You are on page 1of 37

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/254216744

Developments in Solar Still Desalination Systems: A Critical Review

Article  in  Critical Reviews in Environmental Science and Technology · January 2011


DOI: 10.1080/10643389.2011.574104

CITATIONS READS

77 4,803

2 authors:

George Ayoub Lilian Malaeb


American University of Beirut King Abdullah University of Science and Technology
85 PUBLICATIONS   2,407 CITATIONS    18 PUBLICATIONS   1,274 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Solar Still desalination system with enhanced productivity View project

All content following this page was uploaded by George Ayoub on 14 April 2016.

The user has requested enhancement of the downloaded file.


Critical Reviews in Environmental Science and
Technology

ISSN: 1064-3389 (Print) 1547-6537 (Online) Journal homepage: http://www.tandfonline.com/loi/best20

Developments in Solar Still Desalination Systems:


A Critical Review

G. M. Ayoub & L. Malaeb

To cite this article: G. M. Ayoub & L. Malaeb (2012) Developments in Solar Still Desalination
Systems: A Critical Review, Critical Reviews in Environmental Science and Technology, 42:19,
2078-2112, DOI: 10.1080/10643389.2011.574104

To link to this article: http://dx.doi.org/10.1080/10643389.2011.574104

Accepted author version posted online: 20


Oct 2011.

Submit your article to this journal

Article views: 398

View related articles

Citing articles: 12 View citing articles

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


http://www.tandfonline.com/action/journalInformation?journalCode=best20

Download by: [American University of Beirut] Date: 14 April 2016, At: 02:29
Critical Reviews in Environmental Science and Technology, 42:2078–2112, 2012
Copyright © Taylor & Francis Group, LLC
ISSN: 1064-3389 print / 1547-6537 online
DOI: 10.1080/10643389.2011.574104

Developments in Solar Still Desalination


Systems: A Critical Review

G. M. AYOUB1 and L. MALAEB2


1
Department of Civil and Environmental Engineering, American University
of Beirut, Lebanon
2
Water Desalination and Reuse Center, King Abdullah University for Science
Downloaded by [American University of Beirut] at 02:29 14 April 2016

and Technology, Saudi Arabia

Solar still desalination uses a sustainable and pollution-free source


to produce high-quality water. The main limitation is low produc-
tivity and this has been the focus of intensive research. A major
concern while increasing productivity is to maintain economic
feasibility and simplicity. The authors present a critical review of
the research work conducted on solar stills development. Studies
addressing each parameter of concern are grouped together and
results compared. Novelty in design and newly introduced features
are presented. Modeling efforts of flow circulation within the still
and methods to estimate internal heat transfer coefficients are dis-
cussed and future research needs are outlined.

KEY WORDS: design, modeling, solar still, productivity

1. INTRODUCTION

Two challenges relentlessly face humanity today: (a) fresh water is a fi-
nite resource seriously impacted by pollution; (b) fossil fuels are also finite
and their use in fuel-based desalination aggravates environmental pollution
problems. Technologies such as multistage flash, multiple effect, vapor com-
pression, reverse osmosis, ion exchange, electrodialysis, phase exchange,
and solvent extraction are heavy energy consumers often requiring complex
controls and bulky processes (Abu-Arabi and Venkat Reddy, 2005; Fath,
1998; Muller-Holst et al., 1999).

Address correspondence to G. M. Ayoub, Department of Civil and Environmental Engi-


neering, American University of Beirut, Lebanon. E-mail: gayoub@aub.edu.lb

2078
Developments in Solar Still Desalination Systems 2079

Desalination based on renewable energy, on the other hand, presents


a sustainable and a zero-polluting alternative. Renewable resources such
as solar energy allow energy diversification, are cheap and available for
predictable periods of time, and help avoid dependence on external energy
supplies (Garcia-Rodriguez, 2002). Until recently, the availability of fossil
fuels and the tolerated levels of CO2 in the atmosphere slowed down the
widespread reality of solar desalination (Mahmoudi et al., 2009). Today,
however, with the advent of climate change, conditions have changed and
the competitiveness of solar desalination is on the rise.
Solar stills make use of a sustainable and pollution-free source (i.e., the
sun) to produce high-quality water. They are classified among the most in-
expensive desalination methods particularly when used in arid areas where
Downloaded by [American University of Beirut] at 02:29 14 April 2016

sunshine is abundant and fresh water is scarce. Contrary to conventional


plants, desalination plants based on renewable energy are still evolving
and often implemented as pilot plant applications that still face operational
problems (Mathioulakis et al., 2007). No data exist on the accurate scale of
operation of solar stills worldwide and many of the installed plants are ex-
perimental systems for research purposes. Systematic analytical data about
real operational conditions, problems faced, and costs of produced water,
installation, operation, and maintenance in many of these plants are notably
missing. As to the cost of materials, transportation, and construction, this
depends directly on the location and local conditions (Ayoub and Alward,
1996).
Limitations on the use of solar stills include their low productivity per
unit installation area compared with fuel-based desalination technologies,
their high initial costs especially if the land cost is high, the need for large
installation areas, the variability of the energy source, and the limited ex-
perience with large-scale applications. System components and equipment
used to enhance productivity are still expensive often hindering commercial-
ization. Although solar energy is free, the hardware required for capturing
it, converting it to useful forms, and storing it can be expensive (Gude and
Nirmalakhandan, 2010). Another challenge is the variability and intermittent
character of the energy source, leading to limitations concerning the maxi-
mum exploitation capacities per time unit. The geographical distribution of
solar desalination potential may not comply with the water stress intensity
at a local level. A continued interest grows, however, in developing tech-
nologies that are suitable for villages and remote areas. This is especially
true in countries with the greatest water need and where standard fuels
required for large-scale desalination are scarce and expensive (e.g., India,
Egypt, Pakistan; Ayoub and Alward, 1996; Samee et al., 2007). Solar stills may
be economically viable if small water quantities are required and the cost
of pipe work required to supply an arid area with water is high (Kalogirou,
2005). They represent the best technical solution to supply remote villages
2080 G. M. Ayoub and L. Malaeb

or settlements with fresh water without depending on high technology and


expertise (Eltawil et al., 2009), particularly where sunshine is abundant.
Arjunan et al. (2009) described seven large-size solar still installations in
India that vary in capacity from 0.13 to 8 m3/day, mostly treating seawater
and in few cases saline well water. The evaporating areas for these plants
vary from 50 m2 up to 3,110 m2 depending on capacity. Two plants consist-
ing of 240 stills each with a capacity to clean 27.3 m3/day of seawater have
been installed in Balauchistan, Pakistan, and other similar schemes are being
considered to desalinate the brackish to saline groundwater present (Samee
et al., 2007; World Energy Council, 2000).
In their simplest design, they comprise transparently roofed black-
painted basins containing the water to be desalted. Water in the basin gets
Downloaded by [American University of Beirut] at 02:29 14 April 2016

heated and evaporates, and the vapor condenses as it hits the roof and
trickles down reaching a channel, which transmits the flow into a collection
container. The stills can have various forms, shapes, and cover materials
and their operation requires little maintenance besides regularly flushing the
basin to remove accumulated salts. As with other desalination processes, the
resulting brine remains to induce a negative environmental impact. How-
ever, such brine could prove to be an important source of metals that may
be used in various industrial applications (Ayoub et al., 2000). Disposal meth-
ods currently practiced include surface water discharge, disposal to sewers,
deep-well disposal, land application, evaporation ponds, brine concentra-
tors, and spray-drying to solids or crystallizers (i.e., zero discharge option;
Fath, 1998). Discharging brine in turbulent hydrodynamic conditions such as
water bodies featuring waves, currents, and supercritical flows and using dif-
fusers to reduce salinity through mixing and dilution are often recommended
(Medeazza, 2005). Salts recovery (Jegatheesan et al., 2009) may also be an
option to utilize the large mass of concentrated salt components. The ma-
jor waste stream produced by solar stills is not characterized by intentionally
added process chemicals, as is the case with other desalination technologies.
Rather, the concentrate reflects the raw water characteristics, which is of the
same composition but at a more concentrated level (Fath, 1998).
Although environmental impact assessment procedures have been pro-
posed, technical solutions, codes, and standards are still in their early phases
(Medeazza, 2005). The environmental concerns primarily center on the con-
tamination of surface and ground waters, underlying soil, and drinking water
aquifers. Mitigation methods include additional processing to remove or di-
lute any chemicals of concern, use of noncorrosive and nontoxic materials,
use of diffusers, and continuous blending of cleaning wastes and concentrate
or separate disposal of cleaning wastes and concentrates.
The main limitation of solar stills is their low productivity compared with
conventional desalination processes, and this has been the focus of intensive
research. The operating efficiency is low due to main two shortages: the first
Developments in Solar Still Desalination Systems 2081

TABLE 1. Average output of standard simple solar stills

Yield (l/m2/day) Reference

1.72 Tiwari and Tiwari (2006)


3.8 Abu-Arabi and Zurigat (2005)
1.5 Nijmeh et al. (2005)
3.2–3.5 Al-Hayek and Badran (2004)
4.15 Al-Hinai et al. (2002a)
0.98 Aggarwal and Tiwari (1999)
2.4 Akash et al. (1998)
1.58 Kumar and Tiwari (1996)
1.7–3.2 Kamal (1988)
3 Rajvanshi (1981)
Downloaded by [American University of Beirut] at 02:29 14 April 2016

is the rejection to the atmosphere of the latent heat of condensation and the
second is the difficulty in raising evaporation temperature and decreasing
condensation temperature as heating, evaporation, and condensation take
place in one container (He and Yan, 2009). Table 1 presents typical daily
yields for the conventional basin-type still. The yield is usually measured in
liters per unit area per day. Important factors, however, are masked in this
measurement method, which does not reflect the influence of parameters
that directly affect the yield such as available insolation and sunshine hours
in the geographical area where the still is installed.
A major concern in increasing the amount of distillate produced is to
maintain economic feasibility and simplicity in construction, operation, and
maintenance. In this article we present a critical review of the research
work conducted on solar stills development. Important parameters that af-
fect the still productivity include the use of absorbing materials and wicks,
cover cooling, cover slope, still design, brine depth, inlet water flow, and
initial brine temperature. Studies addressing each parameter of concern are
grouped together and results are compared. Novelty in design and newly in-
troduced features and their impact on output are then presented. Modeling
efforts of flow circulation within the still and methods to estimate internal
heat transfer coefficients are discussed next. Finally, costs of solar stills are
presented and future research needs are outlined.

