A Descriptive Review To Access The Most Suitable Rib's Configuration of Roughness For The Maximum Performance of Solar Air Heater
A Descriptive Review To Access The Most Suitable Rib's Configuration of Roughness For The Maximum Performance of Solar Air Heater
A Descriptive Review To Access The Most Suitable Rib's Configuration of Roughness For The Maximum Performance of Solar Air Heater
Review
A Descriptive Review to Access the Most Suitable Rib’s
Configuration of Roughness for the Maximum Performance of
Solar Air Heater
Karmveer 1, *, Naveen Kumar Gupta 1, *, Tabish Alam 2, *, Raffaello Cozzolino 3, * and Gino Bella 3
1 Department of Mechanical Engineering, GLA University, Mathura 281406, Uttar Pradesh, India
2 CSIR—Central Building Research Institute, Roorkee 247667, Uttarakhand, India
3 Department of Engineering, University of Rome Niccolò Cusano, 00166 Roma, Italy; [email protected]
* Correspondence: [email protected] (K.); [email protected] (N.K.G.); [email protected] (T.A.);
[email protected] (R.C.)
Abstract: Solar air heater is considered to be the most popular and widely used solar thermal system.
Solar air heater (SAH) can be used in many applications, ranging from domestic to industrial purposes.
However, it seems that the viability of SAH is not feasible due to the following two reasons: (i) the
low convective heat transfer coefficient at the absorber plate is the reason that causes a low heat
transfer rate to the flowing air, and (ii) the high temperature of the absorber plate insists on high heat
losses, thus, reducing the thermal efficiency. The convective coefficient can be augmented by placing
turbulators/roughness on the absorber plate, which induces turbulence in the flow passage near the
absorber plate by disrupting and destabilizing the laminar sublayer. This comprehensive review has
been presented to summarize the studies on artificial roughness/turbulators geometries to enhance
the heat transfer rate. Various rib configurations (such as grits, grooves, blockages, baffles, winglets,
protrusions, twisted taps, dimples, and mesh wires) and distinct arrangements of rib roughness
(such as inclined, transverse, V shape, with gap) have been reviewed to present heat transfer and
Citation: Karmveer; Gupta, N.K.;
friction characteristics. Additionally, thermal efficiency and thermohydraulic efficiency (in terms of
Alam, T.; Cozzolino, R.; Bella, G. A
Descriptive Review to Access the
net effective efficiency) of various artificial roughnesses and rib configurations are presented under
Most Suitable Rib’s Configuration of distinct operating conditions for comparing purposes. This comparative study has been presented
Roughness for the Maximum to assess the most desirable ribs and their configurations. On the basis of net effective efficiency,
Performance of Solar Air Heater. a multiarc rib with gaps is found to be the best configuration among all and have the highest thermal
Energies 2022, 15, 2800. https:// and net effective efficiency of around 79%.
doi.org/10.3390/en15082800
heating, drying of vegetables, drying of fruits, timber seasoning, crop drying, space heating,
and many other industrial and domestic purposes. It is free from problems related to
freezing and corrosion. SAH can also be used to increase the efficiency and performance of
a conventional drying system by integrating these systems in many applications, such as
conveyer and fluidized bed drying systems.
SAHs are widely used in the world for heating purposes because they are very simple
in design, economical, and almost maintenance-free. The main components are the collector
plate, transparent cover of high transmissivity, back insulations, side edges, and blower/fan,
as shown in Figure 1. The SAH has poor performance due to the low convective heat
transfer coefficient of the absorber plate and high heat losses from the top glass cover.
Due to the development of the laminar sublayer at the absorber plate because of the
thermal resistance generated near the plate retards’ heat transfer. The convective coefficient
can be augmented by using turbulators, which induce turbulence in the duct near the
absorber plate by disrupting and destabilizing the laminar sublayer. The viscous sub-layer
can be disrupted by using irregular-shape obstacles, called artificial roughness, in distinct
form of grits, groove, baffles, ribs, winglet, protrusions, twisted taps, dimples, perforation,
mesh wire, and so on [1]. The roughness may be in a square, rectangular, triangular,
conical, spherical, or chamfered shape or in a hybrid shape. The turbulence promoters also
increase friction losses, which leads to higher power consumption in the pumping of air,
which is not desirable. Therefore, the turbulence zone should be within the region of the
viscous sublayer.
The objective of this review paper is to summarize the works carried out for the
performance enfacement of SAH, exploiting the artificial roughness/turbulators of differ-
ent configurations. Additionally, a performance analysis of various rib configurations is
presented to assess the best rib configuration. This study will be helpful to readers who are
working in heat transfer enhancement. In the following subsection, the concept of artificial
roughness and their corresponding effect are discussed. Additionally, the performance of
SAH having artificial roughness is discussed in detail.
Energies 2022, 15, 2800 3 of 46
2. Performance of SAH
2.1. Heat Transfer Performance
It can be evaluated with the help of an equation proposed by Bliss [2], which is given as:
qu = FR { I (τ.α) − Uo ( Ti − Ta )} (1)
the point of reattachment of the flow and position of the maximum convection coefficient
is quite close to each other. Cortes and Piacentini [6] developed a performance assess-
ment parameter of the SAH with the help of a numerical model called effective efficiency.
They also showed that the effective efficiency of a collector may be enhanced with the
help of perturbations. The obstacles also increase air pumping power requirements. Niku-
radse [7] studied the turbulent flow of fluids with a distinct degree of relative roughness
and developed the temperature and velocity distributions of a sand-grain-roughened pipe
for a large span of roughness Re (e+ ).
r
+ e f
e = Re (5)
D 2
Roughness Re is a parameter that combines both the roughness height and Re. Niku-
radse [7] identified three regions of fluid flow based on variation in f for a wide span of e+ ,
which are shown in Figure 2.
In the hydrodynamically smooth flow region, the value of f for a small range of
roughness Re (5 > e+ > 0) is the same for a smooth and rough pipe, and the roughness exists
entirely within the laminar sublayer. In the transitionally rough flow region, the magnitude
of f increases with the rise in roughness Re as the value of f is a function of roughness Re
in this range (70 > e+ > 5). The thickness of the laminar sublayer is the same as the height
of the roughness element in this region. Additionally, in the fully rough flow region, the
value of f is not depending on the roughness Re in this range (e+ > 70). The friction factor
follows the quadratic law of resistance. The height of the roughness extends through the
laminar sublayer.
Dipprey and Sabersky [8] developed a wall similarity law that is applicable for both
smooth and rough tubes. Webb et al. [9] developed a correlation of heat transfer and f for a
repeated rib roughness in case of turbulent flow within the tube. The correlation of heat
transfer is based on heat momentum transfer analogy as previously used by Dipprey and
Sabersky [8], and friction correlations are based on the wall similarity law as previously
used by Nikuradse [7] in the case of the sand grain type of artificial roughness.
( f /2St − 1)
g(e+ , Pr) = ue + + p (6)
f /2
The ability of the heat transfer of an artificially rough surface is shown by g(e+ , Pr).
Webb et al. [10] further extended the range of validity of these correlations and pro-
posed that the justified value of e+ is greater than 35. Vilemas and Simonis [11] experi-
mentally investigated friction and convective heat transfer in inner rough tubes for Re
numbers 5000 to 500,000. They also developed correlations for friction and heat transfer as
given below;
For a fully rough flow condition with rectangular roughness in annular channels:
For a partially rough flow condition with rectangular roughness in annular channels
at e/De ≥ 0.0025:
Nu = (0.0053 − 0.14e/De )Re0.95+7e/D Pr0.6 ψn (8)
where √
n = −(0.29 + 0.03e51 (e/De )
Re−24e/De (1 − e−0.16x/De )
The friction factor for channels with rectangular roughness:
Han et al. [12] experimentally investigated the effect of a distinct rib roughness of
various shapes, sizes, and cross sections on heat transfer and f. Both St and f have the best
Energies 2022, 15, 2800 5 of 46
value corresponding to p/e = 10. The heat transfer and f correlation are developed based
on heat momentum transfer analogy and the wall similarity law, as previously used by
Dipprey and Sabersky [8]. They studied four angles of attack between rib roughness and
mainstream flow as 20◦ , 45◦ , 75◦ , and 90◦ . The value of α = 45◦ has superior convective
performance in comparison with others. The correlation for the friction factor:
" #
b
+ + a ◦ 0.35 0.57 10
Re = 4.9(e /35) / ( ϕ/90 ) (α/45) ( ) (10)
p/e
where,
a = −0.4 when e+ < 35 and a = 0 when e+ ≥ 35
b = −0.13 when p/e < 10 and b = 0.53(α/90◦ )0.71 when p/e ≥ 10
where,
i = 0 when e+ < 35 and i = 0.28 when e+ ≥ 35
j = 0.5 when α < 45◦ and j = −0.45 when α ≥ 45◦
Prasad and Saini [13] experimentally studied that for specific values of p/e and e/Dh ,
a geometrically similar roughness has the same effect on the convection coefficient and
f. The convective coefficient decreases, and the value of f increases as e/Dh increases.
Gupta et al. [14] experimentally investigated that as the relative height of artificial rough-
ness escalates, the optimum flow rate shifts to a lower value. They studied that the Stanton
number increases up to Re = 12,000 and decreases after further increment in the value of Re.
orientations. According to Lewis [3], the heat transfer and momentum loss function of
roughness depend on various parameters of artificial roughness and duct.
g = g(e+ , Pr, p/e, e/b, Cd , e/Dh , c/e, α,shape, duct cross − sec tion) (12)
R = R(e+ , p/e, Cd , e/b, e/Dh , c/e, α,shape, duct cross − sec tion) (13)
The effects of distinct types of rib roughness and various parameters investigated by
various researchers are discussed below.
