You are on page 1of 113

THE NUMERICAL STUDY OF THE FLOW AND HEAT TRANSFER BETWEEN TWO HORIZONTAL CYLINDERS

HUO YUNLONG (B. Eng.) DEPARTMENT OF MECHANICAL ENGINEERING

A THESIS SUBMITTED FOR THE DEGREE OF MASTER OF ENGINEERING NATIONAL UNIVERSITY OF SINGAPORE
2004

ACKNOWLEDGEMENT

The author wishes to record here his indebtedness and gratitude to many who have contributed their time, knowledge and effort towards the fulfillment of this work. Particularly, he would like to express his heartfelt gratitude and thanks to Dr. T. S. Lee and Dr H. T. Low for their invaluable guidance, supervision, encouragement and patience throughout the course of the investigation. The technical staffs of the Fluid Mechanics Laboratory are also to be thanked for their assistance during the phase of the investigation. Special thanks are also due to my classmates in the Fluid Mechanics Laboratory who give me great help in plotting the figures and using the Tecplots. The author also wants to express his appreciation to his last grandfather whose spirit encourages him, supports him and assists him throughout this period. Finally, the author wished to express his gratitude to those who have directly or indirectly contributed to this investigation.

SUMMERY

The flow and convective heat transfer in concentric and eccentric horizontal annuli with isothermal wall conditions are studied numerically using two-dimensional finite-difference and finite-element models. The Stream-Function Vorticity and primitive variable formulations are applied to the finite different and finite element methods respectively. The structure mesh is obtained to simulate the buoyancy driven flow. Since the complex geometry configuration of the studied cases, the cylindrical and bipolar coordinates are introduced to solve problems of the finite difference method. The model is also designed by the Galerkin finite element method with Penalty Function Approach. The effects of various parameters such as the radius ratio of the annulus, the eccentricity of the annulus, the Rayleigh number and Reynolds number of the rotation of the inner cylinder are investigated at the Prandtl number of 0.701. Overall heat transfer results are obtained. For the case of concentric cylinders, the numerical results obtained are in good agreement with the similar results of other investigators. In the extension of the numerical work done here, rotating the inner cylinder and outer cylinder individually and both are also considered. For the eccentric cases, comparison with available experimental results with the present two-dimensional numerical model is good at relatively low Reynolds numbers in the range of 0-800. The effects of Prandtl number on the flow and heat transfer are also briefly studied in the present investigations. All of the kidney cells, the stagnant zone and the thermal plume are changed with different directions and velocities of the rotation, Rayleigh number, eccentric ratio and Prandtl number.

II

TABLE OF CONTENTS Page ACKNOWLEDGEMENT ABSTRACT NOMENCLATURE LIST OF FIGURES I II V VIII

CHAPTER 1 INTRODUCTION 1.1 Background 1.2 Literature survey 1.3 Flow description 1.4 Objectives and scope 1 2 9 12

CHAPTER 2 PROBLEM FORMULATION 2.1 Derivations of the governing equations 2.2 Coordinate system for finite difference method 2.3 The governing equation and boundary condition in finite element method 2.4 Investigated geometric and physical parameters 2.5 Boundary conditions in the finite difference method 14 18 19 22 24

CHAPTER 3 NUMERICAL METHODS 3.1 The finite-difference and finite-element approaches 3.2 The solution procedure 3.3 Finite difference methods for solving the equations 3.4 Finite element methods for solving the equations 27 28 30 37

III

CHAPTER 4 ANALYSIS OF RESULTS FOR THE FINITE ELEMENT METHOD 4.1 The effect of Rayleigh numbers 4.2 The effect of the radius ratio 4.3 The effect of eccentricity 41 42 43

4.4 The effect of rotating the inner and/or outer cylinders between concentric cases 45 4.5 The effect of rotating the inner and/or outer cylinders between eccentric cylinders 47

CHAPTER 5

ANALYSIS OF RESULTS FOR THE FINITE DIFFERENCE METHOD

5.1 The effects of Rayleigh numbers 5.2 The effects of radius ratio 5.3 The effect of the inner cylinder rotating 5.4 The effects of eccentricity 5.5 The effects of rotating inner cylinder in the eccentric annulus

51 52 53 55 57

CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS REFERENCES FIGURES

60 62 68

IV

NOMENCLATURE A C Area Constant in the transformation equations from Bipolar coordinate systems to Cartesian Coordinate System D E
e

Diameter Magnitude of the eccentricity vector, e , e = | e |


Eccentricity Vector, e = ( eh , ev )

er
e

Magnitude of the eccentricity ratio vector e , er = | e |


r r

Eccentricity ratio vector, e = e /L


r

Gravity acceleration, g=| g |

Gravity vector

h , h Metric or scale factors in the Bipolar coordinate system ig

Unit vector in the direction of the gravity vector, i g = g /g

k
e k eq

Thermal conductivity Overall equivalent thermal conductivity Local equivalent thermal conductivity mean clearance between the two cylinders, L= ro ri Nusselt number based on the diameter of the heated cylinder, Nu D = h D / k Pressure Prandtl number, Pr = / radius

k eql
L
Nu D

p, P Pr r

Ra D

Rayleigh

number

based

on

the

diameter

of

the

heated

cylinder,

Ra D =

gT ' D 3 gT ' D 3
ri i D

Ral

Reyleigh number based on the mean clearance L, Ra l =

Re D

Reynolds number based on the diameter of the heated cylinder, Re D =


ri i L

Rel t T

Reynolds number based on the mean clearance L, Rel = time temperature Velocity vector coordinate variables in the Cartesian coordinate system

x, y Greek

Thermal diffusivity Coefficient of thermal expansion Angle measured clockwise from the upward vertical through the center of the heat transfer

Total emissivity of a surface Vorticity

Coordinate variables in the Bipolar coordinate system Angular coordinate in the Bipolar coordinate system Kinematic viscosity Density Stefan-Boltzmanns constant

VI

Angular position of the gravity vector relative to the negative y-axis measured in the clockwise direction

Stream function Angular speed

Subscripts i stands for inner cylinder; also used as an indexing integer variable for the mesh points o r stands for outer cylinder stands for reference quantity; also used as an indexing integer variable for the mesh points w stands for wall

Superscripts k n number of the inner iteration number of the time step or global iteration

the prime symbol emphasizes the dimensional form of a variable as distinct from its non-dimensional usage

VII

LIST OF FIGURES

Table 3.1 Table 4.1

Shape Functions of Quadrilateral Elements (4-9) Parameters used in calculations

70 70

Fig. 1.1.1 Fig. 2.1.1 Fig. 3.1.1

Geometry of annular region and the gravity direction Mesh used for the numerical computation 2-D view of three kinds of shape function

68 68 69

Fig. 4.1.1 4.5.5

Figures for finite element method

Fig. 4.1.1 4.1.5 Fig. 4.2.1 4.2.2 Fig. 4.3.1 4.3.4 Fig. 4.4.1 4.4.4 Fig. 4.5.5 4.5.5

Flow and temperature fields for various Rayleigh numbers of air (71-73) Flow and temperature fields for various Radius ratio of air (74-75) Flow and temperature fields for various eccentric ratio of air (76-78) Flow and temperature fields for various rotations in concentric annulus (79-82) Flow and temperature fields for various rotations in eccentric annulus (83-87)

Fig. 5.0.1 5.5.6

Figures for finite difference method

Fig. 5.0.1

Fig. 5.1.1 5.1.2 Fig. 5.2.1 5.2.3 Fig. 5.3.1 5.3.2 Fig. 5.4.1 5.4.3 Fig. 5.5.1 5.5.6

Overall heat transfer coefficient versus Rayleigh number at the (88-88) radius ratio of 2.6, er = 0 . Flow and temperature fields for various Rayleigh numbers of air (89-90) Flow and temperature fields for various Radius ratio of air (91-93) Flow and temperature fields for various rotations in concentric annulus (94-95) Flow and temperature fields for various eccentric ratio of air (96-98) Flow and temperature fields for various rotations in eccentric annulus (99-104)

VIII

CHAPTER 1 INTRODUCTION

1.1 Background The flow and thermal fields in enclosed space have received much attention because of theoretical and wide engineering applications, such as thermal energy storage systems, cooling of electronic components, and transmission cables. Fig. 1.1.1 shows the Geometry of annular region and the gravity direction. Given the problem illustrated in Fig. 1.1.1, the inner circular is set to high temperature and the outer part lower temperature under the terrestrial conditions. Due to its simple geometry and welldefined boundary conditions, the system has been studied extensively by researchers such as Bishop et al (1968), Kuehn and Goldstein (1976), Farouk and Guceri (1981), Vafai and Desai (1993), and large number of literatures were published in the past few decades. For concentric and eccentric cases in a horizontal annulus between two circular cylinders, the basic and fundamental configuration, the flow and thermal fields have been well studied. Kuehn and Goldstein (1976) comprehensively studied the concentric case. The experimental and numerical studies of the eccentric case have also been conducted by Kuehn and Goldstein (1978), and Guj and Stella (1995). However, on the whole, little is known about effects of rotation of inner and/or outer cylinders on the flow and the heat transfer in eccentric annuli. There is much scope for further investigation in this aspect of the problem. The numerical method of both of finite element and finite partial difference methods are useful. In present work the author focuses on two parts. In the first part, based on the previous work, the primary aim is to extend knowledge into areas that have not been conducted and examined. The main methods are the finite element method and the finite difference approach. For the finite element method, original governing equations are used to produce the velocity field and

temperature distribution. For the finite difference method, the streamline plot is created through using Stream-Function Vorticity formulation. Compared with each other, both of the two methods are capable of solving the flow and the convective heat transfer in concentric and eccentric horizontal annuli with isothermal wall conditions. Compared with the results of Kuehn and Goldstein (1976), the computed results of the thesis also show that the study may be pushed further to enter the second part. The second part is the knowledge extension that has a little study by other investigators. The primary objectives of this part focus on natural convection in the eccentric horizontal annuli where both inner and outer cylinders are rotated. Both rotated inner and outer cylinders are seldom investigated by other researchers. Different kinds of rotation such as the same direction of the inner and outer cylinders, and the opposite direction of the inner and outer cylinders associated with the eccentric ratio will be discussed in the thesis. Particularly, when the Reynolds number is large, the unstable situation is produced.

1.2 Literature review Natural convection between horizontal concentric cylinders has been widely studied experimentally and numerically over the past three decades because of the importance of this subject in industries, such as transmission cable cooling systems, latent energy storage systems, nuclear reactor designs, etc. The problem was first investigated experimentally by Beckmann (1931) with air, hydrogen and carbon dioxide as the test fluids to obtain overall heat transfer coefficients. A large part of the experimental work was devoted to finding the overall heat transfer between the cylinders using the non-dimensional parameter defining the temperature difference between the cylinders. Particularly Kuehn and Goldstein (1976) investigated the problem experimentally and numerically. A Mach-Zehnder interferometer was used to

determine the temperature distribution and the local heat transfer coefficients in air and water. With water, they demonstrated that the flow remained steady even though the Rayleigh number was well over the critical value obtained experimentally with air, which suggests that the Prandtl number affected the transition characteristics. Unlike the case of natural convection in concentric annulus, similar experimental studies for eccentric annulus are few. The effect of vertical and horizontal eccentricities on the overall heat transfer coefficient was first experimentally investigated by Zagromov and Lyalikov (1966) using air as the working fluid. Using optical interferometry, Kuehn and Goldstein (1978) studied the local and overall heat transfer coefficients for both horizontal and vertical eccentricities of magnitude er up to about 2/3. They found that although the distribution of the local heat transfer coefficient was substantially altered by eccentricity, the overall heat transfer coefficient did not change by more than 10% from the concentric value at the same Rayleigh number. The effect of moving the inner cylinder downwards is to cause the overall heat transfer to increase while moving the inner cylinder upwards has the opposite effect. Yeo (1984) use the same method as Kuehn and Goldstein (1976, 1978) to verify the overall heat transfer coefficients predicted by the numerical model. His experimental results are in good agreement with the experimental results of Kuehn and Goldstein (1978) obtained using nitrogen as test fluid and fit the present numerical curve very well with deviations typically less than 5%. Lee (1991) performed the numerical experiments to study rotational effects on the mixed convection of low-Prandtl-number fluids enclosed between the annuli of concentric and eccentric horizontal cylinders. For the range of Prandtl numbers considered here, numerical experiments showed the mean Nusselt number increases with increasing Rayleigh number for both concentric and eccentric stationary inner cylinders. At a Prandtl number of order 1.0 with a fixed

Rayleigh number, when the inner cylinder is rotated, the mean Nusselt number decreases throughout the flow. Dennis and Sayavur (1998) analytically and experimentally investigated the flow in eccentric annuli of drilling fluids commonly used in oil industry. The expression for azimuthal velocity as a function of eccentricity ratio and rheological parameters of the fluid has been obtained based on the linear fluidity model. Velocity profiles were measured for a Newtonian glycerol/water mixture and a non-Newtonian oil field spacer fluid in eccentric annuli using the stroboscoptic flow visualization method. Because of the limitations of the analytical approach and encouraged by the availability of large computing machines, numerical methods were applied to solve the equations which govern the flow and heat transfer in the annulus. The first numerical solution was obtained by Crawford and Lemlich (1962) using a Gauss-Seidel iterative method. Abbot (1962) used a matrix inversion technique to obtain solutions for the case of narrow annuli. Powe et al. (1971) applied numerical method to determine the Rayleigh number for the onset instability in the flow at relatively low radius ratios and obtained reasonably good qualitative agreement with the earlier experimental results of Powe et al. (1969) on the delineation of the flow regimes. Their numerical results seem to indicate that the onset of multicellular flow at low radius ratios does not affect the overall heat transfer significantly. Charrier-Mojtabi et al. (1980) gave numerical solutions using the alternating direction implicit (ADI) method for three cases: a wide annulus (R=2.26) for Pr=0.7, a narrow annulus (R=1.2) for Pr=0.7 and a wide annulus (R=2.5) for Pr=0.02. On treating the problem numerically at high Rayleigh numbers, Jischke and Farshch (1980) divided the flow field of an annulus into five regions which include an inner boundary layer near the inner cylinder, an outer boundary layer near the outer cylinder, a vertical plume region

above the inner cylinder, a stagnant region below the inner cylinder and a core region surrounded by these four regions; they applied the boundary layer approximation to obtain the temperature distribution and heat transfer rates. A numerical parametric study was carried out by Kuehn and Goldstein (1980), in which the effects of the Prandtl number and the radius ratio on heat transfer coefficient were investigated. Tsui and Tremblay (1984) presented the results of mean Nusselt numbers for both transient and the steady natural convection. San Andres (1984) found the size of the separation eddy and the position of the points of separation and reattachment to be Reynolds number dependent in the numerical study of flow between eccentric cylinders. The separation point moves in the direction of rotation upon increasing the Reynolds number, in contradiction of the first-order inertial perturbation theory of Ballal and Rivlin (1976). The numerical methods employed in their study include Galerkins procedure with B-spline test function. Galpin and Raithby (1986) assessed the impact of the standard treatment of the T-V coupling and proposed an improved method. Newton-Raphson linearization was investigated as a means of accelerating the convergence rate of control volume-based predictions of natural convection flow. It is found that repeated solutions of the Newton-Raphson linear set converge monotonically for a much wider range of relaxation, and the maximum convergence rate can be significantly higher than that corresponding to the standard linear set. Lee and Yeo (1985) developed a numerical model to study the effects of rotation on the fluid motion and heat-transfer processes in the annular space between eccentric cylinders when the inner cylinder is heated and rotating. The overall equivalent thermal conductivity ( K eq ) is obtained for Rayleigh numbers Ra up to 10 6 with rotational Reynolds number Re variations of 0-1120. Investigation shows that, for Re up to the order of 102, the numerical model shows promising results when Ra is

