Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

Journal Pre-proof

Effects of transversely twisted-turbulators on heat transfer and pressure


drop of a channel with uniform wall heat flux

A. Alimoradi (Writing - review and editing) (Funding acquisition), M.


Fatahi (Investigation) (Formal analysis), S. Rehman (Software)
(Project administration), M. Khoshvaght-Aliabadi (Supervision)
(Writing - original draft), S.M. Hassani<ce:contributor-role>Data
Curation)

PII: S0255-2701(20)30488-8
DOI: https://doi.org/10.1016/j.cep.2020.108027
Reference: CEP 108027

To appear in: Chemical Engineering and Processing - Process Intensification

Received Date: 15 February 2020


Revised Date: 30 June 2020
Accepted Date: 2 July 2020

Please cite this article as: Alimoradi A, Fatahi M, Rehman S, Khoshvaght-Aliabadi M, Hassani
SM, Effects of transversely twisted-turbulators on heat transfer and pressure drop of a
channel with uniform wall heat flux, Chemical Engineering and Processing - Process
Intensification (2020), doi: https://doi.org/10.1016/j.cep.2020.108027
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Effects of transversely twisted-turbulators on heat transfer and pressure drop

of a channel with uniform wall heat flux

Alimoradi a,b, M. Fatahi c, S. Rehman d, M. Khoshvaght-Aliabadi c,*, S.M. Hassani e

a
Department for Management of Science and Technology Development, Ton Duc Thang University, Ho Chi Minh

City, Vietnam

of
b
Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City, Vietnam
c
Department of Chemical Engineering, Shahrood Branch, Islamic Azad University, Shahrood, Iran

ro
d
Center for Engineering Research, The Research Institute, King Fahd University of Petroleum and Minerals, Dhahran-

31261, Saudi Arabia


e

-p
Chemical Engineering Department, Faculty of Engineering, Ferdowsi University of Mashhad, Mashhad, Iran
re
*(Corresponding author): E-mail 1: mkhaliabadi@gmail.com E-mail 2: mkhaliabadi@iau-shahrood.ac.ir
lP

Phone: +98 9151811311, Fax: +98 5147244818, Postal address: 36199-43189.


na

Abstract

Air flow and heat transfer through a rectangular channel equipped with transversely twisted-
ur

turbulators are studied experimentally. Three twist-angles and two pitch-ratios are considered to
Jo

explore the effects of design factors. To create a better understanding of possible mechanisms, a

numerical simulation is also carried. The experimental data show that decreasing twist-angle and

pitch-ratio increases both h and ∆p. Among the studied models, the 0° turbulators give the highest

values, followed by the 90° and 180° turbulators come in the second and third, respectively. The

use of 0° turbulators enhances h by about 141.2% at pitch-ratio of 1.875 and 109.4% at pitch-ratio

1
of 3.75, but these enhancements for the models with 90° and 180° turbulators are 133.2%–102.7%

and 54.6%–42.1%, respectively. The numerical results display that the 0° turbulators are more

effective in fluid dispersion towards hot walls and generate stronger swirl flows. However, the

best overall performance is seen for the 90° turbulators. It is also found that at min. Re, the models

with lower pitch-ratio show a better overall performance, while at the max. Re, an inverse outcome

is found. The highest performance index of 1.62 is recorded for the 90° turbulators at the Reynolds

of
number of 1643.

ro
Keywords: Experimental study; Heat transfer enhancement; Transversely twisted-turbulator; Uniform wall heat flux.

1. Introduction

-p
re
Most thermal systems consist of several channels through which a fluid passes as heat

transfer media. Several external devices (or inserts) have been employed as flow turbulators (or
lP

vortex generators) to promote the fluid mixing and enhance the heat transfer inside channels. The

noticeable point is that the heat transfer and pressure drop of the channels are significantly
na

dependent on the configuration and design factors of utilized inserts. In general, both thermal and

hydraulic characteristics must be checked simultaneously to evaluate the overall performance of


ur

these passive enhancement techniques.

Abraham and Vedula [1] conducted experimental efforts on a square converging channel
Jo

with V and W rib turbulators. It was reported that the difference between V and W turbulators was

not noticeable, however, the V shape provided better local thermal performance compared to the

W shape. Numerical simulations were carried out by Fawaz et al. [2] on a square channel equipped

with 45° in-line V shape turbulators. In this study, the influences of blockage-ratio and pitch-ratio

were also examined. Higher values of Nusselt number were obtained for models with larger

2
blockage-ratios and smaller pitch-ratios. Recently, Promvonge and Skullong [3] evaluated V

shaped winglet turbulators in a tubular heat exchanger. The performance index up to 2 was

recorded for this geometry, and empirical correlations were proposed.

In some studies [4–7], helical wire and strip helical screw tape were applied as turbulators

in tubes. The results disclosed that these types of turbulators can be used in curved flow paths with

meaningfully enhanced heat transfer associated with a certain penalty in pumping power. Ibrahim

of
et al. [8] utilized conical ring turbulators inside a circular tube to enhance heat transfer. Three

configurations of convergent, convergent/divergent, and divergent were proposed. The best

ro
performance index of 1.29 was detected for the divergent shape at the Reynolds number of 6000.

Gallegos and Sharma [9] performed experiments on flag shape turbulators in a rectangular channel.

-p
It was seen that the use of this geometry can increase the flow unsteadiness inside the channel and
re
enhance the Nusselt number between 1.34 and 1.62 times. The effects of flag shapes, i.e. straight,

arched, and wavy, were also studied by Hosseinirad et al. [10]. It was reported that the shape and
lP

arrangement of flags can affect both fluid flow and heat transfer, however, the best performance

belonged to the arched shape with the forward arrangement. Khoshvaght-Aliabadi et al. [11]
na

apprised the performance of different turbulators inside a circular tube at the constant temperature

condition. The best efficiency was found for the model equipped with square wings and the twisted
ur

shape comes in the second.

In addition to the channels or tubes, the turbulators were applied in other heat exchange
Jo

devices, such as solar heaters [12–16], plate-fin heat exchangers [17–18], heat sinks [19–22], etc.

For instance, Abdullah et al. [12] evaluated the performance of a double pass solar air heater with

and without turbulators. They attached aluminum cans as turbulators to both the upper and the

lower of the absorber plate. The maximum daily efficiency of 68% was reported for the staggered

3
of these turbulators. Rajaseenivasan et al. [13] studied a single pass solar air heater with circular

and V-type turbulators at six different geometries. It was found that the efficiency of the solar air

heater can enhance with increasing the number of turbulators. However, the enhancement factor

of the system decreased with increasing the Reynolds number. Acır et al. [14,15] examined circular

ring turbulators in a new type solar air heater at different pitch ratios and hole numbers. Similar to

other turbulators, the circular ring can enhance the thermal performance of the solar air heater, so

of
that the enhancement was up to 229%. Further, Bhattacharyya et al. [16] utilized twisted-tape

turbulators in a solar heater at different entrance angles. They reported that the use of single twist

ro
twisted-tape can enhance considerably heat transfer and pressure drop when compared with the

conventional form. It was also found that higher entrance angle can result to better thermal

performance in the solar heater.


