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

Effect of rotation on anisotropic scattering suspension of phototactic algae

S. K. Rajput [email protected] Department of Mathematics, PDPM Indian Institute of Information Technology Design and Manufacturing, Jabalpur 482005, India.
Abstract

In this article, the effect of rotation on the onset of phototactic bioconvection is investigated using linear stability theory for a suspension of forward-scattering phototactic algae in this article. The suspension is uniformly illuminated by collimated flux. The bio-convective instability is characterized by an unstable mode of disturbance that transitions from a stationary (overstable) to an overstable (stationary) state as the Taylor number varies under fixed parameters. It is also observed that the suspension has significant stabilizing effect due to rotation of the system.

I INTRODUCTION

Bioconvection is a widespread phenomenon in fluid dynamics that involves the formation of patterns in suspensions of living microorganisms, such as algae and bacteria Platt (1961); Pedley and Kessler (1992); Hill and Pedley (2005); Bees (2020); Javadi et al. (2020). The term "bioconvection" was first introduced by Platt in 1961. Typically, these microorganisms have a higher density than the surrounding fluid on a small scale and collectively swim in an upward direction. However, there are cases in natural environments where density differences are not necessary for pattern formation, and the microorganisms exhibit different behaviours. Non-living microorganisms do not exhibit pattern formation behaviour. Microorganisms respond to external stimuli and exhibit behavioural changes in their swimming direction, known as taxis. Examples of taxis include gravitaxis, which is influenced by gravitational acceleration, gyrotaxis, which is influenced by both gravitational acceleration and viscous torque when microorganisms have bottom-heavy structures, and phototaxis, which is the movement of microorganisms towards or away from a light source. This article focuses specifically on the effect of phototaxis.

The phenomenon of bioconvection patterns in algal suspensions has been the subject of extensive research, highlighting the significant role of light in shaping fluid dynamics and cellular distribution Wager (1911); Kitsunezaki, Komori, and Harumoto (2007). When exposed to intense light in well-mixed cultures of photosynthetic microorganisms, dynamic patterns can either persist or undergo disruption. Strong light can hinder the formation of stable patterns, and it can also disrupt pre-existing  Kessler (1985); Williams and Bees (2011); Kessler (1989) The response of algae to light of varying intensities is a key determinant in the modulation of bioconvection patterns. Moreover, the interaction of light with microorganisms involves processes of light absorption and scattering. Algal light scattering can be characterized as either isotropic or anisotropic, the latter further classified into forward and backward scattering based on cell properties such as size, shape, and refractive index. Algae, due to their size, predominantly exhibit forward scattering of light within the visible wavelength range.

The phototaxis model proposed by Rajput is being employed to investigate the bioconvection system. In this study, the suspension of phototactic microorganisms is assumed to undergo rotation around the z-axis at a constant angular velocity, while being illuminated from above by collimated flux. Given that many motile algae rely on photosynthesis for their energy needs, they exhibit distinct phototactic behavior. To ensure a comprehensive understanding of their response to light, it is essential to examine the phototaxis model in the context of a rotating medium. Unlike previous models that considered scattering effects in an isotropic manner, our approach incorporates anisotropic scattering. This allows for a more precise representation of the swimming behaviour of microorganisms as they react to light stimuli. By considering the influence of anisotropic scattering, we aim to enhance the accuracy of the model and provide valuable insights into the intricate dynamics of the bioconvection phenomenon.

Refer to caption
Figure 1: Formation of sublayer in rotating medium.

In this experimental setup, the mathematical formulation of the problem involves considering the behavior of a forward scattering algal suspension confined between two parallel boundaries in the (y, z) plane. The suspension is assumed to be uniformly illuminated from above by collimated flux. The goal is to study the behavior of the suspension under the influence of light.

To analyze the system, a realistic phototaxis model is employed, which takes into account the forward scattering of light by the algae. The equilibrium solution of the system is determined by considering a steady-state where the flow velocity is zero, indicating no fluid motion. In this state, the balance between phototaxis and diffusion leads to the formation of a horizontal sublayer where cells accumulate, dividing the suspension into two distinct regions.

The critical intensity GcG_{c} plays a crucial role in determining the position of the sublayer. Above the sublayer (G>GcG>G_{c}), the light intensity is sufficient to suppress cell movement, while below the sublayer (G<GcG<G_{c}), cells exhibit upward motion towards the sublayer due to phototaxis.

The lower region beneath the sublayer is considered unstable and has the potential to exhibit fluid motion and penetrate the stable upper region through penetrative convection if the system becomes unstable. Penetrative convection is a phenomenon observed in various convection problems.

Over the years, researchers have made significant advancements in modeling phototaxis and bioconvection, employing various approaches such as linear stability theory and numerical simulations. For instance, Vincent and Hill Vincent and Hill (1996) investigated the initiation of phototactic bioconvection and identified non-stationary and non-oscillatory disturbance modes. Ghorai and Hill Ghorai and Hill (2005) conducted a two-dimensional simulation of bioconvective flow patterns based on Vincent and Hill’s model, albeit without considering algae scattering. Ghorai et al. Ghorai, Panda, and Hill (2010) examined the onset of bioconvection in an isotropic scattering suspension and observed a steady-state profile with a sublayer located at different depths due to isotropic scattering. Panda and Ghorai Panda and Ghorai (2013) conducted a linear stability analysis in a forward scattering algal suspension, observing a transition from non-oscillatory to non-stationary modes during bioconvective instability caused by forward scattering, while Panda and Singh Panda and Singh (2016) numerically explored the linear stability of PBC using Vincent and Hill’s continuum model in a two-dimensional setting. Previous studies did not account for the effects of diffuse irradiation. Panda et al. Panda et al. (2016) proposed a model that examined how diffuse radiation influenced isotropic scattering algal suspensions, revealing that diffuse irradiation significantly stabilizes such suspensions. Panda Panda (2020) further investigated the impact of forward anisotropic scattering on the onset of PBC, considering both diffuse and collimated irradiation. However, these studies did not account for the influence of oblique collimated flux. Panda et al. Panda, Sharma, and Kumar (2022) introduced oblique collimated flux and discovered that bioconvective solutions switch from non-oscillatory to overstable states and vice versa during bioconvective instability induced by oblique collimated flux. Kumar Kumar (2022) examined the effect of oblique collimated irradiation on isotropic scattering algal suspensions, revealing the existence of both stationary and overstable solutions within certain parameter ranges. Kumar Kumar (2023a) investigated the effect of rotation on the phototactic bioconvection in a non-scattering medium and found a significant stabilizing impact due to rotation. More recently, Panda and Rajput Panda and Rajput (2023) investigate the combined impact of oblique and diffuse irradiation on the onset of bioconvection.

However, there is still a need to investigate phototactic bioconvection in a rotating forward scattering algal suspension using a realistic phototaxis model. Therefore, the objective of this study is to examine bioconvective instability in such a system and provide insights into the complex dynamics of bioconvection in a light-induced pattern formation in a dilute algal suspension.

The article follows a structured approach, starting with the mathematical formulation of the problem. It derives the equilibrium solution and perturbs the governing system of bioconvection with small disturbances. The linear stability problem is formulated and solved using numerical methods. The obtained results are then presented and discussed. The article concludes by highlighting the novelty and significance of the proposed model in advancing the understanding of bioconvective instability in rotating forward scattering algal suspensions.

II MATHEMATICAL FORMULATION

The research focuses on studying the behaviour of a forward scattering algal suspension within the (y, z) plane, which is confined between two parallel boundaries. The boundaries are assumed to be infinite in extent, and there is no reflection of light from the top and bottom boundaries. This setup allows for the investigation of the suspension’s behaviour under the influence of light.

The suspension is uniformly illuminated from above by collimated flux. Collimated flux refers to light that is travelling in parallel rays, indicating that the incident light rays are not scattered before reaching the suspension. This uniform illumination from above serves as the driving force for the system and influences the phototactic behaviour of the algae.

II.1 THE AVERAGE SWIMMING DIRECTION

The Radiative Transfer Equation (RTE) is utilized to determine the light intensity in the system. This equation provides a mathematical description of how light propagates and interacts with the medium. The RTE can be expressed as follows

dL(𝒙,𝒓)dr+(a+σs)L(𝒙,𝒓)=σs4π04πL(𝒙,𝒓)Ξ(𝒓,𝒓)𝑑Ω,\frac{dL(\bm{x},\bm{r})}{dr}+(a+\sigma_{s})L(\bm{x},\bm{r})=\frac{\sigma_{s}}{4\pi}\int_{0}^{4\pi}L(\bm{x},\bm{r^{\prime}})\Xi(\bm{r},\bm{r^{\prime}})d\Omega^{\prime}, (1)

where aa, and σs\sigma_{s} are the absorption coefficient, and scattering coefficients respectively. Ξ(𝒓,𝒓)\Xi(\bm{r},\bm{r^{\prime}}) is the scattering phase function, which describes how light is distributed in different directions when scattered from 𝒓\bm{r^{\prime}} to 𝒓\bm{r}. In this study, we assume that the scattering phase function follows linear anisotropy with azimuthal symmetry, specifically Ξ(𝒓,𝒓)=1+Acosθcosθ\Xi(\bm{r},\bm{r^{\prime}})=1+A\cos{\theta}\cos{\theta^{\prime}}. Here, AA represents the anisotropic coefficient, which determines the degree of anisotropy. Positive values of AA indicate forward scattering, negative values indicate backward scattering and A=0A=0 represents isotropic scattering.

