Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Modeling film flows down a rotating slippery cylinder

Souradip Chattopadhyay schatto5@ncsu.edu (ORCID: 0000-0002-4418-6201) Department of Mathematics, North Carolina State University, Raleigh, North Carolina 27695, USA    Amar K. Gaonkar amar.gaonkar@iitdh.ac.in Department of Mechanical, Materials and Aerospace Engineering, IIT Dharwad, Karnataka 580011, India    Hangjie Ji hangjie_ji@ncsu.edu (corresponding author) Department of Mathematics, North Carolina State University, Raleigh, North Carolina 27695, USA
Abstract

This study investigates the nonlinear stability and dynamics of gravity-driven viscous films on a vertical rotating cylinder, considering both outer and inner surface flows with slip conditions at the cylinder wall. We develop an asymptotic model for the combined effects of rotation and wall slippage. Linear stability analysis indicates that wall slippage enhances instability on both surfaces, while rotation has differing impacts: it amplifies instability due to slip for outer surface flow but reduces it for inner surface flow. A weakly nonlinear stability analysis is then conducted to explore the combined impact of rotation and wall slip on flow stability beyond the linear regime, including the bifurcation of the nonlinear evolution equation for both surfaces. The traveling wave solution of the model is analyzed, showing how rotation affects nonlinear wave speed with a slippery wall. A stability analysis of the traveling wave solutions is also performed. Numerical simulations of the nonlinear evolution of the free surface reveal that increasing slip length enhances the choke phenomenon in inner surface flow, while rotation can delay this effect. Additionally, simulations show that for flow along the outer surface of a slippery rotating cylinder, the film tends to break up into droplets in the presence of rotation.

I Introduction

Coating non-flat objects is crucial for manufacturing various products, such as aerospace components [1] and three-dimensional printed parts [2], to protect their surfaces. Understanding the behavior of a thin liquid film flowing down a vertical cylinder’s inner or outer surface provides valuable insights into this process. This investigation reveals complex interfacial dynamics, including droplet formation and traveling wave patterns [3, 4], which are critical for applications like coating insulation on wires [5], gas absorption [6], desalination [7], and surface patterning on cylindrical substrates [8]. The flow on the inner surface of cylindrical geometries, such as within tubes, is also important in applications such as lung airways [9].

When a thin liquid film is applied to the surface of a fiber or cylinder, it typically experiences capillary instability [10, 11, 4]. Theoretical and experimental studies have examined the dynamics of liquid film flow along a vertical cylindrical geometry [4, 3], providing a detailed understanding of the underlying physics. Unlike films on flat substrates, those on cylindrical surfaces exhibit inherent instability due to the additional azimuthal curvature of the free surface [3], resulting from the interaction of two distinct instability mechanisms. The first is the Rayleigh–Plateau (RP) instability, influenced by the gravity-driven flow in the axisymmetric film coating a cylindrical fiber. The second is the hydrodynamic instability of a falling film due to inertia, known as the Kapitza mode of instability [12]. The challenge becomes more pronounced when the cylinder rotates around its vertical axis, as gravitational, inertial, viscous, surface tension and centrifugal forces interact, potentially leading to instabilities and coating failures. When the cylinder is horizontal, various studies have reported that rotating cylinders effectively control coating flow stability and dynamics, which is crucial in numerous industrial applications for surface control and protection [13, 14, 15]. Experimental investigations by Thoroddsen & Mahadevan [16], as well as Kozlov & Polezhaev [17], have demonstrated how rotation influences flow dynamics in case of horizontal scenarios. Compared to horizontal cylinders, there are limited studies on liquid film flow along a rotating vertical cylinder or fiber in the current literature. Experimental results from rotating horizontal cylinders inspire us to focus on existing models of vertical cylinders under rotation, examining flow along the outer and inner surfaces. By extending these studies, we aim to understand the underlying physics better.

Frenkel [20] examined the case of a large cylinder radius (resulting in a thin film) and derived a Benney-type equation for the interface. However, their model was not derived asymptotically, and the bead velocity was overestimated. Kliakhandler et al. [21] and Craster & Matar [10] investigated the instability of a uniform film at higher flow rates by regularly wetting the fiber from the top. They reported the nonlinear dynamics of axisymmetric waves occurring far from the inlet. For a thin liquid film on the outer surface of a vertical cylindrical fiber, Kim et al. [22] studied a positivity-preserving finite difference method. A recent study by Taranets et al. [23] proved the existence of weak solutions for the flow of a thin liquid film outside a vertical fiber and established the presence of a traveling wave solution for the system. Novbari & Oron [24] employed the energy integral method to investigate the nonlinear dynamics of a thin film flowing on the outer surface of a vertical cylinder. Various models have been reported to understand complex dynamical phenomena of liquid film flow along a vertical fiber/cylinder [25, 26]. Typical long-wave models either assume that the film thickness is much smaller than the radius of the cylinder [4, 18] or that the film radius is much smaller than its characteristic length [10, 27]. These models simplify the Navier-Stokes equations into single evolution equations for the film thickness, making them suitable for analyzing thin-film flows with low Reynolds numbers. More advanced models, such as the integral-boundary-layer model [28] and the weighted-residual integral-boundary-layer model [11], have been developed for scenarios involving a moderate Reynolds number. These models formulate evolution equations for film thickness and volumetric flow rates.

The above-mentioned models assumed the conventional no-slip boundary condition at the cylinder/fiber wall interface. However, numerous studies have explored the potential role of slip at the solid-liquid interface, which can enhance free-surface instability. Haefner et al. [29] investigated the influence of slip on the RP instability in a film along a fiber. Their study, combining experiment and theory, highlighted that wall slip significantly affects the growth rate of instabilities. Applying lubrication theory, the model proposed by Halpern & Wei [18] concluded that slip length enhances drop formation along a vertical fiber. Ji et al. [19] presented a study combining experimental and numerical results on thin film flows along the outer surface of a slippery fiber. Their lubrication model incorporated slip boundary conditions, nonlinear curvature terms, and a film stabilization term. Chao et al. [30] explored the combined influence of wall slip and thermocapillary effects on thin film flow along a vertical cylinder, demonstrating that both factors enhance the RP instability. Chattopadhyay [31] extended this investigation to include the impact of a chemical reaction. His results indicated that an endothermic reaction intensifies the RP instability in the presence of a slippery wall compared to the no-slip case, whereas an exothermic reaction reduces it. The studies [29, 18, 19, 30, 31] examined the influence of slip length on the stability and dynamics of thin film flow along the outer surface of a slippery cylinder/fiber. A recent study by Schwitzerlett et al. [32] developed an asymptotic model examining flow within a vertical tube with a slippery wall. Their findings revealed that wall slippage destabilizes flow instability, similar to what is observed on the outer surface of a vertical cylinder/fiber. Additionally, they concluded that the presence of slip length enhances the formation of plugs.

Although numerous studies have examined the dynamics and stability of vertical cylindrical geometry, very few have addressed the effect of cylinder rotation. Dávalos-Orozco & Ruiz-Chavarria [33] investigated the linear stability of a fluid layer flowing inside and outside of a rotating vertical cylinder. Their analysis incorporated two additional approximations: small wavenumber and small Reynolds number. Using a long-wave expansion method, Chen et al. [34] developed a model to examine the stability and dynamics of a condensate liquid film along the outer surface of a rotating cylinder. They performed both linear and weakly nonlinear stability analyses. They found that the rotation increases the instability of the flow. Rietz et al. [35] conducted an experimental investigation into developing a thin film on the outer surface of a vertical rotating cylinder. Their study specifically addressed the generation of 2D and 3D waves, the formation of rivulets, and dripping phenomena. Liu & Ding [36] examined the dynamics of a coating flow over a fiber that rotates about its axis. They formulated an evolution equation for the surface using the long-wave theory. Their findings indicated that rotation has a destabilizing effect. Furthermore, their spatiotemporal stability analysis revealed that rotation enhances absolute instability. Farooq et al. [37] investigated the dynamics of a falling film on the inner surface of a rotating cylinder. They utilized numerical simulations with the VOF method to explore the effects of rotation on interfacial dynamics in two and three dimensions. Their findings highlight rotation’s significant influence on interfacial waves’ structure. Mukhopadhyay et al. [38] examined how the dynamics of thin liquid films behave on the outer surface of a vertical rotating cylinder under non-isothermal conditions. They found that rotation intensifies flow instability when thermal effects are present.

In this study, we examine the flow of a thin film over the surface of a vertical cylinder under gravity, where the cylinder rotates about its axis. The primary motivation is to develop a mathematical model that explores two distinct scenarios: one where the film flows along the outer surface of the cylinder and another where it flows along the inner surface. Previous studies by Ji et al. [19], Chao et al. [30], and Schwitzerlett et al. [32] have shown that wall slippage enhances flow instability in similar configurations. This work focuses on how cylinder rotation impacts flow dynamics and stability exacerbated by wall slippage on both outer and inner surfaces. We extend the stability analysis beyond the linear regime, focusing on bifurcation scenarios involving the two key parameters: wall slip and rotation. Additionally, we investigate the influence of these parameters on the propagation speed of traveling wave solutions. Further contributions include a comprehensive stability assessment of these traveling waves for both outer and inner surface flows. Finally, we explore the effect of cylinder rotation on the breakup dynamics of liquid flows along the outer surface and on the plug formation in flows along the inner surface, a phenomenon reported by Liu & Ding [39] and Schwitzerlett et al. [32]. Our study aims to provide insights into delaying such plug formation under varying rotation and slip length conditions.

We organize the present paper as follows. Section II introduces a model incorporating wall slip and rotation for a rotating vertical cylinder. Section III presents a linear stability analysis for the flow. Section IV presents a weakly nonlinear stability analysis, exploring the combined effects of wall slip and rotation beyond the linear regime. Section V focuses on investigating traveling wave solutions and their stability in the presence of slip and rotation. In Section VI, we perform nonlinear simulations of the evolution equation. We discuss the influence of rotation on the choke behavior when the flow is along the inner surface of the slippery cylinder and on the breakup behavior when the flow is along the outer surface of the slippery cylinder. Finally, Section VII summarizes the study’s key findings.

II Model

We consider a Newtonian liquid with density ρ𝜌\rhoitalic_ρ, dynamic viscosity μ𝜇\muitalic_μ, and surface tension σ𝜎\sigmaitalic_σ, flowing down a slippery vertical cylinder with radius r=b𝑟𝑏r=bitalic_r = italic_b under gravity g𝑔gitalic_g. The vertical (axial) coordinate is z𝑧zitalic_z (positive downward). We have focused on two scenarios: one where the liquid film flows along the outer surface of the slippery cylinder and the other where it flows along the inner surface. Additionally, we consider the cylinder is rotating about its axis (the z𝑧zitalic_z-axis) at an angular speed ΩΩ\Omegaroman_Ω. These systems are displayed in FIG. 1. The thickness of the liquid film at any instant is denoted by h=h(z,t)𝑧𝑡h=h(z,t)italic_h = italic_h ( italic_z , italic_t ). For liquid flowing along the outer surface of the cylinder, the liquid-air interface is given by r=b+h(z,t)𝑟𝑏𝑧𝑡r=b+h(z,t)italic_r = italic_b + italic_h ( italic_z , italic_t ), and for liquid flowing along the inner surface of the cylinder, the liquid-gas interface is given by r=bh(z,t)𝑟𝑏𝑧𝑡r=b-h(z,t)italic_r = italic_b - italic_h ( italic_z , italic_t ). Thus, we can generally express the equation of the free surface as r=S(z,t)=b+mh(z,t)𝑟𝑆𝑧𝑡𝑏𝑚𝑧𝑡r=S(z,t)=b+mh(z,t)italic_r = italic_S ( italic_z , italic_t ) = italic_b + italic_m italic_h ( italic_z , italic_t ), where m=±1𝑚plus-or-minus1m=\pm 1italic_m = ± 1. The positive sign indicates liquid flow along the outer surface of the cylinder, while the negative sign indicates flow along the inner surface.

b𝑏bitalic_bh(z,t)𝑧𝑡h(z,t)italic_h ( italic_z , italic_t )g𝑔gitalic_gr𝑟ritalic_rz𝑧zitalic_zz𝑧zitalic_zb𝑏bitalic_br𝑟ritalic_rΩΩ\Omegaroman_ΩΩΩ\Omegaroman_ΩOuterInner
Figure 1: Model diagram for liquid films flowing along the outer (left) and inner (right) surfaces of a rotating cylinder

We consider the two-dimensional hydrodynamic problem, and the governing equations are

r1(ru)r+wz=0,superscript𝑟1subscript𝑟𝑢𝑟subscript𝑤𝑧0r^{-1}\left(ru\right)_{r}+w_{z}=0,italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_r italic_u ) start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 , (1)
ρ(ut+uur+wuzr1v2)=pr+μ[r1(rur)r+uzzr2u],𝜌subscript𝑢𝑡𝑢subscript𝑢𝑟𝑤subscript𝑢𝑧superscript𝑟1superscript𝑣2subscript𝑝𝑟𝜇delimited-[]superscript𝑟1subscript𝑟subscript𝑢𝑟𝑟subscript𝑢𝑧𝑧superscript𝑟2𝑢\rho\left(u_{t}+uu_{r}+wu_{z}-r^{-1}v^{2}\right)=-p_{r}+\mu\left[r^{-1}\left(% ru_{r}\right)_{r}+u_{zz}-r^{-2}u\right],italic_ρ ( italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_u italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_w italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = - italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_μ [ italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_r italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT - italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_u ] , (2)
ρ(wt+uwr+wwz)=pz+μ[r1(rwr)r+wzz]+ρg.𝜌subscript𝑤𝑡𝑢subscript𝑤𝑟𝑤subscript𝑤𝑧subscript𝑝𝑧𝜇delimited-[]superscript𝑟1subscript𝑟subscript𝑤𝑟𝑟subscript𝑤𝑧𝑧𝜌𝑔\rho\left(w_{t}+uw_{r}+ww_{z}\right)=-p_{z}+\mu\left[r^{-1}\left(rw_{r}\right)% _{r}+w_{zz}\right]+\rho g.italic_ρ ( italic_w start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_u italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_w italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = - italic_p start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_μ [ italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_r italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ] + italic_ρ italic_g . (3)

Here p𝑝pitalic_p represents pressure, and 𝒰=(u,v,w)𝒰𝑢𝑣𝑤\mathcal{U}=(u,v,w)caligraphic_U = ( italic_u , italic_v , italic_w ) represents the radial, azimuthal, and axial fluid velocity components. We assume the azimuthal velocity v𝑣vitalic_v remains constant v=bΩ𝑣𝑏Ωv=b\Omegaitalic_v = italic_b roman_Ω [38, 34].

At the surface of the cylinder, i.e., at r=b𝑟𝑏r=bitalic_r = italic_b, we apply the Navier-slip and no-penetration boundary conditions as

u=0,w=mδ^wr,formulae-sequence𝑢0𝑤𝑚^𝛿subscript𝑤𝑟u=0,\quad w=m\widehat{\delta}w_{r},italic_u = 0 , italic_w = italic_m over^ start_ARG italic_δ end_ARG italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , (4)

where δ^^𝛿\widehat{\delta}over^ start_ARG italic_δ end_ARG represents the dimensional slip length.

The boundary conditions at the free surface r=S(z,t)𝑟𝑆𝑧𝑡r=S(z,t)italic_r = italic_S ( italic_z , italic_t ) are the balance of stresses (tangential and normal) and the kinematic condition, which are given below

(uz+wr)(1Sz2)+2(urwz)Sz=0,subscript𝑢𝑧subscript𝑤𝑟1superscriptsubscript𝑆𝑧22subscript𝑢𝑟subscript𝑤𝑧subscript𝑆𝑧0\left(u_{z}+w_{r}\right)\left(1-S_{z}^{2}\right)+2\left(u_{r}-w_{z}\right)S_{z% }=0,( italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) ( 1 - italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + 2 ( italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 , (5)
pap+2μ[ur(uz+wr)Sz+wzSz2](1+Sz2)1=σ𝒦,subscript𝑝𝑎𝑝2𝜇delimited-[]subscript𝑢𝑟subscript𝑢𝑧subscript𝑤𝑟subscript𝑆𝑧subscript𝑤𝑧superscriptsubscript𝑆𝑧2superscript1superscriptsubscript𝑆𝑧21𝜎𝒦p_{a}-p+2\mu\left[u_{r}-\left(u_{z}+w_{r}\right)S_{z}+w_{z}S_{z}^{2}\right]% \left(1+S_{z}^{2}\right)^{-1}=\sigma\mathcal{K},italic_p start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT - italic_p + 2 italic_μ [ italic_u start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - ( italic_u start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] ( 1 + italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = italic_σ caligraphic_K , (6)
u=St+wSz,𝑢subscript𝑆𝑡𝑤subscript𝑆𝑧u=S_{t}+wS_{z},italic_u = italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_w italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , (7)

where 𝒦=m[Szz(1+Sz2)1r1](1+Sz2)1/2𝒦𝑚delimited-[]subscript𝑆𝑧𝑧superscript1superscriptsubscript𝑆𝑧21superscript𝑟1superscript1superscriptsubscript𝑆𝑧212\mathcal{K}=m\left[S_{zz}\left(1+S_{z}^{2}\right)^{-1}-r^{-1}\right]\left(1+S_% {z}^{2}\right)^{-1/2}caligraphic_K = italic_m [ italic_S start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ( 1 + italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] ( 1 + italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT is the curvature and pasubscript𝑝𝑎p_{a}italic_p start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is the atmospheric pressure.