-p
re
Samadifar and Toghraie [17] proposed new types of turbulators including a simple

rectangular vortex generator, rectangular trapezius vortex generator, angular rectangular vortex
lP

generator, Wishbone vortex generators, intended vortex generator, and wavy vortex generator for

utilization in plate-fin heat exchangers. However, it was seen that the enhanced heat transfer by
na

the simple rectangular vortex generator was more than other ones. Further enhancements were

found by increasing the height of the vortex generators, while the best angle of attack for the
ur

installation of vortex generator was found to be 45°. Khoshvaght-Aliabadi et al. [18] introduced

delta-winglet on corrugated plate-fin for the first time and tested different compositions of
Jo

water/ethylene glycol mixture as working fluid. It was found that the simultaneous utilization of

vortex generators and corrugations can enhance noticeably the thermal performance of plate-fin

heat exchangers. Also, correlations were proposed for the studied models and working fluids.

4
Li et al. [19] examined the application of delta-winglet turbulators in a heat sink at various

Reynolds numbers, attack angles, turbulators pitches, and turbulators height. The results disclosed

that the thermal resistance of the heat sink decreased as the Reynolds number was increased, and

the attack angle of 30° was a better arrangement for the turbulators. Lu and Zhai et al. [20] analyzed

a heat sink with the combination of turbulators and dimples. The range of heat transfer

enhancement was between 23.4% and 59.8% with a penalty in the friction factor between 22.1%

of
and 54.4%. Chamanroy and Khoshvaght-Aliabadi [21] tested straight and wavy turbulators inside

direct and in-direct channels. It was reported that the thermal performance of both the direct and

ro
the in-direct heat sinks equipped with turbulators was always higher than those of without

turbulators. The enhanced cases showed up to 2.3 times enhancement in the heat transfer

-p
coefficient. In the other study [22], the effects of rectangular turbulators were investigated through
re
a wavy heat sink. In the studied ranges, the heat transfer coefficient and pressure drop for the

enhanced heat sinks were 4–128% and 8–185% higher than those for the original case leading to
lP

a hydrothermal performance index between 0.93 and 1.75. Some studies were also carried out on

water-based hybrid nanofluid and magneto-bio-fluid through porous enclosure and vertical ciliated
na

channel by Shah et al. [23] and Farooq et al. [24], respectively. It was explained that augmenting

the porousness factor can enhance the thermal performance.


ur

Twisted-turbulator is the other shape of turbulators that can effectively be applied. A

review of past studies shows that longitudinal forms of twisted-turbulator were analyzed
Jo

extensively [25–30], and some modifications were also proposed and studied [31–35]. However,

studies on cross shapes, i.e. transverse and normal, of this geometry are very scarce [36]. Hence,

further studies are necessary to apprise the performance of transversely and normally twisted-

turbulators as compared to that of longitudinally twisted-turbulators. This is the main motivation

5
behind the current study, which is divided into experimental and numerical parts. In the

experimental part, heat transfer and pressure drop characteristics are evaluated, while in the

numerical part, the streamlines and velocity/temperature contours are analyzed. In fact, the

application of these turbulators placed in the channel is expected to generate specific flow patterns

and result in efficient heat transfer. The transversely twisted-turbulators are evaluated inside a

channel with uniform wall heat flux. Three different twist-angles of 0°, 90°, and 180° are examined

of
at two pitch-ratios of 1.875 and 3.75. Experiments are carried out over a transient flow regime of

air as heat transfer fluid (Re = 1663, 3286, and 4929). In fact, the transversely twisted-turbulators

ro
can be recognized as a novel passive technique to enhance the heat transfer inside the channels.

Also, the current results can provide useful data for other researchers and designers of different

-p
heat exchangers, such as air conditioning systems, solar air heaters and collectors, etc.
re
2. Turbulators and design factors
lP

In this study, new transversely twisted-turbulators are examined experimentally inside an

aluminum channel having rectangular cross-section with side dimensions of 20 mm × 30 mm,


na

thickness of 1.5 mm, and length of 500 mm. All turbulators are also prepared from the aluminum

sheets with thickness of 0.8 mm. This thickness makes it possible to form the considered
ur

geometries, and also it may avoid additional frictions. Also, the length of sheets is 30 mm, which

is equal to the length of the channel cross-section, but the width of sheets is 10 mm to decrease the
Jo

blockage effects. Note that, it is not possible to generate twisted-turbulators with wider sheets. As

shown in Fig. 1(a), the turbulators are fabricated at three different twist-angles of 0°, 90°, and 180°.

In order to create an insert, i.e. bunch of turbulators, a hole with diameter of 4 mm is generated at

the center of each element. A bolt with the diameter of 4 mm is passed through the holes and some

6
nuts are used to fix the turbulators. The turbulators in each twist-angle are prepared at two pitch-

ratios (distance between two turbulators / hydraulic diameter) of 1.875 and 3.75, which correspond

to element numbers of 11 and 6, respectively. Preliminary studies have shown that using lower

pitch-ratios lead to sharp increases in the pressure drop, and higher pitch-ratios cannot effectively

enhance the heat transfer inside the channel. Two cross-shaped tapes are also employed as supports

at the upstream and downstream sections, which can fix the generated insert at the center part of

of
the channel. The generated inserts are provided in Fig. 1(b).

ro
Fig. 1. (a) Details of fabricated turbulators (b) Assembled form of inserts.

3. Experimental setup and procedure

-p
re
The experimental setup is an open-loop, and the heat transfer fluid is air. The ambient air

is forced to the loop by means of a centrifugal blower. The air velocity is adjusted using an inverter
lP

according to the test purposes and measured using a digital anemometer (AR836, Smart Sensor).

The considered range for the air velocity is between 1 and 3 m/s, which is selected based on the
na

accuracy of the digital anemometer and the power of the centrifugal blower. An air conditioner is

utilized in the room, which can keep the inlet temperature of the air at the preferred value. The air
ur

is passed through an in-line moisture filter trap, then it is directed to the test section. Hence, in

addition to velocity and temperature, the humidity of the air is precisely controlled. As shown in
Jo

Fig. 1(b), the test section is a rectangular channel, which is heated using a flexible electrical wire

heater. A variable digital voltage regulator transformer (3PN1010B, DAM) supplies the required

power, and a multifunction meter (mfm 3430, Ziegler) measures the input power to the electrical

heater. It should be noted that to eliminate the entrance and end effects, two sections are connected

at the upstream and downstream of the test section. These sections are of equal length to the test

7
section, so the whole length of three sections, i.e. entrance, test, and exit, is 150 cm. Four K-type

thermocouples are embedded on the side walls of the test section to evaluate the wall temperature.

They are located at the same axial distance but different circumferential locations. A groove with

depth of 1 mm is generated on each side of the channel and the thermocouples are fixed. An

appropriate paste is applied to protect them from direct contact with the heater. Then, layers of

glass wool with a thickness of about 20 mm are used to insulate the test section from the ambient.

of
This insulation can reduce the heat loses to the ambient by less than 5%. Two RTD sensors are

located at the upstream and downstream of the test section to measure the inlet and outlet

ro
temperatures. Finally, the pressure drop through the test section is evaluated by means of a digital

micro-manometer (MP-100, Kimo). Details of the experimental setup is provided in Fig. 2. Also,

-p
range and accuracy of the applied instruments are tabulated in Table 1. It should be noted that the
re
temperature measuring equipment (K-type thermocouples and RTD sensors) is calibrated by using

a standard thermometer, and the accuracy of other equipment is reported based on the manufacturer
lP

claim.
na

Fig. 2. Experimental test setup.

Table 1. Range and accuracy of applied instruments.


ur

The experiments are performed in the following steps:


Jo

(1) The air conditioner in the room is switched on, and the temperature of the ambient air

is fixed at a suitable value, which signifies the inlet temperature of the air as heat transfer

fluid.

