cell

A solar cell, or photovoltaic cell is an electrical device that converts the energy of light directly into electricity by the photovoltaic effect, which is a physical and chemical phenomenon.[2] It is a form of photoelectric cell, defined as a device whose electrical characteristics, such as current, voltage, or resistance, vary when exposed to light. Solar cells are the building blocks of photovoltaic modules, otherwise known as solar panels.

Solar cells are described as being photovoltaic irrespective of whether the source is sunlight or an artificial light. They are used as a photodetector (for example infrared detectors), detecting light or otherelectromagnetic radiation near the visible range, or measuring light intensity.

The operation of a photovoltaic (PV) cell requires 3 basic attributes:

  • The absorption of light, generating either electronhole pairs or excitons.
  • The separation of charge carriers of opposite types.
  • The separate extraction of those carriers to an external circuit.

In contrast, a solar thermal collector supplies heat by absorbing sunlight, for the purpose of either direct heating or indirect electrical power generation from heat. A “photoelectrolytic cell” (photoelectrochemical cell), on the other hand, refers either to a type of photovoltaic cell (like that developed by Edmond Becquerel and modern dye-sensitized solar cells), or to a device that splits water directly into hydrogen and oxygen using only solar illumination.

Applications[edit]

From a solar cell to a PV system. Diagram of the possible components of a photovoltaic system

Assemblies of solar cells are used to make solar modules which generate electrical power from sunlight, as distinguished from a “solar thermal module” or “solar hot water panel”. A solar array generates solar power using solar energy.

Cells, modules, panels and systems[edit]

Main article: Photovoltaic system

Multiple solar cells in an integrated group, all oriented in one plane, constitute a solar photovoltaic panel or solar photovoltaic module. Photovoltaic modules often have a sheet of glass on the sun-facing side, allowing light to pass while protecting the semiconductor wafers. Solar cells are usually connected in series and parallel circuits or series in modules, creating an additive voltage. Connecting cells in parallel yields a higher current; however, problems such as shadow effects can shut down the weaker (less illuminated) parallel string (a number of series connected cells) causing substantial power loss and possible damage because of the reverse bias applied to the shadowed cells by their illuminated partners. Strings of series cells are usually handled independently and not connected in parallel, though as of 2014, individual power boxes are often supplied for each module, and are connected in parallel. Although modules can be interconnected to create an array with the desired peak DC voltage and loading current capacity, using independent MPPTs (maximum power point trackers) is preferable. Otherwise, shunt diodes can reduce shadowing power loss in arrays with series/parallel connected cells.[citation needed]

Typical PV system prices in 2013 in selected countries (USD)
USD/W Australia China France Germany Italy Japan United Kingdom United States
 Residential 1.8 1.5 4.1 2.4 2.8 4.2 2.8 4.9
 Commercial 1.7 1.4 2.7 1.8 1.9 3.6 2.4 4.5
 Utility-scale 2.0 1.4 2.2 1.4 1.5 2.9 1.9 3.3
Source: IEA – Technology Roadmap: Solar Photovoltaic Energy report, 2014 edition[3]:15
Note: DOE – Photovoltaic System Pricing Trends reports lower prices for the U.S.[4]

History[edit]

The photovoltaic effect was experimentally demonstrated first by French physicist Edmond Becquerel. In 1839, at age 19, he built the world’s first photovoltaic cell in his father’s laboratory. Willoughby Smith first described the “Effect of Light on Selenium during the passage of an Electric Current” in a 20 February 1873 issue of Nature. In 1883 Charles Fritts built the first solid state photovoltaic cell by coating the semiconductor selenium with a thin layer of gold to form the junctions; the device was only around 1% efficient.

In 1888 Russian physicist Aleksandr Stoletov built the first cell based on the outer photoelectric effect discovered by Heinrich Hertz in 1887.[5]

In 1905 Albert Einstein proposed a new quantum theory of light and explained the photoelectric effect in a landmark paper, for which he received the Nobel Prize in Physics in 1921.[6]

Vadim Lashkaryov discovered pn-junctions in Cu{displaystyle _{2}}_{2}O and silver sulphide protocells in 1941.[7]

Russell Ohl patented the modern junction semiconductor solar cell in 1946[8] while working on the series of advances that would lead to the transistor.

The first practical photovoltaic cell was publicly demonstrated on 25 April 1954 at Bell Laboratories.[9] The inventors were Daryl Chapin, Calvin Souther Fuller and Gerald Pearson.[10]

Solar cells gained prominence with their incorporation onto the 1958 Vanguard I satellite.

Improvements were gradual over the next two decades. However, this success was also the reason that costs remained high, because space users were willing to pay for the best possible cells, leaving no reason to invest in lower-cost, less-efficient solutions. The price was determined largely by the semiconductor industry; their move to integrated circuits in the 1960s led to the availability of larger boules at lower relative prices. As their price fell, the price of the resulting cells did as well. These effects lowered 1971 cell costs to some $100 per watt.[11]

Space Applications[edit]

Solar cells were first used in a prominent application when they were proposed and flown on the Vanguard satellite in 1958, as an alternative power source to the primary battery power source. By adding cells to the outside of the body, the mission time could be extended with no major changes to the spacecraft or its power systems. In 1959 the United States launched Explorer 6, featuring large wing-shaped solar arrays, which became a common feature in satellites. These arrays consisted of 9600 Hoffman solar cells.

By the 1960s, solar cells were (and still are) the main power source for most Earth orbiting satellites and a number of probes into the solar system, since they offered the best power-to-weight ratio. However, this success was possible because in the space application, power system costs could be high, because space users had few other power options, and were willing to pay for the best possible cells. The space power market drove the development of higher efficiencies in solar cells up until the National Science Foundation “Research Applied to National Needs” program began to push development of solar cells for terrestrial applications.

In the early 1990s the technology used for space solar cells diverged from the silicon technology used for terrestrial panels, with the spacecraft application shifting to gallium arsenide-based III-V semiconductor materials, which then evolved into the modern III-Vmultijunction photovoltaic cell used on spacecraft.

Price reductions[edit]

solar cell

Dr. Elliot Berman testing various solar arrays manufactured by his company, Solar Power Corporation.

