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

The steady state of the inclined problem

Xiaoding Yang
Abstract

The inclined problem is a problem that describes an open fluid flowing over an angled wall. It has broad applications in science and engineering. In this paper, we study the steady state of the inclined problem in two dimensions. The steady-state solution is depicted by the Euler-Lagrange equation of a given energy functional with a fixed contact angle as the boundary condition. By choosing a suitable maximal point to parameterize the surface of the fluid, we can construct a solution to this Euler-Lagrange equation via a shooting method in terms of the volume of the fluid. The construction works for any contact angle and any arbitrary inclined angle.

1 Introduction

Fluid mechanics is a fundamental branch of physics and engineering that deals with the behavior of fluids in motion and at rest. In recent years, an increasing number of math studies have focused on this field. One particular area of interest is fluid that flows over inclined surfaces, which is so-called fluid inclined problems. These problems arise in various natural and industrial applications such as the movement of water on sloped terrains, oil transport through pipelines, and the design of spillways and drainage systems[1].

When a fluid interacts with an inclined surface, several forces influence its motion, such as gravity, pressure, and surface tension. It is crucial to understand those effects if we want to predict flow behavior, optimize engineering designs, and improve efficiency in fluid-based systems. Even for the steady state (which means that the velocity of the fluid equals zero everywhere), discussing the balancing between different forces acting on fluids is quite interesting and important. It can give us some information about the moving fluid. Because of this, considerable effort has been devoted to analyze the shape of the equilibrium and steady state by mathematicians and engineers. But there is still a lot of work to do, including theoretical and computational problems.

Based on the practical problems we aim to study, the inclined problem typically falls into two distinct settings. The first involves the motion of a fluid flowing down an inclined surface—for instance, a thin film of fluid sliding down a tilted plane under the influence of gravity. The second setting concerns the motion of an open fluid flowing over an angled wall, as illustrated in Figure 1. This scenario can also be interpreted as a fluid mechanics problem within a triangular container, such as the cross-section of water in a channel. Both settings are common in everyday life and are important to discuss. In this paper, we focus exclusively on the second setting.

1.1 Formulation: Based on the discussion in the Background part, the purpose of this paper is to study the static viscous incompressible fluid occupied in an open-top vessel in two dimensions. The vessel is modeled as an open, connected, bounded triangular subset Γ2\Gamma\subset\mathbb{R}^{2} that obeys the following pair of assumptions. See Figure 1. First, we posit that two inclined angles θ1\theta_{1} and θ2\theta_{2} that are defined as angles between the wall of the vessel and the horizontal plane are both angles between 0 and π\pi. Second, we assume that θ1+θ2<π\theta_{1}+\theta_{2}<\pi. Finally, it is assumed that the fluid occupies a subset Ω\Omega of the vessel Γ\Gamma, resulting in a free boundary where the fluid meets the air above the vessel.

In this article, we only discuss the steady state, which is equivalent to finding a static boundary surface. In the scenario depicted in Figure 1, the surface can be represented as a graph of x, which is denoted by v(x). However, this is not always the case; for instance, the boundary curve shown in Figure 3 cannot be represented as a graph of x. This observation forces us to describe the surface in polar coordinates. We choose the intersection point of two vessel walls (denoted point O in Figure 1) as the original point and then use ρ(θ)\rho(\theta) to represent the surface. Then we need to determine the equation that the boundary surface satisfies.

Refer to caption
Figure 1:

We first write down the energy functional with the conservation of the total volume in the polar coordinates.

E(ρ(θ))=13gθ2πθ1ρ3sinθdθ+θ2πθ1σρ2+ρ2𝑑θ[[γ]](ρ(θ1)+ρ(θ2)){}E(\rho(\theta))=\frac{1}{3}g\int_{\theta_{2}}^{\pi-\theta_{1}}\rho^{3}\sin\theta d\theta+\int_{\theta_{2}}^{\pi-\theta_{1}}\sigma\sqrt{\rho^{2}+\rho^{\prime 2}}d\theta-[[\gamma]](\rho(\theta_{1})+\rho(\theta_{2})) (1)
V=𝒱(ρ(θ))=12θ2πθ1ρ2(θ)𝑑θ{}V=\mathcal{V}(\rho(\theta))=\frac{1}{2}\int_{\theta_{2}}^{\pi-\theta_{1}}\rho^{2}(\theta)d\theta (2)

The first term of the functional E defined in (1) is the gravity energy and the second term is the surface tension energy. The third term is the free boundary energy caused by the interface between the fluid and the vessel. The second equation is the conservation of the total energy, where V is an arbitrary given fixed number. To make the model meaningful, we should assume that 1[[γ]]σ1-1\leq\frac{[[\gamma]]}{\sigma}\leq 1.

Using the variation method for this energy functional E, we can write down the Euler-Lagrange equation:

{gρ2sinθ+σρρ2+ρ2σθ(ρρ2+ρ2)P0ρ=0ρρ2+ρ2(θ2)=[[γ]]σρρ2+ρ2(πθ1)=[[γ]]σ{}\begin{cases}g\rho^{2}\sin\theta+\sigma\frac{\rho}{\sqrt{\rho^{2}+\rho^{\prime 2}}}-\sigma\partial_{\theta}(\frac{\rho^{\prime}}{\sqrt{\rho^{2}+\rho^{\prime 2}}})-P_{0}\rho=0\\ \frac{\rho^{\prime}}{\sqrt{\rho^{2}+\rho^{\prime 2}}}(\theta_{2})=\frac{[[\gamma]]}{\sigma}\\ \frac{\rho^{\prime}}{\sqrt{\rho^{2}+\rho^{\prime 2}}}(\pi-\theta_{1})=-\frac{[[\gamma]]}{\sigma}\end{cases} (3)

P0P_{0} is the Lagrange multiplier that depends only on the given volume VV. The purpose of the paper is to construct a solution to the Euler-Lagrange equation (3). Our approach involves selecting a special point and utilizing equation (3) to construct a curve parametrized by the coordinate of this special point, which we designate as special parameters. We then represent the total volume 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)) and pressure P0P_{0} in terms of these chosen parameters. We finally determine the value of special parameters via a shooting method by setting 𝒱=V\mathcal{V}=V, thereby obtaining the desired solution function.

We notice that the only unknown parameter in the equation system (3) is P0P_{0}. We introduce the following shift to eliminate it:

(x,y)(x,yP0g){}(x,y)\rightarrow(x,y-\frac{P_{0}}{g}) (4)

We will see how this shift works in the following sessions.

To better understand the system of equations (3), we define a new variable ψ\psi as follows.

{sinψ(θ)=sinθρ+ρcosθρ2+ρ2cosψ(θ)=ρsinθρcosθρ2+ρ2{}\begin{cases}\sin\psi(\theta)=\frac{\sin\theta\rho^{\prime}+\rho\cos\theta}{\sqrt{\rho^{2}+\rho^{\prime 2}}}\\ \cos\psi(\theta)=\frac{\rho\sin\theta-\rho^{\prime}\cos\theta}{\sqrt{\rho^{2}+\rho^{\prime 2}}}\end{cases} (5)

which represents the slope angle in Cartesian coordinates. Due to the complexity of its formula, the definition may be difficult to grasp. But figure 2 can explicitly express the geometric meaning of ψ\psi.

1.3 The structure: We introduce the following definition.

Refer to caption
Figure 2:
sinψ1\displaystyle\sin\psi_{1} :=sinψ(πθ1)=sin(γ+π2θ1)=([[γ]]σ)sinθ1+1[[γ]]2σ2cosθ1\displaystyle:=\sin\psi(\pi-\theta_{1})=\sin(\gamma+\frac{\pi}{2}-\theta_{1})=(-\frac{[[\gamma]]}{\sigma})\sin\theta_{1}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{1} (6)
sinψ2\displaystyle\sin\psi_{2} :=sinψ(θ2)=sin(γπ2+θ2)=([[γ]]σ)sinθ2+1[[γ]]2σ2cosθ2\displaystyle:=\sin\psi(\theta_{2})=\sin(-\gamma-\frac{\pi}{2}+\theta_{2})=(\frac{[[\gamma]]}{\sigma})\sin\theta_{2}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{2} (7)

where ψ\psi is defined in (5) and γ[π2,π2]\gamma\in[-\frac{\pi}{2},\frac{\pi}{2}] is defined as

sinγ=[[γ]]σ{}\sin\gamma=-\frac{[[\gamma]]}{\sigma} (8)

Then we divide our problem into two cases based on the sign of ψ1\psi_{1} and ψ2\psi_{2}.

1.3.1 ψ1\psi_{1} and ψ2\psi_{2} have different sign: In Section 2 we show that ψ\psi is a global variable for the boundary curve. See figure 3. Then we consider the equation system (10) which is equivalent to the original equation in (3).

{dxdψ=σcosψgyP0dydψ=σdydxdxdψ=σtanψcosψgyP0=σsinψgyP0sinψ(πθ1)=sinψ1=([[γ]]σ)sinθ1+1[[γ]]2σ2cosθ1sinψ(θ2)=sinψ2=([[γ]]σ)sinθ2+1[[γ]]2σ2cosθ2{}\begin{cases}\frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy-P_{0}}\\ \frac{dy}{d\psi}=\sigma\frac{dy}{dx}\frac{dx}{d\psi}=\sigma\tan\psi\frac{\cos\psi}{gy-P_{0}}=\sigma\frac{\sin\psi}{gy-P_{0}}\\ \sin\psi(\pi-\theta_{1})=\sin\psi_{1}=(-\frac{[[\gamma]]}{\sigma})\sin\theta_{1}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{1}\\ \sin\psi(\theta_{2})=\sin\psi_{2}=(\frac{[[\gamma]]}{\sigma})\sin\theta_{2}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{2}\end{cases} (9)

where:

{y(ψ)=(ρsinθ)(ψ)x(ψ)=(ρcosθ)(ψ)\begin{cases}y(\psi)=(\rho\sin\theta)(\psi)\\ x(\psi)=(\rho\cos\theta)(\psi)\end{cases}

Then we apply the shift we introduced in (5) and the system (9) becomes:

{dxdψ=σcosψgydydψ=σdydxdxdψ=σtanψcosψgy=σsinψgysinψ(πθ1)=([[γ]]σ)sinθ1+1[[γ]]2σ2cosθ1sinψ(θ2)=([[γ]]σ)sinθ2+1[[γ]]2σ2cosθ2{}\begin{cases}\frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy}\\ \frac{dy}{d\psi}=\sigma\frac{dy}{dx}\frac{dx}{d\psi}=\sigma\tan\psi\frac{\cos\psi}{gy}=\sigma\frac{\sin\psi}{gy}\\ \sin\psi(\pi-\theta_{1})=(-\frac{[[\gamma]]}{\sigma})\sin\theta_{1}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{1}\\ \sin\psi(\theta_{2})=(\frac{[[\gamma]]}{\sigma})\sin\theta_{2}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{2}\end{cases} (10)

To derive a formula for the total volume of Ω\Omega, we first consider the equation in system (10) without the boundary condition, which is the following equation:

{dxdψ=σcosψgydydψ=σdydxdxdψ=σtanψcosψgy=σsinψgy{}\begin{cases}\frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy}\\ \frac{dy}{d\psi}=\sigma\frac{dy}{dx}\frac{dx}{d\psi}=\sigma\tan\psi\frac{\cos\psi}{gy}=\sigma\frac{\sin\psi}{gy}\end{cases} (11)

From the discussion in Chapter two, we can see that any solution satisfying the system (10) reaches the maximum at an interior point. We define them as follows.

um\displaystyle u_{m} =supψ2ψψ1y(ψ)=y(0)\displaystyle=\sup_{\psi_{2}\leq\psi\leq\psi_{1}}y(\psi)=y(0) (12)
xm\displaystyle x_{m} =x(ψ=0)=x(0)\displaystyle=x(\psi=0)=x(0) (13)

Using the existence and uniqueness theorem for ODE solution, we can obtain a solution curve by combining (11),(12) and (13) together. Then we have the following theorem for the total volume 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)).

Refer to caption
Figure 3:
Refer to caption
Figure 4:
Theorem 1.1.

The functional 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)) and pressure P0P_{0} can be represented as a function of umu_{m}. And we have the following two asymptotic properties of 𝒱\mathcal{V}:

limum0𝒱(um)\displaystyle\lim_{u_{m}\rightarrow 0}\mathcal{V}(u_{m}) =+\displaystyle=+\infty (14)
limum𝒱(um)\displaystyle\lim_{u_{m}\rightarrow-\infty}\mathcal{V}(u_{m}) =0\displaystyle=0 (15)

Consequently, for any V, there exists a umu_{m} and a solution to equation system (10) such that 𝒱(ρ(θ))=𝒱(um)=V\mathcal{V}(\rho(\theta))=\mathcal{V}(u_{m})=V. Furthermore, 𝒱(um)\mathcal{V}(u_{m}) is a monotone increasing function of umu_{m} when [[γ]]<0[[\gamma]]<0 which implies that for any given V, there exists a umu_{m} and a solution to (10) such that 𝒱(um)=V\mathcal{V}(u_{m})=V. And when [[γ]]<0[[\gamma]]<0, the choice of umu_{m} and the solution are unique.

Note: From Theorem 1.1, we can get a solution (x(ψ),y(ψ))(x(\psi),y(\psi)) to (10) and P0P_{0} for any given constant V. Therefore, (x(ψ),y(ψ)+P0g)(x(\psi),y(\psi)+\frac{P_{0}}{g}) is the solution to the original system.

1.3.2: Case two:ψ1\psi_{1} and ψ2\psi_{2} have the same sign

In this case we can show that the surface curve is a graph of x. See Figure 4. So we discuss the problem in Cartesian coordinates and use u(x)u(x) to represent the surface curve. We have the equation system below which is equivalent to the system (3):

{σx(xv1+|xu|2)=gvP0xv1+|xv|2(x1)=sinψ1xv1+|xv|2(x2)=sinψ2{}\begin{cases}\sigma\partial_{x}(\frac{\partial_{x}v}{\sqrt{1+|\partial_{x}u|^{2}}})=gv-P_{0}\\ \frac{\partial_{x}v}{\sqrt{1+|\partial_{x}v|^{2}}}(x_{1})=\sin\psi_{1}\\ \frac{\partial_{x}v}{\sqrt{1+|\partial_{x}v|^{2}}}(x_{2})=\sin\psi_{2}\end{cases} (16)

Here, ψ1\psi_{1} and ψ2\psi_{2} are given by equations (6) and (7). Also:

(x,y)\displaystyle(x,y) =(ρcosθ,ρsinθ)\displaystyle=(\rho\cos\theta,\rho\sin\theta) (17)

Then we apply the shift (5) again to change the system (18) to:

{σx(xu1+|xu|2)=guxu1+|xu|2(x1)=sinψ1xu1+|xu|2(x2)=sinψ2{}\begin{cases}\sigma\partial_{x}(\frac{\partial_{x}u}{\sqrt{1+|\partial_{x}u|^{2}}})=gu\\ \frac{\partial_{x}u}{\sqrt{1+|\partial_{x}u|^{2}}}(x_{1})=\sin\psi_{1}\\ \frac{\partial_{x}u}{\sqrt{1+|\partial_{x}u|^{2}}}(x_{2})=\sin\psi_{2}\end{cases} (18)

We mainly discuss this system in Chapter 3.

Just like in the previous case, we examine the equation without boundary conditions first:

σx(sinψ)=σx(xu1+|xv|2)=gu{}\sigma\partial_{x}(\sin\psi)=\sigma\partial_{x}(\frac{\partial_{x}u}{\sqrt{1+|\partial_{x}v|^{2}}})=gu (19)

ψ\psi is defined by the equation (5). Then we choose another special point. We can show that the slope angle ψ\psi reaches its maximum value at some point. We have the following definitions:

{ψm:=supθ2θθ1ψ(θ)um:=u(θm)xm:=x(θm){}\begin{cases}\psi_{m}:=\sup_{\theta_{2}\leq\theta\leq\theta_{1}}\psi(\theta)\\ u_{m}:=u(\theta_{m})\\ x_{m}:=x(\theta_{m})\end{cases} (20)

Here θm\theta_{m} is defined to be the angle such that ψ(θ)=ψm\psi(\theta)=\psi_{m}. Then we choose the point and angle defined by (20) as the initial conditions and get a solution by combining these conditions with (19). We will show the following theorem in Section 3:

Theorem 1.2.

We have two different circumstances.

  • When θm=πθ1\theta_{m}=\pi-\theta_{1} or θm=θ2\theta_{m}=\theta_{2} (without loss of generality, we can assume that θm=πθ1\theta_{m}=\pi-\theta_{1}), we have ψm=ψ1\psi_{m}=\psi_{1} is a constant. We can choose umu_{m} as the independent variable. Then 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)) and P0P_{0} can be represented as functions of umu_{m}. We have the following result:

    limum𝒱(um)=0\lim_{u_{m}\rightarrow-\infty}\mathcal{V}(u_{m})=0

    and:

    V1:=maxum(,0)𝒱(um)V_{1}:=\max_{u_{m}\in(-\infty,0)}\mathcal{V}(u_{m})

    is a bounded constant. Therefore, there exist a umu_{m} such that 𝒱(um)=V\mathcal{V}(u_{m})=V(defined in equation (2)) and a solution function of system (18) for any VV1V\leq V_{1}.

  • When θm(θ2,πθ1)\theta_{m}\in(\theta_{2},\pi-\theta_{1}), we can show um=0u_{m}=0 and we choose ψm\psi_{m} as the independent variable. Then 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)) and P0P_{0} can be expressed as functions of ψm\psi_{m} and we can show the following result:

    limψm0𝒱(ψm)=+\lim_{\psi_{m}\rightarrow 0}\mathcal{V}(\psi_{m})=+\infty

    and:

    Vm:=minψm(max(ψ1,ψ2),0)𝒱(ψm)V_{m}:=\min_{\psi_{m}\in(\max(\psi_{1},\psi_{2}),0)}\mathcal{V}(\psi_{m})

    is a positive constant. Consequently, there exist a ψm\psi_{m} such that 𝒱(ψm)=V\mathcal{V}(\psi_{m})=V and a solution function of system (18) for any VVmV\geq V_{m}

Finally, we have the following result:

Theorem 1.3.

V1VmV_{1}\geq V_{m}, which means that there exists a solution function of the system (18) for any V0V\geq 0 and the solution is not unique when V1>VmV_{1}>V_{m}.

1.4 previous work: The study of the functional defined by (1) and (2) begins with analyzing the Euler-Lagrange equation, which is derived by Young[18], Laplace[13], and Gauss[3] along with the boundary condition of the contact angle using variational methods.

For fixed contact angle, Robert Finn investigated the existence of solutions in specific cases[2]. Finn proved the existence of an equilibrium of the sessile drop for any given volume V. In addition, he proved the nonexistence of the equilibrium of a liquid drop on an inclined wall. However, to our knowledge, general contact line problems involving different slopes remain largely unexplored.

In recent work, Guo-Tice[8, 7] studied the dynamic stability of the steady state when θ1=θ2=π2\theta_{1}=\theta_{2}=\frac{\pi}{2}. The well-posedness theory of this problem is established by [9, 6, 19], while [4, 5] demonstrates certain decay properties of the water wave in a periodic domain and whole space, which indirectly imply the existence of the steady state. Beyond the contact line problem, Tice has analyzed the dynamics and stability of water waves in various settings related to the contact line problem in his papers [12, 10, 11, 16, 14, 15]. In particular, Tice-Wu[17] examined the dynamic stability of the steady state of the sessile drop which can be viewed as the flat version of the contact line problem where θ1=θ2=0\theta_{1}=\theta_{2}=0. Tice and Wu also showed the existence and stability of the steady state in their paper [17]. All of these problems involve dynamic contact points and contact angles, making them challenging to solve. Their work has inspired us to study a generalized version of the steady-state problem and has provided valuable insight. In this paper, our objective is to show the existence of a steady state for any given volume V and any incline angle θ1\theta_{1} and θ2\theta_{2}.

2 The first case: ψ1ψ2<0\psi_{1}\cdot\psi_{2}<0

In this chapter, we discuss the case that ψ1\psi_{1} and ψ2\psi_{2} have opposite signs. We begin by discussing the possible shape of the curve in Subsection 2.1 and then compute the total volume using the chosen parameter umu_{m}. Without loss of generality, we can assume that ψ1>0\psi_{1}>0 and ψ2<0\psi_{2}<0 are as shown in Figure 3 and θ1>θ2\theta_{1}>\theta_{2}.

