You are on page 1of 41

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

net/publication/312153623

Preliminary Study on Floating Photovoltaic Systems on Dams

Working Paper · November 2016


DOI: 10.13140/RG.2.2.36168.11523

CITATIONS READS

0 2,212

1 author:

Michele Barbuscia
University of Lisbon
2 PUBLICATIONS   1 CITATION   

SEE PROFILE

All content following this page was uploaded by Michele Barbuscia on 08 January 2017.

The user has requested enhancement of the downloaded file.


Preliminary Study on Floating Photovoltaic Systems on Dams
November 2016

Michele Barbuscia

1
Table of Contents

1. Introduction ......................................................................................................................... 3
1.1 Photovoltaic Technology ............................................................................................................... 3
1.2 Rating of PV modules .................................................................................................................... 4
1.3 Collecting Sunlight: tilt angle, orientation and shading ................................................................ 5
1.4 Classification of Photovoltaics ....................................................................................................... 7
2. Photovoltaic Global Market................................................................................................... 9
2.1 Recent PV LCOE Compared to Other Classic Technology ............................................................ 10
2.2 Operation & Maintenance Costs ................................................................................................. 11
3. Floating Installations ........................................................................................................... 12
3.1 Historical Revue ........................................................................................................................... 13
3.2 Types of Structures ...................................................................................................................... 17
3.2.1 Plane unique ......................................................................................................................... 17
3.2.2 Modular ................................................................................................................................ 18
3.2.3 Tilted resistant ...................................................................................................................... 19
3.2.4 Semi-submerged................................................................................................................... 20
3.3 Efficiency benefit due to contact with water .............................................................................. 22
3.3.1 Panels Completely Submerged............................................................................................. 22
3.3.2 Panels on Floating Structures ........................................................................................ 27
4 Installations on dams .......................................................................................................... 29
4.1 Portugal Hydropower Statistics ................................................................................................... 29
4.2 The Opportunity of Coupling PV and Hydropower ..................................................................... 30
4.3 Problems Related to Environmental Conditions ......................................................................... 32
5. RETScreen simulation .......................................................................................................... 32
6. Conclusions......................................................................................................................... 34
References ................................................................................................................................. 35
Appendix 1 ................................................................................................................................. 36

2
ABSTRACT – Solar photovoltaic is probably the most common renewable energy in the World, with a
global cumulative installed power of 229 GW by the end of 2015 and production prices that in 2016
reached for the first time that of on-shore wind and in most countries are permanently lower than that
of conventional energies. Due to its continuously decreasing price, the scientific community starts to
look at other possible applications for it, which could improve its efficiency or avoid some of the
problems this technology still have (such as its low efficiency and high need of space), applications that
until few years ago had no chance to be economically convenient. This is the case of floating
photovoltaic (FPV), whose possible configurations the author will describe in this paper, with particular
attention for FPV systems installed on dams’ lakes.

1. Introduction
1.1 Photovoltaic Technology

Solar photovoltaic is the technology that generates direct current (DC) electrical power from
semiconductor when they are illuminated by photons. Solar cells are typically made of semiconductor
materials, which have weakly bonded electrons occupying the valence energy band. When energy
exceeding a certain threshold, called bandgap energy, is applied to a valence electron, the bonds are
broken and the electron is free to move around in a new energy band, the conduction band, where it
is able to conduct electricity through the material. This energy needed to free the electrons can be
supplied by light photons.

Fig. 1: Schematic of a solar cell

Figure 1 shows the idealized circuit run by the electrons from the valence band (ground state)
excited by a photon moving to the conduction band, where a specially made selective contact collects
them and drives them to the circuit to supply an external load; once lost the energy needed to cover
the cycle they come back to the conduction band. The potential at which the electrons are delivered
to the external world is less than the threshold energy that excited the electrons, and it is independent

3
of the energy of the photon that created it. Thus, in a material with a 1 eV bandgap, electrons excited
by a 2 eV (red) photon or by a 3 eV (blue) photon will both still have a potential voltage of slightly less
than 1V.
At the heart of almost any solar cell is the pn junction. This pn junction results from the “doping”
that produces conduction-band or valence-band selective contacts with one becoming the n-side (lots
of negative charge), the other the p-side (lots of positive charge).
For power production, the solar cells are interconnected (usually in series) and encapsulated into
units called PV modules, which are the product sold to the consumer.

Advantages of photovoltaics:
• Fuel source is widely accessible and essentially infinite
• No emissions, combustion or radioactive waste (does not contribute perceptibly to global
climate change or air/water pollution)
• Low operating costs (no fuel)
• No moving parts (no wear); theoretically everlasting
• Ambient temperature operation (no high-temperature corrosion or safety issues)
• High reliability of solar modules (manufacturers’ guarantees over 30 years)
• Rather predictable annual output
• Modular (small or large increments)
• Can be integrated into new or existing building structures
• Can be very rapidly installed at nearly any point-of-use

Disadvantages of photovoltaics:
• Fuel source is diffuse (sunlight is a relatively low-density energy)
• High initial (installed) costs
• Unpredictable hourly or daily output
• Lack of economical efficient energy storage

1.2 Rating of PV modules

While conventional power generators are rated in watts because they are designed to produce this
level of power continuously, PV modules are rated in watts of peak power (Wp). This is the power the
module would deliver to a perfectly matched load when illuminated with 1 kW/m 2 of incident solar

4
radiation of a certain standard spectrum while the cell temperature is fixed at 25°C. These “standard
test conditions”, or STC, are universally applied to rate peak output of a solar cell in a laboratory or a
module out in the field, but rarely occur in real outdoor applications. Generally the irradiance is smaller
and the temperature much higher. Both factors reduce the generated power. In some cases also the
load is not well matched, or the modules among themselves, reducing further the power output. To
enable useful predictions, the energy in kWh generated in one year (or one month or one average day)
is obtained by multiplying the rated power in kWp times the number of “effective hours” of irradiance
falling on the generator in one year (one month, one average day) times the performance ratio (PR).
The PR accounts for losses above mentioned in real operation, those in the wiring, the inverter (whose
efficiency may be 0.90–0.97), etc. Time for maintenance is also included here. The PR in well-designed
installations can vary from 0.7 to 0.8, but may be even lower in warmer climates because the efficiency
of the cell decreases with the temperature [2]. Since the rating power is 1 kW/m2, the number of
“effective” sun hours is the number of equivalent hours that should be applied to a plane with the
same orientation of the PV generator at rating power in order to it receive the same amount of energy
(example: if in real conditions in a day an 1 m2 module receives 6 KWh during 9 hours, the effective
number of hours will be 6 h/d).

1.3 Collecting Sunlight: tilt angle, orientation and shading

The tilt angle is the angle between the plane of the surface in question and the horizontal, and it
has values between 0° and 180°. Usually, the tilt angle to optimize yearly production for fixed non-
tracking arrays of PV modules is some few degrees below the local latitude (there is more insulation in
the summertime). However, annual output is only weakly dependent on tilt. For example, at mid-
latitudes, the difference in annual average effective hours as the tilt angle of the modules varies from
horizontal (0°) to the latitude angle is 10% ca. Thus, at the Portuguese latitude, which is approximately
40°, the difference in annual effective hours between a horizontal surface (∼4.4 effective hours per
day) and a 40° tilted surface (∼4.6 h/day) is between 5% and 10% [2]. The reason is that the sun’s angle
at that latitude varies from 27◦ to 72◦ between winter solstice and summer solstice at this latitude. In
winter, a steeper surface will have more output than a shallow slope, and vice versa in summer, so the
difference between flat and tilted averages out somewhat during the year.

