Mixed convection flow near an axisymmetric stagnation point on a vertical cylinder

Abstract

The mixed convection flow near an axisymmetric stagnation point on a vertical cylinder is considered. The equations for the fluid flow and temperature fields reduce to similarity form that involves a Reynolds number R and a mixed convection parameter λ, as well as the Prandtl number σ. Numerical solutions are obtained for representative values of these parameters, which show the existence of a critical value λ c  = λ c (Rσ) for the existence of solutions in the opposing (λ < 0) case. The variation of λ c with R is considered. In the aiding (λ > 0) case solutions are possible for all λ and the asymptotic limit λ → ∞ is obtained. The limits of large and small R are also treated and the nature of the solution in the asymptotic limit of large Prandtl number is briefly discussed.

Authors

C. Revnic
Tiberiu Popoviciu  Institute of Numerical Analysis Cluj, Romanian Academy

T. Grosan
Applied Mathematics, Babes-Bolyai University Cluj, Romania

J. Merkin
Department of Applied Mathematics, University of Leeds, Leeds, LS2 9JT, UK

I. Pop
Applied Mathematics, Babes-Bolyai University, Cluj, Romania

Keywords

Asymptotic solutions; Axisymmetric stagnation flow; Boundary layers; Dual solutions; Mixed convection

References

Paper coordinates

C. Revnic, T. Grosan, J. Merkin, I. Pop, Mixed convection flow near an axisymmetric stagnation point on a vertical cylinder, Journal of Engineering Mathematics, 64 (2009), pp. 1–13,
doi: 10.1007/s10665-008-9248-9

PDF

About this paper

Print ISSN

0022-0833

Online ISSN

1573-2703

Google Scholar Profile

soon

[1] Hiemenz K (1911) Die Grenzschicht an einem in den gleich formigen Flussigkeitsstrom eingetacuhten geraden Kreisszylinder. Dinglers Polytech J 326: 321–324
Google Scholar

[2] Eckert ERG (1942) Die Berechnung des Wärmeüberganges in der laminaren Grenzschicht um stromter Korper. VDI – Forchungsheft 416: 1–24
MathSciNet Google Scholar

[3] Gorla RSR (1976) Heat transfer in an axisymmetric stagnation flow on a cylinder. Appl Sci Res 32: 541–553
Article Google Scholar

[4] Hommann F (1936) Der Einfluss grosser Zähigkeit bei der Stromung um den Cylinder und um die Kugel. J Appl Math Phys (ZAMP) 16: 153–164
Google Scholar

[5] Smith FT (1974) Three dimensional stagnation point flow in a corner. Proc R Soc Lond A 344: 489–507
ADS Google Scholar

[6] Wang CY (1974) Axisymmetric stagnation flow on a cylinder. Quart Appl Math 32: 207–213
MATH Google Scholar

[7] Gorla RSR (1978) Nonsimilar axisymetric stagnation flow on a moving cylinder. Int J Eng Sci 16: 392–400
Article Google Scholar

[8] Weidman PD, Putkaradze V (2003) Axisymmetric stagnation flow obliquely impinging on a circular cylinder. Eur J Mech B/Fluids 22: 123–131
Article Google Scholar

[9] Weidman PD, Mahalingam S (1997) Axisymmetric stagnation-point flow impinging on a transversely oscillating plate with suction. J Eng Math 31: 305–318
MATH Article MathSciNet Google Scholar

[10] Gorla RSR (1979) Unsteady viscous flow in the vicinity of an axisymmetric stagnation point on a circular cylinder. Int J Eng Sci 17: 87–93
Article Google Scholar

[11] Ramachandran N, Chen TS, Armaly BF (1988) Mixed convection in stagnation flows adjacent to vertical surface. J Heat Trans 110: 173–177
Google Scholar

[12] Gorla RSR (1993) Mixed convection in an axisymmetric stagnation flow on a vertical cylinder. Acta Mech 99: 113–123
MATH Article Google Scholar

[13] Kuiken HK (1974) The thick free-convective boundary-layer along a semi-infinite isothermal vertical cylinder. J Appl Math Phys (ZAMP) 25: 497–514
MATH Article Google Scholar

[14] Naraian IP, Uberoi MS (1972) Combined forced and free convection heat transfer from thin needles in a uniform stream. Phys Fluids 15: 1879–1882
Article ADS Google Scholar

[15] Naraian IP, Uberoi MS (1973) Combined forced and free convection over thin needles. Int J Heat Mass Trans 16: 1505–1511
Article Google Scholar

[16] Chen ILS (1987) Mixed convection flow about slender bodies of revolution. J Heat Trans 109: 1033–1036
Article Google Scholar

[17] Wang CY (1990) Mixed convection on a vertical needle with heated tip. Phys Fluids A 2: 622–625
Article ADS Google Scholar

[18] Merkin JH (1985) On dual solutions occurring in mixed convection in a porous medium. J Eng Math 20: 171–179
Article MathSciNet Google Scholar

[19] Merkin JH, Mahmood T (1989) Mixed convection boundary layer similarity solutions: prescribed wall heat flux. J Appl Math Phys (ZAMP) 40: 51–68
MATH Article MathSciNet Google Scholar

[20] Stewartson K, Jones LT (1957) The heated vertical plate at high Prandtl number. J Aeronaut Sci 24: 379–380
Google Scholar

[21] Kuiken HK (1968) The heated vertical plate at high Prandtl number free convection. J Eng Math 2: 355–371
MATH Article Google Scholar

[22] Slater LJ (1960) Confluent hypergeometric functions. Cambridge University Press, Cambridge
Google Scholar

[23] Wilks G, Bramley JS (1981) Dual solutions in mixed convection. Proc R Soc Edinb 87A: 349–358
MathSciNet Google Scholar

Paper (preprint) in HTML form

Mixed convection flow near an axisymmetric stagnation point on a vertical cylinder

Cornelia Revnic • Teodor Grosan •
John Merkin • Ioan Pop
Abstract

The mixed convection flow near an axisymmetric stagnation point on a vertical cylinder is considered. The equations for the fluid flow and temperature fields reduce to similarity form that involves a Reynolds number RR and a mixed convection parameter λ\lambda, as well as the Prandtl number σ\sigma. Numerical solutions are obtained for representative values of these parameters, which show the existence of a critical value λc=λc(R,σ)\lambda_{c}=\lambda_{c}(R,\sigma) for the existence of solutions in the opposing ( λ<0\lambda<0 ) case. The variation of λc\lambda_{c} with RR is considered. In the aiding ( λ>0\lambda>0 ) case solutions are possible for all λ\lambda and the asymptotic limit λ\lambda\rightarrow\infty is obtained. The limits of large and small RR are also treated and the nature of the solution in the asymptotic limit of large Prandtl number is briefly discussed.

