You are on page 1of 16

Numerical Heat Transfer, Part A, 52: 71–85, 2007

Copyright # Taylor & Francis Group, LLC


ISSN: 1040-7782 print=1521-0634 online
DOI: 10.1080/10407780601114981

TRANSIENT RESPONSE OF A CO-CURRENT HEAT


EXCHANGER TO AN INLET TEMPERATURE VARIATION
WITH TIME USING AN ANALYTICAL AND NUMERICAL
SOLUTION

M. R. Ansari and V. Mortazavi


Mechanical Engineering Department, Tarbait Modarres University, Tehran,
Islamic Republic of Iran

In this article, we examine the transient response of a co-current double-pipe heat exchan-
ger with respect to the inlet temperature change. For this investigation, a method whose
main feature is analytical solution of the energy equation for one element of the heat
exchanger is developed. The governing equations are linearized with respect to space and
time in one element, and the temperature distribution for the element is obtained. Finally,
the obtained temperature distribution for one element is utilized numerically for the whole
length of the heat exchanger. Various boundary conditions (ramp, exponential, and sinus-
oidal) are applied, and the responses are investigated with respect to various parameters. A
comparison of the results with data from previous studies shows that the method applied in
this article is capable of predicting transient behavior accurately. In addition, because the
energy equation is solved analytically, the present method has higher accuracy than those
previously utilized, and the calculation time declines noticeably.

1. INTRODUCTION
In recent years, the transient behavior of heat exchangers has been a topic of
interest and study in industry and research. Process control applications, heat trans-
fer coefficients determination, and an understanding of the effect of the unsteady
state on heat exchanger tubes to prevent adverse thermal stresses are some of the rea-
sons for this interest [1, 2]. Various methods have been used to study the transient
behavior of heat exchangers. In some methods, the heat exchanger response to the
inlet temperature change or the inlet mass flow rate of one or both fluids were inves-
tigated using an analytical solution of the governing equations [1–4]. Because of the
presence of partial derivatives, these governing equations either have no exact ana-
lytic solution or the analytic solution is extremely complicated and difficult to use [1].
Other methods using numerical solutions have problems with some limitations, such
as numerical instability, numerical diffusion, and the fact that convergence could not
be obtained [1]. For example, prior methods were used to explain the dynamical
response of a heat exchanger to temperature change (see, e.g., Roppo and Ganic
[5] and Correa and Marcheti [6]); their results are also relevant to this study.

Received 13 April 2006; accepted 13 October 2006.


Address correspondence to M. R. Ansari, Mechanical Engineering Department, Tarbait Modarres
University, P. O. Box 14155-4838, Tehran, Islamic Republic of Iran. E-mail: mra_1330@modares.ac.ir

71
72 M. R. ANSARI AND V. MORTAZAVI

NOMENCLATURE

A cross-sectional area, m2 DQ heat transferred, kJ


B dimensionless coefficient t time, s
½ðDx=LÞ  ð1  eDx=L Þ Dt time interval, s
c specific heat capacity, kJ=kg K T temperature, K
D time coefficient, s u(x) step function fu½t  ðD=LÞx ¼ 1
E dimensionless
 coefficient if t ðD=LÞx  0g
ðDx=LÞeDx=L W velocity, m=s
þ½ðDt=DÞ  1ð1  eDx=L Þg x length, m
F dimensionless coefficient Dx longitudinal segment, m
Z dimensionless time ð¼ tL=DX Þ
( b dimensionless coefficient
D1 L1 1
  1eDt=DW
DW Dx DW " #
"    Dt 1  ðDw =DÞ2  
h1 U1 h2 U2  þ1þ 1  eDx=L
þ b1 þ D 1  ðDw =DÞ
h1 U1 þh2 U2 h1 U1 þh2 U2
  # ) Dx   ðDw =DÞ
 1 þ eDx=L 
L2 D2 L1 D1 L 1  ðDw =DÞ
 b2 
L1 D1 Dx DW Dx ðDt=Dw Þ ðDw =DÞ3
 e 
L ð1  Dw =DÞ2
 