5
Annual solar available energy
640
620
600
MWh

580
560
540
520
0 5 10 15 20 25 30 35 40 45 50 55 60
TILT ANGLE [°]

Fig. 2: Influence of the Tilt Angle on the annual solar available energy at Portuguese latitude
(Image from software RETScreen with NASA insolation data)

For solar installations in the northern hemisphere, the optimum orientation for fixed non-tracking
arrays is true south. But again, it is not very sensitive to minor deviations. An array oriented to the
southeast will get more sunlight during the morning and less in the afternoon. Thus, for an array
installed at 40°N latitude with 40° tilt and oriented 45° east (or west) from the true south, the annual
output will be only 6% less compared to the optimum true south orientation [2]. A way to improve an
array in this sense is installing a tracking system, which could increase the sunlight collected by 15-20%
for single-axis trackers and by 25-40% for double-axis trackers. Tracking systems are generally
employed only in large utility-scale ground-mounted arrays, as their costs are higher than for fixed
arrays.
Another important issue to take into consideration when installing a PV array is the shadow. The
first obvious reason is that the shaded parts produce negligible energy because, although the PV is
able to operate with diffuse light, the amount of energy in diffuse light is rather smaller. But there are
other effects that are more insidious. Even a slight shadow on the corner or edge of a module could
dramatically reduce the output of the shadowed module but also of the entire array. This is because
the modules are connected in series, restricting the flow of current in one cell will restrict the current
(hence the output power) of all other cells in that module and thus in all modules connected in series
with it. The use of bypass diodes in series strings reduces these losses to very acceptable values.

6
1.4 Classification of Photovoltaics

Photovoltaics cells are available in many formats and materials, as can be seen in Figure 3. However,
due to high efficiency and mature technology, silicon photovoltaics are the most popular and
commercially available, with three main product streams: monocrystalline PV, polycrystalline PV and
amorphous (uncrystallized) silicon PV. Amorphous silicon (a-Si) is the most popular format adopted for
thin film technology, and exists in single-, double- and triple-junction, with the triple-junction being
most efficient, ranging from 8% to 10%. The crystallized silicon PVs, due to their physical properties
are available only in a rigid format. Efficiencies for crystalline PV modules are significantly higher, at
18-21% for monocrystalline and 13-14% for polycrystalline. A good alternative to silicon for crystalline
cells is the GaAs, a compound semiconductor formed by gallium (Ga) and arsenic (As) having a similar
structure as silicon, which shows high efficiency, low weight and a high heat resistance, but higher cost
as compared to mono- and polycrystalline silicon. The solar cells based on the less pure materials such
as amorphous inorganic or organic materials, or combination of these have less performance but
significantly lower costs.

Fig. 3: PV material chart [3]

One of the reasons for crystalline silicon to be dominant in PV comes from the fact that
microelectronics has developed silicon technology to a great extent. Not only has the PV community
benefited from the accumulated knowledge, but also silicon feedstock and second-hand equipment
have been acquired at reasonable prices. On the other hand, microelectronics has taken advantage of
some innovations and developments proposed in photovoltaics.
Furthermore, after oxygen silicon is the second most abundant material on Earth and, hence, is
available in almost infinite quantities. It is present not in a pure form, but in chemical compounds,
most commonly with oxygen in form of quartz or sand. The undesired oxygen has to be separated from
the silicon dioxide. To do this, the quartz sand is heated together with carbon powder, coke and

7
charcoal monoxide in an electrical arc furnace to a temperature of 1800°C 1900°C. This produces
carbon monoxide and what is known as metallurgical silicon, which is about 98 per cent pure. This is
still not enough for electronic applications. Only billionths of a percent of impurities are acceptable for
photovoltaics, which falls to ten times less for the semiconductors industry. The raw silicon is therefore
purified further in chemical processes.
For floating applications, crystalline silicon will be considered for the pontoon based installations,
due to its easy availability and good efficiency, while amorphous silicon will be considered for the
flexible systems.

Fig. 4: Evolution of efficiencies of the most performing PV technologies in time (source: NREL)

Efficiency Cost [USD/W] Power [W/m2]


Monocrystalline (Si) 15-25% 0,40-0,60 100-190
Polycrystalline (Si) 13-21% 0,38-0,56 75-150
Thinfilm (a-Si) 6-14% 0,38-0,58 55-100

Table 1: Averaged values of efficiency, power for squared meter and cost per watt of monocrystalline, polycrystalline and
thinfilm modules. Values of efficiency and power per square meter are averages from a list of PV modules built by different
industries (Appendix 1) ; values of cost are from the websites www.pv.energytrend.com and www.pvinsight.com .

8
The values in the table are only indicative, but show as monocrystalline modules have higher
efficiencies for higher prices if compared to the polycrystalline ones, while thin film modules, having
much lower efficiencies for almost the same price, would be convenient only due to their flexibility.

2. Photovoltaic Global Market

The cumulative World’s installed solar PV power capacity increased 29% year-on-year to 229 GW
by the end of 2015 [4]. In only 5 years, from 2010 to 2015, the total global PV capacity jumped over
450% from less than 41 GW. Looking back 10 years, solar’s development has been even more
impressive - from 5 GW of total commissioned PV capacity at the end of 2005 the market has grown
45 times in just one decade. In 2015 Asia fully caught up with European solar pioneers. Out of the 229
GW installed and commissioned at the end of 2015, Europe still accounted for the major global share
at 97 GW (but has already reached the 100 GW mark at the beginning of 2016), but the Asia-Pacific
(APAC) countries had almost reached the same level at 96 GW. America (including both North and
South America) still lags behind at 31 GW. Middle East/Africa (MEA) had only 3 GW of PV at the end of
2015.
For the first time in a decade, Germany in 2016 is no longer the country with the largest cumulative
solar capacity: In addition to installing the largest solar volumes per year, China has now also taken the
lead regarding total solar power capacities - reaching 43.5 GW, equal to 19% of the global market share
at the end of 2015. Germany, now ranked second, has yet to reach the 40 GW mark. Japan scored third
place with 34.3 GW connected to the grid, ahead of the USA with 25.6 GW and Italy with 18.6 GW.
Besides these five countries, no other national market touched the 10 GW level by the end of 2015. In
the first quarter of 2016, the UK, which had 9.1 GW installed by end of 2015, exceeded 10 GW as well.
While India (5.1 GW) could be the next to reach the 10 GW level, potentially already this year (2016),
it will take a few more years for others to get there. At the end of 2015, France had a total solar capacity
of 6.5 GW, Spain was at a level of 5.4 GW and Australia at 5.1 GW.

9
Fig. 5: Evolution of global total solar PV installed capacity [4].
(*)APAC without China
(MEA=Middle East Asia; RoW=Rest of the World)

2.1 Recent PV LCOE Compared to Other Classic Technology

2015 resulted in many new record-low solar power supply prices. The most remarkable contract
awarded was for a 100 MW tender in Dubai (UAE) in early 2015. A record-low 58.4 USD/MWh bid lead
the Dubai Energy and Water Authority (DEWA) to double the original size of the project to 200 MW. In
the meantime, several lower bids were awarded in different regions and countries, often without
financial incentives. The 48 USD/MWh in Peru in early 2016 was another milestone, as well as the 36
USD/MWh in Mexico, but everything was beaten by the 29.9 USD/MWh price offered in the third
round of the Dubai tender. While this lowest bid for the 800 MW project outshines everything seen in
the solar and wind sector so far, even the competing bids in this DEWA tender came in at very low
levels - from 36.5 to 45 USD/MWh. A good indication how quickly solar power prices have fallen can
be seen in the offer from the consortium that won the earlier 200 MW Dubai tender at 58.4 USD/MWh,
and which was bidding in the latest tender at 39.5 USD/MWh – that’s a 30% price decrease within one
and a half years.