Table 1. Values of e/Dh at which the highest rate of heat transfer for distinct roughness geometries is
investigated in the SAH duct.
the value of p/e is more than 8. The convection coefficient has the highest values near the
reattachment point. The values of p/e at which highest rate of heat transfer was achieved
for distinct roughness geometries are shown in Table 2.
Table 2. Values of p/e at which the highest rate of heat transfer for distinct roughness geometries is
investigated in the SAH duct.
5. Artificial Roughness/Turbulators
Many researchers have investigated distinct types of artificial roughness arrangements,
which have remarkable effects on the performance of the SAH. The distinct artificial
arrangement used in different SAH ducts are discussed as follows:
respectively. Additionally, the maximum effective efficiency is 44.25% and 69.15%, respec-
tively. Prasad and Mullick [58] studied the effect of a small-diameter wire as roughness
in transverse direction to the direction of the flow. The protruding wires as a roughness
element augment the convection coefficient. The collector efficiency factor increases from
0.68 to 0.72 corresponding to the enhancement in thermohydraulic performance equal to
14% at Re = 40,000.
Karwa et al. [59] studied the effect of repeated chamfered ribs (Figure 7) on the
performance of a SAH at Re ranging from 3750 to 16,350. They applied artificial roughness
in transverse direction to the flow from one sidewall to the other end of the duct. Nu and
the friction factor enhanced from 50% to 120% and 80% to 290%, respectively. The thermal
efficiency of the SAH increased from 10% to 40%. Verma and Prasad [60] experimentally
investigated the thermohydraulic performance of a SAH by using small-diameter wires
as roughness elements in the transverse direction with p/e values varying from 10 to
40 and e/Dh values from 001 to 0.03 at Re values of 5000 to 20,000. The value of e+
varied from 8 to 42. They obtained an optimum performance of 71% at e+ value of 24.
Sahu and Bhagoria [20] studied the effects of a transverse broken rib (Figure 8) on the
thermohydraulic performance of the SAH with e/Dh = 0.0338 and a p/e range of 10–30. The
height of the roughness was 15 mm, and values of Re = 3000–12,000. They reported that
the value of Nu is the maximum for p/e = 20 and also studied that the smooth duct shows
better thermal performance at low Re (i.e., below 5000 in comparison with a roughened
duct). The heat transfer was augmented in the range of 1.25 to 1.40.
Behura et al. [62] investigated the effects of a three-side transverse type of roughness
on the thermohydraulic performance of the SAH by using a wire with 20, 22, and 24 SWG
dimensions at Re ranging from 5000 to 13,000. The heat transfer factor increased by 0.3%
to 0.4%.
Energies 2022, 15, 2800 11 of 46
Singh et al. [19,67] analyzed the performance of a roughened duct by using a V-shaped
down discrete rib (Figure 13) type of roughness. The experimental parameters Re, e/Dh,
p/e, α, g/e, and d/x varied in the range of 3000–15,000, 4–12, 0.015–0.043, 30◦ –75◦ , 0.5–2.0,
and 0.20–0.80, respectively. The highest augmentation in friction factor Nu was 3.11 and
3.04 times over the smooth duct, which occurred at α equal to 60◦ . The highest value of
THPP was 2.06. The effective efficiency was enhanced up to 91% as that of the SAH with
a smooth duct. According to Alam et al. [51], the geometrical parameter of roughness
has significant effects on the thermohydraulic performance of the SAH. Singh et al. [68]
further investigated the effect of a V-shaped down rib with a gap on the performance of the
SAH. The highest values of f and Nu were achieved at p/e = 8. The value of THPP ranged
from 1.27 to 1.93. Karwa et al. [45,69] analyzed the effect of a repeated rib in a rectangular
cross-section V-discontinuous and V-discrete arrangement. The relative roughness length
and α varied from 3 to 6 and 45◦ to 60◦ , respectively. The discrete V-shaped ribs with α = 60◦
showed better performance in comparison with discrete V-shaped ribs with α = 45◦ . The V-
down arrangement of ribs showed better performance in comparison with the V-up pattern.
Hans et al. [28] also studied the effect of a continuous multi-V-rib shown in Figure 14.
Maithani and Saini [25] analyzed effect of a V-shaped rib with multiple symmetrical
gaps (Figure 15) on the thermohydraulic performance of the SAH. The thermohydraulic
performance of the SAH was studied for Re 18,000, whereas p/e varied from 6 to 12 with
Energies 2022, 15, 2800 13 of 46
symmetrical gaps in one limb of V-rib varying from 1–5, g/e varied from 1–5, and α varied
from 30◦ to 75◦ , however, e/Dh is kept fixed at 0.043. The thermohydraulic performance
of the SAH strongly depends on the number of symmetrical gaps and g/e in the limbs.
The f and Nu was augmented 3.67 and 3.6 times, respectively. The highest value of Nu
is achieved at three symmetrical gaps and further, incremented in the number of gaps,
degrading the thermal performance. The value of Nu increased up to g/e equal to 4 and
thereafter decreased. Deo et al. [26] analyzed the effect of a multigap V-shaped down-rib
combined with a staggered rib (Figure 16) on the performance of the SAH.
Figure 16. Multiple-gap V-shapes down-rib combined with a staggered rib [26].
The experimental parameter ranges were Re = 4000–12,000; α = 40◦ –80◦ , and p/e = 4–14.
The parameters w/e = 4.5 and g/e = 1 were fixed. The maximum increments in THPP and
Nu were 2.45 and 3.34 times, respectively, over the smooth duct. The best value of THPP
was achieved near a Re of 12,000. Patil et al. [27] analyzed the effect of V-shaped broken
ribs combined with a staggered rib (Figure 17) on the performance of the SAH. The studies
encompassed parameter ranges of Re = 3000–17,000; p’/p = 0.2–0.8; and r/e = 1–2.5 for a
fixed value of e/Dh = 0.043, g/e = 1, α = 60◦ , and p/e = 10. The highest augmentation in
Nu was found up to 3.18 times over the smooth duct. Jain and Lanjewar [70] analyzed the
thermohydraulic performance of the SAH by using a V-shaped rib with symmetrical gaps
combined with staggered ribs. The parameters p/e and Re varied from 10 to 16 and 3000
to 14,000, respectively, for a fixed value of p’/p = 0.65, α = 60◦ , and r/e = 4. The highest
augmentations in the friction factor and Nu were 3.13 and 2.30, respectively, for p/e = 12.
Kumar and Kim [71,72] studied the effect of various ribs in V-form on the performance
of the SAH in the Re range of 5000–20,000. The various V-form ribs include V-rib. They used
parameter values of α = 60◦ and e/Dh = 0.040. They reported that a V-shaped rib combined
with grooves shows the highest value of THPP as that of other investigated V-ribs of
different arrangements. Nidhul et al. [73] used CFD and exergy analysis to study the effect
of a secondary flow in the duct generated due to a V-shaped rib on the thermohydraulic
performance of an SAH duct. Re ranged from 5000 to 20,000. The value of e/Dh = 0.05,
Energies 2022, 15, 2800 15 of 46
and p/e = 10. The reported value of the highest augmentation in Nu was 2.41 times over
the smooth surface at Re = 7500 and α = 45◦ , and the maximum value of f was 2.53 times
over the smooth surface at Re = 17,500 and α = 60◦ . Mishra et al. [74] used CFD analysis
for a V-shaped down rib with multigap and turbulence promoters (Figure 18) to study the
performance of a triangular-duct SAH in the Re range of 4000–20,000. The investigation
covered an α range of 45◦ to 60◦ and a p/e range of 8 to 14. The maximum thermal perfor-
mance was achieved at p/e = 10 and α = 45◦ . The THPP increased as Re increased from
4000 to 10,000, then thereafter decreased in the upper range of Re. Patel and Lanjewar [75]
experimentally and numerically studied the effect of novel V-shaped ribs on the perfor-
mance of the SAH. The study parameter varied as p/e = 6–14 whereas other parameters
viz. p’/p = 4, r/e = 4, g/e = 4, α = 60◦ , e/Dh = 0.043 and Ng = 3 with Reynolds number in
the range of 4000 to 14,500. The highest augmentation took place in Nu = 1.55 to 2.26 and
the friction factor = 2.63 to 3.40 at p/e equaling 10 and 8, respectively, in comparison with a
smooth surface. The highest value of THPP = 1.59 was achieved at p/e = 10 and Re = 12,364.
Further, Patel and Lanjewar [76] analyzed the effect of a V-shaped roughness geometry with
staggered elements using additional gaps in symmetrically arranged elements of roughness
(Figure 19) on the performance of the SAH. The distinct experimental parameters varied as
e/Dh equaled 0.043, g/e equaled 4, p/e equaled 10, p0 /p equaled 0.4, Ng equaled 4, d/w
equaled 0.25 to 0.85, and Re equaled 4000–15,000. The highest value of the THPP parameter
was 1.82 at d/w equaling 0.65 and Re = 12,524. Alam et al. [77–80] experimentally analyzed
the effect of V-shaped perforated blocks (Figure 20) on the performance of the SAH. The
study encompassed parameter ranges of e/H = 0.4 to 1.0, p/e = 4 to 12, α = 60◦ , and
Re = 2000 to 20,000. The highest augmentations in Nu and f were 6.76 and 28.84 over the
smooth duct at e/H = 0.8 and p/e = 8. Further, Alam et al. [81] experimentally studied the
effects of different types of perforation shapes on the performance of the SAH. They used
square, rectangular, and circular types of perforation shapes in a 1–0.6 range of circularity.
