R11 Chattopopadhyay Etal Analyasis of Heat Transfer

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

International Communications in Heat and Mass Transfer 33 (2006) 475 – 481

www.elsevier.com/locate/ichmt

Analysis of heat transfer in simultaneously developing pulsating


laminar flow in a pipe with constant wall temperature☆
Himadri Chattopadhyay a,⁎, Franz Durst b , Subhashis Ray c
a
Central Mechanical Engineering Research Institute, Durgapur – 713209, India
b
Lehrstuhl für Strömungsmechanik, Universität Erlangen, 91058 Erlangen, Germany
c
Department of Mechanical Engineering, Jadavpur University, Kolkata – 700 032, India
Available online 17 January 2006

Abstract

Numerical studies of flow and heat transfer in a circular tube under pulsating flow condition were carried out in the laminar
regime. The flow at the inlet consists of a fixed part and a pulsating component, which varies sinusoidally in time. The flow was
both thermally and hydrodynamically developing while the tube wall was kept at a uniform temperature. The solution of two-
dimensional Navier–Stokes equation was performed using the SIMPLE algorithm with the momentum interpolation technique of
Rhie and Chow. By analysing the data generated from the simulation, it is observed that in the range of present study (frequency:
0–20 Hz; amplitude: b1.0), pulsation has no effect on time-averaged heat transfer, although the Nu distribution varies in time in the
near-entry region of the pipe.
© 2005 Elsevier Ltd. All rights reserved.

Keywords: Simultaneously developing; Pulsating flow; Heat transfer; Laminar; Circular tube; Numerical

1. Introduction

Heat transfer under pulsating inlet condition is of interest to researchers as it is frequently encountered in different
engineering practices as well as in bio-fluid systems. Although flow development and heat transfer in channels and
ducts is quite well studied, transport phenomena in ducts with pulsating flow condition is yet to be fully understood. A
review of literature, a synopsis of which follows immediately, revealed that there is a wide confusion about the impact
of flow pulsation on the heat transfer in ducts.
Earlier works include the investigations Cho and Hyun [1] and Kim et al. [2]. In Cho and Hyun [1], the effect of
pulsation in a pipe using laminar boundary layer equations was numerically investigated. They have observed that at
the fully established downstream region, the Nusselt number may increase or decrease depending upon the frequency
parameter. The trend is amplified as amplitude increases. Kim et al. [2] studied the thermally developing channel flow
with developed inlet profile and isothermal channel wall. The outcome of the study was that the difference in time-
averaged heat flux for pulsating and corresponding steady-state condition was rather small.


Communicated by A.R. Balakrishnan and S. Jayanti.
⁎ Corresponding author.
E-mail address: chimadri@gmail.com (H. Chattopadhyay).

0735-1933/$ - see front matter © 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.icheatmasstransfer.2005.12.008
476 H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481

Nomenclature
A Amplitude
Cf Skin friction coefficient (2τ/ρw2)
d Diameter of pipe
h Heat transfer coefficient
Ga Graetz number
L Length of domain in x direction
Nu Time-averaged Nusselt number (hR/κ)
p Pressure
Pr Prandtl number (υ/α)
r Radial coordinate
R Radius of the pipe
Re Reynolds number (uavR/ν)
St Strouhal number (fR/uav)
t Time
T Temperature
Tp Time of a cycle
u Velocity component in radial direction
w Velocity component in axial direction
Wo Womersley number (R(w/υ)0.5)
z Axial coordinate

Greek symbols
α Diffusivity
η Enhancement ratio of Nu for pulsating flow
ν Kinematic viscosity
ρ Density
τ Shear stress
κ Thermal conductivity
ω Angular velocity of pulsation cycle

