You are on page 1of 20

Numerical Heat Transfer, Part A, 45: 171–190, 2004

Copyright # Taylor & Francis Inc.


ISSN: 1040-7782 print=1521-0634 online
DOI: 10.1080/10407780390244399

HEAT TRANSFER OF SOLID–LIQUID PHASE-CHANGE


MATERIAL SUSPENSIONS IN CIRCULAR PIPES:
EFFECTS OF WALL CONDUCTION

C. J. Ho, J. F. Lin, and S. Y. Chiu


Department of Mechanical Engineering, National Cheng Kung University,
Tainan, Taiwan 70101, Republic of China

This article considers the problem of conjugate heat transfer in circular pipes with finite
heated length to examine the effects of wall conduction on the heat transfer characteristics
of solid–liquid phase-change material suspension flow. A mixture continuum approach is
adopted in the formulation of the energy equation, with an approximate enthalpy model
describing the phase-change process in the phase-change material particles. From
numerical simulations via the finite-volume approach, it was found that the conduction heat
transfer propagating along the pipe wall results in significant preheating of the suspension
flow in the nondirectly heated region upstream of the heated section, where melting of the
particles may occur and therefore the contribution of the latent heat transfer to convection
heat dissipation over the heated section is markedly attenuated. Contributions of the sensible
and latent heat transfer to the total heat transfer rate of the suspension flow over the heated
section were delineated quantitatively for various sets of the relevant dimensionless para-
meters, including the particle volumetric concentration, the modified Stefan number, the
Peclet number of suspending fluid, the wall thickness ratio, and the wall-to-fluid thermal
conductivity ratio.

INTRODUCTION
With the recent advances in technologies of microencapsulation and emulsion,
a concept of solid–liquid phase-change material (PCM) slurries or suspensions has
been proposed [1] in which PCM particles (typically with diameter of 1–20 mm) are
dispersed in a suspending fluid. In contrast to conventional particulate suspensions,
the PCM suspensions have been suggested to serve as a dual-functional medium for
thermal energy transport and=or storage [2, 3]. The problem of forced convection
heat transfer of PCM suspensions in laminar pipe flows has been the subject of
several investigations [4–7], in which the heat transfer enhancement in the presence
of melting PCM particles is demonstrated.

Received 4 June 2003; accepted 4 June 2003.


This research was supported by the National Science Council (NSC) of the Republic of China in
Taiwan through Project Nos. NSC88-2212-E006-023 and NSC89-2212-E006-160. The authors thank
the National Center for High Performance Computing of the NSC for the computing facility and time.
The constructive comments of reviewers are sincerely appreciated.
Address correspondence to C. J. Ho, National Cheng Kung University, Department of Mechanical
Engineering, Tainan, Taiwan 70101, Republic of China. E-mail: cjho@mail.ncku.edu.tw

171
172 C. J. HO ET AL.

NOMENCLATURE
þ
a parameter tþ
w pipe wall thickness (¼ rþ o  ri Þ
Ap surface area of particle T temperature
b parameter u dimensionless axial velocity (¼ uþ =uþ mÞ
Bip Biot number of PCM particle uþ velocity in the axial direction
(¼ Urþ p =kp ) U interstitial overall heat transfer
cp specific heat coefficient
cp;bf specific heat ratio (¼ cp;b =cp; f ) 
V volume
cp;pf specific heat ratio (¼ cp;p =cp; f ) x dimensionless axial coordinate
cp;wf specific heat ratio (¼ cp;w =cp; f ) [¼ xþ =ðrþ i Pe)]
cv volumetric fraction of PCM particles xþ axial coordinates
dp dimensionless particle diameter a thermal diffusivity
(¼ dpþ =rþ i ) y dimensionless temperature
dþp diameter of PCM particle [¼ ðT  Tm Þ=ðq00o rþ i =kb Þ]
2
Fo Fourier number [¼ ab t=ðrþ i Þ ] n kinematic viscosity
h enthalpy of suspension x volumetric phase fraction in a particle
hls latent heat of fusion r density
k thermal conductivity rbf density ratio (¼ rb =rf )
kbf thermal conductivity ratio (¼ kb =kf ) rpf density ratio (¼ rp =rf )
kpf thermal conductivity ratio (¼ kp =kf ) rwf density ratio (¼ rw =rf )
kwf thermal conductivity ratio (¼ kw =kf ) tlag thermal inertia time-lag parameter

d length of downstream section tlag dimensionless time-lag parameter
ld dimensionless length of downstream Vp =Ap Þ2 =ap g
ftlag =½ðcp;pf cp;bf Þð
section [¼ ldþ =ðrþ i PeÞ]
lþ length of heated section Subscripts
h
lh dimensionless length of heated section b bulk fluid
[¼ lhþ =ðrþ bf bulk to fluid ratio
i PeÞ]
c concentration

u length of upstream section
lu dimensionless length of upstream d downstream
section [¼ lþ þ f suspending fluid
u =ðri PeÞ]
Pe Peclet number (¼ 2rþ þ h heated section
i um =ab )
i inner surface of pipe
Pef Peclet number (¼ 2rþ þ
i um =af )
Pr Prandtl number (¼ nb =a b ) in inlet
q00 local heat flux ‘ liquid phase of PCM
lat latent heat
q00 local dimensionless heat flux
m melting point
(¼ q00 =q00o )
rþ radial coordinate max maximum value
rþ radius of PCM particle o outer surface of pipe
p
p particle
r dimensionless coordinate (¼ rþ =rþ i )
Re Reynolds number (¼ 2rþ þ pf particle to fluid ratio
i um =nb )
Sb inlet subcooling factor s solid phase of PCM
[¼ ðTm  Tin Þ=ðq00o rþ sen sensible heat
i =kb Þ]
u upstream
Ste* modified Stefan number
[¼ cp;b ðq00o rþ w wall of pipe
i =kb Þ=hls 
Steb modified bulb Stefan number wf wall to fluid ratio
[¼ ðrp =rb ÞSte =cv ] Superscripts
t time — surface-averaged quantity
tw dimensionless pipe wall thickness * ratio of quantities
þ
ð¼ tþw =ri ) þ dimensional quantity