The value of α varied from 30◦ to 75◦ . The highest values of Nu and f were achieved at
α = 60◦ . The noncircular shapes of perforation showed a higher thermal performance than
that of circular shapes.
Figure 17. V-shaped broken rib combined with staggered ribs [82].
Energies 2022, 15, 2800 16 of 46
Figure 18. V-shaped rib with multiple gaps and turbulators [74].
The friction factor and Nu were augmented 6.13 and 6.32 times, respectively, in comparison
with a smooth duct. The best value of THPP was achieved at g/e = 1 and d/x = 0.69.
Further, Kumar et al. [84,85] studied the performance of this artificial roughness with the
parameter’s W/w = 6, W/e = 12, e/Dh = 0.0433, and g/e = 1.0. The value of α ranged from
30◦ to 75◦ . They reported that f and Nu are strong functions of α, and also, they have a
maximum value at α = 60◦ . Jin et al. [86–88] numerically analyzed the effect of an inline and
staggered multi-V-shaped rib on the performance of the SAH. The staggered arrangement
had highest enhancement of 26% and 18% in Nu and THPP, respectively, over the inline
arrangement of ribs. The maximum value of THPP was 2.43. Promvonge and Skullon [89]
studied the effect of V-shaped flap-baffle and chamfered-grove vortex generators (VG)
on the performance of a roughened duct, as shown in Figure 22. Both the flap baffle and
VG were at α equal to 45◦ , and the experiment was performed in both the apex-up and
apex-down pattern of a V-shaped flap baffle in the range of Re = 5290–22,600. The apex-up
pattern had better performance in comparison with the apex-down pattern of V-shaped
flap baffles. Nu and the friction factor were enhanced remarkably by using this type of
roughness. They reported that the maximum value of TEF was 2.68 at Re = 5290.
The augmentations in f and Nu were reported to be 2.44 and 2.63 times as that of
the smooth duct. The corresponding values over the continuous arc-shaped rib were
1.19 and 1.14 times, respectively. Ghritlahre et al. [93] studied the performance of an arc-
shaped artificially roughened duct with an apex-down and apex-up flow of air. They took
a values of p/e equal to 10, arc angle = 60◦ , and e/Dh = 0.0395. The highest efficiency
of the apex-up pattern was achieved at 73.2%, and for the apex-down pattern, it was
69.2%. The apex-up pattern performed 33.2% better in comparison with the apex-down
arrangement. The thermal performance of the apex-up pattern was better as compared with
the apex-down pattern of an arc-shaped artificially roughened duct. Yadav and Prasad [94]
theoretically studied the effect of arc-shaped wire roughness on the performance of a
parallel-flow SAH. The thermal efficiency of a roughened parallel-flow SAH was 8% to
10% higher in comparison with a smooth SAH. Ambade and Lanjewar [95] experimentally
studied the effect of a symmetrical gap with arc-shaped roughness and a staggered element
on the performance of the SAH at Re equal to 3000–15,000 and p/e equal to 6–14. The fixed
parameters were p’/p = 3, g/e = 4, r/e equal to 4, Ng equal to 3, and α’ equal to 30◦ . They
compared this geometry with the smooth duct and the duct having broken-arc-shaped rib
roughness with staggered elements. The arc-shaped roughness with a new symmetrical gap
augmented the friction factor and Nu up to 4.15 and 2.04 times, respectively, over the broken-
arc-shaped roughness with a staggered element, while augmentation in Nusselt number
and friction factor were 2.18 and 3.88 times, respectively with corresponding smooth duct.
Azad et al. [96] investigated the effect of a discrete-symmetrical arc type of rib roughness
on the performance of the SAH. The values of the experimental parameters covered in the
study were e/Dh = 0.045, g/e = 2–5, Ng = 3, p/e = 10, α’ = 30◦ , and Re = 3000 to 14,000.
The value of g/e had a remarkable effect on the performance. The highest enhancement
reported in Nu was equal to 3.88 over the smooth duct at g/e = 4. The value of THPP
ranged from 1.4 to 1.68 at g/e = 4, and the best value of THPP was 1.68 at Re equal to
14,000. Gill et al. [97] analyzed the effect of staggered broken-arc hybrid-rib roughness
on the thermohydraulic performance of the SAH. The parameters of the study ranged as
e/Dh = 0.022–0.043, α’ = 15–75◦ , p/e = 4–12, and Re = 2000–16,000. The f and Nu were
augmented 2.57 and 3.16 times over the broken-arc roughened duct. The highest value of
THPP was reported to be 2.33. Sureandhar et al. [98] studied the effect of arc-rib fin-type
roughness on the performance of the SAH. The experimental parameters of the study
varied as e/Dh = 0.04222–0.0541, α/90 = 0.3333, and p/e = 10. Nu and THPP increased as
the mass flow rate increased, while the friction factor decreased.
a multiarc dimple-shaped roughness (Figure 26) on the performance of the SAH. They
covered experimental parameters of p/e equal to 4–16, W/w equal to 1–5, e/Dh equal
to 0.018–0.036, α’ equal to 30◦ –75◦ , and Re equal to 2000–18,000. They reported that the
highest value of Nu was 3.19 to 5.56 times that of the smooth duct at p/e equal to 12. The
value of Nu increased up to e/Dh equal to 0.036; after that, it decreased. The enhancement
in the friction factor equaled 1.36 to 2.27 times in comparison with the smooth duct at e/Dh
equal to 0.045. Agrawal et al. [104] experimentally investigated the effect of a double-arc
reverse rib with even gaps on the performance of a solar collector. The values of the fixed
parameters were e/Dh = 0.027, W/H = 8, and I = 1000 W/m2. The value of the variable
parameter varied as p/e equal to 10, α equal to 30◦ –75◦ , and Re equal to 3000–14,000. The
maximum augmentations in f and Nu were found to be 2.42 and 2.85 times over the smooth
surface. The highest value of THPP equal to 2.41 was noticed at p/e = 10 and e/Dh = 0.0270.
as shown in Figure 28. The effect of L-shaped roughness was studied in the ranges of
parameters of p/e = 7.14–17.86 and Re = 3800–18,000; however, the value of e/Dh was fixed
at 0.042. The highest enhancement in Nu was found up to 2.827 times as that of the smooth
duct at p/e = 7.14 and Re = 15,000. The highest augmentation in the friction factor was
found up to 3.424 times over the smooth duct at e/Dh = 0.042, Re = 3800, and p/e = 7.14.
THPP was found to be in the range of 1.92 to 1.90 by using this repeated roughness.
Figure 29. Arc-shaped wire ribs arranged in an ‘S’ type of pattern [40].
of the SAH with spherical dimple-shaped roughness at the absorber plate at Re vary-
ing from 1900 to 6000. The maximum value of convection coefficient was 20.23 W/m2 K,
and the instantaneous thermal efficiency was 23.45–35.50% higher in comparison with a
smooth duct.
perforated and nonperforated surfaces of the absorber. The experimental parameters varied
as p/g = 3.4–3.8 and Re = 3000–27,000. The secondary flow generated along the surface of
the fin and the mixing of the secondary flow developed in the duct with the mainstream of
flow enhanced the level of turbulence remarkably. In the case of perforated surface, the
friction factor Nu was augmented 5.34 and 2.67 times as that of the smooth duct at p/g = 3.8.
In the case of a nonperforated surface, the friction factor and Nu were augmented 5.93 and
2.61 times over the smooth duct.
Figure 40. V-shaped rib corrugated surface integrated with twisted tape [134].
augmentations in f and Nu were 6.93 and 2.05, respectively, at α = 75◦ for a staggered
pattern. Xiao et al. [139] numerically investigated the effect of inclined trapezoid-shape
turbulators on the thermohydraulic performance of the SAH. Nu increased significantly,
and energy efficiency was augmented by 24% and exergy efficiency was augmented by
31% over the smooth duct.
Table 3. Values of α at which the highest heat transfer rate for distinct geometries of roughness is
investigated in the SAH duct.
Figure 41. Transverse rib with square wave type of profile [42].
Manjunath et al. [140] numerically investigated the effect of a sinusoidal profile type of
a duct surface on the performance of the SAH. Re varied from 4000 to 24,000. The sinusoidal
type of the surface enhanced the level of turbulence significantly, which led to a remarkable
increment in convection coefficient. The thermal efficiency increased up to 12.5% over the
smooth surface at the aspect ratio of the duct equaling 1.5. Bezbaruah et al. [141] studied
the effect of a conical vortex generator type of artificial roughness on the performance
of the SAH. They used experimental parameters as e/Dh = 0.17–0.34, Re = 3000–16,000,
and p/e = 8–15. The highest value of Nu = 142.4 was achieved at e/Dh = 0.34, p/e = 8,
and Re = 16,000 as that of a smooth duct. The highest value of f equaled 0.167 reported
at p/e = 8, e/Dh = 0.34, and Re = 3500. The heat transfer performance was augmented
by 192.2%. Dong et al. [142] numerically investigated the effect of an incline-grove ripple
type of roughness on the performance of the SAH at Re in the range of 12,000–24,000.
The incline-grove on the ripple surface enhanced the level of turbulence remarkably. The
optimum value of α was near 45◦ to 60◦ .
