Open access

Solar Spectrum Conversion for Photovoltaics Using Nanoparticles

Written By

W.G.J.H.M. van Sark, A. Meijerink and R.E.I. Schropp

Submitted: 25 November 2011 Published: 16 March 2012

DOI: 10.5772/39213

From the Edited Volume

Third Generation Photovoltaics

Edited by Vasilis Fthenakis

Chapter metrics overview

12,020 Chapter Downloads

View Full Metrics

1. Introduction

The possibility to tune chemical and physical properties in nanosized materials has a strong impact on a variety of technologies, including photovoltaics. One of the prominent research areas of nanomaterials for photovoltaics involves spectral conversion. Conventional single-junction semiconductor solar cells only effectively convert photons of energy close to the semiconductor band gap (E g ) as a result of the mismatch between the incident solar spectrum and the spectral absorption properties of the material (Green 1982, Luque and Hegedus 2003). Photons with an energy E ph smaller than the band gap are not absorbed and their energy is not used for carrier generation. Photons with energy E ph larger than the band gap are absorbed, but the excess energy E ph E g is lost due to thermalization of the generated electrons. These fundamental spectral losses in a single-junction silicon solar cell can be as large as 50% (Wolf 1971), while the detailed balance limit of conversion efficiency for such a cell was determined to be 31% (Shockley and Queisser 1961). Several routes have been proposed to address spectral losses, and all of these methods or concepts obviously concentrate on a better exploitation of the solar spectrum, e.g., multiple stacked cells (Law et al. 2010), intermediate band gaps (Luque and Marti 1997), multiple exciton generation (Klimov 2006, Klimov et al. 2007), quantum dot concentrators (Chatten et al. 2003a) and down- and up-converters (Trupke et al. 2002a, b), and down-shifters (Richards 2006a, Van Sark 2005). In general they are referred to as Third or Next Generation photovoltaics (PV) (Green 2003, Luque et al. 2005, Martí and Luque 2004). Nanotechnology is essential in realizing most of these concepts (Soga 2006, Tsakalakos 2008), and semiconductor nanocrystals have been recognized as ‘building blocks’ of nanotechnology for use in next generation solar cells (Kamat 2008). Being the most mature approach, it is not surprising that the current world record conversion efficiency is 43.5% for a GaInP/GaAs/GaInNAs solar cell (Green et al. 2011), although this is reached at a concentration of 418 times.

As single-junction solar cells optimally perform under monochromatic light at wavelength λ opt ~ 1240/E g (with λ opt in nm and E g in eV), an approach “squeezing” the wide solar spectrum (300-2500 nm) to a single small band spectrum without too many losses would greatly enhance solar cell conversion efficiency. Such a quasi-monochromatic solar cell could in principle reach efficiencies over 80%, which is slightly dependent on band gap ( Luque and Martí 2003 ). For (multi)crystalline silicon ((m)c-Si) solar cells λ opt = 1100 nm (with E g = 1.12 eV); for amorphous silicon (a-Si:H) the optimum wavelength is λ opt = 700 nm (with E g = 1.77 eV). However, as these cells only contain a thin absorber layer, the optimum spectrum response occurs at about 550 nm (Schropp and Zeman 1998, Van Sark 2002 ).

Modification of the spectrum by means of so-called down- and/or upconversion or -shifting is presently being pursued for single junction cells (Richards 2006a), as illustrated in Fig. 1, as a relatively easy and cost-effective means to enhance conversion efficiency. In addition, so-called luminescent solar converters (LSC) employ spectrum modification as well (Goetzberger 2008, Goetzberger and Greubel 1977). Downconverters or -shifters are located on top of solar cells, as they are designed to modify the spectrum such that UV and visible photons are converted leading to a more red-rich spectrum that is converted at higher efficiency by the solar cell. Upconverters modify the spectrum of photons that are not absorbed by the solar cell to effectively shift the IR part of the transmitted spectrum to the NIR or visible part; a back reflector usually is applied as well.

Figure 1.

Schematic drawings of a solar cell with down converter (or down shifter) layer on top (left) and a solar cell on top of an upconverter layer.

In case of downconversion (DC) an incident high-energy photon is converted into two or more lower energy photons which can lead to quantum efficiency of more than 100%, therefore it is also termed ‘quantum cutting’ (Timmerman et al. 2008, Wegh et al. 1999); for upconversion (UC) two or more low energy photons (sub band gap) are converted into one high-energy photon (Strümpel et al. 2007), see also Fig. 2. Downshifting (DS) is similar to downconversion where an important difference is that only one photon is emitted and that the quantum efficiency of the conversion process is lower than unity (Richards 2006a), although close to unity is preferred to minimize losses. Downshifting is also termed photoluminescence (Strümpel et al. 2007). DC, UC and DS layers only influence solar cell performance optically. As DC and DS both involve one incident photon per conversion, the intensity of converted or shifted emitted photons linearly scales with incident light intensity. UC involves two photons; therefore the intensity of converted light scales quadratically with incident light intensity.

Figure 2.

Energy diagrams showing photon adsorption and subsequent downconversion, downshifting, and upconversion.

In this chapter a review is presented on the use of nanometer sized particles (including quantum dots) in solar spectrum conversion. Modification of the spectrum requires down- and/or upconversion or -shifting of the spectrum, meaning that the energy of photons is modified either to lower (down) or higher (up) energy. Nanostructures such as quantum dots, luminescent dye molecules, and lanthanide-doped glasses are capable of absorbing photons at a certain wavelength and emitting photons at a different (shorter or longer) wavelength. We will discuss down- and upconversion and -shifting by quantum dots, luminescent dyes, and lanthanide compounds, and assess their potential in contributing to ultimately lowering the cost per kWh of solar generated power.

Advertisement

2. Downconversion

2.1. Principles

Downconversion was theoretically suggested first by Dexter in the 1950s (Dexter 1953, 1957), and shown experimentally 20 years later using the lanthanide ion praseodymium Pr3+ in an yttrium fluoride YF3 host (Piper et al. 1974, Sommerdijk et al. 1974). A VUV photon (185 nm) is absorbed in the host, and its energy is transferred into the 1S0 state of the Pr3+ ion, from where two photons (408 and 620 nm) are emitted in a two-step process (1S03PJ at 408 nm followed by 3PJ3F2 at 620 nm). In this way a single absorbed high energy photon results in the emission of two visible photons and a higher-than-unity quantum efficiency is realized. Another frequently used ion is gadolinium Gd3+ (Wegh et al. 1997), either single or co-doped (Wegh et al. 1999). These lanthanide ions are characterized by a rich and well-separated energy level structure in the so-called Dieke energy diagrams (Dieke 1968), and have been identified as perfect “photon managers” (Meijerink et al. 2006). The energy levels arise from the interactions between electrons in the inner 4f shell. Trivalent lanthanides have as electronic configuration [Xe]4fn5s25p6. Inside the filled 5s and 5p shells, there is a partially filled 4f shell where the number of electrons (n) can vary between 0 and 14. The number of possible arrangements for n electrons in 14 available f-orbitals is large (14 over n), which gives rise to a large number of different energy levels that are labelled by so-called term symbols. Transitions between the energy levels give rise to sharp absorption and emission lines. Energy transfer between neighbouring lanthanide ions is also possible and helps converting photons that are absorbed to photons of different energy. Based on their unique and rich energy level structure, lanthanide ions are promising candidates to realize efficient down conversion and recent research in this direction will be discussed below.

Downconversion in solar cells was theoretically shown to lead to a enlarged conversion efficiency of 36.5% (Trupke et al. 2002a) for nonconcentrated sunlight when applied in a single junction solar cell configuration, such as shown in Fig. 1. Note that detailed-balance calculations for single junction c-Si cells lead to a maximum efficiency of 31% (Shockley and Queisser 1961). These calculations were performed as a function of the band gap and refractive index of the solar cell material, for a 6000K blackbody spectrum. The efficiency limit is reached for a band gap of 1.05 eV, and asymptotically approaches 39.63% for very high refractive indices larger than 10. For c-Si, with refractive index of 3.6, the limit efficiency is 36.5%. Analysis of the energy content of the incident standard Air Mass 1.5 Global (AM1.5G) spectrum (ASTM 2003) and the potential gain DC can have shows that with a DC layer an extra amount of 32% is incident on a silicon solar cell (Richards 2006a), which can be converted at high internal quantum efficiency. Figure 3 illustrates the potential gains for DC and UC.

Figure 3.

Potential gain for down- and upconversion for a silicon solar cell. The green part reflects the energy conversion of the absorbed part of the solar spectrum for a c-Si solar cell, the red part reflects the extra energy conversion if every photon with an energy higher than 2Eg results in two NIR photons and the beige part reflects the energy gain if every pair of photons with an energy between 0.5Eg and Eg is converted to one NIR photon. Note that the figure considers no other losses than spectral mismatch losses (Courtesy of F. Rabouw, Utrecht University).

2.2. State of the art

The most promising systems for downconversion rely on lanthanide ions. The unique and rich energy level structures of these ions allow for efficient spectral conversion, including up- and downconversion processes mediated by resonant energy transfer between neighboring lanthanide ions (Auzel 2004, Wegh et al. 1999). Considering the energy levels of all lanthanides, as shown in the Dieke energy level diagram (Dieke 1968, Peijzel et al. 2005, Wegh et al. 2000) it is immediately evident that the energy level structure of Yb3+ is ideally suited to be used in down conversion for use in c-Si solar cells. The Yb3+ ion has a single excited state (denoted by the term symbol 2F5/2) some 10,000 cm-1 above the 2F7/2 ground state, corresponding to emission around 1000 nm. The absence of other energy levels allow Yb3+ to exclusively ‘pick up’ energy packages of 10,000 cm-1 from other lanthanide ions and emitting ~1000 nm photons that can be absorbed by c-Si. Efficient downconversion using Yb3+ as acceptor requires donor ions with an energy level around 20,000 cm-1 and an intermediate level around 10,000 cm-1. From inspection of the Dieke diagram one finds that potential couples are (Er3+, Yb3+), (Nd3+, Yb3+) and (Pr3+, Yb3+) for a resonant two-step energy transfer process. Also, cooperative sensitization is possible where energy transfer occurs from a high excited state of the donor to two neighboring acceptor ions without an intermediate level.

The first report on efficient downconversion for solar cells was based on cooperative energy transfer from Tb3+ to two Yb3+ ions in YbxY1-xPO4:Tb3+ (Vergeer et al. 2005), as shown schematically in Fig. 4. The 5D4 state of Tb3+ is around 480 nm (21,000 cm-1) and from this state cooperative energy transfer to two Yb3+ neighbors occurs, both capable of emitting a 980 nm photon.

Figure 4.

Cooperative energy transfer from Tb to two Yb ions. From the 5D4 state of one Tb ion two neighbouring Yb ions are excited to the 2F5/2 state from where emission of 980 nm photons can occur (after Meijerink et al. 2006, Vergeer et al. 2005).

