Hindawi Publishing Corporation Advances in Mechanical Engineering Volume 2014, Article ID 631081, 9 pages http://dx.doi.org/10.1155/2014/631081
Research Article Experimental and Numerical Investigations of Heat Transfer Characteristics for Impinging Swirl Flow Sami D. Salman,1,2 Abdul Amir H. Kadhum,1 Mohd S. Takriff,1 and Abu Bakar Mohamad1 1
Department of Chemical and Process Engineering, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, 43600 Bangi, Selangor, Malaysia 2 Biochemical Engineering Department, Al-Khwarizmi College of Engineering, University of Baghdad, Baghdad 47024, Iraq Correspondence should be addressed to Sami D. Salman;
[email protected] Received 28 February 2014; Revised 14 April 2014; Accepted 19 April 2014; Published 20 May 2014 Academic Editor: Jinjia Wei Copyright Β© 2014 Sami D. Salman et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. This paper reports experimental and computational fluid dynamics (CFD) modelling studies to investigate the effect of the swirl intensity on the heat transfer characteristics of conventional and swirl impingement air jets at a constant nozzle-to-plate distance (πΏ = 2π·). The experiments were performed using classical twisted tape inserts in a nozzle jet with three twist ratios (π¦ = 2.93, 3.91, and 4.89) and Reynolds numbers that varied from 4000 to 16000. The results indicate that the radial uniformity of Nusselt number (Nu) of swirl impingement air jets (SIJ) depended on the values of the swirl intensity and the air Reynolds number. The results also revealed that the SIJ that was fitted with an insert of π¦ = 4.89, which corresponds to the swirl number Sw = 0.671, provided much more uniform local heat transfer distribution on the surface. The CFD-predicted results help to explain the experimental measurements in terms of the turbulence intensity. Furthermore, the predicted and measured local Nusselt numbers were consistent with each other.
1. Introduction Heat exchangers with convective heat transfer are widely used in many engineering applications. The augmentation of heat transfer in all types of thermotechnical apparatus significantly saves the primary energy and reduces the size and weight of the heat exchangers. In general, heat transfer augmentation techniques are classified as passive and active. The passive methods require special surface geometries or swirl/vortex flow devices, whereas the active techniques require external forces, such as an electric field, surface vibration, or jet impingement. Jet impingement is one of the active techniques that is widely used in several engineering applications because of its promising characteristics, such as high heat transfer rate and low cost. The applications include heating steel plates, cooling turbine blades and electronic components, tempering glass, and drying papers and food products. A number of experimental and numerical investigations have been attempted regarding the effects of single and multiple jets and impinged surface geometries on the mass and the heat transfer [1β10]. Introducing a swirl
motion into an impinging jet is considered an effective means to expand the heat transfer area compared to the conventional jet. Bakirci and Bilen [11] investigated the temperature distribution and the heat transfer rate on impingement. The surface was maintained at a constant wall temperature boundary condition for the swirling (SIJ), multichannel (MCIJ), and conventional impinging jets (CIJ) using the liquid-crystal technique. Swirl generators with swirl angles of 0β , 22.5β , 41β , and 50β were used to change the direction and strength of the swirl in the air flow that exited the jet nozzle. According to the results, the radial uniformity of the local and average Nusselt numbers improved when the swirl angle and/or jet spacing was increased. Ianiro and Cardone [12] performed an experimental analysis of the heat transfer distribution on a flat plate under arrays of swirling impinging jets using infrared thermography and with heated thin-foil sensor for the measurement. Helical inserts were used to generate five swirl number (Sw) values (0, 0.2, 0.4, 0.6, and 0.8) for five nozzle-to-plate distances (π§) (2, 4, 6, 8, and 10 π·). The results indicated that the multichannel jet enhanced heat transfer unlike the common circular impinging jet. However, it was