2.1 The shape of the curve

In this subsection, we first assume that the equation system (10) has a solution function. We rewrite the system as follows.

{dxdψ=σcosψgydydψ=σdydxdxdψ=σtanψcosψgy=σsinψgysinψ(πθ1)=([[γ]]σ)sinθ1+1[[γ]]2σ2cosθ1sinψ(θ2)=([[γ]]σ)sinθ2+1[[γ]]2σ2cosθ2\begin{cases}\frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy}\\ \frac{dy}{d\psi}=\sigma\frac{dy}{dx}\frac{dx}{d\psi}=\sigma\tan\psi\frac{\cos\psi}{gy}=\sigma\frac{\sin\psi}{gy}\\ \sin\psi(\pi-\theta_{1})=(-\frac{[[\gamma]]}{\sigma})\sin\theta_{1}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{1}\\ \sin\psi(\theta_{2})=(\frac{[[\gamma]]}{\sigma})\sin\theta_{2}+\sqrt{1-\frac{[[\gamma]]^{2}}{\sigma^{2}}}\cos\theta_{2}\end{cases}

We first use it to study the shape of the solution curve. In fact, we want to show that the slope angle ψ\psi(defined by equation (5)) is a global legal variable for this solution curve. In addition, the y component of the curve reaches its maximum at an angle θm(θ2,πθ1)\theta_{m}\in(\theta_{2},\pi-\theta_{1}) since ψ1>0\psi_{1}>0 and ψ2<0\psi_{2}<0. We let umu_{m} equal the maximum of the y component of the curve(the definition is given by (12)) and we have the following theorem for ψ\psi and umu_{m}.

Theorem 2.1.

(The shape of the surface curve) We define ψ\psi as in equation (5) and umu_{m} is the maximum of the y component of the solution curve. Then um<0u_{m}<0 and ψ\psi are global legal parameters for the curve, meaning that the boundary curve can be represented as a function of ψ\psi.

Proof.

We first start looking at the shape of the curve. To begin with, we study the curve near the boundary point (x1,y1)=(x(πθ1),y(πθ1))(x_{1},y_{1})=(x(\pi-\theta_{1}),y(\pi-\theta_{1})).

Case 1 0ψ(x1)<π20\leq\psi(x_{1})<\frac{\pi}{2}:

In this setting, we know that sinψ(x1)\sin\psi(x_{1}) and cosψ(x1)\cos\psi(x_{1}) are both greater than 0. From the first and second equation in the system (10):

{dydψ=σsinψgydxdψ=σcosψgy{}\begin{cases}\frac{dy}{d\psi}=\sigma\frac{\sin\psi}{gy}\\ \frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy}\end{cases} (21)

We know that both dudψ\frac{du}{d\psi} and dxdψ\frac{dx}{d\psi} are greater than 0 at the point ψ(x1)\psi(x_{1}) due to our assumption. Therefore, the curve is locally a map of ψ\psi. We want to show that y1=y(ψ1)=y(πθ1)<0y_{1}=y(\psi_{1})=y(\pi-\theta_{1})<0. We prove this by using a contradiction argument. We suppose that y10y_{1}\geq 0 and we separate the main proof into four steps.

Step 1: The local map of ψ\psi can be extended to a map of ψ(ψ1,2πψ1)\psi\in(\psi_{1},2\pi-\psi_{1}).

Using the assumption that u10u_{1}\geq 0, we know from (21) that dudψ\frac{du}{d\psi} and dxdψ\frac{dx}{d\psi} are equal to or greater than 0 near ψ1\psi_{1}. So u(ψ)u(\psi) and x(ψ)x(\psi) are both an increasing function of ψ\psi near ψ=ψ1\psi=\psi_{1}. Using a bootstrap argument, we can show that u(ψ)0u(\psi)\geq 0 and dudψ(ψ)0\frac{du}{d\psi}(\psi)\geq 0 for any ψ(ψ1,π2)\psi\in(\psi_{1},\frac{\pi}{2}). Then we want to extend the curve across the vertical point ψ=π2\psi=\frac{\pi}{2}.

We can solve the first equation in the system (21) by integrating both sides of the equation from ψ1\psi_{1} to any arbitrary angle ψ\psi to get the following equation:

12gy2(ψ)12gy12=σcosψ(x1)σcosψ{}\frac{1}{2}gy^{2}(\psi)-\frac{1}{2}gy_{1}^{2}=\sigma\cos\psi(x_{1})-\sigma\cos\psi (22)

So when ψ=π2\psi=\frac{\pi}{2}, y(π2)y(\frac{\pi}{2}) is finite. Therefore, from the (21) we know that:

dydψ(π2)=1gy(π2)0\frac{dy}{d\psi}(\frac{\pi}{2})=\frac{1}{gy(\frac{\pi}{2})}\neq 0

So, the solution function (x(ψ),u(ψ))(x(\psi),u(\psi)) of the equation system (21) can be extended as a solution of the same system across the point ψ=π2\psi=\frac{\pi}{2}.

When π2<ψπ\frac{\pi}{2}<\psi\leq\pi, we have cosψ<0\cos\psi<0 and sinψ>0\sin\psi>0. Then x(ψ)x(\psi) monotone decreasing and y(ψ)y(\psi) monotone increasing. When ψ=π\psi=\pi, we define x0:=x(π)x_{0}:=x(\pi).

When π<ψ<32π\pi<\psi<\frac{3}{2}\pi we have x(ψ)x(\psi) monotone decreasing and y(ψ)y(\psi) monotone decreasing if y(ψ)>0y(\psi)>0. We will prove that y(ψ)=y(2πψ)y(\psi)=y(2\pi-\psi) in step 2, which means that the curve is symmetric and y(ψ)>0y(\psi)>0 for any π<ψ<32π\pi<\psi<\frac{3}{2}\pi. Therefore, y(ψ)>0y(\psi)>0 for any ψ(π,32π)\psi\in(\pi,\frac{3}{2}\pi), which implies that y(ψ)y(\psi) monotone decreasing. See Figure 3. Again, the curve can be extended across the point ψ=3π2\psi=\frac{3\pi}{2} due to the same reason as the point π2\frac{\pi}{2} and finally it can be extended to the point ψ=2πψ1\psi=2\pi-\psi_{1}.

Refer to caption
Figure 5:

Step 2: The curve constructed in step 1 exhibits the property y(ψ)=y(2πψ)y(\psi)=y(2\pi-\psi), which implies the symmetry of the curve.

Applying the equation (22), We have the following equation:

12gy2(γ)12gy12=σcosψ1σcosγ{}\frac{1}{2}gy^{2}(\gamma)-\frac{1}{2}gy_{1}^{2}=\sigma\cos\psi_{1}-\sigma\cos\gamma (23)

for any γ(ψ1,2πψ1)\gamma\in(\psi_{1},2\pi-\psi_{1}). We set ψ=2πγ\psi=2\pi-\gamma in equation (22) to obtain:

12gy2(2πψ)12gy12=σcosψ1σcos(2πγ)=σcosψ1σcosγ{}\frac{1}{2}gy^{2}(2\pi-\psi)-\frac{1}{2}gy_{1}^{2}=\sigma\cos\psi_{1}-\sigma\cos(2\pi-\gamma)=\sigma\cos\psi_{1}-\sigma\cos\gamma (24)

We combine equations (23) and (24) together to show that:

y(ψ)2=y2(2πψ){}y(\psi)^{2}=y^{2}(2\pi-\psi) (25)

which means that y(ψ)=y(2πψ)y(\psi)=y(2\pi-\psi) or y(ψ)=y(2πψ)y(\psi)=-y(2\pi-\psi). We assume that there exists a ψ0(π,2πψ1)\psi_{0}\in(\pi,2\pi-\psi_{1}) such that y(ψ0)=y(2πψ0)y(\psi_{0})=-y(2\pi-\psi_{0}). Then 2πψ0(ψ1,π)2\pi-\psi_{0}\in(\psi_{1},\pi). However, we know that y(2πψ0)>0y(2\pi-\psi_{0})>0 from the discussion in step 1. Then we should have y(ψ0)<0y(\psi_{0})<0. Therefore, there is a ψ7(2πψ0,ψ0)\psi_{7}\in(2\pi-\psi_{0},\psi_{0}) such that y(ψ7)=y(2πψ7)=0y(\psi_{7})=y(2\pi-\psi_{7})=0. We know that either ψ7(ψ1,π)\psi_{7}\in(\psi_{1},\pi) or 2πψ7(ψ1,π)2\pi-\psi_{7}\in(\psi_{1},\pi), which means that y(2πψ7)>0y(2\pi-\psi_{7})>0 or y(ψ7)>0y(\psi_{7})>0. This contradicts the definition of ψ7\psi_{7}. So we always have:

y(ψ)=y(2πψ)y(\psi)=y(2\pi-\psi) (26)

which is the result we want.

Step 3 The curve we constructed in Step 1 cannot be a solution curve to our system (10).

Using equation (10), we know that the curve should meet with the right wall with the contact angle ψ2\psi_{2}. Because of this, we need to consider the point ψ=ψ(x2)+2π=ψ2+2π\psi=\psi(x_{2})+2\pi=\psi_{2}+2\pi. We suppose that the curve meets the right wall at the point ψ=2π+ψ2\psi=2\pi+\psi_{2}. We want to show that this assumption is not true.

From equation (7), we know that ψ3:=2π+ψ(x2)=2πγπ2+θ2\psi_{3}:=2\pi+\psi(x_{2})=2\pi-\gamma-\frac{\pi}{2}+\theta_{2} is an angle between π\pi and 2π2\pi. More specifically, we have:

ψ3=2πγπ2+θ2<2π(γ+π2θ1)=2πψ1\psi_{3}=2\pi-\gamma-\frac{\pi}{2}+\theta_{2}<2\pi-(\gamma+\frac{\pi}{2}-\theta_{1})=2\pi-\psi_{1}

where we used the fact that θ1>θ2\theta_{1}>\theta_{2} and equation (6) and (7). This implies that ψ3(π,2πψ1)\psi_{3}\in(\pi,2\pi-\psi_{1}). We suppose that the corresponding point is (x3,u3)=(x(ψ3),y(ψ3))(x_{3},u_{3})=(x(\psi_{3}),y(\psi_{3})). See Figure 5. We want to prove that the curve we constructed does not meet the right wall when ψ(ψ1,ψ3)\psi\in(\psi_{1},\psi_{3}).

By definition:

ψ3π=2π+ψ(x2)π=π2γ+θ2>θ2\psi_{3}-\pi=2\pi+\psi(x_{2})-\pi=\frac{\pi}{2}-\gamma+\theta_{2}>\theta_{2}

We can show that for any angle ψ(π+θ2,ψ3)\psi\in(\pi+\theta_{2},\psi_{3}), (x(ψ),u(ψ))(x(\psi),u(\psi)) is on the left hand side of the right wall. Then using the property θ1>θ2\theta_{1}>\theta_{2}, we can show that:

ψ4:=2πψ3=ψ2=γ+π2θ2>γ+π2θ1=ψ1\psi_{4}:=2\pi-\psi_{3}=-\psi_{2}=\gamma+\frac{\pi}{2}-\theta_{2}>\gamma+\frac{\pi}{2}-\theta_{1}=\psi_{1}

which means that ψ4(ψ1,π)\psi_{4}\in(\psi_{1},\pi). In addition, we have u(ψ4)=u(ψ3)u(\psi_{4})=u(\psi_{3}) by using step 2. See Figure 5. We should have x(ψ4)>x(ψ3)x(\psi_{4})>x(\psi_{3}) to avoid self-intersection. So the point (x(ψ4),u(ψ4))(x(\psi_{4}),u(\psi_{4})) should be on the right hand side of the right wall. Therefore the curve intersects the wall at a point (x(ψ5),u(ψ5))(x(\psi_{5}),u(\psi_{5})) where ψ4<ψ5<ψ3\psi_{4}<\psi_{5}<\psi_{3} which contradicts to the assumption.

Therefore, by applying the contact angle condition (7), the curve does not meet the right wall when ψ1<ψ2πψ3\psi_{1}<\psi\leq 2\pi-\psi_{3}. However, when ψ6=2πψ1\psi_{6}=2\pi-\psi_{1}, we can use the Step 2 again to obtain y(ψ6)=y(ψ1)y(\psi_{6})=y(\psi_{1}). In addition, the curve needs to satisfy the condition x(ψ6)<x(ψ1)x(\psi_{6})<x(\psi_{1}) to avoid self-intersection. This implies that the point lies left to the left wall. So the curve meets the left wall and even passes across it at some angle between ψ1\psi_{1} and 2πψ12\pi-\psi_{1}, which is impossible. Hence, we conclude that y1<0y_{1}<0.

Step 4: um<0u_{m}<0.

From Step 1 to Step 3, we have y1<0y_{1}<0. Using the equation (19), we observe that ψ(x)\psi(x) decreases as y(x)<0y(x)<0. Suppose there is an angle 0<ψ8<ψ10<\psi_{8}<\psi_{1} such that y(ψ8)=0y(\psi_{8})=0. Applying the same discussion as for the y10y_{1}\geq 0, we find that this assumption leads to a contradiction. So at the maximal point for u where sinψ=0\sin\psi=0, we must have um=y(0)<0u_{m}=y(0)<0. Consequently, for every ψ(ψ2,ψ1)\psi\in(\psi_{2},\psi_{1}) we obtain y(ψ)<0y(\psi)<0 ensuring that the equation (21) remains well defined for every ψ(ψ2,ψ1)\psi\in(\psi_{2},\psi_{1}). Thus ψ\psi is a well-defined global parameter for the surface curve. This completes the proof of Case 1.

Case 2 π2ψ(x1)<π\frac{\pi}{2}\leq\psi(x_{1})<\pi. In this case, we also do the same discussion as above. Actually, it is even easier since we do not need to consider the extending problem when ψ=π2\psi=\frac{\pi}{2}.

In conclusion, we have shown that ψ\psi is a legal parameter and um<0u_{m}<0.

2.2 The existence of the solution function

From the previous subsection, we have the a priori discussion for the boundary curve, which gives us an idea about the shape of the steady state. In this subsection, we discuss the existence of the steady state.

Using Theorem 2.1, we now use ψ\psi to parameterize the curve (ψ\psi is given by equation (5)). So, the original equation system (3) is equivalent to the equation system (10) after shifting (x,y)(x,yP0g)(x,y)\rightarrow(x,y-\frac{P_{0}}{g}). And we focus on analyzing the system (10) in this subsection.

Using the assumption that ψ1>0\psi_{1}>0 and ψ2<0\psi_{2}<0, the curve reaches its maximum at a point (xm,um)=(x(ψm),y(ψm))(x_{m},u_{m})=(x(\psi_{m}),y(\psi_{m})). At this point, we must have ψ(θm)=0\psi(\theta_{m})=0. See figure 3. Using the existence and uniqueness theorem of ODEs, the solution of this system is uniquely determined once the initial conditions are specified. In other words, we construct the curve using the equation (11) with the initial conditions (12) and (13). Then we use these two new parameters xmx_{m} and umu_{m} to derive a formula for the total volume instead of the unknown parameter P0P_{0}.

Since the equation system (10) includes two boundary conditions, there is effectively only one independent variable. As a result, both the total volume 𝒱\mathcal{V} and P0P_{0} can be expressed as functions of this variable, and then we set 𝒱=V\mathcal{V}=V to determine this variable.

A key challenge in applying the boundary condition in (10) is determining the position of the original point O (the junction of the two walls) after the coordinate change. To be specific, the y-component of its coordinate remains unknown. See the figure 3. In this figure, we know that the y component of O must be P0g-\frac{P_{0}}{g} to make the system compatible with the shift (x,y)(x,yP0g)(x,y)\rightarrow(x,y-\frac{P_{0}}{g}). However, since P0P_{0} is still unknown at this point, we cannot use geometry relations to derive the coordinates of two contact points. Fortunately, we know that the curve hits the left wall and the right wall with contact angles ψ1\psi_{1} and ψ2\psi_{2}, which are fixed angles. So once the curve is constructed by the equation (10) with the chosen initial conditions, we can set ψ=ψ1\psi=\psi_{1} or ψ=ψ2\psi=\psi_{2} to derive their coordinates. This leads to the following theorem.

Theorem 2.2.

( Boundary Points) We suppose that the boundary curve contacts two walls at two contact points (x1,u1)(x_{1},u_{1}) and (x2,u2)(x_{2},u_{2}). We have the following computation:

u1=um2+2σg(1cosψ1)\displaystyle u_{1}=-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})} (27)
u2=um2+2g(1cosψ2)\displaystyle u_{2}=-\sqrt{u_{m}^{2}+\frac{2}{g}(1-\cos\psi_{2})} (28)
x1=xm0ψ1cosγgum2+2σg(1cosγ)𝑑γ\displaystyle x_{1}=x_{m}-\int_{0}^{\psi_{1}}\frac{\cos\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma (29)
x2=xm0ψ2cosγgum2+2σg(1cosγ)𝑑γ\displaystyle x_{2}=x_{m}-\int_{0}^{\psi_{2}}\frac{\cos\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma (30)
r(ψ):=|x(ψ)xm|=|0ψcosγgum2+2σg(1cosγ)𝑑γ|\displaystyle r(\psi):=|x(\psi)-x_{m}|=|\int_{0}^{\psi}\frac{\cos\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma| (31)
Proof.

Step 1 Proof of (27):

Using the equation (22), we have:

12gy2(ψ)12gu12=σ(cosψ1cosψ){}\frac{1}{2}gy^{2}(\psi)-\frac{1}{2}gu_{1}^{2}=\sigma(\cos\psi_{1}-\cos\psi) (32)

We know that u1=y(ψ1)u_{1}=y(\psi_{1}) and ψm=0\psi_{m}=0. So we set ψ=ψm=0\psi=\psi_{m}=0 in the equation (32) to get:

12gu1212gum2=σ(1cosψ1)\frac{1}{2}gu_{1}^{2}-\frac{1}{2}gu_{m}^{2}=\sigma(1-\cos\psi_{1})

which shows that:

12gu12=12gum2+σ(1cosψ1){}\frac{1}{2}gu_{1}^{2}=\frac{1}{2}gu_{m}^{2}+\sigma(1-\cos\psi_{1}) (33)

Using Theorem 2.1, we know that um<0u_{m}<0. Since (xm,um)(x_{m},u_{m}) is the maximum point of the y component of the curve, it follows that u1<0u_{1}<0 as well. Taking square root on both sides of the equation (33), we obtain:

u1=um2+2σg(1cosψ1)u_{1}=-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}

which is exactly (27). We can also use the similar computation to show equation (28)

Step 2 Proof of (29):

Using the first equation in the system (10), we have:

dxdψ=σcosψgy(ψ){}\frac{dx}{d\psi}=\sigma\frac{\cos\psi}{gy(\psi)} (34)

From equation (32) in the step 1 we can obtain:

y(ψ)=um2+2σg(1cosψ){}y(\psi)=-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)} (35)

We plug equation (35) back into the equation (34) to show that:

dxdψ=cosψgum2+2σg(1cosψ){}\frac{dx}{d\psi}=\frac{\cos\psi}{-g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}} (36)

Then we integrate both sides of the equation from 0 to ψ1\psi_{1} then we to derive the following formula:

x1xm=0ψ1cosγgum2+2σg(1cosγ)𝑑γx_{1}-x_{m}=-\int_{0}^{\psi_{1}}\frac{\cos\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma

which is exactly what we want. In addition, we can use the same computation to derive (30).

Finally, the proof of (31) is just another simple application of the equation (36).

We can then express the total volume in terms of umu_{m} and xmx_{m} to. Before we derive this formula, we first establish a relationship between xmx_{m} and umu_{m} using the geometry relation so that we only have one independent parameter.

Theorem 2.3.

