Вы находитесь на странице: 1из 27

The current issue and full text archive of this journal is available at www.emeraldinsight.com/0961-5539.

htm

HFF 23,2

248
Received 15 August 2010 Revised 9 March 2011 28 May 2011 Accepted 8 June 2011

Natural convection of nanouids ow with nanouid-oriented models of thermal conductivity and dynamic viscosity in the presence of heat source
George C. Bourantas
Department of Medical Physics, School of Medicine, University of Patras, Patras, Greece

Eugenios D. Skouras
Department of Chemical Engineering, University of Patras, Patras, Greece and Institute of Chemical Engineering and High Temperature Chemical Processes, Foundation for Research and Technology, Patras, Greece

Vassilios C. Loukopoulos
Department of Physics, University of Patras, Patras, Greece, and

George C. Nikiforidis
Department of Medical Physics, School of Medicine, University of Patras, Patras, Greece
Abstract
Purpose The purpose of this paper is to make a numerical study of natural convection of water-based nanouids in a square cavity when a discrete heat source is embedded on the bottom wall, applying a nanouid-oriented model for the calculation of the effective thermal conductivity (Xu-Yu-Zou-Xus model) and the effective dynamic viscosity ( Jang-Lee-Hwang-Chois model). Another motivation is the numerical solution of the equations of the ow with a meshless method. Design/methodology/approach A meshless point collocation method with moving least squares (MLS) approximation is used. A test validation study of the numerical method takes place for pure water ow, as well for water/Al2O3 nanouids. The inuence of pertinent parameters such as Rayleigh number (Ra), the non-uniform nanoparticle size keeping the mean nanoparticle diameter xed, the volume fraction of nanoparticles and the location of heat source on the cooling performance are studied. Findings The presence of a discrete heat source, as well as the various thermal boundary conditions affects the characteristics of the nanouid ow and heat transfer. When the ratio of minimum to maximum nanoparticle diameter is increased, the local Nusselt number is increased and the heat source temperature is decreased. The increase of solid volume fraction of nanoparticles causes the heat source maximum temperature to decrease and the Nusselt Number to increase. Originality/value The present study constitutes an original contribution to the nanouid ow and heat transfer characteristics when a discrete heat source is presence. Nanouid-oriented models are used for the calculation of the effective thermal conductivity and dynamic viscosity. Keywords Flow, Viscosity, Thermal conductivity, Mathematical analysis, Nanouids, Effective thermal conductivity, Effective dynamic viscosity, Meshfree point collocation method Paper type Research paper

International Journal of Numerical Methods for Heat & Fluid Flow Vol. 23 No. 2, 2013 pp. 248-274 q Emerald Group Publishing Limited 0961-5539 DOI 10.1108/09615531311293452

1. Introduction A nanouid is a mixture consisting of nanometer-sized particles and bers dispersed in a liquid. This mixture has as result the alteration of physical properties of the base uid, such as viscocity, density and heat transfer, among others. The concept of nanouid, in its primitive form, was rst reported in the middle of the eighteenth century, but it was possible to put it into practice only after the tremendous development of nanotechnologies during the last decades (Choi, 1995). Choi was the rst who introduced the thermal conductivity enhancement and, measured the thermal conductivity of nanouids. Since the pioneer work of Choi, a large number of experimental and theoretical studies have been carried out by numerous researchers (Xie et al., 2002; Xuan and Le, 2000; Keblinski et al., 2002). Compared to a conventional liquid and a conventional two-phase mixture, nanouids have some interesting properties, namely, they have higher thermal conductivity, they do not block ow channels and, they induce a very small pressure drop. Furthermore, nanoparticles resist sedimentation, as compared to larger particles, due to Brownian motion and inter-particle forces. A great interest appeared for natural convection heat transfer characteristics of nanouids, due to the wide applications in electronic cooling, heat exchangers, double pane windows, etc. Nanouids are used in a wide range of engineering applications, although the development of the eld faces several challenges such as the lack of agreement between experimental (or between experimental and theoretical) results provided by different groups, the poor performance of suspensions and, the lack of theoretical understanding of the mechanisms. In order to identify innovating applications for these elds, further theoretical and experimental research investigations are needed to understand the heat transfer characteristics of nanouids (Wang and Mujumdar, 2007). Regarding the lack of agreement between experimental (or even experimental and theoretical) results provided by different groups, a representative case is the following. Authors in Khanafer et al. (2003) presented a 2D numerical simulation of natural convection of nanouids in a vertical rectangular enclosure, utilizing a dispersion model similar to that for the ow through porous media. They concluded that as the volumetric fraction of the copper nanoparticles in water was increased, at any Grashof number, heat transfer across the enclosure was increased also. However, contradictory results have been observed in the experimental studies (Pruta et al., 2003; Wen and Ding, 2005). Over the last years, numerous theoretical investigations and several models have been proposed for the effective thermal conductivity to include various mechanisms, so that to predict the anomalously high thermal conductivity of nanouids. More precisely, many possible mechanisms for the anomalous increase in nanouids heat transfer have been consider, including mechanisms as, Brownian motion of the nanoparticles, liquid layering at the liquid/particle interface, nature of the heat transport in the nanoparticles, and, the effect of nanoparticle clustering (Murshed et al., 2008). Furthermore, the increased surface area due to suspended nanoparticles, the increased thermal conductivity of the uid, the interaction and collision among particles, the intensied mixing uctuation and turbulence of the uid, and nally, the dispersion of nanoparticles, are some physical properties that may play crucial role to the determination of the thermo-physical properties of nanouids (Wang and Mujumdar, 2007). Additional to these factors the non-uniform nanoparticle size and mean