The same efficient down conversion process was observed in other host lattices: Zhang et al. (Zhang et al. 2007) observed cooperative quantum cutting in (Yb,Gd)Al3(BO3)4:Tb3+ and quantum efficiencies up to 196% were reported. Cooperative downconversion for other couples of lanthanides was also claimed. For the couple (Pr, Yb) and (Tm, Yb) co-doped into borogermanate glasses a decrease of the emission from higher energy levels of Pr and Tm was observed (Liu et al. 2008). In case of Pr3+ the starting level for the cooperative down conversion is the 3P0 level while for Tm3+ the 1G4 is at twice the energy of the 2F5/2 level of Yb3+. The efficiency of the downconversion process was estimated from the decrease of the lifetime of the donor state. For the Tm3+ emission a decrease from 73 µs (for the sample doped with Tm only) to 45 µs (for a sample co-doped with 20% of Yb) was observed, giving a transfer efficiency of 38%. For the same glass co-doped with Pr and Yb a 65% decrease of the lifetime was observed, implying a 165% quantum yield. Similar results were found for the (Pr, Yb) couple in aluminosilicate glasses (Lakshminarayana et al. 2008). More recent work on the (Pr, Yb) couple confirmed the presence of downconversion, but the mechanism involved was pointed out to be a resonant two-step energy transfer rather than a cooperative energy transfer (Van der Ende et al. 2009a, b). For the (Pr, Yb) couple an intermediate level (1G4) is available around 10,000 cm-1, which makes a two-step resonant energy transfer process possible. This first-order process will have a much higher probability than the second-order (cooperative) transfer process. The results demonstrated efficient quantum cutting of one visible photon into two NIR photons in SrF2:Pr3+, Yb3+. Comparison of absorption and excitation spectra provided direct evidence that the downconversion efficiency is close to 200%, in agreement with a two-step energy transfer process that can be expected based on the energy level diagrams of Pr3+ and Yb3+. This first order energy transfer process is effective at relatively low Yb3+ concentrations (5%), where concentration quenching of the Yb3+ emission is limited. Comparison of emission spectra, corrected for the instrumental response, for SrF2:Pr3+ (0.1%) and SrF2:Pr3+ (0.1%), Yb3+ (5%) revealed an actual conversion efficiency of 140%.

An important aspect for downconversion is the incorporation in a solar cell. The solar spectrum has to be converted before entering the solar cell. A promising design is incorporation of downconverting material in a thin (10-100 nm) transparent layer on top of the solar cell. After downconversion the isotropically emitted 1000 nm need to be directed towards the solar cell. This can be achieved by a narrow-band reflective coating for 1000 nm NIR on top of the downconversion layer, thereby avoiding absorption in the other parts of the spectrum. The narrow band emission of the Yb3+ emission allows for a narrow band reflective coating. To realize a transparent downconversion layer, the downconverting lanthanide ion couples need to be incorporated in nanocrystals with dimensions smaller than ~50 nm to prevent light scattering. Chen et al. reported downconversion in transparent glass ceramics with embedded Pr3+/Yb3+:β-YF3 nanocrystals (Chen et al. 2008a). Quantum cutting downconversion was shown in borate glasses using co-doping of Ce3+ and Yb3+ (Chen et al. 2008b). A UV photon (330 nm) excites a Ce3+ ion, and cooperative energy transfer between the Ce3+ and Yb3+ ion leads to the emission of two NIR photons (976 nm) with a 174% quantum efficiency, based on a 74% energy transfer efficiency calculated from the luminescence decay curves and assuming cooperative energy transfer. Recent work on the (Ce, Yb) couple suggests that the energy transfer mechanism is not cooperative energy transfer, but single step energy transfer via a metal-to-metal charge transfer state. If this is the case, the quantum efficiency is not above 100%. Efficient downconversion for other lanthanide ion couples has been reported, including the (Er, Yb) couple in low phonon hosts (Eilers et al. 2010). Research on new downconversion couples continues. It is not straightforward to establish the efficiency of the conversion to NIR photons and careful studies providing convincing evidence for the energy transfer mechanism and efficiency are required (Van Wijngaarden 2010).

Timmerman et al. (Timmerman et al. 2008) demonstrated so-called space-separated quantum cutting within SiO2 matrices containing both silicon nanocrystals and Er3+. Energy transfer from photo exited silicon nanocrystals to Er3+ had been observed earlier (Fujii et al. 1997), but Timmerman et al. show that upon absorption of a photon by a silicon nanocrystal a fraction of the photon energy is transferred generating an excited state within either an erbium ion or another silicon nanocrystal. As also the original silicon NC relaxes from a highly excited state towards the lowest-energy excited state, the net result is two electron-hole pairs for each photon absorbed. Recent quantum yield measurements as a function of excitation wavelength confirmed the occurrence of carrier multiplication in Si nanocrystals (Timmerman et al. 2011).

Advertisement

3. Downshifting

3.1. Principles

Downshifting (DS) or photoluminescence is a property of many materials, and is similar to downconversion, however, only one photon is emitted and energy is lost due to non-radiative relaxation, see Fig. 2. Therefore the quantum efficiency is lower than unity. DS can be employed to overcome poor blue response of solar cells (Hovel et al. 1979), due to, e.g., non-effective front surface passivation for silicon solar cells. Shifting the incident spectrum to wavelengths where the internal quantum efficiency of the solar cell is higher than in the blue can effectively enhance the overall conversion efficiency by ~10% ( Van Sark et al. 2005 ). Improvement of front passivation may make down shifters obsolete, or at least less beneficial. Downshifting layers can also be used to circumvent absorption of high-energy photons in heterojunction window layers, e.g., CdS on CdTe cells (Hong and Kawano 2003). A recent review is presented by Klampaftis et al. (2010).

Downshifting was suggested in the 1970s to be used in so-called luminescent solar concentrators (LSCs) that were attached on to a solar cell (Garwin 1960, Goetzberger and Greubel 1977). In these concentrators, organic dye molecules absorb incident light and re-emit this at a red-shifted wavelength. Internal reflection ensures collection of all the re-emitted light in the underlying solar cells. As the spectral sensitivity of silicon is higher in the red than in the blue, an increase in solar cell efficiency was expected. Also, it was suggested to use a number of different organic dye molecules of which the re-emitted light was matched for optimal conversion by different solar cells. This is similar to using a stack of multiple solar cells, each sensitive to a different part of the solar spectrum. The expected high efficiency of ~30% (Smestad et al. 1990, Yablonovitch 1980) in practice was not reached as a result of not being able to meet the stringent requirements to the organic dye molecules, such as high quantum efficiency and stability, and the transparency of collector materials in which the dye molecules were dispersed (Garwin 1960, Goetzberger and Greubel 1977).

Nowadays, new organic dyes can have extremely high luminescence quantum efficiency (LQE) (near unity) and are available in a wide range of colors at better re-absorption properties that may provide necessary UV stability. Quantum dots (QDs) have been proposed for use in luminescent concentrators to replace organic dye molecules: the quantum dot concentrator (Barnham et al. 2000, Chatten et al. 2003a,b, Gallagher et al. 2007). Quantum dots are nanometer-sized semiconductor crystals of which the emission wavelength can be tuned by their size, as a result of quantum confinement (Alivisatos 1996, Gaponenko 1998). Recently, both QDs and new organic dyes have been evaluated for use in LSCs (Van Sark et al. 2008a). QDs have advantages over dyes in that: (i) their absorption spectra are far broader, extending into the UV, (ii), their absorption properties may be tuned simply by the choice of nanocrystal size, and (iii) they are inherently more stable than organic dyes (Bruchez Jr. et al. 1998). Moreover, (iv) there is a further advantage in that the red-shift between absorption and luminescence is quantitatively related to the spread of QD sizes, which may be determined during the growth process, providing an additional strategy for minimizing losses due to re-absorption (Barnham et al. 2000). However, as yet QDs can only provide reasonable LQE: a LQE > 0.8 has been reported for core-shell QDs (Peng et al. 1997). Performance in LSCs has been modeled using thermodynamic as well as ray-trace models (Burgers et al. 2005, Chatten et al. 2003b, 2004a, b, Chatten et al. 2005, Gallagher et al. 2004, Kennedy et al. 2008), and results are similar and also compare well with experimental values (Kennedy et al. 2008, Van Sark et al. 2008a). Calculated efficiencies vary between 2.4% for an LSC with mc-Si cell at certain mirror specifications to 9.1% for an LSC with InGaP cell for improved specifications.

Alternatives for dye molecules used in LSCs are luminescent ions. Traditionally, efficient luminescent materials rely on the efficient luminescence of transition metal ions and lanthanide ions. In case of transition metal ions intraconfigurational 3dn transitions are responsible for the luminescence, while in case of lanthanide ions both intraconfigurational 4fn-4fn transitions and interconfigurational 4fn-4fn-15d transitions are capable of efficient emission. In most applications efficient emission in the visible is required and emission from lanthanide ions and transition metal ions is responsible for almost all the light from artificial light sources (e.g. fluorescent tubes, displays (flat and cathode ray tube, CRT) and white light emitting diodes (LEDs) (Blasse and Grabmaier 1995). For LSCs to be used in combination with c-Si solar cells efficient emission in the NIR is needed. The optimum wavelength is between 700 and 1000 nm, which is close to the band gap of c-Si and in the spectral region where c-Si solar cells have their optimum conversion efficiency. Two types of schemes can be utilized to achieve efficient conversion of visible light into narrow band NIR emission. A single ion can be used if the ion shows a strong broad band absorption in the visible spectral range followed by relaxation to the lowest excited state from which efficient narrow band or line emission in the NIR occurs. Alternatively, a combination of two ions can be used where one ion (the sensitizer) absorbs the light and subsequently transfers the energy to a second ion (the activator), which emits efficiently in the NIR. Both concepts have been investigated for LSCs by incorporating luminescent lanthanides and transition metal ions in glass matrices. The stability of these systems is not a problem, in contrast to LSCs based on dye molecules, however, the quantum efficiency of luminescent ions in glasses appeared to be much lower than in crystalline compounds, especially in the infrared, thus hampering use for LSCs.

3.2. State of the art

Chung et al. (Chung et al. 2007) reported downshifting phosphor coatings consisting of Y2O3:Eu3+ or Y2O2S:Eu3+ dispersed in either polyvinyl alcohol or polymethylmethacrylate on top of mc-Si solar cells: an increase in conversion efficiency was found of a factor of 14 under UV illumination by converting the UV radiation (for which the response of c-Si is low) to 600 nm emission from the Eu3+ ion. The used solar cells used were encapsulated in an epoxy that absorbs photons with energy higher that ~3 eV. This protective coating can remain in place and down converters or shifters can easily be added. A 2.9% relative increase in efficiency has been reported for Eu3+ containing ethyline-vinyl-acetate layers (EVA) on top of c-Si cells; this is very relevant as EVA nowadays is the standard encapsulating matrix used for c-Si technology.

