You are on page 1of 16

Desalination 385 (2016) 178–193

Contents lists available at ScienceDirect

Desalination

journal homepage: www.elsevier.com/locate/desal

Engineering advance

The effects of design parameters on productivity performance of a solar


still for seawater desalination: A review
Mohammed Shadi S. Abujazar a, S. Fatihah a,b,⁎, A.R. Rakmi b, M.Z. Shahrom a
a
Civil and Structural Engineering Department, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, 43600 Bangi, Selangor, Malaysia
b
Research Centre for Sustainable Technology (CESPRO), Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, 43600 Bangi, Selangor, Malaysia

H I G H L I G H T S

• Performance of the solar stills is affected by environmental, design and operation parameters.
• Design parameters can be improved to increase the solar still productivity.
• Stepped solar still for enlarging the evaporation area is used to increase the productivity.
• Besides, high thermal conductivity materials increase the workability and productivity.

a r t i c l e i n f o a b s t r a c t

Article history: This paper aims to investigate the different parameters that affect solar still productivity when the solar still pro-
Received 4 September 2015 ductivity is very low compared with other desalination systems, such as other thermal processes or membrane
Received in revised form 1 February 2016 processes. These parameters include environmental, design and operational parameters. The results show that
Accepted 17 February 2016
productivity was highly affected by environmental parameters due to the unpredictability of metrological fac-
Available online xxxx
tors. When design and operational parameters were varied, increases in productivity were observed. The results
Keywords:
indicated that the solar still was inversely affected by evaporation area, water depth, a minimization in the angle
Desalination of the solar still cover during summer seasons and maximization during winter seasons. Stepped solar still tech-
Solar still niques enhanced the productivity, and the addition of wicks in a stepped solar still was proposed. Insulating the
Stepped solar still sun tracking system positively affected the productivity because the insulation increased the heating capability
Climatic parameters and evaporative effects inside the still basin.
Design parameters © 2016 Elsevier B.V. All rights reserved.
Operation parameters

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
2. Productivity issue of solar still application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1. Environmental parameters effects on solar still desalination productivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1.1. Solar radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1.2. Ambient temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1.3. Relative humidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1.4. Wind velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
2.1.5. Cloud and dust cover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3. Design parameters effects on solar still productivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.1. Evaporation area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.2. Water depth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.3. Condensing cover angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
3.4. Thermal storage materials and additives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
3.5. Solar tracking and reflector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

⁎ Corresponding author at: Civil and Structural Engineering Department, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, 43600 Bangi, Selangor,
Malaysia.
E-mail address: fati@ukm.edu.my (S. Fatihah).

http://dx.doi.org/10.1016/j.desal.2016.02.025
0011-9164/© 2016 Elsevier B.V. All rights reserved.
M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 179

3.6. Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187


4. Operation parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
4.1. Water feeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
4.2. Position of stepped solar still . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
4.3. Maintenance of the solar still . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5. Solar still application in large scale plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
6. Cost analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7. Mathematical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.1. Convective heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.2. Radiative heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
7.3. Evaporative heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

of the water constitutes salt water and is not suitable for drinking [1].
Nomenclature Apart from human consumption, water is also needed for industrial
and agricultural purposes. With increases in population and industrial
i annual interest rate, % growth, the corresponding increase in fresh water consumption has in-
N lifetime of still evitably led to a worldwide imbalance between demand and supply of
As area of basin liner (m2) fresh water. This is made worse by the fact that most of the available
Ass area of solar still sides (m2) water is naturally impure or not drinkable and requires treatment [2].
L latent heat of vaporization (JKg−1) In addition to these water crises, energy consumption is another
M mass, kg problem. The increase in the consumption of fresh water, primarily
mw distillate output from still (Lm−2d−1) due to modern lifestyles and population growth, has led to increases
S salvage value of the system (i.e. $) in energy consumption. The high population number and energy con-
t time (day). sumption rates have inevitably resulted in further critical issues, such
Tw basin water temperature (°C) as environmental pollution and fuel shortages [3]. Fossil fuel resources
Tg cover temperature (°C) are fixed, non-renewable resources and their use is responsible for glob-
Pw partial pressure at basin water temperature (N/m2) al warming. Hence, huge shortages of fossil fuel, crude oil and energy re-
Pg partial pressure at cover temperature (N/m2) sources have contributed to the urgency of replacing expensive energy
ϭ Stefan–Bolzman constant (5.67 × 10–8 W/m2K4) sources with renewable energies.
h1g total heat transfer coefficient from cover to atmosphere For a remote area where fresh water is lacking, but with an abundant
(m2/C) amount of solar radiation, the use of solar energy is the preferred alter-
h1w total heat transfer coefficient from water to cover native energy source. Therefore, solar energy is the most suitable form
(W/m2C) of energy that can be tapped to provide cheap potable water using
hrw radiative heat transfer coefficient from water to decentralized methods. This is a fundamental solution because most re-
cover(W/m2C) mote and rural areas, especially those in arid regions, have abundant
hew evaporative heat transfer coefficient from water to sources of solar radiation that can be used to supply fresh water [2].
cover (W/m2C) Sampathkumar et al. [4] states that various advanced technology de-
I(t) solar radiation on the horizontal surface (W/m2) salination techniques using fossil fuels or electrical energy derived from
αw absorptivity of water and basin linear (−) fossil fuels, are used worldwide, such as reserve osmosis, electrodialysis,
αg solar flux absorbed by cover ozone, UV, vapor compression and active carbon filtration to obtain po-
αb solar flux absorbed by basin table water. Although these techniques are capable of desalinating
qrg radiative heat loss from cover to ambient (W/m2) water, they directly contribute to global warming and incur high costs
qeg convective heat loss from cover to ambient, (W/m2) [5]. Additionally, these technologies are commercially suited for large
qcb heat transfer from base to ambient by conduction cities and have high efficiencies and productivity rates when non-
(W/m2) natural energy sources are used [6]. Therefore, to provide fresh water
qcw convective heat transfer from water to cover (W/m2) to arid and semi-arid areas where sunshine is readily available, relative
qrw radiative heat transfer from water to cover (W/m2) to non-natural energy, these methods are not suitable.
qew evaporative heat transfer from water to cover (W/m2) Solar stills have been reviewed as alternative techniques for using
qs side heat loss to ambient by conduction (W/m2) solar radiation to supply potable water for low coastal areas [7]. Solar
hcw convective heat transfer coefficient from water to cover stills can provide potable water to these isolated communities at low
(W/m/C) costs compared with conventional energy sources, which are scarce
(MC)w heat capacity of water mass in basin, (J/m2/C) and expensive [2]. Zarasvand Asadi et al. [8] claimed that a solar still is
V wind speed in (m/s) an environmental friendly energy or green energy process that uses
η efficiency of solar still (%) the free natural energy of sunlight to purify contaminated water to pro-
duce clean water. The solar still process is able to replace fossil fuels by
using solar energy to gain the power needed for purification. A solar still
system has many advantages over other distillation systems. For in-
stance, a solar still system uses free and clean energy and is environ-
1. Introduction mental friendly. Furthermore, no moving parts making it easy to
construct and maintain. A solar still system does not require skilled
Fresh water is indispensable for the survival of humankind. Al- manpower can be manufactured locally and can be easily repaired
though water covers approximately 70% of the world, more than 90% using natural building materials. Finally, a solar still system requires
180 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

minimal investments, minimizes waste and is highly efficient in distill- Nafey et al. [21] reported that increased solar radiation leads to
ing even saline water (i.e., seawater) [9]. increased solar still productivity, i.e., higher solar radiation intensities
Solar desalination is a process for the evaporation and condensation result in increased water mass temperature and facilitates the evapora-
of water and is similar to a natural hydrologic cycle. A solar still uses a tion process. Therefore, the condensing cover material plays an impor-
green energy process that use free natural radiation to purify contami- tant role in the solar radiation absorbing process and thereby affects
nated water into distilled water using the energy transfer process the productivity of the system [22].
shown in Fig. 1. Energy transfer process [6]. Energy transfer process El-Sebaii [23] reported that solar still productivity increased with
stored [6]. In a solar still, the system is filled with impure water is heated wind velocity. Therefore, the solar still evaporation rate is dependent
by sunlight penetrating a condensing cover and heats the water mass up on the water mass temperature in the basin, the temperature of the
to the water evaporation point. The resultant water vapor is separated glass cover and on the difference between the two. This finding supports
from contaminants and impurities and subsequently cooled and con- the claim by [24], who reported that solar still productivity is influenced
densed on the inclined condensing cover. Water droplets flow toward by solar radiation intensity. Solar stills operate based on the heating,
a trough collection channel and are stored [6,10]. Therefore, the incident evaporation and condensing of water to produce fresh water. This indi-
solar radiation is considered the main environmental parameter that af- cates that decreased solar radiation intensity would lower the system
fects the productivity rate of a solar still system [11]. productivity.

2. Productivity issue of solar still application 2.1.2. Ambient temperature


A number of studies have been carried out to determine the effects
The main obstacle to using solar still desalination techniques is a low of climatic parameters on the performance of solar stills. As shown in
productivity rate. Solar stills have low productivity rates compared with Fig. 3, these parameters have direct effects on steeped solar still produc-
other desalination systems, such as other thermal processes or mem- tivity [20,21,25]. For example, an increase in temperature will increase
brane processes. A low productivity refers to three primary parameters the productivity of a solar still [16,18,26–28]. Xiao et al. [6] claimed
(i.e., environmental, design, and operation parameters). Despite being a that the production rate of a solar still increased with the temperature
suitable solution for a lack of a fresh water supply, its low productivity difference between the basin and the condensate cover.
impedes the application of a solar still. A number of research efforts
[12] to improve the production efficiency rate of a solar still (which 2.1.3. Relative humidity
ranges between 2.5–5 l/m2/d) have been carried out by varying a num- Atmospheric humidity cannot be controlled by humans because it is
ber of parameters affecting its efficiency rate [9]. Hence, this review pri- a meteorological parameter [1]. Koffi et al. [29] asserted that 40% up to
marily focuses on different design parameters that affect the rate of 65% variations in humidity have been investigated in previous studies.
solar still productivity. An average humidity of 65% during rainy seasons and humidity rates
of 40% to 55% during dry seasons have been examined as illustrated in
Fig. 4. This has also been examined by other studies [30,31], which
2.1. Environmental parameters effects on solar still desalination have shown that increases in relative humidity lead to increases in sys-
productivity tem yield.

