Sie sind auf Seite 1von 43

Journal Pre-proof

Enhanced heat transfer in rectangular duct with punched winglets

Pongjet Promvonge, Sompol Skullong

PII: S1004-9541(19)30877-8
DOI: https://doi.org/10.1016/j.cjche.2019.09.012
Reference: CJCHE 1572

To appear in: Chinese Journal of Chemical Engineering

Received date: 17 January 2019


Revised date: 8 September 2019
Accepted date: 15 September 2019

Please cite this article as: P. Promvonge and S. Skullong, Enhanced heat transfer in
rectangular duct with punched winglets, Chinese Journal of Chemical Engineering(2019),
https://doi.org/10.1016/j.cjche.2019.09.012

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.

© 2019 Published by Elsevier.


Journal Pre-proof

Enhanced heat transfer in rectangular duct with punched winglets

Pongjet Promvongea, Sompol Skullongb,*


a
Department of Mechanical Engineering, Faculty of Engineering,

King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand


b
Energy Systems Research Group, Department of Mechanical Engineering,

Faculty of Engineering at Sriracha, Kasetsart University Sriracha Campus,

of
199 M.6, Sukhumvit Rd., Sriracha, Chonburi 20230, Thailand

*Corresponding author, E-mail address: sompol@eng.src.ku.ac.th, sfengsps@src.ku.ac.th

ro
Abstract
-p
Thermal performance of a heat exchanger duct with punched winglets (PWs) mounted on the upper duct
re
wall has been examined for Reynolds number (Re) ranging from 4100 to 25,500. In the present experiment,
lP

two types of PWs: punched delta- and elliptical-winglets (P-DW and P-EW) with four punched-hole sizes

were tested at a fixed attack angle, optimal relative pitch and height. Also, data of solid delta- and elliptical-
na

winglets (DW and EW) were included for comparison. The investigation has shown that the P-DW yields

higher thermal-performance enhancement factor () than the P-EW. Although the solid DW and EW with no
ur

punch have the highest heat transfer and friction loss, the PWs yield better  than the solid ones. For PWs, the
Jo

P-DW with smaller hole size has the peak heat transfer and friction loss around 5.7 and 40 times over the

smooth duct, respectively but the optimum  of 2.17 is seen for the one with a certain hole size. The PWs

provides  at about 5–8% above the solid winglets.

Keywords: Heat exchanger; Winglet; Flow resistance; Thermal-performance enhancement factor; Vortex

generator

Nomenclature

A surface area of heat transfer in duct [m2]

1
Journal Pre-proof

AR aspect ratio of duct [=W/H]

b height of winglet [mm]

Cp,a specific heat of air [J/kgK]

d hole diameter on winglet [mm]

Dh hydraulic diameter of duct [mm]

f friction factor [-]

h mean heat transfer coefficient [W/m2K]

H height of duct [mm]

of
I electric current [amp]

ro
ka thermal conductivity of air [W/mK]

Lc
length of test duct [mm]

length of calm section [mm] -p


re
m rate of mass airflow [kg/s]
lP

Nu Nusselt number [=hDh/ka]

Pl longitudinal pitch of winglet [mm]


na

Pt transverse pitch of winglet [mm]

Pr Prandtl number [=Cp/ka]


ur

ΔP pressure drop [Pa]

Reynolds number [=UDh/]


Jo

Re

Q heat transfer rate [W]

T temperature [K]

U average velocity of air [m/s]

V volume flow rate [m3/s]

V electric voltage, [volt]

W width of duct [mm]

Abbreviation

2
Journal Pre-proof

LVG longitudinal vortex generator

SAH solar air heater

VG vortex generator

WVG winglet vortex generator

Greek letters

α angle of attack [°]

ρ density [kg/m3]

of
 kinematic viscosity [m2/s]

ro
 absolute viscosity [kg/ms]

 thermal-performance enhancement factor [-]


-p
re
Subscripts

0 smooth/plain duct
lP

air air
na

b bulk

conv convection
ur

i inlet

o out
Jo

pp pumping power

Re Reynolds number of WVG-enhanced duct

s surface of duct

1. Introduction

The rapid increase in energy utilization from the fossil fuels has resulted in severely environmental

impacts to human being such as global warming and pollution emissions. To reduce those impacts, numerous

researches have therefore been focused on the energy savings as well as the reduction of weight, size and cost

3
Journal Pre-proof

of the thermal systems. A heat exchanger is a device facilitating the convective heat transfer of fluid inside the

ducts/channels and is extensively used in many engineering/industrial applications, such as solar thermal

systems, thermal power plants, air conditioning equipment, chemical processing plants, radiators for

automobiles, and refrigerators. Hence, high performance heat exchanger systems utilized in many industries is

needed to use energy source efficiently to meet the requirements. In general, the flat or smooth surface

employed in the conventional heat exchanger provides poor performance due to the development of thermal

boundary layer. Vortex generating devices, namely, rib/baffles, grooves, coiled-wires, twisted tapes, winglets

etc. are the examples of those devices commonly used for augmenting heat transfer in heating or cooling

of
passages, such as, solar air heater [1–6] and heat exchanger systems [7–19]. A vortex-flow device mounted

ro
on the surface of a heat exchanger is one of the passive techniques and is utilized to induce the