2 reported that the swirl motion decreased the heat transfer rate but improved the radial uniformity of the heat transfer. Nanan et al. [13] empirically investigated the convective heat transfer that was associated with swirling impinging jets on an impinged plate using twisted tape inserts. The effect of the twist ratio (π¦/π), jet-to-plate spacing (πΏ/π·), and jet Reynolds number (Re) on the local/average Nusselt number on the impingement plate was reported. The results showed that the maximum local/average Nusselt number on the impingement plate was achieved at the largest twist ratio, the smallest jet-to-plate spacing, and the highest jet Reynolds number. Nuntadusit et al. [14] studied the effect of the swirl number on the flow and heat transfer characteristics in an impinging jet on an impinged surface using straight and twisted tape inserts. They visualized the flow pattern and the temperature distribution and evaluated the heat transfer rate of the swirling impinging jet using oil film and thermochromic liquid-crystal techniques. Nuntadusit et al. [15] also reported the flow and the heat transfer in multiple swirling impinging jets with 3 Γ 3 in-line arrangement on a flat, heated surface using the same techniques for the flow and heat measurements. The outcome signified that the multiple swirling impinging jets offered a higher heat transfer rate on impinged surfaces than the multiple conventional impinging jets for all jet-to-jet distances. In the present work, the effect of swirl jets on a heated impinged surface plate using classic twist tape inserts with twist ratios π¦ = 2.93, 3.91, and 4.89 is reported. The performance of the insert is evaluated in terms of the measured increase in local Nusselt number values and the uniformity of their radial distributions on the impinged surface. The operational performance of these inserts is experimentally and theoretically compared with that obtained for conventional jets.
2. Experimental Work and Uncertainty Analysis The experimental setup schematic is illustrated in Figure 1. An air jet is supplied to the system through a pressure regulator, using a compressor, condenser, air filter, and flow meter. The air filter was installed between the air tank and the flow meter to clean the air from oil and other impurities. The air was directed toward the impingement surface passing through a rubber hose and a circular, 1 m long aluminum pipe with 21 mm inner diameter and 3.5 mm wall thickness. Because the ratio of the pipe length to the nozzle diameter was 47, a fully developed flow was used at the pipe nozzle entrance. The applied pipe nozzle was 21 mm in inner diameter and 300 mm long. Classical twist tapes with different twist ratios (π¦ = 2.93, 3.91, 4.89) served as inserts to generate the swirl motion; these inserts are made from aluminum strips with width of π€ = 20 mm and thickness of πΏ = 0.8 mm. The twist ratio was obtained by uniformly winding the strip for over 300 mm. The twist ratio βπ¦β is defined as the ratio of the length of one full twist (360β ) to the width of the twist [16]. Figure 2 shows an illustration and the details of the twisted tapes that were used in this study.
Advances in Mechanical Engineering 1
2 3
4
5
6
8
7
14 10 11
9
12
13
Figure 1: Schematic diagram of the experimental setup. (1) Air compressor; (2) air tank; (3) open-close valve; (4) pressure regulator; (5) heat exchanger; (6) air filter; (7) flow meter; (8) jet pipe; (9) thermocouple; (10) heat flux sensor; (11) thermocouple; (12) AC power supply; (13) data logger; (14) impingement plate.
Table 1: Twist ratio with modified swirl number. Twist ratio (π¦) 2.93 3.91 4.89
Number of pitches (π) 10.238 7.672 6.109
Swirl number (Sw) 1.125 0.843 0.671
The swirl intensity for a classical twisted tape is varied by changing the number of pitches per twisted tape length. The swirl intensity is represented by the swirl number, which is defined as follows: Sw =
πΊπ , π
πΊπ
(1)
where πΊπ is the axial flux of the azimuthal momentum, πΊπ is the axial flux of the axial momentum, and π
is the radius of the pipe nozzle. In this study, the effect of the boundary layer in the pipe nozzle is ignored; the axial velocity and the angular velocity are assumed to be constant along the cross section of the nozzle exit. Thus, πΊπ and πΊπ can be expressed as π
πΊπ = 2π β« πππ2 π ππ = ππππ2 π
2 , 0
ππππ ππ
4 πΊπ = 2π β« πππ ππ π ππ = , 2 0 π
(2)
2
where ππ , ππ , and π€ are the azimuthal velocity, the axial velocity, and the angular velocity, respectively. The angular velocity can be calculated from the twisted pitch number (π) and the length of the twisted tape (π = 300 mm) as follows: π=
2ππππ . β
(3)
From (1)β(3), a modified swirl number is obtained and defined as Sw =
πππ
. β
(4)
The modified swirl numbers (Sw) with different twist ratios are provided in Table 1.