Svrcek et al. (Svrcek et al. 2004) demonstrated that silicon nanocrystals incorporated into spin-on-glass (SOG) on top of c-Si solar cells are successful as downshifter leading to a potential absolute efficiency enhancement of 1.2%, while they experimentally showed an enhancement of 0.4% (using nanocrystals of 7 nm diameter with a broad emission centered around 700 nm). McIntosh et al. (McIntosh et al. 2009) recently presented results on encapsulated c-Si solar cells, of which the PMMA encapsulant contained down-shifting molecules, i.e., Lumogen dyes. These results indicate a ~1% relative increase in the module efficiency, based on a 40% increase in external quantum efficiency for wavelengths < 400 nm. Mutlugun et al. (Mutlugun et al. 2008) claimed a two-fold increase in efficiency applying a so-called nanocrystal scintillator on top of a c-Si solar cell; it comprises a PMMA layer in which CdSe/ZnS core-shell quantum dots (emission wavelength 548 nm) are embedded. However, the quality of their bare c-Si cells is very poor. Yuan et al. (2011) presented a 14% improved internal quantum efficiency of a silicon oxide layer containing silicon nanocrystals on top of c-Si cells.

Stupca et al. (2007) demonstrated the integration of ultra thin films (2-10 nm) of monodisperse luminescent Si nanoparticles on polycrystalline Si solar cells. 1-nm sized blue emitting and 2.85-nm red emitting particles enhanced the conversion efficiency by 60% in the UV, and 3 to 10% in the visible for the red and blue emitting particles, respectively, similarly to what has been predicted ( Van Sark et al. 2005 ). An 8% relative increase in conversion efficiency was reported (Maruyama and Kitamura 2001) for a CdS/CdTe solar cell, where the coating in which the fluorescent coloring agent was introduced increased the sensitivity in the blue; a maximum increase in efficiency was calculated to be 30-40%. Others showed results that indicate a 6% relative increase in conversion efficiency (Maruyama and Bandai 1999) upon coating a mc-Si solar cell. The employed luminescent species has an absorption band around 400 nm and a broad emission between 450 and 550 nm. As QDs have a much broader absorption it is expected that in potential the deployment of QDs in planar converters could lead to relative efficiency increases of 20-30%. Downshifting employing QDs in a polymer composite has been demonstrated in a light-emitting diode (LED), where a GaN LED was used as an excitation source for QDs emitting at 590 nm (Lee et al. 2000). Besides QDs, other materials have been suggested such as rare earth ions and dendrimers (Serin et al. 2002). A maximum increase of 22.8% was calculated for a thin film coating of KMgF3 doped with Sm on top of a CdS/CdTe solar cell, while experimental results show an increase of 5% (Hong and Kawano 2003).

Recent efforts to surpass the historical 4% efficiency limit of LSCs (Goetzberger 2008, Wittwer et al. 1984, Zastrow 1994), albeit for smaller area size, have been successful. For example, Goldschmidt et al. (2009) showed for a stack of two plates with different dyes, to which four GaInP solar cells were placed at the sides, that the conversion efficiency is 6.7%; the plate was small (4 cm2), and the concentration ratio was only 0.8. It was argued that the conversion efficiency was limited by the spectral range of the organic dyes used, and that if the same quantum efficiency as was reached for the 450-600 nm range could be realized for the range 650-1050 nm an efficiency of 13.5% could be within reach. They also discuss the benefits of a photonic structure on top of the plate, to reduce the escape cone loss (Goldschmidt et al. 2009). The proposed structure is a so-called rugate filter; this is characterized by a varying refractive index in contrast to standard Bragg reflectors, which suppresses the side loops that could lead to unwanted reflections. The use of these filters would increase the efficiency by ~20%, as was determined for an LSC consisting of one plate and dye. Slooff et al. (2008) presented results on 50x50x5 mm3 PMMA plates in which both CRS040 and Red305 dyes were dispersed at 0.003 and 0.01 wt%, respectively. The plates were attached to either mc-Si, GaAs or InGaP cells, and a diffuse reflector (97% refection) was used at the rear side of the plate. The highest efficiency measured was 7.1% for 4 GaAs cells connected in parallel (7% if connected in series).

As stated above, quantum dots are potential candidates to replace organic dye molecules in an LSC, for their higher brightness, better stability, and wider absorption spectrum (Barnham et al. 2000, Chatten et al. 2003a, b, Gallagher et al. 2007). In fact, the properties and availability of QDs started renewed interest in LSCs around 2000 (Barnham et al. 2000), with the main focus on modeling, while more recently some experimental results have been presented. Schüler et al. (2007) proposed to make LSCs by coating transparent glass substrates with QD-containing composite films, using a potentially cheap sol-gel method. They reported on the successful fabrication of thin silicon oxide films that contain CdS QDs using a sol-gel dip-coating process, whereby the 1-2 nm sized CdS QDs are formed during thermal treatment after dip-coating. Depending on the anneal temperature, the colors of the LSC ranged from green for 250°C to yellow for 350°C and orange for 450°C. Reda (2008) also prepared CdS QD concentrators, using sol-gel spin coating, followed by annealing. The annealing temperature was found to affect absorption and emission spectrum: luminescent intensity and Stokes’ shift both decreased for 4 weeks outdoor exposure to sunlight, which probably was caused by aggregation and oxidation. It is known that oxidation leads to blue-shifts in emission ( Van Sark et al. 2002 ). Blue-shifts have also been observed by Gallagher et al. (2007) who dispersed CdSe QDs in several types of resins (urethane, PMMA, epoxy), for fabrication of LSC plates.

Bomm et al. (2011) have addressed several problems regarding incorporation of QDs in an organic polymer matrix, viz. phase separation, agglomeration of particles leading to turbid plates, and luminescence quenching due to exciton energy transfer (Koole et al. 2006). They have synthesized QDSCs using CdSe core/multishell QDs (Koole et al. 2008) (QE = 60%) that were dispersed in laurylmethacrylate (LMA), see also Lee et al. (2000) and Walker et al. (2003). UV-polymerization was employed to yield transparent PLMA plates with QDs without any sign of agglomeration. To one side of this plate, a mc-Si solar cell was placed, and aluminum mirrors to all other sides. Compared to the bare cell (5x0.5 cm2) that generated a current density of 40.28 mA/cm2 at 1000 W AM1.5G spectrum, the best QDC made generated a current of 77.14 mA/cm2, nearly twice as much. The QDC efficiency is 3.5%. In addition, exposure to a 1000 W sulphur lamp for 280 hours continuously showed very good stability: the current density decreased by 4% only, on average. However, re-absorption may still be a problem, as is demonstrated by a small red shift in the emission spectrum for long photon pathways; also absorption by the matrix is occurring. Besides QDs also nanorods (NRs) have been dispersed in PLMA, showing excellent transmittance of 93%; for long rods (aspect ratio of 6) a QE of 70% is observed, which is only slightly smaller than the QE in solution, implying that these rods are stable throughout the polymerization process. In addition, these NRs have also been dispersed in cellulosetriacetate (CTA), and a ~10 μm thin film on a glass substrate was made showing bright orange luminescence (Bomm et al. 2010).

Hyldahl et al. (2009) used commercially available CdSe/ZnS core/shell QDs with QE=57% in LSCs, both liquid (QDs dissolved in toluene, between two 6.2x6.2x0.3 cm glass plates) and solid (QDs dispersed in epoxy), and they obtained an efficiency of 3.98% and 1.97%, respectively. They also used the organic dye Lumogen F Red300, and obtained efficiency of 2.6% in toluene.

Advertisement

4. Upconversion

4.1. Principles

Upconversion was, like DC also suggested in the 1950s, by Bloembergen (1959), and was related to the development of IR detectors: IR photons would be detected through sequential absorption, as would be possible by the arrangement of energy levels of a solid. However, as Auzel (2004) pointed out the essential role of energy transfer was only recognized nearly 20 years later. Several types of upconversion mechanisms exist (Auzel 2004), of which the APTE (addition de photon par transferts d’energie) or, in English, ETU (energy transfer upconversion) mechanism is the most efficient; it involves energy transfer from an excited ion, named sensitizer, to a neighbouring ion, named activator (see Fig. 5).

Figure 5.

Schematic representation of two upconversion processes and the characteristic time response of the up converted emission after a short excitation pulse. (a), (b) Ground state absorption (GSA) followed by excited state absorption (ESA). This is a single ion process and takes place during the excitation pulse. (c), (d) Upconversion is achieved by GSA followed energy transfer between ions and the delayed response is characteristic of the energy transfer up conversion (ETU) (from Suyver et al. 2005a).

Others are two-step absorption, being a ground state absorption (GSA) followed by an excited state absorption (ESA), and second harmonics generation (SHG). The latter mechanism requires extremely high intensities, of about 1010 times the sun’s intensity on a sunny day, to take place (Strümpel et al. 2007). This may explain that research in this field with focus on enhancing solar cell efficiency was started only recently (Shalav et al. 2007).

Upconverters usually combine an active ion, of which the energy level scheme is employed for absorption, and a host material, in which the active ion is embedded. The most efficient upconversion has been reported for the lanthanide ion couples (Yb, Er) and (Yb, Tm); the corresponding upconversion schemes are shown in Fig. 6. The first demonstration of such an UC layer on the back of solar cells comprised an ultra thin (3 μm) GaAs cell (band gap 1.43 eV) that was placed on a 100 μm thick vitroceramic containing Yb3+ and Er3+ (Gibart et al. 1996): it showed 2.5% efficiency upon excitation of 256 kW/m2 monochromatic sub-band-gap (1.391 eV) laser light (1 W on 0.039 cm2 cell area), as well as a clear quadratic dependence on incident light intensity.

Figure 6.

Up conversion by energy transfer between Yb3+ and Er3+ (left) and Yb3+ and Tm3+ (right). Excitation is around 980 nm in the Yb3+ ion and by a two-step energy transfer or three-step energy transfer process higher excited states of Er3+ or Tm3+ are populated giving rise to visible (red, green or blue) emission. Full, dotted, and curly arrows indicate radiative, nonradiative energy transfer, and multiphonon relaxation processes, respectively (from Suyver et al. 2005a).

Upconversion in solar cells was calculated to potentially lead to a maximum conversion efficiency of 47.6 % (Trupke et al. 2002b) for nonconcentrated sunlight using a 6000K blackbody spectrum in detailed-balance calculations. This optimum is reached for a solar cell material of ~2 eV band gap. Applied on the back of silicon solar cells,the efficiency limit would be ~37% (Trupke et al. 2002b). The analysis of the energy content of the incident AM1.5G spectrum presented in Fig. 3 revealed that cells with an UC layer would benefit from an extra amount of 35% light incident in the silicon solar cell (Richards 2006a).

4.2. State of the art

Lanthanides have also been employed in upconverters attached to the back of bifacial silicon solar cells. Trivalent erbium is ideally suited for up conversion of NIR light due to its ladder of nearly equally spaced energy levels that are multiples of the 4I15/2 to 4I13/2 transition (1540 nm) (see also Fig. 3). Shalav et al. (2005) have demonstrated a 2.5% increase of external quantum efficiency due to up conversion using NaYF4:20% Er3+. By depicting luminescent emission intensity as a function of incident monochromatic (1523 nm) excitation power in a double-log plot, they showed that at low light intensities a two-step up conversion process (4I15/24I13/24I11/2) dominates, while at higher intensities a three-step up conversion process (4I15/24I13/24I11/24S3/2 level) is involved.