We use the following scaling to derive the dimensionless governing equations and boundary conditions

z=z,(S,r,b)=(S,r,b),t=(/𝒱)t,(u,w)=𝒱(ϵu,w),p=pa+𝒫p,formulae-sequence𝑧superscript𝑧formulae-sequence𝑆𝑟𝑏superscript𝑆superscript𝑟superscript𝑏formulae-sequence𝑡𝒱superscript𝑡formulae-sequence𝑢𝑤𝒱italic-ϵsuperscript𝑢superscript𝑤𝑝subscript𝑝𝑎𝒫superscript𝑝z=\mathcal{L}z^{*},~{}(S,r,b)=\mathcal{H}(S^{*},r^{*},b^{*}),~{}t=\left(% \mathcal{L}/\mathcal{V}\right)t^{*},~{}(u,w)=\mathcal{V}\left(\epsilon u^{*},w% ^{*}\right),~{}p=p_{a}+\mathcal{P}p^{*},italic_z = caligraphic_L italic_z start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , ( italic_S , italic_r , italic_b ) = caligraphic_H ( italic_S start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_r start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_b start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_t = ( caligraphic_L / caligraphic_V ) italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , ( italic_u , italic_w ) = caligraphic_V ( italic_ϵ italic_u start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_p = italic_p start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT + caligraphic_P italic_p start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , (8)

where \mathcal{H}caligraphic_H is the mean thickness of the liquid film, 𝒱𝒱\mathcal{V}caligraphic_V is the velocity scale, \mathcal{L}caligraphic_L is the characteristic length in the axial direction and 𝒫𝒫\mathcal{P}caligraphic_P is the pressure scale and ‘*’ denotes dimensionless variables. We set 𝒫=ρg𝒫𝜌𝑔\mathcal{P}=\rho g\mathcal{L}caligraphic_P = italic_ρ italic_g caligraphic_L, 𝒱=ρg2/μ𝒱𝜌𝑔superscript2𝜇\mathcal{V}=\rho g\mathcal{H}^{2}/\mucaligraphic_V = italic_ρ italic_g caligraphic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_μ, and the aspect ratio ϵ=/1italic-ϵmuch-less-than1\epsilon=\mathcal{H/L}\ll 1italic_ϵ = caligraphic_H / caligraphic_L ≪ 1 is a small parameter.

We use (8) in (2)-(7) with =σ/(ρg)𝜎𝜌𝑔\mathcal{L}=\sigma/\left(\rho g\mathcal{H}\right)caligraphic_L = italic_σ / ( italic_ρ italic_g caligraphic_H ), and obtain the following leading order dimensionless system of equations:

pr=r1b2Ro,r1(rwr)r=pz1,formulae-sequencesubscript𝑝𝑟superscript𝑟1superscript𝑏2𝑅𝑜superscript𝑟1subscript𝑟subscript𝑤𝑟𝑟subscript𝑝𝑧1p_{r}=r^{-1}b^{2}Ro,~{}r^{-1}(rw_{r})_{r}=p_{z}-1,italic_p start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R italic_o , italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_r italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT - 1 , (9)
u=0,w=mδwratr=b,formulae-sequence𝑢0formulae-sequence𝑤𝑚𝛿subscript𝑤𝑟at𝑟𝑏u=0,\quad w=m\delta w_{r}\quad\text{at}\quad r=b,italic_u = 0 , italic_w = italic_m italic_δ italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT at italic_r = italic_b , (10)
wr=0,p=m(r1ηSzz),u=St+wSzatr=S,formulae-sequencesubscript𝑤𝑟0formulae-sequence𝑝𝑚superscript𝑟1𝜂subscript𝑆𝑧𝑧formulae-sequence𝑢subscript𝑆𝑡𝑤subscript𝑆𝑧at𝑟𝑆w_{r}=0,\quad p=m\left(r^{-1}-\eta S_{zz}\right),\quad u=S_{t}+wS_{z}\quad% \text{at}\quad r=S,italic_w start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 0 , italic_p = italic_m ( italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_η italic_S start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ) , italic_u = italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_w italic_S start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT at italic_r = italic_S , (11)

where Ro=Ω22/(g)𝑅𝑜superscriptΩ2superscript2𝑔Ro=\Omega^{2}\mathcal{H}^{2}/\left(g\mathcal{L}\right)italic_R italic_o = roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_g caligraphic_L ) is the rotation number, δ=δ^/𝛿^𝛿\delta=\widehat{\delta}/\mathcal{H}italic_δ = over^ start_ARG italic_δ end_ARG / caligraphic_H is the dimensionless slip parameter and η=ϵ2𝜂superscriptitalic-ϵ2\eta=\epsilon^{2}italic_η = italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. To derive (9)-(11), we have considered Ro𝑅𝑜Roitalic_R italic_o and δ𝛿\deltaitalic_δ are of order unity.

We solve the system of equations (9)-(11) and obtain the leading order solutions as

w=(1pz)[S22ln(rb)+b2r24+mδ2b(S2b2)],p=m(1SηSzz)b2Roln(Sr).formulae-sequence𝑤1subscript𝑝𝑧delimited-[]superscript𝑆22𝑟𝑏superscript𝑏2superscript𝑟24𝑚𝛿2𝑏superscript𝑆2superscript𝑏2𝑝𝑚1𝑆𝜂subscript𝑆𝑧𝑧superscript𝑏2𝑅𝑜𝑆𝑟w=\left(1-p_{z}\right)\left[\frac{S^{2}}{2}\ln\left(\frac{r}{b}\right)+\frac{b% ^{2}-r^{2}}{4}+\frac{m\delta}{2b}\left(S^{2}-b^{2}\right)\right],\quad p=m% \left(\frac{1}{S}-\eta S_{zz}\right)-b^{2}Ro\ln\left(\frac{S}{r}\right).italic_w = ( 1 - italic_p start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) [ divide start_ARG italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG roman_ln ( divide start_ARG italic_r end_ARG start_ARG italic_b end_ARG ) + divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG + divide start_ARG italic_m italic_δ end_ARG start_ARG 2 italic_b end_ARG ( italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ] , italic_p = italic_m ( divide start_ARG 1 end_ARG start_ARG italic_S end_ARG - italic_η italic_S start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ) - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R italic_o roman_ln ( divide start_ARG italic_S end_ARG start_ARG italic_r end_ARG ) . (12)

When the liquid flows along the outer surface of the slippery fiber (m=1)𝑚1(m=1)( italic_m = 1 ) in the absence of rotation (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), then (12) agrees well with Ji et al. [19].

Defining the flow rate q𝑞qitalic_q as q(S)=bSrw𝑑r𝑞𝑆superscriptsubscript𝑏𝑆𝑟𝑤differential-d𝑟q(S)=\int_{b}^{S}rwdritalic_q ( italic_S ) = ∫ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT italic_r italic_w italic_d italic_r, we obtain the flow rate as given below

q=(1pz)16[4S4ln(Sb)3S4+4S2b2b4+4mδb(S2b2)2].𝑞1subscript𝑝𝑧16delimited-[]4superscript𝑆4𝑆𝑏3superscript𝑆44superscript𝑆2superscript𝑏2superscript𝑏44𝑚𝛿𝑏superscriptsuperscript𝑆2superscript𝑏22q=\frac{\left(1-p_{z}\right)}{16}\left[4S^{4}\ln\left(\frac{S}{b}\right)-3S^{4% }+4S^{2}b^{2}-b^{4}+\frac{4m\delta}{b}\left(S^{2}-b^{2}\right)^{2}\right].italic_q = divide start_ARG ( 1 - italic_p start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) end_ARG start_ARG 16 end_ARG [ 4 italic_S start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_ln ( divide start_ARG italic_S end_ARG start_ARG italic_b end_ARG ) - 3 italic_S start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 4 italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + divide start_ARG 4 italic_m italic_δ end_ARG start_ARG italic_b end_ARG ( italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (13)

Equation (13) demonstrates that the influence of rotation affects the flow rate via the pressure solution.

Considering α=1/b𝛼1𝑏\alpha=1/bitalic_α = 1 / italic_b, we obtain the interfacial evolution equation from the mass conservative form of the kinematic boundary condition (7) as

(h+mα2h2)t+[(h)(1[𝒵(h)ηhzz]z)]z=0,subscript𝑚𝛼2superscript2𝑡subscriptdelimited-[]1subscriptdelimited-[]𝒵𝜂subscript𝑧𝑧𝑧𝑧0\left(h+\frac{m\alpha}{2}h^{2}\right)_{t}+\left[\mathcal{M}(h)\left(1-\left[% \mathcal{Z}(h)-\eta h_{zz}\right]_{z}\right)\right]_{z}=0,( italic_h + divide start_ARG italic_m italic_α end_ARG start_ARG 2 end_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + [ caligraphic_M ( italic_h ) ( 1 - [ caligraphic_Z ( italic_h ) - italic_η italic_h start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) ] start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 , (14a)
where the expressions of 𝒵(h)𝒵\mathcal{Z}(h)caligraphic_Z ( italic_h ) and the mobility function (h)\mathcal{M}(h)caligraphic_M ( italic_h ) are given below:
𝒵(h)=mα1+mαhRoα2ln(1+mαh),𝒵𝑚𝛼1𝑚𝛼𝑅𝑜superscript𝛼21𝑚𝛼\mathcal{Z}(h)=\frac{m\alpha}{1+m\alpha h}-\frac{Ro}{\alpha^{2}}\ln\left(1+m% \alpha h\right),caligraphic_Z ( italic_h ) = divide start_ARG italic_m italic_α end_ARG start_ARG 1 + italic_m italic_α italic_h end_ARG - divide start_ARG italic_R italic_o end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_ln ( 1 + italic_m italic_α italic_h ) , (14b)
and
(h;α,δ)=h33ϕ(mαh)+h24(2+mαh)2δ.𝛼𝛿superscript33italic-ϕ𝑚𝛼superscript24superscript2𝑚𝛼2𝛿\mathcal{M}(h;\alpha,\delta)=\frac{h^{3}}{3}\phi(m\alpha h)+\frac{h^{2}}{4}(2+% m\alpha h)^{2}\delta.caligraphic_M ( italic_h ; italic_α , italic_δ ) = divide start_ARG italic_h start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG italic_ϕ ( italic_m italic_α italic_h ) + divide start_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ( 2 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ . (14c)
In (14c), ϕitalic-ϕ\phiitalic_ϕ is the shape factor and has the following form
ϕ(Y)=316Y3([4ln(1+Y)3](1+Y)4+4(1+Y)21).italic-ϕ𝑌316superscript𝑌3delimited-[]41𝑌3superscript1𝑌44superscript1𝑌21\phi(Y)=\frac{3}{16Y^{3}}\left(\left[4\ln(1+Y)-3\right](1+Y)^{4}+4(1+Y)^{2}-1% \right).italic_ϕ ( italic_Y ) = divide start_ARG 3 end_ARG start_ARG 16 italic_Y start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ( [ 4 roman_ln ( 1 + italic_Y ) - 3 ] ( 1 + italic_Y ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 4 ( 1 + italic_Y ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) . (14d)

When Y0𝑌0Y\rightarrow 0italic_Y → 0, we have ϕ(Y)=1+Y+(3/20)Y2+O(Y3)italic-ϕ𝑌1𝑌320superscript𝑌2𝑂superscript𝑌3\phi(Y)=1+Y+(3/20)Y^{2}+O\left(Y^{3}\right)italic_ϕ ( italic_Y ) = 1 + italic_Y + ( 3 / 20 ) italic_Y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_O ( italic_Y start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ).

Equation (14) represents a fourth-order nonlinear partial differential equation (PDE) governing the evolution of the thickness h(z,t)𝑧𝑡h(z,t)italic_h ( italic_z , italic_t ). This model incorporates the influences of surface tension, gravity, and azimuthal instabilities while disregarding the effects of inertia and streamwise viscous dissipation [40]. When the cylinder is fixed (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), and the thin film flows over the outer surface of a non-slippery (m=1,δ=0)formulae-sequence𝑚1𝛿0(m=1,\delta=0)( italic_m = 1 , italic_δ = 0 ) vertical cylinder, (14) coincides with that obtained by provided by Craster & Matar [10]. It is important to note that various forms of 𝒵(h)𝒵\mathcal{Z}(h)caligraphic_Z ( italic_h ) are used in the literature to represent the azimuthal curvature of the film [41, 10, 11]. For a more comprehensive discussion on the different forms of 𝒵(h)𝒵\mathcal{Z}(h)caligraphic_Z ( italic_h ), we refer to the study by Ji et al. [19].

If we rescale the time scale as tt/ϕ(mα)𝑡𝑡italic-ϕ𝑚𝛼t\rightarrow t/\phi(m\alpha)italic_t → italic_t / italic_ϕ ( italic_m italic_α ), then the mobility function (14c) takes the form

(h;α,δ)=h3ϕ(mαh)3ϕ(mα)+h2(2+mαh)24ϕ(mα)δ.𝛼𝛿superscript3italic-ϕ𝑚𝛼3italic-ϕ𝑚𝛼superscript2superscript2𝑚𝛼24italic-ϕ𝑚𝛼𝛿\mathcal{M}(h;\alpha,\delta)=\frac{h^{3}\phi(m\alpha h)}{3\phi(m\alpha)}+\frac% {h^{2}(2+m\alpha h)^{2}}{4\phi(m\alpha)}\delta.caligraphic_M ( italic_h ; italic_α , italic_δ ) = divide start_ARG italic_h start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_ϕ ( italic_m italic_α italic_h ) end_ARG start_ARG 3 italic_ϕ ( italic_m italic_α ) end_ARG + divide start_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 2 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_ϕ ( italic_m italic_α ) end_ARG italic_δ . (15)

In (15), setting m=1𝑚1m=1italic_m = 1 makes the mobility function (h)\mathcal{M}(h)caligraphic_M ( italic_h ) exactly match the one found in [19].

In this model, we mainly focus on two parameters: wall slip (δ)𝛿(\delta)( italic_δ ) and rotation (Ro)𝑅𝑜(Ro)( italic_R italic_o ). We estimate these parameters’ ranges to understand better the significance of the physical mechanisms involved and their interaction. For applications involving flows over hydrophobic surfaces with microscopic texture, the slip length can reach up to the scale of micrometers [42], typically ranging from tens of micrometers to approximately 1 mm [43]. Consequently, the scaled slip length, δ𝛿\deltaitalic_δ, would range up to O(101)𝑂superscript101O\left(10^{-1}\right)italic_O ( 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ), consistent with values indicated by Chattopadhyay & Ji [44] and Ding et al. [45]. This order of magnitude for δ𝛿\deltaitalic_δ also applies when the substrate is made of a porous material. Here, κ/BJ𝜅subscript𝐵𝐽\sqrt{\kappa}/B_{J}square-root start_ARG italic_κ end_ARG / italic_B start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT gives the dimensional slip length, where κ𝜅\kappaitalic_κ is the permeability and BJsubscript𝐵𝐽B_{J}italic_B start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT is the Beavers-Joseph constant. According to Sadiq et al. [46], with κ=8.58×105𝜅8.58superscript105\sqrt{\kappa}=8.58\times 10^{-5}square-root start_ARG italic_κ end_ARG = 8.58 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT m and =2.4×1032.4superscript103\mathcal{H}=2.4\times 10^{-3}caligraphic_H = 2.4 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT m, , and considering the range of BJsubscript𝐵𝐽B_{J}italic_B start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, the dimensionless slip length ranges from 0.0089 to 0.3575. On the other hand, Ji et al. [19] found the slip length to be in the range of 0.012-0.014, considering the fiber radius b𝑏bitalic_b and \mathcal{H}caligraphic_H within 0.10.2150.10.2150.1-0.2150.1 - 0.215 mm and 0.4940.5890.4940.5890.494-0.5890.494 - 0.589 mm, respectively. However, they considered the slip length from 0 to 1 for theoretical discussions. Similarly, for thin film flows along the outer surface of a porous vertical fiber, Ding & Liu [47] estimated the dimensionless permeability parameter to be between 0.1 and 0.4. Therefore, we consider the dimensionless slip length δ𝛿\deltaitalic_δ to range from 0.1 to 0.4 in the present model.

Next, let us determine the range for the second key parameter Ro𝑅𝑜Roitalic_R italic_o, defined as Ro=Ω22/(g)𝑅𝑜superscriptΩ2superscript2𝑔Ro=\Omega^{2}\mathcal{H}^{2}/\left(g\mathcal{L}\right)italic_R italic_o = roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_g caligraphic_L ). In our model, we define \mathcal{L}caligraphic_L, the length scale in the streamwise direction, as =/ϵ=σ/(ρg)italic-ϵ𝜎𝜌𝑔\mathcal{L}=\mathcal{H}/\epsilon=\sigma/\left(\rho g\mathcal{H}\right)caligraphic_L = caligraphic_H / italic_ϵ = italic_σ / ( italic_ρ italic_g caligraphic_H ). This implies ϵ=ρg2/σ=Boitalic-ϵ𝜌𝑔superscript2𝜎𝐵𝑜\epsilon=\rho g\mathcal{H}^{2}/\sigma=Boitalic_ϵ = italic_ρ italic_g caligraphic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_σ = italic_B italic_o, where Bo𝐵𝑜Boitalic_B italic_o represents the Bond number. Typically, Bo𝐵𝑜Boitalic_B italic_o is reported to be small in experiments, approximately 0.3 according to Craster & Matar [10] for a liquid film coating along a vertical fiber. Experimental data from Tan et al. [48] for silicon oil indicates ρ=103𝜌superscript103\rho=10^{3}italic_ρ = 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT kg m-3 and σ=20.8×103𝜎20.8superscript103\sigma=20.8\times 10^{-3}italic_σ = 20.8 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT kg s-2. We take =104superscript104\mathcal{H}=10^{-4}caligraphic_H = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT m and Ω=2πfΩ2𝜋𝑓\Omega=2\pi froman_Ω = 2 italic_π italic_f (with f𝑓fitalic_f denoting the frequency in Hertz). With ϵ=0.3italic-ϵ0.3\epsilon=0.3italic_ϵ = 0.3 and using these values, the value of Ro𝑅𝑜Roitalic_R italic_o is approximately Ro0.0001207291f2𝑅𝑜0.0001207291superscript𝑓2Ro\approx 0.0001207291f^{2}italic_R italic_o ≈ 0.0001207291 italic_f start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Allowing f𝑓fitalic_f to vary from 3 to 30, Ro𝑅𝑜Roitalic_R italic_o will be approximately 0.001, 0.003, 0.011, 0.027, and 0.108 for f=3,5,10,15,30𝑓35101530f=3,5,10,15,30italic_f = 3 , 5 , 10 , 15 , 30, respectively. Therefore, for our present model, we will consider Ro𝑅𝑜Roitalic_R italic_o to be in the range of 0.001 to 0.14 for theoretical discussion.

