Sie sind auf Seite 1von 8

Meshless inverse method to determine temperature and heat ux at

boundaries for 2D steady-state heat conduction problems


Guang Xu Yu
a,b,
, Jie Sun
a
, Hua Sheng Wang
a,
, Pi Hua Wen
a
, John W. Rose
a
a
School of Engineering and Materials Science, Queen Mary, University of London, Mile End Road, London E1 4NS, UK
b
DENSO Marston Ltd., Otley Road, Shipley, West Yorkshire BD17 7JR, UK
a r t i c l e i n f o
Article history:
Received 27 March 2013
Received in revised form 11 September
2013
Accepted 11 September 2013
Available online 18 September 2013
Keywords:
Meshless inverse method
Method of fundamental solution
Heat conduction
Heat transfer
Measurement
a b s t r a c t
Inverse determination of temperature and heat ux at an inaccessible surface of a solid has been widely
employed in recent years. In this paper, a meshless inverse method, i.e. the method of fundamental solu-
tion (MFS), has been developed to determine the temperature eld and hence the local boundary temper-
ature and heat ux distributions for a 2D steady-state heat conduction problem based on temperature
measurements at interior sample points in the wall of the boundary. A case study showed that MFS pre-
dicts the boundary temperature and heat ux with about the same accuracy as the Becks function spec-
ied method but consumes signicantly less computing time. Error analysis was carried out regarding
uncertainty in location and accuracy of temperature measurement to demonstrate the reliability of the
proposed method.
2013 Elsevier Inc. All rights reserved.
1. Introduction
The inverse heat conduction problem (IHCP) is used to deter-
mine temperature and heat ux at boundaries utilizing tempera-
tures measured at interior points. Various methods of solving the
IHCP have been reviewed in [1]. The mesh-based methods include,
for example, nite difference method (FDM) [2,3], nite element
method (FEM) [4,5], nite volume method (FVM) [6,7] and bound-
ary element method (BEM) [8,9]. FDM, FEM and FVM are used to
discretize the whole domain, whilst BEM to boundary only. Re-
cently, the meshless methods have been developed as alternatives
to classical discretization methods including the Kansas method
based on radial basis functions [10], method of fundamental solu-
tion (MFS) [11], local Petrov-Galerkin method [12], boundary knot
method (BKM) [13] and Monte Carlo method [14]. These meshless
methods require neither domain nor boundary discretization,
therefore the computational efciency can be improved signi-
cantly. Optimum procedures have been introduced using the least
squares method [15], least squares method with regularization
terms [16], conjugate gradient methods [17], steepest descent
method [18] and adaptive iterative lter method [19].
The meshless inverse method used in the present work is MFS,
which was rstly proposed by Kupradze and Aleksidze in 1960s
[20]. Compared with other meshless inverse methods, the imple-
mentation of MFS is relatively simple due to the fact that the solu-
tion is approximated by linear combination of fundamental
solutions corresponding to ctitious heat source points outside
the domain considered. This feature makes MFS suitable for solving
IHCP even with complicated geometries and guarantees accurate
results at small computational cost and consequently has been
developed rapidly within the last decade. Hon and Wei [21] suc-
cessfully developed MFS to solve transient IHCP with implementa-
tion of Tikhonov regularization technique and the L-curve method
to obtain a stable numerical approximation to the solution. They
later extended their work to multidimensional IHCP and the effec-
tiveness of this method was veried by several numerical exam-
ples [22]. Their work was also extended by Dong et al. [23] for
2D IHCP in an anisotropic medium and the implementation of
the truncated singular value decomposition and the L-curve crite-
rion to solve the resulting matrix equation. Jin et al. [24,25] pro-
posed a scheme for solving IHCP based on MFS, in conjunction
with a regularization method. The efciency, accuracy, conver-
gence and stability of the scheme were demonstrated by numerical
results. They pointed out potential extensions of the method, such
as ill-posed Cauthy problems, inhomogeneous problems, inverse
source problems associated with other elliptic partial differential
equations. Other recent works have further extended the applica-
tions of MFS [2631].
The present work stems from the need for a method to accu-
rately measure local temperature eld and heat ux along the ow
0894-1777/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.expthermusci.2013.09.006