(2) An insert is embedded in the test section, and the experimental setup is turned on.

8
(3) The inlet velocity of the air is adjusted at a desired value, and it is measured by using

the digital anemometer.

(4) The heat source of the electrical heater is switched on and adjusted at the desired value

by means of the variable digital voltage regulator transformer and multifunction meter.

(5) All required data are recorded from the RTD sensors, data logger, and micro-

manometer at the steady-state condition, when all parameters remain constant for about 30

of
min. The experiment for the fixed inlet velocity is finished.

(6) The inlet velocity of the air is changed by using the inverter, and the step (5) is repeated

ro
to acquire the data of this inlet velocity.

(7) After all values of the inlet velocity are tested, the test section is equipped with another

insert and all the steps are repeated.


-p
re
4. Data analysis and uncertainties
lP

In the current study, the difference between the inlet and the outlet temperatures of the air

and the average wall temperature as well as the pressure drop are utilized to evaluate, respectively,
na

the thermal and hydraulic characteristics of different turbulators inside the channel. An equivalent

thermal parameter namely heat transfer coefficient is defined for the thermal analysis, but the
ur

pressure drop is directly applied to apprise the hydraulic analysis.

The rate of heat absorbed by the air is calculated as follows,


Jo

Q  V c p T o u t  T in  (1)

where V is the volumetric flow rate, ρ and cp are, respectively, the density and specific heat of the

air, and Tin and Tout are the inlet and outlet temperatures. The convective heat transfer coefficient

is calculated from the heat transfer rate and the following equation,

9
Q
h  (2)
A t T w  T m 

where At is the total surface area, Tw is the average wall temperature, and Tm is the mean

temperature as follows,

4
1
Tw 
4
T wi
(3)
1

T in  T o u t
Tm  (4)

of
2

The non-dimensional parameter of the Nusselt number is also used to evaluate the thermal

ro
performance of the enhanced models as compared to the smooth case,

-p
hD
Nu  h
(5)
k

where Dh is the hydraulic diameter and k is the thermal conductivity of the air,
re
4Ac L
D h
 (6)
lP

At

in which L is the length of the test section.

The pressure drop through the channel is estimated by using the following formula,
na

 p  p in  p o u t (7)
ur

Also, the friction factor is also considered to evaluate the aerodynamic of the studied

models,
Jo

2D h p
f  (8)
Lv
2

The Reynolds number is calculated based on the inlet velocity and physical properties of

the air as follows,

10
v D
Re  (9)
h

where μ is the dynamic viscosity of the air.

Finally, the performance criterion proposed by Webb [37] is employed as overall

hydrothermal performance index in this study,

 N u enh N u sm o 
  (10)
 f enh 
1 3
f sm o

of
where the subscripts enh and smo refer to the test section with and without the turbulators,

respectively.

ro
The uncertainties inferred from the above equations are caused by measuring of the

-p
parameters and their inaccuracies defined in Table 1. The procedure proposed by Kline and

McClintock [38] is applied to evaluate the corresponding values. Details of the uncertainty analysis
re
calculations are presented in Appendix A. The maximum uncertainties in the obtained values for
lP

Nusselt number, friction factor, and Reynolds number are found to be 1.6%, 0.7%, and 0.9%,

respectively.
na

5. Numerical simulation

In order to create a better insight on possible hydrothermal mechanisms and also reasonable
ur

discussions on the experimental data, a full 3D numerical simulation is also carried out on the
Jo

studied models. As shown in Fig. 3, all sections introduced in Fig. 2, i.e. entrance, test, and exit,

are considered as computational domain. However, some assumptions, such as Newtonian and

incompressible fluid, continuous and steady-state flow without radiation and natural mechanisms,

are considered. Similar to the experimental part, the adiabatic boundary condition is applied on the

walls of the entrance and exit sections, but the constant heat flux boundary condition is utilized on

11
the walls of the test section. The same values of inlet velocity (1, 2, and 3 m/s) and temperature

(298.15) are introduced to have a comparable operating condition between experimental

investigation and numerical simulation. Finally, the outflow boundary condition is applied at the

outlet of the computational domain.

To solve the governing equations and boundary conditions, the finite volume solution

procedure is adopted based a commercial software (CFD package of Fluent V6). According to the

of
iterative technique, the solutions are continued until all variables reach steady regimes, and the

residuals drop under 10−5 for all variables. The semi implicit method for pressure linked equation

ro
(SIMPLE) algorithm is used to discretize the pressure equation, while the second-order upwind

scheme is employed for the momentum and energy equations.

-p
re
Fig. 3. Numerical models.
lP

Due to the complex configuration of the current models, unstructured grids with tetrahedral

scheme is used to discretize the computational domain. The cells become denser around each
na

turbulator where both the velocity and the temperature gradients are significant. The grid

independency study is also performed by employing different sets of cell density. It is found that
ur

for all model, the results are grid-independent when the number of cells is greater than 1160000.

It ensures us that the numerical results are not sensitive to the number of cells. As an example, the
Jo

sensitivity of numerical results to the number of cells for the channel equipped with 0° turbulators

at the pitch-ratio of 1.875 is shown in Fig. 4(a). Furthermore, in order to check the accuracy of

numerical simulations, the results obtained for the heat transfer coefficient and pressure drop are

compared with the corresponding data recorded in the experiments. The comparison shows an

excellent agreement between the numerical results and the experimental data with the maximum

12
deviation of 4.1% for the heat transfer coefficient and 5.7% for the pressure drop. The comparison

between the experimental data and the numerical results are displayed in Fig. 4(b). It implies that

the numerical simulations are reliable, so they can be used to create logical discussions in the

current study.

Fig. 4. (a) Sensitivity of numerical results to number of cells (b) Comparison between experimental data and numerical

results.

of
ro
6. Results and discussion

In order to validate the current experimental procedure, the clear rectangular channel

-p
without installing any turbulators, i.e. smooth case, is tested initially. The corresponding results of

Nusselt number and friction factor are compared with data obtained from some well-known
re
empirical correlations proposed for both laminar and turbulent flow regimes. These correlations
lP

provided from Refs. [39,40] are as follows,

 Laminar
na

Hausen

0 .0 6 6 8  D / L  Re Pr
N u  3 .6 6  2 3
(11)
1  0 .0 4   D / L  R e P r 
ur

Sieder–Tate
Jo

1 3 0 .1 4
 D    
N u  1 .8 6  R e P r 
1 3
    (12)
 L   w 

Shah

1 9 .7 0 2
f  (13)
Re

13
 Turbulent

Dittus–Boelter

N u  0 .0 2 3 R e
0 .8 0 .4
Pr (14)

Gnielinski

N u  0 .0 2 1 4  R e  100  Pr
0 .8 0 .4
(15)

Petukhov

of
2
f   1 .8 2 lo g 1 0 R e  1 .6 4  (16)

ro
Blasius
 0 .2 5
f  0 .3 1 6 4 R e (17)

-p
The deviations of Nusselt number and friction factor between the present results and the

previous data are listed in Table 2. It can be seen that the Nusselt number of this study is in good
re
agreement with both Hausen and Sieder–Tate correlations at the laminar region and Dittus–Boelter
lP

correlation at the turbulent region. The deviation of friction factor are relatively are large at the

laminar region, but it is in reasonable conformity at the turbulent region. However, based on the
na

uncertainty analysis, this deviation is not significant, indicating that the results of the smooth

channel are reliable, and they can be used as baseline for the performance assessment of the
ur

channel equipped with turbulators, i.e. enhanced cases.


