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

Mixed Riemann-Hilbert boundary value problem
with simply connected fibers

Miran Černe111 The author was supported in part by grants Analiza in geometrija P1-0291, Kompleksna in geometrijska analiza J1-3005, Holomorfne parcialne diferencialne relacije N1-0237 and Nelinearni valovi in spektralna teorija N1-0137 from ARRS, Republic of Slovenia. [email protected]
Abstract

We study the existence of solutions of mixed Riemann-Hilbert or Cherepanov boundary value problem with simply connected fibers on the unit disk Δ{\rm\Delta}. Let LL be a closed arc on Δ\partial{\rm\Delta} with the end points ω1,ω1\omega_{-1},\omega_{1} and let aa be a smooth function on LL with no zeros. Let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a smooth family of smooth Jordan curves in {\mathbb{C}} which all contain point 0 in their interiors and such that γω1\gamma_{\omega_{-1}}, γω1\gamma_{\omega_{1}} are strongly starshaped with respect to 0. Then under condition that for each wγω±1w\in\gamma_{\omega_{\pm 1}} the angle between ww and the normal to γω±1\gamma_{\omega_{\pm 1}} at ww is less than π10\frac{\pi}{10}, there exists a Hölder continuous function ff on Δ¯\overline{{\rm\Delta}}, holomorphic on Δ{\rm\Delta}, such that

Re(a(ξ)¯f(ξ))=0onLandf(ξ)γξonΔL̊.{\rm Re}(\overline{a(\xi)}f(\xi))=0{\rm\ on\ }L\ \ \ \ \ {\rm and}\ \ \ \ \ f(\xi)\in\gamma_{\xi}{\rm\ on\ }\partial{\rm\Delta}\setminus\mathring{L}.
keywords:
Boundary value problem, mixed Riemann-Hilbert problem, Cherepanov problem
MSC:
[2020] 35Q15, 30E25
\affiliation

organization=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 Δ={z;|z|<1}{\rm\Delta}=\{z\in{\mathbb{C}};|z|<1\} be the open unit disc in the complex plane {\mathbb{C}} and let Δ={ξ;|ξ|=1}\partial{\rm\Delta}=\{\xi\in{\mathbb{C}};|\xi|=1\} be the unit circle. Let LL be a closed arc on Δ\partial{\rm\Delta}, let L̊\mathring{L} denote its interior with respect to Δ\partial{\rm\Delta}, and let a:LΔa:L\rightarrow\partial{\rm\Delta} be a smooth function.

Recall that the interior Int(γ){\rm Int}(\gamma) of a Jordan curve γ\gamma\subseteq{\mathbb{C}} is the bounded component of γ{\mathbb{C}}\setminus\gamma. We orient γ\gamma positively with respect to Int(γ){\rm Int}(\gamma). Jordan curve γ\gamma\subset{\mathbb{C}} is starshaped with respect to 0, if for any point ww in the interior of γ\gamma the line segment which connects points 0 and ww lies in the interior of γ\gamma, and it is strongly starshaped with respect to 0, [9], if there exists a positive continuous function RR on the unit circle such that

γ={w;|w|=R(w|w|)}\gamma=\left\{w\in{\mathbb{C}};|w|=R\left(\frac{w}{|w|}\right)\right\} (1)

and

Int(γ)={w{0};|w|<R(w|w|)}{0}.{\rm Int}(\gamma)=\left\{w\in{\mathbb{C}}\setminus\{0\};|w|<R\left(\frac{w}{|w|}\right)\right\}\cup\{0\}. (2)

Let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a smooth family of smooth Jordan curves in {\mathbb{C}} which all contain point 0 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 ff on Δ¯\overline{{\rm\Delta}}, holomorphic on Δ{\rm\Delta}, such that

Re(a(ξ)¯f(ξ))=0forξL{\rm Re}(\overline{a(\xi)}f(\xi))=0{\rm\ for\ }\xi\in L (3)

and

f(ξ)γξforξΔL̊.f(\xi)\in\gamma_{\xi}{\rm\ for\ }\xi\in\partial{\rm\Delta}\setminus\mathring{L}. (4)

That is, ff solves a linear Riemann-Hilbert problem on LL and a nonlinear Riemann-Hilbert problem with simply connected fibers on ΔL̊\partial{\rm\Delta}\setminus\mathring{L}. See also [1, 2, 13, 14].

The problem with circular fibers γξ\gamma_{\xi} and LL 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 L=ΔL=\partial{\rm\Delta} or L=L=\emptyset. If L=ΔL=\partial{\rm\Delta}, we consider a homogeneous linear Riemann-Hilbert problem. In this case the essential information on the problem is given by the winding number W(a)W(a) of function aa. It is well known [11, 17, 18] that if the winding number W(a)W(a) is nonnegative, the space of solutions of (3) is a vector subspace of Aα(Δ)A^{\alpha}({\rm\Delta}), 0<α<10<\alpha<1, of real dimension 2W(a)+12W(a)+1.

Remark 1.1

The linear Riemann-Hilbert problem can also be considered in the case of a nonorientable line bundle over Δ\partial{\rm\Delta}, that is, in the case where at some point ξ0Δ\xi_{0}\in\partial{\rm\Delta} we have a(ξ0)=a(ξ0+)a(\xi_{0}^{-})=-a(\xi_{0}^{+}). Then the winding number of function a2a^{2} or the Maslov index of the problem is an odd integer. In this case it holds that if W(a2)1W(a^{2})\geq-1, or, with a little bit of abuse of notation, if W(a)12W(a)\geq-\frac{1}{2}, then the space of solutions of (3) is a vector subspace of Aα(Δ)A^{\alpha}({\rm\Delta}) of real dimension 2W(a)+12W(a)+1, see [3, 4, 15, 18].

If LL is empty, we have a nonlinear Riemann-Hilbert problem with smooth simply connected fibers which all contain 0 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 mm zeros on Δ{\rm\Delta}, m{0}m\in{\mathbb{N}}\cup\{0\}, forms a manifold in space Aα(Δ)A^{\alpha}({\rm\Delta}) of dimension 2m+12m+1, and this manifold is compact if and only if m=0m=0. We assume from now on that neither L=L=\emptyset nor L=ΔL=\partial{\rm\Delta}.

Theorem 1.2

Let k3k\geq 3. Let a:L{0}a:L\rightarrow{\mathbb{C}}\setminus\{0\} be a Ck+1C^{k+1} function and let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a CkC^{k} family of Jordan curves in {\mathbb{C}} which all contain point 0 in their interiors. Let ω1\omega_{1} and ω1\omega_{-1} be the first and the last point of arc LL with respect to the positive orientation of Δ\partial{\rm\Delta}. Let Jordan curves γωj\gamma_{\omega_{j}}, j=±1j=\pm 1, be strongly starshaped with respect to 0 and such that for each wγωjw\in\gamma_{\omega_{j}} the angle between ww and the outer normal to γωj\gamma_{\omega_{j}} at ww is less than π10\frac{\pi}{10}. Let wjw_{j}, j=±1j=\pm 1, be the intersection of γωj\gamma_{\omega_{j}} and the line Re(a(ωj)¯w)=0{\rm Re}(\overline{a(\omega_{j})}w)=0 of the form λ(ia(ωj))\lambda(-ia(\omega_{j})), λ>0\lambda>0, and let πβj\pi\beta_{j} be the oriented angle of intersection of the line Re(a(ωj)¯w)=0{\rm Re}(\overline{a(\omega_{j})}w)=0 with the fiber γωj\gamma_{\omega_{j}} at point wjw_{j}, where β1(0,1)\beta_{1}\in(0,1) and β1(1,0)\beta_{-1}\in(-1,0). Let

0<β<min{β1,1β1,|β1|,1|β1|}.0<\beta<\min\{\beta_{1},1-\beta_{1},|\beta_{-1}|,1-|\beta_{-1}|\}. (5)

Then there exists a unique fAβ(Δ)f\in A^{\beta}({\rm\Delta}) with no zeros on Δ{\rm\Delta} which solves (3-4) for which f(ω1)=w1f(\omega_{1})=w_{1} and f(ω1)=w1f(\omega_{-1})=w_{-1}.

Remark 1.3

Here β1>0\beta_{1}>0, if the tangent vector ia(ω1)-ia(\omega_{1}) to Re(a(ω1)¯w)=0{\rm Re}(\overline{a(\omega_{1})}w)=0 is rotated counterclockwise by angle πβ1\pi\beta_{1} to get a positive tangent vector to γω1\gamma_{\omega_{1}} at point w1w_{1}, and β1<0\beta_{-1}<0, if a positive tangent vector to γω1\gamma_{\omega_{-1}} at w1w_{-1} is rotated clockwise by angle π|β1|\pi|\beta_{1}| to get tangent vector ia(ω1)-ia(\omega_{-1}) to Re(a(ω1)¯w)=0{\rm Re}(\overline{a(\omega_{-1})}w)=0.

Remark 1.4

Observe that conditions in Theorem 1.2 imply ||βj|12|<110||\beta_{j}|-\frac{1}{2}|<\frac{1}{10}, j=±1j=\pm 1, and hence one could choose β=25\beta=\frac{2}{5}.

Remark 1.5

In the cases considered in [5, 14] all boundary curves were circles with center at point 0. Hence |βj|=12|\beta_{j}|=\frac{1}{2}, j=±1j=\pm 1, and the maximal regularity we got was β<12\beta<\frac{1}{2}.

Corollary 1.6

Let a1,,anΔa_{1},\dots,a_{n}\in{\rm\Delta} be a finite set of points with given multiplicities. Then under the assumptions of Theorem 1.2 there exists β(0,1)\beta\in(0,1) and fAβ(Δ)f\in A^{\beta}({\rm\Delta}) which has zeros exactly at points a1,,anΔa_{1},\dots,a_{n}\in{\rm\Delta} with the given multiplicites and which solves (3-4).

2 Function spaces, Hilbert transform and
defining functions

Let 0<α<10<\alpha<1 and let GG\subset{\mathbb{C}} be a compact subset. We denote by Cα(G)C^{\alpha}(G) the algebra over {\mathbb{C}} of Hölder continuous complex functions on GG and by Cα(G)C_{{\mathbb{R}}}^{\alpha}(G) the algebra over {\mathbb{R}} of real Hölder continuous functions on GG. Using the norm

fα=maxzG|f(z)|+supz,wG,zw|f(z)f(w)||zw|α\|f\|_{\alpha}=\max_{z\in G}|f(z)|+\sup_{z,w\in G,z\neq w}\frac{|f(z)-f(w)|}{|z-w|^{\alpha}} (6)

the algebras Cα(G)C^{\alpha}(G) and Cα(G)C_{{\mathbb{R}}}^{\alpha}(G) become Banach algebras. For G=Δ¯G=\overline{{\rm\Delta}} or G=ΔG=\partial{\rm\Delta} and k{0}k\in{\mathbb{N}}\cup\{0\} we also define spaces Ck,α(G)C^{k,\alpha}(G) and Ck,α(G)C_{{\mathbb{R}}}^{k,\alpha}(G) of kk times continuously differentiable functions on GG, whose all kk-th derivatives belong to space Cα(G)C^{\alpha}(G) or space Cα(G)C_{{\mathbb{R}}}^{\alpha}(G).

We also need some algebras of holomorphic functions on Δ{\rm\Delta}. By A(Δ)A({\rm\Delta}) we denote the disc algebra, that is, the algebra of continuous functions on Δ¯\overline{{\rm\Delta}} which are holomorphic on Δ{\rm\Delta}, and by Aα(Δ)=A(Δ)Cα(Δ¯)A^{\alpha}({\rm\Delta})=A({\rm\Delta})\cap C^{\alpha}(\overline{{\rm\Delta}}) the algebra of Hölder continuous functions on the closed disc which are holomorphic on Δ{\rm\Delta}. Using appropriate norms, that is, the maximum norm \|\cdot\|_{\infty} for A(Δ)A({\rm\Delta}) and the Hölder norm α\|\cdot\|_{\alpha} for Aα(Δ)A^{\alpha}({\rm\Delta}), these algebras become Banach algebras. Similarly we define Ak,α(Δ)=A(Δ)Ck,α(Δ¯)A^{k,\alpha}({\rm\Delta})=A({\rm\Delta})\cap C^{k,\alpha}(\overline{{\rm\Delta}}) (k{0},0<α<1)(k\in{\mathbb{N}}\cup\{0\},0<\alpha<1).