There holds:

xm=r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2whenθ1π2\displaystyle x_{m}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}}\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ \theta_{1}\neq\frac{\pi}{2} (37)
xm=r1whenθ1=π2\displaystyle x_{m}=r_{1}\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ \theta_{1}=\frac{\pi}{2} (38)
P0=g(x2tanθ2u2)\displaystyle P_{0}=g(x_{2}\tan\theta_{2}-u_{2}) (39)

where:

r1\displaystyle r_{1} :=r(ψ1)=xmx1\displaystyle:=r(\psi_{1})=x_{m}-x_{1} (40)
r2\displaystyle r_{2} :=r(ψ2)=x2xm\displaystyle:=r(\psi_{2})=x_{2}-x_{m} (41)

Remark: Equations (37), (38), and (39) imply that u1u_{1}, u2u_{2}, r1r_{1}, and r2r_{2} are all functions of umu_{m}. Consequently, by applying this theorem, xmx_{m} can also be expressed as a function of umu_{m}.

Proof.

We can show that xm=r1x_{m}=r_{1} when θ1=π2\theta_{1}=\frac{\pi}{2} just by definition (31). So we only need to prove (37) and (39).

We suppose that the two walls intersect at the point OO. The x component of O is zero by our construction of the coordinate, while y component of the point, denoted as y0y_{0}, is still unknown. Since the point (x1,u1)(x_{1},u_{1}) lies on the left wall which is inclined at θ1\theta_{1}, we can derive the following result by using geometry relation:

y0=u1(x1)tanθ1{}y_{0}=u_{1}-(-x_{1})\tan\theta_{1} (42)

Also, since (x2,u2)(x_{2},u_{2}) lies on the right wall which inclined at angle θ2\theta_{2}, we have:

y0=u2x2tanθ2{}y_{0}=u_{2}-x_{2}\tan\theta_{2} (43)

We then combine (32) and (33) together to get:

u1(x1)tanθ1=u2x2tanθ2{}u_{1}-(-x_{1})\tan\theta_{1}=u_{2}-x_{2}\tan\theta_{2} (44)

Then we plug the definition of r1r_{1} and r2r_{2} (40) and (41) into the equation(44) to get:

u1(r1xm)tanθ1=u2(r2+xm)tanθ2u_{1}-(r_{1}-x_{m})\tan\theta_{1}=u_{2}-(r_{2}+x_{m})\tan\theta_{2}

which implies the following relation:

xm(tanθ1+tanθ2)=u2u1+r1tanθ1r2tanθ2x_{m}(\tan\theta_{1}+\tan\theta_{2})=u_{2}-u_{1}+r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}

which means that:

xm=r1tanθ1r2tanθ2(u1u2)(tanθ1+tanθ2)x_{m}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{(\tan\theta_{1}+\tan\theta_{2})}

This is exactly the equation (37).

Also, using (42), we know that the coordinate of the connecting point O is (0,y0)=(0,u2x2tanθ2)(0,y_{0})=(0,u_{2}-x_{2}\tan\theta_{2}). And we know that after the shift (x,y)(x,y+P0g)(x,y)\rightarrow(x,y+\frac{P_{0}}{g}), coordinates of this point become (0,0)(0,0). So we should have:

0=P0g+(u2x2tanθ2)0=\frac{P_{0}}{g}+(u_{2}-x_{2}\tan\theta_{2})

which implies that:

P0=g(x2tanθ2u2)P_{0}=g(x_{2}\tan\theta_{2}-u_{2})

This finishes the proof. ∎

Now we have the representation for the P0P_{0} and xmx_{m}. In order to continue to compute the area, we also care about the sign of xmx_{m}. We have the following lemma:

lemma 2.4.

When θ1=θ2\theta_{1}=\theta_{2}, xm=0x_{m}=0. When θ2<θ1<π2\theta_{2}<\theta_{1}<\frac{\pi}{2} and [[γ]]0[[\gamma]]\leq 0, we fix the value of θ2\theta_{2} and umu_{m}. Then xmx_{m} is an increasing function of θ1\theta_{1}.

Proof.

Using Theorem 2.3, we have the formula for xmx_{m}:

xm=r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2{}x_{m}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}} (45)

Case one θ1=θ2\theta_{1}=\theta_{2}:

We use f(θ1)f(\theta_{1}) to denote the numerator of the right hand side of equation (45):

f(θ1)=r1tanθ1r2tanθ2(u1u2){}f(\theta_{1})=r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2}) (46)

When θ1=θ2\theta_{1}=\theta_{2}, we know from (6) and (7) that ψ1=ψ2\psi_{1}=-\psi_{2}, which implies that cosψ1=cosψ2\cos\psi_{1}=\cos\psi_{2}. Substituting this into the four equations (The equation (27)-(30)) in theorem 2.2, we obtain:

u1=u2andr1=r2u_{1}=u_{2}\leavevmode\nobreak\ \leavevmode\nobreak\ and\leavevmode\nobreak\ \leavevmode\nobreak\ r_{1}=r_{2}

So we should have xm=0x_{m}=0 by using equation (37).

Case two θ1θ2\theta_{1}\neq\theta_{2}:

Then we want to compute the general case that θ1>θ2\theta_{1}>\theta_{2}. To prove this we need to compute dfdθ1\frac{df}{d\theta_{1}} where f is defined as in equation (46).

First by using equations (27)-(30) in Theorem 2.2 we can show that:

r1=xmx1=0ψ1cosψdψgum2+2σg(1cosψ){}r_{1}=x_{m}-x_{1}=\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}} (47)
r2=x2xm=ψ20cosψdψgum2+2σg(1cosψ){}r_{2}=x_{2}-x_{m}=\int_{\psi_{2}}^{0}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}} (48)

and:

u1u2=um2+2σg(1cosψ1)+um2+2σg(1cosψ2){}u_{1}-u_{2}=-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}+\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{2})} (49)

Then we plug (47), (48) and (49) into equation (46) to compute the first derivative of f with respect to θ1\theta_{1}:

df(θ1)dθ1=\displaystyle\frac{df(\theta_{1})}{d\theta_{1}}= 1cos2θ1r1+dr1dθ1tanθ1du1dθ1\displaystyle\frac{1}{\cos^{2}\theta_{1}}r_{1}+\frac{dr_{1}}{d\theta_{1}}\tan\theta_{1}-\frac{du_{1}}{d\theta_{1}}
=\displaystyle= 1cos2θ10ψ1cosψdψgum2+2σg(1cosψ)+tanθ1cosψ1gum2+2σg(1cosψ1)dψ1dθ1\displaystyle\frac{1}{\cos^{2}\theta_{1}}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}+\tan\theta_{1}\frac{\cos\psi_{1}}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}\frac{d\psi_{1}}{d\theta_{1}}
+sinψ1gum2+2σg(1cosψ1)dψ1dθ1\displaystyle+\frac{\sin\psi_{1}}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}\frac{d\psi_{1}}{d\theta_{1}} (50)

To continue to the next step, we need to introduce an inequality below :

0ψ1cosψdψgum2+2σg(1cosψ)0ψ1cosψdψgum2+2σg(1cosψ1){}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}\geq\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}} (51)

We use Lemma 2.5 which we are going to prove after the main part of the theorem to show this inequality.

We apply inequality (51) to the (50)\eqref{equ:2.2.21} and obtain:

dfdθ1\displaystyle\frac{df}{d\theta_{1}}\geq 1gcos2θ1um2+2σg(1cosψ1)0ψ1cosψdψ\displaystyle\frac{1}{g\cos^{2}\theta_{1}\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}\int_{0}^{\psi_{1}}\cos\psi d\psi
+tanθ1cosψ1gum2+2σg(1cosψ1)dψ1dθ1+sinψ1gum2+2σg(1cosψ1)dψ1dθ1\displaystyle+\tan\theta_{1}\frac{\cos\psi_{1}}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}\frac{d\psi_{1}}{d\theta_{1}}+\frac{\sin\psi_{1}}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}\frac{d\psi_{1}}{d\theta_{1}}

Also we know from (6) that ψ1=γ+π2θ1\psi_{1}=\gamma+\frac{\pi}{2}-\theta_{1} where γ\gamma is the contact angle. So we have dψ(x1)dθ1=1\frac{d\psi(x_{1})}{d\theta_{1}}=-1. We plug it back into the inequality above and obtain:

df(θ1)dθ1\displaystyle\frac{df(\theta_{1})}{d\theta_{1}}\geq sinψ(x1)gcos2θ1um2+2σg(1cosψ(x1))tanθ1cosψ(x1)gum2+2σg(1cosψ(x1))\displaystyle\frac{\sin\psi(x_{1})}{g\cos^{2}\theta_{1}\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}-\tan\theta_{1}\frac{\cos\psi(x_{1})}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}
sinψ(x1)gum2+2σg(1cosψ(x1))\displaystyle-\frac{\sin\psi(x_{1})}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}
=\displaystyle= 1gum2+2σg(1cosψ(x1))(sinψ(x1)cos2θ1tanθ1cosψ(x1)sinψ(x1))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}(\frac{\sin\psi(x_{1})}{\cos^{2}\theta_{1}}-\tan\theta_{1}\cos\psi(x_{1})-\sin\psi(x_{1}))
=\displaystyle= 1gum2+2σg(1cosψ(x1))(sinψ(x1)tan2θ1tanθ1cosψ(x1))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}(\sin\psi(x_{1})\tan^{2}\theta_{1}-\tan\theta_{1}\cos\psi(x_{1}))
=\displaystyle= 1gum2+2σg(1cosψ(x1))tanθ1(sinψ(x1)tanθ1cosψ(x1))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}\tan\theta_{1}(\sin\psi(x_{1})\tan\theta_{1}-\cos\psi(x_{1}))
=\displaystyle= 1gum2+2σg(1cosψ(x1))tanθ1(sin(π2θ1+γ)tanθ1cos(π2θ1+γ))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}\tan\theta_{1}(\sin(\frac{\pi}{2}-\theta_{1}+\gamma)\tan\theta_{1}-\cos(\frac{\pi}{2}-\theta_{1}+\gamma))
=\displaystyle= 1gum2+2σg(1cosψ(x1))tanθ1(cos(θ1γ)tanθ1sin(θ1γ))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}\tan\theta_{1}(\cos(\theta_{1}-\gamma)\tan\theta_{1}-\sin(\theta_{1}-\gamma))
=\displaystyle= 1gum2+2σg(1cosψ(x1))tanθ1cos(θ1γ)(tanθ1tan(θ1γ))\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}\tan\theta_{1}\cos(\theta_{1}-\gamma)(\tan\theta_{1}-\tan(\theta_{1}-\gamma)) (52)

Since [[γ]]0[[\gamma]]\leq 0, we can use the definition of γ\gamma (8) to show that:

sinγ=[[γ]]σ0\sin\gamma=-\frac{[[\gamma]]}{\sigma}\geq 0

which means that γ[0,π2)\gamma\in[0,\frac{\pi}{2}). In addition, we have the condition 0<θ1<π20<\theta_{1}<\frac{\pi}{2}. So we should have cos(θ1γ)>0\cos(\theta_{1}-\gamma)>0 and tanθ1tan(θ1γ)>0\tan\theta_{1}-\tan(\theta_{1}-\gamma)>0. We plug these results back into equation (52) and obtain:

df(θ1)dθ11gum2+2g(1cosψ(x1))tanθ1cos(θ1γ)(tanθ1tan(θ1γ))0\frac{df(\theta_{1})}{d\theta_{1}}\geq\frac{1}{g\sqrt{u_{m}^{2}+\frac{2}{g}(1-\cos\psi(x_{1}))}}\tan\theta_{1}\cos(\theta_{1}-\gamma)(\tan\theta_{1}-\tan(\theta_{1}-\gamma))\geq 0

which finishes the proof. ∎

Then it remains to prove the inequality (51). We have the following lemma:

lemma 2.5.
0ψ1cosψdψgum2+2σg(1cosψ)0ψ1cosψdψgum2+2σg(1cosψ(x1))\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}\geq\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}}
Proof.

First, if ψ1=ψ(θ1)(0,π2)\psi_{1}=\psi(\theta_{1})\in(0,\frac{\pi}{2}), the lemma is straightforward to prove. Since cosψ>0\cos\psi>0 for any ψ(0,ψ1)\psi\in(0,\psi_{1}) and cosψcosψ1\cos\psi\geq\cos\psi_{1} for any ψ(0,ψ1)\psi\in(0,\psi_{1}), we can show that:

cosψgum2+2σg(1cosψ)cosψgum2+2σg(1cosψ(x1)){}\frac{\cos\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}\geq\frac{\cos\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}} (53)

Then we integrate on both sides of the inequality (53) from 0 to π\pi and we have the inequality (51) proved.

If ψ(x1)(π2,π)\psi(x_{1})\in(\frac{\pi}{2},\pi), we split the integral into three parts:

0ψ1cosψdψgum2+2σg(1cosψ)=\displaystyle\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}= 0πψ1cosψdψgum2+2σg(1cosψ)\displaystyle\int_{0}^{\pi-\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}
+πψ1π2cosψdψgum2+2σg(1cosψ)+π2ψ1cosψdψgum2+2σg(1cosψ)\displaystyle+\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}

For the first term above, we use the fact that cosψ0cos(ψ1)\cos\psi\geq 0\geq\cos(\psi_{1}) for any ψ(0,πψ1)\psi\in(0,\pi-\psi_{1}) to obtain that:

0πψ1cosψdψgum2+2σg(1cosψ)0πψ1cosψdψgum2+2σg(1cosψ1){}\int_{0}^{\pi-\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}\geq\int_{0}^{\pi-\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}} (54)

For the second term and the third term we combine them together and using the fact cos(πψ)=cos(ψ)\cos(\pi-\psi)=-\cos(\psi) to obtain:

πψ1π2cosψdψgum2+2σg(1cosψ)+π2ψ1cosψdψgum2+2σg(1cosψ)\displaystyle\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}
=\displaystyle= πψ1π2cosψdψgum2+2σg(1cosψ1)+π2ψ1cosψdψgum2+2σg(1cosψ1)\displaystyle\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}
+πψ1π2cosψdψgum2+2σg(1cosψ)cosψdψgum2+2σg(1cosψ1)\displaystyle+\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}-\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}
+πψ1π2cosψdψgum2+2σg(1cosψ1)cosψdψgum2+2σg(1+cosψ)\displaystyle+\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}-\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1+\cos\psi)}}
=\displaystyle= πψ1π2cosψdψgum2+2σg(1cosψ1)+π2ψ1cosψdψgum2+2σg(1cosψ1)\displaystyle\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}} (55)
+πψ1π2cosψdψgum2+2σg(1cosψ)cosψdψgum2+2σg(1+cosψ)\displaystyle+\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}-\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1+\cos\psi)}} (56)

Here we use the fact that cosψ=cos(πψ)\cos\psi=\cos(\pi-\psi)The terms in (55) are what we are going to keep and we want to show (56) is greater than zero. Using the fact that:

cosψ>0whenψ(πψ1,π2)\cos\psi>0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ \psi\in(\pi-\psi_{1},\frac{\pi}{2})

we can show that:

gum2+2σg(1cosψ)<gum2+2σg(1+cosψ)g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}<g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1+\cos\psi)}

which implies that:

πψ1π2cosψdψgum2+2σg(1cosψ)cosψdψgum2+2σg(1+cosψ)>0\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}-\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1+\cos\psi)}}>0

Therefore, we obtain the following result:

πψ1π2cosψdψgum2+2σg(1cosψ)+π2ψ1cosψdψgum2+2σg(1cosψ)\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}
πψ1π2cosψdψgum2+2σg(1cosψ1)+π2ψ1cosψdψgum2+2σg(1cosψ1){}\geq\int_{\pi-\psi_{1}}^{\frac{\pi}{2}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}+\int_{\frac{\pi}{2}}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}} (57)

We combine the inequality (57) with the inequality (54) and we can obtain the final result. ∎

Remark: Based on Theorem 2.4 we know that xm>0x_{m}>0 when θ2<θ1<π2\theta_{2}<\theta_{1}<\frac{\pi}{2} and [[γ]]<0[[\gamma]]<0. In addition, when θ1>π2\theta_{1}>\frac{\pi}{2}, xmx_{m} is obviously positive by definition (Equation (13)). So we always have xm>0x_{m}>0 when θ1>θ2\theta_{1}>\theta_{2} and [[γ]]<0[[\gamma]]<0

Also, Theorem 2.3 and Theorem 2.4 have already guaranteed that we can use umu_{m} to represent P0P_{0}. Then we only need to compute the total volume 𝒱(um)\mathcal{V}(u_{m}). We split the total volume into five parts. The first part is the region enclosed by y=u1y=u_{1}, x=xmx=x_{m}, and the surface curve (x(ψ),y(ψ))(x(\psi),y(\psi)) where ψ(0,ψ1)\psi\in(0,\psi_{1}). And the second part is the region enclosed by y=u2y=u_{2}, x=xmx=x_{m} and the surface curve (x(ψ),y(ψ))(x(\psi),y(\psi)) where ψ(ψ2,0)\psi\in(\psi_{2},0). The third part is the triangular area enclosed by y=u1y=u_{1}, x=mx=m and the left wall. The fourth part is the triangular area enclosed by y=u2y=u_{2}, x=0x=0 and the right wall. The fifth part is the square area enclosed by y=u1y=u_{1}, x=0x=0,y=u2y=u_{2} and x=xmx=x_{m}. We use Ω1\Omega_{1} to Ω5\Omega_{5} to denote them and also use 𝒱1\mathcal{V}_{1} to 𝒱5\mathcal{V}_{5} to represent the volume. See Figure 6.

Refer to caption
Figure 6:

In addition, there might be some difference between the case θ1=π2\theta_{1}=\frac{\pi}{2} and θ1π2\theta_{1}\neq\frac{\pi}{2} from Theorem 2.3. Therefore, we split our computation into two parts.

2.2.1 θ1π2\theta_{1}\neq\frac{\pi}{2}

We assume that θ1π2\theta_{1}\neq\frac{\pi}{2}. Then we can compute the volume as follows:

𝒱1\displaystyle\mathcal{V}_{1} =u1umr(v)𝑑y=0ψ1r(ψ)dydψ𝑑ψ=0ψ1r(ψ)sinψgy𝑑ψ\displaystyle=\int_{u_{1}}^{u_{m}}r(v)dy=\int_{0}^{\psi_{1}}r(\psi)\frac{dy}{d\psi}d\psi=\int_{0}^{\psi_{1}}r(\psi)\frac{\sin\psi}{gy}d\psi
=0ψ10ψcosγdγgum2+2σg(1cosγ)sinψgum2+2σg(1cosψ)𝑑ψ\displaystyle=\int_{0}^{\psi_{1}}\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}\frac{\sin\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi (58)

Here r(ψ)r(\psi) is defined by equation (31).

Using the same computation, we have the following relations:

𝒱2=ψ200ψcosγdγgum2+2σg(1cosγ)sinψgum2+2σg(1cosψ)𝑑ψ{}\mathcal{V}_{2}=\int_{\psi_{2}}^{0}\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}\frac{\sin\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi (59)

Also, we can know from the geometry that:

𝒱3\displaystyle\mathcal{V}_{3} =12(r1xm)2tanθ1\displaystyle=\frac{1}{2}(r_{1}-x_{m})^{2}\tan\theta_{1} (60)
𝒱4\displaystyle\mathcal{V}_{4} =12(r2+xm)2tanθ2\displaystyle=\frac{1}{2}(r_{2}+x_{m})^{2}\tan\theta_{2} (61)
𝒱5\displaystyle\mathcal{V}_{5} =xm(u1u2)\displaystyle=x_{m}(u_{1}-u_{2}) (62)

Then we want to find the relation between 𝒱=𝒱1+𝒱2+𝒱3+𝒱4+𝒱5\mathcal{V}=\mathcal{V}_{1}+\mathcal{V}_{2}+\mathcal{V}_{3}+\mathcal{V}_{4}+\mathcal{V}_{5} and umu_{m}.

Remark: Although Figure 6 only shows the case where θ1(0,π2)\theta_{1}\in(0,\frac{\pi}{2}), the representation of the total volume 𝒱\mathcal{V} has the same form when θ1(π2,π)\theta_{1}\in(\frac{\pi}{2},\pi).

Theorem 2.6.

The two volume functional 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} are both increasing functions of umu_{m}. And when um0u_{m}\rightarrow 0, 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} converge to ++\infty. When umu_{m}\rightarrow-\infty, 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} converge to 0.

Proof.

Since 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} have similar forms. We only need to prove this theorem for 𝒱1\mathcal{V}_{1}.

