Next Article in Journal
Choice Behavior of Autonomous Vehicles Based on Logistic Models
Previous Article in Journal
Sustainability as a Key Factor in Tourism Competitiveness: A Global Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of the Use of Solar Desalination Distillers to Produce Fresh Water in Arid Areas

ISEP—School of Engineering, Polytechnic of Porto, 4200-072 Porto, Portugal
*
Author to whom correspondence should be addressed.
Sustainability 2020, 12(1), 53; https://doi.org/10.3390/su12010053
Submission received: 27 November 2019 / Revised: 11 December 2019 / Accepted: 17 December 2019 / Published: 19 December 2019

Abstract

:
Water is an important resource for human beings, yet there are inhabited places tormented by the scarcity of it. The present study is concerned with places where, seemingly, the best way to get water is through solar distillers. These places should have, typically, high values of solar irradiation and a lack of human and economic resources to build and operate complex equipment. A set of sites scattered around the world was chosen, and then the presumed productivity and thermal efficiency that solar distillers would have if they were installed at these places was calculated. The mathematical model used with this purpose assumes steady state operation; the values of mass of water distilled and distiller efficiency were calculated for every hour, but the results presented are annual averages. Then, an economic study was made based on local costs of construction materials for the distillers, the work force, and the prices of water to predict the payback time of solar distillers. Finally, a study on environmental impact, particularly in terms of greenhouse gas (GHG) emissions, was made to compare reverse osmosis (RO) with solar distillation. For the sites studied, typical values of annual water output are in the range of 414 dm3/m2, for Évora, up to 696 dm3/m2, for Faya Largeau; the minimum efficiency was found for Évora, as 11.5%, and the maximum efficiency was found for Tessalit, as 15.2%. Payback times are very high, regardless of the areas of the globe where solar distillers are implanted. Regarding the GHG emissions, solar distillation is preferable to RO.

1. Introduction

1.1. Overview

Needless to say, water is of paramount importance to almost all forms of life. Some think tanks concerned with geostrategic matters have even predicted that in the near future, wars will be waged because of lack of water.
The regions of the planet already distressed by water shortage are located in vast deserts in the north of Africa, Namibia, the outback of Australia, central Asia, North and South Americas, and some tiny islands in the middle of oceans. Moreover, because of climate change such places are appearing with increasing frequency.
It is not simply water that worries human beings, since water, namely, brackish water, is abundant and widespread; what worries human beings is the scarcity of fresh water. Yet, there are several ways to get fresh water from brackish water.
Desalination through distillation is a way to obtain fresh water from brackish or salted water. There are several methods to desalinate water, grouped by the authors of [1] as thermal and non-thermal; the current paper focuses only on the solar distillers, which follow a thermal method. The reason for this choice stems on the obvious link between water shortage and poverty, and so it seems realistic to desalinate water through simple and cheap methods, resorting to long-lasting facilities with almost no maintenance requirements, constructed with materials found locally; a solar distiller may last up to 25 years without major repairs [2].
Ultimately, a solar desalination distiller consists of a cover made of transparent material, placed over a basin filled with salted water (see Figure 1). According to Belessiotis et al. [1], the cover is normally roof-shaped, double-shaped, and symmetric, but can also be asymmetric or single sloped, and there can be two covers. Solar radiation successively crosses the cover; energy is absorbed by the water in the basin; water evaporates; and vapor rises to the underside of the cover, cooler than the upper side of the cover, condenses, and then slides, preferably without falling back into the basin, to the gutters bordering the distiller. The productivity of such installation is assessed by the mass of fresh water yielded by unit area of the distiller and unit of time.
The cover is usually of glass, and less often of plastic films. Glass covers have the drawback of being breakable and, according to the authors of [3], should have a hematite content lower than 0.01%. Plastic covers can be made of treated polyvinylchloride and tetraphalate-polyethylene; their durability is less than 15 years, and their adaptability to any desired shape of the distiller is the highest of their class [1]. Phadatare and Verma [4] found that the productivity of distillers covered with glass is 30% to 35% higher than the productivity for distillers covered with Plexiglas.
The slope of the cover must guarantee that the condensed droplets of water, due to their surface tension, do not fall back into the basin but slide to the gutters, to be collected; this slope is normally under 20° [2]. The gutters have a slight slope too, just enough to prompt the flow of the distilled water into a reservoir that feeds consumers of such water. If the basin is shallow, its depth ranges from 10 to 20 mm; if the basin is deep, its depth goes up to 100 mm [4]. The widths are around 2 m and the lengths can reach 100 m [4].
The floor of the basin bottom liner must be waterproof, free of toxic components, withstand temperatures up to 90 °C, have a thermal absorptivity above 0.95 (see Table 1), and a polished surface. From Belessiotis et al. [1], it is normally made of one of these materials: concrete, wood impregnated with epoxy resins, aluminum, magnesium alloys, hard plastic ultraviolet resistant as Plexiglas, etc. To increase its absorption of solar heat, the upper side of the floor of the basin is covered by one of these materials: black paint (without volatile components that could evaporate with the water), spongy materials, butyl rubber, black polyethylene, asphalt, jute, etc.
The frame that supports the glass/plastic cover has to be robust and cling to the ground to withstand inclement weather, such as sand storms, but some of its parts must be removable to allow cleaning and repairs in the distiller.
The bottom and sidewalls of the basin are insulated from the ground to lower, as much as possible, the heat losses from the distiller. The most suitable insulation materials have low conductivity, are easily applicable to this kind of basin, withstand the operation of the distiller without deformations, do not release chemicals harmful into water up to 90 °C, and are waterproof. Pumps, storage tanks, pipes, fittings, etc., are considered ancillary components. Pipes and fittings can be of the same material, such as polyvinyl chloride (PVC) or any other plastic material not damaged by temperatures up to 90 °C; pipes and fittings either of galvanized steel or copper for brackish water, and stainless steel for distilled water, can be used as well. Normally, plastic materials are cheaper than stainless steel.
The sealing between cover and basin is necessary because the volume limited by the surface of the water, the underside of the cover, and the tops of the distiller must be air tight to reduce convection losses. Good sealing materials are easily applied, non-toxic, and have very low absorptivity.
The maintenance requirements of such distillers are cheap and easy. The maintenance consists mostly in removing leaves, dust, and sand from the cover; occasionally glasses break, and then have to be replaced. Some parts must be removable to allow cleaning and repairs in the basin.
Such distillers must endure bad weather conditions, such as strong winds and sand storms.
Solar desalination distillers operate in either batch or continuous mode. In batch mode, the amount of water evaporated from the basin during one day is offset, at the next sunrise, by new salt water. To avoid scaling or algae, every two or three days the basin is emptied and then refilled with new saltwater. As the temperature of the water in the basin increases, the productivity of the distiller increases; this is the reason why during the morning and the night such productivity is low but near noon it is at a maximum. Besides, shallow basins do not store sensible heat as do deep distillers. The productivity of distilled water by shallow distillers is in phase with the value of irradiation received by the distiller and so, from sunset to sunrise, distillation stops; however, as deep distillers store heat, distillation extends after sunset, although at low rate. In continuous mode, the basin is continuously fed by saltwater in a way to keep the amount of water in the basin constant.