Corresponding authors.
E-mail addresses: gyu@denso-marston.co.uk (G.X. Yu), h.s.wang@qmul.ac.uk
(H.S. Wang).
Experimental Thermal and Fluid Science 52 (2014) 156163
Contents lists available at ScienceDirect
Experimental Thermal and Fluid Science
j our nal homepage: www. el sevi er . com/ l ocat e/ et f s
and channel surface during condensation in microchannels (Wang
and Rose [32]). It is very difcult to obtain the local surface tem-
perature and heat ux at such an inaccessible boundary surface.
Su et al. [33] shows that available experimental heat transfer data
for condensation in microchannels were largely scattered. Vapor-
side, heat-transfer coefcients had generally been inferred from
overall measurements by subtraction of thermal resistances and/
or using Wilson plot techniques and therefore have high uncer-
tainty. Rose [34] discussed in detail on heat-transfer coefcients,
Wilson plots and accuracy of thermal measurements. However, it
is relatively easy to measure the temperatures at interior points
in the wall of the boundary and then to determine local surface
temperatures and heat uxes using the inverse method. In order
to maintain the generalization in the present study, MFS is pro-
posed to solve a general IHCP using measured interior wall sample
temperatures as input. It should be noted that the present method
can be applied to a range of heat transfer problems and that the in-
verse method is not limited to any specic IHCP.
2. Method of fundamental solution (MFS)
A 2D steady-state heat conduction case was analyzed to illus-
trate MFS as shown schematically in Fig. 1. The actual problem
with domain X is transformed into an indirect problem in an in-
nite body and the outer boundary C conditions including temper-
ature and heat ux are modeled by assuming ctitious point heat
sources collocated on the ctitious boundary C
0
. This general prob-
lem in a nite domain can be transformed into one in an innite
plane in order to obtain the solution in an indirect way. The points,
where either temperature or heat ux is known, are dened as
eld points in the domain. The total number of the heat sources
is n
sp
and they are arranged out of the domain in the actual prob-
lem. Each of heat sources contributes to the temperature and heat
ux at all eld points in the domain and at the boundary surfaces.
The distance d between the heat source and the point in the do-
main is also shown in Fig. 1. For a 2D steady-state heat conduction
problem, the governing equation with a heat source is given by:
r
2
u

D x
0
; x 1
where u

represents temperature, x
0
is the location of the heat
source point and x is the eld or collocation point. D x
0
; x is the Dir-
ac delta function, which is a mathematical description of the point
source excitation and satises
R
D(x
0
, x)dV = 1. The solution for Eq.
(1) is [35]:
u


1
2p
ln1=r 2
where r is the modular of the distance vector~r between eld point x
and heat source x
0
. Heat ux is obtained by:
q


@u

@~n
3
where ~n is the outwards unit normal vector to the boundary. Substi-
tuting Eq. (2) into Eq. (3) gives:
q


1
2pr
2
r
x
n
x
r
y
n
y
4
where r
x
and r
y
are the components of ~r in the x and y directions,
respectively; n
x
and n
y
are the components of normal in the x and
y directions, respectively.
By the principle of superposition for linear problems, the tem-
perature at eld or collocation point can be represented by:
T
X
nsp
i1
b
i
u

i
5
where b
i
are unknown coefcients of the heat sources (density of
the heat source). For eld point j, temperature can be written as
X
nsp
i1
b
i
u

x
i
; x
j
T
j
; j 1; . . . ; n
T
fp
6
where T
j
is the temperature measured, n
T
fp
is the number of points
with temperature measured, and
X
nsp
i1
b
i
q

x
i
; x
j
Q
j
; j 1; . . . ; n
Q
fp
7
where Q
j
is the heat ux and n
Q
fp
is the number of points with heat
ux measured. Eqs. (6) and (7) constitute an ill-conditioned linear
system of algebraic equations and the unknown coefcients can
Nomenclature
~n outwards unit normal vector to the boundary
n
x
, n
y
the components of ~n in the x, y direction respectively
n
sp
number of heat source
n
Q
fp
number of eld point with known heat ux
n
T
fp
number of eld point with known temperature
Q superpositioned heat ux
q