Subscripts
∞ Ambient
b Bulk
av Average
in Inlet
w Pipe wall

In the analytical solution of Moschandreou and Zamir [3] for the constant heat flux problem in a tube flow
pulsation was found to enhance the heat transfer in the frequency range corresponding to 5 b ϖ b 25. The enhance-
ments were more pronounced at higher Pr. However, later, Hemida et al. [4] demonstrated that the analytical
solution in Moschandreou and Zamir [3] was not correct. The analytical solution from Hemida et al. [4] yielded that
pulsation produces little change in Nu and the change is always negative, i.e. heat transfer decreases for the pulsating
systems. The effect of pulsation was more visible in thermally developing region as compared to the fully developed
region.
Guo and Sung [5] studied numerically the flow in a pipe with fully developed profile and constant heat flux
condition at the wall. They slowed that for small amplitude, heat transfer enhancement is observed within a zone
of pulsation frequency. At very large amplitudes of pulsation, the heat transfer due to pulsation is always
augmented.
Habib et al. [6] performed experimental studies on laminar pulsating air flow at Re between 780 and 2000 and the
frequency of 1–30 Hz. The tube wall was kept at constant heat flux condition. The heat transfer was augmented at
H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481 477

particulars zones of the pulsation frequency. The augmentation was as high as 30% at Re = 1400 and f = 1 Hz. The
improvement in heat transfer is more pronounced in the entrance region of the tube. Noticeably, for turbulent flow,
Gbadebo et al. [7] noticed a limited range of frequency where heat transfer enhancement could be evidenced.
In another experimental study (Kearney et al. [8]), the developing laminar regime was investigated using CARS
(Coherent Anti-Stokes Raman Scattering) and cold wire anemometry. The primary impact of reversed flow was found
to be an increase in the instantaneous thermal boundary layer thickness and thus a period of reduced instantaneous
Nusselt number. However, at certain pulsation range, heat transfer enhancement by a factor of two was discerned. The
enhancement was more pronounced at low values of thermal Graetz number.
From a review of current literature, it is found that there is a need for systematic assessment of the impact of flow
pulsation in heat transfer from circular ducts at constant wall temperature. A recently reported analytical study for a
constant heat flux case [9] concluded that pulsation does not enhance heat transfer in laminar tube flows. While there
is experimental evidence of heat transfer enhancement at certain frequency range, the analytical solutions are
predicting otherwise. Although such a study, it is also possible to assess whether the developing region has more
potential for enhancement. Above all, the present flow configuration represents a canonical problem for many heat
exchanger applications.
The present study is an attempt to investigate the heat transfer in a circular isothermal duct with imposed flow
pulsation at the inlet. The flow is hydrodynamically as well as thermally developing. The present study focuses within
the frequency range of 1–20 Hz and amplitude of b1.0 at a Reynolds number of 200. While we have defined the
Strouhal number as non-dimensional frequency, it is worth mention that in many cases Womerslay number (Wo) is the
standard non-dimensional parameter. However, Wo and St are related as Wo = (2πStRe)0.5. For the present values of
interest, Wo lies around 100 for St = 1. Systematic investigations in this region of parameters are lacking in open
literature. It may be mentioned here that the present range of Wo is particularly relevant in biological fluid mechanics
related to blood flow [10].

2. Problem formulation

The flow and heat transfer in pipe flows are governed by two-dimensional Navier–Stokes and energy
equations. These equations were non-dimensionalized and presented below. The length has been non-dimensio-
nalized by R (radius), velocity by average inlet velocity wav, time by R/wav and temperature is non-dimensiona-
lized as T = (T* − Tin) / (Tw* − Tin*). Here * is used to denote dimensional values.
Continuity equation:
 
