You are on page 1of 36

2nd Reading

January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

International Journal of Computational Methods


Vol. 15, No. 1 (2018) 1850072 (36 pages)
c World Scientific Publishing Company
DOI: 10.1142/S021987621850072X

Modeling of Nanofluid-Fluid Two-Phase Flow


and Heat Transfer

Guan Qiangshun∗ and Yap Yit Fatt†


Department of Mechanical Engineering, The Petroleum Institute
A Part of Khalifa University of Science and Technology
P. O. Box 2533, Abu Dhabi, UAE
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

∗guqiangshun@pi.ac.ae
Int. J. Comput. Methods Downloaded from www.worldscientific.com

†yfatt@pi.ac.ae

Li Hongying
Institute of High Performance Computing
A*STAR, 1 Fusionopolis Way, #16-16 Connexis
138632 Singapore, Singapore
lih@ihpc.a-star.edu.sg

Che Zhizhao
State Key Laboratory of Engines, Tianjin University
92 Weijin Rd, Nankai Qu, Tianjin Shi
Tianjin 300072, P. R. China
chezhizhao@tju.edu.cn

Received 22 May 2017


Accepted 21 November 2017
Published 4 January 2018

This paper presents a model for two-phase nanofluid-fluid flow and heat transfer. The
nonuniform nanoparticles are transported using Buongiorno model by convection, Brow-
nian diffusion and thermophoresis. This is the first attempt to employ Buongiorno model
for two-phase nanofluid-fluid flow. The moving interface between the nanofluid and the
immiscible fluid is captured using the level-set method. The model is first verified and
then demonstrated for coupled flow and heat transfer in (1) a water–alumina nanofluid-
filled cavity with a rising silicone oil drop and (2) stratified flow of water–alumina
nanofluid, pure water and silicone oil in a channel.

Keywords: Two-phase flow; nanofluid; Buongiorno model; level-set method.

1. Introduction
With modern nanotechnology, highly uniform particles with an average size of
50 nm can be produced efficiently. These nanoparticles find very diverse engineering

† Corresponding author.

1850072-1
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

applications. From a heat transfer perspective, addition of small amount of nanopar-


ticles in the form of stably dispersed suspension into a base fluid, i.e., called
nanofluid, alters the base fluid transport properties and favorably enhances heat
transfer performance [Yu et al. (2008)]. Besides heat transfer, other applications of
nanofluid have also been explored [Taylor et al. (2013)].
Given its favorable heat transfer characteristics, numerous experimental and
numerical modeling efforts on nanofluid were performed to further understand the
underlying physics. Generally, the thermodynamic properties of nanofluid depend
strongly on the concentration of nanoparticles. These properties have been deter-
mined theoretically, but some of them still rely heavily on empiricism given the
very fact that a unified consensus has not yet been achieved [Khanafer and Vafai
(2011)]. In particular, in most numerical modeling works of a single-phase nanofluid
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

flow and heat transfer [Alloui et al. (2011); Chamkha and Abu-Nada (2012); Bachok
Int. J. Comput. Methods Downloaded from www.worldscientific.com

et al. (2012); Garoosi et al. (2013); Cianfrini et al. (2014); Turkyilmazoglu (2014)],
the assumption of having uniform concentration of nanoparticles in the base fluid
is frequently made. However, due to various transport processes, nonuniform con-
centration of nanoparticles in the base fluid can in fact arise due to various particle
transport mechanisms [Buongiorno (2005)]. As a result, the local thermodynamic
properties of nanofluid vary both spatially and temporally. This effectively alters
the transport processes in particular heat transfer in a fully coupled manner. Model-
ing of single-phase nanofluid flow with local variation of nanoparticle concentration
accounted for has been conducted using, among others, the Eulerian model (in which
the base fluid and the nanoparticles are assumed as separate continuous phase with
momentum and energy transfers) [Akgül and Pakdemirli (2016)] and Buongiorno
model (in which the nanofluid is assumed as a continuous phase with the dominant
Brownian and thermophoretic effects) [Kuznetsov and Nield (2010); Yang et al.
(2013); Malvandi and Ganji (2014); Corcione et al. (2015); Garoosi et al. (2015);
Elshehabey and Ahmed (2015); Mohyud-Din et al. (2017)].
Frequently in various applications, two-phase flow, where one phase is a
nanofluid and the remaining phase is an immiscible fluid, is encountered, for exam-
ple, nanofluid flow boiling in enhanced heat transfer [Lee and Mudawar (2007);
Cheng and Liu (2013); Duursma et al. (2015)] and drag reduction in slug two-phase
flow of air and water (nanofluid) [Pouranfard et al. (2015)] and fluid properties alter-
ation in enhanced oil recovery [Ogolo et al. (2012)]. However, numerical/analytical
modeling of these two-phase flows involving a nanofluid is very limited in exist-
ing literatures given the complex nature of coupled mass, momentum and energy
transports of the two phases at the evolving interface. Furthermore, the unknown
interface is required to be either tracked or captured.
Analytical modeling of fully developed stratified two-phase nanofluid-fluid flow
was considered [Van Gorder et al. (2012); Van Gorder (2013); Farooq and Lin
(2014)]. The nanoparticles are transported by both Brownian and thermophoretic
diffusions and therefore not uniform within the nanofluid. Steady-state developing
stratified two-phase nanofluid-fluid flow was modeled numerically in Abbasi and

1850072-2
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

Baniamerian [2014] with the assumption of uniformly distributed nanoparticles in


the nanofluid. In these studies, a simplified interfacial profile is assumed. For more
general flow configurations, the interface is not known in advance. As such, general-
ization of these works is challenging. Therefore, a more general numerical model is
required for two-phase nanofluid-fluid flows. It is stressed once more that the adjec-
tive “two-phase” in the current context refers to a nanofluid and an immiscible fluid.
It is not to be confused with “two-phase nanofluid” where nanofluid itself is consid-
ered to consist of a base fluid and solid nanoparticles [Sert et al. (2016)]. To the best
knowledge of the authors, there are only three existing two-phase nanofluid-fluid
flow models [Wang et al. (2014, 2015)]: Eulerian-Eulerian, volume-of-fluid (VOF)
and coupled level-set/volume-of-fluid (CLSVOF) models. In the Eulerian-Eulerian
model, a three-phase solid (i.e., nanoparticles)–gas–liquid mixture is assumed. The
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

dynamics of each phase is governed by its respective conservation equations with


Int. J. Comput. Methods Downloaded from www.worldscientific.com

coupled momentum and energy exchanges between phases accounted for. Thus, the
nanoparticles are transported by various mechanisms and therefore generally not
uniform in the nanofluid. Both VOF and CLSVOF models are named according to
the approach employed in capturing the interface. In these two models, the nanopar-
ticles are assumed uniformly distributed in the nanofluid. Therefore, these are reg-
ular two-phase flow models with modified uniform properties for the nanofluid.
It was highlighted in Wang et al. [2015] that, of the three models, the Eulerian-
Eulerian model is more appropriate for prediction of flow and heat transfer char-
acteristics as the model accounts for the heterogeneity of nanoparticle distribution
within the nanofluid. However, it should be noted that prediction of regular two-
phase flow models, e.g., VOF and CLSVOF, can be improved by incorporating
essential nanoparticle transport mechanisms leading to nonuniform nanoparticle
distribution within the nanofluid. This is attempted in this paper.
Therefore, this paper presents a two-phase flow model involving a nanofluid
interacting dynamically with another immiscible fluid in the presence of heat trans-
fer [Yap and Li (2015)]. The nanoparticles are transported by various mechanisms
using the model of Buongiorno [2005], leading to a nonuniform concentration and
therefore variable nanofluid properties. To the best knowledge of the authors, this
is the first attempt to employ Buongiorno model for two-phase nanofluid-fluid flow.
The transports of mass, momentum, nanoparticles and energy are fully coupled
to the interface between the two fluids. The interface between the two fluids is
captured using a level-set method. The model is applicable for low nanoparticle
concentration where there is negligible particle–particle interaction. Fortunately,
this is usually the case with nanofluid requiring generally low nanoparticle fraction
of within a few percent. Absorption (or desorption) of nanoparticles onto (out of)
interface is not considered. The governing equations are then solved using a finite
volume framework. The model is first verified by comparing to existing results of
a single-phase nanofluid flow with heat transfer. Finally, application of the model
to the case of a rising silicone oil drop in a water–alumina nanofluid-filled cavity is
demonstrated.

