Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Journal Pre-proofs

Solar-LP Gas Hybrid Plant for Dehydration of Food

N.M. Ortiz-Rodríguez, O. García-Valladares, I. Pilatowsky-Figueroa, C.


Menchaca-Valdez

PII: S1359-4311(20)32978-1
DOI: https://doi.org/10.1016/j.applthermaleng.2020.115496
Reference: ATE 115496

To appear in: Applied Thermal Engineering

Received Date: 4 October 2019


Revised Date: 17 April 2020
Accepted Date: 14 May 2020

Please cite this article as: N.M. Ortiz-Rodríguez, O. García-Valladares, I. Pilatowsky-Figueroa, C. Menchaca-
Valdez, Solar-LP Gas Hybrid Plant for Dehydration of Food, Applied Thermal Engineering (2020), doi: https://
doi.org/10.1016/j.applthermaleng.2020.115496

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Elsevier Ltd. All rights reserved.


Solar-LP Gas Hybrid Plant for Dehydration of Food

N. M. Ortiz-Rodríguez, O. García-Valladares*, I. Pilatowsky-Figueroa, C. Menchaca-


Valdez

Instituto de Energías Renovables, Universidad Nacional Autónoma de México, Privada


Xochicalco s/n, Centro, CP 62580, Temixco, Morelos
*Corresponding author: ogv@ier.unam.mx
Abstract
This paper presents the thermal and energy analysis of a Solar-LP gas hybrid drying plant built in
Zacatecas, Mexico. It is focused on the dehydration of agricultural products. The drying system is a
forced convection hot air type with a drying chamber. For the air heating required in the drying
process, there are two solar thermal systems: a direct air heating system (48 solar air heaters with an
area of 111.1 m2) and an indirect air heating system (40 solar water collectors with an area 92.4 m2
and 6000 liters of thermal storage). Also, there is a backup system of conventional energy (L.P gas
burner). Hybrid tests were carried out to dry Nopal (Opuntia ficus-indica) initially operating with the
direct system, after that with the indirect system and complementing the energy demand of drying
with the conventional system. The effective drying time was between 14.3 and 11.7 h, removing a
total of 179 and 183 kg of water, respectively. The air temperature at the entrance of the drying
chamber was around 59 °C. The solar fraction obtained for this hybrid mode of operation was around
80 %. The thermal analysis of the direct and indirect air heating system, together with the conventional
system, are presented and discussed. The performance of the solar systems is finally compared with
a conventional LP gas drying system via a cost and return on investment analysis.

Keywords: Solar energy; Solar drying; Return of investment; Nopal drying; Indirect solar dryer

1. Introduction
The loss (that takes place at the production, storage, processing, and distribution stages) and
waste of food harm the environment. It is due to the use of water, energy, and other natural resources
to produce food that nobody will consume. According to the Food and Agriculture Organization (FAO)
studies, it is estimated that around 30 % of cereals are lost and wasted each year: 40-50 % of tubers,
fruits and vegetables, 20 % of oilseeds, meat and dairy products, and 35 % of fish. Food losses and
waste depend on the specific conditions and local situation of each country or culture [1].

One of the main areas of action to reduce food losses and waste is the improvement of
conservation technologies. However, solutions to minimize losses usually involve greater energy use,
especially in the conservation of food products. Of course, from an environmental point of view, the
negative impacts of measures to reduce food losses and waste should be less than the benefits.
Therefore, the technological proposals to reduce the losses and waste of food should be focused on
integral and sustainable solutions, such as solar drying that uses the energy from the sun to remove
moisture from the products by heat and mass transfer mechanisms [2]. In spite of some barriers,
renewable energy technologies have been predicted to be a suitable strategy for food drying which
brings some benefits such as low cost, high efficiency, increasing employment opportunities, etc. [3].
Drying food products by solar energy can reduce up to 27% -80% of the cost of fossil fuels [4].

Among the various types of solar dryers, distributed (indirect) forced convection dryers have been
reported to be superior in drying speed, drying quality [5] [6] and are the most suitable for drying large
amount of food [7]. However, solar dryers have the inconvenient of depending on the intensity of
solar radiation that is not always available and is intermittent. Hybrid solar dryers, wherein solar
energy is combined with other sources of energy, such as fossil fuel, biomass solid fuel, and electrical
energy, are used as an alternative solar energy source to address the abovementioned disadvantage
[8]. Additionally, solar drying systems can incorporate solar thermal storage systems to reduce
consumption of fuel, helping to decouple food prices from the fluctuating prices of finite fossil fuels.

A literature review on indirect type solar dryers for food drying can be found in ([6], [9]). Various
types of hybrid solar dryers have been designed and developed using electric heaters ([10], [11],[12]),
biomass ([13], [14], [15], [16]) and LP or natural gas burners ([17], [18], [19], [20], [21], [21], [22], [23])
to dry various food products. Hybrid solar dryers that integrate a thermal storage system have also
been developed ([24], [25], [26], [27], [28], [29], [30]). The previous works are characterized in that
the majority of solar dryers have a low load capacity (< 50 kg) and a solar collection area of less than
20 m2. Furthermore, most of them did not evaluate the operation of product drying continuously
during the hours without sun. Consequently, such designs are not appropriate for food processing on
a semi-industrial scale, where the energy supply must be constant to ensure continuous operation
and high production. Moreover, thermal performance of a single convective food dryer under
different combination of heating sources is a missing part as well [23].

It is recognized that drying technologies have been extensively investigated. However, very little
information is available on the development and research of demonstrative solar drying systems with
a focus on high capacity agro-industrial applications.

Table 1 shows the summary of some hybrid forced convection solar drying systems with areas of
solar collection greater than 30 m2, reported in the technical literature from 1993 to 2016.
Furthermore, none of the works presented incorporates a thermal storage solar system to supply
energy in continuous drying operations. The reviews carried out reaffirm that the solar dryer on an
industrial scale has not been sufficiently studied and further research to optimize appropriate solar
dryers is required [31].

Table 1 Semi-industrial scale solar drying systems with energy backup and batch drying
Type and Average
Type solar Chamber Energy Solar
Designer/s area of solar chamber
dryer capacity backup fraction
collector temperature
Arata and 1000 kg of
SAH at the entrance:
Sharma (1993) Distributed peaches LP gas 82.5 %
129.6 m2 61 °C
[32] (cabinet-type)
Arata and 340 kg of
Greenhouse at the entrance:
Sharma (1993) Distributed peaches LP gas 80.9 %
180 m2 50.2 °C
[32] (cabinet-type)
Palaniappan and SAH (Roof-
30 ton/month
Subramanian Distributed integrated) --- Coal 25 %
of tea (tray-
(1998) [33] 212 m2
type or fluid
bed-type)
Janjai (2012) Greenhouse 1000 kg of
Integral 55°C LP gas ≈ 80 %
[34] 160 m2 of floor tomato (table)
Fudholi et al. SWH ≈ 33.6 m 2 200 kg of 44 °C with 626
Mixed Electric 60 %*
(2015) [35] -Greenhouse catfish W/m2.
Mixed-Solar 231 kg of
SWH 31.5 m2 -
Misha et al. assisted solid crushed oil
transparent 50 °C Electric 66 %
(2016) [36] desiccant palm fronds
roof
dryer (cabinet-type)
* Considering electric consumption, SAH: Solar air heater, SWH: Solar water heater

Given the preceding, a demonstration pilot hybrid solar-LP gas drying plant was developed and
installed. The plant was designed to test the technical and economic feasibility of the systems for
commercial use, exploiting national manufacturing base. The plant is unique due to its size and the
use of two different solar thermal systems (solar air heaters and flat-plate solar water heaters) and a
conventional LP gas support. The systems can work independently or coupled to be able to evaluate
the performance of the plant under different ways of operation (see Fig. 1).

Fig. 1 Photography of the solar thermal plant for the dehydration of agricultural foods.

The plant is monitored in real-time (in time intervals of 1 minute) recording process variables
(humidity, temperature, mass flow rates, irradiance, air velocity, etc.) for the evaluation of the energy
balances in each system. In previous work, a detailed assessment of the thermal performance of the
direct air heating system of this plant was presented [37]. The present article focuses on the evaluation
of the hybrid operation of the solar-LP gas pilot plant. It is evaluated the thermal and fluid dynamic
behavior of both solar technologies (direct and indirect heating systems) and the conventional one (LP
gas backup) to ascertain their thermal efficiency and their effectiveness in drying Nopal. The Nopal
(Opuntia ficus-indica) is a vegetable native to the desert areas of northwestern Mexico and is available
throughout the year, its post-harvest waste is very high, around 50 % [38]. Nopal cladodes can be
stored canned and consumed as juices, or stored as dehydrated powder, which is high in dietary fiber
[39]. Andreu-Coll et al. concluded that cladodes have a beneficial fatty acid profile, and a high potential
for their use as food and for pharmaceutical products [40]. Therefore, solar drying offers a convenient
alternative to avoid waste and give added value to the product. Some works report drying by forced
convection at the laboratory level ([38], [41]) and small-scale solar drying [42]. However, there are no
works that evaluate the solar drying of the Nopal from an energy point of view that provides data for
its subsequent industrial implementation, such as this work. The performance of the solar systems is
finally compared here with a conventional LP gas drying system via a cost and return on investment
analysis.

2. Materials and methods


2.1. Description of the plant
The plant has a distributed-type forced convection hybrid solar drying system with the following
components:
 Drying chamber
 Direct air heating system
 Indirect air heating system
 Conventional system

Next, the elements of the distributed-type forced convection hybrid solar drying system of the
plant are described.

