Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Energy Conversion and Management 47 (2006) 3473–3486

www.elsevier.com/locate/enconman

Evaluation of heat transfer and exergy loss in a concentric


double pipe exchanger equipped with helical wires
Ebru Kavak Akpinar *

Mechanical Engineering Department, Firat University, TR-23119 Elazig, Turkey

Received 24 December 2004; received in revised form 14 July 2005; accepted 24 December 2005
Available online 21 February 2006

Abstract

In this study, the effects on heat transfer, friction factor and dimensionless exergy loss were investigated experimentally
by mounting helical (spring shaped) wires of different pitch in the inner pipe in a double pipe heat exchanger. In the exper-
iments, hot (air) and cold (water) fluids flowed through the inner pipe and annulus, respectively. The experiments were
performed for both parallel and counter current flow modes of the fluids at Reynolds numbers between 6500 and
13,000. An augmentation of up to 2.64 times in Nusselt number compared to the empty pipe was obtained in the helical
system. The increase in friction factor was about 2.74 times that of the empty pipe, depending on Reynolds number and the
pitch or helical number. An augmentation of up to 1.16 times in the dimensionless exergy loss compared to the empty pipe
was obtained in the helical system. Some empirical correlations expressing the results were also derived and discussed.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Concentric double pipe heat exchanger; Helical wires; Heat transfer; Exergy loss; Friction factor

1. Introduction

In the past decade, heat transfer enhancement technology has been developed and widely applied to heat
exchanger applications; for example, refrigeration, automotive, process industry, solar water heater, etc. [1].
The aim of augmenting heat transfer is to accommodate high heat fluxes (or heat transfer coefficients). To
date, there have been a large number of attempts to reduce the size and costs of heat exchangers. The most
significant variables in reducing the size and cost of a heat exchanger are basically the heat transfer coefficient
and pressure drop. An increase in the heat transfer coefficient generally leads to another advantage of reducing
the temperature driving force, which increases the second law efficiency and decreases entropy generation.
Thus, research in this area has captivated the interest of a number of researchers [1].
Heat transfer enhancement in heat exchangers may be achieved by numerous techniques, and these tech-
niques can be classified into three groups: passive, active and compound techniques [2,3]. In the active

*
Tel.: +90 424 237 0000/5325; fax: +90 424 241 5526.
E-mail addresses: eakpinar@firat.edu.tr, kavakebru@hotmail.com

0196-8904/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2005.12.014
3474 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

Nomenclature

A heat transfer area (m2)


Ai inner pipe inside surface area (m2)
Ao inner pipe outer surface area (m2)
Cp specific heat capacity (J/kg C)
Cmax maximum heat capacity (W/K)
Cmin minimum heat capacity (W/K)
Cr heat capacity ratio (dimensionless)
d diameter of inner pipe (m)
D diameter of outer pipe (m)
e dimensionless exergy loss rate (dimensionless)
Ex exergy
Ex_ loss
exergy loss rate (W)
f friction factor (dimensionless)
h specific enthalpy (J/kg)
Hc average heat transfer coefficient of cold fluid (W/m2 K)
Hh average heat transfer coefficient of hot fluid (W/m2 K)
hn helical number
k thermal conductivity of fluid (W/m K)
L length of pipe (m)
m_ mass flow rate (kg/s)
n number constants
N number of observations
Nu Nusselt number (dimensionless)
NTU number of heat transfer units (dimensionless)
Pr Prandtl number (dimensionless)
p pitch (mm)
Q_ heat transfer rate (W)
Q_ j time rate of heat transfers at location on boundary of control volume where instantaneous tem-
perature is Tj (W)
R correlation coefficient
Re Reynolds number (dimensionless)
RMSE root mean square error
s specific entropy (J/kg K)
T temperature (C)
Tr temperature ratio (dimensionless)
Tj instantaneous temperature (C)
U overall heat transfer coefficient (W/m2 K)
Um average fluid velocity (m/s)
W total uncertainty in measurement
W_ work rate (W)
W_ cv time rate of energy transfer to control volume by work (W)
Yexp experimental value
Ypre predicted value
DP pressure drop (Pa)
DTli logarithmic temperature difference between wall and air temperature (K)
DTlo logarithmic temperature difference between wall and water temperature (K)
DTL logarithmic mean temperature difference (K)
q density of fluid (kg/m3)
E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3475

