You are on page 1of 22

Article 1

Numerical Study of the Effects of Roughness Coupled with Inclination 2

on a Turbulent Flow Around an Obstacle 3

Amine Brahimi 1,2 , Redha Rebhi 1,2 and Mounir Alliche1,2,* 4

1 LERM - Renewable Energy and Materials Laboratory, University of Medea, Medea, 26000, Algeria; 5
ami_brahimi@yahoo.fr 6
2 Department of Mechanical Engineering, Faculty of Technology, University of Medea, Medea 26000, Algeria; 7
rebhi.redha@gmail.com 8
* Correspondence: alliche_m@yahoo.fr 9

Abstract: In this study, we simulate the cooling of a microprocessor by thermal convection in three 10
different shapes: a square, a trapezoidal, and a triangular shape. The latter is improved by a variety 11
of types of roughness, including square roughness, triangular roughness (Type 1), triangular 12
roughness (Type 2), and triangular roughness (Type 3). The microprocessors are kept at a constant 13
temperature, the air flow is constant, and the geometry is fixed. The physical phenomenon is sim- 14
ulated by the ANSYS software. The findings relate to streamlines, dynamic pressure (max), mean 15
velocity, temperature field, mean Nusselt number (Nu/Nu0) profiles, local coefficient of friction 16
(Cf/f0), mean coefficient of friction (f/f0) profiles, mean velocity field with roughness and fluid 17
temperature field with roughness. The study's goal is to demonstrate how the geometry of the ob- 18
struction and the roughness parameter interact. 19

Keywords: Forced convection; obstacle; CFD ; roughness; Nusselt numbers; friction. 20


21

1. Introduction 22
Currently, forced or free convection environments are used to cool personal com- 23
puters. The most common cooling methods for greater or lesser heat flows involve fans 24
attached to the central processing unit. However, due to a number of factors, including 25
the size of the workspace, the noise, and the quality of the limit cooling, these cooling 26
systems have some limitations for electronic components that operate at higher frequen- 27
cies (over 1000 megahertz). As a result, novel cooling techniques must be used to dissi- 28
pate thermal energy greater than 100 W/cm2 from an electronic component's surface 29
Citation:To be added by editorial
while keeping the device (or component) at acceptable temperatures, usually below 30
staff during production.
85°C. In order to accelerate the rate of heat transmission, impediments are placed in the 31
Academic Editor: Firstname Last- channel's duct to induce turbulence. Baffles and fins are positioned in the forced flow to 32
name create secondary currents or recirculation zones in order to produce turbulence. These 33
Received: date are employed to enhance heat transfer in a number of engineering applications, such as 34
Revised: date solar channels and heat exchangers [1]. The improvement of heat transmission and 35
Accepted: date pressure drop caused by connecting barrier elements of various shapes, sizes, inclina- 36
Published: date tions, and orientations to the channel walls has been investigated by several researchers 37
[2-6] using experimental approaches. Their findings demonstrate that the Nusselt num- 38
ber and thermal efficiency were raised by experimental and numerical analyses to im- 39
Copyright:© 2023by the authors. prove thermal performance, The characterization of the air pressure and velocity field 40
Submitted for possible open access inside a rectangular duct with two baffle plates installed on opposing sides was provided 41
publication under the terms and by Demartini et al. [7]. Baffle plates are found in shell-and tube heat exchangers, and the 42
conditions of the Creative Commons geometry of the problem is a simplification of that geometry. The extent of the low 43
Attribution (CC BY) license pressure regions in the downstream regions and the high pressure regions generated 44
(https://creativecommons.org/license upstream of both baffle plates are the most significant aspects seen. Cao et al. [8] and 45
s/by/4.0/).

Processes2023, 11, x. https://doi.org/10.3390/xxxxx www.mdpi.com/journal/processes


Processes 2023, 11, x FOR PEER REVIEW 2 of 22

Saedodin et al. [9] using the Lattice Boltzmann method, Abchouyeh et al. [10] carried out 46
a numerical investigation of a horizontal sinusoidal baffle channel containing a water/Cu 47
nanofluid. They claimed that when the nanoparticle fraction and obstacle space in- 48
creased, so did the average Nusselt number values. According to the findings, the wedge 49
ribs significantly improved heat transfer and pressure drop. Chompookham et al. [11] 50
offer two types of triangular ribs. Using inclined solid and perforated baffles with a 51
Reynolds number of 12000 to 41000, two equally sized baffles were introduced into a 52
rectangular channel by Dutta and Hossain [12], increasing the local heat transfer charac- 53
teristics as well as the friction through the channel. The findings show that the local 54
Nusselt number has a considerable impact on the orientation, location, and shape of the 55
second deflector. A flat plate solar air heater's energy and exergy analyses were shown by 56
Kalaiarasi et al. [13]. In order to increase thermal efficiency, Peng et al. [14] demonstrated 57
a new kind of solar air collector that uses finned pins on its absorber. Selimefendigil et al. 58
[15] performed a computational analysis of the hydrothermal properties of the convective 59
laminar nanofluid Fe3O4 through a bifurcating type channel with a changeable magnetic 60
field based on the finite element method in addition to experimental methods. Through a 61
square pipe with various baffles, including trapezoidal, triangular and square shapes, 62
Kamali and Binesh [16] explored turbulent heat transmission and examined friction 63
properties. A baffle was inserted by Nasiruddin et al. [17] to enhance heat transfer inside 64
a circular tube. The effect of the baffles' size and position on the channel's thermal per- 65
formance was investigated. Pin fins were installed by Wang et al. [18] to enhance heat 66
transfer in a rectangular channel. The properties of forced convection heat transfer in the 67
presence of transverse grooves on the bottom surface of a two-dimensional channel were 68
reported by Eiamsa-ard and Promvonge [19]. With the use of circular section rings with 69
various spacing’s, Ozceyhan et al. [20] simulation of heat transmission and friction in a 70
tube was accomplished. FLUENT was used to run the simulation, and the Reynolds 71
values ranged from 4475 to 43725. In a square channel with discrete type "V" baffles, 72
Promvonge et al. [21] simulate a turbulent flow and heat transfer in three dimensions. 73
This simulation made use of the finite volume approach and the SIMPLE algorithm. In a 74
two-dimensional channel, the effects of introducing diamond-shaped baffles with vari- 75
ous tip angles on thermal enhancement were examined statistically by Sripattanapipat 76
and Promvonge [22]. The Lattice Boltzmann method was used by Pirouz et al. [23] to 77
model conjugate heat flow inside a rectangular channel with associated barriers. Mo- 78
hammedi et al. [24] investigated how to increase heat transfer efficiency and outlet air 79
temperature in a finned and baffled solar water heater. The enhancement of heat transfer 80
inside a square duct fitted with oblique horseshoe baffles was explored by Skullong et al., 81
[25].When airflow and an absorber installed with corrugated fins were present, Priyam 82
and Chand [26] investigated the effect of flow on the efficiency of the solar collector. To 83
assess the internal flow and heat transfer properties of a solar air collector with baffles, 84
Hu et al. [27] used numerical analysis. Amraoui and Aliane [28] used the CFD technique 85
to simulate the research of fluid flow and heat transfer in a solar channel with obstacles, 86
while Amraoui [29] used CFX software to simulate the study of air flow around a circular 87
obstacle field in an air CSP. The same author, Amraoui [30], conducted a 88
three-dimensional evaluation of two different solar collector types more recently. To 89
complete their research quickly and cheaply, they used the ANSYS simulation code.In 90
order to improve efficiency, Amraoui and Benosman [31] incorporated square barriers. 91
The height and pitch of the roughness element are the two most significant characteris- 92
tics, although many other factors contribute to the arrangement and shape of the rough- 93
ness components. Reynolds number, rib cross section, angle of attack, and combined 94
turbulence promoters are further parameters. Using direct numerical simulations, 95
Leonardi et al. [32] investigated the ordered motion of a turbulent flow in a channel with 96
a series of square bars on the lower and upper sides. For the examination of heat transfer 97
in a rough-walled conduit with sand grains and two-dimensional periodic square section 98
ribs on a wall surface for turbulent flow, Miyake et al. [33] performed direct numeri- 99
Processes 2023, 11, x FOR PEER REVIEW 3 of 22