Natural convection of nanouids ow 249

HFF 23,2

250

nanoparticle diameter were taken into account in Xu et al. (2006). Besides these mechanisms, authors in Murshed et al. (2008), believe that the effects of particle surface chemistry and particles interaction for nanometer-sized particles could be signicant in enhancing the thermal conductivity of nanouids. Since the classical models were found to be unable to predict the anomalously high thermal conductivity of nanouids, many theoretical studies have been carried out to predict the anomalously increased thermal conductivity of nanouids (modern or nanouidsoriented models). A detailed summary of all classical and recently developed models for the prediction of the effective thermal conductivity of nanouids is provided in Murshed et al. (2008). The need for validation of theoretical results and the need of less expensive experiments led to the development of sophisticated numerical methods in order to facilitate the so-called numerical experiments. On the other hand, theoretical scientist needed numerical methods that could provide accurate results with low computational cost. In the eld of nanouids, the most popular numerical method used is the well established nite volume method. In recent years, research on meshless (or meshfree) methods has made signicant progress, particularly in the area of computational mechanics (Liu, 2002). Meshfree methods are a particular class of numerical simulation algorithms for the simulation of physical phenomena. Meshfree methods eliminate some or all of the traditional mesh-based view of the computational domain and rely on a particle (either Lagrangian or Eulerian) view of the physical problem (Atluri and Shen, 2002; Liu and Gu, 2005). A goal of meshfree methods is to facilitate the simulation of increasingly demanding problems, arising in engineering and science and, that require the ability to treat large deformations, advanced materials, complex geometry, nonlinear material behavior, discontinuities and singularities. A survey on some of the most signicant studies which use different types of meshless numerical methods on the solution of the Navier-Stokes equations has been given in Bourantas et al. (2010). Regarding of numerical simulations, as it is mentioned in Khanafer et al. (2003) and Wang and Mujumdar (2007), two approaches have been adopted in the literature in order to investigate the heat transfer characteristics of nanouids, a continuum and a two-phase model approach. In details, the rst approach assumes that the continuum assumption is still valid for uids with suspended nanosize particles, while the other approach uses a two-phase model for better description of both the uid and the solid phases. Additionally, the single phase model is much simpler and computationally more efcient and it is common in the open literature. Nevertheless, the great number of possible mechanism, which maybe explains the anomalously high thermal conductivity of nanouids, is difcult to be described (Wang and Mujumdar, 2007). In Aminossadati and Ghasemi (2009) presented a numerical study of natural convection cooling of a heat source embedded on the bottom wall of an enclosure lled with nanouids. Therein, the top and vertical walls of the enclosure were maintained at a relatively low temperature. The transport equations for a Newtonian uid were solved numerically with a nite volume approach using the simple algorithm. In this study Brinkmans model had been used for the effective dynamic viscosity, while Maxwells model was used for the effective thermal conductivity. The results indicate that adding nanoparticles into pure water improves its cooling performance especially at low Rayleigh numbers. Additionally, in Jang et al. (2007), thermal characteristics of natural convection

in a rectangular cavity heated from below with water-based nanouids containing alumina (Al2O3 nanouids) were theoretically investigated with Jang and Chois model for predicting the effective thermal conductivity of nanouids and various models for the effective viscosity. The theoretical results were compared with experimental results presented there. It was shown that the experimental results were between a theoretical line derived from Jang and Chois model and Einsteins model and a theoretical line from Jang and Chois model and Pak and Chos correlation. Authors in Lin and Violi (2010) analyzed the heat transfer and uid ow of natural convection in a cavity lled with Al2O3/water nanouid that operates under differentially heated walls. The Navier-Stokes and energy equations were solved numerically, coupling Xus model (Xu et al., 2006) for calculating the effective thermal conductivity and Jangs model (Jang et al., 2007) for determining the effective dynamic viscosity, with the slip mechanism in nanouids. The heat transfer rates were examined for parameters of non-uniform nanoparticle size, mean nanoparticle diameter, nanoparticle volume fraction, Prandtl number, and Grashof number. Enhanced and mitigated heat transfer effects due to the presence of nanoparticles were identied and highlighted. Based on these insights, they determined the impact of uid temperature on the heat transfer of nanouids. In the present work we investigate numerically the natural convection of water-based nanouids in a vertical cavity when a discrete heat source is embedded on the bottom wall, applying a nanouid-oriented model for the calculation of the effective thermal conductivity (Xu-Yu-Zou-Xus model), (Xu et al., 2006) and for the determination of the effective dynamic viscosity ( Jang-Lee-Hwang-Chois model), (Jang et al., 2007). The type of nanoparticles that is taken into consideration is the aluminium oxide (Al2O3). We examine the inuence of pertinent parameters such as Rayleigh number (Ra), the non-uniform nanoparticle size keeping the mean nanoparticle diameter xed, the volume fraction of nanoparticles and the location of the heat source on the cooling performance. The effects of various thermal boundary conditions on the nanouid ow formations for different locations of heat source are also studied. Moreover, we solve the equations of the ow with a meshless method. To our attention, the meshless methods have not been used until now for the calculation of natural convection of nanouid (in rectangular enclosures). One of our goals is to use the meshless point collocation method (MPCM) with moving least squares (MLS) approximation, in order to obtain numerical results for nanouids ow and verify the accuracy and the efciency of the method. For validation purposes results are compared with previous published results. The rest of the paper is organized as follows. Section 2 is referred to the physical problem description. In Section 3, we describe the main steps of the meshless point collocation numerical procedure, which is used and an algorithm validation is realized. Numerics for nanouids ow in the presence of a heat source take place in Section 4, where a natural convection problem in a rectangular enclosure lled with nanouids is investigated. Finally, at Section 5 the conclusions are presented. 2. Problem description Figure 1 shows the spatial domain of the 2D enclosure considered in the present study. A heat source is located on the thermally insulated bottom wall of the enclosure. In this gure, b is the length of heat source and d is the distance of heat source from the left wall. The horizontal upper wall is adiabatic and impermeable to mass transfer.

Natural convection of nanouids ow 251

HFF 23,2

252
Nanofluid TH TC L

y b

Figure 1. A schematic diagram of the physical model

x d q''

The vertical walls of the enclosure are maintained at a relatively high temperature (TH) the left wall and a relatively low temperature (TC) the right wall. Furthermore, another set of boundary conditions are also applied and, in order to demonstrate the crucial role of thermal boundary conditions to the thermal properties of nanouids. The thermo-physical properties of the base uid and the nanoparticles are given in Table I. As far as the nanouid ow boundary conditions, the no-slip and no penetration assumptions are imposed on the walls (Lin and Violi, 2010). Furthermore, several assumptions are made regarding the mathematical equations that describe the physical model. In details: . The nanouid in the cavity is Newtonian, incompressible, and the ow is steady, laminar and non-isothermal. . The thermo-physical properties are constant except for the density in the buoyancy force (Boussinesqs hypothesis). . The uid phase and nanoparticles are in a thermal equilibrium state. . It is assumed that the nanoparticles of Al2O3 are uniformly suspended in water; there is no aggregation of nanoparticles in the uid medium, as demonstrated
r (kg/m3)
Table I. Thermo-physical properties of water and nanoparticles Pure water Alumina (A12O3) 997.1 3,900 Cr (J/kgK) 4,179 850 k (W/mK) 0.613 46

b 102 s (K2 1)
21 1.67

Source: Lin and Violi (2010)

experimentally in Jang et al. (2009) with the use of transmission electron microscopy (TEM). Actually, there are few agglomerated nanoparticles and most nanoparticles suspended in a nanouid would not aggregate. Nanoparticles are spherical.

