Model BVP Problem for the Helmholtz equation in a nonconvex angle with periodic boundary data
Abstract
In the presented work, we solve the Dirichlet boundary problem for the Helmholtz equation in an exterior angle with periodic boundary data. We prove the existence and uniqueness of solution in an appropriate funcional class and we give an explicit formula for it in the form of the Sommerfeld integral. The method of complex characteristics [17] is used.
1 Introduction
We consider the following model boundary value problem (BVP) for the Helmholtz equation in a plane angle of magnitude with a complex frequency :
(1.1) |
Here for are the sides of the angle , or ( is the exterior normal to ), are given functions which can be distributions, see Fig. 1

Problems of this type arise is many areas of mathematical physics.
We list some of them.
Firstly, such BVP describe the diffusion of a desintegrating gas [37, Ch. VII, cf. 2].
Secondly, diffraction problems by the wedge are reduced to such problems for a lossy medium [36] or for a slightly conducting medium [7].
Thirdly, time-dependent diffraction by wedges [34, 11, 10, 35, 30, 9, 40, 39, 20]
is reduced to a problem of this type after the Fourier-Laplace (F-L) transform with respect to time [14, 15, 19, 8].
Fourthly, the problem of scattering of waves emitted by a point source by is reduced
to this problem, with
Let us describe this scattering problem in more detail since it was precisely this problem which was the starting point of the present paper.
Let us consider the following time-dependent scattering problem:
(1.2) |
Here . After the F-L transform (1.2) becomes equivalent to
(1.3) |
see Fig. 2

We reduce this problem to a homogeneous Helmholtz equation with nonhomogeneous boundary conditions. Let be s.t.
(1.4) |
Passing to the Fourier transform in (1.4), we obtain
Hence
Define
By (1.3), (1.4) this implies that satisfies the problem of type (1.1):
(1.5) |
Let us calculate on . We have
where ; .
Hence,
.
Thus (1.5) is equivalent to
(1.6) |
Therefore it seems natural to solve first the following model problem, corresponding to problem (1.6)
(1.7) |
where .
Note that this problem is similar to the BVP in [14, (23)] arising in the problem of time-dependent diffraction:
(1.8) |
However, there are substantial differences. The exponents in the right-hand side of
this problem are complex () and, thus, the corresponding functions decrease exponentially when since .
Moreover the structure of boundary conditions in (1.8) is connected with the first equation through the common parameter . This gives a unique opportunity to reduce problem (1.8) to a difference equation which is solved easily in an explicit form.
In contrast, problem (1.7) has periodic boundary conditions which are independent of the first equation. This results in the fact that the corresponding difference equation cannot be solved as easily as in the previous case except for the case when (see Section 5).
In turn, problem (1.7) splits into two problems for such that by linearity:
(1.9) |
(1.10) |
Here .
This paper is devoted to solving the model problem (1.9). Solution of (1.10) is obtained from (1.9) by a simple change of variable, (see (2.4)).
Note that the BVP in a right angle or in its complement and in other particular angles whose magnitudes are commensurate with , were considered in many papers [24, 25, 26, 27, 28, 31, 2, 3, 4, 5, 6].
In those papers exact results were obtained by means of operator methods. Boundary data in those papers belong to Sobolev spaces , . We consider another type of boundary data, namely, periodic functions. We obtain exact solutions in explicit form, namely, in the form of Sommerfeld type integrals. We use the method of automorphic functions (MAF) on complex characteristics [17]. This method was developed by A.Komech for in [12] and then was extended to in [17, Section 1.2 and part 2]. It allows us to find all distributional solutions of the BVP for the Helmholtz equation in arbitrary angles with general boundary conditions. It was applied, in particular, to time-dependent diffraction problems by angles [19, 22, 23, 21, 16, 13].
It should be noted that there is a very effective Sommerfeld-Malyzhinetz method of constructing solutions of diffraction problems in angles; by means of this method many important results were obtained [1]. This method allows one to obtain the solution in the form of the Sommerfeld integral. However, this method does not allow one to prove uniqueness which usually is proved on the basis of physical considerations.
We also obtain solution in the Sommerfeld integral form, using the MAF which also allows us to prove uniqueness in an appropriate functional space (see e.g. [24]).
The paper is organized as follows: in Section 2 we formulate the main result, the Sections 3-10 are devoted to its proof. In Section 3 we reduce boundary value problem to a difference equation and we prove the necessary and sufficient conditions for the existence of solution. In Sections 4 and 5 we find solution of the difference equation for and respectively. In Section 6 we prove the asymptotics of the integrand for the Sommerfeld’s representation of the solution. In Sections 7 and 8 we give the Sommerfeld-type representation of solution and we prove the boundary conditions. In Section 9 we prove the existence and uniqueness of the solution. En Appendices we prove some technical assertions.
2 The main result
We will construct the solutions of problems (1.9) and (1.10) in the form of the well-known Sommerfeld integrals which have the form
where is a certain contour on the complex plane and the correct construction of the factor , which ensures that (10.4) satisfies the boundary conditions, is the main difficulty of the problem.
To formulate the main result we need to describe the integrand of the Sommerfeld integral. The construction of this integrand is the main contents of this paper.
Consider given by (3.23) where is given by (4.27) for and by (5.13) for with , given by (3.24).
Let be the polar coordinates in ,
(2.1) |
and be the Sommerfeld double-loop contour, see Fig. 11.
