Volume 2010, Issue 1 736039
Research Article
Open Access

Mixed Convection Boundary Layer Flow from a Solid Sphere with Newtonian Heating in a Micropolar Fluid

M. Z. Salleh

M. Z. Salleh

Faculty of Industrial Science and Technology, Universiti Malaysia Pahang, Lebuhraya Tun Razak, 26300 UMP Kuantan, Pahang, Malaysia ump.edu.my

Search for more papers by this author
R. Nazar

Corresponding Author

R. Nazar

School of Mathematical Sciences, Faculty of Science and Technology, Universiti Kebangsaan Malaysia, 43600 UKM Bangi, Selangor, Malaysia ukm.my

Search for more papers by this author
I. Pop

I. Pop

Faculty of Mathematics, University of Cluj, CP 253 3400 Cluj, Romania ubbcluj.ro

Search for more papers by this author
First published: 02 March 2010

Abstract

The steady mixed convection boundary layer flow from a solid sphere in a micropolar fluid, generated by Newtonian heating in which the heat transfer from the surface is proportional to the local surface temperature, is considered. The governing boundary layer equations are first transformed into a system of nondimensional equations via the non-dimensional variables, and then into nonsimilar equations before they are solved numerically using an implicit finite-difference scheme known as the Keller-box method. Numerical solutions are obtained for the skin friction coefficient, wall temperature and heat transfer coefficient, as well as the velocity and temperature profiles with several parameters considered, namely the mixed convection parameter, the material or micropolar parameter, and the Prandtl number.

1. Introduction

The essence of the theory of micropolar fluid flow lies in the extension of the constitutive equation for Newtonian fluid, so that more complex fluids such as particle suspensions, liquid crystal, animal blood, lubrication, and turbulent shear flows can be described by this theory. The theory of micropolar fluid was first proposed by Eringen [1]. Extensive review of the theory and applications can be found in the review article by Ariman et al. [2] and quite recent papers by Rees and Bassom [3], Pop et al. [4], Nazar et al. [58], and so forth. This clearly shows the fast development of the theory of micropolar fluid. In summary, all the papers above considered the boundary condition of either a constant wall temperature or constant heat flux.

Generally, in modeling the boundary layer flow and heat transfer of these problems, the boundary conditions that are usually applied are either a constant wall temperature (CWT) or a constant wall heat flux (CHF). However, there is a class of boundary layer flow and heat transfer problems in which the surface heat transfer depends on the surface temperature. Perhaps the simplest case of which is when there is a linear relation between the surface heat transfer and surface temperature. This situation arises in conjugate heat transfer problems (see, e.g., Merkin and Pop [9]), and also when there is Newtonian heating of the convective fluid from the surface. The latter case has been first discussed in detail by Merkin [10]. The situation with Newtonian heating arises in what are usually termed conjugate convective flows, where the heat is supplied to the convecting fluid through a bounding surface with a finite heat capacity. This results in the heat transfer rate through the surface being proportional to the local difference in the temperature with the ambient conditions. However, the Newtonian heating conditions have been used only recently by Lesnic et al. [1113] and Pop et al. [14] to study the free convection boundary layer over vertical and horizontal surfaces as well as over a small inclined flat plate from the horizontal surface embedded in a porous medium. The asymptotic solution near the leading edge and the full numerical solution along the whole plate domain have been obtained numerically, whilst the asymptotic solution far downstream along the plate has been obtained analytically. Chaudhary and Jain [15, 16] studied the unsteady free convection boundary layer flow past an impulsively started vertical infinite flat plate with Newtonian heating.

This configuration occurs in many important engineering devices, for example, in heat exchanger where the conduction in solid tube wall is greatly influenced by the convection in the fluid flowing over it. Furthermore, for conjugate heat transfer around fins where the conduction within the fin and the convection in the fluid surrounding it must be simultaneously analyzed in order to obtain the vital design information and also in convection flow setup when the bounding surfaces absorb heat by solar radiation [15, 16]. This results in the heat transfer rate through the surface being proportional to the local difference in the temperature with the ambient conditions.