III Linear stability analysis

In this section, we will perform the linear stability analysis of the problem to get a first insight into the underlying dynamics. We perturb the PDE (14) as follows:

h=1+h^=1+ζexp[ikz+Λt],1^1𝜁𝑖𝑘𝑧Λ𝑡h=1+\widehat{h}=1+\zeta\exp\left[ikz+\Lambda t\right],italic_h = 1 + over^ start_ARG italic_h end_ARG = 1 + italic_ζ roman_exp [ italic_i italic_k italic_z + roman_Λ italic_t ] , (16)

where k𝑘kitalic_k is the wavenumber, ΛΛ\Lambda\in\mathbb{C}roman_Λ ∈ blackboard_C is the growth rate of the perturbation, i=1𝑖1i=\sqrt{-1}italic_i = square-root start_ARG - 1 end_ARG and ζ1much-less-than𝜁1\zeta\ll 1italic_ζ ≪ 1 is the disturbance amplitude.

We substitute (16) in (14) and ignore the higher-order terms in ζ𝜁\zetaitalic_ζ. Consequently, we obtain the dispersion relation as follows

Λ=ck(α,δ)ik+k23(1+mα)(α2(1+mα)2+mRoα(1+mα)ηk2)χ(α,δ),Λsubscript𝑐𝑘𝛼𝛿𝑖𝑘superscript𝑘231𝑚𝛼superscript𝛼2superscript1𝑚𝛼2𝑚𝑅𝑜𝛼1𝑚𝛼𝜂superscript𝑘2𝜒𝛼𝛿\Lambda=-c_{k}\left(\alpha,\delta\right)ik+\frac{k^{2}}{3\left(1+m\alpha\right% )}\left(\frac{\alpha^{2}}{\left(1+m\alpha\right)^{2}}+\frac{mRo}{\alpha\left(1% +m\alpha\right)}-\eta k^{2}\right)\chi(\alpha,\delta),roman_Λ = - italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ) italic_i italic_k + divide start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 ( 1 + italic_m italic_α ) end_ARG ( divide start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_m italic_R italic_o end_ARG start_ARG italic_α ( 1 + italic_m italic_α ) end_ARG - italic_η italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_χ ( italic_α , italic_δ ) , (17a)
ck(α,δ)=1(1+mα)(ϕ(mα)+mαϕ(mα)3)+(2+mα)δ,χ(α,δ)=ϕ(mα)+34(2+mα)2δ.formulae-sequencesubscript𝑐𝑘𝛼𝛿11𝑚𝛼italic-ϕ𝑚𝛼𝑚𝛼superscriptitalic-ϕ𝑚𝛼32𝑚𝛼𝛿𝜒𝛼𝛿italic-ϕ𝑚𝛼34superscript2𝑚𝛼2𝛿c_{k}\left(\alpha,\delta\right)=\frac{1}{\left(1+m\alpha\right)}\left(\phi(m% \alpha)+\frac{m\alpha\phi^{\prime}(m\alpha)}{3}\right)+(2+m\alpha)\delta,\quad% \chi(\alpha,\delta)=\phi(m\alpha)+\frac{3}{4}(2+m\alpha)^{2}\delta.italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ) = divide start_ARG 1 end_ARG start_ARG ( 1 + italic_m italic_α ) end_ARG ( italic_ϕ ( italic_m italic_α ) + divide start_ARG italic_m italic_α italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_m italic_α ) end_ARG start_ARG 3 end_ARG ) + ( 2 + italic_m italic_α ) italic_δ , italic_χ ( italic_α , italic_δ ) = italic_ϕ ( italic_m italic_α ) + divide start_ARG 3 end_ARG start_ARG 4 end_ARG ( 2 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ . (17b)

While linearizing (14), if we use the mobility function (h)\mathcal{M}(h)caligraphic_M ( italic_h ) given by (15) instead of (14c), then the expressions of ck(α,δ)subscript𝑐𝑘𝛼𝛿c_{k}\left(\alpha,\delta\right)italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ) and χ(α,δ)𝜒𝛼𝛿\chi(\alpha,\delta)italic_χ ( italic_α , italic_δ ) in (17b) will take the following form

ck(α,δ)=1(1+mα)(1+mαϕ(mα)3ϕ(mα))+(2+mα)ϕ(mα)δ,χ(α,δ)=1+3(2+mα)24ϕ(mα)δ.formulae-sequencesubscript𝑐𝑘𝛼𝛿11𝑚𝛼1𝑚𝛼superscriptitalic-ϕ𝑚𝛼3italic-ϕ𝑚𝛼2𝑚𝛼italic-ϕ𝑚𝛼𝛿𝜒𝛼𝛿13superscript2𝑚𝛼24italic-ϕ𝑚𝛼𝛿c_{k}\left(\alpha,\delta\right)=\frac{1}{\left(1+m\alpha\right)}\left(1+\frac{% m\alpha\phi^{\prime}(m\alpha)}{3\phi(m\alpha)}\right)+\frac{\left(2+m\alpha% \right)}{\phi(m\alpha)}\delta,\quad\chi(\alpha,\delta)=1+\frac{3\left(2+m% \alpha\right)^{2}}{4\phi(m\alpha)}\delta.italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ) = divide start_ARG 1 end_ARG start_ARG ( 1 + italic_m italic_α ) end_ARG ( 1 + divide start_ARG italic_m italic_α italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_m italic_α ) end_ARG start_ARG 3 italic_ϕ ( italic_m italic_α ) end_ARG ) + divide start_ARG ( 2 + italic_m italic_α ) end_ARG start_ARG italic_ϕ ( italic_m italic_α ) end_ARG italic_δ , italic_χ ( italic_α , italic_δ ) = 1 + divide start_ARG 3 ( 2 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_ϕ ( italic_m italic_α ) end_ARG italic_δ . (18)

When the cylinder is stationary (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ) and the flow is on the outer side of cylinder (m=1)𝑚1(m=1)( italic_m = 1 ), the dispersion relation (17a) with (18) agrees well with the dispersion relation obtained by Ji et al. [19]. We note that to derive the dispersion relation (17a), we have considered the uniform film solution h=11h=1italic_h = 1 in (16). One can also repeat the above analysis for a uniform thin layer of arbitrary thickness h=h0subscript0h=h_{0}italic_h = italic_h start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [19].

We multiply (17) by i𝑖iitalic_i and use the scaling iΛck(ck/𝒴)1/3Λ~𝑖Λsubscript𝑐𝑘superscriptsubscript𝑐𝑘𝒴13~Λi\Lambda\rightarrow c_{k}\left(c_{k}/\mathcal{Y}\right)^{1/3}\widetilde{\Lambda}italic_i roman_Λ → italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / caligraphic_Y ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT over~ start_ARG roman_Λ end_ARG and k(ck/𝒴)1/3k~𝑘superscriptsubscript𝑐𝑘𝒴13~𝑘k\rightarrow\left(c_{k}/\mathcal{Y}\right)^{1/3}\widetilde{k}italic_k → ( italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / caligraphic_Y ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT over~ start_ARG italic_k end_ARG, where 𝒴=ηχ(α,δ)/(1+mα)𝒴𝜂𝜒𝛼𝛿1𝑚𝛼\mathcal{Y}=\eta\chi(\alpha,\delta)/\left(1+m\alpha\right)caligraphic_Y = italic_η italic_χ ( italic_α , italic_δ ) / ( 1 + italic_m italic_α ). Consequently, (17) is transformed to the standard form [20]

Λ~=k~+ik~23(β^k~2),~Λ~𝑘𝑖superscript~𝑘23^𝛽superscript~𝑘2\widetilde{\Lambda}=\widetilde{k}+\frac{i\widetilde{k}^{2}}{3}\left(\widehat{% \beta}-\widetilde{k}^{2}\right),over~ start_ARG roman_Λ end_ARG = over~ start_ARG italic_k end_ARG + divide start_ARG italic_i over~ start_ARG italic_k end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG ( over^ start_ARG italic_β end_ARG - over~ start_ARG italic_k end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (19a)
β^=αη(1+mα)(α(1+mα)+mRoα2)(ck𝒴)2/3.^𝛽𝛼𝜂1𝑚𝛼𝛼1𝑚𝛼𝑚𝑅𝑜superscript𝛼2superscriptsubscript𝑐𝑘𝒴23\widehat{\beta}=\frac{\alpha}{\eta\left(1+m\alpha\right)}\left(\frac{\alpha}{% \left(1+m\alpha\right)}+\frac{mRo}{\alpha^{2}}\right)\left(\frac{c_{k}}{% \mathcal{Y}}\right)^{-2/3}.over^ start_ARG italic_β end_ARG = divide start_ARG italic_α end_ARG start_ARG italic_η ( 1 + italic_m italic_α ) end_ARG ( divide start_ARG italic_α end_ARG start_ARG ( 1 + italic_m italic_α ) end_ARG + divide start_ARG italic_m italic_R italic_o end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ( divide start_ARG italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG caligraphic_Y end_ARG ) start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT . (19b)
According to Duprat et al. [49], the system’s instability becomes absolute when β^>β^ca[(9/4)×(17+77)]1/31.507^𝛽subscript^𝛽𝑐𝑎superscriptdelimited-[]941777131.507\widehat{\beta}>\widehat{\beta}_{ca}\equiv\left[(9/4)\times\left(-17+7\sqrt{7}% \right)\right]^{1/3}\approx 1.507over^ start_ARG italic_β end_ARG > over^ start_ARG italic_β end_ARG start_POSTSUBSCRIPT italic_c italic_a end_POSTSUBSCRIPT ≡ [ ( 9 / 4 ) × ( - 17 + 7 square-root start_ARG 7 end_ARG ) ] start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT ≈ 1.507. Therefore, we obtain the following condition on the rotation number Ro𝑅𝑜Roitalic_R italic_o
mRo>α[η(1+mα)β^ca(ck𝒴)2/3α2(1+mα)],𝑚𝑅𝑜𝛼delimited-[]𝜂1𝑚𝛼subscript^𝛽𝑐𝑎superscriptsubscript𝑐𝑘𝒴23superscript𝛼21𝑚𝛼mRo>\alpha\left[\eta\left(1+m\alpha\right)\widehat{\beta}_{ca}\left(\frac{c_{k% }}{\mathcal{Y}}\right)^{2/3}-\frac{\alpha^{2}}{\left(1+m\alpha\right)}\right],italic_m italic_R italic_o > italic_α [ italic_η ( 1 + italic_m italic_α ) over^ start_ARG italic_β end_ARG start_POSTSUBSCRIPT italic_c italic_a end_POSTSUBSCRIPT ( divide start_ARG italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG caligraphic_Y end_ARG ) start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT - divide start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + italic_m italic_α ) end_ARG ] , (19c)
above which the system’s instability is absolute.

The real part of ΛΛ\Lambdaroman_Λ, Λr=Re(Λ)subscriptΛ𝑟ReΛ\Lambda_{r}=\text{Re}(\Lambda)roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = Re ( roman_Λ ), describes the effective growth rate for the system. For Λr<0subscriptΛ𝑟0\Lambda_{r}<0roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT < 0, the amplitude of the linear growth rate of perturbation decreases, and the system is stable. On the other hand, for Λr>0subscriptΛ𝑟0\Lambda_{r}>0roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0, the amplitude of the linear growth rate of perturbation increases, and the system is unstable. The expression of ΛrsubscriptΛ𝑟\Lambda_{r}roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT for the present study is given by

Λr=k23(1+mα)[α2(1+mα)2+mRoα(1+mα)ηk2]χ(α,δ),subscriptΛ𝑟superscript𝑘231𝑚𝛼delimited-[]superscript𝛼2superscript1𝑚𝛼2𝑚𝑅𝑜𝛼1𝑚𝛼𝜂superscript𝑘2𝜒𝛼𝛿\Lambda_{r}=\frac{k^{2}}{3\left(1+m\alpha\right)}\left[\frac{\alpha^{2}}{\left% (1+m\alpha\right)^{2}}+\frac{mRo}{\alpha\left(1+m\alpha\right)}-\eta k^{2}% \right]\chi(\alpha,\delta),roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = divide start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 ( 1 + italic_m italic_α ) end_ARG [ divide start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_m italic_R italic_o end_ARG start_ARG italic_α ( 1 + italic_m italic_α ) end_ARG - italic_η italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_χ ( italic_α , italic_δ ) , (19d)

where χ(α,δ)𝜒𝛼𝛿\chi(\alpha,\delta)italic_χ ( italic_α , italic_δ ) is given by (18). Equation (19d) shows that the effective growth rate ΛrsubscriptΛ𝑟\Lambda_{r}roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is influenced by both wall slip δ𝛿\deltaitalic_δ and rotation number Ro𝑅𝑜Roitalic_R italic_o.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Figure 2: Variation of ΛrsubscriptΛ𝑟\Lambda_{r}roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT with k𝑘kitalic_k under different δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o, with α=1/6𝛼16\alpha=1/6italic_α = 1 / 6, η=0.04𝜂0.04\eta=0.04italic_η = 0.04. Top panel: Flow along outer surface (m=1𝑚1m=1italic_m = 1), Bottom: Flow along inner surface (m=1𝑚1m=-1italic_m = - 1)

In FIG. 2, we present the linear growth rate curves as functions of the wavenumber k𝑘kitalic_k. The top panel shows the flow along the outer surface of the cylinder, while the bottom panel illustrates the flow along the inner surface. To analyze the effect of wall slippage, we select three typical δ𝛿\deltaitalic_δ values: δ=0,0.1𝛿00.1\delta=0,0.1italic_δ = 0 , 0.1, and 0.20.20.20.2. Our observations indicate that wall slippage promotes flow instability for the cylinder’s outer and inner surfaces. This destabilizing influence of wall slippage aligns with findings from previous studies by Ji et al. [19], Chao et al. [30], Chattopadhyay [31], and Schwitzerlett et al. [32]. Further, FIGs. 2b and 2c illustrate the impact of rotation on the flow along the outer surface of the cylinder. To explore the effects of rotation, we select two typical Ro𝑅𝑜Roitalic_R italic_o values: Ro=0.002𝑅𝑜0.002Ro=0.002italic_R italic_o = 0.002 and 0.0040.0040.0040.004. By comparing FIGs. 2a to 2c, we observe that as Ro𝑅𝑜Roitalic_R italic_o increases, flow instability also increases, indicating that rotation destabilizes the flow along the outer surface of the cylinder. This finding aligns with the results presented by Liu & Ding [36] and Mukhopadhyay et al. [38]. In contrast, FIGs. 2e and 2f examine the effect of rotation on the flow along the inner surface of the cylinder, with all parameters matching those in the top panel of FIG. 2. Comparing FIGs. 2d to 2f reveals that increasing Ro𝑅𝑜Roitalic_R italic_o decreases flow instability, demonstrating that rotation stabilizes the flow along the inner surface of the cylinder. Thus, our linear stability analysis concludes that rotation exerts completely different influences depending on whether the film flows along the outer or inner surface of the cylinder/fiber. Physically, when the film flows along the outer surface, the centrifugal force destabilizes the flow by pushing fluid away from the cylinder wall. Conversely, when the film flows along the inner surface, the centrifugal force stabilizes the flow by drawing fluid toward the cylinder wall. Moreover, the linear growth rate profiles show that for given δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o, a critical wavenumber exists below which instability increases, after which it decreases.

