You are on page 1of 20

CFD calculation of convective heat transfer coefficients and validation Part 2:

Turbulent flow

Annex 41 Kyoto, April 3rd to 5th, 2006

Adam Neale1, Dominique Derome1, Bert Blocken2 and Jan Carmeliet2,3


1) Dep. of Building, Civil and Environmental Engineering, Concordia University, 1455 de
Maisonneuve blvd West, Montreal, Qc, H3G 1M8, corresponding author e-mail:
aneale@sympatico.ca
2) Laboratory of Building Physics, Department of Civil Engineering, Katholieke Universiteit Leuven,
Kasteelpark Arenberg 40, 3001 Heverlee Belgium

3) Building Physics and Systems, Faculty of Building and Architecture, Technical University
Eindhoven, P.O. box 513, 5600 MB Eindhoven, The Netherlands

Abstract
When a fluid flows over a wall and heat is exchanged, the boundary layer (BL)
velocity profile will to a large extent determine the value of the convective heat
transfer coefficient (hc). Using the Computational Fluid Dynamics (CFD) software
Fluent, the turbulent flow field and heat transfer is simulated for forced convection
over a smooth flat plate and the performance of the most commonly used turbulence
models is evaluated. The BL velocity and temperature profiles are compared with
semi-empirical near-wall data and the heat transfer coefficients calculated in Fluent
are validated by comparison with correlations from literature. In addition, a
simulation of natural convection is provided as an example of when wall function
theory fails.

1. Introduction
The surface coefficients for heat and mass transfer (hc and hm, respectively)
are parameters that are generally not easily calculated analytically and difficult to
derive from experimental measurements. The values of surface coefficients depend
on many variables flow field, boundary conditions, material properties, etc. In
addition, despite the fact that the two transfer processes are mutually dependent,
they are often solved as uncoupled phenomena. Finally, although existing
correlations relating hc and hm are only valid for specific cases, such correlations are
applied widely throughout literature.
This paper is the second part of a two part study of the option to solve for hc
using Computational Fluid Dynamics (CFD). In Part I, the CFD code Fluent was

1
validated for convective heat transfer in the laminar regime for fully developed flow
between parallel plates. Part II is a validation and comparative study of heat transfer
coefficients calculated using different turbulence models implemented in Fluent. The
validity of using wall functions for natural convection cases is also examined.
The paper starts with a brief outline of the methodology that will be used to
validate the results of the simulations. A brief overview of wall function theory is
provided as a necessary basis for the following sections. In Section 4 the heat
transfer theory is presented for forced convection over a smooth flat plate. The
simulation results for both forced and natural convection are in Sections 5 and 6,
respectively, followed by some conclusions and remarks in Section 7.

2. Methodology
In the interest of validating the turbulent models within Fluent, it was
necessary to obtain experimental data to use as a basis for comparison. While
experiments have been performed in this area, it is often difficult to establish whether
the simulation truly matches all of the experimental parameters. However, one area
that has been focused on extensively in the past is the universal law-of-the-wall
that describes turbulent boundary layer flow. Through analytical derivations of
equations and experimental data fitting, the boundary layer velocity profile (and
temperature profile, if applicable) has been subdivided into three regions: the laminar
sublayer, the buffer region, and log-law region (Chen & Jaw 1998, Blocken 2004).
Semi-empirical relationships have been developed for the laminar sublayer and log-
law regions, and empirical equations exist for the buffer region as well (e.g. Spalding
1961). The (semi-)empirical equations will be used to validate the simulation results
from Fluent.

