Open Access

Oscillatory flow of a Casson fluid in an elastic tube with variable cross section


Cite

Introduction

Recently, the study of flow with periodic variations has attracted much attention of researchers due to its various physiological and engineering applications. It helps to understand the characteristics of the blood flow through arteries. Oscillatory motion of a viscous liquid in a thin-walled elastic tube is investigated by Womersley (1955). Further, Womersley (1957) studied the elastic tube theory of pulse transmission and oscillatory flow in mammalian arteries. Rubinow and Keller (1972) analyzed the flow of a viscous fluid in an elastic tube with application to blood flow. Ramachandra Rao and Devanathan (1973) studied the pulsatile flow in tubes of varying cross section. Taylor and Gerrard (1977) presented a mathematical model to analyze the blood flow through arteries and expressed the different pressure radius relationships for elastic tube. Kaimal (1981) analyzed viscoelastic properties on oscillatory flow with consideration of pulsatile nature of tube wall. Furthermore, the significant effect of elasticity of the fluid and pulsation of the tube wall on flow characteristics was observed.

Ramachandra Rao (1983) investigated the oscillatory flow in an elastic tube of variable cross section. Furthermore, Ramachandra Rao (1983) studied the unsteady flow with attenuation in a fluid-filled elastic tube with a stenosis. An analytical solution by the method of linear approximation to describe the velocity fields for laminar periodic flow through porous walls is proposed by Chang et al. (1989). Sankar and Jayaraman (2001) discussed the nonlinear analysis of oscillatory flow in the annulus of an elastic tube. Misra and Ghosh (2003) considered the viscous fluid in a porous elastic vessel with variable cross section to explain the application of haemodynamic flows. In addition, they noticed that the velocity distribution in a small vessel depends significantly on geometry of the wall and its elastic nature.

Vajravelu et al. (2011) considered the case of inserting a catheter into an elastic tube to observe the changes in blood flow pattern by taking Herschel–Bulkley fluid. Unsteady flow of a Jeffrey fluid in an elastic tube with a stenosis was studied by Sreenadh et al. (2012). Omer and Staples (2012) analyzed the dynamics of pulsatile flows through elastic micro-tubes. Sochi (2014) proposed the expression for the volumetric flow as a function of pressure in elastic tube using two pressure area constitutive relationships. Recently, Vajravelu et al. (2016) analyzed the peristaltic transport of Casson fluid in an elastic tube.

Motivated by the abovementioned studies, oscillatory flow of a Casson fluid in an elastic tube with variable cross section is studied in the current paper. Slowly varying cross section of the tube wall and radial displacement are taken into consideration. Expressions for axial velocity and mass flux in terms of pressure gradient are derived. Different pressure****radius relationships are considered to understand the pressure variation along the different tube geometries. The influence of physical parameters such as elastic parameter, Womersley number and Casson parameter on the variation of excess pressure is presented graphically.

Mathematical Formulation

Consider the oscillatory flow of a Casson fluid through elastic tube with variable cross section. The cylindrical coordinate system (r, θ, z), where z is taken along the flow direction and r is perpendicular to it. The rheological equation for the Casson fluid is given as (Nakamura and Sawada (1988), Elabde et al., (2001))

τij=μb+PyPy2π2π2eij,π>πcμb+PyPy2πc2πc2eij,π<πc$$\begin{array}{} \displaystyle {\tau _{ij}} = \left\{ \begin{array}{l} \left( {{\mu _b} + {{{P_y}} \mathord{\left/ {\vphantom {{{P_y}} {\sqrt {2\pi } }}} \right. } {\sqrt {2\pi } }}} \right)2{e_{ij}},\;\;\;\;\;\pi \gt {\pi _c}\\ \left( {{\mu _b} + {{{P_y}} \mathord{\left/ {\vphantom {{{P_y}} {\sqrt {2{\pi _c}} }}} \right. } {\sqrt {2{\pi _c}} }}} \right)2{e_{ij}},\;\;\;\;\pi \lt {\pi _c} \end{array} \right. \end{array}$$

where τij=eijeijandPy=μB2πβ$\begin{array}{} \displaystyle {\tau _{ij}} = {e_{ij}}{e_{ij}} ~\text{and}~ {P_y} = \frac{{{\mu _B}\sqrt {2\pi } }}{\beta } \end{array}$

here, τij = (i, j)th component of stress tensor, eij = (i, j)th component of deformation rate, π = Product of component of deformation rate with itself, πc = Critical value of this product depends on the non-Newtonian fluid, μb is the plastic dynamic viscosity of the non-Newtonian fluid, β = Casson parameter and Py = Yield stress of the fluid.