Natural convection of nanouids ow 253

Fluid ow is governed by the Navier-Stokes equations, a well-known coupled set of non-linear partial differential equations expressing the conservation of mass, linear momentum, and energy. The governing equations in terms of the primary variables formulation are: 7 u * 0; u * 7 u *   1  1 mnf 2 * rb nf g T * 2 T* 7p * 7u ; 2 / rnf rnf rnf u * 7 T * anf 72 T * ; 1 2 3

where u * is the dimensional velocity vector, g is the acceleration of gravity, p* is the pressure, T * is the temperature, while the effective density of the nanouid is given as:

rnf 1 2 wrf wrs

and w is the solid volume fraction of nanoparticles. The thermal diffusivity of the nanouid is:

anf

knf ; rcp nf

where, the heat capacitance of the nanouid is given as: rcp nf 1 2 wrcp f w rcp s : 6

Additionally the thermal expansion coefcient of the nanouid can be determined by: rbnf 1 2 wrbf w rbs : 7

The subscript f is referred to the base uid, while the subscript s to the solid nanoparticles. In order to explain the observed phenomena, many theoretical studies on the effective thermal conductivity in nanouids have been proposed, with the various models grouped in two main categories (Murshed et al., 2008). As far as the rst category, it considers stationary nanoparticles in multiphase systems, including a static model for heat conductivity and the second category include a dynamic model for heat conductivity. Recently, Xu et al. (2006) derived a new model for predicting the thermal conductivity of nanouids by taking into account the fractal distribution of nanoparticle sizes and heat convection between nanoparticles and liquids due to the Brownian motion of nanoparticles in uids. The proposed model is expressed as a function of the average size of nanoparticles, fractal dimension, concentration of

HFF 23,2

nanoparticles, temperature and properties of uids. In this way, the total dimensionless thermal conductivity of nanouids is expressed as: knf ks 2kf 2 2wkf 2 ks Nup df 2 2 Df Df dmax =dmin 12Df 2 1 1 c : 8 kf Pr ks 2kf wkf 2 ks 1 2 Df 2 d max =dmin 22Df 2 1 d p

254

The aforementioned model of Xu et al. (2006) has been chosen in the present study for the description of thermal conductivity of nanouids. The rst term is the Hamilton and Crosser model (Hamilton and Crosser, 1962), which considers the nanoparticles in the liquid as stationary, and the second term is the thermal conductivity based on heat convection due to Brownian motion. c is an empirical constant, which is relevant to the thermal boundary layer and dependent on different uids (e.g. c 85 for the de-ionized water and c 280 for ethylene glycol) but independent of the type of nanoparticles. Nup is the Nusselt number for liquid owing around a spherical particle and equal to Nup 2 for a single particle in this work. The uid molecular diameter df is taken as 4.5 102 10 m for water in present study, as in Lin and Violi (2010). Pr is the Prandtl number, w and dp are the nanoparticle volume fraction and mean nanoparticle diameter, respectively. The fractal dimension Df is determined by: Df 2 2 ln w ; lndp;min =dp;max 9

where dp,max and dp,min are the maximum and minimum diameters of nanoparticles, respectively. With the given/measured ratio of (dp,min /dp,max ) R, the minimum and maximum diameters of nanoparticles can be obtained with mean nanoparticle diameter dp from the statistical property of fractal media:   Df 2 1 dp;min 21 Df 2 1 ; dp;min d p 10 dp;max d p d p;max Df Df On the other hand, regarding the effective dynamic viscosity of water/Al2O3 nanouids, the model of Jang and Choi is used ( Jang et al., 2007), respectively. This model was presented by Jang and Choi for a uid, which contain a dilute suspension of small rigid spherical particles and it accounts for the slip mechanism in nanouids, capturing the new rheological features of nanouids. Thus, the effective dynamic viscosity of nanouids in the above equations is: " #  221 dp 2=311 mnf mf 1 2:5w 1 h w : 11 H The empirical constant 1 and c are 2 0.25 and 280 for Al2O3, respectively. Introduce the following dimensionless quantities: x x* ; L y y* ; L u u *L ; af v v *L ; af p p *L 2 ; ra2 f

* T * 2 TC * 2 T* TH C

12

where L is the dimensional cavity width. Then, the continuity, momentum, and energy equations (1)-(3), in non-dimensional form, in Cartesian system of coordinates are given as: 7 u 0; u 7 u 2 1 7p 1 2 w w rs =r f 1 2:5w1 h dp =H 221 w 2=3 1 1Pr 2 7u 1 2 w w rs =rf 1 2 w w rs bs =rf bf RaPr uj ; 1 2 w w rs =rf knf =kf 72 u : 1 2 w w rcs =rcf 14 13

Natural convection of nanouids ow 255

u 7 u

15

Taking the curl of the two sides for the above equation we get: 7 v 27 2 u ; 1 2:5w1 h d p =H 221 w 2=3 1 1Pr 2 7v 1 2 w w rs =rf 1 2 w w rs bs =rf bf RaPr u k; x 1 2 w w rs =rf knf =kf 72 u ; 1 2 w w rcs =rcf 16

2v 7 u u 7 v

17

u 7u

18

where the Reynolds, Prandtl and Rayleigh numbers of the ow are dened as: Re

rf UD ; mf

Pr

mf ; a f rf

Ra

g b H 3 T H 2 T L : na

19

For 2D, v vk and u (u, v) and the above equations can be written: 72 u 2

v ; y

20

72 v

v ; x

21

HFF 23,2

v v 1 2:5w1 hdp =H 221 w 2=3 1 1Pr v u x y 1 2 w w rs =rf 1 2 w w rs bs =rf bf RaPr u ; x 1 2 w w rs =rf u u knf =kf u v x y 1 2 w w rcs =rcf