In late 1969 Elliot Berman joined Exxon‘s task force which was looking for projects 30 years in the future and in April 1973 he founded Solar Power Corporation, a wholly owned subsidiary of Exxon at that time.[12][13][14] The group had concluded that electrical power would be much more expensive by 2000, and felt that this increase in price would make alternative energy sources more attractive. He conducted a market study and concluded that aprice per watt of about $20/watt would create significant demand.[12] The team eliminated the steps of polishing the wafers and coating them with an anti-reflective layer, relying on the rough-sawn wafer surface. The team also replaced the expensive materials and hand wiring used in space applications with a printed circuit board on the back, acrylic plastic on the front, and silicone glue between the two, “potting” the cells.[15] Solar cells could be made using cast-off material from the electronics market. By 1973 they announced a product, and SPC convinced Tideland Signal to use its panels to power navigational buoys, initially for the U.S. Coast Guard.[13]

Research into solar power for terrestrial applications became prominent with the U.S. National Science Foundation’s Advanced Solar Energy Research and Development Division within the “Research Applied to National Needs” program, which ran from 1969 to 1977,[16] and funded research on developing solar power for ground electrical power systems. A 1973 conference, the “Cherry Hill Conference”, set forth the technology goals required to achieve this goal and outlined an ambitious project for achieving them, kicking off an applied research program that would be ongoing for several decades.[17] The program was eventually taken over by the Energy Research and Development Administration (ERDA),[18] which was later merged into the U.S. Department of Energy.

Following the 1973 oil crisis oil companies used their higher profits to start (or buy) solar firms, and were for decades the largest producers. Exxon, ARCO, Shell, Amoco (later purchased by BP) and Mobil all had major solar divisions during the 1970s and 1980s. Technology companies also participated, including General Electric, Motorola, IBM, Tyco and RCA.[19]

Declining costs and exponential growth[edit]

solar cell
Price per watt history for conventional (c-Si) solar cells since 1977
Swanson’s law – the learning curve ofsolar PV
Growth of photovoltaics – Worldwide total installed PV capacity

Adjusting for inflation, it cost $96 per watt for a solar module in the mid-1970s. Process improvements and a very large boost in production have brought that figure down 99 percent, to 68¢ per watt in 2016, according to data from Bloomberg New Energy Finance.[20] Swanson’s law is an observation similar to Moore’s Law that states that solar cell prices fall 20% for every doubling of industry capacity. It was featured in an article in the British weekly newspaper The Economist.[21]

Further improvements reduced production cost to under $1 per watt, with wholesale costs well under $2. Balance of system costs were then higher than the panels. Large commercial arrays could be built, as of 2010, at below $3.40 a watt, fully commissioned.[22][23]

As the semiconductor industry moved to ever-larger boules, older equipment became inexpensive. Cell sizes grew as equipment became available on the surplus market; ARCO Solar’s original panels used cells 2 to 4 inches (50 to 100 mm) in diameter. Panels in the 1990s and early 2000s generally used 125 mm wafers; since 2008 almost all new panels use 150 mm cells. The widespread introduction of flat screen televisions in the late 1990s and early 2000s led to the wide availability of large, high-quality glass sheets to cover the panels.

During the 1990s, polysilicon (“poly”) cells became increasingly popular. These cells offer less efficiency than their monosilicon (“mono”) counterparts, but they are grown in large vats that reduce cost. By the mid-2000s, poly was dominant in the low-cost panel market, but more recently the mono returned to widespread use.

Manufacturers of wafer-based cells responded to high silicon prices in 2004–2008 with rapid reductions in silicon consumption. In 2008, according to Jef Poortmans, director of IMEC‘s organic and solar department, current cells use 8–9 grams (0.28–0.32 oz) of silicon per watt of power generation, with wafer thicknesses in the neighborhood of 200 microns.

First Solar is the largest thin film manufacturer in the world, using a CdTe-cell sandwiched between two layers of glass. Crystalline silicon panels dominate worldwide markets and are mostly manufactured in China and Taiwan. By late 2011, a drop in European demand due to budgetary turmoil dropped prices for crystalline solar modules to about $1.09[23] per watt down sharply from 2010. Prices continued to fall in 2012, reaching $0.62/watt by 4Q2012.[24]

Global installed PV capacity reached at least 177 gigawatts in 2014, enough to supply 1 percent of the world’s total electricity consumption. Solar PV is growing fastest in Asia, with China and Japan currently accounting for half of worldwide deployment.[25]

Subsidies and grid parity

Solar-specific feed-in tariffs vary by country and within countries. Such tariffs encourage the development of solar power projects. Widespread grid parity, the point at which photovoltaic electricity is equal to or cheaper thangrid power without subsidies, likely requires advances on all three fronts. Proponents of solar hope to achieve grid parity first in areas with abundant sun and high electricity costs such as in California and Japan.[26] In 2007 BP claimed grid parity for Hawaii and other islands that otherwise use diesel fuel to produce electricity. George W. Bush set 2015 as the date for grid parity in the US.[27][28] The Photovoltaic Association reported in 2012 that Australia had reached grid parity (ignoring feed in tariffs).[29]

The price of solar panels fell steadily for 40 years, interrupted in 2004 when high subsidies in Germany drastically increased demand there and greatly increased the price of purified silicon (which is used in computer chips as well as solar panels). The recession of 2008 and the onset of Chinese manufacturing caused prices to resume their decline. In the four years after January 2008 prices for solar modules in Germany dropped from €3 to €1 per peak watt. During that same time production capacity surged with an annual growth of more than 50%. China increased market share from 8% in 2008 to over 55% in the last quarter of 2010.[30] In December 2012 the price of Chinese solar panels had dropped to $0.60/Wp (crystalline modules).[31]

Theory[edit]

Working mechanism of a solar cell

Main article: Theory of solar cells

The solar cell works in several steps:

  • Photons in sunlight hit the solar panel and are absorbed by semiconducting materials, such as silicon.
  • Electrons are excited from their current molecular/atomic orbital. Once excited an electron can either dissipate the energy as heat and return to its orbital or travel through the cell until it reaches an electrode. Current flows through the material to cancel the potential and this electricity is captured. The chemical bonds of the material are vital for this process to work, and usually silicon is used in two layers, one layer being bonded with boron, the other phosphorus. These layers have different chemical electric charges and subsequently both drive and direct the current of electrons.[2]
  • An array of solar cells converts solar energy into a usable amount of direct current (DC) electricity.
  • An inverter can convert the power to alternating current (AC).

The most commonly known solar cell is configured as a large-area p–n junction made from silicon.