casimulations. The impact of repeated ribs on the thermal efficiency of a flat absorber 100
plate solar air channel was investigated by Ansari and Bazargan [34]. When designing a 101
ribbed rough surface for a channel, Kim and Kim [35] presented a study on the numerical 102
optimization technique combined with RANS analysis of flow and heat transmission. A 103
two-passage square channel with and without square section parallel ribs at a 90° angle 104
on a wall surface was subjected to a 3D numerical simulation by Jang et al. [36]. With 105
angular ribs extruded on two opposing surfaces, Kim and Kim [37] carried out a numer- 106
ical simulation of fluid flow and heat transfer for channel shape optimization. In a tri- 107
angular duct with ribs, Kumar et al. [38] looked into the flow and heat transmission 108
characteristics. The properties of heat transmission and friction in a square duct with 109
differently shaped ribs mounted on a wall transverse to the main flow direction were 110
characterized by Kamali and Binesh [39].In the context of surface roughness effects, Ryu 111
et al. [40] looked into the heat transfer characteristics of turbulent flow in channels with 112
two-dimensional ribs and three-dimensional blocks. In order to analyze the heat transfer 113
and hydraulic behavior of the fluid in a rectangular channel flow with periodic ribs in- 114
stalled on one of the major walls, Liu et al. [41] conducted a numerical and experimental 115
investigation. To forecast the thermal-hydraulic performance of a solar air channel 116
roughened with conical and spherical ribs, Alam and Kim [42] used numerical simula- 117
tions. For the numerical optimization of the energy and exergy efficiency of a solar air 118
channel with twisted rib roughness on a heating plate, Kumar and Layek [43] conducted 119
a stochastic analysis. Skullong et al. [44] used corrugated grooves embedded with pairs 120
of trapezoidal fins (ATs) positioned on the absorber plate to investigate the heat transfer 121
properties in a solar air channel. In order to explore the heat transfer, friction factor, and 122
thermal-hydraulic performance characteristics of flow in a rough rectangular duct, Deo et 123
al. [45] carried out an experimental investigation. The thermal and hydraulic perfor- 124
mance of a roughened, dual-flow solar air channel with several C-shaped ribs was ex- 125
perimentally investigated by Gabhane and Kanase-Patil [46]. An experimental investiga- 126
tion on heat transmission and coefficient of friction for a synthetically rough solar air 127
duct was presented by Kumar et al. [47]. For the range of tested roughness parameters, it 128
was discovered that rough solar air duct performs better than smooth duct. The experi- 129
mental research of heat transmission and friction factor in a countercurrent double-pass 130
solar air channel with a V-shaped discrete rib roughness and shifted over two substantial 131
surfaces of the heated plate was reviewed by Ravi and Saini [48]. 132
The numerical analysis of the turbulent air flow in forced convection around a bar- 133
rier will be of particular interest to us in this work. In order to validate this model, the 134
outcomes will be compared to those reported in the literature under similar circum- 135
stances. Our project uses fluid calculation code to mimic the cooling of a microprocessor 136
in two dimensions. 137

2. Methodology 138
The investigation revolved around a heat exchanger. It has a horizontal, rectangular 139
channel, an isothermal upper wall, a thermally insulated lower wall, and extended sur- 140
faces in the form of intermittently spaced-apart obstacle. Figure 1 presents a precise ge- 141
ometric illustration. 142
143
144
145
146
147
148
149
150

Figure 1. Geometry under investigation and boundary conditions. 151


Processes 2023, 11, x FOR PEER REVIEW 4 of 22

In order to determine the ideal setting for enhanced heat transfer, we primarily 152
concentrate on an essential geometric variable associated withan obstacle in the first sec- 153
tion of this study, namely the effect of the slope of their leading edges. Figure 2 shows 154
three cases that can occur depending on the leading edge of the obstacle's angle of incli- 155
nation: 156
• Square obstacles in the case where: 𝜃 = 0° 157

• Trapezoidal obstacles in the case where: 0° < 𝜃 < 45° 158

• Triangular obstacles in the case where: 𝜃 = 45° 159


160
161
162
163
164
165
166
167

Figure 2. Different angles of inclination of the straight sides of obstacles: (a) 𝜃 = 0°: square obsta- 168
cle, (b) 0 < 𝜃 < 45°: trapezoidal obstacle, and (c) 𝜃 = 45°: triangular obstacle. 169

170
To improve the performance of the heat exchanger channel, rough walls are used in 171
the presence of fins and triangular baffles. The third model, figure 2(c) is compared to the 172
performances of four channels in the presence of four roughness modes: square, figure 173
3(a), triangular type 1, Figure 3(b), triangular type 2, Figure 3(c), and triangular type 3, 174
Figure 3(d).In all the cases proposed, the roughness is present only on the upper hot wall 175
of the channel, exactly in line with the last obstacle. 176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191

Figure 3.Various cases of roughness shapes. 192

The specific parameters of the geometry and the thermo-physical data of the fluid 193
(air) and of the solid (walls) are indicated in Tables 1 and 2. 194

Table 1. Geometric parameters of heat exchanger channel with 𝐻 = 0.01𝑚 195

Dimension H
Channel length 25H
Channel height 2H
Channel hydraulic diameter 2H
Processes 2023, 11, x FOR PEER REVIEW 5 of 22

Obstacle width 1H
Obstacle height 1H
Separation distance from obstacles 3H
Roughness width 0.5H
Roughness height 0.5H
Distance between two roughness units 0.5H
Distance between the channel entrance and the left side of the first obstacle 5H
Distance between the right side of the last obstacle and the exit of the conduit 7H
Distance from the right side of the last obstacle to the first roughness unit 1H
Distance from the right side of the last roughness unit to the exit of the channel 0.5H

