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

Bifurcation for a free boundary problem modeling a small arterial plaque

Xinyue Evelyn Zhao Bei Hu Department of Applied and Computational Mathematics and Statistics,
University of Notre Dame, Notre Dame, IN 46556, USA,
[email protected], [email protected]
Abstract

Atherosclerosis, hardening of the arteries, originates from small plaque in the arteries; it is a major cause of disability and premature death in the United States and worldwide. In this paper, we study the bifurcation of a highly nonlinear and highly coupled PDE model describing the growth of arterial plaque in the early stage of atherosclerosis. The model involves LDL and HDL cholesterols, macrophage cells, and foam cells, with the interface separating the plaque and blood flow regions being a free boundary. We establish finite branches of symmetry-breaking stationary solutions which bifurcate from the radially symmetric solution. Since plaque in reality is unlikely to be strictly radially symmetric, our result would be useful to explain the asymmetric shapes of plaque.

keywords:
free boundary problem, atherosclerosis, bifurcation, symmetry-breaking.

1 Introduction

Atherosclerosis, known as an inflammatory disease, is a major cause of disability and premature death in the United States and worldwide. It occurs when fat, cholestrol, and other substances build up in and on the artery walls. These deposits are called plaques, which harden and narrow the arteries over time. The plaque can rupture, triggering a blood clot which restricts blood flow. During this process, a heart attack, stroke, or sudden cardiac death may occur. Every year about 735,000 Americans have a heart attack, and about 610,000 people die of heart diseases in the United States — that is 1 in every 4 deaths (cf.,[1, 22]).

There are several mathematical models that describe the growth of plaque in the arteries (see [2, 3, 8, 9, 13, 18, 19]). All of these models recognize the critical role of the “bad” cholesterols, low density lipoprotein (LDL), and the “good” cholesterols, high density lipoprotein (HDL), in determining whether plaque will grow or shrink. Recently, a free boundary PDE model was proposed in [13], in which a risk-map was generated for any pair values of (LDL, HDL), showing the important influence of LDL and HDL on plaque formation. Later, on the foundation of the model, Hao and Friedman added the impact of reverse cholesterol transport (RCT) in [8]. In addition, the existence of a small radially symmetric stationary plaque and its stability condition were theoretically established for a simplified free boundary model in [9]. Nevertheless, there is no theoretical work to analyze the bifurcation of plaque model. As the plaque in reality is unlikely to be radially symmetric, it is necessary to investigate the non-radially symmetric solutions. Hence in this paper we shall carry out the bifurcation for the plaque model proposed in [9] (also see [7, Chapters 7 and 8]).

Refer to caption
Figure 1: The cross section of an artery.

The process of plaque formation is as follows: when a lesion develops in the inner surface of the arterial wall, it enables LDL and HDL to move into the intima and become oxidized by free radicals. Oxidized LDL triggers endothelial cells to secrete chemoattractant proteins that attract macrophage cells (M) from the blood. Macrophage cells can engulf oxidized LDL, they then become foam cells (F), and the accumulation of foam cells results in the formation of plaque. The effect of oxidized LDL on plaque growth can be reduced by the good cholesterols, HDL: HDL can remove harmful bad cholesterol out from the foam cells and revert foam cells back into macrophage cells; moreover, HDL also competes with LDL on free radicals, decreasing the amount of radicals that are available to oxidize LDL. In the model, we let

L = concentration of LDL,H = concentration of HDL,\displaystyle L\text{ = concentration of LDL,}\hskip 50.00008ptH\text{ = concentration of HDL,}
M = density of macrophage cells,F = density of foam cells.\displaystyle M\text{ = density of macrophage cells,}\hskip 20.00003ptF\text{ = density of foam cells.}

Assuming the artery is a very long circular cylinder with radius 1 (after normalization), we consider a circular cross section of the artery. As can be seen in Fig. 1, the cross section is divided into two regions: blood flow region Σ(t)\Sigma(t) and plaque region Ω(t)\Omega(t), with a moving boundary Γ(t)\Gamma(t) separating these two regions (since plaque can either grow or shrink). The variables L,H,M,FL,H,M,F satisfy the following equations in the plaque region {Ω(t),t>0}\{\Omega(t),t>0\} (cf., [7, Chapters 7 and 8] and [9]):

LtΔL=k1MLK1+Lρ1L,\displaystyle\frac{\partial L}{\partial t}-\Delta L=-k_{1}\frac{ML}{K_{1}+L}-\rho_{1}L, (1.1)
HtΔH=k2HFK2+Fρ2H,\displaystyle\frac{\partial H}{\partial t}-\Delta H=-k_{2}\frac{HF}{K_{2}+F}-\rho_{2}H, (1.2)
MtDΔM+(Mv)=k1MLK1+L+k2HFK2+F+λMLγ+Hρ3M,\displaystyle\frac{\partial M}{\partial t}-D\Delta M+\nabla\cdot(M\vec{v})=-k_{1}\frac{ML}{K_{1}+L}+k_{2}\frac{HF}{K_{2}+F}+\lambda\frac{ML}{\gamma+H}-\rho_{3}M, (1.3)
FtDΔF+(Fv)=k1MLK1+Lk2HFK2+Fρ4F,\displaystyle\frac{\partial F}{\partial t}-D\Delta F+\nabla\cdot(F\vec{v})=k_{1}\frac{ML}{K_{1}+L}-k_{2}\frac{HF}{K_{2}+F}-\rho_{4}F, (1.4)

where ρ1\rho_{1}, ρ2\rho_{2}, ρ3\rho_{3} and ρ4\rho_{4} denote the natural death rate of LL, HH, MM, and FF, respectively. In equations ((1.1))((1.4)), the aforementioned transitions between macrophage cells (MM) and foam cells (FF) are included: k1MLK1+Lk_{1}\frac{ML}{K_{1}+L} accounts for the fact that MM becomes foam cell by combining with LL, k2HFK2+Fk_{2}\frac{HF}{K_{2}+F} describes the removal of foam cell by HH, and the extra term λMLγ+H\lambda\frac{ML}{\gamma+H} in equation ((1.3)) models the effects that oxidized LL attracts MM while HH decreases this impact by competing for free radicals.

We assume that the density of cells in the plaque is approximately a constant, and take

M+FM0in Ω(t).M+F\equiv M_{0}\hskip 20.00003pt\text{in }\Omega(t). (1.5)

Since there are cells migrating into and out of the plaque, the total number of cells keeps changing and, under the assumption ((1.5)), cells are continuously “pushing” each other. This gives rise to an internal pressure among the cells which is associated with the velocity v\vec{v} in ((1.3)) and ((1.4)). We further assume that the plaque texture is of a porous medium type, and invoke Darcy’s law

v=p(the proportional constant is normalized to 1),\vec{v}=-\nabla p\hskip 20.00003pt(\mbox{the proportional constant is normalized to 1}), (1.6)

where pp is the internal pressure relative to the outside pressure (and therefore can admit positive or negative sign). Combining ((1.3))((1.6)), we derive

Δp=1M0[λ(M0F)Lγ+Hρ3(M0F)ρ4F].-\Delta p=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F)L}{\gamma+H}-\rho_{3}(M_{0}-F)-\rho_{4}F\Big{]}. (1.7)

Due to the assumption ((1.5)), we can decrease the number of equations by 1, and replace MM by M0FM_{0}-F in ((1.1))((1.4)), hence we shall have 4 PDEs, for LL, HH, FF and pp, respectively. In particular, combining with ((1.7)), we write the equation for FF in the following form

FtDΔFFp=k1(M0F)LK1+Lk2HFK2+FλF(M0F)LM0(γ+H)+(ρ3ρ4)(M0F)FM0.\frac{\partial F}{\partial t}-D\Delta F-\nabla F\cdot\nabla p=k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-k_{2}\frac{HF}{K_{2}+F}-\lambda\frac{F(M_{0}-F)L}{M_{0}(\gamma+H)}+(\rho_{3}-\rho_{4})\frac{(M_{0}-F)F}{M_{0}}. (1.8)

We now turn to the boundary conditions. We assume no flux condition on the blood vessel wall (r=1r=1) for all variables (no exchange through the blood vessel):

Lr=Hr=Fr=pr=0at r=1;\frac{\partial L}{\partial r}=\frac{\partial H}{\partial r}=\frac{\partial F}{\partial r}=\frac{\partial p}{\partial r}=0\hskip 20.00003pt\text{at }r=1; (1.9)

while on the free boundary Γ(t)\Gamma(t), we take

L𝐧+β1(LL0)=0\displaystyle\frac{\partial L}{\partial{\bf n}}+\beta_{1}(L-L_{0})=0\hskip 20.00003pt on Γ(t),\displaystyle\text{on }\Gamma(t), (1.10)
H𝐧+β1(HH0)=0\displaystyle\frac{\partial H}{\partial{\bf n}}+\beta_{1}(H-H_{0})=0\hskip 20.00003pt on Γ(t),\displaystyle\text{on }\Gamma(t), (1.11)
F𝐧+β2F=0\displaystyle\frac{\partial F}{\partial{\bf n}}+\beta_{2}F=0\hskip 20.00003pt on Γ(t),\displaystyle\text{on }\Gamma(t), (1.12)
p=κ\displaystyle p=\kappa\hskip 20.00003pt on Γ(t),\displaystyle\text{on }\Gamma(t), (1.13)

where 𝐧{\bf n} is the outward unit normal for Γ(t)\Gamma(t) which points inward towards the blood region (as shown in Fig. 1), and κ\kappa is the corresponding mean curvature in the direction of 𝐧{\bf n} (i.e., κ=1R(t)\kappa=-\frac{1}{R(t)} if Γ(t)={r=R(t)}\Gamma(t)=\{r=R(t)\}). The cell-to-cell adhesiveness constant in front of κ\kappa is normalized to 1. The flux boundary conditions ((1.10)) and ((1.11)) are based on the fact that the concentrations of LL and HH in the blood are L0L_{0} and H0H_{0}, respectively; and the meaning of ((1.12)) is similar: there are, of course, no foam cells in the blood.

Furthermore, we assume that the velocity is continuous up to the boundary, so that the free boundary Γ(t)\Gamma(t) moves in the outward normal direction 𝐧{\bf n} with velocity v\vec{v}; based on ((1.6)), the normal velocity of the free boundary is defined by

Vn=p𝐧on Γ(t).V_{n}=-\frac{\partial p}{\partial{\bf n}}\hskip 20.00003pt\text{on }\Gamma(t). (1.14)

In [9], Friedman et al. analyzed the system ((1.1))((1.14)) in the radially symmetric case and established the existence of a unique radially symmetric steady state solution in a ring-region 1ε<r<11-\varepsilon<r<1 with ε\varepsilon being small. It is, however, unreasonable to assume plaque is of strictly radially symmetric shape, hence we’d like to investigate the symmetric-breaking bifurcation for the system. To do that, we study the corresponding stationary problem of ((1.1))((1.14)):

ΔL=k1(M0F)LK1+Lρ1L\displaystyle-\Delta L=-k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-\rho_{1}L\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega, (1.15)
ΔH=k2HFK2+Fρ2H\displaystyle-\Delta H=-k_{2}\frac{HF}{K_{2}+F}-\rho_{2}H\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega, (1.16)
DΔFFp=k1(M0F)LK1+Lk2HFK2+FλF(M0F)LM0(γ+H)+(ρ3ρ4)(M0F)FM0\displaystyle-D\Delta F-\nabla F\cdot\nabla p=k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-k_{2}\frac{HF}{K_{2}+F}-\lambda\frac{F(M_{0}-F)L}{M_{0}(\gamma+H)}+(\rho_{3}-\rho_{4})\frac{(M_{0}-F)F}{M_{0}}\hskip 5.0pt in Ω,\displaystyle\text{in }\Omega, (1.17)
Δp=1M0[λ(M0F)Lγ+Hρ3(M0F)ρ4F]\displaystyle-\Delta p=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F)L}{\gamma+H}-\rho_{3}(M_{0}-F)-\rho_{4}F\Big{]}\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega, (1.18)
Lr=Hr=Fr=pr=0\displaystyle\frac{\partial L}{\partial r}=\frac{\partial H}{\partial r}=\frac{\partial F}{\partial r}=\frac{\partial p}{\partial r}=0\hskip 20.00003pt r=1,\displaystyle r=1, (1.19)
L𝐧+β1(LL0)=0,H𝐧+β1(HH0)=0,F𝐧+β2F=0\displaystyle\frac{\partial L}{\partial{\bf n}}+\beta_{1}(L-L_{0})=0,\hskip 10.00002pt\frac{\partial H}{\partial{\bf n}}+\beta_{1}(H-H_{0})=0,\hskip 10.00002pt\frac{\partial F}{\partial{\bf n}}+\beta_{2}F=0 on Γ,\displaystyle\text{on }\Gamma, (1.20)
p=κ\displaystyle p=\kappa on Γ,\displaystyle\text{on }\Gamma, (1.21)
Vn=p𝐧=0\displaystyle V_{n}=-\frac{\partial p}{\partial{\bf n}}=0\hskip 20.00003pt on Γ.\displaystyle\text{on }\Gamma. (1.22)

In recent years, considerable research works have been carried out on bifurcation analysis for various tumor growth models (see [5, 6, 10, 11, 15, 14, 16, 17, 20, 23, 24, 25, 27, 21]), where the Crandall-Rabinowitz theorem (will be mentioned in Section 2) is a primary tool. Compared with tumor growth models, our system ((1.15))((1.22)) contains more equations which are highly nonlinear and coupled together, therefore it is a formidable task to analyze our model. Besides, the absence of an explicit stationary solution presents a big challenge to verify the Crandall-Rabinowitz theorem. Even though the problems in [26, 27] do not admit explicit representations, the structure of the problem studied here is very different. To overcome it, we establish a lot of sharp estimates in Section 4. To the best of our knowledge, this is the first paper on the study of bifurcation for the system ((1.1))((1.14)). Our main result is stated as follows:

For convenience we shall use μ=1ε[λL0ρ3(γ+H0)]\mu=\frac{1}{\varepsilon}[\lambda L_{0}-\rho_{3}(\gamma+H_{0})] as our bifurcation parameter. We will keep all parameters fixed except L0L_{0} and ρ4\rho_{4}, and vary μ\mu by changing L0L_{0}.

Theorem 1.1.

For each integer n2n\geq 2, we can find a small E>0E>0 and for each 0<ε<E0<\varepsilon<E, there exists a unique μn=(γ+H0)n2(1n2)+O(n5ε)\mu_{n}=(\gamma+H_{0})n^{2}(1-n^{2})+O(n^{5}\varepsilon) such that if μn>μc\mu_{n}>\mu_{c} (μc\mu_{c} is defined in (2.9)), then μ=μn\mu=\mu_{n} is a bifurcation point of the symmetry-breaking stationary solution of the system ((1.15))((1.22)). Moreover, the free boundary of this bifurcation solution is of the form

r=1ε+τcos(nθ)+o(τ),where |τ|ε.r=1-\varepsilon+\tau\cos(n\theta)+o(\tau),\hskip 10.00002pt\text{where }\hskip 10.00002pt|\tau|\ll\varepsilon.
Remark 1.1.

Unlike tumor protrusions which are usually unstable and may cause metastases, the protrusions of plaques are towards the blood region with limited spatial freedom. As nn gets bigger, μn\mu_{n} becomes negative with larger absolute value. By the definition of μn\mu_{n}, this means that the concentration of the good cholesterol (HDL) must be substantially larger than the concentration of the bad cholesterol (LDL) for the bifurcation to occur. The more protrusions, the larger H0H_{0} over L0L_{0} will be required to balance the protrusion forces. Based on the stability results from [9], it is likely to have some stable bifurcation branches.

The structure of this paper is as follows. In Section 2, we give some preliminaries; in section 3, we rigorously justify some expansions which will be needed in applying the Crandall-Rabinowitz theorem; and then we carry out our proof of Theorem 1.1 in Section 4. Some well-known results are collected in the Appendix.

2 Radially symmetric stationary solution

2.1 A small radially symmetric stationary solution

We consider a radially symmetric stationary solution in a small ring-region Ω={1ε<r<1}\Omega_{*}=\{1-\varepsilon<r<1\}, and denote the solution by (L,H,F,p)(L_{*},H_{*},F_{*},p_{*}). Based on ((1.15))((1.22)), the solution satisfies

ΔL=k1(M0F)LK1+Lρ1L\displaystyle-\Delta L_{*}=-k_{1}\frac{(M_{0}-F_{*})L_{*}}{K_{1}+L_{*}}-\rho_{1}L_{*} in Ω,\displaystyle\text{in }\Omega_{*}, (2.1)
ΔH=k2HFK2+Fρ2H\displaystyle-\Delta H_{*}=-k_{2}\frac{H_{*}F_{*}}{K_{2}+F_{*}}-\rho_{2}H_{*} in Ω,\displaystyle\text{in }\Omega_{*}, (2.2)
DΔFFrpr=k1(M0F)LK1+Lk2HFK2+FλF(M0F)LM0(γ+H)+(ρ3ρ4)(M0F)FM0\displaystyle-D\Delta F_{*}-\frac{\partial F_{*}}{\partial r}\frac{\partial p_{*}}{\partial r}=k_{1}\frac{(M_{0}-F_{*})L_{*}}{K_{1}+L_{*}}-k_{2}\frac{H_{*}F_{*}}{K_{2}+F_{*}}-\lambda\frac{F_{*}(M_{0}-F_{*})L_{*}}{M_{0}(\gamma+H_{*})}+(\rho_{3}-\rho_{4})\frac{(M_{0}-F_{*})F_{*}}{M_{0}} in Ω,\displaystyle\text{in }\Omega_{*}, (2.3)
Δp=1M0[λ(M0F)Lγ+Hρ3(M0F)ρ4F]\displaystyle-\Delta p_{*}=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F_{*})L_{*}}{\gamma+H_{*}}-\rho_{3}(M_{0}-F_{*})-\rho_{4}F_{*}\Big{]}\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega_{*}, (2.4)
Lr=Hr=Fr=pr=0,\displaystyle\frac{\partial L_{*}}{\partial r}=\frac{\partial H_{*}}{\partial r}=\frac{\partial F_{*}}{\partial r}=\frac{\partial p_{*}}{\partial r}=0,\hskip 20.00003pt r=1,\displaystyle r=1, (2.5)
Lr+β1(LL0)=0,Hr+β1(HH0)=0,Fr+β2F=0,\displaystyle\hskip 10.00002pt-\frac{\partial L_{*}}{\partial r}+\beta_{1}(L_{*}-L_{0})=0,\hskip 10.00002pt-\frac{\partial H_{*}}{\partial r}+\beta_{1}(H_{*}-H_{0})=0,\hskip 10.00002pt-\frac{\partial F_{*}}{\partial r}+\beta_{2}F_{*}=0,\hskip 10.00002pt r=1ε,\displaystyle r=1-\varepsilon, (2.6)
p=11ε,\displaystyle p_{*}=-\frac{1}{1-\varepsilon},\hskip 10.00002pt r=1ε,\displaystyle r=1-\varepsilon, (2.7)
pr=0,\displaystyle\frac{\partial p_{*}}{\partial r}=0, r=1ε.\displaystyle r=1-\varepsilon. (2.8)

Viewing pr\frac{\partial p_{*}}{\partial r} as v-v, and following Theorem 3.1 in [9], for every H0=O(1)H_{0}=O(1) and ε\varepsilon small, we can find a unique L0L_{0} and a constant KK_{*}, such that there is a unique classical solution to the above system with |λL0ρ3(H0+γ)|<Kε|\lambda L_{0}-\rho_{3}(H_{0}+\gamma)|<K_{*}\varepsilon. The existence theorem for radially symmetric solution of this form, however, is not good enough for the bifurcation theorem.

There are many parameters in our system. We need to choose one as the bifurcation parameter. We let μ=1ε[λL0ρ3(γ+H0)]\mu=\frac{1}{\varepsilon}[\lambda L_{0}-\rho_{3}(\gamma+H_{0})] to be our bifurcation parameter. We can vary μ\mu by, say, keeping λ,γ,ρ3,H0\lambda,\gamma,\rho_{3},H_{0} and ε\varepsilon fixed while changing L0L_{0} only. For simplicity, we shall assume all the parameters are fixed and of order O(1)O(1) except L0L_{0} and ρ4\rho_{4}. With these settings, varying L0L_{0} corresponds to varying μ\mu. In the rest of this paper, we shall thus use μ\mu and ρ4\rho_{4} as our parameters.

Here is our existence theorem for the radially symmetric solutions. We define

μc=ρ3β1{(γ+H0)(λk1M0λK1+ρ3(γ+H0)+ρ1)ρ2H0}.\mu_{c}=\frac{\rho_{3}}{\beta_{1}}\Big{\{}(\gamma+H_{0})\Big{(}\frac{\lambda k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+{\rho_{1}}\Big{)}-\rho_{2}H_{0}\Big{\}}. (2.9)
Theorem 2.2.

For every μ>μc\mu^{*}>\mu_{c} and μc<μ<μ\mu_{c}<\mu<\mu^{*}, we can find a small ε>0\varepsilon^{*}>0, and for each 0<ε<ε0<\varepsilon<\varepsilon^{*}, there exists a unique ρ4\rho_{4} such that the system ((2.1))((2.8)) admits a unique solution (L,H,F,p)(L_{*},H_{*},F_{*},p_{*}).

Proof.

The proof is similar to that in [9] but much more involved. Following Lemma 3.1 of [9], for all parameters of order O(1)O(1), the system ((2.1))((2.7)) admits a unique solution for small ε\varepsilon. In order for this solution to be the solution of our problem, we need to verify ((2.8)). We shall do so by keeping all parameters fixed except ρ4\rho_{4}.

Note that ((2.8)) is equivalent to

Φ(ρ4,ε,μ)=0,where Φ(ρ4,ε,μ)1ε1[λ(M0F)Lγ+Hρ3(M0F)ρ4F]rdr.\Phi(\rho_{4},\varepsilon,\mu)=0,\hskip 20.00003pt\text{where }\Phi(\rho_{4},\varepsilon,\mu)\triangleq\int_{1-\varepsilon}^{1}\Big{[}\lambda\frac{(M_{0}-F_{*})L_{*}}{\gamma+H_{*}}-\rho_{3}(M_{0}-F_{*})-\rho_{4}F_{*}\Big{]}r\mathrm{d}r. (2.10)

As in [9, (3.29)–(3.32)], recalling also (see Appendix 5.1) ξ(r)=1r24+12logr=O(ε2)\xi(r)=\frac{1-r^{2}}{4}+\frac{1}{2}\log r=O(\varepsilon^{2}) (the formulas in [9, (3.23)–(3.25), (3.26)–(3.28), (3.29)] are all missing minus signs; as a result, the corrected [9, (3.29)] should read:

L(r)=L0(k1M0L0K1+L0+ρ1L0)(ξ(r)+εβ1)+Constε2+O(ε3),L_{*}(r)=L_{0}-\Big{(}\frac{k_{1}M_{0}L_{0}}{K_{1}+L_{0}}+\rho_{1}L_{0}\Big{)}\Big{(}\xi(r)+\frac{\varepsilon}{\beta_{1}}\Big{)}+\text{Const}\cdot\varepsilon^{2}+O(\varepsilon^{3}), ([9,(3.29)])

and [9, (3.30),(3.31)] should be corrected in a similar manner; this correction does not change the proof in [9]), we can establish the following:

L(r)\displaystyle L_{*}(r) =\displaystyle= L0εβ1(k1M0L0K1+L0+ρ1L0)+O(ε2)\displaystyle L_{0}-\frac{\varepsilon}{\beta_{1}}\Big{(}\frac{k_{1}M_{0}L_{0}}{K_{1}+L_{0}}+\rho_{1}L_{0}\Big{)}+O(\varepsilon^{2})
=\displaystyle= ρ3(γ+H0)λ+ε[μλρ3(γ+H0)β1(k1M0λK1+ρ3(γ+H0)+ρ1λ)]+O(ε2)\displaystyle\frac{\rho_{3}(\gamma+H_{0})}{\lambda}+\varepsilon\Big{[}\frac{\mu}{\lambda}-\frac{\rho_{3}(\gamma+H_{0})}{\beta_{1}}\Big{(}\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+\frac{\rho_{1}}{\lambda}\Big{)}\Big{]}+O(\varepsilon^{2})
\displaystyle\triangleq ρ3(γ+H0)λ+εL1+O(ε2),\displaystyle\frac{\rho_{3}(\gamma+H_{0})}{\lambda}+\varepsilon L_{*}^{1}+O(\varepsilon^{2}),
H(r)\displaystyle H_{*}(r) =\displaystyle= H0ερ2H0β1+O(ε2)H0+εH1+O(ε2),\displaystyle H_{0}-\varepsilon\frac{\rho_{2}H_{0}}{\beta_{1}}+O(\varepsilon^{2})\;\triangleq\;H_{0}+\varepsilon H_{*}^{1}+O(\varepsilon^{2}), (2.12)
F(r)\displaystyle F_{*}(r) =\displaystyle= εβ2k1M0L0D(K1+L0)+O(ε2)\displaystyle\frac{\varepsilon}{\beta_{2}}\frac{k_{1}M_{0}L_{0}}{D(K_{1}+L_{0})}+O(\varepsilon^{2})
=\displaystyle= ερ3(γ+H0)β2Dk1M0λK1+ρ3(γ+H0)+O(ε2)εF1+O(ε2).\displaystyle\varepsilon\;\frac{\rho_{3}(\gamma+H_{0})}{\beta_{2}D}\;\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+O(\varepsilon^{2})\;\triangleq\;\varepsilon F_{*}^{1}+O(\varepsilon^{2}).

