Sie sind auf Seite 1von 99

MASTER OF SCIENCE THESIS

AUTONOMOUS PHOTOVOLTAIC
POWERED SEAWATER REVERSE
OSMOSIS FOR REMOTE
COASTAL AREAS

EVANGELIA GKEREDAKI

16th June 2011

AUTONOMOUS PHOTOVOLTAIC-POWERED REVERSE


OSMOSIS FOR REMOTE COASTAL AREAS

EVANGELIA GKEREDAKI

For the degree of:

Master of Science in Sustainable Energy Technology

Date of defense: 16th June 2011

THESIS REVIEW COMMITTEE:


PROF. DR. IR. L.C. RIETVELD
DR. IR. S.G.J. HEIJMAN
PROF. DR. F.M. MULDER

DELFT UNIVERSITY OF TECHNOLOGY


SANITARY ENGINEERING SECTION
DELFT UNIVERSITY OF TECHNOLOGY
SANITARY ENGINEERING SECTION
DELFT UNIVERSITY OF TECHNOLOGY
RRR/FUNDAMENTAL ASPECTS OF MATERIALS AND ENERGY

SANITARY ENGINEERING SECTION, DEPARTMENT OF WATER MANAGEMENT


FACULTY OF CIVIL ENGINEERING AND GEOSCIENCES
DELFT UNIVERSITY OF TECHNOLOGY, DELFT

A BSTRACT

This master thesis is part of the venture Drinking with the sun conceived by the section Sanitary
Engineering in CiTG. It is a part of the ongoing research in the section with regards to water
purification using renewable energy technologies.
The project Drinking with the Sun involves system integration of solar photovoltaics with a reverse
osmosis desalination system with the ultimate aim of producing safe but sustainable drinking water
from sea water. Such a technology may be ideally suited for application in remote coastal areas,
facing a shortage of fresh water, but endowed with ample insolation.
Solar Energy is, by nature, fluctuating, whereas existing reverse osmosis systems are designed for
continuous operation. In most autonomous renewable energy powered systems, batteries are used
for energy storage. However, in the last few years, research interest has been towards elimination of
batteries due to the financial and often environmental costs involved with their maintenance and
replacement. Making a batteryless system which will be economic and reliable at the same time is a
challenge for seawater PVRO desalination.
Another possibility for improving efficiency of PVRO units lies in utilizing some of the energy from the
waste high pressure stream of the RO unit. The present project involves testing a device called the
Pearson Pump manufactured by Spectra Watermakers, which is unique in the way that it
incorporates an innovative energy recovery mechanism in a small scale system.
The project work involved sizing the system components, building and testing the prototype in a real
scenario. For the purposes of the experimentation, the island of Crete (Greece) was chosen as the
area of the PVRO installation and cooperation with the Technical University of Crete was established
for the successful implementation of the project and for the investigation of its further sustainability.
The scientific approach followed in the course of the experimentation lies in the worst-case scenario
testing of the real prototype, in real insolation and feed water conditions.
The preliminary results of the system positively indicate the technical feasibility of the system, and
offer recommendations oriented towards improving the reliability of the system in practical field
conditions.

ii

A CKNOWLEDGEMENTS
First and foremost, I would like to thank my supervisor dr.ir. S.G.J Heijman for giving me the
opportunity to carry out this project as my master thesis and for facilitating me to conduct my
experiments in Greece. The flexibility, feedback and support I was given allowed me to explore my
capabilities, and develop myself. Secondly, I would like to thank prof.dr.ir. L.C. Rietveld for the
guidance and feedback review for this thesis as well as prof.dr.ir. F.M. Mulder for agreeing to be a
member of my review committee. Special thanks to professor E.Diamantopoulos and the Laboratory
of Environmental Engineering and Management of Technical University of Crete, Greece for hosting
me during my experiments and for all the generous help, space and equipment provision, which
ensured the smooth completion of the experimental phase of the project.
Last but not least, I would like to thank house no.5, permanent and guests, my family and friends for
giving comfort in moments of stress and hard times and providing me with distraction when I needed
it. My sincerest gratitude and thanks to Yash, for all the unconditional and endless support and help
during my studies and for being always there for me no matter what.
Lastly, I would like to dedicate this thesis to the memory of my beloved grandmother, a woman of
great strength and courage who passed away just before the completion of this report.

iii

C ONTENTS
ABSTRACT .............................................................................................................................I
ACKNOWLEDGEMENTS ...................................................................................................... III
1. WATER SCARCITY AND DESALINATION .......................................................................... 1
1.1. FRESH WATER SCARCITY AROUND THE WORLD ............................................................................... 1
1.2. WATER GUIDELINES AND HEALTH ISSUES RELATED TO WATER SCARCITY ................................................ 2
1.3. DESALINATION - THE SOLUTION TO FRESH WATER SCARCITY .............................................................. 2
1.4. DESALINATION TECHNOLOGIES................................................................................................... 4
1.5. REVERSE OSMOSIS - THE LEADER IN DESALINATION ......................................................................... 5
1.6. DESALINATION SYSTEMS - SYNERGY WITH RENEWABLE ENERGY ......................................................... 6
2. SOLAR DESALINATION AND PV-RO SYSTEMS ............................................................... 11
2.1. INTRODUCTION .................................................................................................................... 11
2.2. PV-POWERED RO SYSTEM COMPONENT ANALYSIS......................................................................... 12
2.3. SYSTEM REVIEWS ................................................................................................................. 22
2.4. ECONOMICS AND MARKET ISSUES ............................................................................................. 25
3. SIZING THE PV-RO SYSTEM COMPONENTS ................................................................... 29
3.1. INTRODUCTION .................................................................................................................... 29
3.2. SIZING APPROACH AND ASSUMPTIONS ........................................................................................ 29
3.3. CRITERIA FOR PRODUCTION OF SAFE WATER ............................................................................... 30
3.4. SIZING STRATEGY................................................................................................................. 32
3.5. RESULTS ............................................................................................................................ 34
3.6. THEORETICAL COST ANALYSIS ................................................................................................. 35
4. EXPERIMENTATION ....................................................................................................... 39
4.1. SELECTION OF THE AREA OF STUDY AND PROJECT PARTNERS ............................................................ 39
4.2. LOCATION CHARACTERISTICS AND METEOROLOGICAL DATA .............................................................. 40
4.3. OBJECTIVES ........................................................................................................................ 41
4.4. EXPERIMENTAL SETUP............................................................................................................ 42
4.5. SYSTEM START UP ................................................................................................................ 47
4.6. SYSTEM OPERATION .............................................................................................................. 48
4.7. PROBLEMS FACED / TROUBLESHOOTING ...................................................................................... 49
5. RESULTS AND DISCUSSION ........................................................................................... 53
5.1. A SAMPLE DAY ..................................................................................................................... 53
5.2. GENERAL COMMENTS ON EXPERIMENTATION................................................................................ 56
5.3. ROLE OF LINEAR CURRENT BOOSTER (LCB) ................................................................................ 59
5.4. OBSERVATION OF SUDDEN QUALITY DETERIORATION...................................................................... 65
5.5. LIMITING FACTORS................................................................................................................ 67
5.6. DIFFERENT MODES OF OPERATION ............................................................................................ 68
5.7. COMPARISON OF THEORETICAL AND PRACTICALLY OBTAINED WATER YIELDS ......................................... 69
6. CONCLUSIONS & RECOMMENDATIONS ......................................................................... 71
BIBLIOGRAPHY .....................................................................................................................I
APPENDICES ........................................................................................................................ V

T ABLE

OF

F IGURES

Figure 1: The alarming extent of water scarcity across the world is detailed in a map compiled by the
International Water Management Institute (IWMI) [6] .................................................................... 1
Figure 2: Total installed desalination capacity by country (2006)[17]................................................ 3
Figure 3: Global installed desalination capacity by feed water sources (in million m3/d)(2006) [17] .... 3
Figure 4: Main desalination processes [7] ....................................................................................... 4
Figure 5: Global installed desalting capacity by process (2006) [17] ................................................. 4
Figure 6: Principle Of Osmosis And Reverse Osmosis Process [26] ................................................... 5
Figure 7: Possible Uses Of Renewable Energy Sources With Desalination Systems [3] ....................... 7
Figure 8: Distribution of RE-powered desalination technologies [3] .................................................. 8
Figure 9: Average 3-year insolation from calibrated data collected by the NASA International Satellite
Cloud Climatology Project (ISCCP) (1991-1993) [29] ..................................................................... 11
Figure 10: Solar Peak Hours on an optimally titled surface during the worst months of the year [30] 11
Figure 11: Simplified general design of a PV-RO desalination plant [16] ......................................... 13
Figure 12: Solar Cell I-V curve with MPPT in varying sunlight [35] ................................................. 15
Figure 13: Spectra Clark Pump and Spectra Pearson Pump [48] ..................................................... 18
Figure 14: Design of the Pearson Pump [48] ................................................................................ 19
Figure 15: SEC for seawater and brackish water PV-RO systems [16] ............................................. 23
Figure 16: Indicative shares of total costs in conventional seawater desalination [3] ....................... 25
Figure 17: Photovoltaics industry from 1999 - 2008....................................................................... 25
Figure 18: PV Module price versus cumulative module production from 1979 - 2009 ....................... 26
Figure 19: Cost versus Conversion Efficiency in 2008 .................................................................... 26
Figure 20: Power Consumption versus Product Flow Rate and Water Quality [1] ............................. 31
Figure 21: Monthly Optimal Inclination Angles for Chania, Greece .................................................. 32
Figure 22: Irradiation during a typical day in Chania in December and its possible Exploitation using
20 PV Panels............................................................................................................................... 34
Figure 23: Variation of Annual Water Production Costs .................................................................. 36
Figure 24: Variation of Water Cost with Interest Rates .................................................................. 36
Figure 25: Contributions to the Total Cost of a PV-RO System ....................................................... 37
Figure 26: Location of the Technical University of Crete, Chania .................................................... 40
Figure 27: Meteorological Data for Chania: Annual Temperature, Precipitation, Wind Speed and Water
Temperature............................................................................................................................... 40
Figure 28: Total solar energy (Monthly Values) over 2010 in Chania [Footnote 11] ......................... 41
Figure 29:The Experimental Setup in the WaterLab, CiTG, TU Delft [1] .......................................... 42
Figure 30: Roof on which the PV Panels were Installed.................................................................. 44
Figure 31: Solar PV Panels Installed on the Roof ........................................................................... 44
Figure 33: Electrical Connections of the Main Components of the System ....................................... 45
Figure 32: Linear Current Booster (LCB) ....................................................................................... 45
Figure 34: Instrumentation systems (from Left to Right): Analogue Pressure Gauges / FLowMeters,
Digital Conductivity meters, pyranometer ..................................................................................... 46
Figure 35: RO unit cabinet on field and Desk Station on field ......................................................... 47
Figure 36: Simplified Process Diagram of PV-RO System................................................................ 47
Figure 37: Relation between TDS and Electrical Conductivity of the feedwater used ........................ 48
Figure 38: Tank Configuration for Permeate Collection .................................................................. 50
Figure 39: Change in Feed Water Temperature during Daily Operation ........................................... 51
Figure 40: Final Process Flow Diagram and instrumentation ........................................................... 52
Figure 41: Solar Irradiation on a Sample Day (5/10/2010) ............................................................. 53
Figure 42: Variation of Operation Parameters in Time on Sample Day: 5/10/2010 ........................... 55
Figure 43: Relation between water quality, power consumption and flow on 5/10/2010 .................. 56
Figure 44: Worst-Case Solar Irradiation Patterns during experimentation period at chania ............... 57
Figure 45: Irradiation vs Feed Flow on 16/09/2010 ....................................................................... 57

Figure 46: Relation between Permeate Production and Power Consumption on a mostly clear day:
17/09/2010 ................................................................................................................................ 58
Figure 47: Power Consumption vs Irradiation on a mostly clear day: 17/09/2010 ............................ 58
Figure 48: Irradiation over time on 17/09/2010 and on 06/10/2010 ............................................... 60
Figure 49: Comparison of Variation in Motor Voltage for one clear day with the lcb and one clear day
without the lcb (Case Study 1) ..................................................................................................... 61
Figure 50: Comparison of Variation in Motor Current for one clear day with the lcb and one clear day
without the lcb (Case Study 1) ..................................................................................................... 61
Figure 51: Comparison of Permeate Flow for one clear day with the lcb and one clear day without the
lcb (Case Study 1) ....................................................................................................................... 62
Figure 52: Comparison of Variation in Permeate Quality for one clear day with the lcb and one clear
day without the lcb (Case Study 1) .............................................................................................. 62
Figure 53: Daily Irradiation Values on 12/10/2010 ........................................................................ 63
Figure 54: Daily Irradiation Values on 11/10/2010 ........................................................................ 64
Figure 55: PV-RO Parameters as a function of Time (Quality Deterioration) .................................... 66
Figure 56: Layout of the RO membranes ...................................................................................... 66
Figure 57: Difference in Solar Irradiation for South and Southwest Facing Panels ........................... 68

List of abbreviations
MENA
SEMI
RO
RES
BW
SW
DW
PV
PVRO
SWPVRO
HPP
LPP
TDS
TMP
LCB
MPPT
PLC
gpm

ER

Middle East and North Africa


South European Mediterranean islands
Reverse osmosis
Renewable energy source
Brackish water
Seawater
Demineralised water
Photovoltaics
Photovoltaic reverse osmosis
Seawater Photovoltaic Reverse osmosis
High Pressure Pump
Low Pressure Pump
Total dissolved solids
Trans membrane pressure
Linear Current Booster
Maximum power point tracker
Programmable Logic Controller
Gallons per minute
Pump efficiency
Energy recovery

1. W ATER

SCARCITY AND

1.1. F RESH W ATER

D ESALINATION

SCARCITY AROUND THE WORLD

Fresh water scarcity poses a big problem in remote regions especially in the Middle East and North
African countries (MENA), the southern European Mediterranean Islands (SEMI) and isolated
communities in deserts [2]. Pollution and exploitation of groundwater aquifers and surface water
have contributed to the decrease in quantity and/or quality of available natural water resources in
those areas [3]. The availability of electricity networks in those areas is often as limited as the
availability of safe drinking water, even as technologies that are able to remove pathogens and
dissolved contaminants require substantial amounts of energy [4].
One of the most conspicuous phenomena of water-quality degradation, particularly in arid and semiarid zones, is salinization of water and soil resources. Salinization is a long-term phenomenon due to
which many aquifers and river basins have become unsuitable for human consumption owing to high
levels of salinity during the course of the last century. The salinity problem has numerous grave
economic, social, and political consequences, particularly in cross-boundary basins that are shared by
different communities [5].

F IGURE 1: T HE

ALARMING EXTENT OF WATER SCARCITY ACROSS THE WORLD IS DETAILED IN A MAP COMPI LED
BY THE

I NTERNATIONAL W ATER M ANAGEMENT I NSTITUTE (IWMI) [6]

For the Arab countries potable water is becoming as critical a commodity as electricity [7]. The per
capita share of total annual renewable water resources has dropped well below the UN threshold for
water poverty (1000m3/yr) with most of the Gulf Arab countries reaching below 200m 3/yr per capita
[8]. The Aegean Archipelago islands in Greece have restricted water resources and for some of the
islands, salt water intrusion into aquifers has contributed to the deterioration of the quality of life of
the inhabitants. Especially during the summer, the majority of small and medium-sized Aegean
Archipelago islands have a significant clean water deficit and in several cases almost 50-80% of the
1

fresh water needed is transferred at a very high cost [9]. In July and August 2010, the cost for
transferring drinking water to 17 arid Greek islands rose to 12.5 /m 3, which is an extravagant cost as
compared to previous years. Transfer of water to arid areas in Greece has been happening since
1980 with the Greek government spending 7.5 million Euros for this year alone [10].
According to the World Water Council, today about three billion people around the world have no
access to clean drinking water, while, by 2020, the world will be about 17% short of the fresh water
needed to sustain the world population. Owing to the foreseen growth of the worlds population
(especially in the developing countries), the problem is expected to become more and more critical
over the next two decades, bringing the lack of potable water to the top of the international agenda
[3, 6].

1.2. W ATER

GUIDELINES

AND

HEALTH

ISSUES

RELATE D

TO

WATER

SCARCITY
Salinity in water is usually defined by the chloride (Cl-) content (mg/L or ppm) or total dissolved
solids content (TDS, mg/L or ppm), although the chloride comprises only a fraction of the total
dissolved salts in water. The Cl/TDS ratio varies from 0.1 in non-marine saline waters to 0.5 in
marine-associated saline waters. Water salinity is also defined by electrical conductivity (EC).
Based on salinity, water can be classified into three groups: Fresh Water - less than 1,000 mg/L,
Brackish Water -between 1,000 and 25,000 mg/L, Seawater - greater than 25,000 mg/L [11]. Most of
the water available on earth has salinity up to 10,000 ppm whereas seawater normally has salinity in
the range of 35,000-45,000 ppm in the form of total dissolved salts. According to World Health
Organization (WHO) latest guidelines (2008), water is considered to be good for consumption when
the TDS level is less than 600 ppm and for special cases goes up to 1000 ppm (based on taste
consideration). However, no health-based guideline value for TDS has been proposed due to the fact
that no reliable data on possible health effects associated with the ingestion of TDS in drinking water
were available [12].
Particularly for European countries, in the European directive 98/83/EC on the quality of drinking
water, TDS is no longer used as an indicator since it is very dependent on the time, season and water
source, and mostly affects the taste. Alternatively, Electrical Conductivity is given as a guideline and
the limit given for drinking water is 2500 S/cm [13].
The consumption of brackish water has been linked to poor health, including diarrhea [11], kidney
and gastric disorders as well as possibly diabetes [14]. High levels of salinity are associated with high
concentrations of other inorganic pollutants (e.g., sodium, sulphate, boron, fluoride), and
bioaccumulated elements (e.g., selenium, and arsenic). In some parts of Africa, China, and India, for
example, high fluoride content is associated with saline groundwater and has been found to cause
severe dental and skeletal fluorosis [5].

1.3. D ESALINAT ION -

THE SOLUTION TO FRE S H WATER SCARCITY

Under these circumstances, an increased trend towards use of desalination is observed around the
world as a means to reduce current or future water scarcity [3]. Although the costs for desalination
can be high because of its intensive use of energy, the cost to desalinate saline water is less than
other alternatives that may exist or be considered for the future in many arid areas of the world [7].
Effectively, this has led to the expansion of the desalination markets and the expectation to their
continuous expansion in the coming years particularly in the Mediterranean, Middle East and North
2

African (MENA) regions [15, 16]. Especially the use of desalination technologies to soften mildly
brackish water is increasing rapidly in various parts of the world [7]. A visual impression can be given
in the following map (Figure 2) where the distribution of installed desalination plants worldwide is
shown.

F IGURE 2: T OTAL

INSTALLED DESALINATION CAPACITY BY COUNTRY

(2006)[17]

The total capacity of desalination plants around the globe was 44.1 million m3/d by the end of 2006
according to the 20th GWI/IDA Worldwide Desalting Plant Inventory [17] whereas according to
the 22nd GWI/IDA Worldwide Desalting Plant Inventory the total installed capacity rose to
59.9 million m/d in 2009 corresponding to more than 14,451 desalination plants worldwide [18]. It is
estimated that this will be more than double, increasing to about 107 million m3/day by 2016 [19].
Nearly half of the current global desalination capacity is located in the Middle East led by the Gulf
Cooperation Council (GCC) countries, with the remaining capacity distributed throughout North
America, Europe, Australia and Asia. New markets are opening in China, India and the USA [20]. Until
2009, desalinated water is used as a main source of municipal water supply in many areas of the
Caribbean, North Africa, and the Middle East, a fact which proves the dependence of these areas on
desalination as a highly reliable, non-conventional source of freshwater [16]. Figure 3 outlines the
installed global desalination capacity for various feed water sources. Seawater desalination is being
applied at 58% of installed capacity worldwide, followed by brackish water desalination accounting for
23% of installed capacity [7, 8, 21].

F IGURE 3: G LOBAL

INSTALLED DESA LINATION CAPACITY BY FEED WATER SOURCES


M

( IN

MILLION

/ D )(2006) [17]

1.4. D ESALINAT ION

TECHNOLO GIES

Generally, desalination processes can be categorized into two major types: (1)phase-change
(thermal) and (2)membrane separation (concept of filtration). Some of the phase-change processes
include multi-stage flash (MSF), multiple effect distillation (MED), vapour compression (VC) which can
be thermal (TVC) or mechanic (MVC), freezing, humidification-dehumidification and solar stills.
Membrane based processes include reverse osmosis (RO), membrane distillation (MD) and
electrodialysis (ED)[7, 22]. The main desalination processes are shown in Figure 4.