Recall that Hilbert transform HH assigns to a real function uu on Δ\partial{\rm\Delta} a real function HuHu on Δ\partial{\rm\Delta} such that the harmonic extension of f=u+iHuf=u+i\,Hu to Δ{\rm\Delta} is holomorphic on Δ{\rm\Delta} and real at 0. It is known that HH is a bounded linear operator on Ck,α(Δ)C_{{\mathbb{R}}}^{k,\alpha}(\partial{\rm\Delta}) (k{0},0<α<1)(k\in{\mathbb{N}}\cup\{0\},0<\alpha<1), [18, §1.6.11], and hence the harmonic extension of f=u+iHuf=u+i\,Hu to Δ{\rm\Delta} belongs to Ak,α(Δ)A^{k,\alpha}({\rm\Delta}). Also, [18, §1.6.11], the Hilbert transform is a bounded linear operator on the Sobolev space Wpk(Δ)W^{k}_{p}(\partial{\rm\Delta}) of kk times generalized differentiable functions with derivatives in Lp(Δ)L^{p}(\partial{\rm\Delta}) (k{0},1<p<)(k\in{\mathbb{N}}\cup\{0\},1<p<\infty) equipped with the norm

fWpk=(j=0kDjfp)1p.\|f\|_{W_{p}^{k}}=\left(\sum_{j=0}^{k}\|D^{j}f\|_{p}\right)^{\frac{1}{p}}. (7)

Recall, [18, §1.6.14], that if Δ=T1T2\partial{\rm\Delta}=T_{1}\cup T_{2} is a partition of Δ\partial{\rm\Delta} in two subarcs T1T_{1} and T2T_{2} and if T0T1T_{0}\subseteq T_{1} is a compactly contained subarc of T1T_{1}, then for k{0}k\in{\mathbb{N}}\cup\{0\}, 1<p<1<p<\infty, 0<α<10<\alpha<1 there exists a constant C=C(k,p,α)C=C(k,p,\alpha) such that

HuWpk(T0)C(uWpk(T1)+uL1(T2))\|Hu\|_{W^{k}_{p}(T_{0})}\leq C(\|u\|_{W^{k}_{p}(T_{1})}+\|u\|_{L^{1}(T_{2})}) (8)

and

HuCk,α(T0)C(uCk,α(T1)+uL1(T2)).\|Hu\|_{C^{k,\alpha}(T_{0})}\leq C(\|u\|_{C^{k,\alpha}(T_{1})}+\|u\|_{L^{1}(T_{2})}). (9)

We will also need compact embedding result, [18, §1.1.8],

Wp1(Δ)Cβ(Δ)Cα(Δ)W^{1}_{p}(\partial{\rm\Delta})\hookrightarrow C^{\beta}(\partial{\rm\Delta})\hookrightarrow C^{\alpha}(\partial{\rm\Delta}) (10)

for 0<α<β<11p0<\alpha<\beta<1-\frac{1}{p}, 1<p<1<p<\infty, which holds on arcs in Δ\partial{\rm\Delta} as well.

Since LΔL\neq\partial{\rm\Delta} we can extend aa to Δ\partial{\rm\Delta} as a nowhere zero function of class Ck+1C^{k+1} so that the winding number W(a)=0W(a)=0. Therefore, [18, p. 25], we can write a¯\overline{a} in the form

a¯=reh,\overline{a}=re^{h}, (11)

where r>0r>0 is a positive Ck,αC^{k,\alpha} function on Δ\partial{\rm\Delta} and hAk,α(Δ)h\in A^{k,\alpha}({\rm\Delta}). Thus the original problem (3-4) is equivalent to the problem

Im(f(ξ))=0forξL{\rm Im}(f_{\ast}(\xi))=0{\rm\ \ for\ \ }\xi\in L (12)

and

f(ξ)γξforξΔL̊,f_{\ast}(\xi)\in\gamma^{\ast}_{\xi}{\rm\ \ for\ \ }\xi\in\partial{\rm\Delta}\setminus\mathring{L}, (13)

where f=iehff_{\ast}=ie^{h}f and γξ=ieh(ξ)γξ\gamma^{\ast}_{\xi}=ie^{h(\xi)}\gamma_{\xi}. Observe that the number of zeros of ff_{\ast} and ff are the same and that 0 belongs to the interiors of all curves γξ\gamma^{\ast}_{\xi}, ξΔL̊\xi\in\partial{\rm\Delta}\setminus\mathring{L}. Also, since for each ξΔ\xi\in\partial{\rm\Delta} the transfomation

wieh(ξ)ww\longmapsto ie^{h(\xi)}w (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 L={ξΔ;Im(ξ)0}L=\{\xi\in\partial{\rm\Delta};{\rm Im}(\xi)\leq 0\} is the lower semicircle. From now on we will consider problem (12-13) with the addition that LL is the lower semicircle and instead of ff_{\ast} and {γξ}ξΔL̊\{\gamma^{\ast}_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} we will still write ff and {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}}.

Remark 2.1

One can also create the ’double’ of the boundary value problem. Using a biholomorphism one can replace the unit disc Δ{\rm\Delta} with the upper half-disk Δ+={ξΔ;Im(ξ)>0}{\rm\Delta}_{+}=\{\xi\in{\rm\Delta};{\rm Im}(\xi)>0\} and LL by the interval [1,1][-1,1].

By the reflection principle we see that problem (12-13) is equivalent to the nonlinear Riemann-Hilbert problem on Δ{\rm\Delta}, where the boundary curves {γξ}ξΔ+L̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}_{+}\setminus\mathring{L}} are symmetrically extended and defined on the lower semicircle so that we have

γξ=γξ¯¯\gamma_{\xi}=\overline{\gamma_{\overline{\xi}}} (15)

for every ξΔ{1,1}\xi\in\partial{\rm\Delta}\setminus\{1,-1\}. In general this symmetrical extension of Jordan curves {γξ}ξΔ+L̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}_{+}\setminus\mathring{L}} to the lower semicircle produces boundary data which are not continuous at points 11 and 1-1. Because the biholomorphism from Δ{\rm\Delta} to the upper semidisc is in A12(Δ)A^{\frac{1}{2}}({\rm\Delta}), 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 {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} in {\mathbb{C}}. Let kk\in{\mathbb{N}}. The family of Jordan curves {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} is a CkC^{k} family parametrized by ξΔL̊\xi\in\partial{\rm\Delta}\setminus\mathring{L} if there exists a function ρCk((ΔL̊)×)\rho\in C^{k}((\partial{\rm\Delta}\setminus\mathring{L})\times{\mathbb{C}}) such that

γξ={w;ρ(ξ,w)=0}andInt(γξ)={w;ρ(ξ,w)<0},\gamma_{\xi}=\{w\in{\mathbb{C}};\rho(\xi,w)=0\}\ \ \ {\rm and}\ \ \ {\rm Int}(\gamma_{\xi})=\{w\in{\mathbb{C}};\rho(\xi,w)<0\}, (16)

and the gradient ρw¯(ξ,w)=ρw¯(ξ,w)0\frac{\partial\rho}{\partial\overline{w}}(\xi,w)=\rho_{\overline{w}}(\xi,w)\neq 0 for every ξΔL̊\xi\in\partial{\rm\Delta}\setminus\mathring{L} and wγξw\in\gamma_{\xi}. We call ρ\rho a defining function for CkC^{k} family of Jordan curves {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}}. We will consider only bounded families of Jordan curves which all lie in some fixed disc Δ(0,R)¯\overline{{\rm\Delta}(0,R)}, R>0R>0, and the space Ck((ΔL̊)×Δ(0,R)¯)C^{k}((\partial{\rm\Delta}\setminus\mathring{L})\times\overline{{\rm\Delta}(0,R)}) is equipped with the standard CkC^{k} norm.

Since we assume that γ±1\gamma_{\pm 1} are strongly starshaped Jordan curves, we also assume that for ρ\rho, the defining function for Jordan curves {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}}, and j=±1j=\pm 1 we have

ρ(j,w)=|w|2Rj2(w|w|)\rho(j,w)=|w|^{2}-R_{j}^{2}\left(\frac{w}{|w|}\right) (17)

for some positive CkC^{k} functions Rj(z)R_{j}(z) on {\mathbb{C}}.

Using parametrization θeiθ\theta\mapsto e^{i\theta} of the unit circle we will also use the notation γθ\gamma_{\theta}, ρ(θ,w)\rho(\theta,w) and ρθ(θ,w)\rho_{\theta}(\theta,w) instead of γξ\gamma_{\xi}, ρ(ξ,w)\rho(\xi,w) and ρξ(ξ,w)\rho_{\xi}(\xi,w). Also, for a function hh on Δ\partial{\rm\Delta}, we will write either h(ξ)h(\xi) or h(θ)h(\theta), where ξ=eiθ\xi=e^{i\theta}. Observe that if hh is holomorphic on Δ{\rm\Delta} with well defined derivative on Δ\partial{\rm\Delta}, then hθ(θ)=iξh(ξ)\frac{\partial h}{\partial\theta}(\theta)=i\xi h^{\prime}(\xi) for ξ=eiθ\xi=e^{i\theta}.

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 ρ\rho given as

ρ(ξ,w)=ρ(ξ¯,w¯)\rho(\xi,w)=\rho(\overline{\xi},\overline{w}) (18)

for every ξΔ{1,1}\xi\in\partial{\rm\Delta}\setminus\{1,-1\} and every ww\in{\mathbb{C}}.

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 ρCk((ΔL̊)×)\rho\in C^{k}((\partial{\rm\Delta}\setminus\mathring{L})\times{\mathbb{C}}) (k3)(k\geq 3).

Let fA(Δ)f\in A({\rm\Delta}) be a solution of (12-13). It is well known [6, 7, 8, 18] that ff restricted to Δ{1,1}\partial{\rm\Delta}\setminus\{-1,1\} is in Ck1,αC^{k-1,\alpha} for any 0<α<10<\alpha<1. Hence we need information on the regularity of ff near points ξ=±1\xi=\pm 1. For j=±1j=\pm 1 we denote f(j)=wjγjf(j)=w_{j}\in{\mathbb{R}}\cap\gamma_{j}.

Using Möbius tranformation from the unit disc Δ{\rm\Delta} to the upper half-plane ={z;Im(z)>0}{\cal H}=\{z\in{\mathbb{C}};{\rm Im}(z)>0\} we consider the case where ff is bounded and continuous on ¯\overline{\cal H} and holomorphic on {\cal H}. Also, point ξ=1\xi=1 is mapped into t=0t=0 and point ξ=1\xi=-1 into \infty. Now ff solves the problem

Im(f(t))=0fort0{\rm Im}(f(t))=0{\rm\ for\ }t\leq 0 (19)

and

f(t)γtfort0.f(t)\in\gamma_{t}{\rm\ for\ }t\geq 0. (20)

Also, using translation, we will assume that f(0)=0γ0f(0)=0\in{\mathbb{R}}\cap\gamma_{0}.