ðDx=LÞðD=Dw 1Þ
G dimensionless coefficient  e  1 eðDt=Dw Þ

 2  
Dt
þ D  Dt 1  eDx=L h time variant temperature coefficient,
2D
"    # K=s
Dx=L Dx D Dx 2 q density, kg=m3
þe Dt  D 
L 2 L s dimensionless temperature
½ðT  Ti Þ=ðTmax  Ti Þ
Subscripts
h convective heat transfer coefficient,
i initial
kW=m2 K
M average value at element length
L spatial coefficient, m
max maximum
m
_ mass flow rate, kg=s
w wall
NTU number of transfer units
1 fluid 1
P pipe circumference, m
2 fluid 2
Q heat, kJ

However, the heat capacity effect of the wall was not considered in the study of
Roppo and Ganic [5]. Both of the methods of [5] and [6] are based on the finite-
difference algorithm. Rotzel and Xuan [7, 8] solved the governing equations using
the Laplace transform. In this method, because it is impossible to invert the solution
from the transformed domain to the original domain, the numerical inverse of the
Laplace transform was used. This method can be problematic, especially when the
boundary condition is in the form of a periodic function. Thus, the previously
obtained result is not reliable for the longer period of time [1, 7]. Alotaibi et al. [9]
applied finite control volumes for transient numerical solution of cross-flow heat
exchangers. Their results showed that the control behavior can be simulated numeri-
cally, and this control methodology is effective. Asinari [10] and Corberan et al. [11]
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 73

also used numerical methods for discritization of governing equations for complex
heat exchangers.
In recent years, due to the spreading application of three-dimensional numeri-
cal methods, some authors have investigated the dynamic behavior of heat exchan-
gers with complex arrangements. Tutar et al. [12] provided a numerical study on
three-dimensional time-dependent modeling of unsteady flow over single-and multi-
row plate fin-and-tube heat exchangers. Saha et al. [13] studied the unsteady three-
dimensional flow in parallel-plate heat exchangers numerically.
The method suggested by Dolezal [14] is used in this article. It is applied to the
transient response calculation of a co-current heat exchanger for the first time. This
method is based on a combination of analytical and numerical methods. In other
words, the energy equation is solved analytically for a space and time element,
and then the obtained temperature distribution for this element is applied to the
whole system, using a numerical algorithm.

2. THE APPLIED METHOD AND THE GOVERNING EQUATIONS


The transient behavior of a co-current double-pipe heat exchanger is demon-
strated in Figure 1. It is assumed that the hot fluid in the outer tube (subscript 1)
and the cold fluid in the inner tube (subscript 2) are flowing. The energy conser-
vation equations for the two fluids and the wall were written with the following
assumptions:

The heat transfer by conduction along the length of the wall is neglected.
The wall heat resistance is negligible in comparison with the convection resistance.
The shroud is adiabatic.

With respect to the above assumptions, the following equations are obtained:

qT1 qT1
c1 q1 A1 ¼ h1 P1 ðT1  Tw Þ  c1 m
_1 ð1Þ
qt qx
qTw
cw qw Aw ¼ h1 P1 ðT1  Tw Þ  h2 P2 ðTw  T2 Þ ð2Þ
qt

Figure 1. Co-current concentric double-pipe heat exchanger.