-p
impingement/reattachment flows on the heated wall apart from the disruption of the boundary layer and fast

fluid mixing. In designing the surface of heat exchanger channel/duct, the use of turbulence promoters (the so-
re
called “turbulator”) such as groove [20], rib [21], fin [22,23], baffle [24,25], wing [26] and winglet [27] is
lP

frequently found in the literature for enhancing the heat transfer rate and thermal performance in those

heating/cooling systems.
na

Detailed reviews have been extensively studied by several investigators. Jin et al. [28] conducted an

experiment to study the heat transfer enhancement by pulsating flow in a triangular-grooved duct. Based on
ur

the experimental results, they claimed that the rate of heat transfer was up to 350% compared with the steady
Jo

flow case. Eiamsa-ard and Promvonge [29] investigated numerically the turbulent heat transfer and flow

behaviors in a channel with periodic transverse grooves on the lower channel wall. They revealed that the

grooved channel provided the increase in heat transfer around 158% above the smooth channel and its

thermal-performance enhancement factor was up to 1.33. Tang et al. [30] numerically studied thermal

behaviors and turbulent flow structure in a channel with three groove geometries, namely, P-, V- and W-type

grooves and indicated that the optimum thermal performance was at the case of P-typed groove. Numerical

analyses of heat transfer and flow resistance using a longitudinal vortex generator (LVG) in a dimpled

channel were carried out by Xia et al. [31] who showed that the LVG yielded higher thermal performance

than the dimpled one at a similar flow condition. Lui et al. [32] predicted the flow and thermal characteristics

4
Journal Pre-proof

of a channel with combined dimples and secondary protrusions. They found that the channel with the

combined turbulators yielded higher thermal performance than the conventional dimpled channel. Promvonge

and Thianpong [33] examined thermal performance of a channel with different shaped ribs such as triangular

(isosceles), wedge (delta) and rectangular ribs. They proposed that the isosceles rib with staggered

arrangement yielded the best thermal performance. For a system with artificial roughness wall, there were

several investigations [34–37] on the modified ribbed/baffled surface of a heat exchanger system, including

the developing correlations for friction factor and heat transfer for such a system.

The vortex generating device, namely, winglet vortex generator (WVG) is employed for producing

of
longitudinal vortex flows that can assist to promote the turbulence intensity leading to higher the rate of heat

ro
transfer, albeit with a minimal pressure drop penalty. Recent investigations [38] revealed that the average heat

-p
transfer by using WVGs mounted on the upstream of a flat/smooth surface was increased around 50–60%

over the flat surface alone while the increase in friction loss was at 30–40%. Zhou and Ye [39] reported
re
thermal characteristics of four WVGs: curved trapezoidal, rectangular, trapezoidal, and delta winglets. They
lP

indicated that the curved trapezoidal winglet yielded the best thermohydraulic performance in turbulent region,

while the delta winglet provided the best in laminar and transitional flow region. Zhou and Feng [40]
na

investigated experimentally on heat transfer augmentation by curved and plane winglets with perforated holes

and suggested the merit of the curved winglet with perforated holes in enhancing the heat transfer. Skullong
ur

and Promvonge [41] examined thermal behaviors of a solar air heater (SAH) channel with two delta-winglet

(DW) arrangements: one placed on the lower wall entrance (DW-E) and the other on the absorber (DW-A).
Jo

Their results revealed that the DW-A with b/H = 0.4 yielded the best thermal performance. The influence of

combined ribs and winglets on thermal characteristics for turbulent flow through a SAH channel was

investigated by Promvonge et al. [42]. Chompookham et al. [43] conducted an experiment to examine the

combined wedge-rib and WVG effect on thermal behaviors in a channel and showed that the combined

devices provided thermal-performance enhancement factor around 17–20% higher than a single use of both

devices. Promvonge et al. [44] explored thermal behaviors in a rectangular duct with combined ribs and DWs

and their results displayed that the noticeable heat transfer augmentation was achieved downstream of the DW

because of stronger vortex flow motion.

5
Journal Pre-proof

With regard to the literature cited earlier, there are several types of vortex-flow devices such as

groove/protrusion/dimple/rib/fin/baffle/wing/winglet utilized for increasing thermal performance in heat

exchangers. In scrutiny of all the VGs mentioned above, the WVG provides the best thermal performance.

The present work emphasizes on using the punched-winglets placed on the surface for increasing the

performance of a heat exchanger system in order to give more information on thermal-system improvements.

The key winglet parameters having a significant effect on thermal performance are the winglet pitch and

height; and the attack angle. Previous studies [21, 37] have shown that for WVGs, the attack angle () of 30

is the optimum condition to achieve higher thermal performance, compared to  = 45 and 60 due to lower

of
friction loss while the relative pitches and heights are about PR  1.0 – 1.5 and BR  0.3 – 0.5 for winglets

ro
mounted on a single wall of ducts. Hence, the 30 winglets are initially offered in the present work while their

-p
PR and BR values are optimized. Then, the 30 solid winglet at optimal PR and BR is selected to be punched for

further improvement. To achieve the optimum condition, the effect of different punched-hole sizes of the PWs
re
on the performance is presented in the current work.
lP

2. Detail of experimental program


na

2.1. Detail of WVG

A schematic of PWs placed on the top surface of a heat exchanger duct is shown in Fig. 1. The aluminum
ur

plate was employed as the test duct of air heating surface in the present work. The inner dimensions of the
Jo

aluminum test duct made of a 10-mm thick plate were 200 mm (W)  800 mm (L)  25 mm (H). The overall

duct length was 3500 mm which was divided into three sections: (1) test section of 800 mm, (2) exit section

of 300 mm and (3) calm section of 2000 mm (Lc). Since a fully developed turbulent flow in smooth ducts

takes place at Lc/Dh  10, [45], the calm section with length, Lc=2000 mm or Lc/Dh=45 was used to make sure

that the flow becomes fully developed as it enters the test section. All WVGs formed using 0.5-mm thick

aluminum strips were accomplished by means of wire-EDM (electrical discharge machine) machining and

periodically placed on the test duct using superglue with a fixed attack angle () of 30° and a transverse pitch

spacing, Pt=H, as depicted in Fig. 1. Two types of PWs were selected for comparative tests: punched delta

winglet (P-DW) and punched elliptical winglet (P-EW) as displayed in Fig. 2. The WVGs composed of three

6
Journal Pre-proof

strip sizes: 14, 12, and 10 mm in height (b) equivalent to relative roughness height (BR=b/H= 0.68, 0.48 and

0.40) and relative roughness pitch (PR=Pl/H= 1.0, 1.5, and 2.0) were first introduced to find the optimal

performance. Owing to the highest performance, both the DW and EW with BR = b/H = 0.48 were further

selected for the punch. The punch on the winglet was displayed in terms of relative hole sizes (dR) defined as

a ratio of hole diameter (d) to winglet height. Holes were punched on the selected location (see Fig. 2) of both

the winglets (called the punched winglet, PW) by using high precision drilling machines with different hole

diameters: 0, 3, 5, and 7 mm equivalent to relative hole sizes, dR = d/b = 0, 0.25, 0.417 and 0.583. The

experimental work was carried out for the Re range of 4100 to 25,500 covering the fully turbulent region.

of
Different configurations of the WVGs and parameter ranges used in the investigation are provided in Table 1.

ro
-p
re
lP
na
ur
Jo

Fig. 1. Test section mounted with PWs.