Efficiency[edit]

The Shockley-Queisser limit for the theoretical maximum efficiency of a solar cell. Semiconductors with band gap between 1 and 1.5eV, or near-infrared light, have the greatest potential to form an efficient single-junction cell. (The efficiency “limit” shown here can be exceeded bymultijunction solar cells.)

Main article: Solar cell efficiency

Solar cell efficiency may be broken down into reflectance efficiency, thermodynamic efficiency, charge carrier separation efficiency and conductive efficiency. The overall efficiency is the product of these individual metrics.

A solar cell has a voltage dependent efficiency curve, temperature coefficients, and allowable shadow angles.

Due to the difficulty in measuring these parameters directly, other parameters are substituted: thermodynamic efficiency, quantum efficiency, integrated quantum efficiency, VOC ratio, and fill factor. Reflectance losses are a portion of quantum efficiency under “external quantum efficiency”. Recombination losses make up another portion of quantum efficiency, VOC ratio, and fill factor. Resistive losses are predominantly categorized under fill factor, but also make up minor portions of quantum efficiency, VOC ratio.

The fill factor is the ratio of the actual maximum obtainable power to the product of the open circuit voltage and short circuit current. This is a key parameter in evaluating performance. In 2009, typical commercial solar cells had a fill factor > 0.70. Grade B cells were usually between 0.4 and 0.7.[32] Cells with a high fill factor have a low equivalent series resistance and a high equivalent shunt resistance, so less of the current produced by the cell is dissipated in internal losses.

Single p–n junction crystalline silicon devices are now approaching the theoretical limiting power efficiency of 33.7%, noted as the Shockley–Queisser limit in 1961. In the extreme, with an infinite number of layers, the corresponding limit is 86% using concentrated sunlight.[33]

In December 2014, a solar cell achieved a new laboratory record with 46 percent efficiency in a French-German collaboration.[34]

In 2014, three companies broke the record of 25.6% for a silicon solar cell. Panasonic’s was the most efficient. The company moved the front contacts to the rear of the panel, eliminating shaded areas. In addition they applied thin silicon films to the (high quality silicon) wafer’s front and back to eliminate defects at or near the wafer surface.[35]

In September 2015, the Fraunhofer Institute for Solar Energy Systems (Fraunhofer ISE) announced the achievement of an efficiency above 20% for epitaxial wafer cells. The work on optimizing the atmospheric-pressure chemical vapor deposition (APCVD) in-line production chain was done in collaboration with NexWafe GmbH, a company spun off from Fraunhofer ISE to commercialize production.[36]

For triple-junction thin-film solar cells, the world record is 13.6%, set in June 2015.[37]

solar cell

Reported timeline of solar cell energy conversion efficiencies (National Renewable Energy Laboratory)

Materials[edit]

Global market-share in terms of annual production by PV technology since 1990

Solar cells are typically named after the semiconducting material they are made of. These materials must have certain characteristics in order to absorb sunlight. Some cells are designed to handle sunlight that reaches the Earth’s surface, while others are optimized for use in space. Solar cells can be made of only one single layer of light-absorbing material (single-junction) or use multiple physical configurations (multi-junctions) to take advantage of various absorption and charge separation mechanisms.

Solar cells can be classified into first, second and third generation cells. The first generation cells—also called conventional, traditional or wafer-based cells—are made of crystalline silicon, the commercially predominant PV technology, that includes materials such as polysilicon and monocrystalline silicon. Second generation cells are thin film solar cells, that include amorphous silicon, CdTe and CIGS cells and are commercially significant in utility-scale photovoltaic power stations, building integrated photovoltaics or in small stand-alone power system. The third generation of solar cells includes a number of thin-film technologies often described as emerging photovoltaics—most of them have not yet been commercially applied and are still in the research or development phase. Many use organic materials, often organometallic compounds as well as inorganic substances. Despite the fact that their efficiencies had been low and the stability of the absorber material was often too short for commercial applications, there is a lot of research invested into these technologies as they promise to achieve the goal of producing low-cost, high-efficiency solar cells.

Crystalline silicon[edit]

Main article: Crystalline silicon

By far, the most prevalent bulk material for solar cells is crystalline silicon (c-Si), also known as “solar grade silicon”. Bulk silicon is separated into multiple categories according to crystallinity and crystal size in the resulting ingot, ribbon or wafer. These cells are entirely based around the concept of a p-n junction. Solar cells made of c-Si are made from wafers between 160 and 240 micrometers thick.

Monocrystalline silicon[edit]

Monocrystalline silicon (mono-Si) solar cells are more efficient and more expensive than most other types of cells. The corners of the cells look clipped, like an octagon, because the wafer material is cut from cylindrical ingots, that are typically grown by theCzochralski process. Solar panels using mono-Si cells display a distinctive pattern of small white diamonds.

Epitaxial silicon[edit]

Epitaxial wafers can be grown on a monocrystalline silicon “seed” wafer by atmospheric-pressure CVD in a high-throughput inline process, and then detached as self-supporting wafers of some standard thickness (e.g., 250 µm) that can be manipulated by hand, and directly substituted for wafer cells cut from monocrystalline silicon ingots. Solar cells made with this technique can have efficiencies approaching those of wafer-cut cells, but at appreciably lower cost.[38]

Polycrystalline silicon[edit]

Polycrystalline silicon, or multicrystalline silicon (multi-Si) cells are made from cast square ingots—large blocks of molten silicon carefully cooled and solidified. They consist of small crystals giving the material its typical metal flake effect. Polysilicon cells are the most common type used in photovoltaics and are less expensive, but also less efficient, than those made from monocrystalline silicon.

Ribbon silicon[edit]

Ribbon silicon is a type of polycrystalline silicon—it is formed by drawing flat thin films from molten silicon and results in a polycrystalline structure. These cells are cheaper to make than multi-Si, due to a great reduction in silicon waste, as this approach does not require sawing from ingots.[39] However, they are also less efficient.