From (19d), we observe that if

α3+mRo(1+mα)αη(1+mα)2ηk2,superscript𝛼3𝑚𝑅𝑜1𝑚𝛼𝛼𝜂superscript1𝑚𝛼2𝜂superscript𝑘2\frac{\alpha^{3}+mRo\left(1+m\alpha\right)}{\alpha\eta\left(1+m\alpha\right)^{% 2}}\leq\eta k^{2},divide start_ARG italic_α start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_m italic_R italic_o ( 1 + italic_m italic_α ) end_ARG start_ARG italic_α italic_η ( 1 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ≤ italic_η italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (19e)

the long-wave instability can be completely impeded by rotation. Consequently, we can define a sufficient condition for the system to be stable in the long-wave range as

mRoα3(1+mα).𝑚𝑅𝑜superscript𝛼31𝑚𝛼mRo\leq-\frac{\alpha^{3}}{\left(1+m\alpha\right)}.italic_m italic_R italic_o ≤ - divide start_ARG italic_α start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + italic_m italic_α ) end_ARG . (19f)

We further obtain the cut-off wavenumber (k=kc)𝑘subscript𝑘𝑐\left(k=k_{c}\right)( italic_k = italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) and the most unstable mode (k=km)𝑘subscript𝑘𝑚\left(k=k_{m}\right)( italic_k = italic_k start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) as

kc=α3+mRo(1+mα)αη(1+mα)2,km=kc2.formulae-sequencesubscript𝑘𝑐superscript𝛼3𝑚𝑅𝑜1𝑚𝛼𝛼𝜂superscript1𝑚𝛼2subscript𝑘𝑚subscript𝑘𝑐2k_{c}=\sqrt{\frac{\alpha^{3}+mRo\left(1+m\alpha\right)}{\alpha\eta\left(1+m% \alpha\right)^{2}}},\quad k_{m}=\frac{k_{c}}{\sqrt{2}}.italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_α start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_m italic_R italic_o ( 1 + italic_m italic_α ) end_ARG start_ARG italic_α italic_η ( 1 + italic_m italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , italic_k start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = divide start_ARG italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG . (19g)

Equation (19g) demonstrates that the slip length δ𝛿\deltaitalic_δ does not influence the critical wavenumber kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for both outer and inner surface flow of the vertical cylinder, and consequently kmsubscript𝑘𝑚k_{m}italic_k start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. This observation aligns with the findings of Chao et al. [30], who studied the flow along the outer surface of a slippery cylinder under non-isothermal conditions. They reported that the critical wavenumber is independent of the slip length without thermal effects.

Refer to caption
(a)
Refer to caption
(b)
Figure 3: Influence of Ro𝑅𝑜Roitalic_R italic_o on the most unstable mode with η=0.01𝜂0.01\eta=0.01italic_η = 0.01. (a) Flow along outer surface (m=1𝑚1m=1italic_m = 1) and (b) Flow along inner surface (m=1𝑚1m=-1italic_m = - 1)

In FIG. 3, we illustrate how the fastest growing mode k=km𝑘subscript𝑘𝑚k=k_{m}italic_k = italic_k start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is influenced by rotation (Ro𝑅𝑜Roitalic_R italic_o) as the cylinder radius (b=1/α)𝑏1𝛼(b=1/\alpha)( italic_b = 1 / italic_α ) changes. FIG. 3a depicts the flow along the outer surface of the cylinder, while FIG. 3b shows the flow along the inner surface. The Ro𝑅𝑜Roitalic_R italic_o values are consistent with those used in FIG. 2. FIG. 3a demonstrates that as Ro𝑅𝑜Roitalic_R italic_o increases, the unstable zone expands, whereas FIG. 3b shows that as Ro𝑅𝑜Roitalic_R italic_o increases, the unstable zone contracts. These findings confirm the dual role of rotation Ro𝑅𝑜Roitalic_R italic_o, depending on whether the film flows along the outer or inner surface of the cylinder. Additionally, both figures reveal that the unstable region consistently enlarges with increasing α𝛼\alphaitalic_α. In other words, flow instability is enhanced as α𝛼\alphaitalic_α increases (i.e., the radius of the cylinder decreases). This observation aligns with previous studies by Craster & Matar [10] and Mukhopadhyay et al. [38]. This phenomenon occurs regardless of the presence of rotation.

The imaginary part of ΛΛ\Lambdaroman_Λ, Λi=Im(Λ)subscriptΛ𝑖ImΛ\Lambda_{i}=\text{Im}(\Lambda)roman_Λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = Im ( roman_Λ ), is Λi=kck(α,δ)subscriptΛ𝑖𝑘subscript𝑐𝑘𝛼𝛿\Lambda_{i}=-kc_{k}\left(\alpha,\delta\right)roman_Λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - italic_k italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ), where ck(α,δ)subscript𝑐𝑘𝛼𝛿c_{k}\left(\alpha,\delta\right)italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ) is given by (18). The rotation does not influence ΛisubscriptΛ𝑖\Lambda_{i}roman_Λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT but is affected by the wall slippage δ𝛿\deltaitalic_δ for outer and inner flow cases. We further obtain the linear wave speed clinsubscript𝑐linc_{\text{lin}}italic_c start_POSTSUBSCRIPT lin end_POSTSUBSCRIPT as clin=Λi/k=ck(α,δ)subscript𝑐linsubscriptΛ𝑖𝑘subscript𝑐𝑘𝛼𝛿c_{\text{lin}}=-\Lambda_{i}/k=c_{k}\left(\alpha,\delta\right)italic_c start_POSTSUBSCRIPT lin end_POSTSUBSCRIPT = - roman_Λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_k = italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_α , italic_δ ). Therefore, the linear wave speed depends only on the wall slippage, not the rotation. However, whether the rotation influences the nonlinear wave speed for outer and inner surface flows remains unknown. we will discuss the impact of rotation on nonlinear wave speeds in Section V.

IV Weakly nonlinear stability analysis

The linear study fails to capture the accurate flow behavior when the perturbed waves grow to a finite amplitude. Therefore, a weakly nonlinear stability analysis is necessary to determine the behavior of the nonlinear wave near criticality. In this section, we focus on understanding the influence of wall slippage and rotation on the weakly nonlinear stability of the model (14), following [51, 44, 52, 53, 54, 55, 50]. We express the PDE (14) as:

ht+γ1(h)hz+γ2(h)hzz+γ3(h)hzzzz+γ4(h)hz2+γ5(h)hzhzzz=0,subscript𝑡subscript𝛾1subscript𝑧subscript𝛾2subscript𝑧𝑧subscript𝛾3subscript𝑧𝑧𝑧𝑧subscript𝛾4superscriptsubscript𝑧2subscript𝛾5subscript𝑧subscript𝑧𝑧𝑧0h_{t}+\gamma_{1}(h)h_{z}+\gamma_{2}(h)h_{zz}+\gamma_{3}(h)h_{zzzz}+\gamma_{4}(% h)h_{z}^{2}+\gamma_{5}(h)h_{z}h_{zzz}=0,italic_h start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_h ) italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_h ) italic_h start_POSTSUBSCRIPT italic_z italic_z end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_h ) italic_h start_POSTSUBSCRIPT italic_z italic_z italic_z italic_z end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ) italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ( italic_h ) italic_h start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_z italic_z italic_z end_POSTSUBSCRIPT = 0 , (20a)
where the coefficients γi,i=1,,5formulae-sequencesubscript𝛾𝑖𝑖15\gamma_{i},i=1,\ldots,5italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_i = 1 , … , 5 are as follows:
γ1(h;α,δ)=αΘ16m(1+mαh),γ2(h;α,δ,Ro)=α2θ(1+mαh)2(Ro16α2+γ1Θ),γ3(h;α,δ)=γ1ηθΘ,formulae-sequencesubscript𝛾1𝛼𝛿𝛼Θ16𝑚1𝑚𝛼formulae-sequencesubscript𝛾2𝛼𝛿𝑅𝑜superscript𝛼2𝜃superscript1𝑚𝛼2𝑅𝑜16superscript𝛼2subscript𝛾1Θsubscript𝛾3𝛼𝛿subscript𝛾1𝜂𝜃Θ\gamma_{1}(h;\alpha,\delta)=\frac{\alpha\Theta}{16m(1+m\alpha h)},\quad\gamma_% {2}(h;\alpha,\delta,Ro)=\frac{\alpha^{2}\theta}{(1+m\alpha h)^{2}}\left(\frac{% Ro}{16\alpha^{2}}+\frac{\gamma_{1}}{\Theta}\right),\quad\gamma_{3}(h;\alpha,% \delta)=\frac{\gamma_{1}\eta\theta}{\Theta},italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_h ; italic_α , italic_δ ) = divide start_ARG italic_α roman_Θ end_ARG start_ARG 16 italic_m ( 1 + italic_m italic_α italic_h ) end_ARG , italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_h ; italic_α , italic_δ , italic_R italic_o ) = divide start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG ( 1 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_R italic_o end_ARG start_ARG 16 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG roman_Θ end_ARG ) , italic_γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_h ; italic_α , italic_δ ) = divide start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_η italic_θ end_ARG start_ARG roman_Θ end_ARG , (20b)
γ4(h;α,δ,Ro)=α2[Θ+mα(hΘθ)](1+mαh)3(Ro16α2+γ1Θ)α4θ16(1+mαh)4,γ5(h;α,δ)=ηγ1,formulae-sequencesubscript𝛾4𝛼𝛿𝑅𝑜superscript𝛼2delimited-[]Θ𝑚𝛼Θ𝜃superscript1𝑚𝛼3𝑅𝑜16superscript𝛼2subscript𝛾1Θsuperscript𝛼4𝜃16superscript1𝑚𝛼4subscript𝛾5𝛼𝛿𝜂subscript𝛾1\gamma_{4}(h;\alpha,\delta,Ro)=\frac{\alpha^{2}\left[\Theta+m\alpha\left(h% \Theta-\theta\right)\right]}{(1+m\alpha h)^{3}}\left(\frac{Ro}{16\alpha^{2}}+% \frac{\gamma_{1}}{\Theta}\right)-\frac{\alpha^{4}\theta}{16(1+m\alpha h)^{4}},% \quad\gamma_{5}(h;\alpha,\delta)=\eta\gamma_{1},italic_γ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ; italic_α , italic_δ , italic_R italic_o ) = divide start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ roman_Θ + italic_m italic_α ( italic_h roman_Θ - italic_θ ) ] end_ARG start_ARG ( 1 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG italic_R italic_o end_ARG start_ARG 16 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG roman_Θ end_ARG ) - divide start_ARG italic_α start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_θ end_ARG start_ARG 16 ( 1 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG , italic_γ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ( italic_h ; italic_α , italic_δ ) = italic_η italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , (20c)
θ(h;α,δ)=mαh(2+mαh)[mαh(2+mαh)(4mαδ3)2]+4(1+mαh)4ln(1+mαh)α4,𝜃𝛼𝛿𝑚𝛼2𝑚𝛼delimited-[]𝑚𝛼2𝑚𝛼4𝑚𝛼𝛿324superscript1𝑚𝛼41𝑚𝛼superscript𝛼4\theta(h;\alpha,\delta)=\frac{m\alpha h\left(2+m\alpha h\right)\left[m\alpha h% (2+m\alpha h)\left(4m\alpha\delta-3\right)-2\right]+4\left(1+m\alpha h\right)^% {4}\ln(1+m\alpha h)}{\alpha^{4}},italic_θ ( italic_h ; italic_α , italic_δ ) = divide start_ARG italic_m italic_α italic_h ( 2 + italic_m italic_α italic_h ) [ italic_m italic_α italic_h ( 2 + italic_m italic_α italic_h ) ( 4 italic_m italic_α italic_δ - 3 ) - 2 ] + 4 ( 1 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_ln ( 1 + italic_m italic_α italic_h ) end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG , (20d)
Θ(h;α,δ)=8(1+mαh)[mαh(2+mαh)(2αδm)+2m(1+mαh)2ln(1+mαh)]α3.Θ𝛼𝛿81𝑚𝛼delimited-[]𝑚𝛼2𝑚𝛼2𝛼𝛿𝑚2𝑚superscript1𝑚𝛼21𝑚𝛼superscript𝛼3\Theta(h;\alpha,\delta)=\frac{8(1+m\alpha h)\left[m\alpha h(2+m\alpha h)(2% \alpha\delta-m)+2m(1+m\alpha h)^{2}\ln(1+m\alpha h)\right]}{\alpha^{3}}.roman_Θ ( italic_h ; italic_α , italic_δ ) = divide start_ARG 8 ( 1 + italic_m italic_α italic_h ) [ italic_m italic_α italic_h ( 2 + italic_m italic_α italic_h ) ( 2 italic_α italic_δ - italic_m ) + 2 italic_m ( 1 + italic_m italic_α italic_h ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_ln ( 1 + italic_m italic_α italic_h ) ] end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG . (20e)

To investigate the nonlinear effect of the present model, we use the method of multiple scales as

tt+j=1ψjtj,zz+j=1ψjzj,h^=j=1ψjh^j.formulae-sequencesubscript𝑡subscript𝑡subscript𝑗1superscript𝜓𝑗subscriptsubscript𝑡𝑗formulae-sequencesubscript𝑧subscript𝑧subscript𝑗1superscript𝜓𝑗subscriptsubscript𝑧𝑗^subscript𝑗1superscript𝜓𝑗subscript^𝑗\partial_{t}\rightarrow\partial_{t}+\sum_{j=1}\psi^{j}\partial_{t_{j}},\quad% \partial_{z}\rightarrow\partial_{z}+\sum_{j=1}\psi^{j}\partial_{z_{j}},\quad% \widehat{h}=\sum_{j=1}\psi^{j}\widehat{h}_{j}.∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT → ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT , ∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT → ∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT , over^ start_ARG italic_h end_ARG = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT over^ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT . (21)

where ψ=kkc𝜓𝑘subscript𝑘𝑐\psi=k-k_{c}italic_ψ = italic_k - italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is a small perturbation parameter.

Inserting (16) in (20), considering terms up to O(h^3)𝑂superscript^3O\left(\widehat{h}^{3}\right)italic_O ( over^ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) and using (21) gives the following Landau-type equation for the perturbation amplitude ζ𝜁\zetaitalic_ζ [51, 44, 52, 53, 54, 55]:

ζt2+iJ0ζz1+J12ζz12ψ2Λrζ+(J2+iJ4)|ζ|2ζ=0,𝜁subscript𝑡2𝑖subscript𝐽0𝜁subscript𝑧1subscript𝐽1superscript2𝜁superscriptsubscript𝑧12superscript𝜓2subscriptΛ𝑟𝜁subscript𝐽2𝑖subscript𝐽4superscript𝜁2𝜁0\dfrac{\partial\zeta}{\partial t_{2}}+iJ_{0}\dfrac{\partial\zeta}{\partial z_{% 1}}+J_{1}\dfrac{\partial^{2}\zeta}{\partial z_{1}^{2}}-\psi^{-2}\Lambda_{r}% \zeta+\left(J_{2}+iJ_{4}\right)|\zeta|^{2}\zeta=0,divide start_ARG ∂ italic_ζ end_ARG start_ARG ∂ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + italic_i italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG ∂ italic_ζ end_ARG start_ARG ∂ italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG + italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ζ end_ARG start_ARG ∂ italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - italic_ψ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_ζ + ( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_i italic_J start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) | italic_ζ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ζ = 0 , (22)

We refer to the studies by Oron & Gottlieb [52] and Chattopadhyay et al. [56] to get a detailed derivation of (22) from (20). The coefficients J0subscript𝐽0J_{0}italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, J1subscript𝐽1J_{1}italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and J4subscript𝐽4J_{4}italic_J start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT are given below:

J0=2k(Γ12Γ3k2)ψ1,J1=Γ16Γ3k2,J2=k2(𝒱32Γ2′′+32Γ3′′k2+Γ4Γ5k2)Γ1,formulae-sequencesubscript𝐽02𝑘subscriptΓ12subscriptΓ3superscript𝑘2superscript𝜓1formulae-sequencesubscript𝐽1subscriptΓ16subscriptΓ3superscript𝑘2subscript𝐽2superscript𝑘2𝒱32superscriptsubscriptΓ2′′32superscriptsubscriptΓ3′′superscript𝑘2superscriptsubscriptΓ4superscriptsubscriptΓ5superscript𝑘2superscriptsubscriptΓ1J_{0}=2k\left(\Gamma_{1}-2\Gamma_{3}k^{2}\right)\psi^{-1},\quad J_{1}=\Gamma_{% 1}-6\Gamma_{3}k^{2},\quad J_{2}=k^{2}\left(\mathcal{E}\mathcal{V}-\frac{3}{2}% \Gamma_{2}^{\prime\prime}+\frac{3}{2}\Gamma_{3}^{\prime\prime}k^{2}+\Gamma_{4}% ^{\prime}-\Gamma_{5}^{\prime}k^{2}\right)-\Gamma_{1}^{\prime}\mathcal{F},italic_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 italic_k ( roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 2 roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_ψ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , italic_J start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 6 roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( caligraphic_E caligraphic_V - divide start_ARG 3 end_ARG start_ARG 2 end_ARG roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 2 end_ARG roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - roman_Γ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) - roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT caligraphic_F , (23a)
𝒱=5Γ2+17Γ3k2+4Γ410Γ5k2,J4=Γ1k+Γ1′′2k+𝒱k2,formulae-sequence𝒱5superscriptsubscriptΓ217superscriptsubscriptΓ3superscript𝑘24subscriptΓ410subscriptΓ5superscript𝑘2subscript𝐽4superscriptsubscriptΓ1𝑘superscriptsubscriptΓ1′′2𝑘𝒱superscript𝑘2\mathcal{V}=-5\Gamma_{2}^{\prime}+17\Gamma_{3}^{\prime}k^{2}+4\Gamma_{4}-10% \Gamma_{5}k^{2},\quad J_{4}=\Gamma_{1}^{\prime}k\mathcal{E}+\frac{\Gamma_{1}^{% \prime\prime}}{2}k+\mathcal{V}k^{2}\mathcal{F},caligraphic_V = - 5 roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + 17 roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 roman_Γ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT - 10 roman_Γ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_J start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k caligraphic_E + divide start_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_k + caligraphic_V italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_F , (23b)
=Γ2Γ3k2+Γ4Γ5k24(4Γ3k2Γ2),=Γ14(4Γ3k2Γ2).formulae-sequencesuperscriptsubscriptΓ2superscriptsubscriptΓ3superscript𝑘2subscriptΓ4subscriptΓ5superscript𝑘244subscriptΓ3superscript𝑘2subscriptΓ2superscriptsubscriptΓ144subscriptΓ3superscript𝑘2subscriptΓ2\mathcal{E}=\frac{\Gamma_{2}^{\prime}-\Gamma_{3}^{\prime}k^{2}+\Gamma_{4}-% \Gamma_{5}k^{2}}{4\left(4\Gamma_{3}k^{2}-\Gamma_{2}\right)},\quad\mathcal{F}=% \frac{-\Gamma_{1}^{\prime}}{4\left(4\Gamma_{3}k^{2}-\Gamma_{2}\right)}.caligraphic_E = divide start_ARG roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_Γ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 ( 4 roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG , caligraphic_F = divide start_ARG - roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG 4 ( 4 roman_Γ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG . (23c)
Here, the ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and its derivatives are evaluated from γisubscript𝛾𝑖\gamma_{i}italic_γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT at h=11h=1italic_h = 1.