increased. Numerical solutions for laminar, fully developed, forced convective heat transfer in eccentric annuli were presented by Manglik and Fang (1995). With an insulated outer surface, two types of thermal boundary conditions had been considered: constant wall temperature (T), and uniform axial heat flux with constant peripheral temperature (H1) on the inner surface of the annulus. Velocity and temperature profiles, and isothermal Re, Nui,j and Nui,H values for different eccentric annuli ( 0 * 0.6 ) with varying aspect ratios ( 0.25 r * 0.75 ) are presented in their paper. The

eccentricity is found to have strong influence on the flow and temperature fields. The flow trends to stagnate in the narrow section and has higher peak velocities in the wide section. The flow maldistribution is found to produce greater nonuniformity in the temperature field and degradation in the average heat transfer. Yoo (1998) numerically investigated dual steady solution in natural convection in an annulus between two horizontal concentric cylinders for a fluid of Prandtl number 0.7. It is found that when the Rayleigh number based on the gap width exceeds a certain critical value, dual steady two-dimensional (2-D) flows can be realized: one being the crescent-shaped eddy flow commonly observed and the other the flow consisting of two counterrotating eddies and their mirror images. The critical Rayleigh number decreases as the inverse relative gap width increases. Mohamed etc. (1998) numerically studied the effect of radiation on unsteady natural convection in a two-dimensional participating medium between two horizontal concentric and vertically eccentric cylinders by using a bicylindrical coordinates system, the stream function, and vorticity. Original results are obtained for three eccentricities, Rayleigh number equal to 104, 105 and a wide range of radiationconduction parameter. Shu and Yeo (2000) applied the global method of polynomialbased differential quadrature (PDQ) and Fourier expansion-based differential

quadrature (FDQ) to simulate the natural convection in an annulus between two arbitrarily eccentric cylinders. Their approach combined the high efficiency and accuracy of the differential quadrature (DQ) method with simple implementation of pressure single value condition. The result confirmed the finding by Guj and Stella (1995). Escudier et al. (2000) concerned a computational and experimental study of fully developed laminar flow of a Newtonian liquid through an eccentric annulus with combined bulk axial flow and inner cylinder rotation. Their results were reported for calculation of the flow field, wall shear stress distribution and friction factor for a range of values of eccentricity , radius ratio and Taylor number Ta. More recently, Lee et al. (2002) used GDQ method due to the eccentricity of the inner and outer cylinders studied the nett fluid circulation around the inner cylinder and the effects of rotation of the inner cylinder with a radius ratio of 2.6. Discrepancies among the results reported in the literature for narrow annuli are found by Rao et al. (1985), Fant et al. (1989), Cheddadi et al. (1992), Kim and Ro. (1994). Large differences are shown not only for the Ra values at which bifurcation occur but also in regard to a possible existence of hysteresis phenomena. For example, Kim and Ro (1994) and Fant et al. (1989) found a hysteresis numerically, whereas Rao et al. (1985) show only one type of multicellular flow. Cheddadi et al. (1992) presented two numerical solutions at the same Ra that depends on the initial conditions: the crescent base flow and a multicellular one. However, they failed to obtain multicellular flows experimentally. Rao et al. (1985) and Kim and Ro (1994) supported numerically the general trends presented by Powe et al. (1969); that is, the appearance of multicellular flow patterns in the upper part of narrow annuli. Furthermore, Rao et al. (1985) reported a transition of the steady upper cells to oscillatory motion at moderate Rayleigh numbers. Using a linear stability analysis of steady two-dimensional natural

convection of a fluid layer confined between differentially heated vertical plane walls, Korpela et al. (1973) reported that the flow is primarily unstable against purely hydrodynamic steady waves in the limit of zero Prandtl number. These secondary shear-driven instabilities are crossing cells called cats eyes. Increases in Prandtl number lead to the appearance of buoyancy-driven oscillatory instabilities. The critical value of Pr determining which type of instabilities appears has been numerically determined which type of instabilities appears has been numerically determined to be around Pr=12.7 by many authors. In slots of finite ratio of height over width the vertical temperature gradient is an additional results and linear stability analysis, Roux et al. (1980) have demonstrated the existence of a zone of limited extent in the (Ra, A)-plane inside which steady cats eyes can develop. This zone is only for aspect ratios larger than about A=11 for air-filled cavities. This result was confirmed by the numerical studies of Lauriat (1980), Lauriat and Desrayaud (1985), and more recently by Le Quere (1990) and Wakitani (1997). As Ra is further increased, a reverse transition from multicellular flow to unicellular flow occurs and this has been numerically and experimentally demonstrated by Roux et al. (1980), Lauriat (1980), Desrayaud (1987), and Chikhaoui et al. (1988). Cadiou et al (1998, 2000) studied numerically the flow structure which develops both in horizontal and vertical regions of narrow air-filled annuli and devoted some part of their paper to the thermal instabilities observed in the top of the annulus and clarify, found in the literature. Yoo (1998) numerically investigated natural convection in a narrow horizontal cylindrical annulus for fluids for Pr 0.3 . For Pr 0.2 , hydrodynamic instability induces steady or oscillatory flows consisting of multiple like-rotating cells. For Pr=0.3, thermal instability creates a counter-rotating cell on the top of annulus.

Until very recently, most numerical studies have been limited to flows in the steady laminar regime. Farouk and Guceri (1982) applied the k turbulence model to study the turbulent natural convection for high Rayleigh numbers ranging from
10 6 to 10 7 with a radius ratio of 2.6. A comparison of Nusselt numbers between the

results obtained numerically and those obtained experimentally by other investigators showed a good agreement. Kenjeres and Hanjalic (1995) studied natural convection in horizontal concentric and eccentric annuli with heated inner cylinder using several variants of single-point closure models at the eddy-diffusivity and algebraic-flux level. Their results showed that the application of the algebraic model for the turbulent heat flux derived from the differential transport equation and closed with the low-Reynolds number form of transport equations for the kinetic energy , its dissipation rate , and temperature variance 2 , reproduced well a range of Rayleigh numbers, for different overheatings and inner-to-outer diameter ratios.

1.3 Flow description From a theoretical point of view, natural convection in horizontal annuli has been one of the focuses of research heat transfer on account of the large variety of flow structures encountered in this configuration according to the value of the radius ratio. A comprehensive review of steady two-dimensional (2-D) convection was presented in the work of Kuehn and Goldstein (1976), in which experimental and numerical studies were performed to determine velocity and temperature distribution and local heat transfer coefficients for convective flows of air (Pr 0.7) and water (Pr 6) within a horizontal concentric annulus. In 1978 Kuehn and Goldstein investigated natural convection heat transfer in concentric and eccentric horizontal cylinders through experiments. And then parametric study of Prandtl number and diameter ratio effects

were done in horizontal cylindrical annuli. Powe et al. (1971) and Rao et al. (1985) investigated flow patterns for air. They found that free convective flow of air could be categorized into four basic types: a steady two-dimensional oscillatory flow, a threedimensional oscillatory flow, and a two-dimensional multicellular flow. Recently, Yoo (1998) investigated the existence of dual steady states for a fluid of Pr=0.7. The basic two-dimensional steady flow that is observed at low Rayleigh numbers is characterized either by two crescent-shaped or by two kidney-shaped cells according to the value of the radius ratio R. The first pattern is observed for narrow annuli whereas the latter is found only for large radius ratios (Bishop and Carley, 1966). These two patterns present symmetry with respect to the vertical centerline. The main difference between these two basic flow fields is in the shape of the central flow regions that become istorted into a kidney shape for the second flow structure. From their experimental work, Powe et al. (1971) depicted flow regime transitions for airfilled annuli and were the first to present a chart for the prediction of the nature of the flow according to the Rayleigh number and radius ratio. This chart shows the limit between the base flow and the two- or three-dimensional flow patterns, stationary or oscillatory, which follow the named pseudo-conduction regime. The effect of heating the inner cylinder is the fluid follows an upward stream along the hot inner cylinder and finally reaches the top of the annular space. The fluid goes then downwards along the cold cylinder and reaches the almost quiescent bottom portion of the annulus. At low Rayleigh number, conduction is the major mode of heat transfer between the hot and cold cylinders. As the Rayleigh number is increased, the center of rotation of the cells moves upwards and a thermal plume starts to form at the upper part of the annulus with an impingement region at the outer cylinder. The shape of isotherms shows that the

10

largest part of the heat convected within the annulus is extracted from the lower part of the inner cylinder. The buoyancy force is proportional to the temperature difference between surfaces. Therefore, at a higher temperature difference between the two cylinders, the strength or circulation of the convection cells is greater. The rate at which heat is being transferred or convected by faster moving fluid is therefore increased. The flow and temperature fields around the inner cylinder greatly resemble that of a heated cylinder convecting to still ambient air. The position of the inner cylinder relative to the outer cylinder is an important geometric parameter that deserves studies because it may either enhance or suppress the development of these flow cells and thus affects the rate of heat transfer. Such situations where the annular region becomes eccentric are encountered in actual practice; as in the snaking of high voltage underground power cable caused by thermal expansion of the cable. If eccentricity does affect the heat transfer in a significant manner, it could be employed as a design factor to either enhance or reduce the amount of heat transfer between the cylinders as the application may require. This aspect thus requires detailed investigation. The rotation of inner and/or outer cylinders will affect the flow in the annular region and thus the heat transfer. The flow generated in an annulus due to rotation of the inner cylinder in the absence of bulk axial flow is one of the most widely investigated topics in the fluid mechanics. Of the hundreds of papers published to date, the majority have been concerned with the Taylor vortices which arise above a critical Taylor number Ta c . Lockett (1992) showed that the occurrence of Taylor vortices is inhibited by eccentricity of the inner cylinder, his numerical calculations being confirmed by the recent experimental work of Escudier and Gouldson (1997) as well as by earlier

11

experiments reported by Kamal (1966), Cole (1968), Vohr (1968) and Castle and Mobbs (1968). The flow separation and the recirculating eddy or vortex which occurs above a critical eccentricity for a given radius has also received widespread attention (Kamal, 1966; Ballal and Rivlin, 1976; San Andres and Szeri, 1984; Siginer and Bakhtiyarov, 1998). In the present study, the effects of various system parameters such as the temperature difference between the surfaces of the two cylinders, the geometry of the annulus, the properties of the fluid and the rotation rate of the inner cylinder on the flow and the heat transfer in the annular spaces are investigated. Because of its common occurrence in practice, the inner cylinder is considered to be the hotter cylinder. As a useful idealization, it is further assumed that the two cylinders are kept isothermal. The author has also restricted his study mainly to flow in the laminar regime. However, the unstable and turbulence is introduced when the Rayleigh and Reynolds number is increased. Some cases are analyzed by using analytical method. Fig. 1.1.1 shows schematically a typical annular region being studied and the physical quantities involved.

1.4 Objectives and scope The natural convection phenomena in concentric and eccentric annuli are numerically studied. The physical behavior of the buoyancy driven flow is investigated through using the two-dimensional numerical model of the finite element and finite difference methods. The computational results can provide the important parameters for the industrial and engineering applications. In present study, both of the methods have advantage and disadvantage. It is difficult for the finite difference method to solve the vorticity-stream function formation with the moving boundary condition. Some special

12

approaches such as the single value condition are proposed to deal with the problem. Corresponding to the finite difference method, the finite element method enables to combine the boundary condition to the matrix and solve complex geometry more easily and produces more accurate results. However, the large sparse matrix created by the finite element method need more compute storage and executed time. Different kinds of the rotation associated with eccentric ratios and radius ratios are discussed, such as the same direction of the inner and outer cylinders rotation, the opposite direction of the inner and outer cylinders rotation. The overall heat transfer coefficients are investigated.

13

CHAPTER 2 PROBLEM FORMULATION

The flow and the heat transfer in the annular space between two horizontal circular cylinders with parallel axes is the main problem that is being studied. The cylinders are assumed to be isothermal with the inner cylinder being held at a higher temperature. The annular space may either be concentric and eccentric. Fig. 1.1.1 and 2.1.1 show the geometrical and mesh configuration of a typical problem. In the finite difference method, effects of such parameters such as radius ratio, eccentricity of the annular, the Rayleigh number, the Prandtl number and the rotation of the inner cylinder expressed in the form of a Reynolds number are of interest to the present investigation. In the finite element method, such primate parameters as the temperature, the radius of the cylinders, the pressure gradient, the density, the thermal diffusivity and the coefficients of thermal expansion are investigated by the author. The primate parameters in the finite element method are compared with the system parameters in the finite difference method. In this chapter, the governing equations are put in a form suitable for subsequent numerical studies. The underlying assumptions of the formulation are stated. The nondimensionalization of the governing equations helps to identify the forms of the relevant parameters in the finite difference method.

2.1 Derivations of the governing equations The fluid flow and the heat transfer in the annular region are governed by the equations of momentum, mass and energy conservation. These equations may be found in standard texts such as Eckert and Drake (1981) and Parker, Boggs and Blick (1969).

14

2.1.1 Simplifying governing equations for finite element method These governing equations in their original and complete form are highly complex. In formulating the actual equations used in this study, several simplifying assumptions are made:

The flow is assumed to be effectively invariant along the axial direction of the cylinders. This leads automatically to a two-dimensional model. The twodimensional approximation is a good representation of the real flow in a long finite length annulus away from the ends provided there are no threedimensional instabilities (Kuehn and Goldstein 1976).

The flow is assumed to be laminar. This is an essential assumption because unless some form of turbulence modeling is used, the governing equations in their usual form will break down when the flow becomes turbulent.

The Boussinesq approximations are adopted. With these simplifying assumptions above, the governing equations are:

Momentum Conservation Equation:


P U ' ' ' + (U )U ' = + (1 (T ' Tr' )) g + r 2 U ---------------------------------(2.1) ' t r Continuity Equation U = 0 -------------------------------------------------------------------------------------(2.2)
'

Energy Conservation Equation:

T ' ' + (U )T ' = r 2T ' -------------------------------------------------------------------(2.3) t ' (Because of the repeated use of some symbols in both dimensional and dimensionless forms, the prime symbol
'

is used to emphasize the dimensional form of the variables

as distinct from its dimensionless usage.)