Recently, Salleh et al. [17, 18] employed an implicit finite-difference scheme, namely the Keller-box method to obtain numerical solutions for mixed convection boundary layer flow near the lower stagnation point of a solid sphere and forced convection boundary layer flow at a forward stagnation point with Newtonian heating in viscous fluids, respectively.

Therefore, the aim of the present paper is to study the problem of mixed convection boundary layer flow from a solid sphere with Newtonian heating in a micropolar fluid. The governing boundary layer equations are first transformed into a system of non-dimensional equations via the non-dimensional variables, and then into non-similar equations before they are solved numerically by the Keller-box method, as described in the books by Na [19] and Cebeci and Bradshaw [20]. To the best of our knowledge, this present problem (for the case of Newtonian heating) has not been considered before, so that the reported results are new.

2. Analysis

Consider a heated sphere of radius a, which is immersed in a viscous and incompressible micropolar fluid of ambient temperature T, which is subjected to a Newtonian heating (NH) as shown in Figure 1. The convective forced flow is assumed to be moving upward, while the gravity vector g acts downward in the opposite direction, where the coordinates and are chosen such that measures the distance along the surface of the sphere from the lower stagnation point and measures the distance normal to the surface of the sphere. We assume that the equations are subjected to a Newtonian heating of the form proposed by Merkin [10]. Under the Boussinesq and boundary layer approximations, the basic equations are (Nazar et al. [6, 8])

Details are in the caption following the image
Physical model and coordinate system.
()
subject to the boundary conditions
()
where and are the velocity components along the and directions, respectively, is the microrotation component normal to the plane, γ is the spin gradient viscosity, κ is the vortex viscosity, T is the local temperature, Tw is the wall temperature, g is the gravity acceleration, α = ν/Pr is the thermal diffusivity, β is the thermal expansion coefficient, ν = μ/ρ is the kinematic viscosity, μ is the dynamic viscosity, ρ is the density, Pr is the Prandtl number, k is the thermal conductivity, qw is the constant heat flux from the wall, and hs is the heat transfer parameter. It is worth mentioning that in boundary conditions (2), n is a constant and 0 ≤ n ≤ 1. The value n = 0, which indicates at the wall, represents concentrated particle flows in which the particle density is sufficiently great that microelements close to the wall are unable to rotate or called as “strong” concentration of microelements [21, 22]. The case corresponding to n = 1/2 results in the vanishing of antisymmetric part of the stress tensor and represents “weak” concentration of microelements [23]. In this case, the particle rotation is equal to fluid vorticity at the boundary for fine particle suspension. When n = 1, we have flows which are representative of turbulent boundary layer [24]. The case of n = 1/2 is considered in the present study.

Let be the radial distance from the symmetrical axis to the surface of the sphere and the local free stream velocity, which are given by

()
where j is the micro-inertia per unit mass.

We introduce now the following non-dimensional variables:

()
where Re = U  a/ν is the Reynolds number. Using (4), the system of (1) then becomes
()
()
()
()
subject to the boundary conditions
()
where λ is the mixed convection parameter and K is the material parameter which are given by
()
with Gr = gβTa3/ν2  (NH) or Gr = gβ(TwT)a3/ν2 (CWT) or Gr = gβ(aqw/k)a3/ν2 (CHF) being the Grashof number. It is worth mentioning that in both cases of CWT and CHF, λ > 0 is for the aiding or assisting flow (heated sphere) and λ < 0 is for the opposing flow (cooled sphere), while for the present case of NH, the value of λ considered is only for λ > 0.

Further, the stream function ψ is introduced and it is defined as

()
Thus, since sin x/x → 1 as x → 0, it is appropriate to introduce the transformation
()

Equations (6) to (8) then become

()
subject to the boundary conditions
()

At the lower stagnation point of the sphere, x ≈ 0, (13) reduces to the following ordinary differential equations:

()
and the boundary conditions (14) become
()
where primes denote differentiation with respect to y.

The quantities of physical interest are the skin friction coefficient, Cf and the local wall temperature distribution, θw(x), which are given by

()
where and is the wall skin friction or wall shear stress.