F IGURE 4: M AIN

F IGURE 5: G LOBAL

DESALINATION PROCESSES

[7]

INSTALLED DESA LTING CAPACITY BY PR OCESS

(2006) [17]

Based on installed capacity for all source water types included, more than 80% of the world's
desalination capacity is provided by two technologies: Multi-stage flash (MSF), and reverse osmosis
(RO) (Figure 5). The MSF process represents more than 93% of the thermal process production while
RO process represents more than 88% of membrane processes production [3].
From Figure 5 for all sourcewater types, it can be seen that RO is the prevalent desalination process.
It accounts for slightly more than half (51% or 22.4Mm 3/day) of the globalcapacity.
Forty percent or 17.7Mm 3/day of the global production of desalinated water comes from
distillation plants, either using the MSF or the MED process. The picture changes if one distinguishes
between the different source water types. Thermal desalination processses acocount for 61% (17.2
Mm3/d) of the production in all desalination plants that use seawater as raw water source, of which
50% is produced in MSF plants. Only 35% of the water comes from ro seawater desalination plants.
On the contrary, RO accounts for 84% and 79% of the production in brachkish water and in
wastewater applications, respectively.

1.5. R EVERSE

OSMOSIS

THE LEADER IN DESALI NATION

Among desalination technologies, reverse osmosis (RO) is rapidly overtaking thermal desalination in
terms of market shares and is forecast that this will continue with 59% of the total new built capacity
being membrane based (mentioned in literature of 2004) [3, 22].
One of the reasons for this is that reverse osmosis is commercially available in a range of sizes and is
one of the most efficient technologies having much lower specific energy consumption (SEC) (about
310 kWh of electric energy per m3 of fresh water produced from seawater) than the average of
desalination technologies (compared to MED and VC) [22, 23]. On top of that, the high share of
recovered product water (up to 55%), the modularity of the systems, the low unit investment costs
and the flexibility in site location, start-up and shut-down all add to the advantages of RO process
[23, 24], making it the best alternative, especially for applications in remote, often off-grid, areas
with small and medium local water demand, such as islands or isolated villages in coastal areas [9,
25].
During the past decade especially, two improvements have helped reduce the operating costs of RO
plants and thus the cost of water producedthe development of membranes that can operate
efficiently at lower pressures, and the use of energy recovery devices [7].
For all the above reasons, reverse osmosis is becoming the technology of choice with continued
advances being made to reduce the total energy consumption and lower the cost of water produced.
[3]

1.5.1. R E VE R S E

O S M O S IS P R I N C IP LE

F IGURE 6: P RINCIPLE O F O SMOSIS A ND R EVERSE O SMOSIS P ROCESS [26]

Osmosis is a natural process of flow through a semi-permeable membrane. When pure water of the
same temperature is present on both sides of a membrane and the pressure on both sides is also
equal, no water will flow through the membrane. However, when the salt on one side is dissolved into
the water, a flow through the membrane from the pure water to the water containing salts will occur.
(Figure 6).Nature tries to equalize concentration differences [26].
Reverse Osmosis (RO) is a pressure-driven process that under high pressure forces salt water against
semi-permeable membranes so that water molecules can pass through membranes while the salts are
retained, when pressure is exerted on the side where the salts are added (Figure 6). The result is two
different flows one of freshwater permeate and one of concentrated brine. The system flow rate is
5

proportional to the difference between the applied - osmotic pressure differential between the brine
and the dilute sections [3]. The membrane itself represents a major pressure differential to the flow
of fresh water and as it is stated in the literature, commercially available RO membranes can retain
about 98-99.5% of the salt dissolved in the feed water [3]. The major energy requirement is for the
initial pressurization of the feed water [22].
Based on different literature and applications, operating pressures required for the desalination differ
with the salt concentration of the feed flow. Thus, for brackish water (BW) desalination the operating
pressures range from 15-30 bar [22, 24] (other literature mentions a range of 10-15 bar [3]), and for
seawater desalination (SW) from 55-70 bar [22] (while other literature gives a range from 55-80 bar
[24]). As desalinated water permeates across the membrane, the feed water becomes more and
more concentrated.
The amount of freshwater that can be recovered from the feed is limited by osmotic pressure and
scaling. Overall water recovery rates for seawater RO systems are typically 45-50% [3], whereas in
other references, recoveries are mentioned to be from 25 to 45% [2]. Brackish water RO plants have
recovery rates as high as 70% or even 90% if antiscalants are used [16, 22].
The disadvantages of the RO process are the sensitivity of the membranes to fouling, the high costs
of maintenance and repair, the risk of disruptions in supply, and the lower product water quality
(compared with thermal processes) [24].
Two major factors controlling the energy requirements of an RO system are membrane properties
and salinity of the feed water. Higher water salinity requires more energy to overcome the osmotic
pressure, whereas the RO system needs only mechanical power to raise the pressure of feed water
[22].

1.6. D ESALINAT ION S YSTEMS - S YNERGY

WITH

R ENEWABLE E NERGY

Despite the advances in desalination, the energy required to run these plants remains a drawback,
especially when it is supplied using conventional energy sources. However, the large ecological
footprint of fossil fuels and their fast on-going depletion in present times have led to a growing
interest in renewable energy sources. These technologies have certainly advanced technically over
the last quarter century to the point where they should now be considered clean-energy alternatives
to fossil fuels.
Many countries have already initiated the transition of their electricity supply schemes to higher
renewable energy shares, by supporting market introduction and expansion of those technologies.
The European Union set a goal to double its renewable energy share until 2010, and the
intergovernmental panel on climate change has recommended a worldwide reduction of 75% of
carbon emissions by the end of this century in order to avoid dangerous, uncontrolled effects on
climate and on the world's economy.
Fortuitously, renewable energy (RE) has unique synergies in regions where desalination is needed.
Many places all over the world which experience water scarcity like the Egyptian desert [35], rural
areas of Jordan [36], remote communities in Australia (e.g., [37]), Sicily [38], Ireland [86], and India
[87] have a high solar and/or wind energy potential [22]. It is expected that RE systems will flourish
in future as this synergy is exploited and play an important role in brackish and seawater desalination
in developing countries due to low operating and maintenance costs [3].

1.6.1. RE

S Y S T EM S IN D ES A L IN AT I O N

Renewable energies for use in desalination processes can include wind, solar thermal, photovoltaic
and geothermal. Among the several possible combinations of desalination and renewable energy
technologies (Figure 7), some are more (or less) promising than others in terms of economic and
technological feasibility.
Their applicability strongly depends on the local availability of renewable energy resources and the
salinity of feedwater, as well as on the remoteness of the region in consideration, accessibility to the
grid and technical infrastructure [22]. Moreover, the size of the plant plays significant role; some
combinations are better suited for large size plants, whereas some others are better suited for smallscale applications.

F IGURE 7: P OSSIBLE U SES O F R ENEWABLE E NERGY S OURCES W ITH D ESALINATION S YSTEMS [3]

In 2010, in the framework of ProDes Project financed by the Intelligent Energy for Europe
programme [27] an updated information collection of 131 RE-desalination systems installed from
1974-2009 was presented. Some of them were installed as pilot installations and have been dismantled after some years of operation, but most of the installations are providing drinking water and
are used by the local populations.
Based on this collection, ProDes [27] made an overview of the most common or promising REdesalination technologies, including typical capacities, energy demand, estimated water generation
cost and the development stage (Table 1Error! Reference source not found.). As is claimed by
he authors, most technologies have already been tested extensively and the water generation costs
are estimated based on operational experience and real data.
RE-desalination systems are currently acknowledged as the most promising for remote regions in
most literature sources studied, where connection to the public electrical grid is either not cost
effective or feasible, and where water scarcity is severe.
In references of 2008 [22] it is mentioned, that RE-RO is most often chosen as it is one of the most
efficient in terms of energy consumption. Some RO plants are particularly suited for small
communities in remote locations, although others may find large-scale applications. Figure 8 shows
the distribution of renewable energy powered desalination technologies [3]. As can be seen from the
graph, photovoltaic-powered reverse osmosis (PV-RO) is the most popular design option and is

considered by many one of the more promising approaches, particularly for small systems where
other technologies are less competitive [24].

T ABLE 1: O VERVIEW

OF THE MOST COMMON

RE - D ESALINATION T ECHNOLOGIES [27]

Solar Still

Typical
capacity
(m3/d)
<0.1

solar MEH

1-100

thermal

2-5

Solar MD
Solar/CSP
MED

0.15-10

thermal

8-15

applications
applications/advanced
R&D
advanced R&D

>5000

thermal

1.8-2.2

advanced R&D

PV-RO

<100

electrical

PV-EDR

<100

electrical

applications/advanced
R&D
advanced R&D

Wind-RO

50-2000

electrical

Wind-MVC
Wave-RO

<100
1000-3000

electrical
pressurised water

BW: 5-7
SW: 9-12
BW: 8-9
Units < 100 m3/d:
BW: 3-5
SW: 5-7
Units of 1000 m3/d:
1.5-4
4-6
0.5-1 (prospective)

energy demand

water production
cost (/m3)

technical
development stage

solar passive

1-5

WMVC
5%

Other
15%

applications/advanced
R&D
basic research
basic research

Solar MED
13%
Solar MSF
6%

PV-RO
32%

Wind RO
19%
PV-ED
6%

F IGURE 8: D ISTRIBUTION

1.6.2. RE- D E S AL IN A T IO N

OF

RE- POWERED

Hybrid
4%

DESALINATION TECHNOLOGIES

[3]

D I S AD V A N T AG ES

As stated in [7], until 2002 desalination systems using renewable energy sources (RE) had been
scarce and of limited capacity representing about 0.02% of the total desalination capacity. Although a
very large amount of work has been conducted in this field because of significant financial - social
benefits of RE-desalination systems (including plant design and implementation, mathematical
models, and economic feasibility) - only a few are currently being used. Most of the desalination
plants are proposed for the purpose of providing drinking water to small communities, especially
remote ones.
8

The reasons for this are related to various, often correlated, aspects such as:
(i)
(ii)

(iii)

(iv)

Availability, where the geographical distribution of RE potential does not always comply with
the water demand intensity at a local level,
Costs, where the initial capital installation costs and various system components are still
expensive. Even though prices decrease continuously still in many cases they are still
prohibiting for commercialization,
Technologies involving the combination of energy conversion and the desalination systems
are still faced with the challenge of optimal design offering high efficiency at the required
volumes, thus decreasing costs,
Sustainability, where in most of the cases, the maturity of the associated technologies does
not match the low level of infrastructures which often characterizes places with severe water
stress. Experience has shown that several attempts to integrate advanced desalination
solutions in isolated areas failed due to lack of reliable technical support (adapted from [28]).
Conversion of renewable energies, including solar, requires high investment cost and though
the intensive R&D effort technology is not yet mature enough to be exploited through largescale applications [3, 28].

The real problem in RE-desalination technologies is the optimum economic design and evaluation of
the combined plants in order to be economically viable for remote or arid regions. The slow
implementation of renewable energy projects especially in the developing countries is mostly due to
the government subsidies of conventional fuels products and electricity. The economic analyses
carried out so far have not been able to provide a strong basis for comparing economic viability of
each desalination technology. The economic performances expressed in terms of cost of water
production have been based on different system capacity, system energy sources, system
component, and water source. These differences make it difficult, if not impossible, to assess the
economic performance of a particular technology and compare it with others. [3].
Nevertheless, the cost reduction of renewable energy systems has been significant during the last
decades. It is estimated according to literature that future reductions as well as rise in fossil fuel
prices will make seawater desalination driven by renewable energies more competitive in the years to
come [3].

10

2. S OLAR

DESALINATION AND

PV-RO

SYSTEMS

2.1. I NTRODUCTION
2.1.1. H IG H

WA T ER S CA R C IT Y I N AR E AS W I TH H IG H IN SO LA T IO N

As mentioned before, high water scarcity is a problem faced in remote regions in the Middle East and
North African countries (MENA), the southern European Mediterranean Islands (SEMI) and isolated
communities in deserts like in Australia. Fortunately, the climatic conditions in these areas are very
favourable for potential exploitation of solar energy. This fact can be seen from the following two
maps (Figure 9: Average 3-year insolation from calibrated data collected by the NASA International
Satellite Cloud Climatology Project (ISCCP) (1991-1993) [29] (Figure 9 and Figure 10) which depict
the spatially resolved solar irradiance in W/m 2 and peak sun hours.

F IGURE 9: A VERAGE 3- YEAR INSOLATION FROM CALIBRATED DATA COLLECTED BY THE NASA
I NTERNATIONAL S ATELLITE C LOUD C LIMATOLOGY P ROJECT (ISCCP) (1991-1993) [29]

F IGURE 10: S OLAR P EAK H OURS

ON AN OPTIMALLY TITLED SURFACE DURING THE WORST MONTHS OF THE


YEAR

[30]

11

Consequently, solar desalination systems and pilot plants have been installed around many of the
darker coloured places in the above maps like Saudi Arabia, Brazil, Australia, Florida, Gran Canaria,
Egypt, Gaza, Jordan, Eritrea, Sicily, Greece, USA, Mexico, Tunisia etc.

2.1.2. PV- P O W ER ED RO

S Y S TE M S M O S T CO M M O N I N

RE- D ES A LI N A T IO N

The feasible solar energy-desalination technologies combinations have been described previously. The
combination of RO membranes and arrays of photovoltaic (PV) modules is the design option that has
been implemented most frequently in solar-driven RO desalination systems. As shown in Figure 8 PVRO systems account for 32% of the total installed capacity, while in PRoDes project [27] it is
estimated that PV-RO systems comprise 31% among combinations of 131 RE-desalination plants
reviewed in 2009.
However, the main problems nowadays are that RO presents a significant requirement for chemicals
and spare parts, and that PV panels still represent significant capital investment. Most importantly
however, PV-RO systems require some degree of technical skill to operate, involving understanding of
means to protect the membranes from fouling and to maintain pumps. For these reasons, even as the
feasibility of PV-powered RO systems as a valid means of desalination has been proven, they still
cannot compete favourably with fossil fuel-based desalination systems [31]. Until some years ago, the
cost for photovoltaics was a major constraint against PVRO systems. Yet, this situation has changed
with advance in PV technology.
Despite the above disadvantages, from a technical point of view today, PV as well as RO are mature
and commercially widely available technologies. RO is modular and compact and has proved to be the
lowest energy consuming technique, using nearly half the energy needed for thermal processes. This,
supplemented by the modular nature of PVs, their low environmental impact as well as the ease of
operation and maintenance are incentives for this combination of technologies to be used, especially
for application in remote areas [32].
The ADIRA project [31], funded by the European Union, examined the application of Autonomous
Desalination Systems (ADS) supplied by RE in several rural and other remote areas. Because of the
high cost of PVs, PVRO systems were only recommended for low quantities of output water either
from brackish or seawater feed.
The extent to which PV energy is competitive with conventional energy depends on the plant
capacity, on the distance to the electric grid and on the salt concentration of the feed [33]. Based on
that, many authors agree that photovoltaic powered reverse osmosis (PV-RO) systems may be the
only technically and economically competitive alternative for small and standalone applications with a
capacity up to 50 m3/day to provide drinking water in remote areas where access to fuel, electricity,
and technical expertise is not available [7, 24, 31, 33, 34]. The continuously decreasing capital cost of
PVs and RO units is also helping the feasibility of such systems.

2.2. PV- POWERED RO S YSTEM C OMPONE NT

ANALYSIS

It has already been mentioned that the PV-RO system is regarded as the most promising approach
for desalination particularly for small systems in remote areas where other technologies are less
competitive.
Although research has been conducted concerning the PV-RO system design since the 1970s, a
standard design approach has so far not emerged. Design solutions may either include or omit energy
storage, inverters, energy recovery devices, etc. Despite the lack of standardization, however, several
components are common to all design approaches [16].

12

Generally, the major common components of PV-RO plants are [31]:


Photovoltaic modules
Pre-treatment units, which remove the dissolved and large suspended solids from the
feedwater prior to flowing through the membranes. This is done to protect the membranes
and to reduce salt deposits that can diminish the efficiency
high pressure pumps, which increase the feed water pressure on the membrane to the point
that it exceeds the osmotic pressure, thereby providing enough energy to move water across
the membrane
Reverse Osmosis Membranes
Post treatment units, in which the acidity of the water is neutralized and chlorine added to
disinfect it
A simplified, general design scheme for PV-RO desalination systems as given in literature [16] is
depicted in Figure 11. Blue lines show the water flow whereas the red line shows the recovery of the
concentrate stream. Dashed lines identify components and connections that may be absent.

F IGURE 11: S IMPLIFIED

GENERAL DESIGN OF A

PV-RO

DESALINATION PLANT

[16]

The most important components used by PV-RO plants will be discussed a bit more elaborately in the
following sections:

2.2.1. PV

M O D U L ES

In PV-RO desalination, the direct current (DC) electrical energy generated in the solar cells by silicon
or other semi-conductors is useddirectly or after regulationto power the pumps that generate the
pressure required for the feed water to permeate across the RO membranes.
Characteristics
PV panels today constitute the fastest growing renewable energy market. Present industrial
production focuses primarily on the production of (mono- and poly-) crystalline silicon and thin film
amorphous silicon cells. Over the course of several technological developments, the conversion
efficiencies of PV modules have reached 13-16% for poly- and mono- crystalline silicon cells and 610% for thin film solar cells. Until recently, the high price for PV modules was also prohibitive in their
application in PV-RO systems. However, the prices of PV modules have been falling rapidly in recent
times year by year, a fact which means that the capital cost of PVs are no longer a major constraint.
PV modules need hardly any maintenance.

13

Until now, both mono-crystalline and multi-crystalline silicon modules have been used in PV-RO
experimental units but no thin film solar cells have yet been used. This fact has mostly to do with thin
film cells having; lower efficiency in normal light conditions (30% less efficient that single crystal)
and uncertain durability. Lower efficiency means that more space and mounting hardware are
required to produce the same power output, and this is a problem when space is a constraint. Also,
thin film materials tend to be less stable than crystalline ones, degrading over time.

2.2.2. I N V E R T ER
Although PV panels supply DC current, many of the PV-RO systems use AC motors and require
inverters to convert DC current to AC. Reliability issues have been reported with inverters and power
conditioning equipment in grid connected and AC systems. Unscheduled site visits for maintenance
are required in such cases, in order to replace the complete inverter, or fuses, reset circuit breakers
etc [31].

2.2.3. T R A C K IN G

SYSTEM

Tracking and module orientation has been widely recognised as an important factor in determining
the electrical power output and thus the overall performance of the desalination plant. While modules
with fixed axes are tilted at a constant angle, modules with adjustable axes can be manually
repositioned based on seasonal changes, or, if a tracking system with controller and drive motor is
installed, the modules can automatically follow the sun's daily path in the sky. It has been estimated
by researchers that utilizing the seasonal tilt angle variation increases the yearly average permeate
flow of a PV-RO desalination plant in Saudi Arabia from 15 to 17 m 3/d [16]. In Jordan gains in
electrical power output and permeate flow of 25% and 15% respectively have been reported when a
one-axis automatic tracking system was used rather than a fixed tilt plate [16]. In some applications
in Australia [4, 15] 30% more solar radiation was measured due to the single-axis tracker while in
other cases 60% higher permeate flow was produced because of tracking [16]. The model of the
complete PV-RO system developed in CREST research institute in Loughborough, UK [28] compared
the gains between a single-axis and a dual axis tracking system. The model showed that single-axis
tracking would increase the annual freshwater production by some 33% while dual-axis tracking
would provide a further increase of only 3%. Furthermore, it was concluded that only the single-axis
tracker is economically justified since it doesnt require automation like the dual-axis mechanism.
All in all, the advantages of the mounting of PV modules on trackers are proven. Nevertheless, the
high initial investment costs required to install tracking systems, have so far limited their use in PVRO desalination.

2.2.4. MPPT

AN D

LCB

Maximum power point tracker (MPPT) circuits are basically DC-DC converters, generally used [4, 14,
28] to control the current drawn from the PV array and to maintain the system operation at a voltage
that achieves maximum power output (Figure 12) under varying conditions of irradiance and module
temperature [16, 21]. It is obvious from the graph that this voltage corresponds to a lower current
than the maximum current and the lower the irradiation, the lower the operating voltage. So, in case
the irradiance is very low, the DC/DC converter called (MPPT) will find, ideally, the optimal operation
point for that irradiation. This means that the motor will operate between certain voltage limits which
will correspond to the maximum power.
The MPPT algorithm has been used in many desalination pilots, and ensures efficiency under
fluctuating conditions. In 2002, A. Schafer and B. Richards [14] successfully tested a small batteryless
PV-RO prototype (power consumption 150 W and water production of just 500 l/d in a remote region
in Australia with the use of MPPT). It was the first time that an MPPT was used without use of
batteries but the authors did not specifically mention any information regarding its usage. At the
14

same year, CREST institute in Loughborough, UK [28] developed an MPPT algorithm specifically for
use with standard industrial inverters and without need of any additional sensors or calibration, in
Matlab/Simulink.