2.1.1. Drying chamber


The drying chamber is a model manufactured in Mexico originally designed to operate with hot
air produced by an LP gas burner (drying conventional system). The chamber is tunnel type with racks,
and the load of the product to be dried can be done semi continuous or in batch. It has a maximum
capacity of 10 drying racks. Each drying rack has a maximum loading capacity of 120 perforated trays
(45.5 cm x 65.5 cm) equally distributed in three vertical sections. The inside of the tunnel, the racks,
and the trays are made of food-grade stainless steel. Fig. 2 shows the dimensions of the drying
chamber and its monitoring instrumentation.

Fig. 2 Dimensions and measurement instrumentation of the drying chamber.

2.1.2. Direct air heating system


48 air heaters integrate the direct air heating system with an aperture area of 111.1 m2 in an
arrangement of three collectors in series by 16 in parallel [37]. The heaters (type: back-pass covered-
plate [43]) are provided with a tempered glass cover. The surface of the metal absorber has a selective
surface (selective paint) that increases the absorption of solar radiation. An air gap separates the
transparent cover and the absorber plate. The lower surface and the sidewalls of the collector are
insulated with a polyisocyanurate foam plate. The air to be heated flows through channels between
the inner surface of the absorption plate and the insulation layer, with heat transfer through the
backside of the absorption plate.

The ambient air enters the solar collectors by forced convection of the suction of an axial fan (7.5
hp, 5.6 kW) with a frequency converter that regulates airflow. With it, the temperature of the air at
the entrance of the drying chamber. It enables different food products to be dried in the chamber.

The entrance of each row of collectors has an air filter to prevent access to dust, insects and
unwanted material. The hot air from the field of solar heaters is transported by the duct system inland
of the drying chamber. The duct system was insulated with 1 inch (25.4 mm) thickness of fiberglass to
prevent heat losses to the environment. In addition, the insulation layer was covered with corrugated
aluminum sheets to withstand the weather conditions.

2.1.3. Indirect air heating system


The indirect air heating system is integrated by: i) field of flat plate solar water collectors; ii) liquid-
liquid plate heat exchanger; iii) thermal storage tank; iv) water-air finned tube heat exchanger; v)
centrifugal fan; and vi) auxiliary equipment.

The field of flat plate solar collectors for indirect air heating is composed of 40 collectors,
equivalent to an aperture area of 92.4 m2. They are distributed in four rows in parallel; each one
consisted of two arrays of five collectors in parallel connected in series (10 collectors per row).

The field of solar collectors operates by a forced circulation system composed of a closed primary
circuit and an open secondary circuit (see Fig. 3). Each loop works with a 1.5 hp (1.12 kW) pump. The
heat transfer between both circuits is done through a plate heat exchanger; the design heat exchanger
capacity was 40 kW. The indirect heating system can operate in two ways: a) the hot water coming
from the solar collector systems goes through the water-air heat exchanger, or b) the hot water can
be stored in a thermo tank for its later use. A horizontal atmospheric thermal tank with a nominal
capacity of 6000 liters is used for the storage of hot water. The indirect heating of the air is carried out
in a water-air fin and tubes heat exchanger with the nominal heat transfer capacity of 60 kW. To
extract the water stored in the tank and supply it to the water-air heat exchanger is used a pump of ¾
hp (559 W). This pump is connected to a frequency inverter that allows regulating the mass flow rate
of hot water that passes through the heat exchanger. A centrifugal fan (10 hp – 7.45 kW motor) sucks
the air to be heated at the outlet of the water-air heat exchange.
Fig. 3 Indirect air heating system with the instrumentation of its components.

2.1.4. Conventional system


The system consists of a traditional hot air generation unit located in the upper part of the tunnel,
composed of: a) a centrifugal fan and b) a natural or LP gas burner from 122 to 249 kW. The air is
heated by direct mixing with the combustion gases. The conventional hot air generating unit has an
automatic control system that regulates the temperature required at the entrance of the drying
chamber. The hot air generation unit can work as an energy backup system when continuous drying
is needed, or the intermittency of the sun does not allow it. In this case, it is operated in hybrid mode.

2.2. Operation modes


The integration of the different technologies offers a system of a generation capable of adapting
quickly and efficiently to different ways of operation. Following are some of the operation schemes of
the thermal power generation system: conventional, solar, and hybrid. The traditional mode of
operation is the energy backup system. The solar mode consists in taking advantage of the solar
energy captured by the direct and indirect heating systems of air, either independently or in
combination. The power of one or both solar systems is combined with a traditional system in the
hybrid mode; it allows to perform a continuous drying process.

2.3. Monitoring and instrumentation


The operation variables of the solar drying system are monitored simultaneously using a data
logger. The measurement instruments used, as well as their measurement range and accuracy are
reported in Table 2.

Table 2 Instrumentation used to measurement of variables of the solar drying system.


Variable Instrument / sensor Range Accuracy
Liquid volumetric flow Impeller flowmeters: Seametrics,
2 to 150 l/min ± 1 % FSO
meter model SPT-100, and SPX-100.
Hotwire anemometer: TSI, model
Air velocity 0 to 30 m/s ±3%
8345
Diaphragm Meter – Elster model AL-
LP gas meter 0 to 41.4 m3/h ±1%
425.
Precision balances
Mass 0 to 4200 g ±0.2 g
Nimbus, model NBL 4201e
Mass Weighing scale: Torrey, model PCP. 0 to 250 kg ± 0.1 kg
Moisture analyzer - OHAUS, model
Moisture content 0.01 to 100 % ± 0.01 %
MB45
Fluke, model 323 True RMS Clamp 0 to 600 V ± 1.5 %
Voltage and current
Meter 0 to 400 A ±2%
Pyranometer: Kipp & Zonen, model 0 to 2 000
Irradiance ±2%
CMP3 W/m2
±2%
Temperature and -20 to 60 °C and
Ibérica, model PCE-P18 ± 0.5 % and
relative humidity 0 to 100 %
±2%
Temperature (data Water-Proof Thermometer Cole-
5 to 80 °C ± 0.1 °C
loggers) Parmer, model 90205-22
Temperature RTD, class PT-1000 -50 to 750 °C ± 0.5 °C
Full-Scale Output (FSO) is the resulting output signal or displayed reading produced when the maximum measurement for a given device is applied.

The pyranometer was installed on the solar collector plane to measure the solar irradiance. Fig. 4
shows the distribution of some of the sensors used to monitor the different variables.

Fig. 4 Schematic representation of the distribution of some of the measurement sensors installed in the plant.

2.4. Methodology for the energy evaluation of the system


Experimental data obtained from fluid flow temperatures (air and water) in different places, mass
flow rates, the moisture removed from the drying product and the energy inputs (solar, combustion
and electrical) were used to perform the energy analysis of the drying system. The mass flow rate of
water is determined directly with mass flow meters. However, the air mass flow rate was determined
indirectly by the measurement of the average air velocity inside ducts. The volumetric airflow (𝑽) is
constant if the fan rotor operates at a constant speed of rotation. Considering that there are no leaks
in the system, the volumetric flow entering the collector system is the same as the output, 𝑽 = 𝑉𝑖𝑛 =
𝑉𝑜𝑢𝑡. The air mass flow rate was determined via the following expression:

𝒎𝒂𝒊𝒓 = 𝝆𝒂𝒊𝒓 𝝊𝒂𝒊𝒓 𝑨𝒄 ― 𝒔 (1)


Where 𝝊𝒂𝒊𝒓 is the measured air velocity at the exit of the air heaters system, 𝑨𝒄 ― 𝒔 is the duct cross
section area and 𝝆𝒂𝒊𝒓 is the air density.The operation of the field of solar collectors involves the
incident solar energy and the electrical consumptions (fans or pumps). The incident solar energy on
the collector plane was evaluated via the following expression:
𝑵

𝑬𝒊𝒏𝒄 = 𝑨𝑪 𝚫𝒕 ∑𝑰
𝒊=𝟏
𝒊 (2)

Where 𝐴𝐶 is the aperture area of solar collectors, 𝐼 is the solar irradiance on the collector plane,
𝚫𝒕 is the time interval of each measurement, and 𝑵 is the number of time intervals used. The electrical
energy (𝑬𝒆𝒍𝒆𝒄) consumed by the motors coupled to the auxiliary equipment was determined by:

𝑬𝒆𝒍𝒆𝒄 = 𝟑 ∙ 𝑨 ∙ 𝑽 ∙ 𝑷𝑭 ∙ 𝒕𝒐𝒑 (3)

Where 𝑨 is the electrical current, and 𝑽 is the electrical voltage that feeds the three-phase motor,
𝑷𝑭 is the power factor, and 𝒕𝒐𝒑 is the fan or pump operating time. The energy gain transferred to a
thermal fluid (useful energy gain) by the field of solar collectors was determined via the following
expression:
𝑵

𝑬𝒖 = 𝚫𝒕 ∑ [𝒎
𝒊=𝟏
𝑻 𝑪𝒑𝒎 (𝑻𝒐𝒖𝒕 ― 𝑻𝒊𝒏)]𝒊 (4)

Where 𝐶𝑝𝑚 is the average heat capacity at a constant pressure of the fluid, 𝑇𝑖𝑛 and 𝑇𝑜𝑢𝑡 are the
inlet and outlet temperatures of the field of solar collectors, respectively. The instantaneous thermal
efficiency of the field of solar collectors was defined as the ratio between the useful heat gain and the
incident solar power on the collector plane and was calculated via the following expression:

𝒎 𝑪𝒑𝒎 (𝑻𝒐𝒖𝒕 ― 𝑻𝒊𝒏)


𝜼𝒕𝒉 = (5)
𝑰 𝑨𝑪

The total energy inflows are the sum of the solar energy incident on the collector plane and the
electrical energy consumed by auxiliary equipment. The following expression determined the overall
energy efficiency of solar energy collection systems considering the consumption of electrical energy
of auxiliary equipment (fan or pump):
𝑬𝒖
𝜼𝒐,𝒆 = (6)
𝑬𝒊𝒏𝒄 + 𝑬𝒆𝒍𝒆𝒄

The total mass of water evaporated (𝒎𝒘) of the drying product was calculated as follows:

𝒎𝒑(𝑴𝒊 ― 𝑴𝒇)
𝒎𝒘 = (7)
𝟏𝟎𝟎 ― 𝑴𝒇

The theoretical thermal efficiency of the dryer, the ratio of the actual temperature drop to the
maximum possible temperature drop of the drying air in the drying chamber, was evaluated via the
following form [44]:

𝑻𝒊𝒏 ― 𝑻𝒐𝒖𝒕
𝜼𝒕𝒉,𝒅 = (8)
𝑻𝒊𝒏 ― 𝑻𝒂𝒎𝒃
Where 𝑇𝑖𝑛 and 𝑇𝑜𝑢𝑡 are air temperatures at the inlet and outlet section of the drying chamber,
respectively, and 𝑇𝑎𝑚𝑏 is the ambient temperature.