In the given model, the light intensity on the top of the suspension is described by:

L(𝒙b,𝒔)=Ltδ(cosθcosθ0),L(\bm{x}_{b},\bm{s})=L_{t}\delta(\cos{\theta}-\cos{\theta_{0}}),

where 𝒙b=(x,y,H)\bm{x}_{b}=(x,y,H) represents the location on the top boundary surface. LtL_{t} is the magnitude of the collimated flux, θ\theta is the zenith angle, and θ0\theta_{0} represents a specific value of the zenith angle.

Introducing two variables, a=αn(𝒙)a=\alpha n(\bm{x}) and σs=βn(𝒙)\sigma_{s}=\beta n(\bm{x}), in terms of the scattering albedo ω=σs/(a+σs)\omega=\sigma_{s}/(a+\sigma_{s}), the Radiative Transfer Equation (RTE) can be expressed as:

dL(𝒙,𝒓)dr+(α+β)nL(𝒙,𝒓)=βn4π04πL(𝒙,𝒓)(Acosθcosθ)𝑑Ω,\frac{dL(\bm{x},\bm{r})}{dr}+(\alpha+\beta)nL(\bm{x},\bm{r})=\frac{\beta n}{4\pi}\int_{0}^{4\pi}L(\bm{x},\bm{r^{\prime}})(A\cos{\theta}\cos{\theta^{\prime}})d\Omega^{\prime}, (2)

where L(𝒙,𝒓)L(\bm{x},\bm{r}) represents the light intensity at position 𝒙\bm{x} along the ray direction 𝒓\bm{r}, α\alpha is the absorption coefficient, β\beta is the scattering coefficient, nn is the cell concentration, and Ω\Omega^{\prime} represents the solid angle.

The total intensity at a fixed point 𝒙\bm{x} in the medium is given by:

G(𝒙)=04πL(𝒙,𝒓)𝑑Ω,G(\bm{x})=\int_{0}^{4\pi}L(\bm{x},\bm{r})d\Omega,

and the radiative heat flux is given by:

𝒒(𝒙)=04πL(𝒙,𝒓)𝒓𝑑Ω.\bm{q}(\bm{x})=\int_{0}^{4\pi}L(\bm{x},\bm{r})\bm{r}d\Omega. (3)

The average swimming velocity of a microorganism is defined as:

𝑾c=Wc<𝒑>.\bm{W}_{c}=W_{c}<\bm{p}>.

The mean swimming orientation of the cell, denoted as <𝒑><\bm{p}>, is determined by employing the following calculation. In this context, WcW_{c} denotes the average cell swimming speed.

<𝒑>=M(G)𝒒|𝒒|,<\bm{p}>=-M(G)\frac{\bm{q}}{|\bm{q}|}, (4)

where M(G)M(G) is the taxis function that characterizes the response of algae cells to light. It takes the form:

M(G)={0, G(𝒙)Gc,<0, G(𝒙)>Gc.M(G)=\left\{\begin{array}[]{ll}\geq 0,&\mbox{ }G(\bm{x})\leq G_{c},\\ <0,&\mbox{ }G(\bm{x})>G_{c}.\end{array}\right.

The taxis function determines the behaviour of the cells in response to the light intensity.

II.2 GOVERNING EQUATIONS WITH BOUNDARY CONDITIONS

In the proposed model of the suspension of phototactic microorganisms, several equations govern the system. These equations, along with the corresponding boundary conditions, are as follows:

Continuity equation:

𝑼=0,\bm{\nabla}\cdot\bm{U}=0, (5)

Momentum equation under the Boussinesq approximation:

ρ(𝑼t+(𝑼)𝑼+2𝛀×𝑼)=P+μ2𝑼nVgΔρ𝒛^,\rho\left(\frac{\partial\bm{U}}{\partial t}+(\bm{U}\cdot\nabla)\bm{U}+2\bm{\Omega}\times\bm{U}\right)=-\bm{\nabla}P+\mu\nabla^{2}\bm{U}-nVg\Delta\rho\hat{\bm{z}}, (6)

Cell conservation equation:

nt=𝑩,\frac{\partial n}{\partial t}=-\bm{\nabla}\cdot\bm{B}, (7)

Total cell flux equation:

𝑩=nU+nWc<𝒑>𝑫n,\bm{B}=nU+nW_{c}<\bm{p}>-\bm{D}\bm{\nabla}n, (8)

where UU is the average fluid velocity, nn is the number of algal cells per unit volume, VV is the volume of the cells, ρ\rho is the density of water, Δρ/ρ\Delta\rho/\rho represents the small variation in density of the cells, 𝛀\bm{\Omega} is the angular velocity, μ\mu is the dynamic viscosity of the suspension, 𝑫\bm{D} is the cell diffusivity, DIDI represents the constant diffusivity assumption, gg is the acceleration due to gravity, and 𝒑\bm{p} is the direction of cell swimming.

The boundary conditions for the system are as follows:

For the lower and upper boundaries:

𝑼𝒛^=0atz=0,H,\bm{U}\cdot\hat{\bm{z}}=0\quad\text{at}\quad z=0,H, (9)
𝑩𝒛^=0atz=0,H.\bm{B}\cdot\hat{\bm{z}}=0\quad\text{at}\quad z=0,H. (10)

For the rigid boundary:

𝑼×𝒛^=0atz=0,\bm{U}\times\hat{\bm{z}}=0\quad\text{at}\quad z=0, (11)

For the stress-free boundary:

2z2(𝑼𝒛^)=0atz=H.\frac{\partial^{2}}{\partial z^{2}}(\bm{U}\cdot\hat{\bm{z}})=0\quad\text{at}\quad z=H. (12)

Additionally, the upper boundary is subjected to collimated flux, which leads to specific conditions for the intensity at the boundaries:

For the upper boundary:

L(x,y,z=0,θ,ϕ)=Ltδ(𝒓𝒓𝟎),(π2θπ),L(x,y,z=0,\theta,\phi)=L_{t}\delta(\bm{r}-\bm{r_{0}}),\quad\left(\frac{\pi}{2}\leq\theta\leq\pi\right), (13a)
L(x,y,z=0,θ,ϕ)=0,(0θπ2).L(x,y,z=0,\theta,\phi)=0,\quad\left(0\leq\theta\leq\frac{\pi}{2}\right). (13b)

In these equations, LtL_{t} represents the intensity of light at a specific location 𝒓𝟎\bm{r_{0}}, and δ(𝒓𝒓𝟎)\delta(\bm{r}-\bm{r_{0}}) denotes the Dirac delta function. These boundary conditions specify the behaviour of the fluid velocity, cell concentration, and light intensity at the boundaries of the system.

To express the governing equations in a dimensionless form, we introduce the following scales: Length HH, time H2/DH^{2}/D, velocity scale D/HD/H, pressure scale μD/H2\mu D/H^{2}, concentration n¯\bar{n}. Using these scales, we can non-dimensionalize the governing equations

𝑼=0,\bm{\nabla}\cdot\bm{U}=0, (14)
Sc1(𝑼t+(𝑼)𝑼)+Ta(z^×𝑼)=PeRn𝒛^+2𝑼,S_{c}^{-1}\left(\frac{\partial\bm{U}}{\partial t}+(\bm{U}\cdot\nabla)\bm{U}\right)+\sqrt{T_{a}}(\hat{z}\times\bm{U})=-\nabla P_{e}-Rn\hat{\bm{z}}+\nabla^{2}\bm{U}, (15)
nt=𝑩=[𝒏𝑼+𝒏𝑽𝒄<𝒑>𝒏.]\frac{\partial{n}}{\partial{t}}=-\bm{\nabla}\cdot\bm{B}=-{\bm{\nabla}}\cdot[\bm{n{\bm{U}}+nV_{c}<{\bm{p}}>-{\bm{\nabla}}n.}] (16)

In the above equations, Sc1=ν/DS_{c}^{-1}=\nu/D represents the Schmidt number, VcV_{c} denotes the dimensionless swimming speed as Vc=WcH/DV_{c}=W_{c}H/D, R=n~VgΔρH3/νρDR=\tilde{n}Vg\Delta{\rho}H^{3}/\nu\rho{D} is the Rayleigh number, and Ta=4Ω2H4/ν2T_{a}=4\Omega^{2}H^{4}/\nu^{2} is the Taylor number.

After non-dimensionalization, the boundary conditions can be expressed in a dimensionless form as follows:

Uz^=0atz=0,1,U\cdot\hat{z}=0\quad\text{at}\quad z=0,1, (17)
Bz^=0atz=0,1,B\cdot\hat{z}=0\quad\text{at}\quad z=0,1, (18)
U×z^=0atz=0,U\times\hat{z}=0\quad\text{at}\quad z=0, (19)
2z2(Uz^)=0atz=1.\frac{\partial^{2}}{\partial z^{2}}(U\cdot\hat{z})=0\quad\text{at}\quad z=1. (20)

The nondimensional Radiative Transfer Equation (RTE) can be expressed as:

dL(𝒙,𝒓)dr+κnL(𝒙,𝒓)=σn4π04πL(𝒙,𝒓)(Acosθcosθ)𝑑Ω,\frac{dL(\bm{x},\bm{r})}{dr}+\kappa nL(\bm{x},\bm{r})=\frac{\sigma n}{4\pi}\int_{0}^{4\pi}L(\bm{x},\bm{r^{\prime}})(A\cos{\theta}\cos{\theta^{\prime}})d\Omega^{\prime}, (21)

where κ=(α+β)n¯H\kappa=(\alpha+\beta)\bar{n}H and σ=βn¯H\sigma=\beta\bar{n}H are the nondimensional extinction coefficient and scattering coefficient, respectively. The scattering albedo ω=σ/(κ+σ)\omega=\sigma/(\kappa+\sigma) represents the scattering efficiency of microorganisms. The RTE describes the balance between absorption and scattering of light in the medium.