F IGURE 12: S OLAR C ELL I-V

CURVE WITH

MPPT

IN VARYING SUNLIGHT

[35]

MPPT controllers for use in battery-less systems are not available commercially as of yet. MPPT
circuits exist mostly in the market as charge controllers which are connected with batteries. The only
available product for direct connection of MPPT between a PV panel and a pump is made from the
Australian energy research laboratories (AERL), however with very limited current output (16A),
which can be used for very small PV systems [36].
Linear current boosters (LCB) are also DC-DC converters mainly used in solar direct pumping
applications. They can achieve 30% increase in the water pumped than when the motor is connected
directly to the solar panels. An LCB basically exchanges voltage for current providing more current
produced by solar panels in order to use the maximum power for a particular time. LCB thus also
prevents stalling of the motor under less than full sun conditions. As the solar irradiation drops, the
output voltage of the LCB drops instead of the current, thus resulting in slowing down of the motor
(since the speed is proportional to the voltage), but continues to provide the same torque (the torque
being proportional to the current), thereby preventing stalling. By examining the effect of the
magnetic field in the motor, and realizing that magnetic flux is constant, we can arrive at the
following two equations relating the torque and speed output of the motor to the supplied current
and voltage:
E

Kv
Km I

Where: (in S.I. Units)


E = Motor Voltage
= Motor Speed
T = Torque developed by the motor
15

I = Motor Current
These are often known as the transducer equations for a motor, since a motor is really an electromechanical transducer. The constants Kv and Km are dependent on the particular motor. Similarly, the
motor starts running much earlier in the morning at a lower speed (thus providing some useful
output) instead of staying stalled until full sun. This translates into more running time of the motor
where it spends a lot of time working instead of stalled doing no work [37].
Although its application has been rarely found in research publications, it is commonly used in almost
every private PV-direct water pumping application (particularly in America and Canada). LCB has been
noted to have been used by B. Richards and A. Schafer in 2002 in design considerations for a PVpowered desalination system for remote communities in Australia [38]. Although they consider the
use of an LCB or MPPT essential for efficient operation when the PV array is connected directly to the
pump, there is no evidence in literature about the results of a field test with it. In contrast, 2 LCB
units from Solar Converters, Inc. (Canada) have been used in a PV-electrolyser system for hydrogen
production in order to improve the efficiency of the system [39].

2.2.5. P R E - TR E A TM EN T U N I T
A major consideration in the design and operation of any RO system is the avoidance, or at least
management, of fouling and scaling of the membranes, since this determines the frequency of
required membrane cleaning and replacement, since the rate of membrane fouling and scaling is very
dependent upon feed-water quality and pre-treatment [40]. The use of an adapted pre-treatment
minimizes the fouling problems and can provide good protection of the membranes and a longer
lifetime [22]. Particularly, suspended solids and larger particulates are important to be removed as
they can damage the membranes when deposited on the surface of the membranes causing clogging
of spacers and increasing the resistances to the flow [14].
Depending on several parameters which influence the choice of the pre-treatment like dissolved
organic carbon, SDI, turbidity, algae content and their evolution during the seasons, and
temperature, the pre-treatment can comprise different technologies, such as conventional pretreatment (i.e. ballasted sedimentation, air flotation, dual-media filtration, mono-media filtration,
double stage filtration) or advanced technologies including membranes coupled with a conventional
process [22]. Specifically, the pre-treatment stage often consists of either 5 or 20 m filters, or sand
filters. An interesting arrangement is the use of a beach well, which by using the sand of the beach,
can provide pre-filtered water and greatly reduce pre-treatment requirements [40].
Nevertheless, some references suggest an alternative solution for smaller systems, which involves
system operation at low recovery rates to prolong membrane viability and to reduce the costs [16,
24, 41]. It is argued that in systems using brackish feed waters, where scaling is a problem, fouling
can be greatly accelerated by use of excessive water recovery ratios, and this sets a limit on the
recovery ratio. In sea-water fed systems, on the other hand, the recovery ratio is limited primarily by
the osmotic pressure in the last element, where the feed water is most concentrated. Operating at a
lower recovery ratio will have little effect on the biological fouling, which tends to occur in the first
element [41].
UF pre-treatment has been used in some BW applications [4, 15, 25] in Australia. In Coober Pedy in
Australia, the experiments showed that an RO plant utilizing UF pre-treatment is a promising
alternative. Since the feedwater is disinfected physically using ultrafiltration (UF), the brine is free
from bacteria and most viruses and hence can be seen more as a reusable feed stream than a waste
stream with a disposal problem [4]. However, UF pretreatment involves higher investment costs than
conventional pretreatment, but because it removes significant numbers of microorganisms and

16

generally delivers higher qualify RO feed, eliminates the need for membrane disinfection. Also UF
pretreatment may reduce RO membrane cleaning and replacement costs [16].

2.2.6. H IG H

P R ES S U R E P U M P

The high-pressure pump supplies the pressure needed to enable the water to pass through the
membrane and have the salts rejected. There are a vast number of possible configurations of motors
and pumps etc. that have been used in reverse osmosis systems coupled with photovoltaics around
the world. This wide range of choices is due to the different costs, efficiency and performances,
seawater compatibility and level of simplicity and other practicalities depending on the capacity and
on the year that each facility was installed.
Centrifugal pumps are regarded inappropriate to couple with PVRO systems. In order to achieve
optimum efficiency with a centrifugal pump, the rotor speed must be matched to the flow/pressure
operating point. However, this is not the case in PVRO systems where the fluctuations in irradiance
influence the flow and the pressure of the system. Due to their construction, they also cannot
manage the water recovery ratio with that injection system [41].
Due to the above mentioned reasons, positive displacement pumps are used as a rule because of
their higher energy efficiencies with respect to centrifugal pumps at the required flows and pressures
[16]. Based on the references which have been studied in the framework of this report, both rotary
positive displacement pumps (e.g., rotary vane [14, 42, 43] and progressive cavity pumps [4, 15, 25,
28, 41]) and reciprocating positive displacement pumps (e.g., piston [24, 28, 41, 44, 45]and
diaphragm pumps [46]) have been used.
Progressive cavity pumps are often used in PV-powered systems, however this is mostly in BW
systems (low salinity) since they are limited to a certain pressure (10-30 bar). Literature cases of BW
systems have reported this type of pumps to fully meet the targets set. Example of a progressive
cavity pump is the custom-designed pump for the projects of Schafer and Richards [4, 25] from
Mono-pumps manufacturer in Australia). However, in 2009 [15], the above mentioned authors
reported failure of the tested system to meet water quality standards when fed with high salinity
water, due to the pump pressure limitation. Another disadvantage reported in literature, is that this
kind of pumps can suffer from starting problems due to the static friction between the rubber stator
and the metal rotor - once turning, the water being pumped acts as a lubricant. However, it is
claimed that starting should not be a problem in a system employing a variable-frequency inverter
[41].
The amount of power, or more specifically current, required to drive the pumps is directly
proportional to the operating pressure. Higher the pressure, higher is the permeate flow. One of the
main operating characteristics or rotary vane high pressure pumps is that the permeate flow shows
very little dependence on the pressure which leads to essentially constant permeate flow. Dankoff
Solar Slow pumps are rotary vane pumps which have been used in PV-RO brackish water desalination
systems [42]. The pump was reported in literature [42] to work fine coupled directly off solar panels
with very few maintenance problems, and with motor brush replacements only every 3-5 years. The
small capacity (2.2 gpm and maximum pressure 17 bar) was mentioned to be one of the
disadvantages of the pump.
The Clark pump, is a positive displacement reciprocating pump (piston) that was specifically
developed for energy recovery in small desalination systems and was used in several seawater PVRO
applications since 2002 [28, 43, 45, 47]. Pearson pump, the successor of the Clark pump, is
nowadays the state-of-the-art pump technology in small reverse osmosis desalination systems (See
below energy recovery section).

17

F IGURE 13: S PECTRA C LARK P UMP

AND

S PECTRA P EARSON P UMP [48]

Most RO plants usually use AC for the pump, which means that DC/AC inverters are required [7].
However, there are systems that have been reported to have used DC pumps, to avoid energy losses
from the conversion DC-AC-DC [15, 24, 42, 43].
More recently, many PVRO plants chose to couple photovoltaics directly to high pressure pumps. In
general terms it is argued in most of the cases, that direct pump connection to the PVs eliminated the
need for inverters and batteries, resulting in low-cost, simplified and energy-efficient systems [4, 14,
15, 25, 28, 42].

2.2.7. E N E R G Y R E CO V ER Y

D E VI C E S

In conventional small Reverse Osmosis systems the high pressure needed in the membranes is
created and regulated by a back pressure regulating valve in the high pressure concentrate (brine)
discharge line. This pressurization of the saline water accounts for most of the energy consumed by
an RO system. All of the potential energy in the high pressure concentrate is lost in the back pressure
regulator, and the power required producing that energy is wasted. Consequently, the energy
efficiency of a seawater reverse-osmosis system is heavily dependent on recovering energy from the
highly pressurized brine stream.
In some cases, Danfoss axial piston motors have been reported to be used for energy recovery [41,
45]. In large reverse osmosis systems, pelton-wheel turbines, reverse running pumps and pressure or
work exchangers (mostly ERI type) have been used to recover energy [14, 24, 49]. According to the
company, ERI's PX Pressure Exchanger(R) device is a rotary positive displacement pump that
recovers energy from the high pressure reject stream of SWRO systems at up to 98% efficiency with
no downtime or scheduled maintenance [50]. Nevertheless, neither is available in small sizes.
The innovative Clark pump, developed by Spectra Watermakers [48] was the first pump for small
desalination systems which incorporated energy recovery and a pressure amplification innovation.
CREST institute (Loughborough) and Thomson tested the Spectra Clark pump with variable flow and
pressure conditions (using solar irradiation data) and concluded that the Clark Pump is very well
suited to a batteryless PV-powered system [41].
In 2009, Spectra Watermakers developed the Pearson Pump, which was a breakthrough in the
evolution and enhancement of pump design especially suited for small applications mainly because of
the incorporated energy recovery mechanism. It is a positive displacement three-cylinder
reciprocating high pressure pump with the same motors and crankcases used in conventional RO
system feed pumps. The Pearson Pump head delivers water to the membranes in the same way as
conventional feed pumps but is capable of recovering the energy in the concentrate stream. This is
done by returning the concentrate to the Pearson pump at high pressure, where it flows into the
pump cylinders on the undersides of the pistons, transferring its energy to the feed water entering
the membranes (Figure 14). The energy recovered from the concentrate leads to reduced load on the
pump motor, reducing the electrical consumption dramatically (to an impressive 2.64 kWh/m) [51].

18

F IGURE 14: D ESIGN

OF THE

P EARSON P UMP [48]

Fixed recovery ratios is another characteristic of the Spectra technology which differentiates it from
the conventional constant-pressure systems. The creation of the high pressure in a Spectra system is
not based on a back pressure regulating valve at the exit of the membrane, but on the innovative
pump design: A remarkably bigger size of the ceramic plunger occupies a significant part of the
volume in the underside of the piston thus reducing the space available for the returning concentrate
(Figure 14). Since feed water is forced out of the cylinder by the upper side of the piston, and only a
portion of that water can return to the underside, a Hydraulic Lock is created, which induces high
pressure in the membrane. When the pressure rises high enough water is forced through the
membrane as permeate.
The proportion of the feed water discharged by the Pearson pump which permeates the membrane
and becomes product water is fixed by the percentage of the volume of the underside cylinder taken
up by the plunger. If the plunger takes up half the volume of the cylinder, then only 50% of the feed
water discharged by the Pearson Pump can return to the pump, and the other 50% must leave the
system as permeate. Thus, a Pearson Pump based system will be a fixed recovery ratio system, and
the operating pressure will vary with pump speed, feed water salinity, and feed water temperature
but the percentage of feed recovered as permeate will always be constant. Conventional systems are
constant pressure systems, where the back pressure regulator keeps the system at a fixed pressure,
and the recovery ratio will vary according to operating conditions [52].
In published literature however, use of the Spectra Pearson pump with PV-RO systems was not
found. However, the pump has been marketed in many solar and wind-powered land-based
desalination units for military, disaster relief, village level water supply, small eco-resort and remote
home applications [51]. In 2009, Brett Ibbotson used the Spectra Pearson pump to examine the
effects of fluctuating operation on reverse osmosis membranes, as part of his master thesis. After
more than 650 hours of fluctuating operation, he concluded that no deterioration of reverse osmosis
membrane performance was observed. From his experiments, SEC ranged between 2.66-2.86
kWh/m3 and TDS between 518-638 ppm [1].

2.2.8. M EM B R AN E

MODULE

The most common RO membranes used in desalination are spiral wound Thin Film Composites. They
consist of a flat sheet sealed like an envelope and wound in a spiral. In a reverse osmosis system, the
higher the pressure, the higher is the permeate flow and also the energy consumption. PV-RO
desalination systems are hence designed with generous membrane areas since for a fixed recovery
rate they can operate at lower pressures and thus with a lower energy consumption [41].
Nevertheless, although large membrane areas compensate for the reduced product flow with
decreasing operating pressure, they result in decreased water quality.
RO membranes are specified in terms of their rejection of NaCl at a specified pressure. Rejection
ranges vary from 95%-99.8% depending on the source of feed water. In seawater applications,

19

rejections values higher than 99.4% are used [53], while a membrane with rejection 95% is
acceptable for desalination of brackish water.
The most common RO configuration is single pass, in which the membranes are organised in series
within one or more pressure vessels [16].
After the introduction of battery-less PVRO systems, there have been several discussions about the
effects of fluctuating operation on seawater reverse osmosis membranes. Membrane manufacturers
suggest that an important factor for the lifetime expectancy of the membranes or RO plants is the
constant flow and pressure operation of the plant. For this reason, they have been reluctant to give
any warranties regarding their products under variable pressure-flow conditions. From various tests
conducted by different researches, the operation of PVSWRO systems under variable pressure flow
conditions indicated no serious problems [1, 41]. Schafer et al (2008) [25], evaluated the effect on
the performance of a PV-(BW)RO and concluded that the membranes tolerated large fluctuations in a
wide range of solar irradiance (500-1200W/m2), resulting in only small increases in the permeate
conductivity. One important point from the above research was the special care for the feedwater
pre-treatment to avoid fouling of the RO membranes.

2.2.9. B A TT ER I E S
Solar power is intermittent and variable while RO plants are generally designed to run continuously
and at constant load. However, constant power supply can only be maintained through energy
storage or backup systems. Typically conventional lead-acid batteries have been used as energy
buffers during day time or for extending the hours of operation beyond sunset for photovoltaicdesalination systems. There has been a lot of debate over the years about the suitability of batteries
as energy storage in PVRO systems. Many authors and membrane manufacturers often propose
batteries in order to address the problem of variable operating conditions (pressure and flow) that
follow solar fluctuations. On the other hand, others argue that lead-acid batteries (accumulators)
which have been used until now in PVRO systems are very troublesome and associate them with
various problems such as:
Increase environmental risk in case of spillage (because of accidents or improper disposal in
remote areas where recycling is limited)
Worse performance and faster degradation at higher temperatures, a fact which is relevant,
since most potential sites for PV-RO applicability do experience high temperatures. With
increasing operating temperature, the charge efficiency decreases, eventually reducing the
number of battery lifecycles [25].
Because of the relatively low number of discharge-cycles, lead-acid batteries need
replacement every 2 years (on an average) [28]. This offsets the overall sustainability
benefits from the use of PVs.
Conversion losses when charging in/charging out, reducing the system efficiency (efficiency
of Pb-acid battery is about 75-80%) [25]; this leads to a greater power requirements, thus a
bigger PV-array. In literature [24], it has been argued that when batteries are used, the PV
system has to be at least 4 times larger for a 24 hr/day operation.
Extra addition to the capital cost of the system
Significant maintenance which add in costs (absence of careful maintenance is typical in
remotely located systems [16])
Complicated system if all auxiliary components such as charge controller and wiring are
considered [16]
The main advantage of batteryless systems is the simplicity since there are no batteries and charge
controllers, capital and maintenance investment is lower and environmental hazards are less. Batteryless PV-RO systems are based on the idea that water storage is often more efficient and cost20

effective than energy storage [15, 16]. The technical feasibility and short-term operation of variable
speed, batteryless PV-RO systems was tested in a series of studies [4, 15, 25, 28, 42, 47]. However,
long-term performance has not been monitored so far.
Thomson et al. from the CREST institute in Loughborough [28] in 2002 presented a model for a small
batteryless PVSWRO system, based on real data from the country of Eritrea and commercially
available components incorporating the innovative Spectra Clark pump and an MPPT algorithm. Based
on the model, the authors described the system as energy efficient and calculated the production cost
of drinking water to be 2/m3. In 2005, the same team [47], developed a prototype which was tested
in the laboratory in UK. However, results were inconclusive since permeate quality was beyond
guidelines and the reliability of the system was not proven. The authors remarked that system
reliability should be treated with equal importance as energy efficiency in PVRO systems.
In 2008, a comparative study between a batteryless and a battery PV-powered seawater system was
published by Mohamed et al [2]. The results obtained from this comparative experimental study
showed that although the battery system was more efficient, there was no real difference (less than
7%) in the SEC and water production between the battery and the batteryless system. The authors
concluded that economically, the extra costs for battery and charge controller were not justified.
The same year, Richards et al [25] in their research on the effect of energy fluctuations on the
performance of a direct coupled PV-RO system for brackish water, found that the system could
tolerate large fluctuations between 500-1200 W/m2).
In 2009, De Munari et al [15], tested a batteryless PVBWRO system in a remote area in Australia and
noticed that the power variations because of the lack of batteries do not affect the permeate quality
nor water production. However, it was indicated that results would not be the same for high salinity
water.
At last, the question of using batteries or not is very site and case specific. M. Thomson et al in 2008
argued that the underlying subject is identifying what is less troublesome and more cost-effective in
remote locations lacking skilled personnel where PVSWRO systems normally find their application,
replacing reverse osmosis membranes or replacing batteries [45].
Nevertheless, with the advance in battery technology recent years, new possibilities about energy
storage in batteries have evolved. Low-maintenance lead-acid rechargeable batteries called VRLA
batteries (valve-regulated lead-acid battery) with a lifespan of up to 10 years are commercially
available. Because of their construction, VRLA batteries do not require regular addition of water to the
cells. VRLA batteries are classified as absorbed glass mat batteries (AGM) and Gel batteries (gel cell).
These batteries are often colloquially called sealed lead-acid batteries, but they always include a
safety pressure relief valve. As opposed to vented (also called flooded) batteries, a VRLA cannot spill
its electrolyte if it is inverted. Because VRLA batteries use much less electrolyte (battery acid) than
traditional lead-acid batteries, they are also occasionally referred to as an "acid-starved" design.
However, cost is a constraint since they cost twice as the traditional lead-acid batteries while their
discharge cycle is less deep (around 50%).
In the market, hybrid Gel batteries, the 2 nd generation VRLA batteries already exist. Manufacturers
claim that they have better service life, especially deep cycle life and high rates of discharge
compared to standard AGM batteries.
Currently, AGM and gel batteries are used in commercial PVRO systems of Spectra Watermakers [54].

21

2.2.10. F E ED W A T ER I N L ET M E AN S
Spectra systems require that feed water enters the system at a pressure of about 1 bar (gauge). For
the purpose of the trials, it would be easiest to have the feed water tank at the same elevation as the
pump, which leads to requirement of a feed pump. However, in a real system, the pressure head
required at the inlet may also be obtained by placing the feed tank at an elevation (greater than 10m,
correcting for the friction factor) higher than the pump inlet.
An extreme concept was proposed by Murakami [55], regarding a hydro-powered pumped water
storage system for co-generation of electricity and desalinated water. The proposed scheme involved
pumping sea water to an elevation of 589 m (above sea level) using hydroelectricity. Water from the
pumped storage reservoir would be used to regenerate electricity for 4 hours in a day in times of high
demand. For the remainder of the operational period (18-20 hours), the pressure head of the water
would be used to produce a desalinated stream of water by using reverse osmosis membranes. The
elevation can be potentially used to generate a head of ~ 58 bar (disregarding friction losses) which
can be used to overcome the osmotic pressure of sea water.

2.3. S YSTEM R EVIEWS


Ever since the first technological breakthroughs on PV-powered RO desalination in the late 1970s,
vast experience has been gained with PV-RO systems from a large number of experimental units.
Although many authors in references agree that photovoltaic powered reverse osmosis (PV-RO)
systems are a proven technology nowadays, most of the units were lab demonstrations or pilot
systems which after their testing were abandoned [7]. The main reasons for this were the very high
capital cost and the high requirement for technical skills in order to protect the membranes and
maintain the pump [24].
The PV-RO technology has been implemented for the desalination of both brackish water (BW) and
seawater (SW), yet only in small and medium systems (less than 75 m 3). It is interesting to note that
until 2010 (as far as the author understands), no PV-RO experimental study has been done on a
larger scale. Most systems have been developed in different parts of the world with intense fresh
water scarcity but in the same time with abundant solar energy potential (e.g Mediterranean region,
Australia, North Africa and Middle East).
The parameters determining the performance of a PV-RO system depends on the specific energy
consumption (SEC), on the total daily water production and on the salt rejection. SEC fundamentally
depends on the salinity of the water [22]; The higher the salinity, the higher is the SEC and the cost
of the water produced [4, 15]. In addition, small systems operate much less efficiently, which leads to
even higher SEC values, leading to an even higher cost of the system [24].
Because of cost considerations, brackish water solar desalination became mature and proved to be an
economically viable technology very early in contrast with seawater solar desalination [2]. Seawater
desalination is a very energy consuming process and has much higher SEC than brackish water
desalination. Until 2002 research was focusing mostly on PV-(BW)RO systems and demonstration
plants. In addition to this, until 2002, most plants were using batteries to balance the energy
fluctuations from the sun.
The situation changed after 2002 for seawater PV-powered desalination, with the invention of the
Spectra Clark pump, incorporating energy recovery mechanism designed especially for small plants.
Energy recovery reduced the SEC for seawater PVRO desalination. However, seawater desalination
powered by photovoltaics is (potentially) economically competitive only for remote areas with no
22

access to the grid areas (electricity / water) [16]. Indicatively, based on literature [4, 16, 24], the
following table shows values for SEC depending on the feedwater quality.
T ABLE 2: I NDICATIVE SEC
R ECOVERY

VALUES FOR DIFFERENT

Feedwater Type
Brackish

Seawater

F IGURE 15: SEC

F EEDWATER T YPES

WITH AND WITHOUT

Energy recovery

SEC (KWh/m3)

No

<10

Yes

<2

No

>10

Yes

<5

FOR SEAWATER AND BRACKISH WATER

PV-RO

SYSTEMS

E NERGY

[16]

In Figure 15, different SEC values can be compared for brackish and seawater systems. Based on that
list, it is very interesting to see that the lowest seawater SEC value (for a real prototype and not a
theoretical estimate) is in Baja California Sur, Mexico. This case study incorporated the Clark pump by
Spectra Watermakers which tremendously reduced the SEC to 2.6 kWh/m 3.
In the comparison made by Mohamed et al. in 2008 [18] between a seawater PV-RO system with and
without a battery, it was shown that the SEC for a battery system was around 4.3 kWh/m 3 while for
the batteryless a bit higher (4.6 kWh/m 3).
It is difficult to compare between different experimental studies with respect to water costs, since
each one is based on different assumptions (variable depreciation times, currency exchange rates,
frequency of replacement of components, feedwater quality, plant capacity, etc.) and is site and year
specific [3]. In the following table an overview of existing PVRO systems is shown, yet it is not
reliable for conclusions due to the aforementioned reasons. Very old systems for brackish PVRO
(before 1990) have been omitted.

23

T ABLE 3: O VERVIEW
Location and
country

Year

Feed TDS
(mg/L)

OF

PV-RO

SYSTEMS

PV capacity,
(kW)

( ADAPTED
Battery
storage

FROM

[16])

Pump

Production
(m3/d)

SEC

Cost,
(US$/m3
)

Brackish desalination systems


Coite-Pedreiras, BRA
Denver, ITN, USA
Lisbon, INETI, PRT
Hammam Lif, TUN
Concepcion del Oro,
MEX
Mesquite, ITN, USA
Murdoch Univ., AUS
White Cliffs, AUS
Ksar Ghilene, TUN
Aqaba, JOR
Baja California Sur,
MEX
Pine Hill, AUS
Sadous, Riyadh, SAU
Lampedusa, ITA
Coober Pedy, AUS
Central Australia,
AUS
Kuwait, KWT

2000
2003
2000
2003
1978

1200
1600
2549
2800
3000

1.1
0.54
0.1
0.59
2.5

yes
no
no
no
yes

DC/AC
DC
DC
DC
DC

6
1.5
0.02
0.05
0.71

12.8
6.5
10.6
11.6
12.8

2003
2003
2003
2005
2005
2005

3480
3480
3500
3500
4000
4000

0.54
0.06
0.26
10.5
16.8
25

no
no
no
yes
yes
yes

DC
DC
DC
AC
AC
AC

1.28
0.05
0.06
7.0
58.0
11.5

3.6
3.6
9.0
6.5
9.8
9.8

2008
1994
1990
2008
2008

5300
5700
8000
BW
BW

0.6
10.08
100
3
3

no
yes
yes
no
no

DC
AC
AC
DC
DC

1.1
5.7
40
0.76
1.1

2005

8000

0.3

yes

DC

1.0

6.5

3.7
9.6
10.6
3.2
2.3

Seawater desalination systems


Agric. Univ., Athens,
GRC
CREST, GBR
Vancouver, CAN
Doha, QAT
Univ. of Bahrain,
BHR
Pozo Izquierdo, ESP
Jeddah, SAU
Baja California, MEX
CREST, GBR

2006

30000

0.85

no

DC

0.35

9.8

2001
1983
1984
1994

32800
33000
35000
35000

1.54
0.48
11.2
0.11

no
no
no
yes

AC
DC
AC
DC

1.45
0.86
5.7
0.2

3.0
9.0
3.0
2.8

2000
1981
2007
2003

35500
42800
SW
SW

4.8
8

yes
yes
no
no

AC
DC
DC
DC

1.24
3.22
19
3

9.6
6.5

2.4

2.6
3

Although the above comparative table is not reliable to draw conclusions, in general, we can observe
the following trends:
The older the system, greater is the cost of desalination (owing to systems with lesser SEC)
The lower the salinity, lower is the water cost (owing to lesser pressure requirements, and
hence a smaller SEC)
It is interesting to note that as far as seawater PV-RO is concerned there are few plants installed
worldwide and most of them have very high water cost. Reduction in costs of seawater solar
desalination appeared mainly after 2001 with the invention of the Spectra Clark pump incorporating
means of energy recovery, as mentioned in publications.
The cheapest water cost for a PV-(SW)RO system, was estimated by Thomson et al. from the CREST
institute in Loughborough [28]. In his model in 2003, he presented a small batteryless PV-(SW)RO
24

system which was incorporating energy recovery (Clark Pump) and MPPT. The water cost estimation
was 2/m3. However, this calculation was not confirmed by the lab scaled prototype which was
developed in 2005. Cost estimations by models are usually lower than in real experimental units.
Despite the technological advances in recent years in seawater PV-powered desalination, it is still not
competitive with grid-powered or fossil fuel powered desalination and mainly finds applications in
case of small plants in remote areas with no access to the grid or with expensive fuel supply chains.

2.4. E CONOMICS

AND

M ARKET I SSUES

It is difficult to compare the costs of desalination installations since the actual costs are contingent on
situation-specific parameters. However, in general terms energy and capital costs are the main
driving forces in the desalination total cost (Figure 16).
Maintenance
and parts
Consumables
7%
3%

Labor
4%

Electrical
Energy
47%

Capital
39%

F IGURE 16: I NDICATIVE

SHARES OF TOTAL COSTS IN CONVENTIONAL SEAWATER DESALINATION

[3]

From Figure 16, it can be observed that for conventional seawater RO systems the largest cost
reduction potential lies in capital costs and energy. In case of PVRO seawater desalination, the capital
costs are much greater because of the high investment costs for the photovoltaics, whereas the
operational costs for electrical energy are non-existent. Feasibility in PV-RO systems can be reached
when the gain achieved by elimination of operating electricity costs outweigh the loss due to rise in
capital costs. PV technology is fast developing, and currently among the renewable energies, they
constitute the fastest growing market. The industry has almost doubled from 2007 until 2008 and
similar trends are expected to happen the following years (Figure 17):

F IGURE 17: P HOTOVOLTAICS

INDUSTRY FROM

1999 - 2008

25

Extensive market penetration, led to the rapid reduction of the price of PV cells from 1979 until 2009
(Figure 18), and currently the prices stand at 3-4 $/Wp in the US and Australian markets. Similar
trends are also expected in the future, making photovoltaics more promising technology than how it
was viewed in the past.

F IGURE 18: PV M ODULE

PRICE VERSUS C UMULATIVE MODULE PRO DUCTION FROM

F IGURE 19: C OST

VERSUS

C ONVERSION E FFICIENCY

IN

1979 - 2009

2008

26

From another point of view, thin film solar panels are less expensive than the crystalline ones, in
terms of /m2. However, the choice of the right technology lies strongly on the application and
purpose. PV experts generally agree that crystalline silicon will remain the "premium" technology for
critical applications in remote areas. Thin film will be strong in the "consumer" market where price is
a critical factor [56].
In addition to PV cost reduction, RO has seen also steep declines in prices; between 1990 and 2002
membrane costs have fallen 86% [7]. Steeply declining maintenance cost in combination with
relatively low capital cost, seems to offer hope that PV-powered seawater desalination will become
competitive with other water supply sources, in context of remote regions where freshwater deficits
are covered with transportation of water, or electricity from the grid is not an option.
The dilemma around batteries for PVRO systems is still unresolved. Other energy storage possibilities
like super-capacitors could also be investigated. Alternatively, it may be interesting to use a battery
array exclusively for the purpose of regulating the solar energy fluctuations in the power supplied to
the motor, instead of for extending operating hours.
However, following the present research trend in favour of batteryless systems for autonomous cost
effective solutions, the present project shall focus on implementation of a seawater PV-RO setup in
the real scenario with a view to minimizing costs, and understanding the potential obstacles to
implementation with respect to long term performance and reliability.

27

28

3. S IZING

THE

PV-RO

SYSTEM COMPONENTS

3.1. I NTRODUCTION
For the purposes of this project, an already existing RO-unit installed in the WaterLab, CiTG, TUDelft
will be coupled directly with photovoltaics. The RO-unit (SpectraTM LB 1800) is commercially available
from the American company Spectra Watermakers, Inc. and is designed for land-based applications.
The maximum capacity of the system (with continuous operation) is 7.9 m3/d and it comprises of1:
2 spiral wound membrane elements
The state-of-the-art Spectra Pearson Pump (HPP) which incorporates energy recovery.
A Low Pressure Pump (LPP)
A water storage tank of 1 m3
One microfiltration pre-treatment unit
From the literature search, it was found that post- 2002 research interest was driven towards
batteryless PVRO systems (mostly for brackish water) operating in fluctuating conditions of power
input, thus avoiding the use of the batteries due to the financial and often the environmental costs
involved with their maintenance and replacement. The investigation of batteryless PVRO systems
operating in fluctuating conditions has given positive indications in the past. In addition, previous
research on the current existing system gave encouraging results for the use of the system without
an energy buffer under fluctuating operation [1].
The ultimate effect of not including batteries is the system having to cope with variable operation
patterns of flow rates and pressure that follow instantaneous atmospheric conditions. Membrane and
pump manufacturers object to such operation regimes since RO systems are designed for constant
operation. Taking these facts into consideration, it was decided that the research results of this
project would be valuable and they would contribute to new knowledge in the field of seawater PVRO
for small systems, if the system did not incorporate batteries.

3.2. S IZING

APPROACH AND ASSUMPTI ONS

It is assumed that the system is operating under fluctuating conditions with no energy buffer. Sudden
start-stops can seriously cause problems in the pumps as well reduce the lifetime of the membranes.
For this reason, the challenge in the sizing of the components lies at a system design operating
without sudden start-stops but within the power range that the motor operates safely and gives safe
water.
The design of an autonomous desalination plant as well as the economic analysis is site specific and
cannot usually be generalized for applications in other situations. This sizing approach is site specific
to the city of Chania in Greece, with respect to tilt angles, radiation data, and the demand-driven
strategy of water production.

More information about the system configuration is given in chapter 4.

29

It has to be ensured that:

The given power to the system is within the operating range (670 W 1200W) of the
motor2
The given power to the system will always be above a certain limit in order to produce safe
drinking water (<500 mg/l).
The following assumptions were made and used for the sizing calculations:
1. Radiation data is site-specific for Chania, Greece, and represents normalized values for a
typical day of each month3.
2. It should be noted that radiation data does not include stochastic variations (like passing
clouds, showers, etc.)
3. Optimized panel tilt angle for every month based on the latitude of the location. 2
4. The water produced in a year is the mean of the projected annual water production on the
basis of June (summer) and December (winter).
5. The maximum rated production of the facility (7.9 m 3/d) is taken into account.
6. No use of batteries (or any other energy storage device).
7. The constant efficiency of conversion of PV panels used in calculations is equal to its
maximum value.
8. The excess in power production by the PV panels is used to power the controllers; a power
requirement which is assumed to be adequately satisfied at all times.

3.3. C RITERIA

FOR PRODUCTI ON OF

3.3.1. M IN I M U M

AN D

S AFE W ATER

M AX IM U M P O W ER S

Figure 20 correlates the power of the system (W) - the HPP including the LPP pump with the product
flow (l/min) and the water quality (TDS in mg/l). From the relation between the power consumption
and the water quality in TDS, it can be calculated that the minimum power required for the HPP-LPP
system to produce safe drinking water (<500 mg/l) is 793 W.

This operating range has been proved to be safe for the motors in previous experimental investigations on the same setup
(By Brett Ibbotson).
3
http://re.jrc.ec.europa.eu/pvgis/apps/radday.php?lang=en&map=europe

30

F IGURE 20: P OWER C ONSUMPTION

VERSUS

P RODUCT F LOW R ATE

AND

W ATER Q UALITY [1]

From the system specifications, the capacity of the system is 7.9 m 3/d. Using the relation acquired
from the above figure for water production and power consumption, the maximum power which the
system can utilize can be calculated to be 1204 W.

3.3.2. O P T IM AL I N C L IN AT I O N

AN G L E

The angle of inclination of the solar panels is an important criterion affecting power production. For a
given location, the optimal angle for harnessing solar irradiation depends upon the time of the year.
Solar PV systems may hence be designed with a mechanism to tilt the plane of the panels to match
the optimal angle for a particular time of the year.
The optimal panel inclination angle is affected by several other minor factors, hence the following
location specific data about the location is essential:
Location: 35 30'26" North, 24 1'26" East Elevation: 20 m a.s.l
Nearest city: Khania, Greece (0 km away)
Land cover class: agro-forestry areas (CLC244)
Optimal inclination angle is: 28 degrees
Orientation (azimuth) of modules: -1.0o (optimum)
Annual irradiation deficit due to shadowing (horizontal): 0.0 %

The optimal inclination angles for every month (and a normalized yearly optimal angle) in Chania as
given by JRC are shown in Figure 21.

31

F IGURE 21: M ONTHLY O PTIMAL I NCLINATION A NGLES

FOR

C HANIA , G REECE

3.4. S IZING S TRATEGY


The general procedure that is followed in the course of the sizing is as follows:
An area of the solar panels is assumed
Based on the area of the solar panels and their efficiency, the minimum required irradiance
W/m2 is calculated (where the power provided by the panels is more than the minimum
power required for safe water production=793 W)
The usable time frame within the month (June/December) is evaluated
All irradiation values leading to power greater than the maximum power (>1204W) which can
be consumed by the system are set to a fixed maximum.
The irradiation within the usable time frame is integrated over the usable time frame
Using the integral value in the relevant equation the total water production over a single day
is calculated
Assuming the relevant pay-back periods and costs of the system, the cost of water (per m 3)
is calculated
The procedure is iterated to reach a value at which the cost of the system is as low as
possible while the energy wastage is minimized.

3.4.1. PV P AN EL

S EL E C T IO N AN D S I Z IN G

Many different PV panels were compared based on their capacity, efficiency and cost/Wp. The final
selected model of PV panel is the Energy Power Plus which combines high conversion efficiency and
very low cost4:
T ABLE 4: PV P ANEL C HARACTERISTICS

Photovoltaic
type

Brand

model

capacity
(W)

Area
(m2)

Module
efficiency

Weight
(kg)

cost
()

cost
(/W)

cost
(/m2)

Polycrystalline
silicon

Solar Energy
Power Plus

SE230P20/Ac

230

1.62

14.0%

20

437

1.9

379

Referring to: http://www.solar-systems.gr/SPECIFICATIONS-POLY.htm

32

3.4.2. M E T HO D O LO G Y

FO R T H E C A L CU L AT I O N O F T H E D A I LY W AT ER P RO D U C T IO N :

From Figure 20 the relation between water production rate and power is:

0.004 x

0.6691

...

(1)

Where;
y= Water production rate (l/min)
x = Power (W)
Since the system is under fluctuating operation, the water production is not constant during the day.
Thus, the daily water production Y can be calculated by the integration:
t stop

t stop

y dt
t start

(0.004 x

0.6691) dt

t start

...

(2)

The Power x can be translated to the irradiance with the following relation:
x

I A

(3)

Where: x = Power (W), I = Irradiance (W/m 2), A= area covered by the panels (m2), p = max
conversion efficiency (%), tstart=start up time of the system, tstop=stop time of the system
From (2) and (3), the total water production over a single day (Y in liters) can be given from:
t stop

0.004 A

I dt
t start

0.6691 dt

(4)

The water production was calculated for both a typical day in winter (December) and in summer
(June) using the irradiation data from JRC. The average of winter and summer production values
represents the estimation for the daily water production throughout the year.
In order to account for the minimum power to produce safe water (793 W) and for the maximum
power which the system can utilize (1204 W), the original irradiation data needs to be modified. In
order to do this, the irradiation (W/m2) corresponding to the maximum rate of absorption of power
from the solar panel array is first calculated (using equation 3). Following this, all values of irradiance
greater than this value are equated to the maximum value.
The second threshold is that the reverse osmosis plant can be in operation only when the power
produced by the panels is larger than the minimum power required by the plant to produce safe
water (793 W). The first instance of the day when this condition can be satisfied becomes the t start
value, and the last possible time instance satisfying the condition becomes the t stop. The integral is
evaluated between the limits of tstart and tstop.

33

500
450
400

Irradiation (W/m 2)

350
300
250
200
150
100
50
0
6.00

8.00

10.00

12.00

14.00

16.00

18.00

Time of the day (hr)


Solar Irradiation (W/m2)

F IGURE 22: I RRADIATION

Absorbed Power

C HANIA IN D ECEMBER
20 PV P ANELS

DURING A TYPICAL DAY IN

E XPLOITATION

USING

AND ITS POSSIBLE

Taking as an example the use of 20 panels, Figure 22 was created. The area between the two lines in
Figure 22 (plotted for December) indicates the energy which needs to be disposed. A part of this
energy could possibly be used in powering the computer and other utilities. The energy which needs
to be disposed off in summer is even greater than what it will be in winter. The sizing is made on the
basis that the system provides adequate quantity of water at a low cost. If fewer panels are used in
the system, the quantity of water produced in winter decreases greatly. The loss of revenue on this
account may be greater than the savings due to a lesser PV Panel requirement (depending on the
cost of water in the winter months).
Since we have assumed that the system works at its maximum efficiency at all irradiation values, the
energy wastage is a little lesser in the real system. If the energy wasted in the summer was desired
to be reduced even further while maintaining the same cost of the system, there would be no
alternative but to use a battery.

3.5. R ESULTS
Multiple iterations (trial and error) of different PV area values were conducted as described above.
The results of the daily water production and the final water cost within the possible time frame for
each area value were compared. It was seen that between 10 - 20 panels there is a balance between
energy wastage and cost of water. For the purposes of the experiments, it was decided to use 10
panels (area of 16.20 m2) taking into account the practical hindrances of temporary installation.
T ABLE 5: R ESULTS

Area covered by the panels (m2)


Number of panels
Operation time in winter
Daily water production (m3/d) in winter
Operation time in summer
Daily water production (m3/d) in summer
Average annual water production (m3)

OF

S IZING A PPROACH

16.20
10
7.25 hours
1.16
11.25 hours
3.12
781
34

3.6. T HEORETICAL C OST A NALYSIS


The costs for the RO system, PV panels, PV installation and mounting structures as well as tanks and
the rest of the equipment is based on real market prices. The RO unit from Spectra Watermakers,
Inc. comes mainly pre-assembled so the assembly cost is minimal. For the purposes of this cost
analysis, apart from the LCB (Linear Current Booster), an PLC controller (to monitor and control the
system operational parameters and water quality in the real scenario) has been accounted for in the
automation costs. The different components of the system can be assumed to have different
depreciation periods. It is assumed that the value of the component at the end of the depreciation
period is zero.
T ABLE 6: A NNUAL C OST A NALYSIS

Particulars
Capital Costs
RO system+pretreatment+HPP
PV panels
PV installation and mounting structures
LCB Controller
Feed Pump
Tanks and other equipment
Automation Costs

(Total capital cost)


Annual Operational Costs
Membrane + Pre-filter Replacement5
Post-treatment6
Miscellaneous Operational Costs7
Maintenance cost8
Feedpump operational costs9

Cost

13000
4347
1710
335
159
300
1200

Depreciation
Period (years)
10
25
25
10
10
10
10

( 21051)

900
19
260
335
26.13

An important contributing factor to the cost of water production is the interest payment on the capital
investment. This is, however, a rather complex factor to account for in the cost calculations since it is
largely dependent on the economic and political scenario where the installation is to be placed. Due
to this variance, the total annual costs of water production (and hence the cost of the water) is
presented for different cases of real (inflation adjusted) interest rates.

Required every 6 months


Use of calcium hypochlorite tablet. Cost based on a chlorination cost of 2.4 cents per m3, which is in turn based on a calcium
hypochlorite cost of 5 USD per kg, and meeting the required dilution of 5 mg of chlorine equivalents per liter of water
7
Considered equal to 2% the cost of the RO unit
8
Considering the maintenance cost of 0.5% for all civil structures, 2% for Mechanical Elements and 4% for Electrical
Components as recommended by de Moel [57]
9
Considering the feed flow of the system equal to 1.64 m3/h, the boost pressure required to feed the water to the HPP equal
to 1 bar and a pump efficiency 70%,the hydraulic pump power is 0.657 kWh/over 10 hours of operation. As an example, the
cost for the kWh in Greece is 0.11/kWh, which means the feedpump operation costs 0.07 per day.
6

35

Annual Cost of Water Production


(in )

4400
4300
4200
4100
4000
3900
3800
3700
3600
3500
3400
3300
3200
3100
3000
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25
Years

0%

3%

5%

0% Averaged

F IGURE 23: V ARIATION

OF

3% Averaged

5% Averaged

A NNUAL W ATER P RODUCTION C OSTS

Based on Figure 23, the cost of water based on an annual production of 772 m 3 (based on the power
output of 10 PV modules10) is shown as a function of the interest rates in Figure 24, where we can
see a linear rise of the cost of water with the interest rates.

Cost of Water (/m3)

6
y = 15.909x + 4.2634
5.5
5
4.5
4
0%

2%

4%

6%

8%

10%

Interest Rate
F IGURE 24: V ARIATION

OF

W ATER C OST

WITH

I NTEREST R ATES

For a sample case, we consider an interest rate of 3% as advised in [57], with our desired
configuration to understand the various contributions to the total cost of a PV-RO system. As we can
see from Figure 25, the PV panels account for a rather small percentage (6.7%) of the total system

10

The reason for choosing 10 solar PV modules with a total area of 16.2 m 2 has to do with the maximum space availability on
the roof of the testing facility

36

costs in the given setup. At first glance, this is an indicator that the solar PV panels do not constitute
the dominant (or prohibitive) fraction of the total cost of a PV-RO system. It must be emphasized that
the PV-RO system could not be optimized due to practical restrictions of roofing area at the test site.

Interest
Amount,
10.1%

RO
system,
35.9%

Other
Operationa
l, 9.0%

Other
Maintenan
ce, 9.3%
PV panels,
6.7%

Membrane
Replaceme
nt, 24.9%

Automatio
n, 4.2%

F IGURE 25: C ONTRIBUTIONS

TO THE

T OTAL C OST

OF A

PV-RO S YSTEM

With increasing number of solar panels, it is possible to extend the operational period of the
batteryless PV-RO system. It is likely that our constrained limit of 10 panels is not optimized with
respect to the production capacity of the RO unit. In the following study, we try understand the
variation of the average water cost with increasing number of solar panels. It is assumed that the
real interest rate is 3%, and that the cost of the solar panels and fixtures scale linearly with the
installed capacity.
In Table 7, different number of panels and resulting values of water production and water cost are
shown for comparison:
T ABLE 7: V ARIATION

IN

W ATER C OST

WITH

N UMBER

OF

PV P ANELS

Number
of panels

Area
(m2)

Operation
time in
winter (hr)

Operation
time in
summer (hr)

Water
production in
winter (m3/d)

Water
production in
summer (m3/d)

10
15
20
25

16.2
24.3
32.4
40.5

4.25
6.75
7.25
7.75

9.25
10.75
11.25
11.75

1.16
2.10
2.40
2.59

3.12
3.54
3.74
3.91

Average
Water cost
(/m3)
4.74
4.17
3.61
3.55

We can see that as the number of panels increase, the average water costs decreases. Also seen is
that the reduction in water cost follows the law of diminishing returns. An innate assumption in the
above analysis is that the efficiency of the solar cell is constant at all values of irradiation. This in fact
is not true, and practically, there may be a much smaller gain in operational periods due to increasing
the number of panels than what this analysis bares out. Also, there are difficulties in obtaining larger
uni-directional roof surfaces for PV Panel installation, and beyond a point, it may be required to pay
rent for such an area, something that has not been considered in the present cost evaluation.

37

38

4. EXPERIMENTATION

The experimental part of the project was conducted in Greece in the campus of the Technical
University of Crete. The field study was carried out between the 3 September and 3 November 2010.

4.1. S ELECTION

OF THE AREA OF STUDY AND PROJECT PARTNERS

The location in Chania, Crete was selected for various reasons:


Firstly, Crete is one of the southernmost locations in Europe. The mean solar irradiation
values increase as we get closer to the equator which makes it the ideal European location to
experiment with solar PV systems. Furthermore, as it can be seen from graph 1, the
temperate Mediterranean climate in Chania ensures that temperatures do not rise to a point
where they seriously start affecting the performance of the solar cells.
Secondly, the presence of the Technical University of Crete in Chania facilitates the
availability of technological expertise, along with the technical support functions. In
particular, Prof. E. Diamantopoulos who heads the Laboratory of Environmental Engineering
and Management has been actively researching novel environmental engineering techniques
and the project Drinking with the sun fits well in the context of the research profile of the
laboratory.
In the framework of this collaboration, the Drinking with the Sun group was provided with the
following from the abovementioned laboratory:
A room for the facility, the pumps, electronic equipment and the storage tanks,
Suitable open roof for the installation of the solar panels
Electricity for the laptops, datalogger and feed pump
Assistance, guidance, and financial support for spare parts and pipes during the duration of
the experiments
Available workspace, internet and stationary
Lab facilities for the production of demineralised water and TDS analyses.
For experimental purposes, a reverse osmosis unit (SpectraTM LB 1800) originally placed in the
WaterLab, CiTG, TUDelft was transported and installed at the place of interest.

39

F IGURE 26: L OCATION

4.2. L OCAT ION

T ECHNICAL U NIVERSITY

OF

C RETE , C HANIA

CHARACT ERIST ICS AND METE OROLOGIC AL DATA

F IGURE 27: M ETEOROLOGICAL D ATA


AND

OF THE

W ATER T EMPERATURE

FOR

C HANIA : A NNUAL T EMPERATURE , P RECIPITATION , W IND S PEED

11

12

The altitude of the location is 137m above the sea level and with coordinates: Latitude: 35 32' 00"
N and Longitude: 24 04' 09. The location of the experiments is characterized with high insolation in
the summer period. In 2010, Chania had 2902 total accumulated sunshine hours, reaching peak in
July: 12.3 hrs daily sunshine and 383 sunshine hours in total were recorded 13. The average value in
2010 for the daily sunshine hours in Chania was 8.0 hours.

11

The data are acquired from the meteorological station of the Laboratory of Atmospheric Aerosols at the Technical University
of Crete.
12
http://www.zoover.co.uk/greece/crete/chania/weather
13
http://www.weatheronline.co.uk/weather/maps/city

40

For a notion of comparison, in Germany (Essen) 1402 sunshine hours were recorded in total in 2010,
while in Indonesia (Bali) 2405 sunshine hours were recorded in total. Furthermore, the average value
in 2010 for daily sunshine hours was 4.3 hours and 7 hours respectively for 2010.

F IGURE 28: T OTAL

SOLAR ENERGY

(M ONTHLY V ALUES )

OVER

2010

IN

C HANIA [F OOTNOTE 11]

A general observation from the Figure 27 and Figure 28 is that during the summer months (June until
September) the precipitation is negligible leading to water scarcity. However, the solar radiation
values reach its peak these months, fact which reinforces the synergy between photovoltaics and
reverse osmosis desalination systems.

4.3. O BJECTIVE S
The main objective of the experiments was the testing of the batteryless PVRO prototype under
realistic conditions:
- At a place with high irradiance potential during fluctuating solar conditions
- Using real seawater with the highest salinity (39000 psu) in the World Ocean 14.
Specifically, the experimentation results will assist in drawing conclusions about:
Test the integration of the PV panels + LCB controller + RO in one system. Does the system
achieve desalination of water in acceptable quality limits? What is the impact of fluctuating
operation in water quality?
Test the performance of the system with respect to stability and reliability issues. Which is
the response of the system in the absence of an energy buffer?
Test the efficiency of the system. Is it more efficient than similar existing products as
required? Comparing the Specific Energy Consumption of the system with other similar
systems will give an impression of the efficiency of the system.
LCB usability. Does the LCB affect the Specific Energy consumption and does it make a
difference in operational times and water quality in cloudy days?

14

Interconnected system of the Earth's oceanic (or marine) waters; comprises the bulk of the hydrosphere, covering almost
71% of the Earth's surface. http://upload.wikimedia.org/wikipedia/commons/2/28/EU-Glob_opta_presentation.png

41

What is the yield of water during experimentation? How can it be compared to the theoretical
values estimated in the sizing of the system?
Test response of membranes on power fluctuations. Is the lifecycle affected by intermittent
operation?

4.4. E XPERIMENTAL S ET UP
In accordance with the component sizing as described in the previous chapter, the batteryless
experimental setup at Chania comprised of the following components:
The reverse osmosis unit with maximum capacity 7.9 m 3/d including a high pressure and a
low pressure pump
10 Photovoltaic cells (2.3 kW total installed capacity)
A Linear Current Booster Controller (LCB)
A storage tank for the feed (1 m3) and a storage tank for the permeate (200 L).

4.4.1. THE REVERSE OSMOSIS UNIT


The reverse osmosis unit (SpectraTM LB 1800) which was used for the purposes of the project is
commercially available from the company Spectrawatermakers and designed for land-based
applications. The test rig used in these experiments was built in late 2009 by the MSc. student Brett
Ibbotson at the WaterLab in CiTG, TUDelft.
The facility comprises of:
One microfiltration pre-treatment unit (5 micron nominal filter cartridge, 4 x 20)
2 thin-film composite spiral-wound membrane elements for seawater connected in series,
(Filmtec SW30-4040) with salt rejection 99.4 %
The state-of-the-art Spectra Pearson Pump (HPP) which incorporates energy recovery.
(Leeson electric Model C4D18FK6, 2HP, 1800RPM, DC motor 24V, safe operating range: 450
W- 1200 W, rated current: 50 A)
The Spectra LB 1800 system used for the experimentation has a fixed water recovery equal to 20%,
which means that 20% of the feed flow is produced as permeate.

F IGURE 29:T HE E XPERIMENTAL S ETUP


TU D ELFT [1]

IN THE

W ATER L AB , C I TG,

The use of a feed pump


The Pearson pump required the inlet flow to be supplied between a minimum of 0.7 bar and a
maximum of 1.4 bar. The pressurization of the inlet stream can be achieved by passive means such
42

as storing feed water in a tank located more than 12 m above the Pearson pump. However for the
purpose of experimenting in the given scenario (with limitations of space and infrastructure), active
means were used, involving a feed pump 15 along with a pressure relief valve configured to open at
1.2 bar pressure. For the experimentation study, the system of the feed pump was considered
independent of the main system and was not powered by the PV source. This may be justified to an
extent, since the power consumption of the high pressure Pearson pump is much greater than that of
the feed pump.
System configuration
Since it was not practical to continuously consume sea water (due to logistics constraints of
transporting sea water to the site of experimentation), the feed water is recirculated. The
configuration of the experimental setup is such that all the water streams together form a closed
loop. The feed pump delivers water at around 1 bar to the high pressure pump via the 5 MF cartridge
filter. The water after passing through the RO membranes is separated into a concentrate and a
permeate stream. The concentrate stream after passing through the Pearson pump for energy
recovery is discharged back into the feed water reservoir. The permeate stream is also discharged
into the feed water reservoir. Since both the streams are directed back into the feed water reservoir,
the net salinity of the feed water remains unchanged. This operation is said to be the RUN mode as
shown on the control panel of the LB1800.
Flushing operation
After a long cycle of normal operation or after an inactivity period of more than 12 hours, the
membranes need to be flushed with fresh water in order to be rinsed from the accumulated salts.
Although the LB 1800 is provided with a FLUSH mode where water can enter the system via a
separate line, the system was mostly flushed using the normal RUN mode (with manually altered
feed streams) in order to allow the fresh water to also pass through the prefilter.
In order to maintain the concentration of the artificial seawater at a constant level during flushing
with fresh water, the initial concentrated discharge is directed into a bucket to capture the salts being
flushed out. Once the conductivity of this discharged water approaches the conductivity of the feed
water, the remainder is directed to the gutter with the captured water being returned to the
reservoir. Similarly once the flush had been completed and the system is put back into a regular
cycle, the initially discharged water was discarded. Further adjustments to the level and conductivity
of the reservoir water were made as necessary to ensure constant characteristics.

4.4.2. PHOTOVOLTAICS ARRAY


Based on the sizing calculations and after a thorough market research was decided to use 10 Energy
Power Plus panels with capacity 230 Wp each (rated at 30 V and 7.8 A), giving a total peak capacity
of 2.3 kW. Each panel covers an area of 1.62 m2, thus an area of 16.2 m2 in total was used.
The roof of the building where the collaborating laboratory was located was selected for the
installation of the solar panels. The location of the roof was ideal due to the following reasons:
-

15

The building is located on a hill and there is no taller building in the vicinity which can cast a
shadow over any part of the panels during the peak hours of energy production
The access to the roof is restricted by a gate which ensures the security of the array from
theft/damage.

The feed pump used was a Low Pressure Pump (Liquiflo 37R positive displacement magnetic-drive pump).

43

The roof was exactly above the Laboratory and above the room where the desalination
system was installed, leading to short cable requirements and thus less power losses.
The orientation of the roof was almost south, which is ideal for solar panels located in the
Northern hemisphere.

T ABLE 8: R OOF C HARACTERISTICS

Roof characteristics
Length (m)
Inclination from
the horizontal (deg)
Orientation

F IGURE 30: R OOF

ON WHICH THE

PV P ANELS

WERE

18
30
SSW (30 from South)

I NSTALLED

The roof inclination is coincidentally close to the optimal panel inclination angle proposed by Joint
Research Centre (JRC) of the European Commission for the area of Chania (28 for the whole year)
hence no additional means were required to alter the angle of the panels. Furthermore, despite the
fact that tracking mechanisms may lead to increase in efficiency of the system, it was decided that it
would be unjustified with respect to the cost, since the optimal inclination angle during the period of
testing (September-October) is in any case close to the existing angle of the roof (33 for September
and 46 for October) based on JRC data.
Special mounting structures were used for the panels in order to avoid drilling on the roof. The panels
were connected in parallel in order to keep the output Voltage at 30 V (rated) which was required by
the Pearson pump motor (rated at 24 V). Consequently, the maximum current of the solar array was
78 A. Further specifications for the PV panels can be found in the appendix.

T ABLE 9: PV P ANEL C HARACTERISTICS

PV characteristics (per panel)


Capacity (Wp)
Rated Voltage (V)
Rated Current (A)
Area (m2)
Size (mm)
F IGURE 31: S OLAR PV P ANELS I NSTALLED

ON THE

230
29.5
7.8
1.62
1636 (L) x 992 (W) x 50 (H)

R OOF

4.4.3. LINEAR CURRENT BOOSTER (LCB) AND SYSTEM INTEGRATION


From the literature study, it was seen that a DC-DC converter was necessary between the
photovoltaics and the pump to optimize the integration of PVs with the pump motor. Since MPPT
circuits exist mostly as charge controllers, it was decided to integrate a Linear Current Booster (LCB)
in the PVRO system. As it was described in earlier chapters, the LCB is a converter which exchanges
voltage for current preventing the motor from stalling in times of low irradiation. The system
requirements were submitted to the Canadian company Solar Converters Inc and the controller was
custom fabricated for the purposes of the project (Figure 32).
44

The LCB controller was supplied with incorporated various other user controls, LED displays and
protections. Most importantly, the device current limits its output to protect the unit and the load
from jammed motor/ low resistance operation. After the manufacturers suggestion, fuses after the
LCB and before the motor were installed. Further specifications of the LCB can be found in the
appendix. For added security, fuses after the photovoltaics load and before the LCB were
incorporated. As an addition at a later stage, a switch was installed depending on the position of
which either the PV source or the AC/DC rectifier could be used to power the motor. This was very
useful in order to perform tests with constant power for reference and for troubleshooting needs.

F IGURE 32: L INEAR C URRENT B OOSTER (LCB)

The final electrical connections can be seen in Figure 33:

F IGURE 33: E LECTRICAL C ONNECTIONS

OF THE

M AIN C OMPONENTS

OF THE

S YSTEM

4.4.4. INSTRUMENTATION
For the requirements of the experimental part of the project, several measuring devices were used in
order to monitor and control important parameters of the process and acquire useful data for the
analysis of the experiments.
The following measuring instruments were used:
45

Pyranometer for the measurement of the instantaneous irradiation


Two electrical conductivity meters for the feed flow and the permeate flow (indirectly
measuring salinity)
Voltmeter and ammeter for measurement of voltage and current of the motor
Membrane pressure gauge
A flow meter gauge for the product flow
Two analogue pressure gauges for measuring the pressure of the feed at the inlet to the prefilter and at the exit of the membranes in bar
Analogue flow meters for the feed flow and the permeate flow
The analogue meters and gauges located on the Spectra control box were used to monitor the
changes in operating parameters during manual control of the Pearson motor speed. The
pyranometer used gives a linear voltage production until 2000 W/m2 and it has sensitivity 69.3
V/W/m2. Temperature of the feed and permeate water flows could be observed on the display of the
two electrical conductivity meters. Since conductometric measurements are temperature dependent,
the meters use the function of temperature compensation to convert all measured values to a
reference temperature (25 C).

F IGURE 34: I NSTRUMENTA TION SYSTEMS ( FROM L EFT TO R IGHT ): A NALOGUE P RESSURE G AUGES /
FL OW M ETERS , D IGITAL C ONDUCTIVITY METERS , PYRANOMETER

Most of the data variables from the measuring instruments were stored independently using a 16channel digital datalogger. The parameters could be monitored in real time from the datalogger
interface installed on a laptop, while in the same time were recorded. The calibrations of the
analogue data inputs are given in the appendix. Table 10 shows a list of parameters logged by the
instruments.
T ABLE 10: P ARAMETERS

LOGGED BY THE

Parameter

I NSTRUMENTS

Type of meter
2

Instantaneous irradiation at the flat surface (W/m )


Motor Voltage (V)
Motor Current (I)
Membrane pressure (bar)
Permeate flow (l/min)
Electrical Conductivity of feed flow mS/cm
Electrical Conductivity of permeate flow S/cm

Pyranometer
Voltmeter
Ammeter
Pressure meter
Flow meter
Conductivity meter
Conductivity meter

46

F IGURE 35: RO

UNIT CABINET ON FIELD AND

D ESK S TATION

ON FIELD

4.5. S YSTEM S TART U P


The reverse osmosis unit was transported from Netherlands to Chania, Greece together with the feed
pump, tubes and spare filters and membranes, packaged in a specially constructed wooden case. The
initial preparation activities involved setting up the location for the system and the work station,
obtain all necessary tanks and equipment, piping and modifications in valves and fittings, and
transferring water from the sea. Furthermore, the installations of a datalogger, electronic devices and
connections as well as calibrations were also done in preparation for the experimentation.
Since one of the objectives of the exercise was to approach reality as much as possible, it was
decided to use real sea water for the experimental purposes. Sea water was transferred from the
nearest organized beach (Agios Onoufrios), which had excellent water quality (zero tars, mineral oils,
seaweed or microbiological activity. The water was pumped from the sea with a submersible pump
provided by the laboratory group. The seawater was filtered with a 50 MF net to ensure that all
suspended solids are removed. A tank of 1 m3 was procured to store the feedwater at the location of
the RO system. A block diagram of the seawater PV-RO system installed at Chania is shown in Figure
36:

F IGURE 36: S IMPLIFIED P ROCESS D IAGRAM

OF

PV-RO S YSTEM

47

4.5.1. D ET E R M IN AT I O N

OF

TDS

E LE C TR I C A L

AN D

C O N D U CT I V I TY RE L A TI O N

TDS (mg/l)

The TDS content and the electrical conductivity of a water sample are closely related. Conductivity
increases with increasing ion concentration, which means that in most cases it gives a good
approximation of the TDS measurement. As is found in literature 16, there is no one single constant
linking the values of conductivity (in S/cm) and TDS (in ppm). To determine the precise relation
between TDS and electrical conductivity for the sea water used for the experiments, TDS analysis was
carried out in the laboratory. This involved evaporating water from measured filtered solutions of
known dilutions and weighing the mass of the remaining solids. Figure 37 was plotted based on the
results obtained from the TDS analysis.
21000
20000
19000
18000
17000
16000
15000
14000
13000
12000
11000
10000
9000
8000
7000
6000
5000
4000
3000
2000
1000
0

y = 718.23x
R2 = 0.9958
0

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29

Conductivity (mS/cm)

F IGURE 37: R ELATION

BETWEEN

TDS

AND

E LECTRICAL C ONDUCTIVITY

OF THE FEEDWATER USE D

The electrical conductivity value for the seawater used was measured to be 53.4 mS/cm (at 26 C).
Based on the correlation factor obtained from the TDS analysis, it was calculated that the electrical
conductivity of the seawater used in the experiments (53.4 mS/cm) corresponds to TDS= 38353
ppm.

4.6. S YSTEM O PERATION


The motors of the two pumps (low and high pressure pumps) were fully controlled via a computer
when the system was placed in the WaterLab, in Delft. However, since the addition of an automated
controller increased the complexity of the system, it was decided to avoid any motor speed controllers
during the experimentation period in Crete. Since the Pearson pump is powered directly by
photovoltaic cells, the power available for the Pearson pump is varying, depending on the fluctuation
in the incident solar radiation. To prevent the pump from stalling in times of low power availability,
the pump speed needs to be controlled manually (using a knob on the control box). To exploit the
available solar energy to the maximum possible extent (thus maximizing production of fresh water), it
is necessary that the motor speed precisely follows the variations in solar irradiation.

16

http://www.appslabs.com.au/salinity.htm

48

The speed of the feed pump motor was set constant at maximum value while the pressure relief
valve was used to prevent overpressurising the stream line. The system efficiency is not affected by
this apparent loss of feed water, since the feed pump is not powered by the photovoltaics.
The feed water was recirculated during the experiments the same way as in the WaterLab.
For the month of September, starting time operation was around 8:45 and ending time 18:00 for a
clear day while for the month of October, starting time operation 9:15 and ending time 17:30
respectively.
The activities to be done every day for a full cycle of operation can be summarized as follows:
Setting up the laptop and configuring a new logging session. For most of the experiments the
time step was set at 30 s.
Adjusting the configuration of the system and valves for the RUN mode
Starting the PV system and the pumps
The manual intervention required during experimentation involved altering the speed of the
Pearson Pump depending on the solar fluctuations, and noting the values of certain system
parameters which were not being recorded by the data logger.
The system was flushed every evening after the sun had set with the 150 L permeate water
stored in a permeate barrel.

4.7. P ROBLEMS

FACE D

TROUBLESHOOTING

After the first set up of the system, numerous problems or challenges were encountered which led to
various modifications to solve these problems. The main problems that required troubleshooting are
as follows:
Lack of unchlorinated tap water to cover flushing requirements of the system
One of the most important aspects in a reverse osmosis process is the fresh water flushing to reduce
biofouling. However, flushing with freshwater proved to be a great problem on site since the
municipal tap water in Chania is chlorinated for disinfection purposes. The sea water membranes
used (Filmtec SW30 40-40) have zero free chlorine tolerance and for this reason premature
membrane damage due to oxidation could occur in case of directly flushing the membranes with tap
water.
As a first solution, demineralised water was produced at the Laboratory facilities in order to flash the
system. However, soon it was noticed that this was neither a practical nor an economical solution,
since more than 200L of demi water was required to flash thoroughly the system.
As a final solution, permeate water was decided to be used for flushing to overcome this problem.
The following tank configuration was adopted to ensure that at the end of the day adequate quantity
of product water would be available for flushing:
A second tank (200 L) was placed above the 1 m 3 tank, in order to store permeate water. The
intention was that during normal running of the system, the permeate flow would discharge in the
200 L barrel while the concentrate would return back to the 1 m 3 tank. The 200L barrel has been
modified in such a way, that after filling in of 150 L, the permeate stream was overflowing back to
the 1 m3 tank.

49

F IGURE 38: T ANK C ONFIGURATION

FOR

P ERMEATE C OLLECTION

Increase in feedwater concentration because of new tank configuration


During the first running of the system in the two-tank configuration, it was observed that the
salinity in the 1 m3 tank was increasing, since the permeate flow was discharged in the barrel and not
returning to the feedwater tank until the barrel started overflowing. At the point when the barrel
starts overflowing, steady state is reached, with a certain feed water concentration. Starting with an
initial salinity of the sea-water in Chania (38353 mg/l), it is found that at steady state, the salinity in
the feedwater tank had risen up to 43000 mg/l TDS. Hence, the salinity of the feed water which
would have been used in experimentation would have not been the same as the real salinity of the
seawater source at steady state.
To solve this problem, a part of the sea water contained in the 1 m 3 tank was diluted with
demineralised water in order to achieve the desired feed water concentration at steady state
conditions (equal to the original feedwater source around 38353 ppm TDS). The required starting
feedwater concentration and the volume of demineralised water required to replace part of the
original seawater were found by solving simple salt mass balances:

C0

V1 C1

V2 C2
V0

32675 ppm

Vreplace

V0 1

C0
C1

148 L

Where: Vo = Total volume of the feedwater tank (1 m 3)


Co = Starting feedwater concentration
V1 = Volume of the feedwater at steady state = (Vo-V2)
C1 = Desired feedwater concentration (at steady state) = 38535 ppm
V2 = Volume of the permeate water in the barrel (150 L)
C2 = Concentration of the permeate water in the barrel (500 ppm)

Increased water Temperature and overheating of the pump


Overheating of the Pearson Pump motor and of the feedwater in the tank was a main problem during
the experimentation period, especially in September when the ambient temperature was very high
(>30 C). The problem was aggravated due to lack of ventilation within the room. A small table fan
and later a mobile air-conditioner unit were used in order to supplement the cooling effect of the
motor fan.
50

Temperature (C)

In addition to the overheating of the motor of the pump, the water temperature was rising on an
average from 27 C to 32 C during the course of the day. Various simple solutions were used to
tackle the problem, like insertion of icepacks in the tank and use of ice cubes around the water tubes.
However, these simple solutions did not manage to solve the problem effectively. Yet, temperatures
at the end of September decreased significantly and remained low for the rest of the experimentation
period and solutions like a heat exchanger in the tank or use of dry ice were avoided (Figure 39).
32.4
32
32
31.9
31.6
15/09/2010
31.2
19/10/2010
30.8
30.4
30
29.8
29.6
29.2
29.2
29.1
28.9
28.8
28.8
28.6
28.4
28.2
28.2
28
27.6
27.6
27.5
27.4
27.2
27.1
26.8
9:00 9:43 10:26 11:09 11:52 12:36 13:19 14:02 14:45 15:28 16:12 16:55 17:38 18:21

F IGURE 39: C HANGE

IN

F EED W ATER T EMPERATURE

DURING

31.8

D AILY O PERATION

Damage of parts of instrumentation equipment and unavailability of spare parts


Very frequently accidents with the electronics and instrumentation equipment were occurring. One of
the most difficult to solve was the failure of the motor voltage measuring sensor which was basically
a voltage divider comprising of 2 resistors of very low tolerance. Availability of such spare parts was
very limited in Chania, and thus this influenced the data acquisition for about 3 days until the voltage
sensor was restored.
Figure 40 shows the final process and energy flows along with all the modifications made to the
system.

51

F IGURE 40: F INAL P ROCESS F LOW D IAGRAM

AND INSTRUMENTATION

52

5.

RESULTS AND DISCUSSION

5.1. A

SAMPLE DAY

An example of an experiment which combines clear and cloudy conditions is shown in Figure 41.
900

Irradiance (W/m2)

800
700
600
500
400
300
200
100
0
8:00 8:28 8:57 9:26 9:55 10:24 10:52 11:21 11:50 12:19 12:48 13:16 13:45 14:14 14:43 15:12 15:40 16:09 16:38 17:07 17:36 18:04 18:33 19:02

F IGURE 41: S OLAR I RRADIATION

ON A

S AMPLE D AY (5/10/2010)

Operation period for 05/10/2010 was a little bit less than 9 hours. Starting time was around 9:00 and
stopping time was around 18:00. An abrupt fall in irradiance occurred between 11:00-12:00. Similar
trends in operation time were observed during the other experimentation days.
The daily water production and the total energy consumption are calculated by integrating the
instantaneous flow rate and power consumption by the Pearson pump motor.
Total water production per day (m3):
The total water production (Y) can be calculated by integrating the rate of water produced (per
minute) (y) over the operation time period. Since the time step is constant, the total water production
can be calculated by the addition of each instant permeate flow (Qp) and by multiplying with the time
step (30sec).
t stop

18:15:10

y dt
t start

yi
9:00:00

dt sum Qpi

l
min
30 sec
min
60 sec

1.94 m3

Energy consumption (kWh):


Similarly, by integration of the power consumed, the total energy produced over the operation time
period can be calculated.
t stop

18:15:10

Power dt
t start

Poweri

dt sum Poweri W 30 sec

9:00:00

hr
60 60 sec

5381.44 Wh = 6.24 kWh

Specific energy consumption (kWh/m3):


E / Y = 3.22 kWh/m3
53

It is important to note that the instantaneous power delivered to the system is not only dependent on
the instantaneous irradiance value but also on the effect of light diffusion. For example, the power
produced in the morning with direct incident sunlight, with a specific irradiation value gives more
power than what is produced in the afternoon with diffused light having the same irradiance value.
This can be attributed to the characteristics of the photovoltaics.
In table 11 the accumulative results obtained for the above measurement are contained:
T ABLE 11: PV-RO R ESULTS
Date:
5/10/2010

LCB CONNECTED

FOR THE

S AMPLE D AY (5/10/2010)

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(lpm)

Flux
(l/h*m2)

ECp
(S/cm)

max

841.11

30.30

43

49.09

4.98

20.17

1761.20

min

55.05

15.61

1034.7
2
9.10

TDS
permeate
(mg/l)
1264.95

0.33

0.11

0.43

3.66

2.63

25.17

28

673.79

41.46

3.49

14.14

963.10

691.73

average

473.92

Daily water
Production
(m3)

1.94

Total Energy
consumption
(kWh)
SEC
(kWh/m3)

6.24

3.22

The feedwater temperature on 05/10/2010 ranged between 27 C- 29.8 C. The variations of the
above mentioned parameters during the day are shown in Figure 42.

54

180
176
172
168
164
160
156
152
148
144
140
136
132
128
124
120
116
112
108
104
100
96
92
88
84
80
76
72
68
64
60
56
52
48
44
40
36
32
28
24
20
16
12
8
4
0

EC permeate/10 (S/cm)
Power consumption/10 (W)
Pressure (bar)
Current (A)
Permeate Flux (l/h*m2)
Permeate flow (l/min)

F IGURE 42: V ARIATION

OF

O PERATION P ARAMETERS

IN

T IME

ON

S AMPLE D AY : 5/10/2010

Based on the above results and trends, the following observatins regarding general behaviour of the
system can be made:
In the beginning of the system operation, the permeate quality falls abruptly (conductivity
and TDS rises), due to the concentration of salts on the membrane surface. After some time,
conductivity falls rapidly due to the removal of the accumulated salts from the membranes.
With the increase in power (blue line), the cross-flow velocity rises and the permeate flow
(green line), flux (brown line) and pressure (pink line) follow the same increasing trend. The
reason why the permeate flow perfectly follows the motor power is due the fact that the
system has a fixed recovery ratio. The same can be said about the pressure (pink line), since
the reason for pressure rise in the membranes is also the fixed recovery ratio of the system.
The same does not apply however, for the electrical conductivity (purple line), which is also
affected by other factors such as cross-flow velocity, etc. Lower power leads to lower crossflow velocity and permeate flow leading to low dilution of the diffused salts on the permeate
side and potential increase in the concentration polarization.
In general, we can see that a sustained stable operation of the system, leads to a lower value
of electrical conductivity (which translates into better quality of the water output).
Rapid fluctuations in the solar irradiation do not directly affect the electrical conductivity as
they affect the flow and pressure. In contrast, electrical conductivity seems to be influenced
by longer trends in the solar irradiation. Sustained and sharp fluctuations which lead to
stopping of the system for long time (more than 5 min) significantly affects the quality of the

55

water produced. After stopping and starting the system again, salts which are accumulated
on the membrane surface get diffused on the permeate side deteriorating the quality.
As it is seen from the solar irradiation profile, there was a significant cloudiness between 11
am and 12 pm leading to a large drop in solar PV power production. The abrupt drop in
radiation was to such a large extent (to <200 W/m 2) that the system was stopped since it
could not continue running at the low torque. The conductivity and thus the water quality are
dramatically affected by the start-stop of the system. This can be attributed to the
accumulation of salts on the membrane surface due to the sudden stopping of the system in
response to the insufficient power.
At the following graph, the inverse relation between water quality and power (as well as permeate
flow) can be observed more closely:

F IGURE 43: R ELATION

BETWEEN WATER QUALITY , POWER CONSUMPTION AN D FLOW ON

5/10/2010

When the motor runs in higher speed (higher power consumption), the flow through the pump is
increased and thus the permeate flow also increases, while the water quality improves significantly.
This can be attributed to the greater cross-flow velocity resulting from the increased feed flow.

5.2. G ENERAL C OMME NTS

ON

E XPERIMENTATION

From 14/09/2010 to 03/11/2010 several experiments were carried out on location in Chania, both
with and without the LCB connected. The experiments done during the first days were mostly to
observe how the system operates and do the necessary calibrations and modifications for optimizing
the process. For this reason, the experimental results during 14/09-16/09/2010 are not very
representative.

56

Furthermore, the experiments which were conducted during October represent the worst possible
scenario since the solar radiation pattern best resembles the winter season, with lots of passing
clouds. Two examples of the worst solar conditions in the month of October are shown in Figure 44,
during which operation of the pump was impossible.

F IGURE 44: W ORST -C ASE S OLAR I RRADIATION P ATTERNS

DURING EXPERIMENTATION PERIOD AT CHANIA

Various tests were carried out in the beginning in order to settle the multiple parameters calibrations
as well as to acquire trends among parameters and reach the optimum operational conditions. After
numerous measurements, a correlation between irradiation and feed flow was observed as shown in
Figure 45.
Based on these observations, it was discovered that above approximately 600 W/m2, the feed flow is
the maximum possible (6 gpm or 22.7 liters per minute) while for irradiation values lower than 200
W/m2 the feed flow is lower than 3 gpm (or 11.4 liters per minute). System start-up was attempted at
even lower values of incident irradiation; however, at these conditions the pump was clearly unable
to run normally and showed intermittent operation.

F IGURE 45: I RRADIATION

VS

F EED F LOW

ON

16/09/2010

Based on the results of the experiments some general remarks can be made regarding the
relationship between power consumption, permeate flow and irradiation. As can be seen from Figure
46, the relationship between permeate flow and power consumption is nearly linear. One significant
observation is related with the cut-in power value (around 350W) under which the slope of the trend
57

changes (for permeate production less than 2 l/min). Another observation is that above 1000 W the
trend levels off at the maximum permeate flow achieved value of 5.2 l/min. However, it is important
to note that depending on the custom system configuration these boundaries may vary.

F IGURE 46: R ELATION

BETWEEN

P ERMEATE P RODUCTION AND P OWER C ONSUMPTION


DAY : 17/09/2010

ON A MOST LY CLEAR

In Figure 47, measured irradiation values have been plotted with the corresponding power
consumption values for the course of one mostly clear day. The main observation here is that a
certain power value can be achieved at lower irradiation values during the second half of the day than
in the first half of the day. This can be attributed to the light source (diffused or direct) and also to
the SSW orientation of the panels.

Second half of the day:


After-noon

First half of the day:


(Before noon)

F IGURE 47: P OWER C ONSUMPTION

VS

I RRADIATION

ON A MOSTLY CLEAR DA Y :

17/09/2010

Evaluation of all full-day experiments showed the ranges for the important system parameters such
as total permeate production, total energy production and the specific energy consumption.

58

T ABLE 12: O BSERVED V ARIATION

IN

PV-RO S YSTEM P ARAMETERS

Experiments powered by PV in
September

Range

Total Water production (m3)


Total Energy consumption (kWh)
Specific Energy Consumption (kWh/m3)

1.65 - 2.55
1.05 8.58
2.7 - 3.37

Experiments powered by PV in October

Range

Total Water production (m3)


Total Energy consumption (kWh)
Specific Energy Consumption (kWh/m3)

0.33 - 2.15
1.07 7.32
2.73 - 3.50

Depending on the weather conditions and the solar fluctuations, the total daily water production
varies. It can be observed that even in the worst case scenario, the SEC values do not increase
significantly. For comparison measured values during an 8-hour operation directly powered by the
grid can be found in Table 13.
T ABLE 13: RO S YSTEM P ERFORMANCE

WHEN RUNNING ON

Experiment powered by the grid in


October
Total Water production (m3)
Total Energy consumption (kWh)
Specific Energy Consumption (kWh/m3)

5.3. R OLE

OF

G RID P OWER

Value
2.2617
6.39
2.83

L INEAR C URRENT B OOSTER (LCB)

In the last week of September (prior to LCB connection), the sky was heavily clouded (leading to a
(global) irradiation of <200 W/m2 for most time of the day for 4 days) and system operation was
difficult. The motor was abruptly stalling and restarting, leading to significant vibrations.
The motor was starting and stopping almost every 10 minutes. However, that practice is not safe for
the motor over extended time scales, and neither does it produce good results. The situation was
aggravated by the absence of a speed controller for the Pearson pump, the need for which is felt
especially in times when there are large fluctuations in the irradiation levels.
After the connection and operation with the LCB, the following were observed:
Since a greater torque is delivered to the motor at low voltages, the motor doesnt stall in
times of low irradiation
Solar fluctuations smoothened by the LCB. The controller avoids deep and steep changes of
motor speed.
By observing the LCB LEDs which measure the voltage level supplied every moment, it is easy
to predict when the motor speed has to be adjusted (in clear sky conditions).
Significant exchange of voltage with current occurs most of the time, in order to supply with
the maximum instant current possible. However, the limitation of the LCB is 40 A, and current
boosting above that limit is not possible. With the LCB connected, it was observed that the 40

17

The RO unit has a water production capacity of 7.9 m3/day which can be translated to 2.6 m3 for 8 hours. However, due to
capacity limitations of the rectifier used, it was only possible to be produced 2.26 m3 in 8 hours.

59

A corresponded to feed flow of 6 gpm or higher, above which it is not possible to run the
system. Consequently, the limitation of the 40 A does not effectively limit the systems
capacity.
The LCB works better when there are no rapid fluctuations. When irradiation reduces
gradually, then there is better exchange of voltage with current.
The above points can be comprehended better in the following case-studies:

5.3.1. C A S E S T U D Y 1: C O M P A RI SO N B E T WE EN A CL E AR D AY
( W I T HO U T AN LCB) A N D AN A LM O S T CL E AR D AY IN O C TO B ER ( WI T H

IN
AN

S EP T EM B E R
LCB)

Most of the days in October had significant fluctuations in irradiation with different length, depth,
frequency and intensity of cloud coverage. This made the comparison with each other difficult. By
comparison of two clear days it is easier to draw some conclusions about the effect of using the LCB.
A clear day in September, when LCB was not connected, was chosen to be compared with the only
clear day in October with the LCB connected. The two days were compared for a specific irradiance
range (500 W/m2 in the morning until 600 W/m 2 in the afternoon) when no clouds existed to hinder
the irradiance in both days. Although the irradiation in October is lower than the irradiation in
September, useful conclusions can be deduced from the following comparisons:
17 September 2010 NO USE OF LCB: Operating period: 9:58:22 (500W/m2) - 16:04:11 (600 W/m2)
06 October 2010 USE OF LCB:
Operating period: 10:20:29 (500W/m2) - 15:19:29 (600 W/m2)

F IGURE 48: I RRADIATION

OVER TIME ON

17/09/2010

AND ON

06/10/2010

The following observations were made regarding the daily irradiation values for the two days:
First observation is that in September irradiation reaches 500 W/m 2 around 20 min earlier
than it reaches in October.
The maximum irradiation for the day in September was 837 W/m 2 at 13:33, while in October
the maximum irradiation was 746 W/m2 at 13:00.
Finally, at the second half of the day, 600 W/m 2 was reached at 16:04 in September, while in
October the same amount of irradiation was reached almost 40 minutes earlier, at 15:19.
60

The following diagrams were created, to compare parameters for the same irradiation range in both
days:

F IGURE 49: C OMPARISON

OF

V ARIATION

IN

M OTOR V OLTAGE FOR ONE CLEAR


(C ASE S TUDY 1)

DAY WITH THE LCB AND ONE

CLEAR DAY WITHOUT THE LCB

The motor voltage in the day when LCB was connected was considerably lower than the Voltage in
September when LCB was not connected.

F IGURE 50: C OMPARISON

OF

V ARIATION

IN

M OTOR C URRENT FOR ONE CLEAR


(C ASE S TUDY 1)

DAY WITH THE LCB AND ONE

CLEAR DAY WITHOUT THE LCB

Despite the lower instant irradiation in October, the average current supply is much higher in October
due to the work of the LCB.