1850072-3
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

For two-phase nanofluid-fluid flow, both the existing Eulerian-Eulerian model


(with interpenetrating continua assumption) and the current model can account for
nonuniformly distributed nanoparticles as a result of various transport mechanisms.
Compared with the Eulerian-Eulerian model, the current model is simpler as it is
developed from a regular two-phase flow model (using level-set method for inter-
face capturing) coupled to an additional nanoparticle transport equation, i.e., the
Buongiorno model. Besides, other regular two-phase flow models using either VOF
or combined CLSVOF can be extended to the scenario of nonuniform nanoparticle
distribution using similar approach of the current model. Potential modeling appli-
cations include cooling in piston gallery [Wang et al. (2014, 2015)], liquid−liquid
extraction [Saien and Bamdadi (2012)] and phase-change heat transfer, e.g., evap-
oration [Gerken and Oehlschlaeger (2017)], pool boiling [Ciloglu and Bolukbasi
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

(2015)] and flow boiling [Fang et al. (2015)]. In these examples, the nanoparticle
Int. J. Comput. Methods Downloaded from www.worldscientific.com

concentration generally changes both spatially and temporally.

2. Problem Description
Figure 1 shows a domain consisting of two immiscible fluids, i.e., the −ve fluid and
the +ve nanofluid with each occupying the Ω− and the Ω+ regions, respectively.
The +ve nanofluid is a dilute mixture of nanoparticles and a base fluid. Due to the
presence of nanoparticles, it generally has better heat transfer properties compared
with those of the base fluid [Yu et al. (2008); Taylor et al. (2013)]. The Ω− and the
Ω+ regions are separated by an interface Γ, which can be highly irregular and evolves
over time. The level-set function [Osher and Sethian (1988)] is used to represent
the interface mathematically as


 −d, if x ∈ Ω−

φ = 0, if x ∈ Γ (1)


+d, if x ∈ Ω (nanofluid),
+

where d is the shortest distance from the interface. The two fluids interact dynam-
ically via appropriate interfacial conditions at the interface. Transport processes

Fig. 1. Domain of interest with two fluids: the –ve fluid and the +ve nanofluid.

1850072-4
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

involving mass, momentum and energy between these fluids are therefore intimately
coupled.

3. Mathematical Formulation
3.1. Governing transport equations
A one-fluid formulation is employed in this paper. Basically, the domain of interest
is visualized as to consist of a single special fluid. This fluid is special in the sense
that its properties at a given time and location are set to the properties of either
the −ve fluid or the +ve nanofluid depending on whichever occupies that particular
location at that particular time. Any property of the special fluid can be determined
conveniently from the level-set function, as will be presented in Sec. 3.2. Within a
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

one-fluid formulation, the transports of mass, momentum, energy and nanoparticles


Int. J. Comput. Methods Downloaded from www.worldscientific.com

of the special fluid are governed by the following equations:


∇ · u = 0, (2)
∂(ρu)
+ ∇ · (ρuu) = −∇p + ∇ · [µ(∇u + ∇uT )] + ρg − σ(∇ · n̂)n̂δ(φ), (3)
∂t
∂(ρcT )
+ ∇ · (ρcuT ) = ∇ · k∇T, (4)
∂t
 
∂ϕ ∇T
+ ∇ · (uϕ) = ∇ · DB ∇ϕ + DT , (5)
∂t T
where u, p, T and ϕ are, respectively, the velocity vector, the pressure, the temper-
ature and the nanoparticle volume fraction. It should be clear that the density ρ,
viscosity µ, specific heat c and thermal conductivity k of the special fluid vary both
spatially and temporally. In Eq. (3), g is the gravitational acceleration. The fourth
term on the right-hand side of Eq. (3) accounts for the interfacial tension effect
using the continuum surface force model [Brackbill et al. (1992)] at the interface
where the Dirac delta function δ(φ) and the unit normal to the interface n̂ can be
calculated, respectively, as


 1 + cos(πφ/ε) , if |φ| < ε
δ(φ) = 2ε (6a)

0, otherwise,
∇φ
n̂ = . (6b)
|∇φ|
The energy and nanoparticle transport equations (Eqs. (4) and (5)) are due to
Buongiorno [2005]. In the energy equation, heat transfer associated with nanoparti-
cle dispersion is small compared with heat conduction and convection and therefore
neglected. The assumption that the nanoparticles and the base fluid are in local
thermal equilibrium is implicitly made. More important transport mechanisms of
nanoparticles due to Brownian diffusion with coefficient DB and thermophoresis
with coefficient DT are included. For the Ω− region occupied by the −ve fluid,

1850072-5
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

both DB and DT are zero. Nanoparticle absorption onto and desorption from the
interface are not considered.

3.2. Properties
The level-set function provides a convenient means to evaluate the properties of the
special fluid. For this purpose, a smoothed Heaviside function H(φ) over a band of
width 2ε is defined:

0,
 if φ < −ε

  

φ+ε 1 πφ
H(φ) = + sin , if |φ| ≤ ε (7)

 2ε 2π ε


1, if φ > +ε.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Then, any property of interest for the special fluid α can be expressed as either an
arithmetic mean of the form
α(φ) = (1 − H)α− + Hα+ (8a)
or a harmonic mean of the form
1 1−H H
= + , (8b)
α(φ) α− α+
where the subscripts − and + refer to quantity associated with the −ve fluid and
the +ve nanofluid, respectively. The density and specific heat capacity are evaluated
using the arithmetic mean. The viscosity, thermal conductivity, Brownian diffusion
coefficient and thermophoretic diffusion coefficient are evaluated using the harmonic
mean.
For the +ve nanofluid, the density, viscosity [Khanafer and Vafai (2011)], spe-
cific heat capacity and thermal conductivity [Khanafer and Vafai (2011)] are given,
respectively, by
ρ+ = ϕρp + (1 − ϕ)ρbf , (9a)
µbf
µ+ = , (9b)
1 − 34.87(dp /dbf )−0.3 ϕ1.03
(ρc)+ = ϕ(ρc)p + (1 − ϕ)(ρc)bf , (9c)
 10  0.03

T kp
k+ = kbf 1 + 4.4Re0.4 0.66
p Prbf ϕ0.66 , (9d)
Tr,bf kbf
where the subscripts p and bf refer to quantity associated with the nanoparticles
and the base fluid, respectively.
In the evaluation of nanofluid viscosity via Eq. (9b), dp and dbf are, respectively,
the nanoparticle diameter and the equivalent diameter of the base fluid molecule.
In particular, dbf can be evaluated as
 1/3
6M
dbf = 0.1 , (9e)
N πρ0,bf

1850072-6
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

where M , N and ρ0,bf are, respectively, the molar mass of the base fluid, the Avo-
gadro number and the mass density of the base fluid at T0 = 293 K. To calculate
the nanofluid thermal conductivity using Eq. (9d), Tr,bf is the freezing point of the
base fluid. The nanoparticle Reynolds number Rep and Prandtl number of the base
fluid Prbf are given by
2ρbf kB T
Rep = , (9f)
πµ2bf dp
cbf µbf
Prbf = , (9g)
kbf
where kB = 1.3806 × 10−23 J/K is the Boltzmann’s constant. Do note that there
are many theoretical relations and empirical correlations developed for viscosity
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

and thermal conductivity, see, for example, discussed in Khanafer and Vafai [2011]
Int. J. Comput. Methods Downloaded from www.worldscientific.com

and Corcione [2011]. If desired, these can also be employed in place of Eqs. (9b)
and (9d).
Finally, the Brownian diffusion coefficient DB and thermal diffusion coefficient
DT of the nanofluid can be calculated, respectively, as
kB T
DB = , (9h)
3πµbf dp
kbf µbf ϕ
DT = 0.26 . (9i)
2kbf + kp ρbf

3.3. Interface capturing


The interface and therefore the level-set function are both advected by the under-
lying fluid velocity field u. Its evolution is governed by
∂φ
+ u · ∇φ = 0. (10)
∂t
Due to nonuniformity in u and numerical errors incurred in advection of the level-
set function, φ may cease to be a distance function. The distance function property
of φ must be maintained for the interface to be captured accurately.
To maintain φ as a distance function after the advection of φ via Eq. (10),
redistancing [Sussman et al. (1994)], where φ is set to the steady-state solution of
the Eq. (11), is performed.
∂φ
+ sign(φ)(|∇φ | − 1) = 0, (11a)
∂ t̄
where t̄ is a pseudo time for φ and sign(φ) is given by [Peng et al. (1999)]
φ
sign(φ) = (11b)
φ + |∇φ|2 (∆x2 )
2

and is subjected to the following initial condition:


φ (x, 0) = ϕ(x). (11c)

1850072-7
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

3.4. Mass correction


To reduce the inherited mass loss/gain of the level-set method, the global mass
correction of Yap et al. [2006] is employed after redistancing. Correction is achieved
by setting φ to the steady-state solution of Eq. (12):
 
∂φ Md − Mc
= sign(φref ) , (12a)
∂t Md
where t is a pseudo-time and subjected to the following initial condition:
φ (x, 0) = φ(x). (12b)
In Eq. (12a), Md and Mc are the desired mass and the most current mass of the
reference fluid, respectively. Either the +ve or the −ve fluid can be used as the
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

reference fluid.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

For stratified flows, if required, a local mass correction approach [Yap et al.
(2005)] is employed in place of global mass correction to alleviate mass loss/gain
problem of the level-set method. Basically, the flowrate at any given cross-section
instead of mass as in the global mass correction approach is “conserved” by intro-
ducing a correction to the level-set function.