Keywords Asymptotic solutions • Axisymmetric stagnation flow • Boundary layers • Dual solutions • Mixed convection

1 Introduction

Combined forced- and free-convection flows (mixed convection) are encountered in many technological and industrial applications including solar receivers exposed to wind currents, electronic devices cooled by fans, nuclear reactors cooled during emergency shutdown, heat exchanges placed in a low-velocity environment and many more. Two-dimensional stagnation-point flows arise in the vicinity of a stagnation line resulting from a two-dimensional flow impinging on a curved surface at right angles to it and thereafter flowing symmetrically about the stagnation line. Hiemenz [1] was the first to study two-dimensional stagnation-point flows. Later Eckert [2] and Gorla [3] considered the corresponding forced-convection heat-transfer problem. Three-dimensional stagnation-point flows have been studied by Homann [4] and Smith [5] and the axisymmetric stagnation-point flow on a circular cylinder by Wang [6] and Gorla [7]. The three-dimensional flow resulting from an axisymmetric stagnation flow

00footnotetext: C. Revnic Tiberiu Popoviciu Mathematical Institute, P.O. Box 68-1, 400110 Cluj, Romania
T. Grosan • I. Pop Applied Mathematics, Babes-Bolyai University, CP 253, 3400 Cluj, Romania
J. Merkin ( \triangle) Department of Applied Mathematics, University of Leeds, Leeds LS2 9JT, UK
e-mail: amtjhm@maths.leeds.ac.uk

impinging obliquely on a body surface has been treated by Weidman and Putkaradze [8]. The problem of axisymmetric stagnation-point flow acting on a porous flat plate oscillating transversely in its own plane has been investigated by Weidman and Mahalingam [9]. In this case a three-dimensional flow results from a stagnation-point flow on a flat plate oscillating in its own plane. Gorla [10] has studied the unsteady viscous flow in the vicinity of an axisymmetric stagnation point on a circular cylinder.

The steady mixed convection flow near the stagnation region of a vertical flat plate has been studied by Ramachandran etal. [11] and by Gorla [12] for the flow near an axisymmetric stagnation point on a slender impermeable vertical cylinder. Mixed convection flows arise when the buoyancy forces resulting from temperature differences within the flow become comparable to the pressure gradient forces arising from the forced flow. As a consequence, both the flow and thermal fields are significantly affected by the buoyancy forces. The study of thick axisymmetric free-convection boundary layers along slender bodies has been shown by Kuiken [13] to have an unusual structure at large distances along the cylinder. When the boundary-layer variables are scaled so as to be of order unity within the boundary layer, the boundary conditions that hold on the surface of the slender body are given at a value of the independent variable which is close to zero. As a result, when a perturbation analysis is used to obtain the solution at large distances along the cylinder, the body is reduced to a line at the first approximation.

Several papers have been published previously on axisymmetric mixed convection boundary-layer flows along slender bodies. Naraian and Uberoi [14, 15] and Chen [16] have studied the mixed convection boundary layer on a vertical needle. These are bodies of revolution whose diameter is of the same order as the thickness of the velocity or thermal boundary layers that develop on it. By appropriately varying the radius of the needle, the boundary-layer equations admit similarity solutions. Wang [17] found a similarity solution for the mixed convection boundary layer on an adiabatic vertical needle with a heat source at the tip, a situation that arises, for example, for a stick burning at its lower end.

The present paper considers the steady mixed convection flow that develops near an axisymmetric stagnation point on a vertical isothermal cylinder in the case when the boundary layer is thick compared to the radius of the cylinder. We start by describing the governing equations, following closely [9,12] for the forced-convection problem. This results in two ordinary differential equations for the flow and temperature fields that involve, as well as the Prandtl number σ\sigma, the two further parameters RR, which is measure of the forced flow, and a mixed convection parameter λ\lambda. We consider both aiding flows (when the outer flow and the buoyancy forces are in the same direction, λ>0\lambda>0 ) and opposing flows (when the outer flow and the buoyancy forces act in opposite directions, λ<0\lambda<0 ). We start by giving numerical solutions to these equations for representative values of λ\lambda and RR, finding dual solutions for negative λ\lambda with critical points λc<0\lambda_{c}<0, requiring λλc\lambda\geq\lambda_{c} for the existence of a solution. We determine how λc\lambda_{c} varies with RR, before considering the asymptotic limits of λ\lambda\rightarrow\infty (free-convection limit) and RR\rightarrow\infty and R0R\rightarrow 0.

2 Equations

We consider the steady mixed convection flow near an axisymmetric stagnation point on an infinite cylinder. The cylinder is taken as mounted vertically and the flow is assumed to be axisymmetric about the xx-axis, which measures distance along the cylinder in a vertical direction with gravity acting in the negative xx-direction. The stagnation point is at x=0,r=ax=0,r=a, where rr measures distance radially from the centre of the cylinder of radius aa. The ambient fluid has a constant temperature TT_{\infty} and the cylinder is maintained at a temperature Tw(x)=T0(x/a)+TT_{w}(x)=T_{0}(x/a)+T_{\infty}. Having Tw>TT_{w}>T_{\infty} corresponds to assisting flow, with Tw<TT_{w}<T_{\infty} corresponding to opposing flow. The outer flow in this situation, taken directly from [12], is

u=(Ua)(ra2r),v=2U(xa)u=-\left(\frac{U_{\infty}}{a}\right)\left(r-\frac{a^{2}}{r}\right),\quad v=2U_{\infty}\left(\frac{x}{a}\right) (1)

where UU_{\infty} is the planar flow at large distances from the cylinder and where uu and vv are the velocity components in the xx and rr directions, respectively.

Refer to caption
Figure 1: Fig. 1 Plots of (a) f′′(1)f^{\prime\prime}(1) and (b) θ(1)\theta^{\prime}(1) against λ\lambda for R=1,5R=1,5, 10 obtained from the numerical solution of equations (3, 4) subject to boundary conditions (5) for σ=1\sigma=1

Again following [12], we introduce the variables

η=(ra)2,u=Uη1/2f(η),v=2U(xa)f(η),θ(η)=TTTwT,\eta=\left(\frac{r}{a}\right)^{2},\quad u=-U_{\infty}\eta^{-1/2}f(\eta),\quad v=2U_{\infty}\left(\frac{x}{a}\right)f^{\prime}(\eta),\quad\theta(\eta)=\frac{T-T_{\infty}}{T_{w}-T_{\infty}}, (2)

where TT is the temperature of the fluid. Applying (2) in the governing equations and making the standard Boussinesq approximation, we find that our flow is described by the similarity equations, again from [12],

ηf′′′+f′′+R(ff′′+1f2)+λθ=0,\displaystyle\eta f^{\prime\prime\prime}+f^{\prime\prime}+R\left(ff^{\prime\prime}+1-f^{\prime 2}\right)+\lambda\theta=0, (3)
ηθ′′+θ+σR(fθfθ)=0,\displaystyle\eta\theta^{\prime\prime}+\theta^{\prime}+\sigma R\left(f\theta^{\prime}-f^{\prime}\theta\right)=0, (4)

subject to the boundary conditions, from (1), that

f(1)=0,f(1)=0,θ(1)=1,f1,θ0 as η,f(1)=0,f^{\prime}(1)=0,\theta(1)=1,\quad f^{\prime}\rightarrow 1,\theta\rightarrow 0\quad\text{ as }\eta\rightarrow\infty, (5)

(primes denote differentiation with respect to η\eta ) where σ\sigma is the Prandtl number and where

R=Ua2v,λ=gβa2T08UvR=\frac{U_{\infty}a}{2v},\quad\lambda=\frac{g\beta a^{2}T_{0}}{8U_{\infty}v} (6)

are, respectively, a Reynolds number and a mixed convection parameter, with λ>0\lambda>0 corresponding to assisting flow and λ<0\lambda<0 corresponding to opposing flow. In (6) ν\nu is the kinematic viscosity of the fluid, gg the acceleration due to gravity and β\beta the coefficient of thermal expansion.