Energies 2022, 15, 2800 29 of 46
Nu was augmented 1.04 to 1.946 times over the smooth surface. Alam et al. [143,144]
numerically investigated the effects of conical protrusions on the thermohydraulic per-
formance of the SAH. The study parameter varied as e/Dh = 0.02 to 0.044 and p/e = 6 to
12. The conical protrusion type of artificial rib roughness had a remarkable effect on the
effective efficiency of the SAH. They reported the maximum value of effective efficiency
as 70.92% at e/Dh = 0.0289 and p/e = 10. Kumar and Goel [145,146] analyzed the effects
of a distinct type of rib roughness on the thermohydraulic performance of the SAH by
using a triangular cross-section channel. The performance of the SAH strongly depended
on the cross section of artificial roughness and also the cross section of a flow passage. A
rectangular cross-section rib with forwarding chamfering showed the highest THPP at 2.75.
Further, Goel et al. [147] analyzed the effect of the hemispherical dimple cavity type of
roughness on the performance of the SAH by using a triangular cross-section channel. The
leading edge of the dimple-cavity-type roughness showed lower heat transfer than that
of a trailing edge. The highest augmentation value of Nu equal to 5.33 was achieved at
Re = 2160. The value of THPP was equal to 3.48. Xi et al. [148] numerically studied a ribbed
channel for Re ranging from 10,000 to 90,000. The study parameter ranged as e/D varying
from 0.05 to 0.15 and rib angle varying from 30◦ to 90◦ .
Table 4. Cont.
Table 4. Cont.
Table 5. Cont.
Table 5. Cont.
× exp −0.28{ln(α/60)}2
Nu = 0.105Reh0.873 (α/60)−0.081 (e/D
i h)
0.453
× exp −0.77{ln(W/w)}2
−3.9
Nu = 2.1h× 10−88 (Re)1.452 i(S/e)12.94
h (d/D ) ( L/ei)99.2
× exp −77.2{ln( L/e}2 × exp −1.4{ln(S/e)}2
Bhushan and Singh [41] (2011) Protrusions h i
× exp −7.83{ln(d/D }2
f = 2.32(Re)−0.201 (S/e)−0.383 ( L/e)−0.484 (d/D
h )
0.133
i
Nu = 0.009016(e/Dh )−3.1354 × Re0.526 exp −0.5834{ln(e/Dh )}2
Patel et al. [118] (2020) NACA 0040 profile rib h i
f = 0.32449Re1.3728 × (e/Dh )5.6236 × exp 0.943{ln(e/Dh )}2
h i
× exp −0.0875{ln(Re)}2
Gabhane and Patil [119] Nu = 0.20627(Re)0.8087 (α/90)0.2735 ( p/e)−0.03724
Multi-C-shape rib
(2017) f = 0.9123(Re)−0.28379 (α/90)−0.12127 ( p/e)−0.14847
Energies 2022, 15, 2800 34 of 46
Table 5. Cont.
Solar energy I is absorbed by the collector plate, which is further transferred to the
working fluid as heat loss from the bottom, useful heat gain, heat loss from the top cover,
and heat loss from the side edges. The energy loses to the surroundings by convection and
radiation from the top glass cover. The ambient and absorber plate temperatures are Ta
and Tp, respectively. The step-by-step procedure for calculating the thermal efficiency is
given below.
Heat exchange between the top glass cover and absorber plate:
σ(T4 −T4 )
qloss,p1−c1 = hc,c1− p Tp − Tc1 + 1 p1 c1
ε c ε p −1
(14)
= hc,p−c1 + hr,p−c1 Tp − Tc1
where
σ Tp2 + Tc1
2 Tp + T
hr,p−c1 = 1 1
(15)
εc + εp −1
Thermal resistance:
1
R1 =
(hc,c1− p +hr,c1− p ) (16)
Convection coefficient of heat transfer between the top glass cover and the absorber plate:
Nu.k
hc,p−c1 = (17)
L
where the Nusselt number between the glass cover and the absorber plate is given as [154]:
h i
Nu = 1 + + ( RaCosβ/5830)0.33 − 1
n o
+ 1.44[1 − 1708/RaCosβ]+ 1 − 1708(sin1.8.β)1.6 /Racosβ (18)
1
R1 = (21)
(hr,c1−a + hw )
The overall heat loss must be equal to the energy exchange between the plates:
The edge and back heat loss coefficient are calculated as:
1 k 1 (U A)edge
Ub = = ins
0
and Ue = = (25)
( R4 ) L ( R3 ) Ac
Uo = Ut + Ub + Ue (26)
The thermal efficiency of the SAH increases significantly with the increase in Re for all
shapes and sizes of artificial roughness geometry. The thermal efficiencies of a distinct type
of roughness with respect to Re and ∆T/I are shown in Figures 45 and 46, respectively. The
value of thermal efficiency changes from 0.27 to 0.79. The least values of η are recognized for
a twisted tape and delta-shaped vortex generator type of roughness and the highest value
in the case of a staggered broken-arc hybrid rib. The multiple V-rib with gaps, continuous
multiribs, and multiarc ribs with a gap also show higher values of η.
Energies 2022, 15, 2800 38 of 46
Figure 46. Thermal efficiency vs. ∆T/I of different types of roughened surfaces [19,24,26,28,32,33,36,
38,39,83,85,97,108,110,118,123,127,152].
9. Conclusions
An attempt has been made to study the thermal and friction characteristics of various
artificial ribs/turbulators exploited in SAH ducts. The correlations of Nu and the friction
factor for various rib configurations have been presented, and the thermohydraulic perfor-
mance parameters of various rib configurations have been compared for a similar range of
Reynolds number. Based on the literature review carried out in this paper, the following
conclusions have been drawn.
Energies 2022, 15, 2800 39 of 46
• The shape and size of artificial roughness and their pattern of arrangements on the duct
surface are the most important factors for the performance optimization of the SAH.
• The thermohydraulic characteristics of a large number of rib geometries have been
investigated by many researchers. For most of the rib roughness geometries, the
optimum performance has been achieved at the following parameters: p/e = 10,
W/w = 6, α = 60◦ , and e/Dh = 0.043.
• THPP and thermal efficiency show the highest values in the case of staggered broken-
arc type of hybrid rib and least values in the case of metal grit, twisted tape, and
delta-shaped vortex generator type of roughness.
• The multi-V and multiarc-shaped roughnesses show higher thermohydraulic per-
formance over other roughness geometries. The introducing gaps in the limb of
multi-V-ribs enhance the level of turbulence significantly.
• The multi-V-shaped ribs show a higher value of the friction factor, and arc-shaped
circular dimples show a lower value of the friction factor.
• The broken-arc-shaped rib combined with a staggered-arc rib piece has better perfor-
mance than broken-arc-shaped and arc-shaped rib roughness.
• The creation of gaps in the continuous ribs has shown remarkable improvement in
thermohydraulic performance over the continuous ribs. The improvement in Nu
due to the creation of gaps in the continuous ribs ranges from 1.1 to 1.3 times, and
corresponding increase in pumping power requirement ranges from 1.0 to 1.4 times.
• THPP shows higher values in the case of an S-shape rib, multi-V ribs, and arc-shaped
roughnesses with gaps. However, the performance of an S-shaped rib is not consider-
able at low Re, but the performance increases remarkably with the increase in Re.
• The arc arrangement of rib roughness shows lower value pressure losses over the
V-shaped arrangement due to the curved nature of the induced secondary flow along
with the roughness.
• In general, higher roughnesses’ height has a higher Nusselt number; however, higher
roughnesses’ height contributes to higher pressure drop. Therefore, the thermohy-
draulic performance of roughnesses needs to be optimized. In this regard, net effective
efficiency is the best tool to analyze roughnesses. On the basis of net effective efficiency,
a multiarc rib with gaps is found to be best around 79% in comparison with other rib
configurations, which is recommended for overall better performance.
The work presented in this review paper was carried out as a convective heat transfer
coefficient from the absorber plate. Apart from performance improvement by heat transfer
enhancement, there is a tremendous scope to further increase the performance of the
SAH by exploiting more advanced active methods, such as electrodynamics, jet spray,
mechanical aid, surface vibration, and fluid vibration. Further, heat loss through a top glass
cover needs to be minimized by optimizing the natural convection between the glass cover
and the absorber or utilizing the vacuum between them. Additionally, the flow structure
should be studied using flow visualization, such as PIV, LCT, and CFD, to achieve the
optimum rib arrangement.
Author Contributions: Conceptualization, K., T.A. and N.K.G.; methodology, K. and T.A.; software,
K. and T.A.; validation, K. and T.A.; formal analysis, N.K.G. and R.C.; investigation, K. and T.A.;
resources, R.C. and G.B.; data curation, K.; writing—original draft preparation, K. and T.A.; writing—
review and editing K., T.A., N.K.G., R.C. and G.B.; visualization, K., T.A. and N.K.G.; supervision,
T.A., R.C. and G.B.; project administration, T.A., R.C. and G.B. All authors have read and agreed to
the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data are not publicly available due to privacy considerations.
Energies 2022, 15, 2800 40 of 46
Nomenclature
P Density kg/m3
ψ Temperature factor Tw /Tf -
α Angle of attack degree
α’ Arc angle degree
α/90 Relative arc angle -
ϕ Chamfering angle of rib degree
β Slope degree
β’ Thermal expansion coefficient of air 1/K
εg Glass cover emissivity
εp Absorber plate emissivity
υ Kinematic viscosity m2 /s
τ Transmissivity
σ Stefan–Boltzmann constant W/m2 ·K4
References
1. Gupta, N.K.; Karmveer; Alam, T. A Review on Augmentation in Thermal Performance of Solar Air Heater. IOP Conf. Series Mater.