Definition 2.1.
is the space of the functions bounded with its first derivatives in and admitting the following asymptotics at the origin
(2.2) |
Our main result consists in the following statement.
Theorem 2.2.
i) Let . There exists a solution to problem (1.7)
belonging to , where are solutions to (1.9), (1.10), respectively, which admit a Sommerfeld integral representation
(2.3) |
(2.4) |
where in (2.3) is constructed according to the algorithm presented below for , in (2.4) for and
(2.5) |
ii) The solution is unique in .
Remark 2.3.
Remark 2.5.
Note that the solution also admits a slightly different representation where several different Sommerfeld-type contours are used (see (8.3)).
3 Reduction to a difference equation. Necessary and sufficient conditions for the Neumann data
Consider problem (1.9). The MAF permits to reduce this problem to finding the Neumann data of solution , and it consists of several steps. In the following subsections we present these steps.
We assume that the solution .
The first step of the MAF is to reduce the problem to the complement of the first quadrant and to extend the solution to the plane, see [17, 15].
3.1 First step: extension of to the whole plane
Consider the linear transformation
which sends the angle to the right angle . This transformation reduces system (1.9) to the problem (3.1a)-(3.1c) in the complement of the first quadrant for
(3.1a) | ||||
(3.1b) | ||||
(3.1c) |
where
(3.2) |
By [17, Lemma 8.2], if is a solution of equation (3.1a), then there exists an extension of by 0, such that ,
(3.3) |
where and has the form
(3.4) |
.
We will use the extension of the Fourier transform defined on , , to by continuity:
(3.5) |
and denote this extension by the tilde, . Applying this transform to (3.3) and using the fact that , we obtain
(3.6) |
Hence,
(3.7) |
Here is the Fourier transform of (3.4), and for
(3.8) |
where are the Fourier transforms of . Thus, if we know , we know by (3.7), and problem (3.11) is reduced to finding the four functions , , .
Remark 3.1.
Formula (3.4) is obtained by direct differentiation (in the sense of of the discontinuous function
(3.9) |
in the case when . Moreover, the formula
is used for . Obviously, in this case the functions in (3.4) are the Cauchy data of the function :
(3.10) |
It turns out that formula (3.4) and representations (3.10) remain true for distributional solutions. The following two lemmas describe the solution of equation (3.1a) in terms of its Cauchy data.
Lemma 3.2.
Lemma 3.3.
Remark 3.4.
Now we use boundary conditons (3.1b), (3.1c). Let be a solution to (3.1a)-(3.1c) and be its extension by satisfying (3.3); then, by (3.10).
(3.11) |
Since , by the distribution theory, we have, generally speaking,
(3.12) |
for some . Here is defined similarly to (3.9). Obviously where is the Heaviside function. We will find a solution to (3.1a)-(3.1c) for
(3.13) |
Thus we put
(3.14) |
Remark 3.5.
It is not guaranteed a priori that the solution of (3.1a)-(3.1c) exists under condition (3.13) because should satisfy a certain connection equation (see Section 3.2). Nevertheless, it turns out that we are able to construct an explicit solution under the condition (3.13). Solutions which correspond to nonzero values of in (3.12) will only contain additional singularities at the origin, and are not of interest.
Substituting given by (3.14) in (3.3), we obtain
(3.15) |
with containing only two unknown functions and .
The MAF gives the necessary and sufficient conditions for the functions
and , which allow us to find these functions in an explicit form. Substituting these functions in (3.15) we obtain (and so ) by (3.7), (3.4).
In what follows we consider equation (3.15).
3.2 Second step: Fourier-Laplace transform and the lifting to the Riemann surface. Connection equation
In addition to the (real) Fourier transform (3.5) we will use the complex Fourier transform (or Fourier-Laplace (F-L) transform). Let
Then by the Paley-Wiener theorem [38], , admits an analytic continuation and . Since , there exist their F-L transforms
(3.16) |
In particular, from (3.14) we have
(3.17) |
where for , in . Hence, using (3.8) we obtain (since )
(3.18) |
In the MAF, the Riemann surface of complex zeros of the symbol of the operator (3.2) plays an essential role, since a necessary condition for the existence of the solution on can be written in terms of this surface. The symbol of this operator is the polynomial
Obviously, does not have real zeros, but it does have complex ones. Denote the Riemann surface of the complex zeros of by
It is convenient to parametrize the complex surface introducing the parameter .
The Riemann surface admits a universal covering , which is isomorphic to (see [17, Ch. 15]). Let be a parameter on . Then the formulas
(3.19) |
describe an infinitely sheeted covering of onto .
Let us “lift” the functions to . For this we must identify
with regions on . This can be done in many ways. For example, define, for ,
(3.20) |
Obviously, for , . Moreover, For , define
and for , define
For , let us “lift” to . Denote this lifting by . Then
Note that . We choose the connected component of corresponding to the condition as , (see Fig. 3, where are represented for ).

Now we “lift” to , using (3.19). We obtain from (3.17), (3.14)
(3.21) |
Further, are analytic functions in by (3.16). Our aim is to find the unknown functions . Having these functions, we obtain and the solution by (3.7), (3.6).
Note that in the case the function given by (3.18) is not lifted to since is not defined at any point of . In fact, is not defined in , is not defined in since , see Fig. 3. In the case this intersection is not empty and such lifting to is possible [17]. Thus, in that case there exists a connection between and generated by (3.6) since has zeros in and must vanish for .
