Re𝜉>0: Ω+(𝑧,𝜉), analytic in Re𝑧<0 and Ω(𝑧,𝜉), analytic in Re𝑧>0, which satisfy on the contour Re𝑝=0 the relation (2.36). Here, for some fixed point 𝜉, Re𝜉>0, the functions 𝑊(𝑝,𝜉)=𝐾(𝑝)+𝜉𝜉,𝑔(𝑝,𝜉)=𝐾(𝑝)Λ+(𝑝,𝜉)(2.37) are called the coefficient and the free term of the Riemann problem, respectively.
Note that bearing in mind formula (2.33) we can find unknown function Φ(𝑝,𝜉) which involved in the formula (2.27) by the relation Φ(𝑝,𝜉)=𝐾(𝑝)+𝜉Ω𝐾(𝑝)+(𝑝,𝜉)Ω(.𝑝,𝜉)(2.38) The method for solving the Riemann problem 𝐴+(𝑝)=𝜑(𝑝)𝐴(𝑝)+𝜙(𝑝) is based on the Theorems 2.2 and 2.3.
In the formulations of Theorems 2.2 and 2.3 the coefficient 𝜑(𝑝) and the free term 𝜙(𝑝) of the Riemann problem are required to satisfy the Hölder condition on the contour Re𝑝=0. This restriction is essential. On the other hand, it is easy to observe that both functions 𝑊(𝑝,𝜉) and 𝑔(𝑝,𝜉) do not have limiting value as 𝑝±𝑖. The principal task now is to get an expression equivalent to the boundary value problem (2.36), such that the conditions of theorems are satisfied. First, let us introduce some notation and let us establish certain auxiliary relationships. Setting 𝐾1(𝑝)=𝑝2,(2.39) we introduce the function 𝑊(𝑝,𝜉)=𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+(𝑝),(2.40) where for some fixed point 𝑘(𝜉) (Re𝑘(𝜉)>0) 𝑤𝑧(𝑧)=𝑧+𝑘(𝜉)1/2,𝑤+𝑧(𝑧)=𝑧𝑘(𝜉)1/2.(2.41) We make a cut in the plane 𝑧 from point 𝑘(𝜉) to point through 0. Owing to the manner of performing the cut the functions 𝑤(𝑧), 𝐾1(𝑧) are analytic for Re𝑧>0 and the function 𝑤+(𝑧) is analytic for Re𝑧<0.
We observe that the function 𝑊(𝑝,𝜉) given on the contour Re𝑝=0 satisfies the Hölder condition and under the assumption Re𝐾1(𝑝)>0 does not vanish for any Re𝜉>0. Also we have 1Ind.𝑊(𝑝,𝜉)=2𝜋𝑖𝑖𝑖𝑑ln𝑊(𝑝,𝜉)=0.(2.42) Therefore in accordance with Theorem 2.3 the function 𝑊(𝑝,𝜉) can be represented in the form of the ratio 𝑋𝑊(𝑝,𝜉)=+(𝑝,𝜉)𝑋,(𝑝,𝜉)(2.43) where 𝑋±(𝑝,𝜉)=𝑒Γ±(𝑝,𝜉)1,Γ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝑞𝑧ln𝑊(𝑞,𝜉)𝑑𝑞.(2.44) Now we return to the nonhomogeneous Riemann problem (2.36). Multiplying and dividing the expression (𝐾(𝑝)+𝜉)/𝜉 by (1/(𝐾1(𝑝)+𝜉))(𝑤(𝑝)/𝑤+(𝑝)) and making use of the formula (2.43) we get 𝑊(𝑝,𝜉)=𝐾(𝑝)+𝜉𝜉=𝑌+(𝑝,𝜉)𝑌𝐾(𝑝,𝜉)1(𝑝)+𝜉𝜉,(2.45) where 𝑌±(𝑝,𝜉)=𝑋±(𝑝,𝜉)𝑤±(𝑝).(2.46) Replacing in (2.36) the coefficient of the Riemann problem 𝑊(𝑝,𝜉) by (2.45) we reduce the nonhomogeneous Riemann problem (2.36) to the form Ω+(𝑝,𝜉)𝑌+=𝐾(𝑝,𝜉)1(𝑝)+𝜉𝜉Ω(𝑝,𝜉)𝑌1(𝑝,𝜉)𝑌+(𝑝,𝜉)𝐾(𝑝)Λ+(𝑝,𝜉).(2.47) Now we perform the function Λ(𝑧,𝜉) given by formula (2.31) as Λ(𝑧,𝜉)=Λ1(𝑧,𝜉)+Λ2(𝑧,𝜉),(2.48) where Λ11(𝑧,𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝐾1(𝑞)+𝜉̂𝑢0𝐾(𝑞)+1(𝑞)𝑞𝐾̂𝑢(0,𝜉)+1(𝑞)𝑞2̂𝑢𝑥Λ(0,𝜉)𝑑𝑞,(2.49)21(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝐾𝑞𝑧1(𝑞)𝐾(𝑞)𝐾(𝐾(𝑞)+𝜉)1(𝑞)+𝜉̂𝑢0𝜉(𝑞)𝑞̂𝑢(0,𝜉)̂𝑢𝑥(0,𝜉)𝑑𝑞.(2.50) Firstly we calculate the left limiting value Λ+1(𝑝,𝜉). Since there exists only one root 𝑘(𝜉) of equation 𝐾1(𝑧)=𝜉 such that Re𝑘(𝜉)>0 for all Re𝜉>0, therefore, taking limit 𝑧𝑝 from the left-hand side of complex plane, by Cauchy theorem we get Λ+1𝑘(𝑝,𝜉)=(𝜉)𝑝𝑘(𝜉)̂𝑢0𝜉(𝑘(𝜉))𝑝̂𝑢(0,𝜉)̂𝑢𝑥(0,𝜉).(2.51) The last relation implies that (𝐾1(𝑝)+𝜉)Λ+1(𝑝,𝜉) can be expressed by the function Λ3(𝑧,𝜉) which is analytic in Re𝑧>0𝐾1Λ(𝑝)+𝜉+1(𝑝,𝜉)=Λ3(𝑝,𝜉),(2.52) where Λ3(𝑧,𝜉)=𝑘𝐾(𝜉)1(𝑧)+𝜉𝑧𝑘(𝜉)̂𝑢0𝜉(𝑘(𝜉))𝑝̂𝑢(0,𝜉)̂𝑢𝑥(0,𝜉).(2.53) By Sokhotzki-Plemelj formula (2.4) we express the left limiting value Λ+2(𝑝,𝜉) in the term of the right limiting value Λ2(𝑝,𝜉) as Λ+2(𝑝,𝜉)=Λ2(𝑝,𝜉)+̃𝑔1(𝑝,𝜉),(2.54) where ̃𝑔1𝐾(𝑝,𝜉)=1(𝑝)𝐾(𝑝)𝐾(𝐾(𝑝)+𝜉)1(𝑝)+𝜉̂𝑢0𝜉(𝑝)𝑝̂𝑢(0,𝜉)̂𝑢𝑥.(0,𝜉)(2.55) Bearing in mind the representation (2.48) and making use of (2.52), (2.55), and (2.45) after simple transformations we get 𝐾(𝑝)Λ+𝑌=+𝑌Λ3+𝐾1(Λ𝑝)+𝜉2+𝜉Λ+𝑔1(𝑝,𝜉),(2.56) where 𝑔1(𝑝,𝜉)=(𝐾(𝑝)+𝜉)̃𝑔1𝐾(𝑝,𝜉)=1(𝑝)𝐾(𝑝)𝐾1(𝑝)+𝜉̂𝑢0𝜉(𝑝)𝑝̂𝑢(0,𝜉)̂𝑢𝑥(0,𝜉).(2.57) Replacing in (2.47) 𝐾(𝑝)Λ+(𝑝,𝜉) by (2.56), we reduce the nonhomogeneous Riemann problem (2.47) in the form Ω+1(𝑝,𝜉)𝑌+=Ω(𝑝,𝜉)1(𝑝,𝜉)𝑌1(𝑝,𝜉)𝑌+𝑔(𝑝,𝜉)1(𝑝,𝜉),(2.58) where Ω+1(𝑝,𝜉)=Ω+(𝑝,𝜉)𝜉Λ+Ω(𝑝,𝜉),1𝐾(𝑝,𝜉)=1𝜉(𝑝)+𝜉1Ω(𝑝,𝜉)Λ2(𝑝,𝜉)Λ3(𝑝,𝜉).(2.59) In subsequent consideration we will have to use the following property of the limiting values of a Cauchy type integral, the statement of which we now quote. The proofs may be found in [19].Lemma 2.5. If 𝐿 is a smooth closed contour and 𝜙(𝑞) a function that satisfies the Hölder condition on 𝐿, then the limiting values of the Cauchy type integral 1Φ(𝑧)=2𝜋𝑖𝐿1𝑞𝑧𝜙(𝑞)𝑑𝑞(2.60) also satisfy this condition.
Since 𝑔1(𝑝,𝜉) satisfies on Re𝑝=0 the Hölder condition, on basis of this Lemma the function (1/𝑌+(𝑝,𝜉))𝑔1(𝑝,𝜉) also satisfies this condition. Therefore in accordance with Theorem 2.2 it can be uniquely represented in the form of the difference of the functions 𝑈+(𝑝,𝜉) and 𝑈(𝑝,𝜉), constituting the boundary values of the analytic function 𝑈(𝑧,𝜉), given by formula 1𝑈(𝑧,𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+𝑔(𝑞,𝜉)1(𝑞,𝜉)𝑑𝑞.(2.61) Therefore the problem (2.58) takes the form Ω+1(𝑝,𝜉)𝑌+(𝑝,𝜉)+𝑈+Ω(𝑝,𝜉)=1(𝑝,𝜉)𝑌(𝑝,𝜉)+𝑈(𝑝,𝜉).(2.62) The last relation indicates that the function (Ω+1/𝑌+)+𝑈+, analytic in Re𝑧<0, and the function (Ω1/𝑌)+𝑈, analytic in Re𝑧>0, constitute the analytic continuation of each other through the contour Re𝑧=0. Consequently, they are branches of unique analytic function in the entire plane. According to generalize Liouville theorem this function is some arbitrary constant 𝐴. Thus, bearing in mind the representations (2.59) and (2.52) we get Ω+(𝑝,𝜉)=𝑌+𝐴𝑈++𝜉Λ+,Ω𝜉(𝑝,𝜉)=𝐾1𝑌(𝑝)+𝜉(𝐴𝑈Λ)+𝜉+1+Λ2.(2.63) Since there exists only one root 𝑘(𝜉) of equation 𝐾1(𝑧)=𝜉 such that Re𝑘(𝜉)>0 for all Re𝜉>0, therefore, in the expression for the function Ω(𝑧,𝜉) the factor 𝜉/(𝐾1(𝑧)+𝜉) has a pole in the point 𝑧=𝑘(𝜉). Also the function 𝜉Λ+1 has a pole in the point 𝑧=𝑘(𝜉). Thus in general case the problem (2.36) is insolvable. It is soluble only when the functions 𝑈(𝑧,𝜉) and 𝜉Λ+1 satisfy additional conditions. For analyticity of Ω(𝑧,𝜉) in points 𝑧=𝑘(𝜉) it is necessary that Res𝑝=𝑘(𝜉)1𝐾1𝑌(𝑝)+𝜉(𝐴𝑈)+Λ+1=0.(2.64) We reduce (2.64) to the form 𝐴𝑌(𝑘(𝜉),𝜉)𝑌(𝑘(𝜉),𝜉)𝑈𝜉(𝑘(𝜉),𝜉)+𝑘(𝜉)̂𝑢(0,𝜉)+̂𝑢0(𝑘(𝜉))̂𝑢𝑥(0,𝜉)=0.(2.65) Multiplying the last relation by 1/𝑌(𝑘(𝜉),𝜉) and taking limit 𝜉 we get that 𝐴=0. This implies that for solubility of the nonhomogeneous problem (2.36) it is necessary and sufficient that the following condition is satisfied: 𝑌(𝑘(𝜉),𝜉)𝑈𝜉(𝑘(𝜉),𝜉)+𝑘(𝜉)̂𝑢(0,𝜉)̂𝑢0(𝑘(𝜉))+̂𝑢𝑥(0,𝜉)=0.(2.66)
