Você está na página 1de 12

International Journal of Heat and Mass Transfer 88 (2015) 433444

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Numerical models for heat conduction and natural convection


with symmetry boundary condition based on particle method
Liang Yangyang a, Sun Zhongguo a,, Xi Guang a, Liu Ling b
a
b

School of Energy and Power Engineering, Xian Jiaotong University, Xian 710049, China
Department of Mechanical and Aerospace Engineering, Utah State University, Logan, UT 84322, USA

a r t i c l e

i n f o

Article history:
Received 15 February 2015
Received in revised form 29 April 2015
Accepted 29 April 2015
Available online 15 May 2015
Keywords:
Moving particle semi-implicit method
Heat transfer
Heat conduction
Nature convection
Symmetry boundary model

a b s t r a c t
Numerical models of heat conduction and natural convection are developed based on Meshless Moving
particle semi-implicit (MPS) method to study the unsteady complex ow coupling with heat transfer.
Fundamental heat exchange process is systematically investigated in one dimension and extended to
two dimensions without internal heat source. The energy conservation equation is discretized on particles and coupled with ow by momentum change that caused by temperature difference. Both Dirichlet
and Neumann boundary conditions are proposed with mirror image technique and the numerical results
are veried to have high accuracy against benchmark solutions (theoretical or experimental, and or other
numerical solutions) for several representative examples. To improve the calculation efciency, symmetry boundary models are established and applied in transient heat conduction and natural convection
heat transfer. The results using the symmetry boundary are comparable and reliable to the
full-domain results, both of which agree well with the experimental and other numerical solutions.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Unsteady complex ow including free surface ow, phase
change and multiphase ow is usually accompanied by large deformation of the uid phase (such as crushing and merging).
Simulating these problems is usually a challenge for conventional
mesh-based numerical methods, because the grid reconstruction
can compromise computational efciency and accuracy to a large
extent. In addition, if the mentioned unsteady complex ow is coupled with heat transfer, namely the Unsteady Complex Heat-Fluid
Coupling (UC-HFC) problem, the uid ow and heat transfer would
interact with each other, and it would be rather complicated to
solve them integrally by any existed conventional numerical
methods.
However, Lagrangian particle method, such as the smooth particle hydrodynamics (SPH) method [1] and the Moving particle
semi-implicit (MPS) method [2] have been introduced to address
these computational difculties in ows. Based on the movement
of particles and no mesh-dependent, these methods can be used
for simulating uid ow with arbitrary deformation [3,4].
This advantage has made these methods eligible for analyzing
the UC-HFC problem on the unsteady complex ow part, but
on the other part of heat transfer and their coupling relationship,

Corresponding author.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2015.04.105
0017-9310/ 2015 Elsevier Ltd. All rights reserved.

the numerical models of heat transfer and the high-delity implementation that incorporates multi-physical elds (including the
ow, pressure and temperature elds) should be established based
on SPH and MPS. Such a numerical scheme could be applied to
study the UC-HFC issues, such as lm cooling and evaporation
etc. Whereas, the difculties eventually lie on the discrete numerical model of heat transfer, and the integration of the temperature
eld with the ow and pressure elds within the MPS.
Indeed, heat is exchanged via three primary modes, as the
heat conduction, heat convection and radiative heat transfer. In
this paper, only the heat conduction and heat convection are discussed. In heat conduction, no macroscopic movement of physical particles happens but the motion of phonons and electrons
from the high- to the low-temperature domain. In heat convection, heat is transferred by the movement of uids (when hot uids and cold uids are blended). Convection could be further
divided into two forms, as the natural convection and forced convection. Natural convection happens when the uid is driven to
move by internal temperature or concentration difference, while
forced convection is mainly driven by external forces. As the
basic heat transfer modes, the numerical models of conduction
and natural convection will be studied respectively in the following sections.
Among the previous studies based on SPH or MPS, Cleary [5]
explored integral approximants and effective thermal conductivity
to the thermal conduction equation which could estimate

434

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

second order derivatives with continuous heat ux. Chaniotis [6]