heat ux excited by heat source x


0
r modular of the distance vector ~r
~r distance vector between eld point x and heat source x
0
r
x
, r
y
components of ~r in the x, y direction respectively
T superpositioned temperature
T
b
uid temperature owing over the bottom surface
T
t
uid temperature owing over the top surface
u

eld point temperature excited by heat source x


0
x eld point
x
0
heat source
Greek symbols
b superposition coefcient of the heat source
a
b
heat-transfer coefcient at the bottom surface
a
t
heat-transfer coefcient at the top surface
Fig. 1. The method of fundamental solution and ctitious boundary.
G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163 157
be solved by singular value decomposition (SVD) method [36]. To
determine the temperature and heat ux distributions at the top
and bottom surfaces with the vector b
i
, one only needs to choose
a nite number of discrete points at the surfaces to nd the corre-
sponding values of u

and q

by Eqs. (2) and (4). These values can


then be substituted into Eqs. (6) and (7) to obtain the temperature
and heat ux at these points. Since no mesh and iteration are re-
quired computing time is signicantly reduced.
3. Results and discussion
3.1. Code validation
To validate the method, a 2D steady-state heat conduction
problem, as shown in Fig. 2, was set up. In this problem, a copper
wall 440 mm long and 18 mm thick with thermal conductivity
398 W/m K is considered. An adiabatic boundary condition is ap-
plied to the sides of the wall. Thirty-ve interior sample tempera-
tures are measured. On each side wall, ve collocation eld points
are selected to represent the adiabatic boundary condition giving
forty-ve eld points in total. The number of the heat sources in
this particular case is n
sp
= 44, and the number of the eld points
with temperature measured n
T
fp
35 and the number of the collo-
cation points for heat ux n
Q
fp
10, respectively. To determine the
wall temperature eld and heat ux distributions at the top and
bottom surfaces using inverse method, the temperatures at the
eld points are obtained from the solutions of a direct, 2D stea-
dy-state heat conduction problem as shown in Fig. 3. Both side sur-
faces are adiabatic. Convective heat transfer (counter-current
ows) occurs at the top and bottom surfaces. The governing equa-
tion is
r
2
T 0 8
The boundary conditions are
dT
dx
0 at x 0; L 9
and T
t
= 56 C and a
t
= 97.0x
2
93.6x + 30.0 kW/m
2
K are specied
at the top surface while T
t
= 0.42x
2
2.72x + 41.11 C and
a
b
= 2.92 kW/m
2
K are specied at the bottom surface. The problem
is readily solved numerically using the nite difference method.
In the inverse process, the temperatures obtained from the di-
rect solution at the corresponding sample points are taken as
experimental measurements and used as the input to MFS. The
predicted temperatures at the top surface agree well with the input
values, and the predicted heat ux distribution at the top surface
also agrees well with the exact solution as shown in Fig. 4. Results
Fig. 2. Schematic of temperature measurements and heat sources for the method of
fundamental solution (MFS).
Fig. 3. Schematic of the direct 2D steady-state heat conduction problem.
Fig. 4. Comparison of actual and predicted (a) temperature and (b) heat ux at the
top boundary surface.
158 G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163
given by the Becks function specied method are plotted for com-
parison in Fig. 4. Following the principle of the Becks function
specied method described in [15] (see Section 19.3.3), the tem-
perature or heat ux at the top and bottom solid surfaces are spec-
ied as functions of the distance (x coordinator) along the solid
surface. The inverse problem is then converted into a direct prob-
lem. The parameters or coefcients used in the specication func-
tions are determined by minimizing the least-square sum
S
Pn
T
fp
i1
T
mea
i
T
cal
i