Alternatively, the RTE can be formulated using direction cosines, which provide a convenient way to describe the direction of light propagation:

ξdLdx+ηdLdy+νdLdz+κnL(𝒙,𝒓)=ωκn4π04πL(𝒙,𝒓)(Acosθcosθ)𝑑Ω,\xi\frac{dL}{dx}+\eta\frac{dL}{dy}+\nu\frac{dL}{dz}+\kappa nL(\bm{x},\bm{r})=\frac{\omega\kappa n}{4\pi}\int_{0}^{4\pi}L(\bm{x},\bm{r^{\prime}})(A\cos{\theta}\cos{\theta^{\prime}})d\Omega^{\prime}, (22)

In the dimensionless form, the intensity at the boundaries is given by:

L(x,y,z=1,θ,ϕ)=Ltδ(𝒓𝒓𝟎),(π2θπ),L(x,y,z=1,\theta,\phi)=L_{t}\delta(\bm{r}-\bm{r_{0}}),\qquad\left(\frac{\pi}{2}\leq\theta\leq\pi\right), (23a)
L(x,y,z=0,θ,ϕ)=0,(0θπ2).L(x,y,z=0,\theta,\phi)=0,\qquad\left(0\leq\theta\leq\frac{\pi}{2}\right). (23b)

III THE STEADY STATE SOLUTION

The equilibrium state of the system is characterized by the following conditions:

𝑼=0,n=ns(z),ζs=×U=0,andL=Ls(z,θ).\bm{U}=0,~{}~{}~{}n=n_{s}(z),~{}~{}~{}\zeta_{s}=\nabla\times U=0,~{}~{}~{}\text{and}~{}~{}~{}L=L_{s}(z,\theta).

At equilibrium, the total intensity GsG_{s} and radiative flux 𝒒s\bm{q}_{s} can be expressed as:

Gs=04πLs(z,θ)𝑑Ω,and𝒒s=04πLs(z,θ)𝒔𝑑Ω.G_{s}=\int_{0}^{4\pi}L_{s}(z,\theta)d\Omega,\quad\text{and}\quad\bm{q}_{s}=\int_{0}^{4\pi}L_{s}(z,\theta)\bm{s}d\Omega.

The 𝒒s\bm{q}_{s} vector has zero components in xx and yy directions due to the independence of Lsd(z,θ)L_{s}^{d}(z,\theta) on ϕ\phi. Therefore, 𝒒s\bm{q}_{s} can be written as 𝒒s=qs𝒛^\bm{q}_{s}=-q_{s}\hat{\bm{z}}, where qsq_{s} represents the magnitude of 𝒒s\bm{q}_{s}.

The equation governing LsL_{s} at equilibrium is given by:

dLsdz+κnsLsν=ωκns4πν(Gs(z)Aqsν).\frac{dL_{s}}{dz}+\frac{\kappa n_{s}L_{s}}{\nu}=\frac{\omega\kappa n_{s}}{4\pi\nu}(G_{s}(z)-Aq_{s}\nu). (24)

The equilibrium intensity LsL_{s} can be decomposed into two components: the collimated part LscL_{s}^{c} and the diffuse part caused by scattering LsdL_{s}^{d}. The equation governing the collimated component LscL_{s}^{c} is:

dLscdz+κnsLscν=0,\frac{dL_{s}^{c}}{dz}+\frac{\kappa n_{s}L_{s}^{c}}{\nu}=0, (25)

with the boundary condition:

Lsc(1,θ)=Ltδ(𝒓𝒓0),θ[π/2,π],L_{s}^{c}(1,\theta)=L_{t}\delta(\bm{r}-\bm{r}_{0}),\quad\theta\in[\pi/2,\pi], (26)

where LtL_{t} represents the total intensity at the point source location 𝒓0\bm{r}_{0}.

Solving the governing equation for LscL_{s}^{c} with the given boundary condition yields the expression for LscL_{s}^{c}:

Lsc=Ltexp(z1κns(z)ν𝑑z)δ(𝒓𝒓0).L_{s}^{c}=L_{t}\exp\left(\int_{z}^{1}\frac{\kappa n_{s}(z^{\prime})}{\nu}dz^{\prime}\right)\delta(\bm{r}-\bm{r}_{0}). (27)

The diffuse part LsdL_{s}^{d} is governed by the equation:

dLsddz+κnsLsdν=ωκns4πν(Gs(z)Aqsν),\frac{dL_{s}^{d}}{dz}+\frac{\kappa n_{s}L_{s}^{d}}{\nu}=\frac{\omega\kappa n_{s}}{4\pi\nu}(G_{s}(z)-Aq_{s}\nu), (28)

with the boundary conditions:

Lsd(1,θ)=0,θ[π/2,π],L_{s}^{d}(1,\theta)=0,\quad\theta\in[\pi/2,\pi], (29)
Lsd(0,θ)=0,θ[0,π/2].L_{s}^{d}(0,\theta)=0,\quad\theta\in[0,\pi/2]. (30)

In the basic state, the total intensity GsG_{s} is given by:

Gs=Gsc+Gsd=04π[Lsc(z,θ)+Lsd(z,θ)]𝑑Ω=Ltexp(z1κns(z)𝑑z)+0πLsd(z,θ)𝑑Ω,G_{s}=G_{s}^{c}+G_{s}^{d}=\int_{0}^{4\pi}[L_{s}^{c}(z,\theta)+L_{s}^{d}(z,\theta)]d\Omega=L_{t}\exp\left(-\int_{z}^{1}\kappa n_{s}(z^{\prime})dz^{\prime}\right)+\int_{0}^{\pi}L_{s}^{d}(z,\theta)d\Omega, (31)

where GscG_{s}^{c} and GsdG_{s}^{d} represent the collimated and diffuse components of GsG_{s}, respectively.

Similarly, the radiative heat flux 𝒒s\bm{q}_{s} in the basic state can be written as:

𝒒s=𝒒sc+𝒒sd=04π(Lsc(z,θ)+Lsd(z,θ))𝒓𝑑Ω=Ltexp(z1κns(z)dz)𝒛^+04πLsd(z,θ)𝒓𝑑Ω.\bm{q}_{s}=\bm{q}_{s}^{c}+\bm{q}_{s}^{d}=\int_{0}^{4\pi}\left(L_{s}^{c}(z,\theta)+L_{s}^{d}(z,\theta)\right)\bm{r}d\Omega=L_{t}\exp\left(\int_{z}^{1}-\kappa n_{s}(z^{\prime})dz^{\prime}\right)\hat{\bm{z}}+\int_{0}^{4\pi}L_{s}^{d}(z,\theta)\bm{r}d\Omega. (32)

To solve the problem, a new variable τ\tau is defined as:

τ=z1κns(z)𝑑z.\tau=\int_{z}^{1}\kappa n_{s}(z^{\prime})dz^{\prime}.

These equations provide a set of coupled Fredholm integral equations of the second kind. Specifically, the equation for GsG_{s} can be written as:

Gs(τ)=eτ+ω20κGs(τ)E1(|ττ|)𝑑τ+Asgn(ττ)qs(τ)E2(|ττ|),G_{s}(\tau)=e^{-\tau}+\frac{\omega}{2}\int_{0}^{\kappa}G_{s}(\tau^{\prime})E_{1}(|\tau-\tau^{\prime}|)d\tau^{\prime}+Asgn(\tau-\tau^{\prime})q_{s}(\tau^{\prime})E_{2}(|\tau-\tau^{\prime}|), (33)

where E1(x)E_{1}(x) and E2(x)E_{2}(x) represent the exponential integrals of order 1 and 2, respectively.

qs(τ)=eτ+ω20κAqs(τ)E3(|ττ|)𝑑τ+sgn(ττ)Gs(τ)E2(|ττ|).q_{s}(\tau)=e^{-\tau}+\frac{\omega}{2}\int_{0}^{\kappa}Aq_{s}(\tau^{\prime})E_{3}(|\tau-\tau^{\prime}|)d\tau^{\prime}+sgn(\tau-\tau^{\prime})G_{s}(\tau^{\prime})E_{2}(|\tau-\tau^{\prime}|). (34)

The coupled fractional integral equations (FIEs) can be solved using the method of subtraction of singularity, where En(x)E_{n}(x) represents the exponential integral of order nn and sgn(x)sgn(x) denotes the signum function.