Let πβ1\pi\beta_{1} (β1(1,1){0})(\beta_{1}\in(-1,1)\setminus\{0\}) be the oriented angle of intersection of the real axis Im(w)=0{\rm Im}(w)=0 and γ0\gamma_{0} at w=f(0)=0w=f(0)=0. The orientation of the real axis is positive with respect to the upper half-plane and the orientation of γ0\gamma_{0} is positive with respect to the interior of γ0\gamma_{0}. Hence β1>0\beta_{1}>0, if the tangent vector to the real axis is rotated counterclockwise by angle πβ1\pi\beta_{1} to get a tangent vector to γ0\gamma_{0} at point 0, and β1<0\beta_{1}<0, if the tangent vector to the real axis is rotated clockwise by angle π|β1|\pi|\beta_{1}| to get a tangent vector to γ0\gamma_{0} at 0.

The defining function ρ\rho can near (0,0)(0,0) for t0t\geq 0 be written as

ρ(t,w)=ρ(0,0)+ρt(0,0)t+2Re(ρw(0,0)w)+12ρtt(0,0)t2+\rho(t,w)=\rho(0,0)+\rho_{t}(0,0)t+2{\rm Re}(\rho_{w}(0,0)w)+\frac{1}{2}\rho_{tt}(0,0)t^{2}+ (21)
+ρww¯(0,0)|w|2+Re(ρww(0,0)w2+ρtw(0,0)tw)+t2+|w|2g(t,w),+\,\rho_{w\overline{w}}(0,0)|w|^{2}+{\rm Re}(\rho_{ww}(0,0)w^{2}+\rho_{tw}(0,0)tw)+\sqrt{t^{2}+|w|^{2}}\,g(t,w), (22)

where gC1(×)g\in C^{1}({\mathbb{R}}\times{\mathbb{C}}) such that g(0,0)=gt(0,0)=gw(0,0)=gw¯(0,0)=0g(0,0)=g_{t}(0,0)=g_{w}(0,0)=g_{\overline{w}}(0,0)=0.

Recall that ρ(0,0)=0\rho(0,0)=0 and that ρw¯(0,0)\rho_{\overline{w}}(0,0) represents an outer normal to γ0\gamma_{0} at point w=0w=0. So we have

ρw¯(0,0)=iλeiπβ1\rho_{\overline{w}}(0,0)=-i\lambda e^{i\pi\beta_{1}} (23)

for some real λ>0\lambda>0. We may assume λ=12\lambda=\frac{1}{2}.

Because

Re(ieiπβ1w)=Im(eiπβ1w)=Im(eiπ(1β1)w){\rm Re}(ie^{-i\pi\beta_{1}}w)=-{\rm Im}(e^{-i\pi\beta_{1}}w)={\rm Im}(e^{i\pi(1-\beta_{1})}w) (24)

we have

ρ(t,w)=At+Im(eiπ(1β1)w)+Bt2+C|w|2+\rho(t,w)=At+{\rm Im}(e^{i\pi(1-\beta_{1})}w)+Bt^{2}+C|w|^{2}+ (25)
+Re(Dw2)+tRe(Ew)+t2+|w|2g(t,w)+\,{\rm Re}(Dw^{2})+t\,{\rm Re}(Ew)+\sqrt{t^{2}+|w|^{2}}\,g(t,w) (26)

for some A,B,CA,B,C\in{\mathbb{R}} and D,ED,E\in{\mathbb{C}}.

Let us assume that we have a solution ff of the problem (19-20) of the form

f(t)=tsκ(t),f(t)=t^{s}\kappa(t), (27)

where κ\kappa is bounded and continuous on ¯\overline{\cal H}, holomorphic on {\cal H}, and 0<s<10<s<1 to be determined.

For t0t\leq 0 we have t=(1)|t|t=(-1)|t| and from (19) we get

Im(eiπsκ(t))=Im(eiπ(1+s)κ(t))=0.{\rm Im}(e^{i\pi s}\kappa(t))=-{\rm Im}(e^{i\pi(1+s)}\kappa(t))=0. (28)

On the other hand for t>0t>0 we have

1tsρ(t,tsκ(s))=At1s+Im(eiπ(1β1)κ(t))+Bt2s+Cts|κ(t)|2+\frac{1}{t^{s}}\rho(t,t^{s}\kappa(s))=At^{1-s}+{\rm Im}(e^{i\pi(1-\beta_{1})}\kappa(t))+Bt^{2-s}+Ct^{s}|\kappa(t)|^{2}+ (29)
+tsRe(Dκ(t)2)+tRe(Eκ(t))+t22s+|κ(t)|2g(t,tsκ(t))=0.+t^{s}{\rm Re}(D\kappa(t)^{2})+t{\rm Re}(E\kappa(t))+\sqrt{t^{2-2s}+|\kappa(t)|^{2}}\,g(t,t^{s}\kappa(t))=0. (30)

We choose 0<s<10<s<1 so that κ\kappa solves boundary value problem with continuous boundary data. That is, we choose s=1β1s=1-\beta_{1}, if β1>0\beta_{1}>0, and s=β1=|β1|s=-\beta_{1}=|\beta_{1}|, if β1<0\beta_{1}<0.

Thus κ\kappa solves the following Riemann-Hilbert problem

Im(eiπ(1β1)κ(t))=0fort0{\rm Im}(e^{i\pi(1-\beta_{1})}\kappa(t))=0{\rm\ for\ }t\leq 0 (31)

and

ρ~(t,κ(t))=0fort0,\widetilde{\rho}(t,\kappa(t))=0{\rm\ for\ }t\geq 0, (32)

where, if β1>0\beta_{1}>0,

ρ~(t,w)=Atβ1+Im(eiπ(1β1)w)+Bt1+β1+Ct1β1|w|2+\widetilde{\rho}(t,w)=At^{\beta_{1}}+{\rm Im}(e^{i\pi(1-\beta_{1})}w)+Bt^{1+\beta_{1}}+Ct^{1-\beta_{1}}|w|^{2}+ (33)
+t1β1Re(Dw2)+tRe(Ew)+t2β1+|w|2g(t,t1β1w),+\,t^{1-\beta_{1}}{\rm Re}(Dw^{2})+t{\rm Re}(Ew)+\sqrt{t^{2\beta_{1}}+|w|^{2}}\,g(t,t^{1-\beta_{1}}w), (34)

and, if β1<0\beta_{1}<0,

ρ~(t,w)=At1|β1|+Im(eiπ(1β1)w)+Bt2|β1|+Ct|β1||w|2+\widetilde{\rho}(t,w)=At^{1-|\beta_{1}|}+{\rm Im}(e^{i\pi(1-\beta_{1})}w)+Bt^{2-|\beta_{1}|}+Ct^{|\beta_{1}|}|w|^{2}+ (35)
+t|β1|Re(Dw2)+tRe(Ew)+t22|β1|+|w|2g(t,t|β1|w).+\,t^{|\beta_{1}|}{\rm Re}(Dw^{2})+t{\rm Re}(Ew)+\sqrt{t^{2-2|\beta_{1}|}+|w|^{2}}\,g(t,t^{|\beta_{1}|}w). (36)

For such choice of ss are the defining function for problem (31-32)

(t,w){ρ~(t,w)=1tsρ(t,tsw);t0,wIm(eiπ(1β1)w);t0,w(t,w)\longmapsto\left\{\begin{array}[]{rl}\widetilde{\rho}(t,w)=\frac{1}{t^{s}}\rho(t,t^{s}w);&t\geq 0,w\in{\mathbb{C}}\\ {\rm Im}(e^{i\pi(1-\beta_{1})}w);&t\leq 0,w\in{\mathbb{C}}\end{array}\right. (37)

and its partial ww-derivative

(t,w){ρ~w(t,w)=ρw(t,tsw);t0,w12ieiπ(1β1);t0,w(t,w)\longmapsto\left\{\begin{array}[]{rl}\widetilde{\rho}_{w}(t,w)=\rho_{w}(t,t^{s}w);&t\geq 0,w\in{\mathbb{C}}\\ \frac{1}{2i}e^{i\pi(1-\beta_{1})};&t\leq 0,w\in{\mathbb{C}}\end{array}\right. (38)

continuous on ×{\mathbb{R}}\times{\mathbb{C}}.

On the other hand, the partial derivative of defining function (37) with respect to the tt variable is not continuous at t=0t=0, but, as we will see, it still has certain LpL^{p} regularity properties, which will imply regularity conditions on κ\kappa and ff.

We know that κ\kappa is Ck1,αC^{k-1,\alpha} on {0}{\mathbb{R}}\setminus\{0\} and we can differentiate (31-32) on {0}{\mathbb{R}}\setminus\{0\} to get

Im(eiπ(1β1)κ(t))=0fort<0{\rm Im}(e^{i\pi(1-\beta_{1})}\kappa^{\prime}(t))=0{\rm\ for\ }t<0 (39)

and

ρ~t(t,κ(t))+2Re(ρ~w(t,κ(t))κ(t))=0fort>0.\widetilde{\rho}_{t}(t,\kappa(t))+2{\rm Re}(\widetilde{\rho}_{w}(t,\kappa(t))\kappa^{\prime}(t))=0{\rm\ for\ }t>0. (40)

For t>0t>0 and β1>0\beta_{1}>0 we have

ρ~t(t,w)=Aβ1tβ11+B(1+β1)tβ1+(1β1)Ctβ1|w|2+\widetilde{\rho}_{t}(t,w)=A\beta_{1}t^{\beta_{1}-1}+B(1+\beta_{1})t^{\beta_{1}}+(1-\beta_{1})Ct^{-\beta_{1}}|w|^{2}+ (41)
+(1β1)tβ1Re(Dw2)+Re(Ew)+β1t2β11t2β1+|w|2g(t,t1β1w)++(1-\beta_{1})t^{-\beta_{1}}{\rm Re}(Dw^{2})+{\rm Re}(Ew)+\frac{\beta_{1}\,t^{2\beta_{1}-1}}{\sqrt{t^{2\beta_{1}}+|w|^{2}}}\,g(t,t^{1-\beta_{1}}w)+ (42)
+t2β1+|w|2(gt(t,t1β1w)+2Re(gw(t,t1β1w)(1β1)tβ1w))+\sqrt{t^{2\beta_{1}}+|w|^{2}}\,(g_{t}(t,t^{1-\beta_{1}}w)+2{\rm Re}(g_{w}(t,t^{1-\beta_{1}}w)(1-\beta_{1})t^{-\beta_{1}}w)) (43)

and for t>0t>0 and β1<0\beta_{1}<0 we have

ρ~t(t,w)=A(1|β1|)t|β1|+B(2|β1|)t1|β1|+|β1|Ct|β1|1|w|2\widetilde{\rho}_{t}(t,w)=A(1-|\beta_{1}|)t^{-|\beta_{1}|}+B(2-|\beta_{1}|)t^{1-|\beta_{1}|}+|\beta_{1}|Ct^{|\beta_{1}|-1}|w|^{2} (44)
+|β1|t|β1|1Re(Dw2)+Re(Ew)+(1|β1|)t12|β1|t22|β1|+|w|2g(t,t|β1|w)++|\beta_{1}|t^{|\beta_{1}|-1}{\rm Re}(Dw^{2})+{\rm Re}(Ew)+\frac{(1-|\beta_{1}|)\,t^{1-2|\beta_{1}|}}{\sqrt{t^{2-2|\beta_{1}|}+|w|^{2}}}\,g(t,t^{|\beta_{1}|}w)+ (45)
+t22|β1|+|w|2(gt(t,t|β1|w)+2Re(gw(t,t|β1|w)|β1|t|β1|1w)).+\sqrt{t^{2-2|\beta_{1}|}+|w|^{2}}\,(g_{t}(t,t^{|\beta_{1}|}w)+2{\rm Re}(g_{w}(t,t^{|\beta_{1}|}w)|\beta_{1}|t^{|\beta_{1}|-1}w)). (46)

The tt-derivative of defining function (37) is 0 for t<0t<0.