10
Capacity Levelized Variable O&M
Fixed O&M LCOE
Plant Type factor (%) capital cost (including fuel)

Dispatchable Technologies
Conventional Coal 85% 65,7 4,1 29,2 100,1
Advanced Coal 85% 84,4 6,8 30,7 123
Advanced Coal with CCS 85% 88,4 8,8 37,2 135,5
Natural Gas-fired
Conventional Combined Cycle 87% 15,8 1,7 48,4 67,1
Advanced Combined Cycle 87% 17,4 2 45 65,6
Advanced CC with CCS 87% 34 4,1 54,1 93,4
Conventional Combustion Turbine 30% 44,2 2,7 80 130,3
Advanced Combustion Turbine 30% 30,4 2,6 68,2 104,6
Advanced Nuclear 90% 83,4 11,6 12,3 108,4
Geothermal 92% 76,2 12 0 89,6
Biomass 83% 53,2 14,3 42,3 111
Non-Dispatchable Technologies
Wind 34% 70,3 13,1 0 86,6
Wind - Offshore 37% 193,4 22,4 0 221,5

Solar Thermal 20% 214,2 41,4 0 261,5


Hydro 52% 78,1 4,1 6,1 90,3

Table 2: averaged LCOEs of most common power production technologies (prices in USD/MWh)
(data from DOE 2011 pubblication)

2.2 Operation & Maintenance Costs

PV systems are designed and installed to be fully functional for a minimum of 20–30 years. The
following is a list of common scheduled tasks:

• Module cleaning to remove soiling.


• Inspection of wires and electrical connections.
• Verification of proper inverter operation.
• Inspection of mechanical mounting system.
• Replacement of broken or damaged modules.
• Vegetation control.

Typically, residential scale systems do not schedule professional O&M visits, as the annual revenues
are too small. The above list of O&M responsibilities are usually shouldered by the owner. Larger
commercial and utility scale systems usually generate enough revenue to warrant having scheduled

11
O&M performed by PV professionals. O&M costs for 4.5–5.0 MW systems installed by two utilities in
the hot and dry southwestern U.S. were monitored over a seven-year period. It was found that annual
O&M cost, on dollar per kWP per year basis, ranged from 0.1 to 0.5% of initial system cost, depending
mostly on inverter repair/replacement cost. At a typical system cost of $8000 per kWP, and by choosing
a middle value of the O&M range at 0.25%, the resulting annual O&M cost is about $20 per kWP/yr.
Another way to express O&M is on an energy basis rather than on an installed capacity basis. A
common energy basis rule of thumb for larger grid-tied PV systems is that the O&M costs run in the
range of 1–10 cents per kWh of energy generated. This cost increases as the system ages and increased
maintenance is required to ensure good performance. Off-grid systems are typically more expensive
than grid-tied systems to operate and maintain on a relative basis due to the additional components
such as batteries.
Monitoring the performance of a PV system is also important, both for recording the power output
of the system and for determining the system health. Since problems can occur between scheduled
maintenance intervals, it is important to minimize the period of time that the system is either down or
not performing at its optimal level. Monitoring a system can be a valuable part of the O&M regimen.
Monitoring is a two-fold process: the system’s output and the local environmental conditions need to
be recorded. This allows detection of system malfunctions as well as system under-performance. In
some large systems it may make financial sense to have a regular on-site crew member to monitor the
system and be able to fix problems as they arise.
For floating applications, the O&M procedures of module cleaning, mechanical support control and
inspections of wires must be done by boat or under-water, which could reasonably double the cost.
This would mean passing to a range of 10-20 cents per kWh of generated energy.

3. Floating Installations

Examples of PV arrays mounted on floating structures have been proposed and already constructed
all over the world. Almost all of them were installed on little farms’ water reservoirs used for irrigation
purposes. The advantages of this kind of solution are numerous.
The first is related to the increasing demand of energy in the agricultural industry as a consequence
of the modernization plans carried out in the last decades. Although water efficiency has improved in
the agricultural section due to the installation of more efficient irrigation systems, electric power
demand has increased substantially.
Another problem farms often have, especially in hot and insulated areas, is the water evaporation
rate in these agricultural reservoirs. Studies have been made in Spain, Turkey and Australia in order to

12
estimate the water losses in this sense. Bengoechea et al. [8] estimated that in Almeria (south of Spain)
water losses due to evaporation amounted to 17%. Martinez et al. [9] estimated water losses of 60
hm3 for the Segura Basin (Murcia, Spain), which means more than 8% of the available water supply for
irrigation purposes. Craig et al. [10] suggested that evaporation phenomena in agricultural reservoirs
in Queensland (Australia) were the cause of a total water loss of 1000 hm3, i.e. about 40 percent of its
total storage capacity. Gökbulak et al. [11] made similar studies from lakes and dams in Turkey and
estimated potential water savings of more than 20%.
Covering all the surface of the water basin decreases temperature and sunlight, hence the water
evaporation. The decrease of temperature and sunlight on the bed of the reservoir discourages also
the growth of algae.

3.1 Historical Revue

As described by Trapani-Redon [12], the project that has received the most news coverage and is
usually claimed as being the first floating PV project, although one for research purposes had been
installed the previous year in Aichi, Japan, was that of the Far Niente wineries in California, USA by SPG
Solar in 2008. The motivation of the wineries’ owner to install the floating PV system on top of their
water reservoir was not to displace land that was used to grow vines, which is a much more precious
resource for their business. The installation was based on modular crystalline PV panels that were
mounted at an optimal tilt angle on top of individual pontoons.

Fig. 6: TTi’s “Floatovoltaic” modular structure incorporated in the Far Niente floating PV system

13
The mounting structure includes walkways between the rows of panels and along the sides to
facilitate maintenance and cleaning of the panels. The array is based on Thompson Technology
Industries (TTi’s) ‘Floatovoltaic’ (illustrated in Figure 6).
The Aichi project, referred to previously, was developed by a research team from the National
Institute of Advanced Industrial Science and Technology in Japan. The aim of the research was to
compare the performances of an array water cooled and another air cooled. The panels in this case
were installed at a slight tilt (1.3° south face) on top of a foam polystyrene floating board. The biggest
project among the first ones that have been developed is that of Bubano, Italy (500 KWp). Here, the
buoyancy for the installation is maintained with hollow polyethylene cubes at the two opposite edges
of the structure which the panels are fixed to (Figure 7). An interesting feature is that this system was
the first that was exposed to snow and ice. The impact was only on the surface panels, because the
climate was not so cold for the water to freeze. Access to this structure is solely by boat, as the array
is situated far from the shore.

Fig. 7: Schematic design of the Bubano’s installation

Further on the topic of evaporation of water reservoirs, a research team in Spain developed a
24kWp floating PV array on a water treatment reservoir in Negret, in 2009, which was expanded to a
300kWp array later the following year because of its good performance. This project is a collaboration
between the Polytechnic University of Valencia and the company CELEMIN ENERGY. The array is made
up of a number of modules each holding two panels, tilted at 10° (facing south) and fabricated by
rotomoulding using medium density polyethylene (refer to Figure 8). The platform has inserts for the
electrical cables and the metal struts from the top. Its base is smooth and rounded (designed to protect
the reservoir’s geomembrane), and each of the modules are connected together with metal pin links,
creating a flexible elastic system able of deforming to the concave profile of the reservoir according to

14
the changing water levels. Recorded annual electrical yield for the Negret PV array averaged at just
under 30GWh since 2009, at a capacity factor of approximately 13.5%. Economics given by Ferrer-
Gisbert et al. indicate that almost 45% of the costs were pontoon-associated costs. The economic
analysis highlighted in the same paper indicates profitability for the array installation at the Negret
site, with an internal rate of return of 12.65%, which does not include any economic savings from the
reduction in water loss.

Fig. 8: Polytechnic University of Valencia and CELEMIN ENERGY floating PV design

In 2010, also in Italy, at Petra Winery, was realized the first project with a tracking system that
rotates the panels according to the sun’s motion. This was designed and constructed by Terra Moretti
Holdings, while the research group Scienza Industria Tecnologia (SCINTEC) was responsible for the
safety systems. The motivation of this installation is similar to that of the Far Niente winery, because
it is also situated in a winery. The structure of the array is made almost entirely of metal struts, with
buoyancy (and integrated tracking system) underneath maintaining the system afloat. The structure is
designed to hold the crystalline panels at an optimal tilt angle of 40°, while changing the orientation
tracking the solar motion.

