Sei sulla pagina 1di 8

International Journal of Heat and Mass Transfer 63 (2013) 183190

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

A boundary-domain integral equation method for solving convective


heat transfer problems
Xiao-Wei Gao , Hai-Feng Peng, Jian Liu
State Key Laboratory of Structural Analysis for Industrial Equipment, Dalian University of Technology, Dalian 116024, China

a r t i c l e

i n f o

Article history:
Received 12 December 2012
Received in revised form 20 March 2013
Accepted 29 March 2013
Available online 27 April 2013
Keywords:
Convective heat transfer
NavierStokes equations
Energy equation
Boundary element method
Steady incompressible viscous ow

a b s t r a c t
In this paper, a new boundary-domain integral equation for solving energy equation in convective heat
transfer problems is derived. The derived integral equation accounts for uid velocity which is computed
using the velocity boundary-domain integral equations presented in [2,3] (X.W. Gao, 2004, 2005).
Numerical implementation of these integral equations is discussed for steady incompressible uid ows,
in which the velocity and pressure integral equations are uncoupled from the temperature equation, and
thus can be solved separately. By substituting the velocity, velocity gradient, and pressure integral equations into the energy integral equation, temperatures can be computed without iterative processes. Two
numerical examples for 2D problems are given to validate the proposed method.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
The governing equations for convective heat transfer problems
are the continuity, momentum, and energy equations. In general,
they are coupled nonlinear partial differential equations which
are not amenable to analytical solutions except for a few simple
problems [1]. Therefore, in the past half century, many works have
been done to develop efcient numerical techniques to solve these
equations. Currently, the most popular methods for solving these
equations are the nite difference method, nite element method,
and nite volume method. Their common feature is that they are
based on domain variable representations and local interpolation
schemes, resulting in systems of equations that are highly sparse
matrices. Recently, the boundary element method (BEM) has been
used to analyze convective heat transfer problems. Compared to
other numerical methods, the distinct advantages of BEM are that
the velocity and temperature gradient formulations can be explicitly derived from the velocity and temperature integral equations
so that the pressure and heat ux have the same computational
accuracy as velocity and temperature themselves (e.g. [24]), and
the boundary conditions at innity are automatically satised in
BEM formulations.
During the past three decades, BEM formulations for convective
heat transfer problems have been extensively studied [515]. Part
of the works focus on the governing convective diffusion equations
[59]. The rst integral equation analysis for steady convective
Corresponding author. Tel.: +86 411 84706332.
E-mail address: xwgao@dlut.edu.cn (X.-W. Gao).
0017-9310/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2013.03.071

diffusion equations was carried out by Ikeuchi and Onishi [5] using
the fundamental solutions of the diffusion-convection and Laplacian operator to develop direct and iterative boundary element
methods respectively. Shi and Banerjee [6] developed a boundary
element formulation for linearized convective heat transfer problems based on unsteady fundamental solutions. In a different
way, free-space time-dependent convective diffusion fundamental
solutions were used by Grigoriev and Dargush [7] to obtain higherorder BEM integral equations. Then the equations are extended to
solve two-dimensional problems [8], from which numerical examples are given to investigate the performance of the BEM formulations [9]. In these mentioned methods, the viscous dissipation is
not taken into consideration in the convective heat transfer problems since NavierStokes equations are not employed in the uid
ow computation.
A BEM formulation was directly derived from the governing
equations of two-dimensional incompressible viscous uids by
Dargush and Banerjee [10] for solving convective heat transfer
problems. The used energy equation includes the terms of viscous
dissipations as a portion of pseudo sources. The resulting integral
equations are written exclusively in terms of velocities and temperatures, and the computation of velocity and temperature gradients is removed through integration by parts. The same technique
was also employed by Banerjee and Honkala [11] to obtain a set of
integral equations incorporating the buoyancy effect for the analysis of the natural convection heat transfer. Different BEM formulations were also established to solve natural convection problems
by other authors, e.g. Onishi et al. [12], and Tosaka and Fukushima
[13]. In addition, the dual reciprocity boundary element method