Based on a simple observation of equation(58), limum𝒱1=0\lim_{u_{m\rightarrow-\infty}}\mathcal{V}_{1}=0 since the denominator of the term inside the integral converges to \infty.

We then show the monotonicity of the volume 𝒱1\mathcal{V}_{1}. Taking derivative with respect to umu_{m} on both sides of the equation (58), we obtain the following expression:

d𝒱1dum=\displaystyle\frac{d\mathcal{V}_{1}}{du_{m}}= um0ψ10ψcosτdτg(um2+2σg(1cosτ))32sinψgum2+2σg(1cosψ)𝑑ψ\displaystyle u_{m}\int_{0}^{\psi_{1}}\int_{0}^{\psi}\frac{\cos\tau d\tau}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau))^{\frac{3}{2}}}\frac{\sin\psi}{-g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi
+um0ψ10ψcosτdτg(um2+2σg(1cosτ))12sinψg(um2+2σg(1cosψ))32𝑑ψ\displaystyle+u_{m}\int_{0}^{\psi_{1}}\int_{0}^{\psi}\frac{\cos\tau d\tau}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau))^{\frac{1}{2}}}\frac{\sin\psi}{-g({u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)})^{\frac{3}{2}}}d\psi (63)

For the first term in the right hand side of equation (63) .we need to study the sign of the integral:

0ψcosτdτg(um2+2σg(1cosτ))32\int_{0}^{\psi}\frac{\cos\tau d\tau}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau))^{\frac{3}{2}}}

Then we define:

a(ψ)=0ψcosτdτg(um2+2σg(1cosτ))32a(\psi)=\int_{0}^{\psi}\frac{\cos\tau d\tau}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau))^{\frac{3}{2}}}

We introduce lemma 2.9 which we will show the the proof after the main proof of this theorem to derive that a(ψ)a(\psi) is positive. So we obtain that uma(ψ)u_{m}a(\psi) is negative. Also by using lemma 2.9 we can know that:

b(ψ):=0ψcosτdτg(um2+2σg(1cosτ))12b(\psi):=\int_{0}^{\psi}\frac{\cos\tau d\tau}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau))^{\frac{1}{2}}}

is also positive. So umb(ψ)u_{m}b(\psi) is a negative function.

Therefore, we can rewrite the equation (63) to:

d𝒱1dum=0ψ1uma(ψ)sinψgum2+2σg(1cosψ)𝑑ψ+0ψ1umb(ψ)sinψg(um2+2σg(1cosψ))32𝑑ψ\frac{d\mathcal{V}_{1}}{du_{m}}=\int_{0}^{\psi_{1}}u_{m}a(\psi)\frac{\sin\psi}{-g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi+\int_{0}^{\psi_{1}}u_{m}b(\psi)\frac{\sin\psi}{-g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}}d\psi

We have sinψ>0\sin\psi>0, uma(ψ)<0u_{m}a(\psi)<0 and umb(ψ)<0u_{m}b(\psi)<0 by using the lemma 2.9 and the fact that ψ(0,π)\psi\in(0,\pi). Thus, d𝒱1dum\frac{d\mathcal{V}_{1}}{du_{m}} is non-negative. Therefore 𝒱1\mathcal{V}_{1} is a monotone increasing function of umu_{m}.

Also, by using the formula for 𝒱1\mathcal{V}_{1} we should have:

𝒱1=0ψ10ψcosτdτgum2+2σg(1cosτ)sinψgum2+2σg(1cosψ)𝑑ψ\mathcal{V}_{1}=\int_{0}^{\psi_{1}}\int_{0}^{\psi}\frac{\cos\tau d\tau}{-g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\tau)}}\frac{\sin\psi}{-g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi

When um=0u_{m}=0, we then have:

𝒱1=0ψ10ψcosτdτg2σg(1cosτ)sinψg2σg(1cosψ)𝑑ψ\mathcal{V}_{1}=\int_{0}^{\psi_{1}}\int_{0}^{\psi}\frac{\cos\tau d\tau}{-g\sqrt{\frac{2\sigma}{g}(1-\cos\tau)}}\frac{\sin\psi}{-g\sqrt{\frac{2\sigma}{g}(1-\cos\psi)}}d\psi

We can use Taylor expansion of the function inside the integral near ψ=0\psi=0 and τ=0\tau=0 to get:

cosτg2σg(1cosτ)sinψg2σg(1cosψ)14g1sinγsinψsinψ14g1γ\frac{\cos\tau}{-g\sqrt{\frac{2\sigma}{g}(1-\cos\tau)}}\frac{\sin\psi}{-g\sqrt{\frac{2\sigma}{g}(1-\cos\psi)}}\sim\frac{1}{4g}\frac{1}{\sin\gamma}\frac{\sin\psi}{\sin\psi}\sim\frac{1}{4g}\frac{1}{\gamma}

Since we know that:

0ψ14g1γ𝑑γ=+\int_{0}^{\psi}\frac{1}{4g}\frac{1}{\gamma}d\gamma=+\infty

we can show that 𝒱1(um)=+\mathcal{V}_{1}(u_{m})=+\infty when um=0u_{m}=0. And then we conclude 𝒱1+\mathcal{V}_{1}\rightarrow+\infty when um0u_{m}\rightarrow 0 by using Fatou’s lemma and the fact that 𝒱1\mathcal{V}_{1} is a monotone increasing function of umu_{m}.

We can use the same discussion to prove that 𝒱2\mathcal{V}_{2} obeys the same rules. ∎

Remark: In reality, uma(ψ1)-u_{m}a(\psi_{1}) defined in the above theorem represents the derivative of r1r_{1}, implying that r1r_{1} is a monotone increasing function of umu_{m}. Similarly, for the same reason, r2r_{2} is a monotone increasing function of umu_{m}. And the proof in Theorem 2.6 shows that:

limum0ri=+{}\lim_{u_{m}\rightarrow 0}r_{i}=+\infty (64)

and

limumri=0{}\lim_{u_{m}\rightarrow-\infty}r_{i}=0 (65)

for i=1,2i=1,2.

Then we want to build some similar results for 𝒱3\mathcal{V}_{3}, 𝒱4\mathcal{V}_{4}, and 𝒱5\mathcal{V}_{5}. We have the following theorem.

Theorem 2.7.

When umu_{m}\rightarrow-\infty , we have 𝒱i0\mathcal{V}_{i}\rightarrow 0 for i=3,4,5i=3,4,5.

Proof.

We plug (37) into the equation (60) and get:

𝒱3=12(r1tanθ2+r2tanθ2+(u1u2)tanθ1+tanθ2)2tanθ1\mathcal{V}_{3}=\frac{1}{2}(\frac{r_{1}\tan\theta_{2}+r_{2}\tan\theta_{2}+(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{1} (66)

We know from equation (64) and equation (65) that r10r_{1}\rightarrow 0 and r20r_{2}\rightarrow 0 when umu_{m}\rightarrow-\infty. Also, from the fact that:

u1u2\displaystyle u_{1}-u_{2} =um2+2σg(1cosψ(x1))+um2+2σg(1cosψ(x2))\displaystyle=-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}+\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{2}))}
=cosψ(x1)cosψ(x2)um2+2σg(1cosψ(x1))+um2+2σg(1cosψ(x2))\displaystyle=\frac{\cos\psi(x_{1})-\cos\psi(x_{2})}{\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{1}))}+\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi(x_{2}))}}

From the equation above, we can observe that u1u20u_{1}-u_{2}\rightarrow 0 when umu_{m}\rightarrow-\infty. We then return to equation (66). When umu_{m}\rightarrow-\infty we have:

𝒱3=12(r1tanθ2+r2tanθ2+(u1u2)tanθ1+tanθ2)2tanθ10\mathcal{V}_{3}=\frac{1}{2}(\frac{r_{1}\tan\theta_{2}+r_{2}\tan\theta_{2}+(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{1}\rightarrow 0

We can also plug (37) into (61) and obtain that:

𝒱4=12(r1tanθ1+r2tanθ2(u1u2)tanθ1+tanθ2)2tanθ2\mathcal{V}_{4}=\frac{1}{2}(\frac{r_{1}\tan\theta_{1}+r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{2}

Using the same reason, we have 𝒱40\mathcal{V}_{4}\rightarrow 0 when umu_{m}\rightarrow-\infty.

From above computation we can actually show that u1u20u_{1}-u_{2}\rightarrow 0 when umu_{m}\rightarrow-\infty. Combining this property with the relation (65), we can obtain:

x0=r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ20whenumx_{0}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{m}\rightarrow-\infty

Based on this result, we can show that:

limum𝒱5=limum(u1u2)x0=0\lim_{u_{m}\rightarrow-\infty}\mathcal{V}_{5}=\lim_{u_{m}\rightarrow-\infty}(u_{1}-u_{2})x_{0}=0

So we have the theorem proved. ∎

Theorem 2.8.

When θ1π2\theta_{1}\neq\frac{\pi}{2} For any fixed total volume V0V\geq 0 we can find at least one umu_{m} such that 𝒱(um)=𝒱1(um)+𝒱2(um)+𝒱3(um)+𝒱4(um)+𝒱5(um)=V\mathcal{V}(u_{m})=\mathcal{V}_{1}(u_{m})+\mathcal{V}_{2}(u_{m})+\mathcal{V}_{3}(u_{m})+\mathcal{V}_{4}(u_{m})+\mathcal{V}_{5}(u_{m})=V. Therefore, we have at least one solution function of the equation system (10) and thus a solution function of the equation system (3).

Proof.

When um0u_{m}\rightarrow 0, we know that 𝒱(um)𝒱1+𝒱2+\mathcal{V}(u_{m})\geq\mathcal{V}_{1}+\mathcal{V}_{2}\rightarrow+\infty using Theorem 2.6.

When umu_{m}\rightarrow-\infty, we know that 𝒱i(um)0,1i5\mathcal{V}_{i}(u_{m})\rightarrow 0,1\leq i\leq 5 from Theorem 2.6 and Theorem 2.7. So we have:

limum+𝒱(um)=limum+(𝒱1+𝒱2+𝒱3+𝒱4+𝒱5)=0\lim_{u_{m}\rightarrow+\infty}\mathcal{V}(u_{m})=\lim_{u_{m}\rightarrow+\infty}(\mathcal{V}_{1}+\mathcal{V}_{2}+\mathcal{V}_{3}+\mathcal{V}_{4}+\mathcal{V}_{5})=0

Since V is a continuous function of umu_{m}, we know that there is at least one um<0u_{m}<0 such that 𝒱(um)=V\mathcal{V}(u_{m})=V

By using this theorem we can get the existence of the solution function. However, for its uniqueness it is not easy to prove. Although we know that V1V_{1} and V2V_{2} are monotone increasing function of umu_{m}. Also, using the computation in the previous theorem and the previous subsection, we know that u1u2u_{1}-u_{2}, r1r_{1}, and r2r_{2} are increasing functions of vmv_{m}. However, the monotonicity of x0x_{0}, r1x0r_{1}-x_{0}, and r2+x0r_{2}+x_{0} is still unknown. We then need to figure out how to show the monotonicity.

Before proving the final theorem, we first need to prove a lemma which is used to derive the monotonicity of 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} in Theorem 2.6.

lemma 2.9.

We have the following inequalities

a(ψ)=0ψcosγdγ(um2+2σg(1cosγ))320ψcosγdγ(um2+2σg(1cosψ))320{}a(\psi)=\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma))^{\frac{3}{2}}}\geq\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}}\geq 0 (67)

for any ψ(0,ψ1)\psi\in(0,\psi_{1}). And

b(ψ)=0ψcosγdγ(um2+2σg(1cosγ))120ψcosγdγ(um2+2σg(1cosψ))120b(\psi)=\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma))^{\frac{1}{2}}}\geq\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{1}{2}}}\geq 0
Proof.

To prove that the integral is non-negative we only need to use the fact that:

0ψcosγdγ(um2+2σg(1cosψ))32=sinψ(um2+2σg(1cosψ))320\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}}=\frac{\sin\psi}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}}\geq 0

we have the last inequality because of the fact that ψ(0,π)\psi\in(0,\pi). Then we only need to prove that:

0ψcosγdγ(um2+2σg(1cosγ))320ψcosγdγ(um2+2σg(1cosψ))32{}\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\gamma))^{\frac{3}{2}}}\geq\int_{0}^{\psi}\frac{\cos\gamma d\gamma}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}} (68)

We notice that the inequality (68) has a similar form as the inequality (51) which is proved in Lemma 2.5. In fact, we can use the same idea to split the integral into three parts and perform a similar computation as we did in Lemma 2.5. Then we can get the result we want. The proof of function be is the same.

Theorem 2.10.

When [[γ]]<0[[\gamma]]<0 and θ1θ2\theta_{1}\geq\theta_{2},𝒱c=𝒱3+𝒱4+𝒱5\mathcal{V}_{c}=\mathcal{V}_{3}+\mathcal{V}_{4}+\mathcal{V}_{5} is a monotone increasing function of umu_{m}

Proof.

By using the equation (60),(61) and (62) We can get:

𝒱c=𝒱3+𝒱4+𝒱5=12(r1xm)2tanθ1+12(r2+xm)2tanθ2+xm(u1u2)\mathcal{V}_{c}=\mathcal{V}_{3}+\mathcal{V}_{4}+\mathcal{V}_{5}=\frac{1}{2}(r_{1}-x_{m})^{2}\tan\theta_{1}+\frac{1}{2}(r_{2}+x_{m})^{2}\tan\theta_{2}+x_{m}(u_{1}-u_{2})

We then plug equation (37) into the equation above and obtain:

𝒱c=\displaystyle\mathcal{V}_{c}= 12(r1r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2)2tanθ1\displaystyle\frac{1}{2}(r_{1}-\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{1}
+12(r2+r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2)2tanθ2\displaystyle+\frac{1}{2}(r_{2}+\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{2}
+(r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2)(u1u2)\displaystyle+(\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})(u_{1}-u_{2})
=\displaystyle= 12(r1tanθ2+r2tanθ2+(u1u2))2(tanθ1+tanθ2)2tanθ1+12((r2tanθ1+r1tanθ1(u1u2))2(tanθ1+tanθ2)2)tanθ2\displaystyle\frac{1}{2}\frac{(r_{1}\tan\theta_{2}+r_{2}\tan\theta_{2}+(u_{1}-u_{2}))^{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}\tan\theta_{1}+\frac{1}{2}(\frac{(r_{2}\tan\theta_{1}+r_{1}\tan\theta_{1}-(u_{1}-u_{2}))^{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}})\tan\theta_{2}
+(r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2)(u1u2)\displaystyle+(\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})(u_{1}-u_{2}) (69)

Using a rearrangement of these terms in the right hand side we have:

(69)=\displaystyle\eqref{equ:2.2.113}= 12(r1tanθ2+r2tanθ2)2(tanθ1+tanθ2)2tanθ1+12(r1tanθ1+r2tanθ1)2(tanθ1+tanθ2)2tanθ2\displaystyle\frac{1}{2}\frac{(r_{1}\tan\theta_{2}+r_{2}\tan\theta_{2})^{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}\tan\theta_{1}+\frac{1}{2}\frac{(r_{1}\tan\theta_{1}+r_{2}\tan\theta_{1})^{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}\tan\theta_{2}
+(u1u2)(r1+r2)tanθ1tanθ2(tanθ1+tanθ2)2(u1u2)(r1+r2)tanθ1tanθ2(tanθ1+tanθ2)2\displaystyle+(u_{1}-u_{2})(r_{1}+r_{2})\frac{\tan\theta_{1}\tan\theta_{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}-(u_{1}-u_{2})(r_{1}+r_{2})\frac{\tan\theta_{1}\tan\theta_{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}
+12(u1u2)2tanθ1(tanθ1+tanθ2)2+12(u1u2)2tanθ2(tanθ1+tanθ2)2\displaystyle+\frac{1}{2}\frac{(u_{1}-u_{2})^{2}\tan\theta_{1}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}+\frac{1}{2}\frac{(u_{1}-u_{2})^{2}\tan\theta_{2}}{(\tan\theta_{1}+\tan\theta_{2})^{2}}
+(r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2)(u1u2)\displaystyle+(\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})(u_{1}-u_{2})
=\displaystyle= 12(r1+r2)2tanθ1+tanθ2tanθ1tanθ2+12(u1u2)2(tanθ1+tanθ2)\displaystyle\frac{1}{2}\frac{(r_{1}+r_{2})^{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}+\frac{1}{2}\frac{(u_{1}-u_{2})^{2}}{(\tan\theta_{1}+\tan\theta_{2})}
+(u1u2)(r1tanθ1r2tanθ2tanθ1+tanθ2)(u1u2)2tanθ1+tanθ2\displaystyle+(u_{1}-u_{2})(\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}})-\frac{(u_{1}-u_{2})^{2}}{\tan\theta_{1}+\tan\theta_{2}}
=\displaystyle= 12(r1+r2)2tanθ1+tanθ2tanθ1tanθ2+(u1u2)(r1tanθ1r2tanθ2tanθ1+tanθ2)12(u1u2)2tanθ1+tanθ2\displaystyle\frac{1}{2}\frac{(r_{1}+r_{2})^{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}+(u_{1}-u_{2})(\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}})-\frac{1}{2}\frac{(u_{1}-u_{2})^{2}}{\tan\theta_{1}+\tan\theta_{2}} (70)

To move on to the next strep, we need introduce following result which we will prove after we prove the main part of this theorem:

1tanθ1+tanθ2(r1tanθ1(u1u2)){}\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-(u_{1}-u_{2})) (71)

is a positive monotone increasing function of umu_{m}. Using this result, we then transform equation (70) into:

(70)=\displaystyle\eqref{equ:2.2.38}= 12r12+r22tanθ1+tanθ2tanθ1tanθ2+r1r2tanθ1+tanθ2tanθ1tanθ2(u1u2)r2tanθ2tanθ1+tanθ2\displaystyle\frac{1}{2}\frac{r_{1}^{2}+r_{2}^{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}+\frac{r_{1}r_{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}-(u_{1}-u_{2})\frac{r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}
+(u1u2)r1tanθ1tanθ1+tanθ212(u1u2)2tanθ1+tanθ2\displaystyle+(u_{1}-u_{2})\frac{r_{1}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-\frac{1}{2}\frac{(u_{1}-u_{2})^{2}}{\tan\theta_{1}+\tan\theta_{2}}
=\displaystyle= 12r12+r22tanθ1+tanθ2tanθ1tanθ2+r2tanθ2tanθ1+tanθ2(r1tanθ1(u1u2))\displaystyle\frac{1}{2}\frac{r_{1}^{2}+r_{2}^{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}+\frac{r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-(u_{1}-u_{2}))
+u1u2tanθ1+tanθ2(r1tanθ112(u1u2))\displaystyle+\frac{u_{1}-u_{2}}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-\frac{1}{2}(u_{1}-u_{2})) (72)

For the first term in (72), it is monotonically increasing since r12r_{1}^{2} and r22r_{2}^{2} are monotonically increasing function of umu_{m}(from (64) and (65)). Also, we will show that tanθ1tanθ2tanθ1+tanθ2>0\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}>0 in Lemma 2.11. So,

12r12+r22tanθ1+tanθ2tanθ1tanθ2\frac{1}{2}\frac{r_{1}^{2}+r_{2}^{2}}{\tan\theta_{1}+\tan\theta_{2}}\tan\theta_{1}\tan\theta_{2}

is a monotonically increasing function of umu_{m}.

For the second term in (72), we already know that r2r_{2} is increasing monotonically by using the remark after Theorem 2.6. In addition, 1tanθ1+tanθ1(r1tanθ1(u1u2))\frac{1}{\tan\theta_{1}+\tan\theta_{1}}(r_{1}\tan\theta_{1}-(u_{1}-u_{2})) is a positive monotone increasing function using Lemma 2.11. So:

r2tanθ2tanθ1+tanθ2(r1tanθ1(u1u2))\frac{r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-(u_{1}-u_{2}))

is a monotone increasing function of umu_{m} since a positive increasing function times another positive increasing function is still monotone increasing.

Finally, for the third term, it is monotone increasing since u1u2u_{1}-u_{2} is positive monotone increasing and

1tanθ1+tanθ2(r1tanθ112(u1u2))\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-\frac{1}{2}(u_{1}-u_{2}))

is also positive and increasing by using Lemma 2.11. So the third term is increasing.

