You are on page 1of 24

International Journal of Sustainable Energy

ISSN: 1478-6451 (Print) 1478-646X (Online) Journal homepage: http://www.tandfonline.com/loi/gsol20

A review of research and development work on


solar dryers with heat storage

Ashish Agrawal & R.M. Sarviya

To cite this article: Ashish Agrawal & R.M. Sarviya (2016) A review of research and development
work on solar dryers with heat storage, International Journal of Sustainable Energy, 35:6, 583-605,
DOI: 10.1080/14786451.2014.930464

To link to this article: http://dx.doi.org/10.1080/14786451.2014.930464

Published online: 27 Jun 2014.

Submit your article to this journal

Article views: 373

View related articles

View Crossmark data

Citing articles: 4 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=gsol20

Download by: [Univ of Alhussein Bin Talal] Date: 12 November 2017, At: 04:03
International Journal of Sustainable Energy, 2016
Vol. 35, No. 6, 583–605, http://dx.doi.org/10.1080/14786451.2014.930464

A review of research and development work on solar dryers with


heat storage
Ashish Agrawal∗ and R.M. Sarviya
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Department of Mechanical Engineering, Maulana Azad National Institute of Technology, Bhopal,


Madhya Pradesh, India

(Received 28 January 2013; final version received 5 April 2014)

Drying of agricultural food products is one of the most attractive and cost-effective applications of solar
energy. The solar dryer is less reliable due to the intermittent nature of solar energy. This shortcoming
can be overcome to some extent by storing solar energy. Information on sensible and latent heat storage
materials and systems is spread widely in the literature. In this paper, we try to gather information about
the previous and current research works in the field of thermal energy storage technology for solar air
heater and dryer. The relative studies are classified on the basis of the type of storage material used in solar
dryers, i.e. phase change material (PCM), rock, water, etc. Several designs of solar dryers with different
heat storage materials were proposed by researchers. Recent studies focused on PCMs such as Paraffin
and salt hydrate, due to their high heat storage capacity per unit volume.

Keywords: solar dryer; heat storage system; phase change material; solar energy

1. Introduction

In most of the developing countries a large amount of food and grains are spoiled due to the lack
of proper storage. Drying of food is an effective means of extending shelf life, improving quality
and minimising losses during storage since most of the water is taken out of the product during
this process. Hot air in the temperature range of 45–60 ◦ C is required for safe drying. Drying
under a controlled weather condition of temperature and humidity brings the moisture level of
the food products to a safe limit and ensures good quality of products (Cohen and Yang 1995).
Various drying techniques are employed to dry different food products such as open sun drying,
conventional industrial drying based on fossil fuel- and solar-based drying system. Open sun
drying is the cheapest method of drying food, but it suffers from contamination of the product
due to dirt and insects and spoilage due to sudden and unpredicted rain. Industrial drying offers
quality drying, whereas its high cost limits its use. Large quantities of fossil fuels are required for
providing hot air in industrial drying. Solar dryer is a low-cost device, operated by solar energy.
A solar dryer without heat storage provides air with large variations in temperature to the dryer,
and drying of food is not possible during partial clouds and late evening hours. In such cases
a solar dryer with thermal storage is required. Reliability of a solar dryer can be increased by

∗ Corresponding author. Email: er_ashishagarwal@yahoo.com

© 2014 Taylor & Francis


584 A. Agrawal and R.M. Sarviya

storing excess energy during peak time and using it in off-sun hours or when the intensity of solar
energy is not inadequate. Solar energy can be stored in the form of sensible heat, latent heat of
the material or a combination of both.
Energy consumption for drying in developing countries is a major portion of the total energy
consumption. Most of the energy requirement for drying is fulfilled by fossil fuels such as coal
and natural gas. India spends annually around 157 million tonnes of coal, 89 million tonnes of
petroleum products and 233 million tonnes of other traditional conventional energy to meet its
industrial, agro and domestic requirements (Ministry of New and Renewable Energy (MNRE)
Annual Report 2010). There is a significant potential for solar dryer in the agricultural sector
to dry agricultural products such as food grains, vegetables, fruits and medicinal plants, thereby
eliminating dependency on open sun drying and industrial drying, while saving huge quantities
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

of fossil fuels.
India is located in the equatorial sun belt of the earth, thereby receiving abundant radiant energy
from the sun. Most parts of India receive 4–7 kWh of solar radiation per square metre per day
with 250–300 sunny days in a year. The annual global radiation varies from 1600 to 2200 kWh/hr
(Arjunan, Aybar, and Nedunchezhian 2009). Hence, solar drying has a high potential of diffusion
in the domestic and industrial sector. It has advantages like low running cost and superior quality of
food. The main hurdles in its popularity are non-reliability due the intermittent nature of sunshine,
higher initial cost and space availability (Reddy and Painuly 2004). Integration of energy storage
is, therefore, essential for a solar dryer. The energy storage system can minimise the gap between
energy supply and energy demand.

2. Classification of solar dryer

Solar dryers used for drying agricultural products are broadly classified as shown in Figure 1.
This classification is based on the design of system components and mode of utilisation of solar
energy.

3. Thermal energy storage

3.1. Classification of energy storage materials

There are various techniques for storing solar energy in the form of sensible and latent heat.
Thermal energy storage increases the reliability and hours of operation of a solar dryer. An
overview of the major techniques of thermal energy storage is shown in Figure 2. The heat
storage capacity of a material depends on its latent and specific heat values. These values should
be as high as possible to minimise the physical size of the heat storage (Abhat 1983).

3.2. Sensible heat storage

In sensible heat storage (SHS), thermal energy is stored by raising the temperature of a liquid
or solid. These systems utilise the heat capacity of the material. The temperature of the material
increases during charging and decreases during discharging. The amount of heat stored depends
on the amount, specific heat and temperature change of the material. SHS has a low heat storage
capacity per unit volume of the stored material, and heat transfer during charging and discharging
also does not take place at a constant temperature.
International Journal of Sustainable Energy 585
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Figure 1. Classification of crop drying using solar energy.

Figure 2. Classifications of energy storage materials.


586 A. Agrawal and R.M. Sarviya

Water, rock, brick, concrete and engine oil are some of the material that are generally used in
SHS. Water appears to be the most promising SHS material due to its high specific heat and as
it is inexpensive for a mid-temperature range. However, liquid metals, oils and molten salts are
used for storage of thermal energy above 100 ◦ C. Rock bed-type storage materials are used for
air heating application (Hasnain 1998).

3.3. Latent heat storage

Latent heat storage (LHS) is based on the absorption or release of latent heat at a constant
temperature when a storage material undergoes a change of phase from solid to liquid, liquid to
gas or vice versa. Materials used in LHS are the phase change materials (PCMs). LHS may be
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

grouped as solid–liquid, solid–gas, solid–solid and liquid–gas on the basis of the change of phase
during the process. During liquid–gas and solid–gas transformations, large change in volume
takes place so it is not suitable for energy storage. Large change in volume makes the system
more complex and unreliable (Sharma et al. 2009).
Volume change during solid–liquid phase transition is less and latent heat is absorbed at a
constant temperature corresponding to the phase transition temperature of the PCMs; therefore,
it is considered to be particularly attractive for heat storage. In solid–solid transitions, heat is
stored or released as the crystal structure of the material is transformed from one crystalline
state to another. Solid–solid transitions generally require a low latent heat and volume change
during the process is also small. LHS has a high heat storage capacity due to the high value of
latent heat compared with sensible heat and due to minimum heat losses from the system during
charging and discharging of heat. LHS occupied two times less volume than water for the same
heat storage capacity. Farid et al. (2004) reviewed the LHS systems and their materials that have
been used for different research applications during the last 40 years. These materials are paraffins,
non-paraffins, hydrate salts, eutectics, etc.

4. Phase change materials

PCMs are LHS materials. The chemical bonds within the PCM break up as the material absorbs
latent heat and the material changes its state from solid to liquid. They store 5–14 times more heat
per unit volume than sensible storage materials such as water, masonry or rock (Hale, Hoover,
and O’Neill 1971). The desirable thermo-physical properties of PCMs are (1) low supercooling,
(2) high density, (3) small volume change during phase change, (4) high thermal conductivity and
(5) high latent heat of fusion (Garg, Mullick, and Bhargava 1985; Lalit, Santosh, and Naik 2010).
PCMs should be chemically stable, non-toxic and non-flammable (Buddhi and Sawhney 1994).
A large number of PCMs are available in the desired temperature range. A comprehensive
review of the most possible materials that may be used for LHS were proposed by Abhat (1983),
Lane (1983), Zhou, Zhao, and Tian (2012), Garg, Mullick, and Bhargava (1985), Zhang et al.
(2007), Tyagi and Buddhi (2007), Saffa, Blaise, and Wenbo (2013), Sethi and Sharma (2008) and
Verma, Varun, and Singal (2008). A large number of organic, inorganic and artistic materials are
identified as PCMs. These materials are available in temperature range from 0 ◦ C to 150 ◦ C. Most
of the PCMs do not fulfil the criteria for an ideal storage medium. Most of them have low thermal
conductivity and the problem of supercooling. Low thermal conductivity leads to less effective
heat transfer during charging and discharging of heat storage media. Metallic fins can be used
to increase the thermal conductivity of PCMs, supercooling may be suppressed by introducing a
nucleating agent or a ‘cold finger’ in the storage material and incongruent melting can be inhibited
by the use of suitable thickness.
International Journal of Sustainable Energy 587

4.1. Organic phase change materials

Organic PCMs are further divided into Paraffins and Non-paraffins. A significant number of
authors have based their work on organic materials such as alkanes, waxes or paraffins. Organic
PCMs are non-toxic, have high latent heat and negligible supercooling, are non-corrosive and
chemically stable and have a wide range of melting point (Lane 1983). Some disadvantages of
organic PCMs are low thermal conductivity in comparison to inorganic compounds, low density
and high cost (Ghoneim and Klein 1989). The problem of low thermal conductivity of the organic
PCMs can be solved to some extent by the mixing of a filler with high thermal conductivity or
the use of aluminium honeycombs or matrixes (Hoogendoorn and Bart 1992).
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

4.1.1. Paraffins

Paraffin wax consists of a mixture of mostly straight chain n-alkanes CH3 –(CH2 )–CH3 . Crys-
tallisation of the (CH3 )-chain releases a large amount of latent heat. Latent heat of fusion and
melting temperature increase with the number of carbon atoms. Paraffins are available in a wide
variety of melting temperature. They are less expensive, reliable, show little volume change on
melting and have a low vapour pressure. Some undesirable properties of paraffins are low thermal
conductivity, moderate flammability and non-compatibility with plastic containers.