3 Near-wall modelling
Boundary layer velocity and temperature profiles are generally described
using dimensionless parameters. Before the BL regions can be discussed in proper
detail, some dimensionless terms must be introduced:
yu *
y+ (1)

where y+ is the dimensionless distance from the wall, y is the distance from the wall,
is the kinematic viscosity, and u* is the friction velocity defined as:

2
w
u* (2)

where w is the wall shear stress and is the fluid density. The wall shear stress is

based on the velocity gradient in the direction normal to the surface of the wall, or in
equation form:
U
w (3)
y y =0

where U is the fluid velocity along the wall, and is the dynamic viscosity. The
velocity can be described in a dimensionless form as a function of the fluid velocity
and the friction velocity:
U
u+ (4)
u*
For cases with heat transfer, the dimensionless temperature may be calculated using
the following equation:
T wT
T+ (5)
T*
where Tw is the wall temperature at a certain point, T is the fluid temperature, and T*
is defined as
qw
T* (6)
k u*
where is the thermal diffusivity, qw is the wall heat flux and k is the thermal
conductivity.
There are two common near-wall modeling techniques employed in CFD:
Low-Reynolds-number modelling and Wall function theory.

3.1 Low-Reynolds-number modelling (Low-Re)


If the boundary layer is meshed sufficiently fine so that the first cells are
placed entirely in the laminar sublayer of the BL, the approach used is generally
referred to as Low-Re Modelling. The laminar sublayer is valid up to y+ < 5 and, in
dimensionless coordinates, the height of the first cell is generally taken to be
approximately y+ = 1. In the range of 5 < y+ < 30 there exists a buffer region between
the laminar sublayer and the log-law region of the boundary layer. It is generally not
advisable to have meshes where the first cell lies within the buffer region, though

3
often it is unavoidable in CFD. For meshes with a y+ > 30, wall function theory may
be applied.

3.2 Wall function theory


Fluid flow over a smooth flat plate is referred to as the simplest case for
analytical fluid dynamics (Schetz 1993). There has been a significant amount of
work done in experiments for boundary layer flow evaluation (summarized in Bejan
1984, Schlichting 1987, Schetz 1993, Chen & Jaw 1998, etc). That work was
subsequently transformed into the wall function concept (e.g. Spalding 1961). Wall
functions allow CFD models to interpret behaviour near a wall without the need for a
very fine mesh that also discretises the generally quite thin laminar sublayer at the
surface of the wall. The wall function equations are based on an analytical solution
of the transport equations in combination with experimental data fitting. The result is
a reduction in computation time and a relatively accurate representation of what
happens within the BL, at least under the conditions for which the wall functions were
derived. Wall functions are recommended for cases where the domain is that large
and the Reynolds numbers are that high that the low-Re number modelling approach
would lead to a mesh that is too large and can no longer be solved economically. On
the other hand, wall functions may cease to be valid in complex situations.
Nevertheless, they are often used even when not valid for complex calculations,
which can be responsible for considerable errors in near-wall flow and the related
convective heat transfer coefficients (Blocken 2004).
Wall functions are generally described as having two regions: the laminar
sublayer and the log-law layer. It is commonly accepted in CFD that the laminar
sublayer is said to be valid in the region where y+ < (5 to 10) (Chen & Jaw 1998).
The equations for the dimensionless velocity and temperature within this region are
(Fluent Inc. 2003):
u+ = y+ (7)

T + = Pr y + (8)
where Pr is the Prandtl number (Pr = /). The region above the laminar sublayer
(y+ > 30) is the log-law layer, which is generally described in the form of:
n + = A ln y + + B (9)

4
where n is either the dimensionless velocity or dimensionless temperature. The
constants A and B in Equation (9) are usually fitted to experimental data, and are
consistent throughout the literature. For the purpose of this report, the following
equations will be used (Fluent Inc. 2003):
u + = 2.5 ln y + + 5.45 (10)

1
T + = Prt ln( Ey + ) + P (11)

where Prt is the turbulent Prandtl number (= 0.85 for air), E is an experimentally
determined constant (= 9.793), and P is described by the following equation:

3
4 Pr
0.007

Pr
P = 9.24 1 1 + 0.28e Prt
(12)
Prt

Spalding (1961) suggests an equation that will cover the entire y+ range of values for
the dimensionless velocity u+ (including the buffer region):
1
y + = u + + Aexp Bu + 1 Bu + Bu +
2
( ) 2

1
6
( )
3
Bu +
1
24
( )
4
Bu + (13)

where A=0.1108 and B=0.4. The equations for the dimensionless velocity and
temperature in the laminar sublayer, the buffer layer and the logarithmic layer are
illustrated in Figure 1.