In conclusion, we should have (72) is an increasing function of umu_{m}. Thus 𝒱c\mathcal{V}_{c} is an increasing function of umu_{m}. ∎

Then we only need to show the lemma.

lemma 2.11.

We assume that θ1>θ2\theta_{1}>\theta_{2} and [[γ]]<0[[\gamma]]<0. Then the functions g1g_{1} and g2g_{2} below

g1(um)=1tanθ1+tanθ2(r1tanθ1(u1u2)){}g_{1}(u_{m})=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-(u_{1}-u_{2})) (73)
g2(um)=1tanθ1+tanθ2(r1tanθ112(u1u2)){}g_{2}(u_{m})=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}\tan\theta_{1}-\frac{1}{2}(u_{1}-u_{2})) (74)

and:

u1u2{}u_{1}-u_{2} (75)

are all positive monotone increasing functions of umu_{m}.

Proof.

We divide our problem into two cases: The first case where θ1(0,π2)\theta_{1}\in(0,\frac{\pi}{2}), the second case where θ1(π2,π)\theta_{1}\in(\frac{\pi}{2},\pi).

Case 1 θ1(0,π2)\theta_{1}\in(0,\frac{\pi}{2}):

Based on our assumption that θ1θ2\theta_{1}\geq\theta_{2}, we should have tanθ1>0\tan\theta_{1}>0 and tanθ2>0\tan\theta_{2}>0. So tanθ1tanθ1+tanθ2>0\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}>0. We split our proof into three steps.

Step 1: (75) is positive and monotonically increasing.

We compute the derivative of u1u2u_{1}-u_{2} which is defined as in the (27) and (28):

d(u1u2)dum=\displaystyle\frac{d(u_{1}-u_{2})}{du_{m}}= d(um2+2σg(1cosψ1)+um2+2σg(1cosψ2))dum\displaystyle\frac{d(-\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}+\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{2})})}{du_{m}}
=\displaystyle= umum2+2σg(1cosψ2)umum2+2σg(1cosψ1)\displaystyle\frac{u_{m}}{\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{2})}}-\frac{u_{m}}{\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}}
=\displaystyle= um2(cosψ2cosψ1)(um2+2σg(1cosψ2)+um2+2σg(1cosψ1))×\displaystyle u_{m}\frac{2(\cos\psi_{2}-\cos\psi_{1})}{(\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{2})}+\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})})}\times
1gum2+2σg(1cosψ2)um2+2σg(1cosψ1)\displaystyle\frac{1}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{2})}\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1})}} (76)

Since θ1>θ2\theta_{1}>\theta_{2}, we can obtain:

cosψ2=cos(γ+π2θ2)=sin(θ2γ)<sin(θ1γ)=cos(γ+π2θ1)=cosψ1{}\cos\psi_{2}=\cos(\gamma+\frac{\pi}{2}-\theta_{2})=\sin(\theta_{2}-\gamma)<\sin(\theta_{1}-\gamma)=\cos(\gamma+\frac{\pi}{2}-\theta_{1})=\cos\psi_{1} (77)

where we use the definition of ψ1\psi_{1} and ψ2\psi_{2} (6) and (7). Therefore, we have:

um(cosψ2cosψ1)0{}u_{m}(\cos\psi_{2}-\cos\psi_{1})\geq 0 (78)

We can then plug (78) back into the equation (76) and we can obtain that:

(76)0\eqref{equ:2.2.46}\geq 0

which finishes the first step. In fact, this proof also works for Case 2.

Step 2: Computing the first derivative of g1g_{1} and g2g_{2} with respect to umu_{m}.

We take the derivative of function g1g_{1} with respect to umu_{m} and get:

dg1(um)dum\displaystyle\frac{dg_{1}(u_{m})}{du_{m}} =1tanθ1+tanθ2d(r1tanθ1(u1u2))dum\displaystyle=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}\frac{d(r_{1}\tan\theta_{1}-(u_{1}-u_{2}))}{du_{m}}
=tanθ1tanθ1+tanθ2ddumr11tanθ1+tanθ2d(u1u2)dum\displaystyle=\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}\frac{d}{du_{m}}r_{1}-\frac{1}{\tan\theta_{1}+\tan\theta_{2}}\frac{d(u_{1}-u_{2})}{du_{m}} (79)

We first use the equation (31) to compute the first term in (79):

dr1tanθ1dum\displaystyle\frac{dr_{1}\tan\theta_{1}}{du_{m}} =ddumtanθ10ψ1cosψdψgum2+2σg(1cosψ)\displaystyle=\frac{d}{du_{m}}\tan\theta_{1}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g\sqrt{u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}
=umtanθ10ψ1cosψdψg(um2+2σg(1cosψ))32\displaystyle=-u_{m}\tan\theta_{1}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi))^{\frac{3}{2}}}} (80)

Noticing that um<0u_{m}<0 and the integral in equation (79) is positive by Lemma 2.9, this derivative is a positive number. Then we use Lemma 2.9 again to show:

dr1tanθ1dumumtanθ10ψ1cosψdψg(um2+2σg(1cosψ1))32{}\frac{dr_{1}\tan\theta_{1}}{du_{m}}\geq-u_{m}\tan\theta_{1}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}}} (81)

For the second term in (79), we use the relation (77) again in the equation (76) and then obtain the following result:

d(u1u2)dumum2gcosψ2cosψ12(um2+2σg(1cosψ1))32=umgcosψ2cosψ1(um2+2σg(1cosψ1))32{}\frac{d(u_{1}-u_{2})}{du_{m}}\leq u_{m}\frac{2}{g}\frac{\cos\psi_{2}-\cos\psi_{1}}{2(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}}=\frac{u_{m}}{g}\frac{\cos\psi_{2}-\cos\psi_{1}}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}} (82)

We then combine equation (82) and (80)\eqref{equ:2.2.44} together to obtain:

dg(um)dum\displaystyle\frac{dg(u_{m})}{du_{m}} =1tanθ1+tanθ2d(r1tanθ1(u1u2))dum\displaystyle=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}\frac{d(r_{1}\tan\theta_{1}-(u_{1}-u_{2}))}{du_{m}}
1tanθ1+tanθ2(umtanθ10ψ1cosψdψg(um2+2σg(1cosψ1))32umgcosψ2cosψ1(um2+2σg(1cosψ1))32)\displaystyle\geq\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(-u_{m}\tan\theta_{1}\int_{0}^{\psi_{1}}\frac{\cos\psi d\psi}{g(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}}-\frac{u_{m}}{g}\frac{\cos\psi_{2}-\cos\psi_{1}}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}})
=1tanθ1+tanθ2(umtanθ1sinψ11g1(um2+2σg(1cosψ1))32umgcosψ2cosψ1(um2+2σg(1cosψ1))32)\displaystyle=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(-u_{m}\tan\theta_{1}\sin\psi_{1}\frac{1}{g}\frac{1}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}}-\frac{u_{m}}{g}\frac{\cos\psi_{2}-\cos\psi_{1}}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}})
=1tanθ1+tanθ2(umg1(um2+2σg(1cosψ1))32(tanθ1sinψ1(cosψ1cosψ2)))\displaystyle=\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(-\frac{u_{m}}{g}\frac{1}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}}(\tan\theta_{1}\sin\psi_{1}-(\cos\psi_{1}-\cos\psi_{2}))) (83)

Using the fact that um<0u_{m}<0, we have:

1tanθ1+tanθ2(umg1(um2+2σg(1cosψ1))32)>0\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(-\frac{u_{m}}{g}\frac{1}{(u_{m}^{2}+\frac{2\sigma}{g}(1-\cos\psi_{1}))^{\frac{3}{2}}})>0

Therefore, we only need to care about the sign of the term tanθ1sinψ1(cosψ1cosψ2)\tan\theta_{1}\sin\psi_{1}-(\cos\psi_{1}-\cos\psi_{2}) in (83). We call it k(γ)k(\gamma). So now we have the formula of the derivative and only need to discuss the sign of a term which only depends on the inclined angles and contact angles.

Step 3: k(γ)k(\gamma) is positive .

Using the definition of ψ1\psi_{1} and ψ2\psi_{2} (33) and (34), we have:

k(γ)\displaystyle k(\gamma) =tanθ1sinψ1(cosψ1cosψ2)\displaystyle=\tan\theta_{1}\sin\psi_{1}-(\cos\psi_{1}-\cos\psi_{2})
=tanθ1sin(π2θ1+γ)(cos(π2θ1+γ)cos(π2θ2+γ))\displaystyle=\tan\theta_{1}\sin(\frac{\pi}{2}-\theta_{1}+\gamma)-(\cos(\frac{\pi}{2}-\theta_{1}+\gamma)-\cos(\frac{\pi}{2}-\theta_{2}+\gamma))
=tanθ1cos(θ1γ)(sin(θ1γ)sin(θ2γ))\displaystyle=\tan\theta_{1}\cos(\theta_{1}-\gamma)-(\sin(\theta_{1}-\gamma)-\sin(\theta_{2}-\gamma))
tanθ1cos(θ1γ)(sin(θ1γ)+sin(γ))=p(γ)\displaystyle\geq\tan\theta_{1}\cos(\theta_{1}-\gamma)-(\sin(\theta_{1}-\gamma)+\sin(\gamma))=p(\gamma)

Noticing that when γ=0\gamma=0, we have:

p(0)=sinθ1sinθ1+sin0=0p(0)=\sin\theta_{1}-\sin\theta_{1}+\sin 0=0

Then we compute the derivative of function p(γ)p(\gamma) with respect to γ\gamma:

p(γ)\displaystyle p^{\prime}(\gamma) =tanθ1sin(θ1γ)+cos(θ1γ)cosγ\displaystyle=\tan\theta_{1}\sin(\theta_{1}-\gamma)+\cos(\theta_{1}-\gamma)-\cos\gamma
=tanθ1(sinθ1cosγcosθ1sinγ)+cosθ1cosγ+sinθ1sinγcosγ\displaystyle=\tan\theta_{1}(\sin\theta_{1}\cos\gamma-\cos\theta_{1}\sin\gamma)+\cos\theta_{1}\cos\gamma+\sin\theta_{1}\sin\gamma-\cos\gamma
=sin2θ1cosθ1cosγsinθ1sinγ+cosθ1cosγ+sinθ1sinγcosγ\displaystyle=\frac{\sin^{2}\theta_{1}}{\cos\theta_{1}}\cos\gamma-\sin\theta_{1}\sin\gamma+\cos\theta_{1}\cos\gamma+\sin\theta_{1}\sin\gamma-\cos\gamma
=sin2θ1cosθ1cosγ+cosθ1cosγcosγ\displaystyle=\frac{\sin^{2}\theta_{1}}{\cos\theta_{1}}\cos\gamma+\cos\theta_{1}\cos\gamma-\cos\gamma
=1cosθ1cosγcosγ0forγ(0,π2)\displaystyle=\frac{1}{\cos\theta_{1}}\cos\gamma-\cos\gamma\geq 0\leavevmode\nobreak\ for\leavevmode\nobreak\ \gamma\in(0,\frac{\pi}{2})

The last inequality is due to the fact that γ(0,π2)\gamma\in(0,\frac{\pi}{2}) when [[γ]]<0[[\gamma]]<0. So finally we get:

k(γ)p(γ)p(0)=0k(\gamma)\geq p(\gamma)\geq p(0)=0

which implies that (73) is increasing monotonically.

For (74), we know that:

g2(um)=g1(um)+121tanθ1+tanθ2(u1u2)g_{2}(u_{m})=g_{1}(u_{m})+\frac{1}{2}\frac{1}{\tan\theta_{1}+\tan\theta_{2}}(u_{1}-u_{2})

Since g1(um)g_{1}(u_{m}) and u1u2u_{1}-u_{2} are both increasing functions of umu_{m}, g2g_{2} must be an increasing function.

Finally, we only need to show the positivity. We use the following fact:

limumr1=limum(u1u2)=0\lim_{u_{m}\rightarrow-\infty}r_{1}=\lim_{u_{m}\rightarrow-\infty}(u_{1}-u_{2})=0

which implies that

limumg1(um)=limumg2(um)=0{}\lim_{u_{m}\rightarrow-\infty}g_{1}(u_{m})=\lim_{u_{m}\rightarrow-\infty}g_{2}(u_{m})=0 (84)

Then the positivity of g1g_{1} and g2g_{2} follows from (84) and the monotonicity of the function g1g_{1} and g2g_{2}.

Case 2 θ1(π2,πθ2)\theta_{1}\in(\frac{\pi}{2},\pi-\theta_{2}):

In this case, we observe that

tanθ1andtanθ1+tanθ2\tan\theta_{1}\leavevmode\nobreak\ \leavevmode\nobreak\ \operatorname{and}\leavevmode\nobreak\ \leavevmode\nobreak\ \tan\theta_{1}+\tan\theta_{2}

are both negative due to the assumption that π2<θ1<πθ2\frac{\pi}{2}<\theta_{1}<\pi-\theta_{2}. So tanθ1tanθ1+tanθ2\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}} is positive. However,

r1tanθ1(u1u2)r_{1}\tan\theta_{1}-(u_{1}-u_{2})

is negative since u1>u2u_{1}>u_{2}, r1>0r_{1}>0 and tanθ1<0\tan\theta_{1}<0. So g1(um)g_{1}(u_{m}) should be positive and so do g2(um)g_{2}(u_{m}).

Using the Remark after Theorem 2.6 and Step 1 in Case 1 we have r1r_{1} and u1u2u_{1}-u_{2} are monotone increasing function of umu_{m}. So

r1tanθ1(u1u2)andr1tanθ112(u1u2)r_{1}\tan\theta_{1}-(u_{1}-u_{2})\leavevmode\nobreak\ \operatorname{and}\leavevmode\nobreak\ r_{1}\tan\theta_{1}-\frac{1}{2}(u_{1}-u_{2})

are monotone decreasing. Therefore, g1(um)g_{1}(u_{m}) and g2(um)g_{2}(u_{m}) should be monotone increasing since 1tanθ1+tanθ2\frac{1}{\tan\theta_{1}+\tan\theta_{2}} is negative and we have the lemma proved. ∎

2.2.2 θ1=π2\theta_{1}=\frac{\pi}{2}

When θ1=π2\theta_{1}=\frac{\pi}{2}, we should have 𝒱=𝒱1+𝒱2+𝒱4+𝒱5\mathcal{V}=\mathcal{V}_{1}+\mathcal{V}_{2}+\mathcal{V}_{4}+\mathcal{V}_{5}. From Theorem 2.6, we already have the monotonicity of 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}. Then we have the following theorem.

Theorem 2.12.

When θ1=π2\theta_{1}=\frac{\pi}{2} and [[γ]]<0[[\gamma]]<0, we have 𝒱4\mathcal{V}_{4} and 𝒱5\mathcal{V}_{5} monotone increasing functions of umu_{m}.

Proof.

First we show the monotonicity for 𝒱4\mathcal{V}_{4}:

𝒱4=12(r2+xm)2tanθ2\mathcal{V}_{4}=\frac{1}{2}(r_{2}+x_{m})^{2}\tan\theta_{2}

By using Theorem 2.4 we know that xm=r1x_{m}=r_{1}. So:

𝒱4=12(r2+r1)2tanθ2\mathcal{V}_{4}=\frac{1}{2}(r_{2}+r_{1})^{2}\tan\theta_{2}

is monotone increasing since r1r_{1} and r2r_{2} are both positive and monotone increasing.

For 𝒱5\mathcal{V}_{5}, we can show that:

𝒱5=xm(u1u2)=r1(u1u2)\mathcal{V}_{5}=x_{m}(u_{1}-u_{2})=r_{1}(u_{1}-u_{2})

is monotone increasing since r1r_{1} and u1u2u_{1}-u_{2} are both positive and monotone increasing. So the theorem is proved

Then we can show the following theorem:

Theorem 2.13.

When θ1,θ2(0,π)\theta_{1},\theta_{2}\in(0,\pi) and θ1(0,πθ2)\theta_{1}\in(0,\pi-\theta_{2}) and [[γ]]0[[\gamma]]\leq 0, 𝒱(um)\mathcal{V}(u_{m}) is a monotonically increasing function of umu_{m}.

In conclusion, we have the following theorem:

Theorem 2.14.

For any given constant V, there exist a umu_{m} and a solution to the equation system (10) such that

𝒱(ρ(θ))=𝒱(um)=V\mathcal{V}(\rho(\theta))=\mathcal{V}(u_{m})=V

Furthermore, 𝒱(um)\mathcal{V}(u_{m}) is a monotone increasing function of umu_{m} when [[γ]]<0[[\gamma]]<0. Therefore, for any given V, there exists a umu_{m} such that 𝒱(um)=V\mathcal{V}(u_{m})=V. And when [[γ]]<0[[\gamma]]<0, the choice of umu_{m} and the solution function is unique.

When ψ1<0\psi_{1}<0 and ψ2>0\psi_{2}>0, the function curve reaches a minimum at a point (x0,u0)(x_{0},u_{0}). Then we use a similar computation as in the previous case; we can also prove the existence. However, the uniqueness is still unknown in this case.

We then have the following theorem:

Theorem 2.15.

The function we constructed in Theorem 2.14 is exactly the solution function of the system (3) after the shift (x,y)(x,y+P0g)(x,y)\rightarrow(x,y+\frac{P_{0}}{g})

Proof.

Using Theorem 2.14, we now have a solution (x(ψ),y(ψ))(x(\psi),y(\psi)) to (10) and umu_{m} for any given V. We then use equation (39) in Theorem 2.3 to obtain the value of P0P_{0}.

After the shift (x,y)(x,y+P0g)(x,y)\rightarrow(x,y+\frac{P_{0}}{g}), the curve (x(ψ),y(ψ)+P0g)(x(\psi),y(\psi)+\frac{P_{0}}{g}) exactly solves the first two equations (9) since they are the solution of the system (10). The only problem is whether it satisfies the boundary condition. In other words, we need to prove that after the shift (x,y)(x,y+P0g)(x,y)\rightarrow(x,y+\frac{P_{0}}{g}) , the two lines θ=θ2\theta=\theta_{2} and θ=πθ1\theta=\pi-\theta_{1} are the given two walls. Since the incline angles of the two walls are θ1\theta_{1} and θ2\theta_{2}, we only need to show that the original point of the polar coordinate coincides with the meeting points of the two walls after the shift. This statement is true since this is how we use umu_{m} to represent P0P_{0} in Theorem 2.3. So we have the theorem proved. ∎

3 The second case: ψ1ψ20\psi_{1}\cdot\psi_{2}\geq 0

In this section, we discuss the case where ψ1\psi_{1} and ψ2\psi_{2} have the same sign. We begin with discussing a special case before we derive the theorem regarding the shape of the surface curve. And then we can follow similar methods as in the first case to derive the result we want.

3.1 A special case where θ1=π2\theta_{1}=\frac{\pi}{2} and θ2=0\theta_{2}=0

We also want to first discuss the shape of the function. But it is more complicated than the discussion in Section 2. So we begin with discussing a special case:

Theorem 3.1.

When θ1=π2\theta_{1}=\frac{\pi}{2} and θ2=0\theta_{2}=0, we assume that [[γ]]>0[[\gamma]]>0 and the contact angle γ(π2,π4)\gamma\in(-\frac{\pi}{2},-\frac{\pi}{4}), we assume that we have a steady state in this case. Then the slope angle ψ\psi (defined by equation (5)) of the curve is neither a monotone increasing function nor a monotone decreasing function of x when the total volume V is large enough.

Proof.

We first consider the problem in Cartesian coordinates whose origin point is at the point where the two different walls connect with each other. Then we have the equation for the free surface (which is the first equation in (18)):

guσ=P0{}gu-\sigma\mathcal{H}=P_{0} (85)

where:

=θ(xu(x)1+|xu|2)\mathcal{H}=\partial_{\theta}(\frac{\partial_{x}u(x)}{\sqrt{1+|\partial_{x}u|^{2}}})

is the mean curvature of the curve.