The equations governing the motion are

u¯t¯+u¯u¯r¯+w¯u¯z¯=1ρ0p¯r¯+ν1+1β2u¯r¯2+1r¯u¯r¯u¯r¯2+2u¯z¯2$$\begin{array}{} \displaystyle \frac{{\partial \bar u}}{{\partial \bar t}} + \bar u\frac{{\partial \bar u}}{{\partial \bar r}} + \bar w\frac{{\partial \bar u}}{{\partial \bar z}} = - \frac{1}{{{\rho _0}}}\frac{{\partial \bar p}}{{\partial \bar r}} + \nu \left( {1 + \frac{1}{\beta }} \right)\left( {\frac{{{\partial ^2}\bar u}}{{\partial {{\bar r}^2}}} + \frac{1}{{\bar r}}\frac{{\partial \bar u}}{{\partial \bar r}} - \frac{{\bar u}}{{{{\bar r}^2}}} + \frac{{{\partial ^2}\bar u}}{{\partial {{\bar z}^2}}}} \right) \end{array}$$

w¯t¯+u¯w¯r¯+w¯w¯z¯=1ρ0p¯z¯+ν1+1β2w¯r¯2+1r¯w¯r¯+2w¯z¯2$$\begin{array}{} \displaystyle \frac{{\partial \bar w}}{{\partial \bar t}} + \bar u\frac{{\partial \bar w}}{{\partial \bar r}} + \bar w\frac{{\partial \bar w}}{{\partial \bar z}} = - \frac{1}{{{\rho _0}}}\frac{{\partial \bar p}}{{\partial \bar z}} + \nu \left( {1 + \frac{1}{\beta }} \right)\left( {\frac{{{\partial ^2}\bar w}}{{\partial {{\bar r}^2}}} + \frac{1}{{\bar r}}\frac{{\partial \bar w}}{{\partial \bar r}} + \frac{{{\partial ^2}\bar w}}{{\partial {{\bar z}^2}}}} \right) \end{array}$$

here, the radius of the cross section of the tube is r = a(z). The velocity components along radial and axial directions are u, w, respectively, ρ0 is the fluid density, ν is the kinematic viscosity and p is the pressure.

The continuity equation is

u¯r¯+u¯r¯+w¯z¯=0$$\begin{array}{} \displaystyle \frac{{\partial \bar u}}{{\partial \bar r}} + \frac{{\bar u}}{{\bar r}} + \frac{{\partial \bar w}}{{\partial \bar z}} = 0 \end{array}$$

The radial displacement of tube wall ξ is given as (Ramachandra Rao (1983))

2ξ¯t¯2=1hρp2νρ0u¯r¯r=aBρξa2$$\begin{array}{} \displaystyle \frac{{{\partial ^2}\bar \xi }}{{\partial {{\bar t}^2}}} = \frac{1}{{h\rho }}{\left( {p - 2\nu {\rho _0}\frac{{\partial \bar u}}{{\partial \bar r}}} \right)_{r = a}} - \frac{B}{\rho }\frac{\xi }{{{a^2}}} \end{array}$$

where B=E1σ2$\begin{array}{} \displaystyle B = \frac{E}{{\left( {1 - {\sigma ^2}} \right)}} \end{array}$ here, h and ρ are the thickness and density of the material tube, respectively, E is the Youngs modulus and σ is Poisson’s ratio.

The corresponding boundary conditions are

u¯=ξ¯t¯,w¯=0;atr¯=a0S(z)$$\begin{array}{} \displaystyle \bar u = \frac{{\partial \bar \xi }}{{\partial \bar t}},\;\;\;\;\bar w = 0;\;\;\;{\rm{at}}\;\;\;\;\bar r = {a_0}S(z) \end{array}$$

where a0 = Tube radius in the absence of elasticity.