3.5. Local particle redistribution


In the solution process, artificial numerical diffusion of the discretization scheme
drives some of the nanoparticles to diffuse across the interface into the −ve fluid
region. This process is of a purely numerical origin and therefore is not physical. If
left untreated, the amount of nanoparticles in the −ve fluid region will increase with
time nonphysically. To alleviate this problem, the following local particle redistri-
bution procedure is performed. The idea is to redistribute the nanoparticles in the
−ve fluid control volumes (CVs) near the interface uniformly into the adjacent +ve
nanofluid CVs. For example, in a two-dimensional setting shown in Fig. 2, at the

Fig. 2. Local particle redistribution.

1850072-8
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

end of a given time step, the nanoparticles diffused into the −ve fluid CV (i, j) are
distributed back into the neighboring +ve nanofluid CVs of (i − 1, j + 1), (i, j + 1),
(i + 1, j + 1) and (i + 1, j). This can be achieved by setting the nanofluid volume
fraction in these neighboring CVs as:
ϕi,j ∆Vi,j
ϕin,jn = ϕin,jn + fin,jn , (13a)
∆Vin,jn
where
Sin,jn
fin,jn = in=i+1,jn=j+1 , (13b)
in=i−1,jn=j−1 Sin,jn

1, if φi,j = −ve, φin,jn = +ve
Sin,jn = (13c)
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

0, otherwise.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

To complete the procedure, the nanoparticle volume fraction φ for CV (i, j) is set
to φ = 0.

4. Solution Procedure
4.1. Numerical method
The transport equations Eqs. (2)–(5) can be written in the form of a generic tran-
sient convection–diffusion equation:
∂(ρ̃Φ)
+ ∇ · (ρ̃uΦ) = ∇ · (Γ̃∇Φ) + S, (14)
∂t
where ρ̃, Γ̃ and S are the appropriate “density”, “diffusion coefficient” and source
term. The source term contains all other terms that cannot be fitted neatly into the
convection or diffusion terms. This generic equation is solved using a finite volume
method [Patankar (1980); Versteeg and Malalasekera (2007)] on a staggered mesh
arrangement. Scalar variables are defined at the node of the CVs. The staggered
velocity components are defined at the surface of the CVs. The convection term is
modeled using a second-order upwind scheme with SUPERBEE limiter [Roe (1983)]
implemented via a deferred correction manner. A fully implicit scheme is used for
time integration. The velocity–pressure coupling of the Navier–Stokes equations is
handled with the SIMPLER algorithm.
To capture the evolving interface accurately, the level-set method requires higher
order numerical schemes. The evolution of the level-set function (Eq. (10)) and
its redistancing (Eq. (11)) are spatially discretized with WENO5 [Jiang and Peng
(2000)] and integrated using TVD-RK2 [Shu and Osher (1988)]. These schemes are
computationally intensive. To reduce the computational effort, the level-set method
is implemented in a narrow-band procedure [Peng et al. (1999)], in which the level-
set function is solved only within a band of certain thickness from the interface.
This reduces one order of computational effort.

1850072-9
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

4.2. Solution algorithm


The overall solution procedure for the presented method can be summarized as
follows:
(1) Specify the initial conditions (i.e., t = 0) for u, p, T , ϕ and φ.
(2) Advance the time step to t + ∆t.
(3) Solve Eqs. (2) and (3) for u|t+∆t and p|t+∆t .
(4) Solve Eq. (4) for T |t+∆t .
(5) Solve Eq. (5) for ϕ|t+∆t .
(6) Solve Eq. (10) for φ|t+∆t and perform redistancing (Eq. (11)).
(7) Repeat steps (3) to (6) until the solution converges.
(8) Perform global/local mass correction.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

(9) Perform nanoparticle redistribution (Eq. (13)).


Int. J. Comput. Methods Downloaded from www.worldscientific.com

(10) Repeat steps (2) to (9) for all time steps.

5. Results and Discussions


5.1. Properties of fluids and nanoparticle
In this study, silicone oil (−ve fluid) and water–alumina nanofluid (+ve nanofluid)
are used. The properties of silicone oil are set to
Silicone oil:
3
ρ− = 920 kg/m , (15a)

µ− = 0.081 Pa s, (15b)

k− = 0.17 W/mK, (15c)

c− = 1970 J/kg K. (15d)


Water–alumina nanofluid consists of alumina nanoparticles in a water base fluid.
The properties of both water [NIST (2011)] and alumina nanoparticle [Morell
(1987)] are temperature-dependent and can be evaluated from the following cor-
relations as employed in [Ryzhkov and Minakov (2014)]. With these correlations,
the properties of water–alumina nanofluid are then calculated from Eqs. (9) and
plotted in Fig. 3.
Water:
ρbf = 999.86 + 6.1238 × 10−2 T − 8.3131 × 10−3 T 2 + 6.4236 × 10−5 T 3
− 3.9530 × 10−7 T 4 + 1.0808 × 10−9 T 5 , (16a)

µbf = 1.7825 × 10−3 − 5.8439 × 10−5 T + 1.2592 × 10−6 T 2 − 1.6986 × 10−8 T 3


+ 1.2480 × 10−10 T 4 − 3.7458 × 10−13 T 5 , (16b)

1850072-10
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

(a) ρ+ (b) µ+
Int. J. Comput. Methods Downloaded from www.worldscientific.com

(c) ρ+ cp+ (d) k+

(e) DB (f) DT

Fig. 3. Physical properties of water–alumina nanofluid and alumina nanoparticle.

1850072-11
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

kbf = 0.5609 + 1.9488 × 10−3 T − 1.0133 × 10−6 T 2 − 1.2840 × 10−7 T 3


+ 6.2118 × 10−10 T 4 , (16c)

cbf = 4218.79 − 3.1667T + 9.5040 × 10−2 T 2 − 1.3890 × 10−3 T 3

+ 1.0722 × 10−5 T 4 − 3.2042 × 10−8 T 5 . (16d)

Alumina nanoparticle:
dp = 25 nm, (17a)

ρp = 3921.71 − 8.5625 × 10−2 T, (17b)

kp = 5.5 + 34.5e−0.0033T , (17c)


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

cp = 1044.60 + 0.1742(273.15 + T ) − 2.796 × 107 /(273.15 + T )2 .


Int. J. Comput. Methods Downloaded from www.worldscientific.com

(17d)
In a pure water-silicon oil system, interfacial tension depends weakly on temper-
ature [El-Hamouz (2007); Peters and Arabali (2013)]. In El-Hamouz [2007], the
measurement shows an interfacial tension decrease of merely 2% when temperature
varies from 25◦ C to 80◦ C, i.e., ≈0.04%/◦C. In Peters and Peters and Arabali [2013],
it was found to interestingly slightly increase 4% ovet T from 18◦ C to 35◦ C, i.e.,
≈0.2%/◦ C. In a separate interfacial tension measurement for a different system,
i.e., water-titanium oxide nanofluid and mineral oil [Sohel Murshed et al. (2008)],
a decrease of roughly 6% is found over the range of 30◦ C to 45◦ C, i.e., ≈ 0.4%/◦ C.
For problems considered in this paper, the temperature ranges from 32◦ C to 42◦ C.
Variation of interfacial tension due to temperature is therefore expected to be small,
in the range of a few percent. In the absence of complete interfacial tension data
for water–alumina nanofluid and silicone oil system at different temperatures and
nanoparticle volume fraction, a constant value of σ = 0.03 N/m is used. With this,
interfacial tension gradient-induced Marangoni effect is not considered. Neverthe-
less, demonstration including Marangoni effect will also be presented. For the tem-
perature range of 32◦ C (305 K) to 42◦ C (315 K), interfacial tension is assumed to
decrease 5% linearly as
σ = σo − 0.005(T − 305)σo , (18)
where σo = 0.03 N/m and T in K. To account for the Marangoni effect due
to this variation of interfacial tension, an additional source term of (n̂ ×
(dσ/dT )∇T ) × n̂δ(φ) is included into the RHS of the momentum equation (Eq. (3)).

5.2. Verification
To the best knowledge of the authors, the current work is the first attempt to
model two-phase nanofluid-fluid flow with Buongiorno model for nanofluid. There-
fore, there is no existing result of which direct verification is possible. The cur-
rent model consists of two essential components: (1) level-set approach to capture

1850072-12
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

Fig. 4. Schematic of natural convection in a square cavity.