Table 2. Thermo-physical data of Air and Al at 300K. 196

𝒌 (𝑾 /𝒎. 𝑲) 𝑪𝒑 (𝑱 /𝒌𝒈. 𝑲) 𝝆 (𝒌𝒈 /𝒎𝟑 ) µ (𝑷𝒂. 𝒔) 𝑷𝒓


Air 0.0242 1006.43 1.225 1.7894×10 -5 0.71
Al 0.1672 14283 0.08189 …. …
The aerodynamic and thermal boundary conditionsare chosen according to the 197
simulation of Nasiruddin and Kamran Siddiqui [17]as shown in Figure 1. 198
199
At the inlet(𝑥 = 0,0 ≤ 𝑦 ≤ 2𝐻): 200
𝑢(0, 𝑦) = 𝑈𝑖𝑛
𝑣(0, 𝑦) = 0
𝑇(0, 𝑦) = 𝑇𝑖𝑛 (1) 201
2
𝑘(0, 𝑦) = 𝑘𝑖𝑛 = 0.005 × 𝑈𝑖𝑛
2
𝜀(0, 𝑦) = 𝜀𝑖𝑛 = 0.1 × 𝑘𝑖𝑛 }
At the outlet(𝑥 = 25𝐻, 0 ≤ 𝑦 ≤ 2𝐻) : 202
𝑃(𝐿, 𝑦) = 𝑃𝑎𝑡𝑚
𝜕∅ } (2) 203
(𝐿, 𝑦) = 0
𝜕𝑥
Where 𝜙 ≡ (𝑢, 𝑣, 𝑇, 𝑘, 𝜀). 204
At the upper surface (0 ≤ 𝑥 ≤ 25 𝐻, 𝑦 = 2𝐻) : 205

𝑢=𝑣=0
𝑘=𝜀=0 } (3) 206
𝑇 = 𝑇𝑤
At the lower surface (0 ≤ 𝑥 ≤ 25 𝐻, 𝑦 = 0) : 207
𝑢=𝑣=0
𝑘=𝜀=0 } (4) 208
𝜕𝑇
=0
𝜕𝑦

At the solid/fluid interface, the following condition is applied: 209

𝑇𝑠 = 𝑇𝑓
𝜕𝑇𝑠 𝜕𝑇𝑓 } (5) 210
𝜆𝑆 | = 𝜆𝑓 |
𝜕𝑛 𝑁
⃗ 𝜕𝑛 𝑁

Where 𝑁 ⃗ is the vector normal to the considered surface interface,𝜆𝑆 and 𝜆𝑓 are 211
thermal conductivities of solid and fluid respectively. 212
The computational domain fitted with square obstacle, trapezoidal obstacle and 213
triangular obstacle is simulated using the Computational Fluid Dynamics (CFD) com- 214
mercial software ANSYS Fluent. 215
Processes 2023, 11, x FOR PEER REVIEW 6 of 22

The k − ε turbulence is adopted in this study. In Cartesian coordinates, the conti- 216
nuity, momentum, energy, and turbulence equations can be written in the following 217
compact form Patankar [49]: 218

𝜕 𝜕 𝜕 𝜕𝜙 𝜕 𝜕𝜙
(𝜌𝑢𝜙) + (𝜌𝑣𝜙) = [𝛤𝜙 ]+ [𝛤𝜙 ] + 𝑆𝜙 } (6) 219
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑦

Where 𝜙 is a variable that serves to represent quantities such as the velocity com- 220
ponents 𝑢and 𝑣, the turbulent kinetic energy 𝑘 or the rate of turbulent energy dissipa- 221
tion 𝜀, and the temperature 𝑇. However, the diffusion coefficient 𝛤𝜙 and the source term 222
𝑆𝜙 have specific values for the different conservation equations in the case of the stand- 223
ard k − ε turbulence model. 224
225
Continuity equation 226

∅=1
Γ𝜙 = 0 } (7) 227
𝑆𝜙 = 0
Momentum equation in X-direction 228
∅=𝑢
Γ𝜙 = 𝜇𝑒
} (8) 229
𝜕𝑃 𝜕 𝜕𝑢 𝜕 𝜕𝑣
𝑆𝜙 = − + [𝜇𝑒 ( )] + [𝜇𝑒 ( )]
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑥
Momentum equation in Y-direction 230
∅=𝑣
Γ𝜙 = 𝜇𝑒
} (9) 231
𝜕𝑃 𝜕 𝜕𝑢 𝜕 𝜕𝑣
𝑆𝜙 = − + [𝜇𝑒 ( )] + [𝜇𝑒 ( )]
𝜕𝑦 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑥
Conservation of energy in the fluid region 232
∅=𝑇
𝜇
Γ𝜙 = 𝑒 } (10) 233
𝜎𝑇
𝑆𝜙 = 0
Conservation of energy in the solid region 234
∅=𝑇
Γ𝜙 = 𝜆𝑆
} (11) 235
𝜕 𝜕
𝑆𝜙 = (𝜌𝑠 𝑢𝑇) + (𝜌𝑠 𝑣𝑇)
𝜕𝑥 𝜕𝑦
k-Turbulent kinetic energy equation 236
∅=𝑘
𝜇
Γ𝜙 = 𝜇𝑙 + 𝑡 } (12) 237
𝜎𝑘
𝑆𝜙 = −𝜌𝜀 + 𝐺𝑘
ε-Turbulent dissipation rate equation 238
∅=𝜀
𝜇
Γ𝜙 = 𝜇𝑙 + 𝑡 } (13) 239
𝜎𝜀
𝜀
𝑆𝜙 = (𝐶1𝜀 𝐺𝑘 − 𝐶2𝜀 𝜌𝜀)
𝑘
𝐺𝑘 is the rate of kinetic energy production due to energy transfer from the turbulent 240
mean flow; it is given by: 241

𝜇𝑒𝑓𝑓 = 𝜇𝑙 + 𝜇𝑡
𝜕𝑢 2 𝜕𝑣 2 𝜕𝑢 𝜕𝑣 2 } (14) 242
𝐺𝑘 = 𝜇𝑡 {2 [( ) + ( ) ] + ( + ) }
𝜕𝑥 𝜕𝑦 𝜕𝑦 𝜕𝑥

where 𝐶𝜇 = 0.09;𝐶1𝜀 = 1.44;𝐶2𝜀 = 1.92;𝜎𝑘 = 1.0; 𝜎𝜀 = 1.3, 𝜎𝑇 = 0.9are the constants 243
of the model,as proposed by Laundry and Spalding [49]. 244
Processes 2023, 11, x FOR PEER REVIEW 7 of 22

3. Results and Discussion 245


For the numerical simulations presented in this article, we refer to the experimental 246
work of Demartini et al. [7], who studied planar baffles. Figure 4 shows our result and 247
those obtained by Demartini et al.[7]. It depicts the pressure coefficient profiles with 248
some previously studied numerical results for Uin = 7.8 m/s after the second chicane at x = 249
0.405 m. A good agreement between the numerical and experimental results is observed, 250
showing the growth of the pressure coefficient near the walls. The lower pressure values 251
near the tip of the baffles are due to the high velocities in this region. 252