Substituting these expressions into the formula ((2.10)) for Φ\Phi, we find that the O(1)O(1) terms in the bracket [][\cdots] cancel out, and

Φ(ρ4,ε,μ)=1ε1{ε[M0(λL1ρ3H1)γ+H0ρ4F1]+O(ε2)}r𝑑r.\Phi(\rho_{4},\varepsilon,\mu)=\int_{1-\varepsilon}^{1}\bigg{\{}\varepsilon\Big{[}\frac{M_{0}(\lambda L_{*}^{1}-\rho_{3}H_{*}^{1})}{\gamma+H_{0}}-\rho_{4}F_{*}^{1}\Big{]}+O(\varepsilon^{2})\bigg{\}}rdr. (2.14)

A direct computation shows that

M0(λL1ρ3H1)γ+H0\displaystyle\frac{M_{0}(\lambda L_{*}^{1}-\rho_{3}H_{*}^{1})}{\gamma+H_{0}} =\displaystyle= M0γ+H0(μμc).\displaystyle\frac{M_{0}}{\gamma+H_{0}}(\mu-\mu_{c}). (2.15)

It follows that, for small ε\varepsilon, Φ(0,ε,μ)>0\Phi(0,\varepsilon,\mu)>0 and Φ(ρ4,ε,μ)<0\Phi(\rho_{4},\varepsilon,\mu)<0 for large ρ4\rho_{4}, hence there must be a value of ρ4\rho_{4} on which Φ(ρ4,ε,μ)=0\Phi(\rho_{4},\varepsilon,\mu)=0.

To finish the proof, it suffices to show ρ4Φ(ρ4,ε)<0\frac{\partial}{\partial\rho_{4}}\Phi(\rho_{4},\varepsilon)<0; the proof is similar to that of [9, Theorem 3.1] in the second part, but is actually a little easier. ∎

Remark 2.1.

By ODE theories, the solution (L,H,F,p)(L_{*},H_{*},F_{*},p_{*}) can be extended to the bigger region Ω2ε={12ε<r<1}\Omega_{2\varepsilon}=\{1-2\varepsilon<r<1\} while maintaining CC^{\infty} regularity. For notational convenience, we still use (L,H,F,p)(L_{*},H_{*},F_{*},p_{*}) to denote the extended solution.

Remark 2.2.

The case μc<0\mu_{c}<0 is certainly true within reasonable parameter range.

Following the above proof, we also derive

Lemma 2.3.

Let μ>μc\mu>\mu_{c}. Then

ρ4\displaystyle\rho_{4} =\displaystyle= 1F1M0γ+H0(μμc)+O(ε)=β2D[λK1+ρ3(γ+H0)]ρ3k1(γ+H0)2(μμc)+O(ε),\displaystyle\frac{1}{F_{*}^{1}}\;\frac{M_{0}}{\gamma+H_{0}}(\mu-\mu_{c})+O(\varepsilon)\;=\;\frac{\beta_{2}D[\lambda K_{1}+\rho_{3}(\gamma+H_{0})]}{\rho_{3}k_{1}(\gamma+H_{0})^{2}}(\mu-\mu_{c})+O(\varepsilon), (2.16)
ρ4μ\displaystyle\frac{\partial\rho_{4}}{\partial\mu} =\displaystyle= 1F1M0γ+H0+O(ε)=β2D[λK1+ρ3(γ+H0)]ρ3k1(γ+H0)2+O(ε).\displaystyle\frac{1}{F_{*}^{1}}\;\frac{M_{0}}{\gamma+H_{0}}+O(\varepsilon)\;=\;\frac{\beta_{2}D[\lambda K_{1}+\rho_{3}(\gamma+H_{0})]}{\rho_{3}k_{1}(\gamma+H_{0})^{2}}+O(\varepsilon). (2.17)
Remark 2.3.

In contrast to [16, 23, 27], where σ~\widetilde{\sigma} is independent of μ\mu, here the explicit dependence of ρ4\rho_{4} with respect to μ\mu is given in the above lemma.

The following estimates are useful later on:

Lemma 2.4.

The following estimate holds for first derivatives,

|L(r)|+|H(r)|+|F(r)|+|p(r)|Cε,1εr1.|L_{*}^{\prime}(r)|+|H_{*}^{\prime}(r)|+|F_{*}^{\prime}(r)|+|p_{*}^{\prime}(r)|\leq C\varepsilon,\hskip 10.00002pt1-\varepsilon\leq r\leq 1. (2.18)
Proof.

From ((2.1))((2.1)) we derive that |ΔL|C|\Delta L_{*}|\leq C, |ΔH|C|\Delta H_{*}|\leq C, |Δp|C|\Delta p_{*}|\leq C. Using the boundary condition L(1)=0L_{*}^{\prime}(1)=0, we find that

|rL(r)|=|r1(ξL(ξ))𝑑ξ|Cε,1εr1.|rL_{*}^{\prime}(r)|=\Big{|}\int_{r}^{1}(\xi L_{*}^{\prime}(\xi))^{\prime}d\xi\Big{|}\leq C\varepsilon,\hskip 10.00002pt1-\varepsilon\leq r\leq 1.

The estimates for H(r)H_{*}^{\prime}(r) and for p(r)p_{*}^{\prime}(r) are similar. Finally, for F(r)F_{*}^{\prime}(r), using the above estimates we find

|(rF(r))|C+CεDmax1εr1|rF(r)|.|(rF_{*}^{\prime}(r))^{\prime}|\leq C+\frac{C\varepsilon}{D}\max_{1-\varepsilon\leq r\leq 1}|rF_{*}^{\prime}(r)|.

We then integrate over (r,1)(r,1) and use F(1)=0F_{*}^{\prime}(1)=0 to derive

|rF(r)|Cε+Cε2Dmax1εr1|rF(r)|,|rF_{*}^{\prime}(r)|\leq C\varepsilon+\frac{C\varepsilon^{2}}{D}\max_{1-\varepsilon\leq r\leq 1}|rF_{*}^{\prime}(r)|,

which implies |rF(r)|Cε|rF_{*}^{\prime}(r)|\leq C\varepsilon. ∎

2.2 The Crandall-Rabinowitz theorem

Next we state a useful theorem which is critical in studying bifurcations.

Theorem 2.5.

(Crandall-Rabinowitz theorem, [4]) Let XX, YY be real Banach spaces and (,)\mathcal{F}(\cdot,\cdot) a CpC^{p} map, p3p\geq 3, of a neighborhood (0,μ0)(0,\mu_{0}) in X×X\times\mathbb{R} into YY. Suppose

  • (1)

    (0,μ)=0\mathcal{F}(0,\mu)=0 for all μ\mu in a neighborhood of μ0\mu_{0},

  • (2)

    Kerx(0,μ0)\mathrm{Ker}\,\mathcal{F}_{x}(0,\mu_{0}) is one dimensional space, spanned by x0x_{0},

  • (3)

    Imx(0,μ0)=Y1\mathrm{Im}\,\mathcal{F}_{x}(0,\mu_{0})=Y_{1} has codimension 1,

  • (4)

    μx(0,μ0)x0Y1\mathcal{F}_{\mu x}(0,\mu_{0})x_{0}\notin Y_{1}.

Then (0,μ0)(0,\mu_{0}) is a bifurcation point of the equation (x,μ)=0\mathcal{F}(x,\mu)=0 in the following sense: In a neighborhood of (0,μ0)(0,\mu_{0}) the set of solutions (x,μ)=0\mathcal{F}(x,\mu)=0 consists of two Cp2C^{p-2} smooth curves Γ1\Gamma_{1} and Γ2\Gamma_{2} which intersect only at the point (0,μ0)(0,\mu_{0}); Γ1\Gamma_{1} is the curve (0,μ)(0,\mu) and Γ2\Gamma_{2} can be parameterized as follows:

Γ2:(x(ε),μ(ε)),|ε| small, (x(0),μ(0))=(0,μ0),x(0)=x0.\Gamma_{2}:(x(\varepsilon),\mu(\varepsilon)),|\varepsilon|\text{ small, }(x(0),\mu(0))=(0,\mu_{0}),\;x^{\prime}(0)=x_{0}.

3 Bifurcations - preparations

Let’s consider a family of perturbed domains Ωτ={1ε+R~<r<1}\Omega_{\tau}=\{1-\varepsilon+\widetilde{R}<r<1\} and denote the corresponding inner boundary to be Γτ\Gamma_{\tau}, where R~=τS(θ)\widetilde{R}=\tau S(\theta), |τ|ε|\tau|\ll\varepsilon and |S|1|S|\leq 1. Let (L,H,F,p)(L,H,F,p) be the solution of

ΔL=k1(M0F)LK1+Lρ1L\displaystyle-\Delta L=-k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-\rho_{1}L\hskip 10.00002pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.1)
ΔH=k2HFK2+Fρ2H\displaystyle-\Delta H=-k_{2}\frac{HF}{K_{2}+F}-\rho_{2}H\hskip 10.00002pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.2)
DΔFFp=k1(M0F)LK1+Lk2HFK2+FλF(M0F)LM0(γ+H)+(ρ3ρ4)(M0F)FM0\displaystyle-D\Delta F-\nabla F\cdot\nabla p=k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-k_{2}\frac{HF}{K_{2}+F}-\lambda\frac{F(M_{0}-F)L}{M_{0}(\gamma+H)}+(\rho_{3}-\rho_{4})\frac{(M_{0}-F)F}{M_{0}}\hskip 10.00002pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.3)
Δp=1M0[λ(M0F)Lγ+Hρ3(M0F)ρ4F]\displaystyle-\Delta p=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F)L}{\gamma+H}-\rho_{3}(M_{0}-F)-\rho_{4}F\Big{]}\hskip 10.00002pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.4)
Lr=Hr=Fr=pr=0,\displaystyle\frac{\partial L}{\partial r}=\frac{\partial H}{\partial r}=\frac{\partial F}{\partial r}=\frac{\partial p}{\partial r}=0,\hskip 20.00003pt r=1,\displaystyle r=1, (3.5)
L𝐧+β1(LL0)=0,H𝐧+β1(HH0)=0,F𝐧+β2F=0\displaystyle\frac{\partial L}{\partial{\bf n}}+\beta_{1}(L-L_{0})=0,\hskip 10.00002pt\frac{\partial H}{\partial{\bf n}}+{\beta_{1}}(H-H_{0})=0,\hskip 10.00002pt\frac{\partial F}{\partial{\bf n}}+\beta_{2}F=0 on Γτ,\displaystyle\text{on }\Gamma_{\tau}, (3.6)
p=κ\displaystyle p=\kappa on Γτ.\displaystyle\text{on }\Gamma_{\tau}. (3.7)

The existence and uniqueness of such a solution is guaranteed by the following lemma.

Lemma 3.6.

Let SC4+α(Σ)S\in C^{4+\alpha}(\Sigma) (Σ\Sigma denotes the unit closed disk) with SC4+α(Σ)1\|S\|_{C^{4+\alpha}(\Sigma)}\leq 1. For sufficiently small ε\varepsilon and |τ|ε|\tau|\ll\varepsilon, there is a unique solution (L,H,F,p)(L,H,F,p) to the problem ((3.1))((3.7)).

Proof.

We shall use the contraction mapping principle to prove this lemma. Let

={(L,H,F); 0LL0, 0HH0, 0FM0}.\mathcal{M}=\{(L,H,F);\;0\leq L\leq L_{0},\,0\leq H\leq H_{0},\,0\leq F\leq M_{0}\}. (3.8)

Step 1. For each (L,M,F)(L,M,F)\in\mathcal{M}, we define a map :(L,H,F)(L^,H^,F^)\mathcal{L}:(L,H,F)\rightarrow(\widehat{L},\widehat{H},\widehat{F}) as follows: we first solve L^\widehat{L} and H^\widehat{H} from the elliptic equations

ΔL^=k1(M0F)L^K1+Lρ1L^\displaystyle-\Delta\widehat{L}=-k_{1}\frac{(M_{0}-F)\widehat{L}}{K_{1}+L}-\rho_{1}\widehat{L}\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau},
ΔH^=k2H^FK2+Fρ2H^\displaystyle-\Delta\widehat{H}=-k_{2}\frac{\widehat{H}F}{K_{2}+F}-\rho_{2}\widehat{H}\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau},

with the boundary conditions

L^r=H^r=0,\displaystyle\frac{\partial\widehat{L}}{\partial r}=\frac{\partial\widehat{H}}{\partial r}=0,\hskip 20.00003pt r=1,\displaystyle r=1,
L^𝐧+β1(L^L0)=H^𝐧+β1(H^H0)=0\displaystyle\frac{\partial\widehat{L}}{\partial{\bf n}}+{\beta_{1}}(\widehat{L}-L_{0})=\frac{\partial\widehat{H}}{\partial{\bf n}}+{\beta_{1}}(\widehat{H}-H_{0})=0\hskip 20.00003pt on Γτ.\displaystyle\text{on }\Gamma_{\tau}.

By the maximum principle, we clearly have

0L^L0,0H^H0in Ω¯τ.0\leq\widehat{L}\leq L_{0},\hskip 20.00003pt0\leq\widehat{H}\leq H_{0}\hskip 20.00003pt\text{in }\overline{\Omega}_{\tau}. (3.9)

We then define p^\widehat{p} by the solution of the system

Δp^=1M0[λ(M0F)Lγ+Hρ3(M0F)ρ4F]in Ωτ,\displaystyle-\Delta\widehat{p}=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F)L}{\gamma+H}-\rho_{3}(M_{0}-F)-\rho_{4}F\Big{]}\hskip 20.00003pt\text{in }\Omega_{\tau}, (3.10)
p^r|r=1=0,p^|Γτ=κ.\displaystyle\frac{\partial\widehat{p}}{\partial r}\Big{|}_{r=1}=0,\hskip 20.00003pt\widehat{p}\Big{|}_{\Gamma_{\tau}}=\kappa. (3.11)

Since L,H,FL,H,F are all bounded, the right-hand side of ((3.10)) is bounded under supremum norm, i.e.,

|Δ(p^+1)|C.|\Delta(\widehat{p}+1)|\leq C. (3.12)

Also, we use the mean-curvature formula, i.e.,

κ|Γτ=11ε+τ(1ε)2(S+Sθθ)+τ2f1,where f1C1+αCSC3+α(Σ),\kappa|_{\Gamma_{\tau}}=-\frac{1}{1-\varepsilon}+\frac{\tau}{(1-\varepsilon)^{2}}(S+S_{\theta\theta})+\tau^{2}f_{1},\hskip 10.00002pt\text{where }\|f_{1}\|_{C^{1+\alpha}}\leq C\|S\|_{C^{3+\alpha}(\Sigma)}, (3.13)

to derive that

p^+1C1+α(Γτ)Cε.\|\widehat{p}+1\|_{C^{1+\alpha}(\Gamma_{\tau})}\leq C\varepsilon. (3.14)

By the maximum principle,

p^+1L(Ωτ)C(ξ+ε)C(O(ε2)+ε)Cε,\|\widehat{p}+1\|_{L^{\infty}(\Omega_{\tau})}\leq C(\xi+\varepsilon)\leq C(O(\varepsilon^{2})+\varepsilon)\leq C\varepsilon, (3.15)

where ξ\xi is defined in Appendix 5.1. Next we are going to estimate p^C1\|\widehat{p}\|_{C^{1}} and show that it is actually independent of ε\varepsilon and τ\tau. To do that, we shall use the Schauder estimates; but before using the Schauder estimates directly, let’s apply the following transformation:

Tτ:r~=r12(ετS(θ))+1,θ~=θε,T_{\tau}:\widetilde{r}=\frac{r-1}{2(\varepsilon-\tau S(\theta))}+1,\hskip 20.00003pt\widetilde{\theta}=\frac{\theta}{\varepsilon},

and denote p~(r~,θ~)=p^(r,θ)1\widetilde{p}(\widetilde{r},\widetilde{\theta})=\widehat{p}(r,\theta)-1. Clearly, TτT_{\tau} maps Ωτ\Omega_{\tau} to a long stripe region (r~,θ~)[12,1]×[0,2πε](\widetilde{r},\widetilde{\theta})\in[\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}]. Based on the calculations from Appendix 5.2, p~\widetilde{p} satisfies

r~((1+A1)p~r~+A2p~θ~)θ~(A3p~r~+(1+A4)p~θ~)+A5p~r~+A6p~θ~=ε2f2,-\frac{\partial}{\partial\widetilde{r}}\Big{(}(1+A_{1})\frac{\partial\widetilde{p}}{\partial\widetilde{r}}+A_{2}\frac{\partial\widetilde{p}}{\partial\widetilde{\theta}}\Big{)}-\frac{\partial}{\partial\widetilde{\theta}}\Big{(}A_{3}\frac{\partial\widetilde{p}}{\partial\widetilde{r}}+(1+A_{4})\frac{\partial\widetilde{p}}{\partial\widetilde{\theta}}\Big{)}+A_{5}\frac{\partial\widetilde{p}}{\partial\widetilde{r}}+A_{6}\frac{\partial\widetilde{p}}{\partial\widetilde{\theta}}=\varepsilon^{2}f_{2},

where coefficients A1,A2,A3,A4CαA_{1},A_{2},A_{3},A_{4}\in C^{\alpha}, A5,A6A_{5},A_{6} are bounded, and f2=rM0[λ(M0F)Lγ+Hρ3(M0F)ρ4F]f_{2}=\frac{r}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F)L}{\gamma+H}-\rho_{3}(M_{0}-F)-\rho_{4}F\Big{]} is also bounded based on ((3.8)). Applying the interior sub-Schauder estimates (Theorem 8.32, [12]) on the region Ωi0:(r~,θ~)[12,1]×[θi02,θi0+2]\Omega_{i_{0}}:(\widetilde{r},\widetilde{\theta})\in[\frac{1}{2},1]\times[\theta_{i_{0}}-2,\theta_{i_{0}}+2], recalling also ((3.14)), we obtain

p~C1+α([12,1]×[θi01,θi0+1])\displaystyle\|\widetilde{p}\|_{C^{1+\alpha}([\frac{1}{2},1]\times[\theta_{i_{0}}-1,\theta_{i_{0}}+1])} Ci0(ε2f2L(Ωi0)+p~L(Ωi0)+p~C1+α({r~=12}))\displaystyle\leq C_{i_{0}}\Big{(}\varepsilon^{2}\|f_{2}\|_{L^{\infty}(\Omega_{i_{0}})}+\|\widetilde{p}\|_{L^{\infty}(\Omega_{i_{0}})}+\|\widetilde{p}\|_{C^{1+\alpha}(\{\widetilde{r}=\frac{1}{2}\})}\Big{)}
Ci0(ε2f2L([12,1]×[0,2πε])+p^+1L(Ωτ)+p^+1C1+α(Γτ))\displaystyle\leq C_{i_{0}}\Big{(}\varepsilon^{2}\|f_{2}\|_{L^{\infty}([\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}])}+\|\widehat{p}+1\|_{L^{\infty}(\Omega_{\tau})}+\|\widehat{p}+1\|_{C^{1+\alpha}(\Gamma_{\tau})}\Big{)}
C~i0ε,\displaystyle\leq\widetilde{C}_{i_{0}}\varepsilon,

where C~i0\widetilde{C}_{i_{0}} is independent of ε\varepsilon and τ\tau. We use a series of sets [12,1]×[θi01,θi0+1][\frac{1}{2},1]\times[\theta_{i_{0}}-1,\theta_{i_{0}}+1] to cover the whole region [12,1]×[0,2πε][\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}], as a result,

p~C1+α([12,1]×[0,2πε])Cε.\|\widetilde{p}\|_{C^{1+\alpha}([\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}])}\leq C\varepsilon.

We then relate p~\widetilde{p} with p^\widehat{p} to derive

p^+1C1(Ω¯τ)1εp~C1([12,1]×[0,2πε])1εp~C1+α([12,1]×[0,2πε])C,\|\widehat{p}+1\|_{C^{1}(\overline{\Omega}_{\tau})}\leq\frac{1}{\varepsilon}\|\widetilde{p}\|_{C^{1}([\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}])}\leq\frac{1}{\varepsilon}\|\widetilde{p}\|_{C^{1+\alpha}([\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}])}\leq C,

and hence

p^L(Ωτ)C,\|\nabla\widehat{p}\|_{L^{\infty}(\Omega_{\tau})}\leq C, (3.16)

where CC is independent of ε\varepsilon and τ\tau.

Finally, recalling equation ((3.3)), we define F^\widehat{F} as the solution to the equation

DΔF^F^p^=k1(M0F^)LK1+Lk2HF^K2+FλF^(M0F)LM0(γ+H)+ρ3M0(M0F^)Fρ4M0(M0F)F^,-D\Delta\widehat{F}-\nabla\widehat{F}\cdot\nabla\widehat{p}=k_{1}\frac{(M_{0}-\widehat{F})L}{K_{1}+L}-k_{2}\frac{H\widehat{F}}{K_{2}+F}-\lambda\frac{\widehat{F}(M_{0}-F)L}{M_{0}(\gamma+H)}+\frac{\rho_{3}}{M_{0}}(M_{0}-\widehat{F})F-\frac{\rho_{4}}{M_{0}}(M_{0}-F)\widehat{F}, (3.17)

with the boundary conditions

F^r|r=1=0,[F^𝒏+β2F^]|Γτ=0.\frac{\partial\widehat{F}}{\partial r}\Big{|}_{r=1}=0,\hskip 20.00003pt\Big{[}\frac{\partial\widehat{F}}{\partial{\bm{n}}}+\beta_{2}\widehat{F}\Big{]}\Big{|}_{\Gamma_{\tau}}=0. (3.18)

By the maximum principle, F^0in Ω¯τ,\widehat{F}\geq 0\hskip 5.0pt\text{in }\overline{\Omega}_{\tau},and, using this result, we employ the maximum principle again to derive the inequality M0F^0in Ω¯τ.M_{0}-\widehat{F}\geq 0\hskip 5.0pt\text{in }\overline{\Omega}_{\tau}. All together, these two inequalities indicate

0F^M0.0\leq\widehat{F}\leq M_{0}. (3.19)

In the next step, we claim that this bound for F^\widehat{F} can be improved. By ((3.8)) and ((3.19)), the right-hand side of equation ((3.17)) is bounded; assume the bound is constant CC. According to Appendix 5.1, C(ξ(r)+c1(β2,ε)+c2(β2,τ))C(\xi(r)+c_{1}(\beta_{2},\varepsilon)+c_{2}(\beta_{2},\tau)) can be a supersolution for F^\widehat{F}, hence the maximum principle leads to

F^L(Ωτ)C(ξ(r)+c1(β2,ε)+c2(β2,τ))L(Ωτ)C(εβ2+2β2|τ|+O(ε2))Cε.\Big{\|}\widehat{F}\Big{\|}_{L^{\infty}(\Omega_{\tau})}\leq\Big{\|}C\Big{(}\xi(r)+c_{1}(\beta_{2},\varepsilon)+c_{2}(\beta_{2},\tau)\Big{)}\Big{\|}_{L^{\infty}(\Omega_{\tau})}\leq C\Big{(}\frac{\varepsilon}{\beta_{2}}+\frac{2}{\beta_{2}}|\tau|+O(\varepsilon^{2})\Big{)}\leq C\varepsilon. (3.20)

After we show this, we can employ the sub-Schauder estimates on ((3.17))((3.18)) in a similar way as we did for p^\widehat{p} to obtain

F^L(Ωτ)C,\|\nabla\widehat{F}\|_{L^{\infty}(\Omega_{\tau})}\leq C, (3.21)

where CC is a constant which does not depend upon ε\varepsilon and τ\tau.