The parameters perhaps of most physical interest are the skin friction parameter CfC_{f} and the Nusselt number Nu, defined as

Cf=τwρU2,Nu=aqwk(TwT) where τw=μ(vr)r=a,qw=k(Tr)r=aC_{f}=\frac{\tau_{w}}{\rho U_{\infty}^{2}},\quad\mathrm{Nu}=\frac{aq_{w}}{k\left(T_{w}-T_{\infty}\right)}\quad\text{ where }\tau_{w}=\mu\left(\frac{\partial v}{\partial r}\right)_{r=a},\quad q_{w}=-k\left(\frac{\partial T}{\partial r}\right)_{r=a}\text{, } (7)

and where μ\mu and kk are the dynamic viscosity and thermal conductivity, respectively. From (2), we have that

Cf=2(xa)f′′(1),Nu=2θ(1).C_{f}=2\left(\frac{x}{a}\right)f^{\prime\prime}(1),\quad\mathrm{Nu}=-2\theta^{\prime}(1). (8)

The problem given by (3-5) has been considered previously by Gorla [12] who presented a range of numerical results for both aiding and opposing flows. The results given by [12] are only for a specific value of RR, namely R=100R=100, with critical points giving a finite range of existence for opposing flows not being identified in [12]. We start by first describing our numerical solutions to Eqs. (3) and (4) subject to boundary conditions (5) for a range of values of λ\lambda and RR, thus adding to the numerical treatment given by Gorla [12]. Throughout we assume that the Prandtl number σ\sigma is of O(1)O(1), our numerical results are all for the case when σ=1\sigma=1.

Refer to caption
Figure 2: Fig. 2 Plots of (a) f′′(1)f^{\prime\prime}(1) and (b) θ(1)\theta^{\prime}(1) against RR for λ=1,2\lambda=1,-2 obtained from the numerical solution of equations (3,4) subject to boundary conditions (5) for σ=1\sigma=1
Figure 3: Fig. 3 A plot of the critical value λc\lambda_{c} of λ\lambda against RR (for σ=1)\sigma=1). The region in the (λ,R)(\lambda,R) parameter plane where solutions exist is labelled on the figure
Refer to caption

3 Numerical results

Equations (3) and (4) subject to boundary conditions (5) were solved numerically using a standard shooting method for solving boundary-value problems (D02AGF in the NAG library). In Fig. 1 we plot f′′(1)f^{\prime\prime}(1) and θ(1)\theta^{\prime}(1) against λ\lambda for R=1,5,10R=1,5,10 (with σ=1\sigma=1 ). This figure shows that, for λ<0\lambda<0, there is a critical value λc\lambda_{c} with solutions possible only for 0>λλc0>\lambda\geq\lambda_{c} and for λ>λc\lambda>\lambda_{c} there are dual solutions. The value of λc\lambda_{c} decreases as RR is increased, thus giving a greater range of negative λ\lambda for possible solutions. For λ>0\lambda>0 there is only one solution with the values of f′′(1)f^{\prime\prime}(1) increasing and θ(1)\theta^{\prime}(1) decreasing as λ\lambda is increased (for a given value of RR ). For the larger values of λ\lambda, the values of f′′(1)f^{\prime\prime}(1) for R=1R=1 becomes greater than those for the larger values of R(=5,10)R(=5,10), indicating that, for sufficiently large values of λ,f′′(1)\lambda,f^{\prime\prime}(1) increases as RR is decreased. However, the values of θ(1)-\theta^{\prime}(1) increase as RR is increased (for a given value of λ\lambda ).

In Fig. 2 we take values for λ(λ=1,2)\lambda(\lambda=1,-2) representative respectively of aiding and opposing mixed convection and plot the corresponding values of f′′(1)f^{\prime\prime}(1) and θ(1)\theta^{\prime}(1) against RR. We see that in both cases θ(1)\theta^{\prime}(1) decreases as RR is increased. However, for λ=1,f′′(1)\lambda=1,f^{\prime\prime}(1) has a minimum value, of f′′(1)=1.6648f^{\prime\prime}(1)=1.6648 at R=0.272R=0.272, before increasing again for the larger values of RR. For λ=2\lambda=-2 there is a critical value RcR_{c} of R(Rc0.622)R\left(R_{c}\simeq 0.622\right) below which there are no solutions, as might be expected from Fig. 1.

3.1 Critical points

We saw in Fig. 1 the existence of a critical value λc\lambda_{c} of λ\lambda, requring λλc\lambda\geq\lambda_{c} for a solution to exist. We can calculate how λc\lambda_{c} varies with RR using the appproach described in [18,19]. Essentially we perturb about the solution given by (3,4)(3,4) to obtain a linear homogeneous problem. It is then the existence of a nontrivial solution to this homogeneous problem that determines λc\lambda_{c} for a given value of RR (and σ\sigma ). In Fig. 3 we plot λc\lambda_{c} against RR, with the figure showing
that λc\lambda_{c} decreases as RR is increased, seemingly in a linear manner for the larger values of RR, and increases towards zero as RR is reduced. The region in the ( λ,R\lambda,R ) parameter space where solutions exist is labelled on this figure.

Our numerical solutions suggest that considering various limiting forms would give some further insights into the nature of the solution. We now discuss these in more detail, starting with the free-convection limit, λ\lambda\rightarrow\infty.

4 Asymptotic results

4.1λ4.1\lambda large

To obtain a solution valid for λ\lambda large we start by putting
f=λ1/4F,ζ=λ1/4(η1)f=\lambda^{1/4}F,\quad\zeta=\lambda^{1/4}(\eta-1),
leaving θ\theta unscaled. This results in the equations
(1+λ1/4ζ)F′′′+λ1/4F′′+R(FF′′F2)+λ1R+θ=0\left(1+\lambda^{-1/4}\zeta\right)F^{\prime\prime\prime}+\lambda^{-1/4}F^{\prime\prime}+R\left(FF^{\prime\prime}-F^{2}\right)+\lambda^{-1}R+\theta=0,
(1+λ1/4ζ)θ′′+λ1/4θ+σR(FθFθ)=0\left(1+\lambda^{-1/4}\zeta\right)\theta^{\prime\prime}+\lambda^{-1/4}\theta^{\prime}+\sigma R\left(F\theta^{\prime}-F^{\prime}\theta\right)=0,
where primes now denote differentiation with respect to ζ\zeta, subject to the boundary conditions
F(0)=0,F(0)=0,θ(0)=1,Fλ1/2,θ0F(0)=0,F^{\prime}(0)=0,\theta(0)=1,\quad F^{\prime}\rightarrow\lambda^{-1/2},\theta\rightarrow 0 as ζ\zeta\rightarrow\infty.

Equations (10,11)(10,11) suggest looking for a solution by expanding

F=F0+λ1/4F1+,θ=θ0+λ1/4θ1+F=F_{0}+\lambda^{-1/4}F_{1}+\cdots,\quad\theta=\theta_{0}+\lambda^{-1/4}\theta_{1}+\cdots (13)

We can scale the leading-order problem obtained by substituting expansion (13) in (10, 11) by putting
F0=R3/4F¯0,ζ¯=R1/4ζF_{0}=R^{-3/4}\bar{F}_{0},\quad\bar{\zeta}=R^{1/4}\zeta.