15

2.1.2 Stream-Function Vorticity formulation for the finite different method The governing equations (2.1) to (2.3) are recast in the Stream-Function Vorticity form by taking the curl of equation (2.1) and defining two functions ' and

' called respectively the Stream-Function and the Vorticity by the following relations:
U ' = ' -------------------------------------------------------------------------------------(2.4)

' Ux =

' ' and U y = ' -------------------------------------------------------------------(2.5) y ' x

The pressure gradient term in equation (2.1) is eliminated because the curl of a gradient is identically zero. The use of ' , the Stream-Function, ensures that the continuity condition is automatically met. The resultant equations in the Stream-Function Vorticity formulation are: The Vorticity Transport equation:

' + ( ' U ' ) = (T ' To' ) g + r 2 ' --------------------------------------------(2.6) t '


The Stream-Function Vorticity equation:

2 ' = ' -------------------------------------------------------------------------------------(2.7)


which is the definition of Vorticity ' in terms of ' . The Energy conservation equation:

T ' + (T ' U ' ) = r 2T ' -------------------------------------------------------------------(2.8) t '


where the convective terms have been put in the conservative form using the mathematical identity ( f U ) = f ( U ) + U f and the continuity relation (2.2).

Non-dimensionalization The governing equations (6) to (8) are made dimensionless by setting

16

U L t ' r T ' To' x' ' ' L2 x = , t = 2 ,T = ' , U = , = , = L r r r L Ti To' Under this scheme of non-dimensionalization, the governing equation (2.6) to (2.8) assumes the following form: Vorticity Transport Equation:
+ ( U ) = Pr Ral T i + Pr 2 -------------------------------------------(2.9) g t

'

Stream-Function Vorticity Equation:


2 = -------------------------------------------------------------------------------------(2.10)

Energy Conservation Equation:


T + (TU ) = 2T -----------------------------------------------------------------------(2.11) t where Pr and Ral are the dimensionless parameters.

With the non-dimensionalization scheme, a problem is fully specified when the following parameters are known: (a) the radius ratio ro / ri , (b) the eccentricity ratio vector e or er and angle ,
r

(c) the Prandtl number Pr, (d) the Rayleigh number Ral, and (e) the Reynolds number Rel. (a) and (b) specify the geometry of the two-dimensional solution region. The dimensionless parameters Pr and Ral appear explicitly in the governing equations. The Reynolds number Rel, which is a measure of the wall velocity of the inner cylinder, is implemented through the vorticity boundary conditions.

17

2.2 Coordinate system for finite difference method For the eccentric cases in the finite difference method, the bipolar coordinate system, which is a more convenient coordinate system, is adopted. The Bipolar coordinate system gives a fine mesh around the inner cylinder. The proved to be essential for the accurate evaluation of temperature gradient in the heat flux calculation. However, the Bipolar coordinate system cant be used for the concentric cases owing to singularity at er = 0 . A Polar coordinate system is therefore employed for the concentric cases. 2.2.1 Concentric geometry The transformation between the Cartesian x-y coordinate system and the r- Polar coordinate system is given by the following relations:
x = r cos , y = r sin

(0 r < ), (0 2 ) ---------------------------------------(2.12)

The governing equations in the r coordinate system are:


U T cos T 1 + (U r ) + + (U ) + r = Pr 2 + Pr Ral (sin ) -----(2.13) t r r r r r 2 = -----------------------------------------------------------------------------------(2.14)
U T T 1 + (U r T ) + (U T ) + r = 2T ---------------------------------------------(2.15) t r r r

where U r =

2 1 2 1 1 + ,U = and 2 ( 2 + 2 ). 2 r r r r r r

2.2.2 Eccentric geometry The transformation between the Cartesian x-y coordinate system and the Bipolar coordinate system is given by the following relations:

x = c sinh /(cosh cos ) ( < < ) (0 2 ) ------------------------------(2.16) y = c sin /(cosh cos )

18

where c is a scaling factor of the transformation related to the eccentricity ratio er and the radius ratio of the two cylinders. This transformation is conformal and preserves the orthogonality of the grid lines. The Governing Equations in the bipolar coordinate system are:
1 1 (U ) + (U ) (U sin + U sinh ) = Pr 2 + + t h h c T T T T Pr Ral[cos ( A + B ) + sin ( B A )]

----------(2.17)

2 = ------------------------------------------------------------------------------------(2.18) T 1 1 T + (U T ) + (U T ) (U sin + U sinh ) = 2T --------------------(2.19) t h h c where U = 2 1 1 2 1 2 ,U = , 2 ( 2 + ) h h h 2

h = c /(cosh cos ), A = (1 cosh cos ) / c, B = sin sinh / c

is the angle which describes the relation between the eccentricity ratio vector e and
r

the gravity vector g .

2.3 The governing equation and boundary condition in finite element method 2.3.1 The governing equation in finite element method The governing equations for the thermal and fluid fields are solved using the Galerkin finite element method. Since details are well documented in many textbooks, only an outline is given here. In essence, the computational domain is first divided into small elements. Within each element, the dependent u ' , P ' and T ' are interpolated by shape functions of , and . (In section 2.3, , and are shape function for u ' , P ' and T ' ) u i ( x, y, t ) = T U i (t ) -------------------------------------------------------------------------(2.20)

19

P( x, y, t ) = T P(t ) ---------------------------------------------------------------------------(2.21) T ( x, y, t ) Tr = T T(t ) -----------------------------------------------------------------------(2.22)


Where U i (t ) , P(t) and T(t) are column vectors of elements nodal point unknowns. Substituting the above equations into the governing equations, we will obtain the residuals R1 , R2 and R3 which represent the momentum, mass conversion and energy equations respectively. The Galerkin form of the method of Weighted Residuals seeks to reduce these errors to zero, and the shape functions are chosen the same as the weighting functions. Following the procedure, the governing equations for the fluid flow and heat transfer may be rewritten as

( (i T + 2i T / r )dA) U i = p ( T dA)P --------------------------------------(2.23)


1

( p is the penalty parameter) ( C p T dA)


dT + ( C pu T dA)T dt -----------------------------------------------(2.24)

+ ( k T dA)T = 0
( T dA)
1 ^

dU i + ( u T dA) U i dt

( (i + 2i / r ) T dA)P + ( ( T + 2 2i T / r 2 )dA) U i
^ ^

-------------------------------------------(2.25)

+ ( (i )( j )dA) U +
T j

( Y i g )T = 0 (for non rotation)

or n i ds +
1
^

n i ds (for ratation case)

where i is unite vector of ith component, p is the penalty parameter. 2.3.2 The boundary condition (1) Temperature boundary condition:

20

The present study is concerned only with isothermal cylinders. The temperature T at the surfaces of the cylinders for both the concentric and eccentric geometries is the actual value which is applied to the boundary. For example: T |outerwall = 301 K T |innerwall = 304 K (2) The velocity boundary condition: Case 1: No rotation of the inner or outer cylinders
u i | = 0 --------------------------------------------------------------------------------------(2.26) Case 2: with rotation of the inner or outer or both cylinders
un = 0 ------------------------------------------------------------------------------(2.27)

2.3.3 Penalty Function Formulation The penalty function methods can be derived directly from the Stokes viscosity law (Fukumori and Wake, 1991);
2 p = p s ( + ) u -----------------------------------------------------------------------(2.28) 3

where p s denotes the thermodynamic or static component of the pressure, p is the mean pressure and is the second coefficient of viscosity. The basic idea of the penalty method consists in expressing the pressure through the pseudoconstitutive relation: p = p s p u -------------------------------------------------------------------------------(2.29) in which p is a large number. Equation (2.29) is then substituted into the momentum equations. ( (i T + 2i T / r )dA) U i = p ( T dA)P
1

and the continuity equation is no longer necessary.

21

2.4 Investigated geometric and physical parameters 2.4.1. Various parameters for the finite element method In the present study, finite element method is applied to solve the primitive variable form of the incompressible Navier-Stokes equations with Boussinesq approximation. Therefore, the basic parameters: 1) 2) 3) 4) 5)

Reference density Coefficient of thermal expansion Reference temperature Kinematic viscosity Thermal diffusivity

Tr'

r r

2.4.2. Various parameters for the finite difference method The vorticity-stream function formulation in the curvilinear coordinate system is taken as the governing equation. Therefore, the system parameters: 1) The radius ratio r0 / ri . r0 / ri = 5, r0 / ri = 2.6, r0 / ri = 1.25 are studied so that the effects of the radius changes can reflect the correct results. 2) The eccentricity ratio vector er . { er =

=(

eh ev , )} er = 1/3, 2/3 is considered. L L

For the eccentricity ratio vector er , L= r0 ri is the mean clearance of the

annulus. 3) The Rayleigh parameter Ral =

gT ' L3 .

For the Rayleigh number, the strength of the buoyancy force relative to the viscous force is disposed. From the formula it is seen that the Rayleigh number is proportional to T ' . 22

4) The Reynolds number Rel =

i ri L .

The rate of rotation of the inner cylinder is expressed as a Reynolds number based on the surface velocity of the inner cylinder and the mean clearance L. 5) The fluid properties Pr = / . The relevant fluid properties are and , the kinematical viscosity and the thermal diffusivity. Because of its common occurrence in practice, the inner cylinder is considered to be the hotter cylinder. And the two cylinders are kept isothermal. Different parameters values are studied in the present paper and the effects of various system parameters are investigated. At last, the heat transfer coefficients are produced compared with that of Kuehn and Goldstein (1976). Heat transfer coefficients One of the primary objectives of the present study is to investigate how heat transfer, both overall and local are affected by the variations of the above parameters. The rate of heat transfer is expressed in terms of equivalent thermal conductivities as defined below: (i) Overall equivalent thermal conductivity K eq (concentric geometry)

K eq =

heat energy converted per unit length of the annulus 2KT ' / In(ro / ri )

Where the denominator is the heat transfer by pure conduction of a motionless medium having the same thermal conductivity k as the fluid in a concentric annulus of radius ratio ro / ri . In this formula the eccentricity of the annulus geometry is ignored; only the radius ratio is taken into account. (ii)
e Overall equivalent thermal conductivity K eq (eccentric geometry)

23

e K eq =

heat energy converted per unit length of the annulus 2KT ' / In( S )

where S=

(ro + ri ) 2 e 2 + (ro ri ) 2 e 2 (ro + ri ) 2 e 2 (ro ri ) 2 e 2

Here the denominator is the heat transfer by pure conduction of a motionless medium having the same thermal conductivity as the fluid. The eccentricity of the annulus is taken into account in the reference conduction term. When the annulus is
e concentric, K eq is equal to K eq .

(iii)

Local equivalent thermal conductivity K eql

Where r is either ri or ro depending on the surface considered. The denominator is the heat flux per units area at the point on the surface if the heat transfer were by pure conduction through a motionless medium having the same thermal conductivity as the fluid in a concentric annulus of radius ratio ro / ri . The reference term of this formula, as in (i), does not take into account the annulus eccentricity. The heat transfer coefficient K eq was first defined by Beckmann (1931). The use of reference conduction terms, in the denominators of (i) and (iii), which ignore the eccentricity of the annulus, facilitates the comparison of heat transfer at different eccentricities.

2.5 Boundary conditions in the finite difference method 2.5.1 Vorticity boundary conditions From a given distribution of the Stream-Function , the vorticity boundary condition is evaluated directly from its definition:

24

wall = 2 | wall ----------------------------------------------------------------------------(2.30)


Expression (2.30) in generalized orthogonal curvilinear coordinates and using the nonslip flow condition at the wall of the cylinders, equation (2.30) becomes

wall =

1 2 1 h ( ) ----------------------------------------------------(2.31) + 2 2 h h h h

where is constant along the wall and grid lines of constant are perpendicular to the wall. h and h are the scale factors of the transformation. (1) Concentric cylinders For the concentric case, r and may be taken as and respectively so that
h = hr = 1 and h = h = r. Equation (2.31) then becomes

wall = (

2 U + ) | wall ------------------------------------------------------------------(2.32) r r 2

(2) Eccentric cylinders For the eccentric case, from Appendix A, h = h = c /(cosh cos ) = h in the Bipolar coordinate system. Equation (2.31) reduces to

wall =

1 2 | wall ------------------------------------------------------------------------(2.33) h 2 2

2.5.2 Stream-function boundary conditions The Stream Function value o on the outer cylinder may be arbitrarily set to zero. In the present study i is determined using the criterion that the pressure distribution in the solution region is a single-valued function. Similar criteria were used by Launder and Ying (1972) and Lewis (1979) for the numerical studies of isothermal flows in non-simply connected geometries. Mathematically, this criterion implies that the line integral of the pressure gradient P along any closed loop circumscribing the s

25

inner cylinder is zero i.e.

s ds = 0.

P can be evaluated from the momentum s

conservation equations. The numerical implementation of this criterion is described in the next chapter. The Stream-Function boundary conditions for both the concentric and eccentric geometries are of the Dirichlet type as follows:

| outerwall = 0 ----------------------------------------------------------------------------------(2.34) |innerwall = i --------------------------------------------------------------------------------(2.35)


2.5.3 Temperature boundary conditions The present study is concerned only with isothermal cylinders. The dimensionless temperature T at the surfaces of the cylinders for both the concentric and eccentric geometries is as follows:

T | outerwall = 0 T |innerwall = 1

26

CHAPTER 3 NUMERICAL SOLUTIONS

The finite difference method and finite element method were adopted to solve the governing equations. The finite difference method is a well-known method for solving the partial differential equations. A good introduction to this method may be found in Ames (1978). The practical applications of the method to problems in incompressible and compressible fluid dynamics are described in great details in the very important work of Roache (1972). It is difficult for the finite difference method to solve the vorticity-stream function formation with the moving boundary condition. Some special approaches such as the single value condition are proposed to deal with the problem. Therefore, in the present study finite element method is applied to solve the primitive variable form of the incompressible Navier-Stokes equations with Boussinesq approximation. The finite element method (FEM) is a numerical technique to obtain approximate solutions to a wide variety of engineering problems. Although originally developed to study the stresses in frame structures, it has since been extended and applied to the broad field of engineering. The basic idea about the finite element method can be found in Heinrich and Pepper (1999).

3.1 The finite-difference and finite-element approaches 3.1.1 The finite-difference method Separate computer programs were written for the concentric and the eccentric geometries. The system of mesh used in each case is the one natural to the particular coordinate system.

27

Second-order accurate finite difference approximations were used for the discretization of the governing equations whenever possible. The finite difference form of equations (2.13) to (2.15) for the concentric and equations (2.17) to (2.19) for the eccentric case were solved in a time-marching manner until satisfactory convergence was attained. At each time the Stream-Function Vorticity equations must be solved to convergence. Instead of second-order central differencing, upwind differencing was used for the convective terms to obtain good stability. 3.1.2 The finite-element method The basic idea of FEM is that a solution can be analytically modeled or approximated by replacing it with an assemblage of discrete elements. Since these elements can be put together in a variety of ways, they can be used to represent exceedingly complex shapes. The FE discretization procedures reduce continuum problems to one of a finite number of unknowns by dividing the solution region into elements and expressing the unknown field variable in terms of assumed approximating functions (or interpolation functions) within each element. The interpolation functions (or shape functions) are defined in terms of the values of the field variables at specified points called nodes (nodal points). The nodal values of the field variable and the interpolation functions for the elements completely define the behavior of the field variable within the elements. Once the nodal values are found, the interpolation functions define the field variable throughout the assemblage of elements.