74 M. R. ANSARI AND V. MORTAZAVI

qT2 qT2
c2 q2 A2 ¼ h2 P2 ðTw  T2 Þ  c2 m
_2 ð3Þ
qt qx

where T is the temperature, c is the specific heat capacity, q is the density, h is


the convective heat transfer coefficient, A is the cross-sectional area, P is the tube
periphery, and m _ is the mass flow rate through the cross section. The subscript w
is used to represent the wall parameters. The above three equations can be rewritten
as follows:
qT1 qT1
D1 þ L1 þ T1 ¼ Tw ð4Þ
qt qx

qTw h1 P1 T1 þ h2 P2 T2
Dw þ Tw ¼ ð5Þ
qt h1 P1 þ h2 P2

qT2 qT2
D2 þ L2 þ T2 ¼ Tw ð6Þ
qt qx

where D is the time coefficient [s] and L is the spatial coefficient [m], defined as

q1 c1 A1 q2 c2 A2 qw cw Aw
D1 ¼ D2 ¼ Dw ¼
h1 P1 h2 P 2 h1 P 1 þ h2 P 2
c1 m
_1 c2 m
_2
L1 ¼ L2 ¼ ð7Þ
h1 P1 h2 P2

In order to solve Eqs. (4)–(6) for the heat exchanger shown in Figure 1, first,
the heat exchanger length was divided into elements with Dx length. Each of these
elements contains the hot and the cold fluids, and the walls between the two fluids
are of equal length Dx. Now, the equations are solved for this element within the
time period of Dt, i.e., a two-dimensional element (Dx, Dt). The solution of the
equation in the elements is based on the following assumptions.
First, the wall temperature (Tw) inside of the elements varies only with time,
and it has a constant value along Dx. This can be explained by the result of energy
transfer during the period Dt, where T2 increases and T1 decreases; the right-hand
side of Eq. (5) does not change noticeably during time period 0 < t < Dt. Hence,
Doležal et al. [15] suggest that the time variation of the unknown temperature of
the wall elements can be represented by a substitute function, which has the
following form:

TwM ðtÞ ¼ TwM ð0Þ þ Dw hð1  et=Dw Þ ð8Þ

The subscript M denotes the average value along the element, and h is the time vari-
ation coefficient [K=s]. For the validation of the above equation, we add that it has
been shown experimentally [16] that the transient response of a heat exchanger to
any disturbance (either of the inlet temperatures or mass flows) is in the form of an
exponential function in which the power contains a time parameter with negative sign.
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 75

Second, the coefficients D, L, and h are assumed to be constant in the two-


dimensional element. The average value will be used in the event that any of the
coefficients change.
Third, the initial condition is assumed to be constant for the both the fluids
and the wall along the element (Dx). We assume that the boundary condition in
the period 0 < t < Dt can be estimated by a linear function. Therefore, the following
equation is written for each fluid:

Tð0; tÞ ¼ Tð0; 0Þ þ uM  t if 0 < t < Dt ð9Þ

Further, uM is defined as

Tð0; DtÞ  Tð0; 0Þ


uM ¼ ð10Þ
Dt
By substituting Eq. (8) into Eqs. (4) and (6), Eqs. (4) and (6) are changed to two
nonhomogenous differential equations for the two-dimensional elements of (Dx, Dt):

qT1 qT1
D1 þ L1 þ T1 ¼ TwM ð0Þ þ Dw hð1  et=Dw Þ ð11Þ
qt qx

qT2 qT2
D2 þ L2 þ T2 ¼ TwM ð0Þ þ Dw hð1  et=Dw Þ ð12Þ
qt qx
It should be noted that each of these equations has one unknown, and both
equations have similar form for the each fluids. Both Eqs. (11) and (12) with the
initial and boundary conditions explained above are solved with respect to D, L,
and h using the Laplace transform method. The solution of the equations for each
hot and cold fluid is written as shown in Eq. (13). (The related subscript for each
fluid can be added to the answer.)
     