(a) (b)

Fig. 2. Configurations of P-DW (a) and P-EW (b).

7
Journal Pre-proof

Table 1. Configurations of WVGs and range of parameters used in the present work.

WVG Type Attack Height (b) Relative roughness Relative Hole diameter Relative hole
angle ( ) height (BR) roughness pitch (d) size (dR)
configuration
(PR)
Typical DW 30° 10, 12 and 0.4, 0.48 and 0.56 1.0, 1.5 and 2.0 - -
14 mm
WVG
EW 30° 10, 12 and 0.4, 0.48 and 0.56 1.0, 1.5 and 2.0 - -
14 mm
P-DW 30° 12 mm 0.48 1.5 3, 5, and 7 mm 0.25, 0.417
and 0.583
PW
P-EW 30° 12 mm 0.48 1.5 3, 5, and 7 mm 0.25, 0.417

of
and 0.583

ro
2.2. Experimental facility

The experimental procedure of the heat exchanger system is exhibited schematically in Fig. 3. The

-p
system comprises a test section, inverter, blower, orifice plate, variac transformer, plate-type heater, clamp
re
meter (Fluke 376 FC), thermocouples, data acquisition system (Fluke 2680A), personal computer, hot-

wire/vane-typed anemometers (Testo 480), inclined-U-tube/digital manometers and other components. The
lP

test section (top plate) was heated using a plate-typed electrical heater at which its power source obtained

from the variac transformer to maintain a constant wall heat flux condition. Air as the test fluid flowed
na

through the test duct using a 2-kW blower and to obtain the desired volumetric flow rate, the motor speed of

the blower was controlled by an inverter. The inlet air temperature was maintained at about 27 °C. An orifice-
ur

typed flowmeter constructed as per [46] was adopted for airflow measurements. The application of a settling
Jo

tank was to adapt the round duct at the inlet end to become the rectangular duct at the other end, aside from

letting the airflow uniformity. The surface temperature measurements along the axial positions of the test duct

were made using thirty thermocouples (type-T) as illustrated in Fig. 4 whereas four RTD-typed

thermocouples were used for measuring the inlet and the exit air temperatures. Also, a data acquisition system

(Fluke 2680A) was utilized to read all temperature signals from the thermocouples while the measurement of

pressure drop across the test duct for computing friction factor at the isothermal condition was done through a

digital manometer. All fluid properties were obtained at the bulk air temperature by using the average value of

the inlet and exit temperatures. In each test, values of the pressure drop, volume flow rate, surface

temperatures and the bulk air temperature were measured and recorded at a steady state condition. The test

8
Journal Pre-proof

runs were repeated twice and more repeated experiments were needed for the case of unusual data obtained.

The experimental data of the surface and fluid temperatures, pressure drop and volume flow rate were

reproducible within uncertainty ranges.

of
ro
-p
re
lP

Fig. 3. Schematic sketch of experimental system.


na
ur
Jo

Fig. 4. Thermocouple positions on test duct.

9
Journal Pre-proof

3. Experimental data analysis and confirmatory test

3.1. Theoretical analysis

In the present work, experimental results in terms of Reynolds number (Re), heat transfer rate (Q), heat

transfer coefficient (h), Nusselt number (Nu) and friction factor (f) are calculated as follows:

The heat absorbed in the test section equates to the convection heat loss, which is expressed as:

 Cp To  Ti   VI  heat loss


Qair  Qconv  m (1)

in which

  Cp,a  To  Ti 

of
Qair  m (2)

In the thermal equilibrium test, heat (=VI, voltage × ampere) obtained from the sheet-typed electrical heater

ro
was around 3–5% above heat absorbed by air. Therefore, by neglecting natural convection loss to the

-p
surrounding air, the heat transfer rate is employed for calculating the heat transfer coefficients. The heat
re
transfer rate is estimated via


~
Qconv  h  A  Ts  Tb  (3)
lP

in which

Tb  (To  Ti )/2 (4)


na

where Ti  (Ti1  Ti2 )/2 and To  (To1  To2 )/2


ur

The average surface temperature of the duct wall is achieved by averaging the local wall temperatures

 T  T  T   T  T  T   T  T  T 
Jo

~ 1
Ts    1 11 21    2 12 22   ...   10 20 30 
10  3   3   3 

The average local wall temperature at various positions as seen in Fig. 4 can be obtained by

T T T  T T T  T T T 
Ts1   1 11 21 , Ts2   2 12 22 , ... , Ts10   10 20 30 
 3   3   3 

~ 1
Ts   (Ts1  Ts2  ...  Ts10 ) (5)
10
~
where A is the heat transfer surface area, Ts is the mean surface temperature of the local surface temperatures

along the heated wall. The average heat transfer coefficient (h) and Nusselt number (Nu) are evaluated by

10
Journal Pre-proof

m  Cp,a  To  Ti 
h
~
A  Ts  Tb  (6)

The Nu can be obtained by

h  Dh
Nu  (7)
ka

The Reynolds number (Re) is estimated by

  U  Dh
Re  (8)

of
where, Dh is the hydraulic diameter of duct and defined as:

2 W  H

ro
Dh  (9)
(W  H )

f 
ΔP
-p
Using pressure drop (ΔP) and the test duct length (L), the friction factor, f is written via
re
(10)
( ρ  U / 2)  ( L / Dh )
2
lP

To assess the merit of WVGs, their performance is relatively estimated to the smooth duct at similar pumping

power in terms of thermal-performance enhancement factor (). A fruitful comparison between the smooth
na

duct with WVGs and the smooth duct can be made by considering the heat transfer coefficients at equal

pumping power of both cases, since this is relevant to the operation cost. At a similar pumping power [47,48],
ur

it is written as:

VΔP  VΔP
Jo

0 (11)

Terms in Eq. (11) is rewritten in the form of Reynolds number and friction factor.

 fRe    fRe 
3
0
3

where the term with subscript “0” means the plain/smooth duct while the other is for the WVG-enhanced duct.