253
Figure 4. Comparison of pressure coefficient with some previously studied numerical results for 254
𝑈𝑖𝑛 = 7,8𝑚/𝑠 255

The evolution of the streamlines for the various types of obstacles with angles of θ = 256
0°, 0<θ< 45° and θ= 45° is shown in Figure 5. According to the obstacle, we see that the 257
streamlines alternate between the three forms, Figure 5(a)-(c). The streamlines in Figure 258
5(a) (square obstacle) are less contoured as they cross the barrier, causing turbulence that 259
disrupts the flow. The streamlines in Figure 5(b), trapezoidal's obstacle, straighten as they 260
cross it, changing the streamlines' field direction. The streamlines are attached to the ob- 261
stacle in Figure 5(c), triangular obstacle, giving us a powerful streamline field. 262
𝑍𝑜𝑛𝑒 𝐵 263
264
265
266
267
268
𝑍𝑜𝑛𝑒 𝐴
269
(a)
270
271
272

(b) 273
274
275
276
(c) 277
𝑍𝑜𝑛𝑒 𝐶 278

Figure 5. Streamlines for Re = 2500 and different obstacles (a) θ = 0°: square obstacle, (b) 0 <θ< 45°: 279
trapezoidal obstacle; (c) θ = 45°: triangular obstacle. 280

According to Figure 5, complex phenomena including recirculation zones and re- 281
gions of turbulence upstream and downstream of the obstacle are produced by the flows 282
Processes 2023, 11, x FOR PEER REVIEW 8 of 22

in the air veins for the three different types of obstacles. The existence of impediments 283
leads the three air streams to divide into multiple distinct zones above, below, and up- 284
stream of the obstructions. As a result, the flow creates recirculation zones (Zone A) close 285
to the lower obstacles, which intensifies the turbulence upstream of the upper obstacles 286
(Zone B). In addition to the separation of the fluid flow induced by the obstructions 287
(Zone C), there are dead zones upstream of the lower obstacles (Zone C) plus to the sep- 288
aration of the fluid flow caused by the obstacles. We observe that the importance of the 289
dead zone for trapezoidal-shaped obstacles is less than that of square obstacles, and that 290
it is almost zero for the model with triangular obstacles. When compared to previous 291
models that have square and trapezoidal barriers, the model with triangle obstacles has a 292
significantly a large turbulent zone (Zone B). Figure 6 depicts dynamic pressure, which is 293
the flow of momentum per unit volume (the kinetic energy density). The three varieties 294
under study are similar in terms of the overall dynamic pressure setup. 295
The dynamic pressure is uniform at the entrance, immediately below the first upper 296
obstacle zone. Upstream of the upper and lower obstacles, the dynamic pressure ap- 297
proaches zero values. The zones (A, B, C, and D) with the highest dynamic pressure are 298
separated by the sharp upstream edge. However, the flow separates from the obstacle's 299
wall, which results in a depression downwind of these obstacles. Due to the poor circu- 300
lation at this location, the dynamic pressure at the Attack terminals in Figure 6(a) is 301
nearly negligible. The dynamic pressure in Figure 6(b) must be displaced along the pro- 302
file, causing an average flow of kinetic energy. The pressure in Figure 6(c) has a substan- 303
tial value since the profile is ideal, which allowed for a large movement. 304

𝑍𝑜𝑛𝑒 𝐴 305
𝑍𝑜𝑛𝑒 𝐵
306
307
308
309
(a) 𝑍𝑜𝑛𝑒 𝐶 𝑍𝑜𝑛𝑒 𝐷 310
311
312
313
(b) 314
315
316
317
(c)
318

Fig.6. Contours of dynamic pressure fields for Re = 2500 and different angles of obstacles: (a) θ = 0°: 319
square obstacle; (b) 0 <θ< 45°: trapezoidal obstacle; and (c) θ = 45°: triangular obstacle, 320

For the three types of obstacles, 0<θ< 45°, Figure 7 shows the dynamic pressure 321
fluctuation as a function of the Reynolds number (Re = 2500, 5000, 7500 and 10000). Re- 322
member that pressure increases every time the Reynolds number increases. As can be 323
seen, the pressure values for the square and trapezoidal obstacle are extremely close to 324
each other. In addition, compared to the two preceding situations, the pressure value in 325
the triangular case is relatively low. 326
327
328
329
330
331
332
333
334
335
Processes 2023, 11, x FOR PEER REVIEW 9 of 22

336
337
338
339
340
341
342
343
344
345
346
347
348
349
350

Figure 7. The effect of the inclination of the upstream edge of the obstacle on the dynamic pressure 351
(max) for various values of the Reynolds number. 352

The entire distance traveled over a certain period of time divided by that period of 353
time is the average velocity. The vortex is always created upstream of the rectangular 354
obstacle, as seen in Figure 8. On the other hand, at the level of the triangular and trape- 355
zoidal obstacles, the flow consistently adheres to the obstacle's wall and follows it. As a 356
result, the flow has two crucial moments. Figures clearly demonstrate that the ridge 357
downstream of the obstacle has an impact on fluid behavior even upstream of it, altering 358
how the flow interacts with the upper wall of the triangle obstacle. This enables us to 359
claim that the presence of the inclination at the level of the edge upstream of the obstacle 360
not only controls the zone of the triangle but also the area around this obstacle. 361
362
363
364
365
(a) 366
367
368
369
(b)
370
371
372
373
374
(c)
375

Figure 8. Contours of mean velocity fields for Re = 2500 and different obstacle: (a) θ = 0°: square 376
obstacle; (b) 0 <θ< 45°: trapezoidal obstacle; and (c) θ = 45°: triangular obstacle. 377

According to the average intensity of the velocity, different zones are identified in 378
the three models under investigation (low or high). Just upstream of the first obstacle, the 379
average velocities for the three models under study are low. Along the downstream side 380
of square, trapezoidal and triangular obstacles, they are also weak. For the three models 381
under study, the average velocities are very high under both the upper and lower obsta- 382
cle. 383
The distribution of the average velocity as a function of the Reynolds number is 384
shown in Figure 9. Keep in mind that the average velocity increases as the Reynolds 385
number increases. Additionally, we discover that for Reynolds values 2000–6000, the 386
Processes 2023, 11, x FOR PEER REVIEW 10 of 22

average velocity in the three instances was quite similar. Reynolds number 6000 indicates 387
that the average velocity dropped in the case of the triangular obstacle and stayed con- 388
stant in the situations of the square and trapezoidal obstacles. 389
390
391
392
393
394
395
396
397
398
399
400
401
402
403

Figure 9. Average velocity as a function of the Reynolds number in the three cases: square, trape- 404
zoidal, and triangular. 405

In the square, trapezoidal and triangular examples under study, the fluctuation of 406
the axial velocity is depicted in Figure 10. As a result of the obstacle, it can be seen that 407
the flow is actually accelerated. The flow's perfect straightening upstream of a triangular 408
obstacle, axial velocity is crucial. Velocity is affected in the case of the trapezoidal obsta- 409
cle. But because of the vortex that is created in the square case, the velocity is varying. 410
411
412
413
414
415
(a) 416
417
418
419
(b) 420
421
422
423
(c)
424

Figure 10. Contours of the axial velocity fields for Re = 2500 and different angles of obstacles: (a) θ 425
= 0°: square obstacle; (b) 0 <θ< 45°: trapezoidal obstacle; and (c) θ = 45°: triangular obstacle. 426