When there is no spatial modulation (filtered wave), the equation (22) yields

ζt2ψ2Λrζ+(J2+iJ4)|ζ|2ζ=0.𝜁subscript𝑡2superscript𝜓2subscriptΛ𝑟𝜁subscript𝐽2𝑖subscript𝐽4superscript𝜁2𝜁0\dfrac{\partial\zeta}{\partial t_{2}}-\psi^{-2}\Lambda_{r}\zeta+\left(J_{2}+iJ% _{4}\right)|\zeta|^{2}\zeta=0.divide start_ARG ∂ italic_ζ end_ARG start_ARG ∂ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG - italic_ψ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_ζ + ( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_i italic_J start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) | italic_ζ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ζ = 0 . (24)

In this case, the solution for the perturbation amplitude ζ𝜁\zetaitalic_ζ can be expressed in a simple form as [52]: ζ=a(t2)exp[ib(t2)t2]𝜁𝑎subscript𝑡2𝑖𝑏subscript𝑡2subscript𝑡2\zeta=a\left(t_{2}\right)\exp\left[-ib\left(t_{2}\right)t_{2}\right]italic_ζ = italic_a ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) roman_exp [ - italic_i italic_b ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ]. We use this solution of ζ𝜁\zetaitalic_ζ in (24). Separating the real and imaginary parts, we obtain the following

at2=a(Λrψ2J2a2).𝑎subscript𝑡2𝑎subscriptΛ𝑟superscript𝜓2subscript𝐽2superscript𝑎2\frac{\partial a}{\partial t_{2}}=a\left(\frac{\Lambda_{r}}{\psi^{2}}-J_{2}a^{% 2}\right).divide start_ARG ∂ italic_a end_ARG start_ARG ∂ italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG = italic_a ( divide start_ARG roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG italic_ψ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (25)

In (25), the second term on the right-hand side is contributed by system nonlinearities. In the linear stability analysis, the exponential growth of the linear disturbance is decided by the sign of ΛrsubscriptΛ𝑟\Lambda_{r}roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. Beyond the linear regime, the perturbation dynamics depend solely on the sign of J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. A positive J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT indicates a supercritical bifurcation, while a negative J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT indicates a subcritical bifurcation. Therefore, based on the signs of ΛrsubscriptΛ𝑟\Lambda_{r}roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, we identify four zones: subcritical stable (J2<0,Λr<0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}<0,\Lambda_{r}<0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT < 0 ), subcritical unstable (J2<0,Λr>0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}<0,\Lambda_{r}>0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0 ), supercritical stable (J2>0,Λr<0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}>0,\Lambda_{r}<0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT < 0 ) and supercritical unstable (J2>0,Λr>0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}>0,\Lambda_{r}>0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0 ) [52, 46]. These zones are labeled 1, 2, 3, and 4. This discussion highlights that understanding the effects of nonlinear terms on the stability of the flow system requires characterizing various flow states using J2subscript𝐽2J_{2}italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, which can be determined through weakly nonlinear stability analysis.

Moreover, when a𝑎aitalic_a is independent of t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, we can obtain the amplitude for a saturated wave (threshold amplitude) from (25) as [38]

ψa=ΛrJ2.𝜓𝑎subscriptΛ𝑟subscript𝐽2\psi a=\sqrt{\frac{\Lambda_{r}}{J_{2}}}.italic_ψ italic_a = square-root start_ARG divide start_ARG roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG start_ARG italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_ARG . (26)
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 4: Effect of Ro𝑅𝑜Roitalic_R italic_o when the film flows along the outer surface of the cylinder (m=1)𝑚1(m=1)( italic_m = 1 ) for η=0.0225𝜂0.0225\eta=0.0225italic_η = 0.0225. (a-c) Neutral stability curve for (δ,Ro)=(0.2,0)𝛿𝑅𝑜0.20\left(\delta,Ro\right)=(0.2,0)( italic_δ , italic_R italic_o ) = ( 0.2 , 0 ) in (a); (δ,Ro)=(0.2,0.05)𝛿𝑅𝑜0.20.05\left(\delta,Ro\right)=(0.2,0.05)( italic_δ , italic_R italic_o ) = ( 0.2 , 0.05 ) in (b); (δ,Ro)=(0.2,0.1)𝛿𝑅𝑜0.20.1\left(\delta,Ro\right)=(0.2,0.1)( italic_δ , italic_R italic_o ) = ( 0.2 , 0.1 ) in (c). (d) Threshold amplitude in the supercritical unstable region with α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 and δ=0.2𝛿0.2\delta=0.2italic_δ = 0.2

FIG. 4 is plotted for thin film flows along the outer surface of the cylinder. FIG. 4a shows a stationary cylinder (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ) with wall slippage (δ=0.4)𝛿0.4(\delta=0.4)( italic_δ = 0.4 ). while FIG. 4b shows a rotating cylinder (Ro=0.02)𝑅𝑜0.02(Ro=0.02)( italic_R italic_o = 0.02 ) with the same slippage. Rotation expands the subcritical unstable zone (zone 2) and contracts the supercritical stable zone (zone 3). In FIG. 4c, increasing the rotation to Ro=0.04𝑅𝑜0.04Ro=0.04italic_R italic_o = 0.04 further amplifies zone 2 and reduces zone 3, confirming that rotation destabilizes the film flow, with instability increasing with wall slippage. In FIG. 4d, we illustrate the impact of rotation on the threshold amplitude in the supercritical unstable zone (J2>0,Λr>0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}>0,\Lambda_{r}>0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0 ) with a slippery cylinder wall (δ>0)𝛿0(\delta>0)( italic_δ > 0 ). We notice that the threshold amplitude rises with increased rotation. Moreover, for a fixed Ro𝑅𝑜Roitalic_R italic_o, the threshold amplitude initially increases until reaching a critical wavenumber k=k𝑘superscript𝑘k=k^{\dagger}italic_k = italic_k start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT. Beyond ksuperscript𝑘k^{\dagger}italic_k start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, the threshold amplitude decreases.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 5: Effect of Ro𝑅𝑜Roitalic_R italic_o when the film flows along the inner surface of the cylinder (m=1)𝑚1(m=-1)( italic_m = - 1 ) for η=0.0625𝜂0.0625\eta=0.0625italic_η = 0.0625. (a-c) Neutral stability curve for (δ,Ro)=(0.4,0.04)𝛿𝑅𝑜0.40.04\left(\delta,Ro\right)=(0.4,0.04)( italic_δ , italic_R italic_o ) = ( 0.4 , 0.04 ) in (a); (δ,Ro)=(0.4,0.06)𝛿𝑅𝑜0.40.06\left(\delta,Ro\right)=(0.4,0.06)( italic_δ , italic_R italic_o ) = ( 0.4 , 0.06 ) in (b); (δ,Ro)=(0.4,0.08)𝛿𝑅𝑜0.40.08\left(\delta,Ro\right)=(0.4,0.08)( italic_δ , italic_R italic_o ) = ( 0.4 , 0.08 ) in (c). (d) Threshold amplitude in the supercritical unstable region with α=1/2.6𝛼12.6\alpha=1/2.6italic_α = 1 / 2.6 and δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4

In FIG. 5, we illustrate the impact of rotation (Ro)𝑅𝑜(Ro)( italic_R italic_o ) on the supercritical and subcritical stability regions when liquid flows along the inner surface of a slippery (δ>0)𝛿0(\delta>0)( italic_δ > 0 ) cylinder. We set the slip length δ𝛿\deltaitalic_δ to 0.4 and select three typical Ro𝑅𝑜Roitalic_R italic_o values: Ro=0.04,0.06𝑅𝑜0.040.06Ro=0.04,0.06italic_R italic_o = 0.04 , 0.06 and 0.08. Comparing FIGs. 5a to 5c, we observe that as the Ro𝑅𝑜Roitalic_R italic_o value increases, zone 3 (supercritical stable) expands, while zone 2 (subcritical unstable) contracts. In the supercritical stable region (J2>0,Λr<0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}>0,\Lambda_{r}<0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT < 0 ), , finite-amplitude disturbances remain always stable. Conversely, in the subcritical unstable region (J2<0,Λr>0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}<0,\Lambda_{r}>0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0 ), both linear and nonlinear instabilities increase, making the system always unstable. FIGs. 5(a-c) demonstrate that zone 3, where finite-amplitude disturbances are always stable, enlarges with higher Ro𝑅𝑜Roitalic_R italic_o, whereas zone 2, where instability consistently increases, diminishes with higher Ro𝑅𝑜Roitalic_R italic_o. Thus, the stabilizing effect of Ro𝑅𝑜Roitalic_R italic_o is consistent with the results from linear stability analysis. In FIG. 5d, we demonstrate how rotation affects the threshold amplitude in the supercritical unstable zone (J2>0,Λr>0)formulae-sequencesubscript𝐽20subscriptΛ𝑟0\left(J_{2}>0,\Lambda_{r}>0\right)( italic_J start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 , roman_Λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT > 0 ) with a slippery cylinder wall (δ>0)𝛿0(\delta>0)( italic_δ > 0 ). We observe a reduction in the threshold amplitude with increased rotation, contrary to the scenario where the liquid flows along the outer surface of the cylinder (see FIG. 4d). Similarly to FIG. 4d, we observe the existence of a critical wavenumber k=k𝑘superscript𝑘k=k^{\dagger}italic_k = italic_k start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT for a given Ro𝑅𝑜Roitalic_R italic_o. Below ksuperscript𝑘k^{\dagger}italic_k start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, the threshold amplitude increases, whereas above ksuperscript𝑘k^{\dagger}italic_k start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, the threshold amplitude decreases.

V Traveling wave solutions and their stabilities

In this section, we investigate how rotation affects the characteristics of traveling wave (TW) solutions in the presence of wall slippage on both the outer and inner surfaces of the flow governed by (14). Prior works [58, 23, 19, 32] have studied TW solutions for thin films flowing along either the inner or outer surface of a vertical cylinder, with an emphasis on the impact of the substrate geometry and slip length. The linear stability analysis in Section III concludes that the linear wave speed is influenced by wall slippage but not by rotation for outer and inner surface flows. Here, we explore how the slippery length and rotation affect the nonlinear wave speed.

V.1 Traveling wave solutions

We consider the model on a periodic domain 0zL0𝑧𝐿0\leq z\leq L0 ≤ italic_z ≤ italic_L and introduce a change of variables to the reference frame of the traveling wave (TW) following [19],

ξ=zct,s=t,h(z,t)=h~(ξ,s),formulae-sequence𝜉𝑧𝑐𝑡formulae-sequence𝑠𝑡𝑧𝑡~𝜉𝑠\xi=z-ct,~{}s=t,~{}h(z,t)=\widetilde{h}(\xi,s),italic_ξ = italic_z - italic_c italic_t , italic_s = italic_t , italic_h ( italic_z , italic_t ) = over~ start_ARG italic_h end_ARG ( italic_ξ , italic_s ) , (27)

where c𝑐citalic_c denotes the TW speed. Substituting (27) into (14), we obtain the PDE for h~(ξ,s)~𝜉𝑠\widetilde{h}(\xi,s)over~ start_ARG italic_h end_ARG ( italic_ξ , italic_s ) in the moving reference frame,

(h~+mα2h~2)sc(h~+mα2h~2)ξ+[(h~)(1[𝒵(h~)ηh~ξξ]ξ)]ξ=0.subscript~𝑚𝛼2superscript~2𝑠𝑐subscript~𝑚𝛼2superscript~2𝜉subscriptdelimited-[]~1subscriptdelimited-[]𝒵~𝜂subscript~𝜉𝜉𝜉𝜉0\left(\widetilde{h}+\frac{m\alpha}{2}\widetilde{h}^{2}\right)_{s}-c\left(% \widetilde{h}+\frac{m\alpha}{2}\widetilde{h}^{2}\right)_{\xi}+\left[\mathcal{M% }\left(\widetilde{h}\right)\left(1-\left[\mathcal{Z}\left(\widetilde{h}\right)% -\eta\widetilde{h}_{\xi\xi}\right]_{\xi}\right)\right]_{\xi}=0.( over~ start_ARG italic_h end_ARG + divide start_ARG italic_m italic_α end_ARG start_ARG 2 end_ARG over~ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_c ( over~ start_ARG italic_h end_ARG + divide start_ARG italic_m italic_α end_ARG start_ARG 2 end_ARG over~ start_ARG italic_h end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT + [ caligraphic_M ( over~ start_ARG italic_h end_ARG ) ( 1 - [ caligraphic_Z ( over~ start_ARG italic_h end_ARG ) - italic_η over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_ξ italic_ξ end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT ) ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT = 0 . (28)

The TW solution H(ξ)𝐻𝜉H(\xi)italic_H ( italic_ξ ) represents a steady state of the PDE (28) and satisfies the fourth-order nonlinear ODE:

c(H+mα2H2)ξ=[(H)(1[𝒵(H)ηHξξ]ξ)]ξ.𝑐subscript𝐻𝑚𝛼2superscript𝐻2𝜉subscriptdelimited-[]𝐻1subscriptdelimited-[]𝒵𝐻𝜂subscript𝐻𝜉𝜉𝜉𝜉c\left(H+\frac{m\alpha}{2}H^{2}\right)_{\xi}=\left[\mathcal{M}\left(H\right)% \left(1-\left[\mathcal{Z}\left(H\right)-\eta H_{\xi\xi}\right]_{\xi}\right)% \right]_{\xi}.italic_c ( italic_H + divide start_ARG italic_m italic_α end_ARG start_ARG 2 end_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT = [ caligraphic_M ( italic_H ) ( 1 - [ caligraphic_Z ( italic_H ) - italic_η italic_H start_POSTSUBSCRIPT italic_ξ italic_ξ end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT ) ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT . (29)

This forms a nonlinear eigenvalue problem, where the propagation speed c𝑐citalic_c acts as the eigenvalue. We employ centered finite differences and Newton’s method to solve the nonlinear ODE (29), treating the speed c𝑐citalic_c as an unknown. For local uniqueness, we specify H(ξ0)=H0𝐻subscript𝜉0subscript𝐻0H\left(\xi_{0}\right)=H_{0}italic_H ( italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for some 0ξ0L0subscript𝜉0𝐿0\leq\xi_{0}\leq L0 ≤ italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ italic_L and enforce a mass constraint:

0L(H+mα2H2)𝑑ξ=M0,superscriptsubscript0𝐿𝐻𝑚𝛼2superscript𝐻2differential-d𝜉subscript𝑀0\int_{0}^{L}\left(H+\frac{m\alpha}{2}H^{2}\right)~{}d\xi=M_{0},∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT ( italic_H + divide start_ARG italic_m italic_α end_ARG start_ARG 2 end_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_d italic_ξ = italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , (30)

where M0subscript𝑀0M_{0}italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the given mass and L𝐿Litalic_L is the domain size. Using the continuation method, we numerically trace a family of TW profiles parametrized by the rotation parameter Ro𝑅𝑜Roitalic_R italic_o and the slip length δ𝛿\deltaitalic_δ, with a prescribed mass M0subscript𝑀0M_{0}italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, domain size L𝐿Litalic_L, and other system parameters. For simplicity, we focus on cases where a one-period solution fits within the domain.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 6: Traveling wave profiles satisfying the fourth-order ODE (29) with varying δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o values. For films flowing along the inner surface of the cylinder (m=1𝑚1m=-1italic_m = - 1), (a) shows the effect of δ𝛿\deltaitalic_δ with a fixed Ro=0.004𝑅𝑜0.004Ro=0.004italic_R italic_o = 0.004, and (b) shows the effect of Ro𝑅𝑜Roitalic_R italic_o with a fixed δ=0.2𝛿0.2\delta=0.2italic_δ = 0.2. For traveling waves along the outer surface of the cylinder (m=1𝑚1m=1italic_m = 1), (c) compares the TW profiles with a varying δ𝛿\deltaitalic_δ and a fixed Ro=0.02𝑅𝑜0.02Ro=0.02italic_R italic_o = 0.02, and (d) presents the effects of Ro𝑅𝑜Roitalic_R italic_o with a fixed δ=0𝛿0\delta=0italic_δ = 0. The other system parameters are α=1/6𝛼16\alpha=1/6italic_α = 1 / 6, η=0.04𝜂0.04\eta=0.04italic_η = 0.04. For the case m=1𝑚1m=1italic_m = 1, we set (M0,L)=(4.569,5)subscript𝑀0𝐿4.5695(M_{0},L)=(4.569,5)( italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_L ) = ( 4.569 , 5 ). For the case m=1𝑚1m=-1italic_m = - 1, we have (M0,L)=(13.124,15)subscript𝑀0𝐿13.12415(M_{0},L)=(13.124,15)( italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_L ) = ( 13.124 , 15 ).
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 7: The corresponding propagation speeds of the TW solutions shown in FIG. 6. (a) The speed increases as the slip length δ𝛿\deltaitalic_δ increases for films flowing along the cylinder’s inner and outer surfaces. (b - c) A larger rotation number Ro𝑅𝑜Roitalic_R italic_o leads to a decreased (increased) TW speed for the inner (outer) surface flow case. The markers indicate the speeds for TW profiles presented in FIG. 6 in each case, circles for the m=1𝑚1m=-1italic_m = - 1 cases, and triangles for the m=1𝑚1m=1italic_m = 1 cases.

