Sunteți pe pagina 1din 14

A computational study of convective heat transfer to carbon dioxide

at a pressure just above the critical value


S. He
a,
*
, W.S. Kim
a
, J.D. Jackson
b
a
School of Engineering, University of Aberdeen, Aberdeen AB24 3UE, UK
b
Simon Building, University of Manchester, Manchester M13 9PL, UK
Received 29 January 2007; accepted 2 November 2007
Available online 7 November 2007
Abstract
Computational simulations are reported of experiments on convective heat transfer to carbon dioxide at a pressure of 75.8 bar, which
is just above the thermodynamic critical value of 73.8 bar. These have been carried out using a variable property, elliptic computational
formulation incorporating low Reynolds number turbulence models of k e and V2F types. Firstly, the simulations were compared with
the heat transfer measurements and then they were used in developing an understanding of interesting phenomena observed in the exper-
iments. It has been found that the eect of buoyancy on turbulence production and heat transfer in uids at supercritical pressure can be
very signicant even under conditions of relatively low buoyancy parameter based on bulk properties. The eect of buoyancy, although
complex, can be explained by relating it to the large-property-variation (LPV) region, i.e., the region within the ow eld near to the
locations where the uid temperature has the pseudo-critical value. Under certain conditions, a very non-uniform radial distribution
of the buoyancy force may be present and cause some reduction of turbulence in the core but a big increase near the wall, resulting
in much improved heat transfer. It is clear that new heat transfer correlations are needed to account for such eects on heat transfer
to supercritical pressure uids as they come to be used more and more in new energy systems applications such as, advanced water-cooled
nuclear reactors, environmentally friendly air-conditioning and refrigeration systems and high pressure water oxidation plant for waste
processing.
2007 Elsevier Ltd. All rights reserved.
Keywords: Supercritical pressure; Mixed convection; Buoyancy inuence; Computational modelling; Energy systems
1. Introduction
Considerable interest in heat transfer to uids at pres-
sures above the thermodynamic critical value was stimu-
lated during the 1950s and 1960s by the introduction of
fossil-fuelled power plant water steam generators operating
at supercritical pressure. A comprehensive review of such
studies was reported by Jackson and Hall [1]. Recently,
there has been renewed interest in the topic driven by active
consideration of new applications involving uids at super-
critical pressure. These include supercritical pressure water
oxidation systems for waste processing [2], the use of car-
bon dioxide at supercritical pressure in a new generation
of air-conditioning systems for cars and refrigeration sys-
tems [3], the development of very compact gas coolers
and internal heat exchangers [4,5], liquid hydrogenoxygen
fuelled rockets [6] and the development of supercritical
pressure water-cooled nuclear reactors which can compete
in terms of cost, safety and reliability with other types of
power generation systems [79].
The critical temperature and pressure of carbon dioxide
are 31.1 C and 7.38 MPa, respectively. When the pressure
of a uid is below the critical value, phase change may
occur as its temperature or pressure is varied. Above the
critical pressure a uid does not undergo phase change.
However, the physical properties of such uids can vary
1359-4311/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2007.11.001
*
Corresponding author. Tel.: +44 1224 272799; fax: +44 1224 272497.
E-mail address: s.he@abdn.ac.uk (S. He).
www.elsevier.com/locate/apthermeng
Available online at www.sciencedirect.com
Applied Thermal Engineering 28 (2008) 16621675
sharply with temperature at constant pressure in the vicin-
ity of the pseudo-critical value, T
pc
, which is dened as
the temperature where its specic heat at constant pressure
reaches a peak value for the pressure under consider-
ation. Fig. 1 shows the phase diagram and variations of
some of the properties of carbon dioxide above the critical
pressure.
The important characteristics of uids at supercritical
pressure which makes them of particular interest from
the point of view of heat transfer is that their physical
properties vary rapidly with both pressure and tempera-
ture, as seen from Fig. 1b. For example, if the temperature
changes by 0.5 C near the pseudo-critical temperature, the
specic heat (c
p
) may change by up to 600%; other uid
properties, such as, thermal conductivity (k), density (q)
and viscosity (l) change by 50% for CO
2
at 7.58 MPa.
The specic heat variation becomes increasingly sensitive
to pressure near the critical value as can be seen from
Fig. 1c. Its peak value increases rapidly as the critical pres-
sure is approached. Note that the pseudo-critical tempera-
ture rises systematically with increase of pressure.
The strong variation of uid properties with tempera-
ture, especially when coupled with inuences of buoyancy
or thermally-induced ow acceleration, can cause very sig-
nicant localised distortions of the mean ow and turbu-
lence elds, resulting in eects such as laminarization
(reversed transition) [10].
In applications involving heat transfer to uids at super-
critical pressure, the ow passages are often vertical. Lab-
oratory experiments to study such ows have usually
been carried out using tubes or annuli. A wide range of
ow and thermal conditions have been covered with dier-
ent pipe diameters and test section orientation, see, for
example [4,5,1114] and also review papers [10,15]. A vari-
ety of uids have been used including water, carbon diox-
ide, freon, hydrogen, helium, and nitrogen, although the
collections of data for water and carbon dioxide are by
far the most extensive. It is noted that measurements in
most studies have mainly been limited to wall temperatures
(and sometimes pressure drops) for specied values of heat
ux and mass ow rate. Detailed information on uid tem-
perature, velocity and turbulence has rarely been obtained
due to the diculties and costs associated with experimen-
tal work on uids at the high pressures. This has limited the
development of the understanding of heat transfer at super-
critical pressure to some extent. It has long been recognised
that computational modelling studies have the potential to
play a very useful role in improving our understanding of
the physics of turbulent ow and heat transfer and assisting
with developing correlation equations for engineering
applications. A number of such investigations have been
undertaken. Earlier ones (pre-eighties) used relatively sim-
ple mixing-length and eddy diusivity type turbulence
models. Some later studies used more advanced models,
including a number of low Reynolds number type
[2,6,7,16] and in one case a re-normalised group (RNG)
k e model [17].
The present paper is concerned with computational sim-
ulations of some experiments on heat transfer to carbon
dioxide carried out by Fewster [11], see Fewster and Jack-
son [12]. These were conducted at a pressure of 7.58 MPa
which is just above the critical value of 7.38 MPa. As a
result, the uid properties varied sharply with temperature.
Striking features were observed in the wall temperature dis-
tributions which were not easy to explain using the limited
information available from the measurements. The purpose
Nomenclature
Bo buoyancy parameter (Bo = Gr/(Re
3.425
Pr
0.8
))
c
p
specic heat capacity (J kg
1
K
1
)
D pipe diameter (m) (or a damping function in tur-
bulence models as dened in Table 4)
E damping function dened in Table 4
Gr Grashof number (Gr bgD
4
q
00
w
=km
2
)
h heat transfer coecient (q
w
/(T
w
T
b
), enthalpy
(J kg
1
)
k turbulent kinetic energy (m
2
s
2
)
_ m mass ow rate (kg s
1
)
Nu Nusselt number (Nu = hD/k)
Pr Prandtl number (Pr = lc
p
/k)
q
00
w
wall heat ux (W m
2
)
x,r axial and radial coordinates (m)
Re Reynolds number (Re = U
b
D/m)
T temperature (C)
U,V velocity components in the x, r-directions
y distance from the inside surface of the tube nor-
mal to it (m)
y
+
non-dimensional distance from pipe wall, y(s
w
/
q)
1/2
/m
Greek letters
e rate of dissipation of k (m
2
s
3
)
k thermal conductivity (W m
1
K
1
)
l molecular viscosity (kg m
1
s
1
)
l
t
turbulent viscosity (kg m
1
s
1
)
t kinematic viscosity, m = l/q (m
2
s
1
)
q density of uid (kg m
3
)
Subscripts
b bulk
w wall
pc pseudo-critical
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1663
of the work presented here is to study the interesting
behaviour exhibited by these experimental results using
detailed information on the mean ow and thermal elds
provided by the computational simulations.
Temperature (
o
C)
P
r
e
s
s
u
r
e