e effectiveness (dimensionless)
v2 chi-square

Subscripts
c cold
CH chemical
e environmental condition
h hot
i inlet condition or inner
KN kinetic
loss loss
max maximum
min minimum
o outlet condition or outer
PH physical
PT potential

techniques, heat transfer is improved by giving additional flow energy to the fluid. In the passive techniques,
however, this improvement is acquired without providing any extra flow energy. In the compound techniques,
two or more of the active or passive techniques may be utilized simultaneously to produce an enhancement
that is much higher than that of the techniques operating separately [3,4].
The thermal boundary layer is of prime importance in the heat transfer between the fluid and the pipe wall.
It is associated with the type of fluid flow, and its thickness is greater in laminar flow. Thus, heat transfer in
turbulent flow occurs faster than that occurring in laminar flow due to both the smaller thickness of the ther-
mal boundary layer and the eddies transporting thermal energy very quickly from place to place in the turbu-
lent core of the flow.
Exergy analysis can reveal whether or not and by how much it is possible to design more efficient thermal
systems by reducing the sources of existing inefficiencies. Increased efficiency can often contribute in a major
way to achieving energy security in an environmentally acceptable way by the direct reduction of irreversibil-
ities that might otherwise have occurred. This makes exergy one of most powerful tools to provide optimum
conditions [5,6].
Although numerous studies on the energy and exergy analysis of thermal systems and applications have
recently been undertaken by some researchers, e.g., Refs. [5–12], very few papers have appeared on exergy
analyses of concentric heat exchangers with swirl generators, e.g., Refs. [13–15]. The detailed literature review
for the present work has shown that there is no information on the exergy analysis of concentric double pipe
exchangers with spaced helical wires.
The purpose of this work is, differently than in the literature study,

• to increase the heat transfer further with a passive method by mounting helical wires of different pitch or
helical number in the inner pipe in a concentric double pipe heat exchanger,
• to study the effect of helical wires on the heat transfer, friction factor and dimensionless exergy loss.

2. Experimental set-up and procedure

The experimental set-up is shown in Fig. 1. The dimensions of the inner and outer pipes of the heat exchan-
ger are: di = 70 mm, do = 75 mm, Li = 1265 mm, Di = 100 mm, Do = 110 mm and Lo = 1200 mm. Hot air was
passed through the inner pipe, while cold water was flowing through the annulus. Heating of the air was
achieved with an electrical heater at the entrance, and its power was adjusted with a rheostat.
3476 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

Fig. 1. Experimental set-up.

In the experimental work, it is intended to search for the changes in the heat transfer coefficients of the air
side turbulent flow by affecting the regions near the wall of the pipe flow. For this purpose, the helical wire
elements were installed on the air side of the pipe as special resistances. These wires have various pitches to
affect the boundary layer. The wire diameter of the helical wires placed inside the inner pipe to improve
the heat transfer was 4 mm and the pitch was chosen as 9, 15 and 21 mm (helical numbers of 137, 84 and
60 and helical angles of 237 0 , 396 0 and 553 0 ). The helical wire was arranged as shown in Fig. 1. A separate
experiment was also conducted for an empty heat exchanger without using the helical wire.
The inlet and outlet temperatures of the fluids (cold water and hot air), the temperatures of the points on
the inner pipe wall (arithmetical average of five points) and ambient temperature were measured with iron-
constantan thermocouples. The temperatures were read with a multi-channel digital thermometer. Pressure
losses were determined by using U manometers that were filled with water for the air side and with mercury
for the water side. Measurements points of the temperatures and the pressure losses are shown in Fig. 1. The
air was supplied from an air compressor. The flow rates of the air and water were adjusted with valves and
measured with rotameters. In order to keep constant air flow rates at the various levels, a pressure regulator
was also used. The experimental work was repeated for both parallel and countercurrent flow modes at var-
ious fluids Reynolds numbers.