At x=0.07m and for various values of the Reynolds number in each of the cases taken 427
into consideration. Figure 11 shows the fluctuation of the axial velocity as a function of 428
the height of the channel. When the channel height is between 0 and 0.015m, it is shown 429
that the velocity increases, and that the velocity increases as the value of the Reynolds 430
number increases. If we use a Reynolds number to examine the axial velocity profiles for 431
the three models, we find that the square and trapezoidal obstacle models have nearly 432
identical velocities. 433
For the model with triangle obstacles, the velocities take substantial values. The ve- 434
locities are observed to be in the opposite direction of the flow at a height of 0.015m, and 435
the model with triangular obstacles has the highest values of velocities when compared 436
to the other models (Figure 12). 437
Processes 2023, 11, x FOR PEER REVIEW 11 of 22

438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454

Figure 11. Profiles of the axial velocity, u, downstream of the first obstacle at x = 0.07m for different 455
angles of obstacles and various values of the Reynolds number. 456

457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473

Figure 12. Profiles of the axial velocity for different angles of obstacles and various values of the 474
Reynolds number. 475

The profiles of the ratios (u/Uin )at the channel's outflow (x=L) are shown in Figure 476
13. These profiles demonstrate that for the smallest Reynolds number, Re=2500, the ratios 477
(u/Uin )(are similar. The ratios (u/Uin )(at the point (0 up to 0.01m) of the model with a 478
triangular obstacle are low for the maximum Reynolds number, Re = 10000. For the 479
square obstacle model has high ratios (u/Uin ).The model with a triangular obstacle has 480
large ratios (u/Uin )at positions (0.01 to 0.018 m). 481
On the square obstacle model has low ratios (u/Uin ).The other height reports 482
(u/Uin )(for the three models range from 0.018 to 0.02m. The fluid is viscous, thus when 483
we get close to the wall, the velocity is zero. Because of the constant section and the var- 484
ied flow, the various Reynolds numbers result in a variation in velocity. 485
486
487
488
Processes 2023, 11, x FOR PEER REVIEW 12 of 22

489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508

Figure 13. Dimensionless axial velocity profiles (u/Uin) at the channel outlet (x = L), for different 509
hydrodynamic (2500 ≤ Re ≤ 10000) and geometric (0 ≤ θ ≤ 45°) parameters. 510

The change in average axial velocity for the three examples is shown in Figure 14 as 511
a function of Reynolds number. As you can see, the axial velocity increases as the Reyn- 512
olds number increases. However, the three investigations revealed an increase in axial 513
velocity. When the Reynolds number is between 2000 and 4000, the velocity value is 514
roughly constant at 5m/s. 515
From the value of the Reynolds number 4000, it can be seen that the three examples 516
have different velocities, with the triangular obstacle having the smallest increase in ve- 517
locity, followed by the trapezoidal obstacle, and then the square obstacle. 518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533

Figure 14. Maximum axial speeds for the different cases studied. 534

Figure 15 shows how the vertical (transverse) velocity varies in the three cases of the 535
triangle, trapezoid, and square. With the exception of the head of the obstacles, the 536
transverse velocity is homogeneous throughout the duct of the channel. However, at the 537
left extremities of the upper and lower obstacles, the velocity assumes positive and neg- 538
Processes 2023, 11, x FOR PEER REVIEW 13 of 22

ative values, respectively. Upstream of a square obstacle, the velocity is high because of 539
the abrupt pipe section reduction. 540
The straightening of the flow causes a trapezoidal obstacle to reduce velocity. Due to 541
the consistent convergence of the channel segment in the event of a triangular obstacle, 542
the velocity is low. 543
544
545
546
547
(a)
548
549
550
551
552
(b) 553
554
555
556
(c)
557

Figure 15. The transverse velocity field in the three cases: square, trapezoidal and triangular, for Re 558
= 2500 559

For various shapes of obstacles, Figure 16 illustrates how the maximum transverse 560
velocity varies as a function of the Reynolds number. The velocity rises as the Reynolds 561
number rises. Due to the acute deformation of the pipe section, which results in a vertical 562
translation of the fluid particles, the square shape has the fastest velocity. Because of the 563
low flow convergence, the trapezoidal form moves slower than the square impediment. 564
Compared to other obstacle forms, the triangle shape moves at a slower speed, which 565
reduces the vertical component of velocity by half. 566
567
568
569
570
571
572
573
574
575
576
577
578
579
580

Figure 16. Transverse velocity for different Reynolds numbers in the three cases: square, trape- 581
zoidal and triangular 582

The temperature field is depicted in Figure 17. Forced convection causes the air near 583
the absorber to heat up quickly the secondary air in the lower part of the channel down- 584
stream of the upper obstacles in zone (A); this suggests that the temperature is high for 585
the three models under consideration. It has been shown that the square obstacle’s weak 586
flow recirculation results in minimal contact between the cold air and the obstacle. Be- 587
cause of the obstacle's diversion of the flow in the trapezoidal cases, we see that there is 588
Processes 2023, 11, x FOR PEER REVIEW 14 of 22

good turbulence. We get good cooling in the triangular instance because the interaction 589
between the cold air and the impediment is significant downstream of the obstacle. 590
591
592
593
594
595
(a) 596
597
598
599
600
(b) 601
602
603
604
605
(c) 606
607

Figure 17. The field of the transverse temperature in the three cases: square, trapezoidal and tri- 608
angular for Re = 2500 609

The temperature change at the channel's exit is shown in Figure 18 for a range of 610
Reynolds number values and obstacle types. With increasing pipe height, the fluid tem- 611
perature profile rises exponentially with a different curve for each value of the Reynolds 612
number. Due to the substantial heat flux that heats the surface, the triangle obstacle has 613
the maximum heat transfer. The temperature value at the middle of two configurations is 614
what causes the trapezoidal obstacle to an average heat flow (between square and tri- 615
angular). Due to the cooling air temperature, the square obstacle has the weakest heat 616
transmission (poor cooling). It can be concluded that the temperature values at the model 617
channel with triangle obstacles are very high. The three models' temperatures are com- 618
parable when the Reynolds number is raised to 10,000. 619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636

Figure 18. Temperature profile at the outlet of channel in the three cases: square, trapezoidal, and 637
triangular. 638
Processes 2023, 11, x FOR PEER REVIEW 15 of 22

Figure 19 shows how the square, trapezoidal and triangular shapes of obstacles af- 639
fect the evolution of the adimensional Nusselt ratio (Nux /Nu0 )with channel length. De- 640
pending on the width of the channel, the Nusselt number fluctuates sinusoidally between 641
obstacles. 642
The turbulence of the flow results in increased convective heat transfer, which is il- 643
lustrated in the picture as having a significant value for the Nusselt number in the case of 644
the triangle obstacle. Because the heat flow is blended in the trapezoidal enclosure, the 645
heat transfer is better (a solid part and a fluid part). Because there is more heat transfer in 646
the solid in the square example, the Nusselt number is low, emphasizing the significance 647
of conduction. The important values in the model with triangle obstacles are used to 648
calculate the normalized local Nusselt number; the value of the ratio (Nux /Nu0 )reaches a 649
maximum of 1600 at the 0.17m position. This explains why the model with triangle ob- 650
stacles had better heat transfer than the other models analyzed. 651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666

Figure 19. The normalized local Nusselt number (Nux /Nu0 )along the channel in the three cases: 667
square, trapezoidal, and triangular for Re = 2500 668