Mono-like-multi silicon (MLM)[edit]

This form was developed in the 2000s and introduced commercially around 2009. Also called cast-mono, this design uses polycrystalline casting chambers with small “seeds” of mono material. The result is a bulk mono-like material that is polycrystalline around the outsides. When sliced for processing, the inner sections are high-efficiency mono-like cells (but square instead of “clipped”), while the outer edges are sold as conventional poly. This production method results in mono-like cells at poly-like prices.[40]

Thin film[edit]

Main article: Thin-film solar cell

Thin-film technologies reduce the amount of active material in a cell. Most designs sandwich active material between two panes of glass. Since silicon solar panels only use one pane of glass, thin film panels are approximately twice as heavy as crystalline silicon panels, although they have a smaller ecological impact (determined from life cycle analysis).[41] The majority of film panels have 2–3 percentage points lower conversion efficiencies than crystalline silicon.[42] Cadmium telluride (CdTe), copper indium gallium selenide(CIGS) and amorphous silicon (a-Si) are three thin-film technologies often used for outdoor applications. As of December 2013, CdTe cost per installed watt was $0.59 as reported by First Solar. CIGS technology laboratory demonstrations reached 20.4% conversion efficiency as of December 2013. The lab efficiency of GaAs thin film technology topped 28%.[citation needed] The quantum efficiency of thin film solar cells is also lower due to reduced number of collected charge carriers per incident photon. Most recently, CZTS solar cell emerge as the less-toxic thin film solar cell technology, which achieved ~12% efficiency. Thin film solar cells are increasing due to it being silent, renewable and solar energy being the most abundant energy source on Earth.[43]

Cadmium telluride[edit]

Cadmium telluride is the only thin film material so far to rival crystalline silicon in cost/watt. However cadmium is highly toxic and tellurium (anion: “telluride”) supplies are limited. The cadmium present in the cells would be toxic if released. However, release is impossible during normal operation of the cells and is unlikely during fires in residential roofs.[44] A square meter of CdTe contains approximately the same amount of Cd as a single C cell nickel-cadmium battery, in a more stable and less soluble form.[44]

Copper indium gallium selenide[edit]

Main article: Copper indium gallium selenide solar cell
Copper indium gallium selenide (CIGS) is a direct band gap material. It has the highest efficiency (~20%) among all commercially significant thin film materials (see CIGS solar cell). Traditional methods of fabrication involve vacuum processes including co-evaporation and sputtering. Recent developments at IBM and Nanosolar attempt to lower the cost by using non-vacuum solution processes.[45]

Silicon thin film[edit]

Silicon thin-film cells are mainly deposited by chemical vapor deposition (typically plasma-enhanced, PE-CVD) from silane gas and hydrogen gas. Depending on the deposition parameters, this can yield amorphous silicon (a-Si or a-Si:H), protocrystalline silicon ornanocrystalline silicon (nc-Si or nc-Si:H), also called microcrystalline silicon.[46]
Amorphous silicon is the most well-developed thin film technology to-date. An amorphous silicon (a-Si) solar cell is made of non-crystalline or microcrystalline silicon. Amorphous silicon has a higher bandgap (1.7 eV) than crystalline silicon (c-Si) (1.1 eV), which means it absorbs the visible part of the solar spectrum more strongly than the higher power density infrared portion of the spectrum. The production of a-Si thin film solar cells uses glass as a substrate and deposits a very thin layer of silicon by plasma-enhanced chemical vapor deposition (PECVD).
Protocrystalline silicon with a low volume fraction of nanocrystalline silicon is optimal for high open circuit voltage.[47] Nc-Si has about the same bandgap as c-Si and nc-Si and a-Si can advantageously be combined in thin layers, creating a layered cell called a tandem cell. The top cell in a-Si absorbs the visible light and leaves the infrared part of the spectrum for the bottom cell in nc-Si.

Gallium arsenide thin film[edit]

The semiconductor material Gallium arsenide (GaAs) is also used for single-crystalline thin film solar cells. Although GaAs cells are very expensive, they hold the world’s record in efficiency for a single-junction solar cell at 28.8%.[48] GaAs is more commonly used in multijunction photovoltaic cells for concentrated photovoltaics (CPV, HCPV) and for solar panels on spacecrafts, as the industry favours efficiency over cost for space-based solar power.

Multijunction cells[edit]

Dawn‘s 10 kW triple-junction gallium arsenide solar array at full extension

Multi-junction cells consist of multiple thin films, each essentially a solar cell grown on top of each other, typically using metalorganic vapour phase epitaxy. Each layers has a different band gap energy to allow it to absorb electromagnetic radiation over a different portion of the spectrum. Multi-junction cells were originally developed for special applications such as satellites and space exploration, but are now used increasingly in terrestrial concentrator photovoltaics (CPV), an emerging technology that uses lenses and curved mirrors to concentrate sunlight onto small but highly efficient multi-junction solar cells. By concentrating sunlight up to a thousand times,High concentrated photovoltaics (HCPV) has the potential to outcompete conventional solar PV in the future.[49]:21,26

Tandem solar cells based on monolithic, series connected, gallium indium phosphide (GaInP), gallium arsenide (GaAs), and germanium (Ge) p–n junctions, are increasing sales, despite cost pressures.[50] Between December 2006 and December 2007, the cost of 4N gallium metal rose from about $350 per kg to $680 per kg. Additionally, germanium metal prices have risen substantially to $1000–1200 per kg this year. Those materials include gallium (4N, 6N and 7N Ga), arsenic (4N, 6N and 7N) and germanium, pyrolitic boron nitride (pBN) crucibles for growing crystals, and boron oxide, these products are critical to the entire substrate manufacturing industry.[citation needed]

A triple-junction cell, for example, may consist of the semiconductors: GaAs, Ge, and GaInP
2
.[51] Triple-junction GaAs solar cells were used as the power source of the Dutch four-time World Solar Challenge winners Nuna in 2003, 2005 and 2007 and by the Dutch solar cars Solutra (2005), Twente One (2007) and 21Revolution (2009).[citation needed] GaAs based multi-junction devices are the most efficient solar cells to date. On 15 October 2012, triple junction metamorphic cells reached a record high of 44%.[52]

Research in solar cells[edit]

Perovskite solar cells[edit]

Main article: Perovskite solar cell

Perovskite solar cells are solar cells that include a perovskite-structured material as the active layer. Most commonly, this is a solution-processed hybrid organic-inorganic tin or lead halide based material. Efficiencies have increased from below 10% at their first usage in 2009 to over 20% in 2014, making them a very rapidly advancing technology and a hot topic in the solar cell field.[53] Perovskite solar cells are also forecast to be extremely cheap to scale up, making them a very attractive option for commercialisation.

Liquid inks[edit]

In 2014, researchers at California NanoSystems Institute discovered using kesterite and perovskite improved electric power conversion efficiency for solar cells.[54]

Upconversion and Downconversion[edit]

Photon upconversion is the process of using two low-energy (e.g., infrared) photons to produce one higher energy photon; downconversion is the process of using one high energy photon (e.g.,, ultraviolet) to produce two lower energy photons. Either of these techniques could be used to produce higher efficiency solar cells by allowing solar photons to be more efficiently used. The difficulty, however, is that the conversion efficiency of existing phosphors exhibiting up- or down-conversion is low, and is typically narrow band.