Drying efficiency is defined as the ratio of the energy required to evaporate the moisture from the
fresh product to the total energy supplied to the drying system (it takes into account the energy
consumed by the fan),

𝝀𝒎𝒘
𝜼𝒐,𝒅 = (9)
𝑬𝒊𝒏, 𝒄𝒉 + 𝑬𝒆𝒍𝒆𝒄

Where 𝝀 is the heat of vaporization of water (measured via the average air temperature at the
inlet and outlet section of the drying chamber), and 𝑬𝒊𝒏, 𝒄𝒉 is the energy delivered to the entrance of
the drying chamber.

Another parameter used to evaluate drying performance is moisture pick-up efficiency. It is a


direct measure of how efficiently is the capacity of the air to absorb moisture. It is expressed as the
ratio of the moisture absorbed by the air to the theoretical capacity of the air to absorb moisture [45],

(𝒀𝒐𝒖𝒕 ― 𝒀𝒊𝒏)
𝜼𝒑 = (10)
(𝒀𝒂𝒔 ― 𝒀𝒊𝒏)

Where 𝒀𝒐𝒖𝒕 , 𝒀𝒊𝒏 are the absolute humidity at the inlet and outlet section of the drying chamber,
respectively, and 𝒀𝒂𝒔 is the absolute humidity at the inlet section of the dryer at the point of adiabatic
saturation.

2.5. Uncertainty analysis


The uncertainty analysis was developed according to the work of Moffat [46] that covers the
method of describing the uncertainties in an engineering experiment and the necessary background
material. If the uncertainties in the independent variables are knows, then the uncertainty in the result
(𝑅) will be given via the following Eq. (11):
12

[∑ (
𝑁

)]
2
∂𝑅
𝛿𝑅 = ∙ 𝛿𝑥𝑖 (11)
∂𝑥𝑖
𝑖=1

The maximum uncertainties of the essential dependent parameters are given in Table 3.

Table 3 Maximum uncertainty in the dependent parameters.


Parameters Uncertainty
Uncertainty in the air mass flow rate -direct air heating- (kg/s) ± 0.03
Uncertainty in the air mass flow rate -indirect and conventional systems- (kg/s) ± 0.05
Uncertainty in the water mass flow rate (kg/s) ± 0.02
Uncertainty in the useful power -direct air heating- (kW) ± 1.56
Uncertainty in the useful power -indirect air heating- (kW) ± 1.73
Uncertainty in the thermal efficiency of the field of SAHs (%) ± 1.74
Uncertainty in the thermal efficiency of the field of SWHs (%) ± 6.99
Uncertainty in the thermal efficiency of the drying chamber (%) ± 2.49
3. Results and discussion
In 2018, two drying tests were carried out operating in hybrid mode: on May 24 (test 1) and on
September 26 (test 2). Tests were carried out operating the direct air heating system in the morning;
at the same time, the water heating system stored energy that will be used later. When the solar
resource of both solar systems was exhausted during the drying process, air heating was carried out
through the LP gas heating system (conventional system).

3.1. Description of meteorological conditions during the tests


The dehydrating plant is located in the municipality of Morelos, Zacatecas, Mexico at latitude 22°
53' N and longitude 102° 39' W, at an altitude of 2207 m above sea level. The climate is semi-desert
with a mean annual temperature of 14.8 °C, a mean yearly rainfall of 407.7 mm concentrated from
June to September (75.66 % of precipitation relative to the annual and 37 days on average of rain on
this period) [47]. The yearly average wind speed is 4 m/s, predominant in the south and south-east
direction. The mean annual daily irradiance is 520 W/m2 [48]. Therefore, it is one of the places with
the highest solar yearly irradiation in Mexico.

Table 4 shows the meteorological parameters recorded during the tests. During the May test, an
average difference between the maximum and minimum value of the ambient temperature of 21.7
°C was obtained. While in September, it was 15.58 °C. These thermal oscillations, together with the
ambient relative humidity, cause variations of the energy demand for the drying process. A minimum
ambient humidity of 8.41 % was recorded for the May drying test (one of the driest months). While
on September test, it was 40.21 % (a rainy month) both registered around solar noon.

Table 4 Summary of the meteorological parameters during the tests.


Test 1 Test 2
Parameters
23-May 24-May 25-Sep 26-Sep
Average ambient temperature* (°C) 20.37 21.57 17.35 17.66
Maximum temperature* (°C) 29.94 31.04 25.89 26.50
Minimum temperature* (°C) 8.74 8.79 9.51 11.73
Average relative humidity* (%) 16.18 23.97 64.88 64.02
Maximum relative humidity* (%) 28.30 48.98 85.94 86.08
Minimum relative humidity* (%) 9.57 8.41 41.25 40.21
Hours of the sun (h) 13.21 13.23 11.90 11.88
Total average irradiance on the collector plane (W/m2) 606.65 601.81 645.15 637.92
Maximum irradiance on the collector plane (W/m2) 992.51 984.82 1126.49 1176.05
Minimum irradiance on the collector plane (W/m2) 26.28 26.07 27.68 27.37
* Registered 24 hours a day

Sunshine hours for May are higher than in September. However, due to the inclination of the
collectors. On September 26 test, the sun hours (11.88 hours) coincide with the productive hours of
solar incidence on the collector surface; however, for May, the productive hours are reduced to 11.98
hours. The solar collectors are oriented to Ecuador with an inclination of 22.72 ± 0.94 ° (water heating)
and 23.49 ± 0.84 ° (air heating); which is practically the same as the geographical latitude of the place
(22.89 °). In September test, the angles of incidence on the collector plane are almost zero (≈0 °),
around the solar noon; the opposite case for the May test (≈19 °). Therefore, the solar irradiance on
the collector plane was higher for the test carried out in September despite being a little cloudy day,
as it can be seen in Table 4.
3.2. Evaluation solar water heating system
The indirect air heating system began operating one day before the drying test to achieve stored
energy capacity. For both tests, the volume of water contained in the tank was 6150 l. The water
heating system was left in automatic mode using an automatic differential control with the following
parameters: switched on difference 8 °C, switched off difference 4 °C and switched off with a
temperature higher than 90 °C.

The test carried out on May is described below. During the first day, the pumps of the water
heating system operated continuously from 7:35 to 16:05 h. In this time interval, the useful incident
solar energy on the surface of the field of collectors was 2196.86 MJ. However, only 1231.02 MJ was
removed by the thermal fluid (useful energy); the overall thermal efficiency of the field of collectors
was 56.04 %. The temperature of the thermo tank increased 42.71 °C, equivalent to 1098.69 MJ or
50.01 % (overall system efficiency) of the total incident solar energy was stored in the thermo tank
during both days. During the waiting period for the next day's operation (mainly at night), the tank
had a thermal loss of 1.99 °C, equivalent to 51.49 MJ. On the second day of operation, the pumps
operated continuously from 8:42 to 14:22 h, increasing 22.94 °C the temperature of the thermo tank.
It is equivalent to 592.24 MJ; the water heating system during the second day contributed 36.12 % of
the total energy stored in the thermo tank. 14.22 effective hours (pumps on) were required to heat
the water in the thermo tank from 27.76 °C to 91.43 °C. Fig. 5 shows for the water heating system, the
temperature profiles: at the entrance (TW-1), in the series connection of the north row (TW-2) and at
the exit (TW-6) of the field of solar collectors. It also shows the progression of the temperature in the
thermo tank (TW-11), the solar irradiance on the collector plane and the ambient temperature for the
two days of operation of the system. The rate of the increase in the temperature of the thermo tank
is significantly lower on the second day. This behavior is due to higher temperatures generate
significant thermal losses, and the efficiency of the field of collectors decreases.

Fig. 5 Temperature profile in the field of collectors, thermo tank and ambient; as well as solar irradiance during the
days of the May test.
Fig. 6 shows the global instantaneous thermal efficiencies of the field of SWHs; as well as the
energy accumulated in the thermo tank during the two days of operation of the May and September
tests. The instantaneous efficiency was determined by Eq. (5); efficiencies gradually decrease during
water heating because thermal losses increase with increasing temperature stored in the thermo tank.

Fig. 6 Instantaneous thermal efficiencies and energy stored during the two days operation.

Table 5 shows a summary of the relevant operating parameters of the SWHS for the two test days.
It is observed that the water flows on the primary and secondary circuits are significantly higher for
the tests carried out during September, due to modifications in the pipeline. The power of the energy
storage system was similar in both cases: 33.04 kW (test 1) and 33.4 kW (test 2).

Table 5 Operation parameters of the SWHS during the tests.