Sci. Eng. 2021, 1116, 012064. [CrossRef]
2. Bliss, R.W. The derivations of several “Plate-efficiency factors” useful in the design of flat-plate solar heat collectors. Sol. Energy
1959, 3, 55–64. [CrossRef]
3. Lewis, M.J. Optimizing the thermo-hydraulic performance of rough surfaces. Int. J. Heat Mass Transf. 1975, 18, 1243–1248.
[CrossRef]
4. Okamoto, S.; Seo, S.; Nakaso, K.; Kawai, I. Turbulent shear flow and heat transfer over the repeated two-dimensional square ribs
on ground plane. J. Fluids Eng. Trans ASME 1993, 115, 631–637. [CrossRef]
5. Sparrow, E.M.; Kang, S.S.; Chuck, W. Relation between the points of flow reattachment and maximum heat transfer for regions of
flow separation. Int. J. Heat Mass Transf. 1987, 30, 1237–1246. [CrossRef]
6. Cortés, A.; Piacentini, R. Improvement of the efficiency of a bare solar collector by means of turbulence promoters. Appl. Energy
1990, 36, 253–261. [CrossRef]
7. Nikuradse, J. Laws of Flow in Rough Pipes. J. Appl. Phys. 1950, 3, 399.
8. Dipprey, D.F.; Sabersky, R.H. Heat and momentum transfer in smooth and rough tubes at various prandtl numbers. Int. J. Heat
Mass Transf. 1963, 6, 329–353. [CrossRef]
9. Webb, R.L.; Eckert, E.R.G.; Goldstein, R.J. Heat transfer and friction in tubes with repeated-rib roughness. Int. J. Heat Mass Transf.
1971, 14, 601–617. [CrossRef]
10. Webb, R.L.; Eckert, E.R.G.; Goldstein, R.J. Generalized heat transfer and friction correlations for tubes with repeated-rib roughness.
Int. J. Heat Mass Transf. 1972, 15, 180–184. [CrossRef]
11. Vilemas, J.V.; Šimonis, V.M. Heat transfer and friction of rough ducts carrying gas flow with variable physical properties. Int. J.
Heat Mass Transf. 1985, 28, 59–68. [CrossRef]
12. Han, J.C.; Glicksman, L.R.; Rohsenow, W.M. An investigation of heat transfer and friction for rib-roughened surfaces. Int. J. Heat
Mass Transf. 1978, 21, 1143–1156. [CrossRef]
13. Prasad, B.N.; Saini, J.S. Effect of artificial roughness on heat transfer and friction factor in a solar air heater. Sol. Energy 1988, 41,
555–560. [CrossRef]
14. Gupta, D.; Solanki, C.; Saini, J.S. Heat and fluid flow in rectangular solar air heater ducts having transverse rib roughness on
absorber plates. Sol. Energy 1993, 51, 31–37. [CrossRef]
15. Boulemtafes-Boukadoum, A.; Absi, R.; El Abbassi, I.; Darcherif, M.; Benzaoui, A. Numerical investigation of absorber’s roughness
effect on heat transfer in upward solar air heaters. Energy Procedia 2019, 157, 1089–1100. [CrossRef]
16. Singh, S.; Singh, B.; Hans, V.S.; Gill, R.S. CFD (computational fluid dynamics) investigation on Nusselt number and friction factor
of solar air heater duct roughened with non-uniform cross-section transverse rib. Energy 2015, 84, 509–517. [CrossRef]
17. Aharwal, K.R.; Gandhi, B.K.; Saini, J.S. Heat transfer and friction characteristics of solar air heater ducts having integral inclined
discrete ribs on absorber plate. Int. J. Heat Mass Transf. 2009, 52, 5970–5977. [CrossRef]
18. Singh, S.; Chander, S.; Saini, J.S. Investigations on thermo-hydraulic performance due to flow-attack-angle in V-down rib with
gap in a rectangular duct of solar air heater. Appl. Energy 2012, 97, 907–912. [CrossRef]
19. Singh, S.; Chander, S.; Saini, J.S. Heat transfer and friction factor correlations of solar air heater ducts artificially roughened with
discrete V-down ribs. Energy 2011, 36, 5053–5064. [CrossRef]
20. Sahu, M.M.; Bhagoria, J.L. Augmentation of heat transfer coefficient by using 90◦ broken transverse ribs on absorber plate of
solar air heater. Renew. Energy 2005, 30, 2057–2073. [CrossRef]
21. Yadav, A.S.; Bhagoria, J.L. A CFD based thermo-hydraulic performance analysis of an artificially roughened solar air heater having
equilateral triangular sectioned rib roughness on the absorber plate. Int. J. Heat Mass Transf. 2014, 70, 1016–1039. [CrossRef]
22. Yadav, A.S.; Bhagoria, J.L. A numerical investigation of square sectioned transverse rib roughened solar air heater. Int. J. Therm.
Sci. 2014, 79, 111–131. [CrossRef]
Energies 2022, 15, 2800 42 of 46
23. Gupta, D.; Solanki, S.C.; Saini, J.S. Thermohydraulic performance of solar air heaters with roughened absorber plates. Sol. Energy
1997, 61, 33–42. [CrossRef]
24. Ebrahim-Momin, A.M.; Saini, J.S.; Solanki, S.C. Heat transfer and friction in solar air heater duct with V-shaped rib roughness on
absorber plate. Int. J. Heat Mass Transf. 2002, 45, 3383–3396. [CrossRef]
25. Maithani, R.; Saini, J.S. Heat transfer and friction factor correlations for a solar air heater duct roughened artificially with V-ribs
with symmetrical gaps. Exp. Therm. Fluid Sci. 2016, 70, 220–227. [CrossRef]
26. Deo, N.S.; Chander, S.; Saini, J.S. Performance analysis of solar air heater duct roughened with multigap V-down ribs combined
with staggered ribs. Renew. Energy 2016, 91, 484–500. [CrossRef]
27. Patil, A.K.; Saini, J.S.; Kumar, K. Heat transfer and friction characteristics of solar air heater duct roughened by broken V-shape
ribs combined with staggered rib piece. J. Renew. Sustain. Energy 2012, 4, 013115. [CrossRef]
28. Hans, V.S.; Saini, R.P.; Saini, J.S. Heat transfer and friction factor correlations for a solar air heater duct roughened artificially with
multiple v-ribs. Sol. Energy 2010, 84, 898–911. [CrossRef]
29. Kumar, A.; Saini, R.P.; Saini, J.S. Experimental investigation on heat transfer and fluid flow characteristics of air flow in a
rectangular duct with Multi V-shaped rib with gap roughness on the heated plate. Sol. Energy 2012, 86, 1733–1749. [CrossRef]
30. Saini, S.K.; Saini, R.P. Development of correlations for Nusselt number and friction factor for solar air heater with roughened duct
having arc-shaped wire as artificial roughness. Sol. Energy 2008, 82, 1118–1130. [CrossRef]
31. Sethi, M.; Thakur, N.S. Correlations for solar air heater duct with dimpled shape roughness elements on absorber plate. Sol.
Energy 2012, 86, 2852–2861. [CrossRef]
32. Yadav, S.; Kaushal, M.; Varun; Siddhartha. Nusselt number and friction factor correlations for solar air heater duct having
protrusions as roughness elements on absorber plate. Exp. Therm. Fluid Sci. 2013, 44, 34–41. [CrossRef]
33. Hans, V.S.; Gill, R.S.; Singh, S. Heat transfer and friction factor correlations for a solar air heater duct roughened artificially with
broken arc ribs. Exp. Therm. Fluid Sci. 2016, 80, 77–89. [CrossRef]
34. Pandey, N.K.; Bajpai, V.K. Experimental investigation of heat transfer augmentation using multiple arcs with gap on absorber
plate of solar air heater. Sol. Energy 2016, 134, 314–326. [CrossRef]
35. Singh, A.P.; Varun; Siddhartha. Effect of artificial roughness on heat transfer and friction characteristics having multiple arc
shaped roughness element on the absorber plate. Sol. Energy 2014, 105, 479–493. [CrossRef]
36. Lanjewar, A.M.; Bhagoria, J.L.; Sarviya, R.M.; Lanjewar, A.M.; Bhagoria, L. Performance analysis of W-shaped rib roughened
solar air heater. J. Renew. Sustain. Energy 2011, 3, 043110. [CrossRef]
37. Lanjewar, A.; Bhagoria, J.L.; Sarviya, R.M. Experimental study of augmented heat transfer and friction in solar air heater with
different orientations of W-Rib roughness. Exp. Therm. Fluid Sci. 2011, 35, 986–995. [CrossRef]
38. Kumar, A.; Bhagoria, J.L.; Sarviya, R.M. Heat transfer and friction correlations for artificially roughened solar air heater duct with
discrete W-shaped ribs. Energy Convers. Manag. 2009, 50, 2106–2117. [CrossRef]
39. Gawande, V.B.; Dhoble, A.S.; Zodpe, D.B.; Chamoli, S. Experimental and CFD investigation of convection heat transfer in solar
air heater with reverse L-shaped ribs. Sol. Energy 2016, 131, 275–295. [CrossRef]
40. Kumar, K.; Prajapati, D.R.; Samir, S. Heat Transfer and Friction Factor Correlations Development for Solar Air artificially
roughened with ‘S’shape ribs. Heater Duct. Exp. Therm. Fluid Sci. 2016, 82, 249–261. [CrossRef]
41. Bhushan, B.; Singh, R. Nusselt number and friction factor correlations for solar air heater duct having artificially roughened
absorber plate. Sol. Energy 2011, 85, 1109–1118. [CrossRef]
42. Singh, I.; Singh, S. CFD analysis of solar air heater duct having square wave profiled transverse ribs as roughness elements. Sol.