4.1.2. Non-paraffins

Unlike paraffin’s, these materials have highly varied properties. Each material has its own prop-
erties. A number of fatty acids, glycols, alcohols and esters were identified as non-paraffin heat
storage material. Some of the characteristic properties of these materials are low thermal con-
ductivity, inflammability, high heat of fusion and low flash point (Lane and Glew 1975; Herrick
and Golibersuch 1977). Fatty acids have a high heat storage capacity compared with paraffins.
It freezes without any supercooling. The general formula describing all fatty acids is given by
CH3 (CH2 )2n · COOH. The cost of fatty acids is 2–2.5 times greater than that of technical grade
paraffins.

4.2. Inorganic PCMs

Inorganic PCMs are classified into salt hydrate and metallics. An alloy of inorganic salts and water
is called salt hydrate. The general formula of salt hydrate is AB.nH2 O. It is an alloy of inorganic
salts and water. Salt hydrate melts into anhydrous salt and water or into lower hydrate and water.
Supercooling is one of the problems associated with salt hydrate. The rate of nucleation at fusion
temperature is very low. The solution has to be supercooled to increase the rate of nucleation. The
solution to this problem is to add a nucleating agent, which leads to nucleation when crystal for-
mation is initiated. Some characteristics of salt hydrate are low thermal conductivity, less volume
change on melting, slight toxicity, compatibility with plastics and high latent heat of fusion (Lane
1978). Belen et al. (2003) reviewed different substances, eutectics and mixtures (inorganic, organic
and fatty acids) that have been studied by different researchers for their potential use as PCMs.

4.3. Commercial PCMs

Many companies such as EPS Ltd, Climator, Mitsubishi Chemical Corporation, Rubitherm GmbH
and Witco are engaged in the production and development of PCMs. Table 1 shows a list of the
commercial PCMs available in the market and their physical and thermal properties as given by
companies.
588 A. Agrawal and R.M. Sarviya

Table 1. A list of Commercial PCMs available in the market (0–117 ◦ C) (Available from www.pcmproducts.net; Avail-
able from www.cristopia.com; Available from www.teappcm.com; Kakiuchi 2002; Available from www.climator.com;
Available from www.rubitherm.com).

Melting Thermal
temperature Latent Density conductivity
PCM name (◦ C) heat (kJ/kg) (kg/m3 ) (W/m-k) Manufacturer

RT 0 0 225a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)


A4 4 200 766 0.210 EPS Ltd (Available from www.pcmproducts.net)
ClimSel C7 7 130 1400 0.5–0.7 Climator (Available from www.climator.com)
RT8 8 180a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
ClimSel C10 10 126 1400 0.5–0.7 Climator (Available from www.climator.com)
RT10 10 150a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
RT 12 12 150a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

RT15 15 140a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)


Latest 18T 18 175 1480–1500 1 TEAP (Available from www.teappcm.com)
RT21 21 160a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
ClimSel C24 24 152 1380 0.5–0.7 Climator (Available from www.climator.com)
RT24 24 150a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
Latest 25T 25 175 1480–1500 1 TEAP (Available from www.rubitherm.com)
A27 27 207 n.a 1.47 Cristopia (Available from www.cristopia.com)
ClimSel C28 28 162 1420 0.5–0.7 Climator (Available from www.climator.com)
Latest 29T 29 175 1480–1500 1 TEAP (Available from www.teappcm.com)
SP 29 29 160a 1500–1550 0.6 Rubitherm (Available from www.rubitherm.com)
RT31 31 170a 760–880 0.2 Rubitherm (Available from www.rubitherm.com)
Latest 32S 32 200 1450 0.6 TEAP (Available from www.teappcm.com)
ClimSel C32 32 162 1420 0.5–0.7 Climator (Available from www.climator.com)
Latest 34S 34 220 1450 0.6 TEAP (Available from www.teappcm.com)
A36 36 217 790 0.18 EPS Ltd (Available from www.pcmproducts.net)
A39 39 105 900 0.22 EPS Ltd (Available from www.pcmproducts.net)
A40 40 230 810 0.18 EPS Ltd (Available from www.pcmproducts.net)
RT42 42 174a 760–880 0.2 Rubitherm (Available from www.rubitherm.com)
RT47 47 170a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
Latest 48S 48 230 1450 0.6 TEAP (Available from www.teappcm.com)
STL47 47 221 1340 n.a Mitsubishi Chemical (Kakiuchi 2002)
S50 50 100 1601 0.43 EPS Ltd (Available from www.pcmproducts.net)
STL52 52 201 1300 n.a Mitsubishi Chemical (Kakiuchi 2002)
RT55 55 172a 770–880 0.2 Rubitherm (Available from www.rubitherm.com)
STL 55 55 242 1290 n.a Mitsubishi Chemical (Kakiuchi 2002)
ClimSel C58 58 289 1460 0.5–0.7 Climator (Available from www.climator.com)
A60 60 145 910 0.22 EPS Ltd (Available from www.pcmproducts.net)
RT62 62 146a 780–880 0.2 Rubitherm (Available from www.rubitherm.com)
SP 70 71 150a 1300–1500 0.6 Rubitherm (Available from www.rubitherm.com)
ClimSel C70 71 283 1400 0.5–0.7 Climator (Available from www.climator.com)
S72 72 127 1666 0.58 EPS Ltd (Available from www.pcmproducts.net)
RT80 HC 79 240a 800–900 n.a Rubitherm (Available from www.rubitherm.com)
PCM 80 80 231 n.a n.a Mitsubishi Chemical (Kakiuchi 2002)
A82 82 155 850 0.22 EPS Ltd (Available from www.pcmproducts.net)
S83 83 141 1600 0.62 EPS Ltd (Available from www.pcmproducts.net)
PCM 86 86 246 n.a n.a Mitsubishi Chemical (Kakiuchi 2002)
S89 89 151 1550 0.670 EPS Ltd (Available from www.pcmproducts.net)
RT90 HC 90 200a 850–950 n.a Rubitherm (Available from www.rubitherm.com)
A95 95 205 900 0.22 EPS Ltd (Available from www.pcmproducts.net)
S117 117 160 1450 0.7 EPS Ltd (Available from www.pcmproducts.net)

Note: n.a = not available.


a
sensible + latent heat.

5. Drying mechanism

Drying of food and agricultural product comprises heat and mass transfer processes. Heat is
transferred to evaporate liquid in food, and mass is transferred as a liquid or vapour within the solid
International Journal of Sustainable Energy 589

and as vapour from the surface. Heat is transferred to food by conduction, convection, radiation
or a combination of these. The movement of liquid or moisture within the solid depends on the
structure and characteristics of the solid. The different mechanisms of moisture movement during
drying of food are diffusion, capillary forces, shrinkage, and pressure gradients (Myer 2007).
During the early stage drying takes place at a constant rate. Drying proceeds by diffusion of
vapour from the saturated surface of the material across a stagnant air film into the environment.
The magnitude of the constant rate depends upon three factors – (1) the heat or mass transfer
coefficient, (2) the area exposed to the drying medium and (3) the difference in temperature or
humidity between the gas stream and the wet surface of the solid. The internal mechanism of liquid
flow does not affect the constant rate (Keey 1978). Constant rate drying period ends when the
moisture content in the solid is equal to critical moisture content. Drying rate decreases with time
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

and the falling rate period begins. The drying rate is now governed by the rate of internal moisture
movement. The drying rate depends on factors affecting the diffusion of moisture away from the
evaporating surface. The energy requirement for drying depends on the initial and final moisture
content of food products. A list of maximum allowable temperature and energy requirement during
drying of food products is given in Table 2.