Test 1 Test 2
Parameters
23-May 24-May 25-Sep 26-Sep
Start of operation (hh:mm) 07:35 08:28 07:24 08:24
End of operation (hh:mm) 16:05 14:22 15:26 14:36
Effective operation time (h) 8.52 5.70 7.68 5.37
Average water mass flow rate in the
99.96 ± 1.71 99.92 ± 1.33 112.4 ± 1.81 115.46 ± 1.78
primary circuit (kg/min)
Average water mass flow rate in the
64.95 ± 0.84 57 ± 2.03 73.87 ± 0.62 67.76 ± 6.06
secondary circuit (kg/min)
Electric power consumed by the primary
48.01 32.13 43.31 30.25
circuit pump (Eq. (3)) (MJ)
Solar energy incident (Eq. (2)) (MJ) 2196.86 1652.71 2228.21 1830.00
Useful energy gain at the outlet of the field
1231.02 787.97 1175.70 878.59
of SWHs (Eq.(4)) (MJ)
Average instantaneous thermal efficiency
60.75 ±1.47 49.49 ±1.30 56.72±1.27 50.66±3.21
of the field of SWHs (Eq. (5)) (%)*
The overall energy efficiency of the field of
54.84 46.77 51.76 47.23
SWHs (Eq. (6)) (%)
Initial water temperature in thermo tank
27.76 68.48 31.39 69.38
(°C)
Final water temperature in thermo tank
70.48 91.43 71.94 89.76
(°C)
Stored thermal energy (MJ) 1098.69 592.24 1043.40 525.77
The energy lost at night (MJ) 51.49 --- 65.70 ---
SWHS global efficiency (stored thermal
50.01 35.83 46.83 28.73
energy/solar energy incident) (%)
*Evaluated from 11 to 13 h (solar time)

3.3. Evaluation of the drying process


Drying tests were performed on May 24 and September 26. The results obtained in the different
modes of operation of the air heating required in the drying process are presented below.

3.3.1. Direct air heating system


The drying process began with a direct air-heating operation mode. For both tests, the axial fan
was operated at 60 Hz, generating volumetric airflow of 4663.5 m3/h. A criterion to switch on the
system an irradiance over 600 W/m2 on the collector plane. This criterion ensures that the ambient
temperature is increased by ≈ 30 °C during the first minutes of starting. Typically, this criterion allows
around the year begin the operation three hours before solar noon. When the air temperature at the
global output of the field of collectors was below 50 °C, the system switched off. Table 6 shows the
operation parameters of the direct SAHs.

Table 6 Operation parameters of the direct SAHs


Test 1 Test 2
Parameters
24-May 26-Sep
Start of operation (hh:mm) 09:00 09:20
End of operation (hh:mm) 15:52 15:19
Effective operation time (h) 6.88 6.00
Average ambient temperature (°C) 29.66 24.09
Average air temperature at the outlet of field of SAHs (TA-19) (°C) 64.77 63.21
Solar energy incident (Eq. (2)) (MJ) 2305.56 2219.50
Fan electrical power consumption (Eq. (3)) (MJ) 116.01 101.12
Useful energy gain at the outlet of the field of SAHs (Eq.(4)) (MJ) 969.71 980.91
Average the instantaneous thermal efficiency of the field of SAHs (Eq.
42.98±0.58 43.13±2.03
(5)) (%)*
The overall energy efficiency of the field of SAHs (Eq. (6)) (%) 40.04 42.27
Useful power (kW) 39.13 45.41
*Evaluated from 11 to 13 h (solar time)

Fig. 7 shows the instantaneous thermal efficiencies of the field of SAHs (Eq. (5)), the air
temperature profile at the global outlet temperature of the field of collectors (TA-19, see Fig. 4) and
the irradiance during the two drying tests. It can be seen that the efficiencies and the temperatures
at the exit are practically the same for the two tests in the period of stable irradiance (clear sky).
However, the average temperature increment (ΔT) in the stable period was higher in test 2 (44.71 °C)
than in test 1 (39.13 °C). This difference is due to the ambient temperature for test 2 was lower than
that in test 1, as can be seen in Table 6. The useful energy power delivered by the direct air heating
system was higher in September (45.41 kW) due to that the higher irradiance and the increases of the
temperature of the air in a time relatively lower than the May test.
Fig. 7 Instantaneous thermal efficiencies, outlet temperature, and solar irradiance for the two test days.

The instantaneous thermal efficiency of the field of solar collector was determined considering the
aperture area and the total mass flow of the field of solar collector. In the case of the SWHs field, the
average efficiency on May 23 was 60.75 ± 1.47 % with a water mass flow of 1.67 ± 0.02 kg/s and an
average outlet temperature (TW-6) of 67.66 °C (from 61.77 to 72.3 °C, evaluated from 11 to 13 h, solar
time). Meanwhile, for the SAHs the average efficiency on May 24 was 42.98 ± 0.58 % with an air mass
flow of 1.11 ± 0.08 kg/s and an average outlet temperature (TA-19) of 70.85 °C (from 67.66 to 72.11
°C, from 11 to 13 h). Table 7 summarizes the characteristics and operating parameters of solar systems
from some works that evaluate distributed or mixed type solar drying systems with solar collection
areas greater than 30 m2 and with a drying chamber capacity greater than 100 kg of fresh product
(agricultural and food applications). With these characteristics, there is only a solar-assisted drying
system using SWHs (evacuated tubes) reported by Misha et al. ([36], [49]), with efficiencies of the field
of SWHs of 59 % and 56 %, it is similar to the efficiencies found in the present work with flat plate
solar collectors (60.75 ± 1.47 %). Meanwhile, the efficiencies of the field of SAHs found in this work
(42.98 ± 0.58 %) were higher than those reported in previous works, with the exception of that
reported by Sreekumar [50], which was 52.55%.

Table 7 Semi-industrial scale solar drying systems


Field of Temperature of
Collector materials Average
Type and area of collectors fluid at the outlet
Designer/s Type solar dryer (cover/absorber/ chamber
solar collector efficiency of field of solar
insulation/box) temperature
/ flow collectors
Arata and Distributed-1000 kg SAH, front-pass Glass (3mm), at the
31 % ( ≈ T ≈ 61 °C,
Sharma (1993) of peaches (cabinet- covered-plate, corrugated steel plate, entrance:
0.77 kg/s) 30<ΔT<35°C
[32] type) 129.6 m2 polyurethane (3cm) 61.0 °C
Arata and Distributed-1000 kg Glass (3mm), at the
SAH, front-pass 36 % ( ≈ T ≈ 56 °C,
Sharma (1993) of peaches (cabinet- corrugated steel plate, entrance:
covered-plate, 0.77 kg/s) 30<ΔT<35°C
[32] type) polyurethane (3cm) 56.0 °C
129.6 m2
15.37 %
Arata and Distributed-340 kg at the
Greenhouse collector Thermal plastic, black efficiency T ≈ 50 °C,
Sharma (1993) of peaches (cabinet- entrance:
180 m2 plastic sheet of one 30<ΔT<35°C
[32] type) 50.2 °C
collector
Tunnel type chamber
Mixed-300 kg of Polyethylene plastic at noon:
Condori et al. inside of greenhouse ΔTmax = 25 °C at
fresh peppers (150 μm), soil covered Tmax ≈ 60 °C 25 % (1
(2001) [51] type collector noon
(tunnel-type) with black plastic kg/s)
75.4 m2
Transparent corrugated
Distributed- 200 kg SAH, roof-integrated, Tmax ≈ 60 °C at
Janjai and fiberglass cover,
of rosella flowers front-pass covered- noon, 5<ΔT<20 °C
Tung (2005) galvanized iron sheet --- 35 %
and lemon-grasses plate from 9:00 to
[52] painted black, glass
(cabinet-type) 72 m2 17:00 h
wool, iron frames
SAH, roof-integrated, Tempered glass, black
Distributed- 200 kg back-pass covered- chrome copper sheet, Tmax = 67.3 Tmax = 76.6 °C at
Sreekumar 52.55 %
pineapples plate, with baffles rock wool (5cm), °C at noon noon, ΔTmax= 43.3
(2010) [50] (0.63 kg/s)
(cabinet-type) transversal structural frame (930 W/m2) °C
46 m2 aluminium
SWH-Evacuated tube.
Mixed-155 kg of Chamber:
Misha et al. SAH- Chamber with Tmax = 75 °C,
kenaf core fiber Polycarbonate-walls, 44 °C 59 %
(2015) [49] transparent roof Secondary circuit
(cabinet-type) Glass-roof.
31.5 m2
Mixed-231 kg of SWH-Evacuated tube.
Chamber:
Misha et al. crushed oil palm SAH- Chamber with
Polycarbonate-walls, 50 °C 56 % ---
(2016) [36] fronds (cabinet- transparent roof
Glass-roof.
type) 31.5 m2
Alveolar polycarbonate
(4 mm), corrugated
galvanized plate 30 % (0.07 80<Tmax<90 °C at
Distributed-450 kg SAH, Suspended- ≈ 50 °C for 6
Condorí painted in matt black kg/s), noon,
of different plate covered-plate: or 7 h of a
et al.(2017) (0.42 mm), glass wool efficiency 50<ΔTmax<60 °C,
vegetables. (type parallel-pass sunny day
[31] with reinforced of one with minimum air
tunnel) 92 m2 (0.07 kg/s)
aluminum foil (5 cm), collector flow of 0.06 kg/s
galvanized sheet box
for support.
SAH, roof-integrated, Between
Distributed-Garner- suspended-plate Transparent (16.1-
Čiplienė et al. Evaluated T= 30.4 °C,
medicinal and spice covered-plate: polycarbonate, sheet of 61.8) %
(2018) [53] without load 7.5<ΔT<14 °C
plants parallel-pass steel painted in black (1.18-
131 m2 2.21) kg/s

3.3.2. Indirect air heating system


After switching off the direct air heating system, the indirect air heating system was switched on
in the storage mode. For both tests, the volumetric flow of air through the finned and tube heat
exchanger was 6367.78 m3/h, operating the centrifugal fan at 35 Hz. Two ways of extracting the energy
stored in the thermo tank were tested. For test 1, the water flow through the exchanger was varied
gradually from 18 to 45 kg/min during the first 40 minutes of operation. Then, the water flow
remained constant at around 55.26 kg/min (maximum pump flow). The energy extraction stopped
when the air temperature at the outlet of the heat exchanger was ≈ 45 °C. For test 2, the water flow
rate that passes through the finned and tube heat exchanger was regulated. This, with the objective
to air temperature at the outlet of the heat exchanger, was maintained between 50 to 55 °C.