Refer to caption
Figure 2: The change in the overall intensity within a homogeneous suspension by altering the anisotropic scattering coefficient AA in the forward direction from 0 to 0.8 is examined for three distinct scenarios: ω=0.1,0.42\omega=0.1,0.42, and 0.47. The cases are represented as follows: A=0()=0(-), A=0.4()=0.4(---), and A=0.8()=0.8(-\cdot-\cdot-). The governing parameter values of the system, namely Sc=20,k=0.5S_{c}=20,k=0.5, and Lt=1L_{t}=1, remain constant.

The mean swimming direction in the basic state is given by

<𝒑𝒔>=Ms𝒒𝒔qs=Ms𝒛^,<\bm{p_{s}}>=-M_{s}\frac{\bm{q_{s}}}{q_{s}}=M_{s}\hat{\bm{z}},

where Ms=M(Gs)M_{s}=M(G_{s}). In the basic state, the cell concentration ns(z)n_{s}(z) satisfies the following equation

dnsdzVcMsns=0,\frac{dn_{s}}{dz}-V_{c}M_{s}n_{s}=0, (35)

which is accompanied by the equation

01ns(z)𝑑z=1.\int_{0}^{1}n_{s}(z)dz=1. (36)

Equations (33) to (36) constitute a boundary value problem that can be solved using numerical techniques such as the shooting method.

To demonstrate the equilibrium state, we fix the values of Sc=20S_{c}=20, Gc=1G_{c}=1, κ=0.5\kappa=0.5, and ω\omega at specific values of 0.1, 0.42, and 0.47. We then investigate the influence of the forward scattering coefficient AA by varying it within the range of 0 to 0.8. The focus is on analyzing the effects of these variations on the total intensity (GsG_{s}) and the fundamental equilibrium solution.

Figure 2 visually illustrates the relationship between the total intensity GsG_{s} and the depth of a uniform suspension, while incrementally increasing the forward scattering coefficient AA from 0 to 0.8 while keeping ω\omega constant. As the value of AA increases, there is a noticeable increase in the intensity of GsG_{s} at lower regions of the uniform suspension. Conversely, at the upper regions, the intensity of GsG_{s} decreases.

Refer to caption
Figure 3: (a) The response curve of taxis for Gc=1G_{c}=1, and (b) changes in the concentration profile of the base by varying the forward scattering coefficient AA from 0 to 0.8 for three different scenarios: ω=0.1,0.42\omega=0.1,0.42 and 0.47. The cases are represented as follows: A=0()=0(-), A=0.4()=0.4(---), and A=0.8()=0.8(-\cdot-\cdot-). The fixed parameter values governing the system are Sc=20,Vc=20,k=0.5S_{c}=20,V_{c}=20,k=0.5, and Lt=1L_{t}=1.

In Figure 3(a), the plot illustrates the phototaxis function as a function of light intensity, considering a critical value of Gc=1G_{c}=1. On the other hand, Figure 3(b) displays the influence of the forward scattering coefficient AA on the equilibrium position of the sublayer, using the same governing parameters.

In the case where ω=0\omega=0, the sublayer forms near the top of the suspension domain in the equilibrium state. However, as AA increases from 0 to 0.8, the sublayer’s equilibrium position gradually shifts towards the lower region.

Similarly, for ω=0.42\omega=0.42 and 0.47, the sublayer’s equilibrium position in the equilibrium state is located approximately at three-quarters and the middle height of the suspension, respectively. Once again, an increase in AA from 0 to 0.8 leads to a downward shift of the sublayer’s equilibrium position in each scenario.

IV Linear stability of the problem

To analyse stability, we employ linear perturbation theory, where a small perturbation with an amplitude ϵ1\epsilon\ll 1 is introduced to the equilibrium state. This perturbation can be mathematically represented by the following equation:

[𝑼,n,ζ,L,<p>]=[0,ns,ζs,Ls,<ps>]+ϵ[𝑼1,n1,ζ1,L1,<𝒑1>]+𝒪(ϵ2)[\bm{U},n,\zeta,L,<p>]=[0,n_{s},\zeta_{s},L_{s},<p_{s}>]+\epsilon[\bm{U}_{1},n_{1},\zeta_{1},L_{1},<\bm{p}_{1}>]+\mathcal{O}(\epsilon^{2}) (37)

By substituting the perturbed variables into equations (14) to (16) and linearizing the equations, we gather terms of o(ϵ)o(\epsilon) around the equilibrium state, leading to the following expression:

𝑼1=0,\bm{\nabla}\cdot\bm{U}_{1}=0, (38)

where 𝑼1=(U1,V1,W1)\bm{U}_{1}=(U_{1},V_{1},W_{1}).

Sc1(𝑼𝟏t)+Ta(z×U1)+Pe+Rn1𝒛^=2𝑼𝟏,Sc^{-1}\left(\frac{\partial\bm{U_{1}}}{\partial t}\right)+\sqrt{T_{a}}(z\times U_{1})+\bm{\nabla}P_{e}+Rn_{1}\hat{\bm{z}}=\nabla^{2}\bm{U_{1}}, (39)
n1t+Vc(<𝒑𝒔>n1+<𝒑𝟏>ns)+W1dnsdz=2n1.\frac{\partial{n_{1}}}{\partial{t}}+V_{c}\bm{\nabla}\cdot(<\bm{p_{s}}>n_{1}+<\bm{p_{1}}>n_{s})+W_{1}\frac{dn_{s}}{dz}=\bm{\nabla}^{2}n_{1}. (40)

If G=Gs+ϵG1+𝒪(ϵ2)=(Gsc+ϵG1c)+(Gsd+ϵG1d)+𝒪(ϵ2)G=G_{s}+\epsilon G_{1}+\mathcal{O}(\epsilon^{2})=(G_{s}^{c}+\epsilon G_{1}^{c})+(G_{s}^{d}+\epsilon G_{1}^{d})+\mathcal{O}(\epsilon^{2}), then the steady collimated total intensity is perturbed as Ltexp(κz1(ns(z)+ϵn1+𝒪(ϵ2))𝑑z)L_{t}\exp\left(-\kappa\int_{z}^{1}(n_{s}(z^{\prime})+\epsilon n_{1}+\mathcal{O}(\epsilon^{2}))dz^{\prime}\right) and after simplification, we get

G1c=Ltexp(z1κns(z)𝑑z)(1zκn1𝑑z)G_{1}^{c}=L_{t}\exp\left(-\int_{z}^{1}\kappa n_{s}(z^{\prime})dz^{\prime}\right)\left(\int_{1}^{z}\kappa n_{1}dz^{\prime}\right) (41)

Similarly, G1dG_{1}^{d} is given by

G1d=04πL1d(𝒙,𝒓)𝑑Ω.G_{1}^{d}=\int_{0}^{4\pi}L_{1}^{d}(\bm{x},\bm{r})d\Omega. (42)

Similarly, for the radiative heat flux q=qs+ϵq1++𝒪(ϵ2)=(qsc+ϵq1c)+(qsd+ϵq1d)+𝒪(ϵ2)q=q_{s}+\epsilon q_{1}++\mathcal{O}(\epsilon^{2})=(q_{s}^{c}+\epsilon q_{1}^{c})+(q_{s}^{d}+\epsilon q_{1}^{d})+\mathcal{O}(\epsilon^{2}), we find

𝒒1c=Ltexp(z1κns(z)𝑑z)(1zκn1𝑑z)z^\bm{q}_{1}^{c}=L_{t}\exp\left(-\int_{z}^{1}\kappa n_{s}(z^{\prime})dz^{\prime}\right)\left(\int_{1}^{z}\kappa n_{1}dz^{\prime}\right)\hat{z} (43)

and

q1d=04πL1d(𝒙,𝒓)𝒓𝑑Ω.q_{1}^{d}=\int_{0}^{4\pi}L_{1}^{d}(\bm{x},\bm{r})\bm{r}d\Omega. (44)

After perturbing the expression for swimming orientation and collecting O(ϵ)O(\epsilon) terms gives the perturbed swimming orientation

<𝒑𝟏>=G1dMsdG𝒛^Ms𝒒𝟏H𝒒𝒔,<\bm{p_{1}}>=G_{1}\frac{dM_{s}}{dG}\hat{\bm{z}}-M_{s}\frac{\bm{q_{1}}^{H}}{\bm{q_{s}}}, (45)

here, 𝒒1H=[𝒒1x,𝒒1y]\bm{q}_{1}^{H}=[\bm{q}_{1}^{x},\bm{q}_{1}^{y}] represents the horizontal component of the perturbed radiative flux 𝒒1\bm{q}_{1}. By substituting the value of <𝒑𝟏><\bm{p_{1}}> from Eq.(45)(\ref{46}) into Eq.(40)(\ref{41}) and simplifying, we obtain:

n1t+Vcz(Msn1+nsG1dMsdG)VcnsMsqs(q1xx+q1yy)+W1dnsdz=2n1.\frac{\partial{n_{1}}}{\partial{t}}+V_{c}\frac{\partial}{\partial z}\left(M_{s}n_{1}+n_{s}G_{1}\frac{dM_{s}}{dG}\right)-V_{c}n_{s}\frac{M_{s}}{q_{s}}\left(\frac{\partial q_{1}^{x}}{\partial x}+\frac{\partial q_{1}^{y}}{\partial y}\right)+W_{1}\frac{dn_{s}}{dz}=\nabla^{2}n_{1}. (46)