The performance of a solar still is expressed as the quantity of water 2.1.4. Wind velocity
distilled in the basin area in a day (l/m2/d) [13] (i.e., a higher distilled The temperature of the upper surface of the cover glass lid is another
water quantity reflects a higher productivity and vice versa). The pro- important factor that affects the yield of a solar still. The effect of wind
ductivity of a solar still is influenced and affected by environmental speed significantly affects solar still productivity. As the cover tempera-
conditions [1], such as solar intensity, ambient temperature, relative hu- ture decreases, the productivity increases. The temperature difference
midity, wind velocity and cloud and dust cover, which are not controlla- between the glass and water increases as the cover temperature de-
ble by humans [12,14,15]. creases, which consequently improves the natural circulation of air
mass within the still. Moreover, the convective heat transfer from the
2.1.1. Solar radiation cover to the atmosphere increases when both evaporative and convec-
Many studies have found that increased solar radiation results in in- tive heat transfer between the basin water increase to compensate for
creased solar still yields due to increased differences between the tem- high wind speeds [32,33]. However, blowing wind increases solar still
perature of the water to be treated and that of the glass lid [4,16–19]. productivity by cooling down the temperature of the glass cover and
Fig. 2 shows that the performance of a solar still is directly affected by consequently increases the condensation rate [34,35]. Hence, any
solar radiation [5], particularly at 13:50, when a maximum production wind obstacles around a solar still cover has a negative effect on the pro-
of 0.5 l/m2/h is observed; at 12:00, a maximum isolation is observed. ductivity [15].
This delay is due to a time lag in the system [20]. The outcomes from According to Muftah et al. [16], when a solar still is under forced con-
previous studies have also shown that the most influential meteorolog- vection conditions when the airflow is parallel to the inclined surface of
ical parameter affecting the productivity of a solar still is the intensity of the cover at multiple wind speeds and at a 10° angle of inclination, the
the incident solar radiation [4]. variation in the output against the water temperature increases as the

Fig. 1. Energy transfer process [6].


M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 181

Fig. 2. The variation of solar intensity and still output [20].

temperature gradient between the water temperature and cover tem-


perature increases. The rate of evaporation increases with the increasing
wind velocity. El-Sebaii [23] studied the effects of wind velocity on the
output of active and multi-effect passive stills and concluded that the Fig. 4. Effect of humidity on solar still productivity water [29].
yield increases with the wind speed. El-Sebaii [36] analyzed the influ-
ence of wind speed on the performance for multiple masses of water
in basins, and he concluded that the performance improves with in- concluded that solar radiation transmission is correlated with dust ac-
creasing wind speeds. However, a conflicting report pertaining to the ef- cumulation. Furthermore, the study found that increases in dust accu-
fects of wind speed on solar still productivity also shows that when the mulation decreases solar transmission and that the tilt angle has a
wind speed increases from 1 to 9 m/s, the total yield of the system de- strong correlation with dust deposition. In another study [16], dust on
creased by 13% [21]. However, increases in solar still productivity has the condensing cover was observed to affect solar radiation transmis-
been reported to be generally attributed with increased wind speed ve- sion by reducing the transmission by an average of 1% at a tilt angle of
locities [23,32,36,37]. 30°. In another study, the glass transmittance drops by 10% during the
El-Sebaii [38] studied the effects of wind speed on solar still yield. summer and 6% during the winter [43].
According to this study, the daily productivity of a solar still increases
with increasing wind speed up to a certain velocity. The effects of 3. Design parameters effects on solar still productivity
wind in the summer time and for larger water masses were more effec-
tive than at other times. Based on the results of previous studies, the The design parameters are the most important parameters that af-
maximum daily productivity of a solar still was obtained at a minimum fect solar still productivity and can be easily controlled, developed and
heat capacity of the basin water. Furthermore, the productivity rate of a designed to improve productivity. These parameters include evapora-
solar still decreases as the water mass increases. Aburideh et al. [39] re- tion area, water depth, cover condensing angle, thermal storage mate-
ported that distilled water was influenced by the presence of wind and rials, additives, solar tracking, reflectors and insulation. Quite a few
the climatic change. These effects decreased the amount of diffused studies have been conducted that confirm the significance of these pa-
solar energy received by brackish water, resulting in 4 l/m2/day yields. rameters [44]. Many studies have examined the productivity of solar
The effect of wind velocity on solar still output is shown in Fig. 5. stills to develop and modify methods to improve the limited yield effi-
ciency. These studies have explored the optimization of different design
2.1.5. Cloud and dust cover parameters to study their effects on solar still productive rates, which
A 70% reduction in solar radiation transmittance annually occurs can be controlled, A number of studies are still in progress [45].
when the condensing cover is not maintained not cleaned [40]. The
solar still productivity is influenced by the cloud cover because it direct- 3.1. Evaporation area
ly affects the absorbed amount of solar radiation [41].
Hegazy [42] studied the effect of dust accumulation on the condens- Xiao et al. [6] reported that the productivity of a solar still or the rate
ing cover on solar transmittance with differing tilt angles. The study of evaporation in a solar still increases with the increase of evaporation

Fig. 3. The effect of ambient temperature on a basin solar still [20].


182 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

Fig. 5. The effect of wind velocity on solar still output [20].

area, i.e., solar still productivity is proportional to the solar still surface compensation of both fins and sponges were used are used and com-
area [11], as shown in Fig. 6. pared with conventional stills as shown in Fig. 9. According to [51],
In a study by Abu-Hijleh & Rababa'h [46], when sponge cups of a the productivity of a tilted-wick type solar still increased by 20–50%
conventional solar still basin were used to maximize the evaporation compared with basin types, i.e., the average tilted-wick still had a higher
area, the results showed that the efficiency level increased by 273%. It efficiency than the basin type, which is especially attractive for in-
is difficult to maintain a minimum depth in a conventional solar still creased distillation performance.
basin because the basin area is large and contains a number of fading
dry spots. A stepped solar still was used to increase the evaporation 3.2. Water depth
area to decrease the water mass by utilizing all trays. Currently, a
stepped solar still technique is used to increase the evaporation area Most studies in the literature concentrate on the effects of water
of basin water for domestic and industrial waste water treatment [8]. depth on solar still productivity rate. Water depth is a design parameter
Table 1 compares variations in stepped solar still productivities with that has gained significant attention and has become an important pa-
those of a non-stepped solar still. (See Table 2.) rameter affecting the production efficiencies of solar stills [52]. These
The evaporation area can be maximized by using joining wicks, jute studies analyzed and investigated different water depths in solar still
cloth, sponges, rubbers and other similar items [47,48]. By integrating basins and found that solar still productivity decreases as water depth
fins, sponges and wicks types in the stepped solar still to increase the increases. In brief, the solar still evaporation rate is inversely propor-
evaporation area, the total productivity was increased by 80%. tional to the water depth [13,53,54].
Velmurugan et al. [35] used fins, sponges, fins and sponges and fins, Many studies have also investigated the effects of different water
sponges and pebbles in stepped solar stills to increase the evaporation depths on solar still productivity rate. Decreased water depths result
area. Increased water temperatures and increased evaporation rates in increased solar still productivity rates [55–57]. Investigations of
were reported. These results show that the productivity increased com- deep basins have revealed that the productivity of a solar still increases
pared with conventional stepped solar stills. Pebbles provided a high with decreasing water depth during the daytime; however, the opposite
volumetric capacity, retained a greater amount of thermal energy and is observed during the nighttime. Furthermore, as water depth increases
increased the water temperature, as shown in Fig. 7. Wicks on the ver- in the solar still basin, the water volume and the water volumetric heat
tical sides of the steps were used by [49], as shown in Fig. 8. The use of capacity increase. These increases in volumetric heat capacity reduce
these wicks resulted in an increased brine evaporation area due to the the water temperature for the same solar radiation input. Therefore, a
absorption of more radiation levels and an increased basin water tem- reduced water depth results in an increased water temperature evapo-
perature, which in turn, yielded 3–5% higher evaporation rates. Stepped ration rate for a constant energy input.
solar still trays at two different depths were designed and investigated In the absence of sunshine, the energy stored in water within the
by [50]. Small fins and sponges were used in the trays to increase the basin is released, and desalinated water production is uniform and con-
evaporation area. These modifications were reflected in increased pro- tinuous even at night. At a minimum water depth in a shallow basin, the
ductivity and the results showed that compared with a conventional heat capacity of the water mass is lower, and the water temperature is
still, there was a 76% increase when the fins were used, a 60.39% in- higher. This results in increased evaporation rate and solar still produc-
crease when the sponges were used, and a 96% increase when a tivity. Any changes to the solar radiation intensity will affect the water
temperature and the solar still productivity. The opposite is true at
night because nocturnal production is lower for a shallow basin still be-
cause little to no heat is stored in a system with a shallow basin with low
water mass [1].
Ahsan et al. [58] developed a low cost solar still for rural and coastal
areas for the desalination of saline water. They evaluated water produc-
tivity by varying water depths (i.e., 1.5, 2.5 and 5 cm) and concluded
that water productivity was inversely proportional to water depth. In
another study, the output of an inverted double basin solar still was ob-
served to increase as the water depth decreased in the basin [59]. In a
study of varying water depth (i.e., 5, 10, and 15 cm), Murugavel et al.
[1] observed higher outputs when the water depths were at 10 cm
Fig. 6. The effect of the variation of evaporation area on distillation yield [97]. and 15 cm for passive and active solar stills, respectively. Similarly, the
M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 183

Table 1
Comparison of the variation of stepped solar still productivities with non-stepped solar still during the daytime of testing per unit area.