Since β1(1,1){0}\beta_{1}\in(-1,1)\setminus\{0\} and κ\kappa is bounded, we have that ρ~t(t,κ(t))\widetilde{\rho}_{t}(t,\kappa(t)) is in Llocp()L^{p}_{\rm loc}({\mathbb{R}}) for

1p<min{1|β1|,11|β1|}.1\leq p<\min\left\{\frac{1}{|\beta_{1}|},\frac{1}{1-|\beta_{1}|}\right\}. (47)

A similar argument can be used for point ξ=1Δ\xi=-1\in\partial{\rm\Delta}. Let πβ1\pi\beta_{-1} (β1(1,1){0})(\beta_{-1}\in(-1,1)\setminus\{0\}) be the orientied angle of intersection of γ1\gamma_{-1} and the real axis Im(w)=0{\rm Im}(w)=0 at point f(1)f(-1). Now β1\beta_{-1} is positive, if a positive tangent vector to γ1\gamma_{-1} at f(1)f(-1) is rotated counterclockwise to get a positive tangent vector to the real axis and negative otherwise. For j=±1j=\pm 1 we define δj=1βj\delta_{j}=1-\beta_{j}, if βj(0,1)\beta_{j}\in(0,1), and δj=|βj|\delta_{j}=|\beta_{j}|, if βj(1,0)\beta_{j}\in(-1,0).

To transfer our observations to the boundary value problem (12-13) on the unit disc, let ΨA12(Δ)\Psi\in A^{\frac{1}{2}}({\rm\Delta}) be a biholomorphic map from Δ{\rm\Delta} to the upper half-disc Δ+{\rm\Delta}_{+}, which maps the lower semicircle LL on [1,1][-1,1] so that Ψ(±1)=±1\Psi(\pm 1)=\pm 1. Let F(x)=12x(3x2)F(x)=\frac{1}{2}x(3-x^{2}). Then F(x)1=12(x1)2(x+2)F(x)-1=-\frac{1}{2}(x-1)^{2}(x+2) and F(x)+1=12(x+1)2(x2)F(x)+1=-\frac{1}{2}(x+1)^{2}(x-2). Hence function ψ(ξ)=F(Ψ(ξ))\psi(\xi)=F(\Psi(\xi)) is real on LL, ψ(±1)=±1\psi(\pm 1)=\pm 1, and C1C^{1} on Δ\partial{\rm\Delta}.

Recall that wjw_{j} is the positive intersection of γj\gamma_{j} and the real axis, j=±1j=\pm 1. Now we consider only those solutions ff of the Cherepanov problem (12-13), which are of the form

f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ)+w11+ψ(ξ)2+w11ψ(ξ)2,f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}, (48)

where κ\kappa is in A(Δ)A({\rm\Delta}).

We will define two (local) defining functions ρ~1(ξ,w)\widetilde{\rho}_{1}(\xi,w) for ξ1\xi\neq-1 and ρ~1(ξ,w)\widetilde{\rho}_{-1}(\xi,w) for ξ1\xi\neq 1. Let

T1(ξ)=(ξ1)i(ξ+1)andT1(ξ)=1T1(ξ)=i(ξ+1)(ξ1).T_{1}(\xi)=\frac{(\xi-1)}{i(\xi+1)}\ \ \ {\rm and}\ \ \ T_{-1}(\xi)=\frac{1}{T_{1}(\xi)}=\frac{i(\xi+1)}{(\xi-1)}. (49)

Then T1(i)=T1(i)=1T_{1}(-i)=T_{-1}(-i)=-1, and T1,T1T_{1},T_{-1} map the upper semicircle to the positive real axis and the lower semicircle to the negative real axis. For j=±1j=\pm 1 and Im(ξ)>0{\rm Im}(\xi)>0 we define

ρ~j(ξ,w)=1Tj(ξ)δjρ(ξ,(ξ1)δ1(ξ+1)δ1w+w11+ψ(ξ)2+w11ψ(ξ)2)\widetilde{\rho}_{j}(\xi,w)=\frac{1}{T_{j}(\xi)^{\delta_{j}}}\rho\left(\xi,(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}w+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}\right) (50)

and for Im(ξ)<0{\rm Im}(\xi)<0 we set

ρ~j(ξ,w)=Im(eiπ(1βj)(ξ1)δ1(ξ+1)δ1Tj(ξ)δjw).\widetilde{\rho}_{j}(\xi,w)={\rm Im}\left(e^{i\pi(1-\beta_{j})}\frac{(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}}{T_{j}(\xi)^{\delta_{j}}}w\right). (51)

As before one can check that ρ~j\widetilde{\rho}_{j} and ρ~jw\widetilde{\rho}_{jw} are continuous on Δ{j}\partial{\rm\Delta}\setminus\{-j\}, j=±1j=\pm 1. Since ff solves the original boundary value problem, we have that ρ~j(ξ,κ(ξ))=0\widetilde{\rho}_{j}(\xi,\kappa(\xi))=0, j=±1j=\pm 1.

Let χ:Δ{i}[0,1]\chi:\partial{\rm\Delta}\setminus\{-i\}\rightarrow[0,1] be a smooth function such that χ(ξ)=1\chi(\xi)=1 for ξ=eiθ\xi=e^{i\theta}, π2<θπ3-\frac{\pi}{2}<\theta\leq\frac{\pi}{3} and χ(ξ)=0\chi(\xi)=0 for ξ=eiθ\xi=e^{i\theta}, 2π3θ<3π2\frac{2\pi}{3}\leq\theta<\frac{3\pi}{2}.

We define a new (global) defining function as ρ~(ξ,w)=χ(ξ)ρ~1(ξ,w)+(1χ(ξ))ρ~1(ξ,w)\widetilde{\rho}(\xi,w)=\chi(\xi)\widetilde{\rho}_{1}(\xi,w)+(1-\chi(\xi))\widetilde{\rho}_{-1}(\xi,w). Then ρ~\widetilde{\rho} and ρ~w\widetilde{\rho}_{w} are well defined continuous function on (Δ{i})×(\partial{\rm\Delta}\setminus\{-i\})\times{\mathbb{C}}. If β1,β1\beta_{1},\beta_{-1} have the same sign, then both function are also continuous at ξ=i\xi=-i, but if β1,β1\beta_{1},\beta_{-1} have the opposite signs, then

ρ~(i,w)=ρ~(i+,w)andρ~w(i,w)=ρ~w(i+,w),\widetilde{\rho}(-i^{-},w)=-\widetilde{\rho}(-i^{+},w)\ \ {\rm and}\ \ \widetilde{\rho}_{w}(-i^{-},w)=-\widetilde{\rho}_{w}(-i^{+},w), (52)

which means that we have a nonorientable bundle as the boundary value data for κ\kappa.

Now locally considered problem (39-40) for κ(t)\kappa(t) and κ(t)\kappa^{\prime}(t) becomes global boundary value problem for κ(θ)\kappa(\theta) and κθ\frac{\partial\kappa}{\partial\theta} (ξ=eiθ)(\xi=e^{i\theta}). Hence κθ\frac{\partial\kappa}{\partial\theta} solves the linear Riemann-Hilbert problem

2Re(ρ~w(θ,κ(θ))κθ)=ρ~θ(θ,κ(θ)),2{\rm Re}\left(\widetilde{\rho}_{w}(\theta,\kappa(\theta))\frac{\partial\kappa}{\partial\theta}\right)=-\widetilde{\rho}_{\theta}(\theta,\kappa(\theta)), (53)

where ρ~w(θ,κ(θ))\widetilde{\rho}_{w}(\theta,\kappa(\theta)) is either a nonzero continuous function on Δ\partial{\rm\Delta} or

ρ~w(i,κ(i))=ρ~w(i+,κ(i))\widetilde{\rho}_{w}(-i^{-},\kappa(-i))=-\widetilde{\rho}_{w}(-i^{+},\kappa(-i)) (54)

and ρ~θ(θ,κ(θ))\widetilde{\rho}_{\theta}(\theta,\kappa(\theta)) belongs to the appropriate Lp(Δ)L^{p}(\partial{\rm\Delta}) space

1p<min{1|β1|,11|β1|,1|β1|,11|β1|}.1\leq p<\min\left\{\frac{1}{|\beta_{1}|},\frac{1}{1-|\beta_{1}|},\frac{1}{|\beta_{-1}|},\frac{1}{1-|\beta_{-1}|}\right\}. (55)
Remark 3.1

In fact ρ~θ(θ,κ(θ))\widetilde{\rho}_{\theta}(\theta,\kappa(\theta)) belongs to LlocpL^{p}_{\rm loc} for

1p<min{1|β1|,11|β1|}1\leq p<\min\left\{\frac{1}{|\beta_{1}|},\frac{1}{1-|\beta_{1}|}\right\} (56)

near ξ=1\xi=1 and to LlocpL^{p}_{\rm loc} near ξ=1\xi=-1 for

1p<min{1|β1|,11|β1|}.1\leq p<\min\left\{\frac{1}{|\beta_{-1}|},\frac{1}{1-|\beta_{-1}|}\right\}. (57)

Let NN be the winding number of function ρ~w(θ,κ(θ))\widetilde{\rho}_{w}(\theta,\kappa(\theta)), that is, 2N2N is the Maslov index of the associated linear Riemann-Hilbert problem. If ρ~w(θ,κ(θ))\widetilde{\rho}_{w}(\theta,\kappa(\theta)) is a continuous function on Δ\partial{\rm\Delta}, Maslov index is an even integer and hence NN is an integer. On the other hand, if ρ~w(i,κ(i))=ρ~w(i+,κ(i))\widetilde{\rho}_{w}(-i^{-},\kappa(-i))=-\widetilde{\rho}_{w}(-i^{+},\kappa(-i)), Maslov index is an odd integer and NN is a half of an odd integer.

Let r(ξ)r(\xi) be the square root function, where we take the branch where {\mathbb{C}} is cut along the negative imaginary axis. Then function ρ~w(θ,κ(θ))\widetilde{\rho}_{w}(\theta,\kappa(\theta)) can be written in the form

ρ~w(θ,κ(θ))=ξNeu+iv(θ),\widetilde{\rho}_{w}(\theta,\kappa(\theta))=\xi^{-N}e^{u+iv}(\theta), (58)

where uu and vv are real continuous functions on Δ\partial{\rm\Delta}, [18, p. 25]. In the case N=2M+12N=\frac{2M+1}{2}, MM\in{\mathbb{Z}}, is a half of an odd integer, we define ξN=ξMr(ξ)\xi^{N}=\xi^{M}r(\xi), which corresponds to the sign changing of ρ~w\widetilde{\rho}_{w} at ξ=i\xi=-i. See also [3, 4, 15]. Hence e±Hve^{\pm Hv} belongs to Lp(Δ)L^{p^{\prime}}(\partial{\rm\Delta}) for any p1p^{\prime}\geq 1, [18, p. 23] and thus

e±i(v+iHv)e^{\pm i(v+iHv)} (59)

belongs to Lp(Δ)L^{p^{\prime}}(\partial{\rm\Delta}) for any p1p^{\prime}\geq 1.

Therefore

Re(ξNei(v+iHv)κθ)=eue(Hv)ρ~θ(θ,κ).{\rm Re}\left(\xi^{-N}e^{i(v+iHv)}\frac{\partial\kappa}{\partial\theta}\right)=-e^{-u}e^{-(Hv)}\widetilde{\rho}_{\theta}(\theta,\kappa). (60)

We conclude that the right-hand side belongs to the same Lp(Δ)L^{p}(\partial{\rm\Delta}) space as function ρ~θ(θ,κ)\widetilde{\rho}_{\theta}(\theta,\kappa). Since Hilbert transform is bounded in Lp(Δ)L^{p}(\partial{\rm\Delta}) spaces, 1<p<1<p<\infty, [18, p. 23], we get that κθ\frac{\partial\kappa}{\partial\theta} is in Lp(Δ)L^{p}(\partial{\rm\Delta}) for the same set (55) of values of pp as function ρ~θ(θ,κ)\widetilde{\rho}_{\theta}(\theta,\kappa). Therefore κ\kappa belongs to L1,p(Δ)L^{1,p}(\partial{\rm\Delta}) for all such values of pp and this implies that κCβ(Δ)\kappa\in C^{\beta}(\partial{\rm\Delta}), [18, p. 10], where

0<β<min{|β1|,1|β1|,|β1|,1|β1|}.0<\beta<\min\{|\beta_{1}|,1-|\beta_{1}|,|\beta_{-1}|,1-|\beta_{-1}|\}. (61)
Remark 3.2

Observe that regularity of κ\kappa and ff could also be expressed locally, that is, near j=±1j=\pm 1 functions κ\kappa and ff belong to Hölder space CβC^{\beta}, where 0<β<min{|βj|,1|βj|}0<\beta<\min\{|\beta_{j}|,1-|\beta_{j}|\}.