Nevertheless, a similar relation between and exists in the case too (see [17, Chap. 21]). Let us describe the corresponding construction. The function is naturally splitted into two summands each of which is extended to and , respectively.
Namely
where
By the Paley-Wiener Theorem the function admits an analytic continuation to , and admits an analytic continuation to , where
Now we can “lift” and to the Riemann surface by formulas (3.19).
We obtain
(3.22) |
(3.23) |
where
(3.24) |
In the case , and have a common domain which is not empty, and thus the connection equation has the following form:
(3.25) |
(see [17, Chap. 10]).
In the case the domain (see Fig. 3). Nevertheless, it turns out that in this case there exists a connection between and such that (3.25) holds in a slightly different sense. Only in this case this equation holds for analytic continuations of and .
Let us formulate precisely the corresponding theorem.
Definition 3.6.
Denote where . Note that , (see Fig. 3).
Theorem 3.7.
3.3 Step 3: Reduction to a difference equation
From (3.26), (3.23) it follows that and admit meromorphic continuations to and
(3.27) |
We will use the following automorphisms on :
(3.28) |
which are symmetries with respect to and , respectively.
Sometimes we will use the notation .
The functions and are automorphic functions with respect to and , respectively:
(3.29) | ||||
(3.30) |
as follows from the fact that depend only on and hence their liftings to satisfies (3.29), (3.30) since satisfy (3.29) and satisfies (3.30).
Thanks to this automorphy we can eliminate one unknown function in the undetermined equation (3.27) and reduce it to an equation with a shift, see [14]. The idea of this method is due to Malyshev [18].
Lemma 3.9.
The proof of this lemma is given in Appendix 10.1.
Now we reduce system (3.31)-(3.32) to a difference equation, which is also called the shift equation. This reduction is the part of MAF which was introduced in [18] for difference equations in angles. It uses the automorphy of on under the automorphisms and the term MAF is due to this observation.
Define, for ,
(3.33) |
For a region in we will denote here and everywhere below the set of meromorphic functions on .
Lemma 3.10.
The proof of this lemma is given in Appendix 10.2.
3.4 Necessary and sufficient condition for
The analyticity condition (3.35), which follows from the connection equation (3.26), imposes certain necessary conditions for the poles of the function , more exactly, of its continuation obtained in Lemma 3.9. This section is devoted to the derivation of these conditions and to the proof of the fact that they are also sufficient for (3.26) to hold.
We will often use the following evident statement.
Lemma 3.11.
Let , and satisfy (3.32). Then

Denote
(3.36) |
(See Fig. 4, where the positions of the curves correspond to the case . We will always assume in the following that this is the case; in the case the construction is similar.) Introduce the following two important parameters:
(3.37) |
In the next proposition we give a necessary and sufficient condition for guaranteeing that condition (3.33) holds.
Proof. First, let us prove the necessity of condition (3.39). By (3.24) the poles of in are
Of all of these poles only belong to (see Fig. 4).
Hence formulas (3.39) follow from (3.35), (3.37). Further, since by (3.35), by (3.32). Hence (3.38) also holds.
Let us prove the sufficiency of conditions (3.38), (3.39). From (3.39) and (3.37) it follows that
Hence, since by (3.38) and belongs to this space too. The proposition is proven.
Remark 3.13.
Condition (3.38) implies .
4 -automorphic solution of difference equation (3.34),
In this section we construct an -automorphic solution of difference equation (3.34) satisfying all conditions of Proposition 3.12 for . This limitation is related to the method of obtaining a solution which uses the Cauchy-type integral. The kernel of this integral must be analytic on the integration contour. In turns out that it is possible to find such a kernel only when . Fortunately, the case does not need an integral of the Cauchy-type since the difference equation (3.34) is solved by elementary methods in this case (see Section 5).
4.1 Poles of and asymptotics.
In this subsection we give the properties of which are necessary for the solution of the main problem. Let
(4.1) |
Obviously, the poles of belong to by (3.33).
Denote
Lemma 4.1.
i) The poles of belonging to are
.
The residues of at these points are
ii) The poles of belonging to are for , for and
iii) The function admits the asymptotics
(4.2) |
uniformly with respect to .
iv)
(4.3) |
Proof. The first three assertions follow directly from (3.33). The last assertion (4.3) follows from the fact that .
Remark 4.2.
In the case ,
However, this will not affect the final results.
4.2 Reduction of problem (3.34), (3.32) to a conjugate problem
Denote
where
We will look for a solution of the following problem: to find an analytic function in whose boundary values on ,
exist and are such that they satisfy the following conditions of conjugation
(4.4) | ||||
(4.5) |
see Fig. 5

4.3 Solution of the conjugate problem (4.4), (4.5) for
We start solving the problem (4.4), (4.5). We will reduce this problem to a Riemann-Hilbert problem. To this end we map conformally onto , where is the Riemann sphere, . For example, define
(4.6) |
Denote the inverse transform to .
Note that when tends to implies that
tends to from above, and when
tends to implies that to from below.
Obviously,
(see Fig.6).

Then problem (4.4), (4.5) is equivalent to the Riemann-Hilbert problem for , which at the same time is the saltus problem (see [17, Ch. 16, 18])
(4.7) |
Here , ; and for , , . From (4.6) and (4.3) it follows that and are continuous on and
(4.8) |
It is well known that a particular solution of (4.7) is given by the Cauchy type integral
(4.9) |
Obviously , and . Moreover, by (4.8) there exists .
In the following lemma we establish an asymptotics of (4.9)
at ; it plays an important role in describing the fact that the
solution belongs to a certain class and hence its uniqueness.
Lemma 4.3.
The function admits the following asymptotics
(4.10) |
where, by , we understand a certain branch that is single-valued on a plane cut along and depends only on ; moreover,
(4.11) |
where , depend only on .
Proof. (4.10) follows from (4.9) (see [29, §16]). Let us find the asymptotics of . (4.9) gives
Represent in the form , where . This is possible since by (4.1) if . Then
Obviously where depends only on ; by (4.8). Further, This implies (4.11).
Now we are able to find a solution to problem (3.34). First we define this solution in and then we extend it to .
Let us define by the formula
(4.12) |
where is given by (4.9) and
(4.13) |
with taken from (4.10).
Obviously satisfies (4.4) and (4.5), since a constant is a solution of the homogeneous equation (4.4) and satisfies (4.5).
Moreover, the same formula (4.12) defines the analytic function in , since satisfies (4.5). Obviously, (4.6) implies that
(4.14) |
Moreover, satisfies (4.4). In fact, for , this follows from (4.7), and for , this follows from (4.7) and (4.3). Further, we extend to by formulas corresponding to the difference equation (3.34):
(4.15) |
and
(4.16) |
This extension is meromorphic by (4.4) and still has property (4.14). Let us prove this. Let (see Fig. 7, 8). Then . By (4.16) and (4.15) we have . But since and by (4.3). Hence in this case. Similarly this is true for .
Proposition 4.5.
For
i) There exists a meromorphic in and analytic in solution of problem (3.34), (3.32) given by (4.15), (4.16).
ii) The function admits the following asymptotics
(4.17) | ||||
(4.18) |
uniformly with respect to .
iii) The solution has poles in for only at and with residues
(4.19) |
For , has poles in only at , and with residues
(4.20) |
Proof. Statement is proved above. The asymptotics (4.17), (4.18) are proved in Appendix 10.5.
Statement follow from the difference equation (3.34), -automorphicity of , Lemma 3.11, Lemma 4.1,
since the function is analytic in by i), see Fig. 7, 8.
As we will see below this asymptotics coincides with the asymptotics of the function in the case . In the following Lemma we describe the poles of the particular meromorphic solution to problem (3.34), (3.32) constructed above.


4.4 Solution of difference equation, case
We want to modify into which will satisfy all the conditions of Proposition 3.12. To this end for we add first to removing the poles and (see (4.19)), since by this proposition must be analytic at these points. It turns out that it is possible to construct in such a way that it produces the pole with the desired residue, as the same proposition requires.
Second, we add producing the pole with the desired residue according to the same proposition.
Consider
(4.21) |
where is given by (3.39), and is defined in Lemma 5.14.
It is easy to see that the function satisfies the following conditions:
(4.22) |
The poles of belonging to are
(4.23) |
(see Fig. 8), and
(4.24) |
Further, we define
(4.25) |
where is defined by (3.37).
Obviously also satisfies (4.22).
The poles of in are only
(4.26) |
Finally, we define
(4.27) |
where is given in Proposition 4.5.
Theorem 4.6.
Let .
i) The function satisfies all the hypothesis of Proposition 3.12.
ii) and it has a unique pole on with residue
(4.28) |
iii) and it has a unique pole at for ,
(4.29) |
Proof. i) satisfies (3.34) and (3.32) by Proposition 4.5, (4.27) and (4.22).
Let us prove (3.38) and (3.39).
Consider . By Proposition 4.5, (4.24), (4.23) and (4.26), the possible poles of in belong to , where is given by (3.36).
Moreover, by (4.19), (4.24) and (4.26). Hence, satisfies (3.38). Moreover, by (4.24),
, . Thus, (3.39) is proven for
, and satisfies all the hypothesis of Proposition 3.12 in this case (see Fig. 7).
Consider . In this case all the poles of in belong to by Proposition 4.5, (4.27). Hence, (3.38) holds. Moreover, from (4.20), (4.26)
Hence, by (3.32).
The equality follows from (4.27), (4.26) and the analyticity of in . Thus satisfies (3.39) too (see Fig. 8).
ii) By Proposition 4.5, which implies . By (4.23), has poles in only at , , , . For none of these poles belong to . has a pole in only in , by (4.25). Further, , and hence holds by (4.27), (4.26).
iii) Consider . By Proposition 4.5, has poles here at for and at for with residues (4.19) and (4.20).
From (4.21) and (4.25) it follows that does not have poles at and has a unique pole at here only for and .
Hence has a pole at for and a possible pole at for .
From (4.27) we obtain
Similarly
Therefore iii) and hence Theorem 4.6 are proven.
Corollary 4.7.
For a unique pole of belonging to is and
(4.30) |
Proof. A unique pole of in is only by Theorem 4.6 ii) with residue (4.28). Hence, has a unique pole at in and (4.30) follows from (3.35), (4.28) and (3.37).
The function is analytic in by Proposition 3.12, since satisfies all the hypothesis of this Proposition by Theorem 4.6.
It remains only to prove that is analytic in . By Theorem 4.6 the function has a unique pole at in with residue (4.29) and the function also has a unique pole at this point with residue by (3.37). Hence, the function is analytic in for and the Corollary is proven.
5 -invariant solution of the difference equation in the case
In the previous sections we have constructed a solution to problem (3.34), (3.32), satisfying all the conditions of Proposition 3.12 for .
It is possible to construct a solution for using the same method. A slight technical inconvenience in this case arises from the fact that the function has a pole on . Nevertheless, one can obtain a solution with the properties indicated in Theorem 4.6.
However, we prefer to find a solution of the problem in the case by another method.
The point is that in this case it is easy to find a solution of the difference equation (3.34) in an explicit form without using the Cauchy-type integral.
Using the Liouville theorem it is easy to show that this elementary solution coincides with the solution obtained by the Cauchy-type integral.
In this section we give a meromorphic -invariant solution of (3.34).
5.1 Meromorphic solution of the difference equation for
In this case the construction of a meromorphic solution of difference equation (3.34) is simpler than in the case and is expressed through elementary functions. By (3.33), for , we have
(5.1) |
Let us solve difference equation (3.34) in this case. First, we solve (3.34) in the class of meromorphic functions. It is easy to guess a solution, using the -periodicity of . Let us define
Then, by (5.1), satisfies (3.34).
Of course, this solution is not unique. All the other solutions differ from it by a -periodic function. Similarly to the case , we will modify this solution in such a way that it will satisfy all the conditions of Proposition 3.12.
Function (5.1) is not automorphic with respect to . Let us symmetrize it.
Define
(5.2) |
Then
(5.3) |
Lemma 5.1.
Proof. i) The assertion follows from a direct substitution of (5.3) into (3.34), and (3.32) follows from (5.2).
ii) The zeros of are
where is defined by (3.36). Obviously, only the poles from belong to , see Fig. 9.
Formulas (5.5) follow from (5.3) and Lemma 3.11.
Now we modify the function in such a way that it will satisfy the conditions (3.38), (3.39) of Proposition 3.12.
To this end we add to an appropriate -periodic function.
Since for , , conditions (3.38), (3.39) take the form
where is given by (3.36), and

5.2 Solution of the difference equation for
By (5.4), the function has 8 poles in belonging to with residues (5.5).
We modify so that (3.38) and (3.39) hold.
To this end we first add to functions and which “correct” the residues at the point and at the point by . Then we add which anihilates the poles and .
So, consider the following functions (the periodic supplements)
(5.7) | ||||
(5.8) | ||||
(5.9) |
where are given by (5.6). Obviously, the functions are -periodic, and -automorphic:
(5.10) |
Finally, define
(5.11) |
From (5.7)-(5.11), it follows directly that the set of the poles of in is given by (5.4) (see Fig. 9) and
(5.12) |
Theorem 5.2.
i) For the function satisfies all the conditions of Proposition 3.12.
ii) The poles of in are
(5.14) |
with the following residues
(5.15) |
Proof. i) Equations (3.34) and (3.32) follow from Lemma 5.1 and (5.13).
From (5.5), (5.6) and (5.12) we obtain that (3.39) holds.
Let us prove (3.38). Since all the poles of in belong to by Lemma 5.1, it suffices to prove that is analytic in .
From (5.5), (5.12), (3.32) and Lemma 3.11 we obtain
Thus (3.38) and i) are proven.
ii) By (5.4) and (5.11) the poles of and in are
From (5.5), (5.6), (5.12) and (5.13) it follows that
does not have poles at and has poles (5.14) with residues (5.15).
Statement ii) is also proven.
Now we establish an important property of the function similar to Corollary 4.7 for the case .
We recall that this function plays the crucial role in the construction of the Sommerfeld-type representation for the solution of the main problem. This representation will be given in the following section.
Corollary 5.3.
For the function given by (3.35) has a unique pole at belonging to , and
Proof. The function by Proposition 3.12 and Theorem 5.2. It suffices to analyze .
First, consider . By Theorem 5.2 ii), (3.24), and -periodicity of , unique poles of and in are and , and the unique pole in of the same function is with residue (5.15) and (3.37). Hence the statement follows from (3.35).
6 Asymptotics of at infinity
We will need to prove (2.2). For this we have to find the asymptotics of the integrand at infinity.
6.1 Asymptotics of at infinity
Lemma 6.1.
For any the function admits the following asymptotics:
(6.1) |
Proof. From (4.21) it follows that admits the following asymptotics
Similarly, admits the same asymptotics by (4.25) and, by (4.27), (4.17), satisfies (6.1) in the case .
Consider the case . From (5.13), (5.3), (5.11) it follows that the asymptotics (6.1) holds in this case too.
Similarly, differentiating (5.13) we obtain (6.1)
Remark 6.2.
The asymptotics of coincide for the cases and .
6.2 Asymptotics of
7 Sommerfeld-type representation of solution to problem (1.9)
In this section we give a Sommerfeld-type representation of solution to problem (1.9). This representation was obtained by A. Sommerfeld and it is widely used in Mathematical Diffraction Theory [35]. This representation is an integral with a specially chosen integrand along a Sommerfeld-type contour. This contour has double-loop form as in Fig. 11.
We define first this curvilinear contour depending on (in contrast to the Sommerfeld contour), and then we reduce it to the rectilinear contour which coincides with the Sommerfeld contour (Fig. 11)).
Define , where
is the segment of the line lying between and , and (see Fig. 10).