15
Fig 9: Some existing plants

The same group, SCINTEC, also developed the installation at Lake Colignola, Italy, in 2011. Also this
has a tracking system similar to the Petra Winery one. The interesting aspect of this development was
the utilization of mirrors to reflect additional solar radiation onto the PV panels. The Lake Colignola’s
project was mounted horizontally rather than tilted, with the mirrors situated on the south and north
faces of each panel. The mirrors are placed at an angle 60° to the water, and this is expected to double
the effective solar radiation on the panels. This also implies higher temperatures, so the proximity of
the PV panels to the water surface is a key aspect in cooling the array in order to maintain decent
efficiencies. Testing estimates a 60-70% increase in annual yield compared with conventional fixed-
ground mounted installations.
SCINTEC took part in other projects, among which that developed in collaboration with the French
company Osesol, interesting because it was mounted on a structure that was constructed entirely from
PVC, which resulted in some cost savings as compared to the installations in Italy, which required metal
struts. Another very interesting project developed by SCINTEC in 2012 was that built in Cheongju,
Korea, in collaboration with Techwin and Koiné Multimedia. This project was especially challenging
because of the climatic conditions it was set in, with the water surrounding the installation subject to
freezing temperatures during the winter months. Special considerations were taken in the choice of
the individual components within the installation, so that they were able of withstanding the seasonal
freezing of the water.

16
Maybe the most active company in the last years is the French Ciel et Terre, whose first installation
was at the Piolenc site, in France. Here, the PV panels are held tilted towards the sun with metal struts,
which also links the array together, and is kept afloat with buoyancy underneath each row within the
array. Ciel et Terre has submitted an application for this same site for a 12 MW installation, which
would make it the largest installation to date. Ciel et Terre has already constructed many floating PV
systems, mainly in India and Japan, among which one of the most recent is that in Okegawa, Japan
(2013), which holds the title for the largest floating installation at almost 1.2 MWp.

Fig. 10: Part of Kagawa’s 696 KWp plant, installed by Ciel et Terre

3.2 Types of Structures

Floating photovoltaic systems can use structures with different shapes, weights and costs. These
can be classified for shape in four categories.

3.2.1 Plane unique

This was the first solution used. It can be composed of cubic (or similar) buoys forming a floating
base, on whose top one can locate the PV modules, as in Aichi, Japan in 2007 or in Solarolo, Italy in
2009, or lateral lines of cubic buoys with a metallic structure supporting the panels, as in Bubano
installation. It is an economic solution, as most of the companies producing floating structures already
have this kind of basic buoys on production, as they are typically used for docks, platforms and piers.
This is also the most rigid structure; it would never work for an offshore marine solution because of
the waves, but it is good for a little lake. It is not easy to foresee a price for this kind of structures, as it
would depend on many variables regarding dimensions, materials used and shape. The company Otto,

17
from Padova, Italy, which produced the floating structures for the first italian plants, presents a price
list that gives a cost of approximately 50 €/m2, without mooring system.

Fig. 11: Example of floating platform built with cubic buoys

3.2.2 Modular

This system is composed of modular floating structures, each supporting one or two PV panels, tied
all together. It is the most ductile structure, as it can have every shape and dimension, and also change
them in time. This could be useful for instance to move the system throughout the lake from summer
to winter to catch more sunlight, in case the available space for the structure is not enough for it to
maintain the previous shape. Every module is typically composed of a main float supporting the
panel(s) and a secondary one for allowing maintenance. Usually the main float that support the panel
has a tilt angle, which increases its performances. It is difficult to give even an indicative price, as the
companies selling this kind of structures sell the entire plant, already mounted with panels on top.

18
Fig. 12: Ciel et Terre High-density polyethylene (HDPE) base structure.

In Figure 12 can be seen Hydrelio, the structure proposed by Ciel et Terre, the French company that
is installing the last major plants in Japan and Europe, such as the Yasugi’s 1098 KWp plant and the
Okegawa’s 1180 KWp plant. On their website some technical characteristics are shown, such as that
Hydrelio can withstand up to 118 mph winds and changes in water levels of up to 20 feet, that it is UV
and corrosion-resistant, and has a Link resistance of 3 000 DaN (equivalent to 3 tons). An indicative
price the company gives for this kind of structures is 100 €/structure.

3.2.3 Tilted resistant

This is a structure used mainly for sea applications, due to its higher costs. It is composed of plastic
buoys, on whose top is mounted a metallic structure supporting the panels with a tilt angle. As
described by Young-Geun Lee et al. [13] , who run some mechanical tests on the structures, the buoys
are made of pultruded fiber reinforced polymeric plastic (PFRP), which has strong characteristics such
as a high corrosion-resistance.

19
Fig. 13: One of the structures during tests

3.2.4 Semi-submerged

Various variants of submerged structures have already been designed and tested, but still none of
them is nowadays used for a full-scale plant. The two big different alternatives are: a rigid structure used
with crystalline panels and a flexible solutions with thinfilm panels.
In the rigid solution, a polyvinyl chloride (PVC) pipe is used as a buoy and a hollow square
aluminium as support for the panel and as heat sink. The aluminium allows a good heat transfer
between the water and the panel, maintaining lower the panel’s working temperature. Studies on this
kind of structures and comparisons of performance with conventional PV modules are performed by
Azmi et al. [14,15].
The flexible solution is made up of a thinfilm panel directly put into the water, with some lateral
buoys to keep it near the water level. This is the best configuration for the module’s performances.
Studies on these kind of structures are described by Trapani and Millar [16, 17].

Fig. 14: Example of flexible thin film system

20
Fig. 15: Scheme of the tested semi-submerged rigid structure

Efficiency
Structure O&M Gain Tilt Angle Cost Panels
Gain

Modular 3% - 5% 0% Yes 100 €/piece (*) mono/poly

Unique 3% - 5% 0% Yes 50 €/m2 (**) mono/poly


Semi-submerged
5% - 7% 0% No 50-70 €/m2 (***) mono/poly
rigid
Suberged panel 7% - 10% ≈10% No 10-30 €/panel thin film

Metallic 0% - 3% 0% Yes 200 €/panel(****) mono/poly


Table 3: Pro and cons of the different types of structures with estimated costs and compatibility with panels
(*) Ciel et Terre selling price (other more expensive models available on common selling website)
(**) calculation based on a structure example proposed by the company Otto
(***) avg cost of pvc pipes= 13€/m; avg cost of aluminium sheet= 30 €/m
(****) price of the portuguese company Lindley

21
3.3 Efficiency benefit due to contact with water

3.3.1 Panels Completely Submerged

Two main effects increase the efficiency of a commercial PV panel placed in water:

 reduction of light reflection

 absence of thermal drift

The light reflection on a commercial PV panel is related to the material used to shield the PV active
material. In most panels, this is glass with a refraction index n ≈ 1.55. An intermediate layer of water,
which has a refraction index n ≈ 1.33 changes the reflected fraction of an incoming perpendicular ray
approximately from 4.5% to 2.6%. This effect of course is enhanced if light is not perpendicular and
increases for wide incidence angles.
Thermal effects are much more important and are characterized by a set of temperature
coefficients. In general six parameters are given:

1. short-circuit current (Isc)

2. maximum power current (Imp)

3. open-circuit voltage (Voc)

4. maximum power voltage (Vmp)

5. maximum power Pmp = Imp • Vmp

6. global panel efficiency (ᶯ).

Table 4 shows some of these parameters directly measured at the temperatures of 25°C and 60°C.

Table 4: Average parameters of 72 single crystalline silicon solar cells ASE-100-DGL-SM modules,
measured at cell temperature Tc and irradiance of 840 W/m2.

22
M. Rosa-Clot et al. [18] present a table (Table 5) that summarizes outdoor measurements of
effective temperature coefficients for various commercially available silicon modules. In the table, the
units for the temperature coefficients have been normalized to 1/°C by dividing the coefficient by the
value for the parameter at ASTM Standard Reporting Conditions (1000 W/m2, AM=1.5, 25 °C).

Table 5: Typical effective derivatives temperature coefficients for commercial modules at 1000 W/m2
and AM=1.5 measured outdoors.