When P0P_{0} is less than zero, we integrate both sides of the equation (85) from 0 to x2x_{2} to obtain the following inequality:

gV+σsin(ψ1)σsin(ψ2)=x2P0gV+\sigma\sin(\psi_{1})-\sigma\sin(\psi_{2})=x_{2}P_{0}

Since we assume V is big enough, we have the left-hand side of the equation is greater than zero. However, the term on the right-hand side of the equation is non-positive by using the assumption that P00P_{0}\leq 0. So P0P_{0} should be greater or equal to zero.

Since u(x2)=0u(x_{2})=0, we can derive the following relation from (85):

σx(sinψ)|x=x2=P0>0-\sigma\partial_{x}(\sin\psi)|_{x=x_{2}}=P_{0}>0

which implies that:

x(sinψ)|x=x2<0\partial_{x}(sin\psi)|_{x=x_{2}}<0

So if ψ\psi is a monotone function of xx, then it should be monotone decreasing. But this can not be true since:

sin(ψ)|x=0=sinψ1=sin(γ)<sin(π2γ)=sinψ2=sin(γ)|x=x2{}\sin(\psi)|_{x=0}=\sin\psi_{1}=\sin(\gamma)<\sin(-\frac{\pi}{2}-\gamma)=\sin\psi_{2}=\sin(\gamma)|_{x=x_{2}} (86)

where we used the definition of ψ1\psi_{1} and ψ2\psi_{2} ((6) and (7)). We also used the assumption γ<π4\gamma<-\frac{\pi}{4} to obtain the inequality sin(γ)<sin(π2γ)-sin(\gamma)<-sin(\frac{\pi}{2}-\gamma).From this, we know that ψ\psi can not be monotone decreasing. Therefore, when V is large enough, ψ\psi is not a global legal parameter for the boundary curve. ∎

Theorem 3.1 implies that the shape of the curve differs when [[γ]]>0[[\gamma]]>0 and [[γ]]<0[[\gamma]]<0. We now aim to derive a theorem for this special case, which may provide insight into the curve’s shape in the general case.

Theorem 3.2.

The shape of the function

When θ1=π2\theta_{1}=\frac{\pi}{2} and θ2=0\theta_{2}=0, we assume that the contact angle γ(π2,π4)\gamma\in(-\frac{\pi}{2},-\frac{\pi}{4}), and we also assume that the steady surface exists.

1: When gV+σsin(γ)σsin(π2γ)0gV+\sigma sin(\gamma)-\sigma sin(-\frac{\pi}{2}-\gamma)\leq 0, ψ\psi is a monotone increasing of x.

2:When gV+σsin(γ)σsin(π2γ)>0gV+\sigma sin(\gamma)-\sigma sin(-\frac{\pi}{2}-\gamma)>0. ψ\psi reaches to a maximal value ψ3\psi_{3} such that ψ3(max(γπ2,γ)\psi_{3}\in(max(\gamma-\frac{\pi}{2},-\gamma).

Proof.

proof of (1): We integrate both sides of the equation (85) from 0 to x2x_{2} and get

gV+σsin(π2γ)σsin(γ)=x2P0gV+\sigma\sin(\frac{\pi}{2}-\gamma)-\sigma\sin(\gamma)=x_{2}P_{0}

Using the condition gV+σsin(ϕ)σsin(π2ϕ)0gV+\sigma sin(\phi)-\sigma sin(\frac{\pi}{2}-\phi)\leq 0, we have P00P_{0}\leq 0. Then the equation (85) becomes:

σxsinψ=gv(x)P0>0foranyx\sigma\partial_{x}sin\psi=gv(x)-P_{0}>0\leavevmode\nobreak\ for\leavevmode\nobreak\ any\leavevmode\nobreak\ x

So we know that ψ\psi is an increasing function of x and it is a legal global parameter for the curve.

  

proof of (2): From theorem 3.1, we know P00P_{0}\geq 0 when gV+σsin(ϕ)σsin(π2ϕ)>0gV+\sigma sin(\phi)-\sigma sin(\frac{\pi}{2}-\phi)>0 then γ\gamma is not monotone decreasing or increasing and it can reach a maximal value at some point since it is continuous. So we only need to prove that this maximum point ψ3<0\psi_{3}<0.

Step 1 ψ30\psi_{3}\leq 0

We use a contradiction argument. We suppose that ψ3>0\psi_{3}>0. Then we want to first prove the following property:

σddxsin(ψ)|x=0=gv(x1)P0>0\sigma\frac{d}{dx}sin(\psi)|_{x=0}=gv(x_{1})-P_{0}>0

We use a contradiction argument and first assume that:

σddxsin(ψ)|x=0=gv(x1)P00{}\sigma\frac{d}{dx}sin(\psi)|_{x=0}=gv(x_{1})-P_{0}\leq 0 (87)

We then take derivative with respect to x on the left hand side of the equation (85) to obtain:

σd2dx2sin(ψ)|x=0=gxv|x=0=gtanψ|x=0=gtan(γ)<0{}\sigma\frac{d^{2}}{dx^{2}}sin(\psi)|_{x=0}=g\partial_{x}v|_{x=0}=g\tan\psi|_{x=0}=gtan(-\gamma)<0 (88)

which implies that:

σd2d2xsin(ψ)<0\sigma\frac{d^{2}}{d^{2}x}sin(\psi)<0

and:

sinψ(x)<0\sin\psi(x)<0

for x near zero. We choose x0x_{0} such that:

x0=min{x:d2dx2sinψ=0}x_{0}=\min\{x:\frac{d^{2}}{dx^{2}}\sin\psi=0\}

This is well-defined since ψ\psi is not a monotone decreasing function of x from the assumption. From the computation in equation (88) we should have tanψ=0\tan\psi=0 when x=x0x=x_{0}. So sinψ(x0)>sinψ(0)\sin\psi(x_{0})>\sin\psi(0). We combine this with the assumption (87) to show that ddx(sinψ)=0\frac{d}{dx}(\sin\psi)=0 at some x1(x,x0)x_{1}\in(x,x_{0}). However, by using the definition of x0x_{0}, we know that d2dx2(sinψ)<0\frac{d^{2}}{dx^{2}}(\sin\psi)<0 when 0<x<x00<x<x_{0}. So ddx(sinψ)<0\frac{d}{dx}(\sin\psi)<0 for any x(0,x0)x\in(0,x_{0}). This contradicts to the existence of x1x_{1}. So (87) is not true.

Therefore, we have the property that:

σddψsinψ|x=0=gv(0)P0>0{}\sigma\frac{d}{d\psi}\sin\psi|_{x=0}=gv(0)-P_{0}>0 (89)

Suppose that ψ3>0\psi_{3}>0 is a maximum value, we have:

σddψsinψ|x=x3=gv(x3)P0=0\sigma\frac{d}{d\psi}\sin\psi|_{x=x_{3}}=gv(x_{3})-P_{0}=0

We then use (88) to show:

σd2dψ2sin(ψ)|x=x3=gtanψ3>0{}\sigma\frac{d^{2}}{d\psi^{2}}sin(\psi)|_{x=x_{3}}=g\tan\psi_{3}>0 (90)

Since ψ3\psi_{3} is the maximum we should have:

σd2dψ2sin(ψ)|ψ=ψ3=gtanψ30\sigma\frac{d^{2}}{d\psi^{2}}sin(\psi)|_{\psi=\psi_{3}}=g\tan\psi_{3}\leq 0

which contradicts to (90). So we have ψ30\psi_{3}\leq 0.

Step 2: ψ30\psi_{3}\neq 0.

We use a contradiction argument again. We suppose that ψ3=0\psi_{3}=0 then we can know from the first and second equation of the system (9) we can know that:

dψdx=gvP0σcosψ{}\frac{d\psi}{dx}=\frac{gv-P_{0}}{\sigma\cos\psi} (91)
dvdψ=σsinψgvP0{}\frac{dv}{d\psi}=\frac{\sigma\sin\psi}{gv-P_{0}} (92)

Then we we integrate both sides of the equation (92) from ψ\psi to ψ3\psi_{3} to show:

v=P0±P022gσ(cosψC)g{}v=\frac{P_{0}\pm\sqrt{P_{0}^{2}-2g\sigma(cos\psi-C)}}{g} (93)

Where CC is just a constant such that u(ψ3)=P0gu(\psi_{3})=\frac{P_{0}}{g}.

We plug the equation (93) back to the ODE (91) and get:

dψdx=P022gσ(cosψC)σcosψ=F(ψ){}\frac{d\psi}{dx}=\frac{\sqrt{P_{0}^{2}-2g\sigma(cos\psi-C)}}{\sigma\cos\psi}=F(\psi) (94)

Taking derivative on function F with respect to ψ\psi near ψ3\psi_{3}, we can show that:

dF(ψ)dψ=sinψP022gσ(cosψC)σcos2ψ+dP022gσ(cosψC)dψ{}\frac{dF(\psi)}{d\psi}=-\frac{\sin\psi\sqrt{P_{0}^{2}-2g\sigma(cos\psi-C)}}{\sigma\cos^{2}\psi}+\frac{d\sqrt{P_{0}^{2}-2g\sigma(\cos\psi-C)}}{d\psi} (95)

For the first term of equation (95), it is a smooth function and converges to zero when ψ\psi converges ψ3=0\psi_{3}=0, which means that this function is bounded in a neighborhood of ψ3=0\psi_{3}=0. Then we look at the second term.

dP022gσ(cosψC)dψ=122gσsinψP022gσ(cosψC){}\frac{d\sqrt{P_{0}^{2}-2g\sigma(cos\psi-C)}}{d\psi}=-\frac{1}{2}\frac{2g\sigma\sin\psi}{\sqrt{P_{0}^{2}-2g\sigma(\cos\psi-C)}} (96)

Then we take Taylor expansion of the right hand side of equation (96) near ψ=0\psi=0 to show:

dP022gσ(cosψC)dψ\displaystyle\frac{d\sqrt{P_{0}^{2}-2g\sigma(\cos\psi-C)}}{d\psi} =122gσsinψP022gσ(cosψC)\displaystyle=-\frac{1}{2}\frac{2g\sigma\sin\psi}{\sqrt{P_{0}^{2}-2g\sigma(\cos\psi-C)}}
=gσψ+O(ψ2)2gσP022gσ+Ccosψ\displaystyle=-g\sigma\frac{\psi+O(\psi^{2})}{\sqrt{2g\sigma}\sqrt{\frac{P_{0}^{2}}{2g\sigma}+C-\cos\psi}} (97)

We know from the definition of constant C:

P022gσ+Ccosψ3=0\frac{P_{0}^{2}}{2g\sigma}+C-\cos\psi_{3}=0

So we can have the Taylor expansion of the denominator which appears in (97):

P022gσ+Ccosψ=1cosψ=22ψ+O(ψ2)\sqrt{\frac{P_{0}^{2}}{2g\sigma}+C-\cos\psi}=\sqrt{1-\cos\psi}=\frac{\sqrt{2}}{2}\psi+O(\psi^{2})

We plug it back into the derivative term (97) and get:

(97)\displaystyle\eqref{equ:3.1.12} =gσψ+O(ψ2)2gψP02+C2gψcosψ\displaystyle=-g\sigma\frac{\psi+O(\psi^{2})}{\sqrt{2g\psi}\sqrt{\frac{P_{0}^{2}+C}{2g\psi}-\cos\psi}}
=gσψ+O(ψ2)2gσ1cosψ\displaystyle=-g\sigma\frac{\psi+O(\psi^{2})}{\sqrt{2g\sigma}\sqrt{1-\cos\psi}}
=gσψ+O(ψ2)gσψ+O(ψ2)\displaystyle=-g\sigma\frac{\psi+O(\psi^{2})}{\sqrt{g\sigma}\psi+O(\psi^{2})}
=gσ(1+O(ψ))\displaystyle=-\sqrt{g\sigma}(1+O(\psi)) (98)

So for our ODE (94):

dψdu=P022gσ(cosψC)σcosψ=F(ψ)\frac{d\psi}{du}=\frac{\sqrt{P_{0}^{2}-2g\sigma(\cos\psi-C)}}{\sigma\cos\psi}=F(\psi)

We know from (95) and (98) that F(ψ)F(\psi) has bounded first derivative and thus is a Lipschitz function near the point ψ=ψ3=0\psi=\psi_{3}=0. So by using the uniqueness of ODE we have ψ=0\psi=0 is the unique solution of the ODE, which contradicts our assumption. ∎

3.2 General case when θ1π2\theta_{1}\leq\frac{\pi}{2}

After discussing this special case, we want to discuss the general case. We first need to assume that 0θ2θ1π20\leq\theta_{2}\leq\theta_{1}\leq\frac{\pi}{2}. From the previous subsection, we may infer that ψ\psi can serve as a valid global parameter only when VV is small. This makes the analysis more straightforward in such cases, as we can apply a similar approach to the first case. Therefore, our primary focus should be on the scenario where VV is large.

The idea:

From Section 3.1, we know that ψ\psi might not be a valid parameter for the boundary curve and attains its maximum value ψm\psi_{m} at a point (xm,vm)=(x(θm),v(θm))(x_{m},v_{m})=(x(\theta_{m}),v(\theta_{m})). After applying the shift (x,y)(x,yP0g)(x,y)\rightarrow(x,y-\frac{P_{0}}{g}). We should have um=vmP0g=0u_{m}=v_{m}-\frac{P_{0}}{g}=0. Our goal is to use ψm\psi_{m} to express the coordinates of two contact points. See Figure 7.

In addition, the curve seems to be a function of x which intrigues us to integrate both sides of the first equation in the system (18) from x1x_{1} to x2x_{2} and get the following:

P0=g𝒱(ρ(θ))+12gx12tanθ1+12gx22tanθ2σsin(ψ2)+σsin(ψ1)x2x1{}P_{0}=\frac{g{\mathcal{V}}(\rho(\theta))+\frac{1}{2}gx_{1}^{2}\tan\theta_{1}+\frac{1}{2}gx_{2}^{2}\tan\theta_{2}-\sigma\sin(\psi_{2})+\sigma\sin(\psi_{1})}{x_{2}-x_{1}} (99)

Where 𝒱(ρ(θ))\mathcal{V}(\rho(\theta)) is defined by the equation (2) and can be rewritten as the following equation in Cartesian coordinates:

𝒱(ρ(θ))=x1x2v(x)𝑑x12(v(x1))21tanθ112(v(x2))21tanθ2{}\mathcal{V}(\rho(\theta))=\int_{x_{1}}^{x_{2}}v(x)dx-\frac{1}{2}(v(x_{1}))^{2}\frac{1}{\tan\theta_{1}}-\frac{1}{2}(v(x_{2}))^{2}\frac{1}{\tan\theta_{2}} (100)

It is the total volume enclosed by the boundary curve and two walls. Then we only need to determine P0P_{0} to make equation (99)\eqref{equ:3.3.14} the relation between 𝒱\mathcal{V} and ψ0\psi_{0}. Our approach is to first show that x serves as a global variable for the curve and then try to use the specially chosen point to express P0P_{0}.

We begin with deriving a theorem concerning the shape of the curve. We assume that V is large enough such that ψ\psi is not a legal global parameter to parameterize the curve. Under this assumption, we establish the following lemma and theorems:

lemma 3.3.

When 0θ2θ1π20\leq\theta_{2}\leq\theta_{1}\leq\frac{\pi}{2} and the sign of ψ1\psi_{1} is the same as the sign of ψ2\psi_{2}, then we have ψ1<0\psi_{1}<0 and [[γ]]>0[[\gamma]]>0.

Proof.

First we need to prove that [[γ]]>0[[\gamma]]>0 in this case. By definition we have:

ψ1=π2θ1+γ\psi_{1}=\frac{\pi}{2}-\theta_{1}+\gamma

and:

ψ2=γπ2+θ2\psi_{2}=-\gamma-\frac{\pi}{2}+\theta_{2}

Then if [[γ]]0[[\gamma]]\leq 0 we should have γ(0,π2)\gamma\in(0,\frac{\pi}{2}) by using the equation (8). Since θ1(0,π2)\theta_{1}\in(0,\frac{\pi}{2}), we should have ψ1=π2θ1+γ0\psi_{1}=\frac{\pi}{2}-\theta_{1}+\gamma\geq 0 and ψ2=γπ2+θ2<0\psi_{2}=-\gamma-\frac{\pi}{2}+\theta_{2}<0. So ψ1\psi_{1} and ψ2\psi_{2} have different signs, which contradicts our assumption. So [[γ]]>0[[\gamma]]>0

To prove the theorem, we suppose that ψ1=π2θ1+γ>0\psi_{1}=\frac{\pi}{2}-\theta_{1}+\gamma>0. Then γ<π2θ1-\gamma<\frac{\pi}{2}-\theta_{1}. Since θ2θ1π2\theta_{2}\leq\theta_{1}\leq\frac{\pi}{2}, we should have γ<π2θ1π2θ2-\gamma<\frac{\pi}{2}-\theta_{1}\leq\frac{\pi}{2}-\theta_{2}, which means that ψ2=γπ2+θ2<0\psi_{2}=-\gamma-\frac{\pi}{2}+\theta_{2}<0. So ψ1\psi_{1} and ψ2\psi_{2} have different signs, which contradicts our assumption again.

So we should have ψ1\psi_{1} and ψ2\psi_{2} are both less than zero. ∎

By using Lemma 3.3 we can get a theorem about the shape of the curve:

Theorem 3.4.

We suppose that 0θ2θ1π20\leq\theta_{2}\leq\theta_{1}\leq\frac{\pi}{2} and that the boundary curve u is not a function of ψ\psi. We also assume that there exists a solution function for our equation system (10). Then ψ(θ)\psi(\theta) increases from ψ1=ψ(πθ1)\psi_{1}=\psi(\pi-\theta_{1}) to a negative maximum ψm=ψ(θm)\psi_{m}=\psi(\theta_{m}). And then decreases from the maximal point to ψ(θ2)\psi(\theta_{2}). Also, the curve is a graph of x.

Proof.

Since ψ1<0\psi_{1}<0, the curve is a function of x near θ=πθ1\theta=\pi-\theta_{1}. We can use the first equation in the system (18). It can be turned into the following form:

x(sinψ)=gu{}\partial_{x}(\sin\psi)=gu (101)

We first suppose that u10u_{1}\leq 0. Then sinψ\sin\psi is a decreasing function in a small neighborhood of (x1,u1)(x_{1},u_{1}). Since ψ1<0\psi_{1}<0, we can easily show ψ<ψ1<0\psi<\psi_{1}<0 when x(x1,x1+δ)x\in(x_{1},x_{1}+\delta). Then we can repeat the discussion for the point (x1+δ,ψ(x1+δ))(x_{1}+\delta,\psi(x_{1}+\delta)) and prove that ψ(x)<ψ1\psi(x)<\psi_{1} for any x(x,x+2δ)x\in(x,x+2\delta). Finally, we can prove that ψ(x)<ψ1\psi(x)<\psi_{1} for any x(x1,x3)x\in(x_{1},x_{3}) where x3x_{3} is a point such that ψ(x3)=π2\psi(x_{3})=-\frac{\pi}{2}. This part of the curve cannot touch the right wall due to the assumption that ψ\psi is not a global variable.

Similarly to Theorem 2.1, we can extend the solution function to ψ(π2,32π)\psi\in(-\frac{\pi}{2},-\frac{3}{2}\pi) and similarly to ψ(32π,2πψ1)\psi\in(-\frac{3}{2}\pi,-2\pi-\psi_{1}). Then using a similar discussion as in Theorem 2.1, we can know that the curve will self-intersect or pass through the left wall, leading to a contradiction.

Thus, we must have u1>0u_{1}>0 and ψ\psi increases to the maximal value ψm\psi_{m}. Since ψm\psi_{m} is the maximum value, we obtain um=0u_{m}=0. We also know from Theorem 3.2 that ψm<0\psi_{m}<0(Although we only proved Theorem 3.2 in the special case, the proof also works for the general case). Examining equation (101), we see that ψ\psi decreases when u(x)<0u(x)<0, implying that ψ\psi decreases monotonically to ψ2\psi_{2} as a function of x. This completes the proof. Moreover, we establish ψ(π2,0)\psi\in(-\frac{\pi}{2},0) for any θ\theta, confirming that the curve represents a graph of x.

Refer to caption
Figure 7:

So from Theorem 3.4, we can divide the surface curve into two segments: the first extending from x1x_{1} to xmx_{m} and the second from xmx_{m} to x2x_{2}. See the figure 7. And the system (18) provides a global description of the curve. And we use v(x)v(x) to represent the curve and use u(x)u(x) to represent the curve after shifting. By making the separation, we establish that each segment of the curve can be parameterized by ψ\psi, which is defined by (5), allowing us to use the equation (11) for each part.