184

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

was used by Rahaim and Kassab [14] to solve incompressible laminar viscous uid ows and heat transfer, in which all nonlinear
terms were lumped into the forcing term.
As to the transient convective heat transfer, Vajravelu et al. [15]
developed a one-dimensional boundary integral equation to analyze the free heat transfer in a viscous, electrically conducting
and heat-generating uid past a vertical porous plate in the presence of free stream oscillations with the use of time-dependent
fundamental solutions.
Recently, a set of general boundary-domain integral equations
for solving full NavierStokes equations using primitive variables
was proposed by Gao [24]. The BEM formulations are expressed
in terms of velocity, traction, and pressure, and are valid for steady,
unsteady, compressible, and incompressible ows. Although these
integral equations are derived under isothermal conditions, they
can also be applied to analyze non-isothermal viscous uids. However, as done in this paper, an integral equation from the energy
equation is required to close the integral equation system.
This paper presents a new boundary-domain integral equation
for computing temperatures, which is derived from the full energy
equation without heat source rates. In this equation, velocities,
velocity gradients and pressures are included, which can be computed using the BEM formulations presented in Refs. [24]. Together with these integral equations, a complete set of general
boundary-domain integral equations for solving convective heat
transfer problems is established. These equations are valid for steady, unsteady, compressible, and incompressible viscous uid ows.
For compressible uids, the number of unknown variables is one
more than those provided by the integral equations. Therefore,
an equation of state is required to close the equation system. However, for incompressible viscous uid ows, the boundary-domain
integral equation set given in this paper is closed, and the temperature integral equation is uncoupled from the velocity and pressure integral equations. After velocities, velocity gradients, and
pressures are computed through evaluating the correspondent
integral equations, the temperatures can be solved by using the derived temperature integral equation. Details of the numerical
implementation involved are provided for steady incompressible
viscous ows. Finally two numerical examples are given to validate
the formulations derived in this paper.

wi rij uj  qEui

4b

where e is the internal energy per unit mass, and for the calorically
perfect gas (constant specic heats)

e cv T

where cv is the specic heat at constant volume. This equation is


sometimes labeled the caloric equation of state. For incompressible
viscous ows, cv = cp (here cp is the specic heat at constant pressure). For Newtonian uids, the constitutive relationship between
the stresses and velocities based on Stokes hypothesis can be expressed as:

rij pdij l



@ui @uj
2 @uk
 l

dij
3 @xk
@xj @xi

in which, p is the pressure, l the dynamics coefcient of viscosity


(constant here), and dij the Kronecker delta symbol.
On the uid surface with outward normal nj, the relationship
between the stress and the traction ti (force per unit area) can be
written as:

ti rij nj

Based on these equations, a complete system of boundary integral


equations for viscous uid mechanics can be well established.

3. Review of velocity boundary-domain integral equations for


viscous ows
The velocity boundary-domain integrals for 2D and 3D uid
ows formulated with respect to primary variables (velocities,
tractions and pressure) can be written as [2,3]:

ui x

2. Governing differential equations in viscous uid mechanics

Z
uij x; yt j ydCy  t ij x; yuj ydCy
C
C
Z

 uij x; ynk yqyuj yuk ydCy
ZC
uij;k x; yqyuj yuk ydXy
ZX
Z
uij;j x; ypydXy uij x; yqybj ydXy
X
ZX
@ quj
 uij x; y
dXy
@t
X

In order to analyze the uid ow and heat transfer in uid


mechanics, one needs to consider the conservation laws of the
mass, momentum and energy. After some operations on these conservation laws, the governing differential equations in viscous uid
mechanics can be expressed as:

where X denotes the domain of the problem and C the boundary of


the domain X. x is referred to as the source point and y as the eld
point, and ;k @ =yk . The fundamental solutions appearing in
integral Eq. (8) can be expressed as:

@ q @ qui

0
@t
@xi

uij x; y

@ qui @ qui uj @ rij

qbi
@t
@xj
@xj

tij x; y



@ qE
@
@T
@wi

k
qbi ui
@t
@xi
@xi
@xi

In the above equations, q is the uid density, t the time, ui the ith
velocity component, xi the ith component of spatial coordinates at
point x, bi the body force per unit mass (e.g. the gravity force), rij
the stress tensor, E the total energy per unit mass, T the temperature, k the thermal conductivity, and the repeated subscripts stand
for summation. Other quantities included in Eq. (3) can be determined using