1 AðruÞ Aw
þ ¼0 ð1Þ
r Ar Az
Momentum equation in radial direction:
 
Au Au Au Ap 1 u2
þu þw ¼− þ j u−
2
ð2Þ
At Ar Az Ar Re r
Momentum equation in axial direction:
Aw Aw Aw Ap 1
þu þw ¼ − þ j2 w ð3Þ
At Ar Az Az Re
Energy equation:
AT AT AT 1
þu þw ¼ j2 T ð4Þ
At Ar Az Red Pr
The usual no-slip boundary conditions are used at the wall and the symmetry condition at the centerline. At the
inlet, the velocity profile is found by adding the uniform velocity profile with a sinusoidal pulsation. Thus, the inlet
velocity profile is given by
 
w ¼ wav 1 þ Ad sinð2pSt tÞ ; ð5Þ

where A is the non-dimensional amplitude and St (the Strouhal number) is defined as fR/wav.
478 H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481

At the outlet the continuous boundary condition, where the second derivative of the primitive variables are set at
zero, is employed. This ensures that there is no abrupt transition at the outlet [11].
The governing equations were solved using the SIMPLE algorithm, a finite volume formulation of Patankar [12].
Despite the popularity of staggered-grid for such computations, a collocated grid arrangement for the primitive
variables was used. While the application of collocated grid is quite straightforward, many researchers used the
staggered grid to avid the ‘checkered-board’ pattern of pressure field. However, in the present work, the momentum
interpolation method due to Rhie and Chow [13] was used, which can satisfactorily overcome the problem of pressure
oscillation. The technique involves appropriate interpolation of velocity field at the interface of the cell faces. It can be
shown that such momentum interpolation is equivalent to solving a continuity equation with an added fourth
derivative of pressure and hence is often referred as added dissipation scheme. This scheme does not change the
formal second-order accuracy of the basic discretization process.
The tube surface was assumed to be isothermal. The heat transfer coefficient for this type of configuration is
generally calculated using bulk temperature Tb. In that case, the local heat transfer coefficient h is given by q = h(Tw
− Tb). However, as under the present circumstances, different inlet conditions (e.g. amplitude and frequency) are to be
imposed, it is more meaningful to compute Nu with respect to inlet temperature instead of the bulk temperature.
The value of Nui(x,t) is thus given by:

1 AT 
Nui ¼ ð6Þ
Tw −Tin Ar w
Here suffix i denotes instantaneous value.
The time and space averaged values can be obtained by integrating
Z L Z TP
Nuav ¼ Nui ðx;tÞdxdt=ðLTP Þ ð7Þ
0 0

Under the present non dimensional scheme, Cf is computed from the velocity gradient at the wall:

Aw
Cf d Re ¼ 2d
Az

3. Results and discussions

The numerical model was initially tested by computing developing flow in a tube without pulsation at the inlet. As such,
the results for such a situation are widely covered in literature. The length of the computational domain was varied for two

20
18
Re=100
16 Re=1000

14
12
Nu

10
8
6
4
2
0.05 0.1 0.15 0.2
X/RePr=1/Gz

Fig. 1. Distribution of Nu at simultaneously developing pipe flow.


H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481 479

-0.00012
-0.00013
-0.00014
-0.00015
u
-0.00016
-0.00017
-0.00018
-0.00019
0.12 0.14 0.16 0.18 0.20 0.22 0.24 0.26 0.28
w

Fig. 2. Phase plot of velocity (u,w) at an arbitrary location.