(
M
P
a
)
Solid
Liquid
Gas
Supercritical
Critical Point
7.38
Experimental condition (P=7.58 MPa, T=0-300
o
C)
31.1 -56.4
0.5
0
(a) Phase diagram for CO
2
0
0.2
0.4
0.6
0.8
1
1.2
1.4
10 15 20 25 30 35 40 45 50
T (
o
C)


[
1
0
3
(
k
g
/
m
3
)
]
,

C
p

[
1
0
5
(
J
/
k
g
K
)
]
,
k

[
1
0
-
1
(
W
/
m
K
)
]
,


[
1
0
-
4
(
k
g
/
m
s
)
]
Conductivity
Density
Viscosity
Specific heat
(b) Properties near pseudo-critical temperature at 7.58Mpa
0
20
40
60
20 25 30 35 40 45 50
Temperature (
o
C)
S
p
e
c
i
f
i
c

h
e
a
t

(
k
J
/
k
g
-
C
)
8.0 MPa
9.5 MPa
7.6 MPa
(c) Pressure dependence
Fig. 1. Carbon dioxide phase diagram and uid property variations near the pseudo-critical temperature condition.
1664 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
2. Special features of heat transfer to uids at supercritical
pressure
2.1. Inuence of property variation
As a consequence of the strong dependence of physical
properties on temperature, the process of heat transfer in
a uid at a pressure just above the critical value is generally
more complex than for ordinary uids. Heat transfer coef-
cients often vary signicantly with axial location along a
heated passage due to the non-uniformity of physical prop-
erties, especially where the uid temperature is near the
pseudo-critical value. In order to illustrate this, the varia-
tion of heat transfer coecient (htc) with bulk temperature
at a owrate of 0.041 kg/s along a tube of diameter 0.01 m
is shown in Fig. 2. The htc was calculated using the Dittus
Boelter heat transfer correlation:
Nu 0:023 Re
0:8
Pr
0:4
1
which was reorganized to give
h
0:028 _ m
0:8
D
1:8
_ _
k
0:6
l
0:4
_ _
c
0:4
p
_ _
2
It can be seen from Fig. 2 that htc peaks at the pseudo-
critical temperature (32.2 C). It increases by almost ve
times as the bulk uid temperature varies by only a few de-
grees. It is interesting to note that the value of the htc is
similar on either side of T
pc
, although the other properties
involved fall signicantly as uid temperature varies from
below T
pc
to above it. Clearly there is a cancelling eect. It
is clear that the peak value of htc is largely due to the
sharp rise in specic heat near the pseudo-critical temper-
ature. The eect of the variation of uid properties with
temperature in a real ow system can be much more com-
plex. This will be demonstrated to some extent later in the
present paper.
2.2. Inuence of buoyancy
The inuence of buoyancy can cause the eectiveness of
heat transfer to be either improved or impaired, through
enhancement or suppression of the production of turbu-
lence. There are two basic mechanisms [18] by which buoy-
ancy can modify turbulence, namely, the direct (structural)
eect, through buoyancy-induced production of turbulence
kinetic energy, and the indirect (external) eect, through
the modication of the mean ow.
For mixed convection in a vertical channel, the indirect
eect is normally the dominant one (see later). In such a sit-
uation, the semi-empirical theory of Jackson and Hall [1]
can be used to explain mixed convection heat transfer
behaviour and to correlate experimental data. For buoy-
ancy-opposed ows (downward ow in a heated passage),
onset of inuence of buoyancy causes the velocity to reduce
near the wall and to increase in other regions. This modi-
cation of the mean ow eld causes an enhancement of tur-
bulence production and, as a result, turbulent diusion is
improved and the eectiveness of heat transfer is improved.
In the buoyancy-aided case (upward ow in a heated pas-
sage), the inuence of buoyancy causes a reduction in the
velocity gradient over most of the ow, except in the region
very close to the wall. As a result, turbulence production is
reduced and the turbulent diusion of heat is impaired. If
the inuence of buoyancy is progressively increased, by
reducing the ow rate and/or increasing the heating rate,
the impairment of turbulence production and the deterio-
ration of heat transfer become more and more marked. A
stage is reached where turbulence production in the near-
wall region virtually ceases. This is sometimes described
as laminarization of the ow (or reversed transition). With
further increase of buoyancy inuence, negative values of
shear stress are generated in the core region and turbulence
can then be readily produced there. Consequently, the
eectiveness of heat transfer recovers. As the buoyancy
inuence becomes increasingly stronger, heat transfer can
become more eective than under conditions of forced con-
vection in the absence of buoyancy inuence. It has been
established (see, for instance [19]) that with normal uids
such as air or water at atmospheric pressure the eect of
buoyancy inuence is negligible if the following criterion
is satised:
Bo
Gr
Re
3:425
Pr
0:8
< 5:6 10
7
3
2.3. Inuence of ow acceleration
Turbulent heat transfer in channels can also be signi-
cantly aected by bulk ow acceleration caused by thermal
expansion of the uid due to strong heating (particularly in
the case of gases). As in the case of acceleration caused by
other means (spatial or temporal) the velocity prole is
modied in such a way that turbulence production is sup-
pressed and turbulent diusion is impaired [16].
0
10
20
30
40
50
60
28 29 30 31 32 33 34 35 36
T (