D D
Tðx; tÞ ¼ TM ð0Þ  et=D 1  u t  x þ Tð0; 0Þ þ uM t  x
L L
  h
D
 ex=L u t  x þ TwM ð0Þ 1  et=D  ex=L ð1  et=Dþx=L Þ
L
  (
D Dw h
u t  x þ D  ð1  et=D Þ  Dw  ð1  et=Dw Þ:
L D  Dw
h i
ex=L D  ð1  et=Dþx=L Þ  Dw ð1  et=Dw þD=Dw x=L Þ
 )
D
u t  x ð13Þ
L

TM represents the initial condition for the element. The step function in the above
equation is defined as
 
D D
u t x ¼1 if t  x  0 ð14Þ
L L
76 M. R. ANSARI AND V. MORTAZAVI

This equation represents the thermal wave position. The thermal wave is propagated
along the heat exchanger tube with the following velocity:

L m
_
W¼ ¼ ð15Þ
D qA

The only unknown in Eq. (13) is h, which can be obtained from the energy
equation for the wall. The energy equation for the wall at the (Dx, Dt) element
can be presented as
DQw ¼ DQ1  DQ2 ð16Þ

where DQw is the wall energy equation, DQ1 is the amount of energy transferred from
fluid 1 to the wall, and DQ2 is the amount of energy transferred from the wall to fluid
2. They can be evaluated from the following equations:
Z Dt
d
DQw ¼ cw qw Aw Dx TwM ðtÞ dt ð17Þ
0 dt

Z Dx Z Dt
DQ1 ¼ h1 U1 ½T1 ðx; tÞ  TwM ðtÞ dt dx ð18Þ
0 0

Z Dx Z Dt
DQ2 ¼ h2 U2 ½TwM ðtÞ  T2 ðx; tÞ dt dx ð19Þ
0 0

If Eq. (13) is substituted in the above equations for T1 and T2, then, h is obtained
after the necessary calculations as follows:
     
h1 P 1 h2 P2 L2 D2
h¼F a1 þ a2 ð20Þ
h1 P 1 þ h2 P 2 h1 P1 þ h2 P2 L1 D1

where a in the above equation is defined as

a ¼ B½TM  TwM ð0Þ þ E½Tð0; 0Þ  TwM ð0Þ þ G  uM ð21Þ

In [14], an equation has been declared for h when only the hot fluid is flowing
through an isolated tube.
Now, having obtained the temperature distribution in one element as described
above, the entire temperature distribution for the whole heat exchanger in each time
interval can be obtained. The procedure can be summarized as follows:
The temperature change is calculated for every element (beginning with the
first element) in every time internal. First, the coefficients of Eq. (20) are calculated,
so h can be evaluated. Once h and the initial and boundary conditions are known, the
temperature distribution can be estimated for both fluids and the wall of the element.
The temperature distribution for the last point of the calculated element (x ¼ Dx) is
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 77

used as the boundary condition for the next element. This procedure is repeated
for the all of the time intervals and continued up to establishment of a new steady
condition.

3. SPECIFICATION OF THE APPLIED METHOD


This method uses a numerical procedure in general. However, as the tempera-
ture distribution for the element with respect to time and space is obtained directly
from the analytical solution of the energy equation, it is anticipated that it produces
results with higher accuracy. In addition, the calculation is conducted only once, so
repetition is not required. Other difficulties, such as numerical instability or diver-
gence of the results, do not appear. Next, the comparison between the present
method and the results obtained from the finite-volume method is shown. In this
comparison, all of the parameters are assumed to be constant and the boundary
condition is in the form of a linear function.
The convergency of results was verified. For the size of elements, it was
obtained when

Dx < W  Dt ð22Þ

In this case, the result was consistent with the physics of the subject. With decreasing
element size, the results converged to a defined value. W is obtained from Eq. (15). In
the case of different velocities between fluids, Eq. (22) is used for the lower velocity.
Selecting element size without respect to Eq. (22) may lead to unrealistic physical
results. In this case, h has a negative value. The proper size of elements will be dis-
cussed in the results section along with the assumptions that were made in order to
obtain Eq. (13). There is a possibility that the method can be generalized when the
mass flow rate, convective heat transfer coefficient, or thermal capacity of fluids are
varied through the entire passage. Pursuant to the above discussion, the capability of
the applied method is greater when compared with the works of Roetzel and Xuan
[7, 8]. Some other applications and specifications of this method can be found in [17].
Furthermore, the manner of solution of the energy equation for one element makes
it possible that we could consider an element with two cross flows and then expand it
to shell-and-tube and cross-flow heat exchangers. The generalized method and its
applications for analysis of transient behavior of shell-and-tube and cross-flow heat
exchangers can be obtained elsewhere.