Proposition 3.3

Let k3k\geq 3. Let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a CkC^{k} family of Jordan curves in {\mathbb{C}}. Let wjw_{j}, j=±1j=\pm 1, be an intersection of γj\gamma_{j} and the real axis and let πβj\pi\beta_{j}, βj(1,1){0}\beta_{j}\in(-1,1)\setminus\{0\}, be the oriented angle of intersection of γj\gamma_{j} with the real axis at point wjw_{j}. Let

0<β<min{|β1|,1|β1|,|β1|,1|β1|}.0<\beta<\min\{|\beta_{1}|,1-|\beta_{1}|,|\beta_{-1}|,1-|\beta_{-1}|\}. (62)

Then for every solution ff of (12-13) of the form

f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ)+w11+ψ(ξ)2+w11ψ(ξ)2,f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}, (63)

where κA(Δ)\kappa\in A({\rm\Delta}), we have f,κAβ(Δ)f,\kappa\in A^{\beta}({\rm\Delta}).

Remark 3.4

Observe that in cases where β1,β1(0,1)\beta_{1},\beta_{-1}\in(0,1), 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 ρ\rho is independent of ξ\xi and βj(0,1)\beta_{j}\in(0,1), we get (1βj)(1-\beta_{j})-regularity. The βj\beta_{j}-regularity comes from ξ\xi-dependence.

Similarly, the expected regularity and the ’order’ of zeros of Riemann maps in the cases where βj(1,0)\beta_{j}\in(-1,0) and which are ξ\xi independent, would be 1+|βj|1+|\beta_{j}|, but ξ\xi-dependence of the defining function ρ\rho changes regularity conditions.

On the other hand, results in [10] show that in the case of nontransversal intersection of the real axis with either γ1\gamma_{1} or γ1\gamma_{-1} solutions might not be of the form (ξ1)δ1κ(ξ)(\xi-1)^{\delta_{1}}\kappa(\xi) or (ξ+1)δ1κ(ξ)(\xi+1)^{\delta_{-1}}\kappa(\xi) for some function κA(Δ)\kappa\in A({\rm\Delta}).

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 LL the lower semicircle. First we consider homogeneous linear problem with piecewise continuous boundary data

Im(f(ξ))=0forξL{\rm Im}(f(\xi))=0{\rm\ for\ }\xi\in L (64)

and

Re(B(ξ)¯f(ξ))=0forξΔL̊,{\rm Re}(\overline{B(\xi)}f(\xi))=0{\rm\ for\ }\xi\in\partial{\rm\Delta}\setminus\mathring{L}, (65)

where BB is a complex nonzero function of class CβC^{\beta} on the upper semicircle. The regularity exponent β(0,1)\beta\in(0,1) is bounded by conditions given in Proposition 3.3. We may assume without loss of generality that |B(ξ)|=1|B(\xi)|=1 for all ξΔL̊\xi\in\partial{\rm\Delta}\setminus\mathring{L}.

Let πβ1\pi\beta_{1}, β1(1,1){0}\beta_{1}\in(-1,1)\setminus\{0\}, be the oriented angle of intersection of the real axis Im(w)=0{\rm Im}(w)=0 and Re(B(1)¯w)=0{\rm Re}(\overline{B(1)}w)=0 at point 0, that is, B(1)=ieiπβ1B(1)=-ie^{i\pi\beta_{1}}. Similarly, let πβ1\pi\beta_{-1}, β1(1,1){0}\beta_{-1}\in(-1,1)\setminus\{0\}, be the oriented angle of intersection of Re(B(1)¯w)=0{\rm Re}(\overline{B(-1)}w)=0 and the real axis Im(w)=0{\rm Im}(w)=0 at point 0, that is, B(1)=ieiπβ1B(-1)=-ie^{-i\pi\beta_{-1}}.

We search for solutions fA(Δ)f\in A({\rm\Delta}) of (64-65) of the form f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ)f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi) for some κAβ(Δ)\kappa\in A^{\beta}({\rm\Delta}). Recall that for j=±1j=\pm 1 we defined δj=1βj\delta_{j}=1-\beta_{j}, if βj(0,1)\beta_{j}\in(0,1), and δj=|βj|\delta_{j}=|\beta_{j}|, if βj(1,0)\beta_{j}\in(-1,0). Hence we also have fAβ(Δ)f\in A^{\beta}({\rm\Delta}).

To define noninteger powers of (ξ1)(\xi-1) and (ξ+1)(\xi+1) we take appropriate branches of the complex logarithm. For (ξ1)δ1(\xi-1)^{\delta_{1}} the complex plane is cut along positive real numbers so that the argument of (ξ1)(\xi-1) for ξΔ\xi\in\partial{\rm\Delta} lies on interval (π2,3π2)(\frac{\pi}{2},\frac{3\pi}{2}), and for (ξ+1)δ1(\xi+1)^{\delta_{-1}} the complex plane is cut along negative real numbers and the argument of (ξ+1)(\xi+1) for ξΔ\xi\in\partial{\rm\Delta} lies on interval (π2,π2)(-\frac{\pi}{2},\frac{\pi}{2}).

An argument similar to the argument in Section 3 shows that κ\kappa solves homogeneous linear Riemann-Hilbert problem

Re(B~(ξ)¯κ(ξ))=0forallξΔ,{\rm Re}(\overline{\widetilde{B}(\xi)}\kappa(\xi))=0\ {\rm for\ all\ }\xi\in\partial{\rm\Delta}, (66)

where B~¯Cβ(D{1})\overline{\widetilde{B}}\in C^{\beta}(\partial D\setminus\{1\}) is defined as

B~(ξ)¯={B(ξ)¯(ξ1|ξ1|)δ1(ξ+1|ξ+1|)δ1,ifIm(ξ)>0±i(ξ1|ξ1|)δ1(ξ+1|ξ+1|)δ1,ifIm(ξ)<0\overline{\widetilde{B}(\xi)}=\left\{\begin{array}[]{rl}\overline{B(\xi)}\left(\frac{\xi-1}{|\xi-1|}\right)^{\delta_{1}}\left(\frac{\xi+1}{|\xi+1|}\right)^{\delta_{-1}},&{\rm if\ }{\rm Im}(\xi)>0\\ \pm i\left(\frac{\xi-1}{|\xi-1|}\right)^{\delta_{1}}\left(\frac{\xi+1}{|\xi+1|}\right)^{\delta_{-1}},&{\rm if\ }{\rm Im}(\xi)<0\end{array}\right. (67)

with the left and the right limits at ξ=±1\xi=\pm 1. The sign for Im(ξ)<0{\rm Im}(\xi)<0 is chosen so that B~\widetilde{B} is continuous at 1-1, that is, we have plus sign, if β1<0\beta_{-1}<0, and minus sign, if β1>0\beta_{-1}>0. At point ξ=1\xi=1 function BB might not be continuous. In general we have B~(1+)=±B~(1)\widetilde{B}(1^{+})=\pm\widetilde{B}(1^{-}). See [19, p. 169-170] for more.

Each factor

ξ1|ξ1|,ξ+1|ξ+1|\frac{\xi-1}{|\xi-1|},\frac{\xi+1}{|\xi+1|} (68)

changes the argument by π\pi when ξ\xi passes Δ\partial{\rm\Delta} once in the positive direction. Hence possible widing number of B~\widetilde{B} 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 B(eiθ)=eiθB(e^{i\theta})=e^{i\theta} for θ[0,π]\theta\in[0,\pi]. In particular we have β1=β1=12\beta_{1}=\beta_{-1}=\frac{1}{2} . Then we get

B~(ξ)={eiπ4B(ξ)ξ¯12=ei2θπ4,if 0θπei3π4ξ¯12=ei3π2θ4,ifπ<θ<2π.\widetilde{B}(\xi)=\left\{\begin{array}[]{rl}e^{-i\frac{\pi}{4}}B(\xi)\,\overline{\xi}^{\frac{1}{2}}=e^{i\frac{2\theta-\pi}{4}},&{\rm if\ }0\leq\theta\leq\pi\\ e^{i\frac{3\pi}{4}}\,\overline{\xi}^{\frac{1}{2}}=e^{i\frac{3\pi-2\theta}{4}},&{\rm if\ }\pi<\theta<2\pi.\end{array}\right. (69)

Hence the winding number W(B~)=0W(\widetilde{B})=0. 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 f(z)=zf(z)=z. The family of (nearby) solutions which are real on the real axis is one-dimensional fa(z)=za1azf_{a}(z)=\frac{z-a}{1-az}, where a(1,1)a\in(-1,1) is a real number.

Example 4.2

Consider the case B(eiθ)=1B(e^{i\theta})=1 for θ[0,π]\theta\in[0,\pi]. In particular we have β1=β1=12\beta_{1}=-\beta_{-1}=\frac{1}{2} . Then we get

B~(ξ)=eiπ+2θ4\widetilde{B}(\xi)=e^{-i\frac{\pi+2\theta}{4}} (70)

and the winding number W(B~)=12W(\widetilde{B})=-\frac{1}{2}. 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 f(z)=1f(z)=1. The family of (nearby) solutions which are real on the real axis is zero-dimensional.

The dimension of the space of solutions in Aβ(Δ)A^{\beta}({\rm\Delta}) depends on the winding number W(B~)W(\widetilde{B}) of function B~\widetilde{B}. It equals 2W(B~)+12W(\widetilde{B})+1 if W(B~)12W(\widetilde{B})\geq-\frac{1}{2}, see [18, p. 25, p. 59] and [3, 4, 15]. We define the winding number of BCα(ΔL̊)B\in C^{\alpha}(\partial{\rm\Delta}\setminus\mathring{L}) as the winding number of B~\widetilde{B}.

Now we can solve appropriate nonhomogeneous linear Riemann-Hilbert problem with piecewise continuous boundary data

Im(f(ξ))=0forξL{\rm Im}(f(\xi))=0{\rm\ for\ }\xi\in L (71)

and

Re(B(ξ)¯f(ξ))=b(ξ)forξΔL̊,{\rm Re}(\overline{B(\xi)}f(\xi))=b(\xi){\rm\ for\ }\xi\in\partial{\rm\Delta}\setminus\mathring{L}, (72)

where BB is as above and bb a real function on Δ\partial{\rm\Delta} of the form

b(ξ)=|ξ1|δ1|ξ+1|δ1b~(ξ)b(\xi)=|\xi-1|^{\delta_{1}}|\xi+1|^{\delta_{-1}}\widetilde{b}(\xi) (73)

for some function b~Cβ(Δ)\widetilde{b}\in C_{{\mathbb{R}}}^{\beta}(\partial{\rm\Delta}) which equals 0 on LL.

To solve (71-72) in the space of functions fAβ(D)f\in A^{\beta}(D) of the form f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ)f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi) for some κAβ(Δ)\kappa\in A^{\beta}({\rm\Delta}) is equivalent to solve the problem

Re(B~¯(ξ)κ(ξ))=b~(ξ)forallξΔ.{\rm Re}(\overline{\widetilde{B}}(\xi)\kappa(\xi))=\widetilde{b}(\xi)\ {\rm for\ all\ }\xi\in\partial{\rm\Delta}. (74)

It is well known that if W(B)=W(B~)12W(B)=W(\widetilde{B})\geq-\frac{1}{2}, then the equation is solvable for any b~Cβ(Δ)\widetilde{b}\in C_{{\mathbb{R}}}^{\beta}(\partial{\rm\Delta}), see [18, p. 25, p. 59] and [3, 4, 15].