3. Experimental uncertainty

All instruments and measurements have certain general characteristics. An understanding of these common
qualities is the first step towards accurate measurement. Errors and uncertainties are inherent in both the
instrument and the process of making the measurement, and too much reliance should not be placed on
any single reading from one affected by the environment. Final accuracy depends on a sound program and
on correct methods for taking readings on the proper instruments. When readings are repeated, they tend
to produce a band of results rather than a point or a line. Errors and uncertainties in the experiments can arise
from instrument selection, instrument condition, instrument calibration, environment, observation, reading
and test planning [16,17]. In the experiments of heat transfer enhancements in a concentric double pipe
exchanger equipped with helical wires, the temperatures, pressure drops and flow rates were measured with
appropriate instruments. During the measurements of the parameters, the uncertainties that occurred are pre-
sented in Table 1. Considering the relative errors in the individual factors as denoted by xn, error estimation
was made using the following equation [18]:
E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3477

Table 1
Uncertainties of the parameters during the experiments of heat transfer enhancements in a concentric double pipe exchanger equipped
with helical wires
Parameter Unit Comment
Uncertainty in temperature measurement
Hot fluid inlet temperature C ±0.380
Hot fluid outlet temperature C ±0.380
Cold fluid inlet temperature C ±0.380
Cold fluid outlet temperature C ±0.380
Inner pipe wall temperature C ±0.628
Ambient temperature C ±0.380
Uncertainty in measurement of pressure drop
Air side mm SS ±0.282
Water side mm Hg ±0.282
Uncertainty in measurement of volume flow rate
Air m3/h ±0.2
Water m3/h ±0.2
Uncertainty in reading values of table (q, cp, k, etc.) % ±0.1–0.2

h i1=2
2 2 2
W ¼ ðx1 Þ þ ðx2 Þ þ    þ ðxn Þ . ð1Þ

4. Analysis

4.1. Heat transfer

The efficiency of a heat exchanger is defined as the ratio of the heat transferred to the maximum heat
transfer:
Q_
e¼ . ð2Þ
Q_ max
The heat transferred from the hot air is given as,

Q_ h ¼ m_ h C ph ðT hi  T ho Þ ¼ H h Ai DT li ; ð3Þ

and the heat received by the cold water is

Q_ c ¼ m_ c C pc ðT co  T ci Þ ¼ H c Ao DT lo . ð4Þ

The difference between these two rates shows the convective heat loss from the exchanger, which may be as-
sumed to be negligible here
Q_ loss ¼ Q_ h  Q_ c . ð5Þ
The Nusselt number, in terms of the air side average heat transfer coefficient, is
H hd i
Nu ¼ . ð6Þ
k
On the other hand, the overall heat transfer coefficient may be defined by the equation
Q_ h ¼ U i Ai DT L ; ð7Þ
where DTL is the logarithmic mean temperature difference over the exchanger. The overall heat transfer coef-
ficient may be calculated using the measured wall temperatures of the inner pipe (arithmetical average of five
points) and Eqs. (3) and (4). Eq. (7) may then be used to display the results of the calculation [15,16,19,20].
3478 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

The following equation can be used to calculate the maximum heat given:
Q_ max ¼ C min ðT hi  T ci Þ ¼ m_ h C ph ðT hi  T ci Þ. ð8Þ
The number of heat transfer units, NTU, is expressed in terms of thermal capacity,
UA
NTU ¼ . ð9Þ
C min

4.2. Exergy analysis

Exergy is the maximum amount of work that can be obtained theoretically at the end of a reversible process
in which complete equilibrium with the environment is attained. According to this definition, the reference
environment conditions must be known to calculate exergy. The temperature of the reference environment
in this work varied between 20 and 22 C (ambient temperature).
A heat exchanger is characterized by two types of losses: from temperature difference and frictional pres-
sure drop in the pipe. These losses refer to irreversibility quantities, and some methods have been devised for
minimizing these losses [21,22]. However, in this study, the exergy analysis does not include friction (or pres-
sure drop) irreversibilities and is based only on heat transfer irreversibilities.
The total exergy of a system Ex can be divided into four components, namely (i) physical exergy ExPH,
(ii) kinetic exergy ExKN, (iii) potential exergy ExPT and (iv) chemical exergy ExCH [23]:
Ex ¼ ExPH þ ExKN þ ExPT þ ExCH . ð10Þ
Physical exergy is the majority of the exergy of a heat exchanger system. Therefore, chemical exergy, poten-
tial exergy, nuclear exergy, magnetic exergy and kinetic exergy (kinetic energy) were neglected in this study.
In this case, the exergy balance in a steady open system can be written as follows [24]:
X X X
_ i
Ex _ oþ
Ex _ product ¼ 0.
Ex ð11Þ
The lost work is being described as the difference between the maximum work and the real work,
W_ lost ¼ W_ max  W_ actual ¼ Ex
_ loss ; ð12Þ
and this expression is equal to the exergy loss rate. Therefore, the exergy loss rate in open systems (control
volume) is
X X X  Te

