Mixed Riemann-Hilbert boundary value problem
with simply connected fibers
Abstract
We study the existence of solutions of mixed Riemann-Hilbert or Cherepanov boundary value problem with simply connected fibers on the unit disk . Let be a closed arc on with the end points and let be a smooth function on with no zeros. Let be a smooth family of smooth Jordan curves in which all contain point in their interiors and such that , are strongly starshaped with respect to . Then under condition that for each the angle between and the normal to at is less than , there exists a Hölder continuous function on , holomorphic on , such that
keywords:
Boundary value problem, mixed Riemann-Hilbert problem, Cherepanov problemMSC:
[2020] 35Q15, 30E25organization=Faculty of Mathematics and Physics, University of Ljubljana, and Institute of Mathematics, Physics and Mechanics, addressline=Jadranska 19, city=Ljubljana, postcode=1 111, country=Slovenia
1 Introduction
Let be the open unit disc in the complex plane and let be the unit circle. Let be a closed arc on , let denote its interior with respect to , and let be a smooth function.
Recall that the interior of a Jordan curve is the bounded component of . We orient positively with respect to . Jordan curve is starshaped with respect to , if for any point in the interior of the line segment which connects points and lies in the interior of , and it is strongly starshaped with respect to , [9], if there exists a positive continuous function on the unit circle such that
(1) |
and
(2) |
Let be a smooth family of smooth Jordan curves in which all contain point in their interiors. In this paper we study the existence and properties of holomorphic solutions of the nonlinear mixed Riemann-Hilbert problem, that is, the Cherepanov boundary value problem with simply connected fibers. The problem asks for a continuous function on , holomorphic on , such that
(3) |
and
(4) |
That is, solves a linear Riemann-Hilbert problem on and a nonlinear Riemann-Hilbert problem with simply connected fibers on . See also [1, 2, 13, 14].
The problem with circular fibers and a finite union of disjoint arcs was considered by Obnosov and Zulkarnyaev in [14], and by the author in [5]. The structure of the family of solutions of problem (3-4) is well known in the cases where either or . If , we consider a homogeneous linear Riemann-Hilbert problem. In this case the essential information on the problem is given by the winding number of function . It is well known [11, 17, 18] that if the winding number is nonnegative, the space of solutions of (3) is a vector subspace of , , of real dimension .
Remark 1.1
The linear Riemann-Hilbert problem can also be considered in the case of a nonorientable line bundle over , that is, in the case where at some point we have . Then the winding number of function or the Maslov index of the problem is an odd integer. In this case it holds that if , or, with a little bit of abuse of notation, if , then the space of solutions of (3) is a vector subspace of of real dimension , see [3, 4, 15, 18].
If is empty, we have a nonlinear Riemann-Hilbert problem with smooth simply connected fibers which all contain in their interiors. This problem was considered and solved in [8, 16, 17, 18]. In particular, it was proved that the family of solutions with exactly zeros on , , forms a manifold in space of dimension , and this manifold is compact if and only if . We assume from now on that neither nor .
Theorem 1.2
Let . Let be a function and let be a family of Jordan curves in which all contain point in their interiors. Let and be the first and the last point of arc with respect to the positive orientation of . Let Jordan curves , , be strongly starshaped with respect to and such that for each the angle between and the outer normal to at is less than . Let , , be the intersection of and the line of the form , , and let be the oriented angle of intersection of the line with the fiber at point , where and . Let
(5) |
Then there exists a unique with no zeros on which solves (3-4) for which and .
Remark 1.3
Here , if the tangent vector to is rotated counterclockwise by angle to get a positive tangent vector to at point , and , if a positive tangent vector to at is rotated clockwise by angle to get tangent vector to .
Remark 1.4
Observe that conditions in Theorem 1.2 imply , , and hence one could choose .