Dividing both sides by f0,Re, gives

 f  f
Re03  0   Re 3 (12)
 f  f 0,Re
 0,Re 

C1
Inserting the like equation of Blasius for plain ducts, f 0  into Eq. (12),
Re m1

11
Journal Pre-proof

 Re0   C1/Re0 1  f
3 m

   
 Re   C1/Re 1  f 0
m
Re

1
3m1
 Re0  f   Re0   f  3m1
        
Rearranging,  Re   f 0  Re or  Re   f 0  Re (13)

h Nu
By the definition of ,   = (14)
h0 pp
Nu0 pp

Substituting Nu-equation like Dittus-Boelter Eq. for plain ducts, Nu0  C2 Re0 m2 Pr n into Eq. (14),

 Nu   Re0m2 

of
Nu
     m 
(C2 Re Pr )(Re / Re )  Nu0  Re
n m2
0
m2 m2  Re 2



(15)

ro
m2
 Nu   f  3m1
Inserting Eq. (13), yields      
(16)
 Nu0  Re  f 0  Re

-p
in which m1, m2, C1, C2 and n are constants. The values of m2 and m1 can be achieved by considering the f and
re
Nu equations of the present plain duct.
lP

3.2. Experimental uncertainty


na

The uncertainties of results were evaluated based upon the analysis of errors in experimental

measurements [49]. The maximum uncertainties of Reynolds number, Nusselt number and friction factor
ur

were estimated to be ±5.2%, ±7.5% and ±9.2% respectively. The uncertainties of axial velocity and pressure

measurement were estimated to be less than ±5% both, while the uncertainty of temperature measurement was
Jo

about ±0.5%.

3.3. Verification of experimental results

To obtain the reliability, the values of Nu and f of the current flat duct are verified with those of

correlations of Dittus-Boelter, Gnielinski; Blasius, and Petukhov [45] for turbulent duct flows.

Gnielinski correlation,

Nu 
 f/8Re  1000 Pr
1  12.7 f/8 Pr 2/3  1
1/2
(17)

Dittus–Boelter correlation,

12
Journal Pre-proof

Nu  0.023Re 0.8Pr 0.4 (18)

Nu-equation for the current plain duct,

Nu0  0.0327Re0
0.755
Pr 0.4 (19)

Petukhov correlation,

f  0.79 ln Re  1.64
2
(20)

Blasius correlation,

f  0.316Re 0.25 (21)

of
f-equation of for the current plain duct,

ro
0.351
f0  (22)
Re 0.262

-p
The current Nu and f of the flat or smooth duct validated with those of Eqs. (17), (18), (20) and (21) are
re
presented in Fig. 5a and b, respectively. As seen from the figure, the discrepancies between Nu and f values

and those from the correlations, are within 5.7% and 6.8%, respectively. Hence, the deviation between the
lP

correlations and experimental data is less than 7% for both Nu and f.
na
ur
Jo

(a) (b)

Fig. 5. Verification of f (a) and (b) Nu for smooth duct.

13
Journal Pre-proof

Hence,  of Eq. (16) using m2= 0.755 from Eq. (19) and m1 = 0.262 from Eq. (22), becomes
0.755
 Nu   f  2.738
     
(16a)
 Nu0  Re  f 0  Re

4. Experimental results

4.1. Application of WVGs

Effects of delta winglet (DW) and elliptical winglet (EW) at three BR (BR= 0.56, 0.48, and 0.40) and PR

values (PR= 1.0, 1.5, and 2.0) on thermal performance are investigated initially and presented below.

of
4.1.1 Effect of PR

ro
The variation of the average Nu/Nu0 with Re for the WVG is displayed in Fig. 6a. A scrutiny of Fig. 6a

reveals that the DW yields the rate of heat transfer around 4.52–6.23 times above the smooth duct acting alone.

-p
Nu/Nu0 has a decreasing tendency with the increase in Re for all the PR cases while it increases with reducing
re
PR. The DW performs much better than the EW for enhancing the heat transfer. This is caused by higher flow

blockage of the DW disturbing the flow leading to higher turbulence intensity in the duct. Another possible
lP

reason is that the present EW arrangement produces stronger the common-flow-up vortices rather than the

common-flow-down ones [37]. Thus, the heat transfer obtained from the vortex-induced impingement is not
na

much compared to the DW case. At PR = 1.0, 1.5 and 2.0, the mean Nu/Nu0 values for the DW are found to be

7.28%, 8.37% and 9% higher than those for the EW, respectively.
ur
Jo

14
Journal Pre-proof

(a) (b)

Fig. 6. Nu/Nu0 (a) and f/f0 (b) versus PR for WVGs.