In FIGs. 6-7, we present numerical solutions for typical TW profiles and their propagation speeds for liquid films flowing along both the inner (m=1𝑚1m=-1italic_m = - 1) and outer (m=1𝑚1m=1italic_m = 1) surfaces of the cylinder. For all the inner surface TW profiles (m=1𝑚1m=-1italic_m = - 1), we set the domain size L=15𝐿15L=15italic_L = 15 and the mass constraint M0=13.124subscript𝑀013.124M_{0}=13.124italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 13.124. For the outer surface TW profiles (m=1𝑚1m=1italic_m = 1), we set the domain size L=5𝐿5L=5italic_L = 5 and the mass constraint M0=4.569subscript𝑀04.569M_{0}=4.569italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.569. FIGs. 6a and 6c show that for both cases with m=±1𝑚plus-or-minus1m=\pm 1italic_m = ± 1, when other system parameters are fixed, stronger slip effects yield single-peak traveling waves with larger amplitudes. This is consistent with the linear stability analysis results, which show that slip effects can lead to enhanced wave instability. The associated speeds of the TW solutions are plotted in FIG. 7a, showing that a larger slip length leads to larger wave speeds for both inner and outer surface TWs.

FIGs. 6b and 6d reveal the different roles of rotation in TW solutions for inner and outer surface flows. With other system parameters identical, a larger rotation number leads to reduced variations in the TW profiles for inner surface flows and enlarged amplitudes for outer surface TW profiles. Correspondingly, FIGs. 7b and 7c show that the rotation effects lead to reduced TW speeds for inner surface flows and induce higher speeds for outer surface flow TW solutions.

V.2 Stability of traveling wave solutions

Previous studies by Ji et al. [19] and Liu and Ding [36] examined the stability of traveling wave (TW) solutions for a vertical cylinder under various conditions, focusing on liquid film flow along the outer surface. However, when the thin film flows along a rotating cylinder’s outer or inner surface in the presence of wall slippage, the TW solutions and their stability remain unexplored. Following [19], we conduct a stability analysis of TW solutions for the present model in this section, focusing on perturbations with the same period.

Let us consider a positive periodic TW solution H(ξ)𝐻𝜉H(\xi)italic_H ( italic_ξ ) over the domain ξ[0,L]𝜉0𝐿\xi\in[0,L]italic_ξ ∈ [ 0 , italic_L ]. We perturb it by setting h~(ξ,s)=H(ξ)+ΥΨ(ξ)eλs~𝜉𝑠𝐻𝜉ΥΨ𝜉superscript𝑒𝜆𝑠\widetilde{h}\left(\xi,s\right)=H(\xi)+\Upsilon\Psi(\xi)e^{\lambda s}over~ start_ARG italic_h end_ARG ( italic_ξ , italic_s ) = italic_H ( italic_ξ ) + roman_Υ roman_Ψ ( italic_ξ ) italic_e start_POSTSUPERSCRIPT italic_λ italic_s end_POSTSUPERSCRIPT, where Υ1much-less-thanΥ1\Upsilon\ll 1roman_Υ ≪ 1 and Ψ(ξ)Ψ𝜉\Psi(\xi)roman_Ψ ( italic_ξ ) is also L𝐿Litalic_L-periodic. We linearize (28) around the steady state H(ξ)𝐻𝜉H(\xi)italic_H ( italic_ξ ) and obtain the O(Υ)𝑂ΥO(\Upsilon)italic_O ( roman_Υ ) equation

λΨ=𝕃Ψ,𝜆Ψ𝕃Ψ\lambda\Psi=\mathbb{L}\Psi,italic_λ roman_Ψ = blackboard_L roman_Ψ , (31)

where the linear operator 𝕃𝕃\mathbb{L}blackboard_L in (31) is obtained as

𝕃Ψc(Ψξ+mαHξΨ1+mαH)+11+mαH[(H)([𝒵(H)]ξΨηΨξξ)ξ]ξ𝕃Ψ𝑐subscriptΨ𝜉𝑚𝛼subscript𝐻𝜉Ψ1𝑚𝛼𝐻11𝑚𝛼𝐻subscriptdelimited-[]𝐻subscriptsubscriptdelimited-[]𝒵𝐻𝜉Ψ𝜂subscriptΨ𝜉𝜉𝜉𝜉\mathbb{L}\Psi\equiv c\left(\Psi_{\xi}+\frac{m\alpha H_{\xi}\Psi}{1+m\alpha H}% \right)+\frac{1}{1+m\alpha H}\left[\mathcal{M}(H)\left(\left[\mathcal{Z}(H)% \right]_{\xi}\Psi-\eta\Psi_{\xi\xi}\right)_{\xi}\right]_{\xi}blackboard_L roman_Ψ ≡ italic_c ( roman_Ψ start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT + divide start_ARG italic_m italic_α italic_H start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT roman_Ψ end_ARG start_ARG 1 + italic_m italic_α italic_H end_ARG ) + divide start_ARG 1 end_ARG start_ARG 1 + italic_m italic_α italic_H end_ARG [ caligraphic_M ( italic_H ) ( [ caligraphic_Z ( italic_H ) ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT roman_Ψ - italic_η roman_Ψ start_POSTSUBSCRIPT italic_ξ italic_ξ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT
11+mαH([(H)]ξ[1(𝒵(H)ηHξξ)ξ]Ψ)ξ.11𝑚𝛼𝐻subscriptsubscriptdelimited-[]𝐻𝜉delimited-[]1subscript𝒵𝐻𝜂subscript𝐻𝜉𝜉𝜉Ψ𝜉-\frac{1}{1+m\alpha H}\left(\left[\mathcal{M}(H)\right]_{\xi}\left[1-\left(% \mathcal{Z}(H)-\eta H_{\xi\xi}\right)_{\xi}\right]\Psi\right)_{\xi}.- divide start_ARG 1 end_ARG start_ARG 1 + italic_m italic_α italic_H end_ARG ( [ caligraphic_M ( italic_H ) ] start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT [ 1 - ( caligraphic_Z ( italic_H ) - italic_η italic_H start_POSTSUBSCRIPT italic_ξ italic_ξ end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT ] roman_Ψ ) start_POSTSUBSCRIPT italic_ξ end_POSTSUBSCRIPT . (32)

In (32), setting m=1𝑚1m=1italic_m = 1 describes the case where the liquid flows over the outer surface of a rotating vertical cylinder with wall slippage. Conversely, setting m=1𝑚1m=-1italic_m = - 1 corresponds to the liquid flowing along the inner surface of the rotating vertical cylinder with wall slippage. The effect of rotation (Ro)𝑅𝑜(Ro)( italic_R italic_o ) is incorporated through the expression 𝒵(H)𝒵𝐻\mathcal{Z}(H)caligraphic_Z ( italic_H ), while the influence of wall slippage (δ)𝛿(\delta)( italic_δ ) is introduced via the mobility function (H)𝐻\mathcal{M}(H)caligraphic_M ( italic_H ). When the flow is along the outer surface of the cylinder (m=1)𝑚1(m=1)( italic_m = 1 ) and the cylinder is not rotating (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), (32) agrees well with Ji et al. [19]. Here, Ψ(ξ)Ψ𝜉\Psi(\xi)roman_Ψ ( italic_ξ ) denotes a normalized eigenfunction associated with an eigenvalue λ𝜆\lambdaitalic_λ, where Ψ2=1subscriptnormΨ21\|\Psi\|_{2}=1∥ roman_Ψ ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1. If there exists an eigenvalue λ𝜆\lambdaitalic_λ with a positive real part in (31), then the periodic TW solution H(ξ)𝐻𝜉H(\xi)italic_H ( italic_ξ ) is unstable.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 8: Stability analysis of traveling wave solutions for inner surface flows (m=1)𝑚1(m=-1)( italic_m = - 1 ). (a) A typical TW solution (solid) with Ro=0.002𝑅𝑜0.002Ro=0.002italic_R italic_o = 0.002, δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, η=0.01𝜂0.01\eta=0.01italic_η = 0.01, M0=13.124subscript𝑀013.124M_{0}=13.124italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 13.124, L=15𝐿15L=15italic_L = 15, and leading eigenfunctions (dashed curves) associated with eigenvalues λ1=0.0051±0.03isubscript𝜆1plus-or-minus0.00510.03i\lambda_{1}=0.0051\pm 0.03\mathrm{i}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.0051 ± 0.03 roman_i, λ2=0.0013±0.015isubscript𝜆2plus-or-minus0.00130.015i\lambda_{2}=0.0013\pm 0.015\mathrm{i}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.0013 ± 0.015 roman_i, and λT=0subscript𝜆𝑇0\lambda_{T}=0italic_λ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0. (b-c) The dependence of the spectrum on system parameters δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 9: Stability analysis of traveling wave solutions for the outer surface flow (m=1)𝑚1(m=1)( italic_m = 1 ). (a) A typical TW solution (solid) with Ro=0.015𝑅𝑜0.015Ro=0.015italic_R italic_o = 0.015, δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, η=0.01𝜂0.01\eta=0.01italic_η = 0.01, M0=4.569subscript𝑀04.569M_{0}=4.569italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.569, L=5𝐿5L=5italic_L = 5 and leading eigenfunctions (dashed curves) associated with eigenvalues λ1=0.01±0.34isubscript𝜆1plus-or-minus0.010.34i\lambda_{1}=0.01\pm 0.34\mathrm{i}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.01 ± 0.34 roman_i and λT=0subscript𝜆𝑇0\lambda_{T}=0italic_λ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0. (b-c) The dependence of the spectrum on δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o.

We conduct numerical investigations into the spectrum and corresponding eigenfunctions of the eigenproblem (31)–(32). In FIG. 8a, we present a typical traveling wave solution H(ξ)𝐻𝜉H(\xi)italic_H ( italic_ξ ) corresponding to a liquid film flowing along the inner surface of the cylinder (m=1𝑚1m=-1italic_m = - 1). This slowly-varying TW profile is obtained by numerically solving the nonlinear ODE problem (29)-(30) with M0=13.124subscript𝑀013.124M_{0}=13.124italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 13.124, L=15𝐿15L=15italic_L = 15, weak rotation Ro=0.002𝑅𝑜0.002Ro=0.002italic_R italic_o = 0.002, δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, and η=0.01𝜂0.01\eta=0.01italic_η = 0.01. The associated leading unstable eigenvalues of this base state are given by complex conjugate pairs λ1=0.0051±0.03isubscript𝜆1plus-or-minus0.00510.03i\lambda_{1}=0.0051\pm 0.03\mathrm{i}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.0051 ± 0.03 roman_i, λ2=0.0013±0.015isubscript𝜆2plus-or-minus0.00130.015i\lambda_{2}=0.0013\pm 0.015\mathrm{i}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.0013 ± 0.015 roman_i, and a translational eigenvalue λT=0subscript𝜆𝑇0\lambda_{T}=0italic_λ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0 that arises from the periodic boundary condition. FIGs. 8(b-c) show the dependence of the spectrum on the parameters δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o. In both figures, the real parts of the dominant unstable eigenmodes decrease as the rotation becomes stronger, indicating the stabilizing effects of rotation for TW solutions. Comparing FIG. 8b and FIG. 8c, we observe that the dominant eigenmodes with a large value of the slip length δ𝛿\deltaitalic_δ are greater than those of the case with a smaller slip length. This indicates that the slip effects destabilize the TW solutions for the case m=1𝑚1m=-1italic_m = - 1.

FIG. 9 presents a similar stability study for typical TW solutions for a liquid film flowing down the outer surface of the cylinder, corresponding to the case with m=1𝑚1m=1italic_m = 1. In FIG. 9a, the slowly-varying TW profile is obtained by solving the ODE problem (29)-(30) with M0=4.569,L=5formulae-sequencesubscript𝑀04.569𝐿5M_{0}=4.569,L=5italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4.569 , italic_L = 5, Ro=0.015𝑅𝑜0.015Ro=0.015italic_R italic_o = 0.015, δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, and η=0.01𝜂0.01\eta=0.01italic_η = 0.01. This TW solution has two unstable eigenmodes, a complex conjugate pair λ1=0.01±0.34isubscript𝜆1plus-or-minus0.010.34i\lambda_{1}=0.01\pm 0.34\mathrm{i}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.01 ± 0.34 roman_i and a translational eigenmode λT=0subscript𝜆𝑇0\lambda_{T}=0italic_λ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0, with the associated eigenfunctions presented in dashed curves in FIG. 9a. In this scenario, FIG. 9(b-c) shows that a positive rotational parameter Ro>0.01𝑅𝑜0.01Ro>0.01italic_R italic_o > 0.01 can lead to an unstable TW profile, and its stability can be further amplified by a larger slip length δ𝛿\deltaitalic_δ. This observation confirms the destabilizing role of rotation (Ro𝑅𝑜Roitalic_R italic_o) and slip parameters (δ𝛿\deltaitalic_δ) for TW solutions in the outer surface flow case.

VI Numerical simulations

To examine the growth of the film instability involving cylinder rotation in the presence of wall slippage, we perform numerical simulations of the PDE (14). Section VI.1 discusses the choke phenomenon for S0𝑆0S\rightarrow 0italic_S → 0. Section VI.2 focuses on the breakup or rupture behavior when Sb𝑆𝑏S\rightarrow bitalic_S → italic_b. Note that, here S𝑆Sitalic_S is related to the film thickness hhitalic_h by S(z,t)=b+mh(z,t)𝑆𝑧𝑡𝑏𝑚𝑧𝑡S(z,t)=b+mh(z,t)italic_S ( italic_z , italic_t ) = italic_b + italic_m italic_h ( italic_z , italic_t ), where m=±1𝑚plus-or-minus1m=\pm 1italic_m = ± 1.

VI.1 Choke phenomenon

In this section, we focus on the “choke phenomenon” [57] or “plug formation” [32, 39] when the thin film flows along the inner surface of the cylinder (m=1)𝑚1(m=-1)( italic_m = - 1 ). For the PDE (14), we consider the initial condition as

S(z,0)=R1+j=1J𝒜jsin(2πLjz)+jcos(2πLjz).𝑆𝑧0𝑅1superscriptsubscript𝑗1𝐽subscript𝒜𝑗2𝜋𝐿𝑗𝑧subscript𝑗2𝜋𝐿𝑗𝑧S(z,0)=R-1+\sum_{j=1}^{J}\mathcal{A}_{j}\sin\left(\frac{2\pi}{L}jz\right)+% \mathcal{B}_{j}\cos\left(\frac{2\pi}{L}jz\right).italic_S ( italic_z , 0 ) = italic_R - 1 + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_sin ( divide start_ARG 2 italic_π end_ARG start_ARG italic_L end_ARG italic_j italic_z ) + caligraphic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_cos ( divide start_ARG 2 italic_π end_ARG start_ARG italic_L end_ARG italic_j italic_z ) . (33a)
In (33a), we set 𝒜j=j=103subscript𝒜𝑗subscript𝑗superscript103\mathcal{A}_{j}=\mathcal{B}_{j}=10^{-3}caligraphic_A start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = caligraphic_B start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and J=10𝐽10J=10italic_J = 10, i.e., 20202020 harmonic modes.

We approximate the spatial solution by the Fourier method as given below

S(z,t)=(N/2)+1N/2S^nexp[2πinzL]𝑆𝑧𝑡superscriptsubscript𝑁21𝑁2subscript^𝑆𝑛2𝜋𝑖𝑛𝑧𝐿S(z,t)=\sum_{-(N/2)+1}^{N/2}\widehat{S}_{n}\exp\left[\frac{2\pi inz}{L}\right]italic_S ( italic_z , italic_t ) = ∑ start_POSTSUBSCRIPT - ( italic_N / 2 ) + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N / 2 end_POSTSUPERSCRIPT over^ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT roman_exp [ divide start_ARG 2 italic_π italic_i italic_n italic_z end_ARG start_ARG italic_L end_ARG ] (33b)

In this case, we use adaptive time stepping and set the relative error to 108superscript10810^{-8}10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT. For the other parameters for investigating the choke phenomenon, we set the other parameters as α=1/3𝛼13\alpha=1/3italic_α = 1 / 3, η=0.0625𝜂0.0625\eta=0.0625italic_η = 0.0625, and N=1024𝑁1024N=1024italic_N = 1024.

Refer to caption
Figure 10: (a-c) Interface profiles near the choke time t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for a slippery fixed cylinder (δ>0,Ro=0)formulae-sequence𝛿0𝑅𝑜0(\delta>0,Ro=0)( italic_δ > 0 , italic_R italic_o = 0 ). (e-g) Interface profiles near choke time t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for a slippery rotating cylinder (δ>0,Ro>0)formulae-sequence𝛿0𝑅𝑜0(\delta>0,Ro>0)( italic_δ > 0 , italic_R italic_o > 0 ). Evolution of the minimum S(z,t)𝑆𝑧𝑡S(z,t)italic_S ( italic_z , italic_t ) for different (d) slip lengths δ𝛿\deltaitalic_δ when Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0 and (h) rotation numbers Ro𝑅𝑜Roitalic_R italic_o when δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4. Here, the domain size L=60𝐿60L=60italic_L = 60.