Advances in Mechanical Engineering
3
H
y = H/w = 2.93
Pitch w
y = H/w = 3.91
Tape
πΏ
Direction of flow
y = H/w = 4.89
D
Pipe nozzle
(a)
(b)
Figure 2: (a) Twisted tapes at various twist ratios and (b) pipe nozzle with twisted tape.
The impinged plate was 300 mm wide and 300 mm long. The electric heater sheet was made from stainless steel with a width and a length of 300 mm and 300 mm, respectively. The heater sheet was imbedded between the impinged plate and a thick 300 Γ 300 mm mica plate to minimize and disregard the lateral heat conduction. The back side of the impinged plate was embedded with brass plated with a chrome-type πΎ thermocouple using a thermally well-conductive paste as an adhesive layer to diminish the disturbance on the thermocoupleβs signal. AC electric current was conducted to the electric heater sheet via uniform heat generation using a power supply unit with a TC4S temperature control system. The temperature and the heat flux of the impinged plate were examined overall using HFX-4 Thin-Film Heat Flux Sensors in a GL220 Data Logger. The surface temperature at different locations on the impingement plate was measured and compared with the temperature on the back side, which showed that the difference was less than 0.5β C under the tested conditions. The experimental data on the heat transfer rate are expressed in terms of the local Nusselt number and the local convective heat transfer coefficient at the impingement plate as follows: Nu =
βπ· , πΎ
πβ , ππ€ β ππ
(6)
where πβ is the heat flux that was measured using the HFX4 Thin-Film Heat Flux Sensors, ππ€ is the impingement wall temperature that was measured using the HFX-4 Thin-Film Heat Flux Sensors, and ππ is the air temperature at the nozzle exit, which was measured using a πΎ-type thermocouple. In addition, the Reynolds number is obtained according to the following equation: Re =
ππ’π· , π
2 ππ
π€π₯π ) ] π€π
= [β ( ππ₯π
1/2
,
(8)
where π€π
is the uncertainty of result π
and π€π₯π is the uncertainty of the independent variables π₯π upon which π
= π
(π₯π ). For the Nusselt number, the π₯π included the heat flux π, the air jet temperature (ππ ), the surface temperature of the impingement plate (ππ ), and the jet nozzle diameter (π·). The final uncertainty equation of the local Nusselt number was derived from (8) as follows:
(5)
where π· is the nozzle diameter, β is the heat transfer coefficient, and π is the conductivity of air based on the reference temperature of 25β C: β=
where π’ is the average velocity at the exit of the nozzle, π· is the nozzle diameter, π is the air density based on the reference pressure and temperature of 1 atm and 25β C, respectively, and π is the air viscosity based on the reference pressure of 1 atm and the reference temperature of 25β C. A standard uncertainty analysis was conducted for each measurement to determine the maximum expected deviation from the experimental results of the local Nusselt number using the uncertainty equation for product functions, which was provided in Holman [17]:
(7)
π€Nu = [(
2 2 πNu πNu πNu π€ ) π€π ) + ( π€π ) + ( ππ ππ πππ ππ
+(
2
2 1/2
πNu πNu π€ππ ) + ( π€ )] πππ ππ· π·
2
(9) .
Using the respective data that are provided from the manufacturers, the accuracy of the HFX-4 Thin-Film Heat Flux Sensor was 0.02% and 0.0005 (ππ +1) for the heat flux and the surface temperature, respectively. The temperature accuracy for the πΎ-type thermocouple was 0.77%, and the nozzle diameter, which was measured using a Vernier calliper, was 0.25%. Consequently, the maximum uncertainty that is associated with the local Nusselt number was calculated to be 2.92% and 2.78% for Reynolds numbers of 4000 and 16000, respectively.