In general, heat transfer enhancement is involved with penalty of increased pressure drop resulting in

higher friction loss. Fig. 6b displays the variation of friction factor ratio, f/f0 with different PR and Re values. It

is seen that f/f0 has the declining tendency with decreasing PR. This can be attributed to the fact that

decreasing PR can induce a stronger vortex strength behind the WVG and larger frontal and surface areas and

higher flow turbulence intensity, leading to much high pressure drop. The EW at PR = 1.0 provides the highest

of
f/f0 while the DW at PR = 2.0 has the lowest. At PR = 1.0, f/f0 is in the range of 50.87–77.05 times above the

ro
smooth duct. Although, there are similar trends between the DW and EW, the PR = 1.0 yields the increase in

-p
f/f0 around 16–20% and 26–31% above the PR = 1.5 and 2.0, respectively. The EW gives higher f/f0 than the

DW around 5–10% at similar operating conditions.


re
4.1.2 Effect of BR
lP

Effects of BR on the heat transfer rate and friction loss are also exhibited in Fig. 7a and b, respectively. It

is clearly observed in Fig. 7a that Nu/Nu0 has the increasing trend with the rise of BR. This is because the
na

arrangement of 30 V-shaped winglets can create the common-flow-down vortices [37] leading to a vortex-

induced impingement effect over the most area of the heated wall, resulting in much higher heat transfer. The
ur

DW performs higher than the EW for the heat transfer at a similar condition. At PR=1.5, Nu/Nu0 values of the
Jo

DW are around 5.48–6.15, 5.31–5.96 and 5.1–5.67 while those of the EW are about 5.1–5.77, 4.8–5.52 and

4.5–5.2 for BR=0.56, 0.48 and 0.40, respectively, depending on Re.

15
Journal Pre-proof

of
ro
(a) (b)

-p
Fig. 7. Nu/Nu0 (a) and f/f0 (b) against BR for WVGs.
re
Figure 7b presents the effect of BR on friction factor ratio, f/f0. As expected, both the DW and EW yield a

significant pressure loss in comparison with the smooth duct alone. Reversing trend with the PR cases, f/f0
lP

values display the uptrend with increasing BR and Re while show the opposite tend for increasing PR as
na

mentioned earlier. At PR=1.5, the f/f0 values of the DW are about 43.1–64.5, 36.9–55.5 and 32.7–49.1 times

whereas those for the EW are around 50.5–77.1, 40.9–62.1 and 35.1–52.3 times for BR=0.56, 0.48 and 0.40,
ur

respectively. The f/f0 for BR=0.56 is approximately 15% and 24% higher than that for BR=0.48 and 0.40,

respectively. The EW gives higher f than the DW owing to larger frontal area.
Jo

4.1.3 Thermal-performance enhancement

Figure 8a and b exhibits the influence of PR and BR on  for air entering the artificially roughness duct.

The DW provides higher  than the EW at a similar operating condition.  for the DW and the EW is seen to

be higher than unity for all cases, suggesting the merit of both devices over the smooth duct alone.  values of

the DW are in the range of 1.966 and 2.01, depending on Re, PR and BR and are around 11.6%, in average,

higher than the EW. The maximum  is seen for the DW with PR = 1.5 and BR = 0.48. Because of giving the

highest , both the DW and EW at PR = 1.5 and BR = 0.48 were only selected and modified with the aim to

16
Journal Pre-proof

reduce the pressure drop by punching a hole on the WVGs (called the punched-winglet vortex generators,

PWs) for further investigation.

of
ro
-p
re
(a) (b)
lP

Fig. 8.  versus PR (a) and BR (b) for WVGs.

4.2. Utilization of PWs


na

To increase the performance of the heat exchanger duct, the pressure drop in the duct with WVGs can be

further reduced by punching a hole on the winglets, PWs. The utilization of PWs placed on the duct wall to
ur

increase the performance of the heat exchanger duct is presented. The appropriate punched-hole size has been
Jo

optimized for reducing pressure drop at a compromise between Nu and f. Related parameters of PWs are three

punched hole sizes on both the P-DW and P-EW: relative hole sizes, dR = 0.25, 0.417, and 0.583, and also

both winglets with no hole (dR = 0) are included for comparison. Please keep in mind that the P-DW and the

P-EW with PR = 1.5 and BR = 0.48 are presented only for the hole size optimization.

4.2.1 Effect on heat transfer

The variation of Nu/Nu0 with dR for P-DW and P-EW is depicted in Fig. 9. It is visible that Nu/Nu0

values tends to decrease with the rise of dR and Re values and show a steep decrease at larger dR. Also, the P-

DW still performs much better than the P-EW for augmenting the heat transfer as mentioned before. Nu/Nu0

17
Journal Pre-proof

of the P-DW with dR = 0.417 is around 7–10% over that of the P-EW with similar hole size. The larger hole

size leads to the decrease in Nu/Nu0 and the P-DW with no hole (dR = 0) provides the highest Nu/Nu0. The

main reason for heat transfer reduction is due to the vortex-strength deterioration by the fluid jet from the

winglet hole, while the heat transfer rate decreases a little for small hole size. Scrutinizing Fig. 9 reveals that

the holes on P-DW and P-EW yield the Nu/Nu0 around 4.2–5.8 times and 3.8–5.6 times, respectively while

the increment of the hole size gives the decline in Nu/Nu0. In comparison with the solid winglet (dR = 0), the

average decreases in Nu/Nu0 for P-DW and P-EW are, respectively, about 4.1%, 10.0% and 18.8%; and 4.0%,

9.1% and 19.3% at dR = 0.25, 0.417 and 0.583. The presence of punched-hole on the winglets leads to the heat

of
transfer reduction at less than 10% for the punched-hole with dR  0.417.

ro
-p
re
lP
na
ur
Jo

Fig. 9. Effect of hole size, dR on Nu/Nu0.

4.2.2 Effect on friction loss

Figure 10 presents the variation of f/f0 with dR values for both PWs. It is clearly seen that f/f0 decreases

considerably with increasing dR but with decreasing Re. It is surprising that the maximum and minimum

pressure loss values are apparent in the P-EW and P-DW cases, respectively and this may imply that the flow

field of the P-DW is obstructed/blocked less than that of the P-EW, apart from lower surface area. The

increase in hole size gives rise to the considerable reduction in friction loss, especially for larger hole size.

18
Journal Pre-proof

This is due to lower surface and flow blockage areas from the punch. Surprisingly, f/f0 of the P-DW appears to

be about 20% less than that of the P-EW.