Above, we have shown that (L^,H^,F^)(\widehat{L},\widehat{H},\widehat{F})\in\mathcal{M}, which means \mathcal{L} maps \mathcal{M} into itself. We shall next prove that \mathcal{L} is a contraction.

Step 2. Suppose that (L^j,H^j,F^j)=(Lj,Hj,Fj)(\widehat{L}_{j},\widehat{H}_{j},\widehat{F}_{j})=\mathcal{L}(L_{j},H_{j},F_{j}) for j=1,2j=1,2, and set

𝒜=L1L2L(Ωτ)+H1H2L(Ωτ)+F1F2L(Ωτ),\displaystyle\mathcal{A}=\|L_{1}-L_{2}\|_{L^{\infty}(\Omega_{\tau})}+\|H_{1}-H_{2}\|_{L^{\infty}(\Omega_{\tau})}+\|F_{1}-F_{2}\|_{L^{\infty}(\Omega_{\tau})},
=L^1L^2L(Ωτ)+H^1H^2L(Ωτ)+F^1F^2L(Ωτ).\displaystyle\mathcal{B}=\|\widehat{L}_{1}-\widehat{L}_{2}\|_{L^{\infty}(\Omega_{\tau})}+\|\widehat{H}_{1}-\widehat{H}_{2}\|_{L^{\infty}(\Omega_{\tau})}+\|\widehat{F}_{1}-\widehat{F}_{2}\|_{L^{\infty}(\Omega_{\tau})}.

Based on our definitions of L^j\widehat{L}_{j}, H^j\widehat{H}_{j}, p^j\widehat{p}_{j}, F^j\widehat{F}_{j} in the first step and recalling ((3.16)) as well as ((3.21)), we derive, for some constant CC^{*},

|Δ(L^1L^2)|C(𝒜+),|Δ(H^1H^2)|C(𝒜+),\displaystyle|\Delta(\widehat{L}_{1}-\widehat{L}_{2})|\leq C^{*}(\mathcal{A}+\mathcal{B}),\hskip 20.00003pt|\Delta(\widehat{H}_{1}-\widehat{H}_{2})|\leq C^{*}(\mathcal{A}+\mathcal{B}),
|F^1|+|F^2|C,|p^1|+|p^2|C,|(p^1p^2)|C𝒜,\displaystyle|\nabla\widehat{F}_{1}|+|\nabla\widehat{F}_{2}|\leq C^{*},\quad|\nabla\widehat{p}_{1}|+|\nabla\widehat{p}_{2}|\leq C^{*},\quad|\nabla(\widehat{p}_{1}-\widehat{p}_{2})|\leq C^{*}\mathcal{A},
|DΔ(F^1F^2)+p^1(F^1F^2)|C(𝒜+).\displaystyle|D\Delta(\widehat{F}_{1}-\widehat{F}_{2})+\nabla\widehat{p}_{1}\cdot\nabla(\widehat{F}_{1}-\widehat{F}_{2})|\leq C^{*}(\mathcal{A}+\mathcal{B}).

Next we shall use the maximum principle to derive bounds for L^1L^2\widehat{L}_{1}-\widehat{L}_{2}, H^1H^2\widehat{H}_{1}-\widehat{H}_{2}, and F^1F^2\widehat{F}_{1}-\widehat{F}_{2}. To do that, we use the function ξ(r)+c1(β1,ε)+c2(β1,τ)\xi(r)+c_{1}(\beta_{1},\varepsilon)+c_{2}(\beta_{1},\tau) defined in Appendix 5.1. As a result,

|L^1L^2|C(𝒜+)(ξ+c1(β1,ε)+c2(β1,τ))L^1L^2L(Ωτ)C(𝒜+)(ε+|τ|),\displaystyle|\widehat{L}_{1}-\widehat{L}_{2}|\leq C^{*}(\mathcal{A}+\mathcal{B})(\xi+c_{1}(\beta_{1},\varepsilon)+c_{2}(\beta_{1},\tau))\Rightarrow\|\widehat{L}_{1}-\widehat{L}_{2}\|_{L^{\infty}(\Omega_{\tau})}\leq C^{**}(\mathcal{A}+\mathcal{B})(\varepsilon+|\tau|),
|H^1H^2|C(𝒜+)(ξ+c1(β1,ε)+c2(β1,τ))H^1H^2L(Ωτ)C(𝒜+)(ε+|τ|),\displaystyle|\widehat{H}_{1}-\widehat{H}_{2}|\leq C^{*}(\mathcal{A}+\mathcal{B})(\xi+c_{1}(\beta_{1},\varepsilon)+c_{2}(\beta_{1},\tau))\Rightarrow\|\widehat{H}_{1}-\widehat{H}_{2}\|_{L^{\infty}(\Omega_{\tau})}\leq C^{**}(\mathcal{A}+\mathcal{B})(\varepsilon+|\tau|),
|F^1F^2|C(𝒜+)(ξ+c1(β2,ε)+c2(β2,τ))F^1F^2L(Ωτ)C(𝒜+)(ε+|τ|),\displaystyle|\widehat{F}_{1}-\widehat{F}_{2}|\leq C^{*}(\mathcal{A}+\mathcal{B})(\xi+c_{1}(\beta_{2},\varepsilon)+c_{2}(\beta_{2},\tau))\Rightarrow\|\widehat{F}_{1}-\widehat{F}_{2}\|_{L^{\infty}(\Omega_{\tau})}\leq C^{**}(\mathcal{A}+\mathcal{B})(\varepsilon+|\tau|),

where both CC^{*} and CC^{**} are independent of ε\varepsilon and τ\tau. The above inequalities imply that

C(𝒜+)(ε+|τ|),\mathcal{B}\leq C^{**}(\mathcal{A}+\mathcal{B})(\varepsilon+|\tau|),

hence we obtain a contraction mapping by taking ε\varepsilon sufficiently small and |τ|ε|\tau|\ll\varepsilon so that

C(ε+|τ|)1C(ε+|τ|)<1.\frac{C^{**}(\varepsilon+|\tau|)}{1-C^{**}(\varepsilon+|\tau|)}<1.

The unique fixed point of the contraction mapping is the unique classical solution to the system ((3.1))((3.7)). ∎

With pp being uniquely determined in the system ((3.1))((3.7)), we define \mathcal{F} by

(τS,μ)=p𝐧|Γτ,\mathcal{F}(\tau S,\mu)=-\frac{\partial p}{\partial{\bf n}}\Big{|}_{\Gamma_{\tau}}, (3.22)

where μ=1ε[λL0ρ3(γ+H0)]\mu=\frac{1}{\varepsilon}[\lambda L_{0}-\rho_{3}(\gamma+H_{0})] is our bifurcation parameter. We know that (L,H,F,p)(L,H,F,p) is a symmetry-breaking stationary solution if and only if (τS,μ)=0\mathcal{F}(\tau S,\mu)=0.

In order to apply the Crandall-Rabinowitz theorem, we need to compute the Fréchet derivatives of \mathcal{F}. For a fixed small ε\varepsilon, we formally write (L,H,F,p)(L,H,F,p) as

L=L+τL1+O(τ2),\displaystyle L=L_{*}+\tau L_{1}+O(\tau^{2}), (3.23)
H=H+τH1+O(τ2),\displaystyle H=H_{*}+\tau H_{1}+O(\tau^{2}), (3.24)
F=F+τF1+O(τ2),\displaystyle F=F_{*}+\tau F_{1}+O(\tau^{2}), (3.25)
p=p+τp1+O(τ2).\displaystyle p=p_{*}+\tau p_{1}+O(\tau^{2}). (3.26)

In the following, we shall first justify ((3.23))((3.26)). The structure of the proofs is similar to that in [5, 16, 17, 20, 23, 27]. However, our problem is much more involved since the system ((3.1))((3.7)) is highly nonlinear and coupled, hence we shall use very delicate estimates and the continuation lemma (see Appendix 5.3) to tackle the problem.

3.1 First-order τ\tau estimates

Lemma 3.7.

Fix ε\varepsilon sufficiently small, if |τ|ε|\tau|\ll\varepsilon and SC4+α(Σ)1\|S\|_{C^{4+\alpha}(\Sigma)}\leq 1, then we have

max{LLL(Ωτ),HHL(Ωτ),FFL(Ωτ),ppL(Ωτ)}C|τ|SC4+α(Σ),\displaystyle\max\{\|L-L_{*}\|_{L^{\infty}(\Omega_{\tau})},\|H-H_{*}\|_{L^{\infty}(\Omega_{\tau})},\|F-F_{*}\|_{L^{\infty}(\Omega_{\tau})},\|p-p_{*}\|_{L^{\infty}(\Omega_{\tau})}\}\leq C|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},
max{(FF)L(Ωτ),(pp)L(Ωτ)}Cε|τ|SC4+α(Σ),\displaystyle\max\{\|\nabla(F-F_{*})\|_{L^{\infty}(\Omega_{\tau})},\|\nabla(p-p_{*})\|_{L^{\infty}(\Omega_{\tau})}\}\leq\frac{C}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},

where CC is independent of ε\varepsilon and τ\tau.

Proof.

We combine ((2.1))((2.7)) and ((3.1))((3.7)) to obtain the equations for LLL-L_{*}, HHH-H_{*}, FFF-F_{*} and ppp-p_{*}. For example, we have

Δ(LL)\displaystyle-\Delta(L-L_{*}) =\displaystyle= k1(M0F)LK1+Lρ1L+k1(M0F)LK1+L+ρ1L\displaystyle-k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-\rho_{1}L+k_{1}\frac{(M_{0}-F_{*})L_{*}}{K_{1}+L_{*}}+\rho_{1}L_{*}
=\displaystyle= [k1(M0F)K1(K1+L)(K1+L)ρ1](LL)+k1LK1+L(FF)\displaystyle\Big{[}-k_{1}\frac{(M_{0}-F)K_{1}}{(K_{1}+L)(K_{1}+L_{*})}-\rho_{1}\Big{]}(L-L_{*})+k_{1}\frac{L_{*}}{K_{1}+L_{*}}(F-F_{*})
\displaystyle\triangleq b1(r)(LL)+b2(r)(FF),\displaystyle b_{1}(r)(L-L_{*})+b_{2}(r)(F-F_{*}),

where b1(r)b_{1}(r) and b2(r)b_{2}(r) are both bounded since 0L,LL00\leq L_{*},L\leq L_{0}, 0H,HH00\leq H_{*},H\leq H_{0}, and 0F,FM00\leq F,F_{*}\leq M_{0} based on Lemma 3.6 and Lemma 3.1 in [9]. In addition, the boundary conditions for LLL-L_{*} are

(LL)r|r=1=0,\displaystyle\frac{\partial(L-L_{*})}{\partial r}\Big{|}_{r=1}=0,
((LL)𝒏+β1(LL))|Γτ=(Lrβ1L)|r=1ε+τS(Lrβ1L)|r=1ε+O(|τS|2).\displaystyle\Big{(}\frac{\partial(L-L_{*})}{\partial{\bm{n}}}+\beta_{1}(L-L_{*})\Big{)}\Big{|}_{\Gamma_{\tau}}=\Big{(}\frac{\partial L_{*}}{\partial r}-\beta_{1}L_{*}\Big{)}\Big{|}_{r=1-\varepsilon+\tau S}-\Big{(}\frac{\partial L_{*}}{\partial r}-\beta_{1}L_{*}\Big{)}\Big{|}_{r=1-\varepsilon}+O(|\tau S|^{2}).

Since L,H,FL_{*},H_{*},F_{*} are all bounded and |L|Cε|L_{*}^{\prime}|\leq C\varepsilon by ((2.18)), we know from the equation ((2.1)) that |L′′||L_{*}^{\prime\prime}| is bounded with a bounded independent of ε\varepsilon and τ\tau. Hence we can find a constant, we denote it by C~\widetilde{C}, which does not depend upon ε\varepsilon and τ\tau, such that

|((LL)𝒏+β1(LL))|Γτ|C~|τ|SC4+α(Σ).\bigg{|}\Big{(}\frac{\partial(L-L_{*})}{\partial{\bm{n}}}+\beta_{1}(L-L_{*})\Big{)}\Big{|}_{\Gamma_{\tau}}\bigg{|}\leq\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}. (3.28)

Similarly, we can write the equations of HHH-H_{*}, FFF-F_{*} and ppp-p_{*} as

Δ(HH)=b3(r)(HH)+b4(r)(FF)\displaystyle-\Delta(H-H_{*})=b_{3}(r)(H-H_{*})+b_{4}(r)(F-F_{*})\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.29)
DΔ(FF)p(FF)=F(pp)+b5(r)(LL)+b6(r)(HH)+b7(r)(FF)\displaystyle\begin{aligned} -D\Delta(F-F_{*})-\nabla p_{*}\cdot\nabla(F-F_{*})=&\nabla F\cdot\nabla(p-p_{*})+b_{5}(r)(L-L_{*})\\ &\quad+b_{6}(r)(H-H_{*})+b_{7}(r)(F-F_{*})\end{aligned}\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.30)
Δ(pp)=b8(r)(LL)+b9(r)(HH)+b10(r)(FF)\displaystyle-\Delta(p-p_{*})=b_{8}(r)(L-L_{*})+b_{9}(r)(H-H_{*})+b_{10}(r)(F-F_{*})\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (3.31)

where bi(r)b_{i}(r), i=3,,10i=3,\cdots,10 are all bounded, and it is shown earlier that FL\|\nabla F\|_{L^{\infty}} and pL\|\nabla p_{*}\|_{L^{\infty}} are bounded; for simplicity, we shall use the same constant C~\widetilde{C} to control FL\|\nabla F\|_{L^{\infty}} and pL\|\nabla p_{*}\|_{L^{\infty}}, namely,

FLC~,pLC~.\|\nabla F\|_{L^{\infty}}\leq\widetilde{C},\hskip 20.00003pt\|\nabla p_{*}\|_{L^{\infty}}\leq\widetilde{C}. (3.32)

Furthermore, the boundary conditions for HHH-H_{*}, FFF-F_{*} and ppp-p_{*} satisfy

(HH)r|r=1=(FF)r|r=1=(pp)r|r=1=0,\displaystyle\frac{\partial(H-H_{*})}{\partial r}\Big{|}_{r=1}=\frac{\partial(F-F_{*})}{\partial r}\Big{|}_{r=1}=\frac{\partial(p-p_{*})}{\partial r}\Big{|}_{r=1}=0, (3.33)
|((HH)𝒏+β1(HH))|Γτ|C~|τ|SC4+α(Σ),\displaystyle\bigg{|}\Big{(}\frac{\partial(H-H_{*})}{\partial{\bm{n}}}+\beta_{1}(H-H_{*})\Big{)}\Big{|}_{\Gamma_{\tau}}\bigg{|}\leq\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.34)
|((FF)𝒏+β2(FF))|Γτ|C~|τ|SC4+α(Σ),\displaystyle\bigg{|}\Big{(}\frac{\partial(F-F_{*})}{\partial{\bm{n}}}+\beta_{2}(F-F_{*})\Big{)}\Big{|}_{\Gamma_{\tau}}\bigg{|}\leq\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.35)
|(pp)|Γτ|C~|τ|SC4+α(Σ),\displaystyle\Big{|}(p-p_{*})|_{\Gamma_{\tau}}\Big{|}\leq\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.36)

where the last inequality is based on the formula of κ\kappa in ((3.13)).

Since the system is highyly coupled, it is not an easy job to prove the estimates in Lemma 3.7. To show that, we use the idea of continuation (Appendix 5.3). We multiply the right-hand sides of ((3.1))((3.31)) by δ\delta with 0δ10\leq\delta\leq 1, and we shall combine the proofs for the case δ=0\delta=0 as well as the case 0<δ10<\delta\leq 1.

We first assume that, in the case δ>0\delta>0, for some M1>0M_{1}>0 to be determined later on

max(LLL(Ωτ),HHL(Ωτ),FFL(Ωτ))2M1|τ|SC4+α(Σ),\displaystyle\max\Big{(}\|L-L_{*}\|_{L^{\infty}(\Omega_{\tau})},\|H-H_{*}\|_{L^{\infty}(\Omega_{\tau})},\|F-F_{*}\|_{L^{\infty}(\Omega_{\tau})}\Big{)}\leq 2M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.37)
(FF)L(Ωτ)2M1Csε|τ|SC4+α(Σ),\displaystyle\|\nabla(F-F_{*})\|_{L^{\infty}(\Omega_{\tau})}\leq\frac{2M_{1}C_{s}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.38)
ppL(Ωτ)3C~|τ|SC4+α(Σ),(pp)L(Ωτ)3CsC~ε|τ|SC4+α(Σ),\displaystyle\displaystyle\|p-p_{*}\|_{L^{\infty}(\Omega_{\tau})}\leq 3\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},\hskip 20.00003pt\|\nabla(p-p_{*})\|_{L^{\infty}(\Omega_{\tau})}\leq\frac{3C_{s}\widetilde{C}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.39)

where C~\widetilde{C} is from ((3.28)), ((3.32)), ((3.34))((3.36)), and CsC_{s} is a scaling factor which comes from applying the C1+αC^{1+\alpha} Schauder estimate as we did in Lemma 3.6; both C~\widetilde{C} and CsC_{s} are independent of ε\varepsilon and τ\tau.

Let’s first show ((3.39)). Based on ((3.37)), for the case δ>0\delta>0, the right-hand side of ((3.31)) is bounded, i.e.,

|Δ(pp)|2M1δ(b8L(Ωτ)+b9L(Ωτ)+b10L(Ωτ))|τ|SC4+α(Σ);|\Delta(p-p_{*})|\leq 2M_{1}\delta(\|b_{8}\|_{L^{\infty}(\Omega_{\tau})}+\|b_{9}\|_{L^{\infty}(\Omega_{\tau})}+\|b_{10}\|_{L^{\infty}(\Omega_{\tau})})|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}; (3.40)

notice that the above estimates is automatically valid in the case δ=0\delta=0 without the assumptions ((3.37)) since the right-hand side is zero. Let

ϕ1(r)=2C~|τ|SC4+α(Σ)cos(1rε),\phi_{1}(r)=2\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\cos\Big{(}\frac{1-r}{\varepsilon}\Big{)}, (3.41)

then

ϕ1(r)=2C~εsin(1rε)|τ|SC4+α(Σ),ϕ1′′(r)=2C~ε2cos(1rε)|τ|SC4+α(Σ),\phi_{1}^{\prime}(r)=\frac{2\widetilde{C}}{\varepsilon}\sin\Big{(}\frac{1-r}{\varepsilon}\Big{)}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},\hskip 20.00003pt\phi_{1}^{\prime\prime}(r)=-\frac{2\widetilde{C}}{\varepsilon^{2}}\cos\Big{(}\frac{1-r}{\varepsilon}\Big{)}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},

and

Δϕ1=[1εcos(1rε)sin(1rε)]2C~ε|τ|SC4+α(Σ),ϕ1|Γτ=2C~cos(1τSε)|τ|SC4+α(Σ),-\Delta\phi_{1}=\Big{[}\frac{1}{\varepsilon}\cos\Big{(}\frac{1-r}{\varepsilon}\Big{)}-\sin\Big{(}\frac{1-r}{\varepsilon}\Big{)}\Big{]}\frac{2\widetilde{C}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},\hskip 20.00003pt\phi_{1}\Big{|}_{\Gamma_{\tau}}=2\widetilde{C}\cos\Big{(}1-\frac{\tau S}{\varepsilon}\Big{)}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},

where C~\widetilde{C}, again, comes from ((3.28)), ((3.32)), ((3.34))((3.36)). Notice that cos10.54>1/2\cos 1\approx 0.54>1/2, and we use the smallness of ε\varepsilon to majorize the right-hand side of ((3.40)) for a supersolution for ppp-p_{*} when ε\varepsilon is small and |τ|ε|\tau|\ll\varepsilon, hence

ppL(Ωτ)2C~|τ|SC4+α(Σ).\|p-p_{*}\|_{L^{\infty}(\Omega_{\tau})}\leq 2\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}.

Using a scaling argument as in ((3.16)), we further have

(pp)L(Ωτ)2CsC~ε|τ|SC4+α(Σ).\|\nabla(p-p_{*})\|_{L^{\infty}(\Omega_{\tau})}\leq\frac{2C_{s}\widetilde{C}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}. (3.42)

In the next step, let’s consider LLL-L_{*} and HHH-H_{*}. It follows from the assumption ((3.37)) that

|Δ(LL)|CM1δ|τ|SC4+α(Σ),|Δ(HH)|CM1δ|τ|SC4+α(Σ),|\Delta(L-L_{*})|\leq CM_{1}\delta|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},\hskip 20.00003pt|\Delta(H-H_{*})|\leq CM_{1}\delta|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}, (3.43)

where CC is some universal constant. Recalling also ((3.32)) and ((3.42)), we have the following estimate for FFF-F_{*},

|Δ(FF)+1Dp(FF)|\displaystyle\Big{|}\Delta(F-F_{*})+\frac{1}{D}\nabla p_{*}\cdot\nabla(F-F_{*})\Big{|}\;\leq 1DF(pp)L+b5(r)D(LL)L\displaystyle\;\;\Big{\|}\frac{1}{D}\nabla F\cdot\nabla(p-p_{*})\Big{\|}_{L^{\infty}}+\Big{\|}\frac{b_{5}(r)}{D}(L-L_{*})\Big{\|}_{L^{\infty}} (3.44)
+b6(r)D(HH)L+b7(r)D(FF)L\displaystyle\hskip 40.00006pt+\Big{\|}\frac{b_{6}(r)}{D}(H-H_{*})\Big{\|}_{L^{\infty}}+\Big{\|}\frac{b_{7}(r)}{D}(F-F_{*})\Big{\|}_{L^{\infty}}
\displaystyle\;\leq (2CsεDC~2+CM1)δ|τ|SC4+α(Σ).\displaystyle\;\;\Big{(}\frac{2C_{s}}{\varepsilon D}\widetilde{C}^{2}+CM_{1}\Big{)}\delta|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}.