Jo

Table 2. Comparison between present results and previous data for smooth channel.

In this section, thermal and hydraulic characteristics of the studied models under the

constant heat flux are provided in forms of heat transfer coefficient and pressure drop versus the

inlet velocity and Nusselt number and friction factor ratios versus the Reynolds number. It should

14
be noted that the quantitative data (h, ∆p, Nu, and f) provided in this section are based on the

experimentations, and the qualitative results (streamlines, velocity vectors, and temperature

contours) are based on the numerical simulations.

Obviously, as shown in Fig. 5(a–b), the channel equipped with different turbulators, i.e.

enhanced cases, gives better thermal performance (higher heat transfer coefficient) than the smooth

case. It can be attributed to complex flow structure and swirl flows produced by the turbulators.

of
For instance, a parts of the air streamlines and tangential velocity vectors between two consecutive

turbulators insight different enhanced models with pitch-ratio of 3.75 is presented in Fig. 6(a–c).

ro
It can be seen that the twist-angle affects both the streamlines and the velocity vectors of the air

flow inside the channel. For the enhanced model equipped with 0° turbulators, the streamlines are

-p
broadcast towards the channel walls where the temperature gradient is high. The velocity vectors
re
show the direction of secondary flows with some swirl flows which are generated near the channel

walls. It seems that these recirculation zones are effective for the heat transfer enhancement
lP

because they can sweep the thermal energy from the thermal boundary more efficiently. It leads

to better thermal performance and higher values of the heat transfer coefficient as shown in Fig.
na

5(a–b). Fig. 6(b) and Fig. 6(c) show that the turbulators with larger twist-angle direct the

streamlines to the core section of the channel and consequently the recirculation zones are
ur

transported to this section. However, it is worth to state that the difference of streamlines between

0° turbulators and 90° turbulators is not noticeable compared to that between 90° turbulators and
Jo

180° turbulators. The velocity and temperature distributions of air flow provided in Fig. 7 confirm

this issue. The same finding can be achieved from the experimental data. At the studied range, the

average deviation of the heat transfer coefficient between 0° turbulators and 90° turbulators is

about 3.4% at the pitch-ratio of 1.875 and 3.1% at the pitch-ratio of 3.75, but the average deviation

15
of the heat transfer coefficient between 90° turbulators and 180° turbulators is about 50.5% at the

pitch-ratio of 1.875 and 42.4% at the pitch-ratio of 3.75. It can also be found that this deviation is

more pronounced at lower pitch-ratio.

Obviously as shown in Fig. 8, all the enhanced models show better thermal performance

than the smooth model. Among enhanced models at the same pitch-ratio, the model with 0°

turbulators gives the highest values of the Nusselt number, and the models with 90° and 180°

of
turbulators come in the second and third, respectively. At the studied range, the use of 0°

turbulators enhances averagely the Nusselt number of the smooth channel by about 141.2% at the

ro
pitch-ratio of 1.875 and 109.4% at the pitch-ratio of 3.75. These enhancements for the models with

90° and 180° turbulators are 133.2%–102.7% and 54.6%–42.1%, respectively. It should be

-p
mentioned that the difference between the enhanced and the smooth models is a function of the
re
inlet velocity (or Reynolds number) and decreases with this operating parameter. For instance, at

the pitch-ratio of 1.875, as the inlet velocity is increased from 1 m/s to 3 m/s, the enhancement of
lP

the Nusselt number diminishes from 172.4% to 119.5% for 0° turbulators, from 164.6% to 111.1%

for 90° turbulators, and from 69.5% to 45.1% for 180° turbulators. However, it is found that the
na

effects of the inlet velocity on the model with 180° turbulators are the highest and the models with

0° and 90° turbulators come in the second and third, respectively. As the inlet velocity is increased
ur

from 1 m/s to 3 m/s, the Nusselt number of the model with 180° turbulators enhances about 1.83

times for the twist-pitch of 1.875 and 1.97 times for the twist-pitch of 3.75. These enhancements
Jo

for the models with 0° and 90° turbulators are 1.82 and 1.67 times and 1.77 and 1.64 times,

respectively. It can be found that the inlet velocity sensitivity decreases as the twist-pitch is

increased. On the other hand, Fig. 5(a–b) demonstrates that the thermal performance is weaken by

increasing the twist-pitch. Actually, at the smaller twist-pitch, the distance between two turbulators

16
is not sufficient to reach a uniform flow with straight streamlines, so the effect of fluid mixing

generated by the former turbulator remains until the next one. This leads to more disruption of the

air flow in the channel thereby better thermal performance.

Fig. 5. Heat transfer coefficient versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.

Fig. 6. Streamlines and velocity vectors of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75 (a) 0° turbulators

(b) 90° turbulators (c) 180° turbulators.

of
Fig. 7. Velocity and temperature contours of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

Fig. 8. Nusselt number ratio versus Reynolds number of different models.

ro
Similarly, the use of turbulators leads to extra pressure drops through the channel, and the

-p
models with the lower twist-angle provide higher values. As shown in Fig. 9(a–b), among
re
enhanced models, the 0° turbulators shows the highest values, and the 90° and 180° turbulators

come in the second and third, respectively. Regarding to the geometry of turbulators, the models
lP

with smaller twist-angle produce more blockage ratio resulting in stronger flow resistance.

Likewise, complex flow structure in these models and superior fluid mixing generated through the
na

flow direction intensify the pressure drop. In other words, the 0° turbulators block and disturb the

flow by their entire frontal surface, but the 90° and 180° turbulators introduce smaller surface for
ur

direct protrusion. Hence, the twisted-turbulators offer lower values of the pressure drop. In order

to create a better insight, the pressure distributions of different models at the inlet velocity of 1 m/s
Jo

and pitch-ratio of 3.75 are displayed in Fig. 10.

As shown in Fig. 11, the use of 0° turbulators augments averagely the friction factor of the

smooth channel by about 6.04 times at the pitch-ratio of 1.875 and 4.11 times at the pitch-ratio of

3.75. These augmentations for the models with 90° and 180° turbulators are 4.79–2.87 times and

17
1.81–1.37 times, respectively. In all the enhanced models, the effects of the pitch-ratio on the

pressure drop (or friction factor) are also obvious. The decrease in the pitch-ratio results in a certain

penalty in the pressure drop, especially in the model with 90° turbulators. For instance, as the

pitch-ratio is decreased from 3.75 to 1.875, the pressure drop of the model with 90° turbulators

augments averagely 66.9%. This value for the models with 0° and 180° turbulators is about 45.5%

and 32.2%, respectively. As previously discussed, the turbulators with the lower twist-angle offer

of
large surface area against the main flow, so the adjacent turbulators display a strong interaction of

flow protrusion leading to higher penalty in the pressure drop. A similar discussion was made by

ro
Nanan et al. [41]. The augmentations in the pressure drop of employing 0°, 90°, and 180°

turbulators with the pitch-ratio of 1.875 as compared to those with the pitch-ratio of 3.75 are

31.2%–62.5%, 64.4%–69.8%, and 30.1%–34.6%, respectively.