1.2. Sites with Distillation Distillers

There is a long record of facilities to desalinate water through solar distillation. Some of these plants are shown in in Table 1.
In 1872, such a desalination plant was built at Chacabuco, (Las Salinas, Chile), which is described in detail by the authors of [5,6,7]. It operated for three decades to provide drinking water for animals used in nitrate mining [5]. Additionally, at Daytona Beach (Florida, USA), a similar facility was built, operated by Battele Memorial Institute of Cleveland [8,9,10]. In addition, numerous studies have been published, such as that by the authors of [11,12,13].
The Table 2 contains some sites, scattered around the world, which are seemingly suitable to solar harnessing facilities in general. The purpose of the present work is to assess the performance of specific distillation plants eventually set up on such sites.

2. Analyses

The solar distillation process to obtain water was analyzed in three respects: thermal efficiency, economics, and carbon emissions.

2.1. Thermal Efficiency

Belessiotis et al., Cooper, Duffie, and Beckman [1,2,5], among others, proposed two energy equations, one for the basin and another for the cover, with almost the same terms.
In the current study, the energy equation applied to the basin is
G τ c α w Q e v a p Q r a d ( b c ) Q c o n v ( b c ) Q g r d = ( M c / A ) b d T b d t ,
where the subscripts c, w, b, and grd stand for cover, water, basin, and ground, respectively; the energy equation applied to the cover, neglecting the variation of internal energy and the solar energy absorbed by it, is
Q e v a p + Q r a d ( b c ) + Q c o n v ( b c ) = Q c o n v ( c a ) + Q r a d ( c a ) ,
where the subscript a stands for the atmosphere. In the right-hand side of Equation (2), the sky is assumed as a plate.
Normally, for shallow and well-insulated basins, a steady-state operation can be assumed and, accordingly, the right-hand side of Equation (1) can be neglected.
According to Cooper [34], the total (all the wavelengths) and directional (angles of incidence of 0°, 30°, 45°, and 60°) values of absorptivity, transmissivity, and reflectivity for the glass cover, the water in the basin and the basin liner are given at Table 3. The angle of incidence, θ, is measured from the normal to the surface.
Dunkle [35] proposes that
Q r a d ( b c ) = 0.9 σ ( T b 4 T c 4 ) ,
assuming that both the water surface and the lower side surface of the cover are diffuse and gray; the value of 0.9 includes the view factor between these surfaces and their emissivities.
For the evaporation
Q e v a p = 9.15 × 10 7 h c ( p w b p w c ) h f g ,
with
h c = 0.884 { ( T b T c ) + [ ( p w b p w c ) / ( 2016 p w b ) ] T b } 1 / 3 ,
and the mass of water evaporated per unit of time is
M ˙ e v a p = 9.15 × 10 7 h c ( p w b p w c ) .
For the convection between the basin and the lower side surface of the cover
Q c o n v ( b c ) = h c ( T b T c ) .
For the convection between the upper surface of the cover and the atmosphere, which is normally a mixture of free and forced convection, McAdams recommends to adopt the convection coefficient as the maximum between 5 Wm−2K−1 and the value obtained with 8.6V0.6Lc−0.4; so
h ( c a ) = m a x . [ 5 ;   8.6 V 0.6 L c 0.4 ] ,
with V representing the velocity of wind, expressed in meters per second, and the characteristic length, Lc, representing the cube root of the volume of the distiller, expressed in meters. Then
Q c o n v ( c a ) = h ( c a ) ( T c T a ) .
The radiation exchange between the upper face of the cover and the sky is
Q r a d ( c a ) = ε c σ ( T c 4 T s k 4 ) .
The temperature of the sky was obtained by
T s k = T a [ 0.711 + 0.0056 T w b + 0.000073 T w b 2 + 0.013 cos ( 15 t ) ] 2 .
The value of Qgrd is negligible, since it depends on the construction quality of the basin, which can be controlled by the builder of the basin; a well-insulated basin is usually assumed.
The performance of the distiller was assessed through the efficiency calculated with
η = M ˙ e v a p / G .