Int. J. Comput. Methods Downloaded from www.worldscientific.com

the interface of a two-phase flow and (2) nanofluid flow with Buongiorno model.
The approach taken in the current paper is to verify these two components sep-
arately. After both components (1) and (2) are verified, the approach is deemed
indirectly verified for two-phase nanofluid-fluid flow. Component (1) was verified in
the authors’ earlier work [Yap and Chai (2012); Li et al. (2015)] and therefore will
not be repeated here. To verify component (2), the limiting case of a single-phase
nanofluid flow with Buongiorno model is attempted. Specifically, verification is con-
ducted for natural convection of a single-phase water–alumina nanofluid in a square
cavity with prescribed temperature gradient (Fig. 4). This problem is selected as
there exists excellent results in Corcione et al. [2013] to be compared with. The left
and right walls are maintained at a temperature of, respectively, Th and Tc . Both
the top and bottom walls are insulated. The flow is essentially buoyancy-driven.
It should be noted that as there is only one phase in the domain, the interface
capturing procedure of the level-set method is temporarily disabled.
Solutions for two cases with different uniform initial nanoparticle volume frac-
tions, i.e., ϕo = 0 and ϕo = 0.04, are obtained. Of course, the case of ϕo = 0 cor-
responds to the situation without alumina nanoparticles, i.e., pure water. Figure 5
shows the present steady-state solutions: streamlines for the case of ϕo = 0.04
(Fig. 5(a)), vertical velocity component v along the horizontal midplane of the
cavity (Fig. 5(b)), dimensionless temperature θ = TTh−T c
−Tc for the case of ϕo = 0.04
(Fig. 5(c)) and temperature T along the horizontal midplane of the cavity (Fig. 5(d))
superimposed on the results of Corcione et al. [2013]. The present solutions, for both
cases of ϕo = 0 and ϕo = 0.04, agree well with those of Corcione et al. [2013]. The
effect of alumina nanoparticles can be carefully observed in Figs. 5(b) and 5(d).
This exercise verifies all components of the present model except the interface
capturing procedure of the level-set method. The implemented interface capturing
procedure of the level-set method has been verified in the authors’ earlier work [Yap
and Chai (2012); Li et al. (2015)] and therefore will not be repeated here.

1850072-13
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

(a) Streamline (ϕo = 0.04) (b) ν


Int. J. Comput. Methods Downloaded from www.worldscientific.com

(c) θ (ϕo = 0.04) (d) T

Fig. 5. Solutions for natural convection of water–alumina nanofluid in a square cavity.

5.3. Water–alumina nanofluid-filled cavity with a rising


silicone oil drop
Figure 6 shows a rectangular cavity with Lo = 0.015 m. The lower half of the
cavity is filled with water–alumina nanofluid encapsulating a silicone oil drop. The
drop has a radius of Ro = 0.25Lo and is initially located at (0.5Lo, 0.5Lo ). Both
the water–alumina nanofluid and silicone oil drop are at an initial temperature
of Th = 315 K. The upper half of the cavity contains pure water at an initial
temperature of Tc = 305 K. All four walls are thermally insulated. As water and
water–alumina nanofluid are miscible in each other, the water and water–alumina
nanofluid are effectively considered as one single phase. Water, nanofluid and silicone
oil drop are then buoyantly driven into motion in the presence of density difference,
where the silicone oil drop rises in the process. At the same time, gradual transport
and migration of alumina nanoparticles into the pure water occur. Clearly, the

1850072-14
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 6. Schematic of nanofluid-filled cavity with a rising drop.

nanoparticles will not be uniformly distributed and therefore justifies the use of
Eq. (5) in determining the temporal and spatial distribution of nanoparticles. Of
course, heat is transferred simultaneously from the hot nanofluid and silicone oil to
pure water.
Given the symmetry of the problem, computations are made for only left half
of the domain. The following initial and boundary conditions apply:
Initial conditions:
u = 0, in whole domain, (19a)

Th , in nanofluid and silicone oil
T = (19b)
Tc , in pure water,

ϕo , in nanofluid
ϕ= (19c)
0, otherwise.

Boundary conditions:


u = 0, ∂v = 0, at symmetric plane
∂x (19d)

u = 0, at walls,

n · k∇T = 0, at walls and symmetric plane, (19e)


 
∇T
n · DB ∇ϕ + DT = 0, at walls and symmetric plane. (19f)
T
Solutions are obtained for three different initial nanoparticle volume fractions, i.e.,
ϕo = 0, 0.02 and 0.04 and plotted, respectively, in Figs. 7, 9 and 10. Depicted in

1850072-15
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

these plots are the velocity, temperature and concentration fields at t = 0, 0.09375,
0.1875, 0.2813, 0.375 and 0.45 s. Note that in order to avoid overcrowding the figure,
only one in every eight velocity vectors is plotted. The first plot in Fig. 7 shows
the grid-independent test conducted. The superimposed solutions were obtained on
two consecutively refined meshes of 100 × 400 CVs with ∆t = 1.875 × 10−4 s and
150 × 600 CVs with ∆t = 9.375 × 10−5 s. A mesh of 100 × 400 CVs with ∆t =
1.875 × 10−4 s is sufficient to capture all the essential features of the solution. This
is also verified for the cases of ϕo = 0.02 and 0.04. Therefore, the same mesh size is
used in these cases.
The effect of variable interfacial tension (Eq. (18))-induced Marangoni effect is
shown in Fig. 8. The interfacial profile and temperature field are almost exactly
identical to those of a constant interfacial tension, except small difference in
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

the interfacial profile at t = 0.45 s. This to a certain extent justifies ignoring


Int. J. Comput. Methods Downloaded from www.worldscientific.com

temperature-induced interfacial tension variation for the cases considered here.


The effect of nanoparticles on the flow and temperature fields is clearly demon-
strated by comparing Fig. 7 to either Fig. 9 or Fig. 10. The presence of nanoparticles
modifies effectively the fluid properties and affects transport process directly in a
fully coupled manner. Generally, the flow structure is very complex with multiple
vortical structures formed trailing the rising silicone oil drop. For a higher initial
nanoparticle volume fraction ϕo , the drop rises more rapidly, revealed upon careful
examination of the drop location and the slightly more deformed diffused nanofluid–
water interface as early as at t = 0.09375 s for the three cases considered. Solely
invoking the Archimedes’ buoyancy principle is sufficient to arrive at a reasonable
explanation. The buoyancy force acting on the silicone oil drop (equal to the weight
of the displaced nanofluid) depends on both the drop volume and the nanofluid den-
sity. The drop volume is identical for all cases and remains so during the process
because of incompressibility. With a denser nanofluid upon increasing ϕo , a larger
induced buoyancy force drives the upward motion of the drop. For example, with
reference to Fig. 3(a), at the early stage, both the nanofluid and silicone oil drop
are at a temperature of around T = 305 K; increasing ϕo = 0−0.04 results in a
roughly 9% denser nanofluid and therefore 9% increase in the buoyancy force. As a
result, the drop rises more rapidly with a higher ϕo . Besides, the more rapidly ris-
ing drop deforms more in the process, attributed to a higher drag force induced by
both higher viscosity (because of a higher ϕo , see Fig. 3(b)) and velocity gradient.
Nonetheless, for all three cases, the drop maintains roughly a circular to an ellip-
tical shape. If the temperature dependence of silicone oil viscosity and interfacial
tension are accounted for (both decrease with a rising temperature), an even larger
deformation is then expected.
Energy and nanoparticle transports are now discussed. The similarity between
the temperature and the nanoparticle volume fraction fields for both cases of
ϕo = 0.02 and 0.04 is overtly striking, suggestive of a similar dominant transport
mechanism. As the drop rises, multiple intricate vortical flow structures form and
evolve. Examination of the velocity, temperature and nanoparticle volume fraction

1850072-16
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 7. Solutions for ϕo = 0 at t = 0, 0.09375, 0.1875, 0.2813, 0.375 and 0.45 s.

1850072-17
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 8. Solutions for constant and variable σ at t = 0.1875, 0.2813, 0.375 and 0.45 s.

fields reveals the close association of energy and nanoparticle transport to these
flow structures. In fact, these flow structures drive dominantly both energy and
nanoparticle transport.
The very sharp features in temperature and nanoparticle volume fraction fields
suggest, at least for the time and length scales considered, a minimal role of other
transport mechanisms, i.e., diffusion in energy transport and Brownian diffusion
and thermophoresis in nanoparticle transport. This is however not unexpected given
the very small coefficients associated with these transport mechanisms, for example,
both DB and DT are in the order of 10−11 m2 /s.
Generally, for these cases, from the perspective of modification in thermal con-
ductivity and heat capacity, the presence of nanoparticles does not help much in heat
transfer enhancement from the hotter lower region to the upper cooler region. The

1850072-18
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 9. Solutions for ϕo = 0.02 at t = 0, 0.09375, 0.1875, 0.2813, 0.375 and 0.45 s.

nanoparticles themselves are bounded by the same dominant transport mechanism


as in energy transfer and therefore have not been transported into upper cooler
region. However, the increase in buoyancy force due to the presence of nanoparti-
cles induces larger convective heat transfer.