different levels of Re, as the developing length is a function of Re. In general, for Pr b 1.0, the development length is given
by the following relation: L/D ∼ 0.1 Re. [14]. Again, different grid distributions were needed for different Re, depending upon
the L/D ratio. Grid independences were checked at all Re for developing flow with comparing the results from the solution
of the Gratez problem and the analytical velocity distribution at the fully developed region. Typically 37 × 301 and 47 × 501
grids were used for Re = 100 and 1000, respectively. Fig. 1 establishes that the present code works well in predicting the
developing laminar flow. The fully developed region shows a Nu value of 3.66, as found in literature [14]. Similarly the
distribution of skin friction coefficient agreed well with literature and corresponding to fully developed region CfRe of 16
was obtained.
In the present work, investigations were performed within the range of 0.1 b St b 20 and 0.1 b A b 1.0 and at ReD = 200. After
performing the grid-independence study, a grid size of 302 × 42 were chosen for all the cases. 200 samples were collected in each
cycle and three consecutive cycles were taken for time averaging. From the sampled data at an arbitrary location (x = 1.0, r = 0.5), the
phase plot for the case of A = 0.5 and St = 10.0 is presented in Fig. 2. The closed curve of the plot establishes that the flow filed is
fully periodic following the periodic pulsation at the inlet.
Fig. 3 shows the instantaneous values of Nu during different phases (at 60° apart) of the cycle. The y-axis shows the ratio of
instantaneous Nu to the value of local Nu calculated for the case without pulsation. We define the enhancement ratio is η = Nu(x) /
Nusteady(x), where Nu is local time-averaged value of local Nusselt number and the suffix steady denotes the case when the inlet
profile is steady. Thus, a value of η over 1.0 indicates enhanced heat transfer at the location. It can be observed from the Fig. 3 that η
may be more or less than united depending upon the phase angle. However, the magnitude of increase and decrease is
complementary and hence when averaged out in time, there is virtually no visible effect of pulsation over heat transfer. It can be
discerned from this plot that the effect of pulsation is limited up to about x = 1.0 and the magnitude of the change is limited within
about ±4%.
From heat transfer point of view, the most important quantity of engineering interest is the time-averaged (and thus cycle-
averaged) Nusselt number, which is a direct measure of heat transfer from the tube surface. Distribution of time-averaged Nu with

1.04 600
1.03 1200
1800
1.02 2400
3000
1.01
1.00
0.99
0.98
0.97
0.96
0 1 2 3 4
x

Fig. 3. Enhancement ratio at different phase angle (A = 0.5, St = 10.0).


480 H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481

40
St=0.1
St=1.0
30 St=5.0
St=10.0

Nu
20

10

0
0 2 4 6 8 10 12 14 16 18 20
x

Fig. 4. Distribution of Nu at different St (A = 0.5).

amplitude of 0.5 at different frequencies is plotted in Fig. 4. It can be readily observed that they do not vary at all in the observed
range of frequency. Thus, no effect of pulsation is visible over the time-averaged value. While Fig. 5 shows that instantaneous Nu
behaved sinusoidally following the pulsation at the cycle, when averaged in time, no enhancement (or even decrease) in Nu is
attained.
It was observed from the computational results (as in Fig. 3) that instantaneous value of local Nusselt number increased
and decreased during the cycle. However, it is important for engineering calculations to obtain the time-averaged value in
the developing region. Table 1 shows the value of time and space-averaged skin friction coefficient and Nusselt number
which can be readily used to assess amount of pressure drop or heat transfer. The Cf value of 4.0 under present scheme of
non-dimensionalization corresponds to a value of 16, when diameter is the reference length. It can be seen from this table
that the value of Cfav decrease with increasing frequency. Similarly, at a particular value of frequency, Cfav reduces with
increasing value of amplitude. It can be mentioned here that a similar finding was reported by Zhao and Chung [15] in
connection with laminar reciprocating pipe flow. They have provided a correlation between A, f and Cf in the following
form:

Cf ¼ 3:722=ðAðRe0:548
k −2:039Þ

where Rek is kinetic Reynolds number based on characteristics length of ϖR.


It is an established fact [16] that under pulsating condition, relatively higher axial flow occurs near the pipe wall and is
known as the annular effect. Thus, the wall gradient of velocity diminishes with increasing A or St. However, while pulsation
directly affects the skin friction, it can be observed from the Table 1 that there is no discernible change in the value of Nuav
over the range of the frequency and amplitude. The average value of Nu of 0.96 based on the temperature difference between

Nui
1.1196

1.1191

1.1186

1.1181

1.1176
0 90 180 270 360

Fig. 5. Distribution of Nui at x = 5.0, during a cycle at St = 10, A = 0.2.