We use

ϕ2(r)=M1|τ|SC4+α(Σ)cos(M2(1r)ε),M2=12min(β1,β2),\phi_{2}(r)=M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\cos\Big{(}\frac{M_{2}(1-r)}{\sqrt{\varepsilon}}\Big{)},\hskip 20.00003ptM_{2}=\frac{1}{2}\min\Big{(}\sqrt{\beta_{1}},\sqrt{\beta_{2}}\Big{)}, (3.45)

as the supersolution with M1M_{1} given by

M1=max(8β1C~,8β2C~,32Csβ1DC~2,32Csβ2DC~2).M_{1}=\max\Big{(}\frac{8}{\beta_{1}}\widetilde{C},\;\frac{8}{\beta_{2}}\widetilde{C},\;\frac{32C_{s}}{\beta_{1}D}\widetilde{C}^{2},\;\frac{32C_{s}}{\beta_{2}D}\widetilde{C}^{2}\Big{)}. (3.46)

Taking derivatives of ϕ2\phi_{2} gives us

ϕ2(r)=M1M2ε|τ|SC4+α(Σ)sin(M2(1r)ε),ϕ2′′(r)=M1M22ε|τ|SC4+α(Σ)cos(M2(1r)ε).\displaystyle\phi_{2}^{\prime}(r)=M_{1}\frac{M_{2}}{\sqrt{\varepsilon}}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\sin\Big{(}\frac{M_{2}(1-r)}{\sqrt{\varepsilon}}\Big{)},\hskip 20.00003pt\phi_{2}^{\prime\prime}(r)=-M_{1}\frac{M_{2}^{2}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\cos\Big{(}\frac{M_{2}(1-r)}{\sqrt{\varepsilon}}\Big{)}.

It is clear that ϕ2(1)=0\phi_{2}^{\prime}(1)=0. Moreover, for the boundary condition at Γτ:r=1ε+τS\Gamma_{\tau}:r=1-\varepsilon+\tau S,

(ϕ2𝒏+β1ϕ2)|Γτ\displaystyle\Big{(}\frac{\partial\phi_{2}}{\partial{\bm{n}}}+\beta_{1}\phi_{2}\Big{)}\Big{|}_{\Gamma_{\tau}} =ϕ2(1ε+τS)+β1ϕ2(1ε+τS)+O(|τS|2)\displaystyle=-\phi_{2}^{\prime}(1-\varepsilon+\tau S)+\beta_{1}\phi_{2}(1-\varepsilon+\tau S)+O(|\tau S^{\prime}|^{2})
=[M2εsin(M2(ετS)ε)+β1cos(M2(ετS)ε)]M1|τ|SC4+α(Σ)+O(|τS|2).\displaystyle=\Big{[}-\frac{M_{2}}{\sqrt{\varepsilon}}\sin\Big{(}\frac{M_{2}(\varepsilon-\tau S)}{\sqrt{\varepsilon}}\Big{)}+\beta_{1}\cos\Big{(}\frac{M_{2}(\varepsilon-\tau S)}{\sqrt{\varepsilon}}\Big{)}\Big{]}M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}+O(|\tau S^{\prime}|^{2}).

Since sinxx\sin x\leq x and cosx1x22\cos x\geq 1-\frac{x^{2}}{2} for x0x\geq 0, we have, for 0<|τ|ε0<|\tau|\ll\varepsilon and ε\varepsilon small,

M2εsin(M2(ετS)ε)M22(1τεS)2M22,cos(M2(ετS)ε)1M222ε(ε2+τ2)34.\frac{M_{2}}{\sqrt{\varepsilon}}\sin\Big{(}\frac{M_{2}(\varepsilon-\tau S)}{\sqrt{\varepsilon}}\Big{)}\leq M_{2}^{2}\Big{(}1-\frac{\tau}{\varepsilon}S\Big{)}\leq 2M_{2}^{2},\hskip 20.00003pt\cos\Big{(}\frac{M_{2}(\varepsilon-\tau S)}{\sqrt{\varepsilon}}\Big{)}\geq 1-\frac{M_{2}^{2}}{2\varepsilon}(\varepsilon^{2}+\tau^{2})\geq\frac{3}{4}.

Then

(ϕ2𝒏+β1ϕ2)|Γτ\displaystyle\Big{(}\frac{\partial\phi_{2}}{\partial{\bm{n}}}+\beta_{1}\phi_{2}\Big{)}\Big{|}_{\Gamma_{\tau}} [2M22+34β1]M1|τ|SC4+α(Σ)+O(|τS|2)\displaystyle\geq\;\;\Big{[}-2M_{2}^{2}+\frac{3}{4}\beta_{1}\Big{]}M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}+O(|\tau S^{\prime}|^{2})
14β1M1|τ|SC4+α(Σ)+O(|τS|2)\displaystyle\geq\;\;\frac{1}{4}\beta_{1}M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}+O(|\tau S^{\prime}|^{2})
  2C~|τ|SC4+α(Σ)+O(|τS|2)C~|τ|SC4+α(Σ).\displaystyle\geq\;\;2\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}+O(|\tau S^{\prime}|^{2})\geq\widetilde{C}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}.

Next we consider the equations ((3.1)), ((3.29)), ((3.30)) in proving ϕ2\phi_{2} is a supersolution. Notice that

Δϕ2=ϕ2′′(r)1rϕ2(r)\displaystyle-\Delta\phi_{2}=-\phi_{2}^{\prime\prime}(r)-\frac{1}{r}\phi_{2}^{\prime}(r) =M1[M22εcos(M2(1r)ε)M2εrsin(M2(1r)ε)]|τ|SC4+α(Σ)\displaystyle=\;\;M_{1}\Big{[}\frac{M_{2}^{2}}{\varepsilon}\cos\Big{(}\frac{M_{2}(1-r)}{\sqrt{\varepsilon}}\Big{)}-\frac{M_{2}}{\sqrt{\varepsilon}r}\sin\Big{(}\frac{M_{2}(1-r)}{\sqrt{\varepsilon}}\Big{)}\Big{]}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}
M1[M22ε342M22r]|τ|SC4+α(Σ)\displaystyle\geq\;\;M_{1}\Big{[}\frac{M_{2}^{2}}{\varepsilon}\frac{3}{4}-\frac{2M_{2}^{2}}{r}\Big{]}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}
M1[M22ε344M22]|τ|SC4+α(Σ),r[1ε+τS,1],\displaystyle\geq\;\;M_{1}\Big{[}\frac{M_{2}^{2}}{\varepsilon}\frac{3}{4}-4M_{2}^{2}\Big{]}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},\hskip 10.00002ptr\in[1-\varepsilon+\tau S,1],

For ((3.43)), it is clear that Δϕ2max{|Δ(LL)|,|Δ(HH)|}-\Delta\phi_{2}\geq\max\{|\Delta(L-L_{*})|,|\Delta(H-H_{*})|\} since the leading order term in Δϕ2-\Delta\phi_{2} is 1ε\frac{1}{\varepsilon} and we can take ε\varepsilon small. Hence ϕ2\phi_{2} is a supersolution for LLL-L_{*} as well as for HHH-H_{*}. For ((3.44)), as is shown, the leading order term in bounding ((3.44)) is 2CsC~2εD|τ|SC4+α(Σ)\frac{2C_{s}\widetilde{C}^{2}}{\varepsilon D}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}; on the other hand,

Δϕ2M1[M22ε342M22r]|τ|SC4+α(Σ)1ε12M1M22|τ|SC4+α(Σ)4CsC~2εD|τ|SC4+α(Σ);-\Delta\phi_{2}\geq M_{1}\Big{[}\frac{M_{2}^{2}}{\varepsilon}\frac{3}{4}-\frac{2M_{2}^{2}}{r}\Big{]}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\geq\frac{1}{\varepsilon}\frac{1}{2}M_{1}M_{2}^{2}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}\geq\frac{4C_{s}\widetilde{C}^{2}}{\varepsilon D}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)};

the extra term 1Dpϕ2\frac{1}{D}\nabla p_{*}\cdot\nabla\phi_{2} is of order O(1/ε)O(1/\sqrt{\varepsilon}) and therefore does not cause a problem. Thus ϕ2\phi_{2} is also a supersolution for FFF-F_{*}.

From the above analysis, we see that the choice of M1M_{1} and M2M_{2} depends only on β1\beta_{1}, β2\beta_{2}, C~\widetilde{C} and CsC_{s}, and is therefore independent of ε\varepsilon and τ\tau. By the maximum principle,

LLL(Ωτ)\displaystyle\|L-L_{*}\|_{L^{\infty}(\Omega_{\tau})} \displaystyle\leq M1|τ|SC4+α(Σ),\displaystyle M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},
HHL(Ωτ)\displaystyle\|H-H_{*}\|_{L^{\infty}(\Omega_{\tau})} \displaystyle\leq M1|τ|SC4+α(Σ),\displaystyle M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},
FFL(Ωτ)\displaystyle\|F-F_{*}\|_{L^{\infty}(\Omega_{\tau})} \displaystyle\leq M1|τ|SC4+α(Σ).\displaystyle M_{1}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}.

Using a scaling argument, we further have

(FF)L(Ωτ)M1Csε|τ|SC4+α(Σ).\|\nabla(F-F_{*})\|_{L^{\infty}(\Omega_{\tau})}\;\leq\;\frac{M_{1}C_{s}}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}.

These estimates are valid in the case δ=0\delta=0 without the assumptions ((3.37))((3.39)) since the right-hand sides are all zero in this case. Conditions (i) and (ii) of Lemma 5.20 are therefore satisfied for the vectors {1M1LLL,1M1HHL,1M1FFL,εM1Cs(FF)L,12C~ppL,ε2CsC~(pp)L,}\Big{\{}\frac{1}{M_{1}}\|L-L_{*}\|_{L^{\infty}},\frac{1}{M_{1}}\|H-H_{*}\|_{L^{\infty}},\frac{1}{M_{1}}\|F-F_{*}\|_{L^{\infty}},\frac{\varepsilon}{M_{1}C_{s}}\|\nabla(F-F_{*})\|_{L^{\infty}},\frac{1}{2\widetilde{C}}\|p-p_{*}\|_{L^{\infty}},\frac{\varepsilon}{2C_{s}\widetilde{C}}\|\nabla(p-p_{*})\|_{L^{\infty}},\Big{\}}. Since condition (iii) is obvious, we finish the proof. ∎

Remark 3.1.

Based on Lemma 3.7, if we further apply the Schauder estimates on the equations for LLL-L_{*}, HHH-H_{*}, FFF-F_{*}, and ppp-p_{*}, we can actually obtain

LLC4+α(Ω¯τ)+HHC4+α(Ω¯τ)+FFC4+α(Ω¯τ)+ppC2+α(Ω¯τ)C|τ|SC4+α(Σ),\|L-L_{*}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}+\|H-H_{*}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}+\|F-F_{*}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}+\|p-p_{*}\|_{C^{2+\alpha}(\overline{\Omega}_{\tau})}\leq C|\tau|\|S\|_{C^{4+\alpha}(\Sigma)},

where CC is independent of τ\tau, but is dependent upon ε\varepsilon.

3.2 Computation of L1L_{1}, H1H_{1}, F1F_{1} and p1p_{1}

In general, if f(y)f(y) (yRNy\in R^{N}, fRMf\in R^{M}) is a C2C^{2} function with bounded second order derivatives, then we have the Taylor’s expansion:

f(y)f(y)=01ddtf(ty+(1t)y)𝑑t=(01f(ty+(1t)y)𝑑t)(yy)=f(y)(yy)+R,\begin{array}[]{rcl}f(y)-f(y_{*})&=&\displaystyle\int_{0}^{1}\frac{d}{dt}f\big{(}ty+(1-t)y_{*}\big{)}dt\;=\;\Big{(}\int_{0}^{1}\nabla f\big{(}ty+(1-t)y_{*}\big{)}dt\Big{)}\cdot(y-y_{*})\\ &=&\nabla f(y_{*})\cdot(y-y_{*})+R,\end{array} (3.47)

where the remainder RR, given by R=01(f(ty+(1t)y)f(y))𝑑t(yy),R=\displaystyle\int_{0}^{1}\Big{(}\nabla f\big{(}ty+(1-t)y_{*}\big{)}-\nabla f\big{(}y_{*}\big{)}\Big{)}dt\cdot(y-y_{*}), satisfies

|R|01D2fL|yy|t𝑑t|yy|=12D2fL|yy|2.|R|\leq\int_{0}^{1}\|D^{2}f\|_{L^{\infty}}|y-y_{*}|\;tdt\cdot|y-y_{*}|=\frac{1}{2}\|D^{2}f\|_{L^{\infty}}|y-y_{*}|^{2}. (3.48)

Thus we have:

Lemma 3.8.

Suppose P\mathscript{P} is a linear operator, P[y]=f(y)\mathscript{P}[y]=f(y), P[y]=f(y)\mathscript{P}[y_{*}]=f(y_{*}). Let y1y_{1} be the linearized solution, i.e., P[y1]=f(y)y1\mathscript{P}[y_{1}]=\nabla f(y_{*})\cdot y_{1}. Then

P[yyτy1]=f(y)(yyτy1)+R,\mathscript{P}[y-y_{*}-\tau y_{1}]=\nabla f(y_{*})\cdot(y-y_{*}-\tau y_{1})+R, (3.49)

where by ((3.47)),

|R|12D2fL|yy|2.|R|\;\leq\;\frac{1}{2}\|D^{2}f\|_{L^{\infty}}|y-y_{*}|^{2}. (3.50)

Later on we shall apply this formula with y=(L,H,F,p)y=(L,H,F,p) and y=(L,H,F,p)y_{*}=(L_{*},H_{*},F_{*},p_{*}). Notice that by Lemma 3.7, |yy|=O(τ)|y-y_{*}|=O(\tau), thus we already have |yy|2=O(τ2)|y-y_{*}|^{2}=O(\tau^{2}). In what follows, we only need to produce correction terms for the linear part of the system, i.e., we shall compute the functions for L1L_{1}, H1H_{1}, F1F_{1} and p1p_{1}. Substituting ((3.23))((3.26)) into ((3.1))((3.7)), and dropping higher order terms of τ\tau, we obtain the linearized system. This is equivalent to taking total differential of the right-hand side ff with respect to L,H,FL,H,F and pp. If we write f=(fL,fH,fF,fp)Tf=(f^{L},f^{H},f^{F},f^{p})^{T}, then, from ((1.15)), fL(L,H,F,p)=k1(M0F)LK1+Lρ1Lf^{L}(L,H,F,p)=-k_{1}\frac{(M_{0}-F)L}{K_{1}+L}-\rho_{1}L, so that

fL(L,H,F,p)(L1,H1,F1,p1)= k1(M0F)K1L1(K1+L)2+k1LF1K1+Lρ1L1,\nabla f^{L}(L_{*},H_{*},F_{*},p_{*})\cdot(L_{1},H_{1},F_{1},p_{1})=\mbox{ $-k_{1}\frac{(M_{0}-F_{*})K_{1}L_{1}}{(K_{1}+L_{*})^{2}}+k_{1}\frac{L_{*}F_{1}}{K_{1}+L_{*}}-\rho_{1}L_{1},$}

and this is the right-hand side of the equation for L1L_{1}. Similar equations are derived for H1H_{1} and p1p_{1}. The right-hand side for F1F_{1} is similar, but we have to take care of the additional gradient terms in the left-hand side. In summary, we obtain the following linearized system on Ω={1ε<r<1}\Omega_{*}=\{1-\varepsilon<r<1\}:

ΔL1=k1(M0F)K1L1(K1+L)2+k1LF1K1+Lρ1L1\displaystyle-\Delta L_{1}=-k_{1}\frac{(M_{0}-F_{*})K_{1}L_{1}}{(K_{1}+L_{*})^{2}}+k_{1}\frac{L_{*}F_{1}}{K_{1}+L_{*}}-\rho_{1}L_{1}\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (3.51)
ΔH1=k2K2HF1(K2+F)2k2FH1K2+Fρ2H1\displaystyle-\Delta H_{1}=-k_{2}\frac{K_{2}H_{*}F_{1}}{(K_{2}+F_{*})^{2}}-k_{2}\frac{F_{*}H_{1}}{K_{2}+F_{*}}-\rho_{2}H_{1}\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (3.52)
DΔF1F1pFp1=k1(M0F)K1L1(K1+L)2k1F1LK1+L+\displaystyle-D\Delta F_{1}-\nabla F_{1}\cdot\nabla p_{*}-\nabla F_{*}\cdot\nabla p_{1}=k_{1}\frac{(M_{0}-F_{*})K_{1}L_{1}}{(K_{1}+L_{*})^{2}}-k_{1}\frac{F_{1}L_{*}}{K_{1}+L_{*}}+\cdots\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (3.53)
Δp1=1M0[λ(M0F)L1γ+HλLF1γ+Hλ(M0F)LH1(γ+H)2+(ρ3ρ4)F1]\displaystyle-\Delta p_{1}=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F_{*})L_{1}}{\gamma+H_{*}}-\lambda\frac{L_{*}F_{1}}{\gamma+H_{*}}-\lambda\frac{(M_{0}-F_{*})L_{*}H_{1}}{(\gamma+H_{*})^{2}}+(\rho_{3}-\rho_{4})F_{1}\Big{]}\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (3.54)
L1r=H1r=F1r=p1r=0\displaystyle\frac{\partial L_{1}}{\partial r}=\frac{\partial H_{1}}{\partial r}=\frac{\partial F_{1}}{\partial r}=\frac{\partial p_{1}}{\partial r}=0\hskip 20.00003pt r=1,\displaystyle r=1, (3.55)
L1r+β1L1=(2Lr2β1Lr)|r=1εS(θ)\displaystyle-\frac{\partial L_{1}}{\partial r}+\beta_{1}L_{1}=\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}S(\theta) r=1ε,\displaystyle r=1-\varepsilon, (3.56)
H1r+β1H1=(2Hr2β1Hr)|r=1εS(θ)\displaystyle-\frac{\partial H_{1}}{\partial r}+\beta_{1}H_{1}=\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}S(\theta) r=1ε,\displaystyle r=1-\varepsilon, (3.57)
F1r+β2F1=(2Fr2β2Fr)|r=1εS(θ)\displaystyle-\frac{\partial F_{1}}{\partial r}+\beta_{2}F_{1}=\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}S(\theta) r=1ε,\displaystyle r=1-\varepsilon, (3.58)
p1=1(1ε)2(S+Sθθ)\displaystyle p_{1}=\frac{1}{(1-\varepsilon)^{2}}(S+S_{\theta\theta}) r=1ε.\displaystyle r=1-\varepsilon. (3.59)

Using the same techniques as in the proof of Lemma 3.7, also recalling Remark 3.1, we can derive L1,H1,F1C4+α(Ω¯τ)L_{1},H_{1},F_{1}\in C^{4+\alpha}(\overline{\Omega}_{\tau}) and p1C2+α(Ω¯τ)p_{1}\in C^{2+\alpha}(\overline{\Omega}_{\tau}); their Schauder estimates may depend on ε\varepsilon, but it is crucial that the LL^{\infty} estimates are independent of ε\varepsilon and τ\tau.

Notice that L1L_{1}, H1H_{1}, F1F_{1} and p1p_{1} are all defined in Ω\Omega_{*}, while LLL-L_{*}, HHH-H_{*}, FFF-F_{*} and ppp-p_{*} are defined in Ωτ\Omega_{\tau}. We would now like to transform L1L_{1}, H1H_{1}, F1F_{1} and p1p_{1} from Ω\Omega_{*} to Ωτ\Omega_{\tau} so that we are able to work on the same domain to derive second-order τ\tau estimates. To do that, we define a transform

Yτ:(r,θ)=Yτ(r¯,θ¯)=((r¯1)(ετS)ε+1,θ)Y_{\tau}:\;(r,\theta)=Y_{\tau}(\overline{r},\overline{\theta})=\Big{(}\frac{(\overline{r}-1)(\varepsilon-\tau S)}{\varepsilon}+1\,,\,\theta\Big{)} (3.60)

and let

L¯1(r,θ)=L1(Yτ1(r,θ)),\displaystyle\overline{L}_{1}(r,\theta)=L_{1}(Y_{\tau}^{-1}(r,\theta)),\hskip 20.00003pt H¯1(r,θ)=H1(Yτ1(r,θ)),\displaystyle\overline{H}_{1}(r,\theta)=H_{1}(Y_{\tau}^{-1}(r,\theta)), (3.61)
F¯1(r,θ)=F1(Yτ1(r,θ)),\displaystyle\overline{F}_{1}(r,\theta)=F_{1}(Y_{\tau}^{-1}(r,\theta)),\hskip 20.00003pt p¯1(r,θ)=p1(Yτ1(r,θ)),\displaystyle\overline{p}_{1}(r,\theta)=p_{1}(Y_{\tau}^{-1}(r,\theta)), (3.62)

for (r,θ)Ωτ(r,\theta)\in\Omega_{\tau}. Similar as using the Hanzawa transformation in [5, 16, 17, 20, 23, 27], the error incurred from applying YτY_{\tau} is less than |τS||\tau S|.

3.3 Second-order τ\tau estimates

The first step in deriving second-order τ\tau estimates is to calculate the equations for LLτL¯1L-L_{*}-\tau\overline{L}_{1}, HHτH¯1H-H_{*}-\tau\overline{H}_{1}, FFτF¯1F-F_{*}-\tau\overline{F}_{1} and ppτp¯1p-p_{*}-\tau\overline{p}_{1}. Here we shall only show the derivations of the equation for FFτF¯1F-F_{*}-\tau\overline{F}_{1}, since the equation for FF is more complex than those for other variables.

Combining the equations for F,F,F_{*},F, and F1F_{1} respectively in ((2.3)) ((3.3)) and ((3.53)), we derive

DΔ(FFτF1)Fp+Fp+τF1p+τFp1=RHS.-D\Delta(F-F_{*}-\tau F_{1})-\nabla F\cdot\nabla p+\nabla F_{*}\cdot\nabla p_{*}+\tau\nabla F_{1}\cdot\nabla p_{*}+\tau\nabla F_{*}\cdot\nabla p_{1}=\text{RHS}. (3.63)

By Lemma 3.8, the right-hand side of ((3.63)) satisfies

RHS=[I]+[II],\text{RHS}=[\text{I}]+[\text{II}],

where I is written as, for bounded functions b11(r),b12(r),b_{11}(r),b_{12}(r), and b13(r)b_{13}(r),

I=b11(r)(LLτL1)+b12(r)(HHτH1)+b13(r)(FFτF1);\text{I}=b_{11}(r)(L-L_{*}-\tau L_{1})+b_{12}(r)(H-H_{*}-\tau H_{1})+b_{13}(r)(F-F_{*}-\tau F_{1});

and II is bounded by |(LL,HH,FF)|2|(L-L_{*},H-H_{*},F-F_{*})|^{2}, hence

IILC|τ|2SC4+α(Σ).\|\text{II}\|_{L^{\infty}}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}.

We then turn to the left-hand side of equation ((3.63)). The terms involving the gradients can be rearranged as

Fp+Fp+τF1p+τFp1\displaystyle-\nabla F\cdot\nabla p+\nabla F_{*}\cdot\nabla p_{*}+\tau\nabla F_{1}\cdot\nabla p_{*}+\tau\nabla F_{*}\cdot\nabla p_{1}
=\displaystyle= p(FFτF1)F(ppτp1)τ(FF)p1.\displaystyle-\nabla p_{*}\cdot\nabla(F-F_{*}-\tau F_{1})-\nabla F\cdot\nabla(p-p_{*}-\tau p_{1})-\tau\nabla(F-F_{*})\cdot\nabla p_{1}.

By Lemma 3.7,

(FF)LCε|τ|SC4+α(Σ);\|\nabla(F-F_{*})\|_{L^{\infty}}\leq\frac{C}{\varepsilon}|\tau|\|S\|_{C^{4+\alpha}(\Sigma)}; (3.64)

furthermore, we can derive from ((3.54)) and ((3.59)) that

|Δ(p1(S+Sθθ))|C,andp1(S+Sθθ)C1+α({r=1ε})Cε,|\Delta(p_{1}-(S+S_{\theta\theta}))|\leq C,\quad\text{and}\quad\|p_{1}-(S+S_{\theta\theta})\|_{C^{1+\alpha}(\{r=1-\varepsilon\})}\leq C\varepsilon,

as SC4+αS\in C^{4+\alpha}; using the same technique as in Lemma 3.6, we shall get

(p1(S+Sθθ))L(Ω)C,\|\nabla(p_{1}-(S+S_{\theta\theta}))\|_{L^{\infty}(\Omega_{*})}\leq C,

hence

p1L(Ω)C,\|\nabla p_{1}\|_{L^{\infty}(\Omega_{*})}\leq C,

for a constant CC which is independent of ε\varepsilon and τ\tau. Together with ((3.64)), we derive

τ(FF)p1LCε|τ|2SC4+α(Σ).\|\tau\nabla(F-F_{*})\cdot\nabla p_{1}\|_{L^{\infty}}\leq\frac{C}{\varepsilon}|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}.