In our case the integrand is the Sommerfeld exponential multiplied by a kernel which was constructed in the previous sections.
Proof of the main Theorem 2.2.
First, we consider the Sommerfeld integral with the contour . We write the integral (2.3) with instead of . We keep the notation for this integral because we will see later that these two integrals coincide.
So, let
(7.1) |
where is defined above. Here and in the following we will use the following estimate: for , with , and
(7.2) |
where . The proof of this estimate is given in Appendix 10.3.
Hence the integral (7.1) converges by the asymptotics (6.2), (3.23), and (3.24), since does not have poles on by Corollaries 4.7 and 5.3 (see Fig. 10, where the exponential decreases superexponentially in the shaded regions).
Let us prove that satisfies the first equation of (1.9). To this end we rewrite (7.1) as
Let us fix . By the Cauchy Theorem
for any sufficiently close to .
Now the differentiation in under the sign of the integral is possible and the first equation in (1.9) follows from the formula .
Finally, boundary conditions (3.1b) and (3.1c) are proved in the next section. The integral (7.1) is transformed into the integral (2.3) over the contour , which no longer depends on (see Fig.11).


8 Proof of the boundary conditions
8.1 Decomposition of the solution into a plane wave and a wave dispersed by the vertex
In this section we decompose the solution of problem (1.9) given by (7.1) into two parts: the first part is the plane wave generated by the first boundary condition (1.9) and the second part is the wave dispersed by the edge of the wedge.

To give this decomposition we recall that a unique pole of lying in is
(8.1) |
as follows from Corollaries 4.7 and 5.3.
Define a plane wave generated by the first boundary condition in (1.9) ;
(8.2) |
where , is given by (3.36), and a “diffracted” wave
The integrand here coincides with the integrand in (2.3), but the contour of integration differs from (see Fig. 12).
It turns out that the solution is decomposed into the sum of and and the corresponding decomposition is more convenient for the proof of boundary conditions.
Proof. By the Cauchy Theorem, defined by (7.1) admits the representation
(8.4) |
where is the contour bounded by , and , .
Let us find the poles of for any , where is the region bounded by .
Let be a pole of . Then is a pole of belonging to .
The function has a unique pole in with the residue (8.1).
Hence, . Obviously
only for .
Calculating the second integral in (8.4) with the help of residues, we obtain (8.3) for .
Therefore, (8.3) holds.
8.2 Boundary values of the solution
We continue to prove the main theorem.
Proof. The fact that satisfies the Helmholtz equation in (1.9) has been proven in Section 6. Let us prove the boundary conditions in (1.9). First, we prove the first condition (1.9) which in the polar coordinates takes the form
Since by (8.2), (3.36), it suffices to prove that satisfies the homogeneous conditions. From (3.35) we have
since is -periodic, the integral of the second summand is equal to because . Thus, it suffices to prove that
(8.5) |
Making the change of the variable
we obtain (8.5) since (after the change) the integrand is an -automorphic function and
Let us prove the second boundary condition in (1.9), (8.3):
From (3.26), (3.22) (see Remark 3.8) we have
Making the change of the variable , we obtain
since is an -automorphic function and
9 The solution belongs to the functional class and is unique
9.1 Behavior at infinity
Lemma 9.1.
The solution is an -function in , bounded in with all its first derivatives.
9.2 Asymptotics of the solution at the origin
We continue to prove the main theorem, namely we prove that given by (2.3) belongs to . It remains only to prove the asymptotics (2.2). Let us prove the first asymptotics in (2.2). Represent the contour as
where
and the contours are shown in Fig. 11. Note that the “finite” part of the integral (2.3), has a “good” asymptotics, since
(9.1) |
by (6.2).
Thus it suffices only to find the principal term of the asymptotics of the “infinite” part of . Since is a -periodic function and is even on , we have
and
(9.2) |
and
where
It suffices to find the asymptotics for , since the asymptotics of is similar.
By (6.3),
(9.3) |
We have
(9.4) |
similarly to (9.1). We need the folowing simple statement whose proof is given in Appendix 10.6.
Lemma 9.2.
Let ,
Then
(9.5) |
By (6.2), noting that , using Lemma 9.2 and the arguments of the proof of (9.3), we obtain
Hence, using (9.4),
Consider
(9.6) |
Similarly to (9.4) and using (6.1), we have
(9.7) |
Similarly to (9.3), we have
Obviously,
Hence, substituting the asymptotics (6.1) for into (9.2), we obtain from (9.6), (9.7) that
Thus,
Similar asymptotics for holds and the second asymptotics of (2.2) is proven.
9.3 Uniqueness
In this section we prove Statement of Theorem 2.2.
Obviously, it suffices to prove the uniqueness of solution of problem (3.1a)-(3.1b) in the same space .
Let , be two solutions of problem (3.1a)-(3.1c) belonging to the space , and , be their Cauchy data (). Then are -automorphic solutions of the difference equation (3.34) and they have the same poles and residues in by Proposition 3.12. Hence, their difference is an analytic solution of the homogeneous equation (3.34), that is, an entire periodic function on .
Moreover, since and admit the same asymptotics (2.2), also admits the asymptotics (2.2). Hence its F-L transform satisfies and hence,
(9.8) |
Since is a periodic function with period , the asymptotics (9.8) holds for . This implies that
(9.9) |
In fact, let us apply the conformal mapping given by (4.6).