One upconversion technique is to incorporate lanthanide-doped materials (Er3+
, Yb3+
, Ho3+
or a combination), taking advantage of their luminescence to convert infrared radiation to visible light. Upconversion process occurs when two infrared photons are absorbed by rare-earth ions to generate a (high-energy) absorbable photon. As example, the energy transfer upconversion process (ETU), consists in successive transfer processes between excited ions in the near infrared. The upconverter material could be placed below the solar cell to absorb the infrared light that passes through the silicon. Useful ions are most commonly found in the trivalent state. Er+
ions have been the most used. Er3+
ions absorb solar radiation around 1.54 µm. Two Er3+
ions that have absorbed this radiation can interact with each other through an upconversion process. The excited ion emits light above the Si bandgap that is absorbed by the solar cell and creates an additional electron–hole pair that can generate current. However, the increased efficiency was small. In addition, fluoroindate glasses have low phonon energy and have been proposed as suitable matrix doped with Ho3+
ions.[55]

Light-absorbing dyes[edit]

Dye-sensitized solar cells (DSSCs) are made of low-cost materials and do not need elaborate manufacturing equipment, so they can be made in a DIY fashion. In bulk it should be significantly less expensive than older solid-state cell designs. DSSC’s can be engineered into flexible sheets and although its conversion efficiency is less than the best thin film cells, its price/performance ratio may be high enough to allow them to compete with fossil fuel electrical generation.

Typically a ruthenium metalorganic dye (Ru-centered) is used as a monolayer of light-absorbing material. The dye-sensitized solar cell depends on a mesoporous layer of nanoparticulate titanium dioxide to greatly amplify the surface area (200–300 m2/g TiO
2
, as compared to approximately 10 m2/g of flat single crystal). The photogenerated electrons from the light absorbing dye are passed on to the n-type TiO
2
and the holes are absorbed by an electrolyte on the other side of the dye. The circuit is completed by a redoxcouple in the electrolyte, which can be liquid or solid. This type of cell allows more flexible use of materials and is typically manufactured by screen printing or ultrasonic nozzles, with the potential for lower processing costs than those used for bulk solar cells. However, the dyes in these cells also suffer from degradation under heat and UV light and the cell casing is difficult to seal due to the solvents used in assembly. The first commercial shipment of DSSC solar modules occurred in July 2009 from G24i Innovations.[56]

Quantum dots[edit]

Main article: Quantum dot solar cell

Quantum dot solar cells (QDSCs) are based on the Gratzel cell, or dye-sensitized solar cell architecture, but employ low band gap semiconductor nanoparticles, fabricated with crystallite sizes small enough to form quantum dots (such as CdS, CdSe, Sb
2S
3
, PbS, etc.), instead of organic or organometallic dyes as light absorbers. QD’s size quantization allows for the band gap to be tuned by simply changing particle size. They also have high extinction coefficients and have shown the possibility of multiple exciton generation.[57]

In a QDSC, a mesoporous layer of titanium dioxide nanoparticles forms the backbone of the cell, much like in a DSSC. This TiO
2
layer can then be made photoactive by coating with semiconductor quantum dots using chemical bath deposition, electrophoretic deposition or successive ionic layer adsorption and reaction. The electrical circuit is then completed through the use of a liquid or solid redox couple. The efficiency of QDSCs has increased[58] to over 5% shown for both liquid-junction[59] and solid state cells.[60] In an effort to decrease production costs, the Prashant Kamat research group[61] demonstrated a solar paint made with TiO
2
and CdSe that can be applied using a one-step method to any conductive surface with efficiencies over 1%.[62]

Organic/polymer solar cells[edit]

Main articles: Organic solar cell and Polymer solar cell

Organic solar cells and polymer solar cells are built from thin films (typically 100 nm) of organic semiconductors including polymers, such as polyphenylene vinylene and small-molecule compounds like copper phthalocyanine (a blue or green organic pigment) andcarbon fullerenes and fullerene derivatives such as PCBM.

They can be processed from liquid solution, offering the possibility of a simple roll-to-roll printing process, potentially leading to inexpensive, large-scale production. In addition, these cells could be beneficial for some applications where mechanical flexibility and disposability are important. Current cell efficiencies are, however, very low, and practical devices are essentially non-existent.

Energy conversion efficiencies achieved to date using conductive polymers are very low compared to inorganic materials. However, Konarka Power Plastic reached efficiency of 8.3%[63] and organic tandem cells in 2012 reached 11.1%.[citation needed]

The active region of an organic device consists of two materials, one electron donor and one electron acceptor. When a photon is converted into an electron hole pair, typically in the donor material, the charges tend to remain bound in the form of an exciton, separating when the exciton diffuses to the donor-acceptor interface, unlike most other solar cell types. The short exciton diffusion lengths of most polymer systems tend to limit the efficiency of such devices. Nanostructured interfaces, sometimes in the form of bulk heterojunctions, can improve performance.[64]

In 2011, MIT and Michigan State researchers developed solar cells with a power efficiency close to 2% with a transparency to the human eye greater than 65%, achieved by selectively absorbing the ultraviolet and near-infrared parts of the spectrum with small-molecule compounds.[65][66] Researchers at UCLA more recently developed an analogous polymer solar cell, following the same approach, that is 70% transparent and has a 4% power conversion efficiency.[67][68][69] These lightweight, flexible cells can be produced in bulk at a low cost and could be used to create power generating windows.

In 2013, researchers announced polymer cells with some 3% efficiency. They used block copolymers, self-assembling organic materials that arrange themselves into distinct layers. The research focused on P3HT-b-PFTBT that separates into bands some 16 nanometers wide.[70][71]

Adaptive cells[edit]

Adaptive cells change their absorption/reflection characteristics depending to respond to environmental conditions. An adaptive material responds to the intensity and angle of incident light. At the part of the cell where the light is most intense, the cell surface changes from reflective to adaptive, allowing the light to penetrate the cell. The other parts of the cell remain reflective increasing the retention of the absorbed light within the cell.[72]

In 2014 a system that combined an adaptive surface with a glass substrate that redirect the absorbed to a light absorber on the edges of the sheet. The system also included an array of fixed lenses/mirrors to concentrate light onto the adaptive surface. As the day continues, the concentrated light moves along the surface of the cell. That surface switches from reflective to adaptive when the light is most concentrated and back to reflective after the light moves along.[72]

Manufacture[edit]

Solar cells share some of the same processing and manufacturing techniques as other semiconductor devices. However, the stringent requirements for cleanliness and quality control of semiconductor fabrication are more relaxed for solar cells, lowering costs.