The thermal efficiency of the water-air heat exchanger was determined via the following Eq. (12):

𝒎𝒂𝒊𝒓 𝑪𝒑𝒎, 𝒂𝒊𝒓 (𝑻𝒐𝒖𝒕 ― 𝑻𝒊𝒏)𝒂𝒊𝒓


𝜼𝒕𝒉,𝒆𝒙𝒄 = (12)
𝒎𝒘𝒂𝒕𝒆𝒓 𝑪𝒑𝒎, 𝒘𝒂𝒕𝒆𝒓 (𝑻𝒐𝒖𝒕 ― 𝑻𝒊𝒏)𝒘𝒂𝒕𝒆𝒓

Fig. 8 shows the thermal efficiencies of the water-air heat exchanger or finned and tube heat
exchanger (ƞ), the temperature profile of the air at the outlet of it (TA-35), the increase in air
temperature (ΔT_A) and the pattern of temperature inside the thermo tank during hot water
extraction (TW-11), as well as the variation of the water flow rate through the heat exchanger (FW-4)
during the two drying tests. An abrupt decrease in the temperature inside the thermo tank can be
observed when the flow of hot water extraction is higher than 30 kg/min. This decrease is due to more
significant mixing of water breaks stratification temperatures inside the thermo tank. It can be seen
in Fig. 8 that during test 2 the stepwise increase in the flow rate of water through the heat exchanger
allowed to maintain more time the temperature of the air at the outlet section of the heat exchanger
around 55 °C.

Fig. 8 Thermal efficiencies and air temperature in the heat exchanger; as well as the temperature in the thermo
tank.

In test 1, 3.05 continuous hours were operated maintaining the air temperature at the outlet of
the exchanger above 50 °C, with a maximum water flow rate of 55 kg/min. In test 2, regulating the
flow rate of water through the exchanger, 3.53 continuous hours were operated with temperatures
above 50 °C increasing air temperature (ΔT_A) above 25 °C (see Fig. 8). Higher air temperatures are
obtained at the outlet of the heat exchanger using the maximum water flow rate. However, this can
be harmful to some products that require drying temperature control. In test 1, air temperatures at
the outlet of the heat exchanger above 70 °C were obtained.

Table 8 shows the operation parameters of the indirect air heating system by extracting the energy
stored in the thermo tank. In test 2, the indirect system operated 1.82 hours less than in test 1 and
with a stepped extraction flow rate. Therefore, in test 1, more energy was extracted from the thermo
tank, and there is a lower temperature inside the thermo tank at the end of the test. The water
temperature in the thermo tank drops, contributing the indirect air heating system with a total energy
of 1006.44 MJ (test 1) and 726.78 MJ (test 2), this is equivalent to 61 % and 48 % of the energy stored
during the two days of operation of the water heating system, respectively. The thermal efficiency in
the water-air heat exchanger was higher when it regulated the flow rate of water through the heat
exchanger because temperatures were maintained lower compared with the maximum water flow
rate.

Table 8 Operation parameters of indirect air heating system.


Test 1 Test 2
Parameters
24-May 26-Sep
Start of operation (hh:mm) 15:57 15:30
End of operation (hh:mm) 21:17 19:01
Effective operation time (h) 5.35 3.53
Initial water temperature in the thermo tank (°C) 90.97 89.20
Final water temperature in the thermo tank (°C) 51.94 61.02
The energy delivered by the thermo tank (MJ) 1006.44 726.79
The energy delivered by the water in the heat exchanger (MJ) 951.91 648.82
Average air temperature in the inlet section of the heat exchanger (°C) 30.74 25.97
Average air temperature at the outlet section of the heat exchanger (°C) 57.86 55.03
Energy received by the air in the heat exchanger (MJ) 790.68 568.85
Average heat exchanger thermal efficiency (Eq. (12)) (%) 82.91±2.95 87.82±3.05
Useful power (kW) 41.23 44.92
Fan electrical power consumption (Eq. (3)) (MJ) 55.19 36.45
Pump electrical power consumption (Eq. (3)) (MJ) 6.61 2.84

3.3.3. Conventional system


Once the energy from the direct air heating system was finished, and without the available solar
resource, the traditional air heating system was switched on. The airflow used by the conventional
system is generated by the same centrifugal fan of the indirect air heating system with a volumetric
airflow of 6367.78 m3/h. LP gas is a mixture integrated in Mexico by approximately 60 % propane and
40 % butane. The lower calorific value of LP gas is equivalent to 46.16 MJ/kg [54]. Table 9 shows the
operation parameters of the conventional air heating system. Not all of the energy available in LP gas
is used to heat the air; this is due to incomplete combustion of gas and heat loss to the outside before
entering the drying chamber. Therefore, the thermal efficiency of the conventional system is around
80 %. The heating of the air in the traditional system is due to the mixing of fresh air with flue gases
of the LP gas. Therefore, the humidity of the air entering the conventional system increases its
moisture content due to the water formed during combustion.

Table 9 Operation parameters of the conventional system


Test 1 Test 2
Parameters
24-May 26-Sep
Start of operation (hh:mm) 22:25 19:09
End of operation (hh:mm) 00:28 21:18
Effective operation time (h) 2.07 2.17
Average air temperature at the heat exchanger (°C) 24.40 22.39
Set point temperature of the LP gas burner (°C) 60 50-54
LP gas consumption (kg) 12.81 11.54
Available energy in the LP gas* (MJ) 589.81 531.30
Total energy delivered by the air (MJ) 464.46 428.72
Useful power (kW) 62.42 54.96
Conventional system efficiency (%) 78.75 80.69
Fan electrical power consumption (Eq.(3)) (MJ) 21.32 22.35
* Calorific power of LP gas (46.16 MJ/kg [54])

3.3.4. Drying chamber


The drying tests were evaluated with Nopal (Opuntia ficus) from the municipality of Villa Nueva
located in the southwest of the state of Zacatecas, in central-northern Mexico. For processing, they
were washed, the thorns removed by passing a knife along the surface of the cladode and then cut
into pieces of maximum 15 cm x 3 cm. After that, they were cut in a food processor using a 5 mm thick
cutting disc. The product was placed in 300 perforated stainless steel trays distributed in 5 drying
racks; the loading density was ≈ 4.72 kg/m2 (fresh product).

The average initial moisture content of the Nopal was 88.27 % (test 1) and 81.29 % (test 2) (wet
base), obtained with a moisture analyzer. The Nopal used in the two tests was from the same harvest
lot. The moisture content of the Nopal of test 2 has lower humidity because it was in storage longer
before the drying process. The average final moisture content of the dried Nopal was 7.24 % (test 1)
and 8.34 % (test 2) (wet base). A total of 179.05 kg of water was evaporated in test 1 and 183.15 kg
in test 2.

Fig. 9 shows that for the direct air heating system, the average and maximum temperatures at the
entrance of the drying tunnel was 65.35 °C (TA-21) and 69.85 °C (close to solar noon) respectively. The
temperatures registered in TA-29 are very different from those recorded in the same cross-section
(TA-22, TA-30, TA-31, and TA-32) during the first hours of drying. This difference disappears as the test
progresses until the four temperatures equalize. Temperatures at the lowest points of the shelf (TA-
31 and TA-32) were similar in all tests. The above results indicate that the airflow is not homogeneous
in the cross-sectional area. The difference in the airflow in the transverse direction represents an
aspect that must be improved in the design of the drying chamber, to obtain a homogeneous
dehydrated product.

The temperatures inside the drying tunnel (TA-22, TA-29, TA-30, TA-31, and TA-32) approach each
other after 6.5 hours of the start of the operation. The indirect air heating system began operating
after 6.88 hours of drying. The average air temperature at the tunnel entrance was 56.69 °C and was
turned off when was close to 44 °C. An inspection of the average moisture content of the product was
carried out on the drying racks. This inspection lasted 1.1 h, after that, the traditional air heating
system (LP gas burner) was switched on to continue the drying process. In this mode of operation, the
average temperature at the tunnel entrance was 61.27 °C.

Fig. 9 Temperature profiles in the drying chamber for May test.


Table 10 shows the operation parameters of the drying chamber during the operation of each air
heating system. At the exit of the drying chamber, the average relative humidity of air is very low (less
than 32 %), and the average temperatures are higher than the adiabatic saturation temperatures (24
°C). The above mentioned indicates that the air in the outlet section still has useful energy to remove
moisture and, therefore, the drying chamber has a low drying efficiency. The energy demand in a
batch drying process is higher at the beginning of the operation and gradually decreases. However, at
the end of the process, it is useful to maintain high permissible temperatures to have high drying rates
and, therefore, shorter drying times. The above mentioned affects the efficiency of drying, since at
the end of the operation, a large part of the energy is not used, and it is discharged to the environment.
In order to increase the drying efficiency, the load of the product can be increased, or it is possible to
replace the dry product with fresh product.

Table 10 Operation parameters in the drying chamber.