E e ui ui =2

4a

1
f7dij
16apl

ln1r r;i r;j g for 2D

1
f7dij
16aplr

r;i r;j g

for 3D


1 
3ni r;j  nj r;i br;i r;j 3dij nk r;k
8apr a

10

uij;k x; y


1 
7dij r ;k  dik r ;j  djk r;i br;i r ;j r ;k
16aplr a

11

uij;j x; y

3r ;i
8aplr a

12

where a = b  1 with b = 2 for 2D and b = 3 for 3D problems, r is the


distance from the source point x to the eld point y, i.e. r = ky  xk,
and

r;i

@r
y  xi
i
@yi
r

13

185

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

Differentiating Eq. (8) with respect to the source points x and using
the singular integral separation technique, the velocity gradient
boundary-domain integral equations can be derived as follows [4]:
Z
Z
@ui x
uij;l x; ytj ydCy t ij;l x; yuj ydCy
@xl
C
C
Z
uij;l x; ynk yqyuj yuk ydCy
ZC
Z
@ quj
 uij;l x; yqybj ydXy uij;l x; y
dXy
@t
X
X
Z
Z
 uij;kl x; yqyuj yuk ydXy  uij;jl x; ypydXy
X

1
3
fb 7cdij dkl  2dik djl djk dil gqxuj xuk x
dil px
8bcl
4bl
14

in which c = b + 2 and the used fundamental solutions are

1 
3ni djl nl dij  nj dil b3nj r;i  ni r;j r;l
8par b

nl r;i r;j  br;m nm dil r;j djl r;i  3dij r;l  cr;i r;j r;l 

t ij;l x; y

uij;kl x;y

15


1
dil djk dik djl 7dij dkl
16aplrb
bdik r;j r;l djk r;i r;l 7dij r;k r;l djl r;i r;k dkl r;i r;j dil r;j r;k

bcr;i r;j r;k r;l
16
3
fdil  br;i r;l g
8aplr b

uij;jl x; y

17

Incorporating the indices i and l in Eq. (14) and substituting the


resulting velocity divergence boundary-domain integral equation
into continuity Eq. (1), the integral equation for evaluating the pressure p can written as [4]:

3
px
4l

Z
uij;i x; ytj ydCy  t ij;i x; yuj ydCy
C
C
Z

 uij;i x; ynk yqyuj yuk ydCy
ZC
uij;ki x; yqyuj yuk ydXy
ZX
Z
@ quj
uij;i x; yqybj ydXy  uij;i x; y
dXy
@t
X
X


3
1
@q @q

qxui xui x 
u x

4bl
qx i @xi @t

other integrals included in Eqs.(8), (14), and (18) are strongly or


weakly singular and are interpreted in the Cauchy principal value
sense. The singularity separation technique [16] is applied to
evaluate the domain integrals appearing in Eq. (14) for kernels
uij;kl and uij;jl , and in Eq. (18) for kernel uij;ki , since they are strongly
singular even for internal collocation points. For the weakly singular integrals, the element/cell sub-division techniques [16] are
used to remove the integral singularities.
Eqs. (8), (14), and (18) are general boundary-domain integral
equations for the analysis of viscous uid ows. For incompressible
ows, the pressures appearing in the discretized algebraic
equations from Eq. (8) can be eliminated by employing Eq. (18),
such that the nal system of equations only contains the velocities/tractions as unknowns. Using the results obtained by Eqs. (8)
and (18), the velocity gradients can be solved through the evaluation of integral Eq. (14). They are only applied to analyze isothermal viscous ows. In order to solve convective heat transfer
problems, the energy equation needs to be considered. For this
purpose, a new boundary-domain integral equation for evaluating
the temperature will be derived in the next section.
4. Temperature boundary-domain integral equation
In order to make the symbols appearing in this section consistent with those used in Eqs. (8), (14), and (18), let yi represent xi
in Eq. (3), and then the energy equation can be expressed as
follows:



@ qE
@
@T
@wi

k
qbi ui
@t
@yi
@yi
@yi

Multiplying both sides of the above equation by a weight function


T and integrating over the domain O bounded with the boundary
C, it is obtained that

T

uij;i x; y

19

3r ;j
8aplra

20

uij;ki x; y

3
8aplr



djk  br;j r;k
b

T

Z
X

21

When the eld points y approaches the source points x, the


fundamental solutions in Eqs. (9)(12), (15)(17), and (19)(21)
are singular. From the kernel functions t ij;l included in Eq. (15)
and tij;i in Eq. (19), it can be seen that the second boundary integrals in Eqs. (14) and (18) are hyper-singular and only exist in
the sense of the Hadamard nite-part. Therefore, integral
Eqs (14) and (18) are valid only for internal collocation points.
For boundary points, the traction-recovery technique as done in
literatures [3,16] is used to evaluate the velocity gradient and
pressure. Except for the above hyper-singular integrals, all the