Polycrystalline silicon wafers are made by wire-sawing block-cast silicon ingots into 180 to 350 micrometer wafers. The wafers are usually lightly p-type-doped. A surface diffusion of n-type dopants is performed on the front side of the wafer. This forms a p–n junction a few hundred nanometers below the surface.

Anti-reflection coatings are then typically applied to increase the amount of light coupled into the solar cell. Silicon nitride has gradually replaced titanium dioxide as the preferred material, because of its excellent surface passivation qualities. It prevents carrier recombination at the cell surface. A layer several hundred nanometers thick is applied using PECVD. Some solar cells have textured front surfaces that, like anti-reflection coatings, increase the amount of light reaching the wafer. Such surfaces were first applied to single-crystal silicon, followed by multicrystalline silicon somewhat later.

A full area metal contact is made on the back surface, and a grid-like metal contact made up of fine “fingers” and larger “bus bars” are screen-printed onto the front surface using a silver paste. This is an evolution of the so-called “wet” process for applying electrodes, first described in a US patent filed in 1981 by Bayer AG.[73] The rear contact is formed by screen-printing a metal paste, typically aluminium. Usually this contact covers the entire rear, though some designs employ a grid pattern. The paste is then fired at several hundred degrees Celsius to form metal electrodes in ohmic contact with the silicon. Some companies use an additional electro-plating step to increase efficiency. After the metal contacts are made, the solar cells are interconnected by flat wires or metal ribbons, and assembled into modules or “solar panels”. Solar panels have a sheet of tempered glass on the front, and a polymer encapsulation on the back.

Manufacturers and certification[edit]

Further information: List of photovoltaics companies

Solar cell production by region[74]

National Renewable Energy Laboratory tests and validates solar technologies. Three reliable groups certify solar equipment: UL and IEEE (both U.S. standards) and IEC.

Solar cells are manufactured in volume in Japan, Germany, China, Taiwan, Malaysia and the United States, whereas Europe, China, the U.S., and Japan have dominated (94% or more as of 2013) in installed systems.[75] Other nations are acquiring significant solar cell production capacity.

Global PV cell/module production increased by 10% in 2012 despite a 9% decline in solar energy investments according to the annual “PV Status Report” released by the European Commission‘s Joint Research Centre. Between 2009 and 2013 cell production has quadrupled.[75][76][77]

China[edit]

Main article: Solar power in China

Due to heavy government investment, China has become the dominant force in solar cell manufacturing. Chinese companies produced solar cells/modules with a capacity of ~23 GW in 2013 (60% of global production).[75]

Malaysia[edit]

In 2014, Malaysia was the world’s third largest manufacturer of photovoltaics equipment, behind China and the European Union.[78]

United States[edit]

Solar cell production in the U.S. has suffered due to the global financial crisis, but recovered partly due to the falling price of quality silicon.[79][80]

See also[edit]

References[edit]