6. Review of solar dryers with thermal heat storage systems

6.1. Solar dryers with rock bed as heat storage material

Rock bed stores heat in the form of sensible heat. It is used widely as an air-based thermal energy
storage material due to its low cost and easy availability. The performance of a solar dryer with
rock bed storage depends on its design and operational parameters such as void fraction within
rock bed, size of rock, thickness of bed, flow rate of air, physical and thermal properties of rock
such as thermal conductivity and specific heat. Rock bed can be placed on the surface that is
exposed to the incident solar radiations to receive the direct heat or it can be placed underground.
Temperature gain of around 24 ◦ C can be obtained in a dryer with rock bed heat storage (Fagunwa,
Koya, and Faborode 2009).
Butler and Troeger (1980) have experimentally evaluated a solar collector-cum-rock bed stor-
age system for peanut drying. The drying time ranged from 22 to 25 h to reduce the moisture
content from 20% to the safe storage moisture level with an air flow rate of 4.9 m3 /s. Theoretical
investigation of a dryer-cum-air-heater with heat storage was made by Chauhan, Choudhury, and
Garg (1996). They studied the drying characteristics of coriander in a deep bed dryer coupled
with a solar air heater and a rock bed storage unit. A rock bed storage device stores heat during
sunshine hours and discharges the heat to ambient air during off-sunshine hours. The theoretical
investigation was made by writing the energy and mass balance equations for different compo-
nents of the dryer-cum-air-heater-cum-storage device and by adopting a finite difference approach
for simulation. The drying performance with and without rock bed storage was compared. The
results revealed that for reducing the moisture content from 28.2% (db) to 11.4% (db), the solar
air heater takes 27 cumulative sunshine hours, i.e. about 3 sunshine days, whereas the solar air
heater and the rock bed storage combined take 31 cumulative hours, i.e. about 2 days and 2 nights
at an air flow velocity of 250 kg/hm2 , and recommended that the heat stored in the rock bed can be
used effectively for heating the inlet (ambient) air for off-sunshine drying of agricultural products.
Choudhury, Chauhan, and Garg (1995) optimised the design and operational parameters of a
rock bed thermal energy storage unit coupled to two solar air heaters. The parameters considered
during the study are charging time, rock bed size, void friction, cross-sectional area and air mass
velocity per bed. The results revealed that air mass velocity per bed cross-sectional area and
cross-sectional area of the bed should be selected depending on the requirement of temperature
590 A. Agrawal and R.M. Sarviya

Table 2. The moisture content of food and agricultural products (Mujumdar 1987; Purohit, Kumar, and Kandpal 2006;
IEA 1998; Sharma, Colnagelo, and Spagna 1993; Brooker, Bakker-Arkema, and Hall 1992).

Product Initial moisture Final moisture Max allowable Energy


Product level (% wb) level (% wb) temperature (◦ C) requirement (MJ/kg)

Apple 80–85 20–24 70 1.502


Apricot 85 15–25 65 1.666
Bananas 70–80 7–15 70 1.679
Barley 15–20 13–14 43 0.055–0.168
Brinjal 95 6 60 –
Cassava 62–75 14–17 150 1.105
Carrots 70 5 75 –
Cabbage 80 4 65 –
Cauliflower 80 6 65 –
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Chillies 75–80 5–14 90 1.610


Cocoa Beans 60–80 25–30 – –
Copra 30 5 – –
Coffee 45–65 9–12 – 0.855–0.865
Corn 24 14 50 –
Garlic flakes 80 4 – –
Garlic 80 4 55 –
Ginger 80 10 – –
Grapes 75–80 15–20 70 1.444
Green peas 80 5 60 –
Green beans 70 5 75 –
Guavas 80 7 65 –
Maize 20–35 8–15 60 0.254–0.565
Mangoes 80–85 12–18 70 1.564
Okra 80 20 65 –
Onions 85 6 55 –
Onion flakes 80 10 55 –
Onion rings 80 10 55 –
Oil seed 20–25 7–9 40–60 –
Pineapple 80 10 65 –
Pepper 71 13 – –
Paddy raw 22–24 11 50 –
Potatoes 70–75 8–13 75–85 1.453
Pulses 20–22 9–10 40–60 –
Rice 20–30 12–18 50 0.218–0.351
Turmeric 80 10 – –
Tea 60–80 25–3 140 1.203
Tobacco leaves 70–85 11–25 70 1.592–1.997
Tomatoes 75 35 60 0.963
Wheat 15–20 13–14 49 0.055–0.193
Wood 40–200 8–10 60 –

for any particular application. Garg et al. (1985) experimentally investigated solar air heater with
rock bed heat storage. The system consists of a flat plate air heater and an integrated rock storage
and collection system. Appropriate rise in temperature of the air above atmospheric temperature
is observed for a small value of mass flow rate. Irrespective of the mass flow rate of air, storage
of heat in rock bed is effective up to 3.30 pm. Heat loss characteristics of the rock bed using night
insulation covers are also reported. Tiwari, Singh, and Bhatia (1994) worked on experimental
simulation of a grain drying system with and without heat storage. The experimental simulation
was conducted to evaluate the drying time for wheat crop. Rocks of average size 5–8 cm diameter
and density 1750 kg/ml were used as storage material. On the basis of experimental simulation,
the following conclusions have been drawn: (1) the fluctuation in temperature is significantly
reduced due to the storage effect, (2) attaining the steady state condition will take a larger time
due to high thermal capacity of the rock bed thermal storage and (3) by using thermal storage,
the maximum temperature of the drying material is reduced within a safe range.
International Journal of Sustainable Energy 591

Fagunwa, Koya, and Faborode (2009) designed and developed a forced convection solar dryer
with black painted gravel as heat storage material for drying Cocoa Beans (Figure 3). The
experimental model dehydrated cocoa beans from 53.4% to 3.6% moisture content (wb) in a
72 h inter-mittent drying process against ambient temperature and relative humidity in the range
25–30 ◦ C and 58–98%, respectively. Temperature variation in the dryer was measured for a 24-h
period, starting from 6 am on the first day of the three-day drying period. The temperature in the
dryer ranged from a minimum of 31 ◦ C at 6 am to a maximum of 54 ◦ C at 3 pm. The corresponding
range of ambient temperature was 25–30 ◦ C. An effect of increasing the thickness and depth of
gravel on the temperature inside the dryer was not significant. The quality of dehydrated beans
was comparable with that of the product from the traditional sun drying.
Dilip and Jain (2004) developed an analytical model for an inclined multi-pass solar air heater
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

with an in-built thermal storage. The air heater was attached to a deep-bed dryer. The deep-bed
drying model was developed to evaluate the performance of a solar air heater for drying the paddy
crop in a deep bed. It was found that the higher rate of drying achieved with a larger depth of bed
and the humidity of the air increases in the drying bed with the increase in depth of the bed. The
efficiency increases with the increase in mass of the grain available in the drying bed. The proposed
mathematical model was useful to predict the moisture content, grain temperature, humidity of
drying air and drying rate in the grain bed. Hourly variation in grain temperature at the varying
mass flow rate of solar collector was analysed. It was found that the grain temperature increases
with a decrease in mass flow rate. There is a drastic drop in grain temperature with an increase
in mass flow rate from 0.014 to 0.042 kg/s. However, by increasing the mass flow rate beyond
0.042 kg/s, only a little drop in grain temperature was observed. Ayensu and Asiedu-Bondzie
(1986) designed and constructed a solar dryer based on convective heat and mass transfer. It
consisted of a single layer of glass as a glazing material, a drying chamber, a chimney and an
air collector-cum-rock bed storage (granite) insulated from the base ground by a thick layer of
straw. It is used for drying cassava leaves, cassava chips and fish. The solar collector is capable of
transferring 118 W/m2 to the drying air at a temperature of 32 ◦ C. The rock storage system stores
1.1 kW/hr energy to enhance drying. The efficiency of the solar collector is 22%.
Babagana, Silas, and Mustafa (2012) designed an inclined solar air heater with thermal storage
for drying food crops (tomato, onion, pepper, spinach, etc.). The storage unit is a compartment
that is located directly below the trays in the drying cabinet. It contains 3/4 in – 3 in rock pebbles
of 21 mm bed thickness. The heated air passes through the storage unit where some of the heat
energy is stored. During the off-sunshine hours, the stored energy is released to the air. An outlet
vent (chimney) was provided towards the upper end at the back of the cabinet to facilitate and
control the convection flow of air through the dryer. The performance of the system was analysed
during natural and forced convection of air. When using the forced mode system, the drying time

Figure 3. Solar dryer with black painted gravel as heat storage (Fagunwa, Koya, and Faborode 2009).
592 A. Agrawal and R.M. Sarviya

of tomato, onion, pepper, okra and spinach were 14, 15, 12, 11 and 1 h, while the drying rate
were 0.20, 0.020, 0.21, 0.22 and 0.77 kg/h, respectively. When using the natural mode system,
the drying time of tomato, onion, pepper and spinach were 24, 27, 25, 21 and 2 h, while the drying
rate were 0.12, 0.08, 0.09, 0.11 and 0.39 kg/h, respectively. The heat absorbed by the collector
was 60.62 W/m2 K. The collector efficiency was 45% and the useful heat of heat storage was
48.9 W/m2 K. The useful heat was utilised during the night hours for about 6 h for drying.
Berinyuy, Tangka, and Weka Fotso (2012) evaluated the performance of a tilted double-pass
natural convection solar tunnel air heater with in-built thermal storage attached with a tunnel dryer
for drying high moisture vegetables and other agricultural products. Heat storage was provided to
the collector absorber and to the drying tunnel floor by crushed basalt rocks (Cp = 880 J/kg ◦ C)
and the whole was covered by two layers of polyethylene sheets. At sunset during the wet har-
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