proposed re-meshing technique which is the periodic
re-initialization of the particle locations being distorted by the ow
map. Typical heat-uid coupling problems (double shear layer,
lid-driven ow in a cavity, natural convection in a differentially
heated cavity) were calculated and discussed. Vishwakarma [7]
studied the heat conduction transfer in irregular geometries (eccentric annulus, a long slab with an array of circular holes inside) by
adopting the radial and rectangular particle arrangement and their
combination. Szewc [8] proposed a more generally SPH approach
for dealing with the natural convection considering the Ga number
in non-isothermal ows. Leroy [9] established more general expression of buoyancy model based on Boussinesq approximation in laminar and turbulent ows. In addition, Mansour [10]
investigated
natural
convection
heat
transfer
in
complex-wavy-wall enclosed cavity lled with nano-uid, the
results shown that the heat transfer performance could be optimized by tuning the wavy-surface geometry parameters. Aly [11]
studied the unsteady natural/mixed convection in a non-Darcy porous cavity, the inuence of Darcy numbers and other parameters on
the ow pattern and heat transfer characteristics were shown.
With the help of MPS method, Chen [12] simulated the single
bubble dynamics during nucleate ow boiling, in which the heat
transfer was expressed in the form of liquid heat exchange and
the latent heat. Tian [13] investigated single steam bubble condensation behaviors in subcooled water, the inuence of liquid subcooling degree on the bubble deformation, size and lifetime was
discussed. Zhang [14,15] studied the RayleighBenard convection
heat transfer problems using the MPS method. In addition, some
improvements for enhancement of stability and accuracy of MPS
method were proposed. Khayyer and Gotoh constructed many
new schemes such as CMPS (gradient model) [16], MPS-HS (higher
order source term for the Pressure Poisson Equation) [17], MPS-HL
(higher order Laplace model) [18,19], MPS-ECS (multi-term source
for Poisson Pressure Equation) [20], MPS-GC (corrective matrix for
gradient model) [20] and MPS-FDS (rst order accurate density
smoothening) [21] to calculate a large number of ow issues. It
is hopeful to introduce these rened differential operator models
to different Heat-Flow cases.
Since the study of this article focuses on the numerical model of
heat transfer based on MPS method, the classical case of the ow
and nature convection heat transfer between concentric cylindrical
annulus is employed. In this work, MPS-based numerical models of
heat conduction and natural convection are developed. The
method is veried by several typical examples including 1D and
2D heat conduction cases in both steady and transient states, and
also the natural convective heat transfer case between concentric
cylinders mentioned above. In addition, in order to improve the
computational efciency for axisymmetrical cases, symmetry
boundary models are proposed, in which only half of the calculation domain is included in the calculation.
2. Numerical method
Originally proposed by Koshizuka in 1996 [2] to simulate
incompressible ow, the MPS method has been largely improved
in recent years and widely applied to solve ow problems [24]
involving large deformation. The basic concept of the MPS method
is to represent a continuous uid as many moving particles that
have their own positions, masses, pressures, momentums, and
energies. The interaction between particles is governed by kernel
function, which is also used to describe the differential operators
in the motion equations. The motion of particles at different time
instants is tracked in the Lagrangian system. With the motion of
individual particles solved, the ow characteristics in the entire
eld can be obtained.

2.1. Governing equation


The original MPS method only has a mass conservation equation
and a momentum conservation equation, i.e. Eqs. (1) and (2). To
enable thermal analysis coupled with uid analysis, a third equation, i.e. Eq. (3), is needed so that energy conservation is enforced.

Dq
qr  u 0
Dt

Du
rp lr2 u qf B
Dt

qcp

DT
kr2 T  q
Dt

Here, q is the uid density, u is the velocity vector, p is the pressure, l is the viscosity, f is the external force vector, cp is the specic heat capacity, T is temperature, k is the thermal conductivity,
and q is the internal heat source. Note that, here, the momentum
conservation Eq. (2) is modied with a new simplied term B,
which accounts for the effect of temperature eld on the momentum eld. The form of B depends on the physic problems.
2.2. Particle interaction models
The interaction between particles can be dened by the kernel
function:

8
2
<
1  rre
r < r e
wr
:
0
r P r e

where r jri  rj j is the distance between the ith and jth particles.
re is a cut-off distance that denes a circular effective area surrounding the target particle. The target particle is then assumed
to interact only with those particles positioned within the effective
area. Compared with the hyperbolic kernel function introduced by
Koshizuka [2], the kernel function described in Eq. (4) has several
advantages and disadvantages. The value of the hyperbolic kernel
function [2] is very large when r gets close to 0, which effectively
prevents particle clustering. However, the absolute value of the
rst-order derivative of such a function is also large when r is close
to 0, which may result in large uctuation of the kernel function. In
contrast, using Eq. (4) could avoid dramatic change of the kernel
function in the vicinity of the target particle, while the relatively
low value of w = 1.0 at r 0 is not effective in preventing particle
clustering. For better simulation results, kernel function needs to
be selected based on the specic problems.
Particle number density is dened based on the kernel function:

hni i

X
wjrj  ri j

ji

Note that uid density should be a constant for incompressible


uid. Therefore, the particle number density should keep as a constant of n0 because uid density is proportional to the particle
number density in the MPS method [2]. Based on the kernel function, the gradient operator at the ith particle should be evaluated
by a weighted average from its neighboring particles to obtain a
conservation form:

hr/ii

d X /j /i
rj  ri wjrj  ri j
n0 ji jrj  ri j2

Here, d equals to 2 for two-dimensional problems. The Laplace


operator r2 / in the governing equations should take the form of:

hr2 /ii

2d X /j  /i
wjrj  ri j
n0 ji jrj  ri j2

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

435

2.3. Algorithm
The MPS method utilizes a semi-implicit time marching
scheme. At each time step, the energy conservation equation is calculated implicitly to obtain the temperature eld, as well as the
source term B which is a force term induced by temperature difference. Then the viscosity term and the source term in the momentum equation are calculated explicitly, leading to the temporal
velocities and positions of the particles. With the updated positions
of particles, a temporal particle number density ni can be calculated. Note that the particle number density should be a constant
for incompressible ow, so ni needs to be modied to n0 by the
mass conservation equation. The pressure Poisson equation is then
obtained which can be solved implicitly. This leads to a solution of
the pressure eld, based on which the velocities and positions of
the particles can be updated. Finally, by tracking the ow characteristics of all particles, information about the ow eld can be
obtained.
3. Numerical models and examples
The accuracy and reliability of the MPS method have already
been veried by solving many cases that involve only uid ow.
However, the new computational algorithm coupling with uid
ow and heat transfer requires combined examples with different
systems. In the following sections, several examples that involve
uid ow, heat conduction, and/or heat convection will be discussed, which includes: (1) one-dimensional steady/transient heat
conduction, (2) two-dimensional steady/unsteady heat conduction, (3) natural convection heat transfer, (4) symmetry boundary
with models above.
3.1. Heat conduction
3.1.1. One-dimensional heat conduction
One-dimensional steady heat conduction, as well as the transient heat conduction is studied as two typical cases in this section.
The status of steady heat conduction is actually the same as the
nal stage of the transient heat conduction. The governing equations for heat conduction can be written in the following form:

DT
ar2 T
Dt

where a is the thermal diffusivity dened by a k=qcp . When DT


is
Dt
equal to zero, the Eq. (8) represents the steady heat conduction.
Based on the Laplace model of Eq. (7), the discrete formula of
Eq. (8) can be written as:
n1
n1
T n1
 T ni
2  a  d X T j  T i
i

wjrn1
 rn1
j
j
i
2
n1
n0
Dt
 rn1
j
ji jrj
i

Fig. 1 shows the calculating settings of these two cases. To meet


the one dimension requirement, a panel with a thickness of 0.1 m
and a height of 1.0 m is employed. Due to the large length/width
ratio
is
10,
both
cases
could
be
considered
as
quasi-one-dimensional issue. For steady case, the temperatures
of the left and right edges are xed to be T1 and T2 respectively,
and for the transient case, the initial temperature of the panel is
assumed T0 and the temperature of the left edge is given constantly
Tw. No internal heat source involved here.
To check the independence of particle size for the simulation,
the steady case is repeatedly calculated by MPS method using different initial particle spacing Dxi from 0.005 m to 0.001 m. Table 1
compares the numerical solutions to the analytical ones at four different locations at x = 0.02, 0.04, 0.06, and 0.08 m. According to Eq.
(8), the steady heat conduction is formulated as a Laplace equation,

Fig. 1. Models of one-dimensional steady heat conduction (left) and transient heat
conduction (right).

1
and the temperature distribution follows Tx T 2 T
x T 1 where
d
Tx denotes the point-wise temperature within the panel. The
results indicate that smaller Dxi (more particles) can achieve
higher calculation accuracy. The absolute error is smaller in the
center of the panel and higher at the two edges. In addition,
the estimated temperature is less than the analytical solution on
the high temperature side but larger on the low side, and the absolute errors distribution presents the symmetric status about the
midline of the panel. The numerical solution eventually agrees well
with the analytical solution in a linear distribution (temperature
with respect to x), and the relative error is less than 0.5% on most
of the points, which shows good accuracy of the MPS method.
The simulation results also illustrate a linear relationship
between the computational error and the initial particle separation
Dxi . Fig. 2 plots the computational absolute error at x = 0.06 m and
x = 0.08 m with respect to Dxi (ve cases are selected), and the relationship of which can be well tted by linear equations as

f e 3:05  105 2:92Dxi and f e 9:32  105 8:75Dxi , respectively. The agreement implying that the MPS method has
rst-order accuracy on one-dimensional steady heat conduction.
Similar to the steady case, Fig. 3 shows the point-wise temperature solutions of transient heat conduction case which are evaluated in time advance from simulation and theory, respectively. The
results show that the high temperature region spreads out along
the x coordinate axis as the simulation proceeds. The temperature
distributions agree well with the theoretical solution in the form of :



Tx; t  T w
x
erf p
T0  Tw
2 at

10

where Tx; t denotes the temperature at time t and location x, a is


the thermal diffusivity, and erf is error function. As noted in a previous work by Zhang et al. [15], the original Laplace model [2]
applied to the energy equation may overestimate thermal diffusivity. Therefore, another improved Laplace model is employed as
shown in equation 7.
Fig. 4 demonstrates the computational error (x = 0.01 m at
t = 2.05.0 s) as functions of the initial particle spacing, Dxi .
Again, linear relationships are identied. Comparing Fig. 4 with
Fig. 2, the computational accuracy is higher in the steady case.
Indeed, error builds up during time stepping, which leads to higher
error in the simulation of unsteady case.
3.1.2. Two-dimensional heat conduction
Since the mechanism of two-dimensional steady heat conduction is mostly the same as the one-dimensional case, and the calculation result has a high accuracy with the analytical solution, we
pay more attention on the two-dimensional unsteady cases with

436

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Table 1
Comparison between the calculation results (Tcal,C) and the analytical results (Tanal,C) and the relative error (e, %) in different initial particle spacing (Dxi , m).
Tanal|x = 0.02 = 16.00

Tanal|x = 0.04 = 12.00

Tanal|x = 0.06 = 8.00

Tanal|x = 0.08 = 4.00

Dxi

Tcal

Tcal

Tcal

Tcal

0.005
0.002
0.001

15.956
15.982
15.991

0.275
0.113
0.056

11.985
11.993
11.997

0.125
0.058
0.025

8.015
8.007
8.003

0.188
0.088
0.038

4.044
4.018
4.009

1.100
0.450
0.225

Fig. 2. Computational error at x = 0.06 m, x = 0.08 m versus the initial particle


spacing (Dxi ). Solid squares and circles: computational errors. Line: tted linear
equations.