By eliminating PeP_{e} and the horizontal component of u1u_{1}, equations (39) and (40) can be simplified to three equations for the perturbed variables: the vertical component of the velocity w1w_{1}, the vertical component of the vorticity ζ1\zeta_{1} (which is defined as ζ𝒛^\zeta\cdot\hat{\bm{z}}), and the concentration n1n_{1}. These variables can be further decomposed into normal modes as:

[W1,ζ1,n1]=[W~(z),Z~(z),N~(z)]exp(σt+i(lx+my)).[W_{1},\zeta_{1},n_{1}]=[\tilde{W}(z),\tilde{Z}(z),\tilde{N}(z)]\exp{(\sigma t+i(lx+my))}. (47)

The equation governing the perturbed intensity L1L_{1} can be written as:

ξL1x+ηL1y+νL1z+κ(nsL1+n1Ls)=ωκ4π(nsM1+Gsn1+Aν(nsqsz^qsn1))κLsn1,\xi\frac{\partial L_{1}}{\partial x}+\eta\frac{\partial L_{1}}{\partial y}+\nu\frac{\partial L_{1}}{\partial z}+\kappa(n_{s}L_{1}+n_{1}L_{s})=\frac{\omega\kappa}{4\pi}(n_{s}M_{1}+G_{s}n_{1}+A\nu(n_{s}q_{s}\cdot\hat{z}-q_{s}n_{1}))-\kappa L_{s}n_{1}, (48)

with boundary conditions

L1(x,y,z=1,ξ,η,ν)=0,θ[π/2,π],ϕ[0,2π],L_{1}(x,y,z=1,\xi,\eta,\nu)=0,\qquad\theta\in[\pi/2,\pi],~{}~{}\phi\in[0,2\pi], (49a)
L1(x,y,z=0,ξ,η,ν)=0,θ[0,π/2],ϕ[0,2π].L_{1}(x,y,z=0,\xi,\eta,\nu)=0,\qquad\theta\in[0,\pi/2],~{}~{}\phi\in[0,2\pi]. (49b)

The Eq. (47)(\ref{48}) implies that L1dL_{1}^{d} can be represented by the following expression:

L1d=Ψd(z,ξ,η,ν)exp(σt+i(lx+my)).L_{1}^{d}=\Psi^{d}(z,\xi,\eta,\nu)\exp{(\sigma t+i(lx+my))}.

From Eqs. (41) and (42), we get

G1c=[Ltexp(z1κns(z)𝑑z)(1zκn1𝑑z)]exp(σt+i(lx+my))=𝔾c(z)exp(σt+i(lx+my)),G_{1}^{c}=\left[L_{t}\exp\left(-\int_{z}^{1}\kappa n_{s}(z^{\prime})dz^{\prime}\right)\left(\int_{1}^{z}\kappa n_{1}dz^{\prime}\right)\right]\exp{(\sigma t+i(lx+my))}=\mathbb{G}^{c}(z)\exp{(\sigma t+i(lx+my))}, (50)

and

G1d=𝔾d(z)exp(σt+i(lx+my))=(04πΨd(z,ξ,η,ν)𝑑Ω)exp(σt+i(lx+my)),G_{1}^{d}=\mathbb{G}^{d}(z)\exp{(\sigma t+i(lx+my))}=\left(\int_{0}^{4\pi}\Psi^{d}(z,\xi,\eta,\nu)d\Omega\right)\exp{(\sigma t+i(lx+my))}, (51)

where 𝔾(z)=𝔾c(z)+𝔾d(z)\mathbb{G}(z)=\mathbb{G}^{c}(z)+\mathbb{G}^{d}(z) is the perturbed total intensity. Similarly from Eqs. (43) and (44), we have

𝒒1=[q1x,q1y,q1z]=[P(z),Q(z),S(z)]exp[σt+i(lx+my)],\bm{q}_{1}=[q_{1}^{x},q_{1}^{y},q_{1}^{z}]=[P(z),Q(z),S(z)]\exp{[\sigma t+i(lx+my)]},

where

[P(z),Q(z)]=04π[ξ,η]Ψd(z,ξ,η,ν)𝑑Ω,[P(z),Q(z)]=\int_{0}^{4\pi}[\xi,\eta]\Psi^{d}(z,\xi,\eta,\nu)d\Omega,

and

S(z)=04π[Ψc(z,ξ,η,ν)+Ψd(z,ξ,η,ν)]ν𝑑Ω.S(z)=\int_{0}^{4\pi}[\Psi^{c}(z,\xi,\eta,\nu)+\Psi^{d}(z,\xi,\eta,\nu)]\nu d\Omega.

Note that the P(z)P(z) and Q(z)Q(z) appear due to scattering. On the other hand, S(z)S(z) appears due to anisotropic scattering which becomes zero in the case of isotropic scattering.

Now Ψd\Psi^{d} satisfies

dΨddz+(i(lξ+mη)+κns)νΨd=ωκ4πν(ns𝔾+GsN~(z)+Aν(nsSqsN~(z)))κνIsN~(z),\frac{d\Psi^{d}}{dz}+\frac{(i(l\xi+m\eta)+\kappa n_{s})}{\nu}\Psi^{d}=\frac{\omega\kappa}{4\pi\nu}(n_{s}\mathcal{\mathbb{G}}+G_{s}\tilde{N}(z)+A\nu(n_{s}S-q_{s}\tilde{N}(z)))-\frac{\kappa}{\nu}I_{s}\tilde{N}(z), (52)

subject to the boundary conditions

Ψd(1,ξ,η,ν)=0,θ[π/2,π],ϕ[0,2π],\Psi^{d}(1,\xi,\eta,\nu)=0,\qquad\theta\in[\pi/2,\pi],~{}~{}\phi\in[0,2\pi], (53a)
Ψd(0,ξ,η,ν)=0,θ[0,π/2],ϕ[0,2π].\Psi^{d}(0,\xi,\eta,\nu)=0,\qquad\theta\in[0,\pi/2],~{}~{}\phi\in[0,2\pi]. (53b)

The stability equations reformed as

(σSc1+k2d2dz2)(d2dz2k2)W~=Rk2N~(z),\left(\sigma S_{c}^{-1}+k^{2}-\frac{d^{2}}{dz^{2}}\right)\left(\frac{d^{2}}{dz^{2}}-k^{2}\right)\tilde{W}=Rk^{2}\tilde{N}(z), (54)
(σSc1+k2d2dz2)Z~(z)=TadW~dz\left(\sigma S_{c}^{-1}+k^{2}-\frac{d^{2}}{dz^{2}}\right)\tilde{Z}(z)=\sqrt{T_{a}}\frac{d\tilde{W}}{dz} (55)
(σ+k2d2dz2)N~(z)+Vcddz(MsN~(z)+ns𝔾dMsdG)iVcnsMsqs(lP+mQ)=dnsdzW~(z),\left(\sigma+k^{2}-\frac{d^{2}}{dz^{2}}\right)\tilde{N}(z)+V_{c}\frac{d}{dz}\left(M_{s}\tilde{N}(z)+n_{s}\mathbb{G}\frac{dM_{s}}{dG}\right)-i\frac{V_{c}n_{s}M_{s}}{q_{s}}(lP+mQ)=-\frac{dn_{s}}{dz}\tilde{W}(z), (56)

with boundary conditions

W~(z)=dW~(z)dz=Z~(z)=dN~(z)dzVcMsN~(z)nsVc𝔾dMsdG=0atz=0,\tilde{W}(z)=\frac{d\tilde{W}(z)}{dz}=\tilde{Z}(z)=\frac{d\tilde{N}(z)}{dz}-V_{c}M_{s}\tilde{N}(z)-n_{s}V_{c}\mathbb{G}\frac{dM_{s}}{dG}=0\quad at\quad z=0, (57)
W~(z)=dW~(z)dz=dZ~(z)dz=dN~(z)dzVcMsN~(z)nsVc𝔾dMsdG=0atz=1.\tilde{W}(z)=\frac{d\tilde{W}(z)}{dz}=\frac{d\tilde{Z}(z)}{dz}=\frac{d\tilde{N}(z)}{dz}-V_{c}M_{s}\tilde{N}(z)-n_{s}V_{c}\mathbb{G}\frac{dM_{s}}{dG}=0\quad at\quad z=1. (58)

n the given context, the non-dimensional wavenumber is represented by kk, which is determined by taking the square root of the sum of the squares of ll and mm. Equations (57)-(58) form an eigenvalue problem, where σ\sigma is described as a function of several dimensionless parameters such as VcV_{c}, κ\kappa, ω\omega, AA, ll, mm, and RR. Equation (56) can be equivalently expressed as follows:

D2N~(z)Λ3(z)DN~(z)(σ+k2+Λ2(z))N~(z)Λ1(z)1zN~(z)𝑑zΛ0(z)=DnsW~,D^{2}\tilde{N}(z)-\Lambda_{3}(z)D\tilde{N}(z)-(\sigma+k^{2}+\Lambda_{2}(z))\tilde{N}(z)-\Lambda_{1}(z)\int_{1}^{z}\tilde{N}(z)dz-\Lambda_{0}(z)=Dn_{s}\tilde{W}, (59)

where

Λ0(z)=VcD(ns𝔾ddMsdG)ιVcnsMsqs(lP+mQ),\Lambda_{0}(z)=V_{c}D\left(n_{s}\mathbb{G}^{d}\frac{dM_{s}}{dG}\right)-\iota\frac{V_{c}n_{s}M_{s}}{q_{s}}(lP+mQ), (60a)
Λ(z)=κVcD(nsGscdMsdG)\Lambda(z)=\kappa V_{c}D\left(n_{s}G_{s}^{c}\frac{dM_{s}}{dG}\right) (60b)
Λ2(z)=2κVcnsGscdMsdG+VcdMsdMDGsd,\Lambda_{2}(z)=2\kappa V_{c}n_{s}G_{s}^{c}\frac{dM_{s}}{dG}+V_{c}\frac{dM_{s}}{dM}DG_{s}^{d}, (60c)
Λ3(z)=VcMs.\Lambda_{3}(z)=V_{c}M_{s}. (60d)