4. RESULTS
In order to verify the method validation, the transient response of a heat
exchanger is studied; the results were analyzed when the inlet temperature varied lin-
early, exponentially, and periodically. For the purpose of a better comparison, it was
assumed that q and m _ are constant for both fluids (incompressible condition) and
thermal capacity, and that the heat convective transfer coefficient does not change
through the whole passage.
78 M. R. ANSARI AND V. MORTAZAVI

The transient response of the system is analyzed for different values of numbers
of transfer units (NTU) under the following condition:

D1 L 1
_ 1 c1 ¼ m
m _ 2 c2 ¼ ¼1
D2 L 2

Dw X X
¼ 0:25 NTU1 ¼ ¼ ð23Þ
D1 2L1 2L1

In the above equation, X is the total length of the tube; NTU1 is the number of
transfer units [18] and can be defined as
 1
1 1 1 X
NTU1 ¼ þ ¼ ð24Þ
h1 P1 X h2 P2 X _ 1 c1 L1 þ L2 ðm
m _ 2 c2 Þ
_ 1 c 1 =m

The boundary conditions in the form of the following functions (periodic,


exponential, and linear) are considered as the input disturbance to the system:

Ti þ 20 sinðZÞ 0  Z  2p
T1 ð0; tÞ ¼ ð25Þ
Ti 2p < Z

T1 ð0; tÞ ¼ Ti þ 20Z  expðZÞ ð26Þ


Ti þ 20Z 0Z2
T1 ð0; tÞ ¼ ð27Þ
Ti þ 40 2<Z

Ti is the initial temperature and Z is the dimensionless time defined as


tL1
Z¼ ð28Þ
D1 X

The temperature distribution for each of the different boundary conditions,


as mentioned above, was obtained for both fluids and the tube wall by varying the
values of NTU1. The results are presented in the form of dimensionless temperature s:
T  Ti
s¼ ð29Þ
Tmax  Ti

where Tmax is the maximum inlet temperature.


Three cases were analyzed in order to verify the results.
First, the response of fluid 1 at x ¼ X is shown in Figure 2 using Eq. (27) for
different values of Dt. As Dt decreased, the results converged to a defined value and
did not change significantly for Dt values smaller than 0.5.
Second, The results obtained from this method were compared with the finite-
volume results at x ¼ 10 m. The finite-volume results were verified in comparison
with experimental cases [19]. See Figure 3. The agreement is good (the results pre-
sented were obtained when Dt ¼ 0.5 and Dx ¼ 0.5). Using the same condition, the
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 79

Figure 2. Comparison of response of fluid 1 to linear inlet temperature change for different values of Dt.

Figure 3. Comparison of results obtained from this method with results obtained from finite-volume
method for linear inlet temperature change.
80 M. R. ANSARI AND V. MORTAZAVI

finite-volume results converged when Dt ¼ 0.02 and Dx ¼ 0.02. This means that the
method used in this article reduces the computational calculation and the CPU time
noticeably.
In addition, in the finite-volume method, the wall temperature prediction (or
one of the fluids) and its correction is required for each increment of time, which,
especially for heat exchangers with complex configurations, increases calculation
time. Figures 2 and 3 are presented for Eq. (27), but for other boundary conditions,
the results were also examined and confirmed.
Third, the results from the analytical solution of the Eqs. (4)–(6) were obtained
for steady condition. The result of the present work with the previously discussed
method at steady condition was also obtained; they confirm each other. See
Figure 4 for the result of the steady condition for fluid 1 at the tube lengths for
different NTU1.
The results obtained for all three different boundary conditions are presented
in Figures 5–7 at x ¼ X. With respect to the conditions described in Eq. (23), it is
possible to compare the effect of change in h1 P1 X and h2 P2 X for each boundary
condition. Figures 5–7 all show common behavior. A comparison of Figures 5a,
6a, and 7a with Figures 5b, 6b, and 7b shows that the energy conservation equation
is verified. With respect to disturbances (i.e., the temperature increase) occurrence on
fluid 1, at lower values of NTU1 (i.e., lower values of h1 P1 X and h2 P2 X Þ, the
maximum temperature increment can be noted for fluid 1. However, fluid 2 has
the lower values of the temperature change. As NTU1 increases, the temperature
change increased for fluid 2 and decreased for fluid 1.