Let (xm,um)(x_{m},u_{m}) be the point where the slope angle reaches its maximum, denoted by ψm\psi_{m}. Since it is the maximum value, we must have gum=0gu_{m}=0, implying that the coordinate of the special point is (xm,0)(x_{m},0) with maximum angle ψm\psi_{m}. The remainder of this chapter focuses on using ψm\psi_{m} to represent the unknown parameters P0P_{0}, x1x_{1}, and x2x_{2}. We first prove the following theorem in terms of the coordinates of the two contact points:

Theorem 3.5.

We have the following formula for the coordinates of two contact points:

u1\displaystyle u_{1} =2σg(cosψmcosψ1)\displaystyle=\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi_{1})} (102)
u2\displaystyle u_{2} =2σg(cosψmcosψ2)\displaystyle=-\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi_{2})} (103)
x1\displaystyle x_{1} =xmψ1ψmcosγg2σg(cosψmcosγ)𝑑γ\displaystyle=x_{m}-\int_{\psi_{1}}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma (104)
x2\displaystyle x_{2} =xmψmψ2cosγg2σg(cosψmcosγ)𝑑γ\displaystyle=x_{m}-\int_{\psi_{m}}^{\psi_{2}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma (105)
r(ψ)\displaystyle r(\psi) =|x(ψ)xm|=|xmψmψcosγg2σg(cosψmcosγ)𝑑γ|\displaystyle=|x(\psi)-x_{m}|=|x_{m}-\int_{\psi_{m}}^{\psi}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma| (106)
Proof.

Since each segment of the curve can be parameterized by ψ\psi, the computation is the same as Theorem 2.2. We just use the equation (10) and the corresponding boundary condition to do these computations. ∎

Unlike the previous section, we now separate the curve into two segments, each of which may exhibit different properties. We need to prove that the second part does not intersect the left wall, which implies that x2>0x_{2}>0.

lemma 3.6.

For any x(xm,x2)x\in(x_{m},x_{2}), the curve u(x)u(x) does not intersect the left wall.

Proof.

Using the construction, we know that the maximal point (xm,um)(x_{m},u_{m}) should be right to the left wall. So we only need to prove that ψ(x)θ1\psi(x)\geq-\theta_{1} when x(xm,x2)x\in(x_{m},x_{2}).

By Theorem 3.4 we know that ψ\psi is a monotonically decreasing function of x. Therefore, it suffices to prove that ψ2θ1\psi_{2}\geq-\theta_{1}. From the definition of ψ2\psi_{2}((7)), we have ψ2=γπ2+θ2\psi_{2}=\gamma-\frac{\pi}{2}+\theta_{2} and ψ1=π2θ1γ\psi_{1}=\frac{\pi}{2}-\theta_{1}-\gamma. Using Lemma 3.3, we know that ψ1<0\psi_{1}<0 and ψ2<0\psi_{2}<0, which allows us to show:

π2θ1<γ<π2θ2\frac{\pi}{2}-\theta_{1}<\gamma<\frac{\pi}{2}-\theta_{2}

which means that :

γπ2+θ2>π2θ1π2+θ2=θ2θ1θ1\gamma-\frac{\pi}{2}+\theta_{2}>\frac{\pi}{2}-\theta_{1}-\frac{\pi}{2}+\theta_{2}=\theta_{2}-\theta_{1}\geq-\theta_{1}

So we have the theorem proved. ∎

Now we have two parameters xmx_{m} and ψm\psi_{m}, we need to find the relation between these two parameters. We can also use the same idea as Section 2 to compute xmx_{m} and get:

xm=r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2whenθ1π2{}x_{m}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}}\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ \theta_{1}\neq\frac{\pi}{2} (107)
xm=r1whenθ1=π2{}x_{m}=r_{1}\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ \theta_{1}=\frac{\pi}{2} (108)

Here r1=xmx1r_{1}=x_{m}-x_{1} and r2=x2xmr_{2}=x_{2}-x_{m}. For the special case θ1=π2\theta_{1}=\frac{\pi}{2}, the computation is straightforward and follows a similar approach to Subsection 2.2.2. Therefore, we assume that θ1π2\theta_{1}\neq\frac{\pi}{2} in the following proof.

We can then compute P0P_{0} by determining the y component of the connect point O (the point where two walls meet each other). After the shifting (x,y)(x,yP0g)(x,y)\rightarrow(x,y-\frac{P_{0}}{g}), we know that the y component of the point O is y0=P0gy_{0}=-\frac{P_{0}}{g}. Also, using the same computation as in Theorem 2.3, we can show y0=((r2+xm)tanθ2u2)y_{0}=-((r_{2}+x_{m})\tan\theta_{2}-u_{2}). We can obtain P0=g((r2+xm)tanθ2u2)P_{0}=g((r_{2}+x_{m})\tan\theta_{2}-u_{2}). Then we use this expression substituting P0P_{0} in the equation (99) and obtain:

g𝒱(xm,ψm)=(x2x1)g((r2+xm)tanθ2u2)12gx12tanθ112gx22tanθ2+σsinψ2σsinψ1{}g{\mathcal{V}}(x_{m},\psi_{m})=(x_{2}-x_{1})g((r_{2}+x_{m})\tan\theta_{2}-u_{2})-\frac{1}{2}gx_{1}^{2}\tan\theta_{1}-\frac{1}{2}gx_{2}^{2}\tan\theta_{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1} (109)

Now we have the expression for the functional 𝒱\mathcal{V} and we want to simplify the representation (109). To achieve this goal, we plug equation (107) into this equation and then we obtain:

g𝒱(ψm)=\displaystyle g{\mathcal{V}}(\psi_{m})= (x2x1)g(r2tanθ1+r1tanθ1tanθ1+tanθ2(u1u2)tanθ2tanθ1+tanθ2u2)\displaystyle(x_{2}-x_{1})g(\frac{r_{2}\tan\theta_{1}+r_{1}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-(u_{1}-u_{2})\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2})
12gx12tanθ112gx22tanθ2+σsinψ2σsinψ1\displaystyle-\frac{1}{2}gx_{1}^{2}\tan\theta_{1}-\frac{1}{2}gx_{2}^{2}\tan\theta_{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1}
=\displaystyle= g((r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)\displaystyle g((r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})
12g(r1xm)2tanθ112g(r2+xm)2tanθ2+σsinψ2σsinψ1\displaystyle-\frac{1}{2}g(r_{1}-x_{m})^{2}\tan\theta_{1}-\frac{1}{2}g(r_{2}+x_{m})^{2}\tan\theta_{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1}
=\displaystyle= g((r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)\displaystyle g((r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})
12gtanθ1tanθ2tanθ1+tanθ2(r12+r22)12g1(tanθ1+tanθ2)(u1u2)2+σsinψ2σsinψ1\displaystyle-\frac{1}{2}g\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}^{2}+r_{2}^{2})-\frac{1}{2}g\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(u_{1}-u_{2})^{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1}
=\displaystyle= g(12(r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)\displaystyle g(\frac{1}{2}(r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})
+12g(r1+r2)2tanθ1tanθ2tanθ1+tanθ212gtanθ1tanθ2tanθ1+tanθ2(r12+r22)\displaystyle+\frac{1}{2}g(r_{1}+r_{2})^{2}\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-\frac{1}{2}g\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}(r_{1}^{2}+r_{2}^{2})
12g1(tanθ1+tanθ2)(u1u2)2+σsinψ2σsinψ1\displaystyle-\frac{1}{2}g\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(u_{1}-u_{2})^{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1}
=\displaystyle= g(12(r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)\displaystyle g(\frac{1}{2}(r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})
+gr1r2tanθ1tanθ2tanθ1+tanθ212g1(tanθ1+tanθ2)(u1u2)2+σsinψ2σsinψ1\displaystyle+gr_{1}r_{2}\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-\frac{1}{2}g\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(u_{1}-u_{2})^{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1} (110)

We call the right-hand side of (110) f(ψm)f(\psi_{m}). Then we have the following theorem:

Theorem 3.7.

If we view r1r_{1} and r2r_{2} as functions of ψm\psi_{m}, then the function f defined below converges to ++\infty when ψm0\psi_{m}\rightarrow 0.

f(ψm)=\displaystyle f(\psi_{m})= g(12(r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)\displaystyle g(\frac{1}{2}(r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})
+gr1r2tanθ1tanθ2tanθ1+tanθ212g1(tanθ1+tanθ2)(u1u2)2+σsinψ2σsinψ1\displaystyle+gr_{1}r_{2}\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-\frac{1}{2}g\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(u_{1}-u_{2})^{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1} (111)
Proof.

Step 1 r1r_{1} and r2r_{2} converge to ++\infty when ψm0\psi_{m}\rightarrow 0.

Using equation (104) we know that:

r1=ψ1ψmcosγg2σg(cosψmcosγ)𝑑γr_{1}=\int_{\psi_{1}}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma

When ψm=0\psi_{m}=0, we can use Taylor expansion and get:

cosγ=112γ2+O((γ)3)\cos\gamma=1-\frac{1}{2}\gamma^{2}+O((\gamma)^{3})

We apply this expansion to the integral above:

r1(0)ψ1012σgcosγγ𝑑γ=+{}r_{1}(0)\geq\int_{\psi_{1}}^{0}\frac{1}{\sqrt{2\sigma g}}\frac{\cos\gamma}{-\gamma}d\gamma=+\infty (112)

From a simple observation, we know that r1r_{1} is a increasing function of ψm\psi_{m} when ψm(ψ1,0)\psi_{m}\in(\psi_{1},0). Therefore, using Fatou’s lemma, we can show that:

=ψ10cosγg2σg(1cosγ)𝑑γlimψm0ψ1ψmcosγg2σg(cosψmcosγ)𝑑γ\infty=\int_{\psi_{1}}^{0}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma\leq\lim_{\psi_{m}\rightarrow 0}\int_{\psi_{1}}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma

which is equivalent to the fact that:

limψm0r1=+\lim_{\psi_{m}\rightarrow 0}r_{1}=+\infty

And we can use the same method to prove that r2+r_{2}\rightarrow+\infty when ψm0\psi_{m}\rightarrow 0, which completes the proof in step 1

Step 2 The behavior of u1u_{1} and u2u_{2} as functions of ψm\psi_{m}.

Using (102) and (103), we can obtain that u12σg(1cosψ1)u_{1}\rightarrow\sqrt{\frac{2\sigma}{g}(1-\cos\psi_{1})} when ψm0\psi_{m}\rightarrow 0. and u22σg(1cosψ2)u_{2}\rightarrow\sqrt{\frac{2\sigma}{g}(1-\cos\psi_{2})} when ψm0\psi_{m}\rightarrow 0. Both of u1u_{1} and u2u_{2} are bounded no matter what ψm\psi_{m} is.

Step 3: The asymptotic behavior of f.

We examining equation (111). For the first row we have:

g(12(r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2)(r1+r2)+g(\frac{1}{2}(r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}})(r_{1}+r_{2})\rightarrow+\infty

whe ψm0\psi_{m}\rightarrow 0. This is because

12(r2+r1)tanθ2tanθ1tanθ1+tanθ2u1tanθ2tanθ1+tanθ2u2tanθ1tanθ1+tanθ2+\frac{1}{2}(r_{2}+r_{1})\frac{\tan\theta_{2}\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}-u_{1}\frac{\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-u_{2}\frac{\tan\theta_{1}}{\tan\theta_{1}+\tan\theta_{2}}\rightarrow+\infty

by using step 1 and step 2 and:

r2+r1+r_{2}+r_{1}\rightarrow+\infty

by using step 2.

For the second row in the equation in (111) we can show the following result:

gr1r2tanθ1tanθ2tanθ1+tanθ212g1(tanθ1+tanθ2)(u1u2)2+σsinψ2σsinψ1+{}gr_{1}r_{2}\frac{\tan\theta_{1}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}-\frac{1}{2}g\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(u_{1}-u_{2})^{2}+\sigma\sin\psi_{2}-\sigma\sin\psi_{1}\rightarrow+\infty (113)

due to the fact that the first term in equation (113) converges to \infty when ψm0\psi_{m}\rightarrow 0. The other terms are bounded for any ψm(ψ1,0)\psi_{m}\in(\psi_{1},0).

In conclusion, we have f(ψm)+f(\psi_{m})\rightarrow+\infty when ψm0\psi_{m}\rightarrow 0

So from the previous theorem we know that when ψm0\psi_{m}\rightarrow 0, the total volume 𝒱(ψm)\mathcal{V}(\psi_{m}) enclosed by the boundary curve and two walls converges to ++\infty. Then we want to study the minimum bound of the total volume. We have the following theorem:

Refer to caption
Figure 8:
Theorem 3.8.

The volume functional 𝒱(ψm){\mathcal{V}}(\psi_{m}) is bounded from below by a minimum volume VmV_{m} which is defined by:

Vm:=minψm(max(ψ1,ψ2),0)𝒱(ψm)V_{m}:=\min_{\psi_{m}\in(\max(\psi_{1},\psi_{2}),0)}\mathcal{V}(\psi_{m})

VmV_{m} is positive when ψ1ψ2\psi_{1}\neq\psi_{2}

Proof.

We also divide the area into several parts. We call the area enclosed by y=u2y=u_{2} and two walls Ω1\Omega_{1} and the corresponding area 𝒱1(ψm){\mathcal{V}}_{1}(\psi_{m}). Then we have the total volume 𝒱(ψm)𝒱1(ψm){\mathcal{V}}(\psi_{m})\geq{\mathcal{V}}_{1}(\psi_{m}). See the figure 7 :

From a simple geometry relation, we obtain:

𝒱1(ψm)=12(u2+P0g)2(1tanθ1+1tanθ2){\mathcal{V}}_{1}(\psi_{m})=\frac{1}{2}(u_{2}+\frac{P_{0}}{g})^{2}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}})

Then plug the equation (103) and P0=g((r2+xm)tanθ2)u2P_{0}=g((r_{2}+x_{m})\tan\theta_{2})-u_{2} into the equation above and get:

𝒱1\displaystyle{\mathcal{V}}_{1} =12(r2+xm)2tanθ2(1tanθ1+1tanθ2)\displaystyle=\frac{1}{2}(r_{2}+x_{m})^{2}\tan\theta_{2}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}})
=12(r1tanθ1+r2tanθ1(u1u2)tanθ1+tanθ2)2tanθ2(1tanθ1+1tanθ2)\displaystyle=\frac{1}{2}(\frac{r_{1}\tan\theta_{1}+r_{2}\tan\theta_{1}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}})^{2}\tan\theta_{2}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}})

using Lemma 3.6, we know that r2+xm=x2>0r_{2}+x_{m}=x_{2}>0, which implies that

r1tanθ1+r2tanθ1(u1u2)0r_{1}\tan\theta_{1}+r_{2}\tan\theta_{1}-(u_{1}-u_{2})\geq 0

for any maximal angle ψm\psi_{m}. And it equals zero if and only if x1=0x_{1}=0, which means that (x1,u1)(x_{1},u_{1}) is the original point. Then we plug x1=r1=0x_{1}=r_{1}=0 and u1=u2u_{1}=u_{2} into the equation (107) to obtain that:

0=xm=r2tanθ2tanθ1+tanθ20=x_{m}=\frac{-r_{2}\tan\theta_{2}}{\tan\theta_{1}+\tan\theta_{2}}

which implies that r2=0r_{2}=0. We then use (106) to obtain that:

0=ψ1ψ2cosγg2σg(cosψmcosγ)𝑑γ{}0=\int_{\psi_{1}}^{\psi_{2}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma (114)

The right-hand side of (114) is positive when ψ1ψ2\psi_{1}\neq\psi_{2}. Therefore 𝒱(ψ1)\mathcal{V}(\psi_{1}) positive when ψ1ψ2\psi_{1}\neq\psi_{2}. And 𝒱1(ψm)\mathcal{V}_{1}(\psi_{m}) is positive when ψ(ψ1,0)\psi\in(\psi_{1},0)

Furthermore, since ψm\psi_{m} is the maximum angle, we have ψm[ψ1,0)\psi_{m}\in[\psi_{1},0). We already know that r2+xm+r_{2}+x_{m}\rightarrow+\infty when ψm0\psi_{m}\rightarrow 0 which implies that limψm0𝒱1(ψm)=+\lim_{\psi_{m}\rightarrow 0}\mathcal{V}_{1}(\psi_{m})=+\infty We then conclude that 𝒱1{\mathcal{V}}_{1} is bounded below by some constant. Since 𝒱𝒱1{\mathcal{V}}\geq{\mathcal{V}}_{1}, we have 𝒱{\mathcal{V}} bounded below by a positive constant VmV_{m}. And VmV_{m} is positive when ψ1ψ2\psi_{1}\neq\psi_{2}

Then we only need to consider the case where V is small, such that VVmV\leq V_{m}. In the previous calculation, we assumed that the maximal angle ψ3\psi_{3} is not on the boundary. Therefore, we only need to discuss the scenario in which ψ(x)\psi(x) reaches its maximum at the boundary and is either a monotonous increasing function or a monotone decreasing function. So, ψ\psi serves as a legal parameter for the boundary curve in this case.

The remaining computation is similar to Section 2. We proceed with the assumption that the boundary curve meets the left wall at the point (x1,u1)(x_{1},u_{1}). We then derive the following expression for the other contact point:

Theorem 3.9.

We let (x1,u1)(x_{1},u_{1}) and (x2,u2)(x_{2},u_{2}) be coordinates of two contact points. We have the following results:

1: When ψ1ψ2\psi_{1}\geq\psi_{2}, then ψ(x)\psi(x) is a monotone decreasing function of x and we have their expressions as follows:

x2\displaystyle x_{2} =x1+ψ2ψ1cosγdγgu12+2σg(cosψ1cosψ2)\displaystyle=x_{1}+\int_{\psi_{2}}^{\psi_{1}}\frac{\cos\gamma d\gamma}{g\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(\cos\psi_{1}-\cos\psi_{2})}} (115)
u2\displaystyle u_{2} =u12+2σg(cosψ1cosψ2)\displaystyle=\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(\cos\psi_{1}-\cos\psi_{2})} (116)

2: When ψ2ψ1\psi_{2}\geq\psi_{1}, then ψ(x)\psi(x) is a monotone increasing function of x and we have expressions for x2x_{2} and u2u_{2} as follows:

x2\displaystyle x_{2} =x1+ψ1ψ2cosγdγgu12+2σg(cosψ1cosψ2)\displaystyle=x_{1}+\int_{\psi_{1}}^{\psi_{2}}\frac{\cos\gamma d\gamma}{g\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(\cos\psi_{1}-\cos\psi_{2})}} (117)
u2\displaystyle u_{2} =u12+2σg(cosψ1cosψ2)\displaystyle=\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(\cos\psi_{1}-\cos\psi_{2})} (118)
Proof.

The proof is the same as Theorem 3.5 and Theorem 2.2 ∎

Then for the remaining part of the discussion we suppose that ψ2ψ1<0\psi_{2}\leq\psi_{1}<0. If ψ1<ψ2\psi_{1}<\psi_{2}, we can use similar computation. See the figure 8

Just like the proof of the equation (107), we have:

tanθ1(x1)=x2tanθ2+(u1u2){}\tan\theta_{1}(-x_{1})=x_{2}\tan\theta_{2}+(u_{1}-u_{2}) (119)

We then plug x2=x1+r2x_{2}=x_{1}+r_{2} into the equation (119). and then obtain:

x1=r2tanθ2+(u2u1)tanθ1+tanθ2{}x_{1}=\frac{-r_{2}\tan\theta_{2}+(u_{2}-u_{1})}{\tan\theta_{1}+\tan\theta_{2}} (120)

Next, we express the total volume in terms of x1x_{1} and u1u_{1}. We divide the total volume into two regions: the first one Ω1\Omega_{1}, is the area enclosed by the boundary curve, the left wall, and u=u2u=u_{2}; the second: Ω2\Omega_{2} is the area bounded by two walls and u=u2u=u_{2}. See the figure 8. The corresponding volumes of these regions are denoted by 𝒱1{\mathcal{V}}_{1} and 𝒱2{\mathcal{V}}_{2}, respectively.