Z
@
@T
@wi
dX T 
T
k
dX
@y
@y
@yi
X
X
i
i
Z
T  qbi ui dX
Z

23

Integrating the former two domain integrals on the right-hand side


of Eq. (23) by parts and using Gausss divergence theorem, it follows
that

where

3
fdjl  br;j r;l gnl
4apr b

@ qE
dX
@t

18

t ij;i x; y

22

T



Z
Z
Z
@
@T
@T
@T 
dX
k
T  k dC  Tk
dC Tk
@yi
@yi
@n
@n
C
C
X


@ @T 
dX

@yi @yi
@wi
dX
@yi

T  wi ni dC 

wi

@T 
dX
@yi

24a

24b

If the weight function T is the Greens function T(x, y) which satises the following equation:



@ @T  x; y
dy  x 0
@yi
@yi

25

where d(y  x) is the Dirac delta function of y at the point x, then in


terms of the integration property of the Dirac delta function [17],
the last domain integral of Eq. (24a) can be written as:

Z
X

Tyk



Z
@ @T 
dXy
Tykydy  xdXy kTx
@yi @yi
X
26

Substituting this equation into Eq. (24a) and the result together
with Eq. (24b) into Eq. (23), it follows that

186

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

T ;n x; ykTydCy
Z
Z
T  x; yni ywi ydCy  T ;i x; ywi ydXy
X
ZC
Z
@ qE

T x; yqybi yui ydXy  T  x; y
dXy
@t
X
X
27

kTx 

T  x; yqydCy 

where T ;n x; y @T  x; y=@n, T ;i x; y @T  x; y=@yi and q(y) is the


heat ux, i.e.,

qy k

@Ty
@n

28

The fundamental solution T and its derivatives in Eq. (25) can be


written as:

(


T x; y

1
2p

1
ln rx;y
for 2D

1
4prx;y

for 3D

1
T ;i x; y
r;i
2par a
T ;n x; y

1

r ;i ni
2par a

29



@ui @uj
2 @uk
uj  l

ui  p qui ui =2ui qcv ui T


3 xk
@xj @xi

5. Numerical implementation for steady incompressible ows


Eqs. (8), (14), (18), and (27) are general integral equations for
solving convective heat transfer problems. To be considered in this
section is the numerical implementation of these equations in
steady incompressible ows. In this case, the density q is constant
and the time-related domain integrals in Eqs. (8), (14), (18), and
(27)) disappear. The details of the numerical implementation of
integral Eqs. (8), (14), and (18) can be found in Refs. [2,3]. To
numerically implement Eq. (27), the boundary C and the domain
X of the problem should be discretized into boundary elements
and internal cells, respectively. For each boundary element or
internal cell, the temperature, heat ux and wi can be expressed
in terms of their nodal values through shape functions as follows:

X g g
N T

33a

30
q

X g
N qg

33b

31
wi

The thermal conductivity k is assumed to be constant in the derivation process of integral Eq. (27). If the thermal conductivity to be
considered is the spatial function of coordinates y, a domain integral will appear in the resulting integral equation [18]. As a result,
the complexity of the solution of convective heat transfer problems
will be raised. The varying thermal conductivity problems will be
treated with in the future research.
Eq. (27) is a general energy boundary-domain integral equation
valid for steady, unsteady, compressible, and incompressible viscous uids. It is bounded only for internal collocation points, since
the second integral associated with the kernel function T ;n x; y in
Eq. (27) is strongly singular when the source point x approaches
the eld point y. For boundary collocation points, the rigid body
motion strategy can be used to treat the strong singularity as in
conventional BEM [16]. Except for the second integral, all other
integrals in Eq. (27) are weakly singular and therefore can be evaluated accurately by using the element/cell sub-division technique
[16].
Eq. (27) together with Eqs. (8) and (18) described in the above
section constitute a boundary-domain integral equation system
for analyzing convective heat transfer problems using the primitive variables. For compressible viscous ows, since the density is
not constant, the number of unknowns being velocity/traction,
pressure, temperature, and density is one more than the number
of integral Eqs. (8), (18), and (27). Therefore, an equation of state,
such as p = qRT for perfect gases (here R is the specic gas constant), is required to close the nal system of equations. For incompressible viscous ows, the system of Eqs. (18), (18), and (27) is
closed. Since l is assumed constant as done in Section 2, integral
Eqs. (8) and (18) are uncoupled from the temperature integral
Eq. (27) and thus the velocity and pressure can be solved independently on the temperature, after which the temperature can be
solved from the integral Eq. (27). It is noted that the temperature
integral equation itself is not independent of the velocity and pressure, embodied in Eq. (27) through the term wi. Substituting Eq. (5)
into Eq. (4a), and the result together with Eq. (6) into Eq. (4b)
yields:

wi l

From Eq. (32), it can be seen that the quantity wi appearing in integral Eq. (27) includes the velocity, velocity gradient, pressure, and
temperature.

32

X g g
N wi

33c

where Ng are the shape functions for boundary elements and internal cells [16], and Tg represents the value of temperature T at node
g.
Substituting Eqs. (33) into the discretized form of Eq. (27) and
collocating x for all boundary nodes (the rigid body motion is used
for determination of diagonal terms) yield the following algebraic
matrix equation:

HfTg Gfqg fbg Eb fwg

34

where {T} and {q} are vectors consisting of temperatures and heat
uxes at all boundary nodes, respectively, and {b} is a constant vector from body forces. The vector {w} is given through the evaluation
of Eq. (32) or Eqs. (4) and (5) at all boundary and internal nodes. [Eb]
is the coefcient matrix determined by both the third boundary
integral and the fourth domain integral on the right-hand side of
Eq. (27).
At each boundary node, either temperature or heat ux is specied as a boundary condition. So after applying boundary conditions to Eq. (34) and rearranging the equation, it follows that

Ab fXg fY b g Eb fwg

35

where {X} is a vector consisting of unknown temperatures and unknown heat uxes, and {Yb} is a known vector. The square matrix
[Ab] is invertible and thereby

fXg fY b g Eb fwg

36

where

fY b g Ab 1 fY b g
Eb  Ab 1 Eb 

37

Similarly, for internal nodes, Eq. (27) gives

fT I g AI fXg fY I g EI fwg

38

where {TI} is the vector consisting of all temperatures at internal


nodes. Substituting Eq. (36) into Eq. (38) yields:

fT I g fY I g EI fwg
where

39

187

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

Table 1
The number of boundary elements, internal cells, boundary nodes, internal nodes and
total nodes in four BEM models.

Cell240
Cell320
Cell480
Cell640

Boundary
element

Internal
cell

Boundary
node

Internal
node

Total
node

76
96
92
112

240
320
480
640

76
96
92
112

203
273
435
585

279
369
527
697

Fig. 1. Geometry and boundary conditions for Poiseuille ow.

"
I

fY g fY I g AI fY g

40

EI  EI  AI Eb 

Upon close inspection of Eq. (32), it can be found that the vector
{w} appearing in Eqs. (36) and (39) consists of the constant values
and unknown temperatures over all nodes after the velocities,
velocity gradients, and pressures are computed using integral
equations (8), (14), and (18). Therefore, substituting the computational results into Eq. (32), the vector fwg can be written as:



fT b g
fwg fcg sB; Dt
fT I g

41

where {c} is a constant vector consisting of both the term qcvuiT


with known boundary temperatures and all other terms in Eq.
(32), and {Tb} is the vector consisting of unknown temperatures at
boundary nodes. The coefcient matrices [B] and [D] corresponding
to the vectors {Tb} and {TI}, respectively, are determined by the term
qcvui. It is noted that after using boundary conditions, the columns of matrix [B] corresponding to the known boundary temperature nodes should be zero. Substituting Eq. (41) into Eqs. (36) and
(39), it follows that