-p
re
Fig. 9. Pressure drop versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.
lP

Fig. 10. Pressure distributions of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

Fig. 11. Friction factor ratio versus Reynolds number of different models.
na

It can be concluded that similar to other techniques of enhancing heat transfer, installing

turbulators augments the pressure drop through the channel and subsequently magnifies operating
ur

costs in process. Hence, a proper performance evaluation criterion (PEC) should be utilized to
Jo

detect the best model. In this study, the PEC of Webb [37] introduced in Eq. (10) is applied. This

equation indicates the trade-off between the enhanced heat transfer and the extra blowing power.

The results for both the studied pitch-ratios are provided in Fig. 12. Evidently, the PEC of all the

enhanced models is beyond unity, indicating better overall hydrothermal performance compared

to the smooth model. In other words, the use of the transversely twisted-turbulators is

18
advantageous as a passive enhancement technique and leads to beneficial outcomes from energy

saving point of view.

It can be seen that the PEC decreases by increasing the Reynolds number. It implies that

the use of the proposed technique is more effective at lower flow regimes, and its advantage

decreases at higher flow regimes. In general, the PEC varies between 1.62 and 1.11, depending on

the turbulators geometry and the design factor of pitch-ratio. One can see that in all range of the

of
Reynolds number, the highest values belong to the model with 90° turbulators. It is interesting to

note that at the minimum Reynolds number (laminar regime) and middle Reynolds number

ro
(transient regime), the model with 0° turbulators shows a better performance compared with the

model with 180° turbulators, while at the maximum Reynolds number (turbulent regime), an

-p
inverse result is observed. The other noticeable point is that, (i) at the minimum Reynolds number,
re
the models with the lower pitch-ratio (1.875) propose higher values, (ii) at the middle Reynolds

number, the results of both the pitch-ratios are close, and (iii) at the maximum Reynolds number,
lP

the models with the higher pitch-ratio (3.75) offer higher values. Overall, in the studied cases, the

best model is the channel equipped with 90° turbulators at the pitch-ratio of 3.75. This model offers
na

the maximum value of 1.62 for the PEC at the Reynolds number of 1643.
ur

Fig. 12. Performance evaluation criterion versus Reynolds number.


Jo

Finally, in order to appraise the current technique as compared to other ones, a comparative

study is performed here. The PEC of 90° turbulators at the pitch-ratio of 3.75 as the best model is

compared with that of some other turbulators found in the previous studies. The considered cases

are circular ring and perforated conical ring by Kongkaitpaiboon et al. [42,43], triple twisted-tape

by Bhuiya et al. [44], inclined vortex ring by Promvonge et al. [45], vortex-generator insert by

19
Deshmukh and Vedula [46], and multiple twisted-tape by Chokphoemphun et al. [47]. Note that,

as show in Fig. 13, the comparison is made within the common range of Reynolds number of all

studies. The general trend for all cases implies that the PEC decreases with increasing the Reynolds

number. At the considered range, the 90° turbulators shows better performance as compared to the

circular ring, perforated conical ring, vortex-generator insert, and multiple twisted-tape.

Comparatively, at the lower Reynolds number, the 90° turbulators at the pitch-ratio of 3.75 offers

of
almost the same values with triple twisted-tape and inclined vortex ring, while at the higher

Reynolds number, the former models provide better performance. However, other parameters such

ro
as fabrication cost, installation technique, pressure drop limitations, channel cross-section, etc.

should also be taken into account when selecting a turbulator.

-p
re
Fig. 13. Comparison between current study and some previous literature.
lP

7. Conclusions

Improving thermal performance of heat exchangers is the main concern of researchers in


na

many industries from food to oil, gas, and petrochemical. Among heat transfer enhancement

approaches in channels or tubes, the use of twisted-tapes is one of the most common and effective
ur

method because of its easy installation and low cost. In the current study, the influences of

transversely twisted-turbulators with three different twist-angles and two pitch-ratios are examined
Jo

on the heat transfer and pressured drop of a rectangular channel heated with a uniform heat flux.

As supplementary study, a numerical simulation is also performed to discuss on the fluid flow and

heat transfer mechanisms. The results show that there are noticeable differences in both heat

transfer coefficient and pressure drop of the enhanced models as compared to those of the smooth

model. The highest values of heat transfer coefficient and pressure drop give for the 0° turbulators,

20
and the 90° and 180° turbulators come in the second and third, respectively. These deviations vary

with the pitch-ratio. The use of 0° turbulators enhances averagely the heat transfer coefficient of

the smooth channel by about 141.2% at the pitch-ratio of 1.875 and 109.4% at the pitch-ratio of

3.75. These enhancements for the models with 90° and 180° turbulators are 133.2%–102.7% and

54.6%–42.1%, respectively. Also, the use of 0° turbulators augments averagely the pressure drop

of the smooth channel by about 21.3 times at the pitch-ratio of 1.875 and 15.4 times at the pitch-

of
ratio of 3.75. These augmentations for the models with 90° and 180° turbulators are 18.1–10.5

times and 6.2–4.5 times, respectively. At the studied range, the overall performance index

ro
decreases with the Reynolds number and varies between 1.62 and 1.11, depending on the

turbulators geometry and their pitch-ratio. Finally, at the optimum condition of the pitch-ratio

-p
(3.75) and Reynolds number of 1643, the model equipped with 0° turbulators provides the best
re
hydrothermal performance with the PEC of 1.62. The current results can provide useful guidelines

for engineers and designers whom work on improving thermodynamic features of heat exchangers
lP

with straight channels such as air conditioners, solar heaters, etc.

Credit Author Statement


na

A. Alimoradi: Writing - Review & Editing, Funding acquisition

M. Fatahi: Investigation, Formal analysis,


ur

S. Rehman: Software, Project administration

M. Khoshvaght-Aliabadi: Supervision, Writing - Original Draft


Jo

S.M. Hassani: Data Curation

21
Declaration of competing interest

The authors declared that there is no conflict of interest.

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

of
Acknowledgments

ro
This work is supported by Islamic Azad University of Shahrood Branch and Ferdowsi

University of Mashhad.

-p
re
Appendix A. Uncertainty analysis

The main reason for uncertainties in the calculated parameters (Q, h, Nu, ∆p, f, and Re) is
lP

due to measuring errors of flow rate, temperatures, pressures, and dimensions. The uncertainty

values are estimated based on the accuracy of each instrument and the following equation, which
na

is prescribed by Kline and McClintock [38],

1 2
 M  
2
R 1 R 
   X   (A.1)
ur

R  j 1   X 
j
R
  
 
j
Jo

As mentioned in the text, due to low temperature variations, the thermophysical properties

of the air are assumed constant and their uncertainness are not considered in the current

calculations.