3. Solution Procedure

Equations in (13) subject to the boundary conditions (14) are solved numerically using the Keller- box method as described in the books by Na [19] and Cebeci and Bradshaw [20]. The solution is obtained in the following four steps.

  • (i)

    Reduce (13) to a first-order system.

  • (ii)

    Write the difference equations using central differences.

  • (iii)

    Linearize the resulting algebraic equations by Newton’s method, and write them in the matrix-vector form.

  • (iv)

    Solve the linear system by the block tridiagonal elimination technique (see Salleh et al. [25] for the details of this method).

4. Results and Discussion

Equations in (13) subject to the boundary conditions (14) are solved numerically using an efficient implicit finite-difference scheme with various parameters, namely the mixed convection or buoyancy parameter λ, the material or micropolar parameter K and the Prandtl number Pr. In this paper, the numerical solutions start at the lower stagnation point of the sphere, x ≈ 0, as given in (15) subjects to the boundary conditions (16) and proceed round the sphere up to the separation point xs. The values of the skin friction coefficient, Cf and the local wall temperature distribution, θw(x) have been obtained at different positions 0 ≤ x ≤ 1200 for following values of the material parameter K = 0 (Newtonian fluid), 1, 2, and 3 (micropolar fluid), with various values of the mixed convection parameter λ. It should be noticed that the present results were obtained up to the value of x = 1200, as given by Nazar et al. [6, 8]. Values of Pr considered are Pr = 0.7, 1, and 7 which correspond to air, electrolyte solution and water, respectively.

Due to the decoupled boundary layer equation (15) when the mixed convection parameter λ = 0      and      K = 0, there is a unique value of the reduced skin friction coefficient, f(0) = 2.4104 for all Prandtl number Pr, which is in good agreement with the value f(0) = 2.4151 found by Nazar et al. [68]. The values of f(0), −θ(0) and θ(0) for the cases of CWT, CHF, and NH are shown in Tables 1, 2, and 3, respectively. Some results reported by Nazar et al. [6, 8] for the cases of CWT and CHF and Salleh et al. [17] for the case of NH in viscous fluid (K = 0) are also included in these tables. It is found that the agreement between the previously published results with the present ones is very good. We can conclude that this method works efficiently and we are also confident that the results presented here are accurate.

Table 1. Values of f′′(0)  and−θ(0) for various values of λ with Pr = 7 and K = 1 (CWT).
λ f(0) θ(0)
Nazar et al. [8] Present Nazar et al. [8] Present
−5.0 1.0549 1.0462 1.5834 1.5681
−2.0 1.4465 1.4430 1.6726 1.6581
0.0 1.7042 1.7020 1.7275 1.7127
3.0 2.1018 2.1089 1.8081 1.7944
10.0 2.9568 2.9554 1.9617 1.9434
Table 2. Values of f′′(0)  and  θ(0) for various values of λ with Pr = 0.7 and K = 1 (CHF).
λ f(0) θ(0)
Nazar et al. [6] Present Nazar et al. [6] Present
−3.0 0.3383 0.3035 1.6592 1.6781
−2.0 0.9790 0.9725 1.4674 1.4708
0.0 1.7135 1.7112 1.3255 1.3274
2.0 2.2941 2.2923 1.2455 1.2471
10.0 4.0030 3.9911 1.0889 1.0909
Table 3. Values of f′′(0)  and  θ(0) for various values of λ with fixed Pr = 1 and K = 0 (viscous fluid) (NH).
λ f(0) θ(0)
Salleh et al. [17] Present Salleh et al. [17] Present
0.05 3.5485 3.5485 65.4268 65.4268
1.0 5.2161 5.2162 8.7364 8.7367
5.0 8.1273 8.1273 3.9495 3.9495
7.0 9.1151 9.1151 3.4047 3.4047
10.0 10.3730 10.3730 2.9258 2.9258