Therefore, we need to put in the problem (2.9) one boundary data and the rest of boundary data can be found from (2.66). Thus, for example, if we put 𝑢(0,𝑡)=0 from (2.66) we obtain for the Laplace transform of 𝑢𝑥(0,𝑡),[𝑌](𝑘(𝜉),𝜉)𝐼(𝑘(𝜉),𝜉)1̂𝑢𝑥(0,𝜉)=̂𝑢0(𝑘(𝜉))𝑌1(𝑘(𝜉),𝜉)2𝜋𝑖𝑖𝑖1𝑞𝑘(𝜉)̂𝑢0(𝑞)𝑌+𝐾(𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉𝑑𝑞,(2.67) where 1𝐼(𝑘(𝜉),𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌+𝐾(𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉𝑑𝑞.(2.68) Now we prove that the coefficient of ̂𝑢𝑥(0,𝜉) does not vanish for all Re𝜉>0. We represent the function 𝐼(𝑘(𝜉),𝜉) in the form 𝐼=𝐼1+𝐼2,(2.69) where 𝐼1=12𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌+𝐼(𝑞,𝜉)𝑑𝑞,21=2𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌+(𝑞,𝜉)𝐾(𝑞)+𝜉𝐾1(𝑞)+𝜉𝑑𝑞.(2.70) Since, for Re𝑧0,𝑖𝑖1𝑞𝑧𝑑𝑞=𝜋𝑖sgn(Re𝑧),(2.71) making use of analytic properties of the function ((1/𝑌+(𝑞,𝜉))1) by Cauchy Theorem we have 𝐼1=12𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑑𝑞+2𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌+1(𝑞,𝜉)1𝑑𝑞=2,(2.72) where 𝜉 is some fixed point, Re𝜉>0. To calculate the function 𝐼2 we will use the identity (2.43). Observe that the function 1/𝑌(𝑞,𝜉) is analytic for all Re𝑞>0. Therefore, setting the relation (2.43) into definition of 𝐼2 and making use of Cauchy Theorem we find 𝐼21=2𝜋𝑖𝑖𝑖11𝑞𝑘(𝜉)𝑌1(𝑞,𝜉)𝑑𝑞=𝑌+1(𝑘(𝜉),𝜉)12.(2.73) Thus, from (2.72) and (2.73) we obtain the following relation for the function 𝐼1𝐼(𝑘(𝜉),𝜉)=𝑌(𝑘(𝜉),𝜉)1.(2.74) Substituting this formula into (2.67) we get ̂𝑢𝑥(0,𝜉)=̂𝑢0(𝑘(𝜉))𝑌1(𝑘(𝜉),𝜉)2𝜋𝑖𝑖𝑖1𝑞𝑘(𝜉)̂𝑢0(𝑞)𝑌+𝐾(𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉𝑑𝑞.(2.75) Now we return to problem (2.36). From (2.63) under the conditions 𝑢(0,𝑡)=0 and (2.75) the limiting values of solution of (2.36) are given by Ω+(𝑝,𝜉)=𝑌+𝑈++𝜉Λ+2,Ω𝜉(𝑝,𝜉)=𝐾1(𝑌𝑝)+𝜉𝑈+𝜉Λ2.(2.76) From (2.76) with the help of the integral representations (2.61) and (2.50), for sectionally analytic functions 𝑈(𝑧,𝜉) and Λ2(𝑧,𝜉), making use of Sokhotzki-Plemelj formula (2.3) and relation (2.45) we can express the difference limiting values of the function Ω(𝑧,𝜉) in the form Ω+(𝑝,𝜉)Ω(𝑝,𝜉)=𝑌+𝑈++𝜉𝐾1(𝑌𝑝)+𝜉𝑈Λ+𝜉+2Λ2=𝑌+𝑈+𝜉𝑈𝐾(𝑝)+𝜉Λ+𝜉+2Λ21=2𝐾(𝑝)𝑔𝐾(𝑝)+𝜉1(𝑝,𝜉)𝐾(𝑝)𝑌𝐾(𝑝)+𝜉+1(𝑝,𝜉)2𝜋𝑖𝑃𝑉𝑖𝑖11𝑞𝑝𝑌+𝑔(𝑞,𝜉)1(𝑞,𝜉)𝑑𝑞.(2.77) We now proceed to find the unknown function Φ(𝑝,𝜉) involved in the formula (2.27) for the solution ̂̂𝑢(𝑝,𝜉) of the problem (2.9). Replacing the difference Ω+(𝑝,𝜉)Ω(𝑝,𝜉) in the relation (2.38) by formula (2.77) we get Φ(𝑝,𝜉)=𝐾(𝑝)+𝜉ΩK(𝑝)+(𝑝,𝜉)Ω(1𝑝,𝜉)=2𝑔1(𝑝,𝜉)𝑌+1(𝑝,𝜉)2𝜋𝑖𝑃𝑉𝑖𝑖11𝑞𝑝𝑌+𝑔(𝑞,𝜉)1(𝑞,𝜉)𝑑𝑞.(2.78) It is easy to observe that Φ(𝑝,𝜉) is boundary value of the function analytic in the left complex semi-plane and therefore satisfies our basic assumption for all Re𝑧>0{Φ}=0.(2.79) Having determined the function Φ(𝑝,𝜉) bearing in mind formula (2.27) and conditions 𝑢(0,𝑡)=0 we determine required function ̂̂𝑢̂1̂𝑢(𝑝,𝜉)=𝐾(𝑝)+𝜉̂𝑢0(𝑝)̂𝑢𝑥1(0,𝜉)21𝐾𝑔(𝑝)+𝜉11(𝑝,𝜉)𝑌𝐾(𝑝)+𝜉+(1𝑝,𝜉)2𝜋𝑖𝑃𝑉𝑖𝑖11𝑞𝑝𝑌+𝑔(𝑞,𝜉)1(𝑞,𝜉)𝑑𝑞,(2.80) where the function 𝑔1(𝑝,𝜉) is given by formula (2.55): 𝑔1𝐾(𝑝,𝜉)=1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉̂𝑢0(𝑞)̂𝑢𝑥.(0,𝜉)(2.81) Now we prove that, in accordance with last relation, the function ̂̂𝑢(𝑝,𝜉) constitutes the limiting value of an analytic function in Re𝑧>0.
With the help of the integral representations (2.61), (2.31), and (2.50) for sectionally analytic functions 𝑈(𝑧,𝜉),Λ(𝑧,𝜉), and Λ2(𝑧,𝜉), and making use of Sokhotzki-Plemelj formula (2.3) we have 1𝐾(𝑝)+𝜉̂𝑢0(𝑝)̂𝑢𝑥(0,𝜉)=Λ+Λ,1𝐾𝑔(𝑝)+𝜉1(𝑝,𝜉)=Λ+2Λ2,12𝜋𝑖PV𝑖𝑖11𝑞𝑝𝑌+𝑔(𝑞,𝜉)1(1𝑞,𝜉)𝑑𝑞=2𝑈++𝑈.(2.82) Substituting these relations into (2.80) we express the function ̂̂𝑢 in the following form: ̂Λ̂𝑢=+Λ12Λ+2Λ2121𝐾𝑌(𝑝)+𝜉+𝑈++𝑈.(2.83) If it is taken into account that Λ(𝑧,𝜉)=Λ1(𝑧,𝜉)+Λ2(𝑧,𝜉) by virtue of the relation (2.45), the last expression agrees with formula ̂̂𝑢=Λ+1Λ1+12Λ+2Λ2121𝐾𝑌(𝑝)+𝜉+𝑈+121𝐾1𝑌(𝑝)+𝜉𝑈.(2.84) Expressing the function 𝑈+ in the last equation in terms of 𝑈𝑈+=𝑈+1𝑌+Λ(𝐾(𝑝)+𝜉)+2Λ2,(2.85) we arrive at the following relation: ̂̂𝑢=Λ+1Λ11𝐾1𝑌(𝑝)+𝜉𝑈,(2.86) where by virtue of (2.49) and (2.66), Λ+1Λ1=1𝐾1(𝑝)+𝜉̂𝑢0(𝑝)̂𝑢𝑥.(0,𝜉)(2.87) Thus the function ̂̂𝑢 is the limiting value of an analytic function in Re𝑧>0. Note the fundamental importance of the proven fact that the solution ̂̂𝑢 constitutes an analytic function in Re𝑧>0 and, as a consequence, its inverse Laplace transform vanishes for all 𝑥<0. We now return to solution 𝑢(𝑥,𝑡) of the problem (2.9).
Under assumption 𝑢(0,𝑡)=0 the integral representation (2.61) takes form 1𝑈(𝑧,𝜉)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+𝐾(𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1(𝑞)+𝜉̂𝑢0(𝑞)̂𝑢𝑥(0,𝜉)𝑑𝑞,(2.88) where ̂𝑢𝑥(0,𝜉) is defined by (2.75). Substituting this relation into (2.86) and taking inverse Laplace transform with respect to time and inverse Fourier transform with respect to space variables we obtain 𝑢(𝑥,𝑡)=𝒢(𝑡)𝑢0=0𝐺(𝑥,𝑦,𝑡)𝑢0(𝑦)𝑑𝑦,(2.89) where the function 𝐺(𝑥,𝑦,𝑡) was defined by formula (2.12). Proposition 2.4 is proved.