 2  v 2 v x 2 y 2

22

256

 2 u 2 u : x 2 y 2

23

The boundary conditions, used to solve the equations (20)-(23) are as follows: u v 0 and u 1 for x 0 and 0 # y # 1; u v 0 and u 0 for x 1 and 0 # y # 1; u 0 for y 0 and 0 # x # D 2 0:5B; u v 0 and y u kf 2 u v 0 and for y 0 and D 2 0:5B # x # D 0:5B; 24 y knf u 0 for y 0 and D 0:5B # x # 1; u v 0 and y u u v 0 and 0 for y 1 and 0 # x # 1: y The local Nusselt number on the heat source surface can be dened as: Nus hL ; kf 25

where, h is the convection heat transfer coefcient: h q Ts 2 Tc 26

and Ts is the dimensional temperature of the heat source. Rearranging the local Nusselt number by using the dimensionless parameters (equation (12)) yields: Nus x 1 ; us x 27

where, us is the dimensionless heat source temperature. The average Nusselt number (Num) is determined by integrating Nus along the heat source: 1 Num B Z
D0:5B

Nus x dx:

28

D20:5B

The thermal performance of the nanouid-lled enclosure under different thermal boundary conditions is studied for a range of solid volume fractions (0 # w # 0.2). 3. Algorithm validation An iterative scheme is utilized for the solution of the velocity-vorticity formulation of the Navier-Stokes equations (conservation of mass, linear momentum and energy). In the majority of the incompressible ow problems modeled with Navier-Stokes equations the most natural boundary conditions arises when the velocity is prescribed all over the boundaries of the problem. The vorticity boundary conditions are determined iteratively from computations. Thus, the solution algorithm used for the discretized uid ow set of equations (20)-(23), must ensure coupled satisfaction of all the equations at convergence. For the numerical solution of the equations of the ow, a MPCM with MLS approximation for the eld variables is applied. The method is developed and presented analytically in Bourantas et al. (2010). Moreover, we follow the approach of the so-called MAC algorithm proposed in Harlow and Welch (1965) and used in Bourantas et al. (2010). It permits us to attain solutions for high values of characteristic numbers of the ow. Two representative examples are executed for the validation of the meshless scheme. First validation steady 2D square cavity ows of buoyancy-driven laminar heat transfer The numerical method was implemented in a MATLAB program. The effect of grid resolution was examined in order to select the appropriate grid density. Thus, an extensive mesh testing procedure was conducted to guarantee a grid independent solution by comparisons with the results in the literature on steady 2D square cavity ows of buoyancy-driven laminar heat transfer. In order to ensure the accuracy and the validity of the proposed scheme, we analyze a system composed of pure uid in an enclosure with Pr 0.7 and different values of Rayleigh number. Thus, the present numerical code was validated against the results of other natural convection studies in enclosures developed by Lin and Violi (2010), Tiwari and Das (2007), Davis (1983), Markatos and Pericleous (1984) and Hadjisophocleous et al. (1998). Table II shows the comparison between the results obtained with the MPC method and the values presented in the literature. Second validation steady 2D square cavity ows of buoyancy-driven laminar heat transfer of nanouids without the presence of heat source Different uniform meshes of 41 41, 81 81, 121 121, 161 161 and 201 201 grid points, were examined for Gr 105, Pr 6, w 0.05, dp 5 nm and R 0.001 or 0.007. It was found that a grid size of 161 161 ensures a grid independent solution. The numerical results were compared with those obtained using the Non-Staggered Articial Pressure for Pressure-Linked equation (NAPPLE) algorithm (Lee and Tzong, 1992). Furthermore, the present code was tested against the code used in Lin and Violi (2010). Results are shown in Figure 2. 4. Numerical results The main objective of the present investigation is to explore the heat transfer behavior of natural convection inside a cavity with water/Al2O3 nanouid with a heat source

Natural convection of nanouids ow 257

HFF 23,2
(a) Ra 103 umax y vmax x/Nu (b) Ra 104 umax y vmax x/Nu (c) Ra 105 umax y vmax x/Nu (d) Ra 106 umax y vmax x/Nu

MPC 3.652 0.812 3.700 0.175 1.118 16.191 0.825 19.662 0.118 2.234 34.93 0.887 63.517 0.068 4.508 68.569 0.937 220.824 0.0375 8.8852

Lin and Violi (2010) 3.597 0.819 3.690 0.181 1.118 16.158 0.819 19.648 0.112 2.243 36.732 0.858 68.288 0.063 4.511 66.46987 0.86851 222.33950 0.03804 8.757933

Tiwari and Das (2007) 3.642 0.804 3.7026 0.178 1.0871 16.1439 0.822 19.665 0.110 2.195 34.30 0.856 68.7646 0.05935 4.450 65.5266 0.839 219.7361 0.04237 8.803

Davis (1983) 3.649 0.813 3.697 0.178 1.118 16.178 0.823 19.617 0.119 2.243 34.73 0.855 68.59 0.066 4.519 64.63 0.85 217.36 0.0379 8.799

Markatos and Hadjisophocleous Pericleous (1984) et al. (1998) 3.544 0.832 3.593 0.168 1.108 16.18 0.832 19.44 0.113 2.201 35.73 0.857 69.08 0.067 4.430 68.81 0.872 221.8 0.0375 8.754 3.544 0.814 3.586 0.186 1.141 15.995 0.814 18.894 0.103 2.29 37.144 0.855 68.91 0.061 4.964 66.42 0.897 226.4 0.0206 10.39

258

Table II. Comparison of pure uid solutions with previous works in an enclosure for Pr 0.7 with different Rayleigh numbers

located at the bottom of the enclosure, applying a nanouid-oriented model for the calculation of the effective thermal conductivity and the effective dynamic viscosity. Specically, we will analyze steady state ow elds, temperature elds, and heat transfer rates for various values of the Rayleigh number, the ratio of minimum to maximum nanoparticle diameter keeping the mean nanoparticle diameter xed, the nanoparticle volume fraction and the locations of the heat source. Moreover, the effects of various thermal boundary conditions on the nanouid ow and heat transfer characteristic are examined. In all calculations the Prandtl number is constant and equal to Pr 6.0. 4.1 Effect of Rayleigh number (or the effect of nanouid temperature) Figure 3 shows the streamlines (left) and isotherms (right) for the nanouid (solid line) with solid volume fraction w 0.05 and the pure water (dashed line) at different Rayleigh numbers, where the heat source is located in the middle of the bottom wall (D 0.5), with D being the dimensionless distance of heat source from the left wall (d/L), while the dimensionless length of the heat source B (b/D) is B 0.4. The streamlines ow patterns show a primary re-circulating clockwise vortex that occupies the bulk of the cavity, while the isotherms display different behaviours as Rayleigh number is increased. When conduction dominates the ow regime, that is for Ra 103 and Ra 104, the isotherms tend to be vertical, with a light curvature along the heat source.