C)
H
e
a
t

t
r
a
n
s
f
e
r

c
o
e
f
f
i
c
i
e
n
t
,

h

(
W
/
m
2
-
C
)
From Dittus-Boelter equation
Mass flow rate 0.041 kg/s
Pressure 7.58 MPa
Fig. 2. Eect of uid properties on heat transfer coecient under
conditions of forced convection.
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1665
3. Experimental investigation
In the study reported here, computational simulations
were performed of selected experiments from a comprehen-
sive experimental investigation of mixed convection heat
transfer to carbon dioxide owing through a vertical
heated tube reported by Fewster and Jackson [12]. One
of the test sections used in that investigation was made
from a thin-walled circular tube of internal diameter of
5 mm. It had an unheated ow development section of
length to diameter ratio 75 followed by a uniformly heated
section of length to diameter ratio 150. The test section was
instrumented with chromelalumel thermocouples welded
to the outer surface of the tube at axial locations approxi-
mately one tube diameter apart along the heated length.
The ow circuit was designed so that the uid could pass
either upwards or downwards through the test section. In
the experiments that have been simulated in the present
study, values of mass ow rate were chosen to give Rey-
nolds numbers of 2 10
5
and 4 10
4
. At the highest value
of mass ow rate, the wall temperature distributions for
conditions of upward and downward ow turned out to
be identical. Thus, the eect of buoyancy was negligibly
small, in spite of the large non-uniformities of density
(and other uid properties) present within the uid. When
the mass ow rate was reduced by a factor of two (giving a
Reynolds number of about 9 10
4
) clear dierences
between the wall temperature distributions for upward
and downward ow were found in the downstream region.
With upward ow a condition of impaired heat transfer
developed. Further reduction in ow rate, again by a factor
of two (giving a Reynolds number of about 4 10
4
) caused
a sharp peak in wall temperature to form in the upward
ow case, indicating strong localized deterioration of heat
transfer. In contrast, the distribution of wall temperature
at this Reynolds number in the downward ow case was
rather similar in form to that observed at the higher ow
rates. Table 1 shows the experimental conditions of the
cases simulated in the present paper.
4. Numerical modelling
The computational study was conducted using an in-
house CFD code named SWIRL. It solves the transport
equations for both the mean ow and turbulence in a cylin-
drical coordinate system using a widely used nite volume
scheme [20]. The ow is considered to be axisymmetric, and
the Reynolds-averaged equations for such a system written
in cylindrical coordinates are:
Continuity
1
r
o
ox
qrU
o
or
qrV
_ _
0 4
U-momentum
1
r
o
ox
qrU
2

o
or
qrVU
_ _

op
ox
qg
1
r
2
o
ox
rl
e
oU
ox
_ _ _ _ _

o
or
rl
e
oU
or

oV
ox
_ _ _ __
5
V-momentum
1
r
o
ox
qrUV
o
or
qrV
2

_ _

op
or

1
r
o
ox
rl
e
oV
ox

oU
or
_ _ _ _ _
2
o
or
rl
e
oV
or
_ _ _ __
2
l
e
V
r
2
6
in which q is density and l
e
is eective viscosity, dened by
l
e
= l + l
T
, and l and l
T
are the molecular and turbulent
viscosities, respectively, the latter resulting from the model-
ling of turbulent stresses by the mean to be described below
as the turbulence models are introduced.
The energy equation written in terms of enthalpy, which
is an appropriate form for the case where there is a strong
dependence of uid properties on temperature, is:
Energy
1
r
o
ox
qrUh
o
or
qrVh
_ _

1
r
o
ox
r
l
Pr

l
T
r
T
_ _
oh
ox
_ _

o
or
r
l
Pr

l
T
r
T
_ _
oh
or
_ _ _ _
7
in which Pr is the molecular Prandtl number and r
T
the
turbulent Prandtl number which, in keeping with widely
employed practice, was assigned a value of 0.9. A number
of low Reynolds number, eddy viscosity turbulence models
were incorporated into the computational formulation.
Modelling results will be reported here from simulations
using the AKN (Abe, Kondoh and Nagano, 1994) model
[21] and the V2F (Behnia, Parneix, and Durbin, 1998)
model [22]. The constitutive and transport equations in
the case of the AKN and V2F models can be written as:
AKN turbulence model [21]
Turbulent kinetic energy (k)
oqUk
ox

1
r
orqVk
or
_ _

o
ox
l
l
t
r
k
_ _
ok
ox
_ _

1
r
o
or
r l
u
t
r
k
_ _
ok
or
_ _
P
k
G
k
qe qD 8
Table 1
Experimental conditions
Cases Tin (C) Re (in) q
w
(W/m
2
) Bo (in) X (in)
Run 1 13.2 187,950 318,000 8.05E9 1.26E3
Run 2 20.5 44,046 68,000 4.28E7 1.64E3
Notes: (1) Buoyancy parameter, Bo
Gr
Re
3:425
Pr
0:8
[1], where Gr
gbD
4
q
00
w
km
2
(2) Acceleration parameter, X
4bq
00
w
qc
p
U
b
[16].
1666 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
Dissipation rate (e)
oqUe
ox