2.2. Economics

In order to carry out the feasibility study of a project, it is very important to make an economic analysis of it. Firstly, this requires predicting expenditures (investments, costs, and expenses) and revenues and, secondly, the analysis of some economic indicators (TCO, NPV, IRR, and payback period). The presently studied facilities consist of distillers with an area arbitrarily set as 1 m2, and therefore the costs considered are per square meter of installation. Costs of components were obtained through the CYPE, a price generator digital platform [36].
Total cost of ownership (TCO) is a financial estimate designed to evaluate direct and indirect costs related to purchase and capital expenditures (CAPEX), in addition to operating and maintenance costs (OPEX), and can be expressed through
T C O = C A P E X + O P E X .
CAPEX covers only the cost of building the equipment. OPEX includes maintenance and operating costs, considering an annual inflation rate of 2% [37,38].
Considering that the equipment has a lifespan of 20 years, the cost of water per 1 m3 is given by
c o s t m 3 = T C O / ( M e v a p - y e a r l y × n ) ,
with n representing the operation years and M e v a p - y e a r l y representing the mass of fresh water obtained yearly.
The net present value (NPV) is the balance of all cash inflows and outflows over the estimated useful life of a project, updated to the present moment, considering an inflation rate for the stipulated period. NPV can be calculated using
N P V = n = 0 n F a n R n n = 0 n F a n G n ,
where F a n is the update factor for period n, R n the revenue for period n, and G n the expenditure for period n.
The value of F a n is calculated through
F a n = 1 ( 1 + i ) n ,
where i is the inflation rate and n the years of operation.
The internal rate of return (IRR) represents the return on an investment, expressed as a percentage rate. The IRR can be viewed as the rate required for NPV to be zero, and calculated by
N P V = 0 = I n i t i a l   I n v e s t m e n t + t = 1 n F n ( 1 + I R R ) n ,
where F n is the cash flow for period n.
The payback period represents the time it will take a project to generate returns that equal the investment, and can be calculated using
P a y b a c k = I n v e s t m e n t R e v e n u e / y e a r .

2.3. Carbon Emissions

The parameter analyzed was the CO 2 e emissions, defined by IEA 2018, corresponding to the production of 1 m3 of fresh water. This parameter only encompasses the CO 2 e produced by the operation of getting fresh water. Two alternative scenarios were studied: (i) the production of water through reverse osmosis (RO), currently the most widespread desalination process [39], and (ii) the production of water through solar distillation, the process presently studied. Obviously, RO consumes electric energy unlike solar distillation. Other methods to get fresh water were excluded from this analysis due to their technical complexities or operational costs, which renders them inappropriate for small and poor communities, the main target of this study.
Table 4 contains values of total emissions of CO 2 e and of the total electricity produced by the countries, where the sites studied are located. Specific emissions of CO 2 e were calculated with both these values; these specific emissions depend on the energy mix of the country. Whenever there were no reliable values either of CO 2 e emissions or the amount of electricity generated, counterpart values of the nearest country were assumed, such as in the cases of Atar, Dikhil, and Faya Largeau.
Besides, in Cabo Verde, in the year of 2018 [40], the consumption of electricity to obtain potable water through RO was 3.909 kWh/m3 (31,684,028 kWh to get 8,106,322 m3 of potable water); it is plausible to assume this same value for other countries, whenever RO is concerned.