 Heat transfer rate (Q), i.e. Eq. (1),

Q w ate r  V c p T o u t  T in 

22
1 2

1  Q  
2 2 2
Q   Q   Q
  V     T in     T out   (A.2)
Q Q  V    T in    T out  

 Heat transfer coefficient (h), i.e. Eq. (2),

Q
h 
A T w  T b 

of
1 2

1  h  
2 2 2 2
h   h   h   h
  Q    A    T w    T b   (A.3)
h h   Q   A   T w   T b  

ro
 Nusselt number (Nu), i.e. Eq. (5),

-p
hD
Nu  h

k
re
1 2
 N u  
2 2
N u 1   N u
  h    D h   (A.4)
Nu N u   h   D h  

lP

 Friction factor (f), i.e. Eq. (8),

2D h p
f 
na

L v
2

1 2

1  f 
2 2 2 2
f   f   f   f 
  D h     p    L    v   (A.5)
ur

f f  D    p   L   v  
 h

 Reynolds number (Re), i.e. Eq. (9),


Jo

v D
Re 
h

1 2

1   Re  
2 2
 Re   Re
  v    D h   (A.6)
Re R e  v   D h  

Also, the accuracy of each measuring instrument is tabulated in the following table,

23
Name of instrument Measuring parameter Accuracy
Digital anemometer Inlet velocity ±0.1 m/s
RTD sensor Inlet/outlet temperatures ±0.1 °C
K-type thermocouple Wall temperatures ±0.1 °C
Digital micro-manometer Pressure drop ±1 Pa

Nomenclature

Ac frontal flow area (m2)

of
At total heat transfer area (m2)

cp specific heat (J kg-1 K-1)

ro
Dh hydraulic diameter (m)

heat transfer coefficient (W m-2 K-1)

-p
h

k thermal conductivity (W m-1 K-1)


re
L channel length, (m)

Q heat transfer rate (W)


lP

p pressure (Pa)

∆p pressure drop (Pa)


na

T temperature (K)

v velocity (m s-1)
ur

x, y, z coordinates
Jo

Greek symbols

ρ density (kg m-3)

μ dynamic viscosity (Pa s)

η performance index

24
Subscripts

enh enhanced

in inlet

m mean

out outlet

of
smo smooth

w wall

ro
Dimensionless groups

f Friction factor
-p
re
Nu Nusselt number

Pr Prandtl number
lP

Re Reynolds number
na

Acronyms

PEC performance evaluation criterion


ur
Jo

25
References

[1] S. Abraham, R.P. Vedula, Heat transfer and pressure drop measurements in a square cross-

section converging channel with V and W rib turbulators, Experimental Thermal and Fluid

Science 70 (2016) 208–219.

[2] H.E. Fawaz, M.T.S. Badawy, M.F. Abd Rabbo, A. Elfeky, Numerical investigation of fully

of
developed periodic turbulent flow in a square channel fitted with 45° in-line V-baffle

turbulators pointing upstream, Alexandria Engineering Journal 57 (2) (2018) 633–642.

ro
[3] P. Promvonge, S. Skullong, Thermo-hydraulic performance in heat exchanger tube with

V-shaped winglet vortex generator, Applied Thermal Engineering 164 (2020) 114424.

[4]
-p
D. Panahi, K. Zamzamian, Heat transfer enhancement of shell-and-coiled tube heat
re
exchanger utilizing helical wire turbulator, Applied Thermal Engineering 115 (2017) 607–

615.
lP

[5] S. Khorasani, S. Jafarmadar, S. Pourhedayat, M.A.A. Abdollahi, A. Heydarpour,

Experimental investigations on the effect of geometrical properties of helical wire


na

turbulators on thermal performance of a helically coiled tube, Applied Thermal

Engineering 147 (2019) 983–990.


ur

[6] S.R. Chaurasia, R.M. Sarviya, Thermal performance analysis of CuO/water nanofluid flow

in a pipe with single and double strip helical screw tape, Applied Thermal Engineering 166
Jo

(2020) 114631.

[7] S.R. Chaurasia, R.M. Sarviya, Experimental thermal performance analysis of fluid flow in

a heat exchanger pipe with novel double strip helical screw tape inserts for utilization of

26
energy resources, Energy Sources, Part A: Recovery, Utilization, and Environmental

Effects, https://doi.org/10.1080/15567036.2019.1669741.

[8] M.M. Ibrahim, M.A. Essa, N.H. Mostafa, A computational study of heat transfer analysis

for a circular tube with conical ring turbulators, International Journal of Thermal Sciences

137 (2019) 138–160.

[9] R.K.B. Gallegos, R.N. Sharma, Heat transfer performance of flag vortex generators in

of
rectangular channels, International Journal of Thermal Sciences 137 (2019) 26–44.

[10] E. Hosseinirad, M. Khoshvaght-Aliabadi, F. Hormozi, Effects of splitter shape on thermal-

ro
hydraulic characteristics of plate-pin-fin heat sink (PPFHS), International Journal of Heat

and Mass Transfer 143 (2019) 118586.

[11]
-p
M. Khoshvaght-Aliabadi, H. Shabanpour, A. Alizadeh, O. Sartipzadeh, Experimental
re
assessment of different inserts inside straight tubes: Nanofluid as working media, Chemical

Engineering and Processing: Process Intensification 97 (2015) 1–11.


lP

[12] A.S. Abdullah, M.M.A. Al-sood, Z.M. Omara, M.A. Bek, A.E. Kabeel, Performance

evaluation of a new counter flow double pass solar air heater with turbulators, Solar Energy
na

173 (2018) 398–406.

[13] T. Rajaseenivasan, S. Srinivasan, K. Srithar, Comprehensive study on solar air heater with
ur

circular and V-type turbulators attached on absorber plate, Energy 88 (2015) 863–873.

[14] A. Acır, İ. Ata, M.E. Canlı, Investigation of effect of the circular ring turbulators on heat
Jo

transfer augmentation and fluid flow characteristic of solar air heater, Experimental

Thermal and Fluid Science 77 (2016) 45–54.

[15] A. Acır, İ. Ata, A study of heat transfer enhancement in a new solar air heater having

circular type turbulators, Journal of the Energy Institute 89 (4) (2016) 606–616.

27
[16] S. Bhattacharyya, H. Chattopadhyay, A. Haldar, Design of twisted tape turbulator at

different entrance angle for heat transfer enhancement in a solar heater, Beni-Suef

University Journal of Basic and Applied Sciences 7 (1) (2018) 118–126.

[17] M. Samadifar, D. Toghraie, Numerical simulation of heat transfer enhancement in a plate-

fin heat exchanger using a new type of vortex generators, Applied Thermal Engineering

133 (2018) 671–681.

of
[18] M. Khoshvaght-Aliabadi, M. Khoshvaght, P. Rahnama, Thermal-hydraulic characteristics

of plate-fin heat exchangers with corrugated/vortex-generator plate-fin (CVGPF), Applied

ro
Thermal Engineering 98 (2016) 690–701.

[19] H-Y. Li, W-R. Liao, T-Y. Li, Y-Z. Chang, Application of vortex generators to heat transfer

-p
enhancement of a pin-fin heat sink, International Journal of Heat and Mass Transfer 112
re
(2017) 940–949.

[20] G. Lu, X. Zhai, Analysis on heat transfer and pressure drop of a microchannel heat sink
lP

with dimples and vortex generators, International Journal of Thermal Sciences 145 (2019)

105986.
na

[21] Z. Chamanroy, M. Khoshvaght-Aliabadi, Analysis of straight and wavy miniature heat

sinks equipped with straight and wavy pin-fins, International Journal of Thermal Sciences
ur

146 (2019) 106071.

[22] M. Khoshvaght-Aliabadi, S.M. Hassani, S.H. Mazloumi, Enhancement of laminar forced


Jo

convection cooling in wavy heat sink with rectangular ribs and Al2O3/water nanofluids,

Experimental Thermal and Fluid Science 89 (2017) 199–210.

28
[23] Z. Shah, M. Sheikholeslami, P. Kumam, M. Shutaywi, P. Thounthong, CFD Simulation of

Water-Based Hybrid Nanofluid Inside a Porous Enclosure Employing Lorentz Forces,

IEEE Access (2018) 177177–177186.