4
Advances in Mechanical Engineering 160
160
140
140
120
120 100 Nu
Nu
100 80
80 60
60
40 40 20 20 0
0 β8 0
2
4 r/D
6
β6
β4
β2
8
Present work, Re = 4000, L/D = 2 Nanan et al. [13], Re = 4000, L/D = 4 Choo and Kim [19], Re = 5000, L/D = 4 Gardon and Cobonque [20], Re = 5100, L/D = 6
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Conventional y = 4.89
Figure 4: Effect of the twist ratio (π¦/π) on the local Nusselt number distribution for Re = 4000.
Figure 3: Comparison between the present work and the previous studies.
160 140
Table 2: Geometry, flow, and physical properties.
3. Experimental Results 3.1. Validation of Results. Before performing the experiments with a swirl jet, the heat transfer results that were obtained using the conventional impinging jet (CIJ) were compared with the experimental works of Nanan et al. [13], Gulati et al. [18], and Choo and Kim [19], as shown in Figure 3. 3.2. Effects of the Twisted Tape Ratio and Jet Reynolds Number. The effect of the twist ratio of the twisted tape on the local Nusselt number distribution of a swirling impingement jet is illustrated in Figures 4, 5, 6, and 7. The Nusselt distribution of the conventional impinging jet (CIJ) showed one peak at
120 100 Nu
Boundary conditions Description π· 21 mm π 4.2 mm π = π»/π€ 2.93, 3.91, 4.89 π» Linear distance of the tape for 360β rotation π€ Width of the twisted tape (mm) Re 4,000β16,000 25β C ππ Plate Copper 1.225 kg/m3 πair πair 0.0242 W/mβ
K 1.7894π β 05 kg/m-s πair Specific heat ratio 1.4
80 60 40 20 0 β8
β6
β4
Conventional y = 4.89
β2
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Figure 5: Effect of the twist ratio (π¦/π) on the local Nusselt number distribution for Re = 8000.
the stagnation point (π/π· = 0) and an abrupt change along the radial axis. However, the swirling impingement jets (SIJ) exhibited two peaks, which were related to the impingement of the two jets that were produced from both sides of each twisted tape. Similar findings in multiple jets were reported in [13, 14]. From Figures 4β7, it is obvious that the largest twist ratio (π¦ = 4.89) produced a higher Nusselt number than the other ratios. This result indicates that the axial momentum
Advances in Mechanical Engineering
5
160
Outlet (pressure outlet)
Intlet (velocity inlet)
Tube side (wall) 140
Twist tape insert
120
Nu
100 Outlet (pressure outlet)
80 Plate
60
Figure 8: Swirling impinging jet assembly.
40 20 0 β8
β6
β4
β2
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Conventional y = 4.89
Figure 6: Effect of the twist ratio (π¦/π) on the local Nusselt number distribution for Re = 12000. 160 140 120
Nu
100 80 60 40 20 0 β8
β6
β4
β2
Conventional y = 4.89
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Figure 7: Effect of the twist ratio (π¦/π) on the local Nusselt number distributions for Re = 16000.
distribution that was induced by a large twist ratio was less significant than the one with a small twist.
4. Numerical Simulation The FLUENT-6.3.26 commercial CFD package was used to perform 3D calculations of the convective heat transfer on
the impinged plate, which is associated with conventional and swirling impinging jets. The model helped us predict and explain the experimental observations. The computational domain consisted of the entire pipe-plate assembly as shown in Figure 8: a cylindrical tube that was 21 mm in diameter and 300 mm long with a helical twist configuration, whose twist ratios were 2.93, 3.91, and 4.89. The geometry of the twisted tape inserts with different twist ratios was made by winding a 20 mm wide uniform strip using the twist option in face sweeping. Twist angles of 360β with lengths of 58.6, 78.2, and 97.8 mm for twist ratios of 2.93, 3.91, and 4.89 were generated using a perpendicular type of sweeping for the entire 300 mm. The details of the geometry, physical properties, and boundary conditions are provided in Table 2. The boundary conditions represent the experimental conditions with respect to the temperature and velocity of the impinging jet, the temperature of the ambient air, and the temperature of the plate. To adequately calculate the heat transfer, the turbulent flow fields in a particular layer through which heat is transferred must be accurately predicted. A segregated solver algorithm was selected as a solver with the settings of implicit formulation, time-independent calculation, viscous turbulent models, and energy equation. SIMPLE, which was the pressure-velocity coupling method and the second-order upwind scheme for the energy and momentum equations, were used to solve the model. The turbulent models are based on the Reynolds averaged Navier-Stokes equations for 2D and 3D domains. Furthermore, a default value for the convergence criterion was chosen for all calculated parameters. The following equations calculate the convective heat transfer in terms of the local Nusselt number and the local heat transfer coefficient: βπ· (10) , πΎ where β is the heat transfer coefficient, π· is the nozzle diameter, and π is the air conductivity: Nu =
β=
πβ , ππ€ β πref
(11)
where πβ is the convective heat flux, ππ€ is the impingement wall temperature, and πref is the reference temperature, which
6
Advances in Mechanical Engineering 140
120
120
100
100
80
80 Nu
Nu
140
60
60
40
40
20
20
0
0
2
Experimental data Standard KW SST KW
4 r/D
6
8
0
0
2
4
6
8
r/D
Standard KE RNG KE Realizable KE
Experimental data Standard KW SST KW
Standard KE RNG KE Realizable KE
Figure 9: Nusselt number distribution for the single jet impinging on a flat plate at π»/π· = 2 and Re = 4000.
Figure 10: Nusselt number distribution for the single jet impinging on a flat plate at π»/π· = 2 and Re = 12000.
is defined in the reference values panel that is available in FLUENT. There are two definitions of πref : the local adiabatic wall temperature, which is used by Gardon and Cobonque [20] and Goldstein et al. [21] or the jet exit temperature. In the present study, the jet exit temperature was close to the ambient temperature, ππ β ππ = Β±0.1β C; therefore, it was applied to calculate the local Nusselt number instead of the local adiabatic wall temperature.
that were obtained from this model are consistent with the experiments; the error was within 1β10% for 0.5 < π/π· < 2.
5. Validating the Model Experimental data for a conventional impinging jet were used to validate the CFD calculation models, which are shown in Figures 9 and 10. Five turbulence models were available in the ANSYS release 12.1: standard π-π, RNG π-π, realizable ππ, standard π-π, and shear stress transport (SST) π-π, and they were tested to select an optimum model for simulation. The enhanced wall treatment was used for standard π-π, RNG π-π, and realizable π-π. The enhanced wall treatment requires a fine mesh near the wall; therefore, 10 cells in the inner layer were chosen for the impingement plate surface to make π¦+ β 1. Meanwhile, the standard π-π model and the shear stress transport (SST) π-π model do not have the options of enhanced wall treatment. The simulation results revealed that the Nusselt number behaved differently from the experimental data near the stagnation region regardless of the turbulence model that was used, and this trend was closely reproduced by the turbulence model. The ideal model to predict the Nusselt number distribution along the radial direction is ST π-π. The numerical Nusselt number values
6. Swirling Impinging Jet (SIJ) Simulation 6.1. Local Nusselt Number. Figures 11, 12, 13, and 14 show the numerical Nusselt number distribution values along the radial direction using the ST π-π model of the conventional impinging jet (CIJ) and the ST π-π model for a swirl jet. It is noted that the Nusselt number of the conventional jets differs from the experimental data near the stagnation region (0 < π/π· < 0.5) as previously mentioned, and the maximum Nusselt number value occurs at the stagnation point π/π· = Β±0.5. Moreover, the locations of these points do not appear to change with an increasing Reynolds number. For the swirling impinging jets (SIJ), it is evident that the heat transfer behavior has similar patterns because double peaks of the Nusselt number are identified in the impingement region. The locations of these peaks appear to shift further in the vertical direction when the twist tape ratio increases, which coincides with the experimental results. 6.2. Nusselt Number. The Nusselt number contours for conventional jet and swirl jet via CFD simulation with Re = 12000 are shown in Figures 15, 16, 17, and 18. The figures indicate that the maximum heat transfer rate is found at the stagnation point for a conventional impinging jet, whereas the maximum heat transfer location shifts away from the stagnation point for the swirling jets because of the high tangential velocity components. For all swirling jets, the heat transfer regions are separated because of the blockage by the
7
160
160
140
140
120
120
100
100
80
Nu
Nu
Advances in Mechanical Engineering
60
60
40
40
20
20
0 β8
β6
β4
β2
0 r/D
2
4
6
0 β8
8
y = 3.91 y = 2.93
Conventional y = 4.89
β4
β2
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Figure 13: Effect of the twist ratio (π¦) on the local Nusselt number distribution for Re = 12000.
160
160
140
140
120
120
100
100 Nu
80
80
60
60
40
40
20
20
0 β8
β6
Conventional y = 4.89
Figure 11: Effect of the twist ratio (π¦) on the local Nusselt number distribution for Re = 4000.
Nu
80
β6
β4
Conventional y = 4.89
β2
0 r/D
2
4
6
8
y = 3.91 y = 2.93
0 β8
β6
β4
Conventional y = 4.89
β2
0 r/D
2
4
6
8
y = 3.91 y = 2.93
Figure 12: Effect of the twist ratio (π¦) on the local Nusselt number distribution for Re = 8000.
Figure 14: Effect of the twist ratio (π¦) on the local Nusselt number distribution for Re = 16000.
twisted top ridge, and the separation becomes notable as the twist ratio decreases. The maximum heat transfer rates of the nozzles with twist tapes of π¦ = 4.89 appear to be higher than those obtained with twist tapes of π¦ = 3.91 and π¦ = 2.93. This result could be caused by the subsidizing effect of the axial velocity intention by increasing the spreading rate prior to impingement.
7. Conclusions The present study experimentally and numerically investigated the effects of the twist ratio (π¦), which corresponds to the swirl intensity, and the jet Reynolds number on the local Nusselt number distribution on an impingement plate that was heated to 50β C. Classical twist tapes with three
8
Advances in Mechanical Engineering 1.15e + 02 1.10e + 02 1.04e + 02 9.85e + 01 9.29e + 01 8.74e + 01 8.18e + 01 7.63e + 01 7.08e + 01 6.52e + 01 5.97e + 01 5.42e + 01 4.86e + 01 4.31e + 01 3.76e + 01 3.20e + 01 2.65e + 01 2.10e + 01 1.54e + 01 9.88e + 00 4.34e + 00 Contours of surface Nusselt number
Y Z
Z
9.89e + 01 9.40e + 01 8.92e + 01 8.43e + 01 7.95e + 01 7.46e + 01 6.98e + 01 6.50e + 01 6.01e + 01 5.53e + 01 5.04e + 01 4.56e + 01 4.07e + 01 3.59e + 01 3.10e + 01 2.62e + 01 2.14e + 01 1.65e + 01 1.17e + 01 6.82e + 00 1.98e + 00
Y Z
X
Contours of surface Nusselt number
Figure 16: Map of Nusselt number contours for the swirl jet with π¦ = 2.93. 9.54e + 01 9.07e + 01 8.61e + 01 8.14e + 01 7.68e + 01 7.22e + 01 6.75e + 01 6.29e + 01 5.82e + 01 5.36e + 01 4.90e + 01 4.43e + 01 3.97e + 01 3.50e + 01 3.04e + 01 2.58e + 01 2.11e + 01 1.65e + 01 1.18e + 01 7.19e + 00 2.55e + 00
Y Z
Y
X
Figure 15: Map of Nusselt number contours for the conventional jet.