2. RESEARCH INVESTIGATIONS ON KEY PARAMETERS


OF CONCERN

Several parameters affect the performance of a solar still. These include the
presence of a dye or absorbing material, cover cooling by air or water flow,
cover slope, still design, geographical position, water depth, inlet water flow,
and initial brine temperature. Research investigations on these influencing
factors are discussed in this section.
2082 G. M. Ayoub and L. Malaeb

2.1 Effect of Absorbing Materials


Various approaches for increasing the basin absorptivity have been tested
and found effective in increasing the daily yield of a solar still. These include
the use of charcoal (Naim and Abdel-Kawi, 2002; Okeke et al., 1990; Tiris
et al., 1996), black, and violet dyes, which were found to be more effective
than other dyes, and their effect increases with deeper stills (Sodha et al.,
1980), floating absorber aluminum sheet (Valsaraj, 2002), and black volcanic
rocks, which performed better than metallic wiry sponges (Abdallah et al.,
2009).
Table 2 summarizes the increase in productivity obtained for different
absorbers. These materials can store higher amounts of heat energy and
increase the heat capacity of the basin in addition to increasing the basin ab-
Downloaded by [American University of Beirut] at 02:29 14 April 2016

sorption. As noted in Table 2, the increase in yield, which is usually reported


in liters per day per unit area, varies depending on the specific absorber used

TABLE 2. Effect of different absorbers on productivity

Absorber Increase in yield Reference

Black granite gravel 17–20% Sakthivel and


Shanmugasundaram (2008)
Sponge cubes 14–18% Velmurugan et al. (2008a);
Velmurugan and Srithar
(2007); Bassam et al. (2003)
Sponges and fins in a stepped 96% Velmurugan et al. (2008c)
still
Potassium dichromate 26% Nijmeh et al. (2005)
Charcoal 15% more than Naim and Abd El Kawi (2002)
wick-type stills
Black gravel 19% Nafey et al. (2001)
Baffle suspended absorber 20% El-Sebaii et al. (2000)
Absorbing black-rubber mat 38% Akash et al. (1998)
Black ink 45% Akash et al. (1998)
Black dye 60% Akash et al. (1998)
Rubber material 38% Bilal et al. (1998)
Charcoal 11–18% more than Tiris et al. (1996)
black-paint and
23–92% more than
blackened rock-bed
Soot powder 13–17% Madani and Zaki (1995)
Dyes 10% Dutt et al. (1989)
Dyes 19.8% Rai and Tiwari (1983)
Black napthylamine 17–28.8% Rajvanshi (1981)
(172.5 mg/l)
Black napthylamine (50 mg/l) 15.9%
Red carmoisine (50 mg/l) 7.4%
Red carmoisine (100 mg/l) 19.9%
Dark green dyes (50 mg/l) 14.8%
Dark green dyes (100 mg/l) 12%
Dyes 14.6% Garg and Mann (1976)
Developments in Solar Still Desalination Systems 2083

from as low as 7% for red carmoisine (Rajvanshi, 1981), as high as 60% for
black dyes (Akash et al., 1998), and as high as 96% for sponges and fins (Vel-
murugan et al., 2008c). An accurate comparison of the effects of different ab-
sorbers in these studies is difficult, however, due to the large variation in the
experimental conditions of each study. Such conditions include geographi-
cal location, time of the year, particular still design and construction material
used, brine depth, water temperature, and other site-specific factors. Aybar
et al. (2005) tested an inclined flat solar absorber plate with a black cloth or
a black fleece and showed that the black wick still produced 2–3 times more
water than a bare plate still. The collected data in this study, however, was
limited to a two-week period and tests were performed over 7 hr per day.
According to Murugavel et al. (2008), rubber is the best material to
Downloaded by [American University of Beirut] at 02:29 14 April 2016

improve absorption, storage, and evaporation effects. Nafey et al. (2001),


however, reported that black gravel absorbed and released incident solar
energy faster than black rubber. Jamal (1991) also found that a mixture
of copper sulfate and potassium permanganate solutions was the ideal ab-
sorber for solar energy compared with individual solutions of copper sulfate,
potassium permanganate, potassium dichromate, cobalt chloride, and TMPD
(tetramethyl-p-phenylenediamine). Khalifa and Hamood (2009a) developed
four correlations for the effects of brine depth, dyes, solar radiation, and
cover tilt angle on productivity based on data from several studies. These
correlations apply for passive stills with a galvanized iron body, an insu-
lation thickness of 5–10 cm of polystyrene or other insulation with similar
conductivity, glass cover with 5–45◦ tilt angle, 1–10 cm brine depth, 20–35◦
latitude angle, 50–100 mg/l dye concentration, and under solar radiation of
8–30 MJ/m2/day. They correlated the productivity yD in the presence of a
dye to the productivity y without dye as follows:

yD = 1.2122y1.0467 , r2 = 0.833 (1)

Where r2 is the best root mean square value. These correlations, however,
should be used with care given the assumptions and limitations that they
were derived for and that are stated previously. Further calibration of the
proposed equations may be necessary and the degree of their validity to
similar stills that have not been included in their derivation is unknown.

2.2 Cover Cooling


Water flow over the still cover at a very low rate has been shown to increase
the still output (Dutt et al., 1993) and proper use of film-cooling parameters
may increase efficiency by 20% (Abu-Hijleh and Mousa, 1997). The convec-
tive heat transfer coefficient also increases with wind velocity, leading to
a decrease in the cover temperature and hence an increase in the overall
yield (Dimri et al., 2008). El-Sebaii (2000) found that productivity increases
2084 G. M. Ayoub and L. Malaeb

with wind speed up to a certain value between 8 and 10 m/s for winter
and summer conditions, respectively. This value was independent of the still
shape and brine heat capacity, and the wind was more effective in summer
and for higher water masses. Similarly, Fath and Ghazy (2002) reported that
increasing the air flow rate up to 0.5 m3/s increases productivity with no
further improvement obtained with air flowing beyond this rate. Tiwari and
Rao (1984) reported that the daily distillate production is almost doubled for
stills with water flowing over glass cover at a uniform velocity. On the other
hand, Fath and Hosny (2002) found that glass cover cooling by wet cloth for
1-, 2-, and 6-hr intervals had no effect on productivity and that continuous
cooling is needed to release a significant amount of condensation energy.
No effect on productivity was noted in their study for wind speed in the
Downloaded by [American University of Beirut] at 02:29 14 April 2016

range of 0–5 m/s.

2.3 Cover Slope and Still Geometry


Several studies examined the effect of cover slope (Kumar and Tiwari, 1998;
Murugavel et al., 2008; Singh et al., 1995; Tanaka and Nakatake, 2009a).
For locations with latitude higher than 20◦ , single-sloped stills were found
preferable whereas for lower latitudes, double-sloped stills facing north and
south directions and having a cover inclination equal to the latitude angle
received sun rays close to normal throughout the year (Murugavel et al.,
2008). Khalifa and Hamood (2009a) developed the following correlation
for the effect of cover tilt angle a on productivity y, which applies for the
conditions stated in section 2.1:

y = −0.0025a2 + 0.1562a + 0.843, r2 = 0.743 (2)

The effects of using different shapes (Figure 1) and having multibasins have
also been studied. In multiple-effect basin stills, the floor of the upper com-
partment acts as the condensing surface of the lower compartment. As a
result, efficiency is greater than that of a single-basin still typically by 35%
or more but the cost and complexity are higher (Qiblawey, 2008). Table 3
summarizes the results of different studies conducted on cover slope and
still geometry.

2.4 Use of Wicks


Many studies were conducted on wick-type stills (Janarthanan et al., 2005;
Minasian and Al-Karaghouli, 1995; Shukla and Sorayan, 2005; Velmurugan
et al., 2008a) or capillary film distillers (Bouchekima et al., 1998). Suction
due to capillary action of cloth fibers creates a thin water film that enhances
evaporation, leading to a 4% increase in efficiency over that of a simple basin
Developments in Solar Still Desalination Systems 2085
Downloaded by [American University of Beirut] at 02:29 14 April 2016

FIGURE 1. Spherical solar still (left) and double-basin still (right) (Color figure available
online).

still and a 10% increase in yield when water is flowing over the cover (Dhi-
man and Tiwari, 1990; Kumar and Anand, 1992; Sodha et al., 1981b). Tanaka
et al. (1981) showed that the performance of a tilted-wick still is superior to
the conventional still with an increase in productivity by about 20–50%. The
wick basin type solar still designed by Minasian and Al-Karaghouli (1995) in-
creased productivity by 85% while that studied by Velmurugan et al. (2008a)
gave a 29.6% increase compared with 15% and 45.5% when using sponges
and fins, respectively. Analytical expressions for the thermal efficiency of
evaporative heat loss and heat transfer in these types of stills have also been
derived (Janarthanan et al., 2005; Shukla and Sorayan, 2005).

2.5 Brine Depth


Studies conducted on the effect of water depth in stills have shown that
the highest outputs and efficiencies occur at lower depths (Badran, 2007;
Cooper, 1969; El-Sebaii, 2005; Nafey et al., 2000, 2002; Tiwari and Tiwari,
2007; Phadatare and Verma, 2007). It was observed that nocturnal distillation
is significant in the case of higher water depths because of reduced ambient
and stored energy (Tiwari and Tiwari, 2006). Relative humidity in the still
is important particularly for large water depths and the internal convective
heat transfer coefficient decreases with increasing water depth due to a
decrease in water temperature (Tiwari and Tiwari, 2006; Tripathi and Tiwari,
2005). A summary of the effect of depth on productivity is shown in Table 4,
2086 G. M. Ayoub and L. Malaeb

TABLE 3. Effect of still geometry

Scope of study Main conclusions Reference

Effect of Maximum of 0.36 l/hr obtained for two stages Obikane and Al-Bilbisi
multistages Reverse wind direction reduces efficiency by (2009)
22% while forward wind direction increases
efficiency by 6%.
Effect of (i) 0.8 L/hr for optimal conditions Obikane et al. (2009)
multistages (ii) Triangular shaped multistage still have
1.4–1.5 times more productivity than those
that are rectangular.
Regenerative Similar to a double-effect still: 70% more Abu-Arabi and Zurigat
still productive than the conventional still (2005)
Vertical still Low efficiency of 18.6–33.19% was attributed Boukar and Harmim
to high vapor and condensation heat (2005)
Downloaded by [American University of Beirut] at 02:29 14 April 2016

leakage. Output ranged from 1.02 to 1.91


l/m2/day for an energy input of 13.14 to
13.68 MJ.
Vertical still (i) 0.5–2.3 l/m2 productivity Boukar and Harmim
(ii) Low efficiency is attributed to the low (2004)
temperature difference between cover and
water film on spongy bottom.
Effect of cover Annual yield is at its maximum when the Singh and Tiwari (2004)
inclination condensing glass cover inclination is equal
to the latitude of the place.
Effect of length The change in length for a given height and Tripathi and Tiwari
and height width of a still does not affect the daily (2004)
output but the change in the height of the
north wall for a given height affects the
daily output.
Effect of Double-basin solar stills give better Kumar et al. (1991)
double performance than the single-basin still due
basins to better utilization of latent heat of
vaporization.
Spherical still 30% more efficiency than the conventional Dhiman (1988)
still
Effect of cover Single-sloped stills give better yield in winter Tiwari et al. (1986);
slope while double-slope stills are better in the Yadav and Tiwari
summer (1987)
Effect of cover A double slope multiwick solar still is more Tiwari and Selim (1984)
slope economical and efficient than a simple one
Effect of cover Increasing the tilt angle from 5◦ to 25◦ Al-Jubouri and Khalifa
slope increased the yield from 2.3 to 3.9 l/m2/day (1982)

which reveals a similar decreasing trend in productivity with increasing brine


depth among these studies. Variations in this trend can be attributed to the
particular site conditions and geographical location as well as to the type
of still tested. For the effect of brine depth d, Khalifa and Hamood (2009a)
developed the following correlation with productivity, which applies for the
conditions stated in section 2.1:

y = 3.884e−0.0458d , r2 = 0.832 (3)