Test 1 Test 2
Parameters
Direct Indirect Conventional Direct Indirect Conventional
Thermal energy loss in transport
-39.19 -33.07 -125.34 -46.93 -24.12 -102.57
to the tunnel (MJ)
Average temperature at the inlet
63.35 56.74 63.79 61.31 53.72 56.73
of the tunnel (°C)
Average relative humidity at the 3.25 4.98 6.45 11.01 10.23 10.88
inlet tunnel (%)
Thermal energy delivered at the 930.52 757.61 464.46 933.98 544.73 428.72
inlet of the tunnel (MJ)
Average temperature at the outlet
41.70 51.41 54.20 39.85 46.32 48.27
of the tunnel (°C)
Average relative humidity at the 17.40 5.81 7.24 31.13 14.47 12.74
outlet tunnel (%)
Water removed (kg) 156.92 18.19 3.94 140.33 36.10 6.72
Average thermal efficiency (Eq.
64.27 20.51 24.35 57.66 26.69 24.63
(8)) (%)
Drying efficiency (Eq. (9)) (%) 35.71 5.28 1.92 32.35 14.76 3.55
Average pick-up efficiency (Eq.
33.71 4.29 2.13 43.51 14.19 4.02
(10)) (%)

As is shown in Table 10, the drying efficiencies for the indirect system during test 2 (in which the
energy delivered in the air heating was regulated) were higher than test 1. The thermal efficiency of
drying of the direct air heating system was higher in test 1 than in test 2. This behavior is due to in
both the energy delivered to the tunnel is practically the same, but the amount of water removed in
test 1 is higher than in test 2. This situation is opposite when comparing the pick-up efficiency, in test
2 the capacity to remove water is better used than in test 1; this is due to the air that leaves the drying
chamber still has a higher ability to remove water. Between both efficiencies, the pick-up efficiency
indirectly considers the thermal part required in drying and the capacity of mass removal (dependent
on humidity and air temperature). Therefore, it can be regarded as a complete parameter in the drying
system comparison.

Fig. 10 shows the temperature profiles and absolute humidity of the air at the inlet and outlet
section of the drying chamber for the two tests. The average temperature at the tunnel entrance was
higher than 50 °C. There are two regimes of drying rates, the periods of constant drying rate and
falling drying rate [55]. The highest water removal occurs in the first six hours of drying (constant
drying) and that water removal decreases substantially after eight hours of the process (falling drying).
However, to achieve the desired humidity in the product, it is necessary to continue providing energy
to remove the water of the product. Dependence between air temperature and water removal can be
observed mainly between periods of constant and falling drying rate.

Fig. 10 Temperature and absolute humidity profiles for both tests at the inlet and outlet section of the drying
chamber.

3.4. Summary of the drying process Kumar and Khatak [5] report that the drying
efficiencies of hybrid systems varies from 17 to 29 %, in this work global
values of drying efficiencies varies from 18.09 to 21.1 %.

Table 11 shows the general summary of the operation parameters of the drying tests. In both tests,
the same amount of water was removed. However, the energy delivered to the drying chamber was
much higher in test 1 against test 2 (757.61 vs. 544.73 MJ, see Table 10) due to the more considerable
energy delivered at the entrance of the tunnel by the indirect air heating system. The percentage of
energy provided by the solar systems with respect to the total energy required (solar fraction) was
around 78 % for both tests. This high value is because the drying time of the product is less than 15
hours and the product loading is by batch. When the product load is semi-continuous or continuous;
and also a continuous drying is carried out, the solar fraction could decrease significantly. Arata and
Sharma [32] carried out the drying of peaches in an indirect solar drying system with LP gas backup;
the operation was continuously (25.7 h) and discontinuously (37.2 h for 5 days). They determined that
the solar fraction was 37 % and 82 %, respectively. Therefore, when the product and the production
schedule allow it, it is convenient to operate discontinuously to maximize the use of the solar
infrastructure and reduce fuel consumption. Kumar and Khatak [5] report that the drying efficiencies
of hybrid systems varies from 17 to 29 %, in this work global values of drying efficiencies varies from
18.09 to 21.1 %.

Table 11 Summary of the operation parameters of the drying tests.


Parameters Test 1 Test 2
Average initial moisture content of the product (%) 88.27 81.29
Average final moisture content of the product (%) 7.24 8.34
Water removed (kg) 179.05 183.15
Average temperature at the inlet of the tunnel (°C) 60.94 58.17
Total thermal energy delivered at the inlet of the tunnel (MJ) 2152.60 1907.43
Contribution of the direct mode (%) 43.23 48.97
Contribution of the indirect mode (%) 35.20 28.56
Contribution of the conventional mode (%) 21.58 22.48
Solar fraction (%) 78.42 77.52
Drying efficiency (%). (Eq. (9)) 18.09 21.10
Effective drying time (h) 14.23 11.70

3.5. Cost and return of investment analysis of solar systems


The cost and return of investment analysis of the direct and indirect air heating systems were
based on the following assumptions [37]:

 A 10-year analysis performed on the useful life of the solar systems


 Calorific power of the LP gas 46.16 MJ/kg of LP gas [54].
 An LP gas cost, in Zacatecas in September 2018, of 19.58 pesos/kg (0.98 USD/kg) [56]
 An annual increase in the price of LP gas considering the LP gas price in Zacatecas since
September 2017 of 17.07 pesos/kg (0.85 USD/kg) [56], resulting in a 14.7 % average annual
increase in the price of LP gas.
 Annual inflation, in 2018, of 4.83 % according to Bank of Mexico.
 208 batches of dehydrated product are considered per year.
 An emission factor of 3.0 kg CO2/kg LP gas [54].
 US Dollar exchange rate (March 2019): 20 pesos/1 USD
 Two financial scenarios: a conventional scenario and an income tax benefit scenario.

According to the average energy delivered at the entrance of the tunnel of the two tests by the
direct air heating system (932.3 MJ) and taking into account the burner efficiency (80 %) and the
calorific power of the LP gas, the equivalent savings per batch are 25.2 kg of LP gas.

The total turnkey cost of the direct air heating system was 407229.7 pesos (20361 USD); the
detail of the breakdown cost is in Table 12. In the breakdown cost, the fan used in the direct and
indirect air heating systems is not considered, because it is default equipment in the conventional
system. In the conventional scenario; the cost and return of investment analysis results in 39-month
payback period for the system, with an internal rate of return (IRR) of 39.5 % and a ten-year net
present value (NPV) of 1333011 pesos (66650 USD), and avoided emissions of 15.7 Ton of CO2 per
year.

Table 12 Breakdown the costs of the direct and indirect air heating system.
Item Cost (pesos) %
Direct air heating system
Solar air heaters with support bases 219112.5 53.81
Air ducts with insulation 59115.4 14.52
Diffusers, Filters, packing and other materials 42492.8 21.24
Installation of the system 86508.9 10.43
Total 407229.7 100.00
Indirect air heating system
Solar water collectors 289473.4 36.55
Thermal storage tank (6000 liters) 96369.3 12.17
Water-air finned tube heat exchanger 68585.0 8.66
Pumps (3 units) 51354.4 6.48
Plate heat exchanger 35074.9 4.43
Differential control, expansion tank and other materials 53060.2 6.70
Installation of the system 198143.8 25.02
Total 792061.0 100.00

The second scenario corresponds to an income tax benefit offered in Mexico for companies, which
have invested in renewable energy projects. The tax benefit provides for an accelerated depreciation
up to 100 % during the first year of investment in equipment for the generation of energy from
renewable sources [57]. This accelerated depreciation represents a real decrease in the initial price of
the solar systems of approximately 30 % of its initial cost. Taking into account this income tax benefit,
the total turnkey cost for the solar collector systems will be 285061 pesos (14253 USD), resulting in
a 28-month payback period, with an IRR of 53.8 % and an NPV in ten years of 1455180 pesos (72759
USD).

According to the average energy delivered at the entrance of the tunnel of the two tests by the
indirect air heating system (651.2 MJ) and taking into account the burner efficiency (80 %) and the
calorific power of the LP gas, the equivalent savings per batch are 17.6 kg of LP gas.

The total turnkey cost of the direct air heating system was 792061 pesos (~39553 USD); the detail
of the breakdown cost is in Table 12. In the conventional scenario, the cost and return of investment
analysis result in an 89-month payback period for the system, with an IRR of 12.7 % and an NPV in ten
years of 423352 pesos (~21168 USD), and avoided emissions of 11.0 Ton of CO2 per year.

Taking into account the income tax benefit, the total turnkey cost of solar collector systems was
554443 pesos (~27722 USD), resulting in a 68-month payback period, with an IRR of 20.4 % and an
NPV in ten years of 660970 pesos (~33049 USD).

4. Conclusions
The Solar-LP gas Hybrid Plant was evaluated drying Nopal under different climatic conditions: May
(clear skies and low relative humidity) and September (cloudy skies and high relative humidity). The
main conclusions are presented below:

 For the indirect air heating system, it was necessary two days of operation to reach the
maximum energy capacity in the thermo tank (6150 l of water around 90 °C). On average, 65
% of the total stored energy was generated during the first day of operation. The overall
energy efficiency of the field of solar collector was approximately 50.2 %. The efficiency of the
coupling between the water heating system and storage tank was on average 40.4 %.
 For the direct air heating system, the effective operation time was approximately 6.4 h. The
overall energy efficiency of the field of solar heaters was approximately 41.2% with a useful
power of around 42.3 kW.
 For the conventional system, the thermal efficiency was around 79.72 % with a useful power
of 58.69 kW.
 The solar fraction obtained for this hybrid mode of operation, direct and indirect air heating
systems, was around 80 %.
 More than 75 % of water removal from Nopal occurs during the first 6 hours of drying.
Therefore, it has low drying efficiencies (<15 %) at the end of drying. In order to increase the
drying efficiency, the load of the product can be increased, or it is possible to replace the dry
product with fresh product.
 Taking into account income tax benefits in Mexico, the economic analysis revealed that the
direct air heating system (solar air heaters) has a 28-month payback period. For the indirect
air heating systems has a 68-month payback period.
 According to the results obtained, different kind of food products can be dried on the thermo-
solar plant, resulting in substantial fuel savings and environmental benefits.