We can use the same computation in Section 2 to obtain the following results:

𝒱1(u1)\displaystyle{\mathcal{V}}_{1}(u_{1}) =𝒱1(x1(u1),u1)\displaystyle=\mathcal{V}_{1}(x_{1}(u_{1}),u_{1})
=ψ1ψ2ψψ1cosγgu12+2σg(1cosγ)𝑑γsinψgu12+2σg(1cosψ)𝑑ψ12tanθ1(u2u1)2\displaystyle=\int_{\psi_{1}}^{\psi_{2}}\int_{\psi}^{\psi_{1}}\frac{\cos\gamma}{g\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(1-\cos\gamma)}}d\gamma\frac{\sin\psi}{g\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(1-\cos\psi)}}d\psi-\frac{1}{2\tan\theta_{1}}(u_{2}-u_{1})^{2} (121)
𝒱2(u1)\displaystyle{\mathcal{V}}_{2}(u_{1}) =𝒱2(x1(u1),u1)=12(1tanθ1+1tanθ2)(u2(u1+x1tanθ1))\displaystyle=\mathcal{V}_{2}(x_{1}(u_{1}),u_{1})=\frac{1}{2}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}})(u_{2}-(u_{1}+x_{1}\tan\theta_{1})) (122)

Then we plug equation (120) into equation (122) and we can show that:

𝒱2=121(tanθ1+tanθ2)(1tanθ1+1tanθ2)(r2tanθ2+(u2u1)tanθ2){\mathcal{V}}_{2}=\frac{1}{2}\frac{1}{(\tan\theta_{1}+\tan\theta_{2})}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}})(r_{2}\tan\theta_{2}+(u_{2}-u_{1})\tan\theta_{2})

Based on the computation above, we can now express the total volume 𝒱=𝒱1+𝒱2{\mathcal{V}}={\mathcal{V}}_{1}+{\mathcal{V}}_{2} in terms of u1u_{1}. We combine our ODE (19):

x(sinψ)=gu\partial_{x}(\sin\psi)=gu

with the result in Theorem 3.9 that ψ(x)\psi(x) is monotonically decreasing to obtain:

gu1=x(sinψ)(x1)0gu_{1}=\partial_{x}(\sin\psi)(x_{1})\leq 0

which implies that:

u10u_{1}\leq 0

Thus, we determine the possible range of values for u1u_{1} and proceed to establish the following theorem.

Theorem 3.10.

The total volume 𝒱(u1){\mathcal{V}}(u_{1}) is a bounded function of u1u_{1} when u10u_{1}\leq 0. And the upper bound V1V_{1} is bigger than or equal to VmV_{m}.

Proof.

We first prove that 𝒱\mathcal{V} is a bounded function of u1u_{1}. We notice that it is a continuous function of u1u_{1}. Also when u1u_{1}\rightarrow-\infty, (115) and (116) imply that:

r2=x2x1=ψ2ψ1cosγdγgu12+2σg(cosψ1cosψ2)0whenu1{}r_{2}=x_{2}-x_{1}=\int_{\psi_{2}}^{\psi_{1}}\frac{\cos\gamma d\gamma}{g\sqrt{u_{1}^{2}+\frac{2\sigma}{g}(\cos\psi_{1}-\cos\psi_{2})}}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty (123)
u2u10whenu1{}u_{2}-u_{1}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty (124)

Then by equation (120), we have:

x1=r2tanθ2+(u2u1)tanθ1+tanθ20whenu1{}x_{1}=\frac{-r_{2}\tan\theta_{2}+(u_{2}-u_{1})}{\tan\theta_{1}+\tan\theta_{2}}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty (125)

Based on the expression for 𝒱1\mathcal{V_{1}} and 𝒱2\mathcal{V}_{2}((121) and (122)), we obtain the following result by combining(123),(124) and (125) together:

𝒱10whenu1\mathcal{V}_{1}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty

and:

𝒱20whenu1\mathcal{V}_{2}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty

The above two asymptotic results mean that:

𝒱=𝒱1+𝒱20whenu1\mathcal{V}=\mathcal{V}_{1}+\mathcal{V}_{2}\rightarrow 0\leavevmode\nobreak\ \operatorname{when}\leavevmode\nobreak\ u_{1}\rightarrow-\infty

Based on the discussion above, when u10u_{1}\leq 0, 𝒱\mathcal{V} is a bounded function of u1u_{1}.(Here we only need to care about the point u1=0u_{1}=0. Based on our previous discussion, the integral in the representation (121) is infinity only when ψ1=0\psi_{1}=0. So the volume functional is bounded at this point since ψ1<0\psi_{1}<0.)

We then only need to show that the upper bound V1V_{1} is bigger than or equals VmV_{m}. We notice that when u1=0u_{1}=0, the curve is the same as the curve in Theorem 3.8 when ψm=ψ1\psi_{m}=\psi_{1}. So we have:

V1limu10𝒱(u1)=𝒱(ψ1)VmV_{1}\geq\lim_{u_{1}\rightarrow 0}{\mathcal{V}}(u_{1})={\mathcal{V}}(\psi_{1})\geq V_{m}

which finishes the proof. ∎

We then combine Theorem 3.8 and Theorem 3.10 to get:

Theorem 3.11.

When θ1\theta_{1} and θ2\theta_{2} are both acute angles, we have at least one steady state for any total volume V.

We also have two constant V1V_{1} and VmV_{m} which depend on θ1\theta_{1}, θ2\theta_{2},σ\sigma and [[γ]][[\gamma]] such that:

  • When VVmV\geq V_{m}, we can find a ψm\psi_{m} with a solution function of system (3) which reaches its maximal slope angle ψ\psi in the interval (ψ2,ψ1)(\psi_{2},\psi_{1})(or (ψ1,ψ2)(\psi_{1},\psi_{2}) if ψ1ψ2\psi_{1}\leq\psi_{2}).

  • When VV1V\leq V_{1}, we can find a u1u_{1} with a solution function of system (3) which reaches its maximal slope angle on the boundary.

  • When V1>VmV_{1}>V_{m}, the solution is not unique.

3.3 General case when θ1>π2\theta_{1}>\frac{\pi}{2}

In this case, the only difference compared with Subsection 3.2 is that ψ1\psi_{1} and ψ2\psi_{2} can have the same sign even when [[γ]]<0[[\gamma]]<0.

By definition, we have ψ1=θ1+π2+γ\psi_{1}=-\theta_{1}+\frac{\pi}{2}+\gamma and ψ2=θ2π2γ\psi_{2}=\theta_{2}-\frac{\pi}{2}-\gamma. Then we can observe that ψ2<π2\psi_{2}<-\frac{\pi}{2} when γ>θ2\gamma>\theta_{2}, which means that the boundary curve may no longer be a function of x. Therefore, we cannot use equation (99) to derive the equation for P0P_{0}. We need to use the method in Section 2 to compute the total volume using the coordinates of a special point. This is the only difference of this subsection compared with Subsection 3.2.

In this section, we only discuss the case γ>θ2\gamma>\theta_{2}. For the case γθ2\gamma\leq\theta_{2}, we can just follow the method in Subsection 3.2. We also assume that VVmV\geq V_{m} for a constant VmV_{m} to be determined such that ψ\psi attains its maximum in the interior. And we assume that ψ(x)\psi(x) reaches its maximal point at (xm,0)(x_{m},0) with the maximal angle ψm\psi_{m}.

We suppose that the curve meets two walls at two contact points (x1,v1)(x_{1},v_{1}) and (x2,v2)(x_{2},v_{2}). After the shift, they become (x1,u1)(x_{1},u_{1}) and (x2,u2)(x_{2},u_{2}). See Figure 9. Then we have the computation (It is the same as the computation in Section two):

Refer to caption
Figure 9:
x1\displaystyle x_{1} =xmψ1ψmcosγg2σg(cosψmcosγ)𝑑γ\displaystyle=x_{m}-\int_{\psi_{1}}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma (126)
u1\displaystyle u_{1} =2σg(cosψmcosψ1)\displaystyle=\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi_{1})} (127)
x2\displaystyle x_{2} =xm+ψ2ψmcosγg2σg(cosψmcosγ)𝑑γ\displaystyle=x_{m}+\int_{\psi_{2}}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma (128)
u2\displaystyle u_{2} =2σg(cosψmcosψ2)\displaystyle=-\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi_{2})} (129)

We also use the same method as in the previous Subsection to derive the relation between xmx_{m} and ψm\psi_{m}. We should have:

x1tanθ1x2tanθ2=u1u2-x_{1}\tan\theta_{1}-x_{2}\tan\theta_{2}=u_{1}-u_{2}

which implies that:

xm=r1tanθ1r2tanθ2(u1u2)tanθ1+tanθ2{}x_{m}=\frac{r_{1}\tan\theta_{1}-r_{2}\tan\theta_{2}-(u_{1}-u_{2})}{\tan\theta_{1}+\tan\theta_{2}} (130)

Then we use xmx_{m} and ψm\psi_{m} to express the total volume. By using geometric computation, we have:

𝒱(ψm)=\displaystyle\mathcal{V}(\psi_{m})= (r1u1ψ1ψmψψmcosγg2σg(cosψmcosγ)𝑑γsinψg2σg(cosψmcosψ)𝑑ψ)\displaystyle(r_{1}u_{1}-\int_{\psi_{1}}^{\psi_{m}}\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\frac{-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi)}}d\psi)
+ψ2ψmψψmcosγg2σg(cosψmcosγ)𝑑γsinψg2σg(cosψmcosψ)𝑑ψ\displaystyle+\int_{\psi_{2}}^{\psi_{m}}\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\frac{-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi)}}d\psi
+12(u1+u2)21tanθ1+(u2)r1+12(u2u1x1tanθ1)2(1tanθ1+1tanθ2)\displaystyle+\frac{1}{2}(u_{1}+u_{2})^{2}\frac{1}{-\tan\theta_{1}}+(-u_{2})r_{1}+\frac{1}{2}(u_{2}-u_{1}-x_{1}\tan\theta_{1})^{2}(\frac{1}{\tan\theta_{1}}+\frac{1}{\tan\theta_{2}}) (131)

Then we need to prove the following theorem about the total volume:

Theorem 3.12.

When ψm0\psi_{m}\rightarrow 0, we have 𝒱(ψm)+{\mathcal{V}}(\psi_{m})\rightarrow+\infty.

Proof.

Using the computation in Theorem 3.7, we know that:

ψψmcosγg2σg(cosψmcosγ)𝑑γ+\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\rightarrow+\infty

when ψm0\psi_{m}\rightarrow 0 for any ψ(ψ2,ψm)\psi\in(\psi_{2},\psi_{m}). So we should have:

𝒱(ψm)ψ2ψmψψmcosγg2σg(cosψmcosγ)𝑑γsinψg2σg(cosψmcosψ)𝑑ψ+\mathcal{V}(\psi_{m})\geq\int_{\psi_{2}}^{\psi_{m}}\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\frac{-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi)}}d\psi\rightarrow+\infty

when ψm0\psi_{m}\rightarrow 0, which means that:

𝒱(ψm)+\mathcal{V}(\psi_{m})\rightarrow+\infty

when ψm0\psi_{m}\rightarrow 0, which finishes the proof. ∎

Just like the previous subsection, we want to show that 𝒱(ψm){\mathcal{V}}(\psi_{m}) has a lower bound VmV_{m}.

Theorem 3.13.

The energy functional 𝒱(ψm)\mathcal{V}(\psi_{m}) has a positive lower bound VmV_{m}.

Proof.

Using equation (131), we know that:

V(ψm)ψ2ψmψψmcosγg2σg(cosψmcosγ)𝑑γsinψg2σg(cosψmcosψ)𝑑ψ{}V(\psi_{m})\geq\int_{\psi_{2}}^{\psi_{m}}\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\frac{-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi)}}d\psi (132)

Using the definition of ψ1\psi_{1} and ψ2\psi_{2}( (6) and (7)), we know that ψ1=θ1+π2+γ\psi_{1}=-\theta_{1}+\frac{\pi}{2}+\gamma and ψ2=θ2π2γ\psi_{2}=\theta_{2}-\frac{\pi}{2}-\gamma. First we can show that ψ1(π2,0)\psi_{1}\in(-\frac{\pi}{2},0) since

0>ψ1=θ1+π2+γθ1+π2+θ2>π+π2=π2{}0>\psi_{1}=-\theta_{1}+\frac{\pi}{2}+\gamma\geq-\theta_{1}+\frac{\pi}{2}+\theta_{2}>-\pi+\frac{\pi}{2}=-\frac{\pi}{2} (133)

We also have the following computation:

ψmψ1=θ1+π2γ=θ2θ1ψ2>πψ2{}\psi_{m}\geq\psi_{1}=-\theta_{1}+\frac{\pi}{2}-\gamma=\theta_{2}-\theta_{1}-\psi_{2}>-\pi-\psi_{2} (134)

We can use (133) and (134) to show that cosψm>cosψ2\cos\psi_{m}>\cos\psi_{2}. Therefore, we obtain that:

ψψmcosγg2σg(cosψmcosγ)𝑑γψψmcosγg2σg(cosψm+1)𝑑γ=sinψmsinψg2σg(cosψm+1){}\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\gamma)}}d\gamma\geq\int_{\psi}^{\psi_{m}}\frac{\cos\gamma}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}+1)}}d\gamma=\frac{\sin\psi_{m}-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}+1)}} (135)

The proof of (135) is the same as the proof in Lemma 2.9. We do not write down the details here. Since ψ(ψ2,ψm)\psi\in(\psi_{2},\psi_{m}), we have sinψm>sinψ\sin\psi_{m}>\sin\psi. We combine (135) with the fact that sinψm>sinψ\sin\psi_{m}>\sin\psi and plug them back into the equation (132) to get:

𝒱(ψm)\displaystyle\mathcal{V}(\psi_{m}) ψ2ψmsinψmsinψg2σg(cosψm+1)sinψg2σg(cosψmcosψ)𝑑ψ\displaystyle\geq\int_{\psi_{2}}^{\psi_{m}}\frac{\sin\psi_{m}-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}+1)}}\frac{-\sin\psi}{g\sqrt{\frac{2\sigma}{g}(\cos\psi_{m}-\cos\psi)}}d\psi
>12σg(cosψm+1)ψ2ψmsinψsinψm+sin2ψdψ\displaystyle>\frac{1}{2\sigma g(\cos\psi_{m}+1)}\int_{\psi_{2}}^{\psi_{m}}-\sin\psi\sin\psi_{m}+\sin^{2}\psi d\psi
>14σg(sinψm(cosψmsinψ2)+ψ2ψmsin2ψdψ)\displaystyle>\frac{1}{4\sigma g}(-\sin\psi_{m}(\cos\psi_{m}-\sin\psi_{2})+\int_{\psi_{2}}^{\psi_{m}}\sin^{2}\psi d\psi)
>14σgψ2ψ1sin2ψdψ>C(ψ1,ψ2)>0\displaystyle>\frac{1}{4\sigma g}\int_{\psi_{2}}^{\psi_{1}}\sin^{2}\psi d\psi>C(\psi_{1},\psi_{2})>0

This inequality combined with Theorem 3.12 imply that that 𝒱(ψm){\mathcal{V}}(\psi_{m}) is a positive continuous function of ψm[ψ1,0)\psi_{m}\in[\psi_{1},0). Also when ψm0\psi_{m}\rightarrow 0, 𝒱+{\mathcal{V}}\rightarrow+\infty. So 𝒱(ψm)\mathcal{V}(\psi_{m}) is bounded from below. ∎

Then we want to compute the case when VV is small. Also we assume that ψ1>ψ2\psi_{1}>\psi_{2}. We suppose that the curve intersects the left wall at the contact point (x1,u1)(x_{1},u_{1}). Then we have the representation x2x_{2} and u2u_{2} the same as equation (115) and equation (116).

Actually, the representation for x1x_{1} and total volume 𝒱(u1)\mathcal{V}(u_{1}) is the same as equation (120), equation (121) and (122). Then we have the theorem:

Theorem 3.14.

The total volume V=𝒱1+𝒱2{V}=\mathcal{V}_{1}+\mathcal{V}_{2} is a function of u1u_{1} bounded from above when u1<0u_{1}<0 by a positive constant VaV_{a}. And VaVmV_{a}\geq V_{m}.

Proof.

The proof is the same as Theorem 3.10. ∎

Finally we combine Theorem 3.11, Theorem 3.13, and Theorem 3.14 together to get:

Theorem 3.15.

When ψ1\psi_{1} and ψ2\psi_{2} have the same sign, we have the existence of the solution function of the equation system (3)

However, unlike the case in Section 2, we cannot show the uniqueness of the steady state in this case.

ACKNOWLEDGEMENTS

The author thanks Yan Guo for numerous comments. His mentorship and constructive feedback significantly contribute to the development of this work.

This work is supported in part by NSF Grant DMS-2405051.

References

  • [1] Yunus Cengel and John Cimbala. Ebook: Fluid mechanics fundamentals and applications (si units). McGraw Hill, 2013.
  • [2] Robert Finn. Equilibrium capillary surfaces, volume 284. Springer Science & Business Media, 2012.
  • [3] Carl Friedrich Gauss and Carolo Friderico Gauss. Principia generalia theoriae figurae fluidorum in statu aequilibrii. Springer, 1877.
  • [4] Yan Guo and Ian Tice. Almost exponential decay of periodic viscous surface waves without surface tension. Archive for Rational Mechanics and Analysis, 207:459–531, 2013.
  • [5] Yan Guo and Ian Tice. Decay of viscous surface waves without surface tension in horizontally infinite domains. Analysis & PDE, 6(6):1429–1533, 2013.
  • [6] Yan Guo and Ian Tice. Local well-posedness of the viscous surface wave problem without surface tension. Analysis & PDE, 6(2):287–369, 2013.
  • [7] Yan Guo and Ian Tice. Stability of contact lines in fluids: 2d stokes flow. Archive for Rational Mechanics and Analysis, 227:767–854, 2018.
  • [8] Yan Guo and Ian Tice. Stability of contact lines in fluids: 2d navier–stokes flow. Journal of the European Mathematical Society, 26(4):1445–1557, 2023.
  • [9] Yan Guo, Ian Tice, Lei Wu, and Yunrui Zheng. Global well-posedness of contact lines: 2d navier-stokes flow. arXiv preprint arXiv:2407.17895, 2024.
  • [10] Juhi Jang, Ian Tice, and Yanjin Wang. The compressible viscous surface-internal wave problem: local well-posedness. SIAM Journal on Mathematical Analysis, 48(4):2602–2673, 2016.
  • [11] Juhi Jang, Ian Tice, and Yanjin Wang. The compressible viscous surface-internal wave problem: local well-posedness. SIAM Journal on Mathematical Analysis, 48(4):2602–2673, 2016.
  • [12] Chanwoo Kim and Ian Tice. Dynamics and stability of surface waves with surfactants. SIAM Journal on Mathematical Analysis, 49(2):1295–1332, 2017.
  • [13] M.de Laplace. Celestial mechanics. Vols. I–IV. Translated from the French, with a commentary, by Nathaniel Bowditch Chelsea Publishing Co., Inc., Bronx, N.Y., (1966).
  • [14] Giovanni Leoni and Ian Tice. Traveling wave solutions to the free boundary incompressible navier-stokes equations. Communications on Pure and Applied Mathematics, 76(10):2474–2576, 2023.
  • [15] Noah Stevenson and Ian Tice. Well-posedness of the stationary and slowly traveling wave problems for the free boundary incompressible navier-stokes equations. Journal of Functional Analysis, 287(11):110617, 2024.
  • [16] Ian Tice and Lei Wu. Dynamics and stability of surface waves with bulk-soluble surfactants. Acta Applicandae Mathematicae, 161:35–70, 2019.
  • [17] Ian Tice and Lei Wu. Dynamics and stability of sessile drops with contact points. Journal of Differential Equations, 272:648–731, 2021.
  • [18] Thomas Young. Iii. an essay on the cohesion of fluids. Philosophical transactions of the royal society of London, (95):65–87, 1805.
  • [19] Yunrui Zheng and Ian Tice. Local well posedness of the near-equilibrium contact line problem in 2-dimensional stokes flow. SIAM Journal on Mathematical Analysis, 49(2):899–953, 2017.