vesting period, the temperature inside the dryer was 5 ◦ C above the ambient because of additional
heating due to heat storage. A complete dryer could dry 17 kg of sliced cabbage with 95% moisture
content (wet basis) down to 9% in 5 days in a period characterised by intermittent downpours and
permanent cloud cover. The results further show that heat storage permits continuous drying dur-
ing periods of low sunshine and high relative humidity. The reduction in drying time was between
30% and 50%, depending on the crop, compared with natural sun drying and the final product
was acceptable in taste and colour. Jain (2007) proposed a transient analytical model to study
the natural flow solar crop dryer having a reversed absorber plate type collector with packed bed
thermal storage. The performance of this crop dryer with packed bed was carried out for drying
onions in trays. The thermal energy storage is effective during non-sunshine hours to reduce the
fluctuation in temperature for drying. Heights of packed bed affect the crop temperature. The
proposed analytical model is useful for calculating the crop temperature, drying rate and moisture
content of crops. It is also useful for evaluating the performance of thermal storage and reversed
solar collector.
Abdel-Galil (2007) experimentally investigated solar drying system with and without heat
storage system for drying Poultry Manure. Crushed Limestone was used as the heat storage
material. A drying chamber was attached with three separate solar collectors, namely a traditional
solar collector and two solar collectors with 5- and 10-cm crushed limestone under the absorber
plates. The outlet temperatures of the solar collector with 5-cm crushed limestone were higher
during the day-time than the corresponding values of the solar collectors with 0.0- and 10-cm
crushed limestone. The hourly average total useful energy gained and the thermal efficiency of the
solar collector with 5-cm crushed limestone were 0.281 kWh/m2 and 41.2%, while the values of
both the traditional solar collector (0.0-cm crushed limestone) and the solar collector with 10-cm
crushed limestone were 0.235 kWh/m2 , 34.2% and 0.257 kWh/m2 , 37.2%, respectively. For the
drying unit attached with the collector of 5-cm crushed limestone, the manure dried faster at all
manure depths followed by the drying unit attached with the solar collector of 10-cm crushed and
the drying unit attached with the traditional solar collector.

6.2. Solar dryers with sand as heat storage material

Sand is a suitable SHS material due to its properties such as low-cost, high specific heat and
thermal stability. Fatah (1994) developed a solar heater with sensible and LHS material. A number
of copper tubes filled with thermal storage were used as an absorber. Different sensible and LHS
materials were inserted into the copper tube and tested. Sand, Paraffin wax and Glauber’s salt were
used as heat storage material. The outlet temperature of the air was maintained 5 ◦ C above ambient
temperature for about 16 h. The daily average efficiency with paraffin wax was about 63.35% as
compared with 59% with sand as the storage material for 0.02 kg/s mass flow rate of air. The
effect of mass flow rate of air on the performance of the system was also studied. Increasing the
International Journal of Sustainable Energy 593

air flow rate increases the heat transfer rate, but deceases the effective period. Aboul-Enein et al.
(2000) studied analytically the performance of a solar air heater with and without thermal storage
for crop drying. The thermal storage material was inserted into the box under the absorber plate in
order to improve the performance of the drying process. Water, sand and granite were used as the
thermal storage material. The average temperature of flowing air increases with the increase in
the collector length and width. The outlet temperature of flowing air decreases with the increase
in the gap between the solar collector and cover. An optimum thickness of the storage material
of about 0.12 m was found to be appropriate for different agricultural products. The proposed
mathematical model is useful for evaluating the thermal performance of solar heaters.
Mohanraj and Chandrasekar (2008) designed a flat plate forced convection solar dryer integrated
with heat storage material. One side of the solar dryer was connected with a blower and other side
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

was connected with the dryer cabin. A glass cover of 5 mm thickness was used over the collector
to reduce the losses from the collector. A mixture of sand and aluminium scraps was used as heat
storage material and it was filled in the gap between the absorber and insulation. The gap between
the glass and the absorber surface was maintained at 25 mm for air circulation. It was concluded
that forced convection solar dryer is more suitable for producing high-quality copra for small
holders. About 75% of high-quality copra could be produced. The average thermal efficiency of
the solar air heater was estimated to be about 24%. El-Sebaii et al. (2002) designed and constructed
a natural convection solar dryer. The dryer consists of a flat plate solar air heater connected to
a cabinet that acts as a drying chamber. The air heater was designed to be able to insert various
storage materials under the absorber plate in order to improve the drying process. Sand was
used as the storage material. Drying experiments have been conducted with and without storage
materials for different varieties of fruit and vegetables. It was found that chemical pretreatment of
fruits and vegetables significantly reduces the drying time. The drying parameters such as drying
temperature, ambient temperature, relative humidity, solar irradiance and temperature distribution
in different parts of the system were recorded during this study. The system is capable of drying
10 kg of chemically treated grapes or green peas during 20 h of sunshine. The results showed that
the maximum temperature in the drying chamber was about 60 ◦ C.
A natural convection solar tunnel dryer integrated with the SHS material was developed and
tested for drying copra by Ayyappan and Mayilsamy (2010). Sand was used as the heat storage
material in the solar tunnel dryer. The experiments were carried out with and without the integration
of heat storage materials. The dryer reduces the moisture content of copra from 52% (wb) to 7.2%
(wb) in 52 and 78 h, respectively, with and without the heat storage material. The average solar
tunnel dryer thermal efficiency was estimated to be about 18% in both the drying modes. The
maximum temperature attained inside the solar tunnel dryer was 61 ◦ C and 52 ◦ C with and without
heat storage material, respectively, compared to a maximum ambient temperature of 32 ◦ C.

6.3. Solar dryers with auxiliary heating unit

Madhlopa and Ngwalo (2007) designed and constructed an indirect natural convention solar dryer
with biomass backup heaters. A biomass burner with a rectangular duct and flue gas chimney is
integrated with solar collector storage thermal mass. The thermal mass was placed in the top part
of the biomass burner enclosure. Thermal mass was made of rock pebbles and concrete. Results
show that the thermal mass was capable of storing part of heat from the biomass burner and
absorbed solar energy. The experimental test was conducted in three modes of operation (solar,
biomass and solar–biomass) using fresh pineapples under different weather conditions. A batch
of pineapples was successfully dried using solar energy only on clear days when sufficient solar
energy was available. Drying was continued, even when the intensity of solar energy was low, by
solar–biomass mode of operation. Moisture content of the pineapples was within acceptable safe
limit after drying.
594 A. Agrawal and R.M. Sarviya

6.4. Solar dryers with PCM as heat storage material

Very limited studies have been reported on the development of solar dryers with PCMs. Enibe
(2002) designed, constructed and evaluated the performance of a natural convention solar air heater
with heat storage using the PCM. Paraffin is used as the PCM and encapsulated in Modules. These
modules are made of rectangular channel whose tops are welded to the absorber plate and bottoms
rest on the bottom insulation of the collector and are welded to the solar collector. The underside
of the absorber plate, together with the vertical sides of the PCM modules, serves as an air heating
chamber. Experimental tests were conducted under natural environmental conditions involving
daily global irradiation in the range 4.9–19.9 MJ/m2 and ambient temperature variations in the
range 19–41 ◦ C. The peak cumulative efficiency of the solar dryer was about 50% and peak
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

temperature rise of the heated air was about 15 K. The transient thermal analysis of a previous
design was also studied by Enibe. Energy balance equations were developed for each major
component of the heater and linked with heat and mass balance equations for the heated air
flowing through the system. The maximum predicted airflow rate was 0.01 kg/s, corresponding
to a maximum inlet velocity of 0.33 m/s. The maximum predicted cumulative useful and overall
efficiencies of the system were within the range of 2.5–13 and 7.5–18%, respectively.
Kanyarat and Kiattisak (2010) performed an experimental investigation of a solar dryer inte-
grated with a PCM heat storage system. Paraffin was used as the heat storage material. The study
focused on the determination of the suitable melting point of the PCM and the enhancement of
thermal conductivity. Steel wool matrix was chosen for thermal conductivity enhancement. Two
mixtures of paraffin and kerosene with mass ratios of 2:1 and 1:1 were tested. It was found that
there is a decrease in the melting point of the mixture by adding kerosene. Paraffin was encap-
sulated in 12 galvanised steel tubes and 100 aluminium soft-drink cans. Diameter and length of
steel tubes were 4.5 and 170 cm respectively. The steel PCM tubes were placed directly under the
absorber sheet of solar collector while the aluminum PCM cans were placed on the floor of the
solar collector.
Lalit et al. (2011) designed and developed a solar dryer with a heat storage material. Paraffin wax
was used as the heat storage material. Performance of the solar dryer was considerably improved
by heat storage and drying was possible at steady and moderate temperatures of 40–75 ◦ C. The
solar collector was made of half split bamboo to reduce cost and weight. The temperature was
measured at different points such as inlet, outlet and below the solar panel with free natural
convection. The outlet temperature was sufficient for drying of food materials. Alkilani et al.
(2009) investigated analytically the effect of mass flow rate on outlet air temperature during the
discharge process for the solar air heater integrated with a PCM unit. The PCM unit consisted of
inline single rows of cylinders filled with a compound of paraffin wax and aluminium powder. The
cylinders were placed in the cross-flow of the forced air stream. Aluminium powder was added
to increase the thermal conductivity of the material. Simulation was performed using a Matlab
computer program to determine air temperature along the duct, the time required to discharge
all the thermal energy and freezing time for each cylinder. Simulation was performed for eight
different mass flow rates of air. The mass flow rate of air varies from 0.05 to 0.19 kg/s. The results
showed that the temperature of air decreases with the increase in mass flow rate of air and the
discharge takes a long interval of time for lower flow rates.
Song et al. (2011) worked on experimental simulation of a solar dryer for grape drying with
the LHS system. The solar dryer consisted of a collector, a drying room, a heat storage system,
an auxiliary heater and a control system as shown in Figure 4. Paraffin (melting temperature
48 ◦ C) was used as the heat storage material. Continuous drying was achieved in the solar dryer
by three modes of heating, i.e. collector heating, heat storage heating and auxiliary heating. As the
external condition changes, the solar dryer can achieve automatic conversion within three modes
by the control system. Its average efficiency of energy storage was above 66%. The function
International Journal of Sustainable Energy 595
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Figure 4. Schematic view of the solar dryer with heat storage for grape drying (Song et al. 2011).