Acknowledgments
This work was supported by FORDECYT Project No. 190603 and SECAMPO Zacatecas. Special
acknowledgements to Víctor Manuel García Saldivar and M. Emilio de los Ríos Ibarra for the technical
support provided to this project.

Nomenclature
𝑨 Current, A
𝑨𝑪 Aperture area of solar collector, m2
𝑪𝒑𝒎 Mean heat capacity at constant pressure, J/kg K
𝒅 Internal diameter of the duct, m
𝑬𝒆𝒍𝒆𝒄 Electrical energy of auxiliary equipment (fan or pump), J
𝑬𝒊𝒏, 𝒄𝒉 Energy delivered to the entrance of the drying chamber, J
𝑬𝒊𝒏𝒄 Solar energy incident, J
𝑬𝒖 Useful energy, J
𝑰 Solar irradiance o solar radiation intensity, W/m2
𝒎 Mass, kg
𝒎𝒘 Mass of water, kg
𝒎𝒑 Mass of product, kg
𝒎 Mass flow rate, kg/s
𝒎𝑻 Total mass flow rate, kg/s
𝑴𝒇 Final moisture content (dry base), %
𝑴𝒊 Initial moisture content (dry base), %
𝑵 Number of time intervals
𝑷𝑭 Power factor
𝑺𝑨𝑯 Solar Air Heater
𝑺𝑾𝑯 Solar Water Heater
𝑻 Temperature, °C
𝒕𝒐𝒑 Fan or pump operating time, h
𝒗 Velocity, m/s
𝑽 Voltage, V
𝑽 Volumetric flow, m3/s
𝑿 The dry base moisture content
𝒀 Absolute humidity, kg H2O/ kg dry air
Greek letters
𝜟𝑻 Temperature difference between the inlet and outlet section of the field of solar collectors,
°C
𝜼𝒐,𝒅 Overall energy efficiency of the drying system, %
𝜼𝒐,𝒆 Overall energy efficiency of the field of solar collectors, %
𝜼𝒕𝒉 Instantaneous thermal efficiency, %
𝜼𝒕𝒉,𝒅 Theoretical thermal efficiency of the dryer, %
𝜼𝒕𝒉,𝒆𝒙𝒄 Thermal efficiency of the water-air heat exchanger, %
𝜼𝒑 Moisture pick-up efficiency, %
𝝀 Latent heat of vaporization of the water, J/kg
𝝆 Density, kg/m3
Subscripts
𝒂𝒎𝒃 Ambient
𝒂𝒔 Point of adiabatic saturation
𝒊𝒏 Inlet
𝒎𝒂𝒙 Maximum
𝒐𝒖𝒕 Outlet