From the above analysis, we obtain the equation for FFτF1F-F_{*}-\tau F_{1},

DΔ(FFτF1)p(FFτF1)\displaystyle-D\Delta(F-F_{*}-\tau F_{1})-\nabla p_{*}\cdot\nabla(F-F_{*}-\tau F_{1})
=F(ppτp1)τ(FF)p1+[I]+[II].\displaystyle\hskip 100.00015pt=\nabla F\cdot\nabla(p-p_{*}-\tau p_{1})-\tau\nabla(F-F_{*})\cdot\nabla p_{1}+[\text{I}]+[\text{II}].

Now we recall the transform YτY_{\tau} in ((3.60)) and the change of variables in ((3.61)) and ((3.62)), we can derive the equation for FFτF¯1F-F_{*}-\tau\overline{F}_{1}, namely,

DΔ(FFτF¯1)p(FFτF¯1)\displaystyle-D\Delta(F-F_{*}-\tau\overline{F}_{1})-\nabla p_{*}\cdot\nabla(F-F_{*}-\tau\overline{F}_{1}) (3.65)
=F(ppτp¯1)τ(FF)p¯1+[I]+[II]+τf4,\displaystyle\hskip 100.00015pt=\nabla F\cdot\nabla(p-p_{*}-\tau\overline{p}_{1})-\tau\nabla(F-F_{*})\cdot\nabla\overline{p}_{1}+[\text{I}]+[\text{II}]+\tau f_{4},

where f4f_{4} is generated by the tiny changing of domain from Ω\Omega_{*} to Ωτ\Omega_{\tau} in applying the transformation YτY_{\tau}, and it contains at most second derivatives of τS\tau S, hence

τf4L(Ωτ)|τ|C|τ|SC2+α(Ωτ)C|τ|2SC4+α(Ωτ).\|\tau f_{4}\|_{L^{\infty}(\Omega_{\tau})}\leq|\tau|\cdot C|\tau|\|S\|_{C^{2+\alpha}(\Omega_{\tau})}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Omega_{\tau})}.

Combining with the estimates we derived before, we have

|Δ(FFτF¯1)+1Dp(FFτF¯1)|1DF(ppτp¯1)L\displaystyle\Big{|}\Delta(F-F_{*}-\tau\overline{F}_{1})+\frac{1}{D}\nabla p_{*}\cdot\nabla(F-F_{*}-\tau\overline{F}_{1})\Big{|}\;\leq\;\Big{\|}\frac{1}{D}\nabla F\cdot\nabla(p-p_{*}-\tau\overline{p}_{1})\Big{\|}_{L^{\infty}}
+b11(r)D(LLτL¯1)L+b12(r)D(HHτH¯1)L\displaystyle\hskip 150.00023pt+\Big{\|}\frac{b_{11}(r)}{D}(L-L_{*}-\tau\overline{L}_{1})\Big{\|}_{L^{\infty}}+\Big{\|}\frac{b_{12}(r)}{D}(H-H_{*}-\tau\overline{H}_{1})\Big{\|}_{L^{\infty}}
+b13(r)D(FFτF¯1)L+Cε|τ|2SC4+α(Σ).\displaystyle\hskip 150.00023pt+\Big{\|}\frac{b_{13}(r)}{D}(F-F_{*}-\tau\overline{F}_{1})\Big{\|}_{L^{\infty}}+\frac{C}{\varepsilon}|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}.

Notice that the above inequality present similar structure as ((3.44)), hence we can use the same technique and similar supersolutions to establish

Lemma 3.9.

Fix ε\varepsilon sufficiently small, if |τ|ε|\tau|\ll\varepsilon and SC4+α(Σ)1\|S\|_{C^{4+\alpha}(\Sigma)}\leq 1, then we have

max{LLτL¯1L(Ωτ),HHτH¯1L(Ωτ)}C|τ|2SC4+α(Σ),\displaystyle\max\{\|L-L_{*}-\tau\overline{L}_{1}\|_{L^{\infty}(\Omega_{\tau})},\|H-H_{*}-\tau\overline{H}_{1}\|_{L^{\infty}(\Omega_{\tau})}\}\;\leq\;C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)},
max{FFτF¯1L(Ωτ),ppτp¯1L(Ωτ)}C|τ|2SC4+α(Σ),\displaystyle\max\{\|F-F_{*}-\tau\overline{F}_{1}\|_{L^{\infty}(\Omega_{\tau})},\|p-p_{*}-\tau\overline{p}_{1}\|_{L^{\infty}(\Omega_{\tau})}\}\;\leq\;C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)},
max{(FFτF¯1)L(Ωτ),(ppτp¯1)L(Ωτ)}Cε|τ|2SC4+α(Σ),\displaystyle\max\{\|\nabla(F-F_{*}-\tau\overline{F}_{1})\|_{L^{\infty}(\Omega_{\tau})},\|\nabla(p-p_{*}-\tau\overline{p}_{1})\|_{L^{\infty}(\Omega_{\tau})}\}\;\leq\;\frac{C}{\varepsilon}|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)},

where CC is independent of ε\varepsilon and τ\tau.

Following Remark 3.1, we shall further have

Lemma 3.10.

Fix ε\varepsilon sufficiently small, if |τ|ε|\tau|\ll\varepsilon and SC4+α(Σ)1\|S\|_{C^{4+\alpha}(\Sigma)}\leq 1, then

LLτL¯1C4+α(Ω¯τ)C|τ|2SC4+α(Σ),\displaystyle\|L-L_{*}-\tau\overline{L}_{1}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}, (3.66)
HHτH¯1C4+α(Ω¯τ)C|τ|2SC4+α(Σ),\displaystyle\|H-H_{*}-\tau\overline{H}_{1}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}, (3.67)
FFτF¯1C4+α(Ω¯τ)C|τ|2SC4+α(Σ),\displaystyle\|F-F_{*}-\tau\overline{F}_{1}\|_{C^{4+\alpha}(\overline{\Omega}_{\tau})}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}, (3.68)
ppτp¯1C2+α(Ω¯τ)C|τ|2SC4+α(Σ),\displaystyle\|p-p_{*}-\tau\overline{p}_{1}\|_{C^{2+\alpha}(\overline{\Omega}_{\tau})}\leq C|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}, (3.69)

where CC is independent of τ\tau, but is dependent on ε\varepsilon.

The estimates ((3.66))((3.69)) are uniformly valid for |τ||\tau| small and SC4+α(Σ)1\|S\|_{C^{4+\alpha}(\Sigma)}\leq 1 . By now, we finish the mathematical justification of ((3.23))((3.26)), and we are ready to derive the Fréchet derivatives of \mathcal{F}.

3.4 Fréchet derivative

Introduce the Banach spaces

Xl+α={SCl+α(Σ),S is 2π-periodic in θ},\displaystyle X^{l+\alpha}=\{S\in C^{l+\alpha}(\Sigma),S\text{ is $2\pi$-periodic in $\theta$}\},
X1l+α=closure of the linear space spanned by {cos(nθ),n=0,1,2,} in Xl+α.\displaystyle X^{l+\alpha}_{1}=\text{closure of the linear space spanned by $\{\cos(n\theta),n=0,1,2,\cdots\}$ in $X^{l+\alpha}$}. (3.70)

It can be easily proved that the system ((3.1))((3.7)) is even in variable θ\theta if we assume S(θ)=S(θ)S(\theta)=S(-\theta). Together with ((3.69)), we know that the mapping (,μ):X1l+4+αX1l+1+α\mathcal{F}(\cdot,\mu):X^{l+4+\alpha}_{1}\rightarrow X^{l+1+\alpha}_{1} is bounded when l=0l=0, and the same argument can show that it is also true for any l>0l>0. In order to apply the Crandall-Rabinowitz theorem, we need to verify the continuous differentiability of \mathcal{F}. As will be shown in the following lemma, the differentiablity is eventually reduced to the regularity of the corresponding PDEs, and explicit formula is not needed if we are only interested in differentiability; therefore a similar argument shows that this mapping is Fréchet differentiable in (R~,μ)(\widetilde{R},\mu); furthermore (R~,μ)/R~\partial\mathcal{F}(\widetilde{R},\mu)/\partial\widetilde{R} (or (R~,μ)/μ\partial\mathcal{F}(\widetilde{R},\mu)/\partial\mu) is obtained by solving a linearized problem about (R~,μ)(\widetilde{R},\mu) with respect to R~\widetilde{R} (or μ\mu). By using the Schauder estimates we can then further obtain differentiability of (R~,μ)\mathcal{F}(\widetilde{R},\mu) to any order.

We now proceed to compute those Fréchet derivatives that are crucial in applying the Crandall-Rabinowitz theorem.

Lemma 3.11.

The Fréchet derivatives of (R~,μ)\mathcal{F}(\widetilde{R},\mu) at the point (0,μ)(0,\mu) are given by

[R~(0,μ)]S(θ)=2pr2|r=1εS(θ)+p1r|r=1ε,\displaystyle\Big{[}\mathcal{F}_{\widetilde{R}}(0,\mu)\Big{]}S(\theta)=\frac{\partial^{2}p_{*}}{\partial r^{2}}\Big{|}_{r=1-\varepsilon}S(\theta)+\frac{\partial p_{1}}{\partial r}\Big{|}_{r=1-\varepsilon}, (3.71)
[μR~(0,μ)]S(θ)=μ(2pr2|r=1ε)S(θ)+μ(p1r|r=1ε).\displaystyle\left[\mathcal{F}_{\mu\widetilde{R}}(0,\mu)\right]S(\theta)=\frac{\partial}{\partial\mu}\Big{(}\frac{\partial^{2}p_{*}}{\partial r^{2}}\Big{|}_{r=1-\varepsilon}\Big{)}S(\theta)+\frac{\partial}{\partial\mu}\Big{(}\frac{\partial p_{1}}{\partial r}\Big{|}_{r=1-\varepsilon}\Big{)}. (3.72)
Proof.

Since

pr|r=1ε=0,\frac{\partial p_{*}}{\partial r}\Big{|}_{r=1-\varepsilon}=0,

which implies (0,μ)=0\mathcal{F}(0,\mu)=0. For R~=τS\widetilde{R}=\tau S, it then follows from ((3.69)) that

(τS,μ)=p𝒏|Γτ\displaystyle\mathcal{F}(\tau S,\mu)=-\frac{\partial p}{\partial{\bm{n}}}\Big{|}_{\Gamma_{\tau}} =(p+τp1)r|r=1ε+τS+O(|τ|2SC4+α(Σ))\displaystyle=\frac{\partial(p_{*}+\tau p_{1})}{\partial r}\Big{|}_{r=1-\varepsilon+\tau S}+O(|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)})
=τ[2pr2|r=1εS(θ)+p1r|r=1ε]+O(|τ|2SC4+α(Σ)),\displaystyle=\tau\Big{[}\frac{\partial^{2}p_{*}}{\partial r^{2}}\Big{|}_{r=1-\varepsilon}S(\theta)+\frac{\partial p_{1}}{\partial r}\Big{|}_{r=1-\varepsilon}\Big{]}+O(|\tau|^{2}\|S\|_{C^{4+\alpha}(\Sigma)}),

which leads to the expression of the Fréchet derivative in ((3.71)), and ((3.72)) is a direct consequence of ((3.71)). ∎

4 Bifurcations - Proof of Theorem 1.1

In this section, we shall employ the explicit expression of the Fréchet derivative ((3.71)) to verify the four conditions in the Crandall-Rabinowitz theorem and complete the proof of Theorem 1.1. Unlike [5, 6, 11, 10, 16, 17, 20, 23, 24, 25], we cannot solve pp_{*} and p1p_{1} explicitly, since our model is highly nonlinear and coupled. To meet the challenges, we need to derive various sharp estimates on pp_{*} and p1p_{1}.

Throughout the rest of this paper, CC is used to represent a generic constant independent of ε\varepsilon, which might change from line to line.

4.1 Estimates for pp_{*}

In order to estimate 2p(1ε)r2\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}} in ((3.71)), we start with evaluating ((2.4)) at r=1εr=1-\varepsilon and substituting the boundary condition ((2.8)), hence we obtain

2p(1ε)r2=1M0(λ(M0F)Lγ+Hρ3(M0F)ρ4F)|r=1ε.-\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}}=\frac{1}{M_{0}}\Big{(}\lambda\frac{(M_{0}-F_{*})L_{*}}{\gamma+H_{*}}-\rho_{3}(M_{0}-F_{*})-\rho_{4}F_{*}\Big{)}\Big{|}_{r=1-\varepsilon}. (4.1)

Similar to the proof of Theorem 2.2, we substitute ((2.1))((2.1)) into the above formula and combine with ((2.16)), we find that both O(1)O(1) and O(ε)O(\varepsilon) terms cancel out, thus

2p(1ε)r2=εM0(M0γ+H0(μμc)ρ4F1)+O(ε2)=O(ε2).\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}}=\frac{\varepsilon}{M_{0}}\Big{(}\frac{M_{0}}{\gamma+H_{0}}(\mu-\mu_{c})-\rho_{4}F_{*}^{1}\Big{)}+O(\varepsilon^{2})=O(\varepsilon^{2}). (4.2)

Denote

J1(μ,ρ4)=1ε22p(1ε)r2, i.e., 2p(1ε)r2=ε2J1(μ,ρ4),J_{1}(\mu,\rho_{4})=\frac{1}{\varepsilon^{2}}\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}},\quad\text{ i.e., }\quad\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}}=\varepsilon^{2}J_{1}(\mu,\rho_{4}), (4.3)

it follows from ((4.2)) that J1(μ,ρ4)=O(1)J_{1}(\mu,\rho_{4})=O(1) is bounded. Besides, we claim that dJ1dμ=J1μ+J1ρ4ρ4μ=O(1)\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}=\frac{\partial J_{1}}{\partial\mu}+\frac{\partial J_{1}}{\partial\rho_{4}}\frac{\partial\rho_{4}}{\partial\mu}=O(1) is also bounded. To prove it, we take μ\mu derivative of equation ((4.2)), and derive

2r2(pμ)|r=1ε=ε(1γ+H0F1M0ρ4μ)+O(ε2).\frac{\partial^{2}}{\partial r^{2}}\Big{(}\frac{\partial p_{*}}{\partial\mu}\Big{)}\Big{|}_{r=1-\varepsilon}=\varepsilon\Big{(}\frac{1}{\gamma+H_{0}}-\frac{F_{*}^{1}}{M_{0}}\frac{\partial\rho_{4}}{\partial\mu}\Big{)}+O(\varepsilon^{2}).

By substituting the formula of ρ4μ\frac{\partial\rho_{4}}{\partial\mu} in ((2.17)), we find that the O(ε)O(\varepsilon) terms in the above equation cancel out, hence

dJ1(μ,ρ4(μ))dμ=1ε22r2(pμ)|r=1ε=1ε2O(ε2)=O(1).\frac{\mathrm{d}J_{1}(\mu,\rho_{4}(\mu))}{\mathrm{d}\mu}=\frac{1}{\varepsilon^{2}}\frac{\partial^{2}}{\partial r^{2}}\Big{(}\frac{\partial p_{*}}{\partial\mu}\Big{)}\Big{|}_{r=1-\varepsilon}=\frac{1}{\varepsilon^{2}}O(\varepsilon^{2})=O(1).

To sum up, the properties of J1J_{1} are listed in the following lemma:

Lemma 4.12.

For function J1(μ,ρ4)J_{1}(\mu,\rho_{4}) defined in ((4.3)), there exists a constant CC which is independent of ε\varepsilon such that

|J1(μ,ρ4(μ))|C,|dJ1(μ,ρ4(μ))dμ|C.|J_{1}(\mu,\rho_{4}(\mu))|\leq C,\hskip 20.00003pt\Big{|}\frac{\mathrm{d}J_{1}(\mu,\rho_{4}(\mu))}{\mathrm{d}\mu}\Big{|}\leq C. (4.4)

4.2 Estimates for p1p_{1}

Set the perturbation

S(θ)=cos(nθ),S(\theta)=\cos(n\theta),

as the set {cos(nθ)}n=1\{\cos(n\theta)\}_{n=1}^{\infty} is clearly a basis for the Banach space X1l+αX^{l+\alpha}_{1} defined in ((3.70)). Since the solution to ((3.51))((3.59)) (L1,H1,F1,p1)(L_{1},H_{1},F_{1},p_{1}) is unique, we know if we can find a solution (L1,H1,F1,p1)(L_{1},H_{1},F_{1},p_{1}) of the form

L1=L1ncos(nθ),\displaystyle L_{1}=L_{1}^{n}\cos(n\theta),\hskip 20.00003pt H1=H1ncos(nθ),\displaystyle H_{1}=H_{1}^{n}\cos(n\theta), (4.5)
F1=F1ncos(nθ),\displaystyle F_{1}=F_{1}^{n}\cos(n\theta),\hskip 20.00003pt p1=p1ncos(nθ),\displaystyle p_{1}=p_{1}^{n}\cos(n\theta), (4.6)

then it is the unique solution of ((3.51))((3.59)). Substituting ((4.5)) and ((4.6)) into ((3.51))((3.59)), we need to find (L1n,H1n,F1n,p1n)(L_{1}^{n},H_{1}^{n},F_{1}^{n},p_{1}^{n}) satisfying

2L1nr21rL1nr+n2r2L1n=f5(L1n,H1n,F1n)\displaystyle-\frac{\partial^{2}L_{1}^{n}}{\partial r^{2}}-\frac{1}{r}\frac{\partial L_{1}^{n}}{\partial r}+\frac{n^{2}}{r^{2}}L_{1}^{n}=f_{5}(L_{1}^{n},H_{1}^{n},F_{1}^{n})\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.7)
2H1nr21rH1nr+n2r2H1n=f6(L1n,H1n,F1n)\displaystyle-\frac{\partial^{2}H_{1}^{n}}{\partial r^{2}}-\frac{1}{r}\frac{\partial H_{1}^{n}}{\partial r}+\frac{n^{2}}{r^{2}}H_{1}^{n}=f_{6}(L_{1}^{n},H_{1}^{n},F_{1}^{n})\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.8)
D2F1nr2DrF1nr+Dn2r2F1nF1nrpr=f7(L1n,H1n,F1n)+Frp1nr\displaystyle-D\frac{\partial^{2}F_{1}^{n}}{\partial r^{2}}-\frac{D}{r}\frac{\partial F_{1}^{n}}{\partial r}+\frac{Dn^{2}}{r^{2}}F_{1}^{n}-\frac{\partial F_{1}^{n}}{\partial r}\frac{\partial p_{*}}{\partial r}=f_{7}(L_{1}^{n},H_{1}^{n},F_{1}^{n})+\frac{\partial F_{*}}{\partial r}\frac{\partial p_{1}^{n}}{\partial r}\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.9)
2p1nr21rp1nr+n2r2p1n=f8(L1n,H1n,F1n)\displaystyle-\frac{\partial^{2}p_{1}^{n}}{\partial r^{2}}-\frac{1}{r}\frac{\partial p_{1}^{n}}{\partial r}+\frac{n^{2}}{r^{2}}p_{1}^{n}=f_{8}(L_{1}^{n},H_{1}^{n},F_{1}^{n})\hskip 20.00003pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.10)
L1nr=H1nr=F1nr=p1nr=0\displaystyle\frac{\partial L_{1}^{n}}{\partial r}=\frac{\partial H_{1}^{n}}{\partial r}=\frac{\partial F_{1}^{n}}{\partial r}=\frac{\partial p_{1}^{n}}{\partial r}=0 r=1,\displaystyle r=1, (4.11)
L1nr+β1L1n=(2Lr2β1Lr)|r=1ε\displaystyle-\frac{\partial L_{1}^{n}}{\partial r}+\beta_{1}L_{1}^{n}=\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon} r=1ε,\displaystyle r=1-\varepsilon, (4.12)
H1nr+β1H1n=(2Hr2β1Hr)|r=1ε\displaystyle-\frac{\partial H_{1}^{n}}{\partial r}+\beta_{1}H_{1}^{n}=\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon} r=1ε,\displaystyle r=1-\varepsilon, (4.13)
F1nr+β2F1n=(2Fr2β2Fr)|r=1ε\displaystyle-\frac{\partial F_{1}^{n}}{\partial r}+\beta_{2}F_{1}^{n}=\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon} r=1ε,\displaystyle r=1-\varepsilon, (4.14)
p1n=1n2(1ε)2\displaystyle p_{1}^{n}=\frac{1-n^{2}}{(1-\varepsilon)^{2}} r=1ε,\displaystyle r=1-\varepsilon, (4.15)

where by ((3.51))((3.54)), f5f_{5}, f6f_{6}, f7f_{7}, and f8f_{8} can all be bounded by linear functions of |L1n||L_{1}^{n}|, |H1n||H_{1}^{n}|, and |F1n||F_{1}^{n}|. In particular, f8f_{8} is expressed as

f8=1M0[λ(M0F)L1nγ+HλLF1nγ+Hλ(M0F)LH1n(γ+H)2+(ρ3ρ4)F1n],f_{8}=\frac{1}{M_{0}}\Big{[}\lambda\frac{(M_{0}-F_{*})L^{n}_{1}}{\gamma+H_{*}}-\lambda\frac{L_{*}F^{n}_{1}}{\gamma+H_{*}}-\lambda\frac{(M_{0}-F_{*})L_{*}H^{n}_{1}}{(\gamma+H_{*})^{2}}+(\rho_{3}-\rho_{4})F_{1}^{n}\Big{]}, (4.16)

which will be used later.

Denote the operator L2r2+1rr+n2r2\mathscript L\triangleq\frac{\partial^{2}}{\partial r^{2}}+\frac{1}{r}\frac{\partial}{\partial r}+\frac{n^{2}}{r^{2}}. For this special operator, one can easily verify the following lemmas.

Lemma 4.13.