1
r
orqV e
or
_ _

o
ox
l
l
t
r
e
_ _
oe
ox
_ _

1
r
o
or
r l
l
t
r
e
_ _
oe
or
_ _
C
e1
f
1
1
T
P
k
C
e1
f
1
1
T
G
k
C
e2
f
2
qe
T
qE 9
where
l
t
qC
l
f
l
kT; where T k=e
P
k
l
t
2
oU
ox
_ _
2

oV
or
_ _
2

V
r
_ _
2
_ _

oU
or

oV
ox
_ _
2
_ _
shear production 10
G
k
q
0
u
0
g
x

bl
t
C
1t
k
e
_ _
oU
or

oV
ox
_ _
oT
or
_ _
g
x
gravitational production 11
where g
x
is the acceleration due to gravity in the x-direc-
tion, being, +g or g for downward and upward ows,
respectively.
V2F turbulence model [22]
Turbulent kinetic energy (k)
oqUk
ox

1
r
orqVk
or
_ _

o
ox
l
l
t
r
k
_ _
ok
ox
_ _

1
r
o
or
r l
l
t
r
k
_ _
ok
or
_ _
P
k
G
k
qe qD 12
Dissipation rate (e)
oqUe
ox

1
r
orqV e
or
_ _

o
ox
l
l
t
r
e
_ _
oe
ox
_ _

1
r
o
or
r l
l
t
r
e
_ _
oe
or
_ _
C
e1
f
1
1
T
P
k
C
e1
f
1
1
T
G
k
C
e2
f
2
qe
T
qE 13
Turbulent velocity scale v
2

oqUv
2

ox

1
r
orqV v
2

or
_ _

o
ox
l
l
t
r
k
_ _
ov
2
ox
_ _

1
r
o
or
r l
l
t
r
k
_ _
ov
2
or
_ _
qkf 6qv
2
e
k
14
and the elliptic relaxation equation for f
0
o
ox
of
ox
_ _

1
r
o
or
r
of
or
_ _

1
L
2
f
C
1
1
L
2
2=3 v
2
=k
T

C
2
L
2
1
qk
P
k
G
k

1
L
2
5v
2
=k
T
15
where P
k
and G
k
are dened using Eqs. (10) and (11)
respectively.
l
t
qC
l
v
2
T; where T max
k
e
; 6

m
e
_ _ _
;
L C
L
max
k
3=2
e
; C
g
m
3
e
_ _
1=4
_ _
C
1
1:4; C
2
0:3; C
g
70; C
L
0:23:
Other model constants (C
l
, C
e1
, C
e2
, r
k
, r
e
), damping func-
tions (f
l
, f
1
, f
2
, D, E) and boundary conditions at the wall
are specied in Tables 13.
The complete computational domain, which covered the
whole unheated and heated lengths of the test section and
ranged from the centre of the tube to the inner wall, was
discretized into a mesh of grids, typically, 120 106
(axial radial). The mesh was rened in the radial direction
towards the tube wall. It was also rened in the axial direc-
tion in the region where the heating commenced. The mesh
was adjusted in each individual run to ensure that the near-
wall ow features were properly resolved and that the y
+
value at the rst node of the mesh was always less than 0.5.
The staggered grid arrangement was used to dene the
variables. The scalar parameters were dened at the grid
points and the velocity components were dened on the
control volume surfaces. The QUICK scheme was used
for approximating the convection terms in the momentum
equations and the SMART scheme was used for other
transport equations for reasons of numerical stability.
The SIMPLE scheme was used for coupling the pressure
and the velocity elds. The resultant ve-point coecient
matrix system was solved iteratively using the line-by-line
algorithm TDMA; at any time, variables at a particular
line were solved simultaneously; variables at the neighbour-
ing lines were assumed to be known and values from the
previous iteration were used. The TDMA was used alter-
nately in the axial and radial directions following the alter-
nating-direction-implicit (ADI) approach to accelerate
convergence.
The NIST Standard Reference Database 23 (REF-
PROP) Version 7 was used for calculating the temperature
and pressure dependent properties of carbon dioxide. To
do this, the relevant FORTRAN subroutines supplied with
the Database were incorporated in the CFD code SWIRL.
Table 2
Constants in the turbulence models
Model Code C
l
C
e1
C
e2
r
k
r
e
Abe-Kondoh-Nagano (1994) [20] AKN 0.09 1.50 1.90 1.4 1.4
V2F (1998) [21] V2F 0.22 1.4 1.9 1.0 1.3
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1667
They were used to generate a property table at the begin-
ning of each run with temperature intervals 5 10
4
C
and pressure intervals 1 10
4
Pa. During the iteration,
the properties were updated by interpolation using the
table. This approach was adopted because it was found
that the calculation was unacceptably slow if the NIST sub-
routines were directly used in each iteration.
5. Computational results and discussion
5.1. Wall temperature and Nusselt number
Fig. 3 shows the axial variation of wall and bulk temper-
atures in Run 1 where the Reynolds number (Re) is high
and buoyancy parameter (Bo) is low. It can be seen from
the gure that the measured wall temperatures for the
upward and downward ow cases collapsed onto the same
curve. This conrms that the inuence of buoyancy was
negligible. Also, both the AKN and V2F models give
results which clearly reect this fact. However, the wall
temperatures predicted by both of the models are some-
what higher than the measured values.
Fig. 4 shows comparisons between the predictions and
the measurements of wall temperature for Run 2 with
upward and downward ow. Fig. 5 shows the correspond-
ing variations of predicted Nusselt number. It can be seen
that the experimental wall temperature increases very
slowly along the tube in the downward ow case. However,
in the upward ow case, the wall temperature distribution
exhibits an abrupt increase with distance from the start of
heating. It peaks at an x/d of about 50, then reduces rap-
idly and reaches a value in the downstream region (x/
d > 70) which is only slightly higher than that in the down-
ward ow case. The only dierence between an upward and
downward ow experiment under the same conditions of
ow rate and heating is in the inuence of buoyancy. Con-
sequently this must be responsible for the large localised
increase of wall temperature seen in the upward ow case.
It is interesting to note that the value of buoyancy param-
eter based on the bulk parameters is only 4.3 10
7
, which
according to extensive experimental work with uids such
as water or air at atmospheric pressure could be taken as
indicating that the inuence of buoyancy ought to be insig-
nicant. However, this is clearly not the case under the
experimental conditions considered here. This will be dis-
cussed later in Section 5.4.
As can be seen from Fig. 4, the predictions for upward
ow using both the AKN and V2F models only manage
to capture the general trend of the observed variation of
wall temperature, whereas those for the downward ow
follow the experimental measurements closely. In the case
of upward ow, the simulations using both models do pre-
dict overheating, but signicantly stronger than that exhib-
ited in the experiment. Interestingly, some reduction of wall
temperature in the downstream region is predicted by both
models. In this respect, the AKN model does a somewhat
better job than the V2F model.
5.2. Buoyancy eect
As mentioned earlier, the buoyancy parameter based on
the bulk properties for Run 2 (4.3 10
7
) is lower than the
limiting value of Bo for onset of buoyancy eects of about
6 10
7
specied on the basis of work with air at normal
pressure [19]. However, the measured distributions of wall
temperature with upward and downward ow in that
experiment were distinctly dierent. This apparently anom-
alous result can be explained with the help of the modelling
results.
Fig. 6 shows the predicted radial distributions of uid
temperature, specic heat and density at several axial loca-
tions obtained using the AKN model for upward ow in
Run 2. As already seen in Fig. 1, uid properties vary
extremely sharply with temperature in the region near the
pseudo-critical value. Thus, to identify the region where
signicant variations of properties are present, a band of
temperature specied by T
pc
0.8 C has been marked
on Fig. 6a, which shows the uid temperature proles at
a number of axial locations. To facilitate the present dis-
cussion, this will be referred to as the large-property-varia-
tion (LPV) region. It can be seen from the gure that,
upstream, this is mainly located within a narrow layer very
close to the wall. Proceeding downstream, the LPV region
moves away from the wall and spreads more widely. This
can also be inferred from Fig. 6b where specic heat is
shown, knowing that the location of the maximum value
of specic heat coincides with the centre of the LPV region.
Table 3
Functions in the turbulence models
Code f
l
f
1
f
2
AKN 1
5
Re
0:75
t
exp
Ret
200
_ _
2
_ _ _ _
1 exp
y
14
_ _ _ _
2
1.0 1 0:3 exp
Ret
6:5
_ _
2
_ _ _ _
1 exp
y