Strümpel et al. have identified the materials of possible use in up- (and down-) conversion for solar cells (Strümpel et al. 2007). In addition to the NaYF4:(Er,Yb) phosphor, they suggest the use of BaCl2:(Er3+,Dy3+) (Strümpel et al. 2005), as chlorides were thought to be a better compromise between having a low phonon energy and a high excitation spectrum, compared to the NaYF4 (Gamelin and Güdel 2001, Ohwaki and Wang 1994, Shalav et al. 2007). These lower phonon energies lead to lower non-radiative losses. In addition, the emission spectrum of Dysprosium is similar to that of Erbium, but the content of Dy3+ should be <0.1%, to avoid quenching (Auzel 2004, Strümpel et al. 2007).

NaYF4 co-doped with (Er3+, Yb3+) is to date the most efficient upconverter (Suyver et al. 2005a, Suyver et al. 2005b), with ~50% of all absorbed NIR photons up converted and emitted in the visible wavelength range. However, the (Yb, Er) couple is not considered beneficial for up conversion in c-Si cells, as silicon also absorbs in the 920-980 nm wavelength range. These phosphors can be useful for solar cells based on higher band gap materials such as the Grätzel-cell (O'Regan and Grätzel 1991), a-Si(Ge):H or CdTe. In that case, the 2F5/2 level of Yb3+ would serve as an intermediate step for up conversion (Shalav et al. 2007) and IR-radiation between ~700 and 1000 nm that is not absorbed in these wider band gap solar cells can be converted into green (550 nm) light that can be absorbed. Nanocrystals of NaYF4:Er3+, Yb3+ also show upconversion. An advantage of using nanocrystals is that transparent solutions or transparent matrices with upconverting nanocrystals can be obtained. An excellent recent review on upconverting nanoparticles summarizes the status of a variety of UC materials that are presently available as nanocrystals, mostly phosphate and fluoride NCs (Haase et al., 2011). A problem with upconversion NCs is the lower upconversion efficiency (Boyer et al., 2010). There is a clear decrease in efficiency with decreasing size in the relevant size regime between 8 and 100 nm. The decrease in efficiency is probably related to surface effects and quenching by coupling with by high energy vibrations in molecules attached to the surface.

When applying an upconverter, it is advantageous to use it with a cell that with a rather high band gap, such as amorphous Si (1.7 eV). A typical external collection efficiency (ECE) graph of a standard single junction p-i-n a Si:H solar cell is shown in Fig. 7. These cells are manufactured on textured SnO2:F coated glass substrates and routinely have >10% initial efficiency. Typically, the active Si layer in the cell has a thickness of 300 nm and the generated current is 14.0-14.5 mA/cm2, depending on the light trapping properties of the textured metal oxide and the back reflector. After light-induced introduction of the stabilized defect density (Staebler-Wronski effect (Schropp and Zeman 1998)), the stabilized efficiency is 8.2-8.5%. From Fig. 7 it can be seen that the maximum ECE is 0.8 at ~550 nm, and the cut-off occurs at 700 nm, with a response tailing towards 800 nm. The response is shown with and without the use of a buffer layer at the front of the cell. The buffer layer causes already an improvement at the short wavelengths, and a down converter or shifter may not be beneficial. The purpose of an upconverter is to tune the energy of the emitted photons to the energy where the spectral response shows a maximum. If the energy of the emitted photons is too close to the absorption limit (the band gap edge), then the absorption coefficient is too low and the up converted light would not be fully used.

Figure 7.

The spectral response for cells with and without a buffer layer at the p/i interface in the as deposited state obtained at -1 V bias voltage. Inset: the p-i-n- structure showing the position of the buffer layer (from Munyeme et al. 2004).

The photogenerated current could be increased by 40% if the spectral response was sustained at high level up to the band gap cut off at 700 nm and by even more if light with wavelengths λ>700 nm could be more fully absorbed. These two effects can be achieved with the upconversion layer, combined with a highly reflecting back contact. While the upconversion layer converts sub-bandgap photons to super-bandgap photons that can thus be absorbed, a non-conductive reflector is a much better alternative than any metallic mirror, thus sending back both the unabsorbed super-bandgap photons as well as the “fresh” super-bandgap photons into the cell. It is estimated that the stabilized efficiency of the 8.2-8.5% cell can be enhanced to ~12 %. Besides a-Si, a material denoted as protocrystalline Si could be used; this is an amorphous material that is characterized by an enhanced medium range structural order and a higher stability against light-induced degradation compared to standard amorphous silicon. The performance stability of protocrystalline silicon is within 10% of the initial performance; its band gap is slightly higher (1.8 eV) than amorphous silicon.

De Wild et al. (2011) have demonstrated upconversion for a-Si cells with NaYF4 co-doped with (Er3+, Yb3+) as upconverter. The upconverter shows absorption of 980 nm (by the Yb3+ ion) leading to efficient emission of 653 nm (red) and 520-540 nm (green) light (by the Er3+) after a two-step energy transfer process. The narrow absorption band around 980 nm for Yb3+ limits the spectral range of the IR that can be used for up conversion. An external quantum efficiency of 0.02% at 980 nm laser irradiation was obtained. By using a third ion (for example Ti3+) as a sensitizer the full spectral range between 700 and 980 nm can be efficiently absorbed and converted to red and green light by the Yb-Er couple. The resulting light emission in the green and red region is very well absorbed by the cell with very good quantum efficiency for electron-hole generation.

Upconversion systems consisting of lanthanide nanocrystals of YbPO4 and LuPO4 have been demonstrated to be visible by the naked eye in transparent solutions, however at efficiency lower than solid state up conversion phosphors (Suyver et al. 2005a). Other host lattices (NaXF4, X=Y,Gd,La) have been used and codoping with Yb3+ and Er3+, or Yb3+ and Tm3+ appeared successful, where Yb3+ acts as sensitizer. Nanocrystals of <30 nm in size, to prevent scattering in solution, have been prepared and they can be easily dissolved in organic solvents forming colloidal solutions, without agglomeration. Further efficiency increase is possible by growing a shell of undoped NaYF4 around the nanocrystal; in addition, surface modification is needed to allow dissolution in water, for use in biological labeling.

Porous silicon layers are investigated for use as up converter layers as host for rare-earth ions, because these ions can easily penetrate the host due to the large surface area and porosity. A simple and low-cost dipping method has been reported (Díaz-Herrera et al. 2009), in which a porous silicon layer is dipped in to a nitrate solution of erbium and ytterbium in ethanol (Er(NO3)3:Yb(NO3)3:C2H5OH), which is followed by a spin-on procedure and a thermal activation process at 900 ºC. Excitation of the sample at 980 nm revealed up-conversion processes as visible and NIR photoluminescence is observed; codoping of Yb with Er is essential, doping only with Er shows substantial quenching effects (González-Díaz et al. 2008).

Sensitized triplet-triplet annihilation (TTA) using highly photostable metal-organic chromophores in conjunction with energetically appropriate aromatic hydrocarbons has been shown to be another alternative up conversion system (Singh-Rachford et al. 2008). This mechanism was shown to take place under ambient laboratory conditions, i.e. low light intensity conditions, clearly of importance for outdoor operation. These chromophores (porphyins in this case) can be easily incorporated in a solid polymer so that the materials can be treated as thin film materials (Islangulov et al. 2007). A problem with TTA upconverters is the spectral range. No efficient upconversion of NIR radiation at wavelengths beyond 800 nm has been reported which limits the use to very wide bandgap solar cells (Singh-Rachford et al, 2010).

Advertisement

5. Modeling of spectral conversion

An extension to the models described above was presented in a study by Trupke et al. (2006), in which realistic spectra were used to calculate limiting efficiency values for up conversion systems. Using an AM1.5G spectrum leads to a somewhat higher efficiency of 50.69% for a cell with a band gap of 2.0 eV. For silicon, the limiting efficiency would be 40.2%, or nearly 10% larger than the value of 37% obtained for the 6000K blackbody spectrum (Trupke et al. 2002b). This increase was explained by the fact that absorption in the earths’ atmosphere at energies lower than 1.5 eV (as evident in the AM1.5G spectrum) leads to a decrease in light intensity. Badescu and Badescu (2009) have presented an improved model, that according to them it appropriately takes into account the refractive index of solar cell and converter materials. Two configurations are studied: cell and rear converter (C-RC), the usual up converter application, and front converter and cell (FC-C), respectively. They confirm the earlier results of Trupke et al. (2002b) in that the limiting efficiency is larger than that of a cell alone, with higher efficiencies at high concentration. Also, the FC-C combination, i.e., up converter layer on top of the cell, does not improve the efficiency, which is obvious. Further, by studying the variation of refractive indexes of cell and converter separately, as opposed to Trupke et al. (2002b), it was found that the limiting efficiency increases with refractive index of both cell and up converter. In practice, a converter layer may have a lower refractive index (1.5, for a transparent polymer: polymethylmethacrylate (PMMA) (Richards and Shalav 2005)), than that of a cell (3.4). Using a material with similar refractive index as the cell would improve the efficiency by about 10%.

In a series of papers the groups of Badescu, De Vos and co-workers (Badescu and De Vos 2007, Badescu et al. 2007 , De Vos et al. 2009) have re-examined the model for down conversion as proposed by Trupke et al. (2002a), and have added the effects of non-radiative recombination and radiation transfer through interfaces. Analogous to the model for up converters they studied FC-C and C-RC configurations, with down conversion or shifting properties. First, neglecting non-radiative recombination, Badescu et al. (2007 ) qualitatively confirm the results presented by Trupke and co-workers (2002a). For both configurations, the efficiency of the combined system is larger than that of a single mono-facial cell, albeit that the efficiency is smaller (~26%) due to inclusion of front reflections. Second, including radiative recombination for both converter and cell does only increase the efficiency for high (near unity) radiative recombination efficiency values. Interestingly, they report that in this case the C-RC combination cell-rear converter yields a higher efficiency than the FC-C combination in high-quality solar cells, while for low quality solar cells, this is reversed. More realistic device values and allowing for different refractive indices in cell and converter was studied by ( Badescu and De Vos 2007 ), leading to the conclusion that in reality down converters may not always be beneficial. However, extending the model once more, with the inclusion of anti-reflection coating and light trapping texture, showed a limiting efficiency of 39.9%, as reported by De Vos et al. (2009).

Del Cañizo et al. (2008) presented a Monte Carlo ray-tracing model, in which photon transport phenomena in the converter/solar cell system are coupled to non-linear rate equations that describe luminescence. The model was used to select candidate materials for up- and down-conversion, but was set-up for use with rare-earth ions. Results show that for both converters, the potential gain in short-circuit current is small, and may reach 6-7 mA/cm2 at intensities as high as 1000 suns, in correspondence with earlier work by Shalav et al. (2007).