Developments in Solar Still Desalination Systems 2087

TABLE 4. Effect of brine depth (modified from Khalifa and Hamood, 2009b)

Variation in Decrease in
brine depth Type of still productivity (%) Reference

2, 4, 8, 12, 16, Single sloped, 30◦ 25%; the yield became Tiwari and Tiwari (2007)
and 18 cm constant for depths
>10 cm
2–3.5 cm Single sloped, 32◦ 26% Badran (2007)
1.5–4 cm Single and double 9% Al-Hayek and Badran
sloped (2005)
3–6 cm Single sloped, 15◦ 29% Nafey et al. (2002)
2–30 cm Double sloped 14% Al-Hinai et al. (2002a)
3–6 cm Double sloped, 15◦ 13.5% Akash et al. (2000)
2–7 cm Single sloped, 32◦ 14% Nafey et al. (2000)
2–5 cm Single sloped, 8–13◦ 11% Kamal (1988)
Downloaded by [American University of Beirut] at 02:29 14 April 2016

1–10 cm Single sloped, 10◦ Increase by 34% Tiwari and Madhuri


(1987)
2–8 cm Uninsulated double 13.8% Garg and Mann (1976)
sloped
1.25–30 cm Double sloped 30% Cooper (1969)

Murugavel et al. (2008) reported that for deep basins, water depth is in-
versely proportional to productivity during daylight but the reverse is true
for overnight production. Also, Tripathi and Tiwari (2005) reported that more
yield is obtained during offshine hours as compared to daytime for higher
water depths due to storage effects. As for continuous saline water feeding,
lower mass flow rates were found to produce more output (Aybar et al.,
2005).

2.6 Inlet Water Flow and Temperature in the Basin


Studies have reported that the overall thermal efficiency of a solar still de-
creases with an increase in mass flow rate (Singh and Tiwari, 1992) and stills
fed with hot water at constant rate give a higher yield than those filled with
hot water only once a day (Sodha et al., 1981a; Tiwari et al., 1985). Produc-
tivity is reported to increase with increasing mass flow rate for higher inlet
water temperatures and decrease for inlet temperatures less than the average
ambient temperature (Gupta et al., 1988; Tiwari et al., 1985). Productivity is
also higher for hot water flow during off-sunshine hours than for contin-
uous flow with lower inlet temperature (Tiwari et al., 1985). According to
Yadav and Kumar (1991), water temperature, distillate output. and efficiency
increase with decreasing mass flow rate. In their study, the optimum value
of the mass flow rate through the still basin was found to be 0.00027 l/m2/s.
The internal evaporative heat transfer is also reported to be a strong function
of the initial water temperature (Kumar and Tiwari, 1990).
2088 G. M. Ayoub and L. Malaeb

3. DEVELOPMENTS IN SOLAR STILL DESIGN

A simple passive solar still can produce from 1–4 l/m2/day (Table 1) depend-
ing on location, water depth, time of the year, tilt angle, construction, and
insulation material and solar radiation. According to Khalifa and Hamood
(2009a), solar radiation I (MJ/m2/day) has the highest impact on productivity
and they correlated them as follows:

y = 0.0036I2 + 0.0701I + 0.2475, r2 = 0.762 (4)

This section presents novel developments introduced in the design of solar


stills with the objective of increasing the amount of distillate output.
Downloaded by [American University of Beirut] at 02:29 14 April 2016

3.1 Stills Coupled to Condensers, Reflectors, and External Heaters


Table 5 presents the results of studies on solar stills with additional con-
densers. The use of external condensers can be either through purging or by
allowing natural circulation as illustrated in Figure 2. In externally heated or
active solar stills, water temperature in the basin is increased by integrating
the still with a solar heater, solar concentrator, or waste heat recovery system.
Circulation through the heater or concentrator can also either be through nat-
ural circulation (thermosiphon) or through forced circulation using a pump
as shown in Figure 3. According to Lawrence and Tiwari (1990), operation
under natural circulation mode has been proven to be more advantageous
in terms of simplicity, reliability, and cost effectiveness.
Although their productivity is higher, active solar stills often require ex-
tra space, are more complex, and hence more costly (Khawaji et al., 2008)

FIGURE 2. Still with passive condenser: (a) purging and (b) natural circulation (Color figure
available online).
Downloaded by [American University of Beirut] at 02:29 14 April 2016

TABLE 5. Studies on solar stills with additional condensers

Main Findings
Type of Still of the Research Productivity Basic Features Reference

Separate evaporation Optimizing energy efficiency and 40 l/m2 product water 15 patented interconnected Koning and Thiesen
and condensation approaching the limit of day in day out for elements enhance each (2005)
chambers freshwater output. optimized version others’ performance and
productivity.
Built-in additional Most influential factors are solar 8 l/d/m2 base area (55% Built-in finned condenser tilted Fath and Hosny
condenser and intensity, insulation, basin mass, more than the to be parallel to sun rays (2002)
enhanced evaporation evaporation surface area and reference) and enhanced evaporation
condenser inner reflectivity. using black dyes.
Still with a radiative A delay in response to thermal Maximum condensation The radiative panel, storage Haddad et al. (2000)
cooling system capacity advances the time at is 2 L in 10 hr. tank, and still are designed
which maximum productivity so that the tank is able to
occurs. Productivity increases condense all vapor
with solar radiation but produced during daylight.
efficiency decreases due to
increased thermal losses.
Inverted absorber solar About double the output of the An optimum value for The evaporative heat loss is a Tiwari and Suneja
still conventional still. yield is reached for strong function of (1998, 1999)
seven basins. temperature.
Still with additional Efficiency increased by 45% A condenser in the shaded Fath (1998)
condenser compared to conventional still. region of the still.
Double-condensing Significant enhancement in 35–77% higher output Suneja and Tiwari
chamber still output due to vapor pressure than that of (1997)
difference between the two conventional solar
condensing chambers. still.
Simple basin still with 2–3 l/m2/day Passively cooled condenser Fatani et al. (1994)
condenser placed on the still roof.
Simple basin still with Efficiency = 71% 7.1 l/m2/day Condenser placed under the Lessing (1990)
condenser energy collecting area.
Multiple solar still with a 20% higher yield than the simple 15–25% higher than the Excess vapor is condensed on Reddy et al. (1983);
condenser wick still but under cloudy and noncondenser-type the additional surface and Tiwari et al. (1984)
arrangement low intensity conditions both still. reduces heat load on the
stills show similar performance. cover.
Simple basin still with Efficiency = 42% 4.5 l/m2/day Passively cooled condenser Chandra et al. (1980)

2089
condenser placed inside the still.
2090 G. M. Ayoub and L. Malaeb
Downloaded by [American University of Beirut] at 02:29 14 April 2016

FIGURE 3. Solar still coupled with flat-plate collector (Color figure available online).

than simple passive stills. Table 6 summarizes the basic features and findings
of studies conducted on solar stills that are coupled to external heat collec-
tors and/or reflectors (Figures 3 and 4). Although the increase in productivity
can be significant in some cases, most studies do not show the additional
costs entailed by the proposed design. Similarly, system feasibility and jus-
tifications for increased area requirements and complexities in maintenance
and operation are often not included.

3.2 Novel Developments in Design


Modifying the solar still design with the purpose of maximizing output at
a minimal cost increase has been the subject of intensive research. Exam-
ples of such modifications include utilizing a transportable hemispherical
dome-shaped still (Ismail, 2009), and employing a computerized sun track-
ing device for rotating the solar still with the movement of the sun, which
can increase the productivity by around 22% (Abdallah and Badran, 2008).
Sun-trackers, however, are not recommended for small solar panels because
of their power consumption and high-energy losses in their driving systems
(Mousazadeh et al., 2009). Abdel-Rahim and Lasheen (2005) used a packed
layer of glass balls as a thermal storage system and installed a rotating shaft
to break the boundary layer of the basin water surface, which increased effi-
ciency by 5–7.5% and 2.5–5.5%, respectively. Their study, however, did not
show whether the yield increase justifies the cost of photovoltaic arrays and
DC-motor needed to operate the shaft. Al-Kharabsheh and Goswami (2003)
utilized vacuum conditions to lower the evaporation temperature of water
in a still and hence reduce the required thermal input. The proposed setup,
Downloaded by [American University of Beirut] at 02:29 14 April 2016

TABLE 6. Studies on solar stills with collectors and reflectors

Main findings
Type of still of the research Productivity Basic features Reference

Air-blown tilted wick A major part of the enthalpy of 1 l/m2/hr for double Heat and mass transfer is Aboabboud et al.
double-chamber with the hot saturated air entering glass cover based on the flow rate of (2009)
heat recycling the lower chamber was saturated air pumped in.
recycled.
Tilted wick still with an Optimal reflector’s inclination 15–27% more output An inclined external reflector Tanaka and Nakatake
inclined reflector was 15◦ from the vertical than still with vertical to increase the productivity (2009a)
reflector of a tilted-wick still.
Tilted wick still with a Optimum still orientation is 41% more yield than the The still is rotated once a day Tanaka and Nakatake
vertical flat plate obtained for the equinox and simple basin type at southing of the sun. (2009b)
reflector solstice days.
Vertical multiple-effect Productivity increases by Productivity is 9–17% No contamination contrary to Tanaka and Nakatake
diffusion type solar increasing partition area and higher for optimum inclined wick stills, which (2007)
still, coupled with a the number of partitions and by collector angles than get soaked and where
heat-pipe solar decreasing the saline water the fixed ones partitions deform due to
collector feed rate and the thickness of protruding fibers from
diffusion gaps between wicks.
partitions.
Asymmetrical still with 20% more output than a Use of mirrors on the inside Al-Hayek and Badran
mirrors symmetric greenhouse walls (2004)
still
Coupling with storage Coupling with a hot water storage Thermal coupling with a Productivity is a linear Mathioulakis and
tank and external heat tank leads to a constant tank can double the function of solar radiation Belessiotis (2003)
sources production rate. productivity. and the average values of air
temperatures of previous
days.
(Continued on next page)