[24] A.A. Farooq, Z. Shah, E.O. Alzahrani, Heat Transfer Analysis of a Magneto-Bio-Fluid

Transport with Variable Thermal Viscosity Through a Vertical Ciliated Channel,

Symmetry (2019) 11(10), 1240.

of
[25] Q. Xiong, M. Ayani, A.A. Barzinjy, R.N. Dara, A. Shafee, T. Nguyen-Thoi, Modeling of

heat transfer augmentation due to complex-shaped turbulator using nanofluid, Physica A:

ro
Statistical Mechanics and its Applications 540 (2020) 122465.

[26] M. Farnam, M. Khoshvaght-Aliabadi, M.J. Asadollahzadeh, Heat transfer intensification

-p
of agitated U-tube heat exchanger using twisted-tube and twisted-tape as passive
re
techniques, Chemical Engineering and Processing - Process Intensification 133 (2018)

137–147.
lP

[27] M. Sheikholeslami, M. Jafaryar, A. Shafee, Z. Lie, R. Haq, Heat transfer of nanoparticles

employing innovative turbulator considering entropy generation, International Journal of


na

Heat and Mass Transfer 136 (2019) 1233–1240.

[28] B. Sajadi, M. Soleimani, M.A. Akhavan-Behabadi, E. Hadadi, The effect of twisted tape
ur

inserts on heat transfer and pressure drop of R1234yf condensation flow: An experimental

study, International Journal of Heat and Mass Transfer 146 (2020) 118890.
Jo

[29] H. Zhang, X. Jin, S.S. Nunayon, A. Chi-keung Lai, Hydraulic performance and deposition

enhancement of ultrafine particles for in-duct twisted tapes under stationary and rotating

conditions, Applied Thermal Engineering 165 (2020) 114519.

29
[30] C. Qi, M. Liu, J. Tang, Influence of triangle tube structure with twisted tape on the thermo-

hydraulic performance of nanofluids in heat-exchange system based on thermal and exergy

efficiency, Energy Conversion and Management 192 (2019) 243–268.

[31] M.M.K. Bhuiya, M.M.Roshid, M.M.M. Talukder, M.G. Rasul, P. Das, Influence of

perforated triple twisted tape on thermal performance characteristics of a tube heat

exchanger, Applied Thermal Engineering 167 (2020) 114769.

of
[32] M. Bahiraei, N. Mazaheri, M.S. Mohammadi, H. Moayedi, Thermal performance of a new

nanofluid containing biologically functionalized graphene nanoplatelets inside tubes

ro
equipped with rotating coaxial double-twisted tapes, International Communications in Heat

and Mass Transfer 108 (2019) 104305.

[33]
-p
M.E. Nakhchi, J.A. Esfahani, Performance intensification of turbulent flow through heat
re
exchanger tube using double V-cut twisted tape inserts, Chemical Engineering and

Processing - Process Intensification 141 (2019) 107533.


lP

[34] M. Khoshvaght-Aliabadi, M. Eskandari, Influence of twist length variations on thermal–

hydraulic specifications of twisted-tape inserts in presence of Cu–water nanofluid,


na

Experimental Thermal and Fluid Science 61 (2015) 230–240.

[35] M. Khoshvaght-Aliabadi, S. Davoudi, M.H. Dibaei, Performance of agitated-vessel U tube


ur

heat exchanger using spiky twisted tapes and water based metallic nanofluids, Chemical

Engineering Research and Design 133 (2018) 26–39.


Jo

[36] S. Rashidi, M. Akbarzadeh, N. Karimi, R. Masoodi, Combined effects of nanofluid and

transverse twisted-baffles on the flow structures, heat transfer and irreversibilities inside a

square duct – A numerical study, Applied Thermal Engineering 130 (2018) 135–148.

30
[37] R.L. Webb, Performance evaluation criteria for use of enhanced heat transfer surfaces in

heat exchanger design, International Journal of Heat and Mass Transfer 24 (1981) 715–

726.

[38] S.J. Kline, F.A. McClintock, Describing uncertainties in single-sample experiments,

Mechanical Engineering 75(1) (1953) 3–8.

[39] J.P. Holman, Heat Transfer, (ninth ed.), McGraw-Hill, New York (2002).

of
[40] S. Kakac, R.K. Shah, A.E. Bergles (Eds.), Low Reynolds Number Flow Heat Exchanger,

Hemisphere, New York (1983), pp. 75-108.

ro
[41] K. Nanan, C. Thianpong, M. Pimsarn, V. Chuwattanakul, S. Eiamsa-ard, Flow and thermal

mechanisms in a heat exchanger tube inserted with twisted cross-baffle turbulators,

Applied Thermal Engineering 114 (2017) 130–147.


-p
re
[42] V. Kongkaitpaiboon, K. Nanan, S. Eiamsa-ard, Experimental investigation of convective

heat transfer and pressure loss in a round tube fitted with circular-ring turbulators,
lP

International Communications in Heat and Mass Transfer 37 (5) (2010) 568–574.

[43] V. Kongkaitpaiboon, K. Nanan, S. Eiamsa-ard, Experimental investigation of heat transfer


na

and turbulent flow friction in a tube fitted with perforated conical-rings, International

Communications in Heat and Mass Transfer 37 (5) (2010) 560–567.


ur

[44] M.M.K. Bhuiya, M.S.U. Chowdhury, M. Shahabuddin, M. Saha, L.A. Memon, Thermal

characteristics in a heat exchanger tube fitted with triple twisted tape inserts, International
Jo

Communications in Heat and Mass Transfer 48 (2013) 124–132.

[45] P. Promvonge, N. Koolnapadol, M. Pimsarn, C. Thianpong, Thermal performance

enhancement in a heat exchanger tube fitted with inclined vortex rings, Applied Thermal

Engineering 62 (1) (2014) 285–292.

31
[46] P.W. Deshmukh, R.P. Vedula, Heat transfer and friction factor characteristics of turbulent

flow through a circular tube fitted with vortex generator inserts, International Journal of

Heat and Mass Transfer 79 (2014) 551–560.

[47] S. Chokphoemphun, M. Pimsarn, C. Thianpong, P. Promvonge, Thermal performance of

tubular heat exchanger with multiple twisted-tape inserts, Chinese Journal of Chemical

Engineering 23 (5) (2015) 755–762.

of
Caption of tables

ro
Table 1. Range and accuracy of applied instruments.

Table 2. Comparison between present results and previous data for smooth channel.

Caption of figures -p
re
Fig. 1. (a) Details of fabricated turbulators (b) Assembled form of inserts.
lP

Fig. 2. Experimental test setup.


na

Fig. 3. Numerical models.

Fig. 4. (a) Sensitivity of numerical results to number of cells (b) Comparison between experimental data and numerical
ur

results.
Jo

Fig. 5. Heat transfer coefficient versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.

Fig. 6. Streamlines and velocity vectors of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75 (a) 0° turbulators

(b) 90° turbulators (c) 180° turbulators.

Fig. 7. Velocity and temperature contours of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

Fig. 8. Nusselt number ratio versus Reynolds number of different models.

32
Fig. 9. Pressure drop versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.

Fig. 10. Pressure distributions of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

Fig. 11. Friction factor ratio versus Reynolds number of different models.

Fig. 12. Performance evaluation criterion versus Reynolds number.