Now we collect some preliminary estimates of the Green operator 𝒢(𝑡). Let the contours 𝒞𝑖 be defined as

𝒞1=𝑝𝑒𝑖(𝜋/2+𝜀),00,𝑒𝑖(𝜋/2+𝜀)𝒞,(2.90)2=𝑞𝑒𝑖((𝜋/2)+2𝜀),00,𝑒𝑖((𝜋/2)+2𝜀)𝒞,(2.91)3=𝑞𝑒𝑖((𝜋+𝜀)/2),00,𝑒𝑖((𝜋+𝜀)/2),(2.92) where 𝜀>0 can be chosen such that all functions under integration are analytic and Re𝑘(𝜉)>0 for 𝜉𝒞1.

Lemma 2.6. The function 𝐺(𝑥,𝑦,𝑡) given by formula (2.12) has the following representation: 1𝐺(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑒𝑝𝑦1𝑑𝑝12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥𝑒Γ(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾(𝑝)+𝜉𝐼(𝑝,𝜉,𝑦)𝑑𝑝,(2.93) where 1𝐼(𝑝,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑝)(𝑞𝑘(𝜉))𝑒Γ+(𝑞,𝜉)𝑤+𝑒(𝑞,𝜉)𝑞𝑦1𝑑𝑞,Γ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝑞𝑧ln𝐾(𝑞)+𝜉𝐾1𝑤(𝑞)+𝜉(𝑞)𝑤+(𝑞)𝑑𝑞.(2.94) The functions 𝑤±(𝑞,𝜉),𝑘(𝜉) were defined in formulas (2.13) and (2.10).