y = 11.6994
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Natural convection of nanouids ow 259

R = 0.001 Gr = 104

(a)
y = 13.1188
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

R = 0.007 Gr = 104

y = 23.0660
1 0.9 0.8 0.7

(b)
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

R = 0.001 Gr = 105

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(c)
y = 25.8972
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

R = 0.007 Gr = 105

(d)

Notes: (a) Gr = 104, (knf /kf) = 1.424 and R = 0.001; (b) Gr = 104, (knf /kf) = 1.684 and R = 0.007; (c) Gr = 105, (knf /kf) = 1.424 and R = 0.001; (d) Gr = 105, (knf /kf) = 1.684 and R = 0.007 for Pr = 6, dp = 5 nm and j = 0.05

Figure 2. Streamlines (left) and isotherms (right) for the enclosures lled with nanouids at different Grashof numbers and different values of R

HFF 23,2
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

ymin, nf = 0.9394, ymin, nf = 1.2353


1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

260

Ra = 103

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(a)
ymin, nf = 5.6171, ymin, nf = 6.4593
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

Ra =104

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(b)
ymin, nf = 13.5387, ymin, nf = 14.5594
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

Ra =105

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(c)
ymin, nf = 26.1625, ymin, nf = 28.3153 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

Figure 3. Streamlines (left) and isotherms (right) for the enclosures lled with pure uid (dashed lines) and nanouid (solid lines) at different Rayleigh numbers

Ra =106

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(d)

Notes: (a) Ra = 103; (b) Ra = 104; (c) Ra = 105; (d) Ra = 106, when (knf /kf) = 1.424, R = 0.001, D = 0.5 and B = 0.4; the maximum value of streamfunction is presented

As the Rayleigh number is increased, streamlines exhibit stronger ow patterns and isotherms display more distinguished boundary layers near the vertical left and right edges, while in the middle, isotherms tend to become horizontal. In fact the ow and temperature patterns are inuenced by the presence of nanoparticles comparing them with corresponding of pure uid. Figure 4(b) shows the surface temperature at the bottom wall where the heat source is located. It can be observed that it is not uniform. It has a maximum value located at the left end of the heat source, which is near the hot wall. At this exact point the local Nusselt number is minimum, as it can be seen at Figure 4(a). More precisely, Figure 4(a) shows the prole of local Nusselt number along the bottom wall of the enclosure and along the heat source at different Rayleigh numbers, when the solid volume fraction w is constant and equal to w 0.05 and the ratio of the minimum to maximum nanoparticles diameter R is also constant and equal to R 0.001. As we can see from Figure 4(a) and (b), when the Rayleigh number increases the temperature maximum decreases, while the Nusselt number maximum increases. As it can be observed the reduction in the heat source maximum temperature is a result of the enhancement in the heat removal process. 4.2 Effect of non-uniform nanoparticle diameter keeping the mean nanoparticles diameter xed In this part we examine the effect of fractal distributions on the heat transfer effect in terms of minimum to maximum nanoparticle diameter R on the steady-state variation of the streamlines and isotherms for Ra 105 and Ra 106, while the mean diameter, nanoparticle volume fraction and the Prandtl number are xed at 5 nm, 0.05 and 6, respectively. The heat source is located in the middle of the bottom wall (D 0.5), while the dimensionless length of the heat source is B 0.4. Figure 5 shows the effect of R on the steady-state variation of the streamlines and isotherms for Ra 105 and Ra 106. The intensity of ow patterns are depicted since the values of streamline contours are presented. In this physical model, the ow patterns are characterized by a primary re-circulating clockwise vortex that occupies the bulk of the cavity. As R is increased from R1 0.001 to R2 0.007, the ow patterns remain quite the same while the absolute circulation strength is enhanced due to relatively high velocity of the uid ow.
7 6 5 Nu 4 3 2 1 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 X
Ra = 103 Ra = 104 Ra = 105 Ra = 106

Natural convection of nanouids ow 261

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 X
Ra = 103 Ra = 104 Ra = 105 Ra = 106

(a)

(b)

Figure 4. The prole of local Nusselt number and the local temperature along the bottom wall of the enclosure and along the heat source at different Rayleigh numbers, when w 0.05, (knf /kf) 1.424 and R 0.001

HFF 23,2
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

y = 13.5387
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

262

R = 0.001 Ra = 105

(a)
y = 26.1625
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

R = 0.001 Ra = 106

(b)
y = 15.1083
1 0.9 0.8 0.7 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

R = 0.007 Ra = 105

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(c)
y = 29.2995
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 5. Streamlines and isotherm contours for R1 0.001 and R2 0.007 with (knf /kf) 1.424 and (knf /kf) 1.684, respectively, at different Rayleigh numbers

R = 0.007 Ra = 106

(d)

The local Nusselt number for these conditions is calculated using equation (27) and the result reported in Figure 6 shows that the higher the Ra number, the larger the heat transfer. It can be shown in Figure 6(a) that, as the R is increased the local Nusselt number is increased. As a consequence, nanouids enhance heat transfer in both large and small buoyancy conditions (Figure 6(b)). 4.3 The effects of solid volume fraction In this part of the study, an enclosure lled with water/Al2O3 nanouid with a heat source (B 0.4) located in the middle of the bottom wall (D 0.5) is considered. Furthermore, we analyze the effect of the nanoparticle volume fraction (0 # f # 0.2) on the heat transfer characteristics. Figure 7 shows the streamlines (left) and isotherms (right) for the nanouid, with Ra 105, Pr 6.0, volume fraction having values 0.0 # f # 0.2 and R 0.001. It is obvious that the ow and temperature patterns are strongly inuenced by the presence of nanoparticles. Adding nanoparticles to the base uid (pure water) reduces the strength of ow eld, as in Ho et al. (2008). This reduction is more pronounced at low Rayleigh numbers where conduction heat transfer dominates. It is also apparent that as nanoparticles are added, the maximum dimensionless temperature is increased which is an indication of enhancing the enclosure thermal performance. At Figure 8 the proles of the local Nusselt number and the temperature values are plotted along the heat source for different volume fraction of the nanoparticles. The lowest and the highest local Nusselt number are obtained at both ends of the heat source, respectively. As it can be observed, as the volume fraction increases the local Nusselt number increases thus, increased cooling performance is observed with increasing volume fraction. 4.4 The effect of heat source location and various thermal boundary conditions In this part of the study, an enclosure lled with water/Al2O3 nanouid (w 0.05) with a heat source (B 0.4) is considered. A numerical study has been conducted altering the location of the heat source. Two sets of thermal boundary conditions are applied. First set of boundary conditions. The rst set of thermal boundary conditions is that shown in Figure 1. That is, a heat source is located on the thermally insulated bottom
6.5 6 5.5 5 Nu 4.5 4 3.5 3 2.5 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 x
Ra = 105 R1 Ra = 105 R7 Ra = 106 R1 Ra = 106 R7