In FIG. 10(a-c), we illustrate the evolution of the free surface when the vertical cylinder is stationary (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ) and the inner cylinder’s surface is slippery (δ>0)𝛿0(\delta>0)( italic_δ > 0 ). We use two typical values for δ𝛿\deltaitalic_δ, δ=0.1𝛿0.1\delta=0.1italic_δ = 0.1 and δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, to show the effect of wall slippage on the choke phenomenon. FIG. 10a shows that when δ=0𝛿0\delta=0italic_δ = 0, indicating no slippage on the inner cylinder surface, the choke time (tc)subscript𝑡𝑐(t_{c})( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) is approximately 132.57. In contrast, FIGs. 10b and 10c reveal that the choke time (tc)subscript𝑡𝑐(t_{c})( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) decreases as wall slippage (δ)𝛿(\delta)( italic_δ ) increases. Additionally, we define the minimum of S(z,t)𝑆𝑧𝑡S(z,t)italic_S ( italic_z , italic_t ) as Sminsubscript𝑆minS_{\text{min}}italic_S start_POSTSUBSCRIPT min end_POSTSUBSCRIPT, i.e., Smin=minzS(z,t)subscript𝑆minsubscript𝑧𝑆𝑧𝑡S_{\text{min}}=\min_{z}S(z,t)italic_S start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = roman_min start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_S ( italic_z , italic_t ). In FIG. 10d, Sminsubscript𝑆minS_{\text{min}}italic_S start_POSTSUBSCRIPT min end_POSTSUBSCRIPT is plotted over time for the different slip lengths δ𝛿\deltaitalic_δ shown in FIGs. 10(a-c). It is evident that as time progresses, the choke time t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT shortens with greater wall slippage. Thus, the left column of FIG. 10 demonstrates that wall slip enhances the choking behavior or plug formation. This observation is consistent with Schwitzerlett et al. [32], which also noted that wall slip facilitates plug formation. In FIG. 10(e-g), we explore the impact of rotation on the choke phenomenon when the inner cylinder surface is slippery (Ro>0,δ>0)formulae-sequence𝑅𝑜0𝛿0(Ro>0,\delta>0)( italic_R italic_o > 0 , italic_δ > 0 ). The slip length is fixed at δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, with other parameters matching those in FIG. 10(a-c). We choose three typical values for Ro𝑅𝑜Roitalic_R italic_o: 0.001, 0.005, and 0.01. In FIG. 10e, for Ro=0.001𝑅𝑜0.001Ro=0.001italic_R italic_o = 0.001, the choke time (tc)subscript𝑡𝑐(t_{c})( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) is approximately 37.02, compared to approximately 35.44 in FIG. 10c without rotation. This suggests that rotation increases the choke time (tc)subscript𝑡𝑐(t_{c})( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ). FIGs. 10f and 10g, where Ro=0.005𝑅𝑜0.005Ro=0.005italic_R italic_o = 0.005 and Ro=0.01𝑅𝑜0.01Ro=0.01italic_R italic_o = 0.01 respectively, show that the choke time (tc)subscript𝑡𝑐(t_{c})( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) further increases with higher Ro𝑅𝑜Roitalic_R italic_o. In FIG. 10h, Sminsubscript𝑆minS_{\text{min}}italic_S start_POSTSUBSCRIPT min end_POSTSUBSCRIPT is plotted over time for the different rotation numbers Ro𝑅𝑜Roitalic_R italic_o shown in FIG. 10(e-g), with the slip length fixed at δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4. Comparing FIGs. 10d and 10h for δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, it is evident that as time progresses, the choke time t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases with higher rotation number Ro𝑅𝑜Roitalic_R italic_o. Therefore, the right column of FIG. 10 concludes that while wall slippage promotes choking behavior or plug formation inside a cylinder, introducing rotation can mitigate this effect.

Refer to caption
Figure 11: (a-d) Influence of Ro𝑅𝑜Roitalic_R italic_o on film profiles over time to delay choke behavior with δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4. (e-h) Influence of rotation number Ro𝑅𝑜Roitalic_R italic_o on film profiles at t35.44𝑡35.44t\approx 35.44italic_t ≈ 35.44 to delay choke behavior with slip length δ𝛿\deltaitalic_δ. Here L=30𝐿30L=30italic_L = 30

In FIG. 11, we depict the delayed choking behavior induced by rotation (Ro𝑅𝑜Roitalic_R italic_o) in the presence of wall slippage (δ)𝛿(\delta)( italic_δ ). In the top panel of FIG. 11, we highlight the free surface configurations to emphasize the delay in choke behavior caused by rotation, with the slip length δ𝛿\deltaitalic_δ fixed at 0.4. FIG. 11a shows the surface configuration at time t=29𝑡29t=29italic_t = 29 when Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0, and we observe that the minimum value of S𝑆Sitalic_S approaches zero at approximately tc35.44subscript𝑡𝑐35.44t_{c}\approx 35.44italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 35.44. Starting from the same initial conditions and varying the rotation number Ro𝑅𝑜Roitalic_R italic_o (Ro=0.001,0.002,0.004𝑅𝑜0.0010.0020.004Ro=0.001,0.002,0.004italic_R italic_o = 0.001 , 0.002 , 0.004), we present the surface configurations from time t=29𝑡29t=29italic_t = 29 to t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in FIGs. 11(b-d). It is clear that while choke behavior occurs at the critical time t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0 for a given wall slippage δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, the onset of choking is delayed in the presence of rotation (Ro>0𝑅𝑜0Ro>0italic_R italic_o > 0), requiring more time for the spatial disturbance to develop. In the bottom panel, we portray the profiles when the slip length varies from δ=0𝛿0\delta=0italic_δ = 0 to δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4 with or without rotation. FIG. 11e shows the free surface configuration without rotation (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), beginning with δ=0𝛿0\delta=0italic_δ = 0 (no slippage). We observe the minimum value of S0𝑆0S\rightarrow 0italic_S → 0 at t=tc35.44𝑡subscript𝑡𝑐35.44t=t_{c}\approx 35.44italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 35.44 when δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4. Keeping time constant at t=tc𝑡subscript𝑡𝑐t=t_{c}italic_t = italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, we keep increasing the rotation number Ro𝑅𝑜Roitalic_R italic_o in FIGs. 11(f-h) and show the surface configurations for δ𝛿\deltaitalic_δ values ranging from 0 to 0.4. It is evident that at tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the choke behavior occurs when the cylinder is stationary (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), but this behavior is delayed as the cylinder begins to rotate (Ro>0)𝑅𝑜0(Ro>0)( italic_R italic_o > 0 ).

VI.2 Breakup behavior and transient solutions

This subsection explores how wall slip and rotation affect the nonlinear evolution of a film flowing on the slippery, rotating vertical cylinder’s outer surface (m=1)𝑚1(m=1)( italic_m = 1 ). We mainly focus on the impact of these effects on the breakup behavior and transient dynamic solutions. We set the computational domain as the interval [L/2,L/2]𝐿2𝐿2[-L/2,L/2][ - italic_L / 2 , italic_L / 2 ] and impose a periodic condition to simulate the PDE (14).

Refer to caption
Figure 12: Influence of rotation on film profiles when the film flows along the outer surface of a slippery cylinder. The other fixed parameters are α=1/3𝛼13\alpha=1/3italic_α = 1 / 3, η=0.0025𝜂0.0025\eta=0.0025italic_η = 0.0025 and L=30𝐿30L=30italic_L = 30

When a breakup happens for a given parameter, the model collapses at a finite time (say, t=tb𝑡subscript𝑡𝑏t=t_{b}italic_t = italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT), known as a finite-time blow-up. As ttb𝑡subscript𝑡𝑏t\rightarrow t_{b}italic_t → italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT, the largest interface deformation rapidly grows until it matches the cylinder radius, i.e., when S(z,t)=b𝑆𝑧𝑡𝑏S(z,t)=bitalic_S ( italic_z , italic_t ) = italic_b. At this point, the system reaches a singularity, and solutions cease to exist beyond the time t=tb𝑡subscript𝑡𝑏t=t_{b}italic_t = italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. Therefore, detecting the onset of model breakup for a given parameter involves checking whether the interface Sb𝑆𝑏S\rightarrow bitalic_S → italic_b. Generally, breakup behavior is local and insensitive to initial and boundary conditions. Therefore, to assess breakup behavior concerning the two key parameters, δ𝛿\deltaitalic_δ and Ro𝑅𝑜Roitalic_R italic_o, we maintain L=30𝐿30L=30italic_L = 30 as a fixed value and set the initial condition as

h(z,0)=10.2sin(2πLz),𝑧010.22𝜋𝐿𝑧h(z,0)=1-0.2\sin\left(\frac{2\pi}{L}z\right),italic_h ( italic_z , 0 ) = 1 - 0.2 roman_sin ( divide start_ARG 2 italic_π end_ARG start_ARG italic_L end_ARG italic_z ) , (34)

In this case, we have used 51210245121024512-1024512 - 1024 Fourier modes (depending on the length of the computational domain). The implicit Gear’s method in time is used, with the relative error set to 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT.

In FIGs. 12(a-d), we illustrate the interface profiles at an instant t=50𝑡50t=50italic_t = 50 under the condition where the outer surface of the cylinder is slippery, with a fixed slip length δ=0.1𝛿0.1\delta=0.1italic_δ = 0.1. We consider three representative values to analyze the influence of rotation Ro𝑅𝑜Roitalic_R italic_o: Ro=0.04,0.08,𝑅𝑜0.040.08Ro=0.04,0.08,italic_R italic_o = 0.04 , 0.08 , and 0.10.10.10.1. FIG. 12a shows the scenario without rotation (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ), where no breakup phenomenon occurs; instead, the interface appears as a quasi-steady traveling wave. In FIG. 12b, with Ro=0.04𝑅𝑜0.04Ro=0.04italic_R italic_o = 0.04, rotational effects cause the interface to break into small droplets separated by varying lengths of inter-drop spacing compared to the non-rotating case (see FIG. 12a). In FIG. 12c, where Ro=0.08𝑅𝑜0.08Ro=0.08italic_R italic_o = 0.08, we observe that larger droplets begin to form, and the unstable film between these larger droplets evolves into smaller droplets. Similar time-periodic isolated droplet regime dynamics have been observed in non-rotating fiber coating models [19]. The droplets are irregularly distributed compared to the Ro=0.04𝑅𝑜0.04Ro=0.04italic_R italic_o = 0.04 case in FIG. 12b. In FIG. 12d, where Ro=0.1𝑅𝑜0.1Ro=0.1italic_R italic_o = 0.1, we find that the big droplets are well developed and larger compared to those in the case of Ro=0.08𝑅𝑜0.08Ro=0.08italic_R italic_o = 0.08. For this set of parameters, there are almost four small droplets between two large droplets. Additionally, we observe that the smaller droplets are nearly uniform in shape at Ro=0.04𝑅𝑜0.04Ro=0.04italic_R italic_o = 0.04 (see FIG. 12b). However, at Ro=0.08𝑅𝑜0.08Ro=0.08italic_R italic_o = 0.08 and Ro=0.1𝑅𝑜0.1Ro=0.1italic_R italic_o = 0.1 (see FIGs. 12c and 12d), while the larger droplets exhibit almost uniform shapes, the smaller droplets between them do not have a similar structure.

FIGs. 12(a-d) depict the scenario with a slip length δ=0.1𝛿0.1\delta=0.1italic_δ = 0.1, while FIGs. 12(e-f) illustrate the impact of rotation for a higher slip length, δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, with all other parameters unchanged from FIGs. 12(a-d). In FIG. 12e, with Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0 and δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4, the interface forms a more developed quasi-steady traveling wave compared to FIG. 12a, without any breakup phenomenon observed. FIG. 12f shows film breakup and smaller droplet formation with Ro=0.04𝑅𝑜0.04Ro=0.04italic_R italic_o = 0.04. The droplets are slightly larger than those in FIG. 12b due to the higher slip length. In FIG. 12g, for Ro=0.08𝑅𝑜0.08Ro=0.08italic_R italic_o = 0.08, large droplets form with four smaller droplets between each pair. Unlike the irregular distribution in FIG. 12c with δ=0.1𝛿0.1\delta=0.1italic_δ = 0.1, the droplets in FIG. 12g for δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4 are almost uniformly spaced. Finally, in FIG. 12h, with Ro=0.1𝑅𝑜0.1Ro=0.1italic_R italic_o = 0.1, more small droplets are observed between each pair of large droplets compared to FIG. 12g (Ro=0.08𝑅𝑜0.08Ro=0.08italic_R italic_o = 0.08), and the large droplets are larger.

Refer to caption
Figure 13: Effect of rotation number Ro𝑅𝑜Roitalic_R italic_o on film profiles to enhance breakup behavior with slip length. Here zbsubscript𝑧𝑏z_{b}italic_z start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the breakup location, i.e., zb=z|t=tbsubscript𝑧𝑏evaluated-at𝑧𝑡subscript𝑡𝑏\left.z_{b}=z\right|_{t=t_{b}}italic_z start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = italic_z | start_POSTSUBSCRIPT italic_t = italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUBSCRIPT. The other fixed parameters are α=1/3𝛼13\alpha=1/3italic_α = 1 / 3, η=0.0025𝜂0.0025\eta=0.0025italic_η = 0.0025 and L=30𝐿30L=30italic_L = 30

FIG. 13 provides a clearer depiction of the liquid-air interfacial profiles, showing how the rupture or breakup phenomena are influenced by both rotation (Ro)𝑅𝑜(Ro)( italic_R italic_o ) and slip length (δ)𝛿(\delta)( italic_δ ) for the film flowing along the outer surface of the cylinder. These profiles are plotted at t=50𝑡50t=50italic_t = 50. To analyze the effect of Ro𝑅𝑜Roitalic_R italic_o, we consider the case without rotation (Ro=0)𝑅𝑜0(Ro=0)( italic_R italic_o = 0 ) and plot values up to Ro=0.14𝑅𝑜0.14Ro=0.14italic_R italic_o = 0.14. Both subfigures demonstrate that in the presence of Ro𝑅𝑜Roitalic_R italic_o, a rupture occurs where the film’s minimal thickness approaches zero, and integration rapidly diverges. As Ro𝑅𝑜Roitalic_R italic_o increases, this tendency for breakup becomes more pronounced. Additionally, it has been observed that increased slip along the cylinder wall accelerates the pronounced breakup induced by rotation.

VI.3 Effects of rotation and slip on free surface configurations for a Gaussian-shaped profile

In Sections VI.1 and VI.2, we discussed the impact of rotation and wall slip on film profiles using sinusoidal initial conditions. In this section, we explore the effects of a slippery cylinder under rotation on film profiles, considering a Gaussian-shaped initial profile centered at z=z0𝑧subscript𝑧0z=z_{0}italic_z = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We perturb the interface as

h(z,0)=1+ς^exp[0.5(z10)2],𝑧01^𝜍0.5superscript𝑧102h(z,0)=1+\widehat{\varsigma}\exp\left[-0.5\left(z-10\right)^{2}\right],italic_h ( italic_z , 0 ) = 1 + over^ start_ARG italic_ς end_ARG roman_exp [ - 0.5 ( italic_z - 10 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] , (35)

where ς^^𝜍\widehat{\varsigma}over^ start_ARG italic_ς end_ARG is a small number that denotes the random disturbance, and z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is set to 10. In this case, we have used 51210245121024512-1024512 - 1024 Fourier modes (depending on the length of the computational domain). Again, the implicit Gear’s method in time is used, with the relative error set to 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT.

Refer to caption
Figure 14: Free surface profiles at t=200𝑡200t=200italic_t = 200 for the outer surface of the cylinder (α=1/6𝛼16\alpha=1/6italic_α = 1 / 6) and the inner surface of the cylinder (α=1/4𝛼14\alpha=1/4italic_α = 1 / 4) under the influence of rotation, in the presence of wall slippage. The other fixed parameters are L=20𝐿20L=20italic_L = 20, η=0.01𝜂0.01\eta=0.01italic_η = 0.01 and ς^=0.1^𝜍0.1\widehat{\varsigma}=0.1over^ start_ARG italic_ς end_ARG = 0.1

FIG. 14 presents the free surface configurations for a rotating slippery cylinder (δ>0,Ro>0)formulae-sequence𝛿0𝑅𝑜0(\delta>0,Ro>0)( italic_δ > 0 , italic_R italic_o > 0 ). The first two columns illustrate the film flow along the outer surface of the cylinder, while the third and fourth columns depict the flow along the inner surface. In the first column, with a fixed slip length of δ=0.1𝛿0.1\delta=0.1italic_δ = 0.1, profiles for Ro=0,0.005,0.01,0.015,𝑅𝑜00.0050.010.015Ro=0,0.005,0.01,0.015,italic_R italic_o = 0 , 0.005 , 0.01 , 0.015 , and 0.020.020.020.02 are shown. This column reveals that as the rotation rate increases, the interfacial waves form higher humps, indicating increased instability compared to the non-rotating case (Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0) displayed in FIG. 14a1. Additionally, as Ro𝑅𝑜Roitalic_R italic_o increases, the one-hump solitary-like wave structure becomes more oscillatory, as seen in FIG. 14e1. In the second column, the slip length is increased to δ=0.2𝛿0.2\delta=0.2italic_δ = 0.2, with other parameters remaining the same. Comparing the first and second columns, it is evident that higher wall slippage results in more pronounced surface distortions. In the third column, the slip length is set to δ=0.3𝛿0.3\delta=0.3italic_δ = 0.3 for the inner surface flow of the cylinder. This column shows that with increased rotation, the interfacial waves exhibit lower humps, thereby reducing instability compared to the non-rotating case (Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0) shown in FIG. 14a3. Comparing FIGs. 14a3 and 14e3, it can be observed that with higher Ro𝑅𝑜Roitalic_R italic_o, the highly distorted liquid-gas interface transitions into a one-hump solitary-like wave structure. In the fourth column, the slip length is further increased to δ=0.4𝛿0.4\delta=0.4italic_δ = 0.4. Here, wall slippage amplifies the disturbances; however, higher rotation mitigates these disturbances.