Remark 1.5
2 Function spaces, Hilbert transform and
defining functions
Let and let be a compact subset. We denote by the algebra over of Hölder continuous complex functions on and by the algebra over of real Hölder continuous functions on . Using the norm
(6) |
the algebras and become Banach algebras. For or and we also define spaces and of times continuously differentiable functions on , whose all -th derivatives belong to space or space .
We also need some algebras of holomorphic functions on . By we denote the disc algebra, that is, the algebra of continuous functions on which are holomorphic on , and by the algebra of Hölder continuous functions on the closed disc which are holomorphic on . Using appropriate norms, that is, the maximum norm for and the Hölder norm for , these algebras become Banach algebras. Similarly we define .
Recall that Hilbert transform assigns to a real function on a real function on such that the harmonic extension of to is holomorphic on and real at . It is known that is a bounded linear operator on , [18, §1.6.11], and hence the harmonic extension of to belongs to . Also, [18, §1.6.11], the Hilbert transform is a bounded linear operator on the Sobolev space of times generalized differentiable functions with derivatives in equipped with the norm
(7) |
Recall, [18, §1.6.14], that if is a partition of in two subarcs and and if is a compactly contained subarc of , then for , , there exists a constant such that
(8) |
and
(9) |
We will also need compact embedding result, [18, §1.1.8],
(10) |
for , , which holds on arcs in as well.
Since we can extend to as a nowhere zero function of class so that the winding number . Therefore, [18, p. 25], we can write in the form
(11) |
where is a positive function on and . Thus the original problem (3-4) is equivalent to the problem
(12) |
and
(13) |
where and . Observe that the number of zeros of and are the same and that belongs to the interiors of all curves , . Also, since for each the transfomation
(14) |
is a composition of a dilation and a rotation, the angle conditions from Theorem 1.2 stay the same.
Using a holomorphic automorphism of the unit disc we may even assume that is the lower semicircle. From now on we will consider problem (12-13) with the addition that is the lower semicircle and instead of and we will still write and .
Remark 2.1
One can also create the ’double’ of the boundary value problem. Using a biholomorphism one can replace the unit disc with the upper half-disk and by the interval .
By the reflection principle we see that problem (12-13) is equivalent to the nonlinear Riemann-Hilbert problem on , where the boundary curves are symmetrically extended and defined on the lower semicircle so that we have
(15) |
for every . In general this symmetrical extension of Jordan curves to the lower semicircle produces boundary data which are not continuous at points and . Because the biholomorphism from to the upper semidisc is in , we get that the regularity of solutions of (12-13) is in general a half of the regularity of solutions of the symmetrical Riemann-Hilbert problem.
We will consider smooth families of smooth Jordan curves in . Let . The family of Jordan curves is a family parametrized by if there exists a function such that
(16) |
and the gradient for every and . We call a defining function for family of Jordan curves . We will consider only bounded families of Jordan curves which all lie in some fixed disc , , and the space is equipped with the standard norm.
Since we assume that are strongly starshaped Jordan curves, we also assume that for , the defining function for Jordan curves , and we have
(17) |
for some positive functions on .
Using parametrization of the unit circle we will also use the notation , and instead of , and . Also, for a function on , we will write either or , where . Observe that if is holomorphic on with well defined derivative on , then for .
Remark 2.2
The reflection principle and the symmetric extension to the lower semicircle mentioned in Remark 2.1 is in terms of defining function given as
(18) |
for every and every .
3 Regularity of solutions
In this section we prove regularity of continuous solutions of a specific form of problem (12-13), where the defining function .
Let be a solution of (12-13). It is well known [6, 7, 8, 18] that restricted to is in for any . Hence we need information on the regularity of near points . For we denote .
Using Möbius tranformation from the unit disc to the upper half-plane we consider the case where is bounded and continuous on and holomorphic on . Also, point is mapped into and point into . Now solves the problem
(19) |
and
(20) |
Also, using translation, we will assume that .