61

F IGURE 51: C OMPARISON

OF

P ERMEATE F LOW

FOR ONE CLEAR DA Y WITH THE LCB AND O NE CLEAR DAY

WITHOUT THE LCB

(C ASE S TUDY 1)

Even as one would expect much lower permeate flow rates in October (as compared to September)
due to lower irradiation values and operational times, due to the use of the LCB, the average
permeate flow in October tends to approach the value of the permeate flow in September. (On
17/09/2010 without LCB: Qp=4.41 l/min while on 06/10/2010 with LCB: Qp=4.40 l/min).
The flux follows as expected the same trend as the permeate flow. The permeate flux in September is
generally higher than in October, due to the higher irradiation levels, and consequently the longer
operating period with high flowrates than in October. However, the higher the current, the higher the
permeate flux, and the use of LCB on 06/10/2010 compensates for the difference in irradiation and
operational time (On 17/09/2010 without LCB: average flux=17.87 l/h*m 2 while on 06/10/2010 with
LCB: average Flux=17.84 l/h*m2).

F IGURE 52: C OMPARISON

OF

V ARIATION

IN

P ERMEATE Q UALITY FOR ONE CLEAR


(C ASE S TUDY 1)

DAY WITH THE LCB AND

ONE CLEAR DAY WITHOUT THE LCB

62

As far as the water quality is concerned, the following remarks can be made based on Figure 52:
The TDS obtained using an LCB is consistently lower than what is obtained without using an
LCB.
For higher current, permeate flow and flux are higher. And the higher the permeate flow and
flux, the better the water quality. This is the reason why in October although the irradiation
range is lower, the permeate quality manages to be even better than September.
Comparable TDS levels are achieved at much lower irradiation levels with the use of LCB. For
e.g. by observing the above graph, if the LCB was not used, to get a TDS level lower than
550 mg/l (say), we would need to wait until after mid-day! However, if an LCB is to be used,
that quality of water could be obtained since about 11 oclock in October (in September, this
would have been even earlier). Concluding, the starts up times are lower with the use of LCB
and thus operational times are much higher.
Although SEC on the day when LCB is used is lower than the SEC on the day when it is not
used, this does not apply to all other experimentation results in general, due to the varying
solar conditions.

5.3.2. C A S E S TU D Y 2: C O M P AR I S O N B E TW E EN TW O
O C TO B ER W IT H AN D W IT H O U T T H E U S E O F LCB

CO N S E CU T I V E S IM I L A R D AY S IN

The days of 12/10/2010 (without using an LCB) and 11/10/2010 (using an LCB) will be compared in
order to deduce the effects of using the LCB.
12 OCTOBER 2010 NO USE OF LCB

Operating period: 9:25:10 - 18:24:10


Irradiation over time 12/10/2010

1000

Irradiance (W/m2)

900
800
700
600
500
400
300
200
100
0
9:20

9:48

10:17 10:46 11:15 11:44 12:12 12:41 13:10 13:39 14:08 14:36 15:05 15:34 16:03 16:32 17:00 17:29 17:58 18:27

F IGURE 53: D AILY I RRADIATION V ALUES

ON

12/10/2010

The day started sunny until noon when big clouds caused abrupt fluctuations until late in the
afternoon. The lack of LCB made the motor stall during these fluctuations and caused frequent starts
and stops. Accumulated data can be observed at the following table:

63

T ABLE 14: PV-RO O PERATIONAL P ARAMETERS (12/10/2010)


Date: 12/10/2010

Parameters
Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

TDS
permeate
(mg/l)
806.32

Max

857

33.61

38.31

982

51.08

5.25

21.27

1122.64

min

66

17.74

0.09

0.30

0.11

0.44

3.67

2.63

average

419

28.53

21.33

598

43.42

3.42

13.88

848.28

609.26

1.85 (m3)

Total daily water production


Total Energy consumption

5.38 (kWh)

Specific Energy Consumption


(SEC)

2.91 (kWh/m3)

11 OCTOBER 2010 USE OF LCB

Operating period: 8:55:09 16:49:30


Irradiation over time 11/10/2010

1000
900

Irradiance (W/m2)

800
700
600
500
400
300
200
100
0
8:50

9:18

9:47 10:16 10:45 11:14 11:42 12:11 12:40 13:09 13:38 14:06 14:35 15:04 15:33 16:02 16:30 16:59 17:28 17:57 18:26

F IGURE 54: D AILY I RRADIATION V ALUES

ON

11/10/2010

The day started partially sunny until noon when clouds caused intense fluctuations. After 14:00
serious cloud coverage occurred leading to abrupt fall in irradiance. The cloud coverage was just as
intense until system operation was stopped at around 17:00 since the motor speed was very low
leading to very low feed flows and bad water quality. In addition, operation of the motor in low
speeds is not a safe practice.
T ABLE 15: PV-RO O PERATIONAL P ARAMETERS (11/10/2010)
Date: 11/10/2010

Parameters
Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max

901

30.33

37.96

920

50.81

5.02

20.37

1455.73

TDS
permeate
(mg/l)
1045.55

min

57

11.13

0.00

0.29

0.11

0.43

616.68

442.92

average

442

24.29

27.28

654

45.60

3.80

15.42

774.04

555.94

Total daily water production

1.8 m

Total Energy consumption

5.17 kWh

Specific Energy
consumption
(SEC)

2.87 kWh/m3

64

Some general comparative remarks for the above mentioned two consecutive days are the following:
Despite the intense solar fluctuations during these cloudy days, SEC is significantly low
(<2.91 kWh/m3). Although SEC on 11/10/2010 when LCB is used is lower than the SEC on
12/10/2010 when it is not used, this does not apply to all other experimentation results in
general.
Rapid and small fluctuations do not considerably influence the permeate quality and
conductivity.
In contrast, sustained and sharp fluctuations which lead to stopping of the system for long
time (more than 5 min), significantly affect the quality of the water produced. After stopping
and restarting the system, salts which are accumulated on the membrane surface get
diffused on the permeate side deteriorating the quality.
The difference in irradiation distribution, intensity and fluctuations in these two days makes it
very difficult to compare the two days with each other.
Nevertheless, during operation with LCB (11/10/2010) the current is higher than during
operation without the LCB (12/10/2010), the average permeate flow and flux is higher during
11/10/2010, which leads to average lower TDS (see tables with summarized results for these
two days in appendix).

5.4. O BSERVATION

OF SUDDE N QUALITY DETERIORATIO N

After a long series of experiments in heavily clouded conditions during October, deterioration in the
water quality was observed at around the end of the experimentation period. Experimental sessions
powered by the grid at maximum possible feed flow rate (6 gpm or 23 L/min) were carried out to
observe the systems response under constant power conditions and constant feedwater salinity.
However, despite a steady rise of electrical conductivity, the other operational parameters
(membrane pressure, permeate flow) remained constant (Table 16). During 4 hours operation of the
system the total permeate conductivity rose from 830 S/cm to 900 S/cm (under constant power
regime) (Figure 55). Comparatively in the past before the occurrence of the sudden deterioration,
permeate conductivity was as low as 630 S/cm (for the same max feed flow and power).

T ABLE 16: O BSERVATION

OF

O PERATIONAL P ARAMETERS (Q UALITY D ETERIORATION )

Operational parameters
Membrane Pressure (P)
Feed Flow (Qf)
Permeate flow (Qp)
Water recovery ()
Feed electrical Conductivity (ECf)
Max system salt rejection
Range of permeate conductivity (Ecp)

Value
50 bar
6 gpm (23 l/min)
1.2 gpm (5 l/min)
20 % (fixed)
53.2 mS/cm
98.4 %
830 900 S/cm

65

F IGURE 55: PV-RO P ARAMETERS

AS A FUNCTION OF

T IME (Q UALITY D ETERIORATION )

Since the two membrane elements are in series, they can be represented as shown in Figure 56:

F IGURE 56: L AYOUT

OF THE

RO

MEMBRANES

66

Simultaneous permeate samples from both membranes showed that the element 2 had unjustified
high electrical conductivity (ECp1= 453 S/cm while ECp2= 1736 S/cm). The water recovery ratio of
the first element was measured during operation: 1=Qp1/Qf1=11.6% while the salt rejection was
calculated to be:

r1

Cp1
C f1

mS
cm
mS
53.2
cm

0.453

99.1%

Since the two membranes are in series, the concentrate of the first element is the feed for the second
element. For the calculation of the salt rejection of the second element, the mass balance for salt for
the element 1 was considered:

Qf1 C f1

Qc1Cc1

Qp1Cp1

Assuming that the salt concentration in the permeate stream is negligible, and that Qp1
applies for the salt concentration of the concentrate from element 1: Cc1

C f1

Qf1 , it

=60 mS/cm.
1

Thus, the salt rejection for the 2nd element was calculated to be:

r2

Cp2
C c1

mS
cm
mS
60.1
cm

1.736

97.1%

It is obvious that the salt rejection of the 2 nd membrane has considerable decrease in the salt
rejection (manufacture value is 99.4%). Possibility of fouling was eliminated, because of constant
pressure during operation. It can however be speculated that there is a pin-hole leakage in the
membrane. This could be likely attributed to the frequent start and stops of the system during the
preceding days. Since the deterioration happened during the last days of the planned period, due to
time constraints and lack of training, the case of the membrane was not opened to visually inspect
possible damage. Furthermore, in case of damage, replacement with a new membrane and storing it
with preservative until next spring was not considered to be a wise course of action.

5.5. L IMIT ING

FACTORS

Southwest orientation of the roof


The orientation of the roof being SSW and not S which is optimum for the operation of photovoltaics
in the Northern Hemisphere is responsible for some losses in power production. However, by
comparison of the power production with the optimum south orientation and the available orientation
(SSW, 30), it was shown that the losses can be estimated at 0.191 kWh/d or 2.7% of the maximum
available energy.

67

1000
South facing

Solar Irradiation (W/m 2)

900

South-West facing

800
700
600
500
400
300
200
100
0
5:16

7:40

F IGURE 57: D IFFERENCE

IN

10:04

S OLAR I RRADIATION

12:28

FOR

S OUTH

14:52

AND

17:16

Time (hrs)

S OUTHWEST F ACING P ANELS

Manual control of the high pressure pump


It was decided to keep the system as simplified as possible, for practical reasons. For this reason, no
automation mechanisms or controllers were used on site during experimentation in Crete, in order to
control the speed of the two pumps. Instead, the motor speed of the Pearson pump was regulated
manually by a knob on the control panel of the unit. However, the losses in power utilization and the
resulting drop in total daily water production have to be accounted for since due to the large
fluctuations in the solar radiation in the winter season (due to presence of clouds), it was very
difficult to follow the solar pattern accurately and to manually achieve the maximum flow rates at
every instance.

5.6. D IFFERE NT

MODES OF OP ERATION

5.6.1. S E AW A TE R / B R A C KI S H

W AT ER A S F E ED S O U R C E CO N S ID E R AT I O N S

Photovoltaic powered seawater desalination has found applications in remote communities with no
access to the grid and when the option of brackish water as a feed source is not available. Brackish
water desalination is definitely a much more viable process due to the much lower operation
pressures. With some simple assumptions, an impression can be given about the difference between
the two processes.
A typical seawater-RO system without energy recovery ER (=20%, P=53 bar), and a typical brackish
water RO system without ER (=80%, P=15 bar) are assumed.
The ideal hydraulic power to drive a pump is dependent on the pressure difference and the feed flow
(Hydraulic Pump Power=membrane pressure P x Qf i ) while the water recovery is: =Qp/Qf. To
simplify the calculations, constant efficiency and permeate flow Qp is assumed. It can be shown that
brackish water can be 14 times less energy intensive than seawater:

53 bar
20%
14 . This factor proves that energy recovery is critical for SW but not for BW. Thus,
Phydr ,BW 15 bar
80%
the question arising is how much difference the ER makes to the power consumption in a SWRO
Phydr ,SW

68

system. A SWRO system with ER is considered (a Spectra system similar to the one used in
experimentation: 80% ER, 2 hp motor, 20% fixed water recovery, 7.9 m 3/d product flow). Assuming
a pump efficiency of 70%, operating pressure of 53 bar and knowing that Qf=Qp/, it applies for the
hydraulic pump power:
m3
.
5 53 105 Pa 2423W
d
The shaft power - the power required transferred from the motor to the shaft of the pump - depends
on
the
efficiency
of
the
pump
and
can
be
calculated
as:
Phydr 2423W
Pshaft
3461W .
70%
Phydr

Qf

7.9

However, the Spectra system considered incorporates 80% ER and the maximum power which it can
be run is at 1200W (to pressurise the water to 53 bar). Based on this, only 1200/3461=35% of the
power required by the unit comes from the motor and the rest comes from the concentrate stream.
Consequently, in a SWRO system with ER like the one used in this project, the power consumed is
(14 times x 35%)=4.9 times more than a typical BWRO system (=80%, P=15 bar) without energy
recovery.

5.6.2.

O N -G R ID / O F F - G R ID C O N S ID ER AT I O N S

Even as the present study involves autonomous system operation, it is interesting to compare its
on-grid and off-grid applications.18 The advent of feed-in tariffs in Europe (and in other places of the
world) creates profitable opportunities for projects involving solar PVs. The current feed-in rate in
Greece for solar PV plants generating less than 100 kWp is set at 0.441 /kWh 19. On the other hand,
the average cost of electricity in Greece is approximately 0.11 /kWh 20.
Consequently, it would be beneficial in case of an on-grid system to sell all the power produced by
the PV directly to the grid whenever possible, and operate the RO facility using the electricity from
the grid during times of little or no irradiance. Due to the policy currently in place, this would in fact
lead to generation of revenue by selling electricity in addition to the revenue earned by the sale of
drinking water. It would also enable a system operation under a non-fluctuating constant load, which
would lead to better quality of drinking water, as well as increasing the lifetime of the membranes.
Of course, such a system would only be feasible in case there was access to a grid which could
absorb the power generated by the PV, and not to remote coastal areas in general.

5.7. C OMPARISON

OF T HEORETICA L AND PRACTICALLY OB TAINED WATER

YIELDS
The water cost estimation in the sizing approach, was based on the average water yield of one day in
the winter and one day in the summer, using irradiation values for the area of experimentation. The
results of the sizing approach showed that for 10 panels and an average annual water production of
781 m3 (1.16m3/d in the winter and 3.12 m3/d in the summer), the water cost is estimated to be
around 4.74 /m3. The true validation of the yield can only be done over much longer experimental

18
19

An on-grid system may well be considered autonomous if it returns to the grid the same quantity of energy as it consumes.
http://www.solarfeedintariff.net/greece.html

20

http://www.energy.eu/

69

time scales. However, we can compare the practically obtained water yields during experimentation
with the theoretical expectations for the months of September and October calculated as described in
the sizing (Chapter 3) . The water costs are obtained by dividing the daily cost of water production by
the water yield on a specific day.
T ABLE 17: C OMPARISON

September 2010
October 2010

OF

T HEORETICAL

AND

P RACTICAL W ATER Y IELDS

Theoretical
Maximum
(in m3/day)

Theoretical
Cost of Water
(/m3)

Obtained Range
(in m3/day)

Expected Cost of
Water (/m3)

2.61
2.35

3.8
4.22

1.65 2.55
0.33 2.15

6.01 3.89
30.05 4.61

As can be seen, the range of yields obtained from the RO facility is expectedly lower than the
theoretical maximum. The naturally relates to having an inverse effects on the costs of water
production. The main reasons for this differences are:
1. The theoretical calculations fail to take into account the presence of clouds, which mainly
result in large drops in yields. The maximum of the ranges typically correspond to cloudless
days, which expectedly come close to the theoretical maximum
2. Another important factor not considered in the theoretical calculations is the drop in efficiency
of the PV panels with decreasing irradiation values. This results in an over-estimation of the
energy yield in the theoretical calculations especially during dawn / dusk, which reflect in
apparently high water yields.
3. It is assumed for the theoretical calculations that the panels face the south, whereas in reality
they face about 30 west of south. However, this should not result in a major deviation
(about 2-3%) from the theoretical estimations.

70

6. CONCLUSIONS & RECOMMENDATIONS

The purpose of this project, was to examine the technical feasibility and economical aspects of an
autonomous batteryless seawater PVRO system. The conclusions of the study, along with
recommendations are as follows:
Worst case scenario testing
The field study of the batteryless PV-RO system lasted two months in total (03/09/2010-03/11/2010)
and gave crucial conclusions regarding its operation. It can be noted that most of the results obtained
represent the worst case scenario due to the unexpectedly winter-like conditions prevailing for most
of the time. However, the worst case situation in Crete does not resemble the circumstances in
remote areas with high insolation and high water scarcity e.g. Middle East, Australia, North African
countries. Thus, it can be expected that experimental results in such areas would have been better
with respect to water quality and water production.
For this reason, it is recommended that an extended experimentation during summer period in
Greece would approach more the conditions in remote places where a seawater PVRO system could
find its application.
Average solar irradiation is not the deterministic factor for the systems output
It is important to mention that all results from the previous comparisons are indications and not
deterministic quantities. For wind powered RO systems, the average wind speed is considered to be a
useful criterion for comparison of the output over two different days of operation, or for judging
whether a particular location is suitable for setting up a system. It was believed that similar to this
quantity, the average solar irradiation could be used to make such determinations for a PV powered
RO system. However, it was observed that the average solar irradiation is not the critical factor
deciding the performance of a PV-RO system. This is because factors such as extent of cloudiness,
type of cloudiness, time of the day when the cloudiness occurs has a greater role to play in
determining the system output.
Following the conclusions, it was proved to be difficult to compare two different days by quantitive
means such as average solar irradiation alone or by qualitive means such as clear day or cloudy
day. For a complete study, a developed statistical quantity could help in comparing irradiation values
for two different days depending on the system in consideration and its desired output (for e.g.
electrical energy output for a PV solar farm or desalinated water for a PV-RO system).
Preferred configuration of the system in a real scenario
The preferred embodiment of an autonomous seawater PVRO system for remote coastal areas should
be as less complex as possible. Since for this project it was decided to follow the last years research
interest towards elimination of batteries due to the financial and often environmental costs involved
with their maintenance and replacement, this approach is for a batteryless system.A high pressure
pump with incorporated energy recovery, like the Spectra Pearson pump, is suggested for such a very
energy intensive system. The associated feed pump which is used to pressurise the inlet stream to
1.2 bar can be replaced by a reservoir at 12 m high above the high pressure pump. In such a

71

configuration, a second pump is avoided and no automation/controller is needed to synchronise the


speed of the two pumps.
The feedwater coming from the reservoir should be pre-treated via 20MF and 5 MF filters before
passing from the RO membranes. The water after passing through the RO membranes is separated
into a concentrate and a permeate stream. The concentrate stream which is under high pressure,
before being discharged, passes from the high pressure pump for energy recovery while the
permeate stream is stored separately. In a remote area where there is lack of fresh water, an
adequate amount of permeate should be kept for flushing the system thoroughly after a completed
cycle of operation.
In order to achieve maximum power harvest, the PV-array should be installed on a tracker with the
optimum inclination angle per season and facing the nearest pole. 2-axis tracker, according to
literature is not justified economically. Since it is optimal to have a larger PV array to get long enough
operational times (table 7), there is high chance that a large amount of energy can be wasted during
summer season. This energy may be utilised in pumping the seawater into the 12 meter height
reservoir.
Motor Speed Controller
The power utilization from the PV panels is not maximized in the results of this experimentation due
to the limiting factors described in the previous chapter (SSW roof orientation and lack of motor
speed controller). The absence of the controller was felt stronger during cloudy days with intense
solar fluctuations (especially during October). It is suggested that for periods with sustained cloud
cover, a motor speed controller is used which would optimize the motor operation at regular intervals
with respect to solar fluctuations.
Water quality
The results obtained show that the fluctuating operation of an RO system has an impact on the water
quality of the produced permeate since very few average readings of permeate quality on very cloudy
days showed values under 600 mg/l, a value generally suggested by the World Health Organization.
However, it should be noted that no health-based guideline is proposed by WHO or by the European
directive (98/83/EC); Based on consumer taste considerations, it is suggested that drinking water can
get unpalatable at TDS levels greater than about 1000 mg/l. Thus, it can be suggested that water
quality levels similar to the ones obtained during the experimentation can be acceptable for
communities in remote arid areas.
Using a lower recovery ratio than the 20% which was used in this system could probably lead to
improvement of the quality of the water produced for systems under intermittent operation.
LCB issues
The LCB allows extended operating times for the PV-RO system, since it prevents stalling of the
system at low power, and is especially valuable on cloudy days. These extended hours of operation
lead to greater water production, which will ultimately lead to a decrease in the cost of water
production.
Despite the intense solar fluctuations, the Specific Energy Consumption (SEC) ranged around 2.73 3.50 kWh/m3 which can be considered considerably low for renewable energy powered seawater
desalination systems. For comparison, typical SEC values found in literature for similar seawater PVRO
systems are about 4-5.5 kWh/m3 for systems with ER and 6.3-17.9 kWh/m3 for systems without ER

72

(Table 2, Figure 15). Nevertheless, it was not possible to correlate the effect of the use of the LCB on
the SEC, due to the different daily solar patterns.
Although conclusions about the water production cannot be drawn based on the comparisons, the
LCB does influence the water quality for the better, due to increased cross flow velocity resulting from
the higher flow.
Energy Buffer requirement
The quality of the water, in fact, suffers significantly due to the frequent starting and stopping of the
system on cloudy days. The authors from literature [45] state that the decision of whether to use a
system with or without batteries depends on which option is less troublesome and more costeffective: replacing reverse osmosis membranes or replacing batteries. However, this is under the
assumption that utilizing either of the above options makes no difference to the water quality.
However, the experiments carried out at Crete confirm that intermittent operation (with a batteryless
system) does not affect only the life span of the membranes, but also the quality of the permeate
water. A significant rise in TDS was observed due to frequent start-stops of the system, especially on
days with passing clouds.
Hence, in case a batteryless system is to be implemented, it is recommended that an energy buffer is
used, which can smoothen the power curve to an extent that the number of start-stops of the system
is reduced, which can result in a better water quality. Options which could potentially be investigated
in this course are super-capacitors or a small battery arrays.
Water cost
The water cost estimation in the sizing approach, was based on the average water output between
one day in the winter and one day in the summer, using irradiation values for the area of
experimentation. The results of the sizing approach showed that for 10 panels and an average annual
water production of 781 m3 (1.16m3/d in the winter and 3.12 m 3/d in the summer), the averaged
annual theoretical water cost is estimated to be around 4.74 /m3.
From the experimental results it was found that the water yield, which ultimately determines the price
of water depends largely on climatic conditions. So it is difficult to obtain an accurate estimate of the
water yield without experimenting over long time scales. It is important to note that the numbers for
the cost of water are derived by assuming that the capital costs are to be repaid on a daily basis using
the quantity of water produced on that particular day itself and this cannot give an accurate estimation
of the net cost of water.Nevertheless, an impression of the water cost derived from the experimental
results for the months of September and October are given in Table 17.
Although these prices are much higher than the current mains delivered water e.g. in Netherlands
(0.5 ), it is important to note for comparison that in July and August 2010, the cost for transporting
drinking water to 17 arid Greek islands reached 12.5 /m 3. Furthermore, if we consider the cost of
the bottled water which is 0.5/L or 500/m3, we can easily understand that seawater desalination is
an alternative solution in places where fresh water distribution is not an option.
In conclusion, it should be noted that the viability of RE-desalination systems is very site specific.
Based on this research, it can be suggested that PV-RO seawater desalination is promising mainly for
remote arid regions with high insolation (where water scarcity is covered with transportation of water
and there is no access to the grid) and with no potential for exploiting brackish water resources e.g.
in places around the MENA region, deserts and remote islands.