This results in
F¯0′′′+F¯0F¯0′′F¯02+θ0=0\bar{F}_{0}^{\prime\prime\prime}+\bar{F}_{0}\bar{F}_{0}^{\prime\prime}-\bar{F}_{0}^{\prime 2}+\theta_{0}=0,
θ0′′+σ(F¯0θ0F¯0θ0)=0\theta_{0}^{\prime\prime}+\sigma\left(\bar{F}_{0}\theta_{0}^{\prime}-\bar{F}_{0}^{\prime}\theta_{0}\right)=0,
subject to
F¯0(0)=0,F¯0(0)=0,θ0(0)=1,F¯00,θ00\bar{F}_{0}(0)=0,\bar{F}_{0}^{\prime}(0)=0,\theta_{0}(0)=1,\quad\bar{F}_{0}^{\prime}\rightarrow 0,\theta_{0}\rightarrow 0 as ζ¯\bar{\zeta}\rightarrow\infty,
where primes now denote differentiation with respect to ζ¯\bar{\zeta}. A numerical solution of (15-17) gives, for σ=1\sigma=1, F¯0′′(0)=0.73950,θ0(0)=0.59509\bar{F}_{0}^{\prime\prime}(0)=0.73950,\theta_{0}^{\prime}(0)=-0.59509.

We can continue to the next order by first rescaling
F1=R1F¯1,θ1=R1/4θ¯1F_{1}=R^{-1}\bar{F}_{1},\quad\theta_{1}=R^{-1/4}\bar{\theta}_{1}.

This gives, with (14),
F¯1′′′+F¯0F¯1′′+F¯1F¯0′′2F¯0F¯1+θ¯1=(F¯0′′+ζ¯F¯0′′′)\bar{F}_{1}^{\prime\prime\prime}+\bar{F}_{0}\bar{F}_{1}^{\prime\prime}+\bar{F}_{1}\bar{F}_{0}^{\prime\prime}-2\bar{F}_{0}^{\prime}\bar{F}_{1}^{\prime}+\bar{\theta}_{1}=-\left(\bar{F}_{0}^{\prime\prime}+\bar{\zeta}\bar{F}_{0}^{\prime\prime\prime}\right),
θ¯1′′+σ(F¯0θ¯1+F¯1θ0F¯0θ¯1F¯1θ0)=(θ0+ζ¯θ0′′)\bar{\theta}_{1}^{\prime\prime}+\sigma\left(\bar{F}_{0}\bar{\theta}_{1}^{\prime}+\bar{F}_{1}\theta_{0}^{\prime}-\bar{F}_{0}^{\prime}\bar{\theta}_{1}-\bar{F}_{1}^{\prime}\theta_{0}\right)=-\left(\theta_{0}^{\prime}+\bar{\zeta}\theta_{0}^{\prime\prime}\right),
subject to
F¯1(0)=0,F¯1(0)=0,θ¯1(0)=0,F¯10,θ¯10\bar{F}_{1}(0)=0,\bar{F}_{1}^{\prime}(0)=0,\bar{\theta}_{1}(0)=0,\quad\bar{F}_{1}^{\prime}\rightarrow 0,\bar{\theta}_{1}\rightarrow 0 as ζ¯\bar{\zeta}\rightarrow\infty.

A numerical solution gives, again for σ=1\sigma=1, that F¯1′′(0)=0.06888,θ¯1(0)=0.23452\bar{F}_{1}^{\prime\prime}(0)=0.06888,\bar{\theta}_{1}^{\prime}(0)=-0.23452.

Refer to caption
Figure 4: Fig. 4 Plots of (λ3/R)1/4f′′(1)\left(\lambda^{3}/R\right)^{-1/4}f^{\prime\prime}(1) and (λR)1/4θ(1)-(\lambda R)^{-1/4}\theta^{\prime}(1) obtained from the numerical solution of Eqs. (3, 4) against λ\lambda for R=1,10R=1,10 and σ=1\sigma=1

From ( 9,14 ) and (18) we have

(d2fdη2)η=1(λ3R)1/4(0.7395+0.0689(Rλ)1/4+),\displaystyle\left(\frac{\mathrm{d}^{2}f}{\mathrm{~d}\eta^{2}}\right)_{\eta=1}\sim\left(\frac{\lambda^{3}}{R}\right)^{1/4}\left(0.7395+0.0689(R\lambda)^{-1/4}+\cdots\right), (22)
(dθdη)η=1(Rλ)1/4(0.5951+0.2345(Rλ)1/4+) as λ.\displaystyle\left(\frac{\mathrm{d}\theta}{\mathrm{~d}\eta}\right)_{\eta=1}\sim-(R\lambda)^{1/4}\left(0.5951+0.2345(R\lambda)^{-1/4}+\cdots\right)\quad\text{ as }\lambda\rightarrow\infty.

Expressions (22) show that, for sufficiently large values of λ,θ(1)\lambda,-\theta^{\prime}(1) increases as RR is increased and that f′′(1)f^{\prime\prime}(1) decreases as RR is increased, in line with Fig. 1. In Fig. 4 we plot the values of (λ3/R)1/4f′′(1)\left(\lambda^{3}/R\right)^{-1/4}f^{\prime\prime}(1) and (λR)1/4θ(1)-(\lambda R)^{-1/4}\theta^{\prime}(1) obtained from the numerical solution of (3,4)(3,4) against λ\lambda for R=1,10R=1,10. In both cases the asymptotic limit for λ\lambda large given by (22) is approached as λ\lambda increases, though more slowly for R=10R=10 than for R=1R=1 as might be expected from (22).

4.2R4.2R large

To obtain a solution for RR large, we follow the approach given in [8] for the forced-convection case and start by writing

f=R1/2ϕ,ξ=R1/2(η1)f=R^{-1/2}\phi,\quad\xi=R^{1/2}(\eta-1) (23)

and leaving θ\theta unscaled. This results in the equations

(1+R1/2ξ)ϕ′′′+R1/2ϕ′′+1+ϕϕ′′ϕ2+λR1θ=0,\left(1+R^{-1/2}\xi\right)\phi^{\prime\prime\prime}+R^{-1/2}\phi^{\prime\prime}+1+\phi\phi^{\prime\prime}-\phi^{\prime 2}+\lambda R^{-1}\theta=0, (24)
(1+R1/2ξ)θ′′+R1/2θ+σ(ϕθϕθ)=0,\left(1+R^{-1/2}\xi\right)\theta^{\prime\prime}+R^{-1/2}\theta^{\prime}+\sigma\left(\phi\theta^{\prime}-\phi^{\prime}\theta\right)=0, (25)

subject to the boundary conditions given in ( 5 ) and where primes now denote differentiation with respect to ξ\xi. Equations (24,25)(24,25) suggest an expansion in powers of R1/2R^{-1/2}, the leading-order term just being the forced-convection limit discussed in [8].