Let be the oriented angle of intersection of the real axis and at . The orientation of the real axis is positive with respect to the upper half-plane and the orientation of is positive with respect to the interior of . Hence , if the tangent vector to the real axis is rotated counterclockwise by angle to get a tangent vector to at point , and , if the tangent vector to the real axis is rotated clockwise by angle to get a tangent vector to at .
The defining function can near for be written as
(21) |
(22) |
where such that .
Recall that and that represents an outer normal to at point . So we have
(23) |
for some real . We may assume .
Because
(24) |
we have
(25) |
(26) |
for some and .
Let us assume that we have a solution of the problem (19-20) of the form
(27) |
where is bounded and continuous on , holomorphic on , and to be determined.
For we have and from (19) we get
(28) |
On the other hand for we have
(29) |
(30) |
We choose so that solves boundary value problem with continuous boundary data. That is, we choose , if , and , if .
Thus solves the following Riemann-Hilbert problem
(31) |
and
(32) |
where, if ,
(33) |
(34) |
and, if ,
(35) |
(36) |
For such choice of are the defining function for problem (31-32)
(37) |
and its partial -derivative
(38) |
continuous on .
On the other hand, the partial derivative of defining function (37) with respect to the variable is not continuous at , but, as we will see, it still has certain regularity properties, which will imply regularity conditions on and .
We know that is on and we can differentiate (31-32) on to get
(39) |
and
(40) |
For and we have
(41) |
(42) |
(43) |
and for and we have
(44) |
(45) |
(46) |
The -derivative of defining function (37) is for .
Since and is bounded, we have that is in for
(47) |
A similar argument can be used for point . Let be the orientied angle of intersection of and the real axis at point . Now is positive, if a positive tangent vector to at is rotated counterclockwise to get a positive tangent vector to the real axis and negative otherwise. For we define , if , and , if .
To transfer our observations to the boundary value problem (12-13) on the unit disc, let be a biholomorphic map from to the upper half-disc , which maps the lower semicircle on so that . Let . Then and . Hence function is real on , , and on .
Recall that is the positive intersection of and the real axis, . Now we consider only those solutions of the Cherepanov problem (12-13), which are of the form
(48) |
where is in .
We will define two (local) defining functions for and for . Let
(49) |
Then , and map the upper semicircle to the positive real axis and the lower semicircle to the negative real axis. For and we define
(50) |
and for we set
(51) |
As before one can check that and are continuous on , . Since solves the original boundary value problem, we have that , .
Let be a smooth function such that for , and for , .
We define a new (global) defining function as . Then and are well defined continuous function on . If have the same sign, then both function are also continuous at , but if have the opposite signs, then
(52) |
which means that we have a nonorientable bundle as the boundary value data for .
Now locally considered problem (39-40) for and becomes global boundary value problem for and . Hence solves the linear Riemann-Hilbert problem
(53) |
where is either a nonzero continuous function on or
(54) |
and belongs to the appropriate space
(55) |
Remark 3.1
In fact belongs to for
(56) |
near and to near for
(57) |
Let be the winding number of function , that is, is the Maslov index of the associated linear Riemann-Hilbert problem. If is a continuous function on , Maslov index is an even integer and hence is an integer. On the other hand, if , Maslov index is an odd integer and is a half of an odd integer.
Let be the square root function, where we take the branch where is cut along the negative imaginary axis. Then function can be written in the form
(58) |
where and are real continuous functions on , [18, p. 25]. In the case , , is a half of an odd integer, we define , which corresponds to the sign changing of at . See also [3, 4, 15]. Hence belongs to for any , [18, p. 23] and thus
(59) |
belongs to for any .
Therefore
(60) |
We conclude that the right-hand side belongs to the same space as function . Since Hilbert transform is bounded in spaces, , [18, p. 23], we get that is in for the same set (55) of values of as function . Therefore belongs to for all such values of and this implies that , [18, p. 10], where
(61) |
Remark 3.2
Observe that regularity of and could also be expressed locally, that is, near functions and belong to Hölder space , where .