_ loss ¼
Ex m_ i ðhi  T e si Þ  m_ o ðho  T e so Þ þ Q_ j 1   W_ cv . ð13Þ
Tj
Here, the term Q_ j represents the time rate of heat transfers occurring at the jth location on the boundary of the
control volume where the instantaneous temperature is Tj. W_ cv represents the time rate of exergy transfer by
work other than flow work.
Fig. 2 shows an exergy band diagram for a heat exchanger [15]. The exergy loss rate in an open system heat
exchanger is given as,

Fig. 2. Exergy band diagram for heat exchanger.


E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3479

_ loss ¼ m_ c ðhci  T e sci Þ þ m_ h ðhhi  T e shi Þ  m_ c ðhco  T e sco Þ  m_ h ðhho  T e sho Þ;


Ex ð14Þ

and rearranging this equation gives the following:


_ loss ¼ m_ c ðhci  hco Þ þ m_ h ðhhi  hho Þ þ T e ½m_ c ðsco  sci Þ þ m_ h ðsho  shi Þ.
Ex ð15Þ

In a heat exchanger, the heat given by the hot fluid becomes equal to the heat taken by the cold fluid if the heat
losses Q_ loss are assumed to be negligible,
Q_ ¼ m_ h ðhhi  hho Þ ¼ m_ c ðhci  hco Þ. ð16Þ
By using Eq. (15), the following equation can be derived:
_ loss ¼ T e ½m_ c ðsco  sci Þ þ m_ h ðsho  shi Þ.
Ex ð17Þ
If the entropy changes of the hot and cold fluids are expressed in terms of the specific heats at constant
pressure,
sco  sci ¼ C pc lnðT co =T ci Þ; sho  shi ¼ C ph lnðT ho =T hi Þ; ð18Þ
and inserting Eq. (18) into Eq. (17),
_ loss ¼ T e ½m_ h C ph lnðT ho =T hi Þ þ m_ c C pc lnðT co =T ci Þ.
Ex ð19Þ
The thermal capacities of the hot and cold fluids and the ratio of capacities are,
C h ¼ m_ h C ph ; C c ¼ m_ c C pc ; C r ¼ C min =C max . ð20Þ
If we assume that the thermal capacity of the hot fluids is the minimum, the exergy loss rate can be written as
follows:
_ loss ¼ T e ½C min lnðT ho =T hi Þ þ C max lnðT co =T ci Þ;
Ex ð21Þ
and if the following relations are described:
T hi Q_ C max ðT co  T ci Þ C min ðT hi  T ho Þ
Tr ¼ ; e¼ ¼ ¼ ; ð22Þ
T ci Q_ max C min ðT hi  T ci Þ C min ðT hi  T ci Þ
and inserting Eqs. (20) and (22) into Eq. (21), the exergy loss rate equation is obtained as follows:
    
_ loss ¼ T e C min ln 1  e 1  1
Ex þ C max ln ½1 þ eC r ðT r  1Þ . ð23Þ
Tr
If Eq. (23) is divided by (TeCmin), the dimensionless exergy loss rate expression can be written as follows:
  
_ loss
Ex 1 1
e¼ ¼ ln 1  e 1  þ ln ½1 þ eC r ðT r  1Þ. ð24Þ
T e C min Tr Cr

4.3. Friction factor

The experimental pressure drops in the inner test pipe of the concentric heat exchanger were measured for
the air side conditions and arranged in non-dimensional form by using the following equation:
DP
f ¼ . ð25Þ
ðLi =d i ÞqU 2m =2

5. Regression analysis

The regression analysis was performed using the Statistica computer program. The correlation coefficient
(R), the reduced chi-square (v2) and root mean square error analysis (RMSE) were used to determine the qual-
ity of the developed relation. These parameters can be calculated as follows [15,17]:
3480 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

Pn 2
2 i¼1 ðY exp;i
 Y pre;i Þ
v ¼ ; ð26Þ
N n
" #1=2
1 XN
RMSE ¼ ðY pre;i  Y exp;i Þ2 . ð27Þ
N i¼1