Sl. Types of solar still Productivity Cover Daily solar Ambient Wind Remarks Site location Reference
no. rate angle radiation Temp. speed
(l/m2/d) (°) (W/m2·d) (°C) (m/s)

1. Conventional Single-slope solar still 5.3 35 – – – Evaporation rate decrease with North Cyprus [99]
solar still increase of water depth.
Single basin solar still 4.5 35 – – – Using violet dye increase of Amman, [100]
approximately 29% if evaporation Jordan
rate.
Basin type solar still 6.7 35 – – – Evaporation rate decrease with Amman, [72]
increase of water depth. Jordan
Single slope solar still 3.54 32 1200 33 2–5 Evaporation rate decrease with Amman, [69]
increase of water depth. Jordan
Double-basin glass solar 2.9 – – – – Double-basin solar still is much India [28]
still better than a single-slope solar
still.
Single slope solar still 4.5 15 – – – Increase by 35% evaporation rate India [18]
coupled with flat plat comparing with passive solar still.
collector
Pyramid shaped solar still 4.25 – – – – – Egypt [101]
Single sloped solar still 4.2 16 – 35 4 Evaporation rate decrease with Suez gulf [21]
increase of water depth. region
Active single basin solar 4 – – – – Productivity increased by increase Jeddah, Saudi [102]
still with sand as a storage of sand mass as its works as heat Arabia
material storage material.
Conventional solar still 3.63 30 20–1150 28–35 0.1–5.5 – Egypt [14]
2. Stepped solar Stepped solar still 5.84 30 20–1150 28–35 0.1–5.5 With 5 steppes. Egypt [14]
still Stepped solar still with sun 6.3 30 20–1150 28–35 0.1–5.5 Internal reflector increase Egypt [14]
tracking system productivity by 75% comparing
with conventional still.
6.5 Top reflector increase productivity
by 91% comparing with
conventional still.
6.8 Bottom external reflector increase
productivity by 98% comparing
with conventional still.
7.4 Internal and top external reflector
increase productivity by 108%
comparing with conventional still.
7.24 Internal and bottom external
reflector increase productivity by
113% comparing with
conventional still.
8.1 Internal and top and bottom
external reflector increase
productivity by 125% comparing
with conventional still.
Stepped solar stills with 5.8 – 600 – 1 With addition of pebbles India [48]
mini solar pond 5.9 With addition of baffle plate
6.1 With addition of fins and sponges.
Stepped solar stills 5.84 30 100–1000 28–33 – Without internal reflector. Egypt [103]
6.3 With internal reflector.
Stepped solar still with 5.2 30 100–9000 28–33 – With cotton absorber. Egypt [70]
continuous water 5 With black absorber.
circulation
Stepped solar still coupled 6.3 31 30–1100 27–35 0.1–3.5 85% higher than conventional still. Egypt [62]
with a solar air-heater
Stepped solar still with 5.5 – – – – – Egypt [104]
cooling film
Stepped solar still with – 30 789 23–35 0–2.6 The system was extremely Malaysia [8]
triangular cross section successful in treating deferent
trays wastewater samples.
Stepped solar still with 4.6 – – – – With 57% efficiency than Jeddah, Saudi [105]
paraffin wax as storage conventional still. Arabia
material
Cascade solar still 6.7 – – – – – Iran [106]
Stepped basin solar still 5.63 – – – – Without wicks 57% efficiency than Egypt [49]
conventional still
5.19 With wick 50% efficiency than
conventional still
Stepped solar still 1.27 – – – – With fin India [35]
1.37 With fin and pebble
1.4 With fin and sponge
1.65 With fin and pebble and sponge
3. Wicks types Single slope solar 4 – – – – With 20% efficiency increase than India [107]

(continued on next page)


184 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

Table 1 (continued)

Sl. Types of solar still Productivity Cover Daily solar Ambient Wind Remarks Site location Reference
no. rate angle radiation Temp. speed
(l/m2/d) (°) (W/m2·d) (°C) (m/s)

still with jute cloth conventional still


Wick-basin type – – – – – Wick-basin type still productivity Iraq [108]
is 85% more than the basin- type
and 43% more than the wick-type
still
Solar still with a concave 4.1 – – – – 45% efficiency Egypt [51]
wick evaporation surface
Tilted wick still with – – – – – Productivity daily mount ranges Japan [109]
inclined flat plat external from 15% to 27
reflector
Single basin wick type solar 2.3 – – – – Summer India [110]
3.4 Winter
wick-type solar still – – – – – Still efficiency has decreased from Iraq [111]
38% to 20% with increase salt
concentration
Floating wick type solar – – – – – Using floating perforated plate in Egypt [112]
still solar still at a brine depth of 6 cm
enhanced the productivity by 40%
Single basin solar still with – – – – – Using different types of wicks India [47]
fin increase the overall efficiency
Single basin double slope – – – – – Black cotton cloth was effective India [113]

outputs were higher when the water depths in the solar stills were Stepped solar stills have also been investigated by many researchers.
10 cm and 15 cm during the off shine hours. Their findings indicate that a reduction in seawater or brine depth in
In another study, the productivity of a solar still with varying water solar still basins increased productivity due to increased solar still tem-
depths (i.e., 1, 4, 6, 8 and 10 cm) was observed to be influenced by water peratures. Variations in the depths and widths of basin trays have also
depth and decreased to 48% as the water depth increased, i.e., increased been found to directly affect stepped solar still yields [62]. In a stepped
water depths reduced solar still productivity. However, at nighttime, solar still with a 0.5 m width, 1 m length, high side walls at heights of
the production output increased with increased water depth and con- 450 mm, low side walls at heights of 160 mm, and 5 steps (each
siderably contributed to the total daily output [44,60]. 0.1 m × 1 m), the overall efficiency was 30.4% higher than a convention-
According to previous studies, water depth is inversely proportional al still under the same environmental conditions. Kabeel et al. [49], re-
to solar still productivity. In one study, as water depth increased from 2 ported a stepped solar still with a higher productivity of 57.3% with
to 7 cm, a 14% reduction was observed in the output [21]. In another 5 mm × 120 mm sized trays compared with a conventional solar still.
study, an optimum depth of 4 cm was determined for an optimized a stepped solar still designed for different waste water treatment appli-
solar still production [61]. Moreover, in most studies, the recommended cations had 11 triangular trays to reduce the footprint area and increase
basin water depth ranged from 2 to 6 cm [20], as shown in Fig. 10. the effective evaporation area [8] as shown in Fig. 11.
One study observed that the productivity of a stepped solar still was
Table 2 57% higher than that of a conventional solar still [14]. A 53.3% increase in
Chronology of solar sill desalination large scale plant. efficiency was observed by a study examining two different depths of
Time frame Description References finned trays in the basin of a solar air heater [34,35]. In general, a
stepped solar still provides a higher productivity than a conventional
4th A method to evaporate impure water and then [86,90,91]
condense it to obtain fresh water was discussed by solar still [14,62,63].
Aristotle.
15th century Earliest documented works by Arab alchemist. The [87,89]
Arab alchemist used a concave Damascus mirror to
focus solar radiation onto glass vessel to obtain fresh
water.
1872 First large solar distillation plant was built in Las [90]
Salinas, Chile. With total capacity 22.70 m3/d and
had a total surface area of 4450 m2.
1940 200,000 inflatable plastic stills were mad to be kept [90]
in crafts for US Navy during Second World War.
1953 Small solar still was built with a tray area of 0.18 m2 [91]
using Plexiglass. It was found that the plastic cover
was not suitable due to problems associated with
drop condensation of distilled water on the inner
surface of the plastic cover.
1962 Bay type prototype was developed. The system was [91]
covered with glass and lined with a black
polyethylene Sheet. The tray area of the tested units
varied between 1,5 and 3 m2. From the prototype
design, a solar distillation plant was built in the
Australian desert to supply fresh water.
1970 Solar distillation plant was built on four Greek Island. [86]
The design was implemented by Technical University
of Athens. The capacity of the asymmetric solar still
was in the range from 2044 to 8640 m3/day. The
largest distillation plant was built in Patmos.
Fig. 7. Effects of various modifications in solar still on evaporation heat transfer [35].
M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 185

water droplets into the condensate plate, maximizes the absorption of


solar radiation and insulates the exterior surface to minimize heat losses
[44].
Solar stills with various cover angles of 10°, 15°, 35°, 45° and 50°
have been tested in a number of studies [15,44,65,67,68]. Studies that
examined angles of 33.3° [60], and 35° observed maximum annual
yields in May [44]. Badran [69] identified an optimum angle at 32°,
and another study stated that the optimum inclination angle was 23°.
Stepped solar stills with 30° angles [14,49,70] showed high productivity.
For increasing cover angles [20], solar still productivity increased
during the winter season. However, during the summer season the
solar still productivity decreased with increasing cover angles, as
shown in Fig. 12. A higher yield was obtained at a 45° cover slope during
the winter season [71,72].