1850072-19
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 10. Solutions for ϕo = 0.04 at t = 0, 0.09375, 0.1875, 0.2813, 0.375 and 0.45 s.

Shown in Fig. 11 is the interesting case of replacing the pure water in the upper
region with nanofluid. Now the enclosure contains only nanofluid encapsulating the
silicone oil drop. The volume fraction of nanoparticle is uniform with ϕo = 0.02.
Both the flow and temperature fields have very similar characteristics to that of the
case with pure water encapsulating a silicone oil drop (ϕo = 0, Fig. 7). The vortical

1850072-20
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 11. Solutions for ϕo = 0.02 (nanofluid only with silicone oil drop) at t = 0, 0.09375, 0.1875,
0.2813, 0.375 and 0.45 s.

structure trailing the rising silicone drop is larger though. The drop rises faster than
any cases considered as now the drop is always surrounded by a denser nanofluid
and therefore driven by a larger buoyancy force. Region of low nanoparticle fraction
forms near the trailing edge of the silicone drop at t = 0.2813 s, see the fourth plot
in Fig. 11. Once this region forms, it extends gradually into the flow under the

1850072-21
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 12. Solutions for Ro = 0.15Lo and ϕo = 0.02 at t = 0, 0.09375, 0.1875, 0.2813, 0.375 and
0.45 s.

influence of the vortical flow structure. This also attributes to the dominance of
convective transport with minimal diffusion effect.
The effect of drop radius is now investigated. With drop radius increased from
Ro = 0.15Lo (Fig. 12), 0.25Lo (Fig. 9) to 0.35Lo (Fig. 13), the larger drop is closer
to the wall and therefore experiences a larger wall effect. Increasingly constrained

1850072-22
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 13. Solutions for Ro = 0.35Lo and ϕo = 0.02 at t = 0, 0.09375, 0.1875, 0.2813, 0.375 and
0.45 s.

by the wall, the drop undergoes a larger deformation leading to a higher drag. As a
result, the drop rises slowly. The flow still dominantly drives energy and nanopar-
ticle transport. Both temperature and nanoparticle distribution are least affected
for the case of the smallest drop, i.e., Ro = 0.15Lo, as its motion only generates a
weak vortical flow.

1850072-23
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

Fig. 14. Schematic of a stratified two-phase nanofluid-fluid flow in a channel.


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

5.4. Stratified two-phase nanofluid–fluid flow in a channel


Figure 14 shows a channel with length 4Lo and height Lo . Three streams of fluids,
i.e., pure water, water–alumina nanofluid and silicone oil, all at temperature Tc
flow into the channel at the inlet with a uniform velocity uo . The thickness of these
layers at the inlet are, respectively, 0.3Lo , 0.3Lo and 0.4Lo. Pure water and water–
alumina nanofluid are miscible in each other and therefore are considered as one
phase, whereas silicone oil is the other immiscible phase. Both the upper and the
lower walls are maintained at a higher temperature of Th . The nanoparticles in
the nanofluid are gradually transported into pure water under the driving forces of
convection, diffusion and thermophoresis.
To specify the problem completely, the following initial and boundary conditions
apply:

Initial conditions:

u = 0, in whole domain, (20a)


T = Th , in nanofluid, pure water and silicone oil, (20b)
ϕ = 0, in nanofluid, pure water and silicone oil. (20c)

Boundary conditions:


 u = uo , v = 0, at inlet

 ∂u
= 0, v = 0, at outlet (20d)

 ∂x


u = 0, at walls,


T = Tc , at inlet

 ∂T
= 0, at outlet (20e)

 ∂x


T = Th , at walls,

1850072-24
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer






ϕo , in nanofluid

ϕ = , at inlet
0, otherwise
(20f)

  

 ∇T

n · DB ∇ϕ + DT = 0, at walls and outlet.
T
For demonstration purpose, the governing parameters are set to Lo = 0.0001 m,
uo = 1 m/s, Th = 315 K, Tc = 305 K and ϕo = 0.04. This gives a Reynolds number
of roughly
2ρbf uo Lo
Re = ∼ 200,
µbf
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

well within the limit of a laminar flow. Note that this definition of Re is used only
Int. J. Comput. Methods Downloaded from www.worldscientific.com

as a reference and would be even lower if the properties of silicone oil are used
instead. Four scenarios differ in the location of the nanofluid layer at the inlet are
considered; these are


 0.0 ≤ y/Lo ≤ 0.0


0.0 ≤ y/Lo ≤ 0.3
ϕo = 0.04 for . (21)

 0.0 ≤ y/Lo ≤ 0.6



0.3 ≤ y/Lo ≤ 0.6

Obviously, the first scenario of φo = 0.04 for 0.0 ≤ y/Lo ≤ 0.0 corresponds to the
situation without nanofluid at the inlet, i.e., only pure water as the lower layer and
silicone oil as the upper layer. The last scenario of ϕo = 0.04 for 0.3 ≤ y/Lo ≤ 0.6
corresponds to the situation shown in Fig. 14 in which the water–alumina nanofluid
flows in the inlet as the middle layer ranging from 0.3 ≤ y/Lo ≤ 0.6.
Plotted in Fig. 15 are the steady-state solutions for the scenario of ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.0 (without nanofluid, i.e., pure water and silicone oil only). The
interfacial profiles obtained on two consecutively refined meshes of 320 × 80 CVs
with ∆t = 6.25 × 10−7 s and 640 × 160 CVs with ∆t = 3.125 × 10−7 s are super-
imposed for the purpose of grid-independent study. A mesh of 320 × 80 CVs with
∆t = 6.25 × 10−7 s is sufficient to resolve the solution. The interface quickly flattens
not too far away from the inlet. Computations were also made using a longer chan-
nel with a length of 8Lo . There is no difference observed in the solution. Therefore,
the channel length of 4Lo is sufficient for the cases considered. Also shown in Fig. 15
are the velocity and temperature fields, both with the interfacial profile superim-
posed. Generally, a steeper transverse temperature gradient occurs adjacent to the
lower wall (in contact with pure water), compared with that of the upper wall (in
contact with the silicone oil). Further coupled with a larger thermal conductivity
of pure water in comparison to silicone oil, heat transfer into the water layer at the
lower wall is therefore larger than that into the silicone oil layer at the upper wall.
From the temperature field in Fig. 15, the temperature around the interface along

1850072-25
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 15. Solutions for ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.0 (without nanofluid, i.e., pure water and
silicone oil only).

the channel varies very little. The resulted Marangoni effect is verified to be minute
and can be safely ignored.
Upon careful examination of the interfacial profiles for the scenarios considered
(Eq. (21)), the presence of nanoparticles does not alter noticeably the interfacial
profile and therefore hydrodynamic behavior. These plots for different scenarios are
therefore not presented. For stratified two-phase flow, the interfacial profile is dom-
inantly affected by the fluids’ viscosities. Even though the viscosity of the combined
miscible nanofluid–water layer (i.e., nanofluid and pure water layers) nominally
increases in the presence of nanoparticles, the viscosity of this layer (of the order
of 10−3 –10−4 Pa s) is still at least one order of magnitude smaller than that of the
silicone oil layer (0.08 Pa s). The effect of an increased viscosity of the combined
nanofluid–water layer is therefore negligible. The interfacial profile is still dictated
primarily by the silicone oil viscosity. This leads to very similar interfacial profiles
for all these scenarios.
The nanoparticle distribution for different scenarios is shown in Fig. 16. The-
oretically, nanoparticles can be transported transversely from the nanofluid layer
into the pure water layer. This transverse transport relies on three mechanisms:
convection, Brownian diffusion and thermophoresis. For convection, note that from

1850072-26
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

(a) 0.0 ≤ y/Lo ≤ 0.3


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

(b) 0.0 ≤ y/Lo ≤ 0.6

(c) 0.3 ≤ y/Lo ≤ 0.6

Fig. 16. Nanoparticle distribution for ϕo = 0.04.