3:1
_ _ _ _
2
V2F
v
2
k
1 0:045

k
v
2
_
1.0
Notes: Re
t

k
2
me
, y


y
m
u
e
.
Table 4
Damping terms D and E and wall boundary conditions for k and e
Code D E Wall BC
AKN 0 0 k
w
0; e
w
2m
k
y
2
V2F 0 0 k
w
v
2
w
f
w
0; e
w
2m
k
y
2
1668 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
It is interesting to note that although uid temperature var-
ies sharply near the wall, the variation of specic heat near
the wall at downstream locations is much smaller than that
in the vicinity of the pseudo-critical temperature where the
temperature gradient is relatively small. This is generally
true for the other uid properties as well. Thus, the major
changes of uid properties occur in the vicinity of the loca-
tion of the pseudo-critical temperature, which can be at a
location away from the wall. This is an important feature
of heat transfer with a supercritical pressure uids. With
a normal uid, the maximum variation of uid properties
always occurs near the wall where the maximum tempera-
ture gradient occurs.
The inuence of buoyancy can now be studied by exam-
ining the distribution of the density (Fig. 6c). At x/d = 5,
the density of the uid varies extremely sharply very close
the wall (around y
+
= 3). Such a large change of density
might be taken as implying a strong local buoyancy inu-
ence leading to a signicantly modied mean velocity pro-
le. However, since the inuence is mainly limited to the
viscous sub-layer, turbulence should not be signicantly
modied and therefore no signicant eect on the wall tem-
perature distribution should be expected. Indeed this does
not occur (see Fig. 4).
As the ow proceeds downstream, the LPV region
moves away from the wall, reaching about y
+
= 20 at the
location x/d = 40. Further out, the density increases shar-
ply, leading to a situation where lighter uid occupies the
whole near-wall region of the shear ow with much heavier
uid in the core ow region. Consequently a large buoy-
ancy force is experienced by the uid in the viscous sub-
layer and buer layer regions and a signicant inuence
on turbulence production is expected and therefore a sig-
nicant reduction in the eectiveness of heat transfer. This
is supported by the wall temperature results shown in
Fig. 4, where the wall temperature peaks at this axial loca-
tion. Further downstream (e.g. x/d = 80 & 120), the LPV
region moves further away from the wall and the spread
0
20
40
60
80
0 20 40 60 80 100 120
0 20 40 60 80 100 120
x/D
T
e
m
p
e
r
a
t
u
r
e

(

C
)
wall temperature (Exp.)
upflow and downflow
wall temperature (AKN)
upflow and downflow
bulk temperature (Exp.)
bulk temperature (AKN)
(a) Predictions using the AKN model
0
20
40
60
80
x/D
T
e
m
p
e
r
a
t
u
r
e

(

C
)
wall temperature (Exp.)
upflow and downflow
wall temperature (V2F)
upflow and downflow
bulk temperature (Exp.)
bulk temperature (V2F)
(b) Predictions using the V2F model
Fig. 3. Variations of wall and bulk uid temperature for Run 1 from the present computational simulations and from the experiments of Fewster [11].
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1669
0
50
100
150
200
250
0 20 40 60 80 100 120
0 20 40 60 80 100 120
x/D
W
a
l
l

t
e
m
p
e
r
a
t
u
r
e

(

C
)
upflow (AKN)
upflow (Exp.)
downflow (Exp.) downflow (AKN)
(a) Predictions using the AKN model
0
50
100
150
200
250
x/D
W
a
l
l

t
e
m
p
e
r
a
t
u
r
e

(

C
)
upflow (V2F)
upflow (Exp.)
downflow (Exp.) downflow (V2F)
(b) Predictions using the V2F model
Fig. 4. Variations of wall and bulk uid temperature for Run 2 from the present computational simulations and from the experiments of Fewster [11].
0
50
100
150
200
250
300
0 20 40 60 80 100 120 x/D
N
u
AKN(downflow)
V2F(downflow)
AKN(upflow)
V2F(upflow)
Fig. 5. Predictions of Nusselt number with upward and downward ow for Run 2.
1670 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
is much wider. Here, the density rst increases sharply near
to the wall as a result of the steep gradient of uid temper-
ature, then a slower but still very signicant variation of
density occurs over the LPV region in the vicinity of the
pseudo-critical temperature. Thus, the buoyancy force is
signicant over a wide region although the distribution is
very dierent from that which would occur in a normal
uid. This has an interesting impact on turbulence which
will be discussed in Section 5.4.
Fig. 7 shows the predicted radial proles of the temper-
ature, specic heat, and density at several axial locations
for Run 2 for downward ow. Again there are regions
where uid properties (including density) vary dramatically
over a narrow radial band. This might be expected to cause
some enhancement of turbulence production. But, for the
upstream stations (x/d = 5, 20 & 40), the LPV regions
are mostly restricted to a region very close to the wall,
and therefore the eect on turbulence is actually quite
small. This is consistent with the wall temperature varia-
tions shown earlier where no signicant inuence of buoy-
ancy was observed at those axial locations. For the last two
stations (x/d = 80 & 120), the buoyant band moves into the
buer layer region and a signicant eect on turbulence
[ x/D=5 (o), x/D=20 (+), x/D=40 (*), x/D=80 () , x/D=120 ( ) ]
0 50 100 150 200
20
25
30
35
40
45
50
y+
T