2091
Downloaded by [American University of Beirut] at 02:29 14 April 2016

2092
TABLE 6. Studies on solar stills with collectors and reflectors (Continued)

Main findings
Type of still of the research Productivity Basic features Reference

Still augmented with A double effect arrangement did 2.19 l/m2 versus 3.13 An asymmetric inverted Smyth et al. (2002)
reflector and not improve output. l/m2 for the passive convection suppression
concentrator still circular reflector.
Solar still with stored Efficiency improved by Storing energy during the day Abdel-Rahim, (2001)
excess solar energy 28% and condensing the vapor
inside copper tubes held at
a lower temperature.
Four-effect diffusion still Different modes and 8.7 l/m2/h with energy Two units: A hot storage tank, Rheinlander and
configurations of heat recovery, input of 2.0 kW/m2 coupled to a solar collector, Graeter (2001)
and natural or forced and the distillation unit.
convection in the four
distillation chambers.
A thermal desalination Numerical simulation results and 25 l/m2/day for 4.8 kW A solar collector and a Schwarzer et al. (2001)
unit with a heat experimental data for the unit h/m2/day solar desalination tower of six
recovery system are presented. radiation stages with water circulation.
Still with concentrator Yield increase by 22% External trough type Zaki et al. (1992)
concentrator.
Solar still integrated with Forced circulation mode gives 30–35% increase in yield Using a thermosiphon mode Yadav (1991)
a flat-plate collector 5–10% higher yield than of operation.
thermo-siphon mode.
Single-basin solar still Best performance is for forced 24% higher yield than Disconnecting the collector Rai and Tiwari (1983);
coupled with a circulation and blackened jute for a simple during off-sunshine hours Tiwari (1985); Rai et al.
flat-plate collector cloth floating over the basin single-basin solar still. increases yield. (1990)
water and a small quantity of
black dye.
Developments in Solar Still Desalination Systems 2093
Downloaded by [American University of Beirut] at 02:29 14 April 2016

FIGURE 4. Still with external reflector (Color figure available online).

however, necessitates the presence of a considerable height above the saline


water source and also requires periodic cleaning by flushing and restarting
to remove any noncondensable gases that might accumulate and destroy
the vacuum. Based on theoretical results, it was concluded that the system
performs much better than a simple basin solar still but no quantitative com-
parison is presented. Some of the investigated design modifications, on the
other hand, did not necessarily improve productivity (e.g., a pyramid-shaped
solar still was found to be less efficient and economical than the single-slope
still; Fath et al., 2003).
Other developments involved a complete transformation in the design
such as the use of an elliptical metallic container working under vacuum and
located in the focus of a concave reflector (Nassar et al., 2007) and the use
of horizontal transparent plastic or glass tubes in solar stills with separate
evaporation and condensation sections (Reali and Modica, 2008). In the
latter study, detailed technological features of the proposed prototype are
presented, but field tests and experimental data have not been performed.
Franceschetti (2009) proposed a new still model that combines continuous
2094 G. M. Ayoub and L. Malaeb

flow to eliminate scaling and stagnation with the suction of humid air to
remove condensation on the internal cover. An elongated shape of the still
was built to prolong retention time and accumulate heat so as to evaporate
more of the brackish water before being discharged. The exiting brine is
collected and rerouted to a heat exchanger in countercurrent flow with
the seawater entering the still, gradually increasing the solution’s thermal
energy. Evaporating water is routed in copper tubes drawn by cold salty
solution entering the system. Vapor in copper tubes is liquefied and the
liquefaction’s energy is given back to the cold salty solution. The reported
net output of the system was 5.3–5.8 l/m2/day and was expected to increase
if the still is operated in tropical areas. Kemp (2009) described another novel
still, which uses convoluted tubes, operates on 95–100% renewable energy,
Downloaded by [American University of Beirut] at 02:29 14 April 2016

has very low environmental impact, has minimal operating and maintenance
costs, and is scalable. For the tested conditions, productivity varied between
7 and 9 l/m2/day.

4. FLOW CIRCULATION AND CONVECTION PROCESSES


WITHIN A SOLAR STILL

A limited number of studies have addressed the effects of natural convec-


tion, buoyancy forces, and flow circulation inside a solar still cavity (De Paul,
2002; Hajri et al., 2007; Omri et al., 2005; Porta-Gandara et al., 2004; De Paul,
2009). Some of these studies (De Paul, 2002; Hajri et al., 2007) presented
numerical models and solutions that yield a two-cellular flow field with the
size of cells depending on the Rayleigh number for a fixed Lewis number
(Hajri et al., 2007). These studies also demonstrated that the bottom of a
basin still is not isotherm and the flow structure is sensitive to the cover tilt
angle (Omri et al., 2005). Utilizing visualization techniques via a video cam-
era, laser sheet, and smoke tracer particles, Porta-Gandara et al. (2004) could
identify a periodic development of the convective structure. The flow pattern
formed a repetitive rollover and destruction of horizontal vortices with the
process being highly unstable and appearing random at first sight. It is noted
that the agreement of the obtained value of the Strouhal number with the
constant universal number suggested by Levi (1983) might indicate an under-
lying general instability process not fully understood at present. Levi (1983)
proposed that for energy conservation to be accomplished between a fluid
current and a stagnant fluid body, a constant universal Strouhal number has
to characterize any oscillatory phenomenon in natural fluid motions. De Paul
(2002) reported that the sinking point of flow oscillated along the cover with
frequencies depending on the still geometry, which also governed produc-
tivity and heat transfer. Measured efficiencies were greater than theoretical
values. A strong turbulence was moreover observed in this study causing a
Developments in Solar Still Desalination Systems 2095

fast homogenization of the temperature inside the still and a recirculation of


an important part of the evaporated water to the basin.

5. DEVELOPMENTS IN MATHEMATICAL MODELING

The accurate estimation of internal heat transfer coefficients is a critical step


for the prediction of a solar still performance. The convective heat transfer
coefficient is given by the Nusselt number, which is the ratio of convective
to conductive heat transfer across a boundary (Bejan et al., 1995):

hc L
Nu = = C · (Gr · Pr)n (5)
Downloaded by [American University of Beirut] at 02:29 14 April 2016

K
The Prandtl number expresses the ratio of viscous to thermal diffusion rates
and is only a function of the fluid and the fluid state:
µCp
Pr = (6)
K
The Grashof number Gr approximates the ratio of the buoyancy force to the
viscous force acting on a fluid:

gβρ 2 L3 T
Gr = (7)
µ2

The volumetric thermal expansion coefficient β is a measure of the frac-


tional change in density as temperature changes at constant pressure. The
difference in temperature T is given by

(Pw − Pg )Tw (Pw − Pg )Tw


T = Tw − Tg +   = Tw − Tg + (8)
Ma PT
− Pw 268.9 × 103 − Pw
Ma −Mw

Natural convection is a mechanism of heat transport whereby fluid motion


is generated by a difference in density in the fluid due to a temperature
gradient. The onset of natural convection is determined by the Rayleigh
number:
ρgL3
Ra = Gr · Pr = (9)

The thermal diffusivity D is defined as


K
D= (10)
ρCp
2096 G. M. Ayoub and L. Malaeb

5.1 Estimation of Heat Transfer Coefficients Using Dunkle Relations


Many researchers use the semiempirical relations developed by Dunkle
(1961) to estimate these coefficients. According to these relations, the pa-
rameters C and n in Eq. (1) are equal to 0.075 and 1/3 respectively, and the
convective heat transfer coefficient hc and the evaporation output rate me
are given by
 1/3
 1/3 (Pw − Pg )Tw
hc = 0.844(T ) = 0.884 (Tw − Tg ) + (11)
268.9 × 103 − Pw
hc
me = 0.016273 (Pw − Pg ) (12)
hfg
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Aggarwal and Tiwari (1999) modeled a double-condensing chamber solar


still, which is predicted to be more economical on a large-scale production
than the conventional solar still. Applying the concept of solar fraction to
the thermal analysis of solar stills, Tripathi and Tiwari (2006) found that the
degree of agreement between theoretical and experimental results was larger
for the active than the passive stills and for daytime than nighttime data.
Studying the convective heat transfer coefficient for a passive/active solar
still, Tiwari et al. (2003) concluded that internal heat transfer coefficients
should be determined using inner glass temperature for thermal modeling.
El-Sebaii (2005) developed a transient mathematical model for a triple-basin
solar still based on an analytical solution of the energy-balance equations
using the elimination technique. Rubio-Cerda et al. (2004) developed a model
to predict thermal asymmetries in double-sloped stills treating the condenser
as a two-element system. Dunkle’s correlations were applied and the model
was found to underestimate the predicted temperature differences between
the two covers.