The non-dimensional quantities are

u=u¯εU0,w=w¯U0,t=ωt¯,ξ=ξ¯a0,r=r¯a0,z=εz¯a0,p=εa0p¯ρ0νU0,$$\begin{array}{} \displaystyle u = \frac{{\bar u}}{{\varepsilon {U_0}}},\;\;\;\;w = \frac{{\bar w}}{{{U_0}}},\;\;\;\;t = \omega \bar t,\;\;\;\;\xi = \frac{{\bar \xi }}{{{a_0}}},\\ \displaystyle\,\,\,\, r = \frac{{\bar r}}{{{a_0}}},\;\;\;\;\;z = \frac{{\varepsilon \bar z}}{{{a_0}}},\;\;\;\;p = \frac{{\varepsilon {a_0}\bar p}}{{{\rho _0}\nu {U_0}}}, \end{array}$$

where ε=a0L(1).$\begin{array}{} \displaystyle \varepsilon = \frac{{{a_0}}}{L}( \ll 1). \end{array}$

here, U0 is the characteristic velocity, ω is the frequency of oscillatory flow and L is characteristic length.

Solution of the Problem

Substituting the abovementioned non-dimensional quantities, Eqs. (2)(6) takes the form

ε2α2ut+ε3Ruur+wuz=pr+ε21+1β2ur2+1rurur2+ε41+1β2uz2$$\begin{array}{} \displaystyle {\varepsilon ^2}{\alpha ^2}\frac{{\partial u}}{{\partial t}} + {\varepsilon ^3}R\left( {u\frac{{\partial u}}{{\partial r}} + w\frac{{\partial u}}{{\partial z}}} \right) = - \frac{{\partial p}}{{\partial r}} + {\varepsilon ^2}\left( {1 + \frac{1}{\beta }} \right)\left( {\frac{{{\partial ^2}u}}{{\partial {r^2}}} + \frac{1}{r}\frac{{\partial u}}{{\partial r}} - \frac{u}{{{r^2}}}} \right)\\ \displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad~ + {\varepsilon ^4}\left( {1 + \frac{1}{\beta }} \right)\frac{{{\partial ^2}u}}{{\partial {z^2}}} \end{array}$$

α2wt+εRuwr+wwz=pz+1+1β2wr2+1rwr+ε21+1β2wz2$$\begin{array}{} \displaystyle {\alpha ^2}\frac{{\partial w}}{{\partial t}} + \varepsilon R\left( {u\frac{{\partial w}}{{\partial r}} + w\frac{{\partial w}}{{\partial z}}} \right) = - \frac{{\partial p}}{{\partial z}} + \left( {1 + \frac{1}{\beta }} \right)\left( {\frac{{{\partial ^2}w}}{{\partial {r^2}}} + \frac{1}{r}\frac{{\partial w}}{{\partial r}}} \right)\\ \displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad~\, + {\varepsilon ^2}\left( {1 + \frac{1}{\beta }} \right)\frac{{{\partial ^2}w}}{{\partial {z^2}}} \end{array}$$

ur+ur+wz=0$$\begin{array}{} \displaystyle \frac{{\partial u}}{{\partial r}} + \frac{u}{r} + \frac{{\partial w}}{{\partial z}} = 0 \end{array}$$

ε2ρhρ0a02ξt2=1RSt2p2ε2urr=S1λ12ξS2$$\begin{array}{} \displaystyle {\varepsilon ^2}\frac{{\rho h}}{{{\rho _0}{a_0}}}\frac{{{\partial ^2}\xi }}{{\partial {t^2}}} = \frac{1}{{R{S_t}^2}}{\left( {p - 2{\varepsilon ^2}\frac{{\partial u}}{{\partial r}}} \right)_{r = S}} - \frac{1}{{\lambda _1^2}}\frac{\xi }{{{S^2}}} \end{array}$$

u=Stξtatr=S$$\begin{array}{} \displaystyle u = {S_t}\frac{{\partial \xi }}{{\partial t}}\;\;\;\;\;\;{\rm{at}}\;\;\;\;r = S \end{array}$$

w=0atr=S$$\begin{array}{} \displaystyle w = 0\;\;\;{\rm{at}}\;\;\;\;r = S \end{array}$$

where α=a0ων12,R=U0a0ν,St=ωa0U0,λ12=ω2L2ρ0a0c2ρhandc2=Bρ$\begin{array}{} \displaystyle \alpha = {a_0}{\left( {\frac{\omega }{\nu }} \right)^{\frac{1}{2}}}, R = \frac{{{U_0}{a_0}}}{\nu }, {S_t} = \frac{{\omega {a_0}}}{{{U_0}}},\lambda _1^2 = \frac{{{\omega ^2}{L^2}{\rho _0}{a_0}}}{{{c^2}\rho h}} ~\text{and}~ {c^2} = \frac{B}{\rho } \end{array}$

here, α is the Womersley parameter, R is the Reynolds number and St is the Strouhal number.