The PWs with the smallest hole size give the maximum f/f0 while the increment of hole sizes results in

considerably low f/f0. At dR = 0.25, 0.417 and 0.583, the increases in f/f0 for the P-DW and P-EW are,

respectively, about 27.25–39.98, 22.35–31.47 and 18.17–24.33; and 30.88–46.07, 25.8–36.74 and 20.53–

27.33 times, depending on Re. The f/f0 of PWs with dR = 0.25 is about 18–21% and 34–40% higher than that

of the ones with dR = 0.417 and 0.583, respectively while the PW with similar BR and PR (BR = 0.48 and PR

=1.5) is about 27–56% lower than that of the solid one (dR = 0). Comparing with solid winglet (dR = 0), at dR

of
= 0.25, 0.417 and 0.583, the mean decreases in f/f0 for the P-DW and P-EW are, respectively, about 27.6%,

ro
42.1% and 54.1%; and 25.3%, 39.3% and 53.5%. Thus, the use of punched-holes on the winglet can decline

-p
the friction loss around 25–50% for the hole size studied.
re
lP
na
ur
Jo

Fig. 10. Effect of hole size, dR on f/f0.

4.2.3 Thermal-performance enhancement

Figure 11a and b depicts the variations of  with Re and dR for various PWs, respectively. In Fig.11a, 

of the PWs is found to decrease with rising Re for all hole sizes. In general, the highest  is seen for the P-

DW with dR = 0.417 while the lowest one is for the P-EW with dR = 0.583 (see Fig.11b). Nevertheless, the

PWs with all holes give  higher than the winglet with no hole.  values for the P-DW with dR = 0.25, 0.417

19
Journal Pre-proof

and 0.583 are, respectively, about 1.66–2.14, 1.68–2.17 and 1.64–2.11 and are around 11–13% higher than the

P-EW with a similar hole size. The maximum  of 2.17 can be achieved for the P-DW with dR = 0.417 at

lower Re and is higher than other cases around 1.1–2.8%. Also, the PW is seen to give the  around 5–8%

above the WVG with no hole.

 in the current work is considerably higher than that in the published artificially roughened ducts

mentioned earlier [4, 21, 33, 34, 37, 44] as can be seen in Fig. 12, since its friction loss can be reduced

significantly from using the punched hole on the WVG. Although Nu/Nu0 for the solid winglet (dR = 0) is

of
seen higher than that for the PW but its f/f0 is also extremely higher due to larger frontal regions. This

phenomenon leads to lower  of the solid winglet than that of the PW as disclosed in Fig. 11. The presence of

ro
the punched WVG can considerably reduce the pressure loss because of decreasing the frontal and surface

-p
areas. Hence, if WVG with high BR is selected for practical use, it should be punched to a suitable hole size to

achieve superior thermal performance.


re
lP
na
ur
Jo

(a) (b)

Fig. 11.  versus Re (a) and dR (b) for PWs.

20
Journal Pre-proof

of
ro
-p
Fig. 12. Comparison of  of the present work with previous artificially roughened ducts.
re
4.3. Development of correlations for PWs
lP

The curve-fitted correlations of the experimental Nu and f values for the heat exchanger duct with PWs

performed by a least-square regression method are a function of Prandtl number (Pr), Re and relative hole size
na

(dR) whereas f is free from Pr. As a result, correlations of Nu and f for the artificially roughened duct have

been formulated and given in Table 2 and their predicted data are verified with the measured as illustrated in
ur

Fig. 13a and b, respectively. As seen in the figure, the deviations of Nu and f are within ±7% each. Thus, the

correlations developed for heat transfer and friction loss are reasonably satisfactory in predicting Nu and f for
Jo

parameters investigated.

Table 2. Summary of empirical correlations for PWs.

P-DW Discrepancy

Nu  0.18Re0.723Pr0.4 (d R ) 0.0955 within 5.2%

f  0.996Re0.091(d R ) 0.2651 within 6.5%

P-EW

Nu  0.182Re0.716Pr0.4 (d R ) 0.0988 within 5.4%

f  1.231Re0.088(d R ) 0.2746 within 6.9%

21
Journal Pre-proof

of
ro
(a) (b)

-p
Fig. 13. Comparison of predicted Nu (a) and f (b) with experimental data for PWs.
re
5. Conclusions
lP

An experimental investigation of a heat exchanger duct fitted with PWs has been performed to explore

the influence of PWs: P-DW and P-EW on its thermal performance. Also, an influence of punched hole sizes
na

of the PWs on thermal performance characteristics is examined. The conclusion can be drawn as follows:

 For experimental results, heat transfer enhancement by WVGs are superior to those by PWs and the
ur

DW performs higher than the EW. Moreover, the PWs cause considerably lower friction loss than

the WVG. The use of PWs can reduce a lot the flow resistance in the heat exchanger duct, especially
Jo

for large punched hole.

 The P-DW yields the highest  of 2.17 or about 7.4% in average above the DW while the P-EW

gives the  around 7.0% higher than the EW. Therefore, the PW is considered to be a promising

device for increasing the performance of a heat exchanger.

 The optimum operating condition is at lower Re for the P-DW and P-EW with dR =0.417 hole. The

maximum  around 2.17 is for the P-DW with dR =0.417 hole while that of 1.92 is for the P-EW

with a similar hole size, indicating the improvement of thermal performance above the smooth duct.

22
Journal Pre-proof

Appendix A

To estimate the experimental-facility reliability, the experimental data uncertainties were examined. Based

on the method as reported in Ref. [49], the data uncertainties for the Reynolds number, friction factor and

Nusselt number can be obtained from the equations below.