The general solution of (η\eta is a constant)

L[ψ]ψ′′1rψ+n2r2ψ=η+f(r),1ε<r<1,\displaystyle{\mathscript L}[\psi]\triangleq-\psi^{\prime\prime}-\frac{1}{r}\psi^{\prime}+\frac{n^{2}}{r^{2}}\psi=\eta+f(r),\hskip 20.00003pt1-\varepsilon<r<1, (4.17)
ψ(1)=0\displaystyle\psi^{\prime}(1)=0 (4.18)

is given by

ψψ1={Arn+Brn+K[f](r),where B=A+1nK[f](1)n0,A+K[f](r)n=0,\displaystyle\psi-\psi_{1}=\left\{\begin{array}[]{ll}\displaystyle Ar^{n}+Br^{-n}+K[f](r),\quad\text{where }B=A+\frac{1}{n}K[f]^{\prime}(1)&n\neq 0,\\ A+K[f](r)&n=0,\end{array}\right. (4.21)

where

ψ1(1)=0,\displaystyle\psi_{1}^{\prime}(1)=0,\hskip 20.00003pt ψ1={ηn24(r22nrn)n0,2,η(1r24+12logr)n=0,η(r28r24logr)n=2,\displaystyle\psi_{1}=\left\{\begin{array}[]{ll}\displaystyle\frac{\eta}{n^{2}-4}\Big{(}r^{2}-\frac{2}{n}r^{n}\Big{)}&n\neq 0,2,\\ \displaystyle\eta\Big{(}\frac{1-r^{2}}{4}+\frac{1}{2}\log r\Big{)}&n=0,\\ \displaystyle\eta\Big{(}\frac{r^{2}}{8}-\frac{r^{2}}{4}\log r\Big{)}&n=2,\end{array}\right. (4.25)

and

K[f](r)={rn2nr1sn+1f(s)ds+rn2n1εrsn+1f(s)dsn0,r1(logsr)sf(s)dsn=0.\displaystyle K[f](r)=\left\{\begin{array}[]{ll}\displaystyle\frac{r^{n}}{2n}\int_{r}^{1}s^{-n+1}f(s)\,\mathrm{d}s+\frac{r^{-n}}{2n}\int_{1-\varepsilon}^{r}s^{n+1}f(s)\,\mathrm{d}s&n\neq 0,\\ \displaystyle\rule{0.0pt}{18.0pt}-\int_{r}^{1}\Big{(}\log\frac{s}{r}\Big{)}\;sf(s)\,\mathrm{d}s&n=0.\end{array}\right. (4.28)

The special solution K[f]K[f] satisfies

|K[f](r)|min(ε2n,1n2)fL,|K[f](r)|min(ε2,1n)fL,n1,|K[f](r)|\leq\min\Big{(}\frac{\varepsilon}{2n},\frac{1}{n^{2}}\Big{)}\|f\|_{L^{\infty}},\hskip 20.00003pt|K[f]^{\prime}(r)|\leq\min\Big{(}\frac{\varepsilon}{2},\frac{1}{n}\Big{)}\|f\|_{L^{\infty}},\hskip 20.00003ptn\geq 1, (4.29)

and

|K[f](r)|εfL,|K[f](r)|εfL,n=0.|K[f](r)|\leq\varepsilon\|f\|_{L^{\infty}},\hskip 20.00003pt|K[f]^{\prime}(r)|\leq\varepsilon\|f\|_{L^{\infty}},\hskip 20.00003ptn=0. (4.30)
Proof.

Using the expression in ((4.28)), we clearly have, for 1εr11-\varepsilon\leq r\leq 1 and n1n\geq 1,

|K[f](r)|fL[12nr1(rs)nsds+12n1εr(sr)nsds]ε2nfL,\displaystyle|K[f](r)|\;\leq\;\|f\|_{L^{\infty}}\Big{[}\frac{1}{2n}\int_{r}^{1}\Big{(}\frac{r}{s}\Big{)}^{n}s\,\mathrm{d}s+\frac{1}{2n}\int_{1-\varepsilon}^{r}\Big{(}\frac{s}{r}\Big{)}^{n}s\,\mathrm{d}s\Big{]}\;\leq\;\frac{\varepsilon}{2n}\|f\|_{L^{\infty}}, (4.31)

We can also integrate the expression to obtain

r1(rs)nsds+1εr(sr)nsdsr1rnsn1ds+1εrrnsn1dsrnrnn+rnrnn=2n;\int_{r}^{1}\Big{(}\frac{r}{s}\Big{)}^{n}s\,\mathrm{d}s+\int_{1-\varepsilon}^{r}\Big{(}\frac{s}{r}\Big{)}^{n}s\,\mathrm{d}s\leq\int_{r}^{1}r^{n}s^{-n-1}\,\mathrm{d}s+\int_{1-\varepsilon}^{r}r^{-n}s^{n-1}\,\mathrm{d}s\leq r^{n}\frac{r^{-n}}{n}+r^{-n}\frac{r^{n}}{n}=\frac{2}{n};

combining it with ((4.31)), we deduce

|K[f](r)|min(ε2n,1n2)fL.|K[f](r)|\leq\min\Big{(}\frac{\varepsilon}{2n},\frac{1}{n^{2}}\Big{)}\|f\|_{L^{\infty}}.

Furthermore, it follows from ((4.28)) that

K[f](r)=rn12r1sn+1f(s)dsrn121εrsn+1f(s)ds;K[f]^{\prime}(r)=\frac{r^{n-1}}{2}\int_{r}^{1}s^{-n+1}f(s)\,\mathrm{d}s-\frac{r^{-n-1}}{2}\int_{1-\varepsilon}^{r}s^{n+1}f(s)\,\mathrm{d}s;

similarly, we shall obtain

|K[f](r)|\displaystyle|K[f]^{\prime}(r)| \displaystyle\leq fL[12r1(rs)n1ds+121εr(sr)n+1ds]min(ε2,1n)fL.\displaystyle\|f\|_{L^{\infty}}\Big{[}\frac{1}{2}\int_{r}^{1}\Big{(}\frac{r}{s}\Big{)}^{n-1}\,\mathrm{d}s+\frac{1}{2}\int_{1-\varepsilon}^{r}\Big{(}\frac{s}{r}\Big{)}^{n+1}\,\mathrm{d}s\Big{]}\;\leq\;\min\Big{(}\frac{\varepsilon}{2},\frac{1}{n}\Big{)}\|f\|_{L^{\infty}}.

The case n=0n=0 is similar. ∎

Lemma 4.14.

If in addition to ((4.17)) and ((4.18)) we further assume ψ(1ε)=G\psi(1-\varepsilon)=G, then, for n1n\geq 1,

A=11+(1ε)2n((1ε)n[Gψ1(1ε)](1ε)nK[f](1ε)1nK[f](1)),\displaystyle A=\frac{1}{1+(1-\varepsilon)^{2n}}\Big{(}(1-\varepsilon)^{n}[G-\psi_{1}(1-\varepsilon)]-(1-\varepsilon)^{n}K[f](1-\varepsilon)-\frac{1}{n}K[f]^{\prime}(1)\Big{)}, (4.32)
B=11+(1ε)2n((1ε)n[Gψ1(1ε)](1ε)nK[f](1ε)+(1ε)2nnK[f](1)),\displaystyle B=\frac{1}{1+(1-\varepsilon)^{2n}}\Big{(}(1-\varepsilon)^{n}[G-\psi_{1}(1-\varepsilon)]-(1-\varepsilon)^{n}K[f](1-\varepsilon)+\frac{(1-\varepsilon)^{2n}}{n}K[f]^{\prime}(1)\Big{)}, (4.33)

and for n=0n=0,

A=Gψ1(1ε)K[f](1ε).A=G-\psi_{1}(1-\varepsilon)-K[f](1-\varepsilon). (4.34)
Lemma 4.15.

For n0n\geq 0 and 0<ε<10<\varepsilon<1,

1nε(1ε)n1nε+12n2ε2.\displaystyle 1-n\varepsilon\leq(1-\varepsilon)^{n}\leq 1-n\varepsilon+\frac{1}{2}n^{2}\varepsilon^{2}. (4.35)
Proof.

The function f(ε)(1ε)n1+nεf(\varepsilon)\triangleq(1-\varepsilon)^{n}-1+n\varepsilon satisfies f(0)=0f(0)=0 and f(ε)=n(1ε)n1+n0f^{\prime}(\varepsilon)=-n(1-\varepsilon)^{n-1}+n\geq 0 for 0<ε<10<\varepsilon<1, so that f(ε)0f(\varepsilon)\geq 0 for 0<ε<10<\varepsilon<1.

Similarly, the function f(ε)(1ε)n1+nε12n2ε2f(\varepsilon)\triangleq(1-\varepsilon)^{n}-1+n\varepsilon-\frac{1}{2}n^{2}\varepsilon^{2} satisfies f(0)=f(0)=0f(0)=f^{\prime}(0)=0 and f′′(ε)=n(n1)(1ε)n2n20f^{\prime\prime}(\varepsilon)=n(n-1)(1-\varepsilon)^{n-2}-n^{2}\leq 0 for 0<ε<10<\varepsilon<1, so that f(ε)0f(\varepsilon)\leq 0 for 0<ε<10<\varepsilon<1. ∎

In order to make the boundary conditions ((4.45))((4.47)) homogeneous, let’s instead work with

L1n~(r)=L1n(r)1β1(2Lr2β1Lr)|r=1ε,\displaystyle\widetilde{L_{1}^{n}}(r)=L_{1}^{n}(r)-\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}, (4.36)
H1n~(r)=H1n(r)1β1(2Hr2β1Hr)|r=1ε,\displaystyle\widetilde{H_{1}^{n}}(r)=H_{1}^{n}(r)-\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}, (4.37)
F1n~(r)=F1n(r)1β2(2Fr2β2Fr)|r=1ε.\displaystyle\widetilde{F_{1}^{n}}(r)=F_{1}^{n}(r)-\frac{1}{\beta_{2}}\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}. (4.38)

Accordingly, L1n~(r)\widetilde{L_{1}^{n}}(r), H1n~(r)\widetilde{H_{1}^{n}}(r), F1n~(r)\widetilde{F_{1}^{n}}(r) satisfy the following equations:

2L1n~r21rL1n~r+n2r2L1n~=f~5f5n2β1r2(2Lr2β1Lr)|r=1ε\displaystyle-\frac{\partial^{2}\widetilde{L_{1}^{n}}}{\partial r^{2}}-\frac{1}{r}\frac{\partial\widetilde{L_{1}^{n}}}{\partial r}+\frac{n^{2}}{r^{2}}\widetilde{L_{1}^{n}}=\widetilde{f}_{5}\triangleq f_{5}-\frac{n^{2}}{\beta_{1}r^{2}}\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.39)
2H1n~r21rH1n~r+n2r2H1n~=f~6f6n2β1r2(2Hr2β1Hr)|r=1ε\displaystyle-\frac{\partial^{2}\widetilde{H_{1}^{n}}}{\partial r^{2}}-\frac{1}{r}\frac{\partial\widetilde{H_{1}^{n}}}{\partial r}+\frac{n^{2}}{r^{2}}\widetilde{H_{1}^{n}}=\widetilde{f}_{6}\triangleq f_{6}-\frac{n^{2}}{\beta_{1}r^{2}}\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.40)
D2F1n~r2DrF1n~r+Dn2r2F1n~F1n~rpr=f~7f7+Frp1nrDn2β2r2(2Fr2β2Fr)|r=1ε\displaystyle\begin{array}[]{ll}-D\frac{\partial^{2}\widetilde{F_{1}^{n}}}{\partial r^{2}}-\frac{D}{r}\frac{\partial\widetilde{F_{1}^{n}}}{\partial r}+\frac{Dn^{2}}{r^{2}}\widetilde{F_{1}^{n}}-\frac{\partial\widetilde{F_{1}^{n}}}{\partial r}\frac{\partial p_{*}}{\partial r}=\widetilde{f}_{7}\triangleq f_{7}+\frac{\partial F_{*}}{\partial r}\frac{\partial p_{1}^{n}}{\partial r}\\ \hskip 205.0003pt-\frac{Dn^{2}}{\beta_{2}r^{2}}\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}\end{array}\hskip 10.00002pt in Ω,\displaystyle\text{in }\Omega_{*}, (4.43)
L1n~r=H1n~r=F1n~r=0\displaystyle\frac{\partial\widetilde{L_{1}^{n}}}{\partial r}=\frac{\partial\widetilde{H_{1}^{n}}}{\partial r}=\frac{\partial\widetilde{F_{1}^{n}}}{\partial r}=0 r=1,\displaystyle r=1, (4.44)
L1n~r+β1L1n~=0\displaystyle-\frac{\partial\widetilde{L_{1}^{n}}}{\partial r}+\beta_{1}\widetilde{L_{1}^{n}}=0 r=1ε,\displaystyle r=1-\varepsilon, (4.45)
H1n~r+β1H1n~=0\displaystyle-\frac{\partial\widetilde{H_{1}^{n}}}{\partial r}+\beta_{1}\widetilde{H_{1}^{n}}=0 r=1ε,\displaystyle r=1-\varepsilon, (4.46)
F1n~r+β2F1n~=0\displaystyle-\frac{\partial\widetilde{F_{1}^{n}}}{\partial r}+\beta_{2}\widetilde{F_{1}^{n}}=0 r=1ε,\displaystyle r=1-\varepsilon, (4.47)

where p1np_{1}^{n} is defined by ((4.10)) and ((4.15)).

Lemma 4.16.

For sufficiently small ε\varepsilon, there exist a constant CC which does not depend on ε\varepsilon and nn such that the following inequalities are valid for the above system,

L1n~L(1ε,1)+H1n~L(1ε,1)+F1n~L(1ε,1)C(n2+1)ε,\displaystyle\|\widetilde{L_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}+\|\widetilde{H_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}+\|\widetilde{F_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}\leq C(n^{2}+1)\varepsilon, (4.48)
(p1n)L(1ε,1)2(n3+1).\displaystyle\|({p_{1}^{n}})^{\prime}\|_{L^{\infty}(1-\varepsilon,1)}\leq 2(n^{3}+1). (4.49)
Proof.

To prove ((4.48)) and ((4.49)), we again use the idea of continuation (Appendix 5.3), and multiply the right-hand sides of ((4.39))((4.43)) as well as ((4.10)) by δ\delta. When δ=0\delta=0, it follows from the maximum principle that L1n~=H1n~=F1n~=0\widetilde{L_{1}^{n}}=\widetilde{H_{1}^{n}}=\widetilde{F_{1}^{n}}=0, hence ((4.48)) clearly holds in this case. Furthermore, it can be solved from

2p1nr21rp1nr+n2r2p1n=01ε<r<1,\displaystyle-\frac{\partial^{2}p_{1}^{n}}{\partial r^{2}}-\frac{1}{r}\frac{\partial p_{1}^{n}}{\partial r}+\frac{n^{2}}{r^{2}}p_{1}^{n}=0\hskip 20.00003pt1-\varepsilon<r<1, (4.50)
p1n(1)r=0,p1n(1ε)=1n2(1ε)2,\displaystyle\frac{\partial p_{1}^{n}(1)}{\partial r}=0,\hskip 20.00003ptp_{1}^{n}(1-\varepsilon)=\frac{1-n^{2}}{(1-\varepsilon)^{2}}, (4.51)

that

p1n(r)=1n2(1ε)2[(1ε)n+(1ε)n](rn+rn),p_{1}^{n}(r)=\frac{1-n^{2}}{(1-\varepsilon)^{2}[(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}]}\Big{(}r^{n}+r^{-n}\Big{)},

and hence for 0<ε10<\varepsilon\ll 1,

(p1n)L(1ε,1)\displaystyle\|(p_{1}^{n})^{\prime}\|_{L^{\infty}(1-\varepsilon,1)} =max1εr1|1n2(1ε)2[(1ε)n+(1ε)n]n(rn1rn1)|\displaystyle=\max\limits_{1-\varepsilon\leq r\leq 1}\bigg{|}\frac{1-n^{2}}{(1-\varepsilon)^{2}[(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}]}n\Big{(}r^{n-1}-r^{-n-1}\Big{)}\bigg{|}
n(n21)|1(1ε)3(1ε)n(1ε)n(1ε)n+(1ε)n|2(n3+1).\displaystyle\leq n(n^{2}-1)\bigg{|}\frac{1}{(1-\varepsilon)^{3}}\frac{(1-\varepsilon)^{n}-(1-\varepsilon)^{-n}}{(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}}\bigg{|}\leq 2(n^{3}+1).

Next we consider the case when 0<δ10<\delta\leq 1. We first assume that

L1n~L(1ε,1)+H1n~L(1ε,1)+F1n~L(1ε,1)n2+1,\displaystyle\|\widetilde{L_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}+\|\widetilde{H_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}+\|\widetilde{F_{1}^{n}}\|_{L^{\infty}(1-\varepsilon,1)}\leq n^{2}+1, (4.52)
(p1n)L(1ε,1)3(n3+1).\displaystyle\|({p_{1}^{n}})^{\prime}\|_{L^{\infty}(1-\varepsilon,1)}\leq 3(n^{3}+1). (4.53)

Then clearly |f~5|C(n2+1)|\widetilde{f}_{5}|\leq C(n^{2}+1), and (K[f~5](r)+|K[f~5](1)|(r+1β1)+1β1|K[f~5](1ε)β1K[f~5](1ε)|)\Big{(}K[\widetilde{f}_{5}](r)+\Big{|}K[\widetilde{f}_{5}]^{\prime}(1)\Big{|}\Big{(}r+\frac{1}{\beta_{1}}\Big{)}+\frac{1}{\beta_{1}}\Big{|}K[\widetilde{f}_{5}]^{\prime}(1-\varepsilon)-{\beta_{1}}K[\widetilde{f}_{5}](1-\varepsilon)\Big{|}\Big{)} is a supersolution for L1n~(r)\widetilde{L_{1}^{n}}(r) when n1n\geq 1. It follows that, by Lemma 4.13,

|L1n~(r)|K[f~5](r)+|K[f~5](1)|(r+1β1)+1β1|K[f~5](1ε)β1K[f~5](1ε)|C(n2+1)ε.\Big{|}\widetilde{L_{1}^{n}}(r)\Big{|}\leq K[\widetilde{f}_{5}](r)+\Big{|}K[\widetilde{f}_{5}]^{\prime}(1)\Big{|}\Big{(}r+\frac{1}{\beta_{1}}\Big{)}+\frac{1}{\beta_{1}}\Big{|}K[\widetilde{f}_{5}]^{\prime}(1-\varepsilon)-{\beta_{1}}K[\widetilde{f}_{5}](1-\varepsilon)\Big{|}\leq C(n^{2}+1)\varepsilon.

The case when n=0n=0 can be easily proved. Similarly, we have |H1n~(r)|C(n2+1)ε|\widetilde{H_{1}^{n}}(r)|\leq C(n^{2}+1)\varepsilon. Next let’s prove the estimate for F1n~\widetilde{F_{1}^{n}}. Under our assumptions, by ((2.18)), ((4.52)), and ((4.53)),

f~7LC(n2+1)+Cε(n3+1),\|\widetilde{f}_{7}\|_{L^{\infty}}\leq C(n^{2}+1)+C\varepsilon(n^{3}+1),

so that, we can use ((4.29)) to derive

|K[f~7](r)|C(n+1)ε,|K[f~7](r)|C(n2+1)ε.|K[\widetilde{f}_{7}](r)|\leq C(n+1)\varepsilon,\hskip 20.00003pt|K[\widetilde{f}_{7}]^{\prime}(r)|\leq C(n^{2}+1)\varepsilon.

The function ϕ=1D{K[f~7](r)+|K[f~7](1)|(r+1β2)+1β2|K[f~7](1ε)β2K[f~7](1ε)|+ε}\phi=\frac{1}{D}\Big{\{}K[\widetilde{f}_{7}](r)+\Big{|}K[\widetilde{f}_{7}]^{\prime}(1)\Big{|}\Big{(}r+\frac{1}{\beta}_{2}\Big{)}+\frac{1}{\beta}_{2}\Big{|}K[\widetilde{f}_{7}]^{\prime}(1-\varepsilon)-\beta_{2}K[\widetilde{f}_{7}](1-\varepsilon)\Big{|}+\varepsilon\Big{\}} satisfies,

DL[ϕ]+ϕrpr\displaystyle D\mathscript{L}[\phi]+\frac{\partial\phi}{\partial r}\frac{\partial p_{*}}{\partial r}
=\displaystyle= f~7+1D(K[f~7](r)+|K[f~7](1)|)pr1Dr|K[f~7](1)|+n2Dr2|K[f~7](1)|(r+1β2)\displaystyle\;\;\widetilde{f}_{7}+\frac{1}{D}\Big{(}K[\widetilde{f}_{7}]^{\prime}(r)+\Big{|}K[\widetilde{f}_{7}]^{\prime}(1)\Big{|}\Big{)}\frac{\partial p_{*}}{\partial r}-\frac{1}{Dr}\Big{|}K[\widetilde{f}_{7}]^{\prime}(1)\Big{|}+\frac{n^{2}}{Dr^{2}}\Big{|}K[\widetilde{f}_{7}]^{\prime}(1)\Big{|}\Big{(}r+\frac{1}{\beta}_{2}\Big{)}
+n2Dr2(1β2|K[f~7](1ε)β2K[f~7](1ε)|+ε)\displaystyle\;\;+\frac{n^{2}}{Dr^{2}}\Big{(}\frac{1}{\beta}_{2}\Big{|}K[\widetilde{f}_{7}]^{\prime}(1-\varepsilon)-\beta_{2}K[\widetilde{f}_{7}](1-\varepsilon)\Big{|}+\varepsilon\Big{)}
\displaystyle\geq f~7CεK[f~7]L+n2εf~7C(n2+1)ε2+n2εf~7,\displaystyle\;\;\widetilde{f}_{7}-C\varepsilon\Big{\|}K[\widetilde{f}_{7}]^{\prime}\Big{\|}_{L^{\infty}}+n^{2}\varepsilon\;\geq\;\widetilde{f}_{7}-C(n^{2}+1)\varepsilon^{2}+n^{2}\varepsilon\;\geq\;\widetilde{f}_{7},

for n1n\geq 1, where we also make use of ((2.18)) in deriving the above estimate. Therefore, it follows from the maximum principle that

|F1n~(r)|1D{K[f~7](r)+|K[f~7](1)|(r+1β2)+1β2|K[f~7](1ε)β2K[f~7](1ε)|+ε}C(n2+1)ε.|\widetilde{F_{1}^{n}}(r)|\leq\frac{1}{D}\Big{\{}K[\widetilde{f}_{7}](r)+\Big{|}K[\widetilde{f}_{7}]^{\prime}(1)\Big{|}\Big{(}r+\frac{1}{\beta}_{2}\Big{)}+\frac{1}{\beta}_{2}\Big{|}K[\widetilde{f}_{7}]^{\prime}(1-\varepsilon)-\beta_{2}K[\widetilde{f}_{7}](1-\varepsilon)\Big{|}+\varepsilon\Big{\}}\leq C(n^{2}+1)\varepsilon.

Finally, in order to estimate (p1n)(p^{n}_{1})^{\prime}, we use the explicit formula from Lemma 4.13. Taking η=0\eta=0 and G=(1n2)/(1ε)2G=(1-n^{2})/(1-\varepsilon)^{2}, we obtain from Lemma 4.13 that

(p1n)\displaystyle(p_{1}^{n})^{\prime} =Anrn1Bnrn1+K[f8](r),\displaystyle=Anr^{n-1}-Bnr^{-n-1}+K[f_{8}]^{\prime}(r), (4.54)

where AA and BB are defined in Lemma 4.14. By ((4.52)),

f8LC(n2+1),\|f_{8}\|_{L^{\infty}}\leq C(n^{2}+1),

and together with ((4.29)) in Lemma 4.13, we have

|K[f8](r)|Cn2+1nε,|K[f8](r)|C(n2+1)ε.|K[f_{8}](r)|\leq C\frac{n^{2}+1}{n}\varepsilon,\hskip 20.00003pt|K[f_{8}]^{\prime}(r)|\leq C(n^{2}+1)\varepsilon. (4.55)

Combining ((4.54)) with ((4.32)) ((4.33)) and ((4.55)), we then obtain

|(p1n)|\displaystyle\Big{|}(p_{1}^{n})^{\prime}\Big{|}\;\leq max1εr1|n(rn1rn1)(1ε)n+(1ε)n[Gk[f8](1ε)]|\displaystyle\;\;\max\limits_{1-\varepsilon\leq r\leq 1}\Big{|}\frac{n(r^{n-1}-r^{-n-1})}{(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}}\Big{[}G-k[f_{8}](1-\varepsilon)\Big{]}\Big{|}
+max1εr1|rn1+(1ε)2nrn11+(1ε)2nK[f8](1)|+max1εr1|K[f8](r)|\displaystyle\quad\;+\max\limits_{1-\varepsilon\leq r\leq 1}\Big{|}\frac{r^{n-1}+(1-\varepsilon)^{2n}r^{-n-1}}{1+(1-\varepsilon)^{2n}}K[f_{8}]^{\prime}(1)\Big{|}+\max\limits_{1-\varepsilon\leq r\leq 1}\Big{|}K[f_{8}]^{\prime}(r)\Big{|}
\displaystyle\leq |nG1ε(1ε)n(1ε)n(1ε)n+(1ε)n|+CnK[f8]L+CK[f8]L\displaystyle\;\;\Big{|}\frac{n\,G}{1-\varepsilon}\frac{(1-\varepsilon)^{n}-(1-\varepsilon)^{-n}}{(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}}\Big{|}+Cn\|K[f_{8}]\|_{L^{\infty}}+C\|K[f_{8}]^{\prime}\|_{L^{\infty}}
\displaystyle\leq |n(n21)(1ε)3(1ε)n(1ε)n(1ε)n+(1ε)n|+C(n2+1)ε\displaystyle\;\;\bigg{|}\frac{n(n^{2}-1)}{(1-\varepsilon)^{3}}\frac{(1-\varepsilon)^{n}-(1-\varepsilon)^{-n}}{(1-\varepsilon)^{n}+(1-\varepsilon)^{-n}}\bigg{|}+C(n^{2}+1)\varepsilon
\displaystyle\leq   2(n3+1),\displaystyle\;\;2(n^{3}+1),

hence (p1n)L(1ε,1)2(n3+1)\|({p_{1}^{n}})^{\prime}\|_{L^{\infty}(1-\varepsilon,1)}\leq 2(n^{3}+1) is valid for sufficiently small ε\varepsilon. ∎

Based on ((4.48)) and ((4.49)), the existence and uniqueness of such a solution (L1n,H1n,F1n,p1n)(L_{1}^{n},H_{1}^{n},F_{1}^{n},p_{1}^{n}) to the system ((4.7))((4.15)) can be justified through the contraction mapping principle, hence we have the following lemma.