Proof. We rewrite formula (2.12) in the form 𝐺(𝑥,𝑦,𝑡)=𝐽1(𝑥𝑦,𝑡)+𝐽2(𝑥,𝑦,𝑡)+𝐽3(𝑥,𝑦,𝑡)+𝐽4(𝑥,𝑦,𝑡),(2.95) where 𝐽11(𝑥𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝(𝑥𝑦)𝐾1(𝑝)𝑡𝐽𝑑𝑝,21(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖+𝜀𝑖+𝜀𝑑𝜉𝑒𝜉𝑡Ψ(𝜉,𝑦)𝑖𝑖𝑒𝑝𝑥1𝐾1(𝐽𝑝)+𝜉𝑑𝑝,31(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖+𝜀𝑖+𝜀𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾1Υ(𝑝)+𝜉1𝐽(𝑝,𝜉,𝑦)𝑑𝑝,(2.96)41(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖+𝜀𝑖+𝜀𝑑𝜉𝑒𝜉𝑡Ψ(𝜉,𝑦)𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾1Υ(𝑝)+𝜉1(𝑝,𝜉,0)𝑑𝑝.(2.97) Here Υ11(𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+𝐾(𝑞,𝜉)1(𝑞)𝐾(𝑞)𝐾1𝑒(𝑞)+𝜉𝑞𝑦𝑒𝑑𝑞,Ψ(𝜉,𝑦)=𝑘(𝜉)𝑦𝑌(𝑘(𝜉),𝜉)+Υ1𝑌(𝑘(𝜉),𝜉,𝑦),±=𝑒Γ±𝑤±,1(2.98)Γ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝐾𝑞𝑧ln(𝑞)+𝜉𝐾1(𝑤𝑞)+𝜉(𝑞)𝑤+(𝑞)𝑑𝑞.(2.99) Firstly we consider the sectionally analytic function Υ1(𝑧,𝜉,𝑦) given by Cauchy type integral (2.98).
On basis of the definition (2.98) its limiting value can be represent in the form Υ1(𝑝,𝜉,𝑦)=𝐼1(𝑝,𝜉,𝑦)+𝐼2(𝑝,𝜉,𝑦),(2.100) where 𝐼1(𝑝,𝜉,𝑦)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+𝑒(𝑞,𝜉)𝑞𝑦𝐼𝑑𝑞,2(𝑝,𝜉,𝑦)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+(𝑞,𝜉)𝐾(𝑞)+𝜉𝐾1𝑒(𝑞)+𝜉𝑞𝑦𝑑𝑞.(2.101) Making use of analytic properties of the functions (1/𝑌+(𝑞,𝜉)1), for Re𝑞<0, and 𝑒𝑞𝑦, for Re𝑞>0, by Cauchy theorem we have 𝐼1(𝑝,𝜉,𝑦)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖1𝑒𝑞𝑧𝑞𝑦𝑑𝑞+lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+(𝑞,𝜉)1(𝑒𝑞𝑦1)𝑑𝑞+lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+(𝑞,𝜉)1𝑑𝑞=𝑒𝑝𝑦+lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+(𝑞,𝜉)1(𝑒𝑞𝑦1)𝑑𝑞,(2.102) where 𝜉 is some fixed point, Re𝜉>0.
To calculate the function 𝐼2(𝑝,𝜉,𝑦) we will use the following identity: 1𝑌+(𝑞,𝜉)𝐾(q)+𝜉𝐾1=1(𝑞)+𝜉𝑌.(𝑞,𝜉)(2.103) Observe that the function 1/𝑌(𝑞,𝜉) is analytic for all Re𝑞>0. Therefore, setting the relation (2.103) into definition of 𝐼2(𝑝,𝜉,𝑦) and making use of Cauchy theorem we find 𝐼2(𝑝,𝜉,𝑦)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧𝑌𝑒(𝑞,𝜉)𝑞𝑦1𝑑𝑞=𝑌𝑒(𝑝,𝜉)𝑝𝑦.(2.104) Thus from (2.102) and (2.104) we obtain the following relation: Υ1(𝑝,𝜉,𝑦)=𝑒𝑝𝑦+Υ1(𝑝,𝜉,𝑦)+𝑌𝑒(𝑝,𝜉)𝑝𝑦,(2.105) where 1Υ(𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11𝑞𝑧𝑌+𝑒(𝑞,𝜉)1𝑞𝑦𝑑𝑞.(2.106) In the same way (see also proof of relation (2.74)) we can prove that Υ11(𝑝,𝜉,0)=1+𝑌.(𝑝,𝜉)(2.107) Also we observe that Ψ(𝜉,𝑦)=𝑒𝑘(𝜉)𝑦+Υ(𝑘(𝜉),𝜉,𝑦).(2.108) Inserting into definition (2.96) the expression (2.105) for Υ1(𝑝,𝜉,𝑦) we obtain the function 𝐽3(𝑥,𝑦,𝑡) in the form 𝐽31(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾1(𝑝)+𝜉(𝑒𝑝𝑦+Υ(𝑝,𝜉,𝑦))𝑑𝑝𝐽1(𝑥𝑦,𝑡).(2.109) Replacing in formula (2.97) the functions Ψ(𝜉,𝑦) and Υ1(𝑝,𝜉,0) by (2.108) and (2.107), respectively, we reduce the function 𝐽4(𝑥,𝑦,𝑡) in the form 𝐽41(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖𝑖𝑑𝜉𝑒𝜉𝑡Ψ(𝜉,𝑦)𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾11(𝑝)+𝜉1+𝑌1(𝑝,𝜉)𝑑𝑝=12𝜋𝑖2𝜋𝑖𝑖𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾1(𝑒𝑝)+𝜉𝑘(𝜉)𝑦Υ(𝑘(𝜉),𝜉,𝑦)𝑑𝑝𝐽2(𝑥,𝑦).(2.110) Therefore inserting into definition (2.95) expressions (2.109) and (2.110), for 𝐽3(𝑥,𝑦,𝑡) and 𝐽4(𝑥,𝑦,𝑡), respectively, we obtain the function 𝐺(𝑥,𝑦,𝑡) in the form 1𝐺(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝑖𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)𝐾1Ξ(𝑝)+𝜉1(𝑝,𝜉,𝑦)𝑑𝑝,(2.111) where Ξ1𝑒(𝑝,𝜉,𝑦)=𝑝𝑦𝑒𝑘(𝜉)𝑦(Υ(𝑝,𝜉,𝑦)Υ(𝑘(𝜉),𝜉,𝑦)).(2.112) Also, note that since Υ(𝑝,𝜉,𝑦)Υ(𝑘(𝜉),𝜉,𝑦)=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖11𝑞𝑧1𝑞𝑘(𝜉)𝑌+𝑒(𝑞,𝜉)1𝑞𝑦𝑑𝑞=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖𝑧𝑘(𝜉)(1𝑞𝑧)(𝑞𝑘(𝜉))𝑌+(𝑒𝑞,𝜉)1𝑞𝑦𝑑𝑞,(2.113) we obtain 𝑒𝑝𝑦𝑒𝑘(𝜉)𝑦(Υ(𝑝,𝜉,𝑦)Υ(𝑘(𝜉),𝜉,𝑦))=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖1+1𝑞𝑧𝑒𝑞𝑘(𝜉)𝑞𝑦+1𝑌+𝑒(𝑞,𝜉)1𝑞𝑦𝑑𝑞=lim𝑧𝑝,Re𝑧>012𝜋𝑖𝑖𝑖𝑘(𝜉)𝑧1(𝑞𝑧)(𝑞𝑘(𝜉))𝑌+𝑒(𝑞,𝜉)𝑞𝑦𝑑𝑞.(2.114) So, 1𝐺(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1𝐼(𝑝)+𝜉(𝑝,𝜉,𝑦)𝑑𝑝,(2.115) where 1𝐼(𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑧)(𝑞𝑘(𝜉))𝑌+𝑒(𝑞,𝜉)𝑞𝑦𝑑𝑞.(2.116) Using relation 𝐾(𝑝)+𝜉𝐾1=𝑌(𝑝)+𝜉+𝑌,(2.117) we rewrite last formula in the following form: 1𝐺(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉(𝑝,𝜉,𝑦)𝑑𝑝.(2.118) On the basis of definitions (2.116) and in accordance with the Sohkotzki-Plemelj formula (2.3) we have 𝐼(𝑝,𝜉,𝑦)=𝐼+1(𝑝,𝜉,𝑦)1(𝑝𝑘(𝜉))𝑌+𝑒(𝑝,𝜉)𝑝𝑦.(2.119) So we get 1𝐺(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑒𝑝𝑦1𝑑𝑝12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉+(𝑝,𝜉,𝑦)𝑑𝑝.(2.120) Now we consider for Re𝜉>0Γ+1(𝑝,𝜉)=2𝜋𝑖lim𝑧𝑝Re𝑧<0𝑖𝑖1𝑞𝑧ln𝐾(𝑞)+𝜉𝐾1𝑤(𝑞)+𝜉(𝑞)𝑤+=1(𝑞)𝑑𝑞2𝜋𝑖𝑒𝑖((𝜋/2)+3𝜀)01𝑞𝑝ln𝑖𝑞2+𝜉𝑞2𝑤+𝜉(𝑞)𝑤++1(𝑞)𝑑𝑞2𝜋𝑖0𝑒𝑖((𝜋/2)3𝜀)1𝑞𝑝ln𝑖𝑞2+𝜉𝑞2𝑤+𝜉(𝑞)𝑤+(𝑞)𝑑𝑞+ln𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+(𝑝)=Γ1(𝑝,𝜉)+ln𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+.(𝑝)(2.121) Note that Γ1(𝑝,𝜉) is analytic in domain 0arg𝜉<(𝜋/2)+3𝜀, (𝜋/2)3𝜀<arg𝜉0, and 𝜋/2arg𝑝<(𝜋/2)+3𝜀 and (𝜋/2)3𝜀<arg𝑝𝜋/2. So we get 1𝐺(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑒𝑝𝑦1𝑑𝑝12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑒Γ1(𝑝,𝜉)𝑤(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1(𝐼𝑝)+𝜉+(𝑝,𝜉,𝑦)𝑑𝑝,(2.122) where 𝐼+1(𝑝,𝜉,𝑦)=2𝜋𝑖lim𝑧𝑝Re𝑧<0𝑖𝑖11(𝑞𝑧)(𝑞𝑘(𝜉))𝑤𝑒Γ1(𝑞,𝜉)𝐾1(𝑞)+𝜉𝑒𝐾(𝑞)+𝜉𝑞𝑦𝑑𝑞.(2.123) Changing the contour of integration with respect to 𝜉 by Cauchy theorem we get 1𝐺(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑒𝑝𝑦1𝑑𝑝12𝜋𝑖2𝜋𝑖𝒞1d𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥𝑒Γ1(𝑝,𝜉)𝑤(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1(𝐼𝑝)+𝜉+(𝑝,𝜉,𝑦)𝑑𝑝res𝜉=𝐾(𝑞)𝒞2𝑒𝑝𝑥𝑒Γ1(𝑝,𝜉)𝑤(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1𝐼(𝑝)+𝜉+,(𝑝,𝜉,𝑦)𝑑𝑝(2.124) where res𝜉=𝐾(𝑞)𝒞2𝑒𝑝𝑥𝑒Γ1(𝑝,𝜉)𝑤(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1𝐼(𝑝)+𝜉+=1(𝑝,𝜉,𝑦)𝑑𝑝2𝜋𝑖𝒞2𝑒𝑝𝑥𝑑𝑝𝒞3𝑒𝑑𝑞𝑞𝑦𝐾(𝑞)𝑡𝑒𝑞𝑝𝜙(𝑝,𝑞),𝜙(𝑝,𝑞)=Γ1(𝑝,𝐾(𝑞))𝑤(𝑝)(𝑝𝑘(𝐾(𝑞)))𝐾1𝐾(𝑝)𝐾(𝑞)1(𝑞)𝐾(𝑞)𝑒Γ1(𝑞,𝐾(𝑞))𝑤(.𝑞)(𝑞𝑘(𝐾(𝑞)))(2.125) Since for 𝑞𝒞3,Re(𝐾(𝑞))<0, from (2.121) 𝑒Γ1(𝑞,𝐾(𝑞))=𝑒Γ+(𝑞,𝐾(𝑞))𝐾(𝑞)𝐾(𝑞)𝐾1𝑤(𝑞)𝐾(𝑞)(𝑞)𝑤+(𝑞)=0,(2.126) and therefore 𝜙(𝑝,𝑞)=0.(2.127)
Thus using relation (2.121) we get relation (2.93). Lemma is proved.