Figure 4. Steady-state temperature of fluid 1 for different values of NTU.


TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 81

Figure 5. Response of (a) fluid 1 and (b) fluid 2 to sinusoidal inlet temperature change.

Furthermore, as NTU1 was increased to values greater than 1, the results


(curves) converged and became unique (This uniqueness presented for NTU1 values
of 3, 4, and 5 and is presented in the figures). For [18], the efficiency change of the
heat exchanger was plotted against NTU for different ratios of the thermal capacity
rate of the two fluids. For a parallel-flow heat exchanger with the m _ 1 c1 ¼ m
_ 2 c2
82 M. R. ANSARI AND V. MORTAZAVI

Figure 6. Response of (a) fluid 1 and (b) fluid 2 to exponential inlet temperature change.
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 83

Figure 7. Response of (a) fluid 1 and (b) fluid 2 to linear inlet temperature change.
84 M. R. ANSARI AND V. MORTAZAVI

condition, we demonstrate that when NTU values are greater than 2, the curves
become a straight line. With respect to the above explanations, it can be concluded
that the outlet response of a parallel flow heat exchanger with the m _ 1 c1 ¼ m
_ 2 c2 con-
dition is the same for any disturbance in the system for NTU values greater than 2.
We also conclude from the figures, taken together, that for any disturbances at x ¼ 0
and system response at x ¼ X there is a special time delay. This means that there is
no change up to the time Z  1 for any of the fluids and the wall at x ¼ X. The
implication is that the rate of L1 =D1 ¼ L2 =D2 shows the effect of the thermal
capacity rate of both fluids. The thermal capacity effect of the wall can be controlled
from the Dw =D1 ¼ 0:25 condition. The study of system response for the case of the
condition L2 =D2 > L1 D1 can lead to interesting results. Similar results have been
obtained in [7] for a shell-and-tube heat exchanger. Quantative comparison was
not possible, since detailed information was not presented, but the qualitative com-
parisons verify the agreement between the two methods.

5. CONCLUSION
The transient behavior of a double-pipe concentric heat exchanger with co-cur-
rent flow has been investigated. The advantage of the applied method has been that
the capability is higher for a combination of both numerical and analytical methods.
The results were analyzed for three types of boundary conditions in the forms of lin-
ear, exponential with negative power, and periodic functions for different NTUs. It
was shown that the outlet response of the heat exchanger for a defined change is
similar for NTU1 greater than 2, when the thermal capacity rate of both fluids is
the same. It is noted that there is a time delay between disturbances and system
response; this time delay illustrate the effect of the thermal capacities of both fluids
and the wall. Comparison of the results of the present method to results obtained by
the finite-volume method showed that the agreement between results is good. In
addition, the present model allows for the use of a larger potential range of elements,
causing the computational calculation and the computational CPU time to decline
noticeably.