3. Results

3.1. Thermal Efficiency

Results concerning output were obtained for the locations indicated at Table 2, with climatic data drawn from reference [42], equations (1) and (2), and the assumption of steady state operation for the basin and the cover.
Hourly values of Q e v a p , M ˙ e v a p , Q c o n v ( b c ) , h ( c a ) , Q c o n v ( c a ) , and Q r a d ( c a ) were calculated with hourly values of V, T a , T s k , and T w b drawn from the referred climatic site and with the above-mentioned equations, from (1) to (11). The values of absorptivity of water, transmissivity, and reflexivity of the glass cover were assumed as the averages of the homologous values contained at Table 3. The unknowns were the temperature of the basin, T b , and the temperature of the cover, T c , determined by a solver for nonlinear equations. Lastly, hourly values of the efficiency of the solar distiller, η, were obtained as well. Annual values shown in the following graphs were calculated with the counterpart hourly values.
Figure 2 shows the annual output of distilled water for the sites studied, whereas Figure 3 shows the annual efficiency of solar distillation plants for the same sites.
Among the chosen sites of Table 2, the most and the least suitable for setting up solar distillation plants are, respectively, Faya Largeua and Évora. For both these sites, Figure 4 shows the monthly output of water and includes the monthly average air temperature, whereas Figure 5 shows the monthly output of water and includes the monthly average value of G.
Figure 6 shows for the site with the highest water output, Faya Largeua, the monthly values of water output for two different situations: the minimum value and the actual value of the convection coefficient h ( c a ) (see Equation (8)).
It is clear from previous figures that the productivity and the efficiency of solar distillers are higher for higher values of either irradiation G or average temperature T m . Besides, the productivity and the efficiency of solar distillers increase as the convection coefficient h(c−a) drops. Generally, for the sites studied, when h ( c a ) drops from the actual to the minimum value, the increase of productivity is around 10%: for the case of Faya Largeua, such an increase is 12.3%. Thus, some expedients should be adopted to decrease, or even eliminate, the velocity of the wind over the distillers. Such expedients can be hedgerows, concrete barriers, etc.

3.2. Economics

An economic analysis was made for the sites referred to in Table 2, with the more important results presented in Table 5. The water annual production was drawn from Figure 2. The average costs of water for consumers were obtained from potable water suppliers: TCO, NPV for twenty years, IRR, and payback result, respectively, from Equation (12), Equation (14), Equation (16), and Equation (17).
It is important to stress that the costs of water presented in Table 5 are subsidized by the abovementioned governments.
Values of Table 5 are, at least, discouraging for poor communities of the Third World tormented by a lack of drinkable water.
Clearly, water can be obtained with distillers made with cheap materials found locally and by local unskilled people; such distillers will be affordable by poor communities, but their performance will be poor. Meanwhile, the performance and economic analyses of such distillers are impossible to calculate because (i) there is an enormous array of materials more or less suitable for the construction of distillers, and consequently, more or less expensive, and (ii) the unskillfulness of people is disputable.
So, the results presented in Table 5, however discouraging, were obtained with the underlying scope to obtain distiller with the highest efficiency, manufactured with prices current at Cabo Verde.
Anyway, from the values of payback of Table 5, it is clear that solar distillers are only suitable for domestic water production where other production methods are unavailable.

3.3. Carbon Emissions

Table 6 contains, for the set of sites studied, the values of Greenhouse Gas (GHG) that would be emitted if the production of the water, instead of being through the solar distillers proposed in this study, was by RO. It is considered that the operation of solar distillers does not emit GHG. The water annual production was drawn from Figure 2, and the consumption of electricity to obtain potable water through RO was assumed 3909 kWh/m3, the same as in Cabo Verde.
Clearly the environmental impact of RO is a drawback of this process when compared with solar distillation.

4. Conclusions

The scarcity of water is a problem of a growing number of regions of the world, which cumulatively are the poorest regions. Solar distillation has higher efficiency as irradiation increases and thermal losses decrease, which means that it is an acceptable desalination process for arid and hot climates.
Typical values of annual water output are in the range of 414 dm3/m2 for Évora and up to 696 dm3/m2 for Faya Largeau; the minimum efficiency found for Évora was 11.5%, and the maximum efficiency found for Tessalit was 15.2%.
However, the payback times for high performance solar distillers are very high, which excludes solar distillation when economic criteria are the only factors to be considered.
Solar distillation is far better than RO, when the environmental impact of GHG is considered. For remote and poor communities, solar distillation is to be considered.
So, as guidelines to decide if solar distillation to obtain fresh water it is important to get positive answers to the following questions:
  • Remote places;
  • Salt water available;
  • High solar irradiation;
  • Population with low income;
  • Population with poor skills;
  • Off-grid places.
Finally, the present study addresses the shortage of fresh water in the growing number of places that comply with the previous guidelines.