Fig. 13. Comparison between current study and some previous literature.

of
0° turbulators

ro
0.8 mm
30 mm
4 mm
10 mm

-p
re
90° turbulators
lP
na

180° turbulators
ur
Jo

(a)

33
of
ro
(b)

-p
Fig. 1. (a) Details of fabricated turbulators (b) Assembled forms of inserts.
re
lP
na
ur
Jo

34
A part of assembled test section Cross-section of test section

Location of wall thermocouples

K-type thermocouple

Test section

Data logger Digital micro-manometer

of
Inverter

ro
RTD sensor RTD sensor

-p
Electrical wire heater

Entrance section Test section Exit section


re
Centrifugal blower Digital anemometer

Moisture filter trap


lP

Multifunction meter

Variable digital voltage


regulator transformer
na

Fig. 2. Experimental test setup.


ur
Jo

35
Air
inlet En
Parts of generated grids
tra
nc
e se
cti
on
(A
dia
ba
tic
)

Te
s ts
ec
tio
n(
Co
ns
ta nt
he
a t fl
ux

of
)

Ex
it s

ro
ec
tio
n(
Ad
iab
ati
c)

0° turbulators 90° turbulators


-p 180° turbulators
Air
outlet
re
lP
na

Fig. 3. Numerical models.


ur
Jo

36
50
Set 1
45
Set 2
Set 3 Set 4 Set 5
40
Parameter (h or ∆p)

Pressure drop
35
Absolute Deviation (%) h ∆p
30 Set 1 and Set 2 8.41 8.86
Set 2 and Set 3 7.33 3.64
25 Set 3 and Set 4 1.45 1.51
Set 4 and Set 5 0.10 0.25

of
20 Heat transfer coefficient
Set 3 Set 4 Set 5
Set 2
15 Set 1

ro
10
0 200000 400000 600000 800000 1000000 1200000 1400000 1600000

-p
Number of cells
re
(a)

50 250
Heat transfer coefficient Pressure drop
lP

40 200
x=y
Numerical results

Numerical results

30 x=y 150
na

20 100
ur

10 50

0 0
0 10 20 30 40 50 0 50 100 150 200 250
Jo

Experimental data Experimental data

(b)

Fig. 4. (a) Sensitivity of numerical results to number of cells (b) Comparison between experimental data and

numerical results.

37
40

Heat transfer coefficient [W/m2.K]


35

30

25
Smooth
20
0 degree
15
90 degree
10 180 degree

of
0
1 2 3
Inlet velocity [m/s]

ro
(a)

-p
40
Heat transfer coefficient [W/m2.K]

35
re
30

25
Smooth
20
lP

0 degree
15
90 degree
10 180 degree

5
na

0
1 2 3
Inlet velocity [m/s]
ur

(b)
Jo

Fig. 5. Heat transfer coefficient versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.

38
0° turbulators

Plane A

Plane B

Plane C

of
ro
Plane A

-p
Plane B
re
Plane C
lP
na
ur

(a)
Jo

39
90° turbulators

Plane A

Plane B

Plane C

of
ro
Plane A

-p
Plane B
re
Plane C
lP
na
ur

(b)
Jo

40
180° turbulators

Plane A

Plane B

Plane C

of
ro
Plane A

-p
Plane B
re
Plane C
lP
na
ur

(c)
Jo

Fig. 6. Streamlines and velocity vectors of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75 (a) 0° turbulators

(b) 90° turbulators (c) 180° turbulators.

41
Velocity [m/s] Temperature [K]

0.5

0.75

1.25

298

302

304

306

308
0.25

1.5

1.75

300

310

312

314
0° turbulators

of
ro
90° turbulators

-p
re
lP
na
180° turbulators

ur
Jo

Fig. 7. Velocity and temperature contours of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

42
3
Pitch-ratio = 1.875
Nusselt number ratio [] 0 degree
2.5
90 degree

180 degree
2
Pitch-ratio = 3.75
0 degree

1.5 90 degree

of
180 degree
1

ro
1000 2000 3000 4000 5000
Reynolds number []

-p
Fig. 8. Nusselt number ratio versus Reynolds number of different models.
re
lP
na
ur
Jo

43
250

200

Pressure drop [Pa] 150


Smooth
0 degree
100
90 degree
180 degree
50

of
0
1 2 3
Inlet velocity [m/s]

ro
(a)

-p
250

200
re
Pressure drop [Pa]

150
Smooth
lP

0 degree
100
90 degree
180 degree
50
na

0
1 2 3
Inlet velocity [m/s]
ur

(b)
Jo

Fig. 9. Pressure drop versus inlet velocity (a) pitch-ratio of 1.875 (b) pitch-ratio of 3.75.

44
18
18
0° turbulators
15 15

12 12

Pressure [Pa] 9
9

6 6

3 3

0 0
0 5 10 15 20 25 30 35 40 45 50

of
-3
Channel length [m]

ro
12
12
90° turbulators
10 10

-p
8 8
Pressure [Pa]

6 6
re
4 4

2 2
lP

0 0
0 5 10 15 20 25 30 35 40 45 50
-2
Channel length [m]
na

6 6

180° turbulators
5 5

4 4
ur Pressure [Pa]

3 3

2 2
Jo

1 1

0 0
0 5 10 15 20 25 30 35 40 45 50
-1
Channel length [m]

Fig. 10. Pressure distributions of air flow at inlet velocity of 1 m/s and pitch-ratio of 3.75.

45
8
Pitch-ratio = 1.875
Friction factor ratio [] 7 0 degree

6
90 degree
5
180 degree
4 Pitch-ratio = 3.75
0 degree
3
90 degree

of
2
180 degree
1

ro
1000 2000 3000 4000 5000
Reynolds number []

-p
Fig. 11. Friction factor ratio versus Reynolds number of different models.
re
lP
na
ur
Jo

46
1.7

1.6
Pitch-ratio of 1.875
1.5
0 degree
1.4 90 degree
PEC []

1.3 180 degree


Pitch-ratio of 3.75
1.2
0 degree
1.1 90 degree
180 degree

of
1
1643 3286 4929
Reynolds number []

ro
Fig. 12. Performance evaluation criterion versus Reynolds number.

-p
re
lP
na
ur
Jo

47
1.7 Current study

1.5 Circular ring

1.3 Perforated conical ring


PEC []
1.1 Triple twisted tape

Inclined vortex ring


0.9
Vortex generator insert
0.7
Multiple twisted tape
0.5

of
2000 4000 6000
Reynolds number

ro
Fig. 13. Comparison between current study and some previous literature.

-p
re
lP
na
ur
Jo

48
Table 1. Range and accuracy of applied instruments.
Instrument Parameter Range Accuracy
Centrifugal blower – – –
Inverter – – –
Digital anemometer Inlet velocity 1–3 m/s 0.1 m/s *
Moisture filter trap – – –
Electrical wire heater – – –
Variable digital voltage regulator transformer – 57 V 1V*
Multifunction meter Power 100 W 0.1 W *
RTD sensor Inlet/outlet temperatures 21.9 –37.6 °C 0.1 °C **
K-type thermocouple Wall temperatures 23.5 –51.1 °C 0.1 °C **
Data logger – – –
Digital micro-manometer Pressure drop 8 –247 Pa 1 Pa *
*
Based on manufacturer claim
**
Based on calibration

of
ro
Table 2. Comparison between present results and previous data for
smooth channel.
Correlation Laminar Turbulent

-p
Hausen 8.7% –
Sieder–Tate –7.9% –
Shah 18.2% –

re
Dittus–Boelter 1.4%
Gnielinski – 22.6%
Petukhov – 10.3%
Blasius – 11.1%
lP
na
ur
Jo

49

You might also like