3.4. Thermal storage materials and additives


Fig. 8. Tray on horizontal side and wicks on vertical side of the step [49].
Thermal capacity plays a major role in solar still productivity. The
production efficiency increases with the increase in thermal capacity
3.3. Condensing cover angle of a solar still system for various solar radiation intensities. The efficien-
cy also increases when the latent heat of condensation is recovered [73].
A condensing cover material is a cover that can be made from plastic As the air temperature inside solar still increases, the volume of air
or glass but must have a high transmittance for solar radiation. Brackish above the water surface is circulated to the cover by buoyancy forces
water contained in the basin must absorb and reflect a minimum caused by the variation in heated air density due to temperature differ-
amount of the solar spectrum. However, the condensing cover must ences between the condensing cover and the water surface.
also provide some degree of insulation to prevent excessive thermal ra- According to a prior study [71], the thermal capacity of the condens-
diation heat transfer from the basin liner to the external environment. ing cover maintains a higher temperature inside the solar still. The con-
The condensing cover is supported at the top of the solar still and directs densing cover must allow solar radiation transmittance and at the same
solar radiation to the basin to evaporate water. Modifications to the con- time, dissipate as much heat as possible to the atmosphere to ensure
densing cover to maximize radiation transmittance would increase the condensation. Increases in the thermal capacity will increase water
productivity of the solar still [64,65]. The top of the assembly is covered temperatures inside the basin to enhance the evaporative capacity [16].
with a condensing cover made of a material that allows solar radiation Tiwari et al. [74] claimed that the bottom frame of a solar still can be
to pass through but prevents thermal radiation (emanating from the in- fabricated from plastic, metal sheets, copper, aluminum, steel or some
terior, including the water mass) from escaping [66]. material with a high thermal capacity. The thermal conductivity of steel
Nafey et al. [21] observed that the slope of the condensing cover is relatively low compared with aluminum and copper (k = 48 W/mK,
(i.e., the cover angle) has a direct effect on the productivity of solar stills. k = 200 W/mK, and k = 390 W/mK, respectively). However, steel is
The slope must be at such an angle to ensure sufficient condensing water cheaper than aluminum and copper (i.e., at half the cost of aluminum
runs smoothly to the collector plate without reflecting too much of the and copper) [1]. According to a prior study [75], a modified solar still
solar energy and without forming large droplets that fall into the basin in- made of copper ensures higher evaporation rate and efficiency. In anoth-
stead of the condensate channel [17,65]. A suitable cover angle provides a er study, the efficiency of a copper sheet-built solar still considerably
cool surface for water vapor condensation, facilitates an easy flow of increased [76].

Fig. 9. Schematic diagram of the stepped solar still [50].


186 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

Fig. 10. The effect of water depth on the solar still productivity [20].

Concrete cement, galvanized iron sheets or fiber reinforced plastic a stepped solar still and a single basin solar still were connected in series
(FRP) can be used as a solar still basin bottom [18,77]. In other studies to enhance the productivity of the solar stills. To increase the productiv-
[68,77], the bottom of the basin can be made of 4 mm-thick aluminum ity, pebbles, baffle plates, fins and sponges were used in both solar stills.
as an absorbent. Painting the basin using a black dye can further in- An 80% productivity increase was obtained when fins and sponges were
crease the absorbent capacity of the base [60]. Black polycarbonate used in both solar stills. The second set of experiments utilized a mini
can be used as a material for solar still basin bases [78]. solar pond, a stepped solar still and a wick type solar still connected in
These additive materials act as absorbent materials for solar radia- series. Pebbles, baffle plates, fins and sponges were used in the stepped
tion and as non-adsorbent materials for water and thereby increase solar still for further productivity augmentation. The productivity dur-
the basin temperature to enhance evaporation. Different materials ing the day- and nighttime indicated a maximum productivity of 78%
have been investigated in different studies, and the findings have indi- when the fins and sponges were used in the stepped solar still. The peb-
cated that using finned basin plates and sponges in the trays tended to bles released accumulated thermal energy during the nighttime. As a re-
increase the productivity of stepped solar stills. Compared with an ordi- sult, a higher nighttime productivity was obtained. Additionally, an
nary stepped solar still, as much as 76% and 60.39% increases in produc- industrial effluent was used as the feed.
tivity were observed when fins and sponges were added, respectively; a As shown in Fig. 15, El-Agouz [70] investigated the performance of a
96% increase was observed when both materials were used [34]. modified stepped solar still using a cotton absorber and a storage tank.
In another study, the productivity increases by 53.3%, 68% and 65% For sea and saline water, 43% and 48% productivity increases were re-
once fins, sponges and pebbles were added to a stepped solar still. A spectfully obtained for the modified stepped solar still compared with
98% productivity increase was observed when sponges and fine pebbles a conventional still, and 20% increases in total efficiency were observed.
were combined [35]. Compared with a conventional still, a 53% increase When using a black dye absorber, the daily efficiency for the modified
in a stepped solar still was observed when an aluminum filling was used stepped and conventional solar stills were approximately 61% and 42%
as a simple solar energy storage system under the absorber plate [62]. for sea water and 55% and 37% for saline water, respectively. When
Two different types of experiments were carried out in a study [48], using a cotton absorber, the daily efficiencies of the modified stepped
as shown in Figs. 13 and 14. In the first set of experiments, a solar pond, and conventional solar stills were approximately 61% and 70% for sea

Fig. 11. Cross-section view of the stepped solar still [8].


M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 187

on the basin bottom. Fixed and sun-tracking stills were compared. An


improved single slope solar still was found to increase the solar still pro-
ductivity rate. When conventional solar stills were modified by
installing reflecting mirrors on all interior sides, replacing the flat
basin with a stepped basin, or coupled with a sun-tracking system, the
system productivity improved by 30%, 180% and 380%, respectively
[65]. Another study investigated the effects of reflectors on the basin
base. The efficiency of a solar still were enhanced by 48% when external
and internal reflectors were added [80]. In another study [81], as shown
in Fig. 16, a stepped solar still was modified with internal and external
reflectors, an external condenser, 100 mm-width trays and compressed
steps. The total daily productivity increased by 165% compared with a
conventional solar still.

Fig. 12. The effect of the cover slope angle on the single solar still output for winter and
summer seasons (Jan. & Feb., May & June) [20]. 3.6. Insulation

Insulation is another important physical parameter. The operation of


a solar still depends on absorbing solar radiation to evaporate water and
water and 40% and 48% for saline water, respectively. The maximum ef-
the transfer of thermal energy. Excess thermal energy is stored within
ficiency of the modified stepped still occurred at a feed water flow rate
the water body of the solar still basin. Proper insulation is necessary to
of 1 LPM for sea water and 3 LPM in salt water. The day and night effi-
prevent the release of thermal energy as waste heat from the bottom
ciency increases by 5 and 3.5%, when the still is modified with a storage
and sides of the basin [82]. Bottom insulation is an important design pa-
tank for salt and sea water. In a stepped solar still, the daily productivity
rameter. Thick insulation is often used to reduce heat losses from the
increase from 3% to 5% when wicks on the vertical sides [49] were
still to the environment [4]. A 6 cm-thick insulation layer increased pro-
utilized.
ductivity by 80% compared with 3 cm and 10 cm-thick layers due to a
higher still temperature [79].
3.5. Solar tracking and reflector In view of these findings, an insulating material should be used to re-
duce heat losses from the bottom and the side walls of a solar still. The
A sun-tracking device was used to track the Sun. and to accordingly insulating material is an important parameter design because it mini-
adjust a solar still to enhance the solar still productivity [79]. External mizes heat losses from the basin bottom and side walls [18]. Previous
and internal refractors were used to focus solar radiation absorption studies have revealed that the use of various types of insulating

Fig. 13. Combination of stepped solar still and single basin solar still with pond [48].
188 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

Fig. 14. Combination of stepped solar still and wick type solar still with pond [48].

materials increased solar still productivities [83]. Stepped solar stills 126%, and 130%, respectively. These studies show the influence of insu-
were made of wooden boxes and insulated with sawdust, as shown in lation on solar still productivity.
Fig. 17 [14,15,35,48,49,62]. A stepped solar still was insulated with
glass wool [8]. In another study [84], five solar stills were investigated,
one without insulation and four with different insulation materials 4. Operation parameters
(i.e., an air gap, plywood, hay, and glass wool). The study reported
that the productivities of the solar stills with insulation materials were Operation parameters have direct effects on solar still productivity.
greater than the productivity of non-insulated solar still by 74%, 82%, These parameters include dust accumulation and system maintenance.

Fig. 15. Schematic diagram of the experimental setup [70].


M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 189

[60]. In another study [48], a solar still was oriented toward the south
to absorb and receive maximum solar radiation throughout the year.
This finding was supported by other studies [14,35,49].