73

74

B IBLIOGRAPHY
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

Ibbotson, B., The Effects of Fluctuating Operation on Reverse Osmosis Membranes . 2010,
Delft University of Technology & University of South Australia: Delft.
Mohamed, E.S., et al., A direct coupled photovoltaic seawater reverse osmosis desalination

system toward battery based systems - a technical and economical experimental comparative
study. Desalination, 2008. 221(1-3): p. 17-22.
Eltawil, M.A., Z. Zhengming, and L.Q. Yuan, A review of renewable energy technologies
integrated with desalination systems. Renewable & Sustainable Energy Reviews, 2009. 13(9):
p. 2245-2262.
Schafer, A.I., A. Broeckmann, and B.S. Richards, Renewable energy powered membrane

technology. 1. Development and characterization of a photovoltaic hybrid membrane system.


Environmental Science & Technology, 2007. 41(3): p. 998-1003.
Vengosh, A., Salinization and Saline Environments. Treatise on Geochemistry, 2003. 9.
Insights from the Comprehensive Assessment of Water Management in Agriculture , in World
Water Week, C. International Water Mangement Institute (IWMI), Sri Lanka, Editor. 2006:
Stockholm.
Al-Karaghouli, A., D. Renne, and L.L. Kazmerski, Solar and wind opportunities for water
desalination in the Arab regions. Renewable & Sustainable Energy Reviews, 2009. 13(9): p.
2397-2407.
El-Nashar, A.M., D. Al Gobaisi, and B. Makkawi, Solar energy for desalination in the Arab
world. Proceedings of Ises Solar World Congress 2007: Solar Energy and Human Settlement,
Vols I-V, 2007: p. 213-221.
Kaldellis, J.K., K.A. Kavadias, and E. Kondili, Renewable energy desalination plants for the
Greek islands - technical and economic considerations. Desalination, 2004. 170(2): p. 187203.
Mygiakis, M., (One in six people lack
access to water), in -. 2010, Eleftheros-Typos: Greece. p. 1-2.
Salinity Levels (Official Website of the City of Phoenix, Arizona, USA). [8 November 2010];
Available from: www.phoenix.gov/waterservices/wrc/yourwater/salinity/index.html.
Guidelines for Drinking-water Quality (Third Edition). 2008, World Health Organization:
Geneva.
Council Directive on the quality of water intended for human consumption , in 98/83/EC.
1998: European Union.
Richards, B.S. and A.I. Schafer, Photovoltaic-powered desalination system for remote
Australian communities. Renewable Energy, 2003. 28(13): p. 2013-2022.
De Munari, A., et al., Application of solar-powered desalination in a remote town in South
Australia. Desalination, 2009. 248(1-3): p. 72-82.
Ghermandi, A. and R. Messalem, Solar-driven desalination with reverse osmosis: the state of
the art. Desalination and Water Treatment, 2009. 7(1-3): p. 285-296.
Sabine Lattemann, I.E., Andrea Schafer, Chapter 2 Global Desalination Situation. Vol. 2.
2010: Elsevier.
Total world desalination capacity close to 60 million m/d. 2009 [30/05/2011]; Available
from: http://www.desalination.biz/news/news_story.asp?id=5121.
Water Pumps for the Desalination Market. 2009 [30/05/2011]; Available from:
http://news.lubipumps-na.com/blog/2009/01/water-pumps-for-the-desalination-market/.
Total desalination capacity by country. 2006 [30/05/2011]; Available from:
http://water.nationalacademies.org/basics_part_3.shtml.
Papapetrou, M., C. Epp, and E. Tzen, Autonomous desalination units based on renewable
energy systems - A review of representative installations worldwide. Solar Desalination for the
21st Century, 2007: p. 343-353.
Charcosset, C., A review of membrane processes and renewable energies for desalination.
Desalination, 2009. 245(1-3): p. 214-231.

23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.

Herold, D. and A. Neskakis, A small PV-driven reverse osmosis desalination plant on the
island of Gran Canaria. Desalination, 2001. 137(1-3): p. 285-292.
Banat, F. and H. Qiblawey, Membrane desalination driven by solar energy. Solar Desalination
for the 21st Century, 2007: p. 271-291.
Richards, B.S., D.P.S. Capao, and A.I. Schafer, Renewable energy powered membrane

technology. 2. The effect of energy fluctuations on performance of a photovoltaic hybrid


membrane system. Environmental Science & Technology, 2008. 42(12): p. 4563-4569.
van Dijk, J.C., Verberk, J.Q.J.C., Heijman S.G.J., Rietveld L.C., Drinking water treatment.

2009: Sanitary Engineering Department, Civil Engineering and Geosciences, TUDelft.


Papapetrou, M., M. Wieghaus, and C. Biercamp, Roadmap for the development of
Desalination powered by Renewable Energy, PRODES.
Thomson, M. and D. Infield, A photovoltaic-powered seawater reverse-osmosis system
without batteries. Desalination, 2003. 153(1-3): p. 1-8.
Loster, M. Total Primary Energy Supply - From Sunlight. 2010 [21 November 2010];
Available from: http://www.ez2c.de/ml/solar_land_area/.
World Solar Insolation Map. Available from: http://www.oynot.com/solar-insolation-map.html.
ADIRA Handbook: A Guide to Autonomous Desalination System Concepts : MEDA Water.
Ahmad, G.E. and J. Schmid, Feasibility study of brackish water desalination in the Egyptian
deserts and rural regions using PV systems. Energy Conversion and Management, 2002.
43(18): p. 2641-2649.
Garcia-Rodriguez, L., Renewable energy applications in desalination: state of the art. Solar
Energy, 2003. 75(5): p. 381-393.
Garcia-Rodriguez, L., Seawater desalination driven by renewable energies: a review.
Desalination, 2002. 143(2): p. 103-113.
Solar Cell I-V curve with MPPT. 2011; Available from: http://en.wikipedia.org/wiki/File:SolarCell-IV-curve-with-MPP.png.
Australian Energy Research Laboratories. [22 November 2010]; Available from:
http://www.aerl.com.au/.
Solar Converters - Product Line. [22 November 2010]; Available from:
http://www.solarconverters.com/product_frame.html.
Richards, B.S. and A.I. Schafer, Design considerations for a solar-powered desalination
system for remote communities in Australia. Desalination, 2002. 144(1-3): p. 193-199.
Gibson, T.L. and N.A. Kelly, Optimization of solar powered hydrogen production using
photovoltaic electrolysis devices. International Journal of Hydrogen Energy, 2008. 33(21): p.
5931-5940.
Thomson, A.M., Reverse-Osmosis Desalination of Seawater Powered by Photovoltaics Without
Batteries. 2003, Loughborough University.
Thomson, M., M. Miranda, J. Gwillim, A. Rowbottom, I. Draisey Batteryless Photovoltaic
Reverse-Osmosis Desalination System, in DTI Sustainable Energy Programmes. 2001.
Cheah, S.-F., Photovoltaic Reverse Osmosis Desalination System. 2004, ITN Energy Systems,
Inc.: Littleton, CO, USA.
Mohamed, E.S., et al., The effect of hydraulic energy recovery in a small sea water reverse
osmosis desalination system; experimental and economical evaluation. Desalination, 2005.
184(1-3): p. 241-246.
Desalination Units powered by Renewable Energy Systems: Opportunities and Challenges. in
International ADU-RES Seminar. 2005. Hammamet, Tunisia: WIP & INGREF.
Bermudez-Contreras, A., M. Thomson, and D.G. Infield, Renewable energy powered
desalination in Baja California Sur, Mexico. Desalination, 2008. 220(1-3): p. 431-440.
Joyce, A., et al., Small reverse osmosis units using PV systems for water purification in rural
places. Desalination, 2001. 137(1-3): p. 39-44.
Thomson, M. and D. Infield, Laboratory demonstration of a photovoltaic-powered seawater
reverse-osmosis system without batteries. Desalination, 2005. 183(1-3): p. 105-111.
Spectra Watermakers - Land Based Applications. [22 November 2010]; Available from:
http://www.spectrawatermakers.com/landbased/about.php.
Mohamed, E.S. and G. Papadakis, Design, simulation and economic analysis of a stand-alone
reverse osmosis desalination unit powered by wind turbines and photovoltaics. Desalination,
2004. 164(1): p. 87-97.

II

50.
51.
52.
53.
54.
55.
56.
57.

ERI. ERI Technology Overview. 2011; Available from:


http://www.energyrecovery.com/index.cfm/0/0/33-Overview.html.
Wood, R. (2009) Small-scale desalination pump combines energy-recovery. Desalination &
Water Reuse.
SpectraWatermakers, Pearson High Pressure Pump Application Guide. 2009.
Reverse Osmosis Desalination Membranes. [22 November 2010]; Available from:
http://www.lenntech.com/products/membrane/sea/general/reverse-osmosis-desalinationmembranes.htm.
Spectra - SSW 5800. [22 November 2010]; Available from:
http://www.spectrawatermakers.com/landbased/products_ssw5800.php.
Murakami, M., Hydropowered Reverse-Osmosis (Ro) Desalination for Cogeneration - a
Middle-East Case-Study. Desalination, 1994. 97(1-3): p. 301-311.
Dako Power - Three Photovoltaic Technologies. 2008 [23 November 2010]; Available from:
http://www.dako.co.za/photovoltaic_types.html.
de Moel, P.J., Verberk, J.Q.J.C., van Dijk, J.C., Drinking Water: Principles and Practices. 2007:
World Scientific Publishing Co. Pte. Ltd. .

III

IV

APPENDICES
A) Datalogger calibrations
Channel in
Datalogger

Parameter recorded

Range

Solar irradiation (W/m2)

2
3
4
5

Motor Voltage (V)


Motor Current (A)
Membrane Pressure (P)
Permeate Flow (l/min)
Feedwater Conductivity
(mS/cm)
Permeate water
Conductivity (S/cm)

(00,05V)
(0-10V)
(0-0.3V)
(2-10V)
(0-10V)

6
7

(0-2V)
(0-2V)

Calibration relation
69.3 V/W/m2
[output V] x 4
[output V] / 0.0045
P = 7.5 x [output V] - 15
Qp=0.98344x [output V]+0.053749
ECf=94.663 x [output V]
Ecp= 1002.36 x [output V]

Gain
8
no
8
no
no
4
4

B) Irradiation data used for the sizing approach (acquired


from JRC)

C) Chosen sizing approach excel sheet

VI

VII

D)

Summarized tables of most representative results

Date: 16/09/2010

max
min
average
Daily water Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

Date: 17/09/2010

Max
Min
average
Daily water Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

Date: 22/09/2010

Max
Min
average
Daily water Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

Date: 30/09/2010

Max
Min
average
Daily water Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

NO LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

854.31
158.53
657.06

29.24
27.00
28.58

54
14
35

1133.60
973.57
1031.95

57.90
39.24

5.31
2.14
4.88

21.52
8.69
19.78

1153.30
10.07
700.83

47.23

TDS
permeate
(mg/l)
828.33
353.33
505.06

2.24
-

defect Voltmeter. Manual log of some Voltage data,


but not enough to calculate energy and SEC.

NO LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

836.70
114.50
604.09

33.79
16.99
29.15

42
0
31

1085.07
1.03
891.42

49.16
0.32
45.65

5.17
0.11
4.41

20.95
0.43
17.87

1939.37
747.91
864.64

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

947
123
556

33.21
29
29.73

39.13
5.63
24.25

1060
163.25
717

49.43
33.25

5.12
0.73
3.60

20.75
2.95
14.60

1443.50
667.72
911.12

TDS
permeate
(mg/l)
1392.91
537.17
621.03

2.55
8.58
3.37

NO LCB

Parameters

43.64

TDS
permeate
(mg/l)
1036.77
479.58
654.39

1.65
5.48
3.32

LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

1033
70
491

30.33
8.64
24.26

41.59
0.03
27.98

1028
1
667

49.11
0.33
43.30

5.06
0.11
3.54

20.51
0.43
14.37

1814.90
765.69
970.49

TDS
permeate
(mg/l)
1303.52
549.94
697.03

1.81
5.66
3.14

VIII

Date: 2/10/2010

LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

Max
Min
average

1026
22
320

30.30
11.64
25.60

40.73
0.00
18.55

987
0
448

49.50
0.30

5.06
0.11
2.37

20.51
0.44
9.61

1770.00
901.00
1220.43

TDS
permeate
(mg/l)
1271.27
647.13
876.55

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

0.83

Irradiatio
n (W/m2)

Voltag
e (V)

Curren
t
(A)

Power
(W)

Pressur
e (bar)

Flux
(l/h*m2
)

ECp
(S/cm
)

TDS
permeat
e (mg/l)

841.11

30.30

43

20.17

1761.20

1264.95

15.61
25.17

0
28

49.09
0.33
41.46

4.98

55.05
473.92

1034.7
2
9.10
673.79

Flow
(l/min
)

0.11
3.49

0.43
14.14

3.66
963.10

2.63
691.73

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

746
99
566

30.32
11.78
23.96

44
0
36

1052
0
869

49.55
0.31

5.04
0.11
4.40

20.45
0.43
17.84

1280.21
720.69
815.95

TDS
permeate
(mg/l)
919
518
586

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

2.08

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

1077
44
307

30.33
9.71
25.87

38
0
18

963
5
437

50.17
9.58
37.32

5.15
0.11
2.48

20.88
0.43
10.05

1732.50
227.97
1000.49

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

1.13

Date: 5/10/2010

max
min
average
Daily
water
Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

Date: 6/10/2010

Date: 8/10/2010

28.38

2.62
3.15

LCB

Parameters

1.94
6.24
3.22

LCB

Parameters

46.53

6.84
3.29

LCB

Parameters
TDS
permeate
(mg/l)
1244
164
719

3.33
2.94

IX

Date: 11/10/2010

LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

901
57
442

30.33
11.13
24.29

37.96
0.00
27.28

920
0
654

50.81
0.29

5.02
0.11
3.80

20.37
0.43
15.42

1452.30
615.23
772.21

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

1.80

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

857
66
419

33.61
17.74
28.53

38.31
0.09
21.33

982
3
598

51.08
0.30
43.42

5.25
0.11
3.42

21.27
0.44
13.88

1120.00
654.00
855.52

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

859
68
247

33.56
17.93
29.37

36.12
0.12
11.28

971
2
314

49.81
0.32
32.52

5.02
0.11
1.63

20.35
0.43
6.63

1120.00
654.00
810.14

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

0.33

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

918
53
335

36.12
18.52
30.74

33.26
0.15
11.17

883
5
328

51.15
0.34
34.77

5.14
0.11
2.00

20.83
0.44
8.12

1536.60
630.49
1014.18

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

0.88

Date: 12/10/2010

max
min
average
Daily water Production
(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

Date: 13/10/2010

Date: 15/10/2010

45.60

TDS
permeate
(mg/l)
1043
442
555

5.17
2.87

NO LCB

Parameters
TDS
permeate
(mg/l)
804.42
469.72
614.46

1.85
5.38
2.91

NO LCB

Parameters
TDS
permeate
(mg/l)
804.42
469.72
581.87

1.07
3.21

NO LCB

Parameters
TDS
permeate
(mg/l)
1103.63
452.84
728.42

2.41
2.73

Date: 17/10/2010

LCB

Parameters

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

903
110
573

30.31
0.00
23.43

41.62
0.00
30.92

1008
0
727

50.67
0.30
44.12

5.02
0.11
4.01

20.37
0.44
16.25

1531.10
528.56
1108.30

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

1.35

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

764
53
438

29.07
0.01
23.82

44.73
0.06
31.60

1117
0
750

49.65
0.32
45.47

5.16
0.12
4.03

20.93
0.47
16.32

1944.00
8.85
1160.59

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

2.15

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

TDS
permeate
(mg/l)

max

746

31.84

31.35

991

21.18

936.89

672.90

7
312

31.56
31.61

23.81
26.99

758
853

50.50
43.07
49.73

5.23

min
average

4.44
5.02

18.01
20.35

795.29
890.35

571.20
639.47

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

2.26

Irradiation
(W/m2)

Voltage
(V)

Current
(A)

Power
(W)

Pressure
(bar)

Flow
(l/min)

Flux
(l/h*m2)

ECp
(S/cm)

max
min
average

656
57
503

30.34
18.72
24.39

50.61
0.03
41.84

1216
1
1018

51.50
0.27

5.17
0.11
4.85

20.97
0.44
19.67

987.24
639.65
867.73

TDS
permeate
(mg/l)
709.07
459.42
623.23

Daily water Production


(m3)
Total Energy
consumption (kWh)
SEC (kWh/m3)

2.09

Date: 19/10/2010

Date: 25/10/2010

Date: 03/11/2010

TDS
permeate
(mg/l)
1099.68
379.63
804.55

4.07
3.03

LCB

Parameters
TDS
permeate
(mg/l)
1396
6
834

6.68
3.11

INVERTER

Parameters

6.39
2.83

LCB

Parameters

49.54

7.32
3.50

XI

E) PV panels characteristics

XII

F) 12/24 V @ 40 Amps Linear Current Booster/Pump Driver


Model: PPT 12/24-40 from Solar Coverters .Inc
Nominal Voltage (V)
12
24
Nominal PV Input (V)

12

24

Maximum PV Input (V)

50

50

Maximum Output Voltage (V)**

15

30

Maximum Output (A)

40

40

Maximum Output (A) short term


- 10 sec

60

60

Efficiency

>95%

>95%

Features:
-

Voltage Limited Output to 15 / 30 V


Greatly increases water flow, even under reduced sunlight conditions (over 40% likely)
Transient Protected on the input and output
Float/Dry Switch input for on/off control
Pump Protection Features - Well Dry, Low Sun, Current Limited, Temperature Limited
LED Display of pump performance
Interface for Optional Digital meter MT-3
Ultra High Efficiency >94%
Rainproof NEMA 3R enclosure

The relation can be derived from the main formula for hydraulic pump power: http://www.engineeringtoolbox.com/pumps-power-d_505.html

XIII

Das könnte Ihnen auch gefallen