Tables 4(a) and 4(b) show comparisons on the values of θ(0) and −θ(0) as well as f(0), respectively, for different boundary conditions (CWT, CHF, and NH cases) with fixed Pr = 7 and K = 1 (micropolar fluid). Table 4(a) shows the values of the heat transfer coefficient −θ(0) for the case of CWT, the wall temperature θ(0) for the case of CHF, and the wall temperature θ(0) or the heat transfer coefficient −θ(0) for the case of NH. It is noticed from Table 4(a) that for the case of CWT, as λ increases, the heat transfer coefficient −θ(0) also increases. This table also shows that for the case of CHF, as λ increases, the wall temperature θ(0) decreases. However, for the case of NH, as λ increases, both the heat transfer coefficient and the wall temperature decrease. The trend for NH case is similar to the CHF case but different from the CWT case. On the other hand, Table 4(b) shows the values of the skin friction coefficient f(0) for the cases of CWT, CHF, and NH. It shows that for all the three cases, as λ increases, the skin friction coefficient f(0) also increases. The trend for NH case is similar to the CWT and CHF cases, but the values of f(0) are closer to the CWT case than the CHF case.

λ CWT CHF NH
θ(0) θ(0) θ(0) θ(0)
−2.0 1.6670 0.5878
−1.0 1.6980 0.5847
0.0 1.7272 0.5790 1.2748 2.2748
2.0 1.7805 0.5688 1.2410 2.2410
4.0 1.8286 0.5598 1.1750 2.1750
6.0 1.8725 0.5517 1.1190 2.1190
8.0 1.9159 0.5445 1.0745 2.0745
10.0 1.9509 0.5379 1.0378 2.0378
20.0 2.1101 0.5115 0.9165 1.9165
λ CWT CHF NH
f(0) f(0) f′′(0)
−2.0 1.4748 1.6244
−1.0 1.6179 1.6761
0.0 1.7568 1.7568 1.7568
2.0 2.0236 1.9102 2.1176
4.0 2.2785 2.0547 2.3937
6.0 2.5235 2.1918 2.6367
8.0 2.7603 2.3227 2.8568
10.0 2.9901 2.4483 3.0597
20.0 4.0590 3.0160 3.9136

The values of f(0) and θ(0) for various values of λ, K and Pr are shown in Tables 5 and 6, respectively. From Table 5, it is found that as K increases, the values of f(0) are higher and θ(0) are lower for K = 1 than those for K = 2 and 3. On the other hand, from Table 6, it can be seen from this table that the values of f(0) and θ(0) are higher for Pr = 0.7 than those for Pr = 1 and 7.

Table 5. Values of f′′(0)  and  θ(0) for various values of λ and K with fixed Pr = 7 (NH).
λ K = 1 K = 2 K = 3
f′′(0) θ(0) f′′(0) θ(0) f′′(0) θ(0)
0.05 1.7662 1.3736 1.4597 1.5677 1.2777 1.7382
1.0 1.9332 1.2988 1.5980 1.4970 1.3990 1.6551
5.0 2.4950 1.1527 2.0583 1.3055 1.7992 1.4333
7.0 2.7268 1.1019 2.2469 1.2450 1.9623 1.3643
10.0 3.0386 1.0378 2.5001 1.1755 2.1807 1.2853
Table 6. Values of f′′(0)  and  θ(0) for various values of λ and Pr with fixed K = 2 (NH).
λ Pr = 0.7 Pr = 1 Pr = 7
f(0) θ(0) f(0) θ(0) f(0) θ(0)
0.05 5.3893 582.9966 3.3508 269.9892 1.4597 1.5677
1.0 5.9345 33.9767 4.0100 18.9700 1.5980 1.4970
5.0 7.5188 9.7701 5.5475 6.5872 2.0583 1.3055
7.0 8.1221 7.8312 6.0922 5.4627 2.2469 1.2450
10.0 8.9099 6.2842 6.7908 4.5284 2.5001 1.1755

Tables 7 and 8 show the values of Cf and θw(x) for Pr = 0.7, K = 1 (micropolar fluid) and various values of λ, respectively. It can be seen from these tables that as the parameter λ increases, Cf increases while θw(x) decreases. The actual value of λ = λk, which first gives no separation is difficult to determine exactly as it has to be found from the equations. However, the numerical solution indicates that the value of λk which first gives no separation lies between 0.05 and 1 for fixed Pr = 0.7 and K = 1.