2
, where T
mea
i
and T
cal
i
denote the measured
and calculated temperatures at location i and n is the number of
the temperature sample points in the whole domain. In the appli-
cation of FSM, the nite difference method was used to solve the
Laplace equation (Eq. (8)). It is seen that MFS predicts the temper-
ature and heat ux distributions at the top boundary surface of the
domain as accurately as FSM. The CPU times needed for the whole
computation, using standard FORTRAN intrinsic CPU_TIME, were
compared. It spent 3.766 s by FSM and 0.016 s by MFS (Platform:
3.0 GHz Pentium 4 processor and 3.0 GB of RAM).
3.2. Effects of temperature and position errors of thermocouples
Effects of experimental errors in temperature and position of
thermocouples have been assessed. Two temperature measure-
ment error bounds, 0.05 K and 0.1 K, were considered. Multi-
plied by a normally distributed random number between (1, 1)
at each sample point, the estimated measurement error was then
added to the computed temperature from the direct solution as
the new measurements. Fig. 5 indicates maximum errors of pre-
dicted heat ux, based on thirty-ve temperature measurements
with maximum error of 0.05 K and 0.1 K (assuming no error in
position of thermocouples), are about 5% and 9% of the input
heat ux, respectively.
The diameter of the holes accommodating thermocouples also
affects the prediction accuracy. To analyze thermocouple position
error, each thermocouple measured temperature is chosen ran-
domly from the temperatures of ve extremist positions inside
Fig. 5. Predicted heat ux at the top boundary surface based on thirty-ve
temperature measurements with maximum error of (a) 0.05 K and (b) 0.1 K.
(a)
(b)
Thermocouple
positions considered
Thermocouple
positions considered
Fig. 6. Predicted heat ux at the top boundary surface based on thirty-ve
temperature measurements with thermocouple hole diameter of (a) 0.3 mm and (b)
0.6 mm.
G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163 159
the thermocouple hole, as shown in Fig. 6. The temperatures at
these positions were known from the direct solution of Laplace
equation. The predicted heat ux distributions at the top surface
for thermocouple hole diameters of 0.3 mm and 0.6 mm are shown
in Fig. 6. The results were obtained based on thirty-ve tempera-
ture measurements assuming no errors in them. Fig. 6 shows that
for a smaller diameter hole, 0.3 mm, the maximum heat ux pre-
diction error is less than 5% of the input whilst it is about 5%
of the input for the 0.6 mm diameter hole.
Analysis where errors in temperature and position of thermo-
couple were considered simultaneously was also done. The results
are presented in Fig. 7. It is seen that with both errors considered,
the prediction accuracy of the heat ux has been reduced com-
pared with Figs. 5 and 6. However, with improving conditions from
thermocouple hole of 0.6 mm and temperature error of 0.1 K to
thermocouple hole of 0.3 mm and temperature error of 0.05 K,
the average and maximum errors of the predicted heat ux are
roughly reduced from 13% and 30% to 9% and 15%, respec-
tively. It should be noted that the maximum errors in predicted
(a)
(b)
Thermocouple
positions considered
Thermocouple
positions considered
Fig. 7. Predicted heat ux at the top boundary surface based on thirty-ve
temperature measurements with (a) thermocouple hole diameter of 0.6 mm and
temperature error of 0.1 K, and (b) thermocouple hole diameter of 0.3 mm and
temperature error of 0.05 K.
Fig. 8. Schematic of MFS with nine temperature and three additional heat ux
measurements.
Fig. 9. Predicted (a) temperature and (b) heat ux at the top boundary surface
based on nine temperature measurements with 0.1 K accuracy.
160 G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163
surface temperatures at the top boundary surface in Figs. 57 are
less than 0.1 K (gures are not shown).
3.3. Effect of additional heat ux measurements at boundaries
Another case study was carried out in order to investigate the
effect of additional heat ux measurements at boundaries, where
nine thermocouples arranged by a 3 3 array and three additional
heat ux sensors are set as shown in Fig. 8. Using only the mea-
sured temperatures with 0.1 K accuracy at the nine sample points
(assuming no error in position of thermocouples), MFS gave the
maximum surface temperature prediction error around 0.2 K
and the heat ux error about 23% of the input, as shown in
Fig. 9(a) and (b), respectively. The larger error in predicted heat
ux is due to the fact that heat ux is proportional to the derivative
of temperature. If only temperature measurement is used by MFS,
a small error in the temperature measurement will be amplied
and causes a signicant change in the predicted heat ux. If some
heat ux measurements with known accuracy on one side of the
domain, for example at the bottom surface, are used, more con-
strains are applied when determining the superposition coef-
cients of the heat sources b
i
in Eqs. (5) and (6), the prediction
accuracy, particularly for the heat ux, can be improved.
Fig. 10(a) and (b) are the estimations for the temperature and heat
ux distributions at the top surface with three additional heat ux
measurements at the bottom surface of the domain. It is seen that
when the heat ux measurement error is 5% of the input, the
maximum heat ux prediction error is reduced to 5% and the sur-
face temperature error down to 0.1 K. It may be seen that to ob-
tain the same prediction accuracy, the sample points for
temperature measurement can be greatly reduced from 35 to 9
and the accuracy of temperature measurement required can be re-
duced from 0.05 K to 0.1 K.
Fig. 10. Predicted (a) temperature and (b) heat ux at the top boundary surface
based on nine temperature measurements with 0.1 K accuracy and three
additional heat ux measurements with 5% accuracy.
Fig. 11. Predicted heat ux at the top boundary surface based on nine temperature
measurements with 0.1 K accuracy and various locations of three additional heat
ux measurements with 5% accuracy.
Fig. 12. Predicted heat ux at the top boundary surface based on nine temperature
measurements with 0.1 K accuracy and various numbers of additional heat ux
measurements with 5% accuracy.
G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163 161
It is found that, given a certain number of temperature mea-
surements, the location and number of the additional heat ux
measurements also affect the prediction accuracy. The estimations
with three additional heat ux measurements conducted at vari-
ous positions at the boundary surface are plotted in Fig. 11. Three
positions were randomly selected for each run, and a specic case
corresponding to the locations designated in Fig. 8 (positions A, B
and C) was also included. It is demonstrated that if the arrange-
ment of temperature measurements is xed, the accurate desired
estimations are obtained when the three additional heat ux mea-
surement positions are aligned with the temperature measure-
ment positions (see Fig. 8). Fig. 12 presents the predicted heat
ux at the top surface of the domain when three, four and ve
additional heat ux measurements were conducted. The locations
of the three heat ux measurements are shown in Fig. 8. When
more than three measurements were made, three of them were lo-
cated the same as in Fig. 8 and the rest are randomly positioned. It
is found that with the same temperature measurement positions
the results do not change signicantly after more than three addi-
tional heat ux measurements were conducted. To determine the
minimum additional heat ux measurements required when the
temperature measurement positions are arranged as in Fig. 8,
MFS was applied with one and two additional heat ux measure-
ments. The position of the one heat ux sensor can be any of the
three locations in Fig. 8 or random. The two heat ux sensors can
be positioned at any two of the three locations in Fig. 8 or random.
Fig. 13(a) and (b) conrm that the minimum number of additional
heat ux measurements is three given the temperature measure-
ment positions as shown in Fig. 8. Further simulations of cases
where there are more columns of temperature measurements
showed that the number of additional heat ux measurements re-
quired depends on the positions of temperature measurement and
is the same as the number of columns, and the locations of the heat
ux measurement should be in line with the temperature mea-
surement columns.
The meshless inverse method can also be applied to cases with
multiple materials and non-rectangular geometries [37].
4. Conclusions
An inverse meshless method of fundamental solution (MFS) has
been developed and validated to determine boundary temperature
and heat ux distribution of steady-state heat conduction prob-
lems. The case study showed that with thirty-ve temperature
sample points in the interior wall, temperature measurement
accuracy should be at least 0.05 K and thermocouple hole diame-
ter no bigger than 0.6 mm, to obtain 0.1 K prediction accuracy for
the surface temperature and 5% predicted accuracy of the surface
heat ux. It was also found that to obtain the same prediction accu-
racy fewer temperature measurements and lower accuracy are re-
quired if three additional heat ux measurements with accuracy of
5% are used. For a given temperature measurement arrangement,
the accurate desired estimations are obtained when the additional
heat ux measurement positions are aligned with the temperature
measurement positions and the number of the heat ux measure-
ments should be at least three.
Acknowledgements
The work was nancially supported by the Engineering and
Physical Sciences Research Council (EPSRC) of the UK through re-
search grant of EP/H014349/1 and EU FP7-2010-IRSES-269205.
References
[1] J.V. Beck, B. Blackwell, C.R. Clair, Inverse Heat Conduction: Ill-Posed Problems,
Wiley, New York, 1985.
[2] N. Al-Khalidy, A general space marching algorithm for the solution of two-
dimensional boundary inverse heat conduction problems, Numer. Heat
Transfer Part B Fund. 34 (3) (1998) 339360.
[3] L. Guo, D. Murio, A modied space-marching nite difference algorithm for the
two-dimensional inverse heat conduction problem with slab symmetry,
Inverse Prob. 7 (2) (1991) 247259.
[4] B.H. Dennis, G.S. Dulikravich, Simultaneous determination of temperatures,
heat uxes, deformations, and tractions on inaccessible boundaries, J. Heat
Trans. T ASME 121 (3) (1999) 537545.
[5] H.J. Reinhardt, A numerical method for the solution of two dimensional inverse
heat conduction problems, Int. J. Numer. Methods Eng. 32 (2) (1991) 363383.
[6] B. Wang, G.-A. Zou, P. Zhao, Q. Wang, Finite volume method for solving a one-
dimensional parabolic inverse problem, Appl. Math. Comput. 217 (12) (2011)
52275235.
[7] M. Dehghani, S.M. Hosseini Sarvari, H. Ajam, Inverse estimation of boundary
conditions on radiant enclosures by temperature measurement on a solid
object, Int. Commun. Heat Mass Transfer 38 (10) (2011) 14551462.
[8] K. Kobayashi, K. Onishi, Y. Ohura, On identifying Dirichlet condition for 2D
Laplace equation by BEM, Eng. Anal. Boundary Elem. 17 (3) (1996) 223230.
Fig. 13. Predicted heat ux at the top boundary surface based on nine temperature
measurements with 0.1 K accuracy and various locations of (a) one additional heat
ux measurement and (b) two additional heat ux measurements with 5%
accuracy.
162 G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163
[9] T.J. Martin, G.S. Dulikravich, Inverse determination of boundary conditions and
sources in steady heat conduction with heat generation, J. Heat Trans. T
ASME 118 (3) (1996) 546554.
[10] E.J. Kansa, Multiquadrics a scattered data approximation scheme with
applications to computational uid-dynamics. II. Solutions to parabolic,
hyperbolic and elliptic partial differential equations, Comput. Math. Appl. 19
(89) (1990) 147161.
[11] M.A. Golberg, C.S. Chen, The method of fundamental solution for potential,
Helmholtz and diffusion equation, in: M.A. Golberg (Ed.), Boundary Integral
Methods Numerical and Mathematical Aspects, Computational Mechanics
Publications, Southampton, 1998, pp. 103176.
[12] J. Sladek, V. Sladek, Y.C. Hon, Inverse heat conduction problems by meshless
local Petrov-Galerkin method, Eng. Anal. Boundary Elem. 30 (8) (2006) 650
661.
[13] B.T. Jin, Y. Zheng, Boundary knot method for some inverse problems associated
with the Helmholtz equation, Int. J. Numer. Methods Eng. 62 (12) (2005)
16361651.
[14] A. Hajisheikh, F.P. Buckingham, Multidimensional inverse heat conduction