Author Contributions

The authors have made equivalent contributions. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

List of Symbols
A Area of the basinm2
C O 2 e Carbon dioxide equivalent emissions
c Specific heatJ kg−1 K−1
C A P E X Capital expenditure
F Casflow
G SolaIrradiation W m−2
G H G Grehouse gas
i Inflion rate%
IRRInternal rate of return%
hConvection coefficientW m−2 K−1
hfgLatent heat of evaporationJ kg−1
LcCharacteristic lengthm
MMasskg
M ˙ Mass flow ratekg s−1
NPVNet Present Value
nPeriodyear
OPEXOperational Expenditure
pPressuremmHg
PBPaybackyear
PVCPolyvinyl chloride
QHeat rateW m−2
RRevenue
ROReverse osmosis
TTemperatureK
tTimes
TCOTotal cost of ownership
VWind velocitym s−1
Subscripts
aAtmosphere
bBasin
cCover
convConvection
dDistilled
evapEvaporation
grdGround
mAverage
nPeriodyear
radRadiation
skSky
wWater
wbWet bulb
Greek symbols
α Absorptivity
η Performance
θAngle of incidence
ρ Reflectivity
σStefan-Boltzmann constantJ m−2 s−1 K−4
τ Transmissivity

References

  1. Belessiotis, V.; Kalogirou, S.; Delyannis, E. Thermal Solar Desalination: Methods and Systems, 1st ed.; Academic Press: Cambridge, MA, USA, 2016. [Google Scholar]
  2. Cooper, P.I. The absorption of radiation in solar stills. Sol. Energy 1969, 12, 333–349. [Google Scholar] [CrossRef]
  3. Kumar, K.V.; Bai, R.K. Performance study on solar still with enhanced condensation. Desalination 2008, 230, 51–61. [Google Scholar] [CrossRef]
  4. Phadatare, M.K.; Verma, S.K. Effect of cover materials on heat and mass transfer coefficients in a plastic solar still. Desalin. Water Treat. 2009, 2, 254–259. [Google Scholar] [CrossRef]
  5. Duffie, J.A.; Beckman, W.A. Solar Engineering of Thermal Processes, 4th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013. [Google Scholar]
  6. Harding, J. Apparatus for solar distillation. Minutes Proc. Inst. Civ. Eng. 1883, 73, 284–288. [Google Scholar] [CrossRef]
  7. Telkes, M. Fresh water from sea water by solar distillation. Ind. Eng. Chem. 1953, 45, 1108–1114. [Google Scholar] [CrossRef]
  8. Lof, G.O.G.; Eibling, J.A.; Bloemer, J.W. Energy balances in solar distillers. AIChE J. 1961, 7, 641–649. [Google Scholar] [CrossRef]
  9. Bloemer, J.W.; Eibling, J.A.; Irwin, J.R. A practical basin-type solar still. Sol. Energy 1965, 9, 197–200. [Google Scholar] [CrossRef]
  10. Talbert, S.G.; Eibling, J.A. Manual on Solar Distillation of Saline Water; NTIS: Springfield, VA, USA, 1970. [Google Scholar]
  11. Hull, J.R. Solar ponds using ammonium salts. Sol. Energy 1986, 36, 551–558. [Google Scholar] [CrossRef]
  12. Hull, J.R.; Bushnell, D.L.; Sempsrote, D.G.; Pena, A. Ammonium sulfate solar pond: Observations from small-scale experiments. Sol. Energy 1989, 43, 57–64. [Google Scholar] [CrossRef]
  13. Nielsen, C.E. Salinity-gradient solar ponds. In Advances in Solar Energy; Böer, K.W., Ed.; Springer: Boston, MA, USA, 1988; Volume 4. [Google Scholar]
  14. Morse, R.N. The Development of a Solar Still for Australian Conditions; Institution of Engineers: Barton, Australia, 1967. [Google Scholar]
  15. Morse, R.N. The construction and installation of solar stills in Australia. Desalination 1968, 5, 82–89. [Google Scholar] [CrossRef]
  16. Morse, R.N.; Read, W.R.W. A rational basis for the engineering development of a solar still. Sol. Energy 1968, 12, 5–17. [Google Scholar] [CrossRef]
  17. Morse, R.N.; Read, W.R.W.; Trayford, R.S. Operating experiences with solar stills for water supply in Australia. Sol. Energy 1970, 13, 99–103. [Google Scholar] [CrossRef]
  18. Morse, R.N. The theory of solar still operation. In Proceedings of the UN Solar Distillation Panel Meeting, Palo Alto, CA, USA, 20–23 October 1968. [Google Scholar]
  19. Lawand, T. Engineering and economic evaluation of solar distillation for small communities. Presented at the Annual ASME Meeting, Los Angeles, CA, USA, 16–20 November 1969. [Google Scholar]
  20. Lawand, T. Systems for solar distillation. In Proceedings of the Appropriate Technologies for Semiarid Areas: Wind and Solar Energy for Water Supply, Berlin, Germany, 15–20 September 1975. [Google Scholar]