Introducing a novel variable denoted as

Θ~(z)=1zN~(z)𝑑z,\tilde{\Theta}(z)=\int_{1}^{z}\tilde{N}(z^{\prime})dz^{\prime}, (61)

the linear stability equations can be reformulated as

D4W~(σSc1+k2)D2W~(σSc1+k2)W~=Rk2DΘ~,D^{4}\tilde{W}-\left(\sigma S_{c}^{-1}+k^{2}\right)D^{2}\tilde{W}-\left(\sigma S_{c}^{-1}+k^{2}\right)\tilde{W}=Rk^{2}D\tilde{\Theta}, (62)
D2Z~(z)(σSc1+k2D2)Z~(z)=TaDW~D^{2}\tilde{Z}(z)-\left(\sigma S_{c}^{-1}+k^{2}-D^{2}\right)\tilde{Z}(z)=\sqrt{T_{a}}D\tilde{W} (63)
D3Θ~Λ3(z)D2Θ~(σ+k2+Λ2(z))DΘ~Λ1(z)Θ~Λ0(z)=DnsW~.D^{3}\tilde{\Theta}-\Lambda_{3}(z)D^{2}\tilde{\Theta}-(\sigma+k^{2}+\Lambda_{2}(z))D\tilde{\Theta}-\Lambda_{1}(z)\tilde{\Theta}-\Lambda_{0}(z)=Dn_{s}\tilde{W}. (64)

The boundary conditions remain unchanged except for the flux boundary condition, given by the equation

D2N~(z)Λ2DN~(z)nsVc𝔾dMsdG=0atz=0,1.D^{2}\tilde{N}(z)-\Lambda_{2}D\tilde{N}(z)-n_{s}V_{c}\mathbb{G}\frac{dM_{s}}{dG}=0\quad at\quad z=0,1. (65)

Furthermore, an additional boundary condition is introduced

Θ~(z)=0,atz=1.\tilde{\Theta}(z)=0,\quad at\quad z=1. (66)

V SOLUTION PROCEDURE

To obtain the neutral (marginal) stability curves or the growth rate Re(σ)Re(\sigma) as a function of R in the (k, R)-plane for a fixed parameter set, a fourth-order accurate finite-difference scheme based on Newton-Raphson-Kantorovich (NRK) iterations is utilized to solve equations 62 to 64. By analysing the resulting graph, points where Re(σ)=0Re(\sigma)=0 are identified, representing the marginal (neutral) stability curve. On this curve, if Im(σ)=0Im(\sigma)=0, it indicates a stationary (non-oscillatory) bioconvective solution, while non-zero Im(σ)Im(\sigma) indicates oscillatory solutions. If the most unstable mode remains on the oscillatory branch of the neutral curve, it signifies overstability.

When oscillatory solutions are present, there is a common point (kb)(k_{b}) where the stationary and oscillatory branches intersect, and the oscillatory branch includes points with kkbk\leq k_{b}. Among the points on the neutral curve R(n)(k)R^{(n)}(k) (where n = 1, 2, 3, …), a particular most unstable mode (kc,Rc)(k_{c},R_{c}) is selected. The wavelength of the initial disturbance is then determined as λc=2π/kc\lambda_{c}=2\pi/k_{c}. The bioconvective solution is referred to as mode n if it can be organized into n convection cells such that one cell overlaps another vertically.

VI NUMERICAL RESULTS

To identify the most unstable mode starting from an initial equilibrium solution, we employ a specific set of fixed parameters. These parameters include Sc=20S_{c}=20, Gc=1G_{c}=1, Vc=10V_{c}=10, 1515, 2020, with ω\omega ranging from 0 to 11, κ\kappa values of 0.50.5 and 11, and AA values of 0, 0.40.4, and 0.80.8. The Taylor number, which represents the rotation rate, is varied from lower non-zero values to higher values. To investigate the impact of rotation on a forward scattering algal suspension, we focus on specific values of the scattering albedo ω\omega. For the case of κ=0.5\kappa=0.5, we examine the scenarios where ω\omega takes on the values 0.10.1, 0.420.42, and 0.470.47. Similarly, for the case of κ=1\kappa=1, we consider ω\omega values of 0.10.1, 0.580.58, and 0.600.60. When ω=0.1\omega=0.1, the equilibrium state’s sublayer is positioned near the top of the domain. As the value of ω\omega increases, the sublayer’s equilibrium position shifts from the top to approximately three-quarters height, and then further to the middle height of the domain.

VI.1 WHEN SCATTERING IS WEAK

The primary focus of this study is to investigate how rotation influences the initiation of bioconvection. To explore this, the researchers compare the relative effectiveness of self-shading and scattering, specifically by utilizing a lower scattering albedo value denoted as ω\omega. Additionally, the strength of self-shading is manipulated by varying the extinction coefficient κ\kappa, with a higher value of κ=1\kappa=1 indicating strong self-shading and a lower value of κ=0.5\kappa=0.5 representing weak self-shading. Throughout the study, a consistent critical intensity of Gc=1G_{c}=1 is used. The researchers divide the entire study into three distinct cases based on the position of the equilibrium state, namely at the top, three-quarter, and mid-height of the domain. Suitable values of ω\omega are chosen accordingly for each case.

VI.1.1 Vc=20V_{c}=20

(a) κ=0.5\kappa=0.5

This section delves into the examination of bioconvective instability and its correlation with the Taylor number. The study investigates this relationship by considering three distinct values of the forward scattering coefficient AA and three different values of ω\omega, while keeping the parameters VcV_{c} and κ\kappa fixed at 20 and 0.5, respectively. Similar to previous sections, the study is divided into three cases, each based on the position of the equilibrium state within the domain (at the top, three-quarters, and mid-height). The appropriate value of ω\omega is chosen accordingly for each case.

Refer to caption
Figure 4: The marginal stability curves for (a) A=0A=0, (b) A=0.4A=0.4. Here, the other governing parameter values Sc=20,Vc=20,k=0.5S_{c}=20,V_{c}=20,k=0.5, and ω=0.1\omega=0.1 are kept fixed.

When ω=0.1\omega=0.1, the sublayer is located approximately at the top of the domain when A=0A=0. However, as the value of AA increases, the sublayer at the equilibrium state shifts towards the three-quarter height of the domain. In this scenario, we vary the Taylor number (TaT_{a}) from a lower value (e.g., Ta=10T_{a}=10) to a higher value (e.g., Ta=10000T_{a}=10000). At Ta=10T_{a}=10, a stationary mode of the bioconvective solution is observed. As TaT_{a} is increased up to 10000, the behaviour of the solutions remains the same, but the critical Rayleigh number increases (refer to Fig.4(a)). For A=0.4A=0.4 and A=0.8A=0.8, the mode of the bioconvective solution remains unchanged, and as TaT_{a} is increased, the critical Rayleigh number also increases (refer to Fig.4).

Refer to caption
Figure 5: The marginal stability curves for (a) A=0A=0, (b) A=0.4A=0.4. Here, the other governing parameter values Sc=20,Vc=20,k=0.5S_{c}=20,V_{c}=20,k=0.5, and ω=0.42\omega=0.42 are kept fixed.

When ω=0.42\omega=0.42, the sublayer in the equilibrium state is located at the three-quarter height of the suspension for A=0A=0. At Ta=10T_{a}=10, a stationary mode of the bioconvective solution is observed. As TaT_{a} is increased to 1000, an oscillatory branch splits from the stationary branch of the marginal stability curve. However, the most unstable solution still occurs on the stationary branch of the neutral curve, resulting in a stationary solution and an increased critical Rayleigh number. Similarly, at Ta=10000T_{a}=10000, the same nature of the marginal stability curve and a bioconvective solution is observed, with an increased critical Rayleigh number. As AA increases, the sublayer at the equilibrium state shifts towards the mid-height of the domain. For A=0.4A=0.4 and A=0.8A=0.8, the mode of the bioconvective solution remains unchanged, and as TaT_{a} is increased, the critical Rayleigh number also increases (refer to Fig. 5).

Refer to caption
Figure 6: The marginal stability curves for (a) A=0A=0, (b) A=0.4A=0.4. Here, the other governing parameter values Sc=20,Vc=20,k=0.5S_{c}=20,V_{c}=20,k=0.5, and ω=0.47\omega=0.47 are kept fixed.