5.2 Limitations and Improvements of Dunkle Relations


Dunkle’s relations apply for cavities that have parallel condensing and evap-
orative surfaces, are independent of the cavity volume, and are only valid for
a low operating temperature range (45–50◦ C) and for an equivalent tempera-
ture difference of 17◦ C between evaporative and condensing surfaces (Tiwari
and Tiwari, 2006). Moreover, the nature of the medium within the still, the
geometry of the system and the characteristics of the flow circulation are not
taken into account and the fact that the environment is an enclosed space
does not appear either (De Paul, 2002). In fact, the discrepancies often found
between theoretical models and experimental measurements are attributed
to these assumptions (Hongfei et al., 2002) and it is estimated that Dunkle’s
model might overpredict evaporation rates by about 30% (Shawaqfeh and
Developments in Solar Still Desalination Systems 2097

Farid, 1995). Adhikari et al. (1990) considered the influence of the charac-
teristic space of stills but ignored the fact that the relationship between the
evaporative and free convective heat transfer coefficients will change with
temperature. Kumar and Tiwari (1996) developed a thermal model for so-
lar distillation units based on linear regression analysis to determine these
coefficients for different Grashof number ranges without the limitations of
Dunkle’s model. They used the following relation:
mew
= C(Ra)n = C(Gr · Pr)n (13)
R
Which can be written in the form:
Downloaded by [American University of Beirut] at 02:29 14 April 2016

y = mx + Co (14)

Where:
  
K 3600
R = 0.016273(Pw − Pg ) (15)
d hfg

The coefficients m and Co can then be determined using regression analysis


(Tiwari and Tiwari, 2006). In order to develop improved heat and mass
transfer correlations for a wider range of Rayleigh number and temperature,
Hongfei et al. (2002) used the empirical relation proposed by Chen et al.
(1984):

Nu = Ra0.26 , 3.5 × 103 < Ra < 106 (16)

The following relation for the evaporation rate of solar stills could be devel-
oped and verified by experimenting on a multistage stacked tray solar still
(Hongfei et al., 2002):
hc
me = (ρw − ρg ) (17)
ρf Cpaf Le(1−n)

Koffi et al. (2009) modeled a solar still using an implicit numerical scheme,
the Diabolo-Sablier algorithm. The amount of distilled water generated highly
depended on the ambient temperature. It was noted that further experiments
with the particular conditions and simulated device are needed before sound
conclusions can be made. Tsilingiris et al. (2009) modeled solar stills with
the relaxation of some simplifying assumptions in Dunkle’s model and the
use of a modified form for the Grashof number in order to account for the
difference in water vapor content, which is lighter than dry air between two
regions of humid air mixture:
 
∗ gL3 ρg
Gr = 2 −1 (18)
ν ρs
2098 G. M. Ayoub and L. Malaeb

Where ρ s and ρ g refer to the saturated mixture density at the free brine
surface and the top condensing plate at temperatures Ts and Tg , respec-
tively. This expression is further simplified assuming ideal gas behavior and
a similar form to that of Dunkle’s relation for the convective heat transfer is
obtained with the numerical multipliers being 5.5% lower due to the con-
sideration of humid air thermophysical properties. The expression used for
the evaporative heat transfer assumes that the saturated vapor pressure at
the water surface and the cover is negligible compared to barometric pres-
sure and is valid for an average still temperature of around 50◦ C. Jubran
et al. (2000) modeled a three-stage still with an expansion nozzle and heat
recovery systems. They used a corrected latent heat of vaporization at the
condensation surface temperature Ts as presented by Cengel (1998):
Downloaded by [American University of Beirut] at 02:29 14 April 2016

h∗fg = hfg + 0.68Cp (Tsaturation − Ts ) (19)

The mass transfer coefficient used is calculated by substituting the diffusiv-


ity and Schmidt number into Sherwood and Grashoff number relations. In
comparing and modeling the productivity of single-effect and double-effect
solar stills, Al-Hinai et al. (2002b) noted an increase of 39% in average daily
output of the double-effect solar still over that of the single-effect still. In
their model, they adopted the formulations of Sartori (1996) and used the
following equations to calculate the heat transfer coefficient between the
cover and the air (Yellott, 1982):

hg = (NuL + Nuf )1/3 K/Lg (20)


NuL = 0.288Re0.731 Pr1/3 (21)
Nuth = 0.037Re4/5 Pr1/3 (22)
Nuf = NuL cos h + (1 − cos h)Nuth
2 2
(23)

Other researchers have treated the still as a black box and correlated the
incidence solar radiation with the distillate quantity produced during daytime
operation (Mathioulakis et al., 1999; Voropoulos et al., 2004):

Mout = f1d Hd + f2d (T − Ta ) + f3d (24)

A similar equation is also proposed for nighttime operation excluding the


first term, which accounts for solar radiation. The previous model was found
to reasonably simulate the system’s operation with a maximum deviation of
3% (Voropoulos et al., 2004).
Developments in Solar Still Desalination Systems 2099

6. COST OF SOLAR STILLS

The value of a solar still is ultimately judged on the basis of its economy
(Arjunan et al., 2009). Solar stills are characterized by their high initial costs
and low annual operating costs. If P is the initial investment in the system, i
the interest rate, n the number of useful years, and S the salvage value, then
the system annual cost is determined as (Arjunan et al., 2009, Kumar and
Tiwari, 2009):

Annnual Cost = First Annual Cost + Annual Maintenance Cost


− Annual Salvage Cost
Downloaded by [American University of Beirut] at 02:29 14 April 2016

i(1 + i)n
First Annual Cost = P (25)
(1 + i)n − 1
i
Annual Salvage Cost = S (26)
(1 + i)n − 1

Although research on improving the still productivity are abound, few have
fully addressed the economical aspects. Mukherjee and Tiwari (1986) ana-
lyzed the costs of different types of stills and concluded that roof type con-
crete solar distillation plants are the best choice for large-scale plants. Kumar
and Tiwari (2009) presented a complete life-cycle cost of a single-slope hy-
brid active solar still and estimated its energy payback time as 4.7 years
compared with 2.9 years for a passive still. The cost of distilled water varies
considerably depending on the system design and the year of investigation.
Reported values vary from $30/m3 for a large-size plant for pyramid-shaped
still (Fath et al., 2003), to $20/m3 for a large size plant (Ghoneyem and
Ileri, 1997), to $16.3/m3 for a cluster of 250 simple stills (Al-Hinai, 2002b),
to $2.4/m3 for 50m3 production using porous basin stills (Madani and Zaki,
1995) and down to $0.66/m3 for a system using the novel still described by
Kemp (2009). From this trend, it can be seen that further research can lead
to a water production cost that competes with other technologies. Larger
systems will help decrease the overhead costs of auxiliary components
(Arjunan et al., 2009). Kabeel et al. (2010) performed a cost analysis for differ-
ent still configurations and some of their results are shown in Table 7 along
with cost and productivity data from other studies. Table 7 also presents
cost data for small-scale reverse osmosis systems for comparison. It should
be emphasized that an accurate comparison in terms of desalination costs
and production is difficult. The prices and material performances depend
greatly on the place and year of the project implementation (Charcosset,
2009). It should be noted, however, that the production rate for solar stills
is proportional to the area of the still, which means that the cost per unit of
water produced is nearly the same, regardless of the installation size, which
makes them more attractive for small-sized applications. This is in contrast
Downloaded by [American University of Beirut] at 02:29 14 April 2016

2100
TABLE 7. Cost and productivity data for solar stills and renewable-energy RO

Average capacity Maximum capacity


Reference System description Cost ($/m3) (L/day) (L/day) Country

Delyannis and Delyannis Simple solar still 12 1 — Greece


(1985)
Woto (1987) Simple solar still 12.5 0.8 — Botswana
Yadav and Tiwari (1987) Simple solar still 12 — — India
Nandwani (1990) Simple solar still 12.53 6 — Costa Rica
Barrera (1992) Staircase solar still 10 9 — Mexico
Hasnain and Alajlan (1998) Simple solar still 2.95–5.67 Total 5.8 m3/day for the plant Saudi Arabia
El-Bahi and Inan (1999) Still with a solar pond 52 3.22 4.65 Egypt
Jubran et al. (2000) Multistage with heat recovery 5.63 — 9 Malaysia
Al-Hinai (2002b) Pyramid-shaped still 13.5 5.16 8.39 Oman
Fath et al. (2003) Single-Sloped still 35 5.81 8.39 Egypt
Fath et al. (2003) Pyramid-shaped still 31 5.9 8.52 Egypt
Abdel-Rahim and Lasheen Still with solar concentrator 58 3.8 5.5 Egypt
(2007)
Samee et al. (2007) Single-Sloped still 63 2.25 3.25 Pakistan
Abdallah et al. (2008) Stepped still with sun tracking 71 3.68 5.32 Jordan
Sadineni et al. (2008) Weir-type still 54 3.85 5.56 USA
Velmurugan et al. (2008b) With fins 54 2.77 4 India
Ismail et al. (2009) Transportable hemi-spherical 180 3.95 5.71 Canada
Kemp (2009) Still made of convoluted tubes 0.66 7 9 USA
Karagiannis and Soldatos Brackish water RO 5.6–12.9 Few cubic
(2008) meters/day
Karagiannis and Soldatos Seawater RO 1.5–18.75 2000–2400
(2008)
Al-Karaghouli et al. (2010) Brackish water RO 3.6–14.9 Small scale
Al-Karaghouli et al. (2010) Seawater RO 4.2–33 Small scale
Bourouni et al. (2011) Brackish water RO 2.7–6.7 Small scale
Bourouni et al. (2011) Seawater RO 1.8–27 Small scale
Developments in Solar Still Desalination Systems 2101

to fresh water supplies and to other desalination methods, where the capital
cost of equipment per unit of capacity decreases as the capacity increases
(Kalogirou, 2005).

7. CONCLUSIONS AND FUTURE NEEDS

As presented in this review, the effect of absorbers, cover cooling, brine


depth, initial water conditions and flow rate, and still geometry have sig-
nificant impacts on yield and efficiency. These factors should therefore be
taken into account in novel designs. Modeling of flow circulation within the
still and enhanced methods to estimate internal heat transfer coefficients are
Downloaded by [American University of Beirut] at 02:29 14 April 2016

also invaluable for better prediction of the still performance and for system
optimization. Promoting solar distillation further requires focusing research
efforts on improving existing technology to lower costs and developing more
compact installations that reduce land use. Of key importance are material
optimization, system reliability, design sustainability and longer life span, use
of hybrid systems, and the combination of the effectiveness of individual de-
velopments into one integrated system. Such improvements need to address
the enhancement of thermal storage and insulation, increasing the surface
area available for evaporation, incorporating thermo-optical properties of
the condensing covers, and developing novel geometrical setups to add to
the benefits of this old system from today’s high-tech developments. The
introduced modifications, however, should not defy the basic advantages
of the solar still in being simple, low cost, easy to handle, and requiring
little maintenance. The use of solar panels, collectors, ponds, condensers,
and other productivity-enhancing devices requires considerable space and
thus developing more compact systems with low cost is a main challenge.
Battery storage may be necessary to account for the intermittent nature of
solar radiation but this storage is only practical for small-scale plants, due to
the cost of batteries. The maturity of the introduced technologies should
also take into account the low level of infrastructure, often characteriz-
ing places with severe water stress and that lend themselves to solar stills
application.
Finally, the solar still remains to be the basic technique for the inexpen-
sive production of water and is a low-tech method that can be easily adopted
by local rural people. With the growing global oil crisis, the need for alter-
natives to conventional desalination plants based on fossil fuels grows. The
use of solar stills has so far been restricted to small-scale systems due to
their relatively low thermal efficiency and production rate compared with
the large areas required. They tend to be competitive, however, with other
renewable-desalination technologies in small-scale production due to their
relatively low cost and simplicity. Enhancing solar stills in order to eradicate
their limitations can therefore provide a means to attain self-reliance and
2102 G. M. Ayoub and L. Malaeb

ensure a regular water supply in areas where power is scarce, demand is


low, and salty water is the only available source. The potential for solar
stills strongly exists in developing areas of the world, particularly in loca-
tions where people live near sources of brackish water or seawater. These
areas often lack the expertise needed to operate small-scale systems based
on membrane technologies. The high salinity and low quality of the feed
water is not a concern in solar stills as it is in membrane technology. Mem-
brane systems such as those based on reverse osmosis are highly sensitive to
low-quality feed, seasonal variations, and sudden changes in pH, tempera-
ture, and constituent concentrations. These conditions entail increased areas
and costs to provide the pretreatment necessary for membranes. Frequent
membrane replacement and the need for proper cleaning to prevent fouling
Downloaded by [American University of Beirut] at 02:29 14 April 2016

is another issue. The brine produced by solar stills may also be of less en-
vironmental impact than that produced by other desalination technologies
since it is free from the chemicals, antiscalants, and cleaning waste streams
that constitute a necessary part in operating other desalination systems.