The change in the average normalized Nusselt number (Nux /Nu0 ) as a function of 669
the Reynolds number for various obstacle forms is shown in Figure 20. It has been noted 670
that as the Reynolds number rises, the average Nusselt value also rises (convection is 671
important for a high Reynolds value, and conduction is almost zero for high velocities). 672
The high axial velocity produces a higher convection force than the other examples, by 673
increasing the fluid's recirculation zones, as shown in the picture, which demonstrates 674
that the Nusselt number has a considerable value in the case of the triangle obstruction. 675
The Nusselt number is smaller for the trapezoidal obstacle because the shape itself cre- 676
ates the restriction to the flow. Due to the geometry of the impediment blocking flow 677
velocity in the square case, the Nusselt number is low. 678
679
680
681
682
683
684
685
686
687
688
689
690
691
Processes 2023, 11, x FOR PEER REVIEW 16 of 22

692

Figure 20. Average Nusselt number (Nu/Nu0 )along the hot wall (upper wall) for different Re 693
numbers in the three cases: square, trapezoidal, and triangular. 694

Over the course of the channel's whole hot length, Figure 21 depicts the local varia- 695
tion of the friction coefficient's adimensional ratio (Cf /f0 ).Only at the level of the obsta- 696
cles does the coefficient of friction rise. The figure demonstrates that for the square ob- 697
stacle, the friction coefficient has a high value. At a distance of 0.19m, the value of 698
(Cf /f0 )can be as high as 4800. The average value of the coefficient of friction is zero due to 699
the flow's straightening. Because the air flow is homogeneous in the triangular example, 700
the coefficient of friction is low. 701
This shows how the physical characteristic of fluid flow the increase in heat transfer 702
after the decrease in the local coefficient of friction is significantly influenced by the ge- 703
ometry of the obstacles. 704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719

Figure 21. The profile of the local friction coefficient (Cf /f0 )along the channel (upper wall) in the 720
three cases: square, trapezoidal, and triangular for Re = 2500 721

Figure 22 illustrates the evolution of the friction ratio (f/f0 )according to the different 722
values of the Reynolds number for different obstacle models. On the other hand, it is seen 723
that for the three models examined, the average coefficient of friction rises as the Reyn- 724
olds number rises. When comparing the models for a given Reynolds number, we see 725
that the average coefficient of friction values for the model with square obstacles are 726
crucial due to the significant turbulence they produce. However, because of air recircu- 727
lation, the average coefficient of friction values for the model with triangular obstacles 728
are low. 729
730
731
732
733
734
735
736
737
738
739
740
741
742
Processes 2023, 11, x FOR PEER REVIEW 17 of 22

743

Figure 22.Average coefficient of friction (f/f0 )at different numbers of Re along the hot wall (upper 744
wall) in the three cases: square, trapezoidal, and triangular. 745

To enhance heat transmission between the absorber and the heat transfer fluid, the 746
researched roughnesses square, triangular Type 1, Type 2, and Type 3are situated on the 747
hot top part of the channel (absorber), downstream of the second baffle. In the presence 748
of triangular obstacles, Figure 23 shows the distribution of the average velocity as a 749
function of the shape of the roughness (square roughness and triangle roughness). Due to 750
the driving surface's unprofiled deformation, the average velocity in the case of square 751
roughness is insignificant. Because the subsequent profile lies downstream of the trian- 752
gular roughness (Type 1), the average speed is low at the beginning of the roughness. 753
Because of the flow profiling discontinuities with the triangular roughness (Type 2), the 754
average velocity is relatively high. The average velocity has a significant value when 755
there is triangular roughness (Type 3). 756
757
758
759
760
(a)
761
762
763

(b) 764
765
766
767
(c) 768
769
770
771
772
(d) 773
774

Figure 23. Average velocity fields in various roughnesses for Re = 2500. 775

Figure 24 displays the temperature distribution for each of the four proposed mod- 776
els for the roughness, fin presence, and triangle baffles across the entire examined region. 777
The temperature is quite high where the absorber and obstacles are located, as well as in 778
the roughness asperities. The channel's exit is improved by the addition of roughness. 779
Due to inadequate air recirculation at the roughness's center, the fluid temperature field 780
has low cooling effectiveness in the case of square roughness. Due to the airflow being 781
inserted into the triangle roughness (Type 1), the channel has effective cooling. Because 782
of the good curvature of the triangle roughness (Type 2), the temperature field has good 783
agitation in the region. Due to the best heat flow straightening provided by a triangle 784
roughness (Type 3), the temperature field has ideal cooling in this scenario. 785
786
787
788
789
790
791
792
793
Processes 2023, 11, x FOR PEER REVIEW 18 of 22

794
795
796
797
798
799
800
(a) 801
802
803
804
(b)
805
806
807

(c) 808
809
810
811
(d) 812
813

Figure 24. The temperature field in the four cases of roughness: square roughness, triangular 814
roughness (Type 1), triangular roughness (Type 2), and triangular roughness (Type 3) for Re = 2500. 815

For each of the several situations of roughness considered for the air duct with tri- 816
angular fins, Figure 25 shows the fluctuations in values of (Nu/Nu0 )as a function of the 817
Reynolds number. With an increase in Reynolds number comes an increase in the mean 818
Nusselt number, and high velocity produces high convection. 819
Due to the flow's effective guidance, the figure demonstrates that the average 820
Nusselt number is large in the case of triangular roughness (Type 3). Due to the regular 821
separation of the flow, the Nusselt number is decreased in the case of triangular rough- 822
ness (Type 2). Because the attack surface has straight roughness in the case of triangular 823
roughness (Type 1), the Nusselt number is low. Due to inadequate fluid recirculation, the 824
Nusselt number in the case of square roughness is minimal. The exchange surface is low 825
in the absence of roughness, which causes a very noticeable fall in the Nusselt values. 826
827
828
829
830
831
832
833
834
835
836
837
838
839
840

Figure 25. Normalized average Nusselt number (Nu/Nu0 )as a function of the Reynolds number in 841
various roughnesses. 842

For four various types of roughness, including square and triangular roughness, 843
Figure 26 depicts the evolution of the average ratio (f/f0 )as a function of the Reynolds 844
Processes 2023, 11, x FOR PEER REVIEW 19 of 22

number, which ranges between 2500 and 10000 (Types 1, 2, and 3). As Reynolds number 845
rises, the average coefficient of friction rises as well. Because there is a huge contact sur- 846
face in the case of square roughness, the figure demonstrates that the average coefficient 847
of friction has a high value. Moreover, as the triangular roughness (Type 1) is perpen- 848
dicular to the flow direction, the coefficient of friction is lower. Due to the minimal flow 849
disturbance in the case of a triangular roughness (Type 2), the coefficient of friction is 850
lower. The low flow resistance in the case of a triangular roughness (Type 3) lowers the 851
coefficient of friction. When there is no roughness, the coefficient of friction is low be- 852
cause the flow cannot be impeded by the roughness. 853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868

Figure 26. The average coefficient of friction for various numbers of Re in various situations. 869

4. Conclusion 870
In this article, we presented a numerical study of forced convection cooling of a mi- 871
croprocessor in a channel with obstacles and roughness. The main goal of this study is to 872
provide a better understanding of this phenomenon, while also reflecting on more di- 873
verse cooling techniques and getting as close as possible to actual usage conditions. 874
During our study, we used the fluent computer code ANSYS as well as its gambit 875
mesh generator on several cases, which allowed us to become more familiar with nu- 876
merical simulation. The complexity of the geometric configuration has been mastered by 877
gambit, but the shape intended for us is relatively complicated to achieve and to knit, 878
which requires quality and, above all, a precision work. The importance of digital inves- 879
tigation is to shed light on the physical phenomena described by the theory without go- 880
ing through experience, which is more expensive. 881
882
❖ Several parameters are studied, in particular the hydrodynamic (2500 ≤ Re ≤ 10000) 883
and geometric (0 ≤ θ ≤ 45°) parameters. 884
❖ The three types of obstacle: θ = 0°: square obstacle; 0 < θ < 45°: trapezoidal obstacle; 885
and θ = 45°: triangular obstacle, with Re = 2500, generate complex phenomena such 886
as recirculation zones and turbulence regions upstream and downstream of the ob- 887
stacles and a turbulent zone for the model containing triangular obstacles is very 888
important compared to other models. 889
❖ The upstream sharp edge has a point of separation (wall of the obstacle) where the 890
dynamic pressure is very high, which causes a depression downstream of these ob- 891
stacles. 892
❖ For a triangular obstacle the pressure has a significant value, which gave us a tre- 893
mendous movement. 894
❖ The higher the Reynolds number, the greater the dynamic pressure. 895
Processes 2023, 11, x FOR PEER REVIEW 20 of 22

❖ The average velocity in the three cases (square; trapezoidal and triangular obstacle) 896
was very close for the Reynolds values 2000-6000 on the other hand the average ve- 897
locity decreased in the triangular obstacle. 898
❖ In the case of a triangular obstacle, the contact between the cold air and the obstacle 899
is important downstream of the obstacle which gives us good cooling, and the 900
Nusselt number has an important value, because the agitation of the flow gives us 901
more convective heat transfer, the coefficient of friction is low, due to the air flow is 902
uniform. 903
❖ The comparative study between four models of the triangular Type 1, triangular 904
Type 2 and triangular Type 3 canal shows that in the case of a triangular roughness 905
(Type 3) : 906
Average velocity has great value; 907

• The temperature field has perfect cooling, because of the best straighten- 908
ing of the heat flux by this roughness; 909
• The average Nusselt number has a high value; 910
• The coefficient of friction is lowered, because the flow resistance is low. 911

References 912
1. Kumar, A., & Kim, M. H. (2015). Convective heat transfer enhancement in solar air channels. Applied Thermal Engineering, 89, 913
239-261. 914
2. Liu, J., Gao, J., Gao, T., &Shi, X. (2013).Heat transfer characteristics in steam-cooled rectangular channels with two opposite 915
rib-roughened walls.Applied Thermal Engineering, 50(1), 104-111. 916
3. Ali, M., Zeitoun, O., &Nuhait, A. (2011).Forced convection heat transfer over horizontal triangular cylinder in cross flow. 917
International Journal of Thermal Sciences, 50(1), 106-114. 918
4. Kumar, A., Bhagoria, J. L., &Sarviya, R. M. (2009). Heat transfer and friction correlations for artificially roughened solar air 919
heater duct with discrete W-shaped ribs. Energy Conversion and management, 50(8), 2106-2117. 920
5. Karwa, R., Maheshwari, B. K., &Karwa, N. (2005). Experimental study of heat transfer enhancement in an asymmetrically- 921
heated rectangular duct with perforated baffles. International Communications in Heat and Mass Transfer, 32(1-2), 275-284. 922
6. Kabeel, A. E., Hamed, M. H., Omara, Z. M., &Kandeal, A. W. (2018). Influence of fin height on the performance of a glazed and 923
bladed entrance single-pass solar air heater. Solar Energy, 162, 410-419. 924
7. Demartini, L. C., et al., Numeric and Experimental Analysis of The Turbulent Flow Through a Channel with Baffle 925
Plates, Journal of the Brazilian Society of Mechanical Sciences and Engineering, 26 (2004), pp. 153-15 926
8. Cao, X., Du, T., Liu, Z., Zhai, H., & Duan, Z. (2019).Experimental and numerical investigation on heat transfer and fluid flow 927
performance of sextant helical baffle heat exchangers.International Journal of Heat and Mass Transfer, 142, 118437. 928
9. Saedodin, S. A. H. Z. S., Zamzamian, S. A. H., Nimvari, M. E., Wongwises, S., &Jouybari, H. J. (2017). Performance evaluation 929
of a flat-plate solar collector filled with porous metalfoam: Experimental and numerical analysis. Energy Conversion and Man- 930
agement, 153, 278-287. 931
10. Abchouyeh, M. A., Fard, O. S., Mohebbi, R., &Sheremet, M. A. (2019). Enhancement of heat transfer of nanofluids in the pres- 932
ence of sinusoidal side obstacles between two parallel plates through the lattice Boltzmann method. International Journal of 933
Mechanical Sciences, 156, 159-169. 934
11. Chompookham, T., Thianpong, C., Kwankaomeng, S., &Promvonge, P. (2010).Heat transfer augmentation in a wedge-ribbed 935
channel using winglet vortex generators. International Communications in Heat and Mass Transfer, 37(2), 163-169. 936
12. Dutta, P., & Hossain, A. (2005).Internal cooling augmentation in rectangular channel using two inclined baffles. International 937
Journal of Heat and Fluid Flow, 26(2), 223-232. 938
13. Kalaiarasi, G., Velraj, R., &Swami, M. V. (2016).Experimental energy and exergy analysis of a flat plate solar air heater with a 939
new design of integrated sensible heat storage. Energy, 111, 609-619. 940
14. Peng, D., Zhang, X., Dong, H., &Lv, K. (2010).Performance study of a novel solar air collector. Applied Thermal Engineer- 941
ing, 30(16), 2594-2601. 942
15. Selimefendigil, F., Oztop, H. F., Sheremet, M. A., & Abu-Hamdeh, N. (2019).Forced convection of Fe3O4-water nanofluid in a 943
bifurcating channel under the effect of variable magnetic field.Energies, 12(4), 666. 944
16. Kamali, R., &Binesh, A. R. (2008). The importance of rib shape effects on the local heat transfer and flow friction characteristics 945
of square ducts with ribbed internal surfaces. International Communications in Heat and Mass Transfer, 35(8), 1032-1040. 946
17. Siddiqui, M. K. (2007).Heat transfer augmentation in a heat exchanger tube using a baffle. International Journal of Heat and Fluid 947
Flow, 28(2), 318-328. 948
18. Wang, F., Zhang, J., & Wang, S. (2012).Investigation on flow and heat transfer characteristics in rectangular channel with 949
drop-shaped pin fins. Propulsion and power research, 1(1), 64-70. 950
Processes 2023, 11, x FOR PEER REVIEW 21 of 22

19. Eiamsa-ard, S., &Promvonge, P. (2008).Numerical study on heat transfer of turbulent channel flow over periodic 951
grooves. International Communications in Heat and Mass Transfer, 35(7), 844-852. 952
20. Ozceyhan, V., Gunes, S., Buyukalaca, O., &Altuntop, N. (2008). Heat transfer enhancement in a tube using circular cross sec- 953
tional rings separated from wall.Applied energy, 85(10), 988-1001. 954
21. Promvonge, P., Changcharoen, W., Kwankaomeng, S., &Thianpong, C. (2011). Numerical heat transfer study of turbulent 955
square-duct flow through inline V-shaped discrete ribs. International Communications in Heat and Mass Transfer, 38(10), 956
1392-1399. 957
22. Sripattanapipat, S., &Promvonge, P. (2009).Numerical analysis of laminar heat transfer in a channel with diamond-shaped 958
baffles. International Communications in Heat and Mass Transfer, 36(1), 32-38. 959
23. Pirouz, M. M., Farhadi, M., Sedighi, K., Nemati, H., &Fattahi, E. (2011). Lattice Boltzmann simulation of conjugate heat transfer 960
in a rectangular channel with wall-mounted obstacles. ScientiaIranica, 18(2), 213-221. 961
24. Mohammadi, K., &Sabzpooshani, M. (2013). Comprehensive performance evaluation and parametric studies of single pass 962
solar air heater with fins and baffles attached over the absorber plate. Energy, 57, 741-750. 963
25. Skullong, S., Thianpong, C., Jayranaiwachira, N., &Promvonge, P. (2016). Experimental and numerical heat transfer investiga- 964
tion in turbulent square-duct flow through oblique horseshoe baffles. Chemical Engineering and Processing: Process Intensifica- 965
tion, 99, 58-71. 966
26. Priyam, A., & Chand, P. (2016). Thermal and thermohydraulic performance of wavy finned absorber solar air heater. Solar 967
Energy, 130, 250-259. 968
27. Hu, J., Sun, X., Xu, J., & Li, Z. (2013).Numerical analysis of mechanical ventilation solar air collector with internal baf- 969
fles. Energy and Buildings, 62, 230-238. 970
28. Amraoui, M. A., &Aliane, K. (2018).Three-dimensional analysis of air flow in a flat plate solar collector. PeriodicaPolytechnica 971
Mechanical Engineering, 62(2), 126-135. 972
29. Amraoui, M. A. (2020).Numerical Study of an Air Flow in a Flat Plate Air Solar Collector with Circular Obsta- 973
cles.In International Conference in Artificial Intelligence in Renewable Energetic Systems (pp. 839-846). Springer, Cham. 974
30. Amraoui, M. A. (2021).Three-Dimensional Numerical Simulation of a Flat Plate Solar Collector with Double Paths. International 975
Journal of Heat and Technology, 39(4), 1087-1096. 976
31. Amraoui, M. A., &Benosman, F. (2021). Numerical modeling of a flat air solar collector fitted with obstacles. In E3S Web of 977
Conferences (Vol. 321, p. 04016). 978
32. Leonardi, S., Orlandi, P., Djenidi, L., & Antonia, R. A. (2004).Structure of turbulent channel flow with square bars on one 979
wall. International journal of heat and fluid flow, 25(3), 384-392. 980
33. Miyake, Y., Tsujimoto, K., &Nakaji, M. (2001). Direct numerical simulation of rough-wall heat transfer in a turbulent channel 981
flow. International Journal of Heat and Fluid Flow, 22(3), 237-244. 982
34. Ansari, M., &Bazargan, M. (2018). Optimization of flat plate solar air heaters with ribbed surfaces. Applied Thermal Engineer- 983
ing, 136, 356-363. 984
35. Kim, K. Y., & Kim, S. S. (2002). Shape optimization of rib-roughened surface to enhance turbulent heat transfer. International 985
Journal of Heat and Mass Transfer, 45(13), 2719-2727. 986
36. Jang, Y. J., Chen, H. C., & Han, J. C. (2001).Computation of flow and heat transfer in two-pass channels with 60 deg ribs. J. Heat 987
Transfer, 123(3), 563-575. 988
37. Kim, H. M., & Kim, K. Y. (2006). Shape optimization of three-dimensional channel roughened by angled ribs with RANS 989
analysis of turbulent heat transfer.International Journal of heat and mass transfer, 49(21-22), 4013-4022. 990
38. Kumar, R., Kumar, A., &Goel, V. (2017). A parametric analysis of rectangular rib roughened triangular duct solar air heater 991
using computational fluid dynamics. Solar Energy, 157, 1095-1107. 992
39. Kamali, R., &Binesh, A. R. (2008). The importance of rib shape effects on the local heat transfer and flow friction characteristics 993
of square ducts with ribbed internal surfaces. International Communications in Heat and Mass Transfer, 35(8), 1032-1040. 994
40. Ryu, D. N., Choi, D. H., & Patel, V. C. (2007).Analysis of turbulent flow in channels roughened by two-dimensional ribs and 995
three-dimensional blocks. Part II: Heat transfer. International Journal of Heat and Fluid Flow, 28(5), 1112-1124. 996
41. Liou, T. M., Hwang, J. J., & Chen, S. H. (1993). Simulation and measurement of enhanced turbulent heat transfer in a channel 997
with periodic ribs on one principal wall. International Journal of Heat and Mass Transfer, 36(2), 507-517. 998
42. Alam, T., & Kim, M. H. (2017). Heat transfer enhancement in solar air heater duct with conical protrusion roughness ribs. 999
Applied Thermal Engineering, 126, 458 469. 1000
43. Kumar, A., &Layek, A. (2019).Energetic and exergetic performance evaluation of solar air heater with twisted rib roughness on 1001
absorber plate. Journal of Cleaner Production, 232, 617-628. 1002
44. Skullong, S., Promvonge, P., Thianpong, C., Jayranaiwachira, N., &Pimsarn, M. (2017).Heat transfer augmentation in a solar air 1003
heater channel with combined winglets and wavy grooves on absorber plate.Applied Thermal Engineering, 122, 268-284. 1004
45. Deo, N. S., Chander, S., & Saini, J. S. (2016). Performance analysis of solar air heater duct roughened with multigap V-down 1005
ribs combined with staggered ribs. Renewable Energy, 91, 484-500. 1006
46. Gabhane, M. G., &Kanase-Patil, A. B. (2017). Experimental analysis of double flow solar air heater with multiple C shape 1007
roughness. Solar Energy, 155, 1411-1416. 1008
47. Kumar, K., Prajapati, D. R., & Samir, S. (2017). Heat transfer and friction factor correlations development for solar air heater 1009
duct artificially roughened with ‘S’shape ribs. Experimental Thermal and Fluid Science, 82, 249-261. 1010
Processes 2023, 11, x FOR PEER REVIEW 22 of 22

48. Ravi, R. K., & Saini, R. P. (2018).Nusselt number and friction factor correlations for forced convective type counter flow solar 1011
air heater having discrete multi V shaped and staggered rib roughness on both sides of the absorber plate.Applied Thermal En- 1012
gineering, 129, 735-746. 1013
49. B.E. Launder, D.B. Spalding, Computer Methods inApplied Mechanics and Engineering, 3(1974) 269-289 1014

You might also like