X
Contours of surface Nusselt number
Figure 17: Map of Nusselt number contours for the swirl jet with π¦ = 3.91.
1.04e + 02 9.97e + 01 9.55e + 01 9.12e + 01 8.70e + 01 8.28e + 01 7.85e + 01 7.43e + 01 7.00e + 01 6.58e + 01 6.16e + 01 5.73e + 01 5.31e + 01 4.89e + 01 4.46e + 01 4.04e + 01 3.62e + 01 3.19e + 01 2.77e + 01 2.35e + 01 1.92e + 01
X
Contours of surface Nusselt number
Figure 18: Map of Nusselt number contours for the swirl jet with π¦ = 4.89.
twist ratios (π¦ = 2.93, 3.91, and 4.89) were used as the jet nozzle inserts at a jet-to-plate spacing of πΏ/π· = 2 and four different jet Reynolds numbers (4000, 8000, 12000, and 16000). The results were compared with those for a conventional jet, and, evidently, the conventional jet has a greater maximum local Nusselt number than the swirl jets. Furthermore, the swirl jet with the largest test ratio (π¦ = 4.89) enhanced the heat transfer rate more than the swirl jets with π¦ = 2.93 and 3.91. Numerical simulations of the impinging jet flow were performed to validate and explain the experimental observations of various turbulence models. The computations show that the predicted results of the ST π-π turbulence model are reasonably consistent with the experimental results for conventional jet, whereas the ST-π-π model was more consistent with the swirl jets than the other models. Contour plots of the predicted Nusselt number were additionally presented. The obtained results demonstrate that the conventional-jet heat transfer rates were greater than the heat transfer rate of the nozzle with twisted tape inserts (π¦ = 4.89, 3.91, and 2.93). These results are consistent with the experimental work.
Nomenclature π·: β: π: Nu: πβ : π:
Nozzle diameter, m Local heat transfer coefficient, w/m2 k Thermal conductivity of air, w/mβ
k Local Nusselt number Convective heat flux Radial distance from the stagnation point, m π/π·: Dimensionless radial distance Re: Reynolds number based on the exit diameter, Re = ππ’π·/π π’: Mean velocity at jet exit, m/s π: Density of air, Kg/m3 π: Kinetic viscosity of air, Kg/mβ
s
Advances in Mechanical Engineering π: Temperature, K or β C Ξπ: Temperature difference, K or β C. Subscript π: π: ref: π€:
Ambient Jet Reference Wall.
Conflict of Interests The authors declare that there is no conflict of interests regarding the publication of this paper.
Acknowledgments The authors would like to thank the National University of Malaysia and the Ministry of Higher Education for the financial support (FRGS/1/2013/TK07/UKM/01/1 and DPP2013-114) to perform this investigation.
References [1] K.-C. Wong and N. H. Saeid, βJet impingement cooling of a finite thickness solid layer conjugated with porous medium,β Numerical Heat Transfer A: Applications, vol. 55, no. 7, pp. 706β 720, 2009. [2] M. M. Rahman, C. F. Hernandez, and J. C. Lallave, βFree liquid jet impingement from a slot nozzle to a curved plate,β Numerical Heat Transfer A: Applications, vol. 57, no. 11, pp. 799β821, 2010. [3] T. Demircan and H. Turkoglu, βThe numerical analysis of oscillating rectangular impinging jets,β Numerical Heat Transfer A: Applications, vol. 58, no. 2, pp. 146β161, 2010. [4] D. Jagannatha, R. Narayanaswamy, and T. T. Chandratilleke, βAnalysis of a synthetic jet-based electronic cooling module,β Numerical Heat Transfer A: Applications, vol. 56, no. 3, pp. 211β 229, 2009. [5] T. Yokomine, A. Shimizu, A. Saitoh, and K. Higa, βHeat trans-fer of multiple impinging jets with gas-solid suspensions,β Experimental Thermal