ACKNOWLEDGMENTS

The authors would like to acknowledge the Lebanese National Council for
Scientific Research and the Middle East Desalination Research Center, Oman,
for financially supporting this work.

NOMENCLATURE

C Thermal capacity, J/m2/K


Cp Specific heat, J/kg/K
Cpaf Specific heat capacity of humid air at average temperature,
J/kg/K
Cw Specific heat of water, J/kg/K
d brine depth, cm
D Thermal diffusivity, m2/s
f1d , f2d , f3d Coefficients of linear model for day water production
Gr Grashof number
g Gravitational acceleration, m/s2
hfg Latent heat of vaporization, J/kg
i Interest rate
I Daily radiation (insolation), MJ/m2/day
K Thermal conductivity, W/m/K
L Characteristic length scale of convection
Le Lewis number
mw Mass of basin water, kg
Developments in Solar Still Desalination Systems 2103

mew Distillate output, kg


Ma Molecular weight of dry air, kg/mol
Mw Molecular weight of water vapor, kg/mol
n Number of useful years for the solar still desalination system
Nu Nusselt number
P Initial investment in the system
pT Total pressure of mixture of air and water vapor in still, N/m2
pg Partial vapor pressure at condensation surface temperature,
N/m2
pw Partial vapor pressure at evaporation surface temperature, N/m2
Pr Prandtl number
S Salvation value
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Sc Schmidt number = Dν
Sh Sherwood number = 0.15(Gr × Sc)1/3
T Temperature, K
Ta Ambient temperature, K
y Daily productivity of the still, kg/m2/d
β Expansion factor
 Difference
εeff Effective emissivity
εw Emissivity of water
εg Emissivity of cover
µ Dynamic viscosity
ν Kinematic viscosity
ρ Density
ρv Density of partial pressure of vapor

REFERENCES

Abdallah, S., Abu-Khader, M., and Badran, O. (2009). Effect of various absorbing
materials on the thermal performance of solar stills. Desalination 242, 128–137.
Abdallah, S., and Badran, O. (2008). Sun tracking system for productivity enhance-
ment of solar still. Desalination 220, 669–676.
Abdallah, S., Badran, O., and Abu-Khader, M. (2008). Performance evaluation of a
modified design of a single slope solar still. Desalination 219, 222–230.
Abdel-Rahim, N. (2001). Utilization of new technique to improve the efficiency of
horizontal solar desalination still. Desalination 138, 121–128.
Abdel-Rahim, Z., and Lasheen, A. (2005). Improving the performance of solar de-
salination systems. Renewable Energy 30, 1955–1971.
Abdel-Rahim, Z., and Lasheen, A. (2007). Experimental and theoretical study of a
solar desalination system located in Cairo, Egypt. Desalination 217, 52–64.
Aboabboud, M., Mink, G., and Kudish, A. (2009). Condensation heat recycle in solar
stills. Paper presented at the World Congress on Engineering, London, England,
July 1–3.
2104 G. M. Ayoub and L. Malaeb

Abu-Arabi, M., and Venkat Reddy, K. (2005). R&D Challenges for further cost reduc-
tions in desalination. Paper presented at the MEDRC2 Conference, Middle East
Desalination Research Center, Amman, Jordan.
Abu-Arabi, M., and Zurigat, Y. (2005). Year-round comparative study of three types
of solar desalination units. Desalination 172, 137–143.
Abu-Hijleh, B., and Mousa, H. (1997). Water film cooling over the glass cover of a
solar still including evaporation effects. Energy 22, 43–48.
Adhikari, R., Kumar, A., and Kumar, A. (1990). Estimation of mass-transfer rates in
solar stills. International Journal of Energy Research 14, 737–744.
Aggarwal, S., and Tiwari, G. (1999). Thermal modeling of a double condensing
chamber solar still: an experimental validation. Energy Conversion and Man-
agement 40, 97–114.
Akash, B., Mohsen, M., and Nayfeh, W. (2000). Experimental study of the basin type
Downloaded by [American University of Beirut] at 02:29 14 April 2016

solar still under local climate conditions. Energy Conversion and Management
4, 1883–1890.
Akash, B., Mohsen, M., Osta, O., and Elayan, Y. (1998). Experimental evaluation of
a single-basin solar still using different absorbing materials. Renewable Energy
14, 307–310.
Al-Hayek, I., and Badran, O. (2004). The effect of using different designs of solar
stills on water distillation. Desalination 169, 121–127.
Al-Hinai, H., Al-Nassri, M., and Jubran, B. (2002a). Effect of climatic, design and
operational parameters on the yield of a simple solar still. Energy Conversion
and Management 43, 1639–1650.
Al-Hinai, H., Al-Nassri, M., and Jubran, B. (2002b). Parametric investigation of a
double-effect solar still in comparison with a single-effect solar still. Desalina-
tion 150, 75–83.
Al-Jubouri, A., and Khalifa, A. (1982). Passive and low energy alternatives. The First
International Plea Conference, Bermuda, September 13–15.
Al-Karaghouli, A., Renne, D., and Kazmerski, L. (2010). Technical and economic
assessment of photovoltaic-driven desalination systems. Renewable Energy 35,
323–328.
Al-Kharabsheh, S., and Goswami, Y. (2003). Experimental study of an innovative
solar water desalination system utilizing a passive vacuum technique. Solar
Energy 75, 395–402.
Arjunan, T., Aybar, H., and Nedunchezhian, N. (2009). Status of solar desalination
in India. Renewable and Sustainable Energy Reviews 13, 2408–2418.
Aybar, H., Egelioglu, F., and Atikol, U. (2005). An experimental study on an inclined
solar water distillation system. Desalination 180, 285–289.
Ayoub, G., El-Fadel, M., Acra, A., and Abdallah, R. (2000). Critical density index for
the solar production of bittern from seawater. International Journal of Environ-
mental Studies 58, 85–97.
Ayoub, J., and Alward, R. (1996). Water requirements and remote arid areas: The
need for small-scale desalination. Desalination 107, 131–147.
Badran, O. (2007). Experimental study of the enhancement parameters on a single
slope solar still productivity. Desalination 209, 136–143.
Barrera, E. (1992). A technical and economical analysis of a solar water still in
Mexico. Renewable Energy 2, 489–495.
Developments in Solar Still Desalination Systems 2105

Bassam, A., Hamzeh, A., and Rababah, M. (2003). Experimental study of a so-
lar still with sponge cubes in basin. Energy Conversion and Management 44,
1411–1418.
Bejan, A. (1995). Convection heat transfer (2nd ed.). New York: Wiley.
Bilal, A., Mousa, S., Omar, O., and Yaser, E. (1998). Experimental evaluation of a
single-basin solar still using different absorbing materials. Paper presented at
the 6th Arab International Solar Energy Conference, Bahrain.
Bouchekima, B., Gros, B., Ouahes, R., and Diboun, M. (1998). Performance study
of the capillary film solar distiller. Desalination 116, 185–192.
Boukar, M., and Harmim, A. (2004). Parametric study of a vertical solar still under
desert climatic conditions. Desalination 168, 21–28.
Boukar, M., and Harmin, A. (2005). Performance evaluation of a one-sided vertical
solar still tested in the Desert of Algeria. Desalination 183, 113–126.
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Bourouni, K., Ben M’Barek, T., and Al Taee, A. (2011). Design and optimization of
desalination reverse osmosis plants driven by renewable energies using genetic
algorithms. Renewable Energy 36, 936–950.
Cengel, Y. (1998). Heat transfer: A practical approach. New York: McGraw-Hill.
Chandra, D., Reddy, M., Sjithachandra, H., and Sabbenwall, S. (1980). Preliminary
study of an inclined stepped wick-type solar still with condenser. National Solar
Energy Convention, Indian Institute of Technology, 123–127.
Charcosset, C. (2009). A review of membrane processes and renewable energies for
desalination. Desalination 245, 214–231.
Chen, Z., Ge, X., Sun, X., Bar, L., and Miao, Y. (1984). Natural convection heat transfer
across air layers at various angles of inclination. Engineering Thermophysics
(Special Issue for the U.S.-China, Bination Heat Transfer Workshop) 211–220.
Cooper, P. (1969). Digital simulation of transient solar still processes. Solar Energy
12, 313–331.
De Paul, I. (2002). New model of a basin-type solar still. Journal of Solar Energy
Engineering 124, 311–314.
De Paul, I. (2009). Experimental Evidence of chaotic heat enhancement in a still.
Applied Thermal Engineering 29, 1840–1845.
Delyannis, E., and Delyannis, A. (1985). Economics of solar stills. Desalination 52,
167–176.
Dhiman, N. (1988). Transient analysis of a spherical solar still. Desalination 69,
47–55.
Dhiman, N., and Tiwari, G. (1990). Effect of water flowing over the glass cover of a
multiwick solar still. Energy Conversion and Management 30, 245–250.
Dimri, V., Sarkar, B., Singh, U., and Tiwari, G. (2008). Effect of condensing cover
material on yield of an active solar still: An experimental validation. Desalination
227, 178–189.
Dunkle, R. (1961). Solar water distillation system: the roof type still and a multi-
ple effect diffusion still: International developments in heat transfer ASME. In
Proceedings of International Heat Transfer Part V (pp. 895–902). Boulder, CO:
University of Colorado.
Dutt, D., Kumar, A., Anand, J., and Tiwari, G. (1989). Performance of a double-basin
solar still in the presence of dye. Applied Energy 32, 207–223.
2106 G. M. Ayoub and L. Malaeb

Dutt, D., Kumar, A., Anand, J., and Tiwari, G. (1993). Improved design of a double
effect solar still. Energy Conversion and Management 34, 507–517.
El-Bahi, A., and Inan, D. (1999). Analysis of a parallel double glass solar still with
separate condenser. Renewable Energy 17, 509–521.
El-Sebaii, A. (2000). Effect of wind speed on some designs of solar stills. Energy
Conversion and Management 41, 523–538.
El-Sebaii, A. (2005). Thermal performance of a triple-basin solar still. Desalination
174, 23–37.
El-Sebaii, A. A., Aboul-Enein, S., and El-Bialy, E. (2000). Single basin solar still
with baffle suspended absorber. Energy Conversion and Management 41, 661–
675.
Eltawil, M., Zhengming, Z., and Yuan, L. (2009). A review of renewable energy
technologies integrated with desalination systems. Renewable and Sustainable
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Energy Reviews 13, 2245–2262.