Lemma 2.7. The estimates are true, provided that the right-hand sides are finite: 𝒢(𝑡)𝜙𝐇1𝜙𝐇1,𝜕(2.128)𝑛𝑥𝒢𝜙𝐋𝑠,𝜇𝐶𝑡(1/2)(𝑛+1+𝛿(1/𝑠)𝜇)𝜙𝐋1,𝛿,(2.129)𝒢(𝑡)𝜙𝑡1Λ𝑥𝑡1/2𝑓(𝜙)=min𝑥𝑡1/2𝑡,11𝑎𝜙𝐋1,1+𝑎,𝑎>0,(2.130) where 𝜇>0, (1/𝑠)+𝜇<𝑛+1,𝑛=0,1,𝛿[0,1], and 𝑓(𝜙) is given by (1.9). The function Λ(𝑥𝑡1/2)𝐋(𝐑+),Λ(0)=0 is defined below (see (2.191)).

Proof. From Lemma 2.6 we have 𝐺(𝑥,𝑦,𝑡)=𝐽1(𝑥,𝑦,𝑡)+𝐽2(𝑥,𝑦,𝑡),(2.131) where 𝐽11(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑒𝑝𝑦𝐽𝑑𝑝,(2.132)21(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥𝑤+(𝑝,𝜉)𝑒Γ(𝑝,𝜉)(𝑝𝑘(𝜉))1𝐾(𝑝)+𝜉𝐼(𝑝,𝜉,𝑦)𝑑𝑝,(2.133)𝐼(𝑝,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑝)(𝑞𝑘(𝜉))𝑤+𝑒Γ+𝑒𝑞𝑦1𝑑𝑞,(2.134)Γ(𝑧,𝜉)=2𝜋𝑖𝑖𝑖1𝑞𝑧ln𝐾(𝑞)+𝜉𝐾1(𝑤𝑞)+𝜉(𝑞)𝑤+(𝑞)𝑑𝑞.(2.135) The functions 𝑤±(𝑞,𝜉),𝑘(𝜉) were defined in formulas (2.13) and (2.10). The contours 𝒞1,𝒞2 was defined by (2.90), and (2.91).
For subsequent considerations it is required to investigate the behavior of the function Γ(𝑧,𝜉). Set 𝜙(𝑝,𝜉)=ln𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+(𝑝)0,Re𝑝=0,Re𝜉<0.(2.136) Observe that the function 𝜙(𝑝,𝜉) obeys the Hölder condition for all finite 𝑝 and tends to a definite limit 𝜙(𝜉) as 𝑝±𝑖𝜙(,𝜉)=lim𝑝±𝑖ln𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+(𝑝)=lim𝑝±𝑖||||ln𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝)𝑤+||||(𝑝)+𝑖lim𝑝±𝑖arg𝐾(𝑝)+𝜉𝐾1𝑤(𝑝)+𝜉(𝑝,𝜉)𝑤+𝜋(𝑝,𝜉)=𝑖2.(2.137) Also there can be easily obtained that for large 𝑝 and some fixed 𝜉 the following inequality holds: ||𝜙(𝑝,𝜉)𝜙||||𝜉||(𝜉)𝐶𝜇||𝑝||2𝜇,𝜇>0.(2.138) Therefore ||𝑒Γ±(𝑧,𝜉)||𝐶(2.139) for all 𝜉𝒞1 and 𝑒Γ+(𝑝,𝜉)𝑒1/2||𝜉||=𝑂𝜇||𝑝||2𝜇[].,𝜇0,1(2.140) Denote 𝒥𝑗(𝑡)𝜙=𝜃(𝑥)0+𝐽𝑗(𝑥,𝑦,𝑡)𝜙(𝑦)𝑑𝑦,(2.141) By Plansherel Theorem it is easily to see that 𝒥1(𝑡)𝜙𝐇1𝐶𝜙𝐇1.(2.142)
Now we estimate 𝒥2(𝑡)𝜙.
From the integral representation (2.134), making use the relation 𝑒Γ+(𝑝,𝜉)Γ(𝑝,𝜉)𝑤+𝑤=𝐾(𝑝)+𝜉𝐾1(𝑝)+𝜉(2.143) and the estimate (2.138) we have for 𝑦>0,𝜉𝒞1, and 𝑝𝒞21𝐼(𝑝,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑝)1(𝑞𝑘(𝜉))𝑤+𝑒Γ+𝑒𝑞𝑦𝑑𝑞=lim𝑧𝑝,Re𝑧<012𝜋𝑖𝑖𝑖11(𝑞𝑧)1(𝑞𝑘(𝜉))𝑤𝑒Γ𝐾1(𝑞)+𝜉𝑒(𝐾(𝑞)+𝜉)𝑞𝑦𝑦𝑑𝑞=𝑂1+𝛾𝒞3||𝑞||1+𝛾||||1𝑞𝑝||||.𝑞𝑘(𝜉)𝑑𝑞(2.144) Here 𝛾[0,1),𝐾(𝑞)=𝑞|𝑞|=𝑞2exp(𝑖𝜃),𝜃=arg𝑞 and 𝒞3=𝑞𝑒𝑖((𝜋𝜀)/2),00,𝑒𝑖((𝜋𝜀)/2),𝜀>0.(2.145) After this observation in accordance with the integral representation (2.133) by the Hölder inequality we have arrived at the following estimate for 𝑛=0,1𝜕𝑛𝑥0+𝐽2(,𝑦,𝑡)𝜙(𝑦)𝑑𝑦𝐋2𝐶𝜙𝐇1𝒞1𝑑𝜉𝑒𝐶|𝜉|𝑡𝒞2||𝑝||𝑑𝑝𝑛||||𝑝𝑘(𝜉)||||𝐾(𝑝)+𝜉𝒞3||𝑞||1+𝛾||||1𝑞𝑝||||𝑞𝑘(𝜉)𝑑𝑞𝐶𝜙𝐇1.(2.146) Therefore according to (2.142) and (2.146) we obtain the estimate (2.129) of the lemma. Now we prove the second estimate of the lemma.
We rewrite the function 𝐼(𝑝,𝜉,𝑦) (see (2.134)) for 𝑝𝒞2,𝜉𝒞1 in the form 𝐼(𝑝,𝜉,𝑦)=𝐼11(𝑝,𝜉,𝑦)+1(𝑝𝑘(𝜉))𝑤+(𝑝,𝜉)𝑒Γ(𝑝,𝜉)+𝑒1/21𝑒(𝑝𝑘(𝜉))𝑘(𝜉)𝑦,1(2.147) where 𝐼11(𝑝,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑝)(𝑞𝑘(𝜉))𝑤+(𝑞,𝜉)𝑒Γ+(𝑞,𝜉)𝑒1/2(𝑒𝑞𝑦1)𝑑𝑞.(2.148) Here we used that since 𝑤+(𝑞,𝜉)𝑒Γ+(𝑞,𝜉) is analytic for 𝑝𝒞2,𝜉𝒞1, and Re𝑘(𝜉)>0 by Cauchy theorem, 12𝜋𝑖𝑖𝑖11(𝑞𝑝)(𝑞𝑘(𝜉))𝑤+(𝑞,𝜉)𝑒Γ+(𝑞,𝜉)1𝑑𝑞=1(𝑝𝑘(𝜉))𝑤+(𝑝,𝜉)𝑒Γ+(𝑝,𝜉)𝑒1/22𝜋𝑖𝑖𝑖1(𝑞𝑝)(𝑞𝑘(𝜉))(𝑒𝑞𝑦1)𝑑𝑞=𝑒1/21𝑒(𝑝𝑘(𝜉))𝑘(𝜉)𝑦.1(2.149) Substituting (2.147) into definition of the function 𝐽2(𝑥,𝑦,𝑡) (see (2.133)) we get 𝐽2(1𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥𝑌+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉1(1𝑝,𝜉,𝑦)𝑑𝑝12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥1𝑒𝐾(𝑝)+𝜉𝑑𝑝1/22𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝑒𝑘(𝜉)𝑦1𝒞2𝑒𝑝𝑥𝑌+(𝑝,𝜉)𝐾(𝑝)+𝜉𝑑𝑝.