  21. Telkes, M. Solar Stills; Proceedings World Symposium on Applied Solar Energy: Phoenix, AZ, USA, 1956. [Google Scholar]
  22. Hirschmann, J. Evaporación solar y su aplicación práctica en Chile. Scientia 1968, 136, 16. [Google Scholar]
  23. Lof, G.O.G.; Hunter, J.A.; Sieder, E.N. Technical Cooperation on the Solar Distillation Development Program of Spain. In Distillation Digest; Tomalin, P.G., Ed.; Office of Saline Water: Washington, DC, USA, 1970; Volume 1–2, pp. 241–252. [Google Scholar]
  24. Bloemer, J.W.; Collins, R.A.; Eibling, J.A. Study and Field Evaluation of Solar Sea-Water Stills; United States. Department of Commerce. Office of Technical Services: Washington, DC, USA, 1961.
  25. Bloemer, J.W.; Irwin, J.R.; Eibling, J.A. ‘Second Two-Years’ Progress on Study and Field Evaluation of Solar Sea-Water Stills; Battelle Memorial Institute: Columbus, OH, USA, 1964. [Google Scholar]
  26. Eckstrom, R. Design and construction of the Symi still. Sun Work 1965, 10, 6–8. [Google Scholar]
  27. Delyannis, A. Solar stills provide an island’s inhabitants with water. Sun Work 1965, 10, 6. [Google Scholar]
  28. Eckstrom, R. New solar stills on Greek islands. Sun Work 1966, 11. [Google Scholar]
  29. Delyannis, A.; Delyannis, E.; Piperoglou, E. Solar distillation developments in Greece. Sun Work 1967, 12. [Google Scholar]
  30. Delyannis, A.; Piperoglou, E. The Patmos solar distillation plant. Sol. Energy 1968, 12, 113–115. [Google Scholar] [CrossRef]
  31. Datta, R.L.; Gomkale, S.D.; Ahmed, S.Y.; Datar, D.S.; Intensity, T.X.S.R. Evaporation of sea water in solar stills and its development for. In Proceedings of the First International Symposium on Water Desalination, Washington, DC, USA, 3–9 October 1965; Volume 2, p. 193. [Google Scholar]
  32. Ahmed, S.Y.; Gomkale, S.D.; Datta, R.L.; Datar, D.S. Scope and development of solar stills for water desalination in India. Desalination 1968, 5, 64–74. [Google Scholar] [CrossRef]
  33. Baum, V.A.; Bairamov, R. Prospects of solar stills in Turkmenia. Sol. Energy 1966, 10–11, 38–40. [Google Scholar] [CrossRef]
  34. Cooper, P.I.; Bairamov, R. The maximum efficiency of single-effect solar stills. Sol. Energy 1973, 15, 205–217. [Google Scholar] [CrossRef]
  35. Dunkle, R.V. Solar water distillation: The roof type still and a multiple effect diffusion still. In Proceedings of the International Heat Transfer Conference, Boulder, CO, USA, 28 August–1 September 1961; Volume 5, p. 895. [Google Scholar]
  36. Gerador de Preços Para Construção Civil. Cabo Verde. CYPE Ingenieros, S.A. Available online: http://www.caboverde.geradordeprecos.info/ (accessed on 17 July 2019).
  37. Chad: Inflation Rate from 1984 to 2024. Available online: https://www.statista.com/statistics/529342/inflation-rate-in-chad/ (accessed on 20 July 2019).
  38. Portugal: Inflation Rate from 1984 to 2024. Available online: https://www.statista.com/statistics/529342/inflation-rate-in-portugal/ (accessed on 20 July 2019).
  39. Al-Karaghouli, A.; Kazmerski, L.L. Energy consumption and water production cost of conventional and renewable-energy-powered desalination processes. Renew. Sustain. Energy Rev. 2013, 24, 343–356. [Google Scholar] [CrossRef]
  40. Relatório e Contas 2018. Available online: http://www.electra.cv/index.php/2014-05-20-15-47-04/relatorios-sarl/ (accessed on 20 October 2019).
  41. CO2 Emissions from Fuel Combustion 2018 (with 2016 Data). Available online: https://webstore.iea.org/co2-emissions-from-fuel-combustion-2018 (accessed on 20 October 2019).
  42. Repository of Free Climate Data for Building Performance Simulation. Available online: http://climate.onebuilding.org/ (accessed on 15 July 2019).
  43. Tarifas de Abastecimiento Almeria. Available online: https://www.aqualia.com/documents/373313/374060/Almer%C3%ADa_TARIFAS+ABASTECIMIENTO.pdf/d9a04334-9c2a-425c-924d-4866c7a10802/ (accessed on 17 October 2019).
  44. La Distribution d’eau Potable Dans la Ville de Nouakchott, Mauritanie. Analyse des Points de Vente d’eau. Available online: https://www.pseau.org/outils/ouvrages/gret_distribution_eau_potable_ville_nouakchott_analyse_points_d_eau.pdf (accessed on 17 October 2019).