Now, neglecting the higher-order terms of ε and the steady oscillatory flow, we have

(u,w,p,ξ)=eit(u~,w~,p~,ξ~)$$\begin{array}{} \displaystyle (u,w,p,\xi ) = {e^{it}}(\tilde u,\tilde w,\tilde p,\tilde \xi ) \end{array}$$

The Eqs.(8)(13) reduces to

p~r=0$$\begin{array}{} \displaystyle \frac{{\partial \tilde p}}{{\partial r}} = 0 \end{array}$$

1+1β2w~r2+1rw~r+λ2w~=p~z$$\begin{array}{} \displaystyle \left( {1 + \frac{1}{\beta }} \right)\left( {\frac{{{\partial ^2}\tilde w}}{{\partial {r^2}}} + \frac{1}{r}\frac{{\partial \tilde w}}{{\partial r}}} \right) + {\lambda ^2}\tilde w = \frac{{\partial \tilde p}}{{\partial z}} \end{array}$$

u~r+u~r+w~z=0$$\begin{array}{} \displaystyle \frac{{\partial \tilde u}}{{\partial r}} + \frac{{\tilde u}}{r} + \frac{{\partial \tilde w}}{{\partial z}} = 0 \end{array}$$

ξ=λ12RSt2S2pe$$\begin{array}{} \displaystyle \xi = \frac{{\lambda _1^2}}{{R{S_t}^2}}{S^2}{p_e} \end{array}$$

u~=iStξ~atr=S$$\begin{array}{} \displaystyle \tilde u = i{S_t}\tilde \xi \;\;\;\;\;\;\;{\rm{at}}\;\;\;\;r = S \end{array}$$

w~=0,atr=S$$\begin{array}{} \displaystyle \tilde w = 0,\;\;\;\;\;{\rm{at}}\;\;\;\;r = S \end{array}$$

here, λ2 = –2 and pe is the excess pressure on the tube wall.

Solving Eq. (16) using the Eq. (20), we have

w~=1λ2dpedz1J0λβ1+βrJ0λβ1+βS$$\begin{array}{} \displaystyle \tilde w = \frac{1}{{{\lambda ^2}}}\frac{{d{p_e}}}{{dz}}\left( {1 - \frac{{{J_0}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} r} \right)}}{{{J_0}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}}} \right) \end{array}$$

The flux across any cross section of the tube is

Q=0S(z)2πrwdr=πS2λ2eitdpedzJ2λβ1+βSλβ1+βSJ0λβ1+β$$\begin{array}{} \quad\! Q = \int\limits_0^{S(z)} {2\pi rw\;\;dr} \\ = - \frac{{\pi {S^2}}}{{{\lambda ^2}}}{e^{it}}\frac{{d{p_e}}}{{dz}}\left[ {\frac{{{J_2}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}}{{\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S{J_0}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} } \right)}}} \right] \end{array}$$

The equations of continuity for longitudinal motion in non-dimensional form effected by area changes are given as (Lighthill (1978))

ωLU0At+Qz=0$$\begin{array}{} \displaystyle \frac{{\omega L}}{{{U_0}}}\frac{{\partial A}}{{\partial t}} + \frac{{\partial Q}}{{\partial z}} = 0 \end{array}$$

here, A is the cross sectional area at any axial point of the tube. The elasticity of the tube wall is taken in to the consideration by expressing a relationship between the excess pressure pp0 and known area A. The pressure radius relation for thick and thin walled elastic tube are presented by (Taylor and Gerrard (1977)).