(2.150) Since by Cauchy theorem 112𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥11𝐾(𝑝)+𝜉𝑑𝑝=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡𝑑𝑝,(2.151) we obtain the following form for Green function 𝐺(𝑥,𝑦,𝑡) (see (2.131)): 𝐺(𝑥,𝑦,𝑡)=3𝑗=1𝐹𝑗(𝑥,𝑦,𝑡),(2.152) where 𝐹11(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝑝𝑥𝐾(𝑝)𝑡(𝑒𝑝𝑦𝐹1)𝑑𝑝,(2.153)21(𝑥,𝑦,𝑡)=𝑒2𝜋𝑖1/22𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝑒𝑘(𝜉)𝑦1𝒞2𝑒𝑝𝑥𝑌+(𝑝,𝜉)𝐹𝐾(𝑝)+𝜉𝑑𝑝,(2.154)3(1𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2𝑒𝑝𝑥𝑌+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉1(𝐼𝑝,𝜉,𝑦)𝑑𝑝,(2.155)1(1𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑧)(𝑞𝑘(𝜉))𝑌+(𝑞,𝜉)𝑒1/2(𝑒𝑞𝑦1)𝑑𝑞.(2.156) Denote 𝑗𝜙=0+𝐹𝑗(𝑥,𝑦,𝑡)𝜙(𝑦)𝑑𝑦.(2.157) Now we prove estimate 𝜕𝑛𝑥1𝜙𝐋𝑠,𝜇𝐶𝑡(1/2)(𝑛+1+𝛿(1/𝑠)𝜇)𝜙𝐋1,𝛿.(2.158)
We have 𝐹11(𝑥,𝑦,𝑡)=2𝜋𝑖𝑡𝑖𝑖𝑒|𝑝|𝑝𝑒𝑝𝑥𝑡1/2𝑒𝑝𝑦𝑡1/2=11𝑑𝑝𝜋𝑡Re0𝑒𝑖𝑝2𝑖𝑝𝑦𝑡1/2𝑒𝑖𝑝𝑥𝑡1/21𝑒𝑖𝑝𝑦𝑡1/2=1𝑑𝑝𝜋𝑡Re𝑦/𝑡0𝑒𝑖𝑝2𝑒𝑖𝑝𝑥𝑡1/2𝑒𝑖𝑝𝑦𝑡1/2+11𝑑𝑝𝜋𝑡Re𝑒𝑖𝜙𝑦/𝑡𝑒𝑖𝑝2𝑖𝑝𝑦𝑡1/2𝑒𝑖𝑝𝑥𝑡1/21𝑒𝑖𝑝𝑦𝑡1/2𝑑𝑝=𝐼11+𝐼12,(2.159) where 𝜙>0 is such that Re𝑖𝑝2>0.
By Bonnet Theorem we get ||𝜕𝑛𝑥𝐼11||𝐶𝑡(1/2)(1+𝑛)|||1𝑒𝑖y2𝑡1||||||||𝑦/𝑡𝑘𝑒𝑖𝑝2𝑒𝑖𝑝𝑥𝑡1/2𝑝𝑛|||||𝑑𝑝𝐶𝑡(1/2)(1+𝑛)|||1𝑒𝑖𝑦2𝑡1|||||||𝑘𝑒𝑖𝑝2𝑒𝑖𝑝𝑥𝑡1/2𝑝𝑛𝑑𝑝𝑦/𝑡𝑒𝑖𝑝2𝑒𝑖𝑝𝑥𝑡1/2𝑝𝑛||||𝑑𝑝𝐶𝑦𝛿𝑡(1/2)(1+𝛿+𝑛)𝑒𝑖𝜙0𝑒𝐶|𝑝|2𝑒𝐶|𝑝|𝑥𝑡1/2||𝑝||𝑛||||,𝑑𝑝(2.160) where 𝑘[0,𝑦/𝑡] is intermediate point. Therefore 𝜕𝑛𝑥𝐼11𝐋𝑠,𝜇𝐶𝑦𝛿𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝑒𝑖𝜙0𝑒𝐶|𝑝|2||𝑝||𝑛(1/𝑠)𝜇𝑑𝑝,(2.161) where 1𝑠+𝜇<𝑛+1.(2.162) Now we get estimate for 𝐼12 from (2.159).
Using for 𝑝=|𝑝|𝑒𝑖𝜙,|𝑝|>𝑦𝑡1/2,𝐶1=sin𝜙<1,|||𝑒𝑖𝑝2𝑖𝑝𝑦𝑡1/2|||1𝐶|||||𝑝||2𝐶1||𝑝||𝑦𝑡1/2|||𝜇1,𝜇10,(2.163) we get ||𝜕𝑛𝑥𝐼12||(𝑥,𝑦,𝑡)𝐶𝑡(1/2)(1+𝑛)𝑦𝑡1/21||𝑝2𝐶1𝑝𝑦𝑡1/2||𝜇1𝑒𝐶|𝑝|𝑥𝑡1/2||𝑝||𝑛|||𝑒𝐶|𝑝|𝑦𝑡1/2|||1𝑑𝑝.(2.164) Therefore from (2.170) we obtain 𝜕𝑛𝑥I12(,𝑦,𝑡)𝐋𝑠,𝜇𝐶𝑡(1/2)(𝑛+1(1/𝑠)𝜇+𝛿)𝑦𝛿𝑦𝑡1/2𝑝𝑛(1/𝑠)𝜇+𝛿||𝑝2𝐶𝑝𝑦𝑡1/2||𝜇1𝑑𝑝𝐶𝑡(1/2)(𝑛+2(1/𝑠)𝜇)𝑦𝛿10𝑝𝑛+𝛿(1/𝑠)𝜇𝑑𝑝+1𝑝𝑛+𝛿(1/𝑠)𝜇𝑝2𝜇1𝑑𝑝𝐶𝑡(1/2)(𝑛+1(1/𝑠)𝜇+𝛿)𝑦𝛿,(2.165) where 1𝑠+𝜇<𝑛+1+𝛿.(2.166) Therefore 𝜕𝑛𝑥12𝜙𝐋𝑠,𝜇𝐶𝑡(1/2)(𝑛+1+𝛿(1/𝑠)𝜇)𝜙𝐋1,𝛿.(2.167) From (2.161) and (2.167) it follows estimate (2.158).
Now we estimate 2𝜙.
Making the change of variables 𝜉=𝑞𝑡 and 𝑝=𝑡𝑧 in (2.154) we get ||𝜕𝑛𝑥𝐹2||(𝑥,𝑦,𝑡)𝐶𝑡(1/2)(1+𝛿+𝑛)𝒞1𝑑𝑞𝑒𝐶|𝑞|𝑂(𝑘(𝑞)𝑦)𝛿𝒞2𝑒𝐶|𝑧|𝑥𝑡1/2|𝑧|𝑛{|𝑧|}1/2||𝐾||||||.(𝑧)+𝑞𝑑𝑧(2.168) Here we used the following estimations: 𝑒𝑘(𝜉)𝑦1=𝑂(𝑘(𝜉)𝑦)𝛿[],||𝑤,𝛿0,1+𝑒Γ(𝑝,𝜉)||||𝑝||𝐶1/2.(2.169) Since 𝑒𝐶|𝑧|𝑥𝑡1/2𝐋𝑠,𝜇𝐶0+𝑥𝜇𝑠𝑒𝐶𝑠𝑡1/2|𝑧|𝑑𝑥1/𝑠𝐶|𝑧|𝑡1/2(1/𝑠)𝜇0+𝑥1𝜇𝑠𝑒𝐶𝑠𝑥1𝑑𝑥11/𝑠,(2.170) we get 𝐹2𝐋𝑠,𝜇𝐶𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝑦𝛿𝒞1𝑒𝐶|𝑞|𝑘𝛿||𝑞||𝑑𝑞𝒞2|𝑧|𝑛(1/𝑠)𝜇{|𝑧|}1/2||𝐾||||||(𝑧)+𝑞𝑑𝑧𝐶𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝑦𝛿,(2.171) where 1𝑠3+𝜇<𝑛+2[].,(𝛿,𝜇)0,1(2.172) From the estimate (2.171) we have under condition (2.172) 𝜕𝑛𝑥2𝜙𝐋𝑠,𝜇𝐶𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝜙𝐋1,𝛿,(2.173) Now we estimate 3𝜙. From estimate (2.138) for 𝛾(0,1) we have 1𝑒Γ+𝑤+𝑒(1/2)||𝜉||=𝑂𝛾||𝑝||2𝛾,(2.174) and therefore for 𝑝𝒞2||𝐼+1||(𝑝,𝜉,𝑦)𝐶𝑦𝛿𝑖𝑖||𝑞||𝛿||||||||||𝜉||𝑞𝑝𝑞𝑘(𝜉)𝛾||𝑞||2𝛾𝑑𝑞.