25 25
u+ Equation (7)
u+ Equation (10)
20 u+ Spalding Equation (13) 20
T+ Equation (8)
T+ Equation (11)
15 15
u+ T+
10 10

5 5

0 0
1 10 100 1000
y+

Figure 1. Wall function dimensionless velocity and temperature distributions

5
It is important to note that Fluent uses a slightly different dimensionless height
y*, but for the cases that were examined the values of y+ and y* are equivalent
(Fluent Inc. 2003).

4. Heat transfer in turbulent flow


Convective heat transfer coefficients are generally expressed in the form of
dimensionless correlations that are based on experimental data. There are a
number of works in literature that summarize the numerous correlations that exist for
different types of flows (e.g. Bejan 1984, Saelens 2002, Lienhard & Lienhard 2006).
For the purpose of this paper, two correlations were selected from Lienhard &
Lienhard (2006) that correspond to the geometry and flow conditions for the forced
convection cases that were simulated. They are given by Equations (14) and (15)
below. The dimensionless parameters required for the equations are as follows
(further described in Appendix C):
U x
Reynolds Number: Re x


Prandtl Number: Pr

hc
Stanton Number: St
c pU
hc x
Nusselt Number: Nu x
k
0.455
Skin Friction Coefficient: C fx = (White 1969)
[ln(0.06 Re x )]2
C fx 2
St x = ; Pr > 0.5 (14)
( )
1 + 12.8 Pr 0.68 1 C fx 2

Nu x = 0.032 Re 0x.8 Pr 0.43 (15)

The heat transfer coefficients for the natural convection case are not compared in
the framework of this paper.

5. Turbulent Forced Convection Simulations


5.1 Computational domain

6
The domain used to represent fluid flow over a flat plate is shown in Figure 2.
The boundary condition (BC) for the top of the domain was chosen to be a symmetry
condition in order to reduce the computation time of the simulation. If a pressure
outlet BC is chosen instead of symmetry it can lead to convergence problems when
modelling turbulence. The height of the domain was selected to be high enough to
reduce the influence of the symmetry condition on the boundary layer.

Symmetry BC

U = 0.5m/s Velocity Pressure


H = 1m
Y Inlet BC Outlet BC
T = 283 K
Wall BC - Constant Heat Flux
X 2
qw = 10 W/m
Note : Not to scale.
L = 5m

Figure 2. Computational domain and boundary conditions (BC)

As explained in Section 3.1, the Low-Re modelling approach recommends


that the first grid cell has a dimensionless height of y+ 1, which means that it is
submerged in the laminar sublayer. (Fluent Inc. 2003). For simulations with Wall
Functions (WF), a y+ between 30 and 60 is recommended. The y+ value is based on
the flow conditions at the surface (see Equation (1)) and therefore requires an
iterative procedure to properly size the first cell. After a number of grid adjustments
the mesh fulfilled the requirements for Low-Re modelling, and is shown below in
Figure 3. The mesh used for wall function cases is shown in Figure 4. An
exponential relationship was used to mesh the vertical direction and a uniform
spacing was used for the horizontal direction. The grid dimensions are shown in
Table 1.

Table 1. Grid parameters and dimensions forced convection case


Grid #Cells in #Cells in Smallest Cell Smallest Cell Total number
X-direction Y-direction Width Height of cells
-3
Low-Re 500 100 0.01 m 1.285x10 m 50000
-2
WF 100 13 0.03 m 4.653x10 m 1300