and Fluid Science, vol. 26, no. 6-7, pp. 617β626, 2002. [6] L. Thielen, H. J. J. Jonker, and K. HanjaliΒ΄c, βSymmetry breaking of flow and heat transfer in multiple impinging jets,β International Journal of Heat and Fluid Flow, vol. 24, no. 4, pp. 444β453, 2003. [7] L. Thielen, K. HanjaliΒ΄c, H. Jonker, and R. Manceau, βPredictions of flow and heat transfer in multiple impinging jets with an elliptic-blending second-moment closure,β International Journal of Heat and Mass Transfer, vol. 48, no. 8, pp. 1583β1598, 2005. [8] T. Wang, J. L. Gaddis, and X. Li, βMist/steam heat transfer of multiple rows of impinging jets,β International Journal of Heat and Mass Transfer, vol. 48, no. 25-26, pp. 5179β5191, 2005. [9] K. Kanokjaruvijit and R. F. Martinez-Botas, βHeat transfer correlations of perpendicularly impinging jets on a hemisphericaldimpled surface,β International Journal of Heat and Mass Transfer, vol. 53, no. 15-16, pp. 3045β3056, 2010. [10] J. Badra, A. R. Masri, and M. Behnia, βEnhanced transient heat ransfer from arrays of jets impinging on a moving plate,β Heat Transfer Engineering, vol. 34, no. 4, pp. 361β371, 2013.
9 [11] K. Bakirci and K. Bilen, βVisualization of heat transfer for impinging swirl flow,β Experimental Thermal and Fluid Science, vol. 32, no. 1, pp. 182β191, 2007. [12] A. Ianiro and G. Cardone, βHeat transfer rate and uniformity in multichannel swirling impinging jets,β Applied Thermal Engineering, vol. 49, pp. 89β98, 2012. [13] K. Nanan, K. Wongcharee, C. Nuntadusit, and S. Eiamsa-ard, βForced convective heat transfer by swirling impinging jets issuing from nozzles equipped with twisted tapes,β International Communications in Heat and Mass Transfer, vol. 39, pp. 844β 852, 2012. [14] C. Nuntadusit, M. Wae-hayee, A. Bunyajitradulya, and S. Eiamsa-ard, βHeat transfer enhancement by multiple swirling impinging jets with twisted-tape swirl generators,β International Communications in Heat and Mass Transfer, vol. 39, no. 1, pp. 102β107, 2012. [15] C. Nuntadusit, M. Wae-hayee, A. Bunyajitradulya, and S. Eiamsa-ard, βVisualization of flow and heat transfer characteristics for swirling impinging jet,β International Communications in Heat and Mass Transfer, vol. 39, pp. 640β648, 2012. [16] G. Pathipakka and P. Sivashanmugam, βHeat transfer behaviour of nanofluids in a uniformly heated circular tube fitted with helical inserts in laminar flow,β Superlattices and Microstructures, vol. 47, no. 2, pp. 349β360, 2010. [17] J. P. Holman, Experimental Methods for Engineers, McGraw Hill, New York, NY, USA, 8th edition, 2011. [18] P. Gulati, V. Katti, and S. V. Prabhu, βInfluence of the shape of the nozzle on local heat transfer distribution between smooth flat surface and impinging air jet,β International Journal of Thermal Sciences, vol. 48, no. 3, pp. 602β617, 2009. [19] K. S. Choo and S. J. Kim, βHeat transfer characteristics of impinging air jets under a fixed pumping power condition,β International Journal of Heat and Mass Transfer, vol. 53, no. 1β3, pp. 320β326, 2010. [20] R. Gardon and J. Cobonque, βHeat transfer between a flat plate and jets of air impinging on it,β in Proceedings of the International Heat Transfer Conference, pp. 454β460, University of Colorado, 1962. [21] R. J. Goldstein, A. I. Behbahani, and K. K. Heppelmann, βStreamwise distribution of the recovery factor and the local heat transfer coefficient to an impinging circular air jet,β International Journal of Heat and Mass Transfer, vol. 29, no. 8, pp. 1227β1235, 1986.