Lemma 4.17.

For each nonnegative nn and sufficiently small ε\varepsilon, the system ((4.7))((4.15)) admits a unique solution (L1n,H1n,F1n,p1n)(L_{1}^{n},H_{1}^{n},F_{1}^{n},p_{1}^{n}).

By ((4.49)) we already derived the estimate

|p1n(r)r|2(n3+1),1εr1.\Big{|}\frac{\partial p_{1}^{n}(r)}{\partial r}\Big{|}\leq 2(n^{3}+1),\hskip 20.00003pt1-\varepsilon\leq r\leq 1.

This estimate, however, is not enough; we need a sharper bound for p1n(1ε)r\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}. To do that, we start with rewriting ((4.36))((4.38)) in the same way as in ((2.1))((2.1)).

Evaluating ((2.1)) at r=1εr=1-\varepsilon, and using ((2.1))((2.1)), we obtain

2L(1ε)r2\displaystyle\frac{\partial^{2}L_{*}(1-\varepsilon)}{\partial r^{2}} =(k1(M0F)LK1+L+ρ1L)|r=1ε11εL(1ε)r\displaystyle=\Big{(}k_{1}\frac{(M_{0}-F_{*})L_{*}}{K_{1}+L_{*}}+\rho_{1}L_{*}\Big{)}\Big{|}_{r=1-\varepsilon}-\frac{1}{1-\varepsilon}\frac{\partial L_{*}(1-\varepsilon)}{\partial r}
=ρ3(γ+H0)(k1M0λK1+ρ3(γ+H0)+ρ1λ)11εL(1ε)r+O(ε).\displaystyle=\rho_{3}(\gamma+H_{0})\Big{(}\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+\frac{\rho_{1}}{\lambda}\Big{)}-\frac{1}{1-\varepsilon}\frac{\partial L_{*}(1-\varepsilon)}{\partial r}+O(\varepsilon).

Recall that the boundary condition for LL_{*} is

L(1ε)r=β1(L(1ε)L0)=β1(ρ3(γ+H0)λ+O(ε)ρ3(γ+H0)λεμλ)=O(ε).\frac{\partial L_{*}(1-\varepsilon)}{\partial r}=\beta_{1}(L_{*}(1-\varepsilon)-L_{0})=\beta_{1}\Big{(}\frac{\rho_{3}(\gamma+H_{0})}{\lambda}+O(\varepsilon)-\frac{\rho_{3}(\gamma+H_{0})}{\lambda}-\frac{\varepsilon\mu}{\lambda}\Big{)}=O(\varepsilon).

We combine the above two equations to derive

1β1(2Lr2β1Lr)|r=1ε=ρ3(γ+H0)β1(k1M0λK1+ρ3(γ+H0)+ρ1λ)+O(ε).\displaystyle\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=\frac{\rho_{3}(\gamma+H_{0})}{\beta_{1}}\Big{(}\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+\frac{\rho_{1}}{\lambda}\Big{)}+O(\varepsilon).

Similarly, we can also get

1β1(2Hr2β1Hr)|r=1ε=ρ2H0β1+O(ε),\displaystyle\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=\frac{\rho_{2}H_{0}}{\beta_{1}}+O(\varepsilon),
1β2(2Fr2β2Fr)|r=1ε=ρ3(γ+H0)β2Dk1M0λK1+ρ3(γ+H0)+O(ε).\displaystyle\frac{1}{\beta_{2}}\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=-\frac{\rho_{3}(\gamma+H_{0})}{\beta_{2}D}\;\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+O(\varepsilon).

Comparing with the definitions of L1L_{*}^{1}, H1H_{*}^{1} and F1F_{*}^{1} in ((2.1))((2.1)), we find that

ρ3(γ+H0)β1(k1M0λK1+ρ3(γ+H0)+ρ1λ)=μλL1,\displaystyle\frac{\rho_{3}(\gamma+H_{0})}{\beta_{1}}\Big{(}\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}+\frac{\rho_{1}}{\lambda}\Big{)}=\frac{\mu}{\lambda}-L_{*}^{1},
ρ2H0β1=H1,ρ3(γ+H0)β2Dk1M0λK1+ρ3(γ+H0)=F1.\displaystyle\frac{\rho_{2}H_{0}}{\beta_{1}}=-H_{*}^{1},\hskip 20.00003pt-\frac{\rho_{3}(\gamma+H_{0})}{\beta_{2}D}\;\frac{k_{1}M_{0}}{\lambda K_{1}+\rho_{3}(\gamma+H_{0})}=-F_{*}^{1}.

Therefore, the above equations indicate

1β1(2Lr2β1Lr)|r=1ε=μλL1+O(ε),\displaystyle\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}L_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial L_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=\frac{\mu}{\lambda}-L_{*}^{1}+O(\varepsilon), (4.56)
1β1(2Hr2β1Hr)|r=1ε=H1+O(ε),\displaystyle\frac{1}{\beta_{1}}\Big{(}\frac{\partial^{2}H_{*}}{\partial r^{2}}-\beta_{1}\frac{\partial H_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=-H_{*}^{1}+O(\varepsilon), (4.57)
1β2(2Fr2β2Fr)|r=1ε=F1+O(ε).\displaystyle\frac{1}{\beta_{2}}\Big{(}\frac{\partial^{2}F_{*}}{\partial r^{2}}-\beta_{2}\frac{\partial F_{*}}{\partial r}\Big{)}\Big{|}_{r=1-\varepsilon}=-F_{*}^{1}+O(\varepsilon). (4.58)

After we show ((4.56))((4.58)), we can combine them with ((4.36))((4.38)) as well as ((4.48)) to claim that

L1n\displaystyle L_{1}^{n} =\displaystyle= μ/λL1+O((n2+1)ε),\displaystyle\mu/\lambda-L_{*}^{1}+O((n^{2}+1)\varepsilon), (4.59)
H1n\displaystyle H_{1}^{n} =\displaystyle= H1+O((n2+1)ε),\displaystyle-H_{*}^{1}+O((n^{2}+1)\varepsilon), (4.60)
F1n\displaystyle F_{1}^{n} =\displaystyle= F1+O((n2+1)ε).\displaystyle-F_{*}^{1}+O((n^{2}+1)\varepsilon). (4.61)

With ((4.59))((4.61)), we are able to derive a more delicate estimate for p1n(1ε)r\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}. Substituting ((2.1))((2.1)) and ((4.59))((4.61)) all into ((4.16)), recalling also ((2.15)) and ((2.16)), we obtain

f8=μγ+H01M0(M0(λL1ρ3H1)γ+H0ρ4F1)+O((n2+1)ε)=μγ+H0+O((n2+1)ε),f_{8}=\frac{\mu}{\gamma+H_{0}}-\frac{1}{M_{0}}\Big{(}\frac{M_{0}(\lambda L_{*}^{1}-\rho_{3}H_{*}^{1})}{\gamma+H_{0}}-\rho_{4}F_{*}^{1}\Big{)}+O((n^{2}+1)\varepsilon)=\frac{\mu}{\gamma+H_{0}}+O((n^{2}+1)\varepsilon), (4.62)

and we are ready to establish the following lemma.

Lemma 4.18.

For each nonnegative nn and small 0<ε10<\varepsilon\ll 1, the following inequality holds:

|p1n(1ε)rεμγ+H0n[(1ε)2n1](1ε)[(1ε)2n+1]G|C(n2+1)ε2,\bigg{|}\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}-\frac{\varepsilon\mu}{\gamma+H_{0}}-\frac{n[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)[(1-\varepsilon)^{2n}+1]}G\bigg{|}\leq C(n^{2}+1)\varepsilon^{2}, (4.63)

where G=(1n2)/(1ε)2G=(1-n^{2})/(1-\varepsilon)^{2}, and the constant CC is independent of ε\varepsilon and nn.

Proof.

The estimate ((4.63)) shall be established by using the explicit formula from Lemma 4.13. Specifically, we take η=μγ+H0\eta=\frac{\mu}{\gamma+H_{0}} and f(r)=f8ηf(r)=f_{8}-\eta. From ((4.62)), we have

fL=f8ηLC(n2+1)ε;\|f\|_{L^{\infty}}=\|f_{8}-\eta\|_{L^{\infty}}\leq C(n^{2}+1)\varepsilon;

we then combine it with Lemma 4.13 to derive

|K[f](r)|C(n+1)ε2,|K[f](r)|C(n2+1)ε2.|K[f](r)|\leq C(n+1)\varepsilon^{2},\hskip 20.00003pt|K[f]^{\prime}(r)|\leq C(n^{2}+1)\varepsilon^{2}. (4.64)

Following Lemmas 4.13 and 4.14, we can explicitly solve p1np_{1}^{n} as

p1n(r)=ψ1(r)+Arn+Brn+K[f](r).p_{1}^{n}(r)=\psi_{1}(r)+Ar^{n}+Br^{-n}+K[f](r).

For ψ1(r)\psi_{1}(r), we use ((4.25)) and Lemma 4.15 to obtain

ψ1(1ε)=ηn(n+2)+O(ε2n),ψ1(1ε)=2ηεn+2+O(ε2),n0,\psi_{1}(1-\varepsilon)=\frac{\eta}{n(n+2)}+O\Big{(}\frac{\varepsilon^{2}}{n}\Big{)},\hskip 20.00003pt\psi_{1}^{\prime}(1-\varepsilon)=\frac{2\eta\varepsilon}{n+2}+O(\varepsilon^{2}),\hskip 20.00003ptn\neq 0,

and

ψ1(1ε)=O(ε2),ψ1(1ε)=ηε+O(ε2),n=0.\psi_{1}(1-\varepsilon)=O(\varepsilon^{2}),\hskip 10.00002pt\psi_{1}^{\prime}(1-\varepsilon)={\eta\varepsilon}+O(\varepsilon^{2}),\hskip 10.00002ptn=0. (4.65)

Together with ((4.32)) ((4.33)) as well as ((4.64)) the first derivative of p1np_{1}^{n} at r=1εr=1-\varepsilon evaluates to

p1n(1ε)r\displaystyle\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r} =ψ1(1ε)+An(1ε)n1Bn(1ε)n1+K[f](1ε)\displaystyle=\;\psi_{1}^{\prime}(1-\varepsilon)+An(1-\varepsilon)^{n-1}-Bn(1-\varepsilon)^{-n-1}+K[f]^{\prime}(1-\varepsilon)
=ψ1(1ε)+n[(1ε)2n1](1ε)[(1ε)2n+1](Gψ1(1ε))+O((n2+1)ε2)\displaystyle=\;\psi_{1}^{\prime}(1-\varepsilon)+\frac{n[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)[(1-\varepsilon)^{2n}+1]}\Big{(}G-\psi_{1}(1-\varepsilon)\Big{)}+O((n^{2}+1)\varepsilon^{2})
=ηε+n[(1ε)2n1](1ε)[(1ε)2n+1]G+O((n2+1)ε2),n0,\displaystyle=\;\eta\varepsilon+\frac{n[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)[(1-\varepsilon)^{2n}+1]}G+O((n^{2}+1)\varepsilon^{2}),\hskip 20.00003ptn\neq 0,

which is equivalent to ((4.63)). It is clear from ((4.65)) that the above formula is also valid for n=0n=0. ∎

Like in ((4.3)), we denote

J2n(μ,ρ4)=1ε2[p1n(1ε)rεμγ+H0n(1n2)[(1ε)2n1](1ε)3[(1ε)2n+1]],J_{2}^{n}(\mu,\rho_{4})=\frac{1}{\varepsilon^{2}}\bigg{[}\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}-\frac{\varepsilon\mu}{\gamma+H_{0}}-\frac{n(1-n^{2})[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2n}+1]}\bigg{]}, (4.66)

which indicates

p1n(1ε)r=εμγ+H0+n(1n2)[(1ε)2n1](1ε)3[(1ε)2n+1]+ε2J2n(μ,ρ4).\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}=\frac{\varepsilon\mu}{\gamma+H_{0}}+\frac{n(1-n^{2})[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2n}+1]}+\varepsilon^{2}J_{2}^{n}(\mu,\rho_{4}). (4.67)

From Lemma 4.18, we immediately obtain that there exists a constant CC which is independent of nn and ε\varepsilon such that

|J2n(μ,ρ4)|C(n2+1).|J_{2}^{n}(\mu,\rho_{4})|\leq C(n^{2}+1).

In addition, we also need to estimate dJ2ndμ\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}. To do that, we take μ\mu derivative of equation ((4.67)) to obtain

dJ2ndμ=J2nμ+J2nρ4ρ4μ=1ε2[r(p1nμ)|r=1εεγ+H0].\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}=\frac{\partial J_{2}^{n}}{\partial\mu}+\frac{\partial J_{2}^{n}}{\partial\rho_{4}}\frac{\partial\rho_{4}}{\partial\mu}=\frac{1}{\varepsilon^{2}}\Big{[}\frac{\partial}{\partial r}\Big{(}\frac{\partial p_{1}^{n}}{\partial\mu}\Big{)}\Big{|}_{r=1-\varepsilon}-\frac{\varepsilon}{\gamma+H_{0}}\Big{]}. (4.68)

In order to estimate the right-hand side of ((4.68)), we differentiate the whole system ((4.7))((4.15)) in μ\mu and follow the same procedures as in Lemmas 4.16 and 4.18. Consequently, a similar result as ((4.63)) can be obtained, i.e.,

|r(p1nμ)|r=1εεγ+H0|C(n2+1)ε2.\bigg{|}\frac{\partial}{\partial r}\Big{(}\frac{\partial p_{1}^{n}}{\partial\mu}\Big{)}\Big{|}_{r=1-\varepsilon}-\frac{\varepsilon}{\gamma+H_{0}}\bigg{|}\leq C(n^{2}+1)\varepsilon^{2}.

Combined with ((4.68)), it follows that |dJ2ndμ|C(n2+1)\Big{|}\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{|}\leq C(n^{2}+1). Therefore we have the following lemma.

Lemma 4.19.

For function J2n(μ,ρ4)J_{2}^{n}(\mu,\rho_{4}) defined in ((4.66)), there exists a constant CC which is independent of ε\varepsilon and nn such that

|J2n(μ,ρ4(μ))|C(n2+1),|dJ2n(μ,ρ4(μ))dμ|C(n2+1).|J_{2}^{n}(\mu,\rho_{4}(\mu))|\leq C(n^{2}+1),\hskip 20.00003pt\Big{|}\frac{\mathrm{d}J_{2}^{n}(\mu,\rho_{4}(\mu))}{\mathrm{d}\mu}\Big{|}\leq C(n^{2}+1). (4.69)

At this point, we are finally ready to prove our main result Theorem 1.1.

Proof of Theorem 1.1.

Substituting ((4.6)) into ((3.71)), we obtain the Fréchet derivative of (R~,μ)\mathcal{F}(\widetilde{R},\mu) in R~\widetilde{R} at the point (0,μ)(0,\mu), namely,

[R~(0,μ)]cos(nθ)=(2p(1ε)r2+p1n(1ε)r)cos(nθ);[\mathcal{F}_{\widetilde{R}}(0,\mu)]\cos(n\theta)=\Big{(}\frac{\partial^{2}p_{*}(1-\varepsilon)}{\partial r^{2}}+\frac{\partial p_{1}^{n}(1-\varepsilon)}{\partial r}\Big{)}\cos(n\theta);

we then combine the above formula with ((4.3)) and ((4.67)) to derive

[R~(0,μ)]cos(nθ)=(εμγ+H0+n(1n2)[(1ε)2n1](1ε)3[(1ε)2n+1]+ε2(J1+J2n))cos(nθ).[\mathcal{F}_{\widetilde{R}}(0,\mu)]\cos(n\theta)=\Big{(}\frac{\varepsilon\mu}{\gamma+H_{0}}+\frac{n(1-n^{2})[(1-\varepsilon)^{2n}-1]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2n}+1]}+\varepsilon^{2}(J_{1}+J_{2}^{n})\Big{)}\cos(n\theta). (4.70)

For fixed nonnegative nn,

(1ε)2n1(1ε)[(1ε)2n+1]=nε+O(n2ε2),\frac{(1-\varepsilon)^{2n}-1}{(1-\varepsilon)[(1-\varepsilon)^{2n}+1]}=-n\varepsilon+O(n^{2}\varepsilon^{2}),

when ε\varepsilon is sufficiently small so that nε<1n\varepsilon<1. In this case, the equation [R~(0,μ)]cos(nθ)=0[\mathcal{F}_{\widetilde{R}}(0,\mu)]\cos(n\theta)=0 is satisfied if and only if

U(μ,ε)μγ+H0n2(1n2)+ε(J1+J2n)+O(n5ε)=0.U(\mu,\varepsilon)\triangleq\frac{\mu}{\gamma+H_{0}}-n^{2}(1-n^{2})+\varepsilon(J_{1}+J_{2}^{n})+O(n^{5}\varepsilon)=0. (4.71)

Notice that both J1J_{1} and J2nJ_{2}^{n} contain μ\mu, it is impossible to solve μ\mu explicitly from equation ((4.71)). However, we are able to claim that for each small ε\varepsilon, ((4.71)) admits a unique solution μ\mu. To prove it, we first find that U((γ+H0)n2(1n2),0)=0U((\gamma+H_{0})n^{2}(1-n^{2}),0)=0; in addition, if we take partial μ\mu derivative on both sides of ((4.71)) and evaluate the value at (μ,0)(\mu,0), we have

μU(μ,0)=[1γ+H0+ε(dJ1dμ+dJ2ndμ)]|ε=0=1γ+H0>0.\frac{\partial}{\partial\mu}U(\mu,0)=\Big{[}\frac{1}{\gamma+H_{0}}+\varepsilon\Big{(}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}+\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{)}\Big{]}\Big{|}_{\varepsilon=0}=\frac{1}{\gamma+H_{0}}>0.

Therefore, it follows from the implicit function theorem that, for each small ε\varepsilon, there exists a unique solution, which is close to (γ+H0)n2(1n2)(\gamma+H_{0})n^{2}(1-n^{2}), such that equation ((4.71)) is satisfied; we denote the unique solution by μn\mu_{n}. In what follows, we shall justify that μ=μn\mu=\mu_{n} with n2n\geq 2 and μn>μc\mu_{n}>\mu_{c} is a bifurcation point for the system ((1.15))((1.22)) when ε\varepsilon is sufficiently small.

What we need to do is to verify the four assumptions of the Crandall-Rabinowitz theorem (Theorem 2.5) at the point μ=μn\mu=\mu_{n}. To begin with, the assumption (1) is naturally satisfied due to Theorem 2.2. To be more specific, for each μn>μc\mu_{n}>\mu_{c}, we can find a small ε>0\varepsilon^{*}>0, such that for 0<ε<ε0<\varepsilon<\varepsilon^{*}, there exists a unique radially symmetric stationary solution, hence (0,μn)=0\mathcal{F}(0,\mu_{n})=0. Next let’s proceed to verify the assumption (2) and (3) for a fixed small ε\varepsilon. It suffices to show that for every mm, mnm\neq n,

[R~(0,μn)]cos(mθ)0,mn,[\mathcal{F}_{\widetilde{R}}(0,\mu_{n})]\cos(m\theta)\neq 0,\hskip 20.00003ptm\neq n, (4.72)

or equivalently,

W(m)εμnγ+H0+m(m21)[1(1ε)2m](1ε)3[(1ε)2m+1]+ε2(J1(μn,ρ4)+J2m(μn,ρ4))0,mn.W(m)\triangleq\frac{\varepsilon\mu_{n}}{\gamma+H_{0}}+\frac{m(m^{2}-1)[1-(1-\varepsilon)^{2m}]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2m}+1]}+\varepsilon^{2}\Big{(}J_{1}(\mu_{n},\rho_{4})+J_{2}^{m}(\mu_{n},\rho_{4})\Big{)}\neq 0,\hskip 10.00002ptm\neq n. (4.73)

To establish statement ((4.72)) (or statement ((4.73))), we split the proof into three cases:

Case (i) m>max{2n,m0}m>\max\{2n,m_{0}\} and mε12m\varepsilon\leq\frac{1}{2}, where m0m_{0} will be determined later. Using the inequality mε12m\varepsilon\leq\frac{1}{2}, together with Lemma 4.15, we deduce that

(1ε)2m12mε+2m2ε212mε+mε1mε,(1-\varepsilon)^{2m}\leq 1-2m\varepsilon+2m^{2}\varepsilon^{2}\leq 1-2m\varepsilon+m\varepsilon\leq 1-m\varepsilon,

hence (recall that n2n\geq 2 so that m>4m>4 in this case)

m(m21)[1(1ε)2m](1ε)3[(1ε)2m+1]ε2m2(m21).\frac{m(m^{2}-1)[1-(1-\varepsilon)^{2m}]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2m}+1]}\geq\frac{\varepsilon}{2}m^{2}(m^{2}-1). (4.74)

In addition, by Lemma 4.12 and Lemma 4.19, there exists a constant CC which does not depend on ε\varepsilon and mm such that,

|J1+J2m||J1|+|J2m|C(m2+1).|J_{1}+J_{2}^{m}|\leq|J_{1}|+|J_{2}^{m}|\leq C(m^{2}+1). (4.75)

Substituting ((4.74)) and ((4.75)) into ((4.73)), we derive

W(m)εμnγ+H0+ε2m2(m21)Cε2(m2+1)=ε[μnγ+H0+12m2(m21)Cε(m2+1)].W(m)\geq\frac{\varepsilon\mu_{n}}{\gamma+H_{0}}+\frac{\varepsilon}{2}m^{2}(m^{2}-1)-C\varepsilon^{2}(m^{2}+1)=\varepsilon\Big{[}\frac{\mu_{n}}{\gamma+H_{0}}+\frac{1}{2}m^{2}(m^{2}-1)-C\varepsilon(m^{2}+1)\Big{]}.

It is clear that W(m)>0W(m)>0 for large mm as the leading order term in the brackets is m42\frac{m^{4}}{2}; hence we can find m0>0m_{0}>0 such that when m>m0m>m_{0},

W(m)>0.W(m)>0.

Case (ii) m>max{2n,m0}m>\max\{2n,m_{0}\} and mε>12m\varepsilon>\frac{1}{2}. In this case, we have

(1ε)2m(112m)2me1,(1-\varepsilon)^{2m}\leq\Big{(}1-\frac{1}{2m}\Big{)}^{2m}\leq e^{-1},

and hence (since m>max{2n,m0}m>\max\{2n,m_{0}\}, we also have m>4m>4 in this case)

m(m21)[1(1ε)2m](1ε)3[(1ε)2m+1]1e12m(m21).\frac{m(m^{2}-1)[1-(1-\varepsilon)^{2m}]}{(1-\varepsilon)^{3}[(1-\varepsilon)^{2m}+1]}\geq\frac{1-e^{-1}}{2}m(m^{2}-1).

