Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Solar Energy 215 (2021) 44–63

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Performance of a hybrid compound parabolic concentrator solar dryer for


tomato slices drying
Hossein Ebadi a, b, Dariush Zare a, *, Masoud Ahmadi a, Guangnan Chen c
a
Biosystems Engineering Department, Shiraz University, P.O. Box. 71441-6518, Shiraz, Iran
b
MAHTEP Group, Dipartimento Energia “Galileo Ferraris” (DENERG), Politecnico di Torino, 10129 Turin, Italy
c
Faculty of Health, Engineering and Sciences, University of Southern Queensland, Toowoomba, QLD 4350, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, a hybrid drying system incorporating a compound parabolic concentrator (CPC) solar collector and
Hybridization an electric auxiliary heater was developed to make advances in sustainable tomato drying. The proposed dryer
Vitamin C destruction was designed to operate in a combined mode with solar energy as the prime source of energy, and the auxiliary
Solar drying
unit was only used in the cases with the absence of solar radiation or limitation in generated solar power.
Shrinkage
CPC solar collector
Experimental tests were carried out to investigate the drying performance at different levels of sample thickness,
airflow rate, and drying temperature. Using image processing and high-performance liquid chromatography
(HPLC) techniques, the quality of the dried products was also evaluated in terms of the changes in color,
shrinkage, and vitamin C content. Results indicated that the average drying time was about 231 min, while the
shortest time was obtained as 83 min, suggesting an improved performance compared to similar works. It was
found that drying temperature is the key factor influencing the rate of color changes, and shrinkage is only
affected by the thickness of tomato slices. The experimental data further showed that the destruction of vitamin C
was mostly influenced by the drying air temperature and sample thickness. The maximum energy and exergy
efficiencies of the solar collector were determined as 25 and 6% at the maximum and minimum airflow rates,
respectively.

1. Introduction drying cherry tomatoes (Nabnean et al., 2016), tomato slices (Dorouzi
et al., 2018), and tomato pomace (Badaoui et al., 2019). Gallali et al.
Food drying has emerged as a promising technique to meet the (2000) compared the quality of the dried tomato under natural and solar
purpose of food preservation. With a lack of suitable storage capacity, drying. Results indicated that solar drying could affect the quality of the
agro-industry often suffers from a substantial post-harvest loss, severe dried product through its texture, color, and flavor characteristics.
impacting on farmers’ revenues. Tomato fruit, as an important vegetable Moreover, the sample preparation method was also an important factor,
crop, mainly consists of water and needs special care in harvesting, where thin tomato slices showed more thermal sensitivity than samples
handling, and storage. The application of drying can bring significant in lobe formation, especially under the temperatures above 60–70 ◦ C.
benefits to communities with inadequate and improper storage facilities Sacilik et al. (2006) modeled a solar tunnel dryer for thin-layer drying of
when tomato preservation is important and necessary. Such commu­ tomato samples. They reported that the system was able to reduce to­
nities also often suffer from poor access to conventional power networks matoes’ moisture levels from 93.35 to 11.50% w.b in four days
for providing the required energy. Therefore, if the drying technology is compared to five days needed by the conventional sun drying method.
to be targeted in these areas, alternative sources of energy with low-cost Manaa et al. (2013) investigated a tomato dryer integrated with a flat
and high efficiency must be used. Solar drying has been demonstrated to plate solar collector and an auxiliary heater. Results exhibited that when
be very effective in tomato preservation by providing heat through the the drying temperature was less than 40 ◦ C then the drying rate was
collection of solar thermal energy and utilizing it to remove the moisture nearly constant for different tomato cultivars, while if the drying tem­
from the product in a hygienic way. perature exceeded 40 ◦ C, the drying curve varied significantly for
Many attempts have been made to develop different solar dryers for different verities. A commercial large-scale cabinet dryer assisted with

* Corresponding author at: Shiraz University, Iran.


E-mail address: dzare@shirazu.ac.ir (D. Zare).

https://doi.org/10.1016/j.solener.2020.12.026
Received 8 May 2020; Received in revised form 5 November 2020; Accepted 12 December 2020
Available online 6 January 2021
0038-092X/© 2020 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 1. Major components used in the experimental setup; (a) CPC solar collector, (b) air fan, (c) drying chamber, (d) drying tray, (e) tomato slices.

16 m2 solar collector was also introduced by Nabnean et al. (2016). With et al. (2020) presented a passive mixed solar dryer and reported that the
the load capacity of 100 kg cherry tomatoes in each batch, the developed constructed dryer reduces tomato drying time by 9 h compared to an
system showed superior performance compared to natural sun drying. indirect solar dryer.
From the assessments, collector efficiency was obtained from 21 to 69%, One of the techniques that can improve the amount of solar ab­
and the payback period was calculated as 1.37 years. Another large-scale sorption upon solar collectors is the utilization of reflectors or concen­
model was designed by Djebli et al. (2019) using a mixed greenhouse trators (Ratismith et al., 2017). This modification can augment the
type solar dryer for tomato fruits cut in two different shapes. The performance of solar dryers coupled with a solar collector. Stiling et al.
experimental investigation revealed that the drying operation could last (2012) showed that the concentrating panels might increase the internal
from 5 to 21.25 h depending on the thickness in flat shape slices and 8.5 temperature by 10 ◦ C and decrease the tomato drying time by 27%.
to 28.25 h for slices with edges shape. Furthermore, the integration of a parabolic trough collector for drying of
In order to improve the performance of solar dryers used in tomato apple samples revealed that the moisture level could drop to 8% after 11
dryings, researchers have developed novel designs to achieve higher h drying (Ullah and Kang, 2017). Using concave solar concentrators for a
efficiency, reduced drying time, and improved quality of final products. passive solar tomato dryer also showed that in addition to a 21%
Samimi-Akhijahani and Arabhosseini (2018) tested the effects of a PV- reduction in drying time, there could not be any apparent sacrifice in
assisted tracking system on an active solar dryer and reported that the pH, titratable acidity, color, Brix, lycopene, and vitamin C (Ringeisen
proposed modification could result in 16.6 to 36.6% reduction in drying et al., 2014).
time without having negative impacts on the quality of the dried to­ Compound Parabolic Concentrators (CPC) collectors with operating
matoes. The integration of a regenerating descant wheel to a tomato temperatures ranging from 100 to 150 ◦ C is a promising technology for
drying process was also introduced by Dorouzi et al. (2018). Using a PV- drying purposes (Lillo et al., 2017). Therefore, scientists (Lee et al.,
powered IR lamp inside the drying chamber, researchers found that the 2007) have tested an evacuated CPC collector for agricultural drying
color of final products varied with operating parameters. Erick César applications and demonstrated the viability of the proposed system. The

Table 1
Details of the experimental test conditions.
Test number Date Airflow rate (m3/s) Drying temperature (◦ C) Daily mean temperature (◦ C) Average wind velocity (m/s) Average solar radiation (W/m2)

1 2017/06/ 0.04 55 25.1 0.9 886.59


21
2 2017/06/ 0.04 65 24 1.5 889.38
23
3 2017/06/ 0.04 75 25.5 0.9 880.35
27
4 2017/07/ 0.025 55 22.8 0.6 920.53
19
5 2017/07/ 0.025 65 22 0.7 879.82
22
6 2017/07/ 0.025 75 23.3 0.6 866.66
23
7 2017/08/ 0.01 55 23 1 900.37
05
8 2017/08/ 0.01 65 25.7 0.9 905.38
06
9 2017/08/ 0.01 75 23.7 0.8 897.76
08

45
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 2. Schematics of the developed hybrid CPC dryer: (1) CPC collector, (2) air fan, (3) auxiliary heater, (4) drying cabinet, (5) drying samples, (6) temperature
sensors, (7) anemometer, (8) Power controller, (9) temperature controller, (10) pyranometer, (11) ambient temperature sensor, (12) data logger.

Table 2 velocities, and tomato slice thicknesses on the drying efficiency were
Variance analysis of drying time for tomato slices dried in the hybrid CPC dryer. studied. The maximum fraction of solar contribution with respect to air
Variables Degree of Sum of Mean square F value velocity was obtained as 25.07%. This showed that the proposed system
freedom squares error needs further amendments to increase solar contribution and decrease
Airflow rate (Q) 2 244279.87 122139.94 2111.42** electrical fraction. On the other hand, conducting TRANSYS modeling
Drying 2 110891.36 55445.68 958.48** and environmental analysis, Lamrani et al. (2019) reported that the
temperature integration of a CPC solar dryer with an auxiliary unit could result in a
(Td) 34% reduction in annual CO2 emitted in drying operations. Thus, the
Sample thickness 2 435477.64 217738.82 3764.03**
(S)
utilization of CPC collector can reduce the consumption of the alterna­
(Q)×(Td) 4 7629.87 1907.47 32.97** tive source and put the solar energy as the dominant and prime source.
(Q)×(S) 4 24738.97 6184.74 106.91** Despite the significant potential of CPC systems, overall, research
(Td)×(S) 4 18265.19 4566.3 78.94** articles revolving around solar CPC drying and its application for
(Q)×(Td)×(S) 8 3508.26 438.53 7.58**
various products are still scarce and inadequate. To the best of our
Error 54 3123.75 57.85
Total 81 knowledge, published literature lacks applied information for the com­
**
bination of CPC dryer with auxiliary heater and the effects of different
Significant at 1% probability.
operating modes on the quality of the final product. Therefore, in the
present research, a new hybrid CPC dryer was developed for drying
integration of CPC collectors into a drum drying process was also studied tomato slices. The effects of mass flow rate, operating temperature, and
by Milczarek et al. (2017). In this research, they used a 98.3 m2 CPC sample thickness on the system’s performance, Vitamin C degradation,
array with 40 kW heating power to dry prune and tomato pomaces shrinkage, and color changes were also investigated using energetic and
where some minor losses were observed in the nutrition and color of exergetic analysis.
final products.
Since the solar drying operation relies on the weather conditions and 2. Materials and methods
irradiance disturbance can mitigate the expected performance, hybrid
solar dryers with an auxiliary heating source would bring more system 2.1. Experimental setup
operation reliability (Amer et al., 2010). In this regard, many models
have been developed using gas (Amjad et al., 2020; López-Vidaña et al., An experimental test rig consisting of a solar CPC collector, a fan, an
2013; Murali et al., 2020; Yassen and Al-Kayiem, 2016; Zoukit et al., electric motor, a drying cabinet, three drying trays, and an auxiliary
2019), biomass (Bosomtwe et al., 2019; Prasad and Vijay, 2005), heating unit, was developed in Biosystems Engineering Department at
geothermal (Ananno et al., 2020), and electricity (Azmi et al., 2012; Shiraz University, Iran. The solar collector was comprised of five indi­
Sekyere et al., 2016) as the alternative energy sources. In the case of the vidual CPC troughs with 1.6 m length, 0.3 m aperture width, and 0.381
electrical back-up system, Lee et al. (2007) incorporated a water storage m absorber diameter, which resulted in a 2.5 concentration ratio (Fig. 1.
tank, CPC reflectors, auxiliary heater, and water to an air heat exchanger a). All absorber pipes were connected to a suction fan in the parallel
to provide hot air inside the drying chamber. Although results demon­ arrangement (Fig. 1.b). A one-phase electric motor (Techtop, Italy) with
strated that the proposed arrangement reduces the drying period by half, 0.5 hp power was utilized to induce the fan and provide hot air inside the
the implementation of the heat exchanger before the drying chamber drying cabinet. After reaching the desired temperature, the heated air
with limited efficiency may decrease the amount of useful energy was entered into the drying chamber constructed in plywood structure
absorbed by the collector. To overcome this problem, Boughali et al. (Fig. 1.c). Inside the dryer enclosure, three trays were positioned to form
(2009) presented a conventional indirect solar dryer integrated with a a thin layer drying process and facilitate the moisture removal phe­
3.75 kW electric heater to dry tomato slices. In this research, the air was nomenon (Fig. 1.d). Each tray consisted of a 0.15 m2 metallic grid
passed through a flat plate collector and subsequently was introduced to reinforced by a wooden frame which, let the air pass through the sample
electric resistors reaching the desired drying temperature before slices and absorb moisture (Fig. 1.e). The mist air left the drying
entering the drying area. The effects of different drying temperatures, air chamber through an exhaust pipe located at the top of the drying

46
H. Ebadi et al. Solar Energy 215 (2021) 44–63

cabinet. In order to minimize the amount of thermal loss, several com­ locations of measuring instruments incorporated in the experiments.
ponents, including all connecting pipes and channels, and the entire
collector were insulated using elastomeric insulation sheets.
2.4. Data analysis
In this work, a control unit for the operation of the auxiliary heater
was designed and implemented. The amount of consumed electric power
2.4.1. Drying
was optimized and proportional to the required value. To this end, a
The study of drying kinetics was carried out through the experi­
temperature controller (TK4L, Autonics Inc., South Korea) was used to
mental investigation of drying rates during various tests. Moisture
measure the air outlet temperature of the solar collector and compare it
content on a wet basis (Mw) can be determined by the ratio of the
with the set value. If the measured temperature was less than the desired
remained water inside the sample to the total weight of the initial
value, a proportional current within the range of 4 to 20 mA depending
sample (Djebli et al., 2019). Therefore, the drying rate (DR) was ob­
on the net difference, was sent to a power controller (SPC1-5, Autonics
tained using the following expression (Dorouzi et al., 2018)
Inc., South Korea). Consequently, based on the received current, the
power controller induced a proportional voltage ranging from 0 to 220 V Mw =
wi − wd
× 100 (1)
to the electric resistors and provided the least necessary heat for desired wi
air temperature. This configuration not only protects the electric heater
from sudden on/off shocks (extending its lifetime in the long-term where wi is the wet matter or initial mass (g) of the sample before drying
application) but also reassures the operation with the minimum elec­ and wd indicates the mass of dry matter (g).
tricity consumption, which can be economically beneficial. Mw+dt − Mw
DR = (2)
dt
2.2. Experimental procedure
2.4.2. Collector performance
Experimental tests were carried out under outdoor conditions during Energy efficiency is defined as the amount of useful energy collected
June, July, and August 2017 to evaluate the performance of the pro­ by the solar collector to the total amount of energy received from the
posed system and the quality of the final product (Ahmadi, 2018). In sun. This parameter is often used to indicate the viability and efficiency
each morning, tomatoes were procured from the local market, and of the system for energy conversion and utilization. Collector energy
samples with the same physical features and without any damage were efficiency (ηen ) can be determined as (Biondi et al., 1988):
selected for tests. Afterward, random samples in the range of 5 to 10 g
ṁCp ΔT
were chosen to be tested for moisture measurement. The mean initial ηen = (3)
IAc
moisture was obtained based on AOAC standards (Mertens, 2005) as a
94% wet basis, and the final moisture of dried material was found as a where Cp indicates specific heat capacity (J/kgK) of fluid, ṁ is the fluid
15% wet basis. Before each test, the dryer was operated for 1 h to reach a mass flow rate (kg/s),ΔT is the temperature difference between inlet and
stable condition with a constant temperature. Then to analyze the per­ outlet of the fluid, I is the solar radiation (W/m2), and Ac is the collector
formance of the developed dryer, experiments were conducted at three area (m2).
levels of drying temperature (55, 65, and 75 ◦ C), three levels of sample Exergy analysis is also a useful tool for efficiency assessment, which
thickness (4, 6, and 8 mm), and three levels of airflow rate (0.01, 0.025, considers the irreversibility of a thermal process and indicates the
0.04 m3/s), which were selected after several trials. For the moisture amount of useful work after degradation and loss through entropy
measurement, during the drying process, trays were removed from the generation. Thus, the value of exergy efficiency (ηex ) was obtained by the
cabinet regularly (12–18 times in each test), and the sample’s weight following equation (Sun et al., 2010)
was measured in less than 30 s to avoid any test error. Finally, dried [( ) ]
samples were photographed for color assessment and transferred to the ṁCp Tout − Tin − Tamb ln TTout − ΔP
lab for vitamin C evaluation. Table 1 shows the details of experimental ηex =
in ρ
(4)
conditions at each test day. IA(1 − Tamb /TS )

2.3. Instrumentation and testing apparatus where ΔP represents the pressure difference between the fluid inlet and
outlet, ρ is the fluid density (kg/m3), and Ts denotes sun’s temperature,
Temperature measurement was achieved using K type sensors which is assumed to be 4350 K (Farahat et al., 2009).
installed in several positions to record air inlet temperature (Tin), air
outlet temperature (Tout), glass cover temperature (Tg), absorber tem­ 2.4.3. Shrinkage measurement
perature (Tabs), ambient temperature (Tamb), and air temperature inside The implementation of MATLAB software and developing commands
the drying chamber (Td). In order to avoid measurement errors, each based on the previous works (Barzegar et al., 2015) for shape features
sensor was calibrated before the tests following standard instructions. extraction leads to the determination of surface area, perimeter,
The intensity of solar radiation was also measured with a pyranometer maximum diameter, minimum diameter, and equivalent diameter.
(Casella CEL, England) mounted inclined and parallel to the collector Shrinkage is a crucial factor for the study and can be evaluated by the
aperture. The employment of a digital data logger (CMC-99, SIMEX, ratio of changes in the surface area to the initial surface area (Hafezi
Poland) made it possible to record the data of temperatures and solar et al., 2015) as given by
radiation at 10 s intervals. Using a digital hotwire anemometer (Testo A1 − A2
435, Germany), airflow rate values were obtained via several mea­ %Shrinkage = × 100 (5)
A1
surements inside the air duct and at different depths. A digital electronic
scale (AND EK-2000, South Korea) was also used to measure the weight where A1 and A2 are the surface areas of tomato slices before and after
variation of the sample before, during, and after the drying process. each test, respectively.
Moreover, to evaluate the initial moisture content, random samples
were dried inside a 70 l oven (Behdad 3487, Iran). For the photography 2.4.4. Color assessment
purpose, images were taken from the dried samples inside an illuminator Collected images were analyzed for color assessment in RGB (L*a*b)
enclosure, also known as the cloudy sky, using a digital CCD camera format with developed MATLAB coding following the instructions
(IXUS, Canon, Japan), which was located 0.45 m above the samples. indicated in the literature (Barzegar et al., 2015). In this method, L*
Fig. 2 depicts the schematic view of the developed system and the indicates the lightness, a* varies from green to red, and b* shows blue to

47
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 3. Comparison of drying temperature at different levels of ; (a) airflow rate and sample thickness, (b) drying temperature and sample thickness, (c) airflow rate
and drying temperature.

48
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 4. Drying curves obtained at; (a) different levels of sample thickness when Td = 75 ◦ C and Q = 0.01 m3/s, (b) different levels of drying temperature when S = 4
mm and Q = 0.04 m3/s, (c) different levels of airflow rate when S = 4 mm and Td = 75 ◦ C.

49
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Table 3 (Td) at a 1% level of probability. As it can be seen, each variable has


Variance analysis of shrinkage percentage for tomato slices dried in the hybrid shown a significant effect on the drying time. Moreover, the interactions
CPC dryer. between different testing factors were also found to be influential on the
Variables Degree of Sum of Mean square F value drying time.
freedom squares error Based on Duncan’s test at a 5% level of probability, the influence of
Airflow rate (Q) 2 44.09 22.04 2.32ns each factor was further studied, as shown in Fig. 3. According to Fig. 3.a,
Drying 2 25.77 12.88 1.39ns as the sample thickness increased, the drying time also increased
temperature (Td) significantly. In each level of thickness, as the airflow rate increased
Sample thickness 2 301.47 150.73 16.35**
from 0.01 to 0.025 m3/s, the drying time also declined significantly.
(S)
(Q)×(Td) 4 10.83 2.7 0.29ns However, when the flow rate reached 0.4 m3/s, the reduction in drying
(Q)×(S) 4 13.45 3.36 0.36ns time became insignificant. As Fig. 3.b represents, drying time increased
(Td)×(S) 4 64.42 16.1 1.74ns significantly with the growth of sample thickness at each drying tem­
(Q)×(Td)×(S) 8 130.03 16.25 0.1ns perature. However, in the 4 mm thickness, the effect of temperature
Error 54 497.6 9.21 factor on the drying time was not significant. In the 6 mm thickness, only
Total 81 when the temperature increased from 55 to 75 ◦ C, drying time decreased
**
Significant at 1% probability. significantly. In 8 mm thickness, drying time was significantly reduced
when drying temperature was 75 and 65 ◦ C compared to 55 ◦ C. Fig. 3.c
yellowness. By tracking the variations in each component ranging from also shows that when drying temperature was 55 or 65 ◦ C, an increase in
− 120 to +120 (Yam and Papadakis, 2004), the rate of changes in color airflow rate from 0.01 to 0.025 m3/s could result in a significant
can be determined by the following relation (Ashebir et al., 2009) reduction in drying time. Yet the effect of 75 ◦ C drying temperature on
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ the drying time was not significant. Moreover, in each of the tempera­
ΔE = (L1 − L2 )2 + (a1 − a2 )2 + (b1 − b2 )2 (6) ture level, a rise in airflow rate from 0.025 to 0.04 m3/s could reduce the
drying time. Thus, it can be concluded that the airflow rate and thickness
have more significant effects on drying time than other drying
2.4.5. Vitamin C evaluation
parameters.
The process of vitamin C determination was achieved by high-
performance liquid chromatography (HPLC). In this step, 50 CC of
3.1.2. Drying rate
diluted metaphosphoric acid (1–10 g water) was prepared and mixed
The variations of moisture ratio under various operating parameters
with 5 g dried samples in the mixer. A 50 CC falcon was used with the
are shown in Fig. 4. In all the experiments, the moisture ratio started to
developed homogenous mixture through a centrifuge device to achieve a
fall at a constant level, losing the free water that existed on the samples’
transparent solution after 30 min centrifuge at 4 oC temperature. The
surface. As the moisture extraction proceeded, it reached the critical
extracted solution was further extruded through a Syringe filter (0.45
moisture initiating the second phase (falling rate period) with a
µm) and was added to the still water with 1 to 10 g ratio. Finally, 40 µl of
descending rate in water removal. This behavior can be attributed to the
the final product was injected into the HPLC device (Knauer, Germany),
water extraction from the deeper layers of the product, which requires a
where the evaluation was obtained using UV indicators in the wave­
higher amount of heat to evaporate bonded water. As Fig. 4.a shows, in a
length of 242 nm with sphere image (C18 column) operating at 25 oC.
given drying temperature of 75 ◦ C and airflow rate of 0.01 m3/s, samples
The detailed instruction of the utilized method can be found in the
with 4 mm thickness needed a shorter drying process compared to 6 and
literature (Georgé et al., 2011).
8 mm. Moreover, as the drying process reached the critical moisture
content, the gap between the obtained values became wider, indicating
2.4.6. Statistical analysis
that when tomato slices got thicker, water must travel a longer distance
Completely randomized design factorial experiments were designed
to reach the surface, which requires more drying time. As Fig. 4.b il­
and carried out under three replications to investigate the impacts of the
lustrates, in a given sample thickness (4 mm) and airflow rate (0.04 m3/
testing variables on drying and solar collector performance. The facto­
s), higher levels of drying temperature resulted in reduced drying time.
rial design and analysis of variance were achieved with SPSS 24 software
This can be attributed to the higher potential of water absorption in
using the Duncan test in 5% of probability.
drying temperature of 75 ◦ C than 65 ◦ C and 55 ◦ C. However, rapid
drying with higher temperatures resulted in several deteriorative
3. Results and discussions
changes in tomatoes, leading to quality degradation (Coelho et al.,
2013). Fig. 4.c also shows that when the drying temperature was 75 ◦ C
3.1. Drying performance
and the sample thickness was 4 mm, an increase in airflow rate from
0.01 to 0.025 m3/s could decrease the drying time by nearly 35%
3.1.1. Drying time
reaching 72 min, while the minimum drying temperature of 55 min was
Table 2 shows the analysis of variance for the effects of factors
achieved at the airflow rate of 0.04 m3/s. The obtained results demon­
including sample thickness (S), airflow rate (Q), and drying temperature
strated an improved drying performance compared to similar works
(Stiling et al., 2012). The drying curves of other test conditions are
presented in Appendix A. It can also be concluded that the sample
thickness promotes higher retardation of drying process than two other
parameters, especially at lower moisture levels. According to Zanoni
et al. (2000), the potential of oxidative damage with dried tomato slices
becomes higher in low moisture content (≤12%, w. b). Therefore, it is
expected that sample thickness plays a critical role in product quality
and shows a high impact on vitamin C degradation.

3.2. Quality evaluation

3.2.1. Shrinkage
Fig. 5. Comparison of shrinkage values at different levels of sample thickness. With the implementation of the image processing technique, photos

50
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 6. Comparison of changes in color components using Dunkan’s test (p < 0.05) at different drying temperatures; (a) L* values, (b) b* values, (c) ΔE vlues. (similar
letters implies no significant difference between the gap).

were taken before and after each test to compare the amount of area niche market. Two main causes of the color alteration in dried tomatoes
shrunk after the drying process. Table 3 presents the analysis of variance are pigment degradation (especially carotenoid pigments) (Lee and
for the effects of factors on the shrinkage percentage at a 1% level of Coates, 1999) and non-enzymatic browning reaction (oxidation of
probability. It can be inferred that sample thickness was the only factor ascorbic acid) (Cernîşev, 2010). In this study, the images of tomato slices
with significant effect, while other variables and the interactions had no were analyzed in L*a*b space, which led to the determination of color
significant impact on the shrinkage percentage. changes (ΔE) in dried samples. Statistical analysis was conducted to
Based on Duncan’s test at a 5% level of probability, Fig. 5 shows that investigate the effect of each factor on L*, a*, b*, and ΔE. Based on the
if the sample thickness increases, the shrinkage percentage inclines, variance analysis tables presented in Appendix B, it was found that the
where the maximum growth was found at S = 8 mm. This fact was due to influenced parameters (p < 0.01) were L*,b*, and ΔE, where the drying
the higher difference in moisture levels that existed between the su­ temperature was the only factor showing a significant difference in color
perficial and core layers. Moreover, the shrinkage increments found in parameters between the raw and dried samples.
the thicker samples might indicate the considerable retardation in dry­ From Fig. 6, it was found that the parameter of L*, which indicates
ing time as observed in Fig. 4.a. In this case, higher shrinkage percent led the brightness of dried samples was significantly decreased at 55 ◦ C. This
to a smaller effective area between the air and the slices, decreasing fact was attributed to the higher drying time at 55 ◦ C (92 min) than those
moisture exchange rate and water pick-up efficiency. at 65 and 75 ◦ C (72 and 34 min, respectively). As a result, higher drying
temperatures with reduced drying time caused less damage, where an
3.2.2. Color specification increase in drying temperature from 55 to 65 ◦ C improved brightness (L*
Color is a major criterion for consumers considering the quality of value) by 52.52% (Fig. 6.a). The observed behavior follows that re­
dried products, which can affect the final product’s acceptance in the ported by other scientists (Toor and Savage, 2006). Concerning b*,

51
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 7. Comparison of vitamin C degradation percentage observed at different levels of (a) drying temperature and airflow rate, (b) drying temperature and
thickness, (c) airflow rate and thickness.

52
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 8. Vitamin C degradation obtained in all test modes.

which represents the color shift from blue to yellowness, Fig. 6.b shows times higher vitamin C retention than 15 mm samples, the thickness was
that an increase in drying temperature from 55 to 65 ◦ C enhanced the found to be influential on vitamin C content during the drying process.
yellowness of tomato slices by 86.9%, indicating significant growth in Nonetheless, this effect could be regarded as indirect and may be
b*. The attained data were in good agreement with those obtained by dependent on other parameters such as peeling, which increases the
Ashebir et al. (2009) in which b* value varied between 21.9 and 28 for exposition to oxygen (Marfil et al., 2008). Fig. 7.c depicts that at a
drying temperatures of 55, 65, and 75 ◦ C and different tomato cultivars. constant flow rate when tomato slices are thicker, the vitamin C content
As Fig. 6.c shows, the changes in the color of dried samples became is less, while this deterioration rate diminishes with increasing the
significantly affected when the drying temperature inclines from 55 to airflow rate. In this case, a 191.2% growth in loss of vitamin C was seen
65 ◦ C (ΔE), while further increase in temperature from 65 to 75 ◦ C led to at the maximum airflow rate, and when samples were doubled in
no significant change in results. This trend was consistent with the idea thickness. This indicates that the airflow rate can be an important var­
introduced by other researchers (Barreiro et al., 1997; Shi et al., 1999), iable in controlling the retention of vitamin C, and the optimum value
which states that ΔE grows with an increase in the drying temperature. can preserve the quality of the final product despite high temperatures.
However, the overall color change with an average value of 30.19 was From Fig. 8, it can be found that in terms of vitamin C preservation,
higher than 14.6, 16.6, and 16.9, as found by Ashebir et al. (2009) for the least degradation of 8.3% was devoted to the test operated at drying
Amoroso, Berlinto, and Messina cultivars, respectively. This demon­ temperature of 55 ◦ C with a thickness of 4 mm and the highest airflow
strated the effects of the sample thickness on the color properties of rate of 0.04 m3/s. The worst-case with 70.8% damage in vitamin C was
dried materials (Coelho et al., 2013), where a higher average value of 6 obtained at the drying temperature of 65 ◦ C with 8 mm thickness and
mm slices used in this study could result in more color changes than 5 airflow rate of 0.01 m3/s.
mm slices investigated by Ashebir et al. (2009).
3.3. Performance of solar collector
3.2.3. Vitamin C degradation
The high-performance liquid chromatography technique was used to Alongside the evaluation of the performance of the developed dryer,
evaluate the amount of vitamin C in the samples. Fig. 7 depicts the the solar CPC collector was investigated from energetic and exergetic
variations in vitamin C and the fraction degraded by applying different points of view during the experiments. Fig. 9 displays the temperature of
test parameters. According to Fig. 7.a, increasing the airflow rate the collector’s components, including absorber tube and glass cover as
decreased the fraction of vitamin C degraded. This can be an effect of the well as the ambient, air inlet and outlet temperatures with respect to
reduction in time of being exposed to hot air. Santos and Silva (2008) solar radiation and within the operating hours. It can be seen that the
also indicated that higher drying temperatures improve the degradation maximum outlet temperatures were observed at solar noon with the
reaction when the drying temperature inclines from 55 to 65 ◦ C, vitamin maximum incident on the collector aperture. Moreover, reducing the
C degradation increased by 7%, reaching 53.2%. Although a further airflow rate from 0.04 to 0.025 and 0.01 m3/s increased the air tem­
increase in drying temperature from 65 to 75 ◦ C would result in more perature difference achieved from the collector by 33, and 166.7%,
deterioration rate with respect to the higher drying temperature, a 4% respectively. In this case, a lower airflow rate gave the passing air
reduction was seen in vitamin C degradation. This shows that at the enough time to extract heat from the inner walls of the absorber tube,
temperature of 75 ◦ C, the positive effect of reduced drying time lessens which consequently produced higher outlet temperature. Therefore, the
the impacting role of the temperature on degradation rate and preserves maximum Tout was found as almost 70 ◦ C at the airflow rate of 0.01 m3/s
more vitamin C compared to 65 ◦ C. However, this amount was still and solar intensity of 1000 W/m2. The temperature of the absorber tube
higher than the value obtained in the drying temperature of 55 ◦ C. Fig. 7. was the highest value recorded among other parameters where the
b represents that in a given temperature, slices with higher thickness maximum amount was attained at Q = 0.01 m3/s, which implies a
undergo greater damage in vitamin C, which was a result of a rise in the greater heat radiation loss from the collector.
drying time. The maximum degradation rise was recorded as 110% in In order to evaluate the system’s performance, the energy and exergy
the condition when the sample thickness increases from 4 to 8 mm, and efficiencies of the solar collector at different airflow rates are depicted in
the drying temperature is 55 ◦ C. Similar to the findings by Adom et al. Figs. 10, and 11, respectively. As Fig. 10 shows, the thermal efficiency
(1997) in which dried Okra slices with 5 mm thickness showed three increased as the day proceeded to the solar noon, reaching the highest

53
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 9. Temperature variation for different collector’s components with respects to solar radiation during operating hours.

value at solar noon and fell with a reduction in the solar incident. Despite the energy efficiency, the maximum exergy efficiency was
Furthermore, thermal efficiency increased with growth in airflow rate, observed at the lowest airflow rate (0.01 m3/s) as 6%. This fact can be
suggesting a better performance with less thermal loss, where the attributed to larger irreversibility at higher flow rates where the tur­
maximum value was achieved as 25% at Q = 0.04 m3/s. This reflects bulent flow regimes grew, and the leakage increased through the system.
that in higher air velocities, the turbulent flow regimes become higher This behavior has also been reported by other researchers in the liter­
inside the absorber tube, improving the Nusselt number and conse­ ature (Alta et al., 2010; Gupta and Kaushik, 2008). The low exergy value
quently enhancing the internal convection coefficient. As a result, a indicates that the solar collector is an inefficient device in terms of
higher heat transfer rate can be achieved between the passing air and the exergy efficiency, and more optimizations are therefore needed to
absorber. With the use of the second law of thermodynamics, exergetic reduce exergy loss and auxiliary energy consumption. One of the
efficiency of the proposed collector with respect to the airflow rate was important parameters in the reduction of exergy loss is the air temper­
obtained, as shown in Fig. 11. ature difference (Kurtbas and Durmuş, 2004), and the obtained results

54
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 10. Energy efficiency of the CPC collector for different airflow rates during operating hours.

Fig. 11. Exergy efficiency of the CPC collector for different airflow rates during operating hours.

echoed this fact by reaching the minimum exergy loss in the airflow rate 3.4. Performance of hybrid system
(0.01 m3/s) with the highest temperature difference (32 ◦ C). Exergy
destruction is the other factor contributing to low exergy efficiencies, In the previous work (Ebadi and Zare, 2020) published by the au­
which is based on the temperature difference between the sun and the thors, the viability of the used system was tested for thermal processes
absorber, the pressure drop in the ducts, and the temperature difference and supported from a thermo-economical point of view. In the present
between the absorber and the fluid (Kalogirou et al., 2016). From the study, the potential was evaluated for drying purposes with fresh air as
observations, the lowest exergy efficiency was recorded in the airflow the working fluid. In this regard, fractions of heat generated by solar
rate (Q = 0.04 m3/s) with the maximum temperature difference be­ (Hsol) and electric (Helc) units were calculated and compared for each
tween the absorber and the sun. In contrast, maximum efficiency was testing mode, as shown in Fig. 12, and Appendix C. According to Fig. 12.
attained when the air and absorber temperatures were very close (at Q a to Fig. 12.c, three different testing modes, including respectively
= 0.01 m3/s). Ge et al. (2014) highlighted the importance of such factors minimum load (Q = 0.01 m3/s and Td = 55 ◦ C), medium load (Q =
and found that the temperature difference between the absorber and the 0.025 m3/s and Td = 65 ◦ C), and maximum load (Q = 0.04 m3/s and Td
sun has a major role in the exergy loss by 72.86% of the total exergy rate. = 75 ◦ C), are presented to study various behaviors of the combined
Farahat et al. (2009) showed that the maximum exergy efficiency system. Fig. 12.a shows the distribution of power generation inside the
varies with the solar collector area and airflow rate. According to the system operating at the lowest energy demand where the solar fraction
results of this study, in the given collector area (2.4 m2), the optimum could completely meet the needed power. There was just a mere 10%
operating condition can be achieved in the airflow rates lower than 0.01 fraction of electricity consumption at the beginning minutes, which may
m3/s, which follows the same result obtained by the scientists (Farahat account for the time collector needed to reach its constant condition. As
et al., 2009). Therefore, exergy analysis must be employed to achieve a can be observed in Fig. 12.b and Fig. 12.c, when the energy demand
cost-effective design in solar drying systems with the highest possible inclines, the contribution of electric heat becomes dominant while solar
thermodynamic efficiencies. As a result, the system would be able to collector still provides a reasonable amount of heat. As a result, the
receive energy at the maximum rate without auxiliary heating. maximum solar contribution reached 50 and 30% in both medium and
high load operations.
The other assessment was to test the accuracy of the responses pro­
vided by the auxiliary and control unit in the operations under unstable

55
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 12. The fractions of contribution in power generation by solar and electric sources for sunny days under; (a) minimum load, (b) medium load, (c)
maximum load.

weather conditions. Therefore, as Fig. 13 illustrates, the performance of compensate for the required energy proportionally. This trend was
the system in terms of fractions generated by each source was obtained found to be precise and instantaneous throughout the entire test,
as a function of received solar radiation. In the early hours and with an demonstrating the accuracy of the hybrid system.
increase in solar radiation, electric fraction decreased gradually. How­
ever, as a sharp fall in a solar incident occurred at 12:00 PM, the useful
energy of the collector plummeted, requiring the electric heater to

56
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. 13. The indication of responses provided by the auxiliary system based on the disturbance in solar radiation for a typical cloudy day.

4. Uncertainty analysis Table 4


Measurement errors and the results of uncertainty analysis.
The uncertainties are based on experimental errors coming from Variable Uncertainty Value
independent variables and their relation with the main dependent fac­
Solar radiation ΔI ± 5%
tors. The following expressions describe the uncertainties for dependent
Airflow velocity Δυair ± (0.03 m/s + 5% measured value)
functions (Holman, 2012)
Air inlet temperature ΔTin ± 0.5 K
R = R(x1 , x2 , x3 , ⋯, xn ) (7) Air outlet temperature ΔTout ± 0.5 K
Ambient temperature ΔTamb ± 0.5 K
If the result R is a given function of the independent variables x1, x2,
Maximum energy efficiency Δηen 6.0%
x3, …, xn, thus
Maximum exergy efficiency Δηex 6.7%
[ ]12
∂R ∂R ∂R
WRe = ( w1 )2 + ( w2 )2 + ⋯ + ( wn )2 (8)
∂x1 ∂x2 ∂xn • The highest level of vitamin C preserved during the drying procedure
was 375.941 mg/100 g with 4 mm thickness of slices dried at 0.04
where WR is the result of uncertainty and w1, w2, …, wn are the un­
m3/s and 55 ◦ C.
certainties of the independent variables. Here the main uncertainty
• The highest energy and exergy efficiencies were calculated as 25.04
variables are the fluid mass flow rate (airflow velocity), solar radiation,
and 6.00% respectively at airflow rates of 0.04 and 0.01 m3/s.
ambient, inlet, and outlet temperatures, where the relative equations for
• The hybrid system showed excellent responses to changes in solar
uncertainties in energy and exergy efficiencies are given in Eqs. (9) and
radiation. The minimum fraction of CPC solar collector was achieved
(10). Table 4 provides the experimental errors and uncertainty results.
in the highest load operation with nearly 30%.
[ ]12
∂η ∂η ∂η ∂η
Wηen = ( en wI )2 + ( en wṁ )2 + ( en wTin )2 + ( en wTout )2 (9) Overall, this study demonstrated the viability of the CPC collectors
∂I ∂ṁ ∂Tin ∂Tout
for the implementation in drying systems. It showed that a hybrid sys­
[ ]12 tem could outperform conventional solar dryers with reduced drying
∂η ∂η ∂η ∂η ∂η time, preservation of the quality of the final product, and conserving the
Wηex = ( ex wI )2 + ( ex wṁ )2 + ( ex wTin )2 + ( ex wTout )2 +( ex Tamb )2
∂I ∂ṁ ∂Tin ∂Tout ∂Tamb energy. The results also indicated that the use of concentrators in solar
(10) drying technology could assist toward the goal of sustainable develop­
ment in agro-industry. For this purpose, the deployment of concen­
5. Conclusions trating systems must be performed cautiously due to the delicacy of
agricultural products, which requires careful quality assessments in
In the present study, the performance of a hybrid drying system terms of nutritional values, and marketability. With the exergy analysis,
equipped with solar CPC collector was investigated, and the quality of optimization can also play a crucial role in identifying the best operating
dried tomatoes was determined using image processing and HPLC condition to ensure economic profits, high quality in dried materials,
techniques. Tests were conducted under different levels of airflow rate, and maximum efficiency. The result of this study can further be scaled-
slice thickness, and drying temperature at the outdoor conditions. The up for larger integration in terms of the constant drying capacity per
main conclusions are listed as follows; collector area (Raitila and Tsupari, 2020). To this end, further studies on
economic analysis are still necessary to achieve the annual gross margin
• The minimum drying time was achieved as 83 min for samples of 4 of the smaller dryer, taking into account of the costs of the solar col­
mm thickness with 0.04 m3/s airflow rate and 75 ◦ C drying lector, energy consumed during operation, interest rate, and the price of
temperature. raw and dried products. Then the result values can be multiplied by the
• The maximum shrinkage was recorded in samples with 8 mm ratio of solar collector surface areas. Future studies could also investi­
thickness. gate the hybridization with other sources of energy, and integration with
• The drying temperature was the only factor influencing the rate of thermal storage systems, and other assisting drying technologies (e.g.,
changes in samples’ color, where the maximum alteration was ach­ ultrasound and infrared) with CPC dryers to reach the best configuration
ieved at 75 ◦ C. in industrial scale.

57
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Declaration of Competing Interest

The author declare that there is no conflict of interest.

Appendix A

See Fig. A1

Fig. A1. Drying curves obtained for various test parameters.

58
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. A1. (continued).

59
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Fig. A1. (continued).

60
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Appendix B

See Table B1-B4

Table B1
Results of variance analysis for changes in L* after the drying process.
Variables Degree of freedom Sum of squares Mean square error F value

Airflow rate (Q) 2 16.507 8.254 1.750ns


Drying temperature (Td) 2 1028.436 514.218 109.044**
Sample thickness (S) 2 0.651 0.326 0.069ns
(Q)×(Td) 4 7.602 1.9 0.403ns
(Q)×(S) 4 0.501 0.125 0.027ns
(Td)×(S) 4 14.867 3.717 0.788ns
(Q)×(Td)×(S) 8 44.879 5.610 1.19ns

Error 54 254.647 4.716


Total 81
**
Significant at 1% probability.

Table B2
Results of variance analysis for changes in a* after the drying process.
Variables Degree of freedom Sum of squares Mean square error F value
ns
Airflow rate (Q) 2 9.511 4.756 0.237
ns
Drying temperature (Td) 2 17.094 8.547 0.425
ns
Sample thickness (S) 2 4.58 2.290 0.114
ns
(Q)×(Td) 4 49.888 12.472 0.620
ns
(Q)×(S) 4 83.571 20.893 1.039
ns
(Td)×(S) 4 20.530 5.133 0.255
ns
(Q)×(Td)×(S) 8 123.025 15.378 0.765

Error 54 1085.533 20.102


Total 81
**
Significant at 1% probability.

Table B3
Results of variance analysis for changes in b* after the drying process.
Variables Degree of freedom Sum of squares Mean square error F value

Airflow rate (Q) 2 20.056 10.028 0.474 ns


Drying temperature (Td) 2 2707.755 1353.878 63.928**
Sample thickness (S) 2 42.171 21.085 0.996ns
(Q)×(Td) 4 11.331 2.833 0.134ns
(Q)×(S) 4 59.324 14.831 0.7ns
(Td)×(S) 4 61.281 15.320 0.723ns
(Q)×(Td)×(S) 8 55.415 6.927 0.327ns

Error 54 1143.618 21.178


Total 81
**
Significant at 1% probability.

Table B4
Results of variance analysis for ΔE after the drying process.
Variables Degree of freedom Sum of squares Mean square error F value

Airflow rate (Q) 2 1.465 0.733 0.044ns


Drying temperature (Td) 2 691.708 345.854 20.832**
Sample thickness (S) 2 35.157 17.578 1.059ns
(Q)×(Td) 4 27.490 6.872 0.389ns
(Q)×(S) 4 65.158 16.29 0.98ns
(Td)×(S) 4 53.989 13.497 0.813ns
(Q)×(Td)×(S) 8 127.830 15.979 0.962ns

Error 54 896.509 16.602


Total 81
**
Significant at 1% probability.

61
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Appendix C

See Fig. C1.

Fig. C1. The fractions of contribution in power generation by solar and electric sources under various operation modes and fair weather.

References Engineering. IOP Publishing, p. 012010. https://doi.org/10.1088/1757-899X/36/1/


012010.
Badaoui, O., Hanini, S., Djebli, A., Haddad, B., Benhamou, A., 2019. Experimental and
Adom, K.K., Dzogbefia, V.P., Ellis, W.O., 1997. Combined effect of drying time and slice
modelling study of tomato pomace waste drying in a new solar greenhouse:
thickness on the solar drying of okra. J. Sci. Food Agric. 73, 315–320. https://doi.
evaluation of new drying models. Renew. Energy 133, 144–155. https://doi.org/
org/10.1002/(SICI)1097-0010(199703)73:3<315::AID-JSFA718>3.0.CO;2-N.
10.1016/j.renene.2018.10.020.
Ahmadi, M., 2018. Evaluation of Non-Evacuated Compound Parabolic Solar Collector for
Barreiro, J.A., Milano, M., Sandoval, A.J., 1997. Kinetics of colour change of double
a Cabinet Dryer (a Case Study, Tomato Slices). Unpublished MS Thesis. Shiraz
concentrated tomato paste during thermal treatment. J. Food Eng. 33, 359–371.
University.
https://doi.org/10.1016/s0260-8774(97)00035-6.
Alta, D., Bilgili, E., Ertekin, C., Yaldiz, O., 2010. Experimental investigation of three
Barzegar, M., Zare, D., Stroshine, R.L., 2015. An integrated energy and quality approach
different solar air heaters: energy and exergy analyses. Appl. Energy 87, 2953–2973.
to optimization of green peas drying in a hot air infrared-assisted vibratory bed
https://doi.org/10.1016/j.apenergy.2010.04.016.
dryer. J. Food Eng. 166, 302–315. https://doi.org/10.1016/j.jfoodeng.2015.06.026.
Amer, B.M.A., Hossain, M.A., Gottschalk, K., 2010. Design and performance evaluation
Biondi, P., Cicala, L., Farina, G., 1988. Performance analysis of solar air heaters of
of a new hybrid solar dryer for banana. Energy Convers. Manag. 51, 813–820.
conventional design. Sol. Energy 41, 101–107. https://doi.org/10.1016/0038-092X
https://doi.org/10.1016/j.enconman.2009.11.016.
(88)90120-X.
Amjad, W., Ali Gilani, G., Munir, A., Asghar, F., Ali, A., Waseem, M., 2020. Energetic and
Bosomtwe, A., Danso, J.K., Osekre, E.A., Opit, G.P., Mbata, G., Armstrong, P., Arthur, F.
exergetic thermal analysis of an inline-airflow solar hybrid dryer. Appl. Therm. Eng.
H., Campbell, J., Manu, N., McNeill, S.G., Akowuah, J.O., 2019. Effectiveness of the
166, 114632 https://doi.org/10.1016/j.applthermaleng.2019.114632.
solar biomass hybrid dryer for drying and disinfestation of maize. J. Stored Prod.
Ananno, A.A., Masud, M.H., Dabnichki, P., Ahmed, A., 2020. Design and numerical
Res. 83, 66–72. https://doi.org/10.1016/j.jspr.2019.05.011.
analysis of a hybrid geothermal PCM flat plate solar collector dryer for developing
Boughali, S., Benmoussa, H., Bouchekima, B., Mennouche, D., Bouguettaia, H.,
countries. Sol. Energy 196, 270–286. https://doi.org/10.1016/j.
Bechki, D., 2009. Crop drying by indirect active hybrid solar – electrical dryer in the
solener.2019.11.069.
eastern Algerian Septentrional Sahara. Sol. Energy 83, 2223–2232. https://doi.org/
Ashebir, D., Jezik, K., Weingartemann, H., Gretzmacher, R., 2009. Change in color and
10.1016/j.solener.2009.09.006.
other fruit quality characteristics of tomato cultivars after hot-air drying at low final-
Cernîşev, S., 2010. Effects of conventional and multistage drying processing on non-
moisture content. Int. J. Food Sci. Nutr. 60, 308–315. https://doi.org/10.1080/
enzymatic browning in tomato. J. Food Eng. 96, 114–118. https://doi.org/10.1016/
09637480903114128.
j.jfoodeng.2009.07.002.
Azmi, M.S.M., Othman, M.Y., Sopian, K., Ruslan, M.H., Majid, Z.A.A., Fudholi, A., Yasin,
Coelho, K., Costa, B.R., Pinto, L.A. de A., 2013. Evaluation of lycopene loss and colour
J.M., 2012. Development of a solar assisted drying system using double-pass solar
values in convective drying of tomato by surface response methodology. Int. J. Food
collector with finned absorber, in: IOP Conference Series: Materials Science and
Eng. 9, 233–238. https://doi.org/10.1515/ijfe-2012-0202.

62
H. Ebadi et al. Solar Energy 215 (2021) 44–63

Djebli, A., Hanini, S., Badaoui, O., Boumahdi, M., 2019. A new approach to the Mertens, D., 2005. AOAC official method 922.02. Plants preparation of laboratuary
thermodynamics study of drying tomatoes in mixed solar dryer. Sol. Energy 193, sample. Chapter 3 Official Methods of Analysis 20877–22417.
164–174. https://doi.org/10.1016/j.solener.2019.09.057. Milczarek, R.R., Ferry, J.J., Alleyne, F.S., Olsen, C.W., Olson, D.A., Winston, R., 2017.
Dorouzi, M., Mortezapour, H., Akhavan, H.R., Moghaddam, A.G., 2018. Tomato slices Solar thermal drum drying performance of prune and tomato pomaces. Food
drying in a liquid desiccant-assisted solar dryer coupled with a photovoltaic-thermal Bioprod. Process. 106, 53–64. https://doi.org/10.1016/j.fbp.2017.08.009.
regeneration system. Sol. Energy 162, 364–371. https://doi.org/10.1016/j. Murali, S., Amulya, P.R., Alfiya, P.V., Delfiya, D.S.A., Samuel, M.P., 2020. Design and
solener.2018.01.025. performance evaluation of solar - LPG hybrid dryer for drying of shrimps. Renew.
Ebadi, H., Zare, D., 2020. Performance evaluation and thermo-economic analysis of a Energy 147, 2417–2428. https://doi.org/10.1016/j.renene.2019.10.002.
non-evacuated CPC solar thermal hybrid system: an experimental study. Int. J. Nabnean, S., Janjai, S., Thepa, S., Sudaprasert, K., Songprakorp, R., Bala, B.K., 2016.
Sustain. Energy 1–25. https://doi.org/10.1080/14786451.2020.1748028. Experimental performance of a new design of solar dryer for drying osmotically
Erick César, L.V., Ana Lilia, C.M., Octavio, G.V., Isaac, P.F., Rogelio, B.O., 2020. Thermal dehydrated cherry tomatoes. Renew. Energy 94, 147–156. https://doi.org/10.1016/
performance of a passive, mixed-type solar dryer for tomato slices (Solanum j.renene.2016.03.013.
lycopersicum). Renew. Energy 147, 845–855. https://doi.org/10.1016/j. Prasad, J., Vijay, V.K., 2005. Experimental studies on drying of Zingiber officinale,
renene.2019.09.018. Curcuma longa l. and Tinospora cordifolia in solar-biomass hybrid drier. Renew.
Farahat, S., Sarhaddi, F., Ajam, H., 2009. Exergetic optimization of flat plate solar Energy 30, 2097–2109. https://doi.org/10.1016/j.renene.2005.02.007.
collectors. Renew. Energy 34, 1169–1174. https://doi.org/10.1016/j. Raitila, J., Tsupari, E., 2020. Feasibility of solar-enhanced drying of woody biomass.
renene.2008.06.014. BioEnergy Res. 13, 210–221. https://doi.org/10.1007/s12155-019-10048-z.
Gallali, Y.M., Abujnah, Y.S., Bannani, F.K., 2000. Preservation of fruits and vegetables Ratismith, W., Favre, Y., Canaff, M., Briggs, J., 2017. A non-tracking concentrating
using solar drier: A comparative study of natural and solar drying, III; Chemical collector for solar thermal applications. Appl. Energy 200, 39–46. https://doi.org/
analysis and sensory evaluation data of the dried samples (grapes, figs, tomatoes and 10.1016/j.apenergy.2017.05.044.
onions). Renew. Energy 19, 203–212. https://doi.org/10.1016/s0960-1481(99) Ringeisen, B., Barrett, M.D., Stroeve, P., 2014. Concentrated solar drying of tomatoes.
00032-4. Energy Sustain. Dev. 19, 47–55. https://doi.org/10.1016/j.esd.2013.11.006.
Ge, Z., Wang, H., Wang, H., Zhang, S., Guan, X., 2014. Exergy analysis of flat plate solar Sacilik, K., Keskin, R., Elicin, A.K., 2006. Mathematical modelling of solar tunnel drying
collectors. Entropy 16, 2549–2567. https://doi.org/10.3390/e16052549. of thin layer organic tomato. J. Food Eng. 73, 231–238. https://doi.org/10.1016/j.
Georgé, S., Tourniaire, F., Gautier, H., Goupy, P., Rock, E., Caris-Veyrat, C., 2011. jfoodeng.2005.01.025.
Changes in the contents of carotenoids, phenolic compounds and vitamin C during Samimi-Akhijahani, H., Arabhosseini, A., 2018. Accelerating drying process of tomato
technical processing and lyophilisation of red and yellow tomatoes. Food Chem. 124, slices in a PV-assisted solar dryer using a sun tracking system. Renew. Energy 123,
1603–1611. https://doi.org/10.1016/j.foodchem.2010.08.024. 428–438. https://doi.org/10.1016/j.renene.2018.02.056.
Gupta, M.K., Kaushik, S.C., 2008. Exergetic performance evaluation and parametric Santos, P.H.S., Silva, M.A., 2008. Retention of Vitamin C in drying processes of fruits and
studies of solar air heater. Energy 33, 1691–1702. https://doi.org/10.1016/j. vegetables—a review. Dry. Technol. 26, 1421–1437. https://doi.org/10.1080/
energy.2008.05.010. 07373930802458911.
Hafezi, N., Sheikhdavoodi, M.J., Sajadiye, S.M., Arkiyan, A.H., 2015. Study on shrinkage Sekyere, C.K.K., Forson, F.K., Adam, F.W., 2016. Experimental investigation of the
characteristic of potato slices based on computer vision. Agric. Eng. Int. CIGR J. 17, drying characteristics of a mixed mode natural convection solar crop dryer with back
287–295. up heater. Renew. Energy 92, 532–542. https://doi.org/10.1016/j.
Holman, J.P., 2012. Experimental methods for engineers. McGraw-Hill/Connect Learn renene.2016.02.020.
Succeed. Shi, J., Maguer, M. Le, Kakuda, Y., Liptay, A., Niekamp, F., 1999. Lycopene degradation
Kalogirou, S.A., Karellas, S., Braimakis, K., Stanciu, C., Badescu, V., 2016. Exergy and isomerization in tomato dehydration. Food Res. Int. 32, 15–21. https://doi.org/
analysis of solar thermal collectors and processes. Prog. Energy Combust. Sci. 56, 10.1016/S0963-9969(99)00059-9.
106–137. https://doi.org/10.1016/j.pecs.2016.05.002. Stiling, J., Li, S., Stroeve, P., Thompson, J., Mjawa, B., Kornbluth, K., Barrett, D.M., 2012.
Kurtbas, İ., Durmuş, A., 2004. Efficiency and exergy analysis of a new solar air heater. Performance evaluation of an enhanced fruit solar dryer using concentrating panels.
Renew. Energy 29, 1489–1501. https://doi.org/10.1016/j.renene.2004.01.006. Energy Sustain. Dev. 16, 224–230. https://doi.org/10.1016/j.esd.2012.01.002.
Lamrani, B., Khouya, A., Draoui, A., 2019. Energy and environmental analysis of an Sun, W., Ji, J., He, W., 2010. Influence of channel depth on the performance of solar air
indirect hybrid solar dryer of wood using TRNSYS software. Sol. Energy 183, heaters. Energy 35, 4201–4207. https://doi.org/10.1016/j.energy.2010.07.006.
132–145. https://doi.org/10.1016/j.solener.2019.03.014. Toor, R.K., Savage, G.P., 2006. Effect of semi-drying on the antioxidant components of
Lee, G.-H., Kang, W.S., Park, C.H., 2007. Drying System Using CPC Evacuated Tubular tomatoes. Food Chem. 94, 90–97. https://doi.org/10.1016/j.
Solar Collector, in: ASABE Annual International Meeting, Minneapolis, 17 - 20 June. foodchem.2004.10.054.
American Society of Agricultural and Biological Engineers, St. Joseph, MI, p. 1. Ullah, F., Kang, M., 2017. Impact of air flow rate on drying of apples and performance
https://doi.org/10.13031/2013.23515. assessment of parabolic trough solar collector. Appl. Therm. Eng. 127, 275–280.
Lee, H.S., Coates, G.A., 1999. Thermal pasteurization effects on color of red grapefruit https://doi.org/10.1016/j.applthermaleng.2017.07.101.
juices. J. Food Sci. 64, 663–666. https://doi.org/10.1111/j.1365-2621.1999. Yam, K.L., Papadakis, S.E., 2004. A simple digital imaging method for measuring and
tb15106.x. analyzing color of food surfaces. J. Food Eng. 61, 137–142. https://doi.org/
Lillo, I., Pérez, E., Moreno, S., Silva, M., 2017. Process heat generation potential from 10.1016/S0260-8774(03)00195-X.
solar concentration technologies in Latin America: the case of Argentina. Energies Yassen, T.A., Al-Kayiem, H.H., 2016. Experimental investigation and evaluation of
10, 383. https://doi.org/10.3390/en10030383. hybrid solar/thermal dryer combined with supplementary recovery dryer. Sol.
López-Vidaña, E.C., Méndez-Lagunas, L.L., Rodríguez-Ramírez, J., 2013. Efficiency of a Energy 134, 284–293. https://doi.org/10.1016/j.solener.2016.05.011.
hybrid solar-gas dryer. Sol. Energy 93, 23–31. https://doi.org/10.1016/j. Zanoni, B., Pagliarini, E., Foschino, R., 2000. Study of the stability of dried tomato halves
solener.2013.01.027. during shelf-life to minimise oxidative damage. J. Sci. Food Agric. 80, 2203–2208.
Manaa, S., Younsi, M., Moummi, N., 2013. Solar drying of tomato in the arid area of https://doi.org/10.1002/1097-0010(200012)80:15<2203::AID-JSFA775>3.0.CO;
TOUAT (Adrar, Algeria). In: Energy Procedia. Elsevier Ltd, pp. 511–514. https://doi. 2-W.
org/10.1016/j.egypro.2013.07.058. Zoukit, A., El Ferouali, H., Salhi, I., Doubabi, S., Abdenouri, N., 2019. Simulation, design
Marfil, P.H.M., Santos, E.M., Telis, V.R.N., 2008. Ascorbic acid degradation kinetics in and experimental performance evaluation of an innovative hybrid solar-gas dryer.
tomatoes at different drying conditions. LWT – Food Sci. Technol. 41, 1642–1647. Energy 189, 116279. https://doi.org/10.1016/j.energy.2019.116279.
https://doi.org/10.1016/j.lwt.2007.11.003.

63

You might also like