Natural convection of nanouids ow 263

0.34 0.32 0.3 0.28 0.26 0.24 0.22 0.2 0.18 0.16

Ra = 105 R1 Ra = 105 R7 Ra = 106 R7 Ra = 106 R1

0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 x

Figure 6. Local Nusselt number and temperature distributions along the heat source for R1 0.001 and R2 0.007

HFF 23,2
1 0.9 0.8

y = 14.5594
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

264
j = 0.0

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

(a) y = 14.0046
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

j = 0.05

(b) y = 15.0026
1 0.9 0.8 0.7 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

j = 0.1

0.6 0.5 0.4 0.3 0.2 0.1 0

Figure 7. Streamlines (left) and isotherms (right) for the enclosures lled with water/Al2O3 nanouid, when Ra 105, R 0.001, D 0.5 and B 0.4 at different volume fractions

(c)

(continued )

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

y = 15.9367

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

Natural convection of nanouids ow 265

j = 0.15

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(d) y = 18.8820
1 0.9 0.8 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

j = 0.2

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

(e)

Notes: (a) j = 0.0; (b) j = 0.05; (c) j = 0.1; (d) j = 0.15; (e) j = 0.2

Figure 7.

6 5.5 5 Nu 4.5 4 3.5 3

= 0.00 = 0.05 = 0.1 = 0.15 = 0.2

0.36 0.34 0.32 0.3 0.28 0.26 0.24 0.22 0.2 0.18

= 0.00 = 0.05 = 0.1 = 0.15 = 0.2

2.5 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 x

0.16 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 x

Figure 8. Local Nusselt number and temperature distributions along the heat source at different volume fractions

HFF 23,2

266

wall of the enclosure. The horizontal upper wall is adiabatic and impermeable to mass transfer. The vertical walls of the enclosure are maintained at a relatively high temperature (TH) the left wall and a relatively low temperature (TC) the right wall. Figures 9 and 10 show the streamlines (left) and the isotherms (right) for different locations of the heat source on the bottom wall, while Figures 11 and 12 show the local Nusselt number and the temperature variation at the heat source, considering a Rayleigh number Ra 105, volume fraction w 0.05, R 0.001 or R 0.007 and ((knf/kf) 1.424) or ((knf/kf) 1.684), respectively. In both cases, the ow patterns are characterized by a primary re-circulating clockwise vortex that occupies the bulk of the cavity. Even though the ow patterns are the same, their strength is changed when R is varied. Concerning the isotherms, they show that as the heat source moves away from the left hot wall, the maximum ow temperature increases, for both cases. This can be explained by the distance that the uid needs to travel in the circulating cell to exchange the heat between the heat source and the left hot wall. It is clear that the isotherms follow the movement of the heat source. From Figures 11 and 12 we see that the closer the heat source is to the left hot wall, the lower heat removal and the higher heat source maximum temperature is achieved. The maximum heat removal is at the right end of heat source, while the maximum heat source temperature is at the left end of the heat source for all locations of the heat source. Second set of boundary conditions. Since no important difference appeared to the streamlines due to the different location of the heat source, different boundary conditions are used and examined for water/Al2O3 nanouid. By this, it can be obtained the inuence of the heat source location to the cooling properties of the nanouid used. The boundary conditions used and the non-dimensionalization procedure are that depicted in Aminossadati and Ghasemi (2009), were Cu-nanoparticles are used and all the walls have steady temperature, while the heat source is located at the bottom wall, being adiabatic. Figure 13 shows the streamlines (left) and the isotherms (right) for different locations of the heat source on the bottom wall, considering a Rayleigh number Ra 105, w 0.05, R 0.001 and (knf/kf) 1.424. Results are comparable with them obtained in Aminossadati and Ghasemi (2009), where classical models, for the thermal conductivity and dynamic viscosity, were used. Namely, Brinkmans model was used for the effective dynamic viscosity and Maxwells model was used for the effective thermal conductivity. As it can be seen, when the heat source is located close to the left wall two unsymmetrical circulating cells with unequal strengths are formatted. The strength of the circulating cells increases as the heat source moves away from the left wall. When the heat source reaches the middle of the bottom wall, the two circulating cells in the enclosure have the same intensities. As far as the isotherms, we see that the maximum ow temperature increases as the heat source moves away from the cold wall. In Figure 14 the vertical velocity proles along the midsection of the enclosure is presented. The results are much closed to them in Aminossadati and Ghasemi (2009), and verify the existence of two counter-rotating circulating cells for all different locations of the heat source. Unsymmetrical velocity proles indicate unbalanced circulating cells within the enclosure for the cases, where the heat source is not located in the middle. As the heat source is moving towards the middle, the maximum velocity

ymax = 13.5207
1 0.9 0.8 0.7

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

Natural convection of nanouids ow 267

D = 0.3

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(a) ymax = 13.5445


1 0.9 0.8 0.7 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

D = 0.4

0.6 0.5 0.4 0.3 0.2 0.1 0

(b) ymax = 13.5387


1 0.9 0.8 0.7 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

D = 0.5

0.6 0.5 0.4 0.3 0.2 0.1 0

(c)

Notes: (a) D = 0.3; (b) D = 0.4; (c) D = 0.5 for Ra = 105, j = 0.05, B = 0.4, R = 0.001 and (knf /kf) =1.424

Figure 9. Streamlines (left) and isotherms (right) for the enclosures lled with water/Al2O3 nanouid for different locations of the heat source

HFF 23,2

1 0.9 0.8 0.7

Ymax = 15.0869

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

D = 0.3

268

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(a)
1 0.9 0.8 0.7

Ymax = 15.1127

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

D = 0.4

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(b)
1 0.9 0.8 0.7

Ymax = 15.1083
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

D = 0.5

0.6 0.5 0.4 0.3 0.2

Figure 10. Streamlines (left) and isotherms (right) for the enclosures lled with water/Al2O3 nanouid for different locations of the heat source