e bg
Ab fXg e
E b fT I g f Y

42

e Ig
E I fT I g f Y
AI fXg e

43

where

8
b
b
>
>
< A  I  E B
e
E b  Eb D
>
>
: eb
f Y g fY b g Eb fcg

44

8 I
I
>
< A  E B
e
E I  I  EI D
>
: I
e g fY I g EI fcg
fY

45

Eb
Ab  e
I
A  e
EI 

fXg
fT I g

e bg
fY
e Ig
fY

)
46

The above equation is the nal system of equations with boundary


unknowns {X} and internal temperatures {TI} as unknowns. Since
Eq. (46) is a linear equation about {X} and {TI}, it can be solved without iterative computation.
6. Numerical examples
To verify the correctness of the proposed method in this paper,
two numerical examples for two-dimensional steady incompressible ows are presented in this section. The rst one is the wellknown Poiseuille ow which has analytical solutions to verify,
and the second one is a 2D pipe ow with curvature and the comparison is made to FLUENT CFD results.
6.1. Poiseuille ow
It is assumed that the uid ows between two parallel plates
without body forces (Fig. 1).
The no-slip condition is applied to the upper and lower sides.
The left and right ends are insulated, while the upper and lower
sides are imposed with the temperatures of 40 and 10, respectively. Other boundary conditions applied are t x p 50; uy 0
on left end and tx p 0; uy 0 on right end. Therefore the viscous uid moves with a uniform x-component of velocity ux, and
the velocity and temperature vary only in the y-direction. Referring
to literature [19], the analytic solutions for the horizontal velocity
and temperature can be expressed as:

ux 

in above equations, [I] is the identity matrix.


Combining Eqs. (42) and (43) yields

#

T

p0
yH  y
2l

p2 2y  H4 15
p2 H4

2y  H 25
H
192lk
192lk

47a

47b

where H = 1 is the distance between the upper and lower sides,


p0 = dp/dx = 10 the gradient of pressure, k = 1 is the thermal conductivity and l = 1.

Fig. 2. BEM models with the different number of internal cells: (a) 240; (b) 320; (c) 480; (d) 640.

188

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

Table 2
Computed temperatures along y-direction at x = 2.5.
y

0.125

0.25

0.375

0.5

0.625

0.75

0.875

Cell240
Cell320
Cell480
Cell640
Analytical

14.118778
14.119476
14.108114
14.108603
14.106038

18.010649
18.011593
17.992470
17.993120
17.988281

21.797131
21.798141
21.774462
21.775145
21.768799

25.551184
25.552202
25.527060
25.527737
25.520833

29.297131
29.298143
29.274578
29.275251
29.268799

33.010651
33.011597
32.992710
32.993340
32.988281

36.618771
36.619469
36.608498
36.608940
36.606038

Table 3
Relative errors (%) between the numerical results and analytical solutions along ydirection at x = 2.5.
y

0.125

0.25

0.375

0.5

0.625

0.75

0.875

Cell240
Cell320
Cell480
Cell640

0.0903
0.0953
0.0147
0.0182

0.1243
0.1296
0.0233
0.0269

0.1301
0.1348
0.0260
0.0292

0.1189
0.1229
0.0244
0.0271

0.0968
0.1003
0.0197
0.0220

0.0678
0.0707
0.0134
0.0153

0.0347
0.0367
0.0067
0.0079

Fig. 5. BEM model for 2D pipe with curved part.

35

Temperature

Fig. 3. Computed temperatures along y-direction at x = 2.5.

Current

30

Fluent

25

20

15

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Length of centreline
Fig. 6. Computational temperature along the center line of the pipe from the inlet
to the outlet.

Fig. 4. Geometry and boundary conditions for 2D pipe ow.

To use the formulations presented in the paper, the computational domain and its boundary are required to be discretized into
the internal cells and boundary elements, respectively. As shown in
Fig. 2, four BEM models with the different number of internal cells
(240, 320, 480 and 640) are considered to examine the inuences
of the number of internal cells on the computational results. The
number of boundary elements, internal cells, boundary nodes,
internal nodes, and total nodes for each BEM model is given in Table 1. Table 2 lists the computed temperatures and analytical solutions along y-direction of x = 2.5, the relative errors of which are
shown in Table 3. Fig. 3 is a plot of the temperature prole. The

189

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

0.3

0.4

0.5

Y-coordinate

2426
22

Fig. 7. Computational temperature on the outlet of 2D pipe.

30

20

Temperature

Current
25

12

Fluent

34
32 30

14

28
26
24 22
16

20

18

16

14
12

0.2

20

0.1

18

10
0.0

2 8 30 3 2

15

12

28 30

20

2426
22
18

34

25

36

Temperature

Fluent

20

32

Current
30

14 1 6

35

16
14

12

20

Fig. 9. Contour plot of temperature computed by the current method.