Modelling downshifting layers on solar cells was also extended for non-AM1.5G spectra, including varying air mass between 1 and 10, and diffuse and direct spectra ( Van Sark 2005 ). Here, the PC1D model (Basore and Clugston 1996) was used to model quantum dots dispersed in a PMMA layer on top of a multi-crystalline silicon cell (mc-Si) as function of the concentration of quantum dots. Annual performance has been modelled by using modelled spectra from the model SEDES2 (Houshyani Hassanzadeh et al. 2007, Nann and Riordan 1991); these spectra can be considered realistic as actual irradiation data is used as input. It was found that the simulated short current enhancement, which varies between about 7% and 23%, is linearly related with the average photon energy (APE) of the spectra, based on hourly spectra of four typical and other randomly selected days throughout the year, and of monthly spectra. The annual short circuit increase was determined at 12.8% using the annual distribution of APE values and their linear relation (Van Sark 2007), which is to be compared with the 10% increase in case of the AM1.5G spectrum. For mc-Si cells with improved surface passivation and a concomitant improved blue response the relative short current increase has been calculated to be lower (Van Sark 2006).

Modeling large area LSCs has indicated the importance of top-surface losses that occur through the escape cone (Chatten et al. 2007) both through primary emission and through emission of luminescence that has been reabsorbed and might otherwise have been trapped via total internal reflection or by mirrors. As an example, for an idealized (perfectly transparent host, LQE = 1), mirrored (perfect mirrors on one short and two long edges, and the bottom surface), 40×5×0.5 cm LSC doped with CdSe/ZnS core-shell QDs with emission matched to a GaInP cell, 24% of the AM1.5G spectrum is absorbed. The photon concentration ratio, C, which is the ratio of the concentrated flux escaping the right-hand surface of the LSC to the flux incident on the top surface, is 4.18. However, since the concentrated flux escaping the right-hand surface is a narrow-band matched to the spectral response of the cell it can all be converted and the idealized LSC would produce 8 times the current compared to the cell alone exposed to AM1.5G. However, 78% of the luminescence is lost through the large top-surface area and only 22% may be collected at the right-hand surface. This lead to the use of wavelength-selective cholesteric liquid crystal coatings applied to the top surface in order to reduce the losses (Debije et al. 2006), as these coatings are transparent to incoming light but reflect the emitted light.

Ray-tracing for LSCs uses basic ray-tracing principles, which means that a ray representing light of a certain wavelength travelling in a certain direction, is traced until it leaves the system e.g. by absorption or reflection at the interface (Burgers et al. 2005, Gallagher et al. 2004). The main extension to the standard ray-tracing model is the handling of the absorption and emission by the luminescent species in the LSC. Ray-tracing has been used to perform parameter studies for a 5x5 cm2 planar LSC to find attainable LSC efficiencies. The concentrator consists of a PMMA plate (refractive index n=1.49, absorption 1.5 m-1) doped with two luminescent dyes, CRS040 from Bayer and Lumogen F Red 305 from BASF, with a FQE of 95% (Slooff et al. 2006). With the ray-tracing model the efficiency of this plate together with a mc-Si solar cell was determined to be 2.45%. Results for attaining efficiencies for other configurations are shown in Table 1 (Van Sark et al. 2008a). Replacing the mc-Si cell by a GaAs cell or an InGaP cell, will increase the efficiency from 3.8% to 6.5% and 9.1%, respectively (based on Voc (FF) values of 0.58 V (0.83), 1.00 V (0.83), and 1.38 V (0.84), for mc-Si, GaAs, InGaP, respectively). Thus, the use of GaAs or InGaP cells will result in higher efficiencies, but these cells are more expensive. A cost calculation must be performed to determine if the combination of the luminescent concentrator with this type of cells is an interesting alternative to mc-Si based solar technology.

mc-SiGaAsInGaPparameters
2.44.25.9fixed mirrors, 85% reflectivity, dyes with 95% LQE
2.95.17.197% reflectivity “air-gap mirrors” on sides, and 97% reflectivity Lambertian mirror at bottom
3.45.98.3reduce background absorption of polymer matrix from 1.5 m-1 to 10-3 m-1
3.86.59.1increase of refractive index from 1.49 to 1.7

Table 1.

Calculated efficiencies (in %) for an LSC for various optimized configurations and parameters (Van Sark et al. 2008a). A Lambertian mirror is one that reflects radiation isotropically.

Currie et al. (2008) projected conversion efficiencies as high as 6.8%, for a tandem LSC based on two single LSCs that consist of thin layer of deposited organic dye molecules onto a glass plate to which a GaAs cell was attached. Using CdTe or Cu(In,Ga)Se2 solar cells conversion efficiencies of 11.9% and 14.5%, respectively, were calculated.

Annual performance has been modeled using an LSC of which the properties and geometry resulted from a cost-per-unit-of-power optimization study by Bende et al. (2008). A square plate of 23.7x23.7x0.1 cm3 was used and in the ray-trace model attached to four c-Si solar cells (18.59% efficiency) on all sides. Using a Lumogen F Red 305 (BASF) dye it was calculated that 46.5% of all photons in the wavelength range of 370-630 nm were collected, leading to an LSC efficiency of 4.24% (Van Sark et al. 2008b). The annual yield of this LSC was determined using realistic spectra representative for the Netherlands from Houshyani Hassanzadeh and collaborators (2007) and amounted to 41.3 kWh/m2, and this is equivalent to an effective annual efficiency of 3.81%.

Advertisement

6. Discussion and outlook

Two issues remain to be solved before downconverters can be applied in solar cells: the absorption strength needs to be increased as the transitions involved for the trivalent lanthanides are sharp and weak (parity forbidden). A second issue is concentration quenching. High Yb-concentrations are needed for efficient energy transfer as every donor (Tb, Pr or Tm) needs to have two Yb-neighbours for energy transfer. For these high concentrations, energy migration over the Yb-sublattice occurs and trapping of the migrating excitation energy by quenching sites strongly reduces the emission output. At present research is conducted to resolve these issues. The limited absorption can be solved by the inclusion of a sensitizer for the 3P0 level of Pr3+, which is able to absorb efficiently over a broad wavelength range (300-500 nm) and subsequent energy transfer to the 3P0 level of Pr3+. In principle, such sensitization can be realized in an efficient and cost-effective manner by the inclusion of a sensitizer ion. The 4f-5d luminescence of Ce3+, for example, is often used to sensitize the Tb3+ luminescence in phosphors for fluorescent tubes. Concentration quenching may be limited by optimization of the synthesis conditions (less quenching sites) while also the synthesis of nanocrystals may be beneficial. In nanocrystals the volume probed by energy migration is limited (due to the small size of the nanocrystal) and for defect-free nanocrystals high quantum yields can be expected, similar to the increase in quantum yield observed for quantum dots vs. bulk semiconductors.

For upconverters based on lanthanides absorption strengths need also to be increased and quenching decreased. In addition, upconversion could be useful for solar cells with band gap higher than that of crystalline silicon; thus, research is directed toward optimum matching of NIR absorption and visible emission with the band gap of the solar cell to which the up converter is attached.

Modelling studies on incorporation of conversion layers on top (down converter or down shifter) or at the bottom (upconverter) of single junction solar cells have shown that the conversion efficiency may increase by about 10% (Glaeser and Rau 2007, Richards 2006b, Strümpel et al. 2007, Trupke et al. 2006, Van Sark 2005 , Van Sark et al. 2005 ). This is still an experimental challenge. Also, stability of converter materials is a critical issue, as their lifetime should be >20 years to comply with present solar module practice.

The usefulness of down- and upconversion and downshifting depends on the incident spectrum and intensity. While solar cells are designed and tested according to the ASTM standard (ASTM 2003), these conditions are rarely met outdoors, as discussed by Makrides and collaborators in Chapter 8 of this book. Spectral conditions for solar cells vary from AM0 (extraterrestrial) via AM1 (equator, summer and winter solstice) to AM10 (sunrise, sunset). The weighted average photon energy (APE) (Minemoto et al. 2007) can be used to parameterize this; the APE (using the range 300-1400 nm) of AM1.5G is 1.674 eV, while the APE of AM0 and AM10 are 1.697 eV and 1.307 eV, respectively. Further, overcast skies cause higher scattering leading to diffuse spectra, which are blue-rich, e.g., the APE of the AM1.5 diffuse spectrum is calculated to be 2.005 eV, indeed much larger than the APE of the AM1.5 direct spectrum of 1.610 eV. As DC and DS effectively red-shift spectra, the more blue an incident spectrum contains (high APE) the more gain can be expected ( Van Sark 2005 , 2008). Application of DC layers will therefore be more beneficial for regions with high diffuse irradiation fraction, such as Northwestern Europe, where this fraction can be 50% or higher. Thus, luminescent solar concentrators are expected to be deployed successfully in such regions (Van Sark et al. 2008b). In contrast, solar cells with UC layers will be performing well in countries with high direct irradiation fractions or in early morning and evening due to the high air mass resulting in low APE, albeit that the non-linear response to intensity may be limiting.

The variation of the incident spectra are of particular concern for series-connected multiple junction cells, such as triple a-Si:H (Krishnan et al. 2009) and GaAs-based cells. Current-mismatch due to spectral differences with respect to the AM1.5G standard leads to lowering of the conversion efficiency, as the cell with the lowest current determines the total current. It has been shown that the calculated limiting efficiency of a GaAs-based triple cell is reduced to 32.6% at AM5 while its AM1.5G efficiency was 52.5% (Trupke et al. 2006). A single cell with an UC layer, having an AM1.5G efficiency of 50.7%, does not suffer that much from an increase in air mass: at AM5 the efficiency is lowered to 44.0% (Trupke et al. 2006). This so-called spectral robustness is due to the current matching constraints, which are much more relaxed in the cell/UC layer case.

Successfully optimizing absorption strength and quenching in lanthanides based down converters will bring the theoretical limits within reach. This also holds for up conversion, where, in addition triplet-triplet annihilation and mixed transition metal/lanthanide systems (Suyver et al. 2005a) constitute new material systems.

Future use of DS layers on top of solar cells may be limited as blue response of present cells will advance to higher levels (Van Sark 2006). On the other hand, these improvements might require additional expensive processing, while application of a DS layer is expected to be low-cost, as it only involves coating of a plastic with dispersed luminescent species.

LSC development will focus on material systems that: 1) absorb all photons with wavelength > 950 nm, and emit them red-shifted at ~1000 nm, for use with c-Si solar cells, 2) have as low as possible spectral overlap between absorption and emission spectra to minimize re-absorption losses; 3) have near unity luminescence quantum yield; 4) have low escape cone losses; 5) be stable outdoors for more than 10 years; 6) be easy to manufacture at low cost (Rowan et al. 2008). Much progress has occurred, which is illustrated by the recent efficiency record of 7.1% (Slooff et al. 2008) for organic dyes in PMMA. The present lack of NIR dyes will prohibit further increase of conversion efficiency towards the 30% limit. Here, quantum dots or nanorods may have to be used, as their broad absorption spectrum is very favourable. However, they should be emitting in the NIR at high quantum efficiency, larger than the present ~70-80%, and their Stokes’ shift should be larger. The latter would be possible as the size distribution of a batch of QDs influences the Stokes’ shift (Barnham et al. 2000). Another strategy could be the use of so-called type II QDs, as their Stokes’ shift could be very large (~300 nm), however, but their stability and QE are not good enough yet. Stability could be improved using multishell QDs (Koole et al. 2008), while interfacial alloying can be optimized to obtain type II QDs with desired properties, i.e., a Stokes’ shift of ~50-100 nm, without spectral overlap (Chin et al. 2007). An different approach was presented recently that employs resonance-shifting to circumvent reabsorption losses (Giebink et al. 2011). Also, Eu3+ has been employed succesfully to address self-absorption loss (Wang et al. 2011). Alternatively, the originally proposed three-plate stack (Goetzberger and Greubel 1977) could be further developed using perhaps a combination of organic dyes and nanocrystals, or even rare earth ions (Rowan et al. 2008), with optimized dedicated solar cells for each spectral region.