REFERENCES
1. J. Yin and M. K. Jenson, Analytic Model for Transient Heat Exchanger Response, Int. J.
Heat Mass Transfer, vol. 46, pp. 3255–3264, 2003.
2. M. A. Abdelghani-Idrissi, F. Bagui, and L. Estel, Analytical and Experimental Response
Time to Flow Rate Step along a Counter Flow Double Pipe Heat Exchanger, Int. J. Heat
Mass Transfer, vol. 44, pp. 3721–3730, 2001.
3. F. E. Romie, Response of Counterflow Heat Exchangers to Step Changes of Flow Rates,
Int. J. Heat Transfer, vol. 121, pp. 746–748, 1999.
4. F. E. Romie, Transient Response of Counterflow Heat Exchangers, Int. J. Heat Transfer,
vol. 106, pp. 620–626, 1984.
5. M. N. Roppo and E. N. Ganic, Time-Dependent Heat Exchanger Modeling, Heat Trans-
fer Eng., vol. 4, no. 2, pp. 42–46, 1983.
6. D. J. Correa and J. L. Marchetti, Dynamic Simulation of Shell-and-Tube Heat Exchan-
gers, Heat Transfer Eng., vol. 8, no. 1, pp. 50–59, 1987.
TRANSIENT RESPONSE OF A CO-CURRENT HEAT EXCHANGER 85

7. W. Roetzel and Y. Xuan, Transient Behavior of Multipass Shell-and-Tube Heat Exchan-


gers, Int. J. Heat Mass Transfer, vol. 35, no. 3, pp. 703–710, 1992.
8. W. Roetzel and Y. Xuan, Transient Response of Parallel and Counterflow Heat Exchan-
gers, Int. J. Heat Transfer, vol. 114, pp. 510–512, 1992.
9. S. Alotaibi, M. Sen, B. Goodwine, and K.T. Yang, Numerical Simulation of the Thermal
Control of Heat Exchangers, Numer. Heat Transfer A, vol. 41, no. 3, pp. 229–244, 2002.
10. P. Asinari, Finite-Volume and Finite-Element Hybrid Technique for the Calculation of
Complex Heat Exchangers by Semiexplicit Method for Wall temperature Linked Equa-
tions (SEWTLE), Numer. Heat Transfer B, vol. 45, pp. 221–248, 2004.
11. J. M. Corberan, P. F. DeCordoba, J. Gonzalvez, and F. Alias, Semiexplicit Method for
Wall Temperature Linked Equations (SEWTLE): A General Finite-Volume Technique
for the Calculation Of Complex Heat Exchangers, Numer. Heat Transfer B, vol. 40,
pp. 37–59, 2001.
12. M. Tutar and A. Akkoca, Numerical Analysis of Fluid Flow and Heat Transfer Charac-
teristics in Three-Dimensional Plate Fin-and-Tube Heat Exchangers, Numer. Heat Trans-
fer A, vol. 46, pp. 301–321, 2004.
13. A. K. Saha and S. Acharya, Unsteady Flow and Heat Transfer in Parallel-Plate Heat
Exchangers with in-line- and Staggered Arrays of Posts, Numer. Heat Transfer A,
vol. 45, pp. 101–133, 2004.
14. R. Dolezal, Simulation of Large State Variations in Steam Power Plants, Springer Verlag,
Berlin, 1987.
15. R. Doležal, A. Rolf, M. Klug, and G. Riemenschneider, Solution of the Heat-Exchanger
Equation System, Proc. 1st IFAC Workshop, ‘‘Modeling and Control of Electric Power
Plants,’’ pp. 69–75, Como, Italy, 1983.
16. M. Lachi, N. EL Wakil, and J. Padet, The Time Constant of Double Pipe and One Pass
Shell-and-Tube Heat Exchangers in the Case of Varying Fluid Flow Rates, Int. J. Heat
Mass Transfer, vol. 40, pp. 2067–2079, 1997.
17. M. R. Ansari and V. Mortazavi, Simulation of Dynamical Response of a Countercurrent
Heat Exchanger to Inlet Temperature or Mass Flow Rate Change, Appl. Thermal Eng.,
vol. 26, pp. 2401–2408, 2006.
18. F. P. Incropera and D. P. Dewitt, Introduction to Heat Transfer, Wiley, New York, 2002.
19. V. Mortazavi, Simulation of Once-Through Boiler with IVD Model, M.Sc. thesis, Tarbiat
Modares University, Tehran, Iran, 2004 (in Persian).
All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately.

You might also like