pt=pA.At$$\begin{array}{} \displaystyle \frac{{\partial p}}{{\partial t}} = \frac{{\partial p}}{{\partial A}}.\frac{{\partial A}}{{\partial t}} \end{array}$$

c2(z)=Aρ0pA=S2ρ0pS$$\begin{array}{} \displaystyle {c^2}(z) = \frac{A}{{{\rho _0}}}\frac{{\partial p}}{{\partial A}} = \frac{S}{{2{\rho _0}}}\frac{{\partial p}}{{\partial S}} \end{array}$$

here, c(z) is the local value of the wave speed. From Eqs. (24) and (25), we have

At=Ac2νU0La02ieitpe$$\begin{array}{} \displaystyle \frac{{\partial A}}{{\partial t}} = \frac{A}{{{c^2}}}\frac{{\nu {U_0}L}}{{{a_0}^2}}i{e^{it}}{p_e} \end{array}$$

using Eqs. (22) and (26) in Eq. (23), we get

d2pedz2+2SdSdza1λβ1+βSdpedzw2L2c2a2λβ1+βSpe=0$$\begin{array}{} \displaystyle \frac{{{d^2}{p_e}}}{{d{z^2}}} + \frac{2}{S}\frac{{dS}}{{dz}}{a_1}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)\frac{{d{p_e}}}{{dz}} - \frac{{{w^2}{L^2}}}{{{c^2}}}{a_2}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right){p_e} = 0 \end{array}$$

where

a1λβ1+βS=J12λβ1+βSJ0λβ1+βSJ2λβ1+βS,a2λβ1+βS=J0λβ1+βSJ2λβ1+βS$$\begin{array}{} \displaystyle {a_1}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right) = \frac{{J_1^2\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}}{{{J_0}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right){J_2}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}} ,\,\,\,\, {a_2}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right) = \frac{{{J_0}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}}{{{J_2}\left( {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S} \right)}} \end{array}$$

Equation (27) is similar to the shell equation for a thin walled elastic tube derived by Ramachandra Rao (1983)

d2pedz2+2SdSdza1λSdpedz2Sλ12a2λSpe=0$$\begin{array}{} \displaystyle \frac{{{d^2}{p_e}}}{{d{z^2}}} + \frac{2}{S}\frac{{dS}}{{dz}}{a_1}\left( {\lambda S} \right)\frac{{d{p_e}}}{{dz}} - 2S\lambda _1^2{a_2}\left( {\lambda S} \right){p_e} = 0 \end{array}$$

where λ12=ω2L2ρ0a0c02ρhandc02=BBρρ$\begin{array}{} \displaystyle \lambda _1^2 = \frac{{{\omega ^2}{L^2}{\rho _0}{a_0}}}{{c_0^2\rho h}} ~\text{and}~ c_0^2 = {B \mathord{\left/ {\vphantom {B \rho }} \right. } \rho } \end{array}$

The above relation can also be written as

λa2=2λ12=ω2L2c12$$\begin{array}{} \displaystyle \lambda _a^2 = 2\lambda _1^2 = \frac{{{\omega ^2}{L^2}}}{{c_1^2}} \end{array}$$

here, c12=Bh2ρ0a0$\begin{array}{} \displaystyle {c_1^2} = \frac{{Bh}}{{2{\rho _0}{a_0}}} \end{array}$ is the classical Mones-Korteweg speed.

For thin walled elastic tubes, different pressure radius relationship are given as (Taylor and Gerrard (1977))

(i)p=ρ0c1211S$$\begin{array}{} \displaystyle (i) \,\,\,\,\, p = {\rho _0}c_1^2\left( {1 - \frac{1}{S}} \right) \end{array}$$

(ii)p=83ρ0c1211S2$$\begin{array}{} \displaystyle (ii) \,\,\,\, p = \frac{8}{3}{\rho _0}c_1^2\left( {1 - \frac{1}{{{S^2}}}} \right) \end{array}$$

(iii)p=2ρ0c121S2S1$$\begin{array}{} \displaystyle (iii) \,\,\,\, p = 2{\rho _0}c_1^2\frac{1}{{{S^2}}}\left( {S - 1} \right) \end{array}$$

(iv)p=83ρ0c121S2S1$$\begin{array}{} \displaystyle (iv) \,\,\,\, p = \frac{8}{3}{\rho _0}c_1^2\frac{1}{{{S^2}}}\left( {S - 1} \right) \end{array}$$

(v)p=2ρ03c1211S4$$\begin{array}{} \displaystyle (v) \,\,\,\,\, p = \frac{{2{\rho _0}}}{3}c_1^2\left( {1 - \frac{1}{{{S^4}}}} \right) \end{array}$$

From the above pressure radius relationships, c2(z) at any axial station is represented by c12c12F(S)F(S),$\begin{array}{} \displaystyle {{c_1^2} \mathord{\left/ {\vphantom {{c_1^2} {F(S)}}} \right. } {F(S)}} , \end{array}$ where F(S) is a function of s depends on the selected pressure radius relationship.