Similar as in Case (i), we substitute the above inequality as well as ((4.75)) into ((4.73)), and derive

W(m)εμnγ+H0+1e12m(m21)Cε2(m2+1);W(m)\geq\frac{\varepsilon\mu_{n}}{\gamma+H_{0}}+\frac{1-e^{-1}}{2}m(m^{2}-1)-C\varepsilon^{2}(m^{2}+1);

notice that the leading order term is 1e12m3\frac{1-e^{-1}}{2}m^{3}, we can easily find a bound for ε\varepsilon, denoted by E1E_{1}, such that when ε<E1\varepsilon<E_{1},

W(m)1e14m(m21)>0.W(m)\geq\frac{1-e^{-1}}{4}m(m^{2}-1)>0.

Case (iii) 0mmax{2n,m0}0\leq m\leq\max\{2n,m_{0}\}. From our previous analysis, we know that μn\mu_{n} is close to (γ+H0)n2(1n2)(\gamma+H_{0})n^{2}(1-n^{2}). Since max{2n,m0}\max\{2n,m_{0}\} is a finite number, we can choose ε\varepsilon small and similarly define all μm\mu_{m} for mmax{2n,m0}m\leq\max\{2n,m_{0}\} so that μm\mu_{m} is close to (γ+H0)m2(1m2)(\gamma+H_{0})m^{2}(1-m^{2}); in this case W(m)0W(m)\neq 0 if and only if μnμm\mu_{n}\neq\mu_{m}. To be more specific, we have

limε0μn=(γ+H0)n2(1n2),limε0μm=(γ+H0)m2(1m2).\lim\limits_{\varepsilon\rightarrow 0}\mu_{n}=(\gamma+H_{0})n^{2}(1-n^{2}),\hskip 20.00003pt\lim\limits_{\varepsilon\rightarrow 0}\mu_{m}=(\gamma+H_{0})m^{2}(1-m^{2}).

Since mnm\neq n and n2n\geq 2, it follows that

limε0|μnμm|\displaystyle\lim\limits_{\varepsilon\rightarrow 0}|\mu_{n}-\mu_{m}| min{limε0|μnμn1|,limε0|μnμn+1|}\displaystyle\geq\min\{\lim\limits_{\varepsilon\rightarrow 0}|\mu_{n}-\mu_{n-1}|,\lim\limits_{\varepsilon\rightarrow 0}|\mu_{n}-\mu_{n+1}|\} (4.76)
=(γ+H0)(4n36n2+2n)12(γ+H0).\displaystyle=(\gamma+H_{0})(4n^{3}-6n^{2}+2n)\geq 12(\gamma+H_{0}).

Recall again mm is bounded in this case, we can find a bound for ε\varepsilon, denoted by E2E_{2}, such that when ε<E2\varepsilon<E_{2},

|μnlimε0μn|+max0mmax{2n,m0}|μmlimε0μm|6(γ+H0),\Big{|}\mu_{n}-\lim\limits_{\varepsilon\rightarrow 0}\mu_{n}\Big{|}+\max\limits_{0\leq m\leq\max\{2n,m_{0}\}}\Big{|}\mu_{m}-\lim\limits_{\varepsilon\rightarrow 0}\mu_{m}\Big{|}\leq 6(\gamma+H_{0}),

together with ((4.76)), we obtain

|μnμm|\displaystyle\Big{|}\mu_{n}-\mu_{m}\Big{|} |limε0μnlimε0μm||μnlimε0μn||μmlimε0μm|6(γ+H0)>0.\displaystyle\geq\Big{|}\lim\limits_{\varepsilon\rightarrow 0}\mu_{n}-\lim\limits_{\varepsilon\rightarrow 0}\mu_{m}\Big{|}-\Big{|}\mu_{n}-\lim\limits_{\varepsilon\rightarrow 0}\mu_{n}\Big{|}-\Big{|}\mu_{m}-\lim\limits_{\varepsilon\rightarrow 0}\mu_{m}\Big{|}\geq 6(\gamma+H_{0})>0.

By combining all three cases, the assumptions (2) and (3) in Theorem 2.5 are satisfied when ε\varepsilon is sufficiently small, i.e.,

KerR~(0,μn)=span{cos(nθ)},\displaystyle\text{Ker}\,\mathcal{F}_{\widetilde{R}}(0,\mu_{n})=\text{span}\{\cos(n\theta)\},
Y1=span{1,cos(θ),,cos((n1)θ),cos((n+1)θ),},\displaystyle Y_{1}=\text{span}\{1,\cos(\theta),\cdots,\cos((n-1)\theta),\cos((n+1)\theta),\cdots\},
and Y1Ker=X1l+α,\displaystyle\text{and }Y_{1}\mbox{$\bigoplus$}\text{Ker}=X^{l+\alpha}_{1},

such that the spaces listed here (codimension space, non-tangential space) meet the requirements of the Crandall-Rabinowitz Theorem. To finish the whole proof, it remains to show the last assumption. Differentiating ((4.70)) in μ\mu, we derive

[μR~(0,μ)]cos(nθ)\displaystyle\left[\mathcal{F}_{\mu\widetilde{R}}(0,\mu)\right]\cos(n\theta) =(εγ+H0+ε2(dJ1dμ+dJ2ndμ))cos(nθ)\displaystyle=\Big{(}\frac{\varepsilon}{\gamma+H_{0}}+\varepsilon^{2}\Big{(}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}+\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{)}\Big{)}\cos(n\theta) (4.77)
=ε(1γ+H0+ε(dJ1dμ+dJ2ndμ))cos(nθ).\displaystyle=\varepsilon\Big{(}\frac{1}{\gamma+H_{0}}+\varepsilon\Big{(}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}+\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{)}\Big{)}\cos(n\theta).

By Lemma 4.12 and Lemma 4.19, there exists a constant CC independent of ε\varepsilon and nn such that

|dJ1dμ+dJ2ndμ||dJ1dμ|+|dJ2ndμ|C(n2+1).\Big{|}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}+\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{|}\leq\Big{|}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}\Big{|}+\Big{|}\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{|}\leq C(n^{2}+1). (4.78)

Based on ((4.78)), we can find a bound E3E_{3} (depending on nn), such that when ε<E3\varepsilon<E_{3},

1γ+H0+ε(dJ1dμ+dJ2ndμ)>1γ+H0CE3(n2+1)>0,\frac{1}{\gamma+H_{0}}+\varepsilon\Big{(}\frac{\mathrm{d}J_{1}}{\mathrm{d}\mu}+\frac{\mathrm{d}J_{2}^{n}}{\mathrm{d}\mu}\Big{)}>\frac{1}{\gamma+H_{0}}-CE_{3}(n^{2}+1)>0,

and hence [μR~(0,μ)]cos(nθ)Y1\left[\mathcal{F}_{\mu\widetilde{R}}(0,\mu)\right]\cos(n\theta)\not\in Y_{1}, i.e., the assumption (4) is satisfied.

Combining all pieces together, we take E=min{ε,E1,E2,E3}E=\min\{\varepsilon^{*},E_{1},E_{2},E_{3}\}, then we know that all the assumptions of the Crandall-Rabinowitz theorem are satisfied when ε(0,E)\varepsilon\in(0,E). Hence we conclude that μ=μn\mu=\mu_{n} is a symmetry-breaking bifurcation point. ∎

5 Appendix

5.1 A supersolution

As in [9], we use the function

ξ(r)=1r24+12logr\xi(r)=\frac{1-r^{2}}{4}+\frac{1}{2}\log r

a lot when we apply the maximum principle. Recall that ξ\xi satisfies

Δξ=1,ξr(r)=1r22r,and ξ(r)=O(ε2) when 1ε<r<1.\displaystyle-\Delta\xi=1,\hskip 10.00002pt\xi_{r}(r)=\frac{1-r^{2}}{2r},\hskip 10.00002pt\text{and }\hskip 10.00002pt\xi(r)=O(\varepsilon^{2})\text{ when }1-\varepsilon<r<1.

Next we take

c1(β1,ε)=1β1ε(2ε)2(1ε)ε(2ε)412log(1ε)εβ1+O(ε2),andc2(β1,τ)=2β1|τ|.c_{1}(\beta_{1},\varepsilon)=\frac{1}{\beta_{1}}\frac{\varepsilon(2-\varepsilon)}{2(1-\varepsilon)}-\frac{\varepsilon(2-\varepsilon)}{4}-\frac{1}{2}\log(1-\varepsilon)\equiv\frac{\varepsilon}{\beta_{1}}+O(\varepsilon^{2}),\hskip 10.00002pt\text{and}\hskip 10.00002ptc_{2}(\beta_{1},\tau)=\frac{2}{\beta_{1}}|\tau|.

It is easy to verify that

[ξr+β1(ξ+c1(β1,ε))]|r=1ε=[(ξ+c1(β1,ε))r+β1(ξ+c1(β1,ε))]|r=1ε=0.\Big{[}-\frac{\partial\xi}{\partial r}+\beta_{1}\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)\Big{)}\Big{]}\Big{|}_{r=1-\varepsilon}=\Big{[}-\frac{\partial\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)\Big{)}}{\partial r}+\beta_{1}\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)\Big{)}\Big{]}\Big{|}_{r=1-\varepsilon}=0. (5.1)

Using ((5.1)), also recalling S(θ)C4+α(Σ)1\|S(\theta)\|_{C^{4+\alpha}(\Sigma)}\leq 1, we derive

[(ξ+c1(β1,ε)+c2(β1,τ))𝒏+β1(ξ+c1(β1,ε)+c2(β1,τ))]|r=1ε+τS\displaystyle\bigg{[}\frac{\partial\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)+c_{2}(\beta_{1},\tau)\Big{)}}{\partial{\bm{n}}}+\beta_{1}\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)+c_{2}(\beta_{1},\tau)\Big{)}\bigg{]}\bigg{|}_{r=1-\varepsilon+\tau S}
=\displaystyle= [ξr11+(τS)2+β1ξ]|r=1ε+τS+β1c1(β1,ε)+β1c2(β1,τ)\displaystyle\Big{[}-\frac{\partial\xi}{\partial r}\frac{1}{\sqrt{1+(\tau S^{\prime})^{2}}}+\beta_{1}\xi\Big{]}\Big{|}_{r=1-\varepsilon+\tau S}+\beta_{1}c_{1}(\beta_{1},\varepsilon)+\beta_{1}c_{2}(\beta_{1},\tau)
=\displaystyle= [ξr+β1(ξ+c1(β1,ε))]|r=1ε+[2ξr2+β1ξr]|r=1ετS+2|τ|+O(|τS|2)+O(|τS|2)\displaystyle\Big{[}-\frac{\partial\xi}{\partial r}+\beta_{1}\Big{(}\xi+c_{1}(\beta_{1},\varepsilon)\Big{)}\Big{]}\Big{|}_{r=1-\varepsilon}+\Big{[}-\frac{\partial^{2}\xi}{\partial r^{2}}+\beta_{1}\frac{\partial\xi}{\partial r}\Big{]}\Big{|}_{r=1-\varepsilon}\tau S+2|\tau|+O(|\tau S|^{2})+O(|\tau S^{\prime}|^{2})
=\displaystyle= 0+[1+(1ε)22(1ε)2+β11(1ε)22(1ε)]τS+2|τ|+O(|τS|2)+O(|τS|2)\displaystyle 0+\Big{[}\frac{1+(1-\varepsilon)^{2}}{2(1-\varepsilon)^{2}}+\beta_{1}\frac{1-(1-\varepsilon)^{2}}{2(1-\varepsilon)}\Big{]}\tau S+2|\tau|+O(|\tau S|^{2})+O(|\tau S^{\prime}|^{2})
=\displaystyle= (1+O(ε))τS+2|τ|+O(|τS|2)+O(|τS|2)>0.\displaystyle(1+O(\varepsilon))\tau S+2|\tau|+O(|\tau S|^{2})+O(|\tau S^{\prime}|^{2})>0.

5.2 Transformation TτT_{\tau}

The transformation TτT_{\tau}

Tτ:r~=r12(ετS(θ))+1,θ~=θεT_{\tau}:\widetilde{r}=\frac{r-1}{2(\varepsilon-\tau S(\theta))}+1,\quad\widetilde{\theta}=\frac{\theta}{\varepsilon}

maps Ωτ\Omega_{\tau} to a long stripe region (r~,θ~)[12,1]×[0,2πε](\widetilde{r},\widetilde{\theta})\in[\frac{1}{2},1]\times[0,\frac{2\pi}{\varepsilon}]. Let yy satisfies

Δy=1rr(ryr)1rθ(1ryθ)=f(y)\displaystyle-\Delta y=-\frac{1}{r}\frac{\partial}{\partial r}\Big{(}r\frac{\partial y}{\partial r}\Big{)}-\frac{1}{r}\frac{\partial}{\partial\theta}\Big{(}\frac{1}{r}\frac{\partial y}{\partial\theta}\Big{)}=f(y)\hskip 20.00003pt in Ωτ,\displaystyle\text{in }\Omega_{\tau}, (5.2)

and set y~(r~,θ~)=y(r,θ)y0\widetilde{y}(\widetilde{r},\widetilde{\theta})=y(r,\theta)-y_{0}. Obviously, y~\widetilde{y} should be 2πε\frac{2\pi}{\varepsilon}-periodic in θ~\widetilde{\theta}. Using the chain rule, we obtain:

r=r~r~r=12(ετS)r~=(12ε+O(τS))r~,\displaystyle\frac{\partial}{\partial r}=\frac{\partial}{\partial\widetilde{r}}\frac{\partial\widetilde{r}}{\partial r}=\frac{1}{2(\varepsilon-\tau S)}\frac{\partial}{\partial\widetilde{r}}=\Big{(}\frac{1}{2\varepsilon}+O(\tau S)\Big{)}\frac{\partial}{\partial\widetilde{r}}\,,
θ=r~r~θ+θ~θ~θ=τSετS(r~1)r~+1εθ~=O(τSC1)r~+1εθ~.\displaystyle\frac{\partial}{\partial\theta}=\frac{\partial}{\partial\widetilde{r}}\frac{\partial\widetilde{r}}{\partial\theta}+\frac{\partial}{\partial\widetilde{\theta}}\frac{\partial\widetilde{\theta}}{\partial\theta}=\frac{\tau S^{\prime}}{\varepsilon-\tau S}\Big{(}\widetilde{r}-1\Big{)}\frac{\partial}{\partial\widetilde{r}}+\frac{1}{\varepsilon}\frac{\partial}{\partial\widetilde{\theta}}=O(\tau\|S\|_{C^{1}})\frac{\partial}{\partial\widetilde{r}}+\frac{1}{\varepsilon}\frac{\partial}{\partial\widetilde{\theta}}\,.

Hence we can write the equation of y~(r~,θ~)\widetilde{y}(\widetilde{r},\widetilde{\theta}) as

r~((1+A1)y~r~+A2y~θ~)θ~(A3y~r~+(1+A4)y~θ~)+A5y~r~+A6y~θ~=ε2f~(y~),-\frac{\partial}{\partial\widetilde{r}}\Big{(}(1+A_{1})\frac{\partial\widetilde{y}}{\partial\widetilde{r}}+A_{2}\frac{\partial\widetilde{y}}{\partial\widetilde{\theta}}\Big{)}-\frac{\partial}{\partial\widetilde{\theta}}\Big{(}A_{3}\frac{\partial\widetilde{y}}{\partial\widetilde{r}}+(1+A_{4})\frac{\partial\widetilde{y}}{\partial\widetilde{\theta}}\Big{)}+A_{5}\frac{\partial\widetilde{y}}{\partial\widetilde{r}}+A_{6}\frac{\partial\widetilde{y}}{\partial\widetilde{\theta}}=\varepsilon^{2}\widetilde{f}(\widetilde{y}),

where f~(y~)=rf(y)\widetilde{f}(\widetilde{y})=rf(y), and A1,A2,A3,A4,A5,A6O(ε)A_{1},A_{2},A_{3},A_{4},A_{5},A_{6}\sim O(\varepsilon) are thus bounded. Furthermore, since A1,A2,A3,A4A_{1},A_{2},A_{3},A_{4} contain at most first derivative of SS, they are CαC^{\alpha} if SC1+αS\in C^{1+\alpha}.

5.3 A continuation lemma

The next lemma concerns the continuation of estimates. The proof is standard and we omit the details.

Lemma 5.20.

Let {Qδ(i)}i=1M\{\vec{Q}_{\delta}^{(i)}\}_{i=1}^{M} be a finite collection of real vectors, and define the norm of the vector by |Qδ|max=max1iM|Qδ(i)||\vec{Q}_{\delta}|_{\max}=\max\limits_{1\leq i\leq M}|Q_{\delta}^{(i)}|. Suppose that 0<C1<C20<C_{1}<C_{2}, and

(i) |Q0|maxC1|\vec{Q}_{0}|_{\max}\leq C_{1};

(ii) For any 0<δ10<\delta\leq 1, if |Qδ|maxC2|\vec{Q}_{\delta}|_{\max}\leq C_{2}, then |Qδ|maxC1|\vec{Q}_{\delta}|_{\max}\leq C_{1};

(iii) Qδ\vec{Q}_{\delta} is continuous in δ\delta.

Then |Qδ|maxC1|\vec{Q}_{\delta}|_{\max}\leq C_{1} for all 0<δ10<\delta\leq 1.

Remark 5.1.

If the finite collection is replaced by an infinite collection, then (iii) will need to be replaced by “uniform continuity” in δ\delta.


References

  • [1] Benjamin, E. J., Virani, S. S., Callaway, C. W., Chamberlain, A. M., Chang, A. R., Cheng, S., Chiuve, S. E., Cushman, M., Delling, F. N., Deo, R., de Ferranti, S. D., Ferguson, J. F., Fornage, M., Gillespie, C., Isasi, C. R., Jiménez, M. C., Jordan, L. C., Judd, S. E., Lackland, D., Lichtman, J. H., Lisabeth, L., Liu, S., Longenecker, C. T., Lutsey, P. L., Mackey, J. S., Matchar, D. B., Matsushita, K., Mussolino, M. E., Nasir, K., O’Flaherty, M., Palaniappan, L. P., Pandey, A., Pandey, D. K., Reeves, M. J., Ritchey, M. D., Rodriguez, C. J., Roth, G. A., Rosamond, W. D., Sampson, U. K., Satou, G. M., Shah, S. H., Spartano, N. L., Tirschwell, D. L., Tsao, C. W., Voeks, J. H., Willey, J. Z., Wilkins, J. T., Wu, J. H., Alger, H. M., Wong, S. S., and Muntner, P. Heart disease and stroke statistics-2018 update: a report from the american heart association. Circulation 137, 12 (2018), e67.
  • [2] Calvez, V., Ebde, A., Meunier, N., and Raoult, A. Mathematical modelling of the atherosclerotic plaque formation. In ESAIM: Proceedings (2009), vol. 28, EDP Sciences, pp. 1–12.
  • [3] Cohen, A., Myerscough, M. R., and Thompson, R. S. Athero-protective effects of high density lipoproteins (HDL): An ODE model of the early stages of atherosclerosis. Bulletin of mathematical biology 76, 5 (2014), 1117–1142.
  • [4] Crandall, M. G., and Rabinowitz, P. H. Bifurcation from simple eigenvalues. Journal of Functional Analysis 8, 2 (1971), 321–340.
  • [5] Cui, S., and Escher, J. Bifurcation analysis of an elliptic free boundary problem modelling the growth of avascular tumors. SIAM Journal on Mathematical Analysis 39 (2007), 210–235.
  • [6] Fontelos, M., and Friedman, A. Symmetry-breaking bifurcations of free boundary problems in three dimensions. Asymptotic Analysis 35 (2003), 187–206.
  • [7] Friedman, A. Mathematical Biology, vol. 127. American Mathematical Soc., 2018.
  • [8] Friedman, A., and Hao, W. A mathematical model of atherosclerosis with reverse cholesterol transport and associated risk factors. Bulletin of mathematical biology 77, 5 (2015), 758–781.
  • [9] Friedman, A., Hao, W., and Hu, B. A free boundary problem for steady small plaques in the artery and their stability. Journal of Differential Equations 259, 4 (2015), 1227–1255.
  • [10] Friedman, A., and Hu, B. Bifurcation from stability to instability for a free boundary problem arising in a tumor model. Archive for rational mechanics and analysis 180, 2 (2006), 293–330.
  • [11] Friedman, A., and Reitich, F. Symmetry-breaking bifurcation of analytic solutions to free boundary problems: An application to a model of tumor growth. Transactions of the American Mathematical Society 353 (2000), 1587–1634.
  • [12] Gilbarg, D., and Trudinger, N. Elliptic Partial Differential Equations of Second Order. Springer-Verlag, New York, 1983.
  • [13] Hao, W., and Friedman, A. The LDL-HDL profile determines the risk of atherosclerosis: a mathematical model. PloS one 9, 3 (2014), e90497.
  • [14] Hao, W., Hauenstein, J. D., Hu, B., Liu, Y., Sommese, A. J., and Zhang, Y.-T. Bifurcation for a free boundary problem modeling the growth of a tumor with a necrotic core. Nonlinear Analysis: Real World Applications 13, 2 (2012), 694–709.
  • [15] Hao, W., Hauenstein, J. D., Hu, B., and Sommese, A. J. A three-dimensional steady-state tumor system. Applied Mathematics and Computation 218, 6 (2011), 2661–2669.
  • [16] Huang, Y., Zhang, Z., and Hu, B. Bifurcation for a free-boundary tumor model with angiogenesis. Nonlinear Analysis: Real World Applications 35 (2017), 483–502.
  • [17] Li, F., and Liu, B. Bifurcation for a free boundary problem modeling the growth of tumors with a drug induced nonlinear proliferation rate. Journal of Differential Equations 263 (2017), 7627–7646.
  • [18] McKay, C., McKee, S., Mottram, N., Mulholland, T., Wilson, S., Kennedy, S., and Wadsworth, R. Towards a model of atherosclerosis. University of Strathclyde (2005), 1–29.
  • [19] Mukherjee, D., Guin, L. N., and Chakravarty, S. A reaction–diffusion mathematical model on mild atherosclerosis. Modeling Earth Systems and Environment (2019), 1–13.
  • [20] Pan, H., and Xing, R. Bifurcation for a free boundary problem modeling tumor growth with ECM and MDE interactions. Nonlinear Analysis: Real World Applications 43 (2018), 362–377.
  • [21] Song, H., Hu, B., and Wang, Z. Stationary solutions of a free boundary problem modeling the growth of vascular tumors with a necrotic core. Discrete & Continuous Dynamical Systems - B 22 (2021), to appear.
  • [22] Centers for Disease Control and Prevention and others. Underlying cause of death, 1999-2013. National Center for Health Statistics, Hyattsville, MD (2015).
  • [23] Wang, Z. Bifurcation for a free boundary problem modeling tumor growth with inhibitors. Nonlinear Analysis: Real World Applications 19 (2014), 45–53.
  • [24] Wu, J. Stationary solutions of a free boundary problem modeling the growth of tumors with gibbs-thomson relation. Journal of Differential Equations 260 (2016), 5875–5893.
  • [25] Wu, J., and Zhou, F. Bifurcation analysis of a free boundary problem modelling tumor growth under the action of inhibitors. Nonlinearity 25 (2012), 2971–2991.
  • [26] Zhao, X. E., and Hu, B. The impact of time delay in a tumor model. Nonlinear Analysis: Real World Applications 51 (2020), 103015.
  • [27] Zhao, X. E., and Hu, B. Symmetry-breaking bifurcation for a free-boundary tumor model with time delay. Journal of Differential Equations 269 (2020), 1829–1862.