3.2 The solution procedure 3.2.1 The finite difference method The solution process begins with the establishment of the necessary initial values for , and T at time t=0. Other necessary parameters or constants that are

28

repeatly used in the program are also computed. The governing equations are solved in a cyclic manner. At the beginning of any particular cycle the time is increased by t and the distribution of at the new time t=t+ t is found by solving the Vorticity Transport equation with boundary conditions obtained from the last known distribution of . With known at all the interior points, the Stream-Function Vorticity equation is solved in an iterative manner. The boundary value of on the outer cylinder is always zero. On the inner cylinder, i is found through an iterative procedure. From the latest distribution of , the velocity terms required in the convective terms of the Energy and the Vorticity Transport equations are calculated. The next step in the cycle is to solve the Energy equation for the temperature distribution T. Local and overall heat fluxes may be calculated from the temperature distribution. The last step is to check if the distributions and T have converged and the energy balance is satisfactory. The above cycle is repeated with increment in time t until the convergence criteria are met. 3.2.2 The finite element method There are in general four different routes leading to a FE formulation: direct approach, variational principle approach, weighted residual approach, and energy balance approach. The advantage of the direct approach is that an understanding of the techniques and essential concepts is gained without much mathematical manipulation. But it can be used only for relatively simple problems. The variational principle relies on the calculus of variations and involves extremizing a functional. Weighted residuals approach has an advantage because it becomes possible to extend FEM to the problems where no functional is available. The most applicable weighted residual approach is Galerkins method. Regardless of the approach used to find the element properties, the

29

solution of a continuum problem by the FEM always follows a step-by-step process described as follows. Mesh generation Pre - processesInitial and boundary condition specification Materials properties input
Internal element calculation Boundary element calculation Processes Assemblage Solution
Computation of derived variables Post - processes Plot of solutions

3.3 Finite difference methods for solving the equations 3.3.1 The detail methods The governing equations are of two types, the time-dependent non-linear parabolic type and the linear elliptic type. (1) Parabolic equation The time-dependent parabolic equations for the concentric and eccentric cases are solved using the Alternating Direction Implicit method (abbreviated as ADI). The ADI method, also commonly known as variable direction method, involves the splitting of the time step to obtain a multidimensional implicit method which required only the inversion of tri-diagonal matrices in the case of a rectangular solution region. In the two-dimensional case, which is the primary concern of the present study, this involves splitting the time step t into two halves. The advancement of the solution over t is accomplished in two steps by solving in the -direction and then rdirection. When second-order accurate central difference operators are used, the method is second-order accurate in space and time. The ADI time-splitting of the

30

Vorticity Transport equation for concentric case, equation and eccentric case, is given below:
n+ U n n 1 u u (U rn n ) + (U n 2 ) + ( r ) u = r r r ---------------(3.1) 1 2 2 n+ 1 n 1 T n cos T n n Pr( 2 + + 2 ) 2 ) + Pr Ral (sin + r r r r r 2 r

2(

n+

1 2

n ) / t +

And
n+ U n n +1 u 1 u u (U rn n +1 ) + (U n 2 ) + ( r ) = r r r ----------(3.2) 1 T n cos T n 2 n +1 1 n +1 1 2 n + 2 Pr( 2 + + 2 ) ) + Pr Ral (sin + r r r r r 2 r

2( n +1

n+

1 2

) / t +

2( n + 0.5 n ) 1 u 1 u n n n n (U n ) + (U n n + 0.5 ) (U sin + U sinh ) = + h h c t Pr 2 n + 0.5 2 n T n T n T n T n ( ) + Pr Ral[cos ( A ) + sin ( B )] A +B + h2 2 2 And

--(3.3)

n n 2( n +1 n+ 0.5 ) 1 u 1 u n n (U n +1 ) + (U n n+ 0.5 ) (U sin + U sinh ) = + h h c t


Pr 2 n + 0.5 2 n +1 T n T n T n T n ( ) + Pr Ral[cos ( A ) + sin ( B )] A +B + h2 2 2

(3.4)

Terms with superscript n are treated as known and taken on its last known values at time t n .
n+ 1 2

obtained by solving equation (3.1) and other terms with known

values at t n are substituted into (3.2). The unknown in equation (3.2) are the n+1 . These are solved for in the same manner as the
n+ 1 2

in equation (3.1). The ADI forms

of the other parabolic equations are similar to above equations, and are not shown. The superscript u expresses a special form of differencing called upwind differencing. (2) Elliptic equations

31

Equation of the concentric case is solved with the Strongly Implicit Procedure of Stone (abbreviated to SIP) introduced by H.L.Stone (1968). A series of ten acceleration parameters generated by the method given in was used. The number ten was arbitrarily selected. One of the important merits of the SIP is that the rate of convergence is not so sensitive to the choice of acceleration parameters. This means that suitable parameters can be more easily and reliably estimated from the coefficient matrix than say in the corresponding ADI method which requires accurate knowledge of the minimum eigenvalues of a certain coefficient matrix to obtain good convergence rate. For the eccentric configuration, the ADI method was favored because the eigenvalues are allowed to be computed theoretically. See Birkhoff et al (1962) section 9 on the Helmholtz Equation in a rectangle. The Wachspress parameters, which are the sequence of acceleration parameters used, can hence be obtained with great accuracy. 3.3.2 Boundary conditions This section describes the numerical implementation of the boundary conditions stated before. The boundary conditions are all of the Dirichlet type. (1). Vorticity Transport Equation First-order and second-order finite difference approximations were obtained using Taylor expansion out of the walls. The first-order formulae were found to yield slightly more accurate results. They are as follows: Concentric Case

w =

2( w+1 w + U w r ) U w + ---------------------------------------------------------(3.5) rw ( r ) 2

Eccentric Case

w =

2( w+1 w + U w hw ) ------------------------------------------------------------(3.6) 2 h w ( ) 2

32

where subscript w and w+1 refer to the values of the variables at the wall and one mesh point away from the wall (in the fluid) respectively. U w is the tangential velocity of the wall, which in the case of the outer cylinder wall is zero. The U w is determined from the Reynolds number Rey1. (2). Stream-Function Vorticity Equaiton

o = 0 i = S f
Assuming that a projected or guessed value of i called S is given, the following steps are used to find a Pmax (maximum pressure difference term) associated with S. This is done by first solving the equation 2 = with i = S and o = 0. From the solution , the pressure gradient terms P (concentric case) or

P are computed at all the interior points using the Momentum Balance equation (2.1) in the respective coordinate systems. By integrating P P or along all the closed

circumferential loops of constant r or respectively (from = = 0to = = 2 ), the pressure difference terms of P ' s at all the loops are determined. A P , among the
P ' s , which has maximum absolute value is selected as Pmax i.e. Pmax = P such

that | P | the absolute value of any P ' s . A series of such S called S1 , S 2 , S 3 , etc. are projected at each time step and
1 2 3 their associated Pmax called Pmax , Pmax , Pmax , are computed. The first value S1 in

each time step is projected from Stream Function values in 1 and in 2 at the previous two time-steps. Depending on whether S1 is considered too high or too low a projection

33

1 (this can be seen from the sign of Pmax ), S 2 is projected linearly from the relation

| S 2 S1 |=

1 | S1 in 1 | . 2

Subsequent values say

Sl

are obtained by linear

p interpolation between one S p and one S q (p,q<l) whereby S p has negative Pmax and q S q has positive Pmax ; both being of minimum absolute values in their respective sign

categories. In this manner a S f with almost minimum absolute | Pmax | is obtained after a sufficient number of projections are made and a convergence criterion applying to consecutive value of the projection is satisfied. S f is then the value of i at time step t n . It would be very time-consuming if the Poisson equation 2 = is solved for every projected value of i at each time step. Fortunately, by the linearity of the equation, we need to solve 2 = only once at each time step. Using linear superposition, if ' is the solution to 2 = with i = S1 , the solution to
2 = with i = S 2 can be obtained by adding ( S 2 S1 ) " to ' where " is the

solution to the Laplace equation 2 = 0 with i = 1.0 . Solution to the last equation are trivial in both coordinate system. (3). Energy Equation The two cylinders are maintained at isothermal condition, that is
Ti = 1.0 and To = 0.0

(4). Progressive built-up of boundary conditions Solution can become unstable if the full value of a boundary condition is suddenly imposed. To give a more gradual and stable start, temperature and velocity boundary conditions are imposed over a number of time steps. The number of steps is

34

roughly proportional to the value of the boundary condition. The final solution is independent of the number of steps used to introduce the boundary condition. 3.3.3 Convergence criteria (1). Convergence of the inner iterations Concentric case For the concentric case where the SIP is used, a Residue term Re sij is defined at each interior point as follows:
k n 2 Re s ij = ij + D ij k ------------------------------------------------------------------------(3.7)

where

ij stands for the second-order finite difference operator of the Laplacian, n is

k the global iteration number or time step and k is the inner iteration number. Re sij is a

direct measure of the lack of convergence of k . The Stream Function distribution k is considered to have converged sufficiently when
k n Re s ij < 0.0005[ | rs |] / N ----------------------------------------------------------------(3.8) rs

for all interior points (i,j). N is the total number of interior points. In words, this means that the residue at each interior point must be less than 0.0005 times the average absolute vorticity value. Eccentricity Case Because the Stream Function frequently attains values close to zero, the criterion that
k k k | ij+1 ij | / | ij | ------------------------------------------------------------------------(3.9)

(where is a small real number) holds at all interior points (i,j) is not practical. An alternative criterion is used which involves first finding kpq , the maximum absolute

35

difference between successive iterates k and k +1 , and the grid point (p,q) where is occurs:
k k kpq = max | ij +1 ij | --------------------------------------------------------------------(3.10)
(i , j )

The criterion for the convergence of the inner iterations is that


k kpq /([ | rs |] / N ) < 0.0001 --------------------------------------------------------------(3.11) rs

i.e. the maximum deviation between successive iterates is less than 0.0001 times the average absolute Stream Function value. This criterion is satisfactory as long as is a reasonably well-behaved function which does not have highly localized peak values. (2). Overall convergence A run is deemed to have converged sufficiently when the following three criteria are met: a) b)
n npq /([ | rs |] / N ) < 0.001 rs

| Tijn+1 Tijn | / Tijn < 0.001

at all interior points (i,j). c) Qo / Qi 0.985 where Q stands for heat flux through the cylinders. The convergence of the Vorticity distribution is not included as a convergence criterion because it is directly related to the Stream Function through its definition. The convergence of the Stream Function implies the convergence of the vorticity distribution through the criterion may be different. The degree of heat balance is included as a convergence criterion because it is a fairly good barometer of the accuracy of the solution and the sufficiency of the mesh size. It was found that the temperature solution usually converged more slowly than the

36

Stream Function. The exceptions were runs at high Reyleigh numbers where the lack of convergence of the Stream Function indicated the presence or onset of flow instability. It is not always possible to satisfy the heat balance criterion though qualitatively it is clear the solution has already converged to a far greater extent than required by both criteria (a) and (b). This happens predominantly in cases where the inner cylinder is moved eccentrically downwards. In these cases, the mesh at the top of the annulus is very coarse, while at the same time the temperature gradient at the top of the outer cylinder is very high because of the impinging thermal plume. For such cases, even interpolation is not sufficient to resolve the extremely high gradient encountered.

3.4 Finite element methods for solving the equations 3.4.1 The details solving procedures Once the form of shape functions of , and is specified, the integrals defined in the equations (2.23) to (2.25) can be expressed by the matrix equations. The momentum and energy equations may be combined into a single global matrix equation.
M 0
1 0 U A(U) + K + EM -p1E T + p NT T 0

D T ( U) + L T ------------------------------(3.12) B

U F = T 0 Note that in constructing the above element matrix equation, the penalty formulation has been applied, and P in the momentum equation is substituted by coefficient matrices in the above equation are defined by: 1 EM -p1E T . The p

M P = T dA

M = T dA

37

L T = k T dA

D T ( U) = C pu T dA

^ ^

K ij = ( ( T + 2 2i T / r 2 )dA) ij + (i )( j T )dA

N T = C p T dA

E i = (i + 2i / r ) T dA

A(U) = u T dA

B = T i g

F = nds

where U is a global vector containing all nodal values of u and v. The assembled global matrix equations are stored in the skyline form and solved using the Gaussian elimination method. The successive substitution method is applied for nonlinear iteration and the time derivatives are approximated using the implicit finite difference scheme. 3.4.2 The shape function The element can be chosen from any of the four node, eight node and nine node elements (fig. 3.1.1). The shape function of quadrilateral elements (4-9) nodes may be described in Table 3.1.

38

CHAPTER 4 ANALYSIS OF RESULTS FOR THE FINITE ELEMENT METHOD

The numerical results from the finite element method will be analyzed and discussed in this chapter. The numerical model from the finite element method produces the computational results of the velocity and temperature fields, which show the detailed flow fields and temperature distributions. The computational algorithm developed above is capable of predicting the temperature distribution, the streamline distribution and internal convection in the horizontal annular between two circular concentric or eccentric cylinders. A set of computed results is shown for air, whose physical properties are given in Table 4.1. The mesh independence testing procedure for the computation is considered here and the final mesh used for the computation is determined so that any further refinement of the mesh produces an error smaller than 0.1%. Finally, 221 nine-node elements are adopted to calculate the thermal and fluid flow for half mode, which is used to simulate the natural convection without both inner and outer cylinders rotation, and 578 ninenode elements for whole mode with the cylinders rotating. The penalty formulation is used for the pressure field computation. A convergence criterion of 110-3 is set for relative error associated with unknowns for temperature and velocity fields. The effects of Rayleigh number, radius ratio and eccentric ratio on the flow and heat transfer are separately discussed in sections 4.1, 4.2 and 4.3. Section 4.4 and 4.5 emphasized the effect of the rotation. Section 4.4 focuses on the effect of rotating the inner, outer and both cylinders between the concentric cases and section 4.5 is set for the eccentric cases. The air is selected as the working fluid, which is used for the numerical studies of the present investigation. The present finite element method is

39

validated by comparing its numerical results with the available data in the literature and very good agreement has been achieved.