4.3. Maintenance of the solar still

Because a solar still is generally fixed and has no moving parts, sys-
tem maintenance is simplified. Solar still maintenance essentially en-
tails cleaning the basin to prevent particulate sedimentation. When a
solar still is operated in the dynamic mode, dissolved solids in the
water continuously flow through the solar still basin and flow into the
overflow channel. None of the dissolved solids settle in the trays; there-
fore, the solar still rarely requires cleaning [85].
However, when a solar still is operated in the static mode, solids and
impurities settle and deposit on each tray and are exposed to the sun
rays. The higher temperature and increased residence times ultimately
produces an undesirable, colored hard deposit that is difficult to re-
move. Non-toxic, diluted acid solutions, such as citric acid or oxalic
Fig. 16. The stepped solar still with reflectors and external condenser [81]. acid, are typically used to clean the solar still [85].
As previously discussed, the accumulation of dust on a solar still may
affect the performance of the system. The collection of dust on the con-
The most influential parameters include the means of water feeding and densing cover creates shadows on the basin and prevents effective solar
the position of the solar still. still operation. Hence, solar still tops should be regularly maintained to
remove dust and any other contaminants that may decrease solar radi-
4.1. Water feeding ation intensity. Furthermore, algae and scales may accumulate on the
basin surface. Basins should be flushed daily to remove deposited scales
In general, there are two modes of water feeding: static and dynamic and particulate organic matter. Flushing may also remove salt precipi-
modes. In the static mode, water is fed to a solar still once a day to fill the tates on the black surfaces of the basin.
basin, and in the dynamic mode, the solar still is continuously or inter- After a lengthy period of operation, small droplet masses may form
mittently fed from a reservoir drip [14,49,60,62,78]. To reduce the on the underside of the glass cover of a solar still. These droplets accu-
water depth, a spray system has been used [63]. The 77.35% increased mulate on the glass cover, appear white in color, and reflect solar and
productivity for any operating condition was significantly greater than near infrared radiation. As a result, these residual spots decrease inci-
those for stills reported by other investigators. dent solar radiation intensity and decrease solar still productivity.
These droplets are generated from the build-up of a very thin, molecular
4.2. Position of stepped solar still layer of particulates on the glass surface and prevent the efficient evap-
oration of water. A removable glass cover has been recommended for
A high efficiency solar still is also dependent on its fixed position rel- easier maintenance [78].
ative to incident solar radiation. The production rate of a solar still is max-
imized when it is optimally positioned with respect to incident solar 5. Solar still application in large scale plant
radiation to absorb the maximum radiation [78]. In the southern hemi-
sphere, a solar still should be oriented such that the input end faces Solar stills have a long history and have been used for over
south. In the northern hemisphere, a solar still should be oriented toward 2000 years. “Desalination” or “desalinization” refers to water treatment
the north. A solar still should be aligned in the north–south direction, as processes that remove salts from saline water. The practice of desalina-
determined when the solar still shadow is aligned with the solar still itself tion has often primarily been used to obtain salt rather than fresh water
at noon. This is the orientation that is used when the operation of solar [86–88].
still is in a fixed, stationary position [78]. If the solar still is in the southern Aristotle descried methods of evaporating impure water and saline
hemisphere, the long axis should be aligned in the east–west direction water and subsequently, condensing water vapor to obtain potable

Fig. 17. Schematic diagram of the experimental setup [14].


190 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

water in the fourth century B.C. According to a few studies [87,89], the due to a large footprint (approximately 250 square meters per cubic
earliest documented stills were attributed to Arab alchemists in 1551, meter of distilled water) [92]. According to one study [93], to maintain
who used polished Damascus mirrors for solar distillation. Scientists high solar still productivity, a minimal cost system must take into con-
and naturalists, including Della Porta (1589), Lavoisier (1862), and sideration design and optimal maintenance costs. The main parameters
Mauchot (1869), used stills over the following centuries. in a cost analysis of desalination units include the capital recovery factor
In 1872, a Swedish engineer, Carlos Wilson, constructed the first and (CRF), fixed annual cost (FAC), sinking fund factor (SSF), annual salvage
large conventional solar still in Las Salinas, Chile, to provide fresh water value (ASV), average annual productivity (M), annual cost (AC), annual
for mining communities from brackish, high salinity saltpeter mine ef- maintenance operational cost (AMC) and cost per litter (CPL). AMC is
fluent water (140 g/kg, or 140,000 ppm). The solar still operated for used to calculate maintenance costs for removing salt deposits and
more than 40 years and was abandoned only when a centralized fresh maintenance of system parts. The maintenance cost of solar stills is
water piping water system was established. During operation, the low. Therefore, the cost of producing distilled water is mainly depen-
solar still used woody material with blackened bottoms and was cov- dent on the initial investment and interest rate [94]. Finally, the annual
ered by a condensing glass cover with a 4450 m2 surface area. The still cost per liter (AC/L) can be expressed by Eqs. (1)–(6) [92]. Table 3 pro-
had a 22.7 m3/d production rate, as shown in Fig. 18 [87,89,90]. vides a comparison between different types of solar stills reported in the
After the First World War, several studies were carried out to devel- literature [92].
op solar still desalination devices, including tilted wicks, inclined trays, . .
AC
various cover designs and inflated stills [87], until the Second World ¼ AC ð1Þ
L M
War.
According to one study [91], only a few solar distillation devices AC ¼ FAC þ AMC–ASV ð2Þ
were in use prior to the Second World War. In 1953, the first small
solar still was built with a 0.18 m2 tray area at the University of Bari, FAC ¼ P  CRF ð3Þ
Italy. The still was made of a Plexiglas tray and a Plexiglas evaporation
cover. A glass cover was compared with the plastic cover; the plastic ið1 þ iÞn
cover had issues with condensing distilled water. In September, 1953, CRF ¼ ð4Þ
ð1 þ iÞn  1
another model presented at a trade fair exhibition at Bari was made
from a glass cover and a Plexiglas tray with a 0.25 m2 effective area. A ASV ¼ SFF  S ð5Þ
new solar distillation system made of a wood tray with a 2 m2 surface
area glass cover and cork insulation at the bottom was constructed. i
However, due to water leakage and bad insulation, a better model was SFF ¼  1: ð6Þ
ð1 þ iÞn
built. The tray was placed on perforated bricks that offered some ther-
mal insulation to prevent heat losses. The result showed that a still
could be integrated into a building, e.g., on flat roof areas of a building. 7. Mathematical model
The glass cover can be utilized and used as a collector for rainfall to in-
crease yearly water production. Solar still performance can be calculated based on productivity, effi-
Accordion to a study [90], an inflatable plastic still was designed for ciency, and internal heat and mass transfer coefficients. The perfor-
life crafts for the US Navy. This was the first recorded mass production of mance of a solar still is directly proportional to the internal heat
a large volume of clean water was achieved during the Second World transfer coefficient and distillate output. There are three relevant heat
War. Table 2 shows the chronology of the important discoveries and transfer coefficients: the convective heat transfer, radiative heat transfer
large scale uses in desalination. and evaporative heat transfer coefficients [95].

6. Cost analysis 7.1. Convective heat transfer

The capital and operating costs are influenced by feed water quality, According to a few studies [95,96], the convective heat transfer coef-
fresh water quality, unit size, and site location. The main economic ad- ficient is influenced by the buoyancy force due to density differences in
vantages of solar distillation include (a) a simple infrastructure and the heated air. The convective heat transfer between a water surface to
(b) easy installation, operation and maintenance. However, the main the condensing glass lid is calculated by Eq. (7) [95,96]:
economical limitation of solar distillation includes unit size limitations
 
qcw ¼ hcw T w  T g : ð7Þ

Where the convective heat transfer coefficient hcw can be calculated


by Eq. (8):
    
0:88 T w  T g þ P w  P g ðT w þ 273Þ
hcw ¼  13 ð8Þ
268:9x103  P w

Table 3
Comparison between different types of solar still [93].

Type M L/m2 CPL $/L/m2

Pyramid shape 1533 0.031


Single slope 1511 0.035
A weir type 1001 0.054
Transportable hemispherical 1026 0.18
Sun tracking 250 0.23
Fig. 18. The world-wide first solar distillation plant at Las Salinas, Chile [86,90].
M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 191

7.2. Radiative heat transfer energy balances can be written for the glass cover, basin water and for
the basin, as shown in Fig. 19.
According to a few studies [95,96], radiative heat transfer occurs The heat transfer mechanism begins when solar energy is incident
through solar energy. Radiative heat transfer from a water surface to on the basin water. As shown in Fig. 19, the heat balance equation on
the condensing cover is calculated using Eq. (9): basin water can be written with the following assumptions [95,96]:
  1. There is no vapor leakage in the solar still.
qrw ¼ hrw T w  T g ð9Þ
2. It is an air tight basin; hence, there is no heat loss.
3. The heat capacities of the cover, absorbing material and insulation
where the radiative heat transfer coefficient hrw can be calculated by
are negligible.
Eq. (10):
4. There is no temperature gradient across the basin water and glass
 2 cover of the solar still.
hrw ¼ εeffect σ½ðT w þ 273Þ2 þ T g þ 273 ð10Þ
5. The water level inside the basin is maintained at constant level.
6. Only film type condensation occurs in place of drop type condensation.
where
 1
1 1
ε effect ¼ þ 1 ð11Þ • Energy balance for glass cover:
εg εw
qrg þ qeg ¼ α g Iðt Þ þ ðqew þ qcw þ qrw Þ: ð16Þ
where

εg ¼ εw ¼ 0:9 ð12Þ • Energy balance for basin water:

Tw
α b Iðt Þ þ qw ¼ ðMC Þw þ qrw þ qew þ qcw : ð17Þ
7.3. Evaporative heat transfer dt

According to a few studies [95,96], the incident solar energy inside a • Energy balance for basin:
solar still evaporates water into steam. The evaporative heat transfer

can be calculated by: Ass
α b Iðt Þ ¼ qw þ ðqcb þ qs : ð18Þ
  As
q ew ¼ hew T w  T g ð13Þ
• Heat transfer coefficients:
where the evaporative heat transfer coefficient hew can be calculated
by: h1g ¼ 5:7 þ 3:8 V: ð19Þ
 
Pw  Pg
hew ¼ 16:27  103  hcw   : ð14Þ • Hourly yield of solar still is given by:
Tw  Tg
qew
The total heat transfer coefficient from water to the condensing mw ¼  3600: ð20Þ
L
cover can be calculated by:
• Efficiency of solar still is given by:
h1w ¼ hcw þ hrw þ hew ð15Þ
qew
A solar still has several components, and its performance can be pre- η¼ ð21Þ
I ðt Þ:
dicted by energy balance equations on these components. Separate