(2.175) Thus after the change of variables 𝜉𝑡=𝑞1,𝑝=𝑧𝑡,𝑞=𝑞2𝑡 we get 𝜕𝑛𝑥𝐹3(𝑥,𝑦,𝑡)𝐶𝑦𝛿𝑡(1/2)(1+𝛿+𝑛)𝒞1𝑑𝑞1𝑒𝐶|𝑞1|||𝑞1||𝛾×𝒞2𝑑𝑧𝑒𝐶|𝑧|𝑥𝑡1/2||𝑞𝑧𝑘1|||𝑧|𝑛{|𝑧|}1/2||𝐾(𝑧)+𝑞1||×𝑖𝑖||𝑞2||𝛿2𝛾||𝑞2||||𝑞𝑧2𝑞𝑘1||𝑑𝑞2.(2.176) From (2.170) we obtain for 𝛾[0,(1+𝛿)/2),𝛿[0,1]𝜕𝑛𝑥3𝜙𝐋𝑠,𝜇𝐶𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝑦𝛿𝜙𝐋1×𝒞1𝑒𝐶|𝑞1|||𝑞1||𝛾𝑑𝑞1𝒞2||𝑞𝑑𝑧𝑧𝑘1||||𝐾(𝑧)+𝑞1|||𝑧|𝑛(1/𝑠)𝜇{|𝑧|}1/2×𝑖𝑖||𝑞2||𝛿2𝛾||𝑞2||||𝑞𝑧2𝑞𝑘1||𝑑𝑞2𝐶𝑡(1/2)(1+𝛿+𝑛(1/𝑠)𝜇)𝜙𝐋1,𝛿,(2.177) where 1𝑠3+𝜇<𝑛+2[].,(𝛿,𝜇)0,1(2.178) Thus from (2.171), (2.177), and (2.158) it follows the second estimate (2.129) of the lemma.
Now we prove the asymptotic formula (2.130).
We will use the formula (2.115) from Lemma 2.6: 1𝐺(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝜀+𝑖𝜀𝑖𝑑𝜉𝑒𝜉𝑡𝑖𝑖𝑒𝑝𝑥𝑌(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1𝐼(𝑝)+𝜉(𝑝,𝜉,𝑦)𝑑𝑝,(2.179) where 1𝐼(𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑧)(𝑞𝑘(𝜉))𝑌+(𝑞,𝜉)(𝑒𝑞𝑦1)𝑑𝑞.(2.180) Since for Re𝜉>0𝑖𝑖𝑌(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾1𝐼(𝑝)+𝜉(𝑝,𝜉,𝑦)𝑑𝑝=0,(2.181) we have 𝐺(0,𝑦,𝑡)=0 and therefore 𝐺(𝑥,𝑦,𝑡)=3𝑗=1𝐹𝑗(𝑥,𝑦,𝑡),(2.182) where 𝐹11(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝐾(𝑝)𝑡(𝑒𝑝𝑥1)(𝑒𝑝𝑦𝐹1)𝑑𝑝,21(𝑥,𝑦,𝑡)=𝑒2𝜋𝑖1/22𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝑒𝑘(𝜉)𝑦1𝒞2(𝑒𝑝𝑥𝑒1)Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)𝐾𝐹(𝑝)+𝜉𝑑𝑝,31(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2(𝑒𝑝𝑥𝑒1)Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉1(𝑝,𝜉,𝑦)𝑑𝑝.(2.183) Here the function 𝐼1(𝑝,𝜉,𝑦) was defined by (2.148). Since 𝐼1𝐼(𝑝,𝜉,𝑦)=+1𝑝(𝑝,𝜉,𝑦)𝑦1(𝑝𝑘(𝜉))𝑒Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)𝑒1/2,(2.184) where 𝐼11(𝑧,𝜉,𝑦)=2𝜋𝑖𝑖𝑖11(𝑞𝑧)(𝑞𝑘(𝜉))𝑒Γ+(𝑞,𝜉)𝑤+(𝑞,𝜉)𝑒1/2×(𝑒𝑞𝑦1+𝑞𝑦)𝑑𝑞,(2.185) we have 𝐹31(𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2(𝑒𝑝𝑥𝑒1)Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉+1+1(𝑝,𝜉,𝑦)𝑑𝑝𝑦2𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝒞2(𝑒𝑝𝑥𝑌1)+(𝑝,𝜉)𝑝1𝐾(𝑝)+𝜉𝑒Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)𝑒1/2𝑑𝑝.(2.186) So 𝑦𝐺(𝑥,𝑦,𝑡)=𝑡Λ𝑥𝑡1/2+3𝑘=1𝑅𝑘(𝑥,𝑦,𝑡),(2.187) where Λ𝑥𝑡1/2=𝑒1/22𝜋2𝑖𝒞1𝑑𝜉𝑒𝜉𝒞2𝑒𝑝𝑥𝑡1/2𝑒1Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾𝑅(𝑝)+𝜉𝑑𝑝,11(𝑥,𝑦,𝑡)=2𝜋𝑖𝑖𝑖𝑒𝐾(𝑝)𝑡(𝑒𝑝𝑥1)(𝑒𝑝𝑦𝑅1+𝑝𝑦)𝑑𝑝,21(𝑥,𝑦,𝑡)=𝑒2𝜋𝑖1/22𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡𝑒𝑘(𝜉)𝑦×1+𝑘(𝜉)𝑦𝒞2(𝑒𝑝𝑥𝑒1)Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)𝑅𝐾(𝑝)+𝜉𝑑𝑝,3(1𝑥,𝑦,𝑡)=12𝜋𝑖2𝜋𝑖𝒞1𝑑𝜉𝑒𝜉𝑡×𝒞2(𝑒𝑝𝑥𝑒1)Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐼𝐾(𝑝)+𝜉+1(𝑝,𝜉,𝑦)𝑑𝑝.(2.188) Using 𝑒𝑝𝑦1+𝑝𝑦=𝑂(𝑝1+𝜇𝑦1+𝜇),𝑒𝑘(𝜉)𝑦1+𝑘(𝜉)𝑦=𝑂(𝑘(𝜉)1+𝜇𝑦1+𝜇),𝜇(0,1) by the same method as in the proof of estimate (2.129) we obtain that 3𝑗=1||||0𝑑𝑦𝑅𝑗||||(𝑥,𝑦,𝑡)𝜙(𝑦)𝐶𝑡1(𝜇/2)min𝑥𝑡1/2,1𝜙𝐋1,1+𝜇.(2.189) Since 𝑒Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝐾𝑒(𝑝)+𝜉1/2𝑝2𝜉+1=𝑂𝑝2,𝒞1𝑑𝜉𝑒𝜉𝒞2𝑒𝑝𝑥𝑡1/211𝑝2+1𝑑𝑝=0,(2.190) we rewrite Λ𝑥𝑡1/2=𝑒1/22𝜋2𝑖𝒞1𝑑𝜉𝑒𝜉𝒞2𝑒𝑝𝑥𝑡1/2𝑒1Γ+(𝑝,𝜉)𝑤+(𝑝,𝜉)(𝑝𝑘(𝜉))𝑒𝐾(𝑝)+𝜉1/2𝑝2+1𝑑𝑝.(2.191) So Λ(𝑥𝑡1/2)𝐋(𝐑+) and Λ(0)=0. From (2.189),(2.187) it follows (2.130) and consequently, 𝒢𝜙𝑡1Λ(𝑥𝑡1/2)𝑓(𝜙)𝐋𝐶𝑡1(𝜇/2)𝑥1+𝜇𝜙𝐋1.(2.192) The lemma is proved.