  45. Global Water Intelligence. Available online: https://www.globalwaterintel.com/news/2017/31/egypt-to-slash-subsidies-in-water-and-wastewater?dm_i=36G3,IGWK,2MX9NG,1XJV0,1 (accessed on 17 October 2019).
  46. SEEN. Available online: https://www.seen-niger.com/fr/activites/relation-clientele/nos-tarifs (accessed on 17 October 2019).
  47. Chohin-Kuper, A.; Strosser, P. Water pricing in Europe and around the Mediterranean Sea: Issues and options. Presented at 4th EWA Brussels Conference, Brussels, Belgium, 8 November 2008. [Google Scholar]
  48. NIPA—National Investment Promotion Agency. Available online: http://www.djiboutinvest.com/index.php?option=com_content&view=article&id=289&Itemid= (accessed on 17 October 2019).
  49. Tarifário dos Serviços de Abastecimento de Águas, Saneamento de Águas Residuais e Gestão de Resíduos Urbanos do Município de Évora. 2019. Available online: http://www.cm-evora.pt/pt/servicos/aguas/Paginas/Tarif%C3%A1rios.aspx (accessed on 17 October 2019).
  50. UNHabitat. Available online: http://unhabitat.org/wpcontent/uploads/2015/12/Soci%C3%A9t%C3%A9%20Tchadienne%20des%20Eaux%20(Chad).pdf (accessed on 17 October 2019).
  51. Fixation des Tarifs de L’Eau Potable Applicables. Available online: http://www.creemali.ml/documents/TARIFS_EAU_2013.pdf (accessed on 17 October 2019).
Figure 1. Schematic cross section of a typical solar distillation distiller.
Figure 1. Schematic cross section of a typical solar distillation distiller.
Sustainability 12 00053 g001
Figure 2. Annual output of distilled water for the sites studied.
Figure 2. Annual output of distilled water for the sites studied.
Sustainability 12 00053 g002
Figure 3. Efficiency of solar distillation plants for the sites studied, calculated with annual weather values.
Figure 3. Efficiency of solar distillation plants for the sites studied, calculated with annual weather values.
Sustainability 12 00053 g003
Figure 4. Monthly output of distilled water for Faya Largeua and Évora and the corresponding monthly average air temperature.
Figure 4. Monthly output of distilled water for Faya Largeua and Évora and the corresponding monthly average air temperature.
Sustainability 12 00053 g004
Figure 5. Annual output of distilled water for Faya Largeua and Évora and the corresponding monthly average value of G.
Figure 5. Annual output of distilled water for Faya Largeua and Évora and the corresponding monthly average value of G.
Sustainability 12 00053 g005
Figure 6. Comparison of the monthly output of distilled water for Faya Largeua with h ( c a ) = 5 W/m2/K and h ( c a ) obtained with the local wind speed measured.
Figure 6. Comparison of the monthly output of distilled water for Faya Largeua with h ( c a ) = 5 W/m2/K and h ( c a ) obtained with the local wind speed measured.
Sustainability 12 00053 g006
Table 1. Features of solar distillation plants.
Table 1. Features of solar distillation plants.
CountrySitesConstruction YearBasin Area (m2)Production (L/day)Daily Production Per Area (L/m2)Reference
AustraliaMuresk I19633728332.24[14,15,16]
Muresk II19663728332.24[15,16,17,18]
Coober Pedy1966315963592.01[15]
Caiguna19663727762.09[19]
Hamelin Pool196655712112.17[10,20]
Griffith19674139082.20[10,20]
Cabo VerdeSanta Maria196574321202.85[10,20]
CaribbeanPetit St. Vincente1967170949212.88[19]
Haiti19692237573.39[10,20]
ChileLas Salinas1872445914,7633.31[6,21]
Quillagua19681004014.01[22]
SpainLas Marinas196686925742.96[23]
USADaytona Beach 119592285302.32[24]
Daytona Beach 219592163791.75[24,25]
Daytona Beach 319612465682.31[25]
Daytona Beach 419631496064.07[10,20]
GreeceSymi1964268775712.82[26,27]
Aegina1965149042402.84[28]
Salamis196538810982.83[20]
Patmos1967864026,1193.02[29,30]
Kimolos1968250875713.02[20]
Nisiros1969204460572.96[20]
IndiaBhavnagar19653778332.21[31,32]
MexicoCalifornia1969953793.99[20]
TunisiaChakmou19674395301.21[20]
Mahdia1968130141643.20[20]
USSRTurkmenistan196459916282.72[33]
Table 2. Sites studied at the present work chosen to assess their suitability to desalination through solar distillation plants.