6. Results and discussion

With the values obtained from the experimental data in the inner pipe, the changes in Nusselt number with
Reynolds number were drawn for the contained helical wires at different pitches or helical numbers, as shown
in Fig. 3. The experiments were performed for both counter and parallel flow modes, and the results were com-
pared to those obtained from the empty pipe and the Dittus–Boelter correlation Nu = 0.023Re0.8 Æ Pr0.4
(describes flow in the smooth tube) [25].
It is clear from Fig. 3 that the highest Nusselt number was achieved with the heat exchanger operated in a
counter flow mode and equipped with the helical wires having the pitch of 9 mm or the helical number of

120
p=9 mm, hn=137 Parallel flow
p=15 mm, hn=84
100 p=21 mm, hn=60
Empty pipe
80 Dittus-Boelter correlation
Nusselt number

60

40

20

0
6000 7000 8000 9000 10000 11000 12000 13000 14000
Reynolds number

140
p=9 mm, hn=137 Counter flow
p=15 mm, hn=84
120
p=21 mm, hn=60
Empty pipe
100
Dittus-Boelter correlation
Nusselt number

80

60

40

20

0
6000 7000 8000 9000 10000 11000 12000 13000 14000
Reynolds number

Fig. 3. Variation of Nusselt number with Reynolds number in parallel and counter flow.
E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3481

137. In this run, an increase in the Nusselt number was obtained up to 2.64 times compared to a heat exchan-
ger with an empty pipe. In all cases, the heat exchanger with the helical wires gave higher values of Nusselt
number than those for the empty pipe and the Dittus–Boelter regime. It was determined that the heat trans-
fer rates increased with decreasing pitch or increasing helical number of the helical wires used in the exper-
iments. This increase in Nusselt number can be explained by two mechanisms: Firstly, the surfaces of the
helical wires may act as extended heat transfer surfaces, which may be considered to be insignificant. Sec-
ondly, the helical wires placed in the inner pipe may act as turbulators in the flow medium and cause
increased turbulent intensities. So, the helical wires retard the development of boundary layers in the pipe,
and the velocity and temperature profiles approach those in plug flow. The turbulent intensities in the inner
pipe must be higher with decreasing pitch and with increasing number of helixes of the helical wires. How-
ever, from this figure, it is also seen that the effect of the helical wires on the heat transfer is less for low
values of the Reynolds numbers. Thus, the relative increase in the Nusselt number was low at smaller Rey-
nolds numbers, while it became greater at higher Reynolds numbers. The enhancements in heat transfer rates
decreased with increasing pitch and with decreasing number of helixes and are 5–10% smaller for the pitch of
15 mm and the helical number of 84 than those with the pitch of 9 mm and the helical number of 137 and are
5–10% smaller for the pitch of 21 mm and the helical number of 60 than those with the pitch of 15 mm and
the helical number of 84. The improvements for parallel flow showed a parallel trend, but, the values found
for the parallel flow mode were also 5–10% lower than the corresponding values obtained in the counter flow
mode as seen in Fig. 3.
Fig. 4 shows the changes of the effectiveness with NTU for heat exchangers containing helical wires at dif-
ferent pitches or helical numbers. It was observed from this figure that the effectiveness increased with the
increase of NTU. The highest effectiveness was obtained in a counter flow mode by a helical wire having
the helical pitch of 9 mm or helical number of 137. The increase in the effectiveness was obtained up to
1.16 times compared to the heat exchanger with empty pipe for the counter flow mode. The increase of
NTU was about 2.64 times that of the empty tube.
The pressure drops were found to be about equal for both flow modes and increased with Reynolds
numbers. The friction factors were calculated by using Eq. (25) based on the pressure drops. The friction
factors in the inner test pipe are given in Fig. 5 as a function of the Reynolds number for the air side and
the various helical wires containing helixes at different pitches or helical numbers. A great increase in fric-
tion factor occurred when the helical wire was mounted in the inner pipe in comparison with that in the
inner pipe without the helical wire. This results mainly from the dissipation of the dynamic pressure of
the fluid (i.e., air) due to very high viscous losses near the pipe wall. Moreover, the pressure drop increase
is probably due to secondary flows occurring as a result of the interaction of pressure forces with inertial
forces in the boundary layer. It was seen from Fig. 5 that the friction factors in the inner pipe decreased
with the increase of the Reynolds number and increased with the decrease of the pitch or with the increase
of the helical number. There was an increase up to 2.74 times in the friction factor for the helical wire hav-
ing the pitch of 9 mm or the helical number of 137 when it was compared to a heat exchanger with empty
pipe.
Fig. 6 shows the changes of the dimensionless exergy loss with NTU for helical wires containing helixes at
different pitches or numbers. The exergy loss increased with the increase of NTU. In addition, the exery loss
and NTU increased with the increase of helical number or decreased with the increase of the pitch. The
increase of the dimensionless exergy loss was about 1.16 times that of the empty tube at the highest Reynolds
number for the helical wire having the pitch of 9 mm and helical number of 137.
Linear and non-linear regression models are important tools to find the relationship between different
variables, especially those for which no established empirical relation exists. In this study, the empirical
relations were derived for the variation of Nusselt number with Reynolds number, the variation of effec-
tiveness with NTU, the variation of friction factor with Reynolds number and the variation of the dimen-
sionless exergy loss with NTU. Based on the multiple regression analysis, the accepted relations and
correlation coefficients, RMSE and v2 values are given in the following. It was shown that the accepted
empirical relations describing values of the heat transfer, the effectiveness, the friction factor and the dimen-
sionless exergy loss of concentric heat exchangers with helical wire gave very high R values and low RMSE
and v2 values.
3482 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