  1. Jump up^ Shockley, William; Queisser, Hans J. (1961). “Detailed Balance Limit of Efficiency of p-n Junction Solar Cells” (PDF). Journal of Applied Physics 32 (3): 510. doi:10.1063/1.1736034.
  2. ^ Jump up to:a b Solar Cells. chemistryexplained.com
  3. Jump up^ “Technology Roadmap: Solar Photovoltaic Energy” (PDF). IEA. 2014. Archived from the original on 7 October 2014. Retrieved 7 October 2014.
  4. Jump up^ “Photovoltaic System Pricing Trends – Historical, Recent, and Near-Term Projections, 2014 Edition” (PDF). NREL. 22 September 2014. p. 4. Archived from the original on 29 March 2015.
  5. Jump up^ Gevorkian, Peter (2007). Sustainable energy systems engineering: the complete green building design resource. McGraw Hill Professional. ISBN 978-0-07-147359-0.
  6. Jump up^ “The Nobel Prize in Physics 1921: Albert Einstein”, Nobel Prize official page
  7. Jump up^ Lashkaryov, V. E. (1941) Investigation of a barrier layer by the thermoprobe method, Izv. Akad. Nauk SSSR, Ser. Fiz. 5, 442–446, English translation: Ukr. J. Phys. 53, 53–56 (2008)
  8. Jump up^ “Light sensitive device” U.S. Patent 2,402,662 Issue date: June 1946
  9. Jump up^ “April 25, 1954: Bell Labs Demonstrates the First Practical Silicon Solar Cell”. APS News (American Physical Society) 18(4). April 2009.
  10. Jump up^ Tsokos, K. A. (28 January 2010). Physics for the IB Diploma Full Colour. Cambridge University Press. ISBN 978-0-521-13821-5.
  11. Jump up^ Perlin 1999, p. 50.
  12. ^ Jump up to:a b Perlin 1999, p. 53.
  13. ^ Jump up to:a b Williams, Neville (2005). Chasing the Sun: Solar Adventures Around the World. New Society Publishers. p. 84.ISBN 9781550923124.
  14. Jump up^ Jones, Geoffrey; Bouamane, Loubna (2012). “Power from Sunshine”: A Business History of Solar Energy (PDF). Harvard Business School. pp. 22–23.
  15. Jump up^ Perlin 1999, p. 54.
  16. Jump up^ The National Science Foundation: A Brief History, Chapter IV, NSF 88-16, July 15, 1994 (retrieved 20 June 2015)
  17. Jump up^ Herwig, Lloyd O. (1999). “Cherry Hill revisited: Background events and photovoltaic technology status”. AIP Conference Proceedings. National center for photovoltaics (NCPV) 15th program review meeting. AIP Conference Proceedings 462. p. 785. Bibcode:1999AIPC..462..785H.doi:10.1063/1.58015.
  18. Jump up^ Deyo, J. N., Brandhorst, H. W., Jr., and Forestieri, A. F.,Status of the ERDA/NASA photovoltaic tests and applications project, 12th IEEE Photovoltaic Specialists Conf., 15–18 Nov. 1976
  19. Jump up^ Reed Business Information (18 October 1979). The multinational connections-who does what where. Reed Business Information. ISSN 0262-4079.
  20. Jump up^ http://www.bloomberg.com/features/2016-solar-power-buffett-vs-musk/
  21. Jump up^ “Sunny Uplands: Alternative energy will no longer be alternative”. The Economist. 21 November 2012. Retrieved28 December 2012.
  22. Jump up^ $1/W Photovoltaic Systems DOE whitepaper August 2010
  23. ^ Jump up to:a b Solar Stocks: Does the Punishment Fit the Crime?. 24/7 Wall St. (6 October 2011). Retrieved on 3 January 2012.
  24. Jump up^ Parkinson, Giles. “Plunging Cost Of Solar PV (Graphs)”. Clean Technica. Retrieved 18 May 2013.
  25. Jump up^ “Snapshot of Global PV 1992–2014” (PDF). International Energy Agency — Photovoltaic Power Systems Programme. 30 March 2015. Archived from the original on 30 March 2015.
  26. Jump up^ BP Global – Reports and publications – Going for grid parityat the Wayback Machine (archived June 8, 2011). Bp.com. Retrieved on 19 January 2011.
  27. Jump up^ BP Global – Reports and publications – Gaining on the grid. Bp.com. August 2007.
  28. Jump up^ The Path to Grid Parity. bp.com
  29. Jump up^ Peacock, Matt (20 June 2012) Solar industry celebrates grid parity, ABC News.
  30. Jump up^ Baldwin, Sam (20 April 2011) Energy Efficiency & Renewable Energy: Challenges and Opportunities. Clean Energy SuperCluster Expo Colorado State University. U.S. Department of Energy.
  31. Jump up^ ENF Ltd. (8 January 2013). “Small Chinese Solar Manufacturers Decimated in 2012 | Solar PV Business News | ENF Company Directory”. Enfsolar.com. Retrieved 1 June2013.
  32. Jump up^ “T.Bazouni: What is the Fill Factor of a Solar Panel”. Retrieved 17 February 2009.
  33. Jump up^ Vos, A. D. (1980). “Detailed balance limit of the efficiency of tandem solar cells”. Journal of Physics D: Applied Physics 13 (5): 839. doi:10.1088/0022-3727/13/5/018.
  34. Jump up^ “French-German collaborators claim solar cell efficiency world record”. EE Times Europe. 2 December 2014. Retrieved3 December 2014.
  35. Jump up^ Bullis, Kevin (June 13, 2014) Record-Breaking Solar Cell Points the Way to Cheaper Power. MIT Technology Review
  36. Jump up^ Janz, Stefan; Reber, Stefan (14 September 2015). “20% Efficient Solar Cell on EpiWafer”. Fraunhofer ISE. RetrievedOctober 15, 2015.
  37. Jump up^ Zyg, Lisa (June 4, 2015). “Solar cell sets world record with a stabilized efficiency of 13.6%”. Phys.org.
  38. Jump up^ Rachow, Thomas; Heinz, Friedemann; Steinhauser, Bernd; Janz, Stefan; Reber, Stefan (2014). “Epitaxial n- and p-type Emitters for High Efficiency Solar Cell Concepts”. Journal of Energy and Power Engineering 8 (David Publishing). pp. 1371–1377. Retrieved October 15, 2015.
  39. Jump up^ Kim, D.S.; et al. (18 May 2003). “String ribbon silicon solar cells with 17.8% efficiency” (PDF). Proceedings of 3rd World Conference on Photovoltaic Energy Conversion, 2003 2: 1293–1296. ISBN 4-9901816-0-3.
  40. Jump up^ Wayne McMillan, “The Cast Mono Dilemma”, BT Imaging
  41. Jump up^ Pearce, J.; Lau, A. (2002). “Net Energy Analysis for Sustainable Energy Production from Silicon Based Solar Cells”.Solar Energy (PDF). p. 181. doi:10.1115/SED2002-1051.ISBN 0-7918-1689-3.
  42. Jump up^ Datasheets of the market leaders: First Solar for thin film,Suntech and SunPower for crystalline silicon
  43. Jump up^ Edoff, Marika (March 2012). “Thin Film Solar Cells: Research in an Industrial Perspective”. AMBIO 41 (2): 112–118.doi:10.1007/s13280-012-0265-6. ISSN 0044-7447.PMC 3357764. PMID 22434436.
  44. ^ Jump up to:a b Fthenakis, Vasilis M. (2004). “Life cycle impact analysis of cadmium in CdTe PV production” (PDF). Renewable and Sustainable Energy Reviews 8 (4): 303–334.doi:10.1016/j.rser.2003.12.001.
  45. Jump up^ “IBM and Tokyo Ohka Kogyo Turn Up Watts on Solar Energy Production”, IBM