Proposition 3.3
Remark 3.4
Observe that in cases where , the regularity conditions we get for solutions of the Cherepanov/mixed Riemann-Hilbert problem (3-4) are consistent with results on the regularity of Riemann maps from the unit disc into simply connected domains bounded by Jordan curves which satisfy so called wedge condition, [12]. If the defining function is independent of and , we get -regularity. The -regularity comes from -dependence.
Similarly, the expected regularity and the ’order’ of zeros of Riemann maps in the cases where and which are independent, would be , but -dependence of the defining function changes regularity conditions.
On the other hand, results in [10] show that in the case of nontransversal intersection of the real axis with either or solutions might not be of the form or for some function .
4 Linear Cherepanov boundary value problem
In this section we consider the linear version of problem (12-13), that is, a linear Riemann-Hilbert problem with piecewise continuous boundary data, [19, p. 169], and the lower semicircle. First we consider homogeneous linear problem with piecewise continuous boundary data
(64) |
and
(65) |
where is a complex nonzero function of class on the upper semicircle. The regularity exponent is bounded by conditions given in Proposition 3.3. We may assume without loss of generality that for all .
Let , , be the oriented angle of intersection of the real axis and at point , that is, . Similarly, let , , be the oriented angle of intersection of and the real axis at point , that is, .
We search for solutions of (64-65) of the form for some . Recall that for we defined , if , and , if . Hence we also have .
To define noninteger powers of and we take appropriate branches of the complex logarithm. For the complex plane is cut along positive real numbers so that the argument of for lies on interval , and for the complex plane is cut along negative real numbers and the argument of for lies on interval .
An argument similar to the argument in Section 3 shows that solves homogeneous linear Riemann-Hilbert problem
(66) |
where is defined as
(67) |
with the left and the right limits at . The sign for is chosen so that is continuous at , that is, we have plus sign, if , and minus sign, if . At point function might not be continuous. In general we have . See [19, p. 169-170] for more.
Each factor
(68) |
changes the argument by when passes once in the positive direction. Hence possible widing number of is either an integer (Maslov index of problem (66) is even) or a half of an odd integer (Maslov index of problem (66) is odd).
Example 4.1
Consider the case for . In particular we have . Then we get
(69) |
Hence the winding number . Using identification of the boundary problem (64-65) with the problem on the unit disc with reflected boundary conditions (15), this example corresponds to the linearization of the boundary value problem, where all boundary curves are unit circles and we linearize at . The family of (nearby) solutions which are real on the real axis is one-dimensional , where is a real number.
Example 4.2
Consider the case for . In particular we have . Then we get
(70) |
and the winding number . Using identification of the boundary problem (64-65) with the problem on the unit disc with reflected boundary conditions (15), this example corresponds to the linearization of the problem where all boundary curves are unit circles and we linearize at function . The family of (nearby) solutions which are real on the real axis is zero-dimensional.
The dimension of the space of solutions in depends on the winding number of function . It equals if , see [18, p. 25, p. 59] and [3, 4, 15]. We define the winding number of as the winding number of .
Now we can solve appropriate nonhomogeneous linear Riemann-Hilbert problem with piecewise continuous boundary data
(71) |
and
(72) |
where is as above and a real function on of the form
(73) |
for some function which equals on .
To solve (71-72) in the space of functions of the form for some is equivalent to solve the problem
(74) |
It is well known that if , then the equation is solvable for any , see [18, p. 25, p. 59] and [3, 4, 15].
Remark 4.3
If the winding number is an odd integer, the function on the right-hand side of (74) needs to belong to a special space of Hölder continuous real functions on of the form , where is the principal branch of the square root and is an odd function. Hence we need condition , which is satisfied because in our case we have . See [3, 4] for more information.