(

o
C

)
(a) Fluid temperature
0 50 100 150 200
0
2
4
6
8
10
12
14
x 10
4
y+
c
p

(
J
.
k
g
-
1
.
K
-
1
)
(b) Specific heat
0 50 100 150 200
0
100
200
300
400
500
600
700
800
900
y+


(
k
g
.
m
-
3
)
(c) Density
Fig. 6. Near-wall variations of temperature, specic heat and density for
Run 2 with upward ow predicted using the AKN model.
0 20 40 60 80 100
20
25
30
35
40
45
50
y+
T

(

o
C

)
(a) Fluid temperature
0 20 40 60 80 100
0
2
4
6
8
10
12
14
x 10
4
y+
c
p

(
J
.
k
g
-
1
.
K
-
1
)
(b) Specific heat
0 20 40 60 80 100
0
100
200
300
400
500
600
700
800
900
y+


(
k
g
.
m
-
3
)
(c) Density
[ x/D=5 (o), x/D=20 (+), x/D=40 (*), x/D=80 () , x/D=120 ( ) ]
Fig. 7. Near-wall variations of temperature, specic heat and density for
Run 2 with downward ow predicted using the AKN model.
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1671
would certainly then be expected (see the discussion in Sec-
tion 5.4).
5.3. Mean velocity and turbulence
Fig. 8 shows the predictions using the AKN model of
mean velocity proles at several axial locations for Run 2
with upward ow. The velocity prole at station x/d = 5
is typical of what would be found under conditions of
forced convection (negligible inuence of buoyancy). How-
ever, as the ow proceeds downstream, an eect of buoy-
ancy begins to be evident. At location x/d = 20 the
velocity prole has become attened. Further downstream,
at location x/d = 40, the prole is distorted very signi-
cantly and takes up an M-shape, with inverted form in
the core. This is typical of mixed convection with laminari-
zation due to inuence of buoyancy. Clearly, this is consis-
tent with the wall temperature distribution seen earlier
(Fig. 4). Further downstream (x/d = 80 and 120), the veloc-
ity prole still maintains its M-shape but the variation in
the vicinity of the maxima is more gradual.
Figs. 9 and 10 show predictions using the AKN model
of the radial proles of turbulent shear stress and turbulent
kinetic energy, respectively, at a number of axial stations
for Run 2 with upward ow. The turbulence quantities
are not aected by buoyancy inuence in the early stages
(e.g. x/d = 5). At this stage, the buoyancy force is concen-
trated in the viscous sub-layer thereby, as noted earlier, it
has little inuence on turbulence production. Proceeding
downstream, the eect of buoyancy on turbulence begins
to increase as the extent of its inuence spreads. At station
x/d = 20, both turbulent kinetic energy and shear stress are
reduced signicantly. This is followed by an even stronger
buoyancy inuence at x/d of 40, where the ow condition is
past the laminarization stage and turbulence production
has started to be re-generated in the core region of the ow
(turbulence recovery). This is evident from the proles of
shear stress shown on Fig. 9 where negative values can
be seen, although they are small. Over much of the ow,
the shear stress is close to zero. This is in accordance with
the wall temperature at x/d = 40 (see Fig. 4).
Proceeding even further downstream (x/d = 80 and x/
d = 120), turbulence recovery becomes stronger. The wall
temperature is greatly reduced and at a level which indi-
cates full recovery of turbulence (Fig. 4). It can be seen
from Fig. 9 that over a large region of the core the turbu-
lent shear stress is negative and more turbulent kinetic
energy is produced (Fig. 10) but the magnitude is actually
still quite low in real terms. However, near the wall, posi-
tive shear stress of very high magnitude is present leading
to considerable turbulence production and high values of
turbulent kinetic energy (as can be seen in Fig. 10). This
contrasts with the picture obtained from simulations with
normal uids where turbulence often remains relatively
low in the wall region and is mainly generated within the
core in the recovery regime. This dierent behaviour found
in the present study could be a special feature of convective
heat transfer with uids at supercritical pressures and
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
1.2
1.4
y/R
U
/
U
b
[ x/D=5 (o), x/D=20 (+), x/D=40 (*), x/D=80 () , x/D=120 ( ) ]
Fig. 8. Normalised velocity proles for Run 2 with upward ow predicted
using the AKN model.
0 200 400 600 800 1000
-5
0
5
10
15
20
x 10
-4
y+
u
v

(
m
2
/
s
2
)
[ x/D=5 (o), x/D=20 (+), x/D=40 (*), x/D=80 ( ) , x/D=120 ( ) ]
Fig. 9. Radial distributions of turbulent shear stress for Run 2 with
upward ow predicted using the AKN model.
0 200 400 600 800 1000
0
1
2
3
4
5
6
7
8
x 10
-3
y+
k

(
m
2
/
s
2
)
[ x/D=5 (o), x/D=20 (+), x/D=40 (*), x/D=80 () , x/D=120 ( ) ]
Fig. 10. Radial distributions of turbulent kinetic energy for Run 2 with
upward ow predicted using the AKN model.
1672 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
associated with the particular distribution of buoyancy
force resulting from the condition where the LPV region
extends right into the core of the pipe. It benets heat
transfer since turbulence produced near the wall is of prime
importance in determining the eectiveness of heat trans-
fer. As a result the wall temperature falls to a low level.
This is seen in both the experiment and the simulations.
5.4. Eect of buoyancy on turbulence production
As discussed in the Introduction, there are two mecha-
nisms by which buoyancy can aect turbulence, namely,
the indirect (external) eect, through production by shear,
on which attention has been focussed so far in this paper
and the direct (structural) eect, through production by
buoyancy [18]. The latter is associated with the fact that
density uctuation can contribute directly to turbulence
production/destruction by correlating with the turbulent
uctuation of axial velocity component and creating a tur-
bulent mass ux and hence a transfer of potential energy.
This is represented in the transport equations for the turbu-
lent kinetic energy and its dissipation by means of source
(sink) terms. These terms can be modelled using the simple
gradient diusion hypothesis (SGDH) or the general gradi-
ent diusion hypothesis (GGDH) in the eddy viscosity
model framework [23]. GGDH is generally regarded as a
more soundly-based approach and has, therefore, been
used in the current study.
Fig. 11 shows the predicted turbulence production by
shear and that by buoyancy, in Run 2 for upward and
downward ow at several axial locations using the AKN
model. It can be seen from Fig. 11df that with downward
ow buoyancy production is always positive, acting to
increase turbulence, but the values are two orders of mag-
0
500
1000
1500
2000
2500
3000
0 0.2 0.4 0.6 0.8 1 y/R
T
u
r
b
u
l
e
n
t

s
h
e
a
r

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=20
-30
-20
-10
0
10
20
30
0 0.2 0.4 0.6 0.8 1 y/R
B
u
o
y
a
n
c
y

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=20
(a) Turbulent shear production (x/D=20) (d) Buoyancy production (x/D=20)
0
500
1000
1500
2000
2500
3000
0 0.2 0.4 0.6 0.8 1 y/R
T
u
r
b
u
l
e
n
t

s
h
e
a
r

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=40
-30
-20
-10
0
10
20
30
0 0.2 0.4 0.6 0.8 1 y/R
B
u
o
y
a
n
c
y

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=40
(b) Turbulent shear production (x/D=40) (e) Buoyancy production (x/D=40)
0
1000
2000
3000
4000
5000
6000
7000
0 0.2 0.4 0.6 0.8 1
y/R
T
u
r
b
u
l
e
n
t

s
h
e
a
r

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=120
-30
-20
-10
0
10
20
30
0 0.2 0.4 0.6 0.8 1 y/R
B
u
o
y
a
n
c
y

p
r
o
d
u
c
t
i
o
n

(
k
g
.
m
-
2
)
akn-downward
akn-upward
x/D=120
(c) Turbulent shear production (x/D=120) (f) Buoyancy production (x/D=120)
Fig. 11. Turbulent shear production and buoyancy production predicted for Run 2 using the AKN model.
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1673
nitude smaller than the shear production, thereby making
it a negligible contribution in real terms. The buoyancy
production with the upward ow can be positive or nega-
tive, i.e., causing either turbulence energy production or
destruction, but again the values are much smaller than
the shear production. Therefore the direct eect of buoy-
ancy on turbulence is negligible in real terms. This conclu-
sion is consistent with the observations made in the study
of buoyancy-inuenced ow with normal uids, see Cotton
and Jackson [24].
Thus, the eect of buoyancy on ow and heat transfer
identied earlier in this paper has almost entirely resulted
from the indirect eect, which can be studied by examining
the shear production shown in Fig. 11ac. For downward
ow, the shear production essentially remains unchanged
at x/d locations of 20 and 40. As already seen in Fig. 7,
although the radial variation of density is very large at
those locations, it is restricted to a region very close to
the wall (y
+
< 5), thereby having little inuence on turbu-
lence production. At location x/d = 120, however, the
lighter uid occupies the region up to y
+
10 and there-
fore, the shear production is greatly increased as can be
seen from Fig. 11c.
For upward ow, the turbulence shear production is sig-
nicantly reduced at x/d = 20. It is further reduced (nearly
to zero) at x/d of 40. This is consistent with the buoyancy
force being large at those locations, as seen earlier in Fig. 6.
The shear production at x/d = 120 is interesting: it is lar-
gely damped out in the centre but greatly increased near
the wall. It can be seen from Fig. 6c that the buoyancy
force is very big at this axial location which implies that
with a normal uid the ow would have reached the recov-
ery regime. Under such conditions, with a normal uid,
turbulence would be produced in the core but would
remain low near the wall, which is exactly opposite to what
exhibited in this study of ow and heat transfer with a uid
at supercritical pressure. Clearly, this is due to the particu-
lar distribution of the buoyancy force which can be present
in a supercritical pressure uid which apparently can pro-
mote very eective renewed turbulence production in the
near-wall region. This buoyancy distribution is associated
with the condition where the large-property-variation
(LPV) region is situated in a region away from the wall,
as was illustrated earlier (see Fig. 6a).
6. Conclusions
Both the low Reynolds number k e model and the
V2F model were able to capture the general trends of
the interesting wall temperature behaviour observed
with upward ow in some experiments with a uid at
a pressure just above the critical value. However the
detailed variation of the wall temperature predicted
using each of the models was rather dierent from that
in the experiments. The prediction of the k e model
was the better of the two for the conditions considered.
With downward ow, the measured distribution of wall
temperature did not exhibit the localised non-uniformity
seen with upward ow and the predicted wall tempera-
ture distributions were closely consistent with observed
behaviour.
It has been found that the eect of buoyancy on heat
transfer to uids at supercritical pressure cannot be reli-
ably assessed just using a buoyancy parameter based on
bulk properties. Strong eects of buoyancy may occur
even when the buoyancy parameter is lower than the
value for which such eects have been found to be small
in the case of normal uids.
The eect of buoyancy on turbulence is largely depen-
dent on the extent of the large-property-variation
(LPV) region, i.e., the region in the vicinity of the point
where the uid temperature takes the pseudo-critical
value.
When LPV region is located very close to the wall, buoy-
ancy has little inuence on turbulence production. When
it extends into the buer layer, it causes a large reduc-
tion in turbulence production with upward ow which,
in turn, causes a signicant reduction in the eectiveness
of heat transfer.
Under some conditions, the very non-uniform distribu-
tion of buoyancy force may cause a reduction of turbu-
lence in the core but a signicant enhancement near the
wall leading to enhanced heat transfer. This interesting
phenomenon could prove to be an important feature
of heat transfer to uids at supercritical pressure which
might possibly be exploited in practical systems.
Under the conditions examined in this study, the direct
eect of buoyancy on turbulence energy production was
found to be negligible in comparison with the indirect
eect. Thus, the eect on heat transfer seen was almost
entirely due to the shear production eect caused by
the distortion of the mean ow as a result of the inu-
ence of buoyancy.
Acknowledgements
The authors gratefully acknowledge the funding for this
project provided by UK Engineering and Physical Sciences
Research Council (EPSRC) through Grant GR/S19424/02
and the use of the valuable experimental data produced by
Jonathan Fewster in his PhD work.
References
[1] J.D. Jackson, W.B. Hall, Forced convection heat transfer to uids at
supercritical pressure, Turbulence Forced Convection in Channels
and Bundles, vol. 2, Hemisphere, New York, 1979, pp. 563611.
[2] N. Zhou, A. Krishman, F. Vogel, W.A. Peters, A computational
model for supercritical water oxidation of organic toxic wastes, Adv.
Environ. Res. 4 (2000) 7995.
[3] S.S. Pitla, D.M. Robinson, E.A. Groll, S. Ramadhyani, Heat transfer
from supercritical carbon dioxide in tube ow: a critical review,
HVAC&R Research 4 (1998) 281301.
1674 S. He et al. / Applied Thermal Engineering 28 (2008) 16621675
[4] S.M. Liao, T.S. Zhao, An experimental investigation of convection
heat transfer to supercritical carbon dioxide in miniature tubes, Int. J.
Heat Mass Transfer 45 (2002) 50255034.
[5] P.X. Jiang, Y.J. Xu, J. Lv, R.F. Shi, S. He, J.D. Jackson,
Experimental investigation of convection heat transfer of CO
2
at
supercritical pressures in vertical mini tubes and in porous media,
Appl. Thermal Eng. 24 (2004) 12551270.
[6] B. Youn, A.F. Mills, Flow of supercritical hydrogen in a uniformly
heated circular tube, Numer. Heat Transfer 24 (1993) 124.
[7] S. Koshizuka, N. Takano, Y. Oka, Numerical analysis of deteriora-
tion phenomena in heat transfer to supercritical water, Int. J. Heat
Transfer 38 (1995) 30773084.
[8] Y. Okano, S. Koshizuka, Y. Oka, Safety analysis of a supercritical
pressure, light water cooled and moderated reactor with double tube
water rods, Ann. Nucl. Energy 24 (1997) 14471456.
[9] X. Cheng, T. Schulenberg, D. Bittermann, P. Rau, Design analysis of
core assemblies for supercritical pressure conditions, Nucl. Eng.
Design 233 (3) (2003) 279294.
[10] J.D. Jackson, Some striking features of heat transfer with uids at
pressures and temperatures near the critical point, Keynote paper for
International Conference on Energy Conversion and Application
(ICECA-2001), Wuhan, China, 2001.
[11] J. Fewster, Mixed forced and free convective heat transfer to
supercritical pressure uids owing in vertical pipes, PhD Thesis,
University of Manchester, 1976.
[12] J. Fewster, J.D. Jackson, Some experiments on heat transfer to
carbon dioxide at supercritical pressure in tubes, in: Proceedings of
the International congress on Advances in Nuclear Power Plants
(ICAPP04), Pittsburg, PA USA, 1317 June, 2004.
[13] P.X. Jiang, Z.P. Ren, B.X. Wang, Convective heat and mass transfer
of water at super-critical pressures under heating or cooling condi-
tions in vertical tubes, J. Thermal Sci. 4 (1) (1995) 1525.
[14] S.H. Yoon, J.H. Kim, Y.W. Hwang, M.S. Kim, K.D. Min, Y.C. Kim,
Heat transfer and pressure drop characteristics during the in-tube
process of carbon dioxide in the supercritical region, Int. J. Refrig. 26
(2003) 857864.
[15] I.L. Pioro, H.K. Khartabil, R.B. Duey, Literature survey devoted to
the heat transfer and hydraulic resistance of uids at supercritical and
near-critical pressures, AECL-12137, FFC-FCT-409, AECL, Chalk
River Labs, Ontario, Canada, 2002.
[16] S. He, P.X. Jiang, Y.J. Xu, R.F. Shi, W.S. Kim, J.D. Jackson, A
computational study of convection heat transfer to CO
2
at supercrit-
ical pressures in a vertical mini tube, Int. J. Thermal Sci. 44 (2005)
521530.
[17] L.J. Li, C.X. Lin, M.A. Ebadian, Turbulent heat transfer to near-
critical water in a heated curved pipe under the conditions of mixed
convection, Int. J. Heat Mass Transfer 42 (1999) 31473158.
[18] B.S. Petukhov, A.F. Polyakov, Heat transfer in turbulent mixed
convection, Hemisphere Publishing Corporation, New York, 1988.
[19] D.P. Mikielewicz, A.M. Shehata, J.D. Jackson, D.M. McEligot,
Temperature, velocity and mean turbulence structure in strongly
heated internal gas ows: Comparison of numerical predictions with
data, Int. J. Heat Mass Transfer 45 (2002) 43334352.
[20] S.V. Patankar, A calculation procedure for two-dimensional elliptic
situation, Numer. Heat Transfer 4 (1981) 409425.
[21] K. Abe, T. Kondoh, Y. Nagano, A new turbulence model for
predicting uid ow and heat transfer in separating and reattaching
Flow-I. Floweld calculations, Int. J. Heat Mass Transfer 37 (1994)
139151.
[22] M. Behnia, S. Parneix, P.A. Durbin, Prediction of heat transfer in an
axisymmetric turbulent jet impinging on a at plate, Int. J. Heat Mass
Transfer 41 (1998) 18451855.
[23] N.Z. Ince, B.E. Launder, On the computation of buoyancy-driven
turbulent ows in rectangular enclosures, Int. J. Heat Fluid Flow 10
(1989) 110117.
[24] M.A. Cotton, J.D. Jackson, Vertical air ows in the turbulent mixed
convection regime calculated using a low Reynolds number k e
turbulence model, Int. J. Heat Mass Transfer 33 (1990) 275286.
S. He et al. / Applied Thermal Engineering 28 (2008) 16621675 1675

S-ar putea să vă placă și