40

4.1 The effect of Rayleigh numbers The effect of different Rayleigh number is shown in Fig 4.1.1 4.1.5. Half model is adopted given the symmetry of the concentric annuli. In Fig. 4.1.1, a symmetric pair of counter-rotating kidney is set in the annuli. However, the temperature distribution (see Fig. 4.1.1(a)) seems like that of the natural convection from a heated circular cylinder. The temperature contour suggests that thermal transport in the annuli is primarily due to conduction. Temperature inversion appears to occur when the Rayleigh number is above 104. Fig. 4.1.2-4.1.5 show that the inversion is gradually becoming greater with increasing Rayleigh number. There are two approaches to increase Rayleigh number, enlarging the radius of cylinders or temperature difference between inner and outer cylinders. Fig. 4.1.2 and 4.1.3 separately depict the temperature, streamline and velocity at the Rayleigh number of 9104. Compared with each other, little difference is found so that Rayleigh number is the major parameter in the low Rayleigh number. The section also discussed when the unstable flow set in as a form of twodimensional oscillatory flow about the longitudinal axis. The numerical solution is reasonable in Rayleigh number of 3.09105. However, the streamline plots and the temperature contours yield oscillation, therefore, there is some hint that the unstable flow appears. When the Rayleigh number is 3.85105, one zero streamline is yielded in the old stagnant area. It may be deduced that for Rayleigh numbers about 1.1105, the time-averaged flow behavior and heat transfer of the actual physical flow between the two cylinders are not very different from those of a hypothetical two-dimensional flow without the oscillatory motion, and this lends credence to the possibility that the numerical method may yet yield a reasonable solution. From Fig. 4.1.4 and Fig. 4.1.5,

41

the highly distorted form of the isotherms and streamlines manifests the unstable nature of the flow at so high Rayleigh numbers. The experimental investigation of Powe et al. (1969) described that for radius ratios greater than about 1.8, instability at high Rayleigh number initially sets in as a form of two-dimensional oscillatory flow about the longitudinal axis. The flow in a plane perpendicular to the longitudinal axis exhibits the same kidney-shaped pattern and the patterns in adjacent planes oscillates out of phase with each other resulting in an apparent wave motion along the axis of the annulus. At the Rayleigh number of
3 10 5 , these oscillations were still only intermittent and presumably of small

amplitude. Above the Rayleigh number of about 1.5 10 6 there was no definite frequency in the oscillation and the plume became turbulent at the Rayleigh number of about 2 10 7 . In present the numerical results show that unstable phenomena began to be yielded between Rayleigh number 3.09105 and 3.85105. One circular flow at the stagnation area of the low Rayleigh number is shown from Fig. 4.1.5 (c). The present numerical results agree well with experimental results obtained by Powe et al.

4.2 The effect of the radius ratio This section focuses on the effect of the radius ratio. This part only emphasizes the effect of the radius ratio of 2.6 and 5.0 from the finite element method in the high Rayleigh number. The computed results for radius ratio of 2.6 and 5.0 are shown in Fig. 4.2.1 4.2.2. The radius ratio has a significant effect on the flow and temperature distribution between horizontal cylinders. Fig. 4.2.1 depicted temperature and flow fields of radius ratio 2.6 and 5.0 cases at the Rayleigh number of 1.5105. From Fig. 4.2.1 (b) and (e),

42

the center of kidney cell moved down and the stagnant zone becomes larger when the radius ratio is enlarged from 2.6 to 5.0 at the same Rayleigh number. It is easy for the higher radius ratios to yield temperature inversion. In comparison of computed results between Figure 4.2.1 (a) and (d) it is apparent that the higher radius ratio has a relatively greater temperature inversion. The stagnant temperature zone is increased further for the larger radius ratios of 5.0. The major reason is the boundary layer flow at high radius ratios has a higher momentum which enables it easily to distort the isotherms to create the effect of temperature inversion. The main conclusion is that the convective heat transfer is favored at higher radius ratio. The convective flow around the inner cylinder increases rapidly as the radius ratio is increased. The instability of the two-dimensional oscillatory type sets in at higher Rayleigh number for higher radius ratio. From Figure 4.2.2 (b), the strong oscillation appears between the two horizontal cylinders, however, the temperature distribution has no great change. Therefore, it may be deduced that for Rayleigh numbers between 5105 and 1106, the heat transfer of the actual convective flow between the two cylinders are not different from those of a hypothetical two-dimensional flow without the oscillatory motion, and this lends credence to the possibility that the finite element method may yet yield a reasonable solution for the radius ratio 5.0.

4.3 The effect of eccentricity The effects of eccentricity between the two horizontal cylinders were numerically studied for eccentric ratio 0.75 and 0.5. In this case of a natural convection in annulus between eccentric cylinders, the cylinders intracenter line is defined as the line to connect the centers of the outer cylinder and the inner cylinder as shown in Fig.

43

1.1.1. For the case of eccentric ratio 0.75, Fig. 4.3.1 and 4.3.2 show that the temperature and fluid flow between eccentric cylinders, which intracenter line is parallel to the gravity vector. This case is very similar to the concentric case in such a way that the flow is symmetric with respect to the intracentral line, and only half of the field can be considered as the computational domain. The computational study of the problem when the intracentral line is inclined with respect to the gravity vector is relatively complicated so that Fig. 4.3.3 depicted two kinds of situation whose intracentral line has degree 45o and 450 against the horizontal line. The immediate effect of moving the inner cylinder upwards (Fig. 4.3.2) is to enlarge the size of the stagnant zone and reduce the size of the of the kidney flow cell. At high eccentric ratio 0.75, Fig. 4.3.2(b) depicted that one circular flow cell appeared to occur at the bottom of the annuli and the kidney cell is squashed and connected to the circular flow. In Fig. 4.3.2(a), the thermal plume is squashed to create a pair of symmetrical humps. This indicates that the thermal boundary layer on the inner cylinder has begun to separate from the cylinder before reaching the top of the annulus. There is no district thermal plume as in other cases. A reduction of convective activities can be expected. Moving the inner cylinder downwards (Figure 4.3.1) has the quite opposite effect. The stagnant zone is reduced and convection is increased. In Fig. 4.3.3, the Rayleigh number is about 4.95104 and the eccentricity ratio is 0.5 with 6 k temperature difference between inner and outer cylinders. Figure 4.3.3 (ac) shows temperature, streamline and velocity plots for moving the inner cylinder to angular 450 against the horizontal line, which has characteristics like that of moving the inner cylinders both right and up. Fig. 4.3.4 (1) and (3) depict the temperature

distribution along the shortest and longest lines between the inner and outer cylinders for the case moving the inner cylinder to angular 450 in the anti-clockwise direction.

44

Compared with the concentric case, the thermal plume is a little shifted to the left side and reduced; the kidney shaped flow cell on the left side is enlarged and that on the left side is squeezed to very small size; A large area where the temperature gradient is very small appears in the left down side, which can be viewed in Fig. 4.3.3 (a) and found in Fig. 4.3.4 (3). Corresponding to the small temperature gradient close to the center of the left kidney cell, a small fluid flow is shown in the velocity plot in Fig. 4.3.3 (c). Fig. 4.3.4 (1) describes the temperature distribution in the shortest distance line between inner and outer cylinders, along which a sharp temperature gradient exists. On the other hand, moving the inner cylinder to angular 450 produced a larger thermal plume and smaller stagnation zone compared to that to angular 450. Fig. 4.3.4 (4) and (2) also shows separately a flat temperature distribution and a large temperature gradient along the longest distance line and the shortest distance line between inner and outer cylinders.

4.4 The effect of rotating the inner and/or outer cylinders between concentric cases This section focuses on investigation of the effects of the rotation between horizontal concentric cylinders, which include inner cylinder rotating, outer cylinder rotating and both of them rotating in the same and opposite direction. The Rayleigh number is kept 4.95104 and the eccentricity ratio is 0.5 with the inner and outer cylinders temperature being separately 307 K and 301 K. The tangential velocity is set to 0.1 m/s for only one cylinder rotating and 0.05 m/s for both of them rotating. In this section, the numerical results and the physics of the flow as revealed by the numerically obtained streamline and isothermal plots and temperature distribution along the selected lines are discussed.

45

The numerical results in Fig. 4.4.1, show pictorially the flow and temperature fields when the inner cylinder is rotated (a-b) and the outer cylinder is rotated (c-d). The rotation considered is both in the clockwise direction. The immediate effect of rotating the inner cylinder is to cause the fluid adjacent to the surface of the rotating cylinder to move with it by the virtue of viscous drag as clearly illustrated in Fig. 4.4.1 (b) and (d). For the non-rotating concentric case, the hot fluid adjacent to the inner cylinder rise upwards because of its reduced density and then it is cooled at the top of the annulus so that the cooled fluid falls sideways along the outer cylinder wall. When the inner cylinder is rotated in the clockwise direction, the left-hand hot-fluid flows velocity has the same velocity direction as the rotating cylinder. Therefore, the lefthand cell is elongated in size while the right-hand cell is reduced for the inner rotating. The circulation, defined as the integral

u ds , of the right-hand is thus reduced while

that in the left-hand is increased. On the other hand, rotating the outer cylinder in the clockwise induces the right-hand cooled fluid with the same velocity direction increasing so that it has a large right-hand cell and small left-hand cell. The isotherm plots Fig. 4.4.1 (a) and (c) show that the thermal plume and the stagnant zone are shifted in the direction of the rotating. Corresponding to the non-rotation case, the thermal plume with the inner rotating is driven to an angular position of = 800 (clockwise) from the highest point of the inner cylinder while the lower temperature stagnant zone is only changed a little; the outer cylinder rotating caused the stagnant zone approaching to = 900 (clockwise) from the lowest point of the outer cylinder with slightly shifted thermal plume, which is clearly depicted in Fig. 4.4.1 (a) and (c). The temperature distribution along the line with different angle against the horizontal line from the inner cylinder to the outer cylinder is shown in Figure 4.4.2. The higher temperature along line 1 ( = 00) compared with the lower line 2 ( = 900) and the

46

nearly same temperature of line 3 and 4 numerically reveals the large thermal plume shifting and small lower temperature stagnant zone shifting for inner cylinder rotating. On the other hand, line 5, 6, 7 and 8 described the outer cylinder rotating case. The effects of rotating both the inner and outer cylinders with the same and opposite directions on the flow and heat transfer for the concentric configuration is pictorially and numerically shown in Fig. 4.4.3 and 4.4.4. Fig. 4.4.3 (a-c) illustrate the temperature, the streamline and the velocity with the same rotating direction, which temperature distribution seems like that of only inner cylinder rotating except that the low temperature stagnant zone move further away from the bottom part. The hot fluid flow on the left-hand cell impinge into the right side with the help of the rotating inner cylinder; the cooled fluid flow on the right-hand cell spreads around most area of the outer cylinder (see Fig. 4.4.3 (b)). Lines 1-4 in Fig. 4.4.4 show the temperature distribution along lines shooting from the inner to the outer cylinder with angular position = 00, 900, 1800 and 2700 against the horizontal line. Various parameters plot for the inner and outer cylinder rotating in the opposite direction is shown in Fig. 4.4.3 (d-f). The root of the thermal plume dives into the right-hand part because of the drag force from the rotating inner cylinder while the top of the thermal plume driven by the viscosity of the rotating outer cylinder is shifted to the left-hand side. The low temperature stagnant zone is moved slightly up to the right-hand. Fig. 4.4.3 (e) shows that the left-hand cell is largely elongated that it will engulf the right cell because the viscosity of both the inner and outer cylinders enlarges the left-hand cell.

4.5 The effect of rotating the inner and/or outer cylinders between eccentric cylinders This section numerically simulates the natural convection in the annulus between the vertically eccentric cylinders (the cylinders intracenter line is parallel to

47

the gravity vector) with both inner and outer cylinders rotating. The Rayleigh number is kept at 4.95104 and the inner and outer cylinders temperature is respectively at 307 K and 301 K. The tangential velocity is set to 0.05 m/s for both of them rotating. Several cases with different eccentric ratios (0.5 and 0.7) are discussed in the section. The close proximity of the outer cylinder at the narrow throat exerts a great influence on the fluid flow and temperature distribution in the annulus. In the case of moving the inner cylinder up at the eccentric ratio of 0.5, the narrowness of the throat region makes it difficult for the thermal plume to go through it. Fig. 4.5.1 (a) and (d) respectively show the temperature of both cylinders rotating with the same and opposite directions. Since the narrowness of the throat region reduces the thermal plume, only the low temperature stagnant zone is moved left or right with the outer cylinder rotating. Fig. 4.5.2 shows that the temperature along the shortest line and longest shooting from the inner and outer cylinder with the same cylinders rotation direction is similar to that with opposite direction. Therefore, the right-hand cell evaded and occupied the stagnant zone when both cylinders rotate in the same direction and shrank and left the stagnant zone when both cylinders rotate in the opposite direction. It is interesting to note that the flow in the right-hand cell (see Fig. 4.5.1 (b)) is cut off from the outer cylinder when the inner cylinder has different velocity as the outer cylinder and the left-hand cell (see Fig. 4.5.1 (e)) is cut off from the inner cylinder when both cylinders rotate in the same direction. Similar temperature and fluid flow fields at the eccentricity ratio of 0.75 are given in Fig. 4.5.3. The velocity of both inner and outer cylinder is 0.1 m/s. The narrow throat plays a further important role on both the flow and temperature in the annulus. From Fig. 4.5.3, the rotation of the outer cylinder exerts a great influence. Compared with the results at the eccentricity ratio of 0.5, the faster rotated outer cylinder (0.1m/s)

48

associated with the further narrow throat decided the flow pattern. The inner cylinder is set to rotate in the clockwise direction. If the outer cylinder rotates in the clockwise, the right kidney cell will occupy the whole computed space except the small area close the narrow throat. On the other hand, the anti-clockwise rotation of the outer cylinder produces the opposite activities. Corresponding to the flow pattern, the low temperature stagnant zone follows the rotation direction of the outer cylinder, which is shown in Fig. 4.5.3 (a) and (d). In the case of rotating the inner cylinder down at the eccentric ratio of 0.5, moving the inner cylinder vertically down without rotation produces an increase in the size of the rotating cells accompanied by a reduction in the size of stagnation zone. The thermal boundary layer on the outer cylinder is increased in length from that of the concentric case. When the rotation is applied to both cylinders, the out layer of the thermal plume is moved with the outer cylinder and the root of the thermal plume is shifted by the inner cylinder. The narrow throat is moved downwards, which reduces the effect of the outer cylinders rotation and increases the influence of the inner cylinders rotation because it is difficult for the particles of the cooled fluid driven by the outer cylinder to pass the smaller stagnant part. The temperature and fluid flow fields of both cylinders rotating in the same and opposite direction are respectively presented in Figure 4.5.4 (a-c) and (e-f). Corresponding to Fig. 4.5.2, Fig. 4.5.5 numerically shows the temperature distribution along the shortest line and longest line shooting from the inner cylinder to the outer cylinder.

49

CHAPTER 5 ANALYSIS OF RESULTS FOR THE FINITE DIFFERENCE METHOD

The chapter solves the numerical model through using finite difference method. The numerical model from the finite difference method yields details of the flow and the temperature fields through the streamline and isotherm plots and then computes the local heat transfer values to predict the overall heat transfer coefficients. These plots help to show the physics of the flow and the heat transfer behavior. All the isotherm and streamline plots in this chapter are drawn by the software Tecplot 8.0, so that one grid line can not be deleted because of the mesh divisions. The overall heat transfer coefficients yielded by the numerical model at the radius ratio of 2.6 are compared with similar results of other investigators. The good agreement is a confirmation of the validity of the numerical model (see Fig. 5.0.1). The effects of Rayleigh number, radius ratio and Reynolds number on the flow and heat transfer between concentric cylinders are discussed in Sections 5.1, 5.2 and 5.3 respectively. Section 5.4 focuses on the heat transfer and fluid flow between eccentric cylinders without inner cylinder rotating and Section 5.5 for eccentric cylinders with inner cylinder rotating. In these sections, the discussions of the heat transfer results are preceded by an analysis of the physics of flow as revealed by the streamline and the isotherm plots. The working fluid with a Prandtl number of 0.701 (air) is assumed for the numerical studies of the present investigation.