When ω=0.47\omega=0.47, the sublayer in the equilibrium state is located approximately at the middle height of the domain for A=0A=0. At Ta=10T_{a}=10, a stationary mode of the bioconvective solution is observed. As TaT_{a} is increased to 1000, the mode of the bioconvective instability remains the same, but the critical Rayleigh number increases. As TaT_{a} reaches 10000, an oscillatory branch bifurcates from the stationary branch of the marginal stability curve at k1.6k\approx 1.6 and exists for all k<1.6k<1.6. However, the most unstable bioconvective solution still occurs on the stationary branch, resulting in a stationary solution. As AA increases, the sublayer at the equilibrium state shifts towards the bottom of the domain. For A=0.4A=0.4, the stationary nature of the solution is observed at Ta=10T_{a}=10, and this nature of the solution remains unchanged for all higher values of the Taylor number. However, the critical Rayleigh number increases as TaT_{a} is increased. A similar nature of the bioconvective solution is also observed for A=0.8A=0.8 (see Fig. 6).

The numerical results of this section are summarized in Table 1.

Table 1: The table shows the numerical results of bioconvective solutions for a constant Vc=V_{c}=20 and κ=\kappa=0.5. The values are presented for different increments in TaT_{a}, while keeping all other parameters unchanged. The presence of a star symbol indicates that the R(1)(k)R^{(1)}(k) branch of the neutral curve exhibits oscillatory behaviour.
ω\omega TaT_{a} A=0A=0 A=0.4A=0.4 A=0.8A=0.8
λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma) λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma) λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma)
0 4.23 194.18 0 4.23 193.15 0 4.39 192.07 0
0.1 1000 2.17 379.88 0 2.17 379.37 0 2.25 379.09 0
10000 1.22 988.77 0 1.17 990.33 0 1.18 991.91 0
0 2.98 217.27 0 3.15 222.85 0 3.44 231.89 0
0.42 1000 1.17 452.01 0 1.79 480.72 0 1.80 522.28 0
10000 1.12 1332.85 0 1.13 1417.52 0 1.14 1533.90 0
0 2.53 432.26 0 2.35 616.06 0 2.17 919.01 0
0.47 1000 1.78 679.23 0 1.79 879.18 0 1.76 1207.16 0
10000 1.20 1652.82 0 1.21 1989.33 0 1.23 2524.01 0

(b) κ=1\kappa=1 In this section, we examine the influence of the Taylor number on bioconvective instability for three different values of the forward scattering coefficient AA at three different values of ω\omega. The investigation is conducted with a set of fixed parameters, namely Vc=20V_{c}=20 and κ=1\kappa=1.

Refer to caption
Figure 7: The marginal stability curves for (a) A=0A=0, (b) A=0.4A=0.4. Here, the other governing parameter values Sc=20,Vc=20,k=1S_{c}=20,V_{c}=20,k=1, and ω=0.1\omega=0.1 are kept fixed.

For ω=0.1\omega=0.1, the sublayer occurs approximately at the top of the domain when A=0A=0. As the value of AA increases, the sublayer at the equilibrium state shifts towards the three-quarter height of the domain. In this case, we vary the Taylor number from a lower value of Ta=10T_{a}=10 to a higher value of Ta=10000T_{a}=10000. At Ta=10T_{a}=10, a stationary mode of the bio-convective solution is observed, and as TaT_{a} is increased up to 10000, the same behaviour of solutions is observed with an increase in the critical Rayleigh number (refer to Fig. 7(a)). For A=0.4A=0.4 and A=0.8A=0.8, the mode of the bio-convective solution remains unchanged, and as TaT_{a} is increased, the critical Rayleigh number also increases (see Fig. 4).

Refer to caption
Figure 8: The marginal stability curves for (a) A=0A=0, (b) A=0.8A=0.8. Here, the other governing parameter values Sc=20,Vc=20,k=1S_{c}=20,V_{c}=20,k=1, and ω=0.58\omega=0.58 are kept fixed.

For ω=0.58\omega=0.58, the sublayer in the equilibrium state occurs at the three-quarter height of the suspension when A=0A=0. At Ta=10T_{a}=10, an oscillatory branch splits from the stationary branch of the marginal stability curve at k3k\approx 3 and exists for all k<3k<3. In this case, the most unstable solution occurs on the oscillatory branch of the marginal stability curve, resulting in an overstable solution. As TaT_{a} is increased to 1000, another oscillatory branch splits from the stationary branch, but the most unstable solution occurs on the stationary branch of the neutral curve, leading to a stationary solution and an increase in the critical Rayleigh number. For Ta=10000T_{a}=10000, the same nature of the marginal stability curve and a bioconvective solution is observed, and the critical Rayleigh number continues to increase (refer to Fig. 8(a)). As the value of AA increases, the sublayer at the equilibrium state shifts towards the mid-height of the domain. For A=0.4A=0.4, the nature of the bioconvective solution and the critical Rayleigh number remain unchanged compared to the case when A=0A=0. For A=0.8A=0.8 and Ta=0.4T_{a}=0.4, a stationary nature of the bioconvective instability is observed for all values of the Taylor number, and the critical Rayleigh number increases as TaT_{a} is increased, similar to the previous case (see Fig. 8).

Refer to caption
Figure 9: The marginal stability curves for (a) A=0A=0, (b) A=0.4A=0.4. Here, the other governing parameter values Sc=20,Vc=20,k=1S_{c}=20,V_{c}=20,k=1, and ω=0.6\omega=0.6 are kept fixed.

For ω=0.6\omega=0.6, the sublayer in the equilibrium state occurs at the three-quarter height of the suspension when A=0A=0. At Ta=10T_{a}=10, an oscillatory branch splits from the stationary branch of the marginal stability curve at k3k\approx 3 and exists for all k<3k<3. In this case, the most unstable solution occurs on the oscillatory branch of the marginal stability curve, resulting in an overstable solution. As TaT_{a} is increased to 1000, another oscillatory branch splits from the stationary branch, but the most unstable solution occurs on the stationary branch of the neutral curve, leading to a stationary solution and an increase in the critical Rayleigh number. For Ta=10000T_{a}=10000, the same nature of the marginal stability curve and a bioconvective solution is observed, and the critical Rayleigh number continues to increase (refer to Fig. 9(a)). As the value of AA increases, the sublayer at the equilibrium state shifts towards the mid-height of the domain. For A=0.4A=0.4 and Ta=10T_{a}=10, an oscillatory branch splits from the stationary branch of the marginal stability curve at k2.2k\approx 2.2 and exists for all k<2.2k<2.2. Here, the most unstable solution occurs on the stationary branch of the marginal stability curve, resulting in a stationary solution. As TaT_{a} is increased to 1000 and, 10000, another oscillatory branch splits from the stationary branch, but the most unstable solution occurs on the stationary branch of the neutral curve, leading to a stationary solution and an increase in the critical Rayleigh number (refer to Fig. 9(b)). For Ta=0.8T_{a}=0.8, a stationary nature of the bioconvective instability is observed for all values of the Taylor number, and the critical Rayleigh number increases as TaT_{a} is increased, similar to the previous cases (refer to Table 2).

The numerical results of this section are presented in Table 2.

Table 2: The table shows the numerical results of bioconvective solutions for a constant Vc=V_{c}=20 and κ=\kappa=0.5. The values are presented for different increments in TaT_{a}, while keeping all other parameters unchanged. The presence of a star symbol indicates that the R(1)(k)R^{(1)}(k) branch of the neutral curve exhibits oscillatory behaviour.
ω\omega TaT_{a} A=0A=0 A=0.4A=0.4 A=0.8A=0.8
λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma) λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma) λc\lambda_{c} RcR_{c} Im(σ)Im(\sigma)
0 2.78 239.43 0 2.78 239.06 0 2.76 238.56 0
0.1 1000 1.92 395.58 0 1.92 395.67 0 1.92 395.64 0
10000 1.18 922.98 0 1.19 924.11 0 1.19 924.11 0
0 2.69 312.75 8.79 2.82 304.94 7.65 1.98 326.80 0
0.58 1000 1.48 498.04 0 1.50 496.53 0 1.55 507.05 0
10000 1.03 1168.45 0 1.05 1193.34 0 1.08 1238.14 0
0 3.09 295.61 7.16 2.09 310.62 0 2.05 475.61 0
0.6 1000 1.54 464.30 0 1.61 481.28 0 1.66 644.02 0
10000 1.07 1126.63 0 1.11 1164.71 0 1.15 1376.95 0

VII Conclusion

The proposed phototaxis model investigates the impact of rotation on the initiation of light-induced bioconvection in a forward-scattering algal suspension illuminated by collimated flux. The model incorporates a linear anisotropic scattering coefficient.

When the forward scattering coefficient AA is increased in a uniform suspension, several effects are observed. Firstly, the total intensity of light undergoes changes in the equilibrium state. Specifically, it decreases in the upper half of the suspension and increases in the lower half. This indicates a redistribution of light intensity within the suspension. Furthermore, the critical value of total intensity, which is the threshold for the onset of bioconvection, exhibits variations. In the lower half of the suspension, the critical intensity decreases as AA is increased, implying that bioconvection is more likely to occur at lower light intensities. Conversely, in the upper half of the suspension, the critical intensity increases with increasing AA, indicating a higher threshold for bioconvection initiation. Additionally, the position of the sublayer, where cells accumulate, is influenced by the value of AA. As AA is increased, the sublayer at the equilibrium state shifts towards the bottom of the domain. This suggests that the presence of forward scattering promotes the accumulation of cells in the lower region of the suspension.