Charunyakorn et al. [4] developed a model for forced convection of micro-


encapsulated PCM suspension flow in circular ducts with constant wall temperature
or heat flux. Their numerical simulation showed that heat transfer enhancement of
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 173

2–4 times in comparison with single-phase flow can be achieved. Goel et al. [5]
conducted an experimental study concerning laminar forced-convection heat transfer
characteristics in heating of a suspension containing microencapsulated n-eicosane
flowing in a circular duct with constant wall heat flux. Heat transfer enhancement
incorporating the phase-change suspension was reported by up to 50% reduction of
the wall temperature rise along the heating section. The experimental data were,
however, found to agree qualitatively with numerical predictions from the model of
Charunyakron et al. [4]. There exists a substantial disparity between the predicted
results and experimental data. Zhang and Faghri [6] improved the theoretical ana-
lysis by taking account of the microcapsules crusts, the initial subcooling, as well as
the extended range of phase-change temperature of PCM particles, and were able to
bring the predicted wall temperatures closer to the data measured by Goel et al. [5].
It can be seen from the foregoing literature that the previous theoretical studies have
neglected the thickness and conduction in the pipe wall. Yet closer examination of
the experimental set up of Goel et al. [5] reveals that their test section were con-
structed from a copper circular tube with an internal diameter of 3.14 mm and
an outer diameter of 4.76 mm, which gives a dimensionless tube wall thickness
þ
tw ½¼ 1  ðrþ
o =ri Þ of 0.52. As can be inferred from the literature on conjugate heat
transfer in single-phase pipe flow [8], conduction in the duct wall of such a thickness
can affect significantly the forced convection in a PCM suspension flow, thus leading
to the disparity between the previous theoretical predictions and the experimental
results.
In the above context, the objective of the present effort is to examine nume-
rically the effects of wall conduction on laminar forced-convection heat transfer of
PCM suspension flow through circular pipes of finite thickness. A mixture con-
tinuum approach is adopted in the present simulation with an approximate enthalpy
model describing the phase-change process in the PCM particles.

PROBLEM FORMULATION
The physical configuration of the conjugate heat transfer problem considered,
as illustrated schematically in Figure 1, is a circular pipe of inner radius rþ
i and wall
þ
thickness tþ þ
w (¼ro  ri ). The outer surface of the pipe is exposed to a constant
uniform heat flux q00o over a finite length lþ h . The upstream side of the duct is suffi-
ciently long so that a PCM suspension flows through the pipe with a fully developed
laminar velocity profile. The particles in suspension can undergo solid–liquid phase
change, so they can be solid, liquid, or both. In the mixture continuum formulation
adopted for the thermally developing convection of the suspension fluid, the fol-
lowing assumptions are invoked. (1) The particles are rigid spheres of homogeneous
size that is sufficiently small compared with the characteristic dimension of flow.
(2) The suspension fluid is dilute, homogenous with a volumetric concentration cv
of the particles, and is considered a Newtonian fluid. (3) The dispersed particles
are neutrally buoyant in the suspension fluid. (4) The flow is assumed to be
laminar, incompressible, and axis-symmetric. (5) Local thermodynamic equilibrium
is assumed except during the solid–liquid phase-change process of the particle.
(6) Density change associated with solid–liquid phase change in the particles is
neglected. (7) All physical properties of the suspension fluid are constants, and
174 C. J. HO ET AL.

Figure 1. Schematic diagram of physical configuration and coordinate system.

viscous dissipation in the flow is neglected. Moreover, conduction heat transfer in


the duct wall is assumed to be a constant property.
Based on the foregoing assumptions, the energy equation for the PCM sus-
pension flow in the pipe may be formulated by utilizing the continuum theory of
mixture in the cylindrical coordinates as
     
qh qh 1 q þ qT q2 T
rb þ uþ þ ¼ kb þ þ r þ þ2 ð1Þ
qt qx r qr qrþ qx

where the mixture enthalpy h can be expressed by means of the volumetric fraction cv
and the liquid phase fraction x‘ of the PCM particles in the suspension as
 
rp
h ¼ cp;b T þ cv x‘ hls ð2Þ
rb

Substituting Eq. (2) into Eq. (1), it follows that


        
qT qT 1 q qT q2 T rp hls qxl qx
þ uþ þ ¼ ab þ þ rþ þ þ þ2  cv þ uþ þl
qt qx r qr qr qx rb cp;b qt qx
ð3Þ

þ þ 2
where the velocity profile is uþ ¼ 2uþ m ½1  ðr =ri Þ . The second bracket on the
right-hand side of Eq. (3) represents the latent heat transport associated with
the solid–liquid phase change of the particles in the suspension flow. Meanwhile, the
conduction heat transfer in the pipe wall takes the form
   
qT 1 q qT q2 T
¼ aw þ þ rþ þ þ þ2 ð4Þ
qt r qr qr qx
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 175

In compact form the foregoing energy equations [Eqs. (3) and (4)] can be ex-
pressed in dimensionless form as follows:

  !2  
qy u qy 1 q qy kbf 1 q2 y cv qx‘ u qx‘
a þ ¼ r þ  b þ ð5Þ
qFo 2 qx r qr qr rbf cp;bf Pe2f qx2 Ste qFo 2 qx

where for the fluid region (0  r  1), u ¼ 2ð1  r2 Þ, a ¼ 1, b ¼ rbf =rpf ; and for the wall
region (1  r  1 þ tw ), u ¼ 0, a ¼ ðkbf =kwf Þðrwf =rbf Þðcp;wf =rp;bf Þ, b ¼ 0.
The relevant boundary conditions of the problem are

y ¼ yin ¼ Sb x ¼ lu ; 0  r  1 ð6aÞ

qy
¼0 x ¼ lu ; 1  r  ð1 þ tw Þ ð6bÞ
qx
qy
¼0 x ¼ ðlh þ ld Þ; 0  r  ð1 þ tw Þ ð6cÞ
qx
qy
¼0 r ¼ 0;  lu  x  ðlh þ ld Þ ð6d Þ
qr

qy kbf
¼  r ¼ ð1 þ tw Þ; 0  x  lh ð6eÞ
qr kwf

qy
¼0 r ¼ ð1 þ tw Þ;  lu  x < 0 or lh < x  ðlh þ ld Þ ð6f Þ
qr

Moreover, the continuities of temperature and heat flux at the inner wall of the
pipe are imposed that

   
qy qy
ðyi Þw ¼ ðyi Þb and kwf ¼ kbf r ¼ 1;  lu  x  ðlh þ ld Þ ð6gÞ
qr w qr b

Solid–liquid Phase Change of Particles


The last term on the right-hand side of Eq. (5) accounts for the transport of
latent heat absorption=release associated with the melting=freezing of the PCM
particles, in which closure for the liquid-phase volumetric fraction of the particles x‘
is required. Within each finite volume of the solution domain, the particles are
considered to be represented by a sample particle. Variation of the liquid-phase
volumetric fraction can then be determined by application of energy balance to the
sample particle undergoing solid–liquid phase change. In view of the small size of
the particle, it is assumed that the lumped heat capacitance method is valid during
the phase-change process. Since the local Stefan number for a small phase-changing
particle can be expected to be very small, variation of the sensible heat of the particle
is accordingly negligible. Variation of specific enthalpy of the phase-changing
176 C. J. HO ET AL.

particle is then related to the interstitial heat exchange between the particle and the
surrounding fluid at local temperature T according to
!
Dx‘ Ap U
¼ ðT  Tp Þ ð7Þ
Dt rp 
Vp h‘s

where h‘s denotes the latent heat of fusion and U is the interstitial overall heat
transfer coefficient, accounting for the effects of molecular diffusion and micro
convection around the particle, as well as the thermal resistance of microcapsule
crust, if the PCM is microencapsulated. Moreover, during the phase-change process,
the particle temperature varies between the phase-change temperature Tm ð¼
Tm  DTms Þ, where DTms denotes the degree of supercooling during solidification,
and the local suspension temperature T. It can then be postulated that

qT
Tp  Tm þ tlag ð8Þ
qt

where tlag is a time-lag parameter of the particle, which is related to heat capacity of
the particle.
Incorporating Eq. (8), Eq. (7) takes the form
!  
Dx‘ Ap U  qT
¼ ðT  Tm Þ  tlag ð9Þ
Dt rp 
Vp h‘s qt

The second term in the brackets on the right-hand side of Eq. (9) represents
the thermal inertia due to the interstitial thermal nonequilibrium effect during the
phase-change process. Furthermore, with the assumptions of negligible supercooling
(DTms ¼ 0) and neglecting thermal inertia of the phase change (tlag ¼ 0), Eq. (9)
degenerates to
!
Dx‘ Ap U
¼ ðT  Tm Þ ð10Þ
Dt rp 
Vp h‘s

Under the situation of fully developed pipe flow considered here, Eq. (10) can
be expressed in Eulerian description as
!
qx‘ þ qx‘ Ap U
þu ¼ ðT  Tm Þ ð11Þ
qt qxþ rp 
Vp h‘s

or in the dimensionless form


! !

qx‘ u qx‘ 12 kpf rbf
þ ¼ 2 Ste Bip y ð12Þ
qFo 2 qx dp kbf rpf

It is interesting to note that Eq. (12) appears to be equivalent to the source term used
in [4] under the condition of local thermal equilibrium in the suspension.
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 177

Taken together with the boundary conditions, Eq. (6), the dimensionless
differential equations of (5) and (12) form the mathematical formulation describing
the conjugate forced-convection problem considered, from which the dimensionless
parameters for the problem include the fluid Peclet number, Pef ; the modified
Stefan number, Ste*; the inlet subcooling factor Sb, or the inlet dimensionless
temperature yin ; the dimensionless particle diameter, dp ; the volumetric concentra-
tion of PCM particles, cv ; the property ratios with respect to the suspending fluid,
kwf , kpf , rwf , rpf , cp;wf , cp;pf ; and the dimensionless wall thickness of pipe, tw .
Moreover, the particle Biot number, Bip , the property ratios, kbf , rbf , and cp;bf
which also appear in the foregoing formulation, are evaluated adopting the
relationships given in [4].
Furthermore, several thermal quantities of interest that can be derived from the
solution to the foregoing governing differential equations of the problem are defined
in dimensionless form as discussed below.
Local mean temperature of the suspension fluid, yb , and the mean liquid-phase
or melted fraction of PCM particles, xb , at a given cross section of the pipe are,
respectively, calculated from
R1
0 uyðr; xÞr dr
yb ðxÞ ¼ R1 ð13Þ
0 ur dr

and
R1
0 ux‘ ðr; xÞr dr
xb ðxÞ ¼ R1 ð14Þ
0 ur dr


Local dimensionless heat flux across the inner wall of pipe, q00i , is defined as a
fraction of the heat flux input at the outer pipe wall and is given by
 
 q00 qy
q00i ¼ i00 ¼ ð15Þ
qo qr r¼1

The surface-averaged heat flux at the inner pipe wall over the heated section is
defined as
Z lh
 1 
q00i;h ¼ q00i dx ð16Þ
lh 0

Moreover, the total heat transfer rate dissipated by the suspension flow over
the heated section is normalized as a fraction of the total heat input at the outer pipe
wall and the resulting expression is of the form

Z lhþ 
rþ q00i;h
qi;h ¼ i
þ 00 q00i ðxþ Þ dxþ ¼ ð17Þ

o h qo
l 0 1 þ tw
178 C. J. HO ET AL.

If the axial fluid conduction is further assumed negligible, the total heat
transfer rate by the suspension flow over the heated section, q
i;h , can be alternatively
derived based on an integral energy balance and takes the form
( )
1 rbf cv
q ¼ ½ðyb ðx ¼ lh Þ  yb ðx ¼ 0Þ þ  ½ðx ðx ¼ lh Þ  xb ðx ¼ 0Þ ð18Þ
i;h
4lh ð1 þ tw Þ rpf Ste b

Equation (18) appears to relate the bulk temperature of suspension fluid yb as


well as the mean melted fraction of the particles xb with the total convection heat
transfer rate q
i;h . Specifically, Eq. (18) reveals that the convection heat transfer by the
suspension flow in the pipe may be due to transport of sensible heat as well as of
latent heat, which are denoted by the following expressions, respectively:

1
ðqi;h Þsen ¼ ½yb ðx ¼ lh Þ  yb ðx ¼ 0Þ ð19aÞ
4lh ð1 þ tw Þ

and
!
1 rbf  cv 
ðqi;h Þlat ¼ ½x ðx ¼ lh Þ  xb ðx ¼ 0Þ ð19bÞ
4lh ð1 þ tw Þ rpf Ste b

NUMERICAL METHOD
Spatial discretization of the differential equations was performed on a uniform
mesh for both fluid and the solid domains, incorporating the second-order central
difference scheme for the diffusion terms and the QUICK scheme [9] for the con-
vective terms. Temporal derivative terms in the differential equations are treated
using an implicit approximation with second-order accuracy. The system of alge-
braic discretization equations obtained was solved through the line-by-line appli-
cation of the tri-diagonal algorithm. Steady-state solutions to the governing
equations of the problem were obtained numerically by a pseudo-transient ap-
proach. The temperature field was considered convergent when the maximum var-
iation between the value of computed at the present and previous iterations was less
than 1076. All computations were performed with double-precision arithmetic.
Moreover, global energy balance over the computational domain was checked for
each simulation and found to be always within 0.1%.
From extensive grid-size convergence tests, simulation results shown later have
been obtained for several meshes ranging from 111 (radial direction) 6 501 (axial
direction) to 4816801, depending mainly on the fluid Peclet number Pef, the particle
volumetric concentration cv , the thermal conductivity ratio kwf , and the pipe wall
thickness ratio tw . Of the grid points in the radial direction, 5 to 31 were laid within
the thickness of the pipe wall. Moreover, all calculations were performed for a fixed
heated length, lhþ ¼ 26rþ i , with the upstream distance between the pipe entrance and
the leading edge of the heated section lþ u and the downstream distance between the
flow exit and the exit of the heated section ldþ fixed, respectively, at 13rþ þ
i and 26ri ,
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 179

þ
except for the condition of Pef ¼ 10 and kwf ¼ 100, for which lþ u ¼ 40ri and
þ þ
ld ¼ 40ri were adopted. Numerical tests have been undertaken to ensure that the
solutions were insensitive to further increases in the dimensions of the upstream and
downstream distances.

RESULTS AND DISCUSSION


To validate the present formulation and the numerical method adopted, results
of the forced convection of PCM suspension in a circular pipe with=without con-
sidering the wall conduction effect were obtained. The comparison between the
present results of the axial variation of the inner wall temperature and the experi-
mental data of Goel et al. [5] is typified in Figure 2 for Re ¼ 200, Ste ¼ 0:1, cv ¼ 0:1,
and yin ¼ 0:07. Also included in Figure 2 are the curves for the predicted results of
Goel et al. [5] using the model developed in [4] and by Zhang and Faghri [6]. It can be

Figure 2. Comparison of the predicted wall temperature distribution for forced convection of PCM
suspension in a circular duct with experimental data of Goel et al. [5] and the predicted results of
Charunyakorn et al. [4] and Zhang and Faghri [6].
180 C. J. HO ET AL.

seen that the dimensionless inner wall temperature predicted by the present for-
mulation with zero tube wall thickness (tw ¼ 0) is markedly higher than the pre-
dictions by Goel et al. [5] and by Zhang and Faghri [6]. By taking the wall
conduction into account, the prediction from the present conjugate formulation for
tw ¼ 0:52 and kwf ¼ 648:2, which mimics closely the copper tube used in the
experimental setup of Goel et al. [5], turns out to agree satisfactorily with the
experimental data, thus clarifying the cause of the large disparity between the pre-
vious numerical predictions and that of Goel et al.’s experiment.
Having established validation of the present formulation as well as demon-
strated significance of the wall conduction effect, numerical results have been ob-
tained for the parameters pertinent to the present problem in the following ranges:
Pef ¼ 10–103, cv ¼ 0–10%, Ste* ¼ 0.05–0.2, kwf *
¼ 0.1–100, tw ¼ 0–0.5 for Sb ¼ 0,
*
dp ¼ 0.01, k pf ¼ 0.422, rwf ¼ 8:959, rpf ¼ 0:770, cp;wf ¼ 0:092, and cp;pf ¼ 0:543.
In the following, influence of the parameters kwf , tw , Pef , Ste , and cv will be
presented in terms of results of the axial distributions of the following quantities: the
outer and inner temperature of the pipe wall, yo and yi ; and the local heat flux at the

inner pipe wall, q00i .

Wall Temperatures
To illustrate the wall conduction effect, the axial distributions of the inner and
outer wall temperatures, yi and yo , are presented in Figures 3–5 under various values
of Pef, Ste*, and cv. From an overall examination of the figures, the wall conduction
is demonstrated to have a strong bearing with the pre- and postheating of the wall at
the up- and downstream regions adjacent to the heated section. Over the up- and
downstream nondirectly heating regions, the inner and outer wall temperatures

Figure 3. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner and outer wall
temperatures for various Peclet numbers.
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 181

Figure 4. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner and outer wall
temperatures for various modified Stefan numbers.

display an increasing=decreasing trend with various concave=convex-curvatures,


respectively.
The distributions of the inner and outer wall temperatures are more sensitive to
changes of kwf or tw when Pef is small. As depicted by the curves for Pef ¼ 10 in
Figure 3, with increase of the thermal conductivity ratio or wall thickness ratio the
pre- and postheating effects become increasingly discernible. At kwf ¼ 100, the axial

Figure 5. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner and outer wall
temperatures for various concentrations.
182 C. J. HO ET AL.

propagation of conduction heat transfer along the pipe wall leads to an increase of
more than twofold at the leading edge and a decrease of more than 10% at the exit of
the heated section for the inner as well as the outer wall temperature. With increase
of Pef beyond 100, the intensified convection heat transfer along the suspension
stream results in a drastic decline in the extent of temperature elevation over the
surfaces of the heated section, exhibiting a fairly isothermal wall temperature at
Pef ¼ 1; 000 as shown in Figure 3.
Next, from inspection of the curves shown in Figures 4 and 5 it is evident that
the increase of cv or decrease of Ste for a fixed value of tw or kwf leads to
considerably lower wall temperatures, lending further demonstration of the efficacy
of the phase-change suspension in reducing the wall temperature as reported in the
previous studies [4–7]. On the other hand, as depicted in Figures 4a and 5a, the
enhancing capability of suspension fluid at fixed Ste* or cv can be markedly impeded
by the wall conduction due to the increase of the wall thickness, resulting in con-
siderably higher inner and outer wall temperatures. Moreover, as shown in Figures
4b and 5b, the outer wall temperature appears to be more sensitive to change of kwf .
Increase of kwf up to 10 leads to around 40% reduction of the outer wall temperature
compared with that for kwf ¼ 1. In contrast, the inner wall temperature distributions
appear rather insensitive to the increase of the thermal conductivity ratio except for
the regions adjacent to the leading edge=exit of the heated section, where a noticeable
temperature increase=decrease can be detected, further reflecting the axial conduc-
tion effects on both the pipe wall and the suspension fluid.
Another quantity of interest in the aspect of temperature control=thermal
management is the maximum temperature at the inner wall, yi;max;c , which is listed in
Tables 1 and 2 to delineate, respectively, the effects of the wall thickness and thermal
conductivity ratio for various sets of parameters considered. Also included in the
tables is the ratio of the maximum inner wall temperature for the suspension fluid to
that for the pure suspending fluid (cv ¼ 0), yi;max;c =yi;max;c¼0 , to reflect quantitatively
the effectiveness of the wall temperature control incorporating the suspension fluid.
An overview of the tables reveals clearly that the maximum inner wall temperature
can be greatly reduced, to about one-fifth that of the pure suspending fluid. Increase


Table 1. Maximum inner wall temperature for different wall thickness ratio with kwf ¼ 10

yi;max;c ðyi;max;c =yi;max;c¼0 Þ

Pef Ste* cv tw ¼ 0:02 tw ¼ 0:2 tw ¼ 0:5

10 0.1 5% 9.195 (82.7%) 10.691 (83.0%) 13.325 (83.5%)


100 1.026 (67.5%) 1.208 (69.7%) 1.527 (74.9%)
1,000 0.268 (58.1%) 0.324 (61.0%) 0.421 (64.8%)
100 0.01 5% 0.332 (21.8%) 0.409 (23.5%) 0.554 (26.4%)
0.1 1.026 (67.5%) 1.208 (69.7%) 1.527 (72.7%)
0.5 1.295 (85.1%) 1.484 (85.6%) 1.813 (86.4%)
100 0.1 0% 1.511 (100%) 1.719 (100%) 2.081 (100%)
2% 1.308 (86.6%) 1.509 (87.8%) 1.859 (89.3%)
5% 1.026 (67.9%) 1.208 (70.3%) 1.527 (73.4%)
10% 0.705 (46.7%) 0.850 (49.4%) 1.107 (53.2%)
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 183

Table 2. Maximum inner wall temperature for different thermal conductivity ratio with tw ¼ 0:2

yi;max;c ðyi;max;c =yi;max;c¼0 Þ

Pef Ste cv kwf ¼ 1 kwf ¼ 10 kwf ¼ 100

10 0.1 5% 10.868 (83.2%) 10.691 (83.0%) 10.113 (80.8%)


100 1.252 (70.1%) 1.208 (69.7%) 1.081 (68.0%)
1000 0.331 (61.4%) 0.324 (61.0%) 0.300 (60.0%)
100 0.01 5% 0.428 (24.0%) 0.409 (23.6%) 0.357 (22.5%)
0.1 1.252 (70.1%) 1.208 (69.7%) 1.081 (68.0%)
0.5 1.531 (85.8%) 1.484 (85.6%) 1.352 (85.0%)
100 0.1 0% 1.771 (100%) 1.719 (100%) 1.576 (100%)
2% 1.558 (88.0%) 1.509 (87.8%) 1.371 (87.0%)
5% 1.252 (70.7%) 1.208 (70.3%) 1.081 (68.6%)
10% 0.883 (49.8%) 0.850 (49.4%) 0.744 (47.2%)

of the concentration or decrease of the modified Stefan number can be seen to be


most effective in reducing the maximum wall temperature. The increase=decrease of
the wall thickness=thermal conductivity ratio tends to cause a marked rise in the
maximum wall temperature, thus attenuating the wall temperature control effec-
tiveness of the suspension fluid. Particular attention may be directed to the situations
of high kwf ¼ 100 or large tw ¼ 0:5 for the set of parameters Pef ¼ 10, Ste ¼ 0:1, and
cv ¼ 5%, in which the suspension fluid extracts heat from the heated section without
the effects of latent heat absorption. Under these conditions a reduction of around
20% in the wall temperature still arises relative to that for a pure suspending fluid,
demonstrating the capability of enhancing heat dissipation simply by the presence of
particles without phase change.
Moreover, by means of least-square regression analysis, correlations of the
maximum inner wall temperature with the relevant parameters can be obtained,
respectively, for the pure suspending fluid (cv ¼ 0) and the suspension fluid as
follows:
  0:032
kwf
yi;max;c¼0 ¼ 20:3 Pe0:513
f ð20aÞ
tw
 0:891   0:041
Steb kwf
yi;max;c ¼ 25:2 Pe0:509 ð20bÞ
f
1 þ Steb tw

Here the modified bulb Stefan number, Steb ½¼ Ste rp =ðrb cv Þ, is used as a
correlation parameter. The correlations Eqs. (21a) and (21b) match the corre-
sponding numerical data to within 4.4% and 16.7%, respectively, for Pef ¼ 100–
1,000, Ste* ¼ 0.01–0.5, cv ¼ 0.02–0.1, k*wf ¼ 1–100, and tw ¼ 0.02–0.2.

Heat Transfer Results


The dimensionless local heat flux to the suspension stream through the

inner wall q00i is presented in Figures 6–8 for various values of Pef , Ste , and cv ,
184 C. J. HO ET AL.

Figure 6. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner wall heat flux
for various Peclet numbers.

respectively. In each figure, results are given for a range of the wall thickness ratio
and=or thermal conductivity ratio. From examination of these figures, certain fea-
tures can be observed. A substantial amount of heat transfer may take place along
the nondirectly heated regions of the inner wall, in conformity with the foregoing
pre- and postheating effect due to the combined axial conduction along both the pipe
wall and fluid stream. The increase of the wall thickness or the thermal conductivity
ratio can affect markedly the extent of the pre- and postheating regions, depending
on Pef , Ste , and cv . In particular, as can be detected from Figure 6, at low Pef ( ¼ 10)
the extent of the preheating region becomes greatly enlarged with increasing kwf up
to 100, while the postheating region contracts. Moreover, along the heated section an
iso-flux heat transfer limit can be increasingly achieved when tw or kwf is decreased
for the ranges of Pef , Ste , and cv considered, as typified by the curves for tw ¼ 0:02
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 185

Figure 7. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner wall heat flux
for various modified Stefan numbers.

or kwf ¼ 1 in Figures 6–8. As further shown in Figures 7 and 8, at a fixed tw or kwf ,


the distribution curves of the local heat flux over the heated section appear to be
rather unaffected by the change of Ste or cv . On the other hand, at a fixed value of
Ste or cv the increase of tw yields a considerable increase in the magnitude of the
local heat flux along the heated section, while the increase of kwf induces increasingly
nonuniform distribution of the heat flux.
Finally, further insight of the nature of the heat transport mechanism of the
suspension fluid flow in the pipe can be gained by evaluating the quantities of sen-
sible heat transfer ðqi;h Þsen and latent heat transfer ðqi;h Þlat associated with melting of
the particles in Eq. (19), which have been tabulated in Tables 3–6 for a range of tw
and=or kwf , respectively, at different values of Pef , Ste , and cv . Also included in the

tables are the results of the surface-averaged heat flux q00i;h and the total heat transfer

rate qi;h over the inner wall of the heated section. From examination of these tables,
186 C. J. HO ET AL.

Figure 8. Effects of (a) wall thickness ratio and (b) thermal conductivity ratio on the inner wall heat flux
for various concentrations.

general features of the competing mechanisms of sensible and latent heat transfer can
be identified as below.
First of all, as can be inferred from Tables 3–6, for the sets of parameters
considered the axial fluid conduction over the heated section accounts for generally
less than 2% of the total heat transport in the pipe flow.
The surface-averaged heat flux as well as the total heat transfer rate over the
heated section, as can be expected from Figures 6–8, for the set of parameters
considered, displays a substantial increase with increasing Pef but a marked decrease
with increasing kwf . The increase of tw tends to affect the averaged heat flux and the
total heat transfer rate differently, giving rise to a marked decrease in the total heat
transfer but a significant increase in the averaged heat flux. Moreover, the total heat
transfer rate tends to be noticeably augmented with the increase=decrease of cv =Ste ,
respectively. By contrast, the averaged heat flux appears somewhat insensitive to
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 187

Table 3. Results of heat transfer rate of pure suspending fluid for various wall thickness and thermal
conductivity ratio at different Pef

Pef kwf tw q00i;h;c¼0 qi;h;c¼0 ðqi;h;c¼0 Þsen ððqi;h;c¼0 Þsen =qi;h;c¼0 Þ

10 1 0.2 1.184 0.987 0.968 (98.1%)


10 0.02 1.009 0.989 0.970 (98.1%)
0.2 1.140 0.950 0.934 (98.3%)
0.5 1.332 0.888 0.872 (98.2%)
100 0.2 0.808 0.673 0.665 (98.8%)
100 1 0.2 1.188 0.990 0.976 (98.6%)
10 0.02 1.012 0.992 0.978 (98.6%)
0.2 1.162 0.969 0.956 (98.7%)
0.5 1.412 0.941 0.929 (98.7%)
100 0.2 1.058 0.882 0.871 (98.6%)
1,000 1 0.2 1.190 0.992 0.980 (98.8%)
10 0.02 1.014 0.995 0.983 (98.8%)
0.2 1.175 0.979 0.968 (98.9%)
0.5 1.441 0.960 0.949 (98.9%)
100 0.2 1.109 0.924 0.913 (98.8%)

changes in Ste or cv . It can be further noticed that relative to the corresponding


results of the pure suspending fluid (cv ¼ 0) tabulated in Table 3, the averaged heat
flux as well as the total heat transfer rate of the suspension fluid at the fixed Pef , tw ,
and kwf shows a minute increase (at most near 2%) for the ranges of Ste and cv
considered in Tables 4–6.
To the total heat transfer over the heat section the contributions of the sensible
and latent heat transfer of the suspension flow are found to be functions of Pef , Ste ,

Table 4. Heat transfer results for various wall thickness and thermal conductivity ratio at different Pef
with Ste* ¼ 0.1 and cv ¼ 5%

q00i;h;c qi;h;c ðqi;h;c Þsen ðqi;h;c Þlat
Pef kwf tw 00 
ðqi;h;c =q00i;h;c¼0 Þ ðqi;h;c =qi;h;c¼0 Þ ððqi;h;c Þsen =qi;h;c Þ ððqi;h;c Þlat =qi;h;c Þ

10 10 0.02 1.009 (1.000) 0.990 (1.001) 0.949 (95.9%) 0.034 (3.4%)


0.2 1.147 (1.006) 0.955 (1.005) 0.938 (98.2%) 0.011 (1.2%)
0.5 1.343 (1.008) 0.895 (1.008) 0.888 (99.2%) 0.000 (0.0%)
100 0.02 1.013 (1.001) 0.993 (1.001) 0.575 (57.9%) 0.414 (41.7%)
0.2 1.166 (1.003) 0.972 (1.003) 0.624 (64.2%) 0.344 (35.4%)
0.5 1.419 (1.005) 0.946 (1.005) 0.681 (72.0%) 0.261 (27.6%)
1,000 0.02 1.014 (1.000) 0.995 (1.000) 0.243 (24.4%) 0.750 (75.3%)
0.2 1.177 (1.002) 0.981 (1.002) 0.266 (27.1%) 0.713 (72.7%)
0.5 1.446 (1.003) 0.964 (1.004) 0.304 (31.5%) 0.658 (68.3%)
10 1 0.2 1.185 (1.001) 0.987 (1.000) 0.954 (96.7%) 0.027 (2.7%)
100 0.815 (1.009) 0.680 (1.010) 0.672 (98.8%) 0.000 (0.0%)
100 1 1.188 (1.000) 0.990 (1.000) 0.635 (64.1%) 0.351 (35.5%)
100 1.070 (1.011) 0.892 (1.011) 0.583 (65.4%) 0.305 (34.2%)
1,000 1 1.190 (1.000) 0.992 (1.000) 0.274 (27.6%) 0.716 (72.2%)
100 1.120 (1.010) 0.933 (1.010) 0.233 (25.0%) 0.698 (74.8%)
188 C. J. HO ET AL.

Table 5. Heat transfer results for various wall thickness and thermal conductivity ratio at different Ste
with Pef ¼ 100 and cv ¼ 5%

q00i;h;c qi;h;c ðqi;h;c Þsen ðqi;h;c Þlat
Ste kwf tw 00
ðqi;h;c =qi;h;c¼0 Þ 00 ðqi;h;c =qi;h;c¼0 Þ

ððqi;h;c Þsen =qi;h;c Þ

ððqi;h;c Þlat =qi;h;c Þ


0.01 10 0.02 1.013 (1.001) 0.993 (1.001) 0.035 (3.5%) 0.955 (96.2%)
0.2 1.170 (1.007) 0.975 (1.006) 0.037 (3.8%) 0.936 (96.0%)
0.5 1.430 (1.003) 0.953 (1.013) 0.046 (4.8%) 0.905 (95.0%)
0.1 0.02 1.013 (1.001) 0.993 (1.001) 0.575 (57.9%) 0.414 (41.7%)
0.2 1.166 (1.003) 0.972 (1.003) 0.624 (64.2%) 0.344 (35.4%)
0.5 1.418 (1.004) 0.946 (1.005) 0.681 (72.0%) 0.261 (27.6%)
0.5 0.02 1.012 (1.000) 0.992 (1.000) 0.907 (91.4%) 0.081 (8.2%)
0.2 1.163 (1.001) 0.969 (1.000) 0.904 (93.3%) 0.061 (6.3%)
0.5 1.413 (1.001) 0.942 (1.001) 0.897 (95.2%) 0.041 (4.4%)
0.01 1 0.2 1.189 (1.001) 0.991 (1.001) 0.044 (4.4%) 0.944 (95.3%)
100 1.101 (1.041) 0.917 (1.040) 0.026 (2.8%) 0.889 (96.9%)
0.1 1 1.189 (1.001) 0.991 (1.001) 0.635 (64.1%) 0.351 (35.5%)
100 1.070 (1.011) 0.892 (1.011) 0.583 (65.4%) 0.305 (34.2%)
0.5 1 1.188 (1.000) 0.990 (1.000) 0.918 (92.7%) 0.068 (6.9%)
100 1.060 (1.002) 0.883 (1.001) 0.842 (95.4%) 0.037 (4.2%)

cv , tw , and kwf . As indicated in Table 4, at low Pef ¼ 10 the sensible heat transfer is
the dominant mechanism of the convective transport in the suspension flow and its
contribution tends to increase considerably with the increase of tw or kwf . At higher
Pef ( 100), the amount of the sensible heat transfer is seen to reduce substantially,

Table 6. Heat transfer results for various wall thickness and thermal conductivity ratio at different cv with
Ste* ¼ 0.1 and Pef ¼ 100

q00i;h;c qi;h;c ðqi;h;c Þsen ðqi;h;c Þlat
cv kwf tw 00 
ðqi;h;c =q00i;h;c¼0 Þ ðqi;h;c =qi;h;c¼0 Þ

ððqi;h;c Þsen =qi;h;c ÞÞ

ððqi;h;c Þlat =qi;h;c ÞÞ


2% 10 0.02 1.013 (1.001) 0.992 (1.000) 0.837 (84.4%) 0.149 (15.0%)


0.2 1.164 (1.002) 0.970 (1.002) 0.844 (87.0%) 0.120 (12.4%)
0.5 1.415 (1.002) 0.943 (1.002) 0.852 (90.3%) 0.085 (9.0%)
5% 0.02 1.013 (1.001) 0.993 (1.001) 0.575 (57.9%) 0.414 (41.7%)
0.2 1.166 (1.003) 0.972 (1.004) 0.624 (64.2%) 0.344 (35.4%)
0.5 1.419 (1.005) 0.946 (1.005) 0.681 (72.0%) 0.261 (27.6%)
10% 0.02 1.014 (1.002) 0.993 (1.001) 0.280 (28.2%) 0.711 (71.6%)
0.2 1.169 (1.006) 0.974 (1.006) 0.320 (32.9%) 0.652 (66.9%)
0.5 1.424 (1.008) 0.950 (1.010) 0.382 (40.2%) 0.565 (59.5%)
2% 1 0.2 1.188 (1.001) 0.990 (1.001) 0.858 (86.7%) 0.126 (12.7%)
100 1.063 (1.005) 0.886 (1.005) 0.791 (89.3%) 0.089 (10.0%)
5% 1 1.188 (1.001) 0.992 (1.003) 0.635 (64.0%) 0.351 (35.4%)
100 1.070 (1.011) 0.892 (1.011) 0.583 (65.4%) 0.305 (34.2%)
10% 1 1.189 (1.002) 0.991 (1.002) 0.330 (33.3%) 0.657 (66.3%)
100 1.078 (1.019) 0.898 (1.018) 0.277 (30.8%) 0.619 (68.9%)
HEAT TRANSFER OF SOLID–LIQUID PCM SUSPENSIONS 189

giving way to the latent heat transfer to assume the dominant role in the convective
heat transport of the suspension flow. At Pef ¼ 1; 000 the latent heat transfer can
account for more than 75% of the convective heat transport over the heated section.
From Tables 5 and 6, further note can be taken of the effects of modified Stefan
number and concentration. Resembling that of increasing Pef , the latent heat
transfer becomes increasingly predominant with the increase of cv or the decrease of
Ste at the fixed tw or kwf and other parameters selected. For instance, as revealed in
Table 5, under the condition of Pef ¼ 100, cv ¼ 5%, tw ¼ 0:02, and kwf ¼ 10, nearly
97% of the total heat transport is due to the latent heat transfer as Ste is reduced to
0.01, resulting in a drastic drop in the maximum wall temperature to only 21.8% of
the pure suspending fluid indicated in Table 1. This convincingly demonstrates the
efficacy of incorporating the phase-change suspension in the temperature control
over the heated section.
Finally, the increasingly dominant role played by the latent heat transfer in the
suspension flow with increasing Pef and cv or decreasing Ste may be markedly im-
peded by the wall conduction associated with increasing tw or kwf . The latent heat
transfer displays a somewhat monotonic declining trend with increasing tw , in par-
ticular at low Pef . A limiting situation can be detected in Table 4 that at Pef ¼ 10
with Ste ¼ 0:1 and cv ¼ 5% the latent heat transfer and thus its contribution be-
comes diminished when tw is increased up to 0.5 at kwf ¼ 10 or 0.2 at kwf ¼ 100. On
the other hand, the increase of kwf produces a nonmonotonic influence on the latent
heat transfer, depending on Pef , Ste , and cv . There exists a value of kwf , as indicated
in Tables 4–6 for high values of Pef ( ¼ 1,000) and cv ( ¼ 10%), or low value of Ste
( ¼ 0.01), at which the trend of declining contribution by the latent heat transfer with
increasing kwf at fixed tw is reversed, whereupon the contribution of the latent heat
transfer is promoted with further increase of kwf .

CONCLUDING REMARKS
The problem of conjugate heat transfer of PCM suspension flow in circular
pipes of finite heated length has been investigated numerically. It has been identified
quantitatively from energy budgets that the latent heat transfer arising in the phase-
change suspension flow can make a major contribution to the convective heat
transport and thus effectively suppress the wall temperature rises over the heated
section. For high values of the wall thickness ratio and=or the wall-to-fluid thermal
conductivity ratio, the pre- and postheating of both the wall and the suspension fluid
in the up- and downstream regions adjacent to the heated section, respectively, have
been shown to be most distinctive at low fluid Peclet number. As a consequence,
substantial increase in the inner wall is resulted in the nondirectly heated region
upstream of the heated section where melting of the PCM particles may has started
and=or completed; so that the latent heat transfer associated with the undergoing
melting and thus the efficacy of wall temperature control over the heated section can
be markedly attenuated. This finding underscores the necessity of taking proper
account of the wall conduction effects when incorporating the PCM suspension
system in a pipe flow configuration having high wall-to-fluid thermal conductivity
ratio and=or large wall thickness ratio.
190 C. J. HO ET AL.

REFERENCES
1. R. Hart and F. Thornton, Microencapsulation of Phase Change Materials, Final Report
Contract No. 82–80, Ohio Department of Energy, Ohio, 1982.
2. C. W. Sohn and M. M. Chen, Heat Transfer Enhancement in Laminar Slurry Pipe Flows
with Power Law Thermal Conductivities, ASME J. Heat Transfer, vol. 106, pp. 539–542,
1984.
3. K. E. Kasza and M. M. Chen, Improvement of the Performance of Solar Energy or Waste
Heat Utilization Systems by Using Phase-Change Slurry as an Enhanced Heat-Transfer
Storage Fluid, J. Solar Energy Eng., vol. 107, pp. 229–236, 1985.
4. P. Charunkyakorn, S. Sengupta, and S. K. Roy, Forced Convective Heat Transfer in
Microencapsulated Phase Change Material Slurries: Flow in Circular Ducts, Int. J. Heat
Mass Transfer, vol. 34, no. 3, pp. 819–833, 1991.
5. M. Goel, S. K. Roy, and S. Sengupta, Laminar Forced Convective Heat Transfer in
Microencapsulated Phase Change Material Suspensions, Int. J. Heat Mass Transfer,
vol. 37, no. 4, pp. 593–604, 1994.
6. Y. Zhang and A. Faghri, Analysis of Forced Convective Heat Transfer in Micro-
encapsulated Phase Change Materials, J. Thermophys. Heat Transfer, vol. 9, no. 4,
pp. 727–732, 1995.
7. S. K. Roy and B. L. Avanic, Laminar Forced Convection Heat Transfer with Phase
Change Material Suspensions, Int. Commun Heat Mass Transfer, vol. 28, pp. 895–904,
2001.
8. M. Faghri and E. M. Sparrow, Simultaneous Wall and Fluid Axial Conduction in
Laminar Pipe-Flow Heat Transfer, ASME J. Heat Transfer, vol. 102, pp. 58–63, 1980.
9. B.P. Leonard, A Stable and Accurate Convective Modeling Procedure Based on Quadratic
Upstream Interpolation, Comput. Meth. Appl. Mech. Eng., vol. 19, pp. 59–98, 1979.

You might also like