For lakes and sea the typical range of temperature of the water in summer is between 10°C and
20°C; this numbers have to be compared with the temperature of a standard panel (without forced
cooling) with good weather conditions in summer, which may even reach 80°C. For the sake of
comparison, the average temperature with a good weather condition of 800 W/m2 can be assumed to
be 60°C. In this case it is easy to see that the temperature difference between a submerged PV panel
and a standard panel exposed in air is of about 40°C with a gain in average of 20% for single crystalline
silicon panels, 15% for polycrystalline panels and 10% for amorphous silicon panels.

The strong improvement in efficiency discussed in the above section is contrasted by the change in
solar spectrum at different water depth. As a matter of fact, clean water is a strong light absorber but
fortunately this absorption occurs mainly in the red-infrared region. In Figure 16 it is shown what
happens to the solar spectrum at different depth.

23
Fig. 16: Variation of solar radiation with water depth with superposed silicon
efficiency spectrum in arbitrary units [18]

It has been superimposed the silicon efficiency spectrum in order to evidence that the hindrance of
solar radiation takes place in the region where PV conversion is less effective. It is possible to estimate
the gain or loss of efficiency, with respect to a PV panel in a standard position by studying the solar
spectrum dumping at a given water depth folded with the efficiency spectrum of PV material,
corrected by the temperature drift.
In Figure 17 the behavior of mono-, poly-crystalline and amorphous silicon are given as a function
of water depth. It is evident that the efficiency is increased or reduced, depending on the choice of the
PV material and on the operating depths.

24
Fig. 17: Efficiency of mono-crystalline, poly-crystalline and an amorphous module with the water depth [18].

In their experiment Marco Rosa-Clot et al. measured the performances of three identical mono-
crystalline silicon panels, panel P1 exposed in air, the other two P2 and P3 submerged respectively 5
cm and 40 cm under the water level. Results confirmed what was calculated theoretically:
 P1 and P2 have almost the same output at the very beginning of the experiment but, after a
few minutes, P1 begins to warm up until reaching 40°C, while the water temperature is 15°C.
The calculated gain in efficiency is about 10%, which matches the expected value.
 P3 submerged 40 cm under the water surface sees its efficiency reduced of about 15%, while
the authors expected value was 14%.
 Submerged panels have much less problems for cleaning than the ones exposed in air. In fact,
although both have been periodically cleaned, the non-submerged panel had on long term an
efficiency loss of 10% while the submerged one did not suffer any decreases of efficiency.
 When the panels P1 and P2 have the same temperature, P1 has a slightly lower efficiency by
about 2%.

Another similar study has been performed by Trapani and Millar in 2013 on thin film panels [19].

25
Fig. 18: Thin film panels 570 Wp experimental floating plant (left) and the control array of 190 Wp (right)
used by Trapani et al.

Two PV systems, one ground-mounted of 190 Wp and the other floating of 570 Wp, have been
installed to experience identical diurnal and environmental parameters throughout the whole
experiment period. The floating system was divided in two parts, the two panels on the left hand side
in Figure 18 had normal buoyancy, while the remaining four panels had extra buoyancy in order to
eliminate the layer of water between the panels and the air. A 3 days comparison of the operational
performances was conducted on the three sets of PV panels. In Figure 19 are shown the different
temperatures measured on the two systems.

Fig. 19: Thermal scans of the ground-mounted (left) and the floating system (right).

The experiment results are shown in the chart in Figure 20. The better performing panels were
those from the floating array that had the added buoyancy to help ensure that no water remained on
top of the panels. There was an average 15% variation in output between the two sets of floating PVs,
which shows how detrimental a ~10 mm layer of water can be to the electrical output. The recorded
data shows a maximum improvement of 8% from onshore to offshore, and an average improvement
of 5% in electrical output over the 3 days. This improvement can be attributed to the cooling effect of

26
the water underneath the panels. The simplicity in manufacturing and deploying the floating PV array,
combined with the added cooling benefit could be key in feasibly moving such arrays from ground
mounted to offshore environments (provided favorable economics). The presence of water has though
shown to have detrimental effect on the output of the panels. An initial 1% reduction in efficiency is
due to accumulation of dirt on the panels, which leads to future exploration of better self-cleaning
mechanisms for the array.

Fig. 20: Electricity output from the 3 sets of PV panels, each operating differently (1 set of 2 PV panels floating with extra
buoyancy to allow them to be higher than the waterline, 1 set of 2 PV panels floating just at the waterline and 1 set of 2 PV
panels installed on ground). Variation in electricity conversion resulting from i) sediment accumulation, ii) water
absorption and iii) water cooling.

3.3.2 Panels on Floating Structures

Young-Kwan Choi [20] studied the performances of two floating plants installed on the
Hapcheon dam reservoir, one 100 kW and one 500kW, comparing them to a 1MW on-land plant
installed in Haman, 60 km away from the Hapcheon lake, which is a distance sufficiently low to assume
equal solar irradiance and similar weather conditions. The two systems are mounted on plastic floating
structures, hence the panels are not in direct contact with the water. In both the floating plants, the
panels have a tilt angle of 33°, while in the on-land plant panels are tilted of 30°.

27
Fig. 21: K-Water 100 kW (a) and 500 kW (b) plants in Hapcheon

For an accurate comparison, analysis days with blackouts, maintenance and data errors were
excluded from the comparison. The analysis period consisted of one year starting from February 2012
to January 2013 for the comparison between the 100 kW floating and the 1 MW on-land plants, and
six months starting from October 2012 to March 2013 for the comparison between the 500 kW floating
and the 1 MW on-land. To compare two plants with different capacities the 1 MW plant’s daily average
energy generation has been converted into the capacity of the other plant, 100 kW or 500 kW, before
being compared.
As a result of the first comparison, the capacity factor of the 100 kW and the 1 MW were 17.6%
and 15.5% respectively, and the daily average generated energy (after being converted) were 421
kWh/day and 3’486 kWh/day respectively.
As a result of the second comparison, the capacity factor of the 500 kW and the 1 MW were 17.1%
and 15.5% respectively, and the daily average generated energy (after being converted) were 2’044
kWh/day and 3’491 kWh/day respectively.

Fig. 22: Comparison of daily average generated energy

As can be seen in the graphs in Figure 22, even if the panels are not submerged in the water their
efficiency is improved by the vicinity of the water, which keeps lower their working temperatures.

28
4 Installations on dams

As the purpose of this paper is to study the viability, and hence the economic convenience, of
installing a floating photovoltaic system on dam’s lakes, there are some variables to take into account
during the design of the system that irrigation reservoirs do not present, such as the bathymetry
variation, the incident waves and the wind. Dam’s lakes are usually much bigger than irrigation
reservoirs, and can easily have waves capable of damaging the floaters, due to high wind and variations
of the inlet rivers’ water flow. In addition, both the inconstant turbine’s water flow and the
precipitation falling on the entire watershed providing water to the dam contribute to make the
variations of the lake’s bathymetry unforeseeable, making harder (and more expensive) the mooring
lines design. Another big disadvantage not to be underestimated is the shadowing effect of the
mountains surrounding the lake, or of the dam itself, which, as already explained above, could
represent a relevant production loss. Furthermore, as dams are usually situated in mountainous areas,
there could be complications related to the wind blowing on the structure. All these are key factors
that will increase the cost of the system and could represent the difference between a convenient
project and a non-convenient one.
On the other hand, floating on a dam’s lake has relevant benefits, like taking advantage of a big
unused space and having a grid transmission line already in place. Depending on many variables, there
is also the possibility to couple the photovoltaic and hydro technologies in order for the PV to work as
a ‘virtual storage system’.
Luis E. Teixeira et al. [21] describe how it would work a coupling hydro-photovoltaic, installing both of
them on a dam in Brazil that until now has operated as a drinking water reservoir. That system
reproduction, run with the software Homer, indicates as the optimal solution the installation of a 227
kW hydroelectric plant operating with a 60 kW photovoltaic system. This combination would result in
an initial cost of USD$ 1715.83 per kW and a cost of energy of USD$ 00.59 $/kWh.