Remark 4.3

If the winding number W(B)=W(B~)W(B)=W(\widetilde{B}) 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 Δ{1}\partial{\rm\Delta}\setminus\{1\} of the form b0(r(ξ))b_{0}(r(\xi)), where r(ξ)r(\xi) is the principal branch of the square root and b0Cβ(Δ)b_{0}\in C^{\beta}_{{\mathbb{R}}}(\partial{\rm\Delta}) is an odd function. Hence we need condition b~(1)+b~(1+)=0\widetilde{b}(1^{-})+\widetilde{b}(1^{+})=0, which is satisfied because in our case we have b~(1)=b~(1+)=0\widetilde{b}(1^{-})=\widetilde{b}(1^{+})=0. See [3, 4] for more information.

Proposition 4.4

Let 0<β<10<\beta<1. Let B:ΔL̊{0}B:\partial{\rm\Delta}\setminus\mathring{L}\rightarrow{\mathbb{C}}\setminus\{0\} be a non-vanishing complex function in Cβ(ΔL̊)C^{\beta}(\partial{\rm\Delta}\setminus\mathring{L}) and let W(B)12W(B)\geq-\frac{1}{2}. Then for every real function bb on Δ\partial{\rm\Delta} of the form

b(ξ)=|ξ1|δ1|ξ+1|δ1b~(ξ)b(\xi)=|\xi-1|^{\delta_{1}}|\xi+1|^{\delta_{-1}}\widetilde{b}(\xi) (75)

for some b~Cβ(Δ)\widetilde{b}\in C^{\beta}_{{\mathbb{R}}}(\partial{\rm\Delta}) which equals 0 on LL, there exists a solution ff of the linear Cherepanov problem

Im(f(ξ))=0forξL{\rm Im}(f(\xi))=0{\rm\ for\ }\xi\in L (76)

and

Re(B(ξ)¯f(ξ))=b(ξ)forξΔL̊{\rm Re}(\overline{B(\xi)}f(\xi))=b(\xi){\rm\ for\ }\xi\in\partial{\rm\Delta}\setminus\mathring{L} (77)

of the form

f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ),f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi), (78)

where κAβ(Δ)\kappa\in A^{\beta}({\rm\Delta}). Moreover, the space of solutions of this form is 2W(B)+12W(B)+1 dimensional real subspace of Aβ(Δ)A^{\beta}({\rm\Delta}).

Proposition 4.5

Let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a CkC^{k} (k3)(k\geq 3) family of Jordan curves in {\mathbb{C}} and let ρ0Ck((ΔL̊)×)\rho_{0}\in C^{k}((\partial{\rm\Delta}\setminus\mathring{L})\times{\mathbb{C}}) be its defining function. Let β1,β1\beta_{1},\beta_{-1} and β\beta be as in Proposition 3.3. Let f0f_{0} be a solution of the Cherepanov problem (12), (13) of the form

f0(ξ)=(ξ1)δ1(ξ+1)δ1κ0(ξ)+w11+ψ(ξ)2+w11ψ(ξ)2,f_{0}(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa_{0}(\xi)+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}, (79)

where κ0Aβ(Δ)\kappa_{0}\in A^{\beta}({\rm\Delta}). Then the mapping Φ(κ):Aβ(Δ)Cβ(Δ)\Phi(\kappa):A^{\beta}({\rm\Delta})\rightarrow C^{\beta}_{{\mathbb{R}}}(\partial{\rm\Delta}), for each κ\kappa evaluated at point ξΔ\xi\in\partial{\rm\Delta} as

{ρ0(ξ,(ξ1)δ1(ξ+1)δ1κ(ξ)+w11+ψ(ξ)2+w11ψ(ξ)2),ifIm(ξ)0,±Im((ξ1)δ1(ξ+1)δ1κ(ξ)+w11+ψ(ξ)2+w11ψ(ξ)2),ifIm(ξ)<0\left\{\begin{array}[]{ll}\rho_{0}(\xi,(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}),&{\rm if\ }{\rm Im}(\xi)\geq 0,\\ \pm{\rm Im}((\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)+w_{1}\frac{1+\psi(\xi)}{2}+w_{-1}\frac{1-\psi(\xi)}{2}),&{\rm if\ }{\rm Im}(\xi)<0\end{array}\right. (80)

is differentiable at κ0\kappa_{0} with the derivative (DΦ)(κ0)(D\Phi)(\kappa_{0}) acting on κAβ(Δ)\kappa\in A^{\beta}({\rm\Delta}) as

{2Re(ρ0w(ξ,f0(ξ))(ξ1)δ1(ξ+1)δ1κ(ξ)),ifIm(ξ)0,±Im((ξ1)δ1(ξ+1)δ1κ(ξ)),ifIm(ξ)<0.\left\{\begin{array}[]{ll}2{\rm Re}(\partial\rho_{0w}(\xi,f_{0}(\xi))(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)),&{\rm if\ }{\rm Im}(\xi)\geq 0,\\ \pm{\rm Im}((\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)),&{\rm if\ }{\rm Im}(\xi)<0.\end{array}\right. (81)

The sign for Im(ξ)<0{\rm Im}(\xi)<0 is chosen as in (66-67).

Remark 4.6

Let ΩCk+1((ΔL̊)×)\Omega\subset C^{k+1}((\partial{\rm\Delta}\setminus{\mathring{L}})\times{\mathbb{C}}) be an open subset of defining functions ρ\rho of the families of Jordan curves over ΔL̊\partial{\rm\Delta}\setminus{\mathring{L}} such that the intersection of the corresponding γ1\gamma_{1} and γ1\gamma_{-1} with the real axis at some points w1γ1w_{1}\in\gamma_{1} and w1γ1w_{-1}\in\gamma_{-1} are transversal with the oriented angles of intersection given by β1,β1(1,1){0}\beta_{1},\beta_{-1}\in(-1,1)\setminus\{0\}. Then, at least locally, w1,w1w_{1},w_{-1} and β1,β1\beta_{1},\beta_{-1} smoothly depend on ρ\rho. Let

X={(κ,ρ)Aβ(Δ)×Ω;Im((ξ1)δ1(ξ+1)δ1κ(ξ))=0,ifIm(ξ)<0}X=\{(\kappa,\rho)\in A^{\beta}({\rm\Delta})\times\Omega;{\rm Im}((\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi))=0,{\ \rm if\ }{\rm Im}(\xi)<0\} (82)

which is a Banach submanifold of Aβ(Δ)×ΩA^{\beta}({\rm\Delta})\times\Omega. Also, let

Y={b(ξ)=|ξ1|δ1|ξ+1|δ1b~(ξ);b~Cβ(Δ),b~(ξ)=0,ifIm(ξ)<0}.Y=\{b(\xi)=|\xi-1|^{\delta_{1}}|\xi+1|^{\delta_{-1}}\widetilde{b}(\xi);\widetilde{b}\in C^{\beta}_{{\mathbb{R}}}({\rm\Delta}),\widetilde{b}(\xi)=0,{\ \rm if\ }{\rm Im}(\xi)<0\}. (83)

The mapping Φ:XY\Phi:X\rightarrow Y defined as in (80) has partial derivative with respect to κ\kappa as a map from

Xρ={κAβ(Δ);Im((ξ1)δ1(ξ+1)δ1κ(ξ))=0,ifIm(ξ)<0}X_{\rho}=\{\kappa\in A^{\beta}({\rm\Delta});{\rm Im}((\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi))=0,{\ \rm if\ }{\rm Im}(\xi)<0\} (84)

to YY of the form (81). If the winding number W(B)W(B) of the Cherepanov problem defined by (81) is greater or equal to 12-\frac{1}{2}, then the partial derivative is surjective with 2W(B)+12W(B)+1 dimensional kernel. Hence implicit function theorem applies and there is a neighbourhood of ρ0\rho_{0} in Ω\Omega and a neighbourhood of κ0\kappa_{0} in Aβ(Δ)A^{\beta}({\rm\Delta}) such that for every ρΩ\rho\in\Omega close to ρ0\rho_{0} there is a 2W(B)+12W(B)+1 dimensional family of solutions of (12-13) near κ0\kappa_{0}.

5 A priori estimates

5.1 A priori estimates on function ff

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 Δ{\rm\Delta}, [8, 16, 18]. We follow the approach in [8].

By assumption all Jordan curves {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} contain point 0 in their interiors. Hence the function

(θ,w)wρw(θ,w),(\theta,w)\longmapsto w\rho_{w}(\theta,w), (85)

defined for (θ,w)(\theta,w) such that wγθw\in\gamma_{\theta}, is homotopic to 0 in {0}{\mathbb{C}}\setminus\{0\} and it can be written in the form

wρw(θ,w)=ec(θ,w)+id(θ,w)w\rho_{w}(\theta,w)=e^{c(\theta,w)+id(\theta,w)} (86)

for some Ck1C^{k-1} functions cc and dd, defined for (θ,w)(\theta,w) such that θ[0,π]\theta\in[0,\pi] and wγθw\in\gamma_{\theta}. Observe that for each wγθw\in\gamma_{\theta} function d(θ,w)d(\theta,w) represents the angle between ww and the outer normal to γθ\gamma_{\theta} at point ww.

Remark 5.1

There exists a CkC^{k} isotopy ρt\rho^{t}, t[0,1]t\in[0,1], where ρ0=ρ\rho^{0}=\rho and ρ1(ξ,w)=|w|2R2\rho^{1}(\xi,w)=|w|^{2}-R^{2} for R>0R>0 large enough, such that the gradient ρw¯t\rho^{t}_{\overline{w}} is nonzero on ρt=0\rho^{t}=0 for each tt, [8]. Then one can find Ck1C^{k-1} functions c(t,θ,w)c(t,\theta,w) and d(t,θ,w)d(t,\theta,w) such that (86) holds for each t[0,1]t\in[0,1] and wγθtw\in\gamma^{t}_{\theta}. In addition, the isotopy can be made such that for every t[0,1]t\in[0,1], j=±1j=\pm 1, Jordan curves γjt\gamma^{t}_{j} are strongly starshaped with respect to 0 and that for each wγωjtw\in\gamma^{t}_{\omega_{j}} the angle between ww and the normal to γωjt\gamma^{t}_{\omega_{j}} at ww is less than π10\frac{\pi}{10}.

Instead of solving (12-13) on the unit disc we consider equivalent problem on the upper semidisc Δ+={zΔ;Im(z)>0}{\rm\Delta}^{+}=\{z\in{\rm\Delta};{\rm Im}(z)>0\}, where the role of the lower semicircle LL is replaced by the interval [1,1][-1,1]. Using the reflection principle f(ξ)=f(ξ¯)¯)f(\xi)=\overline{f(\overline{\xi})}) we can holomorphically extend every solution ff of (12-13) to the unit disc such that it solves nonlinear Riemann-Hilbert problem defined by the function ρ\rho which we get as an extension of the original function ρ\rho using the reflection to the lower semicircle as

ρ(ξ,w)=ρ(ξ¯,w¯)forξ±1.\rho(\xi,w)=\rho(\overline{\xi},\overline{w}){\rm\ for\ }\xi\neq\pm 1. (87)

For ξ=±1\xi=\pm 1 function ρ(ξ,w)\rho(\xi,w) has well defined limits as ξ\xi approaches ±1\pm 1 from above and below. Then we have

ρw(ξ,w)=ρw¯(ξ¯,w¯)=ρw(ξ¯,w¯)¯forξ±1\rho_{w}(\xi,w)=\rho_{\overline{w}}(\overline{\xi},\overline{w})=\overline{\rho_{w}(\overline{\xi},\overline{w})}{\rm\ for\ }\xi\neq\pm 1 (88)

and hence

ρw(ξ,w)w=ρw¯(ξ¯,w¯)w=ρw(ξ¯,w¯)w¯¯.\rho_{w}(\xi,w)w=\rho_{\overline{w}}(\overline{\xi},\overline{w})w=\overline{\rho_{w}(\overline{\xi},\overline{w})\overline{w}}. (89)

Therefore c(ξ¯,w¯)=c(ξ,w)c(\overline{\xi},\overline{w})=c(\xi,w) and d(ξ¯,w¯)=d(ξ,w)d(\overline{\xi},\overline{w})=-d(\xi,w). Also, observe that for ww, an intersection of γ1\gamma_{1} with the real axis, we have

ρw(1+,w)=ρw(1,w¯)¯=ρw(1,w)¯\rho_{w}(1+,w)=\overline{\rho_{w}(1-,\overline{w})}=\overline{\rho_{w}(1-,w)} (90)

and similarly for an intersection of γ1\gamma_{-1} with the real axis.