using the monte carlo method, J. Heat Trans. T ASME 115 (1) (1993) 2633.
[15] J.V. Beck, B. Blackwell, Inverse Problems, in: W.J. Minkowycz, E.M. Sparrow,
G.E. Schneider, R.H. Pletcher (Eds.), Handbook of Numerical Heat Transfer,
Wiley, New York, 1988, pp. 786834.
[16] A.N. Tikhonov, V.Y. Aresenin, Solutions of Ill-Posed Problems, Wintson & Sons,
Washington, 1977.
[17] M. Necati zis ik, Heat Conduction, second ed., John Wiley & Sons, New York,
1993.
[18] O.M. Alifanov, Inverse Heat Transfer Problems, Springer, Berlin, 1994.
[19] A.V. Moultanovsky, Computational parameter selection and calculation
features of adaptive iterative lter as inverse problem method, in: 8th
Annual on Inverse Problems in Engineering Seminar, New York, USA, 1997.
[20] V.D. Kupradze, M.A. Aleksidze, The method of functional equations for the
approximate solution of certain boundary value problems, USSR Comput.
Math. Math. Phys. 4 (4) (1964) 82126.
[21] Y.C. Hon, T. Wei, A fundamental solution method for inverse heat conduction
problem, Eng. Anal. Boundary Elem. 28 (5) (2004) 489495.
[22] Y.C. Hon, T. Wei, The method of fundamental solution for solving
multidimensional inverse heat conduction problems, Comput. Model. Eng.
Sci. 7 (2) (2005) 119132.
[23] C.F. Dong, F.Y. Sun, B.Q. Meng, A method of fundamental solutions for inverse
heat conduction problems in an anisotropic medium, Eng. Anal. Boundary
Elem. 31 (1) (2007) 7582.
[24] B.T. Jin, Y. Zheng, L. Marin, The method of fundamental solutions for inverse
boundary value problems associated with the steady-state heat conduction in
anisotropic media, Int. J. Numer. Methods Eng. 65 (11) (2006) 18651891.
[25] B. Jin, L. Marin, The method of fundamental solutions for inverse source
problems associated with the steady-state heat conduction, Int. J. Numer.
Methods Eng. 69 (8) (2007) 15701589.
[26] A. Shidfar, Z. Darooghehgimofrad, M. Garshasbi, Note on using radial basis
functions and Tikhonov regularization method to solve an inverse heat
conduction problem, Eng. Anal. Boundary Elem. 33 (10) (2009) 12361238.
[27] J.A. Kolodziej, M. Mierzwiczak, M. Cialkowski, Application of the method of
fundamental solutions and radial basis functions for inverse heat source
problem in case of steady-state, Int. Commun. Heat Mass Transfer 37 (2)
(2010) 121124.
[28] M. Mierzwiczak, J.A. Kolodziej, Application of the method of fundamental
solutions and radial basis functions for inverse transient heat source problem,
Comput. Phys. Commun. 181 (12) (2010) 20352043.
[29] B.T. Johansson, D. Lesnic, T. Reeve, A comparative study on applying the
method of fundamental solutions to the backward heat conduction problem,
Math. Comput. Model 54 (12) (2011) 403416.
[30] L. Marin, The method of fundamental solutions for inverse problems
associated with the steady-state heat conduction in the presence of sources,
Comput. Model. Eng. Sci. 30 (2) (2008) 99122.
[31] L. Yan, C.-L. Fu, F.-L. Yang, The method of fundamental solutions for the inverse
heat source problem, Eng. Anal. Boundary Elem. 32 (3) (2008) 216222.
[32] H.S. Wang, J.W. Rose, A theory of lm condensation in horizontal noncircular
section microchannels, J. Heat Trans. T ASME 127 (10) (2005) 10961105.
[33] Q. Su, G.X. Yu, H.S. Wang, J.W. Rose, Microchannel condensation: correlations
and theory, Int. J. Refrigeration 32 (2009) 11491152.
[34] J.W. Rose, Heat-transfer coefcients, Wilson plots and accuracy of thermal
measurements, Exp. Thermal Fluid Sci. 28 (2004) 7786.
[35] L.C. Wrobel, The boundary element method, Applications in Thermo-uids and
Acoustics, vol. 1, Wiley, New York, 2002.
[36] W.H. Press, A.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical Recipes in
Fortran 77, second ed., vol. 1, Press Syndicate of the University Cambridge,
1997.
[37] G.X. Yu, P.H. Wen, H.S. Wang, J.W. Rose, An inverse method to determine
boundary temperature and heat ux for a 2D steady state heat conduction
problem, in: Proc. the ASME 2008 Int. Design Eng. Technical Conf. & Computers
and Information in Engineering Conference, IDETC/CIE 2008, August 36, 2008,
New York, USA. Paper No. DETC2008-49811.
G.X. Yu et al. / Experimental Thermal and Fluid Science 52 (2014) 156163 163

Das könnte Ihnen auch gefallen