For convection to have an effect at leading order we require λ\lambda to be large, specifically of O(R)O(R). This leads us to put

λ=μR with μ of O(1).\lambda=\mu R\quad\text{ with }\mu\text{ of }O(1). (26)

The problem for the leading-order terms ϕ0,θ0\phi_{0},\theta_{0} is now

ϕ0′′′+1+ϕ0ϕ0′′ϕ02+μθ0=0,\phi_{0}^{\prime\prime\prime}+1+\phi_{0}\phi_{0}^{\prime\prime}-\phi_{0}^{\prime 2}+\mu\theta_{0}=0, (27)
θ0′′+σ(ϕ0θ0ϕ0θ0)=0,\theta_{0}^{\prime\prime}+\sigma\left(\phi_{0}\theta_{0}^{\prime}-\phi_{0}^{\prime}\theta_{0}\right)=0, (28)

still subject to the boundary conditions that

Refer to caption
Figure 5: Fig. 5 Plots of (a) ϕ0′′(0)\phi_{0}^{\prime\prime}(0) and (b) θ0(0)\theta_{0}^{\prime}(0) against μ=λR1\mu=\lambda R^{-1} for the solution for large RR given by (27-29) for σ=1\sigma=1
Figure 6: Fig. 6 A plot of λcR1\lambda_{c}R^{-1} against RR for σ=1\sigma=1. The asymptotic limit (30) for RR large is shown by the broken line
Refer to caption
ϕ0(0)=0,ϕ0(0)=0,θ0(0)=1,ϕ01,θ00 as ξ.\phi_{0}(0)=0,\phi_{0}^{\prime}(0)=0,\theta_{0}(0)=1,\quad\phi_{0}^{\prime}\rightarrow 1,\theta_{0}\rightarrow 0\quad\text{ as }\xi\rightarrow\infty. (29)

Equations (27-29) have to solved numerically and graphs of ϕ0′′(0)\phi_{0}^{\prime\prime}(0) and θ0(0)\theta_{0}^{\prime}(0) for σ=1\sigma=1 plotted against μ\mu are shown in Fig. 5. From these figures we see that there is a critical value μc\mu_{c} of μ\mu with dual solutions for 0>μ>μc0>\mu>\mu_{c} and no solutions for μ<μc\mu<\mu_{c}. For σ=1\sigma=1, we find that μc=2.3618\mu_{c}=-2.3618, giving

λc2.3618R+ as R.\lambda_{c}\sim-2.3618R+\cdots\quad\text{ as }R\rightarrow\infty. (30)

For μ>0\mu>0 there is only one solution with both ϕ0′′(0)\phi_{0}^{\prime\prime}(0) and θ0(0)-\theta_{0}^{\prime}(0) increasing as μ\mu is increased.
Expression (30) shows a linear increase in |λc|\left|\lambda_{c}\right| with RR, as noted previously about Fig.3, and to confirm this asymptotic behaviour we plot the values of λcR1\lambda_{c}R^{-1} obtained from our numerical integrations to find λc\lambda_{c} against RR in Fig. 6. The numerically determined values approach this asymptotic limit (shown by the broken line), though only slowly as RR is increased, suggesting that the O(R1/2)O\left(R^{-1/2}\right) correction has a significant effect even at moderately large values of RR.

For μ\mu large we can recover expressions (22) by putting

ϕ0=μ1/4ϕ¯0,ξ¯=μ1/4ξ.\phi_{0}=\mu^{1/4}\bar{\phi}_{0},\quad\bar{\xi}=\mu^{1/4}\xi. (31)

When (31) is substituted in equations (27-29) and then μ\mu\rightarrow\infty, we obtain equations (15-17) and, on using (26), the leading-order terms in (22).

4.3R4.3R small

The behaviour of the solution for RR small depends on whether λ\lambda is small or of O(1)O(1). We start with the latter case, assuming that λ>0\lambda>0.

4.3.1 λ\lambda of O(1)O(1)

For this case we start in an inner region where η\eta is of O(1)O(1) and scale
f=A(R)g,θ=1+B(R)hf=A(R)g,\quad\theta=1+B(R)h.

The scaling factors AA and BB are to be determined, though we assume that
A(R)1,B(R)1,RAB1,AR1A(R)\gg 1,\quad B(R)\ll 1,\quad\frac{RA}{B}\ll 1,\quad AR\ll 1.

When (32) is substituted in (3,4)(3,4) and (33)(33) is taken, the leading-order terms g0,h0g_{0},h_{0} are given by, on satisfying the boundary condtions on η=1\eta=1,

g0=a0(ηlogηη+1),h0=b0logηg_{0}=a_{0}(\eta\log\eta-\eta+1),\quad h_{0}=b_{0}\log\eta (34)

for constants a0,b0a_{0},b_{0} to be determined. Before continuing the solution in the inner region, we next consider the outer region.

For the outer region we write
f=R1G,θ=β2RH,Y=βη,f=R^{-1}G,\quad\theta=\frac{\beta^{2}}{R}H,\quad Y=\beta\eta,\quad where β=β(R)1,Rβ1\beta=\beta(R)\ll 1,\quad\frac{R}{\beta}\ll 1
with the scaling factor β(R)\beta(R) also to be determined. When (35) is substituted in (3,4) and the assumptions for β\beta given in (35) applied, we find that the leading-order problem in the outer region is
YG′′′+G′′+GG′′G2+λH=0,YH′′+H+σ(GHGH)=0YG^{\prime\prime\prime}+G^{\prime\prime}+GG^{\prime\prime}-G^{2}+\lambda H=0,\quad YH^{\prime\prime}+H^{\prime}+\sigma\left(GH^{\prime}-G^{\prime}H\right)=0,
subject to the outer boundary conditions at leading order that
G0,H0G^{\prime}\rightarrow 0,\quad H\rightarrow 0\quad as YY\rightarrow\infty.

To find the inner boundary conditions for (36) we need to match with the inner region.
We can express the inner solution (34) as, on using (35),
fA(R)βa0Y[logY+(logβ)1]+f\sim\frac{A(R)}{\beta}a_{0}Y[\log Y+(-\log\beta)-1]+\cdots,
θ[1+b0B(R)(logβ)]+b0B(R)logY+\theta\sim\left[1+b_{0}B(R)(-\log\beta)\right]+b_{0}B(R)\log Y+\cdots.

From (38) we choose
A(R)=βR(logβ),B(R)=1(logβ)=β2R,b0=1A(R)=\frac{\beta}{R(-\log\beta)},\quad B(R)=\frac{1}{(-\log\beta)}=\frac{\beta^{2}}{R},\quad b_{0}=-1.

Hence β(R)\beta(R) is given implicitly by
β2(logβ)=R,\beta^{2}(-\log\beta)=R,\quad with β2R1/2(logR)1/2+\beta\sim\frac{\sqrt{2}R^{1/2}}{(-\log R)^{1/2}}+\cdots\quad as R0R\rightarrow 0.

We note that (39,40)(39,40) are consistent with the assumptions in (33)(33) and (35)(35). Expressions (38)(38) then give, at leading order,
Ga0Y+,HlogY+G\sim a_{0}Y+\cdots,\quad H\sim-\log Y+\cdots\quad as Y0Y\rightarrow 0.

To get higher order terms in (41) we need to consider the inner region again. On using (39) we see that an expansion of the form

g=g0+g1(logβ)++O(β),h=h0+h1(logβ)++O(β)g=g_{0}+\frac{g_{1}}{(-\log\beta)}+\cdots+O(\beta),\quad h=h_{0}+\frac{h_{1}}{(-\log\beta)}+\cdots+O(\beta) (42)

is required, where the O(β)O(\beta) terms in (42) also include terms in
β(logβ)2,β(logβ),β(logβ)1\beta(-\log\beta)^{2},\quad\beta(-\log\beta),\quad\beta(-\log\beta)^{-1}
and h1=b1logηh_{1}=b_{1}\log\eta for some constant b1b_{1}. When ( 42 ) is substituted in ( 3,4 ) and using ( 32,35 ) we find, after some calculation, that the inner boundary condition (41) for the outer region becomes modified to
Ga0Y+λ2Y2logY+12(a025λ2λb1)Y2+G\sim a_{0}Y+\frac{\lambda}{2}Y^{2}\log Y+\frac{1}{2}\left(a_{0}^{2}-\frac{5\lambda}{2}-\lambda b_{1}\right)Y^{2}+\cdots,
HlogY+b1σa0YlogY+σa0(b1+2)Y+H\sim-\log Y+b_{1}-\sigma a_{0}Y\log Y+\sigma a_{0}\left(b_{1}+2\right)Y+\cdots\quad as Y0Y\rightarrow 0.
We can remove the parameter λ\lambda from this problem by writing Y¯=λ1/2Y\bar{Y}=\lambda^{1/2}Y and then the problem in the outer region becomes
Y¯G′′′+G′′+GG′′G2+H=0,Y¯H′′+H+σ(GHGH)=0\bar{Y}G^{\prime\prime\prime}+G^{\prime\prime}+GG^{\prime\prime}-G^{2}+H=0,\quad\bar{Y}H^{\prime\prime}+H^{\prime}+\sigma\left(GH^{\prime}-G^{\prime}H\right)=0,
subject to the conditions that
G0,H0G^{\prime}\rightarrow 0,\quad H\rightarrow 0\quad as Y¯\bar{Y}\rightarrow\infty
and
Ga¯0Y¯+Y¯22logY¯+12(a¯0252b¯1)Y¯2+G\sim\bar{a}_{0}\bar{Y}+\frac{\bar{Y}^{2}}{2}\log\bar{Y}+\frac{1}{2}\left(\bar{a}_{0}^{2}-\frac{5}{2}-\bar{b}_{1}\right)\bar{Y}^{2}+\cdots,
HlogY¯+b¯1σa¯0Y¯logY¯+σa¯0(b¯1+2)Y¯+H\sim-\log\bar{Y}+\bar{b}_{1}-\sigma\bar{a}_{0}\bar{Y}\log\bar{Y}+\sigma\bar{a}_{0}\left(\bar{b}_{1}+2\right)\bar{Y}+\cdots\quad as Y¯0\bar{Y}\rightarrow 0,
where a¯0=λ1/2a0\bar{a}_{0}=\lambda^{-1/2}a_{0} and b¯1=b1+12logλ\bar{b}_{1}=b_{1}+\frac{1}{2}\log\lambda.
The problem given by (44-46) has to be solved numerically to determine the constants a¯0\bar{a}_{0} and b¯1\bar{b}_{1} and we find, for σ=1\sigma=1, that a¯0=1.0016,b¯1=1.2396\bar{a}_{0}=1.0016,\bar{b}_{1}=-1.2396. This then gives for small RR and σ=1\sigma=1,
(d2fdη2)η=1=1β(logβ)2(1.0016λ1/2+)\left(\frac{\mathrm{d}^{2}f}{\mathrm{~d}\eta^{2}}\right)_{\eta=1}=\frac{1}{\beta(-\log\beta)^{2}}\left(1.0016\lambda^{1/2}+\cdots\right),
(dθdη)η=1=1(logβ)(1+1.2396+12logλ(logβ)+)\left(\frac{\mathrm{d}\theta}{\mathrm{d}\eta}\right)_{\eta=1}=-\frac{1}{(-\log\beta)}\left(1+\frac{1.2396+\frac{1}{2}\log\lambda}{(-\log\beta)}+\cdots\right),
with β\beta given in terms of RR by (40). Equations (3,4) were solved numerically with λ=1\lambda=1 for small values of RR. The solution domain increases as RR is decreased, in line with (35), and the outer boundary condition had to be applied at increasingly larger values of η=η\eta=\eta_{\infty} as RR was decreased, the results shown in Fig. 7 were obtained using η=250\eta_{\infty}=250. In Fig. 7 we give plots of f′′(1)f^{\prime\prime}(1) and θ(1)\theta^{\prime}(1) against RR obtained from our numerical integrations, the values obtained from (47) are shown by broken lines. The agreement with the asymptotic forms for small RR is not particularly good, though both solutions are following the same trend. This difference can, perhaps, be explained by the fact that the approach to the asymptotic forms (47) is only very slow, with correction terms of (logβ)1(logR)1(-\log\beta)^{-1}\sim(-\log R)^{-1} and hence RR has to be extremely small for these correction terms to have only a small effect. In practice we probably require values of RR too small for obtaining reasonably accurate numerical solutions. At these very small values of R,ηR,\eta_{\infty} needs to be extremely large leading to errors in the shooting method employed to solve the two-point boundary-value problem.

4.3.2 λ\lambda small

We can see, particularly from ( 36,37 ), that, when λ\lambda is of O(1)O(1), the flow is driven at leading order only by the natural-convection effects. The forced flow enters the solution at higher order. However, when λ\lambda is small, this cannot be the case and to get an estimate on λ\lambda when the forced-convection effects have an influence at leading order, we see from (35) that the buoyancy term included in equation (36) is of O(λβ2R1)O\left(\lambda\beta^{2}R^{-1}\right) whereas the forced-convection term is of O(R)O(R). This suggests, on using (40), that these two effects will be comparable when λR(logR)\lambda\sim R(-\log R) and leads us to put
λ=R(logR)ν\lambda=R(-\log R)\nu with ν\nu of O(1)O(1) for RR small.

Refer to caption
Figure 7: Fig. 7 Plots of (a) f′′(1)f^{\prime\prime}(1) and (b) θ(1)\theta^{\prime}(1) against RR for λ=1,σ=1\lambda=1,\sigma=1 for small RR obtained from the numerical solution of Eqs. (3, 4) subject to boundary conditions (5). The asymptotic expressions (47) are shown by the broken lines

We start our solution for this case in the inner region where we have at leading order, motivated by our previous solution for λ\lambda of O(1)O(1),

f1(logR)c0(ηlogηη+1)+,θ1+1(logR)(logη+d1(logR)logη+)f\sim\frac{1}{(-\log R)}c_{0}(\eta\log\eta-\eta+1)+\cdots,\quad\theta\sim 1+\frac{1}{(-\log R)}\left(-\log\eta+\frac{d_{1}}{(-\log R)}\log\eta+\cdots\right) (49)

for constants c0c_{0} and d1d_{1} to be determined. For the outer region we now put

f=1RG,θ=1(logR)H,Y=Rη.f=\frac{1}{R}G,\quad\theta=\frac{1}{(-\log R)}H,\quad Y=R\eta. (50)

Applying (50) in (3,4) gives, at leading order,

YG′′′+G′′+1+GG′′G2+νH=0,YH′′+H+σ(GHGH)=0,YG^{\prime\prime\prime}+G^{\prime\prime}+1+GG^{\prime\prime}-G^{2}+\nu H=0,\quad YH^{\prime\prime}+H^{\prime}+\sigma\left(GH^{\prime}-G^{\prime}H\right)=0, (51)

now subject to the outer boundary conditions

G1,H0 as YG^{\prime}\rightarrow 1,\quad H\rightarrow 0\quad\text{ as }Y\rightarrow\infty (52)

and, on matching with the inner region,

Gc0Y+,HlogY+d1+ as Y0.G\sim c_{0}Y+\cdots,\quad H\sim-\log Y+d_{1}+\cdots\quad\text{ as }Y\rightarrow 0. (53)

We note that ν\nu cannot be scaled out of problem (51-53) (as it could previously) and that, for ν=0\nu=0 (forced-convection limit)

c0=1,G=Y,H=eσYlog(σY)(1+σY)+(1+σY)σYev(3+v)logv(1+v)3dvc_{0}=1,\quad G=Y,\quad H=-\frac{\mathrm{e}^{-\sigma Y}\log(\sigma Y)}{(1+\sigma Y)}+(1+\sigma Y)\int_{\sigma Y}^{\infty}\frac{\mathrm{e}^{-v}(3+v)\log v}{(1+v)^{3}}\mathrm{~d}v (54)

giving
d1=0ev(3+v)logv(1+v)3dvlogσ=1.4448logσd_{1}=\int_{0}^{\infty}\frac{\mathrm{e}^{-v}(3+v)\log v}{(1+v)^{3}}\mathrm{~d}v-\log\sigma=-1.4448-\log\sigma.
Plots of c0c_{0} against ν\nu obtained by solving (51-53) numerically are given in Fig. 8. This figure shows that there is a critical value vcv_{c} of vv, with vc=0.9832v_{c}=-0.9832 for σ=1\sigma=1, and dual solutions for 0>v>vc0>v>v_{c}. This gives

λc0.9832R(logR)+ as R0.\lambda_{c}\sim-0.9832R(-\log R)+\cdots\text{ as }R\rightarrow 0. (55)

For v>0v>0 there is only one solution and the problem given by (44-46) can be recovered for vv large by putting Y¯=v1/2Y\bar{Y}=v^{1/2}Y in (51-53) and letting vv\rightarrow\infty.

Refer to caption
Figure 8: Fig. 8 A plot of c0c_{0} against ν\nu, defined in (48), for σ=1\sigma=1 arising in the small RR, small λ\lambda problem (5153)
Refer to caption
Figure 9: Fig. 9 Plots of θ(1)\theta^{\prime}(1) against λ\lambda for R=1R=1 and σ=1\sigma=1, 10 obtained from the numerical solution of Eqs. (3,4)(3,4) subject to boundary conditions (5)

5 Solution for σ\sigma large

Our previous numerical results were for the case σ=1\sigma=1 and we expect qualitatively similar behaviour when σ\sigma is of O(1)O(1), typical of gases. For liquids (water) σ\sigma is somewhat larger and it is thus worth briefly considering the large σ\sigma limit. Following the treatment in [20,21] we expect that, for σ\sigma large, the solution to involve a relatively thin thermal inner layer and a thicker outer viscous flow region. In the inner region we have

η=1+σ1/3τ,f=σ2/3Φ\eta=1+\sigma^{-1/3}\tau,\quad f=\sigma^{-2/3}\Phi (56)

with θ\theta left unscaled. At leading order, for σ\sigma large, we obtain

Φ′′′=0,θ′′+R(ΦθΦθ)=0,\Phi^{\prime\prime\prime}=0,\quad\theta^{\prime\prime}+R\left(\Phi\theta^{\prime}-\Phi^{\prime}\theta\right)=0, (57)

subject to

Φ(0)=0,Φ(0)=0,θ(0)=1,θ0 as τ,\Phi(0)=0,\Phi^{\prime}(0)=0,\quad\theta(0)=1,\theta\rightarrow 0\text{ as }\tau\rightarrow\infty, (58)

with the outer condition on Φ\Phi relaxed at this stage and where primes denote differentiation with respect to τ\tau.
Equations (57,58)(57,58) give

Φ=A0τ2\Phi=A_{0}\tau^{2} (59)

for some constant A0=A0(R)A_{0}=A_{0}(R) to be determined and

θ=(23)!3(13)!esU(43;23;s) where s=13(RA0)τ3\theta=\frac{\left(\frac{2}{3}\right)!}{3\left(\frac{1}{3}\right)!}\mathrm{e}^{-s}U\left(\frac{4}{3};\frac{2}{3};s\right)\quad\text{ where }s=\frac{1}{3}\left(RA_{0}\right)\tau^{3} (60)

in terms of confluent hypergeometric functions [22]. Expressions (56) and (60) give

(dθdη)η=1=σ1/3(RA0)1/332/3(23)!22(13)!2+ as σ.\left(\frac{\mathrm{d}\theta}{\mathrm{~d}\eta}\right)_{\eta=1}=-\sigma^{1/3}\left(RA_{0}\right)^{1/3}\frac{3^{2/3}\left(\frac{2}{3}\right)!^{2}}{2\left(\frac{1}{3}\right)!^{2}}+\cdots\quad\text{ as }\sigma\rightarrow\infty. (61)

Expression (61) shows that the heat transfer increases as σ\sigma is increased, being of O(σ1/3)O\left(\sigma^{1/3}\right) for σ\sigma large.
To determine the constant A0A_{0} we need to consider the outer region, in which we can neglect the temperature and write η=1+η¯\eta=1+\bar{\eta}. This leads to

(1+η¯)f′′′+f′′+R(1+ff′′f2)=0(1+\bar{\eta})f^{\prime\prime\prime}+f^{\prime\prime}+R\left(1+ff^{\prime\prime}-f^{\prime 2}\right)=0 (62)

and, on matching with the inner region, that

fA0η¯2+ as η¯0,f0 as η¯.f\sim A_{0}\bar{\eta}^{2}+\cdots\quad\text{ as }\bar{\eta}\rightarrow 0,\quad f^{\prime}\rightarrow 0\quad\text{ as }\bar{\eta}\rightarrow\infty. (63)
Figure 10: Fig. 10 A plot of the critical value λc\lambda_{c} of λ\lambda against σ\sigma for R=1R=1. The region in the ( λ,σ\lambda,\sigma ) parameter plane where solutions exist is labelled on the figure
Refer to caption

Equations (62,63)(62,63) are essentially the forced-convection limit with the solution given by [8]. Their solution then gives A0A_{0} in terms of RR. For RR large [8] show that A00.6163R1/2+A_{0}\sim 0.6163R^{1/2}+\cdots, giving (dθdη)η=1\left(\frac{\mathrm{d}\theta}{\mathrm{d}\eta}\right)_{\eta=1} of O(R1/2σ1/3)O\left(R^{1/2}\sigma^{1/3}\right) for RR and σ\sigma large.