Proposition 4.4
Let . Let be a non-vanishing complex function in and let . Then for every real function on of the form
(75) |
for some which equals on , there exists a solution of the linear Cherepanov problem
(76) |
and
(77) |
of the form
(78) |
where . Moreover, the space of solutions of this form is dimensional real subspace of .
Proposition 4.5
Remark 4.6
Let be an open subset of defining functions of the families of Jordan curves over such that the intersection of the corresponding and with the real axis at some points and are transversal with the oriented angles of intersection given by . Then, at least locally, and smoothly depend on . Let
(82) |
which is a Banach submanifold of . Also, let
(83) |
The mapping defined as in (80) has partial derivative with respect to as a map from
(84) |
to of the form (81). If the winding number of the Cherepanov problem defined by (81) is greater or equal to , then the partial derivative is surjective with dimensional kernel. Hence implicit function theorem applies and there is a neighbourhood of in and a neighbourhood of in such that for every close to there is a dimensional family of solutions of (12-13) near .
5 A priori estimates
5.1 A priori estimates on function
To get existence results using continuity method we need a priori estimates on solutions of (3-4). It is well known that such a priori estimates can only be achieved for the family of solutions with no zeros on , [8, 16, 18]. We follow the approach in [8].
By assumption all Jordan curves contain point in their interiors. Hence the function
(85) |
defined for such that , is homotopic to in and it can be written in the form
(86) |
for some functions and , defined for such that and . Observe that for each function represents the angle between and the outer normal to at point .
Remark 5.1
There exists a isotopy , , where and for large enough, such that the gradient is nonzero on for each , [8]. Then one can find functions and such that (86) holds for each and . In addition, the isotopy can be made such that for every , , Jordan curves are strongly starshaped with respect to and that for each the angle between and the normal to at is less than .
Instead of solving (12-13) on the unit disc we consider equivalent problem on the upper semidisc , where the role of the lower semicircle is replaced by the interval . Using the reflection principle we can holomorphically extend every solution of (12-13) to the unit disc such that it solves nonlinear Riemann-Hilbert problem defined by the function which we get as an extension of the original function using the reflection to the lower semicircle as
(87) |
For function has well defined limits as approaches from above and below. Then we have
(88) |
and hence
(89) |
Therefore and . Also, observe that for , an intersection of with the real axis, we have
(90) |
and similarly for an intersection of with the real axis.
Thus for every solution of (12-13) the absolute value of is well defined and continuous on , whereas
(91) |
and similarly at .
Let be a solution of the symmetrized boundary value problem with no zeros. Hence can be written in the exponential form
(92) |
Remark 5.2
Let us differentiate function to get
(93) |
Since , we get
(94) |
and so
(95) |
From here we get
(96) |
Observe that function
(97) |
extends holomorphically to the unit disc with value at .
We will get a priori estimates on and hence on by getting a priori estimates on function (97). Using Hilbert transform it is enough to get a priori estimates on its real part. Hence we need a priori estimates on the right hand side of (96).
Function
(98) |
is bounded with the bound which does not depend on function but only on the data and defining function . The bound can also be found to be independent of the isotopy , . Hence one needs a priori bound on function
(99) |
Recall that, [18, p. 23], for , such that we have the estimate
(100) |
Let and let be smooth functions on with values in such that on , on , on , and on .
Let us consider the function
(101) |
for and for . Here we used notation and .
We see that and so is a continuous function on . Let be given. By results from [8, p. 881] we can write , where and is a finite sum of terms of the form , , , on which Hilbert transform acts as a bounded nonlinear operator from into .
Therefore for a given solution of (12-13) with no zeros we can write continuous function on in the form
(102) |
and so
(103) |
where the first term is uniformly bounded and for the second we have . Hence
(104) |
where the first factor is uniformly bounded and the second factor is bounded in for a given .
Since for a given we can get bounds on (104), the boundedness of in some is determined by Hilbert transform of the extension of function to .