Fig. 15, the velocity field is almost parallel within the pure water layer (or com-
bined nanofluid–water in other scenarios). Therefore, convection does not transport
much nanoparticles from the nanofluid layer into the pure water layer. Directed
against nanoparticle concentration gradient, Brownian diffusion drives nanoparticles
from high nanoparticle volume fraction region (nanofluid layer) to low nanoparticle
volume fraction region (pure water layer). Finally, directed against temperature gra-
dient, thermophoresis drives nanoparticles from the high temperature region near
the lower wall toward the low temperature region at the channel center. In the
scenario of ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3, thermophoresis tends to drive the
nanoparticles from the lower nanofluid layer into the middle pure water layer. In
contrast, in the scenario of ϕo = 0.04 for 0.3 ≤ y/Lo ≤ 0.6, thermophoresis tends to
drive the nanoparticles back into the middle nanofluid layer. However, with similar
physical conditions as those prevail in Sec. 5.3, transport due to both Brownian
diffusion and thermophoresis is minimal. Therefore, nanoparticles mostly remain

1850072-27
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

within the nanofluid layer while they are transported downstream dominantly by
convection.
In the scenario of ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3 (Fig. 16(a)), the lower
nanofluid layer becomes thinner along the channel as it is squeezed by both the
middle pure water layer and the more viscous upper silicone oil layer. The aver-
age nanoparticle volume fraction within the nanofluid layer is higher than that in
the same layer at the inlet. Similar higher average nanoparticle volume fraction
within the nanofluid layer is also observed to occur in the scenario of ϕo = 0.04
for 0.3 ≤ y/Lo ≤ 0.6 (Fig. 16(b)). For the scenarios of 0.0 ≤ y/Lo ≤ 0.6 and
0.3 ≤ y/Lo ≤ 0.6, there is a thin region adjacent to the interface with higher
nanoparticle volume fraction. This accumulation of nanoparticles is driven by trans-
verse convection pushing the interface downwards as the more viscous silicone oil
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

layer expands during the transient flow development. This thin region of high
Int. J. Comput. Methods Downloaded from www.worldscientific.com

nanoparticle volume fraction remains given weak diffusion transport.


The predicted temperature fields for these scenarios, both in the combined
nanofluid–water layer and silicone oil layer, are very similar and difficult to be
distinguished visually. Therefore, these plots are not presented. For quantification
of heat transfer performance, the Nusselt numbers at the upper (NuU ) and lower
(NuL ) walls are evaluated. These are defined, respectively, as
 
2Lo  ∂T 
NuU = , (22a)
Th − Tb  ∂y 
y=Lo
 
2Lo  ∂T 
NuL = , (22b)
Th − Tb  ∂y y=0
where the bulk temperature Tb is given by
 Lo
ρcp uT dy
Tb = 0 Lo . (23)
0 ρc p udy
A simple verification of the calculation procedure is performed by setting all fluid
layers to have identical properties, i.e., the problem effectively reduces to heat
transfer in a single-phase channel flow with prescribed wall temperature. Under
this condition, both NuU and NuL at the fully developed region are found to be
7.5415. This value is very close to the Nusselt number of a steady-state single-phase
fully developed laminar flow of Nu = 7.54.
The Nusselt number and the bulk temperature along the channel for steady-state
condition are plotted in Fig. 17. Note that in the first (ϕo = 0.04 for 0.0 ≤ y/Lo ≤
0.0) and last (ϕo = 0.04 for 0.3 ≤ y/Lo ≤ 0.6) scenarios, there is no nanoparticle in
the combined nanofluid–water layer directly adjacent to the lower wall (Fig. 16(c)).
However, in the second (ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3) and the third (ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.6) scenarios, there are nanoparticles in the combined nanofluid–
water layer directly adjacent to the lower wall (Figs. 16(a) and 16(b)).
Generally, the Nusselt number at the lower wall NuL decreases along the channel.
NuL numbers for the first (ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.0) and fourth (ϕo = 0.04

1850072-28
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer


by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

(a) NuU and NuL (b) Tb

Fig. 17. Heat transfer performance for ϕo = 0.04.

for 0.3 ≤ y/Lo ≤ 0.6) scenarios are almost identical. However, with the presence
of nanoparticles adjacent to the lower wall (and similar concentration near the
lower wall, see Figs. 16(a) and 16(b)), NuL numbers for the second (ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.3) and third (ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.6) scenarios are
almost identical and most importantly lower than those of the first (ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.0) and fourth (ϕo = 0.04 for 0.3 ≤ y/Lo ≤ 0.6) scenarios.
Note that Nusselt number can be construed as the ratio of convective to conduction
heat transfer. With increased thermal conductivity attributed to nanoparticles, a
higher increase in conduction heat transfer from the wall to the combined nanofluid–
water layer relative to convective heat transfer occurs. In fact, a similar reduction
in Nusselt number is also observed under conditions considered in the work of
Nimmagadda and Venkatasubbaiah [2015].
For the upper wall, NuU does not seem to be affected much in the presence
of nanoparticles in the lower combined nanofluid–water layer. It remains almost
identical for all scenarios. This is not unexpected as the flows in the upper silicone
oil layer are hydrodynamically very similar, and constant thermophysical silicone
oil properties dictating heat transfer at the upper wall are employed resulting in
similar heat transfer characteristics. Generally, NuU decreases rapidly at the inlet
and then flattens, i.e., a profile similar to that of a typical Graetz problem.
The bulk temperature Tb can be used to quantify heat transferred into the fluids.
It increases with the amount of heat absorbed. As the cumulative heat absorbed
from the wall along the channel increases, Tb increases along the channel for all
scenarios. The bulk temperature along the channel is almost identical for the first
(0.0 ≤ y/Lo ≤ 0.0) and the fourth scenarios (0.3 ≤ y/Lo ≤ 0.6). Comparing to
the first scenario (0.0 ≤ y/Lo ≤ 0.0), the presence of nanoparticles in the middle
layer in the fourth scenario (0.3 ≤ y/Lo ≤ 0.6) does not alter the heat transfer
performance. This suggests that augmentation of thermal properties attributed to

1850072-29
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

(a) NuU and NuL (b) Tb

Fig. 18. Heat transfer performance for ϕo = 0.02.

presence of nanoparticles (e.g., increase in thermal conductivity of 15–20% with


ϕo = 0.04, see Fig. 3(d)) contributes minimally to internal heat transfer within
nanofluid.
For the second (0.0 ≤ y/Lo ≤ 0.3) and third scenarios (0.0 ≤ y/Lo ≤ 0.6), the
bulk temperature along the channel is identical and more importantly is higher than
those of the first (0.0 ≤ y/Lo ≤ 0.0) and the fourth scenarios (0.3 ≤ y/Lo ≤ 0.6).
This is attributed to the presence of nanoparticles near the lower wall in the second
(0.0 ≤ y/Lo ≤ 0.3) and third scenarios (0.0 ≤ y/Lo ≤ 0.6). In the third scenario
(0.0 ≤ y/Lo ≤ 0.6), the whole combined nanofluid–water layer contains nanopar-
ticles that promote heat transfer not only from the lower wall into the nanofluid–
water layer but also within the nanofluid–water layer itself. The second scenario
(0.0 ≤ y/Lo ≤ 0.3) is most interesting. Now, the nanofluid–water layer literally has
only half of the amount of nanoparticles in the third scenario, and yet, the second
scenario achieves a Tb that is identical to the third scenario. This implies that heat
transfer enhancement from the lower wall into the combined nanofluid–water layer
is much more dominantly affected by the nanoparticles near the lower wall.
From a practical point of view, in terms of heat transfer enhancement, nanoparti-
cles only need to be present near the wall where heat transfer occurs. The presence
of nanoparticles beyond this region does not alter the heat transfer characteris-
tics much. Therefore, a much smaller amount of nanoparticles is actually required.
Maintaining this concentrated distribution of nanoparticles near the wall is easier
in a laminar flow where convective transport is much stronger than diffusive trans-
port given the very small diffusivity. However, with enhanced turbulent mixing in
a turbulent flow, this task then becomes challenging.
The effect of ϕo is now considered. With all other governing parameters remain
unchanged, the nanoparticle volume fraction is now lowered to ϕo = 0.02. Figure 18
shows the plots of NuU , NuL and Tb along the channel. Note that the scenario of

1850072-30
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

ϕo = 0.02 for 0.0 ≤ y/Lo ≤ 0.0 is exactly identical to that of ϕo = 0.04 for
0.0 ≤ y/Lo ≤ 0.0, as the nanofluid layer is absent and therefore there are only the
pure water and silicone oil layers.
The variation of NuU along the channel is almost identical for all scenarios
considered. From Fig. 18(a), NuL for the scenario of ϕo = 0.02 for 0.3 ≤ y/Lo ≤ 0.6
is identical to the corresponding scenario of ϕo = 0.04 for 0.3 ≤ y/Lo ≤ 0.6. This
reinforces the fact that nanoparticles only minimally enhance heat transfer within
the nanofluid itself. This enhancement is negligible in a comparison with pure water
(base fluid) as the baseline. For the scenarios of ϕo = 0.02 for 0.0 ≤ y/Lo ≤ 0.3
and 0.0 ≤ y/Lo ≤ 0.6, thermal conductivity augmentation is less compared with
the corresponding scenarios of ϕo = 0.04. Therefore, on a relative basis, conduction
heat transfer is weaker leading to a higher NuL .
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