of heat storage and control systems is to ensure continuous drying operation during day and
night. During day time, part of heat which is provided by solar radiations is transferred to the
air for heating materials and the remaining heat for heating the PCM. The dryer was able to
continue drying operation for 10 h in a day and 14 h in the night (drying temperature was 45 ◦ C).
The results of the experiment show that the dryer met the requirement of continuous drying
and the average time for drying 1 kg of grape from 87.9% (wb) to 14% (wb) was 6 days with
heat storage systems. The rate of drying increases by 2 times compared with that of traditional
drying. The average heat storage and instantaneous efficiency of the collector were 66% and 56%,
respectively.
Seshan Ram (2012) designed and experimentally investigated the performance of a solar tunnel
dryer with LHS for drying pineapple. Acetamide (melting point 80 ◦ C) was used as the heat storage
material. A shell and tube heat exchanger was designed as a PCM container. The PCM container
is to be placed very close to the air duct area of the dryer and the heat exchanger pipe is parallel
to the air inflow. The solar radiation penetrates through the toughened glass of the dryer and is
absorbed by the selective coated absorber sheet and thereby converted into thermal energy. As the
air passes through the container, heat is transferred to the air and this hot air is used for drying
during adverse weather conditions. It was concluded that the developed model is preferred to
be well utilised for solar drying applications for partial energy requirement during night hours.
Stalin and Barath (2013) conducted the theoretical study to determine the performance of phase
change in energy storage material for the solar drying system. The PCM was encapsulated in
spherical balls of diameter 10 cm with a cylindrical hole of 3 cm diameter in the centre. One
hundred and forty-four balls of PCM are placed on the surface of collector and arranged in the
compartment as 18 rows and 8 columns. Such design provides more heat transfer between PCM
and air during charging and discharging of PCM. The melting point and latent heat of selected
PCM were 150 ◦ C and 173 kJ/kg, respectively. The capacity of the drying chamber was 25 kg.
Initial and final moisture contents of the product were 50% and 30%, respectively. The drying time
for 25 kg of products and corresponding mass flow rate and other parameter were theoretically
calculated. The performance of the system was analysed for different mass flow rates of air, i.e.
0.020, 0.025 and 0.030 kg/s. The moisture content is minimum for 0.030 kg/s flow rate of air and
it was observed that the drying rate increased with the increase in mass flow rate.
Arunasalam, Srivatsa, and Senthil (2012) developed and tested a solar air heater with magnesium
chloride as the heat storage material. This work involved fabrication and testing of double reflector
and latent heat thermal energy system (LHTES) for the utilisation of solar energy during off-
sunshine hours. The primary dish collector reflects the incident solar rays to the secondary dish.
596 A. Agrawal and R.M. Sarviya

The secondary dish concentrates the reflected rays as point focus on the copper plate of the LHTES
and the heat transfer takes place through the fin to the PCM stored. The utilisation of PCM in the
heat storage is investigated experimentally for convective heat transfer between the PCM and air.
The performance of the system was compared to two different air flow rates of 1 and 2 m3 /min.
The higher mass flow rate showed slightly increased useful heat gain. The experimental results
showed that the system is capable to supply hot air at 80 ◦ C for more than 4 h with a temperature
gradient of air 30 ◦ C.
Cakmak and Yıldız (2011) compared drying characteristics of grape in a novel solar dryer with
and without swirl element. The dryer is coupled to a solar air collector and a solar air collector
with phase change material to receive hot air during sunshine and off-sunshine hours, respectively.
Calcium chloride hexahydrate was used as heat storage material. Through the mirrors mounted
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

onto tripod edges of the collector which had PCM, it was ensured that sun rays could fall better
onto the collector surface. The drying experiments were carried out simultaneously both under
natural conditions and by the dryer with swirl flow and without swirl flow at three different air
velocities (0.5, 1, and 1.5 m/s). Drying experiments with PCM solar collector were performed at
1.5 m/s air velocity. It was observed that the shortest drying time was achieved with 1.5 m/s air
velocity (with swirl). The dryer reduced the moisture content of seeded grape from 3 to 0.09 kg
water/kg drymatter in 56 and 80 h, respectively, with and without the heat storage material with
an air velocity of 1.5 m/s.
Devahastin and Pitaksuriyarat (2006) experimentally investigated the feasibility of LHS using
paraffin wax as PCM for the solar drying system. The heat storage system absorbs excess heat
of air exhausted from the solar air collector and releases it when the solar energy availability is
inadequate or not available, by forcing ambient air through the energy storage to extract the stored
energy. The effect of temperature and velocity of hot air on charge time was determined. During the
experiments velocity and temperature of air ranged between 1–2 m/s and 70–90 ◦ C„ respectively.
During the discharge period only the effect of inlet ambient air velocity was considered. It was
found that melting was dominated by heat conduction followed by free convection; melting took
place from the centre of the LHS to a point far away in the radial direction and took place
from top to bottom points in the axial direction. However, only heat conduction was dominant
in the solidification process. It was found that charge time decreased with increase in inlet air
temperature and velocity. The amount of the energy extractable from the LHS ranged between
1920 and 1386 kJ min/kg and the energy savings ranged between 40% and 34% when using an
inlet ambient air velocity of 1 and 2 m/s, respectively.

6.5. Solar dryers with water as the heat storage material

Water is the most useful material that can be used for heat storage due to low cost, high specific
heat and easy availability. Vlachos et al. (2002) designed a novel low-cost indirect solar tray dryer
equipped with a solar air collector, a heat storage cabinet and a solar chimney. Inside the heat
storage cabinet, there are 25 sealed metallic containers (5 L each) filled with water and painted
black to facilitate heat storage. The design is based on energy balances and on an hourly averaged
radiation data reduction procedure for tilted surfaces. The dryer was tested for different weather
conditions, i.e. sunny, cloudy and rainy. Drying was also tested during night operations. The
heat storage cabinet showed good performance during the night since the moisture content of
the product decreases continuously. Fairly promising results were obtained regarding the solar
dryer’s efficiency. It was found that dryer efficiency can be improved by properly adjusting the
flow rate and temperature of the air entering the drying chamber.
A hybrid solar dryer has been designed and constructed by Amer et al. (2006). It consists of a
solar collector, a drying chamber, a reflector and a heat exchanger cum heat storage unit. During
International Journal of Sustainable Energy 597

sunny day water gains heat from heated air through the heat exchanger and it also absorbs heat
from direct solar radiation. Using the water tank with the solar dryer, about 15 ◦ C can be stored
in water during the time of sunshine, while at night hot water flows through the heat exchanger to
heat the air in the reverse direction, and it was observed that the drying air temperature rose from
25 ◦ C to 35 ◦ C above the ambient air.
Luna, Nadeau, and Jannot (2010) developed solar kiln with SHS using water as the storage
material (Figure 5). It is composed of four main parts: an air solar collector, a water solar collector,
the drying chamber and the storage unit. The energy storage unit was made of two components:
the exchange-storage unit and the heating unit. It uses water as the storage fluid and air as the heat
extraction fluid. The storage unit is connected to a solar water heater and it contains numbers of
vertical tubes where a part of the air is heated. The solar air collector is also a part of the system
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

where air is heated by solar energy. The system is tested with and without heat storage. The
storage system gives a gain in drying time of 50 h during the month of April; this gain increases
to 62 h during August and 65 h during December, which gives around 30% as an average save in
the drying time.
Puiggali and Penot (1983) integrated water as the heat storage material in a natural convection
solar dryer. During sunshine hours the solar collector is used to heat inlet ambient air. This air
passes through the tubes in the storage unit and goes to the drying chamber. After the flow in the
course of the trays, the air goes out by means of the chimney. During this period water is also
heated using the solar energy. In the off-sunshine hours air passes through the tubes in the hot
water container where energy is recovered to the ambient air. Then heated air flows through the
drying chamber and the chimney. It was found that this type of dryer is suitable for drying apples,
mushrooms, peaches and apricots. The results showed that coupling the solar collector to the
chimney increases the efficiency of the collector from around 24 to 44%. The total efficiency of
the solar dryer was around 45%. As well, 50% of the stored energy is directly lost to the ambient
environment during the night and 21% of the stored energy in the day is recovered during the
night with only 4% used for drying.
Tiwari et al. (1997) performed analytical studies of a crop drying cum water heating system.
The water heater below the air heater systems will act as a storage material for drying the crop
during off-sunshine hours. In this study energy balance equations for each component of the
system were used to predict the analytical results. The system can be used to provide hot water

Figure 5. Solar kiln dryer with water as the heat storage material (Luna, Nadeau, and Jannot 2010).
598 A. Agrawal and R.M. Sarviya

in case the drying system is not in operation. The following conclusions have been drawn on the
basis of the analytical studies: (1) there is no significant effect on drying time due to an increase
in water mass after 0.10 m depth, except a decrease in its temperature for large flow rates of air,
and (2) the outlet air temperature becomes constant after 3.5 m length of air collector.