Recall that are strongly starshaped Jordan curves with respect to and we may assume that for we have
(105) |
for some positive function on . A short calculation gives
(106) |
and so
(107) |
which has strictly positive real part on . Functions and represent the argument of (107). By compactness it follows that there exists , such that and on and therefore
(108) |
for every . Also, if there is an open condition on the size of on , such as , , we can, by choosing the supports of functions and small enough, that is, by choosing small enough, assume that the same condition on the size holds for function for all . Observe also that and . and so
(109) |
By (100) and (104) we get that for every fixed such that the estimate
(110) |
holds. Hence we also have a priori estimate on function (97).
Since
(111) |
an estimate on will give an estimate on . We can write
(112) |
By assumptions of Theorem 1.2 we have and we can choose . By Cauchy-Schwarz inequality we then have
(113) |
From here we get a priori estimates on which imply a priori estimates on and in Hölder space for . Recall (Remark 5.2) that this gives Hölder space a priori estimates on solutions with no zeros of the nonsymmetrical problem (12-13) for .
5.2 A priori estimates on function
We also need a priori estimates on function for which it holds
(114) |
In this subsection we again consider the nonsymmetrical case (12-13). We denote by a universal constant, which depends on the data but does not depend on the particular function we consider.
We know that and hence are smooth on and on compact subsets of we get a priori estimates on by expressing it in terms of . Hence we need a priori estimates on near points . Also, we know from Section 3 that if is continuous on , then both functions belong to for
(115) |
Let us fix that we have a priori estimates on function .
Recall (38) and that . Hence is a function with a priori bounds. As in (60) we can globally write
(116) |
where and are real functions with a priori bounds. To get a priori bounds on for some we will get bounds on the right-hand side function , that it, on near . Considering (41-42-43) termwise we get that , , , and all terms with function are bounded for any such that
(117) |
Let us consider terms which are bounded by . Since we have a priori bounds on , we have
(118) |
for some universal constant . Hence
(119) |
and this function is in some , , if . The bound implies that this will be the case for some such if . Similar argument near gives .
If these two conditions are satisfied, we get a priori estimates on for some . This implies a priori estimate on for .
There are natural bounds on , , in terms of , that is,
(120) |
Hence, if and both inequalities needed for a priori estimates will be satisfied. These two inequalities are equivalent to the condition , that is, the angle between and the normal to at is less than .
6 Final remarks
If arc is the lower semicircle, we can state Theorem 1.2 in an equivalent simplified form.
Theorem 6.1
Let be a family of Jordan curves in which all contain point in their interiors. Let Jordan curves , , be strongly starshaped with respect to and such that for each the angle between and the outer normal to at is less than . Let , , be the positive intersection of and the real axis with the oriented angle of intersection , where and . Let
(121) |
Then there exists a unique with no zeros on which solves (3-4) for which and .
To prove Theorem 6.1 one uses continuity method (see also [8]). The starting boundary value problem (3-4) can be, using an isotopy from Jordan curves to circles with center at and fixed radius , embedded in a one parameter family of boundary value problems which all satisfy assumptions of Theorem 6.1. Here, for we have the starting boundary value problem and for circles as the boundary data.
Results in Section 4 (Proposition 4.4, Proposition 4.5) imply that a solution of the boundary value problem (3-4) for curves can be locally perturbed into a solution for the nearby perturbed boundary data. Hence the set of parameters for which there is a solution of (3-4) is open. On the other hand, a priori estimates from Section 5 together with compact embeddings (10) imply that the set of parameters for which there is a solution of (3-4) is closed. Since there is an obvious solution for the case , where all Jordan curves are circles with center at and fixed radius , we get that there is a solution of (3-4) for .
Corollary 6.2
To prove the corollary we search for solutions of the symmetric problem on the unit disc of the form
(122) |
where is a point in the upper half-disc. Then has to solve a modified problem, where the boundary curves are given by
(123) |
Observe that for .
Remark 6.3
In a similar way one can create a zero at a point , that is, on in the original problem. Jordan curves for the modified problem are
(124) |
Then and but conditions of Theorem 1.2 are still satisfied.