Fatani, A., Zaki, G., and Al-Turki, A. (1994). Improving the yield of simple basin
solar stills by passively cooled condensers. Renewable Energy 4, 377–386.
Fath, H. (1998). Solar distillation: A promising alternative for water provision with
free energy, simple technology and a clean environment. Desalination 116,
45–56.
Fath, H., and Ghazy, A. (2002). Solar desalination using humidification-
dehumidification technology. Desalination 142, 119–133.
Fath, H., El-Samanoudy, M., Fahmy, K., and Hassabou, A. (2003). Thermal-economic
analysis and comparison between pyramid-shaped and single slope solar still
configurations. Desalination 159, 69–79.
Fath, H., and Hosny, H. (2002). Thermal performance of a single-sloped basin still
with an inherent built-in additional condenser. Desalination 142, 19–27.
Franceschetti, P. (2009). Improvement of a solar still by suction of the wet air. Paper
presented at the IDA World Congress, Atlantis, the Palm, Dubai, UAE.
Garcia-Rodriguez, L. (2002). Seawater desalination driven by renewable energies: A
review. Desalination 143, 103–113.
Garg, H., and Mann, H. (1976). Effect of climatic, operational and design parameters
on the year round performance of single-sloped and double-sloped solar still
under Indian arid zone conditions. Solar Energy 18, 159–163.
Ghoneyem, A., and Ileri, A. (1997). Software to analyze stills and an experimental
study on the effects of the cover. Desalination 114, 37–44.
Gude, V., and Nirmalakhandan, N. (2010). Sustainable desalination using solar en-
ergy. Energy Conversion and Management 51, 2245–2251.
Gupta, R., Rai, S., and Tiwari, G. (1988). Transient analysis of double basin so-
lar still intermittent flow of waste hot water in night. Energy Conversion and
Management 28, 245–249.
Haddad, O., Al-Nimr, M., and Maqableh, A. (2000). Enhanced solar still performance
using a radiative cooling system. Renewable Energy 21, 459–469.
Hajri, I., Omri, A., and Nasrallah, S. (2007). A numerical model for the simulation of
double-diffusive natural convection in a triangular cavity using equal order and
control volume based on the finite element method. Desalination 206, 579–
588.
Developments in Solar Still Desalination Systems 2107

Hasnain, S., and Alajlan, S. (1998). Coupling of PV-Powered R.O. brackish water
desalination plant with solar stills. Renewable Energy 14, 281–286.
He, T., and Yan, L. (2009). Application of alternative energy integration technology
in seawater desalination. Desalination 249, 104–108.
Hongfei, Z., Xiaoyan, Z., Jing, Z., and Yuyuan, W. (2002). A group of improved heat
and mass transfer correlations in solar stills. Energy Conversion and Manage-
ment 43, 2469–2478.
Ismail, B. (2009). Design and performance of a transportable hemispherical solar
still. Renewable Energy 34, 145–150.
Jamal, M. (1991). A step forward towards an ideal absorber for solar energy. Inter-
national Journal of Energy Resources 15, 367–376.
Janarthanan, B., Chandrasekaran, J., and Kumar, S. (2005). Evaporative heat loss and
heat transfer for open- and closed-cycle systems of a floating tilted wick solar
Downloaded by [American University of Beirut] at 02:29 14 April 2016

still. Desalination 180, 291–305.


Jegatheesan, V., Liow, J., Shu, L., Kim, S., and Visvanathan, C. (2009). The need for
global coordination in sustainable development. Journal of Cleaner Production
17, 637–643.
Jubran, B., Ahmed, M., Ismail, A., and Abakar, Y. (2000). Numerical modeling
of a multi-stage solar still. Energy Conversion and Management 41, 1107–
1121.
Kabeel, A., Hamed, A., and El-Agouz, S. (2010). Cost analysis of different solar still
configurations. Energy 35, 2901–2908.
Kalogirou, S. (2005). Seawater desalination using renewable energy sources. Progress
in Energy and Combustion Science 31, 242–281.
Kamal, W. (1988). A theoretical and experimental study of the basin-type solar
still under the Arabian-gulf climatic conditions. Solar and Wind Technology 5,
147–157.
Karagiannis, I., and Soldatos, P. (2008). Water desalination cost literature: review
and assessment. Desalination 223, 448–456.
Kemp, H. (2009). Solar still breakthrough: New top competitive desalination uses
95% renewable energy. Paper presented at the IDA World Congress, Atlantis,
the Palm, Dubai, UAE.
Khalifa, A., and Hamood, A. (2009a). Performance correlations for basin type solar
stills. Desalination 249, 24–28.
Khalifa, A., and Hamood, A. (2009b). On the verification of the effect of water depth
on the performance of basin type solar stills. Solar Energy 83, 1312–1321.
Khawaji, A., Kutubkhanah, I., and Wie, J. (2008). Advances in seawater desalination
technologies. Desalination 221, 47–69.
Koffi, B., Konan, D., N’guessan, R., Saraka, J., Tanoh, A., Kouacou, M., Yeo, Z., Koffi,
M., and Koua, A. (2009). Modeling of solar still for production of pure water in
the Abidjan Zones. Research Journal of Physics 3, 5–13.
Koning, J., and Thiesen, S. (2005). Aqua Solaris: An optimized small scale desali-
nation system with 40 litres output per square meter based upon solar-thermal
distillation. Desalination 182, 503–509.
Kumar, A., and Anand, J. (1992). Modeling and performance of a tubular multiwick
solar still. Energy Conversion and Management 17, 1067–1071.
2108 G. M. Ayoub and L. Malaeb

Kumar, A., Anand, J., and Tiwari, G. (1991). Transient analysis of a double slope-
double basin solar distiller. Energy Conversion and Management 31, 129–
139.
Kumar, A., and Tiwari, G. (1990). Use of waste hot water in double slope solar still
through heat exchanger. Energy Conversion and Management 30, 81–89.
Kumar, S., and Tiwari, G. (1996). Estimation of convective mass transfer in solar
distillation systems. Solar Energy 57, 459–464.
Kumar, S., and Tiwari, G. (1998). Optimization of collector and basin areas for a
higher yield for active solar stills. Desalination 116, 1–9.
Kumar, S., and Tiwari, G. (2009). Life cycle cost analysis of single slope hybrid
(PV/T) active solar still. Applied Energy 86, 1995–2004.
Lawrence, S., and Tiwari, G. (1990). Theoretical evaluation of solar distillation under
natural circulation with heat exchanger. Energy Conversion and Management
Downloaded by [American University of Beirut] at 02:29 14 April 2016

30, 205–213.
Lessing, L. (1990). A small modular high-efficiency solar still. Solar Energy 9, 35–46.
Levi, E. (1983). A universal Strouhal Law. ASCE Journal of Engineering Mechanics
109, 718–727.
Madani, A., and Zaki, G. (1995). Yield of solar stills with porous basins. Applied
Energy 52, 273–281.
Mahmoudi, H., Abdellah, O., and Ghaffour, N. (2009). Capacity building strategies
and policy for desalination using renewable energies in Algeria. Renewable and
Sustainable Energy Reviews 13, 921–926.
Mathioulakis, E. K., and Belessiotis, V. (2003). Integration of solar still in a multi-
source, multi-use environment. Solar Energy 75, 403–411.
Mathioulakis, E. K. Voropoulos, K., and Belessiotis, V. (1999). Modeling and predic-
tion of long-term performance of solar stills. Desalination 122, 85–93.
Mathioulakis, E. K., Belessiotis, V., and Delyannis, E. (2007). Desalination by using
alternative energy: Review and state of the art. Desalination 203, 346–365.
Medeazza, M. (2005). Direct and socially-induced environmental impacts of desali-
nation. Desalination 185, 57–70.
Minasian, A., and Al-Karaghouli, A. (1995). An improved solar still: The Wick-Basin
type. Energy Conversion and Management 36, 213–217.
Mousazadeh, H., Kehyani, A., Javadi, A., Mobli, H., Abrinia, K., and Sharifi, A. (2009).
A review of principle and sun-tracking methods for maximizing solar systems
output. Renewable and Sustainable Energy Reviews 13, 1800–1818.
Mukherjee, K., and Tiwari, G. (1986). Economic analyses of various designs of
conventional solar stills. Energy Conversion and Management 26, 155–157.
Muller-Holst, H., Engelhardt, M., and Scholkopf, W. (1999). Small-scale thermal sea-
water desalination simulation and optimization of system design. Desalination
122, 255–262.
Murugavel, K., Chockalingam, K., and Srithar, K. (2008). Progresses in improving the
effectiveness of the single basin passive solar still. Desalination 220, 677–686.
Nafey, A., Abdelkader, M., Abdelmotalip, A., and Mabrouk, A. (2000). Parame-
ters affecting solar still productivity. Energy Conversion and Management 41,
1797–1809.
Nafey, A., Abdelkader, M., Abdelmotalip, A., and Mabrouk, A. (2001). Solar still
productivity enhancement. Energy Conversion and Management 42, 1401–1408.
Developments in Solar Still Desalination Systems 2109