References

[1] FAO, Iniciativa mundial sobre la reducción de la pérdida y el desperdicio de alimentos, 2015.
[Online]. Available: http://www.fao.org/3/a-i4068s.pdf.
[2] O. García-Valladares, I. Pilatowsky, N. Ortiz-Rodríguez, and C. Menchaca-Valdez, Solar thermal
dehydrating plant for agricultural products installed in Zacatecas, México, WEENTECH Proc.
Energy, pp. 1–19, Sep. 2019, doi: 10.32438/WPE.1119.
[3] F. Poggi, A. Firmino, and M. Amado, Planning renewable energy in rural areas: Impacts on
occupation and land use, Energy, vol. 155, pp. 630–640, Jul. 2018, doi:
10.1016/j.energy.2018.05.009.
[4] O. Prakash and A. Kumar, Historical Review and Recent Trends in Solar Drying Systems, Int. J.
Green Energy, vol. 10, Aug. 2013, doi: 10.1080/15435075.2012.727113.
[5] M. Kumar and P. Khatak, Progress in solar dryers for drying various commodities, Renew. Sustain.
Energy Rev., vol. 55, pp. 346–360, Mar. 2016, doi: 10.1016/j.rser.2015.10.158.
[6] A. B. Lingayat, V. P. Chandramohan, V. R. K. Raju, and V. Meda, A review on indirect type solar
dryers for agricultural crops – Dryer setup, its performance, energy storage and important
highlights, Appl. Energy, vol. 258, p. 114005, Jan. 2020, doi: 10.1016/j.apenergy.2019.114005.
[7] V. Belessiotis and E. Delyannis, Solar drying, Sol. Energy, vol. 85, no. 8, pp. 1665–1691, Aug. 2011,
doi: 10.1016/j.solener.2009.10.001.
[8] G. Pirasteh, R. Saidur, S. M. A. Rahman, and N. A. Rahim, A review on development of solar drying
applications, Renew. Sustain. Energy Rev., vol. 31, pp. 133–148, Mar. 2014, doi:
10.1016/j.rser.2013.11.052.
[9] V. Shrivastava, A. Kumar, and P. Baredar, Developments in Indirect Solar Dryer: A Review, Int. J.
Wind Renew. Energy, vol. 3, pp. 67–74, Jan. 2014.
[10] A. Nonclercq, L. Spreutels, C. Boey, L. Lonys, B. Dave, and B. Haut, Construction of a solar drying
unit suitable for conservation of food and enhancement of food security in West Africa, Food
Secur., vol. 1, no. 2, pp. 197–205, Jun. 2009, doi: 10.1007/s12571-009-0019-x.
[11] S. Boughali, H. Benmoussa, B. Bouchekima, D. Mennouche, H. Bouguettaia, and D. Bechki, Crop
drying by indirect active hybrid solar – Electrical dryer in the eastern Algerian Septentrional
Sahara, Sol. Energy, vol. 83, no. 12, pp. 2223–2232, Dec. 2009, doi:
10.1016/j.solener.2009.09.006.
[12] W. Wang, M. Li, R. H. E. Hassanien, Y. Wang, and L. Yang, Thermal performance of indirect forced
convection solar dryer and kinetics analysis of mango, Appl. Therm. Eng., vol. 134, pp. 310–321,
Apr. 2018, doi: 10.1016/j.applthermaleng.2018.01.115.
[13] J. Prasad, V. K. Vijay, G. N. Tiwari, and V. P. S. Sorayan, Study on performance evaluation of hybrid
drier for turmeric (Curcuma longa L.) drying at village scale, J. Food Eng., vol. 75, no. 4, pp. 497–
502, Aug. 2006, doi: 10.1016/j.jfoodeng.2005.04.061.
[14] S. Dhanuskodi, V. Wilson, and S. Kumarasamy, Design and thermal performance of the solar
biomass hybrid dryer for cashew drying, Facta Univ. Ser. Mech. Eng., vol. 12, no. 3, pp. 277–288,
Dec. 2014.
[15] T. A. Yassen and H. H. Al-Kayiem, Experimental investigation and evaluation of hybrid
solar/thermal dryer combined with supplementary recovery dryer, Sol. Energy, vol. 134, pp.
284–293, Sep. 2016, doi: 10.1016/j.solener.2016.05.011.
[16] Hamdani, T. A. Rizal, and Z. Muhammad, Fabrication and testing of hybrid solar-biomass dryer
for drying fish, Case Stud. Therm. Eng., vol. 12, pp. 489–496, Sep. 2018, doi:
10.1016/j.csite.2018.06.008.
[17] R. Smitabhindu, S. Janjai, and V. Chankong, Optimization of a solar-assisted drying system for
drying bananas, Renew. Energy, vol. 33, no. 7, pp. 1523–1531, Jul. 2008, doi:
10.1016/j.renene.2007.09.021.
[18] E. C. López-Vidaña, L. L. Méndez-Lagunas, and J. Rodríguez-Ramírez, Efficiency of a hybrid solar–
gas dryer, Sol. Energy, vol. 93, pp. 23–31, Jul. 2013, doi: 10.1016/j.solener.2013.01.027.
[19] D. Gudiño-Ayala and Á. Calderón-Topete, Pineapple Drying Using a New Solar Hybrid Dryer,
Energy Procedia, vol. 57, pp. 1642–1650, Jan. 2014, doi: 10.1016/j.egypro.2014.10.155.
[20] H. Oueslati, S. Ben Mabrouk, and A. Marni, Design and installation of a solar-gas tunnel dryer:
Comparative experimental study of two scenarios of drying, in 2014 5th International Renewable
Energy Congress (IREC), Mar. 2014, pp. 1–6, doi: 10.1109/IREC.2014.6826970.
[21] R. Anum, A. Ghafoor, and A. Munir, Study of the Drying Behavior and Performance Evaluation of
Gas Fired Hybrid Solar Dryer, J. Food Process Eng., vol. 40, no. 2, p. e12351, 2017, doi:
10.1111/jfpe.12351.
[22] A. Zoukit, H. El Ferouali, I. Salhi, S. Doubabi, and N. Abdenouri, Simulation, design and
experimental performance evaluation of an innovative hybrid solar-gas dryer, Energy, vol. 189,
p. 116279, Dec. 2019, doi: 10.1016/j.energy.2019.116279.
[23] W. Amjad, G. Ali Gilani, A. Munir, F. Asghar, A. Ali, and M. Waseem, Energetic and exergetic
thermal analysis of an inline-airflow solar hybrid dryer, Appl. Therm. Eng., vol. 166, p. 114632,
Feb. 2020, doi: 10.1016/j.applthermaleng.2019.114632.
[24] A. Madhlopa and G. Ngwalo, Solar dryer with thermal storage and biomass-backup heater, Sol.
Energy, vol. 81, no. 4, pp. 449–462, Apr. 2007, doi: 10.1016/j.solener.2006.08.008.
[25] M. A. Leon and S. Kumar, Design and Performance Evaluation of a Solar-Assisted Biomass Drying
System with Thermal Storage, Dry. Technol., vol. 26, no. 7, pp. 936–947, Jul. 2008, doi:
10.1080/07373930802142812.
[26] M. A. Hossain, B. M. A. Amer, and K. Gottschalk, Hybrid Solar Dryer for Quality Dried Tomato,
Dry. Technol., vol. 26, no. 12, pp. 1591–1601, Nov. 2008, doi: 10.1080/07373930802467466.
[27] B. M. A. Amer, M. A. Hossain, and K. Gottschalk, Design and performance evaluation of a new
hybrid solar dryer for banana, Energy Convers. Manag., vol. 51, no. 4, pp. 813–820, Apr. 2010,
doi: 10.1016/j.enconman.2009.11.016.
[28] A. Reyes, A. Mahn, and F. Vásquez, Mushrooms dehydration in a hybrid-solar dryer, using a
phase change material, Energy Convers. Manag., vol. 83, pp. 241–248, Jul. 2014, doi:
10.1016/j.enconman.2014.03.077.
[29] R. Daghigh and A. Shafieian, An experimental study of a heat pipe evacuated tube solar dryer
with heat recovery system, Renew. Energy, vol. 96, pp. 872–880, Oct. 2016, doi:
10.1016/j.renene.2016.05.025.
[30] S. Murali, P. R. Amulya, P. V. Alfiya, D. S. A. Delfiya, and M. P. Samuel, Design and performance
evaluation of solar - LPG hybrid dryer for drying of shrimps, Renew. Energy, vol. 147, pp. 2417–
2428, Mar. 2020, doi: 10.1016/j.renene.2019.10.002.
[31] M. Condorí, G. Duran, R. Echazú, and F. Altobelli, Semi-industrial drying of vegetables using an
array of large solar air collectors, Energy Sustain. Dev., vol. 37, pp. 1–9, Apr. 2017, doi:
10.1016/j.esd.2016.11.004.
[32] A. Arata, V. K. Sharma, and G. Spagna, Performance evaluation of solar assisted dryers for low
temperature drying application—II. Experimental results, Energy Convers. Manag., vol. 34, no.
5, pp. 417–426, May 1993, doi: 10.1016/0196-8904(93)90091-N.
[33] C. Palaniappan and S. V. Subramanian, Economics of solar air pre-heating in south Indian tea
factories: a case study, Sol. Energy, vol. 63, no. 1, pp. 31–37, Jul. 1998, doi: 10.1016/S0038-
092X(98)00028-0.
[34] S. Janjai, A greenhouse type solar dryer for small-scale dried food industries: Development and
dissemination, Int J Energy Env., vol. 3, pp. 383–398, Jan. 2012.
[35] A. Fudholi, Kamaruzzaman Sopian, Mohamed Gabbasa, B. Bakhtyar, M. Yahya, Mohd Hafidz
Ruslan, and Sohif Mat, Techno-economic of solar drying systems with water based solar
collectors in Malaysia: A review, Renew. Sustain. Energy Rev., vol. 51, pp. 809–820, Nov. 2015,
doi: 10.1016/j.rser.2015.06.059.
[36] S. Misha, S. Mat, M. H. Ruslan, E. Salleh, and K. Sopian, Performance of a solar-assisted solid
desiccant dryer for oil palm fronds drying, Sol. Energy, vol. 132, pp. 415–429, Jul. 2016, doi:
10.1016/j.solener.2016.03.041.
[37] O. García-Valladares, N. M. Ortiz, I. Pilatowsky, and A. C. Menchaca, Solar thermal drying plant
for agricultural products. Part 1: Direct air heating system, Renew. Energy, Oct. 2019, doi:
10.1016/j.renene.2019.10.069.
[38] L. Medina-Torres, J. A. Gallegos-Infante, R. F. Gonzalez-Laredo, and N. E. Rocha-Guzman, Drying
kinetics of nopal (Opuntia ficus-indica) using three different methods and their effect on their
mechanical properties, LWT - Food Sci. Technol., vol. 41, no. 7, pp. 1183–1188, Sep. 2008, doi:
10.1016/j.lwt.2007.07.016.
[39] P. Inglese, C. Mondragon, A. Nefzaoui, and C. Sáenz, Ecología del cultivo, manejo y usos del nopal.
Food and Agriculture Org., 2018.
[40] L. Andreu-Coll, M. Cano-Lamadrid, E. Sendra, Á. Carbonell-Barrachina, P. Legua, and F.
Hernández, Fatty acid profile of fruits (pulp and peel) and cladodes (young and old) of prickly
pear [Opuntia ficus-indica (L.) Mill.] from six Spanish cultivars, J. Food Compos. Anal., vol. 84, p.
103294, Dec. 2019, doi: 10.1016/j.jfca.2019.103294.
[41] R. López, A. de Ita, and M. Vaca, Drying of prickly pear cactus cladodes (Opuntia ficus indica) in
a forced convection tunnel, Energy Convers. Manag., vol. 50, no. 9, pp. 2119–2126, Sep. 2009,
doi: 10.1016/j.enconman.2009.04.014.
[42] R. Callejas, A. Ita, Mabel Mier, A. Ramos, J. Gomez, H. Terres, and A. Valdivia, Kinetics modeling
of the drying of Opuntia ficus indica with solar energy, Rev. Mex. Fis., vol. 59, pp. 163–167, Feb.
2013.
[43] O. V. Ekechukwu and B. Norton, Review of solar-energy drying systems III: low temperature air-
heating solar collectors for crop drying applications, Energy Convers. Manag., vol. 40, no. 6, pp.
657–667, Apr. 1999, doi: 10.1016/S0196-8904(98)00094-6.
[44] G. D. Saravacos and Z. B. Maroulis, Food Process Engineering Operations. New work: CRC press,
2011.
[45] M. Augustus Leon, S. Kumar, and S. C. Bhattacharya, A comprehensive procedure for
performance evaluation of solar food dryers, Renew. Sustain. Energy Rev., vol. 6, no. 4, pp. 367–
393, Aug. 2002, doi: 10.1016/S1364-0321(02)00005-9.
[46] R. J. Moffat, Describing the uncertainties in experimental results, Exp. Therm. Fluid Sci., vol. 1,
no. 1, pp. 3–17, Jan. 1988, doi: 10.1016/0894-1777(88)90043-X.
[47] INIFAP, Estadísticas climatológicas básicas del estado de Zacatecas (Périodo 1961-2003),
Instituto Nacional de Investigaciones Forestales, Agrícolas y Pecuarias, Zacatecas, México, 2004.
[48] UAZ, Solarimetric Station UAZ Campus Siglo XXI. Universidad Autónoma de Zacatecas, 2015.
[49] S. Misha, S. Mat, M. H. Ruslan, E. Salleh, and K. Sopian, Performance of a solar assisted solid
desiccant dryer for kenaf core fiber drying under low solar radiation, Sol. Energy, vol. 112, pp.
194–204, Feb. 2015, doi: 10.1016/j.solener.2014.11.029.
[50] A. Sreekumar, Techno-economic analysis of a roof-integrated solar air heating system for drying
fruit and vegetables, Energy Convers. Manag., vol. 51, no. 11, pp. 2230–2238, Nov. 2010, doi:
10.1016/j.enconman.2010.03.017.
[51] M. Condorı́, R. Echazú, and L. Saravia, Solar drying of sweet pepper and garlic using the tunnel
greenhouse drier, Renew. Energy, vol. 22, no. 4, pp. 447–460, Apr. 2001, doi: 10.1016/S0960-
1481(00)00098-7.
[52] S. Janjai and P. Tung, Performance of a solar dryer using hot air from roof-integrated solar
collectors for drying herbs and spices, Renew. Energy, vol. 30, no. 14, pp. 2085–2095, Nov. 2005,
doi: 10.1016/j.renene.2005.02.006.
[53] A. Čiplienė, E. Zvicevičius, and A. Raila, Aerial solar collector usage in drying technologies of
medicinal and spice plants, 2018, doi: https://doi.org/10.22616/ERDev2018.17.N506.
[54] INECC, SEMARNAT, Factores de emisión para los diferentes tipos de combustibles fósiles y
alternativos que se consumen en México, Instituto Nacional de Ecología y Cambio Climático,
Ciudad de México, technical report, Dec. 2014. [Online]. Available:
https://www.gob.mx/cms/uploads/attachment/file/110131/CGCCDBC_2014_FE_tipos_combu
stibles_fosiles.pdf.
[55] O. V. Ekechukwu, Review of solar-energy drying systems I: an overview of drying principles and
theory, Energy Convers. Manag., vol. 40, no. 6, pp. 593–613, Apr. 1999, doi: 10.1016/S0196-
8904(98)00092-2.
[56] C. R. de CRE, Precios al público de gas LP reportados por los distribuidores, gob.mx.
http://www.gob.mx/cre/documentos/precios-al-publico-de-gas-lp-reportados-por-los-
distribuidores (accessed Aug. 09, 2019).
[57] COF, Article 34 of the Mexican Income Tax Law of 2014, México, 2018. [Online]. Available:
https://www.cof.org/sites/default/files/leydelimp.pdf.
Highlights:

 Experimental evaluation of an industrial solar-LP gas plant for dehydration of food


 The thermal efficiency of the solar collector field was 42% (SAH) and 60% (SWH)
 The drying efficiency was between 18 and 22% in the tunnel type drying chamber
 The solar fraction obtained was ~80% for a continuous Nopal drying operation
 Direct air heating system has a 28-months payback period
April 17, 2020

Conflicts of Interest Statement

Manuscript title:
Solar-LP Gas Hybrid Plant for Dehydration of Food

The authors whose names are listed immediately below certify that they have NO
affiliations with or involvement in any organization or entity with any financial interest (such
as honoraria; educational grants; participation in speakers’ bureaus; membership, employment,
consultancies, stock ownership, or other equity interest; and expert testimony or patent-
licensing arrangements), or non-financial interest (such as personal or professional
relationships, affiliations, knowledge or beliefs) in the subject matter or materials discussed in
this manuscript

Authors names:

N. M. Ortiz-Rodríguez, O. García-Valladares, I. Pilatowsky-Figueroa, C. Menchaca-


Valdez

You might also like