Acknowledgements
The author is grateful to the referee for his/her valuable suggestions and comments.
Disclosure
No potential conflict of interest was reported by the author.
Funding
The author acknowledges the financial support from the Slovenian Research Agency (grants P1-0291, J1-3005, N1-0237 and N1-0137).
References
- Čerepanov [1962] G. P. Čerepanov. A nonlinear theory-of-functions boundary problem in some elastoplastic problems (Russian). Dokl. Akad. Nauk SSSR, pages 566–568, 1962.
- Čerepanov [1963] G. P. Čerepanov. An elastic-plastic problem under conditions of antiplanar deformation. J. Appl. Math. Mech., pages 1040–1057, 1963.
- Černe [1995] M. Černe. Analytic discs attached to a generating CR-manifold. Ark. Mat., 22:217–248, 1995.
- Černe [1997] M. Černe. Regularity of discs attached to a submanifold of . J. Anal. Math., 72:261–278, 1997.
- Černe [2019] M. Černe. Some remarks on the nonlinear mixed Cherepanov boundary value problem. Complex Var. Elliptic Equ., 65:1601–1611, 2019.
- Čirka [1982] E. M. Čirka. Regularity of the boundaries of analytic sets (Russian). Mat. Sb., 117:291–336, 1982.
- Čirka et al. [1999] E. M. Čirka, B. Coupet, and A. B. Sukhov. On boundary regularity of analytic discs. Michigan Math. J., 46:271–279, 1999.
- Forstnerič [1988] F. Forstnerič. Polynomial hulls of sets fibered over the circle. Indiana Univ. Math. J., 37:869–889, 1988.
- G. Hansen and Moszyńska [2020] H. Martini G. Hansen, I. Herburt and M. Moszyńska. Starshaped sets. Aequat. Math., 94:1001–1092, 2020.
- Kaiser and Lehner [2017] T. Kaiser and S. Lehner. Asymptotic behaviour of the Riemann mapping function at analytic cusps. Ann. Acad. Sci. Fenn. Math., 42:3–15, 2017.
- Koppelman [1959] W. Koppelman. The Riemann-Hilbert problem for finite Riemann surfaces. Comm. Pure Appl. Math., 12:13–35, 1959.
- Lesley [1985] F. D. Lesley. Conformal mappings of domains satisfying a wedge condition. Proc. Amer. Math. Soc., 93:483–488, 1985.
- Mityushev and Rogosin [2000] V. V. Mityushev and S. V. Rogosin. Constructive methods for linear and nonlinear boundary value problems for analytic functions. Theory and applications. Chapman and Hall/CRC Monographs and Surveys in Pure and Applied Mathematics 108, Boca Raton, 2000.
- Obnosov and Zulkarnyaev [2019] Y. Obnosov and A. Zulkarnyaev. Nonlinear mixed Cherepanov boundary-value problem. Complex Var. Elliptic Equ., 64:979–996, 2019.
- Oh [1995] Y.-G. Oh. Riemann-Hilbert problem and application to the perturbation theory of analytic discs. Kyungpook Math. J., 35:39–75, 1995.
- Šnirelman [1972] A.I. Šnirelman. The degree of a quasiruled mapping, and a nonlinear Hilbert problem (Russian). Mat. Sb, 89(131):366–389, 1972.
- Wegert [1987] E. Wegert. Topological methods for strongly nonlinear Riemann-Hilbert problems for holomorphic functions. Math. Nachr., 134:201–230, 1987.
- Wegert [1992] E. Wegert. Nonlinear boundary value problems for holomorphic functions and singular integral equations. Mathematical Research 65, Akademie-Verlag, Berlin, 1992.
- Wen [1992] G. C. Wen. Conformal mappings and boundary value problems. Translations of Mathematical Monographs 106, American Mathematical Society, Providence, RI, 1992.