3. Proof of Theorem 1.1

By Proposition 2.4 we rewrite the initial-boundary value problem (1.1) as the following integral equation:

𝑢(𝑡)=𝒢(𝑡)𝑢0𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏,𝒩(𝑢)=𝑢𝑥𝑢,(3.1) where 𝒢 is the Green operator of the linear problem (2.9). We choose the space

𝐙=𝜙𝐇1𝐑+𝐋1𝐑+𝐋1,1+𝑎𝐑+(3.2) with 0<𝑎<1 being small and the space

𝜕𝐗=𝑛𝑥[𝜙𝐂0,);𝐋2𝐑+(0,);𝐋2,𝑛+(a/2)𝐑+𝜙𝐗<,𝑛=0,1,(3.3) where now the norm 𝜇[0,𝑎]

𝜙𝐗=sup𝑡01𝑛=0𝑡(1𝜇)/4𝑡1/2𝜕𝑛𝑥𝜙(𝑡)𝐋2,𝑛+(𝜇/2)+𝑡(1/2)(𝑛+(3/2))𝜕𝑛𝑥𝜙(𝑡)𝐋2(3.4) reflects the optimal time decay properties of the solution. We apply the contraction mapping principle in a ball 𝐗𝜌={𝜙𝐗𝜙𝐗𝜌}in the space 𝐗 of a radius

1𝜌=𝑢2𝐶0𝐙>0.(3.5) For 𝑣𝐗𝜌 we define the mapping (𝑢) by formula

(𝑢)=𝒢(𝑡)𝑢0𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏.(3.6) We first prove that

(𝑢)𝐗𝜌,(3.7) where 𝜌>0 is sufficiently small. By virtue of Lemma 2.7 we have

𝒢(𝑡)𝜙𝐇1𝜙𝐇1,𝜕𝑛𝑥𝒢𝜙𝐋2,𝜇𝐶𝑡(1/2)(𝑛+𝛿+(1/2)𝜇)𝜙𝐋1,𝛿(3.8) for all 𝑡0,𝜇<𝑛+(1/2). Therefore for 𝑡<1

𝒢𝑢0𝐇1𝑢𝐶0𝐙,(3.9)𝜕𝑛𝑥𝒢𝑢0𝐋2,𝑛+(𝑎/2)𝐶𝑡(1/2)(𝑛+(1/2)𝑛(𝑎/2))𝑢0𝐋1𝐶𝑡(1𝑎)/4𝑢0𝐙.(3.10) For 𝑡>1

𝜕𝑛𝑥𝒢𝑢0𝐋2𝐶𝑡(1/2)(𝑛+(3/2))𝑢0𝐋1,1𝐶𝑡(1/2)(𝑛+(3/2))𝑢0𝐙,𝜕𝑛𝑥𝒢𝑢0𝐋2,𝑛+(𝑎/2)𝐶𝑡(1/2)(𝑛+(3/2)𝑛(𝑎/2))𝜙𝐋1,1𝐶𝑡(3𝑎)/4𝑢0𝐙.(3.11) Thus

𝒢𝑢0𝐗𝑢𝐶0𝐙.(3.12) Also since 𝑣𝐗𝜌 we get for all 𝜏>0

𝒩(𝑢(𝜏))𝐋1,1𝐶𝑢𝐋2𝑢𝑥𝐋2,1𝐶{𝜏}1/4𝜏3/2𝑢2𝐗,𝒩(𝑢(𝜏))𝐋1𝐶𝑢𝐋2𝑢𝑥𝐋2𝐶𝜏2𝑢2𝐗.(3.13) Therefore by Lemma 2.7 we get

𝜕𝑛𝑥𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏𝐋20𝑡/2𝑡𝜏(1/2)(𝑛+(3/2))𝒩(𝑢(𝜏))𝐋1,1𝑑𝜏+𝑡𝑡/2(𝑡𝜏)(1/2)(𝑛+1(1/2))𝒩(𝑢(𝜏))𝐋1𝑑𝜏𝐶𝑢2𝐗0𝑡/2𝑡𝜏(1/2)(𝑛+2(1/2)){𝜏}1/2𝜏(2(1/2))+𝑑𝜏𝑡𝑡/2(𝑡𝜏)(1/2)(𝑛+1(1/2))𝜏2𝑑𝜏𝐶𝑢2𝐗𝑡(1/2)(𝑛+2(1/2))+𝑡2𝑡(1/2)(𝑛+1(1/2))+1𝐶𝑢2𝐗𝑡(1/2)(𝑛+(3/2))(3.14) for all 𝑡0. In the same manner by virtue of Lemma 2.7 we have

𝜕𝑛𝑥𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏𝐋2,𝑛+(𝑎/2)0𝑡/2(𝑡𝜏)(3𝑎)/4𝒩(𝑢(𝜏))𝐋1,1𝑑𝜏+𝑡𝑡/2(𝑡𝜏)1/4𝒩(𝑢(𝜏))𝐋1𝑑𝜏𝐶𝑢2𝐗0𝑡/2(𝑡𝜏)(3𝑎)/4{𝜏}1/4𝜏3/2𝑑𝜏+𝑡𝑡/2(𝑡𝜏)1/4𝜏2𝑑𝜏𝐶𝑢2𝐗𝑡(1/2)(𝑛+2(1/2)((1+𝑎)/2))+𝑡2𝑡(1/2)(𝑛+1(1/2))+1𝐶𝑢2𝐗𝑡((3𝑎))/4{𝑡}1/4(3.15) for all 𝑡0. Thus we get

𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏𝐗𝐶𝑢2𝐗,(3.16) and hence in view of (3.6) and (3.9)

(𝑢)𝐗𝒢𝑢0𝐗+𝑡0𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏𝐗𝑢𝐶0𝐙+𝐶𝑢2𝐗𝜌2+𝐶𝜌2<𝜌(3.17) since 𝜌>0 is sufficiently small. Hence the mapping transforms a ball 𝐗𝜌 into itself. In the same manner we estimate the difference

(𝑤)(𝑣)𝐗12𝑤𝑣𝐗,(3.18) which shows that is a contraction mapping. Therefore we see that there exists a unique solution 𝑢𝐂([0,);𝐋1(𝐑+)𝐋1,𝑎(𝐑+))𝐂((0,);𝐇1) to the initial-boundary value problem (1.1). Now we can prove asymptotic formula

𝑢(𝑥,𝑡)=𝐴1Λ𝑥𝑡1/2𝑡1+min𝑥𝑡1/2𝑂𝑡,11𝛾,(3.19) where 𝐴1𝑢=𝑓00+𝑑𝜏0+𝑦𝑢𝑦𝑢𝑑𝑦.(3.20) Denote

𝐺0(𝑡)=𝑡1Λ𝑥𝑡1/2.(3.21) From Lemma 2.7 we have

𝑡𝒢(𝑡)𝜙𝐺0(𝑡)𝑓(𝜙)𝐋𝐶𝜙𝐙(3.22) for all 𝑡>1. Also in view of the definition of the norm 𝐗 we have

||||𝑓(𝒩(𝑢(𝜏)))𝒩(𝑢(𝜏))𝐋1,1𝐶{𝜏}1/2𝜏3/2𝑢2𝐗.(3.23) By a direct calculation we have for some small 𝛾1>0,𝛾>0

0𝑡/2𝑒𝜏||𝐺0(𝑡𝜏)𝐺0||(𝑡)𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋𝑡1𝐶𝑢2𝐗0𝑡/2𝐺0(𝑡𝜏)+𝐺0(𝑡)𝐋{𝜏}𝛾1𝜏𝛾2𝑑𝜏𝐶𝑡20𝑡/2{𝜏}𝛾1𝜏𝛾2𝑑𝜏𝐶𝑡𝛾1(3.24) and in the same way

𝑡𝛾𝐺0(𝑡)𝑡/2𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋𝐶𝑢2𝐗.(3.25) Also we have

0𝑡/2𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝐺0(𝑡𝜏)𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋+𝑡𝑡/2𝒢(𝑡𝜏)𝒩(𝑢(𝜏))𝑑𝜏𝐋𝐶0𝑡/2(𝑡𝜏)1𝒩(𝑢(𝜏))𝐋1𝑑𝜏+𝐶𝑡𝑡/2𝒩(𝑢(𝜏))𝐋1𝑑𝜏𝐶𝑡1𝛾𝑢2𝐗(3.26) for all 𝑡>1. By virtue of the integral equation (3.1) we get

𝑡𝛾+1𝑢(𝑡)𝐴𝐺0(𝑡)𝐗(𝒢(𝑡)𝑢0𝐺0𝑢(𝑡)𝑓0𝐋+𝑡𝛾+10𝑡/2𝒢(𝑡𝜏)𝒩(𝑎(𝜏))𝐺0(𝑡𝜏)𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋+𝑡𝛾+1𝑡𝑡/2𝒢(𝑡𝜏)𝒩(𝑎(𝜏))𝑑𝜏𝐋+𝑡𝛾+1𝐺0(𝑡)𝑡/2𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋+𝑡𝛾+10𝑡/2𝐺0(𝑡𝜏)𝐺0(𝑡)𝑓(𝒩(𝑢(𝜏)))𝑑𝜏𝐋.(3.27) The all summands in the right-hand side of (3.27) are estimated by 𝐶𝑢0𝐙+𝐶𝑢2𝐗 via estimates (3.24)–(3.26). Thus by (3.27) the asymptotic (3.19) is valid. Theorem is proved.