0.1 0

(c)

Notes: (a) D = 0.3; (b) D = 0.4; (c) D = 0.5, for Ra = 105, j = 0.05, B = 0.4, R = 0.007 and (knf/kf) = 1.684)

increases and the prole becomes symmetrical. Finally, Figure 15 shows the local Nusselt number and the temperature variation along the heat source, respectively. As the heat source moves towards the middle of the bottom wall, the maximum temperature increases while the local Nusselt number decreases. This happens because the heat source is moving away from the cold wall. 5. Conclusions Natural convection in a partially heated enclosure from below, lled with water/Al2O3 nanouid has been numerically investigated. Nanouid-oriented models were applied for the calculation of the effective thermal conductivity (Xu-Yu-Zou-Xus model) and the effective dynamic viscosity (Jang-Lee-Hwang-Chois model) of the nanouid. The inuence of pertinent parameters such as the Rayleigh number, the non-uniform nanoparticle size keeping the mean nanoparticle diameter xed, the volume fraction of nanoparticles and the location of the heat source on the cooling performance, were studied. For the rst set of boundary conditions, we conclude the following: . When Rayleigh number is increased the thermal properties of the nanouid are increased. The surface temperature, at the bottom wall where the heat source is located, has a maximum value located at the left end of the heat source, which is near the hot wall. At this exact point the local Nusselt number present a minimum. Although in Aminossadati and Ghasemi (2009), different boundary conditions
5.5 5 4.5 4 Nu 3.5 3 2.5 2 1.5 0.1 0.2 0.3 0.4 x 0.5 0.6 0.7 0.8
D = 0.3 D = 0.4 D = 0.5

Natural convection of nanouids ow 269

0.65 0.6 0.55 0.5 0.45 0.4 0.3 0.25 0.2 0.1 0.2 0.3 0.4 x 0.5 0.6 0.7 0.8 0.35
D = 0.3 D = 0.4 D = 0.5

Figure 11. Nusselt and temperature along the heat source for R 0.001 and (knf/kf) 1.424

6 5.5 5 4.5 Nu 4 3.5 3 2.5 2 1.5 0.1 0.2 0.3 0.4 x 0.5 0.6 0.7 0.8
D = 0.3 D = 0.4 D = 0.5

0.65 0.6 0.55 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.1 0.2 0.3 0.4 x 0.5 0.6 0.7 0.8
D = 0.3 D = 0.4 D = 0.5

Figure 12. Nusselt and temperature along the heat source for R 0.007 and (knf/kf) 1.684

HFF 23,2
1 0.9

Yleft = 0.4601, |Yright| = 4.0144

qmax = 0.1570
1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

270
D = 0.3

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Yleft = 1.1099, |Yright| = 4.1329


1 0.9 0.8 0.7 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0

qmax = 0.1642

D = 0.4

0.6 0.5 0.4 0.3 0.2 0.1 0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Yleft = 2.8947, |Yright| = 2.8947


1 0.9 0.8 0.7

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

qmax = 0.1667

D = 0.5

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 13. Streamlines (left) and isotherms (right) for different locations of the heat source for Ra 105, w 0.05, B 0.4, R 0.001 and (knf/kf) 1.424

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

20 15 10 5 0 5 10 15 20 0 0.1 0.2 0.3 0.4 0.5 x 0.6 0.7 0.8 0.9 1 v


D = 0.3 D = 0.4 D = 0.5

Natural convection of nanouids ow 271

Note: Ra = 105, j = 0.05, B = 0.4, R = 0.001, (knf /kf) = 1.424


13 12 11 Nu 10 9 8 7 6 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
D = 0.3 D = 0.4 D = 0.5

Figure 14. Vertical velocity proles along the mid-section of the enclosure for different locations of the heat source

0.17 0.16 0.15 0.14 0.13 0.12 0.11 0.1 0.09 0.08 x 0.1 0.2 0.3 0.4 x 0.5 0.6 0.7 0.8
D = 0.3 D = 0.4 D = 0.5

Figure 15. Nusselt and temperature along the heat source for R 0.001 and (knf /kf) 1.424

and classical models for the thermal conductivity and dynamic viscosity, are applied, conclusions similar to that presented there can be derived. That is, increasing the Rayleigh number strengthens the natural convection ows, increasing the Nusselt number and decreasing the heat source temperature. When the ratio of minimum to maximum nanoparticle diameter, R, is increased from 0.001 to 0.007, the heat transfer characteristics of the nanouid can be enhanced, as in Lin and Violi (2010) where there is not a discrete heat source embedded on the bottom wall. As the R is increased the local Nusselt number is increased and the heat source temperature is decreased for both values of Rayleigh number. The ow and temperature patterns are inuenced by the presence of nanoparticles. Adding nanoparticles to the base uid (pure water) can reduce the strength of ow eld. It is also apparent that the increase of solid volume fraction of nanoparticles causes the heat source maximum temperature to decrease and the Nusselt Number