Advertisement

7. Conclusion

The possibility to tune chemical and physical properties in nanosized materials could have a transformational effect on the performance of solar cells and one of the prominent research areas of nanomaterials for photovoltaics involves spectral conversion. In this chapter we reviewed pathways of spectrum modification for application in solar cells by means of down- and upconversion, and downshifting. Nanoparticles, based on semiconductors or lanthanides, embedded in predominantly polymer layers on top or at the back of solar cells are an essential ingredient of spectral conversion mechanisms that may potentially lead to higher solar cell conversion efficiencies.

Increasing the conversion efficiency, while keeping manufacture cost as low as possible is the main R&D target in solar photovoltaic energy for the coming decades. Nanomaterial synthesis and nanotechnology will play a key role in this.

Advertisement

Acknowledgments

The authors gratefully acknowledge numerous colleagues at Utrecht University and elsewhere who contributed to the presented work. This work was financially supported by the European Commission as part of the Framework 6 integrated project FULLSPECTRUM (contract SES6-CT-2003-502620), SenterNovem as part of their Netherlands Nieuw Energie Onderzoek (New Energy Research) programme, Netherlands Foundation for Fundamental Research on Matter (FOM), and Netherlands Organisation for Scientific Research (NWO).

References

  1. 1. Alivisatos A. P. 1996 Perspectives on the Physical Chemistry of Semiconductor Nanocrystals. Journal of Physical Chemistry 100: 13226-13239.
  2. 2. ASTM. 2003 Standard Tables for Reference Solar Spectral Irradiances: Direct Normal and Hemispherical on 37 Tilted Surface, Standard G173-03e1 West Conshohocken, PA, USA: American Society for Testing and Materials.
  3. 3. Auzel, 2004 Upconversion and Anti-Stokes Processes with f and d Ions in Solids. Chemical Reviews 104 139 173 .
  4. 4. Badescu V. Badescu A. M. 2009 Improved model for solar cells with up-conversion of low-energy photons. Renewable Energy 34 1538 1544 .
  5. 5. Badescu V. De Vos A. 2007 Influence of some design parameters on the efficiency of solar cells with down-conversion and down shifting of high-energy photons. Journal of Applied Physics 102: 073102-1- 073102-7.
  6. 6. Badescu V. De Vos A. Badescu A. M. Szymanska A. 2007 Improved model for solar cells with down-conversion and down-shifting of high-energy photons. Journal of Physics D: Applied Physics 40 341 352 .
  7. 7. Barnham K. Marques J. L. Hassard J. O’Brien P. 2000 Quantum-dot concentrator and thermodynamic model for the global redshift. Applied Physics Letters 76 1197 1199 .
  8. 8. Basore P. A. Clugston D. A. 1996 PC1D Version 4 for Windows: from Analysis to Design. In Proceedings of 25th IEEE Photovoltaic Specialists Conference, Eds. E. C. Boes, D. J. Flood, J. Schmid, and M. Yamaguchi, 377 381 . IEEE.
  9. 9. Bende E. E. Burgers A. R. Slooff L. H. Van Sark W. G. J. H. M. Kennedy M. 2008 Cost and Efficiency Optimisation of the Fluorescent Solar Concentrator. In Proceedings of Twenty third European Photovoltaic Solar Energy Conference, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 461 469 . WIP, Munich, Germany.
  10. 10. Blasse G. Grabmaier B. C. 1995 Luminescent Materials. Berlin, Germany: Springer.
  11. 11. Bloembergen N. 1959 Solid state infrared quantum counters. Physical Review Letters 2 84 85 .
  12. 12. Bomm J. Büchtemann A. Fiore A. et al. 2010 Fabrication and spectroscopic studies on highly luminescent CdSe nanorod polymer composites. Beilstein Journal of Nanotechnology 1 94 100 .
  13. 13. Bomm J. Büchtemann A. Chatten A. J. et al. 2011 Fabrication and full characterization of state-of-the-art CdSe quantum dot luminescent solar concentrators. Solar Energy Materials and Solar Cells 95 2087 2094 .
  14. 14. Boyer C. van Veggel F. C. J. M. 2010 Absolute Quantum Yield Measurements of Colloidal NaYF4: Er3+, Yb3+ Upconverting Nanoparticles, Nanoscale 2 1417 1419
  15. 15. Bruchez M. Jr. Moronne M. Gin P. Weiss S. Alivisatos A. P. 1998 Semiconductor Nanocrystals as Fluorescent Biological Labels. Science 281 2013 2016 .
  16. 16. Burgers A. R. Slooff L. H. Kinderman R. van Roosmalen J. A. M. 2005 Modeling of luminescent concentrators by ray-tracing. In Proceedings of Twentieth European Photovoltaic Solar Energy Conference, Eds. W. Hoffmann, J.-L. Bal, H. Ossenbrink, W. Palz, and P. Helm, 394 397 . WIP, Munich, Germany.
  17. 17. Chatten A. J. Barnham K. W. J. Buxton B. F. Ekins-Daukes N. J. Malik M. A. 2003a A new approach to modelling quantum dot concentrators. Solar Energy Materials and Solar Cells 75 363 371 .
  18. 18. Chatten A. J. Barnham K. W. J. Buxton B. F. Ekins-Daukes N. J. Malik M. A. 2003b The Quantum Dot Concentrator: Theory and Results. In Proceedings of Third World Congress on Photovoltaic Energy Conversion (WPEC-3), Eds. K. Kurokawa, L. Kazmerski, B. McNelis, M. Yamaguchi, C. Wronski, and W. C. Sinke, 2657 2660 .
  19. 19. Chatten A. J. Barnham K. W. J. Buxton B. F. Ekins-Daukes N. J. Malik M. A. 2004a Quantum Dot Solar Concentrators. Semiconductors 38 909 917 .
  20. 20. Chatten A. J. Barnham K. W. J. Buxton B. F. Ekins-Daukes N. J. Malik M. A. 2004b Quantum Dot Solar Concentrators and Modules. In Proceedings of 19th European Photovoltaic Solar Energy Conference, Eds. W. Hoffmann, J.-L. Bal, H. Ossenbrink, W. Palz, and P. Helm, 109 112 . WIP, Munich, Germany; ETA, Florence, Italy.
  21. 21. Chatten A. J. Farrell D. J. Bose R. et al. 2007 Thermodynamic Modelling of Luminescent Solar Concentrators With Reduced Top Surface Losses. In Proceedings of Twenty Second European Photovoltaic Solar Energy Conference, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 349 353 . WIP, Munich, Germany.
  22. 22. Chatten A. J. Farrell D. J. Jermyn C. M. et al. 2005 Thermodynamic Modelling of the Luminescent Solar Concentrator. In Proceedings of 31st IEEE Photovoltaic Specialists Conference, Eds. 82 85 . IEEE.
  23. 23. Chen D. Wang Y. Yu N. Huang P. Weng F. 2008a Near-infrared quantum cutting in transparent nanostructured glass ceramics. Optics Letters 33 1884 1886 .
  24. 24. Chen D. Wang Y. Yu Y. Huang P. Weng F. 2008b Quantum cutting downconversion by cooperative energy transfer from Ce3+ to Yb3+ in borate glasses. Journal of Applied Physics 104: 116105-1- 116105-3.
  25. 25. Chin P. T. K. De Mello Donegá C. Van Bavel S. S. et al. 2007 Highly Luminescent CdTe/CdSe Colloidal Heteronanocrystals with Temperature-Dependent Emission Color. Journal of the American Chemical Society 129 14880 14886 .
  26. 26. Chung P. Chung H.-H. Holloway P. H. 2007 Phosphor coatings to enhance Si photovoltaic cell performance. Journal of Vacuum Science and Technology A 25 61 66 .
  27. 27. Currie M. J. Mapel J. K. Heidel T. D. Goffri S. Baldo M. A. 2008 High-Efficiency Organic Solar Concentrators for Photovoltaics. Science 321 226 228 .
  28. 28. De Vos A. Szymanska A. Badescu V. 2009 Modelling of solar cells with down-conversion of high energy photons, anti-reflection coatings and light trapping Energy Conversion and Management 50 328 336 .
  29. 29. De Wild J. Meijerink A. Rath J. K. van Sark W. G. J. H. M. Schropp R. E. I. 2011 Upconverter solar cells: materials and applications. Energy & Environmental Science 4 4835 4848
  30. 30. Debije M. G. Van der Blom R. H. L. Broer D. J. Bastiaansen C. W. M. 2006 Using selectively-reflecting organic mirrors to improve light output from a luminescent solar concentrator. In Proceedings of World Renewable Energy Congress IX.
  31. 31. Del Cañizo C. Tobias I. Pérez-Bedmar J. Pan A. C. Luque A. 2008 Implementation of a Monte Carlo method to model photon conversion for solar cells. Thin Solid Films 516 6757 6762 .
  32. 32. Dexter D. 1953 A theory of sensitized luminescence in solids. Journal of Chemical Physics 21 836 850 .
  33. 33. Dexter D. 1957 Possibility of luminescent quantum yields greater than unity. Physical Review 108 630 633 .
  34. 34. Díaz-Herrera B. González-Díaz B. Guerrero-Lemus R. et al. 2009 Photoluminescence of porous silicon stain etched and doped with erbium and ytterbium. Physica E 41 525 528 .
  35. 35. Dieke G. 1968 Spectra and Energy Levels of Rare Earth Ions in Crystals. New York, NY, USA: Wiley Interscience.
  36. 36. Eilers J. J. Biner D. van Wijngaarden J. T. Kraemer K. Guedel H.-U. Meijerink A. 2010 Efficient visible to infrared quantum cutting through downconversion with the Er(3+)-Yb(3+) couple in Cs(3)Y(2)Br(9), Appl. Phys. Lett. 96: 151106.
  37. 37. Fujii M. Yoshida M. Kanzawa Y. Hayashi S. Yamamoto K. 1997 1.54 µm photoluminescence of Er3+ doped into SiO2 films containing Si nanocrystals: Evidence for energy transfer from Si nanocrystals to Er3+. Applied Physics Letters 71 1198 1200 .
  38. 38. Gallagher S. J. Eames P. C. Norton B. 2004 Quantum dot solar concentrator behaviour predicted using a ray trace approach. Journal of Ambient Energy 25 47 56 .
  39. 39. Gallagher S. J. Rowan B. C. Doran J. Norton B. 2007 Quantum dot solar concentrator: Device optimisation using spectroscopic techniques. Solar Energy 81: 540.
  40. 40. Gamelin D. R. Güdel H. U. 2001 Upconversion Processes in Transition Metal and Rare Earth Metal Systems. Topics in Current Chemistry 214 1 56 .
  41. 41. Gaponenko S. 1998 Optical Properties of Semiconductor Nanocrystals. Cambridge, U.K.: Cambridge University Press.
  42. 42. Garwin R. 1960 The Collection of Light from Scintillation Counters. Review of Scientific Instruments 31 1010 1011 .
  43. 43. Gibart P. Auzel F. Guillaume J.-C. Zahraman K. 1996 Below band-gap IR response of substrate-free GaAs solar cells using two-photon up-conversion. Japanese Journal of Applied Physics 351 4401 4402 .
  44. 44. Giebink N. C. Wiederrecht G. P. Wasielewski M. R. 2011 Resonance-shifting to circumvent reabsorption loss in luminescent solar concentrators. Nature Photonics 5 694 701 .
  45. 45. Glaeser G. C. Rau U. 2007 Improvement of photon collection in Cu(In,Ga)Se2 solar cells and modules by fluorescent frequency conversion. Thin Solid Films 515: 5964.
  46. 46. Goetzberger A. 2008 Fluorescent Solar Energy Concentrators: Principle and Present State of Development. in High-Efficient Low-Cost Photovoltaics- Recent Developments, Eds. V. Petrova-Koch, R. Hezel, and A. Goetzberger, 159 176 . Heidelberg, Germany: Springer.
  47. 47. Goetzberger A. Greubel W. 1977 Solar Energy Conversion with Fluorescent Collectors. Applied Physics 14 123 139 .
  48. 48. Goldschmidt J. C. Peters M. Bösch A. et al. 2009 Increasing the efficiency of fluorescent concentrator systems Solar Energy Materials and Solar Cells 93 176 182 .
  49. 49. González-Díaz B. Díaz-Herrera B. Guerrero-Lemus R. et al. 2008 Erbium doped stain etched porous silicon. Materials Science and Engineering B 146 171 174 .
  50. 50. Green M. Emery K. Hishikawa Y. Warta W. Dunlop E. D. 2011 Solar Cell Efficiency Tables (Version 38). Progress in Photovoltaics: Research and Applications 19 565 572 .
  51. 51. Green M. 1982 Solar Cells; Operating Principles, Technology and Systems Application. Englewood Cliffs, NJ, USA: Prentice-Hall.
  52. 52. Green M. 2003 Third Generation Photovoltaics, Advanced Solar Energy Conversion. Berlin, Germany: Springer Verlag.
  53. 53. Haase M. Schafer H. 2011 Upconverting Nanoparticles, Angewandte Chemie, Int. Ed. 50 5808 5829 .
  54. 54. Hong B.-C. Kawano K. 2003 PL and PLE studies of KMgF3:Sm crystal and the effect of its wavelength conversion on CdS/CdTe solar cell. Solar Energy Materials and Solar Cells 80 417 432 .
  55. 55. Houshyani Hassanzadeh B. De Keizer A. C. Reich N. H. Van Sark W. G. J. H. M. 2007 The effect of a varying solar spectrum on the energy performance of solar cells. In Proceedings of 21st European Photovoltaic Solar Energy Conference, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 2652 2658 . WIP-Renewable Energies, Munich, Germany.
  56. 56. Hovel H. J. Hodgson R. T. Woodall J. M. 1979 The effect of fluorescent wavelength shifting on solar cell spectral response. Solar Energy Materials 2 19 29 .
  57. 57. Hyldahl M. G. Bailey S. T. Wittmershaus B. P. 2009 Photo-stability and performance of CdSe/ZnS quantum dots in luminescent solar concentrators. Solar Energy 83 566 573 .
  58. 58. Islangulov R. R. Lott J. Weder C. Castellano F. N. 2007 Noncoherent Low-Power Upconversion in Solid Polymer Films. Journal of the American Chemical Society 129 12652 12653 .
  59. 59. Kamat P. 2008 Quantum dot solar cells. Semiconductor nanocrystals as light harvesters. Journal of Physical Chemistry C 112: 18737-18753.
  60. 60. Kennedy M. Chatten A. J. Farrell D. J. et al. 2008 Luminescent solar concentrators: a comparison of thermodynamic modelling and ray-trace modelling predictions. In Proceedings of Twenty third European Photovoltaic Solar Energy Conference, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 334 337 . WIP, Munich, Germany.
  61. 61. Klampaftis E. Ross D. McIntosh K. R. Richards B. S. 2009 Enhancing the performance of solar cells via luminescent down-shifting of the incident spectrum: A review. Solar Energy Materials and Solar Cells 93 1182 1194 .
  62. 62. Klimov V. 2006 Mechanisms for Photogeneration and Recombination of Multiexcitons in Semiconductor Nanocrystals: Implications for Lasing and Solar Energy Conversion. Journal of Physical Chemistry B 110 16827 16845 .
  63. 63. Klimov V. I. Ivanov S. A. Nanda J. et al. 2007 Single-exciton optical gain in semiconductor nanocrystals. Nature 447 441 446 .
  64. 64. Koole R. Liljeroth P. De Mello Donegá C. Vanmaekelbergh D. Meijerink A. 2006 Electronic Coupling and Exciton Energy Transfer in CdTe Quantum-Dot Molecules. Journal of the American Chemical Society 128 10436 10441 .
  65. 65. Koole R. Van Schooneveld M. Hilhorst J. et al. 2008 On the Incorporation Mechanism of Hydrophobic Quantum Dots in Silica Spheres by a Reverse Microemulsion Method. Chemistry of Materials 20 2503 2512 .
  66. 66. Krishnan P. Schüttauf J. W. A. Van der Werf C. H. M. et al. 2009 Response to simulated typical daily outdoor irradiation conditions of thin-film silicon-based triple-band-gap, triple-junction solar cells. Solar Energy Materials and Solar Cells 93 691 697 .
  67. 67. Lakshminarayana G. Yang H. Ye S. Liu Y. Qiu J. 2008 Cooperative downconversion luminescence in Pr3+/Yb3+:SiO2-Al2O3-BaF2-GdF3 glasses. Journal of Materials Research 23 3090 3095 .
  68. 68. Law D. C. King R. R. Yoon H. et al. 2010 Future technology pathways of terrestrial III-V multijunction solar cells for concentrator photovoltaic systems Solar Energy Materials and Solar Cells 94 1314 1318 .
  69. 69. Lee J. Sundar V. C. Heine J. R. Bawendi M. G. Jensen K. F. 2000 Full Color Emission from II-VI Semiconductor Quantum Dot-Polymer Composites. Advanced Materials 12 1102 1105 .
  70. 70. Liu X. Qiao Y. Dong G. et al. 2008 Cooperative downconversion in Yb3+-RE3+ (RE = Tm or Pr) codoped lanthanum borogermanate glasses. Optics Letters 33 2858 2860 .
  71. 71. Luque A. Hegedus S. Eds. Handbook of Photovoltaic Science and Engineering. 2003 Wiley: Chichester, U.K.
  72. 72. Luque A. Marti A. 1997 Increasing the Efficiency of Ideal Solar Cells by Photon Induced Transitions at Intermediate Levels. Physical Review Letters 78 5014 5017 .
  73. 73. Luque A. Martí A. 2003 Theoretical Limits of Photovoltaic Conversion. in Handbook of Photovoltaic Science and Engineering, Eds. A. Luque and S. Hegedus, 113 149 . Chichester, U.K.: Wiley.
  74. 74. Luque A. Martí A. Bett A. et al. 2005 FULLSPECTRUM: a new PV wave making more efficient use of the solar spectrum. Solar Energy Materials and Solar Cells 87 467 479 .
  75. 75. Martí A. Luque A. eds 2004 Next Generation Photovoltaics, High Efficiency through Full Spectrum Utilization. Series in Optics and Optoelectronics, ed. R. G. W. Brown and E. R. Pike. 2004, Institute of Physics Publishing: Bristol, UK.
  76. 76. Maruyama T. Bandai J. 1999 Solar Cell Module Coated with Fluorescent Coloring Agent. Journal of the Electrochemical Society 146 4406 4409 .
  77. 77. Maruyama T. Kitamura R. 2001 Transformations of the wavelength of the light incident upon CdS/CdTe solar cells. Solar Energy Materials and Solar Cells 69 61 68 .
  78. 78. McIntosh K. R. Lau G. Cotsell J. N. Hanton K. Bätzner D. L. 2009 Increase in External Quantum Efficiency of Encapsulated Silicon Solar Cells from a Luminescent Down-Shifting Layer. Progress in Photovoltaics: Research and Applications 17 191 197 .
  79. 79. Meijerink A. Wegh R. Vergeer P. Vlugt T. 2006 Photon management with lanthanides Optical Materials 28 575 581 .
  80. 80. Minemoto T. Toda M. Nagae S. et al. 2007 Effect of spectral irradiance distribution on the outdoor performance of amorphous Si//thin-film crystalline Si stacked photovoltaic modules. Solar Energy Materials and Solar Cells 91 120 122 .
  81. 81. Munyeme G. Zeman M. Schropp R. E. I. Van der Weg W. F. 2004 Performance analysis of a-Si:H p-i-n solar cells with and without a buffer layer at the p/i interface. physica status solidi C 9 2298 2303 .
  82. 82. Mutlugun E. Soganci I. M. Demir H. V. 2008 Photovoltaic nanocrystal scintillators hybridized on Si solar cells for enhanced conversion efficiency in UV. Optics Express 16 3537 3545 .
  83. 83. Nann S. Riordan C. 1991 Solar Spectral Irradiance under Clear and Cloudy Skies: Measurements and a Semiempirical Model. Journal of Applied Meteorology 30 447 462 .
  84. 84. O’Regan B. Grätzel M. 1991 A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 353 737 740 .
  85. 85. Ohwaki J. Wang Y. 1994 Efficient 1.5 µm to Visible Upconversion in Er3+ Doped Halide Phosphors. Japanese Journal of Applied Physics 33: L334 -L337.
  86. 86. Peijzel P. S. Meijerink A. Wegh R. T. Reid M. F. Burdick G. W. 2005 A complete 4fn energy level diagram for all trivalent lanthanide ions Journal of Solid State Chemistry 178 448 453 .
  87. 87. Peng X. Schlamp M. C. Kadavanich A. V. Alivisatos A. P. 1997 Epitaxial Growth of Highly Luminescent CdSe/CdS Core/Shell Nanocrystals with Photostability and Electronic Accessibility. Journal of the American Chemical Society 119 7019 7029 .
  88. 88. Piper W. W. De Luca J. A. Ham F. S. 1974 Cascade fluorescent decay in Pr3+-doped fluorides: achievement of a quantum yield greater than unity for emission of visible light. Journal of Luminescence 8 344 348 .
  89. 89. Reda S. 2008 Synthesis and optical properties of CdS quantum dots embedded in silica matrix thin films and their applications as luminescent solar concentrators Acta Materialia 56 259 264 .
  90. 90. Richards B. S. 2006a Enhancing the performance of silicon solar cells via the application of passive luminescence conversion layers. Solar Energy Materials and Solar Cells 90 2329 2337 .
  91. 91. Richards B. S. 2006b Luminescent layers for enhanced silicon solar cell performance: Down-conversion. Solar Energy Materials and Solar Cells 90: 1189.