H. Chattopadhyay et al. / International Communications in Heat and Mass Transfer 33 (2006) 475–481 481

Table 1
Results at a glance
A St Cfav Nuav
0.0 0.0 4.00 0.96
0.2 1.0 3.99 0.96
0.2 5.0 3.97 0.96
0.2 10.0 3.94 0.96
0.5 1 3.99 0.96
0.5 10 3.77 0.96
0.5 20 2.225 0.96
0.8 10 2.225 0.96

the tube wall and inlet corresponds to a value of 3.66, when Nu is based on the bulk temperature calculation and diameter as
the characteristics length.

4. Concluding remarks

Simulations were performed for observing the effect of flow pulsation over the transport processes in the laminar
regime of developing pipe flow. It was observed from the detailed investigation that, while the instantaneous Nu
follows the pulsation cycle within the initial length of about 2R, the pulsation has no positive effect in augmenting heat
transfer within the investigated range of pulsation frequency and amplitude.

Acknowledgement

H. C. acknowledges gratefully the support of DAAD and LSTM, Erlangen for funding his research stay in
Germany.

References

[1] H.W. Cho, J.M. Hyun, Numerical solutions of pulsating flow and heat transfer characteristics in a pipe, Int. J. Heat Fluid Flow 11 (1990)
321–330.
[2] S.Y. Kim, B.H. Kang, J.M. Yuan, Heat transfer in the thermally developing region of a pulsating channel flow, Int. J. Heat Mass Transfer 36
(1993) 4257–4266.
[3] T. Moschandreou, M. Zamir, Heat transfer in a tube with pulsating flow and constant heat flux, Int. J. Heat Mass Transfer 40 (1997) 2461–2466.
[4] H.N. Hemida, M.N. Sabry, A. Abdel-Rahim, H. Mansour, Theoretical analysis of heat transfer in laminar pulsating flow, Int. J. Heat Mass
Transfer 45 (2002) 1767–1780.
[5] Z. Guo, H.J. Sung, Analysis of Nusselt number in pulsating pipe flow, Int. J. Heat Mass Transfer 49 (1997) 2486–2489.
[6] M.A. Habib, A.M. Attya, A.I. Eid, A.Z. Aly, Convective heat transfer characteristics of laminar pulsating pipe air flow, Heat Mass Transfer 38
(2002) 221–232.
[7] S.A. Gbadebo, S.A.M. Said, M.A. Habib, Average Nusselt number correlation in the thermal entrance region of steady and pulsating turbulent
pipe flow, Heat Mass Transfer 35 (1999) 377–381.
[8] S.P. Kearney, A.M. Jacobi, R.P. Lutchi, Time-resolved thermal boundary layer structure in a pulsatile reversing channel flow, J. Heat Transfer
123 (2001) 655–663.
[9] J.C. Yu, Z. Li, T.S. Zhao, An analytical study of pulsatile laminar heat convection in a circular tube with constant heat flux, Int. J. Heat Mass
Transfer 47 (2004) 5297–5301.
[10] S.A Berger, W. Goldsmith, E.R. Lewis, Introduction to Bioengineering, Oxford University Press, 1996.
[11] G. Biswas, H. Chattopadhyay, Heat transfer in a channel with built-in wing type vortex generators, Int. J. Heat Mass Transfer 35 (1994)
803–814.
[12] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York, 1980.
[13] C.M. Rhie, W.L. Chow, Numerical study of the turbulent flow past an airfoil with trailing edge separation, AIAA J. 21 (11) (1983) 1525–1532.
[14] W.M. Kays, M.E. Crawford, Convective Heat and Mass Transfer, McGraw-Hill, New York, 1970.
[15] T.S. Zhao, P. Cheng, The friction coefficient of a fully developed laminar reciprocating flow in a circular pipe, Int. J. Heat Fluid Flow 17 (1996)
167–172.
[16] J.R. Womersley, Method for calculation of velocity, rate of the flow, and viscous drag in arteries when the pressure gradient is known, J. Physiol.
127 (1955) 553–563.

You might also like