Energy 2018, 162, 442–453. [CrossRef]
43. Saini, R.P.; Saini, J.S. Heat transfer and friction factor correlations for artificially roughened ducts with expanded metal mesh as
roughness element. Int. J. Heat Mass Transf. 1997, 40, 973–986. [CrossRef]
44. Karmare, S.V.; Tikekar, A.N. Heat transfer and friction factor correlation for artificially roughened duct with metal grit ribs. Int. J.
Heat Mass Transf. 2007, 50, 4342–4351. [CrossRef]
45. Karwa, R. 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. Commun. Heat Mass Transf. 2003, 30, 241–250.
[CrossRef]
46. Saini, R.P.; Verma, J. Heat transfer and friction factor correlations for a duct having dimple-shape artificial roughness for solar air
heaters. Energy 2008, 33, 1277–1287. [CrossRef]
47. Bhagoria, J.L.; Saini, J.S.; Solanki, S.C. Heat transfer coefficient and friction factor correlations for rectangular solar air heater duct
having transverse wedge shaped rib roughness on the absorber plate. Renew. Energy 2002, 25, 341–369. [CrossRef]
48. Layek, A.; Saini, J.S.; Solanki, S.C. Heat transfer and friction characteristics for artificially roughened ducts with compound
turbulators. Int. J. Heat Mass Transf. 2007, 50, 4845–4854. [CrossRef]
49. Ahn, S.W. The effects of roughness types on friction factors and heat transfer in roughened rectangular duct. Int. Commun. Heat
Mass Transf. 2001, 28, 933–942. [CrossRef]
50. Taslim, M.E.; Li, T.; Kercher, D.M. Experimental heat transfer and friction in channels roughened with angled, V-shaped and
discrete ribs on two opposite walls. ASME J. Turbomach. 1996, 118, 20–28. [CrossRef]
51. Alam, T.; Saini, R.P.; Saini, J.S. Use of turbulators for heat transfer augmentation in an air duct-A review. Renew. Energy 2014, 62,
689–715. [CrossRef]
Energies 2022, 15, 2800 43 of 46
52. Sahu, R.K.; Gandhi, B.K. Numerical Simulation of Heat Transfer Enhancement due to a Gap in an Inclined Continuous Rib
Arrangement in a Solar Air Heater Duct. Int. J. Adv. Mech. Eng. 2014, 4, 687–693.
53. Aharwal, K.R.; Gandhi, B.K.; Saini, J.S. An experimental investigation of heat transfer and fluid flow in a rectangular duct with
inclined discrete ribs. Int. J. Energy Environ. 2010, 1, 987–998.
54. Xi, L.; Xu, L.; Gao, J.; Zhao, Z.; Li, Y. Study on heat transfer performance of steam-cooled ribbed channel using neural networks
and genetic algorithms. Int. J. Heat Mass Transf. 2018, 127, 1110–1123. [CrossRef]
55. Sherry, M.; Lo Jacono, D.; Sheridan, J. An experimental investigation of the recirculation zone formed downstream of a forward-
facing step. J. Wind Eng. Ind. Aerodyn. 2010, 98, 888–894. [CrossRef]
56. Essel, E.E.; Nematollahi, A.; Thacher, E.W.; Tachie, M.F. Effects of upstream roughness and Reynolds number on separated and
reattached turbulent flow. J. Turbul. 2015, 16, 872–899. [CrossRef]
57. Singh, I.; Vardhan, S.; Singh, S.; Singh, A. Experimental and CFD analysis of solar air heater duct roughened with multiple broken
transverse ribs: A comparative study. Sol. Energy 2019, 188, 519–532. [CrossRef]
58. Prasad, K.; Mullick, S.C. Heat transfer characteristics of a solar air heater used for drying purposes. Appl. Energy 1983, 13, 83–93.
[CrossRef]
59. Karwa, R.; Solanki, S.C.; Saini, J.S. Thermo-hydraulic performance of solar air heaters having integral chamfered rib roughness
on absorber plates. Energy 2001, 26, 161–176. [CrossRef]
60. Verma, S.K.; Prasad, B.N. Investigation for the optimal thermo-hydraulic performance of artificially roughened solar air heaters.
Renew. Energy 2000, 20, 19–36. [CrossRef]
61. Xi, L.; Xu, L.; Gao, J.; Zhao, Z.; Li, Y. Numerical analysis and optimization on flow and heat transfer performance of a steam-cooled
ribbed channel. Case Stud. Therm. Eng. 2021, 28, 101442. [CrossRef]
62. Behura, A.K.; Rout, S.K.; Pandya, H.; Kumar, A. Thermal Analysis of Three Sides Artificially Roughened Solar Air Heaters.
Energy Procedia 2017, 109, 279–285. [CrossRef]
63. Aharwal, K.R.; Gandhi, B.K.; Saini, J.S. Experimental investigation on heat-transfer enhancement due to a gap in an inclined
continuous rib arrangement in a rectangular duct of solar air heater. Renew. Energy 2008, 33, 585–596. [CrossRef]
64. Lu, B.; Jiang, P.X. Experimental and numerical investigation of convection heat transfer in a rectangular channel with angled ribs.
Exp. Therm. Fluid Sci. 2006, 30, 513–521. [CrossRef]
65. Alam, T.; Kim, M.-H. A critical review on artificial roughness provided in rectangular solar air heater duct. Renew. Sustain. Energy
Rev. 2017, 69, 387–400. [CrossRef]
66. Patil, A.K. Heat transfer mechanism and energy efficiency of artificially roughened solar air heaters—A review. Renew. Sustain.
Energy Rev. 2015, 42, 681–689. [CrossRef]
67. Singh, S.; Chander, S.; Saini, J.S. Thermal and effective efficiency-based analysis of discrete V-down rib-roughened solar air
heaters. J. Renew. Sustain. Energy 2011, 3, 023107. [CrossRef]
68. Singh, S.; Chander, S.; Saini, J.S. Thermo-hydraulic performance due to relative roughness pitch in V-down rib with gap in solar
air heater duct—Comparison with similar rib roughness geometries. Renew. Sustain. Energy Rev. 2015, 43, 1159–1166. [CrossRef]
69. Karwa, R.; Bairwa, R.D.; Jain, B.P.; Karwa, N. Experimental study of the effects of rib angle and discretization on heat transfer and
friction in an asymmetrically heated rectangular duct. J. Enhanc. Heat Transf. 2005, 12, 343–355. [CrossRef]
70. Jain, P.K.; Lanjewar, A. Overview of V-RIB geometries in solar air heater and performance evaluation of a new V-RIB geometry.
Renew. Energy 2019, 133, 77–90. [CrossRef]
71. Kumar, A.; Kim, M.H. Heat transfer and fluid flow characteristics in air duct with various V-pattern rib roughness on the heated
plate: A comparative study. Energy 2016, 103, 75–85. [CrossRef]
72. Kumar, A.; Kim, M.H. Thermo-hydraulic performance of rectangular ducts with different multiple V-rib roughness shapes: A
comprehensive review and comparative study. Renew. Sustain. Energy Rev. 2016, 54, 635–652. [CrossRef]
73. Nidhul, K.; Kumar, S.; Yadav, A.K.; Anish, S. Enhanced thermo-hydraulic performance in a V-ribbed triangular duct solar air
heater: CFD and exergy analysis. Energy 2020, 200, 117448. [CrossRef]
74. Misra, R.; Singh, J.; Jain, S.K.; Faujdar, S.; Agrawal, M.; Mishra, A. Prediction of behavior of triangular solar air heater duct using
V-down rib with multiple gaps and turbulence promoters as artificial roughness: A CFD analysis. Int. J. Heat Mass Transf. 2020,
162, 120376. [CrossRef]
75. Patel, S.S.; Lanjewar, A. Experimental and numerical investigation of solar air heater with novel V-rib geometry. J. Energy Storage
2019, 21, 750–764. [CrossRef]
76. Patel, S.S.; Lanjewar, A. Heat transfer enhancement using additional gap in symmetrical element of V-geometry roughened solar
air heater. J. Energy Storage 2021, 38, 102545. [CrossRef]
77. Alam, T.; Saini, R.P.; Saini, J.S. Experimental investigation on heat transfer enhancement due to V-shaped perforated blocks in a
rectangular duct of solar air heater. Energy Convers. Manag. 2014, 81, 374–383. [CrossRef]
78. Alam, T.; Saini, R.P.; Saini, J.S. Experimental investigation of thermo hydraulic performance due to angle of attack in solar air
heater duct equipped with V-shaped perforated blockages. Int. J. Renew. Energy Technol. 2015, 6, 164–180. [CrossRef]
79. Alam, T.; Saini, R.P.; Saini, J.S. Experimental investigation of thermo-hydraulic performance of a rectangular solar air heater duct
equipped with V-shaped perforated blocks. Adv. Mech. Eng. 2014, 6, 948313. [CrossRef]
80. Alam, T.; Saini, R.P.; Saini, J.S. Heat transfer enhancement due to V-shaped perforated blocks in a solar air heater duct. Appl.