50

5.1 The effects of Rayleigh numbers At low Rayleigh numbers, the rate of convection of heat energy from the inner cylinder to the outer cylinder is low because the buoyancy force proportional to the temperature difference is very small and the induced flow velocity is low. At
Ral = 2 10 2 (Fig. 5.1.1 (a)), the isotherm plots are almost concentric circles like those

of pure conduction since the convective flow velocity is small. However, the counterrotating pair of kidney-shaped vortices (Fig. 5.1.2 (a)) is already clear. The overall heat transfer coefficient for Rayleigh numbers up to just a few thousands is close to 1.0, indicating that the heat transfer is predominantly conductive. Though the convective flow exits at low Rayleigh numbers, the overall heat transfer is dominated by the conduction. The temperature in the annuli falls monotonically, shooting outwards. With the Rayleigh number increasing, the increased flow velocity on the surfaces of both cylinders caused by the increased buoyancy results in the distortion of the hitherto circle-like isotherms. At Ral = 10 4 (Fig. 5.1.1 (c)), the isotherm plots have lost its circular forms. As the Rayleigh number is increased to be strong enough, a situation of temperature inversion occurs in which cooler fluid is separated from the cooling surface by a layer of warmer fluid. The extent of inversion is greater at higher Rayleigh numbers. True boundary layer flow occurs for Rayleigh numbers above 3 10 4 . As temperature inversion develops with increasing Rayleigh number, the shape of a thermal plume begins to form at the top of the inner cylinder. By 5 10 4 , a welldefined thermal plume can be seen clearly (Fig. 5.1.1 (d)). The thermal plume becomes narrower as the Rayleigh number is increased. The streamline plots of Fig. 5.1.2 reveals that the kidney-shaped cells become progressively elongated and the center of rotation move higher up though further increase in the Rayleigh number merely further distorting the isotherms. At the Rayleigh number of 10 5 , the centers are about 30 0 51

from the upward vertical. The thermal plume becomes narrower as the Rayleigh number is increased. A pair of very weak counter-rotating cells is formed at the stagnation zone of the bottom of the annulus. The effect is manifested by two additional zero-valued streamlines. These cells are formed as more fluid is being drawn from the sides of the annulus at high Rayleigh numbers. The experimental investigation of Powe et al. (1969) manifested that for radius ratios greater than about 1.8, instability at high Rayleigh number initially sets in as a form of two-dimensional oscillatory flow about the longitudinal axis. The flow in a plane perpendicular to the longitudinal axis exhibits the same kidney-shaped pattern and the patterns in adjacent planes oscillates out of phase with each other resulting in an apparent wave motion along the axis of the annulus. At the Rayleigh number of 3 10 5 , these oscillations were still only intermittent and presumably of small amplitude. Above the Rayleigh number of about 1.5 10 6 there was no definite frequency in the oscillation and the plume became turbulent at the Rayleigh number of about 2 10 7 . It may be deduced that for Rayleigh numbers between 10 5 and 10 6 , the time-averaged flow behavior and heat transfer of the actual physical flow between the two cylinders are not very different from those of a hypothetical two-dimensional flow without the oscillatory motion, and this lends credence to the possibility that the numerical method may yet yield a reasonable solution. From Fig. (5.1.1(f)) and Fig. (5.1.2(f)), the highly distorted form of the isotherms and streamlines manifests the unstable nature of the flow at so high Rayleigh numbers.

5.2 The effects of Radius ratio This section focuses on the effects of radius ratio on the flow and the heat transfer in concentric annulus for radius ratios up to 5.0. The higher radius ratios make

52

the effects of the outer cylinder so small that the flow and the heat transfer around the inner cylinder would be almost close to those of a free convecting cylinder. The three radius ratios of 1.25, 2.6 and 5.0 were investigated for the flow and temperature distribution. The radius ratio is one of the important parameters for measuring the heat transfer and the flow between the horizontal annuli. When the radius is low, free movement of the fluid is more difficult and the conductive mode of heat transfer becomes dominant, whereas at the large radius ratios, the heat transfer from the inner cylinder approaches that of natural convection from a heated horizontal cylinder. The isotherm and streamlines for radius ratio of 1.25, 2.6 and 5.0 at various Rayleigh numbers are described in Figs. 5.2.1-5.2.3 and in Figs. 5.1.1-5.1.2. For the larger radius ratios and the smaller radius ratios, the convection and conduction dominate the flow differently. Convective heat transfer which is dependent on the velocity of the flow over the surface is suppressed at the small radius ratios (1.25) because the proximity of the walls forces the opposite-direction flows on the two cylinders to interact strongly with each other (through viscosity) and slow each other down. Therefore, the isotherms remain more nearly circular up to a higher Rayleigh number. At high radius ratios, the flows on the cylinders are separated by greater distances and do not affect each other that the convection have a great effect. The boundary layer flow at high radius ratios thus has a higher momentum which enables it easily to distort the isotherms to create the effect of temperature inversion. It is thus easy for the higher radius ratios to yield temperature inversion.

5.3 The effect of the inner cylinder rotating

53

The effects of the rotation of the inner cylinder on the flow and the heat transfer in concentric annular spaces are the subjects of this section. The computed results are for the radius ratios of 2.6, and as in the previous section, a Prandtl number Pr of about 0.7 is assumed. The actual numerical results and the physics of the flow as revealed by the numerically obtained streamline and isotherm plots are discussed. Figs. 5.3.1-5.3.2 show the streamline plots and the isotherm plots when the inner cylinder is rotated with increasing speed at a fixed Rayleigh number of 5.5 10 4 . The fluid adjacent to the surface of the inner cylinder moves with the inner cylinder for the virtue of viscous drag yielded by the rotating inner cylinder. Rotating the inner cylinder has the effect on rising the hot fluid on the ascending side of the cylinder while on the descending side, the rising hot fluid is slowed down. Near the surface of the inner cylinder on the descending side is a small region of very low flow velocity. The effect of rotation on the flow field at the Reynolds number (Reyl=140) is depicted in Fig. 5.3.1(a). The right-hand cell is elongated while the left-hand cell is reduced. The circulation, defined as the integral U ds , of the left-hand is reduced while that in the

right-hand cell is increased. Since the zero-value streamline has lifted off the surface of the inner cylinder, the thermal plume is shifted in the direction of the rotation which is depicted in the isotherm plot Fig. 5.3.2 (a). The point of impingement on the outer cylinder is correspondingly moved. The right-hand cell grows further in size while the left-hand cell becomes smaller and weaker with increasing the angular speed to the Reynolds number of 280. The left-hand cell is now separated from the inner cylinder. Fig. 5.3.1 (b) showed that there is a zone of Couette-like flow around the inner cylinder. Fig. 5.3.2 (b) depicted that the thermal plume has shifted even further in the direction of rotation and its origin or root, the region near to the surface of the inner cylinder where the thermal plume rises from, is lightly below the horizontal level of the

54

center of the inner cylinder. On the outer cylinder, the point of plume impinge has shifted further down and temperature gradient has dropped. With increasing Reynolds number, the right-hand cell elongates further as the left-hand cell continues to shrink in size. When the Reynolds number rises to 416, the right-hand cell is nearly close to its maximum length. Therefore, the growing Couetteflow region around inner cylinder encroached the cells. At the Reynolds number of 800 and 1120, the kidney-shaped eddy has been changed into a layer of very weak flow pressed against the left-hand surface of the outer cylinder. At the Reynolds number of 1120, the Couette-flow zone has stretched to a third of the clearance. This can been seen in Fig. 5.3.1 (f). The gradual disappearance of the thermal plume can be seen from the Fig. 5.3.2 (c) to 5.3.2 (f). The plume becomes less distinct and the root of the plume, a region of relatively low temperature gradient is extended when the Reynolds number is increased. Therefore, this results in a rapid reduction of the temperature gradient around the inner cylinder. From the Reynolds number of 560 to 800, increased numbers of nearly circular rings of isotherms are formed around the inner cylinder. At the Reynolds number of 1120, the isotherms away from the inner cylinder almost become circular, and the temperature they represent is approaching that of pure conduction. However, this situation is unlike in real flow because it is known that the two-dimensional circular Couette flows at high Reynolds numbers are unstable.

5.4 The effects of eccentricity Eccentricity has a great effect on the flow and heat transfer behavior in the horizontal annulus. The flow field is in fact strongly influenced by the geometry of the annulus as well as its orientation. The effects of eccentricity on the heat transfer were

55

studied at the radius ratios of 2.6 using the numerical model of the finite difference method. The present study focuses on the eccentricities in the angular 450 against the horizontal direction. The streamline and isotherm lines at the eccentricity ratios er of 1/3 and 2/3 and the radius ratio of 2.6 are depicted in Figs. 5.4.1-5.4.3. The Streamline plots depicted that the immediate effect of moving the inner cylinder upwards is to yield a reduction in the size of the two counter-rotating flow cells and to increase the size of the stagnation region at the bottom of the annulus. The thermal boundary layer on the outer cylinder is reduced in length from that of the concentric case. The increase of conduction can be predicted. Moving the inner cylinder downwards has quite the opposite effect. However, For eccentricity ratio e r , the inner cylinder seem to be unaffected by its own position. At the upward eccentricity e = (0, ) , the isotherm plots
r

1 3

2 3

depict that the isotherms around the inner cylinder yield a pair of symmetrical humps. These indicate the inner thermal boundary layer has begun to separate from the cylinder before reaching the top of the annulus. There is no distinct thermal plume as in the other cases. At the downward eccentricity e = (0, ) the outer cylinder has a thermal
r

2 3

boundary layer almost except near the bottom of the annulus. When the Rayleigh number is increased to 10 5 , there are two smaller humps instead of the usual one big hump, and the rising warm fluid in the two plumes yields a pair of relatively weak counter-rotating vortices at the top of the annulus. A pair of very weak cells is formed by viewing the zero-streamline in the stagnation zone. Moving the inner cylinder seems only to cause one of the kidney-shape cells to be squashed in size and the other to grow in size. The size of the stagnation zone is relatively unaffected. The thermal plume is inclined towards the wider side of the annulus, which can be found in the isotherm 56

plots. However, as the increase of the Rayleigh number, the thermal plume tilts towards the side of the narrow throat. The effect of moving the inner cylinder to angular 450 against the horizontal line (x axis see Fig. 1.1.1) at the eccentricity (er = ) has the characteristics like moving the inner cylinder both upward and right. The size of the stagnation region at the bottom of the annulus is increased when the size of the top region is reduced. However, there is still a distinct thermal plume inclined towards the wider side of the annulus. From the streamline plots, one cell appears in the kidney-shaped cells near the narrower throat of the annulus. This is one weak cell. The effect of moving the inner cylinder to the angular -450 at the eccentricity (er = ) has the characteristics like moving the inner cylinder both downward and right. The size of the stagnation region at the top region is increased when the size of the stagnation region at the bottom of the annulus is reduced. The outer cylinder has a thermal boundary layer that occupied great space except the narrower throat of the annulus. (Fig. 5.4.3)
2 3 2 3

5.5 The effects of rotating inner cylinder in the eccentric annulus Fig. 5.5.1 and Fig. 5.5.2 displayed the streamline and isotherm plots respectively at various eccentric positions with the eccentricity ratio er of 1/3. The adopted Reyleigh number and Reynolds number are 5104 and 140 respectively. Compared with the effects of the rotation for the concentric case, it is found that the heat transfer between the two cylinders is strongly determined by what happens in a small region around the inner cylinder, the Couette-flow region. From the streamline and isotherm plots at the eccentricity ratio er of 1/3, the flow and the temperature distribution adjacent to the inner cylinder surface are close to those at the concentric

57

position. The amount of circulating flow through the annulus, as measured by the Stream Function value on the inner cylinder, has fallen from its corresponding concentric value at the same Reynolds number. This is maybe due to the narrower throat in the annular region that present an obstruction to a free flow. Compared with the flow in the concentric annulus, the thermal plumes are shifted to a slightly lesser extent. The root of the thermal plumes at all the eight eccentric positions appears at about the same location on the inner cylinder. For the eccentricity ratio of 2/3 (Fig. 5.5.3-5.5.4), the effect of rotating the inner cylinder is more significant than the eccentricity ratio of 1/3. The narrow throat between the inner and outer cylinders exerts a great influence on both the flow and temperature distribution around the inner cylinder. The amounts by which the root of the thermal plume has moved are different at all the eight eccentric positions, whereas they were about the same at the eccentricity ratio of 1/3. At the eccentricity ratios
2 e = ( ,0) , r 3

the right cell is enlarged significantly so that it reaches most of the annulus

and the left cell is reduced to only a small region near the narrow throat. The narrowness of the throat region makes it difficult for the thermal plume to go through the narrow throat. At the eccentricity ratios e = (0, ) , compared with rotating inner
r

2 3

cylinder in the concentric case, the thermal plume that is shifted by the rotating is close to that of the concentric case. The reason for the phenomena is perhaps that the narrow throat delays the flow velocity to rise. Compared with the stationary eccentric case, the thermal plume is shifted further. At the eccentricity ratios e = (0, ) , the symmetry is
r

2 3

destroyed by the rotation. There is one larger thermal plume on the left side and one small thermal plume on the right. The Figure depicts the left plume and its root are moved a great distance while the right plume has difficulty passing through the throat.

58

At the eccentricity ratios e = ( ,0) the right cell has been squeezed to the left. The
r

2 3

thermal plume and its root are shifted more than at e = ( ,0) , possibly the joint
r

2 3

effect of the rotation and the eccentricity. At the eccentricity ratio ee = 2 / 3 and the inner positions has an angle 450 against the horizontal line, both the narrowness of the throat region and the rotation of the inner cylinder have the effect on the flow and the heat transfer on the annulus. Fig. 5.5.5-5.5.6 separately show the streamline and temperature fields.

59

CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS

The flow and the heat transfer in horizontal concentric and eccentric annuli have been investigated. The two-dimensional model is used to obtain the numerical results from two methods. The results from the finite element method are compared with those from the finite difference method. The finite difference method is easily implemented in computer codes. However, the finite element can provide more accurate results than the finite difference method though the sparse matrix is created, which need more memory and executed time. For the 2-D steady state cases, the author prefers the finite element method. The numerical results are in good agreement with the experimental and numerical results of other investigators such as Kuehn & Goldstein (1976) and Projahn et al. (1981). This is a confirmation of the validity of the numerical model which can be applied with reasonable confidence to obtain useful results in cases where practical experiment could be difficult or time consuming. The effects of parameters such as radius ratio, eccentricity, Rayleigh number, Reynolds number and Prandtl number on the flow and the heat transfer were investigated numerically. The effects of rotating inner cylinder on the flow and the heat transfer in eccentric annuli of the moderate radius ratios were investigated numerically as the important part of the study. Finally, the effects of the rotating the inner cylinder and outer cylinder and both were studied. It is found that the moving boundary condition can increase the internal convection. The rotation can also change the fluid flow distribution such as elongating/reducing the stagnant zone. Particularly, for the concentric and eccentric cases with both cylinders rotating, the outer cylinder rotating can cause the increase of one side of the cooled fluid flow and restrict another side of the cooled fluid flow. It has a great influence on the low temperature stagnant zone.