Now, Eq. (28) can be written in the general form as

d2pedz2+2SdSdza1dpedzλa2F(S)a2pe=0$$\begin{array}{} \displaystyle \frac{{{d^2}{p_e}}}{{d{z^2}}} + \frac{2}{S}\frac{{dS}}{{dz}}{a_1}\frac{{d{p_e}}}{{dz}} - \lambda _a^2F(S){a_2}{p_e} = 0 \end{array}$$

The expression of excess pressure for large Womersley parameter is

d2pedz2+2SdSdz1iiλβ1+βSλβ1+βSdpedz+λa2F(S)12iiλβ1+βSλβ1+βSpe=0$$\begin{array}{} \displaystyle \frac{{{d^2}{p_e}}}{{d{z^2}}} + \frac{2}{S}\frac{{dS}}{{dz}}\left( {1 - {i \mathord{\left/ {\vphantom {i {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S}}} \right. } {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S}}} \right)\frac{{d{p_e}}}{{dz}} + \lambda _a^2F(S)\left( {1 - 2{i \mathord{\left/ {\vphantom {i {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S}}} \right. } {\lambda \sqrt {\frac{\beta }{{1 + \beta }}} S}}} \right){p_e} = 0 \end{array}$$

The excess pressure given by Eq. (36) is complex, by taking pe = pr + ipi and comparing real and imaginary parts, above expression reduced to two coupled second order ordinary differential equations of pr and pi . Now, the obtained two equations are reduced to four first order equations and solved using Mathematica by defining the initial conditions at some point of the z. With these values of pr and pi, we calculate

pe=pr2+pi21122,dpedz=dprdz2+dpidz21122$$\begin{array}{} \displaystyle \left| {{p_e}} \right| = {\left( {p_r^2 + p_i^2} \right)^{{1 \mathord{\left/ {\vphantom {1 2}} \right. } 2}}},\;\;\;\;\;\;\;\left| {\frac{{d{p_e}}}{{dz}}} \right| = {\left( {{{\left( {\frac{{d{p_r}}}{{dz}}} \right)}^2} + {{\left( {\frac{{d{p_i}}}{{dz}}} \right)}^2}} \right)^{{1 \mathord{\left/ {\vphantom {1 2}} \right. } 2}}} \end{array}$$

The modulus of excess pressure and pressure gradient for some geometries such as (i) straight tube defined by S = 1, 0 < z < 10 (ii) tapered tube defined by S(z) = exp (–0.025z), 0 < z < 10 (iii) locally constricted tube defined by S(z) = 1 – 0.5exp (– (z – 6)2), 0 < z < 10 with prescribed initial conditions

pr=0.1,pi=0,dpedz=0,dpidz=0.01atz=0$$\begin{array}{} \displaystyle {p_r} = 0.1,\;\;\;\;\;{p_i} = 0,\;\;\;\;\;\;\frac{{d{p_e}}}{{dz}} = 0,\;\;\;\;\;\frac{{d{p_i}}}{{dz}} = - 0.01\;\;\;\;{\rm{at}}\;\;\;\;z = 0 \end{array}$$

Results and Discussion

In this paper, the oscillatory motion of a Casson fluid through elastic tube with varying cross section is studied. The effects of various pertinent parameters on pressure variation are analyzed through graphs. Figs. 13 represent the pressure distribution along the axial direction of the tube for pressure radius relation defined by Eq. (30). Fig. 1 illustrates the variation in pressure distribution along straight tube for various values of λa. It is found that the pressure oscillates more as λa takes higher values. The influence of elastic parameter on pressure distribution for tapered tube is shown in Fig. 2. It shows that the amplitude of pressure increases for increasing values of λa. Fig. 3 shows the distribution of excess pressure along the locally constricted tube for various values of elastic parameter. We observe that pressure oscillates more for increasing values of λa with the maximum point of constriction at z = 6.