HFF 23,2

to increase. It means that adding nanoparticles into water improves its cooling performance, applying the rst set of boundary conditions. The location of the heat source proved to signicantly affect the heat source maximum temperature, as well as the nanouid ow and heat transfer characteristics, for both set of thermal boundary conditions which are applied. More specic, for the rst set of boundary conditions the closer the heat source is to the left hot wall, the lower heat removal and the higher heat source maximum temperature is achieved. The maximum heat removal is at the right end of heat source, while the maximum heat source temperature is at the left end of the heat source for all locations of the heat source. For the second set of boundary conditions as the heat source moves towards the middle of the bottom wall, the maximum temperature increases while the local Nusselt number decreases. Finally, for the second set of boundary conditions, the heat source maximum temperature continuously increases as the heat source moves from the left wall towards the middle of the bottom wall of the enclosure. The MPCM and MLS approximation were used for the solution of the equations of the ow. A velocity-vorticity formulation, utilizing a velocity-correction method was applied in order to solve the governing equations. The efciency and the validity of the proposed numerical scheme were demonstrated, since the numerical results obtained were compared with those obtained using the nite volume method. A relative dense grid of 161 161 regular distributed nodes was used, providing accurate numerical results. Various grids were used ensuring the grid independence of the solution. A velocity-correction method was implemented, ensuring the continuity equation. Since nanouids are used or could be used in a wide range of engineering applications, although the development of the eld faces several challenges, further theoretical, numerical and experimental research investigations are needed, so that information to be transformed from one way of research to other and to understand thus the mechanism of the heat transfer characteristics of nanouids.
References Aminossadati, S.M. and Ghasemi, B. (2009), Natural convection cooling of a localised heat source at the bottom of a nanouid-lled enclosure, European Journal of Mechanics B/Fluids, Vol. 28, pp. 630-40. Atluri, S.N. and Shen, S.P. (2002), The Meshless Local Petrov-Galerkin (MLPG) Method, Tech Science Press, Encino, CA, p. 440. Bourantas, G.C., Skouras, E.D., Loukopoulos, V.C. and Nikiforidis, G.C. (2010), Meshfree point collocation schemes for 2D steady state incompressible Navier-Stokes equations in velocity-vorticity formulation for high values of Reynolds numbers, CMES: Computer Modeling in Engineering & Sciences, Vol. 59, pp. 31-64. Choi, S. (1995), Enhancing thermal conductivity of uids with nanoparticles, ASME Fluids Engineering Division, Vol. 231, pp. 99-105. Davis, G.D.V. (1983), Natural convection of air in a square cavity: a benchmark solution, Int. J. Numer. Meth. Fluids, Vol. 3, pp. 249-64. Hadjisophocleous, G.V., Sousa, A.C.M. and Venart, J.E.S. (1998), Predicting the transient natural convection in enclosures of arbitrary geometry using a nonorthogonal numerical model, Numerical Heat Transfer, Part A Applications, Vol. 13, pp. 373-92.

272

Hamilton, R.L. and Crosser, O.K. (1962), Thermal conductivity of heterogeneous two component systems, Indus. Eng. Chem. Fund., Vol. 1, pp. 187-91. Harlow, F.H. and Welch, J.E. (1965), Numerical calculation of time-dependent viscous incompressible ow of uid with free surface, Physics of Fluids, Vol. 8, pp. 2182-89. Ho, C.J., Chen, M.W. and Li, Z.W. (2008), Numerical simulation of natural convection of nanouid in a square enclosure: effects due to uncertainties of viscosity and thermal conductivity, International Journal of Heat and Mass Transfer, Vol. 51, pp. 4506-16. Jang, S.P., Lee, J.H., Hwang, K.S. and Choi, S.U.S. (2007), Particle concentration and tube size dependence of viscosities of Al2O3-water nanouids owing through micro- and minitubes, Applied Physics Letters, Vol. 91, pp. 243112/1-243112/3. Jang, S.P., Lee, J.-H., Hwang, K.S., Moon, H.J. and Choi, S.U.S. (2009), Response to comment on particle concentration and tube size dependence of viscosities of Al2O3-water nanouids owing through micro- and minitubes, Applied Physics Letters , Vol. 94, pp. 066102/1-066102/3. Keblinski, P., Phillpot, S.R., Choi, S.U.S. and Eastman, J.A. (2002), Mechanisms of heat ow in suspensions of nanosized particles nanouids, International Journal of Heat and Mass Transfer, Vol. 45, pp. 855-63. Khanafer, K., Vafai, K. and Lightstone, M. (2003), Buoyancy-driven heat transfer enhancement in a two-dimensional enclosure utilizing nanouids, International Journal of Heat and Mass Transfer, Vol. 46, pp. 3639-53. Lee, F.H. and Tzong, R.Y. (1992), Articial pressure for pressure linked equation, International Journal of Heat and Mass Transfer, Vol. 35, pp. 2705-16. Lin, K.C. and Violi, A. (2010), Natural convection heat transfer of nanouids in a vertical cavity: effects of non-uniform particle diameter and temperature on thermal conductivity, International Journal of Heat and Fluid Flow, Vol. 31, pp. 236-45. Liu, G.R. (2002), Mesh Free Methods, Moving Beyond the Finite Element Method, CRC Press, Boca Raton, FL. Liu, G.R. and Gu, Y.T. (2005), An Introduction to Meshfree Methods and Their Programming, Springer, Dordrecht. Markatos, N.C. and Pericleous, K.A. (1984), Laminar and turbulent natural convection in an enclosed cavity, International Journal of Heat and Mass Transfer, Vol. 27, pp. 755-72. Murshed, S.M.S., Leong, K.C. and Yang, C. (2008), Thermophysical and electrokinetic properties of nanouids a critical review, Applied Thermal Engineering, Vol. 28, pp. 2109-25. Pruta, N., Roetzel, W. and Das, S.K. (2003), Natural convection of nanouids, Heat Mass Transfer, Vol. 39, pp. 775-84. Tiwari, R.K. and Das, M.K. (2007), Heat transfer augmentation in a two-sided lid-driven differentially heated square cavity utilizing nanouids, International Journal of Heat and Mass Transfer, Vol. 50, pp. 2002-18. Wang, X.Q. and Mujumdar, A.S. (2007), Heat transfer characteristics of nanouids: a review, International Journal of Thermal Sciences, Vol. 46, pp. 1-19. Wen, D. and Ding, Y. (2005), Formulation of nanouids for natural convective heat transfer applications, International Journal of Heat and Fluid Flow, Vol. 26, pp. 855-64.

Natural convection of nanouids ow 273

HFF 23,2

274

Xie, H., Wang, J., Xi, T., Liu, Y., Ai, F. and Wu, Q. (2002), Thermal conductivity enhancement of suspensions containing nanosized alumina particles, Journal of Applied Physics, Vol. 91, pp. 4568-72. Xu, J., Yu, B., Zou, M. and Xu, P. (2006), A new model for heat conduction of nanouids based on fractal distributions of nanoparticles, Journal of Physics D: Applied Physics, Vol. 39, pp. 4486-90. Xuan, Y. and Le, Q. (2000), Heat transfer enhancement of nanouids, International Journal for Heat and Fluid Flow, Vol. 21, pp. 58-64. Further reading Abu-Nada, E., Masoud, Z. and Hijazi, A. (2008), Natural convection heat transfer enhancement in horizontal concentric annuli using nanouids, International Communications in Heat and Mass Transfer, Vol. 35, pp. 657-65. Corresponding author Vassilios C. Loukopoulos can be contacted at: vxloukop@physics.upatras.gr

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com Or visit our web site for further details: www.emeraldinsight.com/reprints

Вам также может понравиться