Table 2. Sites studied at the present work chosen to assess their suitability to desalination through solar distillation plants.
SitesCountryLatitude (°)Longitude (°)Altitude (m) T m Annual (°C) G Annual (KWh m−2 y−1)
Al FashirSudan13.625.372925.02482
AlmeriaSpain36.8−2.32118.71880
AtarMauritania20.5−13.023127.02404
Abu RudeisEgypt28.933.1823.72223
BilmaNiger18.712.935726.42428
DakhaWestern Sahara23.7−15.91019.72005
DikhilDjibouti11.142.349025.32584
ÉvoraPortugal38.5−7.824616.31849
Faya LargeauChad17.919.123524.42456
MalakalSouth Sudan9.531.639426.72351
MassawaEritrea15.639.41029.32184
MekeleEthiopia13.439.5225416.52386
MogadishuSomalia2.045.3327.02239
SantiagoCabo Verde14.9−23.59524.62245
St. LouisSenegal16.0−16.42.725.192174
TessalitMali20.20.9749427.82440
Table 3. Total and directional values of absorptivity, transmissivity, and reflectivity for the glass cover, the water in the basin, and the basin liner; these values are presented in % [34].
Table 3. Total and directional values of absorptivity, transmissivity, and reflectivity for the glass cover, the water in the basin, and the basin liner; these values are presented in % [34].
Angle of Incidence, θ30°45°60°
Glass cover
Absorptivity, α5555
Transmissivity, τ90908985
Reflectivity, ρ55610
Water in the basin
Absorptivity, α30303030
Transmissivity, τ68686764
Reflectivity, ρ2236
Basin bottom (liner)
Absorptivity, α95959595
Transmissivity, τ0000
Reflectivity, ρ5555
Table 4. Emissions of CO2e.
Table 4. Emissions of CO2e.
Sites CO 2 e Emissions (Mton) Electricity Generated (TWh)Specific Emissions of CO 2 e
kg CO 2 e /kWh
Ref.
Al Fashir6.215.50.400[41]
Almeria66.7273.40.243[41]
Atar--0.625 1-
Abu Rudeis89.3188.20.474[41]
Bilma0.50.60.833[41]
Dakha21.632.80.658[41]
Dikhil--1.000 2-
Évora18.157.70.314[41]
Faya Largeau--0.400 3-
Malakal0.50.51.000[41]
Massawa0.40.41.000[41]
Mekele013.90[41]
Mogadishu--1.000 2-
Santiago0.20.40.490[40]
St. Louis3.04.80.625[41]
Tessalit--0.625 1-
1 Value from Senegal, 2 value from Eritrea, and 3 value from Sudan.
Table 5. Economic analysis summary.
Table 5. Economic analysis summary.
LocationAverage Cost of Water (€/m3)Water Annual Production (L/m2)TCO 20 Years (€/m2)NPV 20 Years (€/m2)IRR (%)Payback (Years)
Al Fashir0.766 1642.6270.1−234.40-369
Almeria0.699 [43]420.1−247.30-619
Atar5.952 2 [44]644.4−177.85−22.147
Abu Rudeis0.167 [45]545.4−251.27-1996
Bilma0.218 [46]651.4−250.27-1280
Dakha0.460 [47]457.9−248.65-810
Dikhil0.448 [48]680.9−247.07-596
Évora0.507 [49]414.4−248.93-865
Faya Largeau0.766 [50]695.8−242.60-341
Malakal0.766 1619.4−243.75-383
Massawa0.448 3680.9−247.07-596
Mekele0.448 3560.5−248.13-724
Mogadishu0.448 3583.0−247.93-696
Santiago4.23 4 [40]562.5−206.40-76
St. Louis5.952 5551.6−188.67-55
Tessalit0.207 [51]683.0−250.28-1286
1 Cost from Chad, 2 supply in 200 L drums, 3 cost from Djibouti, 4 auto tank supply, 5 and cost from Mauritania.
Table 6. Carbon emissions analysis summary.
Table 6. Carbon emissions analysis summary.
LocationWater Annual Production (L/m2)RO Desalinization Energy Consumption (kWh/m3)GHG Avoid 20 Years ( kg CO 2 e / m 2 )
Al Fashir642.6390920.1
Almeria420.18.0
Atar644.431.5
Abu Rudeis545.420.2
Bilma651.442.4
Dakha457.923.6
Dikhil680.953.2
Évora414.410.2
Faya Largeau695.821.8
Malakal619.448.4
Massawa680.953.2
Mekele560.50
Mogadishu583.045.6
Santiago562.521.5
St. Louis551.627.0
Tessalit683.033.4

Share and Cite

MDPI and ACS Style

Monteiro, J.; Baptista, A.; Pinto, G.; Ribeiro, L.; Mariano, H. Assessment of the Use of Solar Desalination Distillers to Produce Fresh Water in Arid Areas. Sustainability 2020, 12, 53. https://doi.org/10.3390/su12010053

AMA Style

Monteiro J, Baptista A, Pinto G, Ribeiro L, Mariano H. Assessment of the Use of Solar Desalination Distillers to Produce Fresh Water in Arid Areas. Sustainability. 2020; 12(1):53. https://doi.org/10.3390/su12010053

Chicago/Turabian Style

Monteiro, Joaquim, Andresa Baptista, Gustavo Pinto, Leonardo Ribeiro, and Hélder Mariano. 2020. "Assessment of the Use of Solar Desalination Distillers to Produce Fresh Water in Arid Areas" Sustainability 12, no. 1: 53. https://doi.org/10.3390/su12010053

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop