Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2024, Volume 215, Issue 9, Pages 1224–1248
DOI: https://doi.org/10.4213/sm9500e
(Mi sm9500)
 

Numerical-analytic construction of a generalized solution to the eikonal equation in the plane case

P. D. Lebedevab, A. A. Uspenskiia

a N. N. Krasovskii Institute of Mathematics and Mechanics of the Ural Branch of the Russian Academy of Sciences, Ekaterinburg, Russia
b Ural Federal University named after the first President of Russia B. N. Yeltsin, Ekaterinburg, Russia
References:
Abstract: In the 1970s Kruzhkov introduced the concept of a generalized solution to the eikonal equation, and, for a medium of constant refractive index, indicated a class of functions containing the generalized solution to the Dirichlet boundary value problem. We present constructive methods for finding such a solution in the plane case. Nonsmooth singularities are known to develop in generalized solutions owing to the presence of pseudovertices, which are singular points of the boundary of the boundary set. Identifying such pseudovertices is related to finding fixed points of mappings formed in a local reparameterization of this boundary. We obtain necessary conditions for the existence of pseudovertices in the case when the curvature of the parametrically defined boundary of boundary sets is not smooth. These conditions are written as an equation for the pseudovertex marker (a numerical measure of the local nonconvexity of the boundary set). This equation, which has the typical structure of constructions involving fixed points, can be reduced to an algebraic equation. The solution of this equation (the marker) is found analytically if the curvature of the boundary of the boundary set has a nonsmooth extremum at a pseudovertex. We also give an example of a numerical-analytic construction of a generalized solution to the boundary value problem, with indication of the singular set and the evolution of wave fronts.
Bibliography: 29 titles.
Keywords: eikonal equation, generalized solution, singular set, pseudovertex, fixed point.
Received: 28.08.2020 and 14.06.2024
Bibliographic databases:
Document Type: Article
Language: English
Original paper language: Russian

Introduction

Nonlinear first-order partial differential equations typically fail to have classical (differentiable) solutions on the whole domain under consideration. The concept of a generalized solution (from respective classes of functions) is known to describe adequately the processes modelled in mechanics, geometric optics, optimal control, differential games, seismology, economics and other theoretical and applied fields involving first-order partial differential equations. There are various approaches towards the definition of a generalized solution to a first-order partial differential equation and the verification of its uniqueness [1]–[4].

Kruzhkov [1], with the help of some constructions from functional analysis and using short-wave approximation machinery, introduced a generalized solution to the eikonal equation, which is the basic equation of geometrical optics. His method has extensively been developed, and applications to the constructions of generalized solutions of partial differential equations of various types were found, thereby replenishing the stock of available vanishing viscosity methods [2]. This approach to the definition of generalized solutions to first-order partial differential equations is related both to problems with a small parameter multiplying the highest derivative and to the properties of their solutions and functions obtained by taking the limit with respect to the small parameter. There is also a different approach, based on constructions in the theory of positional differential games [5]. Using this approach, Subbotin [3] gave the definition of a minimax solution for equations of this type. Note that, in contrast to the vanishing viscosity method, this definition does not involve derivatives of high order. It must also be noted that the minimax approach is capable of constructing solutions to first-order partial differential equations with nonsmooth boundary conditions and with singularities in the differential operators. Methods of the theory of singularities of differentiable mappings [6] have proved quite useful for the analytic identification of nonsmooth singularities in generalized solutions to boundary value problems for first-order partial differential equations with smooth boundary conditions. Using the machinery of group analysis classes of eikonal equations are singled out for which one can obtain explicit formulae for wave fronts with point sources [7].

A transition from theoretical constructions to numerical algorithms should involve correct approximations of nonsmooth functions. Numerical methods for the construction of generalized solutions to Hamilton-type equations and the eikonal equation, which use difference schemes for the numerical solution of differential equations of mathematical physics and constructions from the method of characteristics are being developed overseas (see, for example, [8]) and in Russia (see, for example, [9]). In the framework of the minimax approach it proves possible to create numerical and numerical-analytic algorithms for the construction of nonsmooth solutions in the theory of positional differential games and optimal control problems and of solutions to the Hamilton–Jacobi equations, where constructions from subdifferential calculus and nonsmooth analysis are used (see, for example, [10]).

In the present paper we consider the planar Dirichlet problem for the eikonal equation with constant refractive index. We consider the case where the boundary set has a twice differentiable boundary with singular points, at which the third-order derivatives of the coordinate functions make finite jumps. A numerical-analytic approach towards the construction of a Kruzhkov generalized solution is proposed. The constructions developed in the minimax approach to the Hamilton–Jacobi equation [11]–[14] are used to create a new method for constructing a generalized eikonal, which is based, in particular, on the theory of fixed points [15] of smooth mappings. The key ingredient of this method is the measure of nonconvexity of a closed subset of a Euclidean space [16], and the fundamental relations are derived using approximations provided by the theory of jets [17] (we also use some methods of convex analysis [18]). Thus, for boundary sets with twice smooth boundaries whose curvature has points of nonsmooth extremum we have traversed all links in the ‘theory–methods–algorithms–numerical simulation’ research chain.

§ 1. Object of investigation, main definitions

In his famous paper [1] Kruzhkov investigated the existence of a generalized solution to the first-order partial differential equation

F(x,u,ux)=n(x)
in the set of functions u=u(x), xRk, from a certain stability class. Here ux=(ux1,,uxk) and n(x) is a measurable positive scalar function. The left-hand side of (1.1) is subject to conditions under which this equation can be interpreted as an eikonal equation, a classical equation of geometric optics. A typical representative of this class is the equation with
F(x,u,p)=F(0)(p)=ki=1p2i,piR,i=1,,k.
It is known that this equation has a unique solution (in the sense of the above definition) such that
u|Ω=u(0)
on the boundary Ω of the domain ΩRk, where the boundary function u(0)=u(0)(x) can either be smooth or nondifferentiable. It has been noted that the existence and uniqueness theorem for a generalized solution is not based on any properties of the boundary Ω of ΩRk; in particular, it does not depend on the smoothness of Ω and the geometry of the domain Ω.

After having constructed the nonlocal theory for the general eikonal equation Kruzhkov considered the simpler class of equations

F(0)(ux)=1,
where the Hessian matrix of F(0)=F(0)(p), pRk, is positive definite, F(0)(0)=0, and F(0)(p)+ as p=mi=1p2i+. For a convex domain ΩRk Kruzhkov justified the formula
u(x)=sup
for the generalized solution to the Dirichlet boundary value problem (1.2), (1.3) (which he called the Cauchy–Dirichlet problem)

Formula (1.4) has related Hopf formulae (based on extremal operations and duality theory) [19], which define viscosity solutions (and thus minimax solutions, as they coincide with viscosity solutions) of the corresponding boundary value problems for first-order partial differential equations [20]. It is also worth pointing out that formulae of this kind are used as a basis for the development of difference schemes for the numerical construction of minimax (viscosity) solutions to the Hamilton–Jacobi equation [21].

In what follows we create structural elements for the numerical-analytic construction of the solution to the plane boundary value problem with u^{(0)}(x)\equiv 0. According to [1], in this case the problem is related to the evolution of wave fronts in the domain \Omega \subset \mathbb{R}^k, where the initial front \Gamma=\Gamma (0) coincides with the boundary \partial \Omega and the successive fronts \Gamma (\tau), where the nonnegative parameter \tau plays the role of time, are constructed from the envelops of families of ovaloids.

We consider the Dirichlet boundary value problem for the eikonal equation on the plane of constant refractive index

\begin{equation} \biggl(\frac{\partial u}{\partial x_1}\biggr)^2 +\biggl(\frac{\partial u}{\partial x_2}\biggr)^2=1, \qquad u_\Gamma=0. \end{equation} \tag{1.5}

The boundary condition in (1.5) is set on \Gamma=\partial \Omega , the boundary of the closed set \Omega \subset \mathbb{R}^2, and \Gamma has no points of self-intersection. We assume that \Gamma=\{x=(\gamma_1 (t),\gamma_2 (t))\colon t\in T\}, where \gamma \colon T\to \mathbb{R}^2 is a continuous mapping of an interval T=(\check{t},\widehat{t}\,), -\infty\leqslant\check{t}<\widehat{t}\leqslant\infty, to the plane. We will also consider contours — curves defined on T=[\check{t},\widehat{t}\,], -\infty<\check{t}<\widehat{t}<\infty, such that \gamma(\check{t})=\gamma(\widehat{t}).

Let

\begin{equation*} M=\operatorname{cl}(\mathbb{R}^2\setminus \Omega) \end{equation*} \notag
be the closure of the complement of \Omega \subset \mathbb{R}^2. Then by (1.4) the Kruzhkov generalized solution of (1.5) assumes the form
\begin{equation} \begin{aligned} \, \notag u(x) &=\sup_{y\in \partial \Omega}[-\Phi(y-x)] =- \inf_{y\in \partial\Omega} \Phi(y-x) \\ &=-\inf_{y\in \partial \Omega} \max_{p\colon\| p \|=1} \langle y-x,p\rangle=-\rho(x,M), \end{aligned} \end{equation} \tag{1.6}
where \rho (x,M)=\inf_{y\in \partial \Omega} \| y-x \|=\inf_{y\in M} \| y-x\| is the Euclidean distance from a point x=(x_1,x_2) to M.

Emphasizing the relationship between two different approaches (the minimax approach and that of Kruzhkov’s) to the definition of a generalized solution to a first-order partial differential equation, we note that \widetilde{u}(x)=\rho (x,M) is a minimax solution to the Dirichlet problem for the Hamilton–Jacobi equation (see [11])

\begin{equation} \min_{\nu \colon\| \nu \|\leqslant 1} \biggl(\nu_1\, \frac{\partial \widetilde{u}}{\partial x_1}+\nu_2 \, \frac{\partial\widetilde{u}}{\partial x_2}\biggr)+1=0, \qquad {\widetilde{u}} |_\Gamma=0, \end{equation} \tag{1.7}
with the same boundary conditions as in (1.5). The generalized solutions (obtained via different approaches) to formally distinct problems (1.5) and (1.7) differ only in sign, and therefore these solutions have the same ‘contour maps’. Hence methods of the construction of a generalized solution to one of these problems can be applied to the construction of a generalized solution to the other. In our study the constructions developed for minimax solutions to the Hamilton–Jacobi equations and time-optimal control problems (see [11]–[14]) are tailored and developed for the construction of a generalized eikonal. Following Kruzhkov (see Remark 6.2 in [1]), we also note that in the solution of the model problem (1.5) a certain structural unity arises, which can be used in the construction of a generalized solution to the general eikonal equation (1.1).

In what follows M\subset \mathbb{R}^2 is called the boundary set of problem (1.5), and the closure of its complement \Omega \subset \mathbb{R}^2 is called the domain of definition of a generalized solution to (1.5). In view of the above definitions these sets have the common boundary \Gamma=\partial \Omega=\partial M. The structure of the singular set of the generalized solution to (1.5) depends substantially on the differential properties of \Gamma . Let us indicate some properties of the curve \Gamma describing the boundary set M. By \det(a,b) we denote the determinant of the 2\times 2 matrix whose rows are the vectors a=(a_1,a_2) and b=(b_1,b_2). We assume in what follows that the boundary of the boundary set satisfies the following conditions:

(\Gamma1) \gamma (t)=(\gamma_1 (t),\gamma_2 (t)) is twice continuously differentiable on T;

(\Gamma2) \gamma' (t)\neq(0,0), t\in T;

(\Gamma3) there exists a finite set T^0\subset T of points t_0\in T^0 at which the one-sided third-order derivatives are finite, and at least one of the equalities \gamma_1'''(t_0 -0)=\gamma_1''' (t_0+0) and \gamma_2''' (t_0-0)=\gamma_2''' (t_0+0) fails;

(\Gamma4) \det (\gamma'(t),\gamma''(t))\ne 0, t\in T^0.

Remark 1. Condition (\Gamma1) implies that the curve \Gamma has classical curvature, and Condition (\Gamma2) means that this curve is regular. Condition (\Gamma3) shows that the boundary of the boundary set has points with discontinuous derivatives — these are singular points, which can give rise to branches of the singular set of the solution of (1.5). It follows from conditions (\Gamma2) and (\Gamma4) that at singular points of the curve the coordinate functions are not flat, and so we can apply approximations provided by the theory of jets (see [17] and [21]) at these points.

By \{\Gamma \}_T we denote the set of curves \Gamma (without self-intersections) with the above differential properties (\Gamma1)–(\Gamma4).

Now we give the definition of the main constructive elements of the approach we develop (see also [11]–[14]).

In what follows we study the properties of solutions to the equation

\begin{equation} G(t_1,t_2)=0, \end{equation} \tag{1.8}
which relates the differential eikonal operator to the boundary of the boundary set. The structure of the equation will be specialized below. Here we only mention that G=G(t_1,t_2) is a symmetric function of two variables in the plane of parameters (t_1,t_2)\in \mathbb{R}^2. We fix t_0\in T^0 and the parameters of smallness \delta_1 >0 and \delta_2 >0. By a solution to this equation we mean local diffeomorphisms (see § 1 in [17]) defined on one side of the point under consideration. We say that the local diffeomorphism t_2=t_2 (t_1) defined by (1.8) is left-semicontinuous at t_1=t_0 and maps the left half-neighbourhood of the point t_1=t_0 to a right half-neighbourhood of it if:

(\text{A}1) t_2((t_0 -\delta_1,t_0))=(t_0,t_0+\delta_2), \delta_1, \delta_2 >0;

(\text{A}2) \lim_{t_1 \to t_0 -0} t_2 (t_1)=t_0 .

Accordingly, we say that the local diffeomorphism t_1=t_1(t_2) defined by (1.8) is right-semicontinuous at t_1=t_0 and maps the right half-neighbourhood of t_2=t_0 to a right half-neighbourhood if (similarly to (\text{A}1) and (\text{A}2)):

(\underline{\text{A}1}) t_1((t_0,t_0+\delta_2))=(t_0 -\delta_1,t_0), \delta_1 , \delta_2 >0;

(\underline{\text{A}2}) \lim_{t_2 \to t_0+0} t_1 (t_2)=t_0 .

Remark 2. In addition to semicontinuity, conditions (\text{A}1) and (\text{A}2) (as well as (\underline{\text{A}1}) and (\underline{\text{A}2}) also provide the definitions (in the limit form) of fixed points in the space of parameters. The existence of such points and the existence of local diffeomorphisms giving rise to these points was verified in [12]–[14] using examples of boundary conditions of various orders of smoothness.

We fix t_0\in T such that k(t_0)\neq 0 (see [23]) and select two arbitrary time instants t_1 and t_2 , t_1<t_0 <t_2 , from the neighbourhood O(t_0, \delta)=(t_0-\delta, t_0+\delta), \delta >0. We draw the tangent lines through the points \gamma (t_1) and \gamma (t_2), and construct the two-parameter family of solutions x^*=x^*(t_1, t_2) of the corresponding system of equations

\begin{equation} \begin{cases} (x_1^\ast -\gamma_1 (t_1))\gamma_2' (t_1)=(x_2^\ast -\gamma_2 (t_1))\gamma_1' (t_1), \\ (x_1^\ast -\gamma_1 (t_2))\gamma_2' (t_2)=(x_2^\ast -\gamma_2 (t_2))\gamma_1' (t_2) . \end{cases} \end{equation} \tag{1.9}

The equation

\begin{equation} G(t_1,t_2) \triangleq\rho^2\bigl(\gamma (t_1), x^\ast(t_1,t_2)\bigr) -\rho^2\bigl(\gamma (t_2),x^\ast (t_1,t_2)\bigr)=0, \end{equation} \tag{1.10}
is a meaningful specialization of (1.8). Assume that there exists a local diffeomorphism t_2= t_2 (t_1) defined by (1.10) and satisfying conditions (\text{A}1) and (\text{A}2). Consider the restriction x^*=x^*(t_1,t_2(t_1)) of the family of solutions x^*=x^*(t_1, t_2). The one-parameter set x^*=x^*(t_1,t_2(t_1)) is a line of quasisymmetry. By construction this set consists of the points of intersection of tangents that lie at equal distances from the point of tangency.

Notation 1. A pseudovertex of a curve \Gamma is defined by

\begin{equation*} x^{(0)}=\gamma (t_0) \triangleq\lim_{t_1 \to t_0 -0} x^*(t_1, t_2(t_1)) . \end{equation*} \notag

Geometrically, a pseudovertex is a point of intersection of the curve \Gamma with the closure of the line of quasisymmetry. For conditions of the existence of a pseudovertex on a twice differentiable curve, see [23].

Remark 3. A pseudovertex can also be defined in terms of the local diffeomorphism t_1=t_ 1(t_2) inverse to t_2=t_2 (t_1) and satisfying conditions (\underline{\text{A}1}) and (\underline{\text{A}2}). In what follows t_2=t_2 (t_1) and t_1=t_ 1(t_2) are referred to as the local diffeomorphisms generating the pseudovertex x^{(0)}\in \Gamma.

Let P_M(x) be the metric projection operator onto M, which associates with a point x\in \operatorname{int}\Omega (\operatorname{int}\Omega=\Omega \setminus \Gamma ) the set of its nearest points in M. If the operator P_M(x), x\in \operatorname{int}\Omega , is single valued (that is, \operatorname{card}P_M(x)=1), then the solution to the eikonal equation is differentiable in the domain \Omega (see [24]). In problems under consideration here this is the case of a convex boundary set M. In this case M is a sun in the sense of [25], and the eikonal is a smooth function, which can be constructed by classical methods. The metric projection onto a nonconvex set M is not single valued (that is, there are points at which \operatorname{card}P_M(x)>1).

Notation 2. The bisector of a set Z\subset \mathbb{R}^k is defined by

\begin{equation*} L=\{x\in \operatorname{int}(\mathbb{R}^k\setminus Z)\colon \operatorname{card}P_Z(x)>1\}. \end{equation*} \notag

For problem (1.5) the bisector L of Z=M consists of the points that have at least two nearest elements (orthogonal projections) in M. So, the bisector L of M\subset \mathbb{R}^k consists of the points at which the Euclidean distance function is not smooth, that is, L is the singular set of the eikonal. Any bisector is a symmetry set, whose topological properties are considered in the theory of singularities of smooth mappings (see, for example, [26]).

Notation 3. The branch L(x^{(0)}) of the bisector L of a curve \Gamma, where {x}^{(0)} is a pseudovertex of \Gamma, is the set of points x=({x_1,x_2})\in \mathbb{R}^2 satisfying

\begin{equation} \begin{cases} (x_1 -\gamma_1 (t_1))\gamma_1' (t_1)+(x_2-\gamma_2 (t_1))\gamma_2' (t_1)=0, \\ (x_1 -\gamma_1 (t_2))\gamma_1' (t_2)+(x_2-\gamma_2 (t_2))\gamma_ 2' (t_2)=0, \end{cases} \end{equation} \tag{1.11}
where t_2=t_2 (t_1) is a local diffeomorphism generating the pseudovertex {x}^{(0)} .

Remark 4. System (1.11) is adjoint to system (1.9). Its solutions are the points x\in \mathbb{R}^2\setminus \Gamma that have two nearest points on \Gamma=\partial M=\partial \Omega .

Pseudovertices and branches of a bisector are the main structural elements in the problem of construction of the singular sets of problem (1.5).

Let us define some scalar characteristics of pseudovertices.

Notation 4. The left-hand one-sided derivative

\begin{equation} \lambda\triangleq t_2'(t_0 -0)=\lim_{t_1 \to t_0 -0} \frac{t_2 (t_1)-t_0}{t_1 -t_0} \end{equation} \tag{1.12}
is known as the left-hand marker of a pseudovertex {x}^{(0)}\in \Gamma; here t_2=t_2 (t_1) is the local diffeomorphism generating the pseudovertex {x}^{(0)}.

Notation 5. The right-hand one-sided derivative

\begin{equation} \mu \triangleq t_1' (t_0+0)=\lim_{t_2 \to t_0+0} \frac{t_1 (t_2)-t_0}{t_2 -t_0} \end{equation} \tag{1.13}
is known as the right-hand marker of a pseudovertex {x}^{(0)}\in \Gamma; here t_1=t_1 (t_2) is the local diffeomorphism generating the pseudovertex {x}^{(0)}.

The significant part of (1.10) is symmetric, so if the local diffeomorphism {t_2=t_2 (t_1)}, t_1\in(t_0 -\delta_1,t_0), is a solution to (1.10), then so is the inverse local diffeomorphism t_1=t_1(t_2), t_2\in(t_0,t_0+\delta_2) (see [12]). In addition, the conditions of lower semicontinuity (\text{A}2) and (\underline{\text{A}2}) imply that the graphs of these diffeomorphisms have a common limit point (t_1,t_2)=(t_0,t_0). In this case the one-sided markers are reciprocal,

\begin{equation} \mu=\lambda^{-1}. \end{equation} \tag{1.14}
We also note that \lambda\leqslant0 since t_2'(t_1)<0 for t_1\in(t_0 -\delta_1,t_0), \delta_1 >0. By (1.14) we have \mu\leqslant0.

Remark 5. The range (spectrum) of one-sided markers is the closure \Lambda=[-\infty,0] of the negative half-line. The markers assume the values in \Lambda depending on the order of smoothness of the curve at the pseudovertex (see [12] and [14]). Once the markers are known, one can construct branches of the singular set either on the basis of Definition 3 or using the machinery of integral curves (for details, see [14]).

Remark 6. Problems of the existence of pseudovertices and other structural elements of our theory (branches and extreme points of bisectors, markers of pseudovertices) were considered in other works by these authors (see, for example, [12]–[14]), for curves with various differential properties, where illustrative examples were also given.

§ 2. The main theoretical result

Let us find the ‘place’ of one-sided markers in the spectrum \Lambda=[-\infty,0] in the case of the nonsmooth curvature of the boundary of the boundary set in problem (1.5). The main result is formulated for the left-hand marker of a pseudovertex; the value of the right-hand marker is reciprocal to that of the left-hand marker.

Theorem. In the Dirichlet problem (1.5) let x^{(0)}=(\gamma_1(t_0), \gamma_2(t_0)) be a pseudovertex of the boundary \Gamma=\{\gamma(t)\colon t\in T\}\in \{\Gamma\}_T, let t_0 \in T^0, and let t_2=t_2 (t_1) be a local diffeomorphism generating the pseudovertex x^{(0)}, at which

\begin{equation} k (t_0)\ne 0. \end{equation} \tag{2.1}
If there exists a finite left-hand marker \lambda=t_2'(t_0-0)\leqslant0 , then there exist coefficients of convex combination
\begin{equation*} \alpha_1 (\lambda)\geqslant 0\quad\textit{and} \quad \beta_1 (\lambda)\geqslant 0, \qquad\alpha_1 (\lambda)+\beta_1 (\lambda)=1, \end{equation*} \notag
independent of this marker and such that
\begin{equation} \alpha (\lambda)k' (t_0+0)+\beta (\lambda)k' (t_0 -0)=0. \end{equation} \tag{2.2}
Here k' (t_0+0) and k'(t_0 -0) are the one-sided derivatives of the curvature at the pseudovertex x^{(0)}=(\gamma_1(t_0), \gamma_2(t_0)).

Proof. In the proof of (2.2) we use the transversality conditions (see [12])
\begin{equation} \lim_{t_1 \to t_0 -0} \biggl(\frac{\partial G(t_1,t_2 (t_1))}{\partial t_1} +t_2' \, \frac{\partial G(t_1,t_2(t_1))}{\partial t_2}\biggr)=0 \end{equation} \tag{2.3}
at the points (t_1,t_2)=(t_1,t_2(t_1)) of the graph of the local diffeomorphism {t_2=t_2 (t_1)}, t_1 \in(t_0 -\delta_1,t_0), \delta_1 >0, which generates the pseudovertex x^{(0)}=(\gamma_1(t_0), \gamma_2(t_0)). This local diffeomorphism t_2=t_2 (t_1) is a solution to the equation
\begin{equation} G(t_1,t_2)\triangleq \frac{\gamma_2(t_2)-\gamma_2(t_1)}{\gamma_1(t_2)-\gamma_1(t_1)} -\frac{-\gamma_1'(t_1)\gamma_1'(t_2)+\gamma_2'(t_1)\gamma_2'(t_2)+s(t_1)s(t_2)} {\gamma_2'(t_1)\gamma_1'(t_2)+\gamma_2'(t_2)\gamma_1'(t_1)}=0, \end{equation} \tag{2.4}
to which the original equation (1.10) can be reduced (for a justification, see [13]). Here s(t)=\sqrt {(\gamma_1' (t))^{2}+(\gamma_2' (t))^{2}} is the length of the tangent vector, which is positive in view of Condition (\Gamma2). Effectively, (2.3) describes the condition of transversal intersection of the closure of the graph of the local diffeomorphism t_2=t_2 (t_1) with the graph of the identity diffeomorphism t_2=t_1 at the common limit point (t_1,t_2)=(t_0,t_0). By the symmetry of (2.4) the identity diffeomorphism is a solution to this equation (see [12]); however, it does not satisfy conditions (\text{A}1) and (\text{A}2). The function t_2=t_1 must be considered as an ‘extraneous’ solution to (1.10). Note that the presence of this solutions in the constructions under consideration hinders considerably the use of classical theorems of analysis: for example, one cannot use here the theorem on the existence and uniqueness of an implicitly defined function.

By the assumptions of the theorem the left-hand marker exits and is finite. Hence by (2.3) it can be evaluated as

\begin{equation} t_2' (t_0 -0)=-\lim_{t_1 \to t_0 -0} \biggl(\frac{\partial G(t_1,t_2 (t_1))}{\partial t_1} \cdot\biggl(\frac{\partial G(t_1,t_2 (t_1))}{\partial t_2}\biggr)^{-1}\biggr). \end{equation} \tag{2.5}
Next, using essentially the theory of jets [17], we find approximations of the partial derivatives in (2.5) taken along the diffeomorphism t_2=t_2 (t_1).

To make the presentation more compact, we adhere to the following conventions. In all approximation expansions involving a one-sided direction, the point under consideration is fixed and we have t_2=t_0 or t_1=t_0 (depending on whether we expand to the left or right). For brevity we omit the value t_0 of the arguments t_2 and t_1. First we find approximations for arbitrary positive increments \Delta_1=t_0-t_1 and \Delta_2=t_2-t_0, and then we approximate for constrained increments conditioned by the local diffeomorphism t_2=t_2 (t_1). It is worth pointing out that if a triple of points t_1,t_0,t_2 in T, where t_1 <t_0 <t_2 , is related by t_2=t_2 (t_1), then the increment \Delta_2=\Delta_2(\Delta_1)=t_2(t_1)-t_0 (that is, \Delta_2 ) depends on \Delta_1 , and

\begin{equation} \Delta_2 \,{=}\,\Delta_2(\Delta_1)\,{=}\,t_2(t_1)-t_0\,{=}\,t_2'(t_0-0)(t_1-t_0)+o (t_1-t_0) \,{=}\,-\lambda\Delta_1+o(\Delta_1). \end{equation} \tag{2.6}
This implies, in particular, that
\begin{equation} \Delta \,{=}\,\Delta_1+\Delta_2(\Delta_1)\,{=}\,\Delta_1-\lambda\Delta_1+o(\Delta_1) \,{=}\,(1-\lambda)\Delta_1+o(\Delta_1),\quad\text{where } 1-\lambda\neq0. \end{equation} \tag{2.7}
Here o(\Delta_1^k), k=1,2,3, denotes the class of functions of higher order of smallness than the argument to the left of the point under consideration, that is,
\begin{equation*} \lim_{\Delta_1\downarrow0} \frac{o(\Delta_1^k)}{\Delta_1^k}=0. \end{equation*} \notag
We denote by \varepsilon(\Delta_1) an infinitesimal function in the left half-neighbourhood of the point under consideration; here
\begin{equation*} \lim_{\Delta_1\downarrow0}\varepsilon(\Delta_1)=0. \end{equation*} \notag
By definition, two scalar functions y=q(t) and y=g(t) are equivalent to the left of a point t=t_0\in \mathbb{R} if
\begin{equation*} \lim_{t \to t_0-0}\frac{q(t)}{g(t)}=1. \end{equation*} \notag
In this case we write q(t) \sim g(t), t \to t_0-0 .

For the functions considered below and, in particular, for the coordinate functions \gamma_1(t) and \gamma_2(t), the curvature k(t), the length of the tangent vector s(t), and for their derivatives evaluated at the central node t=t_0 , we drop the argument; in addition, for one-sided derivatives, in place of the notation t_0-0 and t_0+0, we lower the plus or minus sign to the subscript:

\begin{equation*} \gamma_i'=\gamma_i'(t_0), \quad \gamma_i''=\gamma_i''(t_0), \quad \gamma_{i,-}'''=\gamma_i'''(t_0-0), \quad \gamma_{i,+}'''=\gamma_i'''(t_0+0), \qquad i=1,2, \end{equation*} \notag
and so on. Expanding the partial derivative in the first variable, we have
\begin{equation*} \begin{aligned} \, &\frac{\partial G(t_1,t_2)}{\partial t_1} =\frac{\det (\gamma'(t_1),\gamma (t_2)-\gamma (t_1))} {(\gamma_1 (t_2)-\gamma_1 (t_1))^2} -\frac{s^2(t_2)\det (\gamma' (t_1),\gamma''(t_1))}{(\gamma_2'(t_1)\gamma_1'(t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\quad+\frac{s(t_2)[ {s(t_1)(\gamma_2'' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1'' (t_1)) -s'(t_1)(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1'(t_1))}]} {(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}. \end{aligned} \end{equation*} \notag
We begin with the last term. We have
\begin{equation*} s'(t)=\frac{\langle {\gamma'(t),\gamma''(t)} \rangle}{s(t)}=\frac{\gamma_1'(t)\gamma_1''(t)+\gamma_2' (t)\gamma_2'' (t)}{s(t)} \end{equation*} \notag
and, after rearranging, find that
\begin{equation*} \begin{aligned} \, & \frac{s(t_2)[ s(t_1)(\gamma_2'' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1'' (t_1)) -s'(t_1)(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))]} {(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad =\frac{s(t_2)}{s(t_1)}\cdot \biggl(\frac{((\gamma_1' (t_1))^2+(\gamma_2'(t_1))^2) (\gamma_2'' (t_1)\gamma_1' (t_2 )+\gamma_2' (t_2)\gamma_1'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad\qquad -\frac{(\gamma_1' (t_1)\gamma_1'' (t_1)+\gamma_2' (t_1)\gamma_2'' (t_1)) (\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}\biggr) \\ &\qquad =\frac{s(t_2)}{s(t_1)} \cdot\biggl(\frac{((\gamma_1' (t_1))^2\gamma_1' (t_2)\gamma_2'' (t_1)-\gamma_1' (t_1)\gamma_2' (t_1)\gamma_1' (t_2)\gamma_1'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad\qquad +\frac{((\gamma_2' (t_1))^2\gamma_2' (t_2)\gamma_1'' (t_1)-\gamma_2' (t_1)\gamma_2' (t_2)\gamma_1' (t_1)\gamma_2'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}\biggr) \\ &\qquad =\frac{s(t_2)}{s(t_1)}\cdot\biggl( \frac{\gamma_1' (t_1)\gamma_1' (t_2)(\gamma_1' (t_1)\gamma_2'' (t_1)-\gamma_2' (t_1)\gamma_1'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad\qquad +\frac{\gamma_2' (t_1)\gamma_2' (t_2)(\gamma_2' (t_1)\gamma_1'' (t_1)-\gamma_1' (t_1)\gamma_2'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}\biggr) \\ &\qquad =\frac{s(t_2)}{s(t_1)}\cdot \frac{\gamma_1' (t_1)\gamma_1' (t_2) \det (\gamma' (t_1),\gamma'' (t_1))-\gamma_2'(t_1)\gamma_2' (t_2)\det (\gamma' (t_1),\gamma'' (t_1))}{(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad =\frac{s(t_2)\det (\gamma' (t_1),\gamma'' (t_1))(\gamma_1' (t_1)\gamma_1' (t_2)-\gamma_2' (t_1)\gamma_2' (t_2))} {s(t_1)(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}. \end{aligned} \end{equation*} \notag
Therefore,
\begin{equation*} \begin{aligned} \, \frac{\partial G(t_1,t_2)}{\partial t_1} &=\frac{\det (\gamma'(t_1),\gamma (t_2)-\gamma (t_1))}{(\gamma_1 (t_2)-\gamma_1 (t_1))^2} -\frac{s^2(t_2)s^3(t_1)k(t_1)}{(\gamma_2'(t_1)\gamma_1'(t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &\qquad+\frac{s(t_2)s^2(t_1)k(t_1)(\gamma_1' (t_1)\gamma_1' (t_2)-\gamma_2' (t_1)\gamma_2' (t_2))} {(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2} \\ &=\frac{\det (\gamma' (t_1),\gamma (t_2)-\gamma (t_1))}{(\gamma_1 (t_2)-\gamma_1 (t_1))^2} \\ &\qquad+\frac{s(t_2)s^2(t_1)k(t_1)(\gamma_1' (t_1)\gamma_1' (t_2)-\gamma_2'(t_1)\gamma_2' (t_2)-s(t_2)s(t_1))} {(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))^2}. \end{aligned} \end{equation*} \notag
From now on, we assume that the arguments of the functions involved are related by the local diffeomorphism t_2=t_2 (t_1), which generates the pseudovertex; here, t_1 \in (t_0 -\delta_1,t_0), \delta >0. In other words, by (2.4) these functions satisfy the equality
\begin{equation*} \frac{\gamma_1' (t_1)\gamma_1' (t_2)-\gamma_2' (t_1)\gamma_2'(t_2)-s(t_2)s(t_1)} {\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1)} =-\frac{\gamma_2 (t_2)-\gamma_2 (t_1)}{\gamma_1(t_2)-\gamma_1 (t_1)}. \end{equation*} \notag
Using this equality and the formula for the curvature
\begin{equation*} k(t)=\frac{\det (\gamma' (t),\gamma''(t))}{s^3(t)}, \end{equation*} \notag
we simplify the partial derivative as follows:
\begin{equation} \begin{aligned} \, \notag \frac{\partial G(t_1,t_2)}{\partial t_2} &=\frac{\det (\gamma'(t_2),\gamma (t_1)-\gamma (t_2))}{(\gamma_1 (t_2)-\gamma_1 (t_1))^2} \\ &\qquad -\frac{(\gamma_2 (t_2)-\gamma_2 (t_1))s(t_1)\det (\gamma' (t_2),\gamma''(t_2))} {(\gamma_1 (t_2)-\gamma_1 (t_1))s(t_2)(\gamma_2' (t_2)\gamma_1' (t_1)+\gamma_2' (t_1)\gamma_1' (t_2))}. \end{aligned} \end{equation} \tag{2.8}

Now, to make the calculations more compact, we use a moving (Frenet) frame on the plane. We use the natural parameter l\geqslant 0 (the arc length of the curve), where

\begin{equation} dl=|\gamma'(t)|\,dt \triangleq s(t)\,dt. \end{equation} \tag{2.9}
We set \overline \gamma (l)\triangleq\gamma (t(l)). Now the vectors v=\overline \gamma'(l) and w=v'(l), as calculated at l=l_0 \triangleq l(t_0), form the orthonormal moving frame attached to the point \overline\gamma (l_0)=\gamma (t_0).

Expanding the vector function in a Taylor series about this point up to a third-order term, we have

\begin{equation*} \begin{aligned} \, \overline \gamma (l_0 -\Delta l_1) &=\overline \gamma (l_0)-\overline \gamma'(l_0)\Delta l_1 \\ &\qquad+\frac{1}{2}\overline \gamma''(l_0)\Delta l_1^2 -\frac{1}{6}\overline \gamma_-''' (l_0)\Delta l_1^3+o(\Delta l_1^3), \qquad \Delta l_1 >0, \\ \text{and} \qquad \overline \gamma (l_0+\Delta l_2) &=\overline \gamma (l_0)+\overline \gamma'(l_0)\Delta l_2 \\ &\qquad+\frac{1}{2}\overline \gamma''(l_0)\Delta l_2^2 +\frac{1}{6}\overline \gamma_+''' (l_0)\Delta l_2^3 +o(\Delta l_2^3), \qquad \Delta l_2>0. \end{aligned} \end{equation*} \notag
Hence in the moving base v_0=\overline \gamma'(l_0), w_0=v'(l_0) with origin \overline \gamma (l_0) the expansions assume the following form (see (5.9) in [27]):
\begin{equation} \nonumber \overline \gamma (l_0 -\Delta l_1) =\overline \gamma (l_0)+\biggl(-\Delta l_1+\frac{1}{6}\overline k^2(l_0)\Delta l_1^3+o(\Delta l_1^3)\biggr)v_0 \end{equation} \notag
\begin{equation} \qquad +\biggl(\frac{1}{2}\overline k (l_0)\Delta l_1^2-\frac{1}{6}\overline k_-' (l_0)\Delta l_1^3+o(\Delta l_1^3)\biggr)w_0 \end{equation} \tag{2.10}
and
\begin{equation} \nonumber \overline \gamma (l_0+\Delta l_2) =\overline \gamma (l_0)+\biggl(\Delta l_2 -\frac{1}{6}\overline k^2(l_0)\Delta l_2^3+o(\Delta l_2^3)\biggr)v_0 \end{equation} \notag
\begin{equation} \qquad +\biggl(\frac{1}{2}\overline k (l_0)\Delta l_2^2+\frac{1}{6}\overline k_+' (l_0)\Delta l_2^3+o(\Delta l_2^3)\biggr)w_0 . \end{equation} \tag{2.11}
Here \overline k (l) is the curvature of the curve at the corresponding point.

In view of (2.9)(2.11) the original parametrization of the vector function has the following expansions at the extreme points:

\begin{equation} \begin{aligned} \, \gamma (t_1) &=\gamma+\biggl(-s\Delta_1+\frac{1}{6}k^2s^3\Delta_1^3+o(\Delta_1^3)\biggr)v_0 \notag \\ &\qquad+\biggl(\frac{1}{2}ks^2\Delta_1^2 -\frac{1}{6}k_-' s^3\Delta_1^3+o(\Delta_1^3)\biggr)w_0 \end{aligned} \end{equation} \tag{2.12}
and
\begin{equation} \begin{aligned} \, \gamma (t_2) &=\gamma+\biggl(s\Delta_2-\frac{1}{6}k^2s^3\Delta_2^3+o(\Delta_2^3)\biggr)v_0 \notag \\ &\qquad+\biggl(\frac{1}{2}ks^2\Delta_2^2+\frac{1}{6}k_+' s^3\Delta_2^3+o(\Delta_2^3)\biggr)w_0. \end{aligned} \end{equation} \tag{2.13}
Here \Delta l_i=l'(t_0)\Delta_i=s\Delta_i , i=1,2, and \Delta_i \to 0 as \Delta l_i \to 0 . At these nodes the derivatives up to order 2 have the form
\begin{equation} \gamma'(t_1)=\biggl(s-\frac{1}{2}k^2s^3\Delta_1^2+o(\Delta_1^2)\biggr)v_0 +\biggl(-ks^2\Delta_1^+\frac{1}{2}k_-' s^3\Delta_1^2+o(\Delta_1^2)\biggr)w_0, \end{equation} \tag{2.14}
\begin{equation} \gamma'(t_2)=\biggl(s-\frac{1}{2}k^2s^3\Delta_2^2+o(\Delta_2^2)\biggr)v_0 +\biggl(ks^2\Delta_2+\frac{1}{2}k_+' s^3\Delta_2^2+o(\Delta_2^2)\biggr)w_0, \end{equation} \tag{2.15}
\begin{equation} \gamma''(t_1)=(k^2s^3\Delta_1+o(\Delta_1))v_0 +(ks^2-k_-' s^3\Delta_1+o(\Delta_1))w_0 \end{equation} \tag{2.16}
and
\begin{equation} \gamma''(t_2)=(-k^2s^3\Delta_2+o(\Delta_2))v_0+(ks^2+k_+' s^3\Delta_2+o(\Delta_2))w_0 . \end{equation} \tag{2.17}
Here we have used that \Delta_1'=(t_0 -t_1)'=-1 and \Delta_2'=(t_2 -t_0)'=1.

Using (2.10)(2.12) we expand the determinant as follows:

\begin{equation*} \begin{aligned} \, &\det (\gamma' (t_1),\gamma (t_2)-\gamma (t_1)) \\ &{=}{\kern1pt}\Delta {\begin{vmatrix} s-\dfrac{1}{2}k^2s^3\Delta_1^2+o(\Delta_1^2) &-ks^2\Delta_1+\dfrac{1}{2}k_-' s^3\Delta_1^2+o(\Delta_1^2) \\ s\!-\!\dfrac{1}{6}k^2 s^3(\Delta_2^2\!-\!\Delta_1 \Delta_2\!+\!\Delta_1^2)\!+\!o(\Delta_2^2) &\dfrac{1}{2}ks^2(\Delta_2\!-\!\Delta_1)\!+\!\dfrac{1}{6} k_+'s^3\dfrac{\Delta_2^3}{\Delta}\!+\!\dfrac{1}{6}k_-' s^3 \dfrac{\Delta_1^3}{\Delta}\!+\!o(\Delta_2^2) \end{vmatrix}} \\ &{=}\,\Delta \biggl(\frac{1}{2}ks^3(\Delta_2 -\Delta_1)+\frac{1}{6}k_+' s^4\frac{\Delta_2^3}{\Delta} +\frac{1}{6}k_-'s^4\frac{\Delta_1^3}{\Delta}+ks^3\Delta_1 -\frac{1}{2}k_-' s^4\Delta_1^2+o(\Delta_{12}^2)\biggr) \\ &{=}\,s^3\Delta \biggl(\frac{1}{2}k(\Delta_2+\Delta_1)+\frac{1}{6}k_+' s\frac{\Delta_2^3}{\Delta} +\frac{1}{6}k_-'s\frac{\Delta_1^3}{\Delta}-\frac{1}{2}k_-' s\Delta_1^2+o(\Delta_{12}^2)\biggr) \\ &{=}\,s^3\Delta^2\biggl(\frac{1}{2}k+\frac{1}{6}k_+' s\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}k_-' s\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}+o(\Delta_{12})\biggr). \end{aligned} \end{equation*} \notag
The difference is transformed using (2.12) and (2.13):
\begin{equation} \begin{aligned} \, \notag \gamma_1 (t_2)-\gamma_1 (t_1) &=\gamma_1+s\Delta_2 -\frac{1}{6}k^2s^3\Delta_2^3 -\gamma_1+s\Delta_1-\frac{1}{6}k^2s^3\Delta_1^3+o(\Delta_{12}^3) \\ \notag &=s(\Delta_2+\Delta_1)-\frac{1}{6}k^2s^3\Delta_2^3 -\frac{1}{6}k^2s^3\Delta_1^3+o(\Delta_{12}^3) \\ \notag &=s\Delta -\frac{1}{6}k^2s^3(\Delta_2^3+\Delta_1^3)+o(\Delta_{12}^3) \\ &=s\Delta \biggl(1-\frac{1}{6}k^2s^2(\Delta_2^2 -\Delta_1 \Delta_2+\Delta_1^2) +o(\Delta_{12}^2)\biggr) \notag \\ &=s\Delta(1+o(\Delta_{12})) . \end{aligned} \end{equation} \tag{2.18}
In view of the last two equalities, the minuend in (2.8) assumes the form
\begin{equation} \begin{aligned} \, \notag &\frac{\det (\gamma' (t_1),\gamma (t_2)-\gamma (t_1))}{(\gamma_1 (t_2)-\gamma_1 (t_1))^2} \\ \notag &\qquad =\frac{s^3\Delta^2(\frac{1}{2}k+\frac{1}{6}k_+'s\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}k_-' s\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}+o(\Delta_{12}))} {s^2\Delta^2(1+o(\Delta_{12}))} \\ &\qquad = s\biggl(\frac{1}{2}k+\frac{1}{6}k_+' s\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}k_-' s\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}k_-'s\frac{\Delta_1^2}{\Delta}+o(\Delta_{12})\biggr) . \end{aligned} \end{equation} \tag{2.19}

Let us now consider the subtrahend in (2.8). Expanding in succession the factors in its numerator and denominator and using (2.12)(2.17), we find that

\begin{equation*} \begin{gathered} \, \begin{aligned} \, &\gamma_2 (t_2)-\gamma_2 (t_1) \\ &\qquad=\Delta\biggl(\frac{1}{2}ks^2(\Delta_2 -\Delta_1)+\frac{1}{6}k_+' s^3\frac{\Delta_2^3}{\Delta} +\frac{1}{6}k_-'s^3\frac{\Delta_1^3}{\Delta}+o(\Delta_{12}^2)\biggr) \\ &\qquad=s^2\Delta^2\biggl(\frac{1}{2}k\frac{\Delta_2 -\Delta_1}{\Delta} +\frac{1}{6}k_+' s\frac{\Delta_2^3}{\Delta^2}+\frac{1}{6}k_-'s\frac{\Delta_1^3}{\Delta^2}+o(\Delta_{12})\biggr), \end{aligned} \\ \begin{aligned} \, &\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1) \\ &\qquad =\biggl(-ks^2\Delta_1+\frac{1}{2}k_-' s^3\Delta_1^2+o(\Delta_1^2)\biggr) \biggl(s-\frac{1}{2}k^2s^3\Delta_2^2+o(\Delta_2^2)\biggr) \\ &\qquad\qquad +\biggl(ks^2\Delta_2+\frac{1}{2}k_+' s^3\Delta_2^2+o(\Delta_2^2)\biggr)\biggl(s-\frac{1}{2}k^2s^3\Delta_1^2+o(\Delta_1^2)\biggr) \\ &\qquad =-ks^3\Delta_1^+\frac{1}{2}k_-' s^4\Delta_1^2+ks^3\Delta_2 +\frac{1}{2}k_+' s^4\Delta_2^2+o(\Delta_{12}^2) \\ &\qquad =ks^3(\Delta_2 -\Delta_1)+\frac{1}{2}k_-' s^4\Delta_1^2+\frac{1}{2}k_+' s^4\Delta_2^2+o(\Delta_{12}^2), \end{aligned} \\ \begin{aligned} \, \det (\gamma' (t_1),\gamma''(t_1)) &=\begin{vmatrix} s-\dfrac{1}{2}k^2s^3\Delta_1^2+o(\Delta_1^2) &-ks^2\Delta_1+\dfrac{1}{2}k_-' s^3\Delta_1^2+o(\Delta_1^2) \\ k^2s^3\Delta_1+o(\Delta_1) &ks^2-k_-'s^3\Delta_1+o(\Delta_1) \end{vmatrix} \\ &=ks^3-k_-' s^4\Delta_1+o(\Delta_1) =s^3(k-k_-' s\Delta_1+o(\Delta_1)), \end{aligned} \\ s(t_2)=s+\Delta_2 s'+o(\Delta_2)\quad\text{and} \quad s(t_1)=s-s'\Delta_1+o(\Delta_1). \end{gathered} \end{equation*} \notag
Using these expansions and employing (2.18), for the minuend in (2.8) we have
\begin{equation*} \begin{aligned} \, &\frac{(\gamma_2 (t_2)-\gamma_2 (t_1))s(t_2)\det (\gamma' (t_1),\gamma''(t_1))} {(\gamma_1 (t_2)-\gamma_1 (t_1))s(t_1)(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))} \\ &=s^2\Delta^2\biggl(\frac{1}{2}k\frac{\Delta_2 -\Delta_1}{\Delta} +\frac{1}{6}k_+' s\frac{\Delta_2^3}{\Delta^2}+\frac{1}{6}k_-'s\frac{\Delta_1^3}{\Delta^2}+o(\Delta_{12})\biggr) \\ &\quad \times \frac{s^3(k-k_-' s\Delta_1+o(\Delta_1))}{s\Delta(1-\frac{1}{6}k^2s^2(\Delta_2^2 -\Delta_1\Delta_2+\Delta_1^2)+o(\Delta_{12}^2))} \\ &\quad\times\frac{s+\Delta_2 s'+o(\Delta_2)}{(ks^3(\Delta_2 -\Delta_1)\,{+}\,\frac{1}{2}k_-'s^4\Delta_1^2\,{+}\,\frac{1}{2}k_+' s^4\Delta_2^2\,{+}\,o(\Delta_{12}^2))(s\,{-}\,s'\Delta_1\,{+}\,\frac{1}{2}s''\Delta_1^2\,{+}\,o(\Delta_1^2))} \\ &=s^5\Delta^2 \biggl(\frac{1}{2}k^2\frac{\Delta_2 \,{-}\,\Delta_1}{\Delta}\,{+}\,\frac{1}{6}kk_+' s\frac{\Delta_2^3}{\Delta^2} \,{+}\,\frac{1}{6}kk_-' s\frac{\Delta_1^3}{\Delta^2}\,{-}\,\frac{1}{2}kk_-'s\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}\,{+}\,o(\Delta_{12})\!\biggr) \\ &\quad \times \frac{s+\Delta_2 s'+o (\Delta_2)}{s^4\Delta (k(\Delta_2 -\Delta_1)+\frac{1}{2}k_-' s\Delta_1^2+\frac{1}{2}k_+' s\Delta_2^2\,{+}\,o(\Delta_{12}^2)) (s\,{-}\,s'\Delta_1\,{+}\,\frac{1}{2}s''\Delta_1^2\,{+}\,o(\Delta_1^2))} \\ &=s\Delta \biggl(\frac{1}{2}k^2s\frac{\Delta_2 -\Delta_1}{\Delta}+\frac{1}{6}kk_+' s^2\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}kk_-'s^2\frac{\Delta_1^3}{\Delta^2} \\ &\quad \qquad -\frac{1}{2}kk_-' s^2\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta} +\frac{1}{2}k^2ss'\frac{(\Delta_2 -\Delta_1)\Delta_2}{\Delta}+o (\Delta_{12}) \biggr) \\ &\quad \times\frac{1}{ks (\Delta_2 -\Delta_1)+\frac{1}{2}k_-' s^2\Delta_1^2+\frac{1}{2}k_+' s^2\Delta_2^2 -kss' (\Delta_2 -\Delta_1)\Delta_1+o (\Delta_{12}^2)} \\ &=\biggl(\frac{1}{2}k^2s\frac{\Delta_2 -\Delta_1}{\Delta}+\frac{1}{6}kk_+' s^2\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}kk_-'s^2\frac{\Delta_1^3}{\Delta^2} \\ &\quad \qquad -\frac{1}{2}kk_-' s^2\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta} +\frac{1}{2}k^2ss'\frac{(\Delta_2 -\Delta_1)\Delta_2}{\Delta}+o(\Delta_{12}) \biggr) \\ &\quad \times\frac{1}{k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o(\Delta_{12})}. \end{aligned} \end{equation*} \notag
Note that
\begin{equation*} \begin{aligned} \, s(t_1)s'(t_1) &=\frac{s(t_1)\langle \gamma'(t_1),\gamma''(t_1)\rangle}{s(t_1)} =\langle \gamma'(t_1),\gamma''(t_1) \rangle \\ &=\biggl(s-\frac{1}{2}k^2s^3\Delta_1^2+o(\Delta_1^2)\biggr)(k^2s^3\Delta_1+o(\Delta_1)) \\ &\qquad+\biggl(-ks^2\Delta_1+\frac{1}{2}k_-' s^3\Delta_1^2+o(\Delta_1^2)\biggr)(ks^2-k_-' s^3\Delta_1+o(\Delta_1)) \\ &=k^2s^4\Delta_1 -k^2s^4\Delta_1+o(\Delta_1)=o(\Delta_1). \end{aligned} \end{equation*} \notag
By the mean value theorem,
\begin{equation*} s(t_1)s'(t_1)=((s'(\vartheta))^2+s(\vartheta) s''(\vartheta))\Delta_1,\qquad t_0<\vartheta<t_1. \end{equation*} \notag
Hence ss'=C\Delta_1+o(\Delta_1), C=-((s'(\vartheta))^2+s(\vartheta) s''(\vartheta)) . Therefore, the numerator of the fraction under consideration contains the infinitely small quantity
\begin{equation*} \frac{k^2s s'}{2}\frac{(\Delta_2 -\Delta_1)\Delta_2}{\Delta}=o (\Delta_{12}^2) \end{equation*} \notag
of higher order with respect to o (\Delta_1). Discarding this quantity, the fraction simplifies as
\begin{equation} \begin{aligned} \, \notag &\frac{(\gamma_2 (t_2)-\gamma_2 (t_1))s(t_2)\det (\gamma'(t_1),\gamma''(t_1))} {(\gamma_1 (t_2)-\gamma_1 (t_1))s(t_1)(\gamma_2' (t_1)\gamma_1' (t_2)+\gamma_2' (t_2)\gamma_1' (t_1))} \\ &\ =\frac{\frac{1}{2}k^2s\frac{\Delta_2-\Delta_1}{\Delta} +\frac{1}{6}kk_+' s^2\frac{\Delta_2^3}{\Delta^2}+\frac{1}{6}kk_-'s^2\frac{\Delta_1^3}{\Delta^2} -\frac{1}{2}kk_-' s^2\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o(\Delta_{12})} {k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1+o(\Delta_{12})}{\Delta}+o(\Delta_{12})}. \end{aligned} \end{equation} \tag{2.20}
Using (2.19) and (2.20) and applying some algebraic transformations, we can write the partial derivative \dfrac{\partial G(t_1,t_2)}{\partial t_1} for t_2=t_2 (t_1) as follows:
\begin{equation} \begin{aligned} \, \notag &\frac{\partial G(t_1,t_2 (t_1))}{\partial t_1} =\biggl(\frac{1}{2}ks+\frac{1}{6}k_+' s^2\frac{\Delta_2^3}{\Delta^2}+\frac{1}{6}k_-' s^2\frac{\Delta_1^3}{\Delta^2} -\frac{1}{2}k_-'s^2\frac{\Delta_1^2}{\Delta}+o(\Delta_{12})\biggr) \\ \notag &\qquad\qquad -\frac{\frac{1}{2}k^2s\frac{\Delta_2 -\Delta_1}{\Delta}+\frac{1}{6}kk_+' s^2\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}kk_-'s^2\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}kk_-' s^2\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta} +o(\Delta_{12})}{k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o (\Delta_{12})} \\ \notag &=\biggl(\frac{1}{2}ks+\frac{1}{6}k_+' s^2\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}k_-' s^2\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}k_-'s^2\frac{\Delta_1^2}{\Delta}+o (\Delta_{12})\biggr) \\ \notag &\qquad \times \frac{k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o (\Delta_{12})} {k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o (\Delta_{12})} \\ \notag &\qquad\qquad -\frac{\frac{1}{2}k^2s\frac{\Delta_2 -\Delta_1}{\Delta}+\frac{1}{6}kk_+' s^2\frac{\Delta_2^3}{\Delta^2} +\frac{1}{6}kk_-'s^2\frac{\Delta_1^3}{\Delta^2}-\frac{1}{2}kk_-' s^2\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta} +o(\Delta_{12})}{k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o(\Delta_{12})} \\ \notag &=\frac{\frac{1}{6}kk_-' s^2(\frac{3\Delta_1^2}{2\Delta}+\frac{\Delta_1^3 (\Delta_2 -\Delta_1)}{\Delta^3} -\frac{3\Delta_1^2 (\Delta_2 -\Delta_1)}{\Delta^2}-\frac{\Delta_1^3}{\Delta^2} +\frac{3(\Delta_2 -\Delta_1)\Delta_1}{\Delta})}{k\frac{(\Delta_2 -\Delta_1)}{\Delta} +\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+' s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o(\Delta_{12})} \\ &\qquad\qquad +\frac{\frac{1}{6}kk_+' s^2 (\frac{3\Delta_2^2}{2\Delta}+\frac{\Delta_2^3 (\Delta_2 -\Delta_1)}{\Delta^3} -\frac{\Delta_2^3}{\Delta^2})+o(\Delta_{12})} {k\frac{(\Delta_2-\Delta_1)}{\Delta} +\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}+\frac{1}{2}k_+' s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta} +o (\Delta_{12})}. \end{aligned} \end{equation} \tag{2.21}
In the expansion (2.21) of the partial derivative \dfrac{\partial G(t_1,t_2)}{\partial t_1} along the diffeomorphism t_2=t_2 (t_1), we factor out \Delta_1=t_0 -t_1 and evaluate, using (2.6) and (2.7), the limit of the fraction in brackets, cancelling out the common factor k(t_0)\ne 0 in the numerator and denominator (see (2.1)). As a result, we have
\begin{equation} \begin{aligned} \, \notag &\lim_{\Delta_1 \downarrow 0}\biggl[\frac{\frac{1}{6}kk_-'s^2(\frac{3\Delta_1}{2\Delta} +\frac{\Delta_1^2 (\Delta_2 -\Delta_1)}{\Delta^3}-\frac{3\Delta_1 (\Delta_2-\Delta_1)}{\Delta^2} -\frac{\Delta_1^2}{\Delta^2}+\frac{3 (\Delta_2 -\Delta_1)}{\Delta})} {k\frac{(\Delta_2 -\Delta_1)}{\Delta} +\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}+\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta} -ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o (\Delta_{12})} \\ \notag &\qquad\qquad +\frac{\frac{1}{6}kk_+' s^2(\frac{\Delta_2^3 (\Delta_2-\Delta_1)}{\Delta_1 \Delta^3} +\frac{3\Delta_2^2}{2\Delta\Delta_1}-\frac{\Delta_2^3}{\Delta_1 \Delta^2})+\varepsilon(\Delta_{12})} {k\frac{(\Delta_2 -\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2 -\Delta_1)\Delta_1}{\Delta}+o(\Delta_{12})}\biggr] \\ \notag &\qquad =\frac{\frac{1}{6}kk_-' s^2(\frac{3}{2}\frac{1}{1-\lambda}-\frac{1+\lambda}{(1-\lambda)^3} +\frac{3+3\lambda}{(1-\lambda)^2}-\frac{1}{(1-\lambda)^2}-\frac{3+3\lambda}{(1-\lambda)})}{-\frac{1+\lambda}{1-\lambda}k} \\ \notag &\qquad\qquad +\frac{\frac{1}{6}kk_+' s^2(\frac{\lambda^4+\lambda^3}{(1-\lambda)^3}+\frac{3}{2}\frac{\lambda^2}{(1-\lambda)} +\frac{\lambda^3}{(1-\lambda)^2})}{-\frac{1+\lambda}{1-\lambda}k} \\ \notag &\qquad =\frac{\frac{1}{6}k_-' s^2(\frac{3}{2}-\frac{1+\lambda}{(1-\lambda)^2}+\frac{2+3\lambda}{(1-\lambda)}-3-3\lambda) l+\frac{1}{6}k_+' s^2(\frac{\lambda^4+\lambda^3}{(1-\lambda)^2}+\frac{3}{2}\lambda^2+\frac{\lambda^3}{(1-\lambda)})}{-(1+\lambda)} \\ &\qquad =-\frac{1}{12}k_-' s^2\frac{-6\lambda^3+3\lambda^2-1}{(1-\lambda)^2(1+\lambda)} -\frac{1}{12}k_+' s^2\lambda^2\frac{3\lambda^2-2\lambda+3}{(1-\lambda)^2(1+\lambda)} . \end{aligned} \end{equation} \tag{2.22}
This gives us the following linear asymptotics of the function \dfrac{\partial G(t_1,t_2(t_1))}{\partial t_1} to the left of t_0\in \mathbb{R}:
\begin{equation} \frac{\partial G(t_1,t_2 (t_1))}{\partial t_1}\sim C_1 (\lambda)(t_0 -t_1), \qquad t_1 \to t_0-0. \end{equation} \tag{2.23}
Here the coefficient in the asymptotic expression is equal to the limit of (2.22),
\begin{equation*} C_1 (\lambda)=-\frac{1}{12}k_-' s^2\frac{-6\lambda^3+3\lambda^2-1}{(1-\lambda)^2(1+\lambda)} -\frac{1}{12}k_+' s^2\lambda^2\frac{3\lambda^2-2\lambda+3}{(1-\lambda)^2(1+\lambda)}. \end{equation*} \notag
Recalling the symmetry consideration, in the expansion of \dfrac{\partial G(t_1,t_2 (t_1))}{\partial t_2} we use the corresponding components of \dfrac{\partial G(t_1,t_2 (t_1))}{\partial t_1}, where we replace the increment \Delta_{2} by -\Delta_{1}, \Delta_{1} by -\Delta_{2}, \Delta by -\Delta , k_-' by k_+', and k_+' by k_-' . Now the denominator of the fraction in (2.5) assumes the form
\begin{equation} \begin{aligned} \, \notag \frac{\partial G(t_1,t_2 (t_1))}{\partial t_2} &=\frac{\frac{1}{6}kk_+' s^2(-\frac{3\Delta_3^2}{2\Delta} +\frac{\Delta_2^3 (\Delta_2-\Delta_1)}{\Delta^3} -\frac{3\Delta_2^2 (\Delta_2-\Delta_1)}{\Delta^2} +\frac{\Delta_2^3}{\Delta^2}+\frac{3(\Delta_2-\Delta_1)\Delta_2}{\Delta})} {k\frac{(\Delta_2-\Delta_1)}{\Delta}+\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta} +\frac{1}{2}k_+' s\frac{\Delta_2^2}{\Delta}-ks'\frac{(\Delta_2-\Delta_1)\Delta_2}{\Delta}+o(\Delta_{12})} \\ &\qquad + \frac{\frac{1}{6}kk_-' s^2(\frac{\Delta_1^3 (\Delta_2-\Delta_1)}{\Delta^3} -\frac{3\Delta_1^2}{2\Delta}+\frac{\Delta_1^3}{\Delta^2})+o(\Delta_{12})}{-k\frac{(\Delta_2-\Delta_1)}{\Delta} -\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}-\frac{1}{2}k_+' s\frac{\Delta_2^2}{\Delta} -ks'\frac{(\Delta_2-\Delta_1)\Delta_2}{\Delta}+o(\Delta_{12})}. \end{aligned} \end{equation} \tag{2.24}
Proceeding similarly, we factor out \Delta_1=t_0 -t_1 in the expansion (2.24) of the partial derivative \dfrac{\partial G(t_1,t_2)}{\partial t_2} along the diffeomorphism t_2=t_2 (t_1) and take the limit in the fraction as this increment tends to zero, with the use of condition (2.1) and the limit relations (2.6) and (2.7). As a result, we obtain
\begin{equation*} \begin{aligned} \, & \lim_{\Delta_1 \downarrow 0} \biggl[ \frac{\frac{1}{6}kk_+'s^2(-\frac{3\Delta_3^2}{2\Delta \Delta_1} +\frac{\Delta_2^3(\Delta_2-\Delta_1)}{\Delta_1 \Delta^3} -\frac{3\Delta_2^2 (\Delta_2-\Delta_1)}{\Delta_1\Delta^2} +\frac{\Delta_2^3}{\Delta_1 \Delta^2} +\frac{3(\Delta_2-\Delta_1)\Delta_2}{\Delta_1 \Delta})}{-k\frac{(\Delta_2-\Delta_1)}{\Delta} -\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}-\frac{1}{2}k_+'s\frac{\Delta_2^2}{\Delta} -ks'\frac{(\Delta_2-\Delta_1)\Delta_2}{\Delta}+o(\Delta_{12})} \\ &\qquad\qquad +\frac{\frac{1}{6}kk_-' s^2(\frac{\Delta_1^2 (\Delta_2-\Delta_1)}{\Delta^3}-\frac{3\Delta_1}{2\Delta} +\frac{\Delta_1^2}{\Delta^2})+\varepsilon(\Delta_{12})}{-k\frac{(\Delta_2-\Delta_1)}{\Delta} -\frac{1}{2}k_-' s\frac{\Delta_1^2}{\Delta}-\frac{1}{2}k_+' s\frac{\Delta_2^2}{\Delta} -ks'\frac{(\Delta_2-\Delta_1)\Delta_2}{\Delta}+o(\Delta_{12})}\biggr] \\ &\qquad =\frac{\frac{1}{6}kk_+' s^2(-\frac{3}{2}\frac{\lambda^2}{1-\lambda} +\frac{\lambda^3(1+\lambda)}{(1-\lambda)^3}+3\frac{\lambda^2(1+\lambda)}{(1-\lambda)^2} -\frac{\lambda^3}{(1-\lambda)^2}+\frac{3\lambda^2+3\lambda}{1-\lambda})}{\frac{1+\lambda}{1-\lambda}k} \\ &\qquad\qquad +\frac{\frac{1}{6}kk_-' s^2(-\frac{(1+\lambda)}{(1-\lambda)^3}-\frac{3}{2}\frac{1}{(1-\lambda)} +\frac{1}{(1-\lambda)^2})}{\frac{1+\lambda}{1-\lambda}k} \\ &\qquad =\frac{1}{12}k_+' s^2\frac{\lambda^4-3\lambda^2+6\lambda}{(1-\lambda)^2(1+\lambda)} +\frac{1}{12}k_-'s^2\frac{-3\lambda^2+2\lambda -3}{(1-\lambda)^2(1+\lambda)}. \end{aligned} \end{equation*} \notag
This gives us the linear asymptotics of the function \dfrac{\partial G(t_1,t_2(t_1))}{\partial t_2} to the left of {t_0\in \mathbb{R}}:
\begin{equation} \frac{\partial G(t_1,t_2 (t_1))}{\partial t_2}\sim C_2 (\lambda)(t_0 -t_1), \qquad t_1 \to t_0-0. \end{equation} \tag{2.25}
In view of (2.23) and (2.25) equality (2.5) assumes the form \lambda=\dfrac{C_1(\lambda)}{C_2 (\lambda)}, that is,
\begin{equation*} \lambda=-\frac{-\frac{1}{12}k_-' s^2\frac{-6\lambda^3+3\lambda^2-1}{(1-\lambda)^2(1+\lambda)} -\frac{1}{12}k_+' s^2\lambda^2\frac{3\lambda^2-2\lambda+3}{(1-\lambda)^2(1+\lambda)}} {\frac{1}{12}k_+' s^2\frac{\lambda^4-3\lambda^2+6\lambda}{(1-\lambda)^2(1+\lambda)} +\frac{1}{12}k_-' s^2\frac{-3\lambda^2+2\lambda -3}{(1-\lambda)^2(1+\lambda)}}. \end{equation*} \notag
The structure of this equality is typical for fixed points. Cancelling the numerator and denominator by \frac{1}{12}s^2\ne 0 and performing some algebraic transformations, we single out the zeros of the polynomials in the numerators and denominators:
\begin{equation*} \begin{aligned} \, & k_+' \frac{\lambda^5-3\lambda^3+6\lambda^2}{(1-\lambda)^2(1+\lambda)} +k_-' \frac{-3\lambda^3+2\lambda^2-3\lambda}{(1-\lambda)^2(1+\lambda)} \\ &\qquad\qquad -k_-' \frac{-6\lambda^3+3\lambda^2-1}{(1-\lambda)^2(1+\lambda)}-k_+' \frac{3\lambda^4-2\lambda^3 +3\lambda^2}{(1-\lambda)^2(1+\lambda)}=0, \\ & k_+' \frac{\lambda^5-3\lambda^4-\lambda^3+3\lambda^2}{(1-\lambda)^2(1+\lambda)} +k_-' \frac{3\lambda^3-\lambda^2-3\lambda+1}{(1-\lambda)^2(1+\lambda)}=0, \\ & k_+' \frac{\lambda^2(1+\lambda)(\lambda -1)(\lambda -3)}{(1-\lambda)^2 (1+\lambda)}+k_-' \frac{(1+\lambda)(\lambda-1)(3\lambda -1)}{(1-\lambda)^2(1+\lambda)}=0 . \end{aligned} \end{equation*} \notag
Note that 1-\lambda >0 because \lambda\in\Lambda=[ {-\infty,0} ]. In addition, 1+\lambda \ne 0 because the value of the marker \lambda=-1 is realized in the case of a pseudovertex with smooth curvature [13], which we exclude. Canceling the common factors in the numerators and denominators, we have
\begin{equation*} k_+' \frac{\lambda^2(3-\lambda)}{(1-\lambda)}+k_-' \frac{1-3\lambda}{(1-\lambda)}=0. \end{equation*} \notag
Dividing both parts by (1-\lambda)^2\ne 0, we obtain
\begin{equation} k_+' \frac{\lambda^2(3-\lambda)}{(1-\lambda)^3}+k_-' \frac{1-3\lambda}{(1-\lambda)^3}=0. \end{equation} \tag{2.26}
Thus, \alpha (\lambda)k' (t_0+0)+\beta (\lambda)k' (t_0 -0)=0, where
\begin{equation*} \begin{gathered} \, \alpha (\lambda)=\frac{\lambda^2(3-\lambda)}{(1-\lambda)^3}\quad\text{and} \quad \beta (\lambda)=\frac{1-3\lambda}{(1-\lambda)^3}; \\ \alpha (\lambda)\geqslant 0, \qquad \beta (\lambda)\geqslant 0\quad\ \ \text{and} \quad\ \ \alpha (\lambda)+\beta (\lambda)=1. \end{gathered} \end{equation*} \notag
Here we have taken into account that, by the binomial formula, the numerators of the corresponding fractional coefficients give in combination a cubed binomial from the denominators of the fractions:
\begin{equation*} \lambda^2(3-\lambda)+1-3\lambda=1-3\lambda+3\lambda^2-\lambda^3=(1-\lambda)^3. \end{equation*} \notag
In addition, these numerators are nonnegative because \lambda \leqslant 0. Hence so are the coefficients themselves, because their common denominator (1-\lambda)^3 is positive.