60

In the future, the turbulence and instability situation is the objective which researchers need to search for. When the radius of the inner and outer cylinders is increased, or the temperature difference between the two cylinders is higher, or the Reynolds number for the inner or outer cylinder rotation is larger, the instability appears and further the turbulence appears to occur. In the situation, the numerical model is not good to calculate the flow and heat transfer between the two horizontal cylinders. Therefore, new turbulence model needs to be developed to solve the problem.

61

REFERENCES

ABBOT M.R. 1964, A numerical method for solving the equations of natural convection in a narrow concentric cylindrical annulus with a horizontal axis, Quart. J. Mech. Appl. Math., 17, pp 471-481. AMES W.F. 1978, Nonlinear partial differential equations in engineering, Academic Press. BALLAL B.Y. & RIVLIN R.S. 1976, Flow of a Newtonian fluid between eccentric rotating cylinders: Inertial effects, Arch. Rational Mech. Anal. 62, pp 237-294. BISHOP E.H. & CARLEY C.T. & POWE R.E. 1968, Natural convection flow patterns in cylindrical annuli, Int. J. of Heat Mass Transfer, 11, pp 1741-1752. BECKMANN W. 1931, Die Warmeubertragung in zylindrischen Gasschichten bei naturlicher Konvektion, Forsch. Geb. D. Ingenieurwesen, 2(5), pp165-178. BIRKHOFF G., VARGA R.S. & YOUNG D. 1962, Alternating direct implicit methods, Advances in Computers, 3, pp 189-273. CASTLE P. & MOBBS F.R. 1968, Hydrodynamic stability of the flow between eccentric rotating cylinders: Visual observations and torque measurements, In: Proc. ImechE, vol.182, Part 3N. Tribology Convention. CHARRIER-MOJTABI M.C., MOJTABI A. & CALTAGIRONE J.P. 1979, Numerical solution of a flow due to natural convection in horizontal cylindrical annulus, J. Heat Transfer, 101, pp 171-173. CHEDDADI A., CALTAGIRONE J.P., MOJTABI A. & VAFAI K. 1992, Free twodimensional convective bifurcation in a horizontal annulus, ASME Journal of Heat Transfer, Vol. 114, pp. 99-106. CHIKHAOUI A., MARCILLAT J.F. & SANI R.L. 1988, Successive transitions in thermal convection within a vertical enclosure, Natural Convection in Enclosures, ASME HTD-Vol. 99, pp 29-35. COLE J.A. 1968, Taylor vortices with eccentric rotating cylinders, Trans. ASME J. Lub. Technol., pp 285-296. CRAWFORD L. & LEMLICH R. 1962, Natural convection in horizontal concentric cylindrical annuli, Industrial Eng. Chemistry Fundamentals, 1(4), pp 260-264. DENNIS A. SIGINER & SAYAVUR I. BAKHTIYAROV 1998, Flow of drilling fluids in eccentric annuli, J. Non-Newtonian Fluid Mech., Vol. 78, pp 119-132. DESRAYAUD G. 1987, Analyse de stailite lineaire dans un milieu semi-transparent, these de doctorat es sciences physiques.

62

ECKERT E.R.G. & SOEHNGEN E.E. 1948, Studies on heat transfer in laminar free convection with the Zehnder-Mach interferometer, Wright-Patterson AFB Tech. Rep. No. 5747, ATI-44580. ECKERT E.R.G. & DRAKE R.M. 1981, Analysis of Heat and Mass Transfer, McGraw-Hill. ESCUDIER M. P. OLIVEIRA & PINHO 2002, Fully developed laminar flow of purely viscous non-Newtonian liquids through annuli, including the effects of eccentricity and inner-cylinder rotation, Int. J. Heat and Fluid Flow, 23, pp52-73 ESCUDIER M.P. & GOULDSON I.W. 1997, Effects of centrebody rotation on laminar flow through an eccentric annulus, In: Adrian, R.J., Durao, D.F.G., Durst, F., Heitor, ESCUDIER M.P. GOULDSON I. W. OLIVEIRA & PINHO 2000, Effects of inner cylinder rotation on laminar flow of a Newtonian fluid through an eccentric annulus, International Journal of Heat and Fluid Flow, 21, pp 92-103. FAROUK B. & GUCERI S. I. 1981, Natural Convection from Horizontal CylinderLaminar Regime, ASME Journal of Heat Transfer, Vol. 103, pp 522-527. FAROUK B. & GUCERI S. I. 1982 Laminar and turbulent natural convection in the annulus between horizontal concentric cylinders, J. Heat Transfer 104, pp 631-636. FANT D.B., ROTHMAYER A. & PRUSA J., 1989, Natural convective flow instability between horizontal concentric cylinders, Numerical Methods in Laminar and Turbulent Flow, Vol. 6 Part 2, Pineridge Press, Swansea, UK, pp. 1047-1065. FUKUMORI, E. and WAKE, A. 1991, "The linkage between penalty function method and second viscosity applied to navier-stokes equations," Computational Mechanics, Vol. 2, pp. 1385-1389 GALPIN, P. F. AND RAITHBY, G. D. 1986, Treatment of non-linearities in the numerical solution of the incompressible Navier-Stokes equations, International Journal for Numerical Methods in Fluids, 6, pp 409-426 GRIGULL, U. & HARF, W. 1966 Natural convection in horizontal cylindrical annuli, 3rd Int. Heat Transfer Conference, Chicago, pp 182-195. GRAY, D.D. & GIORGINI, A. 1976, The validity of the Boussinesq Approximation for liquids and gases, Int. J. Heat Mass Transfer, 19, pp 545-551. GUI G. & STELLA F. 1995, Natural convection in horizontal eccentric annuli: Numerical study, Numerical Heat Transfer, Part A, 27:89-105. HEINRICH, I. C. & PEPPER, D. W. 1999, The Intermediate Finite Element Method: Fluid Flow & Heat Transfer Applications, Taylor & Francis.

63

HOCKNEY, R.W. 1965, A fast direct solution of Poissons equation using Fourier analysis, J. Assn. Computing Mach., 12(1), pp 95-113. KAMAL, M.M. 1966, Separation in the flow between eccentric cylinders, ASME J. Basic Eng., pp 717-724. KENJERES S. & HANJALIC K. 1995, Prediction of turbulent thermal convection in concentric and eccentric horizontal annuli, Int. J. Heat and Fluid Flow 16:429-439. KIM C.J. & RO S.T. 1994, Numerical investigation on bifurcative natural convection in an air-filled horizontal annulus, 10th Int. Heat Transfer Conference, Vol. 7, pp. 8590. KORPELA S.A., GOZUM D. & BAXI C.B., 1973, On the stability of the conduction regime of natural convection in vertical slot, Int. J. Heat Mass Transfer, Vol. 16, pp. 1683-1690. KUEHN, T.H. & GOLDSTEIN, R.J. 1976, An experimental and theoretical study of natural convection in the annulus between horizontal concentric cylinders, J. of Fluid Mech., 74(4), pp 695-719. KUEHN, T.H. & GOLDSTEIN, R.J. 1978, An experimental study of natural convection heat transfer in concentric and eccentric horizontal cylinders, ASME J. of Heat Transfer, 100, pp 635-640. KUEHN, T.H. & GOLDSTEIN, R.J. 1980, A parametric study of Prandtl number and diameter ratio effects on natural convection heat transfer in horizontal cylindrical annuli, J. of Heat Transfer, Trans, ASME, 102, pp 768-770. LAURIAT G. 1980, Numerical study of natural convection in a narrow cavity: an examination of high order accurate schemes, ASME Paper No. 80-HT-90. LAUNDER, B.E. & YING, W.M. 1972, Numerical solutions of flow between rotating cylinders, J. of Mech. Eng. Sci., 14(6), pp 400-403 LAURIAT G. & DESTRAYAUD G. 1985, Natural convection in air-filled cavities of high aspect ratios: discrepancies between experimental and theoretical results, ASME Paper No. 85-HT-37. LEE T. S. 1992, Laminar fluid convection between concentric and eccentric heated horizontal rotating cylinders for low-Prandtl-number fluids, Int. J. For Num. Methods in Fluids, Vol. 14, pp 1037-1062. LEE T. S., WIJEYSUNDERA N. E. & YEO K. S. 1985, Convection in Eccentric Annuli with Inner Cylinder Rotation, AIAA Journal, Vol.24, pp170-171. LEE T. S., HU G. S. & SHU C. 2002, Application of GDQ Method for the Study of Natural Convection in Horizontal Eccentric Annuli, Numerical Heat Transfer, Part A, Vol. 41, pp. 803 - 815.

64

LEE T. S. HU G. S. & SHU C. 2002, Application of GDQ Method for the Study of Mixed Convection in Horizontal Eccentric Annuli, International Journal of Computational Fluid Dynamics (Accepted for Publication: 13 August 2002). LE QUERE P., 1990 A note on multiple and unsteady solutions in two-dimensional convection in tall cavity, ASME Journal of Heat Transfer, Vol. 112, pp. 965-974. LEONARDI, E. & REIZES, J.A. 1981, Convective flows in closed cavity with variable fluid properties, Numerical Methods in Heat Transfer, Edited by Lewis, Morgan and Zienkiewicz, pp 387-412. LEWIS, E. 1979, Steady flow between a rotating circular cylinder and fixed square cylinder, J. Fluid Mech., 95(3), pp 497-513. LIU, C.Y., MUELLER, W.K. & LANDIS, F. 1961 Natural convection heat transfer in long horizontal annuli, International developments in hear transfer ASME, pp 976-984 LOCKETT, T.J. 1992, Numerical simulation of inelastic non-Newtonian fluid flows in annuli. Ph.D. Thesis. Imperial College of Science, Technology and Medicine. MANGLIK, R. M. & FANG, P. P. 1995, Effects of eccentricity and thermal boundary conditions on laminar and fully developed flow in annular ducts. Int. J. Heat Fluid Flow, 16 (4), pp 298-306. MING-I CHAR & GUEY-CHUEN LEE 1998, Maximum density effects on natural convection of micropolar fluids between horizontal eccentric cylinders, Int. J. Engng. Sci. Vol. 36, No.2 pp 157-169. MOHAMED NACEUR BORJINI, CHEIKH MBOW & MICHEL DAGUENET 1999, Numerical analysis of combined radiation and unsteady natural convection within a horizontal annular space, International Journal of Numerical Methods for Heat & Fluid Flow, Vol.9 No.7, pp 742-763. MOJTABI A. CHARRIER-MOJTABI M. C. 1992, Analytical solution of steady natural convection in an annular porous medium evaluated with a symbolic algebra code. J. Heat Transfer 114, pp.1065-1067. OBERBECK, A. 1879, Uber die Warmeleitung der Flussigkeiten bei Berucksichtigung der Stromungen infolge von Temperatur Differenzen, Annalen der Physik and Chemie, 7, pp 271-292. PARJER, J.D., BOGGS, J.H. & BLICK, E.F. 1969 Introduction to Fluid Mechanics and Heat Transfer, Addison-Wesley. PEACEMAN, D.W. & RACHFORD, H.H. 1955, The numerical solution of parabolic and elliptic differential equations, J.Soc. Indust. Appl. Math., 3(1), pp 28-41. POWE, R.E., CARLEY, C.T. & BISHOP, E.H. 1969, Free convection flow patterns in cylindrical annuli, J. of Heat Transfer Trans. ASME, 91, pp 310-314.

65

POW, R.E., CARLEY, C.T. & CARRUTH, S.L. 1971, A numerical solution for natural convection in cylindrical annuli, J. of Heat Transfer, Trans. ASME, 93(12), pp 210-220. PROJAHN, U., RIEGER, H. & BEER, H. 1981, Numerical analysis of Laminar natural convection between concentric and eccentric cylinders, Numerical Heat Transfer, 4, pp 131-146. PRUSA, J. & YAO, L.S. 1981, Natural convection heat transfer between eccentric horizontal cylinders, ASME paper #82-HT-43. RAO Y. F. & FUKUDA K. 1985, Study of Natural Convection in Horizontal Cylindrical Annuli, Int. J. Heat Transfer, Vol. 28, No.3, pp 705-714. ROACHE P.J. 1972, Computational Fluid Dynamics, Hermosa Publisher. ROUX B., GRONDIN J., BONTOUX P. & DE VAHL DAVIS G. 1980, Heat transition from multicellular to monocellular motion in vertical annulus, Phys. Chen. Hydro., Vol. 3F, pp. 292-297. SAN A. A. & SZERI, A.Z. 1984, Flow between eccentric rotating cylinders, ASME J. Appl. Mech. 51, pp 869-878. SHU C. & YEO K. S. 2000, An efficient approach to simulate natural convection in arbitrarily eccentric annuli by vorticity-stream function formulation, Numerical Heat Transfer, Part A, 38: 739-756. SHU C., XUE H. & ZHU Y. Z. 2001, Numerical study of natural convection in an eccentric annulus between a square outer cylinder and a circular inner cylinder using DQ method, Int. Journal of Heat and Mass transfer 44, 3321-3333. STONE H.L. 1968, Iterative solution of implicit approximations of multidimensional partial differential equations, SIAM J. Num. Anal., 5(3), pp 108-113. SIGINER, D.A. & BAKHTIYAROV, S.I. 1998, Flow of drilling fluids in eccentric annuli, J. Non-Newt. Fluid Mech. 78, pp 119-132. TSUI, Y. T. & TREMBLAY, B. 1984 On transient natural convection heat transfer in the annulus between concentric horizontal cylinders with isothermal surfaces, Int. J. Heat Mass Transfer 27, pp 103-111. VAFAI, K., & DESAI, C.P. 1993 Comparative Analysis of the Finite-Element and Finite-Difference Methods for Simulation of Buoyancy-Induced Flow and Heat Transfer in Closed and Open Ended Annular Cavities, Numerical Heat Transfer, Part A, Vol. 23, pp 35-59. VOHR, J.H. 1968 An experimental study of Taylor vortices and turbulence in flow between eccentric rotating cylinders, Trans. ASME J. Lub. Technol., pp 185-296.

66

WAKITANI S. 1997, Development of Multicellular solutions in heat convection in air-filled vertical cavity, ASME Journal of Heat Transfer Vol. 119, pp 97-101. YEO K. S. 1984, A study of thermally-influenced fluid flows between concentric and eccentric horizontal cylinders, Master Thesis. National University of Singapore. YOO J. S. 1998, Natural convection in a narrow horizontal cylindrical annulus: Pr 0.3 , International J. Heat Transfer, 41, pp 3055-3073.