15

The exact solutions for the heat ux q on the upper and lower
sides of the uid are 25.8333 and 34.1667, respectively, and the
computed results using Cell640 BEM model are 25.8657 and
34.1359. The relative errors are 0.13% and 0.09%, respectively.

10
0.0

0.1

0.2

0.3

0.4

0.5

Length of line L
Fig. 8. Computational temperature along the line L.

6.2. 2D angled pipe ow

computational results for velocity are not given in this paper as detailed description can be found in Ref. [2]. It is noted that the problem is independent of the density q and the specic heat at
constant volume cv (see the analytical solution (47)).
From Tables 2 and 3 and Fig. 3, it can be seen that the computational results with the use of 240, 320, 480, and 640 internal cells
are in good agreement with the analytical solutions. The comparison of the results denoted by Cell240 and Cell480 to the analytical
solutions shows that the results of 480 internal cells is closer to
the analytical solutions than that of 240 internal cells. However,
from the results denoted by Cell240 and Cell320, we can nd that
the relative errors between the results of 240 cells and analytical
solutions are smaller than that of Cell320 results. These mean that
the ning of the internal cells along y-direction (Fig. 2(a) and (c))
can make the numerical results more accurate, while the mesh
renement along x-direction improves computational accuracy
very little (see Fig. 2(a) and (b)). The phenomenon is mainly due
to the uid velocity and temperature varying along y-direction,
not along x-direction.

The second example deals with a convective heat transfer over


2D pipe with a curved part (Fig. 4). The curvature of the curved part
is determined by a radius of R = 0.5. The uid with the property of
l = 1, q = 1 and cv = 0.4930966 is subjected to a pressure of p = 50
on the inlet of the pipe and is pressure free on the outlet. The pipe
is approximated using 140 linear boundary elements with 140
boundary nodes and 600 internal cells with 531 internal nodes
(Fig. 5). The boundary conditions applied are tx = 0, ty = 50,
T = 30 on the inlet of the pipe, tx = ty = 0, q = 0 on the outlet, and
ux = uy = 0, q = 50 on the upper wall and ux = uy = 0, T = 10 on the
lower wall.
Fig. 6 shows the computed temperatures along the center line
of the pipe from the inlet to the outlet. Figs. 7 and 8 are the distributions of temperature along the y-axis on the outlet of the pipe
and line L (see Fig. 4), respectively, and specic results are listed
in Table 4, along with the relative errors. For the purpose of verication, this problem is also computed using the commercial software FLUENT, where a ne mesh with 2000 quadrilateral

Table 4
Temperatures on the outlet of pipe and the line L (see Fig. 4).
Distance

0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5

Outlet

Line L

Current

FLUENT

Error (%)

Current

FLUENT

Error (%)

12.4586
14.9283
17.4042
19.8831
22.3658
24.8533
27.3456
29.8417
32.3413
34.7998

12.5502
15.0785
1735955
20.1103
22.6288
25.1536
27.6841
30.1266
32.7439
34.9444

0.73
0.80
1.09
1.13
1.16
1.19
1.22
1.24
1.20
0.41

11.5245
13.1173
14.7776
16.5162
18.3448
20.2757
22.3215
24.4935
26.7990
29.2373

11.5275
13.1123
14.7654
16.4977
18.3206
20.2463
22.2871
24.4546
26.7568
29.1906

0.026
0.038
0.083
0.112
0.132
0.145
0.154
0.159
0.158
0.160

30
26
24
22
18
32

30
28

14

16

26
24 22
20

12

20

28

2426
22

18

14 16

20

28

34

12

14 16 2
8

X.-W. Gao et al. / International Journal of Heat and Mass Transfer 63 (2013) 183190

32

190

18

14

16

12

Fig. 10. Contour plot of temperature computed by FLUENT.

elements and 2121 nodes is used and pressure boundary conditions are applied on the inlet and outlet. The computational results
using FLUENT are also shown in Figs. 68 and Table 4. Figs. 9 and
10 are the contour plots of the temperature obtained by BEM and
FLUENT, respectively.
From Figs. 68 and Table 4, it can be seen that the BEM results
are close to the results by FLUENT with a good accuracy. However,
there is a small discrepancy between these two results near the
outlet of the pipe. Table 4 shows the maximum error on the outlet
of 2D pipe being 1.24%.
7. Conclusions
A novel boundary-domain integral equation for computing the
temperature in viscous uids is presented based on the conservation form of the energy equation. The derived formulation is general and applicable to steady, unsteady, compressible, and
incompressible ows. Together with the formulations for the
velocity, pressure, and velocity gradient proposed in Refs. [24],
a boundary-domain integral equation system for solving convective heat transfer problems is established. For incompressible
ows, the system of equations is closed and the velocity and pressure integral equations are uncoupled from the temperature integral equation. However, the temperature integral equation
involves the velocity, pressure, and velocity gradient, which need

to be computed in advance. Two numerical examples for convective heat transfer in steady incompressible uids have been provided to demonstrate the correctness of the derived formulations.
Acknowledgements
The authors gratefully acknowledge the National Natural Science Foundation of China under Grant NSFC Nos. 10872050 and
11172055 for nancial supports to this work.

References
[1] A. Bourchtein, Exact solution of the generalized NavierStokes equations for
benchmarking, Int. J. Numer. Methods Fluids 39 (2002) 10531071.
[2] X.W. Gao, A boundary-domain integral equation method in viscous uid ow,
Int. J. Numer. Methods Fluids 45 (2004) 463484.
[3] X.W. Gao, A promising boundary element formulation for three-dimensional
viscous ow, Int. J. Numer. Methods Fluids 47 (2005) 1943.
[4] X.W. Gao, Explicit formulations for evaluation of velocity gradients using
boundary-domain integral equations in 2D and 3D viscous ows, Int. J. Numer.
Methods Fluids 54 (2007) 13511368.
[5] M. Ikeuchi, K. Onishi, Boundary element solutions to steady convective
diffusion equations, Appl. Math. Model. 7 (1983) 115118.
[6] Y. Shi, P.K. Banerjee, Boundary element methods for convective heat transfer,
Comput. Methods Appl. Mech. Eng. 105 (1993) 261284.
[7] M.M. Grigoriev, G.F. Dargush, Boundary element methods for transient
convective diffusion. Part I: General formulations and 1D implementation,
Comput. Methods Appl. Mech. Eng. 192 (2003) 42814298.
[8] M.M. Grigoriev, G.F. Dargush, Boundary element methods for transient
convective diffusion. Part II. 2D implementation, Comput. Methods Appl.
Mech. Eng. 192 (2003) 42994312.
[9] M.M. Grigoriev, G.F. Dargush, Boundary element methods for transient
convective diffusion. Part III: Numerical examples, Comput. Methods Appl.
Mech. Eng. 192 (2003) 43134335.
[10] G.F. Dargush, P.K. Banerjee, A boundary element method for steady
incompressible thermoviscous ow, Int. J. Numer. Methods Eng. 31 (1991)
16051626.
[11] P.K. Banerjee, K.A. Honkala, Development of BEM for thermoviscous ow,
Comput. Methods Appl. Mech. Eng. 151 (1998) 4362.
[12] K. Onishi, T. Kuroki, M. Tanaka, An application of a boundary element method
to natural convection, Appl. Math. Model. 8 (1984) 383390.
[13] N. Tosaka, N. Fukushima, Integral Equation Analysis of Laminar Natural
Convection Problems, BEM VIII, Computation Mechanics Publication, SpringerVerlag, Southampton, Berlin, 1986.
[14] C.P. Rahaim, A.J. Kassab, Pressure correction DRBEM solution for heat transfer
and uid ow in incompressible viscous uids, Eng. Anal. Bound. Elem. 18
(1996) 265272.
[15] K. Vajravelu, A. Kassab, A. Hadjinicolaou, BEM solution to transient free
convection heat transfer in a viscous, electrically conducting, and heat
generating uid, Numer. Heat Transfer Part A: Appl. 30 (1996) 555574.
[16] X.W. Gao, T.G. Davies, Boundary Element Programming in Mechanics,
Cambridge University Press, Cambridge, 2002.
[17] M.D. Greenberg, Application of Greens Functions in Science and Engineering,
Prentice-Hall, Englewood Cliffs, New Jersey, 1971.
[18] X.W. Gao, A meshless BEM for isotropic heat conduction problems with heat
generation and spatially varying conductivity, Int. J. Numer. Methods Eng. 66
(2006) 14111431.
[19] R.A. Granger, Fluid Mechanics, Dover, New York, 1995.

Potrebbero piacerti anche