Nafey, A., Abdelkader, M., Abdelmotalip, A., and Mabrouk, A. (2002). Enhancement
of solar still productivity using floating perforated black plate. Energy Conver-
sion and Management 43, 937–946.
Naim, M., and Abd El Kawi, M. (2003). Non-conventional solar stills. Part 1. Non-
conventional solar stills with charcoal particles as absorber medium. Desalina-
tion 153, 55–64.
Nandwani, S. (1990). Economic analysis of domestic solar still in the climate of Costa
Rica. Solar & Wind Technology 7, 219–227.
Nassar, Y., Yousif, S., and Salem, A. (2007). The second generation of the solar
desalination systems. Desalination 209, 177–181.
Nijmeh, S., Odeh, S., and Akash, B. (2005). Experimental and theoretical study of
a single-basin solar still in Jordan. International Communications in Heat and
Mass Transfer 32, 565–572.
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Obikane, Y., and Al-Bilbisi, H. (2009). A detail design method of triangular cell
multi-stage solar desalination still with wind effects. Paper presented at the IDA
World Congress, Atlantis, the Palm, Dubai, UAE.
Obikane, Y., Ikeo, S., and Al-Bilbisi, H. (2009). A geometrical modification of multi-
stage solar desalination still. Paper presented at the IDA World Congress, At-
lantis, the Palm, Dubai, UAE.
Okeke, C., Egarievwe, S., and Animalu, A. (1990). Effects of coal and charcoal on
solar-still performance. Energy 15, 1071–1073.
Omri, A., Orfi, J., and Nasrallah, S. (2005). Natural convection effects in solar stills.
Desalination 183, 173–178.
Phadatare, M. K., and Verma, S. K. (2007). Influence of water depth on internal heat
and mass transfer in a plastic solar still. Desalination 217, 267–275.
Porta-Gandara, M., Cervantes, J., and Solorio, F. (2004). Periodic enclosed natural
convection in a laboratory solar still. Experiments in Fluids 37, 483–487.
Qiblawey, H., and Banat, F. (2008). Solar thermal desalination technologies. Desali-
nation 220, 633–644.
Rai, S., Dutt, D., and Tiwari, G. (1990). Some experimental studies of a single basin
solar still. Energy Conversion and Management 30, 149–53.
Rai, S., and Tiwari, G. (1983). Single basin solar still coupled with flat plate collector.
Energy Conversion and Management 23, 145–149.
Rajvanshi, A. (1981). Effect of various dyes on solar distillation. Solar Energy 27,
51–65.
Reali, M., and Modica, G. (2008). Solar stills made with tubes for sea water desalting.
Desalination 220, 626–632.
Reddy, M., Chandra, D., Sehgal, H., Sabberwal, S., Bhargava, A., Kumar, A., and
Chandra, D. (1983). Performance of a multiple-wick solar still with condenser.
Applied Energy 13, 15–21.
Rheinlaender, J., and Graeter, F. (2001). Technologies for desalination of typically
10 m3 of water per day. Desalination 139, 393–397.
Rubio-Cerda, E., Fernandez, J., and Porta-Gandara, M. (2004). Modeling thermal
asymmetries in double slope solar still. Renewable Energy 29, 895–906.
Sadineni, S., Hurt, R., Halford, C., and Boehm, R. (2008). Theory and experimental
investigation of a weir-type inclined solar still. Energy 33, 71–80.
2110 G. M. Ayoub and L. Malaeb

Sakthivel, M., and Shanmugasundaram, S. (2008). Effect of energy storage medium


(black granite gravel) on the performance of a solar still. International Journal
of Energy Research 32, 68–82.
Samee, M., Mirza, U., Majeed, T., and Ahmad, N. (2007). Design and performance
of a simple single basin solar still. Renewable and Sustainable Energy Reviews
11, 543–549.
Sartori, E. (1996). Solar still vs. evaporator: A comparative study between their ther-
mal behaviors. Solar Energy 56, 199–206.
Schwarzer, K., Vieira, M., Faber, C., and Muller, C. (2001). Solar thermal desalination
system with heat recovery. Desalination 137, 23–29.
Shawaqfeh, A., and Farid, M. (1995). New development in the theory of heat and
mass transfer in solar stills. Solar Energy 55, 527–535.
Shukla, A., and Sorayan, V. (2005). Thermal modeling of solar stills: an experimental
Downloaded by [American University of Beirut] at 02:29 14 April 2016

validation. Renewable Energy 30, 683–699.


Singh, A., and Tiwari, G. (1992). Performance study of double effect distilla-
tion in a multi-wick solar still. Energy Conversion and Management 33, 207–
214.
Singh, A., Tiwari, G., Sharma, P., and Khan, E. (1995). Optimization of orienta-
tion for higher yield of solar still for a given location. Energy Conversion and
Management 36, 175–187.
Singh, H., and Tiwari, G. (2004). Monthly performance of passive and active solar
stills for different Indian climatic conditions. Desalination 168, 145–150.
Smyth, M., Strong, A., Byers, W., and Norton, B. (2002). Performance evaluation of
several passive solar stills. Proceedings of Second Users Workshop 207–214.
Sodha, M., Kumar, A., and Tiwari, G. (1981a). Utilization of waste hot water for
distillation. Desalination 37, 325–342.
Sodha, M., Kumar, A., Tiwari, G., and Pandey, G. (1980). Effects of dye on the
performance of a solar still. Applied Energy 7, 147–162.
Sodha, M., Kumar, A., Tiwari, G., and Tyagi, R. (1981b). Simple multiple wick solar
still: analysis and performance. Solar Energy 26, 127–131.
Suneja, S., and Tiwari, G. (1998). Optimization of number of effects for higher yield
from an inverted absorber solar still using the Runge-Kutta method. Desalina-
tion 120, 197–209.
Suneja, S., and Tiwari, G. (1999). Effect of water flow on internal heat transfer solar
distillation. Energy Conversion and Management 40, 509–518.
Tanaka, H., and Nakatake, Y. (2007). Numerical analysis of the vertical multiple-
effect diffusion solar still coupled with a flat plate reflector: Optimum reflector
angle and optimum orientation of the still at various seasons and locations.
Desalination 207, 167–178.
Tanaka, H., and Nakatake, Y. (2009a). Increase in distillate productivity by inclining
the flat plate external reflector of a tilted-wick solar still in winter. Solar Energy
83, 785–789.
Tanaka, H., and Nakatake, Y. (2009b). One step azimuth tracking tilted-wick solar
still with a vertical flat plate reflector. Desalination 235, 1–8.
Tanaka, T., Yamashita, A., and Watanabe, K. (1981). Experimental and analytical
study of the tilted wick type solar still. Proceedings of International Solar Energy
Congress 2, 1087.
Developments in Solar Still Desalination Systems 2111

Tiris, C., Tiris, M., and Türe, I. (1996). Improvement of basin type solar still per-
formance: Use of various absorber materials and solar collector integration.
Renewable Energy 9, 758–761.
Tiwari, A., and Tiwari, G. (2006). Effect of water depths on heat and mass transfer
in a passive solar still: In summer climatic condition. Desalination 195, 78–94.
Tiwari, A., and Tiwari, G. (2007). Thermal modeling based on solar fraction and
experimental study of the annual and seasonal performance of a single slope
passive solar still: The effect of water depths. Desalination 207, 184–204.
Tiwari, G. (1985). Enhancement of daily yield in a double basin solar still. Energy
Conversion and Management 25, 49–50.
Tiwari, G., Kupfermann, A., and Aggarwal, S. (1997). A new design for a double
condensing chamber solar still. Desalination 114, 153–164.
Tiwari, G., and Madhuri, G. (1987). Effect of water depth on daily yield of the still.
Downloaded by [American University of Beirut] at 02:29 14 April 2016

Desalination 61, 67–75.


Tiwari, G., Madhuri, G., and Garg, H. (1985). Effect of water flow over the glass
cover of a single basin solar still with an intermittent flow of waste hot water in
the basin. Energy Conversion and Management 25, 315–322.
Tiwari, G., Mukherjee, K., Ashok, K., and Yadav, Y. (1986). Comparison of various
designs of solar stills. Desalination 60, 191–202.
Tiwari, G., and Rao, B. (1984). Transient performance of a single basin solar still
with water flowing over the glass cover. Desalination 49, 231–241.
Tiwari, G., and Selim, M. (1984). Double slope fiber reinforced plastic (FRP) multi
wick solar still. Solar and Wind Technology 1, 229–235.
Tiwari, G., Sharma, S., and Sodha, M. (1984). Performance of a double condensing
multiple wick solar still. Energy Conversion and Management 24, 155–159.
Tiwari, G., Shukla, S., and Singh, I. (2003). Computer modeling of passive/active
solar stills by using inner glass temperature. Desalination 154, 171–185.
Tiwari, G., and Suneja, S. (1998). Performance evaluation of an inverted absorber
solar still. Energy Conversion and Management 39, 173–180.
Tripathi, R., and Tiwari, G. (2004). Performance evaluation of a solar still by using
the concept of solar fractionation. Desalination 169, 69–80.
Tripathi, R., and Tiwari, G. (2005). Effect of water depth on internal heat and mass
transfer for active solar distillation. Desalination 173, 187–200.
Tripathi, R., and Tiwari, G. (2006). Thermal modeling of passive and active solar
stills for different depths of water by using the concept of solar fraction. Solar
Energy 80, 956–967.
Tsilingiris, P. (2009). Analysis of the heat and mass transfer processes in solar stills:
The validation of a model. Solar Energy 83, 420–431.
Valsaraj, P. (2002). An experimental study on solar distillation in a single slope basin
still by surface heating the water mass. Renewable Energy 25, 607–12.
Velmurugan, V., Deenadayalan, C., Vinod, H., and Srithar, K. (2008b). Desalination
of effluent using fin type solar still. Energy 33, 1719–1727.
Velmurugan, V., Gopalakrishnan, M., Raghu, R., and Srithar, K. (2008a). Single basin
solar still with fin for enhancing productivity. Energy Conversion and Manage-
ment 49, 2602–2608.
2112 G. M. Ayoub and L. Malaeb

Velmurugan, V., Kumaran, S., Prabhu, V., and Srithar, K. (2008c). Productivity en-
hancement of stepped solar still: Performance analysis. Thermal Science 12,
153–163.
Velmurugan, V., and Srithar, K. (2007). Solar stills integrated with a mini solar pond:
Analytical simulation and experimental validation. Desalination 216, 232–241.
Voropoulos, K., Mathioulakis, E., and Belessiotis, V. (2004). A hybrid solar desalina-
tion and water heating system. Desalination 164, 189–195.
World Energy Council. (2000). Renewable energy in South Asia: Status and prospects.
London: WEC.
Woto, T. (1987). The experience with small-scale desalinators for remote area dwellers
of the Kalahari Botswana. Kanye, Botswana: Rural Industrial Promotion/Rural
Industries Innovation Center.
Yadav, Y. (1991). Analytical performance of a solar still integrated with a flat plate
Downloaded by [American University of Beirut] at 02:29 14 April 2016

solar collector: thermo-siphon mode. Energy Conversion and Management 31,


255–263.
Yadav, Y., and Kumar, A. (1991). Transient analytical investigations on a single basin
solar still with water flow in the basin. Energy Conversion and Management 31,
27–38.
Yadav, Y., and Tiwari, G. (1987). Monthly comparative performance of solar stills of
various designs. Desalination 67, 565–578.
Yellott, J. (1982). Passive and hybrid cooling research. In Advances in Solar Energy
(Vol. 1, pp. 241–263). New York: Plenum Press.
Zaki, G., Al-Turki, A., and Al-Fatani, M. (1992). Experimental investigation on
concentrator-assisted solar stills. International Journal of Sustainable Energy
11, 193–199.

View publication stats

You might also like