Thus for every solution ff of (12-13) the absolute value of f(θ)ρw(θ,f(θ))f(\theta)\rho_{w}(\theta,f(\theta)) is well defined and continuous on Δ\partial{\rm\Delta}, whereas

d(0+,f(0+))=d(2π,f(2π))d(0+,f(0+))=-d(2\pi-,f(2\pi-)) (91)

and similarly at θ=π\theta=\pi.

Let ff be a solution of the symmetrized boundary value problem with no zeros. Hence ff can be written in the exponential form

f=eg.f=e^{g}. (92)
Remark 5.2

Since the biholomorphic map ψ\psi from Δ{\rm\Delta} to the upper half-disc Δ+{\rm\Delta}_{+} is of class C12C^{\frac{1}{2}}, a CβC^{\beta} estimate on solutions of the symmetrized boundary value problem gives Cβ2C^{\frac{\beta}{2}} estimate on solutions of (12-13).

Let us differentiate function ρ(θ,f(θ))\rho(\theta,f(\theta)) to get

ρθ(θ,f(θ))+2Re(ρw(θ,f(θ))fθ(θ))=0.\rho_{\theta}(\theta,f(\theta))+2{\rm Re}(\rho_{w}(\theta,f(\theta))\frac{\partial f}{\partial\theta}(\theta))=0. (93)

Since f=egf=e^{g}, we get

ρθ(θ,f(θ))+2Re(ρw(θ,f(θ))f(θ)gθ(θ))=0\rho_{\theta}(\theta,f(\theta))+2{\rm Re}(\rho_{w}(\theta,f(\theta))f(\theta)\frac{\partial g}{\partial\theta}(\theta))=0 (94)

and so

ρθ(θ,f(θ))+2Re(ec(θ,f(θ))+id(θ,f(θ))gθ(θ))=0.\rho_{\theta}(\theta,f(\theta))+2{\rm Re}(e^{c(\theta,f(\theta))+id(\theta,f(\theta))}\frac{\partial g}{\partial\theta}(\theta))=0. (95)

From here we get

2Re(ei(d(θ,f(θ))+iHd(θ,f(θ)))gθ(θ))=ρθ(θ,f(θ))ec(θ,f(θ))Hd(θ,f(θ)).2{\rm Re}(e^{i(d(\theta,f(\theta))+iHd(\theta,f(\theta)))}\frac{\partial g}{\partial\theta}(\theta))=-\rho_{\theta}(\theta,f(\theta))e^{-c(\theta,f(\theta))-Hd(\theta,f(\theta))}. (96)

Observe that function

θei(d(θ,f(θ))+iHd(θ,f(θ)))gθ(θ)\theta\longmapsto e^{i(d(\theta,f(\theta))+iHd(\theta,f(\theta)))}\frac{\partial g}{\partial\theta}(\theta) (97)

extends holomorphically to the unit disc with value 0 at 0.

We will get CβC^{\beta} a priori estimates on gg and hence on ff by getting CβC^{\beta} a priori estimates on function (97). Using Hilbert transform it is enough to get CβC^{\beta} a priori estimates on its real part. Hence we need a priori estimates on the right hand side of (96).

Function

θρθ(θ,f(θ))ec(θ,f(θ))\theta\longmapsto-\rho_{\theta}(\theta,f(\theta))e^{-c(\theta,f(\theta))} (98)

is bounded with the bound which does not depend on function ff but only on the data γξΔL̊\gamma_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} and defining function ρ\rho. The bound can also be found to be independent of the CkC^{k} isotopy ρt\rho^{t}, t[0,1]t\in[0,1]. Hence one needs a priori bound on function

θe±Hd(θ,f(θ)).\theta\longmapsto e^{\pm Hd(\theta,f(\theta))}. (99)

Recall that, [18, p. 23], for uL(Δ)u\in L^{\infty}(\partial{\rm\Delta}), such that u<π2p\|u\|_{\infty}<\frac{\pi}{2p} (1p<)(1\leq p<\infty) we have the estimate

eHup(2πcos(pu))1p.\|e^{Hu}\|_{p}\leq\left(\frac{2\pi}{\cos(p\|u\|_{\infty})}\right)^{\frac{1}{p}}. (100)

Let a(0,π5)a\in(0,\frac{\pi}{5}) and let χ0,χπ\chi_{0},\chi_{\pi} be smooth functions on [0,π][0,\pi] with values in [0,1][0,1] such that χ0(t)=1\chi_{0}(t)=1 on [0,a][0,a], χπ(t)=1\chi_{\pi}(t)=1 on [πa,π][\pi-a,\pi], χ0(t)=0\chi_{0}(t)=0 on [2a,π][2a,\pi], and χπ(t)=0\chi_{\pi}(t)=0 on [0,π2a][0,\pi-2a].

Let us consider the function

d~(θ,w)=d(θ,w)χ0(θ)d0(w)χπ(θ)dπ(w)\widetilde{d}(\theta,w)=d(\theta,w)-\chi_{0}(\theta)d_{0}(w)-\chi_{\pi}(\theta)d_{\pi}(w) (101)

for θ[0,π]\theta\in[0,\pi] and d~(θ,w)=d~(2πθ,w¯)\widetilde{d}(\theta,w)=-\widetilde{d}(2\pi-\theta,\overline{w}) for θ[π,2π]\theta\in[\pi,2\pi]. Here we used notation d0(w)=d(0+,w)d_{0}(w)=d(0+,w) and dπ(w)=d(π,w)d_{\pi}(w)=d(\pi-,w).

We see that d~(0,w)=d~(π,w)=0\widetilde{d}(0,w)=\widetilde{d}(\pi,w)=0 and so d~(θ,w)\widetilde{d}(\theta,w) is a continuous function on Δ×\partial{\rm\Delta}\times{\mathbb{C}}. Let 1<p~<1<\widetilde{p}<\infty be given. By results from [8, p. 881] we can write d~=Re(q)+e~\widetilde{d}={\rm Re}(q)+\widetilde{e}, where p~e~<π2\widetilde{p}\|\widetilde{e}\|_{\infty}<\frac{\pi}{2} and qq is a finite sum of terms of the form eijθwme^{ij\theta}w^{m}, jj\in{\mathbb{Z}}, m{0}m\in{\mathbb{N}}\cup\{0\}, on which Hilbert transform acts as a bounded nonlinear operator from A(Δ)A(\partial{\rm\Delta}) into C(Δ)C(\partial{\rm\Delta}).

Therefore for a given solution ff of (12-13) with no zeros we can write continuous function d~(θ,f(θ))\widetilde{d}(\theta,f(\theta)) on Δ\partial{\rm\Delta} in the form

d~(θ,f(θ))=Re(q(θ,f(θ)))+e~(θ,f(θ))\widetilde{d}(\theta,f(\theta))={\rm Re}(q(\theta,f(\theta)))+\widetilde{e}(\theta,f(\theta)) (102)

and so

H(d~)=H(Re(q))+H(e~)H(\widetilde{d})=H({\rm Re}(q))+H(\widetilde{e}) (103)

where the first term is uniformly bounded and for the second we have e~<π2p~\|\widetilde{e}\|_{\infty}<\frac{\pi}{2\widetilde{p}}. Hence

e±Hd~=e±HRe(q)e±He~e^{\pm H\widetilde{d}}=e^{\pm H{\rm Re}(q)}e^{\pm H\widetilde{e}} (104)

where the first factor is uniformly bounded and the second factor is bounded in Lp~(Δ)L^{\widetilde{p}}(\partial{\rm\Delta}) for a given 1<p~<1<\widetilde{p}<\infty.

Since for a given p~(1,)\widetilde{p}\in(1,\infty) we can get Lp~(Δ)L^{\widetilde{p}}(\partial{\rm\Delta}) bounds on (104), the boundedness of e±Hde^{\pm Hd} in some Lp(Δ)L^{p}(\partial{\rm\Delta}) is determined by Hilbert transform of the extension of function χ0(θ)d0(f(θ))+χπ(θ)dπ(f(θ))\chi_{0}(\theta)d_{0}(f(\theta))+\chi_{\pi}(\theta)d_{\pi}(f(\theta)) to [0,2π][0,2\pi].

Recall that γ±1\gamma_{\pm 1} are strongly starshaped Jordan curves with respect to 0 and we may assume that for j=±1j=\pm 1 we have

ρ(j,w)=|w|2Rj2(w|w|)\rho(j,w)=|w|^{2}-R_{j}^{2}\left(\frac{w}{|w|}\right) (105)

for some positive CkC^{k} function Rj(z)R_{j}(z) on {\mathbb{C}}. A short calculation gives

ρw(j,w)=w¯2Rj(12w¯2|w|3(Rj)z¯+121|w|(Rj)z)\rho_{w}(j,w)=\overline{w}-2R_{j}\left(-\frac{1}{2}\frac{\overline{w}^{2}}{|w|^{3}}(R_{j})_{\overline{z}}+\frac{1}{2}\frac{1}{|w|}(R_{j})_{z}\right) (106)

and so

wρw(j,w)=|w|22iRj|w|Im(w(Rj)z)w\rho_{w}(j,w)=|w|^{2}-2i\frac{R_{j}}{|w|}{\rm Im}(w(R_{j})_{z}) (107)

which has strictly positive real part on {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}}. Functions d0d_{0} and dπd_{\pi} represent the argument of (107). By compactness it follows that there exists 0<β0<10<\beta_{0}<1, such that |d0(w)|π2β0|d_{0}(w)|\leq\frac{\pi}{2}\beta_{0} and |dπ(w)|π2β0|d_{\pi}(w)|\leq\frac{\pi}{2}\beta_{0} on {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} and therefore

|χ0(θ)d0(f(θ))+χπ(θ)dπ(f(θ))|π2β0|\chi_{0}(\theta)d_{0}(f(\theta))+\chi_{\pi}(\theta)d_{\pi}(f(\theta))|\leq\frac{\pi}{2}\beta_{0} (108)

for every θ\theta. Also, if there is an open condition on the size of djπd_{j\pi} on γj\gamma_{j}, such as |diπ(w)|<π2β0|d_{i\pi}(w)|<\frac{\pi}{2}\beta_{0}, j=±1j=\pm 1, we can, by choosing the supports of functions χ0\chi_{0} and χπ\chi_{\pi} small enough, that is, by choosing a>0a>0 small enough, assume that the same condition on the size holds for function χ0(θ)d0(f(θ))+χπ(θ)dπ(f(θ))\chi_{0}(\theta)d_{0}(f(\theta))+\chi_{\pi}(\theta)d_{\pi}(f(\theta)) for all θ\theta. Observe also that |d0(f(0))|=π|β112||d_{0}(f(0))|=\pi|\beta_{1}-\frac{1}{2}| and |dπ(f(π))|=π||β1|12||d_{\pi}(f(\pi))|=\pi||\beta_{-1}|-\frac{1}{2}|. and so

||βj|12|<β02,j=±1.\left||\beta_{j}|-\frac{1}{2}\right|<\frac{\beta_{0}}{2},\ \ \ j=\pm 1. (109)

By (100) and (104) we get that for every fixed 1<p<1<p<\infty such that pβ0<1p\beta_{0}<1 the estimate

e±HdpC\|e^{\pm Hd}\|_{p}\leq C (110)

holds. Hence we also have a priori LpL^{p} estimate on function (97).

Since

fθ=fgθ,\frac{\partial f}{\partial\theta}=f\frac{\partial g}{\partial\theta}, (111)

an estimate on gθ\frac{\partial g}{\partial\theta} will give an estimate on fθ\frac{\partial f}{\partial\theta}. We can write

gθ=(ei(d+iHd))(ei(d+iHd)gθ).\frac{\partial g}{\partial\theta}=\left(e^{-i(d+iHd)}\right)\left(e^{i(d+iHd)}\frac{\partial g}{\partial\theta}\right). (112)

By assumptions of Theorem 1.2 we have β015<12\beta_{0}\leq\frac{1}{5}<\frac{1}{2} and we can choose p>2p>2. By Cauchy-Schwarz inequality we then have

gθp2ei(d+iHd)pei(d+iHd)gθp.\left\|\frac{\partial g}{\partial\theta}\right\|_{\frac{p}{2}}\leq\left\|e^{-i(d+iHd)}\right\|_{p}\left\|e^{i(d+iHd)}\frac{\partial g}{\partial\theta}\right\|_{p}. (113)

From here we get Lp2L^{\frac{p}{2}} a priori estimates on gθ\frac{\partial g}{\partial\theta} which imply a priori estimates on gg and ff in Hölder space Cβ(Δ)C^{\beta}(\partial{\rm\Delta}) for 0<β<12p<2(12β0)0<\beta<1-\frac{2}{p}<2(\frac{1}{2}-\beta_{0}). Recall (Remark 5.2) that this gives Hölder space a priori estimates on solutions with no zeros of the nonsymmetrical problem (12-13) for β(0,12β0)\beta\in(0,\frac{1}{2}-\beta_{0}).

5.2 A priori estimates on function κ\kappa

We also need a priori estimates on function κ\kappa for which it holds

f(ξ)=(ξ1)δ1(ξ+1)δ1κ(ξ)+f(1)1+ψ(ξ)2+f(1)1ψ(ξ)2.f(\xi)=(\xi-1)^{\delta_{1}}(\xi+1)^{\delta_{-1}}\kappa(\xi)+f(1)\frac{1+\psi(\xi)}{2}+f(-1)\frac{1-\psi(\xi)}{2}. (114)

In this subsection we again consider the nonsymmetrical case (12-13). We denote by CC a universal constant, which depends on the data but does not depend on the particular function we consider.

We know that ff and hence κ\kappa are Ck1,αC^{k-1,\alpha} smooth on Δ{1,1}\partial{\rm\Delta}\setminus\{-1,1\} and on compact subsets of Δ{1,1}\partial{\rm\Delta}\setminus\{1,-1\} we get a priori estimates on κ\kappa by expressing it in terms of ff. Hence we need a priori estimates on κ\kappa near points ±1\pm 1. Also, we know from Section 3 that if κ\kappa is continuous on Δ¯\overline{{\rm\Delta}}, then both functions belong to Aβ(Δ)A^{\beta}({\rm\Delta}) for

0<β<min{|β1|,1|β1|,|β1|,1|β1|}.0<\beta<\min\{|\beta_{1}|,1-|\beta_{1}|,|\beta_{-1}|,1-|\beta_{-1}|\}. (115)

Let us fix 0<β<12β00<\beta<\frac{1}{2}-\beta_{0} that we have a priori estimates on function ff.

Recall (38) and that tsκ(t)=f(t)t^{s}\kappa(t)=f(t). Hence ρ~w(θ,κ)\widetilde{\rho}_{w}(\theta,\kappa) is a CβC^{\beta} function with a priori bounds. As in (60) we can globally write

Re(rei(v+iHv)κθ)=eue(Hv)ρ~θ(θ,κ),{\rm Re}\left(re^{i(v+iHv)}\frac{\partial\kappa}{\partial\theta}\right)=-e^{-u}e^{-(Hv)}\widetilde{\rho}_{\theta}(\theta,\kappa), (116)

where uu and vv are real CβC^{\beta} functions with a priori bounds. To get LpL^{p^{\prime}} a priori bounds on κθ\frac{\partial\kappa}{\partial\theta} for some p>1p^{\prime}>1 we will get LpL^{p^{\prime}} bounds on the right-hand side function ρ~θ(θ,κ(θ))\widetilde{\rho}_{\theta}(\theta,\kappa(\theta)), that it, on ρ~t(t,κ(t))\widetilde{\rho}_{t}(t,\kappa(t)) near t=0t=0. Considering (41-42-43) termwise we get that tβ11t^{\beta_{1}-1}, tβ1t^{-\beta_{1}}, κ(t)=tβ11f(t)\kappa(t)=t^{\beta_{1}-1}f(t), and all terms with function gg are LpL^{p^{\prime}} bounded for any p>1p^{\prime}>1 such that

p(1β1)<1andpβ1<1.p^{\prime}(1-\beta_{1})<1\ \ \ {\rm and}\ \ \ p^{\prime}\beta_{1}<1. (117)

Let us consider terms which are bounded by tβ1|κ(t)2|=tβ12|f(t)2|t^{-\beta_{1}}|\kappa(t)^{2}|=t^{\beta_{1}-2}|f(t)^{2}|. Since we have β(0,12β0)\beta\in(0,\frac{1}{2}-\beta_{0}) a priori bounds on ff, we have

|f(t)|C|t|β|f(t)|\leq C|t|^{\beta} (118)

for some universal constant CC. Hence

tβ1|κ(t)2|C|t|2β+β12t^{-\beta_{1}}|\kappa(t)^{2}|\leq C|t|^{2\beta+\beta_{1}-2} (119)

and this function is in some LpL^{p^{\prime}}, p>1p^{\prime}>1, if 1<2β+β11<2\beta+\beta_{1}. The bound 0<β<12β00<\beta<\frac{1}{2}-\beta_{0} implies that this will be the case for some such β\beta if 2β0<β12\beta_{0}<\beta_{1}. Similar argument near ξ=1\xi=-1 gives 2β0<1|β1|2\beta_{0}<1-|\beta_{-1}|.

If these two conditions are satisfied, we get LpL^{p^{\prime}} a priori estimates on ρ~θ(θ,κ(θ))\widetilde{\rho}_{\theta}(\theta,\kappa(\theta)) for some p>1p^{\prime}>1. This implies CβC^{\beta^{\prime}} a priori estimate on κ\kappa for β<11p\beta^{\prime}<1-\frac{1}{p^{\prime}}.

There are natural bounds on βj\beta_{j}, j=±1j=\pm 1, in terms of β0\beta_{0}, that is,

12β02<|βj|<12+β02.\frac{1}{2}-\frac{\beta_{0}}{2}<|\beta_{j}|<\frac{1}{2}+\frac{\beta_{0}}{2}. (120)

Hence, if 2β012β022\beta_{0}\leq\frac{1}{2}-\frac{\beta_{0}}{2} and 12+β0212β0\frac{1}{2}+\frac{\beta_{0}}{2}\leq 1-2\beta_{0} both inequalities needed for LpL^{p^{\prime}} a priori estimates will be satisfied. These two inequalities are equivalent to the condition β015\beta_{0}\leq\frac{1}{5}, that is, the angle between ww and the normal to γωj\gamma_{\omega_{j}} at ww is less than π10\frac{\pi}{10}.

6 Final remarks

If arc LL is the lower semicircle, we can state Theorem 1.2 in an equivalent simplified form.

Theorem 6.1

Let {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\in\partial{\rm\Delta}\setminus\mathring{L}} be a CkC^{k} (k3)(k\geq 3) family of Jordan curves in {\mathbb{C}} which all contain point 0 in their interiors. Let Jordan curves γj\gamma_{j}, j=±1j=\pm 1, be strongly starshaped with respect to 0 and such that for each wγωjw\in\gamma_{\omega_{j}} the angle between ww and the outer normal to γωj\gamma_{\omega_{j}} at ww is less than π10\frac{\pi}{10}. Let wjw_{j}, j=±1j=\pm 1, be the positive intersection of γj\gamma_{j} and the real axis with the oriented angle of intersection πβj\pi\beta_{j}, where β1(0,1)\beta_{1}\in(0,1) and β1(1,0)\beta_{-1}\in(-1,0). Let

0<β<min{β1,1β1,|β1|,1|β1|}.0<\beta<\min\{\beta_{1},1-\beta_{1},|\beta_{-1}|,1-|\beta_{-1}|\}. (121)

Then there exists a unique fAβ(Δ)f\in A^{\beta}({\rm\Delta}) with no zeros on Δ{\rm\Delta} which solves (3-4) for which f(1)=w1f(1)=w_{1} and f(1)=w1f(-1)=w_{-1}.

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 {γξ}ξΔL̊\{\gamma_{\xi}\}_{\xi\partial{\rm\Delta}\setminus\mathring{L}} to circles with center at 0 and fixed radius R>0R>0, embedded in a one parameter family of boundary value problems which all satisfy assumptions of Theorem 6.1. Here, for t=0t=0 we have the starting boundary value problem and for t=1t=1 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 {γξt}ξΔL̊\{\gamma^{t}_{\xi}\}_{\xi\partial{\rm\Delta}\setminus\mathring{L}} can be locally perturbed into a solution for the nearby perturbed boundary data. Hence the set of parameters tt 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 t[0,1]t\in[0,1] for which there is a solution of (3-4) is closed. Since there is an obvious solution for the case t=1t=1, where all Jordan curves are circles with center at 0 and fixed radius R>0R>0, we get that there is a solution of (3-4) for t=0t=0.

Corollary 6.2

Let a1,,anΔa_{1},\dots,a_{n}\in{\rm\Delta} be a finite set of points with given multiplicities. Then under the assumptions of Theorem 6.1 there exists β(0,1)\beta\in(0,1) and fAβ(Δ)f\in A^{\beta}({\rm\Delta}) which has zeros exactly at points a1,,anΔa_{1},\dots,a_{n}\in{\rm\Delta} with the given multiplicites and which solves (12-13).

To prove the corollary we search for solutions ff of the symmetric problem on the unit disc of the form

f(z)=za1a¯zza¯1azf~(z),f(z)=\frac{z-a}{1-\overline{a}z}\,\frac{z-\overline{a}}{1-az}\,\widetilde{f}(z), (122)

where aa is a point in the upper half-disc. Then f~\widetilde{f} has to solve a modified problem, where the boundary curves are given by

γ~ξ=1a¯ξξa1aξξa¯γξ.\widetilde{\gamma}_{\xi}=\frac{1-\overline{a}\xi}{\xi-a}\,\frac{1-a\xi}{\xi-\overline{a}}\,\gamma_{\xi}. (123)

Observe that γ~ξ=γξ\widetilde{\gamma}_{\xi}=\gamma_{\xi} for ξ=±1\xi=\pm 1.

Remark 6.3

In a similar way one can create a zero at a point a(1,1)a\in(-1,1), that is, on L̊\mathring{L} in the original problem. Jordan curves for the modified problem are

γ~ξ=1aξξaγξ.\widetilde{\gamma}_{\xi}=\frac{1-a\xi}{\xi-a}\,\gamma_{\xi}. (124)

Then γ~1=γ1\widetilde{\gamma}_{1}=\gamma_{1} and γ~1=γ1\widetilde{\gamma}_{-1}=-\gamma_{-1} 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 n{\mathbb{C}}^{n}. 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.