Along the channel, Tb for the scenarios of ϕo = 0.02 for 0.0 ≤ y/Lo ≤ 0.3 and
Int. J. Comput. Methods Downloaded from www.worldscientific.com

0.0 ≤ y/Lo ≤ 0.6 is slightly lower than that of the corresponding scenarios with
ϕo = 0.04. The bulk temperature differences ∆Tb at the outlet between the scenarios
with (0.0 ≤ y/Lo ≤ 0.3 and 0.0 ≤ y/Lo ≤ 0.6) and without (0.0 ≤ y/Lo ≤ 0.0 and
0.3 ≤ y/Lo ≤ 0.6) nanoparticles adjacent to the lower wall are roughly 0.1 K and
0.05 K, respectively, for the case of ϕo = 0.04 and ϕo = 0.02. Lowering the amount
of nanoparticles near the lower wall indeed reduces heat transfer performance.
By maintaining ϕo = 0.04 and all other governing parameters fixed, the inlet
velocity is now reduced to uo = 0.1 m/s. This reduces Re and Peclet numbers for
both heat and mass transfers. Diffusion transport becomes relatively more impor-
tant if compared with the original scenarios of uo = 1 m/s. Plotted in Fig. 19 are
the solutions of uo = 0.1 m/s with ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3. The increase in
the strength of transverse diffusion transport can easily be identified for nanopar-
ticles (resulting in a more uniform nanoparticle distribution within the nanofluid
layer particularly near the inlet in comparison with Fig. 16(a)) and energy (affect-
ing the fluids temperature well into the center of the channel in comparison with
Fig. 15). As a result, NuL is lower than the corresponding scenarios with uo = 1 m/s
(Fig. 20). Similar observation of having NuL higher for the scenarios of ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.0 and 0.3 ≤ y/Lo ≤ 0.6 than those for the scenarios of ϕo = 0.04
for 0.0 ≤ y/Lo ≤ 0.3 and 0.0 ≤ y/Lo ≤ 0.6 can be seen from Fig. 20.
NuU is however interestingly different from the scenarios with uo = 1 m/s. With
uo = 0.1 m/s, NuU for the scenarios of ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3 and
0.0 ≤ y/Lo ≤ 0.6 is almost identical but higher than those of ϕo = 0.04 for 0.0 ≤
y/Lo ≤ 0.0 and 0.3 ≤ y/Lo ≤ 0.6 (which are themselves almost identical). For
the scenarios of ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3 and 0.0 ≤ y/Lo ≤ 0.6, the
fluids absorb comparatively more heat from the lower wall due to the presence of
nanoparticles near the lower wall. At any given location, the fluids temperature
is generally higher resulting in a smaller effective transverse temperature gradient.
Heat diffusion from the upper wall to upper silicone oil layer decreases. With a lower
diffusion heat transport relatively to the convective heat transport, NuU increases.
With a reduced uo , the resident time for fluids in the channel is higher (roughly

1850072-31
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Fig. 19. Solutions for ϕo = 0.04 for 0.0 ≤ y/Lo ≤ 0.3 for uo = 0.1 m/s.

(a) NuU and NuL (b) Tb

Fig. 20. Heat transfer performance for ϕo = 0.04 and uo = 0.1 m/s.

1850072-32
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

10 times) compared with the corresponding scenarios with uo = 1 m/s. The fluids
absorb more heat in the flow process. Therefore, Tb along the channel and at the
outlet is much higher (Fig. 20).

6. Concluding Remarks
This paper presents a model for two-phase nanofluid-fluid flow with heat transfer.
Nanofluid is modeled using Buongiorno’s approach. The model is verified against
solution for natural convection of nanofluid in a cavity. For demonstration, solu-
tions for flow and transport in a water–alumina nanofluid-filled cavity with a rising
silicone oil drop and stratified flow of water–alumina nanofluid, water and silicone
oil in a channel are presented. In this study, water–alumina nanofluid is employed.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

Nevertheless, the model works regardless of the type of base fluid and nanoparticles,
Int. J. Comput. Methods Downloaded from www.worldscientific.com

provided that their properties are given.


Generally, nanoparticle distribution is mainly driven by convection. Diffusion
transport of nanoparticles is weak. Therefore, nanoparticles within a nanofluid can
only minimally diffuse into adjacent miscible base fluid. From heat transfer perspec-
tive, the presence of nanoparticles near the wall helps in augmenting heat transfer.
However, the nanoparticles within the bulk do not promote internal heat transfer
much. Therefore, only a thin nanoparticle layer is required adjacent to the wall for
better heat transfer performance, as demonstrated in the case study of a stratified
flow.

References
Abbasi, M. and Baniamerian, Z. [2014] “Analytical simulation of flow and heat transfer of
two-phase nanofluid (stratified flow regime),” Int. J. Chem. Eng. 2014, 474865.
Akgül, M. B. and Pakdemirli, M. [2016] “Numerical analysis of mixed convection of
nanofluids inside a vertical channel,” Int. J. Comput. Meth. 13, 1650012.
Alloui, Z., Vasseur, P. and Reggio, M. [2011] “Natural convection of nanofluids in a shallow
cavity heated from below,” Int. J. Therm. Sci. 50, 385–393.
Bachok, N., Ishak, A. and Pop, I. [2012] “Flow and heat transfer characteristics on a
moving plate in a nanofluid,” Int. J. Heat Mass Transf. 55, 642–648.
Brackbill, J. U., Kothe, D. B. and Zemach, C. [1992] “A continuum method for modelling
surface tension,” J. Comput. Phys. 100, 335–354.
Buongiorno, J. [2005] “Convective transport in nanofluids,” J. Heat Transf. 128, 240–250.
Chamkha, A. J. and Abu-Nada, E. [2012] “Mixed convection flow in single- and double-lid
driven square cavities filled with water–Al2 O3 nanofluid: Effect of viscosity models,”
Eur. J. Mech. B-Fluids 36, 82–96.
Cheng, L. and Liu, L. [2013] “Boiling and two-phase flow phenomena of refrigerant-based
nanofluids: Fundamentals, applications and challenges,” Int. J. Refrig. 36, 421–446.
Cianfrini, C., Corcione, M., Habib, E. and Quintino, A. [2014] “Buoyancy-induced con-
vection in Al2 O3 /water nanofluids from an enclosed heater,” Eur. J. Mech. B-Fluids
48, 123–134.
Ciloglu, D. and Bolukbasi, A. [2015] “A comprehensive review on pool boiling of nanoflu-
ids,” Appl. Therm. Eng. 84, 45–63.

1850072-33
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

Corcione, M. [2011] “Empirical correlating equations for predicting the effective thermal
conductivity and dynamic viscosity of nanofluids,” Energ. Convers. Manage. 52, 789–
793.
Corcione, M., Cianfrini, M. and Quintino, A. [2013] “Two-phase mixture modeling of nat-
ural convection of nanofluids with temperature-dependent properties,” Int. J. Therm.
Sci. 71, 182–195.
Corcione, M., Cianfrini, M. and Quintino, A. [2015] “Enhanced natural convection heat
transfer of nanofluids in enclosures with two adjacent walls heated and the two opposite
walls cooled,” Int. J. Heat Mass Transf. 88, 902–913.
Duursma, G., Sefiane, K., Dehaene, A., Harmand, S. and Wang, Y. [2015] “Flow and heat
transfer of single- and two-phase boiling of nanofluids in microchannels,” Heat Transf.
Eng. 36, 1252–1265.
El-Hamouz, A. [2007] “Effect of surfactant concentration and operating temperature on
the drop size distribution of silicon oil water dispersion,” J. Disper. Sci. Technol.
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