Now (2.3) follows from (2.26). It is worth pointing out that the coefficients of the convex combination of the one-sided curvatures are found constructively.

This proves the theorem.

Proposition. Under the assumptions of the theorem in § 2, if the left-hand marker \lambda=t_2' (t_0 -0) is negative, then, for the right-hand marker \mu=t_1' (t_0 +0), there exist

\begin{equation*} \widetilde{\alpha} (\mu)\geqslant0\quad\text{and} \quad \widetilde{\beta} (\mu)\geqslant0, \qquad \widetilde{\alpha} (\mu)+\widetilde{\beta}_1 (\mu)=1 \end{equation*} \notag
(depending on this marker), such that
\begin{equation} \widetilde{\alpha}(\mu)k_+'+\widetilde{\beta}(\mu)k_-'=0. \end{equation} \tag{2.27}

Proof. The one-sided markers are mutually reciprocal (see (1.14)). Substituting \lambda=\mu^{-1}< 0 into (2.26) and making some transformations,
\begin{equation*} \begin{gathered} \, k_+' \frac{\mu^{-2}(3-\mu^{-1})}{(1-\mu^{-1})^3}+k_-' \frac{1-3\mu^{-1}}{(1-\mu^{-1})^3}=0, \\ k_+' \frac{3\mu -1}{(\mu -1)^3}+k_-' \frac{\mu^2(\mu -3)}{(\mu -1)^3}=0, \end{gathered} \end{equation*} \notag
we obtain the required expansion (2.27), where
\begin{equation*} \widetilde{\alpha}(\mu)=\frac{3\mu -1}{(\mu -1)^3}\geqslant 0\quad\text{and} \quad \widetilde{\beta}(\mu)=\frac{\mu^2(\mu -3)}{(\mu -1)^3}\geqslant 0, \qquad \widetilde{\alpha}(\mu)+\widetilde{\beta}(\mu)=1, \end{equation*} \notag
since \mu <0. This proves the proposition.

Remark 7. The fact that convex combinations of the one-sided curvatures vanish, established in the theorem and proposition in § 2 for a pseudovertex of the curve with failure of the smoothness of the curvature, is in line with the result of [13], where the class of C^3-curves was considered and it was shown that the stationarity condition of the curvature is fulfilled at pseudovertices. In this sense, the results of our studies for problem (1.5) reconcile the cases where the boundaries of the boundary sets have a smooth curvature and those with nonsmooth curvature.

§ 3. Formulae for one-sided markers

The formulae for the one-sided markers are of the same type, and there is no significant difference in calculating them. Consider, for example, the problem of finding the left-hand marker. Canceling the normalizing factor \dfrac{1}{(1-\lambda)^3} in (2.26), we obtain

\begin{equation} k_+' \lambda^2(3-\lambda)+k_-' (1-3\lambda)=0. \end{equation} \tag{3.1}
We look at the special case where the one-sided derivatives of the curvature are strictly separated from zero and have different signs, that is,
\begin{equation} \sigma=\frac{k_+'}{k_-'}<0. \end{equation} \tag{3.2}
Here we speak about the situation where the pseudovertex is a point of nonsmooth extremum of the curvature of the boundary of the boundary condition.

Lemma. Under the assumptions of the theorem in § 2, if inequality (3.2) is fulfilled, then the left-hand marker of the pseudovertex can be found by the formula

\begin{equation} \lambda=2\sqrt {\frac{\sigma -1}{\sigma}} \cos \biggl(\frac{1}{3}\arccos\sqrt {\frac{\sigma}{\sigma -1}} -\frac{4\pi}{3}\biggr)+1. \end{equation} \tag{3.3}

Proof. Considering (3.1) as an algebraic equation with respect to the marker, we write it in terms of the coefficient \sigma={k_+'}/{k_-'}:
\begin{equation*} \lambda^3-3\lambda^2+\frac{3}{\sigma}\lambda -\frac{1}{\sigma}=0. \end{equation*} \notag
Let us study its solutions. Applying the Tschirnhaus transformation
\begin{equation} \lambda=y+1 \end{equation} \tag{3.4}
(see § 5.36 in [22]), we write this equation in the reduced form y^3+py+q=0, where p=3({1}/{\sigma}-1) and q=2({1}/{\sigma}-1), that is, as
\begin{equation} y^3+3\biggl(\frac{1}{\sigma}-1\biggr)y+2\biggl(\frac{1}{\sigma}-1\biggr)=0. \end{equation} \tag{3.5}
The constant
\begin{equation*} Q=\biggl(\frac{p}{3}\biggr)^3+\biggl(\frac{q}{2}\biggr)^2 =\biggl(\frac{1}{\sigma}-1\biggr)^3+\biggl(\frac{1}{\sigma}-1\biggr)^2 =\frac{1}{\sigma}\biggl(\frac{1}{\sigma}-1\biggr)^2 \end{equation*} \notag
is proportional to the discriminant of the polynomial. By (3.2), Q<0, and so equation (3.5) has three real solutions, which can be written in the trigonometric form as follows:
\begin{equation*} y_k=2\sqrt {-\frac{p}{3}} \cos \biggl(\frac{1}{3}\arccos \biggl(\frac{3q}{2p}\sqrt {-\frac{3}{p}}\biggr) -\frac{2\pi k}{3}\biggr), \qquad k=0,1,2 \end{equation*} \notag
(see § 1 in [28]). In our case
\begin{equation*} y_k=2\sqrt {\frac{\sigma -1}{\sigma}} \cos \biggl(\frac{1}{3}\arccos \sqrt {\frac{\sigma}{\sigma -1}} -\frac{2\pi k}{3}\biggr), \qquad k=0,1,2. \end{equation*} \notag
Setting \eta=\sqrt {\sigma/(\sigma -1)} for brevity, where 0<\eta<1, we obtain
\begin{equation*} y_k=\frac{2}{\eta}\cos \biggl(\frac{1}{3}\arccos \eta -\frac{2\pi k}{3}\biggr), \qquad k=0,1,2. \end{equation*} \notag
Let us find the possible positions of these roots. Using the classical techniques of analysis, it can easily be shown that the y_k , k=0,1,2, are strictly monotone as functions of \eta \in (0,1). Consider, for example,
\begin{equation*} y_2 (\eta)=\frac{2}{\eta}\cos \biggl(\frac{1}{3}\arccos \eta-\frac{4\pi}{3}\biggr). \end{equation*} \notag
Its derivative is
\begin{equation*} y_2' (\eta)=-2\frac{\varphi'(\eta)\eta\sin\varphi(\eta)+\cos\varphi(\eta)}{\eta^{2}}, \end{equation*} \notag
where
\begin{equation*} \varphi(\eta)=\frac{1}{3} \arccos\eta-\frac{4\pi}{3}\quad\text{and} \quad \varphi'(\eta)=-\frac{1}{3\sqrt{1-\eta^{2}}}. \end{equation*} \notag
Here
\begin{equation*} -\frac{4\pi}{3}<\varphi(\eta)=\frac{1}{3} \arccos\eta-\frac{4\pi}{3}<-\frac{7\pi}{6}, \end{equation*} \notag
for 0 <\eta<1. Consider the normalizing function
\begin{equation*} q(n)=\sqrt{a^2(\eta)+b^2(\eta)}=\frac{1}{3}\sqrt{\frac{9-8\eta^{2}}{1-\eta^{2}}}. \end{equation*} \notag
Calculating the coefficients
\begin{equation*} a(\eta)=\frac{\varphi'(\eta)\eta}{q(n)}=-\frac{\eta}{\sqrt{9-8\eta^{2}}}\quad\text{and} \quad b(\eta)=\frac{1}{q(n)}=3\sqrt{\frac{1-\eta^{2}}{9-8\eta^{2}}}, \end{equation*} \notag
we write the derivative as
\begin{equation*} y_2' (\eta)=-2q(n) \frac{a(\eta)\sin\varphi(\eta)+b(\eta)\cos\varphi(\eta)}{\eta^{2}}. \end{equation*} \notag
We have -1<a(\eta)<0 and 0<b(\eta)<1, a(\eta)^{2}+b(\eta)^{2}=1, and therefore there exists an angle \psi(\eta), {\pi}/{2}<\psi(\eta)<\pi, such that a(\eta)=\cos\psi(\eta) and {b(\eta)=\sin\psi(\eta)}. After some trigonometry we can write this derivative as
\begin{equation*} \begin{aligned} \, y_2' (\eta) &=-2q(n)\frac{\cos\psi(\eta)\sin\varphi(\eta)+\sin\psi(\eta)\cos\varphi(\eta)}{\eta^{2}} \\ &=-2q(n)\frac{\sin(\varphi(\eta)+\psi(\eta))}{\eta^{2}}. \end{aligned} \end{equation*} \notag
In view of the constraints -{4\pi}/{3}<\varphi(\eta)<-{7\pi}/{6} and {\pi}/{2}<\psi(\eta)<\pi the angles range in the interval
\begin{equation*} -\frac{5\pi}{6}<\varphi(\eta)+\psi(\eta)<-\frac{\pi}{6}. \end{equation*} \notag
Hence \sin(\varphi(\eta)+\psi(\eta))<0, for 0<\eta<1. Since -{2q(n)}/{\eta^{2}}<0 , the derivative y_2' (\eta) is positive for all admissible \eta\in(0, 1).

Thus, the function

\begin{equation*} y_2(\eta)=\frac{2}{\eta}\cos \biggl(\frac{1}{3}\arccos \eta- \frac{4\pi}{3}\biggr) \end{equation*} \notag
is strictly increasing, and its range has the boundary points
\begin{equation*} \lim_{\eta \to+0} y_2 (\eta)=-\infty\quad\text{and} \quad \lim_{\eta \to1-0} y_2 (\eta)=y_2 (1)=-1, \end{equation*} \notag
that is,
\begin{equation} -\infty <y_2 <-1. \end{equation} \tag{3.6}
A similar analysis shows that the function
\begin{equation*} y_0 (\eta)=\frac{2}{\eta}\cos \biggl(\frac{1}{3}\arccos \eta\biggr) \end{equation*} \notag
is strictly monotonically decreasing on (0,1), and its range is (2,+\infty), that is,
\begin{equation} 2<y_0 <+\infty. \end{equation} \tag{3.7}
The function
\begin{equation*} y_1 (\eta)=\frac{2}{\eta}\cos \biggl(\frac{1}{3}\arccos \eta -\frac{2\pi}{3}\biggr) \end{equation*} \notag
is also strictly monotonically decreasing and takes values in (-1,-2/3), that is,
\begin{equation} -1<y_1 <-\frac{2}{3}. \end{equation} \tag{3.8}
Now, in view of inequalities (3.6)(3.8) and after the Tschirnhaus transformation (3.4), we see that for each admissible \eta \in (0,1) the original equation has three real roots, two positive and a negative one. The negative root, on which the marker of the pseudovertex is realized, corresponds to the index k=2. Hence
\begin{equation*} \lambda=y_2+1\triangleq 2\sqrt {\frac{\sigma -1}{\sigma}} \cos \biggl(\frac{1}{3}\arccos \sqrt {\frac{\sigma}{\sigma -1}}-\frac{4\pi}{3}\biggr)+1, \end{equation*} \notag
which proves the lemma.