Table 7. Values of the local skin friction coefficient Cf for Pr = 0.7, K = 1 and various values of λ  (NH).
x λ
0.005 0.02 0.05 1 5 7
00 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000
100 2.2261 2.2280 2.2319 2.2995 2.5190 2.5862
200 4.6769 4.6803 4.6870 4.8001 5.1915 5.3111
300 7.0750 7.0797 7.0889 7.2419 7.7920 7.9602
400 9.4615 9.4673 9.4789 9.6676 10.3657 10.5797
500 11.6907 11.6976 11.7113 11.9338 12.7718 13.1966
600 13.7553 13.7631 13.7788 14.0312 15.0807 15.8120
700 15.7534 15.7622 15.7796 16.0600 17.4505 18.2886
800 17.4679 17.4775 17.4965 17.8096 19.4869 20.3704
900 19.0260 19.0362 19.0566 19.4335 21.2685 22.1885
1000 20.2155 20.2370 20.6934 22.5758 23.5257
1100 21.1823 21.6963 23.5557 24.5200
1200 22.1532 23.9418 24.8979
Table 8. Values of the wall temperature distribution θw(x) for Pr = 0.7, K = 1 and various values of λ (NH).
x λ
0.005 0.02 0.05 1 5 7
00 3972.1601 997.6703 402.7344 24.9385 7.7892 6.3404
100 14946.0423 3741.4930 1500.5887 78.2596 18.0482 13.3545
200 15962.9645 3995.3036 1601.7763 82.9188 18.8185 13.8678
300 16578.9876 4149.1161 1663.1461 85.8007 19.3253 14.2146
400 17114.5642 4282.9142 1716.5875 88.3705 19.8096 14.5542
500 17663.7427 4420.1691 1771.4575 91.0592 20.3424 14.8667
600 18250.9445 4566.9709 1830.1793 93.9733 20.9019 15.0610
700 18940.4194 4739.3755 1899.1698 97.4257 21.4484 15.3222
800 19744.1335 4940.3723 1979.6231 101.4657 22.0515 15.7276
900 20692.2273 5177.4964 2074.5533 106.1772 22.8095 16.2955
1000 5460.6204 2187.9111 111.7045 23.7877 17.0402
1100 2324.9537 118.3370 25.0444 17.9891
1200 126.5161 26.6599 19.1931

Tables 9 and 10 present the values of the local skin friction coefficient Cf and the wall temperature distribution θw(x) for λ = 1, K = 1 and various values of Pr, respectively. It is found that, for fixed λ = 1 and K = 1, as Pr increases, both the Cf and θw(x) decreases. On the other hand, for fixed Pr, as x increases, that is, from the lower stagnation point of the sphere, x ≈ 0, and proceeds round the sphere up to the point, x = 1200, both the values of Cf and θw(x) increase.

Table 9. Values of the local skin friction coefficient Cf for λ = 1, K = 1 and various values of Pr  (NH).
x Pr
0.7 1 7
00 0.0000 0.0000 0.0000
100 2.2995 1.5053 0.1490
200 4.8001 3.1609 0.3806
300 7.2419 4.7598 0.5211
400 9.6679 6.3723 0.7668
500 11.9338 7.8549 0.8887
600 14.0312 9.2817 1.0657
700 16.0600 10.7771 1.3207
800 17.8096 12.0495 1.4135
900 19.4335 13.2069 1.6429
1000 20.6934 14.0239 1.6813
1100 21.6963 14.6790 1.8646
1200 22.1532 14.9131 1.9266
Table 10. Values of the wall temperature distribution θw(x) for λ = 1, K = 1 and various values of Pr  (NH).
x Pr
0.7 1 7
00 24.9385 13.7810 1.2988
100 78.2596 47.5798 3.4913
200 82.9188 50.5583 3.8423
300 85.8007 52.3905 4.0603
400 88.3705 54.0104 4.2166
500 91.0592 55.6998 4.3814
600 93.9733 57.4598 4.5348
700 97.4257 59.2523 4.6891
800 101.4657 61.1798 4.8683
900 106.1772 63.4915 5.0758
1000 111.7045 66.3902 5.3383
1100 118.3370 70.0617 5.6635
1200 126.5161 74.7520 6.0838

The velocity and temperature profiles near the lower stagnation point of the sphere, x ≈ 0, for various values of K with fixed Pr = 7 and λ = 1 are plotted in Figures 2 and 3, respectively. From these figures, it is shown that as K increases, the velocity profiles decrease and the temperature profiles increase.