It is easy to show that
and it admits the asymptotics Hence it has a removable singularity at . By the Liouville Theorem, this implies (9.9). Therefore, and hence by (3.27) This implies that and Since and this means that the Dirichlet and Neumann data of are zeros. Hence,
by the uniqueness of solution of elliptic equations.

10 Appendices
10.1 Proof of the algebraic equation
Since is meromorphic in by (3.27) and is automorphic with respect to , we extend by symmetry with respect to to . Namely, define
Obviously, is meromorphic on since is meromorphic on . We will still use the notation for . Thus (3.32) holds for this extension too.
Since is meromorphic in , by (3.27) admits a meromorphic continuation onto which we also denote . Hence
(10.1) |
by the uniqueness of analytic continuation.
Let us extend , to . Similarly to is a meromorphic function in . Now we extend (10.1) to and obtain
Hence
Continuing the process of extension of and by symmetries (3.28), (3.29), we can extend equation (10.1) to and obtain (3.31)-(3.32).
10.2 Proof of Lemma 3.10
Let us apply the automorphism to equation (3.31). Since by (3.23), (3.30),
(3.27) gives
(10.2) |
Hence, subtracting equation (10.2) from equation (3.27), we obtain
where is given by (3.33).
Using (3.29), we can represent as a function with shifted argument. Applying (3.28) we have
where is a shift since Hence, by (3.33),
10.3 Proof of the estimate (7.2)
10.4 Analysis of the solution near the ray
We prove that is continuous in for small . By (8.3) it suffices to prove that
(10.3) |
satisfies
(10.4) |
since does not have poles on by (8.1).
Making the change of the variable in (10.3), using (8.1), and the Sokhotski-Plemelj Theorem, we obtain
where here and in the following denotes the principal value of the integral (10.3).
Similarly,
10.5 Proof of asymptotics (4.17)
10.6 Proof of Lemma 9.2
Proof. Obviously, the principal term admits the asymptotics
Making the change of the variable , we obtain
Hence (9.5) follows.
11 Conclusion
As is known, an angle is one of the few regions where the boundary value problems for the Helmholtz equation admit an explicit solution. As far as we know, this has always been done for decreasing boundary data, where the operator methods are normally used with the exception of a very specific boundary value problem associated with the plane incident wave (Sommerfeld’s diffraction problem [35]). In the presented work, we solve the Dirichlet boundary problem not related to the incidence of a plane wave and we obtain an explicit solution in the form of the Sommerfeld integral. The proposed method is suitable for the Neumann (NN) and Dirichlet-Neumann (DN) boundary conditions, and for angles less then . We hope that the method is suitable for solving such problems with a real wave number in the Helmholtz operator and also for nonstationary problems.
References
- [1] V.M.Babich, M.A.Lyalinov,V.E.Gricurov, The Sommerfeld-Malyuzhinets Tecnique in Diffraction Theory. Alpha Science International, Oxford, 2007
- [2] L. P. Castro, D. Kapanadze, Dirichlet indexDirichlet-Neumann impedance boundary-value problems arising in rectangular wedge diffraction problems, Proc. Am. Math. Soc. 136 (2008), 2113-2123.
- [3] L. P. Castro, D. Kapanadze, Wave diffraction by a 45 degree wedge sector with Dirichlet and Neumann boundary conditions , Mathematical and Computer Modelling 48 (2008), no. 1/2, 114–121.
- [4] L.P. Castro, D. Kapanadze, Wave diffraction by a 270 degrees wedge sector with Dirichlet , Neumann and impedance boundary conditions, Proc. A. Razmadze Math. Inst. 155 (2011), 96–99.
- [5] L. P. Castro, F.-O. Speck F. S. Teixeira, On a class of wedge diffraction problems posted by Erhard Meister, Oper. Theory Adv. Appl. 147 (2004), 213–240.
- [6] L. P. Castro, F.-O. Speck, F. S. Teixeira, Mixed boundary value problems for the Helmholtz equation in a quadrant, Integral Equations Operator Theory 56 (2006), 1–44.
- [7] A.F. Dos Santos and F.S. Teixeira. The Sommerfeld Problem revisited: Solution spaces and the edge condition. Math Anal Appl 143, 341-357 (1989).
- [8] Esquivel Navarrete A, Merzon AE. An explicit formula for the nonstationary diffracted wave scattered on a NN-wedge. Acta Applicandae Mathematicae 2014; 136(1):119–145. DOI:10.1007/s10440-014-9943-7.
- [9] Bernard JML, Pelosi G, Manara G, Freni A. Time domains scattering by an impedance wedge for skew incidence. Proceeding conference ICEAA 1991: 11-14.
- [10] I. Kay. The diffraction of an arbitrary pulse by a wedge, Comm. Pure Appl. Math. 6 (1953), 521–546.
- [11] Keller J, Blank A. Diffraction and reflection of pulses by wedges and corners. Communications on Pure and Applied Mathematics 1951;4(1):75–95.
- [12] A.I. Komech. Elliptic boundary value problems on manifolds with piecewise smooth boundary. Math. USSR Sbornik 21(1), 91-135(1973).
- [13] A. I. Komech, A. E. Merzon, A. Esquivel Navarrete, J. E. De La Paz Méndez, T. J. Villalba Vega, Sommerfeld s solution as the limiting amplitude and asymptotics for narrow wedgesMath. Methods in Appl. Sci. (2018) 1? 14. https://doi.org/10.1002/mma.5075.