  46. Jump up^ Collins, R. W.; Ferlauto, A. S.; Ferreira, G. M.; Chen, C.; Koh, J.; Koval, R. J.; Lee, Y.; Pearce, J. M.; Wronski, C. R. (2003). “Evolution of microstructure and phase in amorphous, protocrystalline, and microcrystalline silicon studied by real time spectroscopic ellipsometry”. Solar Energy Materials and Solar Cells 78: 143. doi:10.1016/S0927-0248(02)00436-1.
  47. Jump up^ Pearce, J. M.; Podraza, N.; Collins, R. W.; Al-Jassim, M. M.; Jones, K. M.; Deng, J.; Wronski, C. R. (2007). “Optimization of open circuit voltage in amorphous silicon solar cells with mixed-phase (amorphous+nanocrystalline) p-type contacts of low nanocrystalline content” (PDF). Journal of Applied Physics 101(11): 114301. doi:10.1063/1.2714507.
  48. Jump up^ Yablonovitch, Eli; Miller, Owen D.; Kurtz, S. R. (2012). “The opto-electronic physics that broke the efficiency limit in solar cells”. 2012 38th IEEE Photovoltaic Specialists Conference. p. 001556. doi:10.1109/PVSC.2012.6317891. ISBN 978-1-4673-0066-7.
  49. Jump up^ “Photovoltaics Report” (PDF). Fraunhofer ISE. 28 July 2014.Archived from the original on 31 August 2014. Retrieved31 August 2014.
  50. Jump up^ Oku, Takeo; Kumada, Kazuma; Suzuki, Atsushi; Kikuchi, Kenji (June 2012). “Effects of germanium addition to copper phthalocyanine/fullerene-based solar cells”. Central European Journal of Engineering 2 (2): 248–252.Bibcode:2012CEJE….2..248O. doi:10.2478/s13531-011-0069-7.
  51. Jump up^ Triple-Junction Terrestrial Concentrator Solar Cells. (PDF) . Retrieved on 3 January 2012.
  52. Jump up^ Clarke, Chris (19 April 2011) San Jose Solar Company Breaks Efficiency Record for PV. Optics.org. Retrieved on 19 January 2011.
  53. Jump up^ “NREL effiiciency chart”.
  54. Jump up^ Researchers use liquid inks to create better solar cells, Phys.org, 17 September 2014, Shaun Mason
  55. Jump up^ Hernández-Rodríguez, M.A.; Imanieh, M.H.; Martín, L.L.; Martín, I.R. (September 2013). “Experimental enhancement of the photocurrent in a solar cell using upconversion process in fluoroindate glasses exciting at 1480nm”. Solar Energy Materials and Solar Cells 116: 171–175.doi:10.1016/j.solmat.2013.04.023.
  56. Jump up^ Dye Sensitized Solar Cells. G24i.com (2 April 2014). Retrieved on 20 April 2014.
  57. Jump up^ Semonin, O. E.; Luther, J. M.; Choi, S.; Chen, H.-Y.; Gao, J.; Nozik, A. J.; Beard, M. C. (2011). “Peak External Photocurrent Quantum Efficiency Exceeding 100% via MEG in a Quantum Dot Solar Cell”. Science 334 (6062): 1530–3.Bibcode:2011Sci…334.1530S. doi:10.1126/science.1209845.PMID 22174246.
  58. Jump up^ Kamat, Prashant V. (2012). “Boosting the Efficiency of Quantum Dot Sensitized Solar Cells through Modulation of Interfacial Charge Transfer”. Accounts of Chemical Research 45(11): 120411095315008. doi:10.1021/ar200315d.
  59. Jump up^ Santra, Pralay K.; Kamat, Prashant V. (2012). “Mn-Doped Quantum Dot Sensitized Solar Cells: A Strategy to Boost Efficiency over 5%”. Journal of the American Chemical Society134 (5): 2508–11. doi:10.1021/ja211224s. PMID 22280479.
  60. Jump up^ Moon, Soo-Jin; Itzhaik, Yafit; Yum, Jun-Ho; Zakeeruddin, Shaik M.; Hodes, Gary; GräTzel, Michael (2010). “Sb2S3-Based Mesoscopic Solar Cell using an Organic Hole Conductor”. The Journal of Physical Chemistry Letters 1 (10): 1524.doi:10.1021/jz100308q.
  61. Jump up^ Solar Cell Research || The Prashant Kamat lab at the University of Notre Dame. Nd.edu (22 February 2007). Retrieved on 17 May 2012.
  62. Jump up^ Genovese, Matthew P.; Lightcap, Ian V.; Kamat, Prashant V. (2012). “Sun-BelievableSolar Paint. A Transformative One-Step Approach for Designing Nanocrystalline Solar Cells”. ACS Nano 6(1): 865–72. doi:10.1021/nn204381g. PMID 22147684.
  63. Jump up^ Konarka Power Plastic reaches 8.3% efficiency. pv-tech.org. Retrieved on 7 May 2011.
  64. Jump up^ Mayer, A.; Scully, S.; Hardin, B.; Rowell, M.; McGehee, M. (2007). “Polymer-based solar cells”. Materials Today 10 (11): 28.doi:10.1016/S1369-7021(07)70276-6.
  65. Jump up^ Lunt, R. R.; Bulovic, V. (2011). “Transparent, near-infrared organic photovoltaic solar cells for window and energy-scavenging applications”. Applied Physics Letters 98 (11): 113305. doi:10.1063/1.3567516.
  66. Jump up^ Rudolf, John Collins (20 April 2011). “Transparent Photovoltaic Cells Turn Windows Into Solar Panels”. green.blogs.nytimes.com.
  67. Jump up^ “UCLA Scientists Develop Transparent Solar Cell”. Enviro-News.com. 24 July 2012.
  68. Jump up^ Lunt, R. R.; Osedach, T. P.; Brown, P. R.; Rowehl, J. A.; Bulović, V. (2011). “Practical Roadmap and Limits to Nanostructured Photovoltaics”. Advanced Materials 23 (48): 5712–27. doi:10.1002/adma.201103404. PMID 22057647.
  69. Jump up^ Lunt, R. R. (2012). “Theoretical limits for visibly transparent photovoltaics”. Applied Physics Letters 101 (4): 043902.doi:10.1063/1.4738896.
  70. Jump up^ Guo, C.; Lin, Y. H.; Witman, M. D.; Smith, K. A.; Wang, C.; Hexemer, A.; Strzalka, J.; Gomez, E. D.; Verduzco, R. (2013). “Conjugated Block Copolymer Photovoltaics with near 3% Efficiency through Microphase Separation”. Nano Letters 13 (6): 130522121011001. doi:10.1021/nl401420s.
  71. Jump up^ “Organic polymers create new class of solar energy devices”. Kurzweil Accelerating Institute. 31 May 2013. Retrieved 1 June2013.
  72. ^ Jump up to:a b Bullis, Kevin (July 30, 2014) Adaptive Material Could Cut the Cost of Solar in Half. MIT Technology Review
  73. Jump up^ Fitzky, Hans G. and Ebneth, Harold (May 24, 1983) U.S. Patent 4,385,102, “Large-area photovoltaic cell”
  74. Jump up^ Pv News November 2012. Greentech Media. Retrieved on 3 June 2012.
  75. ^ Jump up to:a b c Jäger-Waldau, Arnulf (September 2013) PV Status Report 2013. European Commission, Joint Research Centre, Institute for Energy and Transport.
  76. Jump up^ PV production grows despite a crisis-driven decline in investment. European Commission, Brussels, 30 September 2013
  77. Jump up^ PV Status Report 2013 | Renewable Energy Mapping and Monitoring in Europe and Africa (REMEA). Iet.jrc.ec.europa.eu (11 April 2014). Retrieved on 20 April 2014.
  78. Jump up^ “NYTimes”. Retrieved 25 July 2015.
  79. Jump up^ Plunging Cost Of Solar PV (Graphs). CleanTechnica (7 March 2013). Retrieved on 20 April 2014.
  80. Jump up^ Falling silicon prices shakes up solar manufacturing industry. Down To Earth (19 September 2011). Retrieved on 20 April 2014.