Details are in the caption following the image
Velocity profiles for various values of K with fixed Pr = 7    and    λ = 1 (NH).
Details are in the caption following the image
Temperature profiles for various values of K with fixed Pr = 7    and    λ = 1 (NH).

Figures 4 and 5 illustrate the velocity and temperature profiles near the lower stagnation point of the sphere, x ≈ 0, for various values of λ with fixed Pr = 1 and K = 1, respectively. We found that the velocity profiles increase while the temperature profiles decrease when the mixed convection parameter λ increases. Also, it is noticed that there are overshoots of the velocity profiles when λ > 1 where these overshoots take place higher for λ = 10 than for λ = 1.

Details are in the caption following the image
Velocity profiles for various values of λ with fixed Pr = 1    and    K = 1 (NH).
Details are in the caption following the image
Temperature profiles for various values of λ with fixed Pr = 1    and    K = 1 (NH).

Figures 6 and 7 illustrate the velocity and temperature profiles near the lower stagnation point of the sphere, x ≈ 0, for various values of Pr with fixed K = 2 and λ = 1, respectively. It is seen that, as the Prandtl number Pr increases, both the velocity and temperature profiles decrease. At large Pr(≫1), the thermal boundary layer is thinner than at smaller Pr. This is because for small values of Pr  (≪1), the fluid is highly conductive. Physically, if Pr increases, the thermal diffusivity decreases and this phenomena leads to the decreasing of energy ability that reduces the thermal boundary layer. It is also noticed that there are overshoots of the velocity profiles when Pr ≤ 1 where these overshoots are higher for Pr = 0.7 than for Pr = 1.

Details are in the caption following the image
Velocity profiles for various values of Pr with fixed K = 2    and    λ = 1 (NH).
Details are in the caption following the image
Temperature profiles for various values of Pr with fixed K = 2    and    λ = 1 (NH).

5. Conclusions

In this paper, we have numerically studied the problem of mixed convection boundary layer flow from a solid sphere in a micropolar fluid, generated by Newtonian heating. It is shown that the mixed convection or buoyancy parameter λ, the material or micropolar parameter K, and the Prandtl number Pr affect the flow and heat transfer characteristics. We can conclude that (for the case of NH):

  • (i)

    when Pr and K are fixed, an increase in parameter λ leads to the decrease of both the heat transfer coefficient −θ(0) and the wall temperature θ(0);

  • (ii)

    when Pr and λ are fixed, an increase in parameter K leads to the decrease of the skin friction coefficient f(0) and the increase of the wall temperature θ(0);

  • (iii)

    when Pr and K are fixed, an increase in parameter λ leads to the increase of the skin friction coefficient f(0) and the decrease of the wall temperature θ(0);

  • (iv)

    when K and λ are fixed, an increase in parameter Pr leads to the decrease of the skin friction coefficient f(0) and the wall temperature θ(0);

  • (v)

    near the lower stagnation point of the sphere, the velocity profiles decrease while the temperature profiles increase when the parameter K increases;

  • (vi)

    near the lower stagnation point of the sphere, the velocity profiles increase while the temperature profiles decrease when the parameter λ increases;

  • (vii)

    near the lower stagnation point of the sphere, both the velocity and temperature profiles decrease when the Prandtl number Pr increases.

Acknowledgment

The authors gratefully acknowledge the financial supports received from the Ministry of Higher Education, Malaysia (UKM-ST-07-FRGS0036-2009) and a research grant (RDU 090308) from the Universiti Malaysia Pahang.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.