28, 797–804.
Int. J. Comput. Methods Downloaded from www.worldscientific.com

Elshehabey, H. M. and Ahmed, S. E. [2015] “MHD mixed convection in a lid-driven cavity


filled by a nanofluid with sinusoidal temperature distribution on the both vertical walls
using Buongiorno’s nanofluid model,” Int. J. Heat Mass Transf. 88, 181–202.
Fang, X., Wang, R., Chen, W., Zhang, H. and Ma, C. [2015] “A review of flow boiling heat
transfer of nanofluids,” Appl. Therm. Eng. 91, 1003–1017.
Farooq, U. and Lin, Z.-L. [2014] “Nonlinear heat transfer in a two-layer flow with nanofluids
by OHAM,” J. Heat Transf. 136, 021702.
Garoosi, F., Bagheri, G. and Talebi, F. [2013] “Numerical simulation of natural convection
of nanofluids in a square cavity with several pairs of heaters and coolers (HACs) inside,”
Int. J. Heat Mass Transf. 67, 362–376.
Garoosi, F., Jahanshaloo, L., Rashidi, M. M., Badakhsh, A. and Ali, M. E. [2015] “Numer-
ical simulation of natural convection of the nanofluid in heat exchangers using a Buon-
giorno model,” Appl. Math. Comput. 254, 183–203.
Gerken, W. J. and Oehlschlaeger, M. A. [2017] “Modeling nanofluid sessile drop evapora-
tion,” Heat Mass Transf. 53, 2341–2349.
Jiang, G.-S. and Peng, D. [2000] “Weighted ENO schemes for Hamilton–Jacobi equations,”
SIAM J. Sci. Comput. 21, 2126–2143.
Khanafer, K. and Vafai, K. [2011] “A critical synthesis of thermophysical characteristics
of nanofluids,” Int. J. Heat Mass Transf. 54, 4410–4428.
Kuznetsov, A. V. and Nield, D. A. [2010] “Natural convective boundary-layer flow of a
nanofluid past a vertical plate,” Int. J. Therm. Sci. 49, 243–247.
Lee, J. and Mudawar, I. [2007] “Assessment of the effectiveness of nanofluids for single-
phase and two-phase heat transfer in micro-channels,” Int. J. Heat Mass Transf.
50, 452–463.
Li, H. Y., Yap, Y. F., Lou, J. and Shang, Z. [2015] “Numerical modelling of three-fluid
flow using the level-set method,” Chem. Eng. Sci. 126, 224–236.
Malvandi, A. and Ganji, D. D. [2014] “Effects of nanoparticle migration on force convection
of alumina/water nanofluid in a cooled parallel-plate channel,” Adv. Powder Technol.
25, 1369–1375.
Mohyud-Din, S. T., Usman, M. and Bin-Mohsin, B. [2017] “A study of heat and mass
transfer of nanofluids arising in biosciences using Buongiorno’s model,” Int. J. Comput.
Methods 14, 1750018.
Morell, R. [1987] Handbook of Properties of Technical and Engineering Ceramics. Part 2:
Data Reviews. Section 1: High-Alumina Ceramics (Her Majesty’s Stationery Office,
London).

1850072-34
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

Nanofluid-Fluid Two-Phase Flow and Heat Transfer

Nimmagadda, R. and Venkatasubbaiah, K. [2015] “Conjugate heat transfer analysis of


micro-channel using novel hybrid nanofluids (Al2 O3 + Ag/Water),” Eur. J. Mech.
B-Fluids 52, 19–27.
NIST Chemistry Webbook [2011] http://webbook.nist.gov.
Ogolo, N. A., Olafuyi, O. A. and Onyekonwu, M. O. [2012] “Enhanced oil recovery using
nanoparticles,” SPE Saudi Arabia Section Technical Symposium and Exhibition, 8–11
April, Al-Khobar, Saudi Arabia, SPE-160847-MS.
Osher, S. and Sethian, J. A. [1988] “Fronts propagating with curvature-dependent speed:
Algorithms based on Hamilton-Jacobi formulations,” J. Comput. Phys. 79, 12–49.
Patankar, S. V. [1980] Numerical Heat Transfer and Fluid Flow (Hemisphere Publisher,
New York).
Peng, D., Merriman, B., Osher, S., Zhao, H. and Kang, M. [1999] “A PDE-based fast local
level-set method,” J. Comput. Phys. 155, 410–438.
Peters, F. and Arabali, D. [2013] “Interfacial tension between oil and water measured with
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

a modified contour method,” Coll. Surf. A 426, 1–5.


Int. J. Comput. Methods Downloaded from www.worldscientific.com

Pouranfard, A. R., Mowla, D. and Esmaeilzadeh, F. [2015] “An experimental study of drag
reduction by nanofluids in slug two-phase flow of air and water through horizontal
pipes,” Chin. J. Chem. Eng. 23, 471–475.
Roe, P. L. [1983] “Some contributions to the modelling of discontinuous flows,” Proc. 15th
Summer Seminar on Applied Mathematics, 27 June–8 July, La Jolla, CA, pp. 23–34.
Ryzhkov, I. I. and Minakov, A. V. [2014] “The effect of nanoparticle diffusion and ther-
mophoresis on convective heat transfer of nanofluid in a circular tube,” Int. J. Heat
Mass Transf. 77, 956–969.
Saien, J. and Bamdadi, H. [2012] “Mass transfer from nanofluid single drops in
liquid−liquid extraction process,” Ind. Eng. Chem. Res. 51, 5157–5166.
Sert, I. O., Sezer-Uzol, N. and Kakaç, S. [2016] “Numerical approaches for convective
heat transfer with nanofluids,” Chapter 7, in Microscale and Nanoscale Heat Transfer :
Analysis, Design, and Application, eds. Rebay, M., Kakaç, S. and Cotta, R. M. (CRC
Press, Boca Raton), pp. 183–205.
Shu, C.-W. and Osher, S. [1988] “Efficient implementation of essentially non-oscillatory
shock capturing schemes,” J. Comput. Phys. 77, 439–471.
Sohel Murshed, S. M., Tan, S.-H. and Nguyen, N.-T. [2008] “Temperature dependence
of interfacial properties and viscosity of nanofluids for droplet-based microfluidics,”
J. Phys. D: Appl. Phys. 41, 085502.
Sussman, M., Smereka, P. and Osher, S. [1994] “A level set approach for computing solution
to incompressible two-phase flow,” J. Comput. Phys. 114, 146–159.
Taylor, R., Coulombe, S., Otanicar, T., Phelan, P., Gunawan, A., Lv, W., Rosengarten,
G., Prasher, R. and Tyagi, H. [2013] “Small particles, big impacts: A review of the
diverse applications of nanofluids,” J. Appl. Phys. 113, 011301.
Turkyilmazoglu, M. [2014] “Nanofluid flow and heat transfer due to a rotating disk,”
Comput. Fluids 94, 139–146.
Van Gorder, R. A. [2013] “Rare exact solution to a model of two-phase flow consisting of
nanofluid adjacent to a clear fluid,” Int. J. Heat Mass Transf. 61, 201–208.
Van Gorder, R. A., Prasad, K. V. and Vajravelu, K. [2012] “Convective heat transfer in the
vertical channel flow of a clear fluid adjacent to a nanofluid layer: A two-fluid model,”
Heat Mass Transf. 48, 1247–1255.
Versteeg, H. K. and Malalasekera, W. [2007] An Introduction to Computational Fluid
Dynamics: The Finite Volume Method, 2nd Edition (Prentice Education Limited,
England).

1850072-35
2nd Reading
January 3, 2018 14:9 WSPC/0219-8762 196-IJCM 1850072

G. Qiangshun et al.

Wang, P., Lv, J., Bai, M., Wang, Y., Hu, C. and Zhang, L. [2014] “Numerical simulation
on the flow and heat transfer process of nanofluids inside a piston cooling gallery,”
Numer. Heat Trans. A-Appl. 65, 378–400.
Wang, P., Lv, J., Bai, M., Li, G. and Zeng, K. [2015] “The reciprocating motion charac-
teristics of nanofluid inside the piston cooling gallery,” Powder Technol. 274, 402–417.
Yang, C., Li, W. and Nakayama, A. [2013] “Convective heat transfer of nanofluids in a
concentric annulus,” Int. J. Therm. Sci. 71, 249–257.
Yap, Y. F. and Chai, J. C. [2012] “Level-set method for multiphase flows,” Comput. Therm.
Sci. 4, 507–515.
Yap, Y. F. and Li, H. Y. [2015] “Modeling of two-phase nanofluid-fluid flow with heat
transfer,” in The 26th Int. Symp. Transport Phenomena, 27 September–1 October,
Leoben, Austria, Paper 118, pp. 1–10.
Yap, Y. F., Chai, J. C., Toh, K. C., Wong, T. N. and Lam, Y. C. [2005] “Numerical
modeling of unidirectional stratified flow with and without phase change,” Int. J. Heat
by UNIVERSITY OF NEW ENGLAND on 01/07/18. For personal use only.

Mass Transf. 48, 477–486.


Int. J. Comput. Methods Downloaded from www.worldscientific.com

Yap, Y. F., Chai, J. C., Wong, T. N., Toh, K. C. and Zhang, H. Y. [2006] “A global mass
correction scheme for the level-set method,” Numer. Heat Tr. B-Fund 50, 455–472.
Yu, W., France, D. M., Routbort, J. L. and Choi, S. U. S. [2008] “Review and comparison
of nanofluid thermal conductivity and heat transfer enhancements,” Heat Transf. Eng.
29, 432–460.

1850072-36

You might also like