Friction factor:

0.5
1  f  
2 2 2
  f   f   f
2
f
   P    L    Dh     Re 
f f  P    L   Dh    Re  

0.5
 P   2  L  2  3D  2  2 Re  2 

of
     
h
   
 P   L   Dh   Re  

ro
where

(P) h
P

h
and
 Re  m   Dh  
Re
   
2

  
2

 m   Dh  

0.5

-p
re
Nusselt number:
lP

0.5
1  
2 2
  
2
Nu
  Nuh   NuDh     Nuk  
Nu Nu  h   Dh   k  

na

0.5
 h  2  D  2 
     h
 

 h   h  
D
ur

where h  q" /(Ts  Tb ) then


Jo

0.5
h 1  h  
2 2 2
  h   h
  q   Ts    Tb  
h h  q   Ts   Tb  

0.5
 q 2  T 2  T 2 
   
s
 
b
 
 q 
   T  T b  T  Tb  
 s s

0.5   V 2  
where q"   C p Tbo  Tbi 
    m

Dh L   R  


23
Journal Pre-proof

References

[1] I.T. Togrul, D. Pehlivan, The performance of a solar air heater with conical concentrator under forced

convection, Int. J. Therm. Sci. 42 (2003) 571–581.

[2] P. Promvonge, Heat transfer and pressure drop in a channel with multiple 60° V-baffles, Int. Commun.

Heat Mass Transf. 37(7) (2010) 835–840.

[3] S. Skullong, P. Promvonge, C. Thianpong, N. Jayranaiwachira, M. Pimsarn, Heat transfer augmentation

in a solar air heater channel with combined winglets and wavy grooves on absorber plate, Appl. Therm.

Eng. 122 (2017) 268–284.

of
[4] S. Skullong, Performance enhancement in a solar air heater duct with inclined ribs mounted on the

ro
absorber, J. Res. Appl. Mech. Eng. 5 (2017), 55–64.

-p
[5] S. Skullong, P. Promthaisong, P. Promvonge, C. Thianpong, M. Pimsarn, Thermal performance in solar

air heater with perforated-winglet-type vortex generator, Sol. Energy 170 (2018) 1101–1117.
re
[6] A. Kumar, A. Layek, Thermo-hydraulic performance of solar air heater having twisted rib over the
lP

absorber plate, Int. J. Therm. Sci. 133 (2018) 181–195.

[7] S. Eiamsa-ard, P. Promvonge, Influence of double-sided delta-wing tape insert with alternate-axes on
na

flow and heat transfer characteristics in a heat exchanger tube, Chin. J. Chem. Eng. 19(3) (2011) 410–423.

[8] S. Eiamsa-ard, N. Koolnapadol, P. Promvonge, Heat transfer behavior in a square duct with tandem wire
ur

coil element insert, Chin. J. Chem. Eng. 20(5) (2012) 863–869.

[9] S. Skullong, S. Kwankaomeng, C. Thianpong, P. Promvonge, Thermal performance of turbulent flow in a


Jo

solar air heater channel with rib-groove turbulators, Int. Commun. Heat Mass Transf. 50 (2014) 34–43.

[10] P. Promvonge, Thermal performance in square-duct heat exchanger with quadruple V-finned twisted

tapes, Appl. Therm. Eng. 91 (2015) 298–307.

[11] W. Noothong, S. Suwannapan, C. Thianpong, P. Promvonge, Enhanced heat transfer in a heat exchanger

square-duct with discrete V-finned tape inserts, Chin. J. Chem. Eng. 23 (2015) 490–498.

[12] S. Chokphoemphun, M. Pimsarn, C. Thianpong, P. Promvonge, Heat transfer augmentation in a circular

tube with winglet vortex generators, Chin. J. Chem. Eng. 23 (2015) 605–614.

24
Journal Pre-proof

[13] S. Chokphoemphun, M. Pimsarn, C. Thianpong, P. Promvonge, Thermal performance of tubular heat

exchanger with multiple twisted-tape inserts, Chin. J. Chem. Eng. 23 (2015) 755–762.

[14] S. Skullong, P. Promvonge, N. Jayranaiwachira, C. Thianpong, Experimental and numerical heat transfer

investigation in a tubular heat exchanger with delta-wing tape inserts, Chem. Eng. Process. Process

Intensif. 109 (2016) 164–177.

[15] M. Awais, A.A. Bhuiyan, Heat and mass transfer for compact heat exchanger (CHXs) design: A state-of-

the-art review, Int. J. Heat Mass Transf. 127 (2018) 359–380.

[16] M. Awais, A.A. Bhuiyan, Heat transfer enhancement using different types of vortex generators (VGs): A

of
review on experimental and numerical activities, Therm. Sci. Eng. Progress, 5 (2018) 524–545.

ro
[17] I.L. Liu, H.L., Y.L. He, Z.T. Chen, Heat transfer and flow characteristics in a circular tube fitted with

-p
rectangular winglet vortex generators, Int. J. Heat Mass Transf. 126 (2018) 989–1006.

[18] H.L. Liu, C.C. Fan, Y.L. He, D.S. Nobes, Heat transfer and flow characteristics in a rectangular channel
re
with combined delta winglet inserts, Int. J. Heat Mass Transf. 134 (2019) 149–165.
lP

[19] M. Awais, A.A. Bhuiyan, Enhancement of thermal and hydraulic performance of compact finned-tube

heat exchanger using vortex generators (VGs): A parametric study, Int. J. Therm. Sci. 140 (2019) 154–
na

166.

[20] C.W. Leung, S. Chen, T.T. Wong, S.D. Probert, Forced convection and pressure drop in a horizontal
ur

triangular-sectional duct with V-grooved (i.e. orthogonal to the mean flow) inner surfaces, Appl. Energy

66 (2000) 199–211.
Jo

[21] S. Skullong, C. Thianpong, P. Promvonge, Effects of rib size and arrangement on forced convective heat

transfer in a solar air heater channel, Heat Mass Transf. 51 (2015) 1475–1485.

[22] P. Promvonge, S. Skullong, S. Kwankaomeng, C. Thiangpong, Heat transfer in square duct fitted

diagonally with angle-finned tape–Part 1: Experimental study, Int. Commun. Heat Mass Transf. 39,

(2012) 617–624.

[23] A. Priyam, P. Chand, Thermal and thermohydraulic performance of wavy finned absorber solar air heater,

Sol. Energy 130 (2016) 250–259.

25
Journal Pre-proof

[24] D. Sahel, H. Ameur, R. Benzeguir, Y. Kamla, Enhancement of heat transfer in a rectangular channel with

perforated baffles, Appl. Therm. Eng. 101 (2016) 156–164.