8. Conclusion

This study reviews the parameters that influence solar still produc-
tivity. The parameters include climatic, operation and design parame-
ters. The productivity of a solar still is directly correlated with the
amount of solar radiation, the surrounding temperature, relative hu-
midity, wind speed and the maintenance of the cover. These metrolog-
ical parameters cannot be controlled. The most essential controllable
parameters are design parameters. Based on this detailed review, an op-
timum design to achieve a superior system performance can be summa-
rized as: follows (a) a condensing cover angle, which is strongly
dependent on still location, time of operation, and condensing material.
However, it has been generally [14,49,70] suggested that angles of 30°
and 45° [20] and smooth and low thermal capacity glass are optimal;
(b) a still constructed of copper has a higher thermal conductivity
390 W/mK; (c) a still with a stainless steel frame; (d) a black basin
liner; (e) the use of glass wool or sawdust insulation; (f) a recommend-
ed basin water depth of 3 cm; and (g) a dynamic water feeding mode. A
stepped solar still also results in increased evaporation area and in-
creased productivity. Furthermore, a high productivity rate can be
Fig. 19. Energy flow in Single basin single slope solar still [95,98]. achieved by using a combination of additive materials (e.g., fin and
192 M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193

sponge) integrated in a stepped solar still. Internal and external (top [25] A.N. Khalifa, A. Al-jubouri, M.K. Abed, An experimental study on modified simple
solar stills, Energy Convers. Manag. 40 (1999) 1835–1847.
and bottom) reflectors can also increase the productivity of stepped [26] R. Tripathi, G.N. Tiwari, Effect of size and material of a semi-cylindrical condensing
solar stills. Insulation materials, such as saw dust can enhance the pro- cover on heat and mass transfer for distillation, Desalination 166 (2004) 231–241,
ductivity by preventing thermal heat losses. A spray system can also re- http://dx.doi.org/10.1016/j.desal.2004.06.078.
[27] O.O. Badran, H.A. Al-Tahaineh, The effect of coupling a flat-plate collector on the
sult in a high productivity rate. solar still productivity, Desalination 183 (2005) 137–142, http://dx.doi.org/10.
The production rate of a solar still is maximized when it is correctly 1016/j.desal.2005.02.046.
positioned with respect to the direction of solar radiation during opera- [28] T. Arunkumar, K. Vinothkumar, A. Ahsan, R. Jayaprakash, S. Kumar, Experimental
study on various solar still designs, ISRN Renew. Energy 2012 (2012) 1–10,
tion. The maintenance of a solar still system is highly recommended to http://dx.doi.org/10.5402/2012/569381.
ensure high operative efficiencies. [29] B.K. Koffi, D.K. Konnan, R.K. N'guessan, J.K. Saraka, A. Kouacou, A. Tanohm, et al.,
Modelling of solar still for production of pure water in the abidjan zones, res. j.
phys. 3 (2009) 5–13.
References [30] J. Lindblom, B. Nordell, Water production by underground condensation of humid
air, Desalination 189 (2006) 248–260, http://dx.doi.org/10.1016/j.desal.2005.08.
[1] A. Muthu Manokar, K. Kalidasa Murugavel, G. Esakkimuthu, Different parameters 002.
affecting the rate of evaporation and condensation on passive solar still — a review, [31] K. Kalidasa Murugavel, P. Anburaj, R. Samuel Hanson, T. Elango, Progresses in in-
Renew. Sust. Energ. Rev. 38 (2014) 309–322, http://dx.doi.org/10.1016/j.rser.2014. clined type solar stills, Renew. Sust. Energ. Rev. 20 (2013) 364–377, http://dx.
05.092. doi.org/10.1016/j.rser.2012.10.047.
[2] A.A. El-Sebaii, E. El-Bialy, Advanced designs of solar desalination systems: a review, [32] Y.H. Zurigat, M.K. Abu-Arabi, Modelling and performance analysis of a regenerative
Renew. Sust. Energ. Rev. 49 (2015) 1198–1212, http://dx.doi.org/10.1016/j.rser. solar desalination unit, Appl. Therm. Eng. 24 (2004) 1061–1072, http://dx.doi.org/
2015.04.161. 10.1016/j.applthermaleng.2003.11.010.
[3] M.T. Chaibi, An overview of solar desalination for domestic and agriculture water [33] R.M. Reddy, K.H. Reddy, Upward heat flow analysis in basin type solar still, J. Min.
needs in remote arid areas, Desalination 127 (2000) 119–133, http://dx.doi.org/ Metall. 45 (2009) 121–126.
10.1016/S0011-9164(99)00197-6. [34] V. Velmurugan, C.K. Deenadayalan, H. Vinod, K. Srithar, Desalination of effluent
[4] K. Sampathkumar, T.V. Arjunan, P. Pitchandi, P. Senthilkumar, Active solar using fin type solar still, Energy 33 (2008) 1719–1727, http://dx.doi.org/10.1016/
distillation—a detailed review, Renew. Sust. Energ. Rev. 14 (2010) 1503–1526, j.energy.2008.07.001.
http://dx.doi.org/10.1016/j.rser.2010.01.023. [35] V. Velmurugan, K.J. Naveen Kumar, T. Noorul Haq, K. Srithar, Performance analysis
[5] S. Yadav, K. Sudhakar, Different domestic designs of solar stills : a review, Renew. in stepped solar still for effluent desalination, Energy 34 (2009) 1179–1186, http://
Sust. Energ. Rev. 47 (2015) 718–731, http://dx.doi.org/10.1016/j.rser.2015.03.064. dx.doi.org/10.1016/j.energy.2009.04.029.
[6] G. Xiao, X. Wang, M. Ni, F. Wang, W. Zhu, Z. Luo, et al., A review on solar stills for [36] A.A. El-Sebaii, On effect of wind speed on passive solar still performance based on
brine desalination, Appl. Energy 103 (2013) 642–652, http://dx.doi.org/10.1016/j. inner/outer surface temperatures of the glass cover, Energy 36 (2011) 4943–4949,
apenergy.2012.10.029. http://dx.doi.org/10.1016/j.energy.2011.05.038.
[7] A. Kaushal, Varun, solar stills: a review, Renew. Sust. Energ. Rev. 14 (2010) [37] M.A.S. Malik, Van Vi Tran, A simplified mathematical model for predicting the noc-
446–453, http://dx.doi.org/10.1016/j.rser.2009.05.011. turnal output of a solar still, Sol. Energy 14 (1973) 371–385, http://dx.doi.org/10.
[8] R. Zarasvand Asadi, F. Suja, M.H. Ruslan, N.A. Jalil, The application of a solar still in 1016/0038-092X(73)90015–7.
domestic and industrial wastewater treatment, Sol. Energy 93 (2013) 63–71, [38] A.A. El-Sebaii, Effect of wind speed on some designs of solar stills, Energy Convers.
http://dx.doi.org/10.1016/j.solener.2013.03.024. Manag. 41 (2000) 523–538, http://dx.doi.org/10.1016/S0196-8904(99)00119-3.
[9] V. Velmurugan, K. Srithar, Performance analysis of solar stills based on various fac- [39] Hanane Aburideh, A. Deliou, B. Abbad, F. Alaoui, D. Tassalit, Z. Tigrine, An experimental