7
Figure 3. Grid used for simulations with low-Reynolds-number modelling

Figure 4. Grid used for simulations with wall functions

Note that the grids for the Low-Re modelling and wall function cases can have
the same spacing near the symmetry boundary, since the boundary layer solution
will not be not affected by the grid resolution near the top region of the domain.
The simulations were all initialized with a uniform velocity profile of 0.5 m/s.
The simulations were iterated until the scaled residuals for all parameters were
below 10-7. The outlet velocity profile and turbulence conditions were then used as
the new inlet conditions and the simulation was repeated. The thermal conditions at
the outlet were not used as new inlet conditions. Instead, the same inlet profiles were
used for each simulation. This means that the flow was always thermally developing
from the start of the domain. This procedure was continued until the inlet and outlet
velocity profiles were approximately the same, resulting in a fully developed flow

8
profile. The original uniform velocity profile ensured that the bulk velocity was 0.5
m/s for all cases. A sample of the Fluent solution parameters is provided in
Appendix A. The material properties used for the fluid region in the simulations are
provided below in Table 2.

Table 2. Material properties for air


Density 1.225 kg/m3
Dynamic Viscosity 1.7894 x 10-5 kg/ms
Thermal Conductivity k 0.0242 W/mK
Heat Capacity cp 1006.43 J/kgK

5.2 Simulation results for forced convection


The simulation results are compared at x= 4.5 m for all cases. Simulations
were performed with the following turbulence models with Low-Re Modelling:
1) Spalart-Allmaras Model
2) Standard k- Model
3) RNG k- Model
4) Realizable k- Model
5) Standard k- Model
6) SST k- Model
7) Reynolds Stress Model (RSM)

Simulations were performed with the following models with Wall Functions (WF):
1) Standard k- Model
2) Standard k- Model

Note that the Standard k- Model will automatically interpret whether Low-Re or WF
will be used based on the y+ of the first cell. The default settings for each model
were used for all cases unless otherwise specified.

Figures 5 and 6 and 7 compare the empirical and semi-empirical dimensionless


velocity and temperature profiles on one hand with the calculated dimensionless
profiles on the other hand. The empirical and the calculated convective heat transfer
coefficients are shown in Figure 7.

9
25
Semi-Empirical Equation - Laminar Sublayer
Semi-empirical Equation - Log-law
Empirical Equation - Spalding (1961)
k-e standard
20 k-e RNG
k-e realizable
k-w standard
k-w SST
Spalart-Allmaras
15 RSM
WF - ke
u+ WF - kw

10

0
1 10 100 1000

y+

Figure 5. Comparison of empirical, semi-empirical and calculated dimensionless velocity


profiles for the turbulent simulations

16

Semi-empirical Equation - Laminar Sublayer


Semi-empirical Equation - Log-law
k-e standard
k-e RNG
k-e realizable
12 k-w standard
k-w SST
Spalart-Allmaras
RSM
WF - ke
WF - kw
T+ 8

0
1 10 100 1000

y+

Figure 6. Comparison of empirical, semi-empirical and calculated dimensionless


temperature profiles for the turbulent simulations

10
6

k-e standard
k-e RNG
k-e realizable
5
k-w standard
k-w SST
Spalart-Allmaras
RSM
4
k-e WF
k-w WF
hc (W/m K)

Lienhard (2006) Eq. 6.111


2

Lienhard (2006) Eq. 6.115


3

0
0 1 2 3 4 5

X Position (m)

Figure 7. Comparison of empirical and calculated convective heat transfer coefficients


for the turbulent simulations

5.3 Forced convection results summary


The velocity profiles shown in Figure 5 indicate a good agreement between
the simulations and the universal law-of-the-wall relationships and the universal
Spalding curve, which were both developed based on experimental data. The
laminar sublayer and the log-law region are well defined for all of the turbulence
models, though some models (RSM) tend to under predict the velocity near the
upper boundary (for large values of y+). This can be explained by the fact that the
law-of-the-wall relationship ceases to be valid beyond a certain point (roughly y+ >
500, but the actual boundary depends on the situation) (Blocken 2004).
The temperature profiles in Figure 6 are also consistent with the expected
boundary layer profile, though at upper regions of y+ (>200) the curves begin to
diverge from the log-law equation. The same remark can be made here concerning
the failure of the law-of-the-wall theory at high y+-values.
The empirical correlations for heat transfer are shown in red on Figure 7. The
heat transfer coefficients are consistent between the turbulence models and the
correlations, including the solutions using wall functions. However, in the thermally
developing region (approximately 0 m < x < 1 m), the wall function solutions differ

11
from the other curves. The result is an important underprediction of heat transfer for
cases where there is thermally developing flow. This is due to the fact that the wall
function approach is not valid under these conditions.

6. Natural convection simulations


6.1 Computational domain
The domain used to represent a case of natural convection is shown in Figure
8. The boundary conditions (BC) for the four sides of the domain are impermeable
walls. The vertical wall at x= 0 m is maintained at a constant temperature of 298K,
while the opposite wall at x= 3.0 m is maintained at 288K. The remaining two walls
are considered isothermal. The fluid temperature was initialized to 293K.
The domain is divided into two regions denoted as the Hot Side and the
Cold Side. The velocity profile and temperature profile results will be described in
terms of which wall they are incident upon, either the Hot wall of 298K or the Cold
wall of 288K. The simulation results will be compared at three reference heights: y =
0.75 m, y = 1.50 m, and y = 2.25 m.

Isothermal
y = 3m
Air, Tinitial = 288K

y = 2.25m
Hot Side Cold Side

y = 1.50m
Tw1 = 298K Tw2 = 288K

y = 0.75m

Y
Note : Not to scale.
0m
0m X Isothermal X = 3m

Figure 8. Natural convection computational domain

Two meshes were created to represent the domain in Figure 8: one for Low-Re
modelling and one for wall functions. The grid guidelines for Low-Re modelling and
wall functions described in Section 3 were followed where applicable. The grid

12
parameters are outlined below in Table 3, and the two meshes are illustrated in
Figure 9.

Table 3. Grid parameters and dimensions natural convection case


Grid #Cells in #Cells in Smallest Cell Largest Cell Total number
X-direction Y-direction Dimension Dimension of cells
-3 -2
Low-Re 150 150 1.784x10 m 7.580x10 m 22500
WF 30 30 0.1 m 0.1 m 900

Figure 9. Low-Re modelling mesh (left) and wall function mesh (right)

6.2 Simulation results for natural convection


The velocity profiles that are the result of natural convection flows are very
different from those predicted by wall functions. A typical example of a velocity
profile near the surface of a wall (from the Low-Re simulation results) is compared to
the Spalding profile (converted to dimensional units) in Figure 10. The shape of the
simulation velocity profile is consistent with velocity profiles from other natural
convection simulations and experiments (Zitzmann et al 2005). The Spalding profile
represents the velocity profile that would be used when wall functions are employed.

13
0.8
Low-Re Velocity profile at y = 2.25 m , Hot s ide
0.7 Spalding wall function equation

0.6

0.5

Velocity (m/s)
0.4

0.3

0.2

0.1

0
0 0.05 0.1 0.15
Position (m)

Figure 10. Velocity profiles for natural convection (simulation) and universal law-of-the-wall

The two simulation cases resulted in different velocity patterns within the
domain. The velocity fields for both simulation domains are shown in Figure 11.

Figure 11. Velocity contours for the Low-Re simulation (left) and the wall function simulation
(right). (Note that the scale is in m/s)

Using wall functions will impose a velocity gradient along every wall surface more or less
uniformly. In the Low-Re contour plot the boundary layer velocity profile in the upper right
hand corner is clearly not the same as the upper left hand corner. As described in Figure 8,
the velocity profile data was output for three heights at the hot and cold walls. The
profiles were non-dimensionalized with the method described in Section 3, and the results
are shown below in Figure 12. The Spalding curve is plotted on Figure 12 to compare the
simulation results with the law-of-the-wall theory.

14
30
Low-Re: y=0.75m , Hot s ide Low-Re: y=0.75m , Cold s ide
Low-Re: y=1.50m , Hot s ide Low-Re: y=1.50m , Cold s ide
Low-Re: y=2.25m , Hot s ide Low-Re: y=2.25m , Cold s ide
25 WF: y=0.75m , Hot Side WF: y=0.75m , Cold Side
WF: y=1.50m , Hot Side WF: y=1.50m , Cold Side
WF: y=2.25m , Hot Side WF: y=2.25m , Cold Side
Spalding Wall Function Equation
20

u+ 15

10

0
1 10 100 1000 10000
y+

Figure 12. Non-dimensionalized velocity profiles from the Low-Re and wall function (WF)
simulations

6.3 Natural convection results summary


The results for the natural convection simulations indicate that the standard
wall functions are simply not valid for these cases. The velocity profiles along the
hot and cold walls found in the Low-Re simulation are consistent with profiles for
similar cases in literature, both simulated and experimental (Zitzmann et al 2005).
When compared with the velocity profile described by wall functions, there is a
significant difference in both shape and magnitude. In addition to boundary layer
differences, the velocity field in the domain is very different when wall functions are
used (refer to Figure 11). It is counteradvised to use standard wall functions for the
natural convection cases described in this paper.

15
7. Conclusions
Semi-empirical relationships developed using experimental data and
analytical theory were used to validate CFD simulation results for fully developed
forced convection over a smooth flat plate. The results indicate a good agreement
between (semi-)empirical equations and simulation boundary layer velocity and
temperature profiles for all of the turbulence models studied.
The heat transfer coefficients calculated from the forced convection
simulations are consistent for all of the turbulence models studied, and also coincide
closely with selected correlations from literature. The results for simulations with wall
functions indicate that the heat transfer coefficients calculated in the thermally
developing region of the domain are not consistent with the Low-Re simulation
results or with the correlations. It is concluded that the wall functions are not valid for
thermally developing regions.
Two simulations of natural convection were performed to investigate the
validity of wall functions for natural convection simulations. The Low-Reynolds-
number (Low-Re) simulation resulted in flow field and local velocity profiles
consistent with results seen in literature. The local velocity profiles were very
different from profiles obtained by the wall function equations, and the flow field
resulting from the wall function simulation differed significantly from the Low-Re case.
The dimensionless velocity profiles from the wall function simulation were very
different from natural convection results from literature. It is concluded that the
standard wall functions are not valid for cases involving natural convection and that
instead low-Re number modelling should be used.

16
Appendix A: Sample forced convection Fluent solution parameters

k- Turbulent model parameters:

Model Settings
Space 2D
Time Steady
Viscous Standard k-epsilon turbulence model
Wall Treatment Standard Wall Functions
Heat Transfer Enabled
Solidification and Melting Disabled
Radiation None
Species Transport Disabled
Coupled Dispersed Phase Disabled
Pollutants Disabled
Soot Disabled

Equation Solved
Flow yes
Turbulence yes
Energy yes

Numerics Enabled
Absolute Velocity Formulation yes

Relaxation:
Variable Relaxation Factor
Pressure 0.3
Density 1
Body Forces 1
Momentum 0.7
Turbulent kinetic energy 0.8
Turbulent dissipation rate 0.8
Turbulent viscosity 1
Energy 1

Solver Termination Residual Reduction


Variable Type Criterion Tolerance
Pressure V-Cycle 0.1
X-Momentum Flexible 0.1 0.7
Y-Momentum Flexible 0.1 0.7
Turbulence Kinetic Energy Flexible 0.1 0.7
Turbulence Dissipation Rate Flexible 0.1 0.7
Energy Flexible 0.1 0.7

Discretization Scheme
Variable Scheme
Pressure Standard
Pressure-Velocity Coupling SIMPLE
Momentum First Order Upwind
Turbulence Kinetic Energy First Order Upwind
Turbulence Dissipation Rate First Order Upwind
Energy First Order Upwind

Solution Limits
Quantity Limit
Minimum Absolute Pressure 1
Maximum Absolute Pressure 5000000
Minimum Temperature 1
Maximum Temperature 5000
Minimum Turb. Kinetic Energy 1e-14
Minimum Turb. Dissipation Rate 1e-20
Maximum Turb. Viscosity Ratio 100000

17
Appendix B: Sample natural convection Fluent solution parameters

Model Settings
Space 2D
Time Steady
Viscous Standard k-epsilon turbulence model
Wall Treatment Enhanced wall treatment
Heat Transfer Enabled
Solidification and Melting Disabled
Radiation None
Species Transport Disabled
Coupled Dispersed Phase Disabled
Pollutants Disabled
Soot Disabled

Equation Solved
Flow yes
Turbulence yes
Energy yes

Numerics Enabled
Absolute Velocity Formulation yes

Relaxation:
Variable Relaxation Factor
Pressure 0.3
Density 1
Body Forces 1
Momentum 0.7
Turbulent kinetic energy 0.8
Turbulent dissipation rate 0.8
Turbulent viscosity 1
Energy 1

Solver Termination Residual Reduction


Variable Type Criterion Tolerance
Pressure V-Cycle 0.1
X-Momentum Flexible 0.1 0.7
Y-Momentum Flexible 0.1 0.7
Turbulence Kinetic Energy Flexible 0.1 0.7
Turbulence Dissipation Rate Flexible 0.1 0.7
Energy Flexible 0.1 0.7

Discretization Scheme
Variable Scheme
Pressure Body Force Weighted
Pressure-Velocity Coupling SIMPLE
Density First Order Upwind
Momentum First Order Upwind
Turbulence Kinetic Energy First Order Upwind
Turbulence Dissipation Rate First Order Upwind
Energy First Order Upwind

Solution Limits
Quantity Limit
Minimum Absolute Pressure 1
Maximum Absolute Pressure 5000000
Minimum Temperature 1
Maximum Temperature 5000
Minimum Turb. Kinetic Energy 1e-14
Minimum Turb. Dissipation Rate 1e-20
Maximum Turb. Viscosity Ratio 100000

18
Appendix C: Nomenclature

a Order of the discretization error scheme (-)


b Distance between parallel plates (m)
cp Specific heat (J/kgK)
Dh Hydraulic diameter (m)
hc Convective heat transfer coefficient (W/m2K)
k Thermal conductivity (W/m-K)
L Length of domain (m)
NuDh Nusselt number calculated with the hydraulic diameter (-)
P Heated perimeter of the domain (m)
q Heat flux (W/m2)
u Velocity component in the x-direction (m/s)
v Velocity component in the y-direction (m/s)
T Temperature (K)
U Velocity magnitude (m/s)

Greek symbols
Dynamic viscosity (kg/m-s)
Density (kg/m3)

Subscripts
AV Average property
b Bulk property
c Property taken at the centerline of the domain
f Fluid property
i Property of an element i
ref Reference property
x Property taken at a location x
Free stream property

19
References

1. Bejan, A., Convection Heat Transfer, John Wiley & Sons, Inc., 1984.
2. Blocken, B., Wind-driven rain on buildings, Ph.D. thesis, Leuven: K.U.Leuven.,
2004.
3. Chen, C.-J., Jaw, S.-Y., Fundamentals of Turbulence Modeling, Taylor &
Francis, 1998.
4. Fluent 6.1 Users Guide, 2003.
5. Lienhard IV, J.H., Lienhard V, J.H. A Heat Transfer Textbook, Phlogiston Press,
2006.
6. Saelens, D., Energy performance assessment of single story multiple-skin
facades, Ph.D. dissertation, Leuven: K.U.Leuven, 2002.
7. Schlichting, H., Boundary-Layer Theory, McGraw-Hill, 7th Edition, 1987.
8. Schetz, J.A., Boundary Layer Analysis. Prentice Hall, 1993.
9. Spalding, D.B., A single formula for the law of the wall, J. Appl. Mech., Vol 28,
1961, pp. 455-457.
10. White, F.M., A new integral method for analyzing the turbulent boundary layer
with arbitrary pressure gradient, J. Basic Engr., 91: 371-378, 1969.
11. Zitzmann, T. et al, Simulation of steady-state natural convection using CFD,
Ninth Int. IBPSA Conference, pp 1449-1456, 2005.

20

You might also like