  92. 92. Richards B. S. Shalav A. 2005 The role of polymers in the luminescence conversion of sunlight for enhanced solar cell performance. Synthetic Metals 154 61 64 .
  93. 93. Rowan B. C. Wilson L. R. Richards B. S. 2008 Advanced Material Concepts for Luminescent Solar Concentrators. IEEE Journal of Selected Topics in Quantum Electronics 14 1312 1322 .
  94. 94. Schropp R. E. I. Zeman M. 1998 Amorphous and microcrystalline silicon solar cells: Modeling, Materials, and Device Technology. Boston, MA, USA: Kluwer Academic Publishers.
  95. 95. Schüler A. Python M. Valle del Olmo M. de Chambrier E. 2007 Quantum dot containing nanocomposite thin films for photoluminescent solar concentrators. Solar Energy 81 1159 1165 .
  96. 96. Serin J. M. Brousmiche D. W. Frechet J. M. J. 2002 A FRET-Based Ultraviolet to Near-Infrared Frequency Convertor. Journal of the American Chemical Society 124 11848 11849 .
  97. 97. Shalav A. Richards B. S. Green M. A. 2007 Luminescent layers for enhanced silicon solar cell performance: up-conversion. Solar Energy Materials and Solar Cells 91 829 842 .
  98. 98. Shalav A. Richards B. S. Trupke T. Krämer K. W. Güdel H. U. 2005 Application of NaYF4:Er3+ up-converting phosphors for enhanced near-infrared silicon solar cell response. Applied Physics Letters 86: 013505-1- 013505-3.
  99. 99. Shockley W. Queisser H. J. 1961 Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. Journal of Applied Physics 32: 510.
  100. 100. Singh-Rachford T. N. Haefele A. Ziessel R. Castellano F. N. 2008 Boron Dipyrromethene Chromophores: Next Generation Triplet Acceptors/Annihilators for Low Power Upconversion Schemes. Journal of the American Chemical Society 130 16164 16165 .
  101. 101. Singh-Rachford T. N. Castellano F. N. 2010 Low Power Photon Upconversion based on Sensitized Triplet-Triplet Annihilation, Coordination Chemistry Review, 254 2560 2573 .
  102. 102. Slooff L. H. Bende E. E. Burgers A. R. et al. 2008 A luminescent solar concentrator with 7.1% power conversion efficiency. Physica Status Solidi- Rapid Research Letters 2 257 259 .
  103. 103. Slooff L. H. Kinderman R. Burgers A. R. et al. 2006 The luminescent concentrator illuminated. Proceedings of SPIE 6197: 61970k1-8
  104. 104. Smestad G. Ries H. Winston R. Yablonovitch E. 1990 The thermodynamic limits of light concentrators. Solar Energy Materials 21 99 111 .
  105. 105. Soga T. Ed. Nanostructured Materials for Solar Energy Conversion. 2006 Elsevier: Amsterdam, the Netherlands.
  106. 106. Sommerdijk J. L. Bril A. De Jager A. W. 1974 Two photon luminescence with ultraviolet excitation of trivalent praseodymium. Journal of Luminescence 8 341 343 .
  107. 107. Strümpel C. McCann M. Beaucarne G. et al. 2007 Modifying the solar spectrum to enhance silicon solar cell efficiency--An overview of available materials. Solar Energy Materials and Solar Cells 91: 238.
  108. 108. Strümpel C. McCann M. Del Cañizo C. Tobias I. Fath P. 2005 Erbium-doped up-converters on silicon solar cells: assesment of the potential. In Proceedings of Twentieth European Photovoltaic Solar Energy Conference, Eds. W. Hoffmann, J.-L. Bal, H. Ossenbrink, W. Palz, and P. Helm, 43 46 . WIP, Munich, Germany.
  109. 109. Stupca M. Alsalhi M. Al Saud T. Almuhanna A. Nayfeh M. H. 2007 Enhancement of polycrystalline silicon solar cells using ultrathin films of silicon nanoparticle Applied Physics Letters 91: 063107-1- 063107-3.
  110. 110. Suyver J. F. Aebischer A. Biner D. et al. 2005a Novel materials doped with trivalent lanthanides and transition metal ions showing near-infrared to visible photon upconversion Optical Materials 27 1111 1130 .
  111. 111. Suyver J. F. Grimm J. Krämer K. W. Güdel H. U. 2005b Highly efficient near-infrared to visible up-conversion process in NaYF4:Er3+, Yb3+. Journal of Luminescence 114 53 59 .
  112. 112. Svrcek V. Slaoui A. J. Muller C. 2004 Silicon nanocrystals as light converter for solar cells. Thin Solid Films 451-452: 384-388.
  113. 113. Timmerman D. Izeddin I. Stallinga P. Yassievich I. N. Gregorkiewicz T. 2008 Space-separated quantum cutting with silicon nanocrystals for photovoltaic applications. Nature Photonics 2 105 109 .
  114. 114. Timmerman D. Valenta J. Dohnalova K. de Boer W. D. A. M. Gregorkiewicz T. 2011 Step-like enhancement of luminescence quantum yield of silicon nanocrystals. Nature Nanotechnology 6.
  115. 115. Trupke T. Green M. A. Würfel P. 2002a Improving solar cell efficiencies by down-conversion of high-energy photons. Journal of Applied Physics 92 1668 1674 .
  116. 116. Trupke T. Green M. A. Würfel P. 2002b Improving solar cell efficiencies by up-conversion of sub-band-gap light. Journal of Applied Physics 92 4117 4122 .
  117. 117. Trupke T. Shalav A. Richards B. S. Wurfel P. Green M. A. 2006 Efficiency enhancement of solar cells by luminescent up-conversion of sunlight. Solar Energy Materials and Solar Cells 90: 3327.
  118. 118. Tsakalakos L. 2008 Nanostructures for photovoltaics. Materials Science and Engineering: R: Reports 62 175 189 .
  119. 119. Van der Ende B. M. Aarts L. Meijerink A. 2009a Near infrared quantum cutting for photovoltaics. Advanced Materials 21 3073 3077 .
  120. 120. Van der Ende B. M. Aarts L. Meijerink A. 2009b Lanthanide ions as spectral converters for solar cells, Phys. Chem. Chem. Phys. 11 11081 11095 .
  121. 121. Van Sark W. G. J. H. 2002 Methods of Deposition of Hydrogenated Amorphous Silicon for Device Applications. in Thin Films and Nanostructures, Eds. M. H. Francombe, 1 215 . San Diego: Academic Press.
  122. 122. Van Sark W. G. J. H. 2005 Enhancement of solar cell performance by employing planar spectral converters. Applied Physics Letters 87: 151117.
  123. 123. Van Sark W. G. J. H. 2006 Optimization of the performance of solar cells with spectral down converters. In Proceedings of Twentyfirst European Photovoltaic Solar Energy Conference, Eds. J. Poortmans, H. Ossenbrink, E. Dunlop, and P. Helm, 155 159 . WIP, Munich, Germany.
  124. 124. Van Sark W. G. J. H. 2007 Calculation of the performance of solar cells with spectral down shifters using realistic outdoor solar spectra. In Proceedings of Twenty Second European Photovoltaic Solar Energy Conference,, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 566 570 . WIP, Munich, Germany.
  125. 125. Van Sark W. G. J. H. 2008 Simulating performance of solar cells with spectral downshifting layers. Thin Solid Films 516 6808 6812 .
  126. 126. Van Sark W. G. J. H. M. Barnham K. W. J. Slooff L. H. et al. 2008a Luminescent Solar Concentrators- A review of recent results. Optics Express 16 21773 21792 .
  127. 127. Van Sark W. G. J. H. M. Frederix P. L. T. M. Bol A. A. Gerritsen H. C. Meijerink A. 2002 Blinking, Blueing, and Bleaching of single CdSe/ZnS Quantum Dots. ChemPhysChem 3 871 879 .
  128. 128. Van Sark W. G. J. H. M. Hellenbrand G. F. M. G. Bende E. E. Burgers A. R. Slooff L. H. 2008b Annual energy yield of the fluorescent solar concentrator. In Proceedings of Twenty third European Photovoltaic Solar Energy Conference, Eds. G. Willeke, H. Ossenbrink, and P. Helm, 198 202 . WIP, Munich, Germany.
  129. 129. Van Sark W. G. J. H. M. Meijerink A. Schropp R. E. I. Van Roosmalen J. A. M. Lysen E. H. 2005 Enhancing solar cell efficiency by using spectral converters. Solar Energy Materials and Solar Cells 87 395 409 .
  130. 130. Van Wijngaarden J. T. Scheidelaar S. Vlugt T. J. H. Reid M. F. Meijerink 2010 Energy transfer mechanism for downconversion in the (Pr(3+), Yb(3+)) couple Phys. Rev. B 81 (15): 155112.
  131. 131. Vergeer P. Vlugt T. J. H. Kox M. H. F. et al. 2005 Quantum cutting by cooperative energy transfer in YbxY1−xPO4 : Tb3+ Physical Review B 71: 014119-1- 014119-11.
  132. 132. Walker G. W. Sundar V. C. Rudzinski C. M. et al. 2003 Quantum-dot optical temperature probes. Applied Physics Letters 83 3555 3557 .
  133. 133. Wang T. Zhang J. Ma Luo W. Y. et al. 2011 Luminescent solar concentrator employing rare earth complex with zero self-absorption loss. Solar Energy 85 2571 2579 .
  134. 134. Wegh R. Donker H. Meijerink A. Lamminmäki R. J. Hölsä J. 1997 Vacuum-ultraviolet spectroscopy and quantum cutting for Gd3+ in LiYF4. Physical Review B 56 13841 13848 .
  135. 135. Wegh R. T. Donker H. Oskam K. D. Meijerink A. 1999 Visible Quantum Cutting in LiGdF4:Eu3+ Through Downconversion. Science 283 663 666 .
  136. 136. Wegh R. T. Meijerink A. Lamminmäki R. J. Hölsä J. 2000 Extending Dieke’s diagram. 87-89: 1002-1004.
  137. 137. Wittwer V. Stahl W. Goetzberger A. 1984 Fluorescent planar concentrators Solar Energy Materials 11 187 197 .
  138. 138. Wolf M. 1971 New look at silicon solar cell performance. Energy Conversion 11 63 73 .
  139. 139. Yablonovitch E. 1980 Thermodynamics of the fluorescent planar concentrator. Journal of the Optical Society of America 70 1362 1363 .
  140. 140. Yuan Z. Pucker G. Marconi A. Sgrignuoli F. et al. 2011 Silicon nanocrystals as a photoluminescence down shifter for solar cells: 95 1224 1227 .
  141. 141. Zastrow A. 1994 The physics and applications of fluorescent concentrators: a review. Proceedings of SPIE 2255 534 547 .
  142. 142. Zhang Q. Y. Yang G. F. Pan Y. X. 2007 Cooperative quantum cutting in one-dimensional (YbxGd1−x)Al3(BO3)4):Tb3+ nanorods Applied Physics Letters 90: 021107-1- 021107-3.

Written By

W.G.J.H.M. van Sark, A. Meijerink and R.E.I. Schropp

Submitted: 25 November 2011 Published: 16 March 2012