Refer to caption
Figure 15: Free surface profiles at two different instants for the outer (m=1𝑚1m=1italic_m = 1) and inner (m=1𝑚1m=-1italic_m = - 1) surfaces of the slippery cylinder under the influence of rotation. The other fixed parameters are α=1/6𝛼16\alpha=1/6italic_α = 1 / 6, δ=0.2𝛿0.2\delta=0.2italic_δ = 0.2, η=0.04𝜂0.04\eta=0.04italic_η = 0.04, L=40𝐿40L=40italic_L = 40, and ς^=0.01^𝜍0.01\widehat{\varsigma}=0.01over^ start_ARG italic_ς end_ARG = 0.01

In FIG. 14, we present the interfacial shape at t=200𝑡200t=200italic_t = 200. In FIG. 15, we show how the film profiles are influenced by rotation (Ro𝑅𝑜Roitalic_R italic_o) at longer times, t=500𝑡500t=500italic_t = 500 and t=800𝑡800t=800italic_t = 800. The left panel illustrates the film flow along the outer surface of the slippery cylinder, while the right panel shows the flow along the inner surface. The slip length δ𝛿\deltaitalic_δ is fixed at 0.2, and the computational domain length is L=40𝐿40L=40italic_L = 40 for both outer and inner cylinder scenarios. FIGs. 15a and 15c indicate that at later times, oscillations become more significant with higher rotation when the film flows along the outer surface of the slippery cylinder. Conversely, FIGs. 15b and 15d demonstrate that although surface wave instability slightly amplifies over time, there is no significant development of an oscillatory wave structure for Ro=0𝑅𝑜0Ro=0italic_R italic_o = 0. Even at longer times, rotation reduces surface wave instability in the presence of wall slippage. Additionally, for FIGs. 15b and 15d, the free surface profiles are plotted at Ro=0.1𝑅𝑜0.1Ro=0.1italic_R italic_o = 0.1. At this Ro𝑅𝑜Roitalic_R italic_o value, the crest is almost reduced, indicating a high stabilizing effect of rotation on the inner surface flow, even when the cylinder wall is slippery.

VII Summary and conclusions

In this study, we have conducted a detailed theoretical investigation of gravity-driven viscous film flows along a vertical cylinder with a slippery surface. To account the effect of rotation on the flow dynamics and stability, we have considered the vertical cylinder to be rotating around its vertical axis. The slipperiness of the cylinder wall is accounted for by employing the Navier-slip boundary condition (velocity at the wall is assumed to vary proportionally to the normal derivative of velocity at the solid-liquid interface). To describe the dynamics of the film interface, we have assumed that the mean thickness of the film is much smaller than its characteristic length in the axial direction. Our model captures the coupled effect of rotation and wall slippage for the flow along the outer and inner surface of the vertical cylinder.

We have performed a linear stability analysis to investigate the impact of rotation on the stability of axisymmetric disturbances in flows where wall slip occurs on both the inner and outer surfaces. Our findings indicate that while wall slip enhances flow instability on both surfaces, rotation stabilizes the inner surface but destabilizes the outer surface. In addition to the linear analysis, we have conducted a weakly nonlinear stability analysis to explore how nonlinear effects influence the stability thresholds in our model. This investigation revealed the existence of both supercritical and subcritical regimes, influenced by both wall slip and rotation for flows on both surfaces. Furthermore, we have extended our study to consider traveling wave solutions for both outer and inner surface flows, examining how rotation modifies their characteristics and velocities in the presence of wall slip. We have also evaluated the stability of these traveling wave solutions under conditions where the cylinder rotates, and the wall exhibits slip. Our results demonstrate that incorporating rotational effects can partially or completely mitigate interfacial instabilities in traveling waves for the inner surface flow within a fixed domain. Conversely, for the outer surface flow, rotation tends to exacerbate interfacial instabilities in traveling waves. Finally, numerically simulating the evolution equation, we demonstrate that introducing rotation effectively delays the onset of choking phenomena in the inner surface flow, even when wall slippage is present. Conversely, our simulations also reveal that rotation amplifies the breakup process in the system for flow along the outer surface of the cylinder, particularly pronounced in the presence of slip.

This study holds potentially significant implications for industrial coating processes and the lubrication of pipes with Newtonian fluids on slippery solid surfaces. Our analysis reveals that rotation can mitigate slip-induced instability for inner surface flow but not for outer surface flow. Our analysis, relevant to fluids with negligible inertia, could inspire future research to experimentally validate these findings and further explore thin-film flows along slippery cylinders at low to moderate Reynolds numbers. Future investigations will include evaluating non-isothermal effects and examining non-Newtonian films on similar geometries.

Conflict of interest

The authors have no conflicts to disclose.

DATA AVAILABLITY STATEMENT

The data that supports the findings of this study are available within the article.

Acknowledgements

AKG acknowledges the support from Science and Engineering Research Board (SERB) grant SIR/2022/001357, Government of India. HJ acknowledges support from grant NSF DMS 2309774.

Declaration of Generative AI and AI-assisted technologies in the writing process

During the preparation of this work, the authors used ChatGPT in order to improve the language and readability of the article. After using this tool, the authors reviewed and edited the content as needed and take full responsibility for the content of the publication.

References

  • [1] D. Garg, P. N. Dyer, Tungsten carbide erosion resistant coating for aerospace components, MRS Online Proceedings Library (OPL), 168 (1989), 213-220
  • [2] J. Zhu, J. L. Chen, R. K. Lade, W. J. Suszynski, L. F. Francis, Water-based coatings for 3D printed parts, Journal of Coatings Technology and Research, 12 (2015), 889-897
  • [3] D. Quéré, Thin films flowing on vertical fibers, Europhysics Letters, 13 (1990), 721-726
  • [4] S. Kalliadasis, H. C. Chang, Drop formation during coating of vertical fibres, Journal of Fluid Mechanics, 261 (1994), 135-168
  • [5] D. Quéré, Fluid coating on a fiber, Annual Review of Fluid Mechanics, 31 (1999), 347-384
  • [6] J. Grünig, E. Lyagin, S. Horn, T. Skale, M. Kraume, Mass transfer characteristics of liquid films flowing down a vertical wire in a counter current gas flow, Chemical Engineering Science, 69 (2012), 329-339
  • [7] A. Sadeghpour, Z. Zeng, H. Ji, N. Dehdari Ebrahimi, A. L. Bertozzi, Y. S. Ju, Water vapor capturing using an array of traveling liquid beads for desalination and water treatment, Science Advances, 5 (2019), eaav7662
  • [8] H. Gau, S. Herminghaus, P. Lenz, R. Lipowsky, Liquid morphologies on structured surfaces: from microchannels to microchips, Science, 283 (1999), 46-49
  • [9] J. B. Grotberg, Pulmonary flow and transport phenomena, Annual Review of Fluid Mechanics, 26 (1994), 529-571
  • [10] R. V. Craster, O. K. Matar, On viscous beads flowing down a vertical fibre, Journal of Fluid Mechanics, 553 (2006), 85-105
  • [11] C. Ruyer-Quil, P. Treveleyan, F. Giorgiutti-Dauphiné, C. Duprat, S. Kalliadasis, Modelling film flows down a fibre, Journal of Fluid Mechanics, 603 (2008), 431-462
  • [12] C. Duprat, C. Ruyer-Quil, F. Giorgiutti-Dauphiné, Spatial evolution of a film flowing down a fiber, Physics of Fluids, 21 (2009), 042109
  • [13] K. J. Ruschak, L. E. Scriven, Rimming flow of liquid in a rotating cylinder, Journal of Fluid Mechanics, 76 (1976), 113-126
  • [14] J. Ashmore, A. E. Hosoi, H. A. Stone, The effect of surface tension on rimming flows in a partially filled rotating cylinder, Journal of Fluid Mechanics, 479 (2003), 65-98
  • [15] A. E. Hosoi, L. Mahadevan, Axial instability of a free-surface front in a partially filled horizontal rotating cylinder, Physics of Fluids, 11 (1999), 97-106
  • [16] S. T. Thoroddsen, L. Mahadevan, Experimental study of coating flows in a partially-filled horizontally rotating cylinder, Experiments in Fluids, 23 (1997), 1-13
  • [17] V. Kozlov, D. Polezhaev, Flow patterns in a rotating horizontal cylinder partially filled with liquid, Physical Review E, 92 (2015), 013016
  • [18] D. Halpern, H. H. Wei, Slip-enhanced drop formation in a liquid falling down a vertical fibre, Journal of Fluid Mechanics, 820 (2017), 42-60
  • [19] H. Ji, C. Falcon, A. Sadeghpour, Z. Zeng, Y. S. Ju, A. L. Bertozzi, Dynamics of thin liquid films on vertical cylindrical fibres, Journal of Fluid Mechanics, 865 (2019), 303-327
  • [20] A. L. Frenkel, Nonlinear theory of strongly undulating thin films flowing down vertical cylinders, Europhysics Letters, 18 (1992), 583
  • [21] I. L. Kliakhandler, S. H. Davis, S. G. Bankoff, Viscous beads on vertical fibre, Journal of Fluid Mechanics, 429 (2001), 381-390
  • [22] B. Kim, H. Ji, A. L. Bertozzi, A. Sadeghpour, Y. S. Ju, A positivity-preserving numerical method for a thin liquid film on a vertical cylindrical fiber, Journal of Computational Physics, 496 (2024), 112560
  • [23] R. M. Taranets, H. Ji, M. Chugunova, On weak solutions of a control-volume model for liquid films flowing down a fibre, Discrete and Continuous Dynamical Systems-B, (2024), 1-32
  • [24] E. Novbari, A. Oron, Energy integral method model for the nonlinear dynamics of an axisymmetric thin liquid film falling on a vertical cylinder, Physics of Fluids, 21 (2009), 062107
  • [25] A. Oron, S. H. Davis, S. G. Bankoff, Long-scale evolution of thin liquid films, Review of Modern Physics, 69 (1997), 931-980
  • [26] R. V. Craster, O. K. Matar, Dynamics and stability of thin liquid films, Review of Modern Physics, 81 (2009), 1131-1198
  • [27] R. Liu, Q. S. Liu, Thermocapillary effect on the dynamics of viscous beads on vertical fiber, Physical Review E, 90 (2014), 033005
  • [28] G. M. Sisoev, R. V. Craster, O. K. Matar, S. V. Gerasimov, Film flow down a fibre at moderate flow rates, Chemical Engineering Science, 61 (2006), 7279-7298
  • [29] S. Haefner, M. Benzaquen, O. Bäumchen, T. Salez, R. Peters, J. D. McGraw, K. Jacobs, E. Raphaël, K. Dalnoki-Veress, Influence of slip on the Plateau–Rayleigh instability on a fibre, Nature Communications, 6 (2015), 7409
  • [30] Y. Chao, Z. Ding, R. Liu, Dynamics of thin liquid films flowing down the uniformly heated/cooled cylinder with wall slippage, Chemical Engineering Science, 175 (2018), 354-364
  • [31] S. Chattopadhyay, Thin liquid films on a slippery vertical cylinder in presence of chemical reaction, Chemical Engineering Science, 282 (2023), 119211
  • [32] M. Schwitzerlett, H. R. Ogrosky, I. Topaloglu, A long-wave model for film flow inside a tube with slip, Journal of Fluid Mechanics, 974 (2023), A22
  • [33] L. A. Dávalos-Orozco, G. Ruiz-Chavarría, Hydrodynamic instability of a fluid layer flowing down a rotating cylinder, Physics of Fluids A: Fluid Dynamics, 5 (1993), 2390-2404
  • [34] C. I. Chen, C. K. Chen, Y. T. Yang, Perturbation analysis to the nonlinear stability characterization of thin condensate falling film on the outer surface of a rotating vertical cylinder, International Journal of Heat and Mass Transfer, 47 (2004), 1937-1951
  • [35] M. Rietz, B. Scheid, F. Gallaire, N. Kofman, R. Kneer, W. Rohlfs, Dynamics of falling films on the outside of a vertical rotating cylinder: waves, rivulets and dripping transitions, Journal of Fluid Mechanics, 832 (2017), 189-211
  • [36] R. Liu, Z. Ding, Instabilities and bifurcations of liquid films flowing down a rotating fibre, Journal of Fluid Mechanics, 899 (2020), A14
  • [37] U. Farooq, J. Stafford, C. Petit, O. K. Matar, Numerical simulations of a falling film on the inner surface of a rotating cylinder, Physical Review E, 102 (2020), 043106
  • [38] A. Mukhopadhyay, S. Chattopadhyay, A. K. Barua, Stability of thin film flowing down the outer surface of a rotating non-uniformly heated vertical cylinder, Nonlinear Dynamics, 100 (2020), 1143-1172
  • [39] R. Liu, Z. Ding, Stability of viscous film flow coating the interior of a vertical tube with a porous wall, Physical Review E, 95 (2017), 053101
  • [40] C. Ruyer-Quil, S. P. M. J. Trevelyan, F. Giorgiutti-Dauphiné, C. Duprat, S. Kalliadasis, Film flows down a fiber: Modeling and influence of streamwise viscous diffusion, The European Physical Journal Special Topics, 166 (2009), 89-92
  • [41] L. Yu, J. Hinch, The velocity of ‘large’ viscous drops falling on a coated vertical fibre, Journal of Fluid Mechanics, 737 (2013), 232-248
  • [42] J. P. Rothstein, Slip on superhydrophobic surfaces, Annual Review of Fluid Mechanics, 42 (2010), 89-109
  • [43] S. Kalliadasis, C. Ruyer-Quil, B. Scheid, M. G. Velarde, Falling Liquid Films, Applied Mathematical Sciences, 176, Springer, 2012
  • [44] S. Chattopadhyay, H. Ji, Thermocapillary thin film flows on a slippery substrate with odd viscosity effects, Physica D: Nonlinear Phenomena, 455 (2023), 133883
  • [45] Z. Ding, T. N. Wong, Falling liquid films on a slippery substrate with Marangoni effects, International Journal of Heat and Mass Transfer, 90 (2015), 689-701
  • [46] I. M. R. Sadiq, R. Usha, S. W. Joo, Instabilities in a liquid film flow over an inclined heated porous substrate, Chemical Engineering Science, 65 (2010), 4443-4459
  • [47] Z. Ding, Q. Liu, Stability of liquid films on a porous vertical cylinder, Physical Review E, 84 (2011), 046307
  • [48] M. J. Tan, S. G. Bankoff, S. H. Davis, Steady thermocapillary flows of thin liquid layers. I. Theory, Physics of Fluids A: Fluid Dynamics, 2 (1990), 313-321
  • [49] C. Duprat, C. Ruyer-Quil, S. Kalliadasis, F. Giorgiutti-Dauphiné, Absolute and convective instabilities of a viscous film flowing down a vertical fiber, Physical Review Letters, 98 (2007), 244502
  • [50] S. Chattopadhyay, Odd-viscosity-induced instability of a thin film with variable density, Physics of Fluids, 33 (2021), 082102
  • [51] S. Chattopadhyay, A. S. Desai, Dynamics and stability of weakly viscoelastic film flowing down a uniformly heated slippery incline, Physical Review Fluids, 7 (2022), 064007064007064007064007
  • [52] A. Oron, O. Gottlieb, Subcritical and supercritical bifurcations of the first-and second-order Benney equation, Journal of Engineering Mathematics, 50 (2004), 121-140
  • [53] S. Chattopadhyay, Thermocapillary instability in the presence of uniform normal electric field: effect of odd viscosity, Journal of Engineering Mathematics, 131 (2021), 1-22
  • [54] A. Samanta, Stability of liquid film falling down a vertical non-uniformly heated wall, Physica D: Nonlinear Phenomena, 237 (2008), 2587-2598
  • [55] A. S. Desai, S. Chattopadhyay, A. K. Gaonkar, Shear imposed falling liquid films on a slippery substrate with Marangoni effects: Effect of odd viscosity, International Journal of Non-Linear Mechanics, 156 (2023), 104507
  • [56] S. Chattopadhyay, A. Mukhopadhyay, A. K. Barua, A. K. Gaonkar, Thermocapillary instability on a film falling down a non-uniformly heated slippery incline, International Journal of Non-Linear Mechanics, 133 (2021), 103718
  • [57] R. Camassa, H. R. Ogrosky, J. Olander, Viscous film flow coating the interior of a vertical tube. Part 1. Gravity-driven flow, Journal of Fluid Mechanics, 745 (2014), 682-715
  • [58] H. Ji, R. Taranets, M. Chugunova, On travelling wave solutions of a model of a liquid film flowing down a fibre, European Journal of Applied Mathematics, 33 (2022), 864-893
  • [59] Z. Ding, Z. Liu, R. Liu, C. Yang, Thermocapillary effect on the dynamics of liquid films coating the interior surface of a tube, International Journal of Heat and Mass Transfer, 138 (2019), 524-533
  • [60] O. K. Matar, R. V. Craster, S. Kumar, Falling films on flexible inclines, Physical Review E, 76 (2007), 056301
  • [61] S. Chattopadhyay, A. S. Desai, A. K. Gaonkar, A. Mukhopadhyay, Role of odd viscosity on falling films over compliant substrates, Physical Review Fluids, 8 (2023), 064003