tors affecting the productivity — a review, Renew. Sust. Energ. Rev. 15 (2011) study of a solar still: application on the sea water desalination of Fouka, Procedia Eng.
1294–1304, http://dx.doi.org/10.1016/j.rser.2010.10.012. 33 (2012) 475–484, http://dx.doi.org/10.1016/j.proeng.2012.01.1227.
[10] P.K. Abdenacer, S. Nafila, Impact of temperature difference (water-solar collector) [40] A.M. El-Nashar, Seasonal effect of dust deposition on a field of evacuated tube col-
on solar-still global efficiency, Desalination 209 (2007) 298–305, http://dx.doi.org/ lectors on the performance of a solar desalination plant, Desalination 238 (2009)
10.1016/j.desal.2007.04.043. 66–81, http://dx.doi.org/10.1016/j.desal.2008.03.007.
[11] P. Prakash, V. Velmurugan, Parameters in fl uencing the productivity of solar stills — [41] E. Zamfir, C. Oancea, V. Badescu, Cloud cover i n fluence on long-term perfor-
a review, Renew. Sust. Energ. Rev. 49 (2015) 585–609, http://dx.doi.org/10.1016/j. mances of flat plate solar collectors, Renew. Energy 4 (1994) 339–347.
rser.2015.04.136. [42] A.A. Hegazy, Effect of dust accumulation on solar transmittance through glass
[12] A.E. Kabeel, Z.M. Omara, M.M. Younes, Techniques used to improve the perfor- covers of plate-type Collectors, Renew. Energy 22 (2001) 525–540.
mance of the stepped solar still—a review, Renew. Sust. Energ. Rev. 46 (2015) [43] A.M. El-Nashar, The effect of dust accumulation on the performance of evacuated
178–188, http://dx.doi.org/10.1016/j.rser.2015.02.053. tube collectors, Sol. Energy 53 (1994) 105–115, http://dx.doi.org/10.1016/S0038-
[13] K. Kalidasa Murugavel, K. Srithar, Performance study on basin type double slope 092X(94)90610-6.
solar still with different wick materials and minimum mass of water, Renew. Ener- [44] A.J.N. Khalifa, A.M. Hamood, On the verification of the effect of water depth on the
gy 36 (2011) 612–620, http://dx.doi.org/10.1016/j.renene.2010.08.009. performance of basin type solar stills, Sol. Energy 83 (2009) 1312–1321, http://dx.
[14] Z.M. Omara, A.E. Kabeel, M.M. Younes, Enhancing the stepped solar still perfor- doi.org/10.1016/j.solener.2009.04.006.
mance using internal and external reflectors, Energy Convers. Manag. 78 (2014) [45] A.J.N. Khalifa, A.M. Hamood, Effect of insulation thickness on the productivity of
876–881, http://dx.doi.org/10.1016/j.enconman.2013.07.092. basin type solar stills: an experimental verification under local climate, Energy
[15] F.F. Tabrizi, A.Z. Sharak, Experimental study of an integrated basin solar still with a Convers. Manag. 50 (2009) 2457–2461, http://dx.doi.org/10.1016/j.enconman.
sandy heat reservoir, Desalination 253 (2010) 195–199, http://dx.doi.org/10.1016/ 2009.06.007.
j.desal.2009.10.003. [46] B. Abu-Hijleh, H.M. Rababa'h, Experimental study of a solar still with sponge cubes
[16] A.F. Muftah, M.A. Alghoul, A. Fudholi, M.M. Abdul-Majeed, K. Sopian, Factors affect- in basin, Energy Convers. Manag. 44 (2003) 1411–1418, http://dx.doi.org/10.1016/
ing basin type solar still productivity: a detailed review, Renew. Sust. Energ. Rev. S0196-8904(02)00162-0.
32 (2014) 430–447, http://dx.doi.org/10.1016/j.rser.2013.12.052. [47] V. Velmurugan, M. Gopalakrishnan, R. Raghu, K. Srithar, Single basin solar still with
[17] H.Ş. Aybar, F. Egelioǧlu, U. Atikol, An experimental study on an inclined solar water fin for enhancing productivity, Energy Convers. Manag. 49 (2008) 2602–2608,
distillation system, Desalination 180 (2005) 285–289, http://dx.doi.org/10.1016/j. http://dx.doi.org/10.1016/j.enconman.2008.05.010.
desal.2005.01.009. [48] V. Velmurugan, S. Pandiarajan, P. Guruparan, L.H. Subramanian, C.D. Prabaharan, K.
[18] H.N. Panchal, M.I. Patel, B. Patel, M. Doshi, A comparative analysis of single slope solar Srithar, Integrated performance of stepped and single basin solar stills with mini
still coupled with flat plate collector and passive solar still, IJRRAS 7 (2011) 111–116. solar pond, Desalination 249 (2009) 902–909, http://dx.doi.org/10.1016/j.desal.
[19] P. Durkaieswaran, K.K. Murugavel, Various special designs of single basin passive 2009.06.070.
solar still — a review, Renew. Sust. Energ. Rev. 49 (2015) 1048–1060, http://dx. [49] A.E. Kabeel, A. Khalil, Z.M. Omara, M.M. Younes, Theoretical and experimental
doi.org/10.1016/j.rser.2015.04.111. parametric study of modified stepped solar still, Desalination 289 (2012) 12–20,
[20] H. Al-Hinai, M.S. Al-Nassri, B. Jubran, Effect of climatic, design and operational pa- http://dx.doi.org/10.1016/j.desal.2011.12.023.
rameters on the yield of a simple solar still, Energy Convers. Manag. 43 (2002) [50] V. Velmurugan, S. Senthil Kumaran, V. Niranjan Prabhu, K. Srithar, Productivity en-
1639–1650, http://dx.doi.org/10.1016/S0140-6701(03)90659-X. hancement of stepped solar still — performance analysis, Therm. Sci. 12 (2008)
[21] A.S. Nafey, M. Abdelkader, A. Abdelmotalip, A. Mabrouk, Parameters affecting solar 153–163, http://dx.doi.org/10.2298/TSCI0803153V.
still productivity, Energy Convers. Manag. 41 (2000) 1797–1809. [51] A.E. Kabeel, Performance of solar still with a concave wick evaporation surface, En-
[22] K. Kalidasa Murugavel, S. Sivakumar, J.R. Ahamed, K.K.S.K. Chockalingam, K. Srithar, ergy 34 (2009) 1504–1509, http://dx.doi.org/10.1016/j.energy.2009.06.050.
Single basin double slope solar still with minimum basin depth and energy storing [52] A.A. El-Sebaii, M.R.I. Ramadan, S. Aboul-Enein, M. El-Naggar, Effect of fin configura-
materials, Appl. Energy 87 (2010) 514–523, http://dx.doi.org/10.1016/j.apenergy. tion parameters on single basin solar still performance, Desalination 365 (2015)
2009.07.023. 15–24, http://dx.doi.org/10.1016/j.desal.2015.02.002.
[23] A.A. El-Sebaii, Effect of wind speed on active and passive solar stills, Energy [53] R. Tripathi, G.N. Tiwari, Effect of water depth on internal heat and mass transfer for
Convers. Manag. 45 (2004) 1187–1204, http://dx.doi.org/10.1016/j.enconman. active solar distillation, Desalination 173 (2005) 187–200, http://dx.doi.org/10.
2003.09.036. 1016/j.desal.2004.08.03.
[24] N. Rahbar, J.A. Esfahani, Experimental study of a novel portable solar still by utiliz- [54] A.K. Tiwari, G.N. Tiwari, Effect of water depths on heat and mass transfer in a pas-
ing the heatpipe and thermoelectric module, Desalination 284 (2012) 55–61, sive solar still: in summer climatic condition, Desalination 195 (2006) 78–94,
http://dx.doi.org/10.1016/j.desal.2011.08.036. http://dx.doi.org/10.1016/j.desal.2005.11.014.
M.S.S. Abujazar et al. / Desalination 385 (2016) 178–193 193