- [14] A.I. Komech, N.J. Mauser, A.E. Merzon. On Sommerfeld representation and uniqueness in scattering by wedges. Math meth Appl Sci. 2004; 28:147-183. https://doi.org/10.1002/mma.553
- [15] A. I. Komech, A. E. Merzon, Limiting amplitude principle in the scattering by wedges , Mathematical Methods in the Applied Sciences 29 (2006), no. 10, 1147–1185.
- [16] A. I. Komech, A. E. Merzon On uniqueness and stability of Sobolev’s solution in scattering by wedges, J. Appl. Math. Phys. (ZAMP) 66 (2015), no. 5, 2485–2498.
- [17] A. Komech, A. Merzon. Stationary Diffraction by Wedges, Springer, Lecture Notes in Mathematics 2249, Springer (2019).
- [18] V.A.Malyshev, Random Walks, Wiener-Hopf Equations in the Quadrant of Plane, Galois Automorphisms, Moscow University, Moscow, 1970 (in Russian).
- [19] Merzon AE, De la Paz Méndez JE. DN-scattering of a plane wave by wedges. Mathematical Methods in the Applied Sciences 2011;34(15):1843-1872.
- [20] Merzon AE, De La Paz Méndez JE, Villalba Vega TJ. On the Keller-Blank solution to the scattering problem of pulses by wedges. Mathematical Methods in the Applied Sciences 2014. DOI: 10.1002/mma.3202.
- [21] A. I. Komech, A. E. Merzon, J. E. De La Paz Méndez, Time-dependent scattering of generalized plane waves by wedges , Math. Methods in Appl. Sciences 38 (2015), no. 18, 4774–4785.
- [22] A. E. Merzon, A. I. Komech, J. E. De la Paz Mendez, T. J. Villalba, On the Keller-Blank solution to the scattering problem of pulses by wedges , Math. Methods Appl. Sci. 38 (2015), no.10, 2035–2040.
- [23] A. E. Merzon, P. N. Zhevandrov, J. E. De La Paz Mendez, On the behavior of the edge diffracted nonstationary wave in scattering by wedges near the front. Russian Journal of Mathematical Physics 22 (2015), no. 4, 491–503.
- [24] E. Meister, Some solved and unsolved canonical problems of diffraction theory, pp 320–336 in: Lecture Notes in Math. 1285, 1987.
- [25] E. Meister, A. Passow, K. Rottbrand, New results on wave diffraction by canonical obstacles, Operator Theory: Adv. Appl. 110 (1999), 235–256.
- [26] E. Meister, F. Penzel, F. O. Speck, F. S. Teixeira, Some interior and exterior boundary-value problems for the Helmholtz equations in a quadrant, Proc. Roy. Soc. Edinburgh Sect. A 123 (1993), no. 2, 275–294.
- [27] E. Meister, F. Penzel, F.-O. Speck, F. S. Teixeira, Two canonical wedge problems for the Helmholtz equation, Math. Methods Appl. Sci. 17 (1994), 877–899.
- [28] E. Meister, F. O. Speck, F.S. Teixeira, Wiener-Hopf-Hankel operators for some wedge diffraction problems with mixed boundary conditions, J. Integral. Equat. and Applications 4 (1992), no. 2, 229–255.
- [29] N.I. Muskhelishvili, Singular integral equations. Springer. 1958.
- [30] Oberhettinger F. On the diffraction and reflection of waves and pulses by wedges and corners. Journal of Research National Bureau of Standarts. 1958; 61(2):343–365.
- [31] F. Penzel, F.S. Teixeira, The Helmholtz equation in a quadrant with Robin’s conditions, Math. Methods Appl. Sci. 22 (1999), 201–216.
- [32] Sobolev SL. General theory of diffraction of waves on Riemann surfaces. In Selected Works of S.L. Sobolev, Vol. I, Springer:New York, 2006;201–262.
- [33] Sobolev SL. Some questions in the theory of propagations of oscillations, Chap XII. In Differential and Integral Equations of Mathematical Physics, Frank F, Mizes P(eds). Leningrad:Moscow,1937;468-617.[Russian].
- [34] Sobolev SL. Theory of diffraction of plane waves. Proceedings of Seismological Institute, Russian Academy of Science, Leningrad 1934; 41:605–617.
-
[35]
A. Sommerfeld. Mathematical Theory of Diffraction (Progress in Mathematical Physics, 35) R.J. Nagem, M.Zampolli, G. Sandri (translators).
Birkhäuser, 2004. - [36] F.S. Teixeira. Diffraction by a rectangular wedge: Wiener-Hopf-Hankel formulation. Integral Equations Operator Theory. 14 (1991) p. 436-454.
- [37] A.N. Tikhonov, A.A. Samarskii. Equations of mathematical physics. Dover publications, 1990.
- [38] Methods of Modern Mathematical Physics II: Fourier Analysis, Self-Adjointness (Academic, New York, 1975).
- [39] Rottbrand K. Exact solution for time-dependent diffraction of plane waves by semi-infinite soft/hard wedges and half-planes, 1998. Preprint 1984 Technical University Darmstadt.
- [40] Rottbrand K. Time-dependent plane wave diffraction by a half-plane: explicit solution for Rawlins’ mixed initial boundary value problem. Zeitschrift für Angewandte Mathematik und Mechanik 1998; 78(5): 321-335.