6.6. Solar dryers with oil as the heat storage material

Potdukhe and Thombre (2008) designed, developed and experimentally investigated the perfor-
mance of in-built thermal-storage agro solar dryer for agricultural product such as chillies and
fenugreek leaves. Thermic oil was used as the thermal storage material. It reduces the drying
time and improves the quality of dried product. Higher temperature of around 65 ± 3 ◦ C was
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

maintained in the drying chamber due to the integration of thermic oil as an in-built storage with
the absorber. The length of operation of the solar dryer increased by 1–2 h. The in-built storage
also maintains a high temperature for a longer period. This also exhibits excellent control over the
airflow rate and the drying rate by the dryer. The drying efficiency of the solar dryer for chillies
is 21%. The drying period in the solar dryer was decreased by 40% as compared with that using
the conventional dryer. The dryer operates on thermal buoyancy and natural draught, so there is
no requirement of fan to crate the forced draught. Hence the running cost of this dryer is very
less. This dryer is suitable for agro products such as chillies, onion and grapes that are sensitive
to direct exposure to solar radiation. The mathematical model developed can be utilised to predict
the temperature at various locations of the dryer and to optimise absorber plate.

6.7. Solar dryers with desiccant bed as thermal storage

Shanmugam and Natarajan (2007) designed, developed and experimentally investigated the per-
formance of an indirect forced convection solar dryer with desiccant bed as the thermal storage
material. It was composed of mixture of 60% of bentonite, 10% of calcium chloride, 20% of
vermiculite and 10% of cement. The performance of this dryer was carried out by drying green
peas and pineapple slices in trays. Reflective mirrors are used for boosting the performance of the
dryer. The system was operated in two modes, sunshine hours and non-sunshine hours. During
sunshine hours the desiccant bed receives solar radiation directly and through the reflected mirror
and simultaneously hot air from the flat plate collector is forced to the drying chamber for drying
the product. In the non-sunshine hours air inside the drying chamber was circulated through the
desiccant bed by a reversible fan. Air gets heated by the desiccant bed and drying of food continues
during the non-sunshine hours. Due to heat storage, drying time reduces by 10–12 h. The useful
temperature rise of about 10 ◦ C was achieved with mirror. Dryer average efficiency and drying
rate of the desiccant integrated solar dryer were also high compared with that of the normal solar
dryer without thermal storage. The quality of dried food product was also high in terms of colour
and moisture content. Drying was uniform in all trays. Desiccant material reduces the drying time
and improves the quality of dried product.

7. Comparison of performance of solar dryers with thermal heat storage materials

7.1. Comparison of the performance of solar dryers with rock bed heat storage

A brief comparison of performance of various solar dryers (with a rock bed as heat storage)
discussed earlier is tabulated. The observations of a few investigators who have worked on solar
dryers with rock bed heat storage are presented in Table 3.
Table 3. Summary of the performance of various solar dryers using rock bed as the heat storage material.

Heat storage Storage Nature Decrease in Flow Duration Drying


Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

material geometry of work moisture content rate for drying product Performance References

Rock Rock bed Experimental 20% to safe level 4.9 m3 /s Drying time Peanut Collector area was Butler and Troeger
22–25 h 140 m2 ,total mass of (1980)
rock = 103 Mg, rock
storage provides 74%
contribution in total energy
used
Rock Rock bed Theoretical 28.2–11.4% (db) 250 kg/hm2 2 days and 2 nights Coriander Drying time reduced by one Chauhan,

International Journal of Sustainable Energy


day with heat storage Choudhury, and
Garg (1996)
Black painted Store in chamber Experimental 53.4–3.6%, 9.09% 1.02 and 72 h Cocoa beans Inter-mittent drying process Fagunwa, Koya,
gravel attached with and 7.11% (wb) 1.32 m3 /min is possible. Temperature and Faborode
drying chamber (natural above 6 ◦ C from ambient is (2009)
convection) maintained at 6 am
Rock pebbles Alternate layer of Experimental 85–10% (wb) 0.024 m3 /s 72 h Pineapple Dryer efficiency was 15%, Madhlopa and
and concrete rock pebbles (natural 11% and 13% for solar, Ngwalo (2007)
and concrete convection) biomass and solar–biomass
modes of operation,
respectively
Crushed basalt Rock bed Experimental 80.9–10.8% (wb) 9.68 m3 /hr 96 h Vegetables Temperature 5 ◦ C above the Berinyuy, Tangka,
rock (natural (cabbage and ambient temperature at and Weka Fotso
convection) red pepper) sunset, reduction in drying (2012)
time was between 30% and
50% due to heat storage
Rock pebbles Store in chamber Experimental 75% to 20%(wb) 0.0135 m3 /s 14 h (forced Vegetables Useful heat of heat storage Babagana, Silas,
located directly for tomato (natural mode), 24 h (tomato, onion, was 48.9 W/m2 K. Drying and Mustafa
below the trays convection) (natural mode) pepper and time of tomato was 14 h (2012)
in the drying for tomato spinach) (natural mode), 24 h (forced
cabinet. mode)
Crushed Lower section of Experimental 118.8% to 10.6% 0.02 m3 /s 28 h (2 cm manure Poultry manure Useful energy gained and the Abdel-Galil
limestone the collector (db) depth) thermal efficiency of the (2007)
(5-cm and solar collector with 5-cm
10-cm) crushed limestone were
0.281 kWh/m2 and 41.2%

599
600 A. Agrawal and R.M. Sarviya

It is observed that solar dryers with rock bed heat storage take lesser time to reach the safe level
of moisture content when compared with solar drying without heat storage. The introduction of
rock bed in the dryer enhances the rate of drying and reduces the drying time by nearly 30–40%.
Due to heat storage, drying is possible at a steady rate and drying is possible during non-sunshine
hours. Babagana, Silas, and Mustafa (2012) reported that due to heat storage, drying of vegetable
products was possible during the night for about 6 h and the moisture content of tomato reduces
from approx 1.7% to 0.7% (db), under the natural circulation of air. Chauhan, Choudhury, and
Garg (1996) also reported that drying time is reduced by one day due to heat storage
and moisture content of coriander reduces from 28.2% to 11.4% (db) in 2 days and 2
nights.
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

7.2. Comparison of performance of solar dryers with sand as the heat storage material

A few investigators have used sand as heat storage material in solar dryers for drying fruits and
vegetables and their observations are summarised in Table 4. The dryer developed by El-Sebaii
et al. (2002) is technically sound and economical. Provision is made for SHS (sand) which
facilitates the dryer to be used beyond sunshine hours. The initial investment is moderate and
affordable by farmers. Mohanraj and Chandrasekar (2008) designed and developed a forced
circulation solar dryer with sand and Al scrap as the heat storage material. Drying copra in the
dryer reduced its moisture content from about 51.8% to 7.8% and 9.7% in 82 h, for trays at the
bottom and top, respectively. A natural circulation solar tunnel dryer was developed by Ayyappan
and Mayilsamy (2010) with sand as the heat storage material. The maximum temperature gain
due to heat storage was 29 ◦ C and drying time reduced by 26 h due to heat storage. The use of heat
storage material provides continuous drying and the copra obtained from the solar tunnel dryer
is of high quality.

7.3. Comparison of performance of solar dryers with PCMs, water and oil as heat storage
materials

Some studies using PCMs, water and limestone have also been reported. The incorporation of
PCM in solar dryers has grown to be of interest to the researchers due to its high energy storage
capacity per unit weight. Recently Cakmak and Yıldız (2011) experimentally evaluated the effect
of heat storage on drying kinetics of seeded grape and it was found that drying time considerably
reduced due to heat storage and drying is possible even after sunset. It has been determined that
drying time shortens as drying air velocity increases. Recently Song et al. (2011) designed and
tested a solar dryer with LHS for continuous solar drying. The results show that the system met
the requirement of continuous drying. The drying operation can be ensured at 10 h in the day
and 14 h at night. Devahastin and Pitaksuriyarat (2006) experimentally evaluated the feasibility
of using a PCM to conserve energy during drying and its effect on drying kinetics.
A few investigations have been carried out to evaluate the performance of a solar dryer with
water as the heat storage medium. Vlachos et al. (2002) designed a low-cost solar dryer with a heat
storage cabinet. It was found that drying was continued at night, but at a slower rate (5–3.5%). The
dryer was tested in different weather conditions (sunny, cloudy or rainy), and the drying process
reached full completion in all tests at a reasonable rate of dehydration. Recently, Luna, Nadeau,
and Jannot (2010) designed a solar kiln using water as the storage material. Shell and tube-type
heat exchanger was used to transfer and extract heat from water during sunshine and off-sunshine
hours. It was reported that saving in drying time due to heat storage was around 30%. Potdukhe
and Thombre (2008) used thermic oil heat storage in the solar dryer. Length of operation of the
solar dryer increased by 1–2 h and it also maintained high temperature for a longer period due to
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Table 4. Summary of the performance of various solar dryers using sand as the heat storage material.

Heat storage Storage Nature Decrease in Flow Duration Drying


material geometry of work moisture content rate for drying Product Performance References

Sand & Al scrap Lower section of Experimental 51.8–7.8% (wb) 2.4 m3 /s 82 hr Copra Average thermal Mohanraj and
the collector efficiency 24%, Chandrasekar
average drying (2008)

International Journal of Sustainable Energy


air temperature
43 ◦ C, peak
drying air
temperature
63 ◦ C, velocity
of air 1.2 m/s
Sand Lower section of Experimental 80–18% (db) Natural convection 60 and 72 hr with Fruits (Seedless Drying air El-Sebaii et al.
the collector (for Seedless and without grapes, apple) temperatures at (2002)
grapes) heat storage and vegetables the inlet of the
(for Seedless (tomatoes, drying chamber
grapes) green peas) is 45.5–55.5 ◦ C,
storage material
reduces the
drying process
by 12 h
Sand Base of the tunnel Experimental 52–7.2% (wb) Not mentioned 52 and 78 h, Copra Drying time Ayyappan and
dryer respectively, reduced by 26 Mayilsamy
with and hours by heat (2010)
without the heat storage, max
storage material temperature
gain due to heat
storage 29 ◦ C

601
602
Table 5. Summary of the performance of various solar dryers using PCMs, water and limestone as heat storage materials.

Heat storage Storage Nature Decrease in Flow Duration Drying


Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

material geometry of work moisture content rate for drying Product Performance References

Calcium chloride Lower section of Experimental 75–8.5% (wb) Not mentioned 56 and 80 h, Seeded grape The effective moisture Cakmak and
hegzahidrat the collector respectively, diffusivity of seeded Yıldız (2011)
(PCM) with and grapes was found
without the heat to be range between
storage 5.47 × 10−10 and
1.35 × 10−9 m2 /s and
maximum for 1.5 m/s
Paraffin Not mentioned Experimental 87–8% (wb) Not mentioned 6 days Grape Average efficiency of Song et al. (2011)
heat storage is 66%,

A. Agrawal and R.M. Sarviya


maximum heat transfer
coefficient 26.4 w/m2 -
C, energy density
54.5 MJ/m3 the rate of
drying increases by 2
times
Paraffin Cylindrical Experimental 70–13% (wb) Not mentioned 5h Potatoes Maximum energy savings Devahastin and
container was 40% during drying Pitaksuriyarat
of potatoes due to heat (2006)
storage, velocity of air
1 m/s
Water Lower section of Analytical 85–20% (db) Not mentioned 24 h Crop Dryer Drying time decreases Tiwari et al.
the collector with reflector, max (1997)
temperature with
reflector is 83 ◦ C, air
temperature in the
morning is 53 ◦ C and
70 ◦ C without and with
reflector
Water 25 sealed metallic Experimental 85–0% (wb) 200 m3 /h 25 h Polyurethane Products water content Vlachos et al.
containers foam (reference continued to decrease (2002)
(5 L each) and material) in night although at a
painted black lower rate (5–3.5% )
International Journal of Sustainable Energy 603

heat storage. Table 5 summarize and illustrate the performance of solar dryers using PCMs, water
and limestone as heat storage materials.

8. Conclusion

It has been established that solar drying of food is possible during the non-sunshine hours or
night by using a solar dryer having a heat storage system. The use of solar drying for food and
agricultural products has a large potential from the technical and energy-saving point of view.
Different types of materials such as rock, water, sand and granite, metal scrap, pure paraffin wax, a
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

mixture of aluminium power and paraffin wax have been used as storage materials in solar dryers
by researchers over the last few decades. SHS materials such as rocks and oil were used in most
of the solar dryers. Very less information is available regarding the use of LHS materials in solar
dryers. Latent heat can be used when space restrictions limit larger thermal storage units in solar
drying systems. Paraffin wax was used as a heat storage material in most of the previous research
works because of its low cost and easy availability, but the major drawback of paraffin is low
thermal conductivity. This drawback can be minimised by increasing the thermal conductivity by
artificial means such as inserting a metal matrix of high thermal conductivity and adding fins to
the heat transfer surface. For good thermal performance of a solar air heater, a PCM with high
latent heat and with large surface area is required. There is need for developing a compact, high
efficiency, long life and economical solar dryer having a heat storage system. More research on
solar energy collection and thermal storage is required to reduce the total cost, volume and heat
loss.

References

Abdel-Galil, H. S. 2007. “Solar System with Energy Storage for Drying Poultry Manure.” Misr Journal of Agricultural
Engineering. 24 (4): 978–1003.
Abhat, A. 1983. “Low Temperature Latent Heat Thermal Energy Storage: Heat Storage Materials.” Solar Energy 30 (4):
313–332.
Aboul-Enein, S., A. A. El-Sebaii, M. R. I. Ramadan, and H. G. El-Gohary. 2000. “Parametric Study of a Solar Air Heater
with and Without Thermal Storage for Solar Drying Applications.” Renewable Energy 21 (3–4): 505–522.
Alkilani, M. M., K. Sopian, M. Sohif, and M. Alghol. 2009. “Output Air Temperature Prediction in a Solar Air Heater
Integrated with Phase Change Material European.” Journal of Scientometric Research 27 (3): 334–341.
Amer, B. A. A., K. Gottschalk, and J. Hahn.2006. “New Design for Constructing a Proto-Type of Solar Dryer. In Paper
presented at the 15th International Drying Symposium, Szent Istvan University, Budapest, Hungary, August 20–23,
edited by I. Farkas, Vol. B, pp. 790–796. http://fft.szie.hu/events/pp.html
Arjunan, T. V., H. S. Aybar, and N. Nedunchezhian. 2009. “Status of Solar Desalination in India.” Renewable and
Sustainable Energy Reviews 13 (9): 2408–2418.
Arunasalam, A., B. Srivatsa, and R. Senthil. 2012. “Thermal Performance Analysis on Solar Integrated Collector Storage.”
UARJ 1 (2): 2278–1129.
Ayensu, A., and V. Asiedu-Bondzie. 1986. “Solar Drying with Convective Self-Flow and Energy Storage.” Solar and Wind
Technology 3 (4): 273–279.
Ayyappan, S., and K. Mayilsamy. 2010. Solar Tunnel Drier with Thermal Storage for Drying of Copra. Proceedings of
the 37th National & 4th International Conference on Fluid Mechanics and Fluid Power, (FMFP10), IIT Madras,
Chennai, India, NE–06, 1–10.
Babagana, G., K. Silas, and B. G. Mustafa. 2012. “Design and Construction of Forced/Natural Convection Solar Vegetable
Dryer with Heat Storage.” ARPN Journal of Engineering and Applied Sciences 7 (10): 1213–1217.
Belen, Z., M. M. Jose, F. C. Luisa, M. Harald, 2003. “Review on Thermal Energy Storage with Phase Change: Materials,
Heat Transfer Analysis and Applications.” Applied Thermal Engineering 23 (3): 251–283.
Berinyuy, J. E., J. K. Tangka, and G. M. Weka Fotso. 2012. “Enhancing Natural Convection Solar Drying of High Moisture
Vegetables with Heat Storage.” Agricultural Engineering International: CIGR Journal 14 (1): 141–148.
Brooker, D. B., F. W. Bakker-Arkema and C. W., Hall. 1992. Drying and Storage of Grain and Oilseeds. New York: Van
Nostrand Reinhold.
Buddhi, D., and R. L. Sawhney. 1994. Proceeding of Thermal Energy Storage and Energy Conversion. School of Energy
and Environmental Studies. Indore: Devi Ahilya University.
604 A. Agrawal and R.M. Sarviya

Butler, J. L., and J. M. Troeger. 1980. Drying Peanuts Using Solar Energy Stored in a Rockbed. Agricultural Energy, Vol.
I, Solar Energy, Selected Papers and Abstracts. St Joseph, MI: ASAE Publication.
Cakmak, G., and C. Yıldız. 2011. “The Drying Kinetics of Seeded Grape in Solar Dryer with PCM-Based Solar Integrated
Collector.” Food and Bioproducts Processing 89 (2): 103–1108.
Chauhan, P. M., S. C. Choudhury, and H. P. Garg. 1996. “Comparative Performance of Coriander Dryer Coupled to Solar
Air Heater and Solar Air-Heater-Cum-Rock Bed Storage.” Appl Therm Eng. 16 (6): 475–486.
Choudhury, S. C., P. M. Chauhan, and H. P. Garg. 1995. “Economic Design of a Rock Bed Storage Device for Storing
Solar Thermal Energy.” Solar Energy 55 (1): 29–33.
Climator. Accessed on June 21, 2014. Available from www.climator.com
Cohen, J. S., and T. C. S. Yang. 1995. “Progress in Food Dehydration.” Trends in Food Science and Technology 6 (1):
20–25.
Cristopia Energy Systems. Accessed on June 21, 2014. Available from www.cristopia.com
Devahastin, S., and S. Pitaksuriyarat. 2006. “Use of Latent Heat Storage to Conserve Energy During Drying and its Effect
on Drying Kinetics of a Food Product.” Applied Thermal Engineering 26 (14–15): 1705–1713.
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Dilip, J., and R. K. Jain, 2004. “Performance Evaluation of an Inclined Multi-Pass Solar Air Heater with in-Built Thermal
Storage on Deep-Bed Drying Application.” Journal of Food Engineering 65 (4): 497–509.
El-Sebaii, A. A., S. Aboul-Enein, M. R. I. Ramadan, and H. G. El-Gohary. 2002. “Experimental Investigation of an Indirect
Type Natural Convection Solar Dryer.” Energy Conversion & Management 43 (16): 2251–2266.
Enibe, S. O. 2002. “Performance of a Natural Circulation Solar Air Heating System with Phase Change Material Energy
Storage.” Renewable Energy 27: 69–86.
Environmental Process Systems Ltd. Accessed on June 21, 2014. Available from www.epsltd.co.uk
Fagunwa, A. O., O. A. Koya, and M. O. Faborode. 2009. “Development of an Intermittent Solar Dryer for Cocoa Beans.”
Agricultural Engineering International: The CIGR Ejournal. Vol. XI: 1–14 (Article number 1292).
Farid, M. M., A. M. Khudhair, S. A. Razack, and S. Al-Hallaj. 2004. “A Review on Phase Change Energy Storage:
Materials and Applications.” Energy Conversion and Management 45 (9–10): 1597–1615.
Fatah, H. E. S. 1994. “Thermal Performance of a Simple Design Solar Air Heater with Builtin Thermal Energy Storage
System.” Energy Conversion and Management 36 (10): 989–997.
Garg, H. P., S. C. Mullick, and A. K. Bhargava. 1985. Solar Thermal Energy Storage, 154–291. Dordrecht: D. Reidel
Publishing Co.
Garg, H. P., V. K. Sharma, R. B. Mahajan, and A. K. Bhargave. 1985. “Experimental Study of an Inexpensive Solar
Collector Cum Storage System for Agricultural Uses.” Solar Energy 35 (4): 321–331.
Ghoneim, A. A., and S. A. Klein. 1989. “The Effect of Phase Change Material Properties on the Performance of Solar Air
Based Heating Systems.” Solar Energy 42 (6): 441–447.
Hale, D. V., M. J. Hoover, and M. J. O’Neill. 1971. Phase Change Materials Hand Book. Report no. HREC- 5183-2LMSC-
HREC D225138. NASA Marshal Space Flight Center. Alabama.
Hasnain, S. 1998. “Review on Sustainable Thermal Energy Storage Technologies, Part I: Heat Storage Materials and
Techniques.” Energy Conservation and Management 39 (11): 1127–1138.
Herrick, C. S., and D. C. Golibersuch. 1977. “Qualitative Behavior of a New Latent Heat Storage Device for Solar
Heat/Cooling Systems.” General Electric Technical Information Series. Report 77CRD0006, March 1977.
Hoogendoorn, C. J., and G. C. J. Bart 1992. “Performance and Modelling of Latent Heat Stores.” Solar Energy 48 (1):
53–58.
IEA. 1998. Solar Heating and Cooling Programme: Potential for solar drying in the world, Oct 1998.
Jain, D. 2007. “Modeling the Performance of the Reversed Absorber with Packed Bed Thermal Storage Natural Convection
Solar Crop Dryer.” J Food Engineering 78 (2): 637–647.
Joseph Stalin, M and Barath P. 2013. “Effective Utilization of Solar Energy in Air Dryer.” IJMPERD 3 (1): 133–142.
Kakiuchi, H. 2002. Mitsubishi Chemical Corporation, private communication.
Kanyarat, H., and K. Kiattisak. 2010. Using PCM as Thermal Energy Storage in Solar Dryer. The 3rd Technology and
Innovation for Sustainable Development International Conference (TISD2010), Khon Kaen University, Nongkai
Province,Thailand.
Keey, R. B. 1978. Introduction to Industrial Drying Operations. 1st ed. New York: Pergamon Press, Chapter 2.
Lalit M. B., S. Santosh, and S. N. Naik. 2010. “Solar Dryer with Thermal Energy Storage Systems for Drying Agricultural
Food Products: A Review.” Renewable and Sustainable Energy Reviews 14 (8): 2298–2314.
Lane, G. A. 1978. Macro-encapsulation of PCM. Report no. ORO/5217-8. Midland, MI: Dow Chemical Company, 152.
Lane, G. A. 1983. Solar Heat Storage: Latent Heat Materials – Background and Scientific Principles. Vol. I. Boca Raton,
FL: CRC Press.
Lane, G. A., and Glew, D. N. 1975. “Heat of Fusion System for Solar Energy Storage.” Proceedings of the Workshop on
Solar Energy Storage Subsystems for the Heating and Cooling of Buildings, Charlothensville, Virginia, 43–55.
Lalit, M. B., S. Santosh, S.N. Naik, and V. Meda. 2011. “Review of Solar Dryers with Latent Heat Storage Systems for
Agricultural Products.” Renewable and Sustainable Energy Reviews 15 (1): 876–880.
Luna, D., J. P. Nadeau, and Y. Jannot. 2010. “Model and Simulation of a Solar Kiln with Energy Storage.” Renewable
Energy 35 (11): 2533–2542.
Madhlopa, A., and G. Ngwalo. 2007. “Solar Dryer with Thermal Storage and Biomass Backup Heater.” Solar Energy 81
(4): 449–462.
Ministry of New and Renewable Energy (MNRE) Annual Report 2010 (www.mnre.gov.in).
Mohanraj, M., and P. Chandrasekar. 2008. “Drying of Copra in Forced Convection Solar Drier.” Biosystem Engineering.
99 (4): 604–607.
International Journal of Sustainable Energy 605

Mujumdar, A. S. 1987. Advances in Drying, Vol. 4. Washington, DC: Hemisphere Publishing Corporation.
Myer, K. 2007. “Food Drying and Evaporation Processing Operations.” Chap. 11 in Handbook of Farm, Dairy, and Food
Machinery, edited by K. Myer, 303–340. Norwich: William Andrew Publishing.
Potdukhe, P. A., and S. B. Thombre. 2008. “Development of a New Type of Solar Dryer: Its Mathematical Modeling and
Experimental Evaluation.” International Journal of Energy Research 32 (8): 765–782.
Puiggali, J. R., and F. Penot. 1983. “Analyse du Comportement Dynamique et Thermique d’un sé Choir Solaire Constitué
d’un Capteurá Matrice Poreuse Couplé á une Cheminée Solaire.” Revue de Physique Appliquée. 18 (10): 625–633.
Purohit, P., A. Kumar, and T. C. Kandpal. 2006. “Solar Drying vs. Open Sun Drying: A Framework for Financial
Evaluation.” Solar Energy 80 (12): 1568–1579.
Reddy, B. S., and J. P. Painuly. 2004. “Diffusion of Renewable Energy Technologies – Barriers and Stakeholders’
Perspectives.” Renewable Energy 29 (9): 1431–1447.
Rubitherm Technologies GmbH. Accessed on June 21, 2014. Available from www.rubitherm.com
Saffa R., M. Blaise, and F. Wenbo. 2013. “Phase Change Material Developments: A Review.” International Journal of
Ambient Energy, doi:10.1080/01430750.2013.823106
Downloaded by [Univ of Alhussein Bin Talal] at 04:03 12 November 2017

Seshan Ram, M. R. 2012. “Experimental Investigation of Energy Efficient Solar Tunnel Dryer.” Voice of Research 1 (3):
50–56.
Sethi, V. P., and S. K. Sharma. 2008. “Survey and Evaluation of Heating Technologies for Worldwide Agricultural
Greenhouse Applications.” Solar Energy 82 (9): 832–859.
Sharma, V. K., A. Colnagelo, and G. Spagna. 1993. “Experimental Performance of an Indirect Type Solar Food and
Vegetable Dryer.” Energy Conversion Management 34 (4): 293–298.
Shanmugam, V., and E. Natarajan. 2007. “Experimental Study of Regenerative Desiccant Integrated Solar Dryer with and
Without Reflective Mirror.” Applied Thermal Engineering 27 (8–9): 1543–1551.
Sharma, A., V. V. Tyagi, C. R. Chen, and D. Buddhi. 2009. “Review on Thermal Energy Storage with Phase Change
Materials and Applications.” Renewable and Sustainable Energy Reviews 13 (2): 318–345.
Song, M., Y. Songlin, Z. Biguang, and Z. Dong. 2011. “Experimental Research of Grape Drying Using Solar Dryer
with Latent Heat Storage System.” International Conference on Computer Distributed Control and Intelligent
Environmental Monitoring, 740–742, (CDCIEM), Changsha. IEEE. doi:10.1109/CDCIEM.2011.33
Teappcm. Accessed on June 21, 2014. Available from www.teappcm.com
Tiwari, G. N., P. S. Bhatia, A. K. Singh, and R. K. Goyal. 1997. “Analytical Studies of Crop Drying Cum Tyagi, V. V., and
D. Buddhi. 2007. “PCM Thermal Storage in Buildings: a State of Art.” Renewable and Sustainable Energy Review
11 (6): 1146–1166.
Tiwari, G. N., P. S. Bhatia, A. K. Singh, R. K. Goyal. 1997. “Analytical Studies of Crop Drying Cum Water Heating
System.” Energy Conversion and Management 38 (8): 751–759.
Tiwari, G. N.,A. K. Singh, and P. S. Bhatia. 1994. “Experimental Simulation of a Grain Drying System.” Energy Conversion
and Management 35 (5): 453–458.
Tyagi, V. V., and D. Buddhi. 2007. “PCM Thermal Storage in Buildings: A State of Art.” Renewable and Sustainable
Energy Reviews 11 (6): 1146–1166.
Verma, P., P. Varun, and S. K. Singal. 2008. “Review of Mathematical Modeling on Latent Heat Thermal Energy Storage
Systems Using Phase-Change Material.” Renewable and Sustainable Energy Reviews 12 (4): 999–1031.
Vlachos, N. A., T. D. Karapantsios, A. I. Balouktsis, and D. Chassapis. 2002. “Design and Testing of a New Solar Tray
Dryer.” Drying Technology 20 (5): 1239–1267.
Zhang, Y., G. Zhou, K. Lin, Q. Zhang, and H. Di. 2007. “Application of Latent Heat Thermal Energy Storage in Buildings:
State-of-the-Art and Outlook.” Build Environment 42 (6): 2197–2209.
Zhou, D., C-Y. Zhao, and Tian, Y. (Yuan). 2012. “Review on Thermal Energy Storage with Phase Change Materials
(PCMs) in Building Applications.” Applied Energy 92: 593–605. doi:10.1016/j.apenergy.2011.08.025

You might also like