Mech. Mater. 2014, 619, 125–129. [CrossRef]
Energies 2022, 15, 2800 44 of 46
81. Alam, T.; Saini, R.P.; Saini, J.S. Effect of circularity of perforation holes in V-shaped blockages on heat transfer and friction
characteristics of rectangular solar air heater duct. Energy Convers. Manag. 2014, 86, 952–963. [CrossRef]
82. Singh Bisht, V.; Kumar Patil, A.; Gupta, A. Review and performance evaluation of roughened solar air heaters. Renew. Sustain.
Energy Rev. 2018, 81, 954–977. [CrossRef]
83. Singh, A.; Kumar, R.; Singh, P.; Goel, V. Solar air heater having multiple V-ribs with Multiple-Symmetric gaps as roughness
elements on Absorber-Plate: A parametric study. Sustain. Energy Technol. Assess. 2021, 48, 101559. [CrossRef]
84. Kumar, A.; Saini, R.P.; Saini, J.S. Experimental Investigations on Thermo-hydraulic Performance due to Flow- Attack- Angle in
Multiple V-ribs with Gap in a Rectangular Duct of Solar Air Heaters. J. Sustain. Energy Environ. 2013, 4, 1–7.
85. Kumar, A.; Saini, R.P.; Saini, J.S. Development of correlations for Nusselt number and friction factor for solar air heater with
roughened duct having multi-V-shaped with gap rib as arti fi cial roughness. Renew. Energy 2013, 58, 151–163. [CrossRef]
86. Jin, D.; Zhang, M.; Wang, P.; Xu, S. Numerical investigation of heat transfer and fluid flow in a solar air heater duct with
multi-V-shaped ribs on the absorber plate. Energy 2015, 89, 178–190. [CrossRef]
87. Jin, D.; Zuo, J.; Quan, S.; Xu, S.; Gao, H. Thermo-hydraulic performance of solar air heater with staggered multiple V-shaped ribs
on the absorber plate. Energy 2017, 127, 68–77. [CrossRef]
88. Jin, D.; Quan, S.; Zuo, J.; Xu, S. Numerical investigation of heat transfer enhancement in a solar air heater roughened by multiple
V-shaped ribs. Renew. Energy 2018, 134, 78–88. [CrossRef]
89. Promvonge, P.; Skullong, S. Thermal characteristics in solar air duct with V-shaped flapped-baffles and chamfered-grooves. Int. J.
Heat Mass Transf. 2021, 172, 121220. [CrossRef]
90. Sahu, M.K.; Prasad, R.K. Exergy based performance evaluation of solar air heater with arc- shaped wire roughened absorber
plate. Renew. Energy 2016, 96, 233–243. [CrossRef]
91. Gill, R.S.; Hans, V.S.; Saini, J.S.; Singh, S. Investigation on performance enhancement due to staggered piece in a broken arc rib
roughened solar air heater duct. Renew. Energy 2016, 104, 148–162. [CrossRef]
92. Gill, R.S.; Hans, V.S.; Singh, S. Investigations on thermo-hydraulic performance of broken arc rib in a rectangular duct of solar air
heater. Int. Commun. Heat Mass Transf. 2017, 88, 20–27. [CrossRef]
93. Ghritlahre, H.K.; Sahu, P.K.; Chand, S. Thermal performance and heat transfer analysis of arc shaped roughened solar air
heater—An experimental study. Sol. Energy 2020, 199, 173–182. [CrossRef]
94. Yadav, K.D.; Prasad, R.K. Performance analysis of parallel flow flat plate solar air heater having arc shaped wire roughened
absorber plate. Renew. Energy Focus 2020, 32, 23–44. [CrossRef]
95. Ambade, J.; Lanjewar, A. Experimental investigation of solar air heater with new symmetrical GAP ARC GEOMETRY and
staggered element. Int. J. Therm. Sci. 2019, 146, 106093. [CrossRef]
96. Azad, R.; Bhuvad, S.; Lanjewar, A. Study of solar air heater with discrete arc ribs geometry: Experimental and numerical approach.
Int. J. Therm. Sci. 2021, 167, 107013. [CrossRef]
97. Gill, R.S.; Hans, V.S.; Singh, R.P. Optimization of artificial roughness parameters in a solar air heater duct roughened with hybrid
ribs. Appl. Therm. Eng. 2021, 191, 116871. [CrossRef]
98. Sureandhar, G.; Srinivasan, G.; Muthukumar, P.; Senthilmurugan, S. Performance analysis of arc rib fin embedded in a solar air
heater. Therm. Sci. Eng. Prog. 2021, 23, 100891. [CrossRef]
99. Kumar, R.; Goel, V.; Singh, P.; Saxena, A.; Singh, A. Performance evaluation and optimization of solar assisted air heater with
discrete multiple arc shaped ribs. J. Energy Storage 2019, 26, 100978. [CrossRef]
100. Saravanakumar, P.T.; Somasundaram, D.; Matheswaran, M.M. Thermal and thermo-hydraulic analysis of arc shaped rib
roughened solar air heater integrated with fins and baffles. Sol. Energy 2019, 180, 360–371. [CrossRef]
101. Saravanakumar, P.T.; Somasundaram, D.; Matheswaran, M.M. Exergetic investigation and optimization of arc shaped rib
roughened solar air heater integrated with fins and baffles. Appl. Therm. Eng. 2020, 175, 115316. [CrossRef]
102. Agrawal, Y.; Bhagoria, J.L. Proceedings Experimental investigation for pitch and angle of arc effect of discrete artificial roughness
on Nusselt number and fluid flow characteristics of a solar air heater. Mater. Today Proc. 2020, 46, 5506–5511. [CrossRef]
103. Hassan, A.K.; Hasan, M.M.; Khan, M.E. Parametric investigation and correlation development for heat transfer and friction factor
in multiple arc dimple roughened solar air duct. Renew. Energy 2021, 174, 403–425. [CrossRef]
104. Agrawal, Y.; Bhagoria, J.L.; Pagey, V.S. Materials Today: Proceedings Enhancement of thermo-hydraulic performance using
double arc reverse ribs in a solar collector: Experimental approach. Mater. Today Proc. 2021, 47, 5749–6336. [CrossRef]
105. Kumar, A.; Behura, A.K.; Saboor, S.; Gupta, H.K. Comparative study on W-shaped roughened solar air heaters by using booster
mirror. Mater. Today Proc. 2020, 46, 5675–5680. [CrossRef]
106. Singh, I.; Singh, S. A review of artificial roughness geometries employed in solar air heaters. Renew. Sustain. Energy Rev. 2018, 92,
405–425. [CrossRef]
107. Wang, D.; Liu, J.; Liu, Y.; Wang, Y.; Li, B.; Liu, J. Evaluation of the performance of an improved solar air heater with “S” shaped
ribs with gap. Sol. Energy 2020, 195, 89–101. [CrossRef]
108. Baissi, M.T.; Brima, A.; Aoues, K.; Khanniche, R.; Moummi, N. Thermal behavior in a solar air heater channel roughened with
delta-shaped vortex generators. Appl. Therm. Eng. 2019, 165, 113563. [CrossRef]
109. Kumar, A.; Layek, A. Nusselt number and friction characteristics of a solar air heater that has a winglet type vortex generator in
the absorber surface. Exp. Therm. Fluid Sci. 2020, 119, 110204. [CrossRef]
Energies 2022, 15, 2800 45 of 46
110. Kumar, A.; Layek, A. Nusselt number and friction factor correlation of solar air heater having winglet type vortex generator over
absorber plate. Sol. Energy 2020, 205, 334–348. [CrossRef]
111. Kumar, R.; Kumar, S.; Thapa, S.; Sethi, M.; Fekete, G.; Singh, T. Impact of artificial roughness variation on heat transfer and
friction characteristics of solar air heating system. Alex. Eng. J. 2021, 61, 481–491. [CrossRef]
112. Promvonge, P.; Promthaisong, P.; Skullong, S. Numerical heat transfer in a solar air heater duct with punched delta-winglet
vortex generators. Case Stud. Therm. Eng. 2021, 26, 101088. [CrossRef]
113. Mahanand, Y.; Senapati, J.R. Thermo-hydraulic performance analysis of a solar air heater (SAH) with quarter-circular ribs on the
absorber plate: A comparative study. Int. J. Therm. Sci. 2021, 161, 106747. [CrossRef]
114. Gilani, S.E.; Al-kayiem, H.H.; Woldemicheal, D.E.; Gilani, S.I. Performance enhancement of free convective solar air heater by pin
protrusions on the absorber. Sol. Energy 2017, 151, 173–185. [CrossRef]
115. Perwez, A.; Kumar, R. Thermal performance investigation of the fl at and spherical dimple absorber plate solar air heaters. Sol.
Energy 2019, 193, 309–323. [CrossRef]
116. Debnat, S.; Das, B.; Randive, P. Influences of pentagonal ribs on the performance of rectangular solar air heater. Energy Procedia
2019, 158, 1168–1173. [CrossRef]
117. Antony, A.L.; Shetty, S.P.; Madhwesh, N.; Sharma, N.Y.; Karanth, K.V. Influence of stepped cylindrical turbulence generators on
the thermal enhancement factor of a flat plate solar air heater. Sol. Energy 2020, 198, 295–310. [CrossRef]
118. Patel, Y.M.; Jain, S.V.; Lakhera, V.J. Thermo-hydraulic performance analysis of a solar air heater roughened with reverse NACA
profile ribs. Appl. Therm. Eng. 2020, 170, 114940. [CrossRef]
119. Gabhane, M.G.; Kanase-Patil, A.B. Experimental analysis of double flow solar air heater with multiple C shape roughness. Sol.
Energy 2017, 155, 1411–1416. [CrossRef]
120. Saravanan, A.; Murugan, M.; Reddy, M.S.; Ranjit, P.S.; Elumalai, P.V.; Kumar, P.; Sree, S.R. Thermo-hydraulic performance of a
solar air heater with staggered C-shape finned absorber plate. Int. J. Therm. Sci. 2021, 168, 107068. [CrossRef]
121. Kumar, A.; Layek, A. Thermo-hydraulic performance of solar air heater having twisted rib overthe absorber plate. Int. J. Therm.
Sci. 2018, 133, 181–195. [CrossRef]
122. Kumar, A.; Layek, A. Nusselt number and friction factor correlation of solar air heater having twisted-rib roughness on absorber
plate. Renew. Energy 2019, 130, 687–699. [CrossRef]
123. Kumar, A.; Layek, A. Energetic and exergetic performance evaluation of solar air heater with twisted rib roughness on absorber
plate. J. Clean. Prod. 2019, 232, 617–628. [CrossRef]
124. Thakur, D.S.; Khan, M.K.; Pathak, M. Performance Evaluation of Solar Air Heater with Novel Hyperbolic Rib Geometry. Renew.
Energy 2016, 105, 786–797. [CrossRef]
125. Manjunath, M.S.; Karanth, K.V.; Sharma, N.Y. Numerical analysis of the influence of spherical turbulence generators on heat
transfer enhancement of flat plate solar air heater. Energy 2017, 121, 616–630. [CrossRef]
126. Kumar, V.; Murmu, R. Experimental investigation for thermal performance of inclined spherical ball roughened solar air duct.
Renew. Energy 2021, 172, 1365–1392. [CrossRef]
127. Promvonge, P.; Khanoknaiyakarn, C.; Sripattanapipat, S.; Skullong, S. Heat transfer in solar air duct with multi-V-ribbed absorber
and grooved back-plate. Chem. Eng. Res. Des. 2021, 168, 84–95. [CrossRef]
128. Sharma, S.K.; Kalamkar, V.R. Experimental and numerical investigation of forced convective heat transfer in solar air heater with
thin ribs. Sol. Energy 2017, 147, 277–291. [CrossRef]
129. Prakash, C.; Saini, R.P. Heat transfer and friction in rectangular solar air heater duct having spherical and inclined rib protrusions
as roughness on absorber plate. Exp. Heat Transf. 2018, 32, 469–487. [CrossRef]
130. Luo, L.; Wen, F.; Wang, L.; Sundén, B.; Wang, S. On the solar receiver thermal enhancement by using the dimple combined with
delta winglet vortex generator. Appl. Therm. Eng. 2016, 111, 586–598. [CrossRef]
131. Skullong, S.; Kwankaomeng, S.; Thianpong, C.; Promvonge, P. Thermal performance of turbulent flow in a solar air heater
channel with rib-grove turbulators. Int. Commun. Heat Mass Transf. 2014, 50, 34–43. [CrossRef]
132. Kumar, B.V.; Manikandan, G.; Kanna, P.R. Enhancement of heat transfer in SAH with polygonal and trapezoidal shape of the rib
using CFD. Energy 2021, 234, 121154. [CrossRef]
133. Tanda, G.; Satta, F. Heat transfer and friction in a high aspect ratio rectangular channel with angled and intersecting ribs. Int. J.
Heat Mass Transf. 2021, 169, 120906. [CrossRef]
134. Farhan, A.A.; Issam, A.; Ali, M.; Ahmed, H.E. Energetic and exergetic efficiency analysis of a v-corrugated solar air heater
integrated with twisted tape inserts. Renew. Energy 2021, 169, 1373–1385. [CrossRef]
135. Alam, T.; Kim, M.-H. A comprehensive review on single phase heat transfer enhancement techniques in heat exchanger
applications. Renew Sustain. Energy Rev. 2018, 81, 813–839. [CrossRef]
136. Ansari, M.; Bazargan, M. Optimization of Flat Plate Solar Air Heaters with Ribbed Surfaces. Appl. Therm. Eng. 2018, 136, 356–363.
[CrossRef]
137. Alfarawi, S.; Abdel-moneim, S.A.; Bodalal, A. Experimental investigations of heat transfer enhancement from rectangular duct
roughened by hybrid ribs. Int. J. Therm. Sci. 2017, 118, 123–138. [CrossRef]
138. Alam, T.; Kim, M.-H. Numerical study on thermal hydraulic performance improvement in solar air heater duct with semi ellipse
shaped obstacles. Energy 2016, 112, 588–598. [CrossRef]
Energies 2022, 15, 2800 46 of 46
139. Xiao, H.; Dong, Z.; Liu, Z.; Liu, W. Heat transfer performance and flow characteristics of solar air heaters with inclined trapezoidal
vortex generators. Appl. Therm. Eng. 2020, 179, 115484. [CrossRef]
140. Manjunath, M.S.; Karanth, K.V.; Sharma, N.Y. Numerical Investigation on Heat Transfer Enhancement of Solar Air Heater using
Sinusoidal Corrugations on Absorber Plate. Int. J. Mech. Sci. 2018, 138–139, 219–228. [CrossRef]
141. Bezbaruah, P.J.; Das, R.S.; Sarkar, B.K. Experimentally validated 3D simulation and performance optimization of a solar air duct
with modified conical vortex generators. Sol. Energy 2021, 224, 1040–1062. [CrossRef]
142. Dong, Z.; Liu, P.; Xiao, H.; Liu, Z.; Liu, W. A study on heat transfer enhancement for solar air heaters with ripple surface. Renew.
Energy 2021, 172, 477–487. [CrossRef]
143. Alam, T.; Meena, C.S.; Balam, N.B.; Kumar, A.; Cozzolino, R. Thermo-Hydraulic Performance Characteristics and Optimization
of Protrusion Rib Roughness in Solar Air Heater. Energies 2021, 14, 3159. [CrossRef]
144. Alam, T.; Kumar, A.; Balam, N.B. Thermo-Hydraulic Performance of Solar Air Heater Duct Provided with Conical Protrusion Rib
Roughnesses. Adv. Energy Res. 2020, 2, 159–168.
145. Kumar, R.; Kumar, A.; Goel, V. Performance improvement and development of correlation for friction factor and heat transfer
using computational fluid dynamics for ribbed triangular duct solar air heater. Renew. Energy 2018, 131, 788–799. [CrossRef]
146. Kumar, R.; Goel, V. Unconventional solar air heater with triangular flow-passage: A CFD based comparative performance
assessment of different cross- sectional rib-roughnesses. Renew. Energy 2021, 172, 1267–1278. [CrossRef]
147. Goel, V.; Kumar, R.; Bhattacharyya, S.; Tyagi, V.V.; Abusorrah, A.M. A comprehensive parametric investigation of hemispherical
cavities on thermal performance and flow-dynamics in the triangular-duct solar-assisted air-heater. Renew. Energy 2021, 173,
896–912. [CrossRef]
148. Xi, L.; Xu, L.; Gao, J.; Zhao, Z. Study on Conjugate Thermal Performance of a Steam-Cooled Ribbed Channel with Thick
Metallic Walls. In Proceedings of the ASME Turbo Expo 2021: Turbomachinery Technical Conference and Exposition, Online,
7–11 June 2021. [CrossRef]
149. Alam, T.; Kim, M.-H. Performance improvement of double-pass solar air heater—A state of art of review. Renew Sustain. Energy
Rev. 2017, 79, 779–793. [CrossRef]
150. Patil, A.K.; Saini, J.S.; Kumar, K. A comprehensive review on roughness geometries and investigation techniques used in
artificially roughened solar air heaters. Int. J. Renew Energy Res. 2012, 2, 1–15. [CrossRef]
151. Alam, T.; Balam, N.B.; Kulkarni, K.S.; Siddiqui, M.I.H.; Kapoor, N.R.; Meena, C.S.; Kumar, A.; Cozzolino, R. Performance
Augmentation of the Flat Plate Solar Thermal Collector: A Review. Energies 2021, 14, 6203. [CrossRef]
152. Istanto, T.; Danardono, D.; Yaningsih, I.; Wijayanta, A.T. Experimental study of heat transfer enhancement in solar air heater with
different angle of attack of V-down continuous ribs. AIP Conf. Proc. 2016, 1737, 060002. [CrossRef]
153. Singh, A.P.; Varun; Siddhartha. Heat transfer and friction factor correlations for multiple arc shape roughness elements on the
absorber plate used in solar air heaters. Exp. Therm. Fluid Sci. 2014, 54, 117–126. [CrossRef]
154. Karim, M.A.; Perez, E.; Amin, Z.M. Mathematical modelling of counter flow v-grove solar air collector. Renew. Energy 2014, 67,
192–201. [CrossRef]
155. Bondi, P.; Cicala, L.; Farina, G. Performance analysis of solar air heaters of conventional design. Sol. Energy 1988, 41, 101–107.
[CrossRef]
156. Pandey, N.K.; Bajpai, V.K. Thermo-hydraulic performance enhancement of solar air heater (SAH) having multiple arcs with gap
shaped roughness element on absorber plate. Int. J. Eng. Sci. Technol. 2016, 8, 34–42. [CrossRef]