[55] M.S.K. Tarawneh, Effect of water depth on the performance evaluation of solar still, [84] A.Y. Hashim, J.M. Al-asadi, W.A. Taha, Experimental investigation of symmetrical
Jordan J. Mech. Ind. Eng. 1 (2007) 23–29. double slope single basin solar stills productivity with different insulation, J.
[56] M.K. Phadatare, S.K. Verma, Influence of water depth on internal heat and mass KUFA Phys. 1 (2009) 26–32.
transfer in a plastic solar still, Desalination 217 (2007) 267–275, http://dx.doi. [85] M. Gabr, Solar desalination as an adaptation tool for climate change impacts on the
org/10.1016/j.desal.2007.03.006. water resources of Egypt Associate Prof. Mariam Gabr Salem, 2013.
[57] A.K. Tiwari, G.N. Tiwari, Thermal modeling based on solar fraction and experimen- [86] E. Delyannis, Historic background of desalination and renewable energies, Sol. En-
tal study of the annual and seasonal performance of a single slope passive solar ergy 75 (2003) 357–366, http://dx.doi.org/10.1016/j.solener.2003.08.002.
still: the effect of water depths, Desalination 207 (2007) 184–204, http://dx.doi. [87] S.A. Kalogirou, Seawater desalination using renewable energy sources, Prog. Ener-
org/10.1016/j.desal.2006.07.011. gy Combust. Sci. 31 (2005) 242–281, http://dx.doi.org/10.1016/j.pecs.2005.03.001.
[58] A. Ahsan, M. Imteaz, U.A. Thomas, M. Azmi, A. Rahman, N.N.N. Daud, Parameters [88] A.M.K. El-Ghonemy, Water desalination systems powered by renewable energy
affecting the performance of a low cost solar still, Appl. Energy 114 (2014) sources: review, Renew. Sust. Energ. Rev. 16 (2012) 1537–1556, http://dx.doi.
924–930, http://dx.doi.org/10.1016/j.apenergy.2013.08.066. org/10.1016/j.rser.2011.11.002.
[59] S. Suneja, G.N. Tiwari, Effect of water depth on the performance of an inverted ab- [89] A. Mehta, A. Vyas, N. Bodar, D. Lathiya, Design of solar distillation system, Int. J. Adv.
sorber double basin solar still, Energy Convers. Manag. 40 (1999) 1885–1897, Sci. Technol. 29 (2011) 67–74.
http://dx.doi.org/10.1016/S0196-8904(99)00047-3. [90] S. Avvannavar, M. Mani, N. Kumar, An integrated assessment of the suitability of
[60] M. Ali Samee, U.K. Mirza, T. Majeed, N. Ahmad, Design and performance of a simple domestic solar still as a viable safe water technology for India, Environ. Eng.
single basin solar still, Renew. Sust. Energ. Rev. 11 (2007) 543–549, http://dx.doi. Manag. J. 7 (2008) 667–685 (doi:Cited By (since 1996) 2\nExport Date 9 July
org/10.1016/j.rser.2005.03.003. 2012\nSource Scopus).
[61] F. Farshchi Tabrizi, M. Dashtban, H. Moghaddam, K. Razzaghi, Effect of water flow [91] G. Nebbia, Early work on solar distillation in Italy, 1953–1970, Proc. Sol. World
rate on internal heat and mass transfer and daily productivity of a weir-type cas- Congr. (2005) 2709–2713.
cade solar still, Desalination 260 (2010) 239–247, http://dx.doi.org/10.1016/j. [92] H.E. Fath, M. El-Samanoudy, K. Fahmy, A. Hassabou, Thermal-economic analysis
desal.2010.03.037. and comparison between pyramid-shaped and single-slope solar still configura-
[62] A.S. Abdullah, Improving the performance of stepped solar still, Desalination 319 tions, Desalination 159 (2003) 69–79, http://dx.doi.org/10.1016/S0011-
(2013) 60–65, http://dx.doi.org/10.1016/j.desal.2013.04.003. 9164(03)90046-4.
[63] A.M. El-Zahaby, A.E. Kabeel, A.I. Bakry, S.A. El-agouz, O.M. Hawam, Augmentation [93] A.E. Kabeel, S.A. El-Agouz, Review of researches and developments on solar stills,
of solar still performance using flash evaporation, Desalination 257 (2010) Desalination 276 (2011) 1–12, http://dx.doi.org/10.1016/j.desal.2011.03.042.
58–65, http://dx.doi.org/10.1016/j.desal.2010.03.005. [94] S. Kumar, G.N. Tiwari, Life cycle cost analysis of single slope hybrid (PV/T) active
[64] P. Patel, A.S. Solanki, U.R. Soni, A.R. Patel, A review to increase the performance of solar still, Appl. Energy 86 (2009) 1995–2004, http://dx.doi.org/10.1016/j.
solar still : make it multi layer absorber, Int. J. Recent Innov. Trends Comput. apenergy.2009.03.005.
Commun. 2 (2014) 173–177. [95] H.N. Panchal, P.K. Shah, O. Of, Modelling and verification of single slope solar still
[65] S. Abdallah, O. Badran, M.M. Abu-Khader, Performance evaluation of a modified using ANSYS-CFX, Int. J. Energy Environ. 2 (2011) 985–998 (www.IJEE.
design of a single slope solar still, Desalination 219 (2008) 222–230, http://dx. IEEFoundation.org).
doi.org/10.1016/j.desal.2007.05.015. [96] H.N. Panchal, M. Doshi, P. Chavda, R. Goswami, Effect of cow dung cakes inside
[66] S. Kumar, G.N. Tiwari, H.N. Singh, Annual performance of an active solar distillation basin on heat transfer coefficients and productivity of single basin single slope
system, Desalination 127 (2000) 79–88, http://dx.doi.org/10.1016/S0011- solar still, Int. J. Appl. Eng. Res. Dindigul. 1 (2011) 675–690.
9164(99)00194-0. [97] H.S. Kwatra, Performance of a solar still: predicted effect of enhanced evaporation
[67] P. Malaiyappan, N. Elumalai, Review on application of non conventional energy area on yield and evaporation temperature, Sol. Energy 56 (1996) 261–266.
source: solar stills, J. Chem. Pharm. Sci. 7 (2015) 1–4 (www.jchps.com). [98] C.M.O., K.T.K.T., Kiam Beng Yeo, Heat transfer energy balance model of single slope
[68] G.N. Tiwari, A. Kupfermann, S. Aggarwal, A new design for a double-condensing solar still, J. Appl. Sci. 14 (2014) 3344–3348, http://dx.doi.org/10.3923/jas.2014.
chamber solar still, Desalination 114 (1997) 153–164, http://dx.doi.org/10.1016/ 3344.3348.
S0011-9164(98)00007-1. [99] H.Ş. Aybar, H. Assefi, Simulation of a solar still to investigate water depth and glass
[69] O.O. Badran, Experimental study of the enhancement parameters on a single slope angle, Desalin. Water Treat. 7 (2009) 35–40, http://dx.doi.org/10.5004/dwt.2009.
solar still productivity, Desalination 209 (2007) 136–143, http://dx.doi.org/10. 692.
1016/j.desal.2007.04.022. [100] S. Nijmeh, S. Odeh, B. Akash, Experimental and theoretical study of a single-basin
[70] S.A. El-Agouz, Experimental investigation of stepped solar still with continuous solar sill in Jordan, Int. Commun. Heat Mass Transfer 32 (2005) 565–572, http://
water circulation, Energy Convers. Manag. 86 (2014) 186–193, http://dx.doi.org/ dx.doi.org/10.1016/j.icheatmasstransfer.2004.06.006.
10.1016/j.enconman.2014.05.021. [101] A.E. Kabeel, A.M. Hamed, S.A. El-Agouz, Cost analysis of different solar still config-
[71] A.K. Tiwari, G.N. Tiwari, Effect of the condensing cover's slope on internal heat and urations, Energy 35 (2010) 2901–2908, http://dx.doi.org/10.1016/j.energy.2010.
mass transfer in distillation: an indoor simulation, Desalination 180 (2005) 73–88, 03.021.
http://dx.doi.org/10.1016/j. [102] A.A. El-Sebaii, S.J. Yaghmour, F.S. Al-Hazmi, A.S. Faidah, F.M. Al-Marzouki, A.A. Al-
[72] B.A. Akash, M.S. Mohsen, W. Nayfeh, Experimental study of the basin type solar still Ghamdi, Active single basin solar still with a sensible storage medium, Desalina-
under local climate conditions, Energy Convers. Manag. 41 (2000) 883–890, http:// tion 249 (2009) 699–706, http://dx.doi.org/10.1016/j.desal.2009.02.060.
dx.doi.org/10.1016/S0196-8904(99)00158-2. [103] Z.M. Omara, A.E. Kabeel, M.M. Younes, Enhancing the stepped solar still perfor-
[73] A. Bhattacharyya, Solar stills for desalination of Water in rural households, Int. J. mance using internal reflectors, Desalination 314 (2013) 67–72.
Environ. Sustain. 2 (2013) 21–30 (http://www.sciencetarget.com/Journal/index. [104] Y.A.F. El-Samadony, A.E. Kabeel, Theoretical estimation of the optimum glass cover
php/IJES/article/view/326). water film cooling parameters combinations of a stepped solar still, Energy 68
[74] M. Köhl, G. Jorgensen, S. Brunold, B. Carlsson, M. Heck, K. Möller, Durability of poly- (2014) 744–750, http://dx.doi.org/10.1016/j.energy.2014.01.080.
meric glazing materials for solar applications, Sol. Energy 79 (2005) 618–623, [105] A.M. Radhwan, Transient performance of a stepped solar still with built-in latent
http://dx.doi.org/10.1016/j.solener.2005.04.011. heat thermal energy storage, Desalination 171 (2004) 61–76, http://dx.doi.org/
[75] M.K. Gnanadason, P.S. Kumar, V.H. Wilson, A. Kumaravel, Desalination and Water 10.1016/j.desal.2003.12.010.
treatment productivity enhancement of a-single basin solar still, Taylor Fr. 55 [106] F.B. Ziabari, A.Z. Sharak, H. Moghadam, F.F. Tabrizi, Theoretical and experimental
(2014) 1998–2008, http://dx.doi.org/10.1080/19443994.2014.930701. study of cascade solar stills, Sol. Energy 90 (2013) 205–211, http://dx.doi.org/10.
[76] M.K. Gnanadason, P.S. Kumar, V.H. Wilson, A. Kumaravel, B. Jebadason, Comparison 1016/j.solener.2012.12.019.
of performance analysis between single basin solar still made up of copper and GI, [107] M. Sakthivel, S. Shanmugasundaram, T. Alwarsamy, An experimental study on a re-
Int. J. Innov. Res. Sci. Eng. Technol. 2 (2013) 3175–3183. generative solar still with energy storage medium — jute cloth, Desalination 264
[77] B.I. Ismail, Design and performance of a transportable hemispherical solar still, (2010) 24–31, http://dx.doi.org/10.1016/j.desal.2010.06.074.
Renew. Energy 34 (2009) 145–150, http://dx.doi.org/10.1016/j.renene.2008.03. [108] A.N. Minasian, A.A. Al-karaghouli, An improved solar still: the wick-basin type, En-
013. ergy Convers. Manag. 36 (1995) 213–217.
[78] J. Ward, A plastic solar water purifier with high output, Sol. Energy 75 (2003) [109] H. Tanaka, Y. Nakatake, Increase in distillate productivity by inclining the flat plate
433–437, http://dx.doi.org/10.1016/j.solener.2003.07.019. external reflector of a tilted-wick solar still in winter, Sol. Energy 83 (2009)
[79] S. Abdallah, O.O. Badran, Sun tracking system for productivity enhancement of 785–789, http://dx.doi.org/10.1016/j.solener.2008.12.001.
solar still, Desalination 220 (2008) 669–676, http://dx.doi.org/10.1016/j.desal. [110] S.H. Sengar, A.G. Mohod, Y.P. Khandetod, S.P. Modak, D.K. Gupta, Design and devel-
2007.02.047. opment of wick type solar distillation system, J. Soil Sci. Environ. Manag. 2 (2011)
[80] N.Y. Tanaka Hiroshi, Theoretical analysis of a basin type solar still with internal and 125–133.
external reflectors, Desalination 197 (2006) 205–216, http://dx.doi.org/10.1016/j. [111] J.T. Mahdi, B.E. Smith, A.O. Sharif, An experimental wick-type solar still system: de-
desal.2009.02.057. sign and construction, Desalination 267 (2011) 233–238, http://dx.doi.org/10.
[81] Y.A.F. El-Samadony, A.S. Abdullah, Z.M. Omara, Experimental study of stepped 1016/j.desal.2010.09.032.
solar still integrated with reflectors and external condenser, Exp. Heat Transfer [112] A.S. Nafey, M. Abdelkader, A. Abdelmotalip, A. Mabrouk, Enhancement of solar still
28 (2015) 392–404, http://dx.doi.org/10.1080/08916152.2014.890964. productivity using floating perforated black plate, Energy Convers. Manag. 43
[82] G.N. Tiwari, S. Sinha, P. Saxena, S. Kumar, Review of solar distiller in other thermal (2002) 937–946.
applications, Int. J. Sol. Energy 13 (1992) 135–144, http://dx.doi.org/10.1080/ [113] K. Kalidasa Murugavel, K.K.S.K. Chockalingam, K. Srithar, An experimental study on
01425919208909780. single basin double slope simulation solar still with thin layer of water in the basin,
[83] M. Boukar, A. Harmim, Effect of climatic conditions on the performance of a simple Desalination 220 (2008) 687–693, http://dx.doi.org/10.1016/j.desal.2007.01.063.
basin solar still: a comparative study, Desalination 137 (2001) 15–22, http://dx.
doi.org/10.1016/S0011-9164(01)00199-0.

You might also like