Fig. 4. Computational error at x = 0.01 m and t = 2.0 s, 3.0 s, 4.0 s, 5.0 s versus the
initial particle spacing (Dxi ). Solid symbols: computational errors. Lines: tted
linear equations.

Fig. 3. Temperature versus x coordinates. Four time instants are considered, i.e.
0.1 s, 0.5 s, 1.0 s, and 5.0 s. Solid symbols: the MPS solutions. Lines: the analytical
solutions.

added to form a symmetry pair about the upper edge of the panel
(Neumann boundary). The paired two particles are assumed to
0
have the same physical quantities, i.e. /i /i , where i and i0
denote the interior particle and the corresponding mirror particle,
respectively. All the particles including the mirror particles will
join the calculation. This technique will guarantee zero gradients
of physical quantities across the Neumann boundary in the MPS
calculation.
The contour map of the temperature distribution and the temperature variations for typical positions as y = 0.5 m for cases 1 and
y = 1.0 m for case 2 are evaluated in time series in Fig. 7 and Fig. 8,
respectively. Fig. 7 shows the temperature rising process with very
smooth isotherms. In Fig. 8, the solid circles represent MPS results
and the lines represent the previously SPH results [22], apparently
these two methods generate very consistent results. When the system is still in unsteady state (when t is small), parabolic temperature distribution is found in both cases. However, when the system
reaches steady state (t = 0.5 s), temperature is nally uniformly
distributed within the panel to match the prescribed boundary
condition.

typical heat boundary conditions. The boundary condition with


given temperature is actually named as the Dirichlet boundary
condition, while the one given a function of temperature, such as
@T=@y 0, is called Neumann boundary condition. A square panels
side length is L = 1.0 m and the calculating settings are shown in
Fig. 5, with no internal heat source.
Two cases are simulated here, and case 1 is calculated with the
Dirichlet boundary conditions on all the four edges, while case 2 is
mostly the same as case 1 but with Neumann boundary condition
of @T=@y 0 applied on the upper edge.
To address the Neumann boundary condition, a mirror image
technique is introduced as shown in Fig. 6. For each of the interior
particle, there is a corresponding mirror particle (ghost particle)

3.2. Natural convection heat transfer


3.2.1. Natural convective heat transfer between concentric cylinders
The classic natural convection heat transfer process in concentric cylinders is employed in this paper to study the natural convection model of MPS method. Fig. 9(a) schematically depicts the
simulation model as two concentric cylinders with radii, R1 and
R2 , respectively. The computational domain between these two
cylinders is lled with air, and the two cylinders are kept isothermal with a higher temperature of T 1 in the inner cylinder wall and
a lower temperature of T 2 in the outer one. Fig. 9(b) shows the initial particle description of the model with 50283 particles, and the

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Dh @ 2 h @ 2 h

Ds @X 2 @Y 2

437

14

where Pr m=a is a ratio between the uids momentum diffusivity


(viscosity) and heat diffusivity. In this example, Pr is set as a con-

Fig. 5. Settings for two-dimensional heat conduction with different heat boundary
conditions.

stant of 0.71. The Rayleigh number (Ra gbDTD31 =am) is a dimensionless parameter associated with buoyancy lift, the value of
which donates the level of a heat transfer between conduction
and convection. Larger Rayleigh number means more heat convection. Here, g, b, DT, and D1 are the gravity acceleration, volume
expansion coefcient, temperature difference, and the diameter of
the inner cylinder, respectively. Eq. (11) involvesn, the particle
number density, which is proportional to uid density [2] in the
MPS method.
The parameters are all non-dimensional-normalized, as the
radiuses of the cylinders are R1 = 0.5 and R2 = 1.3, besides,
L=D1 = 0.8, T 1 = 1.0, T 2 = 0.0, and Pr m=a = 0.71. Six simulations
have been performed in different RaD1 with respect to
RaL 1  102 , 1  103 , 1  104 , 2  104 , 3  104 and 5  104 , i.e.

Fig. 6. The schematic diagram of Neumann boundary condition by mirror image


method.

initial temperature is assumed to vary linearly in the domain from


the inner wall to the outer wall (Fig 9(c)).
In order to deal with the buoyancy due to temperature difference in the nature convection problem, the Boussinesq approximation is applied, which supplies the term B in Eq. (2). The MPS
simulation then solves several dimensionless equations dened
as below:

Dn
nr  U 0
Ds

11

DU
@P
@2U @2U

Pr

Ds
@X
@X 2 @Y 2
DV
@P
@2V @2V

Pr

Ds
@Y
@X 2 @Y 2

!
12

!
Ra  Pr  h

t=0.01 s

t=0.01 s

13

RaD1 9:76  104 represents the situation of RaL 5  104 [23]


(the subscript D1 and L stand for the characteristic lengths of corresponding cases, respectively).
Similar to the mesh independence study in mesh-based
method, a particle size independent study is needed in the MPS
method. The particle size should be set in a range, in which it is
small enough to make the calculation result stable and independent with the particle size. In this paper, we simulate the natural
convection heat transfer in concentric cylinders with three different particle sizes (as 0.02, 0.016 and 0.01, and denoted as case1,
case2 and case3, respectively). The local equivalent conductivities
in inner and outer cylinders under these three cases are shown
in Fig. 10.
The curves in Fig. 10 demonstrate that the three numerical solutions are all converge to the experimental results [23]; furthermore
the solutions of case3 with smaller particle size are more close to
those of experiment. With the particle size decreasing (from case1
to case3), the average relative error between MPS solution and
experiment is reduced from 7.24% to 2.12% in the inner cylinder,
and from 10.66% to 6.20% in the outer cylinder. The results conrm
that the accuracy of solutions improve with renement of particle
size. Meanwhile, considering the calculation cost, we choose particle size as 0.01 (the same with case3) in the calculations.

t=0.05 s
(a) Case 1

t=0.10 s

t=0.05 s
(b) Case 2

t=0.10s

Fig. 7. Temperature eld at different time instants (t = 0.01 s to 0.10 s).

438

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

The isotherms of the numerical solutions are exhibited in


Fig. 11. The results illustrate that the cases with small Rayleigh
number has a similar temperature distribution as heat conduction,
and whereas, the cases with large Rayleigh number undergo more
uid ow performance of the system, which means more evident
convection. At RaD1 1:95  102 , convective ow is limited and
the temperature eld changes little comparing to the initial distribution of Fig. 9(c). When RaD1 is increased to 1:95  104 , convective
ow is enhanced, and hot uid invades the region initially occupied by cold uid. This hot uid jet transports mass and heat,
and perturbs the temperature eld drastically. The width of the
jet becomes thinner whereas the temperature of it decreases with
larger Rayleigh number RaD1 9:76  104 , which can be explained
by faster convection and diffusion. In all cases, the hot uid ows
upwards along the inner wall until hitting the upper wall, and then
gets cooled and sinks, which agrees with the principle of buoyancy
forces.
Fig. 12 compares the MPS simulation result (RaD1 9:76  104 )
with the experimental results [23]. Radial variation of the dimensionless temperature is evaluated at six different angles.
Denition of angle and R can be found in the inset of Fig. 12. The
MPS results are shown to have good agreement with experimental
results, particularly in regions close to the cylinder walls. It seems
the calculation has higher accuracy in the domain that heat conduction plays the main role over heat convection, which is also
widely found in the simulations by other numerical methods. On
the other hand, some local uctuations are identied in the MPS
solution due to the interpolation method used in the simulation.
Fig. 13 further compares the solutions of local equivalent conductivities (keq) by the MPS method, experiment [23], and the

Fig. 10. Local equivalent conductivities for both inner and outer cylinders under
different particle size.

nite difference method (FDM) [23]. The local equivalent conductivity is dened as a ratio of the actual Nusselt (Nu) number which
includes both convection and conduction over the Nu number with
conduction only. In this example, keq is evaluated for both cylinders at different angles. The MPS results show good consistency
to the FDM results and as well as the experimental results. We note
that for the inner cylinder, the MPS and FDM results diverge from
each other when h > 150 . The FDM curve rises and the MPS curve
drops, while the experimental data show slight declination. The

Fig. 8. Temperature distribution variation in time series along given lines (a) case1, y = 0.5 m. (b) case2, y = 1.0 m. Solid circles: MPS results. Lines: FO-SSPH solutions [22].

(a)

(b)

(c)

Fig. 9. Model and setting of the convective heat transfer problem (a) basic parameters (b) initial particle discrete (c) initial temperature eld.

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

(a) RaD1 = 1.95 102

(c) RaD1 = 1.95 104

439

(b) RaD1 = 1.95 103

(d) RaD1 = 9.76 104

Fig. 11. Isotherms for different RaD1 for nal steady status.

solution by the MPS method is found more consistent with the


experimental results. However, according to Fig. 11, the isotherms
are increasingly dense from 0 to 180 degree in the vicinity of the
inner cylinder, which means temperature gradient in the radial
direction increases from 0 to 180 degree. Based on the increase
of temperature gradient, the value of keq should increase as well,
which, however, does not agree with experimental and the MPS
simulation results. The potential errors associated with the experiment and the FDM computations are not discussed.
Fig. 14 compares the local equivalent conductivities keq calculated by using the MPS and FDM methods [23]. Good agreement
is found particularly in the outer cylinder, and in the inner cylinder
from 0 to 150 degree. However, odd ow happens on the surface of
inner cylinder from 150 to 180 degree where the temperature gradient is extremely high. Higher local resolutions, both on calculation and interpolation model, are required to get accurate
prediction. On the basis of the good agreement between the MPS
and experimental or FDM results, it is possible to determine
heat-transfer characteristics of nature convection in enclosures
using MPS method. In addition, by considering its advantage on
simulating unsteady complex ows, we recognize the MPS method
with heat transfer model eligible for studying the UC-HFC issue.
3.2.2. Natural heat convection in a differentially heated cavity
As another typical verication case, the natural heat convection
in a differentially heated cavity is also discussed in this section. The
left and right walls of the square cavity has constant but different
temperatures, and the top and the bottom walls are in adiabatic
condition, which are depicted in Fig. 15. The Boussinesq approximation is applied to deal with the buoyancy force. The dimensionless governing equations are the same as Eqs. (11)(14).
In the calculation, the Pr number is set as a constant of 0.71. The
Ra number (Ra gbDTL3 =am) is set as 103 and 104 respectively. The

dimensionless particle sizes as l0 = 0.0125, 0.01 and 0.00625 are


used to study the convergence of the calculation and the particle
size independence. As a result, the solution of l0 0:00625 is
selected and the characteristic relative quantities1 are compared
to the benchmark solution [24] in Table 2.
As shown in Table 2, the results in different Ra numbers generally agree well with the data of benchmark, and the relative errors
are almost less than 23%, which demonstrate the accuracy of the
presented MPS models. The noticeable errors, such as the maximum error of 8.7% on YNu-max when Ra 103 and 6.31% on Numin
when Ra 104 are all about Nusselt number and their positions
which are constantly difcult to get high accuracy with numerical
simulations. Even though, the result by MPS method in this paper
is comparable to other solutions [25]. In addition, according to the
experience of the authors, the accuracy of single point could be further improved by increasing the calculation resolution or using
some improved models [18,20,26].
3.3. Symmetry boundary model
3.3.1. Symmetry boundary model on heat conduction
The two cases which are shown in Fig. 5 of Section 3.1.2 are calculated again in this section with the symmetrical axis at x = 0.5 m.
All the settings are the same except only half of the calculation
domain is involved in the simulation. The mirror image technique
1
Umax, YUmax: maximum horizontal velocity and position on the vertical mid-plane
of the cavity; Vmax, XVmax: maximum vertical velocity and its position on the
horizontal mid-plane of the cavity; Nuave, Nu1/2, Nu0: the average Nusselt number
throughout the cavity, on the vertical mid-plane and on the vertical boundary of the
cavity at x = 0; Numax, YNu-max: the maximum value and its position of the local
Nusselt number on the vertical boundary of the cavity at x = 0; Numin,YNu-min: the
minimum value and its position of the local Nusselt number on the vertical boundary
of the cavity at x =0.

440

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Fig. 12. Radial variation of dimensionless temperature at six different angles (Inset denes the angle, h, and the radial position, R, for any point of interest).

Fig. 13. Local equivalent conductivities for both inner and outer cylinders.

is employed here to describe the temperature on the symmetry


boundary (symmetrical axis). The initial particle distributions are
illustrated in Fig. 16, where the blue solid particles represent the
panel interior, the black solid particles stand for the panel edge,
and the green hollow particles represent the ghost particles. In
case2, the red solid particles represent the upper edge with
Neumann boundary condition @T=@y 0.
The temperature description on the symmetry boundary is
essentially the Neumann condition. Fig. 17 shows the setting of
the symmetry boundary which locates on the last layer of blue particles, for each of the interior particles i, there is a corresponding
0
ghost particle i on the other side of the symmetry boundary. The
0
temperature of particles i is equal to particle i to guarantee zero
gradient of @T=@x 0 on the symmetry boundary.
Fig. 18 demonstrates the isotherms of the full- and half- eld
simulations at t = 0.05 s. The left half of the eld is the isotherms
using the whole-eld calculation represented by solid lines, and
the right half is the results of the half-eld calculation represented
by dashed lines. The two results are highly coinciding with each
other, and the curve is very smooth even on the conjunction area
across the symmetrical axis.
The temperature variations on typical lines mentioned in Fig. 8
are compared again quantitatively for the new cases in Fig. 19,
where the solid dots and hollow circles are the full- and

Fig. 14. Local equivalent conductivities for the (a) inner and (b) outer cylinders.
Case1 through Case5 correspond to RaD1 1:95  102 , 1:95  103 , 1:95  104 ,
3:91  104 and 5:86  104 , respectively.

half-eld calculation results, and the lines represent the results


calculated by the SPH method [22]. The maximum relative
difference is less than 0.1% in both cases, and the high accuracy validates the reliability of the symmetry boundary model for heat
conduction.

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

441

Fig. 15. Model of natural convection in a square cavity.

Table 2
Comparison of the MPS Simulation with the Benchmark Solution.[24].
3

Ra 10

Umax
YUmax
Vmax
XVmax
Nuave
Nu1/2
Nu0
Numax
YNu-

Fig. 17. The setting of temperature in the symmetry boundary.

Ra 10

Benchmark
[24]

MPS

Error(%)

Benchmark
[24]

MPS

Error
(%)

3.649
0.813
3.697
0.178
1.118
1.118
1.117
1.505
0.092

3.718
0.827
3.753
0.178
1.124
1.127
1.120
1.556
0.100

1.89
1.72
1.52
0.00
0.54
0.81
0.27
3.39
8.70

16.178
0.823
19.617
0.119
2.243
2.243
2.238
3.528
0.143

15.869
0.830
19.431
0.115
2.298
2.302
2.280
3.656
0.150

1.91
0.85
0.95
3.36
2.45
2.63
1.88
3.63
4.90

0.692
1.000

0.725
1.000

4.77
0.00

0.586
1.000

0.623
1.000

6.31
0.00

max

Numin
YNumin

3.3.2. Symmetry boundary model in natural convection heat transfer


The symmetry boundary model for natural convection heat
transfer is proposed, and only half the domain of the same concentric cylinders is employed (described in Fig. 20). The dash line represents the symmetry boundary, and the expressions of unknown
quantities on this line are U 0, @V=@X 0, @P=@X 0 and
@h=@X 0. Similar to treatment of the last section, the green ghost
particles are employed to implement the conditions in the symmetry line. However, different from the symmetrical boundary for
heat conduction, the symmetry boundary for heat convection
locates in the center between the green particles and blue particles
(as shown in Fig. 21). Based on the particle interaction models in

MPS method, assuming the uniform particles distribution throughout the calculation process, as long as the conditions of U i U i0 ,
V i V i0 , P i P i0 , hi hi0 are met, the expressions of U 0,
@V=@X 0, @P=@X 0, @h=@X 0 on the symmetry line can be
guaranteed.
Fig. 22 illustrates the isotherms at RaD1 9:76  104 using the
full- and half-eld calculations. The two isotherms agree well with
each other except little differences on the upper part which is
marked by an ellipse. The error may come from particle clustering
on the upper domain, and the post-processing of particle method
could be another potential source to generate the deviation.
Again, for the quantitatively comparison, Figs. 23 and 24 show
comparisons on the non-dimensional temperature and local equivalent conductivities within the full-, half-eld and experimental
results [23]. The maximum relative error of the dimensionless temperature between the full-eld and half-eld calculation appears at
the curve of 120. For local equivalent conductivities, the difference
increase to the maximum 4.51% at the location of 30 in inner
cylinder, and it is 3.87% in outer cylinder at the position of 90.
The qualitative and quantitative comparisons in Figs. 2224
demonstrate the effectiveness of symmetry boundary model in this
ow-heat coupled issue. Indeed, the application of symmetric
boundary model can greatly reduce the computational time with
reliable results, which is of great signicance in dealing with
Large-scale problems.

Fig. 16. Particle distributions of symmetry boundaries.

442

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Fig. 18. Comparison of isotherms of full- and half- eld simulation. (t = 0.05 s) (a) case1 (b) case2.

Fig. 19. Comparison of the temperature distribution for full- and half- eld simulation. (a) case1, y = 0.5 m. (b) case2, y = 1.0 m.

(a) computing setting

(b) symmetrical boudanry (c) detail of symmetrical boudanry

Fig. 20. Half-domain model and symmetrical boundary setting for natural convection heat transfer in concentric cylinders.

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Fig. 21. The settings of unknown quantities in the symmetry boundary.

443

Fig. 24. Comparison of local equivalent conductivities for both inner and outer
cylinders.

4. Conclusions

Fig. 22. Comparison of the isotherms between full-(left solid line) and half-eld
(right dash line) calculations.

Within the framework of MPS method, we establish numerical


models of heat conduction and natural convection for analyzing
ow systems coupling with heat transfer. The numerical models
are veried by using several representative examples including
one-dimensional steady and transient heat conduction,
two-dimensional steady and unsteady heat conduction, and the
convective heat transfer. Typical boundary conditions of heat
transfer, such as Dirichlet boundary condition, and Neumann
boundary condition are proposed based on the discrete scheme
of particle method. Mirror image technique is applied to address
different boundaries, and the results show rst-order accuracy
for heat conduction issue. The solution of the ow between concentric cylinders with high accuracy proves the effectiveness of
natural convection model. The symmetry boundary models are
proposed to address the symmetry issues which half-eld calculation is implemented instead of the whole-eld simulations. The
results conclude that the symmetry boundary conditions are accurate, and of great help on saving computational load. Over all, the
accurate and reliable solutions from the heat transfer models
enable the MPS method a probably tool for analyzing unsteady
complex heat-uid coupling ow.

Fig. 23. Comparison of the radial variation of dimensionless temperature at ve different angles.

444

Y. Liang et al. / International Journal of Heat and Mass Transfer 88 (2015) 433444

Conict of interest
None declared.
Acknowledgment
Thanks to Xiao Chen and Wenjin Cao for providing helpful discussions. This work is supported in part by the Natural Science
Foundation of China (NSFC) Project (No. 51106125, 51236006)
and in part by the Fundamental Research Funds for the Central
Universities.
References
[1] R.A. Gingold, J.J. Monaghan, Smoothed particle hydrodynamics: theory and
application to non-spherical stars, Mon. Not. R. Astron. Soc. 181 (1977) 375
389.
[2] S. Koshizuka, Y. Oka, Moving particle semi-implicit method for fragmentation
of incompressible uid, Nucl. Sci. Eng. 123 (3) (1996) 421434.
[3] Z.G. Sun, Y.Y. Liang, G. Xi, Numerical study of the owing sequence of a pouring
liquid, Sci. Chin. Phys. Mech. Astron. 54 (8) (2011) 15141519.
[4] M. Kondo, S. Koshizuka, Improvement of stability in moving particle semiimplicit method, Int. J. Numer. Methods Fluids 65 (6) (2011) 638654.
[5] P.W. Cleary, J.J. Monaghan, Conduction modelling using smoothed particle
hydrodynamics, J. Comput. Phys. 148 (1) (1999) 227264.
[6] A.K. Chaniotis, D. Poulikakos, P. Koumoutsakos, Remeshed smoothed particle
hydrodynamics for the simulation of viscous and heat conducting ows, J.
Comput. Phys. 182 (1) (2002) 6790.
[7] V. Vishwakarma, A.K. Das, P.K. Das, Steady state conduction through 2D
irregular bodies by smoothed particle hydrodynamics, Int. J. Heat Mass
Transfer 54 (13) (2011) 314325.
[8] K. Szewc, J. Pozorski, A. Tanire, Modeling of natural convection with
Smoothed Particle Hydrodynamics: non-Boussinesq formulation, Int. J. Heat
Mass Transfer 54 (2324) (2011) 48074816.
[9] A. Leroy, D. Violeau, M. Ferrand, A. Joly, Buoyancy modelling with
incompressible SPH for laminar and turbulent ows, Int. J. Numer. Methods
Fluids 2015 (online).
[10] M.A. Mansour, M.A.Y. Bakier, Free convection heat transfer in complex-wavywall enclosed cavity lled with nanouid, Int. Commun. Heat Mass Transfer 44
(2013) 108115.

[11] A.M. Aly, S.E. Ahmed, An incompressible smoothed particle hydrodynamics


method for natural/mixed convection in a non-Darcy anisotropic porous
medium, Int. J. Heat Mass Transfer 77 (2014) 11551168.
[12] R.H. Chen, W.X. Tian, G.H. Su, S.Z. Qiu, Y. Ishiwatari, Y. Oka, Numerical
investigation on bubble dynamics during ow boiling using moving particle
semi-implicit method, Nucl. Eng. Des. 240 (11) (2010) 38303840.
[13] W.X. Tian, Y. Ishiwatari, S. Ikejiri, M. Yamakawa, Y. Oka, Numerical
computation of thermally controlled steam bubble condensation using
Moving Particle Semi-implicit (MPS) method, Ann. Nucl. Energy 37 (1)
(2010) 515.
[14] S. Zhang, K. Morita, K. Fukuda, N. Shirakawa, Simulation of three-dimensional
convection patterns in a RayleighBenard system using the mps method,
Memoirs Fac. Eng. Kyushu Univ. 66 (1) (2006) 2937.
[15] S. Zhang, K. Morita, K. Fukuda, N. Shirakawa, An improved MPS method for
numerical simulations of convective heat transfer problems, Int. J. Numer.
Methods Fluids 51 (1) (2006) 3147.
[16] A. Khayyer, H. Gotoh, Development of CMPS method for accurate watersurface tracking in breaking waves, Coastal Eng. J. 50 (2) (2008) 179207.
[17] A. Khayyer, H. Gotoh, Modied moving particle semi-implicit methods for the
prediction of 2D wave impact pressure, Coast. Eng. 56 (4) (2009) 419440.
[18] A. Khayyer, H. Gotoh, A higher order Laplacian model for enhancement and
stabilization of pressure calculation by the MPS method, Appl. Ocean Res. 32
(1) (2010) 124131.
[19] A. Khayyer, H. Gotoh, A 3D higher order Laplacian model for enhancement and
stabilization of pressure calculation in 3D MPS-based simulations, Appl. Ocean
Res. 37 (2012) 120126.
[20] A. Khayyer, H. Gotoh, Enhancement of stability and accuracy of the moving
particle semi-implicit method, J. Comput. Phys. 230 (8) (2011) 30933118.
[21] A. Khayyer, H. Gotoh, Enhancement of performance and stability of MPS meshfree particle method for multiphase ows characterized by high density ratios,
J. Comput. Phys. 242 (2013) 211233.
[22] T. Jiang, J. Ouyang, X.J. Li, L. Zhang, J.L. Ren (The rst order symmetric SPH
method for transient heat conduction problems), Chin. Acta Phys. Sinica 60 (9)
(2011).
[23] T.H. Kuehn, R.J. Goldstein, An experimental and theoretical study of natural
convection in the annulus between horizontal concentric cylinders, J. Fluid
Mech. 74 (4) (1976) 695719.
[24] G.D. Davis, Natural-convection of air in a square cavityA bench-mark
numerical-solution, Int. J. Numer. Method Fluids 3 (1983) 249264.
[25] A.K. Chaniotis, D. Poulikakos, P. Koumoutsakos, Remeshed smoothed particle
hydrodynamics for the simulation of viscous and heat conducting ows, J.
Comput. Phys. 182 (2002) 6790.
[26] K.C. Ng, Y.H. Hwang, T.W.H. Sheu, On the accuracy assessment of Laplacian
models in MPS, Comput. Phys. Commun. 185 (10) (2014) 24122426.

Você também pode gostar