Fig. 1

Pressure along Straight Tube for Different Values of λa with α = 20, β = 0.5

Fig. 2

Pressure along Tapered Tube for Various Values of λa with α = 20, β = 0.5

Fig. 3

Pressure along Locally Constricted Tube for Various Values of λa with α = 20, β = 0.5

The variation in Pressure gradient for different values of elastic parameter for the pressure radius relation defined by Eq. 3.30 is presented in Figs. 46. The influence of elastic parameter on pressure gradient for straight tube is depicted in Fig. 4. It is noticed that the amplitude of the pressure gradient increases along the axial direction with increasing values of λa. The variation in pressure gradient along axial direction for tapered tube is shown in Fig. 5. It is clear that as growing values of λa, the pressure gradient oscillates more. The distribution of dpedz$\begin{array}{} \displaystyle \left| {\frac{{d{p_e}}}{{dz}}} \right| \end{array}$ for different values of λa for locally constricted tube is depicted in Fig. 6. It is observed that the amplitude of pressure gradient increases as elastic parameter increases.

Fig. 4

Pressure Gradient along Straight Tube for Various Values of λa with α = 20, β = 0.5

Fig. 5

Pressure Gradient along Tapered Tube for Various Values of λa with α = 20, β = 0.5

Fig. 6

Pressure Gradient along Locally Constricted Tube for Various Values of λa with α = 20, β = 0.5

Figs. 7 and 8 shows the variation in excess pressure for the different pressure radius relations defined in Eqs.(31)(34). Fig. 7 denotes the pressure distribution along tapered tube for different pressure radius relations for fixed value of λa. It is noticed that there is no much change in amplitude of the pressure for all the mentioned relations. The similar behavior is observed for the case of constricted tube which is given in Fig. 8.

Fig. 7

Pressure along Tapered Tube for Various Pressure-Radius Relations with λa = 0.75, α = 15, β = 0.5

Fig. 8

Pressure along Constricted Tube for Various Pressure-Radius Relations with λa = 0.75, α = 15, β = 0.5

The distribution of pressure along the tapered and constricted tubes for various values of Womersley parameter α for the pressure radius relationship defined by Eq. (30) is presented in Figs. 9 and 10, respectively. From Fig. 9, it is clear that for fixed values of elastic parameter, the amplitude of pressure increases as α increases. The increasing values of Womersley number decreases the amplitude of the pressure for the constricted tube which is shown in Fig. 10.

Fig. 9

Pressure along Tapered Tube for Various Values of α with λa = 0.5, β = 0.5

Fig. 10

Pressure along Constricted Tube for Various Values of α with λa = 0.5, β = 0.5

The influence of Casson parameter on pressure distribution for tapered and constricted tube for pressure radius relation defined by Eq. (30) is represented in Figs.11 and 12 respectively. It is noticed that for fixed values of α and λa, the amplitude of pressure enhances with increasing Casson parameter for tapered tube whereas the opposite behavior is found for constricted tube.

Fig. 11

Pressure along Tapered Tube for Various Values of β with λa = 0.5, α = 15

Fig. 12

Pressure along Constricted Tube for Various Values of β with λa = 0.5, α = 15

Conclusions

In the present chapter, the oscillatory flow of a Casson fluid in an elastic tube with variable cross section is investigated. The analytic expressions for axial velocity and mass flux are derived. The effects of different pertinent parameters on pressure distribution along the tube for different geometries are analyzed graphically. When elastic parameter increases, the amplitude of pressure increasing and pressure oscillates more in all cases such as straight tube, tapered tube and locally constricted tube. For various geometries the pressure gradient increases as elastic parameter increases. There is no much change in the amplitude of pressure for all defined pressure-radius relations. The variation in distribution of pressure for tapered tube increases as Womersley parameter takes higher values. Increasing Casson parameter enhances the amplitude of pressure for tapered tube where the opposite behavior is noticed for constricted tube.

eISSN:
2444-8656
Language:
English
Publication timeframe:
2 times per year
Journal Subjects:
Life Sciences, other, Mathematics, Applied Mathematics, General Mathematics, Physics