§ 4. Numerical-analytic modelling of a generalized solution

Consider the Dirichlet problem (1.5) on a closed unbounded convex set \Omega \subset \mathbb{R}^2 with C^2-boundary

\begin{equation*} \Gamma=\{{\gamma (t)\in \mathbb{R}^2\colon \gamma (t)=(\gamma_1(t),\gamma_2 (t)),\,t\in T} \}, \end{equation*} \notag
where T=\mathbb{R},
\begin{equation*} \gamma_1 (t) =t\quad\text{and} \quad \gamma_2 (t)= \begin{cases} \dfrac{1}{8}(t^2-2t+4),&t\leqslant 0, \\ (t+2)^{-1},&t>0. \end{cases} \end{equation*} \notag

The curve \Gamma=\partial \Omega has a unique pseudovertex x^{(0)}=(\gamma_1(t_0), \gamma_2(t_0)) corresponding to t_0=0; here T^0=\{0 \}. It is easily seen that the curvature of the curve ceases to be smooth at this point because \gamma_{2,-}''' (t_0)=0\ne \gamma_{2,+}''' (t_0)=-\frac{3}{8}. The curvature k(t_0) is \frac{16}{17^{3/2}}\ne 0. Let us find the ratio of the one-sided curvatures at the pseudovertex and compare it with zero. We have

\begin{equation*} \sigma=\frac{k_+'}{k_-'}=\frac{\gamma_2''' (t_0+0)(1+({\gamma_2' (t_0)})^2) -3(\gamma_2'' (t_0))^2\gamma_2' (t_0)}{\gamma_2''' (t_0 -0)(1+(\gamma_2' (t_0))^2)-3(\gamma_2'' (t_0))^2\gamma_2' (t_0)}=-7.5<0. \end{equation*} \notag
Thus, we are under the assumptions of the theorem in § 2, and condition (3.2) is fulfilled. Using (3.3) to evaluate the left-hand marker, we have
\begin{equation*} \lambda\approx-0.2720. \end{equation*} \notag

To the unique pseudovertex of the boundary condition there corresponds a single branch L(x^{(0)}) of the singular set. To construct this branch one has to know at least one diffeomorphism in the pair of reciprocal diffeomorphisms t_{2}= t_{2}(t_{1}) and t_{1}=t_{1}(t_{2}). In the general case these functions can be found analytically only on rare occasions (see [11]). This is why in our further constructions we have to recourse to numerical methods (see [13]). The graphs (plotted numerically) of the mutually inverse diffeomorphisms t_{2}=t_{2}(t_{1}) and t_{1}= t_{1}(t_{2}) (which generate the pseudovertex) are shown in Figure 1.

This figure illustrates the geometry of one-sided markers in the plane of the parameters t_{1} and t_{2}. The left-hand marker \lambda is the slope to the positive direction of the t_{1}-axis of the one-sided left-hand tangent to the graph t_{2}=t_{2}(t_{1}) at the gluing point of the graphs of the mutually inverse diffeomorphisms. Correspondingly, the right-hand marker \mu=\lambda^{-1} is the slope to the positive direction of the t_{2}-axis of the one-sided right-hand tangent to the graph t_{1}=t_{1}(t_{2}) at the same point.

Next, once both the left-hand marker of the pseudovertex and the local diffeomorphism t_{2}=t_{2}(t_{1}) generating this pseudovertex are known, we construct the branch of the singular set by solving the system of equations (1.11). Note that here we can also follow another approach towards the construction of the branch of a singular set, where we integrate an ordinary differential equation for which the initial conditions are controlled by the markers of the pseudovertex and the dynamics is governed by the local diffeomorphisms generating this pseudovertex (see [14]). These authors developed a software package [29] constructing the level curves of the Euclidean distance function to a closed set M. The main step of the algorithm consists in finding the locus of points lying on normal vectors to the curve \Gamma at a fixed distance r>0 from the feet of the normals. Next, we search for the intersection of this set with the singular set L, after which we exclude the parts lying at distances less than r to M.

Figure 2 shows the singular set obtained numerically and the ‘contour map’ of the generalized eikonal (evolution of wave fronts). Given a set of curves \Phi , the software package returns the table of values of the function u(x_1,x_2) on a rectangular grid in a prescribed compact subset of \Omega.

Figure 3 shows an approximation of the graph of the generalized solution to problem (1.5), which loses smoothness at points in the singular set.


Bibliography

1. S. N. Kružkov (Kruzhkov), “Generalized solutions of the Hamilton–Jacobi equations of eikonal type. I. Formulation of the problems; existence, uniqueness and stability theorems; some properties of the solutions”, Math. USSR-Sb., 27:3 (1975), 406–446  mathnet  crossref  mathscinet  zmath  adsnasa
2. M. G. Crandall and P.-L. Lions, “Viscosity solutions of Hamilton–Jacobi equations”, Trans. Amer. Math. Soc., 277:1 (1983), 1–42  crossref  mathscinet  zmath
3. A. I. Subbotin, Generalized solutions of first-order PDEs. The dynamical optimization perspective, Systems Control Found. Appl., Birkhäuser Boston, Inc., Boston, MA, 1995, xii+312 pp.  crossref  mathscinet  zmath
4. O. A. Oleĭnik, “Discontinuous solutions of non-linear differential equations”, Amer. Math. Soc. Transl. Ser. 2, 26, Amer. Math. Soc., Providence, RI, 1963, 95–172  mathnet  crossref  mathscinet  mathscinet  zmath
5. N. N. Krasovskiĭ and A. I. Subbotin, Game-theoretical control problems, Springer Ser. Soviet Math., Springer-Verlag, New York, 1988, xii+517 pp.  mathscinet  zmath
6. V. I. Arnold, Singularities of caustics and wave fronts, Math. Appl. (Soviet Ser.), 62, Kluwer Acad. Publ., Dordrecht, 1990, xiv+259 pp.  crossref  mathscinet  zmath
7. A. V. Borovskikh, “The two-dimensional eikonal equation”, Siberian Math. J., 47:5 (2006), 813–834  mathnet  crossref  mathscinet  zmath
8. J. A. Sethian and A. Vladimirsky, “Fast methods for the Eikonal and related Hamilton–Jacobi equations on unstructured meshes”, Proc. Natl. Acad. Sci. USA, 97:11 (2000), 5699–5703  crossref  mathscinet  zmath  adsnasa
9. S. I. Kabanikhin and O. I. Ktivorotko, “Numerical solution of the eikonal equation”, Sibirsk. Elektron. Mat. Izv., 10 (2013), 28–34 (Russian)  mathnet
10. G. V. Papakov, A. M. Taras'ev and A. A. Uspenskii, “Numerical approximations of generalized solutions of the Hamilton–Jacobi equations”, J. Appl. Math. Mech., 60:4 (1996), 567–578  crossref  mathscinet  zmath  adsnasa
11. P. D. Lebedev, A. A. Uspenskii and V. N. Ushakov, “Construction of a minimax solution for an eikonal-type equation”, Proc. Steklov Inst. Math. (Suppl.), 263, suppl. 2 (2008), S191–S201  mathnet  crossref  mathscinet  zmath
12. P. D. Lebedev and A. A. Uspenskii, “Conditions for the transversality of branches of a solution of a nonlinear equation in the time-optimal problem with circular indicatrix”, Tr. Inst. Mat. Mekh., 14, no. 4, 2008, 82–99 (Russian)  mathnet
13. A. A. Uspenskii, “Necessary conditions for the existence of pseudovertices of the boundary set in the Dirichlet problem for the eikonal equation”, Tr. Inst. Mat. Mekh., 21, no. 1, 2015, 250–263  mathnet  mathscinet
14. P. D. Lebedev and A. A. Uspenskii, “Constructing the solution of a time-optimality control problem under violation of the smoothness of the curvature of the boundary of the target set”, Izv. Inst. Mat. Inform., 53 (2019), 98–114 (Russian)  mathnet  crossref  mathscinet  zmath
15. V. V. Nemytskiĭ, “The fixed point method in analysis”, Amer. Math. Soc. Transl. Ser. 2, 34, Amer. Math. Soc., Providence, RI, 1963, 1–37
16. V. N. Ushakov and A. A. Uspenskii, “\alpha-sets in finite-dimensional Euclidean spaces and their properties”, Vestn. Udmurt. Univ. Mat. Mekh. Komp'yut. Nauki, 26:1 (2016), 95–120  mathnet  crossref  mathscinet  zmath
17. Th. Bröcker, Differentiable germs and catastrophes, Transl. from the German by L. Lander, London Math. Soc. Lecture Note Ser., 17, Cambridge Univ. Press, Cambridge–New York–Melbourne, 1975, vi+179 pp.  crossref  mathscinet  zmath
18. V. F. Dem'yanov and L. V. Vasil'ev, Nondifferentiable optimization, Transl. Ser. Math. Engrg., Optimization Software, Inc., Publications Division, New York, 1985, xvii+452 pp.  mathscinet  mathscinet  zmath  zmath
19. E. Hopf, “Generalized solutions of non-linear equations of first order”, J. Math. Mech., 14:6 (1965), 951–973  mathscinet  zmath
20. M. Bardi and L. C. Evans, “On Hopf's formulas for solutions of Hamilton–Jacobi equations”, Nonlinear Anal., 8:11 (1984), 1373–1381  crossref  mathscinet  zmath
21. A. M. Taras'ev, A. A. Uspenskij and V. N. Ushakov, “Approximation schemes and finite-difference operators for constructing generalized solutions of Hamilton–Jacobi equations”, J. Comput. Syst. Sci. Int., 33:6 (1995), 127–139
22. J. W. Bruce and P. J. Giblin, Curves and singularities. A geometrical introduction to singularity theory, Cambridge Univ. Press, Cambridge, 1984, xii+222 pp.  mathscinet  zmath
23. A. A. Uspenskii and P. D. Lebedev, “Degeneracy of second-order conditions in the construction of pseudovertices of the boundary set for the eikonal equation”, Vestn. Humanities Univ., 2 (2016), 31–46 (Russian)
24. A. R. Alimov and I. G. Tsar'kov, “Connectedness and solarity in problems of best and near-best approximation”, Russian Math. Surveys, 71:1 (2016), 1–77  mathnet  crossref  mathscinet  zmath  adsnasa
25. N. V. Efimov and S. B. Stechkin, “Some properties of Chebyshev sets”, Dokl. Akad. Nauk SSSR, 118:1 (1958), 17–19 (Russian)  mathnet  mathscinet  zmath
26. V. D. Sedykh, “On the topology of symmetry sets of smooth submanifolds in \mathbb{R}^{k}”, Singularity theory and its applications, Adv. Stud. Pure Math., 43, Math. Soc. Japan, Tokyo, 2006, 401–419  crossref  mathscinet  zmath
27. R. N. Shcherbakov and L. F. Pichurin, Differentials aiding geometry, Prosveshchenie, Moscow, 1982, 191 pp. (Russian)
28. A. D. Polyanin, Explicit solutions of differential, integral, functional and other mathematics equations, Publishing House of the Institute for Problems in Mechanics of the Russian Academy of Sciences, Moscow, 2023, 600 pp. (Russian)
29. P. D. Lebedev and A. A. Uspenskii, Software package for constructing wave fronts and the function of Euclidean distance to a nonconvex compact set, State Certificate of Software Registration no. 2017662074 dated 27.10.2017

Citation: P. D. Lebedev, A. A. Uspenskii, “Numerical-analytic construction of a generalized solution to the eikonal equation in the plane case”, Sb. Math., 215:9 (2024), 1224–1248
Citation in format AMSBIB
\Bibitem{LebUsp24}
\by P.~D.~Lebedev, A.~A.~Uspenskii
\paper Numerical-analytic construction of a~generalized solution to the eikonal equation in the plane case
\jour Sb. Math.
\yr 2024
\vol 215
\issue 9
\pages 1224--1248
\mathnet{http://mi.mathnet.ru/eng/sm9500}
\crossref{https://doi.org/10.4213/sm9500e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4837041}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024SbMat.215.1224L}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001375658800006}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85212481775}
Linking options:
  • https://www.mathnet.ru/eng/sm9500
  • https://doi.org/10.4213/sm9500e
  • https://www.mathnet.ru/eng/sm/v215/i9/p99
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:366
    Russian version PDF:11
    English version PDF:13
    Russian version HTML:72
    English version HTML:96
    References:28
    First page:9
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2025