67

Ri Ci Vi 0 er Co R0

g
V0

Figure 1.1.1 Geometry of annular region and the gravity direction

Mesh for finite element method for 2-D computations

Mesh for finite difference method for 2-D computations

Figure 2.1.1 Mesh used for the numerical computation

68

(a)

(b)

(c)

Figure 3.1.1 2-D view of three kinds of shape function: (a) four node element; (b) eight node element; (c) nine node element

69

Table 3.1 Shape Functions of Quadrilateral Elements (4-9) Shape Function


1 = 1 ( 1 )( 1 ) 4 1 ( 1 + )( 1 ) 4 1 ( 1 + )( 1 + ) 4 1 ( 1 )( 1 + ) 4

i=5
1 5 2 1 5 2

i=6

i=7

i=8
1 8 2

i=9
1 9 4 1 9 4

2 =

1 6 2 1 6 2 1 7 2 1 7 2 1 8 2

3 =

1 9 4 1 9 4 1 9 2 1 9 2 1 9 2 1 9 2

4 =

5 =

1 ( 1 2 )( 1 ) 2 1 ( 1 + )( 1 2 ) 2 1 ( 1 2 )( 1 + ) 2 1 ( 1 )( 1 2 ) 2

6 =

7 =

8 =

1 =

1 ( 1 2 )( 1 2 ) 2

Table 4.1 Parameters used in calculations Parameters (kg/m3) (kg/m-s) (W/m-K) CP (J/kg-K) (K-1) Pr Air 1.1774 1.82910-5 2.62410-2 1005.7 3.310-3 0.701

70

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.1.1 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6; inner radius: 0.01 m; outer radius: 0.026 m The inner temperature: 304 K The outer temperature: 301 K Rayleigh number: 1124

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.1.2 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6; inner radius: 0.0431 m; outer radius: 0.112 m The inner temperature: 304 K The outer temperature: 301 K Rayleigh number: 9104

71

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.1.3 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6; inner radius: 0.01 m; outer radius: 0.026 m The inner temperature: 541 K The outer temperature: 301 K Rayleigh number: 9104

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.1.4 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6; inner radius: 0.065 m; outer radius: 0.169 m The inner temperature: 304 K The outer temperature: 301 K Rayleigh number: 3.085105

72

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.1.5 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6; inner radius: 0.07 m; outer radius: 0.182 m The inner temperature: 304 K The outer temperature: 301 K Rayleigh number: 3.85105

73

(a) Temperature

(b) Streamline

(c) Velocity

(d) Temperature

(e) Streamline

(f) Velocity

Fig 4.2.1 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Rayleigh number: 1.5105 (a-c) Radius ratio: 5.0; inner radius: 0.02044 m; outer radius: 0.1022 m The inner temperature: 304 K The outer temperature: 301 K (d-f) Radius ratio: 2.6; inner radius: 0.02044 m; outer radius: 0.053144 m The inner temperature: 347.9 K The outer temperature: 301 K

74

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.2.2 Thermal and fluid flow distribution in concentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 5.0, inner radius: 0.03054 m; outer radius: 0.1527 m The inner temperature: 304 K The outer temperature: 301 K Rayleigh number: 5105

75

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.3.1 Thermal and fluid flow distribution in eccentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6, inner radius: 0.05 m; outer radius: 0.13 m The inner temperature: 305 K The outer temperature: 301 K Rayleigh number: 1.87105 Eccentric ratio: - 0.75

(a) Temperature

(b) Streamline

(c) Velocity

Fig 4.3.2 Thermal and fluid flow distribution in eccentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6, inner radius: 0.05 m; outer radius: 0.13 m The inner temperature: 305 K The outer temperature: 301 K Rayleigh number: 1.87105 Eccentric ratio: 0.75

76

(a) Temperature

(b) Streamline

(c) Velocity

(d) Temperature

(e) Streamline

(f) Velocity

Fig 4.3.3 Thermal and fluid flow distribution in eccentric annuli without rotation The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6, inner radius: 0.028m; outer radius: 0.084m The inner temperature: 307 K; The outer temperature: 301 K Rayleigh number: 4.95104; Eccentric ratio: 0.5 (a-c) for moving the inner cylinder to angular 450 against the horizontal line; (d-e) for moving the inner cylinder to angular -450 against the horizontal line;

77

1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
1. 2. 3. 4. The shortest distance between Ro and Ri of Fig. 4.3.3 (a) The shortest distance between Ro and Ri of Fig. 4.3.3 (d) The longest distance between Ro and Ri of Fig. 4.3.3 (a) The longest distance between Ro and Ri of Fig. 4.3.3 (d)

Fig 4.3.4 Temperature distribution along the shortest line and longest line shooting from the inner cylinder to outer cylinder corresponding to picture in Figure 4.4.3

78

(a) Temperature

(b) Streamline

(c) Temperature Figure 4.4.1

(d) Streamline

Temperature and Streamline velocity plot in the concentric annulus with the inner or outer cylinder rotating with Pr = 0.7, Ro=0.0728m, Ri=0.028m, To=307K, Ti=301K, Ral=4.95104, (a-b) for inner cylinder rotating with tangential velocity 0.1 m/s, (c-d) for outer cylinder rotating with tangential velocity 0.1 m/s

79

1 0.9 0.8 0.7


(T-To)/(Ti-To)

1 6 2 5 8

0.6 0.5 0.4 0.3 0.2 0.1 0 0

1. 00 (Vo= 0.1) 2. 900 (Vo= 0.1) 3. 1800 (Vo= 0.1) 4. 2700 (Vo= 0.1) 5. 00 (Vi = 0.1) 6. 900 (Vi= 0.1) 7. 1800 (Vi= 0.1) 8. 2700 (Vi= 0.1)

4 3 7

0.01

0.02

0.03

0.04

0.05

Distance from Ro to Ri (m)

Figure 4.4.2 Temperature distribution along different angular line shooting from the inner cylinder to outer cylinder corresponding to picture in Figure 4.4.1.

80

(a) Temperature

(b) Streamline

(c) Velocity

(d) Temperature

(e) Streamline Figure 4.4.3

(f) Velocity

Temperature and Streamline velocity plot in the concentric annulus with both inner and outer cylinder rotating with Pr = 0.7, Ro=0.0728m, Ri=0.028m, To=307K, Ti=301K, Ral=4.95104, (a-c) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = 0.05 m/s, (c-d) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = -0.05 m/s.

81

1 0.9 0.8 0.7


(T-To)/(Ti-To)

0.6 0.5 0.4 0.3


3 7

1. 00 (Vo=Vi) 2. 900 (Vo=Vi) 3. 1800 (Vo=Vi) 4. 2700 (Vo=Vi) 5. 00 (Vo = -Vi) 6. 900 (Vo = -Vi) 7. 1800 (Vo = -Vi) 8. 2700 (Vo = -Vi)

2 4 8 5

0.2 0.1 0 0 0.01

0.02 Figure 4.4.4

0.03

0.04

0.05

Distance from Ro to Ri (m)

Temperature distribution along different angular line shooting from the inner cylinder to outer cylinder corresponding to picture in Figure 4.4.3

82

(a) Temperature

(b) Streamline

(c) Velocity

(d) Temperature

(e) Streamline Figure 4.5.1

(f) Velocity

Temperature and Streamline velocity plot in the eccentric annulus of inner cylinder moving up with both inner and outer cylinder rotating with Pr = 0.7, Ro=0.0728m, Ri=0.028m, To=307K, Ti=301K, Ral=4.95104 and er=0.5, (a-c) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = 0.05 m/s, (c-d) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = -0.05 m/s.

83

1 0.9 0.8 0.7


(T-To)/(Ti-To)

3 1

1. 900 (Vo=Vi) 2. 2700 (Vo=Vi) 3. 900 (Vo = -Vi) 4. 2700 (Vo = -Vi)

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Distance from Ro to Ri (m) 2 4

Figure 4.5.2 Temperature distribution along the shortest line and longest line shooting from the inner cylinder to outer cylinder corresponding to picture in Figure 4.5.1.

84

(a) Temperature

(b) Streamline

(c) Velocity

(d) Temperature

(e) Streamline Fig 4.5.3

(f) Velocity

Both of inner and outer cylinder rotating at 0.1m/s The working fluid is air with a Prandtl number of 0.701 Radius ratio: 2.6, inner radius: 0.03m; The inner temperature: 304 K; the outer temperature: 301 K; (a-c) Inner cylinder rotating vel: 0.1m/s; Outer cylinder rotating vel: 0.1m/s; (d-f) Inner cylinder rotating vel: 0.1m/s; Outer cylinder rotating vel: -0.1m/s; Eccentric ratios: 0.75

85

(a) Temperature

(b) Streamline

Velocity

(a) Temperature

(b) Streamline Figure 4.5.4

Velocity

Temperature and Streamline velocity plot in the eccentric annulus of inner cylinder moving down with both inner and outer cylinder rotating with Pr = 0.7, Ro=0.0728m, Ri=0.028, To=307K, Ti=301K, Ral=4.95104 and er=0.5, (a-c) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = 0.05 m/s, (c-d) for inner and outer cylinder rotating with tangential velocity Vi = 0.05 m/s and Vo = -0.05 m/s.

86

1 0.9 0.8 0.7


(T-To)/(Ti-To)

3 1

1. 900 (Vo=Vi) 2. 2700 (Vo=Vi) 3. 900 (Vo = -Vi) 4. 2700 (Vo = -Vi)

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.01


2 4

0.02

0.03

0.04

0.05

0.06

0.07

0.08

Distance from Ro to Ri (m)

Figure 4.5.5 Temperature distribution along the shortest line and longest line shooting from the inner cylinder to outer cylinder corresponding to picture in Figure 5.5.4.

87

6 Overall equivalent heat transfer coefficient

: Present number ro / ri = 2.6 : Kuehn & Goldstein, expt., ro / ri = 2.6 : Grigull & Hauf, expt., ro / ri = 2.45 : Projahn et al., num., ro / ri = 2.6

1 100

1000

10000 Rayleigh number

100000

1000000

Figure 5.0.1 Overall heat transfer coefficient versus Rayleigh number at the radius ratio of 2.6, er = 0 .

88

(a) Ral = 2 10 2

(b) Ral = 10 3

(c) Ral = 10 4

(d) Ral = 5 10 4

(e) Ral = 10 5

(f) Ral = 10 6

Fig. 5.1.1 Isotherm plots at various Rayleigh number, ro / ri = 2.6, er = 0.

89

(a) Ral = 2 10 2

(b) Ral = 10 3

(c) Ral = 10 4

(d) Ral = 5 10 4

(e) Ral = 10 5

(f) Ral = 10 6

Fig. 5.1.2 Streamline plots at various Rayleigh numbers, ro / ri = 2.6, er = 0.

90

(a) Ral = 2 10 2

(b) Ral = 10 3

(c) Ral = 10 4

(d) Ral = 5 10 4

(e) Ral = 10 5

(f) Ral = 10 6

Fig. 5.2.1 Isotherm plots at various Rayleigh numbers, ro / ri = 5.0, er = 0.

91

(a) Ral = 2 10 2

(b) Ral = 10 3

(c) Ral = 10 4

(d) Ral = 5 10 4

(e) Ral = 10 5

(f) Ral = 10 6

Fig. 5.2.2 Isotherm plots at various Rayleigh numbers, ro / ri = 5.0, er = 0.

92

(c) Ral = 10 4

(d) Ral = 10 4

(e) Ral = 10 5 Isotherms

(f)

Ral = 10 5

Streamlines

Fig. 5.2.3 Isotherm and Streamline plots at various Rayleigh numbers, ro / ri = 1.25,
er = 0.

93

(a) Reyl=140

(b) Reyl=280

(c) Reyl=416

(d) Reyl=560

(e) Reyl=800

(f) Reyl=1120

Fig. 5.3.1 Streamline plots at various Reynolds numbers, ro / ri = 2.6 , er = 0,


Ral = 5.5 10 4 .

94

(a) Reyl=140

(b) Reyl=280

(c) Reyl=416

(d) Reyl=560

(e) Reyl=800

(f) Reyl=1120
ro / ri = 2.6 , er = 0,

Fig. 5.3.2 Isotherm plots at various Reynolds numbers,


Ral = 5.5 10 4 .

95

1 (a) e = (0, ) r 3

1 (b) e = (0, ) r 3

1 (c) e = (0, ) r 3

1 (d) e = (0, ) r 3

1 (e) e = ( ,0) r 3

1 (f) e = ( ,0) r 3

Fig. 5.4.1 Isotherm and streamline plots at various eccentric positions, r0 / ri = 2.6,
1 e = , Ral = 5 10 4 . 3

96

2 (a) e = (0, ) r 3

2 (b) e = (0, ) r 3

2 (c) e = (0, ) r 3

2 (d) e = (0, ) r 3

2 (e) e = ( ,0) r 3

2 (f) e = ( ,0) r 3

Streamlines

Isotherms

Fig. 5.4.2 Isotherm and streamlines plots at various eccentric positions, ro / ri = 2.6,
er = 2 , Ral = 5 10 4 . 3

97

(a) e = (
r

2 3

2 1 , ) 2 3 2

(b) e = (
r

2 3

2 1 , ) 2 3 2

(c) e = (
r

2 3

1 2

2 1 ) 3 2

(d) e = (
r

2 3

1 2

2 1 ) 3 2

Streamline

Isotherms

Fig. 5.4.3 Isotherm and Streamline plots at various eccentric positions 450, ro / ri = 2.6,
er = 2 , Ral = 5 10 4 . 3

98

1 (a) e = (0, ) r 3

1 (b) e = (0, ) r 3

1 (c) e = ( ,0) r 3

1 (d) e = ( ,0) r 3

1 Fig. 5.5.1 Streamline plots at eccentric positions, ro / ri = 2.6, er = , Ral = 5.5 10 4 , 3

Rey1=280.

99

1 (a) e = (0, ) r 3

1 (b) e = (0, ) r 3

1 (c) e = ( ,0) r 3

1 (d) e = ( ,0) r 3

1 Fig. 5.5.2 Isotherm plots at eccentric positions, ro / ri = 2.6, er = , Ral = 5.5 10 4 , 3

Rey1=280.

100

(a) e = (0,2 / 3)
r

2 (b) e = (0, ) r 3

2 (c) e = ( ,0) r 3

2 (d) e = ( ,0) r 3 2 , Ral = 5.5 10 4 , 3

Fig. 5.5.3. Streamline plots at eccentric positions, ro / ri = 2.6, er = Rey1=280,

101

2 (a) e = (0, ) r 3

2 (b) e = (0, ) r 3

2 (c) e = ( ,0) r 3

2 (d) e = ( ,0) r 3 2 , Ral = 5.5 10 4 , 3

Fig. 5.5.4 Isotherm plots at eccentric positions, ro / ri = 2.6, e =


r

Rey1=280.

102

(a) e = (
r

2 3

2 1 ) 2 3 2 ,

(b) e = (
r

2 3

1 2

2 1 ) 3 2

(c) e = (
r

2 3

2 1 , ) 2 3 2

(d) e = (
r

2 3

1 2

2 1 ) 3 2

Fig. 5.5.5 Streamline plots at eccentric positions 45 0 ,135 0 , , ro / ri = 2.6, e =


r

2 , 3

Ral = 5.5 10 4 , Rey1=280.

103

(a) e = (
r

2 3

2 1 ) 2 3 2 ,

(b) e = (
r

2 3

1 2

2 1 ) 3 2

(c) e = (
r

2 3

2 1 , ) 2 3 2

(d) e = (
r

2 3

1 2

2 1 ) 3 2

Fig. 5.5.6 Isotherm plots at eccentric positions, ro / ri = 2.6, e =


r

2 , Ral = 5.5 10 4 , 3

Rey1=280.

104

You might also like