The leading-order problem (57,58,62,63) for σ\sigma large does not give a critical value for λ\lambda. In fact, the solution is independent of λ\lambda at leading order, with the buoyancy forces arising at O(σ1/3)O\left(\sigma^{-1/3}\right) in the inner region. This suggests that the critical value λc\lambda_{c} will, for a given value of RR, decrease to large negative values as σ\sigma is increased. We illustrate this in Fig. 9 with a plot of θ(1)\theta^{\prime}(1) against λ\lambda for R=1R=1 and σ=10\sigma=10 to compare with the results for σ=1\sigma=1. This figure clearly shows that the critical values λc\lambda_{c} has decreased for σ=10\sigma=10 over that for σ=1\sigma=1, giving a greater range of λ\lambda for solutions in the opposing case. The values of θ(1)\theta^{\prime}(1) have, for a given λ\lambda, also decreased for σ=10\sigma=10, in line with expression (61). In Fig. 10 we plot λc\lambda_{c} against σ\sigma again for R=1R=1, showing that λc\lambda_{c} decreases relatively slowly as σ\sigma is increased, as can be expected from the above analysis for the large λ\lambda case.

6 Conclusions

We have considered the mixed convection boundary-layer flow around an axisymmetric stagnation point. The equations for the flow and temperature fields reduce to similarity form ( 353-5 ) and involve the three parameters, the Prandtl number σ\sigma, a Reynolds number RR and a mixed convection parameter λ\lambda, as defined in (6). The similarity equations were solved numerically for representative values of the parameters RR and λ\lambda; see Figs. 1 and 2 . The main conclusions from these numerical integrations were that, for λ<0\lambda<0 (opposing flow), there was a critical value λc\lambda_{c} of λ\lambda at which there was a saddle-node bifurcation with dual solutions for λc<λ<0\lambda_{c}<\lambda<0 and no solutions for λ<λc\lambda<\lambda_{c}. For λ0\lambda\geq 0 (aiding flow) there was a single solution for all λ\lambda. An asymptotic solution for λ\lambda large was derived, with the results summarized in (22).

The occurrence of dual solutions for opposing flows is not unexpected and is consistent with many previous studies of similarity solutions in mixed convection, see [18,19,23] for example. The critical value λc\lambda_{c} was seen to depend on RR (and on σ\sigma ), see Fig. 3. The possibilities of solutions having only a limited range of existence and the existence of dual solutions for opposing flows was not noticed by Gorla [12]. The reason for this is that the results in [12] were all for R=100R=100 and our study suggests that the value of λc\lambda_{c} for this value of RR is quite large, well beyond the values for the mixed convection parameter taken in [12]. A solution for RR large was obtained which showed that λc\lambda_{c} is of O(R)O(R) in this case, as given in (30) for σ=1\sigma=1. Thus for strong external flows (large RR ) boundary-layer flows are still possible even when there are strongly opposing buoyancy forces. For RR small the flow was seen to be driven predominantly by the buoyancy forces when λ\lambda is of O(1)O(1), as summarized in (47). However, for λ\lambda small both buoyancy and the external flow have comparable effects, giving a critical value of O(R(logR))O(R(-\log R)) for RR small, as given by (55) for σ=1\sigma=1. The effect of weak external flows is then to severely limit the range of λ\lambda where there can be opposing flows.

The effect of having large values for the Prandtl number σ\sigma is to confine the thermal effects to a thin layer, of thickness O(σ1/3)O\left(\sigma^{-1/3}\right), next to the cylinder with the outer flow being essentially given by forced convection. A consequence of this is, for opposing flows, to decrease λc\lambda_{c} to large negative values, with solutions then being possible for a large range of the mixed convection parameter.

Acknowledgements CR was supported by the CEEX grant No 2-CEx06-11-96/2006. JM and IP were both supported by a Royal Society (London) Joint Project Grant.

References

  1. 1.

    Hiemenz K (1911) Die Grenzschicht an einem in den gleich formigen Flussigkeitsstrom eingetacuhten geraden Kreisszylinder. Dinglers Polytech J 326:321-324

  2. 2.

    Eckert ERG (1942) Die Berechnung des Wärmeüberganges in der laminaren Grenzschicht um stromter Korper. VDI - Forchungsheft 416:1-24

  3. 3.

    Gorla RSR (1976) Heat transfer in an axisymmetric stagnation flow on a cylinder. Appl Sci Res 32:541-553

  4. 4.

    Hommann F (1936) Der Einfluss grosser Zähigkeit bei der Stromung um den Cylinder und um die Kugel. J Appl Math Phys (ZAMP) 16:153-164

  5. 5.

    Smith FT (1974) Three dimensional stagnation point flow in a corner. Proc R Soc Lond A 344:489-507

  6. 6.

    Wang CY (1974) Axisymmetric stagnation flow on a cylinder. Quart Appl Math 32:207-213

  7. 7.

    Gorla RSR (1978) Nonsimilar axisymetric stagnation flow on a moving cylinder. Int J Eng Sci 16:392-400

  8. 8.

    Weidman PD, Putkaradze V (2003) Axisymmetric stagnation flow obliquely impinging on a circular cylinder. Eur J Mech B/Fluids 22:123-131

  9. 9.

    Weidman PD, Mahalingam S (1997) Axisymmetric stagnation-point flow impinging on a transversely oscillating plate with suction. J Eng Math 31:305-318

  10. 10.

    Gorla RSR (1979) Unsteady viscous flow in the vicinity of an axisymmetric stagnation point on a circular cylinder. Int J Eng Sci 17:87-93

  11. 11.

    Ramachandran N, Chen TS, Armaly BF (1988) Mixed convection in stagnation flows adjacent to vertical surface. J Heat Trans 110:173-177

  12. 12.

    Gorla RSR (1993) Mixed convection in an axisymmetric stagnation flow on a vertical cylinder. Acta Mech 99:113-123

  13. 13.

    Kuiken HK (1974) The thick free-convective boundary-layer along a semi-infinite isothermal vertical cylinder. J Appl Math Phys (ZAMP) 25:497-514

  14. 14.

    Naraian IP, Uberoi MS (1972) Combined forced and free convection heat transfer from thin needles in a uniform stream. Phys Fluids 15:1879-1882

  15. 15.

    Naraian IP, Uberoi MS (1973) Combined forced and free convection over thin needles. Int J Heat Mass Trans 16:1505-1511

  16. 16.

    Chen ILS (1987) Mixed convection flow about slender bodies of revolution. J Heat Trans 109:1033-1036

  17. 17.

    Wang CY (1990) Mixed convection on a vertical needle with heated tip. Phys Fluids A 2:622-625

  18. 18.

    Merkin JH (1985) On dual solutions occurring in mixed convection in a porous medium. J Eng Math 20:171-179

  19. 19.

    Merkin JH, Mahmood T (1989) Mixed convection boundary layer similarity solutions: prescribed wall heat flux. J Appl Math Phys (ZAMP) 40:51-68

  20. 20.

    Stewartson K, Jones LT (1957) The heated vertical plate at high Prandtl number. J Aeronaut Sci 24:379-380

  21. 21.

    Kuiken HK (1968) The heated vertical plate at high Prandtl number free convection. J Eng Math 2:355-371

  22. 22.

    Slater LJ (1960) Confluent hypergeometric functions. Cambridge University Press, Cambridge

  23. 23.

    Wilks G, Bramley JS (1981) Dual solutions in mixed convection. Proc R Soc Edinb 87A:349-358

2009

Related Posts