1
Parallel flow
0.9

0.8
Effectiveness

0.7

0.6 p=9 mm, hn=137


p=15 mm, hn=84
0.5 p=21 mm, hn=60
Empty pipe
0.4
0.4 1.4 2.4 3.4 4.4 5.4 6.4 7.4 8.4 9.4
NTU=(UA)/Cmin

1
Counter flow
0.9

0.8
Effectiveness

0.7

0.6 p=9 mm, hn=137


p=15 mm, hn=84
0.5 p=21mm, hn=60
Empty pipe
0.4
0 1 2 3 4 5 6 7 8 9 10 11
NTU=(UA)/Cmin

Fig. 4. Variation of effectiveness with NTU in parallel and counter flow.

0.006
Parallel-counter flow p=9 mm, hn=137
0.005 p=15 mm, hn=84
p=21 mm, hn=60
0.004
Friction factor

Empty pipe

0.003

0.002

0.001

0.000
6000 7000 8000 9000 10000 11000 12000 13000 14000

Reynolds number
Fig. 5. Variation of friction factor with Reynolds number in parallel and counter flow.
E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3483

0.16
Parallel flow
0.14

Dimensionless exergy loss 0.12

0.1
p=9 mm, hn=137
0.08
p=15 mm, hn=84
0.06 p=21 mm, hn=60
Empty pipe
0.04
0 1 2 3 4 5 6 7 8 9
NTU=(UA)/C min

0.16
Counter flow
0.14
Dimensionless exergy loss

0.12

0.1
p=9 mm, hn=137
0.08
p=15 mm, hn=84
0.06 p=21 mm, hn=60
Empty pipe
0.04
0 1 2 3 4 5 6 7 8 9 10 11
NTU=(UA)/Cmin

Fig. 6. Variation of dimensionless exergy loss with NTU in parallel and counter flow.

140
130 y = 0.8373x + 10.572 Parallel flow
Predicted Nusselt number values

120 R = 0.914
y = 1.0036x - 0.312 Counter flow
110
R = 0.985
100
90
80
70
60
50 Parallel flow
40 Counter flow
30
30 40 50 60 70 80 90 100 110 120 130 140
Experimental nusselt number values

Fig. 7. Comparison of experimental and predicted Nusselt number values by Eqs. (28) and (29).
3484 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

For a heat exchanger equipped with helical wires,


Nu ¼ 0:661  ðRe0:876 Þ  ðPr15:028 Þ  ðhn0:374 Þ in parallel flow,
2
R ¼ 0:914; RMSE ¼ 6:844; v ¼ 62:45; ð28Þ
1:137 21:771 0:419
Nu ¼ 0:593  ðRe Þ  ðPr Þ  ðhn Þ in counter flow,
2
R ¼ 0:985; RMSE ¼ 4:063; v ¼ 22:00; ð29Þ
0:589 0:316
e ¼ 1:102  ðNTU Þ  ðhn Þ in parallel flow,
R ¼ 0:885; RMSE ¼ 0:0500; v2 ¼ 0:00308; ð30Þ
0:537 0:266
e ¼ 0:860  ðNTU Þ  ðhn Þ in counter flow,
R ¼ 0:965; RMSE ¼ 0:0273; v2 ¼ 0:000919; ð31Þ
0:514 0:445
f ¼ 0:0333  ðRe Þ  ðhn Þ in parallel and counter flow,
R ¼ 0:964; RMSE ¼ 0:000204; v2 ¼ 5:13  108 ; ð32Þ
0:581 0:311
e ¼ 0:181  ðNTU Þ  ðhn Þ in parallel flow,
R ¼ 0:885; RMSE ¼ 0:00816; v2 ¼ 8:20  105 ; ð33Þ
0:530 0:262
e ¼ 0:141  ðNTU Þ  ðhn Þ in counter flow,
R ¼ 0:965; RMSE ¼ 0:00444; v2 ¼ 2:43  105 . ð34Þ
Here, (hn) is the helical number of the helical wires.

1
y = 0.7826x + 0.153 Parallel flow
R = 0.885
Predicted effectiveness values

0.9
y = 0.9359x + 0.0446 Counter flow
0.8 R = 0.965

0.7

0.6

0.5 Parallel flow


Counter flow
0.4
0.4 0.5 0.6 0.7 0.8 0.9 1
Experimental effectiveness values

Fig. 8. Comparison of experimental and predicted effectiveness values by Eqs. (30) and (31).

0.005
y = 0.9199x + 0.0002 Parallel-counter flow
Predicted friction factor values

R = 0.964
0.004

0.003

0.002

0.001
0.001 0.002 0.003 0.004 0.005
Experimental friction factor values

Fig. 9. Comparison of experimental and predicted friction factor values by Eq. (32).
E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486 3485

Predicted dimensionless exergy loss values


0.16
y = 0.7827x + 0.0253 parallel flow
0.15
R = 0.885
0.14 y = 0.9362x + 0.0073 counter flow
0.13 R = 0.965
0.12
0.11
0.1
0.09
parallel flow
0.08
counter flow
0.07
0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14 0.15 0.16
Experimental dimensionless exergy loss values

Fig. 10. Comparison of experimental and predicted dimensionless exergy loss values by Eqs. (33) and (34).

The accuracy of the established empirical relations was evaluated by comparing the computed data in
any particular experimental conditions with the observed data. The performance of the empirical relationships
of Nusselt number (Eqs. (28) and (29)), the effectiveness (Eqs. (30) and (31)), the friction factor (Eq. (32))
and the dimensionless exergy loss (Eqs. (33) and (34)) at all experimental conditions have been illustrated
in Figs. 7–10 for both flow modes. The predicted data generally banded around the straight line, which showed
the suitability of the empirical relations in describing the Nusselt number, the effectiveness, the friction factor
and the dimensionless exergy loss value in any particular experimental conditions. In addition, as seen
from Figs. 7–10, the measurement results are reasonably close to the empirical results. The causes of the
errors between the empirical and the experimental values can be considered as resulting from the setting
and measurements of temperature, volume flow rate and pressure of air in the system under laboratory
conditions.

7. Conclusions

The present study explored the effect of different helical wires on the heat transfer, friction factor and
dimensionless exergy loss in a double concentric pipe heat exchanger. The key findings from the study may
be summarized as follows:

1. Heat transfer rates increased with decreasing pitch and with increasing helical number of the helical wires
used in the experiments. The highest enhancement was seen to occur in counter current flow mode of the
exchanger with the helical wire having the pitch of 9 mm and the helical number of 137. The heat transfer
rates in this heat exchanger increased up to 2.64 times with the help of the helical wires.
2. The effectiveness values of the heat exchanger that used helical wires in the inner pipe was obtained up to
1.16 times compared to a heat exchanger with the empty pipe for the counter flow mode.
3. Helical wires caused a considerable increase in pressure drop and friction factor. In the case of both counter
and parallel flow modes, the average increase in friction factors at the highest Reynolds number for the heli-
cal wire having the pitch of 9 mm and the helical number of 137 was by a factor of about 2.74 times in
comparison with that for the inner pipe without helical wire.
4. The dimensionless exergy loss and NTU increased with the increase of helical number and decreased with
the increase of pitch. The increase was about 1.16 times that of the empty tube at the highest Reynolds
number for helical wire having the pitch of 9 mm and the helical number of 137. When the increase in
dimensionless exergy loss is compared with the increase in heat transfer, a counter flow heat exchanger with
helical wire was determined to be more advantageous.
3486 E.K. Akpinar / Energy Conversion and Management 47 (2006) 3473–3486

References

[1] Eiamsa-ard S, Promvonge P. Enhancement of heat transfer in a tube with regularly-spaced helical tape swirl generators. Solar Energy
2005;78(4):483–94.
[2] Bergles AE. Heat transfer enhancement—the encouragement and accommodation of high heat fluxes. J Heat Transfer (Trans ASME)
1997;119:8–19.
[3] Yilmaz M, Comakli O, Yapici S, Sara ON. Heat transfer and friction characteristics in decaying swirl flow generated by different
radial guide vane swirl generators. Energy Convers Manage 2003;44:283–300.
[4] Yilmaz M, Comakli O, Yapici S. Enhancement of heat transfer by turbulent decaying swirl flow. Energy Convers Manage
1999;40:1365–76.
[5] Dincer I. Thermodynamic, exergy and environmental impact. Energy Source 2000;22(8):723–32.
[6] Dincer I, Sahin AZ. A new model for thermodynamic analysis of a drying process. Int J Heat Mass Transfer 2004;47(4):645–52.
[7] Rosen MA, Dincer I. Effect of varying dead-state properties on energy and exergy analyses of thermal systems. Int J Thermal Sci
2004;43(2):121–33.
[8] Kanoglu M, Carpinlioglu MÖ, Yildirim M. Energy and exergy analyses of an experimental open-cycle desiccant cooling system. Appl
Therm Eng 2004;24:919–32.
[9] Bayrak M, Midilli A, Nurveren K. Energy and exergy analyses of sugar production stages. Int J Energy Res 2003;27:989–1001.
[10] Midilli A, Kucuk H. Energy and exergy analyses of solar drying process of pistachio. Energy 2003;28:539–56.
[11] Syahrul S, Hamdullahpur F, Dincer I. Exergy analysis of fluidized bed drying of moist particles. Exergy, An Int J 2002;2:87–98.
[12] Paykoc E, Yüncü H. Irreversibility in double pipe exchangers. TIBTD (Turkish J) 1986;9(3):35–41.
[13] Durmus A. Heat exchanger and exergy loss in a concentric heat exchanger with snail entrance. Int Commun Heat Mass Transfer
2002;29(3):303–12.
[14] Durmus A. Heat transfer and exergy loss in cut out conical turbulators. Energy Convers Manage 2004;45(5):785–96.
[15] Akpinar EK, Bicer Y. Investigation of heat transfer and exergy loss in a concentric double pipe exchanger equipped with swirl
generators. Int J Thermal Sci 2005;44(6):598–607.
[16] Akpinar EK, Bicer Y, Yildiz C, Pehlivan D. Heat transfer enhancements in a concentric double pipe exchanger equipped with swirl
elements. Int Commun Heat Mass Transfer 2004;31(6):857–68.
[17] Akpinar E, Midilli A, Bicer Y. Single layer drying behavior of potato slices in a convective cyclone dryer and mathematical modeling.
Energy Convers Manage 2003;44:1689–705.
[18] Holman JP. Experimental methods for engineers. New York: McGraw-Hill; 1971. p. 37–52.
[19] Yildiz C, Bicer Y, Pehlivan D. Effect of twisted strips on heat transfer and pressure drop in heat exchangers. Energy Convers Manage
1998;39(3/4):331–6.
[20] Yildiz C, Bicer Y, Pehlivan D. Heat transfer and pressure drop in a heat exchanger with a helical pipe containing inside wires. Energy
Convers Manage 1997;38(6):619–24.
[21] Ogulata RT, Doba F, Yilmaz T. Second-law and experimental analysis of a cross-flow heat exchanger. Heat Transfer Eng
1999;20(2):20–7.
[22] Ogulata RT, Doba F. Experiments and entropy generation minimization analysis of a cross-flow heat exchanger. Int J Heat Mass
Transfer 1998;41(2):373–81.
[23] Dincer I. The role of exergy in energy policy making. Energy Policy 2002;30:137–49.
[24] Bejan A. Advanced engineering thermodynamics. New York: John Wiley and Sons Inc; 1988.
[25] Cengel YA. Heat transfer: a practical approach. 2nd ed. New York: McGraw-Hill; 1998.

You might also like