[25] P. Promvonge, S. Kwankaomeng, Periodic laminar flow and heat transfer in a channel with 45° staggered

V-baffles, Int. Commun. Heat Mass Transf. 37 (2010) 841–849.

[26] S. Skullong, P. Promvonge, C. Thianpong, M. Pimsarn, Thermal performance in solar air heater channel

with combined wavy-groove and perforated-delta wing vortex generators, Appl. Therm. Eng. 100 (2016)

611–620.

[27] M. Fiebig, Embedded vortices in internal flow: heat transfer and pressure loss enhancement, Int. J. Heat

of
and Fluid Flow 16 (1995) 376–388.

ro
[28] D.X. Jin, Y.P. Lee, D.-Y. Lee, Effects of the pulsating flow agitation on the heat transfer in a triangular

-p
grooved channel, Int. J. Heat Mass Transf. 50 (2007) 3062–3071.

[29] S. Eiamsa-ard, P. Promvonge, Numerical study on heat transfer of turbulent channel flow over periodic
re
grooves, Int. Commun. Heat Mass Transf. 35 (2008) 844–852.
lP

[30] X.-Y. Tang, G. Jiang, G. Cao, Parameters study and analysis of turbulent flow and heat transfer

enhancement in narrow channel with discrete grooved structures, Chem. Eng. Res. Des. 93 (2015) 232–
na

250.

[31] H.H. Xia, G.H. Tang, Y. Shi, W.Q. Tao, Simulation of heat transfer enhancement by longitudinal vortex
ur

generators in dimple heat exchangers, Energy 74 (2014) 27–36.

[32] J. Liu, Y. Song, G. Xie, B. Sunden, Numerical modeling flow and heat transfer in dimpled cooling
Jo

channels with secondary hemispherical protrusions, Energy 79 (2015) 1–19.

[33] P. Promvonge, C. Thianpong, Thermal performance assessment of turbulent channel flows over different

shaped ribs, Int. Commun. Heat Mass Transf. 35 (2008) 1327–1334.

[34] C. Thianpong, T. Chompookham, S. Skullong, P. Promvonge, Thermal characterization of turbulent flow

in a channel with isosceles triangular ribs, Int. Commun. Heat Mass Transf. 36 (2009) 712–717.

[35] R. Karwa, Experimental studies of augmented heat transfer and friction in asymmetrically heated

rectangular ducts with ribs on the heated wall in transverse, inclined, V-continuous and V-discrete pattern,

Int. J. Heat Mass Transf. 30 (2) (2003) 241–250.

26
Journal Pre-proof

[36] N.K. Pandey, V.K. Bajpai, Varun, Experimental investigation of heat transfer augmentation using

multiple arcs with gap on absorber plate of solar air heater, Sol. Energy 134 (2016) 314–326.

[37] S. Tamna, S. Skullong, C. Thianpong, P. Promvonge, Heat transfer behaviors in a solar air heater channel

with multiple V-baffle vortex generators, Sol. Energy 110 (2014) 720–735.

[38] M.C. Gentry, A.M. Jacobi, Heat transfer enhancement by delta-wing vortex generators on a flat plate:

vortex interactions with the boundary layer, Exp. Therm. Fluid Sci. 14 (1997) 231–242.

[39] G. Zhou, Q. Ye, Experimental investigations of thermal and flow characteristics of curved trapezoidal

winglet type vortex generators, Appl. Therm. Eng. 37 (2012) 241–248.

of
[40] G. Zhou, Z. Feng, Experimental investigations of heat transfer enhancement by plane and curved winglet

ro
type vortex generators with punched holes, Int. J. Therm. Sci. 78 (2014) 26–35.

-p
[41] S. Skullong, P. Promvonge, Experimental investigation on turbulent convection in solar air heater

channel fitted with delta winglet vortex generator, Chin. J. Chem. Eng. 22(1) (2014) 1–10.
re
[42] P. Promvonge, T. Chompookham, S. Kwankaomeng, C. Thianpong, Enhanced heat transfer in a
lP

triangular ribbed channel with longitudinal vortex generators, Energy Convers. Manage. 51 (2010) 1242–

1249.
na

[43] T. Chompookham, C. Thianpong, S. Kwankaomeng, P. Promvonge, Heat transfer augmentation in a

wedge-ribbed channel using winglet vortex generators, Int. Commun. Heat Mass Transf. 37 (2010) 163–
ur

169.

[44] P. Promvonge, C. Khanoknaiyakarn, S. Kwankaomeng, C. Thianpong, Thermal behavior in solar air


Jo

heater channel fitted with combined rib and delta-winglet, Int. Commun. Heat Mass Transf. 38 (2011)

749–756.

[45] F. Incropera, P.D. Dewitt, Introduction to Heat Transfer, 5th edition John Wiley & Sons Inc., 2006.

[46] ASME, Standard measurement of fluid flow in pipes using orifice, nozzle and venturi. ASME MFC-3M-

1984, United Engineering Center 345 East 47th Street, New York, (1984) 1–56.

[47] P. Promvonge, S. Skullong, Heat transfer in solar receiver heat exchanger with combined punched-V-ribs

and chamfer-V-grooves, Int. J. Heat Mass Transf. 143 (2019) 118486.

[48] R.L. Webb, Principles of Enhanced Heat Transfer, second edition, Taylor & Francis, New York, (2005).

27
Journal Pre-proof

[49] ANSI/ASME, Measurement Uncertainty, PTC 19. 1-1985, Part I, USA, (1986).

of
ro
-p
re
lP
na
ur
Jo

28
Journal Pre-proof

Graphical Abstract:

 A novel PW is utilized for increasing thermal performance in SAH.

 The PW-roughened duct provides much higher heat transfer than the smooth duct.

 Optimum parameters of PWs and their correlations are reported.

of
ro
-p
re
lP
na
ur
Jo

29
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13

Das könnte Ihnen auch gefallen