The linear stability analysis predicts both stationary and oscillatory disturbances. The nature of the oscillatory bioconvective solutions appears to depend on the location of the sublayer at the equilibrium state. When the sublayer forms near the top of the domain, the bioconvective instability remains stationary for all values of the Taylor number TAT_{A} and forward scattering coefficient AA. However, when the sublayer forms at the three-quarter height and mid-height of the domain, an overstable nature of the solution is observed for lower values of the Taylor number and forward scattering coefficient.

Furthermore, it is observed that the critical wavelength decreases and the corresponding Rayleigh number increases as the Taylor number increases. This suggests that the suspension becomes more stable as the rotation rate increases.

The proposed model appears to be more realistic than the model presented by Panda et al. due to the consideration of an illuminating source composed of both diffuse and oblique collimated flux, which better represents how sunlight interacts with an algal suspension in natural environments. In contrast, Panda et al. used a vertical collimated flux in addition to a diffuse flux, which may not accurately reflect the behaviour of photo-gyro-gravitactic algae in their natural habitat. However, it is important to note that there is currently no quantitative analysis or experimental data available to directly compare with the theoretical developments presented in the proposed model.

Data Availability

The article includes the necessary information and analysis to support the conclusions and results reported.

References

  • Platt (1961) J. R. Platt, “" bioconvection patterns" in cultures of free-swimming organisms,” Science 133, 1766–1767 (1961).
  • Pedley and Kessler (1992) T. J. Pedley and J. O. Kessler, “Hydrodynamic phenomena in suspensions of swimming microorganisms,” Annual Review of Fluid Mechanics 24, 313–358 (1992).
  • Hill and Pedley (2005) N. A. Hill and T. J. Pedley, “Bioconvection,” Fluid Dynamics Research 37, 1 (2005).
  • Bees (2020) M. A. Bees, “Advances in bioconvection,” Annual Review of Fluid Mechanics 52, 449–476 (2020).
  • Javadi et al. (2020) A. Javadi, J. Arrieta, I. Tuval,  and M. Polin, “Photo-bioconvection: towards light control of flows in active suspensions,” Philosophical Transactions of the Royal Society A 378, 20190523 (2020).
  • Wager (1911) H. Wager, “Vii. on the effect of gravity upon the movements and aggregation of euglena viridis, ehrb., and other micro-organisms,” Philosophical Transactions of the Royal Society of London. Series B, Containing Papers of a Biological Character 201, 333–390 (1911).
  • Kitsunezaki, Komori, and Harumoto (2007) S. Kitsunezaki, R. Komori,  and T. Harumoto, “Bioconvection and front formation of paramecium tetraurelia,” Physical Review E 76, 046301 (2007).
  • Kessler (1985) J. O. Kessler, “Co-operative and concentrative phenomena of swimming micro-organisms,” Contemporary Physics 26, 147–166 (1985).
  • Williams and Bees (2011) C. R. Williams and M. A. Bees, “A tale of three taxes: photo-gyro-gravitactic bioconvection,” Journal of Experimental Biology 214, 2398–2408 (2011).
  • Kessler (1989) J. Kessler, “Path and pattern-the mutual dynamics of swimming cells and their environment,” Comments Theor. Biol. 1, 85–108 (1989).
  • Straughan (1993) B. Straughan, Mathematical aspects of penetrative convection (CRC Press, 1993).
  • Ghorai and Hill (2005) S. Ghorai and N. A. Hill, “Penetrative phototactic bioconvection,” Physics of fluids 17, 074101 (2005).
  • Panda and Singh (2016) M. K. Panda and R. Singh, “Penetrative phototactic bioconvection in a two-dimensional non-scattering suspension,” Physics of Fluids 28, 054105 (2016).
  • Vincent and Hill (1996) R. V. Vincent and N. A. Hill, “Bioconvection in a suspension of phototactic algae,” Journal of Fluid Mechanics 327, 343–371 (1996).
  • Ghorai, Panda, and Hill (2010) S. Ghorai, M. K. Panda,  and N. A. Hill, “Bioconvection in a suspension of isotropically scattering phototactic algae,” Physics of Fluids 22, 071901 (2010).
  • Panda and Ghorai (2013) M. K. Panda and S. Ghorai, “Penetrative phototactic bioconvection in an isotropic scattering suspension,” Physics of Fluids 25, 071902 (2013).
  • Panda et al. (2016) M. K. Panda, R. Singh, A. C. Mishra,  and S. K. Mohanty, “Effects of both diffuse and collimated incident radiation on phototactic bioconvection,” Physics of Fluids 28, 124104 (2016).
  • Panda (2020) M. K. Panda, “Effects of anisotropic scattering on the onset of phototactic bioconvection with diffuse and collimated irradiation,” Physics of Fluids 32, 091903 (2020).
  • Panda, Sharma, and Kumar (2022) M. K. Panda, P. Sharma,  and S. Kumar, “Effects of oblique irradiation on the onset of phototactic bioconvection,” Physics of Fluids 34, 024108 (2022).
  • Kumar (2022) S. Kumar, “Phototactic isotropic scattering bioconvection with oblique irradiation,” Physics of Fluids 34, 114125 (2022).
  • Kumar (2023a) S. Kumar, “Effect of rotation on the suspension of phototactic bioconvection,” Physics of Fluids 35 (2023a).
  • Panda and Rajput (2023) M. K. Panda and S. K. Rajput, “Phototactic bioconvection with the combined effect of diffuse and oblique collimated flux on an algal suspension,” Physics of Fluids 35 (2023).
  • Ghorai and Panda (2013) S. Ghorai and M. K. Panda, “Bioconvection in an anisotropic scattering suspension of phototactic algae,” European Journal of Mechanics-B/Fluids 41, 81–93 (2013).
  • Häder (1987) D.-P. Häder, “Polarotaxis, gravitaxis and vertical phototaxis in the green flagellate, euglena gracilis,” Archives of microbiology 147, 179–183 (1987).
  • Hill and Häder (1997) N. A. Hill and D.-P. Häder, “A biased random walk model for the trajectories of swimming micro-organisms,” Journal of theoretical biology 186, 503–526 (1997).
  • Cash and Moore (1980) J. R. Cash and D. R. Moore, “A high order method for the numerical solution of two-point boundary value problems,” BIT Numerical Mathematics 20, 44–52 (1980).
  • Kessler (1986) J. O. Kessler, “The external dynamics of swimming micro-organisms,” Progress in phycological research 4, 258–307 (1986).
  • Kessler and Hill (1997) J. O. Kessler and N. A. Hill, “Complementarity of physics, biology and geometry in the dynamics of swimming micro-organisms,” in Physics of biological systems (Springer, 1997) pp. 325–340.
  • Kage et al. (2013) A. Kage, C. Hosoya, S. A. Baba,  and Y. Mogami, “Drastic reorganization of the bioconvection pattern of chlamydomonas: quantitative analysis of the pattern transition response,” Journal of Experimental Biology 216, 4557–4566 (2013).
  • Mendelson and Lega (1998) N. H. Mendelson and J. Lega, “A complex pattern of traveling stripes is produced by swimming cells of bacillus subtilis,” Journal of bacteriology 180, 3285–3294 (1998).
  • Gittleson and Jahn (1968) S. M. Gittleson and T. L. Jahn, “Pattern swimming by polytomella agilis,” The American Naturalist 102, 413–425 (1968).
  • Khan et al. (2017) N. S. Khan, T. Gul, M. A. Khan, E. Bonyah,  and S. Islam, “Mixed convection in gravity-driven thin film non-newtonian nanofluids flow with gyrotactic microorganisms,” Results in physics 7, 4033–4049 (2017).
  • Hayat, Alsaedi et al. (2021) T. Hayat, A. Alsaedi, et al., “Development of bioconvection flow of nanomaterial with melting effects,” Chaos, Solitons & Fractals 148, 111015 (2021).
  • Incropera, Wagner, and Houf (1981) F. P. Incropera, T. R. Wagner,  and W. G. Houf, “A comparison of predictions and measurements of the radiation field in a shallow water layer,” Water Resources Research 17, 142–148 (1981).
  • Daniel, Laurendeau, and Incropera (1979) K. J. Daniel, N. M. Laurendeau,  and F. P. Incropera, “Prediction of radiation absorption and scattering in turbid water bodies,”   (1979).
  • Hill, Pedley, and Kessler (1989) N. A. Hill, T. J. Pedley,  and J. O. Kessler, “Growth of bioconvection patterns in a suspension of gyrotactic micro-organisms in a layer of finite depth,” Journal of Fluid Mechanics 208, 509–543 (1989).
  • Modest and Mazumder (2021) M. F. Modest and S. Mazumder, Radiative heat transfer (Academic press, 2021).
  • Chandrasekhar (1960) S. Chandrasekhar, “Radiative transfer dover publications inc,” New York  (1960).
  • Ghorai and Singh (2009) S. Ghorai and R. Singh, “Linear stability analysis of gyrotactic plumes,” Physics of Fluids 21, 081901 (2009).
  • Press (1992) W. H. Press, “Numerical recipes in fortran.” The Art of Scientific Computing.  (1992).
  • Kumar (2023b) S. Kumar, “Isotropic scattering with a rigid upper surface at the onset of phototactic bioconvection,” Physics of Fluids 35, 024106 (2023b).