4.1 Portugal Hydropower Statistics

Area: 92’200 sq km
Population: 10’460’000
Installed hydropower capacity: 4’455 MW + 1’343 MW pumped storage (2014)
Hydropower generation: 16.16 TWh (2014)

29
With no fossil fuel resources or reserves of its own, Portugal has had to depend on imports to meet
its domestic demand for oil and gas. This situation, coupled with European Union targets to cut carbon
emissions, has led to substantially increased interest and investment in renewable sources of energy
over the last decades. In 2014, renewables accounted for 62% of the country’s energy mix in terms of
electricity generated. Hydropower amounted to 31% of the mix, occupying half of the total share of
renewables. With its 5902 MW installed (with the pumped storage systems), Portugal is the 11th
country in Europe as hydropower electricity generation.

4.2 The Opportunity of Coupling PV and Hydropower

The Australian company Sunengy, which from 2005 produces floating photovoltaic systems,
proposes the possibility to achieve a ‘virtual storage’ of the solar generated power without changing
the basic hydro turbine at all. Power production from a photovoltaic plant can allow a corresponding
reduction in the water flow through the hydropower turbines [22]. This would preserve that water for
later use, potentially during more valuable hours in a day (such as the early evening peak load that is
typical in many locations).
Most hydropower plants cannot generate electricity at full capacity 24 hours per day because there
is not sufficient annual water flow through the dam catchment to maintain the average flow rate that
is required at full power. Hence, both the turbine and its transmission line are not used at their full
potential. This capacity utilization factor shows worldwide a range from 20% and 60%, mostly around
40% and still lower in dry periods when water is scarce. So, since the annual water flow availability
through a hydro system is the principal limit to the hydropower energy resource, a co-located solar
component of generation can allow increased dispatch-able energy production, making more full
usage of the grid connection line. This can work only in case of installation of big dimension solar PV
plants, and could potentially double the total energy delivered from the original hydropower plant.

Hydropower is typically used as a supplement to fossil fuels as it is limited by the amount of water it
has available to use on a daily and seasonal basis. The chart in Figure 23 shows how coal and hydro are
often combined to provide base and peak loads.

30
Fig. 23: Typical coupling Hydro-Coal

The fossil power plant is typically used to provide the base power load until its rated power, while
the hydro plant starts working when the demand exceeds this value.

Fig. 24: example of coupling Hydro-Solar PV

Chart in Figure 24 shows how an hydropower turbine and a solar PV plant could combined, with
the PV producing during the central hours of the day, when there is sun irradiation, and the hydro
working the remaining time and being available to fill in through cloudy periods. If the solar generation
capacity equals or exceeds the maximum hydroelectric output, almost no water need to be consumed
from 8am to 4pm on most days, as the power could be supplied by the PV system.
This kind of benefit can be achieved with the installation of big PV systems, and their success
anyways depends on the daily load and solar irradiation profiles. However, if well matched, such a

31
combined plant would be a much more reliable power producer than hydropower alone, as the system
would be far less dependent on water supply; in drought conditions it would be possible to continue
daytime solar power generation from the PV with no consumption of water at all.

4.3 Problems Related to Environmental Conditions

As previously mentioned, being located on a dam, a floating photovoltaic system would need a
higher structural resistance if compared to the ones on irrigation lakes. This is due to the more critical
environmental conditions it should work in. The biggest difference is related to the presence of wind,
which in Portuguese lakes could reach 20 knots of velocity, and could damage the structure in two
different ways, directly with its pressure on the panels (especially if tilted), and creating waves up to 1
meter of height. Wind conditions depend on the shape and, in particular, on the length of the lake.
The wind blowing on the structure would require stronger supporting structure for the panel, while
the incident waves would require a better mooring system. Another relevant factor would be the
current, which could reach 5 m/s of velocity, and again would affect the mooring system. Last but not
the least, dams have significant bathymetry variations, which can be up to 5 meters daily and to 20-30
meters yearly. The daily variations are due to the turbine’s flow and cannot be precisely foreseen,
while the yearly ones are related to the precipitations and are gradual and easily foreseeable.

5. RETScreen simulation

The software RETScreen has been used to simulate a floating PV system on a hypothetic dam in the
north of Portugal. It is a 500 kW plant, composed of 2000 panels of 250 W each. The prices of the
floating structures have been set at 200 €/m2, value provided by a Portuguese company producing
buoys and floating pontoons. In Fig. 25 is shown the Cost Analysis resulted from the simulation, while
in Fig. 26 and 27 the Financial Analysis. The LCOE, indicated as Energy production cost in the Financial
Analysis, is about 400 €/MWh.

32
Fig. 25: Cost Analysis

Fig. 26: Project costs and income summary (left) and Yearly cash flows(right).

33
Fig. 27: Financial viability

6. Conclusions

Floating photovoltaic is already a reality for little size water tank and artificial lakes, where there is
no need of a mooring system to the bottom of the basin and environmental conditions are favorable.
Moving this systems to bigger lakes, particularly to dams, adds several problems related to the
environment (such as wind, snow, currents…), water depth, which means raising up the entire system’s
cost. The cost of this kind of systems, assuming that they can be directly connected to the hydropower
transmission line, is mainly affected by the structure and the mooring system. Of course, also the
choice of the panels plays an important role, but it is strictly related to the choice of the structure.
These are the key factors that one must work on for a further study on this matter.

34
References

[1] John A.Duffie, William A. Beckman, Solar Engineering of Thermal Processes, 4th Edition, 3-199.
[2] A. Luque, S. Hegedus, Handbook of Photovoltaic Science and Engineering, 2nd Edition, 1-165.
[3] A.K. Pandey, V.V. Tyagi et al. Recent advances in solar photovoltaic systems for emerging trends
and advanced applications.
[4] Solar Power Europe, Global Market Outlook for Solar Power/ 2016-2020, 13-23.
[5] Christian Breyer, PV LCOE in Europe 2014-30.
[6] Fraunhofer Institut for Solar Energy Systems ISE, Levelized Cost of Electricity Renewable Energy
Technologies Study (November 2013).
[7] Agora Energiewende, Fraunhofer ISE, Current and Future Cost of Photovoltaics (February 2015).
[8] Bengoechea JM, Pérez Cobos J, Pérez Parra J, López Segura JG. Evaluación de las pérdidas de agua
de riego en el Campo de Dalías. In: Symposium sobre el agua en Andalucía; 1991. Córdoba,
España.
[9] Martinez-Granados D, Maestre-Valero JF, Calatrava J, Martínez-Álvarez V. The economic impact of
water evaporation losses from water reservoirs in the Segura basin, SE Spain. Water Resour
Manag 2011; 25:3153-75.
[10] Craig I, Green A, Scobie M, Schmidt E. Controlling evaporation loss from water storages. NCEA
Publication No. 1000580/1; 2005. Queensland, Australia.
[11] Gökbulak F, Özhan S. Water loss through evaporation from water surfaces of lakes and
reservoirs in Turkey. Off Publ Eur Water Assoc; 2006. http://www.ewa-
online.eu/tl_files/_media/content/documents_pdf/Publications/E-
WAter/documents/40_2006_07.pdf.
[12] Kim Trapani, Miguel Redon Santafé, A review of floating photovoltaic installations: 2007-2013
[13] Young-Geun Lee, Soon Jong Yoon, Design and installation of floating type photovoltaic energy
generation system using FRP members.
[14] M. Y. Othman, K. Sopian, Study on Electrical Power Output of Photovoltaic and Conventional
Photovoltaic.
[15] Z. A. A. Majid, M. H. Ruslan, K. Sopian, M. Y. Othman, M. S. M. Azmi, Study on Performance of 80
Watt Floating Photovoltaic Panel.
[16] K. Trapani, D. L. Millar, The thin film flexible floating PV (T3F-PV) array: The concept and
development of the prototype.
[17] K. Trapani, D. L. Millar, Floating photovoltaic arrays to power the mining industry: a case study
for the McFaulds Lake (Ring of Fire).
[18] M. Rosa-Clot, P. Rosa-Clot, Submerged Photovoltaic Solar Panel: SP2.
[19] K. Trapani, D. Millar, The thin film flexible floating PV (T3F-PV) array: The concept and
development of the prototype.
[20] Young-Kwan Choi, A study on power generation analysis of floating PV system considering
environmental impact.
[21] L. E. Texeira, J. Caux, A. Beluco, I. Bertoldo, J. A. S. Louzada, R. C. Eifler, Feasibility study of a
Hydro PV hybrid system operating at a dam for water supply in Southern Brazil.
[22] P. Connor, Sunengy, Liquid Solar Array Synergy with Hydropower.

35
Appendix 1

RenSMART Solar Panel


Comparison Table

Area Watts per Peak


(square square Output
Product Manufacturer meter) meter Efficiency % Technology (Watts)

NE-80EJEA Sharp 0,64 124,15 12,41 Polycrystalline 80


NE-165UC1 Sharp 1,3 126,83 12,68 Polycrystalline 165

NE-170UC1 Sharp 1,3 130,67 13,07 Polycrystalline 170

ND-123UJF Sharp 0,99 123,95 12,39 Polycrystalline 123

ND-130UJF Sharp 0,99 131 13,1 Polycrystalline 130

ND-176UC1 Sharp 1,32 133,33 13,33 Polycrystalline 176

ND-198UC1 Sharp 1,48 133,6 13,36 Polycrystalline 198

ND-224UC1 Sharp 1,63 137,41 13,74 Polycrystalline 224

NU-U230F3 Sharp 1,63 141,09 14,11 Monocrystaline 230

NU-U235F1 Sharp 1,63 144,16 14,42 Monocrystaline 235

NA-V115H5 Sharp 1,42 80,89 8,09 Thin Film 115

NA-V121H5 Sharp 1,42 85,11 8,51 Thin Film 121

NA-V128H5 Sharp 1,42 90,03 9 Thin Film 128

NA-V142H5 Sharp 1,42 99,88 9,99 Thin Film 142

NA-V135H5 Sharp 1,42 94,96 9,5 Thin Film 135

KD210GH-2PU Kyocera 1,49 141,41 14,14 Polycrystalline 210

KD185GH-2PU Kyocera 1,32 139,66 13,97 Polycrystalline 185

KD135GH-2PU Kyocera 1 134,73 13,47 Polycrystalline 135

KD135SX-1PU Kyocera 1 134,73 13,47 Polycrystalline 135

KD95SX-1P Kyocera 0,69 138,01 13,8 Polycrystalline 95

KD70SX-1P Kyocera 0,51 136,32 13,63 Polycrystalline 70

KD50SE-1P Kyocera 0,53 95,19 9,52 Polycrystalline 50

KC16T Kyocera 0,14 110,53 11,05 Polycrystalline 16

KC21T Kyocera 0,19 111,76 11,18 Polycrystalline 21

KC32T Kyocera 0,26 120,89 12,09 Polycrystalline 32

BP350J BP Solar 0,45 110,98 11,1 Polycrystalline 50

BP380J BP Solar 0,65 123,22 12,32 Polycrystalline 80

BP3160N BP Solar 1,26 127,14 12,71 Polycrystalline 160

BP3170N BP Solar 1,26 135,08 13,51 Polycrystalline 170

BP4175N BP Solar 1,26 139,06 13,91 Polycrystalline 175

BP4175T BP Solar 1,25 139,58 13,96 Polycrystalline 175

SX-305M BP Solar 0,07 74,05 7,41 Polycrystalline 5

SX-310J BP Solar 0,12 86,19 8,62 Polycrystalline 10

36
SX-320U BP Solar 0,21 93,74 9,37 Polycrystalline 20

SX-330U BP Solar 0,3 100,44 10,04 Polycrystalline 30

PowerGlaz RG-
SMT5(36)M Romag 0,63 130,83 13,08 Monocrystalline 83

PowerGlaz RG-
SMT6(48)P 648185 Romag 1,31 141,21 14,12 Polycrystalline 185

PowerGlaz RG-
SMT6(48)P 648180 Romag 1,31 137,39 13,74 Polycrystalline 180

PowerGlaz RG-
SMT6(48)P 648175 Romag 1,31 133,58 13,36 Polycrystalline 175

PowerGlaz RG-
SMT6(48)P 648170 Romag 1,31 129,76 12,98 Polycrystalline 170

PowerGlaz RG-
SMT6(48)P 648165 Romag 1,31 125,95 12,59 Polycrystalline 165

PowerGlaz RG-
SMT6(48)P 648160 Romag 1,31 122,13 12,21 Polycrystalline 160

PowerGlaz RG-
SMT6(54)P 654210 Romag 1,47 142,56 14,26 Polycrystalline 210

PowerGlaz RG-
SMT6(54)P 654200 Romag 1,47 135,77 13,58 Polycrystalline 200

PowerGlaz RG-
SMT6(54)P 654195 Romag 1,47 132,37 13,24 Polycrystalline 195

PowerGlaz RG-
SMT6(54)P 654190 Romag 1,47 128,98 12,9 Polycrystalline 190

PowerGlaz RG-
SMT6(60)P 660235 Romag 1,63 144,16 14,42 Polycrystalline 235

PowerGlaz RG-
SMT6(60)P 660230 Romag 1,63 141,09 14,11 Polycrystalline 230

PowerGlaz RG-
SMT6(60)P 660225 Romag 1,63 138,02 13,8 Polycrystalline 225

PowerGlaz RG-
SMT6(60)P 660220 Romag 1,63 134,96 13,5 Polycrystalline 220

PowerGlaz RG-
SMT6(60)P 660210 Romag 1,63 128,82 12,88 Polycrystalline 210

PowerGlaz RG-
SMT6(60)P 660205 Romag 1,63 125,75 12,58 Polycrystalline 205

HIT-240HDE4 Sanyo 1,39 173,13 17,31 HIT 240

HIT-235HDE4 Sanyo 1,39 169,53 16,95 HIT 235

HIT-205DNKHE1 Sanyo 1,41 145,9 14,59 HIT 205

HIT-200DNKHE1 Sanyo 1,41 142,34 14,23 HIT 200

HIT-N220E01 Sanyo 1,26 174,49 17,45 HIT 220

HIT-N215E01 Sanyo 1,26 170,52 17,05 HIT 215

Imerys Roof Tile Imerys 0,65 84,09 8,41 Polycrystalline 55

SCHOTT POLY
225 Schott 1,67 134,47 13,45 Polycrystalline 225

SCHOTT POLY
220 Schott 1,67 131,48 13,15 Polycrystalline 220

SCHOTT POLY
217 Schott 1,67 129,69 12,97 Polycrystalline 217

37
SCHOTT POLY
210 Schott 1,67 125,51 12,55 Polycrystalline 210

PV-TD175MF5 Mitsubishi 1,38 126,56 12,66 Polycrystalline 175

PV-TD180MF5 Mitsubishi 1,38 130,17 13,02 Polycrystalline 180

PV-TD185MF5 Mitsubishi 1,38 133,79 13,38 Polycrystalline 185

PV-TD190MF5 Mitsubishi 1,38 137,41 13,74 Polycrystalline 190

PV-TE130MF5N Mitsubishi 1,01 129,02 12,9 Polycrystalline 130

PV-TE125MF5N Mitsubishi 1,01 124,05 12,41 Polycrystalline 125

PV-TE120MF5N Mitsubishi 1,01 119,09 11,91 Polycrystalline 120

PV-TE115MF5N Mitsubishi 1,01 114,13 11,41 Polycrystalline 115

PV-MF185TD4 Mitsubishi 1,38 133,79 13,38 Polycrystalline 185

PV-MF180TD4 Mitsubishi 1,38 130,17 13,02 Polycrystalline 180

PV-MF175TD4 Mitsubishi 1,38 126,56 12,66 Polycrystalline 175

PV-MF170TD4 Mitsubishi 1,38 122,94 12,29 Polycrystalline 170

PV-MF130TE4N Mitsubishi 1,01 129,02 12,9 Polycrystalline 130

PV-MF125TE4N Mitsubishi 1,01 124,05 12,41 Polycrystalline 125

PV-MF120TE4N Mitsubishi 1,01 119,09 11,91 Polycrystalline 120

PV-MF115TE4N Mitsubishi 1,01 114,13 11,41 Polycrystalline 115

PV-AD190MF5 Mitsubishi 1,38 137,41 13,74 Polycrystalline 190

PV-AD185MF5 Mitsubishi 1,38 133,79 13,38 Polycrystalline 185

PV-AD180MF5 Mitsubishi 1,38 130,17 13,02 Polycrystalline 180

PV-AD175MF5 Mitsubishi 1,38 126,56 12,66 Polycrystalline 175

PV-AE130MF5N Mitsubishi 1,01 129,02 12,9 Polycrystalline 130

PV-AE125MF5N Mitsubishi 1,01 124,05 12,41 Polycrystalline 125

PV-AE120MF5N Mitsubishi 1,01 119,09 11,91 Polycrystalline 120

PV-AE115MF5N Mitsubishi 1,01 114,13 11,41 Polycrystalline 115

PV-MF170EB4 Mitsubishi 1,26 134,49 13,45 Polycrystalline 170

PV-MF165EB4 Mitsubishi 1,26 130,54 13,05 Polycrystalline 165

PV-MF125EA4 Mitsubishi 1 124,73 12,47 Polycrystalline 125

PV-MF120EC4 Mitsubishi 0,92 130,36 13,04 Polycrystalline 120

PV-MF110EC4 Mitsubishi 0,92 119,49 11,95 Polycrystalline 110

PV-MF160EB4 Mitsubishi 1,26 126,58 12,66 Polycrystalline 160

PV-MF130EA4 Mitsubishi 1 129,72 12,97 Polycrystalline 130

YL 165 P-23b Yingli 1,3 127,23 12,72 Polycrystalline 165

YL 170 P-23b Yingli 1,3 131,08 13,11 Polycrystalline 170

YL 175 P-23b Yingli 1,3 134,94 13,49 Polycrystalline 175

YL 180 P-23b Yingli 1,3 138,79 13,88 Polycrystalline 180

YL 185 P-23b Yingli 1,3 142,65 14,26 Polycrystalline 185

YL 260 P-35b Yingli 1,95 133,31 13,33 Polycrystalline 260

YL 265 P-35b Yingli 1,95 135,88 13,59 Polycrystalline 265

YL 270 P-35b Yingli 1,95 138,44 13,84 Polycrystalline 270

38
YL 275 P-35b Yingli 1,95 141 14,1 Polycrystalline 275

YL 280 P-35b Yingli 1,95 143,57 14,36 Polycrystalline 280

YL 235 P-32b Yingli 1,79 131,15 13,11 Polycrystalline 235

YL 240 P-32b Yingli 1,79 133,94 13,39 Polycrystalline 240

YL 245 P-32b Yingli 1,79 136,73 13,67 Polycrystalline 245

YL 250 P-32b Yingli 1,79 139,52 13,95 Polycrystalline 250

YL 255 P-32b Yingli 1,79 142,31 14,23 Polycrystalline 255

YL 210 P-29b Yingli 1,63 128,56 12,86 Polycrystalline 210

YL 215 P-29b Yingli 1,63 131,62 13,16 Polycrystalline 215

YL 220 P-29b Yingli 1,63 134,68 13,47 Polycrystalline 220

YL 225 P-29b Yingli 1,63 137,74 13,77 Polycrystalline 225

YL 230 P-29b Yingli 1,63 140,8 14,08 Polycrystalline 230

YL 235 P-29b Yingli 1,63 143,86 14,39 Polycrystalline 235

YL 190 P-26b Yingli 1,48 128,37 12,84 Polycrystalline 190

YL 195 P-26b Yingli 1,48 131,75 13,18 Polycrystalline 195

YL 200 P-26b Yingli 1,48 135,13 13,51 Polycrystalline 200

YL 205 P-26b Yingli 1,48 138,51 13,85 Polycrystalline 205

YL 210 P-26b Yingli 1,48 141,89 14,19 Polycrystalline 210

YL175Wp Yingli 1,3 134,94 13,49 Polycrystalline 175

YL180Wp Yingli 1,3 138,79 13,88 Polycrystalline 180

YL210Wp Yingli 1,63 128,56 12,86 Polycrystalline 210

YL230Wp Yingli 1,63 140,8 14,08 Polycrystalline 230

YL260P-35b Yingli 1,95 133,31 13,33 Polycrystalline 260

YL265P-35b Yingli 1,95 135,88 13,59 Polycrystalline 265

YL270-35b Yingli 1,95 138,44 13,84 Polycrystalline 270

YL275-35b Yingli 1,95 141 14,1 Polycrystalline 275

YL280-35b Yingli 1,95 143,57 14,36 Polycrystalline 280

315 Solar Panel SunPower 1,63 193,17 19,32 Monocrystalline 315

230 Solar Panel SunPower 1,24 184,88 18,49 Monocrystalline 230

225 Solar Panel SunPower 1,24 180,86 18,09 Monocrystalline 225

215 Solar Panel SunPower 1,24 172,82 17,28 Monocrystalline 215

210 Solar Panel SunPower 1,24 168,8 16,88 Monocrystalline 210

STP210-18/Ud Suntech 1,47 142,84 14,28 Polycrystalline 210

STP205-18/Ud Suntech 1,47 146,24 14,62 Polycrystalline 215

STP225-20-Wd Suntech 1,65 136,36 13,64 Polycrystalline 225

STP220-20-Wd Suntech 1,65 133,33 13,33 Polycrystalline 220

STP195S-24-Ad+ Suntech 1,28 152,74 15,27 Polycrystalline 195

STP190S-24-Ad+ Suntech 1,28 148,83 14,88 Polycrystalline 190

STP090Ts-AA Suntech 1,43 62,94 6,29 Thinfilm 90

STP086Ts-AA Suntech 1,43 60,14 6,01 Thinfilm 86

39
STP082s-AA Suntech 1,43 57,34 5,73 Thinfilm 82

STP380Ts-DA Suntech 5,72 66,43 6,64 Thinfilm 380

STP370Ts-DA Suntech 5,72 64,69 6,47 Thinfilm 370

STP360Ts-DA Suntech 5,72 62,94 6,29 Thinfilm 360

STP350Ts-DA Suntech 5,72 61,19 6,12 Thinfilm 350

STP185Ts-BA Suntech 2,86 64,69 6,47 Thinfilm 185

STP180s-BA Suntech 2,86 62,94 6,29 Thinfilm 180

STP175s-BA Suntech 2,86 61,19 6,12 Thinfilm 175

STP170s-BA Suntech 2,86 59,44 5,94 Thinfilm 170

STP090Ts-AC Suntech 1,45 62 6,2 Thinfilm 90

STP086Ts-AC Suntech 1,45 59,24 5,92 Thinfilm 86

STP082Ts-AC Suntech 1,45 56,49 5,65 Thinfilm 82

STP185Ts-CA Suntech 2,86 64,69 6,47 Thinfilm 185

STP180Ts-CA Suntech 2,86 62,94 6,29 Thinfilm 180

STP175Ts-CA Suntech 2,86 61,19 6,12 Thinfilm 175

STP170Ts-CA Suntech 2,86 59,44 5,94 Thinfilm 170

STP185S-24-Ad Suntech 1,28 144,91 14,49 Monocrystalline 185

STP180S-24-Ad Suntech 1,28 141 14,1 Monocrystalline 180

STP280-24-Vd Suntech 1,94 144,3 14,43 Monocrystalline 280

STP275-24-Vd Suntech 1,94 141,73 14,17 Monocrystalline 275

C21e SolarCentury 0,51 101,48 10,15 Monocrystalline 52

40

View publication stats

You might also like